surface chemistry of molybdena containing...

117
THESIS OF THE DOCTORAL (PhD) DISSERTATION SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written by: HEIDER NASZER M.Sc. environmental engineer Supervisor: Prof. Dr. ÁKOS RÉDEY Doctoral School of Chemical Engineering and Material Sciences INSTITUTIONAL DEPARTMENT OF ENVIRONMENTAL ENGINEERING AND CHEMICAL TECHNOLOGY FACULTY OF ENGINEERING UNIVERSITY OF PANNONIA Veszprém 2008

Upload: trandiep

Post on 30-Apr-2019

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

THESIS OF THE DOCTORAL (PhD) DISSERTATION

SURFACE CHEMISTRY OF MOLYBDENA CONTAINING

CATALYSTS

Written by:

HEIDER NASZER

M.Sc. environmental engineer

Supervisor: Prof. Dr. ÁKOS RÉDEY

Doctoral School of Chemical Engineering and Material Sciences

INSTITUTIONAL DEPARTMENT OF ENVIRONMENTAL ENGINEERING

AND CHEMICAL TECHNOLOGY

FACULTY OF ENGINEERING

UNIVERSITY OF PANNONIA

Veszprém

2008

Page 2: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

DOKTORI (PhD) ÉRTEKEZÉS

MOLIBDÉN TARTALMÚ KATALIZÁTOROK FELÜLETKÉMIAI

TULAJDONSÁGAINAK VIZSGÁLATA

Készítette: NASZER HEIDER

okleveles környezetmérnök

Témavezető:

Dr. RÉDEY ÁKOS

egyetemi tanár

Készült a Pannon Egyetem

Vegyészmérnöki Tudományok és Anyagtudományok Doktori Iskola Keretében

PANNON EGYETEM

MÉRNÖKI KAR

KÖRNYEZETMÉRNÖKI ÉS KÉMIAI TECHNOLÓGIA INTÉZETI TANSZÉK

Veszprém

2008

2

Page 3: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

MOLIBDÉN TARTALMÚ KATALIZÁTOROK FELÜLETKÉMIAI

TULAJDONSÁGAINAK VIZSGÁLATA

Értekezés doktori (PhD) fokozat elnyerése érdekében

Írta:

NASZER HEIDER

Készült a Pannon Egyetem Vegyészmérnöki Tudományok és Anyagtudományok Doktori

Iskolájához tartozóan. Témavezető: Dr. RÉDEY ÁKOS Elfogadásra javaslom (igen / nem) ………………………. (aláírás) A jelölt a doktori szigorlaton…......... % -ot ért el. Az értekezést bírálóként elfogadásra javaslom: Bíráló neve: …........................…................. (igen /nem) ………………………. (aláírás) Bíráló neve: …........................…................. (igen /nem) ………………………. (aláírás) A jelölt az értekezés nyilvános vitáján…..........% - ot ért el. Veszprém, ………………………….

A Bíráló Bizottság elnöke A doktori (PhD) oklevél minősítése…................................. ………………………… Az EDT elnöke

3

Page 4: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

CONTENTS

CHAPTER 1. INTRODUCTION AND LITERARY OVERVIEW 8

1.1. γ-Al2O3 and molybdena-alumina catalysts 8

1.1.1. The surface hydroxyls of γ-Al2O3 and molybdena-alumina 13

1.2. Current status of catalysts preparation 17

1.2.1. The impregnation method 17

1.2.2. Adsorption methods 18

1.3. Molybdena with CeO2 and SnO2 semiconductors 22

1.4. Direct conversion of methane under nonoxidative conditions 27

1.5. Characterization of the catalysts 31

1.5.1. Physical characterization 32

1.5.2. Surface characterization and IR spectroscopy 34

1.5.3. FTIR spectroscopic detection of CO adsorption on the catalysts 37

1.6. OBJECTIVES 40

CHAPTER 2. EXPERIMENTAL 41

2.1. Materials and catalyst preparation 41

2.2. Catalyst characterization methods and techniques 42

CHAPTER 3. RESULTS 44

3.1. Surface texturing 44

3.2. X-ray diffraction 48

3.2.1. XRD patterns of Mo/Al2O3, Ce-Mo/Al2O3 and Mo/CeO2 48

3.2.2. XRD patterns of Sn-Mo/Al2O3 and Mo/SnO2 51

3.3. Thermal analysis 56

3.3.1. TG and DTA of Mo/Al2O3 56

3.3.2. TG and DTA of Ce-Mo/Al2O3 and Mo/CeO2 58

3.3.3. TG and DTA of Sn-Mo/Al2O3 and Mo/SnO2 60

3.4. Electron Spin Resonance (ESR) measurements 62

4

Page 5: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5. In situ DRIFT spectroscopy measurements 65

3.5.1. DRIFT spectra of the calcined samples under vacuum 65

3.5.1.1. DRIFT spectra of γ-Al2O3 and CeO2 65

3.5.1.2. DRIFT spectra of Mo/Al2O3, Mo/CeO2 and Ce-Mo/Al2O3 66

3.5.1.3. DRIFT spectra of SnO2, Mo/SnO2 and Sn-Mo/Al2O3 68

3.5.2. CO chemisorption 70

3.5.2.1. CO chemisorption on Mo/Al2O3 70

3.5.2.2. CO chemisorption on CeO2, Mo/CeO2 and Ce-Mo/Al2O3 72

3.5.2.3. CO chemisorption on SnO2, Mo/SnO2 and Sn-Mo/Al2O3 74

3.5.3. In situ DRIFT results on methane transformation in absence of oxygen 76

3.5.3.1. Methane transformation on Mo/Al2O3 77

3.5.3.2. Methane transformation on Ce-Mo/Al2O3 78

3.5.3.3. Methane transformation on Mo/CeO2 79

3.5.3.4. Methane transformation on Sn-Mo/Al2O3 80

3.5.3.5. Methane transformation on Mo/SnO2 81

3.5.3.6. DRIFT spectra after methane reaction under vacuum 82

CHAPTER 4. DISCUSSION 84

4.1. Surface texturing 84

4.2. X-ray diffraction 84

4.3. Thermal analysis 86

4.4. Electron Spin Resonance (ESR) 87

4.5. In situ DRIFT spectroscopy 88

4.5.1. DRIFT spectra of the calcined samples under vacuum 88

4.5.2. CO chemisorption 89

4.5.3. In situ DRIFT studies on methane transformation in absence of oxygen 93

SUMMARY AND CONCLUSIONS 104

ACKNOWLEDGMENTS 107

REFERENCES 108

THESES 113

PUBLICATIONS 115

5

Page 6: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

KIVONAT

Az impregnálással és mechanikus keveréssel készült Mo/Al2O3, Mo/CeO2, Mo/SnO2,

Ce-Mo/Al2O3 és Sn-Mo/Al2O3 katalizátorok struktúrájának, termikus stabilitásának, a hordozó

és a Mo közötti kölcsönhatásoknak, aktivitásuknak, a Mo ionok diszperziójának és a felületi

tulajdonságoknak vizsgálatára Induktív Csatolású Plazma (ICP), N2 adszorpció/deszorpció

(BET módszer), Termikus analízis (TG-DTA), X-ray diffrakció (XRD), Elektron Spin

Rezonancia (ESR) és Diffúz-reflexiós Fourier Transzformációs Infravörös (DRIFT)

módszereket alkalmaztam. A katalizátorok aktivitásának összehasonlítása CO adszorpció és a

CH4 konverziója alapján történt.

A BET, XRD és DRIFT eredmények összevetése alapján megállapítható, hogy a kalcinálás

hőmérséklete és időtartama, az oldat pH-ja és a hordozó izoelektromos pontja hatással vannak

a Mo ionok felületi diszperziójára. Másrészt, a cérium növelte a polimer Mo ionok felületi

koncentrációját (főleg hordozóként) cérium-molibdénátok formájában, illetve Mo-O-Ce

kötések megjelenésével (630, 875 cm-1), és újabb kettős O=Mo=O (995 és 1035 cm-1) kötések

jelentek meg, amelyek polimer MoO3 alakulatokra jellemzők.

A termikus analízis alapján megállapítható, hogy a Mo/Al2O3 termikusan a legstabilabb

(900°C-ig), ugyanakkor a Mo/CeO2, Mo/SnO2 és Ce-Mo/Al2O3 minták 700°C feletti

hőmérsékleten történő hőkezelése során morfológiai és kristályszerkezeti változások

következnek be, így lehetővé válik a fémionok diffúziója és kation csere játszódhat le köztük.

A szén monoxid 100°C-on történő kemiszorpciója során a karbonátok különböző formái és

fém-karbonil kapcsolódások jelentek meg, amelyek vákuumban stabilak maradtak. Viszont

ezek kialakulásához szükséges a katalizátorok redukált formája, továbbá a redukció során

kialakult koordinatívan telítetlen helyek (CUS), a kristályrácsban lévő oxigén ionok és a

felületi hidroxil csoportok jelentős szerepet játszanak. Másrészt, a Mo0 atomokat tartalmazó

alakulatok 700°C-on történő redukció után tűnnek fel, amelyekhez terminális, illetve

hídkötésű kapcsolódó szénmonoxid DRIFT sávjai jelentek meg (2025, 2002, 1994 cm-1).

A DRIFT vizsgálatok igazolták, hogy 800ºC-on redukált katalizátorokon a metán 700ºC-os

bontása során a felületi karbonátok különböző formái jelentek meg, illetve ezek bomlásából

származó CO és CO2 fejlődése megfigyelhető volt. Ez a rácsbeli oxigén reakcióképességének

és a hidrogén redukció során kialakult Lewis savas helyeknek köszönhető. Azonban, a

Mo/CeO2 és Mo/SnO2 katalizátorok jelentős aktivitást mutattak (főleg Mo/SnO2) a metán

reakcióban, ami feltehetően a nagymértékben diszpergált MoO3 klaszterek, illetve Ce+3/Ce+4

és Sn2+/Sn4+ redox párok együtthatásával magyarázható.

6

Page 7: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

ABSTRACT

Mo/Al2O3, Mo/CeO2, Mo/SnO2, Ce-Mo/Al2O3 and Sn-Mo/Al2O3 catalysts were prepared by

impregnation and co-precipitation methods. The catalysts were characterized by Inductive

Coupled Plasma (ICP), Thermal analysis (TG and DTA), X-ray diffractometry, Electron Spin

Resonance (ESR), Diffuse reflectance Infrared (DRIFT) spectroscopy as well as N2

adsorption/desorption (BET method) techniques to examine their structural characteristics,

thermal stability, mutual interactions between Mo and the support, catalytic activity as well as

the dispersity of Mo ions and surface structure. The samples were compared to determine

what kinds of adspecies participate efficiently during CO adsorption and CH4 decomposition.

Combining the results obtained from BET, XRD and DRIFT investigations one may suggest

that beyond the Mo dispersion the formation of MoO3 clusters was found to respond to the

calcination temperature and time as well as to the solution pH and the isoelectric point of the

solid support. On the other hand, the introduction of ceria resulted in different molecular

formulae with Mo (particularly as a support). This led to the increase of polymerized surface

Mo species so as to forming Mo-O-Ce (bands at 630 and 875 cm-1) linkages, besides the

formation of coupled O=Mo=O bonds at 995 and 1035 cm-1 indicative of polymeric MoO3.

From thermal analysis, it can be inferred that Mo/Al2O3 is the thermally most stable material

in the temperature range used in the experiment (up to 900°C). Whereas Mo/CeO2, Mo/SnO2

and Ce-Mo/Al2O3 samples undergo morphological and structural modifications above 700°C

resulting in lattice defects, which motivate the mobility of metal ions and thus enhance the

possibility of cation exchange between them.

Additionally, the formation of metal-carbonyl species and various types of carbonates through

CO chemisorption at 100°C needs reduced catalysts containing coordinatively unsaturated

sites (CUS), oxygen vacancies and hydroxyl groups. On the other hand, the bands protruding

at 2025, 2002 and 1994 cm-1 are very likely associated with the terminally and bridged CO σ-

bonded to metallic Mo(0) species appearing after reduction at 700°C. Thus, CO being provided

as weakly adsorbed metal-carbonyls migrating towards the oxides through interfacial sites to

form carbonates being stable even under vacuum at room temperature.

Methane is retained at 700ºC by the catalysts reduced at 800ºC and generates various

carbonate species which decompose to CO and CO2 implying the existence of reactive lattice

oxygen in addition to Lewis acid sites possessed by hydrogen reduction. However, the

Mo/CeO2 and Mo/SnO2 materials presented marked activity (especially Mo/SnO2) in CH4

decomposition. This activity is presumably related to highly dispersed MoO3 clusters besides

the Ce3+/Ce4+ and Sn2+/Sn4+ redox couples as further emphasized by means of DRIFT results.

7

Page 8: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

1. INTRODUCTION AND LITERARY OVERVIEW

1.1. γ-Al2O3 and molybdena-alumina catalysts Molybdena containing catalysts have been the subject of numerous studies due to their

potential role in different important reactions such as hydrodesulphurization (HDS),

hydrogenation (HYD), oxidative dehydrogenation (ODH), hydrodeoxygenation (HDO)

hydrodenitrogenation, hydrometallization, isomerization, epoxidation, partial oxidation of

alkanes and alkene metathesis reactions [1-5]. Presently, there is a continuous interest in Mo-

containing catalysts directed toward understanding the catalytic properties, beside extensive

studies have been addressed to characterize the structure of the catalyst surface. The chemical

studies of molybdena-alumina catalysts include reduction or sulphidation. These two

treatments are generally needed to activate the catalyst for most of the catalytic reactions [6].

The sulphidation occurs readily above 300°C, and the extent of sulphidation increases with

temperature. However, the catalyst sulphur content is limited at a given temperature. The

major sulphiding reaction is through the exchange of oxygen associated with molybdenum for

sulphur. The molybdena phase of the sulphided catalyst was found stoichiometrically close to

MoS2 [3, 6-10]. Perhaps, the most detailed analysis of the reduction of molybdena-alumina

catalyst is the works of Hall, Massoth, Wang and Rédey [3-12]. In their work, the overall

average oxidation state was determined from the gas consumed on oxidation or reduction and

an attempt was made to separate this average into percentages of Mo6+, Mo5+, Mo4+ and Mo3+

on the basis of H+ retained and vacancies created upon reduction with H2. It was found that

the reducibility of the catalyst after 2 hours reduction at 500°C under 1 atm H2 increased from

approximately e/Mo = 0.5 to e/Mo = 2 (e/Mo: electrons gained per Mo atom) as the Mo

loading increased from 2% to 10%, while the extent of reduction of 25% Mo loading catalyst

reached e/Mo = 2.7 in just one hour reduction. It was also noted that the extent of reduction at

short times for the 10% and 25% catalysts were actually greater than for bulk MoO3 implying

the presence of a MoO3 phase on the catalysts but with a smaller particle size than the low

surface area of bulk MoO3. In view of all reduction studies, it has been generally supposed

that from the amounts of H2 consumed and water generated in the reduction step, an amount

of hydrogen retained on the reduced catalyst could be calculated. Two types of adsorbed

hydrogen were characterized: (1) reversibly adsorbed (HR) which is removable as H2 by

evacuation above 450°C, and (2) irreversibly retained (HI) which could not be removed by

evacuation. The latter could only be removed as H2O by either evacuation at higher

temperature or by reoxidation. In addition to the retained hydrogen, water was produced

8

Page 9: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

during the reduction. The amount of water increased with increasing the extent of reduction.

The formation of water was attributed to the removal of oxygen from the molybdena and

consequently created anion vacancies on the surface. On reoxidation, two processes occurred:

the anion vacancies were filled and the HI was reacted to H2O. Given these surface processes,

the following material balance equations were written:

[H2] = [WR] + HR +HI (Eq-1)

[O] = [WR] + HI (Eq-2)

[Wo] = HI – WA (Eq-3)

Where WR and Wo are the quantities of water produced in the reduction and reoxidation

steps, respectively, and WA is water retained by alumina after the reoxidation step

(ideally WA =0) [3, 7, 12].

The redox study was extended by Hall and Lo Jacono who calculated HI per molybdenum

(HI/Mo) and vacancies per molybdenum (□/Mo) at various extents of reduction (Hc/Mo) all

measured as atom/atom [13]. The quantity of hydrogen (Hc) consumed during the reduction

process can be expressed by the equation:

Hc = WR + HR + HI (Eq-4)

At low extents of reduction (Hc/Mo ≤ 1), two HI were present per vacancy created. However,

as reduction was further increased, HI reached a limiting level of about HI/Mo = 0.45 while

□/Mo increased more rapidly. The formation of HI and vacancies was interpreted by assuming

that molybdena forms an epitaxial monolayer on Al2O3. Accordingly, HI was taken as a

measure of Mo5+ and each vacancy was equated to one Mo4+. The reduction process was

pictured as follows:

9

Page 10: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 1. The reduction process of molybdena according to Hall and Lo Jacono

Later, FTIR study of the surface hydroxyl groups of the same catalyst was conducted by

Millman [14]. These results led to a minor modification in the reduction process of Mo/Al2O3

(Fig. 2). The spectra from catalysts reduced with H2 were indistinguishable in the OH region

from the alumina support or from the unreduced catalysts. Indeed, all five alumina OH bands

were found and no new ones were detected. Thus, the intensity increased on reduction was

10

Page 11: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

probably due to the recoveries of Al-OH rather than the formation of Mo-OH. Accordingly,

the authors presented a modified picture for the reduction process of Mo/Al2O3:

Fig. 2. The reduction process of molybdena according to Millman

The implication is, however, that reduction reverses, the process of monolayer formation as

evidenced by the increased intensities of the OH region.

Other studies have suggested that patches of molybdena may be present on the surface,

particularly at high Mo loading. Giordano et al. studied molybdena-alumina catalysts

containing up to 30% Mo by various chemical and physical techniques [15]. These included

11

Page 12: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

differential thermal analysis (DTA), thermogravimetric analysis (TGA) infrared and diffuse

reflectance spectroscopy. The authors deduced that below 4% Mo loading the Mo ions are

highly dispersed in the form of MoO42-. At increasing Mo contents (up to 10-15%), a

progressive increase of structures with bridged oxygen atoms in mixed tetrahedral and

octahedral environment (schemes IV and V surrounding Mo6+) were found.

Scheme 1.

They obtained MoO3-rich catalysts at Mo loading higher than 15% prevailing octahedral

configuration attributed to species like schemes V and VI. It was clearly demonstrated by

these authors that the molybdena surface structure could be altered by the Mo concentration

and by the calcination procedures. The Mo concentration was found to play an important role

in the nature of the molybdena surface structure. They suggested that tetrahedral MoO42- is the

main species in the catalysts with less than 5% Mo. Octahedrally coordinated polymolybdates

were formed with more than 5% becoming the dominant species as the loading increased,

while crystalline Al2(MoO4)3 phase was observed in the catalysts calcined at elevated

temperatures (>600°C) and MoO3 clusters were formed when the loading reached 16% and

20%, respectively. The degree of the polymerization becomes higher and higher as the Mo

loading is further increased. After the surface has been saturated (high Mo content), the

formation of bulk MoO3 and Al2(MoO4)3 occurs at different extents depending on the

preparation and calcination conditions.

12

Page 13: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Another surface polymolybdate structure was proposed by Medema et al. to explain their

Raman results:

Scheme 2.

It was suggested that supported molybdena species are aligned as one-dimensional chains

which result in Mo-O-Mo bonds, yielding octahedral coordination for Mo6+ [16, 17].

Despite the variations made by various authors, a consistent pattern has emerged. At low Mo

loading, tetrahedral MoO42- species dominate the surface. Increasing loading is generally

accompanied by the formation of a polymolybdate phase. On the basis of these studies, it is

generally known that the surface of molybdena remains mainly containing monomeric Mo

species. Moreover, polymeric molybdate species and free MoO3 can be formed at high Mo

contents taking into account the calcination temperature and time as well as the solution pH,

the isoelectric point and surface area of the solid support [12-17].

1.1.1. The surface hydroxyls of γ-Al2O3 and molybdena-alumina γ-Al2O3 has a defect spinel structure, a high surface area, a certain degree of acidity, and

forms solid solutions with transition metal oxides such as NiO and CoO. Above 900°C, γ-

Al2O3 is transformed into α-Al2O3, which has hexagonal structure and smaller surface area.

Even at lower temperatures a slow phase transition occurs, which shortens the catalyst

lifetime. Therefore, the incorporation of small amounts (1-2%) of SiO2 or ZrO2 in γ-Al2O3

shifts the γ → α transformation to higher temperature and increases the stability of the catalyst

[1, 2]. The alumina surface hydroxyl groups have been extensively studied by infrared

spectroscopy. Thermal dehydroxylation studies of γ-Al2O3 by IR spectroscopy were first

executed by Peri who concluded that Al-OH groups close to each other could form water, and

desorbed from the surface [18, 19]. The spectra displayed five different OH frequencies at

300°C for Al-OH groups. Assuming that the (100) crystal face is dominantly exposed on

13

Page 14: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

γ-Al2O3 surface, Peri proposed a model predicting the existence of five types of Al-OH

groups. The model attributes the different OH frequencies to differing numbers of

surrounding surface oxide sites. In other terms, the difference of Al-OH frequency depended

upon the number of the nearest neighbour oxygen anions surrounded a particular Al-OH

group. Peri surmised that the high wavenumber band at 3800 cm-1 was an OH group with four

neighbouring oxygen anions. Whereas the low wavenumber band at 3700 cm-1 was an OH

group with no oxygen neighbours, but four aluminium ions. The other three OH groups giving

rise to the intermediate stretching frequencies at 3730, 3745, 3780 cm-1 were supposed to

have one, two and three nearest neighbouring oxygen anions, respectively. As the temperature

increased, the lower frequency associated (H-bonded) hydroxyl groups, being closest

together, were removed first, leaving isolated Al-OH groups to be removed at higher

temperatures after heating to 700°C. Upon heating to 900°C no hydroxyl groups were

detected by IR.

Knözinger and Ratnasamy, assuming that all of crystallite faces of Al2O3 (100, 110 and 111)

have an equal chance of projection on γ-Al2O3 surface, proposed a model that assigned the

different OH frequencies of the five types of Al-OH groups to tetrahedrally and octahedrally

coordinated Al3+sites in terminal and bridged configuration [20, 21]. These assignments are

shown in Fig. 3.

Fig. 3. Assignments of OH groups on γ-Al2O3 surface according to Knözinger and Ratnasamy

14

Page 15: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Despite the studies mentioned above, none has attempted to correlate directly the extent of

dehydroxylation with the production of Lewis acid sites as a function of the treatment

temperature. Nevertheless, in the spirit of the work of Zaki and Knözinger who undertook a

low temperature IR study of the interaction between Al-OH groups and adsorbed CO, five

bands of different types of hydroxyl groups were assigned on Al2O3 dehydroxylated at 500°C,

two types of associated Al-OH were removed first, leaving three types of isolated Al-OH

groups to be removed at higher temperature. They observed two low temperature CO

adsorption sites at -193°C on Al2O3 and they attributed the first adsorption site (2140-2150

cm-1) to CO hydrogen-bonded to Al-OH groups and a physisorbed CO layer, while the second

CO adsorption site (2195-2213 and 2238 cm-1) was assigned to tetrahedrally and octahedrally

coordinated Al+3 sites [22, 23].

Later, John and Ballinger undertook a low temperature (-196°C) IR study of the interaction

between Al-OH groups and adsorbed CO, in order to examine the acid-base properties of the

hydroxyl groups on Al2O3 after dehydroxylation at elevated temperatures [24]. They found an

approximately direct correlation between the elimination of surface hydroxyl groups and the

increase in integrated absorbance of CO on Lewis acid sites of Al2O3 produced during the

dehydroxylation. Furthermore, they observed two Al3+ adsorption sites: the first develops

following mild dehydroxylation, and the second appears only after dehydroxylation at 600°C

and higher due to the presence of a mixture of Lewis acid sites on highly dehydroxylated

Al2O3 surface. From these studies, it has become apparent that the acidity of Al2O3 develops

when it is dehydroxylated, not because of the OH groups themselves. Consequently, the

removal of water and/or hydroxyl groups (surface ligands), coordinatively unsaturated (CUS),

anions (oxygen ions) and cations (exposed Al3+, anion vacancies) are created.

Numerous IR studies reported that both tetrahedrally and octahedrally coordinated Mo6+ are

present in impregnated Mo/Al2O3 catalysts, and the octahedral/tetrahedral ratio increased with

increasing Mo loading. Infrared spectra of molybdena-alumina have also shown that alumina

surface hydroxyls are eliminated by the addition of molybdena [25-29]. These results

provided strong evidence of bonding between molybdena and the alumina surface. The

formation of tetrahedrally coordinated Mo6+ on the surface of alumina was pictured as the

replacement of the terminal hydroxyl groups as shown in the equation below [12, 30]:

15

Page 16: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

(Eq-5)

Rédey and Millman independently measured the loss of OH groups on alumina in the

formation of molybdena-alumina catalyst by determination of the total hydrogen contents by

exchange with D2 using the isotope dilution method [3, 14]. The number of hydroxyls

eliminated in the formation of 8% Mo/Al2O3 catalyst (∆OH/Mo) was determined to be about

1.7 0.6, which is in agreement with the epitaxial monolayer model (Eq-5). ±

Interestingly, the hydroxyls, which are replaced by molybdena when the catalyst is formed

(calcined), reappear when the catalyst is partly reduced in H2.

The fragments arising from the dissociative adsorption of water on the surface of metal oxides

give rise to hydroxyl groups that are potentially more or less active Brönsted acid sites. Such

surface hydroxyl groups can be detected, directly, recording the IR spectra of the oxide

catalyst powders in the region 3800-3000 cm-1, where the O-H stretching modes (νOH's)

occur [12-24]. Although the position and shape of the ν(OH) bands of such surface hydroxyl

groups is informative on their coordination, these data do not give straightforward information

on their Brönsted acidity. In fact, as for example, the position of the OH band over a basic

catalyst like MgO, of a weakly acidic catalyst as amorphous silica and of a strong Brönsted

acidic catalyst like silica-alumina is almost the same (3745 ± 3 cm-1). Moreover, very acidic

catalysts like, for example, sulphated zirconia and titania do not present any definite sharp

ν(OH) band, while others, like zeolite ZSM-5 and silica-alumina, show sharp ν(OH) bands.

These facts are due to the following main reasons:

1. The ν(OH) frequency depends not only on the O-H bond strength, but also on the nature of

the M-O(H) bond, i.e. the element(s) to which the OH is bonded.

2. In any case, even for OH's bonded to the same element, the function ν(OH) versus acidity is

not necessarily linear but can present a maximum.

3. The state of the OH groups on the surface also depends on the basic strength of the oxide

ions. In fact, in very covalent structures, like for silica-alumina and zeolites, where oxygens

are almost not basic, the acidic OH's are responsible for rather sharp and well-defined bands,

while when the nearest oxygens are more or less basic, the acidic OH's give rise to

H-bonding, with a shift down and a broadening of the ν(OH) band [25-31].

16

Page 17: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

1.2. Current status of catalysts preparation Different methods have been described in the literature for the preparation of these catalysts.

Among these methods, the impregnation (IW) and the equilibrium adsorption (EA) methods

have been widely used [1, 2, 3].

1.2.1. The impregnation method On most of the previous studies, the molybdena-alumina catalysts were prepared by the

impregnation method, which is the most widely used preparation technique for supported

heterogeneous catalysts. This procedure involves making a solution of the active component

having a volume equal to the pore volume of the support material [3-15].

The solution is then impregnated into the support material and mixed evenly where it is

completely taken up. The solid is first dried at about 150°C and then calcined at elevated

temperatures, in general, between 400 to 600°C (Fig. 4). Consequently, the resulting

preparations have varied somewhat depending upon the initial molybdenum compounds used,

the initial pH value of the molybdenum solution, the pore structure of the support and the

preparation skill. One obvious disadvantage of this preparation method has been that

molybdena cannot be uniformly distributed over the support surface.

Knözinger et al. studied molybdena-alumina catalysts prepared by impregnation at pH = 6 and

11 and concluded from Raman results that the catalyst prepared at pH = 6 contained a large

amount of bulk MoO3 while that prepared at pH = 11 did not [32]. Moreover, the same results

showed that this technique does not lead to uniform coverage. Accordingly, it can be argued

that this technique is not well defined. For instance, the pH value of the solution is not usually

specified when using this technique in addition to the pre-treatment conditions including

calcination temperature, time and ambient atmosphere. For example, it has been reported by

Hercules and co-workers that increased calcination temperature and time favour the formation

of Al2(MoO4)3, which is believed to be a sub-surface species, while bulk MoO3 was observed

at low calcination temperature and short time [33-35].

In summary, the final catalyst made by the impregnation method may be affected by the

solution pH value, the mesh size of the support used, and other variables such as calcination

temperature and time. Therefore, it is difficult to reproduce the catalyst by this method from

one batch to another.

17

Page 18: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 4. Production of supported molybdena catalysts by impregnation method

1.2.2. Adsorption methods These methods for preparation of molybdena-alumina catalysts were first used by Sonnemans

and Mars in both liquid and gas phases [36]. In the preparation from liquid phase a fresh

solution of 1% ammonium paramolybdate (pH 1-9) was flowed through an alumina bed for a

period of 4 hours (until the concentration of the effluent solution was found the same as that

of the entering reagent). The final Mo content was found to depend on the solution pH value.

It reached 21% in MoO3 when prepared at pH = 1, but decreased as the pH was increased. At

pH = 9, the catalyst contained only 2% Mo. They also reported that the concentration of the

solution had an effect on the amount adsorbed (the concentration range studied was in the

range between 0.2 and 1%). They concluded that Mo/Al2O3 with monolayer coverage could

be achieved by either gas phase or liquid phase adsorption method. The authors suggested that

the pH effect on the molybdenum content resulted from a change in the mean size of the

polymolybdate ions, which is a function of the pH.

In contrast, Iannibello and co-authors who showed the same pH dependency interpreted the

pH effect as due to a higher fraction of protonated alumina hydroxyls at lower pH leading to a

higher adsorption of anions [37]. They observed that the pH decreased rapidly at first and then

increased slowly to a steady value. The authors attributed the fast initial pH decrease to

18

Page 19: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

exchange of ammonium ions with protons of alumina and the slow subsequent increase to

exchange of molybdate ions with surface hydroxyl groups as represented by Eq-6 as [MoO4]2-

was adsorbed, the equilibrium (Eq-7) would shift toward the right:

Al-OH + [MoO4]2- ⇔ Al-O-MoO3- + OH- (Eq-6)

[Mo7O24]6- + 4H2O ⇔ 7[MoO4]2- + 8H+ (Eq-7)

Actually, at pH values above 7 or 8, Mo4+ occurred as the tetrahedral monomeric molybdate

anions, but polymerization occurred at concentration in excess of 10-4 M at somewhat lower

pH values. The two major equilibriums [3, 12, 37], which may occur in the solution, may be

written as follows (Eqs. 7 and 8):

8MoO42- + 12H+ ⇔ Mo8O26

4- + 6H2O (Eq-8)

The major molybdate species at concentration above 10-3 M that are present at various pH

ranges can be roughly illustrated by the scheme below [11, 12]:

Scheme 3.

However, the well-understood principle of colloid chemistry can provide an explanation for

the variations in Mo loading with pH and the reversibility of the adsorption process. The

surface of a solid oxide in an aqueous solution is generally electrically charged. This charge

may be attributed to: (a) dissociation of surface hydroxyl groups or (b) adsorption of protons

formed by hydrolysis of H2O. These two mechanisms can qualitatively explain the pH

dependence of surface charge and the existence of a pH resulting in zero net charge, called the

isoelectric point of the solid (IEPS) or zero point of charge (ZPC) [38-40].

19

Page 20: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Based on the concept of IEPS, the surface chemistry of an oxide with an aqueous solution can

conveniently be expressed as written by Parfitt [41]:

M-OH2+ M-OH ⇔ ⇔ M-O– + H+ (Eq-9)

Decreasing pH IEPS Increasing pH ⇐ ⇒

The equilibrium indicates that at IEPS of the oxide, the net surface charge is zero, although

this does not necessarily mean that all the surface hydroxyls are in the form of M-OH because

they vary in acidity depending on their coordination. The net zero surface charge simply

means the concentrations of M-OH2+ and M-O– are equal at IEPS. When the pH is adjusted

lower than the IEPS, the concentration of M-OH2+ will be higher than M-O–, consequently the

surface will carry a net positive charge:

++ −⇔+− surfsurf MOHaqHMOH 2)( (Eq-10)

The situation is reversed when the solution pH is adjusted higher than the IEPS, and the oxide

surface will carry a net negative charge:

)(

2

aqHMOMOHor

OHMOOHMOH

surfsurf

surfsurf

+−

−−

+−⇔−

+−⇔+−

(Eq-11)

The same factors play an important role in the adsorption of metal oxyanions on the oxide

support. Consequently, the fact that the charge on the surface can be adjusted by varying the

solution pH. Thus, the amount adsorbed can be varied by changing the zeta potential of the

surface. Accordingly, the IEPS has a great influence on the band position of the adsorbed

species. The IEPS of several oxide supports is shown in Table 1.

20

Page 21: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Table 1. Comparison between IEPS of several oxide supports

Oxide supports Isoelectric point (IEPS)

SiO2 1-2

γ-Al2O3 6-8

TiO2 4.7-5 (Rutile), 5.7-6.2 (Anatase)

MgO 12.1-12.7

ZrO2 6.6-7.1

CeO2 (cerianite) 6.7-6.8

SnO2 (cassiterite) 4.2

J. Sarrin and co-workers compared the reducibility and activity of two series of molybdena-

alumina catalysts prepared by impregnation (IW) and adsorption methods (EA) [42]. Their

results showed that for both series of catalysts, the reducibility of the molybdenum species

increases as the Mo loading increases, in agreement with the literature [12-15]. For the

catalysts with similar loading, the (IW) series showed a higher surface coverage of the Mo

phase and higher degree of reduction than for the corresponding (EA) preparations. The

reducibility data were consistent with the catalytic results and oxygen chemisorption results.

The (IW) preparations (with similar Mo loading) were more active in the isomerization of

1-butene and chemisorbed larger amount of oxygen than their (EA) counterparts. The

differences in reducibility can be ascribed to a nonuniform repartition of the molybdenum

species between the external and internal surfaces of alumina for (IW) preparations, which

may contain a greater fraction of easily reducible polymeric Mo species than their more

uniform (EA) counterparts. Another possible explanation may stem from the decoration effect

of the Mo species by Al3+ ions. The latter may arise from the dissolution of alumina, which is

favoured on the (EA) series due to the long contact time between the Mo solution and

alumina.

Other studies indicate that the preparation method does not influence the molecular structure

of the Mo species present on Al2O3. Thus, for a given Mo loading, the nature of the Mo

species is independent of the preparation method [43-47]. However, it is not clear, how a

given preparation method may induce the Mo speciation.

21

Page 22: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

1.3. Molybdena with CeO2 and SnO2 semiconductors Promoters are the subject of great interest in catalyst research due to their remarkable

influence on the activity, selectivity and stability of industrial catalysts. It is sometimes

difficult to define precisely the function of the promoters that has not been elucidated [1, 2].

Ni and Co promoters in Mo/Al2O3 catalysts are well-known for their success in the

hydrodesulphurization (HDS) of petroleum feedstock and coal liquefaction products [48]. The

promoting role of both promoters was found to increase the Mo dispersion and reduction, in

addition to the increase in H2 mobility, an intercalation effect with MoS2, a decrease in

deactivation and an increase in surface segregation of mixed sulphide phases. For instance,

when Co was introduced into Mo/Al2O3, various effects occurred. Free MoO3, as well as

Al2(MoO4)3, was converted into CoMoO4, and the Co addition resulted in a decrease of the

isolated Mo tetrahedral concentration and favoured the formation of the polymeric form.

Moreover, Topsøe et al. reported direct evidence of Co and Mo existing in the active form as

a Co-Mo-S surface phase [49, 50].

The purpose of doping a semi conductive carrier in order to enhance the catalytic activity of

supported metal catalysts has recently been applied in developing “three-way” catalysts [51].

The conductivity of semiconductors is generally low but can be considerably changed by

either incorporating with other oxides or upon pre-treatments. Their crystal lattices tend to

release or take up oxygen. Therefore, alloying the metals with semiconductors can increase or

decrease the activity. This effect has some industrial relevance since can both accelerate

desired reactions and suppress undesired reactions [52, 53]. For instance, the addition of Sn to

Pd gives selective catalysts for the removal of acetylene from ethylene streams [54].

In the case of n-type semiconductors (e.g. ZnO, TiO2, CeO2, SnO2), a pre-treatment in a

reducing atmosphere generates electron-donor levels (oxygen vacancies VO, metal under a

lower oxidation state), which increases the free electron concentration [55-57].

The presence of electron-donor levels gives rise to electronic transitions, which may occur in

the infrared region:

M(n−1)+ → Mn+ + e− (Eq-12)

VO → VO+ + e− (Eq-13)

VO+ → VO

2+ + e− (Eq-14) For instance, on heating or by reaction with reducing gases such as H2, CO for n-type

semiconductors such as CeO2, the release of oxygen can be described as below:

22

Page 23: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

CeO2 ↔ CeO2-x + ½ xO2(g) (Eq-15)

Oox ↔ Vo + ½ O2(g) (Eq-16)

Ce4+ + Vo ↔ Ce3+ + Vo+ (Eq-17)

Where VO, VO

+ and VO2+ are the neutral and ionized oxygen vacancies, respectively, OO

x is

the lattice oxygen [2].

On the other hand, the adsorption of oxygen on a nonstoichiometric oxide, containing oxygen

vacancies VO, generates lattice oxygen OOx. At the same time, metal ions are oxidized at the

surface and the conductivity is lowered for n-type semiconductors because the oxygen acts as

an electron acceptor:

½ O2(g) + Vo ↔ Oox (Eq-18)

Ce3+ + O2 ↔ Ce4+ + O2– (Eq-19)

The adsorption of oxygen results eventually in complete coverage of the surface by O– or O2-

and the heat of the adsorption remains practically constant while the surface becomes

saturated with oxygen and negatively polarized.

Considering the adsorption of hydrogen on n-type semiconductors, it has been shown that H2

mainly undergoes heterolytic dissociation [2, 58-60]:

M2+ + O2- + H2 → M+ H + OH– (Eq-20)

On heating, the hydroxyl ions are decomposed to water and anionic defects, and a

corresponding number of metal cations are reduced to atoms. In this strong chemisorption, a

free electron or positive hole from the lattice is involved in the chemisorptive bonding. This

changes the electrical charge of the adsorption center, which can then transfer its charge to the

adsorbed molecule. Thus, chemisorbed hydrogen acts as an electron donor and increases the

conductivity of n-type semiconductors. Furthermore, the change in the electrical charge

density on the surface can hinder the further adsorption of the same gas. A decrease in the

heat of adsorption with increasing degree of coverage is then observed, and hence a deviation

from Langmuir adsorption isotherm occurs [2, 61-65].

However, when a metal is applied to an appropriate n-type semiconductor, its electron density

increases [1, 2]. The general behaviour of some nonstoichiometric semiconductor oxides is

summarized in Table 2.

23

Page 24: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Table 2. Behaviour of nonstoichiometric semiconductor oxides

n-type p-type

For instance TiO2, SnO2, CeO2 NiO, CoO, FeO

Type of conductivity Electrons Positive holes

Addition of M21+ O oxides Lowers conductivity Increases conductivity

Addition of M23+O3 oxides Increases conductivity Lowers conductivity

Adsorption of O2, N2O Lowers conductivity Increases conductivity

Adsorption of H2, CO Increases conductivity Lowers conductivity The optical absorption of semiconducting oxides arises from five different phenomena: (i)

intrinsic absorption, corresponding to transitions between (full) valence bands and (empty)

conduction bands, which occur often in the UV–visible range and sometimes in the near-

infrared (NIR) (for narrow gap semiconductors); (ii) transitions between valence bands, called

intervalence transitions, only observed in p-type materials, which may appear in the NIR; (iii)

free carrier absorption, arising from transitions within one band; (iv) transitions of an

electron to or from a localized state; (v) lattice vibrational absorption. Their mid-infrared

examination offers special difficulties due to mainly the transition types (iii) and (iv), which

involve the absorbance due to free carriers and electron- or hole-donors, whose concentration

depends on the semiconduction type, the surrounding atmosphere and the temperature [62].

These difficulties are seriously enhanced when the sample under study is a metal supported on

an n-type semiconducting support, the reduction of the support is then greatly favoured by the

metal, e.g. through activation in vacuum and spill over of hydrogen or CO. For example, this

has been observed in the case of metals supported on ZnO and ceria [63-65].

Lanthanide ions of variable valence particularly Ce3+/4+ usually lead to nonstoichiometric

CeO2-x. It has been reported earlier that the latter aspect and the defect structure on ceria was

due to oxygen vacancies accompanied by triply and/or quadruply Ce interstitial to maintain

the electrical neutrality. However, lately, oxygen vacancies have been finally affirmed as the

prevailing defects neglecting the negligible effect of Ce interstitials in such a fluorite-

structured oxide system [66-72]. The availability of these defect sites on the surface is

probably related to their high bulk concentration. The oxygen anions (O2-) on ceria surface

may be one, two or three coordinated to cerium cations. From what has been conferred for

CeO2 in addition to its role as either oxygen storage and release or thermal stabilizer. It has

been used either as a promoter or as a support for metal catalysts in many applications since

24

Page 25: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

ceria has a beneficial effect for CO oxidation and NOx reduction under both stoichiometry

and excess oxygen beside for CO/NO reaction, CO and CO2 hydrogenation. Since the

catalytic oxidation of CO has acquired tremendous attraction lately particularly in connection

with world-wide endeavors to curb the detrimental impacts of automotive emissions on the

atmosphere. Although the detailed mechanism of the reactions mentioned above is still

unknown, the researchers clearly assigned the promotional effect of the catalysts to the role of

ceria in creating Ce3+/4+ redox couple [73-75]. However, due to the limited supply of precious

metals and some impractical properties, an attention has been given to transition metals and

their oxides as catalysts supported on CeO2 or doped with CeO2, since the ability of ceria to

donate oxygen to supported metals is also a key feature in other catalytic reactions like for

example, catalytic combustion and water gas shift reaction [76-78].

M. Mokhtar investigated the influence of ceria on Mo/TiO2 and found that the presence of

ceria leads to increase the concentration of polymerized surface Mo oxide species, and rather

initiated the formation of MoO3 over-layers. Additionally, the involvement of ceria, on the

other hand, retarded the strong association rendered between Mo and Ti and thus stimulated

the formation of discrete amounts of the corresponding oxides. More specifically, ceria was

found to work as a mediator between Mo and Ti [79, 80].

Lucia and co-workers found that the presence of cerium in the Mo-Sn system increases the

rate of ethanol dehydrogenation as well as the selectivity to acetic acid and acetaldehyde. In

addition, it caused changes in the distribution of Mo species and in the textural properties, but

mostly increasing the basicity of the catalyst [81].

Stannic oxide thin films are attractive for many applications due to their unique physical

properties such as high electrical conductivity, high transparency in the visible part of

spectrum, and high reflectivity in the IR region. In particular, tin oxide films are stable at high

temperatures, have excellent resistance to strong acids and bases at room temperature, are

resistant to mechanical wear, and have very good adhesion to many substrates [82-85]. Thus,

transparent and electrically conductive stannic oxide films are widely used for a variety of

applications. Briefly, these applications include: as electrodes in electroluminescent displays,

imaging devices, protective coatings, antireflection coatings, gas and chemical sensors,

transducers applications based on transparent conductors and other optoelectronic devices.

Furthermore, tin oxide films are more stable than the other transparent conducting oxide

(TCO) films such as zinc oxide (ZnO). Moreover, they have a lower material cost. Recently,

the synthesis of ultra fine tin oxide particles is of great technological and scientific interest

25

Page 26: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

owing to their superior physical and chemical properties and their use as either catalysts for

the oxidation of organic compounds or gas sensors [83-86].

As the electrical conductivity of SnO2 derived from the variable valence on the Sn atomic

center is very sensitive to oxidative and reducing atmospheres, tin oxides as gas sensors

detecting a trace amount of the gases have been applied to processes in chemical, heretical

and fermentation industries to control the amount of the harmful wastes discharged from the

plants, the explosion of the combustible gases and incomplete combustion, exhaust gases

from automobiles [87-90]. However, molybdenum–tin thin films seems to be promising gas

sensors. It has recently been stated that addition of MoO3 to SnO2 increases the sensor

response to CO and NO2 [91-94].

All the above properties have led to intense research of SnO2 coatings over the past few

decades. Currently, numerous techniques exist for the preparation of tin oxide films such as

chemical vapor deposition, spray pyrolysis, sputtering, and sol–gel deposition [95-98].

Nevertheless, the ability of SnO2 to generate defects has been only recently shown to induce

interesting performances for supported Pd catalysts, e.g. in deNOx reactions [99]. Conversely,

tin dioxide has received limited attention in the catalysis field and the use of Mo-Sn oxides in

selective oxidation appears to be unique industrial application [100-109].

On the other hand, Mo/SnO2 catalysts have been used for selective oxidation reactions due to

their high activity. Niwa et al. [100] studying the methanol oxidation with several supported

molybdenum catalysts found the following sequence of activity:

Mo/SnO2 > Mo/Fe2O3 > Mo/ZrO2 > Mo/TiO2 > Mo/Al2O3

Goncalves et al. [101] and Medeiros et al. [102] have shown that acetic acid can be obtained

from ethanol oxidation in only one-step with high yield when Mo/SnO2 catalysts prepared by

precipitation procedure are used.

Recently, Liu et al. [103] have shown that Mo/SnO2 catalysts are very active for the oxidation

of dimethyl ether although they are more selective for formaldehyde than Mo/Al2O3 catalyst.

V. Lochar claimed that the activity of MoO3/SnO2 catalyst for methanol oxidation could be

associated with its Brönsted and Lewis acidity as the result of the catalyst reduction [104].

Other catalytic activity results suggest the existence of synergy between the apparently pure

phases of MoO3 and SnO2. Therefore, MoO3 in close contact with SnO2 has shown to be

much more active and selective than the individual pure phases. The high dispersion of

26

Page 27: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

molybdenum species on the highly reducible SnO2 support was suggested to be responsible

for the exceptional activity of these catalysts [105-109].

However, the information available in the literature on the interpretation of infrared and

Raman spectra of Mo/Sn and Mo/Ce compounds prepared on different surfaces is rather

limited. In assigning vibrational spectra, some DFT (Discrete Fourier Transform) calculations

or some vibrational spectroscopic data relating to Mo/Sn and Mo/Ce systems can be relied on

[65, 80, 88, 91, 97, 104-109]. On the other hand, few papers can be found in the literature on

IR and TG studies related to either Mo/Sn or Mo/Ce system in contrast to publications

relating to noble metals with ceria and tin.

Anyhow, a lot of debates concerning the role played by ceria and tin oxides necessitate further

studies in order to explore the influence of CeO2 and SnO2 on the structure and surface

characteristics of molybdena for better understanding the nature, structure and the physico-

chemical properties of these oxides, since the nature of the interactions between metal oxides

and supports are often attributed to the complexity of these systems and differences in the

preparation and experimental conditions adopted.

1.4. Direct conversion of methane under nonoxidative conditions An important task confronting catalytic chemists is how to realize direct conversion of

methane to versatile fuels and valuable chemicals by building up the desired C–C (or C–O)

bond. Thermodynamic constraints on the reactions in which all four C–H bonds of CH4 are

totally destroyed, such as CH4 reforming into synthesis gas or CH4 decomposition into carbon

and hydrogen, are much easier to overcome than the reactions in which only one or two of the

C–H bonds are broken under either oxidative or nonoxidative conditions [110-117]. Direct

conversion of CH4 with the assistance of oxidants is thermodynamically more favourable than

that under nonoxidative conditions. Therefore, the direct conversion of CH4 under the aid of

oxidants has received much more attention than that under nonoxidative conditions, especially

when considering the production of fuels and valuable chemicals from CH4 [118-123].

With the urge to quest for renewable energy and cleaner fuels, it is recognized that hydrogen

energy will inevitably replace fossil fuel energy in the near future due to the fact that the

burning of hydrogen is pollution free. However, it is a practical way to produce H2 from CH4

due to its high H/C atomic ratio and great abundance in reserves. Therefore, the direct

conversion of CH4 under nonoxidative conditions into H2 and/or H2 accompanied with basic

chemicals is closely related to the effective utilization of CH4-containing resources and thus to

sustainable progress and development of the living conditions of humankind [124-127].

27

Page 28: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

The direct conversion of CH4 under nonoxidative conditions is thermodynamically

unfavorable. Nevertheless, as an alternative approach, it has still attracted the attention of

many researchers. In heterogeneous catalysis, various metals have been discovered that can

chemisorb CH4 at moderate temperatures and that can decompose CH4 to C and H2 at higher

temperatures [128-137].

Amariglio and co-workers reported a “two-step” process on Pt, Ru, and Co in isothermal

experiments [128-130]. In a series of publications, the authors suggested that C–C bonding

could take place between H-deficient and CHx formed during the first step of methane

chemisorption, while H2 saturated the alkane precursors in the second step and removed them

from the surface. In view of the fact that hydrogenation at a temperature lower than that of

CH4 chemisorption is favorable for lessening hydrogenolysis. The authors reported a

nonoxidative conversion of methane to higher hydrocarbons through a dual temperature two-

step reaction on Pt/SiO2 and Ru/SiO2 catalysts. Indeed, when chemisorption of methane was

set at a fixed temperature (usually lower than 320°C), the selectivity to heavier alkanes

increased with the lowering of hydrogenation temperature on both catalysts. On the other

hand, when the hydrogenation temperature was less than 120°C, hydrogenolysis was

negligible, and thus the variations of the products can only be attributed to the changes

affected by the adlayer formed during the chemisorption of methane at a certain temperature.

It was discovered that the products of C2+ hydrocarbons at every hydrogenation temperature

displayed a maximum versus the methane chemisorption temperature on both catalysts. In the

case of the Pt/SiO2 catalyst, mainly C2H6 and n-C5H12 were produced during the first minute

of the reaction. This illustrates that C–C bonds could form during CH4 adsorption, and the

authors assumed a surface intermediate of C5 precursor bonded on dispersed and coordinately

unsaturated Pt atoms.

Van Santon et al. suggested that CH4 first dissociated on a precious metal to form carbide and

H2. Then, the carbide was hydrogenated by H2 to produce higher hydrocarbons. C–C bonds

were supposed to be created during the hydrogenation step. Since the reactivity of the CHx

surface intermediates formed from CO and CH4 was quite similar. The authors suggested that

the chain-growth probability would depend on the metal–carbon bond strength and that the

mechanism of C–C bond formation in the two-step route should be related to that occurring in

the Fisher–Tropsch reaction. They also demonstrated that the homologation of olefins (C2H4,

C3H6, etc.) with methane could occur over Ru/SiO2 and Co/SiO2 catalysts [131, 132].

28

Page 29: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

The two-step route is also feasible over a number of oxide- or zeolite-supported transition

metal and bimetal catalysts. Solymosi and Cserenyi illustrated that over a Cu-promoted

Rh/SiO2 catalyst, the enhanced formation of C2H6 and higher hydrocarbons could be observed

in the two-step process [133, 134].

Guczi et al. reported that the chemisorption of CH4 at 250°C and the subsequent

hydrogenation of the CHx species at 250°C over Co–Pt/NaY and Co–Pt/Al2O3 performed the

best of all the catalysts tested. The chemisorbed CHx species had the highest concentration,

and all CHx species were hydrogenated in the second step, giving a selectivity of C2+ close to

84%. They found that there was a correlation between the hydrogen content of the surface

CHx species (the optimum value of x being around 2) and the chain length of the

hydrocarbons produced in the hydrogenation step in their mechanistic study of the two-step

process [135]. Later, they reported that the two-step process could be simplified into a one-

step process with a C2+ hydrocarbons production higher than that obtained in the two-step

process over Co–Pt/NaY bimetallic catalyst. These results could be obtained if the CH4 was

pulsed with H2/He mixture at 250°C [136].

Bradford reported the results of the isothermal, nonoxidative, two-step conversion of CH4 to

C2+ hydrocarbons over supported and unsupported Pt and Ru catalysts at moderate

temperatures and elevated pressures. It was shown that an increase in reaction pressure

increased the branching and molecular weight distribution of the product [137].

Several researchers suggested the preparation of a multifunctional catalyst to avoid the use of

a two-step process. Furthermore, it has been reported that dehydrogenative coupling of CH4

without any oxidant could be carried out over Pt–SO4/ZrO2 catalysts. A steady conversion of

0.2% (the equilibrium conversion of CH4 into C2H6 and H2 is estimated to be 0.6%) was

observed after the catalyst was reduced in H2 at 500°C [138, 139].

On the other hand, in order to overcome the thermodynamic limit and to enhance the

reactivity for obtaining high yields in direct conversion of CH4 under nonoxidative conditions,

plasma excitation has also been attempted. The product distribution is dependent on the

method by which plasma excitation is produced. For example, in pulsed corona discharges at

atmospheric pressure, C2 hydrocarbons (mainly C2H2) were obtained with a high selectivity of

around 70 to 90%. In microwave plasmas, the product distribution shifted from C2H6 to C2H4

and finally to C2H2 with an increase in power density. By introducing a proper catalyst into

the microwave plasma reactor, CH4 could be converted to higher hydrocarbons at atmospheric

pressure. In addition, with a CH4 and H2 mixture as the feed gas, the selectivity to C2H2 was

88% and that to C2H4 was 6% at a CH4 conversion of 76% [140, 141]. Here, again, the main

29

Page 30: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

drawback is the low energy efficiency to drive this thermodynamically unfavorable reaction.

Thermodynamically, the transformation of CH4 under nonoxidative conditions is more

favorable to aromatics than to olefins. The direct conversion of CH4 to aromatics was tested

on several catalysts in either a pulse or a flow reactor. Wang et al. reported on the

dehydroaromatization of methane (MDA) for the formation of aromatics (mainly C6H6) and

H2 under a nonoxidizing condition in a continuous flow reactor on Mo/HZSM-5 catalysts

[142]. More detailed studies on the reaction revealed that the channel structure and acidity of

the HZSM-5 zeolite, as well as the valence and location of the Mo species, are crucial factors

for the catalytic performance of the Mo/HZSM-5 catalysts. In addition, W/HZSM-5 and

Re/HZSM-5 are also reported to be active elements for MDA [142-144].

Solymosi and co-workers [147-152] and Lunsford and co-workers [153-155] characterized

the Mo/HZSM-5 catalyst by means of XPS and found that during the initial induction period,

the original Mo6+ ions in the zeolite were reduced by CH4 to Mo2C, accompanied by the

depositing of carbonaceous cokes. They suggested that Mo2C provides active sites for C2H4

formation from CH4, while the acidic sites catalyze the subsequent conversion to C6H6. The

Mo2C species probably are highly dispersed on the outer surface, and some of them reside in

the channels of the zeolite. Meanwhile, the spectra of Mo/HZSM-5 samples reacted with CH4

at 700°C for one and 24 hrs were basically identical to the Mo2C reference spectrum, except

for a partial contribution from the Mo oxide, given rise by MoOxCy. These authors claimed

that the Mo oxide species dispersed in the HZSM-5 framework might migrate onto the

external surface of the HZSM-5, be converted by CH4 to Mo2C, and disperse on the support

surface. Therefore, the carbonaceous deposits created in MDA are in various forms and play

different roles. First, Mo2C and/or MoOxCy, which are possibly active species for CH4

activation, are formed during the induction period. Second, the formation of the active

intermediates, the CHx species, follows the activation of CH4 on Mo2C and/or MoOxCy. The

last one to be formed is coke leading to the deactivation of the catalyst. It is understandable

that there are some similarities between the carbonaceous species formed in MDA and those

formed in the first step of the two-step process, since both reactions are carried out under

nonoxidative conditions, and Mo2C shows some precious metal-like properties.

In spite of the fact that the reaction is thermodynamically unfavorable under pressurized

conditions and that 10% CO2 added to the feed totally suppresses the activity of the 2 wt%

Mo/HZSM-5 catalyst. Ichikawa and co-workers found that an increase in CH4 pressure and

the addition of small amounts of CO and CO2 (less than 3%) to the CH4 feed enhanced the

catalyst stability in the reaction [156, 157]. By increasing the CH4 pressure, the formation

30

Page 31: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

rates of C6H6 and hydrocarbons could be moderately increased. This kind of pressure

relationship may be related to a sufficient supply of H2 from CH4 and a suitable concentration

of surface carbon species CHx for the formation of aromatic products. By using a CO and CH4

mixture as the feed to conduct the reaction, the authors suggested that CO dissociated on the

Mo sites to form the active carbon species CHx. The dissociated oxygen species [O] from CO

might react with the surface inert carbon species to regenerate CO, resulting in the

suppression of coke formation on the catalyst. These results imply that although the Brönsted

acid sites are necessary, excess Brönsted acid sites are detrimental for the reaction, since

severe coke formation will occur on them [157-159].

Considerable efforts have been devoted to developing active and selective catalysts and

understanding the bifunctionality of Mo/HZSM-5 catalysts and the nature of carbonaceous

deposits formed during the reaction. However, neither new active and selective catalysts nor a

thorough understanding of the mechanism of the reaction has been achieved.

However, despite all substantial research efforts into nonoxidative two-step or one-step CH4

homologation, its low efficiency is the main problem to further developing it as a commercial

process. In any case, these studies enhanced our knowledge in direct conversion of CH4 under

nonoxidative conditions, particularly methane dehydroaromatization, and stimulated chemists

to explore new methane conversion processes.

1.5. Characterization of the catalysts Both of the physical and the chemical properties of a catalyst must be known if relationships

between the structure and activity, selectivity, and lifetime are to be revealed. There are many

techniques commercially available for the analysis of catalysts opening up new possibilities

for fundamental catalyst research. In this section, I will encounter some methods for

characterizing catalysts and not discuss their capabilities and limitations that have been

described in the literature [1, 2].

31

Page 32: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

1.5.1. Physical characterization The significance of the physical surface is based on the fact that the bonding forces of the

atoms in the outer most atomic layers are not saturated. Thus, these atoms experience

structural reorganization (relaxation, reconstruction) and a high chemical reactivity. On the

other hand, the thermodynamic properties of these layers are different from the bulk.

The distribution of pores across the inner and outer surfaces is an important property of the

catalyst. The texture generally refers to the pore structure of the particles including the pore

size distribution, pore volume and pore shape which are determined by gas adsorption

(usually N2, BET method) at relatively low pressures (low values of p/po = pressure/saturation

pressure) for microporous materials [1, 2].

Although the specific surface area is one of the most important parameters of catalysts and

can be determined by the multipoint BET method. However, there is no direct relationship

between catalyst activity and the surface area. Such predictions can only be made by

chemisorption of appropriate gases such as H2, O2, CO, NO at room temperature or above,

respectively [2-4].

However, additional physical techniques can be very useful such as X-ray diffraction (XRD)

that is very essential to determine the crystalline structure and crystallite sizes of the catalysts

despite its limitation to detect particles smaller than about 2 nm. Therefore, the missing of any

discernable diffraction lines does not necessarily prove that the phase in question is absent [1].

Additionally, thermal analysis is an indispensable tool to follow completely the thermal

behaviour of the catalysts, the changes in the composition of the catalysts should be followed

as a function of the temperature, as well. The use of thermogravimetric/differential

thermogravimetric analysis (TGA/DTA) combined with different types of spectroscopic

techniques is essential to reveal the thermal stability of the catalysts. The thermal analysis

combined with mass spectrometry (TG-MS) of different mixed oxide systems can provide

better understanding of the removal and formation of different components in the catalysts.

Thus, knowing the thermal history of the catalysts, tailoring the thermal properties of prepared

catalysts can be facilitated [83, 160-165].

On the other hand, electron paramagnetic resonance (EPR) or electron spin resonance (ESR)

is a technique for studying chemical species that have one or more unpaired electrons, such as

organic and inorganic free radicals or inorganic complexes possessing a transition metal ion.

The basic physical concepts of EPR are analogous to those of nuclear magnetic resonance

(NMR), but it is electron spins that are excited instead of spins of atomic nuclei.

32

Page 33: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Because most stable molecules have all their electrons paired, the EPR technique is less

widely used than NMR. However, this limitation to paramagnetic species also means that the

EPR technique is one of great specificity, since ordinary chemical solvents and matrices do

not give rise to EPR spectra [166-168].

The EPR technique depends on the fact that certain atomic systems have a permanent

magnetic moment. The energy levels of the magnetic system are influenced by the

surrounding atoms and by external magnetic fields. Transitions among the levels can be

detected by monitoring the power absorbed from an alternating magnetic field, just as

ordinary atomic transitions are detected by absorption of light. Comparing the observed

transitions with model calculations then lets us deduce some features of the environment

around the moment [169, 170].

There are a number of ways for condensed matter to retain some magnetic moments, the most

important of which involve certain unusual molecules, transition-group atoms, or particular

point defects in solids. Molecular NO and NO2 both have an odd number of electrons and

hence a permanent magnetic moment. Similarly, many large molecules can exist with an odd

number of electrons. Completing this group, the ground state of O2 happens to be a partially

filled shell with corresponding moment. Transition-group atoms have incomplete 3d, 4d, 5d,

4f or 5f shells. Bonding of these atoms often involves higher-energy p or s electrons, leaving

the unpaired d or f electrons relatively undisturbed. When this occurs, the atom or ion retains

nearly the full atomic moment. Finally, certain defects such as vacancies or foreign atoms in a

crystal may gain or lose an electron relative to the chemically bonded host, thereby producing

a localized moment. The Hamiltonian for the ion is then the sum of several terms [171]:

(Eq-21)

The first term H0 is the usual free-atom Hamiltonian, except for two parts, which are written

explicitly. The spin-orbit coupling is the second term, while the (AI S) term describes the

"hyperfine" coupling of the electronic spin to the nuclear spin I. We use Hcf to represent the

electrical interaction of the paramagnetic species with the neighbouring atoms, including

effects due to bonding. In the simplest approximation, the interaction can be thought of as due

to point charges at the surrounding host sites, hence the common name "crystal field". By

relating Hcf to the observed EPR spectrum we hope to learn something about the surroundings

33

Page 34: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

of the ion. The last term is the Zeeman interaction. In principle, we also should include the

nuclear Zeeman, but in such cases, it is too weak to be of concern [172].

When the hyperfine interaction is present, we must solve a slightly more complicated

problem, since the electronic energy will depend on the orientation of the nucleus as well as

the applied field. For instance, for 3d ions, the spin-orbit interaction is weak, but the crystal

field strength can be comparable to the electron-electron interaction contained in H0. We must

usually, therefore, include Hcf from the beginning and then treat the spin-orbit, hyperfine and

Zeeman terms as perturbations [170-173].

This technique is very sensitive, the detection of very small quantities of paramagnetic

substances (10-6 g) being possible, and it can even distinguish between isolated and

clusterized paramagnetic centers (agglomeration of paramagnetic centers in certain portions).

Furthermore, according to existing literature the purity of the sample can be measured by

assigning the corresponding signals to different ionic impurities in the solid sample [174].

1.5.2. Surface characterization and IR spectroscopy Generally, the features of a surface determine the properties of a material and therefore must

be characterized:

• Topology, morphology

• Elemental composition

• Chemical bonding of elements

• Structure (geometric and electronic)

Since high information content, high spatial resolution and high absolute and relative

detection power are required, that can only be met by physical techniques based on the

interaction of photons, electrons, ions and electrical fields with the material investigated [2].

Complete understanding of catalytic reactions mechanisms, including the nature of adsorbed

intermediates is highly desirable. However, as such should reasonably be expected to provide

major assistance in reaching the goals of better catalysts and improved catalytic processes

from a better fundamental understanding of catalyst surface chemistry. This is an area in

which infrared (IR) spectroscopy undoubtedly makes further major contributions. A variety of

IR techniques can be used to obtain information on the surface chemistry of different solids.

Special meaning have investigations carried out under the reaction conditions. This includes

spectral characteristics of reaction components, surface changes due to temperature treatment

and many others. In principle, all forms of IR spectroscopy, including transmission-

34

Page 35: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

absorption, diffuse reflectance (DRIFT), ATR (attenuates total reflection) and photo acoustic

spectroscopy (PAS), are suitable for in situ measurements. The principal information obtained

with all these techniques is equivalent and local availability and experimental necessities such

as the sample particle size and the molecular extinction coefficient of the sample may

dominate personal choices. For most practical experimental reasons the vast majority of

experiments are currently performed in the transmission-absorption and the diffuse

reflectance mode. This is more related to the design of cells to be used as reactor than to the

principal problems of the other techniques. The IR cell in which the catalyst sample is pre-

treated and subsequently studied is extremely important in surface studies [175, 176].

The cell is normally chosen to suit the purposes of a particular study. Some features are

usually of overriding importance in a given application. Various complex schemes have been

designed to seal the reactor cell with IR-transparent windows so that the IR cell can be

operated at elevated temperatures and pressures. The development of high temperature and

pressure IR cells has permitted the observation of adsorbates under reaction conditions. These

cells may serve as a differential reactor for steady-state reaction, temperature-programmed

reaction or desorption, and unsteady-state reaction studies. Therefore, the IR cell suitable for

investigations of catalyzed reactions must fulfil two requirements: (a) it must allow the

recording of IR spectra under in situ reaction conditions, and (b) its volume and construction

must assure good mixing the gases inside, and the feasible space velocities must allow

flexible variations of the conversion exposing the catalyst to the reactants and products that

can be analysed precisely at the exit of the reactor cell [176].

However, the use of the infrared spectroscopy in heterogeneous catalysis can be classified: (1)

the determination of catalyst bulk structures, (2) identification of adsorbed species and surface

active sites, (3) characterization of surface hydroxyl groups (including Brönsted sites) and (4)

examining the catalyst surface structures (Fig. 5).

35

Page 36: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 5. Applications of IR spectroscopy in catalysis and surface science

From the discussions above and studies of others, IR can provide information concerning the

metal-oxygen vibrations in the region below 1000 cm-1 [177]. Therefore, the metal-oxygen

vibrational frequencies can be roughly divided into five characteristic ranges as shown in the

table below:

Table 3. IR characteristics of metal-oxygen bonds

Vibrational mode Wavenumber (cm-1)

Symmetric and asymmetric stretches of M=O 900-1100

Asymmetric stretches of M-O-M 700-900

Symmetric stretches of M-O-M 500-700

Bending vibrations of M=O 310-400

Deformation vibrations of M-O-M ~ 200

36

Page 37: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Diffuse reflectance spectroscopy (DRIFT) The optical phenomenon known as diffuse reflectance is commonly used in the UV–Vis, NIR,

and MIR regions to obtain molecular spectroscopic information [175-177]. When it is applied

in MIR area with a Fourier transform it is known as diffuse reflectance infrared Fourier

transform spectroscopy (DRIFTS). It is usually used to obtain spectra of powders with

minimum sample preparation. The collection and analysis of surface-reflected

electromagnetic radiation as a function of frequency or wavelength obtain a reflectance

spectrum. Two different types of reflection can occur: regular or specular reflection usually

associated with reflection from smooth, polished surfaces like mirrors, and diffuse reflection

associated with reflection from so-called mat or dull surfaces textured-like powders. In diffuse

reflectance spectroscopy, electromagnetic radiation reflected from dull surfaces is collected

and analysed. If a sample to be analysed is not shiny, and whatever reason is not amenable to

conventional transmission spectroscopy, diffuse reflectance spectroscopy is a logical

alternative. The advent of FTIR spectrometers has led to the widespread application of DRIFT

becoming a valuable technique since hardly any sample preparation is necessary. This implies

that the DRIFT technique is potentially of great value for in situ studies of catalytic systems.

One of the interesting advantages of DRIFT is the prevention of typical transmission

problems at high wavenumbers (due to scattering) and at low wavenumbers (due to strong

absorption of catalyst carriers). Another advantage is the high sensitivity of DRIFT.

Furthermore, the catalyst powder need not be compressed to obtain high quality spectra. This

improves the reproducibility and the activation of the catalyst. Nevertheless, there is a limited

number of commercially available, heatable and evacuable DRIFT cells in combination with a

specially designed optical system that is suitable for the in situ activation and pre-treatment of

catalyst samples at high temperatures and pressures [176-178].

1.5.3. FTIR Spectroscopic detection of CO adsorption on the catalysts As already cited, coordinatively unsaturated cations exposed on the surface of ionic oxides

give rise to surface Lewis acid sites. Consequently, basic molecules can interact with these

sites by forming a new coordination bond, so completing or increasing the overall

coordination at the surface cation. The stronger are the polarizing power of the Lewis acidic

cation (charge to ionic radius ratio) and the basic strength of the adsorbate, the stronger is the

Lewis interaction. Upon this interaction, electrons flow from the basic molecules towards the

catalyst surface. These electronic perturbations as well as the molecular symmetry lowering

37

Page 38: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

arising from this contact are the causes of a vibrational perturbation of the adsorbate. In most

cases, the vibrational perturbation only consists in shifts of the vibrational frequencies, the

more pronounced, the stronger is the interaction, i.e. the greater is the Lewis strength of the

surface site. Accordingly, the shift of the position of some very sensitive bands of the

adsorbate upon adsorption can be taken as a measure of the Lewis acid strength of the surface

sites and the decrease in frequency has been associated with increasing acid strength [175].

As the type of probe molecule chosen will influence the obtained characteristics of the probed

solid and, hence, will also affect the structure–activity relationship derived, the choice of the

appropriate molecule is very important. On the use of carbon and nitrogen monoxides as

probes for the surface cationic centers it is evident that the carbon and nitrogen monoxides are

very weak bases and largely used for the surface characterization of cationic centers on metal

oxide surfaces [176-179].

CO implies a triple bond between C and O, according to the literature, the stretching

frequency for the free molecule in the CO gas is measured at 2143 cm-1 (see the roto-

vibrational absorption band in Fig. 6). In principle, the electrons are distributed symmetrically

between C and O atoms, so that the lower positive charge of the C nucleus with respect to O

implies the formation of a dipole with the negative charge at the C atom in spite of the lower

electronegativity of C with respect to the O nucleus. For this reason, the CO molecule tends to

interact through the C end with cationic centers. This interaction is rather weak, usually

completely reversible by outgassing at room temperature and should be studied at room or

lower temperature (e.g. at liquid nitrogen, 77 K) [177].

According to theoretical calculations, the metal-CO interaction is a simple polarization, with

no formation of a true coordinative σ-bond with the cationic center. This interaction tends to

increase the CO bond order, so that the CO stretching frequency tends to increase upon it.

Accordingly, the experimental measure of the CO stretching frequency for CO interacting

with surface cations can be taken as a measure of the polarizing power of the cation or, in

other terms, of its Lewis acidity. However, when the cation or the metal atom contains,

besides empty orbitals, also full or partly filled d-type orbitals, they can interact with the

empty π*- type orbitals of CO, via a π-type electron backdonation from the metal to CO. This

implies that these antibonding orbitals become partly filled so that the bond order and the CO

stretching frequency are decreased by this last interaction [178-181]. In this case, the

interaction can become very strong and very stable metal-carbonyl complexes (carbonates)

can be formed. The experimental CO stretching frequency in this case is a complex function

of the electron accepting power of the cation (Lewis acidity) and of its π-type electron

38

Page 39: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

donating power. Accordingly, the CO stretching frequency of CO adsorbed on several metal

cations is very informative on the oxidation state of the adsorbing ion. The associated process

can be easily followed by IR spectroscopy [179-182].

In fact, acidic OH groups form very weak hydrogen bonded adducts with both CO and H2

probes and the shift of the stretching frequency induced by the perturbation is proportional to

the charge present on the hydrogen (and hence indirectly upon its Brönsted acidity). On the

contrary, basic OH groups do not show any tendency to interact with CO and H2 [183-187].

Yet, a few have been published on IR study of CO adsorption on Mo-containing catalysts in

which the Mo loading was mostly 8-10 %. For instance, Zs. Németh undertook a low

temperature IR study at -196°C of the interaction between Mo/Al2O3 (8 %) and CO within the

extent of reduction from 500°C up to 900°C. A correlation was found between the extent of

reduction and the increase of adsorbed CO species. Furthermore, the five bands detected

(1991, 2025, 2052, 2159, 2205 cm-1) were assigned to molybdenum having different valence

states as the result of the higher reduction without the detection of adsorbed carbonate species.

Moreover, two bands at 1991 and 2025 cm-1 assigned to metallic Mo0 were observable after

reduction at 700°C [188, 189].

Fig. 6. FTIR roto-vibrational spectra of gaseous CO (left) and NO (right)

39

Page 40: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

1.6. OBJECTIVES

Knowing that a lot of researchers prepared and studied Mo/Al2O3 catalysts with low Mo

loadings (8-10 wt%), not taking into account the isoelectric point of the solid support and the

calcination temperature and time, since as expected free MoO3 clusters could not be produced

at Mo loading lower than 15 wt% on Al2O3 depending on the calcination temperature and

time. For this reason, my first objective was to prepare MoO3-rich catalysts and thus how

affects the final catalyst.

The second objective was to explore the influence of ceria and tin (as either promoters or

supports) on molybdena, the changes with adsorption of molybdates in comparison with the

Mo/Al2O3 catalyst since IR and thermal analysis studies regarding these systems are rather

limited. So that the knowledge gained from the catalyst formation should be helpful in the

characterization of the resulting catalyst.

A complementary approach, to help reducing such uncertainties in composition and bonding,

is to study these catalysts, which are chosen to mimick the real catalysts as closely as possible

in terms of overall chemical composition. Therefore, the third objective was to study these

model catalysts by various characterization techniques. Each characterization technique has

its limitations, thus it is dangerous to rely on one method. Whenever, various techniques were

applied, not only to obtain supplemental information about the catalysts but also to serve

mutual checks. These included the multipoint BET analysis method, X-ray diffraction (XRD),

Thermal analysis (TG-DTA), Electron Spin Resonance (ESR) and Diffuse reflectance Fourier

Transform Infrared (DRIFT) spectroscopy.

Additionally, their reduction characteristics with H2 and their activity towards CO adsorption

and CH4 decomposition were also aimed to provide insight into their surface characteristics.

The overall objective was to seek correlations and differences between these catalysts to

collate the results providing better understanding and contributing to the identification of the

physicochemical characteristics of Mo-containing catalysts that could influence their catalytic

behaviour in many reactions.

40

Page 41: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

2. EXPERIMENTAL

2.1. Materials and catalyst preparation The γ-Al2O3 support (Ketjen CK300), with BET surface area of 200 m2/g, mean pore diameter

determined by the BJH method of 6.37 nm and pore volume of 0.51 cm3/g, was used for

catalyst preparation. Additionally, the CeO2 formed by the thermal decomposition of

Ce(NO3)3·6H2O (Lobo Feinchemie) in air at 600°C for 6 hrs was found to have a surface area

of 80 m2/g with a mean pore diameter of 12.8 nm and pore volume of 0.016 cm3/g. The SnO2

(purity of 99%, Vega Refinery, Romania) with a surface area of 9 m2/g was found to have a

mean pore diameter of 23.54 nm and pore volume of 0.026 cm3/g.

The Mo/Al2O3, Mo/CeO2 and Mo/SnO2 catalysts were prepared by impregnation of the

supports (γ-Al2O3, CeO2, SnO2) with an aqueous ammonium heptamolybdate (NH4)6

Mo7O24.4H2O (Merck) solution at pH = 2, the pH was adjusted by adding 96 v/v% HNO3.

The concentration of heptamolybdate was that required to not only obtain surface

concentrations close to the theoretical monolayer coverage but also go beyond a compact

single lamella of molybdenum oxide structure [12-15]. The contact time for a given

concentration with the solid (γ-Al2O3, CeO2, SnO2) was extended for 72 hrs and the

temperature of the rotary shaker was maintained at 25°C.

In the second method, the CeO2 and SnO2 oxides were mechanically mixed with 20 wt%

Mo/Al2O3 (impregnated and dried) by adding 10 v/v% HNO3 to prepare CeO2-20 wt%

Mo/Al2O3 and SnO2-20 wt% Mo/Al2O3 samples (containing 5 wt% of CeO2 and 5 wt% of

SnO2, designated as Ce-Mo/Al2O3 and Sn-Mo/Al2O3).

For loadings corresponding to more than 15 wt% Mo on alumina, both Al2(MoO4)3 and bulk

MoO3 are reported with their formation being favoured by both increased calcination

temperature (up to 700°C) and increased calcination time [12-15, 32-35]. So that all samples

were dried in air at 150°C for 6 hrs followed by calcination in air at 600°C for 12 hrs.

41

Page 42: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

2.2. Catalyst characterization methods and techniques a) Inductive Coupled Plasma–Atomic Absorption Spectroscopy (ICP-AAS) was

performed in order to verify the amounts of Mo loading. Therefore, the molybdenum loading

was 20 wt% in Mo/Al2O3, 12 wt% in Mo/CeO2 and 4 wt% in Mo/SnO2 samples.

b) Total surface area, SBET, was calculated using the multipoint BET analysis method.

The assessment of mesoporosity was estimated from the adsorption branches via BJH method.

The nitrogen adsorption/desorption isotherms were measured at -196°C using a conventional

volumetric apparatus.

c) The crystalline structure and particle size were determined by X-ray diffraction

(XRD) experiments performed on a Philips PW3710 diffractometer run with Ni-filtered

copper (CuKα) radiation at 50 kV and 10 mA with a scanning speed of 2θ = 2.5°/min and

X-ray wavelength of λ = 0.154056 nm. The XRD patterns were obtained at room temperature

and crystalline phases were identified by comparison with PDF (Powder Diffraction Files)

standards from ICDD.

d) TG and DTA analyses were conducted for 50 mg of uncalcined samples using

automatically recorded Labsys-TG (France) units. The rate of heating was 10°C/min under a

flow of dry air (Messer, Hungary) from room temperature up to 900°C.

e) The Electronic Paramagnetic Resonance (EPR) measurements were carried out with a

Bruker ELEXSYS 500 spectrometer operating at the X-band microwave frequency of 9.29

GHz and magnetic field modulation of 100 kHz, with ER041 XG microwave bridge (a

microwave power of 1 mW and a modulation amplitude of 10 G, recording time of 1 s).

The EPR spectra were recorded and analyzed (using the Win-EPR data acquisition software)

for the calcined samples before and after CO adsorption at room temperature and at -196°C

using a conventional liquid nitrogen flow system.

f) Diffuse Reflectance Fourier Transform Infrared (DRIFT) spectroscopic studies were

performed by Bruker Tensor 27 spectrometer with dry air cooled DTGS detector. In addition,

the sample compartment is always flushed with dry air to avoid detecting of environmental

gases such as CO2 and H2O. In all cases, in situ DRIFT spectra were recorded using

42

Page 43: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Thermo-Nicolet DRIFT cell by co-addition of 126 scans at 4 cm-1 resolution. The cell

equipped with Thermo-Nicolet chamber specially designed for high temperature/vacuum

studies (up to 900°C) was fitted with KBr disks for low temperature and with CaF2 disks for

high temperature measurements. The operating temperature with deviation of ± 3°C was

controlled by using Thermo-Nicolet proportional temperature controller connected to the cell.

In situ DRIFT spectra expressed as absorbance unit versus frequency were collected after

taking the background spectrum for each experiment and can be evaluated accordingly.

Spectral manipulations such as baseline correction, smoothing, normalization, band

component analysis and deconvolution were performed using “OPUS” and “GRAMS”

software packages. In order to enhance the apparent resolution of a spectrum, or to decrease

the line width as well as spectral ranges comprising broad and overlapping lines can thus be

separated into sharp single lines using Fourier Self-Deconvolution function.

Hence, the DRIFT reactor cell was attached to a glass gas handling/ultrahigh vacuum system

and can be evacuated at up to 1.33x10-8 bar. Thus, with the aid of this system the following

gases were used for treatments and reactions:

• Reduction in hydrogen (Messer, purity of 99.995%)

• CO (Messer, purity of 99.99%)

• O2 and N2 (Messer, purity of 99.6%)

• CH4 (Fluka, purity of 99.99%)

• He (99.99% purity)

43

Page 44: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3. RESULTS 3.1. Surface texturing The Mo surface density is expressed as the number of Mo atoms per nanometer square of

surface area (Mo atoms/nm2). It was obtained by the equation [65]:

Mo surface density = (MoO3% x 6.02 x 1023) / (surface area x 144 x 1018) (Eq-22) Where the unit of the surface area is m2/g and 144 is the molar weight of MoO3. The values of

surface areas, cumulative pore volume, Vp (cm3g−1) and the average mesopore diameter for

the calcined materials were estimated and cited in Table 4, whereas Figs. 7, 8, 9 and 10 show

the cumulative mesopore volume and differential pore distribution curves versus pore

diameter plot of the samples calcined at 600°C.

The Mo/Al2O3 catalyst has an average mesopore diameter around 6.2 nm and a continuous

distribution of larger pores that extend well into the macroporous region. As this sample

contains high Mo amount, its texture may be related to Al2O3 because the pore size

distribution curve obtained for Al2O3 (not shown) is very similar. Accordingly, it may be

inferred that the large pores present in the Mo/Al2O3 sample may be related to the interstices

resulting from the packing of the Al2O3 particles, which behave as pores in the N2 adsorption

measurements. The Ce-Mo/Al2O3 catalyst has an average mesopore diameter around 6.9 nm

and being these mesopores of larger diameter in contrast with Mo/Al2O3. More specifically,

upon addition of cerium the average mesopore size increased in Ce-Mo/Al2O3 sample. This

increase in the average mesopore size is responsible for the significant reduction of the

specific surface area of this sample (Table 4).

This finding might indicate the enforced location of Ce species in the pores of this sample

leading to an effective pore widening. Increasing the pore diameter of the former sample

when compared with the others is indicative to the role played by cerium-molybdate and

CeO2 and to their partial deposition inside the micropores creating their own pore system and

thus enhancing the pore diameter. On the other hand, these species might be deposited on

internal surfaces blocking some pores in Ce-Mo/Al2O3 and thus, reducing the surface area

comparatively. Nevertheless, some Mo nanoparticles presumably entrapped in the micropores

of ceria oxide that can also cause expansion in unit cell and pore diameter.

The Sn-Mo/Al2O3 catalyst has an average mesopore diameter around 5.85 nm. Decreasing the

pore radius of this sample when compared with the rest of samples indicates the probability of

the presence of Sn species as separate phases probably as oxides. On the other hand, some Sn

44

Page 45: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

particles might be entrapped in internal surfaces blocking some pores in Mo/Al2O3 and thus,

reducing the surface area (34 m2/g) comparatively.

For Mo/SnO2 catalyst, the cumulative mesopore volume distribution is bimodal (Fig. 10),

with an average mesopore diameter around 22.43 nm. On the other hand, the pore diameter

size decreased slightly for Mo/CeO2 and Mo/SnO2 samples in comparison with CeO2 and

SnO2. This may propose that the diffusion of MoO3 did not proceed into the CeO2 and SnO2

bulk during the preparation.

Table 4. Textural characteristics derived from BET method for the materials calcined at 600°C

Catalyst

Average mesopore

diameter (nm)

Total mesopore volume (cm3/g)

BET Surface

area (m2/g)

Mo density (atom/nm2)

Atom ratio (Mo/M)

γ-Al2O3 6.37 0.51 200 - - 20 wt%

Mo/Al2O36.20 0.225 117.6 7 Mo/Al=0.134

5 wt% Ce-20 wt% Mo/Al2O3

6.9 0.072 29.6 28 Mo/Ce=6.67

CeO2 12.8 0.016 80 - - 12 wt%

Mo/CeO212.5 0.0053 16.2 31 Mo/Ce=0.24

SnO2 23.54 0.026 9 - - 5 wt% Sn-20

wt% Mo/Al2O35.85 0.073 34 24.6 Mo/Sn=6

4 wt% Mo/SnO2 22.43 0.023 3.86 43 Mo/Sn=0.066

45

Page 46: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 7. Cumulative mesopore volume distribution vs. pore diameter of Mo/Al2O3,

Mo/CeO2 and Ce-Mo/Al2O3 samples

Fig. 8. Differential pore distribution vs. pore diameter of Mo/Al2O3, Mo/CeO2 and

Ce-Mo/Al2O3 samples

46

Page 47: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 9. Cumulative mesopore volume distribution vs. pore diameter of Mo/SnO2 and Sn-Mo/Al2O3 samples

Fig. 10. Differential pore distribution vs. pore diameter of Mo/SnO2 and Sn-Mo/Al2O3

samples

47

Page 48: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.2. X-ray diffraction Figs. 11 and 13 show the diffractograms of the samples calcined at 600°C. The values of the

average crystallite size ( d ) of the crystalline phases were calculated by using the

Debye–Scherrer equation [1]:

θβλ

coskd = (Eq-23)

Where λ = 0.154056 nm, is the X-ray wavelength, k the particle shape factor (0.9 for cubic

particles), β the full width of the peak considered at half maximum (in rad), and θ is the half

of the diffraction angle.

3.2.1. XRD patterns of Mo/Al2O3, Ce-Mo/Al2O3 and Mo/CeO2 It can be seen, the XRD patterns of Mo/Al2O3 exhibited diffraction lines and d-spacing values

corresponding to Al2O3 (tetrahedral and octahedral), orthorhombic Al2(MoO4)3 and octahedral

MoO3 crystallites (Fig. 11). Moreover, the profile of XRD peaks of Mo crystallites can be

attributed to different types of particles: (i) huge particles, responsible for the fine part of the

peak with high intensity (Al2(MoO4)3), and (ii) smaller free MoO3 crystallites giving a low

intensity of the peak (broadening/low intensity peaks) from the XRD patterns of crystalline

species. This can be related to a less crystalline phase and well-dispersed MoO3 and/or MoO3

particles are under limit to be detected by XRD technique (< 2 nm). Indeed, the calculated

average crystallite size of MoO3 was 1.7-3.4 nm in Mo/Al2O3 sample (Table 5).

Consequently, the XRD crystallite size calculation takes into account essentially the biggest

crystallites. However, taking into account that the ionic radius of Mo6+ (0.62 Å) is larger than

that of Al3+ (0.50 Å), therefore, the Mo6+ ions do not incorporate into the lattice of Al2O3 to

replace Al3+ ions suggesting that some molybdena supported on the Al2O3 may spread over

the support. Nevertheless, Mo can diffuse into defect sites of alumina, and/or interact with the

surface hydroxyl groups of alumina becoming difficult reducible and strongly bound to form

Mo-O-Al bonds in the form of Al2(MoO4)3 [12-17, 42-47].

On the other hand, the diffractograms of Ce-Mo/Al2O3 and Mo/CeO2 exhibited diffraction

lines and d-spacing values at 3.7, 3.1, 2.7, 1.9 and 1.6 Å that correspond to (110), (111),

(200), (220) and (311) phases, respectively, of the CeO2 cubic structure [53-65]. Furthermore,

the XRD patterns of Ce-Mo/Al2O3 display a noticeable decrease of the line intensities and

crystallite size of MoO3 (1.4-2.7 nm) and Al2(MoO4)3 (2.8-5.6 nm) phases in the latter

material which was perceived comparatively to Mo/Al2O3 patterns upon the addition of ceria.

48

Page 49: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

The diffraction patterns of Mo/CeO2 material exhibited well-ordered MoO3 crystallites at d-

spacing of 6.25, 3.9, 3.4, 2.83 and 1.84 Å. Thus, the increase in crystallinity of the former

sample could be caused by the presence of more cerium-molybdate in contrast with Ce-

Mo/Al2O3 that became predominant on the small particles, whereas different surfaces such as

(110) and (111) CeO2 phases were exposed on large CeO2 particles (Table 5).

Generally, this sample revealed the presence of new lines with d-spacing values pertaining to

various phases including CeO2 and cerium-molybdate (three crystalline structures) those most

probably exposed to the external surface in contrast with Ce-Mo/Al2O3 sample in which only

one crystalline structure of cerium-molybdate (Ce2Mo3O12) was found (Fig. 11).

The value of the average particle size of MoO3 crystallites in Mo/CeO2 was between 1.8-4.6

nm, which was higher (within the experimental error) than that of 1.4-2.7 nm in

Ce-Mo/Al2O3. On the contrary, the average particle size of CeO2 crystallites was lower in

Mo/CeO2 (Table 5). This may demonstrate that incorporating Mo in ceria, in Mo/CeO2,

effectively prevents particles agglomeration allowing the material to maintain its dispersion.

The presence of the residual peaks of (110) and (111) phase of CeO2, which is the dominant

phase forming cubic ceria, indicates that ceria structure is still intact. Ceria crystallizes in a

cubic fluorite structure where each cerium cation is coordinated by eight equivalents nearest

neighbour oxygen anions at the corner of the cube. The models for clean (111) and (110)

surfaces are shown in Fig. 12, respectively, and are fully relaxed under the restriction of fixed

cell parameters and fixed geometry of the lower layers. The CeO2 (111) surface relaxation is

quite small, with the Ce sub-lattice remaining unperturbed and with the [O] sub-lattice

undergoing small changes in inter-layer distances of around 0.03 Å. The displacements show

a slight contraction of the first few layers for the (111) surface. The relaxation of the (110)

surface exhibits a reverse behaviour compared to that of the (111) surface. Specifically, the

oxygen sub-lattice remains essentially unaffected, while the cerium sub-lattice undergoes

larger changes in the inter-planar distances of ±0.23 Å [66-72].

The flat ideal (110) surface becomes slightly rumpled upon relaxation, with Ce atoms shifted

towards the center of the slab by 0.13 Å relative to the surface O atoms, this equals 2.4% of

the ceria bulk lattice constant. In addition, the Ce–O bond length at the surface contracts by

0.04–2.324 Å compared to the bond length in the bulk by 2.36 Å [73-79].

49

Page 50: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 11. XRD patterns of Mo/Al2O3, Ce-Mo/Al2O3 and Mo/CeO2 calcined at 600°C

Table 5. Crystallite sizes derived from XRD data of Mo/Al2O3, Ce-Mo/Al2O3 and Mo/CeO2 materials calcined at 600ºC

The average crystallite sizes (nm) Catalyst CeO2 MoO3 Al2(MoO4)3

Mo/Al2O3 - 1.7-3.4 3.3-6.9 Ce-Mo/Al2O3 17-20 1.4-2.7 2.8-5.6

Mo/CeO2 7-9 1.8-4.6 -

Fig. 12. Slab models for ceria surfaces (111 and 110), where the yellow and red spheres

represent Ce and O atoms.

50

Page 51: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.2.2. XRD patterns of Sn-Mo/Al2O3 and Mo/SnO2 Fig. 13 displays the X-ray diffraction patterns of Sn-Mo/Al2O3 and Mo/SnO2 catalysts

calcined at 600°C. The diffractogram of Sn-Mo/Al2O3 exhibits reflection peaks and d-spacing

values corresponding to Al2O3, orthorhombic Al2(MoO4)3, tetragonal SnO2 (cassiterite) but

does not show reflections related to crystalline MoO3. On the contrary, phase transformations

either MoO3 into monoclinic MoO2 or SnO2 into SnO were observed.

On the other hand, the XRD patterns of Mo/SnO2 show reflection peaks and d-spacing values

attributed only to two phases in the form of MoO3 (crystallite size between 7-10 nm) and

SnO2 (crystallite size between 3.2-5.8 nm) (Table 6). However, assuming for simplicity a

completely ionic structure, Mo6+ ions, with atomic radius of 0.62 Å can migrate to the surface

and from the surface to the sub-layers. The oxygen vacancies of the surface and of the sub-

layers can easily reduce Mo6+ to Mo5+ or Mo4+ with atomic radius of 0.63 and 0.65 Å,

respectively. As Sn4+ has atomic radius of 0.71 Å, Mo6+, Mo5+ and Mo4+ can easily occupy

the tin lattice sites giving rise to a solid solution. Despite this fact, no mixed oxide phase was

observed between Mo and Sn ions in both samples even after the calcination at 600°C. Since

the Mo/SnO2 catalyst obtained by impregnation only showed the presence of the cassiterite

phase and MoO3, while the Sn-Mo/Al2O3 catalyst obtained by co-precipitation showed the

presence of MoO2, SnO2 and SnO without reflection peaks related to MoO3.

Tin oxide SnO2, in its pure form, has the tetragonal crystalline structure at room temperature

and normal pressure [84-101]. SnO2 is a more densely packed crystal where each tin atom is

surrounded by a slightly distorted oxygen octahedron while in SnO the tin atoms sit on the

vertices of pyramids with an oxygen square basis (Fig. 14).

51

Page 52: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 13. XRD patterns of Sn-Mo/Al2O3 and Mo/SnO2 calcined at 600°C

Table 6. Crystallite sizes derived from XRD data of Sn-Mo/Al2O3 and Mo/SnO2 materials

calcined at 600ºC

The average crystallite sizes (nm) Catalyst MoO3 MoO2 SnO2 SnO Al2(MoO4)3Sn-Mo/Al2O3 - 1.7-2.3 1.7-4.5 1.8-2.4 2.8-6.2

Mo/SnO2 7-10 - 3.2-5.8 - -

52

Page 53: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 14 shows atomic configurations of SnO2 and SnO. (a) is the structure of SnO2 unit cell

where the dashed lines link the O atoms forming an octahedron surrounding a tin atom. (b) is

the structure of SnO unit cell where the dashed lines link the O atoms and the tin atom

forming square-based pyramids. (c) shows the projection of the SnO2 unit cell onto a (100)

plane showing the traces of the alternate O and Sn atomic planes and (d) shows the projection

of SnO unit cell onto a (010) plane showing the traces of the O and Sn atomic planes.

Fig. 14. Atomic configurations of SnO2 and SnO: (a) Structure of SnO2 unit cell. (b) Structure of SnO unit cell. (c) Projection of the SnO2 unit cell onto a (100) plane showing the traces of

the alternate O and Sn atomic planes. (d) Projection of SnO unit cell onto a (010) plane showing the traces of the O and Sn atomic planes.

53

Page 54: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

On the other hand, the coordination of the metal atom in MoO3 can be best considered as that

of a markedly distorted octahedron, although it can be easily deduced from the MoO3

tetrahedron as a basic unit. However, the clean (100) MoO3 has a layer structure in which

each layer is built up of MoO6 octahedrons at two levels, connected along the z-axis by

common edges and corners, so as to forming zigzag rows and along x-axis by common

corners only. Moreover, each layer exhibits, in the direction of z-axis, oxygen atoms, which

are common for three different octahedrons. Each octahedron also shares, along x-axis, two

oxygen atoms with two neighbouring octahedrons. Besides, for each MoO6 octahedron there

is only one oxygen atom which is doubly bound (Mo=O) to the molybdenum atom. It

occupies different positions along the y-axis. For each MoO6 octahedron at the higher level

this oxygen atom points up. On the contrary, for each MoO6 octahedron at the lower level this

oxygen atom points down [194-197].

The structure of the MoO6 octahedron is shown in Fig. 15. In addition, the idealized

arrangement of molybdenum and oxygen atoms on the (100) plane of MoO3 is shown in

Fig. 16. The clean (100) plane presents coordinately unsaturated Mo6+ ions with one bridging

O2- ion missing from their coordination sphere. Thus, it acquires the formal uncompensated

charge (+1). This plane also contains unsaturated bridging O2- ions with one Mo6+ missing

and a formal uncompensated charge of (-1). According to this approach, two types of surface

oxygen atoms can be distinguished: tightly bound inactive lattice oxygen atoms and weakly

bound active lattice oxygen atoms that may play an important role in MoO3 activity.

54

Page 55: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 15. The MoO6 octahedron structure. Mo (shaded circles), O (open circles).

Fig. 16. Space fill idealised (100) MoO3 face. The uncompensated charges are also shown.

Mo (shaded circles), O (open circles).

55

Page 56: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.3. Thermal Analysis The thermogravimetric (mass loss, TG), derivative thermogravimetric (rate of mass loss,

DTG) and differential thermal analysis (DTA) curves of 50 mg of uncalcined samples, over

the range from room temperature up to 900°C in the flow of an atmosphere of dry air, are

shown in Figs. 18, 19, 20, 21 and 22.

3.3.1. TG and DTA of Mo/Al2O3 The TG-DTG and DTA curves recorded in argon flow for pure (NH4)6Mo7O24.4H2O (AHMT)

at a scanning speed of 5°C/min up to 600°C are shown in Fig. 17. The TG-DTG exhibit three

decomposition peaks with the maximum located at 127, 220 and 307°C respectively. The TG

shows that AHMT loses weight with heating over three steps.

The first endothermic peak at 100-127°C was followed by 7.3% mass loss corresponding to

the loss of three H2O molecules and two molecules of NH3, which is consistent with the

theoretical value (according to the literature, the theoretical weight loss, assuming release of

three water molecules and two NH3 molecules is 7.1%) [160–162]. The second endothermic

peak at 200-225°C which was accompanied by a mass loss of 4.4% is probably related to the

elimination of (NH4)2O (i.e. 2NH3 + H2O) which is in agreement with the theoretical value

(4.2%). The third endothermic peak takes place within 267-327°C was associated with 7.4%

weight loss (theoretical value 7.1%) and corresponds to the evolution of two NH3 molecules

together with three molecules of H2O. Moreover, the exothermic peak of maxima at 327°C

confirms the crystallization of new phases (probably MoO3). Therefore, the overall mass loss

was 19.1%. These three stages can be explained to proceed according to the following

equations:

(NH4)6Mo7O24.4H2O (NH⎯⎯⎯⎯⎯⎯ °− C127100 →

4)4Mo7O23.2H2O + 3H2O + 2NH3 (Eq-24)

(NH4)4Mo7O23.2H2O (NH⎯⎯⎯⎯⎯⎯ °− C2252004)2Mo7O22.2H2O + H2O + 2NH3 (Eq-25)

(NH4)2Mo7O22.2H2O ⎯⎯⎯⎯⎯⎯ °− C327267 7MoO3 + 3H2O + 2NH3 (Eq-26)

56

Page 57: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 17. TG-DTG and DTA curves for AHMT in argon atmosphere

The TG-DTG and DTA curves recorded in dry air flow for Mo/Al2O3 exhibit a large mass loss

(37.8%) up to 300°C with an endothermic peak of maxima at 83°C in the DTA thermogram

(Fig. 18). This is corresponding to the loss of structural and intercalated water and to the

decomposition of (NH4)6 Mo7O24.4H2O [160, 161].

The DTA thermogram of this sample displays a very big exothermic peak of a maximum at

around 300°C that could be due to the process of the formation of new phases probably to the

crystallization of MoO3 species [162-165].

However, the three decomposition steps of (NH4)6Mo7O24.4H2O are missed on the TG-DTA

curves of Mo/Al2O3 due to several factors such as: first, Mo/Al2O3 is impregnated and dried,

second, the heating atmosphere (dry air) and heating rate (10°C/min) are different from those

(argon flow, 5°C/min) applied for thermal analysis of pure (AHMT).

57

Page 58: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 18. TG-DTG and DTA curves of Mo/Al2O3

3.3.2. TG and DTA of Ce-Mo/Al2O3 and Mo/CeO2 TG-DTG and DTA curves recorded for Ce-Mo/Al2O3 (Fig. 19) indicate that the weight loss

(20.5%) under 300°C is less than that of Mo/Al2O3. This mass loss is due to the removal of

water and to the decomposition of Ce(NO3)3 and (NH4)6 Mo7O24.4H2O producing the

corresponding oxides according to the equation:

(NH4)6 Mo7O24.4H2O + Ce(NO3)3 + 2O2 → 7MoO3 + CeO2 + 3NO2 + 6NH3 + 12H2O

(Eq-27)

This indicates that the weight loss is highly affected by the mode of the preparation that in

turn affects the interaction mode of CeO2–MoO3 compounds. The latter affects by its turn the

thermal stabilities of the produced compounds. In the meanwhile, the position of the

endothermic peak shifts to a higher value with a maximum at 101°C, whereas the big

exothermic peak remained with the maximum at around 300°C comparatively.

58

Page 59: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Of particular interest, the TG and DTA of Ce-Mo/Al2O3 display a large mass loss (23%)

extended between 700 and 900°C with an endothermic peak of maxima at 758°C indicating

that the material undergoes morphological and structural modifications. This may be due to

the high oxygen storage and release of ceria resulting in lattice defects and thus enhancing the

mobility of Mo and Ce ions. Thus, the crystalline structure is progressively modified

comparatively favoring the possible mutual diffusion of the Mo and Ce ions. The last mass

loss in this material may characterize the sublimation of some molybdena produced [65, 80].

The thermal decomposition course presented for Mo/CeO2 is shown to be of multistep and

represented by four endothermic peaks of maxima at 61, 98, 173 and 737°C in the DTA curve

(Fig. 20) besides a big exothermic peak of a maximum at 308°C. As can be seen, this material

contributes a total weight loss around 34.7%. On the other hand, it is worth mentioning that

the decomposition temperature of either Mo or Ce precursor salts in the synthesized materials

are not similar revealing that the interactions between them are varied according to altering

the preparation methods and their content, which indeed affect the mobility of Mo species.

Fig. 19. TG-DTG and DTA curves of Ce-Mo/Al2O3

59

Page 60: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 20. TG-DTG and DTA curves of Mo/CeO2

3.3.3. TG and DTA of Sn-Mo/Al2O3 and Mo/SnO2 On the TG curve of Sn-Mo/Al2O3 a large weight loss step is marked (38.3%) quite similar to

that of Mo/Al2O3 (37.8%) comparatively implying that the addition of SnO2 did not affect the

thermal behaviour of Mo/Al2O3. This step with 38.3% mass loss of the original mass

corresponds to dehydration and decomposition of ammonium heptamolybdate (Fig. 21).

The DTA curve of Sn-Mo/Al2O3 shows three endothermic peaks at 69, 103 and 211°C

indicating the presence of weakly bounded, strongly bonded and structural water species.

The TG curve of Mo/SnO2 is definitely different (Fig. 22). Two mass loss steps are observed.

The rate in the first stage of its mass loss between 70 and 300°C is relatively low (11.4%),

while the mass loss in the second stage is higher (41.7%). The overall weight loss measured

up to 900°C is 53.1%. The DTA curve of Mo/SnO2 shows first an endothermic dehydration

occurring between 70 and 200°C (temperature peaks marked at 71 and 197°C). It is followed

by the exothermic effect starting from 300°C (peaks marked at 327 and 396°C) indicating the

process of crystallization in this sample (probably formation of molybdenum oxide).

60

Page 61: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Nevertheless, the DTA peaks centred at 797 and 846°C explicitly are two strongly

endothermic heat effects of the main degradation presumably due to the release of some

volatile tin and molybdena species and/or due to surface and structure modifications of the

contact between phases leading to lattice defects, and thus enhancing the mobility of the two

ions and mutual interaction between them. These changes indicate that Mo can migrate, some

Mo and Sn species can be destroyed, while others are created or increased. It is, however,

hard to estimate if new segregated phases are formed upon thermal treatment above 750°C

that can be followed more certainly by TG-MS studies [83].

Fig. 21. TG-DTG and DTA curves of Sn-Mo/Al2O3

Fig. 22. TG-DTG and DTA curves of Mo/SnO2

61

Page 62: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.4. Electron Spin Resonance (ESR) measurements The EPR spectra recorded at room and liquid nitrogen temperature for Mo/Al2O3, Mo/CeO2

and Mo/SnO2 samples calcined at 600°C before and after interaction with CO are shown in

Figs. 23, 24, 25 and 26.

The EPR spectra can best be analysed using a spin-Hamiltonian of the form: H = βSgB + SAI,

where the symbols have their usual meaning (Eq-21). It contains the electronic Zeeman term

(β is the Bohr magneton, S = 1/2 the electron spin, g is the g-tensor and B is the applied field),

perturbed by the hyper-fine coupling term between the unpaired electron and the nuclear spin

of molybdenum (I), whereas (A) being the hyperfine structure tensor [166-169].

The EPR spectra of calcined Mo/Al2O3 show a weak hyperfine structure with six lines in both

parallel (g║) and perpendicular (g┴) bands (Fig. 23). The intense central line arises even from 96Mo isotope which has the nuclear spin I = 0, while the lower intensity lines correspond to

the hyperfine structure from the 95Mo and 97Mo isotopes which have the nuclear spin I= 5/2.

However, the nuclear magnetic moments of 95Mo and 97Mo being close, so the isotope

splitting is not resolved in the EPR spectra.

The EPR lines of Mo/Al2O3 between 3400-3600G becoming more prominent at both room

temperature and -196°C after CO reaction due to the presence of Mo5+ paramagnetic centers

(Mo6+ diamagnetic) and hence there is an increase of concentration of paramagnetic centers in

this sample. Accordingly, Mo6+ ions were reduced by CO to Mo5+. It should be mentioned

that Al2O3 does not give rise to EPR lines. Thus, the spectra obtained for Mo/Al2O3 may be

considered as the hyperfine structure typical for isolated Mo5+ ions and its widening is due to

the contribution of the clustered Mo5+ ions coupled by the dipolar interactions between Mo5+

paramagnetic ions (between the electron spin and its surrounding nuclei) leading to the line

broadening [170-174].

The EPR spectra of Mo/CeO2 show (Fig. 24) a complex EPR signal (due to the phase

containing a mixture of oxides) generated by the two Mo5+ and Ce3+ paramagnetic centers

present in the sample after CO interaction. Thus, both the support and the active phase are

reduced by CO. More specifically, the band at 3370G is assigned to the Ce3+ ions, while the

bands between 3400-3600G are assigned to Mo5+ ions (both Mo6+ and Ce4+ are diamagnetic).

On the other hand, the EPR spectra of Mo/SnO2 show the presence of Mo5+ ions after CO

interaction, the EPR spectrum of Mo/SnO2 did not show any resonance signals of O- and O2-

paramagnetic species neither before nor after CO contact (Fig. 25). Meanwhile, Vo oxygen

vacancy was observed in powdered Mo/SnO2 sample at both temperatures (r.t. and –196°C)

62

Page 63: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

after CO interaction. Accordingly, it is attributed to an oxygen defect, the single ionised

oxygen vacancy Vo, produced by CO interaction:

CO + Oo ↔ Vo + CO2 (Eq-28)

Vo ↔Vo+ + e– (Eq-29)

Where Oo is the lattice oxygen, V0 is a neutral oxygen vacancy and e– an electron in the

conduction band. It is worth mentioning that the EPR spectrum of calcined Mo/SnO2 did not

show any resonance signals due to the fact that both Mo6+ and Sn4+ are diamagnetic.

Fig. 23. The EPR spectra of Mo/Al2O3 recorded at –196°C before and after CO contact

Fig. 24. The EPR spectra of Mo/CeO2 recorded at –196°C before and after CO contact

63

Page 64: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 25. The EPR spectra of Mo/SnO2 recorded at –196°C before and after CO contact

Fig. 26. The EPR spectra of the samples recorded at room temperature after CO contact

64

Page 65: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5. In situ DRIFT spectroscopy measurements 3.5.1. DRIFT spectra of the calcined samples under vacuum All DRIFT spectra were recorded at room temperature for samples calcined at 600°C after

evacuation at 1.33x10-8 bar. The IR assignments of metal-oxygen bonds are shown in Table 7.

3.5.1.1. DRIFT spectra of γ-Al2O3 and CeO2 The DRIFT spectra of the γ-Al2O3 and CeO2 supports are depicted in Fig. 27. The spectrum of

γ-Al2O3 displays bands at 416, 485, 690 and 744 cm-1 are tentatively assigned to terminal

νs(Al-O) (416 cm-1), νas(Al-O) (485 cm-1) and bridging νs(Al-O-Al) (690 cm-1), νas(Al-O-Al)

(744 cm-1) vibration modes [12-17]. On the other hand, the bands at 1017 and 1161 cm-1 can

be assigned to the in plane and the out of plane deformation modes of OH groups being in the

bulk of Al2O3. The band at 1633 cm-1 is related to the deformation vibration of water (H-OH).

The spectrum of CeO2 exhibits bands at 415 cm-1 νs(Ce-O) and 480 cm-1 νas(Ce-O) due to

symmetric and asymmetric stretching modes of Ce-O terminal bonds, while the bands at 687

and 740 cm-1 are assigned to νs(Ce-O-Ce) and νas(Ce-O-Ce) bridging bonds [65-73].

Additionally, two bands at 1016, 1153 and 1633 cm-1 can be attributed to the deformation

modes of OH groups and water as mentioned above. Moreover, the broad band in 3200-3800

cm-1 region in the two spectra is assigned to isolated and associated hydroxyl groups bound to

tetrahedrally and octahedrally coordinated Al3+ sites and to CeO2 surface [18-31, 74-80].

Fig. 27. In situ DRIFT spectra of γ-Al2O3 and CeO2 under vacuum

65

Page 66: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.1.2. DRIFT spectra of Mo/Al2O3, Mo/CeO2 and Ce-Mo/Al2O3 The spectrum of Mo/Al2O3 shows bands at 415 cm-1 and 490 cm-1 corresponding to bridging

stretching modes of νs(Mo-O-Al) and νas(Mo-O-Al) indicating the strong association of Mo

species (MoO42-) with Al2O3 support (Fig. 28) that was confirmed by XRD. These two bands

may have stemmed from overlapping with Al-O vibrations since they are broader than 416

and 485 cm-1 bands corresponding to (Al-O) vibrations. On the other hand, the band at 995

cm-1 that corresponds to νas(Mo=O) terminal stretching in bulk MoO3 indicates strong features

for microcrystalline MoO3 species [12-15]. One may notice that the oxygen anion of the

terminal Mo=O bond when exposed to the ambient atmosphere can easily interact with

moisture, resulting in the formation of hydrated surface species.

On the other hand, in order to help reducing uncertainties in composition and bonding of

Mo/Al2O3 the spectrum of 8 wt% Mo/Al2O3 (prepared by others at the Institute of

Environmental Engineering with 196 m2/g surface area) was recorded at the same conditions

to compare with that of 20 wt% Mo/Al2O3. As can be seen (Fig. 29), the bands at 423 and 512

cm-1 can be assigned to νs(Mo-O-Al) and νas(Mo-O-Al) vibration modes, whereas the band at

957 cm-1 is likely corresponding (terminal νasMo=O stretches) either to isolated tetrahedral

MoO42- species or bonded to alumina surface (Eq-5). However, the absence of this band in the

spectrum of 20 wt% Mo/Al2O3 confirms the evidence for the appearance of polymeric MoO42-

that is further emphasized by the band appearing at 749 cm-1 assigned to νs(Mo-O-Mo), which

shifts to higher frequency (757 cm-1) in Ce-Mo/Al2O3 spectrum [12-17].

The spectra of Mo/CeO2 and Ce-Mo/Al2O3 exhibit similar characteristic IR bands and can be

characterized by new bands at 630, 757, 875, and 1035 cm-1 with the absence of the bands at

415 and 490 cm-1 if compared with that of Mo/Al2O3 and CeO2 (Figs. 27 and 28).

However, the fact that the addition of ceria contributes to the formation of new phases so as to

forming Mo-O-Ce linkages that were represented by the bands at 630 cm-1 νs(Mo-O-Ce) and

875 cm-1 νas(Mo-O-Ce), while the bands separated by 40 cm-1 with different relative

intensities at 995 and 1035 cm-1 are associated with the formation of coupled νs(O=Mo=O)

bonds indicative of the presence of polymeric MoO3 implying that IR is more confined to too

small crystallites than XRD did.

Additionally, the proposed interaction between Mo and ceria is further emphasized by the

absence of the vibration modes of Ce-O at 415 and 687 cm-1 observed in CeO2 spectrum.

66

Page 67: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 28. In situ DRIFT spectra of 20% Mo/Al2O3, Ce-Mo/Al2O3 and Mo/CeO2 under vacuum

Fig. 29. In situ DRIFT spectra of 8% Mo/Al2O3 under vacuum

67

Page 68: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.1.3. DRIFT spectra of SnO2, Mo/SnO2 and Sn-Mo/Al2O3 The DRIFT spectra recorded of SnO2, Mo/SnO2 and Sn-Mo/Al2O3 samples are depicted in

Fig. 30. The spectra display bands at 1634, 740, 690, 630, 590 and 480 cm-1.

The bands at 480 νs(Sn-O) and 590 cm-1 νas(Sn-O) are assigned to symmetric and asymmetric

stretching modes of (Sn-O) terminal bonds. The bands centred at 630 and 690 cm-1

correspond to νs(O-Sn-O) and νas(O-Sn-O) vibrations. Moreover, the peak at 740 cm-1 is

assigned to bridged species of νas(Sn-O-Sn), while the discrete peak at 1031 cm-1 is related to

the lattice of different oxygen-bridged species of tin with the shoulder at 1123 cm-1 to

terminal γ(Sn-OH), besides the band at 1634 cm-1 observed in all samples due to the

deformation vibration mode of the H-O-H bond of water.

The spectrum of Mo/SnO2 exhibits additional band at 995 cm-1 attributed to νas(Mo=O)

terminal stretching in bulk MoO3 indicating strong features for free MoO3 clusters.

On the other hand, in the spectrum of Sn-Mo/Al2O3 the band at 761 cm-1 is related to

νs(Mo-O-Mo) associated with polymolybdates but it is noticeable to observe that the

vibrational peak at 995 cm-1 is absent in this spectrum in agreement with XRD results since

no MoO3 reflection peaks were detected by XRD.

Furthermore, no bands arising from Mo–O–Sn vibrations were observed (in line with XRD

results) in both Mo/SnO2 and Sn-Mo/Al2O3 samples that undoubtedly appear at 860-870 cm-1.

However, it is important to stress that the Mo–O–Mo unit is associated with polymolybdates,

while the Mo–O–Sn is related to isolated tetrahedral Mo species. Therefore, FTIR results

allow these two kinds of coordinations at hydrated conditions to be distinguished [91-98].

68

Page 69: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 30. In situ DRIFT spectra of SnO2, Mo/SnO2 and Sn-Mo/Al2O3 under vacuum

Table 7. Assignments of IR characteristics of metal-oxygen bonds

Assignment Wavenumber (cm-1) Reference Terminal νsCe-O, νasCe-O Bridging νsCe-O-Ce, νasCe-O-Ce

415, 480 690, 687, 740 65-80, 180-187

Terminal νsAl-O, νasAl-O Bridging νsAl-O-Al, νasAl-O-Al

416, 485 690, 744 12-31

In plane and out of plane deformation modes of OH groups

1123, 1153,1161 1016, 1017 82, 175-177

Deformation vibration of water 1633, 1634, 1637 25-31, 82 νsMo-O-Al, νasMo-O-Al stretching 415, 423, 490, 512 12-17, 32-37 Terminal νasMo=O 957, 995 12-15, 42-47 Coupled O=Mo=O 995 and 1035 12-15, 34, 65, 80 νsMo-O-Mo 749, 757, 761 12-17 νsMo-O-Ce, νasMo-O-Ce 630, 875 65, 79-81 Terminal νsSn-O, νasSn-O, νsO-Sn-O, νasO-Sn-O Bridging νsSn-O-Sn νasSn-O-Sn

480, 590 630, 690

740, 1031 82, 90-99

OH groups to tetrahedrally and octahedrally coordinated Al3+ sites

3200-3800 18-31

OH groups of SnO2 and CeO2 3200-3800 90-99, 180-187

69

Page 70: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.2. CO chemisorption

Hence, a new sample (50 mg) was used for each experiment and the pre-treatment for all

samples as follows:

1. The sample was mounted in the cell fitted with CaF2 disks (which are IR transparent up to

1000 cm−1) followed by evacuation at room temperature and at 1.33x10-8 bar for 30 min.

2. The sample was reduced at different temperature in hydrogen flow (60 cm3/min, 1 bar,

10°C/min of heating rate) up to 800°C.

3. Then the system was cooled back to room temperature, flushed with N2 and then evacuated

at room temperature and 1.33x10-8 bar for 30 min.

4. Hence, the CO adsorption was carried out at 4x10-2 bar and in temperature range 20-100°C

for one hour. All the DRIFT spectra (MIR) were preserved following evacuation of the cell at

room temperature and at about 1.33x10-8 bar for 30 min.

On the other hand, since no appreciable CO adsorption bands appeared below 100°C only the

spectra of CO adsorption at 100°C are subtracted and deconvoluted. The IR assignments of

metal-CO bonds are shown in Table 8.

3.5.2.1. CO chemisorption on Mo/Al2O3 The spectra were presented after subtracting the spectra of the gas phase and the sample prior

to the adsorption. The spectra of CO adsorbed at 100°C on 20 wt% Mo/Al2O3 reduced at

600°C, 700°C and 800°C are depicted in Fig. 31.

Four different υ(CO) bands can be observed in the spectrum of Mo/Al2O3 reduced at 600°C.

The two bands at νs1388 and νas1497 cm-1 are originated from monodentate carbonate from

CO adsorbed on molybdena, while the other two bands are due to CO species adsorbed on

metal cations. Thus, the CO adsorption band at 2048 cm-1 can be assigned as CO bonded to

Mo4+ and/or Mo3+ ions. Furthermore, the band at 2194 cm-1 is assigned to σ-bonded CO

adsorbed on octahedrally coordinated Al3+ sites which are present on the surface more than

tetrahedral one, having stronger Lewis acidity, since the lower wavenumber has been

associated with increasing acid strength [20-28]. However, this band became more prominent

and shifted to 2197 cm-1 after reduction of Mo/Al2O3 at 800°C due to the increase in the CO

stretching frequency above of the free CO gas molecule frequency (2143 cm-1) with strong

Lewis acid character.

The spectrum of CO adsorbed on Mo/Al2O3 reduced at 700°C exhibits new bands: (i) two

bands at νs1351 and νas1578 cm-1 due to symmetric and asymmetric vibrations of formate

70

Page 71: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

species that is further confirmed by the broad band at about 2685 cm-1 attributed to ν(CH)

stretching, (ii) two bands at νs1485 and νas1550 cm-1 are assigned to vibrations of

carboxylates, (iii) two bands at νs1385 and νas1447 cm-1 can be assigned to vibrations of

monodentate species. In addition, the bands appearing at νas1415, νs1715 and νs1850 cm-1

indicate the characteristics of free carbonate species on the surface. However, the band at

2048 cm-1 appears with a shoulder at 1994 cm-1 that can be assigned to bridged Mo0-CO

indicating the presence of small amounts of metallic Mo0. These two bands shifted to 2050

and 2002 cm-1 on the catalyst reduced at 800°C.

It can be seen that further reduction up to 800°C enhanced the adsorption of CO to form

bridged carbonates (νas1230 and νs1765 cm-1) and bicarbonates (νs1447 and νas1600 cm-1).

The spectrum also reveals bands located at νas1281, νs1653 cm-1 belonging to bidentates,

besides the bands at νs1385, νas1560 cm-1 assigned to formate. On the other hand, the bands

protruding at 2025 and 2050 cm-1 are created upon the interaction between Mo ions and

adsorbed CO molecule. More specifically, the band observed at 2025 cm-1 with a shoulder at

2002 cm-1 are very likely associated with the terminally configured CO σ-bonded to metallic

Mo(0) species. Whereas the band at 2048 cm-1 that was first seen after reduction at 600°C

shifted to 2050 cm-1 due to the presence of lower molybdenum valence states as the result of

the higher reduction at 800°C when the average molybdenum oxidation number was

estimated to be 1,6 [3-9]. This band can be assigned to CO π-bonding to Mo2+ and/or Mo1+ in

harmony with some reported results in the literature [12-15].

Another interesting point of CO adsorption behaviour on Mo/Al2O3 surface reduced at 800°C

is that the band at 2197 cm-1 is more prominent and broader than the band at 2194 cm-1. This

band is probably formed by overlapping the band corresponding to octahedral alumina sites

with a band generated by interaction of CO with molybdena hydroxyls (Mo-OH bonded).

Consequently, the lower charge and higher reduction of Mo lead to a weaker σ-bond and

enhance a π-type electron backdonation to form Mo-CO complexes (carbonates). This

contributes to a CO stretching frequency below the CO gas frequency (2143 cm-1).

71

Page 72: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 31. In situ DRIFT spectra of CO adsorbed on Mo/Al2O3 reduced at different temperature 3.5.2.2. CO chemisorption on CeO2, Mo/CeO2 and Ce-Mo/Al2O3 reduced at 800°C

On the basis of the results that can be found extensively in the literature regarding the

reduction of Mo, Ce and Sn [3-12, 101-106, 181-186] and taking into account the XRD and

TG-DTA results and those obtained upon CO adsorption on 20 wt% Mo/Al2O3 reduced at

different temperatures. It was decided to perform CO adsorption and CH4 reaction on the

catalysts only reduced at 800°C.

In situ DRIFT spectra of CO adsorption on Ce-Mo/Al2O3, Mo/CeO2 and CeO2 catalysts

reduced at 800°C are depicted in Fig. 32. The DRIFT spectra exhibit reactivity patterns

initiating from adsorbed CO and various types of carbonate on reduced catalysts.

The spectrum of Ce-Mo/Al2O3 exhibits an indicative band at 2198 cm-1 of CO adsorbed on

coordinatively unsaturated Al3+ sites (Aloct3+–CO), whereas a band at 2170 cm-1 is attributed

to CO linearly bonded to Ce4+ cations that is observed in all Ce-containing catalysts. On the

basis of spectral investigations of various carbonate compounds, which can be found

elsewhere [175-178], one can suppose that the surface of Ce-Mo/Al2O3 is covered with

various types of carbonate: monodentate (νs1420 and νas1540 cm-1), bidentate carbonate

(νas1320 and νs1680 cm-1) and bicarbonate (νs1460, νas1630 cm-1 and 1220 cm-1 γOH) with the

absence of Ce3+–CO band (2150 cm-1) in this spectrum.

72

Page 73: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

The DRIFT spectra recorded on reduced CeO2 and Mo/CeO2 following CO adsorption show a

band at 2150 cm-1 assigned to CO linearly bonded to Ce3+ together with the appearance of

Ce4+-CO besides the band at 2025 cm-1 due to CO adsorbed to metallic Mo0 as mentioned

previously. Nevertheless, the absence of other Mo-CO bands may imply that Mo is almost

totally reduced and/or Mo-CO bonds were very weak and disappeared under vacuum or

formed carbonates via π-type electron backdonation [176-179].

One can notice that the reduced surface of CeO2 and Mo/CeO2 after CO chemisorption

exposes clearly monodentate carbonate (νs1430 and νas1570 cm-1), bidentate (νas1340, νs1680

and νas1320, νs1690 cm-1), bicarbonate (1220, 1230 cm-1 γOH, νs1460, νs1490 and

νas1630 cm-1), and bridged carbonate (νas1285 and νs1750 cm-1).

Of particular interest, the CO adsorption on Mo/CeO2 shifts the corresponding peak positions

of carbonates to higher wavenumbers implying the increase in CO surface coverage and

different distribution of electrons comparatively.

Fig. 32. In situ DRIFT spectra of CO adsorbed on CeO2, Mo/CeO2 and Ce-Mo/Al2O3 catalysts reduced at 800°C

73

Page 74: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.2.3. CO chemisorption on SnO2, Mo/SnO2 and Sn-Mo/Al2O3 reduced at 800°C

The spectra of CO adsorbed on the samples are depicted in Fig. 33. The spectrum of SnO2

exhibits bands assigned to CO-Sn4+ (2237 cm-1) and CO-Sn2+ (2145 cm-1) entities that also

appear in the spectrum of Mo/SnO2, besides the bands at νas1257 and νs1597 cm-1 belonging

to bidentate species, while the bands at νs1387 and νas1480 cm-1 are likely arising from

monodentate species. The spectrum of Mo/SnO2 displays band positions at νas1247 and

νs1775 cm-1 with 1094 cm-1 γ(COO)– due to the formation of bridged carbonate, whereas the

bands at νs1445 and νas1585 cm-1 with 1197 cm-1 γ(OH) can be assigned to bicarbonate

species. In addition, the peaks appearing at νas1507 and νs1889 cm-1 indicate the presence of

free carbonates on the surface. A band located at 1994 cm-1 is related to bridged CO adsorbed

on metallic Mo(0) species (Mo(0)-CO) that also observed in the spectrum of Sn-Mo/Al2O3,

while the band with low intensity at 2089 cm-1 can be assigned to Mo4+CO and/or Mo3+CO.

On the other hand, the spectrum of Sn-Mo/Al2O3 shows bands at νs1432 and νas1623 cm-1

with 3634 cm-1 ν(OH) and 1227 cm-1 γ(OH) ascribed to bicarbonate species. Furthermore, the

peak located initially at 2057 cm-1 can be assigned to Mo2+CO and/or Mo1+CO sites.

Additionally, the new band at 2210 cm-1 is ascribed to Aloct3+–CO sites whereas the band

corresponding to CO-Sn4+ shifts to 2244 cm-1. However, it is worth mentioning that CO-Sn2+

band (2145 cm-1) is absent in Sn-Mo/Al2O3 spectrum. One may propose that Sn2+ ions (SnO)

are located in the internal surface of Mo/Al2O3 particles as separate phase and/or the CO-Sn2+

bond was weak so as to forming CO complexes (carbonates).

74

Page 75: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 33. In situ DRIFT spectra of CO adsorbed on SnO2, Mo/SnO2 and Sn-Mo/Al2O3

catalysts reduced at 800°C

Table 8. Assignments of IR bands observed upon CO chemisorption on metal cations

Assignment Wavenumber (cm-1) Reference Mo0 1994, 2002, 2025 188, 189

Mo3+/4+ 2048, 2089 12-15, 45, 117, 188, 189 Mo1+/2+ 2050, 2057 12-15, 45, 117, 188, 189 Ce3+/4+ 2150, 2170 80, 180-187

+3octAl 2194, 2197, 2198, 2210 20-28

Sn2+/4+ 2145, 2237, 2244 91-99

75

Page 76: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.3. In situ DRIFT results on methane transformation in absence of oxygen The procedure of nonoxidative CH4 reaction on 50 mg of the catalysts was as follows:

1. The DRIFT reactor cell was attached to the glass gas handling/vacuum system.

2. Reducing in H2 flow (60 cm3/min, 1 bar, 10°C/min of heating rate) up to 800°C.

3. The system was cooled back to room temperature and evacuated at about 1.33x10-8

bar for 30 min.

4. Recording DRIFT spectra

5. Once, introduction of CH4 (110 cm3) into the reaction space (the whole volume

including the DRIFT reactor cell and gas-circulation system was 387.3 cm3) and the

reaction was carried out for 1 hr at 700°C and 4x10-2 bar.

6. CH4 gas circulation with the aid of a glass gas-circulation pump (built in the system

with max. speed of 200 cm3/min). Helium was added to improve the efficiency of the

gas circulation.

7. Recording DRIFT spectra every 20 min.

The interaction of methane with the surface of the catalysts reduced in hydrogen flow at

800°C has been studied for 1 hr in order to reveal surface intermediate species that allow a

better interpretation of the reaction pathway. The same amount of CH4 was introduced into

the reaction space during the experiments. Due to the limitations posed by CaF2 disks, only

bands appearing at wave number > 1000cm-1 can be detected in this experiment.

The samples were compared to determine precisely what kinds of adspecies participate

efficiently to carbon storage during CH4 transformation. Three series of temperature

programmed and time-dependent DRIFT spectra have been recorded under reaction

conditions on each of the following samples: 20 wt% Mo/Al2O3, 12 wt% Mo/CeO2, 5 wt%

Ce-20 wt% Mo/Al2O3, 4 wt% Mo/SnO2, 5 wt% Sn-20 wt% Mo/Al2O3.

76

Page 77: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.3.1. Methane transformation on Mo/Al2O3 reduced at 800°C Fig. 34 shows the time-dependent DRIFT spectra of Mo/Al2O3 surface species formed after

exposure to methane. At 3013 cm-1 a discrete peak is observed due to the CH4 gas phase. The

bands at 3113 and 2920 cm-1 are assigned to νasCH and νsCH stretching, while the bands at

1355 and 1302 cm-1 attributed to asymmetric and symmetric deformation vibration of (CH)

(δasCH and δsCH) in CH4 gas. Additionally, the spectra showed two bands centred at νs1385

and νas1570 cm-1 due to symmetric and asymmetric vibration of surface formate that was

further emphasized by the presence of the band at 2730 cm-1 assigned to ν(CH) stretching.

On the other hand, the spectra exhibit bands at 2110 and 2190 cm-1 that are likely attributed to

roto-vibrational frequencies of CO gas phase resulting from decomposition of formate species

and/or from a reforming reaction between the carbonaceous deposits generated by CH4

decomposition and oxygen vacancies on the surface. Furthermore, the band at 1265 cm-1 is

likely assigned to twisting vibration mode of δe(-CH3 methyl), besides the band at 1113 cm-1

pertaining to γ(OH). More specifically, the intensity of the peaks assigned to CH4 gas phase

decreased slightly as the reaction proceeds implying that methane conversion is low.

Fig. 34. In situ DRIFT spectra of CH4 reaction on Mo/Al2O3 reduced at 800°C

77

Page 78: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.3.2. Methane transformation on Ce-Mo/Al2O3 reduced at 800°C Fig. 35 shows the DRIFT spectra of reduced Ce-Mo/Al2O3 during the interaction with

methane, respectively. In addition to the bands corresponding to CH4 gas phase (3013 cm-1)

and bands assigned to different vibrational modes of (CH) (3113, 2927, 1350 and 1302 cm-1)

as mentioned above, the spectra also show the presence of a small amount of CO2 (band at

2310 cm-1) in the first 20 min of CH4 reaction. In contrast, after 40 min of the reaction the

spectra exhibit two bands at νs1511 and νas1655 cm-1 originating from symmetric and

asymmetric vibration modes of carbonate species besides the bands at 2310 and 2355 cm-1

belonging to CO2 gas phase with higher intensity after 60 min of the reaction.

These carbonates appear in the form of bicarbonates after 60 min, which is confirmed by the

presence of the peaks at 1240 γ(OH) and 3737 cm-1 ν(OH), whereas the bands at 2110 and

2180 cm-1 are very likely due to CO gas phase, while the band at 1095 cm-1 liberated from

deformation vibration of γ(COO)¯ being adsorbed on the surface. However, the thermal

stability of bicarbonates is reportedly lower than that of other types of carbonates, their

decomposition produces CO2 [178-184].

Fig. 35. In situ DRIFT spectra of CH4 reaction on Ce-Mo/Al2O3 reduced at 800°C

78

Page 79: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.3.3. Methane transformation on Mo/CeO2 reduced at 800°C The time-dependent DRIFT spectra of CH4 decomposition over reduced Mo/CeO2 are

depicted in Fig. 36. The spectra exhibit discrete bands at 3113, 3013, 2920, 1355 and 1302

cm-1 corresponding to different vibration modes of CH4 gas phase as mentioned previously.

The infrared spectrum after 40 min of reaction exhibits peaks at νs1480 and νas1553 cm−1,

with a band at 2737 cm−1 ν(CH) due to the formate species, respectively. While those

assigned to bicarbonates are observed at νs1520, νas1657 cm-1 that was further emphasized by

OH peaks at 3737 cm-1 ν(OH) and 1120 cm-1 γ(OH). The carbonate bands are essentially in

the same positions up to the end of the experiment. The spectra exhibit IR features near 2115

and 2192 cm−1 (CO gas phase), besides the bands at 2312 and 2353 cm-1 (CO2 gas phase)

indicating the presence of CO and CO2 during CH4 decomposition. These bands become more

prominent after 60 min implying the high conversion of CH4 that was confirmed by the

gradual decrease of the band intensity corresponding to the CH4 gas phase (3013 cm-1).

However, it is possible that formate species are adsorbed on Mo while bicarbonates on Ce but

the beneficial effects might have occurred either because the cooperation between the partial

CeOx and Mo generated sites with higher activity, and/or because the oxidative properties of

CeO2 increased the dissociation of CH4.

Fig. 36. In situ DRIFT spectra of CH4 reaction on Mo/CeO2 reduced at 800°C

79

Page 80: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.3.4. Methane transformation on Sn-Mo/Al2O3 reduced at 800°C The time-dependent DRIFT spectra of Sn-Mo/Al2O3 during CH4 reaction are shown in Fig.

37. Several bands are due to fundamentals, overtones vibrations of CH4 gas phase (3113,

3013, 2930, 1350 and 1303 cm-1) as mentioned in the previous sections. It can be seen that

CO and CO2 (2100, 2180 and 2360 cm-1 bands) appear after 40 min. These bands are more

prominent after 60 min of CH4 reaction. Nevertheless, after 60 min the spectra present bands

at νs1370, νas1587 and 2860 cm-1 ν(CH) indicative of formate species, while the band at 1060

cm-1 is liberated from deformation vibration of γ(COO)¯, whereas the bands at νs1470, νas1620

and 3673 cm-1 ν(OH) are practically arising from bicarbonate species. However, the presence

of the former species can be evaluated in analogy to what has been observed for CH4

decomposition on Mo/Al2O3 and Ce-Mo/Al2O3. It is worth pointing out that the intensity of

the peaks corresponding to CH4 gas phase decreased slightly till the end of the experiment

implying that the CH4 conversion is low.

Fig. 37. In situ DRIFT spectra of CH4 reaction on Sn-Mo/Al2O3 reduced at 800°C

80

Page 81: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.3.5. Methane transformation on Mo/SnO2 reduced at 800°C Fig. 38 shows the time-dependent DRIFT spectra of surface species formed upon contact of

Mo/SnO2 with CH4. In addition to the bands belonging to CH4 gas phase (3113, 3013, 2940,

1350 and 1303 cm-1) the DRIFT spectra exhibit a discrete band at 1680 cm-1 with the band at

1190 cm-1 γ(H-CO) are typical for ν(C=O) stretching in coordinatively bonded formaldehyde

on the Lewis acid sites. Accordingly, this suggests that the first detectable adsorbed species to

be adsorbed formaldehyde intermediate accompanied with the presence of a large amount of

CO2 (2310 and 2360 cm-1). However, formaldehyde completely disappeared after 40 min of

CH4 reaction while the intensity of CO2 band gradually increased till the end of the reaction.

The spectra also show the liberation of small amounts of CO (2110 and 2187 cm-1) after 40

min. Furthermore, the bands located at νs1540, νas1620 and 3651 cm-1 ν(OH) are associated

with IR characteristic features of adsorbed bicarbonate species.

Since as can be seen, the intensity of CH4 gas phase peaks gradually decreased even after 40

min and completely disappeared by the end of the experiment demonstrating that the CH4

conversion is very high and almost complete CH4 oxidation was achieved.

Fig. 38. In situ DRIFT spectra of CH4 reaction on Mo/SnO2 reduced at 800°C

81

Page 82: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

3.5.3.6. DRIFT spectra after methane reaction under vacuum As can be seen from the spectra obtained after the cell evacuation for 20 min at room

temperature (r.t.) and 1.33x10-8 bar, some formate and carbonate species are still adsorbed on

the surfaces (Figs. 39 and 40). Therefore, the spectrum of Mo/Al2O3 displays bands at νs1455

and 2857 cm-1 ν(CH) assigned to formates, while the band at 1626 cm-1 is due to the

deformation vibration of water (H-OH) with the band appearing at 1116 cm-1 is assigned to

γ(OH) being in the bulk of the catalyst, which shifted to higher wavenumber (1125 cm-1) in

the spectra of Ce-Mo/Al2O3 and Mo/CeO2 (Fig. 39).

In addition, the spectrum of Ce-Mo/Al2O3 exhibits new bands at νs1437 cm-1 assigned to

carbonate, whereas the band at 2171 cm-1 is very likely assigned to Ce4+-CO that also

appeared in the spectrum of Mo/CeO2. In contrast, the spectrum of Mo/CeO2 also reveals

bands located at 1437, 1650, 1734, 1839, 2873 and 3747 cm-1 are closer to the bands positions

for formates (νs1437, νas1650, and ν(CH) 2873 cm-1) and bridged carbonates (νs1734 and

νs1839 cm-1) being adsorbed on different sites (Mo and Ce).

The spectrum of Sn-Mo/Al2O3 exhibits bands at νs1453, νas1587 and 2789 cm-1 ν(CH)

indicative of formate species whereas the spectrum of Mo/SnO2 shows bands at νas1267 and

νs1878 cm-1 assigned to bridged carbonates. Moreover, the band observed at 1053 cm-1 is due

to the deformation vibration of γ(COO)-, while the band at 1127 cm-1 is assigned to γ(OH),

beside the band at 1633 cm-1 observed in the two spectrums is due to the deformation

vibration of water (H-OH) (Fig. 40).

Of particular interest, the broad hydroxyl band reappears in all spectra with lower intensity

comparing with the spectra of the calcined catalysts besides new OH groups (1116, 1125,

1127 and 3747 cm-1) are still present even after evacuation, that may be explained by the

presence of strongly bound hydroxyl groups encapsulated in the oxide bulk (Figs. 39 and 40).

82

Page 83: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Fig. 39. In situ DRIFT spectra of Mo/Al2O3, Mo/CeO2 and Ce-Mo/Al2O3

after CH4 reaction under vacuum

Fig. 40. In situ DRIFT spectra of Mo/SnO2 and Sn-Mo/Al2O3

after CH4 reaction under vacuum

83

Page 84: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

4. DISCUSSION 4.1. Surface texturing Inspecting of the data compiled in the Table 4 reveals the mesoporosity nature of these

materials. In addition, the computed values of mean pore radius of the different samples,

which are comparable to each other show the mesoporosity and decreasing the total pore

volume with increasing the pore diameter of all samples comparatively (Figs. 7 and 9) may

indicate that their pores are narrow as well as they are deep [70-72].

Therefore, it is worth mentioning that the significant decrease of the surface area of the

samples is rather due to blocking pores by the oxides added (MoO3, SnO2, CeO2). On the

other hand, this may also demonstrate that molybdena is well-dispersed on the supports.

4.2. X-ray diffraction The presence of Al2O3 and the Mo dispersion obtained on Al2O3 may suppress stronger

interaction between Mo and Ce and may contribute to particles agglomeration not allowing

the material to maintain its dispersion. This was confirmed by the presence of only one

crystalline structure of cerium-molybdate (Ce2Mo3O12) in addition to the decrease in the

intensity of XRD lines assigned to Al2(MoO4)3 and MoO3 in the Ce-Mo/Al2O3 sample, while

three crystalline structures of cerium-molybdate were identified in Mo/CeO2 sample (Fig. 11).

The substitution of some Ce4+ ions by Mo6+ favours the formation of defects in the ceria–

molybdena lattice that induce a distortion of the oxygen sub-lattice. This is because the ionic

radius of Mo6+ (0.62 Å) is too small compared to that of Ce4+ (1.01 Å) to accommodate all the

oxygen. This distortion increases with the Mo content and is responsible for the progressive

change of the symmetry of solid solutions. It is not possible to identify clearly the

composition at which phase separation occurs, as this depends on several factors including

sample preparation and treatment, although generally ceria-rich composition crystallizes

easily with a cubic symmetry while intermediate and molybdena-rich compositions prefer an

octahedral phase. From this picture, it is expected that solid solutions having the highest

concentration of Mo will show the best redox behaviour. However, this is not completely true,

since a high MoO3 content will reduce the quantity of active redox elements. Therefore, a

detailed balance between structural defects and Ce or Mo content must be reached for

optimum dispersions that in turn affect the structure of these systems [65, 76-79].

Combining the results obtained from XRD and BET investigations, it can be anticipated that

the introduction of cerium ion promotes aggregation of the Mo particles mainly when using as

84

Page 85: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

a support with high Mo loading as emphasized by diminishing in the surface area and total

pore volume besides the formation of different molecular formulae between Mo and Ce.

Thus, the interaction between molybdena and ceria probably due to charge effects and may be

based on the fact that CeO2 possesses a strong basic property. This indeed affects the particle

size of CeO2 and MoO3 crystallites. More specifically, in principle decreasing the particle size

of the two interacting ions helps their diffusion and thus their mutual interaction that is highly

affected by the mode of the preparation that in turn affects the thermal stability and

crystallinity of the produced compounds.

In conformity, IR and XRD studies by M. M. Mohamed et al. have emphasized that the CeO2

substrate promoted the aggregation of Mo ions up to 8 wt% loading when impregnation

method was adopted for preparation [80].

The use of SnO2 as promoter in Mo/Al2O3 leads to surface structure definitely different from

that of Mo/SnO2 (Fig. 13). We may tentatively attribute the discrepancy between the two

solids to sintering through condensation of the external, octahedral molybdenum layers

between different particles during the calcination of Sn-Mo/Al2O3 sample. While the

superficial molybdate ions could inhibit sintering of the SnO2 support during the calcination

of Mo/SnO2. However, stannous and stannic oxide coexist frequently either due to an oxygen

loss associated with the reduction of SnO2, or to the oxidation of SnO. Two main mechanisms

are responsible for sensing: (1) The bulk diffusion of oxygen from outside into the oxide,

compensating an original deficiency of oxygen, which is typical of most oxides. (2) The low

temperature chemisorptions of environmental gases on the surface of multiple grains,

charging the surface state and charge distribution inside the grain. The oxygen characteristics

at the grain surfaces of the porous SnO2 materials are strongly dependent on the surrounding

gas atmosphere [90-99].

Hence, it can be anticipated that no linkages were observed between Mo and Sn ions that

presumably occur at temperature higher than 700°C. Additionally, the high Mo loading leads

to the decrease of SnO2 crystal size (the decreased peaks intensity assigned to SnO2 in the two

catalysts) probably due to polarization effects and the presence of molybdate species strongly

affects the growth of SnO2 crystals in line with some results in the literature indicating that

Mo disturbs the SnO2 crystallization mainly at higher Mo loading [100-107].

In conformity, F. Goncalves and co-workers showed that the presence of the cassiterite phase,

the crystal size and surface area strongly depends on the molybdenum loading. It was shown

that the presence of molybdenum oxides inhibits crystal growth of the SnO2 support, leading

to material with a high specific area and high dispersion of molybdenum and tin at high

85

Page 86: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

molybdenum loading when co-precipitation and impregnation methods were adopted. They

showed a pronounced influence of the preparation method on the crystallinity and catalytic

properties of Mo-Sn system [101].

N. G. Valente and co-workers have shown that no phase transition takes place at temperatures

up to 700°C of MoO3 and SnO2 mixed mechanically. They claimed that there are synergetic

cooperations between SnO2 and MoO3 to explain the catalytic behaviour in the oxidation of

methanol. So that MoO3 crystals would be oriented on the surface of SnO2 creating

catalytically active sites that would not be present in the isolated phases [105].

Anyhow, the fact that doping of transition metal oxides to SnO2 dramatically influences the

defect and the sintering behaviour of tin oxide. Nevertheless, MoO3 with SnO2 exhibits many

interesting features. This is the reason for the numerous studies of their physical and chemical

properties [101-109].

4.3. Thermal Analysis Several authors have studied the mechanisms of the thermal decomposition of ammonium

heptamolybdate (AHMT). The decomposition process took place via the formation of

different intermediate compounds, which decompose readily yielding solid MoO3 (Eqs. 24-

26). However, similar results were obtained by A. Said and S. A. Halawy [160] who found the

three mass losses corresponding to the thermal decomposition of pure AHMT in nitrogen flow

that were slightly different (7.2, 4.3 and 6.8%) from the results obtained in the present work,

respectively. Details of the mechanism of the thermal decomposition of ammonium

heptamolybdate yielding MoO3 have been given elsewhere [80, 160-162].

Some authors studied SnO2 for its capacity to dissolve cations such as Sb and Mo. Okamoto et

al. suggested [84] through quantitative XRD and catalytic study that SnO2 could dissolve

Mo6+ in spite of charge unbalance leading to high catalytic activity of the Mo/SnO2 system in

the oxidative dehydrogenation of sec-butyric alcohol. While other results did not show

significant dissolution of Mo6+ in the SnO2 structure even at high Mo concentrations after

either calcination or reduction up to 700°C [102-108].

Anyhow, it is worth pointing out that the dissolution of some molybdena in SnO2 was

observed by the end of this experiment and H2SO4 was used to clean the sample cup because

of the adhesive form of the sample. This confirms the hypothesis of significant molybdenum

dissolution in tin oxide above 750°C.

86

Page 87: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

However, probable transformation of the Mo-Sn oxide system needs further studies as well as

the formation of new phases (e.g. formation of a new oxide phase associating Mo with Sn)

during the thermal treatment of this system above 700°C.

Comparing the TG-DTA results, it can be anticipated that Mo/Al2O3 is the thermally most

stable material in the temperature range used in the experiment (900°C). Whereas

Ce-Mo/Al2O3, Mo/CeO2 and Mo/SnO2 samples undergo morphological and textural

modifications throughout the thermal behaviour above 700°C resulting in lattice defects

(Schottky and Frenkel defects) which motivate the mobility of Mo, Ce and Sn ions and thus

enhance the possibility of interaction between them (Figs. 18-22).

However, the fact that the five samples showed varying net mass loss implying that they have

not comparable molecular formulae. Accordingly, although the type of active sites is more or

less similar in all samples, different reaction pathways may result in different activities.

4.4. Electron Spin Resonance (ESR) The EPR spectra of the samples showed the presence of Mo5+ and Ce3+ paramagnetic centers

after CO interaction at both temperatures (r.t. and -196°C). One may notice that the two types

of EPR spectra (after CO interaction at both r.t. and -196°C) are quite similar for all samples

but more prominent and in some cases (e.g. for Mo/Al2O3) broader at -196°C (Figs. 23-26).

According to the literature, the increase of the line width is due to the clusterisation of the

paramagnetic centers in the sample. On the other hand, if any oxygen inside the sample tube,

it will condense at the liquid nitrogen temperature. Therefore, the condensed oxygen adsorbs

on the surface of the solid sample. Consequently, the interaction between the oxygen dipoles

adsorbed on the surface may lead to the broadening of the EPR lines [166-174].

In conformity, O. Cozar et al. and others found the six lines of Mo in their ESR study of

molybdenum-lead-phosphate glasses. Furthermore, the existence of three types of symmetry

sites for Mo5+ ions has been shown. The dipole–dipole and super exchange coupled Mo5+ ions

appeared and their number increased with the Mo content in samples heat-treated in nitrogen

atmosphere [166-169].

B. Kamp et al. studied by ESR spectroscopy the oxygen defects on SnO2 in oxygen

atmosphere at elevated temperatures and pressures and it was found that oxygen defects on

SnO2 occur above 700ºC [170, 171].

F. Morazzoni et al. in either SnO2 or Pt/SnO2 samples observed no paramagnetic species

corresponding to SnO2 after treatment with CO/Ar. The EPR spectra showed only a

symmetrical resonance line that was attributed to an oxygen defect (Eqs. 28-29) [172].

87

Page 88: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

4.5. In situ DRIFT spectroscopy 4.5.1. DRIFT spectra of the calcined samples under vacuum Although the band positions of such surface hydroxyl groups can be distinguishable by

recording and deconvoluting IR spectra in 3000-3800 cm-1 region upon adsorption of

molecules such as pyridine or upon dehydroxylation at elevated temperatures. Details of OH

yield on alumina and ceria have been given elsewhere [18-31, 180-187].

For instance, concerning OH groups on CeO2, M. I. Zaki and co-workers identified and

differentiated acidic and basic OH’s at 3583, 3621, 3652 and 3684 cm-1 involving in pyridine

adsorption on CeO2 [180].

The comparison between the spectra of 8 wt% Mo/Al2O3 and 20 wt% Mo/Al2O3 (Figs. 28 and

29) permits to infer that free MoO3 clusters and polymeric molybdate species can only be

formed at higher Mo loading in line with some results in the literature [12-17, 32-35].

As the result of the proposed interaction between molybdena and ceria on the basis of the

strong basicity possessed by CeO2, this interaction is further emphasized by forming

Mo-O-Ce linkages that were represented by the bands at 630 cm-1 νs(Mo-O-Ce) and 875 cm-1

νas(Mo-O-Ce) in addition to the bands at 995 and 1035 cm-1 associated with the formation of

coupled νs(O=Mo=O) bonds (Fig. 28). Accordingly, ceria can attribute to the increase of

polymerized surface Mo species. Nevertheless, the expected competition between Ce and Mo

with alumina hydroxyl groups may suppress stronger interactions between them (decreased

intensity of OH region in Ce-Mo/Al2O3 spectrum).

Furthermore, no bands arising from Mo–O–Sn vibrations were observed in both spectra of

Mo/SnO2 and Sn-Mo/Al2O3 samples (in line with XRD results). This means that the

molybdenum oxide may spread over the tin oxide readily after calcination at 600°C (Fig. 30).

According to literature sources tin can form –Sn(OH)2–O–Sn(OH)2–O– type polymer chains

in which the OH groups remain stable even at elevated temperatures [82-90]. The

concentration of surface hydroxide of tin oxide may be consistent with the formation of solid

acidity, because its concentration was found to be high up to 700°C of the calcination

temperature of tin oxide. Several bands in 3300-3800 cm-1 region due to fundamentals of OH

yield on SnO2 have been given elsewhere [91-99].

88

Page 89: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

4.5.2. CO chemisorption It is reasonable to suggest that the CO adsorption on Mo/Al2O3 reduced at different

temperature occurs now mainly on small crystallites of MoO3. However, the presence of

formate and carboxylate indicates that the adsorbed CO on Mo reacts with the geminal

hydroxyls on the surface to form these species implying the high reactivity of the lattice

oxygen such as Mo=O and Mo–O–Mo. This means that CO adsorption involves the lattice

oxygen of Mo activated at different sites of the surface throughout the reduction pretreatment.

Nevertheless, the formation of formate species may proceed through a surface reaction

between the weakly adsorbed CO to Mo species and a reactive H atom adsorbed on the

surface upon H2 reduction via (LH) mechanism leading to the formation (with the

participation of lattice oxygen of MoO3 and/or hydroxyls) of various carbonate species which

become more notable on the surface of Mo/Al2O3 reduced at 800°C (Fig. 31). Accordingly,

this supports the following reaction steps:

Scheme 4.

Hence, one may suggest that the presence of MoO3 clusters makes the surface eligible to

liberate various carbonates upon CO adsorption at 100°C. It can be anticipated that there is an

approximately linear correlation between the increase of the extent of reduction and the

increasing integrated absorbance of CO adsorbed on 20 wt% Mo/Al2O3 catalyst (Fig. 31).

On the other hand, the reduction of the catalysts to produce lower oxidation states of Mo and

more Ce3+ species activated the surface to be eligible for CO chemisorption as metal-CO as

well as carbonate species. Thus, the chemisorption of CO involves oxygen in the catalysts

such oxygen could be present in Ce–O–Mo and/or Ce–O–Ce associates. The noticed gain in

intensity for the bands in conjunction with the carbonate species is formerly observed. These

species are relatively stable under vacuum at room temperature (Fig. 32).

It has been revealed that the hydrogen reduction of Ce and Mo at 800°C (when Ce and Mo are

almost totally reduced) improves the surface reactivity as well as the stability of carbonates,

89

Page 90: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

and leads to the presence of small amounts of metallic Mo species after reduction at 700°C.

One may suggest that the intervention of the cerium couple (Ce3+/Ce4+) was appreciable in

Mo/CeO2 following the reduction process and CO adsorption probably due to forming

stabilized species of ceria oxide and molybdena-cerium [54, 65, 80].

However, the absence of Ce3+–CO band (2150 cm-1) in Ce-Mo/Al2O3 spectrum (Fig. 32)

indicates that Ce ions are preferentially located in the internal surface of Mo/Al2O3 particles,

and mostly in compensating positions substituting Mo ions in Ce-Mo/Al2O3, whereas for

Mo/CeO2 the majority of Ce ions are present on the external surface as cerium-molybdate and

CeO2 moieties in line with BET and XRD results.

On the other hand, CO-Sn2+ band (2145 cm-1) is also absent in Sn-Mo/Al2O3 spectrum. One

may propose that Sn2+ ions are located in the internal surface of Mo/Al2O3 and/or the CO-Sn2+

bond was very weak so as to forming CO complexes like carbonates (Fig. 33). Consequently,

CO being provided as weakly adsorbed metal-carbonyls migrating towards the oxides through

interfacial sites to form carbonates via a π-type electron backdonation [175-179].

Some infrared spectroscopy (IR) investigations of CO adsorption on ceria have found that CO

not only adsorbed weakly in vertical orientation as on other oxides, but also interacts strongly

with the CeO2 surface, forming carbonate and/or inorganic carboxylate like complexes even at

room temperature [180-183]. The formation of various carbonate species like complexes of

CO on ceria suggests that CO is oxidized by the lattice oxygen of ceria, or in other words, the

CeO2 surface is partially reduced by CO.

Studies by Liu et al. on the adsorption of NO2 and SO2 on zirconium doped ceria found that

the oxygen sites of the Ce(111) surface are able to react with NO2 and SO2 to form surface

sulphates and nitrates at r.t. [184]. This suggests that there is the possibility of forming surface

carbonates after adsorbing CO.

On the other hand, other experimental studies have found no evidence for chemisorption of

CO on ceria or they have found that chemisorption is only observed after surface

pre-treatment. For example, through photoemission studies of CO adsorption on ultra-thin

ordered CeO2 (111) and CeO2 (110) surfaces [185-187].

Berner et al. found that CO exposure does not affect unreduced CeO2 surface [185].

Z. Yang et al. found that the adsorption of CO exists different features on (111) and (110)

surfaces of CeO2. While only weak adsorptions were found on (111) surface, both weak and

strong adsorptions exist on (110) surface [186]. Electrostatic interactions were involved in the

weak interactions, while covalent bonding was developed in the strong adsorptions.

90

Page 91: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

The geometries of the strong interaction modes for CO on Ce(110) surface all involved the

formation of carbonates on the surface. Their calculated results were in agreement with the

experimental results. These confirm experimental observations that CO not only adsorbed

weakly in a linear form (vertical orientation), but also interacts strongly with the CeO2 surface

forming considerable amounts of carbonate species even at room temperature. In their study,

only the stochiometric non-defective (110) and (111) surfaces of ceria have been examined.

Therefore, they concluded that the adsorptions of CO on the (111) and (110) surfaces have

distinctly different properties. On the (111) surface, only weak adsorption (physisorption)

modes exist, whereas, for the (110) surface, not only weak adsorption modes exist but also a

few strong adsorption modes exist that can be classified as chemisorption. In all cases, the

strong adsorption modes involved the formation of carbonate species on the surface.

All these conflicting results in the literature on whether or not surface complexes exist suggest

that the surface properties and the adsorption behaviour of CO on the ceria surfaces are

strongly related to the nature of the surfaces used in the experiments. Therefore, some

experiments have simultaneously observed both weak and strong interactions for the

adsorption of CO on cerium oxide, while others found only weak interactions. Why both

strong and weak interactions of CO on ceria occur in some cases, while only weak

interactions are observed in others remains unclear. Although some theoretical investigations

on stoichiometric bulk ceria and its surfaces have been published, there have been few

theoretical investigations on the adsorption of CO on ceria and no first-principles studies.

Thus, the mechanisms and atomic level understanding of the interaction of CO on ceria

surfaces have not yet been adequately addressed theoretically.

Considering the present results and taking into account some results concerning CO

adsorption on Mo, Ce and Al oxides. It can be anticipated that the interaction between Mo and

supports has an effect on the adsorption properties of CO, which may indicate the different

electronic effects in the catalysts. Within this context, intimate coupling of Mo and Ce ions of

different oxidation states has great facilities for electron exchange interactions. Thus, the

electron-mobile environment necessitated by redox reactions is established that has a great

share in enhancing the CO adsorption. On the other hand, the geometric effects also play an

important role for the Mo–support interactions, the variance of the reducibility may be caused

by the different locations of Mo on the supports or the different orientations of Mo

combination to the supports. The special Ce–Mo–Al interaction may be originated from both

the electronic effect that affects the CO adsorption and the geometric effect that may

contribute to the reducibility variance of MoO3.

91

Page 92: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Furthermore, different methods for sample preparation and pre-treatment will result in

surfaces with different orientations, different oxidation states (e.g. unreduced and reduced),

and varying degrees of surface contamination. Additionally, defects such as step edges,

vacancies and partial surface reconstructions allow the formation of surface carbonates upon

CO adsorption [176-187].

Accordingly, the active sites for CO adsorption in the five samples are not similar and the

relative intensities of the same band positions are different indicating a various distribution of

electrons. Moreover, the pronounced increase in absorbance can be easily interpreted as due

to a decreased concentration of free electrons and of electrons trapped in oxygen vacancies

upon CO adsorption. For instance, the creation of Sn2+ donor levels and oxygen vacancies

upon reduction and CO adsorption as well as the variations of point defects and carbonates

may be described as follows:

Upon H2 reduction:

SnO2 + H2 ↔ SnO + H2O (Eq-30)

SnO2 ↔ SnO2−x + xO (Eq-31)

O2-(lattice) +H2 ↔ VO

2- + H2O (Eq-32)

Sn4+ + VO2- ↔ Sn2+ + VO (Eq-33)

Upon CO adsorption:

CO(ads) + O2-

(lattice) ↔ (COO)– + VO (Eq-34)

The increase in electron concentration arises from:

VO ↔ VO+ + e− (Eq-35)

VO+ ↔ VO

2+ + e− (Eq-36)

Sn2+ ↔ Sn4+ + 2e− (Eq-37)

Where VO, VO+ and VO

2+ are the neutral and ionized oxygen vacancies, respectively,

according to the Kröger–Vink notation [177-179].

92

Page 93: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

These results are in agreement with the literature. For instance, Popescu et al. and others

found that CO adsorption on SnO2 even at room temperature gives rise to CO–Sn2+ and CO–

Sn4+ end-on species and to various carbonate entities [99, 106-109].

Anyhow, the formation of carbonate species throughout the contact with CO implies the

existence of reactive oxygen on the surfaces after the reduction in H2 flow at 800°C, which

creates point defects (native and foreign atoms) that can act as both donors and acceptors.

The interaction of CO gas with these surfaces led to different changes in the lattice oxygen

contents on such surface, this by its turn changes the amount of adsorbed carbonate species.

Some discrepancies concerning CO adsorption on the five samples are practically due to

several factors such as the nature of the sample, the stoichiometry of oxides, the presence of

lattice defects, the size and shape of the particles and hydroxyl groups concentration.

However, the dissolution of Mo was also observed in Mo/SnO2 sample reduced at 800ºC by

the end of this experiment (here, again H2SO4 was used for cleaning).

4.5.3. In situ DRIFT studies on methane transformation in absence of oxygen It is generally accepted that methane is mainly activated on metallic surfaces. The electron

donation from the HOMO of CH4 to the lowest unfilled molecular orbitals of metal surface

should dominate dissociative CH4 adsorption. However, the fact that the carburization of Mo

species by CH4 is thermodynamically possible, therefore CH4 can interact with Mo vacancies

according to:

Mo[ ]Mo + CH4 ↔ Mo[C]Mo + 2H2 (Eq-38)

In this state, carbon can be extracted from the site, producing CO and regenerating the

vacancy, as shown:

(Eq-39)

H)x4(CHMoeCHMoCH x_

44 −+⎯⎯ →⎯+⎯⎯ →⎯ + (Eq-40) (Eq-41)

Mo[C]Mo· · ·OCO ↔ Mo[ ]Mo + 2CO (Eq-42)

93

Page 94: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Where (♦) and (■) represent adsorption sites on Mo oxide. Thus, the formation of CO from

CH4 would not be a simple dual site reaction involving adsorbed carbon and lattice oxygen,

but involves a solid-state reaction. This mechanism would be consistent with the high

activation energy of CH4 conversion on the supported catalyst.

Accordingly, the reduction of Mo by H2 at high temperature provides reactive H atoms able to

form OH groups and oxygen vacancies on the surface, particularly those present in the bulk,

are the driving force for CH4 activation. Furthermore, hydroxyls are also formed during the

activation step due to the reaction of the lattice oxygen with hydrogen as dihydrogen arising

from methane dehydrogenation or as water arising from the initial decomposition of methane.

It comes therefore that these rather basic OH groups are required for formate formation

supporting the following reaction steps:

H2(g) + 2Mo → 2Mo-H (Eq-43)

Mo-H + O2– → Mo + OH– + e– (Eq-44)

CO + OH– ↔ HCOO – (Eq-45)

Based on these results, it can be inferred that the loss of OH groups during CH4 conversion

does not necessarily implicate these species in the reaction to produce CO as the generation of

formate species from CO would also account for the consumption of hydroxyls (Fig. 34).

Lin et al. carried out a comparative FTIR study on the interaction of CH4 with silica, alumina,

and HZSM-5. The results demonstrated that OH groups played a very important role in CH4

adsorption. When an interaction between the OH groups and CH4 took place, the band shift of

the OH groups varied and the strength of the interaction decreased in accordance with the

order of their acidities (Si–OH–Al > Al–OH > Si-OH). The authors considered the possibility

that CH4 is activated by interacting with a proton leading to a heterolytic cleavage of a C–H

bond of CH4 [110].

Recent studies using supported and unsupported Mo compounds indicate that interactions

with methane at temperature around 700°C lead to the formation of Mo2C, which is

considered as the active site for the formation of CH2 and CH3 fragments. These carbides may

be destroyed by reaction with air or CO2 [111-113].

However, it is well-known that alumina support has the lowest oxygen mobility and non-

reducible alumina support is unable to store carbonaceous adspecies, the present results show

that molybdenum oxide can activate methane and oxidize it into surface formates and carbon

94

Page 95: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

monoxide but methane conversion is low. Consequently, this CH4 partial oxidation involves

the interaction of methane with lattice oxygen anions and surface OH groups created by the

previous reduction step and/or upon the initial decomposition of CH4.

Iglesia and co-workers [114] claimed that selective silanation of external acid sites on

HZSM-5 by using large organosilane molecules could decrease the content of acid sites as

well as the number of MoOx species retained on the external surface, which were regarded as

key factors for coke formation during MDA. On samples prepared using silica-modified

HZSM-5, acid sites, MoOx precursors, and active MoCx species formed during the CH4

reaction at 677°C were found to predominately reside within the zeolite channels, where

spatial constraints could inhibit the bimolecular chain-growth pathways. Consequently, the

selectivity of hydrocarbons on a 4% Mo/silica-modified HZSM-5 increased by about 30% in

comparison with that on a 4% Mo/HZSM-5.

On ceria containing catalysts, to explain the fractional presence of CO2 and CO a mechanism

can be proposed (Figs. 35 and 36), as well as taking into account the carbon exchange

between CH4 and the surface. Surface reactions describe the proposed mechanism:

(Eq-46)

Meanwhile, the CO2 could adsorb on the carbon filled-site from the reaction above, extract

the carbon to produce CO, then the carbon is removed from the active surface replenishing a

vacancy according to the reverse Boudouard reaction:

(Eq-47)

Moreover, ceria could simultaneously undergo a redox reaction in the presence of CO2:

COCeO2COOCe 2232 +↔+ (Eq-48)

(Eq-49)

Where (■) represents adsorption sites on ceria, on the other hand, the CO could also reduce

the ceria by reversing reaction, and further undergo CO disproportionation as the consequence

of the occurrence of the Boudouard equilibrium:

(Eq-50)

95

Page 96: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

However, such higher methane conversions are not necessarily related to the larger amounts

of coke and faster deactivation of the catalyst, because deactivation also critically depends on

the rate with which the coke can be removed by CO2 under reaction conditions when CO2 can

react with the CHx species as well as with CH4:

CO2 + CHx → 2CO + (x/2) H2 (Eq-51)

Therefore, in comparison with the CH4 reaction on Mo/Al2O3 the presence of CeO2

contributes to the rapid activation of CH4, thereby accelerating the carbon gasification

reaction to produce CO2. Thus, this supports the conclusion that CO2 is readily dissociated to

CO and adsorbed oxygen over reduced CeO2 by filling up the oxygen vacancies in Ce3+

species. The occurrence of Ce4+/Ce3+ redox couple generates oxygen vacancies and releases

free electrons. Free electrons transfer readily from Ce3+ to π* orbital of CO2 to activate CO2.

The increased CO2 then decomposes to CO and active surface oxygen, reacts with the CHx

species and enhances the catalytic activity of CH4 decomposition since with the aid of

Ce4+/Ce3+ redox couple, CO2 is more readily activated to release more surface oxygen, and the

rate of carbon elimination has been accelerated.

Analogous observations were made by Darujati et al. who found that Ce promotion

dramatically improves the stability of the Mo2C/γ-Al2O3 catalysts. They claimed that Ce acted

to increase the oxidation resistance of Mo2C and avoid coking by CO2 activation via the redox

reaction (Eq-48), thereby helping to prevent oxidation of Mo2C by CO2. From their study the

addition of ceria promoter to Mo2C/γ-Al2O3 catalyst appears to alter the dry methane

reforming (DMR) mechanism proposed earlier for bulk Mo2C catalysts by enhancing

relatively strong CO2 adsorption and the role of ceria was found to influence the redox

reactions on the surface as well as the activity and stability of the catalyst [115, 116].

On the other hand, although the presence of CO and CHX as well as carbonates over Rh/Al2O3

and Rh/Mo/Al2O3 catalysts has been observed by Anderson et al. during IR study on CH4

decomposition at 400°C, which also facilitated carbide formation [117].

R. Wang et al. have reported that an interaction between Rh and CeO2 was induced by high

temperature reduction, which resulted in the creation of oxygen vacancies in ceria. They

concluded that the CO2 activation in CH4/CO2 reforming should be mainly favoured by

availability of Ce4+/Ce3+ redox couple in Rh–CeO2/Al2O3 catalyst, which was rather slow

process on Rh/Al2O3 catalyst. The Ce3+ species readily promoted CO2 dissociation into CO

and surface oxygen. The higher catalytic activity and coke resistance of Rh–CeO2/Al2O3 were

96

Page 97: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

associated with the presence of the two-redox couples favoring the activation of both CH4 and

CO2. Their catalytic test results showed that the promotional effect of CeO2 on CO2

conversion was much higher than that on CH4 conversion [118].

Given that oxygen mobility is high in ceria, allowing substantial reduction of the bulk, and

given that CO2 gas is present under reaction conditions, one need not invoke surface mobility

of adsorbed CO or CO2 in order to explain the apparently high coverage of carbonate.

Formation of CO2 takes place in two distinct moments: a fraction of the CO is rapidly

oxidized to CO2 and the CO2 desorbs readily from the catalysts, while the remaining fraction

of CO/CO2 is slightly adsorbed and accumulated on the catalyst forming various carbonate

species (Eqs. 47-51).

When ceria and molybdena have been reduced at high temperature (>700°C), oxygen

vacancies, particularly those present in the bulk, seem to be the driving force for CH4

activation. Therefore, when CeO2 is exposed to H2 or CO, oxygen vacancies VO2- with two

electrons trapped have been created. Such a vacancy is a neutral entity with respect to the

surface lattice of ceria. It can easily lose an electron by spontaneous ionization becoming

singly positively charged with respect to the solid as follows [118-121]:

VO 2OOHH 222

)lattice(−+→+− (Eq-52)

−+ +−→− eOO VV 22 (Eq-53)

2CeO2 + H2 ↔ Ce2O3 + H2O (Eq-54)

Reduction of CeO2 with hydrogen is generally thought to occur via a stepwise mechanism,

first reduction of the outer most layers of Ce4+ (surface reduction), then reduction of the inner

Ce4+ layers (bulk reduction) at higher temperatures. A few mechanisms have been put forward

to account for this behaviour that comprises sequential steps of: (i) dissociation of

chemisorbed hydrogen with formation of OH groups, (ii) formation of anionic vacancies with

desorption of water by recombination of H and OH (with concomitant reduction of Ce4+ to

Ce3+) and (iii) diffusion of surface anionic vacancies into the bulk. This picture is consistent

with results obtained by Trovarelli and others upon the temperature programmed reduction

with hydrogen of high surface area CeO2 when the TPR profile showed two well-defined

peaks centred at approximately 600 and 800°C [119, 120]. Hence, the ability of ceria to be

97

Page 98: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

easily reduced to nonstoichiometric oxides is related to the properties of fluorite structured-

mixed valence oxides to deviate from stoichiometry (Fig. 12).

When considering the hydrogen conversion, the dissociation is faster and occurs at much

lower temperature (200°C) than for the alkanes (> 600°C). Hence, the cleavage of the H-H

bond is likely fast and is probably not a rate-limiting step such as the cleavage of C-H bond of

the alkanes. It may also appear on different active sites. Therefore, the conversion of

hydrogen is able to maintain its level with time, despite the production of water.

As it was pointed out previously, CH4 transformation over Mo/CeO2 can be explained by

invoking a redox mechanism with a simple redox route for CH4 oxidation, which utilizes

oxygen activated from the support in a typical reduction/oxidation mechanism (Mars Van

Krevelen type) in which the catalyst undergoes a partial reduction by methane [124-127].

Oxygen storage is therefore important because it provides an alternative route for the

oxidation of CH4. An alternative redox route involves oxygen from the support, which reacts

with methane to form adsorbed CO2 in the form of carbonates. Decomposition of carbonates

is then stimulated which provides also reoxidation of the support:

22Ox22 xHxCOeVCeOMoCeOMoCH4 ++−+++−↔−+ −− (Eq-55)

)xy(COCeOVCeOCOeVCeOCO y22Ox222Ox22 ⟨+↔++↔−+++ −−−−

−−++ (Eq-56)

However, Ce4+/Ce3+ redox couple facilitates the elimination of CHx species by partial

oxidation, resulting in higher methane conversion and lower amounts of coke. This partial

oxidation of CHx species over Mo/CeO2 will continuously result in the creation of oxygen

vacancies and Ce4+/Ce3+ redox couple. Thus, CH4 decomposition acts as the supplier of a

hydrogen pool, while Ce4+/Ce3+ redox couple promotes CO2 activation by accepting electrons

and replenishing the oxygen vacancies. This demonstrates that the reoxidized ceria can be

reduced again by methane, regenerating the oxygen vacancies and releasing free electrons, so

the creation of oxygen vacancies of ceria is a reversible process in the reaction atmosphere.

Accordingly, the presence of ceria in the catalysts as either promoter or support leads to

significantly higher methane conversion especially on Mo/CeO2 by decreasing in C storage

capacity and therefore an increase in the CO2 release upon the decomposition step. This effect

is likely to favour carbon trapping via carbonates as shown previously by DRIFT

spectroscopy. However, the mean CO concentration was not affected by the ceria or tin since

it is essentially controlled by the enthalpy of desorption from the metal phase [117].

98

Page 99: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Regarding the intimate atomic mechanism involved in oxidation of carbon, several authors

pointed out the importance of redox properties of the catalyst. That is, the effectiveness of the

catalyst can be related to its ability to deliver oxygen from the lattice to carbon reactant in a

wide temperature range. Recently, it has been reported that the use of supports based on CeO2

confers interesting properties to CH4 decomposition catalysts due to high availability of

surface oxygen and high surface reducibility. Nevertheless, analysis discrepancies in the

outcome of the results from different laboratories derive from synthesis and treatment

procedures [116-121, 145, 146].

For instance, Craciun et al. studied the CH4 + H2O reaction over Rh, Pt and Pd supported on

ceria catalysts [145]. In their study, they proposed a mechanism, which involved a surface

reaction of the adsorbed oxygen on the ceria with the dissociated methane on the surface.

Their study led to the conclusion that oxygen transfer from the ceria to the noble metals was

the rate determining step, where the participation of lattice oxygen and catalyst reducibility

have shown to improve overall performances.

Similar results over Pt/CeO2 were reported by Otsuka et al. who found that the oxidation of

CH4 by CeO2 was thermodynamically available at above 600°C. The reduction degree of

CeO2 was significantly improved from 3.5% to 17.1% in the presence of Pt after the reaction

with CH4 [146].

Concerning the CH4 dissociation on Mo/SnO2, the generation of formaldehyde intermediate

upon CH4 reaction may indicate that the Mo/SnO2 has a high concentration of Lewis acid sites

(Fig. 38). However, since the free CH2O was not detected in the gas phase that undoubtedly

appears at around 1730 cm-1 [175-179], the absence of formaldehyde may be due to its low

concentration and/or due to surface reactions by extracting lattice oxygen to form additionally

CO and CO2. These steps are represented by the following reaction steps:

CH4 + 2MO → CH2O + H2O + 2M (Eq-57)

CH2O + MO → CO + H2O + M (Eq-58)

CH2O + 2MO → CO2 + H2O + 2M (Eq-59)

Where MO represents metal oxide surface site. Moreover, CO and H2 can also be produced

via the pyrolytic decomposition of CH2O:

CH2O → CO + H2 (Eq-60)

99

Page 100: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Meanwhile, exclusion of a direct pathway from CH4 to CH2O allows CO and CO2 to be

treated as a single product (CO)x in the analysis of the kinetics of CH2O formation and

consumption. Since the interconversion of CO and CO2 was not investigated, no attempt was

made to include the effects of this process in the modelling of CH2O formation and

consumption on Mo/SnO2. The following mechanistic reaction pathway can be proposed:

2,32

41

4 COCOOMHCHOlatticeOxCHnabstractioHCH ⎯⎯⎯ →⎯ −⎯⎯⎯⎯ →⎯−⎯⎯⎯⎯ →⎯

Scheme 5.

Accordingly, it is argued that step 1, hydrogen abstraction, is stated to be favoured by lattice

oxygen possessing more negative charge. The general trends observed with formaldehyde

selectivity were explained on the basis of electrophilicity of adsorbed oxygen enhancing the

rate of step 2 with respect to step 3. Consequently, it can be stated that acid-base bifunctional

catalysts would be more effective for formaldehyde production [100-106].

Niwa and Igarashi correlated the acidity and reducibility to the catalytic behaviour of

Mo/SnO2 system in the oxidative dehydrogenation of methanol converted into formaldehyde

selectively. They found that the generation of acid sites is also strongly affected by the

calcination temperature of tin oxide that affects by its turn the formation of acid sites on the

loaded molybdenum oxide [100].

Smith and Ozkan [190, 191] studied the partial oxidation of methane to formaldehyde over

MoO3 samples exposing different relative amounts of (010) and (100) plane areas. Their

experimental characterization studies suggest that the Mo=O sites residing preferentially on

the side planes could be promoting to the formation of formaldehyde, while the bridging sites

Mo–O–Mo mainly on the basal plane were more likely to lead to complete oxidation of CH4.

S. Chempath and A. T. Bell studied the partial oxidation of CH4 at 700ºC on Mo/SiO2

catalyst. They found that CH2O is the only initial product. As the CH4 conversion increased,

the CH2O selectivity decreased and the selectivity for CO and CO2 increased [192].

Finally, the present results permit to infer that Mo/SnO2 showed the highest CH4 conversion

among the catalysts studied leading almost to complete CH4 oxidation (Fig. 38). On the other

hand, the results also demonstrate the activity of the molybdena and tin lattice oxygen and its

participation in the reaction under study conditions.

100

Page 101: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

One may suggest that the high catalytic activity of the Mo/SnO2 system towards CH4 might be

associated with the dissolution of Mo ions in the SnO2 crystals, since this dissolution was

observed at the end of the experiment and H2SO4 was used to clean the sample cup due to the

adhesive form of the sample.

The activation of methane is believed to be the key step in the conversion of methane.

According to Lunsford, methane activation occurs homolytically via the abstraction of a

hydrogen atom by oxygen anions present on oxide surface [153-155]:

[ ] ⋅+−→+−− 3CHOHCatCHOCat 4 (Eq-61)

This first hypothesis has been substantiated by the presence of methyl radicals on the surface.

The methyl radicals so formed on the surface could react with the catalyst and produce

methoxide ions, which on a subsequent reaction with water yield methanol. Further oxidation

of methanol or dehydrogenation of methoxide ion leads to the formation of formaldehyde.

However, the second hypothesis has been proposed by Sokolovskii and co-workers [193].

They suggested that methane activation proceeds heterolytically via the participation of acid-

base centers giving place to a proton detachment and the formation of a metal–methyl

compound, where the methyl results in being negatively charged:

−++−−+ +→+ 22

3422 OHMeCHCHOMe (Eq-62)

The coordinately unsaturated metal and ion paired with a strong nucleophile (O2-) ion may act

as an active center. Further oxidation of these surface methyl anions would lead to methyl

radicals, which then dimerize:

−+⋅+− ++→ eMeCHMeCH 2

32

3 (Eq-63) At the same time, the authors mentioned the possibility that some methyl radicals could

escape to the basal plane and dimerize subsequently in ethane or, in presence of additional

electron–hole pair excitations, to produce formaldehyde. In addition, the barrier to methyl

radical diffusion over O2- sites is approximately the same as the desorption energy (0.4 eV).

Unless CH2O is desorbed, additional electron–hole pair excitations should lead to further

dehydrogenation and surface reduction with the formation of CO and CO2.

101

Page 102: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

Moreover, the results of B. Irigoyen et al. [194] obtained from their theoretical and

computational study of the CH4–MoO3 chemical interaction suggest that while the hydrogen

abstraction requires less energy for each consecutive step the overall process remains

endothermic. The different sequences and sites for hydrogen abstraction from methane,

analyzed over different layers exposing molybdenum and oxygen atoms, allowed them to

conclude that the heterolytic H-abstraction is an energetically more favorable process in

comparison to the homolytic one. Their results indicate that despite an important energy

barrier being necessary for the first C–H bond activation, the overall oxidation process is

kinetically more favoured in the heterolytic mechanism.

Other studies suggest that the activation of methane occurs on super acid catalysts as well as

on organometallic complexes at low temperatures via heterolytic cleavage of the C-H bond of

CH4 [195-197]. In addition, the controlled activation of the C–H bond of methane and the

formation of the C–C bond have been extremely important and common topics in transition

metals and methane chemistry, as well as in homogeneous and heterogeneous catalysis. In

fact, transition metal centers play crucial roles in the recent development of promising

catalytic systems for C–H activation reactions in heterogeneous catalysis, and considerable

mechanistic insight has been gained [123-157, 190-197].

However, the five samples were compared to determine precisely what kinds of adspecies

participate efficiently to C storage upon CH4 transformation. This carbon storage on the

reduced Mo/Al2O3 catalyst occurs through the formation of formate species, which in turn can

either desorb into the CO gas phase or be stored on the surface that can only be formed thanks

to oxygen provided by surface lattice oxygen reacting with carbonaceous species provided by

the methane decomposition. It can therefore be deduced that the formation of formates needs

reduced molybdena containing oxygen vacancies and contributes directly to an efficient C

storage, though it requires the reaction of the surface oxygen with hydrogen atoms arising

either from hydrogen reduction (Mo–H) or from the initial methane decomposition (Eqs. 43-

45). Indeed, no straightforward relationship exists between this capacity and the hydrogen

yield since the former depends on many factors such as the basicity of the catalyst (e.g. to

accommodate carbonate species), the availability of lattice oxygen and/or defects for ensuring

the formation and spill over of the various carbon-containing adspecies. In turn, the hydrogen

yield is essentially controlled by the methane decomposition over the metal phase.

However, CH4 decomposition is associated with the hydroxyl consumption at high

temperatures (disappearance of OH region) in accordance with some results reported in the

literature indicating the involvement of OH groups in CH4 activation [110-118].

102

Page 103: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

The reappearance of OH region (with lower intensity) after CH4 reaction (Figs. 39 and 40)

confirms that some hydrogen provided by the methane dissociation is still adsorbed on the

metal oxide surfaces supporting the mechanism of CH4 dissociation mentioned previously.

Furthermore, the presence of CO and CO2 implying the existence of reactive lattice oxygen

that may be due to the weakening of the covalence of the metal-oxygen bonds and/or the

enhancement of the mobility of lattice oxygen sites in addition to Lewis acid sites induced by

hydrogen reduction (Mo with lower oxidation states, Ce4+/3+ and Sn4+/2+ redox couples).

The fact that the coke provoked by carbon deposition has shown to intervene effectively in

different amounts on all the catalysts during CH4 decomposition because of the black colour

of the catalysts and this can greatly reduce the IR signal intensities.

Nevertheless, in any case, as evaluated from the spectra using Kubelka-Munk function to

obtain quantitative analysis [175-179], there is an approximately linear correlation between

the amount of converted methane and the sum of the amount of CO and CO2 gases plus the

amount of carbon stored on the catalyst, thus confirming a closed carbon mass balance within

the experimental error.

Finally, in contrast with the considerable amount of papers published about the activation of

methane on pure transition metals, the theoretical studies of this reaction on metal oxides have

not yet succeeded in giving a sufficiently clear and complete explanation of this process and

the reaction pathway is still up for debate.

103

Page 104: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

SUMMARY AND CONCLUSIONS The Mo/Al2O3, Mo/CeO2, Mo/SnO2, Ce-Mo/Al2O3 and Sn-Mo/Al2O3 samples prepared by

impregnation and co-precipitation methods showed either different Mo dispersity or catalytic

activity towards CO adsorption and CH4 transformation. These catalysts with high Mo

loadings were prepared at low pH taking into account the calcination temperature (600°C) and

time as well as the solution pH, the isoelectric point and surface area of the solid support.

Two types of surface molybdena species have been identified. These are the surface bound

MoO42- polymeric species due to the concomitant strong interaction revealed between Mo and

support and free MoO3 crystallites that validated by XRD and DRIFT results. Accordingly,

high Mo loadings (>15 wt% on Al2O3) are necessary to obtain considerable amounts of free

MoO3 favorable for keeping the active metal in a higher dispersion state. Free MoO3

crystallites are more easily reducible than MoO42- molybdates with tetrahedral configuration

strongly bound to Al2O3 (Mo-O-Al bonds) in the form of Al2(MoO4)3.

The Mo/Al2O3 material had the highest specific surface area (117.6 m2/g) and rather presented

mesopores of regular distribution. However, this material showed the highest thermal stability

while the Mo/CeO2, Mo/SnO2, Ce-Mo/Al2O3 samples undergo structural modifications above

700°C. This may be due to high oxygen storage and release of ceria and tin resulting in lattice

defects and thus enhancing the mobility of metal ions and mutual interactions between them.

One may indicate that the introduction of cerium promotes aggregation of the Mo particles

(particularly as a support) probably due to charge effects and/or to the strong basic property

possessed by CeO2 as further emphasized by means of DRIFT (Mo–O–Ce linkages) and XRD

(decreased crystallites size of MoO3 species and different molecular formulae between Mo

and Ce). This led to the increase of polymerized surface Mo species besides the formation of

coupled O=Mo=O bonds (bands at 995 and 1035 cm-1) indicative of polymeric MoO3.

In the meanwhile, doping Mo/Al2O3 with SnO2 leads to surface structure definitely different

from that of Mo/SnO2. In the case of Sn-Mo/Al2O3, MoO3 crystals completely disappeared

and transformed into MoO2 with the presence of SnO, whereas Mo/SnO2 was formed only by

the two phases of MoO3 and SnO2 with high crystallinity but in both catalysts no linkages

were observed between Mo and Sn ions after calcination at 600°C and molybdate species

strongly affect the growth of SnO2 crystals. The major point that should be outlined is that the

characteristic of Mo-Sn system is the result of the preparation method employed.

104

Page 105: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

On the other hand, the thermal behaviour of Mo/SnO2 showed the dissolution of molybdena

ions in the SnO2 crystals above 750°C. Thereby its high activity revealed in CH4 conversion

might be associated with the former. It is reasonable to note that the marked dissolution was

observed in all thermal treatments of Mo/SnO2 above 750°C (TG-DTA, CO adsorption and

CH4 reaction after H2 reduction at 800°C) with somewhat different extents.

One may notice that the reduction of the catalysts improves the surface reactivity leading to

the presence of small amounts of metallic Mo after reduction at 700°C. Moreover, further

reduction up to 800°C enhanced the adsorption of CO so as to forming various types of

carbonates, therefore as the reduction continues more coordinatively unsaturated (CUS) sites

will be produced and thus resulting in more CO adsorption sites.

CO chemisorption at 100°C on Mo/Al2O3 reduced at different temperature mainly occurred

on small MoO3 crystallites with only one band appearing at 2197 cm-1 corresponding to

octahedral alumina sites (Aloct+3–CO) implying that Al2(MoO4)3 is difficult reducible.

Furthermore, CO chemisorption at 100°C on Mo/Al2O3 leads to the formation of formates,

carboxylates and carbonates. On the other hand, CO chemisorption on Mo/Al2O3, Mo/CeO2,

Mo/SnO2, Ce-Mo/Al2O3 and Sn-Mo/Al2O3 catalysts reduced at 800°C involves oxygen in the

catalysts such oxygen could be present in Mo=O, Mo–O–Mo, Ce–O–Mo, Ce–O–Ce and

Sn–O–Sn associates to form carbonates. It is reasonable to suggest that the formation of

carbonate species involves reduced catalysts with hydroxyl groups and oxygen vacancies

implying the existence of reactive lattice oxygen that may be due to the weakening of the

covalence of the metal-oxygen lattice bonds and/or the enhancement of the mobility of lattice

oxygen sites. Within this context, intimate coupling of Mo with Ce and Sn ions of different

oxidation states has great facilities for electron exchange interactions. Thus, the electron-

mobile environment necessitated by redox reactions is established that has a great share in

enhancing the CO adsorption and therefore, leading to apparently high coverage of carbonate

species on the catalysts reduced at 800°C.

Concerning the CH4 decomposition on Mo/Al2O3 the results showed that molybdena oxide

could activate methane and oxidize it into surface formates and carbon monoxide but methane

conversion is low. However, for CH4 decomposition on Mo/CeO2 and Ce-Mo/Al2O3 the

results permit to infer that the beneficial effects occurred either because the cooperation

between Ce and Mo interfacial active sites generated with higher activity or because the

oxidative properties of CeO2 increased the dissociation of CH4 resulting in the liberation of

the apparently high coverage of carbonates besides CO and CO2, and as a result the methane

conversion increased mainly on Mo/CeO2.

105

Page 106: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

For CH4 decomposition on Sn-Mo/Al2O3 the CH4 conversion is low with the formation of

carbonate species while Mo/SnO2 revealed the highest CH4 activity (high conversion into

CO2) and selectivity leading to the formation of formaldehyde intermediate and to almost total

CH4 oxidation too. This relevant activity and selectivity is presumably due to the dissolution

of molybdena ions in the SnO2 crystals resulting in more active sites.

Comparatively, the presence of CeO2 and SnO2 contributes to the rapid activation of CH4,

thereby accelerating the carbon gasification reaction to produce CO2 decreasing the coke

could be provoked by carbon deposition that has shown to intervene effectively with

somewhat different rates throughout CH4 decomposition. Accordingly, the following order of

the catalysts can be affirmable in accordance with the increase of CH4 conversion:

Mo/SnO2 > Mo/CeO2 > Ce-Mo/Al2O3 > Sn-Mo/Al2O3 > Mo/Al2O3

Finally, the present results suggest that Mo/CeO2 and Mo/SnO2 reduced at 800°C have the

most likely active species for CO adsorption and CH4 decomposition (especially Mo/SnO2)

due to highly dispersed MoO3 species besides Ce3+/Ce4+ and Sn2+/Sn4+ redox couples that

have high capacity towards oxygen. These species were responsible for their catalytic activity

revealed in CH4 oxidation.

Indeed, these results in correlation with the literature suggested that the higher dispersion of

MoO3 on a highly reducible support leads to more active and selective catalysts by optimizing

the interaction with the support throughout the preparation procedure.

Furthermore, it is inferred that there is an approximately linear correlation either between the

increase of the extent of reduction up to 800°C and the increasing integrated absorbance of CO

adsorbed on the catalysts or between the amount of converted methane and the sum of the

amount of CO and CO2 gases plus the amount of carbon stored on the catalysts.

106

Page 107: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

ACKNOWLEDGMENTS I wish to express my greatest thanks to my supervisor Professor Ákos Rédey for enabling this

research to be feasible and for his friendly guidance and his incredible assistance.

I would like to express my sincere appreciation to the following persons whose support made

this research work possible:

Dr. József Kovács and Dr. Tibor Egyházy for their extraordinary help in making some

measurements.

Prof. Monica Caldararu from the Institute of Physical Chemistry “ilie murgulescu” of the

Romanian Academy of Sciences and Dr. Roman Dula from the Institute of Catalysis and

Surface Chemistry in Krakow for their knowledge and cooperation, guidance and help in the

Electron Spin Resonance (ESR) measurements.

Prof. János Kristóf for a very good course in infrared spectroscopy.

Prof. Dénes Kalló, Prof. Pál Tétényi and Dr. Jenő Hancsók for their advices in preparing the

manuscript.

Dr. Tatiana Yuzhakova, Pál Bui and all my colleagues at the Institutional Department of

Environmental Engineering and Chemical Technology for their friendship and encouragement

received from them throughout my research work.

107

Page 108: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

REFERENCES

1. Szabó Z., Kalló D.: Kontakt katalízis, Akadémiai Könyvkiadó, Budapest 1966. 2. H. Jens: Industrial Catalysis, a practical approach, Wiley-VCH, Weinheim 1999. 3. Rédey Á.: Akadémiai Doktori, Veszprémi Egyetem, 1994. 4. Á. Rédey, W. K. Hall: J. Catal., 108, 185 (1987). 5. Á. Rédey, J. Goldwasser, W. K. Hall: J. Catal., 113, 82 (1988). 6. F.E. Massoth: J. Catal., 36, 164 (1975). 7. F.E. Massoth: J. Catal., 30, 204 (1973). 8. F.E. Massoth: Adv. Catal., 27, 265 (1978). 9. W. K. Hall and F. E. Massoth: J. Catal.: 34, 41 (1974). 10. J. Valyon, W. K. Hall: J. Catal., 84, 216 (1983). 11. L. Wang, W. K. Hall: J. Catal., 77, 232 (1982). 12. L. Wang: PhD dissertation, University of Wisconsin-Milwaukee, 1982. 13. W.K. Hall and M. LoJacono: Proc. VIth intl. Cong. Catal., 1, 246 (1977). 14. W. S. Millman and W. K. Hall: J. Catal., 60, 404 (1979). 15. N. Giordano, J.C.J. Bart, A. Vaghi, A. Castellan, G. Martinotti: J. Catal., 36, 81 (1975) 16. J. Medema: J. Catal., 37, 91 (1975). 17. J. Medema, C. Van Stam, V. H. J. De Beer, A. J. A. Konings, D. C. Koningsberger: J.

Catal. 53, 386 (1978). 18. J. P. Peri: J. Phys. Chem., 86, 1615 (1982). 19. J. P. Peri: J. ACS Symp. Ser. no., 288, 422 (1985). 20. H. Knözinger, P. Ratnasamy: Catal. Rev. Sci. Eng., 17 (1), 31-70 (1978). 21. P. Ratnasamy, H. Knözinger: J. Catal., 54, 155 (1978). 22. M.I. Zaki, H. Knözinger: J. Spectrochim. Acta., 43A, 1455 (1987). 23. M.I. Zaki, H. Knözinger: J. Mater. Chem. Phys., 17, 201 (1987). 24. H. Ballinger, J.T. Yates: J. Langmuir, 7 (12), 3041 (1991). 25. J. Goldwasser, S. M. Fang, M. Houalla, W. K. Hall: J. Catal., 115, 34 (1989). 26. R. Nakamura, D. Pioch, R.G. Bowman, R.L.Jr. Burwell: J. Catal., 93, 388 (1985). 27. E.A. Paukshtis, E.N. Yurchenko: J. Russ. Chem. Rev., 52, 242 (1983). 28. T. P. Beebe, P. Gelin, J. T. Yates: J. of Surf. Sci., 148, 526 (1984). 29. J. Chung, J. P. Zhang, R. L. Burwell: J. of Catal., 116, 506 (1989). 30. M. Dufaux, M. Che, and C. Naccache: J. Chim. Phys., 67, 527 (1970). 31. M. Digne, P. Sautet, P. Raybaud, P. Euzen, H. Toulhoat: J. of Catal., 211, 1-5 (2002). 32. H. Knözinger, H. P. Boehm: J. Cat. Sci. and Techn., Vol. 4, 39 (1983). 33. D. M. Hercules, M. Yamada, J. Yasumaru, M. Houalla: J. Phys. Chem., 95, 7037

(1991). 34. D. M. Hercules, V. Sazo, L. Gonzalez, J. Goldwasser, M. Houalla: Surf. Interface

Anal. 23, 367 (1995). 35. D. M. Hercules, J. Goldwasser, C. Scott, M. J. Perez Zurita, M. Houalla: Catal. Lett.,

32, 273 (1995). 36. J. Sonnemans, P. Mars: J. Catal. 31, 209 (1973). 37. A. Iannibello, F. Trifiro: Z. Anorg: Allg. Chem., 413, 293 (1975). 38. S. Ardizzone and S. Trasatti: Adv. Coll. Inter. Sci., 64, 173 (1996). 39. L. K. Chau and, M. D. Porter: J. Coll. Inter. Sci., 145, 283 (1995). 40. E. McCafferty, J. P. Wightman: J. Coll. Inter. Sci., 194, 344–355 (1997). 41. G. D. Parfitt and D.W. Fuerstenau: J. Coll. Sci. 48, 415 (1976). 42. J. Sarrın, O. Noguera, H. Royo, M. J. Perez Zurita, C. Scott, M.R. Goldwasser, J.

Goldwasser, M. Houalla: J. Mol. Catal. A: Chemical 144, 441–450 (1999). 43. J. Engelhardt, J. Goldwasser, W. K. Hall: J. Catal., 76, 48 (1982).

108

Page 109: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

44. J. Engelhardt, J. Goldwasser, W. K. Hall: J. Mol. Catal., 15, 173 (1982). 45. N. Koizumi, G. Bian, K. Murai, T. Ozaki, M. Yamada: J. Mol. Catal. A: Chemical,

207, 171–180 (2004). 46. E. C. DeCanio, D. A. Storm: J. of Catal., 130, 653 (1991). 47. I. Oliveros, M. J. Perez Zurita, C. Scott, M. R. Goldwasser, J. Goldwasser, S. Rondon,

M. Houalla, D.M. Hercules: J. Catal., 171, 485 (1997). 48. P. Tétényi, T. I. Korányi, M. Dobrovolszky, T. Koltai, K. Matusek, Z. Paál: Fuel Proc.

Tech. 61, 55-71 (1999). 49. N. Y. Topsøe, H. Topsøe: J. Catal., 84, 386 (1983). 50. N. Y. Topsøe, H. Topsøe: J. Catal., 139, 631 (1993). 51. A. J. Brungs, A. P. E. York, J. B. Claridge, C. Marquez-Alvarez, M. L. H. Green:

Catal. Lett., 70, 117 (2000). 52. B. Claridge, A.P.E. York, C. Marquez-Alvarez, A.J. Brungs, J. Sloan, S.C. Tsang,

M.L.H. Green: J. Catal. 180, 85 (1998). 53. A. P. E. York, J. B. Claridge, A. J. Tsang, M. L. H. Green: Chem. Comm., 39, (1997). 54. S. Yokota, K. Okumura, M. Niwa: Catal. Lett., 84, 131 (2002). 55. E. McCafferty, P. M. Natishan, G. K Hubler: Corros. Sci., 35, 239 (1993). 56. E. McCafferty: Corros. Sci., 37, 481 (1995). 57. M. Mantel, J. P.Wightman: Surf. Inter. Anal., 21, 595 (1994). 58. G. Ramis, G. Busca: J. Mol. Struct., 193, 93 (1989). 59. G. Ramis, G. Busca, M. Turco, E. Kotur, R. J.Willey: J. Catal., 157, 523 (1995). 60. G. Ramis, G. Busca, C. Cristiani, L. Lietti, P. Forzatti: Langmuir, 8, 1744 (1992). 61. C. U. I. Odenbrand, J.G.M. Brandin, G. Busca: J. Catal., 135, 505 (1992). 62. M. Trombetta, G. Busca, S. A. Rossini, V. Piccoli: J. Catal., 168, 349 (1997). 63. H. C. Yao: J. Catal., 70, 440 (1981). 64. M. C. Kung, H. H. Kung: J. Catal. Rev.-Sci. Eng., 27 (3), 425 (1985). 65. L. Dong and Y. Chen: Farady Trans., 92, 22 (1996). 66. I. E. Wachs, A. M. Turek: J. Phys. Chem., 96, 5000-5007 (1992). 67. E. Odier, Y. Schuurman, K. Barral, C. Mirodatos: Surf. Sci. Catal., 147, 79 (2004). 68. E. Odier, Y. Schuurman, H. Zanthoff, C. Millet, C. Mirodatos, Stud. Surf. Sci. Catal.,

133, 327 (2001). 69. K. M. S. Khalil: J. Catal., 178, 198 (1998). 70. K. M. S. Khalil, L. A. Elkabee, B. Murphy: Micro. Meso. Mater., 78, 83 (2005). 71. K. M. S. Khalil, L. A. Elkabee, B. Murphy: J. Coll. Interf. Sci., 287, 534 (2005). 72. K. M. S. Khalil: J. Coll. Interf. Sci., 307, 172–180 (2007). 73. F. Boccuzzi, A. Chiorino, S. Tsubota, M. Haruta: J. Phys. Chem., 100, 3625 (1996). 74. M. Manzoli, F. Boccuzzi, A. Chiorino, F. Vindigni, W. Deng, M. F. Stephanopoulos:

J. Catal., 245, 308–315 (2007). 75. F. Menegazzo, M. Manzoli, A. Chiorino, F. Boccuzzi, T. Tabakova, M. Signoretto, F.

Pinna, N. Pernicone: J. Catal., 237, 431 (2006). 76. T. Tabakova, F. Boccuzzi, M. Manzoli, D. Andreeva: Appl. Catal. A, 252, 385 (2003). 77. F. Boccuzzi, A. Chiorino, M. Manzoli, D. Andreeva, T. Tabakova: J. Catal., 188, 176

(1999). 78. M. Manzoli, A. Chiorino, F. Boccuzzi: Appl. Catal. B, 52, 259 (2004). 79. M. M. Mohamed: Appl. Catal. A, 267,135 (2004). 80. M. M. Mohamed, S. M. A. Katib: Appl. Catal. A, 287, 236–243 (2005). 81. G. A. Lucia, F. M. Goncalves, P. R.S. Medeiros: Appl. Catal. A: General 208,

265–270 (2001). 82. J. Kristóf, H. Nasser, R. L. Frost, A. De Battisti, Á. Rédey: Appl. Surf. Sci., 242,

13–20 (2005).

109

Page 110: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

83. J. Kristóf, N. Heider, R. L. Frost and V. Vágvölgyi: J. Therm. Anal. Calorim. Vol. 78, 687–695 (2004).

84. Y. Okamoto, K. E. Oh-Hiraki and S. Teranishi: J. Catal. 71, 99 (1981). 85. G. Kozma, Á. Kukovecz, Z. Kónya: J. Mol. Structure 834, 430–434 (2007). 86. J. Madarász, P. Bombicz, M. Okuyac, S. Kanekoc, Gy. Pokol: J. Anal. Appl. Pyr. 72,

209–214 (2004). 87. R. S. Patil, M. D. Uplane, P. S. Patil: Appl. Surf. Sci. 252, 8050–8056 (2006). 88. S. Mihaiu, G. Postole, M. Carata, M. Caldararu, D. Crisan, N. Dragan, M. Zaharescu:

J. Eur Ceramic Society 24, 963–967 (2004). 89. K. V. R. Chary, K. R. Reddy, G. Kishan, J. W. Niemantsverdriet, Gerhard Mestl: J.

Catal., 226, 283–291 (2004). 90. J. M. Herrmann, F. Villain, L. G. Appel: Appl. Catal. A: General 240, 177–182

(2003). 91. E. Comini, M. Ferroni, V. Guidi, G. Martinelli, G. Sberveglieri: Sens. and Acts. B 95,

157–161 (2003). 92. L. B. Gulina, V. P. Tolstoy: Thin Solid Films 440, 74–77, (2003). 93. Y. Maehara, T. Kimura, H. Kawanowa, Y. Gotoh: J. Crys. Growth 275, 1619–1624

(2005). 94. T. Serin, N. Serin, S. Karadeniz, H. Sarı, N. Tugluoglu, O. Pakma: J. Non-Crys. Solids

352, 209–215 (2006). 95. M. Ivanovskaya, E. Lutynskaya, P. Bogdanov. Sens. and Acts. B 48, 387–391 (1998). 96. A. Chiorino, G. Ghiotti, F. Prinetto, M. C. Carotta, D. Gnani, G. Martinelli: Sens. and

Acts. B 58, 338–349 (1999). 97. A. Chiorino, G. Ghiotti, M. C. Carotta, G. Martinelli: Sens. and Acts. B 47, 205–212

(1998). 98. E. Heracleous, M. Machli, A. A. Lemonidou, I. A. Vasalos: J. Mol. Catal. A:

Chemical 232, 29–39 (2005). 99. D. Amalric-Popescu, F. Bozon-Verduraz: Catal. Today 70, 139–154 (2001).

100. M. Niwa, J. Igarashi: Catal. Today 52, 71-81 (1999). 101. F. Goncalves, P. R. S. Medeiros, J. G. Eon, L. G. Appel: Appl. Catal. A: General 193,

195–202 (2000). 102. P. R. S. Medeiros, L. G. Appel: Appl. Catal. A: General 231, 125–130 (2002). 103. H. Liu, P. Cheung, E. Iglesia: J. Catal. 217, 222-229 (2003). 104. V. Lochar: Appl. Catal. A: General 309, 33–36 (2006). 105. N. G. Valente, L. E. Cadtis, O. F. Gorriz, L. A. Arrtia, J. B. Rivarola: Appl. Catal. A:

General 153, 119-132 (1997). 106. M. Daturi and L. G. Appel: J. Catal., 209, 427–432 (2002). 107. R. Rai, T. D. Senguttuvan, S. T. Lakshmikumar: Comp. Mater. Sci., 37, 15–19 (2006). 108. C. da Silva, L. G. Appel, S. Braun: Appl. Catal. A: General 314, 172–178 (2006). 109. J. Lercher, E. Heracleous, A. Lemonidou: J. Appl. Catal., 264, 73-80 (2004). 110. L. Lin, Y. Xu: Appl. Catal. A, 188, 53 (1999). 111. Y. Yang, W. Li, H. Xu: React. Kinet. Catal. Lett., 77, 155 (2002). 112. T. Armaroli, D. Minoux, S. Gautier, P. Euzen: Appl. Catal. A, 251 241–253 (2003). 113. D. C. LaMont, W. J. Thomson: Chem. Eng. Sci., 60, 3553 – 3559 (2005). 114. E. Iglesia, W. Ding, G. D. Meitzner: J. Catal., 206, 14 (2002). 115. A. R. S. Darujati, D. C. LaMont, W. J. Thomson: Appl. Catal., 253, 397 (2003). 116. A. R. S. Darujati, D. C. LaMont, A. J. Gilligan, A. S. Chellappa, W. J. Thomson:

Appl. Catal., 255, 239 (2003). 117. J. A. Anderson, Z. Wang, C. H. Rochester: J. Catal. 184, 213-223 (1999). 118. R. Wang, H. Xu, X. Liu, Q. Ge, W. Li: Appl. Catal. A, General 305, 204–210 (2006).

110

Page 111: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

119. A. Trovarelli: Catal. Rev. Sci. Eng., 38, 439 (1996). 120. A. Trovarelli, M. Boaro, E. Rocchini, C. de Leitenburg, G. Dolcetti: J. Alloys Compd.,

374, 387 (2004). 121. M. Boaro, F. Giordano, S. Recchia, V. Dal Santo, M. Giona, A. Trovarelli: Appl.

Catal. B: Environ., 52, 225 (2004). 122. H. Jenő: Alternatív motorhajtóanyagok, Veszprémi Egyetemi Kiadó, Veszprém, 2004. 123. G. A. Olah, J. A. Olah: J. Am. Chem. Soc., 93, 1256 (1971). 124. T. Bunluesin, H. Cordatos, R.J. Gorte: J. Catal., 157, 222 (1995). 125. T. Bunluesin, E. S. Putna, R. J. Gorte: Catal. Lett., 41, 1 (1996). 126. T. Bunluesin, G. W. Graham, R. J. Gorte: Appl. Catal. B, 14, 105 (1997). 127. S. Hilaire, X. Wang, T. Luo, R. J. Gorte, J. Wagner: Appl. Catal. A, 215, 271 (2001). 128. H. Amariglio, P. Pareja, A. Amariglio: Catal. Today, 25, 113 (1995). 129. H. Amariglio, M. Belgued, P. Pareja, A. Amariglio: Catal. Lett., 31, 19 (1995). 130. H. Amariglio, A. Amariglio, M. Belgued, P. Pareja: J. Catal., 177, 113 (1998). 131. R. A. van Santon, T. Koerts, M. T. A. G. Deelen: J. Catal., 138, 101 (1992). 132. R. A. van Santon, T. Koerts: J. Mol. Catal., 74, 185 (1992). 133. F. Solymosi, A. Erdohelyi, J. Cserenyi, A. Felvegi: J. Catal., 147, 272 (1994). 134. F. Solymosi, J. Cserenyi: Catal. Lett., 34, 343 (1995). 135. L. Guczi, K. V. Sarma, L. Borko: J. Catal., 167, 495 (1997). 136. L. Guczi, L. Borko: Catal. Today, 64, 91 (2001). 137. M. C. J. Bradford: Catal. Lett., 66, 113 (2000). 138. G. Jacobs, L. Williams, U. Graham, G. A. Thomas, D. E. Sparks, B. H. Davis: Appl.

Catal. A: General, 252, 107–118 (2003). 139. H. Liu, T. Li, B. Tian, Y. Xu: Appl. Catal. A, 213, 103 (2001). 140. D. Ma, D. Wang, L. Su, Y. Shu, Y. Xu, X. Bao: J. Catal., 208, 260 (2002). 141. Y. Xu, X. Bao, L. Lin: J. Catal., 216, 386–395 (2003). 142. L. Wang, R. Ohnishi, M. Ichikawa: J. Catal., 190, 276 (2000). 143. C. Müller, C. Freysoldt, M. Baudin, K. Hermansson: Chem. Phys., 318, 180 (2005). 144. B. Herschend, M. Baudin, K. Hermansson: Chem. Phys., 328, 345–353(2006). 145. R. Craciun, W. Daniell, H. Knözinger: Appl. Catal. A, 230, 153 (2002). 146. K. Otsuka, Ye Wang, E. Sunada, I. Yamanaka: J. Catal., 175, 15 (1998). 147. F. Solymosi, A. Erdohelyi, A. Szoke: Catal. Lett., 32, 43 (1995). 148. A. Szoke, F. Solymosi: Appl. Catal. A, 142, 361 (1996). 149. J. Kiss, F. Solymosi: J. Catal., 179, 277 (1998). 150. F. Solymosi, L. Bugyi, A. Oszko: Catal. Lett., 57, 103 (1999). 151. F. Solymosi, and M. Pasztor: J. Catal. 104, 312 (1987). 152. A. Berko, F. Solymosi: J. Catal., 183, 387 (1999). 153. J. H. Lunsford: Stud. Surf. Sci. Catal., 81, 1 (1993). 154. J. H. Lunsford, D. Wang, M. P. Rosynek: Top. Catal., 3, 289 (1996). 155. J. H. Lunsford, D. Wang, M. P. Rosynek: J. Catal., 169, 347 (1997). 156. M. Ichikawa, Y. Shu, R. Ohnishi: J. Catal., 206, 134 (2002). 157. M. Ichikawa, R. Onishi, S. Liu, Q. Dong, L. Wang: J. Catal., 182, 92 (1999). 158. P. Mériaudeau, L. V. Tiep, V. T. T. Ha, C. Naccache, G. Szabo: J. Mol. Catal. A, 144,

469 (1999). 159. P. Mériaudeau, V. T. T. Ha, L. V. Tiep: Catal. Lett., 64, 49 (2000). 160. A. A. Said and S. A. Halawy: J. Therm. Anal. Cal., Vol. 41, 1075-1090 (1994). 161. L. B. Sharma and S. Batra: J. Therm. Anal. Cal., Vol. 34, 1273-1281 (1988). 162. N. R. E. Radwan, M. Mokhtar and G. A. El-Shobaky: J. Therm. Anal. Cal., Vol. 71,

977–986 (2003). 163. Z. Sarbak and W. K. Józwiak: J. Therm. Anal. Cal., 87, 229 (2007).

111

Page 112: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

164. R. Lyszczek: J. Therm. Anal. Cal., 90, 533 (2007). 165. M. Blachnio, P. Staszczuk, G. Grodzicka, L. Lin and Y. X. Zhu: J. Therm. Anal. Cal.,

88, 601 (2007). 166. O. Cozar, D.A. Magdas, I. Ardelean: J. Non-Crys. Solids 354, 1032–1035 (2008). 167. L. Bih, A. Nadiri, H. Khemakhem, B. Elouadi: J. Mol. Struct. 872, 1–9 (2008). 168. D. Zayachuk, Ye. Polyhach, E. Slynko, O. Khandozhko, C. Rudowicz: Spectr. Acta

Part A 60, 1247–1256 (2004). 169. G. Mitrikas, G. Kordas, G. Fanourakis, E. Simoen: Nuc. Inst. and Meth. in Phy.

Research B 186, 181–185 (2002). 170. B. Kamp, R. Merkle, J. Maier: Sens. and Acts. B 77, 534-542 (2001). 171. B. Kamp, R. Merkl, R. Lauck, J. Maier: J. Solid State Chem. 178, 3027–3039 (2005). 172. F. Morazzoni, C. Canevali, N. Chiodini, C. Mari, R. Ruffo, R. Scotti, L. Armelao,

E. Tondello, L. Depero, E. Bontempi: Mats. Sci. and Eng. C 15, 167–169 (2001). 173. A. Martinez-Arias, J. Soria, J. C. Conesa, X. L. Seoane, A. Arcoya: J. Chem. Soc.

Faraday Trans. 91, 1679 (1995). 174. M. Figaj, K. D. Becker: Solid State Ionics, 141, 507-512 (2001). 175. L. Vien, D. Colthup, N. B. Fateley, W. G. and Grasselli, J. G., “The Handbook of

Infrared and Raman Characteristic Frequencies of Organic Molecules.” Academic Press, San Diego, 1991.

176. J. Ryczkowski: Catal. Today, 68, 263–381 (2001). 177. Guido Busca: Catal. Today, 41, 191-206 (1998). 178. A. Zecchina, C. Lamberti, S. Bordiga: Catal. Today, 41, 169-177 (1998). 179. A. Zecchina, D. Scarano, S. Bordiga, G. Ricchiardi, G. Spoto, F. Geobaldo: Catal.

Today, 27, 403 (1996). 180. M. I. Zaki, M. A. Hasan, F. A. Al-Sagheer, L. Pasupulety: Coll. and Surf. A:

Phys. and Eng. Aspects, 190, 261–274 (2001). 181. H. Zhu, Z. Qin, W. Shan, W. Shen, J. Wang: J. Catal., 225, 267–277 (2004). 182. H. Zhu, Z. Qin, W. Shan, W. Shen, J. Wang: J. Catal., 233, 41–50 (2005). 183. B. Herschend, M. Baudin, K. Hermansson: J. Chem. Phys., 120, 4939 (2004). 184. H. Liu, L. Su, H. Wang, W. Shen, X. Bao, Y. Xu: Appl. Catal. A, 236, 263 (2002). 185. U. Berner, K. Schierbaum, G. Jones, P. Wincott, S. Haq, G. Thornton: Surf. Sci., 467,

201 (2000). 186. Z. Yang, T. K. Woo, K. Hermansson: Chem. Phys. Lett., 396, 384–392 (2004). 187. El. Aneggi, M. Boaro, C. de Leitenburg, G. Dolcetti, A. Trovarelli:

J. Alloys and Compounds, 408–412, 1096–1102 (2006). 188. Németh T. Zs.: Doktori értekezés, Veszprémi Egyetem, 1996. 189. N. Heider, R. Ákos, Y. Tatiana, N. T. Zsolt and O. Tamás: Reac. Kinet. And

Catal. Letts., Vol. 92 No. 2, 329-335, (2007). 190. M. R. Smith, U.S. Ozkan: J. Catal., 141, 124 (1993). 191. M. R. Smith, L. Zhang, S. A. Driscoll, and U.S. Ozkan: Catal. Lett., 19, l-15 (1993). 192. S. Chempath, A. T. Bell: J. Catal., 247, 119–126 (2007). 193. V. D. Sokolosvskii, G. M. Aliev, O. V. Buyesvskaya, A. A. Davydov:

Catal. Today, 4, 293 (1989). 194. B. Irigoyen, N. Castellani, A. Juan: J. Mol. Catal. A, 129, 297–310 (1998). 195. I. A. Fisher, A. T. Bell: J. Catal., 172, 222 (1997). 196. A. Fritz, V. Pitchon: Appl. Catal. B, 13, 1 (1997). 197. X. Zhang, D. He, Q. Zhang, Q. Ye, B. Xu, Q. Zhu: Appl. Catal. A: General, 249,

107–117 (2003).

112

Page 113: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

THESES The main results of the dissertation are summarized in the following thesis points:

1. Taking into account the calcination temperature and time as well as the solution pH,

the isoelectric point and the surface area of the solid support high Mo loadings (>15

wt% on Al2O3) are necessary to obtain considerable amounts of free MoO3 crystallites

favorable for keeping the active metal in a higher dispersion state. Thus, MoO3

clusters are more easily reducible than MoO42- molybdates with tetrahedral

configuration strongly bound to Al2O3 (Mo-O-Al bonds) in the form of Al2(MoO4)3

and the higher activity of Mo/Al2O3 may be associated with the former.

2. For high Mo loadings obtained two types of molybdena species were the predominant

surface species. These are the surface bound MoO42- polymeric species due to the

concomitant strong interaction revealed between Mo and support and free MoO3

crystallites that validated by XRD and DRIFT results.

3. The Mo/Al2O3 material showed the highest thermal stability up to 900°C while the

Mo/CeO2, Mo/SnO2, Ce-Mo/Al2O3 samples undergo morphological and structural

modifications above 700°C resulting in lattice defects, thus enhancing the mobility of

metal ions and the possibility of interactions between them.

4. The introduction of cerium promotes aggregation of the Mo particles (particularly as a

support) probably due to charge effects and/or to the strong basic property possessed

by CeO2 so as to forming different molecular formulae. This led to the increase of

polymerized surface Mo species besides the formation of coupled O=Mo=O bonds

indicative of polymeric MoO3 as further emphasized by means of DRIFT and XRD.

5. Concerning the use of SnO2, the major point that should be outlined:

a) The characteristic of Mo-Sn system is the result of the preparation method adopted.

However, doping Mo/Al2O3 with SnO2 leads to surface structure definitely different

from that of Mo/SnO2. Accordingly, when using SnO2 as promoter the MoO3 crystals

completely disappeared and transformed into MoO2 with the presence of SnO whereas

Mo/SnO2 was formed only by MoO3 and SnO2 oxides. Meanwhile, in both cases, no

linkages were observed between Mo and Sn ions after calcination at 600°C and

molybdate species strongly affect the growth of SnO2 crystals.

b) The thermal behaviour of Mo/SnO2 showed the dissolution of molybdena ions in

the SnO2 crystals above 750°C with somewhat different extents, thereby resulting in

more active sites and thus leading to a high catalytic activity of Mo/SnO2 catalyst.

113

Page 114: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

6. The hydrogen reduction of the catalysts improves the surface reactivity generating

oxygen vacancies and coordinatively unsaturated sites (CUS) leading to the presence

of small amounts of metallic Mo after reduction at 700°C. Moreover, further reduction

up to 800°C enhanced their activity towards CO adsorption and CH4 dissociation.

7. The peculiarities of in situ DRIFT studies of CO adsorption and CH4 transformation

on the catalysts have been achieved under reaction conditions. Thus, the noticed gain

in the intensity for the bands in conjunction with various types of carbonate species

has been observed upon CO adsorption at 100°C and CH4 decomposition at 700°C.

This involves reduced catalysts containing coordinatively unsaturated sites (CUS) with

hydroxyl groups and oxygen vacancies so as to forming various carbonate species

implying the existence of the reactive lattice oxygen in the catalysts such oxygen

could be present in Mo=O, Mo–O–Mo, Ce–O–Mo, Ce–O–Ce and Sn–O–Sn entities.

8. The Mo/CeO2 and Mo/SnO2 catalysts reduced at 800°C have the most likely active

species for CO adsorption and CH4 dissociation. The highly dispersed MoO3 species

besides Ce3+/Ce4+ and Sn2+/Sn4+ redox couples that have high capacity towards oxygen

were responsible for the high catalytic activity revealed by Mo/SnO2 and Mo/CeO2.

Thus, the CH4 conversion increased in accordance with the following order of the

catalysts:

Mo/SnO2 > Mo/CeO2 > Ce-Mo/Al2O3 > Sn-Mo/Al2O3 > Mo/Al2O3

114

Page 115: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

PUBLICATIONS

1. H. Nasser, Á. Rédey, T. Yuzhakova, J. Kovács: Thermal stability and surface

structure of Mo/CeO2 and Ce-doped Mo/Al2O3 catalysts, Journal of Thermal Analysis

and Calorimetry 2008, accepted for publication.

2. H. Nasser, Á. Rédey, T. Yuzhakova, J. Kovács: In situ DRIFT study of nonoxidative

methane reaction on Mo/SnO2 catalyst, Reaction Kinetics and Catalysis letters 2008,

accepted for publication.

3. H. Nasser, Á. Rédey, T. Yuzhakova, Zs. N. Tóth and T. Ollár: FTIR study of CO

adsorption on molybdena-alumina catalysts for surface characterization, Reaction

Kinetics and Catalysis letters, Vol. 92 No. 2, 329-335, 2007.

4. Nasser H., Rédey Á., Yuzhakova T.: Structure and Thermal Stability of Ceria-doped

Mo/Al2O3 Catalysts, Environmental Engineering and Management Journal, Vol. 5 No.

3, 425-432, 2006.

5. Nasser H., Rédey Á., Yuzhakova T: Structure and Surface Chemistry of Ceria-doped

Mo/Al2O3 Catalysts, MicroCAD International Conference Proceedings pp.

65-71, 2007.

6. Nasser H., Rédey Á., Yuzhakova T.: Környezetvédelmi felhasználású cérium tartalmú

Mo/Al2O3 katalizátorok struktúrája és termikus stabilitása, Országos

Környezetvédelmi Konferencia kiadványa pp. 285-293, 2006.

7. Nasser H., Kristóf J., Rédey Á., R. L. Frost, A. De Battisti: IrO2/SnO2 katalizátorok

képződési mechanizmusának vizsgálata és felületkémiai jellemzése, XIX Országos

Környezetvédelmi Konferencia kiadványa pp. 239-246, 2005.

8. J. Kristóf, H. Nasser, E. Horváth, R. L. Frost, A. De Battisti, Á. Rédey: Investigation

of SnO2 thin film evolution by thermoanalytical and spectroscopic methods, Applied

Surface Science, 242, 13–20, 2005.

9. J. Kristóf, H. Nasser, E. Horváth, R. L. Frost and V. Vágvölgyi: Investigation of

IrO2/SnO2 thin film evolution by thermoanalytical and spectroscopic methods, Journal

of Thermal Analysis and Calorimetry, Vol. 78, 687–695, 2004.

10. Yuzhakova T., Rédey Á., Caldararu M., Auroux A., Carata M., Postole G., Hornoiu

C., Nasser H.: Study of Pt/Sn-Al Catalyst for Environmental Application,

Environmental Engineering and Management Journal, Vol. 5 No. 4, 559-568, 2006.

115

Page 116: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

11. Yuzhakova T., Rédey Á., Nasser H., Caldararu M., Strukova L., Gaál Z., Fazakas J.:

The Effect of Pretreatment on the Catalytic Properties of Supported Tin Catalysts,

Conference Proceedings, Paper Number 133, Brisbane, Chemeca 2005.

12. Yuzhakova T., Rédey Á., Holenda B., Domokos E., Nasser H., Caldararu M., Gaál Z.,

Fazakas J.: Surface Chemistry Studies on Molybdena-Alumina Catalysts, Conference

Proceedings, Paper Number 134, Brisbane, Chemeca 2005.

PRESENTATIONS

1. Nasser H., Rédey Á., Yuzhakova T., Caldararu M.: In-situ DRIFT Studies of Mo/SnO2

and Sn-doped Mo/Al2O3 Catalysts, 8th International Conference of the Romanian

Catalysis Society, Bucharest, Romania, June 21-23, 2007.

2. Nasser H., Rédey Á., Yuzhakova T., Caldararu M.: In-situ DRIFT Studies of Mo/SnO2

and Sn-doped Mo/Al2O3 Catalysts, Past and Present in DeNOx Catalysis, DeNOxCAT

2007, Uzlina, Romania, June 17-20, 2007.

3. Nasser H., Rédey Á., Yuzhakova T.: Structure and Surface Chemistry of Ceria-doped

Mo/Al2O3 Catalysts, MicroCAD International Scientific Conference, Miskolc, March

22-23, 2007.

4. Nasser H., Rédey Á., Yuzhakova T.: Structure and Thermal Stability of Ceria-doped

Mo/Al2O3 Catalysts, ICEEM03/EEMJ Conference, Iasi, Romania, September 21-24,

2006.

5. H. Nasser, J. Kristóf, R. L. Frost, A. De Battisti, Á. Rédey: Investigation of SnO2 thin

film evolution by thermoanalytical and spectroscopic methods, 10th EuCheMS

Conference on Chemistry and the Environment, Rimini: September 4-7, 2005.

6. Rédey Á., Yuzhakova T., Nasser H.: Environmental Impact Assessment Theory and

Practice, Key note lecture, ICEEM03/EEMJ Conference, Iasi, Romania, September

21-24, 2006.

7. Rédey Á., Kováts N., Yuzhakova T., Nasser H.: Environmental Impact Assessment,

Theory and Practice, Key note lecture, 1st European Chemistry Congress, Budapest,

Hungary, August 27-31, 2006.

8. Yuzhakova T., Rédey Á., Auroux A., Caldararu M., Carata M., Postole G., Hornoiu

C., Nasser H., Popescu I., Sandulescu I., Strukova L., Fazakas J.: Effect of Pt Loading

on the Surface and Catalytic Behaviour of Tin Containing Catalysts, 1st European

Chemistry Congress Budapest, Hungary, August 27-31, 2006.

116

Page 117: SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTSkonyvtar.uni-pannon.hu/doktori/2008/Naszer_Heider_dissertation.pdf · SURFACE CHEMISTRY OF MOLYBDENA CONTAINING CATALYSTS Written

9. Yuzhakova T., Rédey Á., Holenda B., Domokos E., Nasser H., Caldararu M., Gaál Z.,

Fazakas J.: Surface Chemistry Studies on Molybdena-Alumina Catalysts, Chemeca

2005, Brisbane, Australia, September 25-28, 2005.

10. Yuzhakova T., Rédey Á., Nasser H., Caldararu M., Strukova L., Gaál Z., Fazakas J.:

The Effect of Pretreatment on the Catalytic Properties of Supported Tin Catalysts,

Chemeca 2005, Brisbane, Australia, September 25-28, 2005.

11. Yuzhakova T., Rédey Á., Caldararu M., Auroux A., Scurtu M., Postole G., Nasser H.,

Fazakas J.: Adsorption Capacity of Tin Oxide Supported Catalysts to Capture Air

Pollutants, 2nd International Conference on Thermal Engines and Environmental

Engineering, Galati, Romania, June 7-9, 2007.

12. Yuzhakova T., Caldararu M., Hornoiu C., Auroux A., Rédey Á., Nasser H.:

Correlation Between Physico-Chemical Characteristics and Electrical Capacitance of

Tin Oxide Supported on Alumina Catalysts, 8th International Conference of the

Romanian Catalysis Society, Bucharest, Romania, June 21-23, 2007.

AWARDS: 1. Ceepus mobility grant to the University of Czestochowa, Poland 2005.

2. Ceepus mobility grant to the University of Poznan, Poland 2006.

3. Ceepus mobility grant to the Technical University of Cluj, Romania 2007.

4. Inclusion in Who’s Who in the World, 2009 Edition.

117