Sei sulla pagina 1di 163

HETEROGENEOUS

PHOTOCATALYTIC
SELECTIVE OXIDATION OF
CYCLOHEXANE

Vincenzo Vaiano

Unione Europea

UNIVERSIT DEGLI
STUDI DI SALERNO

Fondo sociale europeo


Programma Operativo Nazionale 2000/2006
Ricerca Scientifica, Sviluppo Tecnologico, Alta Formazione
Regioni dellObiettivo 1 Misura III.4
Formazione superiore ed universitaria

Department of Chemical and Food Engineering


Ph.D. Course in Chemical Engineering
(IV Cycle-New Series)

HETEROGENEOUS PHOTOCATALYTIC
SELECTIVE OXIDATION OF CYCLOHEXANE

Supervisor
Prof. Paolo Ciambelli
Scientific Referees
Prof. Roger I. Bickley
Prof. Leonardo Palmisano
Prof. Salvatore Vaccaro

Ph.D. Course Coordinator


Prof. Ernesto Reverchon

Ph.D. student
Vincenzo Vaiano

To my grandfather

Acknowledgments
Con questa attivit di ricerca ho potuto fare unimportante esperienza di
vita arricchita e impreziosita dal rapporto quotidiano con persone che mi
hanno aiutato moltissimo a superare le difficolt che incontravo.
Sono tante le persone a cui devo il raggiungimento di questo traguardo,
persone che nei modi pi disparati mi hanno aiutato, spronato, consigliato...e
senza dilungarmi troppo nel cercare le parole migliori per ringraziarle, lo
faccio.
Un ringraziamento importante va al Prof. Paolo Ciambelli per il sostegno
dimostratomi fin dallinizio e per lattenzione con la quale ha seguito e reso
possibile la realizzazione di questo lavoro.
Un ringraziamento particolare va al Prof. Vincenzo Palma e alla Dott.
Diana Sannino, mai paghi nel profondermi approvazione e indispensabili
consigli, a cui devo il raggiungimento di questa importante meta.
Voglio inoltre ringraziare il Prof. Salvatore Vaccaro e il Prof. Leonardo
Palmisano per la loro utile collaborazione.
Ringrazio gli Ingegneri Massimo Ricciardi, Antonietta Maria Manna,
Emma Palo, Arianna Ruggiero, Maria Sarno, Giuseppa Matarazzo, Paola
Russo e Caterina Leone che mi hanno fatto capire quanto sia importante
collaborare in piena serenit, instaurando un rapporto di amicizia.
Vorrei inoltre ringraziare il Prof. Roger I. Bickley per lospitalit
offertami presso lUniversit di Bradford e per i suoi preziosi suggerimenti.
Un ringraziamento va al Prof. Eric M. Gaigneaux e al Dott. Frdric Dury
che hanno realizzato le misure XPS presenti nel lavoro.
Ma ringrazio soprattutto mia madre, mio padre, mia sorella Maria Rita e
mia zia Betta per avermi sostenuto in ogni occasione.

Publications list
1. P. Ciambelli, D. Sannino, V. Palma, V. Vaiano; Photocatalytic
selective oxidation of cyclohexane on Mo-exchanged ferrierite. Proc.
of the 6nd National Congress on Science and Technology of Zeolites,
September 20-23, 2003, Vietri sul Mare, Italy, p.125-126.
2. P. Ciambelli, D. Sannino, V. Palma, V. Vaiano; Gas-solid
photocatalytic oxidative dehydrogenation of cyclohexane to benzene
on MoOx/TiO2. Proc.of the 14nd National Congress on Catalysis,
June 6-10, 2004, Lerici, Italy, p.24-25.
3. P. Ciambelli, D. Sannino, V. Palma, V. Vaiano; Heterogeneous
photocatalytic oxidative dehydrogenation of cyclohexane over
supported Mo oxide catalysts. Proc.of the 13nd International
Congress on Catalysis, July 11-16, 2004, Paris, France, p.191.
4. P. Ciambelli, D. Sannino, V. Vaiano, V. Palma, S. Vaccaro;
Catalysed photooxidative dehydrogenation of cyclohexane to
benzene in a fluidized bed reactor. Proc.of the 7nd World Congress
of Chemical Engineering, July, 10-14, 2005, Glasgow, Scotland,

http://chemengcongress.somcom.co.uk/.
5. P. Ciambelli, D. Sannino, V. Palma, V. Vaiano; Cyclohexane
photocatalytic oxidative dehydrogenation to benzene on sulphated
titania supported MoOx; Stud. Surf. Sci. Catal. 155 (2005) 179-187.
6. P. Ciambelli, D. Sannino, V. Palma, V. Vaiano; Photocatalysed
selective oxidation of cyclohexane to benzene on MoOx/TiO2; Catal.
Today. 99 (2005) 143-149.
7. P. Ciambelli, V. Palma, D. Sannino, S. Vaccaro, V. Vaiano;
Selective oxidation of cyclohexane to benzene on molybdena-titania
catalysts in fluidized bed photocatalytic reactor. THE TOCAT 5
Conference on Advanced Catalytic Science and Technology, July,
23-28, 2006, Tokyo, Japan, submitted.

Contents

Introduction................................................................................. 1
I.1

Heterogeneous photocatalysis.............................................. 2
I.1.1

I.2

Basic Principles........................................................... 3

Fundamental Engineering Aspects....................................... 7


I.2.1

Types of Radiation Sources ......................................... 8

I.2.2

Photoreactor Materials and Geometries..................... 9

I.2.3

Efficiency of Heterogeneous Photocatalytic Systems 10

I.2.4

Modeling of Photocatalytic Reactors ........................ 11

I.3

Aim of the Work ................................................................ 12

II

State of the art....................................................................... 15


II.1

Photocatalysis as potential technology for synthesis of

chemicals from hydrocarbon feedstocks.................................................. 15


II.1.1

Photocatalytic oxidation of cyclohexane................... 16

II.1.2

Oxidative dehydrogenation of cyclohexane............... 17

II.2

Influence of catalyst surface acidity on photocatalytic

performances............................................................................................ 18
II.3
III

Photocatalytic fluidized bed reactors ................................. 19


Experimental

Results:

Photocatalytic

oxidation

of

cyclohexane on MoOx/Al2O3 ...................................................................... 23


III.1

Samples preparation....................................................... 23

III.2

Catalysts characterization .............................................. 23

III.2.1

ICP-MS ...................................................................... 24

III.2.2

Thermal analysis (TG-MS) ........................................ 24

III.2.3

Micro Raman spectroscopy ....................................... 27

III.2.4

FT-IR Spectroscopy ................................................... 28

III.2.5

N2 adsorption measurements| .................................... 28

III.2.6

TPD investigations .................................................... 29

III.3

Laboratory apparatus for catalytic test........................... 29

III.3.1

Feed section............................................................... 30

III.3.2

Reaction section......................................................... 31

III.3.3

Analysis section ......................................................... 32

III.4

Photocatalytic tests conditions and typical trend ........... 33

III.4.1
III.5

Thermodynamic analysis ........................................... 36


Results and discussion ................................................... 37

III.5.1

Specific surface area and Chemical analysis ............ 37

III.5.2

Thermal analysis........................................................ 38

III.5.3

FT-IR spectroscopy.................................................... 40

III.5.4

Micro Raman spectroscopy ....................................... 42

III.5.5

Photocatalytic activity tests ....................................... 44

IV

Experimental

Results:

Photocatalytic

oxidation

of

cyclohexane on zeolites supported MoOx ................................................. 47


IV.1

Samples preparation....................................................... 47

IV.2

Thermal analysis ............................................................ 48

IV.3

N2 adsorption measurements.......................................... 51

IV.4

FT-IR spectroscopy........................................................ 52

IV.5

Photocatalytic activity tests............................................ 54

Experimental

Results:

Photocatalytic

oxidation

of

cyclohexane on MoOx/TiO2 ........................................................................ 59


V.1

II

Effect of molybdenum loading........................................... 59

V.1.1

Samples preparation.................................................. 59

V.1.2

Specific surface area and Chemical analysis ............ 59

V.1.3

Thermal analysis ....................................................... 61

V.1.4

FT-IR spectroscopy ................................................... 64

V.1.5

Micro Raman spectroscopy ....................................... 66

V.1.6

Photocatalytic activity tests....................................... 67

V.1.7

Effect of light intensity............................................... 73

V.2

Influence of sulphate content ............................................. 74

V.2.1

Samples preparation.................................................. 74

V.2.2

Specific surface area, chemical analysis and sample

acidity

................................................................................... 75

V.2.3

Thermal analysis ....................................................... 76

V.2.4

FT-IR spectroscopy ................................................... 77

V.2.5

Micro Raman spectroscopy ....................................... 79

V.2.6

Photocatalytic activity tests....................................... 79

V.3

Photocatalytic oxidative dehydrogenation of cyclohexane:

reaction mechanism ................................................................................. 84


V.3.1

Photocatalytic oxidation of cyclohexanol ................. 84

V.3.2

Photocatalytic oxidation of cyclohexene ................... 86

V.3.3

Role of the sulphate in the reaction mechanism ........ 90

VI

Photocatalytic flat-plate reactor.......................................... 93

VI.1

Experimental set up apparatus and photocatalytic tests

conditions

....................................................................................... 93

VI.2

Photocatalytic activity tests ........................................... 94

VII

Photocatalytic fluidized bed reactor ................................... 99

VII.1

Photocatalytic fluidized bed reactor design ................... 99

VII.2

Preliminary results ....................................................... 102

VII.3

MoOx supported on TiO2/Al2O3 sample ...................... 106

VII.3.1

TiO2-Al2O3 sample preparation .............................. 107

VII.3.2

Results..................................................................... 108

VII.4
comparison

Photocatalytic

fixed

and

fluidized

bed

reactor

..................................................................................... 109

III

VII.5

Effect of light intensity ................................................ 119

VII.6

Photocatalytic oxidation of cyclohexane on sulphated

MoOx/-Al2O3 catalysts.......................................................................... 122


VII.6.1

Samples preparation ............................................... 122

VII.6.2

Specific surface area and Thermal analysis ........... 122

VII.6.3

Photocatalytic activity tests .................................... 124

VII.7

Photocatalytic oxidative dehydrogenation of ethylbenzene

to styrene

..................................................................................... 127

VII.7.1
VIII
IX

IV

Experimental results ............................................... 128

Conclusions ..................................................................... 131


References............................................................................ 133

Index of figures

Figure 1 Schematic representation of a photocatalytic process..................... 2


Figure 2 Change in the electronic structure of a semiconductor compound as
the number N of monomeric units increases from unity to clusters of
more than 2000...................................................................................... 4
Figure 3 Illustration of the main processes occurring on a semiconductor
particle under electronic excitation....................................................... 5
Figure 4 Emission spectrum of a 40W actinic lamp....................................... 9
Figure 5 Schematic representation of the modeling of a photocatalytic
reactor ................................................................................................. 11
Figure 6 Schematic representation of light scattering in (A) film fixed bed,
(B) granular fixed bed and (C) fluidized bed ...................................... 20
Figure 7 TGAQ500 Thermogravimetric Analyzer........................................ 25
Figure 8 SDTQ600 Simultaneous DSC/TGA................................................ 25
Figure 9 Pfieffer Vacuum Benchtop Thermostar mass spectrometer........... 25
Figure 10 TG and DTG curves..................................................................... 26
Figure 11 Dispersive MicroRaman (Invia, Renishaw)................................. 27
Figure 12 Bruker IFS 66 FT-IR spectrophotometer..................................... 28
Figure 13 Costech Sorptometer 1040........................................................... 29
Figure 14 Laboratory apparatus for catalytic test ....................................... 30
Figure 15 Mass flow controller .................................................................... 31
Figure 16 Annular gas-solid photocatalytic fixed bed reactor..................... 32

Figure 17 Outlet reactor concentration (a.u.) of cyclohexane, oxygen


benzene and cyclohexene and (ppm) of carbon dioxide as a function of
run time................................................................................................ 34
Figure 18 Photocatalysed cyclohexane oxy-dehydrogenation to cyclohexene
and benzene ......................................................................................... 35
Figure 19 Grafic interface of Gaseq 0.74 program...................................... 36
Figure 20 Effect of temperature on cyclohexane conversion, carbon dioxide
and benzene selectivity ........................................................................ 37
Figure 21 Evolution of the weight loss during decomposition of ammonium
heptamolybdate together with DTG and DSC signals......................... 39
Figure 22 Thermogravimetric analysis of 8MoAl catalyst after calcination
............................................................................................................. 40
Figure 23 IR spectrum of 8MoAl obtained by subtracting the spectrum of

Al ........................................................................................................ 41
Figure 24 IR spectra of Al and 2MoAl .................................................... 42
Figure 25 Raman spectra of Al and 8MoAl............................................... 43
Figure 26 Raman spectra of Al and 2MoAl ............................................. 44
Figure 27 Cyclohexane conversion on 8MoAl (8Mo) and 2MoAl (2Mo) as
a function of illumination time............................................................. 45
Figure 28 Selectivity to benzene cyclohexene and CO2 on 8MoAl (8Mo) and
2MoAl (2Mo) as a function of illumination time............................... 46
Figure 29 Ferrierite structure ...................................................................... 47
Figure 30 Thermogravimetric analysis of AFer and HFer........................... 49
Figure 31 Thermogravimetric analysis of NaY ............................................ 50
Figure 32 Thermogravimetric analysis of HY .............................................. 51
Figure 33 FT-IR spectra of Fer 5MoAFer and 20MoFer ....................... 52
Figure 34 FT-IR spectra of NaY and 20MoNaY........................................... 53
Figure 35 FT-IR spectra of HY and 20MoHY .............................................. 54

VI

Figure 36 Outlet reactor concentration of cyclohexane and carbon dioxide


as a function of run time...................................................................... 55
Figure 37 Outlet reactor concentration (a.u.) of benzene and cyclohexene on
5MoAFer as a function of run time ..................................................... 56
Figure 38 Cyclohexane conversion on 5MoAFer and 20MoAFer as a
function of illumination time ............................................................... 56
Figure 39 Benzene selectivity on 5MoAFer and 20MoAFer as a function of
illumination time.................................................................................. 57
Figure 40 Cyclohexene selectivity on 5MoAFer and 20MoAFer as a function
of illumination time.............................................................................. 57
Figure 41 Variation of specific surface area with MoO3 loading ................ 60
Figure 42 TG-MS results on DT2 sample..................................................... 61
Figure 43 TG-MS results on 2MoDT2 sample ............................................. 62
Figure 44 TG-MS results on 4MoDT2 sample ............................................. 63
Figure 45 TG-MS results on 8MoDT2 sample ............................................. 63
Figure 46 FT-IR spectra of DT2, 2MoDT2, 4MoDT2 and 8MoDT2 ........... 65
Figure 47 Raman spectrum of DT2 sample.................................................. 66
Figure 48 Raman spectra of DT2, 2MoDT2, 4MoDT2, 8MoDT2 and MoO3
samples ................................................................................................ 67
Figure 49 Cyclohexane conversion on DT2 as a function of illumination
time ...................................................................................................... 68
Figure 50 Carbon dioxide concentration on DT2 as a function of
illumination time.................................................................................. 69
Figure 51 Cyclohexane conversion on MoDT2s as a function of illumination
time ...................................................................................................... 70
Figure 52 Selectivity to benzene on MoDT2s as a function of illumination
time ...................................................................................................... 71
Figure 53 Selectivity to CO2 on MoDT2s as a function of illumination time72
Figure 54 Effect of light intensity on cyclohexane conversion..................... 74
Figure 55 FT-IR spectra of MoT0, MoT05, and MoT20 .............................. 78

VII

Figure 56 Raman spectra of MoT0, MoT05 and MoT20 samples ................ 79


Figure 57 Cyclohexane conversion on T0, T05 and T20 as a function of
illumination time.................................................................................. 80
Figure 58 Carbon dioxide concentration on T0, T05 and T20 as a function of
illumination time.................................................................................. 80
Figure 59 Mechanism for oxidation of cyclohexane on titania .................... 81
Figure 60 Cyclohexane conversion on Mo/T0, Mo/T05 and Mo/T20 catalysts
as a function of illumination time ........................................................ 82
Figure 61 Benzene selectivity versus cyclohexane conversion on Mo/T0,
Mo/T05 and Mo/T20............................................................................ 83
Figure 62 Carbon dioxide concentration on 2MoDT2, 4MoDT2 and
8MoDT2 as a function of illumination time......................................... 85
Figure 63 Cyclohexene conversion on 2MoDT2, 4MoDT2 and 8MoDT2 as a
function of illumination time................................................................ 87
Figure 64 Benzene concentration on 2MoDT2, 4MoDT2 and 8MoDT2 as a
function of illumination time................................................................ 87
Figure 65 Carbon dioxide concentration on 2MoDT2, 4MoDT2 and
8MoDT2 as a function of illumination time......................................... 88
Figure 66 Mechanism for oxidation of cyclohexene on bare titania ............ 89
Figure 67 Photocatalytic flat-plate reactor.................................................. 94
Figure 68 Cyclohexane conversion on MoDT2s catalysts as a function of
illumination time.................................................................................. 95
Figure 69 Selectivity to benzene on MoDT2s catalysts as a function of
illumination time.................................................................................. 96
Figure 70 Selectivity to CO2 on MoDT2s catalysts as a function of
illumination time.................................................................................. 96
Figure 71 Photocatalytic fluidized bed reactor .......................................... 100
Figure 72 Schematic picture of the cyclone................................................ 101
Figure 73 Schematic picture of the experimental apparatus...................... 102

VIII

Figure 74 Outlet reactor concentration of benzene on 4MoDT2 catalyst as


function of illumination time ............................................................. 103
Figure 75 Outlet reactor concentration (a.u.) of cyclohexane, oxygen
benzene and cyclohexene and as a function of run time.................... 104
Figure 76 Effect of reaction temperature on cyclohexane conversion and
benzene outlet concentration. UV sources: two eye mercury lamps,
125W.................................................................................................. 106
Figure 77 Comparison between cyclohexane conversion and benzene outlet
concentration obtained on Mo8 and Mo10. Reaction temperature:
120C. UV sources: two eye mercury lamps, 125 W......................... 108
Figure 78 Cyclohexane conversion as a function of illumination time on
8MoDT2 catalyst using the fluidized bed and the annular fixed bed
reactor. .............................................................................................. 109
Figure 79 Comparison of benzene outlet concentration on 8MoDT2 between
fixed bed and fluidized bed reactor. .................................................. 110
Figure 80 Total carbon mass balance in the fixed bed reactor as a function
of illumination time............................................................................ 112
Figure 81 TG-MS results on 4MoDT2 catalyst after photocatalytic results in
the fixed bed reactor.......................................................................... 113
Figure 82 TG-MS results on 8MoDT2 catalyst after photocatalytic results in
the fixed bed reactor.......................................................................... 113
Figure 83 Outlet reactor concentration (MS signal) of cyclohexene, and
benzene and (ppm) of carbon dioxide (NDIR analyzer) as a function of
temperarure on 4MoDT2 after activity measurements in the fixed bed
reactor. .............................................................................................. 114
Figure 84 Outlet reactor concentration (MS signal) of SO2 as a function of
temperature on 4MoDT2 after activity measurements in the fluidized
bed reactor......................................................................................... 115

IX

Figure 85 Difference of Mon+ contributions found by XPS for fresh 8MoDT2


catalyst and that recovered after photocatalytic test both in the fixed
bed reactor and in fluidized bed reactor ........................................... 116
Figure 86 Detailed XPS scan spectrum of S 2p on fresh 8MoDT2............. 117
Figure 87 Detailed XPS scan spectrum of S 2p on 8MoDT2 after
photocatalytic test in the fixed bed reactor........................................ 118
Figure 88 XPS scan spectrum of S 2p on 8MoDT2 after photocatalytic test in
the fluidized bed reactor .................................................................... 119
Figure 89 Schematic picture of the modified photocatalytic fluidized bed
reactor ............................................................................................... 120
Figure 90 Comparison between cyclohexane conversion obtained by using
two and four UV sources. Reaction temperature: 120 C. ................ 121
Figure 91 Comparison between cyclohexane conversion obtained by using
two and four UV sources. Reaction temperature: 120 C. ................ 121
Figure 92 Thermogravimetric curves of 8Mo2S, 8Mo4S and 8Mo6S catalysts
........................................................................................................... 123
Figure 93 Cyclohexane conversion and cyclohexene outlet concentration as
a function of SO42- percentage. Reaction temperature: 120 C. UV
sources: four eye mercury lamps, 125 W........................................... 124
Figure 94 Cyclohexane conversion and cyclohexene outlet concentration as
a function of reaction temperature .................................................... 125
Figure 95 Arrhenius plot for the photoreaction of cyclohexane on 8Mo2S 126
Figure 96 Ethylbenzene conversion and styrene outlet concentration as a
function of illumination time.............................................................. 129

Index of tables

Table 1 List of catalysts with their MoO3 nominal content .......................... 23


Table 2 Specific surface area and MoO3 amount of MoOx/Al2O3 catalysts . 38
Table 3 Characteristics of Na,K-Ferrierite.................................................. 48
Table 4 List of catalysts with their MoO3 nominal content .......................... 48
Table 5 Microporous volume of zeolites based samples .............................. 51
Table 6 List of catalysts with their MoO3 nominal contents ........................ 59
Table 7 List of catalysts and their characteristics ....................................... 60
Table 8 Hydroxyls and surface sulphates density ........................................ 64
Table 9 List of catalysts with their MoO3 nominal content .......................... 75
Table 10 List of catalysts and their characteristics ..................................... 75
Table 11 Hydroxyls and surface sulphates density ...................................... 77
Table 12 Geometrical parameters of the cyclone ...................................... 101
Table 13 Comparison between performance of 4MoDT2 and 8MoDT2
catalysts. Reaction temperature: 70 C. UV sources:two blacklight
blue, 160 W, PHILIPS ....................................................................... 104
Table 14 Effect of UV sources on 8MoDT2 catalyst. Reaction temperature:
100C................................................................................................. 105
Table 15 Emission characteristics in the UVA and UVB range................. 105
Table 16 Comparison between performance of 4MoDT2 and 8MoDT2
catalysts in the fixed bed reactor....................................................... 110
Table 17 Comparison between performance of 4MoDT2 and 8MoDT2
catalysts in the fluidized bed reactor................................................. 111

XI

Table 18 Catalysts and their characteristics.............................................. 122


Table 19 Nominal and effective sulphate content....................................... 124

XII

Abstract

Heterogeneous photocatalytic oxidation has been extensively studied for


environmental applications such as detoxification processes by total
oxidation of organics at room temperature both in water and in air. Until
now, few studies have been performed in order to obtain fine chemicals.
Starting from the discovery of the occurrence of the photocatalytic
oxidative cyclohexane dehydrogenation to benzene and cyclohexene nearly
at ambient temperature and pressure, this work has been dedicated to
improve the knowledge of this new type of photocatalytic reaction by
optimising catalyst formulation, photocatalytic process and operative
conditions
Some process aspects have been developed to explore its potential
application in an industrial process.
For this purpose, the selective photooxidation of cyclohexane has been
investigated in an annular gas-solid continuous flow reactor on molybdenum
based catalysts, by changing molybdenum loading, nature and surface
characteristic of the support.
Among several supports, such as zeolites, and alumina and titania,
highest photoactivity to dehydrogenated products was shown by MoOx/TiO2
catalysts.
In order to elucidate the effect of sulphate content on MoOx/TiO2 activity
and selectivity, cyclohexane photocatalytic oxidation has been also studied
on sulphated and unsulphated titania. These experimental results led to the
formulation of a catalyst highly selective for benzene production, although
with very low conversion of cyclohexane.
These experimental results have been confirmed employing a second type
of fixed bed reactor, a plate photoreactor designed at Bradford University.
In order to improve cyclohexane conversion and to reduce the extent of
the deactivation phenomena observed in fixed bed photoreactors, a
photocatalytic fluidized bed reactor was designed and realized, improving
the exposure of the catalysts to the light irradiation and assuring higher mass
transfer rates. Since titania belongs to the group C of Geldart classification,
it is not so easy to fluidize. Thus, to improve fluidization characteristics of

photocatalysts, new support obtained by sol-gel synthesis and physical


mixture with A group solid have been employed.
Thermodynamic analysis performed for the cyclohexane oxidative
dehydrogenation reaction indicated that temperature increase up to 330 C
gives increasing cyclohexane equilibrium conversion up to total conversion.
Experimental tests performed in the fluidized bed photoreactor by
increasing temperature and light intensity, led to a significant increase in
cyclohexane conversion of about ten times.
On the basis of the photocatalytic activity tests and physical and chemical
characterization tests, a mechanism of oxidative dehydrogenation of
cyclohexane to benzene and cyclohexene has been proposed.
Deeper knowledge of the reaction mechanism led to a formulation of
innovative photocatalysts (sulphated MoOx/-Al2O3) selective for oxidative
mono-dehydrogenation of cyclohexane.
Finally, the effectiveness of ethylbenzene to styrene conversion has been
also checked.

14

Introduction

Volatile organic compounds (VOCs) are widely used in (and produced


by) both industrial and domestic activities Wolf et al. (1991). This
extensive use results in their occurrence in aquatic, soil and atmosphere
environments Shah and Singh (1988). Many VOCs are toxic, and some are
considered to be carcinogenic, mutagenic, or teratogenic Wilkinson
(1987). However, the most significant problem related to the emission of
VOCs is centred on the possible production of photochemical oxidants, for
example ozone and peroxyacetyl nitrate Japer et al. (1991). Tropospheric
ozone, formed in the presence of sun light from NOx and VOC emissions, is
toxic to humans, damaging to crops and is implicated in the formation of
acid rain Cortese (1990), Fisherman (1991). Emissions of VOCs also
contribute to localized pollution problems of toxicity and odor. Many VOCs
are implicated in the depletion of the stratospheric ozone layer and may
contribute to global warming.
As a result of all these problems, VOCs have drawn considerable
attention in the last decade. Approximately 50% of the U.S. Enviromental
Protection agency (EPAs) list of priority pollutants is composed of VOCs.
The Clean Air Act of 1990 calls for a 90% reduction in the emission of 189
toxic chemicals over the next 8 years, 70% of these being VOCs Armor
(1992). Therefore, there is currently a great deal of interest in developing
processes which can destroy these compounds, and since a large number of
the VOCs are oxidizable, chemical oxidation process can be looked upon as
a viable method. The application of heterogeneous catalytic oxidation
technology to air pollution control is well-established. Examples are
automotive exhaust treatments Armor (1992) and catalytic incineration
Kosuko and Nunez (1990). In general, these catalysts operate at elevated
temperatures.
Photocatalytic oxidation of organic compounds in the gas-phase using
TiO2 as a catalyst appeared to be a promising process for remediation of air
and ground-water polluted by VOCs. Heterogeneous photocatalysis using
TiO 2 has several attractions:

Chapter I

(a) TiO2 is relatively inexpensive, (b) it dispenses with the use of other
coadjutant reagents, (c) it shows efficient destruction of toxic contaminants,
(d) it operates at ambient temperature and pressure and (e) the reaction
products are usually CO2 and H2O, or HCl, in the case of chlorinated organic
compounds (Figure 1).

Figure 1 Schematic representation of a photocatalytic process

I.1

Heterogeneous photocatalysis

The term photocatalysis is still the subject of some debate. For


example, it is argued Suppan (1994) that the idea of a photocatalysed
reaction is fundamentally incorrect, since it implies that, in the reaction, light
is acting as a catalyst, whereas it always acts as a reactant which is
consumed in the chemical process. In reality the term photocatalysis is in
widespread use and is here to stay; it is not meant to, nor should it ever be
used to, imply catalysis by light, but rather the acceleration of a
photoreaction by the presence of a catalyst. The term photoreaction is
sometimes elaborated on as photoinduced or photoactivated reaction, all
to the same effect.
The above definition of photocatalysis includes the process of
photosensitization, a process by which a photochemical alteration occurs
in one chemical species as a result of the initial absorption of radiation by
another chemical species called the photosensitizer. It follows from the
above that heterogeneous photocatalysis involves photoreactions which
occur at the surface of a catalyst. If the initial photoexcitation process occurs
in an adsorbate molecule, which then interacts the ground state of the

Introduction

catalyst substrate, the process is referred to as a catalysed photoreaction.


If, on the other hand, the initial photoexcitation takes place in the catalyst
substrate and the photoexcited catalyst then interacts with the ground state
adsorbate molecule, the process is a sensitized photoreaction. In most
cases, heterogeneous photocatalysis refers to semiconductor photocatalysis
or semiconductor-sensitized photoreactions.

I.1.1

Basic Principles

The overall process of semiconductor-sensitized photoreactions can be


summarized as follows:
semiconductor

A+ D

A + D +

light E bg

where Ebg is the band gap energy of the semiconductor. If, in the absence
of semiconductor and light of energy greater than or equal to Ebg, G0 for
reaction (1) is negative, the semiconductor-sensitized photoreaction is an
example of photocatalysis Bard (1980). Alternatively, if G0 for reaction
(1) is positive, the semiconductor-sensitized photoreaction is an example of
photosynthesis Bard (1980).
As indicated in Figure 2, for many compounds, as the number N of
monomeric units in a particle increases the energy necessary to photoexcite
the particle decreases. In the limit when N>>2000, it is possible to end up
with a particle which exhibits the band electronic structure of a
semiconductor, as illustrated in Figure 2, in which the highest occupied band
(the valence band) and lowest unoccupied energy band (the conductance
band) are separated by a bandgap Ebg, a region devoid of energy levels in a
perfect crystal.

Chapter I

N=1

N=2

N=6

N>>2000

Figure 2 Change in the electronic structure of a semiconductor compound


as the number N of monomeric units increases from unity to clusters of more
than 2000
The basic principles of heterogeneous photocatalysis can be summarized
shortly as follows Litter (1999). A semiconductor (SC) is characterized by
an electronic band structure in which the highest occupied energy band,
called valence band (vb), and the lowest empty band, called conduction band
(cb), are separated by a bandgap, i.e. a region of forbidden energies in a
perfect crystal. When a photon of energy higher or equal to the bandgap
energy is absorbed by a semiconductor particle, an electron from the vb is
promoted to the cb with simultaneous generation of a hole (h+) in the vb.
The electron in the cb and the h+ in the vb can recombine on the surface or
in the bulk of the particle in a few nanoseconds (and the energy dissipated as
heat) or can be trapped in surface states where they can react with donor (D)
or acceptor (A) species adsorbed or close to the surface of the particle.
Thereby, subsequent anodic and cathodic redox reactions can be initiated
(Figure 3). The energy level at the bottom of the cb is actually the reduction
potential of photoelectrons and the energy level at the top of the vb
determines the oxidizing ability of photoholes, each value reflecting the
ability of the system to promote reductions and oxidations.

Introduction

Figure 3 Illustration of the main processes occurring on a semiconductor


particle under electronic excitation
The flatband potential, Vfb, locates the energy of both charge carriers at
the SCelectrolyte interface, and depends on the nature of the material and
the system equilibria Serpone (1997). From a thermodynamic point of
view, adsorbed couples can be reduced photocatalytically by cb electrons if
they have redox potentials more positive than the Vfb of the cb, and can be
oxidized by vb holes if they have redox potentials more negative than the Vfb
of the vb. The efficiency of a photocatalyst depends on the competition of
different interface transfer processes involving electrons and holes and their
deactivation by recombination Serpone (1997), Hoffman et al. (1995), Fox
and Dulay (1993), Linsebigler et al.(1995). The position of the flatband of
an SC in solution follows a Nernstian pH dependence, decreasing 59mV per
pH unit Ward (1983), and consequently, the ability of electrons and holes
to enact redox chemistry can be controlled by changes in the pH.
By using titania as semiconductor, the heterogeneous photocatalytic
process is a complex sequence of reactions that can be expressed by the
following set of simplified equations:

Chapter I

A great discussion exists nowadays about the oxidative pathway, which


could be performed by direct hole attack or mediated by HO radicals, in
their free or adsorbed form. The oxidative pathway leads, in many cases, to
complete mineralization of an organic substrate to CO2 and H2O. Generally,
A is dissolved O2, which is transformed in superoxide radical anion (O2)
and can lead to the additional formation of HO:

The acceptor A can also be a metal ion species having a convenient redox
potential to be transformed into a different oxidation state:
Some oxide and chalcogenides have enough bandgap energies to be
excited by UV or visible light, and the redox potentials of the edges of the vb
and cb can promote a series of oxidative or reductive reactions.
From the available semiconductors, ZnO is generally unstable in
illuminated aqueous solutions, especially at low pH values, and WO3,
although useful in the visible range, is generally less photocatalytically
active than TiO2. Among others, CdS, ZnS and iron oxides have been also
tested. However, and without any doubt, TiO2 is so far the most useful
material for photocatalytic purposes, owing to its exceptional optical and
electronic properties, chemical stability, non-toxicity and low cost.
TiO2 exists in three main crystallographic forms, brookite, anatase and
rutile and it is the most widely used semiconductor. The energy bandgaps of
anatase (3.23 eV, 384 nm) and rutile (3.02 eV, 411 nm) combine with the vb
positions to generate highly energetic holes at the interface, giving rise to

Introduction

easy oxidation reactions. Anatase has been found, in most of the cases, to be
photocatalytically more active than rutile. The most popular commercial
form of TiO2 is produced by the German company Degussa under the name
P-25; this sample contains around 80% anatase and 20% rutile and possesses
an excellent activity. Recently, values for the Vfb of the cb and vb of
Degussa P-25 have been calculated as 0.3 and +2.9V (pH = 0), respectively
Martin et al.(1994).
With regards to kinetic aspects, the rate of heterogeneous photocatalytic
processes depends on the global reaction resistance but also on the
concentration of photoproduced electron-hole pairs Palmisano and Sclafani
(1997). The concentration of these latter depends on the intensity of the
radiation of suitable energy impinging on the system and on their
recombination rate. When the maximum concentration of the pairs has been
achieved (steady state), the rate depends on several factors, such as
electronic, chemical and morphological properties of semiconductor,
presence of additives in the reacting system, donor acceptor and acid-base
properties of the solution and of the solid, temperature (high temperatures
generally lead to higher rates because they provoke a more frequent collision
between the substrate and the SC) and pressure. The adsorption of reagents,
the charge transfer from and to reagents and the desorption of the products of
the photoreaction are essential kinetic steps and their role should be
evaluated every time as in catalysis. For example an augmentation to the rate
of oxygen photo-uptake by a TiO2 surface can be created if the surface is
pre-exposed to the vapour of a strongly adsorbing species, such as an
aliphatic alcohol (ROH) Bickley (1997). Alcohols adsorb very strongly up
to monolayer coverage on TiO2 and the adsorbed species (e.g., RO- (ads))
can themselves act as very efficient hole traps. In these circumstances the
rate of uptake of oxygen by the surface under band-gap energy illumination
is greatly enhanced and attains a pseudo-zero order character. In these cases
the surface becomes covered progressively with the oxidised form of the
alcohol (ethanal when ethanol is used; propanone when propan-2-ol is used).
Accordingly, a study of strongly preadsorbed organic species in the
presence only of a pure gaseous species such as O2 (g) is strictly a precursor
to sustained photocatalytic oxidation where, with the availability also of
excess of vapour or liquid phase reagent, a photocatalytic oxidation can
occur with the progressive consumptions of both O2 (g) (or O2 dissolved)
and the organic reagent e.g., propan-2-ol. In many examples, the kinetic of
the photocatalytic reaction are controlled uniquely by the photo-electronic
events induced in the solid in the copresence of the reacting partners.

I.2

Fundamental Engineering Aspects

Photocatalytic reactions are the result of the interaction of photons having


the appropriate wavelength with a semiconductor. When the arriving light

Chapter I

has an energy equal or greater than the semiconductor band gap, radiation is
absorbed and electrons are moved from the valence band to the conduction
band giving rise to the formation of electronhole pairs. These charge
carriers can migrate to the catalyst surface in competition with an exothermic
and normally fast recombination reaction. When they reach the
semiconductor surface they may, once more recombine, or participate in
successive chemical reactions Cassano and Alfano (2000). The main
components of a photocatalytic process are indeed the photoreactor and the
radiation sources Augugliaro et al. (1997). For thermal and catalytic
processes the parameters that affect reactors performance are:
1. the mode of operation;
2. the phases present in the reactor;
3. the flow characteristics;
4. the needs of heat exchange;
5. the composition and the operative conditions of the reacting
mixture.
For selecting the type of heterogeneous photoreactor additional
parameters must be considered since photons are the primary source for
the occurrence of photoreaction. The selection of the construction
material for the photoreactor must be generally done in order to allow
the penetration of radiation into the reacting mixture. The choice of the
radiation source must be made by considering that the absorbed radiation
energy should be equal to or higher than the band gap.

I.2.1

Types of Radiation Sources

Different types of lamps allow generation of radiation with different


ranges of wavelengths. As reported before, the choice of a particular lamp is
made on the basis of the reaction energy requirements Augugliaro et al.
(1997). There are four many types of radiation sources:
a. arc lamps;
b. fluorescent lamps;
c. incandescent lamps;
d. lasers.
In arc lamps the emission is obtained by a gas activated by collisions with
electrons accelerated by an electric potential. Typical activated gases are
mercury and/ or Xenon vapours.
In fluorescent lamps the emissions is obtained by exciting an emitting
fluorescent substance, deposited in the inner side of a cylinder, by an electric
discharge occurring in the gas filling the lamps. Generally these lamps emit
in the visible region, but the actinic type ones have emission in the near
UV-region. Of course the emission spectrum depends on the nature of the
mixture of fluorescent substances user. Their power is quite small (up to
150W). In incandescent lamps, the emission is obtained by heating at very

Introduction

high temperature suitable filaments, of various nature, by current circulation.


A typical emission spectrum of an actinic lamp is reported in Figure 4.

spectral irradiance, W/m2

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
-0.1
250

300

350

400

450

wavelenght, nm

Figure 4 Emission spectrum of a 40W actinic lamp

I.2.2

Photoreactor Materials and Geometries

In order to allow the penetration of radiation into the photoreactor, the


photoreactor wall between the lamp and the reacting mixture must be made
of a material transparent to the radiation. If near-UV radiation is utilised by
the photoreacting mixture, Pyrex glass can be used but if UV radiation is
needed, quartz glass must be used. It must be noted that, by increasing the
thickness of photoreactor walls, the light transmission decreases so that the
size of reactor and its operative conditions (temperature and pressure) are not
independent variables. In order to utilise all the radiation emitted by the
lamp, the transmission or absorption of radiation from photoreactor walls is
avoided by wrapping them with reflecting surfaces such as aluminium foils.
For photoreactors, the geometry and the spatial relation between reactor
and light source are most important. In fact geometry plays an important role
in determining reaction yields as well as reactor operability. The geometrical
configuration of a photoreactor is usually chosen to obtain the maximum
benefit from the pattern of irradiation taking into account shape and cooling
requirements of available commercial lamps useful for wavelength required
by the reaction.

Chapter I

The most used photoreactor geometries are:


a.
cylindrical;
b.
parallel plate;
c.
annular.
The irradiation may be normal or parallel to the reactor surface. In
selecting the reactor geometry configuration, it is necessary to determine the
optical path of the light which will be obtained within the reactor. It is, in
fact, the most important factor affecting the absorption of radiation by the
reaching mixture and therefore it determines the efficiency of the
photoprocess.

I.2.3

Efficiency of Heterogeneous Photocatalytic Systems

In heterogeneous photocatalysis, quantum yield () has come to define


the number of molecules converted per unit time relative to the number of
photons absorbed per unit time as follows:

transformed molecules
unit time
=
absorbed photons
unit time

(1)

In a heterogeneous photocatalytic process, where a solid is involved, the


rates must be referred to the active sites. The difficulty of determining the
number of the active sites is generally overcome by using the BET surface
area of the particle instead of the active sites; the implicit assumption of this
simplification is that the number of active sites is proportional to the surface
area. For the occurrence of a photoreaction it is necessary that photons of
suitable energy are absorbed by the semiconductor, which usually is in the
form of polycrystalline porous particle. A first quantity that must be
measured is therefore the rate of photons absorption (rpa), defined as:

rpa =

absorbed photons
time surface area

(2)

Once radiation is absorbed, electron-hole pairs are generated if the photon


energy equals or exceeds the semiconductor band-gap energy. The pairs can
generate thermal energy (recombination), determine a lattice conversion
(photocorrosion) or be trapped by suitable surface species and initiate a
reaction sequence with adsorbed surface molecules (photocatalysis). On this
ground a second quantity that must be measured is the specific reaction rate
(srr), defined as:

10

Introduction

srr =

reacted molecules
time surface area

(3)

The values may be determined, knowing the rpa and srr parameters,
by the following equation:

I.2.4

srr
rpa

(4)

Modeling of Photocatalytic Reactors

The modeling of photocatalytic reactors requires a complex analysis of


the radiation field in the photoreactor Cassano et al. (1995). This analysis,
linked to the modeling of the fluid-dynamics and the reaction kinetics,
results in integro-differential equations which almost invariably require
demanding numerical computations. Further advances of photocatalytic
oxidation on an industrial scale will be facilitated by the availability of
simpler mathematical models that retain the essential elements of a rigorous
model and that can be easily used for scale-up and design.
Figure 5 shows a schematic representation of the modeling of a
photocatalytic reactor. The development of a reactor model requires the
inclusion of a number of sub-models. These are a radiation emission model,
a radiation absorption-scattering model, a kinetic model and a fluid-dynamic
model.
Radiation emission
model

LVRPA

Fluid-dynamic model

Kinetic model

Material balance

Reactor model

Radiation absorptionscattering model

Figure 5 Schematic representation of the modeling of a photocatalytic


reactor
The central aspect of the modeling procedure is the calculation of the
Local Volumetric Rate of Photon Absorption (LVRPA) at each point of the
reaction space, which requires solving the radiative transfer equation (RTE)
in the reaction space. Due to the complex nature of radiation scattering, this
results in a set of integro-differential equations, which require demanding
numerical computational efforts. In practice, combining a simplified
radiation emission model of the light source with a simplified radiation
absorption-scattering model in the reaction space and performing a radiation

11

Chapter I

balance in the reaction space can simplify the RTE.


The above scheme assumes that the useful photons of a given
photocatalytic reactor, i.e. those photons with energy higher than the bandgap of the semiconductor photocatalyst, are absorbed by the solid
photocatalytic particles only. This assumption removes the interdependence
of the progress of the reaction and the attenuation of the radiation because
the absorbing species do not undergo changes in concentration.
Consequently, the incident radiation flux becomes a function of the reactor
position only and can be obtained independently of the information provided
by the material balance equation.
Once the LVRPA has been calculated, this is normally substituted into
the kinetic equation and into the material balance equation which when
solved with suitable boundary conditions, yields the concentration of a
generic substrate at the reactor outlet.
Three approaches have been proposed in the literature for the calculation
of the LVRPA: 1) The rigorous approach which involves the mathematical
solution of the RTE, although its integro-differential nature makes this
approach considerably complex Cassano et al. (1995); 2) The numerical
approach which involves the Monte Carlo simulation of the radiation field
in the photoreactor, a simple but also a computationally demanding
procedure Pasquali et al. (1996); and 3) the simplified approach which
models the radiation field in the photoreactor using two-flux Raupp et al.
(1997), Brucato and Rizzuti (1997), or six-flux Li Puma et al. (2004)
radiation absorption-scattering models. The two and six-flux models yield a
sensible representation of the LVRPA in the reaction space and allow a
considerable simplification of the mathematical model. Finally the scattering
properties of the photocatalyst, and the geometrical configuration of the
photoreactor, determine in large extent the degree of complexity of a
mathematical model Li Puma et al. (2005).

I.3

Aim of the Work

Heterogeneous photocatalysis has been widely investigated as a novel


technique for environmental applications such as detoxification processes
both in water and in air. In contrast, very few studies have been conducted
concerning photocatalysis as potential technology for the synthesis of
chemicals from hydrocarbon feedstocks. Particularly, photocatalysed
selective oxidation of hydrocarbons such as cyclohexane with gaseous
oxygen is still a challenge.
Concerning this last aspect, in this work photocatalytic selective
oxidation of cyclohexane on molybdenum based catalysts has been
investigated by changing Mo loading, nature and surface characteristic of the
support.

12

Introduction

The photocatalytic activity and selectivity of MoOx supported catalysts


on TiO2, and Al2O3, have been determined with an annular gas-solid
continuous flow reactor. Photocatalytic test were also carried out on MoOx
supported NaY, HY zeolites and Ammonium ferrierite (AFer).
In order to elucidate the effect of sulphate content on MoOx/TiO2 activity
and selectivity, cyclohexane photocatalytic oxidation has been also studied
on sulphated and unsulphated titania.
Activity and selectivity of MoOx/TiO2 catalysts have been also measured
by a photocatalytic plate reactor.
Finally, a photocatalytic fluidized bed reactor has been designed and
realized in order to improve both exposure of the catalysts to light irradiation
and a good contact between reactants and catalyst.

13

II State of the art

II.1 Photocatalysis as potential technology for synthesis of


chemicals from hydrocarbon feedstocks
Over the past 10 years, the semiconductor TiO2 (Ebg = 3.2 eV) as a
photocatalyst has become the focus of numerous studies owing to its
attractive characteristics and applications in the treatment of environmental
contaminants both in water streams than in air streams Yu et al. (1997).
Complete mineralization of a broad variety of organic compounds containing
unsaturated bonds by TiO2 photocatalysis has been reported Gratzel.
(1983).
The photodecomposition of chlorinated Philips and Raupp (1992), Liu et
al. (1997) and fluorinated Ohtani et al. (1990) organics, hydrocarbons
Fu et al. (1996), aldehydes, ketones, alcohols Blake and Griffin (1988),
Sauer and Ollis (1996), Peral and Ollis (1992), aromatics Fu et al. (1995)
or trichloroethylene Dibble and Raupp (1990), Dibble and Raupp (1992)
over UV-irradiated titanium dioxide is well documented in literature. A
drawback of these processes is the formation of partial oxidation products,
sometimes resulting in increased toxicity of the treated stream.
Few studies have been conducted on photocatalysis as potential
technology for synthesis of chemicals from hydrocarbon feedstocks.
Significant examples of photocatalytic processes employed for synthetic
purposes are oxidation and reduction processes, isomerization reactions, C-H
bond activations, and C-C and C-N bond-reforming reactions Maldotti et
al. (2002). For these purposes heterogeneous photocatalysis processes seem
to be advantageous since the use of solar light as a reagent in oxidative
catalysis is particularly relevant to realizing innovative and economically
advantageous processes for conversion of hydrocarbons into oxygenates and,
at the same time, to move toward a sustainable chemistry that has a
minimal environmental impact. The second key reagent employed in the
oxygenation processes considered herein in the molecule of O2. With regard
to this last aspect, it is important the search for new catalysts capable of

Chapter II

inducing the oxofunctionalization of hydrocarbons represents a major target


from the synthetic and industrial points of view Roby and Kingsley
(1996). On the basis of pure thermodynamic considerations, most organic
compounds are not stable with respect to oxidation by O2. There are
however, kinetic limitations. The activation of both O2 and the organic
substrate may be achieved by photochemical excitation with light of the
visible or of the near-ultraviolet regions ( > 300 nm). Moreover, selectivity
is a key issue in the catalysis of fine chemistry. To purpose this objective, all
the steps of the oxidation process must be optimized. For example the use of
heterogeneous and organized systems is a suitable way to control efficiency
and selectivity of catalytic processes through the control of the microscopic
environment surrounding the catalytic center. In particular the nature of the
reaction environment may affect numerous physical and chemical
functionalities of the photocatalytic system such as the absorption of light,
the generation of elementary redox intermediates, the rate of competitive
chemical steps, and the adsorption-desorption equilibria of substrates,
intermediates, and final products.
In the seventies pioneer works reported photo-oxidation of alkanes and
alcohols in heterogeneous gas-solid reactors Walker et al. (1977), Djeghri
and Teichner (1980), Bickley et al. (1973), Formenti et al. (1971). Alkanes
have been oxidised on anatase TiO2 achieving 75% selectivity to acetone at
3% isobutane conversion, 30 % selectivity to butanone from n-butane, but
complete oxidation from 1-butene and 2-butene. In more recent works, few
studies employed gas-solid reactors.
Ethanol was selectively converted to acetaldehyde and formaldehyde on
TiO2. Muggli and Falconer showed that selectivity of ethanol photocatalytic
oxidation is changed as effect of titania sites poisoning by adsorption of
reactants Muggli and Falconer (1998). This seems suggest a way to design
a selective catalyst for photocatalytic oxidation.
The studies of Frei on photocatalysed oxidation of hydrocarbons in
zeolite cages indicate a potential way to achieve high selectivity in gas-solid
systems. Very impressive results (100% selectivity to aldehydes and ketones
from the corresponding alkanes, cycloalkanes, alkenes, aromatics) are
overviewed Blatter et al. (1998).

II.1.1

Photocatalytic oxidation of cyclohexane

The oxidation of cyclohexane to form cyclohexanone is significantly


important reaction since cyclohexanone is an intermediate material to caprolactone which is a raw material for nylon synthesis. The oxidation
process of cyclohexane to produce cyclohexanone has been industrialized
over cobalt-base catalyst with oxygen above 150 C under high pressure
Davis and Kemp (1991). To make the reaction condition milder, new
reaction system has been sought out. Hydrogen peroxide is often used as an

16

State of the art

oxidizing reagent to achieve the mild reaction condition Carvalho et al.


(1999), Armengol et al. (1999). Nevertheless, realization of oxidation by
molecular oxygen is strongly desired.
The most studies of cyclohexane photocatalytic oxidation deal with slurry
systems. Cyclohexanol, cyclohexanone, and polyoxygenates have been
obtained on silica supported vanadia or polyoxytungstate catalysts Molinari
et al. (1998), Giannotti and Richter (1999). It is proposed that
cyclohexanone is obtained from cyclohexane through a light induced radical
mechanism resulting in the selective formation of ketone due to a
confinement effect inside the zeolite cage Blatter and Frei (1998).On
titania cyclohexane photo-oxidation in dichloromethane leads to the
formation of cyclohexene traces Almiquist and Biswas (2001).
Gas-solid heterogeneous photocatalytic oxidation of cyclohexane and
cyclohexene in humidified air was studied at 30 C on P-25 titania powder
Einaga et al. (2002). Deep oxidation to CO2 was essentially obtained
together with formation of carbon deposits on the surface, which is
responsible for catalyst deactivation.

II.1.2

Oxidative dehydrogenation of cyclohexane

Potential useful intermediates and monomers for polyadditions (such as


cyclohexene and cyclohexadiene), together with benzene can be produced by
cyclohexane oxidative dehydrogenation reaction. Cyclohexane is available
in large amounts in naphthas, so it can be recovered from them although
with quite demanding procedures. By means of a refinery distillation tower
(named benzene heartcut tower) where cyclohexane is present in the side
fraction, a benzene-rich fraction (50% of benzene) is separated from a heavy
gasoline bottom and a light gasoline head fraction Blomberg et al. (2002).
This side fraction is used to provide benzene for petrochemical processes.
Due to the important role of benzene as an intermediate in petrochemistry,
processes for conversion of low-value hydrocarbons into benzene, such as
aromatization of light alkanes Nishi et al. (2002), are under development.
In this context, the conversion of cyclohexane, either recovered from the
benzene heartcut tower side fraction, or still in mixture with benzene, can be
a way to enhance the production of benzene and to fulfil the need for
benzene in some cases.
The occurrence of oxidative dehydrogenation of cyclohexane to
cyclohexene and benzene on MoO3 or MoO3/-Al2O3 catalysts in the
temperature range 280-417 C has been reported Alyea and Keane (1996).
Selectivity to dehydrogenated products is influenced by side-reactions such
as combustion or cracking, leading to a heavily selectivity decrease at
increasing temperature. A recent paper investigates the thermal production
of benzene from cyclohexane in vapor phase over several catalysts, such as
V2O5/SiO2, V2O5-Nb2O5/SiO2, Ce-,V-, Fe-phosphates, H-ZSM5, Co-ZSM5

17

Chapter II

Panizza et al. (2003). It was found that gas-phase reaction occurs above
277 C and selectivity to benzene was positively influenced by the presence
of water vapor.
Very recently, heterogeneous oxidative dehydrogenation of cyclohexane
to benzene was found for the first time to occur under photocatalytic
conditions. Initial cyclohexane conversion of 38% and maximum selectivity
to benzene of 41% were achieved on a MoOx/TiO2 catalyst in the presence
of gaseous oxygen at temperature of 35 C under UV irradiation in a gassolid continuous flow reactor. The main by-product was carbon dioxide
Ciambelli et al. (2002).

II.2 Influence of catalyst surface acidity on photocatalytic


performances
Titanium dioxide is widely used in heterogeneous photocatalysis as a
semiconductor photocatalyst because of its long-term stability, no toxicity
and good, often the best, photocatalytic activity. However, TiO2 is active
only under ultraviolet (UV) light, so in recent years, transition metal ions
doping has been widely performed by chemical synthesis and other methods,
in order to improve photoactivity Legrini et al. (1993). Karakitsou and
Verykios (1993) showed that doping with cations having a valence higher
than +4 can increase the photoactivity, whereas Mu et al. (1989) reported
that doping with trivalent or pentavalent metal ions was detrimental to the
photoactivity even in the UV region. Furthermore, according to a systematic
study on the photoactivity and transient absorption spectra of quantum-sized
TiO2 doped with 21 different metals the energy level and d-electron
configuration of the dopants were found to govern the photoelectrochemical
process in TiO2 Yamashita et al. (1999). Even though the effects of metal
doping on the activity of TiO2 have been a frequent topic of investigation, it
remains difficult to make general conclusion.
The presence of sulphate and metal oxides as dopant of titania has been
reported to enhance the photooxidation reactivity of several organic
compounds. It was shown Yu et al. (2002), Muggli and Ding (2001) that
TiO2 treatment with sulphuric acid could increase its photoactivity, although
the mechanism was not clarified. Kozlov et al. (2003) reported that the
treatment of TiO2 with sulphuric acid enhances the photocatalytic activity in
acetone oxidation by 2030%. Colon et al. (2003), studying the
photocatalytic degradation of phenol on sulphated TiO2 in a slurry reactor,
evidenced that the improvement of the photocatalytic activity is related to
the optimisation of the redox step in the photocatalytic process instead of the
acidity properties that could favour the adsorption of the organic substrate.
Fu et al. (1999) studied the structure of SO42-/TiO2 and its activity for room
temperature photocatalytic oxidation (PCO) of CH3Br, C6H6, and C2H4 in
air. For catalysts calcined at 723 K, conversion of CH3Br over SO42-/TiO2

18

State of the art

was six times higher than over TiO2. Moreover, TiO2 deactivated faster than
SO42-/TiO2; after 6 h of PCO, SO42-/TiO2 did not deactivate, whereas on TiO2
conversion decreased from 88 to 20% for C6H6 and from 60 to 12% for
CH3Br. They concluded that the improved rate for SO42-/TiO2 was due to a
greater surface area as well as a larger fraction of the anatase phase of TiO2,
which is more active than rutile for PCO.
Doping TiO2 with metal oxides such as WO3, MoO3, and Nb2O5 increases
surface acidity and PCO activity of TiO2. Cui et al. (1995) studied the
activity of Nb2O5/TiO2 during PCO of 1,4-dichlorobenzene. They found that
surface acidity increased and photocatalytic activity doubled when niobium
oxide was deposited on titania up to a monolayer. However, higher niobium
loading did not increase surface acidity and PCO photocatalytic activity due
to Nb2O5 segregation. They concluded that the same structural feature that
enhances surface acidity increases PCO rate. Similarly, the authors found
correlation between PCO activity and acidity for TiO2 doped with WO3,
MoO3, or Nb2O5 Papp et al. (1994), Okasaki and Okuyama (1983), Lee et
al. (1997).

II.3

Photocatalytic fluidized bed reactors

For photocatalytic degradation processes, two methods of TiO2


application are favoured: (1) TiO2 suspended in aqueous media and (2) TiO2
immobilized on support materials. Fluidized beds are known as good
chemical reactors due to excellent reactants contact, high mass and heat
transfer rate and easy to control the reaction temperature. When a
photocatalytic reaction takes place in a gassolid reactor, it is necessary to
achieve both exposures of the catalysts to light irradiation and a good contact
between reactants and catalyst.
It is believed that fluidized bed can take advantages of better use of light,
easy of temperature control, and good contacting between target compound
and photocatalysts over slurry reactors or fixed bed reactors with
immobilized TiO2.
Activity increase of fluidized bed is partially associated with higher light
absorption due to utilization of scattered light by the catalyst. Figure 6
illustrates light scattering in the three types of beds typically used.

19

Chapter II

Figure 6 Schematic representation of light scattering in (A) film fixed bed,


(B) granular fixed bed and (C) fluidized bed
In the wavelength range of UVA light used, titania particles scatter a part
of incident light. In the film bed, the scattered light never meets catalyst
again. In granular fixed bed, some part of the scattered light meets the
catalyst and is absorbed or scattered again. In fluidized bed, the probability
for collisions of scattered light with titania particles is the highest
Vorontsov et al. (2000).
The fluidized bed of ultrafine particle of photocatalyst was applied to
treat nitrogen oxides (NOx) by photocatalytic oxidation Matsuda et al.
(2001). Three different TiO2 particles with primary particle diameters of 7,
20 and 200 nm were used as the bed material. The fluidized bed of
aggregates of 7 nm crystallites TiO2 exhibited high removal efficiency of
NOx because of its large specific surface area. It was found that the amount
of NOx removal is proportional to the specific surface area. Agglomerates of
7 and 20 nm particles appeared to be so hard that they were not destroyed
during fluidization because of large adhesion forces of particles. In the case
of 7 and 20 nm particle systems, the bed height was found to increase
progressively with the increase in gas velocity, while the bed expansion is
observed to level off for 200 nm particle system. The entrainment rates of 7
and 20 nm particle systems were found to be smaller than that of 200 nm
particle system.
The effects of CuO loading on titania support, reaction temperature, and
surface gas velocity on the photocatalytic reduction of NO have been
determined in an annular flow type and a modified two-dimensional
fluidized bed photoreactor. The optimum CuO loading was found to be 3.3
wt% and NO degradation conversion in the modified two-dimensional
fluidized bed photoreactor was more than 70% at 2.5 times the minimum
fluidization velocity, Umf Lim et al. (2000). The decomposition of NO by
photocatalysis increased with decreasing initial NO concentration and
increasing gas-residence time and the reaction rate increased with increasing
UV light intensity. The photocatalytic oxidation of ethanol vapour was
investigated with an annulus fluidized bed reactor Kim et al. (2004) of

20

State of the art

silica gel powder coated TiO2 catalyst prepared by the sol-gel method. The
UV lamp was installed at the center of the bed as the light source. It was
found that at 1.2 Umf value, about 80% of ethanol with initial
concentration.of 10000 ppm was decomposed and the increase of superficial
gas velocity reduced the reaction rate significantly.
A loading type of Eu/Ti/Si catalyst was synthesized with Eu(NO3)3,
Ti(OC4H9)4 and porous silica by sol-gel method Ping et al. (2004).
Benzene and its compounds were degraded in a three-phase fluidized bed
photo-catalytic reactor and the reaction conditions for photo-catalytic
degradation were investigated. The results indicated that comparing the
binary catalyst Ti and Si and the triple compound catalyst of Eu/Ti/Si, the
latter was more efficient. The photo-catalysis removal efficiency for benzene
and its compounds would reach over 98%, and that was 10-20% higher than
that as Ti/Si catalyst was used. Photocatalytic NH3 synthesis was
successfully performed in a fluidized reactor of parallel-wall configuration
by irradiating Fe-doped TiO2 with near UV light Yue et al. (1983). The
catalyst was prepared in such a way that good quality of fluidization was
obtained. Mixing the catalyst with -alumina affected the fluidization
behavior. NH3 production was increased when the catalyst was suitably
fluidized because of enhanced utilization of light energy.

21

III Experimental Results:


Photocatalytic oxidation of
cyclohexane on MoOx/Al2O3

III.1

Samples preparation

All chemicals used in the experiments were HPLC grade obtained from
Aldrich Co. Al2O3 (Al), (Aldrich) and Al2O3 (Al), (Puralox SBA
150, SASOL S.p.A.), were impregnated with an aqueous solution of
ammonium heptamolybdate (NH4)6Mo7O244H2O.
Powder samples were dried at 120 C for 12 hours and calcined in air at
400 C for 3 hours.
In Table 1 the list of prepared catalysts with their nominal MoO3 content
is reported.
Table 1 List of catalysts with their MoO3 nominal content
Catalyst
Al
2MoAl
Al
8MoAl

III.2

Nominal MoO3 content, wt%


2.0
8.0

Catalysts characterization

To characterize the samples studied in this work the following techniques


were used:

Inductively coupled plasma-mass spectrometry (ICP-MS);


Thermal analysis (TG-MS);

Chapter III

Micro Raman spectroscopy;


Fourier Transform Infrared (FTIR) spectroscopy;
N2 adsorption at -196 C to obtain specific surface area and
porosity characteristics;
Temperature programmed desorption (TPD).

III.2.1

ICP-MS

Inductively Coupled Plasma Mass Spectrometry (ICP-MS) is extensively


used to detect and quantify the presence of the majority of the elements in
the periodic table and is a very powerful tool for trace (ppb-ppm) and ultratrace (ppq-ppb) elemental analysis. In ICP-MS, the plasma is formed from
Argon gas and reaches very high temperatures of up to approximately
7000 K. The plasma is used to atomize and ionize the elements in a sample.
The sample to be analysed is introduced into the plasma as a fine aerosol. As
the sample aerosol passes through the plasma, it collides with free electrons,
argon cations and neutral Argon atoms. The result is that any molecules
initially present in the aerosol are quickly and completely broken down to
charged atoms. The resulting ions are then passed through a series of
apertures (cones) into the high vacuum analyzer. The isotopes of the
elements are identified by their mass-to-charge ratio (m/z) and the intensity
of a specific peak in the mass spectrum is proportional to the amount of that
isotope (element) in the original sample. An ICP-MS Agilent 7500ce
instrument was used for the analysis of Mo.

III.2.2

Thermal analysis (TG-MS)

The performances of samples as a function of temperature were


determined by Air flow thermal analysis (TG-MS). The apparatus used were
a TGAQ500 (Figure 7) thermogravimetric analyzer (TA Instruments) and a
SDTQ600 (Figure 8) simultaneous DSC/TGA (TA Instruments). Both
analyzers can be coupled to a Pfieffer Vacuum Benchtop Thermostar mass
spectrometer (MS) (Figure 9)

24

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

Figure 7 TGAQ500 Thermogravimetric Analyzer

Figure 8 SDTQ600 Simultaneous DSC/TGA

Figure 9 Pfieffer Vacuum Benchtop Thermostar mass spectrometer

25

Chapter III

TGAQ500 measures weight changes in a material as a function of


temperature. The system works in a temperature range of 20-1000 C, and
weight variation resolution is 0.1g . The sample, loaded in a crucible made
of platinum and connected to the balance arm by a small hook, is
progressively heated in the oven. A thermocouple controls the oven
temperature and a second thermocouple reads the sample temperature.
Sample pan loading and furnace movement are totally automated and there is
a touch screen data display to change operating parameters. Typically
measurements are carried out with 20 mg of sample in chromatographic air
flow (60 Ncc/min) with a heating rate of 10 C/min in the temperature range
of 20- 800 C.
The results are displayed as TG curves showing the mass variations as
functions of temperature or time, and DTG curves showing the conversion
rate (mass loss percentage per unit time) as functions of temperature or time.
Figure 10 contains the typical trends of the TG and DTG curves.

100
98

TG (%)

94
1
92
90

DTG (%/min)

96

88
86

100

200

300

400

500

-1
600

Temperature (C)

Figure 10 TG and DTG curves


SDTQ600 provides a simultaneous measurement of weight change
(TGA) and heat flow (DSC) on the same sample from ambient to 1500 C. It
features a proven horizontal dual beam design with automatic beam growth
compensation, and the ability to analyze two TGA samples simultaneously.
DSC heat flow data is dynamically normalized using the instantaneous
sample weight at any given temperature. The sample is loaded in a crucible
made of alumina and heated in the horizontal oven. There are two

26

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

thermocouples to control the oven temperature and the sample temperature.


Measurements are carried out with about 30 mg of sample in
chromatographic air flow (100 Ncc/min) with a heating rate of 10 C/min in
the temperature range of 20- 800 C.
Pfieffer Vacuum Benchtop Thermostar mass spectrometer can measure
the gas evolved from thermal analyzers up to 300 AMU. The evolved gases
are introduced into a heated quartz capillary, which is extremely fine, in
order to produce the necessary high vacuum, when the evolved gases enter
the mass spectrometer. The heated capillary is necessary in order to prevent
condensation of the hot gases on cold surfaces. The analysis of gases is
performed by a very high sensitive quadrupole mass detector. The necessary
high vacuum is obtained through 2 stages of vacuum pumps that are
integrated into a compact housing. First stage is a rotary pump; second stage
is a turbo molecular pump.
Both systems, the Mass Spectrometer and the Thermal Balance, are
connected to a common PC for data acquisition.

III.2.3

Micro Raman spectroscopy

Raman spectroscopy is a technique for the identification and


quantification of the chemical components of specimens.
When light is scattered by any form of matter, the energies of the
majority of the photons are unchanged by the process, which is elastic or
Rayleigh scattering. However, about one in one million photons or less, lose
or gain energy that corresponds to the vibrational frequencies of the
scattering molecules. This can be observed as additional peaks in the
scattered light spectrum. The process is known as Raman scattering and the
spectral peaks with lower and higher energy than the incident light are
known as Stokes and anti-Stokes peaks respectively.
Laser Raman spectra of powder samples were obtained with a Dispersive
MicroRaman (Invia, Renishaw) (Figure 11), equipped with 785 nm diodelaser, in the range 100-2500 cm-1 Raman shift.

Figure 11 Dispersive MicroRaman (Invia, Renishaw)

27

Chapter III

III.2.4

FT-IR Spectroscopy

This technique is based on the analysis of the absorption or of laser light


by specific chemical bonds. Fourier Transform Infrared (FTIR) spectroscopy
was performed in the 4000-400 cm-1 range with a resolution of 2 cm-1.
A Bruker IFS 66 FT-IR spectrophotometer, was used. Samples were
diluted at 1 wt % in KBr. The mixture was ground and a transparent disk of
100 mg was prepared with a press in vacuum. Disks are introduced into the
proper chamber and the scan is carried out at room temperature.

Figure 12 Bruker IFS 66 FT-IR spectrophotometer


The infrared spectrum includes all the radiation of wavelengths ranging
from 0.1 to 1000 m.

III.2.5

N2 adsorption measurements|

In order to study the porosity of catalysts powder, N2 adsorption


measurement were carried out at -196 C with a Costech Sorptometer 1040
(Figure 13).
The measurement was performed by continuous-flow method after
sample pre-treatment at 150 C for 2 h in He flow, in order to measure total
specific surface area (via single and multi-point methods) and micropore
volume (via micropore method).

28

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

Figure 13 Costech Sorptometer 1040

III.2.6

TPD investigations

Temperature programmed desorption (TPD) experiment of catalysts after


activity measurements were carried out in N2 flow (500 Ncc/min) at
atmospheric pressure in a quartz flow reactor, connected on-line with CO,
CO2 (Uras 10E Hartmann & Braun) and with on-line quadrupole mass
detector (MD800, ThermoFinnigan) in the range 20-500 C, with an heating
rate of 10 C/min. For the analysis 1 g of sample was loaded into the
microreactor.

III.3

Laboratory apparatus for catalytic test

Catalytic activity tests were performed using the laboratory apparatus


shown in Figure 14. It consists of three sections:

Feed section;
Reaction section;
Gas composition analysis section.

29

Chapter III

Figure 14 Laboratory apparatus for catalytic test


All the gas pipes ( e.d.) are of Teflon, connections are made with
Swagelok union and two, three and four way Nupro valves. All the
connections are in stainless steel to avoid any corrosion due to the presence
of water.

III.3.1

Feed section

All the gases come from SOL SPA with a purity degree of 99,999%.
Oxygen and nitrogen were fed from cylinders, nitrogen being the carrier
gas for cyclohexane (CH) and water vaporized from two temperature
controlled saturators and by changing temperature and N2 flow, it is possible
to obtain different concentrations in the reaction feed.
To feed an accurately controlled flow, Brooks measured flow controllers
(MFC) are used, able to operate with a maximum pressure drop of 3 atm
(Figure 15).

30

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

Figure 15 Mass flow controller


The working principle of the MFC is heat transport: the temperature
difference in a capillary, where a part of the gas is split, it is measured. This
temperature difference is proportional to the amount of heat adsorbed by the
mass gas for the equation:
.T = K Cp m
where:
T = temperature difference.
Cp = specific heat of the gas.
K = dimensional constant.
m = mass flow.
The instruments temperature detector produces an electrical signal from
0 to 5 V (c.c.); this signal is sent to the control unit (MFC C.U.) which
converts the signal in volumetric flow. This control unit allows the mass
flow of the gases to be regulated.
A rotameter for N2 is used for vaporizing water to feed to the reactor.

III.3.2

Reaction section

In the reaction section a system of valves allows the reactants to go to the


reactor, and the products to the analysis section, or, in the bypass position,
the reactants to the analysis section to verify the reactant composition.
The annular section of the fixed bed photocatalytic reactor (reactor
volume: 7 l) was realised with two axially mounted 500 mm long quartz
tubes of 140 and 40 mm diameter, respectively.

31

Chapter III

Gas Out

Gas In

Figure 16 Annular gas-solid photocatalytic fixed bed reactor


The reactor was equipped with seven 40 W UV fluorescent lamps
providing photons wavelengths in the range from 300 to 425 nm, with
primary peak centred at 365 nm. One lamp (UVA Cleo Performance 40 W,
Philips) was centred inside the inner tube while the others (R-UVA TLK 40
W/10R flood lamp, Philips) were located symmetrically around the reactor.
Both photoreactor and lamps were covered with reflectant aluminum foils. In
order to avoid temperature gradients in the reactor caused by irradiation, the
temperature was controlled to 35 2 C by cooling fans.
The catalytic reactor bed was prepared in situ, by coating quartz flakes
previously loaded in the annular section of a quartz continuous flow reactor
with an aqueous slurry of catalysts powder. The coated flakes were dried at
393 K for 24 hours in order to remove the excess of physisorbed water. This
treatment resulted in uniform coating well adhering to the quartz flakes
surface.

III.3.3

Analysis section

The gas composition is determined by on line analysers connected to a


PC for data acquisition. CO and CO2 concentration is measured by an on line
non dispersive IR analyzer (Uras 10, Hartmann & Braun), working on the
basis of specific adsorption of IR radiation (wavelength from 2 to 8 m).
Oxygen, cyclohexane and reaction products composition is determined by an
on line quadrupole mass detector (MD800, ThermoFinnigan) that can

32

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

analyze the outlet reactor gas, introduced into a heated silica capillary, up to
m/z = 800.

III.4

Photocatalytic tests conditions and typical trend

The amount of deposited catalyst, evaluated by weighing the reactor


before and after the coating treatment was 20 g. Catalytic tests were carried
out feeding 830 Ncc/min N2 stream containing 1000 ppm cyclohexane, 1500
ppm oxygen and adding 1600 ppm water to minimise catalyst
photodeactivation Einaga et al. (2002). Concerning this last aspect, several
authors Vorontsov et al. (2000), Martra et al. (1999), Marci et al. (2003),
Augugliaro et al. (1999) reported the positive influence of water on the
oxidation rate of many hydrocarbons. In fact water increases photocatalytic
activity because it is necessary for production of OH radicals and for
capture of photogenerated Park et al. (1999) holes enhancing the reaction
between O2 and an electron promoted to the cb from the vb to form
superoxide radical anion (O2).
A typical trend of photocatalytic test is reported in Figure 17 with
reference to a sulphated titania (2 wt % as SO3) supported MoOx catalyst
with a nominal Mo load equal to 4.7 wt % (as MoO3).

33

Chapter III

60
Lamp ON

Lamp OFF

m/z = 84

40

m/z = 32
m/z = 78
m/z = 67

30

CO2, ppm

20

CO2, ppm

MS signal, a.u.

50

10

50

100

150

200

0
250

time, min
Figure 17 Outlet reactor concentration (a.u.) of cyclohexane, oxygen
benzene and cyclohexene and (ppm) of carbon dioxide as a function of run
time
At the run starting time the nitrogen stream containing 1000 ppm
cyclohexane, 1500 oxygen and 1600 water was passed through the reactor in
the absence of irradiation at ambient temperature. Dark adsorption of
cyclohexane is observed. Cyclohexane breakthrough time was about 10
minutes. Thereafter cyclohexane outlet concentration slowly increased to
reach the inlet value after about 50 minutes, indicating that adsorption
equilibrium of cyclohexane on the catalyst surface was attained. At that time
the lamps were switched on: the cyclohexane outlet concentration
immediately decreased to about 22% of the inlet value and then progressively
increased with run time, reaching a steady state value corresponding to about
6% cyclohexane conversion after about 110 minutes. In the same figure the
change of oxygen outlet concentration is also reported showing a general
trend similar to that of cyclohexane. The analysis of products in the outlet
stream disclosed the presence of benzene and (much lower amount)
cyclohexene, as identified from the characteristic fragments m/z = 78, 77, 76,
74, 63, 52, 51, 50 (fragment 78 reported in Figure 17) and 82, 67, 54,
respectively (fragment 67 reported in Figure 17), and of carbon dioxide, as
detected by the NDIR analyser (Figure 17).

34

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

It is worthwhile to note that both breakthrough time of the products and


change of concentration with time are different for the different products.
Figure 17 shows that carbon dioxide is formed immediately after lamp
switch on and reaches a maximum concentration value in about 30 minutes,
slowly decreasing during the remaining run time. Instead, the outlet
concentration of benzene progressively increases reaching a maximum value
at greater time with respect to carbon dioxide. Then it decreases to a steady
state value reached after about 220 minutes. A similar trend is shown by
cyclohexene concentration, however the values are very much lower with
respect to benzene.
In order to verify that cyclohexane was converted in a heterogeneous
photocatalytic process, blank experiments were performed. A control test
was carried out with the reactor loaded with uncoated quartz flakes. No
conversion of cyclohexane was detected during this test, indicating the
necessity of the catalyst for the observed reaction. A second test was
performed with the catalyst loaded reactor, but leaving the lamps switched
off even after cyclohexane adsorption equilibrium was reached. In these
conditions the composition of the outlet reactor was identical to that of the
reactor inlet, indicating that no reaction occurred in dark conditions.
These results confirm the occurrence of photocatalysed cyclohexane oxydehydrogenation to cyclohexene and benzene (Figure 18) together with deep
oxidation to carbon dioxide.

Figure 18 Photocatalysed cyclohexane oxy-dehydrogenation to cyclohexene


and benzene
The catalytic performance was evaluated as:
CH % conversion = 100(moles of inlet CH moles of outlet CH)/(moles
of inlet CH),
BE % selectivity = 100(moles of outlet BE)/(moles of inlet CH moles
of outlet CH)
CO2 % selectivity = 100(moles of outlet CO2)/6(moles of inlet CH
moles of outlet CH), where CH is for cyclohexane and BE for benzene.
Total carbon mass balance was evaluated by comparing the inlet carbon
as cyclohexane and the outlet carbon as the sum of unconverted cyclohexane
and outlet benzene, cyclohexene and carbon dioxide.

35

Chapter III

After each photocatalytic test, catalyst was regenerated in air flow under
UV irradiation for 16 h.

III.4.1

Thermodynamic analysis

Thermodynamic analysis has been carried out with Gaseq program that is
written as a Microsoft Windows program with an easy graphic interface as
shown in Figure 19.The basic principle of this program is the minimization
of Gibbs free energy. Gaseq calculates chemical equilibrium in perfect
gases.

Figure 19 Grafic interface of Gaseq 0.74 program


For calculations Gaseq uses the thermodynamic information on species
that are provided in one or more files with the extension .tdd in the library.
A list of the types of calculation which can be performed is obtained by
clicking the down arrow on the Problem Type box in the top left of the
screen. The types available are:
Equilibrium at defined temperature and pressure;
Adiabatic temperature and composition at defined pressure;
Equilibrium at defined temperature and constant volume;

36

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

Adiabatic temperature and composition at constant volume;


Adiabatic compression/expansion;
Equilibrium Constant
For thermodynamic analysis the general reaction used is the following:
0.1C6H12 +0.15O2 +0.16H2O + 99.59N2products
where the main products are C6H12, CO2, O2, N2, H2O, C6H6 and C6H10.
The results of thermodynamic analysis have been reported in Figure 20 in
terms of cyclohexane conversion, benzene and carbon dioxide selectivity.
120
100

80

80
cyclohexane conversion

60
40

carbon dioxide selectivity

60

benzene selectivity

40

20

20

27

127

227

327

427

527

selectivity, %

conversion, %

100

627

T, C
Figure 20 Effect of temperature on cyclohexane conversion, carbon dioxide
and benzene selectivity
Conversion increases with temperature and reaches about 100% at a
temperature of about 330 C. A similar trend is shown by benzene
selectivity. Carbon dioxide selectivity decreases with temperature up to
obout 0 % at a temperature of 327 C. It is possible to see that benzene
selectivity is higher than carbon dioxide selectivity for a temperature above
110 C.

III.5

III.5.1

Results and discussion

Specific surface area and Chemical analysis

Chemical analysis (ICP-MS) was performed after microwave digestion


(Ethos Plus from Milestone) of sample in H3PO4/HNO3 mixtures.

37

Chapter III

Table 2 contains the values of specific surface area and the results of
chemical analysis (ICP-MS) of the samples as a percentage weight of MoO3.
Table 2 Specific surface area and MoO3 amount of MoOx/Al2O3 catalysts
Catalyst

Al
2MoAl
Al
8MoAl

Nominal MoO3
content,
wt%
0
2.0
0
8.0

MoO3
by ICP analysis,
wt %
2.2
7.6

Specific
surface area
(BET), m2/g
10
10
144
147

It can be seen that surface areas of Mo loaded catalysts and unsupported


aluminas are very similar. Moreover nominal MoO3 content is approximately
equal to MoO3 evaluated by ICP-MS analysis. Considering 18% Mo loading
as the upper for monolayer formation, as estimated by Inamura et al. (1998),
a coverage degree of 42% was calculated for 8MoAl, and. 108% was
obtained for 2MoAl Imamura et al. (1998).
Specific surface area results together with MoOx coverage calculation
suggest that molybdenum oxide is well dispersed on the alumina support and
probably as a monolayer.

III.5.2

Thermal analysis

Measurements are carried out with SDTQ600 in air flow (100 Ncc/min)
with a heating rate of 10 C/min in the temperature range of 20- 700 C.
Figure 21 shows the progression of the TG and DTG curves, together
with the heat flow signal (DSC), during the decomposition of ammonium
heptamolybdate used as precursor to impregnate alumina support.

38

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

12
100

10

DSC

DSC, W/g

TG, %

90

TG
80

DTG

6
4

DTG, %/min

-0.5 8

2
-5.5
0

70

200

400

600

-2

Temperature, C

Figure 21 Evolution of the weight loss during decomposition of ammonium


heptamolybdate together with DTG and DSC signals
Up to 600 C four decomposition steps at 127, 200, 295, and 385 C,
characterize the evolution of the weight loss, with continuous weight loss
between the first and the third decomposition steps. The weight loss after the
third decomposition step at 295 C amounts to 17.1%, and after the fourth
step at 385 C, to 18.10%. The latter value is in agreement with the
calculated weight loss of 18.48%, if MoO3 is assumed to be the
decomposition product of the starting material. During the first
decomposition step, an endothermic DSC signal can be distinguished. The
third decomposition step also shows endothermic character, while the second
and fourth steps are exothermic. With the exception of the first step, which
begins with the evolution of water only, all steps are accompanied by the
evolution of water and ammonia Wienold et al. (2003).
Figure 22 displays the loss of mass with temperature during
thermogravimetric analysis on 8MoAl catalyst after calcination.

39

Chapter III
100

0.8

0.6

96

0.4

TG ,%

98

DTG
94

0.2

92

0.0

90

200

400

600

DTG, %/min

TG

-0.2
800

Temperature, C

Figure 22 Thermogravimetric analysis of 8MoAl catalyst after calcination


The sample showed an initial weight loss with a DTG peak at about
66 C due to the adsorbed water loss. Above 200 C up to about 870 C no
significant weight loss was observed. The TG results suggest that the
dehydration and thermal decomposition of the precursor to form
MoO3/Al2O3 can be finished at 400 C which is in agreement with the
literature Wienold et al. (2003). Similar results are obtained on 2MoAl.

III.5.3

FT-IR spectroscopy

It is generally noted that Mo species formed on the surface of alumina are


isolated etrahedral Mo, octahedral polymolybdate, and MoO3. Mo is present
in the forms of heptamolybdate (Mo7O246-) and monomeric MoO42-. MoO42is preferentially adsorbed on the basic hydroxyl groups present on alumina
and, thus, isolated tetrahedral species are formed at very low Mo-loading
Jeziorowski and Knozinger (1979), Diaz and Bussel (1993). After
consumption of the hydroxyl groups octahedral polymolybdate is produced
Jeziorowski and Knozinger (1979), Giordano et al.(1975), Wang and Hall
(1980), Acro et al.(1992), and when the monolayer is completed
Al2(MoO4)3 begins to be formed, followed by subsequent formation of MoO3
crystallites at higher Mo-loading Medema et al.(1978). Spectroscopic
techniques give more direct and useful information on the structures of Mo.
Figure 23 shows the IR spectrum of 8MoAl measured by a transmission
mode. The original spectrum has been subtracted from the spectrum due to
Al alone.

40

Transmittance, a.u.

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

1250

1150

1050

950

850

Wavenumber, cm

750

650

550

-1

Figure 23 IR spectrum of 8MoAl obtained by subtracting the spectrum of


Al
Only one broad absorption band is observed at about 914 cm-1 in the
Mo=O vibrational stretching region. This band is not due to MoO3 (bands at
599, 821, 873 and 995 cm-1) and is attributed to surface-bound Mo species,
like octahedral (Oh) polymolybdate or isolated tetrahedral (Td) species.
Since no band of bridged MoOMo species in the lower wavenumber
region is present, this band is ascribed to isolated TdMo Imamura et al.
(1998).
IR spectra of Al and 2MoAl are contained in Figure 24.

41

Transmittance, a.u.

Chapter III

2Mo
Al

Al

1200

1100

1000

900

800

700

600

500

400

-1

Wavenumber, cm
Figure 24 IR spectra of Al and 2MoAl

Visible IR Mo-related bands in the spectrum of 2MoAl could not be


obtained causing to the strong absorption by Al, and no detailed
information can be attained about MoOx dispersion.In addition, altough Mo
species cover the support surface, sharp bands from MoO3 crystallites are not
observed.

III.5.4

Micro Raman spectroscopy

Better evidence for Mo surface species was obtained by Micro Raman


spectroscopy.
Raman spectra of Al and 8MoAl are reported in Figure 25.

42

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

400000

350000

Counts

Al
300000

8MoAl
250000

200000

150000

400

600

800

Raman shift, cm

1000

1200

-1

Figure 25 Raman spectra of Al and 8MoAl


It can be seen that Al do not exhibit any Raman band due to the low
polarizability of light atoms and the ionic character of the Al-O bonds
Wachs (1996). On 8MoAl catalyst the main Mo=O band is at 949 cm-1.
Other two bands are present as shoulders at 920 and 980 cm-1. The bands at
673, 822, and 999 cm-1 for MoO3 crystallites are absent.
Figure 26 shows the Raman spectra of Al and 2MoAl catalyst in the
range 900-1100 cm-1.

43

Chapter III

120000

2Mo
Al

Counts

100000

80000

Al
60000

40000
900

950

1000

1050

Raman shift, cm-1


Figure 26 Raman spectra of Al and 2MoAl
Similar to Al, also Al do not present Raman signals. On 2MoAl a
band at 925 cm-1 and a main Mo=O band at 955 cm-1 are visible. There is
also a band at 980 cm-1 present as a shoulder. Moreover the presence of
MoO3 crystallites is not detected.
As several authors report Wachs (1996), Bian et al. (1999), on Moalumina catalysts Mo=O bands near 950 cm-1 are characteristic of Mo=O
stretching mode of polymolybdenyl species.

III.5.5

Photocatalytic activity tests

To better understand the contribution of the metal, preliminary tests were


performed on Al and on Al. The obtained results evidenced that the both
alumina supports didnt exhibit any photoactivity.
Figure 27 and Figure 28 show the comparison between cyclohexane
conversion and benzene, cyclohexene and CO2 selectivity as a function of
illumination time on 8MoAl (8Mo) and 2MoAl (2Mo).

44

Experimental Results: Photocatalytic oxidation of ciclohexane on MoOx/Al2O3

cyclohexane conversion, %

0.3
0.25
0.2
8Mo

0.15

2Mo

0.1
0.05
0
0

10

20

30

40

50

illumination time, min


Figure 27 Cyclohexane conversion on 8MoAl (8Mo) and 2MoAl (2Mo) as
a function of illumination time

45

Chapter III

120

selectivity, %

100
80
benzene 8Mo
cyclohexene 8Mo

60

CO2 8Mo
benzene 2Mo

40

cyclohexene 2Mo
CO2 2Mo

20
0
0

10

20

30

40

illumination time, min


Figure 28 Selectivity to benzene cyclohexene and CO2 on 8MoAl (8Mo) and
2MoAl (2Mo) as a function of illumination time
Photocatalytic activity test showed that very low cyclohexane conversion,
about 0.25%, was obtained on both catalysts (Figure 27), but the selectivity
to CO2 was close to 100% (Figure 28). Only few ppm of cyclohexene and
benzene were detected on 8MoAl in the first two minutes of reaction
(Figure 28).

46

50

IV Experimental Results:
Photocatalytic oxidation of
cyclohexane on zeolites
supported MoOx

IV.1

Samples preparation

Zeolitic catalysts was prepared starting from Na,K-Ferrierite, with Si/Al


ratio of 8.4 and NaY, HY (Engelhard). The characteristics of the Na, Kferrierite are shown in Table 3.

Figure 29 Ferrierite structure

Chapter IV

Table 3 Characteristics of Na,K-Ferrierite


Bulk Density, g/cm3

0.40

Porous Diameter,

4.0

SiO2, dry wt %

84.9

Al2O3, dry wt %

8.6

Na2O, dry wt %

1.5

K2O, dry wt %

5.6

K2O/Al2O3

0.7

Na2O/Al2O3

0.28

SiO2/Al2O3

16.8

Ferrierite was ion exchanged to ammonium form with 1 M solution of


ammonium nitrate in order to obtain the NH4 form (AFer). Powdered
catalysts were prepared by wet impregnation of Afer, or HY or NaY with an
aqueous solution of ammonium heptamolybdate (NH4)6 Mo7O244H2O,
drying at 120 C for 12 hours and calcination in air at 550 C for 3 hours.
Table 4 contains the list of samples prepared with indications of the nominal
MoO3 load.
Table 4 List of catalysts with their MoO3 nominal content
Catalyst
AFer
5MoAFer
20MoAFer
NaY
20MoNaY
HY
20MoHY

IV.2

Nominal MoO3 content, wt%


5.0
20.0
20.0
20.0

Thermal analysis

Thermogravimetric analyses were carried out on zeolites used as support.

48

Experimental Results: Photocatalytic oxidation of ciclohexane on zeolites supported MoOx

Figure 30 shows the TG and DTG curves of ferrierite in ammonium form


(AFer) and of hydrogen ferrierite (HFer) obtained after calcination of AFer
at 550 C in air for 2 hours.
102

0.1

DTG HFer

100

TG, %

-0.1

TG AFer

96

-0.2
94

DTG AFer

-0.3

DTG, %/min

98

92
-0.4

90

TG HFer
88

-0.5
25

225

425

625

825

Temperature, C
Figure 30 Thermogravimetric analysis of AFer and HFer
As evidenced by the DTG minima, the two samples showed an initial
weight loss at temperatures lower than 120 C due to the adsorbed water
loss. AFer sample showed a second weight loss in the 200 500 C range,
which is absent in HFER, due to ammonium decomposition. In both
samples, another weight loss is present around 800 C, due to water formed
by hydroxyl condensation with the consequent collapse of the zeolite
structure. Figure 31 shows thermogravimetry of NaY zeolite.

49

100

0.2

96

0.0

TG, %

92
88

DTG

-0.2

TG

-0.4
-0.6

84
-0.8

80

DTG, %/min

Chapter IV

-1.0

76

-1.2

72

-1.4

25

225

425

625

Temperature, C
Figure 31 Thermogravimetric analysis of NaY
The only pronounced peak on the sample maximises at 99 C and is
associated with the removal of water molecules present either in cages or
channels of the zeolite.
Figure 32 shows the TG and DTG curves of HY zeolites.

50

Experimental Results: Photocatalytic oxidation of ciclohexane on zeolites supported MoOx

100

0.2
0

96

TG, %

92

-0.4

DTG
TG

88

-0.6
-0.8

84

-1

DTG, %/min

-0.2

-1.2
80

-1.4

76

-1.6
25

125

225

325

425

525

625

725

Temperature, C
Figure 32 Thermogravimetric analysis of HY
It can be seen that there is only a DTG peak at around 80 C caused by
the loss of zeolite-adsorbed water.

IV.3

N2 adsorption measurements

Surface area and porosity characteristics were obtained by N2 adsorption


at -196 C. The microporous volume was obtained by the Dubinin method
and its value for all the samples are shown in Table 5.
Table 5 Microporous volume of zeolites based samples
Catalyst
AFer
5MoAFer
20MoAFer
NaY
20MoNaY
HY
20MoHY

Microporous volume, cm3/g


0.130
0.045
0.024
0.261
0.084
0.206
0.024

The results obtained evidenced that the addition of molybdenum oxide


molybdenum on zeolite structure, leads to a microporous volume reduction

51

Chapter IV

and in particular for MoOx/AFer samples, the microporous volume decreases


with increase in Mo-loading.

IV.4

FT-IR spectroscopy

FT-IR spectra of HFer, 5MoAFer and 20MoAFer are contained in Figure 33.
On all the samples, typical absorption bands of ferrierite in the range 4001200 cm-1 are visible Corbo et al. (1994), Pirone et al. (1996). Absorption
at 1072 cm-1 is assigned to asymmetric stretching of SiO4, AlO4 tetrahedrons,
while at 802 cm-1 to symmetric vibration. Bands at 594 and 533 cm-1 are
related to double ring vibrations. Pore opening bands are at 461 and 437cm-1.

Transmittance, a.u.

HFer

1200

5MoAFer

20MoAFer

1100

1000

900

800

700

600

500

400

-1

Wavenumber, cm

Figure 33 FT-IR spectra of Fer 5MoAFer and 20MoFer


Both 5MoAFer and 20MoAFer show a characteristic band at around 920
cm probably due to Mo=O vibration of tetrahedral species MoO42- Xu et
al. (1995) and the intensity of this band increases with Mo content. On
20MoAFer catalyst additional bands at 963 cm-1 (assigned to
polymolybdate) at 995 cm-1 and around 868 cm-1 (chracteristics of MoO3
crystallites) can be observed Afanasiev et al. (1994).
FT-IR spectra of NaY and 20MoNaY are reported in Figure 34.
-1

52

Experimental Results: Photocatalytic oxidation of ciclohexane on zeolites supported MoOx

Transmittance, a.u.

NaY

20MoNaY

1200

1100

1000

900

800

700

600

500

400

Wavenumber, cm-1
Figure 34 FT-IR spectra of NaY and 20MoNaY
As it can be seen, with respect to unsupported NaY, FT-IR spectra of
20MoNaY show additional bands in the region 840-945 cm-1.due to the
presence of Mo-species. It is possible to notice, at 921 cm-1, the same band
present in the FT-IR spectrum of 20MoAFer catalyst, assigned to Mo=O
vibration of MoO42- species between Mo and structural oxygen of the zeolite.
Dimeric complexes in tetrahedral coordination show vibrational bands in the
range 873913 cm1 and 919943 cm1 Xu et al. (1995). Therefore,
vibrational bands present around 881 cm1, 894 cm1, 906 cm1 and 941 cm1
are probably due to such species present on catalyst surface. Finally, the
shoulder present at 868 cm-1 indicates the presence of MoO3 crystallites.
FT-IR spectra of HY and 20MoHY are reported in Figure 35.

53

Chapter IV

Transmittance, a.u.

HY

1200

20MoHY

1100

1000

900

800

700

-1

600

500

400

Wavenumber, cm
Figure 35 FT-IR spectra of HY and 20MoHY

With respect HY, FT-IR spectrum of 20MoHY catalyst shows a shoulder


around 929 cm-1 due to Mo=O vibration of MoO42- species. The other bands,
present around 889 cm-1, 902 cm-1, 929 cm-1, are probably due to dimeric
complexes in tetrahedral coordination. The shoulder at 954 cm-1 is assigned
to Mo=O vibrational stretching. Finally the shoulder around 1006 cm-1
indicates the presence of MoO3 crystallites on catalyst surface.
For all prepared catalysts, FT-IR studies have evidenced penetrations of
the Mo species into the zeolite cages during calcination, since it was detected
the presence of Mo oxide species in tetrahedral coordination, characteristic
of the migration of Mo species into zeolite Corma et al. (1988).

IV.5

Photocatalytic activity tests

Cyclohexane and CO2 concentrations as functions of time during a


photocatalytic test on AFer are reported in Figure 36.

54

Experimental Results: Photocatalytic oxidation of ciclohexane on zeolites supported MoOx

1200
lamp on

1000

lamp off

ppm

800

Cyclohexane

600

CO2
400
200
0
0

50

100

150

200

time, min
Figure 36 Outlet reactor concentration of cyclohexane and carbon dioxide
as a function of run time
Loading 20 g of catalyst and feeding 1000 ppm cyclohexane and 3%
oxygen in N2 (total flow rate: 500 Ncc/min) to the reactor in the absence of
irradiation, adsorption of cyclohexane was observed . When the lamps were
switched on, the cyclohexane outlet concentration immediately decreased of
about 20% with respect to the inlet value and then progressively increased
with run time, reaching a steady state value corresponding to about 15% of
cyclohexane conversion after about 120 minutes with a CO2 production of
220 ppm and no partial oxidation product were detected.
Photocatalytic tests were also carried out on 5MoAFer and 20MoAFer. In
these cases, loadings 20 g of catalyst and feeding 100 ppm cyclohexane,
1500 ppm oxygen and 1600 ppm water in N2 (total flow rate: 830 Ncc/min),
in the steady state, cyclohexane conversion reached lower values (less than
1%), but the analysis of the reaction products disclosed the presence of
benzene and cyclohexene, as identified from their characteristic fragments
(Figure 37) with together carbon dioxide. These results demonstrate that
photocatalysed cyclohexane oxy-dehydrogenation to cyclohexene and
benzene occurs on molybdenum-supported ferrierite since AFer alone
exhibits high activity only in total oxidation to carbon dioxide but is not
active for conversion to benzene and cyclohexene.

55

Chapter IV

MS signal, a.u.

m/z = 78
m/z = 67

lamp on

20

40

60

80

100

120

lamp off

140

160

180

200

time, min
Figure 37 Outlet reactor concentration (a.u.) of benzene and cyclohexene on
5MoAFer as a function of run time

cyclohexane conversion, %

The influence of Mo loading on AFer was evaluated and obtained results


are summarized in Figure 38, Figure 39 and Figure 40.
0.8
0.7
0.6
0.5
0.4
0.3

20MoAFer

0.2

5MoAFer

0.1
0
0

10

20

30

40

50

60

70

illumination time, min

Figure 38 Cyclohexane conversion on 5MoAFer and 20MoAFer as a


function of illumination time

56

80

benzene selectivity, %

Experimental Results: Photocatalytic oxidation of ciclohexane on zeolites supported MoOx

100
90
80
70
60
50
40
30
20
10
0

20MoAFer
5MoAfer

10

20

30

40

50

60

70

80

illumination time, min

cyclohexene selectivity, %

Figure 39 Benzene selectivity on 5MoAFer and 20MoAFer as a function of


illumination time

4
3.5
3
2.5
2

20MoAFer

1.5

5MoAFer

1
0.5
0

10

20

30

40

50

60

70

80

illumination time, min

Figure 40 Cyclohexene selectivity on 5MoAFer and 20MoAFer as a function


of illumination time
On both catalysts, after an illumination time of 80 minutes, cyclohexane
conversion obtained was less than 1% (about 0.6 % on 5MoAFer and 0.2 %
on 20MoAFer). It reached a maximum value after about 15 minutes and then
decreased to a steady state value, evidencing a catalyst deactivation.

57

Chapter IV

Selectivity to benzene reached a very great steady state value (about 80%)
on 20MoAFer while this value was 27 % on 5MoAFer. On 20MoAFer
benzene selectivity showed a value higher than 90 % after about 10 minutes
and then progressively decreased with illumination time. Instead on
5MoAFer, there was a progressive increase up to a steady state value.
Similar behaviours were obtained for cyclohexene selectivity and its steady
state value was very low for both catalysts (about 2 % on 5MoAFer and 2.5
for 20MoAFer). In summary, there was a decrease of cyclohexane
conversion and an increase of benzene and cyclohexene selectivity by
increasing Mo loading.
Finally, no photoactivity was obtained on HY, 20MoHY, NaY and
20MoNaY showing that it is important for system selectivity not only the
presence of molybdenum on catalyst surface but also the type of zeolite used
as support because it influences the distribution of Mo species on catalyst
surface (as FT-IR spectra showed) and then its activity.

58

V Experimental Results:
Photocatalytic oxidation of
cyclohexane on MoOx/TiO2

V.1

Effect of molybdenum loading

V.1.1

Samples preparation

TiO2 as anatase phase, containing 2 wt% sulphates (DT51, Rhone


Poulenc) was used as support. The impregnation process was performed by
means of aqueous solutions of ammonium heptamolybdate
(NH4)6 Mo7O244H2O of different concentrations in order to obtain TiO2based catalysts with different molybdenum percentages. Powder samples
were dried at 120 C for 12 hours and calcined in air at 400 C for 3 hours.
The following table reports the list of catalysts with their nominal MoO3
contents.
Table 6 List of catalysts with their MoO3 nominal contents
Catalyst
DT2
2MoDT2
4MoDT2
8MoDT2

V.1.2

Nominal MoO3 content, wt%


2.0
4.7
8.0

Specific surface area and Chemical analysis

Chemical analyses of Molybdenum load were performed after microwave


digestion of catalyst in HNO3/HCl and HF/HCl mixtures. The list of samples
with their surface areas and the results of chemical analyses in comparison

Chapter V

with nominal metal loadings with together theoretical MoOx surface


coverage degree are reported in Table 7.
Table 7 List of catalysts and their characteristics
Catalyst

Nominal
MoO3
content,
wt%

DT2
2MoDT2
4MoDT2
8MoDT2

MoO3
by ICP
analysis,
wt %

2.0
4.7
8.0

1.8
4.4
7.6

Specific
Theoretical
surface
MoOx surface
area
coverage degree,
(BET),
%
m2/g
71
71
21
68
54
63
100

It can be seen that MoO3 loadings evaluated by ICP analysis are


approximately equal to that one calculated for the impregnation process.
MoOx coverage, calculated from the analysed MoO3 loading and assuming a
monolayer capacity of 0.12% (w/w) for MoO3/m2 Ng and Gulari (1985),
Kim et al. (1989), is 54% for 4MoDT2, 21% for 2MoDT2 and 100% for
8MoDT2.
Specific surface areas (S.S.A.) of the catalysts (m2/g) are plotted against
measured MoO3 loading in Figure 41.

75

S.S.A, m2/g

70
65
60
55
50
0

MoO3, wt %
Figure 41 Variation of specific surface area with MoO3 loading

60

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

It can be seen that surface area of Mo loaded catalysts showed a


decreasing trend with increasing Mo content. It is noted that it remained
constant up to 1.8 MoO3 wt% and started to decrease with further increases
in Mo-loading perhaps due to pore blockage.

V.1.3

Thermal analysis

Figure 42 shows TG-MS results on DT2 sample.


0.2

101

TG

TG, %

99

0.1

98

97

DTG, %/min

DTG

100

0.0

96

95

200

400

600

800

-0.1
1000

Temperature, C

Figure 42 TG-MS results on DT2 sample


The first main step of weight loss below 150 C is associated with
hydration water desorption. The second step (present as a shoulder) that
occurred up to about 390 C, is related to the removal of OH- surface groups
of titania. Morishige (1985) reported that TiO2 as the anatase phase
undergoes the removal of most surface OH- groups in a temperature range
extending up to 400 C. The last weight loss, located at temperature higher
than 600 C, is attributed to the decomposition of sulphate species giving
rise to gaseous SO3 as identified from its characteristic fragments m/z = 48
and 64.
Results of TG-MS analysis of calcined 2MoDT2, 4MoDT2 and 8MoDT2
catalysts are reported in Figure 43, Figure 44 and Figure 45.

61

Chapter V

100
0.6

99

TG

0.4

TG, %

98
0.2

97
0.0

DTG, %/min

DTG

96
-0.2

95

200

400

Temperature,C

Figure 43 TG-MS results on 2MoDT2 sample

62

600

800

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

102

0.6

TG

100

0.4

DTG

96

0.2

DTG, %/min

TG, %

98

94
0.0
92

90

200

400

600

800

-0.2

Temperature, C

Figure 44 TG-MS results on 4MoDT2 sample


100

TG

0.3

DTG

98

TG, %

96
0.1
94

DTG, %/min

0.2

0.0
92
-0.1
90

200

400

600

800

Temperature, C

Figure 45 TG-MS results on 8MoDT2 sample


For all MoDT2s, two main complex stages of weight loss, respectively
under and over 400C, can be detected. The first main step (up to about

63

Chapter V

390C) is associated to water and titania hydroxyl group removal, while the
second main step is due to decomposition of different kinds of surface
sulphates (as fragments m/z = 48 and 64 evidenced). Moreover it was found
that samples didnt show the weight loss due to the presence of ammonium
ion. Sulphate stability is affected by the presence of the metal oxide on
titania Ciambelli et al. (1996). The thermogravimetric curves of DT2,
2MoDT2, 4MoDT2 and 8MoDT2 showed that the sulphate weight loss step
in the catalysts containing molybdenum started at temperatures lower than in
sulphated titania.
The hydroxyl groups and surface sulphates (as SO42-) density (evaluated
in the range 180-350C and 400-800 C respectively) are presented in Table
8.
Table 8 Hydroxyls and surface sulphates density
Catalyst

Hydroxyls density,
mmol/m2

SO42- density,
mmol/m2

DT2
2MoDT2
4MoDT2
8MoDT2

0.028
0.015
0.018
0.017

0.0031
0.0029
0.0029
0.0033

It is possible to observe that the introduction of Mo species on titania


caused a decrease of hydroxyls density. All MoDT2s catalysts showed the
same amount of hydroxyls/m2, despite the decrease of the specific surface
area. This probably means that the introduction of molybdenum oxide on the
titania surface occurred with the formation of new surface hydroxyls
Sorrentino et al. (2001). From the data reported in Table 8 it can also be
seen that surface sulphates density is similar for all catalysts.

V.1.4

FT-IR spectroscopy

The FT-IR spectra as a function of Mo-loadings are shown in Figure 46.

64

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

8MoDT2

Transmittance, a.u.

4MoDT2

2MoDT2

DT2

1200

1100

1000

900

800

700

-1

Wavenumber, cm

Figure 46 FT-IR spectra of DT2, 2MoDT2, 4MoDT2 and 8MoDT2


In the spectrograms, the 8001200 cm1 region is interesting as in this
region characteristic bands due to Mo species can be detected. For pure
titania, two broad and strong bands appear at 1049 and 1132 cm1 which are
attributed to S=O vibrations of the free sulphate groups Samantaray and
Parida (2001). FT-IR spectra of TiO2 supported catalysts showed Morelated bands superimposed upon typical absorptions from the support. For
all Mo catalysts no MoO3 crystallites sharp bands (Mo=O stretching
vibrations at 992 cm-1 and bulk vibrations and 820 cm-1) are present
Matralis et al. (1995). A band at 954 cm-1 on 2MoDT2, at 957 cm-1 on
4MoDT and at 963 cm-1 on 8MoDT is observed. The bands near 960 cm-1
can be assigned to the stretching mode of molybdenyl species in hydrated
form Matralis et al. (1995), Lietti et al. (1996). In particular, they have
been assigned to terminal Mo=O stretching of octahedral polymeric surface
species Matralis et al. (1995). The results indicate that the formation of

65

Chapter V

octahedral polymeric molybdate is observed and the amount of this


polymeric octahedral molybdate increases with increase in Mo-loading. Ng
and Gulari (1985) have reported from their IR and Raman experiments that
at lower loading, the majority of the surface molybdates are in tetrahedral
coordination and at monolayer coverage, octahedrally coordinated polymeric
surface species are formed. Beyond a monolayer, bulk molybdenum trioxide
appears. Quincy et al. (1989) have also reported similar results.

V.1.5

Micro Raman spectroscopy

Raman spectrum of DT2 sample after calcination is presented in Figure


47.
300000
250000

Counts

200000
150000
100000
50000
0
100

200

300

400

500

Raman shift, cm

600

-1

700

800

Figure 47 Raman spectrum of DT2 sample


The sample display bands at 144, 396, 514, 637 cm-1 with a very weak
shoulder at 195 cm-1 due to the Raman active fundamentals of anatase
Alemany et al. (1995), Osaka et al. (1978). The supported molybdenum
oxide phase on titania introduced a new Raman feature at Raman shift
around 960 cm-1. The Raman spectra of the MoDT2s catalysts in comparison
with DT2 and MoO3 spectra in the range 700-1200 cm-1 are shown in Figure
48.

66

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2


800000
700000

Counts

600000

8MoDT2

500000

4MoDT2

400000

2MoDT2
DT2

300000

MoO3

200000
700

800

900

1000

-1

1100

1200

Raman shift, cm

Figure 48 Raman spectra of DT2, 2MoDT2, 4MoDT2, 8MoDT2 and MoO3


samples
The weak band at 790 cm-1 is due to the first overtone of the 396 cm-1
band of TiO2 (anatase). MoO3 is characterized by the prominent peaks at 997
and 821 cm-1 which are attributed to the stretching vibrations of the terminal
Mo=O and the brindging Mo-O-Mo bonds respectively Hirata (1989).
Supported molybdenum oxide phase on titania introduced a new Raman
feature at Raman shift around 960 cm-1. In particular, Raman spectra of
MoDT2s catalysts show a main band at 953 cm-1 on 2MoDT2 and at 958
cm-1 on 4MoDT2. On 8MoDT2 a complex band could be due to the
overlapping of 956, 966, 978, 984 and 995 cm-1 peaks. The increasing in
wavenumbers with increasing Mo loading has been attributed to higher
polymerisation degree of Mo species Cheng and Schrader (1979). The
small peak at 995 cm-1 could indicate the incipient formation of segregated
MoO3 crystallites. Therefore, the surface is mostly covered by polymeric
octahedral species anchored at the surface for all MoDT2s catalysts.

V.1.6

Photocatalytic activity tests

Cyclohexane conversion and CO2 concentration obtained on DT2 catalyst


are shown in Figure 49 and Figure 50 respectively.

67

Chapter V

cyclohexane conversion, %

30
25
20
15
10
5
0
0

10

20

30

40

50

60

70

80

90 100 110

illumination time, min

Figure 49 Cyclohexane conversion on DT2 as a function of illumination


time
Maximum cyclohexane conversion was about 25 %, decreasing to 3% in
30 minutes. A steady state condition was obtained after about 90 minutes of
illumination with a conversion of approximately 2 %.

68

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

140
120

CO2, ppm

100
80
60
40
20
0
0

10

20

30

40

50

60

70

80

90

100 110

illumination time, min

Figure 50 Carbon dioxide concentration on DT2 as a function of


illumination time
Carbon dioxide was formed immediately after the lamps were switched
on and reached a concentration of about 115 ppm after an illumination time
of 110 minutes. CO2 and water were the only observed products and no other
reaction products were detected. Thus, photocatalytic oxidation of
cyclohexane on unsupported titania leads to a complete mineralization into
CO2 and H2O without formation of by-products, according to the following
reaction:
C6H12+9O2=6CO2+6H2O
The comparison of cyclohexane conversion over 2MoDT2, 4MoDT2 and
8MoDT2 is shown in Figure 51.

69

cyclohexane conversion, %

Chapter V

50
45
40
35
30
25
20
15
10
5
0

8MoDT2
4MoDT2
2MoDT2

10

20

30

40

50

60

70

80

90

100 110

illumination time, min


Figure 51 Cyclohexane conversion on MoDT2s as a function of illumination
time
On all catalysts a maximum value was reached after about 5 minutes,
then activity decreased approaching a steady state conversion. On 2MoDT2
maximum cyclohexane conversion was about 45 %, decreasing to about 20
% in 30 minutes. On 4MoDT2 maximum conversion was lower (about 21%
after 8 minutes of illumination). It was about 7% after 30 minutes and 6%
after 110 minutes. Increasing Mo loading up to 8 wt% MoO3 the initial
maximum conversion was lower, about 15%, while steady state conversion
was 2.3% after 30 minutes. Therefore the progressive coverage of the titania
surface by MoOx species resulted in decreased initial and steady state
cyclohexane conversions.
While on DT2 the only reaction product was CO2, all MoOx/TiO2
catalysts exhibited unexpected high selectivity to benzene. On 2MoDT2
selectivity to benzene was about 10% after 110 minutes. On 4MoDT2
(Figure 52) maximum selectivity to benzene reached 31%. On 8MoDT2
higher selectivity to benzene was observed (65% after 40 minutes). On all
catalysts the presence of very low amounts of cyclohexene in the reaction
products was detected (steady state selectivity was about 0.4% on 2MoDT2,
0.7% on 4MoDT2 and 1.5% on 8MoDT2).

70

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

benzene selectivity, %

70
60
8MoDT2

50

4MoDT2
40

2MoDT2

30
20
10
0
0

10

20

30

40

50

60

70

80

90

100 110

illumination time, min

Figure 52 Selectivity to benzene on MoDT2s as a function of illumination


time
Selectivity to CO2 on MoDT2s as a function of illumination time is
reported in Figure 53.

71

Chapter V

CO2 selectivity, %

30
8MoDT2

25

4MoDT2
20

2MoDT2

15
10
5
0
0

10

20

30

40

50

60

70

80

90

100 110

illumination time, min


Figure 53 Selectivity to CO2 on MoDT2s as a function of illumination time
It can be seen that selectivity to CO2 decreased with increasing
molybdenum loading. On 2MoDT2, 4MoDT2 and 8MoDT2 maximum
selectivity to carbon dioxide was about 27 %, 8 % and 5 % respectively.
In addition, a catalyst containing 12 wt% of MoO3 nominal content was
prepared. In this case, FT-IR and Raman spectra revealed the presence of
MoO3 cristallytes on catalyst surface and no photocatalytic activity was
obtained.
Photocatalytic activity tests on all MoDT2s catalysts evidenced that the
presence of MoOx species on the surface of titania changes the selectivity of
the catalyst with increasing molybdenum content indicating that the
interaction between titania and supported molybdenum oxide plays an
essential role in changing the catalyst selectivity. FT-IR and Raman
spectroscopy data coupled with photocatalytic activity results showed that
the selective formation of benzene is likely due to the presence of
polymolybdate species supported on the titania surface. It could be argued
that polymolybdate species poison unselective sites of the titania surface
which would lead to total oxidation of cyclohexane.
Moreover, Figure 52 and Figure 53 clearly show the different shapes of
selectivity curves relevant to benzene and carbon dioxide, respectively.
Carbon dioxide is likely formed immediately after lamp switch on mostly on
titania sites, while benzene formation seems to require the formation on
intermediate surface species, whose generation should be related to some
interaction between support and molybdenum oxide surface species. A
possible reaction mechanism will be discussed.

72

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

V.1.7

Effect of light intensity

Light intensity is an important parameter to consider. Increasing the light


intensity affects the rate of the reaction by increasing the number of charge
carriers generated in the semiconductor. Most researchers have found
different effects at different levels of light intensity Rendel (1987):
at low light intensities, the rate increases in proportion to the
light intensity;
at intermediate light intensities the rate only varies with the
square root of intensity Ollis (1991), Okamoto et al. (1985);
at high light intensities the rate of photodegradation is
independent of light intensity.
For photocatalysis, as light intensity increases the rate increases due to
the increased number of oxidising species that are produced. The rate
increases with light intensity (I) to a power n. At low light intensities the
reaction rate increases directly in proportion to light intensity suggesting that
few oxidising species are lost through recombination processes. At high light
intensities, the reaction rate increases in proportion to I to the power of 0, i.e.
the rate becomes independent of light intensity and the expected ratelimiting factor is mass transfer. At intermediate light intensities the rate only
varies with the square root of intensity, Okamoto et al. (1985), Kormann et
al. (1991) and hence efficiency suffers. This was attributed by Egerton and
King (1979) to energy wasting recombination reactions between electrons
and holes. Increased intensity always results in an increase in the volumetric
reaction rate until the mass transfer limit is encountered. However, once
intermediate light intensities are reached any increase in I will not lead to a
proportional increase in rate. The I1 to I0.5 rate transition is said to depend on
the catalyst material Matthews et al. (1990).
The dependence of the cyclohexane conversion on light intensity has
been investigated on 4MoDT2 catalyst. The change of cyclohexane
conversion in the steady state condition as a function of light intensity is
shown in Figure 54.

73

Chapter V

cyclohexane conversion, %

4.5
4
3.5
3
2.5
2
1.5
1
0.5
0
0

50

100

150

200

250

300

light intensity, W
Figure 54 Effect of light intensity on cyclohexane conversion
Cyclohexane was unconverted in the absence of light and its conversion
increased up to 4 % in correspondence of a light intensity equal to 280W. In
all cases selectivity to benzene and carbon dioxide was 27 % and 8 %
respectively. The obtained results suggest that cyclohexane is not converted
in absence of irradiation. Moreover the relationship is linear at low light
intensities up to a certain point (intermediate light intensity) where the
cyclohexane conversion starts to level off (high light intensities). This would
suggest that the cyclohexane conversion is proportional to the light intensity
up to a certain point where it becomes independent by light intensity.

V.2

V.2.1

Influence of sulphate content

Samples preparation

Three titanias were used as supports: two commercial titania samples (DT
and DT51, Rhone Poulenc) with different sulphate contents (respectively 0.5
wt % and 2 wt %) and an ultrafine sulphate-free titania produced by laserpyrolisis Curcio et al. (1991). The samples are named, respectively, T0,
T5 and T20 with reference to the sulphate content. MoOx-based catalysts

74

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

were prepared by wet impregnation of titania with aqueous solution of


ammonium heptamolybdate (NH4)6Mo7O244H2O, followed by drying at 120
C and calcination at 400 C for 3 hours. The list of catalysts is reported in
Table 9
Table 9 List of catalysts with their MoO3 nominal content
Catalyst
T0
MoT0
T05
MoT05
T20
MoT20

V.2.2

Nominal MoO3 content, wt%


4.7
4.7
4.7

Specific surface area, chemical analysis and sample acidity

Table 10 contains the values of specific surface area, the results of


chemical analysis of the samples, the theoretical MoOx surface coverage
degree and point of zero charge (ZPC) values.
Table 10 List of catalysts and their characteristics
Catalyst

ZPC
pH unit

MoO3
by ICP
analysis,
wt %

T0
MoT0
T05
MoT05
T20
MoT20

6
3.8
4.7
1.8
2
1.7

4.5
4.2
4.4

Theoretical
Specific
MoOx surface
surface
area (BET),
coverage
2
m /g
degree,
%
88
88
44
67
68
50
71
68
54

Chemical analysis showed that the Mo loading is similar for all supported
catalysts whereas theoretical MoOx surface coverage increased with sample
sulphate content. Moreover specific surface area data suggest that for MoT0
and MoT05 catalyst, Mo-species are well dispersed on titania surface; for
MoT20 catalyst, there is a small decrease of S.S.A. value between catalyst
and the relevant support.
Mass titration method was used to estimate the acidity of sample
powders. The point of zero charge, which describes the acidity of oxide
materials, may be measured using potentiometric titration, mass titration, or
measurement of the wetting angles. The mass titration method of ZPC
characterization was initially proposed by Noh and Schwarz (1989). Their

75

Chapter V

studies showed that when mass fraction of oxide powder is increased in


aqueous electrolyte then the pH value of suspension approaches the ZPC
value of oxide powder.
The limiting pH value is independent of the initial pH of the electrolyte
alac and Kallay (1992). The electrolyte is usually nitrogen-bubbled to
prevent the influence of carbon dioxide on the pH of the electrolyte. The
electrolyte usually contains only monovalent ions such as Na+, K+, and Cl.
It is assumed that the sample powder is pure and insoluble in the electrolyte
used. If the sample powder is contaminated then it is possible that the
impurities radically influence on the obtained pH values. It is also possible
that there is not a limiting pH value at all Noh and Schwarz (1989), alac
and Kallay (1992). Acidic contaminations tend to decrease pH, whereas
basic contaminations increase pH. Other mass titration experiment have been
performed by Reymond and Kolenda (1999) and alac and Kallay (1992).
Reymond and Kolenda (1999) studied ZPC of several pure and mixed
oxides. The ZPC values obtained for pure oxides were in agreement with the
values determined by other methods (e.g., potentiometric titration). For
impure samples the obtained pH values indicated the chemical nature of
impurities. alac and Kallay (1992) extended the idea of mass titration for
contaminated samples. In addition, for ordinary mass titration, they
performed traditional acidbase titration for sample suspensions, in which
the solid powder content was so high that the limiting pH value had been
achieved. They showed that using acidbase titration it is possible to
determine both the ZPC of the sample powder and the nature and degree of
contamination. In this study the mass titration studies were performed using
procedures described in Noh and Schwarz (1989). Shorter stabilization times
after each powder addition (2 hours in this study) were used to minimize
possible dissolution of sample powders.
Table 10 shows that the value of ZPC of T0 is 6, according to the
amphoteric character of anatase titania. The presence of sulphate increases
surface acidity, leading to ZPC value of 4.7 on T05. As the sulphate load
increases, ZPC decreases to 2. The comparison of ZPC values of catalysts
and the relevant support indicates that the presence of molybdenum confers,
in all cases, strongest acidity than that of the relevant support.

V.2.3

Thermal analysis

Hydroxyls and sulphates amount were evaluated by TG-MS analysis in


the range 180-350C and 400-800C, respectively. The obtained results can
be summarized in the following table in which there are the values of
hydroxyls and sulphates (as SO42-) density on catalysts surface.

76

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

Table 11 Hydroxyls and surface sulphates density


Catalyst

Hydroxyls density,
mmol/m2

SO42- density,
mmol/m2

T0
MoT0
T05
MoT05
T20
MoT20

0.052
0.022
0.034
0.019
0.028
0.018

0.00086
0.00084
0.0031
0.0029

From the data reported in Table 11, it is possible to observe that


MoOx/TiO2 catalysts have a lower hydroxyls density than the corresponding
support, whereas the values of SO42- density are similar and increase with the
surface sulphates amount.

V.2.4

FT-IR spectroscopy

In Figure 55 FT-IR spectra of MoT0, MoT05 and MoT20 in the range


700-1200 cm-1 are reported.

77

Chapter V

Transmittance, a.u

MoT20

MoT05

MoT0

1200

1100

1000

900

800

700

-1

Wavenumber, cm

Figure 55 FT-IR spectra of MoT0, MoT05, and MoT20


FTIR spectra of TiO2 supported catalysts show MoOx species bands
superimposed upon typical absorptions from the support. Sulphate
absorption main bands are located at 1050 and 1132 cm-1 on T05 and, with
enhanced intensity, on T20. For all Mo-based catalysts no evidence for
MoO3 crystallites (sharp bands from Mo=O stretching vibrations at 992 cm-1
and bulk vibrations and 820 cm-1) are present Maity et al. (2001). Bands at
947 cm-1 on Mo/T0, at 951 cm-1 on Mo/T05 and at 957 cm-1 on Mo/T20 are

78

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

observed, assigned to terminal Mo=O stretching of polymeric surface


species Cheng and Schrader (1979).

V.2.5

Micro Raman spectroscopy

Raman spectra of MoT0, MoT05 and MoT20 are contained in Figure 56.
130000
120000

MoT20

Counts

110000
100000
90000

MoT05

80000
70000

MoT0

60000
50000
40000
700

800

900

1000

1100

1200

-1

Raman shift, cm

Figure 56 Raman spectra of MoT0, MoT05 and MoT20 samples


It can be seen that beyond the weak band at 790 cm-1 due to the first
overtone of the 396 cm-1 band of titania, Raman spectra evidenced a main
signal at 950 cm-1 on Mo/T0, at 956 cm-1 on Mo/T05 and at 958 cm-1 on
Mo/T20, all characteristic of octahedral MoOx species. The increasing of
wavenumber can be attributed to a higher degree of polymerisation of these
last species on catalyst surfaces.

V.2.6

Photocatalytic activity tests

In Figure 57, cyclohexane conversion on T0, T05 and T20 as a function


of illumination time is reported. Photocatalytic tests performed on these
samples showed that cyclohexane conversion reached a maximum after
about 5 minutes of illumination time and then decreased for all catalysts to
reach a steady state conversion of about 2%, 5% and 8% on T2, T05 and
T20, respectively.

79

cyclohexane conversion, %

Chapter V

40
35
30
T0

25

T05

20

T20

15
10
5
0
0

20

40

60

80

100

illumination time, min


Figure 57 Cyclohexane conversion on T0, T05 and T20 as a function of
illumination time
In Figure 58 CO2 production on T0, T05 and T20 as a function of
illumination time is reported.
600

CO2, ppm

500
400
300

T0
T05

200

T20
100
0
0

20

40

60

80

100

illumination time, min


Figure 58 Carbon dioxide concentration on T0, T05 and T20 as a function
of illumination time

80

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

For all catalysts, carbon dioxide was the only product detected in the gas
phase and started forming immediately after lamp on, reaching steady state
values (510 ppm, 335 ppm and 110 ppm on T0, T05 and T20 respectively)
after about 110 minutes. Thus cyclohexane conversion and CO2 yield
decreased with the sulphate content.
These last results may be explained taking into account the reaction
mechanism of gas-solid photocatalytic decomposition of cyclohexane to CO2
on titania catalyst (Figure 59) reported by Einaga et al.(2002).

Figure 59 Mechanism for oxidation of cyclohexane on titania


Irradiation of TiO2 with UV light generates highly reactive electron-hole
pairs. The hole subsequently oxidizes the surface hydroxyl groups to form
the OH radicals Einaga et al. (2002). The reactivity of OH radicals toward
hydrocarbons has been well investigated Seinfeld and Pandis (1997). The
OH radicals abstract the H atoms of saturated CH bonds of cyclohexane.
The resulting intermediate radicals are subsequently oxidized by molecular
O2. These intermediates are decomposed to CO2 via the subsequent
oxidation processes. This proposed mechanism evidences that the OH
radicals are the active species for the cyclohexane photocatalytic oxidation
on titania. Moreover it has been reported that the catalytic activity of TiO2
could be maintained indefinitely under an abundance of water vapour since it
can be adsorbed on the catalyst surface to form surface hydroxyl groups
Alberici and Jardim (1997), Obee and Brown (1995). The surface
hydroxyl groups exhibit their important influence on the photoreaction
process by trapping the charge transfer reaching the catalyst surface to
produce very reactive surface hydroxyl radicals. From these considerations it
is possible to observe that rehydroxylation process of the catalyst surface is
essential for activity of titania toward cyclohexane total oxidation. Xie et al.
(2004), studying gas-solid photocatalytic oxidation of heptane on sulphated
and unsulphated titania in presence of water vapour, found that the presence
of SO42 may be detrimental to the rehydroxylation ability of the catalyst.
This last observation could explain the behaviour of cyclohexane conversion
and CO2 production reported in Figure 57 and Figure 58 respectively.

81

cyclohexane conversion, %

Chapter V

45
40
35
30

MoT0

25

MoT05

20

MoT20

15
10
5
0
0

50

100

150

200

illumination time, min


Figure 60 Cyclohexane conversion on Mo/T0, Mo/T05 and Mo/T20 catalysts
as a function of illumination time
Cyclohexane conversion on Mo/T0, Mo/T05 and Mo/T20 is shown in
Figure 60. The analysis of the outlet stream disclosed the presence of
benzene, cyclohexene (less than 1 ppm) and CO2 for all catalysts. A
maximum value of conversion was reached after about 5 minutes, then
activity decreased approaching a steady state conversion. On Mo/T0 the
maximum cyclohexane conversion was about 40 %, decreasing to 17% in 25
minutes, less quickly with respect to Mo/T05 and Mo/T20. On Mo/T20 the
initial maximum conversion was lower (about 21% after 8 minutes of
illumination), reached 7% after 30 minutes and 4% after 220 minutes. On
Mo/T05 the initial maximum conversion was higher with respect to Mo/T20
(about 23%), while steady state conversion was about 1% after 60 minutes.

82

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

benzene selectivity, %

100
90
80
70
60

Mo/T0

50

Mo/T05

40
Mo/T20

30
20
10
0
0

10

20

30

40

cyclohexane conversion, %

Figure 61 Benzene selectivity versus cyclohexane conversion on Mo/T0,


Mo/T05 and Mo/T20
In Figure 61 benzene selectivity as a function of cyclohexane conversion
is reported. Low benzene selectivity and high selectivity to CO2 were
observed on Mo/T0 in the whole conversion range. On Mo/T05 benzene
selectivity ranged from 35 to about 70%. On Mo/T20 benzene selectivity
reached values higher than 80 % (86% with a cyclohexane conversion of
6%). The dependence of benzene selectivity on cyclohexane conversion for
all catalysts is less strong with respect to the usually found effect in
hydrocarbon catalytic partial oxidation processes especially in the case of
Mo/T20, for which the selectivity to benzene is very weakly decreasing with
conversion. Therefore, the presence of sulphate species on titania surfaces
enhanced the benzene yield more at higher sulphate contents.
In the section V.1, it was shown that polymolybdate species change the
photoactivity of titania, the higher polymerisation favouring the formation of
benzene. The FT-IR and Raman results showed that highly polymerised
MoOx species are present on Mo/T20 catalyst that gives the best benzene
yield. Even if, due to the presence of sulphate, the higher polymerisation
degree could be associated to the lower surface area available to
molybdenum oxide species deposition, leading to a higher surface density,
the whole results suggest an active role of sulphate in promoting the
selectivity to benzene. A possible role of the sulphate in the reaction
mechanism will be discussed.

83

Chapter V

V.3 Photocatalytic oxidative dehydrogenation of cyclohexane:


reaction mechanism
Gas-solid partial photooxidation of paraffins on titanium dioxide under
UV irradiation lead to formation of ketones and aldehydes Walker et al.
(1977), Djeghri and Teichner (1980), Bickley et al. (1973), Formenti and
Meriaudeau (1971). An intermediate formation of an alcohol by the
addition of atomic oxygen to the paraffin was postulated. The alcohol may
be either directly oxidized into an aldehyde or ketone or dehydrated into an
olefin which, in turn, is oxidized at the ethylenic bond into aldehydes
(primary carbon) or ketones (secondary carbon).
On the basis of this reaction mechanism, cyclohexene detected in the
outlet stream could be product by a dehydration step involving cyclohexanol
which may be formed by oxidation of cyclohexane. Because of the
possibility of cyclohexanol undergoing this reaction (dehydration into
cyclohexene), it was considered to investigated the photooxidation of
cyclohexanol and cyclohexene in order to elucidate the mechanism of gassolid photocatalytic oxidative dehydrogenation of cyclohexane.
Catalytic tests were performed on MoDT2s catalysts under the same
conditions as for the cyclohexane.

V.3.1

Photocatalytic oxidation of cyclohexanol

Carbon dioxide produced by the photocatalytic oxidation of cyclohexanol


is reported in Figure 62.

84

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

250

CO2, ppm

200

2MoDT2
4MoDT2
8MoDT2

150
100
50
0
0

30

60

90

120

illumination time, min


Figure 62 Carbon dioxide concentration on 2MoDT2, 4MoDT2 and
8MoDT2 as a function of illumination time.
Maximum value of CO2 concentration was reached after about 3 minutes,
then it decreased approaching a steady state value. Carbon dioxide yield
decreased with the increasing molybdenum content: after 120 minutes, it was
64 ppm, 36 ppm and 20 ppm on 2MoDT2, 4MoDT2, 8MoDT2 respectively
and on all these catalysts no formation of benzene and cyclohexene was
observed.
The obtained results evidenced that under UV irradiation of the catalysts,
in the presence of oxygen, cyclohexanol was simply oxidized into CO2
showing that it is not an intermediate compound of cyclohexane oxidative
dehydrogenation reaction. The cyclohexanol oxidation rate decreased with
molybdenum content (Figure 62) probably because Mo-species poison sites
of titania surface which lead to total oxidation of cyclohexanol.
Many authors showed that there is the formation of carboxylic acids as
intermediates during the photooxidation of alcohols both in gas and in liquid
phase Guillard et al. (2002), Pillai and Sahle-Demessie (2002). These
species are strongly adsorbed on titania surface Ekstrom and McQuillan
(1999) and their photodegradation in presence of oxygen leads to carbon
dioxide production Jiang et al. (2004). In particular, under UV irradiation,
CO2 is formed by ethanol through a sequence of two step including
oxidation of the alcohol to form adsorbed acetic acid and its following
decarboxylation with formation of CO2Coronado et al. (2003).
On the basis of this last reaction mechanism, the general scheme of the
photooxidation of cyclohexanol involving titania surface may be expressed
in a schematic way as follows:

85

Chapter V

Cyclohexanol (ads)

+O2

Adsorbed intermediate compounds


(probably carboxylic acids)
-CO2

(g)
Another adsorbed intermediate compound
+O2

Another adsorbed intermediate compound


-CO2

(g)
Etc.
Cyclohexanol may be oxidized into strongly adsorbed intermediate
compounds (probably carboxylic acids) which, in turn, are decarboxylated
into CO2 (which desorbs from catalyst surface) and others adsorbed
intermediate compounds. These last compounds are then oxidized and
decarboxylated by a series of consecutive steps which lead to formation of
CO2 detected in the gas phase during photocatalytic oxidation of
cyclohexanol (Figure 62)

V.3.2

Photocatalytic oxidation of cyclohexene

As cyclohexene has been identified as a principal reaction product and


possible reaction intermediate of photooxidation of cyclohexane,
photocatalytic oxidation of cyclohexene as the reactant was also considered
in the same condition. Cyclohexene conversion (Figure 63) over the three
molybdena systems yielded benzene (Figure 64) and carbon dioxide (Figure
65) as the only detected products.

86

cyclohexene conversion, %

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

12
10
8
8MoDT2

4MoDT2
4

2MoDT2

2
0
0

20

40

60

80

100

120

illumination time, min

Figure 63 Cyclohexene conversion on 2MoDT2, 4MoDT2 and 8MoDT2 as a


function of illumination time
Cyclohexene conversion reached higher value (10% after 110 minutes)
on 8MoDT2 with respect to 2MoDT2 (about 1%) and with respect to
4MoDT2 (about 8%).
120

benzene, ppm

100
80
60

8MoDT2

40

4MoDT2
2MoDT2

20
0
0

20

40

60

80

100

120

illumination time, min

Figure 64 Benzene concentration on 2MoDT2, 4MoDT2 and 8MoDT2 as a


function of illumination time

87

Chapter V

60

CO2, ppm

50
40

2MoDT2

30

4MoDT2

20

8MoDT2

10
0
0

20

40

60

80

100

120

illumination time, min


Figure 65 Carbon dioxide concentration on 2MoDT2, 4MoDT2 and
8MoDT2 as a function of illumination time
Figure 64 shows that benzene production increased with molybdenum
loading. In particularly, after 110 min, benzene concentrations were 7 ppm,
80 ppm and 100 ppm on 2MoDT2, 4MoDT2 and 8MoDT2 respectively.
Carbon dioxide production (Figure 65) showed the opposite trend: it was
higher on 2MoDT2 (50 ppm after 110 min) with respect to 4MoDT2 (12
ppm), and with respect to 8MoDT2 (6 ppm).
The nature of products obtained shows that the reaction is not limited to a
complete oxidation into CO2. In fact, cyclohexene itself is also involved in a
photocatalytic oxidative dehydrogenation reaction giving benzene. It has
been reported that cyclohexene is photooxidized into CO2 by using titania as
photocatalyst Einaga et al. (2002). Moreover a number of examples of
oxidative photocatalysts are based on the use of polyoxometallates, thanks to
their ability to undergo photoinduced electron transfers without irreversible
modifications Ermolenko and Giannotti (1996), Maldotti et al. (1996),
Duncan and Hill (1997), Tanielian (1998) and to control efficiency and
selectivity of the photocatalytic processes Mizuno and Misono (1998). In
particular, Maldotti et al. (2003), showed the photoexcited decatungstate is
able to initiate the selective oxidation of the cyclohexene through hydrogen
abstraction in aqueous media. A key step in the photocatalytic cycle under
aerobic conditions is the subsequent reoxidation of the decatungstate by O2.
Therefore carbon dioxide (Figure 65) detected during photocatalytic tests
on cyclohexene is due to its total oxidation on bare titania whereas oxidative
dehydrogenation to benzene, is probably catalysed by octahedral
polymolybdate on catalyst surface as results obtained by increasing
molybdenum content suggested (Figure 64).

88

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

From these considerations, it is possible to illustrate a probable


mechanism of reaction involving both molybdate and titania.
Total oxidation of cyclohexene on bare titania may be occurred through
the mechanism showed by Einaga et al. (2002) and reported in Figure 66.

Figure 66 Mechanism for oxidation of cyclohexene on bare titania


The hole generated by irradiation of catalyst oxidizes the surface
hydroxyl groups to form the OH radicals which abstract the H atoms of
saturated CH bonds of cyclohexene. The resulting intermediate radicals are
subsequently oxidized by molecular O2. These intermediates are
decomposed to CO2 via the subsequent oxidation processes. Simultaneously,
the photoexcited octahedral molybdate is able to initiate the oxidative
dehydrogenation of adsorbed cyclohexene through hydrogen abstraction (eq.
5) to form benzene which desorbs from catalyst surface (eq. 6).
C6H10(ads) + 4Mo7O246-/TiO2 C6H6(ads) + 4HMo7O246-/TiO2 (5)
C6H6(ads) C6H6(g)

(6)

O2-(lattice) + h+ O-

(7)

HMo7O246-/TiO2 + O- Mo7O246-/TiO2 + OH-

(8)

OH- + h+ OH.

(9)

OH. + OH. H2O2

(10)

H2O2 H2O + O2

(11)

Photoreduced molybdate is than regenerated by O- (eq 8) which is formed


by the reaction between lattice oxygen and positive hole (eq. 7). OH- formed
by reoxidation of reduced molybdate is involved in a series of reactions

89

Chapter V

which lead to formation of water (eq. 9-11). In summary, the formation of


benzene may be occur mainly according to a photocatalytic cycle that
involves the molybdate both in the oxidized and in the photoreduced forms
(Redox mechanism).
On the basis of these results and considerations, it is possible to
hypothesize a mechanism of oxidative dehydrogenation of cyclohexane.
Cyclohexane may be oxidized into CO2 and H2O because of unselective sites
of titania according to mechanism showed in Figure 59 while Mo-species
may be responsible of oxidative dehydrogenation of cyclohexane to
cyclohexene and cyclohexene further oxy-dehydrogenation to benzene
according to the following reactions.
TiO2 + h e- + h+
Mo7O246-/TiO2 + h

(12)
Mo7O246-/TiO2 (polymolybdate photoexcited)
(13)

C6H12(ads) + 2Mo7O246-/TiO2 C6H10(ads) + 2HMo7O246-/TiO2


(14)
C6H10(ads) C6H10(g)

(15)

C6H10(ads) + 4Mo7O246-/TiO2 C6H6(ads) + 4HMo7O246-/TiO2 (5)

V.3.3

C6H6(ads) C6H6(g)

(6)

HMo7O246-/TiO2 + O- Mo7O246-/TiO2+ OH-

(8)

OH- + h+ OH.

(9)

OH. + OH. H2O2

(10)

H2O2 H2O + O2

(11)

Role of the sulphate in the reaction mechanism

A possible active role of sulphate in promoting the selectivity to benzene


was suggested. Sulphate present on catalyst surface could realize a hydrogen

90

Experimental results: Photocatalytic oxidation of cyclohexane on MoOx/TiO2

abstraction from an adsorbed cyclohexane molecule due to its strong basic


properties but the results shown in the Figure 58 reveal that there is no
formation of benzene over sulphated unsupported titania suggesting that
sulphate could interact with octahedral molybdate but not with cyclohexane
according to following reactions:
TiO2 + h e- + h+
Mo7O246-/TiO2 + h

(12)
Mo7O246-/TiO2 (polymolybdate photoexcited)
(13)

O2-(lattice) + h+ O-

(7)

C6H12(ads) + 2Mo7O246-/TiO2 C6H10(ads) + 2HMo7O246-/TiO2


(14)
C6H10(ads) C6H10(g)

(15)

C6H10(ads) + 4Mo7O246-/TiO2 C6H6(ads) + 4HMo7O246-/TiO2 (5)


HMo7O246-/TiO2 + O- Mo7O246-/TiO2+ OHHMo7O246-/TiO2 + SO42-/TiO2

(8)

Mo7O246-/TiO2 + SO32-/TiO2 + OH.


(16)

OH. + OH. H2O2

(10)

H2O2 H2O + O2

(11)

Probably surface sulphate- reoxidizes reduced molybdate (eq. 16) with


together lattice oxygen (eq. 8) increasing the concentration of Mo7O246which increases the rate of the reaction between molybdate and cyclohexane
(eq. 14) improving the production of dehydrogenated compounds
(cyclohexene and benzene).

91

VI Photocatalytic
flat-plate reactor

In recent years, a flat-plate photocatalytic reactor was considered as a


kind of efficient solar photoreactor for commercial application because this
reactor has a high surface-area-to-volume ratio in the photoreactor and can
treat the polluted air and water quickly and efficiently Turchi et al. (1993).
Photocatalytic cyclohexane oxidative dehydrogenation tests were carried
out on MoDT2s catalysts by using a photocatalytic flat-plate reactor in order
to verify the possibility of obtaining the same reaction products with
different configurations of reactor.

VI.1 Experimental set up apparatus and photocatalytic tests


conditions
The photocatalytic plate reactor was realised in steel with a quartz
window (reactor volume: 0.7 l). All the gas pipes ( e.d.) are of Teflon,
connections are made with Swagelok unions and two, and three way Nupro
valves. All the connections are in stainless steel to avoid hydrocarbon
adsorption on the walls and any corrosion due to the presence of water.
Oxygen and nitrogen were fed from cylinders, nitrogen being the carrier gas
for cyclohexane and water vaporized from two temperature controlled
saturators and by changing temperature and N2 flow, it is possible to obtain
different concentrations in the reaction feed. To feed a correct flow,
flowmeters for N2 and O2 are used. The quartz window of the reactor was
illuminated by an UV light source (MEDIUM PRESSURE MERCURY
LAMP, 400W, PHILIPS) in a dark box. The temperature reaction was
120C. In order to control the reaction temperature, a heater system was
installed under the reactor. The gas composition was determined by on line
quadrupole mass detector (Genesys) that can analyze the inlet and outlet
reactor gas, introduced into two different heated capillaries made from silica.
The mass spectrometer was connected to a PC for data acquisition. A

Chapter VI

schematic picture of the photocatalytic flat-plate reactor is reported in Figure


67.

Gas IN

Gas Out

Figure 67 Photocatalytic flat-plate reactor


An aqueous slurry of catalyst powder was used to coat a steel plate. The
coated particles were dried at 90 C for 8 hours in order to remove the excess
of physisorbed water. This treatment resulted in uniform coating well
adhering to the steel plate. The amount of deposited catalyst, evaluated by
weighing the steel plate before and after the coating treatment was 1 g. The
still plate was then located in the reactor. Catalytic tests were carried out
feeding 830 Ncc/min N2 stream containing 1 % cyclohexane, 1.5 % oxygen
and adding 1.6 % water to minimise catalyst photodeactivation. Lamp was
switched on after complete adsorption of cyclohexane on catalyst surface.

VI.2

Photocatalytic activity tests

The analysis of the reaction products in the outlet stream disclosed the
presence of benzene, cyclohexene and CO2. In order to verify that
cyclohexane was converted in a heterogeneous photocatalytic process, blank
experiments were performed. A control test was carried out with the reactor
loaded with uncoated aluminium plate. No conversion of cyclohexane was
detected during this test, indicating the necessity of the catalyst for the
observed reaction. A second test was performed with the catalyst loaded
reactor, without switching on the lamp even after establishing the
cyclohexane adsorption equilibrium. In these conditions the composition of
the outlet reactor was identical to that of the reactor inlet, indicating that no
reaction occurred in dark conditions. The results indicate that the
photocatalysed oxy-dehydrogenation of cyclohexane to cyclohexene and
benzene together with deep oxidation to carbon dioxide occurs irrespective
of the reactor configuration.
The comparison of cyclohexane conversion between 2MoDT2, 4MoDT2
and 8MoDT2 is shown in Figure 68.

94

cyclohexane conversion, %

Photocatalytic flat plate reactor

20
18
16
14
12
10
8
6
4
2
0

2MoDT2
4MoDT2
8MoDT2

10

20

30

40

50

60

70

illumination time, min

Figure 68 Cyclohexane conversion on MoDT2s catalysts as a function of


illumination time
On all catalysts a maximum value was reached after about 3 minutes,
then activity decreased approaching a steady state conversion. On 2MoDT2
maximum cyclohexane conversion was 18 %, decreasing to about 4.5 % in
10 minutes. On 4MoDT2 maximum conversion was lower (about 10%). It
was about 2% after 20 minutes. Increasing Mo loading up to 8 wt% MoO3
the initial maximum conversion was lower, about 15%, while steady state
conversion was 2.3% after 30 minutes. Therefore the progressive coverage
of titania surface by MoOx species resulted in decreased initial and steady
state cyclohexane conversion.
While on DT2 the only reaction product was CO2, all MoOx/TiO2
catalysts exhibited unexpected high selectivity to benzene (Figure 69). On
2MoDT2 selectivity to benzene was about 3% after 35 minutes. On
4MoDT2 maximum selectivity to benzene reached 14%. On 8MoDT2 higher
selectivity to benzene was observed (34% after 15 minutes). On all catalysts
the presence of very low amounts of cyclohexene in the reaction products
were detected.

95

benzene selectivity, %

Chapter VI

40
35
30

2MoDT2

25

4MoDT2

20

8MoD2

15
10
5
0
0

10

20

30

40

50

60

70

illumination time, min

Figure 69 Selectivity to benzene on MoDT2s catalysts as a function of


illumination time
Selectivity to CO2 on MoDT2s catalysts as a function of illumination
time is reported in Figure 70.

CO2 selectivity, %

12
10
8

2MoDT2
4MoDT2

8MoDT2
4
2
0
0

20

40

60

80

illumination time, min

Figure 70 Selectivity to CO2 on MoDT2s catalysts as a function of


illumination time
It can be seen that selectivity to CO2 decreased with molybdenum
loading. On 2MoDT2, 4MoDT2 and 8MoDT2 steady state selectivities to

96

Photocatalytic flat plate reactor

CO2 were about 10 %, 4 % and 2 % respectively. The obtained results


confirmed that the increasing presence of MoOx species on the titania
increase the selectivity of the catalyst indicating that the interaction between
titania and supported molybdenum oxide plays an essential role in changing
the catalyst selectivity.

97

VII Photocatalytic
fluidized bed reactor

VII.1 Photocatalytic fluidized bed reactor design


A two dimensional fluidized bed photoreactor was designed in order to
improve both exposure of the catalysts to light irradiation and a good contact
between reactants and catalyst. The powder used for the design of the two
dimensional photocatalytic fluidized bed reactor was Al2O3, (Aldrich)
with a Sauter average diameter of 50 m; its density is 3970 kg/m3. The
equation 17 was used to compute the minimum fluidization velocity (Umf):

Re mf = (1135.7 + 0.0408 Ar )0.5 33.7

(17)

where:

Re mf =

d p f U mf

= Reynolds number at mimimum fluidization

velocity

Ar =
dp =

d 3p f ( s f ) g

2
1

x
di
pi

= Archimedes number

= Sauter average diameter

f = fluidizing gas density


s = particle density of fluidizing solid
U mf = minimum fluidization velocity

Chapter VII

xi = weight fraction of particles of size d pi


g = acceleration of gravity
= viscosity of fluidizing gas
The cross-section was evaluated by using a gas superficial velocity equal
to 4*Umf.
A schematic picture of the photocatalytic fluidized bed reactor is reported
in Figure 71

Gas OUT

Gas IN
Figure 71 Photocatalytic fluidized bed reactor
The gas (flow rate: 830 Ncc/min) was introduced into the fluidized bed
reactor with 40 mm x 10 mm cross section. The wall is made of pyrex-glass
(2mm in thickness). A bronze filter (5 m size) was used for gas feeding to
provide uniform gas distribution. Its walls, 230 mm in height, were made of
2mm thick pyrex-glass.
During transient condition, some elutriation phenomena were observed.
Thus in order to decrease the amount of transported particles, an expanding
section (50 mm x 50mm cross-section at the top) and a cyclone, specifically
designed, (Figure 72) are located on the top and at the outlet of the reactor
respectively. The cyclone was designed by utilising the Lapple configuration
Lapple (1973) and the diameter of its body was calculated by the
following equation:

D=

100

Q
0.5 0.25 Vi

(18)

Photocatalytic fluidized bed reactor

where:
Q = overall volumetric flow rate
Vi = cyclone inlet velocity = 2 m/s.
The other geometrical parameters (Table 12) of the cyclone were
calculated by the relations of the Lapple configuration Lapple (1973).

De
b
h

S
D

D = cyclone body diameter


a = cyclone entrance length
b = cyclone entrance length width
B = solids eductor diameter
h = cyclone barrel length
H = cyclone height
De = cyclone outlet diameter
S = vortex tube length

B
Figure 72 Schematic picture of the cyclone
Table 12 Geometrical parameters of the cyclone
D, mm
a, mm
b, mm
B, mm
De, mm
h, mm
H, mm
S, mm

7.45
3.73
1.86
1.86
3.73
14.9
29.8
4.66

The fluidized bed reactor is illuminated by two UV light sources


(BLACKLIGHT BLUE, 160W, PHILIPS or EYE MERCURY LAMP,
125W) in a dark box. In order to control the reaction temperature, a PID
controller connected to a heater system was installed near the reactor.
Figure 73 reports a schematic picture of the experimental set up. Catalytic
tests were carried out feeding 830 Ncc/min N2 stream containing 1000 ppm
cyclohexane, 1500 ppm oxygen and 1600 ppm water.

101

Chapter VII

Figure 73 Schematic picture of the experimental apparatus

VII.2 Preliminary results


Titania is typically a solid of group C in the particle distribution proposed
by Geldart (1973). This type of particles exhibit cohesive tendencies, and as
the gas flow is further increased, usually rathole; the gas opens channels
that extend from the gas distributor to the surface. If channels are not
formed, the whole bed will lift as a piston. At higher velocities or with
mechanical agitation or vibration, this type of particle will fluidize but with
the appearance of clumps or clusters of particles. These phenomena play a
negative role on particle light exposition. For testing Mo-titania catalyst, in
order to obtain good fluidization, physical mixtures with Al2O3 at
different percentages of Mo-titania catalysts were experimented. Al2O3
used in the reactor has physical characteristics of group A. When gas is
passed upward through a bed of particles of groups A, B, or D, friction
causes a pressure drop expressed by the Carman-Kozeny fixed-bed
correlation. As the gas velocity is increased, the pressure drop increases until
it equals the weight of the bed divided by the cross-sectional area. This
velocity is called minimum fluidizing velocity, Umf. When this point is
reached, the bed of group A particles will expand uniformly until at some
higher velocity gas bubbles will form (minimum bubbling velocity, Umb).
This process can avoid the appearance of clumps or clusters of titania
particles and improve light exposure. In Figure 74 the outlet benzene

102

Photocatalytic fluidized bed reactor

benzene, ppm

concentrations obtained by mixing 5g of 4MoDT2 with 40g of Al2O3 in


comparison with that obtained mixing 10g of 4MoDT2 are reported.

50
45
40
35
30
25
20
15
10
5
0

10g 4MoDT2

5g 4MoDT2

10

20

30

40

50

60

illumination time, min


Figure 74 Outlet reactor concentration of benzene on 4MoDT2 catalyst as
function of illumination time
It can be seen that benzene production increased by increasing catalyst
amount. It was about 8 ppm with 5g of 4MoDT2 and about 37 ppm with 10g
of the same catalyst. The optimal mixture was composed of a physical
mixture of 14g of catalyst and 63g of Al2O3.
In order to verify that cyclohexane was converted in a heterogeneous
photocatalytic process, a control test was carried out with the reactor loaded
with Al2O3 alone. No conversion of cyclohexane was detected during this
test, indicating the necessity of the catalyst for the observed reaction. A
second test was performed with the catalyst loaded in the reactor, without
switching on the lamps, after cyclohexane adsorption equilibrium was
reached. In these conditions the composition of the outlet reactor was
identical to that of the reactor inlet, indicating that no reaction occurred in
dark conditions.A photocatalytic test at 120C was also performed by mixing
14g of DT2 with 63 g of Al2O3 in which no catalyst activity was found.
A typical trend of fluidized bed photocatalytic test is reported in Figure
75 with reference to 8MoDT2. When the lamps were switched on, the
cyclohexane outlet concentration immediately decreased reaching a steady
state value corresponding to about 4% cyclohexane conversion after about
10 minutes. In the same figure the change of oxygen outlet concentration is
also reported showing a trend similar to that of cyclohexane. The analysis of
products in the outlet stream disclosed the presence of benzene and

103

Chapter VII

cyclohexene, as identified from the characteristic fragments m/z = 78, 77,


76, 74, 63, 52, 51, 50 (fragment 78 reported in ) and 82, 67, 54, respectively
(fragment 67 reported in Fig. 2). No presence of carbon dioxide was
disclosed, as detected by the NDIR analyser. The outlet concentration of
benzene progressively increased reaching a steady state value after about 50
minutes. A similar trend was shown by cyclohexene concentration, however
the values are very much lower with respect to benzene. No deactivation of
catalyst was observed during photocatalytic tests.

m/z = 84

MS signal, a.u.

Lamps on

Lamps off

m/z = 32

m/z = 78
m/z = 67

20

40

60

80

100

120

140

time, min
Figure 75 Outlet reactor concentration (a.u.) of cyclohexane, oxygen
benzene and cyclohexene and as a function of run time
In Table 13 the comparison of photocatalytic performances in the
fluidized bed reactor of 4MoDT2 and 8MoDT2 catalysts is summarised.
Table 13 Comparison between performance of 4MoDT2 and 8MoDT2
catalysts. Reaction temperature: 70 C. UV sources:two blacklight blue,
160 W, PHILIPS
Catalyst

cyclohexane
conversion,
%

benzene
selectivity,
%

carbon
dioxide
selectivity,
%

cyclohexene
selectivity,
%

4MoDT2

67

32.2

0.8

8MoDT2

99

1.0

104

Photocatalytic fluidized bed reactor

On 4MoDT2 selectivity to benzene reached 67%, while that to carbon


dioxide was 32.2 %. On 8MoDT2 about 100% selectivity to benzene was
obtained and no formation of carbon dioxide was detected. With both
catalysts the presence of a very low amount of cyclohexene in the reaction
products was detected (selectivity was about 0.8% on 4MoDT2 and 1% on
8MoDT2). Therefore, the presence of MoOx species on the surface of titania
increases the selectivity of the catalyst as the molybdenum content increases.
The effect of UV sources was evaluated on 8MoDT2 catalyst which was
the most selective to benzene. In Table 14 the obtained results are reported.
Table 14 Effect of UV sources on 8MoDT2 catalyst. Reaction temperature:
100C
UV source
BLACKLIGHT BLUE, 160W
MERCURY LAMP , 125W

cyclohexane
conversion, %

benzene outlet
concentration, ppm

28.5

4.7

46.55

Cyclohexane conversion increased up to 4.7 % by using two mercury


lamps with a power of 125W. In this case benzene outlet concentration was
higher (46.6 ppm). In order to explain this last result, it is useful to examine
the emission characteristics in the UVA and UVB range for both types of
UV source (Table 15).
Table 15 Emission characteristics in the UVA and UVB range
UV source

UVA
emission, W

UVB
emission, W

BLACKLIGHT BLUE, 160W

1.8

0.04

MERCURY LAMP , 125W

4.2

0.1

It can be seen that MERCURY LAMP has both UVA and UVB emission
higher than BLACKLIGHT BLUE LAMP indicating a higher flux of
photons emitted in the UVA region by the first UV source. This last aspect
could explain the better catalytic performances obtained by using
MERCURY LAMP.
The effect of reaction temperature on cyclohexane conversion and
benzene production obtained by loading 14 g of 8MoDT2 catalyst mixed
with 63 g of Al2O3 is shown in Figure 76.

105

Chapter VII

80

70

60
7

50

6
5

40

30

cyclohexane conversion

benzene outlet concentration

benzene, ppm

cyclohexane conversion, %

10

20
10

1
0

40

60

80

100

120

140

160

0
180

T, C
Figure 76 Effect of reaction temperature on cyclohexane conversion and
benzene outlet concentration. UV sources: two eye mercury lamps, 125 W
Cyclohexane conversion and benzene production increased with
increasing reaction temperature; moreover the presence of very low amounts
of cyclohexene in the reaction products was detected. Moreover there is only
a small increment of catalytic activity between 120 C and 160 C indicating
that above a certain value of temperature, the reaction rate off and becomes
independent of temperature. This limit depends on the geometry and on the
working conditions of the photoreactor which probably ensures a total
absorption of efficient photons.

VII.3 MoOx supported on TiO2/Al2O3 sample


An alternative to the physical mixing between titania based catalyst and
Al2O3 is to realize TiO2-Al2O3 mixed oxide catalytic supports. In general,
sol-gel methods have been preferred to produce such mixed oxide systems.
When titanium and aluminium alkoxides are used as support precursors,
small pore diameters of the order of 2-3 nm are obtained Ramirez et al.
(1993). Organic polymers were used to control textural properties of
catalysts supported on single oxides Basmadjian et al. (1962), Trimm and
Stanislaus (1986), as well as in some mixed oxides systems such as Al2O3-

106

Photocatalytic fluidized bed reactor

SiO2 Mascia (1995), Snel (1984). The effect of the type of polymeric
additive, its amount, molecular weight and sequence of addition of the
reactants on the surface area and pore-size distribution in TiO2-Al2O3 mixed
oxides were reported Klimova et al. (1990). Wei et al. (1990) have
examined the effect of the preparation method on the morphology of titaniaalumina support. It is shown that the grafting technique gives the best
dispersion of TiO2 on Al2O3.
Mixed oxides were used as catalyst in photocatalytic processes mainly for
environmental applications. Two kinds of chemical vapour deposition
approaches have been employed for the preparation of the photocatalyst of
titanium dioxide supported on alumina Zhang and Zhou (2005). One was
simultaneous deposition and calcinations (one-step process); the other was
preliminary gas impregnation of the support followed by a decomposition
step (two-step process). The results of characterization indicated that the
structure of the support was destroyed by a two-step process because of pore
blocking. It was found that the one-step process resulted in a superior
photocatalyst; this was attributed to higher external surface concentration
and more perfect crystalline structure of TiO2. Titania, synthesized through a
sol-gel procedure with acetylacetone chelating agent, was immobilized on Al2O3 and deposited with photoreduced Ag Chen et al. (2005). The
prepared catalysts were characterized and applied for decolourization of
methyl orange. The photoactivity of TiO2/-Al2O3 is affected by the H2O/Ti
molar ratio applied in the sol-gel process. Incorporating photoreduced Ag to
TiO2/-Al2O3 leads to reduction reaction of methyl orange in addition. to
oxidation reaction, and yields a significant increment in decolorization
efficiency. The preparation of two sets of polycrystalline photocatalysts
prepared by supporting TiO2 (anatase) on TiO2 (rutile) or TiO2 (anatase) on
Al2O3 is reported Loddo et al. (1999). The powders were prepared by a
wet impregnation method using titanium (IV) isopropylate. Both sets of TiO2
(anatase) supported samples resulted photoactive and the photoactivity
toward 4-nitrophenol photodegradation increased by increasing the content
of the anatase phase.

VII.3.1 TiO2-Al2O3 sample preparation


Titania-alumina support was prepared by dispersing DT2 powder in a
boehmite sol (10 wt% of Condea Pural in bydistilled water, pH< 2 by
HNO3). The sol was gelled by slight heating until it was too viscous to stir.
The gel was thus dried at 120C for 3 hours and then calcined at 500 C for
2 hours. After calcinations the solid was crushed and sieved to recover 50
m diameter particles. (NH4)6 Mo7O244H2O was used as MoO3 precursor to
impregnate the alumina-TiO2 support (30wt% TiO2, 113 m2/g). Finally the
powder was dried at 120 C for 12 hours and calcined in air at 400 C for 3

107

Chapter VII

hours. The calcined catalyst (10MoDTAl) contained 10 wt % of MoO3 as


nominal loading.

VII.3.2 Results
In Figure 77 the comparison between the results obtained by mixing 14g
of 8MoDT2 (Mo8) and 63 g of Al2O3 and by loading 30 g of 10MoDTAl
(Mo10) are reported.
200

180

Cyclohexane conversion Mo8

160

cyclohexane conversion Mo10

140

benzene concentration Mo8

120

benzene concentration Mo10

100

80

60

40

20

0
0

20

40

60

benzene, ppm

cyclohexane conversion, %

10

80

illumination time, min


Figure 77 Comparison between cyclohexane conversion and benzene outlet
concentration obtained on Mo8 and Mo10. Reaction temperature: 120 C.
UV sources: two eye mercury lamps, 125 W
On 10MoDTAl cyclohexane steady state conversion was higher (about
9%) with respect to Mo8 (about 7%). On Mo8 benzene outlet concentration
reached steady state value (70 ppm) after 10 min, while on 10MoDTAl it
was higher (about 90 ppm). This result may come from the fluidization of a
higher catalyst mass and consequently of a higher gas contact time together
with better utilization of UV light than through Al2O3 diluent bed.
Similar results have been reported by Lynette, and Gregory (1992) on silica
support with respect Al2O3.

108

Photocatalytic fluidized bed reactor

VII.4 Photocatalytic fixed and fluidized bed reactor comparison


Figure 78 shows cyclohexane conversion as a function of illumination
time on 8MoDT2 catalyst using the fluidized bed and the annular fixed bed
reactor.

cyclohexane conversion, %

18
16
14
12

fixed bed reactor

10

fluidized bed reactor

8
6
4
2
0
0

20

40

60

80

100

120

illumination time, min


Figure 78 Cyclohexane conversion as a function of illumination time on
8MoDT2 catalyst using the fluidized bed and the annular fixed bed reactor.
In the fixed bed reactor, cyclohexane conversion reached a maximum
value after about 5 minutes, then activity decreased approaching a steady
state conversion. Maximum conversion was about 15 %, decreasing to 3% in
15 minutes. The steady state conversion was 2% after 30 minutes. The
behaviour is completely different when the fluidized bed reactor is used. In
this case, when the lamps were switched on cyclohexane conversion
immediately increased reaching a steady state value corresponding to 4%
cyclohexane conversion after about 10 minutes. This result evidences that
the fluidization enhanced the cyclohexane conversion. In Figure 79 the
comparison of benzene outlet concentration on Mo8 between fixed bed and
fluidized bed photoreactor is reported.

109

Chapter VII

45

benzene, ppm

40
35
30

fixed bed reactor

25

fluidized bed reactor

20
15
10
5
0
0

20

40

60

80

100

120

illumination time, min


Figure 79 Comparison of benzene outlet concentration on 8MoDT2 between
fixed bed and fluidized bed reactor.
It can be seen that, by using the fixed bed reactor, the outlet concentration
of benzene progressively increased reaching a maximum value of about 17
ppm after 37 minutes. Then it decreases to a steady state value (about 14
ppm) reached after about 80 minutes. Instead, in fluidized bed reactor,
benzene outlet concentration progressively increased reaching a steady state
value of 39 ppm (higher than that obtained in the fixed bed reactor) after
about 20 minutes.
In Table 16 and Table 17 the comparison of photocatalytic performances
in the steady state of 4MoDT2 and 8MoDT2 catalysts in the fixed and
fluidized bed reactors is respectively reported.
Table 16 Comparison between performance of 4MoDT2 and 8MoDT2
catalysts in the fixed bed reactor
Catalyst

cyclohexane
conversion,
%

benzene
selectivity,
%

carbon
dioxide
selectivity,
%

cyclohexene
selectivity,
%

4MoDT2

27

0.7

8MoDT2

65

1.5

110

Photocatalytic fluidized bed reactor

Table 17 Comparison between performance of 4MoDT2 and 8MoDT2


catalysts in the fluidized bed reactor
Catalyst

cyclohexane
conversion,
%

benzene
selectivity,
%

carbon
dioxide
selectivity,
%

cyclohexene
selectivity,
%

4MoDT2

67

32.2

0.8

8MoDT2

99

1.0

By using the fixed bed reactor, cyclohexane conversion was 4 % on


4MoDT2 and 2 % on 8MoDT2. On 4MoDT2 selectivity to benzene reached
31%, while selectivity to carbon dioxide was 8%. On 8MoDT2 higher
selectivity to benzene was observed (65%) and selectivity to carbon dioxide
was 5%. With both catalysts the presence of very low amounts of
cyclohexene in the reaction products was detected (selectivity was about
0.7% on 4MoDT2 and 1.5% on 8MoDT2). In the fluidized bed, cyclohexane
conversion was 6% and 4% on 4MoDT2 and 8MoDT2 respectively. On
4MoDT2 selectivity to benzene was 67%, while that to carbon dioxide was
32.2 %. On 8MoDT2 about 100% selectivity to benzene was obtained and
no formation of carbon dioxide was detected. Moreover selectivity to
cyclohexene was about 0.8% on 4MoDT2 and 1% on 8MoDT2.
The obtained results showed that both in the fixed bed reactor and in the
fluidized bed reactor, cyclohexane conversion and selectivity to carbon
dioxide decreased with increasing molybdenum content whereas selectivity
to benzene and cyclohexene increased. Moreover with both catalysts, the
performances obtained in the fluidized bed reactor are better than those
obtained in the fixed bed reactor. Cyclohexane conversion and outlet
benzene concentration characteristics suggested that there was no catalyst
deactivation when the fluidized bed reactor was used. Selectivity data
evidenced that in the fixed bed reactor, total carbon mass balance (evaluated
by comparing the inlet carbon as cyclohexane and the outlet carbon as the
sum of unconverted cyclohexane and outlet benzene, cyclohexene, carbon
monoxide and carbon dioxide) was not close to 100% while the fluidized
bed reactor it was almost complete. Carbon mass balance as function of
illumination time during the catalytic tests in the fixed bed reactor on
4MoDT and 8MoDT catalysts has been reported in Figure 80. The carbon
mass balance is closed to about 99 % on 8MoDT and 90% on 4MoDT in
steady state conditions.

111

total carbon mass balance, %

Chapter VII

100
95
90
85

8MoDT2

80

4MoDT2
75
70
0

20

40

60

80

100

illumination time, min


Figure 80 Total carbon mass balance in the fixed bed reactor as a function
of illumination time
Figure 81 and Figure 82 show TG-MS results of 4MoDT2 and 8MoDT2
catalyst after photocatalytic tests.

112

Photocatalytic fluidized bed reactor

100

0.3

TG
99

0.2

TG ,%

0.1
98
0.0

DTG, %/min

DTG

97
-0.1

96

200

400

-0.2
800

600

Temperature,C

Figure 81 TG-MS results on 4MoDT2 catalyst after photocatalytic results in


the fixed bed reactor
100

0.4

TG
0.2

TG, %

98
0.0
97

DTG, %/min

DTG

99

-0.2
96

95

200

400

600

-0.4
800

Temperature, C

Figure 82 TG-MS results on 8MoDT2 catalyst after photocatalytic results in


the fixed bed reactor
Thermogravimetric analysis on 8MoDT2 and on 4MoDT2 showed the
presence of carbonaceous species adsorbed on catalyst surface since there
was the formation of carbon dioxide in the range 200-430 C (as evidenced

113

Chapter VII

by carbon dioxide characteristic fragments m/z = 44) that could explain the
carbon deficit checked during the reaction in the fixed bed reactor. Moreover
an interesting result was the total absence of surface sulphates evidenced by
the absence of SO2 (g) (fragments m/z = 48 and 64).
A literature summary of catalyst deactivation in gas-solid photocatalysis
was reported by Sauer and Ollis (1996). They showed that deactivation is a
very commonly observed phenomenon, especially for single pass flow
reactor. By using the fixed bed reactor, tests performed on both catalysts
revealed that rapid decays of activity occurrred in the first minutes on stream
to reach a stable value for conversion under steady state conditions. This
initial decrease of activity could be due to a poisoning of the surface by
carbonaceous species. It can be supposed that the effect could be due either
to a strong adsorption of the reactant or the products (benzene, CO2 and
cyclohexene), or to an adsorption of carboxylate or other carbonaceous
species yielding a blocking of a part of the surface sites.
In order to elucidate this last aspect, TPD coupled with Mass
Spectrometer and CO, CO2 continuous analyzers, in nitrogen flow, of both
catalysts after activity measurement in the fixed bed reactor in the range 20500 C were performed (Figure 83).
300
250

CO2

200

MS signal, a.u.

CO2, ppm

m/z = 78
m/z = 67

150
100
50
0
30

130

230

330

Temperature, C

430

Figure 83 Outlet reactor concentration (MS signal) of cyclohexene, and


benzene and (ppm) of carbon dioxide (NDIR analyzer) as a function of
temperarure on 4MoDT2 after activity measurements in the fixed bed
reactor.

114

Photocatalytic fluidized bed reactor

m/z= 64, a.u.

Desorption of cyclohexene (m/z = 67) from 100 C up to 210 C, of


benzene (m/z=78) from 160 C to 260 C, and of CO2 from about 150 C up
to 500 C were observed. No others products were found. Similar results
were obtained for 8MoDT2 sample. Therefore, part of products remained
adsorbed on the surface and could be related to the observed deactivation. It
must be also noted that deactivated samples did not appear brownished, as
found by Einaga et al.(2002). They identified carbon deposits on TiO2
surface in heterogeneous photocatalytic decomposition of benzene, toluene,
cyclohexane and cyclohexene, finding that deactivated TiO2 catalysts were
photochemically regenerated in the presence of water vapour and the carbon
deposits were decomposed to COx. Therefore preliminary characterization of
deactivated catalysts does not give any evidence for carbon deposit
formation.TG-MS analysis on 4MoDT2 and 8MoDT2 catalysts after activity
measurements in fluidized bed reactor evidenced the presence of surface
sulphate and the absence of adsorbed carbonaceous compounds. According
to these results, TPD tests performed on same catalysts (with the same
conditions described above), showed the only desorption of SO2 (m/z = 64)
from 270 C (Figure 84). Desorption of benzene, cyclohexene and CO2 was
not detected

200

250

300

350

400

450

500

Temperature, C
Figure 84 Outlet reactor concentration (MS signal) of SO2 as a function of
temperature on 4MoDT2 after activity measurements in the fluidized bed
reactor
These last results confirm the presence of sulphate on catalysts surface
after photocatalytic tests in the fluidized bed reactor. The non-appearance of

115

Chapter VII

the reaction products on catalyst surface could explain the absence of


catalyst deactivation observed in the fluidized bed reactor. Moreover TG-MS
and TPD results suggest that the deactivation phenomenon in the fixed bed
reactor is probably also due to the disappearance of sulphate species.
The surface chemical states of the samples were examined by X-ray
photoelectron spectroscopy (XPS). This last type of analysis was performed
with a Kratos Axis Ultra instrument working with a monochromatic Al K
radiation. Mo 3d, O 1s and C 1s bands and survey spectra were recorded.
Binding energies were calibrated by fixing the C(CH) contribution of C 1s
at 284.8 eV. Further details on XPS experiments and data treatments are
given elsewhere Dury et al. (2003). Figure 85 summarizes the difference
of Mon+ contributions found by XPS for fresh 8MoDT2 catalyst and that
recovered after photocatalytic test both in the fixed bed reactor and in
fluidized bed reactor.
120
100

80

Mo5+
Mo6+

60
40
20
0

fresh

after fixed bed


reactor

after fluidized
bed reactor

Figure 85 Difference of Mon+ contributions found by XPS for fresh 8MoDT2


catalyst and that recovered after photocatalytic test both in the fixed bed
reactor and in fluidized bed reactor
Fresh 8MoDT2 exhibited mainly Mo6+ and a small amount of Mo5+
species. A surface oxidation of 8MoDT2 occurred during photocatalytic test
in both reactors as an increase of the Mo6+ contribution is observed.
Correspondingly the Mo5+ contribution decreased. Figure 86, Figure 87 and
Figure 88 show the detailed scan spectra of S 2p obtained on fresh 8MoDT2
catalyst and that recovered after photocatalytic test both in the fixed bed
reactor and in fluidized bed reactor.

116

Photocatalytic fluidized bed reactor

Figure 86 Detailed XPS scan spectrum of S 2p on fresh 8MoDT2

117

Chapter VII

Figure 87 Detailed XPS scan spectrum of S 2p on 8MoDT2 after


photocatalytic test in the fixed bed reactor

118

Photocatalytic fluidized bed reactor

Figure 88 XPS scan spectrum of S 2p on 8MoDT2 after photocatalytic test


in the fluidized bed reactor
XPS spectra reveal that sulphur is absent from the surface of sample
recovered after test in the fixed bed reactor whereas its continuing presence
is found with the sample recovered after test in the fluidized bed reactor.
Therefore, taking into account that a higher state of oxidation was found on
8MoDT2 after test both in the fluidized bed reactor and in the fixed bed
reactor with respect to fresh sample, the deactivation of the catalyst may be
correlated with the sulphur disappearance since the sample used in fluidized
conditions remains active and maintains its sulphur content at the surface.

VII.5 Effect of light intensity


As showed in the section VI.1.7, the experimental tests performed by
increasing UV light intensity showed an improvement of catalytic activity in
terms of cyclohexane conversion. Therefore the fluidized bed system has
been modified by introducing the possibility of irradiating the catalyst with
four UV sources simultaneously. The schematic picture of the modified
photocatalytic fluidized bed reactor is reported in Figure 89.

119

Chapter VII

Gas OUT

Gas IN
Figure 89 Schematic picture of the modified photocatalytic fluidized bed
reactor
The catalytic results in terms of cyclohexane conversion and benzene
outlet concentration obtained on 8MoDT2 by irradiating the reactor with two
and four UV light sources (EYE MERCURY LAMP, 125W) are shown in
Figure 90 and Figure 91 respectively.

120

cyclohexane conversion, %

Photocatalytic fluidized bed reactor

16
14
12
10

4 UV sources

2 UV sources

6
4
2
0
0

10

20

30

40

50

60

illumination time, min

70

80

Figure 90 Comparison between cyclohexane conversion obtained by using


two and four UV sources. Reaction temperature: 120 C.
160

benzene, ppm

140
120
100

4 UV sources

80

2 UV sources

60
40
20
0

10

20

30

40

50

60

70

80

illumination time, min


Figure 91 Comparison between cyclohexane conversion obtained by using
two and four UV sources. Reaction temperature: 120 C.
It can be seen that cyclohexane conversion and benzene outlet
concentration were about 7 % and 70 ppm respectively when 2 UV sources
were used and they increased up to about 13 % and 130 ppm in
correspondence of 4 UV sources. The obtained results confirmed that

121

Chapter VII

catalyst activity increases by irradiating the reactor with a greater number of


UV sources. Consequently, photocatalytic activity tests were carried out by
using the modified fluidized bed reactor.

VII.6 Photocatalytic oxidation of cyclohexane on sulphated


MoOx/-Al2O3 catalysts
Since selectivity and yield to benzene increased with the sulphate content
of the support and no formation of benzene was found when molybdenum
was supported on unsulphated - or - alumina in the gas-solid fixed bed
reactor, the photoactivity of MoOx/-Al2O3 catalysts at various sulphation
degrees toward oxidative dehydrogenation of cyclohexane was determined
in the fluidized bed photoreactor.

VII.6.1 Samples preparation


Al2O3 (Puralox SBA 150, SASOL, 144 m2/g) was impregnated with an
aqueous solution of ammonium heptamolybdate (NH4)6 Mo7O244H2O.
Powder samples were dried at 120 C for 12 hours and calcined in air at
400C for 3 hours The calcined catalysts contained 8 wt % of MoO3 (8Mo)
nominal loading. A second impregnation of 8Mo with an aqueous solution of
ammonium sulphate was performed. Powder samples so obtained were dried
at 120 C for 12 hours and calcined in air at 300 C for 3 hours.

VII.6.2 Specific surface area and Thermal analysis


The list of catalysts and their characteristics is reported in Table 18.
Table 18 Catalysts and their characteristics
Catalyst
8Mo
8Mo2S
8Mo4S
8Mo6S

MoO3
nominal
content, wt %
8
8
8
8

SO42nominal
content, wt %
2.4
4.8
7.2

Specific surface area


(BET),
m2/g
144
135
114
111

It can be seen that specific surface area decreased as the sulphate content
increased. Thermogravimetric curves of 8Mo2S, 8Mo4S and 8Mo6S are
reported in Figure 92.

122

Photocatalytic fluidized bed reactor

1.2

TG 8Mo2S
TG 8Mo4S

100

0.8

95

TG, %

DTG 8Mo6S

0.6

90

0.4

TG 8Mo6S

85

0.2

80

DTG, %/min

105

75

DTG 8Mo4S

-0.2

DTG 8Mo2S

70

-0.4

200

400

600

800

T, C
Figure 92 Thermogravimetric curves of 8Mo2S, 8Mo4S and 8Mo6S
catalysts
The first main step (up to about 250C) is associated to water removal,
while the second main step (up to about 380 C) is due to decomposition of
ammonia. The third complex stage of weight loss is associated to different
kind of surface sulphates. DTG curve of powder 8Mo2S showed only one
high temperature peak (centred at about 717 C), probably associated to the
more stable sulphate coordinated on alumina. For 8Mo4S other two
overlapping peaks appeared, centred at 540C and 610C respectively,
probably due to a different sulphate coordinated on alumina. Another DTG
peak, centred at about 430C, appeared. This could be attributed to sulphate
interacting molybdate. This last peak was observed also on 8Mo6S catalyst.
It can be seen that on this last catalyst, four overlapping high temperature
peaks, centred at 523C, 685C and 745C due to different kind of surface
sulphates coordinated on alumina.
Effective surface sulphate content (as SO42-) was evaluated by TG-MS
analysis in the range of temperature in which the sulphate decomposition
occurred. It can be seen that SO42- content measured by thermogravimetric
analysis corresponds to that calculated for the impregnation process (Table
19).

123

Chapter VII

Table 19 Nominal and effective sulphate content


SO42- nominal
content, wt %
2.4
4.8
7.2

Catalyst
8Mo
8Mo2S
8Mo4S
8Mo6S

SO42- effective
content, wt %
2.6
4.9
7.3

VII.6.3 Photocatalytic activity tests


The reaction temperature was 120 C and a physical mixture of 13g of
each catalyst and 58 g of Al2O3 was prepared. In the case of 8Mo sample,
no catalyst activity was found whereas on sulphated catalysts the analysis of
products in the outlet stream disclosed only the presence of cyclohexene and
no formation of benzene and CO2, showing that the only reaction in the
system was the oxidative dehydrogenation of cyclohexane to cyclohexene.
The effect of catalyst sulphate content on both cyclohexane conversion
and cyclohexene outlet concentration is shown in Figure 93
120
cyclohexane conversion

25

cyclohexene outlet
concentration

20

100
80

15

60

10

40

20

0
0

SO42-,wt %

cyclohexene, ppm

cyclohexane conversion, %

30

Figure 93 Cyclohexane conversion and cyclohexene outlet concentration as


a function of SO42- percentage. Reaction temperature: 120 C. UV sources:
four eye mercury lamps, 125 W
As it can be seen, in absence of sulphate, there was found no catalyst
activity. Cyclohexane conversion increased to about 10 % with increasing

124

Photocatalytic fluidized bed reactor

30

120

25

100
cyclohexane conversion
cyclohexene concentration

20

80

15

60

10

40

20

0
80

90

100

110

120

cyclohexene, ppm

cyclohexane conversion,%

SO42- load up to 2.4 wt % and decreased with a further increase in SO42load. A parallel trend can be observed for cyclohexene outlet concentration
behaviour. The effect of reaction temperature on cyclohexane conversion
and cyclohexene outlet concentration on 8Mo2S photocatalyst is shown in
Figure 94.

130

T, C
Figure 94 Cyclohexane conversion and cyclohexene outlet concentration as
a function of reaction temperature
Results reveal that cyclohexane conversion and cyclohexene production
increases with increasing reaction temperature. It has been reported that UV
light irradiation is the primary source of electron-hole pairs that are essential
to initiate the photocatalytic reaction because generally the band gap energy
is too high to be overcome thermally Lim et al. (2000). Therefore, it can
be considered that the increase of cyclohexane conversion may be mostly
due to the increasing collision frequency of the molecule with increasing
reaction temperature.
The results reported above verified the possibility to produce a cyclic
olefin from a cyclic paraffin with a selectivity to cyclic olefin of 100% by
means a photocatalytic process. Light alkanes oxidative dehydrogenation
(ODH) process is today a potential route to obtain olefin Kung (1994),
Cavani and Trifir (1995) in comparison to industrial processes such as
dehydrogenation and steam-cracking. However, catalysts that can be

125

Chapter VII

economically competitive with performances of common catalysts have not


yet been developed. The main problem with most of the redox-type catalysts
studied in ODH is that olefin yields do not exceed typically 30%. The reason
for the low yields is the effective activation of alkenes by these catalysts. In
fact alkenes remain strongly adsorbed on catalyst surface, inducing a high
rate for consecutive deep oxidation of the desired product Kung (1994),
Watling et al. (1996). Recently, a new type of catalyst formulation was
reported to give high yields of olefin in the oxidative conversion of alkanes
Fuchs et al. (2001). In this case olefin yields could reach values above
50%. The composition of these catalysts is similar to that studied for
methane oxidative coupling and ethane ODH Conway et al. (1991) and
contained a basic oxide (such as MgO) mixed with rare-earth oxide and
promoted by alkali metal (Li, Na) oxide and halogen (Cl, Br) Herskowitz
et al. (1996).
It has been reported that the oxidative dehydrogenation of alkanes such as
propane Boisdron et al. (1995) followed first-order kinetics in paraffin
concentration. Therefore apparent first-order rate constant (k) for
photocatalytic oxidative dehydrogenation of cyclohexane to cyclohexene on
8Mo2S catalyst were calculated according to the following equation at
different reaction temperatures:

k=

Q
[ ln (1 x )]
w

(19)

where Q is the overall volumetric flow rate, w is the mass of catalyst and
x is the cyclohexane conversion employed under steady state conditions
Katasanos (1985). The Arrhenius plot for temperature dependence of
apparent kinetic constant is shown in Figure 95.
-6

ln k

-7
-8
-9
-10
-11
0.0025 0.0026 0.0026 0.0027 0.0027 0.0028 0.0028

1/T, K-1

Figure 95 Arrhenius plot for the photoreaction of cyclohexane on 8Mo2S

126

Photocatalytic fluidized bed reactor

The apparent activation energy and the pre-exponential factor were found
to be 26 kcalmol-1 and 25.65 Nm3g-1h-1 respectively. The value of preexponential factor is lower than that in the most of catalytic oxidative
dehydrogenation reactions whereas the apparent activation energy is very
similar to the activation energy required for the same reaction type Comite
et al. (2003).

VII.7 Photocatalytic oxidative dehydrogenation of ethylbenzene


to styrene
Styrene is one of the most important base chemicals in the petrochemical
industry Yoo (1996). Styrene is industrially manufactured by the
dehydrogenation process of ethylbenzene (EB) over iron-based catalysts.
Process is operated by adding excess steam to EB in an adiabatic reactor
under pressurized condition with the reaction temperatures of about 600 C.
The problems associated with the dehydrogenation process of ethylbenzene
are follows: thermodynamic limitation, low conversion rate, high
endothermic energy and deactivation of catalysts by coke formation Cavani
and Trifir (1995). Developments are going on to increase the
concentration level by removal of the reaction product hydrogen and thereby
shifting the thermodynamic equilibrium towards higher conversions. As an
alternative way, the oxidative dehydrogenation of ethylbenzene has been
proposed to be free from thermodynamic limitations regarding conversion,
operating at lower temperatures with an exothermic reaction Hanuza et al.
(1985). The systems most often studied for the oxidative process are based
on promoted inorganic oxide catalysts. Murakami and co-workers have
studied a SnO2-P2O5 catalyst and found it to have moderate activity greater
than 30 % conversion with greater than 80 % selectivity at 550C
Murakami et al. (1981). Their work focused particularly on the role that
promoters play in modifying the acid-base properties of the catalysts. Emig
and Hofmann have reported a ZrO2-P2O5 system which is slightly more
active in the same temperature range-conversion up to 55% with over 80%
selectivity Emig and Hoffmann (1983). Vrieland has similarly worked
with phosphate-promoted catalysts and has recently reported a cerium
pyrophosphate system which exhibits 76% conversion with 90% selectivity
at 605C Vrieland (1988). In addition to focusing on the acid-base
properties of the system, the latter two studies also investigated the role of
carbonaceous overlayers on the surface of the catalyst. It was determined
that each of the various systems studied had an induction period during
which a thin uniform layer of carbon built up on the surface. In another
study, Cadus and coworkers similarly found carbon overlayers on a sodiumdoped alumina catalyst Cadus et al. (1988). They provide a mechanism to
show how the carbonaceous species, not the organic oxide, brings about the
catalytic transformation. The influence of molybdenum, chromium and

127

Chapter VII

cobalt on the oxidation state and redox processes in V-Mg-O catalyst was
studied by means of X-ray diffraction and ESR spectroscopy. The results
obtained were correlated with catalytic activity and selectivity. It is shown
that molybdenum doping increased the selectivity and the chromium and
cobalt doping improved the activity of the catalyst studied Oganowski et al.
(1996). Selectivity to styrene as high as 97% was reported on Sn-Ge mixed
phosphate Turco et al. (1990). Mixed zirconium-tin phosphates gave much
higher ethylbenzene conversion with respect to pure Zr and Sn phosphates,
the selectivity to styrene being likewise high Bagnasco et al. (1991). This
behaviour is a consequence of the higher surface area of mixed phases. No
influence of chemical composition on catalytic activity was found. Surface
acidity of medium-high strength plays a relevant role in the reaction through
the formation of a catalytically active coke. However, the photocatalysed
oxidative dehydrogenation of ethylbenzene to styrene has never been
reported.

VII.7.1 Experimental results


Since an olefinic bond was obtained from a paraffinic bond by means a
photocatalytic oxidative dehydrogenation process (as it was shown in the
section VII.6.3), the possibility to form styrene from ethylbenzene was
investigated. Photocatalytic activity tests were carried out on 8Mo2S sample
by feeding 830 Ncc/min N2 stream containing 1000 ppm ethylbenzene, 1500
ppm O2 and 1600 ppm H2O. The reactor was illuminated by four UV light
sources (EYE MERCURY LAMP, 125W). The reaction temperature and
catalyst weight were 120C and 14g, respectively. The only reaction product
was styrene and no formation of CO2 was detected. Ethylbenzene conversion
and styrene outlet concentration are reported in Figure 96.

128

Photocatalytic fluidized bed reactor

120

ethylbenzene conversion

25

100

styrene outlet concentration

20

80

15

60

10

40

20

styrene, ppm

ethylbenzene conversion, %

30

0
0

20

40

60

illumination time, min

80

0
100

Figure 96 Ethylbenzene conversion and styrene outlet concentration as a


function of illumination time
The steady state value of ethylbenzene conversion was reached after
about 25 minutes and its value was about 11%. Styrene outlet concentration
was 110 ppm after 85 minutes of illumination and increased less quickly
with respect to ethylbenzene conversion. Total carbon mass balance was
closed to 100% and no catalyst deactivation was observed.

129

VIII

Conclusions

The occurrence of photocatalysed heterogeneous oxidative


dehydrogenation of cyclohexane in mild conditions has been studied in a
gas-solid annular photocatalytic fixed bed reactor in the presence of Mosupported catalysts, obtaining interesting selectivity to benzene and
cyclohexene.
Photocatalytic performances of molybdenum active phases loaded on
several supports, such as zeolites, and alumina and titania with different
level of surface sulphates, have been deeply investigated. With respect to
MoOx/AFer, MoOx/Al2O3, MoOx/Al2O3, MoOx/NaY and MoOx/HY,
highest photoactivity to dehydrogenated products was shown by MoOx/TiO2
catalysts.
The formation of polymolybdate species spread on anatase surface
increased by increasing molybdenum load up to that corresponding to the
monolayer formation. In this case, high selectivity to benzene was obtained,
whereas titania alone was 100% selective to carbon dioxide. After
monolayer formation, segregation of MoO3 cristallytes has been observed,
with a significant loss in photooxidative dehydrogenation activity.
Photocatalytic activity tests performed on anatase titania with different
sulphate content revealed that only deep oxidation of cyclohexane to CO2
occurred; steady state cyclohexane conversion decreasing with increasing
sulphate content. By contrast photocatalytic selective cyclohexane oxidation
to benzene and cyclohexene was observed on sulphated titania supported
MoOx. The presence of sulphate species on the surface of titania enhanced
benzene yield more with increases in the sulphate content up to 2 wt%. On
the basis of these last results, the selective formation of benzene is due to the
presence of polymolybdate species supported on the titania surface, but also
in the presence of surface sulphates.
The obtained results evidenced that benzene selectivity can be tuned by
catalyst formulation.
Photocatalytic activity tests performed on Mo-supported sulphated titania
in a photocatalytic flat-plate reactor confirmed the results obtained with the
annular photocatalytic fixed bed reactor.

Chapter VIII

In order to enhance the exposure of the catalysts to light irradiation, to


improve the contact between reactants and catalyst, and to permit easier
realisation of catalytic bed, a two dimensional fluidized bed photoreactor has
been designed and realised. Strong improvements in catalyst light exposure
and solid-gas mass transfer rates have been reached. As a result of these
improvements, a significant reduction in the total volume of reactor has been
achieved, passing from about 7 liters to about 0.1 liters with higher
cyclohexane conversion in similar operative conditions. Moreover it is worth
to note that by using fluidized bed photoreactor, catalyst deactivation
phenomena, probably due to the reaction products strogly adsorbed on active
sites, are absent.
In addition, a thermodynamic analysis of the oxidative dehydrogenation
reaction has been performed, and results indicated that the increase of
temperature up to 330 C gave an increase in cyclohexane conversion. The
effect of the light intensity and of the radiation wavelength was also
optmized.
Physico-chemical characterization tests performed on catalysts after
photocatalytic activity tests both in the fixed and in the fluidized bed reactors
suggested that catalyst deactivation may be correlated also with the sulphur
disappearance. In fact the samples used in fluidized conditions remain active
and maintain their sulphur content at the support surface.
A mechanism for the catalytic photooxidative dehydrogenation of
cyclohexane based on consecutive steps of oxy-dehydrogenation occurring
on active sites in competition with total oxidation on bare titania has been
proposed. This mechanism considers the oxidation of cyclohexane to
cyclohexene and its further oxy-dehydrogenation to benzene occurring on
molybdenum oxides active sites hypothesizing a detailed reaction network.
Deeper knowledge of reaction mechanism led to a formulation of
innovative photocatalysts (sulphated MoOx/-Al2O3) selective for oxidative
mono-dehydrogenation of hydrocarbons.
By changing the catalyst support and active species, it has been
demonstrated that total selectivity to mono- or further dehydrogenated
products can be obtained. The effectiveness has been verified in cyclohexane
conversion to cyclohexene and in ethylbenzene oxy-dehydrogenation to
styrene.

132

IX References

Acro M. D., S. R. G. Garrazan, V. Rives, and J. V.Garcia-Ramos


(1992) J. Mater. Sci., 27, 5921.
Afanasiev P., C. Geantet, M. Breysse, G. Coudurier, J. C. Vedrine
(1994) J. Chem. Soc. Faraday Trans, 90, 193.
Alberici R.M., W.F. Jardim (1997) Appl. Catal. B: Enviromental,
14, 55.
Alemany L., L. Lietti, N. Ferlazzo, P. Forzatti, G. Busca, E.
Giamello, F. Bregani (1995) J. Catal., 155, 117.
Almquist C. B., P. Biswas (2001) Appl. Catal.A, 214, 259.
Alyea E.C., M.A. Keane (1996), J. Catal., 164, 28.
Armengol E., A. Corma, V. Forns, H. Garca, J. Primo (1999)
Appl. Catal. A: General., 181, 305.
Armor J.N. (1992) Appl. Catal., B: Environ., 1, 221.
Augugliaro V., S. Cosuccia, V. Loddo, L. Marchese, G. Martra, L.
Palmisano, M. Schiavello (1999) Appl. Catal. B: Environ., 20, 15.
Bagnasco G., P. Ciambelli, M. Turco, A. La Ginestra, P. Patrono
(1991) Appl. Catal., 68, 69.
Bard A.J. (1980) Science, 207, 139.
Basmadjian D., G. N. Fulford, B. I. Parsons and D. S. Montgomery
(1962), J. Catal., 1, 547.
Bian G., Y. Fu, Y. Ma (1999) Catal. Today, 51, 187.
Bickley R. I., G. Munuera, F. S. Stone (1973) J. Catal., 31, 398.
Blake N.R. and G.L. Griffin (1988) J. Phis. Chem., 92, 5697.
Blatter F., H. Frei (1996) J. Am. Chem. Soc., 118, 6873.
Blatter F., H. Sun, S. Vasenkov, H. Frei (1998) Catal. Today, 41,
297.
Blomberg J., P.J. Schoenmakers and U.A. Th. Brinkman (2002) J.
Chromatogr. A, 972, 137.
Boisdron N., A. Monnier, L. Jalowlecki-Duhamel, Y. Barbaux
(1995) J. Chem. Soc. Faraday Trans., 91, 2899.

Chapter IX

134

Brucato A., L. Rizzuti (1997) Ind. Eng. Chem. Res., 36, 4748.
Cadus L. E., L. A. Arrua, O. F. Gorriz, J. B. Rivarola (1988) Ind.
Eng. Chem. Res., 27, 2241.
Carvalho W.A., M. Wallau, U. Schuchardt (1999) J. Mol. Catal.,
144, 91.
Cassano A. E., O. M. Alfano (2000) Cat. Tod., 58, 167.
Cassano A.E., C.A Martin, R.J Brandi, O.M Alfano (1995) Ind. Eng.
Chem. Res., 34, 2155.
Cavani F., F. Trifido (1995) Appl. Catal. A: General, 133, 219.
Cavani F., F. Trifido (1995) Catal. Today, 24, 307.
Chen L. C., F.R. Tsai, C. M. Huang (2005) J. Photochem. Photobiol.
A: Chem, 170, 7.
Cheng C. P., G. L. Schrader (1979) J. Catal., 60, 276.
Ciambelli P., G. Lisanti, D. Sannino, V. Palma, The TOCAT 4 PreConference Symposium in Kobe Catalysis for the Environment
and New Energy Sources July 12th, 2002 Kobe, Japan
Coln G., M.C. Hidalgo, J.A. Navo (2003) Appl Catal B: Environ.,
45, 39.
Comite A., A. Sorrentino, G. Capannelli, M. Di Serio, R. Tesser, E.
Santacesaria (2003) J. Mol. Catal., 198, 151.
Conway S.J., D.J. Wang, J.H. Lunsford (1991) Appl. Catal. A:
General, 79, L1L5.
Corbo P., M. Gambino, D. Sannino, P. Ciambelli (1994), Mater.
Eng., 5, 237.
Corma A., M.I. Vazquez, A. Bianconi, A. Clozza, J. Garcia, O.
Pollota, J.K. Cruz (1988) Zeolites, 8, 464.
Coronado J. M., S. Kataoka, I. Tejedor-Tejedor, M. A. Anderson
(2003) J. Catal., 219, 219.
Cortese A.D. (1990) Environ. Sci. Technol., 24, 442.
Cui K., K. Dwight, S. Soled, A. Wold (1995) Solid State Chem.,
115, 187.
Curcio F., M. Musci, M. Notaro, G. Quattrini (1991) Mat. Sci. Mon.,
66, 2569.
Davis D.D., D.R. Kemp (1991) in: J.I. Kroschwitz, M. HoweGrant(Eds.), KirkOthmer Encyclopedia of Chemical Technology,
Vol. 1, Wiley, New York, p. 46
Diaz A. L., M. E.Bussell (1993) J. Phys. Chem., 97, 470.
Dibble L.A., G.B. Raupp (1990) Catal. Lett., 4, 345.
Dibble L.A., G.B. Raupp (1992) Environ. Sci. Technol., 26, 492.
Djeghri N., S. J. Teichner (1980) J. Catal., 62, 99.
Duncan D.C., C.L. Hill (1997) J. Am. Chem. Soc., 119, 243.
Dury F., E.M. Gaigneaux, P. Ruiz (2003) Appl. Catal. A: General,
242, 187.

References

Egerton A., C.J. King (1979) J. Oil Colour Chem. Assoc., 62, 386.
Einaga H., S. Futamura, T. Ibusuki (2002) Appl. Catal. B:
Enviromental, 38, 215.
Ekstrom G.N., A.J. McQuillan (1999) J. Phys. Chem. B, 103, 10562.
Emig G., H. Hoffmann (1983) J. Catal, 84, 15.
Ermolenko L.P., C. Giannotti (1996) J. Chem. Soc. Perkin Trans., 2,
1205.
Fisherman J. (1991) Chemosphere, 22, 685.
Formenti M., J. P. Meriaudeau, S. J. Teichner (1971) Chemtech., 1,
680.
Fox M.A., M.T. Dulay (1993) Chem. Rev., 93, 341.
Fu X., L.A. Clark, W.A. Zeltner and M.A. Anderson (1996) J.
Photochem. Photobiol. A., 97, 181.
Fu X., W.A. Zeltner and M.A. Anderson (1995) Appl. Catal. B:
Environ., 6, 209.
Fu X., Z. Ding, W. Su, Chin (1999) J. Catal., 20, 321.
Fuchs St., L. Leveles, K. Seshan, L. Lefferts, A.A. Lemonidou, J.A.
Lercher (2001) Top. Catal. , 15, 169.
Geldart D. (1973) Powder Technol., 7, 285.
Giannotti C., C. Richter (1999) Int. J. Photoenergy, 1, 1.
Giordano N., J. C. J Bart, A. Vaghi, A. Castellan, G. Martinotti
(1975) J. Catal., 36, 81.
Gratzel M. (1983), in: Energy Resources through Photochemistry
and Catalysis, Academic Press, New York,.
Guillard C., P. Thron, P. Pichat, C. Ptrier (2002) Water Reasarch,
36, 4263.
Hanuza J., B. Jesovska-Trzebiatowska, W. Oganovski (1985) J. Mol.
Catal., 29, 109.
Herskowitz M., M. Landau, M. Kaliya (1996), PCT Int. Appl. WO.
9622161. A1. 960725.
Hirata T. (1989) Appl. Surf. Sci., 40, 179.
Hoffmann M.R., S.T. Martin, W. Choi, D. Bahnemann (1995)
Chem. Rev., 95, 69.
Inamura S., H.Sasaki, M. Shono, H. Kani (1998) J. Catal. 177, 72.
Japer S.M., T.J. Wallington, S.J. Rudy, T.Y. Chang (1991) Environ.
Sci. Technol., 25, 415.
Jeziorowski H., H. Knozinger (1979) J. Phys. Chem., 83, 1166.
Jiang D., H. Zhao, S. Zhang, R. Jhon (2004) J. Catal., 223, 212.
Karakitsou K.E., X. E. Verykios (1993) J. Phys. Chem., 97, 1184.
Katasanos N. A. (1985) J. Catal, 94, 376.
Kim D.S., Y. Kurusu, I.E.Wachs, F.D. Hardcastle, K. Segawa
(1989) J. Catal., 120, 325.
Kim M., W. Nam. , G. Y Han (2004). J. Chem. Eng. , 21, 721.

135

Chapter IX

136

Klimova T., E. Carmona, J. Ramirez (1998) J. Mat. Sci. , 33, 1981.


Kormann C., D.W. Bahnemann, M.R. Hoffmann (1991) Environ.
Sci. Technol., 25, 494.
Kosuko M., C.M. Nunez (1990) J. Air Wastes Manage. Assoc., 40,
254.
Kozlov D., D. Bavykin, E. Savinov (2003) Catal Letters, 86, 169.
Kung H.H. (1994) Adv. Catal., 40, 1.
Lapple C. E. (1973) in: Air Pollution Enginering Manual, USEPA,
AP-40, 2nd Ed.
Lee W. I., G. J. Choi, Y. R. Do (1997), Bull. Kor. Chem. Soc., 18,
667.
Legrini O., E. Oliveros, A.M. Braun (1993) Chem.Rev. , 2, 671.
Li Puma G., J.N Khor, A. Brucato (2004) Environ. Sci. Technol., 38,
3737.
Li Puma G., Proc. of the 7th World Congress of Chemical
Engineering, 2005, Glasgow.
Lietti L., P. Forzatti, L.J. Alemany, G. Busca, E. Giamello, F.
Bregani (1996) Catal. Today, 29, 143.
Lim T.H., S.M. Jeong, S.D. Kim, J. Gyenis (2000) J. Photochem.
Photobiol. A: Chemistry, 134, 209.
Linsebigler A.L., G. Lu, J.T. Yates Jr. (1995) Chem. Rev., 95, 735.
Litter M. I. (1999) Appl. Catal., B: Environ., 23, 89.
Liu H., S. Cheng, J. Zhang, C. Cao, W. Jiang (1997) Chemosphere,
35, 2881.
Loddo V., G. Marci, C. Martin, L. Palmisano, V. Rives, A. Sclafani
(1999) Appl. Catal. B: Environmental, 20, 29.
Lynette A.D., B.R. Gregory (1992) Environ. Sci. Technol., 26, 492.
Lyons J. E., G. W. Paashall (1994) Catal. Today, 22, 313.
Maity S. K., M. S. Rana, S. K. Bej (2001) Appl. Catal. A: General,
205, 215.
Maldotti A., A. Molinari, P. Bergamini, R. Amadelli, P. Battioni, D.
Mansuy (1996) J. Mol. Catal., 113, 147.
Maldotti A., A. Molinari, R. Amadelli (2002) Chem. Rev., 102,
3811.
Maldotti A., R. Amadelli, I Vitali, L. Borgatti, A. Molinari (2003) J.
Mol. Catal., 703, 204.
Marc G., M. Addamo, V. Augugliaro, S. Cosuccia, E. Garca
Lopez, V. Loddo, G. Martra, L. Palmisano, M. Schiavello (2003) J.
Photochem. Photobiol. A: Chemistry, 160, 105.
Martin S.T., H. Herrmann, M.R. Hoffmann (1994) J. Chem. Soc.,
Faraday Trans., 90, 3323.
Martra G., S. Coluccia, L. Marchese, V. Augugliaro, V. Loddo, L.
Palmisano, M. Schiavello (1999) Cat. Tod. ,53, 695.

References

Mascia L. (1995) Trends Polym. Sci., 3, 61.


Matralis H., S. Theret, Ph. Sebastians, M. Ruwet, P. Grange (1995)
Appl. Catal. B: Enviromental, 5, 271.
Matsuda S., H. Hatano, A. Tsutsumi (2001) Chem. Eng Journal, 82,
183.
Matthews R.W., M. Abdullah, G.K.-C. Low (1990) Anal. Chim.
Acta, 233, 171.
Medema J., C. Van Stam, V. H. J. De Beer, A. J. A. Konings, and D.
C. Koningsberger (1978) J. Catal., 53, 386.
Mizuno N., M. Misono (1998) Chem. Rev., 98, 199.
Molinari A., A. Maldotti, R. Amadelli, A. Sgobino, V. Parassiti
(1998) Inorg. Chim. Acta, 272, 197.
Morishige K., F. Kamo, S. Ogawara, and S. Sasaki (1985) J. Phys.
Chem., 89, 4404
Mu W., J.M. Herrmann, P. Pichat (1989) Catal. Lett., 3, 73.
Muggli D. S., J. L. Falconer (1998) J. Catal., 175, 213
Muggli D.S. and L. Ding (2001) Appl. Catal. B: Environ., 32, 181.
Murakami Y., K. Iwayama, H. Uchida, T. Hattori, T. Tagawa (1981)
J. Catal, 71, 257.
Ng K.Y.S., E. Gulari (1985) J. Catal., 92, 340.
Nishi K., S. Komai, K. Inagaki, A. Satsuma and T. Hattori (2002)
Appl. Catal. A: Gen., 223, 187.
Noh J., J. Schwarz (1989) J. Colloid Interface Sci., 130,157.
Obee T.N., R.T. Brown (1995) Environ. Sci. Technol., 29, 1223.
Oganowski W., J.Hanuza, H. Drulis,W. Mista, L. Macalik (1996)
Appl. Catal. A: General, 136, 143.
Ohtani B., Y. Ueda, S. Nishimoto, T. Kagiya and H. Hachisuka,
(1990) J. Chem Soc. Perkin Trans., 11, 1972
Okamoto K., Y. Yamamoto, H. Tanaka, M. Tanaka, A. Itaya (1985)
Bull. Chem. Soc. Jpn., 58, 2015.
Okasaki S., T. Okuyama (1983) Bull. Kor. Chem. Soc. Jpn., 56,
2159.
Ollis D.F.(1991) in: E. Pelizetti, E. Schiavello (Eds.), Photochemical
Conversion and Storage of Solar Energy, Kluwer Academic
Publishers, Dordrecht, p. 593.
Oshaka T., F. Izumi, and Y. Fujiki (1978) J. Raman Spectrosc., 7,
321.
P. Ciambelli, M.E. Fortuna, D. Sannino, A. Baldacci (1996) Catal.
Today, 29, 161.
Palmisano L. and A. Sclafani (1997) in: M. Schiavello,
Heterogeneous Photocatalysis, Wiley & Sons: London, p 109.
Panizza M., C. Resini, G. Busca, E. F. Lpez, and V. S. Escribano
(2003), Catal. Lett., 89, L-4.

137

Chapter IX

138

Papp J., S. Soled, K. Dwight, A. Wold (1994) Chem. Mater. 6, 496.


Park D., J. Zhang, K. Ikeue, H. Yamashita, & M. Anpo (1999) J.
Catal., 185, 114.
Pasquali M., F. Santarelli, J. F. Porter, P.L. Yue (1996) AIChE J.,
42, 532.
Peral J. and D.F. Ollis (1992) J. Catal., 136, 554.
Philips L.A. and G.B. Raupp J. Mol. Cat. (1992), 77, 297.
Pillai U. R., E. Sahle-Demessie (2002) J. Catal., 211, 434.
Ping C., N. Qian, Z. Ya-li (2004) CODEN: HYZHEN ISSN: 10010017. Journal written in Chinese, 78, 61
Pirone R., P. Ciambelli, G. Moretti and G. Russo (1996) Appl. Catal
B: Environmental, 8, 197.
Quincy R.B., M. Houalla, A. Proctor, D.M. Hercules (1989), J.
Phys. Chem., 93, 5882.
R. I. Bickley (1997) in: M. Schiavello, Heterogeneous
Photocatalysis, Wiley & Sons: London, p 87.
Ramirez J., L. Ruiz-Ramirez, L. Cedeno, V. Harle, M. Vrinat and M.
Breysse (1993) Appl. Catal., 93, 163.
Raupp G. B., J. A. Nico, S. Annangi, R. Changrani, R. Annapragada
(1997) AIChE J., 43, 792.
Rendell D (1987) in: Fluorescence and Phosphorescence, Analytical
Chemistry by Open Learning, Wiley, New York,.
Reymond J.P., F. Molenda (1999) Powder Technol., 103, 30.
Roby A., J. P. Kingsley (1996), Chemtech, 26, 39.
Samantaray S. K., K. M. Parida (2001) Appl. Catal. A: General, 211,
175.
Sauer M.L., and D.F. Ollis (1996) J. Catal., 158, 570.
Sauer M.L., D.F. Ollis (1996) J. Catal., 163, 215.
Seinfeld J.H., S.N. Pandis (1997) in: Atmospheric Chemistry and
Physics: from Air Pollution to Climate Change, Wiley/Interscience,
New York,.
Serpone N. (1997) J. Photochem. Photobiol. A: Chem., 104, 11.
Shah J. J, H.B. Singh (1988) Environ. Sci. Technol., 22, 1381.
Snel R. (1984) Appl. Catal., 12, 347.
Sorrentino A., S. Rega, D. Sannino, A. Magliano, P. Ciambelli, E.
Santacesaria (2001) Appl. Catal. A: General, 209, 45.
Suppan P. (1994) Chemistry and Light, Royal Society of Chemistry,
Cambridge.
Tanielian C., K. Duffy, A. Jones (1997) J. Phys. Chem. B, 101,
4276.
Trimm D. L. and A. Stanislaus (1986) Appl. Catal., 21, 215.
Turchi C, M. Mehos and J. Pacheco (1993), Design issues for solardriven photocatalytic systems, in Photocatalytic Purification and

References

Treatment of Water and Air, Ed by Ollis DF and Al-Ekabi H,


Elsevier, Amsterdam, The Netherlands, pp 789794
Turco M., G. Bagnasco, P. Ciambelli, A. La Ginestra and G. Russo
(1990) Stud. Surf. Sci. Catal. , 55, 327.
Vorontsov A. V., E. N. Savinov, P. G. Smirniotis (2000) Chem. Eng.
Sci., 55, 5089.
Vrieland G. E (1988) J. Catal, 111, 1.
Wachs I. E. (1996) Catal. Today, 27, 437.
Walker A., M. Formenti, J. P. Meriaudeau, S. J. Teichner (1977) J.
Catal., 50, 237.
Wang L. and W. K. Hall (1980) J. Catal., 66, 251.
Ward M.D., J.M. White, A.J. Bard (1983) J. Am. Chem. Soc., 105,
27.
Watling T.C., G. Deo, K. Seshan, I.E. Wachs, J.A. Lercher (1996)
Catal. Today, 28, 139.
Wei B., Q. Xin, X.X. Guo, P. Grange and B. Delmon (1990), Appl.
Catal., 63, 305.
Wieinold J., R. E. Jentoft, T. Ressler (2003) Eur. J. Inorg. Chem., 6,
1058.
Wilkinson C. F. (1987) Environ. Sci. Technol., 21, 843.
Wolf K., A. Yazdani, P. Yates (1991) J. Air Wastes Manage. Assoc.,
41, 1055.
Xie C., Z. Xu, Q. Yiang, N. Li, D. Zhao, D. Wang, Y. Du (2004) J.
Mol. Catal., 217, 193.
Xu Y., Y. Shu, S. Liu, J. Huang, X. Guo (1995) Catal. Lett., 35, 233.
Yamashita H., Y. Ichihashi, M. Takeuci, S. Kishiguchi, M. Ampo
(1999) J. Synchrotron Rad., 6, 451.
Yoo J.S. (1996) Appl. Catal. A: General, 142, 19.
Yu J.C., J. Lin, R.W.M. Kwork (1997) J. Photochem. Photobiol. A:
Chem., 111, 199.
Yu J.C., J. Yu and J. Zhao (2002) Appl. Catal. B: Environ., 36, 31.
Yue P. L., F. Khan, L. Rizzuti (1983) Chem. Eng. Sci., 38, 1893.
alac S., N. Kallay (1992) J. Colloid Interface Sci., 149, 233.
Zhang X., M. Zhou, L. Lei (2005) Appl Catal., A: General., 282,
285.

139

Potrebbero piacerti anche