You are on page 1of 199

WEI SHEN

ALKYLATION OF ISOBUTANE/1-BUTENE OVER


ACID FUNCTIONALIZED MESOPOROUS
MATERIALS

Thse prsente
la Facult des tudes Suprieures de l'Universit Laval
dans le cadre du programme de doctorat en gnie chimique
pour l'obtention du grade de Philosophiae Doctor (Ph.D)

DEPARTEMENT DE GENIE CHIMIQUE


FACULT DES SCIENCES ET DE GNIE
UNIVERSIT LAVAL
QUBEC

2010

Wei Shen, 2010

Rsum
L'alkylat est un additif de choix de l'essence reformule puisqu'il possde une
densit nergtique et un indice d'octane levs et que de plus sa combustion est propre
et produit moins d'missions. Les catalyseurs liquides commerciaux utiliss pour
l'alkylation isobutane/N-butne crent des dangers significatifs pour l'environnement. De
nombreux acides solides ont t essays pour cette raction. Aucun d'entre eux n'a connu
un succs commercial en raison de la dsactivation rapide du catalyseur. Cette
dsactivation est due 1'adsorption prfrentielle et au remplissage des pores par
l'olfine. C'est la raison pour laquelle l'application d'acides msoporeux solides est
d'intrt.
Dans cette thse, une srie de matriaux acides fonctionnaliss et msostructurs
ont t synthtiss. La force acide a pu tre ajuste par l'immobilisation des groupes
acides propyl-sulfoniques, arne-sulfoniques et perfluoroalkylsulfoniques. Le caractre
hydrophile-hydrophobe de la surface a t ajust par l'utilisation d'une organosilice ou
par le recouvrement des OHs de la surface de la silice. Des matriaux ayant des rseaux
de canaux une dimension et trois dimensions ont t synthtiss et compars. Ces
matriaux sont caractriss gnralement par l'analyse lmentaire, la volumtrie d'azote,
le dosage acido-basique, la RMN MAS du 29Si, FT-IR, TGA, SEM et TEM . Ce sont des
matriaux msostructurs typiques possdant des groupes acides immobiliss.
L'alkylation de l'isobutane par le 1-butne a t effectue sur les matriaux
synthtiss. La force acide des catalyseurs est le facteur dterminant de l'activit
d'alkylation. Les matriaux modifis avec les groupes acides perfluoroalkylsulfoniques
dmontrent

la

meilleure

activit

catalytique.

Les

organosilices

msoporeuses

fonctionnalises acides de la silice silanise, qui possdent une surface hydrophobe, ont
mieux perform que les catalyseurs ayant une surface hydrophile. Nous croyons que cette
performance suprieure est le rsultat d'une concentration leve de la paraffine dans les
pores. Les solides msoporeux acides rseaux tridimensionnels surpassent ceux une
dimension pour la conversion des butnes et

la slectivit en TMPs. La structure

msoporeuse trois dimensions est capable de diffuser facilement les molcules et de


ii

rsister la dsactivation par le colmatage des pores. Il est suggr que les sites acides
forts comme les groupes acides perfluoroalkylsulfoniques, la structure msoporeuse
trois dimensions et la surface hydrophobe doivent tre recherchs pour concevoir un
catalyseur d'alkylation.

Abstract
Alkylate is the preferred blending stock of reformulated gasoline as it is clean
burning, producing less emissions, and possesses high energy and octane. Commercial
liquid catalysts used for isobutane/w-butene alkylation pose significant environment
hazards. Many solid acids have been investigated towards this reaction. None of them
have met commercial success because of a rapid catalyst deactivation. This deactivation
is due to the preferential adsorption and pore filling with the olefin. For this reason, the
application of mesoporous solid acids is favorable.
In this thesis, a series of acid functionalized mesostructured materials was
synthesized. The acid strength was tuned by immobilizing propyl-sulfonic, arene-sulfonic
or perfluoroalkylsulfonic acid groups. The hydrophilic-hydrophobic nature of the surface
was adjusted by using organosilica or by capping the surface OHs of silica. Materials
with one-dimensional and three-dimensional channels were synthesized and compared.
These materials were generally characterized using elemental analysis, nitrogen
adsorption, acid-base titration, XRD, 29Si MAS NMR, FT-IR, TGA, SEM and TEM.
They are typical mesostructured materials with immobilized acid groups.
Alkylation of isobutane with 1-butne was carried out over the synthesized
materials. The acid strength of the catalyst is the factor determining the alkylation activity.
Materials modified with perfluoroalkylsulfonic acid groups exhibited the best catalytic
activity. Acid functionalized mesoporous organosilica and silanized silica which have
hydrophobic surface were shown to outperform catalysts with hydrophilic surface. This
better catalytic performance is believed to be a result of higher paraffin concentration in
the pore system. The three-dimensional mesopores solid acids outperform onedimensional ones both in butnes conversion and TMPs selectivity. Three-dimensional
mesopores structure is capable of diffusing molecules easily and resisting deactivation by
pore plugging. It is suggested that strong acid sites like perfluoroalkylsulfonic acid
groups, three-dimension mesopores structure and hydrophobic surface should be desired
for an alkylation catalyst.

IV

Acknowledgements
First of all, I would like to express my sincere gratitude to my thesis director,
Professor Serge Kaliaguine, for his continuous guidance, support and patience throughout
my Ph.D studies as well as for the great opportunity which he offered me to work in his
lab during the last four years.
I am especially grateful to Madam Guoying Xu for the kindness and help from the
first days I came to Laval University to start my Ph.D study.

I would also like to thank to Professor F reddy Kleitz and Professor Do Trong-On
who gave me advice, support and discussions.
Many thanks go to Mr. Gilles Lemay, Dr. H. Zahedi-Niaki, Dr. Nguyen Tien
Thao, Dr. David Dube and Dr. Michat Rat for their generous help in performing the
thesis work. I want to thank all colleagues in the laboratory of catalysis.
Special gratitude must be given to Professor Hualong Xu and Dongyuan Zhao in
the Chemistry Department of Fudan University for their continuous support.
I would like to express my gratitude to all my Chinese friends in Laval University
for their friendship, support and encouragement.
Finally, I would like to express my gratitude to my family for their support.

Table of Contents
Rsum

ii

Abstract

iv

Acknowledgements

Table of Contents

vi

Index of Tables

xi

Index of Figures

xiii

1. Introduction

2. Mechanism, Catalysts, and Process of Isobutane/Alkene Alkylation

2.1. Introduction
2.2. Isobutane/butene alkylation reaction mechanism on solid acids
2.2.1. Product distribution

5
11
11

2.2.2. Proposed isobutane/butene alkylation reaction mechanism on solid acids .... 14


2.2.2.1. Protonation of the alkene and formation of tert-C-f ion

14

2.2.2.2. Addition of tert-C^ ion to second olefin and isomerization

16

2.2.2.3. Hydride transfer to C% ion

17

2.2.2.4. Oligomerization, disproportionation and cracking

20

2.2.2.5. Self-alkylation

23

2.2.2.6. Deactivation mechanism of solid acid catalysts

24

2.2.2.7. Catalyst requirements

27

2.3. Solid acid alkylation catalysts

29

2.3.1. Zeolites

29

2.3.2. Mounted liquid acids

34

2.3.3. Heteropolyacids

35

2.3.4. Acidic organic polymers

36

2.3.5. Sulfated metal oxides

37
vi

2.4. Effect of process parameters on alkylation

38

2.4.1. Reaction temperature

38

2.4.2. Paraffin/olefin ratio and olefin space velocity

39

2.5. New solid acid catalysts-acid functionalized mesostructured materials

40

2.5.1. Post-synthesis preparation of organosulfonic-functionalized mesostructured


materials

42

2.5.2. Direct-synthesis of organosulfonic-functionalized mesostructured materials 43


2.5.3. Tuning the hydrophobicity of pore surface

45

2.5.4. Tuning the acid strength of anchored sulfonic-acid sites

48

2.5.5. Lewis acids

52

2.6. Further developments


3. Experimental

53
55

3.1. Introduction

55

3.2. Synthesis

55

3.2.1. Preparation of SBA-15 and SBA-15 type PMO

55

3.2.2. Preparation of arene-sulfonic acid functionalized SBA-15

56

3.2.3. Preparation of arene-sulfonic and propyl-sulfonic acid functionalized PMO 57


3.2.4. Preparation of sulfonated polystyrene coated SBA-15

57

3.2.5. Preparation of aluminum chloride grafted SBA-15 and PMO

58

3.2.6. Preparation of perfluoroalkylsulfonic acid functionalized SBA-15 and PM058


3.2.7. Preparation of Nafion functionalized SBA-15 materials (Nafion(X) /SBA-15
andNafion(X)/Me-SBA-15)

59

3.2.8. Preparation of Nafion functionalized SBA-16 materials (Nafion(X) /SBA-16


andNafion(X)/Me-SBA-16)
3.3. Characterization

60
61

3.3.1. Nitrogen adsorption isotherms

61

3.3.2. X-ray diffraction (XRD)

63

3.3.3. Fourier transform infrared spectroscopy (FTIR)

64

3.3.4. Magic angle spinning nuclear magnetic resonance spectroscopy (MAS NMR)
66
3.3.5. Transmission electron microscopy (TEM)

67
vii

3.3.6. Scanning electron microscopy (SEM)

67

3.3.7. Thermogravimetric analysis (TGA)

68

3.3.8. Flame atomic absorption spectroscopy (AAS)

68

3.3.9. Elemental analysis of carbon and sulfur

69

3.3.10. Ion exchange capacity

69

3.3.11. Intelligent gravimetric analyzer (IGA)

69

3.4. Catalytic reaction

70

3.4.1. Reaction system and procedure

70

3.4.2. Calculation

72

4. Preliminary Studies of Isobutane/1-butene Alkylation over Mesostructured Solid


Acids

74

Abstract

74

4.1. Introduction

74

4.2. Experimental

75

4.2.1. Catalyst preparation

75

4.2.2. Catalyst characterization

78

4.2.3. Catalytic performance

79

4.3. Results and discussion


4.3.1. Catalyst characterization

79
79

4.3.1.1 Arene-sulfonic and propyl-sulfonic acid functionalized materials

79

4.3.1.2. Aluminum chloride grafted SBA-15 and PMO

82

4.3.1.3. Sulfonated polystyrene coated SBA-15 (SPS-SBA-15)

84

4.3.1.4. Perfluoroalkylsulfonic acid functionalized SBA-15 and PMO

85

4.3.2. Catalytic performance


4.4. Conclusion

89
94

5. Alkylation of Isobutane/1 -butne over Periodic Mesoporous Organosilica


Functionalized with Perfluoroalkylsulfonic Acid Group

96

Abstract

96

5.1. Introduction

96

5.2. Experimental

98
viii

5.2.1. Catalyst preparation

98

5.2.2. Catalyst characterization

99

5.2.3. Catalyst tests

99

5.3. Results and discussion

100

5.3.1. Catalyst characterization

100

5.3.2. Catalytic performance

101

5.4. Conclusion

103

Reference

103

6. Alkylation of Isobutane/1-butne on Methyl-modified Nafion/SBA-15 Materials


112
Abstract

112

6.1. Introduction

113

6.2. Experimental

115

6.2.1. Catalyst preparation

115

6.2.2. Catalyst characterization

116

6.2.3. Reaction Procedure

117

6.3. Results and discussion

118

6.3.1. Catalyst characterization

118

6.3.2. Catalytic performance

121

6.4.3. Conclusions

125

References

126

7. Alkylation of Isobutane/1-butne on Methyl-modified Nafion /SBA-16 Materials


140
Abstract

140

7.1. Introduction

141

7.2. Experimental

143

7.2.1. Catalyst preparation

143

7.2.2. Catalyst characterization

144

7.2.3. Reaction procedure

145

7.3. Results and discussion

146
ix

7.3.1. Catalyst characterization

146

7.3.2. Catalytic performance

149

7.3.3. Conclusions

153

References
8. Conclusions and Recommendations

154
167

8.1 Summary

167

8.2. Recommendations

170

9. Scientific Contributions

171

9.1 Publications

171

9.2 Oral presentations

171

9.3 Posters

172

10. References

173

Index of Tables
Table 1. Octane number contributions from components of a typical US gasoline. (Miller,
1999)

Table 2. RON and composition of the various alkylates produced with different acidic
catalysts

11

Table 3. Comparison of TMP and DMH selectivity on liquid and solid acid catalysts,
with thermal equilibrium composition

13

Table 4. Nomenclature and chemical shift of Q* and T*

66

Table 5. Physico-chemical properties of ArS-PMO, ArS-SBA-15 and PrS-PMO

80

Table 6. Textural and compositional properties of Al-PMO and Al-SBA-15

83

Table 7. Physico-chemical properties of SBA-15, PSSBA-15 and SPS-SBA-15

85

Table 8. Physico-chemical properties of FS-PMO and FS-SBA-15

86

Table 9. Catalysts used in alkylation reaction and their textural properities

90

Table 10. Initial (TOS = 1.0 min) butnes conversion and product distribution obtained
on different catalysts

91

Table 11. Physical properties of the catalysts in the alkylation of isobutane/1-butne.. 106
Table 12. Textural properties for Nafion-modified mesostructured materials

129

Table 13. Methyl incorporation and acidic-related properties of Nafion modified


mesostructured materials
Table 14. Textural and catalytic properities of catalysts

130
131

Table 15. Influence of I/O ratio on the initial (TOS = 1 min) butnes conversion and
products distribution on Nafion(30)/SBA-15 and Nafion(30)/Me-SBA-15a

132

Table 16. Textural properties for Nafion-modified mesostructured SBA-16 materials. 157
Table 17. Methyl incorporation and acidic-related properties of Nafion modified
mesostructured materials
Table 18. Textural and catalytic properities of catalysts

158
159
xi

Table 19. Variation of the butnes conversion and products distribution at different
reaction temperature and time on stream over Nafion(30)/Me-SBA-15 and
Nafion(30)/Me-SBA-16 catalysts

160

xu

Index of Figures
Figure 1. Isomerization of 2,2,3-TMP+ via methyl and hydride shifts

17

Figure 2. Pathway to oligomerization products with the corresponding rate constants... 20


Figure 3. Different modes of/i-scission

22

Figure 4. Post-synthesis strategy for the preparation of sulfonic-acid-modified


mesostructured materials

43

Figure 5. The effect of spectator groups on activity in a Friedel-Crafts acylation

47

Figure 6. Route to synthesis mesostructured polymer-silica composite materials

48

Figure 7. Perfluoroalkylsulfonic-modified mesostructured materials synthesized by the


grafting technique
50
Figure 8. Preparation of perfluoroalkylsulfonic-modified mesostructured materials by
direct synthesis

51

Figure 9. Synthesis of bismuth sulfonate immobilized on silica

53

Figure 10. Preparation of perfluorosulfonic acid functionalized SBA-15 or PMO

59

Figure 11. Types of sorption isotherms (Sing et al., 1985)

61

Figure 12. Vacuum system used for the adsorption of pyridine

64

Figure 13. Infrared cell for the adsorption of pyridine

65

Figure 14. Flow schematic of the experimental reactor system

71

Figure 15. Preparation of perfluorosulfonic acid functionalized PMO

77

Figure 16. N2 adsorption isotherms and BJH pore diameter distribution for ArS-PMO.. 80
Figure 17. X-ray diffraction pattern of samples

81

Figure 18. Thermogravimetric weight loss curve for ArS-PMO

81

Figure 19. Thermogravimetric weight loss curve for ArS-SBA-15

82

Figure 20. N2 adsorption isotherms and BJH pore diameter distribution for Al-PMO.... 82
Figure 21. Pyridine FTIR spectra (after desorption at 120 C) of Al-PMO and Al-SBA-15
83
xiii

Figure 22.27A1 MAS NMR spectra of Al-PMO

84

Figure 23. Structures of A1C13 grafted on PMO with the 27A1 MAS NMR chemical shifts.
84
Figure 24. XRD pattern and TEM images for FS-PMO

85

Figure 25. N2 adsorption isotherms and BJH pore diameter distribution for FS-PMO.... 86
Figure 26. 29 Si MAS NMR spectrum of the FS-PMO

87

Figure 27. FT-IR spectra of FS-PMO sample heated at different temperatures

88

Figure 28. FTIR spectra of FS-PMO sample evacuated at 100 C before pyridine
adsorption (a), after pyridine adsorption and evacuation at 35 C (b), 50C (c)
and 75 C (d)

88

Figure 29. Butnes conversion with TOS in the alkylation of isobutane/1-butne.


(Reaction condition: T = 343K; olefin WHSV = 2 h" ; isobutane/1-butne
molar ratio = 40.)

93

Figure 30. Butnes conversion and distribution of C5+ with time on stream on FS-PMO
(Reaction condition: T = 343 K; olefin WHSV = 2 h"1; isobutane/1-butne
molar ratio = 40.)

94

Figure 31. Preparation of perfluoroalkylsulfonic functionalized PMO by the grafting


technique

107

Figure 32. XRD pattern and TEM images for FS-PMO

108

Figure 33. 29 Si MAS NMR spectrum of the FS-PMO

109

Figure 34. Butnes conversion with time on stream in the alkylation of isobutane/1butene
Figure 35. Distribution of C5+ products with time on stream on FS-PMO catalyst

110
111

Figure 36. (a) Nitrogen adsorption-desorption isotherms at 77 K and (b) X-ray


diffraction

133

Figure 37. Pore size distributions of the Nafion-modified SBA-15 materials

134

Figure 38. 29 Si MAS NMR spectrum of Nafion(30)/Me-SBA-15

135
xiv

Figure 39. (a) SEM image and TEM images of Nafion(30)/Me-SBA-15 viewed (b) in the
[100] direction and (c) in the [110] direction

136

Figure 40. Catalytic performance of Nafion-based catalysts. Reaction conditions: T =


100 C; olefin WHSV = 2 h"1; isobutane/1-butne (I/O) molar ration = 40... 137
Figure 41. Variation of the conversion of butnes (a) and yield of TMPs (b) with time on
stream over Nafion(30)/Me-SBA-15 under varying experimental conditions.
100 C, WHSV = 1.4 h"\ I/O = 40 ( ) ; 100 C, WHSV = 2h"\ I/O = 40 (A);
70C, WHSV = 2 h"1, I/O = 40 (O); 100 C, WHSV = 2 h"1, I/O = 20 (A). 138
Figure 42. Initial catalytic activity of fresh and regenerated Nafion-containing catalysts.
Reaction conditions: T = 100 C; olefin WHSV = 2 h"1; I/O = 40

139

Figure 43. X-ray diffraction patterns of SBA-16, Me-SBA-16 and their Nafion-modified
samples

161

Figure 44. (a) N2 adsorption/desorption isotherms and (b) pore size distributions
calculated from the adsorption branch by BJH method

162

Figure 45. TEM images of Nafion(30)/Me-SBA-16 showing characteristic planes for a


cubic pore structure: (a) [111], (b) [100], (c) [120]

163

Figure 46. Adsorption isotherms of water in SBA-16, Me-SBA-16, Nafion(30)/SBA-16


and Nafion(30)/Me-SBA-16 at 25 C

164

Figure 47. Catalytic performance of Nafion modified mesoporous materials. Experiment


conditions: T = 100 C; olefin WHSV = 2 h"1; isobutane/1-butne (I/O) molar
ration=40. Nafion(30)/Me-SBA-16 (A), Nafion(30)/SBA-16 ( ) ,
Nafion(15)/SBA-16(0), SAC -13 (),Nafion(15)/SBA-15 (A)

165

Figure 48. Variation of the conversion of butnes (a) and yield of TMPs (b) with time on
stream over Nafion(30)/Me-SBA-16 under varying experimental conditions:
100 C, WHSV = 1.4 h"\ I/O = 40 ( ) ; 100 C, WHSV = 2 h"1, I/O = 40 (A);
70 C, WHSV = 2 h"\ I/O = 40 (O); 100 C, WHSV = 2 h"1, I/O = 20 (A). 166

xv

1. Introduction
Alkylation occurs when isobutane is reacted with a light olefin, typically w-butene.
This process results in highly branched paraffin (called alkylate), mainly isooctanes that
have outstanding properties for use as blending components in reformulated gasoline.
Alkylate possesses high octane numbers and low volatility, has a high heat of combustion,
and is the cleanest burning of all other gasoline products (Albright and Wood, 1997).
Alkylate gasoline is produced in quantities of around 1,000,000 barrels/day in North
America alone and is as such an important component in the North American gasoline
pool, being surpassed only by reformate (composed mostly of aromatic hydrocarbons and
isoparaffins) and catalytic cracker gasoline (composed mostly of olefins). Several
drawbacks are associated with both reformate and catalytic cracker gasoline. The only
way to clean-burning, high-octane gasoline with no limitations imposed by the
specifications of the clean air regulations is to utilize branched alkanes. Therefore, the
demand for alkylate is increasing. It was reported in 2008 that 36 alkylation units are
either being enlarged or being constructed in the world (HPI Construction Boxscore,
2008).
The refinery process of isobutane alkylation with olefins, although a commercial
success, is based on liquid acid catalysts such as hydrofluoric acid and sulfuric acid,
which are very corrosive, and either relatively expensive to recover (H2SO4), or very
dangerous in case of an accidental release (HF). While the process produces a highquality, environmentally benign gasoline component, it suffers from a number of
disadvantages and shortcomings. For this reason, replacement of the liquid acids by solid
catalysts continues to be one of the most important research goals in the field of catalysis.
Researchers from the SunOil laboratories observed in pioneering studies that an
exchanged zeolite was active for alkylation catalysis at temperatures between 25-100 C
(Kirsch et a l , 1968). Since ca. 1980, a huge number of solid acids have been scrutinized
for their potential as catalysts in isobutane/butene alkylation. A major drawback, however,
evidenced in all evaluations of solid acid catalysts for alkylation is their rapid
deactivation (Albright, 1990).
1

Among the solid catalysts, zeolites have received much attention because of their
extreme regularity in pore shape and size, high surface area, and high strength and
amount of active sites generated in the framework. Due to the properties above, they are
successfully used as catalytic agents in petrochemical industry in cracking, reforming,
selective chemical productions and separation treatments, especially for molecules with
kinetic diameter smaller than 10 (Davis and Lobo, 1992; Marcilly, 2003; Degnan, 2003;
Cundy and Cox, 2003). However, despite their catalytically desirable properties, zeolites
and related microporous materials become inadequate if molecules with sizes larger than
the diameters of their pores have to be processed. The rapid catalyst decay appears
inevitable when zeolites are used as alkylation catalysts. It is generally accepted that the
formation and retention of heavy hydrocarbons within the zeolite micropores is the major
reason for deactivation. These heavy products either poison the active sites or block the
catalyst pores. In addition, the high polarity of the zeolites leads to the preferential
adsorption of olefin, resulting in much higher rates of oligomerization. This selective
adsorption is another reason to cause rapid deactivation of zeolites.
In trying to overcome the diffusion limitations in the micropores of zeolites, in
1992, Mobil researchers discovered a new silicate/aluminosilicate mesoporous molecular
sieve family with exceptionally large pores and uniform structures (Kresge et al., 1992;
Beck et a l , 1992). The synthesis of these mesoporous materials was based on a new
concept that a self-assembled molecular aggregate or supramolecular assembly of
surfactant molecules such as cetyltrimethylammonium (CTA) cation, rather than the
conventional single-(amine-) molecules (templates for microporous zeolite structures),
may be used as the structure agents to direct the one dimensional, hexagonal, and uniform
pore structure. Since then, many mesostructured materials including hexagonal SBA-15
(SBA is the acronym of Santa Barbara amorphous type material) with one-dimensional
channels and cubic SBA-16 with three-dimensional large spherical mesopores have been
developed (Zhao et a l , 1998a, 1998b).
These new materials possess very high surface area (about 1000 m /g) and their
pore diameters are in the mesopore range (2 < d < 50 nm), overcoming thus the pore size
constraints of zeolites. It is, however, difficult to obtain mesoporous materials showing

specific catalytic and adsorption properties such as a strong acidity and low polarity of
surface. In order to overcome these drawbacks, the attachment of organic functionalities
to the surface of the silica mesoporous supports has been intensively researched during
the last years. Several strategies have been described to anchor organic acidic groups to a
mesostructured silica surface, including covalent attachment of sulfonic groups on the
pore walls of mesostructured materials using grafting methods (Van Rhijn et al., 1998) or
co-condensation reactions (Huo et a l , 1996; Margolese et a l , 2000), impregnation of
acid resin in pore system (Wang and Guin, 2001; Martinez et a l , 2008), and entrapping
acid resin particles within a porous silica network using a sol-gel technique (Harmer et a i ,
1996). The acid strength can be tuned by choosing different sulfonic acid groups. For
example, the presence of electron-withdrawing fluorine atoms in the structure will
significantly increase the acid strength of the terminal sulfonic-acid groups, which
becomes

comparable

to

that

of

pure

sulfuric

acid.

Concerning

the

the

hydrophilic/hydrophobic character of the catalyst's surface, hydrophobization of the acid


sites microenvironment can be achieved by synthesizing periodic mesoporous
organosilicas (PMOs) from silane precursors (Asefa et a l , 1999; Inagaki et al., 1999) or
grafting trimethylsilyl groups (or similar) onto preformed catalysts (Macquarrie et a i ,
2005).
As mentioned above, alkylation is associated with a rapid catalyst deactivation
because of formation and retention of heavy hydrocarbons within the zeolite micropores.
It is generally accepted that large pore size and high acid strength are critical for
alkylation (Corma and Martinez, 1993). Thus, a catalyst with large pore size which
provides enough space for diffusion, high acid strength which is strong enough to
catalyze hydride transfer, and hydrophobic surface which allows a higher paraffin steady
concentration in the pores is desired. These characters may be achieved through
synthesizing and functionalizing mesoporous materials. Strong acid groups, such as
perfluorosulfonic acid groups, can be grafted on the pore wall, which still leaves enough
pore space. This facilitates the diffusion of reactants to the active sites and alkylation
products out of the pore network. Organosilicon moiety can be introduced to wall, which
tunes the polarity of surface. The acid functionalized mesostructured materials possess
obvious advantages as catalysts for isobutane/butene alkylation.
3

The present Ph.D work has been undertaken with the aim of synthesizing a series
of acid functionalized mesoporous materials and studying the influence of acid strength,
hydrophilic/hydrophobic character and pore structure of the materials on the performance
in alkylation of isobutane/1-butne.
The thesis is divided into eight chapters. A general introduction, including the
objectives of the research was elaborated in Chapter 1. The literature review which is
related to the subject of the dissertation was reviewed in Chapter 2. In this chapter, the
mechanism, catalysts and process of isobutane/-butene alkylation, and synthesis,
functionalization of mesostructured solid acids were reviewed. Chapter 3 describes the
apparatus and experimental procedures that were employed during the performing of this
Ph.D. work.
Chapter 4 describes the synthesis, characterization and catalytic evaluation of a
series of acid functionalized mesoporous materials. The influence of acid strength, acid
nature (Bronsted or Lewis acid sites) and hydrophilic/hydrophobic nature of the walls on
the alkylation reaction was investigated. Chapter 5 reports the synthesis, characterization
of a perfluoroalkylsulfonic acid functionalized PMO (periodic mesoporous organosilica).
The material was evaluated and compared with other solid acids in alkylation reaction.
This paper was published in Catalysis Communications, 10 (2008) 291-294. Chapter 6
reports the synthesis, characterization and catalytic evaluation of a mesostructured SBA15 modified by OH capping and impregnation of Nafion resin. This chapter was accepted
for publication in Applied Catalysis A: General. In Chapter 7, the synthesis,
characterization and hydrophobic property of SBA-16 with three-dimensional pore
structure were reported. Then, the effect of the pore structure and hydrophobic property
on catalytic performance was discussed. This chapter was recently submitted for
publication in Industrial & Engineering Chemistry Research.
Finally, Chapter 8 completes the thesis by summarizing general conclusions.

2. Mechanism, Catalysts, and Process of Isobutane/Alkene


Alkylation
2.1. Introduction
The catalytic alkylation of isobutene with light olefins, usually butnes, leads to
the formation of mixtures of branched alkanes, called alkylate, which is a highly valued
blending component for today's motor gasoline and for tomorrow's reformulated
gasoline. Alkylate has a high content of highly branched alkanes, such as
trimethylpentanes, that gives it a high octane number (-100). Moreover, it has a low
vapor pressure and a narrow distillation range and is free of olefins, aromatics, and sulfur.
Therefore it is an excellent blending component for gasoline.
Table 1. Octane number contributions from components of a typical US gasoline. (Miller,
1999)
Gasoline Component

PCT of Pool

Blending Octane

Contribution

Reformate

41.40

88

36.4

FCC gasoline

26.40

86

22.7

Alkylate

14.30

93

13.3

Isomerate

7.40

87

6.4

Alcohols

3.40

108

3.7

1.90

73

1.4

Hydrocracker gasoline

1.60

81

1.3

MTBE*

1.00

110

1.1

-Butane

1.10

92

1.0

1.20

78

0.9

TAME**

0.30

105

0.3

Total

100.00

Light straight-run gasoline

Light delayed coker gasoline

*MTBE - methyl-ter/-butyl ether **TAME - ter/-amyl-methylether

88.6

From the total US gasoline pool, alkylate constitutes the component having the
third most important contribution to the octane number (see Table 1 ).
Of all gasolines produced in a refinery, alkylates are the cleanest burning and
have the highest octane number. The alkylation unit plays an important role in a modem
fluid catalytic cracking (FCC) based refinery as it converts LPG-range byproducts from
the production of FCC naphtha into alkylate. With the increasing strictness of air
regulations in the E.U. and the U.S.A., the contents of alkenes, sulfur and aromatics,
particularly benzene, in the gasoline will become increasingly restricted. Besides alkylate,
other high octane number components that could constitute alternatives include the
butane fraction, which suffers from a high volatility and is only suitable for cold climates,
and oxygenate components, which possess a lower heat of combustion than hydrocarbons
and are either toxic (methanol), expensive or dangerous to the environment (ethers easily
contaminate aquifers), thus limiting their usefulness. Compared with methyl-tertiarybutyl ether (MTBE) and alcohols, alkylate does triumph over them by a large margin.
MTBE, which is a high octane oxygenate, has been found to cause drinking water to be
malodorous already in ppb concentrations (leaking out from underground storage tanks
into the ground water). In addition, it is also thought to be carcinogenic. As a
consequence, it will be phased out in several countries. Alcohols such as ethanol that
could conceivably replace the ethers as oxygenate source suffer from a very high
blending vapor pressure when mixed into gasoline, thus, limiting their usefulness. The
only way to clean-burning, high-octane gasoline with no limitations imposed by the
specifications is to utilize branched alkanes. Therefore, it is expected that the demand for
alkylate will increase dramatically as it is a very promising alternative to all the abovementioned components. It is non-reactive, nontoxic and has a low volatility.
The first attempt to carry out the alkylation of isoalkanes with alkenes was made
in the 1930s. Ipatieff and Grosse discovered that isoalkanes react with alkenes in the
presence of promoted aluminum chloride to give saturated hydrocarbons (Ipatieff and
Grosse, 1935). Alkylates were first produced in 1938 as a cooperative effort of several
American companies to produce higher-quality gasolines for the warplanes of that time.
These alkylates helped Allied planes perform better than enemy planes during World War
6

II. Sulfuric acid was used as catalyst at the early alkylation plants. Later, several plants
based on HF as catalyst were constructed, which are more flexible regarding the feed
alkenes. Alkylation experienced remarkable increase in the 1940s as a result of the
demand for high octane aviation fuel. After World War II, refiners' interest in alkylation
shifted from the production of aviation fuel to the use of alkylate as a blending gasoline
component. During the 1950s and 1960s, Capacity remained relatively flat during due to
the comparative cost of other blending components. However, the U.S. Environmental
Protection Agency's lead phasedown program in the 1970s and 1980s further increased
the demand for alkylate as a blending component for motor fuel. Furthermore, the
phasedown of methyl-tertiary-butyl ether (MTBE) also boosted the demand for alkylate
to meet the requirements for reformulated gasoline (Hartley et. al., 1999). Hydrocarbon
Processing (HPI Construction Boxscore, 2008) reported that 36 alkylation units are either
being enlarged or being constructed. In the U.S., seven units are being enlarged. Seven
units are being constructed in Asia. The remaining plants are located in Europe, Africa,
South America, Canada, and Mexico. Most of the new capacity will employ units
designed by Stratco, Phillips-Conoco, and UOP.
Since the 1940s, five processes based on liquid acid catalysts have been widely
used. Sulfuric acid processes were designed by Stratco (which was purchased by the
DuPont Co. several years ago), Exxon-Mobil, and M.W. Kellogg Co (Albright, 2003).
No new Kellogg units have been built in the past 35 years, but a significant number of
their plants are still operating. The reactors of ExxonMobil and Kellogg are similar in
design. Phillips Petroleum (now Conoco-Phillips) and UOP designed units using HF as
the catalyst. Their reactors are of a different design. UOP announced in December 2007
that they had purchased the Conoco-Phillips alkylation technology and hence could
broaden their alkylation offerings (Albright, 2009).
In the 1950s, about 75% of the alkylate was produced by sulfuric acid process.
During the next 30 years, the importance of HF grew until about 50% of the alkylate was
produced by HF. Then, two incidents occurred, one in 1986 and the second in 1987, that
clarified the dangers of HF (Albright, 2003). Since then, the relative importance of
alkylations with sulfuric acid has increased. As to the end of 2001, the worldwide
7

production capacity for alkylation gasoline amounted to 74 million ton/year, nearly equal
amounts of alkylate are produced by HF and H2SO4 processes (Stell, 2001).
Although both processes are considered to be mature, a number of drawbacks are
inherent to both of them. The sulfuric acid processes tend to suffer from unusually high
catalyst consumption. It has been estimated that replacement of spent sulfuric acid in
alkylation plants accounts for about 89% of the total volume of alkylation catalysts
(Furimsky, 1996). Furthermore, the spent acid containing conjunct polymers forms a
toxic sludge that must be sent to a recovery facility. The transport of spent and fresh acid
to and from the sulfuric acid regeneration plant has given rise to some concern and
increased the pressure on refiners to establish sulfuric acid regeneration plants near the
alkylation unit. Indeed, the cost of the recovered acid has been estimated to be two to
three times that of sulfuric acid available on the market (Furimsky, 1996). About onethird of the operating cost of H2SO4 process can be attributed to acid consumption
(Albright, 1998). However, no major new improvements have been introduced in sulfuric
acid alkylation technology dealing with the spent acid issue.
The process with HF demonstrates superior activity for production of high octane
alkylate because of the superior solubility of isobutane in HF. In addition, catalyst costs
are much lower for HF processes relative to those involving H2SO4. However, the use of
anhydrous HF as catalyst has caused significant concern because of its toxicity, high
vapor pressure and tendency to form aerosols clouds containing lethal levels of HF that
can drift downwind for at least several miles. An accident at a Texas refinery released
considerable amounts of gaseous HF. As a result, over 1000 individuals in the
neighborhood were evacuated. Since HF from a chemical point of view is a good and
cheap catalyst for isobutane alkylation, attempts have been made to improve the safety of
HF alkylation. In order to minimize these potential problems and reduce the risks, a
number of safety improvements have been developed by refiners who operate
hydrofluoric acid alkylation units. However, these methods have not been completely
effective at eliminating the dangers associated with this acid. Moreover, some
industrialized countries have ceased to license new HF alkylation units.

Since

1990, there has been substantial pressure to develop a more

environmentally friendly alkylation process. Economic advantages also exits in replacing


liquid acid technology. The rate of catalyst consumption for H2SO4 alkylation processing
is two orders of magnitude (by weight) greater than that for catalytic cracking catalysed
by zeolites. It encouraged research in industry and academia to find suitable replacements
for the existing liquid acid catalysts. With heterogenized gaseous and liquid strong acids
(Child et. al., 1990; Cooper et. a l , 1992; Huss and Jonson, 1993), the risk of leaching of
the "immobilized" acid always remains, requiring continuous addition of fresh acid. The
development of true solid acid catalysts receives much recognition. Zeolites, being
noncorrosive, non-toxic and rather inexpensive materials, were the first solid acids tested
as alternatives to sulphuric and hydrofluoric acid in iso-butane/alkene alkylation
(Garwood and Venuto, 1968; Kirsch et. a l , 1972). The main drawback in the use of
zeolites for iso-butane/butene alkylation is its rapid deactivation, which up to date has
impeded its industrial application. Thus, in order to achieve an economically feasible
industrial process, the catalysts must be frequently regenerated. Multiple regenerations
are needed for processes based on solid catalysts to be competitive with existing
processes based on H2SO4 and HF.
Other

materials

studied

are

sulfated

zirconia

and

related

materials,

heteropolyacids, acidic polymers, supported metal halides. On the whole, these materials
also deactivate rapidly and some of them additionally exhibit environmental and health
hazards.
Up to now, none of these solid acid catalysts met any commercial success on
isobutane/butene alkylation. This is associated with a rapid catalyst deactivation for the
generation of isooctane (Walker, 2000). However, several demonstration units based on
solid acids were built. At least three types of catalyst beds have been applied; they are
packed beds, fluidized beds and moving bed. Exelus developed a solid acid-catalyzed
alkylation process using packed beds. The catalyst is reactivated about every 12 h
(Mukhergee et a l , 2007). Two reactors are used: one is on stream while the other is being
regenerated. For the alkylene process of UOP, Roeseler et al. reported that tests indicated
over 9000 cycles in over 9 months (Roeseler et a l , 2002). Calculations indicate that
9

reactivations were made approximately every 0.7-0.8 h. Hydrogen gas is employed for
reactivations. The UOP catalyst has platinum deposited on its surface; hence, it acts as
the catalyst both for hydrogenating the "deactivating species" and also for alkylation.
UOP has reported that, in its units, the catalyst is partially deactivated in a moving-bed
reactor and then completely reactivated in another reactor. Oxygen or air has also been
employed for the reactivation of other catalysts.
We believe that the fast deactivation of solid acid catalysts is due to the
preferential adsorption and pore filling with the olefin. During solid acids catalyzed
isobutane/butene alkylation, a highly unsaturated and highly branched polyolefine is
formed. The polymer strongly adsorbs on the acid sites and completely fills the pores
thus ending the reaction.
The newly discovered mesoporous molecular sieves have opened a new era in the
development of solid catalysts (Trong On et. al., 2003; Taguchi and Schuth, 2005). The
pores of these materials have strictly the same diameter in the range 2 to 50 nm. Many
catalytically active functional groups, such as acid groups, can be grafted on the pore wall,
which still leaves enough pore space. This facilitates the diffusion of reactants to the
active sites and alkylation products out of the pore network. Mesoporous molecular
sieves have very different adsorption behaviors when different silicon source are used,
such as organic/inorganic hybrid materials (Asefa et. al., 1999). Several such materials
have already been synthesized with a composition comparable to that of silicon rubber,
namely siloxane based materials. The carbophilicity of the organosilicon moiety of walls
should allow a higher paraffin steady concentration in the reaction conditions, which may
decrease the formation of high hydrocarbons and prolong the lifetime of solid acid
catalysts. The acid functionalized mesostructured materials should be promising catalysts
for isobutane/butene alkylation.
The purpose of this chapter is to review the processes of isobutane/butene
alkylation reaction using solid acid catalysts and the synthesis of acid functionalized
mesostructured materials.

10

2.2. Isobutane/butene alkylation reaction mechanism on solid acids


2.2.1. Product distribution
Today, sulfuric acid and hydrofluoric acid are the commercially used catalysts for
alkylating isobutane with C3 to C4 olefins. The reaction temperature for H2SO4 catalyzed
isobutane/butene alkylation is lower than 10 C and molar ratio of isoparaffin/olefin (I/O
ratio) is between 5 and 8. The HF catalyzed process applies higher reaction temperature,
between 10 to 40 C and higher I/O ratio in the range of 10 to 15. Octane values of
alkylate produced through liquid acid-catalyzed processes range from 95 to 98.
The reported composition of alkylates produced with different acidic catalysts are
shown in Table 2. It can be seen that the product distribution is similar within a variety of
acidic catalysts, both solid and liquid, and over a wide range of process conditions.
Alkylate is a complex mixture of branched hydrocarbons. Trimethylpentanes (TMPs) are
the preferred products since they possess high octane number (Research Octane Number
> 100). The rest of Cg products are mainly dimethylhexanes (DMHs), which are
secondary products and possess an octane number between 55 and 82. Clearly, higher
TMP/DMH ratio is preferred in alkylation process. Besides Cg, alkylate also contains
compounds of C5-C7 and C^. The complexity of the product illustrates that a set of
parallel and consecutive reaction steps, like isomerization, oligomerization, /7-scission
and hydride transfer have to be involved.
Table 2. RON and composition of the various alkylates produced with different acidic
catalysts
Research
Component (wt%)

Octane
Number

Catalyst
Sulfated

H2SO4

HF

USY

T=258K

T=?

T=323K

T=275K

P/0=5

P/0=12

P/0=15

P/0=15

Zirconia

Isopentane

93

1.2

1.8

15.6

24

n-Pentane

61.8

0.1

0
11

2,2-Dimethylbutane

91.8

0.8

2,3-Dimethylbutane

104.3

1.5

1.4

5.7

4.3

2-Methylpentane

73.4

0.2

1.7

3.5

3-Methylpentane

74.5

0.1

0.1

2.2

1.7

n-Hexane

24.8

2,2-Dimethylpentane

92.8

1.3

0.1

2,4-Dimethylpentane

83.1

0.6

1.3

3.5

5.5

2,2,3-Trimethylbutane

112.1

0.1

0.4

0.3

3,3-Dimethylpentane

80.8

0.3

2,3 -Dimethylpentane

91.1

0.6

0.6

2.9

1.8

2-Methylhexane

42.4

0.1

3-Methylhexane

52

0.2

0.5

0.7

2,2,4-Trimethylpentane

100

30.2

48.7

11.4

25.5

n-Heptane

2,2-Dimethylhexane

72.5

0.4

2,4-Dimethylhexane

65.2

1.2

2.9

0.4

0.8

2,5-Dimethylhexane

55.5

2.1

0.7

2,2,3-Trimethylpentane

109.6

0.8

1.1

3.3

11

2,3,4-Trimethylpentane

102.7

33.9

21.4

8.0

2,3-Dimethylhexane

71.3

1.7

2.1

4.5

0.9

2-Methylheptane

21.7

2,3,3 -Trimethylpentane

106.1

20.4

12.9

8.2

7.4

3,4-Dimethylhexane

76.3

0.2

0.2

4.3

0.4

3-Methylheptane

26.8

Octenes

>90

0.1

1.3

C9+

s 80-85

5.4

2.9

23.1

3.3

Data taken from references (Albright et. al., 1988) for H2SO4, (Hutson and Logan, 1975)
for HF, (Corma et. al., 1994a) for USY and (Corma et. al., 1994c) for sulfated zirconia

12

Primary products should come from the direct addition of the olefin to isobutane.
The product distribution obtained experimentally from liquid and solid acids is far from
equilibrium and depends on conditions. Table 3 gives the comparison of TMP and DMH
selectivity of experiment and thermal equilibrium. The distribution within the
trimethylpentanes is not equilibrated. It illustrates that rearrangement of alkyl carbenium
ions formed occurs rapidly, leading to a wide distribution of different isomers.
Table 3. Comparison of TMP and DMH selectivity on liquid and solid acid catalysts,
with thermal equilibrium composition
Catalyst

HF

H 2 S0 4 REHY

Thermal Equilibrium Composition

Temperature C

n.a.

-10

90

-20

50

90

127

2,2,4-TMP

65.5

47.0

29.8

67.1

63.5

61.4

59.2

2,2,3-TMP

2.5

3.1

5.6

16.8

17.5

16.5

16.6

2,3,4-TMP

17.0

26.5

25.8

8.0

9.0

10.8

11.2

2,3,3-TMP

14.9

23.4

38.8

8.1

10.0

11.3

13.0

2,5-DMH

21.4

64.7

23.9

55.0

30.4

2,4-DMH

34.8

49.2

36.2

25.5

2,3-DMH

38.4

6.0

4.0

4.0

TMP Dist.[%]

DMH Dist.[%]

2,2-DMH

29.4
2.0

22.2

3,3-DMH
3,4-DMH

12.7
5.4

4.0

20.9

4.8

5.1

Data taken from references (Chu and Chester, 1986; Cardona et. al., 1995; Corma and
Martinez, 1993)

13

2.2.2. Proposed isobutane/butene alkylation reaction mechanism on


solid acids
2.2.2.1. Protonation of the alkene and formation of tert-C^ ion
It is generally accepted that isoparaffin-olefin alkylation proceeds via a carbenium
ion mechanism in the presence of a liquid acid catalyst (Corma and Martinez, 1993). The
first step of the mechanism is the protonation of the olefin. This reaction follows
Markovnikov's rule. On solid acids, the first step of the mechanism is the protonation of
the olefin by a Bronsted acid site (H+surf) to form a secondary butyl carbenium ion. The
adsorption of butyl carbenium ions on the basic surface oxygen atoms of solid acids leads
to more or less covalent surface alkoxides rather than adsorbed carbenium ions. The
carbenium ions are transition states (Kazansky, 1999; Rigby et al., 1997). The lowest
energy adsorbed state of an alkyl species on a zeolite is an alkoxide (Haw, 1989;
Derouane et a l , 1999), in which a covalent bond is formed between an oxygen atom of
the zeolite and a carbon atom of the alkyl species. The alkene is "solvated" by the basic
surface oxygen atoms, similar to the solvation through water in aqueous solutions.
When isobutane is alkylated with 1- or 2-butene on liquid or solid acid catalysts,
some common features in product distribution can be observed: same products are
obtained with the distribution slightly varied; Cg is the predominant product in Cs+; TMPs
are predominant in Cg paraffins; the content of -paraffins and olefins is low (Kirsch et
a i , 1972; Walker, 2000). That means interconversion of 1- and 2-butene occurs on solid
acid catalysts under alkylation conditions.
The protonation of isobutane will give a trimethylcarbenium ion and hydrogen or
a dimethylcarbenium ion and methane via a pentacoordinated carbenium ion. This
reaction occurs only at temperature higher than 473K. Under alkylation conditions, the
protonation reaction has to be initiated by an olefin (Sanchez-Castillo et a l , 2002).
When -butnes are used, the initiation will produce a secondary carbenium
ion/butoxide. This species may isomerize via a methyl-shift (reaction 1) or suffer hydride

14

transfer from a molecule of isobutane to give -butane and a more stable tertiary butyl
cation (reaction 2).
Methyl-shift
(Surf)

(1)

(Surf)

Hydride transfer
(Surf)

The methyl-shift in reaction 1 has to go through a transition state, which


resembles a primary carbenium ion. The activation energy of this transition state is about
130 kJ/mol (Boronat et a l , 1998). Hydride transfer to secondary carbenium ions such as
the sec-butyl carbenium ions is known to be much slower than hydride transfer to tertiary
carbenium ions, which explains that except butane, only alkanes with tertiary carbon
atoms can be observed in the products. The transition state for H" transfer between
isobutane and protonated -butne is in the form of two-electron three-atom. The
activation energy is calculated to be negative. Once an active carbonium ion is generated
by strong acids, the reactions occur spontaneously (Wang et a l , 2007). So hydride
transfer is a more likely event to form a more stable, tertiary butyl carbenium ion (tertC4*) and -butane, -butane being formed in substantial amounts in initial reaction stage
confirmed this mechanism . Anyway, hydride transfer to secondary carbenium ions is
very slow. An induction period is observed until a high concentration of tert-C<f has been
generated (Corma and Martinez, 1993).
Another route of tert-C* formation is hydride transfer from isobutane to an
adsorbed tert-R+. Hydride transfer to tert-df is faster than to sec-R+ (Cardona et al.,
1995). This reaction is more important for regeneration of tert-C^ in alkylation stage.

15

2.2.2.2. Addition of tert-Cf ion to second olefin and isomerization


The tert-Ctf can undergo electrophilic addition to a second butne molecule to
form the corresponding Cg+ carbenium ion (see reaction 3, 4, 5). Isomers of butne are
present due to the rapid isomerization of 1-C4~ to 2-C4~ and /-C4~ on solid catalysts (Chu
and Chester, 1986). The addition is exothermic and it contributes most of all the reaction
steps to the overall heat of reaction. The DFT calculation results show that the reactions
proceed easily in the presence of strong acids (Wang et a l , 2007).

+ \

(Surf)

(Surf)

(3)

2,2-DMH+

(Surf)

(4)

2,2,3-TMP +

(Surf)

/X

(5)
2,2,4-TMP+

(Surf)

Using 1-butne as the feed alkene does not primarily lead to dimethylhexanes or a
mixture of 2,2-DMH, 2,2,3-TMP and 2,2,4-TMP as expected, but to a mixture of TMPs.
This is due to a rapid isomerization of the linear butnes almost to equilibrium
compositions. Beside TMPs, some of the isomers of DMHs also produced because of the
isomerization.
The octyl carbenium ion may undergo rapid isomerization via hydride and
methyl shifts to form more stable carbenium ions. Hydride shifts are much faster than
methyl shifts. The rearrangements, which are possible for 2,2,3-TMP+, are illustrated in
16

Figure 1. The energies of 2,2,4-TMP+, 2,3,4-TMP+ and 2,2,3-TMP+ are -314.80211,


-314.82465 and -314.82086 a.u., respectively. These values indicate that the order of
stability is 2,2,4-TMP+ > 2,3,4-TMP+ > 2,2,3-TMP+ (Wang et a l , 2007), which agrees
with the selectivities of TMPs on a typical La-X catalyst (Feller et a l , 2004).

(Surf)
=CH,

2,3,4-TMP+

=CH,
(Surf)

=CH,
(Surf)
2,2,4-TMP+

Figure 1. Isomerization of 2,2,3-TMP+ via methyl and hydride shifts


2.2.2.3. Hydride transfer to Cs+ ion
Hydride transfer involves the shift of a hydride (H-) from a fluid-phase
isobutane to the adsorbed alkyl fragment. The octyl carbenium ion can undergo hydride
transfer from isobutane to form the corresponding isoparaffin. The tert-butyl cation is
regenerated and the chain sequence can continue. The reaction involving 2,2,4-TMP+ is
shown as reaction 6.

(Surf)

(Surf)

(6)

Hydride transfer from isobutane is the key step in the alkylation reaction catalyzed
by acidic catalysts. It prevents long hydrocarbon buildup and keeps the reaction going
along in the desired propagation cycle. The relative rate of hydride transfer to an
adsorbed Cg alkyl species controls catalyst deactivation during alkylation. In general, the
17

hydride transfer between alkanes and alkyl-carbenium ions is the elementary step
responsible for chain propagation of acid-catalyzed transformations of hydrocarbons
(Cumming and Wojciechowski, 1996). On US Y catalyst, relative amounts of -butane
and TMPs in the production during the initial stages of the reaction indicate that the
formation of tert-butyl cation via reaction 6 is approximately 15 times faster than that
produced via reaction 2 (Cardona et al., 1995). Hydride transfer is a relatively rapid
process if both the reactant and the product carbenium ion are tertiary whereas it is much
less favoured if a chemisorbed tertiary carbenium ion reacts with an alkane, which can
only give a secondary or a primary carbenium ion. Activation energies for hydride
transfer between isobutane and secondary acceptor cations were almost twice the
energies for tertiary/secondary hydride transfer (Nowak et al., 1996). This is due to the
decreased stability of the less-substituted secondary carbenium ion transition state
compared with the tertiary carbenium ion transition state discussed earlier. Stabilization
of the carbenium ions by isobutane in the form of a shared-hydride structure is weak, and
thus the driving force for hydride transfer is small. The hydride transfer rate may be
expected to be highest over a solid acid that best stabilizes the carbenium-ion transition
state. However, the relative rate of the hydride transfer step to the alkylation step controls
catalyst deactivation, and stabilization of carbenium ions on the catalyst surface may
enhance the alkylation rate as well.
Simpson et al. fitted a simplified model of the alkylation mechanism to initial
reaction rate data to determine the intrinsic relative rate of alkylation and hydride transfer
over an ultrastable Y-type zeolite (Simpson et al., 1996). The ratio of the rate constant of
alkylation to that of hydride transfer was found to be 8000 at 323 K and 3800 at 373 K,
yielding an activation barrier for alkylation that is 14.9 kJ mof 1 lower than that for
hydride transfer and a preexponential factor for alkylation that is 30 times greater than
that for hydride transfer. A density functional theory study of the alkylation of isobutane
with butne over phosphotungstic acid was made by Janik et al. The calculations show
that the activation barrier for alkylation of a /-butyl alkoxide was 10 kJ mol -1 lower than
that for hydride transfer, in good agreement with the experimental results over zeolite
catalysts (Janik et al., 2006). The interaction of a carbenium ion with an alkene is
18

substantially longer-range than the interaction with an alkane. Thus, the pre-exponential
factor for alkylation would be greater for the alkylation step, because an encounter
between an alkene and a carbenium ion is likely to show a greater sticking coefficient
than an encounter between an alkane and a carbenium ion.
Zeolites have a high initial alkylation activity associated with a rapid catalyst
deactivation. Both the activity and the selectivity to Cg alkanes decrease with time on
stream (Simpson et a l , 1996; de Jong et a l , 1996; Weitkamp and Traa, 1999; Walker,
2000; Yoo and Smimiotis, 2003; Feller et a l , 2004). As the olefin conversion decreases,
the fraction of higher alkenes in the product stream typically increases, and at long
reaction times, only butne oligomerization is catalyzed. This indicates that at long
reaction times, the hydride transfer step ceases, and thus the loss of activity and
selectivity is assumed to result from a decreased ability to promote the hydride transfer
step (Simpson et a l , 1996; Cardona et a l , 1995; Weitkamp and Traa, 1999; Feller et a l ,
2003a; Platon and Thomson, 2005).
Hydride transfer does not proceed exclusively between isobutane and carbenium
ions. Carbenium ions can abstract hydrides from other hydride donors, such as cyclic and
unsaturated compounds presented in conjunct polymers. These species can abstract a
hydride from iso-butane to form the tert-butyl carbenium ion, and they can give a hydride
to a carbenium ion, producing the corresponding alkane, for example the TMPs, as shown
in reactions 7 and 8. However, heavy hydrocarbon molecules trend to deactivate the
catalyst by pore blocking or site blocking.

Hydride transfer

(Surf)

(7)

(8)

19

2.2.2.4. Oligomerization, disproportionate)!! and cracking


Alkylation of butne with isobutane is not the only reaction occurring under
alkylation conditions. A number of side reactions occur simultaneously, which generally
tend to reduce the quality of the alkylate. These reactions including: oligomerization,
disproportionation, cracking, and self-alkylation reactions.
Oligomerization reactions result from the addition of a second olefin to the octyl
carbenium ion, as shown in reaction 9.

(9)

* c +

,2

(Surf)

(Surf)

(Surf)

The C i2+ can continue to react with an olefin to form a larger isoalkyl cation (see
reaction 10).

c,16fi +

12,

(Surf)

(Surf)

(10)

C,6

(Surf)

Oligomerization and cracking are responsible for the formation of light and heavy
ends, i.e., of iso-alkanes with odd carbon numbers and also partially responsible of some
of the Cg compounds produced during the alkylation reaction. A general oligomerization
scheme is presented in Figure 2 (Simpson et a l , 1996).

ci

+
K

KBl

ri+y

Kc2

Cn+y
A1
K

Kc3
--n+2y

A2

'n+2y

cy=
KA3

KB3

B2

Cn+y

'n+2y

Figure 2. Pathway to oligomerization products with the corresponding rate constants.


20

In the oligomerization scheme KA defines the rate of alkene addition, KB defined


the hydride transfer rate and Kc the rate of deprotonation. The ratio between hydride
transfer and the combined olefin addition and deprotonation is one of the main
parameters that determines the lifetime of a catalyst under defined reaction conditions.
Cn+ refers to any carbocation present in the reaction medium coming either from the
classical alkylation reaction or from any other source as, i.e., cracking, that can undergo
oligomerization, alkylation or deprotonation reactions. Cy~ refers to any olefin present in
the reaction medium coming either from the feed or as a product of cracking or
deprotonation reactions.
The overall product distribution of isobutane/butene alkylation is governed by the
relative rates of alkene addition and hydride transfer. Alkene addition is much more facile
than hydride transfer. With sulfuric acid, -butne oligomerization was reported to be
four times faster than hydride transfer (Lee and Harriott, 1977). With zeolites, the
preferential adsorption of olefins over paraffins, as well as the concentration effect in the
small zeolite cavity, favors oligomerization over hydride transfer. Oligomerization was
reported to be two orders of magnitude faster than hydride transfer by de Jong et al. (de
Jong et al., 1996) and three orders of magnitude by Simpson et al. (Simpson et al., 1996).
With too low internal paraffin/olefin ratios the alkenes will oligomerize before they can
be removed via hydride transfer. This is the key problem in solid acid catalyzed
alkylation. A polymer will build up, which will finally block the acid sites.
Disproportionation is the combination of two molecules of the same type to form
two molecules of differing type, one of higher molar mass and the other of lower molar
mass.
2C 8 H 18

C 7 H I6 + C9H20

(n)

Several experiments have been done to measure the relative rate of disproportionation of
some isoparaffins in contact with acids and in absence of olefins (Corma and Martinez,
1993). It was found that 2,3,4-TMP disproportionates most readily and 2,3-DMB the least.
Isopentane, 2,2-DMB, and 2,2,3-TMB were found not to disproportionate.
21

Compared with disproportionation, cracking is much more important to produce


products with carbon numbers which are not multiples of four, such as C5-C7 and C9, C10
and higher hydrocarbons. C racking is connected to oligomerization. The heavy cations
formed by oligomerization have the tendency to fragment, forming C 4-C 16 cations and
alkenes, according to the ^-scission rule (see reaction 12). The produced iso-pentene
either will be protonated or added to another carbenium ion. With a butyl, this would lead
to a nonyl carbenium ion. The formed carbenium ion fragment can receive a hydride and
leave the reaction as a heptane, or possibly add a butne to form a C 11 carbenium. With
hydride transfer, another alkane with an odd carbon number is produced. This small
example shows the huge variety of possible reactions. By means of GC analysis Albright
and Wood (Albright and Wood, 1997) found about 100-200 peaks in the C 9-C 16 region,
regardless of the alkene and acid employed. C racking reactions are favoured by strong
acid catalysts and elevated temperatures (Corma et al., 1994c).

(12)

Cracking of hydrocarbons under typical alkylation conditions is solely produced by


/?-scission. Weitkamp et al. introduced a useful classification of different modes of /?scission, as shown in Figure 3 (Weitkamp and Traa, 1997).
Sv

T\peD:
Ri

u>6

TypeC:

sec.^prirji

'+N

^ R =

R.

sec.sec.
w

R:

^ R =
1

u>7

u>7

n>8

T}TBr

T>TB::

sec.-tert.

*?

-^^-<iN

R.

T>pe.A
*,X

R.
\

u
\x
I

+
1

ten. sec.
R

1
ten ten.
w

R r^ +

K
1

R.

Figure 3. Different modes of ^-scission


22

Type A ^-scission, which is the most rapid, starts from a tertiary carbenium ion to
give again a tertiary carbenium ion and an alkene. Type B ^-scission, which is slower
than type A /7-scission, starts from a tertiary to give a secondary or vice versa. Type C Bscission, which is still slower, starts from a secondary carbenium ion to give another
secondary carbenium ion. Type D /?-scission, which is the slowest mechanism, start from
a secondary carbenium ion and yield a primary carbenium ion.
Besides, other side reactions may occur. The sec-butyl carbenium ion may react with
another 2-butene molecule to form 3,4-dimethylhexane (DMH) cations. These DMHs+
may then undergo any number of hydride or methyl shifts thus leading to all possible
DMH isomers.
2.2.2.5. Self-alkylation
Self-alkylation of isobutene involves the alkylation of isobutane with isobutene,
which is formed from another tert-C4+. The primary product is 2,2,4-TMP. At the same
time, a saturated paraffin of the same carbon number as the tert-C^ is obtained. The
importance of self-alkylation depends on the acid, the alkene and the reaction
temperature. Generally, HF and zeolites show more activity for self-alkylation than
sulfuric acid. An example is shown as follows:

(13)
(Surf)
(Surf)

(Surf)

^V

H-

(14)

(15)
(Surf)

(2,2,4-TMP+)

(Surf)

23

The overall reaction with a linear alkene CH2 as the feed can be written as follows:
2/-C4H10 + CH2n

/-CgH.g + CH2+2

(16)

The crucial step in self-alkylation is decomposition of the butoxy group into a free
Bronsted acid and isobutylene (see reaction 14).
2.2.2.6. Deactivation mechanism of solid acid catalysts
Rapid catalyst deactivation is the key problem of isobutane/butene alkylation. Many
attempts have been made over the past 20 years to use solid acids as catalysts of this
reaction, none of these attempts met any commercial success because of short catalyst
lifetime.
Deactivation appears inevitable on any studied solid acid catalyst system. Almost
all the authors agree that deactivation is due to the formation and retention within the
zeolite micropores of heavy products. Meanwhile, the knowledge of their composition
and of their effect on the active sites (poisoning, blockage of their access) is
indispensable for finding ways to increase the catalyst stability and to design an efficient
regeneration method. Some characterization techniques were applied to study
carbonaceous deposits and the deactivated zeolite catalysts. They are elemental analysis
(Klingmann et a l , 2005), 13C NMR spectroscopy (Weitkamp and Maixner, 1987; Stocker
et a l , 1994; Flego et a l , 1995), FT-IR spectroscopy (Flego et a l , 1995; Pater et a l , 1999;
Diaz-Mendoza et a l , 1998; Nivarthy et a l , 1998a; Feller et a i , 2003b; Petkovic and
Ginosar, 2004; Petkovic et a l , 2005), UV-vis spectroscopy (Klingmann et a l , 2005;
Flego et a l , 1995; Petkovic and Ginosar, 2004; Petkovic et a l , 2005), temperature
programmed oxidation (Petkovic and Ginosar, 2004; Petkovic tal., 2005; Querini, 1997 ;
Querini, 2000), matrix

assisted

laser desorption/ionization

time-of-flight

mass

spectroscopy (MALDI-TOF) (Feller et a l , 2003b) and dissolution of the zeolite in HF


solution with subsequent extraction and analysis of the organic phase (Pater et a i , 1999;
Feller et a l , 2003b) by GC , FT-IR, H NMR, MS and GC /MS coupling. The
characterization of the deactivated catalysts is also indispensable, in particular for
specifying the interaction between the molecules of carbonaceous compounds and the
24

acidic sites. Most of the authors agree with the presence of heavy (up to 30-35 C atoms),
unsaturated highly branched species containing cycles. The atomic H/C ratio was 1.6-1.8.
Aromatic compounds were found by Feller et al and their amount increased with reaction
temperature. The number of unsaturations in the carbonaceous compounds was
concluded to be 2 from the presence of a band at 305 nm in the UV-vis spectrum
originating from monoenylic carbocations (Klingmann et a l , 2005); on the other hand,
from hydrognation of carbonaceous compounds, Pater et al. suggested the presence of
only one double bond (Pater et a l , 1999).
IR spectroscopy analysis of the "coked" zeolite samples confirmed the interaction
between coke molecules and the protonic sites. The bands corresponding to the bridging
hydroxyl groups disappeared at high (130 C) and low (40 C) temperature (Pater et a l ,
1999; Feller et a l , 2003b). This interaction is responsible for the catalyst deactivation.
The formation of heavy hydrocarbons is the major reason for deactivation of
catalysts. It is reasonable to limit the growth of product molecules in order to increase the
catalyst lifetime. Competing reactions produce undesired side products that ultimately
lead to heavy hydrocarbons. The Cg species can desorb as an alkene and subsequently
alkylate an adsorbed t-butyl species rather than undergo hydride transfer. The adsorbed
Cg species can combine with another butne molecule to produce an adsorbed C12 species,
which can further react. Skeletal isomerization and ^-scission reactions lead to a wide
distribution of products. Alkene oligomerization can also occur, leading to the production
of Cg alkenes or other heavier hydrocarbons. The formation of multiple alkylates (C12 and
greater) is directly tied to catalyst deactivation. Heavy hydrocarbons may plug catalyst
pores or block active catalyst sites, leading to deactivation. Ideally, hydride transfer from
isobutane to the adsorbed Cg species occurs before further alkylation or alkene desorption,
thus terminating the growing chain and preventing heavy hydrocarbon buildup. So the
lifetime of the alkylation catalyst is mainly determined by the relative rates of hydride
transfer and of oligomerization. The higher the hydride transfer to oligomerization ratio,
the greater the yield in alkylation products and the smaller the formation of carbonaceous
deposits hence the slower the deactivation. However, the alkylation step is estimated to
occur at a rate two to three orders of magnitude faster than hydride transfer (Simpson et
25

a l , 1996; de Jong et a l , 1996). In order to promote hydride transfer, alkylation is


performed industrially in a large excess of isobutane over butne to maximize the relative
hydride transfer rate (Corma and Martinez, 1993).
Though oligomerization is a much more facile reaction, a stability increase of
hydride transfer can be obtained by an adequate choice of the zeolite catalyst, of the
reactor and of the operating conditions (Weitkamp and Traa, 1999; Feller et a l , 2004; de
Jong et a i , 1996). Hydride transfer requires strong acid sites with in addition a significant
positive effect of their density; moreover, this reaction involves a bulky bimolecular
transition state. As a consequence, a large pore zeolite such as FAU with the greatest
possible concentration of acidic sites (e.g. X rather than Y) will be chosen. TMP
molecules being relatively bulky, limitations in their desorption from the zeolite
micropores could occur with therefore an increase in the secondary reactions of
condensation involved in the formation of carbonaceous deposits. The use of zeolites
with small crystallites will reduce these limitations. Thus, the small size of the HBEA
crystals could be one of the reasons for the good stability of this zeolite (Loenders et a l ,
1998). Lastly the significance of oligomerization which involves two or more molecules
of butne will be lower if -butne concentration is kept very low throughout the reaction.
Reaction temperature was also shown to have a significant effect on the lifetime with a
maximum at 75 C (over a LaX catalyst). Faster deactivation at low temperatures would
be due to diffusion limitations with catalyst deactivation by pore blocking, whereas at
high temperatures it could be due to a faster formation of highly unsaturated compounds
which strongly adsorb on the active sites (site poisoning) (Feller et a l , 2004).
The most common deactivation mechanisms in alkylation processes catalyzed by
solid acids were summarized by Guisnet et al. (Guisnet and Magnoux, 1997):
Poisoning of the active sites by feed components and impurities, or by nondesorbed heavy products (coke).
Blockage of the access of the reactant to the active sites (or of the adsorbate to the
pores) by coke, by extra-framework species resulting from dealumination, etc.
26

Structure alterations.
Sintering of supported metals (e.g., in bifunctional catalysts).
The primary reason for rapid catalyst deactivation is the loss of the hydride transfer
function, which ultimately favours oligomerization and polymerization reactions and
leads to coke formation. Deactivated zeolite catalysts contain large amounts of heavy
hydrocarbons (Querini, 2000).
2.2.2.7. Catalyst requirements
Alkylation is associated with a rapid catalyst deactivation because of coke
formation. Coke formation is directly tied to the preferential adsorption and pore filling
with the olefin. Certain properties can be identified as necessary to achieve a high
alkylation yield and selectivity.
1. Bronsted acid sites of high strength are preferred. Strong acid sites are
necessary to form and stablilize the intermediate carbocations. They faciliate
hydride transfer reaction and surpress oligomerization and polymerization
reaction, which are beneficial for good activity and extended reaction times.
2. A narrow distribution of acid strength is expected. The catalyst should not
promote undesirable side reactions, such as oligomerization, polymerization or
cracking of Cg product. The catalyst requirements for the various reactions are
different. Cracking reactions demand strong acid sites, whereas oligomerization
and polymerization reactions require medium strength acid cites. Double bond
isomerization can be catalyzed by weak acid sites. Even a fully deactivated
zeolite retains some activity for isomerizing butnes (Chu and Chester, 1986).
Correlations between the acidity and the alkylation performance revealed that
the acid strength required decreases in the order: cracking > alkylation
(addition of butne to a tertiary butyl) > dimerization (addition of a butne to a
secondary butyl) (Corma et a l , 1994b).

27

3. Carbophilicity of catalyst surface is desired. An important reason of


deactivation of isobutane/1-butne alkylation on solid catalysts is the
preferential adsorption of alkenes. Although the concentration of alkenes in
liquid phase might be low, they will preferentially adsorb in the pores. The
concentration of alkenes in the pore system is much higher resulting in much
higher relative rates of oligomerization vs. hydride transfer. The different
hydrophilicity/carbophilicity of walls will make the materials have very
different adsorption behaviours. The carbophilic walls should allow a higher
paraffin steady concentration in the vicinity of the action site.
4. Large internal pore size is required. Only large-pore zeolites exhibit sufficient
activity and selectivity for the alkylation reaction. Large pore size can facilitate
the diffusion of reactant to the active site and diffusion of the products out of
the pore network.
It is recommended (Albright, 2009) that refineries considering alkylation units
employing solid catalysts should seek details concerning the following factors: (1) time
required before catalyst needs to be regenerated; (2) qualities and yields of alkylate as a
function of time while the catalyst is being used and hence deactivated; (3) catalyst
information including composition of the catalyst plus possible precious metals, range of
pore size and particle size, and expected longevity of the catalyst; (4) operating
conditions for alkylation with different olefins; (5) predicted quality and yield of alkylate
obtained for various operating conditions and for different olefins; and (6) recommended
operating conditions for reactivation of the catalyst. For example, does reactivation affect
the longevity of the catalyst such as loss of precious metals or size of particles? What
special safety precautions are needed during regeneration? Both hydrogen and oxygen
have safety concerns.

28

2.3. Solid acid alkylation catalysts


2.3.1. Zeolites
In literature, zeolites got much attention as alternative catalysts for alkylation of
isobutane with n-butene. The success of zeolites application in catalytic cracking created
interest in applying them in alkylation process. Large pore zeolites such as H-FAU, HBEA and H-EMT have been tested as alkylation catalysts. However, despite the potential
benefit of using solid acid catalysts, zeolite did not achieve commercialization up to now,
due to their unacceptable fast decay. Understanding the characteristics and behavior of
zeolites during the alkylation reaction is beneficial for developing better solid acid
catalysts.
Compared with the liquid acids, zeolites possess lower proton (acid site)
concentration. 1 g of H2SO4 contains 20* 10"3 moles of protons, whereas 1 g of zeolite
HY (with Si/Al ratio of 5) contains no more than 3*10"3 moles of protons. Furthermore, 1
g of H2SO4 occupies far less volume (i.e., 0.5 cm ) than the equivalent mass of zeolites
(4-6 cm3). Furthermore, the acidic protons on zeolites are not totally accessible as those
in H2SO4.
Liquid acids like H2SO4 and HF have a well-defined acid strength. This is not the
case for zeolites. Acidic sites on zeolites differ substantially in their nature and strength,
depending on the zeolite type, its aluminum content and the ion exchange procedure.
Bronsted and Lewis acid sites with a wide range of strength and concentration are present.
Like mentioned above, isobutane/butene alkylation comprises a set of parallel and
consecutive reaction steps, such as isomerization, oligomerization, ^-scission and hydride
transfer. The different reaction steps require different minimum acid strengths. Double
bond isomerization can be catalyzed by weak acid sites. Even a fully deactivated zeolite
retains

some

activity

for

isomerizing

butnes

(Chu

and

Chester,

1986).

Dimerization/oligomerization does not either require strong acidity. This was concluded
from a study of a series of USY zeolites with different unit cell sizes. Hydride transfer is
only catalyzed by strong acid sites (Corma et al., 1994b).
29

In the reaction steps involved in alkylation, hydride transfer is the crucial step that
determines the product quality and the catalyst lifetime. Thus, it is really important to
know which factors favor hydride transfer. Nivarthy et al. found that the lifetime of
zeolite H-BEA depends on the concentration of Bronsted acid sites (Nivarthy et al,
1998b). In later studies various researchers found that strong Bronsted acid sites are
required for hydride transfer (Feller et a l , 2004; Stocker et a l , 1994; Corma et a l , 1994a;
Rorvik et a l , 1997). Stocker et al. synthesized and tested EMT and FAU samples with
enhanced Si/Ai ratios of 3.5. H-EMT with the highest strong/weak Bronsted acid sites
concentration shows better activity. Dealumination of the H-FAU led to better results
because it generates additional small numbers of very strong acid sites (Stocker et a l ,
1994). Rorvik et al found that La-exchange of H-EMT can increase the strong/weak
Bronsted acid sites ratio, which leads to a better alkylation performance (Rorvik et al,
1997). Corma et al. compared USY, MOR, BEA, ZSM-5 and MCM-22 (MCM is the
acronym of Mobil crystalline of materials). They proposed that the relative decrease of
activity for the formation of TMPs during time on stream was depending on the
concentration of strong Bronsted acid sites in the fresh zeolite (Corma et a i , 1994a). The
ratio of strong Bronsted acid sites to total Bronsted acid sites was proposed as key
property of the catalyst. Weak Bronsted acid sites are not capable of catalyzing hydride
transfer, but some of them are sufficiently strong to catalyze the polymerization of butne
leading to deactivation of the catalyst (Sievers et a l , 2007; Mostad et a l , 1996). Thus,
the stronger sites are necessary to effectively catalyze hydride transfer, the strength of the
Bronsted acid sites determines the conversion of the carbenium ion like species on the
catalyst surface.
The lifetime of zeolitic alkylation catalysts also depends on the concentration of
Bronsted acid sites. Sievers et al. (Sievers et a l , 2008) compared the catalytic
performance of LaX (Si/Al =1.1) and LaY (Si/Al = 2.4) with the same pore structure.
LaX had a much higher concentration of strong Bronsted acid sites. In the reaction over
LaX, the catalyst lifetime was twice as high as in the reaction over LaY. Nivarthy et al
compared a series of zeolite H-BEA with different concentrations of Bronsted centers.
Low concentration of Bronsted acid sites led to a concomitant decrease in the measured
30

catalyst lifetime during alkylation (Nivarthy et a l , 1998b). Generally, with time on


stream, the sites strong enough for catalyzing hydride transfer deactivate first.
The influence of the Si/Al ratio on the catalytic performance of zeolites in
isobutane/-butene alkylation has been the subject of various studies (Weitkamp and Traa,
1999; Yoo and Smimiotis, 2002; de Jong et a l , 1996; Corma et a l , 1994b). In principle,
the number of potential acid sites is given in a zeolite by the number of framework
aluminums. With increasing aluminum concentration in the framework (i.e., with lower
Si/Al ratio), the total concentration of acid sites increases. Though it is believed that the
strength of the acid sites may decrease with increasing aluminum concentration, the
zeolites with lower Si/Al ratio possess more high strength acid sites. At the same time,
high Al concentration facilitates the formation of extraframework species (Scherzer,
1989). Corma and Martinez reviewed the influence of zeolite acidity on isobutane/butene alkylation reaction (Corma and Martinez, 1993). They pointed out that a higher
density of framework Al would be good for both number of cation precursors and high
hydrogen transfer, if the density of framework Al was diminished, hydrogen transfer
decreased, and both conversion and selectivity would decrease. In faujasites, a low Si/Al
ratio provides a long catalyst lifetime. This finding is supported by reports on increasing
hydride transfer activity with decreasing Si/Al ratio for faujasites used as cracking
catalysts (Cumming et a l , 1996).
Unlike liquid acids, zeolites possess both Lewis and Bronsted acidity. Most
researchers believe that the activity of zeolite catalysts in the alkylation reaction is due to
the number and activity of the Bronsted acid sites. Lewis acid sites do not catalyze the
alkylation reaction, but influence the alkylation performance of zeolitic catalysts. DiazMendoza et al claimed that the presence of strong Lewis acid sites promotes the
formation of unsaturated compounds (Diaz-Mendoza et a l , 1998). Flego et al. found
Lewis-acidity favored production of unsaturated carbenium ions. They studied the
deactivation of a La-H-FAU zeolite in isobutane/1-butne alkylation and found that
higher catalyst activation temperatures led to higher Lewis acid site concentrations,
which increased the formation of mono- and dienylic carbenium ions (Flego et a l , 1995).
Besides the ability to increase the rate of formation of unsaturated compounds, Lewis
31

acid sites enhance the local butne concentration in the vicinity of the Bronsted acid sites,
the increased alkene concentration accelerates oligomerization and leads to deactivation
(Nivarthy et a l , 1998b).
Numerous transfer steps obviously occur in the pores of the solid catalysts. Both
isobutane and the feed olefins diffuse inward in the pores, and product molecules diffuse
outward (Albright, 2009). Sherwood et al. outlined several problems that occur in the
porous catalyst (Sherwood et a l , 1975). The rates of transfer or diffusion vary for
different molecules because of differences in size, shape, and molecular weight. In the
pores, numerous reaction steps occur. Conditions in the pores are conductive to forming,
in addition to alkylate, both conjunct polymers and pseudo-alkylate. Diffusion of
conjunct polymer in particular would be slow and probably nonexistent. The conjunct
polymers would likely complex or bond with the inner surfaces of the solid catalyst.
Large internal pore sizes are necessary to facilitate the diffusion of the reactants to the
active sites as well as effectively promote diffusion of the products out of the catalyst
pores. Only large pore zeolites exhibit sufficient activity and selectivity for the alkylation
reaction, as diffusion limitations in medium and small pore zeolites lead to premature
deactivation (Corma et a l , 1994a; Chu and Chester, 1986). It was found that ZSM-5,
which has medium pore size, is inactive under typical alkylation conditions (Chu and
Chester, 1986). The authors attribute this to diffusion limitations in the pores. Corma et al.
studied H-ZSM-5 and H-MCM-22 samples at 50 C. The ZSM-5 exhibited a very low
activity and a rapid deactivation. The products were dominated by dimethylhexanes and
dimethylhexenes. The authors claim that alkylation takes place mainly at the external
surface of the zeolite particles, while dimerization, which is less sterically demanding,
proceeds in the pore system. MCM-22 is a zeolite with larger pore void space than ZSM5, which exhibited a performance in between large and medium pore size zeolites (Corma
et a l , 1994a). MCM-36 is a pillared zeolite based on the structure of MCM-22. It
contains mesopores between layers of MCM-22 crystallites. This structure exhibited
higher activity and stability compared with MCM-22 (He et a l , 1998).
Stocker et al. compared FAU and EMT zeolites in their H- and La-exchanged form
with and without dealumination. EMT was always superior to FAU. It also produced a
32

higher amount of trimethylpentanes compared with the FAU samples. The differences
between the two materials were attributed to the slightly larger supercage in EMT, which
is claimed to reduce the steric constraints on the bulky transition states for hydride
transfer (Stocker et a l , 1994; Stocker et a l , 1996; Rorvik et a l , 1996).
It was shown that zeolites with three dimensional pore networks allow more
efficient diffusion of the products, which leads to an increase of the catalyst lifetime (Yoo
et a l , 2001).
Beside the difference in their nature and strength of acid, another major difference
between acidic zeolites and the liquid acids is their selective and strong chemisorption of
unsaturated compounds. Due to the high polarity of the zeolitic surface, especially in
aluminum rich zeolites, polar molecules will be preferentially adsorbed. Furthermore, the
electrostatic field in the zeolite pores enhances the adsorption of polarizable molecules
(Ramachandran et a l , 1996). Although the concentration of alkenes in the liquid phase is
low, their preferential adsorption in the zeolite surface and pores makes alkenes
concentrate on the external and internal surface. Thus the real alkanes/alkenes ratio on
catalysts is much lower than in the hydrocarbon phase, which results in much higher
relative rates of oligomerization vs. hydride transfer.
The heat of adsorption of hydrocarbons on zeolites increases with increasing chain
length (Eder and Lercher, 1997). Each C-atom contributes equally to the total heat of
adsorption. Because the heat of adsorption is essentially the energy of activation, the
desorption of a C12 molecule is about four orders of magnitude slower than a Cg molecule,
a Ci6 is eight and a C20 is twelve orders of magnitude slower. So once the heavy products
were produced and adsorbed, it is very difficult for them to desorb from the zeolites
surface. That is the major reason why zeolites deactivate rapidly.
The influence of the ion exchange with rare earth cations on the catalytic
performance of zeolites in isobutane/-butene alkylation has been the subject of various
studies (Flego et a l , 1995; Feller et a l , 2004; Josl et a l , 2004; Klingmann et a l , 2005;
Guzman et a l , 2005). At temperatures above 333 K lanthanum cations start to migrate
into the sodalite cages. It has been suggested that the incorporation of La3+ cations leads
33

to an increase in the strength of Bronsted acid sites via polarization of the zeolite
framework (van Bokhoven et a l , 2000). Rare earth cations are involved in the formation
of Bronsted acid sites via hydrolysis. C heetham et al. found that in calcined LaY every
La3+ cation has an affiliated OH group (Cheetham et a l , 1984).
[La(H20)]3+ + H 2 0

[LaOH(H20).,]2+ + H 3 0 +

In lanthanum exchanged faujasites, a high concentration of strong Bronsted acid sites can
be achieved. A high ion exchange degree was essential for a good catalyst lifetime in
isobutane/-butene alkylation (Sievers et a l , 2008).

2.3.2. Mounted liquid acids


Supported solid acids have been studied as candidate alkylation catalysts. Triflic
acid and perfluoroalkanedisulphonic acids were supported on a porous silica carrier
(Clerici et a l , 1993; Clerici et a l , 1995; de Angelis et a l , 2001; Ingallina et a l , 2002).
The authors claimed that supported acid catalysts were able to convert butne completely
with high selectivity to trimethylpentanes. The supported acids are stable to elution when
the catalyst activity is high. Only at the end of the reaction a partial loss of the active
phase is observed. They emphasized there is a strong interaction between the acids and
the support, which prevents leaching of the acid. In pulsed liquid phase (alternated
feeding mode) isobutane/1 -butne alkylation experiments at 298 K, the catalysts
produced a very high quality alkylate.
Chlorinated alumina, obtained by reacting alumina with hydrogen chloride, is a
highly Bronsted-acidic and porous solid. C let et al prepared a series of chlorinated
alumina catalysts modified with Li+ and Na+ cations. These catalysts gave products
resembling an alkylate but with more dimethylhexanes, light- and heavy-end products
(Clet et a l , 2000).
Fluorinated alumina materials were prepared by Moreno et al. to catalyze
alkylation of -butne with isobutane (Moreno et a l , 2003). The addition of fluorine on
alumina leads to highly selective and active catalysts for butne alkylation. They claimed
34

only Bronsted acid sites appear to be involved in the catalytic process. No simple
relationship between the fluorine content and the activity was found. They explained that
above 6.0 wt% of fluorine, a large amount of AIF3 is formed, which leads to inactive
catalysts for butne alkylation.

2.3.3. Heteropolyacids
Heteropolyacids are strongly acidic nonporous solids. The surface area is low.
Supporting materials on highly porous carriers is a way to increase the surface area. This
was done by Blasco et a l , who supported 12-tungstophosphoric acid on various supports
such as silica, aluminosilicate, MCM-41 with different pore size. They found all the
supported heteropolyacid catalysts are active and selective for carrying out the alkylation
of butne with isobutane at relatively low reaction temperatures. A higher alkylation
activity is observed when amorphous silica is used as the support as compared to the
mesoporous aluminosilicate and all-silica MCM-41 materials. A maximum in activity,
TMP's selectivity and catalyst stability is obtained for the silica-based catalyst with 40
wt% acid loading. A partial blockage of the MCM-41 occurs, decreasing the accessibility
of the acid sites to the reactants. The pore blockage is more significant as the amount of
heteropolyacid in the catalyst increases and can be minimized by increasing the mean
pore diameter of the mesoporous molecular sieve. Anyway, all these materials
deactivated rapidly (Blasco et a l , 1998). Heteropolyacids have also been supported on
sulfated zirconia catalysts. The combination was found to be superior to heteropolyacid
supported on pure zirconia and other supports treated with a variety of mineral acids (de
Angelis et a l , 1999). Solutions of heteropolyacids (containing phosphorous or silicon) in
acetic acid have been tested as alkylation catalysts at 50 C by Zhao et al. The system
was sensitive towards the ratio of heteropoly acid/acetic acid and the amount of crystal
water. Similar to conventional liquid acids, on which a polymer designated as red oil is
formed, a polymeric material was obtained, which enhanced the catalytic activity (Zhao
et a l , 2000).

35

2.3.4. Acidic organic polymers


Perfluorinated sulfonic acid resins, such as Nafion, are strongly acidic solid with
acidity comparable to sulfuric acid. The surface area of these materials is very low (about
0.2 m2/g for Nafion). To increase the insufficiently small surface area, supporting the
resin on porous carriers, or incorporating it into silica through a sol-gel method provides
ways to increase the surface area. Both methods have been utilized by Botella et al.
(Botella et a l , 1999), who compared several composite nafion/silica samples with
varying surface area and Nafion loading in iso-butane/2-butene alkylation. The
unsupported resin performed poorly. The composite materials with intermediate surface
areas showed the best performance. The authors explained that an interaction between the
sulfonic groups of the resin with the silanol groups of the silica decreases the acid
strength of the resin. The supported resin showed similar activity and selectivity to the
composite material of same Nafion content. Experiments at different temperature of
32 C to 80 C showed that the material produces oligomers at low temperatures and
saturated products at high temperature.
Kumar et al. studied the conversion of a mixture of pure isobutane and Raffinate II
over Nafion/silica nanocomposites (Kumar et a l , 2006). Raffinate II is the remaining C4
cut of the steam cracker effluent

after removal of butadienes/isobutene and

propane/propene. They claimed that an excess of isobutane (molar ratio i-C^IC^ = 20 and
110) in the reaction mixture resulted in better activity. At temperatures around 70-80 C,
the selectivity toward isooctane had a maximum of up to 62%. Using a step-up procedure
under

batch

conditions,

the

selectivities

of

trimethylpentanes

(TMPs)

and

dimethylhexanes (DMHs) were found to be dependent on the isobutane concentration in


the reaction mixture. The composites are directly recyclable; only low catalyst
deactivation occurs over several regenerations. They also regenerated catalysts by
refluxing them with acetone and then further treatment with aqueous hydrogen peroxide
(15 wt%) at 80 C for 2 h. After the first and second regeneration the results were found
comparable with the once of the fresh catalyst.

36

2.3.5. Sulfated metal oxides


Sulfate-promoted oxides are widely known as solid acid catalysts for hydrocarbon
transformations (Song and Sayari, 1996). Sulfated zirconia and related materials have
attracted a great deal of attention from researchers because it has the highest acid strength
and activity. Corma et al. studied the acitivity, selectivity and deactivation behavior of
sulfated Z1O2, Ti2 and Sn02 in isobutane/2-butne alkylation (Corma et a l , 1996). The
initial activity decreased in the order: S0427Zr02 > S0427Ti02 > S0427Sn02, which was
the same trend reported for acid strength measured by Hammett indicators. Owing to the
higher strength of the acid sites, sulfated Z1O2, Ti2 presented a very high initial cracking
activity, thus decreasing the selectivities of the desired TMPs. A certain loss of sulfates
(0-30% of the initial sulfates) took place during the alkylation reaction. All the sulfated
materials suffered a rapid deactivation with time on stream.
Sulfated alumina possesses lower acidity in comparison with most other sulfated
oxides. Sulfated aluminas with different sulfate contents were prepared and tested in
isobutane/butene alkylation reaction (Smirnova et. al., 2008). It was found that sulfated
aluminas in contrast with SZ are active in the presence of significant amounts of sulfate
containing phases. The catalytic properties of sulfated aluminas in isobutane/butene
alkylation depend slightly on the sulfate precursor and strongly on its concentration. The
catalytic activity appeared when the sulfate content exceeded half of monolayer
concentration and reached a maximum for the sample, whose sulfate concentration
approached the monolayer. The yield of C5+ products as well as C5-C7 products
selectivity increase as the sulfate content increases up to 23 wt% and remain practically
constant with further increase in sulfate content. Increasing the sulfate content up to 14
wt% led to an increase in activity. A reduction in activity was observed when the sulfate
content increased further.
Sulfated zirconia and zeolite BEA was compared in isobutane/2-butne alkylation
(Corma et a l , 1994e). BEA catalyzed mainly dimerization at 0 C, while the sulfated
zirconia exhibited a high selectivity to TMPs. At 50 C zeolite BEA produced more
TMPs than sulfated zirconia, while sulfated zirconia produced mainly cracked products
37

with 65 wt% selectivity. The TMP/DMH ratio was always higher for the sulfated zirconia
sample. The authors attributed it to the much stronger acid sites in sulfated zirconia as
compared to zeolite BEA. The strong sites preferentially catalyzed cracking reactions,
and allowed hydride transfer at lower temperatures than the zeolitic acid sites. The timeon-stream behavior was more favorable for BEA, which deactivated at a slower rate than
sulfated zirconia.

2.4. Effect of process parameters on alkylation


2.4.1. Reaction temperature
The reaction temperature when using H2SO4 as catalyst is around 10 C. HF is
operated at 10-40 C. In principle, solid acids operate at significantly higher reaction
temperatures. This is due to the lower acid strength of solid acids or the lack of solvation
which makes activation energies of individual reaction steps higher. Higher temperatures
favor efficient mobility in the pores of solid catalysts. The optimum temperature for
zeolites is in the range of 50 to 100 C. Nivarthy et al. found a temperature optimum for
zeolite H-BEA at 75 C, at which the highest octane selectivity and the highest
TMP/DMH ratio was achieved. At lower temperatures oligomerization and at higher
temperatures cracking reactions dominated (Nivarthy et a l , 1998a). Kirsch et al. studied
the relationship between acidity and optimum temperature on rare earth exchanged Y
zeolites. A sample with 0.2 wt% residual sodium had a temperature optimum around
40 C, while a sample with 1.0 wt% sodium performed best at 80 C. That means lower
acid strength requires higher reaction temperature (Kirsch et a l , 1972). Corma et al.
studied effects on the product selectivities by using a H-BEA at 50 and at 80 C. At the
higher temperature, the activity was higher, as seen in the increased conversion. The
selectivity to cracked products increased drastically, also the Cg+ selectivity increased.
Within the TMP-fraction, 2,2,4-TMP increased considerably with temperature (Corma et
a l , 1994d). The selectivity phenomena can be explained by the relative rates of the
individual reaction steps. ^-Scission and presumably also alkene addition require higher
activation energies than hydride transfer. An increase in temperature consequently leads
to higher relative rates of secondary products from multiple alkylation and cracking.
38

Cracked products are favored over multiple alkylation products, because the activation
energy is higher for ^-scission than for alkene addition, which is the (exothermic) reverse
reaction. The authors explained the bad performance of zeolites at low reaction
temperatures by the hindered diffusion of bulky molecules under such conditions. The
catalyst will prematurely deactivate by pore blocking. The temperature influence on the
Nafion/silica composites catalyzed isobutane alkylation was studied by Kumar et al.
(Kumar et a l , 2006). The conversion of butne increases with temperature. At around 7580 C, the Cg fraction dominates. At somewhat higher temperatures, the fraction of the
C5-C7 products increases, while at lower temperatures, the C9-C12 products become
favored.

2.4.2. Paraffln/olefin ratio and olefin space velocity


The temperature influences the reaction rates via the activation energies; the feed
composition does through the concentration term of the rate expressions. The crucial
parameter of the paraffin/olefin alkylation on solid acid catalysts is the ratio of the
hydride transfer and oligomerization rates. This ratio should be as high as possible. With
high isobutane concentrations the carbenium ion has a higher probability to react with an
iso-butane molecule to form the desired product via hydride transfer, oligomerization is
restrained, undesired reactions and catalyst deactivation are minimized. The feed
paraffin/olefin (P/O) ratio and the olefin space velocity (OSV) are two process
parameters that influence the hydride transfer/oligomerization ratio. The P/O ratio
determines the concentration of iso-butane in the reactor, which in turn determines the
rate of hydride transfer. It should be pointed out that the preferential adsorption of olefin
in the surface of solid acids makes the olefin concentration on catalytic surface
considerably higher than in the liquid phase. This results in much higher relative rates of
oligomerization vs. hydride transfer. With high feed P/O ratios this problem will be
minimized. Thus, increasing the P/O ratio increases catalytic performance. On the other
hand, at high P/O ratios more iso-butane has to be recycled, which leads to increased
separation costs. A balance has to be found in order to optimize the economical
performance of the unit.

39

The OSV determines the production rate of alkylate, so that high OSV would be
economically favored. This is limited by the capability of catalysts. Higher OSV results
in lower conversion of olefin, lower quality of alkylate and shorter lifetime of catalyst.
Also a balance should be found to optimize the economical performance of the unit.
The alkylation mechanism, the influence of the catalyst type and the reaction
conditions were discussed above. Compared with the liquid acids, solid acid catalysts,
which are more enviroment benign, suffer from rapid deactivation. Some new solid acids
with stronger acidity, larger pore size and different adsorption behavior may minimize
the drawbacks of the already tested catalysts.

2.5. New solid acid catalysts-acid functionalized mesostructured


materials
Above mentioned solid acids are the most widely used solid acids in
petrochemistry and gas phase reactions. The small pore size imposes a severe restriction
for the accessibility of larger molecules to the internal surface of the catalyst.
Mesoporous molecular sieves have opened a new era in the' development of solid
catalysts (Trong On et a l , 2003; Taguchi and Schuth, 2005). These materials have
relatively uniform pore sizes and high void volumes and specific surface areas. The pore
sizes of these materials can be tailored depending on the synthesis method. Many kinds
of catalytically active functional groups can be anchored on the pore wall. For example,
mesoporous silicas can be employed as supports to which strong acid sites can be
covalently anchored. Sulfonic acids are organic compounds that exhibit an acid strength
comparable to those of sulfuric acid and benzenesulfonic acid.
Basically, two methods have been developed to covalently anchor functional
groups to the walls of mesostructured materials: Post-synthesis procedure (grafting
methods) and direct synthesis (co-condensation). Generally, the direct method yields a
uniform surface coverage of functional groups and gives good control over the amount of
incorporated functional groups. Covalent anchoring on silica surfaces is based on the
presence of silanol groups. There are three different types of surface silanols, that is,
40

single, hydrogen-bonded, and geminal hydroxyl groups. Only free ( = Si-OH) and
geminal (=Si(OH)2) silanol groups take part in silylation reactions, and hydrogen-bonded
silanol groups are less accessible to modification because they form hydrophilic networks.
The specific density of the silanol groups per gram of silica depends on many factors
including surface area, procedure of silica formation, incomplete condensation of
neighbouring silanols and thermal pretreatment of the silica at high temperatures. Even
though silanol groups can also be in the interior of the silica particles, silanols together
with silanoxy groups (=Si-0-Si=bridges) are the natural terminal groups of the silicas
at the solid surface. Since the silanol groups are located at the surface, the population of
silanols increases along with the total surface area. Also, silica samples obtained by solgel synthesis have a considerably higher population of silanol groups, due to incomplete
condensation of the molecular precursors, than pyrogenic silicas produced at high
temperatures. Treating silicas at temperature about 900 C produces the condensation of
two neighboring silanol groups to form a silanoxy bridge. Ideally, the maximum
achievable loading of catalytic sites on the solid is desired and, therefore, a large
population of silanol groups is convenient to effect the anchoring of a sufficiently high
amount of catalyst. On the other hand, after having performed the covalent anchoring, the
presence of residual silanol groups may sometimes be negative and detrimental for the
catalytic performance of the solid, and it could be even necessary to mask these free
silanol groups.
The most widely used methodology to anchor on silica surfaces is the reaction of
surface silanol groups with silylating reagents. Badley and Ford anchored sulfonic groups
on the walls of amorphous silica in 1989 (Badley and Ford, 1989). Following their
pioneering work, several mesostructured materials anchored with sulfonic groups on the
pore walls were synthesized in 1998 (Van Rhijn et a l , 1998a; Van Rhijn et a l , 1998b;
Lim et a i , 1998). These sulfonated mesostructured materials are of high surface area,
narrow pore size distribution and high acid loading. Moreover, the sulfonic-acid groups
covalently attached to the walls of mesostructured silica are supposed to be highly stable
enough.

41

2.5.1. Post-synthesis preparation of organosulfonic-functionalized


mesostructured materials
In 1998 Van Rhijn et al. published synthesis strategy for the preparation of
alkanesulfonic-acid-modified mesostructured materials (Van Rhijn et a l , 1998a). MCM
and HMS (acronym of hexagonal mesoporous silica) samples were first functionalized
with

propane-thiol

groups

mercaptopropyltrimethoxysilane

by

reaction

(MPTMS).

of

the

Compared

surface
with

silanols
amorphous

with

3-

silicas,

organosiloxane loadings of mesostructured materials are much higher. It may be because


of the higher surface area and ordered pore structure of mesostructured materials. In order
to get sulfonic groups, thiol groups must be postsynthetically treated with a large excess
of an oxidant (mainly hydrogen peroxide (Van Rhijn et a l , 1998a) or nitric acid (Van
Rhijn et a l , 1998b). The strategy of synthesis is schematized in Figure 4. Some
optimizations were made by several authors. Van Rhijn et al. developed a modified
grafting procedure to enhanch the incorporation of sulfonic groups (Van Rhijn et al.
1998b). Boveri et al. improved the grafting techniques by optimizing extraction condition
(Boveriefr/.,2005).
The presence of electron-withdrawing species close to the sulfonic group is
expected to increase the acid strength of the acid sites. Anchoring arenesulfonic groups
on mesoporous materials will yield strong solid acids. Lindlar et al. synthesized
arenesulfonic-modified MCM-41 materials with pore diameters up to 60 (Lindlar et a l ,
2001). The synthesis was based on grafting of phenyl groups to the silica surface and
subsequent sulfonation with chlorosulfonic acid.
This post-synthesis synthesis strategy has serious drawbacks. Materials prepared in
this way have yielded XRD patterns with lower scattering intensities that indicate
relatively poor long-range ordering in comparison to the starting material. Moreover, a
decrease in the surface area and pore volume is clearly observed and reduces the potential
application of these catalysts. As for the strategy shown in Figure 4, despite the large
excess of oxidant used in the process, in most cases the postoxidation step does not allow
quantitative reaction of thiol groups, and in some cases, leaching of sulfur species is
42

clearly evidenced. The presence of unoxidized sulfur species might have a negative effect
on the catalytic performance of these materials.

Figure 4. Post-synthesis strategy for the preparation of sulfonic-acid-modified


mesostructured materials

2.5.2. Direct-synthesis of organosulfonic-functionalized mesostructured


materials
The post-synthesis methods require a procedure of post oxidation. These
treatments lead to a loss of mesoscopic order and, in general, to damage of the textural
properties and an incomplete oxidation of the thiol groups accompanied by a loss of
sulfur species. Direct synthesis involving co-condensation of siloxane and organosiloxane
species in the presence of different templating surfactants has been shown to be a
promising alternative to the grafting procedures (Huo et a l , 1996; Burkett et a l , 1996;
Macquarrie, 1996; Fowler et a l , 1997; Lim et a i , 1997; Fowler et a l , 1998; Richer and
Mercier, 1998; Lim and Stein, 1999; Babonneau e t a l , 1999; Brown e t a l , 1999; Mercier
43

and Pinnavaia, 2000). A one-step direct synthesis procedure to create periodic ordered
alkanesulfonicfunctionalized msostructurs was developed in 2000 by G.D. Stucky's
group (Margolese et a l , 2000). This new procedure involves a one-step synthetic strategy
based on the co-condensation of tetraethoxysilane (TEOS) and 3-mercaptopropyltrimethoxysilane (MPTMS), in the presence of Pluronic 123 species and H2O2 in HC1
aqueous solutions. The one-step synthesis using acid catalysis in the presence of Pluronic
123 produced SBA-15-modified materials with greater oxidation efficiency (100% vs.
25-77% in postsynthetic treatments) with larger (up to 70 ) and more uniform pores,
higher surface areas (700-800 m2/g), and good long-range order in contrast to
postoxidative methods. Moreover, this outcome was achieved under milder and simpler
synthetic conditions utilizing less time and less material (in-situ oxidation with a
stoichiometric amount of oxidant and acid exchange). The net result was a sulfonic
mesostructured material with acid capacities several times greater than those achieved
with postoxidative methods and thermal stability up to 450 C in air. Finally, one of the
most interesting features of the in-situ syntheses was the observed enhancement of the
long-range order found in the mesoporous products. In-situ formation of sulfonic-acid
moieties apparently introduces variations that promote microphase separation of the
PEO-PPO-PEO copolymer blocks and thus increasing mesoscopic ordering.
Later (Melero et a l , 2002), the same route was used to prepare ordered SBA-15
materials containing arenesulfonic-acid groups, except that MPTMS was substituted with
2-(4-chlorosulfonylphenyl)-ethyltrimethoxy silane (CSPTMS)). Hydrolysis of the chlorosulfonyl groups (-SO2CI) to the corresponding sulfonic-acid groups is achieved under
acidic condensation conditions. Direct synthesis procedure allowed the effective
anchoring of arenesulfonic groups on the pore surface of SBA-15 mesostructured
materials. The resultant materials showed large uniform pore sizes (ca. 60 ) with large
surface areas (ca. 650 m2/g), good mesoscopic ordering, and thick walls, leading to
hydrothermally stable materials. The acid centers displayed excellent thermal resistance
and significant stability against leaching in aqueous and organic medium. No disulfide
species were detected, as confirmed by

13

C CP MAS NMR. Solid acidity of the

synthesized arenesulfonic-acid material was characterized by means of

31

P NMR of

chemisorbed triethylphosphine oxide and compared with other acid solids. The
44

arenesulfonic SBA-15 material presented 31P NMR signal shifts toward higher values in
comparison to those obtained for alkylsulfonic acid sample, confirming enhancement of
the acid strength due to the presence of an electron-withdrawing substituent adjacent to
the sulfonic-acid groups, such as the benzene ring.

2.5.3. Tuning the hydrophobicity of pore surface


The hydrophilic/hydrophobic character of the catalyst's surface has important
effects on the adsorption and diffusion of reactants and products within the mesopores.
Moreover, most solid acids are poisoned by water in reactions where this highly polar
molecule is involved as reactant, product, or even solvent. Thus, the hydrophobization of
the acid sites microenvironment is an important challenge to reduce poisoning by water
molecules. It can be expected that the adjustment of sulfonic-modified catalyst
hydrophobicity will probably have strong implications in the catalytic performance. As
for isobutane/1-butne alkylation, an important reason of deactivation is the preferential
adsorption and pore filling with the olefin. The different hydrophilicity/carbophilicity of
walls makes the materials have very different adsorption behaviours. The carbophilicity
of the organic moiety of its walls should allow a higher paraffin steady concentration in
the reaction conditions. The importance of control over surface chemistry is therefore of
great importance. A new class of materials which allows a combination of the
mesoporous inorganic framework of the MCM and SBA materials with an essentially
organic surface is thus of great interest. Periodic mesoporous organosilicas (PMOs) are
materials which utilise silane precursors containing at least two silane groups (Hatton et
a l , 2005; Hunks and Ozin, 2005). The first papers in 1999 (Asefa et a i , 1999; Melde et
al., 1999; Inagaki et a l , 1999) sparked enormous interest, particularly in the materials
community, where the ability to produce materials with the structural control of
templated silicas, coupled with organic, hydrophobic surfaces has led to many new
opportunities. While the oligo-silanes are often less readily available than the simpler
mono-silanes, the enhanced material properties (including much higher stability towards
water) make application of these materials very attractive. Inagaki has demonstrated that
phenylene bis-silanes can be condensed around a template to give materials which show a
high degree of order not only of the pore system, but in the walls, where phenylene units
45

stack up to give crystallinity (Inagaki et a l , 2002). They incorporated mercaptopropyl


-silane into a phenylene, followed by oxidation, to get sulfonic acid functionalized
material. The material showed good activity for esterification.
Kaliaguine's group reported the synthesis of alkylsulfonic-acid-bearing ethanesilica mesostructured materials (Hamoudi and Kaliaguine, 2003). Different loadings of
functional groups (up to 32 mol%) were successfully incorporated into the hybrid
mesostructured material. Acid capacity was up to 0.7 mmol H+ per gram. Interestingly,
the acidity of the sulfonic-acid sites was investigated by FTIR using pyridine as the probe
molecule. The spectrum of the solids after pyridine adsorption and degassing at different
temperatures showed absorbance bands at 1490, 1548, and 1640 cm"1, attributed to
interaction of the basic molecule with Bronsted-acid sites. These bands were still retained
after evacuation at 100 C, which indicates the existence of relatively strong acid sites.
Later, they proposed a new member in PMO family (Hamoudi et a l , 2004). Its walls are
constituted of ethane silica grafted with ethyl benzene sulfonic acid groups, which
protrude in the pore channel. Its designation is arene sulfonic mesostructured ethane
silica (ArSMES). This is the first report on the synthesis of mesoporous ethane - silica
materials functionalized with arene-sulfonic acid groups. Two different synthesis
procedures involving the use of Brij-56 or PI23 surfactants under mildly acidic
conditions yielded good quality materials with acid capacity (up to 1.38 meq/g) and
y

proton conductivity (up to 1.6><10" S/cm), higher than those obtained for the propyl
-sulfonic acid attached analogues. Both arenesulfonic groups and organics in the walls of
the arene-sulfonic acid functionalized materials exhibited high thermal stability (up to
350 C).
Trimethylsilyl groups (or similar) are often grafted onto preformed catalysts to
modify the hydrophilic/hydrophobic character of the catalyst's surface, thus to modify
the surface adsorption properties. Incorporation of a second organosilane (spectator
groups) in the synthesis of the material is a method we can apply. These non-catalytic
active groups - which modify catalytic activity, presumably by changing the surface
polarity and hence sorption behavior - have been shown to enhance catalytic activity
considerably. D. J. Macquarrie et al. introduced propyl groups as spectator groups to the
46

wall of perfluorosulfonic acid grafted mesostructured material and enhanced the rate of a
Friedel-Crafts acylation dramatically (Macquarrie et a l , 2005). Figure 5 gives a
comparison of the catalysts with and without spectator groups. It appears that the propyl
groups have a positive effect on the structure of the material.

,OMe

spectator group

catalyst

pucoci

- catalyst with spectator groups


,OMe

10

20
30
time (mine)

40

S^

COPh

Figure 5. The effect of spectator groups on activity in a Friedel - Crafts acylation.


Diaz et al. (Diaz et a l , 2000a; Diaz et a l , 2000b) have also demonstrated the
efficacy of methyl groups on the surface of the material in the acid-catalysed
esterification of glycerol with fatty acids. They showed that 2 mmol g _l of methyl groups
trebled the activity of a sulfonic acid founctionalized material. Further increases in the
loading of methyl groups had little effect on rate or conversion, presumably because
optimum surface polarity had been achieved. A titanium-containing ethane bridged
hybrid mesoporous silsesquioxane has been shown to be active and with good selectivity
in the ammoxidation of bulky ketones (Bhawsik et a l , 2003). The hydrophobicity of the
materials is important in determining activity and can be controlled by the nature of the
organic group in the silsesquioxane.
Ryoo's

group

developed

a new

process

to

synthesize

functionalized

mesostructured polymer-silica composite materials (Figure 6) (Choi et a l , 2005). Vinyl


monomer, such as styrene was polymerized after impregnation into mesoporous silicas
with various structures, which were synthesized using polyalkylene oxide-type block
47

copolymers. The location of the polymers was systematically controlled with detailed
structures of the silica framework and the polymerization conditions. Monomers were
selectively adsorbed as a uniform film on silica walls and the polymer-silica composite
structure was obtained by in situ polymerization. The analysis of XRD data and the N2
adsorption isotherms indicates the formation of uniform polymer nanocoating. The
resultant polystyrene-silica composite materials can easily be sulfonated to yield arene
sulfonic mesostructured materials. These materials have very different adsorption
properties compared with mesopores silica because their walls were coated by
hydrophobic polymer.

Figure 6. Route to synthesize mesostructured polymer-silica composite materials

2.5.4. Tuning the acid strength of anchored sulfonic-acid sites


Alkylation reactions are dramatically influenced by the strength of acid sites,
which determines the rate of the reaction and in some cases the feasibility of the catalytic
process. The chance to tune the acid strength of sulfonic-acid groups by close attachment
of different functionalities might enlarge the potential catalytic applications of this type
of material. The presence of electron-withdrawing species close to the sulfonic group will
increase the acid strength of the acid sites. Compared with methylene groups, phenyl
groups are high-electron-withdrawing ones and will increase the acid strength. As
discussed above, Melero et al. (Melero et a l , 2002) prepared ordered SBA-15 materials
containing arene sulfonic-acid groups. The synthesis strategy involved the cocondensation of TEOS and 2-(4-chlorosulfonylphenyl)-ethyltrimethoxy silane (CSPTMS))
using Pluronic 123 as template under acidic conditions. Hydrolysis of the chlorosulfonyl
groups (-SO2CI) to the corresponding sulfonic-acid groups is achieved under acidic
48

condensation conditions. Direct synthesis procedure allowed the effective anchoring of


arenesulfonic groups on the pore surface of SBA-15 mesostructured materials. The
resultant materials showed large uniform pore sizes (ca. 60 ) with large surface areas
(ca. 650 m2/g), good mesoscopic ordering, and thick walls, leading to hydrothermally
stable materials. The acid centers displayed excellent thermal resistance and significant
stability against leaching in aqueous and organic medium. No disulfide species were
detected, as confirmed by

C CP MAS NMR. Solid acidity of the synthesized

arenesulfonic-acid material was characterized by means of

31

P NMR of chemisorbed

triethylphosphine oxide and compared with other solid acids. The arenesulfonic SBA-15
material presented

31

P NMR signal shifts toward higher values in comparison to those

obtained for alkylsulfonic acid sample, confirming enhancement of the acid strength due
to the presence of an electron-withdrawing substituent adjacent to the sulfonic-acid
groups, such as the benzene ring.
Alkane- and arenesulfonic groups are strong acids, but their pKa values are higher
than that of sulfuric acid. The acids or mixture of acids that have lower pKa values than
pure sulfuric acid (pK = -12) are known as superacids (Olah et a l , 1979; Jacquesy,
2004). Some reactions, particularly those that are carried out at low temperatures, require
acid sites reaching the superacidic region. The interest of having solid acids with an
acidity comparable to or even stronger than that of sulfuric acid is obvious considering
that the previously commented advantages of solid acids with respect to liquid or soluble
acids in terms of corrosion, waste minimization and reuse are even more necessary in this
case. The most important example of an insoluble strong solid acid is Nafionperfluorosulfonic acid resin (Olah et a l , 1986). Nafion is developed by copolymerization
of a perfluorosulfonyl ether co-monomer and tetrafluoroethylene. The presence of
electron-withdrawing fluorine atoms in the structure significantly increases the acid
strength of the terminal sulfonic-acid groups, which becomes comparable to that of pure
sulfuric acid and to that of trifluoromethanesulfonic acid (Harmer et a l , 1998; Harmer
and Sun, 2001). However, the catalytic performance of these materials is normally
limited by restricted surface area and low availability of acid sites in the polymeric resin.
To overcome these textural and steric constraints, silica-supported Nafion resin
nanocomposites, such as the commercially available Nafion-SAC-13, have been
49

developed and tested for evaluation in several acid-catalyzed reactions (Molnar, 2008).
Harmer et al. synthesized Nafion-silica composites by entrapping Nafion particles in
porous silica framework to increase the surface area (Heidekum et al., 1998; Harmer et
a l , 1996; Harmer et a l , 1998; Ledneczki et a l , 2005). These nanocomposite materials
have large surface areas (150-500 m2/g) and contain small (< 100 nm) Nafion particles.
However, the materials suffer from limited availability of the acid groups due to the
imperfect dispersion of the resin within the silica pores. The impregnation of Nafion resin
in silica has provided perfluorosulfonic acid materials with higher surface areas and
higher density of available acid sites leading to the enhancement of their acid catalytic
properties. SBA-15 functionalized with perfluorosulfonic acidic Nafion resin using a
post-synthetic

impregnation

method

exhibits better catalytic performance

than

commercial Nafion-SAC-13 in the Friedel-Crafts acylation of anisole (Martinez et a l ,


2008). Wang and Guin reported the incorporation of Nafion to mesoporous MCM-41
silica by sol-gel techniques, concluding that the use of impregnation is more beneficial
than the sol-gel technique for the etherification of olefins (Wang and Guin, 2001).

in dry toluene

-()

-OH

rv

*0

6 hr reflux

SO,H
F p F

Figure 7. Perfluoroalkylsulfonic-modified mesostructured materials synthesized by the


grafting technique
A further development was achieved by anchoring in a single step, and starting
from preformed MCM-41, a perfluoralkanesulfonic group. The resulting material
(denoted as Nafion@MCM-41) exhibits a significant catalytic activity enhancement of
about one order of magnitude with respect to nonzeolitic-silica-supported Nafion (Alvaro
et a l , 2004; Alvaro et a l , 2005). This activity increase is even more remarkable
50

considering that the loading of sulfonic groups of Nafion@MCM-41 has a maximum of


about 1.5 wt%, which is much smaller than that of silica-depolymerized Nafion
composite (15-20 mol% of sulfonic groups). The key point in the preparation of
Nafion@MCM-41 was the reaction of perfluoromethyl /?-sulfone with the free silanol
groups of MCM-41 via direct grafting (Figure 7) (Alvaro et a l , 2004).
However, a problem that still has to be solved is the hydrolysis of the covalent
O-C bond. The problem arises from the hydrolysis of the =Si-0-CF2-bond. This may
require the introduction of a convenient tether between the perfluorinated alkanesulfonic
group and the silyloxy bond. However, the introduction of Si-C bond in the
perfluorosulfonic acid might be better off.

k O s JO*
RO
Et

FF
Precursor

F F

Oft

Silica source

SOjH

Direct Synthesis

Figure 8. Preparation of perfluoroalkylsulfonic-modified mesostructured materials by


direct synthesis.
Harmer et al: prepared mesoporous silica-perfluorosulfonic-acid materials by a
nonoxidative direct-synthesis procedure combining the advantages of the acid strength of
Nafion with the excellent dispersion available with sol-gel templated synthesis
(Macquarrie et a l , 2005). Materials were prepared by co-condensation of the
corresponding silane (precursor of alkylperfluorosulfonic-acid group) and TEOS using
n-dodecylamine as template (Figure 8). The material was tested in the Friedel-Crafts
acylation of anisole with benzoyl chloride in solventless conditions. The catalyst showed
a rapid acylation of anisole and provided high selectivity to the para isomer even after 24
h of reaction at 100 C. The perfluorosulfonic-modified material was also tested in the
reaction of 2-methylfuran and acetone to give the bis-furan and was compared with
propylsulfonic-modified MCM-41 and HMS solids. The activity of the perfluorosulfonic51

acid site was 6 times higher than those obtained with

propylsulfonic-modified

mesostructured materials with selectivities toward bis-furan over 97%.

2.5.5. Lewis acids


All above-mentioned catalysts are Bronsted acids. However, it can be said that, in
principle, each acid-catalyzed reaction can be conducted using a Lewis acid as catalyst,
although with different activity and selectivity. The activity and selectivity also change
depending on the nature of the Lewis acid.
Lewis acids such as BF3, AICI3, SbF5 supported on Si2, AI2O3, cation-exchange
resins, and even pillared layered compounds have also been claimed as alkylation
catalysts (Cooper et a l , 1994; McClure, 1997). Fluorinated silica and alumina have also
been used as alkylation catalysts, but the selectivity to the desired high-octane
trimethylpentanes is low. Ionic liquids formed by AICI3 with quaternary ammonium
halides are also active for isobutane/butene alkylation (Olivier, 1994).
D. Dube et al. synthesized mesostructured strong Lewis acids by grafting
aluminum chloride on mesoporous materials in moisture free conditions (Dube, et a l ,
2005). The performed materials show essentially pure Lewis acid sites. Their catalytic
activity in benzene alkylation by 1-dodecene is comparable to that presented by
aluminum chloride under homogeneous catalysis conditions. The grafted catalysts
recycled up to five times without any change in catalytic activity and selectivity. The
material has been tested as alkylation catalyst by us. It presents good activity at the
beginning. However, AICI3 grafted catalysts are very sensitive to moisture. The activity
decreased dramatically if the catalysts were exposed to ambient air moisture for 5 min.

(MeO, 3 Si-SH

CHjCN

"

- O S H

CV^^S0 3 BiCI 2

C K ^

QSO3H

Q^^^SO#i(OTf)2
^

S*caBi(1)

52

Figure 9. Synthesis of bismuth sulfonate immobilized on silica


P. Sreekanth et al. developed a new route to synthesize strong Lewis acids [SiBi(OTf>2] by immobilizing bismuth sulfonate on silica gel (Figure 9) (Sreekanth et a l ,
2004). The material is not as sensitive to moisture as AICI3 grafted catalysts. It may be a
promising catalyst for isobutene/1-butne alkylation.

2.6. Further developments


As mentioned above, alkylation is associated with a rapid catalyst deactivation
because of coke formation. A catalyst with large pore size which provides enough space
for diffusion, high acid strength comparable to that of sulfuric acid, and hydrophobic
surface which allows a higher paraffin steady concentration in the pores is desired.
PMOs are hybrid organic-inorganic materials containing siloxane moieties bridged
by organic group with well ordered mesopores. The presence of organic groups within
the framework is expected to modify the hydrophilic/carbophilic character of the PMOs.
The carbophilicity of the organic moiety of the walls will allow a higher paraffin steady
concentration under reaction conditions and increase the rate ratio of hydride transfer vs.
oligomerization. Syntheses and applications of different types of PMOs functionalized
with alkylsulfonic and arene sulfonic acid groups have been reported in recent years
(Hamoudi et a l , 2003; Hamoudi et a l , 2004; Melero et a l , 2006). Study on Syntheses of
perfluorinated sulfonic acid functionalized PMOs and alkylation of isobutane/1-butne
over acid functionalized PMOs is still an open field. Such studies however yield original
information about the effect of acid strength and wall polarity of mesostructured solid
acid catalysts. Moreover it is necessary to assess the potential of these materials as
catalysts for alkylation of isobutane/1-butne.
The impregnation of Nafion resin in mesoporous silica provided perfluorosulfonic
acid materials with higher surface areas and more density of available acid sites. The
polarity and structure of mesoporous materials are believed to affect the catalytic
performance. However, there is no report on the effect of polarity and structure on
alkylation activity.
53

Based on these reasons, further study of the synthesis, characterization and activity
measurement of perfluorinated

sulfonic acid functionalized

PMOs and Nafion

impregnated mesoporous materials was chosen as the subject of the present thesis.

54

3. Experimental
3.1. Introduction
This chapter will present details of experimental procedures and instruments used in
this thesis. This section will be divided into three parts, the preparation of catalysts,
characterization and catalytic reaction procedure. In the preparation part, the synthesis of
all catalysts used in Chapters 4 to 7 will be described. In the following part, the
techniques used to analyze materials will be presented. The characterization techniques
which have been used in this thesis include nitrogen adsorption, X-ray diffraction (XRD),
scanning electron microscopy (SEM), transmission electron microscopy (TEM), FTIR of
the chemisorbed pyridine,

29

Si MAS NMR, thermogravimetric analysis (TGA),

intelligent gravimetric analyzer (IGA) and elemental analysis of carbon, sulphur and
aluminum. In the last part, the experimental setups, experimental procedures and
calculations of isobutane/1-butne alkylation reaction will be presented.

3.2. Synthesis
In this thesis, a series of acid functionalized periodic mesostructured silica and
organosilica (ethane-silica) have been synthesized. First, mesostructured SBA-15 silica
and an ethane-silica with similar structure were synthesized. Then, acid functional groups
were grafted by a post synthesis procedure.

3.2.1. Preparation of SBA-15 and SBA-15 type PMO


The original synthetic procedure of mesostructured SBA-15 silica was developed by
Zhao et al. (Zhao et al., 1998). The material was synthesized by the acid-catalyzed
hydrolysis and condensation of tetraethyl orthosilicate (TEOS, Aldrich) using the triblock
copolymer Pluronic PI23 (EO20PO70EO20, BASF) as the structure-directing agent. In a
typical synthesis, 4 g of PI23 was dissolved in 125 g of 2 M HC1 solution under stirring
at 40 C. This was followed by adding 9.12 g of TEOS into the solution as silicon source.
After being stirred vigorously for 24 h at 40 C, the resulting gel was transferred to a
Teflon-lined autoclave and heated at 100 C for an additional 48 h. After cooling to
55

ambient temperature, the solid in the autoclave was recovered by filtering, washing with
deionized water and drying at 80 C. Finally, the solids were calcined at 550 C for 5 h to
remove the organic surfactant.
In this thesis, the PMO stands for "Periodic mesostructured organosilica (ethanesilica)". It was synthesized by the acid-catalyzed hydrolysis and condensation of 1,2-bis
(trimethoxysilyl)ethane (BTME, Gelest) using Pluronic PI23 according to the procedure
adopted by Muth et al. (Muth et a l , 2001). In a typical synthesis, 2 ml of BTME was
added to a mixture of 1.5 g of Pluronic PI23, 15 ml of 12.2 M HC1, and 75 ml of
deionized water under stirring at 40 C. After the mixture was stirred for 24 h, the
resulting gel was aged at 90 C for an additional 24 h. The obtained white precipitate was
separated by filtration, washed thoroughly with deionized water and dried at room
temperature. The surfactant was removed by solvent extraction with anhydrous ethanol in
a Soxhlet apparatus for 48 h.

3.2.2. Preparation of arene-sulfonic acid functionalized SBA-15


Arene-sulfonic acid functionalized SBA-15 was synthesized according to the
procedure described by Melero et al. (Melero et a l , 2002). The material was synthesised
as follows: 4 g of Pluronic 123 were dissolved with stirring in 125 g of 1.9 M HC1 at
room temperature. The solution was heated to 40 C before addition of TEOS. A TEOS
prehydrolysis time of 45 min was used prior to the addition of the 2-(4chlorosulfonylphenyl)-ethyltrimethoxysilane (CSPTMS, Gelest) to the mixture. The
molar composition of the mixture was TEOS, 2; CSPTMS, 0.22; HC1, 12.97; P123,
0.0378; H 2 0, 360. The resultant solution was stirred for 20 h at 40 C, after which time
the mixture was aged at 100 C for 24 h under static conditions. The solid product was
recovered by filtration and air-dried at room temperature overnight. The template was
removed from the as-synthesised material by washing with ethanol under reflux in a
Soxhlet apparatus for 48 h.

56

3.2.3. Preparation of arene-sulfonic and propyl-sulfonic acid


functionalized PMO
Arene-sulfonic acid functionalized PMO (ArS-PMO) was synthesized as described
for the arene-sulfonic acid functionalized SBA-15 above. In the presence of PI23, the
synthesis gel molar composition was BTME, 1; CSPTMS, 0.22; HC1, 12.97; PI23,
0.0378; H 2 0, 360.
As for propyl-sulfonic functionalized ethane silica (PrS-PMO), the synthesis gel
molar composition was BTME, 0.75; MPTMS (3-mercaptopropyltrimethoxysilane,
Gelest), 0.25; P123, 0.034; HC1, 11.7; H 2 0, 326. The surfactant was removed by solvent
extraction with anhydrous ethanol in a Soxhlet apparatus for 48 h. Conversion of the thiol
bearing groups into sulfonic acid moieties was performed on extracted samples via
oxidation using hydrogen peroxide. Typically, 0.2 g of extracted material was contacted
with 8 g of aqueous H2O2 (30%) and the suspension was stirred at room temperature for
24 h. After filtration and washing with deionized water and ethanol separately, the
oxidized samples were acidified in 0.1 M H2SO4 solution (1 wt%) during 2 h.
Subsequently, the samples were washed thoroughly with deionized water until neutral pH,
filtered and vacuum dried at 60 C overnight.

3.2.4. Preparation of sulfonated polystyrene coated SBA-15


Sulfonated

polystyrene

coated

SBA-15

(SPS-SBA-15)

was

prepared

by

polymerization of styrene (80 mol%) and divinylbenzene (Aldrich, 98%) (20 mol%) with
a,f'-azoisobutyronitrile (Aldrich, 98%) (3% relative to total vinyl group) inside SBA-15,
followed by sulfonation using concentrated H2SO4, according to the method adopted by
Choi et al. (Choi et a l , 2005). In a typical synthesis, 1 g of SBA-15 was impregnated
with 0.250 ml of styrene (80 mol%) and dissolved with 0.078 ml of divinylbenzene (20
mol%) and 0.016 g (3% relative to total vinyl group) of a,a'-azoisobutyronitrile in 1.50
ml of dichloromethane. After impregnating the solution, the silica sample was dried at
40 C for 2 h to remove dichloromethane and was heated for 6 h at 60 C for the
polymerization. The resultant sample was washed with chloroform and ethanol and dried
57

at 80 C. The polystyrene-SBA-15 samples were sulfonated in concentrated H2SO4 at


100 C for 6 h. After extensive washing with distilled water, samples were vacuum-dried
at room temperature overnight.

3.2.5. Preparation of aluminum chloride grafted SBA-15 and PMO


The synthesis method of aluminum chloride grafted SBA-15 and PMO was derived
from the one described by Dube et al. (Dube et a l , 2005). 2.5 g of SBA-15 or PMO was
dried at 400 C under vacuum (overnight) and kept under dry argon in a Schlenk tube.
The dried SBA-15 or PMO was added to 100 ml of benzene (sodium dried) in the
reaction tube under a dry argon atmosphere. To this solution was added 1.5 g of
aluminum chloride stored in a Schlenk tube. The mixture was refluxed for 3 h under dry
argon and then filtered and washed three times in a Schlenk tube with dry benzene (100
ml). The grafted SBA-15 or PMO was stored in a Schlenk tube under a dry argon
atmosphere and denoted Al-SBA-15 or Al-PMO.

3.2.6. Preparation of perfluoroalkylsulfonic acid functionalized SBA-15


and PMO
The synthesis of perfluoroalkylsulfonic acid functionalized SBA-15 (FS-SBA-15)
and PMO (FS-PMO) materials occurs in two steps: the preparation of mesostructured
support and the post-synthesis grafting. SBA-15 and PMO were prepared following the
procedure described above. The grafting of functional group was carried out as follows
(Figure 10): 2 g of solvent extracted mesostructured support was pretreated in a Schlenk
tube under vacuum at 120 C for 12 h and stored under dry argon. The support was
introduced into a Schlenk tube kept under argon containing 50 mL of toluene dried over
sodium. To this solution, was added 2 g of l,2,2-trifluoro-2-hydroxy-l- trimethylethane
sulfonic acid /?-sultone (THTS, Synquest) stored at 4 C. The mixture was refluxed for 4
h under magnetic stirring and argon atmosphere. The material was then filtered, washed
three times with toluene (50 mL) and dried at 100 C.

58

PMO

PMO

(SBA-15)

(SBA-15)
F

F3C

VVs03H

VF

THTS
Figure 10. Preparation of perfluorosulfonic acid functionalized SBA-15 or PMO

3.2.7. Preparation of Nafion functionalized SBA-15 materials (Nafion(X)


/SBA-15 and Nafion(X)/Me-SBA-15)
The designation of catalysts reflects characteristics of the material. X indicates the
theoretical wt% of Nafion loading (in this thesis, two Nafion loading were studied, i.e. X
= 15 or 30). Me-SBA-15 indicates that the surface -OHs of SBA-15 were silanized by
ethoxytrimethylsilane. For example, the Nafion(30)/Me-SBA-15 means that 30 wt% of
Nafion resin was supported on the SBA-15 which surface OHs were capped by
silylation.
Silylation of the surface -OHs of SBA-15 was carried out as follows: 2.6 g SBA-15
was pre-dried under vacuum at 200 C for 12 h before adding 3.5 g ethoxytrimethyl
-silane (Aldrich) and 30 ml of dry toluene under Argon. The mixture was refluxed at
100 C for another 12 h. Then, the -OHs capped SBA-15 material was filtered and
washed by toluene and anhydrous ethanol in mm. At last, the solid was dried at 80 C
overnight.
The incorporation of Nafion was carried out by a simple impregnation procedure
using a Nafion 5 wt% water-alcoholic solution as precursor (Aldrich). This
impregnation method consisted in the dispersion of powder SBA-15 and Me-SBA-15
silica with the appropriate amount of Nafion water-alcoholic solution to achieve
theoretical 15 wt% and 30 wt% of resin loading. The impregnation process was carried
out at 60 C and atmospheric pressure under stirring for 6 h. The solid was firstly dried
at room temperature for 12 h in static conditions, and then the water and alcohols was
evaporated thoroughly under vacuum at 60 C for an additional 12 h.
59

3.2.8. Preparation of Nafion functionalized SBA-16 materials (Nafion(X)


/SBA-16 and Nafion(X)/Me-SBA-16)
As described above, X indicates the theoretical wt% of Nafion loading and MeSBA-16 indicates that the surface -OHs of SBA-16 were silanized by ethoxytrimethyl
-silane. For example, the Nafion(30)/Me-SBA-16 means that SBA-16 which surface
OHs were capped by silylation was functionalized by Nafion with loading of 30 wt%.
SBA-16 mesoporous silica sample was synthesized by the acid-catalyzed
hydrolysis and condensation of tetraethyl orthosilicate (TEOS, Aldrich) using the
triblock copolymer F127 (EO106PO70EO106, Aldrich) as the structure-directing agent,
with the aid of K 2 S0 4 (Yu et a l , 2002). In a typical synthesis, 4 g of F127 and 10.48 g
of K2SO4 were dissolved in 120 g of 0.5 M HC1 solution under stirring at 38 C. This
was followed by adding 16.8 g of TEOS into the solution as silicon source. After being
stirred vigorously for 24 h at 38 C, the resulting gel was transferred to a Teflon-lined
autoclave and heated at 100 C for an additional 24 h. After cooling to ambient
temperature, the solid in the autoclave was recovered by filtering, washing and drying at
80 C. Finally, the solids were calcined at 550 C in air for 6 h to remove the organic
surfactant.
Me-SBA-16 was synthesized as described above for the Me-SBA-15. 2.6 g of
SBA-16 was pre-dried under vacuum at 200 C for 12 h before adding 3.5 g
ethoxytrimethylsilane (Aldrich) and 30 ml of dry toluene under Argon. The mixture was
refluxed at 100 C for 12 h. Finally, the solids were filtered, washed and dried at 80 C
overnight.
Nafion was incorporated on SBA-16 and Me-SBA-16 using the impregnation
method described above for Nafion/SBA-15. Materials with theoretical 15 wt% and 30
wt% of Nafion resin loading were synthesized.
Other commercial acid catalysts: Nafion membrane, Nafion-Si02 composite (SAC13), Zeolite Y (Si/Al=10.5) and Amberlyst-15 were supplied by Sigma-Aldrich. Nafion

60

were pretreated with 0.5 M H2SO4, washed with deionized water and dried at 70 C on
vacuum over night. Zeolite Y was heated at 500 C for 12 h before used.

3.3. C haracterization
3.3.1. Nitrogen adsorption isotherms
Nitrogen adsorption is commonly used for characterization of porous materials,
including the determination of specific surface area, pore volume and pore size
distribution. The sorption isotherms of nitrogen measured at its condensation temperature
(-196 C ) reflect the textural characteristics of the materials. In 1985, the IUPAC
developed a standard classification for these sorption isotherms into six main types as
shown in Figure 11 (Sing et a l , 1985). Type I isotherms are usually considered to be
indicative of adsorption in microporous materials like zeolites or monolayer adsorption at
low relative pressure. Type II and III isotherms are the normal form of adsorption by
nonporous or macroporous adsorbents in which the adsorption proceeds via multilayer
formation. Type IV and V isotherms are characteristic of adsorption on mesoporous
materials that proceeds via multilayer adsorption followed by capillary condensation at
higher relative pressure. Type VI isotherm represents stepwise multilayer adsorption on a
uniform nonporous surface.

III

r
IV
Capillaiy .
evaporation | |

w<-
V

VI

1fl

^^'""'Capillaiy '
' \
condensation
Relative pressure &p-

Figure 11. Types of sorption isotherms (Sing et a l , 1985).


61

When nitrogen is sorbed in mesoporous materials, hysteresis loops (the sorption and
desorption branches do not coincide at high relative pressure) are often found with such
type IV isotherms. Normally the isotherm can be divided into three parts under partial
pressures. In the range pip 0 < 0.3, the monolayer and multilayer adsorption will occur on
the surface of the material. This area allows to calculate the specific surface area. In the
range 0.42 < pip 0 < 0.8, the pores may be filled with capillary condensation which is
characteristic with increasing volume under the same partial pressure. This section
provides information on pore size. Finally, when pip 0 > 0.8, the interparticle porosity will
be filled.
The specific surface of a material is determined from the linear part of isotherm by
using the BET (Brunauer Emmet Teller) equation (Eqn. 3.1):
p i Pc.
1
C \
p

=
+
x
V(l-p/p0)
VmC VmC P o

(Eqn.
3.1);
q

where V is the adsorbed volume,


Vm is the volume of gas adsorbed in a complete monolayer,
P is the pressure of the adsorbate,
P 0 is vapor-liquid equilibrium pressure of the adsorbate at the temperature of
measurement,
C is the BET constant, proportional to exp(-(AHd-AHvap)IRT), AHj is the enthalpy of
adsorption in the first layer and AHvap is the heat of vaporization.
The pore diameter of mesostructured materials can be determined using the BJH
(Barret-Joyner-Halenda) method (Barret et al., 1951) based on the Kelvin equation (Eqn.
3.2) in the form:

In

Po

'*- x

R T

(Eqn. 3.2)

n
62

where plp 0 is the relative pressure of vapor in equilibrium with a meniscus having a
radius rk (Kelvin radius), y and Vi are surface tension and molar volume of liquid nitrogen
at its boiling point, R and T are universal gas constant and isothermal temperature of
nitrogen. This method is often used to estimate the mesopore size distribution of the
materials with non-connected cylindrical pore channels (Kruk et a l , 1997; Lukens Jr. et
a l , 1999). For the materials with cagelike pore structures such as SBA-16, this method is
not suitable to evaluate the pore size. BJH pore size calculations assuming a cylindrical
pore model systematically underestimates the pore size of cagelike pores (Kleitz et al.,
2006; Ravikovitch and Neimark, 2002a; Ravikovitch and Neimark, 2002b). However, up
to now, no generally accepted model is available for pore size evaluation of cagelike
pores with a spherical geometry. The pore size of SBA-16 was estimated using BJH
method in this thesis.
The pore volume was determined from the amount of nitrogen adsorbed at -196 C
at the relative pressure of 0.99. At such a pressure the channels of the samples are
assumed to be completely filled with nitrogen.
The analysis of the surface by nitrogen volumetry was performed using an
Omnisorp-100 automatic analyser (Chapter 4 and 5) and a Micromeritics TRISTAR 3000
apparatus (Chapter 6 and 7). The heat treatment of samples was performed at 120 C
under reduced pressure for 6 h. The specific surface area was calculated with the multipoint BET method in the relative pressure range of 0.05 to 0.25. The pore diameter was
estimated from the peak position of the BJH pore size distribution.

3.3.2. X-ray diffraction (XRD)


X-ray diffraction (XRD) is an extensively used technique for identification and
characterization not only of the crystal structure of zeolites or zeolite-like catalysts but
also the structure of amorphous mesostructured materials. The power X-ray diffraction
pattern of microporous zeolite-type solids is generally taken between the 2 lvalues of 5
and 40 because the most intense peaks characteristic of these materials occur within this
range. On the other hand, the presence of distinct reflections in the X-ray diffraction
63

pattern of mesoporous molecular sieves is observed in the low range of the 2 G angle (0 to
5), due to the long range regularity of these materials.
In this research, the structural mesoporous phase of acid functionalized mesoporous
materials was monitored by powder low-angle X-ray diffraction (XRD), recorded on a
German Brucker D4 X-ray diffractometer with Ni filtered

CUKQ

radiation (1 = 1.5406 )

in the range of 2lvalues from 0.5 to 5with a scan step size of 0.02. The tube voltage
was 40 kV, while the current was 40 mA.

3.3.3. Fourier transform infrared spectroscopy (FTIR)


Pressure reading

IR cell

Figure 12. Vacuum system used for the adsorption of pyridine


Fourier transform infrared spectroscopy (FTIR) has been widely used for
characterizing the surface chemistry of heterogeneous catalysts (Ryczkowski, 2001). In
principle, IR spectra arise from transitions between vibrational or rotational energy levels
corresponding to certain normal modes of vibration of molecules or surface group. Useful
information about structure and surface properties of catalysts can be obtained from
framework vibrations by IR spectroscopy in the region between 4000 and 400 cm"1.
Using pyridine as probe molecule, information concerning the nature and acid strength of
acid sites (Bronsted and Lewis) can be obtained. All FTIR spectra were recorded on a
Biorad FTS-60 spectrometer connected to a data acquisition personal computer through
64

the conventional RS-232 interface. Generally, the IR spectra of the framework vibrations
were collected in the transmittance mode at a spectral resolution of 2 cm"1, while IR
spectra of the pyridine adsorbed were collected in the absorbance mode.
The pyridine adsorption was carried out in the vacuum system shown in Figure 12.
The vacuum system has mainly two pumps: a mechanical pump to a primary vacuum and
a diffusion pump to achieve a pressure around 10"5 Torr. It has a trap maintained at 77K
by liquid nitrogen, an input for introduction of pyridine and a section where you can set
the infrared cell for adsorption and desorption of pyridine at different temperatures.
The glass cell (pyrex and quartz) is used to record infrared spectra described in
Figure 13. It has a part that allows samples to be heated in a furnace tube and another for
infrared analysis. The analysis part is fitted with two KBr windows. The cell can be
maintained under vacuum or dry argon atmosphere.

1 ]

\ML

Pellet Support

L
KBr Window

Figure 13. Infrared cell for the adsorption of pyridine


Sample preparation is the most important step in this characterization. To do so, the
samples were weighed (10 mg), pressed into a wafer and installed on the support. Once
the infrared cell has been closed, it was evacuated and maintained at pressure of 10"5 Torr
and temperature of 120 C overnight. After this treatment, the cell was cooled down to
room temperature. In the adsorption/desorption step, pyridine was then admitted in the
sample chamber at the reduced pressure of mechanical pump for 10 min. Once adsorption
was complete, the system was vacuumized by the diffusion pump to remove the excess
and physisorbed pyridine in the system. The system was stabilized for 120 min. A first
infrared reading was taken between 400 cm"1 and 4000 cm"1 in order to check whether the
65

pyridine was well adsorbed or not. Subsequently, the wafers were evacuated again at 10"5
torr and 120 C to keep only the pyridine chemisorbed on acid sites. The spectra were
recorded after cooling at room temperature. The absorption bands of adsorbed pyridine
on acid sites of catalysts were between 1400 cm"1 and 1700 cm"1.

3.3.4. Magic angle spinning nuclear magnetic resonance spectroscopy


(MAS NMR)
NMR spectroscopy is one of the principal techniques used to obtain physical,
chemical, electronic and structural information about molecules due to the chemical shift
effect on the resonance frequencies of the nuclei present in the sample. The magic angle
spinning NMR (MAS NMR) including

29

Si MAS NMR and 27A1 MAS NRM is very

useful to characterize complex aluminosilicate materials as it can distinguish different


9Q

chemical environments of Si and Al atoms. Thus

Si MAS NMR spectra can allow to

characterize the type of grafting and determine the number of Si-O-X (X = C, Si, OH)
bonds around a silicon atom (Fyfe, 1983; Engelhardt and Michel, 1987; Bell and Pines,
1994). The resonances were assigned to Qx and T* where Q indicates that silicon can
make four bonds and x is the number of bonds Si-O-Si between 1 and 4. The symbol T
indicates that silicon is coupled with the neighboring carbon and x is the number of bonds
Si-O-Si between 1 and 3 (Table 4).
Table 4. Nomenclature and chemical shift of Q* and T*
Q2

Q3

Q4

T1

T2

T3

~94ppm

-100 ppm

-109 ppm

-43 ppm

~54ppm

-67 ppm

Si(OSi)2(OH)2

Si(OSi)3OH

Si(OSi)4

RSiOSi(OH)2 RSi(OSi)2OH

RSi(OSi)3

97

On the other hand,

Al MAS NMR spectra of the aluminosilicate materials

generally consist of a resonance corresponding to the tetrahedrally coordinated lattice


aluminum in a relatively narrow range of chemical shifts from 50-68 ppm, while the
66

octahedrally coordinated extra-framework aluminium produces a NMR resonance at


about 0-10 ppm (Fyfe, 1983; Engelhardt and Michel, 1987; Kosslick et a l , 1997; TrongOn and Kaliaguine, 2002). A broad signal at 30-50 ppm is associated to non-framework
aluminum species in a disturbed tetrahedral coordination or a penta-coordinated state,
which is found usually in the amorphous mesoporous materials (Rocha et a l , 1991;
Trong-One/a/., 1998).
In this thesis, solid-state NMR measurements were performed under magic angle
spinning (MAS) condition at room temperature on a Bruker ASX 300 spectrometer with
4 mm cylindrical zirconia rotors at a spin rate of 4 or 8 kHz. 29Si MAS NMR spectra
were recorded at a frequency of 59.60 MHz using 30 pulses of 3 ps duration, 2600 scans
and 30 s recycle delays. 27A1 MAS NMR spectra were collected at a frequency of 78.17
MHz using 90 pulses of 4 ps duration and 50000-60000 scans with a recycle time of 0.2
s.

3.3.5. Transmission electron microscopy (TEM)


Transmission electron microscopy (TEM) is a technique used to examine the surface
and bulk microstructure of materials. In this thesis, TEM images were obtained on a
JEM-2011 transmission electron microscope (JEOL Company) combined with energy
dispersive X-ray spectroscopy (EDX) operating at 200 kV. The samples were prepared
by sonication in ethanol and suspended on holey carbon grids.

3.3.6. Scanning electron microscopy (SEM)


Scanning electron microscopy (SEM) is a powerful technique to study the surface
morphologies (texture, shape and size) of the bulk particles of solid materials. In this
thesis, SEM and EDX microanalyses were obtained using a Philip-XL30 apparatus
operated at 20 kV.

67

3.3.7. Thermogravimetric analysis (TGA)


The thermogravimetric analysis is allowing to measure the weight loss of a sample
as a function of increasing temperature. In this way, by analyzing the weight loss at
different temperatures, it is possible to identify different levels of degradation. In this
thesis, the TGA were recorded by using a Perkin-Elmer TGA 7 from room temperature
up to 800 C at a heating rate of 5 or 10 C/min under air.

3.3.8. Flame atomic absorption spectroscopy (AAS)


The atomic absorption spectrometry is a method of elemental analysis. The
quantitative analysis is based on Beer-Lambert Law (Eqn. 3.3):
A = log (Io/I) = klc

(Eqn. 3.3)

where k is the absorption constant,


/ is the length of the absorbing medium
c is the concentration
According to this equation, the absorption is proportional to the concentration of
absorbent species for given conditions of the instrument. However, when the
concentration and the absorbance increase, the imperfections of the absorption process
will cause a deviation from the standard curve. It is important to dilute the solution in
order to locate the absorption in a linear zone.
The AAS experiment was carried out on a Perkin-Elmer 1100B atomic absorption
spectrometer equipped with a multiunit Perkin-Elmer 30 mA lamp with N20-acetylene
flame. The wavelength for aluminum analysis is 309.3 nm for a linear zone between 0
and 100 ppm. The catalyst samples were dissolved in a mixture of HC1 and HF. Typically,
100 mg of solid sample was introduced into a 100 ml polyethylene digestion bottle,
which contained a binary acid mixture (1 ml of concentrated HF (48-49%) and 25 ml of
diluted HC1 (10%)). A reference solution which contains no sample was also prepared.
68

The bottle was then placed in an oven at 60 C and shaked overnight to completely
dissolve the sample. The digested sample was then transferred to an acid-cleaned
polyethylene volumetric flask and the total volume was adjusted to 100 ml by adding
deionized water.

3.3.9. Elemental analysis of carbon and sulfur


The elemental analysis of carbon and sulfur aims at determining the mass fraction of
mesostructured materials acid functional groups. These measures of content of carbon
and sulfur were performed using a Vario EL III apparatus (CHNS model).

3.3.10. Ion exchange capacity


The ion exchange capacity is a technique to measure the acid site concentration of
acid catalysts. The ion exchange capacity was determined by using solution of sodium
chloride 2 M (NaCl) and sodium hydroxide 0.1 M (NaOH). In a typical experiment, 0.1 g
of solid was added to 20 g of 2 M NaCl solution, and vigorously stirred at room
temperature overnight. The resulting suspension was filtered and washed thoroughly with
a total amount of 80 ml 2 M NaCl to retain the hydrogen ions in the solution, and
thereafter titrated potentiometrically by a drop wise addition of a 0.1 M NaOH.

3.3.11. Intelligent gravimetric analyzer (IGA)


The IGA apparatus is an ultrahigh vacuum system which allows the adsorptiondesorption isotherms and the corresponding kinetics of adsorption or desorption at each
pressure step to be determined. The system consists of a fully computerized microbalance
which automatically measures the weight of the samples as a function of time with the
gas vapor pressure and sample temperature under computer control. The microbalance
had a long term stability of 1 pg with a weighing resolution of 0.1 pg. The equilibrium
pressures were determined by highaccuracy Baratron pressure transducers, and
maintained at the set point by active computer control of the admittance/exhaust valves
throughout the duration of the experiment. The sample temperature was also monitored

69

throughout the adsorption process and regulated to 0.1 C by either a water bath or a
furnace.
Water adsorption isotherms at 25 C were carried out using IGA (IGA-002, Hiden,
UK). Before the measurements, samples (ca. 50 mg each) were outgassed up to a
pressure of <10~5 Pa at 200 C for about 12 h. The adsorption isotherms were measured
by increasing the equilibrium pressure in small stepwise. The mass uptake was measured
as a function of vapor pressure.

3.4. Catalytic reaction


3.4.1. Reaction system and procedure
Liquid phase alkylation of isobutane (> 99% purity) with 1-butne (> 99% purity)
was carried out in an automated stainless-steel fixed-bed continuous reactor. The reaction
system, including the feed section, reaction zone, and sampling system, was computer
controlled, and it is schematized in Figure 14. The reactants are maintained in the liquid
phase in two separate vessels under nitrogen pressure and fed at the desired liquid feed
rate by means of two piston-type pumps. Isobutane and 1-butne are mixed before
entering the reaction section. The feedstock contains 1 wt% of nonane as internal
standard, Mass balances are calculated based on the flow rate of reactants and products.
Over 95% of mass balance is achieved. Nitrogen is used for pretreating catalysts,
increasing pressure and purging the system. Helium is used to carry samples collected in
the sample loops of a 16-loop sampling valve, to the sample valve of a GC for analysis.
The reactor is heated by a single-zone furnace. Reactor pressure is regulated through a
manually controlled back pressure regulator.
In each run, 0.35 g of catalyst pellets with diameters of 0.5 to 0.8 mm, obtained by
compressing the powder into tablets, crushing, and sieving, was loaded in a reactor tube
with internal diameter of 8 mm. The catalyst bed was fixed between two plugs of quartz
wool and the remaining empty volume filled with quartz beads. Before being used, the
catalysts were pretreated in situ at 110 C for 10 h in a flow of N2 of 15 ml/min.
Thereafter the reactor was cooled to reaction temperature and a pressure of 3.2-3.4 MPa
70

I
l/J

a
I

I
I
u

3
c
5

"S
Cfl

I
U.

J3

a.

CM

I'

71

was established with nitrogen by adjusting the back pressure regulator. Once the desired
pressure was achieved, the feedstock consisting of an isobutane/1 -butne mixture with a
molar ratio of 20/1 and 40/1 was delivered to the reactor by switching the three way valve
T2. In order to obtain different initial olefin conversions, the WHSV, referred to 1-butne,
was varied (1.4 and 2 h"1) by changing the flow rate of feedstock. The effluent from the
reactor passes the 16-loop valve sampling system before going through the back pressure
regulator. In this way samples can be taken during the run at intervals of 1 min. The time
zero of the reaction corresponds to the observation of the liquid front of products
observed through the high pressure plastic tube placed next to the 16-loop sampling valve.
After completion of the experiment, helium flow flushes out the loop of the 16-loop
sampling valve and transports the sample to the on-line GC (Thermo Trace GC Ultra) for
analysis by switching the 4-port valve SVI. Reactants and products were separated in a
100-m capillary column (CP SIL PONA CB) and the individual C4-C8 hydrocarbons
identified by means of available reference standards and GC-MS analysis. No effort has
been made to identify individual C9 and higher compounds.

3.4.2. Calculation
Alkylation is catalyzed by sites with strong acidity, while double bond isomerization
is catalyzed by very weak acid sites (Corma et al., 1994c). Even a fully deactivated
zeolite retains some activity for isomerizing butnes. Thus the conversion is calculated
based on disappearance of butnes (1-butne, 2-butene and isobutene). The selectivity of
a product was defined as the ratio between the weight of the product and the weight of
converted butne. For pure isobutane-butene alkylation, the alkylate selectivity value
equals 204 wt%, whereas for butne dimerization it equals 102 wt%. The olefin
conversion and the selectivity of products were calculated as follows (Eqn. 3.4 and Eqn.
3.5):

72

(%wtC~)1 ~(%wtC~f

Butne conversion (%wt) = 100 x -

(Eqn. 3.4)

(%tc-y
Selectivity of j (%wt) = 100 x

_ 1 W t J
(%wtCt\) -(%wtC~y
4

(Eqn. 3.5)

where / and / superscripts stand for "initial" and a " t " times respectively, j stands
for the reaction products.

73

4. Preliminary Studies of Isobutane/1-butne Alkylation over


Mesostructured Solid Acids
Abstract
In this study, a series of acid functionalized mesostructured materials were
synthesized and characterized by BET, XRD, 29Si MAS NMR, FT-IR and TEM etc. The
typical mesostructured materials synthesized were tested in the liquid phase alkylation of
isobutane by 1-butne. The catalytic activity is dramatically influenced by the strength of
acid sites. Only the catalysts with acid sites stronger than amberlyst-15 can catalyze this
alkylation reaction. Hydrophobic surface and large pore size favor catalytic activity. The
perfluoroalkylsulfonic acid functionalized PMO (periodic mesoporous organosilicas)
exhibited excellent catalytic activity as well as catalytic stability compared to zeolitic
catalysts. Its high and stable activity can be attributed to its high acid strength, larger pore
size and carbophilicity of the pore surface.

4.1. Introduction
Alkylation of isobutane/-butene is representative of the refining operation which
produces gasoline from the C3 and C4 hydrocarbons. This operation is currently
performed using strong liquid acids (HF or H2SO4) as the catalysts. Due to environmental
considerations, the use of solid acids as catalysts for the alkylation process has being
investigated over the past 20 years. Many solid acids, such as zeolites, sulfated zirconia
and related materials, heteropolyacids and acidic polymers, have been studied for this
reaction (Corma et a l , 1993; Feller and Lercher, 2004). None of them have met
commercial success because of a rapid catalyst deactivation due to the preferential
adsorption and pore filling with the olefin. Certain properties can be identified as
necessary to achieve a high alkylation yield and selectivity. Bronsted acid sites of high
strength are preferred; strong acid sites are necessary to catalyze alkylation reaction; a
narrow distribution of acid strength is important to good selectivity; carbophilicity of
catalyst surface is desired, which allows a higher paraffin steady concentration in the
vicinity of the active site; large internal pore size is required, it facilitates the diffusion of
74

reactants to the active site and diffusion of the products out of the pore network.
Unfortunately, none of the solid acids mentioned above possesses all these properties.
Mesoporous molecular sieves have provided numerous possibilities in the
development of solid catalysts (Trong On et a l , 2003). Many catalytically active
functional groups, such as acid groups, can be grafted on the pore wall, which still leaves
enough pore space. This facilitates the diffusion of reactants to the active sites and
alkylation products out of the pore network. Surfaces of materials have very different
adsorption behaviours when different silicon sources are used, such as organic/inorganic
hybrid materials (Asefa et a l , 1999). Several such materials have already been
synthesized with a composition comparable to that of silicon rubber, namely siloxane
based materials. The carbophilicity of the organosilicon moiety of walls should allow a
higher paraffin steady concentration in the reaction conditions, which may decrease the
formation of high hydrocarbons by olefin polymerization and prolong the lifetime of
solid acid catalysts. The acid strength of materials can be tuned by grafting different acid
functional groups. The purpose of this work is to synthesize solid acids which have large
surface area, large pore size, hydrophobic surface and strong acid sites. Propyl-sulfonic
acid, arene-sulfonic acid and perfluoroalkylsulfonic acid functionalized SBA-15 and
SBA-15 type PMO, aluminum chloride grafted SBA-15 and PMO, Sulfonated
polystyrene coated SBA-15 and PMO were synthesized and compared as catalysts in
alkylation of isobutane with 1-butne.

4.2. Experimental
4.2.1. Catalyst preparation
SBA-15 mesoporous silica sample was synthesized by the acid-catalyzed hydrolysis
and condensation of tetraethyl orthosilicate (TEOS, Aldrich) using the triblock
copolymer Pluronic PI23 (EO20PO70EO20, BASF) as the structure-directing agent,
according to the procedure adopted by Zhao et al. (Zhao et al., 1998a). The PMO
material was synthesized by the acid-catalyzed hydrolysis and condensation of bis(trimethoxysilyl) ethane (BTME, Gelest) using Pluronic PI23 according to the procedure
75

adopted by Muth et al. (Muth et a l , 2001). After synthesis, the triblock copolymer was
removed by extraction with ethanol. In a typical synthesis, 2 ml of BTME was added to a
mixture of 1.5 g of Pluronic P123, 15 ml of 12.2 M HC1, and 75 ml of deionized water
under stirring at 40 C. After the mixture was stirred for 24 h, the resulting gel was cured
at 90 C for an additional 24 h. The obtained white precipitate was separated by filtration,
washed thoroughly with deionized water and dried at room temperature. The surfactant
was removed by solvent extraction with anhydrous ethanol in a Soxhlet apparatus for 48
h.
Arene-sulfonic and propyl-sulfonic acid functionalized PMOs were synthesized
according to the procedure described by Hamoudi et al. (Hamoudi et a l , 2004). As for
arene-sulfonic acid functionalized ethane silica (ArS-PMO), in the presence of PI23
(EO20PO70EO20, BASF), the synthesis gel molar composition was BTME (1,2bis(trimethoxysilyl)ethane, Gelest), 1; CSPTMS (2-(4-chlorosulfonylphenyl)-ethyltrimethoxysilane, Gelest), 0.22; HC1, 12.97; PI23, 0.0378; H 2 0, 360. An arene-sulfonic
functionalized SBA-15 (ArS-SBA-15) was also synthesized for comparison. As for
propyl-sulfonic functionalized ethane silica (PrS-PMO), the synthesis gel molar
composition was BTME, 0.75; MPTMS (3-mercaptopropyltrimethoxysilane, Gelest),
0.25; P123, 0.034; HC1, 11.7; H 2 0, 326. Conversion of the thiol bearing groups into
sulfonic acid moieties was performed on extracted samples via oxidation using hydrogen
peroxide. Typically, 0.2 g of extracted material was contacted with 8 g of aqueous H2O2
(30%) and the suspension was stirred at room temperature for 24 h. After filtration and
washing with deionized water and ethanol separately, the oxidized samples were acidified
in 0.1 M H2SO4 solution (1 wt%) during 2 h. Subsequently, the samples were washed
thoroughly with deionized water until neutral pH, filtered and vacuum dried at 60 C
overnight.
Perfluoroalkylsulfonic acid functionalized PMO (FS-PMO) was synthesized by
grafting THTS (l,2,2-trifluoro-2-hydroxy-l-trifluoromethylethane sulfonic acid/?-sultone,
Synquest) on PMO (Figure 15). 2.0 g of solvent extracted PMO was pretreated at 120 C
in vacuum for 12 h and cooled down to room temperature. 50 ml dry toluene and 2.0 g of
THTS were then added. The mixture was refluxed for 4 h under Ar atmosphere. The
76

resulting solid was filtered, washed thoroughly with toluene and dried at 100 C
overnight. In order to compare the catalytic activity, perfluoroalkylsulfonic acid group
functionalized SBA-15 (FS-SBA-15) was also prepared according to the method adopted
by Alvaro et al. (Alvaro et a l , 2004).
F

PM0

+F
F

F3C

x^r
F

PMO^

^S03H

THTS
Figure 15. Preparation of perfluorosulfonic acid functionalized PMO
Sulfonated

polystyrene

coated

SBA-15

(SPS-SBA-15)

was

prepared

by

polymerization of styrene (80 mol%) and divinylbenzene (Aldrich, 98%) (20 mol%) with
a,tf'-azoisobutyronitrile (Aldrich, 98%) (3% relative to total vinyl group) inside SBA-15,
followed by sulfonation using concentrated H2SO4, according to method adopted by Choi
et al. (Choi et a l , 2005). In a typical synthesis, 1 g of SBA-15 was impregnated with
0.250 ml of styrene (80 mol%) and dissolved with 0.078 ml of divinylbenzene (20 mol%)
and 0.016 g (3% relative to total vinyl group) of a,a'-azoisobutyronitrile in 1.50 ml of
dichloromethane. After impregnation with the solution, the silica sample was dried at
40 C for 2 h to remove dichloromethane and was heated for 6 h at 60 C for
polymerization. The resultant sample was washed with chloroform and ethanol and dried
at 80 C. The polystyrene-SBA-15 samples were sulfonated in concentrated H2SO4 at
100 C for 6 h. After extensive washing with distilled water, samples were vacuum-dried
at room temperature overnight.
The synthesis method of aluminum chloride grafted SBA-15 or SBA-15 type PMO
was derived from the one described by Dube et al. (Dube et a l , 2005). 2.5 g of SBA-15
or PMO was dried at 400 C under vacuum (overnight) and kept under dry argon in a
Schlenk tube. The dried SBA-15 or PMO was added to 100 ml of benzene (sodium dried)
in the reaction tube under a dry argon atmosphere. To this solution was added 1.5 g of
77

aluminum chloride stored in a Schlenk tube. The mixture was refluxed for 3 h under dry
argon and then filtered and washed three times in a Schlenk tube with dry benzene (100
ml). The grafted SBA-15 or PMO was stored in a Schlenk tube under a dry argon
atmosphere and denoted with Al-SBA-15 or Al-PMO.
Other commercial acid catalysts: Nafion membrane, Nafion-Si02 composite (SAC13), Zeolite Y (Si/Al=10.5) and Amberlyst-15 were supplied by Sigma-Aldrich. Nafion
were pretreated with 0.5 M H2SO4, washed with deionized water and dried at 70 C under
vacuum over night. Zeolite Y was heated at 500 C for 12 h before use.

4.2.2. Catalyst characterization


Nitrogen adsorption measurements were performed using a Coulter Omnisorp 100
gas analyzer. The BET specific surface areas were evaluated from data in the relative
pressure range [0.01 - 0.03]. The pore diameter was estimated from the peak position of
BJH pore size distribution.
Powder XRD patterns were recorded using a Bruker D4 X-ray diffractometer with
nickel-filtered CuKa radiation.
Transmission electron micrographs (TEM) were obtained using a JEOL 2011
microscope operated at 200 kV.
Solid-state 29Si and 27A1 MAS NMR spectra were recorded at room temperature on a
Bruker ASX 300 spectrometer. Tetramethylsilane (TMS) was used as external reference
7Q

for

7Q

Si MAS NMR analyses.

Si were recorded at a frequency of 59.6 MHz and were

spun at 5 kHz. 27A1 MAS NMR were recorded at 78.2 MHz and were spun at 5 kHz.
The sulfur content was determined using a LECO S144DR instrument. The
concentration of surface sulfonic acid groups was determined by titration with 0.1 M
sodium hydroxide solution.
FT-IR spectra were recorded on a BIO-RAD FTS-60 FTIR spectrometer. The
weight loss curves (TGA) of the catalysts were recorded on a Perkin Elmer 7 series
78

thermal analysis system.

4.2.3. Catalytic performance


Liquid phase alkylation experiments were carried out in an automated stainless steel
fixed bed continuous reactor. The powder samples were pelletized, crushed and sieved.
Nafion membrane and SAC-13 were cut and sieved. Particles from 0.42 to 0.84 mm in
size were selected. In each run, 0.35 g of catalyst was used. The feedstock consisting of
an isobutane/1-butne mixture with a molar ratio of 40/1, containing 1 wt% of nonane as
internal standard was delivered to the reactor using a mass flow controller for liquids
from a storage vessel pressurized at 3.5 Mpa with nitrogen.
Samples were collected and stored using a 16-loop sampling valve (Valco), they
were analyzed using a GC equipped with a 60-m capillary column (Wcot fused silica)
and a FID detector.

4.3. Results and discussion


4.3.1. Catalyst characterization
4.3.1.1 Arene-sulfonic and propyl-sulfonic acid functionalized materials
Nitrogen adsorption-desorption isotherms of the arene-sulfonic and propyl-sulfonic
acid functionalized materials were generally of type IV, with clear hysteresis loops
associated with capillary condensation in the mesopores and with regular pore sizes. A
typical isotherm of ArS-PMO is given in Figure 16; structural properties are summarized
in Table 5. The BET surface areas ranged from 490 to 860 m2/g. The corresponding pore
size distributions were narrow and centered around 2.0-5.1 nm. The incorporation of the
sulfonic acid groups was confirmed and measured by means of titration. Powder
diffraction analysis of ArS-PMO, ArS-SBA-15 revealed highly ordered structures with a
single prominent peak centered at 26 = 0.89 and 0.84, respectively (Fig. 17). As for
PrS-PMO, the structure is less ordered because of the H2O2 added. These nitrogen
adsorption-desorption isotherms and XRD patterns indicate that the synthesized arene79

sulfonic and propyl-sulfonic acid functionalized PMO and SBA-15 are mesostructured
materials. Sulfonic acid groups were grafted on these materials. Compared with propylsulfonic acid functionalized PMO made by postoxidation of anchored mercaptopropyl
groups, arene-sulfonic acid functionalized materials synthesized directly are catalysts
with better mesostructure and higher acid loading.
400
5? 300
rn
T3

CD

200

O
tfl

-o

100

C3

_3
O

>

0.2

0.4

0.8

0.6
P/P0

Figure 16. N2 adsorption isotherms and BJH pore diameter distribution for ArS-PMO

Table 5. Physico-chemical properties of ArS-PMO, ArS-SBA-15 and PrS-PMO


C value

Acid Capacity

of BET

(mmolHVgcat)

SBET(m2/g)

PD(nm)

P v (cm7g)

ArS-PMO

497

3.3

0.72

100

0.91

ArS-SBA-15

854

5.1

1.25

120

0.99

PrS-PMO

826

2.0

1.39

97

0.43

Catalyst

Acid capacity exchange determined by titration

80

ArS-PMO
ArS-SBA-15
PrS-PMO
0

2 Theta (Degree)

Figure 17. X-ray diffraction pattern of samples


The thermogravimetric profile of ArS-PMO (Fig. 18) shows a desorption peak
centred at 50 C corresponding to the evaporation of ethanol. The peak at 150C is due to
residual P123 (Margolese et a l , 2000; Melero et a l , 2002; Inagaki et a l , 2002). The
decomposition peak of arene-sulfonic acid group is observed at 500 C (Hamoudi et al.,
2004). The peak at 710 C, which is also found in unfunctionalized ethane-silica material,
is due to decomposition of ethyl group. In the thermogravimetric analysis for ArS-SBA15, two peaks are observed at 40 C and 560 C, corresponding to evaporation of ethanol
and decomposition of arylsufonic acid group (Fig. 19). The results confirm that sulfonic
acid groups were grafted on acid functionalized materials and these materials are stable
up to 400 C.

30

130

230

330

430

530

630

730

Temperature ( C )

Figure 18. Thermogravimetric weight loss curve for ArS-PMO.


81

100

DT6

86

,1

1-

85

rsid ua th anol

100

200

arene-sulfo nie \
acid group

' 300

400

500

TG

600

700

800

Temperature ( C)

Figure 19. Thermogravimetric weight loss curve for ArS-SBA-15.


4.3.1.2. Aluminum chloride grafted SBA-15 and PMO
Al-SBA-15 and Al-PMO exhibit similar XRD patterns to SBA-15, show type IV N2adsorption-desorption isotherms, indicating the mesoporous structure is preserved after
the addition of aluminum chloride. Figure 20 is N2 adsorption-desorption isotherms and
BJH pore diameter distribution for Al-PMO. The textural and compositional properties of
the catalysts are presented in Table 6.

800

CO

s
-a

>

600

400

..-~*

200
/

0
0.0

0.2

0.4

0.6

0.8

1.0

P/PO

Figure 20. N2 adsorption isotherms and BJH pore diameter distribution for Al-PMO

82

Table 6. Textural and compositional properties of Al-PMO and Al-SBA-15


Al Content

Catalyst

SBET<m2/g)

PD(nm)

Al-PMO

259

5.5

2.85

Al-SBA-15

244

5.5

2.62

(mmol/gcat)

The acidity of the catalysts was characterized using FTIR spectra of adsorbed
pyridine (pyridine-FTIR). Spectra of Al-PMO and Al-SBA-15 show two bands (1456
cm"1 and 1496 cm"') attributed to pyridine adsorbed on Lewis acids sites (Figure 21).
None of these two samples display band at 1545 cm"1 which is attributed to pyridine
adsorbed on Bronsted acids sites (Ghanbari-Siahkali et a l , 2001). It can be concluded
that the predominant acid species are Lewis acid sites on Al-PMO and Al-SBA-15.

2.2
3
c

1.7

o
CJ

1.2
sO

ft
<

_Q

0.7
0.2

1550

1500

1450

1400

Wavenumber(cm-1 )

Figure 21. Pyridine FTIR spectra (after desorption at 120 C) of Al-PMO and Al-SBA-15
Al-PMO and Al-SBA-15 show similar 27A1 MAS NMR spectra. Figure 22 gives the
spectrum of Al-PMO and the deconvolution curves. The ascription of peaks is shown in
Figure 23. The peak at 0 ppm is attributed to six-coordinate Al species. It appears from
air contamination when the samples were placed in an NMR cell (Sato and Maciel, 1995).
The peaks indicate that the aluminum chloride is grafted on the surface of PMO and
83

SBA-15. The grafting of electron-withdrawing AICI3 in the structure makes Al-PMO and
Al-SBA-15 to be strong Lewis acids.

85ppm
65ppm

fif

</>
C

70

120

20

-80

-30

ppm
27

Figure 22. Z,A1 MAS NMR spectra of Al-PMO.

OH

HO.

HO>A1

/\

O O
Si

OH

OH

^ A l OH

/\

O O
Si
Si
Oppm

a
"

/l\

HOOOH
Si

Si

Si Si
36ppm

Al

Al

Al

l\

/l\

/ \

O O H O O C

Si Si
65ppm

Si

Si Si
75ppm

Si Si
85ppm

27,
Figure 23. Structures of A1C13 grafted on PMO with the "Al
MAS NMR chemical shifts

4.3.1.3. Sulfonated polystyrene coated SBA-15 (SPS-SBA-15)


Sulfonated polystyrene coated SBA-15 (SPS-SBA-15) exhibits XRD patterns
similar to SBA-15 and type IV N2-adsorption-desorption

isotherms, indicating

mesoporous structure of SBA-15 is preserved after coating of polystyrene and sulfonation.


The surface area and pore size decrease after coating by polystyrene. The physicochemical properties of SBA-15, polystyrene coated SBA-15 (PSSBA-15) and SPS-SBA15 are presented in Table 7.
84

Table 7. Physico-chemical properties of SBA-15, PSSBA-15 and SPS-SBA-15

Catalyst

SBET<m /g)

PD(nm)

,
Pv(cm3/g)

SBA-15

770

7.7

0.94

PSSBA-15

404

6.9

0.55

SPS-SBA-15

331

5.0

0.45

Acid Capacity
(mmolH /gcat)

1.0

4.3.1.4. Perfluoroalkylsulfonic acid functionalized SBA-15 and PMO


XRD patterns, N2-adsorption-desorption isotherms and transmission electron
microscopy (TEM) images of FS-SBA-15 and FS-PMO are similar and reveal that both
of them have the characteristic of 2-d hexagonal (P6mm) structure. Figure 24 gives the
powder X-ray diffraction (XRD) pattern and (TEM) images of FS-PMO. Diffraction
peaks in XRD pattern and regular hexagonal arrangement in TEM images indicate that
the long range ordering of material

is not affected

by the anchoring of

perfluoroalkylsulfonic acid groups. Nitrogen adsorption-desorption isotherms for the FSPMO are generally of Type IV (Figure 25), indicating the mesoporous nature of the
solids.
iffjsmjjjiis
3
CD

:
.

(fl
c

>
ro
ro
o;

01

!
/

l_

2 3 4 5 6 7 8 9

Hfek
10

Hffc

2-Tneta(Degree)

Figure 24. XRD pattern and TEM images for FS-PMO


85

Figure 25. N2 adsorption isotherms and BJH pore diameter distribution for FS-PMO
The BET surface area and pore diameter of the sample are given in Table 8. The C
value of BET is 97 on FS-PMO, while it is only 54 on FS-SBA-15. It indicates that the
adsorption of N2 on FS-PMO is stronger than on FS-SBA-15. The amount of
perfluoroalkylsulfonic acid groups present in the solids was determined by analyzing the
S element concentration. Though same and excessive amount of THTS (l,2,2-trifluoro-2hydroxy-1 -trifluoromethylethane sulfonic acid /?-sultone) was used to synthesize FSSBA-15 and FS-PMO, the maximum sulfur loading was 1.10 mmol/g on FS-PMO and
only 0.42 mmol/g on FS-SBA-15. Compared to mesoporous silicas, more terminal
perfluoroalkylsulfonic acid groups were grafted on PMO materials.
Table 8. Physico-chemical properties of FS-PMO and FS-SBA-15
C value of

S content

BET

(mmol/gcat)

Catalyst

SBET(m2/g)

PD(nm)

Pv(cm3/g)

FS-PMO

539

5.6

1.0

97

1.10

FS-SBA-15

359

4.6

0.6

54

0.42

The presence of ethane carbon atoms covalently linked to Si in the framework of


86

FS-PMO can be attested by 29Si solid state MAS NMR (Fig 26). The signals at -64.0 ppm
and -57.0 ppm can be attributable to T3 [RSi(OSi)3] and T2 [RSi(OSi)2OH] resonance
respectively according to literature (Inagaki et a l , 1999; Hamoudi and Kaliaguine, 2003;
Hamoudi et a i , 2004). There is no occurrence of Q" [Si(OSi)(OH)4_] resonances within
the detection limit of the NMR analysis which reveals that essentially all silicon atoms
were covalently bound to carbon atoms and the synthesized FS-PMO is indeed a material
of periodic mesoporous organosilica material.

ft
5
c

-120

29

Figure 26. zySi MAS NMR spectrum of the FS-PMO


A FT-IR spectrum of FS-PMO is illustrated in Figure 27. Two bands at 1415 cm"1
and 1780 cm"1 were observed. The band at 1415 cm"1 is characteristic of v (S-O)
stretching vibrations of undissociated -SO3H groups previously reported for Nafion films
and triflic acid (Buzzoni et a l , 1995; Zecchina et a l , 1996; Alvaro et a l , 2004). The
other band at 1780 cm"1 can be assigned to the vibrations of protonated water molecules
according to the IR studies on Nafion. FT-IR spectra could indicate that the
perfluoroalkylsulfonic acid is grafted on PMO by covalent attachment. These bands
undergo some shift to higher wavenumbers at higher temperature as observed on
"Nafion"-functionalized MCM-41 (Alvaro et a l , 2005). The intensity of the two bands
was maintained at 150 C indicating thermal stability in the reaction conditions of
alkylation (reaction temperature is lower than 150 C).
87

1415 cm"1
1780 cm-1

1J

=5

A.
A.

etf

150C

o
X)

30C

<

00

2100

1900

>
\

1500

1700

131

Wavelength(crrr! )

Figure 27. FT-IR spectra of FS-PMO sample heated at different temperatures

1450

Figure 28. F TIR spectra of F S-PMO sample evacuated at 100 C before pyridine
adsorption (a), after pyridine adsorption and evacuation at 35 C (b), 50C (c) and 75 C
(d).
88

The kind of acid sites of the FS-PMO was characterized using FTIR spectra of
adsorbed pyridine (pyridine-FTIR). Spectra of the FS-PMO evacuated at 35 C, 50 C
and 75 C show 2 bands (1490 and 1550 cm"1) attributed to pyridine adsorbed on
Bronsted acid sites (Figure 28). The band at 1550 cm"1 is attributed to Bronsted acid sites
and the band at 1490 cm'1 is may belong to both Bronsted and Lewis acid sites
(Ghanbari-Siahkali et a l , 2001). The absence of any line in the 1450 cm"1 area indicates
complete absence of Lewis acid sites. The band at 1440 cm"1 observed before desorption
is associated with physisorbed pyridine. The non-functionalized PMOs did not display
significant pyridine FTIR signals (not shown), indicating that these PMOs have no acidic
properties. For perfluorosulfonic acid functionalized catalyst, the only acid species are
Bronsted acid sites.

4.3.2. Catalytic performance


The synthesized acid functionalized mesoporous materials and certain commercial
solid acids were tested in the liquid phase alkylation of isobutane by 1-butne. The
catalysts used and their textural properties are presented in Table 9. In literature, zeolites
received much attention as solid acid catalysts for this alkylation (Corma et a l , 1993;
Feller and Lercher, 2004). In order to compare with zeolitic catalysts, a zeolite Y sample
with Si/Al ratio of 10.5 was also tested in alkylation.
Alkylation of isobutane with butne is dramatically influenced by the strength of
acid sites. The different reaction steps in alkylation require different minimum acid
strengths to be effectively catalyzed. Double bond isomerization is catalyzed already by
weak acid sites, while hydride transfer is only catalyzed by very strong acid sites. Unlike
liquid acids which have a well defined acid strength with a given composition, zeolites
encompass Bronsted and Lewis acid sites with a wide range of strengths. It is difficult to
clarify the correlations between the acidity and alkylation performance from zeolite
studies. The synthesized acid functionalized materials have essentially one type of acid
with similar strengths. The acid strength can be tuned by the presence of electrowithdrawing species close to the sulfonic group. For example, the electro-withdrawing
fluorine atoms close to the sulfonic group make perfluorinated sulfonic acid a very strong
89

acid with Ho comparable to sulfuric acid. These materials are suitable solid acids to study
the influence of acid strength, structure and polarity of surface on the alkylation
performance
Table 9. Catalysts used in alkylation reaction and their textural properities
Catalyst

SBET(m2/g)

PD(nm)

Acid Capacity
(mmolH+/gSi02)

PrS-PMO

826

4.0

0.41

ArS-PMO

497

6.6

0.89

ArS-SBA-15

854

10.2

1.01

SPS-SBA-15

331

5.0

0.73

Amberlyst-15

45

25.0

4.90

Al-PMO

160

4.5

3.01

Al-SBA-15

334

4.5

2.67

Nafion

0.01

0.94

SAC-13

200

0.14

FS-PMO

539

5.6

1.10

FS-SBA-15

359

4.6

0.42

H-Y

600

0.7

G.D. Stucky and coworkers compared the relative acid strengths of sulfonicftmctionalized SBA-15 with other acid catalysts by

31

P MAS NMR measurements of

chemically adsorbed triethylphosphine oxide (Margolese et al., 2000). The acid strength
decreases in the order: USY > Amberlyst-15 > S03H/(SBA-15 containing organic
groups) > S03H/SBA-15 > A1-MCM-41>SBA-15. The acid strength of catalysts used in
this work ascends in sequence: PrS-PMO < ArS-SBA-15 < ArS-PMO, SPS-SBA-15 <
Amberlyst-15 < AICI3-PMO (A1C13-SBA-15) < Zeolite Y < Nafion (SAC-13, FS-PMO,
FS-SBA-15). The initial (TOS = 1 min) alkylation activity and selectivity over these
materials are shown in Table 10.

90

Table 10. Initial (TOS = 1.0 min) butnes conversion and product distribution
obtained on different catalysts
Catalysts

*Conv.(%)

C5-7/C5+

TMP/C 5+

DMH/C5+

" s - , other is-,


*-8
'*-5+

C9+/C5+

PrS-PMO"

ArS-SBA-15 b

ArS-PMO"

SPS-SBA-15b

Amberlyst-15

24

14

34

52

Al-PMO"

100

34

27

26

Al-SBA-15"

100

20

62

0.5

10

Nafion3

21

15

38

21

22

SAC-13a

65

29

34

21

13

FS-PMOa

100

32

48

14

FS-SBA-153

99

30

40

20

Zeolite Y 3

100

24

50

14

12

T = 343K for a and 373 for b, Olefin WHSV = 2 h"1, /C4/l-C4= = molar ratio of 40
*Conv.(%): Disappearance of 1-butne and its isomers
"pother . C s

other than T M p and

D M R

Double bond isomerization is catalyzed by weak acid sites. Thus the conversion is
calculated based on disappearance of butnes (1-butne and its isomers). Blank tests
performed under identical conditions without any catalyst showed negligible conversion
of 1-butne. With PrS-PMO, ArS-PMO, ArS-SBA-15 and SPS-SBA-15, the reaction
only showed activity of isomerization of 1-butne. No product of C5+ was achieved
(Table 10). It can be concluded that the acid strength of alkylsulfonic and arene-sulfonic
acid functionalized materials is not high enough to catalyze alkene addition and hydride
transfer. Amberlyst-15 is a sulfonic resin with higher acid strength than arene-sulfonic
acid functionalized mesoporous materials (Margolese et a l , 2000). It gives C5+ products
91

when used as catalyst for alkylation of isobutane/1-butne. The products are composed of
C4~ isomers, dimethyl hexane (DMH), octene and heavy-end hydrocarbons. They are the
products of isomerization, dimerization and oligomerization. No TMPs being obtained
indicated that the acidity of Amberlyst-15 is not strong enough to catalyze the alkylation
reaction. AICI3 grafted SBA-15 and PMO materials are catalysts with very strong Lewis
acid sites. No Bronsted acid sites were detected by pyridine-FTIR analysis. Both Al-PMO
and Al-SBA-15 exhibited 100% initial butne conversion. That means strong Lewis acid
sites also catalyze the alkylation reaction, in contradiction with the results of DiazMendoza (Diaz-Mendoza et a l , 1998). These authors thought Lewis acid sites of solid
acid catalysts do not catalyze the alkylation reaction, the presence of strong Lewis acid
sites just promotes the formation of unsaturated compounds. AICI3 grafted SBA-15 and
PMO materials are indeed catalysts with good alkylation activity. However, these
materials are very sensitive to moisture. When AICI3 grafted materials are exposed to air,
the acid strength of these materials decreases dramatically. The exposed Al-PMO and AlSBA-15 even do not catalyze isomerization reaction.
Nafion is an acidic solid with acid strength comparable to that of sulfuric acid. It
gives C5.7 and TMPs which are products of cracking and alkylation. However, the
activity of Nafion is not as good as that of zeolite Y. Nafion produces more octenes and
heavy-end products. The very small surface area of Nafion (0.01 m2/g) strongly limits the
number of available acid sites. The accessibility of the acid sites in Nafion can be
increased by entrapping nanosized particles of Nafion resin within a highly porous silica
network. SAC-13 is a commercial Nafion resin/silica nanocomposites with 13-16 wt% of
Nafion entrapped within a silica porous network. It is much more active than Nafion
granule though the content of Nafion is only 13-16 wt%. This indicates that the acid
strength is the factor determining the alkylation activity while highly porous network and
accessible acid sites are important for good catalytic performance. Perfluorinated sulfonic
acid functionalized mesostructured materials are solid acids with acid strength similar to
SAC-13. Compared with SAC-13, FS-SBA-15 and FS-PMO exhibited much better initial
activity. The high activity could be attributed to their meso-structure, large surface area
and high acid density.
92

10

15
TOS(min)

20

25

Figure 29. Butnes conversion with TOS in the alkylation of isobutane/1-butne.


(Reaction condition: T = 343K; olefin WHSV = 2 h"1; isobutane/1-butne molar ratio =
40.)
Figure 29 shows butnes conversion with time on stream in the alkylation of
isobutane/1-butne over perfluorinated sulfonic acid functionalized materials. Fast
deactivation on catalysts can be observed within 25 min except on FS-PMO. Both FSSBA-15 and FS-PMO have meso-structure and large surface area. The activity of FSPMO is very stable while that of FS-SBA-15 decays obviously. The good catalytic
performance of FS-PMO can therefore be ascribed to its carbophilic surface and higher
acid loading. Compared with zeolite Y, FS-PMO shows better and stable activity for
alkylation of isobutane/1-butne. The conversion of butne reached 100% over 80 min
times on stream (Figure 30) while that of zeolite is less than 5 min in the same condition.
FS-PMO is a much better catalyst for isobutane/-butene alkylation compared with
conventional solid acid catalysts. The results indicate that acid strength of the catalyst
plays a key role on alkylation activity while large surface area, large pore size and
hydrophobic surface of catalyst promote its activity and stability.

93

o-ocoooo-o-o-ooo

20

40

60

80

100

120

TO.S.[min]
Figure 30. Butnes conversion and distribution of Cs+ with time on stream on FS-PMO
(Reaction condition: T = 343 K; olefin WHSV = 2 h"1; isobutane/1-butne molar ratio =
40.)
It was also observed that the morphology of materials affects catalytic performance.
Only well pelletized and sieved FS-SBA-15 and FS-PMO exhibit good activity, very low
activity is achieved when powder of the same materials is used. It is important to load
well pelletized and sieved particles in the reactor. This may be the main reason of abrupt
deactivation of FS-PMO at 80 min times on stream. The reason is still under investigation.

4.4. Conclusion
A series of acid functionalized mesostructured materials such as propyl-sulfonic,
arene-sulfonic and perfluoroalkylsulfonic

acid functionalized SBA-15 and PMO,

sulfonated polystyrene coated SBA-15 and PMO, aluminum chloride grafted SBA-15 and
PMO were synthesized. All these materials have the characteristic of 2-d hexagonal
(P6mm) structure with good textural uniformity. Acid groups were grafted on the pore
surface. Nitrogen adsorption-desorption isotherms for these materials are of Type IV,
indicating the mesoporous nature of the solids. The acid strength increases in the order:
PrS-PMO < ArS-SBA-15 < ArS-PMO, SPS-SBA-15 < Amberlyst-15 < AlCl3-PMO
(AICI3-SBA-I5)

<

Zeolite

<

Nafion

(SAC-13,

FS-PMO,

FS-SBA-15).
94

Perfluoroalkylsulfonic acid functionalized materials are solid acids with acid strength
comparable to that of concentrated sulfuric acid. The predominant acid species on
sulfonic acid functionalized materials are Bronsted acid sites while the predominant acid
species on Al-PMO and Al-SBA-15 are Lewis acid sites.
The synthesized acid functionalized mesoporous materials were evaluated as
catalysts in the alkylation of isobutane/1-butne. The acid strength of the catalysts is the
factor determining the alkylation activity. Only the catalysts with acid strength higher
than that of Amberlyst-15 show activity of alkylation. Not only the Bronsted acidity has
activity of alkylation, strong Lewis acidity also shows good alkylation activity. Large
surface area and high acid loading are very important for alkylation. FS-PMO exhibits the
best catalytic activity in the alkylation of isobutane by 1-butne among the various
catalysts tested in this work. Its high and stable activity can be attributed to its high acid
strength, large surface area, large pore size and carbophilic nature of surface.

95

5. Alkylation of Isobutane/1 -butne over Periodic Mesoporous


Organosilica Functionalized with Perfluoroalkylsulfonic Acid
Group
Wei Shen a b , David Dub b, Serge Kaliaguine b *
"Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Department
of Chemistry, Fudan University, Shanghai 200433, PR China
Department of Chemical Engineering, Laval University, Pavillon A-Pouliot, Sainte Foy,
Canada G1K 7P4
Published in Catalysis Communications 10 (2008) 291-294

Abstract
A periodic mesoporous organosilica (PMO) has been successfully functionalized
with perfluoroalkylsulfonic acid groups by reacting the PMO with 1,2,2-trifluoro-2hydroxy-1 -trifluoromethylethane sulfonic acid /?-sultone. The resulting material exhibits
much higher catalytic activity in the alkylation of isobutane by 1-butne than
perfluoroalkylsulfonic acid functionalized SBA-15, alkylsulfonic acid functionalized
PMO and zeolite Y.

5.1. Introduction
Alkylation of isobutane/butene is a refining operation which produces mixtures of
branched alkanes from the C3 and C4 hydrocarbons produced in cracking and steam
cracking units. The formed product, designated as alkylate, is an excellent blending
component for gasoline because of its high octane number, low Reid vapor pressure, and
freedom from aromatics or olefins [1]. The current alkylation technology makes use of
strong liquid acids (HF or H2SO4). The obvious environmental risks of an unintentional
release upon separation and disposal of these catalysts is a strong motivation for the
petroleum industry to develop a heterogeneous catalyst. Many attempts have been made
96

to use strong acidic solids including sulfated zirconia and related materials [2, 3],
heteropolyacids [4, 5], acid resins [6, 7], chlorinated alumina [8], and acid zeolites [9-17]
as catalysts of this reaction. None of them met any commercial success because of their
unacceptably rapid deactivation. During zeolite-catalyzed isobutane/butene alkylation,
which almost exclusively produces isoalkane products, a highly unsaturated and highly
branched polymer is formed. The polymer is strongly adsorbed on the acid sites and
completely fills the pores at the end of the reaction [18]. The alkylation reaction proceeds
mainly via addition of w-butene to an isobutyl carbenium ion. The resulting octyl
carbenium ion is removed (after possible isomerization) from the active site by hydride
transfer from isobutane leading to trimethylpentanes as ideal products and an isobutyl ion,
which perpetuates the reaction. The high ratio of the rate of hydride transfer vs. the rate of
oligomerization is crucial for good catalytic performance. Strong acid sites are necessary
to effectively catalyze hydride transfer [19]. The electrostatic field in the zeolite pores
enhances the adsorption of polarizable molecules [20]. The preferential adsorption of
alkene on zeolite surface leads to high rate of oligomerization and the small pore size of
zeolite limits the diffusion of reactants to active sites and products out of the pore system.
Thus, a catalyst with large pore size which provides enough space for diffusion, strong
acid strength comparable to that of sulfuric acid, and nonpolar surface which allows a
higher paraffin steady concentration in the pores is desired.
The newly discovered periodic mesoporous organosilicas (PMOs) have opened a
new area in the development of solid catalysts [21, 22]. PMOs are hybrid organicinorganic materials containing siloxane moieties bridged by organic group with well
ordered mesopores. The presence of organic groups within the framework is expected to
modify the hydrophilic/carbophilic character of the PMOs. The carbophilicity of the
organic moiety of the walls will allow a higher paraffin steady concentration under
reaction conditions and increase the rate ratio of hydride transfer vs. oligomerization.
Syntheses and applications of different types of PMOs functionalized with alkylsulfonic
and arene sulfonic acid groups have been reported in recent years [23-25].
Perfluorinated sulfonic-acid resins such as Nafion have been used as catalysts in a
wide range of organic reactions. The presence of electron-withdrawing fluorine atoms in
97

the structure significantly increases the acid strength of the terminal sulfonic-acid groups,
which becomes comparable to that of pure sulfuric acid. However, Nafion presents low
surface areas. Supporting Nafion on high surface area carriers or incorporating Nafion
into silica using a sol-gel technique provided ways to overcome this limitation [6, 7].
These materials are far from optimum due to the imperfect dispersion of the resin within
the silica pores. Mcquarrie et al. immobilised perfluorinated sulfonic acids into
mesoporous silica frameworks by a sol-gel templated synthesis [26]. Corma et al.
anchored perfluoroalkylsulfonic acid groups over MCM-41 and SBA-15 solids by
grafting the precursor (1,2, 2-trifluoro-2-hydroxy-l-trifluoromethylethane sulfonic acid
/?-sultone 1) over the silica surface [27]. However, there is no report on the perfluorinated
sulfonic acid functionalized PMOs and alkylation of isobutene/1-butne over acid
functionalized PMOs. In this communication, we report for the first time the synthesis of
perfluoroalkylsulfonic acid functionalized PMO (FS-PMO) and their application in
alkylation of isobutane/1 -butne.

5.2. Experimental
5.2.1. Catalyst preparation
The parent PMO material was synthesized by the acid-catalyzed hydrolysis and
condensation of bis-(trimethoxysilyl) ethane (BTME, Aldrich) using the triblock
copolymer Pluronic PI23 (EO20PO70EO20, BASF) as the structure-directing agent,
according to the procedure adopted by Muth et al. [28]. After synthesis, the triblock
copolymer was removed by extraction with ethanol. In a typical synthesis, 2 ml of BTME
was added to a mixture of 1.5 g of Pluronic P123, 15 ml of 12.2 M HC1, and 75 ml of
deionized water under stirring at 40 C. After the mixture was stirred for 24 h, the
resulting gel was cured at 90 C for an additional 24 h. The obtained white precipitate
was separated by filtration, washed thoroughly with deionized water and dried at room
temperature. The surfactant was removed by solvent extraction with anhydrous ethanol in
a Soxhlet apparatus for 48 h.
The PMO was functionalized with perfluoroalkylsulfonic acid group by grafting 1
98

(Synquest) on its surface (Scheme 1). 2.0 g of solvent extracted PMO was pretreated at
120 C in vacuum for 12 h and cooled down to room temperature. 50 ml dry toluene and
2.0 g of 1 were then added. The mixture was refluxed for 4 h under Ar atmosphere. The
resulting solid was filtered, washed thoroughly with toluene and dried at 100 C
overnight. In order to compare the catalytic activity, perfluoroalkylsulfonic acid group
functionalized SBA-15 (FS-SBA-15) [27] and propylsulfonic acid group functionalized
PMO (PS-PMO) [23] were also prepared.

5.2.2. Catalyst characterization


Nitrogen adsorption measurements were performed using a Coulter Omnisorp 100
gas analyzer. The BET specific surface areas were evaluated from data in the relative
pressure range [0.01 - 0.03]. The pore diameter was estimated from the peak position of
BJH pore size distribution.
Powder XRD spectra were recorded using a German Bruker D4 X-ray
diffractometer with nickel-filtered CuKa radiation.
Transmission electron micrographs (TEM) were obtained using a JEOL 2011
microscope operated at 200 kV.
Solid-state 29Si MAS NMR spectra were recorded at room temperature on a Bruker
ASX 300 spectrometer. Tetramethylsilane (TMS) was used as external reference for 29Si
MAS NMR analyses.
The sulfur content was determined using a LECO S144DR instrument.
FT-IR spectra were recorded on a BIO-RAD FTS-60 FTIR spectrometer.

5.2.3. Catalyst tests


Liquid phase alkylation experiments were carried out in an automated stainless steel
fixed bed continuous reactor. In each run, 0.35 g of catalyst was used. The feedstock
consisting of an isobutane/1-butne mixture with a molar ratio of 40/1, containing 1 wt%
99

of nonane as internal standard was delivered to the reactor using a mass flow controller
for liquids from a storage vessel pressurized at 3.5 Mpa with nitrogen.
Samples were collected and stored using a 16-loop sampling valve, they were
analyzed using a GC equipped with a 60-m capillary column (Wcot fused silica) and a
FID detector.

5.3. Results and discussion


5.3.1. Catalyst characterization
Figure 32 shows the powder X-ray diffraction (XRD) pattern and transmission
electron microscopy (TEM) images of FS-PMO. Diffraction peaks in XRD pattern and
regular hexagonal arrangement in TEM images reveal that the synthesized material has
the characteristic of 2-d hexagonal (P6mm) structure with good textural uniformity,
which indicates that the long range ordering of PMO is not affected by the anchoring of
perfluoroalkylsulfonic acid groups. Energy-dispersive X-ray analysis showed that
perfluoroalkylsulfonic acid groups were incorporated in the framework. Nitrogen
adsorption-desorption isotherms for the FS-PMO are generally of Type IV, indicating the
mesoporous nature of the solids. The BET surface area and pore diameter of the sample
are given in Table 11.
Figure 33 shows the 29Si solid state MAS NMR spectrum of the FS-PMO material
and the deconvolution curves. The signals at -64.0 ppm and -57.0 ppm attributable to T3
[RSi(OSi)3] and T2 [RSi(OSi)20H] resonance respectively according to literature [21, 23,
24]. There is no occurrence of Q" [Si(OSi)(OH)4_] resonances within the detection limit
of the NMR analysis, which attests that essentially all silicon atoms were covalently
bound to carbon atoms and the synthesized FS-PMO is a material of periodic mesoporous
organosilicas.
Covalent attachment of the perfluoroalkylsulfonic acid chains to the PMO surface
can be confirmed by the FT-IR spectra of the self-supporting wafer of the FS-PMO. Two
bands at 1415 cm"1 and 1780 cm"1 were observed. The band at 1415 cm"1 is characteristic
100

of v (S-O) stretching vibrations of undissociated -SO3H groups previously reported for


Nafion films and triflic acid. The other band at 1780 cm" can be assigned to the
vibrations of protonated water molecules according to the IR studies on Nafion [27, 29,
30]. These bands undergo some shift to higher wavenumbers as observed on "Nafion"fiinctionalized MCM-41 [31]. The intensity of the band was maintained at 150 C
indicating its thermal stability in the reaction condition of alkylation (T lower than 150
C).
The amount of perfluoroalkylsulfonic acid groups present in the solids was
determined by analyzing the S element concentration. Corma et al. anchored
perfluoroalkylsulfonic acid groups over MCM-41 and SBA-15 solids by grafting 1 over
the silica surface [27]. Though excessive amount of 1 was added, Hybrid materials with
maximum acid capacities of 0.5 mmol of H+ per gram of solid were obtained. In our
experiments, the maximum sulfur loading was 1.10 mmol/g on FS-PMO and only 0.42
mmol/g on FS-SBA-15 (Table 11), which indicates that the organic moiety of the walls
improves the reaction between 1 and the surface silanol groups and that the carbophilic
surface environment is beneficial to grafting of terminal perfluoroalkylsulfonic acid
groups to the framework.

5.3.2. Catalytic performance


The perfluoroalkylsulfonic acid functionalized materials FS-PMO, FS-SBA-15,
Nafion and propylsulfonic acid functionalized PMO (PS-PMO) were tested in the liquid
phase alkylation of isobutane by 1 -butne. In literature, zeolites received much attention
as solid acid catalysts for this alkylation [1, 9-17]. In order to compare with zeolitic
catalysts, a zeolite Y sample with Si/Al ratio of 10.5 was used in alkylation. The physical
properties of the catalysts are listed in Table 11.
Double bond isomerization is catalyzed by weak acid sites. Even a fully deactivated
zeolite retains some activity for isomerizing butnes. Thus the conversion is calculated
based on disappearance of butnes. Blank reaction performed under identical conditions
without any catalyst showed negligible conversion of 1-butne. With PS-PMO, the
reaction only showed activity of isomerization of 1-butne. No product of alkylation was
101

achieved (Figure 34). It can be concluded that the acid strength of alkylsulfonic acid
grafted PMO is not high enough to catalyze alkene addition and hydride transfer. Nafion
and FS-SBA are active for alkylation but with relatively low conversion of butnes. The
activity can be attributed to their very strong acid sites. The very low surface area of
Nafion and low acid loading of FS-SBA-15 limit their activity. In the case of FS-PMO,
the conversion of butnes reached 100% over 20 min time on stream. Both Nafion and
FS-PMO have high strength acid sites and carbophilic surface. The high activity of FSPMO could be attributed to its meso-structure and large surface area. Both FS-SBA-15
and FS-PMO have however meso-structure and large surface. The high activity of FSPMO can therefore be ascribed to its carbophilic surface and higher acid loading.
Compared with zeolite Y, FS-PMO shows better activity for alkylation of isobutane/1butene. Its high and stable activity may be attributed to its higher acid strength, higher
carbophilicity of surface and larger pore size compared with that of zeolite Y. The XRD
pattern and BET results of the used FS-PMO are similar to those of the fresh one, which
indicates that the structure of FS-PMO is not changed during the reaction.
It was observed that all catalysts are poisoned by water. The catalysts under study all
indeed lost their activity when exposed to water vapor. Thus the hydrophobic
environment of the acid site is beneficial for catalysts to achieve and keep a high and
stable activity. Although the activity of FS-PMO is more stable than that of zeolite Y, it
gradually decreases after 20 min. This is due to the hydrolysis of Si-O-C bond under
catalytic conditions. The S loading of FS-PMO decreases from 1.1 mmol/g to 0.49
mmol/g over 1 hour.
Figure 35 shows the distribution of C5-1- products with time on stream in the
isobutane/1-butne alkylation on FS-PMO catalyst. C5, Ce, C7, TMP (trimethylpentanes),
DMH (dimethylhexanes) and C9+ are obtained. The loss of terminal perfluoroalkylsulfonic acid groups caused by hydrolysis makes it difficult for the catalyst to
transfer hydrogen, thus causing the ratio TMP/C5+, i.e., alkylation/oligomerization ratio,
to shift toward lower values. The concentration of products related to oligomerization,
such as C9+ and C5-C7 which come from cracking of C9+, shift to higher values.

102

This enhanced stability of both the activity and product selectivity compared to Y
zeolite is not without potential commercial interest. Indeed the problem of heterogenizing
this alkylation catalyst has been so significant that technologies have been proposed in
order to make use of even the fast deactivating Y zeolite catalysts. The corresponding
alkylation processes involve either multiple reactors [32, 33] or a transport reactor [34] to
implement fast reaction/regeneration cycling. The frequency of this cycling is of course
dependant on the catalyst deactivation rate so that a deactivation time on stream of some
tens of minutes as illustrated in Figures 34 and 35 would already simplify the cycling
process compared to the few minutes on stream allowed by the zeolite catalysts.

5.4. Conclusion
A high acid strength, carbophilic surface mesostructured material has been
synthesized by grafting perfluoroalkylsulfonic acid groups on PMO. The resulting PMO
material exhibits better catalytic activity as well as catalytic stability compared to
previously reported hybrid organic-inorganic acid catalysts and zeolitic catalysts.
Moreover the work illustrates the effects of the most important parameters which must be
controlled in order to improve the catalyst stability in this critically important catalytic
process.

Reference
[1]

A. Corma, H. Garcia, Chem.Rev.-Sci.Eng. 35 (1993) 483.

[2]

A. Corma, A. Martinez, C. Martinez, J. Catal. 52 (1994) 149.

[3]

A. S. Chellappa, R. C. Miller, W. J. Thomson, Appl. Catal. A 359 (2001) 209.

[4]

T. Blasco, A. Corma, A. Martinez, P. Martinez-Escolano, J. Catal. 306 (1998)


177.

[5]

V. Sarsani, Y. Wang, B. Subramaniam, Ind. Eng. Chem. Res. 44 (2005) 6491.

[6]

P. Botella, A. Corma, J. M. Lopez-Nieto, J. Catal. 371 (1999) 185.


103

[7]

C. J. Lyon, V. S. R. Sarsani, B. Subramaniam, Ind. Eng. Chem. Res. 43 (2004)


4809.

[8]

S. I. Hommeltoft, Appl. Catal. A 421 (2001) 221.

[9]

A. W. Chester, Y. F. Chu, Zeolite 6 (1986) 195.

[10] A. Corma, A. Martinez, C. Martinez, Catal. Lett. 28 (1994) 187.


[11] M. Stocker, H. Mostad, T. Rorvik, Catal. Lett. 28 (1994) 203.
[12] F. Cardona, N. S. Gnep, M. Guisnet, G. Szabo, P. Nascimento, Appl. Catal. A
128(1995)243.
[13] A. Corma, A. Martinez, P. A. Arroyo, J. L. F. Monteiro, E. F. Sousa-Aguiar,
Appl. Catal. A 142 (1996) 139.
[14] D. E. Sherwood, R. J. Taylor, Appl. Catal. A 155 (1997) 195.
[15] G. S. Nivarthy, Y. He, K. Seshan, J. A. Lercher, J. Catal. 176 (1998) 192.
[16] R. Loenders, P. A. Jacobs, J. A. Martens, J. Catal. 176 (1998) 545.
[17] A. Feller, A. Guzman, I. Zuazo, J. A. Lercher, J. Catal. 224 (2004) 80.
[18] A. Feller, J. Barth, A. Guzman, I. Zuazo, J. A. Lercher, J. Catal. 220 (2003) 192.
[19] A. Corma, A. Martinez, C. Martinez, J. Catal. 185 (1994) 146.
[20] S. Ramachandran, T. G. Lenz, W. M. Skiff, A. K. Rappe, J. Phys. Chem. 100
(1996)5898.
[21] S. Inagaki, S. Guan, Y. Fukushima, T. Ohsuna, O. Terasaki, J. Am. Chem. Soc.
121(1999)9611.
[22] D. Trong On, D. Desplantier-Giscard, C. Danumah, S. Kaliaguine, Appl. Catal.
A 222(2001)299.
104

[23] S. Hamoudi, S. Kaliaguine, Micropor. Mesopor. Mater. 59 (2003) 195.


[24] S. Hamoudi, S. Royer, S. Kaliaguine, Micropor. Mesopor. Mater. 71 (2004) 17.
[25] J. A. Melero, R. V. Grieken, G. Morales, Chem. Rev. 106 (2006) 3790.
[26] D. J. Macquarrie, S. J. Tavener, M. A. Harmer, Chem. Commun. (2005) 2363.
[27] M. Alvaro, A. Corma, D. Das, V. Fornes, H. Garcia, Chem. Commun. (2004)
956.
[28] O. Muth, C. Schellbach, M. Frba, Chem. Commun. (2001) 2032.
[29] R. Buzzoni, S. Bordiga, G. Ricchiardi, G. Spoto, A. Zecchina, J. Phys. Chem. 99
(1995)11937.
[30] A. Zecchina, F. Geobaldo, G. Spoto, S. Bordiga, G. Ricchiardi, R. Buzzoni, G.
Petrini, J. Phys. Chem. 100 (1996) 16584.
[31] M. Alvaro, A. Corma, D. Das, V. Fomes, H. Garcia, J. Catal. 231 (2005) 48.
[32] P.J. Nat, E.H. Van Broekhoven, J.W.M. Sonnemans, V.J. D'amico, M.
Mukherjee, US Patent Application US2003/0220529 Al (2003).
[33] E.H. Van Broekhoven, J.W.M. Sonnemans, S. Zuijdendorp, US Pat., 7,176,340
B2 (2007).
[34] W.H. Radcliffe, W.L. Kiel, CD. Gosling, P.A. Sechrist, P. Anderson, US Pat.,
6,814,943 B2 (2004).

105

Table 11. Physical properties of the catalysts in the alkylation of isobutane/1-butene


Sample

SBET/m g"

Dp*/nm

content/mmolg"1
FS-PMO

1.10

539

5.6

FS-SBA-15

0.42

359

4.6

Nafion

0.89

0.02

PS-PMO

0.43

825

4.0

* Pore size distributions were calculated using the desorption branch of the N2
adsorption/desorption isotherms and BJH method.

106

F,C

PMO-

.OH

F '

Figure 31. Preparation of perfluoroalkylsulfonic functionalized PMO by the


grafting technique

107

JJJ3

Cfl

B
g

CD

>

S
CD

Cd

01

2 3 4 5 6 7 8 9

10

2-Tneta(Degree)

Figure 32. XRD pattern and TEM images for FS-PMO

108

29

Figure 33. x*Si MAS NMR spectrum of the FS-PMO

109

10

15

20

25

TOS(min)

Experimental conditions: T=373K (353K for zeolite Y); olefin WHSV=2 h"1;
isobutane/1-butne molar ratio=40.
Figure 34. Butnes conversion with time on stream in the alkylation of
isobutane/1-butene

110

- * - TMP in CS
- e r - C 8 in C5+

100

- - C 5-C 7 in C5+
- o - C 9+ in C5+

80 %
.
O

<

60

-C

y.

40

-O

in

20
^

o
1

~"

10

15

20

25

TOS(min)
Experimental conditions: T=373 K; olefin WHSV=2 h" ; isobutane/1-butne molar
ratio=40.
Figure 35. Distribution of C s+ products with time on stream on FS-PMO
catalyst

111

6. Alkylation of Isobutane/1-butne on Methyl-modified


Nafion/SBA-15 Materials
Wei Shen u , Yi Gu ', Hualong Xu ', David Dub 2 , Serge Kaliaguine 2*
' Department of Chemistry, Shanghai Key Laboratory of Molecular Catalysis and Innovative
Materials and Laboratory ofAdvanced Materials, Fudan University, Shanghai 200433, PR
China
2

Department of Chemical Engineering, Laval University, Quebec City, QC, Canada G1K 7P4

Accepted for publication in Applied Catalysis A: General

Abstract
Hydrophobicity modification of the intrinsic polarity of the surface of SBA-15
mesoporous by ethoxytrimethylsilane was used in this work to make hybrid organicinorganic mesoporous matrix. This matrix was functionalized with perfluorosulfonic
acidic Nafion resin by a post-synthetic impregnation method. Characterized by N2physisorption, XRD, and TEM, all the materials synthesized were highly ordered.
Elemental analysis, 29Si MAS NMR, TGA, EDX and potentiometric titration showed that
trimethylsilane is grafted on the surface by capping the OHs and the Nafion resin was
incorporated, revealing a strong solid acid with hydrophobic surface. The alkylation of
isobutane/1-butne was thereafter evaluated on each material under specified conditions.
Compared with the polar surface of conventional SBA-15 and commercial Nafion silica
nanocomposite SAC-13, methyl-modified surface of SBA-15 material (denoted as MeSBA-15) is a much better solid acid catalyst for isobutane/1-butne alkylation.
Keywords: Methyl modification, Silylation, Ethoxytrimethylsilane, Nafion, SBA-15,
Alkylation of isobutane/1-butne

112

6.1. Introduction
Alkylation of isobutane with light (C3-C5) olefins is an important refining operation
which produces mixtures of branched alkanes. The formed product, designated as
alkylate, is the highest-quality hydrocarbons for the gasoline pool because of its high
octane number, low Reid vapor pressure, very low sulfur content and freedom from
aromatics or olefins [1]. The current alkylation technology makes use of strong liquid
acids (HF or H2SO4). Both processes, while producing high quality, environmentally
benign gasoline components, suffer from a number of drawbacks.
HF is a corrosive and highly toxic liquid with a boiling point close to room
temperature. Tests indicated that liquid HF can form aerosol clouds containing lethal
levels of HF, which drift downwind at ground level for several kilometers. Two incidents
occurred in 1986 and 1987 [2]. Sulfuric acid is also a dangerous chemical, but the
dangers are relatively localized. The major disadvantage of sulfuric acid is its unusually
high catalyst consumption, which can be as high as 70-100 kg of acid/ton of alkylates [3].
Furthermore, the spent acid containing conjunct polymers forms a toxic sludge that must
be sent to a recovery facility. The transport of spent and fresh acid to and from the
sulfuric acid regeneration plant has given rise to some concern and increased the pressure
on refiners to establish sulfuric acid regeneration plants near the alkylation unit. Indeed,
the cost of the recovered acid has been estimated to be two to three times that of sulfuric
acid available on the market [4]. About one-third of the operating cost of the H2SO4
process can be attributed to acid consumption [5].
The obvious environmental risks of an unintentional release upon separation and
disposal of these catalysts is a strong motivation for the petroleum industry to develop a
heterogeneous catalyst. A variety of strong acidic solids has been tested as alkylation
catalysts, including sulfated zirconia and related materials [6, 7], heteropolyacids [8, 9],
acid resins [10, 11], chlorinated alumina [12], and acidic zeolites [13-22]. None of them
met any commercial success because of their unacceptably rapid deactivation. During
zeolite-catalyzed isobutane/butene alkylation, which almost exclusively produces
isoalkane products, a highly unsaturated and highly branched polymer is formed. The
113

polymer is strongly adsorbed on the acid sites and completely fills the pores at the end of
the reaction [23]. The alkylation reaction proceeds mainly via addition of -butne to an
isobutyl carbenium ion. The resulting octyl carbenium ion is removed (after possible
isomerization) from the active site by hydride transfer from isobutane leading to
trimethylpentanes as ideal products and an isobutyl ion, which perpetuates the reaction. A
high ratio of the rate of hydride transfer vs. the rate of oligomerization is crucial for good
catalytic performance. Strong acid sites are necessary to effectively catalyze hydride
transfer [24]. Hydrophobic surface which allows a higher paraffin steady concentration
in the pore system is important to hydride transfer. The hydrophilicity of the zeolite pores
enhances the adsorption of polarizable molecules [25]. The preferential adsorption of
alkene on zeolite surface leads to high rate of oligomerization. Moreover, the small pore
size of zeolite limits the diffusion of reactants to active sites and products out of the pore
system [26]. Thus, a catalyst with large pore size, strong acid sites and hydrophobic
surface is desired.
Perfluorosulfonic acid Nafion resin has been used as catalysts in a wide range of
organic reactions. The presence of electron-withdrawing fluorine atoms in the structure
significantly increases the acid strength of the terminal sulfonic-acid groups, which
becomes comparable to that of pure sulfuric acid [27]. However, Nafion presents low
surface areas. It performed poorly in isobutene/-butene alkylation [10, 28]. Supporting
Nafion on high surface area carriers provided perfluorosulfonic acid materials with
higher surface areas and higher density of available acid sites, which leads to the
enhancement of their catalytic performance [29]. Martinez et. al. reported that
Nafion/SBA-15 made by impregnation is a very good solid acid catalyst for acylation of
anisole [30]. Wang and Guin reported that impregnation is more beneficial than the solgel technique to make Nafion/silica catalyst for etherification of olefins [31].
Perfluorosulfonic acid grafted mesoporous materials synthesized by post-synthetic
grafting strategies [28, 32] or direct one-step synthesis [33] are catalysts with similar
advantages. However, the limited availability of suitable fluorinated silanes hinders the
applicability of this kind of materials for large scale production. Comparatively,
impregnation of Nafion resin in mesoporous silica such as SBA-15 is a practical way to
produce catalyst with higher surface area, large pore size and high strength acid sites. In
114

addition, the hydrophobic/hydrophilic balance of the catalyst can be tuned by capping the
surface -OHs with alkyl trimefhoxysilane. This is beneficial to increase the
isoparaffin/olefin ratio in the alkylation reaction conditions.
In this work, ethoxytrimethylsilane was used to cap the surface -OHs of SBA-15 to
tune its hydrophobicity. Nafion was impregnated on hydrophobicity modified SBA-15.
The activity of hydrophobicity-modified Nafion/SBA-15 (denoted as Nafion/Me-SBA-15)
materials was evaluated and compared with Nafion/SBA-15 and SAC-13 on the
alkylation of isobutene/1-butne. The influence of the hydrophobic nature of support,
reaction temperature, alkane/alkene ratio and 1-butne space velocity were studied.

6.2. Experimental
6.2.1. Catalyst preparation
SBA-15 mesoporous silica sample was synthesized by the acid-catalyzed hydrolysis
and condensation of tetraethyl orthosilicate (TEOS, Aldrich) using the triblock
copolymer Pluronic PI23 (EO20PO70EO20, BASF) as the structure-directing agent,
according to the procedure adopted by Zhao et al. [34]. In a typical synthesis, 4 g of PI23
was dissolved in 125 g of 2 M HC1 solution under stirring at 40 C. This was followed by
adding 9.12 g of TEOS into the solution as silicon source. After being stirred vigorously
for 24 h at 40 C, the resulting gel was transferred to a Teflon-lined autoclave and heated
at 100 C for an additional 48 h. After cooling to ambient temperature, the solid in the
autoclave was recovered by filtering, washing and drying at 80 C. Finally, the solids
were calcined at 550 C for 5 h to remove the organic surfactant.
Silylation of the surface -OHs of SBA-15 was carried out as follows: 2.6 g of SBA15 was pre-dried under vacuum at 200 C for 12 h before adding 3.5 g
ethoxytrimethylsilane (Aldrich) and 30 ml of dry toluene under Argon. The mixture was
refluxed at 100 C for another 12 h. Then, the hydrophobicity-modified SBA-15 material
was filtered and washed by toluene and anhydrous ethanol in mm. At last, the solid was
dried at 80 C overnight [35].

115

The supported Nafion catalysts were prepared by impregnating Nafion (5 wt%


Nafion in water-alcohol solution, Dupont) on the above prepared SBA-15 (pure SBA-15
or hydrophobicity-modified SBA-15), stirring at 60 C and atmospheric pressure for 6 h.
The solid was firstly dried at room temperature for 12 h in static conditions, and then the
water and alcohols were evaporated thoroughly under vacuum at 60 C for an additional
12 h. The resultant materials were denoted as Nafion(X)/SBA-15 or Nafion(X)/Me-SBA15, where X indicates the theoretical wt% of Nafion loading (in this work, two Nafion
loading were studied, i.e. X = 15 or 30). Me-SBA-15 indicates that the surface -OHs of
SBA-15 were silanized by ethoxytrimethylsilane. For comparison, Nafion resin/silica
composite, SAC 13, with resin content of ca. 13 wt% obtained from Aldrich was also
studied.

6.2.2. Catalyst characterization


Nitrogen adsorption-desorption isotherms at 77 K were performed using a
Micromeritics TRISTAR 3000 apparatus. The samples were degassed at 120 C and high
vacuum prior to the measurements. BET model was used to estimate the surface areas of
the materials. BJH model was performed to calculate the mesopore size using adsorption
branches of isotherms, and the pore diameter was estimated out from the peak position of
BJH pore size distribution.
Powder XRD spectra were recorded using a Bruker D4 X-ray diffractometer with
nickel-filtered CuK, radiation (X = 1.5418 ). The tube voltage was 40 kV, while the
current was 40 mA. Diffraction patterns were recorded with scan step of 0.02 for 2 theta
between 0.5 and 5.
Transmission electron micrographs (TEM) were obtained using a JEOL 2011
microscope operated at 200 kV. Scanning electron microscope (SEM) and energydispersive X-ray (EDX) microanalyses were obtained using a Philip-XL30 apparatus
operated at 20 kV.

116

Solid-state 29Si MAS NMR spectra were recorded at room temperature on a Bruker
ASX 300 spectrometer at a frequency of 59.6 MHz and at 8 kHz spinning rate.
Tetramefhylsilane (TMS) was used as external reference for 29Si MAS NMR analyses.
The ion exchange capacities (corresponding to acid site concentration) of the
Nafion-modified SBA-15 materials were determined using aqueous solutions of NaCl
and titrated potentiometrically by NaOH [36]. In a typical experiment, 0.1 g of solid was
added to 20 g of 2 M NaCl solution, and vigorously stirred at room temperature overnight.
The resulting suspension was filtered and washed thoroughly with a total amount of 80
ml 2 M NaCl to retain the hydrogen ions in the solution, and thereafter titrated
potentiometrically by a drop wise addition of a 0.1 M NaOH.
Thermal gravimetric analysis (TGA) was conducted on a PerkinElmer TGA 7 from
ambient temperature to 800 C at a heating rate of 10 C/min in air.
Sulfur and total organic content were determined by elemental analysis using a
Vario EL III apparatus (CHNS model).

6.2.3. Reaction Procedure


Liquid phase alkylation of isobutane/1-butne experiments were carried out in an
automated stainless steel fixed bed continuous reactor. In each run, 0.35 g of catalyst
pellets with diameters of 0.5 to 0.8 mm, obtained by compressing the powder into tablets,
crushing, and sieving, were loaded in a reactor tube with internal diameter of 8 mm. The
catalyst bed was fixed between two plugs of quartz wool and the remaining empty
volume filled with quartz beads. The catalysts were pretreated in situ at 110 C for 10 h
in a 15 ml/min flow of N2. The reactor was cooled to reaction temperature and a pressure
of 3.2-3.4 MPa was established with nitrogen before feeding the reactants. The feedstock
consisting of an isobutane/1-butne mixture with a molar ratio of 20/1 and 40/1 was
delivered to the reactor using a piston-type pump, and it goes through the catalyst bed
located in the middle zone of the reactor. The product stream coming out of the reactor is
then collected and stored using a 16-loop sampling valve. In this way samples at intervals
of 1 min can be taken during the run and analyzed automatically once the experiment has
117

been finished. The reaction products were analyzed using a GC (Thermo Trace GC Ultra)
equipped with a 100m capillary column (CP SIL PONA CB) and a FID detector. The
individual C4-C8 hydrocarbons identified by means of available reference standards and
GC-MS analysis.

6.3. Results and discussion


6.3.1. Catalyst characterization
Hybrid Nafion/SBA-15 materials combine the merits of Nafion resin and
mesoporous substrates as a whole in these dedicated materials. Nafion resin, a
perfluorosulfonic acid, has an acidity comparable to that of concentrated sulfuric acid,
providing effective active sites in alkylation reactions. Mesoporous SBA-15 support has
uniform pore size, together with large surface and pore void, ensuring more interacting
surface of reactants and meanwhile has a potential for loading more well-dispersed
Nafion resin as active sites. Silylation of the surface -OHs into methylsilane further
contributed to the function of the materials. Hydrophobicity-modified surface of SBA-15
is expected to be suitable for alkylation reaction by changing the surface concentration
ratios of isobutane/1-butne which affects the reaction remarkably.
Figure 36a and b present nitrogen adsorption-desorption isotherms and XRD
patterns of SBA-15, Me-SBA-15 and their Nafion-modified samples. All the materials
displayed type IV isotherms with HI-type adsorption-desorption hysteresis loop. The
narrow HI-type hysteresis loop of SBA-15 is maintained after the grafting of
trimethylsilane (OH capping), indicating that the structure is not affected by silylation. It
becomes broader slightly after impregnation of 15 wt% of Nafion. That means the Nafion
resin that penetrated into the porous framework seems to be homogeneously deposited
along the cylindrical mesoporous channels of materials when the Nafion loading is lower
than 15 wt%. With increasing Nafion loading, the hysteresis loop in the N2 isotherm
becomes broader, suggesting Nafion resin places within the mesoporous structure in
forms of polymeric aggregates [37]. XRD patterns of Nafion/SBA-15 samples show the
characteristic (100), (110) and (200) diffractions clearly, indicating the presence of
118

mesostructure with hexagonal p6mm symmetry. The secondary diffraction peaks of


materials modified by silylation and impregnation decrease in intensity, demonstrating
the deterioration of the mesopore order. However, the sharp peaks of (100) of the
materials synthesized show that mesoporous textures can be maintained in all cases.
Textural parameters deduced from nitrogen isotherms and XRD patterns of the
synthesized materials are summed in Table 12. Me-SBA-15 has textural parameters
similar to those of SBA-15, indicating little influence of the silylation treatment on
mesoporous textures. Impregnation of Nafion induces reduction of BET surface area,
pore diameter and pore volume. BJH analysis of the adsorption branches of the isotherms
reveals that the mesopore diameter decreased from ca. 7.6 nm to 6.6 nm, while the wall
thickness increased systematically from 2.8 nm to 4.5 nm with the incorporation of 30
wt% Nafion loadings. The decrease of pore diameter and increase of wall thickness
confirm the deposition of Nafion resin on the pore walls. Though the pore diameter
decreases after OH capping and Nafion loading, the mean pore size remains beyond 6.6
nm for all the materials, which makes these acid catalysts potential applicants in catalytic
reactions involving bulky molecules. The impregnation of Nafion affects the narrow pore
size distribution of the purely siliceous SBA-15 support, leading to broader distributions
centered at smaller pore sizes (Figure 37). That suggests Nafion resin may be placed
within the mesopores as polymeric aggregates especially at higher loading.
29

Si MAS NMR spectrum of Nafion(30)/Me-SBA-15 is presented in Figure 38. Four

peaks at -110, -101, -90, and -64 are attributable to Q4 [Si(OSi)4], Q3 [OHSi(OSi)3], Q2
[(OH)2Si(OSi)2] and T3 [RSi(OSi)3] resonances respectively [38, 39]. The result indicates
that ethoxytrimethylsilane is grafted on the surface of the SBA-15 catalyst. The surface
ratios of the peak components associated to these silica species are recognized to be
Q4:Q3:Q2:T3 = 57:23:5:15. While there is no occurrence of T3 [RSi(OSi)3] resonances for
Nafion(30)/SBA-15 sample and its Q4:Q3:Q2 is 49:43:8. The increase in T3 and Q4
corresponds to the disappearance of some of the surface OH's upon capping. Result of
elemental analysis (Table 13) indicates that 0.6 mmol/gcat of methyl groups was grafted
on SBA-15.

119

Figure 39 (a) shows a typical SEM image of Nafion(30)/Me-SBA-15 material which


consists of irregular particles. TEM images (Figure 38b and 38c) of this sample show the
hexagonal array of uniform channels with the typical honeycomb appearance of SBA-15
materials with an estimated pore diameter around 6.6 nm which is in agreement with the
pore size obtained by nitrogen adsorption-desorption isotherms.
Table 13 shows several parameters related to the chemical properties including the
acidity and methyl incorporation of the hybrid materials. TG analyses have been used to
evaluate the experimental Nafion loading of each sample. The weight loss between 300
C and 550 C was assigned to the thermo decomposition of the Nafion resin [30]. The
incorporation yield ranges from 84 wt% to 88 wt%. It demonstrated that the Nafion
loading (between 15 wt% to 30 wt%) and polarity of SBA-15 did not make great effect
on incorporation yield. Results of H+ concentrations determined by drop-wise
potentiometric titration and the sulfur content determined by elemental analysis are
consistent with TG result. The comparison between S and [H+] contents in the various
solids suggests that Nafion is protonated and essentially all acid sites can exchange
sodium ions from NaOH solution. S/Si atomic ratios given by EDX also indicate the
trend of Nafion incorporation and accessibilities of acid sites. All these results evidence
that the active sites provided by Nafion resin are highly accessible, revealing the effective
role of the mesostructured silica support.
Methyl capping of surface -OHs, though not providing the active sites as the Nafion
resin, plays an important role in tuning the physical property of the surface of the catalyst.
It changes the relative affinity of the catalyst for the sorption of less-polarized reactant,
i.e. isobutane, which is beneficial to increasing the surface isoparaffin/olefin ratio in the
alkylation reaction conditions. Furthermore, methyl groups improve the catalytic
performance with an efficient removal from the catalyst surface of the water introduced
by the feedstock or moisture that are difficult to remove by thermo heating at low
temperature due to the limitation of the thermostability of the Nafion resin. In the
catalysis result shown below, the benefit of the methyl groups on the overall conversion
of alkylation will be discussed.

120

6.3.2. Catalytic performance


Nafion/silica composites, such as commercially available SAC-13, are reported to be
good catalysts for isobutane/-butene alkylation [40]. Nafion/SBA-15 with similar H+/gcat
is a better solid acid catalyst for acylation [30]. In this part of the work, Nafion
impregnated SBA-15 and Me-SBA-15 catalysts were synthesized and compared with
SAC-13 in the alkylation of isobutane/1-butne reaction. The influence of reaction
temperature, isobutane/1-butne ratio, alkene space velocity was studied. Isobutane/1butene alkylation on acid catalysts yields trimethylpentanes (TMPs) as the primary
products. A set of parallel and consecutive reaction steps can occur, giving mixtures of
hydrocarbons ranging from C5 to C9+. Cs consists of trimethylpentanes (TMPs),
dimethylhexanes (DMHs) and Cg olefins (Cg). The TMPs+ are commonly considered to
be formed by the reaction of /-butyl carbenium ion with butne, they can undergo hydride
transfer from isobutane to form the TMPs. The tert-butyl cation is regenerated and the
chain sequence can continue. The DMHs and Cs~ are believed to be formed by
dimerization of olefin [10] or cracking of heavier compounds [41] depending on the
hydride transfer activity. The product distribution is governed by the relative rates of
alkene addition, isomerization, and hydride transfer. Hydride transfer/alkene addition
determines the selectivity to single and multiple alkylation. A high ratio of hydride
transfer vs. alkene addition retards the buildup of long hydrocarbon chains, leads to high
ratio of TMPs in products and long catalytic lifetime.
Alkylation is catalyzed by sites with strong acidity, while double bond isomerization
is catalyzed by very weak acid sites [6]. Even a fully deactivated zeolite retains some
activity for isomerizing butnes. Thus the conversion is calculated based on
disappearance of butnes.
Three Nafion/SBA-15 materials were tested and compared with SAC-13 at same
reaction conditions. C5, C, C7, TMPs, DMHs, Cg= and C9+ are obtained. TMPs are the
primary products. Figure 40 shows the catalytic performance of each material in terms of
butnes conversion, TMPs yield and C5-C7 selectivity with time on stream, whereas their
relevant textural and catalytic properties are summarized in Table 14.
121

The results

shown

indicate that

SAC-13 gives a higher activity than

Nafion(15)/SBA-15 which has similar H+ concentration and larger surface area. It is


inverse to the results got by Martinez et al. in acylation reaction [30]. This can be
explained by the different acid strength these two reactions required. Alkylation needs
very strong acid sites to catalyze. The interaction between the sulfonic groups of Nafion
and silanol groups of the silica leads to a decrease in acidity. Corma et al. reported that
Nafion/silica composites with the same polymer content but a larger surface area showed
a considerably lower activity in isobutane/2-butne alkylation[10]. Nafion(15)/SBA-15
has a higher number of silanol groups resulting in more interaction between the active
phase and the silica support. In this way, a charge transfer may occur from the protons of
the sulfonic groups to the silanol groups of the silica, decreasing the acid strength of the
resin. Palinko et al. studied the interaction between -SO2OH groups of Nafion and the
silanols of the silica using physical characterization techniques [42]. They found the
interaction leads to a decrease in acidity due to the levelling effect of the hydrating
environment. Thus, the greater the number of silanol groups, the weaker the global
acidity of the active phase. Consequently, the catalytic conversion of alkylation will be
lower. This was also confirmed by the selectivity of C5-C7, which are the result of
cracking. Cracking is catalyzed only by the strongest acid sites [6]. SAC-13 shows higher
initial cracking activity than nafion(15)/SBA-15. This can be attributed to the increase in
acid strength.
With the increase in Nafion loading (from 15 wt% to 30 wt%), Nafion resin is
present within the mesoporous structure under the form of polymeric aggregates. The
negative effect of the interactions between Nafion and silanol groups diminishes. The
number and strength of acid sites which catalyze isobutane/1-butne alkylation increase.
The initial butnes conversion was increased from 54.2 wt% to 92.3 wt% when the
Nafion loading of Nafion/SBA-15 was changed from 15 wt% to 30 wt%. High Nafion
content favors the formation of TMPs and C5-C7 products. The initial yield of TMPs of
Nafion(30)/SBA-15 is not as high as expected because the strong perfluorosulfonic acid
sites show high initial cracking activity. With time on stream, deactivation of the
catalysts, decrease in selectivity to C5-C7 and a higher content of octenes and Cg+ can be
observed. The strongest acid sites are the first ones to be poisoned.
122

The activity of Nafion(30)/SBA-15 is obviously improved after its surface OHs


were capped by ethoxytrimethylsilane. Nafion(30)/Me-SBA-15 gives higher initial
butnes conversion (98.6 wt% vs. 92.3 wt%) and TMPs yields (63.5 wt% vs. 51.7 wt%).
At same time, the selectivity of C5-C7 of two catalysts is almost same. That means the
OH capping does not seem to increase activity through increasing the acid strength.
Catalyst characterization indicated there are no essential structure changes induced by the
capping procedure. This capping is intended to enhance the pore surface hydrophobicity
which is reflected by

Si MAS NMR data indicating a diminution of the surface silanol

density. The higher activity obtained with Nafion(30)/Me-SBA-15 in comparison to


Nafion(30)/SBA-15 can be related to a higher isobutane concentration in the hydrophobic
environment of the capped catalyst comparatively to the hydrophilic silica support. As
mentioned above, for isobutane/1-butne alkylation reaction, the octyl carbenium ion is
removed from the active site by hydride transfer from isobutane leading to
trimethylpentanes as ideal products and an isobutyl ion, which perpetuates the reaction.
The high ratio of the rate of hydride transfer vs. the rate of oligomerization is crucial for a
good catalytic performance. High ratio of isobutane/1-butne is necessary for hydride
transfer. The electrostatic field at the surface of a solid acid catalyst enhances the
adsorption of alkenes. The preferential adsorption of alkene leads to high rate of
oligomerization, which either blocks the pores or poisons the active sites. To enhance
surface carbophilicity by capping the silanols of SBA-15 increases the I/O ratio in the
reaction conditions and thus improves the catalytic performance.
The influence of reaction parameters on the catalytic performance of Nafion/MeSBA-15 was studied. The results are shown in Figure 41. It can be seen that with same 1butene space velocity and isobutane/1-butne ratio, the conversion of butnes and yield
of TMPs increase dramatically with temperature. The initial conversion of butnes
increases from 71.4 wt% to 98.6 wt% when the reaction temperature increases from
70 C to 100 C (WHSV = 2 h'\ I/O = 40). Moreover, this increase in conversion is
accompanied by an increase in yield to TMPs. The fraction of the C5-C7 products is
favoured at higher temperature while C9+ increases at lower temperature (not shown).

123

The influence of WHSV is also shown in Figgure 41. When WHSV of 1-butne
decreases from 2 h'1 to 1.4 h"1 (I/O = 40, T = 100 C), the decay of catalyst is much
slower and the initial conversion of butnes increases slightly (from 98.6 wt% to 100
wt%). As the WHSV of 1-butne increases, the contact time decreases, resulting in an
increase of the Cg and C9+ fractions, which not only decreases the selectivity of TMPs
but also fastens the catalyst decay.
The isoparaffin/olefin ratio is a very important parameter in isobutane/w-butene
alkylation. Lower I/O ratio leads to higher rate of oligomerization. It can be seen in
Figure 41 that I/O shows obvious effect on catalytic performance. The conversion of
butnes decreases much faster when the I/O ratio decreases from 40 to 20 (WHSV = 2 h"1,
T = 100 C). High I/O ratio obviously favours the yield of TMPs. However, high I/O
ratio will leads to high cost of separating products from the excess isobutane, which
limits the I/O ratio that can be used industrially.
The I/O ratio determines the concentration of isobutane in the reactor and thereby
the rate of hydride transfers. The I/O ratio also sets the product concentration, which
affects the rates of the product degradation reactions [23]. The local concentration of 1butene in the pore system is strongly affected by the polarity of surface. Capping the OHs
enhances the hydrophobic nature of the SBA-15 surface, which allows higher
concentrations of isobutane, thereby mitigating the effect of a decrease in I/O ratio. The
effect of I/O ratio on initial catalyst activity with and without OH capping is compared in
Table 15. As the I/O ratio decreases from 40 to 20, the apparent concentration of 1butene is doubled, the activity of Nafion(30)/Me-SBA-15 decreases slightly while that of
Nafion(30)/SBA-15 decreases dramatically. The initial conversion Nafion(30)/Me-SBA15 decreases from 92.3 wt% to 49.8 wt%. The result confirms that OH capping provides
hydrophobic enviroment and allows comparatively higher local paraffin concentration
even at low I/O ratio. This enhanced activity of Nafion(30)/Me-SBA-15 at lower I/O ratio
has potential commercial interest by reducing the cost of separation.
The main reason for deactivation of Nafion based solid acid catalysts is due to the
formation of heavy products that are strongly adsorbed over the active sites [30, 40]. No
124

leaching of Nafion resin was detected by elemental and TG analysis on tested catalysts.
Both Nafion/SBA-15 series and SAC-13 catalysts were regenerated with washing by
acetone and further treatment with HNO3 [30]. The catalytic performance of fresh and
regenerated catalysts is illustrated in Figure 42. About 85% of the initial catalytic activity
was achieved.

6.4.3. Conclusions
Mesostructured

SBA-15

hydrophobically

modified

by

OH

capping

and

functionalized with Nafion resin by means of impregnation was described. The resulting
solid acid catalysts preserve the mesostructure of SBA-15 and show high surface area (ca.
400 m2/g) and narrow pore size distribution centred in the mesoscale range (ca. 1 nm).
Capping of surface OH diminishes the surface silanol density and provides a hydrophobic
environment for the isobutane/1-butne reaction.
Hydrophobically modified Nafion(30)/Me-SBA-15 shows excellent activity and
efficiency in the production of isooctane compared with Nafion/SBA-15 and SAC-13.
The catalytic performance of Nafion/SBA-15 is obviously improved after its surface OHs
were capped. To enhance the surface carbophilicity increases the local I/O ratio in the
reaction conditions and results in a higher ratio of hydride transfer/oligomerization rate
and good catalytic performance. The catalytic activity of hydrophobic Nafion/Me-SBA15 is much superior to that of hydrophilic Nafion/SBA-15 especially at low I/O ratio. No
nafion leaching being detected evidences a high stability of the impregnated resin. The
used catalysts can be regenerated by washing with acetone.

Acknowledgments
This work was supported by NSERC of Canada and Science & Technology
Commission of Shanghai Municipality (08DZ2270500) "

125

References
[I]

A. Corma, and A. Martinez, Catal. Rev.-Sci. Eeg., 35 (1993) 483-570.

[2]

L. F. Albright, Alkylation-Industrial. In Encyclopedia of Catalysis; Howath, I. T.,


Ed.; John Wiley and Sons: New York, Vol. 1, (2003) 226-281.

[3]

J. Weitkamp, Y. Traa, Handbook of Heterogeneous Catalysis; Ertl, G,


Knoezinger, H., Weitkamp, J., Eds.; VCH: Weinheim, Vol. 4, (1997) 20392069.

[4]

E. Furimsky, Catal Today 30 (1996) 223-286.

[5]

L. F. Albright, Chemtech. (1998) 46-53.

[6]

A. Corma, A. Martinez, C. Martinez, J. Catal. 149 (1994) 52-60.

[7]

A. S. Chellappa, R. C. Miller, W. J. Thomson, Appl. Catal. A 209 (2001) 359374.

[8]

T. Blasco, A. Corma, A. Martinez, P. Martinez-Escolano, J. Catal. 177 (1998)


306-313.

[9]

V. Sarsani, Y. Wang, B. Subramaniam, Ind. Eng. Chem. Res. 44 (2005) 64916495.

[10] P. Botella, A. Corma, J. M. Lopez-Nieto, J. Catal. 185 (1999) 371-377.


[II] C. J. Lyon, V. S. R. Sarsani, B. Subramaniam, Ind. Eng. Chem. Res. 43 (2004)
4809-4814.
[12] S. I. Hommeltoft, Appl. Catal. A 221 (2001) 421-428.
[13] A. W. Chester, Y. F. Chu, Zeolite 6 (1986) 195-200.
[14] A. Corma, A. Martinez, C. Martinez, Catal. Lett. 28 (1994) 187-201.
126

[15] M. Stocker, H. Mostad, T. Rorvik, Catal. Lett. 28 (1994) 203-209.


[16] F. Cardona, N. S. Gnep, M. Guisnet, G. Szabo, P. Nascimento, Appl. Catal. A
128(1995)243-257.
[17] A. Corma, A. Martinez, P. A. Arroyo, J. L. F. Monteiro, E. F. Sousa-Aguiar,
Appl. Catal. A 142 (1996) 139-150.
[18] D. E. Sherwood, R. J. Taylor, Appl. Catal. A 155 (1997) 195-215.
[19] G. S. Nivarthy, Y. He, K. Seshan, J. A. Lercher, J. Catal. 176 (1998) 192-203.
[20] R. Loenders, P. A. Jacobs, J. A. Martens, J. Catal. 176 (1998) 545-551.
[21] A. Feller, A. Guzman, I. Zuazo, J. A. Lercher, J. Catal. 224 (2004) 80-93.
[22] M. Mukhergee, J. Nehlsen, Hydrocarbon Process. 86 (2007) 110-114.
[23] A. Feller, I. Zuazo, A. Guzman, J. Barth, J. A. Lercher, J. Catal. 216 (2003) 313323.
[24] A. Corma, A. Martinez, C. Martinez, J. Catal. 146 (1994) 185-192.
[25] S. Ramachandran, T. G. Lenz, W. M. Skiff, A. K. Rappe, J. Phys. Chem. 100
(1996) 5898-5907.
[26] K. P. de Jong, C. M. A. M. Mesters, D. G. R. Peferoen, P. T. M. van Brugge, C.
de Groot, Chem. Eng. Sci. 51 (1996) 2053-2060.
[27] M. A. Harmer, W.E. Farneth, Q. Sun, Adv. Mater. 10 (1998) 1255-1257.
[28] W. Shen, D. Dube, S. Kaliaguine, Catal. Commun. 10 (2008) 291-294.
[29] A. Heidekum, M.A. Harmer, W.F. Holderich, J. Catal. 176 (1998) 260-263.
[30] F. Martinez, G. Morales, A. Martin, R. van Grieken, Appl. Catal. A 347 (2008)
169-178.
127

[31] S. Wang, J.A. Guin, Energy Fuels 15 (2001) 666-670.


[32] M. Alvaro, A. Corma, D. Das, V. Fomes, H. Garcia, Chem. Commun. (2004)
956-957.
[33] D.J. Macquarrie, S.J. Tavener, M.A. Harmer, Chem. Commun. (2005) 23632365.
[34] D. Zhao, J. Feng, Q. Huo, N. Melosh, G. H. Fredrickson, B. F. Chmelka, G. D.
Stucky, Science 279 (1998) 548-552.
[35] S. Parambadath, M. Chidambaram, A.P. Singh, Catal. Today 97 (2004) 233-240.
[36] D. Margolese, J. A. Molero, S. C. Christiansen, B. F. Chmelka, G. D. Stucky,
Chem. Mater., 12 (2000) 2448-2459.
[37] M. Choi, F. Kleitz, D. Liu, H.Y. Lee, W.S. Ahn, R. Ryoo, J. Am. Chem. Soc.
127(2005)1924-1932.
[38] G. Morales, G. Athens, B. F. Chmelka, R. Van Grieken, J. A. Melero, J. Catal.
254(2008)205-217.
[39] S. Hamoudi, S. Royer, S. Kaliaguine, Microp. Meso. Mater., 71, 2004, 17-25.
[40] P. Kumar, W. Vermeiren, J. P. Dath, W. H. Hoelderich, Energy Fuels 20 (2006)
481-487.
[41] L. Lee, P. Harriott, Ind. Eng. Chem. Process Des. Dev. 16 (1977) 282-287.
[42] I. Palinko, B. Torok, G. S. K. Prakash, G. A. Olah, Appl. Catal.A Gen. 174,
(1998) 147-153.

128

Table 12. Textural properties for Nafion-modified mesostructured materials

Sample

SBET (m

/g)

DPore3(nm)

. vPore3
3

b
A
"100

(cm /g)

(nm)

Wall
thickness0
(nm)

SBA-15

702

7.5

0.96

9.3

2.9

Nafion(15)/SBA-15

474

7.0

0.65

9.6

4.1

Nafion(30)/SBA-15

404

6.6

0.40

9.6

4.5

Me-SBA-15

727

7.6

1.05

9.0

2.8

Nafion(30)/Me-SBA-15

363

6.6

0.45

9.5

4.4

Total pore size and pore volume of Nafion-modified SBA-15 materials were calculated
from BJH adsorption branch, and the pore diameter was deduced from the peak
position of BJH pore size distribution.

d (100) spacing, measured from small-angle XRD.

Pore wall thickness calculated as ao - pore diameter with a0 = 2d(l00)l V3

129

Table 13. Methyl incorporation and acidic-related properties of Nafion modified


mesostructured materials

Sample

Nafion

Acidic Properties

Methyl

incorporation
wt%3

Yield"

content
(mmol

(mmol

S/Si

(mmol

S/gca.)"

H+W

(molar

methyl/gca,) 6

ratio)"
Nafion(15)/SBA-15

12.9

86

0.124

0.130

0.007

Nafion(30)/SBA-15

26.3

88

0.260

0.270

0.017

Nafion(30)/Me-SBA-15

25.2

84

0.248

0.258

0.016

0.6 0.1

Percentage of weight loss from 300 to 550 C. Value corrected by deducing the weight

decrease by the dehydration of the corresponding mesoporous silica prepared without


Nafion, i.e. siliceous SBA-15 or Me-SBA-15 [30].
b

Sulfur content evaluated by elemental analysis.

Obtained by cationic-exchange in 2 M NaCl, and titrated potentiometrically with 0.1 M

NaOH after filtration and thorough wash.


d

S/Si atomic ratio given by EDX, average ratio of 20 random ordered regions per sample.

Carbon content evaluated by elemental analysis.

130

Table 14. Textural and catalytic properities of catalysts

Sample

Textural properties

Catalytic properties
mmol S/gca,"

mmol H+/gcatc

0.45

0.248

0.258

6.6

0.40

0.260

0.270

7.0

0.65

0.124

0.130

0.43

0.147

0.133

SBET

Dpore

Vpore

(m2/g)

(nm)

(cm3/g)

Nafion(30)/Me-SBA-15

363

6.6

Nafion(30)/SBA-15

404

Nafion(15)/SBA-15

474

SAC-13

200

Total pore size and pore volume were calculated by BJH method.
b

Sulfur content evaluated by elemental analysis.

Obtained by cationic-exchange with NaCl and titration with NaOH.

131

Table 15. Influence of I/O ratio on the initial (TOS = 1 min) butnes conversion and
products distribution on Nafion(30)/SBA-15 and Nafon(30)/Me-SBA-15a

Catalyst:
I/O ratio:

Nafion(30)/SBA-15

Nafion(30)/Me -SBA-15

40

20

40

20

92.3

49.8

98.6

98.2

c 5 -c 7

41.5

30.8

42.7

36.7

Cg

44.2

52.8

48.2

55.2

C9+

14.3

16.4

9.1

8.1

TMP

62.1

42.1

65.5

50.2

DMH

30.7

40.9

33.4

45.8

7.2

17.0

1.1

4.0

Butnes conversion (wt%)


Distribution of C5+ (wt%)

Distribution of C8 (wt%)

c8
a

Other experimental conditions: T = 100 C, olefin WHSV = 2 h"

132

Z3

(a)

Nafion(3(

(b)

(100)

1300

< 200 >

Nafion(30)/Me-SBA-15

Me-SBA-15

Nafion(30)/SBA-15

Nafion(15)/SBA-15

SBA-15
J.8

P/P.

1.0

2 t h e t a (Hpnrppi

Figure 36. (a) Nitrogen adsorption-desorption isotherms at 77 K and (b) X-ray


diffraction

133

<i

a-

H -

4
I \l

Naiion(30)/Me-SBA-15

1 1
i1

1 -

S"

Me-SBA-15

A
fl

O-0~0o

O-o-O^O-O-o-'-

Nafio)jij(30)/SBA-15
* - - -

3 -

2 -

A A ^ ^

||

Nafion(15)/SBA-15

rHY"ViA l\i\ fl A

A-A-AAAA

1 -

"*.'

SBA-15

D 1

1 i i i

'

1D0

Pore Diameter (nm)

Figure 37. Pore size distributions of the Nafion-modified SBA-15 materials.

134

c/5

C
O

-50

-60

-70

(PPM)

y
Figure 38. '29,
Si MAS NMR spectrum of Nafion(30)/Me-SBA-15.

135

%v

.am;

(a)

JJP*

50 nm

Figure 39. (a) SEM image and TEM images of Nafion(30)/Me-SBA-15 viewed
(b) in the [100] direction and (c) in the [110] direction.

136

100

-r
c-

%
C
O

Nafion(30)/Me-SBA-15 (O)
Nafion(30)/SBA-15(A)
SAC-13 ( )
Nafion(15)/SBA-15(A)

SO
60

> 40
c
o
U
20

J*jJA^

^ ^

20
TOS (min)

10

30

40

80
Nafion(30)/Me-SBA-15 (O)
Nafion(30)/SBA-15(A)
SAC-13 ( )
Nafion(15)/SBA-15(A)

60

40

"S
? 20

10

15

20

25

TOS (min)

30

35

40

50

Nafion(30)/Me-SBA-15 (O)
Nafion(30)/SBA-15(A)
SAC-13 ( )
Nafion(15)/SBA-15(A)

40

%
&

30

>

CJ
'.s

20

C/j

w-,

10

u
10

20

30

40

TOS (min)

Figure 40. Catalytic performance of Nation-based catalysts. Reaction conditions: T


= 100 C; olefin WHSV - 2 h"1; isobutane/1-butene (I/O) molar ration = 40.
137

100
80
NT

60

c
o
U
20

10

10

20
30
TOS (min)

15 20 25
TOS (min)

30

40

35

40

Figure 41. Variation of the conversion of butnes (a) and yield of TMPs (b) with
time on stream over Nation(30)/Me-SBA-15 under varying experimental conditions.
100 C, WHSV = 1.4 h \ I/O = 40 ( ) ; 100 C, WHSV = 2h \ I/O = 40 (A); 70C,
WHSV = 2 h \ I/O = 40 (O); 100 C, WHSV = 2 h ', I/O = 20 (A).

138

100
n Fresh catalyst
After regeneration

80

.2

60

ca
u
o

>

40
ea

20

Nafion(30)
/Me-SBA-15

Nafion(30)
/SBA-15

SAC-13

Figure 42. Initial catalytic activity of fresh and regenerated Nafion-containing


catalysts. Reaction conditions: T = 100 C; olefin WHSV = 2 h"1; I/O = 40.

139

7. Alkylation of Isobutane/1 -butne on Methyl-modified Nafion


/SBA-16 Materials
Wei Shen u , Yi Gu \ Hualong Xu \ David Dub 2 , Serge Kaliaguine 2 *
' Department of Chemistry, Shanghai Key Laboratory of Molecular Catalysis and
Innovative Materials and Laboratory of Advanced Materials, Fudan University,
Shanghai 200433, PR China
2

Department of Chemical Engineering, Laval University, Quebec City, QC, Canada G IK

7P4
Submitted to Industrial & Engineering Chemistry Research

Abstract
Three-dimensional mesoporous SBA-16 silica materials were functionalized with
perfluorosulfonic acidic resin Nafion using an impregnation method. The intrinsic
polarity of SBA-16 surface was tuned by grafting ethoxytrimethylsilane on its surface.
Characterized by N2-physisorption, XRD, and TEM, all the materials synthesized
exhibited ordered three-dimensional Im3m mesoporous structure. Elemental analysis and
water adsorption measured by intelligent gravimetric analyzer (IGA) showed that
trimethylsilane is grafted on the surface by capping the OHs which enhances the
hydrophobicity of SBA-16. Elemental analysis and potentiometric titration showed that
Nafion resin was incorporated, revealing a three-dimensional mesoporous strong solid
acid with hydrophobic surface. The catalytic alkylation of isobutane/1-butne was
thereafter evaluated on each material under specified conditions and compared with the
one-dimensional Nafion/SBA-15 and commercial Nafion silica nanocomposite SAC-13.
The catalyst with three-dimensional mesoporous channels was shown to outperform the
one with one-dimensional channels. The higher activity of Nafion over methyl modified
SBA-16 materials is related to the more hydrophobic surface of support.

140

Keywords:

Alkylation

of

isobutane/1-butne,

Nafion,

SBA-16,

Silylation,

Dimensionality, Pore structure, Hydrophobicity.

7.1. Introduction
Alkylation of isobutane with light (C3-C5) olefins is an important refining operation
which produces mixtures of branched alkanes. The formed product, designated as
alkylate, is the highest-quality hydrocarbons for the gasoline pool because of its high
octane number, low Reid vapor pressure, very low sulfur content and freedom from
aromatics or olefins [1]. The current alkylation technology makes use of strong liquid
acids (HF or H2SO4). Both processes, while producing high quality, environmentally
benign gasoline components, suffer from a number of drawbacks. The obvious
environmental risks of an unintentional release upon separation and disposal of these
catalysts necessitate their replacement. A variety of strong acidic solids has been
investigated, including sulfated zirconia and related materials [2, 3], heteropolyacids [4,
5], acid resins [6, 7], chlorinated alumina [8], and acidic zeolites [9-18] None of them
met any commercial success because of their unacceptably rapid deactivation.
The alkylation reaction proceeds mainly via addition of w-butene to an isobutyl
carbenium ion. The resulting octyl carbenium ion is removed (after possible
isomerization) from the active site by hydride transfer from isobutane leading to
trimethylpentanes as ideal products and an isobutyl ion, which perpetuates the reaction.
On the other side, the octyl carbenium may continue to react with olefins to form a
polymer. The polymer is strongly adsorbed on the acid sites and completely fills the
pores at the end of the reaction [19]. It is generally accepted that a promising catalyst
should be acidic enough to form the intermediate carbocations and catalyze hydride
transfer. Meanwhile, the pores should be large enough to allow the diffusion of reactants
to active sites and products out of the pore system [1]. Moreover, hydrophobic surface
which allows a higher paraffin steady concentration in the pore system is desired [20].
Apart from the above factors, the pore structure of solid acid catalysts also plays very
important roles in alkylation. It has a direct influence on the effective diffusivities [21].
Yoo et al reported [22] that zeolite ZSM-12 which possesses one-dimensional
141

noninterpenetrating channels outperformed USY with three-dimensional channels and


larger pores. The authors thought the linear channels which do not possess any
expansions do not allow the formation of bulky carbonaceous materials that could lead to
pore plugging. This is in contrast to the common belief that three-dimensional zeolites are
less susceptible to pore plugging than one-dimensional zeolites. Corma et al pointed out
that tridirectional structures of zeolites which allowed easier diffusion should be
preferred to monodirectional ones [1]. No research on the influence of pore structure of
acid functionalized mesoporous silica on alkylation reaction was yet reported.
The strength of perfluorosulfonic acid is comparable to that of pure sulfuric acid
[23]. Perfluorosulfonic acid functionalized silica showed good activity in isobutane/butene alkylation [6, 20, 24]. These materials can be made by immobilization of
perfluorosulfonic acid groups on mesoporous materials using post-synthetic grafting
strategies [20, 25] or direct one-step synthesis strategies [26]. However, the limited
availability of suitable fluorinated silanes hinders the applicability of this kind of
materials for large scale production. Supporting perfluorosulfonic acid resin Nafion on
high surface area carriers provided a practical way to produce catalyst with higher surface
area, large pore size and high strength acid sites [27]. Wang and Guin reported that
impregnation is more beneficial than the sol-gel technique to make Nafion/silica catalyst
for etherification of olefins [28]. Recently, we synthesized Nafion modified SBA-15
using an impregnation method [29]. Nafion resin was supported on the one-dimensional
mesopores, which still leave enough space for diffusion of alkylation products. In
addition, the hydrophobic/hydrophilic balance of the catalyst can be tuned by capping the
surface -OHs with alkyl trimethoxysilane. This is beneficial to increase the local
isoparaffin/olefin ratio in the pore system of catalysts.
In this work, Nafion was impregnated on three-dimensional mesoporous SBA-16
silica. The hydrophobicity of SBA-16 surface was enhanced by capping the surface OHs.
The activity of Nafion

modified

SBA-16 (denoted

as Nafion/SBA-16)

and

hydrophobicity-modified Nafion/SBA-16 (denoted as Nafion/Me-SBA-16) materials was


evaluated and compared with Nafion/SBA-15, Nafion/Me-SBA-15 and SAC-13 in the
alkylation of isobutene/1-butne. The influence of the structure, hydrophobic nature of
142

support, reaction temperature, isobutane/1-butne ratio and 1-butne space velocity were
studied.

7.2. Experimental
7.2.1. Catalyst preparation
SBA-16 mesoporous silica sample was synthesized by the acid-catalyzed hydrolysis
and condensation of tetraethyl orthosilicate (TEOS, Aldrich) using the triblock
copolymer Pluronic F127 (EO106PO70EO106, BASF) as the structure-directing agent with
the aid of K2SO4, according to the procedure described by Zhao et al. [30]. Because large
surface area of silica decreases the acidity of the sulfonic groups due to the interaction
between the sulfonic groups of Nafion resin and silanol groups of the silica [6], a
relatively larger TEOS/F127 ratio was used to synthesize SBA-16 with smaller surfaces.
In a typical synthesis, 4 g of F127 was dissolved in 120 g of 0.5 M HC1 solution under
stirring at 38 C. This was followed by adding 16.8 g of TEOS into the solution as silicon
source. After being stirred vigorously for 24 h at 38 C, the resulting gel was transferred
to a Teflon-lined autoclave and heated at 100 C for an additional 48 h. After cooling to
ambient temperature, the solid in the autoclave was recovered by filtering, washing and
drying at 80 C. Finally, the solids were calcined at 550 C for 6 h in air to remove the
organic surfactant.
Silylation of the surface -OHs of SBA-16 was carried out as follows: 2.6 g of SBA16 was pre-dried under vacuum at 200 C for 12 h before adding 3.5 g
ethoxytrimethylsilane (Aldrich) and 30 ml of dry toluene under argon. The mixture was
refluxed at 100 C for another 12 h. Then, the hydrophobicity-modified material was
filtered and washed by toluene and anhydrous ethanol in turn. At last, the solid was dried
at 80 C overnight [31].
The supported Nafion catalysts were prepared by impregnating Nafion (5 wt%
Nafion in water-alcohol solution, Dupont) on the above prepared materials (pure or
hydrophobicity-modified SBA-16), stirring at 60 C and atmospheric pressure for 6 h.
The solid was firstly dried at room temperature for 12 h in static conditions, and then the
143

water and alcohols were evaporated thoroughly under vacuum at 60 C for an additional
12 h. The resultant materials were denoted as Nafion(X)/SBA-16 or Nafion(X)/Me-SBA16, where X indicates the theoretical wt% of Nafion loading (in this work, two Nafion
loading were studied, i.e. X = 15 or 30). Me-SBA-16 indicates that the surface -OHs of
SBA-16 were silanized by ethoxytrimethylsilane. For comparison, Nafion(X)/SBA-15,
Nafion(X)/Me-SBA-15 and the Nafion resin/silica composite SAC 13, with resin content
of ca. 13 wt% obtained from Aldrich were also studied.

7.2.2. Catalyst characterization


Nitrogen adsorption-desorption isotherms at 77 K were performed using a
Micromeritics TRISTAR 3000 apparatus. The samples were degassed at 120 C and high
vacuum prior to the measurements. BET equation was used to estimate the surface areas
of the materials. BJH model was utilized to calculate the mesopore size using adsorption
branches of isotherms, and the pore diameter was estimated from the peak position of the
BJH pore size distribution.
Powder XRD spectra were recorded using a Bruker D4 X-ray diffractometer with
nickel-filtered CuIC radiation (X. = 1.5418 ). The tube voltage was 40 kV, while the
current was 40 mA. Diffraction patterns were recorded with scan step of 0.02 for 2 theta
between 0.5 and 5.
Transmission electron micrographs (TEM) were obtained using a JEOL 2011
microscope operated at 200 kV. The samples were prepared by sonication in ethanol and
suspended on holey carbon grids.
The ion exchange capacities (corresponding to acid site concentration) of the
Nafion-modified SBA-16 materials were determined using aqueous solutions of NaCl
back titrated potentiometrically by NaOH [32]. In a typical experiment, 0.1 g of solid was
added to 20 g of 2 M NaCl solution, and vigorously stirred at room temperature overnight.
The resulting suspension was filtered and washed thoroughly with a total amount of 80
ml 2 M NaCl to retain the hydrogen ions in the solution, and thereafter titrated
potentiometrically by a drop wise addition of a 0.1 M NaOH.
144

Water adsorption isotherms at 25 C were performed using an intelligent gravimetric


analyzer (IGA-002, Hiden, UK). Before the measurements, samples (ca. 50 mg each)
were degassed at 200 C for 12 h in a high-vacuum system. IGA is a very accurate,
completely computer controlled gravimetric technique which can well define the
adsorption behavior of the gas-solid system. The apparatus is an ultrahigh vacuum system.
It consists of a fully computerized microbalance which automatically measures the
weight of the sample as a function of time with the gas vapor pressure and sample
temperature under computer control.
Sulfur and carbon content were determined by elemental analysis using a Vario EL
HI apparatus (CHNS model).

7.2.3. Reaction procedure


Liquid phase alkylation of isobutane/1-butne experiments were carried out in an
automated stainless steel fixed bed continuous reactor. In each run, 0.35 g of catalyst
pellets with diameters of 0.5 to 0.8 mm, obtained by compressing the powder into tablets,
crushing, and sieving, were loaded in a reactor tube with internal diameter of 8 mm. The
catalyst bed was fixed between two plugs of quartz wool and the remaining empty
volume filled with quartz beads. The catalysts were pretreated in situ at 110 C for 10 h
in a 15 ml/min flow of N2. The reactor was cooled to reaction temperature and a pressure
of 3.2-3.4 MPa was established with nitrogen before feeding the reactants. The feedstock
consisting of an isobutane/1-butne mixture with a molar ratio of 20/1 and 40/1 was
delivered to the reactor using a piston-type pump, and it goes through the catalyst bed
located in the middle zone of the reactor. The product stream coming out of the reactor is
then collected and stored using a 16-loop sampling valve. In this way samples at intervals
of 1 min can be taken during the run and analyzed automatically once the experiment has
been finished. The reaction products were analyzed using a GC (Thermo Trace GC Ultra)
equipped with a 100m capillary column (CP SIL PON A CB) and a FID detector. The
individual C4-C8 hydrocarbons were identified by means of available reference standards
and GC-MS analysis. The conversion is calculated based on disappearance of butnes,
i.e., 1-butne, 2-butene and isobutene. The selectivity of a product was defined as the
145

ratio between the weight of the product and the weight of converted butnes. In absence
of any simultaneous reactions with isobutane-butene alkylation, the theoretical alkylate
selectivity value would be 204 wt%.

7.3. Results and discussion


7.3.1. Catalyst characterization
Nafion modified SBA-16 and Me-SBA-16 materials prepared by impregnation
respond to the need of supporting perfluorosulfonic acid resin over materials with threedimensional mesopore structure in order to provide well-dispersed active sites and allow
easier diffusion. Nafion-perfluorosulfonic acid resin has an acidity comparable to that of
concentrated sulfuric acid, providing effective active sites in alkylation reactions.
Mesoporous SBA-16 support has desired three-dimensional pore structure, ensuring
easier diffusion of products and more surface interacting with reactants (Gobin et al.,
2006; Wei et al., 2007). The influence of pore structure on the alkylation reaction can be
revealed by comparing the behaviors over Nafion modified mesoporous materials with
one-dimensional and three-dimensional channels. The surface hydrophobicity of SBA-16
can be adjusted by silylation of the surface -OHs into methylsilane. By this way, the
influence of surface polarity of SBA-16 on alkylation reaction rate can be established.
Powder XRD patterns of SBA-16, Me-SBA-16 and their Nafion-modified samples
are depicted in Fig. 43, showing the effect of silylation of surface OHs and impregnation
of Nafion resin on the mesostructured SBA-16. It is clearly observed that all the samples
show one strong diffraction peak in the range of 0.85-0.90, assigned to (110). The
secondary diffraction peaks of (200) and (211) are slightly showed in the samples of
SBA-16, demonstrating the well defined Im3m mesophase of SBA-16. The intensity of
the secondary diffraction peaks decreased when the materials are modified by silylation
or Nafion impregnation. XRD lines normally decrease in intensity as guest species are
loaded in the mesopores, and such an intensity change can be useful as a means of
judging the location of the guest species (Choi et al., 2005). Therefore, this suggests

146

Nafion resin and methyl groups are located in the mesopores. The obvious peaks of (110)
of all samples indicate that the cubic mesostructure is maintained in all cases.
The N2 adsorption-desorption isotherms and pore size distributions of SBA-16
and Me-SBA-16 with various Nafion loading are shown in Fig. 44. All the samples
exhibit a sharp capillary condensation step on the desorption branch at plpo = 0.45, which
are typical features of ordered Im3m mesostructure. The narrow IV-type hysteresis loop
of SBA-16 material is maintained after the grafting of trimethylsilane, indicating that the
structure is not drastically affected by silylation. The mesostructure is also maintained
after impregnation of 15 wt% of Nafion, which means the Nafion resin having penetrated
into the porous framework seems to be homogeneously deposited along the threedimensional mesoporous channels of the materials when the Nafion loading is lower than
15 wt%. With increasing Nafion loading to 30 wt%, the capillary condensation step on
desorption branch of Nafion(30)/Me-SBA-16 becomes broader while no significant
broadening is found in the sample Nafion(30)/SBA-16. Meanwhile, the pore size
distribution of Nafion(30)/Me-SBA-16

is also a little broader than that of

Nafion(30)/SBA-16. Part of the resin on Nafion(30)/Me-SBA-16 may be within the


mesoporous structure in the form of polymeric aggregates. The textural parameters
deduced from nitrogen isotherms and XRD patterns of the synthesized materials are
summarized in Table 16. Impregnation of up to 30 wt% Nafion on SBA-16 has only
slight effect on BET surface area, pore diameter and wall thickness. Decreases of BET
surface area and pore diameter, and an increase in wall thickness can be observed on MeSBA-16. This is indicative of the presence of Nafion resin aggregates in the channels of
Me-SBA-16. Interestingly, the pore diameter increases from 4.4 nm (SBA-16) to 4.8 nm
(Me-SBA-16) when SBA-16 is silanized by ethoxytrimethylsilane. We believe this is due
to a leaching process during the silylation procedure. This is coherent with the decrease
of wall thickness from 8.0 nm to 7.1 nm after silylation. Though the textural parameters
changed a little after OH capping and Nafion loading, the three-dimensional
mesostructure of SBA-16 was not affected, which makes these materials suitable
catalysts for alkylation.

147

The mesoporous structure of Nafion modified SBA-16 materials was characterized


using transmission electron microscopy (TEM). Fig. 45 shows a typical TEM image of
Nafion(30)/Me-SBA-16. Well-ordered cubic pore structure of ca. 4.5 nm in diameter is
clearly observed, which is in a good agreement with the analysis of XRD and N2
physisorption.
Table 17 shows several parameters related to the chemical properties including the
acidity and methyl incorporation of the hybrid materials. Methyl content calculated from
carbon and sulfur content measured by elemental analysis in all silanized samples is ca.
0.4 mmol/g. It indicates that methyl groups were grafted on SBA-16 and no leaching
happened during impregnation of Nafion. The experimental Nafion loading of each
sample is also evaluated using elemental analysis method. The incorporation yield refers
to the percentage of Nafion resin deposited on support. It ranges from 81 wt% to 85wt%,
which demonstrates that the Nafion loading (between 15 wt% to 30 wt%) and polarity of
SBA-16 did not make great effect on incorporation yield. H+concentrations determined
by drop-wise potentiometric titration are similar to sulfur concentrations. The results
evidence that the active sites provided by Nafion resin are highly accessible, revealing the
effective role of the mesostructured silica support.
Methyl capping of surface -OHs, though not providing the active sites as the Nafion
resin, can play an important role in tuning the physical property of the surface of the
catalyst. It may enhance the hydrophobicity of catalysts, which is beneficial to increasing
the surface isoparaffin/olefin ratio in the alkylation reaction conditions, resulting into
good catalytic performances. To compare the hydrophobicity of the samples, H 2 0
adsorption measurements were carried out using an IGA system. Fig. 46 depicts water
adsorption isotherms of SBA-16, Me-SBA-16 and their Nafion modified samples.
Compared with SBA-16, Me-SBA-16 takes less water over the whole pressure range (010 millibar), indicating the hydrophobicity of methyl-modified material. A different
water adsorption behavior could be observed when these two materials were modified by
Nafion resin. The IGA curves for water adsorbed over Nafion(30)/SBA-16 and
Nafion(30)/Me-SBA-16 can be divided into two parts. The mass uptake of water is
slightly higher over Nafion(30)/Me-SBA-16 than over Nafion(30)/SBA-16 when the
148

water vapor pressure is less than 1.8 millibar. This is due to the introduction of
perfluorosulfonic acid groups. Water molecules interact strongly with acid sites. They
adsorb preferentially on perfluorosulfonic acid groups when the samples are modified
with Nafion resin. Because the interaction between acid groups of Nafion and the silanols
of the silica leads to a decrease in acid strength [6, 33], the acid sites of Nafion(30)/MeSBA-16 are more acidic and can take more water molecules at low vapor pressure.
However, the difference in acidity is not significant; the mass uptake of water over
Nafion(30)/Me-SBA-16 is just slight higher. With the increasing of vapor pressure to
more than 1.8 millibar, the situation is different. More water was retained over
Nafion(30)/SBA-16. The acid sites were saturated and water molecules adsorbed mainly
on surface of SBA-16 or Me-SBA-16. The results indicate that the hydrophobicity is
enhanced when the silica is silanized by ethoxytrimethylsilane.

7.3.2. Catalytic performance


The alkylation of isobutane with -butne over solid acid catalysts has been studied
intensively over the past two decades [1, 34, 35]. Solid acids deactivate by the buildup of
a polymer, which eventually blocks the pores of the catalyst. The pore structure of solid
acid catalysts plays very important roles in alkylation. Numerous transfer steps obviously
occur in the pores of the solid catalysts. Isobutane and olefins diffuse inward in the pores,
and product molecules diffuse outward. The rates of transfer or diffusion vary for
different molecules because of differences in size, shape, molecular weight and
interaction with the pore walls [36]. In the pores, numerous reaction steps occur.
Conditions in the pores are conductive to forming, in addition to alkylate, both conjunct
polymers and pseudo-alkylate. The diffusion of conjunct polymer is slow. The conjunct
polymers would be difficult to diffuse out or may bond with the inner surfaces of the
solid catalyst. Diffusion limitations would lead to faster deactivation. Particular attention
should be paid to the role of the structural characteristics of the catalysts. Nafion
modified mesoporous materials are solid acids with high surface area, large pore size and
high strength acid sites. They are promising solid catalysts for isobutane/w-butene. It is
meaningful to study the effect of pore structure and hydrophilic-hydrophobic nature of
Nafion modified mesoporous materials on the catalytic performance in alkylation.
149

Nafion modified mesoporous materials with different structure and hydrophilichydrophobic nature were tested and compared with SAC-13, which was reported to be a
good catalyst for isobutane/-butene alkylation [24]. Fig. 47 displays the catalytic
performance of catalysts in terms of butnes conversion, TMPs (trimethylpentanes) yield
and C5-C7 (hydrocarbons with 5 to 7 carbon atoms) selectivity as a function of time on
stream. Relevant textural and catalytic properties are summarized in Table 18.
It can be observed the activity of Nafion modified materials with similar H+
concentration (-0.13 mmol H+/gcat) decreases in the order: Nafion(15)/SBA-16 > SAC13

>

Nafion(15)/SBA-15.

The

activity

of

three-dimensional

mesostructured

Nafion(15)/SBA-16 exhibits much higher initial activity (99.1 wt% vs. 54.2 wt%) and
remains at a higher level in comparison with the one-dimensional Nafion(15)/SBA-15,
though Nafion(15)/SBA-15 contains larger mesopores and pore plugging is typically
anticipated to have only negligible effects on large-pore solid acids. We believe that the
good catalytic performance of Nafion(15)/SBA-16 for alkylation is mainly due to its pore
structure. Three-dimensional mesopores structure is beneficial for isobutane and olefins
diffusing inward in the pores, and product molecules diffusing outward. At the same time,
the slightly lower surface area of Nafion(15)/SBA-16 is also beneficial to its activity.
There is more interaction between the Nafion resin and the silica support on a larger
surface. In this way, a charge transfer may occur from the protons of the sulfonic groups
to the silanol groups of the silica, decreasing the acid strength of the resin and activity of
alkylation [6, 33]. The influence of surface area on acid strength can be confirmed by the
initial selectivity of C5-C7, which is the result of cracking and is catalyzed only by the
strongest acid sites [2]. It can be observed in Fig. 47 that the selectivity of C5-C7 increase
in the order: Nafion(15)/SBA-15 < Nafion(15)/SBA-16 < SAC-13. SAC-13 is a material
with the perfluorosulfonic Nafion resin supported over an amorphous silica via sol-gel
techniques. It has the smallest surface area and exhibits the highest initial C5-C7
selectivity. Anyway, Nafion(15)/SBA-16 exhibits much better conversion and TMPs
yield than SAC-13 and Nafion(15)/SBA-15. A better availability of acid sites seems to be
achieved in the mesostructured silica SBA-16 support. It is reasonable to attribute the
good catalytic performance of Nafion(15)/SBA-16 mainly to the three-dimensional
mesopores structure. In a recent study from our laboratory, it was found that the
150

diffusivity of small hydrocarbons was systematically two orders of magnitude higher in


SBA-16 compared to SBA-15 [21]. Three-dimensional structures of mesoporous solid
acids which allowed easier diffusion should be preferred to one-dimensional ones.
With the increase in Nafion loading on SBA-16 (from 15 wt% to 30 wt%), the
number of acid sites which catalyze isobutane/1 -butne alkylation increases. The butnes
conversion of Nafion(30)/SBA-16 remains at a higher level in comparison with the
Nafion(15)/SBA-16. High Nafion content favors the formation of TMPs and C5-C7
products.
It can also be observed in Fig. 47 that the activity and stability of Nafion(30)/SBA16 are obviously improved after its surface OHs were capped by ethoxytrimethylsilane.
Though both materials give very high initial activity value, Nafion(30)/Me-SBA-16 gives
evidently higher butnes conversion and TMPs yields. The improvement of catalytic
performance may be attributed to stronger acidity and three-dimensional pore system.
The acid sites of Nafion(30)/Me-SBA-16 are just slightly stronger than those of
Nafion(30)/SBA-15 which did not induce evident differences on the C5-C7 selectivity. In
other words the OH capping does not seem to increase activity through increasing the
acid strength. Meanwhile, catalyst characterization indicated there are no essential
structure changes induced by the capping procedure. We believe that the improved
activity and stability of OHs capped materials for this reaction is a result of its
hydrophobic surface. The hydrophobic environment of the capped catalyst allows a
higher steady isobutane concentration in the pore system. During isobutane/1-butne
alkylation reaction, the octyl carbenium ion is removed from the active site by hydride
transfer from isobutane leading to trimethylpentanes as ideal products and an isobutyl ion,
which perpetuates the reaction. The high ratio of the rate of hydride transfer vs. the rate
of oligomerization is crucial for a good catalytic performance, while the isobutane/olefin
ratio determines the rate of hydride transfer. To enhance surface carbophilicity by
capping the silanols of SBA-16 increases the I/O ratio in the reaction conditions and
therefore improves the catalytic performance of the catalysts. The fact that the pore
surface hydrophobicity is enhanced by OHs capping was confirmed by IGA results. Less

151

water was adsorbed on the surface of Me-SBA-16 and Nafion(30)/Me-SBA-16 in


comparison to SBA-16 and Nafion(30)/SBA-16.
The influence of pore structure on the activity and selectivity for the OH capped
Nafion modified mesoporous materials is shown in Table 19. On hydrophobic-modified
samples with similar Nafion content but with a different pore structure, the catalyst with
three-dimensional structure deactivates more slowly. At 100 C, the butnes conversion
decreases from 100 wt% at 1 min TOS to 88.6 wt% at 10 min TOS over Nafion(30)/MeSBA-16 while it decreases from 98.6 wt% to 39.8 wt% over Nafion(30)/Me-SBA-15.
Moreover, decrease in selectivity to TMPs is also slower on three-dimensional catalysts.
The TMPs/Cg ratio decreases from 65.7 wt% at 1 min TOS to 44.2 wt% at 10 min TOS
over Nafion(30)/Me-SBA-16 while it decreases from 65.5 wt% to 35.6 wt% over
Nafion(30)/Me-SBA-15. The three-dimensional mesopores solid acid outperforms the
one-dimensional one both in butnes conversion and TMPs selectivity.
The adsorption and diffusion of products and reactants are affected by reaction
temperature. Compared with liquid acids, a higher reaction temperature is required when
solid acids are used. This is attributed to the lower strong acid site concentration of sold
acids or the lack of solvation, resulting in higher activation energies for the individual
reaction steps. Efficient mobility of products and reactants in the pores of solid acids also
requires higher temperature. The poor performance of several solid acids at low reaction
temperature is most likely a consequence of the hindered diffusion of bulky molecules
under such conditions. The catalyst will be prematurely deactivated by pore blocking. In
Table 19, catalytic performances at 100 C and 70 C over Nafion(30)/Me-SBA-15 and
Nafion(30)/Me-SBA-16 are also presented. The inferiority of one-dimensional catalyst is
intensified at lower temperature. Diffusion problem is more severe for one-dimensional
materials. The results obtained suggest that three-dimensional structures, both in
hydrophilic and hydrophobic environment, are capable of diffusing molecules easily and
resisting deactivation by pore plugging. Three-dimension mesopores structure and
hydrophobic surface should be desired for an alkylation catalyst.

152

As with alkylation reaction over other solid acid catalysts, a close interconnection
exists between the catalytic performances of Nafion/Me-SBA-16 and the operation
condition. The most important parameters are the reaction temperature, the olefin space
velocity, and the feed isoparaffin/olefin ratio. Fig. 48 displays the influence of reaction
parameters on the catalytic performance of Nafion(30)/Me-SBA-16.
It can be seen that with same 1-butne space velocity and isobutane/1-butne ratio,
the conversion of butnes increase obviously with temperature. The initial conversion of
butnes increases from 88.6 wt% to 100 wt% when the reaction temperature increases
from 70 C to 100 C (WHSV = 2 h 1 , I/O = 40). The yield of TMPs is presented in the
same manner except the beginning stage (at 1 min TOS). The decline in the TMPs yield
with rising temperature during beginning stage is compensated by an increase in the
cracking selectivity.
The influence of WHSV is also shown in Fig. 48. When WHSV of 1-butne
decreases from 2 h"1 to 1.4 h"1 (I/O = 40, T = 100 C), the catalyst decay is much slower.
As the WHSV of 1-butne increases, the contact time decreases, resulting in an increase
of the C f and C9+ fractions, which not only decreases the selectivity of TMPs but also
fastens the catalyst decay.
The isoparaffin/olefin ratio is a very important parameter in isobutane/-butene
alkylation. Lower I/O ratio leads to higher rate of oligomerization. It can be seen in Fig.
48 that I/O shows obvious effect on catalytic performance. The conversion of butnes
decreases when the I/O ratio decreases from 40 to 20 (WHSV = 2 h"1, T = 100 C). High
I/O ratio obviously favours the yield of TMPs. However, high I/O ratio will leads to high
cost of separating products from the excess isobutane, which limits the I/O ratio that can
be used industrially.

7.3.3. Conclusions
Mesostructured

SBA-16

hydrophobically

modified

by

OH

capping

and

functionalized with Nafion resin by means of impregnation was described. The resulting
solid acid catalysts preserve the three-dimensional mesostructure of SBA-16. Capping of
153

surface OH decreases the materials polarity and provides a hydrophobic environment for
the isobutane/1-butne reaction.
The impact of structure of Nafion modified mesoporous materials for activity and
selectivity during the alkylation of isobutane with 1 -butne was discussed. The threedimensional mesopores solid acid outperforms one-dimensional ones both in butnes
conversion and TMPs selectivity. Three-dimensional mesopores structure is capable of
diffusing molecules easily and resisting deactivation by pore plugging.
Hydrophobically modified Nafion(30)/Me-SBA-16 shows excellent activity and
efficiency in the production of isooctane compared with Nafion modified SBA-15 and
SAC-13. It is suggested that three-dimension mesopores structure and hydrophobic
surface should be desired when designing an alkylation catalyst.

Acknowledgments
This work was supported by NSERC of Canada, Science & Technology Commission of
Shanghai Municipality (08DZ2270500) and "863" key project (2009AA033701).

References
[1]

A. Corma, and A. Martinez, Catal. Rev.-Sci. Eng., 35 (1993) 483-570

[2]

A. Corma, A. Martinez, C. Martinez, J. Catal. 149 (1994) 52-60.

[3]

A. S. Chellappa, R. C. Miller, W. J. Thomson, Appl. Catal. A 209 (2001) 359-374.

[4]

T. Blasco, A. Corma, A. Martinez, P. Martinez-Escolano, J. Catal. 177 (1998)


306-313.

[5]

V. Sarsani, Y. Wang, B. Subramaniam, Ind. Eng. Chem. Res. 44 (2005) 64916495.


154

[6]

P. Botella, A. Corma, J. M. Lopez-Nieto, J. Catal. 185 (1999) 371-377.

[7]

C. J. Lyon, V. S. R. Sarsani, B. Subramaniam, Ind. Eng. Chem. Res. 43 (2004)


4809-4814.

[8]

S. I. Hommeltoft, Appl. Catal. A 221 (2001) 421-428.

[9]

A. W. Chester, Y. F. Chu, Zeolite 6 (1986) 195-200.

[10] A. Corma, A. Martinez, C. Martinez, Catal. Lett. 28 (1994) 187-201.


[11] M. Stocker, H. Mostad, T. Rorvik, Catal. Lett. 28 (1994) 203-209.
[12] F. Cardona, N. S. Gnep, M. Guisnet, G. Szabo, P. Nascimento, Appl. Catal. A 128
(1995)243-257.
[13] A. Corma, A. Martinez, P. A. Arroyo, J. L. F. Monteiro, E. F. Sousa-Aguiar, Appl.
Catal. A 142 (1996) 139-150.
[14] D. E. Sherwood, R. J. Taylor, Appl. Catal. A 155 (1997) 195-215.
[15] G. S. Nivarthy, Y. He, K. Seshan, J. A. Lercher, J. Catal. 176 (1998) 192-203.
[16] R. Loenders, P. A. Jacobs, J. A. Martens, J. Catal. 176 (1998) 545-551.
[17] A. Feller, A. Guzman, I. Zuazo, J. A. Lercher, J. Catal. 224 (2004) 80-93.
[18] M. Mukhergee, J. Nehlsen, Hydrocarbon Process. 86 (2007) 110-114.
[19] A. Feller, I. Zuazo, A. Guzman, J. Barth, J. A. Lercher, J. Catal. 216 (2003) 313323.
[20] W. Shen, D. Dube, S. Kaliaguine, Catal. Commun. 10 (2008) 291-294.
[21] O. C. Gobin, Q. Huang, H. Vinh-Thang, F. Kleitz, M. Eic, S. Kaliaguine, J. Phys.
Chem. C. 111 (2007) 3059-3065.
[22] K. Yoo, E. C. Burckle, P. G. Smirniotis, J. Catal. 211 (2002) 6-18.
155

[23] M. A. Harmer, W.E. Farneth, Q. Sun, Adv. Mater. 10 (1998) 1255-1257.


[24] P. Kumar, W. Vermeiren, J. P. Dath, W. H. Hoelderich, Energy Fuels 20 (2006)
481-487.
[25] M. Alvaro, A. Corma, D. Das, V. Fornes, H. Garcia, Chem. Commun. (2004)
956-957.
[26] D.J. Macquarrie, S.J. Tavener, M.A. Harmer, Chem. Commun. (2005) 2363-2365.
[27] F. Martinez, G. Morales, A. Martin, R. van Grieken, Appl. Catal. A 347 (2008)
169-178.
[28] S. Wang, J.A. Guin, Energy Fuels 15 (2001) 666-670.
[29] W. Shen, Y. Gu, H. Xu, D. Dube, S. Kaliaguine, Alkylation of Isobutane/1-butne
on Methyl-modified Nafion/SBA-16 Materials, Applied Catalysis A, General
(2009) accepted
[30] D. Zhao, Q. Huo, J. Feng, B.F. Chmelka, G.D. Stucky, J. Am. Chem. Soc. 120
(1998) 6024-6036.
[31] S. Parambadath, M. Chidambaram, A.P. Singh, Catal. Today 97 (2004) 233-240.
[32] D. Margolese, J. A. Molero, S. C. Christiansen, B. F. Chmelka, G. D. Stucky,
Chem. Mater., 12 (2000) 2448-2459.
[33] I. Palinko, B. Torok, G. S. K. Prakash, G. A. Olah, Appl. Catal.A Gen. 174, (1998)
147-153.
[34] A. Feller, J. A. Lercher, Adv. Catal. 48 (2004) 229-295
[35] L. F. Albright, Ind. Eng. Chem. Res. 48 (2009) 1409-1413
[36] T. K. Sherwood, N. L. Pigford, C. R. Wilke, In Mass Transfer, McGraw-Hill, Inc.,
New York (1975) pp 17-35, 379-382.
156

Table 16. Textural properties for Nafion-modified mesostructured SBA-16


materials
Textural properties for Nafion-modified mesostructured SBA-16 materials
BET

Pore

Pore

area

diameter3

volume2

(V/g)

(nm)

(cm3/g)

SBA-16

404

4.4

Nafion(15)/SBA-16

392

Nafion(30)/SBA-16

Sample

^110

ao C

Wall
thicknessd

(nm)

(nm)

0.33

10.2

14.4

8.0

4.3

0.20

10.0

14.2

8.0

388

4.4

0.15

10.0

14.2

7.9

Me-SBA-16

409

4.8

0.20

9.7

13.7

7.1

Nafion(15)/Me-SBA-16

356

4.7

0.165

9.9

14.0

7.5

Nafion(30)/Me-SBA-16

234

4.3

0.11

9.8

13.9

7.7

(nm)

Total pore size and pore volume of Nafion-modified SBA-16 materials were calculated
by BJH adsorption branch, and the pore diameter was pointed out from the peak
position of BJH pore size distribution.
b

d (110) spacing, measured from small-angle XRD.

_ ,1/2 ,

ao-2
d

duo

Pore wall thickness = 3l/2ao/2 - pore diameter

157

Table 17. Methyl incorporation and acidic-related properties of Nafion modified


mesostructured materials
Sample

Methyl content

Acidity

Nafion
incorporation
Yield %d

(mmol

(mmol

(mmol

methyl/g)a

H+/gcat)b

S/gcat)c

SBA-16

Nafion(15)/SBA-16

0.132

0.126

84

Nafion(30)/SBA-16

0.264

0.255

85

Me-SBA-16

0.40.1

Nafion( 15)/Me-SBA-16

0.40.1

0.130

0.123

82

Nafion(30)/Me-SBA-16

0.40.1

0.256

0.243

81

Methyl content calculated from elemental analysis.


b

Obtained by cationic-exchange in 2 M NaCl, and titrated potentiometrically with 0.1 M

NaOH after filtration and thorough wash.


c

Sulfur content calculated from elemental analysis..

Yield of incorporation of Nafion was determined by comparing the amount of sulfur

deposited and introduced.

158

Table 18. Textural and catalytic properities of catalysts


Sample

Textural properties
SBET

Dpoie

vPorea

(m2/g)

(nm)

(cm3/g)

Nafion(30)/Me-SBA-16

234

4.3

Nafion(30)/ SBA-16

388

Nafion( 15)/ SBA-16

Catalytic properties
mmol S/gca,"

mmol H7gca, c

0.11

0.243

0.256

4.4

0.15

0.255

0.264

392

4.3

0.20

0.126

0.132

Nafion( 15)/ SBA-15

474

7.0

0.65

0.124

0.130

SAC-13

200

0.43

0.147

0.133

Total pore size and pore volume were calculated by BJH method.

Sulfur content evaluated by elemental analysis.

Obtained by cationic-exchange with NaCl and titration with NaOH.

from ref. [29]

159

Table 19. Variation of the butnes conversion and products distribution at different
reaction temperature and time on stream over Nafion(30)/Me-SBA-15 and
Nafion(30)/Me-SBA-16 catalysts.
Catalyst:
Temperature(C)

Nafion(30)/Me-SBA-16
100

100

70

100

100

70

10

10

54^0

88^6

98^6

39^8

7~

TOS (min)
C4= Conv. (wt%)

Nafion(30)/Me-SBA-15

Distribution of C 5+ (wt%)
C5-C7

42.2

9.5

30.0

42.7

8.8

27.3

C8

51.3

66.8

62.9

48.2

65.9

64.4

C9+

6.6

23.7

7.1

9.1

25.3

8.0

Distribution of C 8 (wt%)
TMP

65.7

44.2

70.8

65.5

35.6

52.8

DMH

30.0

41.4

25.4

33.4

50.0

37.6

Cg"

4.3

14.4

3.7

1.1

14.4

9.6

Other experimental conditions: I/O ratio = 40, olefin WHSV = 2 h"1.

160

SBA-16

Nafion(15)/SBA-16

Nafion(30)/SBA-16

Me-SBA-16
Nafion(15)/Me-SBA-16
Nafion(30)/Me-SBA-16

*_

/_!

Figure 43. X-ray diffraction patterns of SBA-16, Me-SBA-16 and their Nafionmodified samples.

161

(a)

250

SBA-ie

225

Nafion(15)/SBA-16

wv-

200-

SBA-16

-* * ***

Nafion(30)/SBA-16
/jM

Nafion(15)/Me-SBA-16

>
-a

J,
\

f r**"1
/?

Nafion(15)/SBA-16

Nafion(30)/SBA-16
m->i-i-ui-

.^vlarion(30)/Me-SBA-16
IOOOOOOSCSSOXUS

3 0 - 0 0-0

O-OODD-

V
A!
<?

b 4

Nafion(15)/Me-SBA-16
4-44-44

Nafion(30)/Me-SBA-16

^teCODCCD OCXl-O O O
0.0

0.2

0.4

0.6

0.8

4-4-44-4-

O-OO-O-

1.0

rwn

Figure 44. (a) V adsorption/desorption isotherms and (b) pore size


distributions calculated from the adsorption branch by BJH method.

162

Figure 45. TEM images of Nafion(30)/Me-SBA-16 showing characteristic planes


for a cubic pore structure: (a) [111], (b) [100], (c) [120]

163

Water Vapor Pressure (millibars)

Figure 46. Adsorption isotherms of water in SBA-16, Me-SBA-16,


Nafion(30)/SBA-16 and Nafion(30)/Me-SBA-16 at 25 C.

164

4
TOS (min)

10

80
60
&

40

20

10

TOS (min)
50
NT

^*

40

.i.
u
U
U

m
rO
U

30
20
10

10

TOS (min)
Figure 47. Catalytic performance of Nation modified mesoporous materials.
Experiment conditions: T - 100 C; olefin WHSV = 2 h"1; isobutane/1-butene (I/O)
molar ration=40. Nafion(30)/Me-SBA-16 (A), Nafion(30)/SBA-16 ( ) ,
Nafion(15)/SBA-16(0), SAC -13 ( ) , Nafion(15)/SBA-15 (A)
165

100

80

c
c
fe

c
o
O

60

40

20

10

15
TOS (min)

20

25

100

80

60
09
DH

40

20

10
15
TOS (min)

20

25

Figure 48. Variation of the conversion of butnes (a) and yield of TMPs (b) with
time on stream over Nafion(30)/Me-SBA-16 under varying experimental conditions:
100 C, WHSV = 1.4 h \ I/O = 40 ( ) ; 100 C, WHSV = 2 h \ I/O = 40 (A); 70 C,
WHSV = 2 h ', I/O = 40 (O); 100 C, WHSV = 2 h \ I/O = 20 (A).

166

8. Conclusions and Recommendations


8.1 Summary
In this thesis, a series of perfluorosulfonic acid functionalized mesostructured
materials were synthesized and studied as catalysts in the alkylation of isobutane/1butne. The effect of strength of acid sites, structural and hydrophilic -hydrophobic
nature of mesoporous substrates was examined, aiming to develop more effective, time
stable and clean catalysts.
In Chapter 2 of this thesis, an up to date and in-depth overview about alkylation of
isobutane/n-butene on solid acids and synthesis of mesostructured solid acids is given.
Special attention is paid to the mechanism of deactivation and catalyst requirements. The
formation of heavy hydrocarbons is the major reason for deactivation of catalysts. Heavy
hydrocarbons may plug catalyst pores or block active catalyst sites, leading to
deactivation. Ideally, hydride transfer from isobutane to the adsorbed Cg species occurs
before further alkylation or alkene desorption, thus terminating the growing chain and
preventing heavy hydrocarbon buildup. So the lifetime of the alkylation catalyst is mainly
determined by the relative rates of hydride transfer and of oligomerization. The higher the
hydride transfer to oligomerization ratio, the higher the yield in alkylation products and
the smaller the formation of carbonaceous deposits hence the slower the deactivation. In
order to promote hydride transfer, alkylation is performed industrially in a large excess of
isobutane over butne to maximize the relative hydride transfer rate. Hydride transfer
requires strong acid sites with in addition a significant positive effect of their density;
moreover, this reaction involves a bulky bimolecular transition state. As a consequence, a
catalyst with large pores is expected. Thus, a catalyst with large pore size which provides
enough space for diffusion, high acid strength comparable to that of sulfuric acid, and
nonpolar surface which allows a higher paraffin steady concentration in the pores is
desired. The methods to synthesize acid functionalized catalysts with mesoscale pores
and to tune the acid strength and hydrophilicity-carbophilicity of the surface are also
reviewed in Chapter 2.

167

Chapter 3 describes the materials, apparatus and experimental procedures that were
employed during the performing of this Ph.D. work.
In Chapter 4 a series of mesostructured solid acids including arene-sulfonic and
propyl-sulfonic acid functionalized PMO, aluminum chloride grafted SBA-15 (Al-SBA15) and PMO (Al-PMO), sulfonated mesoporous polystyrene coated SBA-15 (PSSBASO3H), perfluoroalkylsulfonic acid functionalized SBA-15 (FS-SBA-15) and PMO (FSPMO) were synthesized. All these materials have the characteristic of 2-d hexagonal
(P6mm) structure with good textural uniformity. Acid groups were grafted on the pore
surface. Nitrogen adsorption-desorption isotherms for these materials are of type IV,
indicating the mesoporous nature of the solids. The acid strength increases in the order:
PrS-PMO < ArS-PMO (ArS-SBA-15, PSSBA-SO3H) < Amberlyst-15 < Al-PMO (AlSBA-15) < Zeolite Y < Nafion (FS-PMO, FS-SBA-15). The predominant acid species on
arene-sulfonic acid, propyl-sulfonic acid and perfluoroalkylsulfonic acid functionalized
PMO are Bronsted acid sites while the predominant acid species on Al-PMO and AlSBA-15 are Lewis acid sites. Acid strength of catalysts is the key factor of activity.
Alkylation of isobutane with w-butene is dramatically influenced by the strength of acid
sites. Only the catalysts with acid sites stronger than amberlyst-15 can catalyze the
alkylation reaction. The materials with weaker acid strength such as PrS-PMO are
inactive for alkylation. Not only the Bronsted acidity has activity for alkylation, strong
Lewis acidity also shows good alkylation activity. Large surface area and high acid
loading are very important for alkylation. Nafion and FS-SBA with very strong acid sites
are active but with relatively low conversion of butnes. The very low surface area of
Nafion and low acid loading of FS-SBA-15 limit their activity. FS-PMO exhibits the best
catalytic activity in the alkylation of isobutane by 1 -butne. Its high and stable activity
can be attributed to its higher acid strength, higher carbophilicity of surface and larger
pore size compared with that of zeolite Y.
Chapter 5 reports the synthesis, characterization and catalytic evaluation of
perfluoroalkylsulfonic acid functionalized PMO (periodic mesoporous organosilica).
Hybrid organic-inorganic PMO was synthesized by the acid-catalyzed hydrolysis and
condensation of BTME (bis-(trimefhoxysilyl) ethane) using the triblock copolymer
168

Pluronic PI23 as the structure-directing agent. PMO was functionalized with perfluoroalkylsulfonic acid group by grafting

l,2,2-trifluoro-2-hydroxy-l-trifluoromethylethane

sulfonic acid yff-sultone on its surface. The resulting material has the characteristic of 2-d
hexagonal (P6mm) structure with good textural uniformity. It exhibits better catalytic
activity as well as catalytic stability compared to conventional solid acid catalysts. The
effects of the most important parameters which must be controlled in order to improve
the catalyst stability in this critically important catalytic process were also discussed.
Chapter 6 was focus on effect of hydrophilicity-carbophilicity of the surface on
alkylation of isobutane/1-butne. Mesostructured SBA-15 hydrophobically modified by
OH capping step and functionalized with Nafion resin by means of impregnation was
described. The resulting solid acid catalysts preserve the mesostructure of SBA-15 and
y

show high surface area (ca. 400 m /g) and narrow pore size distribution centred at the
mesoscale range (ca. 1 nm). Capping of surface OH diminishes the surface silanol
density and provides hydrophobic enviroment
Hydrophobically modified Nafion(30)/Me-SBA-15

for isobutane/1-butne

reaction.

shows excellent activity and

efficiency in the production of isooctane compared with Nafion/SBA-15 and SAC-13.


The catalytic performance of Nafion/SBA-15 is obviously improved after its surface OHs
were capped. Enhanced surface carbophilicity and increased the I/O ratio in the reaction
condition, results in higher ratio of hydride transfer/oligomerization and good catalytic
performance. The catalytic activity of hydrophobic Nafion/Me-SBA-15 is much superior
to that of hydrophilic Nafion/SBA-15 especially at low I/O ratio.
Chapter 7 discusses the impact of the structure of mesoporous materials on activity
and selectivity during the alkylation of isobutane with 1-butne. SBA-15 with one
dimensional channels and SBA-16 with three dimensional channels were selected as
supports. They were modified by OH capping and functionalized with Nafion resin by
means of impregnation. The mesostructure of SBA-15 and SBA-16 was preserved and
OH capping provided a more hydrophobic enviroment for alkylation. Nafion/SBA-16
with or without OH capping exhibited better catalytic performances even though their
surface area and pore diamenter are lower than that of SBA-15. A three dimensional pore
structure effectively enhances the diffusion of reactants to the active site and diffusion of
169

the products out of the pore network. Three dimensional Nafion/Me-SBA-16 is a


promising solid catalyst for alkylation of isobutane/n-butene.

8.2. Recommendations
The following are some recommendations for future works:
1. Further study the synthesis and catalytic application of other mesoporous solid
acids, such as Nafion/Mesoporous carbon, Nafion/Mesoporous polymer.

Perform

covalent attachment of perfluoroalkylsulfonic acid groups on mesoporous materials by


Si-C bond. Optimize the synthesis conditions of mesoporous solid acids with higher acid
site concentration
2. Further study the relationship between acid strength, acid strength distribution and
alkylation activity. Measure the acid strength and acid strength distribution of solid acids
by microcalorimetry.
3. Further investigate the catalyst deactivation process.
4. Comparatively study the adsorption-desorption isotherms of isobutane and 1butene on solid acids with different surface polarity by intelligent gravimetric analysis
method.
5. Further investigate diffusion mechanism, for example, the diffusion path of
isobutane, n-butene and products of alkylation in mesostructured solid acids with
different pore structure.
6. Study the isobutane/alkene alkylation under conditions of possible industrial
application. For example, a mixture of C4 alkenes was used as a source of alkenes instead
of pure alkenes.

170

9. Scientific Contributions
9.1 Publications

"Alkylation

of

Isobutane/1-butne

over

Periodic

Mesoporous

Organosilica

Functionalized with Perfluoroalkylsulfonic Acid Group"


Shen W., Dub D., Kaliaguine S., Catal. Commun., 10, 291 (2008).

"Perfluorinated

Alkylsulfonic

Acid F unctionalized

Periodic

Mesostructured

Organosilica: a New Acidic Catalyst"


Dub D., Rat M., Shen W., Nohair B., Bland F., Kaliaguine S., Appl. Catal. A, 358,
232 (2009).

"Perfluoroalkylsulfonic acid Functionalized Periodic Mesostructured Organosilica: a


Strongly Acidic Heterogeneous Catalyst"
Dub D., Rat M., Shen W., Bland F., Kaliaguine S., J. Mater. Sci. 44, 6683 (2009).

"Alkylation of Isobutane/1 -butne on Methyl-modified Nafion/SBA-15 Materials"


Shen W., Gu Y., Xu H., Dub D., Kaliaguine S., Appl. Catal. A, in press.

"Alkylation of Isobutane/1-butne on Methyl-modified Nafion/SBA-16 Materials"


Shen W., Gu Y., Xu H., Dub D., Kaliaguine S., Ind. Eng. Chem.Res., submitted.

9.2 Oral presentations

"Alkylation of Isobutane/1-butne over F unctionalized Periodic Mesoporous


Organosilicas"
Shen W., Dub D., Bland F., Kaliaguine S., CSC 2008, Kingston, Canada

171

"Acid F unctionalized Hybrid Organic/inorganic Mesostructured Materials: A New


Family of Acid Catalysts"
Dub D., Shen W., Rat M., Bland F., Kaliaguine S., ICCE-15,2007, Haikou, China

9.3 Posters

"Alkylation of Isobutane/1-butne over F unctionalized Periodic Mesoporous


Organosilicas"
Shen W., Dub D., Bland F., Kaliaguine S., CSC 2008, Kingston, Canada

172

10. References
>

Albright, L. F., Spalding, M. A., Faunce, J., Eckert, R. E., Ind. Eng. Chem. Res. 27,
391 (1988).

>

Albright, L.F., Oil Gas J. 12, 79 (1990).

>

Albright, L. F. and Wood, K. V., Ind. Eng. Chem. Res. 36,2110 (1997).

>

Albright, L. F., Chemtech. July, 46 (1998).

>

Albright, L. F., Alkylation-Industrial. In Encyclopedia of Catalysis; Howath, I. T.,


Ed.; John Wiley and Sons: New York, Vol. 1, 226-281 (2003).

>

Albright, L. F., Ind. Eng. Chem. Res. 48, 1409 (2009).

>

Alvaro, M., Corma, A., Das, D., Fomes, V., Garcia, H., Chem. Commun., 956 (2004).

>

Alvaro, M., Corma, A., Das, D., Fornes, V., Garcia, H., J Catal., 231, 48 (2005).

>

Asefa, T., MacLachlan, M.J., Coombs, N., Ozin, G.A., Nature, 402, 867 (1999).

>

Babonneau, F., Leite, L., Fontlupt, S., J. Mater. Chem., 9, 175 (1999).

>

Badley, R. D. and Ford, W.T., J Org. Chem., 54, 5437 (1989).

>

Barrett, E. P., Joyner, L. G., Halenda, P. P., J. Am. Chem. Soc, 73, 373 (1951).

>

Beck, J. S., Vartuli, J. C , Roth, W. J., Leonowicz, M. E., Kresge, C. T., Schmitt, K.
D., Chu, C. T. W., Olson, D. H., Sheppard E. W., McCullen, S. B., Higgins, J. B.,
Schlenker, J. L., J. Am. Chem. Soc, 114, 10834 (1992).

>

Bell A. T. and Pines A., NMR Techniques in catalysis, Marcel Dekker Inc. New
York (1994).

>

Bhawsik, A., Kapoor, M. P., Inagaki, S., Chem. Commun., 470 (2003).

173

>

Blasco, T., Corma, A., Martinez, A., and Martinez-Escolano, P., J. Catal. I l l , 306
(1998).

>

Boronat, M., Viruela, P., Corma, A., J. Phys. Chem. A 102, 982 (1998).

>

Botella, P., Corma, A., Lopez-Nieto, J. M., J. Catal. 185, 371 (1999).

>

Boveri, M., Aguilar-Pliego, J., Perez-Pariente, J., Sastre, E. Catal. Today 868, 107
(2005)

>

Brown, J., Mercier L., Pinnavaia, T. J., Chem. Commun., 69 (1999).

>

Burkett, S. L., Sim, S. D., Mann, S., Chem. Commun., 1367 (1996).

>

Buzzoni, R., Bordiga, S., Ricchiardi, G., Spoto, G., Zecchina, A., J. Phys. Chem. 99,
11937(1995).

>

Cardona F., Gnep N. S., Guisnet M., Szabo G., Nascimento P., Appl. Catal. A 128,
243(1995).

>

Cheetham, A. K., Eddy, M. M., Thomas, J. M., J. Chem. Soc. -Chem. Commun. 1337
(1984).

>

Child, J. E., Chou, T. S., Huss, A. Jr., Kennedy, C. R., Ragonese, F. P., Tabak, S. A.,
U.S. Patent 4,956,518 (1990).

>

Choi, M., Kleitz, F., Liu, D., Lee, H. Y., Ahn, W. S., Ryoo, R., J. Am. Chem. Soc,
127, 1924(2005).

>

Clerici, M. G., Perego, C , de Angelis A., Montanari, L., US. Patent 5,571,762
(1993).

>

Clerici, M. G., de Angelis A., Ingallina, P., Italian Patent 1,272,926 (1995).

>

Clet, G., Goupil, J. M., Szabo, G., Cornet, D., Appl. Catal. A 202, 37 (2000).

>

Chu, Y.F. and Chester, A.W., Zeolites 6 (3), 195 (1986).


174

>

Cooper, M. D., King, D. L., Sanderson, W.A., Patent VO 92/04977 ( 1992).

>

Cooper M. D., King D. L., Sanderson W. A., Patent WO 9402243 (1994).

>

Corma, A.and Martinez, A. Catal. Rev.-Sci. Eeg., 35, 483 (1993).

>

Corma, A., Martinez, A., Martinez, C , Catal. Lett. 28, 187 (1994a).

>

Corma, A., Martinez, A., Martinez, C , J. Catal. 146, 185 (1994b).

>

Corma, A., Martinez, A., Martinez, C , J. Catal. 149, 52 (1994c).

>

Corma, A., Gomez, V., Martinez, A., Appl. Catal. A 119, 83 (1994d).

>

Corma, A., Juan-Rajadell, M. L, Lopez-Nieto, J. M., Martinez, A., Martinez, C , Appl.


Catal. A 111, 175 (1994e).

>

Corma, A., Martinez, A., Martinez, C , Appl. Catal. A 144, 249 (1996).

>

Cumming, K. A. and Wojciechowski, B. W., Catal Rev.-Sci. Eng. 38, 101 (1996).

>

Cundy, C. S. and Cox, P. A., Chem. Rev., 103, 663 (2003).

>

Davis, M. E. and Lobo, R. F., Chem. Mater., 4, 756 (1992).

>

de Angelis, A., Ingallina, P., Berti, D., Montanari, L., Clerici, M. G., Catal. Lett. 61,
45 (1999).

>

de Angelis, A., Flego, C , Ingallina, P., Montanari, L., Clerici, M. G., Carati, C ,
Perego, C , Catal. Today 65, 363 (2001).

>

Degnan, T. F. Jr., J. Catal., 216, 32 (2003).

>

de Jong, K. P., Mesters, C. M. A. M., Peferoen, D. G. R., van Brugge, P. T. M, de


Groot, C , Chem. Eng. Sci. 51, 2053 (1996).

>

Derouane, E.G., He, H., Derouane-Abd Hamid, S.B., Ivanova, I. L, Catal. Lett. 58, 1
(1999).
175

>

Diaz-Mendoza F. A., Pernett-Bilano L., Cardona-Martinez N., Thermochim. Data


312,47(1998).

>

Diaz, I., Marques-Alvarez, C , Mohino, F., Perez-Pariente, J., Sastre, E., J. Catal,
193, 283 (2000a).

>

Diaz, I., Marques-Alvarez, C , Mohino, F., Perez-Pariente, J., Sastre, E., J. Catal,
193,295 (2000b).

>

Dube, D., Royer, S., Trong On, Do, Beland, F, Kaliaguine, S., Micropor. Mesopor.
Mater., 79, 137 (2005).

>

Engelhardt, G. and Michel, D., High-Resolution Solid-State NMR of Silicates and


Zeolites, John Wiley & Sons Ltd., Chichester (1987).

>

Eder, F. and Lercher, J. A., Zeolites 18, 75 (1997).

>

Feller, A., Zuazo, I., Guzman, A., Barth, J.O., Lercher, J.A., J. Catal. 216, 313
(2003a).

>

Feller A., Barth J.O., Guzman A., Zuazo I., Lercher J.A., J. Catal. 220, 192 (2003b).

>

Feller, A.; Guzman, A.; Zuazo, I. and Lercher, J.A. J. Catal., 224, 80-93 (2004).

>

Feller, A. and Lercher, J.A., Adv. Catal., 48, 229 (2004).

>

Flego C , Kiricsi I., Parker Jr. W.O., Clerici M.G., Appl. Catal. A: Gen. 124, 107
(1995).

>

Fowler, C. E., Burkett S. L., Mann, S., Chem. Commun., 1769 (1997).

>

Fowler, C. E., Lebeau, B., Mann, S., Chem. Commun., 1825 ( 1998).

>

Furimsky, E., Catal Today 30,223 ( 1996).

>

Fyfe, C. A.. Solid State NMR for Chemists, C.F.C Press., Ontario ( 1983)

176

>

Garwood, W.E., and Venuto, P.B., J. Catal 11, 175 (1968).

>

Ghanbari-Siahkali, A., Philippou, A., Garforth, A., Cundy, C.S., Anderson, M.W.,
Dwyer, J., J. Chem. Mater. 11, 569 (2001).

>

Gobin, O. C , Wan, Y., Zhao, D. Y., Kleitz, F., Kaliaguine. F., J. Phys. Chem. C, 111,
3053 (2007).

>

Guisnet M., Magnoux P., Catal. Today 36, 477 (1997).

>

Guzman, A., Zuazo, I., Feller, A., Olindo, R., Sievers, C , Lercher, J. A., Micropor.
Mesopor. Mater. 83, 309 (2005).

>

Harmer, M. A., Farmeth, W. E., Sun, Q., J. Am. Chem. Soc, 118, 7708 (1996).

>

Harmer, M. A., Farmeth, W. E., Sun, Q., AdV Mater., 10, 1255 (1998).

>

Harmer, M. A. and Sun, Q., Appl. Catal. A 221, 45 (2001 ).

>

Hamoudi, S. and Kaliaguine, S., Microporous Mesoporous Mater., 59, 195 (2003).

>

Hamoudi, S., Royer, S., Kaliaguine, S., Microporous and Mesoporous Materials, 71 ,
17 (2004).

>

Hartley, W. R., Englande Jr. A. J., Harrington, D. J., Wat. Sci. and Technol. 39, 305
(1999).

>

Hatton, B., Landskron, K., Whitnall, W., Perovic D., Ozin, G. A., Ace. Chem. Res.,
38, 305 (2005).

>

Haw, J. F., Richardson, B. R., Oshiro, I. S., Lazo, N. D., Speed, J. A., J. Am. Chem.
Soc 111,2052(1989).

>

He, J. Y., Nivarthy, G. S., Eder, F., Seshan, K, Lercher, J. A., Microp. Mesop. Mater.
25, 207 (1998).

>

Heidekum, A., Harmer, M. A., Hoelderich, W. F., J. Catal., 176, 260 (1998).
177

>

HPI Construction Boxscore. Hydrocarbon Process. 10, 87 (2008) (Supplement).

>

Hunks, W. J., Ozin, G. A., J. Mater. Chem., 15,3716 (2005).

>

Huo, Q., Margolese D. I., Stucky, G. D., Chem. Mater., 8, 1147(1996).

>

Huss, Jr. A. and Jonson, I. D., U.S. Patent 5,221,777(1993).

>

Hutson, Jr. T. and Logan, R. S., Hydrocarbon Processing Sept. 1975, 107 (1975).

>

Inagaki, S., Guan, S., Fukushima, Y., Ohsuna, T. and Terasaki, O., J. Am. Chem.
Soc, 121,9611(1999).

>

Inagaki, S., Guan, S., Ohsuna, T., Terasaki, O., Nature, 416 (2002).

>

Ingallina, P., de Angelis, A, Parker, Jr. W.O., Clerici, M. G., Catal. Lett. 78, 297
(2002).

>

Ipatieff, V.N. and Grosse,A.V., J. Am. Chem. Soc. 57,1616 (1935).

>

Jacquesy, J.C, Carbocation Chemistry, 359 ( 2004).

>

Janik, M. J., Davis, R. J., Neurock, M., J. Catal. 224, 65 (2006).

>

Josl, R., Klingmann, R., Traa, Y., Glaser, R., Weitkamp, J., Catal. Commun. 5, 239
(2004).

>

Kazansky, V.B., Catal. Today 51, 419 (1999).

>

Kirsch, F W., Potts, J.D., Barmby, D.S., Proceedings - American Petroleum Institute,
Division of Refining 48, 1000 (1968).

>

Kirsch, F. W., Potts, J.D., Barmby, D.S., J. Catal.21 142(1972).

>

Kleitz, F., Czuryszkiewicz, T., Solovyov, L. A., Linden, M., Chem. Mater. 18, 5070
(2006).

178

>

Klingmann, R., Josl, R., Traa, Y., Glaser, R., Weitkamp, J., Appl. Catal. A 281, 215
(2005).

>

Kosslick, H., Lischke, G., Walther, G., Storek, W., Martin, A., Fricke, R., Micropor.
Mater., 9, 13(1997).

>

Kresge, C. T., Leonowicz, M. E., Roth, W. J., Vartuli, J. C , Beck, J. S., Nature, 359,
710(1992).

>

Kruk, M., Jaroniec, M., Sayari, A., Langmuir, 13, 6267 (1997).

>

Kumar, P., Vermeiren, W., Dath, J. P., Hoelderich, W. F., Energy & Fuels 20, 481487 (2006).

>

Ledneczki, I., Daranyi, M., Fulop, F., Molnar, A., Catal. Today, 100, 437 (2005).

>

Lee, L. and Harriott, P., Ind. Eng. Chem. Process Des. Develop. 16, 282 (1977).

>

Lim, M. H., Blanford, C. F., Stein, A., J. Am. Chem. Soc, 119, 4090 (1997).

>

Lim, M. H., Blanford, C. F., Stein, A., Chem. Mater., 10, 467 (1998).

>

Lim, M. H. and Stein, A., Chem. Mater., 11, 3285 (1999).

>

Lindlar, B., Luchinger, M., Haouas, M., Kogelbauer, A., Prins, R., Stud. Surf Sci.
Catal, 135,4828(2001).

>

Loenders, R., Jacobs, P.A., Martens, J.A., J. Catal. 176, 555 (1998).

>

Lukens, Jr. W. W., Schmidt-Winkel, P., Zhao, D., Feng, J., Stucky, G. D., Langmuir,
15,5403(1999).

>

Macquarrie, D. J., Chem. Commun., 1961(1996).

>

Macquarrie, D. J., Tavener, S. J., Harmer, M. A., Chem. Commun., 2363(2005).

>

Marcilly, C , J. Catal., 216, 47 (2003).


179

>

Margolese, D., Melero, J. A., Christiansen, S. C , Chmelka B. F., Stucky, G. D.,


Chem. Mater., 12, 2448(2000).

>

Martinez, F., Morales, G., Martin, A., van Grieken, R., Appl. Catal. A. 347, 169
(2008).

>

McClure J. D., Patent US 4022847 ( 1997).

>

Melde, B. J., Holland, B. T., Blandford, C. F., Stein, A., Chem. Mater., 11, 3302
(1999).

>

Melero, J.A, Stucky, G. D., van Grieken, R., Morales, G., J. Mater. Chem. 12, 1664
(2002).

>

Melero, J.A., Grieken, R.V., Morales, G., Chem. Rev. 106, 3790 (2006).

>

Mercier, L. and Pinnavaia, T. J., Chem. Mater., 12, 188(2000).

>

Miller, K.D. Jr., Alkylates Key components in clean-burning gasoline, presented to


the Clean Air Act Advisory Committee Panel on Oxygenate Use in Gasoline, May 24,
1999. (www.epa.gov/otaq/consumer/fuels/oxypanel/dxmiller.ppt)

>

Molnar, A., Curr. Org. Chem. 12, 159 (2008).

>

Moreno, M., Rosas, A., Alcaraz, J., Hernandez, M., Toppi, S., Costa, P. D., Appl.
Catal. A. 251, 369 (2003).

>

Mostad, H.B., Stocker, M., Karlsson, A., Rorvik, T., Appl. Catal. A. 144, 305 (1996).

>

Mukhergee, M., Nehlsen, J., Hydrocarbon Process. 86, 110 (2007).

>

Muth, O., Schellbach, C , Froba, M., Chem. Commun. 2032 (2001 ).

>

Nivarthy, G., He, Y., Seshan K., Lercher J.A., J. Catal. 176,47 (1998a).

>

Nivarthy, G., Seshan, K., Lercher, J.A., Micropor. Mesopor. Mater. 22, 379 (1998b).

180

>

Nowak, A. K., Mesters, C. M. A. M., Rigby, A. M., Schulze, D., Div. Petr. Chem.,
Am. Chem. Soc. 41, 668 (1996).

>

Olah, G.A., Prakash, G.K.S., Sommer, J., Science, 206, 13 (1979).

>

Olah, G.A., Iyer, P.S., Prakash, G. K. S., Synthesis, 513 (1986).

>

Olivier, H., J. Mol. Catal., 92, 155 ( 1994).

>

Pater J., Cardona F., Canaff C , Gnep N.S., Szabo G., Guisnet M., Ind. Eng. Chem.
Res. 38, 3822 (1999).

>

Petkovic, L.M. and Ginosar, D. M., Appl. Catal. A: Gen. 275, 235 (2004).

>

Petkovic, L., Ginosar D. M., Burch, K. C , J. Catal. 234, 328 (2005).

>

Platon, A. and Thomson, W. J., Appl. Catal. A. 282, 93 (2005).

>

Querini C.A., Appl. Catal. A: Gen. 163, 199 (1997).

>

Querini, CA. Catal. Today 62, 135 (2000).

>

Ramachandran, S., Lenz, T. G., Skiff, W. M., Rappe, A. K., J. Phys. Chem. 100,
5898 (1996).

>

Ravikovitch, P. I. and Neimark, A.V., Langmuir, 18, 1550 (2002a).

>

Ravikovitch, P. I. and Neimark, A.V., Langmuir, 18, 9830 (2002b).

>

Richer R. and Mercier, L., Chem. Commun., 1175(1998).

>

Rigby, A.M., Kramer, G. J., van Santen, R.A., J. Catal. 171,1 (1997).

>

Rocha, J., Carr, S. W., Klinowski, J., Chem. Phys. Letters, 187, 401 (1991).

>

Roeseler, C. M., Shields, D. J., Black, S. M., Gosling, C. D., Improved Solid Catalyst
Alkylation Technology for Clean Fuels: The Alkylene Process. Presented at the
NPRA Annual Meeting, San Antonio, TX, 17, (2002).
181

>

Rorvik, T., Mostad, H., Ellestad, O. H., Stocker, M., Appl. Catal. A 137, 235 (1996).

>

Rorvik, T., Mostad, H.B., Karlsson, A., Ellestad, O.H., Appl. Catal. A. 156, 267
(1997).

>

Ryczkowski, J., Catal. Today, 68, 263 (2001).

>

Sanchez-Castillo, M.A., Agarwal, N., Miller, C , Cortright, R.D., Madon, R.J.,


Dumesic, J.A., J. Catal. 205, 67 (2002).

>

Sato, S., Maciel, G. E., J. Mol. Catal. A. Chem. 101, 153 (1995).

>

Scherzer, J., Catal. Rev.-Sci. Eng. 31,215(1989).

>

Sherwood, T. K., Pigford, N. L., Wilke, C. R., In Mass Transfer, McGraw-Hill, Inc.,
New York, 17,379(1975).

>

Sievers, C , Zuazo, I., Guzman, A., Olindo, R., Syska, H., Lercher, J.A., J. Catal. 246,
315-324(2007).

>

Sievers, C , Liebert, J. S., Stratmann, M. M., Olindo, R., Lercher, J.A., Appl. Catal. A.
190,89-100(2008).

>

Simpson, M. F., Wei, J., Sundaresan, S., Ind. Eng. Chem. Res. 35, 3861 (1996).

>

Smimova, M. Y., Urguntsev, G. A., Ayupov, A. B., Vedyagin, A. A., Echevsky, G.


V., Appl. Catal. A., 344, 107 (2008).

>

Sing, K. S. W., Everett, D. H., Haul, R. A. W., Moscou, L., Pierotti, R. A.,
Rouquerol, J., Siemieniewska, T., Pure Appl. Chem., 57, 603, (1985).

>

Song, X. and Sayari, A. Catal. Rev. -Sci. Eng. 38, 329 (1996).

>

Sreekanth, P., Park, J.K., Kim, J.W., Hyeon, T., Kim, B.M., Catalysis Letters, 96, 3,
(2004).

>

Stell, J., Oil Gas J. 99 (52), 74 (2001)


182

>

Stocker, M., Mostad, H., Rorvik T., Catal. Lett. 28, 203 (1994).

>

Stocker, M., Mostad, H., Karlsson, A., Junggreen, H., Hustad, B., Catal. Lett. 40, 51
(1996).

>

Taguchi, A. and Schuth, F., Micropor. Mesopor. Mater., 11, 1, (2005).

>

Trong-On, D., Zaidi, S. M. J., Kaliaguine, S., Micropor. Mesopor. Mater., 22, 211
(1998).

>

Trong-On, D. and Kaliaguine, S., Angew. Chem. Int. Ed., 41, 1036 (2002).

>

Trong On, D., Desplantier-Giscard, D., Danumah, C , Kaliaguine, S., Appl. Catal. A.
253, 545 (2003).

>

van Bokhoven, J. A., Roest, A. L., Konigsberger, D. C , Miller, J. T., Nachtegaal, G.


H., Kentgens, A. P. M., J. Phys. Chem. B 104, 6743 (2000).

>

Van Rhijn, W. M., De Vos, D. E., Sels, B. F., Bossaert, W. D., Jacobs, P. A. Chem
Commun. 317 (1998a).

>

Van Rhijn, W. M., De Vos, D. E., Bossaert, W. D., Bullen, J., Wouters, B., Grobet,
P., Jacobs, P. A., Stud. Surf. Sci. Catal. 117, 183 (1998b).

>

Walker, G. R. "Alkylation of 1-butene with isobutane using EMT and Y zeolites ",
PhD thesis, University of Waterloo, (2000).

>

Wang, S. and Guin, J.A., Energy Fuels 15, 666 (2001).

>

Wang, P., Wang, D., Xu, C , Gao, J., Appl. Catal. A, 332, 22 (2007).

>

Wei, T. C. and Hillhouse, H. W., Langmuir, 23, 5689 (2007).

>

Weitkamp J. and Maixner S., Zeolites 1, 6 (1987).

>

Weitkamp J. and Traa Y., in "Handbook of Heterogeneous Catalysis " (Ertl G.,
KnzingerH, Weitkamp J., Eds), Vol. 4, p.2039, VCH, Weinheim, (1997).
183

>

Weitkamp, J. and Traa, Y. Catal. Today 49, 193 (1999).

>

Yoo, K., Burckle, E., Smimiotis, P., Catal. Lett. 74, 85 (2001).

>

Yoo, K. and Smimiotis, P.G., Appl Catal. A. 227, 171 (2002).

>

Yoo, K. and Smimiotis, P.G., Appl. Catal. A. 246, 243 (2003).

>

Yu, C , Fan, J., Tian, Stucky, G. D., Zhao, D., J. Am. Chem. Soc. 124, 4556 (2002).

>

Zecchina, A., Geobaldo, F., Spoto, G., Bordiga, S., Ricchiardi, G., Buzzoni, R.,
Petrini, G., J. Phys. Chem. 100, 16584 (1996).

>

Zhao, D., Feng, J., Huo, Q., Melosh, N., Fredrickson, G. H., Chmelka, B. F., Stucky,
G. D., Science, 279, 548 (1998a)

>

Zhao, D., Huo, Q., Feng, J., Chmelka, B. F., Stucky, G. D., J. Am. Chem. Soc, 120,
6024 (1998b).

>

Zhao, Z., Sun, W., Yang, X., Ye, X., Wu, Y., Catal. Lett. 65, 115 (2000).

184

You might also like