144
Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte over Jacobsen’s catalyst immobilized in inorganic porous materials Von der Fakultät für Mathematik, Informatik und Naturwissenschaften der Rheinisch- Westfälischen Technischen Hochschule Aachen zur Erlangung des akademischen Grades eines Doktors der Naturwissenschaften genehmigte Dissertation vorgelegt von M.Sc.-Ing. Jairo Antonio Cubillos Lobo aus El Bagre, Kolumbien Berichter: Universitätsprofessor. Dr. rer. nat. W. F. Hölderich Universitätsprofessor Dr. rer. nat. Carsten Bolm Tag der mündlichen Prüfung: 12. 05. 2005 Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar

Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte over Jacobsen’s catalyst immobilized in

inorganic porous materials Von der Fakultät für Mathematik, Informatik und Naturwissenschaften der Rheinisch-Westfälischen Technischen Hochschule Aachen zur Erlangung des akademischen

Grades eines Doktors der Naturwissenschaften genehmigte Dissertation

vorgelegt von

M.Sc.-Ing.

Jairo Antonio Cubillos Lobo

aus El Bagre, Kolumbien

Berichter: Universitätsprofessor. Dr. rer. nat. W. F. Hölderich Universitätsprofessor Dr. rer. nat. Carsten Bolm

Tag der mündlichen Prüfung: 12. 05. 2005

Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar

Page 2: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Preface

This work has been accomplished from Marz 2002 until April 2005 In the Department of Chemical Technology and Heterogeneous Catalysis of the university of Technology of Aachen (Germany). At first I would like to thank my supervisor, Prof. Dr. Wolfgang F. Hölderich, who gave me the wonderful opportunity to work in his institute, and suggested such an interesting topic for my thesis. My special thank is directed Dr. John Niederer for being an co-advisor of this work and for the useful discussions. I thank Prof. Dr. Carsten Bolm for refereeing the work. This work was supported by the “Graduiertenkolleg 440” program of the university of Technology of Aachen (Germany). I would also like to thank Dr. Hans Schuster, Prof. Dr. Luis Rios and Dr. Patrick Weckes, who helped me to find my self in a new country. For performing numerous GC analyses I want to thank Dr. Heike Hausmann, Marco Gillian, Natalie Mager and Heike Boltz-Fickers. For performing XRD and N2 physisorption measurements my thank goes to Karl Josef Vaeßen. For the thermal analyses measurements I would like to thank Elena Modrogan and Melinda Batorfi. For numerous elemental analyses I want to thank Elke Biener. For numerous spectroscopic analyses by FT-IR and (DR) UV-Vis I thank Elke Biener and Marco Gillian. For the preparation of cis-ethyl cinnamate I want to thank Adrian Crosman. For efficient repairs of defect mechanical, electric and electronic equipment I thank Günter Wirtz. Furthermore I thank all the members of the department for steady assistance and the comfortable working atmosphere and especially to Bernd Müller, Jose-María Menendez-Torre, Marco Gillian and Daniel Egbuniwe. Many Thanks to all my colleagues of the Grupo Catalisis Ambiental of the University of Antioquía (Colombia), who supported me during the writing of this thesis and helped keep me a good mood even in the most difficult moments. Also our scientific discussion were always fruitful. I thank specially to Prof. Dr. Consuelo Montes de

i

Page 3: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Correa and Aída-Luz Villa-Holguín, who spent so many hours for the correction of the text of this thesis. Last but not least, many cordial thank to my dear family. Without their support I would never finish my Ph.D.

ii

Page 4: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Content 1 Scope of thesis 12 Background 52.1 Optically active compounds 5

2.2 Sources of enantiomers 9

2.3 Homogeneous asymmetric epoxidation 14

2.4 Heterogeneous asymmetric epoxidation 23

2.4.1 Zeolites 28

2.4.1.1 Modification of zeolites 33

2.4.1.2 Zeolites as support for asymmetric epoxidation catalysts 41

2.4.2 Mesoporous materials 45

2.4.2.1 MCM-41 as support for asymmetric epoxidation catalysts 48

3 Results and discussion 573.1 Motivation and goal 57

3.2 Preparation and characterization of the catalysts 60

3.2.1 Preparation and caracterization of the support materials 60

3.2.1.1 Highly dealuminated zeolites X and Y 60

3.2.1.2 Al-MCM-41 and Si-MCM-41 63

3.2.2 Characterization of the heterogeneous catalysts 65

3.2.2.1 Heterogeneous catalysts based on modified zeolites X and Y 65

3.2.2.2 Heterogeneous catalysts based on Al-MCM-41 and Si-MCM-41 76

3.3 Catalytic activity 86

3.3.1 NaOCl/4-PPNO as oxidation system 86

3.3.2 m-CPBA/NMO as oxidation system 89

3.3.3 In situ generated DMD as oxidation system 91

4 Conclusions and perspectives 1025 Experimental 1055.1 Chemicals 105

5.2 Catalysts preparation 105

5.2.1 Jacobsen’s catalyst 105

5.2.2 Modification of zeolites Na-X and Na-Y 106

5.2.3 Immobilization of Jacobsen’s catalyst into the generated mesoporous

of the zeolites X and Y

106

iii

Page 5: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

5.2.3.1 According to the method described by Hölderich et al. [213] 106

5.2.3.2 According to the method described by Corma et al. [134] 107

5.2.3.3 According to the method described by Bein et al. [135] 107

5.2.4 Synthesis of Al-MCM-41 108

5.2.5 Synthesis of Si-MCM-41 108

5.2.6 Immobilization of Jacobsen’s catalyst in Al-MCM-41 109

5.2.6.1 Immobilization by ion exchange 109

5.2.7 Immobilization of Jacobsen’s catalyst in Si-MCM-41 109

5.2.7.1 Immobilization by coordination bond between manganese and

nitrogen atom anchored previously on amino propyl-functionalized

Si-MCM-41

109

5.2.7.2 Immobilization by chemical attachment of the salen ligand on

functionalized Si-MCM-41 110

5.2.8 Preparation of cis-ethyl cinnamate 111

5.2.9 Preparation of the chiral epoxides used as references samples 111

5.4 Characterization 113

5.4.1 XRD 113

5.4.2 Sorption of N2 113

5.4.3 Elemental analysis 113

5.4.4 Thermogravimetric analysis 113

5.4.5 UV-Vis 114

5.4.6 FT-IR 114

5.5 Catalytic test reaction 114

5.5.1 NaOCl/4-PPNO as oxidation system 114

5.5.2 m-CPBA/NMO as oxidation system 115

5.5.3 in situ generated DMD as oxidation system 115

5.6 Calculation of catalytic activity and enantioselectivities 116

5.6.1 GC-Analysis 116

5.6.1.1 Partial hydrogenation of ethyl phenylpropiolate 116

5.6.1.2 Asymmetric epoxidation of cis-ethyl cinnamate 116

5.6.1.3 Asymmetric epoxidation of styrene 117

5.6.1.4 Asymmetric epoxidation of 1,2-dihydronaphthalene 117

5.6.2 Calculus of conversions, selectivities and enantiomeric excesses 118

6.0 References 119

iv

Page 6: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

List of figures Figure 1. Schematic representation of one pair of enantiomers 5

Figure 2. The Cahn-Ingold-Prelog convention 8

Figure 3. Stereoisomers isomers of ethyl phenylglycidate. 9

Figure 4. Resolution of the racemates of 1,2-diaminocyclohexane with tartaric

acid 11

Figure 5. Catalytic kinetic resolution 11

Figure 6. Types of asymmetric synthesis 12

Figure 7. Sharples enantioselective epoxidation for primary alcohols 15

Figure 8. Jacobsen-Katsuki homogeneous enantioselective epoxidation catalysts 16

Figure 9. Approximation of the incoming olefin towards the oxometallic species 17

Figure 10. Common stereogenic centers in a Mn(III) salen complex 18

Figure 11. Total synthesis of the Taxol side chain 20

Figure 12. (R,R)-(─)-N,N'-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediamine-

manganese(III) chloride 21

Figure 13. Formation of dimethyldioxyrane from Oxone® and acetone 22

Figure 14. Basic chemical structure of a zeolite [143] 28

Figure 15. Schematic representation of the formation of mesopores 35

Figure 16. Nitrogen adsorption and desorption isotherms on a series of zeolites Y 38

Figure 17. Generation of mesopores inside Faujasite type zeolites [212]. 40

Figure 18. Immobilization of a (salen)Mn(III) complex in zeolite Y by encapsulation 42

Figure 19. Highly diastereoselective epoxidation of α-(–)-pinene over a Co-salen

complex immobilized on partially dealuminated zeolite Y 44

Figure 20. The immobilization of Jacobsen’s catalyst on silica 49

Figure 21. The immobilization of Jacobsen’s catalyst on Si-MCM-41 50

Figure 22. The immobilization of Jacobsen’s catalyst on Si-MCM-41 or silica gel 51

Figure 23. The immobilization of Jacobsen’s catalyst on silica gel 52

Figure 24. The immobilization of Jacobsen’s catalyst on Al-MCM-41 (Frunza’s

method) 53

Figure 25. The immobilization of Jacobsen’s catalyst on Al-MCM-41 (Frunza’s

method) 54

Figure 26. The immobilization of Jacobsen’s catalyst on Si-MCM-41 (Li’s method) 55

v

Page 7: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

List of figures (continuation)

Figure 27. The immobilization of a Cr(III) salen complex on Si-MCM-41 (Zhou’s

methods) 56

Figure 28. XRD diffractograms of the parent material Na-X and the modified zeolite 61

Figure 29. XRD diffractograms of the parent material Na-Y and the modified zeolite 62

Figure 30. Nitrogen sorption isotherms of the parent material Na-X and the

modified zeolite 63

Figure 31. Nitrogen sorption isotherms of the parent material Na-Y and the

modified zeolite Na-Y 63

Figure 32. XRD diffractogram for Al-MCM-41 and Si-MCM-41 64

Figure 33. Immobilization of the Jacobsen’ catalyst on dealuminated zeolite X and

Y by Method 1 65

Figure 34. Immobilization of the Jacobsen’s catalyst on dealuminated zeolite X and

Y by Method 2 66

Figure 35. Immobilization of the Jacobsen’s catalyst on dealuminated zeolite X and

Y by Method 3 67

Figure 36. DTG of the free Jacobsen´s catalyst and the Jacobsen´s catalyst

immobilized on modified zeolites X. 70

Figure 37. DTG of the free Jacobsen´s catalyst and the Jacobsen´s catalyst

immobilized on modified zeolites Y. 71

Figure 38. UV-VIS spectra of the Jacobsen’s salen, Jacobsen’s catalyst, modified

zeolite Y and the Jacobsen´s catalyst on modified zeolites Y. 72

Figure 39. UV-VIS spectra of the Jacobsen’s salen, Jacobsen’s catalyst, modified

zeolite X and the different immobilized materials on modified zeolites X.

73

Figure 40. FT-IR spectra of the Jacobsen’s catalyst and the different immobilized

materials on modified zeolites Y. 74

Figure 41. FT-IR spectra of the Jacobsen’s catalyst and the different immobilized

materials on modified zeolites X. 75

Figure 42. Immobilization by ionic bond between Jacobsen’s cation and

aluminosilicates by Method 4 76

Figure 43. Aminopropyl-functionalized Si-MCM-41 77

Figure 44. Anchoring of the Jacobsen’s catalyst in NH2-Si-MCM-41 by Method 5 78

vi

Page 8: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

List of figures (continuation)

Figure 45. Anchoring of the Jacobsen’s catalyst in NH2-Si-MCM-41 by Method 6 79

Figure 46. DTG of the free Jacobsen´s catalyst and the Jacobsen´s catalyst

immobilized on MCM-41 type mesoporous materials. 82

Figure 47. FT-IR spectra of Jacobsen’s catalyst and the Jacobsen’s catalyst

immobilized by ionic bond. 83

Figure 48. UV-Vis spectra of Jacobsen’s catalyst and the Jacobsen’s catalyst

immobilized on NH2 group modified Si-MCM-41. 85

Figure 49. FT-IR spectra of heterogeneous catalyst (MCM1HC) before and

after of reaction with NaOCl as oxygen source. 89

Figure 50. FT-IR spectra of the heterogeneous catalyst (MCM1HC), before and

after of reaction with m-CPBA as oxygen source 91

Figure 51. FT-IR spectra of the heterogeneous catalyst (MCM1HC), before and

after of reaction with in situ generated DMD as oxygen source 94

Figure 52. FT-IR spectra of the Jacobsen’s catalyst and MCM1HC used once in

the epoxidation of 1,2-dihydronaphthalene with different oxidants 97

Figure 53. Effect of the MCM3HC reuse on the conversion 98

Figure 54. Effect of the MCM3HC reuse on the enantioselectivity 98

vii

Page 9: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

List of tables

Table 1. Combined demand for special chemistry products 1

Table 2. Strong and weak points of homogeneous and heterogeneous catalysts 23

Table 3. Heterogenization strategies for homogeneous catalysts 27

Table 4. Historical events and future application for the zeolites 30

Table 5. N2-sorption data of the materials based on modified zeolites 68

Table 6. Composition of the materials based on modified zeolites. 69

Table 7. Nitrogen sorption data for the heterogenized catalyst on Al-MCM-41 and

Si-MCM-41. 80

Table 8. Composition of the materials based on Al-MCM-41 and Si-MCM-41. 81

Table 9. Results of catalytic activity for the heterogenized catalyst on modified

zeolites, using NaOCl as oxygen donor 87

Table 10. Results of catalytic activity for the heterogenized catalyst on Al-MCM-41,

using NaOCl as oxygen donor 88

Table 11. Results of the catalytic activity for the catalyst heterogenized in different

supports and its homogeneous counterpart, using m-CPBA as oxygen

donor

90

Table 12. Results of the catalytic activity for the catalyst heterogenized in different

supports and its homogeneous counterpart, using in situ generated DMD

as oxygen donor

93

Table 13. Catalytic activity results for the enantioselective epoxidation of three

olefins

95

Table 14. Result of catalytic activity and recyclability of the homogeneous catalyst

and in situ generated DMD as oxidant 99

Table 15. List and nomenclature of the applied catalysts 112

List of schemes

Scheme 1. Heterogeneous catalysts for asymmetric synthesis. 25

viii

Page 10: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Scope of the thesis

1 Scope of the thesis

Optically active compounds are of vital interest for the pharmaceutical, agro-food,

agro-chemical, cosmetic, and perfume industries. A high purity or enantiomeric

excess is a very important factor, as often only one of the two possible enantiomers

shows the desired effects on a particular substrate, whereas the other one is

ineffective at best, or even showes contrary effects [1]. Therefore, the demand for

optically pure compounds has been increasing at a constant annual rate of 9.4 %, as

illustrated in Table 1. It can also be seen that the highest combined demand (this

includes the value of merchant sales as well as captive and contract production) for

chiral compounds with achieve a maximal value of US $ 15.1 billion in 2005 [2].

Table 1. Combined demand for special chemistry products [2]

Compounds Million US $ Growth % annual

1990 2000 2005 1990-2000 2000-2005

Chiral compounds 3938 9640 15115 9.4 9.4 Pharma chemicals 2586 6840 11450 10.2 10.9

Hormones 421 1520 2605 13.7 11.4

Anti-inflammatory 616 1435 2130 8.8 8.2

Cardiovascular 440 1185 1755 10.4 8.2

Central nervous system 460 920 1355 7.2 8.1

Respiratory 211 650 1315 11.9 15.1

Non- pharmaceuticals 1352 2800 3665 7.6 5.5

Others 438 1130 2290 9.9 15.2

Asymmetric synthesis is an appropriate method to obtain these high added value

compounds, as one of the enantiomeric pair can now be obtained in high purities in a

single stage [3]. This has stimulated the design of large scale asymmetric synthesis

processes. For example, the production of L (–)-dopa, a powerful drug for the

treatment of Parkison’s disease, is a well established as a practical industrial process

(approximately 1 t / year, 95% enantiomeric excess) [4]. Other processes have also

been developed at industrial scale, including the asymmetric hydrogenation of double

bond C=C, C=O and C=N. While hydrolysis reactions, followed by oxidation reactions

(mainly dihydroxylation and epoxidation reactions) and C-C coupling reactions have

1

Page 11: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Scope of the thesis

been only developed in either pilot or bench scale. An excellent review on such

industrial processes was published by Blaser and collegues [5].

The most important advances of an industrial asymmetric process are the use of

catalytic amounts of the chiral component (generally expensive), the easy separation

of the catalyst and reaction products and the re-utilization of the catalyst [6]. Under

such consideration also continuous processes, which are more productive and

efficient than batch, should be implemented.

Several mainly chiral organometallic complexes, have been found to be efficient and

successful catalysts for asymmetric syntheses, using cheaper substrates (prochiral

substrates) [7]. However, under reaction conditions these catalysts are soluble

(homogeneous), making it difficult to use conventional separations. The recovery of a

catalyst free pure product is of vital inportance: for example, in pharmaceutical drugs

a metal concentration below 10 p.p.m is required [8]. In addition, the separation of

the catalyst under extreme temperature or pressure conditions, such as distillation

under vacuum, can lead to the decomposition of the catalyst, degradation of the

products, high consumptions of energy and waste volume [9].

The heterogenization of homogeneous catalysts is a very important topic for both the

academic and industrial field. Heterogeneous systems are expected to combine the

excellent catalytic activity of homogeneous catalysts with their re-use [10]. Most of

the developed heterogeneous systems use a solid as support or carrier material.

Organic polymers have also been explored as supports for chiral organometallic

catalysts, but they are often unsuitable, because of their sensitivity under the reaction

conditions [11]. In the particular case of oxidation reactions, the polymeric supports

can easily be degraded both by the oxidizing agent and/or the organic solvent

(swelling) [12].

Inorganic supports such as molecular sieves have become of interest in many

studies in the area of heterogeneous catalysis due to its well defined pore structure,

which permit the immobilization of organometallic complexes within the pore system

using different strategies [13]. The immobilization using covalent bonds is the most

used method, as resulting the catalyst will be very stable to the leaching [14].

2

Page 12: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Scope of the thesis

However, in this method the functionalization of the catalyst and/or support is

necessary [15]. Often this proves to be a difficult method [16]. To circumvent this

drawback, the physical entrapment method in well-defined closed cavities according

to the "ship-in-a-bottle" concept has received great attention [17]. In particular, in

Faujasite type zeolites (X and Y), a cubic structure containing broad cavities (1.3 nm)

interconnected by smaller apertures of 0.74 nm, homogeneous catalysts can be

retained inside the cavities [18]. However, this strategy is limited due to the molecular

size of the organometallic complex and to the dimensions of the zeolite cages.

In our group, a procedure was developed to generate mesopores inside microporous

materials by the removal of framework aluminium in Faujasites type zeolites (X and

Y), in such way that voluminous chiral organometallic catalysts could be immobilized.

These materials were shown to be very active in the heterogeneous stereoselective

epoxidation of α-(–)-pinene and R(+)-limonene [19].

Mesoporous materials (e.g Al-MCM-41) have also been considered for the

immobilization of voluminous chiral catalysts, since these have large pores varying

between 30 and 100 Å [20]. Thus, using this material as support, the diffusion

limitations found for zeolitic supports can be avoided. However, in this case the

immobilization by the ”ship-in-a-bottle” concept cannot be used, as its one-

dimensional mesopore structure is directly connected to the surface of the crystallite,

which facilitates the leaching of the catalyst during the catalytic processes [21].

However, in this case, fixation can be achieved by physical or ionic interactions

between the catalyst and the solid surface. This can obviously result in a weakly

bonded immobilization [22]. Therefore, the retention method using covalent bonds is

applied more commonly. However, using this method the active species may lose its

capacity for the chiral communication with the substrate during the reaction [23].

In our group an immobilization method without the use of chemical bonds was

developed. Thus, several enantioselective catalysts were retained just by sorption

and ionic interaction inside the Al-MCM-41 mesopores [24]. These materials have

been successfully used in the enantioselective hydrogenation of various prochiral

olefins [24].

3

Page 13: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Scope of the thesis

The central goal of this thesis is to study the heterogeneous asymmetric epoxidation

of cis-ethyl cinnamate over Jacobsen’s catalyst immobilized inside the generated

mesopores of X and Y type zeolites using the ”ship-in-a-bottle” approach, as well as

immobilized onto Al-MCM-41 internal surface by ionic bond and Si-MCM-41 by

covalent bond.

(2R,3R)-cis-ethyl phenylglycidate obtained from cis-ethyl cinnamate enantioselective

epoxidation is a very important intermediate for the synthesis of taxol and taxotère,

drugs used for the treatment of breast, lung, and ovarian cancer [25].

This work has been part of the macro-project ”Methods in Asymmetric Synthesis” of

the Graduiertenkolleg 440 of the chemistry department of the RWTH Aachen

University.

4

Page 14: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

2 Background 2.1 Optically active compounds

These are spatial isomers which are not superimposable onto its mirror image [26].

To understand this, it is necessary to know the molecular structure of organic

compounds. These contain the element carbon, which can form four covalent bonds

to other atoms. Typically, these bonds are arranged in the shape of a tetrahedron,

with the carbon atom in the center and each bond pointing to a different corner of the

imaginary tetrahedron. If the four atoms or group of atoms bonded to this carbon are

different, then two distinct ways exist to arrange these four atoms around the central

carbon. The two resulting molecules cannot be rotated in such a way as to

superimpose one over the other (see Figure 1). These molecules are mirror images

of each other as our hands are left and right [26].

Mirror plane

Rotation

Figure 1. Schematic representation of one pair of enantiomers

To predict whether or not a compound is optically active the concept of symmetry is

used. A molecular structure possesses a symmetry plane, if it can be divided in half

in such a way that the two halves are mirror images of each other, or if the mirror

image of the complete molecule is identical (superimposable) to the original molecule

[27].

5

Page 15: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

These molecules are called achiral. Therefore, a molecular structure that possesses

a symmetry plane will not be optically active, as it does not possess stereogenic

centers. A stereogenic center in an organic compound is the structural atom, which

defines an asymmetric plane [27]. The most common stereogenic atom is carbon, but

other atoms such as nitrogen, phosphorous and sulfur, can also be stereogenic

centers [28]. Molecules with at least one asymmetric plane are called enantiomers.

Such enantiomers exist in pairs; a single one is called an optically pure compound

[29].

An achiral molecule can be used to obtain enantiomers via an asymmetric synthesis.

Those are called prochiral molecules [30]. Such molecules for example have double

bonds (C=C, C=O, C=N), where addition reactions can create stereogenic centers

(e.g prochiral olefins). A molecule is also prochiral when the replacement of one of

two equivalent groups produces a stereogenic center (e.g. prochiral diols) [30].

Optically active compounds were first discovered in the middle of the 19th century by

the French chemist Louis Pasteur [31]. He observed that the enantiomers of

potassium tartrate rotated plane-polarized light in opposite directions, but in equal

magnitude [31]. The rotation of the plane-polarized light in opposite directions (optical

activity) is the only property that allows to differentiate an enantiomer from the other

one [32]. Later it has been recognized that optically active compounds are free of its

other enantiomer in large molecules such as proteins and carbohydrates [33].

Many chiral compounds are the active components in drugs, insecticides, herbicides,

flavors, fragrances and food additives [34]. Indeed, considering that large chiral

molecules are the basic components of living organisms, it should not be surprising

to find that enantiomerically pure compounds play a dominant role in life science.

Cinchona alkaloids enantiomers, quinine and quinidine are useful naturals drugs

against malarial and arrhythmic, respectively [35]. In other cases, natural sources are

unknown or insufficient to satisfy the required volume demand. This is the case for

taxol, an effective chiral drug for treating different cancer types (breast, lung and

ovarian). Here, asymmetric synthesis can lead to its commercial production, rather

than its physical extraction from the bark of the western yew, taxus brevifolia, a very

6

Page 16: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

limited natural resource [36]. Therefore, in the last 35 years synthetic chemists have

studied the transfer of optical activity (chiral induction in the case of catalysts) from

natural sources, thus producing optically active synthetic compounds, which are not

readily available in nature.

Enantiomerically pure compounds have also shown interest as specials catalysts to

speed up a reaction, as well as to provide chiral induction in the products [37].

Cinchona alkaloids enantiomers were discovered as active catalysts for the

asymmetric hydrogenation of olefins [38]. In contrast, the catalysts developed by

Knowles, Noyori, Sharpless (winners of the chemistry Nobel prize, 2001) and

Jacobsen and Katsuki were obtained via synthetic routes from natural chiral

compounds.

To evaluate the optical purity of an enantiomeric mixture, the term “enantiomeric

excess” (ee) is used, given by equation (1) [39].

% ee = ( ) ( )( ) ( ) %100×

+−SRSR (1)

In which [R] and [S] are the concentrations of the (R)- and (S)-enantiomer,

respectively. These symbols refer to the distinct configuration of each enantiomer

and their assignments depend on the sign of the optical rotation [40]. However, the

use of the concept of stereogenic centers is preferred, since it is possible to obtain

information about the spatial arrangement of the substituents around the stereogenic

atom [41]. The Cahn-Ingold-Prelog convention was created for the assignation of the

(R) and (S) configuration [42]. Figure 2 illustrates this type of assignation.

7

Page 17: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

C C

(R) (S)

ca

b

d c

a

b

d

Figure 2. The Cahn-Ingold-Prelog convention. For each configuration, the atomic

weight of the substituents linked to the sterogenic carbon follows the order:

Ma>Mb>Mc>Md

Spatial isomers or stereoisomers that are not mirror images of each other are called

diastereoisomers [43]. This definition includes both chiral molecules and achiral cis-

trans geometric isomers. The purity of a diastereomer can be evaluated the same

way as enantiomers. In contrast with enantiomers, the physical and chemical

properties of diastereomers are different as well, and their optical rotations can differ

both in sign and magnitude [44]. Figure 3 illustrates the stereoisomers for ethyl

phenylglycidate. This compound contains two stereogenic centers, which lead to four

spatial isomers according to the 2n rule (n: number of stereogenic carbon atoms) [45].

Among them, there are now two pairs of enantiomers (one pair for each cis-trans

configuration) and four pairs of diastereomers.

8

Page 18: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

C C

O

C

O

OC2H5

H H

C C

O

C

O

OC2H5

H H

C C

O

C

O

OC2H5

H H

C C

O

C

O

OC2H5

H H

123 123

123 123

(2R,3R)-cis ethyl-3-phenylglycidate (2S,3S)-cis ethyl-3-phenylglycidate

(2S,3R)-trans ethyl-3-phenylglycidate (2R,3S)-trans ethyl-3-phenylglycidate

Figure 3. Stereoisomers isomers of ethyl phenylglycidate.

2.2 Sources of enantiomers

There are three methods for the production of enantiomerically pure compounds [46]:

1. The chiral pool approach (enantiomerically pure natural substrates)

2. Separation of enantiomers via classical resolution

3. Asymmetric synthesis

The chirality pool is the rich diversity of natural enantiomers (e.g., carbohydrates,

terpenes, or alkaloids) easily obtained from plant or animal sources by conventional

techniques such as extraction or destillation [47, 48]. If in a given reaction the chiral

component is incorporated in the final product, it is called chiral synthon [49]. In

contrast, if the chiral component does not appear in the chiral product, it is called

chiral auxiliary, which can be used either in catalytic or stoichiometric amounts [49].

The use of the chiral pool, is restricted to the synthesis of chiral compounds with

commercial availability of the starting material [50].

9

Page 19: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Pasteur developed the second method when he separated the enantiomers of

potassium tartrate deposited in barrels filled with maturing wines [51]. Here, the

starting material is an equimolar mixture of the enantiomers, which is called racemate

or racemic mixture [51]. Three categories for the resolution of racemates exist:

preferential crystallization, diastereomer crystallization and catalytic kinetic resolution

(chemical or biological) [52]. The first category is used for racemates which consist of

a mechanical mixture of the two enantiomers. The separation occurs when the

racemic mixture is more soluble in a solvent than either of the enantiomers. Most

racemates form of enantiomeric mixtures that are impossible to separate via

preferential crystallization. In these cases, the racemate resolution via diastereomeric

crystallization and kinetic resolution are used. Diastereomeric crystallization is based

on a reaction between the racemic mixture and a pure enantiomer (resolving agent)

to form a diastereomeric salt mixture, which can be separated by crystallization [53].

The most commonly used resolving agents are based on natural products such as L-

(+)-tartaric acid, D(-)-camphorsulfonic acid and various alkaloid bases [53]. A

pertinent example is the separation of the enantiomers of 1,2-diaminocyclohexane

with l-(+)-tartaric acid and d(-)-tartaric acid (see Figure 4). These compounds are

used in the preparation of Jacobsen’s catalyst in its two configurations ((R,R) and

(S,S)) [54].

Kinetic resolution also uses a chiral component, that may be of chemical (chiral acid

or base or chiral organometal complex) or biological origin (enzyme or

microorganism) [55]. Kinetic resolution can be used when one of the enantiomers of

a racemic mixture is more readily transformed than the other [56]. More specifically,

this was the method used by Pasteur in the fermentation of an aqueous solution of

racemic ammonium tartrate, using a Penicillium glaucum mold as chiral reactant [57].

10

Page 20: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

HO OH

H2N NH2

HO2C

HO OH

CO2H

HO2C CO2H

O2C

H

O2C

H

H3N

HOH3N

HO OHNH3

CO2

NH3

H

OH

CO2

H

H2O / CH3COOH

90 °C 5 °C

1,2 Diaminocyclohexane

l-(+)-tartaric acid

d-(-)-tartaric acid

(S,S)-1,2-Diaminocyclohexanemono-(+)-tartrate salt

(R,R)-1,2-Diaminocyclohexanemono-(-)-tartrate salt

Figure 4. Resolution of the racemates of 1,2-diaminocyclohexane with tartaric acid

Figure 5 shows two more modern examples, the enzymatic kinetic resolution of

racemic N-acetylamino acids and the chemical kinetic resolution of a secondary

allylic alcohol, developed by Chibata et al. [58] and Sharpless et. al. [59],

respectively.

NHAc

H2O

OH

TBHP

DIPT

OH

H

RNH2

SNHAc

H

H O

CH3CHCOOH CH3CHCOOH + CH3CHCOOH L-acylase

(R,S)

*

+Ti(O-iPr)4

(a)

(b)

96 % ee

Figure 5. Catalytic kinetic resolution by (a): biological agent and (b): chemical agent.

11

Page 21: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Although the methods of resolution are responsible for ca. half of the enantio-

enriched drugs, they generally require numerous recycle loops to improve the

enantiomeric purity, and fractional crystallization to eliminate the undesired

enantiomer, which industrially is an expensive process [60].

Asymmetric synthesis involves the creation of a stereogenic center in both an achiral

or a chiral molecule [61]. There are two types of asymmetric synthesis [61]. The most

important method uses an achiral or prochiral molecule as staring material and a

chiral auxiliary, which is economically attractive if it can be used in catalytic

quantities. This method corresponds to an enantioselective synthesis (truly

asymmetric synthesis) [62]. A second type of asymmetric synthesis involves the

preferential formation of a single diastereomer in a chiral molecule [62], the so called

diastereoselective synthesis. This method does not require a chiral catalyst. Thus,

the new stereogenic center is induced by the same substrate and the initial optical

configuration is conserved. In contrast, if a chiral catalyst is used, the optical

configuration may be inverted [62]. Figure 6 illustrates the two types of asymmetric

synthesis applied for the epoxidation of olefins.

HC CH2R + [O] Chiral catalystC CH2

OR

H

R1

H R2+ [O] Achiral catalyst

R2

H R2

CH2

O

(a)

(b)

CH2

Figure 6. Types of asymmetric synthesis; (a) enantioselective synthesis; (b)

diastereoselective synthesis.

Catalytic asymmetric synthesis is generally considered to be most appropriate for

large scale productions, as with using a minor quantity of usually expensive catalyst

12

Page 22: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

large amounts of the desired product can be produced in one single step [63].

Although the easy separation and re-use of the catalyst are the main keys to lead a

commercialized enantioselective process, there are still a number of industrial

processes in which the isolation of the chiral product and catalyst is still difficult

(distillation or crystallization). In some cases the catalyst suffers from degradation,

too [64]. These drawbacks can be overcome by heterogenization of the chiral active

phase on a solid support [65]. Only now one simple filtration or decantation is

needed. Therefore, independent of the reaction type, catalyst and special reaction

conditions, new processes are required using heterogeneous catalysts [66].

There commonly are two types of chiral chemical catalysts used for enantioselective

synthesis: chiral acids or bases and organometallic complexes [67]. All these

catalytic systems have the tendency to be soluble in the reaction medium

(homogeneous catalysts) [68]. Their recovery and recyclability are therefore limited to

such conditions. These catalysts have been studied broadly and many of them give

excellent chemical yields and enantiomeric excesses in various reactions of industrial

interest, such as hydrogenations, hydroformylations, dihydroxilations, epoxidations,

hydrolytic kinetic resolution of epoxydes and many more [69].

At the same time, heterogenization strategies have been developed to improve the

productivity by recycling the catalyst without loss of its initial activity [70]. The first

examples of heterogeneous enantioselective catalysts were chirally modified metallic

catalysts (chiral organic base) on inorganic supports (silica, alumina and carbon),

used for the hydrogenation of β-ketoesters and α,β-diketones [71]. A most recent

industrial application is the heterogenization of the Ru-binap organometallic complex

by grafting on a functionalized organic polymer (cross-linked polystyrene). This

catalytic system was found to be more efficient than chirally modified metallic

catalysts [72].

13

Page 23: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

2.3 Homogeneous asymmetric epoxidation

Optically active epoxides are important starting materials for the synthesis of a wide

variety of biologically and pharmaceutically important compounds, as they are readily

converted to various useful functional groups with regio- and stereoselective routes

[73]. For example, pherromones and self-defensive substances against blast disease

of lepidopteran pests in leukotrienes are well known natural chiral epoxides [74].

The asymmetric hydrogenation of prochiral olefins is the most intensively and

thoroughly investigated and developed process in asymmetric synthesis. Without any

doubt, the first breakthrough in asymmetric epoxidation was the Sharpless

epoxidation for primary allylic alcohols. Thus, 90-99% ee and 70-90% chemical yield

can be achieved [75]. This powerful methodology uses tert-butyl hydroperoxide as

oxygen source and titanium isopropoxide-diethyltartrate (DET) as the chiral

component. Additionally, 3Å or 4Å molecular sieves are introduced for the catalytic

version [76] (see Figure 7). In this reaction, the allylic hydroxyl group plays a

fundamental role, in the coordination of the substrate to the chiral titanium complex.

Thus, a favourable environment for the asymmetric induction can be created by virtue

of the strong coordination of the alcohol to the titanium metal center. The Sharpless

epoxidation has also been applied to the kinetic resolution of chiral secondary allylic

alcohols [78]. However, Sharples’s asymmetric epoxidation resulted to be inefficient

in the case of unfunctionalized olefins [77].

14

Page 24: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Figure 7. Sharples enantioselective epoxidation for primary alcohols

R1

R2

R3

OH

D-(-)-diethyltartrate

L-(+)-diethyltartrate

Ti(OiPr)4, TBHP

CH2Cl2, -20 °C MS 4A

O

O

R1

R2

R3

OH

R1

R2

R3

OH

Yield = 70-90 %ee > 90 %

After 10 years, the research groups of Jacobsen and Katsuki reported the

enantioselective epoxidation of simple conjugated unfunctionalized olefins, catalyzed

by chiral salen-Mn(III) complexes (see Figure 8) with rather good chemical yields and

high enantiomeric excesses [79, 80]. These complexes can easily be synthesized

from a salicylaldehyde derivative, a chiral diamine and a metal ion: some are

commercially available [81]. Since their discovery, various homogeneous epoxidation

studies over these organomanganese complexes were made to find the optimal

reaction conditions [82]. The usual oxygen source is sodium hypochlorite (NaOCl),

which combines a high chemical yield with acceptable environmental conditions [83].

Other typical oxidation agents are iodosylbenzene (PhIO) and m-chloroperbenzoic

acid (m-CPBA) [84]. Some reports have shown the convenient use of hydrogen

peroxide (H2O2) [85], O2/aldehyde [86] and Oxone® (2KHSO5●KHSO4●K2SO4)

derivatives such as dimethyldioxirane (DMD) [87] and tetrabutylammonium

monoperoxosulfate (Bu4NHSO5) [88].

15

Page 25: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

R1

R2

R3

R4

+ [O]Chiral organomanganese complex R1

R2

R3

R4

O

* *

[O]: NaOCl, PhIO, IO3-

Chiral organomanganese complexes:

N

O O

NMn

R R

tButBu

R' R'

(R, R)

C C

R= Ph, -(CH2)4-R'= Me, tBu, OMe, OCMe3, OSi(Pr)3

Cl

Jacobsen's catalysts

N

O O

N

Mn

R R

CC

XPh Ph

H H

Katsuki's catalysts

R= Ph, 3,5-Me2C6H4 , -(CH2)4-R'= Ph, MeX= -OAc, PF6

-

H H

Figure 8. Jacobsen-Katsuki homogeneous enantioselective epoxidation catalysts for

conjugated unfunctionalized olefins.

16

Page 26: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

The Jacobsen-Katsuki asymmetric epoxidation represents the asymmetric version of

the achiral olefin epoxidation by salen-manganese(III) complexes, which was

extensively studied by Kochi and coworkers [89]. These systems are based on the

structural similarity with the metal(III) porphyrin complex (metalloporphyrins), which

are the basic units of cytochrome P-450, an oxidizing agent naturally occuring in

widespread use [90]. Therefore, many results of the catalytic behaviour found with

these systems can be applied to metallic salen complexes. For example, trans-

disubstituted olefins were found to be less suitable substrates than cis-disubstituted

due to the unfavourable steric interaction between one of the alkene substituents and

the organometallic complex plane, which was explained from the side-on approach

model (see Figure 9) [91].

O

M

O

M

(a) (b)

Figure 9. Approximation of the incoming olefin towards the oxometallic species. (a)

cis-olefins and (b) trans-olefins

Nitrogen containing cyclic compounds such as pyridine derivatives were found to

accelerate the reaction rate [92]. Although its action mode continues to be a matter of

debate, it is now generally accepted that the catalytic deactivation by dimerization is

suppressed [93]. In addition, asymmetric induction in this complex type is improved

by the reactivity decrease of the active oxidation species, making low reaction

temperatures favourable [94]. In comparison to metalloporphyrins, the metal-salen

complexes have the advantage of possessing potentially stereogenic centers much

closer to the metallic center; in fact, chiral manganese(III) salen complexes

developed by Jacobsen and Katsuki were based on this concept [95]. In contrast to

the Sharpless epoxidation, the active species involved in the Jacobsen-Katsuki

17

Page 27: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

epoxidation consist of oxometallic(V) moieties, and the asymmetric induction is

governed by nonbonding interactions [96]. It has been suggested that both salen-

Mn(III) and salen-oxo-Mn(V) complexes adopt a non-planar conformation during the

oxygen transfer [97]. Optically active salen manganese(III) complexes have also

been recognized as excellent catalysts for the oxidation of sulfides and the

hydroxylation of a C-H bond [98].

In Figure 8 shows the catalysts developed by Jacobsen and Katsuki, respectively.

Generally, Jacobsen’s catalysts are more active than Katsuki’s catalysts, although

the latter gives the best results for trans-disubstituted olefins [99]. Possibly, the

asymmetric centers at 8* and 8’* in the salen ligand are responsible for the chiral

communication with trans-olefins, whereas one at 1 and 1’ in the amine bridge favour

the chiral induction for cis-olefins (see Figure 10) [100]. In addition, the nature of the

substituents located on the salen ligand play a key role in the appropriate approach

of the incoming olefin towards the stereogenic center [101]. This can be explained by

the steric and electronic features of these substituents, which govern the asymmetric

induction of the catalyst [101]. For example, if they have donor properties, they

stabilize the reactive oxo-manganese intermediate, which lowers the reaction rate of

the oxo-transfer step [101]. This increases the enantiomeric selectivity, as then the

interaction between the olefin and the chiral salen ligand is more intense. On the

other hand, Cr(III) and Ru(III) salen complexes have shown the best results for trans-

olefins; however, in most cases stoichiometric amounts are required [102].

Ph

O

N

H

H

Ph

C

Mn

C

Cl

Ph

H

N

O

H

Ph

1 1'1*

2*

3*4*

6*7*

5*

1'*

2'*

3'*4'*

5'*

6'*7'*

8*8'*

Figure 10. Common stereogenic centers in a Mn(III) salen complex for the

enantioselective epoxidation of olefins.

18

Page 28: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Among the numerous substrates, which have been epoxidized using manganese

salen catalysts, mostly unfunctionalized olefins, a series of cis-cinnamate esters

(poor-electron olefins) were also successfully epoxidized using Jacobsen’s catalyst,

sodium hypochlorite (NaOCl) as oxygen source, 4-phenylpyridine N-oxide (4-PPNO)

as an additive and dichloromethane (CH2Cl2) as solvent [103]. The epoxidation of cis-

ethyl cinnamate afforded 96 % chemical yield and 93 % ee for the cis-epoxide

((2R,3R)-3-phenylglycidate) [103].

This (2R,3R)-3-phenylglycidate has become of great interest, as it is the starting

material for the preparation of (2R, 3S)-3-phenylisoserine, the active component of

the C-13 side chain of the taxol family, which are known to have antitumour activity

[104] A synthetic four step procedure was found by Jacobsen and coworkers (see

Figure 11) [104], starting with a partial hydrogenation of the commercially available

ethyl phenylpropiolate (1), using the Lindlar’s catalyst to give cis-ethyl cinnamate (2).

In the second step the asymmetric epoxidation of (2) is performed over Jacobsen’s

catalyst, NaOCl/4-PPNO as oxidation system and CH2Cl2 as solvent. Subsequently,

the optically active epoxide (3) is treated directly with ammonia in ethanol to generate

3-phenylisoserinamide (4) in a highly regioselective ring opening. In the fourth step,

the hydrolysis of the amide was carried out using aqueous barium hydroxide

(Ba(OH)2) and sulphuric acid (H2SO4), thus obtaining (2R,3S)-3-phenylisoserine (5).

This material is converted to the Taxol side chain (6) by treatment with a mixture of

benzoyl chloride and sodium hydrocarbonate (NaHCO3), followed by hydrochloric

acid [104].

19

Page 29: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Ph

Ph

C

C

C

H2

C

O

COC2H5

O

COC2H5

NaOCl4-PPNO

COH

O

COH

O

OH

H

CH

OH

CH

NH

CH

NH2

CH

C

O

Ph

Ph

Ph

HPh

Ph

H

CH

C

NH2

O

C

CH

OH

H

NH3/EtOH

O

O

COC2H5

COC2H5

+

CH2Cl2, 4 °C, 12 h.

Lindlar's catalyst (Pd/C)

(1)

(2) (3)

100 °C

(4)

1. Ba(OH)22. H2SO4

(5)

1. PhCOCl/NaHCO32. HCl

(6)

Asymmetric step

Jacobsen's catalyst

Figure 11. Total synthesis of the Taxol side chain according to Jacobsen et al. [104]

20

Page 30: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

In this synthesis the asymmetric transformation is the most relevant step, during

which the optically active epoxide is obtained. Here, the tert-butyl groups located on

the salen ligand of the Jacobsen’s catalyst (see Figure 12) are crucial for the chiral

communication, as they conduct the olefin on the top of the amine close to both the

stereogenic center and oxo-Mn(V) active species according to the side-on approach

model [105]. Additionally, Katsuki proposed a deeply stepped conformation of the

oxo-Mn(V) complex, which is formed during the olefin approach [106].

N

O

Mn

H C C

Cl

N

O

H

Figure12. (R,R)-(−)-N,N'-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediamine-

manganese(III) chloride (Jacobsen’s catalyst).

Chiral dioxiranes obtained from chiral ketones and Oxone® (2 KHSO5 ● 1 KHSO4 ● 1

K2SO4) have also been demonstrated as suitable catalysts for the enantioselective

epoxidation of cinnamate esters [107] and cinnamic acid derivatives [108]. These

catalytic systems do not have any metal, and the oxygen transfer occurs via the

formation of peroxocarbocylic species [109]. Whit these catalysts, trans-olefins are

more easily converted than cis-olefins [109]. However, they are less active than

Jacobsen’s catalyst. In addition, chiral dioxyranes are easily destroyed under reaction

conditions due to the Baeyer-Villiger oxidation reaction, which is favoured at high pH

(pH > 10.5) [110]. On the other hand, atomic oxygen sources obtained from Oxone®,

such as dimethyldioxirane (DMD) and tetrabutyl ammonium monoperoxosulfate

(Bu4NHSO5) have shown to be excellent oxidants in consideration with Jacobsen’s

catalyst for the epoxidation of several unfunctionalized and functionalized olefins [87].

21

Page 31: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

DMD obtained from acetone and Oxone® has shown to be a good oxidant for the

asymmetric epoxidation of chromenes and isoflavones [85], whereas as the use of

Bu4NHSO5 has only been investigated for unfunctionalized olefins [85]. For DMD a

mechanism has been proposed in which the formation occurs in a buffered medium

at pH 7-7.5, as depicted in Figure 13 [112].

H3C

H3C

H3C

H3C

H3C

C

C

C

O

OH

O

O -

O

O

H

SO3K

O

O

OH - / H2O

SO3 K

H3C O SO3K

H3C

H3C

H2O

C

H3C

H3C

H3C

H3C

O

O -

DMD

C

C

O

O

O

O

OH

SO3K

O

KHSO4

SO3K

+

+

Oxone as KHSO5Acetone

pH = 7.0-8.0

Figure 13. Formation of dimethyldioxirane from Oxone® (peroxysulfonic acid, KHSO5)

and acetone.

According to our knowledge, the Jacobsen epoxidation method is the most powerful

tool for the homogeneous asymmetric epoxidation of cinnamate esters. Other, similar

catalytic systems have only been tested for the epoxidation of allylic and cyclic

unfunctionalized olefins. On the other hand, environmentally attractive oxygen

sources such as oxygen (O2) and hydrogen peroxide (H2O2), have only been used for

the epoxidation of unfunctionalized olefins [113]. Some studies with these oxidants

require the use of an additional component, which is necessary to activate the

oxidizing agent. However, this may affect the environmental viability. For example,

asymmetric epoxidations with oxygen generally requires the use of stoichiometric

amounts of an aldehyde [114].

22

Page 32: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

2.4 Heterogeneous asymmetric epoxidation

To date, many attempts are devoted to the preparation of heterogeneous chiral

catalysts [115]. Table 2 illustrates some of the strong and weak points both of

homogeneous and heterogeneous catalysts [66]. Such homogeneous catalysts have

a well defined molecular structure, which helps to explain their catalytic behaviour.

However, these catalysts often require special conditions for their handling, and

generally low productivities are obtained due to the irreversible change of the active

species during the catalytic cycle [66]. For the homogeneous epoxidation of cis-ethyl

cinnamate over Jacobsen catalyst, only low Turn Over Numbers (TON) between 20-

45 are obtained [103]. For heterogeneous catalysts the easy separation allows their

reutilization, therefore better productivities can be expected. It is generally accepted

that deactivation routes (e.g. the dimerization of Jacobsen's catalyst) can be

suppressed with the isolation of the homogeneous catalyst on a support material, as

in this way the interaction between the active species is minimized [116]. Although

heterogeneous systems are most complex on a molecular level, there are very

advanced characterization spectroscopic methods that can be used to obtain

information about the interaction of the active species and the support [117]. Another

important aspect in the use of heterogeneous catalysts is that reactants and products

are in a phase different to that of the catalyst. This causes diffusion limitations of

reactants and products within the support material [118].

Table 2. Strong and weak points of homogeneous and heterogeneous catalysts [66]

Homogeneous Heterogeneous

Strong points Molecular structure known on

Molecular level.

Easy separation and recovery

of the catalyst from reaction

products.

Better stability and handling.

Weak points High sensitivity.

From medium to low productivities.

Difficult understanding on

molecular level.

Diffusion limitation to and within

the catalyst.

23

Page 33: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Both biological as well as chemical catalysts are being heterogenized. However,

there are more examples for chemical catalysts than biological ones due to the better

productivities, reproducibility and stability obtained with the first one [5].

Scheme 1 shows an overviews of the different kinds of heterogeneous catalysts for

enantioselective catalysis. The first category includes chirally modified

heterogeneous catalysts, both metallic and metal oxides. Among them, metallic

catalysts modified with low or high molecular weight chiral agents are gaining

industrial importance for hydrogenation reactions [119], as for example, the

hydrogenation of α-ketoacetals over cinchona alkaloids modified with platinum

catalyst [120]. With regard to asymmetric epoxidation, the most representative

catalytic system was a chiral tantalum tartrate complex grafted on silica gel [121].

This heterogeneous system was also useful for the enantioselective epoxidation of

primary allylic alcohols [121]. Ultimately, the development of organometallic

complexes as highly active homogeneous catalysts has stimulated their

heterogenization [122]. Immobilized homogeneous catalysts are the second group in

this classification. Various strategies have been developed for the heterogenization

of homogeneous catalysts, most of them including the use of a solid as support or

carrier material [123]. The supports can for example be an inorganic material or

organic polymers. Heterogenization of a Ru-binap complex and p- toluolsulfonic acid

on an organic matrix (polydimethylsiloxane membrane) by chemical bonding

constitutes a remarkable example in this area. This heterogeneous catalyst has been

applied to the asymmetric hydrogenation of C=O double bonds [124]. Inorganic

carriers have the advantage that they usually are resistant to the reaction conditions

(no swelling) and they are mechanically stable [125]. In addition, they may be

functionalized to introduce reactive sites that serve for the fixation of the catalyst,

either by covalent or coordination bonding [125].

24

Page 34: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Heterogeneous catalysts for asymmetric synthesis

Chirally modified heterogeneous catalysts Immobilized homogeneous catalysts

Metallic catalysts Metal oxide catalysts modified withlow molecular weight chiral auxiliary.

Modified with high molecularweight chiral support

Containing low molecularweight chiral auxiliary

Catalyst: PdSupport: Silk fibroin Application:Hydrogenation of anOxazolinone derivativeResults: ee < 66 % [119 ]

No practical examples for epoxidation

Catalyst: Pt/Al2O3Modifier: Cinchona alkaloidApplication:Hydrogenation of α-ketoacetalsResults: ee ~ 50-97% [120 ]

No practical examples for epoxidation

Catalyst: Ta-complex/silica gelModifier: Tartaric acidApplication:Epoxidation of primary allylicalcoholsResults: ee = 98% [121 ]

Catalyst: Ru-binapSupport: Cross-linked

polystyreneMethod: covalent bindingApplication: Hydrogenation of C=OResult: ee = 97 [124]

No practical examples for epoxidation

Scheme 1. Heterogeneous catalysts for asymmetric synthesis.

Table 3 summarizes the most important strategies for the immobilization of

homogeneous catalysts when both organic and inorganic polymers are used as

carrier support [66]. For each one of the heterogenization methods, the

homogeneous catalyst can be introduced directly or can be prepared in situ. The

immobilization by covalent bonding is the most frequently used strategy as very

stable heterogeneous systems can be obtained [126]. This can be performed in

organic polymers either by copolymerization of functionalized ligands with a suitable

monomer, or by grafting on a preformed polymer [127].

In the case of inorganic supports, it is necessary to functionalize the carrier surface,

and sometimes a functionalization of the homogeneous catalyst is also necessary

[65]. Using this method excellent catalytic systems for enantioselective

hydrogenations were developed. They consist of chiral Rh and Ir-diphosphine

complexes which were immobilized on solid inorganic supports or in organic

polymers [128]. In addition, organic polymer and silica gel supported Co-salen

complexes exhibited excellent performance in the hydrolytic kinetic resolution of

25

Page 35: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

terminal epoxides [129]. On the other hand, heterogenization processes without

chemical bonding have also received great attention, since under these conditions

the catalyst can have a catalytic behaviour very similar to the free unimmobilized

system. Therefore, the stereoselectivity of the products may be controlled [130].

For the latter case, the most popular immobilization method is the physical

entrapment of the homogeneous catalyst by the “ship-in-a-bottle” concept [131].

Here, zeolites with well defined pores and cavities are the most suitable materials,

since the imposed restriction due to their pore size prevents the escape of the

homogeneous catalyst, resulting in their occlusion in the zeolite’s cavities [132]. In

this strategy, the metal complex is introduced directly, or is prepared step by step

inside the well-defined cages of the zeolitic support [133]. Using this idea, Mn

epoxidation catalysts with different salen ligands were assembled inside zeolite Y

and EMT [134,135]. However, the entrapped materials are much less active than

their homogeneous counterparts. On the other hand, when the catalyst is difficult to

make, but is very stable, the support may be build up around the catalyst. A

promising example was reported by Jacobs et al. [136], who entrapped Jacobsen’s

Mn-salen epoxidation catalyst and Noyori’s Ru-binap hydrogenation catalyst in a

polydimethylsiloxane film. Loss of the homogeneous catalyst by leaching was

strongly influenced by the size and the solubility of the metal complex and the

swelling of the polymers [136]. Another strategy without chemical bonds is the

heterogenization by adsorption properties such as ion pair formation and electrostatic

interactions [137]. Here, a specific adsorptive interaction rather than the size of the

metal complex is important for the catalyst fixation in the support [137]. In this

method, mesoporous materials have widely been used. This choice is based on their

large pore size (15 to 100 Å), which offers the possibility to retain voluminous

organometallic catalyst [138].

26

Page 36: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Table 3. Heterogenization strategies for homogeneous catalysts [66]

Immobilization

methods

Covalent binding

Adsorption and ion pair formation

Entrapment or

“Ship in a bottle”

approach

Applicability

Problems

Broad

Preparation

Restricted

Competition with solvents and

substrates

Restricted

Size of substrate

and diffusion

limitation

M

L

L

LL

M

L LM

ML L3

Z ZZ

Other strategies for the heterogenization without a carrier material have also

emerged. Two approaches show some promise. The first is the use of a

functionalized soluble ligand, separable by extraction with a suitable organic solvent,

ultrafiltration, or by precipitation by varying of the reaction temperature or the pH

[139]. The second one uses functionalized catalysts to run the reactions in two non-

miscible liquid phases [140]. However, the separation of the catalyst is not so trivial

and the technical feasibility of these methods not been demonstrated yet [66].

Recently, Kureshy and coworkers [141] have reported an alternative strategy based

on the change of solubility of the catalyst by increasing its molecular weight. These

voluminous catalysts were recovered easily by precipitation with hexane and they

could be reused several times without loss of the catalytic activity [141].

27

Page 37: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

2.4.1 Zeolites

Zeolites are crystalline microporous minerals which are broadly distributed in nature,

which have been known for almost 250 years. It was the Swedish mineralogist Axel

Fredrick Cronsted who in 1756 had the honour to discover the so called stilbite. He

observed that a large amount of steam was obtained upon heating [144]. Therefore,

this material was named “zeolite”, which stems from classical Greek, where ζεω (zeo)

means “to boil” and λįθος (lithos) means “stone”. Zeolite structures consist of silicon

(Si+4) and aluminium (Al+3) cations, which are tetrahedrally coordinated by four

oxygen anions (O-2), thus forming a macromolecular three-dimensional framework in

such a way that uniform voids and channels are created in the crystals, with pore

sizes ranging between 4-12 Å [142]. The aluminium in this polymeric structure

generates a negative charge, which will be located on one of the oxygen anions

connected to each aluminium cation, according to Figure 14 [143].

SiO

Al

H

Figure 14. Basic chemical structure of a zeolite [143]

Commonly, the negative charge is compensated by additional non-framework cations

like sodium (Na+).

Table 4 summarizes the most important discoveries concerning these materials

chronologically. The first relevant application was for the physical separations of

liquid or gas mixtures; therefore, they were included in the family of molecular sieves,

where their well-defined pores system has been recognized to allow separations at

molecular levels with high selectivities [146,147]. It motivated the first investigations

about the structural characterization of zeolites, mainly by X-ray diffraction (XRD)

[148, 149]. At the begining, zeolites were applied as adsorbents in the purification of

28

Page 38: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

gas flows to remove water and volatile organic species, and in the separation of

different isomers and gas-mixtures [154,156]. Nowdays they are also applied in the

cleaning up of radioactive waste [161].

However, the major use of zeolites is as ion exchanger, due to its interchanging

cation (see Figure 14). For example, they remove calcium and magnesium dissolved

in water in laundry detergents [159].

Furthermore, the enormous interest in these materials is due to their wide catalytic

applications within industrial areas such as oil refining, petrochemistry, and the

synthesis of special chemicals [157,158]. Nowadays, zeolites are broadly available

on a large scale and used in a variety of applications [162].

In general the major benefit of the zeolites as catalysts is that they contain a unique

microporous structure [163]. The shape and size of the particular pore system exerts

a steric influence on the reaction, controlling the adsorption of reactants and

desorption of products [164]. It is for that reason that zeolite’s micropores can induce

various kind of shape selectivity (reactant, transition state, and product shape

selectivity) [164]. Besides the highly favourable role in providing shape selectivity, the

presence of micropores in some cases limits the catalytic performance of the zeolites

[165]. For example, titanium silicalate (TS-1) can bring about selective

heterogeneous epoxidation with aqueous hydrogen peroxide (H2O2), but its catalytic

activity is limited by the small pore diameter (c.a 5.5 Å); therefore, the oxidation of

olefins with molecular sizes larger than 5.5 Å has been ineffective [166].

29

Page 39: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Table 4. Historical events and future application for the zeolites

Year Event

1756 Cronsted discovered intumescence of stilbite in blowpipe flame:

zeolite [144]

1862 Claire Deville synthesized levynite [145].

1896 Friedel developed open-framework idea: occlusion various liquids

[146].

1925 Weigel and Steinhoff discovered the molecular sieve effect (they

observed that dehydrated chabazite adsorbed water, methanol,

ethanol, but excluded acetone, ether and benzene [147].

1927 Leonard used XRD for structural identification [148].

1930 Taylor and Pauling determined the first structures [149].

1932 McBain introduced the concept of “Molecular Sieves” [150].

1945 Barrer presented the first classification based on molecular size

considerations [151].

1948 Barrer synthesized synthetic analogue of mineral mordenite [152].

1949-1954 Milton and Breck discovered zeolites A, X, and Y [153].

1954 Union Carbide commercialized zeolites as separation and/or

purification materials [154].

1955 Breck and Reed reported on the structure of zeolite A [155].

1959 Union Carbide developed n/iso-paraffin separation and

isomerization catalyst based on zeolite Y [156].

1962 Mobil Oil used zeolite X as cracking catalyst [157].

1969 Grace described first modification: steaming of zeolite Y to ultra-

stable zeolite Y (US-Y) [158].

1974 Henkel used zeolite A in detergents [159].

1982 Union Carbide described aluminophosphates (AlPO4) [160].

In our days Zeolitic materials are mainly being explored as support materials

for highly active homogeneous catalysts [162].

30

Page 40: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Zeolites are generally obtained by hydrothermal synthesis, which consist of water as

the solvent, a silicon source, an aluminium source, a mineralizing agent and a

structure-directing agent [167]. The latter component plays a very important role in

the control of the resulting structures, and in the prediction of a specific structure

[168]. With regards to this, Lobo et al. [169], have stated some general rules between

the structure-directing agent and the structure of zeolites with a high silica content:

1. In the absence of structure-directing agent dense crystalline and layered materials

are obtained.

2. One-dimensional structures with 10 or 12-ring channels are obtained from linear

structure-directing agents, for example in the case of mordenite (MOR).

3. Multi-dimensional zeolites with pore diameters of 4-7 Å are formed when branched

structure-directing agents are used. Two typical examples are ferrierite (FER) and

zeolites X and Y (FAU). Ferrierite is a two-dimensional zeolite with 10-ring main

channels, which are interconnected via smaller 8-ring side channels. Zeolites X

and Y are three-dimensional structures with large cavities that are interconnected

by 12-ring channels, which means that there are 12 cations (Si+4 and Al+3) and 12

O-2 anions present in the ring.

In this thesis the main focus will be on zeolites X and Y, which will be used as starting

material in order to generate mesopores by removal of framework aluminium. These

modified zeolites will be used as support for the immobilization of Jacobsen’s

catalyst.

Although very stable zeolites can be produced, they have one serious limitation; they

are not able to efficiently process molecules that are larger than their pore diameters

(maximum 12 Å) [170]. Consequently, these have been a long search for synthesis

methods that will increase the pore size, and at the same time retain the crystalline

framework. Many efforts have mainly been made utilising bulky structure-directing

agents, however, for zeolites this strategy has not been very successful until quite

recently. This strategy was also found to be applicable to systems containing

aluminium and phosphorous in the framework, the so-called AlPO4 structures [160].

These materials have pore sizes in the range of 13-15 Å, and examples are cloverite,

VPI-5 and AlPO4-8 [171]. Soon after, related materials such as

31

Page 41: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

silicoaluminophosphates (SAPOs) and metal aluminophosphates (MeAPOs) were

synthesized as well [172]. Common for many of these materials is their lack of

thermal or hydrothermal stability [173].

The most important advance in the preparation of large pore crystalline materials is

probably the synthesis of the high silicate zeolite UTD-1 by Balkus et al [174]. UTD-1

has a one-dimensional pore system consisting of elliptical formed pores with a 7.5х10

Å diameter, and it is extremely stable compared to VPI-5, AlPO4-8 and cloverite

[174]. A major drawback, however, is the structure-directing agent, which is a huge

organocobalt complex [174]. This must be removed, and the remaining Co in the

pores has to be washed out by acids [174]. This will probably restrict the practical

applications of this zeolite. All in all, it can be concluded that large pore crystalline

materials are not yet very applicable in catalytic processes, even though great

progress has been made.

Other aspects associated with the pore size is the diffusional behaviour of reactants

and products in zeolites [175,176]. It is clear that the micropore diameters of zeolites

and the sizes of the molecules for which are largely applied as catalysts have similar

dimensions [177]. The migration of molecular species through zeolites therefore

occurs in close contact with the micropore walls [178]. The driving force for difussion

is a concentration gradient, which depends on the size and nature of the sorbate

molecules, the pore-structure and the applied temperature [178].

The occurrence of diffusion limitation can be used to the benefit of the catalytic

process, to enhance the selectivity of the reaction. This was demonstrated in the

toluene disproportionation and xylene isomerization reaction over ZSM-5, where the

diffusion of the desired product p-xylene is much faster than that of unwanted

isomers, which can only leave the zeolites micropores easily if they are converted to

p-xylene [179]. Consequently, the selectivity to p-xylene improves with increasing

zeolite crystal size and is substantially higher than one would expect from

thethermodynamic equilibrium [179]. However, in most cases the effective low

diffusivity in a zeolite crystal limits the reaction rate and yield rather high values for

the diffusional time constant [180].

32

Page 42: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

2.4.1.1 Modification of zeolites

For catalytic applications, sodium ions are mostly replaced by protons (H+) to form a

bond with the negatively charged oxygen anions of the zeolite [143] (see Figure 14).

This results in Brønsted acid sites, which have proven to be highly active in acid-

catalyzed reactions such as catalytic cracking, synthesis of fuel and isomerization

reactions, whereas K+, Co+2, Ba+2 and Sr+2 exchanged zeolites act as basic catalysts

[181]. The ion exchange depends on the nature, concentration and affinity of the new

cation [181].

In addition, it is also possible to replace framework atoms (Si and Al) by other ones

such as Si, B, Ga, Fe, Ti, V. In this way, the acid properties can be changed, on

active sites for oxidation and/or reduction reactions are generated [182].

Isomorphous substitution of Si by Ti or V in silicalite type zeolites led to the discovery

of powerful oxidation catalysts such as titanium silicalite-1 (TS-1) and vanadium

silicalite-1 (VS-1) [183].

In many cases the low effective diffusivity inside a zeolite crystal limits the reaction

rate and yields rather high values for the diffusional time constant [180]. An excellent

way to overcome diffusion limitations in zeolites micropores is the generation of

mesopores (size in the range 20-500 Å) [184]. In this way, the micropores of the

zeolite are effectively shortened and the accessibility of the crystal is largely

enhanced, because the diffusion in the mesopores is several orders of magnitude

higher than in the micropores [178].

Moreover, the creation of mesopores in a zeolite crystal increases the number of

pore mouths that are exposed to the reactant [185]. The effect of the presence of

mesopores for catalysis is demonstrated for several industrially applied

processesthat make use of zeolite catalysts: the cracking of heavy oil fractions and

synthesis of fine chemicals over zeolite Y, and the production of cumene and

hydroisomerization of alkanes over mordenite [186]. For these processes, the

mesopores ensure an optimal accessibility and transport of reactants and products,

while the zeolites micropores induce the preferred shape selective properties [187].

33

Page 43: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

At the present, the generation of mesopores in faujasites type zeolites (X and Y)

plays a key-role in the immobilization of highly efficient enantioselective catalysts for

the epoxidation of prochiral olefins, using the “ship-in-a-bottle” approach.

The different routes presented in literature for the generation and characterization of

mesopores in zeolites crystals will be reviewed. Mesopores can be created via

several routes, including steaming and acid leaching as the most frequently applied

procedures [188]. The isomorphic substitution of framework Al by Si, using chemical

agents such as silicon tetrachloride (SiCl4), EDTA or (NH4)2SiF6 has also been

reported to obtain a dealuminated zeolite with the formation of few mesopores [189].

However, with this method it is known that the zeolite structure collapses if the rate of

extraction of aluminium is much faster than the migration of silicon into the framework

[190].

The creation of mesopores in zeolites by hydrothermal treatment is possible in the

presence of steam [191]. Thermal treatment without steam can create defects in the

zeolite structure [191]. In addition, the use of steam greatly enhances the mobility of

aluminium and silicon species [192]. Almost all the work on steaming of zeolites has

been performed on zeolite Y, although examples of steaming of other zeolites are

known [193]. Steaming is usually performed at temperatures above 500 °C, with the

zeolite is in the hydrogen form. During the contact with steam, hydrolysis of Al-O-Si

bonds takes place. The aluminium is finally expelled from the framework causing a

vacancy (hydroxyl nest) or partial amorphization of the framework. The amorphous

material is a source of mobile silicon species, which can heal the vacancies in the

framework left by the expelled aluminium atoms [194]. Thus, parts of the vacancies

are filled, while others grow to form mesopores, as depicted in Figure 15. In regions

of high defect concentrations spherical mesopores are formed during the steaming,

which are filled with debris from the partial amorphization of the framework and the

extraction of aluminium from the framework [194].

34

Page 44: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Figure 15. Schematic representation of the formation of mesopores (adapted from

Marcilly [194]). (a). Al extraction, (b). Si migration, (c). zeolite with stable mesopores.

The corners of the squares denote framework silicon and the sides denote the

chemical bonds between the framework atoms (Si and Al).

35

Page 45: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

The extra-framework material in the micro and mesoporous can be extracted by mild

acid leaching [195]. Zeolite Y with very high mesopore volume of almost exclusively

cylindrical mesopores can be obtained after hydrothermal treatment at temperatures

above 100 °C and pressures above 1 bar [196]. This hydrothermal treatment

deviates from steaming, which is performed with gaseous steam at atmospheric

pressures.

Although the created mesopores are beneficial, the main disadvantage of steaming is

the partial amorphization of the zeolite framework [197]. Not only the relative

cristallinity drops with increasing severity of the treatment, leading to a decreased

amount of the active phase, also part of the micropores and mesopores are filled with

amorphous debris, leading to a partial blockage of active sites [198]. Another side

effect of steaming is that the number and nature of the acid sites is changed by the

extraction of aluminium [196]. In addition, if only steaming and no acid leaching is

applied, the bulk Si/Al ratio remains the same, but the framework Si/Al ratio

increases.

Acid leaching is used for the direct creation of mesopores, without the use of steam

or the removal of non-framework material created during the steaming process. The

first method consist of severe treatments with strong inorganic acids, where

aluminium is removed from the zeolite framework [199]. The effectiveness of this

technique depends on the zeolite used [200]. Especially in the case of mordenite,

direct acid treatments are used to generate mesopores [201]. The nature of the acid

used can be of great influence on the final mesopores structures [202]. Treating

calcined mordenite with oxalic acid resulted in a much higher bulk Si/Al ratio and less

mesopores compared to nitric acid [203]. One problem with acid leaching is that the

Si/Al ratio is changed (Si/Al ratio is increasing), resulting in a loss of active sites

[200]. For several purposes one would like to separate the contributions of the

changes of the acid sites and the generation of mesopores in order to study their

influence on catalytic reactions. With this method, both variables are changed at the

same time, thus complicating the interpretation of the activity of the catalyst.

The second method is frequently applied because during steaming material is

extracted from the framework and subsequently deposited in the micro- and

36

Page 46: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

mesopores and on the external surface of the zeolite crystallites [195]. A mild acid-

leaching step, with either inorganic acids such as diluted nitric acid or organic

(complexing) acids such as oxalic acid, dissolves this extra-framework material [195].

In this case the mesopores are not actually formed during the acid leaching process

[200]. Rather, the mesopores formed during the steaming process are emptied,

resulting in a higher mesoporous volume compared to zeolites that have only been

steamed [200].

The most widely applied technique to study the size and surface area of mesopores

in zeolites is nitrogen physisorption. Already from the shape of the nitrogen

adsorption and desorption isotherm valuable information on the presence and shape

of the mesopores can be deduced [195, 199].

In Figure 16, typical curves for the nitrogen adsorption-desorption of some

dealuminated zeolites Y are displayed. The NaY zeolite without mesopores will give a

type I isotherm, whereas after the formation of mesopores a combination of type I

and IV isotherms is found, as is observed for the three different dealuminated zeolite

Y [199]. The existence of a hysteresis loop in the isotherms is an indication for the

presence of mesopores, whereas the shape of the hysteresis loop is related to the

shape of the mesopores [199]. Roughly, a vertical hysteresis loop is indicative of

cylindrical mesopores, whereas a horizontal hysteresis loop indicates ink bottle-type

mesopores [199]. From the presence and shapes of the hysteresis loops in Figure

14, it is obvious that by varying the steaming treatment different amount and forms of

mesopores can be created.

37

Page 47: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Figure 16. Nitrogen adsorption and desorption isotherms on a series of zeolites Y (a):

zeolite NaY, (b): steamed once, (c): steamed twice and acid leached, (d)

hydrothermally treated by gaseous vapour (H2O) at temperature over 100°C and

atmospheric pressure [199].

For the analysis of the nitrogen physisorption data several methods are available.

Very often only the BET surface area of the material is given. Since for microporous

materials the boundary conditions for multilayer adsorption are not fulfilled, the

reported BET areas have no physical meaning [203]. They should be understood as

a number proportional to the total micropore volume rather than the specific surface

area [203].

A valuable tool for the analysis of the external surface area is the t-plot method [204].

The external surface area is the total surface area of all meso- and macropores. If the

crystallite size remains the same during the formation of mesopores the difference

between the external surface areas of the parent and the treated zeolite is the

surface area of the mesopores created [204].

38

Page 48: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Another method that is often applied is the Barret-Joyner-Halenda method (BJH),

which uses the desorption branch of the nitrogen isotherms to calculate the pore size

distribution and the adsorbed volume [205]. Up to approximately 4 nm diameter the

BJH method gives a reasonable good pore size distribution based on the Kelvin

equation [205]. However, below a pore diameter of 2 nm the Kelvin equation is not

valid anymore and, the area between 2 and 4 nm is prone to errors if a hysteresis

loop is present [205]. In that case a huge amount of nitrogen is released at once

between a relative pressure of 0.38 and 0.44, which is visible as the closing of the

hysterisis loop in the isotherm. The BJH method correlates the volume of released

nitrogen via the Kelvin equation to pores of 3-4 nm diameter [205]. This is not always

correct, since ink bottle type of pores with pore necks smaller than 4 nm diameter

also release their nitrogen between a relative pressure of 0.38 and 0.44 [206]. Many

examples are known where people claim that uniform pores of 4 nm diameters are

formed, which do not change during more severe treatments [207]. Sometimes the

BJH method is applied to the adsorption branch of the nitrogen isotherm. However,

many programs treat the adsorption isotherm as if it was a desorption isotherm in

order to apply the BJH method. This results in errors in the calculated pore size

distribution. Therefore, care should be taken when the BJH method is applied to the

adsorption branch [208].

On the other hand, electron microscopy and electron termographic are methods that

are being developed to obtain information on the three-dimensional shape, size and

connectivity of the mesopores [202, 209].

Various applications have been developed to show the positive effect that the

combination of micropores with mesopores has, mostly concerning the use of zeolite

Y cracking catalysts applied in the hydrocarbon processing industries [210].

The synthesis of fine chemicals in mesopores, created mainly in zeolite Y, has been

reported [211]. However, diffusion limitation effects play a major role due to the bulky

size of the molecules involved in the process [13]. With regard to this, in our group,

the beneficial role of mesopores in zeolite Y was confirmed for the selective

isomerization of α-pinene oxide. Dealuminated zeolite Y exhibited high selectivities

towards the preferred campholenic aldehyde.

39

Page 49: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

After exposing this material to a mild acid treatment, a much higher activity as well as

a slightly enhanced selectivity was observed. This was explained by the fact that the

mesopores formed during the steaming process are emptied by the acid treatment,

resulting in a higher mesopore volume compared to the sample that was only

steamed. The mesopores allow for a good diffusional transport of large organic

molecules, even at the low temperatures that are often favored in asymmetric

synthesis [17].

Later, a combination of various steps of dealumination on zeolites X and Y was

performed to obtain mesopores surrounded by micropores as depicted in Figure 17

[212]. In the first step, framework aluminium was removed at 250 °C using silicon

tetrachloride (SiCl4) as silicon source. This isomorphic substitution did not create

mesopores, but was performed to protect the crystalline structure of the zeolite during

the subssequent dealumination with steam at 600 °C and 850 °C during which the

mesopores, were formed. Finally, the extra-framework alumina formed during the

steaming step was removed from micro- and mesopores by mild leaching acid using

diluted hydrochloric acid at pH = 2.7.

Figure 17. Generation of mesopores inside Faujasite type zeolites [212].

In the newly created mesopores, various highly active organometallic complexes for

asymmetric epoxidations were assembled [213]. These organometallic complexes,

which were formed out of their building units (active amine and salicylaldehyde

moieties), are entrapped by the microporous system which surround the created

mesopores according to the “ship-in-a-bottle” approach. Extraction with solvents was

used to remove any immobilized species from the surface of the crystallite. The

activity of these materials for the diastereoselective epoxidation of (–)-α-pinene and

R-(+)-limonene using oxygen and pivalaldehyde as oxidation system was

40

Page 50: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

investigated [214]. The catalytic behaviour for the epoxidation of (–)-α-pinene using

(R,R)-(N,N’)-bis(3,5-di-tert-butylsalycilidene)-1,2-diphenylene- 1,2-diaminocobalt

[214] was comparable with that of the un-immobilized complex (homogeneous

catalyst) was obtained.

2.4.1.2 Zeolites as support for asymmetric epoxidation catalysts

Zeolites are the most popular support to immobilize organometallic complexes, since

they consist of an inert porous structure with highly specific surface areas [215]. In

particular, the well-defined cavities of the faujasites zeolites X and Y, with cavities

almost spherical of 12 Å and EMT zeolite, with hypercages of 13*13*14 Å, are

attractive for the immobilization of active homogeneous catalysts using the “ship-in-a-

bottle” concept [216]. Obviously, this method depends on the respective sizes of the

complex and the cage that serves as host [216]. Therefore, the catalyst is held

sterically, thus preventing leaching during the reaction.

The encapsulation of manganese (III) salen complexes in zeolites was

simultaneously described by two groups, respectively using zeolite Y and EMT

zeolite [134, 135]. The first group reported that a trans-(R,R)-(1,2-

bis(salicyideneamino)-cyclohexane manganese complex, analogous to the

Jacobsen’s catalyst without tert-butyl groups, could be fitted into the cages of zeolite

Y [134]. The preparation of this material consisted of the introduction of Mn+2 ions by

ion exchange, followed by addition of the building units of the chiral ligand

(salicylaldehyde and trans-(R,R)-1,2-diaminocyclohexane), as shown in Figure 18.

The activity of this catalyst in epoxidation of unfunctionalized olefins using sodium

hypochlorite (NaOCl) as the terminal oxidant was investigated [134], and the rates of

the reaction for the heterogeneous catalyst were lower than the rates of the

homogeneous, which was explained with the restrictions imposed on the diffusion of

both substrate and products through the micropores of the zeolites [134]. Despite

this, comparable conversions were obtained for the indene and styrene epoxidations

[134]. In all cases a decrease in enantiomeric excess was observed upon

immobilization, although moderate enantioselectivities could be obtained for cis-β-

methyl styrene (e.e. of 58 % for the trans-epoxide) and indene (e.e. of 50%). The

slightly reduced enantioselectivities were interpreted as a combination of a non-

41

Page 51: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

catalyzed, non-selective epoxidation reaction in the liquid phase and/or the existence

of residual amounts of non-complexed Mn+2 acting as catalytic sites [134].On the

other hand, it has been suggested that the confinement effect of the zeolite may also

play an important role, since the zeolite is a rigid structure and the catalyst is a close

fit within the supercages [217]. This may subtly effect the conformation of the ligand

by preventing changes to ligand geometry during the heterogeneous reaction.

NH2CHONH2OH

HH

NNMn

O OO

_

+

1. Zeolite MnY, CH2Cl2Heat , 12 h.

2. O2

2

+

Figure 18. Immobilization of a (salen)Mn(III) complex in zeolite Y by encapsulation

[134].

The second group used the larger cages of the EMT structure for the accommodation

of a larger complex, carrying methyl groups on positions 5* and 5’*, and tert-butyl

groups on 3* and 3’* in the salen ligand [135]. Here, the activities of the

heterogeneous catalysts were lower, but enantioselectivities were unchanged [135].

Addition of pyridine N-oxide to the zeolites increased both activity and

enantioselectivity significantly [135]. For example, an enantiomeric excess of up to

88% was obtained for cis-β-methyl styrene. In these heterogeneous catalytic systems

42

Page 52: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

it could be proved that a hindered access led to the lower conversions if bulky

oxidants such as iodosylbenzene (PhIO) instead of NaOCl were used [135].

Both groups reported that the reaction takes place in the intrazeolite space and any

leaching was observed [134,135]. However, no recyclability experiments were

reported.

To overcome the problem of complex size and unfavourable interactions between the

complex and the support our group developed a sequential combination of

dealumination steps on zeolites X and Y (see Figure 17) [212]. These catalysts were

tested for the diastereoselective epoxidation of α(–)-pinene and R-(+)-limonene,

using molecular oxygen as oxidizing agent and pivalaldehyde as sacrificial aldehyde

[214]. The best results (100 % conversion, 96 % selectivity and 91% diastereomeric

excess) for α-(–)-pinene were obtained over a (R,R)-(N,N’)-bis(3,5-di-

tertbutylsalycilidene)-1,2-diphelyne-1,2 diaminocobalt complex (see Figure 19).

These catalysts were reusable without an appreciable loss of activity, and the

materials did not suffer from leaching [214]. These catalytic systems represent the

development of a truly heterogeneous processs. However, they were reported to be

inactive for the enantioselective epoxidation of prochiral olefins [218].

43

Page 53: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

CHO

N NC

O O

Co

C C

C

COOH

O

Co-salen complex

O2, RT

Co-salen complex =

Conversion = 100 %Selectivity = 96 %Diastereomeric excess = 91 %

Figure 19. Highly diastereoselective epoxidation of α-(–)-pinene over a Co-salen

complex immobilized on partially dealuminated zeolite Y [218].

In a recent study, a Jacobsen type salen complex, with sulfonato groups (SO3-) in 5*

and 5’* position on the aromatic ring, was immobilized in the interlayer space of the

Zn(II)-Al(III) oxides containing hydrotalcite [219]. This material was active in the

diastereoselective epoxidation of R-(+)-limonene, under identical conditions reported

by our group [218]. In contrast, the stability of the complex was explained by the

electrostatic interaction between sulfonato groups and positively charged support

layers [219].

44

Page 54: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

2.4.2 Mesoporous materials

According to the IUPAC-rules, mesoporous materials have pore diameters ranging

from 20 to 500 Å [220]. The most promising materials in this group were reported in

the late 1980s and early 1990s. In 1988 Yanagisawa et al. [221] reported the

synthesis of ordered mesoporous materials with narrow pore size distributions and

large surface area. These materials were prepared from the layered polysilicate

kanemite [222]. The first formation mechanism included a two-step sequence in

which first the Na+ ions in the kanemite interlayers are ion exchanged by

alkyltrimethyl ammonium cations, and second, the flexible kanemite layers fold and

cross-link to each other, thus forming the inorganic framework [223]. Also, it has been

proposed that in the second step a reorganization of the kanemite layers occurs

under influence of the polar head groups of the surfactants [224].

In 1992 researchers of the Mobil Research and Development Corporation published

the synthesis of a new group of mesoporous materials, designated M41S. These

materials are also characterized by narrow pore size distributions, tunable from 15 to

100 Å and large surface areas [225]. The main groups of this family are MCM-41,

which has a one-dimensional, hexagonally ordered pore structure, MCM-48, which

has a three-dimentional, cubic pore structure, and MCM-50, which has an unstable

lamellar structure. These materials are fundamentally different from zeolites by the

fact that the pore walls are amorphous. The ordering lies in the pore arrangements

[226]. The M41S materials have attracted far more scientific interest than the

materials derived from kanemite, even though studies have revealed that their

physico-chemical properties are quite similar [227].

For the M41S synthesis four basic components are used: the structure-directing

surfactants, a silica source, a solvent, and an additive (an acid or a base) [225]. In

their pioneering work on the M41S materials, the Mobil researchers used

alkyltrimethyl ammonium halides as the structure-directing surfactants and

combinations of sodium silicate, tetraethoxy silicate (TEOS), fumed silica and Ludox

as the silica source [225,226]. Sodium hydroxide or tetraethyl ammonium hydroxide

was used as basic additive to the aqueous synthesis solutions. In the case of

aluminosilicate materials, an aluminium source (normally, aluminium isopropoxide)

45

Page 55: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

was added as well. The synthesis solutions were kept at temperatures ranging from

100 to 150 °C for 24-144 h. The solid product recovered by filtration is calcined at

540 °C under a gas-flow of nitrogen followed by air, resulting in porous structures

[225]. The Mobil researchers found that the relative concentrations of the species

present in the synthesis solution were of great importance for the final pore

structures. They also showed that the pore diameter of MCM-41 increases as the

chain length of the surfactant increases [228]. Later on, modified synthesis methods

that improved the long-range structure of MCM-41 were suggested [229, 230].

The synthesis mechanism of M41S materials has been suggested to follow a liquid

crystal templating (LCT) mechanism due to its remarkable similarity with the lyotropic

liquid crystalline phase and the fact that the pore sizes were tunable from 15 to 100

Å, either by varying the chain length of the surfactants or by adding an organic agent.

In the model two different reaction pathways were proposed [231]:

1. The silica or aluminiumsilica framework is formed around pre-existing liquid

crystalline phases.

2. The addition of inorganic species somehow affects the assembly of surfactants

into ordered arrays.

Other studies claimed that surfactants in solution conduct the ordering of the

materials, but the type of interactions between the surfactants and inorganic species

are viewed differently [232]. Stucky and co-workers [233] have done the most

extensive studies on the formation mechanism of the mesoporous structure. They

developed a cooperative model that describes the formation of surfactant-inorganic

materials [233]. This is a dynamic model and no preorganized organic arrays are

necessary. If such an array is present, another array configuration can be expected in

the final structure. Generally, the synthesis was assumed follow an organization of

hydrophobic and hydrophilic entities into a biphasic composite material [233].

Therefore, three types of interactions are involved in this cooperative process:

inorganic-organic, organic-organic and inorganic-inorganic interactions [233].

For catalytic applications the siliceous mesoporous materials do not have enough

intrinsic acidity to be catalytically active. However, catalytic activity can be created by

46

Page 56: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

modification of the siliceous framework by other elements (often metals). This may be

done in several ways. One method is direct addition of extra precursors to the

synthesis solution. Hence, the modifying elements are incorporated into the structure.

Another possibility in this method is to modify the materials after a desired structure

has been synthesized [170]. A second method is the deposition of metal in the

structure by the so-called incipient wetness impregnation technique. 26-30 % of the

silicon atoms were connected to hydroxyl groups to form so-called silanol groups

(SiOH) [234], which are attractive anchor sites for the attachment of organic

compounds. Generally, silylation procedures are performed to find the appropriate

linker for the immobilization of catalytic organometallic structures (chemical bonding

method) [235].

Most studies on modified M41S materials are based on aluminium containing

materials due to its acidic properties similar to zeolitic materials [236]. For Al-MCM-

41, it is generally accepted that tetrahedral aluminium is incorporated into the wall

structure, while octahedral aluminium is regarded as extra framework species [237].

In most cases, cationic surfactants as structure-directing agents have been applied in

the synthesis of these materials. After calcination, they are replaced by Na+ cations

[238]. This fact has permitted the incorporation of other non-framework cations, thus

creating new catalytic properties similar to zeolites [236]. On the other hand,

incorporation of transition metals into the wall structure is necessary for the

preparation of mesoporous redox-catalysts. Due to the success of titanium-containing

zeolites in selective oxidation reactions, titanium modified mesoporous materials

were found to be very interesting for similar processing of larger molecules [239]. Ion

exchange is an easy strategy to incorporate metallic and organometallic catalysts,

but the stability of the incorporated species is very weak compared to directly in the

structure incorporated species [240].

The most valuable methods for structure elucidation of mesoporous materials are

powder X-ray diffraction (XRD), transmission electronic microscopy (TEM) and

adsorption-desorption measurements [170]. XRD provides direct information of the

pore architecture [241]. For mesoporous materials, the diffraction patterns only have

reflections in the low-angle range, meaning 2θ less than 10, without reflections at

higher angles [170]. It can therefore be concluded that the pore walls are mainly

47

Page 57: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

amorphous. TEM is a powerful tool to visualize different pore ordering [242].

Adsorption-desorption techniques are used to determine the porosity and surface

area. The most common adsorbate is N2 at 77 K [243]. According to IUPAC

definitions, mesoporous materials exhibit a type IV adsorption-desorption isotherm

[220]. At low relative pressure (p/po) the adsorption only occurs as a thin layer on the

walls (monolayer coverage). Depending on the pore size, a sharp increase is seen at

relative pressures between 0.25 to 0.5, which corresponds to capillary condensation

of N2 in the mesopores. The sharpness of the inflection reflects the uniformity of the

pore sizes and the height indicates the pore volume. A hysterisis effect is often

observed for adsorption-desorption isotherms when the pore diameter is larger than

approximately 40 Å [170]. Another important characterization technique for

mesoporous materials is solid-state nuclear magnetic resonance (NMR). 29Si MAS

NMR spectroscopy has been used to study the degree of condensation in the pore

walls. Their structure is found to be quite similar to amorphous silica, and usually two

or three broad signals (at –89/-92, -98/-102 and –108/-111 ppm) are observed. Which

can be assigned to Si(OSi)2(OH)2-, Si(OSi)3(OH)- and Si(OSi)4 groups, respectively

[244].

27Al MAS NMR spectroscopy can be used to distinguish between tetrahedrally and

octahedrally coordinated framework aluminium (at 50 and 0 ppm, respectively) [245].

Hence, dealumination processes can be evaluated [246]. Also, diffuse reflectance

FT-IR and UV/vis have been used to obtain information on the incorporation of other

elements into the siliceous framework, or those incorporated as compensation

cations [247]. Information about the oxidation state and local coordination geometry

of the transition metals can be obtained by X-ray Photoelectron (XPS), X-ray

Absorption Near Edge Structure(XANES) and Extended X-ray Absorption Fine

Structure Spectroscopy (EXAFS) [247].

2.4.2.1 MCM-41 as support for asymmetric epoxidation catalysts

Mesoporous materials like siliceous and Al-MCM-41 are popular supports to

immobilize bulky catalytic species, since these materials have pore sizes in the range

of mesopores (15 to 100 Å) [122]. The three-dimensional pore system of MCM-48

has the advantageous that the interior of the particles is more readily accessible

48

Page 58: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

because the pore openings are not restricted to one direction; therefore, pore

plugging problems are probably less [248]. However, for this material, few practical

examples about the immobilization of organometallic catalyst are known.

Covalent attachment is the method mostly applied to immobilize catalytically active

organometallic complexes on the Si-MCM-41 internal walls [65]. The most important

drawback of such heterogenized systems is the fact that the catalyst’s ligand has to

be functionalized, which can lead to an unfavourable geometrical conformation for

the chiral communication of the active species with the substrate [65]. This shows

that the tert-butyl groups of Jacobsen’s catalyst that not only serve to derivatize the

complex and subsequently link it to a support, but they are also crucial in inducing

enantioselectivity in the epoxidation reactions [105].

The immobilization of Jacobsen’s catalyst by covalent tethering has been studied

extensively [65]. First, a covalent link is formed between the silica support carrying

Si-H groups and the vinylfunctionalized ligand (see Figure 20). In this study, the

choice of oxidizing agent was crucial. A considerable amount of chlorinated by-

products were obtained when using sodium hypochlorite (NaOCl) as oxygen source,

whereas meta-chloroperbenzoic acid (m-CPBA) only resulted in moderate

enantioselectivities. The best catalytic activities were obtained using iodosylbenzene

(PhIO), although there is a slight decrease after regeneration [249].

OH

OH

O SiNN

Mn

O OCl

H H

CH3

CH3

Silica

Figure 20. The immobilization of Jacobsen’s catalyst on silica according to the

Janssen et al. [249].

49

Page 59: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

A considerable number of papers have included the functionalization of the MCM-41

internal surface by silylation as previous step to the immobilization. It consists of a

reaction between the surface silanol groups and a silylation agent in order to provide

the appropriate linker which forms a covalent bond with the catalyst [250]. The most

popular silylation agent is 3-aminopropyl trimethoxysilane (APTES) [251]. For

example, Kim et al. [252] reported a salen complex immobilized on the surface of

mesoporous MCM-41 in a step by step manner. Reacting APTES with MCM-41

molecular sieves resulted in a modified MCM-41 with a free amino group, which was

used to connect the salen ligand. A multifunctional group aldehyde 2,6-diformyl-4-

tert-butylphenol was attached to MCM-41 through imine bond formation. Thus, with

different chiral diamines and different substituents on the benzene ring, a variety of

Mn(III)-salen complexes (including Jacobsen’s catalyst) were attached to molecular

sieves (see Figure 21), and were tested for the asymmetric epoxidation of styrene at

–78 °C, using m-CPBA/NMO as oxidant. The best results were 89 % ee at 92 %

conversion after 4h of reaction [252]. In comparison, 84% ee at 97% conversion after

45 min was observed over the homogeneous catalyst. The authors reported that the

catalyst could be reused at least 4 times without activity loss, although experimental

data of recyclability were not provided [252].

O

O

O Si N C

Si-MCM-41

NNMn

O OCl

H H

Figure 21. The immobilization of Jacobsen’s catalyst on Si-MCM-41 according to Kim

et al. [252].

50

Page 60: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

More recently, Biggi et al. [253] reported the heterogenization of Jacobsen’s catalyst

using a longer linker in order to improve the access of the olefin to the metal center

during the reaction (see Figure 22). This material was tested in the asymmetric

epoxidation of various unfuntionalized olefins using m-CPBA/NMO as oxidation

system. Again, heterogenization of the catalyst did not cause a significant decrease

in enantioselectivity, although the conversion was affected tremendously [253]. The

lower yield could be attributed to diffusion limitations inside mesoporous structure.

The catalyst supported on non-mesoporose silica showed similar enantiomeric

excesses, and in addition a positive effect in the catalyst recycling [253].

O

O

O

Si

N

NN

C8H17O O

H

NNMn

O OCl

H H

Si-MCM-41 or silica gel

N

Figure 22. The immobilization of Jacobsen’ catalyst on Si-MCM-41 or silica gel

according to Bigi et al. [253].

In earlier work, Pini et al. [254] studied the immobilization of Jacobsen’s catalyst on

silica gel by forming a sulphide link between the ligand and an organically modified

silica, using the two methods showed in Figure 23. 1,2-dihydronaphthalene, indene

and 1-phenyl cyclohexane were epoxidised with enantioselectivities of 31%, 38% and

58%, respectively. However, neither comparison with homogeneous system nor

recyclability results were reported [254].

51

Page 61: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

OH

OH

(MeO)3Si SH

R

HN

OMe

H

O

N

O

R: (CH2)S(CH2)3Si(OMe)3

Mn

Si

Cl

O

O

SH

R

OOMe

OSi S

H

N

H

N

O

Mn

ClO S

OMe

SiO

O

Silica gel

TolueneReflux 40 hr

R: CH=CH2AlBN, CHCl3Reflux 44 hr

Figure 23. The immobilization of Jacobsen’ catalyst on silica gel according to Pini et

al. [254].

The presence of Al in Al-MCM-41 gives the possibility to incorporate highly active

catalytic species by ion exchange. In particular, their larger pore size is attractive to

retain organometallic complexes [122]. Frunza et al. [255] reported a series of

molecular sieves, including Mn(III) salen complexes obtained by simple impregnation

on mesopores structures (see Figure 24). Guest/host interactions, mainly between

aromatic rings of the complex and the internal silanol groups of the walls of the

mesoporous structures, were claimed to be stable. The epoxidation of 1,2-

dihydronaphthalene with NaOCl proceeded with exactly the same enantioselectivity,

but with lower conversions than the homogeneous reaction. The conversion

reduction was explained by the lower content of catalyst on the support. The

reusability of the catalyst was not tested [255].

52

Page 62: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Figure 24. The immobilization of Jacobsen’ catalyst on Al-MCM-41 according to

Frunza et al. [255].

Also, Hutchings et al. [256] reported an immobilization method for this complex based

on ion exchange. The salen ligand was connected with a Mn exchanged Al-MCM-41

by refluxing in dichloromethane to give the salen incorporated molecular sieves (see

Figure 25). The asymmetric epoxidation of cis-stilbene and iodosylbenzene as

oxygen donor gave slightly lower enantioselectivities than the homogeneous

reaction.

53

Page 63: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

Figure 25. The immobilization of Jacobsen’ catalyst on Al-MCM-41 according to

Hutchings et al. [256].

In the last two reports, recycle experiments were not provided. In general, it is

accepted that the immobilization by physical interaction is too weak, therefore, the

homogeneous catalyst will readily leach into solution as the catalyst will reach

equilibrium between the surface adsorbed species and solution species [65].

A combination of covalent attachment and electrostatic interaction has also been

used to immobilize Jacobsen’s catalyst (see Figure 26). This material was used for

the asymmetric epoxidation of α-methylstyrene, using NaOCl as oxidant and

dichloromethane (CH2Cl2) as solvent. Similar enantioselectivities and less conversion

than the homogeneous reaction were obtained. The conversion could be improved

54

Page 64: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

when polar solvents such as acetone and ethanol were used. It was shown that the

catalytic species are mainly anchored in the mesopores of MCM-41, since reaction

with 1-phenyl cyclohexane (substrate too large) was not observed. This material

could be reused three times, without a notable loss in activity [257].

Figure 26. The immobilization of Jacobsen’ catalyst on Si-MCM-41 according to Li et

al. [257]

O O OEt

Si

Si-MCM-41

O

MnN N

O O

ClHH

C C

Zhou et al. [258] reported another similar strategy for the immobilization of a Cr(III)

salen complex (see Figure 26). In this work, the unmodified Cr-binaphthyl Schiff base

was coordinated to an amine ligand, which was previously attached covalently to the

walls of Si-MCM-41. The activity of this material was studied with the epoxidation of a

series substituted styrenes, using iodosylbenzene as the oxidant. The best results

were achieved with 4-chloro styrene (65% yield and 65% ee). Under identical

conditions the heterogenized catalyst gave significantly higher enantioselectivities

than the free catalyst. This was attributed to the enhanced stability of the chromium

55

Page 65: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Background

complex upon immobilization, or the unique spatial micro-environment constituted by

both the chiral binaphthyl Schiff base ligand and the surface of the support. The

catalyst was recycled 4 times, both with yield and ee decreasing strongly after the

second reuse [258].

N NC C

O O

Cr

Cl

Br

Br Br

Br

O O O

Si

NH2

Si-MCM-41

Figure 27. The immobilization of a Cr(III) salen complex on Si-MCM-41 according to

Zhou et al. [258].

56

Page 66: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

3 Results and discussion

3.1 Motivation and goal

According to our knowledge the heterogeneous enantioselective epoxidation of

functionalized olefins (electron poor substrates) has only been studied for olefins

containing hydroxyl groups (allylic alcohols), over catalytic systems as developed by

Sharples and Katsuki [75]. In their studies re-use of the catalyst was not investigated.

Another important type of functionalized olefins, cinnamate ester derivatives, have

been efficiently homogeneously epoxidized with high enantioselectivities, using a

chiral Mn(III) salen complex developed by Jacobsen et al., commonly denoted

Jacobsen’s catalyst [104]. However, this process is commercially not viable due to

the high cost of the used catalytic system [5]. Therefore, the catalyst should be

recovered and re-used without a loss of its catalytic properties [65]. This often proves

to be difficult. A method for overcoming this difficulty is by immobilizing the

asymmetric catalyst on a non-soluble support, thus creating a chiral heterogeneous

catalyst, which can easily be recovered from reaction mixtures [122].

The goal of this thesis is to study the heterogeneous asymmetric epoxidation of cis-

ethyl cinnamate, using the catalytic system developed by Jacobsen. The epoxide

(2R,3R)-ethyl 3-phenylglycidate is an important intermediate for the synthesis of the

(2R,3S)-3-phenylisoserinamide, which is used for the total synthesis of the C-13 side

chain of taxol [25]. Taxol is an important chiral drug against different types of cancer

[25]. Furthermore, some other unfunctionalized olefins (styrene and 1,2-

dihydronaphthalene) have been included to determine correlations between the

catalyst activity and recyclability with substrate’s nature. Zeolites modified by

dealumination [212] and MCM-41 (with and without Al) are used as the supports. The

“ship-in-a-bottle” approach was chosen to be the appropriate method in modified

zeolites, whereas immobilization by chemical bonds (ionic or covalent bond) was

used in MCM-41.

Porous inorganic solids, such as zeolites and mesoporous materials, were chosen as

appropriate supports for the immobilization of homogeneous catalysts due to their

high chemical and mechanical stability [14]. On the other hand, their well-known pore

57

Page 67: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

structures not only allow possible shape selectivity, but they can also offer the

catalyst an appropriate environment to form the required geometrical conformation

which leads to an efficient stereochemical communication between the active species

and substrate [65]. Recently, our group, reported on a Co(II)-salen-complex

entrapped inside the intrazeolitic space of a USY-zeolite according to the “ship-in-the-

bottle” method, adopting a distorted geometrical conformation, which was responsible

for the observed high asymmetric induction in the transhydrogenation of

acetophenone at higher temperatures (25 °C), while the homogeneous system has to

be cooled down to -10°C [259].

Most works on the heterogenization of Jacobsen type catalytic systems was done for

the unfunctionalized olefin epoxidation [260]. This is the first investigation on the

heterogeneous epoxidation of a functionalized olefin (cis-ethyl cinammate).

Generally, it is accepted that electron poor olefins are less reactive than

unfunctionalized olefins [261]. Various immobilization methods have been studied

and summarized in recent reviews [65, 260]. Most of the reported methodologies

claim the retention of the catalytic propeties of the homogeneous catalysts. However,

only in a few cases the successful reuse of the catalyst could be shown [65]. In this

context, reaction conditions can be an important factor in order to combine the

excellent catalytic behaviour of the homogeneous catalysts with the reuse of its

heterogeneous counterpart. Concerning Jacobsen’s catalyst, two routes of catalytic

deactivation have been identified. The first one is the formation of oxo-

manganese(IV) dimeric species, which can be prevented either by the presence of a

nitrogen base [261] or by isolation of the catalyst by a solid support [116]. The

second one is the degradation of the salen ligand due to the prolonged oxidation

conditions. The latter has been identified as the main cause of non-reusability in

many heterogeneous systems [260]. Sodium hypochlorite has been chosen as the

suitable oxidizing agent in the homogeneous phase [261]. In contrast, for

heterogeneous systems it has been proposed that the use of NaClO as oxygen

source is not convenient, since the exchangeable cation (Na+) can replace the Mn

containing in inorganic supports such as zeolites and mesoporous materials [262]. To

prevent this kind of catalytic deactivaction, organic oxygen sources such as

iodosylbenzene (PhIO) and m-chloroperoxobenzoic acid (m-CPBA) are preferred.

However, in many cases it has been reported that when using these oxidants, the

58

Page 68: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

catalyst is not reusable due to the salen ligand degradation [263]. Oxygen sources

obtained from Oxone® (2 KHSO5 ● 1 KHSO4 ● 1 K2SO4), such as dimethyldioxiranes

(DMD) and tetrabutyl ammonium monoperoxosulfate (Bu4NHSO5), have emerged as

appropriate oxidizing agents for the asymmetric epoxidation of various prochiral

olefins, when using chiral organomanganese complexes as catalysts [88]. For

example, the combination of Jacobsen’s catalyst and isolated DMD prepared from

Oxone® and acetone has been demonstrated to be a useful oxidation system for the

homogeneous enantioselective epoxidation of 2,2-dimethyl-2H-chromenes and

electron poor olefins such as isoflavone derivatives [111]. In situ generated DMD has

only been shown to be efficient in the case of non-asymmetric epoxidations [265]. In

this case, the pH control plays a fundamental role in the formation of DMD. It is

generally accepted that DMD is formed at a pH between 7.0 and 7.5 [112]. At a pH

below 7.0 no reaction is observed, whereas at pH above 8.0 the Oxone®

decomposition is favoured [112]. Unfortunately, oxygen sources of high

environmental value, such as molecular oxygen (O2), hydrogen peroxide (H2O2) and

organic hydroperoxides, do not present good results for the asymmetric epoxidation

of prochiral olefins over Jacobsen’s catalyst [113]. Most of them are using an

additional compound for the activation of the oxygen source [114]. In many cases,

this reagent is not environmentally compatible. For example, when using molecular

oxygen as oxygen donor, the addition of an aldehyde is necessary to form the

intermediate reactive species (in situ formation of a peroxyacid), which oxidizes the

organomanganese complexe to obtain the oxo-manganese active species [114].

On the other hand, the use of in situ generated dioxiranes from Oxone® and a chiral

ketone as catalysts has been shown to be a remarkably promising oxidation reagent

[264]. However, the destruction of the chiral ketone due to the Beayer-Villiger

reaction has limited its reuse [110].

As already mentioned, the presence of nitrogen containing compounds (generally

pyridine derivatives) has a positive effect on the catalytic activity, which is remarkably

more so for unreactive olefins such as cinnamate esters derivatives [261]. For

Jacobsen’s catalyst under biphasic reaction conditions, 4-phenylpyridine N-oxide (4-

PPNO) is used as additive, while 4-methylmorpholine-N-oxide (NMO) is preferred in

organic reaction media [266]. The function of the additive is to act as an axial ligand

59

Page 69: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

to the catalytic center to stabilize the oxo-manganese catalytic species, thus

preventing the formation of inactive dimeric µ-oxo manganese(IV) species [261].

additionally, it has been proposed that the additive plays a role as a phase transfer

catalyst (PTC) in a water-organic biphasic medium, for example when NaOCl is used

as oxygen donor [261]. Normally, dichloromethane or acetonitrile have normally

been found as the suitable reaction solvents [261]. In a recent study, Li et al. [257]

found that polar solvents such as acetone and ethanol have a favourable effect on

the re-use of the catalyst. Usually, low reaction temperatures are optimal under

homogeneous conditions, since it minimizes the reactivity of the oxomanganese

species, favouring the enantioselective route [261]. However, under these conditions

the heterogeneous counterpart suffers a reduction in the reaction rate due to the

diffusion limitations imposed by the porous system of the support [134,135].

3.2 Preparation and characterization of the catalysts

3.2.1 Preparation and characterization of the support materials

3.2.1.1 Highly dealuminated zeolites X and Y

X and Y zeolites (Na-X and Na-Y) were modified according to the method reported by

Hölderich et al. [218]. This method basically consisted in the extraction of framework

aluminium from the zeolitic structure by combination of several dealumination steps

(Figure 17). In the first step Na-X and Na-Y were treated with silicon tetrachloride

(SiCl4) to protect the zeolite crystalline structure against hydrothermal processes at

high temperatures (higher than 873 K), and to increase the Si/Al ratio, which favours

the formation of mesopores in the steaming steps [218]. The obtained materials were

converted to the ammonium form by ion exchange with ammonium salts (NH4Cl and

CH3CO2NH4). Subsequently, these materials were treated with H2O vapour

(steaming), first at 873 K and then at 1123 K. Finally, the resulting materials were

treated with hydrochloric acid (pH = 2.7) at room temperature to remove the extra-

framework material (aluminium moieties), which were formed during the steaming

process. These zeolites, modified by the dealumination processes were denoted JCX

and JCY. The dealuminated zeolites clearly show the diffraction pattern of the

corresponding parent material (Figures 28 and 29). Even though there was some

60

Page 70: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

loss of crystallinity, zeolite Y was more stable than zeolite X. Zeolite Y is of great

commercial importance in petrochemical processes due to its high thermal stability

[184]. On the other hand, the shift of the reflections to lower d-values, which becomes

more apparent at smaller d-values, could be attributed to the framework

dealumination and subsequent framework rearrangement, which caused a lattice

expansion [194].

Figure 28. X-ray diffractograms of the parent material Na-X and the modified X (JCX)

61

Page 71: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 29. X-ray diffractograms of the parent material Na-Y and the modified Y (JCY)

The formation of mesopores inside the crystalline structure of the dealuminated

zeolites could be shown with N2-sorption analysis. The adsorption-desorption

isotherms (Figures 30 and 31) are a mixture of type I and IV isotherms, which means

that the remaining micropores and created mesopores coexist in the modified zeolite

structures. These results showed that the generated mesopores are completely

surrounded by remaining micropores [213]. These new hosts are very attractive for

the immobilization of voluminous homogeneous catalysts such as Jacobsen type

catalysts, using the “ship-in-a-bottle” concept. For these catalysts, conventional

zeolites are not suitable due to the limited size of their intrazeolitic space [214]. It is

important to note that the mesopore formation leads to a tremendous reduction in the

total pore volume (see Table 5). This reduction is most remarkable in dealuminated

zeolite X. The reduction of the total pore volume may affect the preparation of the

catalyst inside the created mesopores.

62

Page 72: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 30. Nitrogen sorption isotherms of the parent material Na-X and the modified

zeolite (JCX)

Figure 31. Nitrogen sorption isotherms of the parent material Na-Y and the modified

zeolite (JCY)

63

Page 73: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

3.2.1.2 Al-MCM-41 and Si-MCM-41

Al-MCM-41 and Si-MCM-41 were prepared by conventional methods with some

modifications [270, 226]. Upon removal of the structure-directing agent

(tetradecyltrimethyl ammonium bromide) by calcination at 823 K, the obtained

mesoporous materials were characterized by XRD and N2-sorption at 77 K. The

materials were dried overnight in vacuum at 413 K and then left to cool down to room

temperature under argon prior to further treatments. XRD revealed a strong signal at

an 2θ = 2.4 and 1.8 for Al-MCM-41 and Si-MCM-41, respectively. It signal is

associated with the reflection d100. Also, both materials showed two less intense

signals at 2θ = 3.5 and 3.8 for Al-MCM-41 and 2θ = 2.4 and 2.8 for Si-MCM-41.

These signals are associated with the reflections d110 and d200 respectively. These

signals confirm the hexagonally ordered structure in the prepared materials [241]

(see Figure 32).

Figure 32. X-ray diffractograms for Al-MCM-41 and Si-MCM-41

64

Page 74: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

3.2.2 Characterization of the heterogeneous catalysts

3.2.2.1 Heterogeneous catalysts based on modified zeolites X and Y

The Jacobsen catalyst was prepared inside the generated mesopores of zeolites X

and Y by three different procedures. In the first one (method 1 see Figure 33),

manganese ions are introduced in the carrier material by ion exchange with sodium

ions. Subsequently, the salen ligand was built up around the Mn+2 ions by first adding

the active amine ((1R,2R)-(+)-1,2-diaminocyclohexane L-tartrate), followed by the

salicylaldehyde derivative (3,5-di-tert-butyl-2-hydroxybenzaldehyde) [213]. The

materials were denotet JCXSB1 and JCYSB1. In this method has been suggested

that the support material provides negative charges to stabilize the

organomanganese cation. These charges are associated with framework oxygen

atoms [134].

Figure 33 . Immobilization of the Jacobsen’ catalyst on dealuminated zeolite X and Y

using the method 1 [213]

In method 2 (see Figure 34), Mn atoms are also incorporated in the support by ion

exchange, but in this method the preparation of the salen ligand by addition of

components (active amine and salicylaldehyde derivative) was in the opposite order

[134]. In addition, an oxidation step was included to oxidize Mn+2 to Mn+3 [134]. The

65

Page 75: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

materials were denoted JCXSB2 and JCYSB2. Also, in this method the host structure

has been claimed to stabilize the Jacobsen’s cation [134].

Figure 34. Immobilization of Jacobsen’ catalyst on dealuminated zeolite X and Y

using method 2 [134]

In method 3 (see Figure 35), the salen ligand was first synthesized [135].

Subsequently, the metal is incorporated by complexation with the preformed salen

ligand, followed by an oxidation step and the addition of a chloride source [135]. The

materials were named JCXSB3 and JCYSB3 [135]. In this case, chloride anion

neutralize the positive charge generated by the Jacobsen’s cation as in the

homogeneous catalyst.

All materials were washed thoroughly using dichloromethane (80°C, 24 h) and

toluene (110°C, 24 h) in a Soxhlet apparatus until no color in the washing solvents

was observed. After that, the solids were dried at room temperature.

66

Page 76: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 35. Immobilization of Jacobsen’ catalyst on dealuminated zeolite X and Y

using method 3 [135]

Table 5 depicts the N2-sorption results for the various materials, using the “ship-in-a-

bottle” concept. For the support materials (modified zeolites JCY and JCX) obtained

by dealumination steps [212], the reduction in the surface area is in accordance with

the loss of total micropore volume. The third column in Table 5 indicates the

mesopore volume according to the BJH method. From these data the mesopore

formation in the modified zeolites can also be verified. This is in accordance with the

adsorption-desorption isotherms (see Figures 30 and 31). After immobilization both

the micropore and mesopore volume decreased for all three procedures used,

showing that organic species occupied both types of pores. The reduction of total

pore volume was mostly remarkable for the heterogeneous catalysts based on

modified zeolite X.

67

Page 77: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Table 5. N2-sorption data of the materials based on modified zeolites

Material BET surface area (m2/g)

Micropore volume (cm3/g)

BJH pore volume (cm3/g)

Na-Y 721.6 0.328 0 JCY 597.3 0.245 0.165 JCYSB1 401.4 0.160 0.100 JCYSB2 396.2 0.158 0.098 JCYSB3 393.0 0.153 0.102 Na-X 623.1 0.283 0 JCX 351.3 0.108 0.201 JCXSB1 95.6 0.073 0.045 JCXSB2 59.8 0.020 0.037 JCXSB3 40.0 0.001 0.077 The framework SiO2/Al2O3 ratio (determined with FT-IR) and surface area

(determined with N2-sorption) for the dealuminated zeolites decreased considerably,

as was reported on earlier in our group [212] (see Table 6), revealing the removal of

framework aluminium. However, the total aluminium content determined by ICP-AES

is almost similar to the parent material, which means that a considerable amount of

extra-framework aluminium was not removed after of the treatment with diluted

hydrochloric acid (pH = 2.7). This was also demonstrated with 27Al MAS NMR [268].

Here, the spectra of the dealuminated zeolite indicated the presence of both

tetrahedrally coordinated aluminium (signal around 60 ppm) and non-framework

octahedrally coordinated aluminium (signal around 5 ppm) [268]. Also in Table 6, the

composition of the materials based on modified zeolites is presented. The difference

in the Mn/ligand ratios can be due to the different preparation methods. Although

general correlations cannot be found easily, it can be concluded that method 3

always leads to materials with an slightly excess of ligand, which is frequently

reported when zeolites are used as carrier material [134, 135].

68

Page 78: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Table 6. Composition of the materials based on modified zeolites.

Material SiO2/Al2O3a

(mol/mol) Mn loadingb

(w%) Salen loadingc

(w%) Mn/ligand (mol/mol)

Na-Y 6.3 - - - JCYSB 188.7 - - - JCYSB1 188.7 0.49 6.4 0.77 JCYSB2 188.7 0.79 4.40 1.78 JCYSB3 188.7 1.13 13.70 0.71 Na-X 3.0 - - - JCXSB 188.6 - - - JCXSB1 188.6 1.20 7.3 1.6 JCXSB2 188.6 0.57 3.11 1.8 JCXSB3 188.6 0.83 13.01 0.63 aDetermined with FT-IR from the symmetric Al-O stretch vibration [267]. bDetermined with ICP-AES. cDetermined with TA. The identity and purity of the retained complex was first investigated by

thermogravimetric analysis (see Figures 36 and 37). For comparison reasons, the

region between 30-120 °C, associated with the removal of physisorbed H2O and

immobilization solvents is not included. From these curves, it can be seem that the

non-immobilized Jacobsen’s catalyst first shows a very small loss of mass at 130°C,

which can be associated with the loss of chloride as hydrogen chloride (the chloride

is the counterion of Jacobsen’s cation) [255]. The main decomposition temperature

occurred at 320 °C. Another remarkable loss of mass appears at 480 °C, which can

be associated to the destruction of the more stable groups (tert-butyl groups located

in the aromatic rings of the ligand salen) [255]. Finally, above 500 °C, metal oxides

are obtained. For the heterogeneous catalysts, lower intensities were obtained. The

samples prepared by the method 3 (JCYSB3 and JCXSB3) show similar curves as

found for the free catalyst, whereas for the other samples, losses of mass near to the

mass loss of the free complex were obtained. These results suggest that the samples

obtained by method 1 and 2 contain different organomanganese complexes,

including the Jacobsen’s catalyst in very low concentrations.

69

Page 79: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 36. DTG of the free Jacobsen´s catalyst and of Jacobsen´s catalyst immobilized on modified zeolite X.

70

Page 80: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 37. DTG of the free Jacobsen´s catalyst and Jacobsen´s catalyst immobilized

on modified zeolite Y.

The UV-Vis spectra are depicted in Figures 38 and 39. The spectrum of the salen

ligand shows typical absorption bands at approximately 250 nm and 330 nm,

attributable to the salen ligand n-π* and π-π* charge transfer bands [255]. For the

Jacobsen’s catalyst these bands are shifted to 325 and 440 nm, due to the salen-to-

manganese charge transfer [255]. Additionally, at around 500 nm, a weak signal

appears. This is associated to the d-d transition in the manganese ions. The highly

dealuminated zeolites (JCY and JCX) did not show any signal, which indicates that

no organic material is present prior to the immobilization process. Upon

heterogenization, the corresponding bands are slightly shifted. The obtained samples

71

Page 81: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

with methods 1 and 2 had the least intense absorption bands, whereas intense

signals were obtained in the samples prepared by method 3.

Figure 38. UV-Vis spectra of the Jacobsen’s salen, Jacobsen’s catalyst, modified

zeolite Y and Jacobsen´s catalyst on modified zeolite Y.

72

Page 82: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 39. UV-VIS spectra of the Jacobsen’s salen, Jacobsen’s catalyst, modified

zeolite X and the different immobilized materials on modified zeolites X.

The FT-IR spectra of the catalyst heterogenized on modified zeolites X and Y are

shown in Figures 40 and 41. For the non-immobilized Jacobsen’s catalyst, the

absorption band at 1540 cm-1 indicates the complexation of the manganese ion by

the salen ligand [255]. The signal is commonly used as reference to verify the

immobilization of organomanganese complexes on solid supports. As in the previous

analysis methods, lower intensities were obtained for the heterogeneous catalysts.

However, the typical band of Jacobsen’s catalyst is present in all samples, having the

highest intensity for the one obtained by method 3.

73

Page 83: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 40. FT-IR spectra of Jacobsen’s catalyst and the different immobilized

materials on modified zeolite Y.

74

Page 84: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 41. FT-IR spectra of Jacobsen’s catalyst and the different immobilized

materials on modified zeolite X.

From the results obtained by TA, UV-Vis and FT-IR it becomes obvious that method 3 led more clearly to the heterogenization of Jacobsen’s catalyst. This is not

surprising, since in this method all the preparation steps as for the homogeneous

catalyst were applied [135]. In method 1 the synthesis of the ligand was carried out

at room temperature, no oxidation step for the manganese ion was performed and

the anion source to stabilize the Jacobsen’s cation was omitted. In method 2, the

ligand synthesis was carried out at the same temperature as for the free

homogeneous catalyst and the oxidation step was included, but no anion source was

added. In methods 1 and 2, it is proposed that the zeolite structure provides the

anionic moieties to neutralize the Jacobsen cation [134, 212].

75

Page 85: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

3.2.2.2 Heterogeneous catalysts based on Al-MCM-41 and Si-MCM-41

Al-MCM-41 and Si-MCM-41 with high surface areas were synthesized. These solids

were also used as carrier material. Solids with high surface area are preferred for

heterogeneous processes, which include either contact between liquid-solid or gas-

solid phases [170].

Three methods were used to incorporate Jacobsen’s catalyst. In the first one

(method 4 see figure 42), immobilization by ionic bonding using Al-MCM-41 as

support was performed. Manganese ions (Mn+2) were first incorporated by ion

exchange as in zeolites. After calcinations, the material obtained was denoted Mn-Al-

MCM-41. This material was treated with Jacobsen’s salen in CH2Cl2 at reflux

temperature, yielding a slight brown solid [256]. This solid was denoted MCM1HC.

Figure 42. Immobilization by ionic bonding between Jacobsen’s cation and the

aluminosilicate anion of the Al-MCM-41 framework (method 4) [256].

76

Page 86: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

For methods 5 and 6, the amino group functionalized Si-MCM-41 was used as

support. The modification of Si-MCM-41 was carried out by treatment with 3-

aminopropyltriethoxysilane (3-APTES) according to Figure 43 [274]. The obtained

material was denoted NH2-Si-MCM-41.

OH OH OH

H2N

Si-MCM-41Si

EtOOEt OEt

O

H2N

O

NH2-Si-MCM-41

Si

O

3 EtOH+

3-APTES

Dry toluene, reflux, 24 h, Ar+

Figure 43. Preparation of aminopropyl-functionalized Si-MCM-41 [274].

In method 5 (see Figure 44) the Jacobsen’s catalyst was treated with dry silver

perchlorate to change the [Jacobsen cation]+-Cl- ionic bond to a [Jacobsen cation]+-

ClO4-. In this way, the catalyst is allowed to form a coordination bond with the

nitrogen atom of the amino group previously anchored on Si-MCM-41 [274]. This

solid was denoted MCM2HC.

77

Page 87: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

H

O

N

Cl

HN

MnO

O

H

O

N

H2N

O

MCM2HC

Si

HN

MnO

O

AgClO4

H

O

NH

NMn

ClO4

ODry CH3CN, rt, 16h

+

1. NH2-Si-MCM-41 CH2Cl2, Ar, rt, 24 h

2. Soxhlet extraction CH2Cl2, 70 °C

+

Jacobsen's catalyst

Figure 44. Anchoring of Jacobsen’s catalyst on NH2-Si-MCM-41 through the

manganese atom (method 5) [274]

In the method 6 (see Figure 45) the catalyst was built up on NH2-Si-MCM-41 and

immobilized by covalent bonding of the salen ligand. This method started by the

reaction between 2,6-diformyl-4-tert-butylphenol and (1R,2R)-(+)-1,2-

diaminocylohexane in the presence of NH2-Si-MCM-41 to give half of the salen ligand

(NH2-MCM-41(1)). This solid was treated with 3,5-di-tert-butyl salicylaldehyde to give

the salen ligand linked to the nitrogen atom by chemical bonding (NH2-MCM-41(2)).

Finally, NH2-MCM-41(2) was treated with a manganese source (manganese acetate,

(CH3CO2)2Mn●4H2O) and chloride source (lithium chloride, LiCl) to give Jacobsen’s

catalyst [252]. This solid was denoted MCM3HC.

78

Page 88: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

+

HC

OH

CH

H H

H2N NH2

O

H2NH

O

Si

HN

N

HO

O

C

Reflux, EtOH, Ar, 14 h

NH2-Si-MCM-41

2,6-Diformyl-4-tert-butylphenol

(1R,2R)-(+)-1,2- Diaminocyclohexane

+

NH2-MCM-41(1)

OO COH

OH

3,5-Di-tert-butylsalicylaldehyde

HN

OH

HN

N

HO

C

O O O

Si

O

H

O

N

MCM3HC

O

Si

Cl

Mn

HN

N

O

O

C

Reflux, EtOH, Ar, 18 h

NH2-MCM-41(2)

1. Mn(CH3CO2)2.4H2O, LiCl, reflux, Air, 4 h.

2. Soxhlet extraction CH2Cl2, 70 °C.

Figure 45. Anchoring of Jacobsen’s catalyst in NH2-Si-MCM-41 via covalent bonding

of the ligand to the linker (method 6) [252].

79

Page 89: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Upon immobilization, both the surface area and mesopores volume decreased, the

most remarkable for the material obtained by immobilization method 4 (see Table 7).

However, in all the methods a significant fraction of the pore volume remains

accessible to be used in the catalytic process. In contrast, when zeolites are used the

remaining pore volume is too small and the reactant cannot approach the catalytically

active sites. Using Al-MCM-41 the diffusion limitation can be reduced.

Table 7. Nitrogen sorption data for the heterogenized catalysts over Al-MCM-41 and

Si-MCM-41.

Material BET surface

area (m2/g) BJH pore

volume (cm3/g) Al-MCM-41 981 0.978 MCM1HC (method 4) 890 0.550 Si-MCM-41 968 0.685 NH2-MCM-41 408 0.260 MCM2HC (method 5) 332 0.100 MCM3HC (method 6) 237 0.041

Al-MCM-41 was obtained with a Si/Al ratio equal to 40 (determinated by ICP-AES),

which was maintained after immobilization (see Table 8). This indicates that no

leaching of aluminium occured during the immobilization process. In contrast, in the

case of dealuminated zeolites, the extraframework aluminium formed during the

dealumination process was not completely removed from the zeolite crystal. This fact

may affect the immobilization process due to blocking of the pore system in zeolites

by aluminium oxide species [114].

The difference in the manganese load and ligand content obtained in the

heterogenized catalysts can be due to the different preparation methods, as was

case for the zeolites. Obviously, a Mn/ligand ratio equal to 1 is obtained when

Jacobsen’s catalyst is formed prior to the immobilization (method 5). The samples

obtained both by treatment of the salen ligand on Mn modified Al-MCM-41

(MCM1HC); and building up the salen ligand (MCM3HC); reveal an excess of metal.

Although the amount of salen ligand increases to reduce the free manganese

content, the Mn/salen ratio did not vary significantly. This indicates that some Mn

80

Page 90: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

sites in the mesoporous structure are not readily accessible for the salen ligand.

Another possibility to reduce the Mn/salen ratio near to 1 is to reduce the manganese

concentration, but in this way the catalytic activity can be reduced, too. On the other

hand, the difference in the catalyst load can be due to the different preparation

methods.

Table 8. Composition of the materials based on Al-MCM-41 and Si-MCM-41.

Catalyst Si/Ala (mol/mol)

Mn loadinga

(mmol/g) Salen loadingb

(mmol/g) Mn/salen

(mmol/mmol) Complex loading

(mmol/g) Al-MCM-41 40 - - - - MCM1HC 40 0.228 0.089 2.561 0.089c Si-MCM-41 - - - - - MCM2HC - 0.035 N.C 1.000 0.035d MCM3HC - 0.439 0.134 3.276 0.134c aDetermined by ICP-AES. bDetermined by TGA. cCalculated from the salen ligand

content. dCalculated from the Mn content. N.C: Not determined.

The identity of the retained complex was also investigated with thermogravimetric

analysis (see Figure 46). The mass losses of the homogeneous catalyst were already

discussed. The loaded complex shows a different decomposition behaviour due to

the guest/host interaction. All samples show a first weight loss at 130 °C, which

indicates that the positive charge of the organic cation is balanced by chloride and

not by the anionic host [255]. However, this is not totally clear, since in method 4 and 5 no chloride source is involved. In addition, it is expected that the host structure

is the anion when the immobilization is carried out by ion exchange (method 4).

81

Page 91: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 46. DTG of the free Jacobsen´s catalyst and Jacobsen´s catalyst immobilized

on MCM-41 type mesoporous materials.

The identity and purity of the organomanganese complex immobilized in the MCM-41

type mesoporous systems was also investigated by spectroscopic analysis with FT-

IR and (DR) UV-Vis. Figure 47 shows the FT-IR spectra for the salen ligand,

Jacobsen’s catalyst and heterogeneous catalyst MCM1HC. The C=N imine vibration

signal in the salen complex at 1630 cm-1 is shifted to the lower wave number of 1610

cm-1, when the complexation with the metal is occurred [255]. This signal should not

be taken in account to evaluate the immobilization process, since the non-loaded

82

Page 92: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

support (Al-MCM-41) shows a broad signal centered at 1640 cm-1. This signal is

attributable to the presence of H2O in inorganic materials [13]. However, the most

characteristic signal from the point of view of the structure of metal salen complexes

is the presence of the band appearing near 1540 cm-1 [255]. This signal is also

obtained in sample MCM1HC, which indicates that the Jacobsen’s catalyst did not

suffer any change in its chemical structure during the immobilization process (see

Figure 47).

Figure 47. FT-IR spectra of Jacobsen’s catalyst and Jacobsen’s catalyst immobilized

by ionic bonding.

83

Page 93: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

No immobilization on Si-MCM-41 was observed using method 4, indicating that

Jacobsen’s catalysts is mainly retained by ionic interactions rather than adsorptive

interactions with silanols groups. It is well known that the presence of framework

aluminium generates a negative charge, which permit the incorporation of other

cations different to sodium or potassium [143].

In contrast, the signal assigned to the metal salen complex could not be found clearly

in the spectrum of the MCM2HC and MCM3HC samples, because several signals

from amino propyl groups appear in the same zone of the IR spectrum (1450-1550

cm-1). In these cases (DR) UV-Vis was chosen as the most suitable spectroscopic

method (see Figure 48). The UV-Vis spectra of the Jacobsen’s salen shows typical

signals near at 260 and 330 nm, attributable to ligand n–π* and π–π* charge transfer

bands [275]. Jacobsen’s catalyst exhibits a broad signal centered around 425 nm

attributable to ligand-to-metal charge transfer band similar to metal salen compounds

[275]. NH2-Si-MCM-41 did not show any signal, as was observed for the

dealuminated zeolite. After immobilization, this signal appears with lower resolution

than for the homogeneous catalyst. However, the UV-Vis spectra of the

heterogeneous catalysts are very similar to the spectrum of the Jacobsen’s catalyst

in solution, revealing that the organomanganese complex is loaded into the

mesoporous structure. On the other hand, no signal was obtained when using

unmodified Si-MCM-41.

84

Page 94: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 48. UV-Vis spectra of Jacobsen’s catalyst and the Jacobsen’s catalyst

immobilized on NH2 group modified Si-MCM-41.

In comparison with the materials based on the dealuminated zeolites, the

immobilization of Jacobsen catalyst on Al-MCM-41 has two differences. First, the

characterization methods are easier to elucidate the nature of the retained catalyst.

Second, a higher concentration of catalyst is obtained.

85

Page 95: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

3.3 Catalytic activity

3.3.1 NaOCl/4-PPNO as oxidation system

The homogeneous enantioselective epoxidation of cis-ethyl cinnamate using

NaOCl/4-PPNO as oxidation system is well known [103], and is generally conducted

at low temperature (4°C). However, in the present study the homogeneous

epoxidation of cis-ethyl cinnamate was performed at room temperature, a yielding

almost the same results as at 4 °C (see Table 9). On the other hand, the procedure

at 4°C for the heterogenized catalysts showed very low conversions, probably

associated to diffusion limitations, which are most remarkable at low temperatures

[134, 135]. Therefore, the catalytic activity of the heterogeneous catalysts was

studied at room temperature. Blank experiments indicated no appreciable influence

on the activity by the support or uncomplexed manganese moieties.

The highly dealuminated zeolites heterogenized Jacobsen’s catalyst showed only a

poor racemic epoxidation. The low conversions can be explained by the diffusion

limitations imposed by the porous system of the support, which are most remarkable

in highly dealuminated zeolites due to the reduction of their total pore volume. In

presence of the heterogenized catalyst, mainly chlorinated compounds were found

(detected by GC-MS), therefore very low selectivities to the epoxides were obtained.

In addition, with this oxidation system no asymmetric induction was observed, which

indicates that the diffusion limitations might prevent the chiral communication

between the substrate and the catalyst. Using the heterogenized catalyst, the

cis/trans ratio also changed and in most cases the trans-epoxide was favoured, as

was reported in the diastereoselective epoxidation of α-(–)-pinene and R-(+)-

limonene [214]. In other related work with dealuminated zeolites it was found that

lower enantioselectivities than over the homogeneous catalyst were obtained for the

enantioselective epoxidation of 1,2-dihydronaphthalene [268]. In this study, calcium

hypochlorite led to the best results (40%ee). However, the enantioselectivity was

tremendously affected after the first reuse (10%ee), suggesting catalyst degradation

[268].

86

Page 96: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Table 9. Catalytic activity for the modified zeolites heterogenized catalyst, using

NaOCl as oxygen donora

Catalyst Conversion(%)

Epoxide Selectivity(%)

Cis/trans epoxide

eecis (%)

eetrans (%)

None 0 - - - - Homogeneous catalystb 98 99 4.1 97 70 Homogeneous catalyst 97 96 4.0 97 74 JCY 8 15 0.1 - - Mn-JCY 10 16 Only trans - - JCYSB1 10 10 1.0 0 0 JCYSB2 20 12 0.4 0 0 JCYSB3 25 15 0.6 0 0 JCX 5 0 - - - Mn-JCX 3 0 - - - JCXSB1 35 20 Only trans 0 0 JCXSB2 4 0 - 0 0 JCXSB3 15 10 Only trans 0 0 aCis-ethyl cinnamate = 0.5 mmol; NaOCl (0.55 M in Na2HPO4, pH = 10.2) = 2.7

mmol; 4-PPNO = 0.15 mmol; catalyst = 0.05 mmol (30 mg homogeneous or 300-500

mg heterogeneous catalyst); CH2Cl2 = 4 ml; T = 23-25 °C; t = 12 h. bT = 4 °C.

On the other hand, using the catalyst immobilized on Al-MCM-41, lower conversions

and selectivities than in the case of the homogeneous catalyst were obtained (see

Table 10). Therefore, generally NaOCl/4-PPNO as oxidation system led to low

chemical yields, when inorganic carriers such as zeolites and Al-MCM-41 are used

for the epoxidation of cis-ethyl cinnamate with Jacobsen’s catalyst. Diffusion

limitations through the pore system and the preferential formation of chlorinated

compounds were the main reasons for the obtained low yields. By increasing the

concentration of NaOCl, no change in the activity was observed.

When using Al-MCM-41 as support, the best enantioselectivities were obtained,

although these were lower than for the homogeneous catalyst. In comparison with

the materials based on dealuminated zeolites, the pore volume size was not sufficient

to allow for the chiral communication between the substrate and active species. The

presence of the nitrogen additive (4-PPNO) was necessary to obtain enantiomeric

excesses.

87

Page 97: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Table 10. Catalytic activity for the Al-MCM-41 heterogenized catalyst, using NaOCl

as oxygen donora

Catalyst Conversion(%)

Epoxide Selectivity(%)

Cis/trans epoxide

eecis (%)

eetrans (%)

None 0 - - - - Homogeneous catalystb 98 99 4.1 97 70 Homogeneous catalyst 97 96 4.0 97 74 Al-MCM-41 4 0 - - - Mn-Al-MCM-41 12 0 - - - NH2-Si-MCM-41 10 0 - - - MCM1HC 41 25 2.0 70 35 MCM2HC 17 27 2.0 58 25 MCM3HC 5 43 3.0 25 5 MCM1HCc 39 20 1.1 0 0 MCM1HCd 0 - - - - aCis-ethyl cinnamate = 0.5 mmol; NaOCl (0.55 M in Na2HPO4, pH = 10.2)= 2.7 mmol;

4-PPNO = 0.15 mmol; catalyst = 0.05 mmol (30 mg homogeneous or 300-500 mg

heterogeneous catalyst); CH2Cl2 = 4 ml; T = 23-25 °C; t = 12 h; bT = 4 °C; c without 4-

PPNO; dcatalyst reused once using the same reaction conditions as in a.

According to Table 10, the best catalyst was MCM1HC; however, upon its reuse no

catalytic activity was observed anymore. Mn determination by ICP-AES the fresh and

reused catalyst revealed an approximally 25% leaching of the metal. In addition the

deactivation of the catalyst was also due to the degradation of the salen ligand, which

was confirmed by the absence of the absorption band at 1540 cm-1 in the IR spectra

of the reused catalyst (see Figure 49). The deactivation by degradation under

prolonged oxidation conditions has also been observed using NaOCl as oxygen

source in the presence of cationic clays as carrier material [263].

88

Page 98: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 49. FT-IR spectra of MCM1HC before and after reaction with NaOCl as

oxygen source.

3.3.2 m-CPBA/NMO as oxidation system

This oxidation system is very popular for enantioselective epoxidations of prochiral

olefins at very low temperatures (-80 to 0°C). The results with this oxidant are shown

in Table 11. As in the previous oxidation system the homogeneous catalyst shows

the same activity at room temperature. In this case, only the conversion is a bit higher

when the reaction is conduced at 4 °C. Both the inorganic support and the free

manganese ions were catalytically inactive. Racemic epoxidation as with NaOCl was

obtained for the materials based on modified zeolites. Here, only the catalytic activity

of JCYSB3 is shown (for the other ones similar results were obtained).

In contrast with Al-MCM-41 as support, very similar activities as with the free

homogeneous catalyst was obtained. In this case high selectivities to the epoxide

were found. Again, MCM1HC showed the best results, but as in the previous case

this catalyst suffered deactivation due to leaching of the metal (55 %) and to salen

ligand degradation. Also, in this case, the presence of a nitrogen additive (N-methyl

89

Page 99: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

morpholine N-oxide, NMO) was necessary to promote the asymmetric induction. The

degradation of the salen ligand becomes obvious from Figure 50, where the absence

of the band at 1540 cm-1 in the reused catalyst suggests that the salen ligand

structure was changed. The reused and reloaded catalyst could recover the initial

activity, which indicates that the mesoporous structure is not destroyed. It confirms

that the activity loss of the reused material was mainly due to degradation of the

salen ligand and the loss of metal by leaching.

Table 11. Results of the catalytic activity for the catalyst heterogenized in different

supports and its homogeneous counterpart, using m-CPBA as oxygen donora

Catalyst Conversion(%)

Epoxide selectivity(%)

Cis/trans epoxide

eecis (%)

eetrans (%)

None 0 - - - - Homogeneous catalystb 17 99 6.5 97 90 Homogeneous catalyst 13 99 5.0 95 80 JCY 0 - - - - Mn-JCY 3 86 0.8 - - JCYSB3c 10 98 0.6 0 0 Al-MCM-41 0 - - - - Mn-Al-MCM-41 2 69 1.1 - - NH2-SI-MCM-41 1 0 - - - MCM1HC 14 95 4.5 88 66 MCM2HC 8 97 3.0 48 36 MCM3HC 9 95 3.0 21 15 MCM1HCd 10 95 2.0 0 0 MCM1HCe 2 73 1.0 0 0 MCM1HCf 12 95 3.7 85 65 aCis-ethyl cinnamate = 0.5 mmol; m-CPBA = 2.0 mmol; NMO = 3.0 mmol; catalyst =

0.05 mmol (30 mg homogeneous or 300-500 mg heterogeneous catalyst); CH2Cl2 = 4

ml; T = 23-25 °C; t = 2 h; bt = 4 °C. ctime = 24 h; dreaction without NMO; ecatalyst

reused once using the same reaction conditions as in a. fCalcined material and re-

loaded after first use.

90

Page 100: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 50. FT-IR spectra of the heterogeneous catalyst MCM1HC before and after

reaction using m-CPBA as oxygen source

3.3.3 In situ generated DMD as oxidation system

This source of the oxygen donor has been less studied than the previous oxidation

systems. In situ chiral dioxyranes obtained from chiral ketones and Oxone® (2

KHSO5●1KHSO4●1K2SO4) are very popular oxidation reagents for homogeneous

enantioselective epoxidations of prochiral olefins [265]. However, their application is

limited due to the degradation of the chiral ketone (catalyst) by the Baeyer-Villiger

reaction [110]. Achiral DMD obtained from acetone and Oxone® has proved to be a

convenient oxidizing agent for the homogeneous enantioselective epoxidation of

functionalized olefins, such as isoflavone derivatives [111], when using Jacobsen

catalyst. However, the use of in situ generated DMD in asymmetric epoxidation

reactions has not been reported yet.

In Table 12 the catalytic activity is presented. As in the previous oxidation systems,

the homogeneous catalyst gave very similar catalytic activities at room temperature

as at 4 °C. In this case, the enantioselectivities increase with decreasing reaction

91

Page 101: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

temperature. It has been found that at low reaction temperatures, the combination m-

CPBA/NMO preferably oxidizes the organomanganese complex instead of the

substrate [98]. However, as already explained low reaction temperatures resulted in

low conversions over the heterogenized catalyst.

In contrast to the previous oxidation systems, blank experiments indicate that both

the support and the metal catalyzed the racemic epoxidation of the substrate (Table

12). No asymmetric induction was observed with the heterogeneous catalysts based

on zeolites; only racemic epoxidation as in the abcense of a catalyst was observed.

This confirms that for this support type, the substrate and reactants do not reach the

active catalytic chiral species.

However, over the Al-MCM-41 heterogenized catalyst very similar enantioselectivities

as for the homogeneous catalyst where obtained. This suggests that the racemic

route did not influence the catalytic behaviour of the heterogenized chiral catalyst. As

in the previous case, similar conversions and high selectivities as for the free catalyst

were found. Here, 4-PPNO used as additive improved the catalytic activity only a bit.

During the reaction the pH control was very important to obtain A good asymmetric

induction. It is known that the formation of DMD takes place at a pH between 7.0-7.5

[112]. At pH < 7.0 DMD is not formed, therefore no epoxidation reaction was

observed. At pH > 7.0, the decomposition of the oxone® occurs preferably.

In this study the catalytic tests were stopped after the total addition of the oxygen

source as at this point the reaction rate reached its maximal value, and to avoid

degradation of the salen ligand. In addition, the pH was increasing, which suggests

that the oxygen source is mainly being decomposed [112]. This is further evidence

which confirms that the epoxidation reaction rate did change anymore. On the other

hand, a poor asymmetric induction was only observed, when other solvents such as

CH2Cl2 and CH3CN were used, which confirms that the formation of DMD was

responsible for good enantioselectivities.

92

Page 102: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Table 12. The catalytic activity for the catalyst heterogenized in different supports and

its homogeneous counterpart, using in situ generated DMD as oxygen donora

Catalyst Conversion(%)

Epoxide selectivity(%)

Cis/trans epoxide

eecis (%)

eetrans (%)

None 0 - - - - Homogeneous catalystb 40 97 4.6 86 72 Homogeneous catalyst 45 98 4.2 77 60 JCY 40 95 4.5 - - Mn-JCY 43 99 4.2 - - JCYSB3 45 98 4.5 0 0 Al-MCM-41 38 96 4.0 - - Mn-Al-MCM-41 40 95 3.5 - - NH2-Si-MCM-41 37 98 3.3 - - MCM1HC 32 95 3.8 75 50 MCM2HC 33 99 5.0 74 55 MCM3HC 30 96 3.3 40 32 MCM1HCc 30 96 2.8 65 38 MCM1HCd 30 95 1.4 5 0 MCM1HCe 35 95 4.1 68 45 aCis-ethyl cinnamate = 0.25 mmol; Oxone® (2 KHSO5●KHSO4●K2SO4) = 0.5 mmol;

catalyst = 0.06 mmol (30 mg homogeneous or 300-500 mg heterogeneous catalyst);

4-PPNO = 0.075 mmol; CH3COCH3 = 4 ml; NaHCO3 = 3.5 mmol; T = 23-25 °C; the

pH was kept between 7.0 and 7.5 with aqueous NaHCO3 (5 %w); bT = 20-25 °C; creaction without 4-PPNO; dcatalyst reused once using the same reaction conditions

as in a; eCalcined material and re-loaded after first use.

For the reusability test, the reused heterogeneous catalyst (MCM1HC) was first

washed with sufficient H2O to eliminate inorganic salts (KHSO4 and NaHCO3).

However, the recycling of the catalyst was not also possible due to leaching (20 %)

and degradation of the catalyst (see Figure 51).

93

Page 103: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 51. FT-IR spectra of the heterogeneous catalyst MCM1HC before and after

reaction with in situ generated DMD as oxygen source

The previous results indicate that the oxidative degradation of the salen ligand can

avoid the re-use of against leaching stable heterogeneous catalysts (immobilization

by covalent bonding). On the other hand, the conversions that were too low in spite

of the excess of the oxygen source. Under these conditions, the prolonged oxidation

conditions are favoured, therefore, the degradation of the salen ligand took place.

For this reason, we decided to study the enantioselective epoxidation of

unfunctionalized olefins (styrene and 1,2-dihydronaphthalene) which are more easy

to oxidize than functionalized olefins.

The catalytic activities are presented in Table 13. For the new substrates, racemic

epoxidation in absence of catalyst was also observed. The use of CH2Cl2 or CH3CN

instead of acetone gave only a poor asymmetric epoxidation (conversions up to 5%),

demonstrating that the in situ formation of DMD lead to the oxidation active oxo-

manganese(V) species.

94

Page 104: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Table 13. Results for the enantioselective epoxidation of three olefinsa.

Substrate Catalyst Conversionb

% Selectivityb

% (ee)

%d,i

Homogeneous 38 99 (5.0)c 86e (72)f MCM1HC 32 97 (5.0)c 75e (50)f MCM2HC 25 98 (4.7)c 78e (60)f

MCM3HC 30 95 (4.5)c 35e (30)f

Homogeneous 100 100 40g MCM1HC 100 100 38g MCM2HC 95 100 35g

MCM3HC 97 100 12g

Homogeneous 94 89 37h MCM1HC 95 88 32h

MCM2HC 96 90 34h MCM3HC 95 90 8h

C C

H H

C O

O

C2H5

Cis-ethylcinnamate

CH

Styrene

CH2

1,2-Dihydronaphthalene

aReaction condition: substrate = 0.25 mmol; Oxone® (2 KHSO5●KHSO4●K2SO4) =

0.25 mmol; catalyst = 30 mg homogeneous or 300-500 mg heterogoneous catalyst;

4-PPNO = 0.075 mmol; CH3COCH3= 4 ml; NaHCO3 = 3.5 mmol; T = 23-25 °C; the

pH was kept between 7.0 and 8.0 using aqueous NaHCO3 (5%w, pH = 8.0), while the

Oxone® slowly was added; t = 25-50 min. bDetermined by standard GC.ccis/trans

epoxide ratio dDetermined with chiral GC. eOptical configuration: 2R,3R (cis-epoxide). fOptical configuration: 2S,3R (trans-epoxide). gOptical configuration: R. hOptical

configuration: 1S, 2R. iThe optical configuration was assigned by comparison with

those reported in literature [102, 253].

The difference in the conversions and enantioselectivities for a given catalyst

depends on the nature of the olefin (Table 13). Thus, cis-ethyl cinnamate, an

unreactive olefin (electron poor substrate), yielded the highest enantioselectivities

compared to the more reactive substrates styrene and 1,2-dihydronaphthalene. In

comparison with the homogeneous catalytic activity, the heterogeneous catalysts

showed only a slight reduction in the conversion of cis-ethyl cinnamate, but for the

unfunctionalized olefins the similar conversions were obtained. In all cases, the high

selectivity exhibited by the homogeneous catalyst was found for the heterogeneous

systems, too. In addition, the cis/trans-epoxide ratio for the cis-ethylcinnamate

epoxidation was almost unchanged. All these results indicate that diffusion

95

Page 105: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

limitations, side reactions catalyzed by the support and immobilized complexes in the

external surface did not influence the catalytic behaviour of Jacobsen’s catalyst

immobilized into the mesopores of Al-MCM-41. With MCM1HC and MCM2HC the

best enantioselectivities were obtained, comparable to the homogeneous phase.

However, these materials showed moderate leaching (10% and more). This can be

explained by the very weak interaction between catalyst and the support. On the

other hand, MCM3HC proved to be very stable against leaching. In this case the

enantioselectivties decreased tremendously. The reduced mobility of the immobilized

complex in the MCM3HC sample might very well influence geometric conformation

needed for the chiral induction.

This could also be demonstrated with FT-IR spectroscopy on the once reused

catalyst MCM1HC, when in situ generated DMD and other common oxidants

(NaOCl/4-PPNO and m-CPBA/NMO) were used for the asymmetric epoxidation of

1,2-dihydronaphthalene. Figure 52 clearly shows that the chemical structure of the

immobilized catalyst did not suffer any change, when using in situ generated DMD as

oxidizing agent, whereas the use of the common oxidants, such as NaOCl and m-

CPBA, destroyed the catalyst after reaction. This can clearly be seen by the

presence or absence of the signal assigned to the metal salen complex (1540 cm-1).

It is important to note that the degradation of the salen ligand depends on the

substrate reactivity. Thus, using olefins which are easier to oxidize, the degradation

of the catalyst can be prevented by stopping the reaction prior to complete

conversion.

96

Page 106: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

Figure 52. FT-IR spectra of Jacobsen’s catalyst and MCM1HC used once in the

epoxidation of 1,2-dihydronaphthalene with different oxidants

Finally, the recyclability of catalyst MCM3HC was shown. Figures 53 and 54 illustrate

the results for the investigated substrates. Both the conversion and enantiomeric

excesses could be maintained up to three recycles. Using this catalyst, the ligand

salen deactivation could be suppressed. Thus, MCM3HC is stable against leaching

by using a covalent attachment on the salen ligand, and dimerization of the active

species is minimized by the presence of the additive 4-PPNO, and degradation of the

salen ligand is reduced by using in situ generated DMD as oxidizing agent.

97

Page 107: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

0

10

20

30

40

50

60

70

80

90

100

Cis-ethyl cinnamate Styrene 1,2-Dihydronaphthalene

Substrate

Con

vers

ion % Fresh catalyst

First cycleSecond cycleThird cycle

Figure 53. Effect of the MCM3HC reuse on the conversion

0

5

10

15

20

25

30

35

40

Cis-ethyl cinnamate Styrene 1,2-Dihydronaphthalene

Substrate

ee %

Fresh catalystFirst cycleSecond cycleThird cycle

Figure 54. Effect of the MCM3HC reuse on the enantioselectivity

98

Page 108: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

An interesting observation was made in the homogeneous phase. Jacobsen’s

catalyst could be recovered partially (80 %) by precipitation in the cis-ethylcinnamate

epoxidation. Encouraged by this result, the recyclability of the catalystwas

investigated. We could prove that the catalyst retained the high selectivity and

asymmetric induction as obtained by the fresh catalyst (see Table 14). However, a

conversion loss was observed, which can be associated with the incomplete

separation of the catalyst.

Table14. Result of catalytic activity and recyclability of the homogeneous catalyst

using in situ generated DMD as oxidanta

aReaction condition: Cis-ethyl cinnamate = 0.25 mmol; Oxone® (2

KHSO5●KHSO4●K2SO4) = 0.33 mmol; catalyst = 0.06 mmol 4-PPNO = 0.01255

mmol; CH3COCH3 = 4 ml; NaHCO3 = 3.5 mmol; T = 0-5 °C; the pH was kept between

7.5 and 8.0 using aqueous NaHCO3 (5% wt); bCalculated from the Mn content by

ICP-AES, before and after each test.

Non-immobilized catalyst

Conversion %

(recovered catalyst %)b

Epoxide Selectivity

%

Cis/Trans epoxide ratio

eecis %

eeTrans %

Fresh catalyst 41(80) 97 3.6 86 58 First recycle 33 (70) 95 3.2 84 53 Second recycle 21 (75) 96 3.4 86 54 Third recycle 16 (55) 96 4.2 82 56

A first control experiment using dichloromethane instead acetone afforded a lower

conversion (5 %). Additionally, the catalyst turned from dark brown to slight yellow,

suggesting that catalyst degradation occurred. These results indicate that Oxone® by

itself is a poor oxidant due to its low solubility in organic solvents and that ketone

compounds are the best parent materials for the regeneration of dioxyranes. These

observations are in agreement with various previous reports related with the

behaviour of dioxyranes [112, 276].

In contrast, the enantioselective epoxidation of non-functionalized olefins can be

catalyzed efficiently by a dimeric Jacobsen’s catalyst using Oxone® as oxidant [277].

Another control experiment was made to explain the partial precipitation of the

catalyst during the reaction. It could be shown that the catalyst could be recovered

99

Page 109: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

completely in acetone (0.25 mmol/4 ml, pH = 9.6) by the dropwise addition of an

aqueous solution of NaHCO3 up to pH = 10.4. However, it was necessary to use a

big volume of the buffer solution to keep the pH near the precipitation pH when

reaction conditions are applied. Thus, a very diluted liquid system was obtained,

which is very tedious to manipulate. Additionally, the increase of the reaction pH

mainly favours the Oxone® decomposition instead of the precipitation of the catalyst.

The last control experiment was made in the absence of a chiral catalyst. In this

case, racemic epoxidation took place with the same conversion and epoxide

selectivity as obtained in the presence of Jacobsen’s catalyst. Here, it is important to

note that the enantioselectivity observed in the presence of the chiral catalyst is due

to the formation of DMD, which preferentially oxidizes the catalyst to the active oxo-

manganese(V) species instead of the direct oxidation of the olefin. Figure 55 shows a

possible asymmetric route for the catalytic process described here.

100

Page 110: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Results and discussion

KHSO5

H3C C CH3

O

KHSO4

H3C C CH3

O O

N N

O OMn

Cl

N N

O OMn

Cl

O

C C

Ph C

O

O C2H5

H HC C

Ph C

O

O C2H5

H H

O

* ** *

* *

Figure 55. Proposed catalytic cycle for the enantioselective epoxidation of cis-

ethylcinnamate using Jacobsen’s catalyst and in situ generated DMD as oxidant.

101

Page 111: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Conclusions and outlook

4.0 Conclusions and outlook • The epoxidation of cis-ethyl cinnamate in the presence of Jacobsen’s catalyst,

according to different methods immobilized on various porous inorganic supports

was investigated.

• Mesopores completely surrounded by micropores were created inside the

crystallites of zeolite X and Y. This modified zeolites, Al-MCM-41 and Si-MCM-41

were used as supports for the immobilization of Jacobsen’s catalyst.

• Jacobsen´s catalyst was immobilized on modified zeolites using three different

procedures. The best result was obtained synthesizing the organometallic complex

under the same conditions as used for the homogeneous catalyst. However, no

asymmetric induction was observed in the catalytic test.

• Diffusion limitations of the substrate in the porous system of the modified zeolites

lower the reaction rate for the asymmetric epoxidation and prevent the appropriate

conformation of the catalyst for the chiral communication with the substrate. In

addition, the epoxide selectivity is strongly reduced with NaOCl as oxygen source.

In this case the support catalyzes the addition of chloride to the aromatic ring of

the substrate.

• Jacobsen´s catalyst was immobilized on Al-MCM-41, using ionic exchange

method. In this case the immobilization method clearly proved the retention of the

catalyst. The fixation of the catalyst can be explained by ionic interactions between

the Jacobsen’s cation and the anionic aluminium center of the support framework.

• Jacobsen’s catalyst was immobilized on NH2 group functionalized Si-MCM-41

(NH2-Si-MCM-41). Two methods were performed. In the first one, the catalyst is

retained by coordination bond between the manganese and nitrogen atoms. The

second one, salen ligand is attached to the solid surface by covalent bonding.

• In contrast to the heterogeneous catalysts based on modified zeolites, asymmetric

induction could be obtained when MCM-41 type mesoporous solids were used as

102

Page 112: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Conclusions and outlook

support. With m-CPBA, enantioselectivities very similar to the unimmobilized

catalyst were found, whereas when using NaOCl and in situ generated DMD a

slight reduction was found. On the other hand, NaOCl led to a tremendous

reduction in the epoxide selectivity, as was also observed in the presence of

modified zeolites.

• The methods of immobilization without covalent attachment of the salen ligand

(method 4 and method 5) showed moderate leaching, although the same

enantioselectivities as in the homogeneous phase were found.

• The stability towards leaching could be improved by covalent attachment of the

catalyst on the modified surface of NH2-Si-MCM-41 (method 6). However, the

enantioselectivities decreased tremendously, probably to the lesser mobility of the

catalyst, which produces a poor chiral communication with the substrate.

• The asymmetric epoxidation of various prochiral olefins using Jacobsen’s catalyst

and in situ generated DMD as oxidizing agent was investigated. Under

heterogeneous systems based on Al-MCM-41 and NH2-Si-MCM-41 as support,

the catalyst has proven to be very stable to the oxidative degradation of the salen

ligand.

• The use of in situ generated DMD as oxidant instead NaOCl or m-CPBA

minimizes the oxidative destruction of the salen ligand. Under these conditions,

the non-immobilized Jacobsen‘s catalyst could be separated partially and recycled

at least three times without enantioselectivity loss. The decrease in the conversion

can be anticipated due to the incomplete precipitation of the catalyst rather than to

the oxidative degradation of the salen ligand. Also, the physical loss of catalytic

material during the recovery process cannot be discarded totally.

• For future works, it is recommended to study the enantioselective epoxidation of

unfunctionalized olefins that have shown to give the best results in the

homogeneous phase, such as cis-β-methylstyrene and chromene derivatives.

103

Page 113: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Conclusions and outlook

• It is interesting to study the substitution of Oxone® by H2O2 as oxygen source. Shi

and co-workers reported the asymmetric epoxidation of various olefin types by in

situ generated dioxiranes from H2O2 and chiral ketones [269].

104

Page 114: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.0 Experimental

5.1 Chemicals

Chemicals were purchased from Aldrich, Fluka, Riedel-de-Haen and Merck in p.a.

purity and used without further purification.

5.2 Catalyst preparation

5.2.1 Jacobsen’s catalyst

Jacobsen´s catalyst was prepared according to Jacobsen et al. [54]. Typically, 0.112

mol (1R,2R)-(+)-1,2-diaminocyclohexane L-tartrate [54] and 0.225 mol K2CO3 were

dissolved in 150 ml water, after which 600 ml ethanol was added. The resulting

mixture was refluxed for 2 h and a solution of 0.225 mol 3,5-di-tert-butyl-2-

hydroxybenzaldehyde [54] dissolved in 250 ml ethanol was added over 30 minutes.

The funnel was rinsed with 50 ml ethanol, and refluxing was continued for 2 h.

Subsequently, 150 ml water was added, and the mixture was cooled to 5 °C for 3 h.

After filtering and washing with 100 ml ethanol, the solid was dissolved in 500 ml

CH2Cl2, washed with water (2*300 ml) and 100 ml saturated aqueous NaCl and dried

over Na2SO4. The material obtained was denoted Jacobsen´s ligand. Subsequently,

0.27 mol Mn(CH3CO2)2.4H2O was refluxed in 600 ml ethanol, after which a solution of

0.091 mol Jacobsen´s ligand dissolved in 250 ml toluene was added over 45 min.

The funnel was rinsed with 50 ml toluene and the mixture was refluxed for 2 h,

followed by bubbling air through the reaction mixture for 1 h (20 ml/min). After

addition of 100 ml of a saturated NaCl solution, the mixture was cooled to room

temperature and rinsed with 200 ml toluene. The organic phase was washed with

water (3*500ml), saturated aqueous NaCl (500 ml) and dried over Na2SO4. Solvent

removal in vacuo yielded a brown solid, which was denoted Jacobsen´s catalyst.

105

Page 115: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.2.2 Modification of zeolites Na-X and Na-Y [212]

Typically, 10 g of zeolite Na-X or Na-Y were dried in a deep bed at 723 K for 12 h

under a steady flow of dry nitrogen (3 l/h) in a horizontal tubular quartz reactor (3 cm

I.D.). Subsequently, the nitrogen was saturated with SiCl4 at room temperature, and

passed through the zeolite sample at 523 K for 1 h. The SiCl4-treatment was then

stopped and the reactor was purged with nitrogen to remove residual silicon

tetrachloride while the sample was allowed to cool down to room temperature. The

obtained material was ion exchanged three times with NH4+ salts (10 ml/g) at 353 K

for 24 h to replace sodium with ammonium. For the first ion-exchange a 10 M

NH4CO2CH3/ 2 M NH4Cl solution was used. The second and the third ion exchanges

were done using a 2 M NH4Cl solution. After washing and drying, the zeolite was

then treated in the same horizontal tubular quartz reactor with water vapor (nitrogen

flow saturated with water at 353 K) at 873 K (heating rate: 10 K/min) for 3 h. After

cooling to room temperature the zeolite sample was ion exchanged with a 2 M

(NH4)Cl solution and treated with water vapor at 1123 K (heating rate: 10 K/min) for 3

h. Some of the extra framework aluminium was removed by treatment with aqueous

hydrochloric acid at pH 2.7 for 72 h at room temperature. Subsequently, the zeolite

was washed with sufficient H2O and finally calcined at 823 K. The obtained materials

were denoted JCX and JCY, respectively.

5.2.3 Immobilization of Jacobsen’s catalyst into the generated mesopore of zeolites X

and Y

5.2.3.1 According to the method described by Hölderich et al. [213] (method 1)

JCX and JCY were ion exchanged with a solution of Mn(CH3CO2)2.4 H2O (0.1 M, 10

ml/g) at 353 K for 24 h. The ion-exchanged materials were dried under vacuum at

433 K and cooling to room temperature under an argon atmosphere. Subsequently,

the ligand was synthesized at room temperature under an inert atmosphere by the

addition of a with Mn stoichiometric amount of (1R,2R)-(+)-1,2-diaminocyclohexane

L-tartrate dissolved in CH2Cl2. After stirring for 24 h, the appropriate amount of

salysaldehyde (3,5-di-tert-butyl-2-hydroxybenzaldehyde) was added, after which

stirring was continued for another 24 h. The solvent was removed and the recovered

106

Page 116: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

solid was dried for 12 h at 413 K under vacuum. Finally, to remove surface adsorbed

species the obtained solid was soxhlet-extracted with CH2Cl2 (353 K, 24 h) and

toluene (433 k, 24 h), respectively. The extraction steps were repeated until FT-IR

analysis of the washing solution no longer showed bands attributable to the ligand or

the organometallic complex. The materials obtained were denoted JCXSB1 and

JCYSB1.

5.2.3.2 According to the method described by Corma et al. [134] (method 2)

After JCX and JCY were ion-exchanged with Mn, dried under vacuum at 433 K for

12 h and left to cool to room temperature under argon, salicylaldehyde dissolved in

CH2Cl2 was added. The thus obtained slurry was stirred and refluxed (353 K) for 12

h. Subsequently, the active amine was added using the same conditions as for the

salicylaldehyde. The inert atmosphere was replaced with air in order to oxidise Mn+2

to Mn+3. The obtained solid was treated similarly as described in the previous

method, and the materials were denoted JCXSB2 and JCYSB2.

5.2.3.3 According to the method described by Bein et al. [135] (method 3)

In contrast to the previous methods, the heterogenization process started with the

ligand synthesis. First, the active amine, dissolved in tetrahydrofuran, was added to

the support, after which the obtained slurry was refluxed at 353 K for 12 h under

argon. Next, salicylaldehyde was added under the same conditions as for the active

amine. Subsequently, the appropriate amount of the manganese source was added,

and the inert atmosphere was replaced by flowing air to the oxidize Mn+2 to Mn+3.

The obtained brownish slurry was refluxed at 80 K for 15 h, after which the

appropriate amount of LiCl (Cl- source as counter ion for the Jacobsen cation) was

added to the slurry, which was refluxed at 353 K for 24 h. Finally, the obtained solid

was treated similarly as in the previous methods; the materials were denoted

JCXSB3 and JCYSB3.

107

Page 117: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.2.4 Synthesis of Al-MCM-41 [270]

42 g tetraethyl ammonium hydroxyde (TEAOH, 40 %wt) and 0.84 g sodium aluminate

(NaAlO2) were dissolved in 122.4 g H2O and stirred for 1 h. Then, 40 g tetradecyl

trimethyl ammonium bromide(TDTMABr) were added; the resulting mixture was

stirred for 4 h. Subsequently, 60.92 g Ludox-HS were added dropwise over 1 h,

followed by vigorous stirring for 4 h. The gel had following composition:

1.0 NaAlO2 : 40 SiO2 : 11.6 TDTMABr : 28 TEAOH : 800 H2O

The gel was autoclaved in a teflon-lined static autoclave at autogeneous pressure at

110 °C for 7 days. During the crystallization the pH of the gel was maintained at 10.5

using aqueous acetic acid (10% wt). After filtering and washing the solid was calcined

first in flowing nitrogen (100 mL/min) and then in flowing air (100 ml/min). In both

cases the temperature program was:

1. Heat from room temperature to 413 K with 1 °C/min

2. Isothermal temperature at 413 K for 3 h

3. Heat from 413 K to 813 K with 1 °C/min

4. Isothermal temperature at 540 °C for 3 h

5. Cool down

5.2.5 Synthesis of Si-MCM-41 [226]

3.6 g, 25% solution ammonium hydroxide was diluted with 25.0 g water and added to

9.6 g TDTMABr while stirring. To this mixture, 4.08 g of tetraethyl ammonium

hydroxide (TMAOH) dissolved in 25 g water was added, followed by the addition of

27.2 g of tetramethyl ammonium silicate. The mixture was stirred for 20 min. Fumed

silica (6.2 g) was added slowly while stirring, which was continued for 1 h. The pH of

the final gel was maintained at 10.5 using aqueous acetic acid (10% wt). The molar

composition of the synthesis gel:

1.0 SiO2 : 0.0086 (NH4)2 : 0.089 (TDTMA)2 : 0.155 (TMA)2 : 40 H2O

108

Page 118: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

The gel was then transferred to a stainless steel autoclave and heated to 383 K for 5

days. The obtained solid was filtered, dried and calcined using the same temperature

program as Al-MCM-41.

5.2.6 Immobilization of Jacobsen’s catalyst in Al-MCM-41

5.2.6.1 Immobilization by ion exchange [256] (method 4)

A method similar to the one reported by Hutchings et al. [29] was used. In a typical

procedure 3 g of calcined and dry Al-MCM-41 was stirred in a manganese (II) acetate

tetrahydrate (Mn(CH3CO2)2.4H2O) aqueous solution (100 ml, 0.2 M) for 24 h at room

temperature. The material was then filtered, washed, dried and then stirred again in a

fresh Mn(CH3CO2)2.4H2O solution for a further 24 h. This process was repeated

twice. Subsequently the exchanged material was calcined at 823 K for 6 h. The

obtained material was denoted Mn-Al-MCM-41. Dry Mn-Al-MCM-41 was treated with

0.82 g (R,R)-(–)-N,N’-bis(3,5-di-tert-butylsalicylidene)cyclohexane-1,2-diamine

(Jacobsen’s salen) in 20 ml dry refluxing CH2Cl2 under argon atmosphere for 24 h.

After that, the inert atmosphere was replaced by an air flow and stirring was

continued for a further 6 h. The mixture was cooled to room temperature and then the

solvent was removed by reduced pressure at room temperature. The obtained solid

was Soxhlet extracted using CH2Cl2 at 343 K the washing solvent remained

colourless. The obtained material was denoted MCM1HC.

5.2.7 Immobilization of Jacobsen’s catalyst in Si-MCM-41

5.2.7.1 Immobilization by coordination bond between manganese and nitrogen atom

anchored previously on amino propyl-functionalized Si-MCM-41 (method 5)

A similar method reported by García et al. [274] was used. First, Si-MCM-41 is

functionalized using 3-aminopropyltriethoxysilane (3-APTES). In a typical procedure,

5 g of calcined and dry Si-MCM-41 were added to a solution containing 1.84 g of 3-

APTES dissolved in 15.33 g of dry toluene. The resulting suspension was stirred at

reflux temperature under argon for 24 h. The modified Si-MCM-41 was recovered by

filtration and Soxhlet extracted with CH2Cl2 for 24 h. After that, the obtained solid was

109

Page 119: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

dried at 318 K under vacuum prior to the next step. This process was performed

using (SiO2/APTES) = 1 and (toluene/SiO2) = 2 molar ratios, which produces a

nitrogen load equal to 2.0 mmol / g [274]. The obtained material was denoted NH2-Si-

MCM-41.

On the other hand, 5.30 (R,R)-(−)-N,N'-bis(3,5-di-tert-butylsalicylidene)-1,2-

cyclohexanediamine manganese(III) chloride (Jacobsen’s catalyst) was treated with

1.73 g of silver perchlorate (AgClO4) in 25 ml of dry CH3CN at room temperature for

16 h. After filtration through Celite and washing with dry CH3CN, the filtrate was

concentrated by reduced pressure. Dry CH2Cl2 and NH2-Si-MCM-41 were added and

the obtained suspension was stirred at room temperature under argon for 24 h. The

obtained solid was washed in a Soxhlet extraction apparatus using CH2Cl2 at 413 K

until the washing solvent retained colourless. The obtained material was denoted

MCM2HC

5.2.7.2 Immobilization by chemical attachment of the salen ligand on amino propyl-

functionalized Si-MCM-41 (method 6)

In this case, Jacobsen’s catalyst was prepared and modified during the

immobilization process using a method similar to that reported by Kim et al. [252]. In

a typical procedure, 3 g of the support were reacted with 0.30 g of 2,6-diformyl-4-tert-

butylphenol [278] and 0.38 g of (1R, 2R)-(+)-1,2-diaminocyclohexane in refluxing dry

ethanol for 14 h under argon. After cooling, the obtained solid was collected by

filtration and washed with CH2Cl2 and methanol, followed by drying under vacuum.

The obtained solid (NH2-MCM-41(1)) was treated with 0.34 g of 2,4-di-tert-butyl

salicylaldehyde in refluxing dry ethanol for 18 h under argon. The solid (NH2-MCM-

41(2)) recovered in the previous step was treated with 0.36 g of Mn(CH3CO2)2●4H2O

and 0.062 g of LiCl in ethanol at 343 K in the presence of air for 4 h. The obtained

solid was exposed to vacuum and washed in a Soxhlet extraction apparatus using

CH2Cl2 at 343 K until the washing solvent remained colourless. The obtained material

was denoted MCM3HC.

110

Page 120: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.2.8 Preparation of cis-ethyl cinnamate [103]

62 mmol ethyl phenylpropiolate was dissolved in 540 ml of a 7/2 hexane/hexene

mixture under dry argon. Subsequently, 86 mmol quinoline and 3.6 g Lindlar´s

catalyst (5% Pd) were added, after which 1 atm of hydrogen was added. After

completion of the reaction the mixture was filtered through a celite pad. The filtrate

was washed with a 10% acetic acid solution (4*500 ml), water (3*500 ml) and a

saturated NaHCO3 solution (4*500 ml) and dried over Na2SO4. The solvent was

removed under reduced pressure. The residue was determined to be a mixture of

92% cis-ethyl cinnamate (85 % yield), 5 % over-reduced ethyl phenylpropiolate and

3% trans-ethyl cinnamate by GC.

5.2.9 Preparation of the chiral epoxides used as references samples

To evaluate the enantiomeric excess of the chiral epoxides derived from cis-ethyl

cinnamate, a commercial sample of its racemic epoxide (3-ethyl phenylglycidate,

cis/trans = 10/90) was used. For the case of styrene and 1,2-dihydronaphthalene,

their chiral epoxides were prepared by reaction with m-CPBA in CH2Cl2 (substrate/

m-CPBA = 1/1, 4 ml CH2Cl2) at room temperature for 30 min. (R)-styrene oxide was

obtained using (R,R)-Jacobsen’ catalyst, (S)-styrene oxide using (S,S)-Jacobsen’

catalyst. (1S,2R)-1,2-dihydronaphthalene oxide was obtained using (R,R)-Jacobsen’

catalyst and (1R,2S)-1,2-dihydronaphthalene oxide using (S,S)-Jacobsen’ catalyst.

111

Page 121: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

Table 15. List and nomenclature of the materials

Material Description

Homogeneous catalyst

Jacobsen’s catalyst

(R,R)-(−)-N,N'-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediamine manganese(III) chloride [54]

Carrier material JCY Modified zeolite Y by dealumination processes according to the

method reported by Schuster et al. [212] (commercially available zeolite NaY as parent material)

JCX Modified zeolite X by dealumination processes according to the method reported by Schuster et al. [212] (commercially available zeolite NaX as parent material)

Al-MCM-41 Obtained material according to the method reported by Kresge et al. [270]

Si-MCM-41 Obtained material according to the method reported by Beck et al.

[226]

Immobilized homogeneous catalyst JCYSB1 Obtained material on JCY according to the method reported by

Hölderich et al. [213] (method 1) JCYSB2 Obtained material on JCY according to the method reported by

Corma et al. [134] (method 2) JCYSB3 Obtained material on JCY according to the method reported by

Bein et al. [135] (method 3) JCXSB1 Obtained material on JCX according to the method reported by

Hölderich et al. [213] (method 1) JCXSB2 Obtained material on JCX according to the method reported by

Corma et al. [134] (method 2) JCXSB3 Obtained material on JCX according to the method reported by

Bein et al. [135] (method 3) MCM1HC Obtained material in Al-MCM-41 according to the method

reported by Hutchings et al [256] (method 4)

Obtained material in Al-MCM-41 according to the method

reported by García et al [274] (method 5)

MCM3HC Obtained material in Al-MCM-41 according to the method

reported by Kim et al. [252] (method 6)

MCM2HC

112

Page 122: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.4 Characterization

5.4.1 XRD

X-ray powder diffractograms were collected on a Siemens D-5000, equipped with a

secondary monochromator, a variable diaphragm V 20 and a Nickel filter with CuKα

radiation.

5.4.2 Sorption of N2

Physisorption of N2 at 77 K was carried out on a volumetric apparatus (Asap 2010

Micromeritics) in a static mode. The carrier samples were evacuated prior to

measurement at 673 K and 10-2 Pa for 12 h, whereas the heterogenized catalysts

were treated at 373 K. The microporosity has been calculated according to the

method of Dubinin [272] and compared with the uptake at p/p0= 0.05. Mesoporosity

was computed according to the method of Barrett, Joyner and Halenda (BJH) [273]

and compared to the difference between the total uptake and microporous volume.

5.4.3 Elemental analysis

Elemental composition was determined by ICP-AES (Spectroflame D, Spectro). In a

typical procedure, 30 mg sample was dissolved in 500 µl HF (40 %), 1 ml H2SO4 and

49 ml deionized.

5.4.4 Thermogravimetric analysis

Measurements were performed on a Netzsch 209/2/E equipped with a STA 409

Controller. The heating rate was 2 K/min, using Al2O3-crucibles with α-Al2O3 as

reference material.

113

Page 123: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.4.5 UV-Vis

The spectra were collected on a Lambda 7 (Perkin Elmer). The samples were diluted

with BaSO4, which was also used as the standard. The recording speed was 60

nm/min, the date interval 1.0 nm and the slit width 4.0 nm.

5.4.6 FT-IR

FT-IR was done on a Nicolet Protégé 460 at room temperature using KBr techniques

for the homogeneous catalyst, whereas the heterogeneous catalysts were prepared

as self-supporting wafers.

5.5 Catalytic test reactions

5.5.1 NaOCl/4-PPNO as oxidation system

0.5 mmol of cis-ethyl cinnamate and 0.15 mmol 4-phenylpyridine N-oxide (4-PPNO)

were dissolved in 4.0 ml CH2Cl2. Afterwards, the catalyst (0.05 mmol soluble or 300

mg insoluble catalyst) was added. The solution was kept at the temperature indicated

in the Tables 9 and 10 . Afterwards, 2.7 mmol of NaOCl (4.7 ml, 0.55 M in Na2HPO4,

pH = 10.3) was added while stirring. The reaction was stopped after 12 h. The

heterogeneous catalyst was separated by filtration; the liquid was removed via

distillation under vacuum (378 K and 10-4 mbar) for the homogeneous reaction. In

both cases, the liquid mixture without catalyst was extracted with methyl-tert-butyl-

ether (MTBE) and the aqueous phase was discarded. The organic phase was dried

over sodium sulphate (Na2SO4) and analyzed by chiral GC on a γ-chiraldex column.

The recovered heterogeneous catalyst was washed with CH2Cl2 in a Soxhlet

extraction 343 K for 24 h and dried at room temperature prior to reuse. The

recovered un-immobilized catalyst was dissolved in CH2Cl2 and washed with H2O.

Subsequently, the solvent was removed under vacuum and the catalyst was re-

dissolved in CH2Cl2 (4 ml) prior to reuse.

114

Page 124: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.5.2 m-CPBA/NMO as oxidation system

0.5 mmol cis-ethyl cinnamate, 0.05 mmol soluble or 300 mg insoluble catalyst and 3

mmol 4-methylmorpholine N-oxide (NMO) were dissolved in 4.0 ml dry CH2Cl2. The

obtained mixture was stirred at the temperature as indicated in the Table 11, after

which 2.0 mmol m-chloroperbenbezoic acid (m-CPBA) was added in 4 portions for 4

minutes. After 2 h reaction time, the catalysts were isolated as in the previous

method. In this case, the liquid mixture containing one organic phase was treated the

same way as in the previous method and its content analysed by chiral GC on a γ-

chiraldex column. Also, the recovered catalysts were treated as in the previous

method before reuse.

5.5.3 in situ generated DMD as oxidation system

In a typical procedure, 4 ml acetone was added to a mixture of 0.25 mmol of the

olefin, 0.075 mmol 4-PPNO, 1.2 mmol NaHCO3 and 0.06 mmol soluble or 300 mg

insoluble catalyst. Next, a solution of 0.25 mmol Oxone (2KHSO5•1KHSO4•1K2SO4)

in 4 ml water was added slowly under continuous stirring and with pH control in the

range 7.0-8.0 using aqueous NaHCO3 (5 %wt, pH = 8.0) as buffer solution.

Generally, the addition of the oxygen source took place between 20 and 40 minutes.

Subsequently, the reaction was stopped and the catalyst was isolated by filtration.

Both the solid and the liquid mixture were retained for further use. An aliquot of the

liquid mixture was taken for Mn analysis by ICP-AES. The liquid mixture was

extracted with ethyl acetate (CH3CO2C2H5) and the aqueous phase was separated.

The content of the organic phase was analyzed with GC on a γ-chiraldex column.

The recovered catalyst was washed with sufficient H2O for 24 h to remove inorganic

salts resulting from Oxone® (KHSO4 and K2SO4) and the buffer solution (NaHCO3).

Afterwards, the catalyst was washed with CH2Cl2 in a Soxhlet extraction apparatus at

323 K for 24 h and dried at room temperature prior to reuse.

For the unimmobilized catalyst, the precipitated mixture which contains the catalyst

was recovered by filtration. Both the solid and liquid mixture were used for further

treatments. The liquid mixture was extracted with ethyl acetate (4 ml) and then the

organic layer was analyzed by GC on a γ-chiraldex column. The precipited mixture

115

Page 125: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

was washed with sufficient H2O (500 ml) to dissolve the inorganic salts. The brown

solid (recovered catalyst) was separated by filtration, dried at 373 K and then

dissolved in 4 ml of acetone prior to first reuse.

5.6 Calculation of catalytic activity and enantioselectivities

5.6.1 GC-Analysis

5.6.1.1 Partial hydrogenation of ethyl phenylpropiolate

Device: Siemen RGC 202

Column: FS-SE54, 60 m CS 25288-6

Temperature program: 318 K, 8 K/min, 543 K, 20 min isothermal

Sample volume: 0.2 µL

Carrier gas: He

Detector type: FID

Detector temperature: 573 K

Detector gas: H2/air (40/450 mL/min)

Injector temperature: 573 K

Pressure: 1 bar

5.6.1.2 Asymmetric epoxidation of cis-ethyl cinnamate

Device: HP6890 Series

Column: FS-Chirasil-Dex 25 m х 0.25 mm

Temperature program: 393 K isothermal

Sample volume: 0.2 -1.0 µ L

Carrier gas: Helium

Detector type: FID

Detector temperature: 573 K

Detector gas: H2/air (40/450 mL/min)

Injector temperature: 573 K

Pressure: 0.5 bar

Total flow: 17.1 mL/min

Split ratio: 1/30

116

Page 126: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.6.1.3 Asymmetric epoxidation of styrene

Device: HP6890 series

Column: Varian Chrompack Chirasil-Dex CB 25 mm х 0.25 mm

Temperature program: 383 K (10 min iso),-433 K (10 min iso)- 10 K/min

Sample volume: 1µL

Carrier gas: Helium

Detector type: FID

Detector temperature: 553 K

Detector gas: H2/air (40/450 mL/min)

Injector temperature: 523 K

Pressure: 1.5 bar

Total flow: 87.1 mL/min

Split ratio: 60/1

5.6.1.4 Asymmetric epoxidation of 1,2-dihydronaphthalene

Device: Siemens RGC 202

Column: Varian Chrompack Chirasil-Dex CB 25 mm х 0.25 mm

Temperature program: 383 K (10 min iso)-433 K (10 min iso)-463 K, 10 K/min

Sample volume: 0.1-0.3 µL

Carrier gas: Helium

Detector type: FID

Detector temperature: 523 K

Detector gas: H2/air (40/450 mL/min)

Injector temperature: 523 K

Pressure: 1.0 bar

Total flow: 72.9 mL/min

Split ratio: 60/1

117

Page 127: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

Experimental

5.6.2 Calculation of conversions, selectivities and enantiomeric excesses

Conversion, total selectivity to epoxides, cis/trans epoxide ratio and enantiomeric

excesses were calculated using the GC areas analysis in the following equations:

Conversion = ( )( ) %100×∑∑AsAp

Total selectivity to epoxides = ( )

%100×∑ApAepoxide

cis/trans epoxide ratio = ( )( )epoxideAtrans

epoxideAcis−

Enantiomeric excess for cis-epoxide (ee)cis = ( )( ) %100

,,

,, ×+

SSRR

SSRR

AAAA

Enantiomeric excess for trans-epoxide (ee)trans = ( )( ) %100

,,

,, ×+

RSSR

RSSR

AAAA

With

Ap : Product area

As : Substrate area

Aepoxides : Total epoxide area (cis + trans)

epoxideAcis − : Total cis-epoxide area (racemic mixture)

epoxideAtrans − : Total trans-epoxide area (racemic mixture)

JIA , : Enantiomerically pure epoxide area of optical configuration I,J

118

Page 128: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

6.0 References

1. Sheldon, R. A., Chirotechnology, Marcel Dekker, Inc., New York, (1993).

2. http://www.bioportfolio.com /reports/freedonia_chiral.htm.

3. Morrison, J. D (Eds)., Asymmetric Synthesis, Vol 5, Chiral Catalysis, Academic Press, (1985).

4. Knowles, W. S., Chem. Ind. 1996, 68, 141.

5. Blaser, H. U.; Spindler, F.; Studer., M in Hölderich, W. F., (Ed). Appl. Catal A: General 2001, 221, 119.

6. Lin, G-Q.; Li, Y-M.; Chan, A. S. C., Principles and Applications of Asymmetric Synthesis, Wiley/Interscience, New York, (2001).

7. Jacobsen, E. N.; Yamamoto, H.; Pfaltz (Eds). Comprehensive Asymmetric Catalysis, Springer, Berlin (1999).

8. Collins, A. N.; Sheldrake, G. N.; Crosby, J.; (Eds). Chirality in Industry, John Wiley, Chichester (1997).

9. Hölderich, W. F; Heitmann, G.; Catal. Today 1997, 38, 227.

10. Hölderich, W. F.; Appl. Catal. A: General 2001, 221, 1.

11. Bellocq, N.; Abramson, S.; Laspéras, M.; Brunel, D.; Moreau, P.; Tetrahedron: Asym. 1999, 10, 3229.

12. Baiker, A.; J. Mol. Catal. A Chem. 1997, 115, 473.

13. Hölderich, W. F.; Appl. Catal. A: General 1994, 113, 115.

14. Vankelecom, I. F. J.; Jacobs, P. A.; Chiral Catalyst Imobilization and Recycling., De Vos, D. E.; Vankelecom, I. F. J.; Jacobs, P. A (Eds) (2000).

15. Bergbreiter, D.; Organic Polymers as a Catalyst Recovery Vehicle., De Vos, D. E.; Vankelecom, I. F. J.; Jacobs, P. A (Eds) (2000).

16. Katsuki, T.; J. Mol. Catal A: Chem. 1996, 113, 87.

119

Page 129: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

17. Hölderich, W. F.; Röseler, J.; Heitmann, G.; Liebens, A.; Catal. Today. 1997, 37, 353.

18. Thomas, J. M.; Ramdas, S.; Millward, G. R.; Klinowski, J.; Audier, M.; Conzalec-Calbet, F.; Fyfe, C. A.; J. Solid. State Chem. 1982, 45, 368.

19. Schuster, C.; Möllmann, E.; Tompos, A.; Hölderich, W. F.; Catal. Lett. 2001, 74, 69.

20. Corma, A. Martinez, A.; Martinez-Soria, V.; Monton, J. B.; J. Catal. 1995, 153, 25.

21. Karandikar, P.; Dhanya, K. C.; Deshpande, S.; Chandwadkar, A. J. Sivasanker, S.; Catal. Commun. 2004, 5, 69.

22. Bein, T.; Curr. Opin. Solid State Mater. Sci. Catal. 1991, 59, 73.

23. Hodge, P.; Chem. Soc. Rev. 1997, 26, 417.

24. Wagner, H. H.; Hausmann, H.; Hölderich, W. F.; J. Catal. 2001, 203, 150.

25. Nicolau, K. C.; Dai, W-M.; Guy, R. K. Angew. Chem., Int. Ed. Engl. 1994, 33, 15.

26. Gawley, R. E.; Aubé.; J. Principles of Asymmetric Synthesis. Elsevier Science, Oxford (1996).

27. Eliel, E. L.; Wilen, S. H.; Mander, L. N.; Stereochemistry of Organic Compounds. John Wiley and Sons Inc., New York (1994).

28. Nogradi, M.; Stereochemistry. Basic Concepts and Applications, Pergamon, Oxford (1981).

29. Testa, B.; Priciples of Organic Stereochemistry, Marcel Dekker, New York (1979).

30. Boner, W. A.; Origins of Chiral Homogeneity in Nature, Topics in stereochemistry, Eliel, E.L.; Wilen, S.H. (Eds), Vol 18, Wiley, New York (1988).

31. Pasteur, L.; C. R. Acad. Sci. 1848, 26, 535.

32. Pasteur, L.; C. R. Hebd. Seances Acad. Sci. 1853, 37, 162.

33. Cushny, A. R.; Biological Relantionships of Optically Active Substance, Bailliere, Tindall and Cox, London (1926).

120

Page 130: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

34. Orgel, L. The Origins of Life. Molecules and Natural Selection, Chapman and Hall, London. (1973).

35. Winberg, H.; Topics in Stereochemistry. 1986, 16, 87.

36. Wani, M. C.; Taylor, H. L.; Wall, M. E.; J. Am. Chem. Soc. 1971, 93, 2325.

37. Brown, J. M.; Chem. Ind. (London). (1988), 612.

38. Bredig, G.; Biochem. Z. 1932, 249, 241.

39. Pirkle, W. H.; Pochapsky, T. C.; Advan. Chromatogr. 1983, 22, 71.

40. Fischer, E.; Ber. Dtsch. Chem. Ges. 1919, 524, 129.

41. Cahn, R. S.; Ingold, C. K. Prelog, V.; Experientia 1956, 12, 81.

42. Cahn, R. S.; Ingold, C. K. Prelog, V.; Angew. Chem. Int. Ed. Engl. 1966, 5,385.

43. Mislow, K.; Raban, M.; Topics in stereochemistry 1967, 1, 1.

44. Hirschmann, H.; Hanson, K. R.; J. Org. Chem. 1971, 36, 3293.

45. Pirkle, W. H.; Finn, J.; In: Asymmetric Synthesis, Morrison, J. D. (Ed), Academic Press, New York (1983).

46. Scott, J. S.; In: Asymmetric Synthesis, Morrison, J. D. (Ed), Academic Pres, Orlando, Vol 4. (1984).

47. Röper, H. In: Carbohydrates as Organic Raw Materials. Lichtenthaler, F. W. (Eds), Verlag Chemie, Weinheim. (1991).

48. Van Bekkum, H.; In: Carbohydrates as Organic Raw Materials, Lichtenthaler, F. W. (Eds), Verlag Chemie, Weinheim. (1991).

49. Sheldon, R. A.; Stud. Surf. Sci. Catal. 1991, 59, 33.

50. Blaser, H. U.; Pugin, B.; Spindler, F.; Applied Homogeneous Catalysis by Organometallic Complexes. Cornils, B. Herrmann, W. A (Eds), Verlag Chemie, Weinheim, (1996).

51. Vallery-Radot, R.; The life of Pasteur, Garden City Publishing Co. Inc., New York (1926).

52. Vries, T.; Wynberg, H.; Van Echten, E.; Koek, J.; Ten Hoeve, W.; Kellog, R. M.; Broxterman, Q. B.; Angew. Chem. 1998, 110, 2491.

121

Page 131: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

53. Sheldon, R. A.; Hulshof, L. A.; Bruggink, A.; Leusen, F. J. J.; Van der Haest, A. D.; Winberg, H. Chem. Today 1991, 9, 23.

54. Larrow, J. F.; Jacobsen, E. N.; J. Org. Chem. 1994, 59, 1939.

55. Kagan, H. B.; Fiaud, J. C.; In: Topics in Stereochemistry, Eliel, E. L.; Wilen, S. H (Eds), Vol 18, Wiley, New York (1988).

56. Mislow, K.; Introduction to Stereochemistry, Benjamin, New York (1965).

57. Rety, J.; Robinson, J. A.; Stereospecifity in Organic Chemistry and Enzimology, Verlag Chemie, Basel (1982).

58. Chibata, I.; Pure. Appl. Chem. 1978, 50, 667.

59. Martin, V. S.; Woodard, S. S.; Katsuki, T.; Yamada, Y.; Ikeda, M.; Sharpless, K. B.; J. Am. Chem. Soc. 1981, 103, 6237.

60. Blaser, H-U.; Chem. Commun. 2003, 293.

61. Bassindale, A.; The Third Dimension in Organic Chemistry, Wiley, New York, (1984).

62. Scott, J. W.; Topics in Stereochem. 1989, 19, 209.

63. Ohkuma, T.; Kitamura, M.; Noyori, R. In: Catalytic Asymmetric Synthesis, 2nd. Ojima, I.; Wiley, V. C. H (Eds), New York, (2000).

64. Davies, M.; Microporous Mesoporous Mater. 1998, 21, 173.

65. McMorn, P.; Hutching, G.; J. Chem. Soc. Rev. 2004, 33(2), 108.

66. Blaser, H. U.; Pugin, B.; Studer, M.; Enantioselective Heterogeneous Catalysis: Academic and Industrial Challenges In Chiral Catalyst Immobilization and Recycling, De Vos, D. E.; Vankelecom, I. F. J.; Jacobs, P. A (Eds) (2000).

67. Winberg, H.; Staring, E. G. J.; J. Am. Chem. Soc. 1982, 104, 166.

68. Blaser, H. U.; Müller, M.; Stud. Surf. Sci. Catal. 1991, 59, 73.

69. Baiker, A.;Blaser, H. U. Handbook of Heterogeneous Catalysis. Ertl, G.; Knötzinger, H.; Weitkamp, J (Eds) (1997).

122

Page 132: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

70. Hölderich, W. F.; Wagner, H.; Valkenberg, M. H.; Immobilized catalyst and their use in the synthesis of fine and intermediate chemicals. Royal Society of Chemistry, Supported Catalysts and Their Applications Proceedings of the 4th International Symposium on Supported Reagents and Catalysts in Chemistry held on 2-6 July 2000 at the University of St. Andrews, UK, p. 76-93.

71. Webb, G.; Wells, P. B.; Catal. Today 1992, 12, 319.

72. Ter Halle, R.; Schulz, E.; Spagnol, M.; Lemaire, M.; Tetrahedron Lett. 2000, 41, 7407.

73. Schürig, V.; Betschinger, F.; Chem. Rev. 1992, 92, 873.

74. Jhonson, R. A.; Sharpples, K. B.; Asymmetric Epoxidation of Allylic Alcohols in Catalytic Asymmetric Synthesis. Ojima, I (Ed). VCH: New York, (1993).

75. Katsuki, T.; Sharpless, K. B.; J. Am.Chem. Soc. 1980, 102, 5974.

76. Hanson, R. M.; Sharpless, K. B.; J. Org. Chem. 1986, 51, 1922.

77. Finn, M, G.; Sharpless, K. B.; In: Asymmetric Synthesis, Vol 5, Morrison, J. D (Ed), Academic Press, New York (1985).

78. Martin, V. S.; Woodward, S. S.; Katsuki, T.; Yamada, Y.; Ikeda, M.; Sharpless, K. B.; J. Am. Chem. Soc. 1981, 103, 6237.

79. a) Zhang, W.; Loebach, J. L.; Wilson, S. R.; Jacobsen, E. N.; J. Am. Chem. Soc. 1990, 112, 2801. b) Jacobsen, E. N.; Asymmetric Catalytic Epoxidation of unfunctionalized olefins in Catalytic Asymmetry Synthesis, Ojima, I (Ed), VCH, New York (1993).

80. a) Irie, R.; Noda, K.; Ito, Y.; Matsumoto, N.; Katsuki, T.; Tetrahedron Lett. 1990, 31 7345. b) Katsuki, T.; Coord. Chem. Rev. 1995, 140, 189.

81. Katsuki, T.; Synlett 2003, 281.

82. Dalton, C. T.; Ryan, K. M.; Wall, V. M.; Bousquet, C.; Gilheany, D. G.; Top. Catal. 1998, 5, 75.

83. Jacobsen, E. N.; Zhang, W.; Muci, A. R.; Ecker, J. R.; Deng, L.; J. Am. Chem. Soc. 1991, 113, 7063.

84. a) Irie, R.; Noda, K..; Ito, Y.;Matsumoto, N.; Katsuki, T.; Tetrahedron: Asym. 1991, 2, 481. b) Palucki, M.; Mc Cormick, G. J.; Jacobsen, E. N.; Tetrahedron Lett. 1995, 67, 2248.

85. Pietikäinen, P.; Tetrahedron 1998, 54, 4319.

123

Page 133: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

86. Yamada, T.; Imakawa, K..; Nakata, T.; Mukaiyama, T.; Bull. Chem. Soc. Jpn. 1994, 67, 2248.

87. Adam, W.; Jeko, J.; Levai, A.; Nemes, C:; Patonay, T.; Sebok, P.; Tetrahedron Lett. 1995, 35, 3669.

88. Pietikäinen, P.; Tetrahedron. 2000, 56, 417.

89. Srinivasan, K..; Michaud, P.; Kochi, J. K.; J. Am. Chem. Soc. 1985, 107, 7606.

90. Groves, J. T.; Myers, R. S.; J. Am. Chem. Soc. 1983, 105, 5791.

91. Groves, J. T.; Nemo, T. E.; J. Am. Chem. Soc. 1983, 105, 5786.

92. Guilmet, E. Meunier, B.; Tetrahedron Lett. 1982, 23, 2449.

93. Hughes, D, L.; Smith, G. B.; Liu, J.; Dezeny, G. C. Senanayake, C. H.; Larsen, R. D.; Verhoeven, T. R.; Reider, P. J.; J. Org. Chem. 1997, 62, 2222.

94. Katsuki, T. J. Mol. Cat. A: Chemical. 1996, 113, 87.

95. Nishinaga, A.; Yamato, H.; Abe, T.; Maruyama, K..; Marsura, T.; Tetrahedron Lett. 1988, 6309.

96. Brandt, P.; Norrby, P-O.; Daly, A. M.; Gilgeany, D. G.; Chem. Eur. J. 2002, 8, 4299.

97. Hamada, T.; Fukuda, T.; Imanishi, H.; Katsuki, T.; Tetrahedron 1996, 52, 515.

98. Jacobsen, E. N.; Palucki, M.; Hanson, P.; Tetrahedron Lett. 1992, 33, 7111.

99. Bonini, C.; Righi, R.; Tetrahedron 2002, 58, 5201.

100. Mikame, D.; Hamada, T.; Irie, R.; Katsuki, T.; Synlett 1995, 827.

101. Strassner, T.; Houk, K. N.; Org. Lett. 1999, 1, 2060.

102. a) Daily, A: M.; Renehan, M. F.; Gilheany, G.; Org. Lett. 2001, 3, 663. b) Takeda, T.; Irie, R.; Shinoda, Y.; Katsuki, T.; Synlett 1999, 1157.

103. Jacobsen, E. N. Deng, L.; Furukawa, Y.; Martínez, L. E.; Tetrahedron 1994, 50, 4323.

104. Deng, L.; Jacobsen, E. N.; J. Org. Chem. 1992, 57, 4323.

105. Jacobsen, E. N.; Zhang, W.; Muci, A. R.; Ecker, J. R.; Deng, L.; J. Am. Chem. Soc. 1991, 113, 7063.

124

Page 134: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

106. Hosoya, N.; Hatayama, A..; Irie, R.; Sasaki, H.; Katsuki, T.; Tetrahedron 1994, 50, 4311.

107. a) Wang, Z.-X.; Tu, Y.;Shi, Y.; J. Org. Chem. 1997, 62, 8622. b) Armstrong, A. Hayter, B. R.; Chem. Commun. 1998, 621.

108. Bortolini, O.; Fogagnolo, M.; Fantin, G.; Maietti, S.; Medici, A.; Tetrahedron: Asym. 2001, 12, 1113.

109. Wang, Z-X.; Tu, Y.; Frohn, M.; Zhang, J-R.; Shi, Y.; J. Am. Chem. Soc. 1997, 119, 11224.

110. Wang, Z-X.; Miller, S. M.; Anderson, O. P.; Shi, Y.; J. Org. Chem. 1999, 64, 6443.

111. a) Adam, W.; Jekö, J.; Levai, A.; Nemes, C.; Patonay, T.;Sebök, P.; Tetrahedron Lett. 1995, 36, 3669. b). Adam, W.; Fell, R. T.; Levai, A.; Patonay, T.; Peters, K.;Simon, A.; Tóth, G.; Tetrahedron: Asymm. 1998, 9, 1121.

112. Webb, K. S.; Ruszkay, S. J.; Tetrahedron 1998, 54, 401.

113. Kragl, U.; Dwars, T.; Trends in Biotechnology 2001, 19, 442.

114. Kureshy, R. I.; Khan, H. R.; Abdi, S. H. R. Iyer, P.; Patel, S. T.; Polyhedron 1999, 18, 1773.

115. Hölderich, W. F. Heterogeneous Catalysts for the Environmentally Benign Production of Fine and Intermediate Chemicals. Green Chemistry Series N°1, Collection of Lectures of the Summer School on Green Chemistry, Venice 1998-2004, part 3.

116. Pugin, B.; J. Mol. Catal. A: Chem. 1996, 107, 273.

117. Thomas, J. M.; Catlow, R. A.; Sankar, G.; Chem. Commun. 2002, 2921.

118. Pugin, B.; Müller, M.; Stud. Surf. Sci. Cat. 1993, 107.

119. Osawa, T.; Harada, T.; Tai, A.; Catal. Today 1997, 37, 465.

120. Török, B.; Felfödi, K.; Balazsik, K.; Bartok, M.; Chem. Commun. 1999, 1725.

121. Meunier, D.; Piechaczyk, A.; de Mallmann, A.; Basset, J-M.; Angew. Chem. 1999, 111, 3738.

122. Hölderich, W. F.; Wagner, H. Immobilization of Chiral Homogeneous Catalyst and their Use for Oxidation and Hydrogenation Reaction, in Catalysis by Unique Metal Structure in Solid Matrices II, Mathematics, Physics and Chemistry, Vol 13, (2001), 279.

125

Page 135: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

123. Shuttleworth, S. J.; Allin, S. M.; Sharma, P. K.; Synthesis 1997, 1217.

124. Tas, D.; Thoelen, C.; Vankelecom, I. F. J.; Jacobs, P. A.; Chem. Commun. 1997, 2323.

125. Brinker, C. J.; Scherer, G. W.; Sol-Gel Science, The Physics and Chemistry of Sol-gel Processing, Academic Press (1990).

126. Hodge, P.; Chem. Soc. Rev. 1997, 26, 417.

127. De, B. B.; Lohray, B. B.; Sivaram, S.; Dhal, P. K.; Tetrahedron: Asym. 1995, 40, 6827.

128. Nagel, U.; Leipold, J.; Chem. Ber. 1996, 129, 815.

129. Tas, D.; Jeanmart, D.; Parton, R. F.; Jacobs, P. A.; Stud. Surf. Sci. Cat. 1997, 493.

130. Schulz-Ekloff, G.; Meyer, G.; Wöhrle, D.; Mohl, M.; Zeolites 1984, 4, 30.

131. Jacobs, P. A.; Parton, R.; de Vos, D.; in: Zeolite microporous solids: synthesis, structure and reactivity, Derovane, E. G (Ed), (1992).

132. Hölderich, W. F.; Stud. Surf. Sci. Cat. 1991, 68, 257.

133. Csicsery, S. M.; Zeolite 1984, 4, 202.

134. Sabater, M. J.; Corma, A.; Domenech, A.; Fornes, V.; García, H.; Chem. Commun. 1997, 1285.

135. Ogunwami, S. B.; Bein, T.; Chem. Commun. 1997, 901.

136. Vankelecom, I. F, J.; Tas, D.; Parton, R. F.; Van de Viyer, V. S.; Jacobs, P. A.; Angew. Chem. Int. Ed. Eng. 1996, 35, 1346.

137. Selke, R.; Capka, M.; J. Mol. Catal. 1990, 63, 319.

138. Zhao, X. S.; Lu, G. Q.; Millar, G. J.; Ind. Eng. Chem. Res. 1996, 35, 2075.

139. a) Joó, F.; Katho, Á.; J. Mol. Catal. A: Chem. 1997, 116, 3. b) Felder, M.; Giffels, G.; Wandrey, C.; Tetrahedron: Asym. 1997, 8, 1975. c) Bergbreiter, D. E.; Liu, Y. S.; Tetrahedron Lett. 1997, 38, 3703.

140. Pozi, G.; Cinato, F.; Montanari, F.; Quici, S.; Chem. Commun. 1998, 877.

141. a) Kureshy, R. I.; Khan, N-U. H.; Abdi, S. H. R.; Ahmad, I.; Singh, S.; Jasra, R. V.; J. Mol. Catal. A: Chem. 2003, 203, 69. b) Kureshy, R. I.; Khan, N-U. H.; Abdi, S. H. R.; Ahmad, I.; Jasra, R. V.; Vyas, A. P.; J. Mol. Catal. A: Chem. 2004, 224, 229.

126

Page 136: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

142. Ward, J.; W. J. Catal. 1967, 9, 225.

143. Haag, W. O.; Lago, R. M.; Weisz, P. B.; Nature 1984, 309, 589.

144. Crønsted, A. F.; Akad. Handl. Stockholm. 1756, 18, 120. [Source: Flanigen, E. M, Introduction to Zeolite Science and Practice. Bekkum, H. V.; Flanigen, E. M.; Jacobs.; Jansen, J. C (Eds), 2nd edition, Elsevier Science, Amsterdam (2001).

145. Rabo, J. A.; Zeolite Chemistry and Catalysis, Washington, DC: American Chemical Chemical Society (1976).

146. Meier, M. W.; Molecular Sieves, Society for Chemical Industry, London, 10, (1968).

147. Barrer, M.; Hydrotermal Chemistry of Zeolites, Academic Press (1982).

148. Meier, W. M. Moeck, H. J.; J. Sol. State Chem. 1979, 349, 27.

149. Barrier, R. M.; Pure Appl. Chem. 1979, 51, 1091.

150. McBain, W. The sorption of gases and vapors by solids, Rutledge, London, Ch 5 (1932).

151. Barrer, R. M.; Kanellopoulos, A. G.; J. Chem. Soc., Ser. A. 1970, 765, 775.

152. Barrer, R. M.; J. Chem. Soc. 1948, 2158.

153. Breck, D. W.; Zeolite Molecular Sieves, John Wiley and Sons, Inc., New York, (1974).

154. Zhdanov, S. P.; ACS Advances in Chem. Ser. 1971, 101, 20.

155. Breck, D. W.; Eversole, G.; Milton, R. M.; J. Am. Chem. Soc. 1956, 78, 2338.

156. Dessau, R. M.; Kerr, G, T.; Zeolites 1984, 4, 315.

157. Beyer, H. K.; Balenykaja.; Catalysis by Zeolites, Elsevier (1980), 203.

158. Lohse, U.; Alsdorf, F. E.; Stach, H.; Z. Anorg. Allg. Chem. 1978, 477, 64.

159. Davis, M. E.; Ind. Eng. Chem. Res. 1991, 30, 1675.

160. Wilson, T.; Lok, B. M.; Flanigen, E. M.; U. S. Patent 4,310,440 (1982)

161. Weitkamp, J.; Solid State Ion. 2001, 131, 175.

127

Page 137: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

162. Hölderich, W. F.; New Horizons in Catalysis using Modified and Unmodified Pentasil Zeolites in Murakami, Y.; Jijima, A.; Ward, J. W (Eds). Proc 12th Int. Zeolite Conf. (1986) 827.

163. Hölderich, W. F.; Acid and Base Catalysts in the Synthesis of Intermediate and Bulk Product. Proceedings of the international symposium on Acid-Base Catalysis, Saporo, Japan, (1989), 1-20.

164. Marcilly, C. R.; Top. Catal. 2000, 13, 357.

165. Kärger, J.; Ruthven, D. M.; Diffusion in Zeolites and other Microporous Materials, John Wiley and Sons, New York (1992).

166. Bellusi, G.; Carati, A.; Clerici, M. G. Maddinchi, G.; Millini, R.; J. Catal. 1992, 133, 220.

167. Barrer, M.; Denny, P. J.; J. Chem. Soc. 1961, 971.

168. Venuto, P. B.; Landis, P. S.; Adv. Catal. 1968, 18, 259.

169. Lobo, R. F.; Jones, S. I.; Davis, M. E.; J. Inclusion Phenom. Mol. Recogn. 1995, 21, 47.

170. Øye, G.; Sjöblom, J.; Stöker, M.; Advances in Colloid and Interface Science 2001, 89-90, 439-466.

171. Anderson, J. R. Jackson, W. R, Hay, D.; Yang, C. P.; Campi, E. M.; Zeolites 1996, 16, 15.

172. Wilson, S. T.; Flanigen, E. M.; ACS Symp. Ser. 1989, 398, 329.

173. Merruche, A.; Patarin, J.; Kessler, H.;Soulard, M.; Delmotze, L.; Goth, J.; Jolly, J. F.; Zeolites 1992, 12, 226.

174. Freyhardt, C. C.; Tsapatsis, M.; Lobo, R. F.; Balkus, K. J.; Davis, M. E.; Nature 1996, 381, 295.

175. Nagy, J. B.; Bodart, P.; Hannus, I.; Kiricsi, I.; Characterization and Use of Zeolitic Microporous Materials, Deagen, Szeged, (1998).

176. Corma, A.; Chem. Rev. 1997, 97, 23736.

177. Ruthven, D. M.; Post, M. F. M.; in: Introduction to Zeolites Science and Practice Bekkum, H. V.; Flanigen, E. M.; Jacobs, P. A.; Jansen, J. C (Eds), 2nd edition, Elsevier Science, Amsterdam (2001).

178. Weisz, P. B.; Chemtech 1973, 3, 498.

128

Page 138: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

179. Chen, N. Y.; Kaeding, W. W:; Dwyer. J. Am. Chem. Soc. 1979, 101, 6783.

180. Rajagopalan, K.; Peters, A. W.; Edwards, G. C.; Appl. Catal. 1986, 23, 69.

181. Hölderich, W. F.; in: Acid-Base Catalysis. Tanabe, K.; Hattori, H.; Yamaguchi, T.; Tanaka, T (Eds), Tokyo (1988).

182. Wilson, S. T.; Flanigen, E. M.; ACS Symp. Ser. 1989, 398, 329.

183. Tundo, P.; Continuos Flow Methods in Organic Synthesis, Horwood, Y (Ed), New York (1991).

184. Hölderich, W. F.. Stud. Surf. Sci. Cat. 1991, 68, 257.

185. Corma, A.; Chem. Rev. 1997, 97, 2373.

186. a) Yaluris, G.; Rekoske, R. E.; Aparicio, L. M.; Madon, R. J.; Dumesic, J. A.; J. Catal. 1995 b) Carvill, B. T.; Lerner, B. A.; Adelmar, B. J.; Tomczak, D. C.; Sachtler, W. M. H.; J. Catal. 1993, 44, 1.

187. Tromp, M.; van Bokhoven, J. A.; Ostenbrink Garriga, M. T.; Bitter, J. H.; de Jong, K. P.; J Catal. 2000, 190, 209.

188. Lutz, W.; Löffler, E.; Zibrowius, B.; in: Zeolites: Refined Tool for Designing Catalytic Sites. Bonneviot, L.; Kaliaguine, S (Eds), Elsevier Science B. V. (1995).

189. Bartl, P.; Zibrowius, B; Hölderich, W. F.; Chem. Commun. 1996, 1611.

190. Kerr, G. T.; Chester, A. W.; Olson, D. H.; Catal. Lett. 1994, 25, 401.

191. Sato, K.; Nishimu7ra, Y.; Shimada, H.; Catal. Lett. 1999, 60, 83.

192. Lohse, U.; Mildebrath, M.; Z. Anorg. Allg. Chem. 1994, 25, 401.

193. Scherzer, J.; ACS Symposium Series. 1984, 248, 157.

194. Marcilly, C. R.; Petrole et Techniquews 1986, 328, 12.

195. Lynch, J.; Raatz, F.; Dufresne, P.; Zeolites 1987, 7, 333.

196. Cooper, D. A.; Hastings, T. W. Hertzenberg, E. P.; U. S. Patent N° 5601798 (1997).

197. Sasaki, Y.; Suszuki, T.; Takamura, Y.; Saji, A.; Saka, H.; J. Catal. 1998, 178, 94.

198. Dutartre, R.; Menorval, L. C. D.; di Renzo, F.; McQueen, D.; Fajula, F.; Schulz, P.; Micropor. Mater. 1996, 6, 311.

129

Page 139: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

199. Morin, S.; Ayrault, P.; Gnep, N, S.; Guisnet, M.; Appl. Catal. A General 1998, 166,

281.

200. Benazzi, E.; Lynch, J.; Gola, A.; Lacombe, S.; Marcilly, C. R.; Proc 12th Int. Zeolite Conf. (1999) 2735.

201. Olken, M. M.; Garces, J. M.; Proc 9th Int. Zeolite Conf. (1993) 559.

202. Giudici, R.; Kouwenhoven, H. W.; Prins, R.; Appl. Catal. A: General 2000, 203, 101.

203. van Santen, R. A.; van Leeuwen, P. W. N. M.; Moulijn, J. A.; Averill, B. A.; Stud Surf. Sci. Catal. 1999, 123, 543.

204. Hudec, P.; Novansky, J.; Silhar, S.; Trung, T. N:; Zubek, M.; Madar, J.; Ads. Sci Tech. 1986, 3, 159.

205. Groen, J. C.; Peffer, L. A. A.; Perez-Ramirez, J.; Micropor. Mesopor. Mater. 2002, 51, 75.

206. Cartlidge, S.; Nissen, H. -U.; Wessicken, R.; Zeolites 1989, 9, 346.

207. Ogura, M.; Shinomiya, S.; Tateno, J.; Nara, Y.; Nomura, M., Kikuchi, E.; Matsukata, M.; Appl. Catal. A Gen. 2001, 219, 33.

208. Marcilly, C. R.; Top. Catal. 2000, 13, 357.

209. Choi-Feng, C.; Hall, J. B.; Huggins, B. J.; Beyerlein, R. A.; J. Catal. 1993, 140, 395.

210. Nace, D. M.; Ind. Eng. Chem. Prod. Res. Dev. 1970, 9, 203.

211. Hölderich, W. F.; van Bekkum, H.; Stud. Surf. Sci. Catal. 1991, 58, 631.

212. Schuster, C.; Hölderich, W.F.; Catal. Lett. 1999, 58, 75.

213. Schuster, C.; Hölderich, W.F.; Catal. Today 2000, 60, 193.

214. Schuster, C.; Möllmann, E.; Tompos, A.; Hölderich, W. F.; Catal. Lett. 2001, 74, 69.

215. Hölderich, W. F.; Götz, N.; Zeolite Catalyzed Rearrangements of Epoxides and Glycidic Esters, In Proc 9th Int. Zeolite Conf. (1992) 309.

130

Page 140: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

216. Röhrlich, N.; Löffler, E.; Zibrowius, B.; Chassot, L. Hölderich, W. F.; J. Chem. Soc. Faraday Trans. 1998, 94(4), 609.

217. Kahlen, W. ; Wagner, H. H.; Hölderich, W. F.; 1998, 54, 85.

218. Schuster, C.; Stereoselektive Oxidation an Übergangsmetall-salen-Komplexen Immobilisiert in Nanoporösen Materialien, PhD thesis, RWTH- Aachen (2000).

219. Bhattacharjee, S.; Anderson, J. A.; Chem. Commun. 2004, 554.

220. IUPAC. Manual of Symbols and Terminology. Pure Appl. Chem. 1978, 314, 578.

221. Shimuzu, T.; Yanagisawa, T.; Kuroda, K.; Kato, C.; Annual Meeting of the Chemical Society of Japan Abstract N°1 XII D42, (1988)1761.

222. Iganaki, S.; Fukushima, Y.; Kuroda, K. J. Chem. Soc., Chem. Commun. 1993, 8, 680.

223. Iganaki, S.; Fukushima, Y.; Kuroda, K.; Stud. Surf. Sci. Catal. 1994, 84, 125.

224. Chen, C. Y. Xiao, S. Q.; Davis, M. E.; Micropor. Mater. 1995, 4, 1.

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

226. Beck, J. S.; Vartuli, J. C.; Roth, W. J.; J. Am. Chem. Soc. 1992, 114, 10834.

227. Ciesla, U.; Schüth, F.; Micropor. Mater. 1994, 6, 2317.

228. Vartuli, J. C.; Schmitt, K. D.; Kresge, C. T.; Chem. Mater. 1994, 6, 2317.

229. Lin, H. P.; Cheng, S.; Mou, C. Y.; Micropor. Mater. 1997, 10, 111.

230. Kruk, M.; Jaroniec, M.; Sayori, A.; Micropor. Mesopor. Mater. 1999, 27, 217.

231. Beck, J. S.; Vartuli, J. C.; Kennedy, G. J.; Kresge, C. T.; Roth, W. J.; Schramm, S. E.; Chem. Mater. 1999, 27, 217.

232. Alfredsson, V.; Keung, M.; Monnier, A.; Stucky, G. D.; Unger, K. K.; Schüth, F.; J. Chem. Soc., Chem. Commun. 1994, 8, 921.

233. Firouzi, A.; Atef, F.; Dertli, A G.; Stucky, G. D.; Chmelka, B. F.; J. Am. Chem Soc. 1997, 119, 3596.

131

Page 141: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

234. Feng, X.; Fryxell, G. E.; Wang, L. Q.; Kim, A. Y.; Liu, J.; Kemmer, K. M.; Science 1997, 276, 923.

235. Soai, K.; Watanabe, M.; Yamamoto, A.; J. Org. Chem. 1990, 55, 4832.

236. Hunger, M.; Schenk, U.; Breuninger, M.; Gläser, R. Weitkamp, J.; Micropor. Mesopor. Mater. 1999, 27, 261.

237. ∅ye, G.; Sjöblom, J.; Stöcker, M.; Micropor. Mesopor. Mater. 1999, 27, 171.

238. Hitz, S. Kogelbauer, A.; Lindlar, B.; Prins, R.; Stud. Surf. Sci. Catal. 1998, 117, 519.

239. Nießen, T. E. W.; Niedeerer, J. P. M.; Gjervan, T.; Hölderich, W. F.; Micropor. Mesopor. Mater. 1998, 21, 67.

240. Alfredsson, V.; Anderson, M. W.; Ohsuna, T.; Terasaki, O.,Jacob, M.; Bojrup, M.; Chem. Mater. 1997, 9, 2066.

241. Hölderich, W. F.; Mross, W. D.; Gallei, E.; Arab. J. Sci. Eng. 1985, 10, 407.

242. Feuston, B. P.; Higgins, J. B.; J. Phys. Chem. 1994, 98, 4459.

243. Sing, K. S. W.; Adv. Colloid Interface. Sci. 1998, 76/77, 3.

244. Schmidt, R.; Akporiaye, D.; Stöcker, M.; Ellestad, O. H.; Stud. Surf. Sci. Catal. 1994, 84, 61.

245. Schmidt, R.; Akporiaye, D.; Stöcker, M.; Ellestad, O. H. J. Chem. Soc. Chem. Commun. 1994, 12, 1493.

246. Corma, A.; Fornés, V.; Navarro, M. T.; Peréz-Parriente, J.; J. Catal. 1994, 148, 569.

247. ∅ye, G.; Axelrod, E.; Feldman, Y.; Sjöblom, J.; Stöcker, M.; Colloid. Poly. Sci. 2000, 278, 517.

248. Monnier, A. C.; Schüth, F.; Huo, Q.; Science 1993, 261, 1299.

249. Janssen, K. PhD Thesis. K. U. Leuven, Belgium (1999).

250. Bellocq, N.; Abramson, S.; Laspéras, M.; Brunel, D.; Moreau, P.; Tetrahedron: Asym. 1999, 10, 3229.

251. Choplin, A.; Quignard, F.; Coor. Chem. Rev. 1998, 178-180, 1679.

252. Kim, G.-J.; Shin, J.-H.; Tetrahedron Lett. 1999, 40, 6827.

132

Page 142: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

253. Bigi, F.; Moroni, L.; Maggi, R.; Sartori, G.; Chem. Commun. 2002, 716.

254. Pini, D.; Mandoli, A.; Orlandi, S.; Salvatori, P.; Tetrahedron: Asym. 1999, 10, 3883.

255. Frunza, L.; Kosslick, H.; Landmesser, H.; Höft, E.; Fricke, R.; J. Org. Catal A, 1997, 123, 179.

256. Piaggio, P.; Morn, P.; Murphy, D.; Bethell, D.; Pages, P: C.; Hancock, F. E.; Sly, C.; Kerton, O. J.; Hutchings, G. J.; J. Chem. Soc., Perkin. Trans. 2000, 2, 2008.

257. Li, C.; Xin, Q.; Zhang, Y.; Xiang, S.; Chem. Commun. 2002, 2696.

258. Zhou, X. -G.; Yu, X. -Q.; Huang, J. -S.; Li, S. -G.; Li, L. -S.; Che, C. -M.; Chem. Commun. 1999, 1789.

259. Möllmann, E.; Tomlinson, P.; Hölderich, W. F.; J. Mol. Cat A. 2003, 206, 253.

260. Fan, Q-H.; Li, Y-M.; Chan, A. S. C.; Chem. Rev. 2002, 102, 3385.

261. Jacobsen, E. N.; “Asymmetric Catalytic Epoxidation of Unfunctionalized Olefins,” in: Catalytic Asymmetric Synthesis, Ojima, I., Ed.; VCH: New York, 1993; Chapter 4.2.

262. Piaggio, P.; McMorn, P.; Langham, C.; Bethell, D.; Bulman-Page, P. C.; Hancock, F. E.; Hutchings, G. J.; New. J. Chem. 1998, 1167.

263. Fraile, J. M.; García, J. I.; Massam, J.; Mayoral, J. A.; J. Mol. Catal A. 1998, 136, 47.

264. Ferraz, H. M. C.; Muzzi, R. M.; Vieira, T.; Viertler, H.; Tetrahedron Lett. 2000, 41, 5021.

265. Frohn, M.; Shi, Y.; Synthesis 2000, 14, 1979.

266. Canali, L.; Sherrington, D. C.; Chem. Soc. Rev. 1999, 28, 85.

267. Sohn, J. R.; Decanio, S. J.; Zeolites 1986, 6, 265.

268. Möllmann, E.; Synthese von Immobilisierten Übergangsmetall-salen Komplexen, deren Charakterisierungn und Einsatz in der Enantioselektiven Epoxidierung von 1,2- Dihydronaphthalin. thesis, RWTH university Aachen (2002).

269. Shu, L.; Shi, Y.; Tetrahedron Lett. 1999, 40, 8721.

270. Kresge, C. T.; Leonowicz, M. E.; Roth, J. C.; Vartuli, J. C.; Vartuli, J. S.; Nature 1992, 359, 710.

133

Page 143: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

References

271. Kim, G.-J.; Kim, S-H.; Catal. Lett. 1999, 57, 139.

272. Dubinin, M, M.; J. Coll. Inter.Sci. 1967, 23, 487.

273. Barret, E. P.; Joyner, L. Y.; Halenda. P, P.; J. Chem. Soc., Chem. Comm. 1951, 73, 373.

274. Baleizão, C.; Gigante, B.; Sabater, M. J.; García, H.; Corma, A.; J. Appl. Catal A: General. 2002, 228, 279.

275. Karandikar, P.; Dhanya, K. C.; Deshpande, S.; Chandwadkar, A. J.; Sivasanker, S.; Agashe, M.; Catal. Commun. 2004, 5, 69.

276. Kachasaku, P.; Assabumrungrat, S.; Praserthdam, P.; Pancharoen, U.; Chemical Engineering Journal, 2003, 92, 131.

277. Kureshy, R. I.; Khan, N. H.; Abdi, S. H. R.; Ahmed, I.; Singh, S.; Vir-Jasra, R.; J. Mol. Catal A: Chem. 2003, 203, 69.

278. 2,6 diformyl-4-tert-butylphenol was synthesized according to: Chang, H-R.; Larsen, S. K.; Boyd, D. W.; Pierpont, C, G.; Hendrckson, D. N. J. Am. Chem. Soc. 1998, 110, 4565.

134

Page 144: Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte

LEBENSLAUF Persönliche Daten Name Jairo Antonio Cubillos Lobo Geburtsdatum 26.07.72 Geburtsort El Bagre (Kolumbien) Familienstand ledig Staatsangehörigkeit Kolumbianisch Schulausbildung 01/84 - 11/90 IDEM Caucasia 1° agrupacion, Caucasia, Kolumbien Hochschulausbildung 09/91 - 12/97 Studium des Chemieingenieurwesens an der Universidad de

Antioquia, Medellín, Kolumbien Abschluss: Chemieingenieur Thema: „Catalytic oxidation of methane over Pd/ZSM-5 at low

temperatures“ Note: 4,5 (von 5) entspricht „sehr gut“ 01/99 - 03/02 Studie der Umweltwissenschaften an der Universidad de

Antioquia, Medellín, Kolumbien Abschluss: Master in Environmental Engineering

Thema: „ Catalytic epoxidation of R-(+)-Limonene over Ti containing catalysts“

Note: Accepted entspricht „gut“ Praktische Erfahrungen

01/99 - 03/02 Lehrauftrag für chemische Verfahrenstechnik, am Institut für

chemische Verfahrenstechnik, Universidad de Antioquia, Medellín, Kolumbien

seit 03/02 Wissenschaftlicher Angestellter am Institut für Brennstoff-chemie RWTH Aachen

Sprachen Spanisch - Muttersprache Englisch - Fließend im Wort und Schrift Jairo Antonio Cubillos Lobo