15.02.2014 Views

School of Engineering and Science - Jacobs University

School of Engineering and Science - Jacobs University

School of Engineering and Science - Jacobs University

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Preparation <strong>and</strong> Characterization <strong>of</strong> Pt <strong>and</strong> Pd Nanoclusters Modified with<br />

Chiral Lig<strong>and</strong>s: Examination <strong>of</strong> Catalytic Activity in Hydrogenation <strong>of</strong><br />

Ethyl Pyruvate<br />

by<br />

Alex<strong>and</strong>er Kraynov<br />

A thesis submitted in partial fulfilment<br />

<strong>of</strong> the requirements for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy<br />

in Chemistry<br />

Approved, Thesis Committee<br />

Pr<strong>of</strong>. Dr. Ryan M. Richards<br />

Name <strong>and</strong> title <strong>of</strong> chair<br />

Pr<strong>of</strong>. Dr. Thomas Nugent<br />

Name <strong>and</strong> title <strong>of</strong> committee member<br />

Pr<strong>of</strong>. Dr. Stefan Tautz<br />

Name <strong>and</strong> title <strong>of</strong> committee member<br />

Dr. N. Waldoefner<br />

Name <strong>and</strong> title <strong>of</strong> committee member<br />

27 September 2006<br />

<strong>School</strong> <strong>of</strong> <strong>Engineering</strong> <strong>and</strong> <strong>Science</strong>


Table <strong>of</strong> Contents<br />

Acknowledgment<br />

Abstract<br />

IV<br />

V<br />

Chapter 1<br />

Introduction to the enantioselective catalysis 1<br />

1.1 General aspects: optical activity <strong>and</strong> chirality 2<br />

1.2 Chirality in our life 4<br />

1.3 Obtaining pure enantiomers 4<br />

1.3.1 Separation <strong>of</strong> enantiomers 4<br />

1.3.2 Asymmetric synthesis 7<br />

1.3.3 Asymmetric catalysis 7<br />

Chapter 2<br />

Introduction to nanoscale materials 18<br />

2.1 General information <strong>and</strong> definitions 18<br />

2.2 Methods <strong>of</strong> colloid <strong>and</strong> supported nanoparticles preparation 21<br />

2.3 Preparation <strong>of</strong> nanoclusters on a heterogeneous support 24<br />

Chapter 3<br />

Current thesis in the EU-COST project: aims <strong>of</strong> the work 25<br />

3.1 Introduction <strong>and</strong> goals setting 25<br />

Chapter 4<br />

Cinchonidine modified Pt colloidal nanoparticles: characterization <strong>and</strong><br />

catalytic properties 28<br />

4.1 Introduction 28<br />

4.2 Experimental 29<br />

4.3 Results <strong>and</strong> discussion 31<br />

4.3.1 General characterization 31<br />

4.3.2 FTIR investigation <strong>of</strong> cinchonidine adsorbed on Pt 35<br />

4.3.3 Investigation <strong>of</strong> catalytic behavior <strong>of</strong> cinchonidine<br />

modified Pt nanoclusters 43<br />

4.4 Summary 53<br />

Chapter 5<br />

Cinchonidine modified Pt colloidal nanoparticles immobilized on a<br />

heterogeneous support 55<br />

5.1 Introduction 55<br />

5.2 Experimental 56<br />

5.3 Results <strong>and</strong> discussion 57<br />

5.3.1 Role <strong>of</strong> catalyst activation in obtaining enantiomeric<br />

excess 57<br />

5.3.2 Comparison <strong>of</strong> catalysts <strong>and</strong> proposed model <strong>of</strong><br />

catalyst activation 58<br />

5.3.3 Catalyst stability <strong>and</strong> reuse tests 61<br />

5.3.4 IR investigation <strong>and</strong> proposed model <strong>of</strong> activation 63<br />

ii


5.4 Summary 66<br />

Chapter 6<br />

Cinchonidine modified Pd colloidal nanoparticles 67<br />

6.1 Introduction 67<br />

6.2 Experimental 68<br />

6.3 Results <strong>and</strong> discussion 69<br />

6.4 Summary 74<br />

Chapter 7<br />

Quiphos modified Pt <strong>and</strong> Pd colloidal nanoparticles 75<br />

7.1 Introduction 75<br />

7.2 Experimental 76<br />

7.3 Results <strong>and</strong> discussion 77<br />

7.4 Summary 85<br />

Chapter 8<br />

“Quiphos-spider” modified Pt colloidal nanoparticles 86<br />

8.1 Introduction 86<br />

8.2 Experimental 86<br />

8.3 Results <strong>and</strong> discussion 87<br />

8.4 Summary 89<br />

Chapter 9<br />

Diphos <strong>and</strong> diop modified Pt <strong>and</strong> Pd colloidal nanoparticles 90<br />

9.1 Introduction 90<br />

9.2 Experimental 90<br />

9.3 Results <strong>and</strong> discussion 91<br />

9.4 Summary 102<br />

Chapter 10<br />

Binap <strong>and</strong> synphos modified Pt colloidal nanoparticles 103<br />

10.1 Introduction 103<br />

10.2 Experimental 103<br />

10.3 Results <strong>and</strong> discussion 104<br />

10.4 Summary 111<br />

References 112<br />

iii


Acknowledgment<br />

I extend my sincere gratitude <strong>and</strong> appreciation to many people who made this doctoral<br />

thesis possible. Special thanks are due to my supervisor Pr<strong>of</strong>. Ryan Richards for his<br />

timely advices <strong>and</strong> cooperation.<br />

Thanks are also due to my former <strong>and</strong> current lab mates Mr. A. Suchopar for help with<br />

experiment organisation, Dr. K. Zhu <strong>and</strong> Dr. L. D’Souza <strong>and</strong> Dr. F. Pozgan, for many<br />

useful advices <strong>and</strong> directing, Dr. J. C. Hu <strong>and</strong> Mrs. L. F. Chen for general help <strong>and</strong><br />

cooperation in lab work, Mr. L. Zhi <strong>and</strong> Mrs. Helga van Eys-Schaefer with Mrs. Kirsten<br />

Töbe (Fraunh<strong>of</strong>er-Institut für Fertigungstechnik und Angew<strong>and</strong>te Materialforschung,<br />

Bremen) for very important <strong>and</strong> high quality TEM photographs.<br />

Author thanks Pr<strong>of</strong>. Gerard Buono <strong>and</strong> Dr. Didier Nuel <strong>of</strong> Université Aix-Marseille III,<br />

France for providing the quiphos <strong>and</strong> quiphos-spider lig<strong>and</strong>s <strong>and</strong> for useful discussion<br />

concerning the quiphos modifier, Dr. Virginie Ratovelomanana-Vidal (Laboratoire de<br />

Chimie Organique, E.N.S.C.P., UMR) for providing the synphos lig<strong>and</strong> <strong>and</strong> her<br />

colleague Dr. Véronique Michelet for useful discussion about synphos lig<strong>and</strong>.<br />

I would like to thank Pr<strong>of</strong>. U. Kortz (International <strong>University</strong> Bremen, Bremen) <strong>and</strong> Dr.<br />

E. Kirilina (Physikalisch-Technische Bundesanstalt, Berlin) for useful discussions<br />

about NMR spectra, Pr<strong>of</strong>. Thomas Nugent (International <strong>University</strong> Bremen, Bremen)<br />

<strong>and</strong> Pr<strong>of</strong>. S. Tautz (International <strong>University</strong> Bremen, Bremen) for useful discussions<br />

about organic chemistry <strong>and</strong> surface science, correspondingly, as well as for being my<br />

work supervisors, Dr. T. Mallat <strong>and</strong> Pr<strong>of</strong>. A. Baiker (Swiss Federal Institute <strong>of</strong><br />

Technology, Zurich) for useful discussions about cinchonidine on Pd <strong>and</strong> Pt systems,<br />

Dr. Achim Gelessus (International <strong>University</strong> Bremen, Bremen) for help with work<br />

with Gaussian 03 s<strong>of</strong>tware, Dr. L. G. Gordeeva (Boreskov Institute <strong>of</strong> Catalysis,<br />

Novosibirsk) for discussions about surface chemistry <strong>of</strong> alumina <strong>and</strong> silica materials,<br />

Dr. A. G. Okunev (Boreskov Institute <strong>of</strong> Catalysis, Novosibirsk) for discussion about<br />

kinetics <strong>of</strong> heterogeneous reactions, Dr. Serguei Soubatch (International <strong>University</strong><br />

Bremen, Bremen) for discussions about chemical physics <strong>of</strong> surfaces.<br />

I am highly indebted to Dr. Yu.I.Aristov (Boreskov Institute <strong>of</strong> Catalysis, Novosibirsk),<br />

Pr<strong>of</strong>. S. A. Dzuba <strong>and</strong> Mrs. R. I. Ratushkova (Chair <strong>of</strong> Chemical <strong>and</strong> Biological<br />

Physics, Novosibirsk State <strong>University</strong>, Novosibirsk) for introducing me to the world <strong>of</strong><br />

science <strong>and</strong> right directing <strong>and</strong> general support.<br />

My gratitude goes to the International <strong>University</strong> Bremen for providing me a financial<br />

support during my Ph.D. studies <strong>and</strong> making this work possible.<br />

Finally, my special gratitude goes to my wife Mrs. Liliya Kulishova <strong>and</strong> my family for<br />

their cooperation, support <strong>and</strong> encouragement.<br />

iv


Abstract<br />

Studies <strong>of</strong> the interactions <strong>of</strong> organic substrates with nanoscale metals <strong>and</strong> metal oxides<br />

are <strong>of</strong> fundamental importance for both fundamental scientific underst<strong>and</strong>ing <strong>and</strong> the<br />

development <strong>of</strong> commercial applications. Additionally, the development <strong>of</strong><br />

heterogeneous (solid/liquid or solid/gas) enantioselective catalysts is <strong>of</strong> great interest<br />

for the pharmaceutical industry. Here, the adsorption <strong>of</strong> chiral lig<strong>and</strong>s on metal surfaces<br />

to develop heterogeneous asymmetric catalysts is presented. Functionalizing metal<br />

nanoparticles with chiral lig<strong>and</strong>s facilitates increased surface coverage <strong>and</strong> the observed<br />

activity <strong>of</strong> these catalysts is significantly higher than their conventional phase<br />

counterparts.<br />

The first two chapters are an introduction to the enantioselective catalysis <strong>and</strong><br />

nanosized materials. As described in the chapter 3, the primary objectives <strong>of</strong> this work<br />

are<br />

1. to synthesize Pt <strong>and</strong> Pd nanoclusters modified by specific chiral lig<strong>and</strong>s by<br />

`bottom up' chemical synthesis,<br />

2. to determine the geometric orientation <strong>of</strong> adsorbed modifier with respect to<br />

the metal surface via DRIFTS (Diffuse Reflectance Infra-red Fourier<br />

Transform Spectroscopy) <strong>and</strong> molecular modeling,<br />

3. test enantioselective activity <strong>of</strong> the obtained samples in the hydrogenation <strong>of</strong><br />

ethyl pyruvate. The special effort has been given for investigation <strong>of</strong><br />

cinchonidine as modifier (chapter 4, 5, 6).<br />

Among the following chiral lig<strong>and</strong>s: quiphos (chapter 7), “quiphos-spider” (chapter 8),<br />

binap <strong>and</strong> synphos (chapter 10), only diop (chapter 9) was found to be able to induce<br />

low enantioselectivity, that was qualitatively explained through comparison with<br />

cinchonidine/Pt system.<br />

v


Chapter 1<br />

Introduction to the enantioselective<br />

catalysis<br />

1.1 General aspects: optical activity <strong>and</strong> chirality<br />

Any material that rotates the plane <strong>of</strong> polarized light is said to be optically active <strong>and</strong><br />

molecule <strong>of</strong> this materials is nonsuperimposable on its mirror image (Fig. 1-1). If a<br />

molecule is superimposable on its mirror image, the compound does not rotate the<br />

plane <strong>of</strong> polarized light; it is optically inactive. The property <strong>of</strong> nonsuperimposability <strong>of</strong><br />

an object on its mirror image is called chirality. If a molecule is not superimposable on<br />

its mirror image, it is chiral. Inverse logic is also true, i.e. if a molecule is<br />

superimposable on its mirror image, it is achiral. This is a necessary <strong>and</strong> sufficient<br />

requirement.<br />

Fig. 1-1. Superimposable (up) <strong>and</strong> nonsuperimposable (bottom) models <strong>of</strong> molecules.<br />

If a molecule is nonsuperimposable on its mirror image, the mirror image must be a<br />

different molecule, since superimposability is the same as identity. Both real molecule<br />

<strong>and</strong> its mirror image are called enantiomers. Enantiomers have identical (without<br />

considering interaction between elemental particles i.e. weak interaction [1]) physical<br />

<strong>and</strong> chemical properties except in two important respects:<br />

1. They rotate the plane <strong>of</strong> polarized light in opposite direction, though in equal<br />

amounts.<br />

2. They react at different rate with other chiral compounds. These rates may be so<br />

close together that the distinction is practically useless, or reaction rate <strong>of</strong> one<br />

enantiomer might be significantly far from another one. This is the reason that


many compounds are biologically active while their enantiomers are not active<br />

or lead to different reaction results.<br />

The mixture <strong>of</strong> enantiomers in equal amounts is called racemic mixture or racemate.<br />

The separation <strong>of</strong> a racemic mixture into its two enantiomers is called resolution. It is<br />

important to mention that the presence <strong>of</strong> optical activity always proves that a given<br />

compound is chiral, but its absence does not prove that the compound is achiral. A<br />

compound that is optically inactive may be achiral, or it may be a racemic mixture.<br />

Optically active compound may be classified into several categories.<br />

1. Compound with a chiral carbon atom. If there is only one carbon atom with<br />

four different substitutes (so called chiral carbon atom), the molecule must be<br />

optically active. The optical activity has been detected even for compounds,<br />

where one group is hydrogen <strong>and</strong> another deuterium i.e. 1-butanol-1-d ( Fig 1-2)<br />

[2].<br />

C 2 H 5<br />

H<br />

C<br />

D<br />

Fig. 1-2. 1-butanol-1-d [3].<br />

2. Compound with other quadrivalent chiral atoms. A molecule containing an<br />

atom that has four bonds pointing to the corner <strong>of</strong> a tetrahedron will be optically<br />

active if the four groups are different. Among atoms in this category are Si [4],<br />

Ge [5], Sn [5] <strong>and</strong> N (in quaternary salts or N-oxides).<br />

3. Compounds with tervalent chiral atoms. Atoms with pyramidal bonding might<br />

be expected to be optically active if the atom is connected to three different<br />

groups, since the unsaturated pair <strong>of</strong> electrons is analogous to a fourth group.<br />

For example, a secondary or tertiary amine (Fig. 1-3) where X, Y <strong>and</strong> Z are<br />

different. However due to the rapid oscillation <strong>of</strong> the lone pair from one side <strong>of</strong><br />

the XYZ plane to the other <strong>and</strong> following conversion <strong>of</strong> the molecule into its<br />

enantiomers it is very difficult to isolate <strong>and</strong> even detect an enantiomer. The<br />

inversion is less rapid in substituted ammonias [6-8] <strong>and</strong> phosphates [9-12] <strong>and</strong><br />

in suitable cases the inversion energetic barrier can be increased <strong>and</strong> such<br />

enantiomers could be detected <strong>and</strong> separated at room temperature [13, 14].<br />

OH<br />

Z<br />

N<br />

X Y<br />

Fig. 1-3. Chiral amine [3].<br />

4. Suitably substituted adamantine. Adamantines bearing four different substitutes<br />

at the bridgehead positions are chiral <strong>and</strong> optically active <strong>and</strong> could be resolved<br />

[15]. This type <strong>of</strong> molecule (Fig. 1-4) is a kind <strong>of</strong> exp<strong>and</strong>ed tetrahedron <strong>and</strong> has<br />

the same symmetry properties as any other tetrahedron.<br />

CH 3<br />

H<br />

COOH<br />

Br<br />

Fig. 1-4. Example <strong>of</strong> adamantane [3].<br />

2


5. Restricted rotation giving rise to perpendicular dissymmetric planes. Some<br />

compound what do not have asymmetric atoms are nevertheless chiral because<br />

their structure could be represented as two perpendicular planes neither <strong>of</strong><br />

which can be bisected by a plane <strong>of</strong> symmetry (Fig. 1-5). There are few<br />

examples for illustration below. Biphenyls containing four large groups can not<br />

freely rotate about the central bond because <strong>of</strong> steric hindrance (Fig. 1-6). In<br />

such compound the two rings are in perpendicular planes, that fits to the<br />

structure in Fig. 1-5.<br />

Fig. 1-5. Two perpendicular planes [3].<br />

There are few examples for illustration below.<br />

Cl<br />

O 2 N<br />

COOH<br />

HOOC<br />

NO 2<br />

Cl<br />

O 2 N<br />

COOH<br />

HOOC<br />

NO 2<br />

A<br />

P(Ph) 2<br />

P(Ph) 2<br />

(Ph) 2 P<br />

(Ph) 2 P<br />

Mirror<br />

Fig. 1-6. Chiral 3'-chloro-2',6'-dinitrobiphenyl-2,6-dicarboxylic acid (A) <strong>and</strong> binap (B)<br />

[3].<br />

6. Chirality due to a helical shape. Some molecules existing in helical shape with<br />

left- or right-h<strong>and</strong>ed orientation (Fig. 1-7). The entire molecule is usually less<br />

that one full turn <strong>of</strong> the helix, but this does not alter the possibility <strong>of</strong> left- <strong>and</strong><br />

right-h<strong>and</strong>edness.<br />

B<br />

3


Fig. 1-7. Hexahelicene [3].<br />

7. Chirality caused by restriction rotation <strong>of</strong> other types. In this case (e.g. Fig. 1-8.)<br />

chirality results because the benzene ring can not rotate in such a way that the<br />

carboxyl group goes through the alicyclic ring.<br />

HOOC<br />

(CH 2 ) 10<br />

Fig. 1-8. Example <strong>of</strong> chiral molecule with restricted rotation [3].<br />

1.2 Chirality in our life<br />

Chemical compound like drugs, fragrances, agrochemicals, etc. are needed in<br />

enantiopure form, because the all living organisms are built from chiral molecules, e.g.<br />

DNA. The enantiopure drugs, for instance, thus can work more efficiently than in<br />

racemate form because they influence on specific biochemical reactions by a lock-key<br />

mechanism. When one enantiomer is needed as an active compound, the other one is<br />

ballast, in the best case, or can induce a poison effect, like the (S)-thalidomide (where it<br />

was in racemic mixture with (R)- thalidomide), which caused birth defects in the<br />

beginning <strong>of</strong> 1960s. In present days drugs manufacturers need enantiopure chemical<br />

compound for producing enantiopure drugs. Thus the market related to the singleenantiomer<br />

drug business reached 40 % <strong>of</strong> total drug market in 2000. Whereas it<br />

covered 4 % only in 1985.<br />

1.3 Obtaining pure enantiomers<br />

Since huge number <strong>of</strong> natural compound are (e.g. cinchonidine etc.) chiral thus one <strong>of</strong><br />

the simplest way for obtaining chiral compounds is through extraction from nature i.e.<br />

from plants, living organisms etc. However it is clear that not all needed chiral<br />

compounds are available from the natural “chiral pool”. At the same time difficulties in<br />

obtaining chiral compound in sufficient amounts might lead to increases <strong>of</strong> products<br />

cost. Therefore methods for obtaining required enantiopure compounds have to be<br />

developed.<br />

1.3.1 Separation <strong>of</strong> enantiomers<br />

The resolution <strong>of</strong> the racemic mixture could be performed through removal or<br />

decomposition partially or completely <strong>of</strong> one enantiomer.<br />

Mechanical separation<br />

In 1848 L. Pasteur has shown an opportunity <strong>of</strong> mechanical separation <strong>of</strong> crystals <strong>of</strong> Na<br />

salt <strong>of</strong> racemic tartaric acid. In this example crystals <strong>of</strong> enantiomers form separately<br />

<strong>and</strong> visually look different, like mirror reflections <strong>of</strong> each other (Fig. 1-9), obtaining<br />

4


crystals are called enantiomorphic crystals. This method is known to as spontaneous<br />

resolution by crystallization.<br />

Fig. 1-9. Enantiomorphic crystals. Abstract model [16].<br />

Preferential crystallization by inoculation<br />

The selective crystallization <strong>of</strong> a single enantiomer is possible to perform in the<br />

presence <strong>of</strong> traces <strong>of</strong> the same enantiomer that is to be separated. In this case an<br />

oversaturated racemic solution is "seeded" in small amount <strong>of</strong> pure enantiomer <strong>and</strong> the<br />

resulting precipitate will have domination <strong>of</strong> the enantiomer in the seed. However, with<br />

the exception <strong>of</strong> the separation <strong>of</strong> glutamic acid <strong>and</strong> <strong>of</strong> a copper complex <strong>of</strong> DLaspartic<br />

acid, this method has been found to be impractical or resulted in portion<br />

resolution only.<br />

Biochemical processes<br />

The chiral compounds that react at different rates with the two enantiomers may be<br />

present in a living organism. For example, in 1858 Pasteur discovered that aqueous<br />

solution <strong>of</strong> racemic tartaric acid after contact with the mould fungus (Penicillium<br />

glaucum) slowly became enantioenriched. The mould was preferentially metabolizing<br />

the right rotatory enantiomer. This method is limited, since it is necessary to find the<br />

proper organism <strong>and</strong> since one <strong>of</strong> the enantiomer is destroyed in the process. However,<br />

when the proper organism is found, the method leads to a high extent <strong>of</strong> resolution<br />

since biological processes are usually very stereoselective. Another disadvantage <strong>of</strong> this<br />

method is that only dilute solutions can be used.<br />

Conversion in diastereomers<br />

If the racemic mixture to be resolved contains a carboxyl group (<strong>and</strong> no strongly basic<br />

group), it is possible to form a salt with an optically active base. Fig. 1-10 illustrates<br />

reaction <strong>of</strong> S-brucine with racemic hydroxypropanoic acid produced a mixture <strong>of</strong> two<br />

salts having the configuration SS <strong>and</strong> RS. Although the acids are enantiomers, the salts<br />

are diastereomers <strong>and</strong> have different properties e.g. solubility. The mixture <strong>of</strong><br />

diastereometric salts is allowed to crystallize from a suitable solvent.<br />

5


R<br />

H<br />

COOH<br />

C OH<br />

CH 3<br />

H<br />

COO - Bricine-H +<br />

C OH<br />

CH 3<br />

S-brucine<br />

R<br />

S<br />

S<br />

COO - Bricine-H +<br />

COOH<br />

HO C H<br />

HO C H<br />

CH 3<br />

CH 3<br />

S S<br />

Fig. 1-10 formation <strong>of</strong> diastereometric salts [3].<br />

Since the solubilities are different, the initial crystals formed will be richer in one<br />

diastereomer. Once the two diastereomers have been separated, it is easy to convert the<br />

salt back to the free acids <strong>and</strong> the recovered base can be used again. Advantages <strong>of</strong> this<br />

approach are that naturally occurring optically active bases (mostly alkaloids: brucine,<br />

ephedrine, morphine, etc.) are readily available. However, unfortunately, the difference<br />

in solubilities is rarely if ever great enough to effect total separation with one<br />

crystallization.<br />

Differential absorption<br />

When a racemic mixture is placed on a chromatographic column consisting <strong>of</strong> chiral<br />

phase, then, in principal the enantiomers should move along the column at different<br />

rates <strong>and</strong> should be separated. This technique is widely in use in different<br />

chromatography applications: GC, HPLS.<br />

Chiral recognition<br />

In order to separate enantiomers from racemic mixture the use <strong>of</strong> chiral host to form<br />

diastereometric inclusion is possible. Some hosts can form an inclusion with one<br />

enantiomer only or this formation is faster then with another enantiomer. For example,<br />

chiral host (Fig. 1-11-A) can partially resolve the racemic amine salt [17] (Fig. 1-11-B).<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

Ph<br />

CH 3<br />

C<br />

+<br />

NH 3<br />

-<br />

PF 6<br />

H<br />

A<br />

B<br />

Fig. 1-11. Chiral host-A <strong>and</strong> guest molecule –B [3].<br />

Kinetic resolution<br />

Since enantiomers react with chiral compounds at different rates, it is sometimes<br />

possible to effect a partial separation by stopping the reaction before completion. An<br />

example is the resolution <strong>of</strong> allylic alcohol [18] (Fig. 1-12-A) with one enantiomer <strong>of</strong> a<br />

chiral epoxidizing agent, in this case the discrimination was extreme. One enantiomer<br />

was converted to the epoxide <strong>and</strong> the other was not, the ratio being more than 100.<br />

Reactions catalyzed by enzymes can be utilized for this kind <strong>of</strong> resolution.<br />

6


Bu<br />

O<br />

Bu<br />

Bu<br />

OH<br />

OH<br />

OH<br />

racemic (R)-enantiomer (S)-enantiomer<br />

A B<br />

Fig. 1-12. Kinetic resolution <strong>of</strong> racemic allylic alcohol –A, yielding its (S) enantiomer<br />

(B) [3].<br />

1.3.2 Asymmetric synthesis<br />

There are two basic ways to synthesize a chiral compound. The first way is to begin<br />

with single stereoisomer <strong>and</strong> to use a synthesis that does not affect the chiral centre.<br />

The other basic method is called asymmetric synthesis. The definition <strong>of</strong> asymmetric<br />

synthesis is as follows: "Asymmetric synthesis is a reaction in which an achiral unit in<br />

an ensemble <strong>of</strong> substrate molecules is converted by a reactant into a chiral unit in such<br />

a manner that the stereoisomeric products are formed in unequal amounts" [19].<br />

1.3.3 Asymmetric catalysis<br />

As was mentioned above in order to synthesize a chiral compound another chiral<br />

substance is needed. Before such a chiral substance was considered as a reactant,<br />

however it can also be a chiral catalyst in asymmetrical synthesis. According to the<br />

definition catalyst influences on reaction but keeps its structure unchanged in the end <strong>of</strong><br />

reaction. Thus, a small amount <strong>of</strong> specific chiral catalyst can induce synthesis <strong>of</strong> chiral<br />

product in large scales. This strategy attracts much attention from industry <strong>and</strong> science,<br />

in fact, the number <strong>of</strong> publication on asymmetric syntheses has been increasing<br />

exponentially every year since the 80s [20].<br />

Willian S. Knowles, Ryoji Noyori <strong>and</strong> Barry Sharpless were awarded the Nobel Prize<br />

for chemistry in 2001 for their contribution in asymmetrical catalysis, sealing the<br />

importance <strong>of</strong> this field in contemporary chemistry <strong>and</strong> chemical technology.<br />

Asymmetrical reactions catalyzed with natural <strong>and</strong> man-made chiral catalysts becomes<br />

main-stream in different fields <strong>of</strong> industry (chemical, pharmaceutical, etc.).<br />

Asymmetrical catalysis can be divided into two main areas: homogeneous <strong>and</strong><br />

heterogeneous. It is clear from terms that in the case <strong>of</strong> homogeneous catalysis the<br />

important reactions are occurring in one phase, e.g. molecules in a solution. Whereas,<br />

in the case <strong>of</strong> heterogeneous catalysis reactions are occurring on the interface <strong>of</strong> two<br />

phases e.g. solid-liquid, solid-gas or in other words, on the surface <strong>of</strong> solid bodies.<br />

It is important to mention here briefly the important requirements [21] for<br />

enantioselective catalysts, since they are applied either for heterogeneous <strong>and</strong><br />

homogeneous catalysts.<br />

Enantioselectivity, expressed as enantiopurity <strong>of</strong> a system in terms <strong>of</strong> enantiomeric<br />

excess (ee, %). A catalyst has to be able to induce enantioselectivity in the range <strong>of</strong> 99<br />

% for further pharmaceutical applications. However ee’s > 80 % are acceptable also if<br />

further enantiopurification could be easily performed, e.g. via recrystallization.<br />

Catalyst productivity, characterized in terms <strong>of</strong> turnover number (TON). For example,<br />

in case <strong>of</strong> homogeneous enantioselective hydrogenation reactions the following values<br />

<strong>of</strong> TON can be used for the catalyst productivity classification. TON’s > 1000 for<br />

small-scale, high-value products <strong>and</strong> > 50000 for large-scale or less expensive products.<br />

7


Catalyst activity, expressed in the turnover frequency (TOF). For large-scale synthesis<br />

TOF > 10000 1/h are acceptable (under high pressure).<br />

Separation, should be able to perform by a simple low cost operation e.g. settling,<br />

filtration, distillation <strong>and</strong> at least 95 % <strong>of</strong> catalyst should be recovered.<br />

Stability, catalyst should be chemically <strong>and</strong> mechanically stable without loss <strong>of</strong> activity<br />

<strong>and</strong> enantioselectivity.<br />

Price <strong>of</strong> catalyst, consisting <strong>of</strong> price <strong>of</strong> expensive chiral lig<strong>and</strong> (homogeneous<br />

catalysis) <strong>and</strong> depending on type <strong>of</strong> catalyst in heterogeneous catalysis.<br />

Availability <strong>of</strong> the catalyst at the right time in an appropriate quantity.<br />

These criteria determine the choice <strong>of</strong> chiral catalyst for the specific reaction, as was<br />

illustrated in the synthesis <strong>of</strong> HPB (Ethyl (R)-2-Hydroxy-4-Phenylbutyrate) ester [22].<br />

In the following paragraphs the description <strong>of</strong> the most important aspects <strong>and</strong> ideas is<br />

summarized, however more detailed reviews can be found in the references that are<br />

given below.<br />

Homogeneous asymmetric catalysis<br />

The homogeneous enantioselective catalysis is widely used in industry covers the<br />

following asymmetric reaction [20]: hydrogenation, hydrosilylation (<strong>and</strong> related<br />

reaction), isomerisation <strong>of</strong> allylamines, carbometallation, oxidation (<strong>and</strong> related),<br />

carbonylation, polymerization <strong>and</strong> other that are not mentioned here.<br />

The origin <strong>of</strong> homogeneous asymmetric catalysis is in chiral organometallic complex<br />

catalyzing enantioselectivly specific types <strong>of</strong> reactions performed in one phase, e.g. in<br />

solutions. The electronic properties <strong>of</strong> the transition metals promote the formation <strong>of</strong><br />

organometallic complex (e.g. binap-Ru Fig. 1-13-A) with lower activation barrier for<br />

the reaction, whereas steric properties <strong>of</strong> used lig<strong>and</strong> create a chiral environment around<br />

the metal, leading to the specific reaction pathway <strong>and</strong> ending in enantioselectivity. The<br />

model <strong>of</strong> the reaction pathway for the C=O bond hydrogenation in the original work <strong>of</strong><br />

Noyori [23], here we only mention that formation <strong>of</strong> the transition state Fig 1-13-B<br />

which is responsible for enantioselectivity results in excellent, up to 98-99 % <strong>of</strong> ee [23].<br />

H<br />

Ar 2 R<br />

H 1 O<br />

P Cl 2<br />

CH H H<br />

N R 2<br />

H<br />

Ru<br />

N<br />

P Cl<br />

N R 3<br />

X(R 3 P) 2 Ru<br />

H 2<br />

Ar 2<br />

R 4<br />

H 2 N<br />

Ar=C 6 H 5 ; R 1 =R 4 =H; R 2 =R 3 =C 6 H 5<br />

X=H, OR, etc<br />

A<br />

B<br />

Fig. 1-13 Chiral binap-Ru complex (A) <strong>and</strong> based on it transition state (B) in<br />

mechanism <strong>of</strong> asymmetrical hydrogenation <strong>of</strong> C=O bond [23].<br />

Binap-Ru complex is not only one example <strong>of</strong> chiral homogeneous catalysis, in general,<br />

the suitable combination <strong>of</strong> a metal species <strong>and</strong> chiral lig<strong>and</strong> might be able to induce<br />

enantioselectivity. Here (Fig. 1-14) we give some examples <strong>of</strong> lig<strong>and</strong>s: quiphos [9-12],<br />

diop [24], BPPFA [25], synphos [26-35], PPCP [36], BPPM [37] which were<br />

successfully applied in corresponding asymmetrical reactions. Some <strong>of</strong> these lig<strong>and</strong>s<br />

were used as chiral modifiers <strong>of</strong> surface <strong>of</strong> Pt <strong>and</strong> Pd nanoparticles in the current thesis,<br />

please see corresponding chapters below.<br />

8


N<br />

N<br />

P<br />

Fe P(C 6H 5 ) 2<br />

O<br />

P<br />

P<br />

N<br />

O O<br />

P(C 6 H 5 ) 2<br />

Quiphos<br />

BPPFA: X= (CH 3 ) 2 N<br />

Diop<br />

O<br />

O PPh 2<br />

(C 6 H 5 ) 2 P<br />

P(C 6 H 5 ) 2<br />

O PPh 2<br />

P(C 6 H 5 ) 2<br />

P(C 6 H 5 ) N<br />

2<br />

O<br />

X<br />

BPPM: X=(CH 3 ) 3 COCO<br />

PPCP<br />

Synphos<br />

Fig. 1-14. Some chiral lig<strong>and</strong>s used in homogeneous asymmetrical catalysis.<br />

In spite <strong>of</strong> the fact that chiral homogeneous catalysts demonstrate excellent<br />

enantioselectivity, due to the separation criterion mentioned above, the disadvantage<br />

this type <strong>of</strong> catalysts appears. In fact, these soluble catalyst are more difficult to<br />

separate <strong>and</strong> h<strong>and</strong>le that the heterogeneous ones. One promising strategy to combine<br />

the best properties <strong>of</strong> the two catalyst types in the heterogenization or immobilization <strong>of</strong><br />

active metal complexes on supports or carriers, which may be separated by filtration or<br />

precipitation. However we leave this approach out <strong>of</strong> the current thesis, but keeping it<br />

in EU-COST project, please see below. Consequently, emphasis has to be given to the<br />

design <strong>of</strong> stable heterogeneous catalysts that are capable <strong>of</strong> high enantioselectivity.<br />

Heterogeneous asymmetric catalysis<br />

To date there have been numerous approaches to the design <strong>of</strong> heterogeneous<br />

asymmetric catalysts, since Schwab <strong>and</strong> coworkers first demonstrated that Cu <strong>and</strong> Ni<br />

could be supported on chiral silica surfaces [38, 39] <strong>and</strong> that the resulting catalysts<br />

could give low enantioselection in the dehydration <strong>of</strong> butan-2-ol.<br />

There are several strategies for the design <strong>of</strong> chiral heterogeneous catalysts, the three<br />

mentioned below, lead to stable catalysts that allow high enantioselectivity.<br />

• Modification <strong>of</strong> metal interfaces with chiral molecules;<br />

• Tethering <strong>of</strong> active homogeneous catalysts to a three-dimensional structure;<br />

• Electrostatic interaction as a means <strong>of</strong> immobilization.<br />

An overview <strong>of</strong> the about three strategies can be found in the work <strong>of</strong> Hutchings [40],<br />

whereas the current thesis focuses more on modification <strong>of</strong> the surface <strong>of</strong> the metal<br />

nanoclusters by chiral molecules.<br />

In contrast to homogeneous enantioselective catalysis the heterogeneous one is rather<br />

young <strong>and</strong> has a limited success in applications. Several factors contributed to this<br />

situation: a) the more complex structure <strong>of</strong> the heterogeneous catalyst surface on which<br />

coexist centers with different catalytic activity <strong>and</strong> selectivity, which can lead to<br />

undesired secondary reactions, b) an increased difficulty to create an effective<br />

asymmetric environment <strong>and</strong> to accommodate it with the multitude <strong>of</strong> reactions that are<br />

interesting to be carried out under enantioselective restrictions.<br />

X<br />

9


In fact, there are only two heterogeneous catalysts that reliably give high<br />

enantioselectivities (90 % e.e. or above). These are Raney nickel (or Ni/SiO 2 ) system<br />

modified with tartaric acid or alanine for hydrogenation <strong>of</strong> β-ketoesters [41] <strong>and</strong> β-<br />

diketones [41], methyl ketones [42, 43] <strong>and</strong> others summarized in the reviews <strong>of</strong> Blaser<br />

et al. [44] <strong>and</strong> platinum or platinum-on-support modified with cinchona alkaloids first<br />

reported by Orito [45-47] for the hydrogenation <strong>of</strong> α-ketoesters (Orito reaction), α-<br />

ketolactones [48-52], α-diketones [53-57], α-keto acetals [58, 59], α- α- α-<br />

trifluoroketones [60-63] <strong>and</strong> linear <strong>and</strong> cyclic α-ketoamides [64-66]. Cinchonidine<br />

modification <strong>of</strong> Pd surface was found to be successful in enantioselective<br />

hydrogenation <strong>of</strong> C=C bonds [67].<br />

Below we give basic aspects <strong>and</strong> models concerning behavior <strong>of</strong> cinchonidine based<br />

heterogeneous catalysts <strong>and</strong> origin <strong>of</strong> its enantioselectivity, since significant part <strong>of</strong> the<br />

current thesis is devoted to cinchonidine modified Pt <strong>and</strong> Pd systems, more detailed<br />

reviews can be found in work <strong>of</strong> Baiker [50, 68], Blazer [44], Murzin [69], Hutchings<br />

[40] <strong>and</strong> others [49].<br />

Conformations <strong>of</strong> cinchonidine molecule<br />

Several geometrical conformations were found for the cinchonidine molecule. The most<br />

important ones are in the relative orientation <strong>of</strong> quinoline <strong>and</strong> quinuclidine moieties<br />

(Fig. 1-15) via changing torsion angles τ 3’,4’,9,8 <strong>and</strong> τ 4’,9,8,N . The two types <strong>of</strong> major<br />

conformers are called closed (Fig. 1-16-A) <strong>and</strong> open (Fig. 1-16-B). The terms “open”<br />

<strong>and</strong> “closed” indicate whether the lone pair <strong>of</strong> quinuclidine nitrogen point away or<br />

towards the quinoline ring, respectively.<br />

Chiral center<br />

8<br />

HO<br />

9<br />

4'<br />

3'<br />

N<br />

N<br />

Quinuclidine<br />

Quinoline<br />

Fig. 1-15. Scheme <strong>of</strong> cinchonidine molecule<br />

It was shown by NMR spectroscopy [70], that cinchonidine molecule can exist in open<br />

<strong>and</strong> closed conformations at room temperature in different solvents. It has to be noted<br />

that there are other conformations (e.g. closed 1, 2, 4 they are not shown here) <strong>of</strong><br />

cinchonidine which belong to the open or closed types.<br />

10


Fig. 1-16. Closed –A <strong>and</strong> open –B types <strong>of</strong> cinchonidine conformers. Hydrogens are<br />

not shown for clarity.<br />

The relative population <strong>of</strong> these conformers depends on polarity <strong>of</strong> the solvent <strong>and</strong><br />

possible chemical interaction (cinchonidine-solvent), in fact as was show by<br />

calculations <strong>and</strong> NMR spectroscopy the open 3 conformer is most abundant in the<br />

solvent with low dielectric constant.<br />

Fig. 1-17. Dependence <strong>of</strong> relative abundance <strong>of</strong> conformer (Open 3) on solvent polarity<br />

determined by NMR spectroscopy <strong>and</strong> calculated using reaction field model [41].<br />

Solvents: (1) benzene, (2) toluene, (e) ethyl ether, (4) tetrahydroguran, (5) aceton, (6)<br />

dimethylformamide, (7) dimethyl sulfoxide, (8) water, (9) ethanol. The calculated<br />

model, does not take into account specific interaction such as the hydrogen boning<br />

between cinchonidine <strong>and</strong> ethanol. The figure is adapted from review <strong>of</strong> Baiker [50].<br />

It is important to note, that in the hydrogenation <strong>of</strong> ketopantolactone over cinchonidinemodified<br />

Pt ee decreases with increasing the dielectric constant <strong>of</strong> a solvent in the very<br />

11


similar way as relative abundance <strong>of</strong> the Open 3 conformer. Such relation is in<br />

agreement with suggestion that the Open 3 plays a crucial role in the mechanism <strong>of</strong> the<br />

enantioselectivity, see below.<br />

Fig. 1-18. Combined plot <strong>of</strong> the enantiomeric excess achieved in the hydrogenation <strong>of</strong><br />

ketopantolactone over cinchonidine-modified Pt [52, 71] (left axis) <strong>and</strong> the population<br />

<strong>of</strong> conformer Open(3) as calculated by density functional theory in combination with a<br />

reaction field model (P Open(3) , right axis) versus the dielectric constant <strong>of</strong> the solvent.<br />

The axis scale is arbitrarily chosen. The solvents are (1) cyclohexane, (2) hexane, (3)<br />

toluene, (4) diethyl ether, (5) tetrahydr<strong>of</strong>urane, (6) acetic acid, (7) ethanol, (8) water,<br />

(9) formamide. The picture is adapted from [70].<br />

Adsorption modes on Pt <strong>and</strong> Pd macro support<br />

Adsorption <strong>of</strong> quinoline (part <strong>of</strong> cinchonidine) <strong>and</strong> dihydrocinchonidine on the surface<br />

<strong>of</strong> Pt (single or polycrystal) was studied by XPS [72, 73], LEED [72], NEXAFS [74],<br />

H/D exchange experiments [75] in vacuum. It was found that adsorption <strong>of</strong> these<br />

compounds takes place preferentially in the flat mode through the π-aromatic system at<br />

room <strong>and</strong> lower temperatures, whereas even at 50 °C on average the quinoline ring was<br />

considerably tilted with respect to the surface [74]. Since experimental conditions there<br />

were far from used in the reactions, Ferri <strong>and</strong> Bürgi [76] performed ATR studies <strong>of</strong> the<br />

cinchonidine adsorption on solid-liquid interface, where 100 nm layer <strong>of</strong> Pt was<br />

deposited on alumina <strong>and</strong> cinchonidine in different concentration in dichloromethane<br />

was used. It was found that at low concentration (10 -6 M) a predominantly flat<br />

adsorption mode prevails (Fig. 1-19-1), at increasing coverage two different tilted<br />

species, α-H abstracted (Fig. 1-19-2) <strong>and</strong> N lone pair bonded (Fig. 1-19-3) cinchonidine<br />

were observed, at higher (10 -3 – 10 -4 M) concentrations all three species coexist on the<br />

Pt surface, this was also confirmed by the RAIRS experiments <strong>of</strong> Zaera <strong>and</strong> et al. [77].<br />

Adsorption <strong>of</strong> cinchonidine on Pt (111) was also studied by STM in vacuum <strong>and</strong> with<br />

presence <strong>of</strong> hydrogen, where mobility <strong>of</strong> cinchonidine was observed at high hydrogen<br />

pressure.<br />

12


Fig. 1-19. Three schematically favored adsorbed orientations <strong>of</strong> cinchonidine on wide<br />

Pt support [76]. Species 1: “π-bonded”, 2: “α-H abstracted”, 3: “N lone pair bonded”.<br />

Hydrogen atoms are omitted.<br />

It is interesting to note, that maximum observed ee values were found at low<br />

cinchonidine concentration, for example, in the works <strong>of</strong> LeBlond [78] <strong>and</strong> Bartok [51,<br />

79].<br />

The Fig. 1-20 illustrates dependence <strong>of</strong> ee on concentration <strong>of</strong> cinchonidine above the<br />

Pt/Al 2 O 3 catalyst.<br />

Fig. 1-20. Effect <strong>of</strong> cinchonidine concentration <strong>of</strong> the ee <strong>and</strong> the initial rate <strong>of</strong><br />

hydrogenation. Filled squares - ee , filled cicles - initial rate in ethyl pyruvate<br />

hydrogenation, open squares - ee <strong>and</strong> open cycles initial rate in bute-2,3-dione<br />

hydrogenation. Reaction conditions: dichlormethane (12.5 ml), reactant (66 mmol),<br />

catalyst 0.25 g, 50 bar H 2 , 20 C, 1200 rpm stirring. The picture is adapted from [80].<br />

The observed dependence <strong>of</strong> ee on cinchonidine concentration can be explained in the<br />

following way, in fact, at low concentration <strong>of</strong> modifier <strong>and</strong> small part <strong>of</strong> Pt surface is<br />

modified, whereas on unmodified sites reaction goes racemically (ee =0). When the<br />

13


surface concentration <strong>of</strong> cinchonidine becomes optimum ([cinchonidine]/[Pt sur ] ≈ 1/20<br />

[78, 81]) the maximum <strong>of</strong> ee is reached. Adsorption <strong>of</strong> cinchonidine in tilted modes<br />

becomes dominated at further increase <strong>of</strong> cinchonidine’s concentration in solution, that<br />

leads to decrease <strong>of</strong> relative concentration <strong>of</strong> π-bonded cinchonidine (active chiral site,<br />

see below) <strong>and</strong> thus decrease <strong>of</strong> ee.<br />

With respect to the adsorption on Pt, cinchonidine demonstrates differences <strong>and</strong><br />

similarities in adsorption on the Pd/Al 2 O 3 (as a thin film) surface [82]. In fact,<br />

cinchonidine was found in adsorption mode with the quinoline moiety nearly parallel to<br />

the Pd surface, likely through the π-system (flat adsorption mode) <strong>and</strong> in the mode with<br />

interaction with Pd through the lone pair <strong>of</strong> the quinoline N. Compared to Pt, the<br />

relative abundance <strong>of</strong> the two species under similar conditions was found to be different<br />

on the two surfaces, such that the π-bonded species are less stable on Pd than on Pt.<br />

Based on the comparison with adsorbed pyridine on Pd <strong>and</strong> Pt, the most obvious<br />

difference in the adsorption <strong>of</strong> cinchonidine on the two metals is that the : “α-H<br />

abstracted” species observed on Pt are absent on Pd. The difference in the diffuseness<br />

<strong>of</strong> the d orbital <strong>of</strong> the metal was proposed to be the origin <strong>of</strong> the different behavior with<br />

respect to the adsorption <strong>of</strong> cinchonidine.<br />

Origin <strong>of</strong> enantioselectivity <strong>and</strong> rate enhancement <strong>and</strong> two-cycle model<br />

Through the systematic variation <strong>of</strong> the cinchona alkaloid structure it was found that<br />

changing the absolute configuration at C-8 (S) <strong>and</strong> C-9 (R) <strong>of</strong> cinchonidine, i.e.,<br />

substituting cinchonidine by the diastereomer cinchonine, alters the chirality <strong>of</strong> the<br />

product from (R)- to (S)-lactate. The most important is the finding that the<br />

enantioselectivity is completely lost upon alkylation <strong>of</strong> the quinuclidine nitrogen atom<br />

which indicates that this center plays a crucial role in the mechanism <strong>of</strong><br />

enantioselection [83].<br />

According to the modern model, based on theoretical studies using quantum chemistry<br />

techniques, at both ab initio <strong>and</strong> semiemperical levels, <strong>and</strong> molecular mechanics [84-<br />

86] enantioselectivity is assumed to be thermodynamically controlled, i.e. by the<br />

difference in the energy <strong>of</strong> adsorption <strong>and</strong> complex formation ΔG 0 ad, <strong>of</strong> the pro-(R),<br />

[EP-CD] R ad <strong>and</strong> pro-(S), [EP-CD] S ad (Fig. 1-21). When adsorption equilibrium is<br />

established the relative surface coverage <strong>of</strong> [EP-CD] R ad <strong>and</strong> [EP-CD] S ad can be<br />

0<br />

ΔΔGad<br />

θ * −(<br />

)<br />

R<br />

RT<br />

expressed as = e . If no kinetic factor controls, i.e. if activation energies <strong>and</strong><br />

θ *<br />

S<br />

pre-exponential factors <strong>of</strong> the subsequent hydrogen addition are the same for pro-(R)<br />

[ R]<br />

[ θ * ]<br />

R<br />

<strong>and</strong> pro-(S) complexes, the ratio <strong>of</strong> surface coverage = . Thus, under these<br />

[ S]<br />

[ θ * ]<br />

S<br />

conditions the origin <strong>of</strong> enantioselectivity should be determined by stability <strong>of</strong> the<br />

intermediate pro-(R) <strong>and</strong> pro-(S) transition states. The question regarding whether the<br />

mechanism is kinetically or thermodynamically controlled, or whether the<br />

thermodynamic <strong>and</strong> kinetic factors are working in concert with each other is not yet<br />

defined.<br />

Enantioselectivity <strong>and</strong> rate enhancement (in factor <strong>of</strong> 10-100 compared with racemic<br />

EP hydrogenation) occur always together in this system. The H-bonding (in Pro-R <strong>and</strong><br />

Pro-S state, Fig. 1-21) would stabilize the half-hydrogenated state, increasing its life<br />

time <strong>and</strong>, hence, its concentration, thereby increasing the observed reaction rate [87].<br />

14


Fig. 1-21. The optimized structures <strong>of</strong> the proposed transition state formed on<br />

interaction <strong>of</strong> cinchonidine’s protonated nitrogen <strong>and</strong> the pyruvate ester [88].<br />

It has to be noted, that in the transition states the α-ketoesters are stabilized in their<br />

half-hydrogenated form [88] <strong>and</strong> cinchonidine (in the flat adsorption mode) is<br />

represented in open 3 conformation in these transition states as was suggested from the<br />

investigation <strong>of</strong> obtained ee with conformation analysis.<br />

Today it is practically accepted that enantioselective hydrogenation proceeds through<br />

the “two-cycle” mechanism as presented in Fig. 1-22. It is assumed, that the active<br />

chiral sites associated with adsorbed cinchonidine [CD] ad on Pt surface. Ethyl pyruvate<br />

from the liquid phase adsorb reversibly on these sites in its two enanti<strong>of</strong>acial<br />

configurations forming the diastereomeric intermediate complexes [EP-CD] R ad <strong>and</strong><br />

[EP-CD] S ad, which upon hydrogenation afford the (R)- <strong>and</strong> (S)- ethyl lactate,<br />

respectively. It has been suggested that the adsorbed cinchonidine interacts with the<br />

adsorbed ethyl pyruvate via hydrogen bonding between quinuclidine N <strong>and</strong> O atom <strong>of</strong><br />

the α-carbonyl moiety [73, 84]. Blaser et al. [89] suggested that for hydrogenation on<br />

unmodified sites, leading to the racemic products, addition <strong>of</strong> the first hydrogen is ratedetermining<br />

whereas for hydrogenation on modified sites, leading to the major<br />

enantiomer, the rate determining step is the addition <strong>of</strong> the second hydrogen.<br />

Fig. 1-22. Two-cycle mechanism suggested for enantiomeric hydrogenation <strong>of</strong> ethyl<br />

pyruvate [EP] over Pt modified by cinchonidine. The [CD] ad present an active chiral<br />

15


site on Pt surface formed by a adsorbed cinchonidine, [EP-CD] R ad <strong>and</strong> [EP-CD] S ad,<br />

represent half-hydrogenated complexes formed by interaction <strong>of</strong> chiral sites with EP,<br />

affording (R)- <strong>and</strong> (S)-ethyl lactate (EL). The picture is adapted from [50].<br />

It has to be noted, that the alternative models (e.g. the Chemical Shielding Model [90-<br />

93]) explaining the origin <strong>of</strong> enantioselectivity were developed, but it was found to be<br />

less successful with respect to the model mentioned above. The detailed comparison <strong>of</strong><br />

existing models with underlining their strong <strong>and</strong> weak point can be found in the work<br />

<strong>of</strong> Vargas [16].<br />

Several features with cinchonidine based systems<br />

Here we briefly mention some features <strong>and</strong> parameters <strong>of</strong> cinchona alkaloid based<br />

heterogeneous catalyst <strong>and</strong> give corresponding references for detailed examination,<br />

whereas some <strong>of</strong> the below mentioned features <strong>and</strong> reaction parameters (metal particle<br />

size, type <strong>of</strong> metal, pressure, conversion dependence on ee <strong>and</strong> catalyst reuse) <strong>and</strong> their<br />

role in enantioselectivity will be under investigation <strong>and</strong> discussion in the following<br />

paragraphs <strong>and</strong> chapters.<br />

Instead <strong>of</strong> Pt, Ir [94, 95] is also a suitable metal for enantioselective hydrogenation <strong>of</strong><br />

α-ketoesters. The Rh/Al 2 O 3 [94] gave 30 % lower ee than Pt/Al 2 O 3 in the hydrogenation<br />

<strong>of</strong> α-ketoesters. The Ru/Al 2 O 3 [96] afforded no enantioselectivity with cinchonidine in<br />

the hydrogenation <strong>of</strong> methyl pyruvate. The use <strong>of</strong> Pd will be discussed in a later<br />

chapter.<br />

The most convenient support material, like Al 2 O 3 , SiO 2 <strong>and</strong> carbon, have been reported<br />

to be suitable for ethyl pyruvate hydrogenation <strong>and</strong> have no particular effect <strong>of</strong><br />

enantioselection [97]. However an ee <strong>of</strong> 72 % was reported in the hydrogenation <strong>of</strong> E-<br />

α-phenylcinnamic support texture <strong>and</strong> a large effect on enantioselectivity [98, 99].<br />

The hydrogen pressure in the range <strong>of</strong> 1-100 bar was found to effect on reaction rate as<br />

the first order dependence, whereas ee increases with plateau [44].<br />

Hydrogenation <strong>of</strong> ethyl pyruvate started to decrease at temperature <strong>of</strong> 70 °C [89],<br />

whereas the drop in ee has been observed above 50 °C [100]. Moreover the adsorption<br />

mode <strong>of</strong> the quinoline ring is changed above 50-60 °C towards Pt (111) surface [89].<br />

Effect <strong>of</strong> concentration <strong>of</strong> ethyl pyruvate induces changes in the reaction order 1 → 0<br />

(saturation) or sometimes maximum [44, 89, 101].<br />

The increase <strong>of</strong> ee in the beginning <strong>of</strong> the reaction is the so called initial transient<br />

period. The nature <strong>of</strong> the effect is still not well understood [40], whereas its<br />

underst<strong>and</strong>ing might be useful in developing a catalyst which induces constant high<br />

enantiomeric excess during reaction, at low hydrogen pressure. This effect will be<br />

considered in details in the chapter 4.<br />

Additivies can effect on enantioselective hydrogenation in different ways, e.g., act as<br />

poison, interact with the products, change the chemical nature <strong>of</strong> modifier or interact<br />

with the modifier. All this effects are summarized in the review <strong>of</strong> Murzin [69].<br />

The very interesting work has been done by Baiker <strong>and</strong> co-workers through variation <strong>of</strong><br />

the modifier structure, briefly the new alternative to quinuclidine chiral moisture<br />

attached to the anchor part was designed [50, 102-104], the new obtained modifier<br />

results in high ee in the hydrogenation <strong>of</strong> α-ketoesters. At the same time it was show<br />

that the anthracene anchor <strong>of</strong> the new chiral lig<strong>and</strong> results in ~ 10 % higher ee with<br />

respect to naphthalene or quinoline anchors.<br />

Quasi-homogeneous catalysis<br />

16


The term quasi-homogeneous (it is also known as colloidal or soluble heterogeneous or<br />

soluble hybrid heterogeneous-homogeneous) was introduced first by Schmid [105]. It<br />

means that the catalyst forms the same (homogeneous) phase with the reaction mixture<br />

i.e. it is soluble in the used solvent however catalyst is not consisting <strong>of</strong> molecules or<br />

complexes like e.g. mentioned above binap-Ru catalyst is. The quasi-homogeneous<br />

catalysis consists <strong>of</strong> small particles (generally few nm in size) which are soluble since<br />

their surface is modified by a specific lig<strong>and</strong>.<br />

Colloidal catalysts are <strong>of</strong> interest for two reasons. Firstly the support effect is<br />

eliminated <strong>and</strong> secondly, it is possible to better control the morphology (size <strong>and</strong> shape)<br />

<strong>of</strong> the metal particles compared to classical supported catalysts. Among several studies,<br />

where Pt <strong>and</strong> other colloids were used for the cinchonidine-modified hydrogenation <strong>of</strong><br />

pyruvates, diketones [106-112] <strong>and</strong> trifluoroacetophenone [113] we would like to mark<br />

the works <strong>of</strong> Bönneman [114, 115] since in his work dihydrocinchonidine was used as<br />

chiral modifier <strong>and</strong> stabilizer, simultaneously, that has allowed to investigate this<br />

system <strong>and</strong> compare with conventional supported Pt catalysts. Others interesting works<br />

[106, 116, 117] where Pt colloidal catalyst stabilized by PVP (poly-N-vinyl-2-<br />

pyrrolidone) (about stabilization see the next chapter) demonstrated the maximum<br />

obtained 98 % ee in methyl pyruvate hydrogenation. Very recently Pd colloids<br />

modified with specific chiral lig<strong>and</strong>s demonstrated up to 98 % <strong>of</strong> ee in allylic alkylation<br />

reaction [118].<br />

Similar rate acceleration [109] <strong>and</strong> solvent effect [106] which are known for<br />

conventional Pt/Al 2 O 3 were observed for colloidal systems.<br />

The very interesting questions arise, wherever the size <strong>of</strong> Pt nanoparticles influence on<br />

ee or not <strong>and</strong> how adsorption <strong>of</strong> cinchonidine on small ~1-2 nm nanoparticles is<br />

different from adsorption <strong>of</strong> big macroscopic Pt support (plane). Thus, in the current<br />

thesis these questions will be under focus <strong>of</strong> our attention.<br />

17


Chapter 2<br />

Introduction to the nanoscale materials<br />

2.1 General information <strong>and</strong> definitions<br />

Introduction<br />

Colloidal metal (originally called "sols") particles are known at lease since work [119]<br />

<strong>of</strong> Hans Heinrich Helcher (1718) about gold colloids. In the Middle Ages glass-blower<br />

workers used gold <strong>and</strong> other metals in colloidal form in order to make coloured glass<br />

(Fig. 2-1) which were used e.g. as stained-glass window in churches or as service<br />

dishes. In 1857, Faraday reported the formation <strong>of</strong> deep red solutions <strong>of</strong> colloidal Au by<br />

reduction <strong>of</strong> an aqueous solution <strong>of</strong> AuCl 4 using phosphorus in CS 2 (a two-phase<br />

system) in a well-known work [120].<br />

Fig. 2-1. Red coloured glass based on gold colloids. Both pictures were found via<br />

google.com<br />

Now the field <strong>of</strong> nanomaterials is booming with scientific investigations <strong>and</strong><br />

application <strong>of</strong> the different areas <strong>of</strong> technology. In fact, for example the catalytic<br />

properties <strong>and</strong> the electronic structure <strong>of</strong> the nanomaterials can be considered from<br />

changing the cluster size, composition <strong>and</strong> structure [121, 122]. Herein, we will provide<br />

an overview <strong>of</strong> colloidal methods <strong>and</strong> properties as well as the approaches being used<br />

to bridge the fields <strong>of</strong> nanotechnology <strong>and</strong> catalysis. In our opinion, it is better to start<br />

acquaintance with nanomaterials from definitions developed by Schmid (Table 2-1)<br />

[123] as well st<strong>and</strong>ardized by IUPAC [124].<br />

Classification<br />

Nanoparticles are constituted <strong>of</strong> several tens or hundreds <strong>of</strong> atoms or molecules <strong>and</strong> can<br />

have a variety <strong>of</strong> sizes <strong>and</strong> morphologies (amorphous, crystalline, spherical, needles,<br />

etc.). According to the IUPAC definition the term colloidal refers to a state <strong>of</strong><br />

subdivision, implying that the molecules or polymolecular particles dispersed in a<br />

medium have at least in one direction a dimension, roughly between 1 nm <strong>and</strong> 1µm, or<br />

that in a system discontinuities are found at distances <strong>of</strong> that order [124].<br />

Table 2-1. Classification <strong>of</strong> nanoscale materials.<br />

18


Discrete, single Metal clusters, Traditional metal Bulk metal<br />

metal a complexes including<br />

nanoclusters<br />

colloids<br />

1-10 nm > 10 nm<br />

Comments: A - the classification can be used not only for metal materials.<br />

Another useful classification <strong>of</strong> variety form <strong>of</strong> colloids is illustrated in the Table 2-2,<br />

adapted form online encyclopaedia (Wikipedia) [125].<br />

Table 2-2. Examples <strong>of</strong> variety <strong>of</strong> colloids.<br />

Dispersed medium<br />

Gas Liquid Solid<br />

Continu Gas None (all gases are Liquid Aerosol Solid Aerosol<br />

ous<br />

soluble)<br />

(fog, mist) (smoke, dust)<br />

media Liquid Foam (cream) Emulsion (milk, Sol (paint, ink)<br />

blood)<br />

Solid Solid foam (aerogel, Gel (gelatine, Solid sol (ruby glass)<br />

pumice)<br />

cheese)<br />

Full shell clusters model<br />

According to the Schmid [123], metal clusters which have a complete, regular outer<br />

geometry are designated as full-shell, or “magic number”, clusters (Table 2-3).<br />

It is believed, that full-shell clusters have added stability, as their densely-packed<br />

structures provide the maximum number <strong>of</strong> metal-metal bonds [126, 127].<br />

Full-shell clusters are constructed by successively packing layers – or shells- <strong>of</strong> metal<br />

atoms around a single metal atom. The total number <strong>of</strong> metal atoms, y, per n th shell is<br />

given by the equation y 2 n<br />

= A ⋅ n + 2 (n>0). In this formula parameter A depends on<br />

geometry <strong>of</strong> the current crystal, the values <strong>of</strong> parameter A are given in the Table 2-4 for<br />

several types <strong>of</strong> crystals.<br />

Table 2-3. Full shell model on the example <strong>of</strong> cuboctahedrone crystal. The table is<br />

adapted from [128].<br />

Table 2-4. Full shell model applied for different polyhedrons [129, 130].<br />

Polyhedron<br />

Illustrated A values<br />

index a<br />

Truncated tetrahedron 1 14<br />

Cuboctahedron or twinned cuboctahedron 2 10<br />

Truncated octahedron 3 30<br />

Truncated cube 4 46<br />

19


Triangular prism 5 4<br />

Comments: A – The geometry <strong>of</strong> the corresponding polyhedron can be found in the Fig.<br />

2-3.<br />

Fig. 2-3. Geometrical models <strong>of</strong> some polyhedrons. Numeration corresponds to<br />

illustration index from Table 2-4.<br />

The full shell model is very useful in calculation <strong>of</strong> concentration <strong>of</strong> surface atoms for a<br />

nanocluster with known size, that can be considered as number <strong>of</strong> active catalytic sites<br />

e.g. for estimation <strong>of</strong> initial rate values normalized on number <strong>of</strong> active sites.<br />

Stabilization <strong>of</strong> nanoparticles<br />

Colloidal particles generally are not thermodynamically stable, because <strong>of</strong> high free<br />

surface energy. As any physical system, nanoparticles always have a tendency to<br />

minimize their full energy that leads to their agglomeration <strong>and</strong> losing nanoscale size, it<br />

should be noted that Van der Waals attraction (see below) can also take part in inducing<br />

an agglomeration. In order to prevent agglomeration particles have to be stabilized.<br />

However before considering stabilization principals it makes sense to mention different<br />

type <strong>of</strong> interactions between colloidal particles [125], they are summarized below.<br />

Excluded Volume Repulsion: This refers to the impossibility <strong>of</strong> any overlap between<br />

hard particles.<br />

Electrostatic interaction: Colloidal particles <strong>of</strong>ten carry an electrical charge <strong>and</strong><br />

therefore attract or repel each other. The charge <strong>of</strong> both the continuous <strong>and</strong> the<br />

dispersed phase, as well as the mobility <strong>of</strong> the phases are factors affecting this<br />

interaction.<br />

Van der Waals forces: This interaction is due to induced dipole-dipole interaction. Even<br />

if the particles don't have a permanent dipole, fluctuations <strong>of</strong> the electron gas give rise<br />

to a temporary dipole, meaning that Van der Waals forces are always present, although<br />

possibly at a much lower magnitude than others.<br />

Entropic forces: According to the second law <strong>of</strong> thermodynamics, a system progresses<br />

to a state in which entropy is maximized. This can result in effective forces even<br />

between hard spheres.<br />

Steric forces between polymer-covered surfaces or in solutions containing nonadsorbing<br />

polymer can modulate interparticle forces, producing an additional repulsive<br />

steric stabilization force or attractive depletion force between them.<br />

Steric stabilization <strong>and</strong> electrostatic stabilization are the two main mechanisms for<br />

colloid stabilization. Electrostatic stabilization is based on the mutual repulsion <strong>of</strong> like<br />

20


electrical charges; in fact the total electric potential is the sum <strong>of</strong> electrostatic <strong>and</strong> Van<br />

der Waals potentials <strong>and</strong> results in energy barrier. Electrostatic repulsion (Fig. 2-4 left)<br />

is generally caused by an electrical double (really multi) layer [131] formed by ions<br />

adsorbed on the particle surface (e.g. halides) <strong>and</strong> the corresponding counter ions (e.g<br />

tetraalkylammonium). Small particle sizes lead to enormous surface areas, <strong>and</strong> this<br />

effect is greatly amplified in colloids. In a stable colloid, the mass <strong>of</strong> a dispersed phase<br />

is so low that its buoyancy or kinetic energy is too little to overcome the electrostatic<br />

repulsion between charged layers <strong>of</strong> the dispersing phase. The charge on the dispersed<br />

particles can be observed by applying an electric field: all particles migrate to the same<br />

electrode <strong>and</strong> therefore must all have the opposite sign charge with respect to the<br />

electrode’s charge.<br />

Fig. 2-4. Schematic image <strong>of</strong> two electrostatically-stabilized nanoparticles (top-left) <strong>and</strong><br />

plot <strong>of</strong> electrical potential space distribution (bottom-left). Two sterically-stabilized<br />

nanoparticles (top-right) <strong>and</strong> caused a large energy barrier (bottom-right) against<br />

particle interaction. Pictures are adapted from [128, 132].<br />

Steric stabilization (Fig. 2-4-right) is achieved by the coordination <strong>of</strong> sterically<br />

dem<strong>and</strong>ed molecules that act as protective shields on the metallic surface. In this way,<br />

nanometallic cores are separated from each other <strong>and</strong> agglomeration is prevented.<br />

2.2 Methods <strong>of</strong> colloid <strong>and</strong> supported nanoparticles<br />

preparation<br />

Nanoparticles can be obtained by two general approaches: “top down” <strong>and</strong> “bottom up”.<br />

In “top down” method bulk materials are mechanically ground to the nanosized scale<br />

<strong>and</strong> stabilized by a suitable molecule [133, 134]. The problem with this method is<br />

difficulty in achieving the narrow size distribution <strong>and</strong> control both shape <strong>and</strong> average<br />

size <strong>of</strong> the particles.<br />

Moreover, bimetallic nanoparticles with core shell structures cannot be obtained by this<br />

method. In “bottom up” methods, nanoparticles are obtained by starting with molecular<br />

21


precursors <strong>and</strong> building-up. This method allows control <strong>of</strong> size <strong>and</strong> shape <strong>of</strong><br />

nanoparticles <strong>and</strong> their size distribution. Taking into account these reasons we will<br />

focus on “bottom up” methods only.<br />

The “bottom up” method can be classified conditionally as methods occurring in<br />

condensed phase <strong>and</strong> gas (or vacuum) phase. In condensed phase synthesis,<br />

nanoparticles are prepared by means <strong>of</strong> chemical synthesis; this method is also called as<br />

wet chemical preparation. For example, in order to prepare metal nanoparticles, metal<br />

salts are usually reduced in the presence <strong>of</strong> suitable capping agents (stabilizer) that<br />

stabilize the nanoparticles from agglomeration. In the gas phase synthesis, metal is<br />

vaporized <strong>and</strong> then vaporized metal atoms are condensed. The same idea can be<br />

realized in vacuum where the metal <strong>of</strong> interest is vaporized with high energy ions or by<br />

heating or by laser beam (the latter method can also be used in liquid media) <strong>and</strong> thus<br />

generated metal vapour is deposited on a support. Below we describe methods for<br />

preparation metal (due to their particular interest in the current thesis) nanoparticles in<br />

condensed <strong>and</strong> briefly in gas phase.<br />

Wet chemical preparation method<br />

Wet chemical reduction procedures have been widely used to stabilize the different<br />

types <strong>of</strong> metal nanoparticles with different kinds <strong>of</strong> organic as well as inorganic<br />

stabilizers.<br />

As was mentioned above, metal precursors are usually reduced in the presence <strong>of</strong> a<br />

stabilizer in order to prevent the nanoparticles from agglomeration. The choice <strong>of</strong> metal<br />

precursor, stabilizer, reducing agent <strong>and</strong> reaction conditions (relative<br />

precursor/stabilizer concentration, solvent, temperature <strong>and</strong> even stirring rate) strongly<br />

influence in properties <strong>of</strong> obtained nanoparticles, i.e. average particle size.<br />

The main classes <strong>of</strong> stabilizer predominant in the literature are macromolecules<br />

containing P, N, <strong>and</strong> S (e.g. phosphanes, amines, thioethers) atoms [105, 135-145].<br />

Such solvents as THF [135, 146] or THF/MeOH [147] also can act as stabilizers.<br />

Molecules like ionic surfactants stabilize the nanoparticles by means <strong>of</strong> both<br />

electrostatic repulsion <strong>and</strong> steric hindrance. Their polar group forms an electrostatic<br />

layer around them <strong>and</strong> the long organic moiety gives steric hindrance. In general,<br />

lipophilic protecting agents give metal colloids that are soluble in organic media<br />

(“organosols”), while hydrophilic agents yield water-soluble colloids (“hydrosols”).<br />

Reducing agents play an important role in directing the size <strong>and</strong> shape <strong>of</strong> particles.<br />

Among different kinds <strong>of</strong> reducers (NaBH 4 , R-CHO, etc.) we would like to underline<br />

hydrogen. In fact, it has been applied in the synthesis <strong>of</strong> various metal clusters. The<br />

advantage <strong>of</strong> using hydrogen lies in the avoidance <strong>of</strong> by-products, the disadvantage is a<br />

weak reducing agent that is unable to reduce many metal complexes <strong>and</strong> salts.<br />

However, the use <strong>of</strong> hydrogen makes possible to perform reduction under mild<br />

condition that might be important for sensitive stabilizers.<br />

It is interesting to note, that reduction <strong>of</strong> metal salts e.g. Na 2 PtCl 4 by microwave<br />

radiation, thus in presence <strong>of</strong> stabilizer gives 20-30 Pt nanoparticles [148].<br />

The decomposition <strong>of</strong> organometallic complexes with null valence metal atom(s) in<br />

presence <strong>of</strong> stabilizer can also be applied as a method <strong>of</strong> preparation <strong>of</strong> nanoparticles.<br />

This decomposition is typically initiated by an external force e.g. heat, ultrasonic or by<br />

a chemical agent capable to induce decomposition e.g. H 2 or CO. Classical example is<br />

in thermal decomposition <strong>of</strong> iron Fe(CO) 6 [149] or cobalt complex Co 2 (CO) 8 [150] at<br />

130-170 °C in presence <strong>of</strong> polymer yielding 45 nm Co nanoparticles [151].<br />

22


The advantage <strong>of</strong> use <strong>of</strong> organometallic route in preparation <strong>of</strong> nanoclusters is obvious<br />

when it is needed to stabilize (modify) nanoparticles by hydrophobic surfactant or in<br />

other words, when required stabilizer is water insoluble.<br />

An interesting method <strong>of</strong> preparation <strong>of</strong> metal nanoparticles is through electrochemical<br />

approach established by Reetz [152]. In this method, an anode is made up <strong>of</strong> the<br />

specific metal (required for nanoparticles) <strong>and</strong> a cathode is <strong>of</strong> any other metal. Under<br />

the suitable applied current density the anode sacrificially dissolves in the electrolyte,<br />

the metal ions migrate towards the cathode <strong>and</strong> reduction occurs. The nucleation <strong>and</strong><br />

growth <strong>of</strong> reduced metal atoms occurs at the electrode surface. The stabilizing agent in<br />

the reaction vessel arrests the growth <strong>of</strong> the nanoparticles <strong>and</strong> facilitates the formation<br />

<strong>of</strong> stable nanostructures. The final step is dissolution <strong>of</strong> nanoparticles from the<br />

electrode surface into the bulk <strong>of</strong> the solution. The advantages <strong>of</strong> the electrochemical<br />

pathway are that the contamination with by-products resulting from chemical reducing<br />

agents is avoided <strong>and</strong> that the products are easily isolated from the precipitate. Further,<br />

the electrochemical preparation allows for size-selective particle formation. The<br />

particle size obtained by the electrochemical route depends on many factors, the<br />

distance between the electrodes, reaction time, temperature, <strong>and</strong> polarity <strong>of</strong> the solvent<br />

contribute to the particle size. Experiments have also shown that the applied current<br />

density also has a major influence on the particle size.<br />

Vapor phase synthesis<br />

As was mentioned above metal atoms can be obtained not only through reduction <strong>of</strong><br />

metal salts, but also through solid to gas phase transformation <strong>of</strong> metal. The Solvated<br />

metal atom dispersion (SMAD) techniques is based on this method. Typically, the metal<br />

<strong>of</strong> interest is heated in a crucible at elevated temperature evaporated under vacuum <strong>and</strong><br />

then co-condensed with a specified amount <strong>of</strong> liquid lig<strong>and</strong> substrate on the liquid<br />

nitrogen cooled walls <strong>of</strong> the reactor vessel. The reactor is removed from the liquid<br />

nitrogen <strong>and</strong> the reactor is allowed to warm slowly. After the warming stage, particles<br />

can be stabilized either sterically (by solvation) or electrostatically (by incorporation <strong>of</strong><br />

negative charge). The SMAD techniques has been successfully applied e.g. in<br />

preparation <strong>of</strong> thiol stabilized gold colloid [153]. The main advantage <strong>of</strong> this “clean”<br />

method <strong>of</strong> preparing nanoclusters in gram scale is in fact, that it does not involve<br />

starting materials containing counter ions like nitrate, chlorides, sulfates which are<br />

common in reduction synthetic methods. However, the use <strong>of</strong> SMAD method is limited<br />

because <strong>of</strong> difficulties in the operation <strong>of</strong> the apparatus <strong>and</strong> it is difficult to obtain<br />

narrow particle size distributions.<br />

Another interesting method is through the laser ablation <strong>of</strong> bulk metal target crystal<br />

being in liquid phase. The recent example is in preparation <strong>of</strong> Pt nanoparticles by laser<br />

ablation <strong>of</strong> Pt crystal in water [154-156]. Due to the laser induced evaporation <strong>of</strong> Pt<br />

atoms into the liquid phase with formation <strong>of</strong> 1-15 nm nanoparticles, it was found that<br />

the yield <strong>of</strong> nanoparticles depends on the wavelength <strong>of</strong> the used laser beam whereas it<br />

has no significant influence on the average size <strong>of</strong> produced nanoparticles. The laser<br />

ablation method has the same important advantages as SMAD method demonstrates,<br />

moreover there is not need to work with very high temperatures (for metal evaporation)<br />

<strong>and</strong> keep system under high vacuum. However with respect to the chemical methods<br />

the laser ablation has the following disadvantages. There is more insight about the<br />

reaction at plasma <strong>and</strong> liquid interface <strong>and</strong> it is relatively more expensive to obtain the<br />

large amount <strong>of</strong> products, in fact the rate <strong>of</strong> nanoparticles production was about 4.4<br />

mg/h.<br />

23


2.3 Preparation <strong>of</strong> nanoclusters on a heterogeneous<br />

support<br />

Metal clusters especially in nanoscale size are widely used in heterogeneous catalysis.<br />

Heterogeneous catalysts are traditionally obtained by impregnation methods consisting<br />

<strong>of</strong> immersing a solid support like Al 2 O 3 , SiO 2 or ZrO 2 in a solution <strong>of</strong> metal salts <strong>of</strong><br />

interest <strong>and</strong> inducing deposition. After some specified time, the solid substance is<br />

filtered <strong>and</strong> dried. The dried product is calcinated at high temperature in the presence <strong>of</strong><br />

gaseous reducing agents such as H 2 or CO to yield the final catalytic material. The<br />

problem with impregnation methods is the difficulty in controlling the particle size <strong>and</strong><br />

that the metal particle formed after the reduction may be located where it is not<br />

accessible for further application.<br />

An alternative method for the preparation <strong>of</strong> heterogeneous catalysis is the “precursor<br />

concept“ which was originally developed by Turkevich in 1970 [157] <strong>and</strong> was further<br />

developed in the 1990's by many groups [135, 158-161].<br />

This method consists <strong>of</strong> depositing the pre-prepared metal colloid on a solid support; by<br />

applying this method it is possible to control the size <strong>and</strong> shape <strong>of</strong> the nanoparticles.<br />

Application <strong>of</strong> the precursor concept also allows full pre-characterization <strong>of</strong> the catalyst<br />

components (support <strong>and</strong> colloid) prior to deposition, thus facilitating the possibilities<br />

for comprehensive study <strong>of</strong> heterogeneous catalyst support/metal effects. Moreover,<br />

almost all <strong>of</strong> the particles on the inert support are accessible because the preformed<br />

colloids generally do not “fill” cracks <strong>and</strong> other defects <strong>of</strong> the support’s surface as<br />

molecular precursors <strong>of</strong>ten do.<br />

Metal nanoparticles have a high active metal surface area (50-100 m 2 /g for 6-3 nm<br />

particles) that facilitates proper surface modification (roughly, due to the high activity<br />

<strong>of</strong> “freshly” prepared nanoclusters) <strong>and</strong> sometimes unique size dependence properties.<br />

The catalytic properties <strong>of</strong> supported <strong>and</strong> not supported metal nanoclusters were found<br />

in many important reactions such as oxidation (Co, Pt, Pd, Ag <strong>and</strong> Au), hydrogenation<br />

(Pt, Pd, Pt/Pd, Au/Pt, Au/Pd <strong>and</strong> Rh) <strong>and</strong> others. The detailed overview can be found in<br />

the work <strong>of</strong> D’Souza [162].<br />

24


Chapter 3<br />

Current thesis in the EU-COST project:<br />

aims <strong>of</strong> the work<br />

3.1 Introduction <strong>and</strong> goals setting<br />

The idea to use metal nanoclusters modified with chiral lig<strong>and</strong>s is a much promising<br />

strategy in research <strong>and</strong> development <strong>of</strong> enantioselective catalysts. In fact, making<br />

nanoparticles stabilized with chiral lig<strong>and</strong>s it is possible to reach maximum chiral<br />

coverage <strong>of</strong> the metal surface, that is very important for obtaining high<br />

enantioselectivity. Since Pt <strong>and</strong> Pd (generally alumina or silica supported) were found<br />

to be the best metals for heterogeneous enantioselective hydrogenation, in this thesis,<br />

we focused on the chiral modification <strong>of</strong> the surface <strong>of</strong> Pt <strong>and</strong> Pd nanoclusters.<br />

The current thesis is a part <strong>of</strong> the “EU COST D24 WORKING GROUP 7<br />

D24/0007/02” working project, which besides International <strong>University</strong> Bremen (Pr<strong>of</strong>.<br />

R. Richards) also involves research groups from the Marseille <strong>University</strong> (Pr<strong>of</strong>. G.<br />

Buono), Ecole Nationale Supérieure de Chimie de Paris Institute (Pr<strong>of</strong>. J.-P. Genet),<br />

<strong>University</strong> <strong>of</strong> Ljubljana (Pr<strong>of</strong>. M. Kocevar) <strong>and</strong> <strong>University</strong> <strong>of</strong> Bucharest (Pr<strong>of</strong>. V. I.<br />

Parvulescu).<br />

The working scheme (Fig. 3-1) shows that the groups from Paris <strong>and</strong> Marseille provide<br />

us with specific chiral lig<strong>and</strong>s (synphos <strong>and</strong> quiphos with quiphos-spider, respectively).<br />

At the International <strong>University</strong> Bremen we have prepared Pt <strong>and</strong> Pd nanoclusters<br />

modified with obtained from our colleagues (<strong>and</strong> purchased from Fluka <strong>and</strong> Aldrich)<br />

chiral lig<strong>and</strong>s for chiral modification <strong>of</strong> them on the surface <strong>of</strong> metal nanoclusters. The<br />

groups in Paris, Ljubljana <strong>and</strong> Bucharest study catalytic activity (as well as<br />

enantioselectivity) over obtained samples in corresponding reactions. The details about<br />

their work can be found in the description <strong>of</strong> the EU-COST project [163] <strong>and</strong> from the<br />

corresponding research groups.<br />

25


Paris<br />

Marseille<br />

Ljubljana<br />

Synphos<br />

Quiphos<br />

Quiphos - spider<br />

IUB catalysts<br />

preparation <strong>and</strong><br />

characterization<br />

Bucharest<br />

Aldrich<br />

Fluka<br />

Cinchonidine,<br />

binap, diop, diphos<br />

Fig. 3-1. The working scheme <strong>of</strong> EU-COST project.<br />

The all used in the current work lig<strong>and</strong> for the surface modification are summarized in<br />

the Fig. 3-2.<br />

O<br />

HO<br />

N<br />

N<br />

N<br />

P<br />

N<br />

O<br />

PPh 2<br />

O<br />

PPh 2<br />

PPh 2<br />

O PPh 2<br />

N<br />

Cinchondine<br />

Quiphos<br />

(2R, 5S)-3-Phenyl-2-(8-quinolinoxy)<br />

-1,3-diaza-2-phosphabicyclo-octane<br />

Binap<br />

2,2'-Bis(diphenylphosphino)-<br />

1,1'-binaphthyl<br />

O<br />

Synphos<br />

(2,3,2',3'-tetrahydro-5,5'-bi(1,4-<br />

benzodioxin)<br />

-6,6'-diyl)-bis(diphenylphosphane)<br />

P<br />

P<br />

P<br />

P O<br />

P<br />

O<br />

O<br />

Diphos<br />

Bis(diphenylphosphino)ethane<br />

Diop<br />

4,5-Bis(diphenylphosphinomethyl)-<br />

2,2-dimethyl -1,3-dioxolane<br />

Quiphos-Spider<br />

4-isopropyl-2 methyl cyclohexyl-1methylphosphonite<br />

-2-2'-1,5 binaphtyl<br />

Fig. 3-2. Lig<strong>and</strong>s used for modification <strong>of</strong> surface <strong>of</strong> Pt <strong>and</strong> Pd nanoclusters.<br />

26


Besides preparation <strong>of</strong> chiral lig<strong>and</strong> modified Pt <strong>and</strong> Pd nanoclusters another aim <strong>of</strong><br />

the current thesis is in determination the geometrical models <strong>of</strong> adsorption <strong>of</strong> chiral<br />

lig<strong>and</strong>s (Fig. 3-2) on the surface <strong>of</strong> Pt <strong>and</strong> Pd nanoclusters via DRIFTS (Diffuse<br />

Reflectance Infra-red Fourier Transform Spectroscopy) <strong>and</strong> molecular modeling. Since<br />

the ability <strong>of</strong> the adsorbed cinchonidine to induce enantioselectivity e.g. in<br />

hydrogenation <strong>of</strong> α-ketoesters is well known in heterogeneous enantioselective<br />

catalysis we placed high emphasis on the examination <strong>of</strong> cinchonidine modified Pt <strong>and</strong><br />

Pd nanoclusters in ethyl pyruvate hydrogenation. The work is also focused on<br />

obtaining active, highly enantioselective stable catalyst for the performing<br />

hydrogenation under mild conditions that is very important from application point <strong>of</strong><br />

view. Furthermore we investigated catalytic <strong>and</strong> enantioselective activity <strong>of</strong> all<br />

synthesized samples in typical Orito reaction.<br />

27


Chapter 4<br />

Cinchonidine modified Pt colloidal<br />

nanoparticles: characterization <strong>and</strong><br />

catalytic properties<br />

4.1 Introduction<br />

Cinchonidine modified transition metal surfaces (primarily Pt) represent one approach<br />

towards asymmetric heterogeneous catalysis. Since the first report by Orito [45, 46],<br />

numerous experimental <strong>and</strong> theoretical investigations into the cinchona on Pt <strong>and</strong> Pd<br />

systems have revealed a great deal <strong>of</strong> insight into the system <strong>and</strong> its applicability. As<br />

was mentioned above, adsorption geometry <strong>of</strong> the cinchonidine molecule on metal<br />

surface has a significant role in the enantioselectivity. It is known from the work <strong>of</strong><br />

Ferri [76], Zaera [77] <strong>and</strong> Wells [74], that cinchonidine can adsorb on the surface <strong>of</strong><br />

macroscopic Pt (plane, single crystal with sizes are much bigger than the size <strong>of</strong> the<br />

cinchonidine molecule) in three different orientations (Fig. 1-19): “π-bonded”, “α-H<br />

abstracted” <strong>and</strong> “N lone pair bonded”. It is also considers, that only cinchonidine in the<br />

flat “π-bonded” orientation is able to induce enantioselective hydrogenation. However<br />

adsorption mode <strong>of</strong> cinchonidine on the surface <strong>of</strong> nanosized Pt crystals has never been<br />

investigated <strong>and</strong> it attracts special attention, for instance, because the maximum<br />

reported enantiomeric excess (98 %, in the hydrogenation <strong>of</strong> methyl pyruvate) was<br />

obtained on small (1.2 nm) Pt colloids. Thus the beginning <strong>of</strong> this chapter is dedicated<br />

to the determination <strong>of</strong> adsorption modes <strong>of</strong> cinchonidine on Pt nanoclusters.<br />

The initial transient period (ITP) is an interesting phenomenon which consists <strong>of</strong> an<br />

increase <strong>of</strong> enantiomeric excess in the beginning <strong>of</strong> e.g. ethyl pyruvate hydrogenation.<br />

The ITP effect was observed in the examination <strong>of</strong> enantioselectivity <strong>of</strong> cinchonidine<br />

modified Pt nanoclusters <strong>and</strong> discussed in the second part <strong>of</strong> the current chapter. The<br />

initial transient period (ITP) for the Pt/Al 2 O 3 system has been reported <strong>and</strong> investigated<br />

by numerous groups <strong>and</strong> has been the subject <strong>of</strong> intense discussions [164, 165]. ITP<br />

was observed by Baiker <strong>and</strong> co-workers [166] <strong>and</strong> was attributed to an impurity effect<br />

(from destructive adsorption <strong>of</strong> EP <strong>and</strong> alcoholic solvents on Pt) which could be<br />

eliminated by pre-activation with hydrogen. Blackmond <strong>and</strong> co-workers proposed the<br />

“reaction-driven equilibration" explanation [167] <strong>and</strong> demonstrated the significant<br />

influence <strong>of</strong> mass transport (gas-liquid diffusion <strong>of</strong> hydrogen) limitation on the<br />

enantioselectivity [168]. Recently, it was shown that ITP is not only the effect <strong>of</strong><br />

impurities (lactates in ethyl pyruvate) but could also be a result <strong>of</strong> competitive<br />

adsorption <strong>of</strong> reactant, modifier <strong>and</strong> solvent [169]. From the work <strong>of</strong> Hutchings et al.<br />

[170], we know, that the ITP effect was found with cinchonidine premodified<br />

conventional Pt/Al 2 O 3 in ethanol <strong>and</strong> dichloromethane solution <strong>and</strong> explanation through<br />

“removal <strong>of</strong> adsorbed oxygen species from the surface” was proposed.<br />

Here, we investigate the behavior <strong>of</strong> enantiomeric excess (ee) during hydrogenation <strong>of</strong><br />

ethyl pyruvate with conventional Pt/Al 2 O 3 <strong>and</strong> with cinchonidine Pt nanoclusters<br />

prepared by two different methods, an aqueous (first reported by Bönnemann [114])<br />

<strong>and</strong> a novel organometallic method through decomposition <strong>of</strong> Pt 2 (DBA) 3 (DBA is bisdibenzylidene<br />

acetone) complex. The nature <strong>of</strong> the increase in ee at initial reaction<br />

time, further decreases <strong>and</strong> differences in the behavior <strong>of</strong> conventional Pt/Al 2 O 3 <strong>and</strong><br />

28


“quasi-homogeneous” cinchonidine stabilized Pt nanoclusters are discussed below.<br />

Further, the experimental conditions selected for these experiments were chosen to be<br />

mild: room temperature <strong>and</strong> low hydrogen pressure (2-10 bar) since these are likely to<br />

be the most relevant conditions for hydrogenation <strong>of</strong> α-ketoesters on an industrial scale.<br />

4.2 Experimental<br />

Materials<br />

Unless otherwise mentioned the following materials: Pt on alumina (Aldrich 5% Pt),<br />

cinchonidine (> 98%, Fluka), acetic acid (99.8% Fluka), ethyl pyruvate (98%, Aldrich),<br />

THF (Applichem, 99.8%), formic acid (Aldrich, 99%), bis-dibenzylidene acetone<br />

(Fluka, ≥96.0%), K 2 PtCl 4 (Acros Organics) <strong>and</strong> ethanol (Applichem, 99.9%) were used<br />

as received in all chapters.<br />

The 5 % Pt on alumina (E4759) catalyst was kindly provided by Engelhard company by<br />

request.<br />

Sample preparation<br />

In order to activate the conventional Pt/Al 2 O 3, it was heated at 300 °C under vacuum<br />

(10 -3 mbar) for 1 hour, then kept for 3 hours under flowing hydrogen (100 ml/min) <strong>and</strong><br />

cooled to room temperature. Then the required amount was weighed <strong>and</strong> transferred<br />

into a glass reactor with acetic acid, ethyl pyruvate <strong>and</strong> cinchonidine.<br />

Chiral modification <strong>of</strong> conventional Pt/Al 2 O 3 . 40 mg <strong>of</strong> conventional Pt/Al 2 O 3 was<br />

heated to 400 ºC under vacuum (10 -3 mbar) for 1 hour, then the mixture <strong>of</strong> N 2 (93%)<br />

<strong>and</strong> H 2 (7%) gases were passed (100 ml min -1 ) over the sample for 3 hours. Then, under<br />

flowing gases (H 2 <strong>and</strong> N 2 ), the support was cooled to room temperature <strong>and</strong> the solution<br />

<strong>of</strong> cinchonidine in CHCl 3 (6.0 ml, 3 mM) was added by syringe through a stopper. The<br />

system was kept at 10-15 ºC for 24 hours, then washed 3 times with CHCl 3 <strong>and</strong> finally<br />

dried under vacuum at room temperature.<br />

Preparation <strong>of</strong> 10.11-dihydrocinchonidine (DHCIN) has been done according to the<br />

protocol from work <strong>of</strong> Morawsky [171]. 10 g cinchonidine (33 mmol) was dissolved in<br />

100 ml 0.5 M H 2 SO 4 <strong>and</strong> placed in a metal reactor. 0.5 g Pd/C (5 % wt, Degussa<br />

E101/0/W) was added to the solution. The H 2 pressure was set to 6 bars, reaction<br />

started by stirring. After 1.5 h reaction was finished. The catalyst was removed via a G4<br />

filter, 50 ml 2 M NaOH was added to precipitate the hydrated alkaloid. The white<br />

precipitate was washed with 500 ml H 2 O <strong>and</strong> recrystallized in ethanol. It has to be<br />

noted that resulted product had some amount <strong>of</strong> cinchonidine, as was detected as<br />

presence <strong>of</strong> attenuated peak at 1636 cm -1 in IR spectrum <strong>of</strong> the product, which<br />

corresponds to the stretching <strong>of</strong> C=C bond.<br />

The preparation <strong>of</strong> Pt 2 (DBA) 3 organometallic complex has been done according to the<br />

literary protocol [172, 173]. Typically, to a solution <strong>of</strong> 2.8 g sodium acetate<br />

<strong>and</strong> 2.36 g <strong>of</strong> bis-dibenzylidene acetone (dba) in 60 mL <strong>of</strong> ethanol at 50 °C was added a<br />

solution <strong>of</strong> 1.40 g K 2 PtCl 4 in 12 mL <strong>of</strong> freshly distilled water. The initial pale yellow<br />

suspension redissolved while the mixture was heated to the refluxing temperature (90<br />

°C). After refluxing for 1 h a dark violet precipitate formed. The mixture was allowed<br />

to settle overnight. The solution was eliminated by filtration <strong>and</strong> the solid washed three<br />

times with 200 mL <strong>of</strong> water, dried overnight under vacuum, further washed three times<br />

with 200 mL <strong>of</strong> pentane <strong>and</strong> finally dried under vacuum overnight.<br />

A water soluble Pt nanoclusters modified with DHCIN was prepared based on the<br />

protocol published by Bönnemann [114]. Typically, 0.208 g (0.6 mmol) PtCl 4 in 160 ml<br />

distilled water was reduced with 0.1 M aqueous formic acid (15 ml) with dissolved<br />

29


dihydrocinchonidine (355 mg or 1.2 mmol) under reflux, at ~95 ºC. After 10-15<br />

minutes the product was precipitated into 200 ml <strong>of</strong> aqueous saturated NaHCO 3<br />

solution, then the precipitate was filtered with a G4 fritted glass filter <strong>and</strong> washed first<br />

with a half-saturated NaHCO 3 (400 ml) solution followed by water. Finally, the black<br />

powder was taken up from the filter using an acetic acid solution. The acetic acid – Ptcinchonidine<br />

system was dried at room temperature under vacuum. This catalyst was<br />

found to be redispersible in water, acetic acid <strong>and</strong> toluene. This catalyst will be referred<br />

to as B – henceforth in this chapter. Physical state: black powder.<br />

The cinchonidine modified Pt nanosized catalyst was prepared by decomposition <strong>of</strong> the<br />

organometallic complex Pt 2 (DBA) 3 under hydrogen in the presence <strong>of</strong> cinchonidine.<br />

Pt 2 (DBA) 3 (402 mg or 0.37 mmol) <strong>and</strong> an excess <strong>of</strong> cinchonidine (4.5 g or 15 mmol)<br />

were dissolved in 160 ml THF <strong>and</strong> flushed with Ar for 20 min. The mixture was kept in<br />

the glass reactor under 2 bar <strong>of</strong> H 2 with constant stirring at 300 rpm. After 18 hours the<br />

black precipitate was collected <strong>and</strong> washed first with 50 ml pentane, then with a<br />

pentane/THF mixture (5:1) <strong>and</strong> chlor<strong>of</strong>orm until all hydrogenated DBA <strong>and</strong> excess<br />

cinchonidine was removed as monitored by IR <strong>and</strong> XRD spectroscopy. Thus prepared<br />

catalysts were found to be redispersible in acetic acid <strong>and</strong> THF <strong>and</strong> will be referred as<br />

D – henceforth in this chapter. Physical state: black powder.<br />

In order to purify ethyl pyruvate, 20 ml <strong>of</strong> it was added to 80 ml <strong>of</strong> water/n-pentane<br />

(1/1 by volume) mixture <strong>and</strong> kept for two days in a separation funnel. The n-<br />

pentane/ethyl pyruvate fraction was separated <strong>and</strong> n-pentane was evaporated in a rotary<br />

evaporator. The procedure was repeated two times <strong>and</strong> the collected purified ethyl<br />

pyruvate was found to contain 0.05 % R, S ethyl lactate (from the initial 0.5%) by GC.<br />

Characterization <strong>and</strong> experiments<br />

Hydrogen chemisorption on conventional Pt/Al 2 O 3 (Aldrich) catalyst was performed on<br />

the Autosorb instruments supplied by Quantachrome, Germany. Typically the sample<br />

was preheated at 400 °C for 1 h under vacuum (10 -2 mm. Hg.) <strong>and</strong> for 3 hour under<br />

flow <strong>of</strong> hydrogen then chemisorption has started at 40 °C in the 0-586 mm. Hg.<br />

hydrogen pressure interval.<br />

Transmission electron microscopy (TEM) micrographs were obtained by using a<br />

TECNAI F20 instrument in the IFAM Bremen. Specimens were prepared by placing a<br />

drop <strong>of</strong> the colloidal onto a copper grid with a perforated carbon film <strong>and</strong> then allowing<br />

the solvent to evaporate. Around 100-150 particles were considered for size distribution<br />

calculations.<br />

X-ray diffraction (XRD): Powder XRD analyses were performed using a Siemens<br />

D5000 instrument with Cu Kα radiation (wavelength 1.54·10 -10 m) operated at 40 kV<br />

<strong>and</strong> 40 mA.<br />

Gas chromatography: (GC) Measurements were performed at 88 °C (isothermal) on a<br />

Varian GC 3900, FID with a Lipodex-E chiral column.<br />

Catalytic Reactor: The catalytic reactor was supplied by Parr Instrument (Germany)<br />

GmbH. The apparatus consists <strong>of</strong> five major components; reactor, H 2 reservoir, oil<br />

reservoir, a temperature <strong>and</strong> pressure controller <strong>and</strong> computer. The batch type reactor<br />

(300 ml) is made <strong>of</strong> double walled borosilicate glass with a maximum pressure rating <strong>of</strong><br />

10 bars. A regulator situated between the reactor <strong>and</strong> reservoir maintains a constant<br />

hydrogen pressure inside the reactor. The reactor chamber also consists <strong>of</strong> a water<br />

cooler, an electric stirrer <strong>and</strong> thermocouple.<br />

Hydrogenation <strong>of</strong> ethyl pyruvate was carried out in the described above 300 ml glass<br />

reactor under a pressure <strong>of</strong> either 2 or 10 bars <strong>of</strong> hydrogen at room temperature (22±1<br />

30


°C). The stirring speed was 300 rpm. The ethyl pyruvate was added to 70 ml <strong>of</strong> acetic<br />

acid, then the mixture was sonicated for 10 min, Ar was then flushed through the<br />

solution for 10 min to remove dissolved air. The required amount <strong>of</strong> catalyst was<br />

added, the mixture was sonicated for 5 min, transferred into the reactor, flushed with Ar<br />

for 10 min <strong>and</strong> finally after the pressure <strong>of</strong> hydrogen had been stabilized for 10-15 sec<br />

the initial time (t=0) was set <strong>and</strong> the reaction kinetics were monitored. Around one ml<br />

aliquots <strong>of</strong> the reaction mixture were taken at certain time intervals, filtrated with a G4<br />

filter <strong>and</strong> 1 μl was injected into the gas chromatograph (GC).<br />

FTIR spectra <strong>of</strong> D type <strong>of</strong> catalyst were measured with a “Thermo Nicolet Avatar”<br />

spectrometer in transmission mode, resolution 4 cm -1 , number <strong>of</strong> scans 500. IR pellets<br />

were prepared by mixing 100 mg KBr <strong>and</strong> 1.5 mg sample then pressing into a pellet<br />

under a pressure <strong>of</strong> 10 tons. For IR investigation <strong>of</strong> the B type <strong>of</strong> catalyst a “Thermo<br />

Nicolet 4700 FT-IR” spectrometer with liquid nitrogen cooled detector <strong>and</strong> DRIFT<br />

accessory was used with the following parameters: 3000 scans, 600-2000 cm -1 scan<br />

range, 4 cm -1 resolution. The DRIFT spectra were chosen for analysis because it is the<br />

more applicable technique for observing adsorbed species on finely divided metal<br />

powder [174].<br />

The carbon <strong>and</strong> nitrogen elemental analysis measurements were done on a Carlo Erba<br />

NA2500 C/N analyzer in the <strong>School</strong> <strong>of</strong> Geo<strong>Science</strong>s at Grant Institute in the <strong>University</strong><br />

<strong>of</strong> Edinburgh. Measurements were repeated several times with an error ~1%. The metal<br />

content was found from thermogravimetric Analysis (TGA) in 30-1200 °C temperature<br />

interval under nitrogen 100 ml/min with an SDTQ 600 instrument from TA<br />

Instruments.<br />

The cumulative enantiomeric excess (ee) was calculated according to the formula<br />

[ R]<br />

− [ S]<br />

ee = ⋅100%<br />

.<br />

[ S]<br />

+ [ R]<br />

The actual enantiomeric excess (ee*) was calculated by the previously established<br />

[169] formula<br />

ee y − ee y<br />

ee<br />

*<br />

i+<br />

1 i+<br />

1 i i<br />

i+<br />

1 / 2<br />

=<br />

,<br />

yi+<br />

1<br />

− yi<br />

where y i - mol <strong>of</strong> both ethyl lactates, ee –cumulative (observed) enantiomeric excess.<br />

Computational methods: Geometry optimization <strong>and</strong> calculation <strong>of</strong> IR spectra <strong>of</strong><br />

cinchonidine <strong>and</strong> quinoline were carried out using the Gaussian 03 program [175].<br />

Density functional theory (DFT) calculations were performed by using the B3LYP/6-<br />

31G method <strong>and</strong> basis set. The vibrations <strong>and</strong> orientations <strong>of</strong> dipole moment were<br />

visualized with GaussView <strong>and</strong> MOLDEN programs.<br />

4.3 Results <strong>and</strong> discussion<br />

4.3.1 General characterization<br />

Choice <strong>of</strong> conventional catalyst<br />

The Engelhard 4759 (5% Pt/Al 2 O 3 ) is known from 90 th to be one <strong>of</strong> the most studied<br />

catalyst for Orito reaction, however the Engelhard company does not produce one any<br />

more <strong>and</strong> it is not widely available. Due to this reason <strong>and</strong> the fact that under the same<br />

reaction conditions (Table 4-1) catalyst from Aldrich demonstrates higher<br />

enantioselectivity <strong>and</strong> reaction rate, we chose it to be the basic conventional catalyst for<br />

further comparisons.<br />

31


Table 4-1. Comparison between 5 % Pt/Al 2 O 3 catalysts from Aldrich <strong>and</strong> Engelhard<br />

under the same reaction conditions A .<br />

Supplier Initial rate, Conversion, % after Ee, % (dominated<br />

mmol·(min·g) -1 (time, min) enantiomer)<br />

Aldrich 28 80 (60) B 91 (R)<br />

Engelhard 2 34 (250) B 70 (R)<br />

Comments: A - To 4 ml ethyl pyruvate (36 mmol) <strong>and</strong> 20 mg cinchonidine (1 mM) 20<br />

mg thermally activated catalyst (please find the activation procedure in the chapter 5)<br />

was added, the mixture purged with Ar for 10 min <strong>and</strong> hydrogen pressure <strong>of</strong> 10 bar was<br />

set under the constant stirring (1000 rpm), B - according to the curve pr<strong>of</strong>ile<br />

(conversion vs. time) reaction can go further.<br />

From the hydrogen chemisorption experiment (Fig. 4-1) with conventional 5% Pt/Al 2 O 3<br />

(here <strong>and</strong> further Aldrich catalyst is mentioned as a conventional) it was determined<br />

that the metal dispersion is 29 %, active metal surface area is 3.6 m 2 g -1 <strong>and</strong><br />

concentration <strong>of</strong> active sites is 37 μmol g -1 , average crystal size is 3.8 nm.<br />

Fig. 4-1. Combined, weak <strong>and</strong> their difference (strong) curves <strong>of</strong> hydrogen<br />

chemisorption data on conventional 5% Pt/Al 2 O 3 .<br />

X-ray diffraction phase analysis <strong>of</strong> the cinchonidine modified Pt nanoclusters was<br />

carried out using powder XRD to ascertain the polycrystalline nature <strong>of</strong> the particles.<br />

32


800<br />

111<br />

Intensity, a.u.<br />

400<br />

200<br />

220 311<br />

0<br />

2Θ<br />

40 60 80<br />

Fig. 4-2. XRD spectrum <strong>of</strong> B sample.<br />

600<br />

111<br />

Intensity, a.u.<br />

400<br />

200<br />

200<br />

220<br />

0<br />

40 50 60 70<br />

2Θ<br />

Fig. 4-3. XRD spectrum <strong>of</strong> D sample.<br />

As can be seen from XRD spectra <strong>of</strong> sample B (Fig. 4-2) <strong>and</strong> D (Fig. 4-3) they both<br />

have dominant peak at ≈ 39 °, which corresponds to the (111) crystal face, as well as<br />

peak at 66 °, which corresponds to (220) crystal phase. The peak at 46 ° (corresponding<br />

to the (220) crystal phase) is clearly seen in the spectra <strong>of</strong> D sample, however it is<br />

present as poorly observable shoulder in the spectra <strong>of</strong> B sample, probable due to the<br />

widening <strong>of</strong> the dominant peak at 39 ° <strong>and</strong> their overlapping. The assignment <strong>of</strong> the<br />

peaks was based on the work <strong>of</strong> Davey [176]. The quantitative value <strong>of</strong> a peak<br />

33


widening can be used in estimation <strong>of</strong> the average crystal size, according to the<br />

Scherrer equation [177]. The obtained values can be found in the Table 4-1 below.<br />

K ⋅λ<br />

d hkl<br />

= , where K is a constant whose value depends on the shape <strong>of</strong> the<br />

Δ ⋅cos(Θ)<br />

particle <strong>and</strong> the method used to measure the peak width, in this work K was taken to be<br />

0.94, according to the work <strong>of</strong> Mukherjee [148], Δ – Peak width, taken on a half height.<br />

TEM images <strong>of</strong> the B <strong>and</strong> D samples (Fig. 4-4) <strong>and</strong> their corresponding size distribution<br />

histograms (Fig. 4-5) clearly show that D type <strong>of</strong> sample contains Pt nanoparticles with<br />

average size <strong>of</strong> 2.3 nm, whereas average size <strong>of</strong> nanoparticles <strong>of</strong> B sample is 1.4 nm.<br />

Fig. 4-4. TEM <strong>of</strong> B <strong>and</strong> D-sample.<br />

The very similar values <strong>of</strong> the average crystal size were estimated from the Scherrer<br />

equation based on XRD data are summarized in the Table 4-2 together with data<br />

obtained from TEM.<br />

Table 4-2. The average size <strong>of</strong> cinchonidine modified Pt nanoclusters obtained by<br />

aqueous (B) <strong>and</strong> organometallic (D) methods determined via TEM <strong>and</strong> XRD<br />

measurements.<br />

Type <strong>of</strong> sample (TEM), nm. (XRD), nm.<br />

B 1.4 1.2<br />

D 2.3 2.3<br />

34


Number <strong>of</strong> particles<br />

30 Types <strong>of</strong> sample<br />

D<br />

B<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

1.0 1.5 2.0 2.5 3.0 3.5<br />

Particle size, nm<br />

Fig. 4-5. Particles size distribution for B <strong>and</strong> D types <strong>of</strong> sample.<br />

From the elemental analysis it was found, that the B-series sample contains 43.5 % C,<br />

4.3 % N or 45 % cinchonidine, based on nitrogen content. From the TGA, the Pt<br />

content was found to be 25 % thus differing from the 40 % metal content in the original<br />

work [11] for 1.5 nm sized Pt cluster, probably because <strong>of</strong> the presence <strong>of</strong><br />

moisture/CO/CO2 <strong>and</strong> possibly acetic acid (30 %).<br />

Elemental analysis <strong>of</strong> the D-catalysts before reaction (please see below) showed 16.6 %<br />

C <strong>and</strong> 1.8 % N or 20 % cinchonidine, by making average between C <strong>and</strong> N. 63 % Pt<br />

was found from TGA. The remaining content was attributed to moisture/CO/CO 2 . The<br />

content <strong>of</strong> C <strong>and</strong> N on the D catalyst after 50 min <strong>of</strong> reaction was found to be 5.3 % <strong>and</strong><br />

0.8 %, correspondingly or 7 % <strong>of</strong> cinchonidine.<br />

4.3.2 FTIR investigation <strong>of</strong> cinchonidine adsorbed on Pt<br />

FTIR investigation <strong>of</strong> cinchonidine adsorbed on Pt was performed with cinchonidine<br />

modified conventional Pt/Al 2 O 3 (Fig. 4-6) <strong>and</strong> cinchonidine modified Pt nanoclusters B<br />

(Fig. 4-7) <strong>and</strong> D (Fig. 4-8) samples. The use <strong>of</strong> different Pt sample allows comparison<br />

<strong>of</strong> adsorbed modes <strong>of</strong> the cinchonidine molecule <strong>of</strong> Pt crystals with different sizes,<br />

starting from 1.2 (B-sample), passing 2.3 nm (D-sample) with 3.8 nm (conventional<br />

alumina supported Pt) <strong>and</strong> ending with comparison <strong>of</strong> adsorption mode <strong>of</strong> cinchonidine<br />

on microscopic Pt plane from the work <strong>of</strong> Ferri [76] <strong>and</strong> Zaera [77, 178]. In the Fig. 4-6<br />

the region below 1300 cm -1 is not accessible <strong>of</strong> IR investigation, due to the contribution<br />

<strong>of</strong> alumina background, whose FTIR spectrum can be found as an inset in the Fig. 5-5<br />

from the chapter 5.<br />

35


1<br />

Absorbance<br />

0.1<br />

2<br />

1600 1400 1200 1000 800<br />

Wavenumber, cm -1<br />

Fig. 4-6. DRIFT spectra <strong>of</strong> the cinchonidine modified conventional Pt/Al 2 O 3 (1) <strong>and</strong><br />

free cinchonidine (2), intensity <strong>of</strong> that spectrum is decreased for better presentation.<br />

Absorbance<br />

0.13<br />

1<br />

2<br />

1600 1400 1200 1000 800<br />

Wavenumber cm -1<br />

Fig. 4-7. DRIFT spectra <strong>of</strong> the cinchonidine modified Pt nanoclusters (1, sample B) <strong>and</strong><br />

free cinchonidine (2), intensity <strong>of</strong> that spectrum is decreased for better presentation.<br />

It is important to note, that the absence <strong>of</strong> Al 2 O 3 (normally used as a support for Pt<br />

[76]) allows investigation down to the range <strong>of</strong> 700-900 cm -1 thus facilitating<br />

assignment <strong>of</strong> the important out <strong>of</strong> plane vibrations <strong>of</strong> aromatic rings.<br />

36


Absorbance<br />

0.05<br />

1<br />

2<br />

1600 1400 1200 1000<br />

Wavenumber, cm -1<br />

Fig. 4-8. FTIR spectra <strong>of</strong> the cinchonidine modified Pt nanoclusters (1, sample D) <strong>and</strong><br />

free cinchonidine (2), intensity <strong>of</strong> that spectrum is decreased for better presentation.<br />

The spectra measured in transmission mode with KBr pressed pellet.<br />

The 3-9 cm -1 blue <strong>and</strong> red shifts in the wavenumber <strong>of</strong> the corresponding peaks in the<br />

spectra <strong>of</strong> free <strong>and</strong> adsorbed modifier clearly indicate on the fact that cinchonidine is<br />

adsorbed on the Pt cluster. The quality <strong>of</strong> shown spectra increases significantly with<br />

absence <strong>of</strong> a support due <strong>of</strong> the mentioned background contribution <strong>and</strong> because <strong>of</strong><br />

higher cinchonidine concentration <strong>of</strong> surface <strong>of</strong> Pt for B <strong>and</strong> D samples with respect to<br />

modified conventional Pt/Al 2 O 3 . However, despite <strong>of</strong> the fact that not all peaks in the<br />

spectrum <strong>of</strong> cinchonidine modified conventional Pt/Al 2 O 3 (Fig. 4-6) are resolved well<br />

(due to the alumina background effect) <strong>and</strong> some corresponding peaks from<br />

cinchonidine modified Pt nanoclusters (Fig. 4-7, 4-8) presence in the Fig. 4-6 as<br />

shoulders, it is possible to conclude, that the spectra <strong>of</strong> adsorbed cinchonidine on all<br />

three samples have no principal difference. And only difference is in quality <strong>of</strong> the<br />

spectra. Through comparison <strong>of</strong> all three spectra is also possible to conclude that the<br />

cinchonidine molecule is still on the Pt surface <strong>and</strong> did not decompose after samples<br />

preparation, that is especially important to note for B <strong>and</strong> D sample. Since the spectrum<br />

quality <strong>of</strong> B type <strong>of</strong> nanosized sample is better that the quality <strong>of</strong> modified conventional<br />

Pt/Al 2 O 3 <strong>and</strong> D sample, this spectrum had been chosen for further detailed analysis,<br />

where as brief analysis <strong>of</strong> two left spectra can be found below in this chapter (sample<br />

B) <strong>and</strong> in the chapter 5 (cinchonidine modified conventional Pt/Al 2 O 3 ).<br />

The DRIFT spectrum <strong>of</strong> sample B (Fig. 4-7) has many peaks <strong>and</strong> shoulders, thus it<br />

gives difficulties to find corresponding peaks in the spectrum <strong>of</strong> free cinchonidine<br />

unambiguously for all <strong>of</strong> them. This is due to the adsorption selection rule [179] by<br />

which intensities <strong>of</strong> some vibration bonds <strong>of</strong> adsorbed molecules are changed. That is<br />

why only some <strong>of</strong> the peaks <strong>and</strong> shoulders (which are well separated <strong>and</strong>/or with high<br />

intensity) were chosen for further analysis, they are summarized in Table 4-3. Some <strong>of</strong><br />

the peaks from Table 4-3 could be easily assigned to the ‘flat’ or/<strong>and</strong> ‘tilted’ orientation<br />

mode <strong>of</strong> cinchonidine by using data from the works <strong>of</strong> Ferri [76] <strong>and</strong> Zaera [77].<br />

However, not all vibrations from Table 4-3 could be correlated with prior work due to<br />

37


the focus <strong>of</strong> these prior experiments on selective regions <strong>of</strong> the spectra. In the work <strong>of</strong><br />

Ferri <strong>and</strong> co-workers the interval 1321-1635 cm -1 is particularly covered <strong>and</strong> in the<br />

work <strong>of</strong> Zaera <strong>and</strong> co-workers, the description <strong>of</strong> many peaks is given, but some <strong>of</strong><br />

them, which were observed in our experiments, are not mentioned. This why the<br />

assignment <strong>and</strong> interpretation <strong>of</strong> summarized in the Table 4-3 peaks <strong>of</strong> adsorbed<br />

cinchonidine has been performed with help <strong>of</strong> modelling <strong>of</strong> the IR vibrations <strong>of</strong> the<br />

molecule <strong>and</strong> metal adsorption selection rule [179]. The metal adsorption selection rule<br />

[179] states, if a vibration <strong>of</strong> a molecule adsorbed on a metal surface produces a dipole<br />

moment orientated perpendicularly to the surface, the intensity <strong>of</strong> this vibration is<br />

enhanced (or can be observed) <strong>and</strong> vice versa; i.e. if the dipole moment is orientated<br />

parallel to the metal surface, then intensity <strong>of</strong> the corresponding vibration is reduced (or<br />

not present). It means that dipole moments corresponding to all vibrations from Table<br />

4-3 are orientated mostly perpendicularly to the metal surface.<br />

Fig. 4-9. Illustration <strong>of</strong> the metal adsorption selection rule. Image charges are drawn<br />

dotted.<br />

It is important to note, that intensity <strong>of</strong> the IR radiation adsorbed by a molecule is<br />

proportional to the first time derivative <strong>of</strong> dipole moment in power <strong>of</strong> two, or, roughly,<br />

it is proportional to the squared dipole moment. At the same time the absolute value <strong>of</strong><br />

electric field induced by any dipole moment is inversely proportional to the distance in<br />

power <strong>of</strong> three. Thus contribution <strong>of</strong> the electric field from dipole moment <strong>of</strong> the<br />

molecule adsorbed on the neighbor nanoparticle is significantly less (approximately in<br />

the factor <strong>of</strong> 10 3 ) that electric field from dipole moment <strong>of</strong> the molecule adsorbed on<br />

the current nanoparticle. Here we took into account, that size <strong>of</strong> the molecule is ~ 10-20<br />

% <strong>of</strong> the size <strong>of</strong> nanocluster <strong>and</strong> typical distance between nanoparticles (in the solid<br />

state) can be estimated as a size <strong>of</strong> nanoparticle. Since, an intensity <strong>of</strong> any<br />

electromagnetic field is proportional to the squared amplitude <strong>of</strong> electric field, we can<br />

conclude, that observed IR peaks belong to the molecule adsorbed on the current metal<br />

nanoparticle only.<br />

In the case <strong>of</strong> cinchonidine molecule, its optimized geometry was chosen in the open 3<br />

conformation [77] <strong>and</strong> vibration modes with corresponding dipole moments orientation<br />

were found from DFT calculation (with correlation coefficient dω theory /dω exp = 1.02).<br />

Then every observed vibrational frequency from Table 4-3 was analyzed with respect<br />

to the metal adsorption selection rule. The analysis results are also summarized in the<br />

38


Table 4-3 <strong>and</strong> the ‘flat’ <strong>and</strong> ‘tilted’ modes <strong>of</strong> adsorbed cinchonidine are indicated by a<br />

vibration b<strong>and</strong> e.g. at 768 cm -1 <strong>and</strong> at 1514 cm -1 are illustrated in the Fig. 4-10 <strong>and</strong> Fig.<br />

4-11, respectively.<br />

Fig. 4-10. Flat adsorbed cinchonidine demonstrates vibration mode at 768 cm -1 with<br />

dipole moment (big arrow) orientated mostly perpendicularly to the metal surface. The<br />

most intensive displacement <strong>of</strong> atoms is shown by small arrows.<br />

39


Fig. 4-11. Tilted adsorbed cinchonidine demonstrates vibration mode at 1514 cm -1 with<br />

dipole moment (big arrow) orientated mostly perpendicularly to the metal surface. The<br />

most intensive displacement <strong>of</strong> atoms is shown by small arrows.<br />

It should be noted, that observed “blue <strong>and</strong> red shifts” in frequencies (see Table 4-3)<br />

indicate the changes in the bond force constants <strong>and</strong> thus changes in the skeletal<br />

geometry <strong>of</strong> an adsorbed molecule with respect to the free species. However, for the<br />

purpose <strong>of</strong> this work, we will leave these considerations aside, since they do not hinder<br />

determination <strong>of</strong> the orientation <strong>of</strong> adsorbed cinchonidine <strong>and</strong> other used molecules<br />

(see below). Calculated (correlation coefficient 1.06) <strong>and</strong> measured IR spectra <strong>of</strong> free<br />

quinoline (mostly medium <strong>and</strong> strong peaks were chosen) were used as a particular<br />

reference for facilitating assignment <strong>of</strong> the spectra <strong>of</strong> the cinchonidine (as well as<br />

quiphos with aniline from the next chapter), Fig. 4-12.<br />

3<br />

A<br />

Relative units<br />

2<br />

B<br />

1<br />

C<br />

0<br />

D<br />

1600 1400 1200 1000 800 600<br />

Wavenumber cm -1<br />

Fig. 4-12 Comparison <strong>of</strong> IR peaks between free cinchonidine (A), quinoline (B),<br />

quiphos (C) <strong>and</strong> aniline (D).<br />

Table 4-3. Assignment <strong>of</strong> some frequencies (cm -1 ) from IR spectrum <strong>of</strong> adsorbed <strong>and</strong><br />

free cinchonidine <strong>and</strong> free quinoline.<br />

Cinchonidine<br />

on Pt<br />

Free<br />

cinchonidine<br />

Quinoline Cinchonidine<br />

theoretical<br />

Vibration<br />

description<br />

Orientation<br />

<strong>of</strong> adsorbed<br />

cinchonidine<br />

1634 (sh) 1636 (med) - 1709<br />

19,20 C strch. Flat<br />

1613 (sh) 1619 (wk3) 1620<br />

(med)<br />

1594 (med) 1590 (str4) 1596<br />

(med)<br />

1670 C-Qu.<br />

strch., H-<br />

Qu. b.<br />

1643 C-Qu.<br />

strch., H-<br />

Qu. b.<br />

Tilted<br />

Tilted<br />

40


1576 (sh) 1567 (med) 1571<br />

(med)<br />

1622 C-Qu.<br />

strch., H-<br />

Qu. b.<br />

1514 (str) 1507 (str) 1502 (str) 1556 C-Qu.<br />

strch., H-<br />

Qu. b.<br />

1463 (str) 1452 (str) - 1494<br />

1427 (sh) 1420 (med) 1431 (wk) 1445<br />

15,17,22 CH 2 -<br />

QL. sci.<br />

2,3 C strch.,<br />

C-H-Qu.,<br />

ip. def.,<br />

15 CH 2 ,<br />

HO 2 C sci.,<br />

H 2 C b.<br />

1208 (med) 1207 (med) - 1240 C-H-Qu.<br />

ip. b., 2,3 C<br />

strch., HO,<br />

2 CH b.,<br />

QL. comp.<br />

1179 (wk) 1180 (wk) - 1223<br />

1168 (wk) 1163 (wk) - 1189<br />

30 H-Qu.,<br />

23,24 H b.,<br />

H- 17 C-Ql.<br />

comp.<br />

25,26,30 H-<br />

Qu. ip. b.,<br />

23,24 H b.,<br />

QL. comp.<br />

1145 (wk) 1133 (wk) - 1143 H-Qu. ip.<br />

b.,<br />

23,24 H<br />

b., Ql.<br />

comp.<br />

1118 (med) 1099 (wk) - 1128 H-Qu. ip.<br />

b., O 2 C<br />

strch., QL.<br />

comp.<br />

856 (med) 862 (wk) - 873<br />

29,10 H-Qu.<br />

oop. wag.,<br />

QL. comp.<br />

831 (med) 823 (med) 807 (str) 843 H-Qu. oop.<br />

wag, H 20 2 C<br />

sci., QL.<br />

Comp.<br />

809 (med) 803 (med) 787 (str) 824 H-Qu. oop.<br />

Wag.,<br />

40,41 H 20 C<br />

sci., QL.<br />

comp.<br />

782 (str) 778 (wk) 761 (wk) 791 H-Qu. oop.<br />

wag.,<br />

40,41 H 20 C<br />

Tilted , flat<br />

Tilted<br />

Tilted<br />

Flat<br />

Tilted<br />

Flat<br />

Flat, tilted<br />

Flat<br />

Flat<br />

Flat<br />

Flat<br />

Flat<br />

Flat<br />

41


sci., QL.<br />

comp.<br />

768 (str) 756 (str) 740 (str) 775<br />

25,26,27,28 H- Flat<br />

Qu. oop.<br />

wag.<br />

Comments: sh. – shoulder, med. – medium, wk – weak, st. – strong., 15C –carbon atom<br />

number 15. Qu – quinoline, QL – quinuclidine. C-Qu. – only carbons in quinoline, H-<br />

Qu. – only hydrogens in quinoline, skel. – skeleton, comp. – complex: more then one<br />

type <strong>of</strong> vibrations (stretching, wagging, rocking, <strong>and</strong> scissoring) are present, def. –<br />

deformation, C-H-QL. – carbons <strong>and</strong> hydrogens in quinuclidine, b. – bending, strch. –<br />

stretching, sci. – scissoring, wag. – wagging, ip.- in plane, oop. – out <strong>of</strong> plane vibration.<br />

As it is seen from the Table 4-3, cinchonidine on Pt nanoclusters was found in both the<br />

‘flat’ <strong>and</strong> ‘tilted’ adsorption modes. Here, we did not differentiate between the two<br />

types <strong>of</strong> tilted adsorption modes, the so called “N-lone pair bonded” <strong>and</strong> “α-H<br />

abstracted" because we found assignment <strong>of</strong> the corresponding species is somewhat<br />

ambiguous <strong>and</strong> since the flat adsorption mode is more determinant in enantioselective<br />

catalysis.<br />

It is very important to compare the adsorbed modes <strong>of</strong> cinchonidine on macroscopic<br />

<strong>and</strong> nanocluster surfaces <strong>of</strong> Pt. In fact, through comparison <strong>of</strong> IR investigations <strong>of</strong><br />

cinchonidine adsorbed on bulk Pt crystals (ATR, RAIRS) <strong>and</strong> on Pt nanoclusters<br />

(DRIFTS) we found no principal difference in the spectra as well in the spectra <strong>of</strong><br />

sample D <strong>and</strong> cinchonidine modified conventional Pt/Al 2 O 3 (the only difference we<br />

found was in the quality <strong>of</strong> the spectra). However, in the spectra <strong>of</strong> cinchonidine on Pt<br />

nanoclusters small shifts <strong>of</strong> the spectra from free cinchonidine were found <strong>and</strong> between<br />

the spectra <strong>of</strong> cinchonidine adsorbed on Pt. For example: in case <strong>of</strong> free cinchonidine,<br />

the peak at 1593 cm -1 in work <strong>of</strong> Ferri [76] was observed at 1590 cm -1 in the current<br />

thesis <strong>and</strong> 1591 cm -1 in the work <strong>of</strong> Zaera [77]; in case <strong>of</strong> cinchonidine on Pt, the peak<br />

at 1458 cm -1 in work <strong>of</strong> Ferri was found at 1463 cm -1 in the current work <strong>and</strong> at 1463<br />

cm -1 in the work <strong>of</strong> Zaera. It seems most likely that these small shifts are due to the role<br />

<strong>of</strong> solvents/purity <strong>of</strong> cinchonidine used <strong>and</strong> other secondary effects. Another difference<br />

is that the spectra <strong>of</strong> the lig<strong>and</strong>s on nanoclusters contains less noise <strong>and</strong>, mostly, have<br />

better quality with respect to the spectra <strong>of</strong> lig<strong>and</strong>s on bulk metal. The reason for this<br />

may be due to the high surface area <strong>of</strong> nanoclusters <strong>and</strong> thus higher concentration <strong>of</strong><br />

adsorbed lig<strong>and</strong> in the IR beam. If size <strong>of</strong> the cinchonidine supported Pt nanoclusters<br />

(2.3 nm sample D, or 4 nm conventional Pt/Al 2 O 3 ) is bigger than typical size <strong>of</strong><br />

molecule (about 0.5 nm, the length <strong>of</strong> the quinoline anchor), the support can be<br />

considered as a plane <strong>and</strong> molecule behaves similarly as to being adsorbed on a<br />

continuous surface <strong>of</strong> a (bulk) metal crystal.<br />

However, if the size <strong>of</strong> the clusters becomes comparable with the size <strong>of</strong> the molecule,<br />

as we have in case <strong>of</strong> cinchonidine on Pt (1.2 nm in sample B) the lig<strong>and</strong> might behave<br />

differently (especially in terms <strong>of</strong> enantioselectivity) <strong>and</strong> would be expected to<br />

experience less steric hindrance due to the curvature <strong>of</strong> the surface.<br />

In fact, Yang <strong>and</strong> co-authors [106] reported enantiomeric excess in hydrogenation <strong>of</strong><br />

methyl pyruvate to be 90 % over 1.2 nm Pt cluster <strong>and</strong> 85 % over 3.4 nm under 40 bar<br />

<strong>of</strong> H 2 pressure. They explained this difference as a size effect resulting from<br />

cinchonidine <strong>and</strong> methyl pyruvate forming a half-hydrogenated complex which is<br />

adsorbed on different crystal faces. However, these authors did not focus on stability <strong>of</strong><br />

Pt nanoclusters during hydrogenation <strong>and</strong> changes in size, especially after removal <strong>of</strong><br />

42


PVP stabilizer. In the work <strong>of</strong> Bönnemann [114, 115] , no such dependence was found<br />

(ee = 76% over 1.5 nm Pt cluster <strong>and</strong> 78 % over 3.9 nm, under 1 bar <strong>of</strong> H 2 pressure).<br />

In our opinion, one <strong>of</strong> the primary effects causing enhanced ee is a higher concentration<br />

<strong>of</strong> active chiral sites (“π-bonded” cinchonidine) on the catalyst surface. In the work <strong>of</strong><br />

Yang <strong>and</strong> co-authors cinchonidine adsorbs on the Pt surface from solution (optimum<br />

concentration <strong>of</strong> cinchonidine [80, 102, 180]) whereas high hydrogen pressure (40 bar)<br />

promotes a surface cleaning (from PVP, AcOH <strong>and</strong> etc.) for obtaining more space for<br />

flat adsorbed cinchonidine ( cinchonidine in any tilted mode needs, definitely, less free<br />

space for the adsorption). The catalyst prepared according to Bönnemann’s procedure<br />

(sample B) possesses adsorbed cinchonidine in flat <strong>and</strong> in two tilted modes as<br />

demonstrated in this manuscript <strong>and</strong> a high pressure <strong>of</strong> hydrogen was not applied.<br />

Moreover ee=94-95 % was obtained under high (100 bar) H 2 pressure with<br />

conventional 5R94 [180] <strong>and</strong> Engelhard 4759 [181]- Pt/Al 2 O 3 catalysts (average<br />

particles size 3.6 nm [182]) where the above mentioned size effect seems to be less<br />

probable.<br />

Based on these facts, we would like to conclude here, that it was found that the size <strong>of</strong><br />

Pt clusters (in range 1.4- ∞ nm) has no observed influence on the adsorption mode <strong>of</strong><br />

the cinchonidine (at least in case <strong>of</strong> sample B (1.4 nm), sample D (2.3 nm),<br />

conventional Pt/Al 2 O 3 (4 nm) <strong>and</strong> macroscopic samples from work <strong>of</strong> Zaera [77] <strong>and</strong><br />

Ferri [76]). This is important because it has been reported in the literature that cluster<br />

size does influence both catalytic activity <strong>and</strong> the observed ee [106].<br />

Thus, it was found that both ‘flat’ <strong>and</strong> ‘tilted’ modes <strong>of</strong> adsorbed cinchonidine are<br />

present on Pt nanoclusters <strong>and</strong> could be an explanation <strong>of</strong> the fact that this type <strong>of</strong><br />

catalyst does not provide the maximum possible enantiomeric excess (99 %).<br />

4.3.3 Investigation <strong>of</strong> catalytic behaviour <strong>of</strong> cinchonidine<br />

modified Pt nanoclusters<br />

Investigation <strong>of</strong> catalytic behavior <strong>of</strong> cinchonidine modified Pt nanoclusters was<br />

studied with B, D <strong>and</strong> conventional modified Pt/Al 2 O 3 in the hydrogenation <strong>of</strong> ethyl<br />

pyruvate.<br />

Initial transient period<br />

Under all reaction conditions, summarized in Table 4-4, enantiomeric excess was found<br />

to depend on the reaction time in a manner similar to that shown in the Fig. 4-13.<br />

During the first moments <strong>of</strong> hydrogenation (the period <strong>of</strong> time termed the initial<br />

transient period) enantiomeric excess increases to a maximum value (ee max ) <strong>and</strong> then<br />

decreases until the entire amount <strong>of</strong> ethyl pyruvate is converted in the reaction mixture.<br />

43


Conversion, %<br />

100<br />

80<br />

60<br />

40<br />

ITP time<br />

ee max<br />

80<br />

70<br />

Actual enantiomeric excess, %<br />

20<br />

0<br />

Y =83.7-0.19 X<br />

60<br />

0 25 50 Time, min75 100 125<br />

Fig. 4-13. Typical catalytic behaviour <strong>of</strong> cinchonidine modified Pt nanoclusters during<br />

ethyl pyruvate hydrogenation.<br />

Effect <strong>of</strong> hydrogen pressure<br />

The effect (Exp. 1 <strong>and</strong> 2, see Table 4-4) <strong>of</strong> changing the hydrogen pressure in the<br />

reaction system is shown in the Fig. 4-14 (other graphs are similar, thus they are not<br />

shown in further, important features are mentioned in the text, in the corresponding<br />

paragraphs <strong>and</strong> in the Table 4-4). With an increase <strong>of</strong> hydrogen pressure from 2 to 10<br />

bar the initial rate increases by a factor <strong>of</strong> ten from 14 min -1 to 140 min -1 An increase <strong>of</strong><br />

initial rate could be qualitatively explained as an increase <strong>of</strong> hydrogen migration from<br />

the gas to the liquid phase, but the fact that the maximum enantiomeric excess reached<br />

significantly increases from 36 % to 78 % seems to indicate a specific hydrogen –<br />

catalyst interaction, as a consequence <strong>of</strong> this, the catalyst demonstrates increasing<br />

enantioselectivity. A decrease by 4 fold in the time required to reach ee max is observed<br />

by changing the pressure from 2 (110 min) to 10 (30 min) bar.<br />

44


80<br />

Conversion, %<br />

100<br />

80<br />

60<br />

40<br />

Y =83 - 0.19 X<br />

Hydrogen pressure<br />

10 bar<br />

2 bar<br />

70<br />

60<br />

50<br />

40<br />

Actual enantiomeric excess, %<br />

20<br />

30<br />

0<br />

20<br />

0 50 100 150 200 250 300<br />

Time, min<br />

Y =44 - 0.07 X<br />

Figure 4-14. Effect <strong>of</strong> hydrogen pressure.<br />

Experiment 1: 5 mg (13 μmol Pt) D- type <strong>of</strong> catalyst, [EP 0 ] =2.571 mol l -1 , no free<br />

cinchonidine, P H2 = 2 bar. Full squares (left axis) – ethyl pyruvate conversion, crossed<br />

squares (right axis) – actual EE, line – linear approximation <strong>of</strong> actual ee decrease.<br />

Experiment 2: 4 mg (13 μmol Pt) D- type <strong>of</strong> catalyst, [EP 0 ] = 2.571 mol l -1 , no free<br />

cinchonidine, P H2 = 10 bar. Full triangles (left axis) – ethyl pyruvate conversion,<br />

crossed triangles (right axis) – actual ee, line – linear approximation <strong>of</strong> actual EE<br />

decrease.<br />

Type <strong>of</strong><br />

[EP ],<br />

Table 4-4. Details <strong>of</strong> catalysts <strong>and</strong> reaction parameters.<br />

Exp. P H2 , Pt surf.<br />

ee * max, Cinchonidine, Time slope Initial<br />

0<br />

N. bar Atoms, catalyst % concentration, ee * max, min -1 mol l - rate b ,<br />

μmol a <strong>and</strong><br />

mmol l -1<br />

1<br />

min<br />

min -1<br />

loading,<br />

mg<br />

1 2 5.8 D, 4 36 0 110 0.07 2.571 14<br />

2 10 5.8 D, 4 78 0 30 0.07 2.571 140<br />

3 c 10 5.8 D, 4 76 0 20 0.19 0.514 99<br />

4 c 10 29 D, 20 65 0 9 0.22 0.514 37<br />

5A c 10 1.11 Pt/Al 2 O 3 , 76 1 - 0.24 0.257 128<br />

20<br />

5B c 10 0.56 Pt/Al 2 O 3 ,<br />

10<br />

76 1 - 0.21 0.257 132<br />

45


6A c 10 19.5 B, 20 61 0 3 0.18 0.514 24<br />

6B c 10 3.9 B, 4 76 0 - - 0.514 -<br />

7A d 10 2.2 D, 1.5 79 1 30 - 0.514 e 130<br />

7B 10 2.2 D, 1.5 78 1 30 - 0.514 120<br />

8 c 10 5.8 D, 4 76 1 55 0.04 0.514 121<br />

9 c 10 3.9 B, 4 74 1 34 0.07 0.514 42<br />

Comments: a – The number <strong>of</strong> Pt surface atoms was estimated from the medium<br />

particles size according to “full shell” model <strong>of</strong> Pt nanoclusters <strong>and</strong> from chemisorption<br />

on conventional Pt/Al 2 O 3 , b – Initial rate values [min -1 ] were calculated as<br />

normalization a reaction rate [mmol min -1 ] on number <strong>of</strong> Pt surface atoms [mmol] in<br />

the reaction mixture, c – The graph is not shown; important parameters <strong>of</strong> reaction are<br />

presented, d – The purified EP was used.<br />

Effect <strong>of</strong> catalyst loading<br />

The decrease <strong>of</strong> initial rate from 99 min -1 to 37 min -1 with an increase <strong>of</strong> catalyst<br />

loading from 4 mg (13 μmol Pt) to 20 mg (65 μmol Pt) was found in Exp. 3 <strong>and</strong> 4.<br />

Also, the maximum <strong>of</strong> actual enantiomeric excess (ee* max ) has also decreased from 76<br />

% to 65 %. A possible explanation for this is that the time necessary to reach the<br />

maximum enantioselective activation level may not be the same for various catalyst<br />

loadings <strong>and</strong> the full conversion <strong>of</strong> ethyl pyruvate has already been accomplished. In<br />

other words, the number <strong>of</strong> chiral sites (π-bonded cinchonidine) did not reach the<br />

maximum possible value under the 10 bars <strong>of</strong> hydrogen pressure. It should be noted,<br />

that this effect can not be explained as mass transport phenomena in liquid phase, since<br />

according to Minder, Baiker <strong>and</strong> co-workers [183] the kinetically controlled region lays<br />

below 15 g l -1 <strong>of</strong> catalyst concentration while in the Exp. 3 <strong>and</strong> 4 the maximum catalyst<br />

concentration is 0.28 g l -1 . In fact, we performed the reactions with 20 mg (Exp. 5A,<br />

ee max = 76 %, initial rate = 128 min -1 ) <strong>and</strong> 10 mg (Exp. 5B, ee max = 76 %, initial rate =<br />

132 min -1 ) with conventional Pt/Al 2 O 3 <strong>and</strong> concluded that in our experiments mass<br />

transport effect in liquid phase does not limit the reaction. The rate <strong>of</strong> migration <strong>of</strong><br />

hydrogen from gas to the liquid phases was the same, since both experiments were<br />

performed under the same hydrogen pressure (10 bar).<br />

A similar decrease <strong>of</strong> ee max with an increase <strong>of</strong> catalyst loading was found for the B<br />

type catalyst (see Exp. 6A <strong>and</strong> 6B in Table 4-4).<br />

By comparing these results with those reviewed by Studer, Blaser <strong>and</strong> Exner in 2003<br />

[44] we can see that there are intrinsic differences between the colloids stabilized by<br />

chiral lig<strong>and</strong>s <strong>and</strong> the conventional Pt/Al 2 O 3 heterogeneous catalysts which are<br />

unmodified prior to introduction to the reaction vessel. In fact, decrease <strong>of</strong> initial rate<br />

<strong>and</strong> maximum enantiomeric excess with an increase <strong>of</strong> catalyst loading are in<br />

contradiction to the results observed for the conventional Pt/Al 2 O 3 system which may<br />

be attributed to the fact, that our catalysts have adsorbed cinchonidine already on them,<br />

whereas the catalyst systems reviewed in the literature are exposed to cinchonidine<br />

during the reaction (in situ).<br />

Role <strong>of</strong> the ethyl pyruvate in nature <strong>of</strong> ITP<br />

The ITP time was found to be almost independent <strong>of</strong> the initial concentration <strong>of</strong> ethyl<br />

pyruvate (Exp. 2 <strong>and</strong> 3). The ITP was found to be 20-30 min in Exp. 2 (20 ml EP) <strong>and</strong><br />

20-25 min in the case <strong>of</strong> Exp. 3 (4 ml EP). Ee* max is almost the same value (76-78 %)<br />

in both experiments, but initial rate increased by ≈ 30 % with an increase <strong>of</strong> the initial<br />

concentration <strong>of</strong> ethyl pyruvate.<br />

46


In order to clarify the role in ee <strong>of</strong> the initial presence <strong>of</strong> ethyl lactate (0.5 % each S <strong>and</strong><br />

R isomers) as an impurity in EP, the catalytic hydrogenation reaction with the D type <strong>of</strong><br />

catalyst was performed with both the unpurified <strong>and</strong> purified EP. Fig. 4-15 (Exp. 7A,<br />

7B) shows that in this experiment the ee observed is similar for both the purified <strong>and</strong><br />

unpurified experiments with respect to the reaction time <strong>and</strong> both demonstrate the same<br />

ITP = 30 min <strong>and</strong> initial rate = 120-130 min -1 .<br />

80<br />

Conv. 47 %<br />

Enantiomeric excess, %<br />

76<br />

72<br />

Purified EP<br />

Normal EP<br />

Conv. 4 %<br />

68<br />

0 20 40 60 80<br />

Time, min<br />

Fig. 4-15. Role <strong>of</strong> the impurities in the ethyl pyruvate on EE.<br />

Experiment 7A: 1.5 mg (4.8 μmol Pt) D type <strong>of</strong> catalyst, [EP 0 ] = 0.514 mol l -1<br />

(purified), [Cinchonidine 0 ] =1 mmol l -1 , P H2 = 10 bar.<br />

Experiments 7B: 1.5 mg (4.8 μmol Pt) D type <strong>of</strong> catalyst, [EP 0 ] = 0.514 mol l -1 ,<br />

[Cinchonidine 0 ] =1 mmol l -1 , P H2 = 10 bar.<br />

However, the initial presence <strong>of</strong> R, S ethyl lactates gives a systematic shift in ee. This<br />

shift has a maximum value (≈ 3 %) in initial times (at low conversion) <strong>and</strong> decreases<br />

with time (with increase <strong>of</strong> conversion). Taking into account the initial concentration <strong>of</strong><br />

both ethyl lactates, the values <strong>of</strong> the shift lie within the estimated limit.<br />

In order to show that the presence <strong>of</strong> EP does not play a significant role in increasing<br />

ee, for example as a reagent in side reactions (aldol condensation) reaction [184] we<br />

performed a blank experiment in the absence <strong>of</strong> EP with free cinchonidine under<br />

reaction conditions (acetic acid, 10 bar <strong>of</strong> H 2 ). However, both B <strong>and</strong> D samples are<br />

unstable with prolonged hydrogen exposure especially at high pressure. Thus, to<br />

improve stability, we prepared cinchonidine modified Pt nanoclusters deposited on<br />

alumina. This catalyst was found to be stable under the reaction conditions for a long<br />

time <strong>and</strong> demonstrates (P H2 = 10 bar, [EP 0 ] = 0.514 mol l -1 , [cinchonidine] = 1 mmol l -<br />

1 ) ee = 79 %, however, after pretreating the catalyst for 16 hours in acetic acid it<br />

demonstrates ee = 84 %. This experiment obviously shows that the ethyl pyruvate does<br />

not influence significantly to the increase <strong>of</strong> ee after exposing catalyst under reaction<br />

condition. Another interesting fact is that under optimized conditions (hydrogen<br />

47


pressure, exposing time, cinchonidine concentration, etc.) the maximum obtained ee<br />

was found to be 88 %.<br />

In order to keep clarity in description <strong>of</strong> the samples <strong>and</strong> their characterization, it is<br />

recommended to the reader to see Chapter 5 for further information about cinchonidine<br />

modified Pt nanoclusters deposited on alumina, where we give detailed models <strong>of</strong> effect<br />

<strong>of</strong> ee increasing, whereas here we would like to continue description <strong>of</strong> observed ITP<br />

effect.<br />

Role <strong>of</strong> free cinchonidine in ITP<br />

The addition <strong>of</strong> 20 mg (0.068 mmol) <strong>of</strong> free cinchonidine into the reaction mixture<br />

(Exp. 3 <strong>and</strong> 8) increases ITP by a factor <strong>of</strong> almost three (from 20 min to 55 min) while<br />

ee* max ≈ 76 % remained nearly constant <strong>and</strong> the initial rate increases only slightly. The<br />

stability in ee with addition <strong>of</strong> free cinchonidine was also found for B type <strong>of</strong> catalyst in<br />

out experiments <strong>and</strong> in original work <strong>of</strong> Bönnemann [114]. This fact might be<br />

explained by competitive formation <strong>and</strong> disappearance <strong>of</strong> new chiral sites, as is<br />

discussed below.<br />

Post ITP decrease <strong>of</strong> enantiomeric excess<br />

The decrease <strong>of</strong> ee during the hydrogenation <strong>of</strong> ethyl pyruvate has been previously<br />

observed both at high (40-100 bar) [50, 185] <strong>and</strong> low (~1 bar) [186] pressures <strong>of</strong><br />

hydrogen <strong>and</strong> has been shown to result from the partial hydrogenation <strong>of</strong> the quinoline<br />

(anchor) portion <strong>of</strong> the cinchonidine. The partially hydrogenated quinoline ring is<br />

unable to function as an anchor any longer due to the loss <strong>of</strong> the π-aromaticity <strong>and</strong> as a<br />

result, the cinchonidine desorbs from the surface resulting in an unmodified catalyst<br />

which is not capable <strong>of</strong> inducing enantioselectivity. Our results support the possibility<br />

that this effect also takes place with ‘quasi-homogeneous’ catalysts in the 2-10 bar<br />

range.<br />

To the best <strong>of</strong> our underst<strong>and</strong>ing <strong>and</strong> characterization <strong>of</strong> the effect, the decay <strong>of</strong> the<br />

actual ee after ITP was linearly approximated <strong>and</strong> the absolute value <strong>of</strong> the line’s slope<br />

was used to determine the stability <strong>of</strong> catalyst enantioselectivity with respect to the<br />

catalyst type <strong>and</strong> reaction conditions. Thus, in experiments with the same initial<br />

concentration <strong>of</strong> ethyl pyruvate we can easily see where no additional cinchonidine was<br />

added (Exp. 3, 4, slope ~ 0.18-0.22 min -1 ) the rate <strong>of</strong> decrease <strong>of</strong> actual enantiomeric<br />

excess was observed to be 3-4.5 times greater than with (Exp. 8, 9) free (20 mg or 1<br />

mmol l -1 ) cinchonidine (slope ~ 0.04-0.07 min -1 ).<br />

Qualitatively, this effect can be explained by maintaining a more or less constant π-<br />

bonded cinchonidine concentration on Pt surface due to the replacement <strong>of</strong> “old”<br />

molecules (with partially hydrogenated quinoline anchor) by “new” molecules <strong>of</strong><br />

cinchonidine from solution.<br />

Comparison <strong>of</strong> D, B ‘quasi-homogeneous’ <strong>and</strong> conventional<br />

heterogeneous catalysts<br />

Comparison <strong>of</strong> catalytic behavior<br />

The important principal difference in catalytic behavior <strong>of</strong> D, B <strong>and</strong> conventional<br />

Pt/Al 2 O 3 catalyst systems is the negligibly small increase <strong>of</strong> actual (non-cumulative)<br />

enantiomeric excess during the reaction in the case <strong>of</strong> the previously non-modified<br />

48


Pt/Al 2 O 3 (Exp. 5A). This increase <strong>of</strong> enantiomeric excess was found to be significantly<br />

smaller than in the series <strong>of</strong> experiments with pre-modified catalysts. This indicates the<br />

ITP effect is nearly non-existent for the systems which are modified in situ under our<br />

reaction conditions. That the ITP effect is hardly observed in acetic acid for catalysts<br />

without cinchonidine pre-modification is in agreement with the findings <strong>of</strong> Balazsik<br />

<strong>and</strong> Bartok [169]. In our system, the conventional Pt/Al 2 O 3 shows ee* max = 76 % which<br />

is the same value as was found in Exp. 3 <strong>and</strong> 8 with the D type <strong>of</strong> catalysts <strong>and</strong> also<br />

demonstrated similar values <strong>of</strong> initial rate. However, the rate <strong>of</strong> decrease <strong>of</strong> the actual<br />

enantiomeric excess is 3-6 times higher for the conventional catalyst than for the B <strong>and</strong><br />

D types <strong>of</strong> catalyst with the presence <strong>of</strong> the same amount <strong>of</strong> free cinchonidine in the<br />

reaction mixture (Exp. 8 <strong>and</strong> 9). Also, the D type <strong>of</strong> catalyst gave slightly higher<br />

maximum enantiomeric excess <strong>and</strong> initial rate than the B type under the same reaction<br />

conditions (compare Exp. 3 with 6B, Exp. 4 with 6A <strong>and</strong> Exp. 8 with 9) possibly due to<br />

the different preparation methods (used solvents, washing procedure, preparation<br />

temperature, etc.).<br />

The two-cycle mechanism proposed by Blaser, Garl<strong>and</strong> <strong>and</strong> Jallet [81, 180], based on<br />

the assumption that the reaction goes with [R]/[S] = 1 on unmodified sites <strong>and</strong> with<br />

[R]/([R]+[S]) = s on modified sites (near flat adsorbed cinchonidine), explains the<br />

observed correlation between ee max <strong>and</strong> initial rate (R) as ee max = A-B*R -1 , where<br />

A=(2s-1)k m /(k m -k u ), B=A*k u , k u <strong>and</strong> k m are pseudo-first-order rate constants <strong>of</strong><br />

unmodified <strong>and</strong> modified cycles. In fact, the graph ee max vs. R -1 (Fig. 4-16) shows a<br />

good linear correlation for all our experiments. This means that on modified sites the<br />

ratio [R]/[S]=(1/s)-1 is the same for all catalysts <strong>and</strong> does not depend (within the<br />

margin <strong>of</strong> error) on reaction conditions, this seems reasonable because the [R]/[S] ratio<br />

is determined by the nature <strong>of</strong> the interactions between molecules <strong>of</strong> cinchonidine, ethyl<br />

pyruvate <strong>and</strong> hydrogen on Pt surface. However the ratio <strong>of</strong> modified/unmodified sites<br />

might strongly depend on experimental conditions, hydrogen pressure (cleanliness <strong>of</strong><br />

the surface), type <strong>of</strong> catalyst, concentration <strong>of</strong> modifier, etc.<br />

80<br />

ee max<br />

, %<br />

60<br />

ee max<br />

= A - B*R -1<br />

40<br />

0.00 0.02 0.04 0.06 0.08<br />

Rate -1 , min.<br />

49


Fig. 4-16. Linear relationship between maximum reached enantiomeric excess <strong>and</strong><br />

inversed initial rate for Exp.1-9.<br />

Comparison <strong>of</strong> the stability <strong>of</strong> samples<br />

It is obvious that for the high performance, long activity <strong>and</strong> reuse <strong>of</strong> the catalyst the<br />

stability <strong>of</strong> its structure is very important aspect to focus on. Here we qualitatively<br />

compare stability <strong>of</strong> B type <strong>of</strong> colloidal sample before (Fig. 4-16-1) <strong>and</strong> after (Fig. 4-<br />

16-2) ethyl pyruvate hydrogenation. Precipitation <strong>of</strong> Pt particles due to the<br />

agglomeration says about low stability <strong>of</strong> the sample under exposure to hydrogen for a<br />

long time (approximately 2-3 hours, 10 bars <strong>of</strong> H 2 , no free cinchonidine). It should be<br />

notes, that sample can be kept in colloidal form in acetic acid solution for long (~1<br />

year) time if no hydrogen is applied.<br />

Fig. 4-17. Soluble (stable) cinchonidine modifier colloidal Pt nanoparticles in acetic<br />

acid before hydrogenation reaction (1) <strong>and</strong> precipitated (2) after exposure under 10 bars<br />

<strong>of</strong> hydrogen (in AcOH) for 2 hours.<br />

IR investigation<br />

The orientation <strong>of</strong> the cinchonidine species on Pt during the reaction was investigated<br />

by measuring the IR spectra <strong>of</strong> D type catalyst prior to <strong>and</strong> after 50 min <strong>and</strong> 24 hours<br />

under reaction conditions, at 10 bar. After 50 min, the reaction in Exp. 4 (P H2 = 10 bar,<br />

[cinchonidine] = 0 ml l -1 , [EP] = 0.514 mmol l -1 ) was terminated, the hydrogen pressure<br />

was released <strong>and</strong> the reactor opened to ambient conditions. After 1 hour the colloidal<br />

catalyst precipitated <strong>and</strong> a black powder was collected, washed three times with acetic<br />

acid, dried at 30 ºC under vacuum <strong>and</strong> finally was used for IR investigation. To insure<br />

that the washing procedure does not influence results, a control experiment was<br />

conducted both before <strong>and</strong> after washing B <strong>and</strong> D type catalysts (without exposure to<br />

the catalytic reaction conditions). Because the spectra revealed the presence <strong>of</strong> both the<br />

tilted <strong>and</strong> flat orientations <strong>of</strong> cinchonidine for all samples it was determined that the<br />

washing procedure does not influence the results.<br />

The reaction in Exp. 4 was repeated <strong>and</strong> when finished, the reaction mixture was kept<br />

in the reactor under 10 bar <strong>of</strong> hydrogen pressure for 24 hours after which the colloid<br />

solution formed a black precipitate that was collected, washed three times with acetic<br />

acid <strong>and</strong> dried at 30 ºC under vacuum <strong>and</strong> finally was used for IR investigation.<br />

For better visualization, intensity <strong>of</strong> the spectra (Fig. 4-18) <strong>of</strong> free cinchonidine was<br />

decreased. Intensities <strong>of</strong> the spectra <strong>of</strong> the D type catalyst are presented as measured.<br />

To aid in spectral analysis, the observed vibration frequencies are summarized in Table<br />

4-4.<br />

50


0.16<br />

0.12<br />

1<br />

1 Free cinchondine<br />

2 Before reaction<br />

3 After 1 h<br />

4 After 24 h<br />

Absorbance<br />

0.08<br />

2<br />

0.04<br />

3<br />

4<br />

0.00<br />

1700 1600 1500 1400 1300 1200 1100 1000<br />

Wavenumber, cm -1<br />

Fig. 4-18. FTIR spectra <strong>of</strong> the D type <strong>of</strong> catalyst at different reaction time <strong>and</strong> spectra<br />

<strong>of</strong> free cinchonidine.<br />

Table 4-4. Vibration b<strong>and</strong> assignment (cm -1 ) <strong>of</strong> the cinchonidine spectra <strong>of</strong> the free<br />

cinchonidine, <strong>and</strong> D type <strong>of</strong> catalyst before reaction, at 50 min <strong>and</strong> after 24 h under<br />

reaction condition.<br />

Free<br />

Before reaction At 50 min At 24 h Assignment a<br />

cinchonidine<br />

1162 1152 1152 1152 F: 1<br />

1207 1209 1209 1209 T: 2, 3<br />

1383 1383 - 1382 F: 1, T: 2, 3,<br />

acetic acid<br />

1420 1402 1402 - F: 1, acetic acid<br />

1451 1456 - - T: 2, 3<br />

1507 1507 - - T: 3<br />

1569 1570 1567 - F: 1, T: 2, 3<br />

1590 1590 - - T: 3<br />

Comments: a – F: 1 – flat adsorbed cinchonidine, T: 2 <strong>and</strong> T: 3 – cinchonidine adsorbed<br />

in tilted mode, according to the Fig. 1-19. Assignment <strong>of</strong> peaks had been done<br />

according to the work <strong>of</strong> Ferri <strong>and</strong> Burgi [76], Zaera [77, 178, 187] <strong>and</strong> Table 4-3 .<br />

Analysis <strong>of</strong> the IR spectra <strong>of</strong> the D system shows that prior to reaction, cinchonidine is<br />

adsorbed in the flat (1) <strong>and</strong> in two tilted modes: α-H abstracted (2) <strong>and</strong> in N-lone pair<br />

bonded (3). It was found that after 50 min under 10 bar <strong>of</strong> hydrogen the peaks at 1456<br />

cm -1 <strong>and</strong> 1383 cm -1 corresponding to the tilted (2 <strong>and</strong>/or 3) <strong>and</strong> peaks at 1407 cm -1 <strong>and</strong><br />

1590 cm -1 tilted (3) form <strong>of</strong> cinchonidine are absent <strong>and</strong> only those IR peaks at 1152<br />

cm -1 , 1209 cm -1 , 1402 cm -1 <strong>and</strong> 1567 cm -1 <strong>and</strong> indicating the flat-mode (1) <strong>and</strong> tilted<br />

mode (2) are observed. After 24 hours under 10 bar <strong>of</strong> hydrogen only peaks at 1152 cm -<br />

51


1 , 1209 cm -1 <strong>and</strong> 1382 cm -1 remain. The peaks at 1152 cm -1 <strong>and</strong> 1209 cm -1 could be<br />

assigned to flat (1) <strong>and</strong> tilted (2) modes, whereas the peak at 1382 cm -1 corresponds<br />

here to the acetic acid.<br />

The absence <strong>of</strong> the tilted (3) mode <strong>of</strong> the adsorbed cinchonidine (absence <strong>of</strong> the peaks<br />

at 1590 cm -1 , 1507 cm -1 <strong>and</strong> possibly 1456 cm -1 ) after approximately one hour <strong>of</strong><br />

hydrogen exposure indicates that N-lone pair bonded cinchonidine on Pt nanoclusters<br />

has a weak adsorption strength with respect to the flat (π-bonded) <strong>and</strong> α-H abstracted<br />

species. This fact has been demonstrated previously in the work <strong>of</strong> Ferri <strong>and</strong> Burgi [76]<br />

for a thin film <strong>of</strong> Pt. Further, we have recently demonstrated above that the adsorption<br />

modes for cinchonidine for both the thin film <strong>and</strong> nanocluster systems are similar<br />

(please the beginning <strong>of</strong> this chapter).<br />

Collectively, these experimental data <strong>of</strong> the system have led us to propose the model<br />

illustrated in the Fig. 4-19 for the observed phenomena. During ITP for our Pt<br />

nanocluster systems, desoprtion <strong>of</strong> weakly bound cinchonidine molecules (tilted<br />

orientation) is occurring as a result <strong>of</strong> interaction with hydrogen <strong>and</strong> possibly solvent,<br />

making free space for the formation <strong>of</strong> new chiral sites via adsorption <strong>of</strong> cinchonidine<br />

in the ‘flat’ conformation thus leading to an increase <strong>of</strong> ee. When cinchonidine<br />

coverage <strong>of</strong> the Pt surface reaches some minimum (due to hydrogenation <strong>and</strong>/or<br />

desorption) the nanoclusters become unstable <strong>and</strong> agglomeration takes place as was<br />

observed with B <strong>and</strong> D catalysts.<br />

cinchonidine<br />

in solution<br />

H 2<br />

1507, 1590 cm -1<br />

N<br />

N<br />

HO<br />

H<br />

N<br />

OH<br />

H<br />

H<br />

H<br />

N<br />

N<br />

H H CO/CO 2 /H 2 O<br />

OH<br />

H<br />

H<br />

N<br />

1 2 3<br />

Pt surface<br />

Fig. 4-19. Proposed model <strong>of</strong> behaviour <strong>of</strong> cinchonidine on Pt under hydrogen<br />

exposure. 1-“π-bonded”, 2- “α-H abstracted”, 3- “N lone pair bonded”.<br />

The addition <strong>of</strong> free cinchonidine leads to an increase <strong>of</strong> enantioselective stability <strong>of</strong><br />

these catalysts <strong>and</strong> increase <strong>of</strong> ITP likely due to the replacement “old” to “new”<br />

cinchonidine molecules on Pt.<br />

It is interesting to compare adsorption energies <strong>of</strong> hydrogen <strong>and</strong> cinchonidine on Pt<br />

surface in order to check a principal ability <strong>of</strong> hydrogen to replace cinchonidine from<br />

the surface. It is well known that hydrogen adsorb on Pt surface with further<br />

dissociation to two hydrogen atoms (here we consider strong hydrogen adsorption<br />

only), however adsorption energy <strong>of</strong> hydrogen atom on Pt depends on many factors<br />

such as type <strong>of</strong> Pt crystal plane, adsorption site (atop, bridge, etc.), size <strong>of</strong> the Pt cluster,<br />

nature <strong>of</strong> a support (Al 2 O 3 etc.), presence <strong>of</strong> electric charge, etc. The experimental<br />

values for hydrogen adsorption on Pt (111) range from 29 to 90 kJ mol -1 [188-190] .<br />

52


We do not consider an extreme values, <strong>and</strong> would like to take the average 45 kJ mol -1<br />

values for rough comparison, since it is considered for the uncharged, unsupported Pt<br />

(111) plane [191]. The adsorption energy <strong>of</strong> cinchonidine on Pt continuous plane<br />

deposited on Al 2 O 3 support in CH 2 Cl 2 was estimated to be 31 kJ mol -1 [76], note that<br />

this is average value assigned for all three species <strong>of</strong> cinchonidine adsorption. Thus, we<br />

can conclude, that hydrogen atoms, in principal, can remove all adsorbed species <strong>of</strong><br />

cinchonidine from Pt surface, however there is a high chance, that weakly adsorbed<br />

mode <strong>of</strong> cinchonidine (N-lone pair bonding) will be removed (desorbed) faster that<br />

others.<br />

4.4 Summary<br />

The orientation <strong>of</strong> the chiral modifier cinchonidine on the surfaces <strong>of</strong> Pt nanoclusters<br />

has been investigated for the first time using a combination <strong>of</strong> DRIFTS <strong>and</strong> molecular<br />

modeling methods. It was found that the size <strong>of</strong> Pt clusters (in range 1.2- ∞ nm) has no<br />

observed influence on the adsorption mode <strong>of</strong> the cinchonidine. Thus, it was found that<br />

both ‘flat’ <strong>and</strong> ‘tilted’ modes <strong>of</strong> adsorbed cinchonidine are present on Pt nanoclusters<br />

<strong>and</strong> could be an explanation <strong>of</strong> the fact that this type <strong>of</strong> catalyst does not provide the<br />

maximum possible enantiomeric excess (98 %).<br />

For the first time, the ITP was observed <strong>and</strong> studied for cinchonidine modified Pt<br />

‘quasi-homogeneous’ systems. The ITP was observed at room temperature (22 ± 1 ºC),<br />

in acetic acid media with catalysts prepared by two different methods (D <strong>and</strong> B type),<br />

with different catalyst loadings (1.5 - 20 mg), with different ethyl pyruvate<br />

concentration (0.257, 0.514 <strong>and</strong> 2.57 mol l -1 ) under 2 <strong>and</strong> 10 bar <strong>of</strong> hydrogen pressure<br />

both with (1 mmol l -1 ) <strong>and</strong> without addition <strong>of</strong> free cinchonidine into the reaction<br />

mixture.<br />

The ITP was found to be dependent on hydrogen pressure <strong>and</strong> catalyst loading,<br />

increasing from 30 min to 100 min with a decrease <strong>of</strong> hydrogen pressure from 10 bar to<br />

2 bar. Further, ITP decreases from 30 min to 9 min with an increase <strong>of</strong> catalyst (D-type)<br />

loading from 4 to 20 mg. ITP was also found to be independent <strong>of</strong> the initial<br />

concentration <strong>of</strong> ethyl pyruvate in the range <strong>of</strong> 0.257-2.57 mol l -1 . It was also found that<br />

cinchonidine modified Pt nanoclusters on alumina catalyst demonstrates (5-10%)<br />

higher ee after treatment under reaction condition without EP.<br />

Addition <strong>of</strong> free cinchonidine (1 mmol l -1 ) shows almost no detectable influence on the<br />

maximum value <strong>of</strong> ee <strong>and</strong> slight increase <strong>of</strong> initial rate, but increases ITP by a factor <strong>of</strong><br />

3 <strong>and</strong> enantioselective stability <strong>of</strong> catalyst by nearly 5 fold. The highest ee was also<br />

found to be independent on the initial concentration <strong>of</strong> the ethyl pyruvate <strong>and</strong> decreases<br />

with an increase <strong>of</strong> the catalyst loading from 78 % to 65 % in the range <strong>of</strong> 4 mg to 20<br />

mg, probably due to the incomplete chiral activation <strong>of</strong> the catalyst.<br />

The racemic ethyl lactates as an impurity in EP does not play a significant role in<br />

behavior <strong>of</strong> enantiomeric excess, however they make systematical down shift in ee,<br />

which decreases with conversion.<br />

From IR investigation, it was found that after approximately one hour under 10 bar <strong>of</strong><br />

hydrogen pressure in acetic acid, the weakly bound species which correspond to<br />

cinchonidine adsorption in the tilted (“N lone pair bonded”) mode are not observed <strong>and</strong><br />

only the strongly bound species corresponding to cinchonidine adsorption in the flat<br />

<strong>and</strong> in tilted (“α-H abstracted”) mode remain. In other words, “N lone pair bonded”<br />

species <strong>of</strong> the adsorbed cinchonidine are desorbed due to the influence <strong>of</strong> hydrogen <strong>and</strong><br />

possibly solvent.<br />

53


Our results are in line with previously published work <strong>of</strong> Balazsik <strong>and</strong> Bartok [169] in<br />

which the authors assigned the nature <strong>of</strong> the ITP to competitive adsorption <strong>of</strong> reactants,<br />

modifier <strong>and</strong> solvent <strong>and</strong> supplement it in desorption <strong>of</strong> weakly bonded cinchonidine<br />

adsorbed in tilted “N lone pair bonded” mode. It also has to be noted, that the hydrogen<br />

effect <strong>of</strong> cleaning Pt surface from previously adsorbed impurities: H 2 O/CO/CO 2 <strong>and</strong><br />

oxygen plays a major role in ITP [166].<br />

54


Chapter 5<br />

Cinchonidine modified Pt colloidal<br />

nanoparticles immobilized on a<br />

heterogeneous support<br />

5.1 Introduction<br />

Orito [45, 46] first reported the system <strong>of</strong> cinchonidine on Pt/Al 2 O 3 as a catalyst for<br />

enantioselective hydrogenation <strong>of</strong> α-ketoesters. The ability <strong>of</strong> Pt modified with<br />

cinchonidine to induce enantioselectivity has great fundamental <strong>and</strong> practical interest<br />

[21], for example in the synthesis <strong>of</strong> HPB ester (Ethyl (R)-2-hydroxy-4-phenylbutyrate)<br />

[22]. In order to obtain high enantiomeric excess, the reaction conditions were<br />

optimized by Orito <strong>and</strong> Blaser <strong>and</strong> summarized in a review <strong>of</strong> Studer, Blaser <strong>and</strong> Exner<br />

in 2003 [44]. As can be seen from this review, that high (≥ 40 bar) hydrogen pressure<br />

<strong>and</strong> thermal pretreatment (generally at 400 °C under flow <strong>of</strong> H 2 ) <strong>of</strong> the conventional<br />

Pt/Al 2 O 3 are critical requirements to obtain high ee. However, the use <strong>of</strong> high hydrogen<br />

pressure <strong>and</strong> thermal pretreatment <strong>of</strong> the Pt/Al 2 O 3 catalyst can induce technical<br />

difficulties, safety issues <strong>and</strong> increase production cost especially on an industrial scale.<br />

The available literature dealing with mild conditions (no thermal treatment, low (≤ 10<br />

bar) hydrogen pressure) are limited [51, 78, 192, 193].<br />

In fact, ee <strong>of</strong> 94 % was obtained in ethyl pyruvate hydrogenation by Sun <strong>and</strong> coworkers<br />

[78], under 5.8 bar H 2 , 17 °C, however, this system required maintaining low<br />

dihydrocinchonidine concentration through dosing <strong>of</strong> modifier (otherwise ee<br />

significantly decreases) throughout the reaction <strong>and</strong> a very specific reaction procedure<br />

which included “add catalyst <strong>and</strong> AcOH to reactor, hydrogenate for 1 h under the<br />

hydrogenation condition (e.g. 17 °C, 5.8 bar). Evacuate H 2 <strong>and</strong> add the modifier to the<br />

reactor under a N 2 purge. Apply H 2 to desired pressure <strong>and</strong> quickly add pyruvate to<br />

reactor under H 2 pressure using a syringe pump”. Despite the fact that the authors<br />

obtained 94 % ee without thermal pretreatment <strong>of</strong> the catalyst, such a reaction<br />

procedure seems to present its own set <strong>of</strong> difficulties.<br />

In another interesting work <strong>of</strong> Bartok <strong>and</strong> coworkers [51], 97 % ee was obtained in<br />

ethyl pyruvate hydrogenation under 10 bar <strong>of</strong> H 2 with ultrasonic activation (20 kHz, 30<br />

W) in a closed hydrogen atmosphere, the activated catalyst thus requires h<strong>and</strong>ling<br />

without exposure to air. Further, using this ultrasonic activation it was reported that<br />

changes in the size <strong>and</strong> morphology <strong>of</strong> Pt nanoclusters on alumina support occurred.<br />

It is also should be noted, that the “CatASium F214” [192] catalyst developed by<br />

Degussa <strong>and</strong> Solvias does not require thermal activation <strong>and</strong> can induce e.g. 94 % ee in<br />

ethyl pyruvate hydrogenation under 60 bar <strong>of</strong> H 2 . “Alternatively, the reaction can also<br />

be carried out at 1.1 bar hydrogen pressure in a glass shaker <strong>and</strong> the hydrogen can be<br />

supplied from a balloon. However, activity <strong>and</strong> selectivity don't reach an optimum”<br />

[193].<br />

Herein, we present the detailed preparation <strong>of</strong> a novel heterogeneous catalyst:<br />

cinchonidine modified Pt nanoclusters deposited on non-porous alumina which<br />

demonstrates 80-88 % ee in hydrogenation <strong>of</strong> ethyl pyruvate under relatively low<br />

hydrogen pressure (2.5-10 bar). The reaction procedure does not require thermal or<br />

ultrasonic activation <strong>of</strong> the catalyst <strong>and</strong> allows storage <strong>of</strong> the catalyst in air before<br />

55


eaction. The stability <strong>of</strong> the catalyst structure under reaction <strong>and</strong> preparation<br />

conditions is considered as well as catalyst recycling <strong>and</strong> reuse. The influence <strong>of</strong> the<br />

catalyst nature towards obtaining high enantiomeric excess in this pressure range is also<br />

considered.<br />

5.2 Experimental<br />

Materials<br />

5% Pt on alumina (Aldrich, dispersion 29 %, Pt particles size 3.9 nm), cinchonidine<br />

(98%, Fluka), acetic acid (99.8% Fluka), ethyl pyruvate (98%, Aldrich), formic acid<br />

(Aldrich, 99%), acetone <strong>and</strong> aluminum oxide (4 m 2 g -1 ) 90 active neutral (0.063-0.2<br />

mm) by Applichem were used as received.<br />

Samples preparation<br />

Adsorption <strong>of</strong> PtCl 4 on alumina was performed according to the following procedure. 4<br />

g Al 2 O 3 was heated at 180 °C under vacuum (10 -3 mbar) for 16 hours <strong>and</strong> then was<br />

cooled to room temperature under vacuum. To this alumina, 16 mg PtCl 4 dissolved in<br />

40 ml <strong>of</strong> acetone solution was added by syringe. After 5-10 minutes stirring, the<br />

initially dark supernatant acetone solution became colorless <strong>and</strong> an additional 10 ml<br />

aliquot <strong>of</strong> acetone (with 30 mg PtCl 4 dissolved) was added by syringe. This procedure<br />

was repeated several times, until 480 mg PtCl 4 was added, at which point the<br />

supernatant no longer became colorless. The mixture was then refluxed for 12 hours<br />

<strong>and</strong> the supernatant was evaporated under vacuum. Thus 480 mg PtCl 4 on 4000 mg<br />

alumina (~11 % PtCl 4 ) was obtained.<br />

In order to reduce Pt The 600 mg <strong>of</strong> alumina with adsorbed PtCl 4 was transferred into a<br />

round bottom flask with distilled water (100 ml). The mixture was refluxed with<br />

constant stirring. 160 mg dihydrocinchonidine dissolved in 25 ml aqueous solution <strong>of</strong><br />

formic acid 0.1 mol·L -1 was added rapidly by syringe <strong>and</strong> Pt was reduced. The heat was<br />

removed after 10 min. The mixture was allowed to stir for 24 hours, the black powder<br />

was isolated by filtration <strong>and</strong> washed with water, then with acetic acid <strong>and</strong> finally dried<br />

in air at room temperature. Thus prepared catalysts will be referred to as A1 henceforth<br />

in this manuscript.<br />

For catalyst activation140 mg <strong>of</strong> the A1 catalyst was transferred into a metal reactor<br />

with 70 ml acetic acid <strong>and</strong> 20 mg cinchonidine. The mixture was maintained under<br />

hydrogen pressure (10 bar) for 1 hour with constant stirring (700 rpm). The pressure<br />

was then released <strong>and</strong> the black powder was collected, washed 3 times with acetic acid<br />

(100 ml), <strong>and</strong> then dried in air at room temperature. Thus prepared catalysts will be<br />

referred as A2 henceforth in this manuscript.<br />

Unless otherwise mentioned, the conventional Pt/Al 2 O 3 was heated to 400 °C under<br />

vacuum (10 -3 mbar) for 1 hour, then kept for 3 hours under a flowing mixture <strong>of</strong> N 2<br />

(93%) <strong>and</strong> H 2 (7%) gases (100 ml min -1 ) <strong>and</strong> subsequently cooled to the room<br />

temperature. The required amount was then weighed (in open atmosphere) <strong>and</strong><br />

immediately transferred into a glass reactor with acetic acid, cinchonidine <strong>and</strong> ethyl<br />

pyruvate.<br />

Modification <strong>of</strong> conventional Pt/Al 2 O 3 has been done, according to the following<br />

procedure. 40 mg <strong>of</strong> conventional Pt/Al 2 O 3 was heated to 400 ºC under vacuum (10 -3<br />

mbar) for 1 hour, then the mixture <strong>of</strong> N 2 (93%) <strong>and</strong> H 2 (7%) gases were passed (100 ml<br />

min -1 ) over the sample for 3 hours. Then, under flowing gases, the support was cooled<br />

to room temperature <strong>and</strong> the solution <strong>of</strong> cinchonidine in CHCl 3 (6.0 ml, 3 mM) was<br />

56


added by syringe through a stopper. The system was kept at 10-15 ºC for 24 hours, then<br />

washed 3 times with CHCl 3 <strong>and</strong> finally dried under vacuum at room temperature.<br />

Characterization <strong>and</strong> experiments<br />

Transmission electron microscopy (TEM) micrographs were taken using a TECNAI<br />

F20 instrument operating at 200 kV. Specimens were prepared by placing a drop <strong>of</strong> an<br />

ethanolic sample suspension onto a copper grid with a perforated carbon film <strong>and</strong> then<br />

allowing the ethanol to evaporate.<br />

For IR investigation a “Nicolet 4700 FT-IR” spectrometer with liquid nitrogen cooled<br />

detector <strong>and</strong> nitrogen purged DRIFT accessory was used with the following parameters:<br />

200 scans, 600-4000 cm -1 scan range, 4 cm -1 resolution. KBr background was acquired<br />

before spectra measurements. The DRIFT spectra were chosen for analysis because it is<br />

the more applicable technique for observing adsorbed species on finely divided metal<br />

powders [174].<br />

Hydrogenation <strong>of</strong> ethyl pyruvate<br />

Hydrogenation <strong>of</strong> ethyl pyruvate was carried out in a metal reactor (Parr Instrument<br />

GmbH) under a pressure <strong>of</strong> 2.5-10 bar <strong>of</strong> hydrogen at room temperature (22±1 °C). The<br />

stirring speed was 1000 rpm.<br />

Typical conditions were such that 4 ml (36 mmol) ethyl pyruvate was added to 20 mg<br />

(68 µmol) cinchonidine in acetic acid (70 ml) solution. 20 mg <strong>of</strong> catalyst was added<br />

into the mixture <strong>and</strong> then transferred into the reactor, flushed with Ar for 10 min <strong>and</strong><br />

finally, after the pressure <strong>of</strong> hydrogen had been stabilized for 10-15 sec, the initial time<br />

(t=0) was set <strong>and</strong> the reaction kinetics were monitored. About one ml aliquots <strong>of</strong> the<br />

reaction mixture were taken at certain time intervals, filtrated with a G3/G4 glass filter<br />

<strong>and</strong> 1 μl was injected into the gas chromatograph (GC). GC measurements were taken<br />

using a Varian 3900 instrument with FID detector <strong>and</strong> Lipodex-E column.<br />

Catalyst stability <strong>and</strong> reuse tests<br />

In order to test the stability <strong>of</strong> the catalysts under the typical reaction conditions, 30 mg<br />

A1 was transferred into the reactor with cinchonidine (250 mg) in 70 ml acetic acid<br />

solution <strong>and</strong> left with continuous stirring under 10 bar <strong>of</strong> hydrogen pressure for 16<br />

hours. The pressure was then released <strong>and</strong> samples were collected for TEM analysis.<br />

In the case <strong>of</strong> catalyst reuse experiments, after the reaction is finished (conversion<br />

reached 100 %), hydrogen pressure was released <strong>and</strong> the reactor was unplugged. The<br />

black powder was collected through centrifugation <strong>and</strong> washed three times with acetic<br />

acid, then was dried in air at room temperature; the catalyst was then used again in the<br />

identical manner to that described above.<br />

5.3 Results <strong>and</strong> discussion<br />

5.3.1 Role <strong>of</strong> catalyst activation in obtaining enantiomeric<br />

excess<br />

The A1 catalyst (Table 5-1) demonstrates 79-80 % <strong>of</strong> ee which is similar to the data<br />

reported by Bönnemann [114, 115] for alumina supported cinchonidine modified Pt<br />

nanoclusters. The same catalyst after activation (A2) demonstrates 85 % <strong>of</strong> ee <strong>and</strong> an<br />

increase <strong>of</strong> reaction rate by a factor <strong>of</strong> ~9. The A2 catalyst demonstrates stable ee <strong>and</strong><br />

100 % <strong>of</strong> conversion after 20-25 min (Fig. 5-1).<br />

57


Table 5-1. Ee <strong>and</strong> reaction rate induced by cinchonidine modified catalysts before <strong>and</strong><br />

after activation.<br />

Catalyst A Ee, % Reaction rate, mmol<br />

(min·g) -1<br />

A1- Before activation 79-80 3<br />

A2 -After activation 85 28<br />

100<br />

100<br />

Conversion, %<br />

80<br />

60<br />

40<br />

20<br />

80<br />

60<br />

40<br />

20<br />

Enantiomeric excess, %<br />

0<br />

0 5 10 15 20<br />

Time, min<br />

0<br />

Fig. 5-1. Evolution <strong>of</strong> conversion (squares) <strong>and</strong> enantiomeric excess (triangles) for A2<br />

catalyst under 10 bar <strong>of</strong> hydrogen pressure.<br />

A similar effect <strong>of</strong> increased ee (10-20%) after catalyst exposure to reaction condition<br />

was found by Bartok <strong>and</strong> Balàzsik [194] for conventional Pt/Al 2 O 3 (E4759) in toluene<br />

but was not found in acetic acid, probably, due to the differences in the catalysts. The<br />

proposed origin <strong>of</strong> ee <strong>and</strong> rate enhance for our catalyst is discussed below.<br />

5.3.2 Comparison <strong>of</strong> catalysts <strong>and</strong> proposed model <strong>of</strong><br />

catalyst activation<br />

In order to compare cinchonidine modified catalysts A1 <strong>and</strong> A2 with modified <strong>and</strong><br />

unmodified conventional catalyst (Aldrich) (the latter is modified with cinchonidine in<br />

situ) we tested them in ethyl pyruvate hydrogenation in the range <strong>of</strong> 2.5-10 bar <strong>of</strong><br />

hydrogen pressure.<br />

The major difference in selectivity between cinchonidine modified conventional<br />

Pt/Al 2 O 3 (see modification procedure in experimental part) <strong>and</strong> A1 catalyst was<br />

observed in the ethyl pyruvate hydrogenation without addition <strong>of</strong> free cinchonidine. In<br />

fact, the A1 catalyst induces 71 % <strong>of</strong> ee, whereas cinchonidine modified conventional<br />

Pt/Al 2 O 3 demonstrates only 40 % ee.<br />

It should be noted that without free cinchonidine, the A1 catalyst shows 71 % <strong>of</strong> ee <strong>and</strong><br />

ee decreases with reaction time (Fig. 5-2), whereas in the experiment with the addition<br />

58


<strong>of</strong> free cinchonidine, the catalyst demonstrates 80 % <strong>of</strong> ee <strong>and</strong> no decrease <strong>of</strong> ee was<br />

found during reaction. Another important fact is that just 10 % <strong>of</strong> conversion after 80<br />

min (Fig. 5-2) was reached in the ethyl pyruvate hydrogenation without free<br />

cinchonidine, whereas conversion <strong>of</strong> 70 % after 80 min (not shown) was reached with<br />

free cinchonidine with the same catalyst <strong>and</strong> under the same conditions. This<br />

phenomena can be explained as sorption/desorption <strong>of</strong> cinchonidine between Pt <strong>and</strong><br />

solution. Desorption <strong>of</strong> cinchonidine from Pt surface has been already observed even<br />

under 1 bar <strong>of</strong> hydrogen <strong>and</strong> explained as partial hydrogenation <strong>of</strong> quinoline anchor<br />

[169].<br />

10<br />

72<br />

Conversion, %<br />

8<br />

6<br />

4<br />

70<br />

Enantiomeric excess, %<br />

2<br />

68<br />

0<br />

0 20 40 60 80<br />

Time, min<br />

Fig. 5-2. Evolution <strong>of</strong> conversion (squares) <strong>and</strong> enantiomeric excess (triangles) for A1<br />

catalyst without free cinchonidine, hydrogen pressure 10 bar.<br />

Based on IR investigations, we would like to suggest here that A1 (<strong>and</strong> A2) possess a<br />

higher concentration <strong>of</strong> adsorbed cinchonidine with respect to modified conventional<br />

Pt/Al 2 O 3 . Despite the fact that both catalysts have Pt particles <strong>of</strong> similar size (3 (TEM)<br />

<strong>and</strong> 3.9 nm (chemisorption)) the A1 catalyst (made by reduction <strong>of</strong> the Pt salt) is more<br />

active than conventional Pt/Al 2 O 3 , most probably due to the fact the latter has a lower<br />

level <strong>of</strong> cleanliness <strong>of</strong> the Pt surface (even after treatment with H 2 at 400 °C) with<br />

respect to freshly reduced <strong>and</strong> modified Pt.<br />

For ethyl pyruvate hydrogenation with the addition <strong>of</strong> free cinchonidine (Fig. 5-3) it<br />

was found that in the range <strong>of</strong> 2.5- 10 bar <strong>of</strong> hydrogen pressure, A2 demonstrates<br />

approximately 7-8 % higher enantiomeric excess <strong>and</strong> a higher reaction rate by a factor<br />

<strong>of</strong> 1.7-2.3 than the conventional (non modified in prior) Pt/Al 2 O 3 at the same<br />

conditions.<br />

59


Enantiomeric excess, %<br />

85<br />

80<br />

75<br />

A2<br />

Conventional Pt/Al 2<br />

O 3<br />

40<br />

30<br />

20<br />

Reaction rate, mmol/min*g cat<br />

10<br />

70<br />

2 4 6 8 10<br />

Hydrogen Pressure, bar<br />

Fig. 5-3. Influence <strong>of</strong> hydrogen pressure on enantiomeric excess (filled symbols) <strong>and</strong><br />

reaction rate (open symbols) for conventional Pt/Al 2 O 3 (triangles) <strong>and</strong> A2 catalysts<br />

(squares).<br />

Our proposal that this is probably due to higher surface concentration <strong>of</strong> active chiral<br />

sites for the A2 catalyst is based on the fact that the A2 catalyst has already adsorbed<br />

cinchonidine (in flat <strong>and</strong> in tilted adsorption mode, see section 5.3.4 for IR<br />

investigation) <strong>and</strong> adsorbs additional cinchonidine (in both modes) from solution <strong>and</strong> at<br />

the same time the conventional Pt/Al 2 O 3 (which is previously unmodified) adsorbs<br />

cinchonidine only from solution (also in both modes) during reaction (in situ).<br />

An increase <strong>of</strong> ee with increased hydrogen pressure was found for both A2 <strong>and</strong> the<br />

conventional catalyst. This increase could be explained through the hydrogen cleaning<br />

effect. In fact, it was shown that hydrogen cleans the Pt surface <strong>of</strong> different impurities<br />

from solvent, ethyl pyruvate, <strong>and</strong> air (O 2 , CO/CO 2 ) [166, 195]. It is known, that the<br />

likelihood <strong>of</strong> the occurrence for flat adsorption mode <strong>of</strong> the cinchonidine (active chiral<br />

site) depends on the concentration <strong>of</strong> modifier in the solution [76, 187], however it is<br />

clear, that cinchonidine adsorbed in flat mode requires more free space on the Pt<br />

surface with respect to any tilted adsorption mode. Therefore, we suppose that with an<br />

increase <strong>of</strong> concentration <strong>of</strong> free sites (due the hydrogen cleaning effect) the chance <strong>of</strong><br />

flat adsorption increases.<br />

A1 after activation (see experimental part) is referred to as A2 was found to have a<br />

lower concentration <strong>of</strong> CO contamination on the Pt surface (see IR investigation section<br />

5.3.4) with respect to its state before activation, <strong>and</strong> demonstrates a plateau in the plot<br />

<strong>of</strong> ee vs P H2 (Fig. 5-3). The presence <strong>of</strong> the plateau for the A2 catalyst <strong>and</strong> its absence<br />

for conventional catalysts indicates differences in the processes associated with each<br />

system.<br />

It should be noted that, since all catalysts were kept in air <strong>and</strong> thermal treatment at 400<br />

°C is not possible for those catalysts modified during preparation, these catalysts as<br />

well as the conventional Pt/Al 2 O 3 catalyst were used without treatment at high<br />

60


temperature in all <strong>of</strong> the above mentioned experiments. However, thermally treated<br />

(with hydrogen according to procedure described above) conventional Pt/Al 2 O 3<br />

demonstrates 91 % ee (10 bar H 2 ) which is 3 % higher than maximum obtained with A2<br />

catalyst under the same reaction conditions. This significant (~14 %) increase in ee<br />

after catalyst treatment clearly demonstrates importance <strong>of</strong> surface cleaning.<br />

It is known, that platinum can be oxidized by O 2 <strong>and</strong> form PtO or PtO x [196] under red<br />

heat <strong>and</strong> Pt is able to dissociatively chemisorb oxygen at room temperature, other<br />

proposed models <strong>of</strong> oxygen adsorption on Pt can be found in the works <strong>of</strong> Smith [197]<br />

<strong>and</strong> Matsushima [198] <strong>and</strong> others [199-201]. In our experiments with <strong>and</strong> without<br />

thermal pretreatment (catalyst was transferred to reactor in air at room temperature) <strong>of</strong><br />

the Pt catalyst, chemisorption <strong>of</strong> oxygen on Pt likely occurs. However, it does not<br />

prevent cinchonidine <strong>and</strong> ethyl pyruvate adsorption with the presence <strong>of</strong> hydrogen in a<br />

solvent (in situ) <strong>and</strong> min. 77 % <strong>of</strong> ee (conventional Pt/Al 2 O 3 ) was observed due to the<br />

surface cleaning effect, induced by hydrogen [76, 82, 195, 202]. The significant<br />

increase <strong>of</strong> ee (up to 91 %, in our work) with thermally treated Pt/Al 2 O 3 has been<br />

known since report <strong>of</strong> Orito et al. [45] <strong>and</strong> was explained as further surface cleaning<br />

[203] e.g. from organic species, H 2 O, CO, CO 2 , etc.<br />

The global role <strong>of</strong> oxygen in the reaction is rather ambiguous [69], for example, it was<br />

shown that oxygen induces a cleaning effect through CO removal [195, 204] <strong>and</strong><br />

oxygen was to reduce cinchonidine coverage on Pt in such a way that a larger fraction<br />

<strong>of</strong> the Pt surface would be available for ethyl pyruvate hydrogenation [205].<br />

5.3.3 Catalyst stability <strong>and</strong> reuse tests<br />

The TEM investigation <strong>of</strong> sample A1 before the stability test (Fig. 5-4) shows that<br />

alumina globules are fully covered with Pt nanoclusters with an average size about 2.7<br />

nm. The small increase <strong>of</strong> average particles size (maximum <strong>of</strong> 0.5 nm) was observed<br />

from the sample following the stability test; in fact the average size <strong>of</strong> Pt nanoclusters<br />

after the stability test was found to be about 3.2 nm. It should be noted that the surface<br />

<strong>of</strong> an alumina globule after the stability test is still fully covered with Pt nanoclusters<br />

<strong>and</strong> ultrasonic treatment (at least for half an hour) does not influence the size <strong>and</strong><br />

stability <strong>of</strong> alumina supported Pt nanoclusters.<br />

61


35<br />

30<br />

Before<br />

After<br />

Percent <strong>of</strong> particles, %<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5<br />

Size, nm<br />

Fig. 5-4. TEM images <strong>of</strong> fully covered alumina globule before (above) <strong>and</strong> after<br />

(below) stability test <strong>and</strong> a histogram representing the Pt particle size distribution<br />

before <strong>and</strong> after catalyst stability test.<br />

Since, there is no significant change in the size <strong>of</strong> Pt nanoclusters after the reaction time<br />

was exceeded; the possibility <strong>of</strong> recycling was explored. The A2 catalyst (see Table 5-<br />

2) was found to be fully reusable (at least four times) without loss <strong>of</strong> activity <strong>and</strong><br />

enantioselectivity. Moreover, a slight increase <strong>of</strong> ee was observed, which could be<br />

explained as the result <strong>of</strong> additional activation after first reaction.<br />

Table 5-2: A2 catalyst reuse data under 10 bar <strong>of</strong> hydrogen pressure.<br />

Number <strong>of</strong> use Ee, % Reaction rate, mmol<br />

(min·g) -1<br />

1 85 28<br />

2 87 25<br />

3 86 29<br />

4 88 30<br />

It is known from the work <strong>of</strong> Reschetilowski <strong>and</strong> co-workers [206] (home-made<br />

Pt/zeolite) <strong>and</strong> work <strong>of</strong> Bartok <strong>and</strong> co-workers [194] (conventional Pt/Al 2 O 3 ), that ee<br />

could be constant (Reschetilowski) or decrease <strong>of</strong> 5 % in catalyst reuse experiments,<br />

where “fresh” (dihydro)cinchonidine was added in every reuse. However, a significant<br />

decrease <strong>of</strong> reaction rate was detected in both works, which might indicate a decrease<br />

<strong>of</strong> catalyst activity, whereas the A2 catalyst does not demonstrate a decrease <strong>of</strong> ee or<br />

62


eaction rate, moreover, the ee <strong>and</strong> reaction rate were slightly increased, probably due<br />

to the further activation.<br />

5.3.4 IR investigation <strong>and</strong> proposed model <strong>of</strong> activation<br />

As we have shown already (chapter 4) cinchonidine demonstrates similarity in<br />

adsorption on Pt nanoclusters (1.2 nm size) with respect to the macroscopic Pt support.<br />

In fact, cinchonidine was found in both the tilted <strong>and</strong> flat adsorption modes in both<br />

cases. By using previously obtained data from modeling <strong>of</strong> IR vibrations <strong>of</strong> the<br />

cinchonidine molecule in open 3 conformation (chapter 4) <strong>and</strong> data from the work <strong>of</strong><br />

Ferri [76] <strong>and</strong> Zaera [77] the assignment <strong>of</strong> the observed peaks has been accomplished.<br />

Fig. 5-5 shows the presence <strong>of</strong> peaks <strong>and</strong> shoulders that are in correlation with the<br />

spectrum <strong>of</strong> free cinchonidine. Here, modification <strong>of</strong> both: conventional catalysts <strong>and</strong><br />

A1 prepared as described above is clearly established. It should be noted that, in the<br />

case <strong>of</strong> A1, peaks corresponding to the adsorbed cinchonidine are better resolved <strong>and</strong><br />

have higher intensity with respect to cinchonidine modified conventional Pt/Al 2 O 3 due<br />

to the higher concentration <strong>of</strong> adsorbed cinchonidine on the Pt surface.<br />

1640<br />

0.5<br />

D<br />

F<br />

A - 71 % ee<br />

1500 1000<br />

Adsorbance<br />

0.25<br />

B - 40 % ee<br />

C<br />

1700 1600 1500 1400<br />

Wavenumber,<br />

1300<br />

cm<br />

1200 1100 1000<br />

Fig. 5-5. DRIFTS <strong>of</strong> cinchonidine modified A1 (A), conventional Pt/Al 2 O 3 (B) <strong>and</strong> free<br />

cinchonidine (C). The unmodified Pt/Al 2 O 3 (Aldrich) <strong>and</strong> alumina presented in inset as<br />

F <strong>and</strong> D, correspondingly. Numbers near symbols A <strong>and</strong> B show maximum obtained<br />

enantiomeric excess with respective catalyst without addition <strong>of</strong> free cinchonidine<br />

under 10 bar <strong>of</strong> H 2 . The intensity <strong>of</strong> spectra <strong>of</strong> free cinchonidine was decreased for<br />

better comparison.<br />

A DRIFTS investigation <strong>of</strong> cinchonidine modified Pt nanoclusters on alumina support<br />

was performed before (A1) <strong>and</strong> after catalyst activation (A2). The practical absence <strong>of</strong><br />

the peaks at ~2100 cm -1 (Fig. 5-6) which corresponds to adsorbed CO [202, 207] after<br />

63


catalyst activation seems to indicate a Pt surface cleaning effect. This cleaning effect<br />

was also observed by Baiker <strong>and</strong> co-authors for similar systems [166, 202].<br />

Absorbance<br />

0.004<br />

A1<br />

A2<br />

2300 2200 2100 2000 1900<br />

Wavenumber, cm -1<br />

Fig.5-6. DRIFT spectra <strong>of</strong> CO adsorbed on A1 <strong>and</strong> A2 catalysts.<br />

It should be noted, that the spectral range below 1200 cm -1 is not accessible for<br />

investigation, due to the alumina background (Fig. 5-5, 5-7) which covers the frequency<br />

range below 1200 cm -1 where the most useful (for unambiguous assignment <strong>of</strong> the flat<br />

adsorbed species <strong>of</strong> the cinchonidine) out <strong>of</strong> plane vibrations <strong>of</strong> quinoline ring are<br />

located. Further, the presence <strong>of</strong> alumina “dilutes” the cinchonidine in the IR beam <strong>and</strong><br />

lowers peak resolution (especially around 1580 cm -1 ) with respect to cinchonidine<br />

stabilized Pt nanoclusters (support free) (please see chapter 4). Both <strong>of</strong> these effects<br />

hinder unambiguous interpretation <strong>of</strong> the change in spectra. However, some significant<br />

changes are clearly observed, they are summarized in the Table 5-4 <strong>and</strong> we provide the<br />

following interpretation.<br />

64


A2 -88 % ee<br />

Adsorbance<br />

1709<br />

1613<br />

1583<br />

1572<br />

1470<br />

1463<br />

1422<br />

1422<br />

1371<br />

1340 1336<br />

1280<br />

A1 -80 % ee<br />

C<br />

0.5<br />

1700 1600 1500 1400 1300 1200 1100 1000<br />

Wavenumber, cm -1<br />

Fig. 5-7. DRIFT spectra <strong>of</strong> A1 <strong>and</strong> A2 catalysts <strong>and</strong> free cinchonidine (C). Numbers<br />

near symbols A <strong>and</strong> B show enantiomeric excess obtained with respective catalyst with<br />

addition <strong>of</strong> free cinchonidine (1 mol L -1 ) under 10 bar <strong>of</strong> H 2 . The intensity <strong>of</strong> spectra <strong>of</strong><br />

free cinchonidine was decreased for better performance.<br />

The appearance <strong>of</strong> the shoulder at 1613 cm -1 after activation could be explained as an<br />

increase <strong>of</strong> concentration <strong>of</strong> cinchonidine in the “N lone pair bonded” adsorption mode.<br />

The peak at 1572 cm -1 before <strong>and</strong> 1583 cm -1 after activation are super positions <strong>of</strong><br />

peaks/shoulders (which could be seen separately in DRIFT spectra <strong>of</strong> unsupported<br />

cinchonidine on Pt nanocluster samples (chapter 4)) corresponding to 1619, 1591 <strong>and</strong><br />

1568 cm -1 peaks in the spectra <strong>of</strong> free cinchonidine. Such a shift as that from 1572<br />

before to 1583 cm -1 after activation can be explained as partial interconversion <strong>of</strong> one<br />

or more adsorption mode (flat 1, tilted 2, tilted 3) adsorption mode to tilted 3. Also, an<br />

increase <strong>of</strong> the relative intensity <strong>and</strong> blue shift in wavenumber from 1463 (wk) to 1470<br />

(med) cm -1 allows interpretation as an interconversion <strong>of</strong> adsorption modes (flat 1,<br />

tilted 2, tilted 3) between each other <strong>and</strong> taking into the account the assumption that a<br />

larger shift in wavelength results from stronger adsorption, thus it is possible to propose<br />

that the concentration <strong>of</strong> tilted 2 mode decreased with respect to flat 1 or/<strong>and</strong> tilted 3<br />

modes. No significant change in intensity or in wavenumber was found for the peak at<br />

1422 cm -1 , corresponding to the flat 1 adsorption mode, whereas an increase <strong>of</strong><br />

intensity at 1371 cm -1 led to the appearance <strong>of</strong> the shoulder which corresponds to the<br />

flat 1 adsorption mode. The decrease <strong>of</strong> intensity at 1340 cm -1 before activation <strong>and</strong> red<br />

shift to 1336 cm -1 after activation is difficult to unambiguously explain since there is<br />

not a clear assignment <strong>of</strong> this vibration b<strong>and</strong> with an adsorption mode. The situation is<br />

similar with the peak at 1264 cm -1 before activation <strong>and</strong> the weak shoulder at 1264 cm -1<br />

after activation. The latter is almost covered by the broad peak at 1275 cm -1 which can<br />

be assigned to acetic acid.<br />

65


Table 5-3. Assignment <strong>of</strong> some vibrations (cm-1) <strong>of</strong> cinchonidine adsorbed on Pt<br />

before (A1) <strong>and</strong> after activation (A2).<br />

A1 A2 Free Assignment <strong>and</strong> orientation <strong>of</strong><br />

cinchonidine. adsorbed cinchonidine<br />

- 1709 (st. wd.) - Adsorbed acetic acid<br />

- 1613 (sh) 1619 (wk) Tilted 3<br />

1572 (st) 1583 (st) 1591 (st), Tilted 3<br />

1568 (med) Flat 1, tilted 2 <strong>and</strong> tilted 3<br />

1463 (wk) 1470 (med) 1452 (str) Flat 1, tilted 2 <strong>and</strong> tilted 3<br />

1422 (med) 1422 (med) 1422 (med) Flat 1<br />

- 1371 (sh) 1355 (med) Flat 1<br />

1340 (med) 1336 (sh) 1327 (med), -<br />

1337 (med) Flat 1<br />

- 1275 (wd) - Adsorbed acetic acid<br />

1264 (med) 1264 (wk, sh.) - -<br />

Comments: st. – strong, wd. – wide, sh. – shoulder, wk. – weak, med. – medium, flat: -1<br />

“π-bonded” adsorption mode, tilted: 2 - “α-H abstracted” adsorption mode, tilted: 3 -<br />

“N lone pair bonded” adsorption.<br />

In order to assess the ability <strong>of</strong> the cinchonidine to adsorb on used alumina (pure<br />

alumina passed all preparation <strong>and</strong> activation steps) the control experiments were<br />

performed <strong>and</strong> it was found from DRIFT spectra that no cinchonidine peaks were found<br />

on alumina after washing with acetic acid. Further, no conversion <strong>of</strong> ethyl pyruvate was<br />

observed by using cinchonidine <strong>and</strong> alumina without Pt.<br />

Based on the data described above we would like to propose a model for catalyst<br />

activation. Thus, activation results from a decrease <strong>of</strong> CO coverage <strong>of</strong> Pt surface <strong>and</strong> an<br />

increase <strong>of</strong> the concentration <strong>of</strong> flatly bound cinchonidine (active chiral site), probably<br />

through reorganization between adsorbed cinchonidine species (partial desorption <strong>of</strong><br />

weakly bound cinchonidine in tilted mode to solution) <strong>and</strong> additional adsorption <strong>of</strong><br />

cinchonidine from solution onto the produced free sites as well <strong>of</strong> adsorption <strong>of</strong> acetic<br />

acid, which could be competitive.<br />

5.4 Summary<br />

High enantiomeric excess in ethyl pyruvate hydrogenation was obtained at low – 2.5-10<br />

bar hydrogen pressure with conventional Pt/Al 2 O 3 - 91 % ee (after thermal pretreatment<br />

with hydrogen) <strong>and</strong> 80-88 % ee (without thermal pretreatment) with cinchonidine<br />

modified Pt nanoclusters deposited on nonporous alumina (after activation with<br />

hydrogen <strong>and</strong> cinchonidine in acetic acid at room temperature) without requirement <strong>of</strong><br />

avoiding catalyst contact with air. The stability <strong>of</strong> the catalyst under reaction conditions<br />

<strong>and</strong> recycling <strong>and</strong> reuse <strong>of</strong> the catalyst was successfully demonstrated.<br />

The positive effect <strong>of</strong> activation under reaction conditions was demonstrated <strong>and</strong><br />

explained as removal <strong>of</strong> contaminants from the Pt surface <strong>and</strong> reorganization <strong>of</strong> the<br />

adsorbed cinchonidine species <strong>and</strong> adsorption <strong>of</strong> a new portion <strong>of</strong> lig<strong>and</strong> from solution.<br />

These effects led to an increase <strong>of</strong> the concentration <strong>of</strong> active chiral sites on Pt surface<br />

<strong>and</strong> thus obtained ee <strong>and</strong> rate enhancement.<br />

66


Chapter 6<br />

Cinchonidine modified Pd colloidal<br />

nanoparticles<br />

6.1 Introduction<br />

It is known, that Pd is a suitable catalyst for hydrogenation <strong>of</strong> double bonds.<br />

Cinchonidine modified Pd system was found to be able to induce enantioselectivity in<br />

hydrogenation <strong>of</strong> substituted alkenes. In 1985 Perez <strong>and</strong> co-workers reported the<br />

hydrogenation <strong>of</strong> (E)-α-phenyl-cinnamic acid to 2,3-diphenylpropionic acid over<br />

cinchonidine-modified Pd/C; the enantiomeric excess was 30 % in favour <strong>of</strong> the S-<br />

product [208]. The role <strong>of</strong> a support for Pd, metal loading, dielectric constant <strong>of</strong> the<br />

solvent, concentration <strong>of</strong> cinchonidine <strong>and</strong> substrate, hydrogen pressure <strong>and</strong><br />

temperature has been studied by Nitta <strong>and</strong> co-workers [98, 209-213]. Enantioselective<br />

hydrogenation <strong>of</strong> the carbon-nitrogen double bond on an example <strong>of</strong> Ph(C=NOH)Me<br />

(ee 18 % ) is known from reports <strong>of</strong> Blaser [214] <strong>and</strong> Nakamura [215]. More examples<br />

for enantioselective hydrogenations <strong>of</strong> double carbon <strong>and</strong> carbon-nitrogen bond can be<br />

found from the review <strong>of</strong> Wells [87].<br />

In this chapter we focus on hydrogenation activated ketones over cinchonidine<br />

modified Pd catalysts. Hydrogenation <strong>of</strong> ethyl (or methyl) pyruvate on cinchonidine<br />

modified Pd catalyst leads to the interesting phenomena, so called inversed<br />

enantioselectivity. In fact, in 1988 Blaser <strong>and</strong> co-workers reported that cinchonidinemodified<br />

Pd/carbon was mildly enantioselective for the hydrogenation <strong>of</strong> ethyl<br />

pyruvate, the reaction giving an enantiomeric excess <strong>of</strong> 4% in favour <strong>of</strong> the S-<br />

enantiomer [94], where as hydrogenation over cinchonidine modified Pt catalyst results<br />

in enantiomeric excess in favour <strong>of</strong> R-ethyl lactate (see chapters 4 <strong>and</strong> 5). Closer<br />

examination showed that reaction over conventionally supported Pd differed very<br />

significantly from the corresponding Pt system [49, 216]. In fact, rate enhancement was<br />

absent over cinchonidine modified Pd, enantioselective reaction was half order in<br />

hydrogen whereas it was first-order for Pt, reactions over Pd were only achieved in<br />

certain solvents <strong>and</strong> Pd was enantioselective only if reduced at low temperature.<br />

It is known, that in the case <strong>of</strong> Pt catalyst the use <strong>of</strong> cinchonidine modifier results in<br />

excess <strong>of</strong> R enantiomer <strong>and</strong> cinchonine gives S lactate. However very interesting<br />

phenomenon takes place with use <strong>of</strong> Pd based catalyst, in fact, the use <strong>of</strong> cinchonidine<br />

over Pd catalysts leads to positive (excess <strong>of</strong> R enantiomers) or negative (excess <strong>of</strong> S<br />

enantiomers) ee, at the same time cinchonine over Pd catalysts is able to induce<br />

negative <strong>and</strong> positive ee [49]. Wells <strong>and</strong> Whyman [217] first reported this phenomenon,<br />

they found, that the direction <strong>of</strong> the enantioselectivity being solvent <strong>and</strong>/or substrate<br />

(methyl <strong>and</strong> ethyl pyruvate) dependent.<br />

In this chapter we define the following aims: to investigate orientation <strong>of</strong> the<br />

cinchonidine molecule being adsorbed on Pd nanoclusters (~1.4 nm, see below) via<br />

DRIFTS similarly to the Pt nanoclusters (please see the chapter 4) <strong>and</strong> to show<br />

differences in the direction <strong>of</strong> enantioselectivity between modified with cinchonidine in<br />

situ conventional Pd/Al 2 O 3 catalyst <strong>and</strong> cinchonidine modified Pd nanoclusters in ethyl<br />

pyruvate <strong>and</strong> ethyl benzoyformate hydrogenation reactions. The choice <strong>of</strong> ethyl<br />

benzoyformate was caused because it can not, in principal, form an enolate tautomer<br />

state (please find details below).<br />

67


6.2 Experimental<br />

Materials<br />

Cinchonidine (>98% Fluka), 5 % Pd/Al 2 O 3 (Fluka), Pd(DBA) 2 (Aldrich), THF (> 99.5<br />

% Fluka), cyclohexene (> 99% Fluka) <strong>and</strong> ethyl benzoylformate (95 % Aldrich) were<br />

used as received.<br />

Samples preparation<br />

Cinchonidine modified Pd nanoclusters were prepared according to the following<br />

procedure. 1084 mg Pd(DBA) 2 (1.88 mmol Pd) <strong>and</strong> 1658 mg cinchonidine (5.64 mm)<br />

in 70 ml THF were transferred into the metal reactor. After purging with Ar for 10 min,<br />

3 bar <strong>of</strong> hydrogen pressure was applied under constant stirring. After 3 hours pressure<br />

was released <strong>and</strong> the mixture was transferred into 500 ml cyclohexene <strong>and</strong> stirred for<br />

24 h, then stirring was stopped <strong>and</strong> after 10 hours a precipitate was collected <strong>and</strong><br />

washed with cyclohexene (200 ml), then dried in air at room temperature. The washing<br />

procedure was repeated one more time using 50 ml THF only, for complete removal <strong>of</strong><br />

side products. Obtained black precipitate was dried in air <strong>and</strong> transferred into 40 ml<br />

AcOH (sample was completely dissolved). After stirring (30 min), the mixture was<br />

transferred (Bönnemann’s procedure) into the 500 ml saturated NaHCO 3 water<br />

solution, after 2 h filtrated with G4 glass filter, then washed with 200 ml half saturated<br />

water solution <strong>of</strong> NaHCO 3 , then with water. Then it was picked up by AcOH (10-15<br />

ml) (with few drops <strong>of</strong> water) <strong>and</strong> left for drying on the glass plate in air at room<br />

temperature. Finally a black powder was scratched from the glass.<br />

The thermal activation <strong>of</strong> conventional Pd/Al 2 O 3 was performed at 400 °C under<br />

vacuum (10 -3 mbar) for 1 hour, then kept for 3 hours under a flowing mixture <strong>of</strong> N 2<br />

(93%) <strong>and</strong> H 2 (7%) gases (100 ml min -1 ) <strong>and</strong> subsequently cooled to the room<br />

temperature. The required amount was then weighed (in open atmosphere) <strong>and</strong><br />

immediately transferred into a glass reactor with acetic acid, cinchonidine <strong>and</strong> ethyl<br />

pyruvate.<br />

Characterization<br />

Molecular modelling, GC, XRD spectra measuring <strong>and</strong> DRIFTS investigations have<br />

been done as it is described e.g. in the chapter 4 <strong>and</strong> 5.<br />

Hydrogenation <strong>of</strong> ethyl pyruvate (EP) <strong>and</strong> ethyl benzoylformate (EB) was conducted in<br />

the metal reactor under 10 bars <strong>of</strong> hydrogen. Typically, to 70 ml AcOH with 36 mmol<br />

substrate <strong>and</strong> (unless otherwise mentioned) 20 mg cinchonidine (1 mM) <strong>and</strong> 20 mg<br />

catalyst was added. In the case <strong>of</strong> colloidal catalyst the mixture was sonicated for 10<br />

min. Finally it was transferred into the metal reactor, purged with Ar for 10 min <strong>and</strong><br />

hydrogen pressure (10 bar) has been set, the stirring frequency was 1000 rpm. In the<br />

case <strong>of</strong> the conventional Pt/Al 2 O 3 (was used for comparison) <strong>and</strong> Pd/Al 2 O 3 the catalyst<br />

was thermally activated under hydrogen (as described in details in the e.g. chapter 5)<br />

before use.<br />

68


O<br />

OH<br />

OH<br />

O<br />

O<br />

O<br />

Catalyst, AcOH<br />

O<br />

P H2 = 10 bar<br />

O<br />

O<br />

O<br />

OH<br />

OH<br />

O<br />

O<br />

O<br />

Ph<br />

Ph<br />

Ph<br />

O<br />

O<br />

O<br />

Fig. 6-1. Scheme <strong>of</strong> ethyl pyruvate <strong>and</strong> ethyl benzoylformate hydrogenation.<br />

6.3 Results <strong>and</strong> discussion<br />

XRD spectrum <strong>of</strong> cinchonidine modified Pd nanoclusters (Fig. 6-2) shows strong peaks<br />

at ~40° which corresponds to the (111) crystal face <strong>and</strong> small peak at 67° (220) face,<br />

according to work <strong>of</strong> Haglund [218].<br />

400<br />

300<br />

200<br />

100<br />

0<br />

-100<br />

20 30 40 50 60 70 80<br />

2Θ<br />

Fig. 6-2. XRD spectrum <strong>of</strong> cinchonidine modified Pd nanoclusters.<br />

The average crystal size has been estimated to be 1.6 nm based on the Scherrer’s model<br />

[177] from the width on a half height <strong>of</strong> the strongest peaks (2Θ=40°). The TEM image<br />

(Fig. 6-3) shows particles with 1.4 <strong>and</strong> 2.9 nm size. It is important to note, that at close<br />

examination it is seen that “big” particles consist <strong>of</strong> at least two “small” particles. The<br />

average size <strong>of</strong> “small” particles is very similar to the size obtained from XRD analysis<br />

<strong>and</strong> in further we consider 1.4 nm as the average size <strong>of</strong> basic Pd nanoclusters.<br />

69


Fig. 6-3. TEM image <strong>of</strong> cinchonidine modified Pd nanoclusters.<br />

Determination <strong>of</strong> adsorption geometry <strong>of</strong> cinchonidine on Pd nanoclusters via DRIFTS<br />

It is convenient to perform FTIR investigations <strong>of</strong> cinchonidine modified Pd<br />

nanoclusters through direct comparison with the spectrum <strong>of</strong> cinchonidine modified Pt<br />

nanocluster (considered in the chapter 4). Fig. 6-4 shows that DRIFT spectra <strong>of</strong><br />

cinchonidine modified Pt <strong>and</strong> Pd nanoclusters are very similar, however some<br />

distinctions are present <strong>and</strong> they are discussed below. The most important peaks <strong>and</strong><br />

shoulders (from both DRIFT spectra) for determination the adsorption modes <strong>of</strong><br />

cinchonidine are summarised in the Table 6-1.<br />

Table 6-1. Assignment <strong>of</strong> some frequencies (cm -1 ) from DRIFT spectra <strong>of</strong> cinchonidine<br />

modified Pt <strong>and</strong> Pd nanoclusters.<br />

Cinchonidine on Pd Cinchonidine on Pt Free cinchonidine Direction <strong>of</strong> dipole<br />

moment A .<br />

n.o. 1637 (sh) 1636 (med) (0, 0, 1)<br />

1613 (sh) 1613 (sh) 1619 (wk) (-1, 1, 0)<br />

1586 (med) 1594 (med) 1590 (st) (1, 1, 0)<br />

1570 (sh) 1576 (sh) 1567 (med) (1, -1, 0)<br />

1515 (st) 1514 (st) 1507 (st) (-1, 1, 0)<br />

1462 (med) 1463 (st) 1452 (st) (-1, 1, 0)<br />

811 (wk) 812 (med) 806 (med) (1, -1, -1)<br />

768 (med) 766 (str) 758 (st) (0, 0, 1)<br />

687 (st) 697 (med) 665 (wk) (-1, 1, 0)<br />

652 (sh) 655 (st) 648 (wk) or<br />

638 (med)<br />

(1, 1, -1) or<br />

(1, 0, -1)<br />

Comments: A – Orientations <strong>of</strong> dipole moments were characterized as directions <strong>of</strong> the<br />

unit vector (X, Y, Z) in the coordinate system having X <strong>and</strong> Y axes along the long <strong>and</strong><br />

short axis <strong>of</strong> the quinoline <strong>and</strong> Z axis along the normal <strong>of</strong> the quinoline plane towards<br />

to quinuclidine.<br />

70


The presence <strong>of</strong> in plane vibrations <strong>of</strong> cinchonidine on Pd nanoclusters e.g. at 1515 cm -<br />

1 clearly indicates the tilted adsorption mode. The out <strong>of</strong> plane vibrations <strong>of</strong> quinoline<br />

moiety e.g. 811 cm -1 were also observed on Pd that confirms the flat adsorption mode.<br />

The proposed adsorption modes (schematically shown in Fig. 4-10 <strong>and</strong> 4-11) <strong>of</strong> the<br />

cinchonidine on Pd nanoclusters (1.4 nm) were found to be very similar to the<br />

adsorption modes determined from the work <strong>of</strong> Ferri, Burgi <strong>and</strong> Baiker [82] over<br />

macroscopic film <strong>of</strong> Pd on Al 2 O 3 .<br />

1609<br />

1640<br />

3<br />

2<br />

1<br />

1600 1400 1200 1000 800<br />

Wavenumber, cm -1<br />

Fig. 6-4. DRIFT spectra <strong>of</strong> free (1) <strong>and</strong> adsorbed on Pd (2) <strong>and</strong> Pt (3) nanoclusters<br />

cinchonidine. Intensity <strong>of</strong> spectrum <strong>of</strong> free cinchonidine was decreased for better<br />

performance.<br />

Comparison <strong>of</strong> adsorption modes <strong>of</strong> cinchonidine on Pt <strong>and</strong> Pd nanoclusters<br />

The relative intensity <strong>of</strong> the peaks (811 cm -1 <strong>and</strong> 768 cm -1 ) corresponding to the out <strong>of</strong><br />

plane vibrations <strong>of</strong> quinoline ring is significantly lower with respect to the cinchonidine<br />

on Pt system, probably due to the lower concentration <strong>of</strong> the π-bonded cinchonidine<br />

with respect to the tilted mode. In fact, in the preparation <strong>of</strong> the samples the used ratio<br />

cinchonidine/Pd was 3, whereas in the case <strong>of</strong> Pt based sample the ratio cinchonidine/Pt<br />

was 2, at the same time it is known that with increase <strong>of</strong> concentration <strong>of</strong> cinchonidine<br />

the chance <strong>of</strong> flat adsorption decreases, as was shown for Pt [76] <strong>and</strong> Pd [82]. It is also<br />

can be expressed as a favour <strong>of</strong> tilted adsorption on Pd with respect to Pt, due to the<br />

nature <strong>of</strong> the adsorption <strong>of</strong> cinchonidine on metal.<br />

It is interesting to note the absence <strong>of</strong> the peaks (or shoulder) at 1640 cm -1 which<br />

corresponds to the stretching <strong>of</strong> C=C bond, that might indicate to its hydrogenation.<br />

However the parallel to the Pd surface orientation <strong>of</strong> this bond due to the specific<br />

adsorption mode <strong>of</strong> cinchonidine can not be fully excluded, especially at high<br />

concentration <strong>of</strong> modifier, whose adsorption modes are less studied in high<br />

concentration range.<br />

Hydrogenation <strong>of</strong> α-ketones<br />

71


From the results <strong>of</strong> EP <strong>and</strong> EB hydrogenations which are summarized in the Table 6-2,<br />

it is seen that presence <strong>of</strong> the cinchonidine (entry 2) results in 14 % <strong>of</strong> conversion (in<br />

contrast with 0 % without cinchonidine, entry 1) <strong>and</strong> in excess <strong>of</strong> R lactate (37 % ee). It<br />

is clear that cinchonidine induces a reaction rate <strong>and</strong> positive ee (excess <strong>of</strong> R<br />

enantiomer) similarly to the Pt catalyst, but reaction is significantly less intensive over<br />

Pd/Al 2 O 3 , In fact only 14 % <strong>of</strong> conversion has been achieved after 19 hours (entries 1<br />

<strong>and</strong> 2), whereas under the same conditions Pt/Al 2 O 3 results 91 % ee (R lactate) <strong>and</strong><br />

80% conversion after ~60 min (chapter 5). The different situation was found with<br />

cinchonidine modified Pd nanoclusters (entry 3), this catalyst without presence <strong>of</strong> free<br />

cinchonidine demonstrates 11 % enantiomeric excess in favour <strong>of</strong> S lactate <strong>and</strong> high<br />

conversion after 60 min. Surprisingly the normalized on number <strong>of</strong> metal surface atoms<br />

reaction rates were found to be almost the same as for conventional Pd/Al 2 O 3 , however<br />

the direct comparison between colloidal <strong>and</strong> thermally activated catalyst is laboured<br />

due to the presence <strong>of</strong> impurities blocking active sites.<br />

Conv. B , %<br />

EP ≈0 ≈0 0 D (75) -<br />

EP ≈0.2 ≈0.1 1.3 D (90) -<br />

EP ≈0 ≈0 0 D (60) -<br />

Table 6-2. Results <strong>of</strong> ethyl pyruvate <strong>and</strong> ethyl benzoylformate hydrogenation over Pd<br />

based catalysts.<br />

Run Catalyst Substra Initial rate, Initial<br />

Ee, %<br />

(min) -1 (time, enantiomer)<br />

te mmol·(min·g) -1 rate A after (dominated<br />

min)<br />

1<br />

C<br />

Pd/Al 2 O 3 EP ≈0 ≈0 ~0 (16 h) -<br />

2 Pd/Al 2 O 3 EP 0.36 9.8 14 D (19 h) 37 (R)<br />

3 Cinchonidine EP 13 9.6 59 E (60) 11 (S)<br />

on Pd coll. C<br />

4 Cinchonidine EP 9 6.7 71 E (120) 19 (S)<br />

on Pd coll.<br />

5 Diphos on<br />

Pd coll. C,F<br />

6 Diop on Pd<br />

coll. C,F<br />

7 Quiphos on<br />

Pd coll. C,F<br />

8 Cinchonidine EB 3.9 2.9 61 E (280) 0<br />

on Pd coll.<br />

9 Pt/Al 2 O 3 EB ≈0 ≈0 0 D (19 h) -<br />

Comments: A – Values were estimated based on TEM, metal content <strong>of</strong> the samples<br />

<strong>and</strong> chemisorption data, B - initial amount <strong>of</strong> racemic ethyl lactate (as impurity in<br />

commercial ethyl pyruvate) was taken into account, C – no free cinchonidine was<br />

added, D – according to the curve pr<strong>of</strong>ile (conversion vs. time) reaction can not go<br />

further, E - according to the curve pr<strong>of</strong>ile (conversion vs. time) reaction can go further,<br />

F – the sample was show for comparison, detailed information about it can be found in<br />

the corresponding chapter.<br />

Addition <strong>of</strong> free cinchonidine (entry 4, 1 mM) increases the enantiomeric excess to the<br />

maximum reported value <strong>of</strong> 19 % with respect to the S enantiomer, but practically does<br />

not influence on the reaction rate. The change <strong>of</strong> direction <strong>of</strong> enantioselectivity towards<br />

to the excess <strong>of</strong> S enantiomer over cinchonidine modified Pd nanoclusters in particular<br />

(entries 3 <strong>and</strong> 4) <strong>and</strong> over different types <strong>of</strong> Pd catalysts in general [217] can be<br />

explained through formation <strong>of</strong> enol intermediate state in a rate-determining step (Fig.<br />

72


6-5) [87], the latter was verified after reaction with deuterium yielded the product<br />

exchanged at the methyl group near α-keto group, i.e. giving the product<br />

CX 3 CX(OX)COOCH 3 (where X = H or D) [219]. It is assumed that due to the<br />

intermolecular repulsion between the >C=C< <strong>and</strong> >C=O moiety <strong>of</strong> the adsorbed state 3<br />

(Fig. 6-5) can be released by rotation around the C-C bond to give species 3´ which<br />

yields S enantiomer upon hydrogenation. According to this mechanism, these is not any<br />

supposed enhancement <strong>of</strong> the reaction rate (it is assumed that intermediate state 2<br />

which is involved in the rate=determining step 2-3, can not participate in hydrogen<br />

bonding making semi hydrogenated complex with cinchonidine). However Pd colloids<br />

modified with other chiral <strong>and</strong> non chiral lig<strong>and</strong>s (entries 5, 6 <strong>and</strong> 7) demonstrate very<br />

low (practically zero) reaction rate. The similar effect <strong>of</strong> enhancement <strong>of</strong> reaction rate<br />

was observed by Wells <strong>and</strong> co-workers [217] over Pd colloids but was not explained. In<br />

the current thesis we also do not give exact explanation, because this problem requires<br />

addition investigation that can not be covered by current PhD thesis having other aims,<br />

we give only an idea <strong>of</strong> interpretation <strong>of</strong> the effect which can be found below.<br />

O OEt<br />

1<br />

*<br />

H 3 C O<br />

*<br />

+H -H<br />

2<br />

O<br />

*<br />

H 2 C<br />

OEt<br />

* O<br />

+H slow<br />

3<br />

HO OEt<br />

H 2 C * * O<br />

+2H<br />

fast<br />

HO<br />

H<br />

H 3 C<br />

OEt<br />

O<br />

*<br />

(R)<br />

fast<br />

H 2 C OEt<br />

+2H<br />

H OEt<br />

HO<br />

(S)<br />

3' *<br />

HO O<br />

*<br />

fast<br />

H 3 C O<br />

Fig. 6-5. Mechanism for the enantioselective hydrogenation <strong>of</strong> the EP over Pd, adapted<br />

from [87].<br />

It is very interesting to note, that in our experiments (entries 2 <strong>and</strong> 4) under the same<br />

reaction conditions enantiomeric excess in favour <strong>of</strong> R enantiomer (37 %) <strong>and</strong> in favour<br />

<strong>of</strong> S enantiomer (19 %) was found for unmodified in advance (modified in situ)<br />

Pd/Al 2 O 3 <strong>and</strong> already modified with cinchonidine Pd nanoclusters, respectively. The<br />

similar inverse <strong>of</strong> ee with cinchonidine Pd system has been observed for different type<br />

<strong>of</strong> Pd catalysts, solvents <strong>and</strong> substrates [217]. This clearly indicates on the fact that<br />

there are competing mechanisms <strong>of</strong> hydrogenations e.g. the presence <strong>of</strong> both enols <strong>and</strong><br />

keto forms on the Pd surface. The keto- mechanism results in excess <strong>of</strong> R product (in<br />

the similar way to the cinchonidine-Pt system, probably with rate enhancement) <strong>and</strong><br />

enol- mechanism gives excess <strong>of</strong> S enantiomer as it is shown in Fig. 6-5 without rate<br />

73


enhancement. However the exact factor (nature <strong>of</strong> solvent, Pd catalyst (e.g. presence <strong>of</strong><br />

Pd δ+ ), substrate, hydrogen pressure or present <strong>of</strong> chiral or non chiral impurities) playing<br />

a major role in domination <strong>of</strong> one mechanism with respect to another one is not known<br />

still. As was mentioned above the origin <strong>of</strong> the rate enhancement is also unknown <strong>and</strong> it<br />

can not be explained through domination <strong>of</strong> the keto – mechanism because it will give<br />

an excess <strong>of</strong> R lactate. Our idea is based on another adsorption mode <strong>of</strong> cinchonidine<br />

on Pd surface i.e. via its double carbon bond (Fig. 6-6-right).<br />

Fig. 6-6. Proposed adsorption modes: π-bonded (left) <strong>and</strong> via C=C bond (right) <strong>of</strong><br />

cinchonidine on Pd surface results in different asymmetrical environments caused by H<br />

<strong>and</strong> OH groups.<br />

In fact, in this adsorption mode (opposite to adsorption via quinoline moiety)<br />

cinchonidine might behave in the similar manner to cinchonine, i.e. having replaced<br />

OH <strong>and</strong> H groups, thus induces excess <strong>of</strong> S enantiomer <strong>and</strong> rate enhancement (third<br />

mechanism). The proposed adsorption mode probably is not stable on the metal surface<br />

due the C=C bond hydrogenation. Probably because <strong>of</strong> the addition <strong>of</strong> free cinchonidine<br />

(entry 4) ee has increased from 11 to 19 % in favour <strong>of</strong> S enantiomer.<br />

Interestingly that hydrogenation <strong>of</strong> ethyl benzoylformate (where formation <strong>of</strong> enole<br />

state is not possible in principal, thus enol- mechanism is forbidden) with the same<br />

cinchonidine modified Pd nanoclusters results in zero enantiomeric excess. This means<br />

that the reaction was performed racemically (on unmodified sites) or equal contribution<br />

from keto- mechanism (yielding excess <strong>of</strong> R lactate) <strong>and</strong> the third mechanism (excess<br />

<strong>of</strong> S lactate). At the same time, conventional activated Pt/Al 2 O 3 (entry 9) with presence<br />

<strong>of</strong> free cinchonidine (1 mM) showed no conversion <strong>of</strong> EB in contrast with cinchonidine<br />

modified Pd colloids under 10 bars <strong>of</strong> hydrogen.<br />

The observations show that the Pd catalyzed enantioselective hydrogenation <strong>of</strong><br />

pyruvate esters is much more complicated system than the corresponding Pt catalyzed<br />

reaction <strong>and</strong> co adsorption <strong>of</strong> modifier, solvent <strong>and</strong> substrate [217] <strong>and</strong> possible<br />

impurities on Pd surface might be critical in determining the direction <strong>of</strong><br />

enantioselectivity. The Pd cinchona system requires more investigation for concluding<br />

reaction mechanisms especially in the light <strong>of</strong> recent work <strong>of</strong> Garl<strong>and</strong> et al. [220] where<br />

74


inverse <strong>of</strong> ee in hydrogenation <strong>of</strong> EB <strong>and</strong> EP was observed even over Pt/Al 2 O 3 <strong>and</strong> Pt/C<br />

catalysts.<br />

6.4 Summary<br />

The orientation <strong>of</strong> the cinchonidine on the surfaces <strong>of</strong> 1.4 nm Pd nanoclusters has been<br />

investigated for the first time using a combination <strong>of</strong> DRIFTS <strong>and</strong> molecular modeling<br />

methods. It was found that both ‘flat’ <strong>and</strong> ‘tilted’ modes <strong>of</strong> adsorbed cinchonidine are<br />

present on Pd nanoclusters having reduced relative concentration <strong>of</strong> flat bonded<br />

cinchonidine with respect to the cinchonidine modified Pt nanoclusters having the<br />

similar size (1.4 nm).<br />

The highest known enantiomeric excess <strong>of</strong> 19 % with respect to the S ethyl lactate was<br />

reported over cinchonidine modified Pd nanoclusters whereas under the same reaction<br />

conditions enantiomeric excess induced by conventional Pd/Al 2 O 3 modified with<br />

cinchonidine in situ was found to be 37 % with excess <strong>of</strong> R ethyl lactate.<br />

The inverse <strong>of</strong> enantioselectivity with respect to the Pt catalyzed hydrogenation was<br />

explained with literature model through the enol- mechanism, whereas rate<br />

enhancement effect <strong>and</strong> dependence <strong>of</strong> direction <strong>of</strong> enantiomeric excess on the nature<br />

<strong>of</strong> Pd catalyst <strong>and</strong> reaction conditions require additional investigation.<br />

Chapter 7<br />

Quiphos modified Pt <strong>and</strong> Pd colloidal<br />

nanoparticles<br />

7.1 Introduction<br />

The Quinoline diazaphospholidine ‘quiphos’ (Fig. 7-1) family <strong>of</strong> lig<strong>and</strong>s represents a<br />

possible alternative to the cinchonidine class <strong>of</strong> lig<strong>and</strong>s for application in<br />

enantioselective catalysis. Quiphos is a chiral molecule because <strong>of</strong> the chiral<br />

phosphorus atom due to the inversion high barrier at room temperature. In this work we<br />

used enantiopure quiphos that was confirmed from synthesis <strong>and</strong><br />

31 P-NMR<br />

spectroscopy, for the further information about its synthesis we recommend to contact<br />

to Dr. Didier Nuel <strong>of</strong> Université Aix-Marseille III, France.<br />

These lig<strong>and</strong>s have attracted attention as a result <strong>of</strong> their ease <strong>of</strong> synthesis <strong>and</strong> stability<br />

to air <strong>and</strong> moisture as well as possessing the same quinoline anchor as the cinchonidine<br />

systems <strong>and</strong> thus may be able to assume a flat “π-bonded” geometrical orientation<br />

which is known to be necessary for inducing enantioselectivity in the cinchonidine<br />

system.<br />

N<br />

N<br />

N<br />

P<br />

O<br />

Fig. 7-1. Quiphos molecule.<br />

75


Further, this family <strong>of</strong> man-made lig<strong>and</strong>s <strong>of</strong>fers numerous possibilities for structural<br />

modification which may enable ‘tuning’ <strong>of</strong> the resulting system <strong>and</strong> exploration <strong>of</strong> a<br />

wide variety <strong>of</strong> derivitization. This family <strong>of</strong> lig<strong>and</strong>s has also been successfully applied<br />

to a variety <strong>of</strong> transition metal catalysts for homogeneous enantioselective synthesis [9,<br />

11, 12].<br />

7.2 Experimental<br />

Materials<br />

Quiphos lig<strong>and</strong> was provided by Pr<strong>of</strong>. Gerard Buono <strong>and</strong> Dr. Didier Nuel <strong>of</strong> Université<br />

Aix-Marseille III, France. Other chemicals were purchased <strong>and</strong> used as received.<br />

Samples preparation<br />

The modification <strong>of</strong> conventional Pt/Al 2 O 3 with quiphos was performed in the similar<br />

way, how it was described for cinchonidine in the chapter 4. Typically, 40 mg Pt/Al 2 O 3<br />

was heated at 200-220 ºC under vacuum (~10 -2 mm) for 2-3 hours, then the gases N 2<br />

(93%) <strong>and</strong> H 2 (7%) were passed (50 ml/min) over the support for 2-3 hours without<br />

contact with air at the same temperature. While still under flowing gases, the support<br />

was cooled to room temperature <strong>and</strong> lig<strong>and</strong> was added as a solution in CHCl 3 (6.0 ml, 3<br />

mM). Then the system was kept at 10-15 ºC for one day, washed 3 times with CHCl 3<br />

<strong>and</strong> finally dried under vacuum at room temperature.<br />

For the preparation <strong>of</strong> quiphos stabilized Pt nanoclusters 70 ml <strong>of</strong> chlor<strong>of</strong>orm was<br />

flushed with nitrogen for 10 minutes, then 405 mg <strong>of</strong> Pt 2 (DBA) 3 (0.74 mmol <strong>of</strong> Pt) was<br />

added together with 775 mg <strong>of</strong> quiphos (2.22 mmol). The solution was flushed with Ar<br />

for 15 min under stirring. Then the contents were transferred into the glass reactor <strong>and</strong><br />

3 bar <strong>of</strong> hydrogen pressure was applied under constant stirring. After 18 hours a black<br />

reaction mixture was precipitated by adding the reaction mixture to 230 ml <strong>of</strong><br />

cyclohexene. The precipitate was filtrated out using a G4 glass filter. The precipitate<br />

was first washed with cyclohexene in chlor<strong>of</strong>orm (1/1 by volume) mixture then with<br />

chlor<strong>of</strong>orm (300 ml) until the product was free from hydrogenated DBA <strong>and</strong> unreduced<br />

Pt 2 (DBA) 3 as determined by IR spectroscopy. The sample is readily redispersible in a<br />

solution <strong>of</strong> phenol/ethanol (1:1) with 10% water.<br />

In order to prepare quiphos stabilized Pd nanoparticles 1084 mg Pd(DBA) 2 (1.9 mmol<br />

Pd) <strong>and</strong> 659 mg quiphos (1.9 mmol) were dissolved in nitrogen flushed THF (80 ml).<br />

Then the content was transferred into the metal reactor <strong>and</strong> 5 bar <strong>of</strong> hydrogen pressure<br />

was applied for 4 hours under constant stirring. Then the hydrogen pressure was<br />

released, reactor unplugged <strong>and</strong> the mixture was transferred into 300 ml cyclohexene.<br />

After 24 hours <strong>of</strong> settling, black precipitate was filtrated by G4 filter <strong>and</strong> washed with<br />

THF : cyclohexene (1:5) mixture until washing solvents mixture no longer became<br />

yellow. Finally sample was washed with cyclohexene (200 ml) <strong>and</strong> black powder was<br />

dried in air at room temperature for 12 hours.<br />

Characterization<br />

The samples were characterized by XRD, TEM, elemental analysis <strong>and</strong> DRIFTS in the<br />

same ways as it is described e.g. in the chapter 4. However due to the limitations in<br />

accessibility to TEM <strong>and</strong> elemental analysis apparatuses the corresponding<br />

experiments: TEM characterization <strong>of</strong> quiphos modified Pd nanoclusters <strong>and</strong> elemental<br />

analysis <strong>of</strong> this sample are missing.<br />

Hydrogenation <strong>of</strong> ethyl pyruvate was carried out in a metal reactor (Parr Instrument<br />

GmbH) under a pressure <strong>of</strong> 10 bar <strong>of</strong> hydrogen pressure at room temperature (22±1<br />

°C). The stirring speed was 1000 rpm.<br />

76


Typical conditions were such that 4 ml (36 mmol) ethyl pyruvate was dissolved in 70<br />

ml THF solvent, when mentioned 24 mg (68 mmol) <strong>of</strong> free quiphos was added. 20 mg<br />

<strong>of</strong> catalyst was added <strong>and</strong> mixture was sonicated for 10 min then transferred into the<br />

reactor, flushed with Ar for 10 min <strong>and</strong> finally, after the pressure <strong>of</strong> hydrogen had been<br />

stabilized for 10-15 sec, the initial time (t=0) was set <strong>and</strong> the reaction kinetics were<br />

monitored. About one ml aliquots <strong>of</strong> the reaction mixture were taken at certain time<br />

intervals, filtrated with a G4 glass filter <strong>and</strong> 1 μl was injected into the gas<br />

chromatograph (GC). GC measurements were taken using a Varian 3900 instrument<br />

with FID detector <strong>and</strong> Lipodex-E column.<br />

Activation <strong>of</strong> conventional Pt/Al 2 O 3 has bee performed in the same was as is described<br />

in the chapter 5.<br />

7.3 Results <strong>and</strong> discussion<br />

From the elemental analysis <strong>of</strong> quiphos modified Pt sample it has been found that the<br />

sample has 4.5 % N <strong>and</strong> 28.3 % C or 37 % <strong>of</strong> quiphos as an average by carbon <strong>and</strong><br />

nitrogen <strong>and</strong> 56 % <strong>of</strong> metal (based on TGA), whereas metal content for quiphos<br />

modified Pd sample was found to be 35 % (TGA).<br />

XRD spectra <strong>of</strong> quiphos modified Pt <strong>and</strong> Pd nanoclusters clearly demonstrate Pt <strong>and</strong> Pd<br />

crystalline nature <strong>of</strong> the samples. In fact, both spectra have a strong peak at 39°<br />

corresponding (111) crystal face. In case <strong>of</strong> Pt small peaks at 46° <strong>and</strong> 67° correspond to<br />

(200) <strong>and</strong> (220) faces, according to work <strong>of</strong> Swanson et al. [221]. The asymmetrical<br />

peak at ~35° is difficult to assign unambiguously, since it appears not only in the<br />

spectra <strong>of</strong> this sample but also in case <strong>of</strong> diphos on Pd nanoclusters (see chapter 9) we<br />

believe that it is not due to the instrument fault. Here we would like to assign it to the<br />

(222) Pt crystal face base on work <strong>of</strong> Hull [222].<br />

350<br />

300<br />

250<br />

1<br />

200<br />

150<br />

100<br />

50<br />

2<br />

0<br />

-50<br />

20 30 40 50 60 70 80<br />

2Θ<br />

Fig. 7-1. XRD <strong>of</strong> quiphos modified Pt (1) <strong>and</strong> Pd (2) nanoclusters.<br />

XRD spectrum <strong>of</strong> quiphos modified Pd nanoclusters besides <strong>of</strong> mentioned already peak<br />

at 39° ((111) face) shows wide peaks at 22° (200) [222] <strong>and</strong> 71° (620) [176].<br />

77


Fig. 7-2. TEM <strong>of</strong> quiphos modified Pt nanoclusters.<br />

TEM image <strong>of</strong> quiphos modified Pt nanoparticles (Fig. 7-2) shows that the average<br />

particle’s size is 15 ± 0.5 nm nm, that differs (Table 7-1) from average size estimated<br />

from the Scherrer equation (4.3 nm) <strong>and</strong> from the average size <strong>of</strong> quiphos modified Pd<br />

nanoparticles (1.2 nm), probably due to the partial agglomeration <strong>of</strong> Pt nanoparticles.<br />

Table 7-1. Comparison <strong>of</strong> the sizes <strong>of</strong> quiphos modified Pt <strong>and</strong> Pd particles found via<br />

TEM <strong>and</strong> XRD.<br />

Sample , nm (TEM) , nm (XRD)<br />

Quiphos on Pt 15 4.3<br />

Quiphos on Pd - 1.2<br />

FTIR investigation <strong>of</strong> quiphos adsorbed on Pt <strong>and</strong> Pd nanoparticles<br />

Two conformations <strong>of</strong> the quiphos molecule were found through molecular modelling<br />

(Fig. 7-3). The energy difference between the more stable conformation (Fig 7-3-1) <strong>and</strong><br />

the second (Fig. 7-3-2) was found to be 11.3 kcal/mol in air <strong>and</strong> 5.2 kcal/mol in THF.<br />

The geometry <strong>of</strong> the first conformation is in agreement with single-crystal XRD data<br />

[11] <strong>and</strong> was chosen for analysis.<br />

Fig. 7-3. Skeleton geometry <strong>of</strong> quiphos molecule in two conformations, hydrogen<br />

atoms are omitted. Conformation “1” is energetically more preferred, ΔE 1,2 = -11.3<br />

kcal/mol (in air).<br />

78


The DRIFT spectra <strong>of</strong> the quiphos adsorbed on conventional Pt/Al 2 O 3 (Fig. 7-4)<br />

demonstrates shifted peaks <strong>and</strong> shoulders whose analogs can be found in the spectra <strong>of</strong><br />

free quiphos. However due to the low quality (not all important peaks can be found) <strong>of</strong><br />

the spectra <strong>of</strong> this sample, probably because <strong>of</strong> low degree <strong>of</strong> surface coverage, this<br />

spectra was not taken in to account for further interpretation in favor <strong>of</strong> spectra <strong>of</strong><br />

quiphos modified nanoparticles (please see below).<br />

Relative units<br />

1<br />

2<br />

1600 1400 1200 1000<br />

Wavenumber, cm -1<br />

Fig. 7-4. DRIFT spectra <strong>of</strong> quiphos adsorbed on conventional Pt/Al 2 O 3 (1) <strong>and</strong> free<br />

quiphos (2).<br />

Adsorbed quiphos on Pt <strong>and</strong> Pd produces IR spectra (Fig. 7-5) for which the peaks are<br />

summarized in Table 7-2. Here, only peaks which can be unambiguously assigned to<br />

free quiphos are presented.<br />

In the same way as it is done in case <strong>of</strong> the cinchonidine, the DFT geometry<br />

optimization <strong>and</strong> calculation <strong>of</strong> the IR spectrum <strong>of</strong> quiphos have been done to obtain<br />

the orientation <strong>of</strong> the corresponding dipole moment. To facilitate spectra assignment,<br />

the experimental <strong>and</strong> calculated IR spectra <strong>of</strong> quiphos (correlation coefficient is 1.03)<br />

quinoline <strong>and</strong> aniline were compared with each other; see Fig. 7-6 <strong>and</strong> Table 7-2.<br />

Aniline was chosen since quiphos also contains a C 6 H 5 N unit. The IR spectrum <strong>of</strong><br />

aniline was calculated with correlation coefficient 1.09.<br />

All observed peaks corresponding to IR vibrations <strong>of</strong> quiphos adsorbed on Pt<br />

demonstrate two orientations <strong>of</strong> the dipole moment: mostly perpendicular to the<br />

quinoline ring <strong>and</strong> mostly perpendicularly to the phenyl ring. This leads us to propose<br />

two orientations <strong>of</strong> adsorbed quiphos on Pt. In the first mode, quiphos is adsorbed on Pt<br />

via the quinoline anchor (Fig. 7-7), in the similar way as cinchonidine does, but,<br />

possibly more tilted; in the second mode - via the phenyl group (Fig. 7-8) <strong>and</strong> possibly<br />

via 3,4 N <strong>and</strong>/or P atoms. Some degree <strong>of</strong> tilt is also possible in both cases. In contrast to<br />

the cinchonidine, quiphos can not be adsorbed in a tilted orientation mode, for example,<br />

through the nitrogen in quinoline part without significant changes in geometry.<br />

The DRIFT spectra <strong>of</strong> adsorbed quiphos on Pt <strong>and</strong> Pd nanoparticles do not have<br />

principal differences. In fact, positions <strong>of</strong> some peaks (see Table 7-2) are slightly<br />

shifted that can be interpreted in slight differences in force constants <strong>of</strong> chemical bonds<br />

79


<strong>of</strong> quiphos adsorbed on Pt <strong>and</strong> Pd <strong>and</strong> in the same time in differences in adsorption<br />

strength. It is also can be seen (Fig. 7-5 <strong>and</strong> Table 7-2) that some new peaks appeared<br />

or became stronger for quiphos being adsorbed on Pd with respect to Pt. This<br />

phenomenon can be explained as increase <strong>of</strong> quality <strong>of</strong> the spectra (for example due to<br />

the differences in size <strong>of</strong> Pt <strong>and</strong> Pd nanoparticles, or other secondary effects) or as<br />

differences in relative concentration <strong>of</strong> spices adsorbed via the quinoline <strong>and</strong> phenyl<br />

ring. It is difficult to conclude unambiguously which explanation is more valid here<br />

without addition investigation in which we decided to do not focus on here.<br />

1<br />

Relative units<br />

2<br />

3<br />

1600 1400 1200 1000 800<br />

Wavenumber, cm -1<br />

Fig. 7-5. IR spectrum <strong>of</strong> quiphos adsorbed on Pd (1) <strong>and</strong> Pt (2) nanoparticles <strong>and</strong> free<br />

quiphos (intensity is reduced for clarity) (3).<br />

80


3<br />

A<br />

Relative units<br />

2<br />

B<br />

1<br />

C<br />

0<br />

D<br />

1600 1400 1200 1000 800 600<br />

Wavenumber cm -1<br />

Fig. 7-6. Comparison <strong>of</strong> IR peaks between free cinchonidine (A), quinoline (B),<br />

quiphos (C) <strong>and</strong> aniline (D).<br />

Fig. 7-7. Flat adsorbed quiphos demonstrates vibration mode at 1604 cm -1 with dipole<br />

moment (big arrow) orientated mostly perpendicularly to the metal surface. The most<br />

intensive displacement <strong>of</strong> atoms is shown by small arrows.<br />

81


Fig. 7-8 Flat adsorbed quiphos via phenyl ring demonstrates vibration mode at 696 cm -1<br />

with dipole moment (big arrow) orientated mostly perpendicularly to the metal surface.<br />

The most intensive displacement <strong>of</strong> atoms is shown by small arrows.<br />

Table 7-2. Assignment <strong>of</strong> frequencies (cm -1 ) <strong>of</strong> IR spectra <strong>of</strong> adsorbed <strong>and</strong> free<br />

quiphos, free quinoline <strong>and</strong> free aniline.<br />

Quiphos<br />

on Pt<br />

Quiphos<br />

on Pd<br />

Free<br />

quiphos<br />

Quiphos<br />

theoretical<br />

Quinoline Aniline Description Orientation<br />

<strong>of</strong><br />

adsorbed<br />

- 1610 1652 1620 - C-Qu.<br />

strch., H-<br />

Qu. b.<br />

1604<br />

(st)<br />

1580<br />

(sh)<br />

1558<br />

(med)<br />

1503<br />

(st)<br />

1601<br />

(st)<br />

1574<br />

(med)<br />

1593 1650 - 1600<br />

6,7,9,10 C-A.,<br />

3,5 C strch.,<br />

ip. def.<br />

n.o. 1639 1596 - C-Qu.<br />

strch., H-<br />

Qu. b.<br />

- 1568 1598 1572 - C-Qu.<br />

strch., H-<br />

Qu. b.<br />

1534<br />

(med)<br />

1500<br />

(st)<br />

1538 - Unharmonic<br />

vibr.<br />

1498 1542 1501 1495 C-Qu.<br />

strch., H-<br />

Qu. b.<br />

quiphos<br />

-<br />

Quinoline<br />

flat<br />

Phenyl flat<br />

Phenyl flat<br />

-<br />

Phenyl flat<br />

1468 1468 1464 1502 1471 - H-Qu. Phenyl flat<br />

(med) (st)<br />

asym. ip. b.<br />

1422 1427 1422 1477 1432 1479 H-Qu. sym. Phenyl flat<br />

(wk) (sh)<br />

ip. b.<br />

1390 1388 1383 1434 1394 -<br />

37,42 H-Q. Phenyl flat<br />

82


(wk) (med) sym. ip. b.,<br />

C,<br />

C 16 N-<br />

Qu. strch.<br />

1379<br />

(sh)<br />

1375<br />

(med)<br />

1369 1408 1371 - H,C-Qu.<br />

comp.,<br />

30,31,32 H b.<br />

- 1324 1361 1314 1304 C-Qu.<br />

strch., H-<br />

Qu. ip. b.,<br />

C-A. strch.<br />

1309<br />

(st)<br />

1259<br />

(sm)<br />

1319<br />

(st)<br />

1258<br />

(sm)<br />

1311 1337 - 1273<br />

1247 1292 - -<br />

3 C 5 N strch.,<br />

C,H-A.<br />

comp.<br />

H 11,13,14 2 C<br />

b.<br />

1 O 23 C<br />

strch.,<br />

40,41,42 H ip.<br />

b., H 14 2 C b.<br />

- 1225 1258 - - H 11,12,13,14 2 C<br />

b.<br />

- 1202 1240 - 1171<br />

26,16 H-A,<br />

28,29 H-A<br />

sci.,<br />

H 11,12,13,14 2 C<br />

b.<br />

- 1087 1113 1118 - H,C-Qu.<br />

comp.,<br />

11,12 C strch.,<br />

- 1027 1027 - 992 H-Qu. oop.<br />

wag.,<br />

C 7 C,<br />

8,10 C 9 C ip.<br />

b.<br />

- 988 1014 - - C-A. sym.<br />

expan.,<br />

H 11,12,13,14 2 C<br />

b.<br />

- 971 989 - -<br />

- 954 951 - -<br />

13,14 C,<br />

12 C 4 N –<br />

syn. strch.,<br />

H 11,12,14 2 C<br />

b.<br />

13,14 C,<br />

12 C 4 N –<br />

asyn. strch.,<br />

H 11,12,14 2 C<br />

b.<br />

- 901 926 - - H,C-A., -<br />

Phenyl flat<br />

-<br />

Quinoline<br />

flat<br />

Phenyl flat<br />

-<br />

-<br />

-<br />

-<br />

-<br />

-<br />

-<br />

83


comp.,<br />

H 11,12,13,14 2 C<br />

b., 2 P 11 O 3 N,<br />

N 14 C 13 C b.<br />

- 887 912 - 871 C-Qu.<br />

comp. def.,<br />

H-A. oop.<br />

wag.<br />

- 864 884 - -<br />

12,13 C 14 C<br />

comp.,<br />

11,12 C strch.,<br />

3,4 NP asym.<br />

strch., C-A.<br />

comp.<br />

833 823 829 849 805 - C,H-Qu.<br />

(med) (med)<br />

oop. wag.<br />

-<br />

-<br />

Quinoline<br />

flat<br />

- 807 835 - - C-Qu. ip. b. -<br />

- 803<br />

(sm)<br />

793 812 786 - C,H-Qu.<br />

oop. wag.<br />

Quinoline<br />

flat<br />

- 760 780 739 - C,H-Qu. -<br />

oop. wag.<br />

758 (st) 755 (st) 746 782 - 745 C,H-A. oop. Phenyl flat<br />

wag.<br />

696 (st) 697 (st) 690 712 - 688 C,H-A. oop. Phenyl flat<br />

wag.<br />

- 669 682 - -<br />

7,9 C 8 C, -<br />

6,10 C 5 C ip.<br />

b. , 3 NP<br />

strch.<br />

Comments: The same as for the Table 4-3, plus the following n.o.- not observed, expan.<br />

– expansion, asym. – asymmetrical, syn. – synchronously, asyn. – asynchronously, A-<br />

aniline.<br />

The results <strong>of</strong> ethyl pyruvate hydrogenation with quiphos modified Pt, Pd nanoparticles<br />

<strong>and</strong> activated conventional Pt/Al 2 O 3 are summarized in the Table 7-3. As can be seen<br />

from the Table 7-3 the addition <strong>of</strong> free quiphos stops the reaction completely. In other<br />

words quiphos can be considered as a poison for this reaction with Pt or Pd based<br />

catalysts. Since conversion was not observed at all, ee can not be defined.<br />

Table 7-3. Results <strong>of</strong> ethyl pyruvate hydrogenation with quiphos modified Pt <strong>and</strong> Pd<br />

nanoclusters.<br />

Sample<br />

name<br />

Catalyst<br />

loading, mg.<br />

Reaction rate,<br />

mmol·(min·g) -<br />

Initial rate<br />

(min) -1 Conversion A , [Free<br />

% after time, quiphos],<br />

1<br />

min<br />

mM<br />

B<br />

Pt/Al 2 O 3 20 0.6 16.5 12.5 (345) C 0<br />

Pt/Al 2 O 3 20 ~0 ~0 ~0 (90) 1<br />

Quiphos on<br />

Pt<br />

nanoparticles<br />

20 ~0 ~0 ~0 (60) 0<br />

84


Quiphos on 20 ~0 ~0 ~0 (90) 0<br />

Pd<br />

nanoparticles<br />

Comments: A - initial amount <strong>of</strong> racemic ethyl lactate (as impurity in commercial ethyl<br />

pyruvate) was taken into account. B- racemic hydrogenation without a modifier, C –<br />

according to the curve development (conversion vs. time) reaction can go further.<br />

7.4 Summary<br />

It has been found that quiphos adsorbs on Pt nanoclusters in a similar mode as<br />

cinchonidine through the quinoline moiety. An additional adsorption mode <strong>of</strong> the<br />

quiphos is via the phenyl ring with possibly contribution from the 3,4 N <strong>and</strong>/or P atoms.<br />

The presence <strong>of</strong> quiphos in the reaction mixture induces strong poison effect that stops<br />

reaction completely, probably because <strong>of</strong> occupation <strong>of</strong> active surface sites by quiphos<br />

molecule.<br />

85


Chapter 8<br />

“Quiphos-spider” modified Pt colloidal<br />

nanoparticles<br />

8.1 Introduction<br />

The “quiphos-spider” (4-isopropyl-2 methyl cyclohexyl-1methyl-phosphonite -2-2'-1,5<br />

binaphtyl) compound is the second <strong>of</strong> three lig<strong>and</strong>s which is provided by our<br />

collaborators (Marseille). The chirality <strong>of</strong> this lig<strong>and</strong> is conditioned by the fact that both<br />

couple <strong>of</strong> aromatic rings do not lie in the one plane, making ~32° <strong>and</strong> the image mirror<br />

can not be superimposed with the original molecule.<br />

P O<br />

Fig. 8-1. Quiphos-spider.<br />

Here we investigate catalytic activity <strong>of</strong> Pt nanoparticles modified with quiphos-spider<br />

lig<strong>and</strong> in the ethyl pyruvate hydrogenation <strong>and</strong> explain obtained results based on<br />

proposed models <strong>of</strong> the molecule adsorption geometry.<br />

8.2 Experimental<br />

Materials<br />

Quiphos-spider was provided by Pr<strong>of</strong>. Gerard Buono <strong>and</strong> Dr. Didier Nuel <strong>of</strong> Université<br />

Aix-Marseille III, France. Other chemicals were purchased <strong>and</strong> used as received.<br />

Samples preparation<br />

Preparation <strong>of</strong> quiphos-spider modified Pt nanoclusters had been done according to the<br />

following procedure. 321 mg (0.73 mmol) quiphos-spider was dissolved in argon<br />

purged 70 ml THF with 400 mg Pt 2 (DBA) 3 (0.37 mmol Pt) the mixture was transferred<br />

in to the glass reactor <strong>and</strong> hydrogen pressure <strong>of</strong> 2 bar was applied under constant<br />

stirring at room temperature for 19 hours. Then the mixture was transferred into 400 ml<br />

cyclohexene, after 24 hours black-brown precipitate was collected <strong>and</strong> washed with<br />

(THF/cyclohexene mixture 1/3) 200 ml <strong>and</strong> then with 200 ml cyclohexene, finally<br />

sample was dried at 40°C over night.<br />

Due to the limited amount <strong>of</strong> provided quiphos-spider lig<strong>and</strong> Pt modification <strong>of</strong> Pt<br />

nanoclusters was done only.<br />

Characterization<br />

FTIR spectra were recorded on “Thermo Nicolet AVATAR – 370” instrument, with<br />

DRIFTS accessories, measuring KBr background first.<br />

TEM images were recorded according to the already described in the chapter 4<br />

procedure.<br />

86


Molecular modelling consisting <strong>of</strong> geometry optimization <strong>and</strong> IR spectra calculation<br />

has been done with help <strong>of</strong> G03 program [175] for quiphos-spider <strong>and</strong> naphthalene.<br />

Ethyl pyruvate hydrogenation (including activation <strong>of</strong> conventional Pt/Al 2 O 3 ) was<br />

performed according to the procedure described in the chapter 6. Important parameters<br />

are summarized in the Table 8-1.<br />

8.3 Results <strong>and</strong> discussion<br />

The average size <strong>of</strong> Pt nanoclusters was found to be 1.9±0.3 nm from TEM image (Fig.<br />

8-2). Particles were found to be not uniformly distributed; however no significant<br />

agglomeration was found.<br />

Fig. 8-2. TEM image <strong>of</strong> quiphos-spider modified Pt nanoclusters.<br />

As can be seen from comparison <strong>of</strong> DRIFT spectra <strong>of</strong> free <strong>and</strong> adsorbed on Pt<br />

nanoclusters quiphos-spider lig<strong>and</strong> relative intensities <strong>and</strong> positions <strong>of</strong> some peaks have<br />

changed. Since quiphos-spider molecule consists <strong>of</strong> two (not identical, due to their<br />

orientation within skeleton geometry <strong>of</strong> the lig<strong>and</strong>) naphthalene rings, assignment <strong>of</strong><br />

the main adsorption peaks can be done through comparison IR spectra <strong>of</strong> naphthalene<br />

<strong>and</strong> quiphos-spider. In fact in plane deformation <strong>of</strong> naphthalene ring(s) at 1595 cm -1 ,<br />

1389 cm -1 , 1267 cm -1 having dipole moment orientated along short axis <strong>of</strong> naphthalene<br />

<strong>and</strong> at 1509 cm -1 , 1209 cm -1 <strong>and</strong> 1013 cm -1 with dipole moment orientated along long<br />

its axis, as well as out <strong>of</strong> plane vibrations at 956 cm -1 , 790 cm -1 <strong>and</strong> 617 cm -1 have<br />

dipole moment along the normal vector <strong>of</strong> molecule’s plane. Since quiphos-spider<br />

molecule have two naphthalene planes <strong>and</strong> reduced symmetry, the one peak in the<br />

spectrum <strong>of</strong> naphthalene corresponds to two or more peaks (shoulders) in the spectrum<br />

<strong>of</strong> quiphos-spider. Due to the fact that vibrations <strong>of</strong> free lig<strong>and</strong>s having dipole moment<br />

orientated along short naphthalene axis <strong>and</strong> along long naphthalene axis <strong>and</strong> along<br />

normal vector <strong>of</strong> naphthalene planes are present as strong corresponding peaks in the<br />

spectrum <strong>of</strong> adsorbed quiphos-spider, we would like to propose the following<br />

adsorption models <strong>of</strong> the molecule (Fig. 8-5). In fact, both these models allow having<br />

non zero projection <strong>of</strong> every mentioned orientation <strong>of</strong> dipole moment along normal<br />

vector <strong>of</strong> the metal surface. It has to be mentioned that we do not exclude adsorption on<br />

corners or edges <strong>of</strong> a nanocluster however we think that molecules adsorb mostly on a<br />

87


flat crystal surface due to the higher number <strong>of</strong> atoms (making adsorption sites) on<br />

planes <strong>of</strong> nanocluster than on corners.<br />

Relative units<br />

1<br />

2<br />

1600 1400 1200 1000 800 600<br />

Wavenumber, cm -1<br />

Fig. 8-3. Drift spectra <strong>of</strong> quiphos-spider on Pt nanoclusters (1) <strong>and</strong> (2) free quiphosspider.<br />

1<br />

2<br />

3<br />

1600 1400 1200 1000 800 600 400<br />

Wavenumber, cm -1<br />

Fig. 8-4. Comparison <strong>of</strong> experimentally obtained IR spectra <strong>of</strong> free quiphos-spider (1),<br />

naphthalene (2) <strong>and</strong> calculated IR spectra (correlation coefficient 0.96) <strong>of</strong> free<br />

naphthalene (3). The huge peak at ~800 cm -1 is not shown fully. The similar vibration<br />

modes are marked.<br />

88


Fig. 8-5. Proposed adsorption models <strong>of</strong> quiphos-spider on Pt nanoclusters.<br />

The results <strong>of</strong> ethyl pyruvate hydrogenation with quiphos-spider modified Pt<br />

nanoparticles <strong>and</strong> activated conventional (unmodified) Pt/Al 2 O 3 are summarized in the<br />

Table 8-1. Quiphos-spider modified Pt nanoclusters sample demonstrates slow racemic<br />

hydrogenation <strong>of</strong> ethyl pyruvate, but higher normalized on gram <strong>of</strong> a catalyst reaction<br />

rate than unmodified activated with hydrogen (see experimental section) conventional<br />

Pt/Al 2 O 3 , however it shows lower reaction rate normalized on number <strong>of</strong> surface metal<br />

atoms, probably because some <strong>of</strong> them are “poisoned” by quiphos-spider lig<strong>and</strong>.<br />

Zero enantiomeric excess was found probably due to the inability <strong>of</strong> lone pair <strong>of</strong><br />

phosphorus to “catch” the adsorbed ethyl pyruvate (as it is required for cinchonidineethyl<br />

pyruvate system) because lone pair <strong>of</strong> phosphorus might be involved into the<br />

adsorption binding.<br />

Despite <strong>of</strong> the fact that the reaction rate normalized on gram <strong>of</strong> a catalyst was higher for<br />

colloidal sample, reaction rate normalized on estimated mole <strong>of</strong> Pt surface atoms was<br />

higher for conventional Pt/Al 2 O 3 , probably because the latter one was thermally<br />

activated under hydrogen.<br />

Initial rate A Conversion B , %<br />

Table 8-1. Results <strong>of</strong> ethyl pyruvate hydrogenation with quiphos-spider modified Pt<br />

nanoclusters.<br />

Sample name Catalyst Reaction rate,<br />

loading, mg. mmol·(min·g) -1 (min) -1 after time, min<br />

C<br />

Pt/Al 2 O 3 20 0.6 16.5 12.5 (345) D<br />

Quiphos-spider 17 2 3.3 17 (160) D<br />

on Pt<br />

nanoparticles<br />

Comments: A – values were estimated based on chemisorption data (Pt/Al 2 O 3 ), size <strong>of</strong><br />

Pt nanoclusters <strong>and</strong> metal content (30%), B - initial amount <strong>of</strong> racemic ethyl lactate (as<br />

impurity in commercial ethyl pyruvate) was taken into account. C- racemic<br />

hydrogenation without a modifier, D – according to the curve pr<strong>of</strong>ile (conversion vs.<br />

time) reaction can go further.<br />

8.4 Summary<br />

The adsorption models <strong>of</strong> quiphos spider on Pt nanocluster based on DRIFTS <strong>and</strong><br />

molecular modeling investigations has been proposed. According to these models<br />

quiphos-spider adsorbs via its P atom <strong>and</strong> <strong>of</strong> one naphthalene ring. This system was not<br />

found to be able to induce enantioselectivity in the hydrogenation <strong>of</strong> ethyl pyruvate<br />

(racemic reaction was observed) probably due to the inability <strong>of</strong> adsorbed quiphosspider<br />

to make semi-hydrogenated transition state by analogy with adsorbed<br />

cinchonidine-hydrogen-ethyl pyruvate system.<br />

89


Chapter 9<br />

Diphos <strong>and</strong> diop modified Pt <strong>and</strong> Pd<br />

colloidal nanoparticles<br />

9.1 Introduction<br />

In this chapter we used diphos <strong>and</strong> diop organophosphorus lig<strong>and</strong>s as modifiers for Pt<br />

<strong>and</strong> Pd nanoclusters. The chiral lig<strong>and</strong> diop attracts attention since it has been<br />

successfully used in enantioselective homogeneous catalysis [24], whereas non chiral<br />

diphos molecule has rather similar <strong>and</strong> simple structure that facilitates DRIFTS<br />

investigation <strong>of</strong> lig<strong>and</strong>s adsorbed on nanoclusters <strong>and</strong> relative comparisons e.g.<br />

catalytic properties <strong>of</strong> diphos <strong>and</strong> diop modified Pt <strong>and</strong> Pd nanoclusters.<br />

P<br />

P<br />

P<br />

P<br />

O<br />

O<br />

9.2 Experimental<br />

Diphos<br />

Diop<br />

Fig. 9-1. Diphos <strong>and</strong> Diop molecules.<br />

Materials<br />

Ethylene bis(diphenylphosphine) (diphos) 96 % (Aldrich), (4S,5S)-4,5-<br />

Bis(diphenylphosphinomethyl)-2,2-dimethyl-1,3-dioxolane ( or (S,S)-Diop) <strong>and</strong><br />

(4R,5R)-4,5-Bis(diphenylphosphinomethyl)-2,2-dimethyl-1,3-dioxolane ( or (R,R)-<br />

Diop) 98 % (Aldrich) <strong>and</strong> Pd(DBA) 2 (Aldrich) were used as received, Pt 2 (DBA) 3 was<br />

synthesize as described in the chapter 4.<br />

Samples preparation<br />

In order to prepare diphos on Pt nanoclusters, 560 mg Pt 2 (DBA) 3 (1 mmol Pt) in 40 ml<br />

THF was added to 400 mg diphos (1 mmol) in 30 ml THF. The mixture was purged<br />

with Ar for 10 min in metal reactor <strong>and</strong> hydrogen pressure <strong>of</strong> 4.5 bar was applied for<br />

2.5 hours under constant stirring. After that time, hydrogen pressure was realized <strong>and</strong><br />

the mixture was transferred into 400 ml cyclohexene, after 30 hour supernatant was<br />

decanted <strong>and</strong> precipitated black substance was transferred into 400 ml<br />

THF:cyclohexene mixture (1:7), stirred for 10 min <strong>and</strong> after 4 hour precipitate was<br />

collected <strong>and</strong> this procedure has been repeated until no trace <strong>of</strong> yellow DBA solution<br />

left. Finally black powder was washed with 400 ml cyclohexene <strong>and</strong> dried in air at 65<br />

°C.<br />

The diphos on Pd nanoclusters sample was prepared according to the following<br />

procedure. 1084 mg Pd(DBA) 2 (1.9 mmol Pd) in 40 ml THF was added to 752 mg (1.9<br />

mmol) diphos in 20 ml THF. The mixture was purged with Ar for 10 min in the metal<br />

reactor <strong>and</strong> then hydrogen (5 bar) was applied for 20 hours under constant stirring.<br />

90


After that time the mixture was transferred to 450 ml cyclohexene <strong>and</strong> after 10 hour<br />

black-brown precipitated was collected by G4 glass filter. 400 ml THF: cyclohexene<br />

(1:1) mixture <strong>and</strong> then 200 ml cyclohexene were used for washing. Finally sample was<br />

dried under vacuum at 40°C.<br />

In order to prepare diop on Pd nanoclusters 400 mg (0.8 mmol) (S, S) diop in 40 ml<br />

THF was mixed with 460 mg Pd(DBA) 2 (0.8 mmol Pd) in 40 ml THF in metal reactor.<br />

The mixture was purged with Ar for 10 min <strong>and</strong> then 5 bar <strong>of</strong> hydrogen pressure was<br />

applied for 6.5 hours. After that time, reaction mixture was transferred into 400 ml<br />

cyclohexene, after 16 hours black precipitated was collected <strong>and</strong> washed with 200 ml<br />

THF: cyclohexene mixture (1:4) <strong>and</strong> with 200 ml cyclohexene. Finally it was dried<br />

under vacuum for during 10 hour at 40 °C.<br />

(R, R) <strong>and</strong> (S, S) enantiomers <strong>of</strong> diop were used as modifiers for Pt nanocluster. 250<br />

mg (0.5 mmol) e.g. (S, S) diop in 35 ml THF was transferred into the metal reactor with<br />

274 mg Pt 2 (DBA) 3 in 35 ml THF. After 10 min for Ar purging hydrogen pressure <strong>of</strong> 4<br />

bar has been set for 5.5 hours. After that time the mixture was transferred into 400 ml<br />

cyclohexene, collected after 20 hours precipitate was washed with 400 ml THF:<br />

cyclohexene mixture (1:4) <strong>and</strong> with 200 ml cyclohexene. Finally sample was dried at<br />

40 °C under vacuum.<br />

Characterization <strong>and</strong> experiments<br />

Molecular modelling including geometry optimisation <strong>and</strong> IR spectra calculation has<br />

been performed with Gaussian 03 s<strong>of</strong>tware [175] for diop <strong>and</strong> diphos molecules by<br />

using the DFT method (b3lyp) <strong>and</strong> (6-31G) basis set. Visualisation <strong>of</strong> the geometry <strong>and</strong><br />

IR vibrations was performed using Gaussview s<strong>of</strong>tware. In order to facilitate assigment<br />

<strong>of</strong> the peaks <strong>of</strong> the experimental IR spectrum for the free molecule, its IR spectrum was<br />

compared (in terms <strong>of</strong> intensities <strong>of</strong> the signals <strong>and</strong> their vibrational frequencies) with<br />

calculated IR spectra using MS Excel <strong>and</strong> Origin (intensities <strong>of</strong> both spectra were<br />

normalized, frequencies <strong>of</strong> the calculated spectra were multiplied by a correlation<br />

coefficient). The correlation c<strong>of</strong>ficient dωtheory/dωexp <strong>of</strong> 1.02 was found to provide<br />

the best fit.<br />

XRD spectra were measured in the way described in the chapter 4.<br />

Hydrogenation <strong>of</strong> ethyl pyruvate <strong>and</strong> DRIFTS investigations have been performed<br />

according to the procedure described in details in the chapter 7 <strong>and</strong> 4 correspondingly.<br />

Due to the absence <strong>of</strong> TEM instrument in the IUB <strong>and</strong> limited access to the instrument<br />

outside the university, TEM investigation <strong>of</strong> diop modified Pt nanoclusters has been<br />

done only.<br />

9.3 Results <strong>and</strong> discussion<br />

XRD analysis <strong>of</strong> diop <strong>and</strong> diphos modified Pt <strong>and</strong> Pd nanoclusters clearly demonstrate<br />

Pt <strong>and</strong> Pd crystalline nature <strong>of</strong> the samples. In fact, all spectra have a strong peak at 39°<br />

corresponding (111) crystal face (Fig. 9-2). The small peak at 46° (diop on Pt, Fig. 9-2,<br />

curve 1) corresponds to (200) crystal face [221], where as in case <strong>of</strong> diphos on Pt<br />

sample (same Fig. curve 4 this peak is not seen. The asymmetrical peak at ~35° (diphos<br />

on Pd, curve 3) is difficult to assign unambiguously it probably might be assigned to<br />

(311) Pd crystal face base on work <strong>of</strong> Hull [222].<br />

Based on Scherrer equation [177] the average size <strong>of</strong> metal crystals was estimated<br />

considering width on half height <strong>of</strong> the strongest peak at 39°. The corresponding sizes<br />

<strong>of</strong> Pt <strong>and</strong> Pd nanoclusters are summarized in the Table 9-1 together with data obtained<br />

from the TEM image (Fig. 9-3 ).<br />

91


Table 9-1. Average crystal size <strong>of</strong> diphos <strong>and</strong> diop modified Pt <strong>and</strong> Pd nanoclusters<br />

estimated according to the Scherrer model.<br />

Sample<br />

, nm (Scherrer , nm (TEM)<br />

model)<br />

Diop on Pt 1.7 2<br />

Diop on Pd 1.4 -<br />

Diphos on Pt 1.2 -<br />

Diphos on Pd 1.8 -<br />

Estimated values <strong>of</strong> the size <strong>of</strong> crystals lie within 1-2 nm range. We will use these<br />

values for estimation <strong>of</strong> initial rate in further. Metal content in all samples was found to<br />

be 43±5 % from TGA.<br />

1<br />

2<br />

3<br />

4<br />

30 40 50 2Θ 60 70<br />

Fig. 9-2. XRD spectra <strong>of</strong> diop modified Pd (1), Pt (2), diphos modified Pd (3) <strong>and</strong> Pt (4)<br />

nanoclusters.<br />

92


Fig. 9-3. TEM image <strong>of</strong> diop modified Pt nanoclusters.<br />

Conformation analysis<br />

As was found from molecular modelling the diphos molecule can exist in several<br />

geometrical conformations (Fig. 9-4) at room temperature. The presence <strong>of</strong> different<br />

conformations is conditioned by free rotation around Ph-P <strong>and</strong> P-C bonds.<br />

93


Fig. 9-4. Conformation <strong>of</strong> diphos molecules.<br />

At least three conformations <strong>of</strong> diphos were found with the energy gaps ΔE 2,1 =0.014<br />

eV, ΔE 2,3 =0.003 eV that are less that energy corresponding to the heat motion <strong>of</strong><br />

molecules (K b T=0.023 eV, T = 300 °K), thus these three (<strong>and</strong> probably more)<br />

conformation coexist at room temperature.<br />

The very similar situation was found with diop molecule, in fact, the energy gap<br />

ΔE 2,1 =7.5 meV results in coexisting <strong>of</strong> both conformations (Fig. 9-5) at room<br />

temperature.<br />

Fig. 9-5. Conformation <strong>of</strong> diop molecule. For classification <strong>of</strong> dipole moment<br />

orientation please the text below.<br />

DRIFTS investigation<br />

94


Since the orientation <strong>of</strong> phenyl rings <strong>of</strong> diphos <strong>and</strong> diop molecules is not rigidly fixed<br />

due to the free rotation around mentioned bonds, we have to do not take into account<br />

the orientations <strong>of</strong> dipole moment induced by any vibration <strong>of</strong> any phenyl ring <strong>of</strong><br />

diphos <strong>and</strong> diop in determination <strong>of</strong> adsorption mode from analysis <strong>of</strong> DRIFT spectra<br />

<strong>of</strong> corresponding samples. In fact, for example, in the case <strong>of</strong> conformation 1 <strong>of</strong> diop<br />

the out <strong>of</strong> plane vibration <strong>of</strong> right upper phenyl ring (Fig. 9-5 left) induces dipole<br />

moment orientated almost along the axis connecting two phosphorous atoms in the<br />

molecule, whereas in the case <strong>of</strong> conformation 2 (Fig. 9-5 right) the dipole moment<br />

induced by the same phenyl ring is orientated in almost perpendicular direction (out <strong>of</strong><br />

paper). However the vibrations produced by phenyl ring(s) are very strong <strong>and</strong> can be<br />

use for comparison between DRIFT spectra <strong>of</strong> diop <strong>and</strong> diphos based samples.<br />

Since diop <strong>and</strong> diphos have similar structure, their DRIFT spectra have very similar<br />

groups <strong>of</strong> peaks (Fig. 9-6). In fact, for example, two strong peaks at 1479 cm -1 , 1432<br />

cm -1 can be found in spectra <strong>of</strong> diop <strong>and</strong> diphos. It is important to note, that each <strong>of</strong><br />

them is a observed superposition <strong>of</strong> low resolved (due to the very close positions) group<br />

<strong>of</strong> single peaks (red sticks) corresponding to in plane vibration <strong>of</strong> every four phenyl<br />

ring, as is demonstrated from theoretical IR modelling <strong>of</strong> diphos molecule (Fig. 9-7).<br />

Taking this fact into account, we would like to say, that it is not possible to answer on<br />

the following question exactly. Which one <strong>of</strong> single peaks (red sticks, Fig. 9-7) does<br />

correspond to the observed peak (1484 cm -1 ) in the spectra <strong>of</strong> adsorbed (Fig. 9-8,<br />

curves 2 <strong>and</strong> 3) molecule diphos (as well as diop)?<br />

2<br />

1400 1200 1000 800<br />

Wavenumber, cm -1 1<br />

Fig. 9-6. DRIFT spectra <strong>of</strong> free diphos (1) <strong>and</strong> diop (2).<br />

It is also possible that the observed peak (1484 cm -1 ) in the spectrum <strong>of</strong> adsorbed<br />

molecule consists <strong>of</strong> several (or all) singe peaks (some <strong>of</strong> them present as shoulders)<br />

<strong>and</strong> it is not clear which one <strong>of</strong> them dominates. At the same time the geometrical<br />

orientations <strong>of</strong> phenyl rings are not fixed in the space due to the conversion between<br />

conformations. The same objections are valid for other in plane vibrations (in the range<br />

1160-1000 cm -1 ) <strong>and</strong> out <strong>of</strong> plane vibrations (in the range 846-675 cm -1 ) <strong>of</strong> phenyl<br />

95


ings. Due to these reasons we can not successfully apply the metal adsorption selection<br />

rule for determining the geometry <strong>of</strong> adsorption. This is why only some vibrations <strong>of</strong><br />

molecule which do not affect vibrations <strong>of</strong> phenyl rings were taken into the account.<br />

Among these vibrations (at 1276 cm -1 , 1099 cm -1 <strong>and</strong> 907 cm -1 ) the vibration at 1099<br />

cm -1 inducing dipole moment orientated (for all three conformations) along the line<br />

connecting two phosphor atoms (Fig. 9-4 or 9-9) is the most strongest one. Since this<br />

vibration presences in the spectrum <strong>of</strong> diphos adsorbed on Pt <strong>and</strong> Pd nanoclusters as a<br />

strong peak, (both at 1105 cm -1 ) (Fig. 9-8) we propose that the axis connecting two<br />

phosphor atoms is not parallel to the metal surface, i.e. it either parallel to the surface<br />

normal vector, either has a significant angle <strong>of</strong> tilt, as schematically shown in Fig. 9-9.<br />

We think that the adsorption on corners <strong>of</strong> Pt <strong>and</strong> Pd nanocluster can not be excluded,<br />

however, in order to facilitate analysis <strong>and</strong> larger number <strong>of</strong> atoms (making adsorptions<br />

sites) are on the plane <strong>of</strong> crystals we consider adsorption occurring on a flat surface<br />

only.<br />

Since no principal differences were found between spectra <strong>of</strong> adsorbed diphos on Pt<br />

<strong>and</strong> Pd nanoclusters except that some shoulders in the spectrum <strong>of</strong> diphos on Pt are<br />

represented as peaks in the spectrum <strong>of</strong> diphos on Pd <strong>and</strong> vice versa we think that there<br />

are no principal differences in adsorption geometry. However some degree <strong>of</strong> tilt <strong>of</strong> the<br />

adsorbed diphos is possible that might be a reason for slight difference in wavenumbers<br />

<strong>of</strong> the corresponding peaks e.g. (out <strong>of</strong> plane vibration <strong>of</strong> phenyl ring at 754 cm -1 in<br />

case <strong>of</strong> Pt <strong>and</strong> 747 cm -1 in case <strong>of</strong> Pd nanoparticles), that can also be explained as a<br />

different adsorption strength.<br />

Fig. 9-7. Fragment <strong>of</strong> calculated (low frequency axis) IR spectrum <strong>of</strong> diphos.<br />

Corresponding frequencies <strong>of</strong> experimental observed peaks are written above, see also<br />

text for details.<br />

96


1106<br />

1106<br />

3<br />

2<br />

1099<br />

1<br />

1400 1200 1000 800<br />

Wavenumber, cm -1<br />

Fig. 9-8. Drift spectra <strong>of</strong> free (1) <strong>and</strong> adsorbed on Pt (2) <strong>and</strong> Pd (3) nanoparticles<br />

diphos. Intensity <strong>of</strong> spectrum <strong>of</strong> free diphos is decreased for better performance.<br />

We think that the lone pair <strong>of</strong> phosphor atom does not bind to the metal surface because<br />

<strong>of</strong> steric hinder induced by both neighbour phenyl rings. The possible distance between<br />

metal surface <strong>and</strong> phosphor atom can be roughly (without considering changes in the<br />

geometry <strong>of</strong> adsorbed molecule) estimated as a half size <strong>of</strong> phenyl ring (~1.4 Å,<br />

hydrogens were ignored).<br />

The same logic was applied in the analysis <strong>of</strong> DRIFT spectra <strong>of</strong> adsorbed diop on Pt<br />

<strong>and</strong> Pd nanoclusters (Fig. 9-10), thus only those vibrations which do not affect<br />

vibrations <strong>of</strong> phenyl rings were considered. Among them the strongest vibrations<br />

(marked in the Fig. 9-10) were chosen <strong>and</strong> analyzed as shown in the Table 9-2. As can<br />

be seen from orientation <strong>of</strong> corresponding dipole moment the main axis (line<br />

connecting both P-atoms) <strong>of</strong> the diop molecule can not have only exactly perpendicular<br />

or parallel to the normal <strong>of</strong> surface orientations, otherwise all these vibrations<br />

(absorption peaks) will not be observable simultaneously, in one spectrum. This means,<br />

that the molecule has more then one adsorption mode, or what we think is the most<br />

probable, the axis <strong>of</strong> the molecule has a significant tilt with respect to the all three<br />

coordinate axis’s (one normal <strong>and</strong> two tangent vectors <strong>of</strong> the plane), schematically<br />

adsorption mode is shown in the Fig. 9-11. Here again we would like to note that the<br />

adsorption on the edges <strong>and</strong> corners <strong>of</strong> nanoparticle is also not excluded, however due<br />

to the mentioned above reasons we consider adsorption on a plane.<br />

97


Fig. 9-9. Proposed orientation <strong>of</strong> adsorbed diphos on Pt <strong>and</strong> Pd nanoclusters.<br />

1087<br />

3<br />

1385<br />

1371<br />

1251<br />

1216<br />

891<br />

2<br />

1400 1200 1000<br />

Wavenumber, cm -1 800<br />

1<br />

Fig. 9-10. DRIFT spectra <strong>of</strong> free (1) <strong>and</strong> adsorbed diop on Pt (2) <strong>and</strong> Pd (3)<br />

nanoclusters.<br />

Table 9-2. Assignment <strong>of</strong> some frequencies (cm -1 ) in DRIFT spectra <strong>of</strong> free, adsorbed<br />

on Pt <strong>and</strong> Pd nanoclusters diop.<br />

Free diop Diop on Pt Diop on Pd Diop theoretical A<br />

1385 (st) 1376 B (st) 1381 (st)<br />

C<br />

1461 ↕ 1,5<br />

1371 (st) 1376 B (st) 1372 (sh) 1447 ● 1,2<br />

1251 (st) C 1236 (st) 1240 (st)<br />

B<br />

1282 ↕ 1,6<br />

B<br />

1275 ↕ 1,5<br />

98


1216 (st) 1217 (sh) -<br />

B<br />

1210 ↔ 1,4<br />

1087 (st) 1101 (st) 1104 (med)<br />

B<br />

1016 ↕ 1,5<br />

891 (st) 885 (st) 889 (st)<br />

B<br />

874 ↔ 1,4<br />

Comments: A – Type <strong>of</strong> the vibration summarized in this table <strong>and</strong> corresponding<br />

orientation <strong>of</strong> the dipole moment are the same for all conformations <strong>of</strong> diop, B -<br />

unambiguous assignment is difficult, C – orientation <strong>of</strong> the dipole moment was<br />

characterized as “↔” - horizontal (in plane <strong>of</strong> the figure),“↕” – vertical <strong>and</strong> “●” -<br />

horizontal (perpendicular to the figure’s plane) with numerical subscript according to<br />

scheme on Fig. 9-5.<br />

No principal differences between DRIFT spectra <strong>of</strong> diop adsorbed on Pt <strong>and</strong> Pd<br />

nanoparticles were found. However the relative intensity <strong>of</strong> some peaks is slightly<br />

different, some peaks have their analogs as shoulders, small shifts in position <strong>of</strong> some<br />

peaks probably indicate on some space deviation <strong>of</strong> main axis <strong>of</strong> diop being adsorbed<br />

on Pd with respect to Pt nanoclusters or/<strong>and</strong> different adsorption strength. The peaks at<br />

721 cm -1 (out <strong>of</strong> plane vibration <strong>of</strong> phenyl ring) in the spectrum <strong>of</strong> diop on Pd has no<br />

corresponding peak in case <strong>of</strong> Pt, probably due to the mentioned above reasons: some<br />

degree <strong>of</strong> tilt <strong>of</strong> molecular axis <strong>and</strong> different bonding strength.<br />

Here we again propose that the lone pair <strong>of</strong> closest to the metal surface phosphorous<br />

atom does not take part in binding to metal (at least in case <strong>of</strong> adsorption on flat<br />

surface), because <strong>of</strong> “big” distance (~1.4 Å, estimated in the same way as have been<br />

done for diphos) to surface that is caused by space shielding <strong>of</strong> phenyl rings.<br />

Fig. 9-11. Proposed adsorption model <strong>of</strong> diop on Pt <strong>and</strong> Pd nanoclusters, hydrogens are<br />

not shown.<br />

99


The model <strong>of</strong> adsorption <strong>of</strong> diop on Pt <strong>and</strong> Pd nanoclusters is very similar to model <strong>of</strong><br />

diphos adsorption. This is not surprising because both molecules have the same (Ph) 2 P-<br />

X-P(Ph) 2 structure.<br />

Results <strong>of</strong> ethyl pyruvate hydrogenation with diop <strong>and</strong> diphos modified Pt <strong>and</strong> Pd<br />

catalysts are summarized in the Table 9-3.<br />

Table 9-3. Results <strong>of</strong> ethyl pyruvate hydrogenation with diop <strong>and</strong> diphos modifiers <strong>of</strong><br />

Pt nanoclusters.<br />

Sample<br />

name<br />

Catalyst<br />

loading,<br />

mg.<br />

Reaction rate,<br />

mmol·(min·g) -1<br />

Conversion B ,<br />

Initial<br />

(min) -1 min<br />

rate A % after time,<br />

Ee, %<br />

(dominated<br />

enantiomer)<br />

Diphos on 20 2 1.2 11 C (90) 0<br />

Pt<br />

Diphos on 20 ≈0 ≈0 0 D (75) -<br />

Pd<br />

(S,S)Diop 20 13 10 80 C (300) 4.4 (S)<br />

on Pt<br />

(R,R)Diop 20 10 7 50 C (90) 4.4 (R)<br />

on Pt<br />

Diop on Pd 20 ≈0 ≈0 0 D (90) -<br />

E<br />

Pt/Al 2 O 3 20 ≈0 ≈0 0 D (240) -<br />

F<br />

Pt/Al 2 O 3 20 0.6 16 2.3 C (90) 0<br />

Comments: A – Values were estimated based on average crystal size (XRD) <strong>and</strong> metal<br />

content (TGA) <strong>of</strong> the samples, B - initial amount <strong>of</strong> racemic ethyl lactate (as impurity<br />

in commercial ethyl pyruvate) was taken into account, C – according to the curve<br />

pr<strong>of</strong>ile (conversion vs. time) reaction can go further, D - according to the curve pr<strong>of</strong>ile<br />

(conversion vs. time) reaction can not go further E- reaction was performed with<br />

presence <strong>of</strong> 34 mg (S,S)diop (1 mM), F- racemic hydrogenation without a modifier.<br />

As can be seen from the Table 9-3, surprisingly, diop modified Pt nanoclusters were<br />

found to be able to induce low (4.4 %) but reproducible enantioselectivity in ethyl<br />

pyruvate hydrogenation reaction. Moreover this type <strong>of</strong> catalyst demonstrates<br />

enhancement <strong>of</strong> specific reaction rate <strong>and</strong> thus enhancement <strong>of</strong> the initial rate values in<br />

factor <strong>of</strong> 7-8 with respect to Pt nanoclusters with similar average size modified by non<br />

chiral diphos.<br />

No conversion was found in ethyl pyruvate hydrogenation with activated (400°C,<br />

H 2 /N 2 , 2 hours) conventional Pt/Al 2 O 3 with presence <strong>of</strong> free diop (1 mM), probably due<br />

to the differences in surface chemistry (surface contaminations <strong>and</strong> external impurities)<br />

between diop modified nanoclusters <strong>and</strong> conventional Pt/Al 2 O 3 .<br />

Interestingly, that (S,S)diop leads to excess <strong>of</strong> S lactate <strong>and</strong> (R,R)diop to R lactate <strong>and</strong><br />

corresponding catalysts demonstrate comparable reaction rate. These facts clearly<br />

indicate the importance <strong>of</strong> chiral nature <strong>of</strong> modifier on yielded direction <strong>of</strong><br />

enantioselectivity.<br />

Taking into the account the model <strong>of</strong> enantioselectivity developed for cinchonidine on<br />

Pt system we would like to propose the following model explaining the sense <strong>of</strong><br />

enantioselectivity. Author does not pretend to absolute rightfulness <strong>of</strong> the model,<br />

however finds it to be sufficient for qualitative explanation <strong>of</strong> enantioselectivity, its<br />

inverse with using opposite enantiomers <strong>of</strong> diop <strong>and</strong> enhancement <strong>of</strong> the reaction rate.<br />

100


According to the determined adsorption geometry <strong>of</strong> diop, the lone pair <strong>of</strong> phosphorous<br />

is not involved into binding to the metal surface (at least for adsorption on a flat<br />

surface), thus it is possible that the phosphorous atom (blue marked) (Fig. 9-10 right)<br />

can “catch” adsorbed on Pt ethyl pyruvate molecule via hydrogen bond (rate<br />

enhancement), like unproronated cinchonidine does [50] in non-ionic solvents (toluene,<br />

THF, etc.). In both cases α-carbonyl group <strong>of</strong> ethyl pyruvate molecule forms halfhydrogenated<br />

state with modifier, is in asymmetric environment produced by red<br />

marked OH <strong>and</strong> H groups <strong>of</strong> a modifier. It is known for cinchona alkaloids on Pt that<br />

reverse <strong>of</strong> positions <strong>of</strong> OH <strong>and</strong> H groups leads to inverse <strong>of</strong> enantioselectivity<br />

(cinchonidine “became” cinchonine) [87]. The same (or very similar) fact was found<br />

for diop obtaining S lactate by using (S,S)diop <strong>and</strong> R lactate with (R,R) diop.<br />

O<br />

H<br />

N<br />

OH<br />

H<br />

H<br />

P<br />

H<br />

O<br />

O<br />

O<br />

O<br />

*<br />

= X<br />

C<br />

X<br />

C<br />

X<br />

P<br />

O<br />

X<br />

N<br />

HHO<br />

O H<br />

N<br />

O<br />

P<br />

H<br />

H<br />

O<br />

X<br />

Pt surface<br />

Fig. 9-12. Models <strong>of</strong> half hydrogenated state in case <strong>of</strong> literature known [84-86] [ethyl<br />

pyruvate-cinchonidine] ad complex (left bottom) on Pt <strong>and</strong> proposed [ethyl pyruvatediop]<br />

ad complex (right bottom) on Pt. Schematic positions <strong>of</strong> OH <strong>and</strong> H groups <strong>of</strong><br />

modifiers with respect to α-carbonyl C=O bond <strong>of</strong> ethyl pyruvate for cinchonidine (left<br />

up) <strong>and</strong> diop (right up) modifier.<br />

It is interesting to note, that due to the geometry <strong>of</strong> modifiers OH <strong>and</strong> H groups are<br />

located significantly closer to “caught” C=O group <strong>of</strong> ethyl pyruvate in case <strong>of</strong><br />

cinchonidine than for diop. Probably because <strong>of</strong> remoteness <strong>of</strong> OH <strong>and</strong> H groups their<br />

asymmetrical influence on C=O bond is not as strong as it is in the case <strong>of</strong> cinchona<br />

system, that results in low ee. It is also possible that the different nature <strong>of</strong> H-N <strong>and</strong> H-<br />

P hydrogen bonds also influences on ee. The electric (electrostatic) influence <strong>of</strong> closed<br />

to the metal surface phenyl rings, in principal, can also cause asymmetrical influence <strong>of</strong><br />

C=O bond, since one <strong>of</strong> them in tilted. However we do not think their influence <strong>and</strong><br />

resulted asymmetric environment are significant, because they have very similar nature<br />

(with respect to OH, H pair), in fact, by keeping (Ph) 2 P group in the same state <strong>and</strong><br />

reversing positions <strong>of</strong> OH <strong>and</strong> H groups (using another enantiomer <strong>of</strong> diop) only, we<br />

change direction <strong>of</strong> enantioselectivity <strong>of</strong> the system.<br />

101


The calculation <strong>of</strong> energy <strong>of</strong> forming pro-R <strong>and</strong> pro-S half-hydrogenated diop-ethyl<br />

pyruvate complexes can confirm or disprove this model; however we leave this out <strong>of</strong><br />

the subject <strong>of</strong> the current thesis.<br />

9.4 Summary<br />

The models <strong>of</strong> adsorption geometry <strong>of</strong> diop <strong>and</strong> diphos molecules on Pt <strong>and</strong> Pd<br />

nanoclusters have been proposed. According to these models both molecules are<br />

adsorbed via their phenyl rings pointing molecular axis (a line connecting two<br />

phosphorous atoms) out from the surface, some degree <strong>of</strong> tilt <strong>of</strong> the axis is also<br />

possible. In this model lone pair <strong>of</strong> phosphorous does not take part in adsorption,<br />

probably due to this fact <strong>and</strong> asymmetrical (with respect to C=O bond) location <strong>of</strong> OH<br />

<strong>and</strong> H groups <strong>of</strong> diop result in observed enantioselectivity in hydrogenation <strong>of</strong> ethyl<br />

pyruvate. Diop modified Pt nanoclusters demonstrate enhancement <strong>of</strong> the reaction rate<br />

with respect to Pt nanoclusters modified with non chiral lig<strong>and</strong> in a factor <strong>of</strong> 7-8. The<br />

absence enhanced reaction rate <strong>and</strong> enantioselectivity in ethyl pyruvate hydrogenation<br />

with conventional Pt/Al 2 O 3 with presence <strong>of</strong> diop shows on differences between<br />

catalysts, probably in surface modification.<br />

102


Chapter 10<br />

Binap <strong>and</strong> synphos modified Pt<br />

colloidal nanoparticles<br />

10.1 Introduction<br />

In the current chapter we use chiral binap <strong>and</strong> synphos molecules (Fig. 10-1) as<br />

modifiers for Pt (synphos, binap) <strong>and</strong> Pd (binap) nanoclusters.<br />

O<br />

PPh 2 O PPh 2<br />

PPh 2 O PPh 2<br />

Binap<br />

Synphos<br />

Fig. 10-1. Binap <strong>and</strong> synphos molecules.<br />

As it was already mentioned binap attracts attention due to the numerous successful<br />

applications <strong>of</strong> it in enantioselective homogeneous catalysis [223]. The Synphos<br />

molecule is similar to binap structure <strong>and</strong> synphos based organometalic complex as a<br />

homogeneous catalyst has been found to be even more efficient in some <strong>of</strong> reactions<br />

[33]. Immobilization <strong>of</strong> this type <strong>of</strong> chiral lig<strong>and</strong> on a surface <strong>of</strong> metal nanoparticles is<br />

one strategy for heterogenation <strong>of</strong> enantioselective homogeneous catalysts. Recently,<br />

binap was immobilized on Au <strong>and</strong> Pd nanoclusters with size 1.7 <strong>and</strong> 2 nm,<br />

correspondingly, interestingly, that the latter system was found to be able to induce 95<br />

% enantiomeric excess in hydrosilylation <strong>of</strong> styrene [224]. Another example is the<br />

immobilization <strong>of</strong> Ru complex with phosphonic acid substituted binap on Fe 3 O 4<br />

nanoparticles with 10 nm size [225]. The obtained quasi-homogeneous catalyst<br />

demonstrated even higher ee’s in hydrogenation <strong>of</strong> aromatic ketones. It is very<br />

important to mention that magnetic properties <strong>of</strong> the nanoparticles allow their<br />

separation from reaction products by permanent magnet.<br />

In the current chapter we test binap <strong>and</strong> synphos modified Pt nanocluster in<br />

hydrogenation <strong>of</strong> ethyl pyruvate only, whereas the further test <strong>of</strong> these (<strong>and</strong> all other)<br />

samples is supposed to be conducted by our collaborators from Paris <strong>and</strong> Bucharest,<br />

according to the mentioned in the chapter 3 EU-COST project.<br />

10.2 Experimental<br />

Materials<br />

Synphos lig<strong>and</strong> was provided by Virginie Ratovelomanana-Vidal from the Laboratoire<br />

de Synthese Selective Organiquie et Produits Naturels, UMR (Ecole Nationale<br />

Superieure de Chemie de Paris), binap was purchased from Aldrich.<br />

Samples preparation<br />

O<br />

103


For the preparation <strong>of</strong> synphos modified Pt nanoclusters 273 mg Pt 2 (DBA) 3 (0.51 mmol<br />

<strong>of</strong> Pt) was dissolved in nitrogen flushed THF (40 ml), 300 mg (0.47 mmol) synphos<br />

dissolved in 20 ml nitrogen flushed THF. Both mixtures were transferred into the metal<br />

reactor, purged with argon for 10 min <strong>and</strong> kept under 3 bar <strong>of</strong> hydrogen pressure. After<br />

five hours the pressure was released <strong>and</strong> the mixture was transferred into 300 ml<br />

cyclohexene, then additional 250 ml cyclohexene was added. The brown-back<br />

precipitate was collected by G4 filter. The precipitate in the filter was washed three<br />

times with 100 ml THF/cyclohexene (2/3 by volume) mixture, then with 300 ml<br />

cyclohexene. The brown-black powder was collected mechanically from the filter <strong>and</strong><br />

finally fried on air at 35°C for 24 hours. The sample was found to be redispersible in<br />

e.g. chlor<strong>of</strong>orm or THF.<br />

Due to the limited amount <strong>of</strong> synphos, immobilization <strong>of</strong> it on Pd nanoclusters has not<br />

been done.<br />

In order to prepare binap modified Pt nanoclusters 560 mg Pt 2 (DBA) 3 (0.51 mmol Pt)<br />

in nitrogen purged for 10 min 40 ml THF was added to nitrogen purged 30 ml THF<br />

with 635 mg (1.02 mmol) binap. The mixture was transferred into the metal reactor <strong>and</strong><br />

after purging with Ar for 10 min hydrogen pressure <strong>of</strong> 4 bar has been applied. After 5<br />

hours the mixture was transferred into 500 ml cyclohexene. After 24 hours black<br />

precipitate was collected <strong>and</strong> washed with 400 ml THF: cyclohexene mixture (1:4) <strong>and</strong><br />

with 200 ml cyclohexene. Finally sample was dried at 40 °C under vacuum.<br />

Characterization<br />

The free (not adsorbed) lig<strong>and</strong> <strong>and</strong> quasi-homogeneous synphos modified Pt colloidal<br />

sample (3 mg) were dissolved (separately) in 1.3 ml <strong>of</strong> CDCl 3 <strong>and</strong> 31 P-NMR<br />

measurements were performed on JEOL ECX 400 MHz instrument making 20 <strong>and</strong><br />

20000 scans, respectively.<br />

Molecular modelling, XRD spectra measuring, DRIFTS investigations <strong>and</strong><br />

hydrogenation <strong>of</strong> ethyl pyruvate have been performed as it is described e.g. in the<br />

chapter 9.<br />

Due to the absence <strong>of</strong> TEM instrument in the IUB <strong>and</strong> limited access to the instrument<br />

outside the university, TEM investigation <strong>of</strong> prepared samples is missing.<br />

10.3 Results <strong>and</strong> discussion<br />

31 P-NMR investigation<br />

NMR investigation <strong>of</strong> free lig<strong>and</strong> <strong>and</strong> lig<strong>and</strong> modified colloidal Pt nanoclusters is<br />

illustrated on example <strong>of</strong> synphos.<br />

Synphos molecule has two phosphorous atoms; however 31 P-NMR spectrum (Fig. 10-2<br />

up) has only one peak at -14.8 ppm due to the specific structure <strong>of</strong> the synphos<br />

molecule (identical surrounding <strong>of</strong> P atoms). The absolute absence <strong>of</strong> this peak in the<br />

spectrum <strong>of</strong> colloidal sample unambiguously says that the obtained after written above<br />

purification procedure sample contains no trace <strong>of</strong> free (not adsorbed) synphos.<br />

Whereas, in case <strong>of</strong> adsorbed synphos, significantly shifted peak to 8.2 ppm says about<br />

changes in chemical environment <strong>of</strong> phosphorous atoms.<br />

The obtain spectrum is very similar to the spectra obtained by Hyeon <strong>and</strong> co workers<br />

[226] for binap (<strong>and</strong> other phosphor organic compounds) adsorbed on colloidal Pd<br />

nanoclusters.<br />

104


-14.8<br />

Free synphos<br />

8.2<br />

Synphos on Pt<br />

40 30 20 10 0 -10 -20 -30 -40<br />

PPM<br />

Fig. 10-2. 31 P-NMR spectra <strong>of</strong> free synphos (up) <strong>and</strong> synphos modified colloidal Pt<br />

nanoclusters (bottom). Both samples were dissolved in CDCl 3 .<br />

It is important to note, that due to the small size <strong>of</strong> colloidal Pt nanoclusters estimated<br />

as 1 nm based on XRD (see below) <strong>and</strong> 2 nm based on dynamic light scattering (data<br />

are not included) it was estimated that r<strong>and</strong>om rotation <strong>of</strong> single nanocluster (in<br />

solution) is fast enough to do not consider anisotropic NMR effects <strong>and</strong> the system can<br />

be considered within liquid phase NMR theory, i.e. free r<strong>and</strong>om rotation <strong>of</strong> nanocluster<br />

corresponds to free r<strong>and</strong>om rotation <strong>of</strong> a molecule. In fact, rotation correlation time <strong>of</strong><br />

an object with radius R (2 nm) in a solvent with viscosity η (0.58 cp or 5.8·10 -4<br />

kg/(m·sec)) at room temperature T (300 K) can be determined from the following<br />

equation, based on Debye-Stokes-Einstein formula.<br />

4 3 η<br />

−<br />

τ = ≈ 5 ⋅10<br />

9<br />

c<br />

πR<br />

sec<br />

3 KT<br />

From another h<strong>and</strong> the chemical shift anisotropy (CSA) effect can be roughly estimated<br />

as a 10-15 ppm, that is ≈6 kHz in terms <strong>of</strong> frequencies or τ SCA = 0.2·10 -3 sec. Thus, the<br />

rotation correlation time is much less than typical CSA time that verifies our mentioned<br />

above assumption.<br />

The absent <strong>of</strong> the peak corresponding to the free state <strong>of</strong> binap in the spectrum <strong>of</strong> the<br />

binap (as well as e.g. diop) modified Pt colloidal sample again clearly demonstrate that<br />

used preparation <strong>and</strong> purification procedures allow having lig<strong>and</strong> in adsorbed state only<br />

<strong>and</strong> no trace <strong>of</strong> free lig<strong>and</strong>. However due to the technical <strong>and</strong> organisation problems<br />

with NMR instrument it was not possible to perform required continuous measurement<br />

(≈ 20000 scans, ~3 days) for obtaining sufficient signal/noise ratio for resolving peak<br />

corresponding to the adsorbed lig<strong>and</strong> for all samples having P atoms.<br />

XRD spectra <strong>of</strong> binap <strong>and</strong> synphos modified Pt nanoclusters demonstrate crystalline<br />

nature <strong>of</strong> the samples.<br />

105


1<br />

2<br />

20 40 60 80<br />

2Θ<br />

Fig. 10-3. XRD spectra <strong>of</strong> binap (1) <strong>and</strong> synphos (2) modified Pt nanoclusters. Intensity<br />

<strong>of</strong> (1) was decreased for better performance.<br />

In fact, all spectra have a strong peak at 39° corresponding (111) crystal face (Fig. 10-<br />

3). In case <strong>of</strong> binap on Pt sample peaks at 36°, 46° <strong>and</strong> 67° correspond to (222), (200)<br />

<strong>and</strong> (220) faces, according to work <strong>of</strong> Swanson et al. [221]. The last tree peaks are not<br />

resolved in the spectrum <strong>of</strong> synphos on Pt sample, probably, due to the low sample<br />

loading. Based on the width <strong>of</strong> the peak at 39° the average crystal size was estimated<br />

with help <strong>of</strong> Scherrer model, data are summarized in the Table 10-1.<br />

Table 10-1. Average crystal size <strong>of</strong> binap <strong>and</strong> synphos modified Pt estimated according<br />

to the Scherrer model.<br />

Sample<br />

, nm<br />

Binap on Pt 1.6<br />

Synphos on Pt 1.0<br />

Estimated values <strong>of</strong> the size <strong>of</strong> crystals lie within 1-2 nm range. We will use these<br />

values for estimation <strong>of</strong> initial rate in further. Metal content in all samples was found to<br />

be 43±5 % from TGA.<br />

Conformation analysis<br />

It was found from molecular modelling that at least two conformations <strong>of</strong> synphos<br />

molecule are possible due two rotation around P-Ph <strong>and</strong> C-P(Ph) 2 bonds, like it was<br />

found for diphos <strong>and</strong> diop molecules.<br />

106


Fig. 10-4. Conformations <strong>of</strong> synphos.<br />

The energy difference between conformations was found to be 0.027 eV, that is almost<br />

the same value as energy <strong>of</strong> thermal motion (KT=0.026 eV) is, thus both conformations<br />

can coexist, especially in a solution at room temperature. Because <strong>of</strong> this fact, we do<br />

not take into the consideration molecular vibrations induced by phenyl rings.<br />

Regardless <strong>of</strong> the fact that such type <strong>of</strong> conformations were not found for binap<br />

molecule, we ignore any vibration <strong>of</strong> a phenyl ring in interpretation <strong>of</strong> DRIFT spectrum<br />

<strong>of</strong> adsorbed binap as well.<br />

DRIFTS investigation<br />

From analysis <strong>and</strong> comparison <strong>of</strong> DRIFT spectra <strong>of</strong> free synphos, binap <strong>and</strong> diphos, all<br />

(in <strong>and</strong> out <strong>of</strong> plane) vibrations <strong>of</strong> phenyl rings were excluded from analysis due to the<br />

mentioned above reasons.<br />

1<br />

2<br />

1600 1400 1200 1000 800<br />

Wavenumber, cm -1<br />

Fig. 10-5. DRIFT spectra <strong>of</strong> free binap (1) <strong>and</strong> synphos (2).<br />

107


The DRIFT spectra <strong>of</strong> synphos (Fig. 10-6) <strong>and</strong> modified Pt nanoclusters clearly show<br />

shifts in wavenumbers <strong>of</strong> 2-10 cm -1 that indicates to slight changes in skeleton<br />

geometry or/<strong>and</strong> changes <strong>of</strong> force constants <strong>of</strong> bonds. The relative intensity <strong>of</strong> some<br />

corresponding peaks is also changed. In order to facilitate spectra analysis, frequencies<br />

<strong>of</strong> some peaks allowing unambiguous assignment in terms <strong>of</strong> orientation <strong>of</strong> induced<br />

dipole moments are summarized in the Table 10-2. Left peaks were not considered due<br />

to the fact that some peaks are superposition <strong>of</strong> two or more vibrations, as was<br />

illustrated in Fig. 9-5.<br />

2<br />

1600 1400 1200 1000 800 600<br />

Wavenumber, cm -1 1<br />

Fig. 10-6. DRIFT spectra <strong>of</strong> free (1) <strong>and</strong> adsorbed on Pt nanoclusters (2) synphos.<br />

Intensity <strong>of</strong> free synphos was decreased for better performance.<br />

Table 10-2. Assignment <strong>of</strong> some frequencies (cm -1 ) <strong>of</strong> DRIFT spectra <strong>of</strong> free <strong>and</strong><br />

adsorbed on Pt nanoclusters synphos.<br />

Free synphos<br />

(experimental)<br />

Synphos on Pt Calculated synphos A Dipole moment<br />

orientation B<br />

1452 (str) 1463 (wk. sh.) 1501 (str) ↔ 1,4<br />

1433 (str) 1441 (str) 1485 (str) ↕ 1,5<br />

1401 (med) 1404 (med) 1449 (wk) ↔ 1,3<br />

1289 (str) 1301 (str) 1316 (str) ↔ 1,4 or ↕ 1,5<br />

1255 (med) 1257 (med) 1281 (med) ↔ 1,4<br />

1150 (str) 1159 (med) 1172 (med) ↔ 2,4<br />

938 (med) 939 (med) 975 (med) ↕ 1,5<br />

873 (med) 877 (med) 906 (med) C<br />

902 (wk) C ↔ 1,4<br />

↕ 1,5<br />

Comments: A - conformation from Fig. 10-4 (left) was used, calculated frequencies for<br />

other conformation were just slightly (± max. 10cm -1 ) different, B - orientation <strong>of</strong> the<br />

dipole moment was the same for both conformations <strong>and</strong> was characterized as “↔” -<br />

horizontal (in plane <strong>of</strong> the figure),“↕” – vertical <strong>and</strong> with numerical subscript according<br />

to scheme in Fig. 10-4, C – unambiguous assignment is difficult.<br />

108


Below in the text we assume that only one adsorption mode <strong>of</strong> synphos exists on Pt<br />

surface <strong>and</strong> adsorption takes place on a flat surface.<br />

Based on the data from the Table 10-2 it follows, that the synphos molecule can not<br />

adsorb (see denied examples N1, 2 <strong>and</strong> 3 in Fig. 10-7) on the flat surface without<br />

significant tilt (with respect to depicted in Fig. 10-4 orientation, assuming that the metal<br />

surface has a normal parallel to ↕ 1,5 direction). At the same time it allows a coexistence<br />

<strong>of</strong> non zero projection to the surface’s normal <strong>of</strong> all orientations <strong>of</strong> dipole moment from<br />

this table. However “denied” orientations become possible if they coexist on the Pt<br />

surface, e.g. 1 coexist with 2 or 3.<br />

Due to the isotropic shape <strong>of</strong> the molecule it is not possible to determine the orientation<br />

<strong>of</strong> the adsorbed molecule unambiguously; in fact, at least two adsorption species (see<br />

allowed examples (4 <strong>and</strong> 5) in Fig. 10-7-bottom) can be proposed, they can also coexist<br />

on the surface together.<br />

Fig. 10-7. Schematic illustration <strong>of</strong> denied mono modes (up) <strong>and</strong> allowed mono <strong>and</strong><br />

multi modes (bottom) <strong>of</strong> adsorption <strong>of</strong> synphos molecule on Pt nanoclusters. Hydrogen<br />

atoms are not shown.<br />

It is interesting to note that proposed adsorption mode N5 (Fig. 10-7) is similar to<br />

adsorption modes found for diphos <strong>and</strong> diop molecules. In fact, all three molecules<br />

have P(Ph) 2 functional group which can be considered as a common anchor for them, in<br />

the similar way as quinoline is an anchor for cinchonidine <strong>and</strong> quiphos molecules.<br />

In spite <strong>of</strong> similarity in the structure <strong>of</strong> synphos <strong>and</strong> binap, IR spectrum <strong>of</strong> the latter is<br />

difficult for full analysis because it has a lot <strong>of</strong> peaks which have no analogs in the IR<br />

spectrum <strong>of</strong> synphos (probably due to the enhanced inharmonic vibrations for binap).<br />

Unfortunately, almost every strong peak in the spectrum <strong>of</strong> binap which can be<br />

unambiguously assigned has to be excluded from consideration due the mentioned<br />

above reasons (different conformations <strong>and</strong>/or contribution from vibration <strong>of</strong> the<br />

similar parts <strong>of</strong> the molecule, e.g. another phenyl rings). From another h<strong>and</strong>, those<br />

109


peaks which can be used for analysis <strong>of</strong> orientation <strong>of</strong> adsorbed binap can not be<br />

assigned unambiguously.<br />

However ambiguity in assignment <strong>and</strong> interpretation <strong>of</strong> spectra very <strong>of</strong>ten results in<br />

difficulties (impossibilities) in discrimination between ↕ 1,5 <strong>and</strong> ↔ 1,4 orientations <strong>of</strong><br />

induced dipole moment. For example the strong peaks at 816 cm -1 in the spectrum <strong>of</strong><br />

binap on Pt (Fig. 10-8) corresponds to the peak at 818 cm -1 in the spectrum <strong>of</strong> free<br />

binap which consists <strong>of</strong> two components (due to the symmetrical <strong>and</strong> asymmetrical out<br />

<strong>of</strong> plane bending <strong>of</strong> hydrogens in naphthalene rings, both are not shown). The<br />

symmetrical vibration has dipole moment along the ↕ 1,5 direction, whereas<br />

asymmetrical one has it along the ↔ 1,4 direction. The similar situation is with all other<br />

strong peaks.<br />

816<br />

2<br />

1600 1400 1200 1000 800<br />

Wavenumber, cm -1 1<br />

Fig. 10-8. DRIFT spectra <strong>of</strong> free (1) <strong>and</strong> adsorbed on Pt nanoclusters (2) binap.<br />

Intensity <strong>of</strong> free binap was decreased for better performance.<br />

Since spectra <strong>of</strong> adsorbed binap on Pt nanoclusters allows interpretation in terms <strong>of</strong><br />

orientation <strong>of</strong> dipole moment in ↕ 1,5 or/<strong>and</strong> ↔ 1,4 directions, we would like to use the<br />

same logic for determination <strong>of</strong> orientations <strong>of</strong> adsorbed binap. This results to the same<br />

adsorption modes like for synphos molecule (Fig. 10-7).<br />

In hydrogenation <strong>of</strong> ethyl pyruvate both catalysts: binap <strong>and</strong> synphos did not show<br />

detectable enantioselectivity. The specific reaction rate <strong>and</strong> initial rate (normalized on<br />

the number <strong>of</strong> metal surface atoms) were similar for both samples. Reaction rate<br />

normalized on number <strong>of</strong> surface Pt atoms was higher for conventional Pt/Al 2 O 3 ,<br />

probably due to the thermal activation <strong>of</strong> the latter under hydrogen. The reason <strong>of</strong><br />

absence <strong>of</strong> enantioselectivity with synphos <strong>and</strong> binap modified Pt nanoclusters<br />

probably lies in principal difference in their structure, i.e. both molecules do not meet<br />

the specific requirements developed for cinchona on Pt system, whereas, we think, diop<br />

molecule (see chapter 9) fits partly, that results in low ee.<br />

Table 10-3. Results <strong>of</strong> ethyl pyruvate hydrogenation with binap <strong>and</strong> synphos modifiers<br />

<strong>of</strong> Pt nanoclusters<br />

110


Initial rate A Conversion B , %<br />

Sample name Catalyst Reaction rate,<br />

loading, mg. mmol·(min·g) -1 (min) -1 after time, min<br />

Binap on Pt 20 5 3.7 30 (120) D<br />

Synphos on Pt 10 4.5 2.8 16 (120) D<br />

C<br />

Pt/Al 2 O 3 20 0.6 16.5 12.5 (345) D<br />

Comments: A – values were estimated based on chemisorption (Pt/Al 2 O 3 ), size <strong>of</strong> Pt<br />

nanoclusters <strong>and</strong> metal content (43%) data, B - initial amount <strong>of</strong> racemic ethyl lactate<br />

(as impurity in commercial ethyl pyruvate) was taken into account. C- hydrogenation<br />

without a modifier, D – according to the curve pr<strong>of</strong>ile (conversion vs. time) reaction<br />

can go further.<br />

10.4 Summary<br />

The models <strong>of</strong> adsorption geometry <strong>of</strong> synphos <strong>and</strong> binap molecules on 1 <strong>and</strong> 1.6 nm Pt<br />

nanoclusters have been proposed. According to these models both molecules are<br />

adsorbed via their phenyl rings, however existence <strong>of</strong> other adsorption modes is also<br />

possible. In any model <strong>of</strong> adsorption on a flat surface, the lone pair <strong>of</strong> phosphorous<br />

does not take part in adsorption, due to steric hinder <strong>of</strong> induced by e.g. phenyl rings. No<br />

enantioselectivity was found with binap <strong>and</strong> synphos modified Pt nanoclusters in the<br />

hydrogenation <strong>of</strong> ethyl pyruvate, probably due the fact that both molecules do not meet<br />

specific requirements which were developed based on cinchona on Pt system.<br />

111


References<br />

[1] G.E. Tranter, J. Chem. Soc., Chem. Commun. (1986) 60-61.<br />

[2] A. Streitwieser, W.D. Schaeffer, J. Am. Chem. Soc. 78 (1956) 5597-5599.<br />

[3] J. March, Advanced Organic Chemistry, John Wiley & Sons, New Your, 1992.<br />

[4] Corriu, Guerin, Moreau, The Chemistry <strong>of</strong> Organic Silicon Compounds, John<br />

Wiley & Sons, New York, 1998.<br />

[5] M. Gielen, Top. Curr. Chem. 104 (1982) 57-105.<br />

[6] Stackhou.J, R.D. Baechler, K. Mislow, Tetrahedron Lett. (1971) 3437.<br />

[7] Stackhou.J, R.D. Baechler, K. Mislow, Tetrahedron Lett. (1971) 3441.<br />

[8] J.D. Andose, J.M. Lehn, K. Mislow, J. Wagner, J. Am. Chem. Soc. 92 (1970)<br />

4050.<br />

[9] G. Delapierre, J.M. Brunel, T. Constantieux, G. Buono, Tetrahedron:<br />

Asymmetry 12 (2001) 1345-1352.<br />

[10] G. Delapierre, T. Constantieux, J.M. Brunel, G.R. Buono, Eur. J. Org. Chem.<br />

(2000) 2507.<br />

[11] J.M. Brunel, T. Constantieux, G. Buono, J. Org. Chem. 64 (1999) 8940-8942.<br />

[12] J.M. Brunel, T. Constantieux, A. Lab<strong>and</strong>e, F. Lubatti, G. Buono, Tetrahedron<br />

Lett. 38 (1997) 5971-5974.<br />

[13] S.J. Brois, J. Am. Chem. Soc. 90 (1968) 506.<br />

[14] S.J. Brois, J. Am. Chem. Soc. 90 (1968) 508.<br />

[15] H. Hamill, M.A. McKervey, J. Chem. Soc., Chem. Commun. (1969) 864.<br />

[16] A. Vargos, Phd Thesis, Zurich, 2004.<br />

[17] V. Prelog, M. Kovacevic, M. Egli, Angew. Chem., Int. Ed. Engl. 28 (1989)<br />

1147-1152.<br />

[18] V.S. Martin, S.S. Woodard, T. Katsuki, Y. Yamada, M. Ikeda, K.B. Sharpless,<br />

J. Am. Chem. Soc. 103 (1981) 6237-6240.<br />

[19] J.D. Morrison, H.S. Mosher, Asymmetric Organic Reactions, Englewood Cliffs,<br />

New Jersey, 1971.<br />

[20] I. Ojima, Catalytic Asymmetric Synthesis, Willey-VCH, 1999.<br />

[21] D.E. De Vos, I.F.J. Vankelecom, P.A. <strong>Jacobs</strong>, Chiral catalyst immobilization<br />

<strong>and</strong> recycling, Weinheim, 2000.<br />

[22] H.-U. Blaser, M. Eissen, P. Fauquex, K. Hungerbühler, E. Schmidt, G.<br />

Sedelmeir, M. Studer, in: H.U. Blaser, E. Schmidt (Eds.), Asymmetric Catalysis<br />

on Industrial Scale: Challenges, Approaches <strong>and</strong> Solutions, WILEY-VCH,<br />

Weinheim, 2004, pp. 91-103.<br />

[23] R. Noyori, M. Koizumi, D. Ishii, T. Ohkuma, Pure Appl. Chem. 73 (2001) 227-<br />

232.<br />

[24] T.P. Dang, H.B. Kagan, J. Chem. Soc., Chem. Commun. (1971) 481-481.<br />

[25] T. Hayashi, M. Kumada, Acc. Chem. Res. 15 (1982) 395-401.<br />

[26] S. Jeulin, N. Champion, P. Dellis, V. Ratovelomanana-Vidal, J.P. Genet,<br />

Synthesis-Stuttgart (2005) 3666-3671.<br />

[27] R. Touati, T. Gmiza, S. Jeulin, C. Deport, V. Ratovelamanana-Vidal, C. Ben<br />

Hassine, J.P. Genet, Synlett (2005) 2478-2482.<br />

[28] R. Le Roux, N. Desroy, P. Phansavath, J.P. Genet, Synlett (2005) 429-432.<br />

[29] C. Mordant, P. Dunkelmann, V. Ratovelomanana-Vidal, J.P. Genet, Eur J Org<br />

Chem (2004) 3017-3026.<br />

[30] C. Mordant, P. Dunkelmann, V. Ratovelomanana-Vidal, J.P. Genet, Chem<br />

Commun (2004) 1296-1297.<br />

112


[31] S. Jeulin, S.D. de Paule, V. Ratovelomanana-Vidal, J.P. Genet, N. Champion, P.<br />

Dellis, P Natl Acad Sci USA 101 (2004) 5799-5804.<br />

[32] J.P. Genet, Accounts Chem Res 36 (2003) 908-918.<br />

[33] S.D. de Paule, S. Jeulin, V. Ratovelomanana-Vidal, J.P. Genet, N. Champion, P.<br />

Dellis, Eur. J. Org. Chem. (2003) 1931-1941.<br />

[34] S.D. de Paule, S. Jeulin, V. Ratovelomanana-Vidal, J.P. Genet, N. Champion,<br />

G. Deschaux, P. Dellis, Org Process Res Dev 7 (2003) 399-406.<br />

[35] S.D. de Paule, S. Jeulin, V. Ratovelomanana-Vidal, J.P. Genet, N. Champion, P.<br />

Dellis, Tetrahedron Lett 44 (2003) 823-826.<br />

[36] K. Inoguchi, K. Achiwa, Synlett (1991) 49-51.<br />

[37] K. Achiwa, J. Am. Chem. Soc. 98 (1976) 8265-8266.<br />

[38] G.M. Schwab, F. Rost, L. Rudolph, Kolloid Z 68 (1934) 157-159.<br />

[39] G.M. Schwab, L. Rudolph, Naturwisserschaft 20 (1932) 363-364.<br />

[40] G.J. Hutchings, Annu. Rev. Mater. Res. 35 (2005) 143-166.<br />

[41] A. Tai, T. Kikukawa, T. Sugimura, Y. Inoue, S. Abe, T. Osawa, T. Harada, Bull.<br />

Chem. Soc. Japan 67 (1994) 2473-2477.<br />

[42] T. Harada, T. Osawa, in: G. Jannes, V. Dubois (Eds.), Chiral Reactions in<br />

Heterogeneous Catalysis, Plenum Press, New York, 1995, p. 83.<br />

[43] T. Osawa, A. Tai, Y. Imachi, S. Takasaki, in: G. Jannes, V. Dubois (Eds.),<br />

Chiral Reactions in Heterogeneous Catalysis, Plenum Press, New York, 1995, p.<br />

75.<br />

[44] M. Studer, H.U. Blaser, C. Exner, Adv. Synth. Catal. 345 (2003) 45-65.<br />

[45] Y. Orito, S. Imai, S. Niwa, J. Chem. Soc. Jpn. (1979) 1118.<br />

[46] Y. Orito, S. Imai, S. Niwa, J. Chem. Soc. Jpn. (1980) 670.<br />

[47] Y. Orito, S. Imai, S. Niwa, Nippon Kagaku Kaishi (1979) 1118-1120.<br />

[48] N. Kunzle, R. Hess, T. Mallat, A. Baiker, J. Catal. 186 (1999) 239-241.<br />

[49] P.B. Wells, A.G. Wilkinson, Top. Catal. 5 (1998) 39-50.<br />

[50] A. Baiker, J. Mol. Catal. A: Chem. 163 (2000) 205-220.<br />

[51] B. Törok, K. Felfoldi, G. Szakonyi, K. Balazsik, M. Bartok, Catal. Lett. 52<br />

(1998) 81-84.<br />

[52] M. Schürch, N. Künzle, T. Mallat, A. Baiker, J. Catal. 176 (1998) 569-571.<br />

[53] O.J. Sonderegger, T. Burgi, A. Baiker, J. Catal. 215 (2003) 116-121.<br />

[54] E. Toukoniitty, P. Maki-Arvela, J. Kuusisto, V. Nieminen, J. Paivarinta, M.<br />

Hotokka, T. Salmi, D.Y. Murzin, J. Mol. Catal. A: Chem. 192 (2003) 135-151.<br />

[55] E. Toukoniitty, P. Maki-Arvela, M. Kuzma, A. Villela, A.K. Neyestanaki, T.<br />

Salmi, R. Sjoholm, R. Leino, E. Laine, D.Y. Murzin, J. Catal. 204 (2001) 281-<br />

291.<br />

[56] J.A. Slipszenko, S.P. Griffiths, P. Johnston, K.E. Simons, W.A.H. Vermeer,<br />

P.B. Wells, J. Catal. 179 (1998) 267-276.<br />

[57] W.A.H. Vermeer, A. Fulford, P. Johnston, P.B. Wells, J. Chem. Soc., Chem.<br />

Commun. (1993) 1053-1054.<br />

[58] B. Torok, K. Felfoldi, K. Balazsik, M. Bartok, Chem. Commun. (1999) 1725-<br />

1726.<br />

[59] M. Studer, S. Burkhardt, H.-U. Blaser, Chem. Commun. (1999) 1727-1728.<br />

[60] M. von Arx, T. Mallat, A. Baiker, Catal. Lett. 78 (2002) 267-271.<br />

[61] M. von Arx, T. Mallat, A. Baiker, Tetrahedron: Asymmetry 12 (2001) 3089-<br />

3094.<br />

[62] M. von Arx, T. Mallat, A. Baiker, J. Catal. 193 (2000) 161-164.<br />

[63] K. Balazsik, K. Szori, K. Felfoldi, B. Torok, M. Bartok, Chem. Commun.<br />

(2000) 555-556.<br />

113


[64] A. Szabo, N. Kunzle, T. Mallat, A. Baiker, Tetrahedron: Asymmetry 10 (1999)<br />

61-76.<br />

[65] N. Kunzle, A. Szabo, M. Schurch, G.Z. Wang, T. Mallat, A. Baiker, Chem.<br />

Commun. (1998) 1377-1378.<br />

[66] G.Z. Wang, T. Mallat, A. Baiker, Tetrahedron: Asymmetry 8 (1997) 2133-2140.<br />

[67] W.R. Huck, T. Mallet, A. Baiker, J. Catal. 193 (2000) 1-4.<br />

[68] M. von Arx, T. Mallat, A. Baiker, Top. Catal. 19 (2002) 75-87.<br />

[69] D.Y. Murzin, P. Maki-Arvela, E. Toukoniitty, T. Salmi, Catal. Rev. - Sci. Eng.<br />

47 (2005) 175-256.<br />

[70] T. Bürgi, A. Baiker, J. Am. Chem. Soc. 120 (1998) 12920-12926.<br />

[71] M. Schürch, O. Schwalm, T. Mallat, J. Weber, A. Baiker, J. Catal. 169 (1997)<br />

275-286.<br />

[72] A.F. Carley, M.K. Rajumon, M.W. Roberts, P.B. Wells, Journal Of The<br />

Chemical Society-Faraday Transactions 91 (1995) 2167-2172.<br />

[73] K.E. Simons, P.A. Meheux, S.P. Griffiths, I.M. Sutherl<strong>and</strong>, P. Johnston, P.B.<br />

Wells, A.F. Carley, M.K. Rajumon, M.W. Roberts, A. Ibbotson, Recueil Des<br />

Travaux Chimiques Des Pays-Bas-Journal Of The Royal Netherl<strong>and</strong>s Chemical<br />

Society 113 (1994) 465-474.<br />

[74] T. Evans, A.P. Woodhead, A. Gutierrez-Sosa, G. Thornton, T.J. Hall, A.A.<br />

Davis, N.A. Young, P.B. Wells, R.J. Oldman, O. Plashkevych, O. Vahtras, H.<br />

Agren, V. Carravetta, Surf. Sci. 436 (1999) L691.<br />

[75] G. Bond, P.B. Wells, J. Catal. 150 (1994) 329.<br />

[76] D. Ferri, T. Burgi, J. Am. Chem. Soc. 123 (2001) 12074-12084.<br />

[77] W. Chu, R.J. LeBlanc, C.T. Williams, J. Kubota, F. Zaera, J. Phys. Chem. B 107<br />

(2003) 14365-14373.<br />

[78] C. LeBlond, J. Wang, J. Liu, A.T. Andrews, Y.K. Sun, J. Am. Chem. Soc. 121<br />

(1999) 4920.<br />

[79] M. Bartok, B. Torok, K. Balazsik, T. Bartok, Catal. Lett. 73 (2001) 127-131.<br />

[80] X.B. Li, N. Dummer, R. Jenkins, R.P.K. Wells, P.B. Wells, D.J. Willock, S.H.<br />

Taylor, P. Johnston, G.J. Hutchings, Catal. Lett. 96 (2004) 147-151.<br />

[81] M. Garl<strong>and</strong>, H.-U. Blaser, J. Am. Chem. Soc. 112 (1990) 7048-7050.<br />

[82] D. Ferri, T. Burgi, A. Baiker, J. Catal. 210 (2002) 160-170.<br />

[83] H.-U. Blaser, H.P. Jalett, D.M. Monti, A. Baiker, J.T. Wehrli, Stud. Surf. Sci.<br />

Catal. 67 (1991) 147.<br />

[84] O. Schwalm, B. Minder, J. Weber, A. Baiker, Catal. Lett. 23 (1994) 271-279.<br />

[85] O. Schwalm, J. Weber, B. Minder, A. Baiker, Theochem-Journal Of Molecular<br />

Structure 330 (1995) 353-357.<br />

[86] O. Schwalm, J. Weber, B. Minder, A. Baiker, Int. J. Quantum Chem. 52 (1994)<br />

191-197.<br />

[87] P.B. Wells, R.P.K. Wells, in: D.E. De Vos, I.F.J. Vankelecom, P.A. <strong>Jacobs</strong><br />

(Eds.), Chiral Catalyst Immobilization <strong>and</strong> Recycling, Wiley-VCH, Weinheim,<br />

2000, pp. 123-138.<br />

[88] G. Bond, P.A. Meheux, A. Ibbotson, P.B. Wells, Catal. Today 10 (1991) 371.<br />

[89] H.-U. Blaser, H.P. Jalett, M. Garl<strong>and</strong>, M. Studer, H. Thies, A. Wirth-Tijani, J.<br />

Catal. 173 (1998) 282.<br />

[90] J.L. Margitfalvi, M. Hegedus, J. Mol. Catal. A: Chem. 107 (1996) 281-289.<br />

[91] J.L. Margitfalvi, M. Hegedus, E. Tfirst, 11th International Congress On<br />

Catalysis - 40th Anniversary, Pts A And B, 1996, pp. 241-250.<br />

[92] J.L. Margitfalvi, M. Hegedus, E. Tfirst, Tetrahedron: Asymmetry 7 (1996) 571-<br />

580.<br />

114


[93] J.L. Margitfalvi, E. Talas, E. Tfirst, C.V. Kumar, A. Gergely, Appl. Catal., A<br />

191 (2000) 177.<br />

[94] H.-U. Blaser, H.P. Jalett, D.M. Monti, J.F. Reber, J.T. Wehrli, Stud. Surf. Sci.<br />

Catal. 41 (1988) 153.<br />

[95] K.E. Simons, A. Ibbotson, P. Johnston, H. Plum, P.B. Wells, J. Catal. 150<br />

(1994) 321-328.<br />

[96] E. Toukoniitty, P. Maki-Arvela, A.N. Villela, A.K. Neyestanaki, T. Salmi, R.<br />

Leino, R. Sjoholm, E. Laine, J. Vayrynen, T. Ollonqvist, P.J. Kooyman, Catal.<br />

Today 60 (2000) 175-184.<br />

[97] H.-U. Blaser, H.P. Jalett, M. Müller, M. Studer, Catal. Today 37 (1997) 441-<br />

463.<br />

[98] Y. Nitta, K. Kobiro, Chem. Lett. (1996) 897-898.<br />

[99] Y. Nitta, Y. Okamoto, Chem. Lett. (1998) 1115-1116.<br />

[100] P.A. Meheux, A. Ibbotson, P.B. Wells, J. Catal. 128 (1991) 387-396.<br />

[101] Y. Nitta, T. Imanaka, S. Teranishi, J. Catal. 80 (1983) 31-39.<br />

[102] A. Baiker, J. Mol. Catal. A: Chem. 115 (1997) 473-493.<br />

[103] M. Schürch, T. Heinz, R. Aeschimann, T. Mallat, A. Pfaltz, A. Baiker, J. Catal.<br />

173 (1998) 187.<br />

[104] T. Heinz, Phd Thesis, Universität Basel, Basel, 1997.<br />

[105] H.A. Wierenga, L. Soethout, J.W. Gerritsen, B.E.C. V<strong>and</strong>eleemput, H.<br />

Vankempen, G. Schmid, Advanced Materials 2 (1990) 482-484.<br />

[106] X.B. Zuo, H.F. Liu, D.W. Guo, X.Z. Yang, Tetrahedron 55 (1999) 7787-7804.<br />

[107] A. Kockritz, S. Bisch<strong>of</strong>f, V. Morawsky, U. Prusse, K.D. Vorlop, J. Mol. Catal.<br />

A: Chem. 180 (2002) 231-243.<br />

[108] A. Gamez, J. Kohler, J. Bradley, Catal. Lett. 55 (1998) 73-77.<br />

[109] J.U. Kohler, J.S. Bradley, Langmuir 14 (1998) 2730-2735.<br />

[110] J.U. Kohler, J.S. Bradley, Catal. Lett. 45 (1997) 203-208.<br />

[111] J.S. Bradley, J. Kohler, D. deCaro, Abstracts Of Papers Of The American<br />

Chemical Society 212 (1996) 14-CATL.<br />

[112] Y.L. Huang, J.R. Chen, H. Chen, R.X. Li, Y.Z. Li, L.E. Min, X.J. Li, J. Mol.<br />

Catal. A: Chem. 170 (2001) 143-146.<br />

[113] J. Zhang, X.P. Yan, H.F. Liu, J. Mol. Catal. A: Chem. 175 (2001) 125-130.<br />

[114] H. Bönnemann, G.A. Braun, Angew. Chem., Int. Ed. 35 (1996) 1992-1995.<br />

[115] H. Bönnemann, G.A. Braun, Chem. Eur. J. 3 (1997) 1200-1202.<br />

[116] X.B. Zuo, H.F. Liu, M.H. Liu, Tetrahedron Lett. 39 (1998) 1941-1944.<br />

[117] X.B. Zuo, H.F. Liu, J.F. Tian, J. Mol. Catal. A: Chem. 157 (2000) 217-224.<br />

[118] S. Jansat, M. Gomez, K. Philippot, G. Muller, E. Guiu, C. Claver, S. Castillon,<br />

B. Chaudret, J. Am. Chem. Soc. 126 (2004) 1592-1593.<br />

[119] H.H. Helcher, J. Herbord Klossen, 1718.<br />

[120] M. Faraday, Philos. Trans. R. Soc. London 147 (1857) 145.<br />

[121] U. Kolb, S.A. Quaiser, M. Winter, M.T. Reetz, Chem. Mater. 8 (1996) 1889-<br />

1894.<br />

[122] N. Toshima, Y. Wang, Langmuir 10 (1994) 4574-4580.<br />

[123] G. Schmid, Endeavour 14 (1990) 172-178.<br />

[124] DEFINITION AND CLASSIFICATION OF COLLOIDS, IUPAC,<br />

http://www.iupac.org/reports/2001/colloid_2001/manual_<strong>of</strong>_s_<strong>and</strong>_t/node33.ht<br />

ml<br />

[125] Classification <strong>of</strong> colloids, Wikipedia, http://en.wikipedia.org/wiki/Colloid<br />

[126] M.A. Watzky, R.G. Finke, Chem. Mater. 9 (1997) 3083-3095.<br />

[127] M.A. Watzky, R.G. Finke, J. Am. Chem. Soc. 119 (1997) 10382-10400.<br />

115


[128] J.D. Aiken III, R.G. Finke, J. Mol. Catal. A: Chem. 145 (1999) 1-44.<br />

[129] B.K. Teo, N.J.A. Sloane, Inorg. Chem. 24 (1985) 4545-4558.<br />

[130] N.J.A. Sloane, B.K. Teo, J. Chem. Phys. 83 (1985) 6520-6534.<br />

[131] A.J. Bard, L.R. Faulkner, Electrochemical Methods, Wiley, New York, 1980.<br />

[132] A. Roucoux, J. Schulz, H. Patin, Chemical Reviews 102 (2002) 3757-3778.<br />

[133] A. Amulyavichus, A. Daugvila, R. Davidonis, C. Sipavichus, Fiz. Met.<br />

Metalloved. 85 (1998) 111-117.<br />

[134] E. Gaffet, M. Tachikart, O. ElKedim, R. Rahouadj, Mater. Charact. 36 (1996)<br />

185-190.<br />

[135] H. Bönnemann, W. Brijoux, R. Brinkmann, R. Fretzen, T. Joussen, R. Koppler,<br />

B. Korall, P. Neiteler, J. Richter, J. Mol. Catal. 86 (1994) 129-177.<br />

[136] R. Houbertz, T. Feigenspan, F. Mielke, U. Memmert, U. Hartmann, U. Simon,<br />

G. Schon, G. Schmid, Europhys. Lett. 28 (1994) 641-646.<br />

[137] G. Schmid, S. Peschel, New J. Chem. 22 (1998) 669-675.<br />

[138] U. Simon, R. Flesch, H. Wiggers, G. Schon, G. Schmid, J. Mater. Chem. 8<br />

(1998) 517-518.<br />

[139] M. Bardaji, O. Vidoni, A. Rodriguez, C. Amiens, B. Chaudret, M.J. Casanove,<br />

P. Lecante, New J. Chem. 21 (1997) 1243-1249.<br />

[140] G. Schmid, H. West, J.O. Malm, J.O. Bovin, C. Grenthe, Chem. Eur. J. 2 (1996)<br />

1099-1103.<br />

[141] M.N. Vargaftik, V.P. Zagorodnikov, I.P. Stolarov, Moiseev, II, D.I. Kochubey,<br />

V.A. Likholobov, A.L. Chuvilin, K.I. Zamaraev, J. Mol. Catal. 53 (1989) 315-<br />

348.<br />

[142] G. Schmid, B. Morun, J.O. Malm, Angew. Chem., Int. Ed. Engl. 28 (1989) 778-<br />

780.<br />

[143] G. Schmid, A. Lehnert, Angew. Chem., Int. Ed. Engl. 28 (1989) 780-781.<br />

[144] G. Schmid, Polyhedron 7 (1988) 2321-2329.<br />

[145] G. Schmid, Applied Homogeneous Catalysis with Organometallic Compounds,<br />

Wiley-VCH, 1996, p. 677.<br />

[146] R. Franke, J. Rothe, J. Pollmann, J. Hormes, H. Bönnemann, W. Brijoux, T.<br />

Hindenburg, J. Am. Chem. Soc. 118 (1996) 12090-12097.<br />

[147] O. Vidoni, K. Philippot, C. Amiens, B. Chaudret, O. Balmes, J.O. Malm, J.O.<br />

Bovin, F. Senocq, M.J. Casanove, Angew. Chem., Int. Ed. 38 (1999) 3736-<br />

3738.<br />

[148] K. Patel, S. Kapoor, D.P. Dave, T. Mukherjee, J. Chem. Sci. 117 (2005) 311-<br />

316.<br />

[149] S.H. Sun, C.B. Murray, D. Weller, L. Folks, A. Moser, <strong>Science</strong> 287 (2000)<br />

1989-1992.<br />

[150] P.H. Hess, P.H. Parker, J. Appl. Polym. Sci. 10 (1966) 1915.<br />

[151] Y. Kato, S. Sugimoto, K. Shinohara, N. Tezuka, T. Kagotani, K. Inomata,<br />

Mater. Trans. 43 (2002) 406-409.<br />

[152] M.T. Reetz, W. Helbig, J. Am. Chem. Soc. 116 (1994) 7401-7402.<br />

[153] S.I. Stoeva, B.L.V. Prasad, S. Uma, P.K. Stoimenov, V. Zaikovski, C.M.<br />

Sorensen, K.J. Klabunde, J. Phys. Chem. B 107 (2003) 7441-7448.<br />

[154] W.T. Nichols, T. Sasaki, N. Koshizaki, Euromat, Prague, 2005.<br />

[155] K.M. Beck, T. Sasaki, N. Koshizaki, Chem. Phys. Lett. 301 (1999) 336-342.<br />

[156] X.Y. Zeng, N. Koshizaki, T. Sasaki, A. Yabe, F. Rossignol, H. Nagai, Y.<br />

Nakata, T. Okutani, M. Suzuki, Appl. Surf. Sci. 140 (1999) 90-98.<br />

[157] Turkevic.J, G. Kim, <strong>Science</strong> 169 (1970) 873.<br />

116


[158] H. Bönnemann, G. Braun, W. Brijoux, R. Brinkmann, A.S. Tilling, K. Seevogel,<br />

K. Siepen, J. Organomet. Chem. 520 (1996) 143.<br />

[159] H. Bönnemann, W. Brijoux, Active Metals, VCH, Weinheim, 1996.<br />

[160] H. Bönnemann, R.M. Richards, Eur. J. Inorg. Chem. (2001) 2455-2480.<br />

[161] H. Bönnemann, W. Brijoux, R. Brinkmann, E. Dinjus, T. Joussen, B. Korall,<br />

Angew<strong>and</strong>te Chemie-International Edition In English 30 (1991) 1312-1314.<br />

[162] L. D'Souza, Phd Thesis, International <strong>University</strong> Bremen, Bremen, 2005.<br />

[163] COST Chemistry Action D24, Sustainable Chemical Processes: Stereoselective<br />

Transition Metal-Catalysed Reactions,<br />

http://costchemistry.epfl.ch/load/d24_mid_term_report.pdf<br />

[164] T. Mallat, A. Baiker, J. Catal. 176 (1998) 271-274.<br />

[165] D.G. Blackmond, J. Catal. 176 (1998) 267-270.<br />

[166] T. Mallat, Z. Bodnar, B. Minder, K. Borszeky, A. Baiker, J. Catal. 168 (1997)<br />

183-193.<br />

[167] U.K. Singh, R.N. L<strong>and</strong>au, Y.K. Sun, C. Leblond, D.G. Blackmond, S.K.<br />

Tanielyan, R.L. Augustine, J. Catal. 154 (1995) 91-97.<br />

[168] Y.K. Sun, J. Wang, C. Leblond, R.N. L<strong>and</strong>au, D.G. Blackmond, J. Catal. 161<br />

(1996) 759-765.<br />

[169] K. Balazsik, M. Bartok, J. Catal. 224 (2004) 463-472.<br />

[170] X.B. Li, R.P.K. Wells, P.B. Wells, G.J. Hutchings, Catal. Lett. 89 (2003) 163.<br />

[171] V. Morawsky, Phd Thesis, Braunschweig, 2002.<br />

[172] T.O. Ely, C. Pan, C. Amiens, B. Chaudret, F. Dassenoy, P. Lecante, M.J.<br />

Casanove, A. Mosset, M. Respaud, J.M. Broto, J. Phys. Chem. B 104 (2000)<br />

695-702.<br />

[173] K. Moseley, P.M. Maitlis, J. Chem. Soc. D (1971) 982-983.<br />

[174] P. Yuan, D.Q. Wu, H.P. He, Z.Y. Lin, Appl. Surf. Sci. 227 (2004) 30-39.<br />

[175] M. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R.<br />

Cheeseman, V.G. Zakrzewski, J.A. Montgomery, Jr., R.E. Sratmann, J.C.<br />

Burant, S. Dapprich, J.M. Millam, A.D. Daniels, R. Cammi, B. Mennucci, C.<br />

Pomelli, C. Adamo, S. Clifford, J. Ochterski, G.A. Petersson, P.Y. Ayala, Q.<br />

Ciu, K. Morokuma, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B.<br />

Foresman, J.P. Cioslowski, I. Komaromi, R. Gompets, R.L. Martin, D.J. Fox, T.<br />

Keith, M. Al-Laham, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, J.L.<br />

Andres, C. Gonzalez, M. Head-Gordon, E.S. Replogle, J.A. Pople, Gaussian<br />

Inc., Pittsburgh, 2003.<br />

[176] W.P. Davey, Physical Review 25 (1925) 763.<br />

[177] P. Scherrer, Göttingen Nachrichten 2 (1918) 98.<br />

[178] J. Kubota, Z. Ma, F. Zaera, Langmuir 19 (2003) 3371-3376.<br />

[179] R.G. Greenler, J. Chem. Phys. 44 (1966) 310.<br />

[180] H.-U. Blaser, M. Garl<strong>and</strong>, H.P. Jallet, J. Catal. 144 (1993) 569-578.<br />

[181] H.-U. Blaser, H.P. Jalett, J. Wiehl, J. Mol. Catal. 68 (1991) 215-222.<br />

[182] E. Orglmeister, T. Mallat, A. Baiker, J. Catal. 233 (2005) 333-341.<br />

[183] B. Minder, T. Mallat, A. Baiker, G. Wang, T. Heinz, A. Pfaltz, J. Catal. 154<br />

(1995) 371.<br />

[184] D. Ferri, T. Burgi, K. Borszeky, T. Mallat, A. Baiker, J. Catal. 193 (2000) 139-<br />

144.<br />

[185] V. Morawsky, U. Pruße, L. Witte, K.D. Vorlop, Catal. Commun. 1 (2000) 15.<br />

[186] K. Balazsik, M. Bartok, J. Catal. 224 (2004) 463.<br />

[187] J. Kubota, F. Zaera, J. Am. Chem. Soc. 123 (2001) 11115.<br />

[188] B.E. Spiewak, R.D. Cortright, J.A. Dumesic, J. Catal. 176 (1998) 405-414.<br />

117


[189] K. Umezawa, T. Ito, M. Asada, S. Nakanishi, P. Ding, W.A. Lanford, B.<br />

Hjorvarsson, Surf. Sci. 387 (1997) 320-327.<br />

[190] M.R. Voss, H. Busse, B.E. Koel, Surf. Sci. 414 (1998) 330-340.<br />

[191] J. Fearon, G.W. Watson, J. Mater. Chem. 16 (2006) 1989-1996.<br />

[192] Asymmetric Hydrogenation <strong>of</strong> Activated Ketones with catASium® F214, R.<br />

Hartung, J.G.E. Krauter, S. Seebald, P. Kukula, U. Nettekoven, M. Studer,<br />

http://www.degussa-catalysts.com/asp/documents/79_Document.pdf<br />

[193] CatASium® F214 Data Sheet heterogeneous catalyst for the enantioselective<br />

hydrogenation <strong>of</strong> -functionalised ketones, http://www.degussacatalysts.com/asp/documents/125_Document.pdf<br />

[194] K. Balazsik, M. Bartok, J. Mol. Catal. A: Chem. 219 (2004) 383-389.<br />

[195] Z. Ma, J. Kubota, F. Zaera, J. Catal. 219 (2003) 404-416.<br />

[196] C.P. Hwang, C.T. Yeh, J. Mol. Catal. A: Chem. 112 (1996) 295-302.<br />

[197] C.E. Smith, J.P. Biberian, G.A. Somorjai, J. Catal. 57 (1979) 426-443.<br />

[198] T. Matsushima, D.B. Almy, J.M. White, Surf. Sci. 67 (1977) 89-108.<br />

[199] H.H. Rotermund, J. Lauterbach, G. Haas, Appl. Phys. A: Mater. Sci. Process. 57<br />

(1993) 507-511.<br />

[200] H. Niehus, G. Comsa, Surf. Sci. 93 (1980) L147-L150.<br />

[201] P.R. Norton, R.L. Tapping, J.W. Goodale, J. Vac. Sci. Technol. 14 (1977) 446-<br />

451.<br />

[202] D. Ferri, T. Burgi, A. Baiker, J. Phys. Chem. B 105 (2001) 3187-3195.<br />

[203] B. Minder, M. Schurch, T. Mallat, A. Baiker, T.P.A. Heinz, J. Catal. 160<br />

(1996) 261.<br />

[204] C. Keresszegi, T. Burgi, T. Mallat, A. Baiker, J. Catal. 211 (2002) 244-251.<br />

[205] R.L. Augustine, S.K. Tanielyan, J. Mol. Catal. A: Chem. 118 (1997) 79-87.<br />

[206] U. Bohmer, F. Franke, K. Morgenschweis, T. Bieber, W. Reschetilowski, Catal.<br />

Today 60 (2000) 167-173.<br />

[207] X.H. Li, X. You, P.L. Ying, J.L. Xiao, C. Li, Top. Catal. 25 (2003) 63-70.<br />

[208] J.R.G. Perez, J. Malthete, J. Jacques, Comptes Rendus De L Academie Des<br />

<strong>Science</strong>s Serie Ii 300 (1985) 169-172.<br />

[209] Y. Nitta, A. Shibata, Chem. Lett. (1998) 161-162.<br />

[210] Y. Nitta, K. Kobiro, Y. Okamoto, Stud. Surf. Sci. Catal. 108 (1997) 191-198.<br />

[211] Y. Nitta, K. Kobiro, Chem. Lett. (1995) 165-165.<br />

[212] Y. Nitta, Y. Ueda, T. Imanaka, Chem. Lett. (1994) 1095-1098.<br />

[213] Y. Nitta, K. Kobiro, Y. Okamoto, Proc. 70th Ann. Meeting Chem. Soc. Japan,<br />

1996, p. 573.<br />

[214] H.-U. Blaser, M. Müller, Stud. Surf. Sci. Catal. 59 (1991) 73.<br />

[215] Y. Nakamura, Bull. Chem. Soc. Japan 16 (1941) 367.<br />

[216] T.J. Hall, P. Johnston, W.A.H. Vermeer, S.R. Watson, P.B. Wells, 11th<br />

International Congress On Catalysis - 40th Anniversary, Pts A And B, 1996, pp.<br />

221-230.<br />

[217] P.J. Collier, T.J. Hall, J.A. Iggo, P. Johnston, J.A. Slipszenko, P.B. Wells, R.<br />

Whyman, Chem. Commun. (1998) 1451-1452.<br />

[218] J. Haglund, A.F. Guillermet, G. Grimvall, M. Korling, Physical Review B 48<br />

(1993) 11685-11691.<br />

[219] T.J. Hall, P. Johnston, W.A.H. Vermeer, S.R. Watson, P.B. Wells, Stud. Surf.<br />

Sci. Catal. 101 (1996) 221.<br />

[220] F. Gao, L. Chen, M. Garl<strong>and</strong>, J. Catal. 238 (2006) 402-411.<br />

[221] Swanson, Tatge, Natl. Bur. St<strong>and</strong>. (U.S.), Circ. 539 (1953) 31.<br />

[222] Hull, Physical Review 17 (1921) 571.<br />

118


[223] R. Noyori, H. Takaya, Acc. Chem. Res. 23 (1990) 345-350.<br />

[224] M. Tamura, H. Fujihara, J. Am. Chem. Soc. 125 (2003) 15742-15743.<br />

[225] A.G. Hu, G.T. Yee, W.B. Lin, J. Am. Chem. Soc. 127 (2005) 12486-12487.<br />

[226] S.U. Son, Y. Jang, K.Y. Yoon, E. Kang, T. Hyeon, Nano Lett. 4 (2004) 1147-<br />

1151.<br />

119


I declare that the current Ph.D. thesis has been done by myself.<br />

Alex<strong>and</strong>er Kraynov,<br />

Date: 24.07.2006<br />

120

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!