ruthenium-catalyzed formylations using

179
Research Collection Doctoral Thesis Ruthenium-catalyzed formylations using carbon dioxide as carbon source ans solvent Author(s): Rohr, Markus Publication Date: 2005 Permanent Link: https://doi.org/10.3929/ethz-a-005114149 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library

Upload: phamlien

Post on 12-Jan-2017

239 views

Category:

Documents


1 download

TRANSCRIPT

Research Collection

Doctoral Thesis

Ruthenium-catalyzed formylations using carbon dioxide ascarbon source ans solvent

Author(s): Rohr, Markus

Publication Date: 2005

Permanent Link: https://doi.org/10.3929/ethz-a-005114149

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

Diss. ETH No. 16281

Ruthenium-catalyzed Formylations Using

Carbon Dioxide as Carbon Source

and Solvent

A dissertation submitted to the

Swiss Federal Institute of Technology Zurich (ETH)

for the degree of Doctor of Technical Sciences

presented by

Markus Rohr

Dipl. Chem.-Ing. ETH

born April 10, 1972

citizen of Hunzenschwil (AG)

accepted on the recommendation of

Prof. Dr. A. Baiker, examiner

Prof. Dr. G. Consiglio, co-examiner

2005

Dedicated to myparents

Myrtaf and Fritz Rohr-Deubelbeiss

for all the love, trust, and support

Acknowledgments

Firstly, I would like to express my sincere gratitude to Prof. Dr. A. Baiker

for his support, both personally and professionally, and the opportunity to

complete my doctoral studies in his group.

Moreover, I would like to thank Prof. Dr. G. Consiglio for accepting the

task of co-examiner for this thesis.

A special thank is due to Dr. J.-D. Grunwaldt for his support and

contributions to this thesis. I really appreciated the assistance, scientific

discussions and efforts during the last years.

Furthermore, I would like to thank Dr. C. Beck and W. Kunz for their

valuable help and the many scientific and non-scientific discussions, and

M. Günther for his contributions to this work during his diploma work.

Thanks are also due to my office-mates M. Ramin, Dr. M. Burgener,

M. Caravati, S. Diezi, F. Jutz, and N. van Vegten for sharing a lot of good

times and unforgettable moments, and to the entire Baiker group for their

company and support.

I thank U. Krebs, M. Kupfer, R. Mäder, and P. Trüssel for their help with

fine mechanics and M. Wohlwend for electronic support.

Thanks are due to HASYLAB, DESY in Hamburg, Germany for providing

beamtime and to the lab's beamline staff, M. Herrmann and J. Wienold, for

their support, and also to H. Emerich and W. van Beek of the Swiss-

Norwegian Beamline at ESRF, Grenoble, France for offering beamtime and

for support during the measurements. The Bundesamt für Energie is to be

thanked for its financial support.

Finally, I would like to thank my family and my friends for their love and

support throughout all these years of my education.

Table of Contents

Acknowledgments v

Table of Contents ix

Summary xv

Zusammenfassung xix

1 Introduction 1

1.1 General Aspects of Carbon Dioxide Utilization 1

1.2 Physical and Chemical Properties of Carbon Dioxide 2

1.3 Carbon Dioxide as a Ci-Building Block 4

1.3.1 Synthesis Using Carbon Dioxide 4

1.3.2 Formic Acid and its Derivatives 7

1.3.3 Further Reactions Using Carbon Dioxide as

Ci-Building Block 9

1.4 Carbon Dioxide as a Solvent 12

1.5 Scope of the Thesis 14

1.6 References 16

2 Experimental 27

2.1 Experimental Equipment 27

2.1.1 High-pressure Batch Reactor Cell 27

X

2.1.2 Apparatus for Synthesis of Catalysts in the

Presence of UV-Light 29

2.1.3 View-cell for Phase Behavior Studies 30

2.2 Characterization Methods 30

2.2.1 X-ray Absorption Spectroscopy Cell for Catalyst

Studies in Liquid Phase 31

2.2.2 High-pressure XAS Cell for Studying Simultaneously

the Liquid Phase and the Solid/Liquid Interface 33

2.3 Catalytic Tests 37

2.3.1 Experimental Procedure 37

2.3.2 Analysis 37

2.3.3 Evaluation of Catalytic Tests 38

2.4 References 39

3 Catalyst Characterization under Reaction Conditions

using XANES and EXAFS 41

3.1 Summary 41

3.2 Introduction 42

3.3 Basics of XANES and EXAFS 43

3.4 High-pressure In Situ X-ray Absorption Spectroscopy 47

3.5 Case Study: Ruthenium-catalyzed Formylation of

3-Methoxypropylamine with Hydrogen and

Supercritical Carbon Dioxide 48

3.5.1 Introduction 48

3.5.2 Experimental 49

3.5.3 Results 50

3.5.4 Discussion 52

3.5.5 Conclusions 53

xi

3.6 References 53

4 A Simple Route to Highly Active Ruthenium Catalysts

for Formylation Reactions with Hydrogen and

Carbon Dioxide 59

4.1 Summary 59

4.2 Introduction 60

4.3 Experimental 61

4.4 Results and Discussion 63

4.5 Conclusions 69

4.6 References 69

5 Formylation with Supercritical Carbon Dioxide over

RU/AI2O3 Modified by Phosphines: Heterogeneous or

Homogeneous Catalysis? 73

5.1 Summary 73

5.2 Introduction 74

5.3 Experimental 75

5.3.1 Catalyst Materials 75

5.3.2 Catalytic Formylation of 3-Methoxypropylamine 76

5.3.3 Characterization ofIn Situ Formed Homogeneous

Complex Using ICP-OES 77

5.3.4 Ex Situ and In Situ X-ray Absorption Spectroscopic

Studies 77

5.4 Results 79

5.4.1 Modification of RU/AI2O3 with dppe and its Performance

in the Formylation of 3-Methoxypropylamine 79

5.4.2 Identification of the Catalytically Active Species 82

Xll

5.4.3 Spectroscopic Study of the In Situ Formation of the

Homogeneous Ruthenium Catalyst 86

5.4.4 Structural Identification of the Homogeneous

Ruthenium Complex 88

5.5 Discussion 94

5.6 Conclusions 98

5.7 References 98

6 Solvent-free Ruthenium-catalyzed Vinyl Carbamate

Synthesis from Phenylacetylene and Diethylamine in

Supercritical Carbon Dioxide 103

6.1 Summary 103

6.2 Introduction 104

6.3 Experimental 105

6.4 Results and Discussion 106

6.5 Conclusions 110

6.6 References 110

7 Evaluation of Strategies for the Immobilization of Bidentate

Ruthenium Phosphine Complexes Used for the Formylation

of Amines in Supercritical Carbon Dioxide 113

7.1 Summary 113

7.2 Introduction 114

7.3 Experimental 116

7.3.1 Preparation of the Covalently Bound RuCl2(bspe)2

Catalyst 116

7.3.2 Preparation of the Adsorbed Catalyst RuCl2(dppe)2 on

Aminopropyl-modified Silica 119

Xlll

7.3.3 Preparation of Further Catalytic Materials, Covalently

Bound RuCl2(bspp)2 and In Situ Generated Catalyst

from "Si"-NH2-RuCl3 and dppe 120

7.3.4 Catalytic Formylation of Amines with Supercritical

Carbon Dioxide 121

7.3.5 Physico-chemical Characterization Using X-ray

Absorption Spectroscopy (XANES, EXAFS) and

Further Characterization Methods (XPS, Nitrogen

Physisorption, NMR, and DRIFTS) 121

7.4 Results 125

7.4.1 Preparation and Characterization of

Heterogeneous Catalysts 125

7.4.2 Catalytic Formylation of Amines Using

Immobilized RuCl2(bspe)2 130

7.4.3 Comparison of the Immobilized Ruthenium

Complex and the Adsorbed Ruthenium Complex 132

7.4.4 Structural Analysis of the Solid Catalysts and

the Soluble Species after Reaction 133

7.5 Discussion 138

7.6 Conclusions 142

7.7 References 143

Final Remarks 147

List of Publications 151

Curriculum Vitae 155

Summary

The use of carbon dioxide as a carbon source and solvent is an attractive

approach to the synthesis of formic acid and its derivatives. For that reason,

this strategy has been used in a number of former and present studies.

Formamides can be synthesized using group (VIII) transition metals as

catalysts for the formylation of amines with carbon dioxide and hydrogen.

Applicable amines range from simple primary and secondary amines

(dimethylamine, diethylamine) to cyclic amines, such as, piperidine,

pyrrolidine, and morpholine. Ruthenium phosphine complexes constitute

efficient catalysts for the formylation of amines. In this reaction, carbon

dioxide acts both as reactant and solvent, resulting in a "solventless"

process.

The main focus of this doctoral thesis was the development of new and

simpler homogeneous and heterogeneous catalysts for formylation

reactions, complemented by spectroscopic investigations under ex situ and

in situ conditions. A very simple catalyst system, not investigated

previously, consists of commercially available RUCI3 in the presence of

l,2-bis(diphenylphosphino)ethane (dppe). Catalytic and spectroscopic

investigations using X-ray absorption near edge structure (XANES) and

XVI

extended X-ray absorption fine structure (EXAFS) spectroscopy revealed

that a RuCl2(PPli3)3-like complex formed in situ.

The in situ formed complex was highly active and exhibited turnover rates

comparable to the stepwise synthesized RuCl2(dppe)2 complex. Both of

these catalysts were more active than a RuCl2(PPli3)3 catalyst.

In our search for a simple heterogeneous catalyst system for the

formylation of amines, we modified a commercial RU/AI2O3 catalyst by

addition of dppe. Although this catalyst afforded high activity, closer

inspection revealed that the catalytically active species was a homogeneous

Ru-complex. This could be confirmed with the help of a novel

spectroscopic in situ batch reactor cell for X-ray absorption spectroscopy.

By investigating both the solid and the liquid phase, it could be proven that

a homogeneous complex formed as soon as dppe was added. The solid

RU/AI2O3 catalyst was reduced as a consequence of the presence of

hydrogen. Further catalytic studies and elemental analysis by ICP-OES

showed that the formed ruthenium complex acts as a homogeneous catalyst

and is present in only small amounts in the reaction mixture (50 -

200 ppm). Its structure could be further elucidated using a specially

designed EXAFS cell with a long path length to record transmission

EXAFS spectra in liquid solution even at very low concentrations. The

structure of the active species was found to be similar to that of a

Ru(dppe)2X2 complex.

In a final step, different immobilization procedures for the presently most

active RuCl2(dppe)2 complex were explored. Coordinatively anchored

complexes on specially modified silica did not prove to be sufficiently

stable, probably because of the harsh reaction conditions and leaching of

the complex. The best results were obtained using covalently immobilized

XV11

ruthenium complexes. This catalyst also deactivated significantly less than

the coordinatively anchored complexes. The deactivation was attributed to

the destruction of the immobilized catalyst, as also evidenced by

spectroscopic studies.

Apart from these new catalyst preparation routes, the formylation of

primary and secondary amines was extended to substrates with other

functional groups, such as additional ether or alcohol groups (3-methoxy-

propylamine, 2-ethylaminoethanol, 2-methylaminoethanol, and morpho¬

line) in order to extend the scope of possible formylation reactions in

supercritical carbon dioxide.

Finally, the synthesis of vinyl carbamate from phenylacetylene,

diethylamine, and carbon dioxide was targeted in our attempts to extend the

scope of reactions which utilize carbon dioxide as a Ci-building block and

as solvent. RuCl2(C5H5N)4 and RuCl2(/76-C6H6)PMe3 catalysts were

found to be suitable for this reaction, making it a "green" alternative to the

synthesis of vinyl carbamates using phosgene.

Zusammenfassung

Die Verwendung von Kohlendioxid als Kohlenstoffbaustein und Lösungs¬

mittel ist ein interessanter Ansatz zur Herstellung von Ameinsensäure und

ihrer Derivate. Im Speziellen können Formamide aus Kohlendioxid und

Wasserstoff durch übergangsmetallkatalysierte Formylierungen an Aminen

synthetisiert werden. Primäre, sekundäre sowie auch zyklische Amine sind

für Formylierungen zugänglich. Beispiele dafür sind Diethylamin, Piperi-

din, Pyrrolidin und Morpholin. Ruthenium-Phosphin-Komplexe zeigen

exzellente katalytische Eigenschaften für die Formylierung in über¬

kritischem Kohlendioxid, das gleichzeitig als Edukt und Lösungsmittel

dient.

Der Fokus in dieser Doktorarbeit wurde auf die Entwicklung neuartiger

und einfacherer homogener sowie heterogener Katalysatoren für

Formylierungen in überkritischem Kohlendioxid gelegt. Informationen

über die lokale Struktur und das Verhalten der Katalysatoren unter

Reaktionsbedingungen wurden mit ex situ und in situ Röntgenabsorptions-

spektroskopie gewonnen.

Ein einfaches - bisher noch nicht betrachtetes - katalytisches System geht

von handelsüblichem RUCI3 und l,2-Bis(diphenylphosphino)ethane (dppe)

aus. Katalytische und spektroskopische Untersuchungen mit Röntgen-

XX

absorptionsspektroskopie (X-ray absorption near edge structure, XANES,

und extended X-ray absorption fine structure, EXAFS) belegten die in situ

Bildung eines hochaktiven Komplexes, dessen Struktur ähnlich derjenigen

von RuCl2(PPh3)3 ist.

Auf der Suche nach einem einfachen heterogenen Katalysatorsystem für

Formylierungen von Aminen wurde ein kommerziell erhältlicher RU/AI2O3

Katalysator durch Zugabe von dppe modifiziert. Röntgenabsorptions-

spektroskopische Untersuchungen zeigten die Bildung homogener

Ruthenium-Komplexe sofort nach Zugabe von dppe. Dazu wurde eine

Batch-Reaktorzelle konstruiert, die gleichzeitige in situ Messungen an der

festen (Katalysator) und an der flüssigen Phase (Reaktionsmischung)

ermöglichte. Der anwesende Wasserstoff reduzierte dabei den an der Luft

teilweise oxidierten RU/AI2O3 Katalysator. Weitere katalytische Studien

und Elementaranalysen mittels ICP-OES zeigten, dass der gebildete

Ruthenium-Komplex als homogener Katalysator vorliegt und in der

Reaktionsmischung nur in sehr kleinen Mengen vorhanden ist (50-

200 ppm). Um die Strukturaufklärung bei diesen kleinen Konzentrationen

zu ermöglichen, wurde eine EXAFS-Flüssigkeitszelle mit grösserer

Weglänge für die Röntgenstrahlen entworfen. Mit EXAFS Messungen in

Transmissionsgeometrie wurden Spektren der katalytisch aktiven Spezies

in der flüssigen Phase erhalten. Die Analyse der Spektren zeigte, dass die

Struktur dieser Partikel einem Ru(dppe)2X2-Komplex gleicht.

Verschiedene Immobilisierungsverfahren wurden für die hochaktiven

RuCl2(dppe)2-Komplexe entwickelt und verglichen. Auf Amin-

modifiziertem Silika koordinativ am Zentralatom gebundene Komplexe

zeigten keine ausreichende Stabilität. Die besten Resultate wurden mit

kovalent immobilisierten Ruthenium-Komplexen erreicht. Diese

XXI

Katalysatoren zeigten eine wesentlich geringere Desaktivierung als die

koordinativ verankerten Komplexe.

Abgesehen von diesen neuen Varianten der Katalysatorherstellung wurde

die Palette möglicher Formylierungen durch Amine mit zusätzlichen

funktionellen Gruppen erweitert. Amine mit Ether- oder Alkoholgruppen,

wie 3-Methoxypropylamin, 2-Ethylaminoethanol, 2-Methylaminoethanol

und Morpholin wurden erfolgreich zu den entsprechenden Formamiden

umgesetzt.

Als weitere Reaktion, die ebenfalls Kohlendioxid als Ci-Baustein und

Lösungsmittel verwendet, wurde die Synthese von Vinylcarbamat aus

Phenylacetylen, Diethylamin und überkritischem Kohlendioxid untersucht.

RuCl2(C5H5N)4 and RuCl2(/76-C6H6)PMe3 waren in dieser Reaktion

katalytisch aktiv, die eine „grüne" Alternative zur Synthese von Vinyl¬

carbamat aus Phosgen darstellt.

Chapter

Introduction

1.1 General Aspects of Carbon Dioxide Utilization

The large amount of carbon dioxide released into the atmosphere every

year has been the topic of countless heated discussions about the role this

greenhouse gas plays on the future state of the climate. In energy

production, carbon dioxide is emitted by combustion processes, based on

fossil raw materials, such as crude oil, natural gas, and coal. Other sources

are, for example, chemical processes (e.g. hydrogen production) and the

cement industry.

The concentration of carbon dioxide in the atmosphere has increased in the

last 200 years from 250 ppm to 360 ppm [1], mostly due to the growing

demand for energy and to the industrial production of chemicals and

pharmaceuticals. As a consequence, the natural CO2 equilibrium is

expected to become noticeably unbalanced by human industrial activities.

In nature, the CO2 equilibrium is maintained by a closed loop involving the

fixation of atmospheric carbon dioxide in organic material by the

photosynthesis process, the refeeding by animal respiration, and the

degradation of organic material. In the photosynthesis process, one of the

most important chemical processes on earth, CO2 and water are converted

to glucose and oxygen with the effect of sunlight.

1

2 Chapter 1

Strategies to avoid further increases of CO2 emission, and to extend the

availability of the non-renewable fossil raw material sources are of

tremendous importance to the development of earth's climate and, with it,

the future of mankind. Promising ways to implement such strategies exist,

including energy saving measures, increasing efficiency in energy

production, optimizing consumption processes, and increasingly making

use of non-fossil renewable energy sources.

Another attractive strategy is the use of carbon dioxide as a source of

carbon for chemical synthesis. The abundance of carbon dioxide released

by the processes described above constitutes a source of cheap and easily

available Ci-building blocks. Carbon dioxide is non-toxic and has the

capability to replace other toxic Ci -building blocks, such as carbon

monoxide and phosgene [2-5].

In addition, the excellent properties of supercritical carbon dioxide with

regard to miscibility, diffusion, and mass transfer coefficients make it a

preferred solvent in chemical processes, at the same time offering a

replacement for environmentally hazardous organic solvents [6-8].

1.2 Physical and Chemical Properties of Carbon Dioxide

Carbon dioxide is a non-flammable, non-toxic, colorless, and odorless gas,

which is highly abundant and renewable at low costs [8, 9]. Carbon dioxide

exhibits its supercritical point at relatively mild conditions, one of the most

important advantages of using carbon dioxide in chemical reactions as

solvent. Figure 1-1 shows a schematic phase diagram of carbon dioxide.

The solid, liquid and gaseous states of CO2 are separated by the

corresponding phase equilibrium curves. The liquid-gas vapor pressure

curve, for example, starts at the triple point (p = 5.11 bar, T = 216.8°C) and

Introduction 3

ends at the critical point (p = 73.75 bar and T = 31.0°C) [10]. At higher

pressure and temperature, carbon dioxide is in its supercritical state, in

which the gas phase and the liquid phase are indistinguishable. Substances

in the supercritical state exhibit both gas and liquid-like properties [9]. The

supercritical fluid shows liquid-like densities suitable to dissolve

chemicals, and its diffusivities and viscosities resemble that of gases, with

the advantage of much better mass transport [11]. Near the critical point,

small changes of the parameters pressure and temperature cause drastic

variations of density, viscosity, and diffusivity, allowing the tuning of these

properties as desired by the chemical reaction at hand.

supercriticalregion

temperature

Figure 1-1. Schematic phase diagram of carbon dioxide. Phase equilibrium curves

define the regions in which carbon dioxide is in the gaseous, liquid, solid, and

supercritical state. Triple point (tp) at 5.11 bar, 216.8°C and critical point (cp) at

73.75 bar, 31.0°C.

4 Chapter 1

With these properties, supercritical fluids constitute an attractive replace¬

ment for conventional and maybe even toxic industrial solvents [12-14].

The solvent can be easily separated from the products after reaction by

decreasing the pressure, eliminating the need for expensive separation

procedures. The decaffeination of coffee with supercritical carbon dioxide,

the replacement of perchloroethylene, or the applications in semiconductor

processing [14-16] illustrate only some of the many industrial uses of

supercritical fluids [9, 17]. The use of supercritical fluids in industry is not

restricted to extractions, as indicated by a growing interest in the fields of

polymerization reactions [18], catalysis [6, 19], and material science [20].

1.3 Carbon Dioxide as a Ci-Building Block

1.3.1 Synthesis Using Carbon Dioxide

Carbon dioxide is not only an interesting solvent, it has also been used as a

reactant. One of the first, simple processes using CO2 as a reactant was the

Solvay process for the production of sodium carbonate, the basis of the

glass and ceramic industry. In this process, carbon dioxide and ammonia

are fed into a saturated solution of sodium chloride. The precipitated

sodium bicarbonate is calcinated and sodium carbonate is produced as

depicted in Scheme 1-1.

2 NaCI + 2 C02 + 2 NH3 + 2 H20 2 NaHCQ3 + 2 NH4CI

2NH4CI + CaO - 2 NH3 + CaCL, + H20

2 NaHCO^ Na2C03 + H20 + C02

Scheme 1-1. Solvay process for the production of sodium carbonate from carbon

dioxide, developed 1865 by Ernest Solvay.

Introduction 5

In organic chemistry, several reactions involve carbon dioxide as a

synthesis block, in which carbon dioxide replaces toxic chemicals, such as

phosgene or carbon monoxide. Scheme 1-2 shows an overview of some

synthetic routes starting from CO2, recently reviewed by Arakawa et al.

[21]. Valuable homogeneous catalyzed routes include the production of

carbonates [22, 23], carbamates [24, 25], urethanes [26-28], lactones [29-

31], pyrones [32, 33], formic acid and its derivatives [3, 7], and aldehydes

[34-36].

H

<OR

R

O R

A.o o

R-^R

f^' ^CT

Scheme 1-2. Possible chemical syntheses startingfrom carbon dioxide as a Ci-building

block [21].

An important application of the direct hydrogénation of carbon dioxide is

the synthesis of methanol, particularly in view of its possible future

6 Chapter 1

application in the area of hydrogen storage. Industrial production of

methanol is based on synthesis gas CO/H2, with either Zn/Q-203 acting as

catalyst in the high-pressure process (400°C / 200 bar), or Cu/Zn/Al acting

as catalyst in the ICI process (250°C / 50 bar) [37], as depicted in

Scheme 1-3. In both cases, carbon dioxide is added [38,39]. The direct

synthesis of methanol from CO2 instead of synthesis gas, as well as

efficient heterogeneous catalysts, such as Cu/ZrC>2, Cu/SiC>2, Cu/ZnO, and

Q1/AI2O3, has been reported as well (see [40-42]).

1) CO + 2H2 « CH3OH

2) C02 + 3H2 « CH3OH + H20

3) CO + H20 » C02 + H2

Scheme 1-3. Industrial synthesis of methanol 1) starting from synthesis gas or

2) directly from carbon dioxide. 3) CO2 can be producedfrom CO by way of the water

gas shift reaction.

The production of aspirin is an excellent example of an organic reaction

using CO2 as a building block. The synthesis of the intermediate salicylic

acid is achieved industrially by the Kolbe-Schmitt process, carboxylating

sodium phenolate with carbon dioxide [43], as illustrated in Scheme 1-4.

co, +

H+rrc^^XOONa ~-^ "COOH

Ur

Scheme 1-4. Production ofsalicylic acid by the "Kolbe-Schmitt" reaction.

Carbon dioxide is rather inert and difficult to activate, probably the main

reason why toxic carbon monoxide is still the favored Ci -synthesis block in

industrial processes [5,44,45]. However, in certain cases catalysts allow

Introduction 7

the use of carbon dioxide in the production of chemicals, such as urea and

its derivatives [45-47], organic carbonates [45,48-51], and salicylic acid

[45].

For the catalytic activation of carbon dioxide, the use of the supercritical

state of CO2 offers some advantages [5, 44, 52]. In addition to the liquid¬

like density, high diffusivity, and high mass transfer coefficients as

discussed in section 1.2, it has the desirable property that no additional

organic solvent is needed for the reaction. Two different strategies exist to

activate the relatively inactive carbon dioxide: 1) the use of highly reactive

reactants, e.g. epoxides forming organic carbonates, and 2) the use of

another "oxygen sink", e.g. water, since, from a thermodynamic point of

view, CO2 is an oxygen sink by itself. In the latter case, the reaction of

carbon dioxide with alcohols or amines in the presence of hydrogen leads

to formic acid derivatives, which is particularly interesting in connection

with solar or hydrothermal hydrogen production [53].

1.3.2 Formic Acid and its Derivatives

Formic acid is used in the silage of animal feed, for tanning and dyeing in

the textile industry, as a detergent, and as an intermediate in chemical

synthesis of pharmaceuticals, synthetic sweetening, and pesticide

production. The annual industrial production of 300 000 t of formic acid is

based on the hydrolysis of methyl formate that is formed from methanol

and carbon monoxide in the presence of sodium methoxide as catalyst. The

resulting by-product methanol is recycled.

An important alternative route to formic acid and its derivatives is the

transition-metal-catalyzed hydrogénation of carbon dioxide [54-60].

8 Chapter 1

Addition of a base stabilizes the metastable formic acid and lowers the

reaction enthalpy, thus thermodynamically favoring the products [4].

Ruthenium complexes with phosphine ligands show excellent activities in

the formation of formic acid, especially RuCl2(P(CH3)3)4 [61]. The

authors applied the system to the production of alkylformamides by adding

a primary or secondary amine like dimethylamine, diethylamine,

propylamine [61-63]. The proposed mechanism then proceeds via the

formation of the dialkylammonium dialkylcarbamate, as shown in

Scheme 1-5.

2 C02 + 2 H2 * 2 HCOOH

HNR2 + C02 .« R2NCOOH

R2NCOOH + HNR2 « [H2NR2][02CNR2]

2 HCOOH + [H2NR2][02CNRJ 2 R2NCHO + 2 H20 + C02

C02 + H2 + HNR2 R2NCHO + HzO

Scheme 1-5. Proposed reactions involved in the synthesis offormamides with amine,

hydrogen, and carbon dioxide.

Up to now, different homogeneous catalysts have been used in the

formylation of primary and secondary amines. High conversion with a

selectivity of 100% could be achieved with RuH2(P(CH3)3)4,

RuCl2(P(CH3)3)4, RuH2(P(C6H5)3)4, and RuCl(02CCH3)(P(CH3)3)4

[60,61,64], or with in situ formed complexes starting from

[RuCl2(C6H6)]2, RuCl2(DMSO)4, and [RuCl2(COD)]n [65]. Using

catalysts like RuCl2(PPh3)3, Ru3(CO)i2, and RuCl2(dppe)2 [62, 63], the

suitability of functionalized amines, such as pyrrolidine, piperidine,

piperazine, aniline, morpholine, and a-methylbenzylamine [66], for

carbamate

formation

Introduction 9

formylation reactions have been demonstrated. It could be further shown

that bidentate ruthenium phosphine catalysts of the type RUCI2L2 (L =

dppm, dppe, dppp) achieved the highest conversion and selectivity in the

formylation of amines with supercritical carbon dioxide and hydrogen [62,

66,67].

The development of heterogeneous catalysts for the formylation in

supercritical carbon dioxide is an attractive strategy with regard to

continuous processes. In particular, hybrid catalysts allow combining the

advantages of heterogeneous catalysts (stability, separation, and reuse) with

the excellent catalytic behavior of the homogeneous analogs.

Heterogeneous catalysts were investigated by Kröcher et al. [68-70],

incorporating Ru, Ir, Rh, Pt, and Pd complexes into silica gels. Hybrid

catalysts containing bidentate ruthenium phosphine complexes turned out

to give best performance [71], even higher than the most active

homogeneous catalysts previously known [60, 72].

1.3.3 Further Reactions Using Carbon Dioxide as Ci-Building Block

There exist several additional synthetic routes using carbon dioxide as

reactant, as summarized in Table 1-1. The synthesis of dimethyl carbonate

from CO2 and methanol, the synthesis of cyclic carbonates, and the

hydroformylation of alkenes are particularly interesting and will therefore

be discussed in more detail in this section.

Toxic and corrosive methylating or carbonylating agents like phosgene and

dimethyl sulfate may be replaced by the non-toxic dimethyl carbonate

which can be synthesized from carbon dioxide and methanol [73, 74], as

depicted in Scheme 1-6. Organotin compounds, Sn(IV) and Ti(IV)

alkoxides, and metal acetates were used as catalysts [75-77]. Note that in

10 Chapter 1

these cases the product yield is limited by the reaction equilibrium to less

than 5%. For high selectivity, weak Broensted acidity and base functions

are required, which may be satisfied, for example, by bifunctional catalysts

like ZrC>2, ZrC>2 in combination with H3PO4, CeC>2, and Ce02-Zr02-

Table 1-1. Heterogeneous catalyzed reactions using carbon dioxide both as Ci-building

block and as solvent in the supercriticalfluid region.

Product Reactant Catalyst Solvent p/bar T/°C Refs.

Dimethyl

carbonate

MethanolCe02-Zr02

Pd/ß-Ga203C02

60-210

10

70-170

50-450

[74]

[78]

Propylene

carbonate

Propylene

oxide

Mg-Al-0

different zeolities

co2/

DMF

50-140 150-200 [79-83]

Propylene

carbonate

Propylene

oxide

ZnBr2(Py)2

immobilizedC02 40-50 100-140 [84]

Cyclic

carbonateEpoxide ZnBr2(Poly-Py)2 C02 35 100 [85]

Styrene

carbonateEpoxide

Mg-based and

uncatalyzedC02 20-80 100-150 [79]

Dimethyl Epoxides

carbonate and KI/ZnO C02 165 150 [86]

and glycols methanol

Dimethyl

carbonatesGlycols

Mg-based

smectite catalystsC02 80 150

[87]

[88]

Aldehydes,

alcohols

Alkenes Ru complexesC02/co-

solvent

60-80 140-160 [89-91]

Aldehydes,

alcohols

Alkenes Rh complexesC02/

toluene

184 175 [92]

Introduction 11

2 CH3OH + C02 (CH30)2CO + H20

Scheme 1-6. Schematic illustration of the synthesis ofdimethyl carbonatesfrom carbon

dioxide and methanol.

Starting with carbon dioxide, cyclic carbonates can be synthesized from

epoxides [93], using Lewis acids, transition metal or organometallic

complexes, and polysiloxane-supported metal halides under high-pressure

conditions [94, 95], as illustrated in Scheme 1-7. High yields were obtained

with alkali metal salts in the presence of phase transfer agents such as

crown ethers or quarternary ammonium salts [96]. A new catalytic system,

based on organotin halides with quaternary ammonium or phosphonium

salts, was found by Baba et al. [97]. It delivers high yields under high-

pressure conditions (p = 50 bar, T = 40°C). Insoluble polystyrene beads

containing quaternary ammonium or phosphonium salts [98], polymer-

supported quaternary onium salts [98], immobilized cobalt complexes on

silica [99], and polymer-supported zinc complexes [85], act as

heterogenized catalytic systems, but show significant lower yields than the

homogeneous catalyzed reactions.

o

A

Scheme 1-7. Synthesis of propylene carbonate from propylene epoxide and carbon

dioxide.

The cycloaddition of CO2 to propylene oxide [84] may serve as an

example of the cyclic carbonate production. Its result is propylene

12 Chapter 1

carbonate, an intermediate for polycarbonates or other polymers, which is

also used as an aprotic solvent in the production of lithium batteries,

polyurethanes, and cosmetics [93].

Hydroformylation of alkenes is an important commercial route to

aldehydes and alcohols, the production currently exceeding six million tons

per year [100]. Classically, the synthesis is based on carbon monoxide and

alkenes as reactants, catalyzed by homogeneous transition metal

complexes. Tominaga et al. [89,90] found, that ruthenium carbonyl

clusters and alkaline metal halides catalyze the hydroformylation of alkenes

with carbon dioxide. In a multi-step process, the CO2 is probably first

reduced to CO and used in situ for the hydroformylation of alkenes with

catalysts such as H4Ru4(CO)i2, Ru3(CO)i2, and [Ru(CO)3Cl2]2 [91, 92].

The ratio of the products, aldehydes, alcohols, and alkanes, strongly

depends on the reaction conditions. Note that in all these reactions

homogeneous catalysts were applied predominantely.

1.4 Carbon Dioxide as a Solvent

The fact that carbon dioxide is rather inert is an advantage even when only

used as a solvent. In addition, the physical properties are quite favorable,

see chapter 1, section 1.2. The adjustability of the diffusion rate, mass and

heat transfer - in contrast to liquids - is desirable in heterogeneous cataly¬

sis, since mass transfer by diffusion often turns out to be a limiting step.

The tunability of the solvent properties by varying pressure and co-

solvents, and the elimination of gas/liquid and liquid/liquid mass transfer

resistances, are important features of the solvent. Heterogeneously

catalyzed reactions in supercritical fluids thus promise improvements in

terms of the reaction rate, control of the selectivity, and deactivation of the

Introduction 13

catalyst [101-104]. Compared to other supercritical fluids, such as

ammonia, ethane, ethene, methanol, and water, supercritical carbon dioxide

is often used as solvent, due to its moderate critical pressure of 73.8 bar, its

critical temperature of 31.1°C, its relatively high inertness, and its low

production costs.

Hydrogénation [105-111], partial oxidation [112-118], and isomerization

[119-122] reactions are some of the heterogeneously catalyzed reactions in

supercritical carbon dioxide that have received considerable attention.

Hydrogénation is typically a fast reaction that is diffusion-limited in

conventional solvents because of the low solubility of hydrogen in the

solvents and the mass transport limitations due to the gas/liquid mass

transfer. Interesting areas in the field of heterogeneously catalyzed

hydrogénations in supercritical carbon dioxide are the hydrogénation of

building blocks for fine chemicals and pharmaceuticals [123-127], as well

as the enantioselective hydrogénation on chirally modified surfaces, e.g.

the enantioselective hydrogénation of ethyl pyruvate on Pt/Al203 in the

presence of a cinchona alkaloid [128-130].

Because of the non-flammability, the miscibility, and the chemically inert

behavior, carbon dioxide commends itself as solvent for oxidation reactions

with molecular oxygen [103, 131, 132]. Apart from these properties, the

high diffusion constants of supercritical carbon dioxide were supposed to

improve the desorption of the partial oxidation products, which are strongly

adsorbed on the catalyst surface during reaction [133, 134]. Furthermore,

the higher heat capacity of supercritical carbon dioxide, as compared to

gases, and the higher mass transfer coefficient, as compared to liquid

solvents, is an advantage in these chiefly exothermic oxidations. In the field

of hydrocarbon activation, partial oxidation reactions were performed with

14 Chapter 1

propane [113], propylene [135], or in the epoxidation of olefins [136], for

example. In the area of partial oxidation of alcohols, noble metal catalyzed

oxidations of water-insoluble alcohols to carbonyl compounds [137, 138]

received particular interest.

Finally, preventing catalyst coking by extracting coke precursors in situ due

to its liquid-like densities and enhanced mass transfer properties favor

supercritical CO2 as solvent in isomerizations processes such as 1-hexene

or n-butane isomerizations [139, 140].

1.5 Scope of the Thesis

The aim of this doctoral thesis is to develop new and simpler catalytic

systems in the field of homogeneously and heterogeneously catalyzed

formylation reactions, combined with spectroscopic investigations under ex

situ and in situ conditions. The envisaged strategies are depicted in

Figure 1-2. The catalytic activity of the homogeneous catalysts

RuCl2(dppe)2 and RuC^(dppp)2 (route 1) are well known from earlier

work and are used in this work for comparison. To simplify the synthesis of

active catalysts, the in situ formation of catalysts (route 2) is evaluated. For

this purpose, in situ generated catalysts starting from a simple ruthenium

salt, RUCI3, in the presence of phosphine ligands (PPI13 and dppe) are

tested. Additionally, the in situ modified solid RU/AI2O3 catalyst with dppe

is viewed as a promising approach. On the way towards structurally stable

heterogeneous catalysts, two groups of heterogenized catalysts attract

special interest: adsorbed catalysts (RuCl2(dppe)2/Si02 and "Si"-NH2-

RUCI3 + dppe) in route 3, and hybrid-gel catalysts with covalently

anchored Ru-complexes (RuCl2(bspe)2/Si02 and RuCl2(bspp)2/Si02) in

Introduction 15

route 4. Beside their catalytic performance, the structural stability is of

particular interest.

Homogeneous route

0

In situ route

©

Adsorption route

©

RuCI Ru/Al,0

Ph. Ph.

.P. 2CI PJ I

OK)p p

Ph, Ph,

n+cr

- dppe

cPh, Ph,P. CI Pj

Ph. Ph.

^Pv CI P^\l/ ^1 <\!/>Ru > Ru </IS/1 </i\>P CI P P CI P

Ph, Ph, Ph. Ph.

H„ CO dppe H,, C02

RuCI2(dppe)2 Ru(dppe)2X2

dppe

O I O-Si-0

.s; °s /b\

i /°

o-

O ! O-Si-0' ° /b\

/ ô °--Si-0

\

/°Immobilization route

\/—Si-

JSi-O.

'o-

\l

0'S|-Si-O.

\

Ov I

.Sk0 o

1 I

Sl Si!o

©

.OH HO^ l

Ph PIV\ .Si

;r, ci p ^-^

\

"OI .ov ,

'Si Si.I l

Ox ,0Si

C X )^p ci p^^Ph Pfî"

I „O-Si—.Si, I

0 0 Ll „O-SiSu \

so' "o

ill,

\/ ^o-—Si-O, I

I .sks.

I 0 0

Si-O^ I I „OH

/ .Su „Si.

.SiHO l

O.

-Sl-n'l

/Oo

o' "9„O^ I „O-Si—

? /

oV

o'

-s,."o"

— Si-

-sr

\

o

Si Sll l OH

\ Os „O\ Si

o

HO^ l

Ph PIV\ .Si

P. CI P ^~^

\YRu

/ I \P CI P.Ph Ph" .Si

HO l

O.

I ,0-Si—•Sk I

0 I.

1 .O-Si

..Sk \

o ox

.s',CO' -o

-Si-

/ °-0

,0V I ,0-Si-a /

! %0-S|-^

Figure 1-2. Classification of the catalysts synthesized and investigated in this work.

1) the homogeneous catalysts RuCÏ2(dppe)2 and RuCÏ2(dppp)2, 2) the in situ formed

catalysts from RuCl$ and Ru/A^O^ with dppe, 3) the adsorbed catalysts

RuCl2(dppe)2/Si02 and RuCl^/Si02 with dppe, 4) the immobilized catalysts

RuCl2(bspe)2/Si02andRuCl2(bspp)2/Si02.

In order to gain more insight into the structure of the catalysts under

reaction conditions, and in order to determine whether the reactions are

homogeneously or heterogeneously catalyzed, the catalysts are investigated

in situ and ex situ using XANES and EXAFS. For this purpose, suitable

spectroscopic cells are required.

16 Chapter 1

Finally, the extension of possible substrates as well as the reaction concept

are investigated. The formylation of different amines with additional

functional groups is treated as an important extension to the commonly

investigated scope of substrates.

In addition, the synthesis of vinyl carbamate from phenylacetylene, amines,

and carbon dioxide is studied using different homogeneous ruthenium

catalysts. As before, the aim is to use CO2 both as reactant and as solvent.

1.6 References

[I] A. Behr, Carbon Dioxide Activation by Metal Complexes, VCH,

Weinheim, New York (1988).

[2] A. Behr, Angew. Chem. Int. Ed. Engl. 27 (1988) 661.

[3] W. Leitner, Angew. Chem. Int. Ed. Engl. 34 (1995) 2207.

[4] P. G. Jessop, T. Ikariya, R. Noyori, Chem. Rev. 95 (1995) 259.

[5] A. Baiker, Appl. Organomet. Chem. 14 (2000) 751.

[6] A. Baiker, Chem. Rev. 99 (1999) 453.

[7] P. G. Jessop, T. Ikariya, R. Noyori, Chem. Rev. 99 (1999) 475.

[8] P. G. Jessop, W. Leitner, Chemical Synthesis Using Supercritical

Fluids, Wiley-VCH, Weinheim (1999).

[9] L. T. Taylor, Supercritical Fluid extraction (Techniques in

Analytical Chemistry Series), Wiley, New York (1996).

[10] T. Andrews, Phil. Trans. 159 (1869) 575.

[II] C. A. Eckert, B. L. Knutson, P. G. Debenedetti, Nature 383 (1996)

313.

Introduction 17

12

13

14

15

16

17

18

19

20

J. M. DeSimone, Z. Guan, C. S. Elsbernd, Science 257 (1992) 945.

J. M. DeSimone, E. E. Maury, Y. Z. Menceloglu, J. B. McClain, J.

Romack, J. R. Combes, Science 265 (1994) 356.

D. Bradley, New Scientist 6 (1994) 32.

K. Zosel, Angew. Chem. 90 (1978) 702.

J. W. King, L. L. Williams, Curr. Opin. Sol. State Mat. Sei. 7 (2003)

413.

M. McHugh, V. J. Krukonis, Supercritical Fluid Extraction:

Principles and Practice, Butterworth-Heinemann, Boston (1994).

J. L. Kendall, D. A. Canelas, J. L. Young, J. M. DeSimone, Chem.

Rev. 99 (1999) 543.

J.-D. Grunwaldt, R. Wandeler, A. Baiker, Catal. Rev. - Sei. Eng. 45

(2003)1.

Y.-P. Sun, Supercritical Fluid Technology in Materials Science and

Engineering, Marcel Dekker, New York (2002).

211 H. Arakawa, M. Aresta, J. N. Armor, M. A. Barteau, E. J. Beckman,

A. T. Bell, J. E. Bercaw, C. Creutz, E. Dinjus, D. A. Dixon, K.

Domen, D. L. DuBois, J. Eckert, E. Fujita, D. H. Gibson, W. A.

Goddard, D. W. Goodman, J. Keller, G. J. Kubas, H. H. Kung, J. E.

Lyons, L. E. Manzer, T. J. Marks, K. Morokuma, K. M. Nicholas, R.

Periana, L. Que, J. R. Rostrupp-Nielsen, W. M. H. Sachtler, L. D.

Schmidt, A. Sen, G. A. Somorjai, P. C. Stair, B. R. Stults, W.

Tumas, Chem. Rev. 101 (2001) 953.

[22] M. Aresta, E. Quaranta, Chemtech (1997) 32.

18 Chapter 1

[23

[24

[25

[26

[27

[28

[29

[30

[31

[32

[33

[34

[35

[36

[37

M. Aresta, A. Dibenedetto, J. Mol. Catal. A 182 (2002) 399.

K. Takeshita, A. Kitamoto, J. Chem. Eng. Jpn. 21 (1988) 411.

C. Bruneau, P. H. Dixneuf, Carbon Dioxide Fixation and Reduction

in Biological and Model Systems, C.-I. Bränden, G. Schneider,

University Press, Oxford (1994) 131.

W. D. McGee, D. P. Riley, Organomet. 11 (1992) 900.

W. D. McGee, D. P. Riley, M. E. Christ, K. M. Christ, Organomet.

12(1993)1429.

W. D. McGee, Y. Pan, D. P. Riley, Chem. Commun. (1994) 699.

A. Behr, K.-D. Juszak, W. Keim, Synthesis (1983) 574.

A. Behr, K.-D. Juszak, J. Organomet. Chem. 255 (1983) 263.

A. Behr, R. He, K.-D. Juszak, C. Krüger, Y.-H. Tsai, Chem. Ber. 119

(1983)991.

M. T. Reetz, W. Konen, T. Strack, Chimia 47 (1993) 493.

K. Buchmuller, N. Dahmen, E. Dinjus, D. Neumann, B. Powietzka,

S. Pitter, J. Schon, Green Chem. 5 (2003) 218.

K.-I. Tominaga, Y. Sasaki, J. Mol. Catal. A 220 (2004) 159.

S. Jääskeläinen, M. Haukka, Appl. Catal. A 247 (2003) 95.

K.-I. Tominaga, Y. Sasaki, Chem. Lett. 33 (2004) 14.

H. Beyer, W. Walter, Lehrbuch der organischen Chemie, S. Hirzel

Verlag, Stuttgart (1991).

[38] G. C. Chinchen, P. J. Denny, D. G. Parker, M. S. Spencer, D. A.

Whan, Appl. Catal. 30 (1987) 333.

[39

[40

[41

[42

[43

[44

[45

[46

[47

[48

[49

[50

[51

Introduction 19

K. C. Waugh, Catal. Today 15 (1992) 51.

C. Schild, A. Wokaun, R. A. Koppel, A. Baiker, J. Catal. 130 (1991)

657.

I. A. Fisher, H. C. Woo, A. T. Bell, Catal. Lett. 44 (1997) 11.

J. Wambach, A. Baiker, A. Wokaun, Phys. Chem. Chem. Phys. 1

(1999)5071.

A. S. Lindsey, H. Jeskey, Chem. Rev. 57 (1957) 583.

J. M. DeSimone, W. Tumas, Green Chemistry Using Liquid and

Supercritical Carbon Dioxide, University Press, Oxford (2003).

K. Weissermehl, H.-J. Arpe, Industrial Organic Chemistry, Wiley-

VCH, Weinheim (2003).

R. Nomura, Y. Hasegawa, M. Ishimoto, T. Toyosaki, H. Matsuda, J.

Org. Chem. 57 (1992) 7339.

J. Fournier, C. Bruneau, P. H. Dixneuf, S. Lécolier, J. Org. Chem. 56

(1991)4456.

H. Kisch, R. Millini, I.-J. Wang, Chem. Ber. 119 (1986) 1090.

W. Dümler, H. Kisch, Chem. Ber. 123 (1990) 277.

W. D. McGee, D. P. Riley, J. Org. Chem. 60 (1995) 6205.

J.-L. Dubois, K. Sayama, H. Arakawa, Chem. Lett. (1992) 1115.

[52] M. Halmann, Chemical Fixation of Carbon Dioxide: Methods for

Recycling CO2 into Useful Products, CRC Press, Boca Raton,

Florida (1993).

[53] J. R. Rostrupp-Nielsen, Appl. Catal. A 255 (2003) 3.

20 Chapter 1

[54] Y. Inoue, H. Izumida, Y. Sasaki, H. Hashimoto, Chem. Lett. (1976)

863.

[55] M. M. T. Khan, S. B. Halligudi, N. N. Rao, J. Mol. Catal. 51 (1989)

161.

[56] J.-C. Tsai, K. M. Nicholas, J. Am. Chem. Soc. 114 (1992) 5117.

[57] E. Graf, W. Leitner, J. Chem. Soc, Chem. Commun. (1992) 623.

[58] W. Leitner, E. Dinjus, F. Gassner, J. Organomet. Chem. 475 (1994)

257.

[59] F. Gassner, W. Leitner, J. Chem. Soc, Chem. Commun. (1993)

1465.

[60] P. G. Jessop, T. Ikariya, R. Noyori, Nature 368 (1994) 231.

[61] P. G. Jessop, Y. Hsiao, T. Ikariya, R. Noyori, J. Am. Chem. Soc. 118

(1996)344.

[62] O. Kröcher, R. A. Koppel, A. Baiker, Chem. Commun. 5 (1997) 453.

[63] S. Schreiner, J. Y. Yu, L. Vaska, Inorg. Chim. Acta 147 (1988) 139.

[64] C. A. Thomas, R. J. Bonilla, Y. Huang, P. G. Jessop, Can. J. Chem.

79(2001)719.

[65] C.-C. Tai, J. Pitts, J. C. Linehan, A. D. Main, P. Munshi, P. G.

Jessop, Inorg. Chem. 41 (2002) 1606.

[66] L. Schmid, A. Canonica, A. Baiker, Appl. Catal. A: Gen. 255 (2003)

23.

[67] L. Schmid, M. S. Schneider, D. Engel, A. Baiker, Catal. Lett. 88

(2003) 105.

[68

[69

[70

[71

[72

[73

[74

[75

[76

[77

Introduction 21

O. Kröcher, R. A. Koppel, A. Baiker, J. Chem. Soc, Chem.

Commun. (1996) 1497.

O. Kröcher, R. A. Koppel, A. Baiker, Process Technology

Proceedings, Ph. Rudolf von Rohr, Ch. Trepp, Amsterdam 12 (1996)

91.

O. Kröcher, R. A. Koppel, M. Fröba, A. Baiker, J. Catal. 178 (1998)

284.

L. Schmid, O. Kröcher, R. A. Koppel, A. Baiker, Micropor.

Mesopor. Mater. 35-36 (2000) 181.

P. G. Jessop, T. Ikariya, R. Noyori, Science 269 (1995) 1065.

W. J. Peppel, Ind. Eng. Chem. 50 (1958) 767.

K. Tomishige, Y. Furusawa, Y. Ikeda, M. Asadullah, K. Fujimoto,

Catal. Lett. 76(2001)71.

S. Sakai, T. Fujinami, T. Yamada, S. Furusawa, Nippon Kagaku

Kaishi 10 (1975) 1789.

J. Kizlink, Collect. Czech. Chem. Commun. 58 (1993) 1399.

J. Kizlink, I. Pastucha, Collect. Czech. Chem. Commun. 60 (1995)

687.

[78] S. E. Collins, A. Baltanas, A. L. Bonivardi, J. Catal. 226 (2004) 410.

[79] H. Kawanami, Y. Ikushima, Chem. Commun. (2000) 2089.

[80] K. Yamaguchi, K. Ebitani, T. Yoshida, H. Yoshida, K. Kaneda, J.

Am. Chem. Soc. 121 (1999) 4526.

[81] S. Fujita, B. M. Bhanage, Y. Ikushima, M. Shirai, K. Toriib, M.

Arai, Catal. Lett. 79 (2002) 95.

22 Chapter 1

[82] H. Yasuda, L.-N. He, T. Sakakura, J. Catal. 209 (2002) 547.

[83] M. Tu, R. J. Davis, J. Catal. 199 (2001) 85.

[84] M. Ramin, J.-D. Grunwaldt, A. Baiker, J. Catal. 234 (2005) 256.

[85] H. S. Kim, J. J. Kim, H. N. Kwon, M. J. Chung, B. G. Lee, H. G.

Jang, J. Catal. 205 (2002) 226.

[86] Y. H. Chang, T. Jiang, B. X. Han, Z. M. Liu, W. Z. Wu, L. Gao, J.

C. Li, H. X. Gao, G. Y. Zhao, J. Huang, Appl. Catal. A 263 (2004)

179.

[87] B. M. Bhanage, S. Fujita, Y. Ikushima, M. Arai, Appl. Catal. A 219

(2001)259.

[88] B. M. Bhanage, S. Fujita, Y. Ikushima, K. Toriib, M. Arai, Green

Chem. 5(2003)71.

[89] K.-I. Tominaga, Y. Sasaki, Catal. Commun. 1 (2000) 1.

[90] K.-I. Tominaga, Y. Sasaki, J. Mol. Catal. A: Chem. 220 (2004) 159.

[91] S. Jääskeläinen, M. Haukka, Appl. Catal. A: Gen. 247 (2003) 95.

[92] O. Hemminger, A. Marteel, M. R. Mason, J. A. Davies, A. R. Tadd,

M. A. Abraham, Green Chem. 4 (2002) 507.

[93] A.-A. G. Shaikh, S. Sivaram, Chem. Rev. 96 (1996) 951.

[94] R. J. De Pasquale, J. Chem. Soc, Chem. Commun. (1973) 157.

[95] R. Nomura, A. Ninagawa, H. Matsuda, J. Org. Chem. 45 (1980)

3735.

[96] G. Rokicki, W. Kuran, Bull. Chem. Soc. Jpn. 57 (1984) 1662.

Introduction 23

97] A. Baba, T. Nozaki, H. Matsuda, Bull. Chem. Soc. Jpn. 60 (1987)

1552.

98] T. Nishikubo, A. Kameyama, J. Yamashita, M. Tomoi, W. Fukuda,

J. Polym. Sei. Part A 31 (1993) 939.

99] X.-B. Lu, J.-H. Xiu, R. He, K. Jin, L.-M. Luo, X.-J. Feng, Appl.

Catal. A 227 (20047) 537.

100] C. Botteghi, M. Marchetti, S. Paganelli, Transition Metals for

Organic Synthesis, M. Beller,C. Bolm, VCH, Weinheim 1 (1998) 23.

101] B. Subramaniam, M. A. McHugh, Ind. Eng. Chem. Des. Dev. 25

(1986)1.

102] H. Tiltscher, H. Hofmann, Chem Eng. Sei. 42 (1987) 959.

103] G. Musie, M. Wei, B. Subramaniam, D. H. Busch, Coord. Chem.

Rev. 219 (2001) 789.

104] J. R. Hyde, P. Licence, D. Carter, M. Poliakoff, Appl. Catal. A 222

(2001)119.

105] V. Arunajatesan, B. Subramaniam, K. W. Hutchenson, F. E. Herkes,

Chem. Eng. Sei. 56 (2001) 1363.

106] R. Tschan, M. M. Schubert, A. Baiker, W. Bonrath, H. Lansink-

Rotgerink, Catal. Lett. 75 (2001) 31.

107] B. M. Bhanage, Y. Ikushima, M. Shirai, M. Arai, Catal. Lett. 62

(1999) 175.

108] M. Chatterjee, F. Y. Zhao, Y. Ikushima, Appl. Catal. A 262 (2004)

93.

24 Chapter 1

[109] F. Y. Zhao, R. Zhang, M. Chatterjee, Y. Ikushima, M. Arai, Adv.

Synth. Catal. 346(2004)661.

[110] M. B. O. Andersson, J. W. King, L. G. Blomberg, Green Chem. 2

(2000) 230.

[111] M.-B. Macher, J. Högberg, P. Moller, M. Härröd, Fett-Lipid 101

(1999)301.

[112] A. Martin, B. Kerler, Chem. Ing. Technik 72 (2000) 382.

[113] A. Martin, B. Kerler, Chem. Eng. Technol. 24 (2001) 41.

[114] U. Armbruster, A. Martin, Q. Smejkal, H. Kosslick, Appl. Catal. A

265 (2004) 237.

[115] J.-D. Grunwaldt, M. Caravati, M. Ramin, A. Baiker, Catal. Lett. 90

(2003)221.

[116] M. Caravati, J.-D. Grunwaldt, A. Baiker, Catal. Today 91-92 (2004)

1.

[117] M. Caravati, J.-D. Grunwaldt, A. Baiker, Phys. Chem. Chem. Phys.

7 (2005) 278.

[118] R. Ciriminna, S. Campestrini, M. Pagliaro, Adv. Synth. Catal. 346

(2004)231.

[119] B. Sander, M. Thelen, B. Kraushaar-Czarnetzki, Ind. Eng. Chem.

Res. 40 (2001) 2767.

[120] V. I. Bogdan, T. A. Klimenko, L. M. Kustov, V. B. Kazansky, Appl.

Catal. A 267 (2004) 175.

[121] M. G. Hitzler, F. R. Smail, S. K. Ross, M. Poliakoff, Chem.

Commun. (1998) 359.

Introduction 25

122] M. C. Clark, B. Subramaniam, Ind. Eng. Chem. Res. 37 (1998) 1243.

123] M. G. Hitzler, F. R. Smail, S. K. Ross, M. Poliakoff, Org. Proc Res.

Developm. 2 (1998) 137.

124] L. Devetta, P. Canu, A. Bertucco, K. Steiner, Chem. Eng. Sei. 52

(1997)4163.

125] L. Devetta, A. Giovanzana, P. Canu, A. Bertucco, B. J. Minder,

Catal. Today 48 (1999) 337.

1261 E. P. Martins, D. A. G. Aranda, F. L. P. Pessoa, J. L. Zotin, Braz. J.

Chem. Eng. 17(2000)361.

127] F. Zhao, Y. Ikushima, M. Chatterjee, M. Shirai, M. Arai, Green

Chem. 5(2003)114.

1281 B. Minder, T. Mallat, K. H. Pickel, K. Steiner, A. Baiker, Catal. Lett.

34(1995)1.

129] R. Wandeler, N. Künzle, M. S. Schneider, T. Mallat, A. Baiker,

Chem. Commun. (2001) 673.

130] T. Bürgi, A. Baiker, Ace Chem. Res. 37 (2004) 909.

131] E. J. Beckman, Environ. Sei. Technol. 37 (2003) 5289.

132] E. J. Beckmann, J. Supercrit. Fluids 28 (2004) 121.

1331 M. Ghezai, M. Wei, B. Subramaniam, D. H. Busch, Coord. Chem.

Rev. 219-221 (2001) 789.

134] B. R. Müller, A. Martin, B. Lücke, J. Supercrit. Fluids 23 (2002)

243.

135] G. Jenzer, T. Mallat, M. Maciejewski, F. Eigenmann, A. Baiker,

Appl. Catal. A 208 (2001) 125.

26 Chapter 1

[136] F. Loeker, W. Leitner, Chem. Eur. J. 6 (2000) 2011.

[137] G. Jenzer, M. S. Schneider, R. Wandeler, T. Mallat, A. Baiker, J.

Catal. 199(2001)141.

[138] A. M. Steele, J. Zhu, S. C. Tsang, Catal. Lett. 73 (2001) 9.

[139] B. Subramaniam, Appl. Catal. A 212 (2001) 199.

[140] V. I. Bogdan, T. A. Klimenko, L. M. Kustov, V. B. Kazansky, Appl.

Catal. A 267 (2004) 175.

Chapter

Experimental

In this chapter, general information about the experimental procedures is

given, with the focus set on the description of the basic experimental setups

and catalytic sequences. In addition, some general features of the applied

methods and techniques are described. Detailed information on specific

experimental procedures can be found in the corresponding chapters. Note

that the experiments described in this work involve the use of high pressure

and require equipment with the appropriate pressure rating.

2.1 Experimental Equipment

2.1.1 High-pressure Batch Reactor Cell

All the catalytic reactions in this work were carried out in a high-pressure

autoclave of the type Medimex No. 128 as depicted in Figure 2-1. The

stainless steel batch reactor with metal sealing features a reaction volume

of 500 ml and can be operated at temperatures up to 400°C and at pressures

up to 700 bar. The reactor is equipped with a magnetic stirrer, using a

conventional 6-blade turbine (Medimex type SR) with manual adjustment

of the stirring frequency. Heating is provided by a copper jacket with

electric heating elements, and cooling by water running through an outer

aluminum jacket. Both the batch sheath temperature and the core

2

28 Chapter 2

temperature are controlled by a PID controller operated in the cascade

mode. A piezoelectric pressure gauge records the reaction pressure. In

order to provide a means to depressurize after reaction, a purge valve was

installed, which in turn was connected to the purge lines of the high-

pressure laboratory.

Icomputer ,

'

t

control, t

moniîoraig

Figure 2-1. Schematic illustration of the setup used for the catalytic experiments.

1) autoclave, 2) cooling water, 3) rupture disk, 4) carbon dioxide cylinder with dip tube,

5) compressor with 6) pump and 7) cooling unit, 8) compressed air for compressor

control (3-5 bar), 9) massflow controller, and 10) hydrogen cylinder.

Hydrogen was directly transferred from a gas cylinder to the autoclave and

quantified using a pressure gauge. Liquid carbon dioxide, stored in a gas

cylinder with dip tube, was compressed to the reaction pressure with a fluid

compressor (New Ways of Analysis, PM-101), consisting of a pump unit

and a cooling unit to stabilize the temperature of the compressed medium

below its boiling point. The amount of carbon dioxide filled into the reactor

was quantified with a mass flow controller (Rheonik RHM 01) measuring

Experimental 29

the Coriolis force. Catalyst and reactant were weighed gravimetrically and

poured into the reactor before it was closed and flushed with hydrogen.

2.1.2 Apparatus for Synthesis of Catalysts in the Presence of UV-Light

The catalyst synthesis described in chapter 7 required a specially designed

vessel to allow the irradiation of the reactants with ultraviolet light

(Figure 2-2). For this purpose, an ultraviolet lamp made by Osram

(Ultramed FDA R7s, 400 W) was placed in a UV-light permeable double

wall quartz flask. The ultraviolet lamp was cooled directly by pressurized

air as well as by cooling water running between the two walls. This quartz

flask was enclosed by a second flask with an inlet and an outlet to allow

flushing with argon during the air-sensitive reaction. The whole UV-light

reactor was operated in a metallic box designed to protect the eyes against

ultraviolet light.

Electrical power and

cooling air for UV lamp fIIssssssssssœssssssssJiylL

Figure 2-2. Schematic illustration of the apparatusfor ultraviolet irradiation (left) and

a photograph of the complete apparatus (right) with the protecting box behind the

power unit and the temperature control unit.

30 Chapter 2

Prior to the experiment, the quartz flask was kept in a drying oven for 24 h.

The flasks were installed in the protecting metal box and were evacuated

and recharged with argon three times. Syringes were used to fill the

reactants in the reactor. During the exposure time, the reaction mixture was

flushed permanently by a small argon stream while stirring the reaction

mixture slowly. After irradiation, the product mixture was removed with a

syringe and stored in argon in a Schlenk flask for further processing.

2.1.3 View-cell for Phase Behavior Studies

Phase behavior studies of the reaction mixture were performed in a

computer-controlled high-pressure view-cell of variable volume (23 -

63 ml) from SITEC, in combination with online digital video imaging and

recording [1]. The magnetically stirred view-cell consisted of a horizontal

stainless steel cylinder equipped with a sapphire window.

Digital video imaging allows observations of even minor volumes of

gaseous and liquid phases, facilitating reliable static measurements of

phase behavior at temperatures and pressures up to 200°C and 200 bar.

Typically, the reactants were first filled into the view-cell, then hydrogen

and carbon dioxide were added. The temperature was adjusted by a

thermostated oil bath, and the pressure by changing of the volume [1].

2.2 Characterization Methods

To gain information on the structure of the catalysts, they were investigated

by !H-, 29Si-, 31P-cross-polarization magic angle spinning (CP-MAS)

NMR, nitrogen physisorption (BET), elemental analysis (EA), X-ray

photoelectron spectroscopy (XPS), attenuated total reflection infrared

spectroscopy (ATR-IR), diffuse reflectance infrared Fourier transform

Experimental 31

spectroscopy (DRIFTS), and X-ray absorption spectroscopy (XANES and

EXAFS). The metal content of the catalysts and liquid reaction mixtures

were measured by inductively coupled plasma optical emission

spectroscopy (ICP-OES). The structural and analytical methods are well

described in the pertinent literature [2-14] and the reader is referred to these

sources for further information. Since XANES and EXAFS were used in a

number of experiments in this work, they are described in chapter 3. The

following paragraphs describe special experimental setups and

spectroscopic cells, which were specifically developed for this work.

2.2.1 X-ray Absorption Spectroscopy Cell for Catalyst Studies in

Liquid Phase

To identify the structure of homogeneous catalysts formed under reaction

conditions, the liquid reaction mixtures were investigated by XANES and

EXAFS after reaction. For this purpose, a specially designed stainless steel

XAS cell was used for transmission experiments (Figure 2-3). Because of

the low ruthenium concentration in the reaction mixtures in the range of

50-100 ppm, a long path length of 4 cm was required, using windows

with a 6 mm x 11 mm cross section. On both sides of the cell,

exchangeable Kapton windows were employed.

The cell can be filled from the top and has a volume of 2 ml. A 1 mm x

10 mm X-ray beam was aligned with the center of the spectroscopic cell

using an x, y, 9-table from Newport.

In addition, an EXAFS cell suitable for both transmission and fluorescence

mode was constructed with similar technical dimensions as those chosen

for the XAS cell for liquid samples (Figure 2-4). The cell was equipped

32 Chapter 2

with a 10 mm x 20 mm large window positioned at a 90° angle with respect

to the X-ray beam.

window equippedwith Kapton

Figure 2-3. Stainless steel XAS cell for structural identification of the catalytically

active speciesformed in the liquidphase during reaction.

This allowed recording EXAFS spectra in the fluorescence mode at a 90°

angle to the beam. Fluorescence spectra at the Ru K-edge were recorded

using a 13-element Ge fluorescence detector (Canberra) at the Swiss

Norwegian Beamline in Grenoble (for further details, cf. chapter 7,

section 7.3.5).

Kapton

Figure 2-4. Stainless steel XAS cell for fluorescence and transmission mode for

structural identification of the catalytically active species formed in the liquid phase

during reaction.

Experimental 33

2.2.2 High-pressure XAS Cell for Studying Simultaneously the Liquid

Phase and the Solid/Liquid Interface

While the spectroscopic cells in section 2.2.1 allowed recording spectra on

homogeneous catalytic species after reaction, in many cases, in situ

monitoring of structural changes of a catalyst are required that occur at the

solid/liquid interface and in the bulk fluid phase. This approach is

advantageous if homogeneous catalytic species formed from a

heterogeneous catalyst and some ligands are dissolved in the reaction

mixture. Therefore, a spectroscopic batch reactor cell was designed that

allows in situ monitoring of the changes occurring in the bulk liquid as well

as at the liquid/solid interface of heterogeneous reactions at elevated

pressure and temperature. Hence, the dimensions of the high-pressure cell

were chosen to monitor the liquid/solid interface at the bottom of the cell

with a path length of 4 mm, and the liquid part with a path length of 15 mm

at a height of 10 mm above the bottom of the reactor cell (Figure 2-5). The

total volume of the batch reactor is 10 ml.

The inner part of the cell consists of a PEEK (polyetheretherketone, density

1.3 g/cm3) container, which has also high chemical resistance against acids,

organic solvents, and alkaline media. As Figures 2-5 and 2-6 show, this

container is embedded in a stainless steel cell. Apart from its corrosion-

resistance, PEEK was chosen for its high transparency for X-rays above

9 keV, which is better than many other chemically resistant thermoplastics

such as teflon. At 250°C, the operating temperature of this material is

rather high for a thermoplastic. As depicted in Figures 2-5 and 2-6, four

windows of 5 mm x 1 mm size inside the stainless steel container serve to

let the X-ray beam pass through the batch reactor cell at two positions (see

X-ray paths in Figure 2-6). In order to prevent damages of the PEEK

34 Chapter 2

container at high pressure, 0.5 mm thick Be disks (7.8 mm diameter,

Brushwellman) were placed between the PEEK container and the windows

in the stainless steel container. Note that the PEEK polymer itself has a

high tensile strength and therefore Be disks are not necessary at

temperatures up to 150°C.

Gas inlets, thermocouple

pressure transducer

Top cover with

PEEK inset

"eflon O-ring

EEK inset

Stainless steet bodywith cartridgee windowsX-ray beam at

two positions

Magnetic stirrer

with motor

Figure 2-5. A schematic representation of the apparatus designed for in situ X-ray

absorption spectroscopy studies with two pathways to monitor the soliddiquid interface

ofheterogeneous catalysts and the liquidphase or soliddiquid reactions.

The temperature of the batch reactor cell is adjusted using two 160 Watt

cartridge heaters (SUVAG) inserted at both sides of the stainless steel cell.

Temperature control is carried out by a commercial KS 20-1 controller

(PMA, Omni Ray) by help of a NiCr/Ni thermocouple inside the stainless

mantle. The reaction temperature is measured within the reactor. Stirring is

provided by a magnetic stirrer. In the interest of a mechanically compact

design, the thermally insulated motor, together with the magnet, was

Experimental 35

installed below the batch reactor cell. Finally, a burst plate was installed for

safety reasons.

Figure 2-6. Three-dimensional views ofthe spectroscopic batch reactor cell with PEEK

container and magnetic stirrer.

Another crucial point is the alignment of the cell in the X-ray beam. For

this purpose, the whole cell is mounted on a x, z, 9-table as shown in

Figure 2-7. This is very similar to the principle we used for a high-pressure

continuous flow cell [15]. The apparatus can be moved between the two

positions using the z-translation. Another important part of the sample cell

is the sealing. This is achieved using an appropriate cover and a

polytetrafluoroethylene (PTFE) sealing. Again, PEEK is used inside the

cover, but two different designs were brought into action, depending on the

experiment to be performed. Note that the whole setup was placed inside a

safety box (Figure 2-7).

36 Chapter 2

Figure 2-7. View ofthe batch reactor cell in the experimental hall ofthe beamline XI at

HASYLAB (DESY, Hamburg), details see text.

The cover shown in Figures 2-5 - 2-7 was adapted for reactions with

supercritical carbon dioxide and gases being present in the reaction. It has

an inlet and an outlet, both equipped with a valve, a burst plate (Swagelok,

190 bar), a pressure transducer (Wika), and the appropriate gas or liquid

supply system. For gases, the appropriate gas mixture was filled into the

batch reactor cell to the desired gas pressure. Liquid carbon dioxide was

dosed using a C02-compressor unit (NWA PM-101) with a Rheonik flow

controller (RHMO15).

Experimental 37

2.3 Catalytic Tests

2.3.1 Experimental Procedure

The amount of the amine and the catalyst were weighed and added to the

batch reactor before closing and flushing three times with hydrogen. Then,

60 bar of hydrogen was added and heated to 100°C while stirring at

300 min-1. As soon as the temperature was stabilized, the hydrogen

pressure was adjusted to the desired value of 80 bar, if not specified

differently. The reaction was started by adding liquid carbon dioxide,

quantified by a mass flow controller. To stop the reaction, the autoclave

was cooled to room temperature and vented slowly. Note that special care

had to be taken during cleaning. The reactor cleaning procedure involved

two washing steps, one with distilled water and one with acetone, followed

by drying completely.

2.3.2 Analysis

The product mixture at the bottom of the reactor was analyzed by a gas

Chromatograph (HP-6890) equipped with a HP-5 capillary column (30 m x

0.32 mm x 0.25 jiim) and a flame ionization detector (FID). Three drops of

the reaction mixture were dissolved in ethanol, and 1.0 jlxI of the resulting

solution was injected in the GC inlet at 190°C with a split ratio of 20:1. For

3-methoxypropylamine, the oven temperature was held at 35°C for 5 min,

heated to 200°C at a rate of 10°C/min, and held constant for another 5 min.

The final temperature of 250°C was reached at a rate of 75°C/min, with an

additional holding time of 8 min. The FID was set to 250°C. For

quantitative analysis, a calibration curve was recorded to calculate the

product/reactant ratio from the measured peak area. This analysis

38 Chapter 2

procedure was optimal for the reaction mixture of 3-methoxypropylamine.

It was modified slightly in the case of other amines.

Formamide products were identified with GC-MS, lH- and 13C-nuclear

magnetic resonance (NMR).

2.3.3 Evaluation of Catalytic Tests

The conversion X in the formamide synthesis was calculated as the ratio of

the amount of consumed reactant and the amount of amine charged

(Equation 2-1). In each equation, n represents the amount of amine or

formamide in terms of mol, and t is the time in h.

fl — fly amine amine /O 1 \

n°amine

The selectivity was calculated as the ratio of the amount of the desired

product and the amount of converted amine (Equation 2-2).

ci formamide /r\ r\ \

n° —namm e a mm e

The turnover number (TON) and the turnover frequency (TOF) are used to

quantify the activity of a catalyst. TON is defined as the ratio of the amount

of produced formamide to the amount of used catalyst (Equation 2-3). The

turnover frequency (TOF) in h-1 is the number of molecules reacting per

active site per h at the conditions of the experiment, according to

Equation 2-4.

n

Tf)M —

formamide O-li}

ncatalyst

Experimental 39

rps-\ y-i formamide /r\ a \

nt1 t-t

catalyst

2.4 References

[I] R. Wandeler, N. Künzle, M. S. Schneider, T. Mallat, A. Baiker, J.

Catal. 200(2001)377.

[2] A. Baiker, Chimia 35 (1981) 408.

[3] J. W. Niemantsverdriet, Spectroscopy in Catalysis, VCH, Weinheim

(1995).

[4] G. Ertl, H. Knözinger, J. Weitkamp, Handbook of Heterogeneous

Catalysis, Wiley-VCH, Weinheim (1997).

[5] B. Imelik, J. C. Védrine, Catalyst Characterization: Physical

Techniquesfor SolidMaterials, Plenum Press, New York (1994).

[6] B. M. Weckhuysen, In-situ Spectroscopy of Catalysts, American

Scientific Publishers, Stevenson Ranch, CA (2004).

[7] H. Friebolin, Ein- und zweidimensionale NMR-Spektroskopie, VCH,

Basel (1992).

[8] R. Voelkel, Angew. Chem. 100 (1988) 1525.

[9] K. L. Walther, A. Wokaun, A. Baiker, Mol. Phys. 71 (1990) 769.

[10] R. K. Harris, Analyst 110 (1985) 649.

[II] M. Hesse, H. Meier, B. Zeeh, Spektroskopische Methoden in der

organischen Chemie, Thieme, Stuttgard (1987).

[12] J. F. Watts, J. Wolstenholme, An Introduction to Surface Analysis by

XPSandAES, Wiley, Chichester (2003).

40 Chapter 2

[13] D. C. Koningsberger, R. Prins, X-Ray Absorption: Principles,

Applications, Techniques of EXAFS, SEXAFS, and XANES, Wiley,

New York (1988).

[14] H. Günzler, H.-U. Gremlich, IR Spectroscopy: An Introduction,

Wiley-VCH, Weinheim (2002).

[15] J.-D. Grunwaldt, M. Caravati, M. Ramin, A. Baiker, Catal. Lett. 90

(2003)221.

Chapter

Catalyst Characterization under Reaction

Conditions Using XANES and EXAFS

3.1 Summary

In the first part of this chapter, X-ray absorption spectroscopy for catalyst

characterization in terms of X-ray absorption near edge structure (XANES)

and extended X-ray absorption fine structure (EXAFS) is introduced. Using

EXAFS, information on the local structure of the absorber atom, such as

the kind of its nearest neighbor atoms, their distances and coordination

numbers can be extracted. XANES gives a "fingerprint" of the electronic

structure and the symmetry. Hence, X-ray absorption spectroscopy can

provide information on the structure of heterogeneous catalysts, even if the

material is X-ray amorphous. Another important advantage of

XANES/EXAFS is the possibility to gain this information in situ, even

under the high-pressure conditions required in this work.

In the second part of this chapter, the potential of such in situ studies

employing a batch-like reactor cell that allows measurements of both the

solid catalyst and the dense liquid-like phase (the reaction mixture) under

high-pressure conditions is shown. For this purpose, one X-ray path probes

the bottom, while the other X-ray path penetrates the center of the in situ

cell. The principles of X-ray absorption spectroscopy are explained and

3

42 Chapter 3

illustrated with the help of examples from the ruthenium-catalyzed

formylation of amines in supercritical carbon dioxide in the presence of

hydrogen.

3.2 Introduction

X-ray absorption spectroscopy (XAS) using synchrotron radiation is a

powerful tool for the characterization of solid materials [1-3]. Together

with other complementary techniques, such as X-ray diffraction (XRD),

infrared spectroscopy (IR), electron spin resonance (ESR, EPR), or Raman

spectroscopy, it is a well-established characterization method in hetero¬

geneous catalysis [3-13]. As shown in Figure 3-1, two regions can be

distinguished, the X-ray absorption near edge structure (XANES) and the

extended X-ray absorption fine structure (EXAFS). The spectrum in the

near edge region is characterized by electronic transitions and multiple

scattering effects. Therefore, it provides a means to obtain information

about the electronic properties and the local structure (symmetry) of the

absorbing atom. The EXAFS region is dominated by single-scattering

events of the outgoing electron wave at the neighboring atoms, giving

information about the local atomic structure around the absorber atom. The

fact that the samples under investigation can be amorphous leads to

important advantages of using XANES/EXAFS in the field of the

heterogeneous catalysis [1,2], since they involve primarily large surface

areas and often amorphous materials.

Another advantage of the XANES/EXAFS technique is the possibility to

investigate the solid catalyst under in situ reaction conditions, as well as

any species dissolved in the liquid phase owing to the good penetration

characteristics of hard X-rays. Only a few techniques, such as IR [14, 15],

Catalyst Characterization In Situ Using XANES and EXAFS 43

NMR [16,17], UV-vis [18,19], Raman [15,20], and EPR [21,22]

spectroscopy, allow in situ studies on heterogeneous catalysts in super¬

critical fluids. In the past decades, only a small number of studies in gas-

solid reactions [3, 23-29] and on solid-liquid interfaces [30-33] at high

temperatures and pressures have taken advantage of the potential of X-ray

absorption spectroscopy. High-pressure reactions of this kind involve not

only a liquid phase, but also solids encountered in various fields of

chemistry, including heterogeneous catalysis [34-37]. For the elucidation of

the reaction mechanism, information on the solid and possibly the

dissolved species under reaction conditions is required concurrently. X-ray

absorption spectroscopy is a valuable method to gain this kind of

information under in situ conditions.

3.3 Basics ofXANES and EXAFS

The basis of X-ray absorption spectroscopy is the absorption of X-rays by

excitation of a core electron to an empty state or continuum (Figure 3-1).

The absorption of the X-rays can be described with Lambert-Beer's law

and is a function of the energy of the exciting X-rays. If the energy is high

enough to excite the core electrons, X-ray absorption is observed, which

leads to an absorption edge. In solid materials, metallic complexes and

clusters in solution, not only an absorption step is visible, but also an

oscillatory fine structure as depicted in Figure 3-1 at 21 keV at the Ru K-

edge.

In 1923, Kronig [38, 39] observed the first extended X-ray absorption fine

structure that was explained in 1971 by Sayers, Stern, and Lytle [40, 41].

The fine structure above the edge can be attributed to interferences from

the outgoing electron wave and the electron wave that is backscattered at

44 Chapter 3

the neighboring atoms. The resulting characteristic oscillations, called the

EXAFS function x(k), exhibits maxima and minima caused by positive and

negative interferences. The EXAFS function depends on the kind of the

neighboring atoms, their distances and their number.

Figure 3-1. Principle of X-ray absorption: schematic representation for a) the

excitation of a core electron and b) the scattering of the electron wave at the

neighboring atom and c) the spectrum of RuClß at the K-edge with the XANES and

EXAFS region.

With the single scattering approximation, the EXAFS function x(k) can be

calculated using Equation 3-1, assuming that the photoelectron scatters

only once during its lifetime [42].

%(k) = TNSl{k)F(k)e-^kle-^^k) sin(-2krs+Mk»(3.1)

Yj represents the distance between the absorber atom i and the neighboring

atoms in the shell j, Nj is the number of neighboring atoms in the shell j,

F;(k) is the backscattering amplitude, §[\(k) is the phase shift experienced

Catalyst Characterization In Situ Using XANES and EXAFS 45

by the photoelectron in the scattering process. So2(k) is the amplitude

reduction factor, causing some damping of the signal. The exponential

terms with the Debye-Waller factor a2 and the mean free path length X are

additional damping factors. The mean free path length X of only a few Â

makes EXAFS an ideal local sampling technique.

In practice, the white X-rays typically used at a synchrotron are

monochromized using a double crystal such as Si(lll) or Si(311). In this

way, a spectrum can be taken by scanning the energy region step by step

around the element-specific near edge region of the sample. The intensity

before and after traversing the sample can be recorded with two ionization

chambers, located before and after the sample. Energy calibration may be

achieved by placing a reference foil between the second and an additional

third ionization chamber. Aside from this transmission experiment, data

can also be taken in the fluorescence mode [2, 3]. The fluorescence mode is

the preferred choice if the concentration of the atoms of interest is low and

no reasonable signal-to-noise ratio can be obtained in transmission

mode [43].

In summary, XANES gives a "fingerprint" of the structure, such as

oxidation state and symmetry around the absorber atom, while the EXAFS

region can reveal the structure of the nearest neighbors, such as distances

and coordination numbers. The result of the Fourier transformation of the

EXAFS data is a radial distribution function.

As an example, Figure 3-2 shows the EXAFS spectra of RuCl2(dppe)2 and

RuCl2(PPh3)3 and the corresponding Fourier-transformed EXAFS

functions, taken in the liquid phase around the Ru K-edge at 22.117 keV.

Due to the high energy of the Ru K-edge, the contribution of the solvent to

the absorption is very small. The structure of Ru phosphine complexes can

46 Chapter 3

also be identified even at very low concentrations (-50 ppm). In these

cases, the spectra are preferably taken in the fluorescence mode [44]. As

demonstrated in chapter 7, section 7.4.4, even the transmission mode is

suitable under certain circumstances, down to a concentration of about

200 ppm.

22 10 22 12 22 14 22 16 22 18 22 20

E/keV3

R/Â

Figure 3-2. a) EXAFS spectra of RuCE(dppe)2 and RuCEfPPhj)^ with b) the

corresponding Fourier-transformed spectra (taken in transmission mode).

In the example shown in Figure 3-2, the spectrum of RuCl2(dppe)2 differs

significantly from that of RuCl2(PPli3)3, both in the XANES and in the

EXAFS region, the reason being the different symmetry and the different

number of phosphine ligands using dppe and PPI13, respectively. Also, the

peaks in the Fourier-transformed spectra are slightly different, but both

structures exhibit a backscattering peak at 1.8 and 2.5 Â. Note that the

spectra presented here are not phase-shift-corrected. The backscattering

contribution of RuCi2(dppe)2 is much smaller than that of RuCi2(PPli3)3.

Analysis of the EXAFS data - by fitting with Ru-P and Ru-Cl shells using

FEFF 6.0 [45] and WinXAS 3.0 [46] - yields the type, the distance, and the

number of the nearest neighbors. The fitting procedure is a multiparameter

problem requiring a careful data analysis procedure. The structural data

Catalyst Characterization In Situ Using XANES and EXAFS 47

obtained for RuCl2(PPli3)3 (see Table 4-1 in chapter 4) are in accordance

with literature [47]. Three different Ru-P bond lengths at 2.23, 2.37, and

2.41 Â and two Ru-Cl bond lengths at 2.387 and 2.388 Â were reported by

single crystal X-ray diffraction. In our studies, only one Ru-P distance at

2.22 Â and a contribution at a higher value of 2.93 Â were found. This can

be attributed to the fact that the measured catalyst was dissolved in the

reaction mixture, and that the EXAFS technique cannot differentiate

between neighbors that are close to each other. In RuCl2(dppe)2, dissolved

in the reaction mixture, a Ru-P distance of 2.32 Â was found in close

accordance with single crystal X-ray diffraction data reported in ref. [48].

3.4 High-pressure In Situ X-ray Absorption Spectroscopy

With XANES and EXAFS, in situ studies on the structure of catalysts

under reaction conditions can be carried out. The penetration of X-rays at

the Ru K-edge both through the solvent and through window materials such

as Be is quite good. Note that CO2 results in 25% absorption over a

distance of 10 mm and Be absorbs even less [42]. Such studies are of great

importance in heterogeneous catalysis, since in situ experiments can

elucidate information about the catalyst structure, which may be very

different from the structure the catalyst exhibits after air exposure. If the

catalyst structure changes under reaction conditions, the relationship

between structure and activity can hardly be determined with ex situ

experiments. Considering the dramatic influence of the high-pressure

reaction conditions on the structure of the catalyst, in situ studies are of

particular importance. Monitoring both the solid catalyst under reaction

conditions and the catalytic species dissolved in the liquid reaction phase at

the same time can lead to a better understanding of the role of possible

48 Chapter 3

dissolved catalyst species. Therefore, this approach can show whether the

reaction is homogeneously or heterogeneously catalyzed (see next section).

In order to get some answers in this context, a specially designed high-

pressure in situ EXAFS cell was constructed, which allows determining the

best compromise between the spectroscopic technique and the reaction

conditions for catalysis. For details of the cell, see chapter 2, section 2.2.2.

The capabilities of this approach are illustrated in the following case study.

3.5 Case Study: Ruthenium-catalyzed Formylation of

3-Methoxypropylamine with Hydrogen and Supercritical

Carbon Dioxide

3.5.1 Introduction

Formylation of amines with supercritical carbon dioxide (SCCO2) and

hydrogen using ruthenium based phosphine complexes is a well-known

"green" process [49-55]. Important products such as A^/V-dimethyl-

formamide and TV-formylmorpholine can be produced in this way. Usually,

homogeneous catalysts were used, and only a few heterogeneous catalysts

have been reported for this reaction [56, 57].

Recently, we found that RU/AI2O3 in the presence of 1,2-bis(diphenyl-

phoshino)ethane (dppe) also shows good activity [58]. In order to get more

insight into the role and the mechanism of this new catalyst, the new in situ

batch reactor cell was used and tested in the formylation of

3-methoxypropylamine with carbon dioxide and hydrogen.

For this purpose, both the solid catalyst and the liquid phase were first

investigated during the formylation without dppe, and then, in a second

step, in the presence of dppe.

Catalyst Characterization In Situ Using XANES and EXAFS 49

3.5.2 Experimental

X-ray Absorption Spectroscopy Measurements and Data Analysis. The in

situ X-ray absorption spectroscopy experiments were performed at

beamline XI of the Hamburger Synchrotron Laboratory (HASYLAB) at

Deutsches Elektronen-Synchrotron (DESY, Hamburg, Germany). The

beam size was cut to 5 x 1 mm and the in situ batch reactor cell was aligned

using the x, z, 9-table. The beamline is equipped with Si(311) and Si(lll)

double crystal monochromators which were detuned to 60% to reduce the

amount of higher harmonics. Three ionization chambers were used to

measure the incident and outcoming X-ray intensities, located before and

after the in situ cell, as well as after a suitable reference foil for energy

calibration. The ionization chamber gases (Ar, N2, Kr, depending on the

energy of the absorption edge) and their pressure were adjusted in such a

way that about 10% were adsorbed in the first ionization chamber and 40%

in the second and third ionization chambers.

Typical scans at the Ru K-edge were recorded in the following way. For

the solid catalyst or the liquid phase under static conditions or after

equilibration, spectra between 21.900 and 23.500 keV were taken in the

step scanning mode. Faster scans in the continuous scanning mode

(QEXAFS) were recorded between 22.065 and 22.665 keV (0.15 s/eV)

during dynamic changes of the structure. As reference for energy

calibration, RUCI3 (pressed as self-supporting wafer with BN) and Ru foil

were used. 5 wt-% RU/AI2O3 (Engelhard) served as catalyst material, and

l,2-bis(diphenylphosphino)ethane served as ligand. Both were used

unaltered, as received by the supplier.

The raw data were energy-calibrated with the help of the respective

reference sample/foil, background-corrected and normalized using the

50 Chapter 3

WINXAS 3.0 software [46]. Linear combination analysis was performed

with this software package as well.

In situ experiment. 100 mg RU/AI2O3 together with 3 ml 3-methoxypropyl-

amine were filled into the batch reactor cell, and hydrogen was added until

a final pressure of 70 bar was reached. Then 4.2 g CO2 was added to the

reaction mixture and the cell was heated to 100°C. After 4 h, in a second

step after cooling to 40°C and depressurizing, about 5 mg dppe were added,

and the same procedure was repeated in the presence of the phosphine.

3.5.3 Results

Figure 3-3 shows the in situ X-ray absorption spectra at the Ru K-edge

recorded a) in the lower part of the cell (mainly solid phase) and b) in the

upper part of the reaction cell (liquid phase). No absorption at the Ru K-

edge was detected in the liquid phase during the first step (without dppe)

while some structural changes of the solid phase were observed.

22.10 22.15 22.20 22.25 22.30 22.10 22.15 22.20 22.25 22.30

E / keV E / keV

Figure 3-3. Monitoring ofthe fate of the solid Ru/Al20^ catalyst and the appearance of

soluble Ru species in the liquidpart of the cell a) at the bottom of the reactor and b) in

the center ofthe spectroscopic batch reactor cell.

Catalyst Characterization In Situ Using XANES and EXAFS 51

The so-called whiteline at 22.12 keV decreased, indicating that the solid

ruthenium-based catalyst was in oxidized state at the beginning of the

experiment, but was reduced in the presence of hydrogen.

Reconstructing the spectra from the initial spectrum (oxidized ruthenium

species) and a reference Ru foil (chosen energy range for linear

combination of the spectra: 22.07 - 22.26 keV) allowed to quantify the

fraction of metallic Ru species in the catalyst. The results are shown in

Figure 3-4.

0 9-

5 jg 0 8-

100 is I'œ0 7-

10-^b)

60 M

ceu°-

b .0 0 4-<0 r- =

40 £o 3 0 3-

II 02-

it 0 1-

0 0-

\

300 400 500

Time / min

0 100 200 300 400 500

Time / min

Figure 3-4. a) Temperature and pressure during the experiment as function of time

(dppe was added after 450 min); b) Quantification of the reduced (bullets) and oxidized

(squares) ruthenium species on the solid catalyst determined by linear combination

analysis (left), and the relative concentration of the dissolved Ru species in the liquid

phase (triangles, final concentration ca. 200 ppm) as determined from the X-ray

absorption step.

It is obvious that the reduction occurred steadily during heating to 100°C.

A further increase of the reaction temperature to 120°C led to slightly more

reduction. Nevertheless, the absorption at the Ru K-edge in the liquid phase

is marginal. Only at very high stirring rates (between t = 250 and 350 min,

Figure 3-4), a small signal of Ru species in the solution could be found,

probably because some of the catalyst formed a suspension. However, as

52 Chapter 3

soon as dppe was added to the reaction mixture, a significant signal in the

upper part of the cell - which reflects the liquid part of the reaction

mixture - was observed. Two spectra are shown in Figure 3-3b. These

results clearly show that the catalyst was partially dissolved, probably due

to the strongly complexing and thus corrosive action of the phosphine

ligand. Interestingly, the concentration of about 200 ppm is sufficient to

catalyze the formylation of 3-methoxypropylamine and the role of these

dissolved Ru species is discussed in more detail in chapter 5.

3.5.4 Discussion

A high-pressure reactor cell suitable for X-ray absorption spectroscopy and

its application in a case study of catalysis has been described. The

technique makes use of the applicability of X-ray absorption spectroscopy

to solid amorphous materials and to structural identification in liquid phase.

In this way, reactions can be studied where both the solid/liquid interface

and the liquid phase need to be monitored. In catalysis, the question

frequently arises, whether the reaction is homogeneously or hetero¬

geneously catalyzed. Because of the high sensitivity and the element-

specificity of XAS, even small concentrations can be detected in solution.

In the case study, it was found that a homogeneous Ru complex formed

upon addition of a phosphine (dppe) ligand, and the resulting catalyst is

thus homogeneous. Recently, Köhler et al. [59] observed that a soluble Pd

catalyst was formed from supported palladium after a short induction time

during the Heck coupling of bromo benzene with styrene. In this case, one

of the reactants served as ligand to dissolve the heterogeneous catalyst.

When the reaction was finished, the catalytically active complex

decomposed again. In that study, the homogeneous species were merely

identified by elemental analysis of the liquid mixture. The present approach

Catalyst Characterization In Situ Using XANES and EXAFS 53

using X-ray absorption spectroscopy allows the identification of both the

structure and the concentration of the catalytically active species in situ -

this may prove to be especially important if the homogeneous species are

observed under reaction conditions only.

3.5.5 Conclusions

The case study illustrates that the new in situ cell is well-suited to follow

heterogeneously catalyzed reactions in SCCO2 or similar solvents at

pressures of up to 200 bar. The cell fills the gap between recently designed

setups for heterogeneously catalyzed reactions in continuous flow cells

[60, 61] and homogeneously catalyzed reactions [62].

3.6 References

[1] B. B. Theo, EXAFS: Basic Principles and Data Analysis, Springer,

Berlin (1986).

[2] D. C. Koningsberger, R. Prins, X-Ray Absorption: Principles,

Applications, Techniques of EXAFS, SEXAFS, and XANES, Wiley,

New York (1988).

[3] Y. Iwasawa, X-ray Absorption Fine Structure for Catalysts and

Surfaces, World Scientific, Singapore 2 (1996).

[4] J. M. Thomas, G. N. Greaves, C. R. A. Catlow, Nucl. Instr. Meth.

Phys. Res. B 97 (1995)1.

[5] G. Vlaic, D. Andreatta, P. E. Colvita, Catal. Today 41 (1998) 261.

[6] B. S. Clausen, Catal. Today 39 (1998) 293.

[7] B. S. Clausen, H. Topsoe, R. Frahm, Adv. Catal. 42 (1998) 315.

[8] J. M. Thomas, Angew. Chem. Int. Ed. 38 (1999) 3588.

54 Chapter 3

[9] D. Bazin, C. Mottet, G. Treglia, J. Lynch, Appl. Surf. Sei. 164

(2000) 140.

[10] G. Sankar, J. M. Thomas, C. R. A. Catlow, Topics Catal. 10 (2000)

255.

D. Bazin, C. Mottet, G. Treglia, Appl. Catal. A 200 (2000) 47.

J. M. Thomas, G. Sankar, J. Synchr. Rad. 8 (2001) 55.

D. Bazin, L. Guczi, Appl. Catal. A 213 (2001) 147.

G. Laurenczy, F. Lukacs, R. Roulet, Anal. Chim. Acta 359 (275)

275.

M. Poliakoff, S. M. Howdle, S. G. Kazarian, Angew. Chem. Int. Ed.

Engl. 34 (1995) 1275.

J. Jonas, High Pressure NMR, P. Diehl, E. Fluck, H. Günther, R.

Kosfeld, J. Seelig, Springer Verlag, Berlin, Heidelberg 24 (1990).

I. T. Horvath, J. M. Millar, Chem. Rev. 91 (1991) 1339.

S. Kim, K. P. Johnston, Ind. Eng. Chem. Res. 26 (1987) 1206.

J. Lu, B. Han, H. Yan, Phys. Chem. Chem. Phys. 1 (1999) 3269.

S. BrunsgaardHansen, R. W. Berg, E. H. Stenby, Appl. Spectr. 55

(2001)745.

S. Ganapathy, T. W. Randolph, C. Carlier, J. A. O'Brien, Int. J.

Thermophys. 17 (1996) 471.

J. L. deGrazia, T. W. Randolph, J. A. O'Brien, J. Phys. Chem. A. 102

(1998) 1674.

J.-D. Grunwaldt, L. Basini, B. S. Clausen, J. Catal. 200 (2001) 321.

Catalyst Characterization In Situ Using XANES and EXAFS 55

[24] K. K. Bando, T. Saito, K. Sato, T. Tanaka, F. Dumeignil, M.

Imamura, N. Matsubayashi, H. Shimada, J. Synchr. Rad. 8 (2001)

581.

[25] J.-D. Grunwaldt, B. S. Clausen, Top. Catal. 18 (2002) 37.

[26] T. Ressler, J. Wienold, R. E. Jentoft, T. Neisius, M. M. Günther,

Topics Catal. 18(2002)45.

[27] R. Revel, D. Bazin, A. Seigneurin, P. Barthe, J. M. Dubuisson, T.

Decamps, H. Sonneville, J. J. Poher, F. Maire, P. Lefrancois, Nucl.

Instr. Meth. Phys. Res. B 155 (1999) 183.

[28] P. Kappen, J.-D. Grunwaldt, B. S. Hammershoi, L. Tröger, G.

Materlik, B. S. Clausen, J. Catal. 198 (2001) 56.

[29] J.-D. Grunwaldt, P. Kappen, L. Basini, B. S. Clausen, Catal. Lett. 78

(2002) 13.

[30] G. Sankar, F. Rey, J. M. Thomas, G. N. Greaves, A. Corma, B. R.

Dobson, A. J. Dent, Chem. Commun. (1994) 2279.

[31] A. P. Markusse, B. F. M. Küster, D. C. Koningsberger, G. B. Marin,

Catal. Lett. 55(1998)141.

[32] H. H. C. M. Pinxt, B. F. M. Küster, D. C. Koningsberger, G. B.

Marin, Catal. Today 39 (1998) 351.

[33] A. Edelmann, W. Schiesser, H. Vinek, A. Jentys, Catal. Lett. 69

(2000)11.

[34] B. Subramaniam, M. A. McHugh, Ind. Eng. Chem. Des. Dev. 25

(1986)1.

[35] A. Baiker, Chem. Rev. 99 (1999) 453.

56 Chapter 3

[36] J.-D. Grunwaldt, R. Wandeler, A. Baiker, Catal. Rev. -Sei. Eng. 45

(2003)1.

[37] P. G. Jessop, W. Leitner, Chemical Synthesis Using Supercritical

Fluids, Wiley-VCH, Weinheim (1999).

[38] R. D. Kronig, Z. Phys. 70 (1931) 317.

[39] R. D. Kronig, Z. Phys. 75 (1932) 468.

[40] D. E. Sayers, E. A. Stern, F. W. Lytle, Phys. Rev. Lett. 27 (1975)

4836.

[41] F. W. Lytle, J. Synchr. Rad. 6 (1999) 123.

[42] J.-D. Grunwaldt, A. Baiker, Phys. Chem. Chem. Phys. 7 (2005)

3526.

[43] J. Jaklevic, J. A. Kirby, M. P. Klein, A. S. Robertson, G. S. Brown,

P. Eisenberger, Solid State Commun. 23 (1977) 679.

[44] S. G. Fiddy, J. Evans, T. Neisius, X. Z. Sun, Z. Jie, M. W. George,

Chem. Commun. 6 (2004) 676.

[45] S. I. Zabinsky, J. J. Rehr, A. Ankudinov, R. C. Albers, M. J. Eller,

Phys. Rev. B. 52 (1995) 2995.

[46] T. Ressler, J. Synchrotron Rad. 5 (1998) 118.

[47] S. J. LaPlaca, J. A. Ibers, Inorg. Chem. 4 (1965) 778.

[48] T. S. Lobana, R. Singh, E. R. T. Tiekink, J. Coord. Chem. 21 (1990)

225.

[49] P. G. Jessop, T. Ikariya, R. Noyori, Nature 368 (1994) 231.

Catalyst Characterization In Situ Using XANES and EXAFS 57

[50] P. G. Jessop, Y. Hsiao, T. Ikariya, R. Noyori, J. Am. Chem. Soc. 118

(1996)344.

[51

[52

[53

[54

[55

[56

[57

[58

[59

[60

[61

[62

O. Kröcher, R. A. Koppel, A. Baiker, Chimia 51 (1997) 48.

C.-C. Tai, J. Pitts, J. C. Linehan, A. D. Main, P. Munshi, P. G.

Jessop, Inorg. Chem. 41 (2002) 1606.

Y. Kayaki, Y. Shimokawatoko, T. Ikariya, Adv. Synth. Catal. 345

(2003) 175.

L. Schmid, A. Canonica, A. Baiker, Appl. Catal. A: Gen. 255 (2003)

23.

L. Schmid, M. S. Schneider, D. Engel, A. Baiker, Catal. Lett. 88

(2003) 105.

O. Kröcher, R. A. Koppel, M. Fröba, A. Baiker, J. Catal. 178 (1998)

284.

O. Kröcher, R. A. Koppel, A. Baiker, J. Mol. Catal. A: Chem. 140

(1999) 185.

M. Rohr, J.-D. Grunwaldt, A. Baiker, J. Catal. 229 (2005) 144.

S. S. Pröckl, W. Kleist, M. A. Gruber, K. Köhler, Angew. Chem. Int.

Ed. 43(2004)1881.

J.-D. Grunwaldt, M. Caravati, S. Hannemann, A. Baiker, Phys.

Chem. Chem. Phys. 6 (2004) 3037.

J.-D. Grunwaldt, M. Caravati, M. Ramin, A. Baiker, Catal. Lett. 90

(2003)221.

S. L. Wallen, D. M. Pfund, J. L. Fulton, C. R. Yonker, M. Newville,

Y. Ma, Rev. Sei. Instrum. 67 (1996) 2843.

Chapter

A Simple Route to Highly Active Ruthenium

Catalysts for Formylation Reactions with

Hydrogen and Carbon Dioxide

4.1 Summary

A simple route to ruthenium catalysts suitable for formamide production

from amines, hydrogen, and carbon dioxide is reported. The formylation of

3-methoxypropylamine has been employed as a test reaction. Highly active

and selective ruthenium based catalysts were formed in situ under reaction

conditions from solid rutheniumtrichloride (RUCI3) in the presence of

triphenylphosphine (PPI13) and l,2-bis(diphenylphosphino)ethane (dppe).

While RUCI3 does not catalyze the reaction efficiently, the addition of

phosphines led to a near five-fold increase in rate. The achieved turnover

frequencies are comparable to those of synthesized reference Ru-phosphine

complexes. As a consequence of the high activity, only very small amounts

(-300 ppm) of both RUCI3 and the phosphine are necessary to catalyze

effectively the formylation reaction. In spite of the very low concentration

of the Ru complex, the structure of the in situ formed active complex was

uncovered by X-ray absorption near edge structure (XANES) and extended

X-ray absorption fine structure (EXAFS) spectroscopy. Both indicated

4

60 Chapter 4

similar local structures for the in situ formed complex and a Ru-reference

complex after reaction.

4.2 Introduction

The simultaneous use of carbon dioxide in chemical synthesis as a Ci-

building block and as a solvent is an attractive strategy for green chemistry.

Toxic chemicals such as phosgene and carbon monoxide as well as

environmentally harmful organic solvents may be potentially replaced [1-

4]. One interesting route is the formylation of amines with hydrogen and

carbon dioxide in the presence of ruthenium phosphine complexes, cf.

chapter 1, section 1.3.2 [2,5,6]. For example, the ruthenium catalyzed

synthesis of dimethylformamide from dimethylamine, hydrogen and carbon

dioxide showed good results in homogeneous and heterogeneous catalysis

[3,7,8]. Apart from dimethylamine, other amines, such as propylamine

[9, 10], diethylamine [11] and morpholine [12, 13], have been investigated.

Up to now, the homogeneous catalysts used to be prepared directly,

involving complexes like RuH2(P(CH3)3)4, RuCl2(P(CH3)3)4,

RuH2(P(C6H5)3)4, and RuCl(02CCH3)(P(CH3)3)4 [9, 14, 15], or by the in

situ formation of the active complexes from [RuCl2(CgH6)]2,

RuCl2(DMSO)4, and [RuCl2(COD)]n [16]. However, the use of a simple

ruthenium salt, RUCI3, in the presence of phosphine ligands has not been

reported and tested in the solventless formylation of amines using carbon

dioxide as a Ci-synthesis block, see chapter 1, section 1.3. To our

knowledge, the only studies hinting in this direction are the formylation of

amines in hexane in the presence of a RUCI3-PPI13-THF solution covered in

the patent of Kiso et al. [17], and the synthesis of formic acid in the

A Simple Route to Ruthenium-catalyzed Formylations 61

presence of RUCI3/PPI13 in a mixture of ethanol and water as the solvent

[18].

In this chapter, it is shown that highly active Ru catalysts can be prepared

in situ by adding a suitable phosphine, l,2-bis(diphenylphosphino)ethane

(dppe) or triphenylphosphine (PPI13), to the reaction mixture containing

RUCI3 as a pre-catalyst (route 2, Figure 1-2 in chapter 1, section 1.5). We

investigated this simple in situ preparation of the catalyst using the

formylation of 3-methoxypropylamine as a test reaction. It has been shown

that the catalytic performance of the in situ generated homogeneous

catalysts is comparable to that of RuCl2(dppe)2 and RuCl2(PPli3)3

reference complexes. Both strategies lead to highly active and virtually

100% selective catalysts. Inductively coupled plasma optical emission

spectroscopy, X-ray absorption near edge structure, and extended X-ray

absorption fine structure spectroscopy were applied to gain information

about the concentration and the structure of the catalyst formed in situ from

RUCI3 and the phosphine.

4.3 Experimental

The formylation of 3-methoxypropylamine (mpa, Scheme 4-1) was chosen

as a test reaction. The reaction was carried out in a temperature-controlled

500 ml high-pressure stainless steel autoclave with a dosing system for

gases [19]. In a typical procedure, mpa and the catalyst (50 jiimol Ru,

adding the corresponding stoichiometric amount of phosphine in the case

of RUCI3) were filled into the reactor before closing and flushing with

hydrogen. The reactor was then filled with hydrogen and adjusted to a

pressure of 80 bar at 100°C. Finally, 100 g carbon dioxide was added,

resulting in a total pressure of 220 bar at 120°C. The stirring rate was fixed

62 Chapter 4

at 300 min-1. After a certain reaction time, the reactor was cooled down

and depressurized by opening the outlet valve. The reaction mixture

consisted of two phases, a dense liquid phase containing the reactant amine

and the products (formamide and water), and a less dense supercritical

carbon dioxide phase. The reaction mixture was analyzed using a gas

Chromatograph (HP-6890) equipped with a HP-5 capillary column (30 m x

0.32 mm x 0.25 jiim) and a flame ionization detector (FID). Product

identification was achieved with a gas Chromatograph (HP-6890) coupled

to a mass spectrometer (HP-5973).

scCO,, H, -H,0 H

RuCI3 in situ formed

+ — homogeneous

dppe Ru catalyst

Scheme 4-1. Formylation of 3-methoxypropylamine with hydrogen, supercritical

carbon dioxide and in the presence ofhighly active and selective ruthenium catalysts.

The X-ray absorption spectroscopy studies were performed at beamline XI

at HASYLAB (DESY). A Si(311) double-crystal monochromator was used

and higher harmonics were effectively removed by detuning the crystals to

70% of the maximum intensity. Additional experiments were performed at

SNBL (ESRF, Grenoble). EXAFS spectra were taken in the step-scanning

mode around the Ru K-edge (22.117 keV) between 21.900 and 22.800 keV

using a RUCI3 pellet as reference. Due to the low concentration of

ruthenium, a transmission EXAFS cell with a long path length of 4 cm was

used as described in chapter 2, section 2.2.1. The raw data were energy-

calibrated (Ru K-edge energy of RUCI3: 22.120 keV [20], first inflection

point), background-corrected and normalized using the WINXAS3.0

software [21]. Fourier transformation for the EXAFS data was applied to

A Simple Route to Ruthenium-catalyzed Formylations 63

the k1-weighted functions in the interval k = 3-12.5Â_l. Data fitting was

performed in R-space. Typical deviations are ±0.5 for the coordination

number and ±0.02 Â for the distance.

4.4 Results and Discussion

RUCI3 used as pre-catalyst showed only a low activity with a conversion of

12% and a turnover frequency (TOF, see definition in chapter 2,

section 2.3.3) of 19 h-1 under standard reaction conditions as denoted in

Figure 4-1. After adding dppe, the activity of the in situ formed catalyst

increased, affording a TOF of 90 Ir1 and a conversion of 95%. A similar

catalytic performance was observed when adding PPI13 instead of dppe

(Figure 4-1). The catalytic behavior of the homogeneous catalysts

RuCl2(dppe)2 and RuCl2(PPli3)3 used as references was in the same range.

The selectivity to the corresponding formamide was 100% in all reactions.

100 100

dppe

RuCI, RuCl^dppe), RuCI,(PPh,),+

PPh

Figure 4-1. Comparison ofthe performance ofcatalysts generated in situfrom RuCE +

dppe or PPh$ and related homogeneous reference catalysts RuC^fPPh^ß and

RuCE(dppe)2 in the formylation of mpa. Selectivity to formamide was 100% in all

cases. Reaction conditions: 100 mmol mpa, 50 jumol Ru catalyst, 100 g CO2, 120°C,

totalpressure 220 bar, 20 h.

64 Chapter 4

Consequently, this in situ generation offers a very simple and cheap

procedure for preparing highly active homogeneous Ru-catalysts. To

confirm the formation of complexes acting as homogeneous catalysts

formed from RUCI3 and dppe or PPI13, firstly the Ru-concentration in the

reaction mixtures after filtration and centrifugation was quantified with

ICP-OES, and secondly, pertinent tests according to Sheldon et al. [22]

were performed. Specifically, the reaction was stopped after a certain time,

the reaction mixture was filtered off, and it was checked whether the

catalytic reaction could be continued after this procedure. No significant

loss of activity could be observed. This confirmed the supposition that all

Ru-complexes formed were dissolved in the amine rich phase. After 3 h of

reaction time, 49% of the amine was converted with the catalyst RUCI3 +

dppe. Only about 300 ppm ruthenium was present in the reaction mixture.

After filtration and continuation of the reaction for a total time of 20 h, the

conversion reached 90%. In an experiment without filtration, virtually the

same conversion (88%) was found after 20 h, which proved that the

ruthenium based complex is a homogeneous catalyst formed in situ under

reaction conditions. Note that the conversion of RUCI3 + PPI13 after 3 h was

25% compared to 49% for the RUCI3 + dppe system.

This is in line with literature, where dppe complexes were found to be more

active than monophosphine complexes [8], and it also supports our

hypothesis that the Ru-phosphine complex act as homogeneous catalyst.

To gain some information about the structure of the in situ formed Ru-

complexes, the product mixtures were investigated by XANES and EXAFS

using an appropriate stainless steel cell constructed for transmission

experiments (chapter 2, section 2.2.1). As a result of the low Ru-

concentration in the product mixture, only the use of the long path length

A Simple Route to Ruthenium-catalyzed Formylations 65

(4 cm) in this cell allows structural identification even at low

concentrations. Because of the high energy of the Ru K-edge, the

contribution of the solvent to the absorption is very small [23]. This shows

that the structure of Ru-phosphine complexes can be identified even at low

concentrations. Studies on the local structure of ruthenium in systems with

such low concentrations of the target species are difficult to perform by

means of any other technique [24]. Essentially, EXAFS spectra can be

taken in two different modes: fluorescence detection or transmission mode.

Whereas Fiddy et al. [25] proposed the use of fluorescence detection, it

seemed that in our case transmission EXAFS is a viable alternative. Both

the Ru K-edge and Rh K-edge are found at rather high energies, so that the

absorption at 22 keV is only ud = 0.08. At 300 ppm of Ru concentration, an

edge jump of 1.5 was found, leading to a sufficient signal/noise ratio.

22'10'

22'12'

22'14'

22'16'

22'18'

22'20E/keV

Figure 4-2. X-ray absorption near edge structure at the Ru K-edge ofRuClß, RuClß +

dppe, RuCE(dppe)2 and RuCEfPPh^)^ All spectra were taken from the liquidproduct

mixture.

Figure 4-2 shows the XANES region of the catalysts RUCI3 + dppe and the

references RUCI3, RuCl2(dppe)2, and RuCl2(PPli3)3. Upon dppe addition,

the spectrum of RUCI3 changed significantly. From these changes, the

66 Chapter 4

comparison of the near edge structure of the in situ formed complex and the

Ru-phosphine reference complexes, RuCl2(dppe)2, and RuCl2(PPli3)3, we

infer that the local structure of the in situ formed complex is similar to

RuCl2(PPfi3)3 after reaction.

Figure 4-3 shows the corresponding kl-weighted Fourier-transformed

EXAFS spectra. All catalysts exhibit backscattering peaks at 1.8 Â and

2.5 Â; RUCI3 shows a supplementary peak at 3.25 Â. The peaks are slightly

shifted to lower R values because the spectra are not corrected for the phase

shift. The backscattering contribution of RUCI3 and RuCl2(dppe)2 at 2.5 Â

is much smaller than that of RUCI3 + dppe and RuCl2(PPli3)3. The Fourier-

transformed spectra were fitted with Ru-Cl and Ru-P shells calculated by

FEFF 6.0 [26].

=> 0 0035--

cc

X 0 0030-

73CD

£ 0 0025-

I 0 0020-

^-0 0015-

CD

J 0 0010-

E

^ 0 0005-

I- 0 0000-Fü- 0

Figure 4-3. The corresponding Fourier-transformed EXAFS spectra (k1-weighted) at

the Ru K-edge ofRuClj, RuCE + dppe, RuCE(dppe)2, andRuCEfPPhj)^

The structural data obtained for RuCl2(PPfi3)3 are in good accordance with

literature [27]. Three different Ru-P bond lengths (2.23 À, 2.37 À, 2.41 À)

and two Ru-Cl bond lengths (2.387 À, 2.388 À) were reported (Table 4-1).

Only one Ru-P distance, 2.22 Â, and one contribution at higher values at

2.93 Â, were found in our studies (Table 4-2). This originates from the fact

RuCI3 + dppe

RuCI3

RuCI2(PPh3)3— RuCI2(dppe)2

A Simple Route to Ruthenium-catalyzed Formylations 67

that we measured the catalyst dissolved in the reaction mixture. A Ru-P

distance of 2.32 Â was found in RuCl2(dppe)2 in the reaction mixture that

is also documented in literature [28].

Table 4-1. Structural data from X-ray diffraction of Ru-phosphine complexes as

reported in literature

Compound Bond length/Â Bond ang;les/° Reference

RuCl2(PPh3)3 Ru-P(l) 2.374 P(l)-Ru-P(2) 156.4 [27]

Ru-P(2) 2.412 P(l)-Ru-P(3) 101.1

Ru-P(3) 2.230 P(2)-Ru-P(3) 101.4

Ru-Cl(l) 2.387

Ru-Cl(2) 2.388

RuCl2(dppm)2 Ru-P(l)

Ru-P(2)

Ru-Cl

2.340

2.367

2.426

P(l)-Ru-P(2) 71.39 [29]

RuCl2(dppe)2 Ru-P(l)

Ru-P(2)

Ru-Cl

2.389

2.369

2.436

P(l)-Ru-P(2) 82.1 [28]

RuBrCl(dppe)2 Ru-P(l)

Ru-P(2)

Ru-Cl

Ru-Br

2.388

2.362

2.432

2.496

P(l)-Ru-P(2) 81.43 [30]

RuCl2(dppp)2 Ru-P(l)

Ru-P(2)

Ru-Cl

2.416

2.441

2.435

P(l)-Ru-P(2) 86.79 [31]

dppm: bis(diphenylphosphino)methane; dppe: l,2-bis(diphenylphosphino)ethane; dppp: 1,3-bis-

(diphenylphosphino)propane.

68 Chapter 4

With a Ru-P bond length of 2.27 Â and a contribution at 2.82 Â, RUCI3 +

dppe shows a local structure that is similar to RuCl2(PPfi3)3 in solution.

The excess of chloride ligands in the RUCI3 + dppe system underlines more

that structure and even leads to the formation of dimers which could

explain the contribution at higher R-values. Note that the EXAFS fit only

resulted in a Ru-P shell, but not in a Ru-Cl contribution as reported in refs.

[32,33].

Table 4-2. Structural data extracted from EXAFS spectra of different ruthenium

phosphine complexes as solid and in the liquid reaction mixture (after reaction)

Catalyst Ru-P Ru-Cl and Ru-Ru

N R/Â Aa2/Â2 N R/Â Aa2/Â2

RuCl2(PPh3)3a 1.0

2.0

2.23

2.38

0.0025

0.0025

2.3 2.44 0.0027

RuCl2(dppe)2a 4.0 2.29 0.0013 2 2.42 0.0006

RuCl2(PPh3)3b 3.0 2.24 0.0048 2 2.93 0.0060

RuCl2(dppe)2b 3.6 2.32 0.0063 - - -

RuCl3 + dppeb 3.9 2.27 0.0046 2 2.82 0.0091

N Coordination number; R: distance of the corresponding neighbor; Aa2: Debye-Waller factor.

a Solid catalyst.b

Liquid solution measured after reaction with 3-methoxypropylamine (mpa).

This may be attributed to the similar distances of Ru-P and Ru-Cl in the

formed complexes. However, the Ru-P distance is usually shorter than the

Ru-Cl distance [32, 33], indicating the formation of phosphine complexes.

Note that one of the limitations of EXAFS is that it is an averaging

technique. We can assume that several species exist in solution, but these

are averaged out by the EXAFS technique. Nevertheless, the XANES and

A Simple Route to Ruthenium-catalyzed Formylations 69

EXAFS data show that the structure of the in situ formed complex of the

RUCI3 + dppe mixture is similar to RuCl2(PPfi3)2 after reaction. In a next

step, it would be desirable to identify in more detail the hydrido species,

which is known to be easily formed in the presence of hydrogen [34, 35].

Finally, it seems likely that the extension of the in situ catalyst preparation

to other transition metals and other phosphines may provide catalysts with

even higher efficiency than those shown in this work. A systematic study

towards this aim could be rewarding.

4.5 Conclusions

The in situ generation of highly active catalysts from solid RUCI3 and a

suitable phosphine (dppe or PPI13) for the formylation of mpa from

hydrogen and carbon dioxide has been demonstrated by catalytic

measurements, in combination with X-ray absorption spectroscopy and

ICP-OES. The activity and the structure of the formed catalyst is similar to

RuCl2(PPfi3)3. The use of this simple method for catalyst preparation may

open new possibilities for high-throughput screening and lead to more

economical and greener formylation processes requiring only very low

amounts of an easily accessible catalyst.

4.6 References

[1] A. Behr, Angew. Chem. Int. Ed. Engl. 27 (1988) 661.

[2] W. Leitner, Angew. Chem. Int. Ed. Engl. 34 (1995) 2207.

[3] P. G. Jessop, T. Ikariya, R. Noyori, Chem. Rev. 95 (1995) 259.

[4] A. Baiker, Appl. Organomet. Chem. 14 (2000) 751.

[5] O. Kröcher, R. A. Koppel, A. Baiker, Chimia 51 (1997) 48.

70 Chapter 4

[6] P. G. Jessop, T. Ikariya, R. Noyori, Chem. Rev. 99 (1999) 475.

[7] S. Schreiner, J. Y. Yu, L. Vaska, Inorg. Chim. Acta 147 (1988) 139.

[8] O. Kröcher, R. A. Koppel, A. Baiker, Chem. Commun. 5 (1997) 453.

[9] P. G. Jessop, Y. Hsiao, T. Ikariya, R. Noyori, J. Am. Chem. Soc. 118

(1996)344.

[10] L. Schmid, A. Canonica, A. Baiker, Appl. Catal. A: Gen. 255 (2003)

23.

[11] L. Schmid, M. Rohr, A. Baiker, Chem. Commun. (1999) 2303.

[12] L. Schmid, M. S. Schneider, D. Engel, A. Baiker, Catal. Lett. 88

(2003) 105.

[13] G. Süss-Fink, M. Langenbahn, T. Jenke, J. Organomet. Chem. 368

(1989)103.

[14] P. G. Jessop, T. Ikariya, R. Noyori, Nature 368 (1994) 231.

[15] C. A. Thomas, R. J. Bonilla, Y. Huang, P. G. Jessop, Can. J. Chem.

79(2001)719.

[16] C.-C. Tai, J. Pitts, J. C. Linehan, A. D. Main, P. Munshi, P. G.

Jessop, Inorg. Chem. 41 (2002) 1606.

[17] Y. Kiso, K. Saeki, Japan. Kokai Tokkyo Koho (1977) 36617.

[18] J. Z. Zhang, Z. Li, H. Wang, C. Y. Wang, J. Mol. Catal. A: Chem.

112(1996)9.

[19] O. Kröcher, R. A. Koppel, A. Baiker, J. Mol. Catal. A: Chem. 140

(1999) 185.

A Simple Route to Ruthenium-catalyzed Formylations 71

[20] K. Okamoto, T. Takahashi, K. Kohdate, H. Kondoh, T. Yokoyama,

T. Ohta, J. Synchrotron Rad. 8 (2001) 689.

[21] T. Ressler, J. Synchrotron Rad. 5 (1998) 118.

[22] R. A. Sheldon, M. Wallau, I. W. C. E. Arends, U. Schuchardt, Ace.

Chem. Res. 31(1998)485.

[23

[24

[25

[26

[27

[28

[29

[30

[31

J.-D. Grunwaldt, A. Baiker, Phys. Chem. Chem. Phys. 7 (2005)

3526.

P. G. Jessop, W. Leitner, Chemical Synthesis Using Supercritical

Fluids, Wiley-VCH, Weinheim (1999).

S. G. Fiddy, J. Evans, T. Neisius, X. Z. Sun, Z. Jie, M. W. George,

Chem. Commun. 6 (2004) 676.

S. I. Zabinsky, J. J. Rehr, A. Ankudinov, R. C. Albers, M. J. Eller,

Phys. Rev. B. 52 (1995) 2995.

S. J. LaPlaca, J. A. Ibers, Inorg. Chem. 4 (1965) 778.

T. S. Lobana, R. Singh, E. R. T. Tiekink, J. Coord. Chem. 21 (1990)

225.

A. R. Chakravarty, F. A. Cotton, W. Schwotzer, Inorg. Chim. Acta

84(1984)179.

E. P. Perez, R. M. Carvalho, R. H. A. Santos, M. T. P. Gambardella,

B. S. Lima-Neto, Polyhedron 22 (2003) 3289.

M. R. M. Fontes, G. Oliva, L. A. C. Cordeiro, A. A. Batista, J.

Coord. Chem. 30 (1993) 125.

72 Chapter 4

[32] E. Lindner, S. Al-Gharabli, I. Warad, H. A. Mayer, S. Steinbrecher,

E. Plies, M. Seiler, H. Bertagnolli, Z. Anorg. Allg. Chem. 629 (2003)

161.

[33] O. Kröcher, R. A. Koppel, M. Fröba, A. Baiker, J. Catal. 178 (1998)

284.

[34] M. G. Basallote, J. Duran, Inorg. Chem. 38 (1999) 5067.

[35] B. R. James, A. Pacheco, S. J. Rettig, I. S. Thorburn, R. G. Ball, J.

A. Ibers, J. Mol. Catal. 41 (1987) 147.

Chapter

Formylation with Supercritical Carbon

Dioxide over RU/AI2O3 Modified by

Phosphines: Heterogeneous or

Homogeneous Catalysis?

5.1 Summary

The formylation of 3-methoxypropylamine with hydrogen and supercritical

carbon dioxide over ruthenium-based catalysts was studied. In this

solventless process, carbon dioxide acts both as a reactant and as a solvent.

Interestingly, Ru/Al203 modified by the phosphine 1,2-bis(diphenyl-

phosphino)ethane (dppe) showed a high formylation activity at 100%

selectivity, comparable to those of the homogeneous catalysts

RuCl2(dppe)2 and RuCl2(PPli3)3. Analysis of the reaction mixture using

ICP-OES and structural studies by in situ X-ray absorption spectroscopy

revealed that the presence of the phosphine modifier did not lead to the

modification of the Ru surface, but to the formation of a homogeneous

ruthenium catalyst.

5

74 Chapter 5

5.2 Introduction

A step toward "green" formylation of amines is the use of carbon dioxide

and hydrogen as formylation agents instead of toxic compounds such as

carbon monoxide and phosgene [1-5]. This approach provides the

opportunity to use supercritical carbon dioxide acting as both Ci -building

block and solvent (solvent-free process) [6-9]. A variety of ruthenium

complexes have been reported [10, 11] to be active and selective for such

formylation reactions, including RuCl2(PMe3)4 in the formylation of

hydrogen with carbon dioxide to formic acid [7], RuCl2(PPli3)3,

Ru3(CO)i2 [12] and RuCl2(dppe)2 [13] in the formylation of

dimethylamine, propylamine [14], and morpholine [15]. Among these

ruthenium-based complexes, RuCl2(dppe)2 was found to be superior with

respect to activity and stability [13, 14, 16].

A further step toward a "green" process is the application of heterogeneous

catalysts because of their intrinsic advantages concerning catalyst

separation, reuse and handling. An interesting approach is the

immobilization of the corresponding homogeneous complexes

functionalized by silyl ether groups in an inert silica matrix, as shown

previously for RuCl2X3 (X = Ph2P(CH2)2Si(OEt)3, Me2P(CH2)2Si(OEt)3)

[8, 17] and RuCl2(dppp)2 (dppp = Ph2P(CH2)3PPh2) [18]. However, the

preparation of these Ru-silica hybrid gels is rather demanding, and more

easy accessible heterogeneous catalysts would be desirable. A simpler

approach might involve the modification of a supported Ru-catalyst with a

suitable phosphine, which is the focus of this study. The modification of

metal catalysts by adsorbed auxiliaries (modifiers) has been successfully

applied to improve the catalytic properties of a variety of metals, as

covered in a recent review [19]. A related example is the use of phosphine

Formylation over Ru/Al203 Modified by Phosphines 75

modifiers in partial oxidation reactions on Pt/alumina catalysts [20]. This

prompted us to apply this strategy to the formylation using a 5 wt-%

Ru/Al203 catalyst modified with a phosphine (dppe or PPI13, see route 2,

Figure 1-2 in chapter 1, section 1-5). We show that these catalysts exhibit

good activity in the formylation of 3-methoxypropylamine at 100%

selectivity. Extensive in situ EXAFS studies using a specially designed

high-pressure cell revealed that the observed catalytic behavior has to be

attributed to a highly active homogeneous Ru complex formed from the

Ru/alumina catalyst during reaction.

5.3 Experimental

5.3.1 Catalyst Materials

The 5 wt-% Ru/Al203 catalyst (9001, escat 44, powder, prereduced) was

used as received from Engelhard. Note that the ruthenium catalyst was

partially oxidized, but we denote it as Ru/Al203. The 5 wt-% Ru/Al203

catalyst was modified with dppe in two different ways. On the one hand,

dppe was added before the reaction by reflux in THF during 2 h and dried

under high vacuum; on the other hand, it was simply added during the

reaction (section 5.3.2).

The homogeneous RuCl2(dppe)2 catalyst was synthesized according to the

method of Mason et al. [21]. A suspension of 1.00 g RuCl2(PPli3)3 in 20 ml

acetone was mixed with 0.85 g dppe under Ar atmosphere and fast stirring.

A yellow precipitate was observed after 2 min. It was separated with a

suction filter after 10 min of stirring at room temperature. The yellow

powder produced was washed with acetone and methanol and dried in

vacuum, and its structure was confirmed with lH-, 31P-NMR and elemental

76 Chapter 5

analysis; calculated (%) for C52H48Cl2P4Ru (968.8 g/mol): C 64.47, H

4.99, P 12.79, CI 7.32, Ru 10.43, found: C 64.30, H 5.19, P 12.94.

RuCl2(PPli3)3, tris(triphenylphosphine)ruthenium(II)dichloride (Fluka,

purum) was used as received after structural analysis by lH-, 31P-NMR and

elemental analysis; calculated (%) for C54H45Cl2P3Ru (958.8 g/mol): C

67.64, H 4.73, P 9.69, CI 7.39, Ru 10.54; found: C 67.36, H 4.84, P 9.72,

CI 7.56.

5.3.2 Catalytic Formylation of 3-Methoxypropylamine

The catalytic studies were performed using a 500 ml high-pressure stainless

steel autoclave (see chapter 2, section 2.1.1) with temperature control, a

rupture disk, and a dosing system for gases [22]. The chemicals 3-methoxy¬

propylamine (mpa, Fluka, >99%) and l,2-bis(diphenylphosphino)ethane

(dppe, Fluka, 98%) were used as received, liquid carbon dioxide (99.995%)

and hydrogen gas (99.999%) were supplied by Pangas.

In a typical procedure, mpa, dppe, and Ru/Al203 were poured into the

reactor before it was closed and flushed with hydrogen. The reactor was

filled with hydrogen (p ~ 60 bar). After the reactor temperature reached

100°C, the hydrogen pressure was adjusted to 80 bar, and 100 g carbon

dioxide was added. This resulted in a total pressure of about 200 bar. The

stirring rate was fixed at 300 min-1. After a certain reaction time, the

reactor was cooled down and depressurized by opening the outlet valve.

The reaction mixture was analyzed with a gas Chromatograph (HP-6890)

equipped with a HP-5 capillary column (30 m x 0.32 mm x 0.25 ixm) and a

flame ionization detector (FID). Product identification was achieved with a

gas Chromatograph (HP-6890) coupled to a mass spectrometer (HP-5973).

In a parametric study, the temperature and the amounts of catalyst, mpa,

Formylation over Ru/Al203 Modified by Phosphines 77

hydrogen and carbon dioxide were varied in order to find the optimal

reaction conditions.

5.3.3 Characterization of In Situ Formed Homogeneous Complex

Using ICP-OES

After reaction, selected samples were filtrated and centrifuged to quantify

the ruthenium and phosphorus content of the liquid phase with inductively

coupled plasma optical emission spectroscopy (performed by ALAB AG in

Urdorf, Switzerland).

5.3.4 Ex Situ and In Situ X-ray Absorption Spectroscopic Studies

The extended X-ray absorption fine structure (EXAFS) and X-ray

absorption near edge structure (XANES) experiments presented here were

mainly performed at the beamline XI, HASYLAB at DESY in Hamburg,

Germany. The storage ring typically operates at 4.45 GeV and a ring

current between 80 and 120 mA. A Si(311) double-crystal monochromator

was used, and higher harmonics were effectively removed by detuning the

crystals to 70% of the maximum intensity. Three ionization chambers filled

with Ar were used to record the intensity of the incident and the transmitted

X-rays. The samples were located between the first and second ionization

chamber, and a reference sample (RUCI3 pellet) was placed between the

second and the third ionization chamber. Some additional experiments

were performed at the Swiss-Norwegian beamline (SNBL) at the European

Synchrotron Radiation Facility (ESRF) in Grenoble, France. At SNBL, a

Si(lll) crystal was used as a channel-cut monochromator, and a double-

bounce gold-coated mirror system helped to reject the higher harmonics.

The three ionization chambers were filled with Ar, N2 or Kr in different

combinations (I0 Ar, It 30% Kr and 70% N2, Iref 30% Kr and 70% N2). At

78 Chapter 5

both beamlines (XI and SNBL), EXAFS spectra were taken under

stationary conditions in the step-scanning mode around the Ru K-edge

(22.117 keV) between 21.900 and 22.800 keV, with a RuCl3 pellet as a

reference. Fast QEXAFS scans were recorded in the continuous scanning

mode, usually between 22.065 and 22.665 keV (0.15 s/eV). The raw data

were energy-calibrated (Ru K-edge energy of the RUCI3 pellet: 22.120 keV

[23], first inflection point), background-corrected, and normalized with

WINXAS 3.0 software [24]. The EXAFS data was Fourier-transformed to

the kl-weighted functions in the interval k = 3-12.5Â_l and fitted to R-

space. Typical deviations for the coordination number were within ±0.5,

and those for the distance were within ±0.02 Â.

To identify the structure of the ruthenium complexes formed, the product

solution was investigated by EXAFS on liquid samples after reaction as

well. For this purpose, the special stainless steel EXAFS cell, described in

chapter 2, section 2.2.1, was used for transmission experiments. Because of

the low concentration of Ru in the product mixture (50- 100 ppm) the

EXAFS cell with the long path length of 4 cm was chosen.

After calibration with a sample of known absorption step u-d, the

concentration of the solution could also be determined independently of the

ICP-OES measurements.

In addition, in order to perform spectroscopic studies under reaction

conditions, the in situ batch reactor described in chapter 2, section 2.2.2

was used, since it allows investigation of the reaction volume at two

locations [25]. Both the solid catalyst sample at the bottom and the liquid

phase 10 mm above the bottom could be probed. For the experiments,

100 mg of 5 wt-% Ru/Al203 was loaded into the batch reactor cell, 3.0 ml

of mpa was added, and the reactor was closed. After cell alignment, normal

Formylation over Ru/Al203 Modified by Phosphines 79

EXAFS spectra of the solid catalyst and the solution were taken at room

temperature, while the solution was stirred with a magnetic stirrer.

Hydrogen (purity 99.999%) was added to a pressure of 70 bar, and 4 g of

liquid C02 was introduced with the help of a C02 compressor (NWA PM-

101) and a Rheonik mass flow controller (RHM015). While changes in the

solid material were monitored by QEXAFS, the mixture was heated to

120°C. Then the reaction mixture was cooled down to room temperature.

The treatment (addition of hydrogen and carbon dioxide and heating to

100°C) was repeated, but this time after adding 50 mg dppe.

5.4 Results

5.4.1 Modification of RU/AI2O3 with dppe and its Performance in the

Formylation of 3-Methoxypropylamine

Table 5-1 gives an overview of the catalytic results obtained during the

formylation under different reaction conditions. The dppe-modified

Ru/Al203 catalyst system showed good catalytic performance, whereas

runs without either catalyst or dppe (run 1-3) were unsuccessful. A highly

active catalyst was thus formed by modification of the Ru/Al203 surface

with dppe in a molar dppe:Ru ratio of 1:1. The addition of dppe before the

catalytic reaction under reflux in THF (premodification), or its addition

directly during the reaction itself led to similar catalytic performance.

Consequently, most of the experiments were carried out without pre¬

modification. Table 5-1 provides some information about the influence of

parameters like temperature, the amount of carbon dioxide, the initial

partial pressure of hydrogen, and the amounts of Ru/Al203 and dppe on the

conversion. As expected, the increase of the temperature led to a higher

reaction rate (run 4-6).

80 Chapter 5

Table 5-1. Catalytic formylation of 3-methoxypropylamine (mpa) with carbon dioxide

and hydrogen using dppe-modifiedRu/Al2Oß as catalyst0

Runi Amount

of dppe

Amount

ofC02

H2 initial

pressure

Temperature Time Conversion

/umol /g /bar /°C /h /%

lb 0 100 80 100 3 0

2 0 100 80 100 3 0

3b 50 100 80 100 3 0

4 50 25 80 80 3 18

5 50 25 80 100 3 24

6 50 25 80 120 3 35

7 50 50 80 100 3 23

8 50 100 80 100 3 48

9 50 25 40 100 3 15

10 50 25 140 100 3 38

11 50 25 180 100 3 39

12 14 100 80 100 20 16

13 25 100 80 100 20 47

14 37 100 80 100 20 67

15 50 100 80 100 20 82

16 100 100 80 100 20 63

a Conditions, if not otherwise stated: 100 mmol mpai, 50 |amol Ru.

b 0 |amol Ru.

The conversion was also affected by the amount of carbon dioxide added

(run 5, 7, 8). Between 50 and 100 g C02, the conversion increased about

two times. The best conversion (48%) was achieved with 100 g carbon

dioxide (run 8). Previous studies of the formylation of various amines

Formylation over Ru/Al203 Modified by Phosphines 81

[14, 15] under similar conditions showed that the reaction mixture is made

up of two distinguishable phases: A dense liquid-like phase containing

mainly the amine, the product formamide and water, and on top of that a

supercritical, more gas-like phase containing mainly hydrogen and the

carbon dioxide (note that the term supercritical is exactly defined only for

pure substances [26]). Increasing the initial hydrogen pressure between 40

and 140 bar, resulted in higher conversions (Table 5-1, run 5, 9 - 11). At

40 bar hydrogen pressure, 15% conversion was reached within 3 h, whereas

at 140 bar, 38% and at 180 bar, 39% conversion was achieved. The amount

of dppe added affected the reaction as well (Table 5-1, run 12-16).

Conversion increased with the addition of dppe up to 82%, the P/Ru-ratio

(mol/L Ru / mol/L P) in this optimal case was 2:1. When dppe was added

in a P/Ru-ratio of 4:1, the conversion decreased to 63%.

800

600

sz

400 OI-

D200

0

Figure 5-1. Comparison of the activity of the homogeneous complexes RuCE(dppe)2,

RuCÏ2(PPh3)3 with dppe-modified Ru/Al2Oß in 3-methoxypropylamine formylation.

Reaction conditions: 100 mmol mpa, 50 jumol Ru, 100 g CO2, 80 bar H2 initial partial

pressure, 3 h and 100°C.

82 Chapter 5

By way of comparison, related homogeneous ruthenium catalysts were

tested in the formylation reaction; results are shown in Figure 5-1. The

homogeneous complexes RuCl2(dppe)2 and RuCl2(PPli3)3 showed a

turnover frequency of 425 tr1 and 235 h-1, respectively, at 100%

selectivity. Dppe-modified Ru/Al203 showed a conversion of 48% under

the same reaction conditions. This conversion is higher than that observed

with the homogeneous catalyst RuCl2(PPli3)3, but lower than that observed

with RuCl2(dppe)2. The higher activity of RuCl2(dppe)2 compared with

RuCl2(PPfi3)3 is in line with earlier studies [14], where RuCl2(dppe)2 has

been found to be most active.

5.4.2 Identification of the Catalytically Active Species

Figure 5-2 shows the ex situ X-ray absorption spectra of the solid catalyst

under different conditions: 1) Ru/Al203 untreated, 2) Ru/Al203 treated

with dppe and amine before the reaction, and 3) dppe-modified Ru/Al203

after the reaction. A clear decrease in the whiteline at 22.12 keV is apparent

after step 3), which reveals that ruthenium constituent was partly reduced

during the reaction. No change in the EXAFS spectra was observed during

modification of the Ru/Al203 catalysts by treatment with amine and dppe

in advance. In the EXAFS region (Fourier-transformed data, not shown), a

clear change in the Ru-0 and the Ru-Ru backscattering was found.

However, ruthenium was not completely reduced.

To elucidate the role of possible corrosion of the Ru constituent and the

formation of active ruthenium complexes in solution under reaction

conditions in the presence of dppe, we performed appropriate tests that

allowed us to distinguish between heterogeneous and homogeneous

catalysis [27]. The reaction was stopped after a certain time, the solid

Formylation over Ru/Al203 Modified by Phosphines 83

catalyst was filtered off, and we inspected it in order to determine whether

the reaction could be continued in the filtered reaction mixture (Table 5-2).

Note that almost the same conversion was observed, irrespective of

whether the Ru/Al203 catalyst was filtered off after 3 h or whether it was

present during the whole reaction time. This implies that during the first 3 h

of the reaction, a highly active Ru-based homogeneous catalyst must have

been formed.

22 1 22 2 22 3

E/keV

22 4 22 5

Figure 5-2. XAFS spectra of 1) Ru/A^O^, compared to 2) dppe-modified Ru/A^O^

before reaction (pretreated with dppe and amine under reflux), and 3) dppe-modified

RU/AI2O3 after reaction.

In addition, the ruthenium and phosphorus contents in the liquid phase of

the samples were measured by ICP-OES. Table 5-3 gives an overview. The

ruthenium concentration in the liquid reaction mixture was about 50 ppm

(50 mg/kg, mass based) after 3 h (run 20) under standard reaction

conditions, as described above. The results show that after 20 h, 66 ppm

(0.689 mmol/L) ruthenium was dissolved in the liquid phase (run 21),

compared with 60 ppm (0.631 mmol/L) if the catalyst was filtered off after

3 h (run 22).

84 Chapter 5

Table 5-2. Test series of the formylation of 3-methoxypropylamine with dppe-modified

Ru/Al20ß to check the possibility oftheformation ofa homogeneous catalyst0

Run Catalytic system Time/h Conversion/%

17 dppe-modified Ru/Al203 3 48

18 Filtered reaction mixture of run 17 20b 78

19 dppe-modified Ru/Al203 20 82

a Conditions: 100 mmol mpa, 50 mmol Ru on A1203, 50 mmol dppe, 100 g C02, 80 bar H2 at

100°C. After 3 h (run 17), the solid catalyst was removed and the reaction was continued in the

filtered reaction mixture. The conversion after 20 h (run 18) was equal to the one observed with

the catalyst present during 20 h (run 19).b Catalyst was filtered off after 3 h.

The ruthenium concentrations were confirmed by X-ray absorption

spectroscopy (edge jump at the Ru K-edge). Application of UV-vis

spectroscopy was hampered by the very low Ru-concentration, which

pushed the detection capabilities of the technique to its limits. Note that the

ruthenium concentration in the liquid phase was very similar, irrespective

of whether the reaction was stopped after 3 h, or whether the Ru/Al203

catalyst was removed or remained present in the reactor for 20 h. The

background concentrations of ruthenium and phosphorus were significantly

lower during blank runs (Table 5-3, run 23, 24), and no conversion was

observed. If we relate the TOF to the amount of ruthenium dissolved in the

liquid phase, a very high rate of 3148 fr1 is estimated based on a ruthenium

concentration of 51 ppm in the reaction mixture. By way of comparison,

the Ru concentration of the homogeneous catalyst RuCl2(dppe)2 was also

determined. After a reaction time of 3 h, 133 ppm Ru (corresponding to

1.386 mmol/L in run 26) was found, which increased to 252 ppm

(2.577 mmol/L in run 27) after 20 h, reflecting the time dependence of the

solubility of the catalyst in the amine-rich liquid-like phase. Additionally,

Formylation over Ru/Al203 Modified by Phosphines 85

in the case of the homogeneous catalyst RuCl2(PPli3)3, a concentration of

385 ppm Ru and 375 ppm P (according to a P/Ru ratio of 3) was detected

(run 25). Note that the concentration of Ru was very small, but sufficient to

catalyze the formylation of mpa.

Table 5-3. Solubility of the in situformed catalystfrom Ru/Al2Oß and dppe in the liquid

phase, compared to RuCE(PPh3)3 andRuCE(dppe)2

Run Catalyst Reaction dataa Dissolved in reac

Ru P

;tion mixture

Time Conv. TOFb P/Ru

/h /% /h-1 /ppm /ppm /(mol/mol)

20 dppe-modified

Ru/Al203

3 48 3148 50.5 251 16.2

21 dppe-modified

Ru/Al203

20 82 559 66.4 256 12.6

22 dppe-modified

Ru/Al203

20 78 585 60.6 294 15.8

23c Blank test 3 0 0 0.60 21.8 118

24d dppe 3 0 0 1.89 268 463

25 RuCl2(PPh3)3 3 36 315 385 375 3

26 RuCl2(dppe)2 3 64 425 133 173 4.2

27 RuCl2(dppe)2 20 91 91 252 319 4.1

Conditions: 50 |amol Ru, 100 mmol mpa, 100 g C02, 80 bar H2 partial pressure, 100°C.

TOF based on Ru dissolved in reaction mixture.

0 mol Ru, 0 mol dppe.0 mol Ru, 50 |amol dppe.

86 Chapter 5

5.4.3 Spectroscopic Study of the In Situ Formation of the

Homogeneous Ruthenium Catalyst

To understand the formation of the homogeneous catalyst, the changes in

the solid Ru/Al203 catalyst and in the liquid phase were investigated. For

this purpose, the specially designed batch reactor cell - described in

chapter 2, section 2.2.2-was used. It allowed in situ EXAFS measure¬

ments both at the bottom (solid phase) and in the center (liquid) of the

batch reactor [25]. In a first step, the catalyst was investigated in the

presence of mpa, hydrogen, and carbon dioxide. The results are shown in

Figure 5-3. Figure 5-3a shows the treatment of the catalyst in the reaction

mixture without dppe. Figure 5-3b depicts the spectra taken during the

reaction in the presence of dppe.

22 2 22 3

E/keV

22 1 22 2 22 3 22 4 22 5

E/keV

Figure 5-3. a) Reduction of Ru/A^O^, monitored in situ by X-ray absorption

spectroscopy, solidphase 1) before reaction at 40°C as reference, and at 120°C after 2)

0 min, 3) 30 min, 4) 60 min, 5) 150 min reaction time, b) In situ monitoring during

reaction over Ru/A^O^, solidphase 1) before reaction at 40°C as reference, and with

dppe at 100°C after 2) 150 min, 3) 160 min, 4) 250 min. Conditions: 10 ml batch

reactor cell, 3 ml mpa, 70 bar H2, totalpressure 120 bar, rest CO2-

Formylation over Ru/Al203 Modified by Phosphines 87

With increasing temperature and in the absence of dppe (Figure 5-3a), Ru

was reduced, as indicated by the decrease in the whiteline at 22.12 keV.

The structural changes in ruthenium were limited to the solid phase, and no

Ru species were observed in the liquid phase.

After a reaction time of 4.5 h (without dppe), the concentration of

ruthenium in the liquid phase was still negligible. Only after an increase in

the stirring rate (Figure 5-4, spectrum 2) some Ru was visible in the X-ray

absorption spectrum. Part of the catalyst was probably suspended in the

upper part of the cell, and thus a small edge jump was observed. Note that

the XANES scans around the Ru K-edge were taken directly around the

edge only (spectra 2 and 3), and the absorption step was related to the

maximum intensity observed at the end of the experiment (Figure 5-4,

spectrum 4, absorption edge of 1.0).

E/keV

Figure 5-4. Solubility and formation of the catalytic active species, monitored by

XANES in liquidphase during reaction. 1) at 0 min, without dppe, 2) at 4.5 h, without

dppe, increased stirring rate, 120°C, 3) at 30 min, with dppe, 120°C, 4) after reaction,

with dppe, 25°C, 5) ex situ spectrum after reaction with dppe, 25°C. Spectra 1) - 4) are

calibrated to the normalized spectrum 4), simultaneously recorded to the data in

Figure 5-3.

88 Chapter 5

Therefore, during this first step no catalyst was dissolved in the liquid-like

amine-rich phase. However, as soon as dppe was added, an abrupt increase

in the Ru concentration and a further increase during heating was observed

(Figure 5-4, traces 3 and 4). During this step, almost no changes in the solid

Ru-catalyst (Figure 5-3B) could be noticed. The in situ formation of the

active species in the liquid-like amine-rich phase as a consequence of

adding dppe (Figure 5-4) can be related to the tendency of this ligand to

corrode the ruthenium and form Ru complexes. Supporting this view was

the fact that no catalytic activity was observed in the absence of dppe, and

no significant amount of Ru was found in the liquid solution with ICP-OES

(Table 5-3). Note that the in situ spectrum after the reaction (Figure 5-4,

spectrum 4) and the ex situ spectrum after the reaction (Figure 5-4, trace 5)

look quite similar. No substantial change could be observed while cooling

and depressurizing occurred.

5.4.4 Structural Identification of the Homogeneous Ruthenium

Complex

To gain insight into the structure of the active species formed in situ, ex situ

XANES and EXAFS spectra were recorded. Because of the low

concentration of 50-100 ppm in the reaction mixture, the liquid cell with

4 cm path length was used, as described in chapter 2, section 2.2.1. The top

of the cell was open in order to eliminate possible gas bubbles in the highly

viscous product mixture, and Kapton foils were used as X-ray transparent

windows on both sides of the liquid cell to keep the liquid within the beam.

Figure 5-5 compares the XANES region of the dppe-modified Ru/Al203

catalyst in the liquid product mixture with the homogeneous catalyst

RuCl2(dppe)2 in the liquid phase and as a solid (see corresponding catalytic

Formylation over Ru/Al203 Modified by Phosphines 89

results in Table 5-1). For structural identification, RuCl2(PPli3)3 in the

liquid reaction mixture was also investigated. The analytical results

indicate that the structure of the catalyst formed in situ from Ru/Al203 and

dppe is similar to the structure of RuCl2(dppe)2 under reaction conditions,

whereas the structure of the solid RuCl2(dppe)2 seems to be different,

probably as a consequence of the replacement of the chloride by hydrogen

and by the amine under reaction conditions.

— Ru/Al203 + dppe- RuCI2(dppe)2- RuCI2(PPh3)3— RuCI2(dppe)2, solid

10 22 12 22 14 22 16 22 18 22 20

E/keV

Figure 5-5. X-ray absorption near edge structure at the Ru K-edge of RuCE(dppe)2,

liquid product mixture (dotted black line), dppe-modified Ru/Al2Oß, liquid product

mixture (solid black line), RuCE(dppe)2, solid complex (dotted gray line), and

RuCÏ2(PPh3)3, liquid product mixture (solid gray line). Reference spectrum: solid

RuCÏ2(dppe)2-

Note that a change in the XANES region is also caused by the change in

multiple scattering paths, which, in addition, affect the near edge structure

[28, 29]. The variation of the XANES region of RuCl2(PPh3)3 indicates

that the structure of the catalyst formed in situ from Ru/Al203 and dppe

resembles more closely the structure of RuCl2(dppe)2 than that of

RuCl2(PPli3)3. Only a slight shift in the Ru K-edge is found, suggesting

that the oxidation states of the two complexes are similar [23].

90 Chapter 5

13

nj 0 0030-

^g 0 0025-

sz

.3> 0 0020-CD

-^ 0 0015-

-S 0 0010-13

D) 0 0005-

,_0 0000-

0 12 3 4 5

R/A

Figure 5-6. Comparison of the Fourier-transformed EXAFS spectra (k1 weighted) at

the Ru K-edge of RuCE(dppe)2 in reaction mixture from reaction with mpa (dotted

black line), dppe-modified Ru/A^O^ in reaction mixture (solid black line),

RuCE(dppe)2, solid phase (dotted gray line) and RuCEfPPhj)^, in reaction mixture

(solid gray line).

Figure 5-6 shows the corresponding ki-weighted Fourier-transformed

extended X-ray absorption fine structure (EXAFS) spectra. As in the case

of the near edge spectra, significant differences of the as-prepared solid

catalysts RuCl2(PPfi3)3 and RuCl2(dppe)2 and the corresponding Ru

complexes after reaction (liquid product mixture) are evident. The

backscattering is lower for both Ru-Cl and Ru-P, and the peak is shifted to

lower R-values (spectra not corrected for the phase shift). The spectra of

the catalysts exhibit backscattering peaks at 1.8 and 2.5 Â for all catalysts.

Whereas RuCl2(dppe)2 in reaction mixture and dppe-modified Ru/Al203

show similar backscattering from the nearest neighbor atoms, the solid

RuCl2(dppe)2 exhibits a large peak at 1.8 Â and only a very small one at

2.5 Â. The corresponding spectrum of RuCl2(PPfi3)3 is quite different

(Figure 5-6).

Formylation over Ru/Al203 Modified by Phosphines 91

_3

cc0 06-

, v

(1

ci> 0 04-

±.

m

ni 0 02-

5

'v 0 00-

CD-0 02-

-1-^

ni-0 04-

CD

2-0 06-

HLL

0

R/A R/A

Figure 5-7. Fourier-transformed and k1-weighted Ru K-edge EXAFS spectra (k^xfi))

of a) dppe-modifed Ru/A^O^ (20 h) and b) RuCE(dppe)2 after reaction: radial

distributionfunction together with the imaginary part, measured data (solid line), fitted

data (dotted line).

The Fourier-transformed spectra were fitted with Ru-Cl and Ru-P shells

calculated by FEFF 6.0 [30]; the results are listed in Table 5-4. Some of the

experimental and fitted data are compared in Figure 5-7. The structural data

obtained for the solid RuCl2(PPfi3)3 (entry 1 of Table 5-4) are in good

accordance with the literature [31]. For the solid homogeneous catalyst,

three different Ru-P bond lengths, 2.23 Â, 2.37 Â, and 2.41 Â were

reported. Only two Ru-P distances (2.23 Â and 2.38 Â, Table 5-4) could be

distinguished with EXAFS, and only one Ru-Cl bond length was detected

at 2.44 Â by EXAFS, a consequence of the limitations of the method used

in our studies. While the typical uncertainty is estimated to be ±0.02 Â for

the distance and about ±0.5 for the coordination number, the superposition

of two shells of the same backscatterer with similar distances, in particular,

could not be resolved satisfactorily. The results from the EXAFS spectra of

solid RuCl2(dppe)2 show that the different structure is appropriately

represented. Both Ru-P and Ru-Cl distances are found. The bond length of

Ru-Cl (2.42 À) in the solid RuCl2(dppe)2, calculated from the EXAFS fit,

92 Chapter 5

is in the range (2.436 Â) of the value published in literature [32]. In

addition, we found only one averaged Ru-P bond, as opposed to the two

different Ru-P bonds (2.389 À and 2.369 À) reported.

Table 5-4. Structural data extracted from EXAFS spectra of different ruthenium

phosphine complexes as solid and in liquidphase. Ru concentration in the liquidphase

~50 100 ppm, N: coordination number, R: distance of the corresponding neighbor,

Ao2: Debye-Wallerfactor

Catalyst Ru-P

N

Ru-Cl

R/Â Aa2/Â2

Residuald

N R/Â Aa2/Â2

RuCl2(PPh3)3a 1.0

2.0

2.23

2.38

0.0025

0.0025

2.3 2.44 0.0027 1.4

RuCl2(dppe)2a 4 2.29 0.0013 2 2.42 0.0006 2.0

RuCl2(PPh3)3b 4.1 2.22 0.0044 2 2.93 0.0060 7.2

RuCl2(dppe)2b 3.6 2.32 0.0063 - - 9.4

dppe-modified 3.3 2.35 0.0060 - - 3.2

Ru/Al203a Solid catalyst.b

Liquid solution measured after reaction with mpa.c Additional shell: Ru-N or Ru-O: N = 0.8, R = 2.07 Â, Aa2 = 0.006.

dQuality of the fit according to ref [24].

Significant structural changes were detected for the complexes formed

during the formylation reaction of mpa. After the use of RuCl2(PPfi3)3 in

the reaction, a Ru-P bond length of 2.22 Â, as well as an additional

contribution that could be fitted with a Ru-Ru distance at 2.93 Â, were

found, corresponding to the two peaks observed in Figure 5-6. This shows

that there are only slight changes in the Ru-P neighbors in liquid (after

reaction) or solid phase, but the Ru-Cl coordination changes. All bond

Formylation over Ru/Al203 Modified by Phosphines 93

lengths of the complexes RuCl2(dppm)2 [33], RuCl2(dppe)2 [32],

RuCl2(dppp)2 [34], RuBrCl(dppe)2 [35], and RuClH2(dppp)2PF6 [36]

reported in the literature are between 2.340 Â and 2.441 Â, depending on

the ligand used. Ru-Cl bond lengths range from 2.407 Â to 2.436 Â. In the

liquid phase, the bond length of Ru-P in RuCl2(dppe)2, 2.32 Â, is in the

expected range (the decrease may result from additional oxygen/nitrogen

neighbors in the complex), but no Ru-Cl bond is indicated in the spectra.

Thus, a change of the structure of RuCl2(dppe)2 in the liquid phase is

observed during the reaction. There are two plausible explanations for this

observation: The Cl-atoms could have been replaced by hydrogen to form a

hydrido complex [37-39], or a partial amino (or hydroxo-) complex was

formed, with hydrogen and mpa present in the liquid phase. Another reason

might be the similar Ru-P and Ru-Cl distances in the complexes formed.

However, the Ru-P distance is usually shorter than the Ru-Cl distance (see

references above and [40, 41]). Note also, that the catalyst was investigated

after air exposure. Finally, EXAFS analysis as an averaging technique has

its limitations when several species are involved.

The reaction mixture containing dppe-modified Ru/Al203 exhibited a Ru-P

bond length of 2.35 Â. This shows that a homogeneous Ru-dppe based

catalyst forms, as could already be concluded from the near edge region

and the Fourier-transformed EXAFS spectra. As expected, no Ru-Cl is

observed. The structure of this catalyst formed in situ seems to resemble

that of the observed RuCl2(dppe)2 in the liquid phase, but amino or oxygen

neighbors were also found in the first shell of the dppe-modified Ru/Al203

system. The EXAFS fit of the spectrum improved significantly when

nitrogen/oxygen neighbors (being neighbors in the periodic table, the two

atoms have similar backscattering amplitudes) were used additionally. The

94 Chapter 5

formation of such a complex is reasonable because the catalyst is dissolved

in the amine-rich phase during reaction.

5.5 Discussion

Ru/Al203 modified by dppe was found to be highly active for the

formylation of 3-methoxypropylamine (mpa), in conjunction with carbon

dioxide and hydrogen. The catalyst shows good activity and 100%

selectivity, comparable to the performance of the homogeneous ruthenium-

based phosphine complexes RuCl2((PCH3)3)4 and RuCl2(dppe)2. The

study indicates that the formylation of amines with hydrogen and carbon

dioxide can be extended to amines containing an ether group.

The reaction rates for mpa in terms of the TOF are comparable to or better

than those achieved with other primary amines, such as propylamine in the

presence of RuCl2(P(CH3)3)4 with 52 Ir1 [11]. Turnover frequencies of

secondary amines like diethylamine are between 440 h-l [3] and

360000 h"1 [13], depending on the catalyst used and the experimental

conditions.

Carbon dioxide acts as both reactant and solvent. C02 is not present as a

single supercritical phase, but exists in two phases, an amine-rich C02

phase with H2 dissolved, and a C02/H2 phase [14]. The volume of the

amine-rich C02 phase with dissolved H2 increases with the amount of

carbon dioxide in the system. Higher partial pressure of hydrogen affects

the dissolved amount of H2 in the amine-rich C02 phase and results -

together with the higher density - in a better mixing of the reactants in the

amine-rich C02 phase and, ultimately, in better conversion and TOF.

Interestingly, it was found that dppe-modified Ru/Al203 showed good

catalytic performance in the formation of 3-methoxypropylformamide,

Formylation over Ru/Al203 Modified by Phosphines 95

similar to the homogeneous catalysts RuCl2(dppe)2 and RuCl2(PPli3)3.

This behavior can be traced to the formation of active free Ru complexes as

a result of the interaction of the supported Ru particles with the strongly

complexing phosphine (dppe). This interaction leads to some corrosion of

the Ru particles.

The formation of a soluble complex during reaction could be proven in

different ways. On the one hand, X-ray absorption spectroscopy and ICP-

OES showed the presence of dissolved Ru in the product mixture. On the

other hand, the reaction could be resumed if the solid material was filtered

off after a certain reaction time (Table 5-2). Only a small fraction of the

ruthenium is dissolved, resulting in a concentration of 50-100 ppm in the

product mixture. This shows that the homogeneous Ru-based catalyst

formed is highly active (TOF based on the amount of dissolved Ru is about

3100 h-l, Table 5-3, run 20).

The formation of the ruthenium complex, which is acting as homogeneous

catalyst, could be monitored in situ by X-ray absorption spectroscopy. In

principle, there are only a few techniques, such as EXAFS spectroscopy,

UV-vis spectroscopy, NMR spectroscopy, and infrared spectroscopy that

render such studies feasible [42-46]. Because of the low concentration of

the homogeneous catalyst, structural studies using UV-vis were not

successful. The concentration of the target species was also too low for

NMR and IR spectroscopy. On the other hand, owing to the good

penetration capability of X-rays at 20 keV, EXAFS studies in the

transmission mode could be accomplished. The advantage of X-ray

absorption spectroscopy lies in the fact that the concentration and the

structure of the Ru species can be determined element-specifically and

under reaction conditions. Such studies are rarely done under such harsh

96 Chapter 5

conditions (150 bar, 120°C) as applied here and have only recently been

reported for solid [47] and homogeneous catalysts [48, 49]. In our case, we

found that the X-ray absorption technique, combined with an appropriately

high-pressure cell, allowed the simultaneous monitoring of the solid

catalyst at the bottom, and identification of the in situ formed complex in

the liquid phase in the upper part of the reactor.

During reaction, the ruthenium constituent in the solid catalyst is reduced.

However, ruthenium can only be found in the liquid phase if dppe is added

to the system. This indicates corrosion of the ruthenium particles in the

presence of hydrogen and dppe. Presumably it is the structure of the dppe

ligand helps in forming a highly stable chelate complex, favoring the

corrosion of the Ru and thus the in situ formation of the Ru complex.

Such an in situ formation of a homogeneous catalyst from a supported

metal catalyst has only rarely been reported. Köhler et al. recently

described the in situ formation of a Pd catalyst in Heck reactions [50]. In

this case, ligands/reactants in the solution led to the formation of an active

homogeneous catalyst. This example emphasizes the necessity for

appropriate in situ spectroscopic tools to help decide between

homogeneous complexes and surface modification on metal particles as

active species for catalysis.

The amount of dissolved ruthenium in the liquid phase was quantified by

ICP-OES and confirmed by XAS. The concentration of ruthenium was very

low, in the range of 50-100 ppm, and no redeposition as reported in ref.

[50] was observed. The TOF (3148 fr1) related to the dissolved ruthenium

species turned out to be significantly higher than the TOF (324 Ir1) based

on the total amount of ruthenium present in the reactor. We can conclude

that the ruthenium species formed in situ has a higher intrinsic activity than

Formylation over Ru/Al203 Modified by Phosphines 97

the homogeneous catalyst RuCl2(dppe)2. The reason for the better activity

may be found in the different structures of dissolved RuCl2(dppe)2 and the

species formed in situ from Ru/Al203 and dppe. Furthermore, the solubility

of RuCl2(dppe)2 in the reaction mixture may affect its catalytic

performance.

Structural analysis by EXAFS spectroscopy of the Ru complex formed

indicates that a Ru(dppe)2X2 catalyst is formed that may contain either

additional hydrido- or amino/hydroxo-ligands. The formation of a hydrido

complex was previously indicated in other studies [37-39, 51], and the

mechanism for the Ru-catalyzed formylation of amines proposed by Jessop

et al. [11] can be further confirmed. Based on these findings, we can

assume that a hydrido complex is probably formed from Ru/Al203 and

dppe in the liquid phase under reaction conditions. C02 is activated by

insertion into the metal-hydrogen bonding of the hydrido complex. In the

presence of mpa, the complex keeps reacting to form the product

3-methoxypropylformamide, and water is formed as a by-product.

The present study shows that the modification of supported transition-metal

catalysts by a phosphine may corrode the metal and result in free metal

phosphine complexes, disguising the intrinsic catalytic properties of the

modified supported metal catalyst. Similar behavior has also been observed

when another phosphine, triphenylphosphine, was applied. Thus it seems to

be of paramount importance to test for this possibility when supported

metal catalysts are applied in conjunction with phosphines.

98 Chapter 5

5.6 Conclusions

Modification of Ru/Al203 by the addition of a phosphine (dppe) has been

shown to result in homogeneous Ru catalysts that are highly active in the

formylation of 3-methoxypropylamine with carbon dioxide and hydrogen.

The catalytically active species were not the phosphine modified ruthenium

particles, but rather a homogeneous chlorine-free Ru-dppe complex that

formed under reaction conditions. The presence of dppe led to the

dissolution (corrosion) of ruthenium and probably to the formation of a

highly active homogeneous Ru(dppe)2X2 complex, which was sufficiently

active to achieve high conversion even at very small concentrations

(50 ppm Ru in reaction mixture).

X-ray absorption spectroscopy proved to be a valuable technique for

identifying the role of solid and liquid Ru species in the catalytic reaction.

The formation of the homogeneous catalyst could be monitored under

reaction conditions, and the structural identification of the homogeneous

Ru-dppe complex by transmission EXAFS was feasible down to a

concentration of 100 ppm and less in the product mixture.

5.7 References

[1] M. Halmann, Chemical Fixation of Carbon Dioxide: Methods for

Recycling CO2 into Useful Products, CRC Press, Boca Raton,

Florida (1993).

[2] W. Leitner, Angew. Chem. Int. Ed. Engl. 34 (1995) 2207.

[3] P. G. Jessop, T. Ikariya, R. Noyori, Chem. Rev. 95 (1995) 259.

[4] M. Aresta, E. Quaranta, Chemtech (1997) 32.

[5] A. Baiker, Appl. Organomet. Chem. 14 (2000) 751.

Formylation over Ru/Al203 Modified by Phosphines 99

6] A. Behr, Angew. Chem. Int. Ed. Engl. 27 (1988) 661.

7] P. G. Jessop, T. Ikariya, R. Noyori, Nature 368 (1994) 231.

8] O. Kröcher, R. A. Koppel, A. Baiker, Chimia 51 (1997) 48.

9] P. G. Jessop, W. Leitner, Chemical Synthesis Using Supercritical

Fluids, Wiley-VCH, Weinheim (1999).

10] P. G. Jessop, Y. Hsiao, T. Ikariya, R. Noyori, J. Am. Chem. Soc. 116

(1994)8851.

11] P. G. Jessop, Y. Hsiao, T. Ikariya, R. Noyori, J. Am. Chem. Soc. 118

(1996)344.

12] S. Schreiner, J. Y. Yu, L. Vaska, Inorg. Chim. Acta 147 (1988) 139.

13] O. Kröcher, R. A. Koppel, A. Baiker, Chem. Commun. 5 (1997) 453.

14] L. Schmid, A. Canonica, A. Baiker, Appl. Catal. A: Gen. 255 (2003)

23.

15] L. Schmid, M. S. Schneider, D. Engel, A. Baiker, Catal. Lett. 88

(2003) 105.

16] C.-C. Tai, J. Pitts, J. C. Linehan, A. D. Main, P. Munshi, P. G.

Jessop, Inorg. Chem. 41 (2002) 1606.

17] O. Kröcher, R. A. Koppel, A. Baiker, J. Chem. Soc, Chem.

Commun. (1996) 1497.

18] L. Schmid, M. Rohr, A. Baiker, Chem. Commun. (1999) 2303.

19] T. Mallat, A. Baiker, Appl. Catal. A: Gen. 200 (2000) 3.

20] T. Mallat, C. Brönnimann, A. Baiker, Appl. Catal. A: Gen. 149

(1997)103.

100 Chapter 5

[21

[22

[23

[24

[25

[26

[27

[28

[29

[30

[31

[32

[33

[34

R. Mason, D. W. Meek, G. R. Scollary, Inorg. Chimia Acta 16

(1976) LU.

O. Kröcher, R. A. Koppel, A. Baiker, J. Mol. Catal. A: Chem. 140

(1999) 185.

K. Okamoto, T. Takahashi, K. Kohdate, H. Kondoh, T. Yokoyama,

T. Ohta, J. Synchrotron Rad. 8 (2001) 689.

T. Ressler, J. Synchrotron Rad. 5 (1998) 118.

J.-D. Grunwaldt, M. Ramin, M. Rohr, A. Michailovski, G. R. Patzke,

A. Baiker, Rev. Sei. Instrum. 76 (2005) 054104.

R. Wandeler, N. Künzle, M. S. Schneider, T. Mallat, A. Baiker, J.

Catal. 200 (2000) 377.

R. A. Sheldon, M. Wallau, I. W. C. E. Arends, U. Schuchardt, Ace.

Chem. Res. 31(1998)485.

O. Sipr, G. Dalba, F. Rocca, Phys. Rev. B. 69 (2004) 134201.

D. Bazin, J. J. Rehr, J. Phys. Chem. B 107 (2003) 12398.

S. I. Zabinsky, J. J. Rehr, A. Ankudinov, R. C. Albers, M. J. Eller,

Phys. Rev. B. 52 (1995) 2995.

S. J. LaPlaca, J. A. Ibers, Inorg. Chem. 4 (1965) 778.

T. S. Lobana, R. Singh, E. R. T. Tiekink, J. Coord. Chem. 21 (1990)

225.

A. R. Chakravarty, F. A. Cotton, W. Schwotzer, Inorg. Chim. Acta

84(1984)179.

M. R. M. Fontes, G. Oliva, L. A. C. Cordeiro, A. A. Batista, J.

Coord. Chem. 30 (1993) 125.

[35

[36

[37

[38

[39

[40

[41

[42

[43

Formylation over Ru/Al203 Modified by Phosphines 101

E. P. Perez, R. M. Carvalho, R. H. A. Santos, M. T. P. Gambardella,

B. S. Lima-Neto, Polyhedron 22 (2003) 3289.

B. Chin, A. J. Lough, R. H. Morris, C. T. Schweitzer, C. D'Agostino,

Inorg. Chem. 33 (1994) 6278.

M. K. Whittlesey, R. N. Perutz, M. H. Moore, Organometallics 15

(1996)5166.

M. G. Basallote, J. Duran, Inorg. Chem. 38 (1999) 5067.

K. A. Lenero, M. Kranenburg, Inorg. Chem. 42 (2003) 2859.

E. Lindner, S. Al-Gharabli, I. Warad, H. A. Mayer, S. Steinbrecher,

E. Plies, M. Seiler, H. Bertagnolli, Z. Anorg. Allg. Chem. 629 (2003)

161.

O. Kröcher, R. A. Koppel, M. Fröba, A. Baiker, J. Catal. 178 (1998)

284.

M. Hunger, J. W. Weitkamp, Angew. Chem. Int. Ed. Engl. 40 (2001)

2954.

D. Bazin, H. Dexpert, J. Lynch, X-ray Absorption Fine Structure for

Catalysts and Surfaces, World Scientific, Singapore 2 (1996).

[44] G. Sankar, J. M. Thomas, Topics Catal. 8 (1999) 1.

[45] J.-D. Grunwaldt, B. S. Clausen, Topics Catal. 18 (2002) 37.

[46] J.-D. Grunwaldt, R. Wandeler, A. Baiker, Catal. Rev. -Sei. Eng. 45

(2003)1.

[47] J.-D. Grunwaldt, M. Caravati, M. Ramin, A. Baiker, Catal. Lett. 90

(2003)221.

102 Chapter 5

[48] S. L. Wallen, D. M. Pfund, J. L. Fulton, C. R. Yonker, M. Newville,

Y. Ma, Rev. Sei. Instrum. 67 (1996) 2843.

[49] S. G. Fiddy, J. Evans, T. Neisius, X. Z. Sun, Z. Jie, M. W. George,

Chem. Commun. 6 (2004) 676.

[50] R. G. Heidenreich, J. G. E. Krauter, J. Pietsch, K. Köhler, J. Mol.

Catal. A: Chem. 182-183 (2002) 499.

[51] L. Costella, A. Del Zotto, A. Mezzetti, E. Zangrando, P. Rigo, J.

Chem. Soc. Dalton Trans. (1993) 3001.

Chapter

Solvent-free Ruthenium-catalyzed Vinyl

Carbamate Synthesis from Phenylacetylene

and Diethylamine in Supercritical

Carbon Dioxide

6.1 Summary

As shown in the previous chapters, formylation reactions of amines with

hydrogen and carbon dioxide in the presence of ruthenium catalysts exhibit

a promising route of using carbon dioxide as a reactant and as a solvent. In

addition, other reactions - for example the synthesis of vinyl carbamate

from phenylacetylene, diethylamine, and carbon dioxide - show the

immense potential of using C02 as a raw material. This is the subject of the

present chapter. An environmentally friendly route of vinyl carbamate

synthesis, which is not only based on carbon dioxide as substrate, but also

on using supercritical carbon dioxide as the solvent is described. The

results show that the new method also provides enhanced reaction rates

compared to organic solvent-based systems.

6

104 Chapter 6

6.2 Introduction

The discovery that metal vinylidene complexes are readily available via the

reaction of late transition metal complexes with terminal alkynes spawned

a new field of catalytic chemistry utilizing this functionality [1]. Among

these transformations, the ruthenium catalyzed synthesis of vinyl

carbamates from carbon dioxide, secondary amines, and terminal

acetylenes has proven particularly interesting (Scheme 6-1) [2]. Vinyl

carbamates are versatile synthetic intermediates used in the pharmaceutical

and agrochemical industries [3]. In the past, all preparative routes to this

class of compounds have involved the highly toxic chemical phosgene [4].

o

Et2N^(OHC= CHPh

(1a) (E)

Ph-C = CH+HNEt +scCO, + (1b) (z)

O

EtN^ //CH2

Ph

C

(2)

CI

[Ru] = RuCI, xH20 PyvRiTPy Ru(C6H6)(PMe,)CI2

py'^py

(3) (4) (5)

Scheme 6-1. Vinyl carbamate synthesis from phenylacetylene, diethylamine and

supercritical carbon dioxide with ruthenium catalysts [Ru].

In the late 80's, Dixneuf et al. reported the homogeneous ruthenium(II)-

catalyzed synthesis of vinyl carbamates from secondary amines, a terminal

alkyne and carbon dioxide in organic solvents, effectively replacing

phosgene with non-toxic carbon dioxide [5]. The high natural abundance

Solvent-free Ru-catalyzed Vinyl Carbamate Synthesis 105

and ease of handling of carbon dioxide has prompted considerable interest

in its utilization in chemical processes [6-8]. Recent developments in this

field demonstrate that carbon dioxide is a useful reactant which can

simultaneously be used as a reaction medium [9-11]. As shown in the

previous chapters, C02 as a raw material is successfully used in the

formylation of amines, such as 3-methoxypropylamine [12, 13],

morpholine [14, 15], or diethylamine [16]. In these reactions, ruthenium

complexes were investigated regarding to their catalytic activity.

Specifically, RuCl2(dppe)2 as a homogeneous catalyst, in situ generated

catalysts from RUCI3 and phosphines (chapter 4) or from Ru/Al203 and

dppe (chapter 5), and heterogenized silica-based hybrid catalysts

RuCl2(bspe)2/Si02 and RuCl2(bspp)2/Si02 (see chapter 7) have turned out

as highly active catalysts. In the presence of similar ruthenium complexes,

the catalytic vinyl carbamate (1) synthesis (Scheme 6-1) is investigated

with respect to the use of carbon dioxide as a source of carbon and in the

absence of any other solvent, simplifying product and catalyst separation.

6.3 Experimental

Catalytic tests were run in a 250 ml stainless steel autoclave similarly

configured as the batch reactor described in chapter 2, section 2.1.1. The

reactor was charged with phenylacetylene (10 mmol), diethylamine

(80 mmol) and catalyst (0.036 mmol, 0.027 mmol and 0.018 mmol for

catalysts (3), (4) and (5), respectively). Carbon dioxide was poured into the

reactor using a compressor. The catalysts RuCl2(C5H5N)4 (4) [17] and

RuCl2( 776-CgH6)PMe3 (5) [18] were prepared according to literature

methods, and RUCI3 • x H20 (3) was used as supplied by ABCR,

(Karlsruhe, Germany). /?-[(diethylcarbamoyl)oxy]styrene (1) was isolated

106 Chapter 6

and identified by comparison with available spectroscopic data [5]. After

the reaction, the autoclave was cooled to room temperature and vented.

Reaction samples were dissolved in toluene with pentadecane added as the

standard. Analysis was performed on a HP-6890 gas Chromatograph

equipped with a FID detector and HP-5 column (length 30.0 m, diameter

320 urn, film thickness 0.25 urn). Turnover frequencies (TOFs) were

calculated as [moles (l)/(moles catalyst x time)]. Phase behavior of the

reaction was studied in a computer-controlled high-pressure variable

volume (23 -63 ml) view-cell [19] equipped with an online digital video

camera. The magnetically stirred cell comprised a horizontal cylinder fitted

with a sapphire window covering the entire diameter, opposite a

horizontally moving piston equipped with an illuminated sapphire window.

The computer-based approach with video imaging has been described

elsewhere [20]. For constructing details, see chapter 2, section 2.1.3.

6.4 Results and Discussion

A comparison of conversion and selectivity for RuCl2(C5H5N)4 (4) and

RuCl2( r^-CßüßjPMQ^ (5) catalyzed vinyl carbamate (1) synthesis in

supercritical C02 and toluene is presented in Figure 6-la and 6-lb. In

general, conversions are significantly higher in supercritical C02 than in

toluene (Figure 6-la), with RUCI3 • x H20 (3) (not shown) and (4) results

being similar, and (5) displaying the highest activity. The corresponding

TOFs reflect this behavior and are approximately three times higher in

supercritical C02 (92 fr1) than in toluene (31 Ir1), as demonstrated for (5)

at 90°C and 90 bar. For catalysts (4) and (5), conversion increased

significantly from 60 up to 120°C at 90 bar, whereas selectivity reached a

maximum at 90°C. The highest conversion was reached at a total pressure

Solvent-free Ru-catalyzed Vinyl Carbamate Synthesis 107

of 90 bar and 120°C. For (5), at a constant temperature of 90°C, selectivity

increased, whereas conversion was only slightly affected by increasing

pressure. Favorably, reaction conditions (90°C, 90 bar) for selectivity in

ruthenium-catalyzed vinyl carbamate (1) synthesis are similar for the

reaction in supercritical C02 and toluene. Previously used solvents include

toluene, THF and acetonitrile [5]. Toluene was selected here for

comparison purposes based on the similarity of solvent properties with

those of supercritical C02.

a) 120 bar

90 bar

100-

80-

3e

.1 60-

o 40-O

SO tar

20^

1

b)

100-

80-

k? 60 -

jt

o

S

g 40-

20-

126 bar

iObar

60 bar

«aC 80 "C 1»*C

Iwc MX 120 "C

Figure 6-1. Influence of temperature and pressure on phenylacetylene a) conversion

and b) selectivity to vinyl carbamate (1). Data for catalysts (4) and (5) are shown in

light and dark gray, respectively. Experiments with toluene (50 ml) as solvent are

depicted with striped bars, all other bars represent experiments in scCO^ The reaction

time was 3 h in all cases.

The positive influence of supercritical C02 as reaction medium manifests

itself as a considerable increase in conversion. For catalyst (5), 95%

conversion is achieved in just 3 h at 120°C and 90 bar as opposed to 20 h

required in earlier work [5]. The increased catalytic activity in supercritical

C02 is not unprecedented. A similar behavior was observed in the

108 Chapter 6

ruthenium-catalyzed synthesis of Af/V-diethylformamide, where TOFs were

also much greater in supercritical C02 [16].

The Markovnikov product (2) (Scheme 6-1) was formed in negligible

amounts in agreement with Dixneufs work [5]. Z/E ratios for vinyl

carbamate (1) were in the range of 3.8-8.3:1 with a mean of 6:1, as

compared to 4:1 reported previously [5]. Interestingly, at 90°C and 90 bar

the Z/E ratios were considerably different for (4) and (5), being higher for

the latter. The catalysts also showed opposite behaviors; for example, at

90 bar the Z/E ratio increased for (4) and decreased for (5) with increasing

temperature. Therefore, it appears that although (4) is active for the

catalytic synthesis of (1), the reaction may proceed on a slightly different

pathway than is the case for (5). In general, increasing the amount of base

to about 100 mmol had no significant effect on Z/E ratios or conversion.

Increasing the amount further led to reduced conversions, probably due to

competitive attack of the base and reactants at the metal center. Under

similar conditions, RuCl2(PPli3)3 - used also as a reference catalyst in the

formylation of 3-methoxypropylamine in chapter 4 and 5 - displayed signi¬

ficantly lower conversion and selectivity than (4) and (5). This may be a

consequence of the sterically more demanding phosphine substituents

preventing access of the reactant to the catalytically active center.

Vinyl carbamate synthesis and phenylacetylene polymerization share the

same ruthenium-vinylidene intermediate. Therefore, selectivity to (1)

depends at least partly on the relative rates of nucleophilic attack of the

carbamate anion (Et2NC02~) and phenylacetylene incorporation at the a-

carbon of the vinylidene intermediate [2]. Indeed, for catalysts (4) and (5)

the selectivity was < 10% with regard to the acetylene insertion products

l,4-diphenyl-3-buten-l-yne and l,l'-(l-butan-3-yne-l,4-diyl)dibenzene.

Solvent-free Ru-catalyzed Vinyl Carbamate Synthesis 109

Therefore, it seems likely that higher molecular weight polymers, which

are not detected by gas chromatography, constituted the remaining by¬

products, as evidenced by the formation of viscous reaction residues.

Figure 6-2. Video images of the reaction mixture a) at the beginning and b) after 2 h of

the vinyl carbamate (1) synthesis in supercritical carbon dioxide at 90 bar and 90°C.

The increase in volume of the liquidphase is highlighted and shown schematically next

to the images. The illuminated sapphire window manifests itselfas the bright round spot

at the far end of the cell. The outer bright circle is caused by reflection of light at the

outerflange of the sapphire window. The vertical thermocouple and magnetic stirrer at

the bottom ofthe cell are also visible.

Examination of the reaction mixture using a custom-built view-cell [19]

combined with video imaging [20] revealed a complex phase behavior

(chapter 2, section 2.1.3). At the beginning of the reaction two phases were

present: a small amount of a liquid phenylacetylene/diethylamine phase

saturated in C02 at the reactor base, and an upper supercritical C02-rich

phase (Figure 6-2a). Note that the term supercritical describes the dense

C02-rich phase at temperatures beyond its critical point, irrespective of any

other liquid phases present in the system. As the reaction proceeds (at about

2 h), the liquid phase increases in volume at the bottom of the reactor

110 Chapter 6

(Figure 6-2b). This increase can be traced to the lower volatility of the

vinyl carbamate (1) leading to product condensation.

6.5 Conclusions

In conclusion, considerably higher reaction rates are observed in the

solvent-free supercritical C02 system compared to toluene, even though the

reaction conditions have not been optimized. Further improvements of the

reaction performance may be achieved by increasing the density of

supercritical C02, and - as a result - the solubility of reactants and

catalysts in the latter, eventually leading to a supercritical single-phase

system. The present synthetic route to vinyl carbamates abandons the use

of phosgene and solvents. It is another example for the use of carbon

dioxide as a reactant and a solvent, in addition to the formylation reactions

discussed in the chapters 4, 5, and 7 of this work.

The catalysts RuCl2(C5H5N)4 (4) and RuCl2(/7^-C6H6)PMe3 (5) used in

these experiments are highly active and selective. In the previous chapters,

the effort to simplify and understand the ruthenium catalysts for the

formylation of amines has been demonstrated. It could be applied in a

similar way to the present catalyst/reaction system.

6.6 References

[1] M. I. Bruce, Chem. Rev. 91 (1991) 197.

[2] C. Bruneau, P. H. Dixneuf, Ace. Chem. Res. 32 (1999) 311.

[3] R. J. Khur, H. W. Dorough, Carbamate Insectizides Chemistry,

Biochemistry and Toxicology, CRC Press, Cleveland, Ohio (1976).

[4] P. J. Stang, G. H. Anderson, J. Org. Chem. 46 (1981) 4585.

Solvent-free Ru-catalyzed Vinyl Carbamate Synthesis 111

[5] R. Mahé, Y. Sasaki, C. Bruneau, P. H. Dixneuf, J. Org. Chem. 54

(1989)1518.

[6] A. Baiker, Appl. Organomet. Chem. 14 (2000) 751.

[7] A. Behr, Angew. Chem. Int. Ed. Engl. 27 (1988) 661.

[8] M. Aresta, Quim. Nova 22 (1999) 269.

[9] P. T. Anastas, T. C. Williamson, ACS Symposium Series 626,

American Chemical Society, Washington, DC (1996).

[10] P. G. Jessop, T. Ikariya, R. Noyori, Chem. Rev. 99 (1999) 475.

[11] A. Baiker, Chem. Rev. 99 (1999) 453.

[12] M. Rohr, J.-D. Grunwaldt, A. Baiker, J. Mol. Catal. A: Chem. 226

(2005)253.

[13] M. Rohr, J.-D. Grunwaldt, A. Baiker, J. Catal. 229 (2005) 144.

[14] G. Süss-Fink, M. Langenbahn, T. Jenke, J. Organomet. Chem. 368

(1989)103.

[15] L. Schmid, M. S. Schneider, D. Engel, A. Baiker, Catal. Lett. 88

(2003) 105.

[16] L. Schmid, M. Rohr, A. Baiker, Chem. Commun. (1999) 2303.

[17] N. S. Al-Zamil, E. H. M. Evans, R. D. Gillard, D. W. James, T. E.

Jenkins, R. J. Lancashire, P. A. Williams, Polyhedron 1 (1982) 525.

[18] M. A. Bennett, M. L. Bruce, T. W. Matheson, Comprehensive

Organometallic Chemistry, G. Wilkinson, F. G. A. Stone, E. W.

Abel, Pergamon, Oxford, UK (1982) 796.

112 Chapter 6

[19] R. Wandeler, N. Künzle, M. S. Schneider, T. Mallat, A. Baiker, J.

Catal. 200 (2000) 377.

[20] R. Wandeler, A. Baiker, Chimia 53 (1999) 566.

Chapter

Evaluation of Strategies for the Immobilization

of Bidentate Ruthenium Phosphine Complexes

Used for the Formylation ofAmines in

Supercritical Carbon Dioxide

7.1 Summary

Different routes for the immobilization of highly active bidentate

ruthenium-phosphine complexes (RuCl2(dppe)2 and RuCl2(dppp)2) on

silica have been explored, and the resulting heterogeneous catalysts were

evaluated with regard to activity and stability when applied to the

formylation of functionalized amines, such as 3-methoxypropylamine, in

supercritical carbon dioxide. Two fundamentally different immobilization

strategies were applied: covalent linkage and coordinative anchoring of the

active complex. The latter immobilization method resulted in strong

leaching of the complex under reaction conditions. Structural and textural

properties of the heterogeneous catalysts were characterized using various

methods, including X-ray absorption spectroscopy (XAS), X-ray

photoelectron spectroscopy (XPS), diffuse reflectance infrared Fourier

transform spectroscopy (DRIFTS), and inductively coupled plasma optical

emission spectroscopy (ICP-OES). Particular attention was given to a

7

114 Chapter 7

thorough characterization of the catalysts when subjected to reuse. XAS in

fluorescence and transmission mode provided deeper insight into the

observed deactivation phenomena. Despite the low concentration of Ru in

solution (ca. 50 ppm) the structural changes could be resolved by both

detection modes.

7.2 Introduction

Because of its large scale availability, low cost, and low toxicity, carbon

dioxide represents an interesting starting material and reaction medium for

the production of base and fine chemicals, especially under supercritical

conditions [1-7]. As one of the first reactions, the formylation of

dimethylamine to dimethylformamide using carbon dioxide as reactant and

solvent gained notable attention [2, 3, 8-13]. In recent years, more amines

such as propylamine [13], diethylamine [13, 14], morpholine [15, 16], and

3-methoxypropylamine [17] have been investigated.

In the synthesis of formamides by formylation of amines using carbon

dioxide in organic solvents or dense carbon dioxide, various group VIII

transition metal complexes were found to be efficient homogeneous

catalysts [2-4,8-12]. Among them, ruthenium-based systems with

phosphine ligands proved to be most promising [4, 8, 9]. Moreover, the use

of supercritical carbon dioxide instead of organic solvents leads to a

"solventless" process [2, 13, 18]. RuCl2(dppe)2 with the bidentate

phosphine ligand l,2-bis-(diphenylphosphino)ethane (dppe) has been

reported to be one of the best homogeneous catalysts in the solvent-free

formylation of dimethylamine using carbon dioxide as Ci -building block

[19, 20]; see also chapter 4 and chapter 5. A corresponding heterogeneous

catalyst is therefore an interesting target since the advantages of a

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 115

heterogeneous catalyst (stability, separation and reuse) could be combined

with the excellent catalytic behavior of the homogeneous complex. One of

the most promising routes towards heterogeneous catalysts is the

immobilization of highly active and selective homogeneous transition

metal complexes on silica.

C X )P CI P

Ph2 Ph2

homogeneous

RuCI2(dppe)2(1)

—Sr

"Si-CX/

O'

o

o^Sl-I

Sk„

o

'OI „OHSl „-

O'

HOx !

-Ph E^x/Sl

o

-Si-O

\

\-Si-

oI „o.

"SiI

Si

,R. ci p:

Ru

k / ! \

/\ „,P CI P^ ^\Sl ^-^Ph Ph"-^ „Si

I OH HO !

.0 O.

O

OI „O

-Suo

O'

.Ox I

SiI

„o'Si

\/-Si—

o

lxn-Sr

.O-Si

•o\

I/

Sl-o.O-Si-

/

/

o

immobilized RuCI2(bspe)2 (2)

Ph, Ph,

P. 2CI P.2

f Rù

V*. p-

Ph, Ph.

CI P.

CI"

O' I "O-Si-0

-Si

/ ô °— Si-O

/b\

.0/

adsorbed RuCI2(dppe)2on ammopropyl-modifiedsilica (3)

Ph, Ph,P. CI P^\vRu

/ I \

P CI P

Ph2 Ph2

—-Sro

.Sko o

Si-O^ I I „OHSk „Sl

"O ^-^!o

-kO

-Si-O

\

\-Si-

'O

Si SkI I OH

O^ „OSi

O

Ph

Px CI P.

\yRu

/ ! \

P CI p.Ph Ph

o' \/I „O-Si—

.Sk Ioo L

HO^ I I „0SiPiv\. „Sk „Skp ^-^ O O

\

O'i:

„-\ „Ovl„0-Sr„Si Si /

HO I I'

Si /.

Si

I'

oo-

\

RuCI, + dppe

NH2

.SkO I O-Si-0

-Si

/ ô °--Si-O

/b\

/

homogeneous

RuCI2(dppp)2 (4)

immobilized RuCI2(bspp)2 (5) in situ generated catalyst (6)from "Si"-NH2-RuCI3 + dppe

Figure 7-1. Structure of the homogeneous RuCE(dppe)2 catalyst (1) and the

immobilized catalysts RuCÏ2(bspe)2/Si02 (2), RuCE(dppe)2 adsorbed on aminopropyl-

modified silica (3), the homogeneous catalyst RuCE(dppp)2 (4) with the immobilized

RuCl2(bspp)2/Si02 (5), and the in situ prepared catalyst from "Si "-NH2-RUCI3 and

dppe (6).

116 Chapter 7

Both adsorbed Ru complexes on a modified silica surface [21] and the

covalent anchoring of ruthenium-phosphine complexes have been proposed

[14,22]. A crucial property of these "heterogenized" homogeneous

catalysts is their stability with repetitive use. Considering these points, we

critically evaluated the two different strategies for the immobilization of

bidentate ruthenium-phosphine complexes on silica: covalent bonding via

functionalized ligands using RuCl2(bspe)2 and RuCl2(bspp)2, and

coordinative anchoring of RuCl2(dppe)2 and analogs.

The different catalysts investigated are depicted in Figure 7-1. Various

methods, including X-ray absorption spectroscopy, XPS, nitrogen

adsorption, DRIFTS, and ICP-OES were applied in order to characterize

the catalysts and to uncover the structural changes the catalysts underwent

with repetitive use.

7.3 Experimental

7.3.1 Preparation of the Covalently Bound RuCl2(bspe)2 Catalyst

All preparations were carried out under oxygen-free conditions employing

usual Schlenk techniques. Solvents were used in spectrophotometric purity,

degassed, and stored under argon. Scheme 7-1 shows the synthesis strategy

of the Ru-complex (2), covalently linked to the silica matrix. Starting from

l,2-bis(diphenylphosphino)ethane (dppe), the ligand l,2-bis[2'-(triethoxy-

silyl)ethylphenylphosphino] ethane (bspe) was synthesized with the

corresponding silyl groups that lead to the covalent linkage to the silica

matrix. Finally, the corresponding Ru complex was introduced into the

silica matrix by a solution sol-gel process.

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 117

1) Li, THF

2)H20

hi)

Kr ^

Si(OEt)3

J„O^I„0N

Si

I

„Sil

SkI

o

Os I

.\Ph Ph^\ „Si

.P. CI P. ^~^

Ru

k. / I \„

„,P CI PC ^\ „O"-^Ph Ph""-^ „Si

O I

RuCI2(PPh3),acetone

r-° k J °^ ..Sk

O I o

o.

1)TE0S, THF

oxalic acid

2)THA, THF

1

\/ ^O—Si-O^ I

I .Sl

Si-O^ I

/ „Sio' o'

oI „OH

-Sk /

O'HOv I

Ph PI>-\ „.Si

;pv CI P. ^-^

9 ^o—SiOU „O /\

\ Si SkNI I OH

\ 0 0Si

C X )„P CI Pk

-^Ph pFT-

oI „o-

-Sko

.SiHO I

O.

-Si-

/

O'

SiI

Si

\/-Si

.O-Si'

•o\

i y

•Sl~o.O-Si-

/

o o-

/-Si"

\

Scheme 7-1. Synthesis of the immobilized RuCl2(bspe)2 hybrid catalyst (2). Starting

from l,2-bis(diphenylphosphino)ethane (dppe), the ligand l,2-bis[2'-(triethoxysilyl)-

ethylphenylphosphino]ethane (bspe) was synthesized via 1,2-bis(phenylphosphino)-

ethane (ppe). After complexation with bspe in a ligand exchange process from

RuCÏ2(PPh3)3, catalyst (2) was obtained.

f2-Bis(phenylphosphino)ethane (ppe). The phosphine (ppe) was prepared

in a similar manner as described in [23], using the following procedure:

100 ml tetrahydrofuran (THF), freshly distilled over sodium, was degassed

and stored in a 500 ml three-way flask under argon atmosphere. 4 g of

118 Chapter 7

lithium wire in pieces of 2 cm was added. In an ice bath at 0°C, 21.35 g of

l,2-bis(diphenylphosphino)ethane in THF was added dropwise under

heavy stirring, resulting in a tawny solution. After 2 h of refluxing at 80°C,

the color changed again to dark brown. The solution was separated from

the lithium and cooled with ice to 0°C. For the precipitation of lithium

hydroxide, 15 ml water in 60 ml THF was added before drying in vacuum.

The product l,2-bis(phenylphosphino)ethane (10.32 g, 78%) was extracted

with 100 ml diethylether after addition of 200 ml water, followed by

vacuum distillation at 0.13 mbar and 160°C. 1H-NMR: (500 MHz, C6D6,

28°C) ô = 1.5 - 1.9 (m, 4H, CH2), 3.83 (m, 1H, PH), 4.25 (m, 1H, PH), 7.0

(m, 6H, ArH), 7.25 (m, 4H, ArH); 31P-NMR: (200 MHz, C6D6, 28°C) ô =

-12.93. IR: (neat) 3068, 3053, 2281 (s,v(P-H)), 1481, 1435, 730, and

694 cm-1.

f2-Bis[2'-(triethoxysilyl)ethylphenylphosphino]ethane (bspe). In the next

step, the biphosphine (ppe) was further modified to introduce silyl groups.

4.829 g (19.6 mmol) l,2-bis(phenylphosphino)ethane and 7.46 g

(39.2 mmol) vinyltriethoxysilane were exposed to UV-light in a water-

cooled fused quartz vessel for 72 h (12.2 g bspe, 99%), in a similar way as

reported in [24]. 1H-NMR: (500 MHz, C6D6, 28°C) ô = 0.62 - 0.84 (m,

4H, PCH2C7/2Si), 1.10 (t, J = 7.0, 18H, SiOCH2C//5), 1.72 - 2.4 (m, 8H,

VCH2CH2V, PC7/2CH2Si), 3.71 (q, J = 6.98, 12H, SiOC7/2CH3), 7.0 - 7.5

(m, 10H, ArH); 31P-NMR: (200 MHz, C6D6, 28°C) ô = -12.93. IR: (neat)

3064, 2974, 2926, 2889, 1482, 1434, 1390, 1293, 1261, 1167, 1103, 1080,

958, 773, 743, 725, and 697 cm-1. No signal was observed for v(P-H).

RuCl2(bspe)2 and the synthesis of the hybrid gel. Finally, the transition

metal was introduced and the hybrid gel was prepared. For this purpose,

2.40 g (3.83 mmol) bspe was added to a suspension of 1.86 g (1.94 mmol)

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 119

RuCl2(PPli3)3 in 20 ml acetone and stirred for 12 h. A dark brown oily

material was obtained after the solvent had been removed. Analysis of the

oily product: 1H-NMR: (500 MHz, C6D6, 28°C) ô = 0.90 - 0.95 (m, 4H,

PCH2C7/2Si), 1.10 (t, J = 7.0, 18H, SiOCH2C//5), 1.2 - 2.4 (m, 8H,

VCH2CH2V, PC7/2CH2Si), 3.71 (q, J = 6.98, 12H, SiOC7/2CH3), 7.0 - 7.6

(m, 10H, ArH); 31P-NMR: (200 MHz, C6D6, 28°C) ô = -4.08. IR: (neat)

3064, 2973, 2924, 2886, 1476, 1433, 1390, 1165, 1101, 1075, 958, 741,

and 692 cm-1.

To 20.05 g tetraethoxysilan in 10 ml THF, 3.00 g RuCl2(bspe)2 complex

was added slowly. 10 ml of 0.001 M oxalic acid in 30 ml THF was added

dropwise with stirring. The moderate stirring of the solution was continued

for 24 h. The addition of 4.07 g trihexylamine in 15 ml THF caused the

gelation within a few minutes. The gel was kept under argon and aged for

4 days. During this time, tetrahydrofuran, which was released by the gel,

was replaced by ethanol [25].

Finally, the resulting RuCl2(bspe)2/Si02 (2) was crushed after evaporation

of the solvent, washed with ethanol, and dried in vacuum at 60°C. 3lP-

MAS-NMR of yellowish catalyst (2): (162 MHz, 28°C) ô = 33.46; 29si-

MAS-NMR: (80 MHz, 28°C) ô = -93.5 (Si(OH)2(OH)2), -101.38

(Si(OH)(OSi)3), -109.49 Si(OSi)4.

7.3.2 Preparation of the Adsorbed Catalyst RuCl2(dppe)2 on Amino-

propyl-modified Silica

RuCl2(dppe)2 (1) was synthesized from 1.00 g RuCl2(PPfi3)3 in a 20 ml

acetone suspension by ligand exchange with 0.85 g dppe under argon

atmosphere [26]. After filtration, washing with acetone and methanol, the

yellow precipitate was dried in vacuum and analyzed with lH- and 31P-

120 Chapter 7

NMR and elemental analysis; calculated (%) for C52H4gCl2P4Ru

(968.8 g/mol): C 64.47, H 4.99, P 12.79, CI 7.32, Ru 10.43; found: C

64.30, H 5.19, P 12.94; 1H-NMR and 3lP-NMR as in ref. [26].

In a next step, 0.1002 g (0.1034 mmol) RuCl2(dppe)2 (1) was dissolved in

8.712 g (100 mmol) morpholine (Fluka, > 99.0%) and stirred for 2 h under

80 bar hydrogen (Pangas, 99.999%) atmosphere. 0.5 g aminopropyl-

modified silica was added and stirred for 2 h. After filtration and drying in

vacuum, 0.49 g of a white powder was obtained that contained 1070 ppm

Ru.

7.3.3 Preparation of Further Catalytic Materials, Covalently Bound

RuCl2(bspp)2 and In Situ Generated Catalyst from "Si"-NH2-RuCl3

and dppe

The immobilized catalyst RuCl2(bspp)2 (5) was synthesized according to

[19]. A mixture of l,3-bis[phenylphosphino]propane and vinyltriethoxy-

silane was exposed to UV-light for 4 days to prepare l,3-bis[2'-(triethoxy-

silyl)ethylphenylphosphino]propane (bspp). After ligand exchange of

RuCl2(PPli3)3 with bspp and applying a solution sol-gel process,

catalyst (5) was dried in vacuum.

The in situ generated catalyst (6) "Si"-NH2-RuCl3 + dppe from "Si"-NH2-

RUCI3 and dppe was prepared in a similar way as reported by Zhang et

al. [21], using a mixture of hydrated RUCI3 and "Si"-NH2 in ethanol.

Before the formylation in the high-pressure batch reactor, dppe was added

to the mixture of amine and "Si"-NH2-RuCl3, forming the active catalyst

from "Si"-NH2-RuCl3 + dppe (6) during reaction.

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 121

7.3.4 Catalytic Formylation of Amines with Supercritical Carbon

Dioxide

In a typical procedure, the corresponding amine and the catalyst were

poured into a 500 ml high-pressure stainless steel autoclave (see chapter 2,

section 2.1.1). After closing, the reactor was flushed three times with

hydrogen (Pangas, 99.999%), then heated to 100°C, adjusting the hydrogen

pressure to 80 bar. 100 g liquid carbon dioxide (Pangas, 99.995%),

quantified by a Rheonik mass flow controller (RHM 01, RHE 02), was

added using a compressor from a C02 gas cylinder equipped with a dip

tube. The end of this filling process marked the starting time of the

reaction, which was carried out at a constant stirring rate of 300 min-1. The

final total pressure at the start was about 200 bar at 100°C. A gas

Chromatograph (HP-6890) equipped with a HP-5 capillary column (30 m x

0.32 mm x 0.25 urn) and a flame ionization detector (FID) was used for

analyzing the reaction mixture after filtering (0.45 urn pore size). The

identification of the products was carried out on a gas Chromatograph (HP-

6890) coupled to a mass spectrometer (HP-5973). The ruthenium content of

the liquid phase and the ruthenium loadings of the solid catalysts were

determined by inductively coupled plasma optical emission spectroscopy

(ICP-OES).

7.3.5 Physico-chemical Characterization Using X-ray Absorption

Spectroscopy (XANES, EXAFS) and Further Characterization

Methods (XPS, Nitrogen Physisorption, NMR, and DRIFTS)

X-ray absorption near edge structure and extended X-ray absorption fine

structure experiments were performed at the Swiss-Norwegian beamline

(SNBL) at the European Synchrotron Radiation Facility (ESRF) in

122 Chapter 7

Grenoble, France. The storage ring operates at 6 GeV with a ring current of

200 mA. The monochromator used consists of Si(lll) channel-cut. Higher

harmonics were effectively removed by a double-bounce gold-coated

mirror system. Each of the three ionization chambers was filled with a

combination of the gases Ar, N2 and Kr (Iq At, It and Iref 30% Kr and 70%

N2) to record the intensity of the incident and the transmitted X-rays. The

samples were placed between the first and second ionization chamber, with

the RUCI3 reference pellet placed between the second and third ionization

chamber. For low-concentration samples, the X-ray absorption spectra

were in addition measured in fluorescence mode. The EXAFS spectra were

taken under stationary conditions in the step-scanning mode around the Ru

K-edge (22.117 keV) from 21.900 keV to 22.800 keV with a RuCl3 pellet

as reference. The raw data were energy-calibrated with the Ru K-edge

energy of the RUCI3 pellet at 22.120 keV [27] at the first inflection point,

background-corrected and normalized using WINXAS 3.0 software [28].

The EXAFS data was Fourier-transformed to the kl-weighted functions in

the interval k = 3-13.0Â_1 and fitted to R-space. Phase shifts and

backscattering amplitudes were calculated using FEFF 6 [29]. Deviations

of the distances were within ±0.02 Â, and those for the coordination

numbers were within ±0.5.

For the identification of the structure of catalyst (2) and catalyst (3),

pressed pellets of the corresponding materials were used in which the metal

content was adjusted to yield an absorption ixd of about 1.5 at the Ru K

absorption edge. For fluorescence EXAFS, a pellet of about 3 mm

thickness was used at a 45° angle to the beam, complemented by the 13-

element fluorescence detector (13-element Ge solid-state detector,

Canberra). The fluorescence signals from the 13 detector elements were

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 123

pre-amplified directly and then fed into the XIA digital pulse processing

electronics (DXP 2X). The digital electronics features full 8k MCA capa¬

bility for each detector element. The Ru Ka-fluorescence regions were

selected and calibrated on the basis of a full spectrum for each channel,

collected prior to the measured scans. The signal of interest (Ru Ka-

fluorescence) versus the (elastic/matrix) background was optimized with

the aid of filters. The resulting 13 SCA data were added up after each

measurement. The corresponding liquid reaction mixtures were

investigated in the stainless steel EXAFS cell with Kapton windows, both

in transmission and in fluorescence mode, as depicted in Figure 7-2 and

described in detail in chapter 2, section 2.2.1. The volume is about 2 ml,

and a long path length of 4 cm was chosen for the transmission EXAFS

experiments because of the low ruthenium concentration in the samples

(50 - 100 ppm). For the same reason, fluorescence EXAFS spectra were

recorded at a 90° angle, too. Figure 7-2 shows that both fluorescence and

transmission EXAFS spectra could be taken at the same time. The 0.5 mm

x 5 mm beam was focused on the center of the spectroscopic cell by use of

a moveable x, z, 0-table from Newport.

X-ray photoelectron spectroscopy (XPS). Surface analytical information

was obtained by XPS using a Leybold Heraeus LHSll MCD instrument

[30] and Mg Ka (1253.6 eV) radiation. The catalyst powder was pressed

into a sample holder, evacuated slowly in a load lock to 10-6mbar and

transferred to the analysis chamber (typical pressure < 10"9mbar). The

peaks were energy-shifted to the binding energy of C(ls) to correct the

charging of the material. The surface composition of the catalysts was

determined from the peak areas of C(ls), 0(1 s), Si(2p), Ru(3d), which

were computed after subtraction of the Shirley-type background by

124 Chapter 7

empirically derived cross-section factors [31]. The relative error of the

analysis is estimated to be within ±5%.

Combined

transmission/fluorescence

EXAFS cell

4 cm

Reference

sample

X-raybeam

5 mm x 0 5 mmfl

Ionization chamber 1 Ionization chamber 2 Ionization chamber 3

Cr-filter

13 element

fluorescence

Ge-detector

Figure 7-2. Schematic representation of the experimental setup with the stainless steel

EXAFS cell for simultaneous measurements in transmission andfluorescence mode at

the Ru K-edge, as performed at SNBL (ESRF, Grenoble). The distance of the 13-

element Ge solid-state detector is adjustable, and the use ofthe filterfoil depends on the

kind ofsample used (for details, see text and chapter 2, section 2.2.1).

Nitrogen physisorption. The BET surface areas Sbet were determined

from nitrogen physisorption at 77 K using a Micromeritics TriStar 3000

instrument. Prior to the measurements, the samples were heated in vacuum

at 60°C for 6 h. BET surface areas were calculated in a relative p/po

pressure range from 0.05 to 0.2, assuming a cross-sectional area of

0.162 nm2 for the adsorbed nitrogen molecule.

Nuclear magnetic resonance spectroscopy (NMR). IH-NMR spectra were

taken on a Bruker Avance 500 instrument at an operation frequency of

500.12 MHz. Liquid 31P-NMR spectra were measured on a Bruker

Avance 500 at 202.4 MHz. Solid-state 31P-MAS-NMR spectra of

catalyst (2) and the corresponding 29Si-MAS-NMR spectra were carried

out on a Bruker AMX 400 at 162.0 MHz and at 79.5 MHz for the 29Si-

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 125

MAS-NMR, respectively. Magic angle spinning (MAS) was applied at

10 kHz for 29Si and at 12 kHz for 3lP, respectively.

Infrared spectroscopy. Neat liquids were measured as thin films in the

transmission mode using KBr windows. The Ru-complexes (RuCl2(bspe)2

and RuCl2(PPh)3) were measured in the attenuated total reflection (ATR)

mode using a ZnSe crystal accessory mounting from Harrick. Diffuse

reflectance infrared Fourier transform spectra (DRIFTS) of immobilized

Ru complexes were obtained at 25°C after dilution with KBr using a

commercial cell (Harrick). Spectral acquisition was achieved by adding

100 scans at 4 cm-1 resolution with an Equinox 55 spectrometer (Bruker

Optics) purged with dry air. The MCT detector was selected for DRIFTS

measurements. KBr and the pure ZnSe crystal were used as the reference

for transmission IR/DRIFTS and for ATR-IR measurements, respectively.

7.4 Results

7.4.1 Preparation and Characterization of Heterogeneous Catalysts

Different strategies for the immobilization of Ru-based catalysts were

applied, as summarized in Figure 7-1. RuCl2(dppe)2 (1) was used as

reference and for the adsorption on aminopropyl-modified silica resulting

in catalyst (3). Catalyst (5) as a heterogeneous analog of RuCl2(dppp)2 (4)

was prepared and characterized according to [19]. Special attention was

given to the new hybrid-catalyst RuCl2(bspe)2/Si02 (2) since the

homogeneous analog RuCl2(dppe)2 (1) exhibits outstanding catalytic

activity [19, 20]. Different techniques were used to elucidate the textural

and structural properties of the heterogeneous catalysts. iH-NMR, 31P-

NMR, and 29Si-NMR, as well as DRIFTS, were used to gain information

on the ligand and silica structure during immobilization, EXAFS/XANES

126 Chapter 7

to identify the structure around the Ru absorber atom, and nitrogen

adsorption for surface area and pore structure analysis.

The iH-NMR spectrum of the complex RuCl2(bspe)2 showed a similar

ligand structure as the uncomplexed ligand bspe, only the peak in the 31P-

NMR was shifted to -4.08, because of electronic effects of the ruthenium-

phosphorus bond. The 31P-MAS-NMR spectrum of catalyst (2) in

Figure 7-2 indicated that no significant change of the geometric structure in

the coordination sphere of the phosphorus atoms had taken place during the

immobilization process. Beneath the shifted main peak at 33.46 ppm, side¬

bands were observed in the spectra of the solid material, caused by the

sample rotation. The condensation degree of catalyst (2) was determined by

applying 29Si-MAS-NMR (Figure 7-3). Three distinguishable peaks

corresponding to different moieties of silicon atoms with four oxygen

neighbors were discernible (Figure 7-3), the chain- (Q2, Si(OH)2(OSi)2) at

-93.5 ppm, the bridgehead- (Q3, Si(OH)(OSi)3) at -101.38 ppm, and the

completely condensed hydroxysilanes (Q4, Si(OSi)4) at -109.49 ppm.

The DRIFT spectrum of catalyst (2) further confirmed the completed

condensation. The typical strong signals at 2974, 2926 and 2889 cm-1 of

the asymmetric and symmetric stretchings of CH2 and CH3 of the ethoxy

groups of the neat ligand disappeared upon immobilization. Moreover, a

band of medium intensity at 3691 cm-1 revealed the presence of free silanol

groups. New signals attributable to the CH2 groups of the immobilized

ligand appeared at 2956, 2931, 2870 and 2859 cm-1 and were accompanied

by a weak signal at 3060 cm-l, originating from the v(C-H) modes of the

phenyl rings. Additional signals at 747 and 695 cm-1 (out-of-plane ô(C-H)

and ring deformation) and at 724 cm-1 (v(P-CH2)) [32] demonstrated the

presence of methylenephenylphosphine groups on the silica support.

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 127

b) 29Si-MAS-NMR

a)31P-MAS-NMR side bands

^«^»^S»*^»»»^

(ppm)

HOv ,OSlSl

Sio' "0S|

SlOv ,OSlSl

sid' "osi

V /

-100

(ppm)

Figure 7-3. Solid state a) 31P-MAS-NMR and b) 29Si-MAS-NMR of the immobilized

catalyst RuCEßspe)2 (2).

The XANES and the corresponding k1-weighted Fourier-transformed

EXAFS spectra of the covalently bound catalyst (2) and the adsorbed

catalyst (3) are shown in Figure 7-4. For comparison, the spectrum of the

covalently bound catalyst (5) is also shown. The XANES spectra are

similar for all three immobilized catalysts. Note that variations were found,

in correspondence with a previous study of RuCl3(PPfi3)3 and RUCI3 [33].

Even slight changes in the symmetry and ligands usually affect the near

edge structure of such complexes due to different multiple scattering paths

[34]. The backscattering amplitude of the first nearest neighbors (Figure 7-

128 Chapter 7

4b) shows a variation in both the covalently immobilized and adsorbed

complexes. This may be attributed to a slight variation in the local structure

of the immobilized complexes. In principle, two covalent linkages are

possible for each ligand. Depending on its incorporation in the silica

matrix, the Ru-P distance may vary to a certain extent, a consequence of

the structurally slightly different species in the silica matrix.

RuCI2(bspe)2/Si02 (2)

RuCI2(dppe)2/Si02 (3)

^RuClibsppySiO, (5)0 0000'

RuCI2(bspe)2/Si02 (2)

RuCI2(dppe)2/Si02 (3)^RuCL(bspp),/SiO,(5)

22 1 22 2 22 3

E/keV

22 4 22 5 3 4 5

R/A

Figure 7-4. a) X-ray absorption near edge structure at the Ru K-edge of the

immobilized RuCl2(bspe)2 (2), the adsorbed RuCE(dppe)2 (3), and the immobilized

RuCl2(bspp)2 (5) catalysts; and b) the corresponding Fourier-transformed spectra of

the k1-weightedEXAFSfunctions.

Nevertheless, the structural data extracted from EXAFS spectra of the

immobilized phosphine complexes and from RuCl2(dppe)2 (1) as reference

show that the immobilized catalysts (2) and (5) have similar local structures

(Table 7-1). Like the homogeneous catalyst (1) and the adsorbed

catalyst (3), they show backscattering peaks at 2.35 Â for the Ru-P distance

and at 2.43 Â for the Ru-Cl bond length. The Ru-P and the Ru-Cl distances

of both catalysts are in agreement with the Ru-P and Ru-Cl distances of the

well-characterized complex RuCl2(dppm)2 [35]. The altered coordination

numbers found in catalyst (3) may be explained by a certain coordination to

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 129

the amino group, but also to water molecules or OH-groups. As Figure 7-4

shows, the spectrum of catalyst (3) is quite similar to RuCl2(dppe)2 (1),

which indicates that the structure resembles that of the homogeneous

catalyst (1) [33, 36]. Therefore, hardly any interaction with the nitrogen

neighbor can thus be found.

Table 7-1. Structural data extractedfrom EXAFS spectra of the different immobilized

ruthenium phosphine complexes

Catalyst Ru-P Ru-Cl

N R/Â Ao2/Â2 N R/Â Ao2/Â2

RuCl2(bspe)2/Si02 (2) 3.3 2.35 0.0053 1.9 2.43 0.0022

RuCl2(bspp)2/Si02 (5) 4.9 2.35 0.0100 1.5 2.43 0.0089

RuCl2(dppe)2/Si02 (3) 5.1 2.36 0.0062 2.6 2.40 0.0049

RuCl2(dppe)2(l) 1.9

2.0

2.37

2.39

0.0043

0.0048

1.3 2.43 0.0043

N coordination number.

R distance from the corresponding neighbor.Aa2 Debye-Waller factor. Residual as quality of the fit 1.1 - 2.5 according to ref [28]).

The BET surface of catalyst (2) was 476 m2/g, whereas catalyst (3) showed

only a small surface area of 46 m2/g, and XPS uncovered that the surface

loading of ruthenium was rather low. Catalyst (2) contained 0.1% Ru at the

surface, catalyst (5) 0.2%, while catalyst (3) showed the highest Ru loading

with 0.3%. The chlorine content of the catalysts (RuCl2(bspe)2/Si02 (2):

0.1%, RuCl2(bspp)2/Si02 (5): 0.1%, adsorbed RuCl2(dppe)2/Si02 (3):

0.2%) was in the same range as the Ru loading. Also, the phosphorus

content was quite low and no surface enrichment could be observed. ICP-

OES measurements gave a ruthenium loading of about 0.5 wt-% for

130 Chapter 7

catalyst (2), 0.15 wt-% for catalyst (5), and 0.10 wt-% for catalyst (3),

respectively.

7.4.2 Catalytic Formylation of Amines Using Immobilized

RuCl2(bspe)2

The catalytic activity of the immobilized RuCl2(bspe)2 (2) was tested in the

formylation of various amines with hydrogen and carbon dioxide

(Table 7-2). Note that the substrates contained an additional functional

group (ether or hydroxyl). The highest catalytic activity related to the

catalyst mass was achieved for formylmorpholine synthesis - 164 mmol

product per g catalyst - corresponding to a TON of 3416.

Table 7-2. Catalytic formylation of functionalized amines with supercritical carbon

dioxide using the immobilizedRuCE(bspe)2/SiO'2 catalyst (2)

Run Amine Catalytic activity TON

/(mmol product/g catalyst)

1 Morpholine 164 3416

2 3-Methoxypropylamine 148 3078

3 2-Ethylaminoethanol 48 974

4 2-Methylaminoethanol 37 760

Conditions: 100 mmol amine, 100 g C02, 80 barH2 at 100°C, 0.5 g catalyst, 100°C, 20 h.

Selectivity: about 100%, no by-product identified.

The formylation of the primary amine 3-methoxypropylamine, with a

catalytic activity of 148 mmol per g catalyst in 20 h and a TON of 3078,

was about as efficient as that of morpholine. The formylation of secondary

amines containing a free hydroxyl-group, 2-ethylaminoethanol (48 mmol/g,

TON of 974) and 2-methylaminoethanol (37 mmol/g, TON of 760)

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 131

occurred considerably slower. As expected, catalyst (2) was more active

than the reference catalyst (5) (TON of 1270 for (5), TON of 3078 for (2)).

Therefore, further attention was mainly given to catalyst (2).

CZI RuCI2(bspe)2/Si02 (2)^ RuCI2(dppe)2/Si02 (3)

< 0 1 mmol/g cat

-7ZZZA I I I I2 3 4

exp. no.

Figure 7-5. a) Changes of the conversion in mmolproduct per g catalyst and turnover

number (TON) of the immobilized RuCl2(bspe)2 (2) catalyst during four recycling

loops. Reaction conditions: 100 mmol mpa, 0.5 g catalyst, 100 g CO2, 80 bar H2 initial

partial pressure, 6 h and 100°C; b) variation of the catalytic activity during four

recycling loops ofthe immobilized catalyst (2) compared with catalyst (3).

In order to test its behavior during recycle, catalyst (2) was reused four

times in the formylation of 3-methoxypropylamine. After each run, the

reaction mixture was filtered off and the catalyst was washed with 20 ml

distilled water and finally dried in vacuum for 24 h. Water was used in this

step, since it is a reaction by-product. Figure 7-5a shows the catalytic

activity related to the catalyst mass in the synthesis of 3-methoxy-

propylformamide and the corresponding turnover number after reuse of

catalyst (2). The formamide yield decreased with repeated use of the

catalyst. After four cycles, the initial catalytic activity related to the catalyst

mass of about 40 mmol/g dropped to 20 mmol/g, a behavior also reflected

by the turnover numbers.

132 Chapter 7

7.4.3 Comparison of the Immobilized Ruthenium Complex and the

Adsorbed Ruthenium Complex

A completely different behavior after reuse was observed for catalyst (2)

and catalyst (3). Figure 7-5b shows the results for the formylation of

3-methoxypropylamine. With 128 mmol/g catalyst and a resulting turnover

number of about 12073 mol product per mol ruthenium, the productivity of

catalyst (3) was significantly higher during the first cycle than that of

catalyst (2) (catalytic activity related to the catalyst mass of 3-methoxy-

propylformamide 41 mmol/g). However, with catalyst (3) the catalytic

activity dropped dramatically right after the first recycling step. After the

second recycling step, no further formylation activity was found for

catalyst (3). The in situ generated "Si"-NH2-RuCl3 + dppe catalyst (6)

showed a catalytic activity related to the catalyst mass of 59 mmol product

per g catalyst and a TON of 735 during the first use for 3 h at standard

experimental conditions. Also, a leaching test as suggested by Sheldon et

al. [37] was positive, as the remaining species in the filtrate led to further

reaction.

This prompted us to investigate the loss of ruthenium per recycling step for

both catalyst (2) and catalyst (3). In Table 7-3, the ruthenium content of the

catalysts and the ruthenium concentrations in the reaction mixtures of

catalyst (2) are listed. The freshly synthesized catalyst contained

0.048 mmol ruthenium per g catalyst. During the first reaction, this

ruthenium content decreased to 0.034 mmol/g. In each of the following

recycling steps, the ruthenium loading of the catalyst decreased, but also

the ruthenium concentrations in the corresponding solutions dropped from

0.7 mmol ruthenium per g solution to 0.1 mmol/g. Catalyst (3) contained

0.0011 mmol ruthenium per g catalyst after synthesis. After its first use,

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 133

only 2.1 jLimol ruthenium per g catalyst remained in the catalyst, the

catalytic activity related to the catalyst mass decreased to 4.8 mmol/g, and

after a second recycling step, a very low ruthenium concentration of

0.6 jLimol/g catalyst was measured and no more production of formamide

could be observed.

Table 7-3. Ruthenium concentrations of the immobilized RuCl2(bspe)2 catalyst (2) in

each recycling step during formylation of 3-methoxypropylamine with supercritical

carbon dioxide

Experiment number Ru in solid catalysta Ru in solution

(recycling step) /(mmol/g) /(umol/g)

0 0.048

1 0.034 0.706

2 0.036 0.267

3 0.028 0.281

4 0.017 0.123

a note: determination of Ru concentration in solid is less reliable than in solution.

7.4.4 Structural Analysis of the Solid Catalysts and the Soluble Species

after Reaction

In a similar manner as for the as-prepared Ru-catalysts, Ru K-edge XANES

and EXAFS spectra of catalysts (2) and (3) were taken after reaction

(Figure 7-6). Essentially, the spectra of catalysts (2) and (3) in as-prepared

state as well as after their reuse were similar. However, the shoulder at

1.2 Â was more prominent. The main backscattering peak was shifted to

slightly lower R-values and a shoulder appeared at 2.5 Â. Moreover, in the

case of catalyst (3) the concentration of Ru with 0.02 wt-% was

134 Chapter 7

significantly lower after first use, causing the statistics to be worse despite

of fluorescence detection.

1 4

1 2

1 0

08

06

04

0 2-f

00

- RuCI2(bspe)2/Si02 (2)- RuCI2(bspe)2/Si02 (2), after use

- RuCI2(dppe)2/Si02 (3)- RuCI (dppe) /SiO (3), after use

0 005-

S 0 004-

S 0 003-

^0 002H

T3

| 0 001-

CO

^ 0 000-r—'

b)

RuCI2(bspe)2/Si02 (2)after use

A

Mi \\i \

'i RuCI2(dppe)2/Si02 (3)after use

22 1 22 2 22 3

E/keV

22 4 22 5 4

R/A

Figure 7-6. a) X-ray absorption near edge structures at the Ru K-edge of catalysts (2)

and (3) before and after reaction, and b) the corresponding Fourier-transformed

spectra ofthe kl-weightedEXAFSfunctions.

In addition, DRIFT spectra of the used catalyst (2) after recovery from the

autoclave were compared to that before and after immobilization. The

results shown in Figure 7-7 indicate that additional bands appear at 2046,

1982, ca. 1950, 1859 and 1661 cm-1, compared to the corresponding

freshly prepared immobilized Ru complex. The latter signal is attributed to

the presence of adsorbed products (v(C=0) of amide groups). On the other

hand, the features above 1800 cm-1 are related to Ru carbonyl species, if

compared to organometallic complexes of Ru containing CO [38, 39]. Note

also that the Ru-H stretching bond was reported between 1800 and

2000 cm-l [40]. This may explain the appearance of the shoulders in the

EXAFS spectra in Figure 7-6. Significant changes in the filtered catalyst

are also observed in the 850 - 650 cm-l spectral region for the bands

attributed to the phenylphosphine groups.

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 135

Figure 7-7. DRIFT spectra of the RuCl2(bspe)2 catalyst a) before immobilization, b)

immobilized RuCl2(bspe)2 (2) before reaction, and c) immobilized RuCE(bspe)2 (2)

after reaction.

As mentioned before, after the first use of catalyst (2), soluble ruthenium

species were found in the product solution. With a low concentration of

about 0.7 jLimol/g solution, generally fluorescence mode is favored [7, 41].

In a recent study, we carried out transmission EXAFS experiments at a

concentration as low as 2 jiimol/g solution [33]. These experiments inspired

us to compare both fluorescence and transmission EXAFS data. In

Figure 7-8, the spectra obtained from the liquid solution of catalyst (2) are

depicted in the fluorescence and in the transmission mode. Both methods

led to the same whiteline and near edge structure. Note that in both cases

the Ru concentration was about 70 ppm and the spectra were taken during

1 h. It seems that due to the long path length of the liquid cell (4 cm) and

136 Chapter 7

the high energy of the Ru K-edge, transmission EXAFS data are of similar

quality as the fluorescence EXAFS data.

1.0-

0.8-

d

cd 0.6-

^ j I

Fluorescence EXAFS

Transmission EXAFS

22.1 22.2 22.3 22.4 22.5

E/keV

Figure 7-8. Comparison offluorescence EXAFS data taken at the Ru K-edge in the

fluorescence mode (dashed line) and in the transmission mode (solid line), taken from

the solution after the reaction of catalyst (2). The concentration of the catalyst in the

liquidphase was 71 ppm. Total scanning time was about 1 hour.

To gain more information about the structure of the soluble species after

reaction, the liquid reaction mixtures after reaction under standard

conditions with catalyst (2) and (3) were compared to those from the

homogeneous catalyst (1) after similar reaction exposure. The spectrum of

catalyst (2) was measured during 5 h, whereas for catalyst (3) 3 h were

sufficient due to the higher Ru-concentration. As depicted in Figure 7-9,

the spectrum resulting from the product solution of catalyst (2, solid black

line) differs strongly from those of catalyst (3, dashed black line) and

catalyst (1, solid gray line). It seems that the catalytically active species on

catalyst (3) exhibit a structure similar to that of the homogeneous

RuCl2(dppe)2 (1) after reaction. This implies that catalyst (3) desorbs

during reaction, homogeneously catalyzes the reaction, and remains in the

rv^-*-/--

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 137

liquid product mixture. The structure of the soluble part of catalyst (2),

however, was much different. The peak at 1.8 Â was shifted to significantly

lower values and also was less intense. This indicates that other ligands,

such as amino or hydroxyl groups, are coordinated to the Ru center.

RuCI2(bspe)2/Si02 (2)

RuCI2(dppe)2/Si02 (3)— free RuCI (dppe)

22 1 22 2 22 3

E/keV

22 4 22 5

RuCI2(bspe)2/Si02 (2

RuCI2(dppe)2/Si02 (3)— free RuCI2(dppe)2

22 1 22 2

E/keV

22 3

k/A-1

to

^ 0 0015"

d)

k50 0010-

^ I \

Magnitude0

0

0

0

0

0

0

0

O

Ol

Ji\I \^^^^^^— r*^

LL

R/A

Figure 7-9. Structural comparison of the Ru-complexes formed during the formylation

of 3-methoxypropylamine in the presence of hydrogen and carbon dioxide with

immobilized RuCE(bspe)2 (2, solid black line), adsorbed RuC'I^(dppe)2 (3, dashed line)

and RuCE(dppe)2 (1, solid gray line) using X-ray absorption spectroscopy at the Ru K-

edge; data taken in transmission mode; a) XAS spectra, b) XANES regions, c) k1-

weighted xßj^-functions and d) Fourier-transformed ^-weighted EXAFS functions

(k = 3.5- 13Â-1).

138 Chapter 7

This may also explain the change in the near edge structure, supporting the

conclusion that the modified phosphine ligands (catalyst (2)) are better

incorporated into the matrix than in the case of catalyst (3), where they are

evidently dissolved in the reaction mixture, resulting in XANES and

EXAFS spectra that are similar to catalyst (1) after reaction. For the spectra

of ruthenium complexes in the liquid phase, see chapter 4 and chapter 5.

7.5 Discussion

Two strategies for the immobilization of highly active Ru-phosphine

complexes - the coordination of the Ru to amine-modified silica, and the

covalent anchoring on silica via functionalized ligands - have been

evaluated with respect to their potential for the preparation of

heterogeneous Ru catalysts for the formylation of amines in supercritical

carbon dioxide. This encompasses the characterization of the prepared

grafted complexes, the catalytic behavior, the reuse, and the determination

of possible leached species.

The adsorption of homogeneous complexes on modified silica has been

reported in a number of studies [42] as a simple strategy. Zhang et al. [21]

recommended this procedure as well applicable to the hydrogénation of

carbon dioxide. The active catalyst was formed in situ from amine-

modified silica-supported ruthenium complexes ("Si"-NH2-RuCl3) by

addition of a phosphine (dppe or PPI13). As typical examples of the

adsorption strategy, we prepared this catalyst (6) and additionally grafted

the RuCl2(dppe)2 complex on an aminopropyl-modified silica surface

(catalyst (3)). EXAFS and XPS were applied to prove the successful

deposition of the Ru-complex on aminopropyl-modified silica.

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 139

For the covalent linkage of the phosphine ligand, we prepared the new

covalently immobilized catalyst (2) RuCl2(bspe)2/Si02 with improved

activity compared to RuCl2(bspp)2/Si02 (5). Tai et al. [20] showed the bite

angle of the phosphine to be a decisive criterion, concluding that dppe and

dmpe as phosphine ligands are much more active than dppp, dppb and the

monophosphines in the formylation of amines with CO2. In chapter 5, the

Ru(dppe)2X2 catalyst proved to be the most active one as well. This

strengthens the motivation to immobilize a complex with a small bite angle

as encountered in catalyst (2). Different routes for the synthesis of

catalyst (2) using covalent linkage of the phosphine ligand were

considered. The route described in Scheme 7-1 afforded high yield, and all

the intermediates could be confirmed by NMR, IR, and/or EXAFS

spectroscopy. Also, the structure of the desired catalyst (2) was evidenced

using NMR, IR, EXAFS and XPS. On an alternative route, an amino group

instead of the triethoxysilane group was used for immobilization,

employing a nucleophilic substitution with the halogenalkyl group of a

correspondingly pre-treated silica support [43]. However, the reaction of

1,2-bis(phenylphosphino)ethane (ppe, see section 7.3.1 and Scheme 7-1)

and 3-aminopropene to l,2-bis(3'-aminopropylphenylphosphino)ethane in

the presence of benzene, UV-light, and AIBN as a radical starter [44]

resulted in a much lower yield of the desired ligand, and several by¬

products were evidenced by NMR.

Catalysts (2), (3), (5), as well as the in situ generated "Si"-NH2-RuCl3 +

dppe catalyst (6), proved to be catalytically active in the formylation of

differently functionalized amines. In contradiction to the experiments of

Zhang et al. [21], the adsorbed complexes in catalysts (3) and (6) were not

stable in the formylation of amines and hardly active after first reuse.

140 Chapter 7

Therefore, this simple procedure is not applicable under relatively harsh

reaction conditions with temperatures around 100°C and a pressure of

about 200 bar, and ICP-OES studies confirmed the strong leaching

behavior of as-prepared catalysts (3) and (6).

In comparison with these adsorbed complexes, the immobilized complex

RuCl2(bspe)2 (catalyst (2)) is more stable. As expected, the new

catalyst (2) with smaller bite angle of the phosphine was more active than

catalyst (5). The heterogenized catalysts (2) and (5) were not as active as

their homogeneous analogs, probably due to lower accessibility of the

active centers and the properties of the support. This has also been reported

in other comparative studies [42]. In order to gain more information about

the structure of the catalyst under reaction conditions, the deactivation

mechanism, and the structure of the leached species in the liquid reaction

mixture, X-ray absorption spectroscopy complemented leaching studies by

catalytic activity and ICP-OES.

XANES/EXAFS analysis of catalyst (3) after use, as well as of the liquid

solution, confirmed the strong leaching behavior under the reaction

conditions applied in this work. The structure of the remaining adsorbed

Ru-species was similar to that in the as-prepared catalyst, and thus the

conversion strongly decreased during reuse as a consequence of the

leaching of the catalytic species and possibly the formylation on the

anchoring amino groups. There is also another indication that the

catalytically active species acted as a homogeneous Ru-catalyst. According

to EXAFS, the structure of the formed homogeneous catalyst was similar to

that from RuCl2(dppe)2, which is known to act as efficient catalyst [17,

33]. Also re-adsorption during reaction as reported in [45, 46] could hardly

be observed.

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 141

Although the loss of catalytic activity of the covalently immobilized

RuCl2(bspe)2 catalyst (2) during reuse was much less pronounced than that

of the adsorbed complexes (catalysts (3) and (6)), even in this case some

deactivation was found. This may be due to insufficient immobilization,

incomplete complexation, or the relatively harsh reaction conditions during

formylation, which are demanding with respect to catalyst stability. In fact,

the leached Ru complex in the liquid phase of the reaction mixtures

(Figure 7-8) exhibited a strikingly different XANES and EXAFS spectrum

compared to the immobilized Ru complexes, the complex formed from

RuCl2(dppe)2, and the Ru-complex adsorbed on aminopropyl-modified

silica (catalyst (3)). The structure of the solid part of catalyst (2) remained

similar to the original structure. Hence, the main reason for the loss in

activity of catalyst (2) is the partial dissolution by destruction of the

immobilized complex rather than by the leaching of the intact complex. To

illustrate this, compare the corrosion of the RU/AI2O3 catalyst in chapter 5

with the resulting in situ formation of active ruthenium phosphine

complexes during the reaction.

In general, deactivation of grafted complexes may occur by too weak

anchoring of the complexes, by destruction of the active complexes on the

silica matrix due to the relatively harsh reaction conditions, or by the

blocking of the active sites due to the deposition of carbonaceous

species/catalyst poison. Discrimination of these phenomena requires the

use of complementary methods for leaching studies [37] (catalytic studies,

ICP-OES) and spectroscopic techniques. In this study, both transmission

and fluorescence EXAFS were found to be well-suited characterization

tools to uncover the structure of both the liquid species and the

immobilized catalysts. Even though the concentration was rather low

142 Chapter 7

(50 ppm), structural information on the liquid Ru species could be gained.

Due to the high absorption edge of ruthenium (Ru K-edge at 22.12 keV),

EXAFS in transmission mode using a long path length cell was applicable,

giving similar results as fluorescence EXAFS. For the catalyst (3) used,

fluorescence EXAFS was the preferred method, since with 0.02%, the

concentration was small, and only a limited amount of sample after a

catalytic test was available. In addition, important information on the

covalently immobilized catalysts could be gained both by EXAFS and

DRIFTS indicating the formation of Ru-carbonyl complexes.

7.6 Conclusions

Different strategies for the immobilization of highly active Ru-phosphine

complexes applied for the formylation of amines in supercritical carbon

dioxide have been examined regarding their potential for the preparation of

corresponding heterogeneous catalysts. Best results with regard to activity

and stability during reuse were obtained with catalysts where the Ru-

complexes were covalently anchored on silica. Simple impregnation proved

to be very unstable, resulting in strong leaching under the reaction

conditions applied. An in situ generated catalyst from "Si"-NH2-RuCl3 and

dppe, as well as a catalyst derived by adsorption of RuCl2(dppe)2 on

aminopropyl-modified silica (3), failed since the catalysts were rather

inactive during reuse in the formylation. The immobilization of the

catalytic active ruthenium-phosphine complex RuCl2(bspe)2 on silica

support via silyl anchoring groups led to a heterogeneous catalyst (2) with

high activity. The fairly demanding reaction conditions in the presence of

an amine at 100°C and 200 bar require a strong linkage to the support.

Hence, among the tested strategies only the first immobilization route with

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 143

covalently linked ligands seems to be a feasible approach for continuous

processes, in which leaching can be minimized.

XANES, EXAFS and DRIFT spectra of catalyst (2) were largely

unchanged before and after reaction, indicating the structural stability of

the anchored catalytically active species. A similar behavior was also

observed with catalyst (3). In contrast, the leached catalyst in the liquid

reaction mixtures exhibited a different XANES and EXAFS spectrum

compared to the immobilized Ru complexes (2) and (5) and the Ru

complex adsorbed on aminopropyl-modified silica (3). The application of

X-ray absorption spectroscopy both in fluorescence and transmission mode

with suitable sample cells turned out to be a powerful approach to identify

the structure of the transition metal in the as-prepared and used catalyst, as

well as in the remaining product mixture (even down to a concentration of

50 ppm).

7.7 References

[1] M. Halmann, Chemical Fixation of Carbon Dioxide: Methods for

Recycling CO2 into Useful Products, CRC Press, Boca Raton,

Florida (1993).

[2] P. G. Jessop, T. Ikariya, R. Noyori, Chem. Rev. 95 (1995) 259.

[3] W. Leitner, Angew. Chem. Int. Ed. Engl. 34 (1995) 2207.

[4] A. Behr, Angew. Chem. Int. Ed. Engl. 27 (1988) 661.

[5] M. Aresta, E. Quaranta, Chemtech (1997) 32.

[6] A. Baiker, Appl. Organomet. Chem. 14 (2000) 751.

144 Chapter 7

7] J.-D. Grunwaldt, R. Wandeler, A. Baiker, Catal. Rev. -Sei. Eng. 45

(2003)1.

8] P. Haynes, L. H. Slaugh, J. F. Kohnle, Tetrahedron Lett. 5 (1970)

365.

9] Y. Kiso, K. Saeki, Japan. Kokai Tokkyo Koho (1977) 36617.

10

11

12

13

14

15

16

17

18

19

20

K. Kudo, H. Phala, N. Sugita, Y. Takezaki, Chem. Lett. (1977) 1495.

H. Phala, K. Kudo, N. Sugita, Bull. Inst. Chem. Res. Kyoto Univ. 59

(1981)88.

S. Schreiner, J. Y. Yu, L. Vaska, Inorg. Chim. Acta 147 (1988) 139.

P. G. Jessop, Y. Hsiao, T. Ikariya, R. Noyori, J. Am. Chem. Soc. 118

(1996)344.

L. Schmid, M. Rohr, A. Baiker, Chem. Commun. (1999) 2303.

G. Süss-Fink, M. Langenbahn, T. Jenke, J. Organomet. Chem. 368

(1989)103.

L. Schmid, M. S. Schneider, D. Engel, A. Baiker, Catal. Lett. 88

(2003) 105.

M. Rohr, J.-D. Grunwaldt, A. Baiker, J. Mol. Catal. A: Chem. 226

(2005)253.

P. G. Jessop, Y. Hsiao, T. Ikariya, R. Noyori, J. Am. Chem. Soc. 116

(1994)8851.

O. Kröcher, R. A. Koppel, A. Baiker, Chem. Commun. 5 (1997) 453.

C.-C. Tai, J. Pitts, J. C. Linehan, A. D. Main, P. Munshi, P. G.

Jessop, Inorg. Chem. 41 (2002) 1606.

Evaluation of Immobilization Strategies of Ru Phosphine Complexes 145

[21] Y. Zhang, J. Fei, Y. Yu, X. Zheng, Catal. Lett. 93 (2004) 231.

[22] O. Kröcher, R. A. Koppel, A. Baiker, J. Mol. Catal. A: Chem. 140

(1999) 185.

[23] J. Dogan, J. B. Schulte, G. F. Swiegers, S. B. Wild, J. Org. Chem. 65

(2000)951.

[24] L. Schmid, O. Kröcher, R. A. Koppel, A. Baiker, Micropor.

Mesopor. Mater. 35-36 (2000) 181.

[25] M. Schneider, A. Baiker, Catal. Rev. - Sci. Eng. 37 (1995) 515.

[26] R. Mason, D. W. Meek, G. R. Scollary, Inorg. Chimia Acta 16

(1976) LU.

[27] K. Okamoto, T. Takahashi, K. Kohdate, H. Kondoh, T. Yokoyama,

T. Ohta, J. Synchrotron Rad. 8 (2001) 689.

[28] T. Ressler, J. Synchrotron Rad. 5 (1998) 118.

[29] S. I. Zabinsky, J. J. Rehr, A. Ankudinov, R. C. Albers, M. J. Eller,

Phys. Rev. B. 52 (1995) 2995.

[30] J.-D. Grunwaldt, U. Göbel, A. Baiker, J. Anal. Chem. 358 (1997) 96.

[31] C. D. Wagner, L. E. Davis, M. V. Zeller, J. A. Taylor, R. H.

Raymond, L. H. Gale, Surf. Interface Anal. 3 (1981) 211.

[32] H. Günzler, H.-U. Gremlich, IR Spectroscopy: An Introduction,

Wiley-VCH, Weinheim (2002).

[33] M. Rohr, J.-D. Grunwaldt, A. Baiker, J. Catal. 229 (2005) 144.

[34] M. Tromp, J. A. van Bokhoven, G. P. F. van Strijdonck, P. W. N. M.

van Leeuwen, D. C. Koningsberger, D. E. Ramaker, J. Am. Chem.

Soc. 127 (2005) 777.

146 Chapter 7

[35] A. R. Chakravarty, F. A. Cotton, W. Schwotzer, Inorg. Chim. Acta

84(1984)179.

[36] T. S. Lobana, R. Singh, E. R. T. Tiekink, J. Coord. Chem. 21 (1990)

225.

[37] R. A. Sheldon, M. Wallau, I. W. C. E. Arends, U. Schuchardt, Ace.

Chem. Res. 31(1998)485.

[38] P. Chutia, M. Sharma, P. Das, N. Kumari, J. D. Woollins, A. M. Z.

Slawin, D. K. Dutta, Polyhedron 22 (2003) 2725.

[39] K. Dallmann, R. Buffon, J. Mol. Catal. A: Chem. 185 (2002) 187.

[40] J. Chart, R. G. Hayter, J. Chem. Soc. (1961) 2605.

[41] S. G. Fiddy, J. Evans, T. Neisius, X. Z. Sun, Z. Jie, M. W. George,

Chem. Commun. 6 (2004) 676.

[42] M. H. Valkenberg, W. F. Hölderich, Catal. Rev.-Sci. Eng. 44 (2002)

321.

[43] C. Baleizäo, B. Gigante, M. J. Sabater, H. Garcia, A. Corma, Appl.

Catal. A: Gen. 228 (2002) 279.

[44] R. Uriarte, T. J. Mazanec, K. D. Tau, D. W. Meek, Inorg. Chem. 19

(1980)79.

[45] R. G. Heidenreich, J. G. E. Krauter, J. Pietsch, K. Köhler, J. Mol.

Catal. A: Chem. 182-183 (2002) 499.

[46] S. S. Pröckl, W. Kleist, M. A. Gruber, K. Köhler, Angew. Chem. Int.

Ed. 43(2004)1881.

Final Remarks

In this doctoral thesis, some advances pertaining to the design and the

application of ruthenium catalysts in formylation of amines have been

described. Starting with homogeneous ruthenium phosphine catalysts, a

simple access route to this catalyst class was found by in situ generation

from RUCI3 and the corresponding phosphines. This strategy is also

promising in the area of catalyst screening, where the phosphine ligand

may be varied, for example. The in situ formation of active species from

RU/AI2O3 and l,2-bis(diphenylphosphino)ethane (dppe) has demonstrated

the strongly complexing character of this ligand, even leading to the

corrosion of Ru from the solid catalyst.

In this context, X-ray absorption spectroscopy turned out to be an

important tool for elucidating the local structure of the catalysts and for

investigating the role of the solid catalysts and corresponding catalytic

species dissolved in the reaction mixture under high-pressure conditions.

For most purposes, X-ray absorption spectroscopy proves to be a technique

well applicable to the optimization of catalysts mentioned in this thesis, but

also of catalysts used in related areas. In order to gain additional insight

into the mechanism of formylation reactions and the active sites, the

formed catalytic intermediates should be investigated using additional

148 Final Remarks

spectroscopic methods, such as IR- and high-pressure NMR-spectroscopy.

Incidentally, the latter method recently helped to evidence the formation of

ruthenium hydride complexes during reaction.

Special efforts were made to find simplified preparation procedures of

heterogeneous catalysts. One solution was the modification of Ru/Al203

with dppe, another one the synthesis of coordinatively adsorbed ruthenium

complexes on modified silica surfaces. Although these simplified routes

resulted in high conversion and selectivity, a strong trend toward leaching

was observed. Only covalently anchored ruthenium phosphine complexes

on silica fulfilled the requirements for heterogeneous catalysts. In this case,

hardly any active homogeneous catalyst was found in solution. We suggest,

therefore, that simplified routes toward heterogeneous Ru catalysts should

concentrate on new strategies for developing simple covalent linkages to

the silica support.

In this work, the expansion of the use of carbon dioxide both as reactant

and as solvent was realized in two ways: firstly, the formylation of amines

with carbon dioxide could be extended to amines containing functional

groups, such as 3-methoxypropylamine, 2-ethylaminoethanol, 2-methyl-

aminoethanol, and morpholine; secondly, the extension to further reactions,

such as the synthesis of organic carbonates, but also to more sophisticated

reactions like the synthesis of vinyl carbamate from phenylacetylene,

amine, and carbon dioxide, as has been demonstrated in this work.

Most reactions described here involve hydrogen in the same quantities as

the consumed carbon dioxide. Today's hydrogen production, however, is

mainly based on fossil raw material, resulting in the fact that hydrogen

production entails the production of an equal amount of carbon dioxide as a

by-product. Therefore, it is utterly important to find ways decoupling

Final Remarks 149

hydrogen production from fossil sources. Only by achieving this aim can it

be hoped that reactions such as formylations with carbon dioxide may not

only lead to safer processes, but also help in discovering new materials and

energy loops.

List of Publications

List of publications related to the thesis

"A mesoporous ruthenium silica hybrid aerogel with outstanding catalytic

properties in the synthesis of A^,TV-diethylformamide from CO2, H2, and

diethylamine"

L. Schmid, M. Rohr, A. Baiker, Chem. Commun. 22 (1999) 2303.

"Solvent-free ruthenium-catalyzed vinyl carbamate synthesis from

phenylacetylene and diethylamine in supercritical carbon dioxide"

M. Rohr, C. Geyer, R. Wandeler, M. S. Schneider, E. F. Murphy, A.

Baiker, Green Chem. 3 (2001) 123.

(Chapter 6)

"A simple route to highly active ruthenium catalysts for formylation

reactions with hydrogen and carbon dioxide"

M. Rohr, J.-D. Grunwaldt, A. Baiker, J. Mol. Catal. A: Chem. 226 (2005)

253.

(Chapter 4)

152 List of Publications

"Formylation with supercritical carbon dioxide over RU/AI2O3 modified by

phosphines: heterogeneous or homogeneous catalysis?"

M. Rohr, J.-D. Grunwaldt, A. Baiker, J. Catal. 229 (2005) 144.

(Chapter 5)

"High-pressure in situ X-ray absorption spectroscopy cell for studying

simultaneously the liquid phase and the solid/liquid interface"

J.-D. Grunwaldt, M. Ramin, M. Rohr, A. Michailovski, G. R. Patzke, A.

Baiker, Rev. Sei. Instrum. 76 (2005) 054104.

(Chapter 3)

"Evaluation of strategies for the immobilization of bidentate ruthenium

phosphine complexes used for the reductive amination of carbon dioxide"

M. Rohr, M. Günther, F. Jutz, J.-D. Grunwaldt, H. Emerich, W. Van Beek,

A. Baiker, Appl. Catal. A: Gen. 296 (2005) 238.

(Chapter 7)

List of Conference Contributions

The following is a list of posters presented by the author at various

conferences.

"Solvent-free synthesis of vinyl carbamate from alkyne, amine, and carbon

dioxide"

C. Geyer, M. Rohr, A. Baiker, International Conference on Carbon Dioxide

Utilization, Karlsruhe, Germany (1999).

List of Publications 153

"Outstanding catalytic properties of Ru-silica hybrid-aerogel in the

synthesis of A^A^-diethylformamide from carbon dioxide, hydrogen, and

diethylamine"

L. Schmid, M. Rohr, A. Baiker, Fall Meeting of the New Swiss Chemical

Society, Lausanne, Switzerland (2001).

"Solvent-free synthesis of 3-methoxypropylformamide from amine,

hydrogen, and carbon dioxide"

M. Rohr, J.-D. Grunwaldt, A. Baiker, Fall Meeting of the New Swiss

Chemical Society, Lausanne, Switzerland (2003).

"Synthesis of 3-methoxypropylformamide using supercritical carbon

dioxide both as solvent and reactant in an autoclave at 200 bar"

M. Rohr, J.-D. Grunwaldt, A. Baiker, 42nd European High-Pressure

Research Group Meeting (EHPRG'42), Lausanne, Switzerland (2004).

ccIn Situ formation of a ruthenium catalyst in the formylation of 3-

methoxypropylformamide with carbon dioxide as reactant and solvent"

M. Rohr, J.-D. Grunwaldt, A. Baiker, Fall Meeting of the New Swiss

Chemical Society, Zürich, Switzerland (2004).

"Formylation of amines with carbon dioxide over dppe-modified RU/AI2O3

monitored by X-ray absorption spectroscopy: homogeneous or hetero¬

geneous catalysis?"

M. Rohr, J.-D. Grunwaldt, A. Baiker, User Meeting ESRF, Grenoble,

France (2004).

154 List of Publications

"Formylation of amines with carbon dioxide over dppe-modified RU/AI2O3

monitored by X-ray absorption spectroscopy: homogeneous or hetero¬

geneous catalysis?"

M. Rohr, J.-D. Grunwaldt, A. Baiker, User Meeting HASYLAB, DESY,

Hamburg, Germany (2005).

"Formylation of amines over phosphine-modified RU/AI2O3 in

supercritical carbon dioxide: heterogeneous or homogeneous catalysis?"

M. Rohr, J.-D. Grunwaldt, A. Baiker, International Conference of Energy

Technologies for a Sustainable Future (ETSF 5), PSI, Villigen, Switzerland

(2005).

Curriculum Vitae

Name Markus Rohr

Date of Birth April 10, 1972

City Holderbank (AG)

Citizen of Hunzenschwil (AG)

Nationality Swiss

Education

1988-1992

1992-1993

1993-1999

1999-2005

Alte Kantonsschule Aarau

Graduation with Matura Type C

Military service (artillery)

Rank: Captain ofNBC Defense (2002)

ETH Zürich, Chemistry Department

Chemical Engineering Studies

Graduation as Dipl. Chem.-Ing. ETH

ETH Zürich, Institute for Chemical and Bioengineering

Doctoral Thesis under the Supervision of

Prof. Dr. A. Baiker