Sie sind auf Seite 1von 34

Report

A literature study of
New Hydroformylation Techniques

Student: Ferdy de Leeuw


Studentnumber: 0579091
Education: Chemical engineering (TU/e)
Course: 6vk00

Date: 08 – 01 – 06

Pages: 30
Appendices: 2

Tutor: P. v/d Broeke


Preface
This thesis is writing for the course 6vk00 for the short education (VKO) of the Chemical
Engineering department at the TU in Eindhoven. This course is an obligate course in the
bachelor curriculum of the education. The report is a result of a literature study that has
been done and my own opinion about the studied subject.
The total size of the project equals 3 ECTS (approximately 84 hours of work).
The subject of the literature study is: New techniques of hydroformylation, the research
are performed for the Process Development group. The purpose is to describe some of the
new techniques of hydroformylation, which are investigated at research laboratory all
over the world. These new techniques are compared to the existing industrial
hydroformylation techniques.

The report is created for the Process Development group. The subject is provided and
accompanied by L.J.P. v/d Broeke. I would like to thank him and anybody else of the
Process Development group who advised me doing my research for this report.

2
Summary
The subject of this report is hydroformylation reactions. In a hydroformylation reaction
linear and branched aldehydes are formed out of alkenes (olefins) and syngas. There are
two kind of basic processes in the industry that are used for hydroformylation reactions.
Homogeneous processes, where cobalt and rhodium catalysts are used and a biphasic
process where rhodium is used as catalyst. The rhodium catalyst is modified by ligands to
form an active catalyst complex; phospine ligands used in industry are
triphenylphosphine oxide (TPPO) for homogeneous systems and meta-trisulfonated
triphenylphosphine (TPPTS) for biphasic systems.

The current industrial processes have limitations; the most important are mentioned here.
In case of the homogeneously processes, separation of the product mixture and the
catalyst complex is difficult because the mixture is in the same phase. Separation of long
chain olefins is even more difficult because distillation at high temperatures causes
degradation of the catalyst.
It is also hard to get high conversions and control selectivity’s of the products (linear
aldehydes is mostly the desired product). Especially conversion of higher olefins in
biphasic systems is difficult because of the low solubility in the aqueous phase.
Another problem is that organic solvents, used in hydroformylation, are environmental
unfriendly. Other solvent are therefore preferred.

In this thesis a description is given of new developed techniques of hydroformylation.


The goal is to find techniques that provide solutions for the most important limitations
and problems of the current industrial processes. The new developed techniques
described are:
Hydroformylation with scCO2 as solvent, this has the advantage that no organic solvents
are used. Also distillation is not needed, because a catalyst attached to a polymer is easy
to separate with a membrane. Disadvantage is that the system operates under high
pressure, which is expensive.
A modification of the biphasic system is by addition of surfactant. This creates a micellar
system, which has more potential in hydroformylation of long chain olefins. The micellar
system provides large interfacial area and larger interfacial catalyst concentration by
cationic surfactant addition.
Another modification of the biphasic system is: the supported aqueous phase catalyst,
where the catalyst is supported phase by silica nanoparticles. This makes the separation
of the catalyst easier.
A different approach of the normal homogeneously process is provided by using a
polyether phosphite ligand. A property of phospites is that it is slightly soluble in apolar
solvents. By cooling the product/catalyst mixture the catalyst can be easily separated.
These techniques all have benefit in comparison of the industrial processes, but al lot of
research is still needed to scale the processes op to commercial profitable processes.

At the end three techniques that are less well-known are described. These techniques are
hydroformylation under high pressure, with a ruthenium catalyst and with a sulfonated
xantphos ligand. The first two are not that likely to be used in commercial industry.

3
Table of contents
PREFACE ..................................................................................................................................................... 2
SUMMARY................................................................................................................................................... 3
1. INTRODUCTION .............................................................................................................................. 5
2. HYDROFORMYLATION IN GENERAL....................................................................................... 6
2.1 HYDROFORMYLATION REACTION ................................................................................................. 6
2.2 PRODUCTS AND MARKETS ........................................................................................................... 7
3. COMMERCIALLY APPLIED PROCESSES ................................................................................. 8
3.1 COBALT BASED CATALYST ........................................................................................................... 8
3.1.1 BASF process.......................................................................................................................... 8
3.1.2 Exxon process ......................................................................................................................... 8
3.1.3 Shell process ........................................................................................................................... 9
3.2 RHODIUM-BASED PROCESSES ....................................................................................................... 9
3.2.1 UCC process........................................................................................................................... 9
3.2.2 Ruhrchemie/Rhône-Poulenc process .................................................................................... 10
4. NEW TECHNIQUES AND RESEARCH AREAS ........................................................................ 11
4.1 INTRODUCTION........................................................................................................................... 11
4.2 HYDROFORMYLATION IN SUPERCRITICAL CARBON DIOXIDE (SCCO2) ........................................ 11
4.2.1 General experimental setup.................................................................................................. 12
4.2.2 Results................................................................................................................................... 12
4.2.3 Hydrofromylation in scCO2 using a polymer based catalyst ................................................ 14
4.3 MICELLAR SYSTEMS ................................................................................................................... 16
4.3.1 Experimental setup ............................................................................................................... 16
4.3.2 Results................................................................................................................................... 16
4.4 SUPPORTED RHODIUM COMPLEXES ON SILICA NANOPARTICLES ................................................. 19
4.4.1 Experimental setup ............................................................................................................... 19
4.4.2 Results................................................................................................................................... 19
4.4.3. Silica-immobilized catalyst in scCO2 .................................................................................... 21
4.5 POLYETHER PHOSHITE AS LIGAND .............................................................................................. 22
4.5.1 Experimental setup ............................................................................................................... 22
4.5.2 Results................................................................................................................................... 22
4.5.3 OPGPP in combination with ruthenium catalyst.................................................................. 24
4.6 REMAINING TECHNIQUES ........................................................................................................... 25
4.6.1 Ruthenium catalyst ............................................................................................................... 25
4.6.2 High-pressure hydroformylation .......................................................................................... 25
4.6.3 Sulfonated Xantphos ligand.................................................................................................. 25
5. DISCUSSION .................................................................................................................................... 27
5.1 HYDROFORMYLATION IN SCCO2 ................................................................................................ 27
5.2 MICELLAR SYSTEMS ................................................................................................................... 27
5.3 SILICA SUPPORTED CATALYST .................................................................................................... 28
5.4 POLYETHER PHOSPITE AS LIGAND .............................................................................................. 28
5.5 REMAINING TECHNIQUES ........................................................................................................... 28
REFERENCE LIST ................................................................................................................................... 30
APPENDIX I: PROCESS FLOW DIAGRAMS OF INDUSTRIAL PROCESSES ..................... 31
APPENDIX II: RHODIUM CATALYST CYCLE ........................................................................... 34

4
1. Introduction
The subject of this term paper is to study new techniques of hydroformylation reactions.
The goal is to project an overview of new (catalytic) processes and developments of these
types of reactions.

Hydroformylation is a reaction where olefins react with synthesis gas (CO and H2) in
presence of homogeneous catalysts to form linear and/or branched aldehydes. This
homogeneous catalyst process is one of the few applied on commercial and industrial
scale.

There is a big market for linear aldehydes and there are several problems in the industrial
applied processes. The most important problems are:
The use of organic solvents to dissolve the catalyst, organic solvents cause environmental
pollution and are toxic.
The reactant, products and catalyst are in the same phase, the catalyst must be recovered
and separated from the product after the reaction, which is difficult and is energy
consuming because the mixture is in the same phase.
Recovery of the catalyst is very important because it is an expensive metal complex bases
on rhodium, cobalt and sometimes platinum. There is also loss in activity and degradation
of the catalyst when they are used multiple times in the reaction process.
In the current processes is it hard to get high conversions and control selectivity’s of the
products, because both linear and branched aldehydes are formed, mostly the linear form
is desired. Especially conversion of higher olefins is difficult.
There is also very little knowledge of the kinetics of the reaction.
In order to improve the industrial processes and overcome the biggest problems
mentioned above, there is a lot of research and development on hydroformylation
reactions.

The report contains the following subjects. In chapter 2 a brief description of general
themes, such as the hydroformylation reaction and markets for the products is given.
In chapter 3 the most important existing industrial processes will be discussed, with their
performances and limitations.
In the 4th chapter the recent developments in hydroformylation reactions will be
introduced and discussed, by means of explanations of experiments. In chapter 5, the
final chapter, the described techniques will be compared with the existing industrial
processes and an opinion is given, whether these newly developed techniques can or will
be scaled up to commercially achievable processes or remain laboratory research.

5
2. Hydroformylation in general
In this chapter first the reaction mechanism will be described firstly and an overview of
the products will be given.

2.1 Hydroformylation reaction


The hydroformylation reaction (or oxo syntheses) is a reaction where aldehydes are
formded out of olefins, carbon mono-oxide and hydrogen. The reaction only takes place
with the use of a homogenous catalyst, usually a rhodium or cobalt based complex. An
overview of the reaction is given in figure 2.1.

Figure 2.1: hydroformylation reaction

During the homogeneous catalysis the catalyst is dissolved (often modified with ligands)
in a suitable organic solution. Another possibility is a biphasic system where the catalyst
is dissolved in one phase (aqueous) and the reactants in another phase (organic). More
details of these processes will be reported in the chapter ‘Commercially applied
processes’.

Catalysts that are used for industrial hydroformylation processes are cobalt and rhodium
based metal complexes. Also ruthenium and platinum are used as catalyst, but not for
industrial applications, they have much lower activity. Typical complexes are HCo(CO)4,
HCo(CO)3, HCo(CO)3PBu3 and HRh(CO)(PR3)3. The catalysts complexes are modified
by ligands.
The processes with cobalt and rhodium catalyst use phosphines as electron-donating
ligand. In the biphasic system a water soluble phosphine is used as a ligand, transferring
the reaction into the aqueous phase. The only classes of ligands used in industrial
hydroformylation are phosphines PR3 (R = C6H5, n-C4H9), triphenylphosphine oxide
(TPPO), meta-trisulfonated triphenylphosphine (TPPTS) and in some experimental cases
phospites P(OR)3.
The modified catalyst complexes are i.e. HRh(CO)2(TPP)2, HRh(CO)2(TPPTS)3, etc.
Many of these complexes are in equilibrium with each other in a complex catalytic cycle.
The different complexes cause different product, i.e. HRh(CO)(TPP)2 causes linear
aldehydes and HRh(CO)2TPP causes branched products. These complexes are formed out
of the key intermediate HRh(CO)2(TPP)2 as in figure 2.2 which is a part of the Wilkinson
hydroformylation cycle for modified rhodium catalysts (dissociative and associative
mechanisms, ref 10).

6
Figure 2.2: Initial equilibria forming the active catalyst species; L=TPP

2.2 Products and Markets


Since the mid 50s hydroformylation reactions have been performed on commercially
scale and over the recent years a steady and continuous growth in production capacity of
aldehydes has taken place.
In table 1 the estimate capacities for aldehydes of ethylene, propene and higher olefins
are shown.

Table 2.1: Production capacity of aldehydes by hydroformylation (estimate for 1993)


Capacity (*1000)
Region C3 C4 C5-C13 >C13
Europe (west) 25 1600 535 85
Europe (east) 785
North America 75 970 450 270
Latin America 120 55
Far East 1040 140 30
Total 100 4515 1180 385
% 2 73 19 6

The main product is butanal with a share of 73%. The formed aldehydes are used in the
bulk and specialty chemical business. They can be used to form other compounds like
amines, alcohols, carboxylic acids and ethers for instance. Main consuming industries of
aldehydes are the polymer- (especially PVC) and detergent industry.
The most desired products are in the linear form, but some products are desired in the
branched i.e. hydroformylation of styrene into a branched aldehyde is especially
important because it is a main step in ibuprofen production.

7
3. Commercially applied processes
There are three main types of commercially applied processes, the cobalt based catalyst,
the rhodium bases catalyst and the biphasic homogeneous hydroformylation.
Gas-phase hydroformylation with heterogeneous catalysts plays no role today.

Some of the important industrial processes will be discussed shortly in this paragraph.
The most important parameters will be shown. The process flow diagrams of the
described processes are added in appendix I.

3.1 Cobalt based catalyst


Cobalt-catalyzed hydroformylation is used since the 50s. Cobalt processes are mostly
used in the production of medium- to long chain olefins, because rhodium catalysts
dominate the hydroformylation of propene. The presently applied cobalt processes have
reached a high standard of performance. Most of these processes are pretty similar, the
main difference between the cobalt processes are the separation of product and catalyst.
Three of these processes are that of BASF, Exxon and Shell, they will be discussed
shortly and the most important process parameters will be shown in table 3.1.

3.1.1 BASF process


The BASF hydroformylation process of propene or higher olefins occurs under high
pressure. The catalyst is in the form of HCo(CO)4. This catalyst will be separated from
the liquid product by addition of oxygen and formic or acetic acid, leading to an aqueous
solution which contains the cobalt mainly as formate or acetate. The organic products are
withdrawn in a phase separator (6) and the cobalt solution is concentrated afterwards and
sent to the carbonyl generator (2). The cobalt losses are compensated. In the reactor (3)
stirring is important to achieve thorough mixing of the olefins and the aqueous catalyst
solution. The best selectivity to linear aldehydes is claimed for low temperatures.

3.1.2 Exxon process


The Exxon process is designed to convert olefins in the range of C6 – C12. Recovery of
the catalyst is different than in process where the catalyst is recovered by oxidation (like
the BASF process). In the Exxon process the ‘Kuhlmann catalyst cycle technology is
applied. This involves two main steps: the recovery of sodium carbonylate and its
regenerative conversion into cobalt carbonyl hydride. The HCo(CO)4 catalyst reacts with
syngas in the reactor under normal hydroformylation conditions. After the reactor the
product mixture is treated with aqueous alkali, to convert HCo(CO)4 to water-soluble
NaCo(CO)4, which is extracted as aqueous solution from the organic product phase. Then
the catalyst is regenerated by addition of H2SO4.

The benefit of the Exxon process is that the catalyst does not undergo decomposition and
enters the reactor in its most active/effective form. A big disadvantage is that catalyst
separation and recovery must be carried out under CO pressure to preserve the catalyst.

8
3.1.3 Shell process
In the Shell process reactants in range of C7 – C14 are converted using a phosphine-
modified cobalt catalyst. In this process the product mixture is distillated, the organic
products leave the distillation column at the top and the catalyst is recovered at the
bottom. Before re-entering the reactor the catalyst recycle is upgraded with catalyst and
phosphine ligand. The benefits of this process are the high n/i ratio, the low pressure and
the direct formation of alcohols. The disadvantage is the low activity of the ligand-
modified catalyst, which requires a large reactor volume.

Table 3.1: process parameters of cobalt catalyzed processes


Basf process Exxon process Shell process
Parameter Unit Range Range Range
o
Temperature C 120 – 160 175 150 – 190
Pressure MPa 27 – 30 29 – 30 4–8
CO/H2 ratio 1:1 1:1.16
phosphine/cobalt ratio 1-3 : 1
Feedstock 1-octene Propene trimer
Co concentration wt% of feed 0.4 – 0.7 0.3 0.5 – 1.0
n/i ratio Higher at low T. 85 – 90 %
Product composition wt%
n-1-octene 8 – 12
Light products 3–5 11 – 13
C9 aldehydes 70 – 75 72 – 74
C9 alcohols 6 – 10
Heavy ends 4–6 13 – 17

3.2 Rhodium-based processes


Rhodium catalyst processes are used since the 70s and dominate hydroformylation of
propene. Rhodium catalysts are more expensive than cobalt catalysts and have higher
activity, but have lower activity in case of branched olefins. Problems using rhodium in
long chain olefins is the catalyst recovery. For example: the product aldehydes beyond
C7/C8 have increasing molecular weight and higher boiling points, separation by
distillation becomes commercially unattractive and loss of stability due to thermal strain
results in degradation and loss of catalyst.
A very elegant solution is the two-phase system, where the catalyst and products are in
two different immiscible phases. The main advantage of this process is that the products
and catalyst can be separated easily due to immiscibility of aqueous and organic phase.
The most important industrial processes will be discussed. Process parameters are listed
in table 3.2.

3.2.1 UCC process


The Union Carbide Corporation (UCC) commercially applies the hydroformylation of
propene in a liquid-recycle process. These plants are called LPO (Low Pressure Oxo)
plants. The reaction takes place in a stainless steel reactor where the gas and propene are
introduced via a feed line and a gas-recycle. The catalyst is dissolved in high-boiling
aldehyde condensation products. The product mixture out of the reactor consists of

9
dissolved gas, aldehydes, rhodium-phospine complex, free phosphine ligand and the
higher-boiling aldehyde condensation products. The product mixture enters a separator
and a flash evaporator, where the major part of inerts and unconverted reactants is taken
overhead. The flashed-off gases are returned to the reactor. The liquid stream is heated
and is fed to two distillation columns in series. The gaseous aldehydes are sent over the
top and will be condensed and separated from the CO/H2 gas, which is recycled. At the
bottom the catalyst solution is separated and recycled in the reactor. If the feed of the
process has a sufficient purity the catalyst may last more than a year. The BASF and
Mitsubishi rhodium based processes are quite parallel to the UCC process and are
therefore not further discussed.

3.2.2 Ruhrchemie/Rhône-Poulenc process


The RCH/RP (Ruhrchemie/Rhône-Poulenc) process is an example of a two phase system.
Propene, n-pentanal and n-butene are produced on a commercial scale in a continuously
stirred tank reactor followed by a phase separator and a strip column. The reactor
contains the aqueous catalyst and is fed with propene and syngas, TPPTS is used as a
ligand for the rhodium catalyst complex. After the reactor the crude aldehyde product is
degassed and separated into the aqueous catalyst solution and the organic aldehyde phase.
The heat of the aqueous phase is then used to produce steam in a heat exchanger. After
separation the organic phase is passed through a stripping column, where the unreacted
olefins are separated and sent back to the reactor. The product mixture is then distilled
into n- and isobutanal (linear/branched). The produced steam from the reactor is used in
the reboiler of the distillation unit, which is a big advantage. This is the cleanest
hydroformylation industrial process because of the straightforward separation of the
organic products from the catalyst. However the system is limited by the solubility of
organic substrates in aqueous phase. The rate of higher olefins hydroformylation drops
dramatically because of their low solubility in water.

Table 3.2: process parameters of rhodium catalyzed processes


UCC process RCH/RP process basf process
Parameter Unit Range Range Range
o
Temperature C 85 – 90 110 – 130 110
Pressure MPa 1.8 4–6 1.5 – 1.7
CO/H2 ratio 1:1.07 1: 0.98-1.03 45:55
Rh concentration ppm 240 - 270 160 – 190
TPP wt% 11 – 12 3.5 – 4.5
Propene conversion % 85 – 89 85 – 99 84 – 86
n/i ratio 93/7 – 97/3 84/16
Productivity Mol/L.h 1.5 – 2.0 0.063 – 0.0721
Selectivity to C4 aldehydes wt% 99
1
productivity in kg/L.h

10
4. New techniques and research areas
4.1 Introduction
In this chapter some of the main scientific researches on hydroformylation reactions are
described. Three different kinds of research topics for hydroformylation can be pointed
out. In the first group, studies dealing with the properties of the solutions (used to run the
reaction) are described. In the second and third group research to different catalyst
complexes and ligands are described.

Current industrial processes have reached high standards of performance but uses
traditional techniques, researchers try to find novel techniques which leads to more
sustainable processes. The most important reasons to do research on hydroformylation
reaction are already pointed out in the previous chapters, but they will be mentioned
again for the complete overview.
The current solutions used in the industrial processes are organic solutions, which cause
environmental pollution. Therefore scientists try to find different solutions; preferable
non organic solutions i.e. super critical CO2. Also organic solutions are still studied trying
to improve reaction rates, selectivity and the hydroformylation of higher olefins.
A very important problem is the recovery of the catalyst. The rhodium and cobalt
catalysts are very expensive, so it is very important to maintain the catalyst. The current
processes can recover the catalyst, but several separation units (mostly distillation) are
needed which have limitations and are energy consuming. Research of new catalyst,
ligands and solution and/or new kind of processes is done to have better and easier
catalyst recovery.
The hydroformylation of higher olefins is more difficult then propene for instance. In the
current processes the higher olefin hydroformylation have limitations. In case of the
biphasic process, the heavier olefins have less solubility in the aqueous catalyst phase,
which cause a dramatically drop of reaction rate. Mostly hydroformylation of higher
olefins is carried out in a single organic phase. The catalyst is separated by distillation,
which may result in catalyst decomposition, because of high boiling points of the heavier
products. Therefore much research is done to improve reactions of the higher olefins.

In this chapter new solutions for hydroformylation are discussed. The research subjects
described are: super critical carbondioxide, micellar systems

4.2 Hydroformylation in supercritical carbon dioxide (scCO2)


A major problem of hydroformylation reactions are the use of organic solvents, which
cause environmental pollution. Another problem is the catalyst recovery after the
reaction. Hydroformylation in supercritical CO2 as a solvent is one of the possibilities
that can solve these problems.

Advantages of using scCO2 as solvent in homogeneous catalysis are; it is inert to most


reactions, non-toxic, cheap, non-flammable and it is readily available. Both cobalt and
rhodium catalyst are used in the experimental research.

11
First a description is given of the hydroformylation of propylene in scCO2 using a cobalt
catalyst (ref 3). The general setup of the experiment is explained in the next paragraph
and in figure 4.1.

4.2.1 General experimental setup


The reaction in scCO2 occurs under high pressures, experimental data show pressures of
approximately 100 to 241 bar. So the main part of the system is a stirred high pressure
reactor (4) as in figure 4.1 of app. 100-400 mL. The gasses (or liquid in case of higher
olefins) needed for the reaction are: CO and H2 and propylene (9,10) which are of high
purity, CO2 is used as solvent (1), these components are introduced into the reactor using
high pressure pumps and valves. The reactor is pressurized by pumping (2) CO2 into the
reactor. To control the experiments the reactor is equipped with a pressure gauge (6), a
temperature monitor (12) and a gas sample trap (8). The catalyst in the described
experiments is Co2(CO)8 (octacarbonyldicobalt), which is in reality the precatalyst.

Figure 4.1: experimental setup of hydroformylation in scCO2

4.2.2 Results
Some important experimental results are shown in this paragraph. The pressure and
temperature are important parameters of the reaction and influence the conversion. In
figure 4.2 the conversion is shown as a function of reaction time at constant temperature
of 88 oC with different pressures. 1 g/batch of catalyst was used, in 282 mL reactor
volume. The result is an increase of the conversion with increasing pressure at a given
reaction time. This increase in conversion levels off at about 2100 psig (144.8 bar).

In the experiments of figure 4.3 the opposite is done, by changing the temperature at
constant pressure of 2400 psig. The result is an increase of the conversion by increasing
the temperature.

12
Figure 4.2: Propylene conversion at different Figure 4.3: Propylene conversion at different
pressures, T= 88 oC. temperatures, P= 2400 psig

The experiments prove that homogeneously catalyzed hydroformylation occurs in scCO2.


Most commercial aldehydes are desired in the linear form, the next logical dependence to
examine is, whether a good selectivity is obtained and whether the selectivity can be
controlled. This means that the linear to branched ratio (L/B) must be high. In table 4.1
the result are shown. The best L/B ratios are obtained at low temperatures and high
pressures.

Table 4.1: Reaction selectivity at different pressures and temperatures with 1.0 g of catalyst/batch (ref. 3)
Pressure Temperature L/B ratio
o
Psig C
2400 78 4.2
2400 88 4.1
2400 98 3.1
2400 108 2.7
1350 88 2.7
1650 88 3.0
2100 88 4.2
2700 88 4.3
During the experiments a possible reaction rate constant is found out of proposed kinetics
from the literature (ref. 3). There is a pseudo-first order rate constant observed, the rate
constant at a constant temperature is a function of hydrogen- and carbon mono-oxide
pressure and the present catalyst. A proposed empirical reaction rate equation is:
PH
r = K ⋅Cp K = kWcat 2
PCO
Where Cp is the concentration of propylene and k is the reaction rate constant, which
depends on pressure. In literature different kind of kinetic models are used to describe
reaction rates, see ref. 4 for another example.

Reaction rate and selectivity can be changed by tuning the pressure and temperature, but
theoretical explanations of the observed results are not given.
There is also another problem. There is still no solution presented to separate the product
and recover the catalyst. In paragraph 4.2.3 and 4.4.3 possible solutions are described.
4.2.3 Hydrofromylation in scCO2 using a polymer based catalyst
In the previous paragraphs it is proven that hydroformylation in scCO2 works. Still there
is now convenient solution to separate the catalyst from the product mixture.
A new approach of catalyst recovery in scCO2 systems is the use of a polymer based
catalyst. In this case the catalyst complex is chemically bound to a polymer. This polymer
based complex is easy to separate by a membrane because of its large size compared to
the products and remaining reactants.

The case discussed in this paragraph is the hydroformylation of styrene into a branched
aldehyde, with a polymer supported rhodium catalyst (ref. 5) This process is a main step
in the production of ibuprofen.

4.2.3.1 Catalyst preparation


The polymer chosen to support the Rh-complex is a fluoroacrylate polymer, which is
used in most of the experiments. This polymer is soluble in scCO2. A possible catalyst
employed in these experiments is a Wilkinson-type catalyst. First, the polymer is
synthesized by polymerization of 1H,1H,2H,2H-heptadecafluorodecyl acrylate monomer
(zonyl TAN) and N-acrylosuccinimide (NASI), the former increasing the solubility in
scCO2 the latter providing attachment sites for the catalyst. Then, the succinimide group
is exchanged by phosphines using NH2(CH2)3PPh2 (DPPA). During this step, 15–30% of
the phosphines are oxidized due to an uncontrolled reaction, as was estimated by NMR
analysis. Finally, this precursor is reacted with [RhCl(COD)]2 to obtain the catalyst. The
catalyst composition will be about 4 – 5 x10-5 mol of Rh g-1 of catalyst. The ideal
molecular structure of the catalyst is presented in Fig. 4.4 for 13:1 ratio of zonyl Tan to
NASI.
Hydroformylation occurs with different catalysts i.e. Rh(TAN7DPPA)3Cl or
Rh(TAN13DPPA)3Cl, where TAN is the fluoracrylate component of the polymer and
DPPA is the phospine ligand.

Figure 4.4: catalyst structure Figure 4.5: Formation of the catalytic active
species for the hydroformylation of olefins.

14
During the reaction several reactions have to occur before a catalytic cycle starts and the
key intermediate HRh(CO)2L2 (fig. 2.2) is formed. The overall process is given in figure
4.5. It is proposed that the reaction mainly takes place through the dissociative reaction
mechanism, see appendix II.

4.2.3.2 Results
The experimental setup to perform the reaction is comparable with figure 4.1 only a
membrane is placed behind the reactor, to separate the catalyst.
The experiments are reproducible and have a conversion of about 90%, with high
selectivity to branched aldehydes as shown in figure 4.6. In rhodium catalyzed
hydroformylation of styrene, selectivity to branched aldehyde increases as the
temperature decreases. This is due to the stability of the benzylic rhodium species. The
selectivity to branched aldehyde was about 85% at 348 K and nearly 100% at 323 K, both
at 241 bar. An increase of pressure also induces an increase of branched aldehyds with
constant temperature. In most of the used reaction conditions (T: 323 – 348 K, P: 172 –
241 bar) the catalyst was totally soluble, TAN13 more than TAN7.
At total overview of the reaction studies is summarized in table 4.2.

Figure 4.6: styrene hydroformylation reproducibility at 348 K and 172 bar using Rh(TAN13DPPA)3Cl
catalyst.

Table 4.2: Hydroformylation of styrene catalyzed by Rh(TAN13DPPA)3Cl complex in scCO2


Entry P (bar) T (oC) Time (h) Conversion (%) BA1 LA2 EB TOF (h-1)3
1 241 50 7.5 100 1.0 - - 16
2 241 75 24 98.3 0.86 - 0.14 15
3 207 50 23.5 98.7 1.0 - - 13
4 207 75 10 97.5 0.88 0.08 0.08 23
54 207 75 2.5 80.6 0.99 - 0.01 64
6 172 50 23 78.0 0.96 0.04 - 8
7 172 75 31 98.0 0.99 - 0.03 6
Reaction conditions: styrene/rhodium molar ratio: 200, initial mole fractions: CO/H2=0.127, styrene= 4.13x10-4, CO2=0.872
1
BA: branched aldehyde (moles of branched aldehyde produced.mol of styrene reacted)
2
LA: Linear aldehyde
3
Average turnover frequency, mole of aldehyde per mole of rhodium per hour
4
Initial mole fractions: styrene= 8.22x10-4, CO/H2= 0.227, CO2= 0.772

Hydroformylation of linear olefins with polymer based catalyst are also studied with
good results. Conversions are up to 100% and there is no activity decrease and rhodium
leaching. For further details see ref 6, this is not further described in this report.

15
4.3 Micellar systems
Hydroformylation in a biphasic system is already used on industrial scale. Still a lot of
research and development is performed on biphasic systems, one of those developed
techniques are micellar systems. The technique consists in adding appropriate
amphiphilic agent (surfactant or detergent) to a biphasic liquid system. The aqueous and
organic phases form a single stable micellar medium (micro emulsion) with a continues
and dispersed phase. The amphiphilic molecules have a hydrophobic carbon chain and a
hydrophilic head group. Using an amphiphilic ligand it is possible to dissolve organic
molecules into water by creating micelles. The technique is based on transfer of organic
groups into the aqueous phase or to the interface where the catalyst is present.
This has two major advantages; firstly, micelles can dissolve molecules which are
normally not soluble in the bulk solvent. Secondly, there is an enormous interface
between continuous and dispersed phase due to the very small size of micelles (10 – 100
nm). On the other hand there is also a problem, because the stable micellar phase must be
separated into two phases at the end of the reaction, in order to recover the aqueous
catalytic phase.
The experimental results of rhodium based hydroformylation of 1-dodecene with TPPTS
as ligand (ref. 7) are described.

4.3.1 Experimental setup


A typical experimental setup for the reaction is with a magnetic stirred stainless steel
autoclave of 100 – 500 ml. The autoclave contains the following species: rhodium
catalyst, TPPTS ligand, anionic-, cationic- or non-ionic surfactant or alcohol, water, 1-
dodecene (or other olefins) and heptane. Before the experiments the autoclave is purged
with syngas for several times, before it is heated to the desired temperature. The pressure
is kept constant during the reaction runs by adding syngas. After a given reaction time,
stirring is stopped and the autoclave is quickly cooled. Products are analyzed by gas
chromatography and surface tensions are measured by maximum bubble pressure
method.

4.3.2 Results
The experimental results are described in this paragraph, three kinds of experiments can
be performed; the first with the different surfactants, the second with alcoholic solvents
instead of surfactants and the third without any additives. The results are listed in table
4.3. Hydroformylation did not occur in absence of surfactant or alcoholic solvent.
The addition of cationic surfactants or alcohol improved the reaction rate of 1-dodecene.
Non-ionic surfactants did not enhance the reaction rate very obvious; this is in contrast of
some other reports in literature. Anionic surfactants inhibit the reaction.

The increase of the reaction rate by adding alcohols is probably caused by changes in
solubility. The solubility of olefins in the aqueous phase or of the catalyst complex in the
organic phase is than increased. A disadvantages effect is that a part of the catalyst can
not be recovered, when it leached into the organic phase.

16
Increase in reaction rate by addition of cationic surfactants can be described by two
factors. The first one is the formation of micelles, which cause an increase of the
interfacial area between the phases and it brakes the phase barriers. The second reason is
that the cationic ends of the micelle are orientated to the aqueous phase; this forms a
positive charged ion layer. This layer attracts the anionic active rhodium/TPPTS
complex, HRh(CO)[P(m-C6H4SO3)3]33n- (figure 4.7). This means that the catalyst is
highly concentrated at the interface and more easily coordinated with the olefins in the
micelle. Other experiments are therefore performed with cationic surfactant CTAB,
which is pretty common in other researches (ref 10 i.e.).

Micelles are formed in addition of anionic and non-ionic surfactants, causing increase of
interfacial area, only no high reaction rates were established. In case of non-ionic
surfactant the catalyst concentration at the interface does not increase. In case of the
anionic surfactant the negative charged ion layer of the micelle could not contact with the
negative charged catalytic active species as result of the static repulsion. This causes a
very low catalytic concentration in the interfacial layer.

Table 4.3: Biphasic hydroformylation of 1-dodecene with different surfactants and alcoholic solvents.
Additive Nil. Tween Span SDS CTAB DTAB BDAC Diglycol ethanol propanol
20 40 methylether
Yield (%)1 0 3.8 3 0 61.3 60.9 44.8 12 24.5 42.3
n/i ratio - - - - 6.1 5.6 6.3 3.9 3.4 3.6
Reaction conditions: Rh =6.4x10-4 mol/l, TPPTS/Rh =16, 1-dodecene: 22.6 mmol, heptane: 5 ml, water 20 ml,
surfactant: 5.5x10-3 mol/l, T=100oC, P=0.5MPa, reaction time 120 min. (with alcohols 10 ml of water)
1
conversion of olefin to aldehydes
CTAB: cetyltrimethylammonium brom., DTAB: dodecyltrimethylamm. brom., BDAV: benzyltetradecyldimethylamm. Chloride.
Tween20: polyoxyethylene(20) sorbitan monolaurate, Span40: sorbitan monopalmitate, SDS: sodium dodecyl sulfate.

Figure 4.7: sketch map of catalytic active species in the interfacial layer of cationic micelle

The addition of cationic surfactants is also favourable for linear aldehydes. The micelle
structure creates an orderly and compact micro environment where olefins are solubilized
and coordinated with catalyst complexes, favourable for linear aldehydes. This

17
hypothesis is controlled with the following explanation. The stirring rate is changed,
because this would cause variation in the reaction micro circumstances, which could
influence the reaction rate and selectivity. Vigorous stirring will disturb the meta-stable
micelle structure. This hypothesis is confirmed in an experiment, see table 4.4. The rate
constant k of the reaction increases with increasing stirring rate, the selectivity however
decreases. Higher stirring rate is better for the transfer of substrates to the interface, but is
a disadvantage for the formation of micelles.

Table 4.4: Stirring rate influence


TPPTS/Rh 30 15
Stirring rate (rpm) 300 400 600 400 600
k x103 (/min) 6.33 8.26 8.66 9.40 9.69
n/i ratio 7.6 5.4 4.1 4.6 3.4
Reaction conditions: Rh =9.6x10-4 mol/l, 1-dodecene: 56.5 mmol, heptane: 12.5 ml, water 50 ml, CTAB surfactant:
5.5x10-3 mol/l, T=90oC, P=1 MPa

The cationic surfactant (CTAB) concentration is an important parameter on the activity


and selectivity of the hydroformylation (figure 4.8). In this example toluene is used
instead of heptane, because the heptane caused emulsification with higher CTAB
concentrations. Until a CTAB concentration of about 4.5x10-3 mol/l there is a clear
increase of conversion and linear/branched ratio.

: conversion
: n/i

Figure 4.8: Influence of CTAB concentration.


Reaction conditions: Rh =3.53x10-4 mol/l, P/Rh= 30, 1-dodecene: 22.6 mmol, toluene: 5 ml, PH2:PCO = 1.2 T=90oC,
P=0.5 MPa, time: 60 min.

The surface tension of the reaction mixture is also measured; at a CTAB concentration of
about 0.6x10-3 mol/l a major decrease in surface tension is measured. This corresponds
with the critical micelle concentration (CMC) and explains the increase in conversion as
seen in figure 4.8.

An important issue of this system is that afterwards organic product phase and aqueous
catalyst phase must be separated. Thus the emulsification of the micellar phase must be
reversible. In the research of micellar hydroformylation systems with the common used
surfactant CTAB the descriptions of solutions for the catalyst product separation are very
limited in literature.
Surfactants that fulfil that requirement are polyoxyethylenes (ref 9). The separation
occurs when the mixture is cooled. The stable emulsion separates into an aqueous phase
and organic phase, when the system is cooled.

18
4.4 Supported rhodium complexes on silica nanoparticles
Another solution to separate the catalyst in hydroformylation is provided by
nanoparticles. However literature does not provide much research, it can be a good
alternative for the current processes. In this case the catalyst is supported by non-porous
fumed-silica nanoparticles for SAPC (supported aqueous phase catalyst), see ref. 11,12
and 13. Highly dispersed fumed silica’s are normally used as fillers, additives and
medicinal- and industrial adsorbents.
Supported aqueous-phase catalysis shows high capacity in the conversion, selectivity and
also the easy recovery of the catalyst from the organic phase. In this system, the catalyst
is immobilized in a thin water layer adhered within the pores of high surface area
silicates. Using this system, one can convert higher olefins at relative high rates, without
metal leaching (loss of catalyst).
An important problem however is the hydration of the catalyst. The fumed silica particles
have low amounts of hydroxyl groups. In an aqueous suspension the active sites of the
fumed silica form strong hydrogen bonds with water molecules.

4.4.1 Experimental setup


During the experiments a rhodium catalyst is used with TPPTS as ligand, the active
complex HRh(CO)(TPPTS)3 was synthesized (ref. 11). Interaction between water-soluble
and/or its corresponding rhodium complex and silica surfaces occurs due to hydrogen
bonding of =SiOH. Two kinds of commercially available silicas (non-porous fumed- and
porous granular-SiO2) have been used as supports. The SAPC was prepared with the
known method of self-assembly in an autoclave.
In a typical run, the catalyst complex, TPPTS , SiO2 and deionised water were mixed and
placed in the autoclave. Then 1-hexene and toluene were added before closing the
autoclave. The reactor is than pressurized with CO/H2 (1:1 vol%) and heated to the
reaction temperature. After the reaction the catalyst and the liquid phase were separated
by filtration. The organic phase was analysed with a gas chromatograph.

4.4.2 Results
Several experiments have been performed in literature and the most interesting results are
highlighted in this paragraph (ref. 11).
The performances of 1-hexene hydroformylation using SAPC are shown in table 4.5. For
comparison, the results of the conventional biphasic catalyst system have been listed also.
Specifications of the experiments are given in the table.

After reaction there existed a slurry aqueous catalyst layer, containing non-porous fumed
silicas, Rh-complexes, uncoordinated TPPTS and water. The upper layer of products was
separated from the slurry catalyst layer by decantation.
In all cases, the selectivity towards heptanal was approximately 100%. The use of fumed-
SiO2 SAPC shows good catalytic performance, it is comparable with the conventional
biphasic system, with CTAB surfactant. The silica particles are defined as FA, GA or
GB, the properties of these particles are listed in table 4.5. Porous SiO2 (GB) and large
particles with relative small specific surfaces (SiO2-GA) however shows a negative effect
on the reaction rate.

19
The rhodium content in the organic phase in the case of the fumed-SiO2, on the other
hand was about one magnitude lower than that in the conventional system (compare first
and last experiment, rhodium leaching).
The experiments are performed at an agitation speed of 400 rpm. The reaction rate
increased linearly with agitation speeds between 100 – 400 rpm. Above 400 rpm the
reaction rate is not influenced.

Table 4.5: Hydroformylation of 1-hexene catalyzed by (un)supported aqueous phoshing-Rh complex


Support Pressure (MPa) Time (h) Conv. (%) n/i ratio TOF (h-1) Rh leaching (mol.l-1)
1
SiO2-FA 4.0 4 72.9 3.2 454.4 4.9 x 10-5
1.0 2 44.8 5.6 560.4 -
SiO2-GA2 4.0 4 5.0 2.6 31.1 -
SiO2-GB3 4.0 4 34.5 3.1 215.1 4.9 x 10-6
1.0 2 22.2 5.7 277.4 -
Silica colloidal4 4.0 4 7.3 2.6 45.6 -
5
4.0 4 16.2 5.5 101.4 7.8 x 10-6
5
+ethanol 4.0 4 66.3 3.2 414.6 1.9 x 10-5
5
+CTAB 4.0 4 95.1 3.1 594.1 1.6 x 10-4
Reaction conditions: Rh: 2×10−5 mol.g−1 SiO2, TPPTS/Rh: 50 mol/mol, support hydration = 64.0 wt.%, olefine/[Rh]: 2500 mol/mol,
toluene/1-hexene: 80/20 (wt/wt), T = 373 K, CO/H2: 1/1 (v/v), 4.0MPa, agitation speed = 400 rpm.
1
SBET = 361 m2 g−1, particle mean diameter: 16 nm.
2
SBET = 781 m2 g−1, particle diameter: 110–180 nm.
3
SBET = 400 m2 g−1, pore volume: 1.09 ml g−1, particle diameter: (1–1.5) × 105 nm.
4
Colloidal silica contains 25 wt.% SiO2.
5
Biphasic catalyst system, water/oil: 3/1 (v/v), H2O = 36 ml (ethanol: 17.1 ml, CTAB: 0.121 g)

Further testing is performed with the non-porous fumed silica. The influence of structure
and particle size was investigated. Smaller particle size and higher surface areas are
beneficial for the reaction rate and the linear/branched ratio, table 4.6.
Table 4.6: comparison of 1-hexene hydroformylation with several kinds of fumed-SiO2-SAPC
Support SBET (m2/g) Mean diameter (nm) Conversion (%) n/i STY (mmol/h gRh) TOF (h-1)
SiO2-FA 361 16 69.3 3.3 4210.5 433.3
SiO2-FB 193 12 62.2 3.0 3777.2 388.7
SiO2-FC 393 14 72.2 3.3 4399.0 452.7
SiO2-FD 361 17 47.7 3.4 2894.6 297.9
SiO2-FE 232 13 53.3 3.1 3234.8 332.9
SiO2-FF 189 16 51.2 3.3 3112.5 320.3
Reaction conditions: T=373K, CO/H2= 1/1 (v/v), P= 4 Mpa, [Rh]= 4x10-5 mol, TPPTS/Rh = 10 mol/mol, agitation speed= 400 rpm,
SiO2= 2 gr. support hydration = 50 wt%, 1-hexene = 0.1 mol, toluene = 38.0 ml, reaction time= 4 h.

Other important parameter are the olefin/Rh molar ratio (maximum reaction rate is
reached 2500 at 373 K ant 4 MPa) and the pressure. The olefin/Rh ratio has a small
influence on the n/i aldehyde ratio.
This suggests that the adsorption ability of the fumed silica in respect to the TPPTS-Rh
complexes and TPPTS would be conceivably high.
A drop of the total pressure cause the decrease of syngas concentration and the relative
ratio of syngas/phospine concentration in the catalyst layer, which is an advantage for the
competitive coordination of phosphine rather than CO on rhodium atom. This is
favourable for the formation of active rhodium species with more phosphine ligands.
Lower pressure thus results in lower conversion but better micro-circumstances to form
linear aldehydes, as in figure 4.9.

20
Reaction conditions: L/Rh = 50 mol/mol, P=1
MPa, reaction time 2 h., others same as in table
4.6.

:conversion, :n/i.

Figure 4.9: 1-hexene hydroformylation performance of fumed-SiO2-FA-SAPC as a function total pressure.

Higher catalytic performance comparable to that of the conventional biphasic system was
achieved by using non-porous fumed SiO2-SAPC for 1-hexene hydroformylation. The
rhodium content (by catalyst leaching) in the organic phase was about one magnitude
lower than that in biphasic system.
The SAPC system provides a possibility that the catalyst can operate efficiently in a wide
range of support hydration range from 40 to 64 wt%, keeping a high reaction rate and
selectivity.
The structure of support has been confirmed to be the key issue in the performance of
SAPC. The SAPC prepared by using fumed SiO2 with smaller particle size and larger
surface area provided beneficial contribution to the catalytic performance.

4.4.3. Silica-immobilized catalyst in scCO2


There are also a few reports of a silica-immobilized catalyst used in a scCO2 system. In
this report no detailed description is given, because too little information is found to give
a complete overview of this research area. In reference 13 a small description is given of
these kind of experiments.
The catalyst used in the experiment is the rhodium complex of N-(3-trimethoxysilane-n-
ptpropyl)-4,5-bis(diphenylphosphino)-phenoxazine immobilized on silica (particle size
200–500 nm), as in figure 4.10. Typically 1 g of silica with a 0.4% Rh loading was
loaded into a 5 ml supercritical flow reactor.

Figure 4.10: Ligand complex used in scCO2 silica-immobilized catalyst


4.5 Polyether phoshite as ligand
In the most hydroformylation processes TPPO or TPPTS (for biphasic systems) are used
as ligands. A new approach for the hydroformylation of higher olefins is with polyether
phospite (OPGPP) as ligand in a non aqueous system. The use of phosphite has received
much attention because of its great improvement in activity compared to the general used
TPP, especially for long linear olefins. The hydroformylation of 1-decene, with a
rhodium catalyst and OPGPP will be described in this chapter (ref 14).

The research group as in ref. 15 reported a novel polyether phosphite (OPGPP) to


facilitate aqueous 2-phase hydroformylation of water immiscible higher olefins and
styrene. But there was great loss in activity after a couple of successive runs, because of
hydrolysis of phosphite in water. Therefore they started a research of hydroformylation of
higher olefins with OPGPP in the absence of water (ref. 14). That means that they studied
a homogeneously organic reaction. Polyether phosphines are slightly soluble in some
apolar solvents such as toluene and heptane at room temperature. This provides a good
approach for catalyst separation by decantation.

4.5.1 Experimental setup


The hydroformylation experiments were performed in a 75 ml stainless steel autoclave
equipped with magnetic stirrer. Hydroformylation catalysts were formed in situ from
catalyst precursor, ligand, solvent and internal standard. The system was checked for
leaks and was pressurized with CO/H2 (1/1) and brought to the required temperature in a
thermostatic oil bath. After appropriate reaction time, samples were taken and analyzed.
The prepared OPGPP ligand has a structure as in figure 4.11.

Figure 4.11: OPGPP ligand

4.5.2 Results
The main goals are to maintain a high hydroformylation activity and a good separation of
the catalyst after the reaction. Therefore the solubility of different ligands in organic
solvents is determined at room temperature. These results are summarized in table 4.7.

Table 4.7: Solubility of OPGPP in heptane and toluene at ambient temperature1


Polyether chain length of Solvent Temperature (oC) Solubility ( g P/ml Ligand solubility in
OPGPP(n) solvent solvent (%)2
12 Heptane 16 154.5 14.72
12 Toluene 16 - Fully soluble
19 Heptane 16 42.41 5.55
19 Toluene 16 - Largely soluble
34 Heptane 16 9.0375 Partly soluble
58 Heptane 16 0.3250 0.045
58 Toluene 16 171.7625 27.35
58 CH2Cl2 16 - Fully soluble
58 THF 16 - Fully soluble
1
2.054x10-4 mol OPGPP/6.0 ml organic solvent.
2
Ligand solubility: P content en solvent/total P addition

22
It can be concluded that OPGPP with increasing molecular weight (with R = 58, see
figure 4.11) has decreasing solubility in heptane. From the point of catalyst recovery,
apolar solvent such as heptane is preferred for non-aqueous homogeneous catalysis.
This is confirmed in experiments with other solvent than heptane. The solvents toluene,
1,4-dioxane and anisol give high conversions as heptane. The colour of the reaction
mixture with toluene and 1,4-dioxane and anisol turned however out to be slightly
yellow, indicating loss of catalyst in the reaction mixture. This confirms the choice of
heptane as solvent. This is probably the result of different solubility of phosphines in the
solvents.

At higher temperatures the solubility of OPGPP increases from 70 oC, at a normal


hydroformylation temperature of 90 oC, 3.33% of OPGPP dissolves in heptane. An
increase in reaction temperature therefore induces an increase in reaction rate, table 4.8.
An interesting observation is that in time the ligand dissolves fully in heptane with the
formation of aldehydes. Somehow homogeneous and heterogeneous catalysis takes place
in competition at the beginning of the reaction, but when aldehydes accumulate to some
amount the reaction becomes homogeneously
Table 4.8: Effect of temperature on the reaction
Temperature (oC) Conversion (%) Aldehyde yield (%) Aldehyde n/iso
60 31.4 25.2 1.50
70 80.0 77.7 1.31
80 95.1 94.2 1.04
90 99.4 97.6 0.94
100 99.7 95.3 0.72
Reaction conditions: 1-decene 2.0 ml., heptane 4.0 ml., internal standard 0.3 m, Rh(acac)(CO)2 1.058x10-5 mol, OPGPP/Rh =13,
P(CO)/P(H2) = 1:1, syngas pressure 5.0 MPa, reaction time 4 h.

Another important parameter is the ratio phosphine/rhodium. The best results for the
hydroformylation of 1-decene were obtained at a molar ratio >13. Below 13 the
conversion decreases and there will be loss of catalyst, indicated by the colour of the
mixture. Addition of TPP (triphenylphosphine) caused an increase in the n/iso ratio,
which is preferred for most products. This seems no good solution to raise the n/iso ratio
because a slight loss of catalyst was induced, because TPP dissolves more in the solution.
See table 4.9 for the results.

Table 4.9: Effect of OPGPP/Rh (mol) ratio and triphenylphosphine on the reaction
OPGPP/Rh (mol) PPh3/Rh Conversion (%) Aldehyde yield (%) n/iso ratio
3 75.3 73.5 0.94
7 90.0 88.3 0.96
13 99.4 97.6 0.94
30 99.7 98.9 0.99
3 3 54.3 53.2 0.98
13 4 94.0 93.1 1.3
13 13 95.5 94.6 1.8
Reaction conditions: 1-decene 2.0 ml., heptane 4.0 ml., internal standard 0.3 m, Rh(acac)(CO)2 1.058x10-5 mol, P(CO)/P(H2) = 1:1, T
=90 oC, syngas pressure 5.0 MPa, reaction time 4 h.

The goal is to recycle the catalyst; therefore successive reaction runs have been
employed. Separation takes place by cooling the mixture to room temperature and use

23
decantation to separate the insoluble solvent and catalyst/ligand complex. Loss of
phosphite in the organic phase was determined by ICP technique.
In the experiments the separated catalyst was reused up to six times without significant
decrease in activity. In the first two reaction runs there was a loss of P of 7.8 and 7.2 %.
In the following runs the loss decreased, which is showed in table 4.10.

Table 4.10: recycling of OPGPP(n=58)/Rh catalyst


Reaction runs Conversion Aldehyde yield n/iso ratio P in product phase P loss in product1
(%) (%) (µg/ml) (%)
1 99.4 97.6 0.91 60.4350 7.8
2 99.3 97.8 0.95 55.2800 7.2
3 99.2 95.0 0.95 23.3150 3.05
4 96.3 93.0 0.92 18.6350 2.4
5 97.3 96.3 0.93 17.0000 2.22
6 94.3 90.5 0.97 15.1350 1.98
7 93.0 92.4 0.93 7.0650 0.92
Reaction conditions: same as in table 4.9
1
Loss of ligand/employed ligand (based on P content).

This decrease in loss of P is ascribed to the weight distribution of the polyether group in
the ligand molecule. Polyether derivatives are mixtures of different weight compounds,
according to table 4.7 phosphates of lower weight have higher solubility.

Not only 1-decene can be converted in an aldehyde using OPGPP ligand, but also other
olefins. This results in increasing aldehyde yield and n/iso ratio for olefins with
decreasing chain length.

4.5.3 OPGPP in combination with ruthenium catalyst


Another possibility is to use another catalyst with the OPGPP ligand. Ruthenium
complexes have received little attention in the hydroformylation because of the lower
activity. Experiments are performed with a combination of polyether phosphite ligand,
ruthenium catalyst and heptane as solvent. The reaction occurs with the following
catalyst complex: OPGPP/Ru3(CO)12. Under a higher temperature then Rh-catalyst, 130
o
C, high conversion is reached and a reasonable activity is retained after some successive
runs, see table 4.11. The results also show considerable hydrogenation of aldehydes into
alcohols, which is not an advantage, because a second separation step is needed, although
alcohols are made out of aldehydes, as said in chapter 2.2.

Table 4.11: OPGPP(n=58)/Ru3(CO)12 catalyzed hydroformylation and recovery


Reaction runs Conversion (%) Aldehyde yield (%) Alcohol (%)
1 96.3 70.0 25.0
2 95.4 68.3 26.0
3 89.0 63.7 23.1
4 87.1 64.5 22.0
Reaction conditions: 1-decene 1.0 ml., heptane 2.0 ml., internal standard 0.3 m, Ru3(CO)12 0.0072 gr, OPGPP 0.064 gr, P(CO)/P(H2) =
1:1, T =130 oC, syngas pressure 5.0 MPa, reaction time 9 h.

24
4.6 Remaining techniques
In this paragraph some of the remaining interesting techniques are discussed shortly. The
reason to mention these techniques shortly is because little information is found in
literature or the articles are a bit old.
Techniques described are: hydroformylation with ruthenium catalyst, high pressure
effects of homogeneously catalyzed hydroformylation and sulfonated xantphos ligand.

4.6.1 Ruthenium catalyst


As already described most catalyst are based on cobalt and rhodium, both have their
disadvantages. For cobalt for instance it is hard to control selectivity towards the linear
products, and rhodium is very expensive and needs a large excess of phospine to reach
high linearity. Trying to find a solution, research is done to the less active ruthenium
catalyst, compared with cobalt and rhodium (ref. 16). Ruthenium is used in combination
with 1.10- phenenthroline as a ligand and amide as a solvent. The active catalyst complex
used in the reaction is Ru3(CO)12/1.10-phenanthroline. Different reaction conditions are
used in the experiments, the temperature is about 120 – 130 oC and the pressure is 80
atm. of syngas, which is high compared to normal hydroformylation. Mainly the
hydroformylation of propylene and 1-octene are described. The experiments took a long
time (20 hours) and the yield of C4-aldehydes varied from 65 – 93 % and of C9-aldehydes
from 49 – 55 %, thus no very interesting results so far. However, the linearity of the
aldehydes were very high, 95% and higher, which is much better than in existing
hydroformylation. Unfortunately the data from the report is almost ten years old and no
further useful articles have been found.

4.6.2 High-pressure hydroformylation


The current hydroformylation processes operate with pressures below 30 MPa. The
described hydroformylation processes in this case (ref. 17) operate with pressures and
temperatures up to 700 MPa and 200 oC. The reason to perform such experiments is
based on the Le-Chatelier principle that reactions with negative entropy should preferably
be carried out at high pressures if gasses are involved.
The process is rhodium based, the active complex used is Rh(cod)(PPh3)2BF4, which is a
precatalyst. The most important result is that yields of aldehydes increase at higher
pressures, but the linear to branched ratio decreases at higher pressures.
These types of experiments are very rare for homogeneously catalysis. For commercial
application the process economy of this process should by high enough, because high
pressure technology is cost intensive. Therefore only high value products seem to be
suited for this type of process.

4.6.3 Sulfonated Xantphos ligand


Sulfonated xantphos is a new kind of ligand used in hydroformylation. The ligand is used
in a biphasic system, according to reference 18 this type of hydroformylation is
cyclodextrin based. Cyclodextrins are used as mass-transfer promoters in aqueous phase
rhodium based catalysis. The trick is to find cyclodextrins that interact with the
sulfonated xanthphos/rhodium complex, so mass-transfer can be promoted. Promotion of
mass transfer results in higher reaction rates, because the catalyst complex can move

25
faster through the aqueous phase and have therefore more interaction with the organic
reactant phase. Experiments are performed at normal hydroformylation conditions, only
the pressure is higher, 50 bar. The results using interacting cyclodextrins are very
promising for 1-octene hydroformylation; the conversion is 90 % with aldehyde
selectivity to 99 %. Also the linear to branched ratio is better than in normal
hydroformylation, l/b is about 20. Reaction rate however are not yet very satisfying,
because experiment took 24 hours.
The conclusion is that sulfonated xantphos ligand appears to be a valuable ligand in
combination with a biphasic system with addition of the right cyclodextrins. Separation
of the catalyst is also possible because it is a biphasic system. Methods to separate are
described earlier in this thesis, but not in the found literature about xantphos ligands.

26
5. Discussion
In this final chapter the described new techniques in hydroformylation will be discussed.
In the literature and articles that are used for this report, a lot of different research and
developments are described. All these different techniques are presented and tested on
laboratory scale. Little is known whether these methods can be used on commercial scale.
Therefore advantages and disadvantages of the described systems in chapter 4 will be
compared with the current processes described in chapter 3.

5.1 Hydroformylation in scCO2


The biggest benefit of supercritical CO2 as solvent in hydroformylation processes is that
no organic solvents are used. From an environmental point of view this is a major
improvement compared with the current industrial processes. CO2 is also a very safe non-
reactive solvent
The conversions for cobalt and rhodium catalysts are up to 100% which is better than or
equal to the current processes (compare with table 3.1 and 3.2). Only the linearity (80%,
table 4.1) of the products is a bit lower than in the Shell process for instance (85 -90 %),
but this can be solved by fine-tuning the system by further research.
A main difference is that scCO2 hydroformylation is a high pressure process with
pressures up to 240 bar, this is a lot higher than the current processes. High pressure
equipment is therefore needed, which have high costs. This does not necessarily have to
be a major problem. When using a polymer based catalyst, the product and catalyst
mixture can be easily separated with a membrane. This means no large separation train of
distillation and flash columns are needed. The equipment of scCO2 hydroformylation is
therefore more expensive, but there is less equipment. Hydroformylation of long chain
olefins is probably easier to achieve in scCO2, because no high temperature distillation is
needed, which cause degradation of the catalyst.
There are major benefits using scCO2, but still a lot of research has to be performed
before industrial application is possible. All lab scale experiments are performed as batch
processes, while continues processes are preferred.

5.2 Micellar systems


The replacement of organic solvent by water is beneficial for environmental, safety and
economical reasons. There are some examples of biphasic systems with acceptable
results, the RCH/RP process described in 3.3.2. This is already a better process to recover
the catalyst, because there is an organic product phase and an aqueous catalyst phase.
Major disadvantage of this system is that long chain olefins have low conversion, because
of their low solubility in the aqueous phase.
A micellar system provides a solution for the hydroformylation of long chain olefins. By
adding a surfactant, micelles are formed creating a large interfacial area.
The best surfactant to add is a cationic surfactant, because the negative charged ends of
micelles are orientated to the aqueous phase, which attract the positive charged catalyst.
This means that the catalyst is highly concentrated at the interface and more easily
coordinated with the olefins in the micelle.
Conversion of long chain olefins is much higher in the micellar system, with almost the
same process conditions as the RCH/RP process and no distillation units are needed as
with long chain olefins in homogeneously hydroformylation.

27
Small disadvantages of the system are the micellar phase and the separation of the
catalyst. Behaviour of the micellar phase (stability, size and micelle shape) is rather
complex and depends on many parameters. A lot of research is needed to obtain optimal
selectivity in combination with reaction rates.
Separating the product and catalyst can be achieved by cooling the mixture and use
decantation to separate organic and aqueous phase. The presence of surfactants
complicates the phase separation, because the separation with surfactant is not as neat as
with ordinary biphasic systems. This causes a slight loss of catalyst or product.

5.3 Silica supported catalyst


Another convenient solution to separate the catalyst is by supporting the catalyst by
fumed silica nanoparticles, also in a biphasic system. The catalyst is in the aqueous phase
as normal. This modified biphasic system operates at low pressures of about 4 MPa and
temperatures of about 100 oC, almost similar as the existing RCH/RP process. Also the
conversion (up to 95 %) is comparable, but the linearity is lower.
This system is a small modification of the RCH/RP process and could therefore be used
in practice. The process only needs to be tested on continues basis. In the researches the
silica catalyst complex is separated by simple filtration, for practical use a convenient
solution has to be found to recover the particles.

5.4 Polyether phospite as ligand


Polyether is a new ligand used in homogeneous hydroformylation. Polyether phosphite
shows more activity in long chain hydroformylation than the common ligands.
Hydroformylation with a rhodium catalyst, OPGPP ligand and heptane as solvent shows
good result in reaction rate and conversion (up to 99%) only with a low linear/branched
ratio. In the experiments the temperature and pressure of the process is about 90 oC and 5
MPa, which is quite common for the industrial processes.
The benefit of phosphines is that they are slightly soluble in apolar solvents at room
temperature. Cooling the catalyst/product mixture is a good approach of separation by
decantation. Loss of the catalyst however can be quite large as described in 4.5.2. Further
optimization of the linear/branched ratio and the recovery (without a distillation unit) is
needed before it can be used in practice.

5.5 Remaining techniques


In this last paragraph the possible practical application of the three shortly described
techniques are discussed.

Using another catalyst like ruthenium is not very probable, because the whole industry is
based on cobalt and rhodium. Ruthenium has also lower activity although the better
linearity. Other provided techniques are more probable than this one.

The hydroformylation with high pressures up to 700 MPa is not yet a good alternative,
because hydroformylation does not occurs significant better. Costs most likely will be
higher than the benefits. Only when there is a market for high value products made by
high pressure hydroformylation has a future.

28
In contrary of the other two, the use of sulfonated xantphos ligand in combination with
cyclodextrins in a biphasic system has a future. This is due to the high conversions and
linearity. This research is yet only at the beginning and it will take same time to develop.

29
Reference list
1. Cornelis, B., Herrmann, W.A., Applied Homogeneous Catalysis with Organometallic
Compounds, part I: Applied homogeneous catalysis. Wiley-VCH, Weinheim 2002.
2. Bahrmann, H., Bach, H., Oxo Synthesis. Ullmann’s Encyclopedia of Industrial
Chemistry. Wiley-VCH Verlag GmbH & Co, 2002.
3. Akgerman, A., Guo, Y., Hydroformylation of Propylene in Super Critical Carbon
Dioxide. Ind. Eng. Chem. Res., 36, 4581 – 4585 (1997).
4. Erkey, C., Palo, D., Kinetics of the Homogeneous Catalytic Hydroformylation of 1-
octene in Supercritical Carbon Dioxide with HRh(CO)[P(p-CF3C6H4)3]3. Ind. Eng. Chem.
Res., 38, 3786 – 3792 (1999).
5. Kani, I., Flores, R., at all, Hydroformylation of Styrene in Supercritical Carbon Dioxide
with Fluoroacrylate Polymer Supported Rhodium Catalysts. J. of Supercritical Fluids, 31,
287 – 294 (2004).
6. Lopez-Castillo, Z., Flores R., at all, Evaluation of Polymer-Supported Rhodium Catalysts
in 1-octene Hydroformylation in Supercritical Carbon Dioxide. Ind. Eng. Chem. Res., 42,
3893 – 3899 (2003).
7. Chen, H., Li, Y., at all. Micellar Effect in High Olefin Hydroformylation Catalyzed by
Water-Soluble Rhodium Complex. J. of Molecular Catalysis, 149, 1 – 6 (1999).
8. Vyve, F. van, Renken, A., Hydroformylation in Reverse Micellar Systems. Catalysis
Today, 48, 237 – 243 (1999).
9. Miyagawa, C., Kupka, J., Schumpe, A., Rhodium-Catalyzed Hydroformylation of 1-
octene in Micro-Emulsions and Micellar Media. J. of Mol. Cat. A: Chemical, 234, 9 – 17
(2005).
10. Riisager, A., Hanson, B., CTAB Micelles and the Hydroformylation of Octene with
Rhodium/TPPTS Catalysts. Evidence for the Interaction of TPPTS with Micelle Surfaces.
J. of Mol. Cat. A: Chemical, 189, 195 – 202 (2002).
11. Li, Z., Peng, Q., Yuan, Y., Aqueous Phoshine-Rh Complexes Supported on Non-Porous
Fumed-Silica Nanoparticles for Higher Olefin Hydroformylation. Applied Catalysis A:
General, 239, 79 – 86 (2003).
12. Sandee, A., Reek, J., at all, A Silica-Supported, Switchable, and Recyclable
Hydroformylation–Hydrogenation Catalyst. J. Am. Chem. Soc., 123, 8468 – 8476 (2001).
13. Leeuwen, P. van, Sandee, J., at all, Xantphos-Based, Silica-Supported, Selective and
Recyclable Hydroformylation Catalysts: a Review. J. of Mol. Cat. A: Chemical, 182 –
183, 107 – 123 (2002).
14. Liu, X., Li, H., at all, Polyether Phosphite for Hydroformylation of Higher Olefins in
Non-Aqueous System and Catalyst Recovery. J. of Organom. Chem., 654, 83 – 90
(2002).
15. Liu, X., Chen, R., Jin, Z., J. Organom. Chem. 571, 201 (1998).
16. Mitsudo, T., Suzuki, N., at all, Ru3(CO)12/1.10-Phenanthroline-Catalyzed
Hydroformylation of Olefins. J. of Mol. Cat. A: Chemical, 109, 219 – 225 (1996).
17. Albers, J., Dinjus, E., at all, High-Pressure Effects in the Homogeneously Catalyzed
Hydroformylation of Olefins. J. of Mol. Cat. A: Chemical, 219, 41 – 46 (2004).
18. Leclercq, L., Reek, J., at all, Sulfonated Xantphos Ligand and Methylated Cyclodextrin: a
Winning Combination of Rhodium Catalyzed Hydroformylation of Higher Olefins in
Aqueous Medium. Organometallics, 24, 2070 – 2075 (2005).

30
Appendix I:Process flow diagrams of industrial processes

Cobalt based processes

BASF process

Exxon process

31
Shell process

Rhodium based processes

RCH/RP process

32
UCC process, liquid recycle

33
Appendix II: Rhodium catalyst cycle

Dissociative mechanism of the hydroformylation cycle for modified rhodium catalysts.

34

Das könnte Ihnen auch gefallen