Está en la página 1de 7

Ind. Eng. Chem. Res.

2005, 44, 9853-9859 9853

Stability and Kinetic Studies of Supported Ionic Liquid Phase


Catalysts for Hydroformylation of Propene
Anders Riisager,† Rasmus Fehrmann,† Marco Haumann,‡ Babu S. K. Gorle,‡ and
Peter Wasserscheid*,‡
Department of Chemistry and Center for Sustainable and Green Chemistry, Technical University of Denmark,
Building 207, DK-2800 Kgs. Lyngby, Denmark, and Lehrstuhl für Chemische Reaktionstechnik,
Universität Erlangen-Nürnberg, Egerlandstrasse 3, D-91058 Erlangen, Germany

Supported ionic liquid phase (SILP) catalysts have been studied with regard to their long-term
stability in the continuous gas-phase hydroformylation of propene. Kinetic data have been
acquired by variation of temperature, pressure, syngas composition, substrate concentration,
and residence time. The activation energy was determined to be 63.3 ( 2.1 kJ mol-1, which is
in good agreement with known results from biphasic hydroformylation. The results from the
kinetic studies confirmed previously published results on the homogeneous nature of the
heterogenised Rh-SILP catalyst. Long-term stability exceeded 200 h time on stream with no
loss in selectivity. A small decrease in activity could be compensated by a vacuum procedure
regaining the initial activity.

Introduction pressure which makes them attractive as alternative


solvents for homogeneous catalysis. Their polar nature
The hydroformylation of alkenes, discovered in 1938 allows the stabilization of ionic transition metal com-
by Otto Roelen at Ruhrchemie, Oberhausen, Germany,1 plexes as well as complexes being susceptible to hy-
is one of the largest applications of homogeneous transi- drolysis. The use of ionic liquids in catalysis has been
tion metal catalysis in industry. Worldwide production reviewed recently.8
capacities exceeded 8 mio t/year in 1995.2 Industrial
However, besides their advantageous properties as
hydroformylation catalysts are based on cobalt and
solvents, ionic liquids are relatively expensive, even
rhodium complexes exclusively, with the rhodium pro-
though being commercially available now.9 Further-
cesses operating at much milder conditions (typically
more, ionic liquids are normally highly viscous solvents,
80 °C, 15 bar) than the cobalt processes (typically 200
°C, 300 bar). The aldehydes formed by the addition of thus having low diffusion coefficients for hydrogen and
hydrogen and carbon monoxide to the olefinic double carbon monoxide.10 In case of fast chemical reactions
bond are important precursors for, e.g., plasticizers and this may result in mass transfer limitations, where only
detergent alcohols. Besides the desired linear n-alde- the catalyst in the diffusion layer of the biphasic system
hydes, branched iso-aldehydes are obtained in a parallel is utilized. In an ideal system, a thin film of catalyst
reaction as depicted in Scheme 1. The selectivity of the containing ionic liquid the size of the diffusion layer
catalyst can be influenced by the use of sterically would make use of all of the catalyst and ionic liquid
demanding ligands such as sulfoxantphos 1 (see Scheme (see Figure 1).
1) and is referred to as n:iso ratio or linearity (in Such a film can be immobilized by physisorption,
percentage). tethering or covalent anchoring of ionic liquid fragments
In 1984, a biphasic propene hydroformylation process, on a highly porous material, thus creating a large
introduced by Ruhrchemie/Rhône-Poulenc (RCH/RP), reaction surface. Furthermore, a solid catalyst that can
made use of the immiscibility between an aqueous phase be applied in conventional fixed-bed reactors is preferred
containing the water-soluble Rh complex and the or- from an industrial point of view. Such supported ionic
ganic product phase.3 Simple product separation and liquid-phase (SILP) catalyst systems have recently been
quantitative catalyst recovery were achieved in a de- reported by us and other researchers for hydroformyl-
cantation unit, thus overcoming the drawbacks of ations,11 hydrogenations,12 hydroaminations,13 and Heck
homogeneous catalysis. Ever since, a variety of im- reactions.14 The SILP catalytic concept offers a very
mobilization concepts for hydroformylation catalysts efficient ionic liquid utilization and provides relatively
have been studied, ranging from biphasic fluorous short diffusion distances of reactants compared to
phase4 over microemulsions5 to supported catalysis.6 conventional two-phase organic-ionic liquid catalyst
Among the biphasic catalysis concepts the use of ionic systems. In addition, the negligible vapor pressure, large
liquids has gained significant interest over the past liquid range, and thermal stability of ionic liquids
decade both in academia and industry.7 Ionic liquids ensure that the solvent is retained on the support in
consist entirely of ions and have no measurable vapor its fluid state at common reaction conditions, e.g.,
elevated temperatures, which makes SILP catalysts
highly suitable for continuous processing. Moreover, the
* To whom correspondence should be addressed. Phone:
(+49) 9131-8527420. Fax: (+49) 9131-8527421. E-mail: ionic liquid solvent properties are tuneable by the choice
peter.wasserscheid@crt.cbi.uni-erlangen.de. of the cation/anion combination, which further leads to
† Technical University of Denmark. advantages of the SILP concept such as, e.g., compat-
‡ Universität Erlangen-Nürnberg. ibility with hydrolytically labile catalyst complexes,
10.1021/ie050629g CCC: $30.25 © 2005 American Chemical Society
Published on Web 11/11/2005
9854 Ind. Eng. Chem. Res., Vol. 44, No. 26, 2005

Scheme 1. Propene Hydroformylation Using Rh-Sulfoxantphos Catalyst Complex

specific solubility properties, and possible catalyst ac- air at 450 °C for 24 h and stored under argon prior to
tivation by liquid-catalyst interaction. use. The ligand sulfoxantphos 1 has been synthesized
We have recently demonstrated15 the applicability of by sulfonation of xanthene (Aldrich) according to lit-
selective (n:iso ratios up to 23.7, i.e., 96.0% linearity) erature procedures.16 Rh(CO)2(acac) (0.0512 g, 0.2 mmol)
unmodified silica gel SILP rhodium catalysts (Rh-SILP) (Aldrich), was dissolved in 20 mL of dried MeOH and
containing the bisphosphine sulfoxantphos 1 (Figure 1) stirred for 10 min. Ligand 1 (1.57 g, 2 mmol) were
with physisorbed hexafluorophosphate and n-octyl sul- added, and the orange solution was stirred for another
fate ionic liquids [BMIM][X] (X ) PF6 or n-C8H17O- 10 min. Afterward, 1 mL (1.06 g) of ionic liquid [BMIM]-
SO3) for the gas-phase propene hydroformylation in a [n-C8H17O-SO3] (pH ) 8.3, H2O < 500 ppm, Cl- < 100
continuous-flow process. The ionic liquid film proved to ppm, Solvent-Innovation) was added to the solution.
be a prerequisite for an active, selective, and stable After being stirred for 30 min, 10 g of calcinated silica
catalyst. Experiments without the ionic liquid revealed were added and the solution was stirred for 60 min. The
a deactivation of the catalyst within 10 h time on MeOH was removed in vacuo, and a pale red powder
stream, indicated by a loss of both activity and selectiv- was obtained. The supported ionic liquid-phase catalyst
ity. The homogeneous nature of the Rh-SILP catalyst was stored under argon until further use.
was verified by in situ FTIR experiments. A proper Kinetic Experiments. The long-term stability and
thermal pretreatment, reducing the content of surface kinetic studies were carried out in a continuous fixed
silanol groups of the support material, has been identi- bed reactor setup depicted in Chart 1. The SILP catalyst
fied to be crucial for long-term stability of the catalyst. was filled into the reactor (4), and the complete rig was
NMR measurements on the Rh-SILP catalyst indicated evacuated at room temperature. The rig was pressurized
that more than 50% of the excess ligand was bound to with 50 bar helium and left under pressure for 30 min
the remaining acidic sites of the support, thus making while monitoring the pressure. If no pressure drop was
a higher L:Rh ratio than in biphasic catalysis a prereq- observed, the reactor was heated to reaction tempera-
uisite for obtaining a stable and selective catalyst ture under helium pressure. The complete setup was
system. evacuated and flushed with helium three times before
In this work we present further evidence for the syngas and propene were fed into the system.
homogeneous nature of the Rh-SILP catalyst based on Propene (2.8, Linde) was taken out of a reservoir in
kinetic results. The results strongly support the previ- the liquid state and fed into a heated evaporator (2) via
ous spectroscopic results. Furthermore, the catalyst a HPLC pump (1) (Knauer) to control the molar flow of
performance has been studied for more than 200 h time the substrate. Both carbon monoxide (3.5, Linde) and
on stream in order to verify the long-term stability of hydrogen (5.0, Linde) flows were adjusted by means of
the Rh-SILP catalyst. mass-flow controllers (MFC, 5850 S series, Brooks
Instruments). The preheated gases were combined with
Experimental Section the propene in the mixing unit (3), which was filled with
glass beads in order to ensure proper mixing and
Catalyst Preparation. All syntheses were carried isothermal conditions. The gas mixture could then either
out using standard Schlenk techniques under prepuri- enter the reactor (4) or exit the system via a bypass.
fied argon. The silica gel 100 (particle size 0.06-0.2 mm, The reactor (4) consisted of a stainless steel tube (10
pore volume 0.966 mL g-1, Merck), was calcinated in mm diameter, 220 mm length) equipped with a bronze
sinter plate for catalyst placement. After the reactor,
the gas mixture passed a 7-µm filter in order to avoid
Chart 1. Flow Scheme of the Continuous
Hydroformylation Reactor Setup

Figure 1. Schematic principle of SILP catalysis.


Ind. Eng. Chem. Res., Vol. 44, No. 26, 2005 9855

decontamination of the tubing with catalyst or solid Table 1. Rh-SILP Hydroformylation of Propene at
particles. A back-pressure regulator valve (5) (Samson) Various Substrate Concentrationsa
was used to maintain the desired reaction pressure and temperature ppropene conversion TOF n-butanal
outlet gas flow. After the regulator valve, the gas stream entry (°C) (bar) (%) (h-1) (%)
was split and a minor flow was passed through a 134- 1 65 0.9 0.86 4 95.9
µL sampling loop mounted on a gas chromatograph (HP 2 65 1.8 0.91 8 95.9
5890 II plus). Samples were taken at regular intervals 3 65 3.2 0.93 17 95.8
by injecting the volume of the sampling loop via a 6-port 4 82 0.9 2.13 9 95.7
valve into the gas chromatograph. 5 82 1.8 2.09 19 95.6
6 82 3.2 2.54 45 95.4
Gas Chromatography (GC). The conversion of 7 100 0.9 6.54 29 95.2
propene as a function of process conditions was mea- 8 100 1.8 7.48 67 95.2
sured using on-line GC technique. A HP 5890 GC 9 100 3.2 9.82 175 95.1
equipped with a Pona column (50 m, 0.2 mm diameter, 10 120 0.9 20.5 91 94.1
0.25 µm coating) and a flame ionization detector (FID) 11 120 1.8 21.1 188 93.8
12 120 3.2 21.4 381 93.5
precalibrated for propene, n-butanal, and iso-butanal
13 130 0.9 27.6 123 93.4
(allowing the peak areas to be transferred into propene 14 130 1.8 32.8 292 92.9
conversion) was applied: Injector temperature 150 °C, 15 130 3.2 28.1 501 90.1
split ratio 43:1, helium carrier gas flow 2.4 mL min-1, 16 140 0.9 21.0 94 92.5
detector temperature 250 °C. To detect possible high- 17 140 1.8 26.9 240 92.3
boiling byproducts (heavies), the following temperature 18 140 3.2 28.1 500 87.7
program was used: initial temperature 50 °C, initial a Standard conditions: 10 bar syngas pressure, H :CO ) 1:1,
2
time 5 min, heating ramp of 50 °C min-1, final temper- nRhodium ) 8.17 × 10-5 mol, τ ) 0.9 s.
ature 150 °C, final time 3 min, cooling ramp 50 °C
min-1, final temperature 50 °C.
GC mass spectroscopy data have been recorded on a
Varian Saturn 2100T equipped with a CP 8410 au-
tosampler, ion trap detector, and Varian factor four
capillary column (15 m, 0.25 mm diameter, 0.25 µm
coating): Injector temperature 220 °C, split ratio 50:1,
helium carrier gas flow 1 mL min-1. The following
temperature program was used: initial temperature 50
°C, initial time 5 min, heating ramp of 10 °C min-1, final
temperature 220 °C, final time 10 min.
Inductively Coupled Plasma (ICP) Analysis.
During the long-term stability experiments the exit gas
streams from the GC and reactor were combined and
collected in a cold trap filled with liquid nitrogen. The
liquid samples were subsequently analyzed by ICP.
Calculations. The residence time τ was calculated
from the total amount of Rh in the catalyst divided by
the total molar flow Ftotal of substrates
Figure 2. Differential analysis of the hydroformylation at various
nRhodium propene partial pressures and temperatures. 10 bar total pressure,
τ) [s] H2:CO ) 1:1, nrhodium ) 8.17 × 10-5 mol, τ ) 0.9 s.
Ftotal
stream, no change in activity and selectivity was
The turn-over frequency (TOF) was calculated from the observed. On the basis of this preliminary result, the
molar flow of aldehydes divided by the amount of kinetic data were acquired within the first 36 h of a
rhodium freshly installed catalyst sample.
Variation of Propene Pressure. The hydroformyl-
Faldehyde FpropeneX -1 ation has been performed at propene partial pressures
TOF ) ) [h ]
nRhodium nRhodium of 0.9, 1.8, and 3.2 bar partial pressure and in the
temperature range 65-140 °C with a residence time of
The rate of reaction was calculated from the converted 0.9 s. Table 1 compiles the results of these variations.
propene and the residence time At 65 and 82 °C, the reaction rates observed were low
with TOFs in the range of 4-45 h-1 (entries 1 to 6).
∆ppropene ppropene,0X bar The highest TOFs were obtained at high propene partial
r)
τ
)
τ [ ]
s
pressure and high temperature (entries 15 and 18).
However, the activities at 140 °C were not as high as
Results and Discussion could be expected assuming a normal Arrhenius tem-
perature dependence. This might be attributed to the
Stability. In a first set of experiments, the Rh-SILP formation of high-boiling byproducts, namely, 2-ethyl-
catalyst system has been applied for the hydroformyl- hexanal and 2-ethyl-hexanol (“heavies”), which can
ation of propene at 100 °C and 10 bar syngas pressure dissolve in the ionic liquid layer, thus lowering the
(H2:CO 1:1) over a period of 36 h time on stream. TOFs effective catalyst concentration (for more details, see
between 71 and 74 h-1 and selectivities for n-butanal later Deactivation-Reactivation Studies). A differential
of 95% (n:iso ) 19) were obtained in good agreement analysis of all data was made by plotting ln(rate)
with previously published data.15 After 36 h time on against ln(ppropene). The reaction order n of propene was
9856 Ind. Eng. Chem. Res., Vol. 44, No. 26, 2005

Figure 3. Integral analysis of propene hydroformylation at


different residence times and temperatures. 10 bar syngas pres- Figure 5. Activity of the Rh-SILP catalyst at different syngas
sure, H2:CO ) 1:1, nrhodium ) 8.17 × 10-5 mol, ppropene ) 2.1 bar. compositions (H2:CO ratios) and total pressures of 10, 15, 20, and
30 bar. 100 °C, nrhodium ) 7.98 × 10-5 mol, τ ) 0.8 s.

Figure 4. Activity of the Rh-SILP catalyst as a function of partial


pressure of CO. 100 °C, nrhodium ) 7.98 × 10-5 mol, τ ) 0.8 s, Figure 6. Arrhenius plots based on propene pressure and
phydrogen ) 1.3 bar, ppropene ) 1.8 bar (balanced with He). residence time variations.

By assumption of a first-order reaction with respect to


calculated from the slope of the curves and the rate propene, integration of the rate expression
constant k from the intercept. Figure 2 shows the results
of the differential analysis. dppropene
The hydroformylation of propene with the Rh-SILP r)- ) kpn

catalyst under investigation was found to be first order
in propene and the activation energy was calculated resulted in
from an Arrhenius plot (see Figure 6) to be 61.9 kJ mol-1
with a collision factor k0 of 3.7 × 107 s-1. ln(ppropene) - ln(ppropene,0) ) -kτ
Variation of Residence Time. At a constant pro-
pene partial pressure of 2.1 bar, the residence time Plotting ln(ppropene) - ln(ppropene,0) against the residence
inside the catalyst bed was varied by altering the total time τ resulted in a linear dependence as shown in
reactant flow. From the results in Table 2, it can be seen Figure 3.
that shorter residence times generally resulted in lower The rate constants were determined from the slopes
conversions, whereas at longer residence times the of the graphs, and the activation energy was found to
conversions were higher. Under nondifferential condi- be 64.8 kJ mol-1 in good agreement with the one
tions at higher conversions (entries 29-38), the ob- obtained from variation of substrate partial pressure
served TOFs decreased slightly with longer residence (see Figure 6). The collision factor k0 of 7.1 × 107 s-1
was higher than the one determined by substrate
time representing the lower mean level of propene
pressure variation.
present in the reactor.
Variation of Total Pressure and Syngas Com-
The selectivity for the desired linear aldehyde de- position. At 100 °C, the partial pressure of CO was
creased only slightly from 96.2-93.6% by increasing the varied between 0.7 and 5.2 bar, while the hydrogen and
residence time and temperature. At temperatures of 120 propene partial pressures were kept constant at 1.3 and
°C, TOFs between 320 and 470 h-1 could be achieved. 1.8 bar, respectively. Helium was used as inert gas to
Ind. Eng. Chem. Res., Vol. 44, No. 26, 2005 9857

Table 2. Rh-SILP Hydroformylation of Propene at rhodium-catalyzed hydroformylations using sulfonated


Various Residence Timesa ligands in aqueous media and supported aqueous media.
temperature residence conversion TOF n-butanal The activation energy of propene using Rh-SILP
entry (°C) time (s) (%) (h-1) (%) catalysts is slightly lower than the one determined by
19 65 2.0 0.72 7 95.3 Mao et al. for HRh(CO)(TPPTS)3 in water.20 Compared
20 65 1.5 0.96 12 95.0 with HRh(CO)(TPPTS)3 in supported aqueous phase
21 65 1.2 1.02 16 95.4 (SAP) catalysis of higher alkenes,22,23 the activation
22 65 1.0 0.82 16 96.2
energy for propene hydroformylation is in good agree-
23 65 0.8 0.68 17 96.2
24 84 2.0 4.66 44 95.4 ment. This comparison provides additional evidence that
25 84 1.5 3.67 46 95.5 the Rh-SILP catalyst truly acts as a homogeneous
26 84 1.2 3.11 49 95.4 catalyst in the ionic film immobilized on the silica.
27 84 1.0 2.13 48 95.4 Deactivation-Reactivation Studies. The stability
28 84 0.8 1.75 44 95.7
29 100 2.0 10.3 97 95.2
of the Rh-SILP catalyst was initially studied at 100
30 100 1.5 8.57 108 95.2 °C and 10 bar syngas pressure (H2:CO ) 1:1) over a
31 100 1.2 7.31 115 95.1 period of 36 h with no loss in activity or selectivity. This
32 100 1.0 5.46 103 95.1 time on stream was extended to 180 h to test the long-
33 100 0.8 5.57 140 95.1 term stability of the system. Figure 7 shows the results.
34 120 2.0 34.3 324 94.5
35 120 1.5 29.1 367 94.4 The selectivity toward n-butanal remained constant
36 120 1.2 25.5 401 93.9 around 95% (n:iso ) 19) during the reaction whereas,
37 120 1.0 23.1 437 93.8 the TOFs slightly decreased from 74 to 61 h-1, corre-
38 120 0.8 18.7 471 93.6 sponding to a total loss in activity of 17% or 0.1% per
a Standard conditions: 10 bar syngas pressure, H :CO ) 1:1,
2
hour. Since the selectivity remained unchanged, it was
nRhodium ) 8.17 × 10-5 mol, ppropene ) 2.1 bar. assumed that the catalyst did not decompose, as ob-
served in previous studies when using low L:Rh ra-
maintain constant volume flow and total pressure of 10 tios.11,15 All ICP analyses of the exit gas streams
bar. The activity decreased with increasing pCO as condensed in liquid nitrogen showed rhodium contents
depicted in Figure 4, while the selectivity for n-butanal below the detection limit of 3 ppm. Instead, the forma-
remained unchanged at 95%. tion of high boiling side products dissolving in the ionic
The total syngas pressure was increased to 15, 20, liquid layer is expected to be the cause of the slow
and 30 bar with pCO varied in the range between 1.0 deactivation. The effective rhodium concentration would
and 7.8 bar (ptotal ) 15 bar, pH2 ) 2.0 bar, ppropene ) 1.3 be lowered by the dissolved byproducts. Furthermore,
bar), 1.3 and 10.4 bar (ptotal ) 20 bar, pH2 ) 2.6 bar, the film thickness might be increased and smaller pores
ppropene ) 1.8 bar), and 2.0 to 15.6 bar (ptotal ) 30 bar, can be flooded, which will lead to a lower reaction
pH2 ) 3.9 bar, ppropene ) 2.7 bar), in all cases balanced surface. To confirm this hypothesis, the gas flow was
with Helium gas. As can be seen from Figure 4, the stopped after 180 h time on stream and the setup was
activity increased with increasing total pressure. From evacuated for 10 min at 100 °C using a vacuum pump.
the slopes of the linear fitting a negative order with When the experiment was continued after this proce-
respect to pCO of -0.4 can be calculated according to dure, the activity had indeed increased by 80% from 60
to 108 h-1. Within the next 20 h, the TOFs decreased
r ) kpCOx (r expressed as TOF) again from 108 to 76 h-1 and the selectivity reestab-
lished at 95% n-butanal. A second vacuum period of 10
In Figure 5, the activity, expressed as ln(TOF), has been min resulted in improved TOFs of 116 h-1, as depicted
plotted against ln(ptotal). The slope of the linear fitting in Figure 8. In both cases, the observed “overshooting”
should give the order m with regard to total syngas of the activity by evacuation might be caused by
pressure according to simultaneous removal of CO ligand of the Rh complex
leading to higher activity. A lower selectivity is expected
r ) kptotalm (r expressed as TOF) and is indeed observed for a short while after evacua-
tion. Thereafter, the catalyst solution is again saturated
For pH2:pCO ratios of 0.25, 0.33, 0.67, and 2 an order m with CO gas, and both the activity and the selectivity
) 0.4 has been calculated from the fitted graphs in are approaching the initial levels.
Figure 5. Both the negative order in CO17 and the first These findings confirm the interpretation that the
order in substrate concentration18 are known from observed slight deactivation over time is not due to
literature and can be derived from Wilkinson’s mecha- catalyst decomposition. Our previously published results
nism for modified rhodium-catalyzed hydroformyla- clearly indicated that catalyst deactivation due to ligand
tion.19 degradation is accompanied with loss of selectivity,15
In Figure 6, the ln(k) was plotted against T-1 in an which is not observed in this case. In this context, we
Arrhenius-type diagram based on substrate and resi- expected the formation of 2-ethyl-hexanal and 2-ethyl-
dence time variation. hexanol to be of relevance. At higher temperatures,
The activation energies calculated from the indepen- traces of these high-boiling side products were observed
dent experiments are similar and can be accounted to in the GC chromatograms of the gas outlet streams.
be 63.3 ( 2.1 kJ mol, clearly indicating that the Rh- To further support this interpretation of our results,
SILP catalyst is operating under kinetically controlled the deactivated catalyst was removed from the reactor
reaction conditions. The collision factors k0 determined and the pale-yellow catalyst particles were placed in two
from the Arrhenius plots range between 2.7 × 107 s-1 round-bottom flasks containing (i) cyclohexane and (ii)
(based on residence time) and 9.8 × 107 s-1 (based on ethanol. Immediately, the ethanol solution became
propene pressure). In Table 3, the results from this orange in color, whereas the cyclohexane solution
study are compiled and compared with results for remained colorless. As expected the ethanol wash leads
9858 Ind. Eng. Chem. Res., Vol. 44, No. 26, 2005

Table 3. Comparison of Results for Rh-SILP Hydroformylation with Literature Results


EA
substrate catalyst (kJ mol-1) solvent support ref
Propene HRh(CO)2(1) 63.2 ILa SiO2 this work
Propene HRh(CO)(TPPTS)3 77.0 water 20
1-octene HRh(CO)(TPPTS)3 65.9 water 21
1-octene HRh(CO)(TPPTS)3 71.0 water SiO2 22
Linaloolb HRh(CO)(TPPTS)3 60.7 water SiO2 23
1-dodecene HRh(CO)(TPPTS)3 72.8/70.9c waterd 24
a IL ) [BMIM][ n-C H O-SO ]. b Linalool ) 3,7-dimethyl-1,6-octadien-3-ol. c Two different algorithms were used for data analysis.
8 17 3
d Cetyltrimethylammoniumbromide surfactant used.

(corresponding to the CdO stretching frequency), clearly


indicating the presence of aldehydes in the catalyst.

Conclusion
In this work, we have demonstrated the long-term
stability of a homogeneous rhodium hydroformylation
catalyst, which has been immobilized by the use of the
new SILP technique. No loss in selectivity for linear
butanal was observed during the experiments. The
activity decreased by 17% within 180 h time on stream
due to formation of high boiling side-products. These
heavies could easily be removed from the catalyst by a
vacuum procedure, after which the initial activity could
be regained.
The activation energy of 63.3 ( 2.1 kJ mol-1 added
Figure 7. Long-term hydroformylation stability experiment. 100 evidence that the catalyst is indeed a homogeneous
°C, 10 bar syngas pressure (H2:CO ) 1:1), nrhodium ) 3.53 × 10-5 complex dissolved in an ionic liquid film on a support.
mol, ppropene ) 1.8 bar, τ ) 0.38 s. Furthermore, the Rh-SILP catalyst performed very
similar to a homogeneous catalyst with regard to
variation in syngas composition. This work further
confirms the high technical potential of the SILP
catalysis concept as it allows combining molecular
defined homogeneous catalysis with heterogeneous,
fixed-bed technology.
In perspective, we believe that the knowledge gained
in this and previously published work will accelerate
significantly the successful development of new SILP
catalysts and future SILP catalysis applications (e.g.,
other C-C couplings, hydrogenations, etc.). The SILP
concept is most advantageous for continuous gas-phase
reactions, where the combination of well-defined cata-
lyst complexes, nonvolatile ionic liquids, and solid,
porous supports can enhance the process economics with
respect to product separation and catalyst recovery. If
mechanical and chemical removal of the ionic liquid film
Figure 8. Reactivation of Rh-SILP catalyst by consecutive can be prevented, the SILP technology can also be
application of vacuum. 100 °C, 10 bar syngas pressure (H2:CO ) applied in multiphase slurry reactions. The higher
1:1), nrhodium ) 3.53 × 10-5 mol, ppropene ) 1.8 bar, τ ) 0.38 s.
utilization of the catalyst species and the ease of catalyst
recycling by filtration will offer advantages compared
to the complete removal of the ionic catalyst phase from to classical biphasic systems. We therefore anticipate
the support, while the cyclohexane wash does not. The that the SILP catalysis concept will help bridging the
ethanol solution was distilled in order to separate gap between homogeneous and heterogeneous catalysis
nonvolatile ionic liquid fragments and the distillate was and will lead to improved catalytic processes in the
analyzed by means of GC-MS. The cyclohexane solution future.
was directly analyzed by GC-MS. In both cases, small
amounts of byproducts 1-butanol, 2-ethyl-hexanal and
2-ethyl-hexanol were detected (the GC signals were Acknowledgment
further confirmed by injection of the respective pure This work was supported by the framework “Smart
compounds as references). These results indicate that ligands - smart solvents” of ConNeCat financed by the
relatively unpolar heavies were washed or extracted German Federal Ministry for Education and Research
from the Rh-SILP catalyst by both ethanol and cyclo- (BMBF). Financial support (M. Haumann) by the Deut-
hexane. sche Forschungsgemeinschaft and by the Danish Re-
Additionally, infrared analysis of both used and search Council for Technology and Production is grate-
unused Rh-SILP catalysts was carried out. The used fully acknowledged. The authors thank Dipl.-Ing. Helmut
catalyst showed an absorbance band at 1730 cm-1 Gerhard for support regarding the GC analysis, Ber-
Ind. Eng. Chem. Res., Vol. 44, No. 26, 2005 9859

thold Melcher for doing the syngas variation experi- Consorti, C. S.; Spencer, J. J. Braz. Chem. Soc. 2000, 11, 337; (d)
ments, and Dr. Christine Ernst for support with the Sheldon, R. Chem. Commun. 2001, 2399; (e) Olivier-Bourbigou,
catalyst synthesis. Dipl. Chem. Guido Henze (Univer- H.; Magna, L. J. Mol. Catal., A: Chem. 2002, 182-183, 419; (f)
Stenzel, O.; Raubenheimer, H. G.; Esterhysen, C. J. Chem. Soc.,
sität Dortmund, Lehrstuhl für Technische Chemie A) Dalton Trans. 2002, 1132; (g) Zhao, H.; Malhotra, S. V. Aldrichim.
is acknowledged for the ICP analyses. Acta 2002, 35, 75.
(8) (a) Wasserscheid, P.; Keim, W. Angew. Chem., Int. Ed. 2000,
Nomenclature 39, 3772; (b) Gordon, C. M. Appl. Catal., A 2001, 222, 101; (c) Zhao,
D.; Wu, M.; Kou, Y.; Min, E. Catal. Today 2002, 74, 157; (d)
acac ) Acetylacetonate Dupont, J.; de Souza, R. F.; Suarez, P. A. Z. Chem. Rev. 2002,
BMIM ) n-Butyl-methyl-imidazole cation 102, 3667.
CO ) carbon monoxide (9) Selection of ionic liquid suppliers: (a) Acros Organics
EA ) Arrhenius activation energy (kJ mol-1) (www.acros.com); (b) Fluka (www.fluka.com); (c) Merck
F ) Molar flow (mol s-1) (www.merck.com); (d) Sigma-Aldrich (www.sigma-aldrich.com);
(e) Solvent Innovation (www.solvent-innovation.com); (f) Strem
GC ) gas chromatography (www.strem.com); (g) Wako (www.wako-chem.co.jp).
GC MS ) gas chromatography coupled with mass spec- (10) (a) Brennecke, J. F.; Anthony, J. L.; Maginn, E. J. Gas
troscopy solubilities in ionic liquids. In Ionic Liquids in Synthesis; Wass-
H2 ) hydrogen erscheid, P., Welton, T., Eds.; Wiley-VCH: Weinheim, 2003; (b)
ICP ) Inductively Coupled Plasma Ohlin, C. A.; Dyson, P. J.; Laurenczy, G. Chem. Commun. 2005,
IR ) infrared spectroscopy 201.
k ) rate constant (for nth order reaction) ((mol L-1)1-n s-1) (11) (a) Riisager, A.; Eriksen, K. M.; Wasserscheid, P.; Fehr-
k0 ) collision factor (for nth order reaction) ((mol L-1)1-n mann, R. Catal. Lett. 2003, 90, 149; (b) Riisager, A.; Wasserscheid,
s-1) P.; van Hal, R.; Fehrmann, R. J. Catal. 2003, 219, 252; (c) Riisager,
m ) reaction order with respect to total syngas pressure A.; Fehrmann, R.; Wasserscheid, P.; van Hal, R. In Ionic Liquids
IIIB: Fundamentals, Progress, Challenges, and Opportunities -
MFC ) mass-flow controller Transformations and Processes; Rogers, R. D., Seddon, K. R., Eds.;
MeOH ) methanol ACS Symposium Series, 2005, 902, 334; (d) Mehnert, C. P.; Cook,
n ) reaction order with respect to propene R. A.; Dispenziere, N. C.; Afeworki, M. J. Am. Chem. Soc. 2002,
n-butanal ) linear butanal 124, 12932; (e) Mehnert, C. P. Chem.-Eur. J. 2005, 11, 50.
n:iso ) linear-to-branched ratio of products (12) (a) Mehnert, C. P.; Mozeleski, E. J.; Cook, R. A. Chem.
NMR ) nuclear magnetic resonance spectroscopy Commun. 2002, 3010; (b) Wolfson, A.; Vankelecom, I. F. J.; Jacobs,
nrhodium ) molar amount of rhodium in reactor (mol) P. A. Tetrahedron Lett. 2003, 44, 1195.
p ) pressure (bar) (13) Breitenlechner, S.; Fleck, M.; Müller, T. E.; Suppan, A. J.
PF6 ) hexafluorophosphate anion Mol. Catal. A: Chem. 2004, 214, 175.
r ) reaction rate (mol L-1 s-1) (14) Hagiwara, H.; Sugawara, Y.; Isobe, K.; Hoshi, T.; Suzuki,
T. Org. Lett. 2004, 6, 2325.
RCH/RP ) Ruhrchemie/Rhône-Poulenc (now part of Eu-
(15) Riisager, A.; Fehrmann, R.; Flicker, S.; van Hal, R.;
ropean Oxo and Oxeno) Haumann, M.; Wasserscheid, P. Angew. Chem., Int. Ed. 2005, 44,
Rh-SILP ) Rhodium sulfoxantphos immobilized by SILP 185.
SILP ) supported ionic liquid phase (16) Goedheijt, M. S.; Kamer, P. C. J.; van Leeuwen, P. W. N.
τ ) residence time (s) M. J. Mol. Catal. A: Chem. 1998, 134, 243.
T ) temperature (°C) (17) (a) Moser, W. R.; Papite, C. J.; Brannon, D. A.; Duwell, R.
TOF ) turn-over frequency (h-1) A.; Weininger, S. J. J. Mol. Catal. 1987, 41, 271; (b) Evans, D.;
x ) reaction order with respect to partial pressure of CO Osborn, J. A.; Wilkinson, G. J. Chem. Soc. 1968, 3133; (c) Brown,
C. K.; Wilkinson, G. J. Chem. Soc. 1970, 2753; (d) Gregorio, G.;
Montrasi, G.; Trampieri, M.; Cavalieri d’Oro, P.; Pagani, G.;
Literature Cited Andreetta, A. Chim. Ind. 1980, 62, 389; (e) Cavalieri d’Oro, P.;
(1) Chemische Verwertungsgesellschaft Oberhausen m.b.H (O. Raimondi, L.; Pagani, G.; Gregorio, G.; Montrasi, G.; Andreetta,
Roelen), DE 849,548 (1938/1952); US Patent 2,327,066, 1943. A. Chim. Ind. 1980, 62, 572.
(2) Frohning, C. D.; Kohlpaintner, C. W. In Applied Homoge- (18) (a) Strohmeier, W.; Michel, M. Z. Z. Phys. Chem. Neue
neous Catalysis with Organometallic Compounds; Cornils, B., Folge 1981, 124, 23; (b) Divekar, S. S.; Desphande, R. M.;
Herrmann, W. A., Eds.; VCH: Weinheim, 1996. Chaudhari, R. V. Catal. Lett. 1993, 21, 191; (c) Van Rooy, A.; de
(3) (a) Rhône-Poulenc Ind. FR 2,230,654, 1983; FR 2,314,910, Bruijn, J. N. H.; Roobeek, C. F.; Kamer, P. C. J.; van Leeuwen, P.
1975; FR 2,338,253, 1976; FR 2,349,562, 1976; FR 2,366,237, 1976; W. N. M. J. Organomet. Chem. 1996, 507, 69.
FR 2,473,505, 1979; FR 2,478,078, 1980; FR 2,550,202, 1983; FR (19) (a) Young, J. F.; Osborn, J. A.; Jardine, F. A.; Wilkinson,
2,561,650, 1984; (b) Cornils, B.; Kuntz, E. J. Organomet. Chem. G. J. Chem. Soc. Chem. Commun. 1965, 131; (b) Evans, D.; Osborn,
1995, 502, 177; (c) Wiebus, E.; Cornils, B. Chem.-Ing.-Tech. 1994, J. A.; Wilkinson, G. J. Chem. Soc. 1968, 3133; (c) Evans, D.;
66, 916. Yagupsky, G.; Wilkinson, G. J. Chem. Soc. 1968, 2660.
(4) (a) Horváth, I. T.; Rábai, J. Science 1994, 266, 72; (b) Exxon (20) Yang, C.; Mao, Z. S.; Wang, Y.; Chen, J. Catal. Today 2002,
Research and Eng. Co. (Horváth, I. T.; Rábai, J.), EP 0,633,062 74, 111.
A1 1995 and US 5,463,082 1995; (c) Bebefice-Malouet, S.; Blancou, (21) Desphande, R. M.; Purwanto, Delmas, H.; Chaudhari, R.
H.; Commeyras, A. J. Fluorine Chem. 1985, 30, 171. V. Ind. Eng. Chem. Res. 1996, 35, 3927.
(5) (a) Vyve, F. V.; Renken, A. Catal. Today 1999, 48, 237; (b) (22) Jáuregui-Haza, U. J.; Pardillo-Fontdevila, E.; Kalck, P.
Fell, B.; Schobben, C.; Papadogianakis, G. J. Mol. Catal. A: Chem. Catal. Today 2003, 79-80, 409.
1995, 111, 179; (c) Haumann, M.; Koch, H.; Hugo, P.; Schomäcker, (23) Benaissa, M.; Jáuregui-Haza, U. J.; Nikovc, I.; Wilhelm,
R. Appl. Catal., A 2002, 225, 239. A. M.; Delmas, H. Catal. Today 2003, 79-80, 419.
(6) (a) Arhancet, J. P.; Davis, M. E.; Merola, J. S.; Hanson, B. (24) Zhang, Y.; Mao, Z. S.; Chen, J. Catal. Today 2002, 74, 23.
E. Nature (London) 1989, 339 (6224), 454; (b) Hjortkjaer, J.;
Scurrell, M. S.; Simonsen, P.; Svendsen, H. J. Mol. Catal. 1981, Received for review May 30, 2005
12, 179; (c) Arhancet, J. P.; Davis, M. E.; Merola, J. S.; Hanson, Revised manuscript received August 19, 2005
B. E. J. Catal. 1990, 121, 327. Accepted September 27, 2005
(7) (a) Holbrey, J. D.; Seddon, K. R. Clean Prod. Proc. 1999, 1,
223; (b) Welton, T. Chem. Rev. 1999, 99, 2071; (c) Dupont, J.; IE050629G

También podría gustarte