Vous êtes sur la page 1sur 43

Chapter 3

Hydroformylation of 1,4-diacetoxy-
2-butene using water-soluble
rhodium complex catalysts in a
biphasic medium
Chapter-3

3.1. Introduction
As already discussed in Chapter-1. the rhodium-catalyzed hydroformylation of alkenes
continues to attract considerable attention because of its wider ranging applications in the
manufacture of industrial chemicals in addition to the unique catalytic chemistry. The
chemistry associated with this reaction is rich and complex and despite the fact that it has
been extensively studied, many fundamental challenges in catalyst design remain open. A
striking drawback of this process is the difficulty in separating and recovering catalysts
from the mixture of reactants, products and solvents, especially for expensive catalysts
made from noble metals and non-volatile products. It is observed that the hydroformylation
products are non-volatile and thermally unstable in many cases and their separation from
catalysts poses a serious challenge. Different techniques have been used to heterogenize
the homogeneous catalysts retaining their high activity and selectivity, a relevant literature
review on which is presented in Chapter-1.

The biphasic hydroformylation of olefins catalyzed by water-soluble rhodium


complexes has attracted great attention in recent years due to a simple but efficient
separation of the catalyst from products. It was after the work of Kuntz2 on the synthesis of
triphenylphosphine trisulfonate (TPPTS) ligand and its application in the hydroformylation
of olefins, that the research on water-soluble catalysis gained momentum. The biphasic
(aqueous-organic) hydroformylation of propylene has been successfully commercialized to
produce the butyraldehyde using water-soluble rhodium complex as the catalyst. Although
hydroformylation has been applied to several long chain olefins, for the manufacture of a
range of aldehydes/alcohols (oxo-alcohols), the biphasic catalytic process has so far been
applied only to hydroformylation of propylene (Ruhrchemie / Rhone-Poulenc process)' and
butylenes , which have significant solubility in aqueous phase. This technique provides a
simple method for separation of reaction products in the organic phase from the catalyst in
the aqueous phase immediately after the reaction.'^ Besides development of new catalyst
present and ligands, the understanding of the overall kinetics of biphasic catalytic reactions
is an equally important aspect in the evolution of an economical process. Significant efforts
have been made to enhance the rates of biphasic catalytic reactions for long chain substrates
like I-dodecene, which have a relatively lower solubility in the aqueous phase,' by
developing novel active metal complex catalysts and ligands7 or by improving the

158
Chapter-3

miscibilitv and solubility of reactants. mainly by using amphophilic ligands1. co-ligands


like triphenylphosphine10 or cyclodextrins (SAPC)" or by adding co-solvents" and
surfactants '.
The difficulties encountered in catalyst-product separation for hydroformylation of 1.4-
diacetoxy-2-butene (DAB) [to Diacetoxy-formyl butane (DAFB) and 2-formyl-4-acetoxy-
2-butene] have been discussed in Chapter-2 with the objective of developing a biphasic
catalytic system for this reaction to take advantage of its easy separation, a detailed study
was undertaken using Rh-TPPTS catalytic system. A detailed investigation on variation of
activity, selectivity and stability for DAB hydroformylation using water soluble
[Rh(COD)Cl]2 / TPPTS catalyst has been reported in this chapter. Hydroformylation of
olefins using water-soluble catalysis is an example of a gas-liquid-liquid catalytic reaction,
in which reaction of two gaseous reactants (carbon monoxide and hydrogen) with liquid-
phase olefin occurs in the presence of a water-soluble catalyst in a liquid-liquid dispersion.
The reaction of dissolved gases and olefins occurs in the aqueous ("bulk") phase or organic
-aqueous interface. Overall rate of reaction in such a multiphase catalytic system depends
on gas-liquid and liquid-liquid mass transfer, the solubility of gas -phase reactants in the
organic and aqueous phases, the liquid-liquid equilibrium properties and intrinsic kinetics
of the reaction in the aqueous phase. Depending on the fractional hold-up of the aqueous
phase, it will be either a continuous aqueous phase with dispersed organic droplets or a
dispersed aqueous phase in a continuous organic medium. The coupled influence of mass
transfer with chemical reaction is expected to be quite different in these two situations. In
such cases, the importance of g-1 and 1-1 mass transfer effects is very crucial to establish a
kinetic regime. In this chapter, we also report a detailed investigation on the analysis of
mass transfer effects followed by kinetic study on DAB hydroformylation using water
soluble [Rh(COD)Cl]2 / TPPTS catalyst, in a temperature range of 338-358K.

3.2. Experimental
3.2.1. Materials
Rhodium trichloride (RhCl3.3H20) (Aldnch, U.S.A), triphenylphosphine (PPh3) (Loba
India), Butenediol (Merck. India), Acetic anhydride (Merck, India) were used as received
without further purification. Sulphuric acid, dimethyl formamide, acetyl acetonate and

l 59
Chapter-3

Sodium hydroxide (SD fine chemicals, India) were used as received. Oleum of 65 o (w \v
of SO;, in IFSO4) strength was prepared. Distilled degassed water was used in all
operations. Solvents, toluene, water, ethanol, cyclohexane were freshly distilled and
degassed prior to use. Hydrogen, nitrogen supplied by Indian Oxygen Ltd. Bombay, and
carbon monoxide (> 99.8 % pure) from Matheson Gas Co., U.S.A. were used directly from
cylinders. The syn gas with 1:1 ratio of lL: CO was prepared by mixing H2 and CO in a
reservoir. All operations were performed under argon atmosphere.

3.2.2. Instrumentation

'H NMR spectra were obtained on a Bruker AC-200 Spectrometer in CDCI3 at room
3I
temperature. P NMR spectra were obtained on a Bruker AC-200 or MSL-300
spectrometer in CDCI3 at room temperature. The peak positions are reported with positive
shifts in ppm downfield of TMS as calculated from the residual solvent peaks or downfield
of external H3PO4. FT-IR spectra were recorded on the Bio-Rad Spectrophotometer. The
analysis of reactants and products was carried out by a gas chromatographic method on a
5% phenylmethyl siloxane capillary column. For this purpose, HP 6890 gas chromatograph
was used.

3.2.3. Synthesis
(a) Synthesis of TPPTS
Sulfonated phosphmes are useful ligands for preparing water-soluble catalysts,
which can be used in biphasic mode. Triphenyl phosphine trisulfonate (TPPTS) forms
complexes with most of the transition metals, just like triphenyl phosphine (TPP)' 4 and
hence can be widely used for a variety of reactions. Moreover, it has a higher solubility in
water (1.1 kg/1) as compared to other phosphines \ For the synthesis of TPPTS a procedure
standardized by Bhanage"' was used.

Sulfonation Reactor Set Up


For synthesis of triphenyl phosphine trisulphonate (TPPTS), a double-jacketed one-liter
glass reactor equipped with a high-speed half moon stirrer with provision for monitoring
bulk liquid temperature was used. This reactor was designed so that operations under argon

160
Chapter-3

atmosphere were possible. The temperature was controlled by circulation of a constant


temperature fluid (50% ethanol-water mixture v/v) using a cryostat.

Reaction Charge for the synthesis of TPPTS


SO, TPP =12 (molar ratio)
SO,/(S(>, - H2SO4 + TPP) = 34% (w/w)
H2SO4/ (SO, + H2SO4 + TPP) = 56%(w/w)
Reaction temperature: 22C

200 g of 98% pure sulfuric acid was introduced into the reactor. The acid was cooled under
constant stirring to 12-15C by means of a cryostat. 50 g of triphenyl phosphine (TPP)
(190.75 mmol) was introduced slowly at 15C over a period of 30-45 min. This gave a
homogeneous yellow coloured solution of TPP in sulfuric acid. 280 g (141.48 ml, d=l .98
at 35C) of 65% oleum (SO3 content: 2.275 mol) was transferred into the addition funnel
from the oleum receiver. This oleum was then introduced in the sulfonation reactor
containing TPP solution in sulfuric acid, over a period of 40-45 min maintaining a
maximum temperature of 15C with good stirring. The temperature of the reaction mixture
was then raised to 22C and was maintained for 76 hrs. This step of keeping the reaction
mixture at 22"C is critical for optimum yield of TPPTS. [A cryostat was used for this
purpose]. Thereafter, the temperature of the reaction was lowered to ~ 10C and 50 g of
distilled water was introduced while maintaining the temperature at 10C. This addition is
necessary to quench the excess SO3 present after the reaction is complete. The addition of
water is highly exothermic, and hence temperature was maintained at < 10C while adding
water. This gave a solution of sulfonated triphenyl phosphine in sulphuric acid. This
reaction mixture was further diluted to approximately 800 ml under cooling (10C). This
diluted solution was then transferred into a neutralization reactor under Argon atmosphere.
The reactor was similar to the reaction vessel but of a larger capacity (approx. 3 liters). A
50% (w/w) degassed sodium hydroxide solution was added to this reactor maintaining
10"C temperature. At neutralization point, the solution turns yellow; under slightly alkaline
conditions the colour of the solution was distinct yellow. Neutralization was earned out to
a point where the solution was colourless (slightly acidic). [Note: Na2SC>4 precipitated out

161
Chapter-3

much before neutralization was complete and the reaction mixture obtained was in a slurry
form]. The neutralized mixture was then filtered and the filtrate was evaporated under
reduced pressure at 80-90"C till the volume reduced to about 250 ml from the initial two
liters. 1.5 liters of methanol was added to the above solution and the mixture was refluxed
under argon atmosphere for 2 hours. TPPTS dissolved in methanol completely, which was
filtered hot, with residue comprising only sodium sulphate. The filtrate was evaporated
under reduced pressure. The solid TPPTS obtained was weighed and stored under argon
atmosphere. The yield was found to be 80-85%. j l P NMR analysis was comparable with
that reported in the literature,17 showing approximately 90% TPPTS and 10% OTPPTS
formation. No further purification of TPPTS was undertaken for use as a ligand for the
preparation of water-soluble catalyst precursors.

(b) Synthesis of [Rh(COD)CI]2


[Rh(COD)Cl]2 was prepared according to a procedure reported by Chatt and Venanzi18
Rhodium trichloride trihydrate (1 x 10~3 kg) in ethanol (3 x 10"5 m3) was boiled under reflux
with a solution of 1,5 cyclo-octadiene (COD) (2 x 10"6 m3) for 3 hours. The solution was
cooled to obtain an orange solid material, which was filtered, washed with ethanol, dried
and recrystallized from acetic acid. (Yield=60%) The complex was characterized by the IR
spectroscopy. The elemental analysis for [Rh(COD)Cl]2 was found- C: 38.95, H: 4.9, CI:
14.5; Calculated: C: 39.05, H: 5.0; CI: 14.4.

(c) Synthesis of 1,4-diacetoxy-2-butene (DAB)


DAB was prepared according to a procedure described by Brevet and co-workers'9. Acetic
anhydride was added to a solution of 2-butenel,4-diol (10 x 10"3 kg, 114 mmol) in
anhydrous pyridine (50 x 10"' m ) with stirring and cooling in an ice bath. The stirring was
continued for 14 h at room temperature and the mixture was poured into ice water (150 x
10"(> m3) and extracted with diethyl ether [4 x (130 x 10"6 m3]. The diethyl ether solution
was washed sequentially with saturated aqueous CuS0 4 [2 x (60 x 10"6 m3)], saturated
aqueous NaHC0 3 (50 x 10"6 m3), H 2 0 [2 x (40 x 10'6 m3)], brine [2 x (50 x 10"(l m3)] and
treated with dried MgSOj and concentrated in vacuo. The residual liquid was
chromatographed on silica gel, to give l,4-diacetoxy-2-butene as a colourless oily liquid in

162
Chapter-3

the purest form. 1.4-diacetoxy-2 butene was characterized by using H NMR, IR. Elemental
analysis and was confirmed by GC-MS (Mass Peak at 172). For spectroscopic details refer
to the Experimental section of Chapter-2.

(d) Preparation of Rh(CO)2(acac)


Rh(CO)2(acac) was prepared by a method described by Varshavskii and Cherkasova" .
Acetylacetonate (12 x 10"'' m3) was added to a solution of RhCb (3.Ox I0"3 kg) in DMF (60
x 10"6 m') with stirring. The solution was refluxed for 30 minutes and then cooled. It was
diluted to twice the volume with distilled water. Addition of water resulted in a voluminous
crimson precipitate. The precipitate was filtered and washed with alcohol and ether. The
complex was recrystallized from hexane solution. The red green crystals of needle shape
were obtained by slow cooling of the hexane solution. The yield was about 70%. The
elemental analysis of Rh(acac)(CO)2 was found- C: 23.31, H: 2.68 and Calculated: C:
23.25, H: 2.71. Characteristic IR shifts for this complex are at 2068 cm"1, 1992 cm"' and
1518 cm"1.

(e) Preparation of HRh(CO)(TPPTS)3


A TPPTS solution of 400 x 10"6 kg (0.704, mmol) in 1 ml water was added to 50 x 10"6
kg (0.194 mmol) acetylacetonato dicarbonyl Rhodium (I), [Rh(CO)2(acac)], under stirring
to a small two neck flask kept under argon. The solution turned maroon coloured after
dissolution of Rh(CO)2(acac). Syngas (H2/CO 1:1) was introduced to the flask at room
temperature and atmospheric pressure. The colour of solution changed from maroon to
yellow in approximately 5 mm. The contents were stirred further for 6h under H2/CO (1:1)
atmosphere. Then the solution was filtered under positive flow of argon to remove small
amounts of rhodium metal. To the solution, 8 x 10"6 m3 of absolute alcohol saturated with
Hi/CO (1:1) was added to yield a yellow precipitate. The solid collected was washed with
absolute alcohol, and vacuum dried. 380 mg of HRh(CO)(TPPTS)3 was obtained. The
complex was characterised by 31P NMR (43.1 ppm with jRh-p of 150 Hz) and found to be
comparable to that reported in the literature21.

163
Chapter-3

3.2.4. Experimental Set Up


All the Hydroformylation experiments were carried out in a 50 ml (50 x 10"' nr) micro
reactor, made of stainless steel (Maximum pressure capacity of 20.7 MPa at 548 K),
supplied by Pair Instrument Company, USA. The reactor was provided with arrangements
for sampling of liquid and gaseous contents, automatic temperature control and variable
agitation speed. The reactor was designed for a working pressure of 20.4 MPa and
temperature up to 523K. The consumption of CO and H2 at a constant pressure was
monitored by observation of the pressure drop in the gas reservoir as a function of time. A
schematic diagram of the experimental set up is shown in Figure-2.5, in Chapter-2.

3.2.5. Experimental Procedure ([Rh(COD)CI]2/ TPPTS Catalyst System)

In a typical experiment, a solution of [Rh(COD)Cl]2 (1.622 x 10"3 kmol/m ) in toluene


(15 x 10"6 m3) was added to a solution of TPPTS (16.22 x 10~3 kmol/m3) in water (10 x 10"6
nr). The rhodium complex in toluene was extracted with TPPTS in the aqueous phase and
the organic phase (toluene phase) was discarded (after ensuring that the entire rhodium
complex had been extracted as indicated by decolourisation of the organic phase). This
aqueous catalyst solution containing Rh-TPPTS complex was charged into the reactor along
with DAB (0.116 kmol/m3) in toluene (15 x 10"6 m3), which comprises the organic phase
for the reaction. The contents were flushed with nitrogen and then with a mixture of CO
and H2. Heating was started to attain a desired temperature, and then a mixture of CO and
H2 (in a required ratio, 1:1) was introduced into the autoclave up to the desired pressure (6.8
MPa). A sample of the liquid mixture was withdrawn, and the reaction started by switching
the stirrer on. The reaction was then continued at a constant pressure of CO + Hi (1:1) by
supply of syngas from the reservoir vessel through a constant pressure regulator. Since, in
this study the major product formed was an aldehyde, supply of CO + H2 at a ratio of 1:1
(as per stoichiometry) was adequate to maintain a constant composition of H2 and CO in the
reactor as introduced in the beginning. This was confirmed in a few cases by analysis of CO
content in the gas phase at the end of reaction.

164
Chapter-3

3.2.6. Analytical methods


Liquid samples were analyzed using a Hewlett Packard 6890 Series GC controlled by the
HP Chemstation software and equipped with an auto sampler unit, by using an HP-5
capillary column (30 m x 320 um x 0.25 urn film thickness with a stationary phase of
phenyl-methyl siloxane). The quantitative analysis was earned out external standard
method by constructing a calibration-table for reactants and products in the range of
concentrations studied. The conversion of DAB, selectivities of products (DAFB and FAB),
and catalytic activity defined as TON and TOP (hr~ ) were calculated using the formulae
given in Chapter-2. The standard GC conditions for the analysis of reactants and products
of different reactions are the same as given in Chapter-2, Table-1.1. Since, in the present
case two phases are involved (Aqueous and Organic phases), the two phases were separated
and then analyzed separately.

3.2.7. Solubility Data

For kinetic study, a knowledge of the solubility of H2 and CO in organic phase and
aqueous phase is required. The solubility data for H2-water and CO-water systems (at 338-
358 K temperature range) were estimated by extrapolation of the previously reported
literature solubility data by Chaudhari and coworkers22 and Delmas and coworkers23 and
Hand-book data2"'' "\ The solubility data obtained from literature and the extrapolated
values at different temperatures are shown in Figure-3.1 and the extrapolated values for the
temperature range (338-358K) required for this work are shown in Table-3.1.

165
Chapter-3

Temperature, K
Figure-3.1: Solubility ofH2 and CO in water (Plot of literature values)

Table-3.1 Solubility ofH2 and CO in water


Temperature, K HA, HB,
3
MPa m /kmol MPa rnVkmol
338 142.5 156.7
348 136.9 158.0
358 131.6 159.14

The solubility of CO and H2 at 338-358 K in toluene and toluene-DAB mixtures was


experimentally determined and the results are presented in Tables-3.2 and 3.3. The
solubility of CO and H2 in toluene obtained experimentally at 338-358 K was also
compared with the literature26 (extrapolated) values and is shown in Figure-3.2. It was
found that the experimental values compare very well with the literature extrapolated
values, as shown in Figure-3.2.

1M>
Cliapter-3

140 140
o
120
CO (Literature) 120
E
n - - H (Literature)
E *- CO (Experimental)
TO 100 100
Q. *- H (Experimental)
s 80 80

c 60 60
o
o
40 40
c
I 20- 20

335 340 345 350 355 360


Temperature, K

Figure-3.2: Solubility ofH? and CO in Toluene (Plot of literature values)

Table-3.2: Henry's constants of CO in toluene and toluene-DAB mixture


HB HB HB
Gas, Solvent MPa mVkmol MPa nrVkmol MPa m3/kmol
at 338 K at 348 K at 358 K
CO in Toluene 19.2 21.6 24.5
CO in 4% DAB in toluene (v/v) 19.25 21.62 24.6
CO in 8% DAB in toluene (v/v) 19.31 21.73 24.62

Table-3.3: Henry's constants of Hi in toluene and toluene-DAB mixture


HA HA HA
3 3
Gas, Solvent MPa m /kmol MPa m /kmol MPa m3/kmol
at 338 K at 348 K at 358 K
ITT in Toluene 69.9 84.7 100.8
H2 in 4% DAB in toluene (v/v) 70.1 84.81 100.9
H2 in 8% DAB in toluene (v/v) 70.15 84.9 100.9

167
Chapter-3

The solubility of DAB in the aqueous phase was experimentally determined and also
validated by doing calculations using the UNIFAC approach as described by Prausnitzz .
Applicability of this method has already been demonstrated wherein the comparison of the
experimental data with the predictions showed a good agreement28. The validity of this
method has also been demonstrated by Chaudhari and coworkers" for prediction of liquid-
liquid equilibrium data for 1-octene-ethanol-water system at 323 K. The experimental data
for liquid-liquid equilibrium for DAB in toluene-water mixture is shown in Table 3.4. It
was observed that temperature had only a marginal effect on the solubility of DAB in the
aqueous phase.

Table-3.4: Liquid-Liquid equilibrium data for DAB-water-toluene system


DAB in aqueous Ratio of aqueous
Temperature, DAB in organic
phase x 10 , to organic cone,
K phase, kmol/m3
kmol/m 3
MDxl03
338 0.116 1.02 8.8
348 0.116 1.08 9.3
358 0.116 1.10 9.5

The Henry's constants for hydrogen (HA) and carbon monoxide (HB) (given in Table-
3.2 and 3.3) and the partition coefficients for DAB (given in Table-3.4) were used to
evaluate concentrations of H2 and CO and DAB in the aqueous phase required for rate
analysis.

3.3. Results and Discussion


The objective of this work was to study the kinetics of hydroformylation of DAB using
water soluble [Rh(COD)Cl]2/ TPPTS complex catalyst. For this purpose, the effect of
different parameters such as catalyst concentration in aqueous phase, hydrogen and carbon
monoxide partial pressures and DAB concentration on the initial rate of hydroformylation
was studied in a temperature range of 338-358 K. The results are discussed in the following
sections.

168
Chapter-3

3.3.1. P r o d u c t Distribution and Selectivity


The complex [RhCl(l,5-COD)]2 (where l, 5-COD is 1,5-cyclo-octadiene) is highly
insoluble in water, however, displacement of the COD ligand with the TPPTS affords a new
complex [RhCl(TPPTS),,] which is highly soluble in water'". This complex was
successfully used as a catalyst precursor for the preparation of butyraldehyde from
propylene by Ruhrchemie ^ and thereafter in most of the published reports' on biphasic
hydrofonnylation of olefins. Literature reports on hydrogenation reactions show that the
starting from the water soluble analogue of the Wilkinson's complex [RhCl(TPPTS)3], a
dimeric species is obtained in equilibrium, which is similar to that obtained from the
catalyst precursor [Rh(COD)Cl]?,as shown in Scheme-3.132.

C l s JPPTS _TppTS C l
JPPTS
s
,Rh . Rh
TPPTS TPPTS TPPTS' "Solvent

TPPTSX ,CI TPPTS ^. ,CI /-.


>h >h ^ = ^ C^Rh > h ^ D
V /
TPPTS CI TPPTS - CI

H 2 /CO

Active Catalyst

Scheme-3.1: Active catalyst precursor formation from [Rh(COD)Cl]2 and TPPTS

It has already been demonstrated in Chapter-2, that the major products formed during
the hydrofonnylation of DAB using homogeneous HRh(CO)(PPh3)3 catalyst, are DAFB
and FAB, the latter being due to elimination of acetic acid from DAFB. Several
preliminary experiments on hydrofonnylation of DAB using water soluble [Rh(COD)Cl]2-
TPPTS catalyst in a two-phase system were carried out to assess stability of biphasic
system, material balance and product distribution. The stoichiometric reaction is shown in
Scheme-3.2.

\W
Chapter-3

Homogeneous System
HRh(CO)(PPh,U ? CHO
-HOAc J
CH,OCO / ^ / \ L i . CH3OCO 1 / \
\ / \ / OCOCH3 H2/CO \y\X OCOCH3 " \ s OCOCH3
DAB 1000psig/75C DAFB FAB

^""""o^^^ Biphasic system ^^^**

[Rh(COD)CI] 2 / TPPTS

Scheme-3.2: Hydroformylation of DAB using water soluble [Rh(COD)Cl]2-TPPTS catalyst


in a two-phase system.

The results (Concentration-Time Profile) of a typical experiment carried out at 348 K and
6.8 MPa total pressure of CO/H2 (1:1) are shown in Figure-3.3. These results represent
concentrations in organic phase. It was observed that the concentration of DAB and FAB
were negligibly small over the entire batch run in the aqueous phase. This indicates a
complete extraction of the products in the organic phase after reaction in aqueous phase.
The analysis in the organic phase accounted for more than 95% of the material balance of
DAB and FAB.
It was observed that almost complete conversion of DAB.was achieved in 60 minutes
and material balance of CO or H2 and DAB consumed was in good agreement with the total
amount of aldehydes formed. In the range of conditions investigated, the only product
formed was the deacetoxylated aldehyde, formyl acetoxy butene (FAB). No other side
products were formed during the hydroformylation reaction.

170
Chapter-3

1
' 1 i ' i
0 12-
DAB --
0.10- FAB ^ T ^
o
o
E
c
o
0.08-
*K E

0.06-
. -o
0)
n.
L.
c o
o 0.04-
-- 1 w
c
o o
u 0.02- o

0.00- 1 ' i i - i > i i i

10 20 30 40 50 60
Time, Min

Figure-3.3: A typical C-T profile for the hydroformylation of DAB using soluble
[Rh(COD)Cl]2-TPPTS catalyst in a two-phase system

Reaction Conditions: DAB: 0.116 kmol/m3, Two Phases: Toluene and Water, s (aqueous phase hold
up) = 0.4, [Rh(COD)Cl]2: 1.622 x 10 kmol/m3, TPPTS: 16.22 x 10"3 kmol/m3, Agitation Speed =
18.3 Hz, Total volume: 2.5 xlO"5 m3 ,T =348K., P c o = 3.4 MPa, PH2 = 3.4 MPa

3.3.2. Effect of Solvents


For the aqueous biphasic hydroformylation of DAB, different organic solvents were
screened and the results are presented in Table-3.5. Toluene was selected as the organic
solvent mainly because of the ease of separation, low volatility, non-reacting and non-toxic
nature. As seen in Table-3.5, higher rates of reaction were obtained for hexane and
cyclohexane as the solvents. The reason for not selecting these solvents for further detailed
kinetic investigation was the partial solubility of the product formed (FAB) in these
solvents during the hydroformylation reaction.

171
Chapter-3

Table-3.5: Effect of solvents on the activity and selectivity ofBiphasic hydroformylation


DAB

S. No. Solvent Initial Product % Product in /o


Rate of Distribution the organic Rhodium
the phase Leaching
reaction,
kmol/m /s

1 Toluene 7.93 x 10 5 100% FAB 100% < 0.03%

2 Hexane 8.2 xlO" 5 100% FAB 65-70% -2%

3 Cyclohexane 9xl0"5 100% FAB 65-70% 5%

4 Ethanol 5.23 x 10"5 FAB + Acetals . 10-12 %

Reaction Conditions: DAB: 0.116 kmol/m3, Two Phases: Organic Solvent and Water, e = 0.4,
[Rh(COD)Cl]2 : 1.622 x 10'3 kmol/m3, TPPTS: 16.22 x 10"3 kmol/m3, Agitation Speed = 18.3 Hz,
Total volume: 2.5 xl0~5 m3, T =348K., PC0 = 3.4 MPa, PH2 = 3.4 MPa

3.3.3. Test of rhodium Leaching

The major advantage of biphasic catalysis is the efficient and simple recycle of the
aqueous catalyst phase. This is achieved by a simple phase separation of two immiscible
liquid phases. For the recycle experiments, the aqueous phase catalyst was used in several
successive experiments with fresh charge of the organic phase each time. The catalyst was
recycled almost four times without loss of activity as shown in Figure-3.4. The rates as
well as the selectivity behavior towards aldehyde products were found to be unaffected in
recycle experiments indicating the reusability of the catalyst. The initial rates of
hydroformylation were calculated from the observed CO or H2 consumption-time profile.
Under the conditions chosen for the kinetic study, no side reactions were found to occur and
hence, these data represent the overall hydroformylation of DAB to the corresponding
aldehyde (FAB).

172
Chapter-3

H Rate Selectivity to FAB


m v v w 100
CO

E
O 8 80 ^
CO
Tf <
o UL
T- 6- 60 o
+
X
< &
">
4- -40 ~
TO <u
Of 0)
to
to o- - 20
+3
|E
n
U n i i i -0
Recycle-1 Recycle-2 Recycle-3 Recycled

Figure-3.4: Catalyst Recycle Experiment

Reaction Conditions: DAB: 0.116 kmol/m3, Two Phases: Toluene and Water, e = 0.4,
[Rh(COD)Cl]2: 1.622 x 10 kmol/m3, TPPTS: 16.22 x 10 3 kmol/m3, Agitation Speed = 18.3 Hz,
Total volume: 2.5 xlO"5 m3, T =348K., P c o = 3.4 MPa, PH2 = 3.4 MPa

Another important observation in case of biphasic hydroformylation of DAB is the


sequence of addition of the reactants and the catalyst to make the reaction charge. It was
observed that when [Rh(COD)Cl]2 was extracted with water soluble TPPTS first, followed
by addition of DAB in Toluene to the aqueous layer, the reaction proceeds smoothly and
there was no leaching of rhodium in toluene-water biphasic system (Scheme-3.3, Case-A).
On the other hand, as reported in several publications for other olefins \ if the initial
reaction charge was made from [Rh(COD)Cl]2-TPPTS-DAB and toluene followed by
catalyst extraction and phase separation (as shown in Scheme-3.3, Case-B), there is a
considerable leaching of the [Rh(COD)Cl]2 in the organic layer. This could be because of

173
Chapter-3

the very strong affinity of DAB to form a coordinated Rhodium complex in contrast to the
formation of Rh-TPPTS water-soluble complex in the absence of water.

[Rh(COD)CI]+Toluene
[Rh(COD)CI] 2 + T o l u e n e + DAB

PTS
PTS

Toluene Phase Rh-TPPTS in Water

DAB in Toluene Toluene Phase Water Phase


(Toluene phase contains
Rh complex)
Reaction charge

Case-A (No Leaching is observed) Case-B (Leaching of rhodium complex


in the Organic Phase)

Scheme-3.3: Leaching effect caused by different Sequence of addition

3.3.4. Physical description of the Biphasic system

Hydroformylation of DAB in a biphasic medium is a multiphase reaction since, it


involves three phases (gas-liquid-liquid) containing two gaseous reactants H2 and CO with
DAB as a liquid reactant in the organic medium and the catalyst in the aqueous medium.
The overall reaction rate depends on transport of H2 and CO to the liquid phase, followed
by transport of dissolved H2, CO and DAB from organic phase to aqueous phase and
reaction in the aqueous catalyst phase. A schematic representation of biphasic
hydroformylation reaction is shown in Figure-3.5.

174
Chapter-3

H2 00
Caseous Keactant

J^y^a
Organic Phase
(Reactant+Products+Solvent)
fi(a<g) B(org) D(ag) Rap)

It t It t
/Kaq) B(aq) D(aq) P(aq)

FWTPFTB catalyst
Aqueous Phase (Catalyst)

Figure-3.5: Schematic representation ofbiphasic hydroformylation of DAB

The factors affecting the rate behavior of gas-liquid-liquid catalytic reactions have been
discussed by Chaudhari and coworkersj4. The overall performance of such reactions would
depend on a number of parameters such as reaction kinetics, gas-liquid and liquid-liquid
mass transfer, solubility of gases in organic and aqueous phases, liquid-liquid equilibrium
and the complex hydrodynamics of the dispersed liquid droplets. The various
parameters/steps governing the rate of gas-liquid-liquid reactions are shown in Figure-3.6.
In order to understand the rate behaviour, a detailed analysis of these phenomena with
relevant experimental studies are necessary.
Chapter-3

-liquid-liquid -of the gaseous reactants in


(in the presence of a the aqueous phase
dispersed gas phase) -of the organic reactants in the
catalytic phase
-gas-liquid
(in die presence of a dispersed -dependence on pressure
second liquid liquid phase)
-influence of eosolvent

Inter-phase Solubility
mass transfer

_ Overall reaction rate ^_


Of biphasic gas-liquid-liquid reaction

Thermodynamic Intrinsic
phase equilibria kinetics

-Influence of concentration
- determination of reactant and
(reactant, metal, ligand),
product concentrations in the
partial and total pressure,
aqueous and organic phase
temperature

- determination of reaction
network and kinetic parameters

Figure-3.6: Parameters/steps in gas-liquid-liquid reactions .

3.3.5. Mass transfer effects in Gas-Liquid-Liquid systems


For investigation of intrinsic kinetics of multiphase reactions, it is important to ensure
that the rate data obtained are in the kinetic regime. A careful consideration to the
significance of mass transfer steps must be given for kinetic analysis of a complex
multiphase reaction involving gas-liquid-liquid system. For such a case, the overall rate
would depend on the gas-liquid, liquid-liquid mass transfer and the intrinsic kinetics of the
reaction in the aqueous phase.

3.3.5.1. Agitation and Aqueous Phase hold up effect


A few experiments were carried out to investigate the effect of agitation speed and
aqueous catalyst phase hold up on the rate of reaction. These experiments were important to

I 76
Chapter-3

understand the role of mass transfer and ensure the kinetic regime. The experiments were
carried out in a temperature range of 338-358 K, The following reaction conditions were
used for these experiments: DAB: 0.116 kmol/m\ Two Phases: Toluene and Water.
(Aqueous phase hold up) 8 = 0.4, [Rh(COD)Cl]2 : 1.622 x 10"' kmol/m3, TPPTS: 16.22 .\
10"3 kmol/nr, Agitation Speed = 18.3 Hz, Total volume: 2.5 xl0~5 m \ T =348K.
Figure-3.7 shows that the rate of reaction increases with agitation speed, up to 700 rpm
but for agitation speed greater than 700 rpm, the rate of reaction does not change. This
observation indicates that the mass transfer effects are not important above 700 rpm.

|
o
E

400 600 800 1000 1200 1400


Agitation, rpm

Figure-3.7: Effect of agitation speed on initial rate of hydroformylation


Reaction Conditions: DAB: 0.116 kmol/m\ Two Phases: Toluene and Water, E = 0.4,
lRh(COD)Cl]2: 1.622 x 10" kmol/m3, TPPTS: 16.22 x 10~3 kmol/m'. Agitation Speed = 18.3 Hz,
Total volume: 2.5 xlO5 nr ,T =348K, P c o = 3.40 MPa, P2 = 3.40 MPa.

Figure-3.8 shows the effect of aqueous catalyst phase hold-up (e) on the initial rate of
hydroformylation of DAB at various agitation speeds. It was observed that the phase
inversion takes place at aqueous phase hold up of about 0.6. For aqueous phase hold-up less
than 0.6, the aqueous phase is the dispersed phase, as shown schematically in Figure-3.9
(a). In this case, the liquid-liquid interfacial area is determined by aqueous phase hold-up

177
Chapter-3

(<;). For aqueous phase hold-up greater than 0.6, the organic phase is the dispersed phase, as

shown schematically in Figure-3.9 (b). In this case, the liquid-liquid interfacial area will be

determined by the organic phase hold-up (1-e).

25 H

E
=5 20 H
o -1100 rpm
E - 800 rpm
o 15H

DC
10-^
TO
DC

e 5-

1.0
Aqueous Phase Hold Up

Figure-3.8: Effect of aqueous phase aqueous phase holdup on initial rate of


hydroformylation

Reaction Conditions: DAB: 0.116 kmol/m, Two Phases: Toluene and Water, E = 0.4,
[Rh(COD)Cl] 2 : 1.622 x 10"3 kmol/m3, TPPTS: 16.22 x 10 3 kmol/m3, Agitation Speed = 18.3 Hz,
Total volume: 2.5 xlO"5 m3 ,T =348K, P c 0 = 3.40 MPa, PH2 = 3.40 MPa.

At agitation speed of 1100 rpm (18.3 Hz), a plot of initial rate vs. aqueous (catalyst)
phase hold-up (Figure-3.8) shows that the rate first increases with increase in aqueous phase
hold up and then passes through a maximum. In kinetic regime, the rate per unit volume of
the aqueous phase is expected to remain constant with linear dependence on aqueous phase
hold up. However, in the case where the'reaction occurs essentially at liquid-liquid
interface, it would depend on the liquid-liquid interfacial area, which is governed by both
agitation speed and hold up of the dispersed phase. The results at aqueous phase hold up
less than 0.5 indicate a kinetic regime. For higher aqueous phase hold up, the decreasing
rate is a result of phase inversion with organic phase dispersed in continuous aqueous

178
Chapter-3

phase. In order to understand this effect a more detailed analysis of mass transfer in
biphasic hydroformylation is necessary. For the purpose of kinetic studies the data below
aqueous phase hold up of 0.5 (at 0.4) were used, wherein kinetic regime prevails.

jGas-liquid boundary

Organic phase (continuous)


r Aqueous phase (continuous)

Aqueous phase droplets Organic phase droplets

(a) Aqueous phase dispersed (b) Organic phase dispersed


(0<s<0.5) (0.5<e<1.0)
Figure-3.9: Schematic presentation of two different physical situations prevailing in the
reactor depending upon the phase hold up

3.3.5.2. Gas-liquid mass transfer effect

The significance of gas-liquid mass transfer resistance was analyzed by comparing the
initial rate of reaction and maximum possible rate of gas-liquid mass transfer. The gas-
liquid mass transfer resistance is negligible if a factor a/ defined as follows is less than 0.1.
R,
3.1
k G C
t B A,X

R,
a, 3.2

Where, RA is the observed rate of hydroformylation (kmol/m3), kLaB the gas-liquid mass
transfer coefficient and Q and CR represent the saturation solubility of reacting gases i.e.
CO and HT in equilibrium with the gas phase concentration at the reaction temperature
(kmol/m"). The equilibrium solubilities for the gases given in the respective tables (Table-
3.1-3.4) were used. The gas-liquid mass transfer coefficient (kLas) used in above equations

179
Chapter-3

was calculated using a correlation proposed by Chaudhari and coworkers''6 for a reactor
similar to that used in this work.

k,aH = 1.48 x 10"3 (A'f IH


x {\\,, VL )' 8 S x {d,/dTYA x ( h,/h )'"' 3.3

The terms involved in above Equation are described in Table-3.6 along with the respective
values obtained from the reactor and charge used in the present case. The ki_aB value for
1100 rpm(18.3Hz) was evaluated as 0.41 s"1.

Table-3.6: Parameters used for kian calculations by Eq-3.3

Parameter Description Value


vg Gas volume (mJ) 65 x 10"
N Agitation Speed (Hz) 18.3
VL Liquid volume (mJ) 25.4 x 10"6
di Impeller diameter (m) 1.9 x 10"2
dT Tank diameter (m) 3.6 x 10"2

h, Height of the impeller from the bottom (m) 1.3 x 10~2

h2 Liquid height (m) 3.3x10"-'

The factor ot|,A and ctiB for both hydrogen and carbon monoxide were calculated and found
to be 9.3152 x 10"4and 3.824 x 10"4, respectively. Since, the values of aA and an are very
much less than 0.1 for both the gaseous reactants, gas-liquid mass transfer resistance can be
assumed to be negligible.

3.3.5.3. Liquid-liquid mass transfer effect


To ensure that the liquid-liquid mass transfer of DAB is not rate limiting, the maximum
rate of reaction (RA) was compared with the maximum rate of mass transfer (KHCDAB)- The
liquid-liquid mass transfer resistance is negligible if a factor a2 defined as following is less
than 0.1.

180
Chapter-3

R
i
a, = \ 3.4
K
n Qua
Where, Kn is the Liquid-Liquid mass transfer coefficient and is given by Lq-3.5
K
n =an xk
i.aq 3.5
an, the liquid interfacial area is given by Eq-3.6.
6e
a = 3.6
dP
Where, dp is the maximum drop diameter (m) and e is the dispersed phase hold up.
If we consider that the organic (dispersed) phase is present in the form of spherical droplets
with a negligible slip velocity with respect to the continuous phase, K|-aq can be estimated
(by using Sherwood number, Sh = 2.0 ) by Eq-3.7
2D
V, - 3.7
dp
Therefore,
\2Dwe
A=-^- 3.8
dp'

Where D A B is the molecular diffusivity of DAB (m2/s) in toluene calculated using Wilke
and Chang correlation37 given by Eq-3.11
For our experiments on hydroformylation, the system (toluene +water) was same as used by
Lagisetty and coworkers . So the correlation given by Eq-3.9 was directly used for
determining dp and liquid-liquid interfacial area an by Eq-3.6

^ = 0.083 (1 + C e)2 We" 3.9


d,
Where di is the impeller diameter (m), C constant (C=4)3 dependent on the geometry of the
vessel and agitator; and We is the Weber number calculated by Eq-3.10

We = N'd,3p 3.10
a
1 lere. p is the Density of the continuous phase and a is the Interfacial tension

Diffusivity of DAB (DAB) was calculated using Wilke-Chang Correlation37 given in Eq-
Chapter-3

7.4x10 s 7"(#/ /> ,)' " , ,,


/) 1 .1.11
/'' I

Where, <I> is the association parameter of the solvent; u.B is the viscosity of B (T or kg/m.s),
Mi, is the molecular weight of the solvent B; T is temperature (K), VA is the solute molar
volume. For agitation speed of 1100 rpm at which all experiments have been carried out,
the values of cr> were found to be in a range of 2.34 x 10"2- 0.46 x 10", which indicates a
kinetic regime.

3.3.6. Kinetics of Hydroformylation of DAB

3.3.6.1. Experimental data


The kinetics of hydroformylation of only a few functionalized olefins is studied using
Rh/TPPTS two-phase catalysis. The kinetic studies provide valuable information for
understanding the mechanism of the reaction as well as for reactor design. The kinetics of
hydroformylation of DAB was studied using a [Rh(COD)Cl]2 and TPPTS catalyst system.
For this purpose, several experiments were carried out in the range of conditions given in
Table-3.7. The effect of catalyst concentrations, DAB concentration, Partial pressures of
CO and lb was studied in the given range of conditions (given in Table-3.7) at 1100 rpm to
ensure a kinetic regime.

Table-3.7 Range of conditions used for the kinetic studies


Parameter/ Units Quantity

Concentration of catalyst [Rh(COD)Cl]2 (kmol/m3) (0.811-3.242) x 10"3


Concentration of TPPTS (kmol/m3) 16.22 x 10"3
Concentration of DAB (kmol/m3) 0.058-0.232
Partial pressure of hydrogen, MPa 2.040 - 4.760
Partial pressure of carbon monoxide, MPa 2.040-4.760
Temperature, K 338-358K
Solvent Toluene-Water
Aqueous phase hold up, r. 0.400
Agitation speed, rpm 1100

1X2
Chapter-3

Reaction volume, irf 2.500 x 10

Effect of catalyst concentration

The effect of catalyst loading on the rate of hydroformylation of DAB was investigated at a
CO and H : (1:1) pressure of 6.80 MPa, DAB concentration of 0.116-kmol/m 3 and TPPTS
concentration of 16.22 x 10 kmol/m"1. The result is shown in Figure-3.10.

Catalyst Concentration x 103, kmol/m3

Figare-3.10: Effect of catalyst Loading variation on the Biphasic hydroformylation of DAB.


Reaction Conditions: DAB: 0.116 kmol/nr, Two Phases: Toluene and Water, z -= 0.4, TPPTS:
16.22 x 10"3 kmol/m1, Agitation Speed = 18.3 Hz, Total volume: 2.5 x 10"5 m \ T =348K, P c o =
3.40 MPa. P H: = 3.40 MPa.

The catalyst concentration effect studied at three temperatures shows an increase in the
initial rate of the reaction with an increase in the catalyst concentration as expected for
kinetic regime.

1 S3
Chapter-3

Effect of partial pressure of hydrogen (PH2)

The elTcct of partial pressure of hydrogen on the rate of hydrofbrmylation of DAB was
investigated at a constant CO partial pressure of 3.40 MPa, DAB concentration of 0.1 16
k m o l m 3 . [Rh(COD)Cl] 2 concentration of 1.622 x 10 kmol/m' and TPPTS concentration of
16.22 x 10"' kmol/m'. The results are shown in Figure-3.11.
c

i | i | i | . |
kmol/m Is

338K
b

348K / * '
A 358K yS
- 15-
O
X
<
CH i.o- s' ^^****'*^
Initial Rate

s* ^ ^ * ^
en
o

^ -""""
0.0 Hl r I 1 . | i | i | i
1 2 3 4
Partial Pressure of H MPa

I''ignre-3.I J: Effect of Partial Pressure of Hydrogen on the biphasic hydroformylation of


DAB

Reaction Conditions: DAB: 0.116 kmol/m", Two Phases: Toluene and Water, z = 0.4,
[RhCODCl], : 1.622 x 10"3 kmol/m3, TPPTS: 16.22 x 10"3 kmol/m3, Agitation Speed = 18.3 Hz,
Total volume: 2.5x1 (f m 3 , T =348K, P., = 3.40 MPa.

The rate of reaction was found to be first order with PHZ- AS per the reported mechanism
(See Figure-3.12) for the hydroformylation of olefins, the oxidative addition of hydrogen to
the acyl carbonyl rhodium species is the rate-limiting step, leading to increase in rates at
higher partial pressures of hydrogen. Similar observations are reported for homogeneous
41 A2 A
as well as biphasic ' ' hydroformylation systems.

184
Chapter-3

[Rh(1,5 cyclo-octadiene)CI]2

H.,. CO " TPPTS

HRh(CO)(TPPTS)3
TPPTS
(OH)Rh(CO)(TPPTS)2
R-CH2-CH,-CHO H2

HRh(CO (TPPTS) 2 " " OH,


.CO

HRh(CO),(TPPTS).
'2\ ' /2, w

(H2)(R-CH2-CH2CO)Rh(CO)(TPPTS), R-CH^CH,

CH,3
r C3a
R-CH=CH,
(H2)(R-CHCO)Rh(CO)(TPPTS7
C2a ;HRh(CO)2(TPPTS)2

C2b
(R-CH2-CH2CO)Rh(CO)(TPPTS)2 R-CH2-CH2Rh(CO)2(TPPTS)2
CH,3 + ^ "" CH,3 +
I I
(R-CHCO)Rh(CO)(TPPTS)2 R-CHRh(CO)2(TPPTS)2
CO CO

(R-CH2-CH2CO)Rh(CO)2(TPPTS)2
CH, +
(R-CHCO)Rh(CO)2(TPPTS)2
CO co C

(R-CH2-CH2CO)Rh(CO)3(TPPTS)2
CH, +
(R-CHCO)Rh(CO)3(TPPTS)2

Figure-?*. 12: Mechanism for hydroformylation of olefins using water soluble


HRh(CO)(TPPTS)}

1X5
Chapter-3

Effect of partial pressure of CO (PCo)

The effect of I\ 0 on the rate of hydroformylation of DAB was studied at a constant


hydrogen partial pressure of 3.40 MPa, DAB concentration of 0.116 kmol/m''
[Rh(COD)Cl] 2 concentration of 1.622 x 10 ;' kmol/m 3 and TPPTS concentration of 16.22 x
10 kmol/m 1 , See Figure-3.13.

Figure-3.13: Effect of Partial Pressure of Carbon Monoxide on the biphasic


hydroformylation of DA B
Reaction Conditions: Partial pressure of Hydrogen: 3.40 MPa, DAB: 0.116 kmol/m''. Two Phases:
Toluene and Water, e = 0.4, [Rh(COD)Cl]2: 1.622 x 10"3 kmol/m\ TPPTS: 16.22 x 10"'1 kmol/m\
Agitation Speed = 18.3 Hz, Total volume: 2.5 x 10 nr .

The rate first increased with increasing P C o and then passed through a maximum, with
substrate-inhibited kinetics at higher Pco- Initially the increase in the partial pressure of CO
will cause an enhancement of Species-C 2 and C 3 (See Figure-3.12). Any further increase in
CO will also cause the formation of inactive dicarbonyl and tricarbonyl rhodium species-C'4
(See Figure-3.12). This will cause a decrease in the concentration of the active catalytic

1X6
Chapter-3

species, and hence lower rates of reaction will be observed. The negative effect of CO
concentration on the rate of hydroformylation has been previously well established, for
homogeneous as well as biphasic systems. Substrate inhibited kinetics for the homogeneous
hydroformylation has been reported for 1-hexene using HRh(CO)(PPh3)3 catalyst by
Deshpande and Chaudhari' , homogeneous hydroformylation of 1-dodecene using
HRh(CO)(PPh3)3 catalyst by Bhanage and Chaudharr4, homogeneous hydroformylation of
vinyl acetate using HRh(CO)(PPh3)3 catalyst by Deshpande and Chaudhari40, homogeneous
hydroformylation of allyl alcohol using HRh(CO)(PPh.3)3 catalyst by Deshpande and
Chaudhari ', homogeneous hydroformylation of styrene by Nair and Chaudhari42,
homogeneous hydroformylation of ethene using PPli3 modified Rh(acac)(CO)2 by Kiss and
co-workers4'' Substrate inhibited kinetics at higher PCo has been reported for biphasic
hydroformylation of 1-octene using [Rh(COD)Cl]2 / TPPTS complex using ethanol as a co-
4
solvent by Deshpande and Chaudhari and Purwanto and Delmas2j, biphasic
hydroformylation of ethylene using [Rh(COD)Cl]2 / TPPTS complex catalyst system by
Deshpande and Chaudharr".

Effect of DAB concentration

The effect of concentration of DAB on the rate of hydroformylation of DAB was studied at
a total pressure of 6.80 MPa (CO:H 2 = 1:1), [Rh(COD)Cl]2: 1.622 x 10 3 kmol/m3, TPPTS:
16.22 x 10 kmol/m . The results are shown in Figure-3.14.
The rate was found to be independent of DAB concentration. The addition of olefin to form
Rh-olefin complex (Species C2 to C3 in Scheme-3.12), is an equilibrium reaction. As DAB
is a strongly coordinating olefin [As can be seen in Section-3.3.3], the equilibrium may be
attained even at lower concentration of DAB leading to zero order dependence.

1X7
Chapter-3

| - 338K
o 1.0-
- 348K
E -* 358K

<

S
+- 0.5-
TO

0.0-
0.00 0.05 0.10 0.15 0.20 0.25

DAB Concentration, kmol/m

Fignre-3.14: Effect of DAB Concentration on the biphosic hydroformylation of DAB

Reaction Conditions: Two Phases: Toluene and Water, e = 0.4, [Rh(COD)Cl]2 : 1.622 x 10"-
kmol/m\ TPPTS: 16.22 x 10"" kmol/m3. Agitation Speed = 18.3 Hz, Total volume: 2.5 xlO"5 nr,T
348 K, \\;(,: 3.40 MPa, P ll: : 3.40 MPa.

3.3.6.2. Kinetic Modeling

For the purpose of kinetic modeling, only those experiments were chosen wherein the
criteria for kinetic regime were satisfied. Mainly the data at higher aqueous phase hold up
and higher catalyst concentrations were excluded, to eliminate any possibility of mass
transfer effects.
The mechanism of biphasic hydroformylation is well known (see Figure-3.12), and
hence, it was thought more appropriate to consider the molecular level approach for kinetic
modeling of water soluble catalytic hydroformylation instead of the empirical rate forms
considered in earlier reports"' " . The elementary steps in the mechanism shown in Figure-
3.12 can be simplified as shown in Scheme-3.4.

iss
Chapter-3

HRh(CO)L2 + CO ^ = ^ HRh(CO)2L2 (3.12)


C1 C2
K2
HRh(CO) 2 L 2 +RCH=CHR =^= Rh(CO)L2(COCHRCH2R) (3.13)
C2 DAB C3

Rh(CO)L2(COCHRCH2R) + H2 J X HRh(CO)L2 + RCH.CHRCHO (3.14)


C3 C1 Products

Rh(CO)L2(COCHRCH2R) + CO - ^ Rh(CO)2U(COCHRCH2R) (3.15)


C3 K C4

Rh(CO)2L2(COCHRCH2R) + CO ^ = ^ Rh(CO)3L(COCHRCH2R) (3.16)


C3 C5

Scheme-3.4: Simplified steps in the Mechanism of Hydroformylation. Here L = TPPTS

The intra-molecular rearrangement steps were considered as fast. Herrmann and


coworkers "* have shown that HRh(CO)(TPPTS)3 in aqueous medium can form
HO-Rh(TPPTS) 2 (CO), a species in equilibrium with the HRh(CO)(TPPTS) 3 . However, it
lias also been shown that under hydroformylation conditions, the hydroxy species is
quantitatively reconverted to HRh(CO)(TPPTS);? and does not interfere the
45
hydroformylation reaction . Considering the addition of H2 to acyl rhodium carbonyl
species was considered as a rate-determining step (as per the accepted hydroformylation
mechanism), the rate of the hydroformylation reaction was given as:

*,A-,/f 2 C 0 C,[CO][//,] 3 ,7
2
(\ + Kl[CO\ + K,K2[CO]Cl +KlK2Ki[CO] Cl + K.K.K.KJCOf C,

As mentioned earlier in the chapter-2, the rate of deacetoxylation is given by


r: =k2 3.18

The change in concentrations of DAB ( d ) , DAFB (C 2 ) and FAB (C-,) as a function of time
in a batch reactor is given by the following equations (Liq- 3.19-3.22):

IS1)
Chapter-3

</C
'' r 3.19
<lt ~ '

c/C\
3.20

c/C,
3.21
<//
with initial conditions :
t = 0, Ci = C | 0 , C 2 = 0, C3 = 0 3.22

Here, the rate terms, n and r2 represent the reaction rates for the hydroformylation and
deacetoxylation steps respectively.
In case oi'biphasic hydroformylation of DAB using [Rh(COD)Cl]2 / TPPTS as the catalyst
system, only one product is formed. This is the deacetoxylated product FAB. The rate of
deacetoxylation is very high as compared to the rate of hydroformylation. Secondly, there is
no other product formed other than the deacetoxylated product, FAB. This is also
confirmed by the formation of an equivalent amount of acetic acid in the reaction mixture.
Therefore, the overall rate of the reaction is attributed to the rate of hydroformylation
another rate constant k2 for the deacetoxylation step is not considered. Under these
conditions, the rate equations reduce to the following:

*,*,A:,C,,C,[CO][//,] 3 2 3

(\ + K][CO] + K[K2[CO]Cl +K,K2Ki\CO]2C, + K,KlKlK^\CO'\i C,

Where r, is the rate of hydroformylation reaction.


And the change in concentrations of DAB (C|) and FAB (C2) as a function of time in a
batch reactor is given by the following equations:

3.24
clt

3.25

w ith initial conditions:

1W
Chapter-3

1 - 0 . C, =C,,C:=--() 3.26

Various forms of rate equations were examined to fit the experimental concentration-time
data by solving the above set of model equations in combination with an optimization
programme. The details of the procedure followed for model discrimination, statistical
analysis and parameter estimation was the same as described earlier in Chapter-2. Based on
the statistical criteria, the rate equations derived from the mechanistic model were found to
represent the concentration-time data using a non-linear least square regression analysis,
obtained over a wide range of conditions as shown for in Figures-3.15 (a) and 3.15 (b). The
agreement between the model prediction and the experimental data was found to be
excellent. The rate parameters were calculated for the mechanistic model and the results are
presented in Table-3.8. The activation energy was calculated to be 30.1 kJ/mol.

0.15 0.15

0.10 0.10
o
E

o
' ^

ro
k_

c<D
+
0.05- 0.05
U
c
o
o

0.00 + 0.00
30 60
Time, min
Figure-3.15 (a): Comparison of experimental and predicted Concentration-Time profiles
for 348K

191
Chapter-3

0 . 1 5 -A 1 ' 1 1 ' 1 ' ' ' 1 0.15

0 5 10 15 20 25 30
Time, min

Figure-3.15 (b): Comparison of experimental and predicted Concentration-Time profiles


for 358K

Table-3.8: Rate constants evaluated from the rate model

Temp k, K, K2 K3 K4
3 3
K m kmol/s m / kmol m /kmol m /kmol m7kmol
338 2.82 600.09 4.0 7.0 12.0
348 3.80 600.06 4.0 7.0 12.0
358 5.03 600.87 4.0 7.0 12.0
Chapter-3

3.4. Conclusions

The hydroformylation of DAB using [Rh(COD)Cl]2 / TPPTS complex catalyst in a two


phase system (Toluene and Water) was studied in a batch reactor at 348K and 6.8 MPa syn
gas pressure at s = 0.4. It was observed that the solvents had a significant effect on the
activity and product distribution in the aqueous and the organic phases. Jt was observed that
toluene was the best solvent for DAB biphasic hydroformylation, as it did not show any
rhodium leaching in the aqueous phase and retained the entire product formed in the
organic phase. The leaching in case of hydroformylation of DAB using [Rh(COD)Cl]? /
TPPTS complex catalyst was also dependent on the sequence of addition of the reactants.
The best form of avoiding leaching is by adding DAB to the preformed rhodium aqueous
phase catalyst.

Kinetics of hydroformylation of DAB using [Rh(COD)Cl]2 / TPPTS water soluble


complex was studied in the temperature range of 338-358K. The reaction was found to be
first order with respect to the catalyst concentration and hydrogen partial pressure, while the
reaction was zero order with respect to DAB concentration. The rate Vs. Partial pressure of
CO showed a substrate-inhibited kinetics at higher CO pressures. The rate equations were
proposed and it was observed that the mechanistic model very well represented this kinetic
study. Based on the proposed rate equations, the kinetic parameters were estimated.

193
List of Notations

R\ Observed rate of hydroformylation, kmol/nr 1


C'I The saturation solubility of CO, kmol/nr 1
('/; The saturation solubility of H2, kmol/nr
ki an Gas-liquid mass transfer coefficient, s"
Kn Liquid-liquid mass transfer coefficient
an Liquid-Liquid interfacial area, nT/nT5
Sh Sherwood number
DAB Molecular diffusivity of DAB, m 2 /s
dp Maximum drop diameter, cm
C Constant dependent on the geometry of the vessel and agitator
N Impeller Speed, rev/sec
We Weber number
T Temperature, K
Mn Molecular weight of the solvent B
\ A Solute molar volume at its normal boiling point, nrVkg mol (For further
calculations refer Transport Processes and unit operations bv Christie J.
Geankoplis, third edition, 1997)
C'i Concentration of DAB, kmol/W
C'2 Concentration of DAFB, kmol/m'1
C;, Concentration of FAB, kmol/m"
i'i Reaction rate for the hydroformylation step, kmol/nr/s
i"; Reaction rate for the deacetoxylation step, kmol/nr/s
k| Rate constant for the hydroformylation step
k; Rate constant for the deacetoxylation step
Co Concentration of Catalyst, kmol/m'
C, Concentration of DAB, kmol/m'1
IL Concentration of If, kmol/nr 1
CO Concentration of CO, kmol/nr 1
Y Gas volume, nr
N Agitation Speed, Hz
V| Liquid volume, mJ
d| Impeller diameter, m
dr Tank diameter, m
lii Height of the impeller from the bottom, m
h; Liquid height, m

Greek Symbols

p Density of the continuous phase, k g / m 3

a lnterfacial tension, dynes/cm

E Dispersed phase hold u p

(p Association parameter of the solvent

|.iB viscosity B in Kg/m.s

UI.A Parameter defined by Eq-3.1

CXI.B Parameter defined by Eq-3.2

u; Parameter defined by Eq-3.4


Chapter-3

References:
1 P. W. N. M. van Leeuwen, C. Claver (Eds), Rhodium catalyzed hvdroformvlation
B R . James, P.W.N.M. van Leeuwen (Eds), Catalysis by Metal Complexes Series.
vol. 22, Kluwer Academic Publishers, Dordrecht, 2000 and C. D. Frohnmg, C. W.
Kohlpamtner, H. W. Bohnen, Second Ed., B. Comils, W.A. Herrmann (Eds.),
Applied Homogeneous Catalysts with Organometallic Compounds, vol. 1, Wiley-
VCH. Weinheim, 2002, pp. 31-103 (Chapter 2.2.1)

2 (a) E. Kuntz, FR 2,3,14,910 (1975), US 4,248.802, 1981 (b) E. Kuntz, CHEMTECH


1970,570
3 (a) B. Cormls, E.G. Kuntz, ./ Organomet. Chem. 1995, 502, 177 (b) B. Cornils,
W.A. Herrmann, R.W Eckl, J. Mol. Catal A, 1997, 116, 27
4 B. Cornils, W.A. Herrmann (Eds.), Applied Homogeneous Catalysts with
Organometallic Compounds, Wiley/VCH, Weinheim/New York, 2000
5 (a) E.G. Kuntz, CHEMTECH, 1987, 17, 570 (b) E. Wiebus, B. Cornils,
Hydrocarbon Process. 1996, 66. 63
6 B. Cornils, Angew. Chem. 1995, 34, 1575
7 (a) B. Breit, W. Seiche, Synthesis, 2001, 1 (b) W.A. Herrmann. C.W. Kohlpaintner,
Angew. Chem. Int. Ed. Engl.,1993, 32, 1524
8 (a) S. Bischoff, M Kant, hid. Eng. Chem. Res. 2000, 39, 4908 (b) S. Bischoff, M
Kant, Cata. Today, 2001, 66, 183 (c) M.S. Goedheijt, B.E. Hanson, J.N.H. Reek, P.
C. J. Kamer, P.W.N.M. van Leeuwen, J. Am. Chem. Soc, 2000, 122, 1650

9 P. Kalck, M Dessoudeix, S. Schewarz, J. Mol. Catal. A, 1999, 143,41


10 R. V. Chaudhari, B. M. Bhanage, R. M. Deshpande, H. Delmas, Nature (London)
1995,373,50]
11 J. P. Arhancet, M.E. Davis, J. S. Merola, B. E. Hanson, Nature, 1989, 339, 865 (b)
P. Kalck, L. Miquel, M. Dessoudeix, Catal. Today. 1998, 42, 431
12 (a) R. M. Deshpande, Purwanto, H. Delmas, R. V. Chaudhari, ./. Mol. Catal A,
1997, 126, 133 (b) R. M. Deshpande, P. Purwanto, H. Delmas, R. V. Chaudhari,
Ind.Eng. Chem. Res., 1996, 35, 3927 (c) S Paganeli, M. Zanchet, M . Marchetti, G.
Mangano,,/. Mol. Catal. A, 2000. 157, 1

196
Chapter-3

13 (a) F. Van Vyve, A. Rcnken, Catal. Today, /999, 48, 237 (b) II. Chen. Y.Z. Li, J.R.
Chen. Y.E. He, X.J. Li, J. Mol. Catal. A, 1999, 149, 1 (c) Y.Z. Li, H. Chen, J. R.
Chen. P.M. Cheng. J.Y. Hu, X.J. Li, Chin.J. Chew. 2001, 19, 58 (d) Y. Q. Zhang,
Z.S. Mao, J.Y. Chen, Catal. Today, 2002, 74, 23 (e) Y. Q. Zhang, Z.S. Mao, J.Y.
Chen, /W. /zg. OK</;?. /?<?.V. 2001, 40, 4496, (f) B. Fell, C. Schobben, G.
Papadogianakis, J. Mol. Catal. A 101 (1995) 179; (g) H. Chen, Y. Z. Li. J. R. Chen,
P.M. Cheng, Y. E. He, H.J. Li, J. Mol. Catal. A, 1999, 149, 1381; (h) M. Haumann,
H. Koch, P. Hugo, R. Schomacker, Appl. Catal. A, 2002, 225, 239; (d) Y. Zhang, Z.
Mao, J. Chen, Catal. Today, 2002, 74, 23 (i) H. Chen, Y. Li, J. Chen, P. Cheng, X.
Li, 2002, 74 131; (j) C. Yang, X. Bi, Z. Mao, J. Mol. Catal. A, 2002, 187, 35; (k) M.
Haumann, H. Yildiz, H. Koch, R. Schomacker, Appl. Catal. A, 2002, 236, 173

14 (a) IV. A. Herrmann, J. Kellner, H. RiepI,./. Orgcmomet. Chem. 1990, 389. 103 (b)
W. A. Herrmann, J. A. Kulpe, J. Kellner, H. Riepl, H. Bahrmann, W. Konkol,
Angew. Chem. Int. Ed. Engl., 1990, 32, 391
15 E. G. Kuntz, CHEMTECH, 1987, 17, 570
16 B. M. Bhange, Studies in hydroformylation of olefins using transition metal complex
catalysts. Ph.D. Thesis, University of Pune, 1995
17 T. Bartik, B. Bartik, B. E. Hanson, T. Glass, W. Bebout, Inorg. Chem. 1992, 31,
2667
18 J. Chatt and L. M. Venanzi, J. Chem. Soc. A, 1957, 4735
19 J. L. Brevet, K. Mori, Synthesis, 1992, 1007
20 Y. S. Varshavskiy T. G. Cherkasova, Russian Journal of Inorganic Chemistry,
1967, 72 899
21 (a) 1. T. Harvath. R. V. Kastrup, A. A. Oswald, E, L. Mozeleski, Catalysis Lett.
1989, 2, 85 (b) J. P. Arhancet, M. E. Davies, J. S. Merola and B. E. Hanson, J. of
Catalysis, 1990, 121,321
22 R. M. Deshpande, B. M. Bhanage, S. S. Divekar, S. Kanagasabapathy and R. V.
Chaudhari, bid. Eng. Chem. Res, 1998, 37, 2391
23 P. Purwanto and H. Delmas, Catalysis Today. 1995, 24, 135
24 R. VYiebe, V. L. Gaddy, J. Am. Chem. Soc. 1934, 56, 76

197
C'hapter-3

25 ,|. A. Dean, Table 10-1 in Lunge's Hundhaok of'Chemistry. 13th lid.. Mc. Graw. Hill

Publications. 1972
2d B. M. Bhanage. S. S. Divekar, R. M. Dcshpande. R. \'. Chaudhan, ./. Mol.
Catulvsis-A: Chemical. 1997. 115,241
27 A. Fredenslund, R. L. Jones, J. M. Prausnitz, AlChKJ, 1975, 21, 1086
28 I. Hablot, J. Jenck, G. Casamatta, H. Delmas, Chem. Engg. Sci., 1992, 47, 1267
29 P. Purwanto, R. M. Dcshpande, R. V. Chaudhan, H. Delmas, J. Chem. Ling. Data,
1996.41. 1414
30 E.G. Kuntz, French Patent No. 2,314,910, 1975 (to Rhone Poulenc Industries)
31 (a) R. M. Deshpande, Purwanto, H. Delmas, R. V. Chaudhari, J. Mol. Catalysis-A:
Chemical, 1997, 126, 133 (b) P. Purwanto, H. Delmas, Catalysis Today, 1995, 24,
135
32 (a) W. A. Hermann, C. W. Kohlpaintner, Angew. Chem. Int. Ed., 1993, 32, 1524 (b)
R. M. Deshpande, Purwanto, H. Delmas, R. V. Chaudhari, J. Mol. Catalysis A:
Chemical, 1997, 126, 133
33 (a) B. M. Bhanage, S. S. Divekar, R. M. Deshpande and R. V. Chaudhan, J. Mol.
Catal.A, 1997, 115, 247. (b) R. M. Deshpande, R. V. Chaudhari, Ind. Eng. Chem.
Res. 1998,27, 1996
34 R. V. Chaudhari, A. Bhattacharya, B. M. Bhanage, Catal. Today, 1995, 24, 123
35 P. Claus, M. Baems, Aqueous-Phase Organometallic Catalysis, Cornils B.,
Hermann W. A. (Eds.) Wiley- VCH, New York, 1998
36 R. V. Chaudhan, R. V. Gholap, G. Emig and H. Hoffmann, Can. J. Chem. Eng.
1987, 65,744
37 C. R. Wilke, P. Chang, AlCheE. J, 1955, 1, 264
38 J. S. Lagisetty, P. K. Das, R. Kumar, K. S. Gandhi, Chemical Eng. Science, 1986,
41. 65
39 R. M. Deshpande, R. V. Chaudhan, Ind. Eng. Chem. Res. 1998, 27, 1996
40 R. M. Deshpande and R. V. Chaudhari,./. Mol. Catal.A, 1989, 57, 177
41 R. M. Deshpande and R. V. Chaudhari, J. Catalysis, 1989. 115, 326.
42 Y. S. Nair, S. P. Mathew and R. Y. Chaudhari../. Mol. Catal.A, 1998. 143. 99

o,\
Chapter-3

43 G. Kiss, E.J. Mozeleski, K.C. Nadler. E. Van Driessche and C. De Roover. J. Mol.
Catalysis.4: Chemical, 1999, 138, 155
44 R. M. Deshpande. Purwanto, H. Delmas, and R. V. Chaudhari, hid. Eng. Chem. Res.
1996. J J, 3927
45 W.A. Herrmann. J. Kellner, H. Riepl, J. Organomet. Chem. 1990, 389, 103

199

Vous aimerez peut-être aussi