You are on page 1of 25

~1~ ~ A

CP
AP
TALE
ILYDSS
I
A: GENERAL
ELSEVIER Applied Catalysis A: General 154 (1997) 29-53

Synthesis of bisphenol-A: comparison of efficacy of


ion exchange resin catalysts vis-a-vis heteropolyacid
supported on clay and kinetic modelling
G.D. Yadav*, N. K i r t h i v a s a n
Chemical Engineering Division, University Department of Chemical Technology (UDCT), University of
Mumbai, Matunga. Mumbai 400 019, India

Received 27 March 1996; received in revised form 23 September 1996; accepted 4 October 1996

Abstract

The production of bisphenol-A from phenol and acetone over ion exchange resins has been
practised industrially, with the selectivity to the product largely depending on the type of catalyst
and reaction conditions. This study reports the use of a novel catalyst based on heteropolyacid
supported on clay, particularly dodecatungstophosphoric acid (DTP) supported on K-10 clay and its
comparison with commercially available resins such as Amberlyst-15, Amberlyst-31 and
Amberlyst-XE-717p. DTP/K-10 is an efficient and re-usable catalyst which could be employed
at higher temperatures. This catalyst is characterised by XRD, BET, SEM, FTIR and the cation
exchange capacity (CEC) has been determined. The kinetics of the reaction with DTP/K-10 have
shown interesting features among which the formation of intermediate isopropenyl phenol was
found to be the rate determining step with the Eley-Rideal type of mechanism.

Keywords: Bisphenol-A; Ion-exchange resins; Heteropolyacids

1. Introduction

Bisphenol-A is a very important raw material for the synthesis of epoxy


resins and other polymers. Currently two grades of bisphenol-A are manufactured;
one is the epoxy grade with a purity of 95%, and the other is the polycarbonate
grade with a purity >99%. Conventionally bisphenol-A is manufactured by the acid
catalysed condensation of phenol and acetone. Ion exchange resins are the
preferred catalysts, but these have temperature limitations. High molar proportions

* Corresponding author. E-mail gdy@udct.ernet.in; Fax: (+91-22) 4145614.

0926-860X/97/$17.00 ~()
:'" 1997 Elsevier Science B.V. All rights reserved.
PII S0926-860X(96)00364-X
30 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

of phenol to acetone are required for a facile progress of the reaction as well as
to suppress the side reactions of acetone. A large number of side-products are
generated depending upon reaction conditions and the type of catalyst. An ideal
catalyst for this reaction should be moderately acidic and shape selective, whereby
lesser quantities of the by-products would be formed. Heteropoly acids are an
interesting class of catalysts which are being used in a number of reactions [1,2].
Recently we reported a novel catalyst based on heteropoly acid supported on clay,
which was found to be superior to several others including Amberlyst-15, for the
synthesis of methyl tert-butyl ether (MTBE) from tert-butanol and methanol [3]. It
was thought worthwhile to test its potential in the synthesis of bisphenol-A from
phenol and acetone. Therefore, a systematic investigation was undertaken to
compare the efficacy of dodecatungstophosphoric acid (DTP) supported on K-
10 clay vis-~t-vis Amberlyst-15, Amberlyst-31 and Amberlyst-XE-717p, including
its characterisation and the kinetics of the reaction, which are reported in this paper.

2. Literature review

Due to its commercial significance, considerable information on bisphenol-A


is available in the patent literature. The use of divalent sulphur compounds such
as mercaptans and glycolic acids to increase the reaction rate is also indicated.
Sulphonated polystyrene-divinylbenzene ion-exchange resins with a portion of
the sulphonic acid groups converted into mercaptan functionality were found
better catalysts than unmodified resin [4,5]. The use of zeolites coated with
mercaptoamine at 120-180°C has been reported [6]. Singh [7] has discussed in
detail the synthesis of bisphenol-A over zeolite catalysts such as H-ZSM-5,
H-mordenite, H-Y and RE-Y vis-a-vis Amberlyst-15 and shown that zeolites
with larger openings are more selective for this process although ion exchange
resins are more active than zeolites. However, the general trend shows that
modified ion exchangers are the catalysts used worldwide for an optimum yield
of bisphenol-A. Alkylation of propenyl halide with phenol using Friedel
Crafts catalysts for the synthesis of bisphenol-A has been reported [8]. Conversions
to the extent of 60% are obtained in the above mentioned process. Some
monographs mention the use of commercial acid treated clays for the synthesis
of bisphenol-A [9,10].

3. Reaction chemistry

The acid catalysed condensation of phenol and acetone in homogeneous medium


leads to several products; in particular with strong sulphuric and hydrochloric acids
almost 28 products have been identified. Strong acids such as 70% sulphuric acid,
hydrochloric acid or sulphonated polystyrene divinylbenzene cation exchange
G.D. Yadav, N. Kirthivasan/Applied Catalysis A." General 154 (1997) 29-53 31

Desbtd Reaction

OH

2 + H3C-- C--
o
II
CH 3

.o©_?@_o. CH 3
÷ ti20

Pl~noi Acetone
Bisphenol-A

Res~tJon MeeJMnbm

O OH
II ~ I (a)
H3C--C--CH 3 ~ H3C--~)--CH3

(h)
I + I _- HO-
CH3 CH3 ~ H

C,H3 _ _
H "+ - ® + FI20 (c)
\. CH 3

C,H3 CH 3

(d)

CH3 ~ H CH3

Protonated IsOl~p~yl plmlol Bi~heml- A

Fig. 1. Reaction scheme and mechanism for homogenous acid catalysis.

resins are industrially preferred. Fig. 1 shows the reaction scheme. In a strong acid
medium, acetone is protonated to a stable carbenium ion as shown by reaction a in
Fig. 1. In the following steps, the carbenium ion adds to the limiting quinonoid
structure of phenol to yield a protonated carbinol (reaction b). The carbinol
rearranges to release water and yields protonated isopropenyl phenol (reaction
c). The isopropenylphenol adds to a second phenol molecule to yield bisphenol-A
(reaction d). Several side products are formed as shown in Fig. 2, the major one
32 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

CH3
I
O=C
OH O I
I II H+ C//CH H20 Col
H3C--.~--CH3 + H3C--C--CH 3 /\
H3C CH3
Mesityl Oxide
HO
O OH ~H 3

+ H 3 C - - C - - CH3 ~ --OH H2() Ifl

Acetone CH3
Phenol
o,p'-Binphenol-A

CH3
HO I CH3
c.zCH H+ OH 1t20 (g)
/\
H3C CH3 H3C CH3
Phenol Mesityl Oxide
1,1,3-1rim~yl,3,4L hydmxy phenyl, 5-irdanol

CH3 O CH3
OH O=C
H+ L ~ ~ CH3 + H20 (hi
+ C~CH
' H3 C
/\
H3C CH3

Phenol Mesityi Oxide OH


2,2,4-tria3ethy~
4-(41-hydroxypheny~el~ornan

OH CH3
o=c H+ o

+ c~CH = ~ C H 3 + H20 (i)


/\
Phenol H3C CH3 CH3
Mesityl Oxide 2,2,4- ixi~Ethylchromen
Fig. 2. By products of bisphenol-A synthesis.

being the o,p 1 isomer formed by the reaction of p-isopropenyl-phenol with the
phenol in the ortho position (reaction f). Other side products are the chroman
derivatives formed by the reaction of phenol with mesityl oxide, which itself is a
product of self-condensation of acetone followed by dehydration (reaction e,
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 33

Fig. 2). Out of the 28 products mentioned, those which are formed in relatively
larger quantities are 2,2,4-trimethyl chromen, 1,1,3-trimethyl-5-indanol, 9,9-
dimethylxanthane and dimethyl hydroxy biphenyl. The remaining by products
constitute to less than 0.2% of the reaction mixture. It is obvious from Fig. 2 that
the formation of byproducts can be suppressed if the self- condensation of acetone
followed by dehydration leading to mesityl oxide is avoided. One of the ways
would be to employ excess of phenol over acetone in batch experiments or semi-
batch mode of operation with continuous addition of acetone making the phenol to
acetone ratio very high.

4. Experimental

4.1. Chemicals and catalysts

The preparation of alumina, zirconia, chromia exchanged clays and heteropoly


acid supported on clays has been detailed elsewhere [3,11]. Amberlyst-15,
Amberlyst-31 and Amberlyst XE-717p were procured from Rohm and Haas
(USA). K-10 and Filtrol-24 were purchased from Fluka (Switzerland) and
Engelhardt (Germany), respectively.
Phenol and acetone were obtained from M/s s.d. Fine Chemicals Pvt. Ltd.,
Mumbai.

4.2. Preparation of dodecatungstophosphoric acid (dtp) supported


on clay (DTP/K-IO)

Heteropoly acids (HPA) have been supported on silica, titania, alumina, etc. and
the preparatory procedures are well standardised [ 12]. We reported for the first time
the use of a novel catalyst heteropoly acid supported on clay [3], the preparation of
which is given below.
Approximately 10 g of K-10 clay was dried in an oven at 120°C for 1 h of which
8 g were weighed for subsequent experiment. 2 g of dry DTP was also weighed.
The HPA was dissolved in 8 ml of dry methanol. This volume of solvent was
approximately equal to the pore volume of the catalyst. The solution was added in
aliquots of 1 ml each to the clay under constant stirring with a glass rod or kneading
it properly. The solution was added at time intervals of 30 sec. Initially and up to
the addition of 6 ml of the DTP solution, the clay was in the powdery form but upon
subsequent addition the clay formed a paste. Further kneading of the paste for
10 min. yielded a dry free flowing powder. The preformed catalyst was dried in an
oven at 120 °C for 1 h and then calcined at 275°C for 3 h.
Two factors were found to be important in the preparation of supported
heteropoly acids: calcination temperature and heteropoly acid loading on the
clay, the details of which we have discussed earlier [3].
34 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

5. Characterisation of catalysts

5.1. X-ray diffraction studies

The X-ray diffractograms were recorded with a Rigaku automated diffract-


ometer. The X-ray target was the Cu K s (1=1.5402 ,~). The X-ray pattern of K-10
is given in Fig. 3a. It shows an ill-defined peak at 5 ° correspondin~ to a d-spacing
of 17 ,~. This diffraction band extends from about 10 A to 17-19 Awith a plateau
inclined towards 17 ,~ which has been discussed by Laszlo [9]. The exact reason for
such a peak is not known though it could be speculated that it was caused due to the
presence of lattice debris produced during the acid treatment of montmorillonite.
The d(001) spacing is attributed to the sharp peak observed at 20=9 °. The peak at
20=18 ° is the d(003) spacing as suggested by Slaughter and Milne [13]. There are
several other sharp peaks which are indicative that K-10 is still crystalline. The
peak at 20=27 ° is assigned to c~-quartz.
In the diffractogram of DTP/K-10 the peaks are less intense indicating that in the
impregnation process, the clay has lost some of its crystallinity compared to K-10.
The intensity of the d(001) peak at 20=10 ° is also subdued. Most of the
diffractogram for DTP/K-10 is similar to that of K-10. (Fig. 3b)

5.2. Surface area measurement

The surface areas of the K-10 and DTP/K-10 were measured by the BET method
using the nitrogen adsorption technique on an indigenously made apparatus and
were found to be 230 m2/g and 107 m2/g, respectively. The catalysts were pre-
dried at 275°C overnight prior to the analysis.
The surface area of DTP/K-10 was found to be 50% lower than that of K-10.
Since both the catalysts are pre-treated in a similar way prior to analysis the
reduction in surface area in the case of DTP/K-10 could primarily be due to the

,~0
' 35' 3b 25' 20' 1 1'0
20

Fig. 3. X-ray diffractograms (a)K-10 (b)DTP/K-10.


G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 35

blockage of the smaller pores by the HPA crystals. It appears that the active species
are held in a few junctions of such dimensions from where the access to smaller
pores is denied, thereby leading to reduction in accessible surface area. The
particle size of K-10 and DTP/K-10 was determined by image analysis (Tracor
Northern TN 8500) as 80-100 lam.

5.3. Scanning electron microscopy

The SEM was performed on a CAMECA (France) instrument. Fig. 4(a) and (b)
show the micrographs of K-10 and DTP/K-10, respectively. A distinct difference
between the two is found. In the case of the DTP/K-10 there are a lot of dispersed
particles on the surface of the support K-10.

5.4. Cation exchange capacities (CEC)

The CEC of the catalysts were determined photometrically by methylene blue


adsorption, as described by Robertson and Ward [14] and were found to be 35 and
39 meq/100 g for K-10 and DTP/K-10, respectively.
The CEC of K-10 closely agrees with the values given by Clark et al. [15]. The
CEC of DTP/K-10 exhibits a slightly higher value than K-10. It could be due to
some additional surface protons which come by way of heteropoly acid impreg-
nation which may play a role on account of easy availability for the exchange
reaction.

5.5. Fourier transform infra-red spectra (FTIR)

FTIR studies of the catalysts were conducted by using a Bruker IFS-66 single
channel Fourier transform spectrophotometer. Thin wafers were prepared by
mixing 3 mg of the catalysts with 50 mg of KBr. The wafers were subjected to
200 scans after which the spectra were recorded.
Fig. 5(a) shows the FTIR scan of K-10. Very prominent bands are observed in
the region between 600 and 3000 cm -1. The broad band stretching from 900-
1200 cm -1 is due to the Si-O stretching vibration and the shoulder at 935 cm -1 is
indicative o f - ~ ) H groups bonded to the A1 atoms. Bands in the range 533-
807 cm -1 are assigned to Si-O and Si-O-A1 vibrations. The band at 1642 cm -~ is
attributed to - O H bending frequency of water molecules and the band at
3431.6 cm-~ is attributed to the structural hydroxy stretching vibration.
Fig. 5 (b) shows the scan of DTP. The scan matches very well with that
of Kozhevnikov et al. [16] and Sharpless and Munday [17]. The following
prominent bands are observed. 1087.8cm I (P-O), 989.5cm -l (W=O),
898.3 cm -1, stretching vibration of W - O - W bridges between comer sharing
octahedra and 800 cm 1 stretching vibrations of W - O - W bridges between edge
sharing octahedra [ 18].
36 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

a) K-IO

b) DTP/K-IO
Fig. 4. Scanning electron micrographs.

The scan of DTP/K-10, (Fig. 5c) is similar to that of K-10 except that there is a
prominent peak at 821.1 cm -1 which is indicative of W - O - W vibrations. How-
ever, the shift observed for this peak when compared to pure DTP could be
attributed to the interaction between DTP and K-10.

5. 6. Stability of DTP on clay

It was necessary to study the stability of DTP on K-10 in order to reuse the
catalyst. Any leaching of the active sites from the catalyst would render it
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 37

' "7

e-

I--

• I
0'~ o

t
r::

03

,<5
I I I I
4000 3000 2000 1000
Wavenumbers {cm-1 )
Fig. 5. FFIR Spectra (a)K-10 (b)DTP (c)DTP/K-10.

commercially unattractive. Our reaction studies for the synthesis of MTBE from t-
butanol and methanol showed that the DTP/K-10 showed constant activity for a
minimum of four runs even in the presence of a polar solvent such as methanol and
also 1,4 dioxane [3]. It was therefore decided to verify the possibility of leaching of
DTP from the support under severe conditions.
38 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

HPA could be analysed by the heteropoly blue colour which is observed when it
is made to react with a mild reducing agent such as ascorbic acid [19]. This
property was utilised for the quantitative determination of leaching, if any.
Standard samples amounting to 1, 2, 3, 4 and 5% of DTP in water were prepared.
To 10 ml of the above samples 1 ml of 10% ascorbic acid was added. The mixture
was diluted to 25 ml. The absorbance of the resulting solution was measured at
Amax of 785 c m - 1 . A standard calibration curve was obtained by plotting the
absorbance with of DTP in aqueous solution.
10 g of DTP/K-10 was taken in 25 ml of methanol and the mixture was refluxed.
1 ml aliquot of the refluxing solution was withdrawn to which 1 ml of 10% ascorbic
acid solution was added. The solution was diluted to 25 ml and the absorbance was
measured at the Area x value of 785 cm -1.
It was observed that up to a period of 3 h there was a loss of 3% of the loaded
DTP. The above tests were repeated to find out if the leaching was indeed
negligible. Leaching would have taken place if the heteropoly acid were loosely
bound to the clay matrix. Additionally, the heat treatment does not modify the
characteristics of DTP. The absence of coloration could be attributed only to the
absence of any free HPA in the solution. It was therefore concluded that the DTP
was chemically adsorbed on the catalyst surface and no leaching was observed. In
the presence of acetone and phenol thus no leaching was expected in the presence
of acetone and phenol.

6. Experimental

6.1. Reaction procedure

All the experiments were carried out in a 100 ml Parr autoclave equipped with a
four bladed turbine impeller. The temperature was maintained within i 0 . 5 ° C of
the desired value. The vessel was also equipped with a speed regulator that could
maintain the desired speed at + 1% of the set value.
Predetermined quantities of reactants and the catalyst were charged into the
autoclave and the temperature was raised to the desired value. Once the tempera-
ture was attained the initial sample was withdrawn at time=0 and the stirrer was
started. Further samples were withdrawn at definite time intervals.
A typical experiment consisted of 5.28 g (0.091 gmol) of acetone, 42.77 g
(0.451 gmol) of phenol, 1.25 g of catalyst (loading of catalyst, 0.0268 g/cm 3 of
the liquid phase). The reaction temperature was maintained at 100°C and 135°C for
the ion exchange resins and inorganic catalysts, respectively.

6.2. Analysis

The samples were analysed in a gas chromatograph (Perkin-Elmer Model 8500)


equipped with a flame ionisation detector. A 2 × 0.003 m column packed with 10%
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 39

OV-17 supported on chromosorb WHP was used. Calibration curves were


prepared by using standard samples and synthetic mixtures in order to quantify
the data.

7. Results and discussion

7.1. Comparison of the activity of the catalysts

A maximum of six products were detected in the reaction mixture. The products
were identified as 1,1,3-trimethyl-5-indenol, 2,2,4-trimethyl chromen, 9,9-
dimethyl xanthane, dimethyl hydroxy biphenyl, o,p-l-bisphenol and bisphenol-
A. In the case of Amberlyst-31 and Ambedyst-XE-717p only o,p-l-bisphenol and
bisphenol-A were formed, whereas in the case of Amberlyst-15, all the by products
were formed in large quantities resulting in a lower selectivity to bisphenol-A.
Since it was desired to test the efficacy of dodecatungstophosphoric acid
supported on K-10, a few experiments were also conducted with the support clay
K-10 as well as Filtrol-24 as shown in Table 1. Both K-10 and Filtrol-24 were
rather ineffective. As was expected, Lewis acid type catalysts, such as alumina
exchanged K-10, zirconia exchanged K-10, chromia exchanged K-10 and
sulphated zirconia calcined at 650°C did not show any conversion at 135°C after
4 h of reaction The Bronsted acid type catalyst were more effective.
In the absence of diffusional effects the performance of the various catalysts is
commented upon in what follows. K-10, an acid treated clay, is obtained by the
acid treatment of the Bavarian montmorillonite Tonsil-13. This catalyst has some
amount of residual acidity which may catalyse many reactions. Exchanged clays
are modified forms of the precursor K-10. During the preparation of exchanged
clays, the precursor K-10 was washed free of any residual acidity making it less

Table 1
Activity of different catalysts for the synthesis of bisphenol-A

Catalyst Conversion (%) Selectivity (%)

K-10 2 0
Filtrol-24 7 3
Alumina Ex.K-10 0 0
Zirconia Ex.K-10 0 0
Chromia Ex.K- 10 0 0
Amberlyst- 15 10 30
DTP/K- 10 23 60
Amberlyst-31 24 82
Amberlyst XE-717p 60 90
TGA/DTP/K- 10 13 83

Acetone: 0.091 gmol, phenol: 0.451 gmol, catalyst: 1.25g, temperature: 135°C, Mole ratio acet-
one:PhOH=l : 5. In the case of ion exchangers the temperature was 100°C, time: 4 h.
40 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

acidic [ 11 ]. As a result the extent of activity observed over K-10 was not noticed in
the exchanged clays. DTP/K-10 has a higher amount of Bronsted acidity by way of
the additional acidity conferred by the supported heteropoly acid. This catalyst is
highly active in a number of reactions which we have studied including the one
under current investigation. Amberlyst-15 is a sulphonated polystyrene divinyl-
benzene having a CEC of 400 me/ql00 gm of resin [20]. Literature mentions that
ion-exchangers promoted with mercaptans have enhanced selectivity to bisphenol-
A. Amberlyst-31 is a polystyrene cross-linked with divinylbenzene where the
extent of cross-linking is 4%. This relatively low cross-linking enables the ion-
exchanger to swell to a larger extent in the reaction medium allowing larger
molecules to enter the pore network. This catalyst is specifically used for sucrose,
lactose and bisphenol-A synthesis and has a greater CEC (490 meq / 100 g resin)
than Amberlyst- 15 [20].
Amberlyst-XE-717p is a promoted ion-exchanger where 17% of the acid sites
are promoted with an undisclosed molecule [21]. The activity of the catalyst and
selectivity for bisphenol-A were very high. The exchange with any sulphur
compound reduces the acidity of the catalyst, whereby the yield of bisphenol-
A is increased. The main side reactions were of acetone, particularly the formation
of mesitylene and its subsequent reactions. The mesitylene reactions are promoted
by strong acidic sites. The sulphur compounds selectively poison the strong acidic
sites which are responsible for side reactions. The activity and selectivity of
Amberlyst-15 are much lower than over DTP/K-10 and Amberlyst-31, both of
which having similar activities; but the latter being more selective. Amberlyst XE-
717p is more active and selective than both DTP/K-10 and Amberlyst-31.
However, Amberlyst-31 and Amberlyst-XE-717p, like any other ion exchanger
have very poor thermal stability. They can be used at a maximum temperature of
120°C. By contrast, the clay modified catalysts can be used at temperatures as high
as 300°C. The lower selectivity of bisphenol-A with DTP/K-10 could be due to its
high acidity giving rise to a variety of by-products. If the acidity of the catalyst is
controlled by reducing the activity of some of the sites; for instance, by doping with
sulphur compounds the formation of byproducts could be lowered. Further, some
reactions can occur on the external surface of the catalysts without any shape
selectivity.
It was thought desirable to study the kinetics of this reaction with DTP/K-10 as
the catalyst. The effect of various parameters was studied under otherwise similar
conditions at a standard temperature of 135°C.

7.2. Effect of speed of agitation

Fig. 6 shows the conversion of acetone at different time intervals. The conver-
sions were found to remain practically the same at speeds beyond 1000 rpm
thereby indicating absence of solid-liquid mass transfer resistance. Further reac-
tions were conducted at a speed of 1000 rpm. Since acetone was taken as a limiting
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 41

30

25

20-

1oz

O i i i i i i i l ~ l l l l l J l l l l ~ l l T 1 1 1 1 1 1 1 1 ~ l l q l , I
O 2 4 6 B 10 12 14 16
T i m e x 10 -a, s

Fig. 6. Effect of speed of agitation. Temperature:135"C, Catalyst loading:5.36x10 -2 g/cm 3, Mole ratio of
PhOH to acetone=5 : 1, Cat.:20% DTP/K-10.(O) 1000 rpm and beyond,(*) 700 rpm, (Z~) 500 rpm.

reactant it could be concluded that any external resistance to its transfer from the
bulk liquid phase to the external surface of the catalyst was absent.

7.3. Effect of catalyst loading

Fig. 7 shows the plot of initial rate of reaction of acetone ( r o i , gmol/cm3/s)


against catalyst loading (w, g/cm3). It indicates that the rate of reaction increases
linearly up to a loading of 2.6z 10 -2 g/cm 3 and thereafter remains constant even
though the loading is almost doubled. Since the number of acidic sites available in
the reaction medium is proportional to the available intra-particle surface, the
initial rate of reaction of acetone should be directly proportional to catalyst loading
(mass/volume) if there are no intra-particle diffusion limitations.
This indicates the following:
• An intra-particle diffusion limitation was set in for the transfer of acetone from
the exterior surface of catalyst beyond a solid loading of 2.6x l0 2 g/cm 3,
acetone being the limiting reactant (CAo << Cpo).
• The side reactions were also significant due to extra available active sites.
It clearly demonstrated that mass transfer limitations were set in and not all
internal surface area was utilised for the reaction. In fact, the acetone concentration
would become zero at a certain distance from the centre of the particle and the
reaction would become intra-particle mass transfer controlled. Further experi-
ments were therefore done at catalyst a loading of 2.6 x l0 -2 g/cm 3 in the absence
of intra particle resistance.
42 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

(>-

c)

c)

0 iJiir~ll r~,l,,,ir iil,,,,ll ,,ir,llll ,rll,,~ll I i,,,,,i,,


0 1 2 3 4 5 6
w x 10 ~, g / c m s

Fig. 7. Plot of initial rate of reaction (roi) against catalyst loading (w). Temperature: 135°C, Speed of agitation:
1000 rpm, mole ratio of PhOH to acetone=5 : 1.

7.4. Effect of temperature

Fig. 8 shows plots of conversion of acetone with time at temperatures of 120,


135 and 150°C. The rate of reaction increases as the temperature increases. The
conversions at 120°C are linear with respect to time showing zero-order depen-
dence on acetone concentration. The other two lines indicate non-zero order
dependence which will be discussed later in the kinetic interpretation.

7.5. Effect of mole ratio

In the industrial processes for the manufacture of bisphenol-A, phenol to acetone


molar ratio of 10 : 1 is normally maintained because at higher molar ratios the
formation of by-products is lowered. All the standard runs mentioned above were
carried out with phenol to acetone molar ratios of 5 : 1. In addition, a mole ratio of
3 : 1 was also studied. Fig. 9 shows the conversion data with time for the different
mole ratios studied. The conversions at a mole ratio of 10 : 1 are marginally better
than those at a ratio of 5 : 1.

7.6. Effect of addition of thioglycolic acid to the reaction medium

The use of thioglycolic acid as a promoter in homogeneous catalysis has been


reported in literature [22,23]. The role of the promoter is to form a more stable
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 43

40

. .-- f ~
30
i
/-
/ I
/
t~
~ 20- - 4t"

,/
• /"
10- /

/
/ •
0 I,-,, ...... ~ ......... i ......... , . . . . . . . . .

@ 4 8 12 16
T i m e x 10 -s, s

Fig. 8. Effect of temperature. Catalyst loading: 2.6x 10 -2 gmol/cm3, speed of agitation: 1000 rpm, catalyst:
20% DTP/K-10, mole ratio of PhOH to acetone=5 : 1. (A) 150°C, ('~&-)135°C, (,) 120°C.

40

30 Z
g -

20

10.

~ ( I I I I I I I I [ I I I I I I I I I [ I I I I I I I I I I I ~ [ ! [ I

0 4 8 12 16,
Time x 1 0 -S, s

Fig. 9. Effect of mole ratio. Catalyst loading:2.6xl0-2,g/cm 3, Temperature:135°C, Speed of agitation:


1000 rpm, Cat.: 20% DTP/K-10. Mole ratio of PhOH:acetone.((~) 10: 1,(+) 5 : 1,(*) 3 : 1.

ionic intermediate with acetone, namely,


SCH2COOH OH
H3C--C--CH 3 I H3C_6_CH 3
+ instead o f +
44 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

100 1 ~0

80
F
~e
-6~
o

-4~ -~

20- 20
+ x ~ x .............

. l l , l ~ l l l l , i r ~ l l , l l i r , l l l l l , , l l l r l , l ~ ~
0 4 8 12 16
Time x lO -s, s

Fig. 10. Effect of addition of thio-glycolic acid(TGA) on the catalyst 50% w/w DTP/K-10. Speed of agitation:
1000 rpm, Temperature: 135°C.
Cat. loading w TGA
g/cm 3
((>) 2.6-10 -2 50% by moles
(+) 2.6.10 2 50% by moles
(×) 1.3.10 -2 50% by moles

instead of which is normally formed with a proton in homogeneous catalysis. The


idea of using thioglycolic acid was to enhance the selectivity and to suppress the
side reactions of acetone. It was desirable to confirm whether the above complex
was formed or not. Therefore thioglycolic acid was added in two different modes.
In the first case, thioglycolic acid was added homogeneously to the reaction
mixture when no reaction was found to occur and in the second case, a known
equivalence of thioglycolic acid amounting to 50% by mole of the heteropoly acid
was loaded on the catalyst surface by the incipient wetness impregnation (Table 1).
The purpose of loading thioglycolic acid was to subdue the activity of the catalyst
by poisoning some of the more acid sites. Thus the selectivity to the desired
product could be greatly enhanced by controlling the acidity. It was found that at a
suitable loading of thioglycolic acid the selectivity to bisphenol-A increased but
the loadings above a critical point made the catalyst totally inactive. Fig. 10 shows
the plot of conversions with time when thioglycolic acid was loaded on the catalyst.

8. Reaction mechanism and kinetics

8.1. Development of mechanistic model


The kinetics of the reaction of phenol and acetone in heptane as solvent
and in the presence of sulphonated poly(styrene-divinyl benzene) to give
G.D. Yadav, iV. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 45

bisphenol-A and water has been measured and represented by the following
equation [23,24].

kKACAK2C 2
rate =
(1 + KACA + KwCw)(1 + KpCp -}- KHCH) 2

where A is acetone, P is phenol, W is water and H is the solvent, n-heptane, which


acts as an inhibitor, K's are the respective adsorption equilibrium constants and k is
the rate constant. The above equation also suggests that there are two types of sites
on the polymer surface whereupon acetone and water compete for one type, and
phenol and heptane for another. Acetone and water are more polar and they bond to
-SO3H groups, whereas heptane and the aromatic ring of phenol have affinity for
the hydrocarbon backbone and aromatic rings of the polymer.
There could be several possible reaction products, but those in Fig. 2 were
detected at the end of the reaction in the current work.
The initial rate data can be analysed on the basis of Langmuir-Hinshelwood-
Hougen-Watson (LHHW) or Eley-Rideal mechanisms. Since the selec-
tivities were higher over for DTP/K-10, Amberlyst-31 and Amberlyst XE-
717p, a simple model is presented below. Due to the presence of heptane, Renicker
and Gates [24] assumed two types of sites, which are likely to be present in a
polymeric support. This situation may not exist if no solvent is used or when
phenol reacts from solution with adsorbed acetone in DTP/K-10 which is an
inorganic catalyst.
In the case of the present studies, the kinetics of the reaction is established for the
DTP/K-10 catalyst assuming the same type of sites (S) as shown in Fig. 11. The
conversion data points were so selected that the formation of by-products was
minimum and the integral form of the rate equation was used. For initial rate data,
the following analysis is the most appropriate.
Adsorption of acetone (A) on a vacant site S is given by:
/CA
A+S ~ AS. (1)

Adsorption of phenol (P) is represented by:


Kp
P+S ~ PS. (2)

But adsorbed phenol does not react. It only poisons the sites. It is the phenol
from the liquid phase that reacts with AS, according to the Eley-Rideal
mechanism.
Reaction of AS with P from the liquid phase in the pore space gives an
intermediate:

AS+P ~ I~S. (3)


46 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

i ~;i LHI i E ~ -~ ':"


~~-(X)- ........... --=:.....-'\J-- c~,
CH3--C--CH3-- ,CH3--~ACH3 ~ ~"~""]

2 ....
/ ~__~____
......:: ~ / ' ~
CH3 ~ (P)
~o-c---=/
I+
~=o.
+

. . . . . . . . n

(IiS) OH

oq
(a)

HO--CI OH H + ~ H3
CH3 ~ ~ I O--C /f~--OH
L_____/

(I2S)
Fig. 11. Proposedmechanismfor the formationof bisphenol-Aon DTP/K-10.

The intermediate I1S desorbs to form 11 which again gets chemisorbed to form
species I2S as given in Fig. 11:
1/Kn
I1S ~ I1+S. (4)

gi2

I1 + S ~-- 12 + S. (5)

Although the above two steps could be combined in one equation with an
equilibrium constant of KI(= Kn/KI2), they are shown separately to represent the
mechanism in Fig. 11.
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 47

Now the second molecule of phenol reacts with I2S, also according to the Eley-
Rideal mechanism:
KR2
I2S q- P ~- BS + W. (6)

The desorption of bisphenol-A from the site:


1/KB
BS ~ B+S. (7)

Poisoning of sites by water is given by:


Kw
W+S ~ WS.

If it is assumed that the intermediates 11 and 12 are too weakly adsorbed, then
total site balance is given by:

Ct = Cs + CAS + Ces + CBS + Cws. (9)


or

Ct = Cs -~ KACACs + gpCpCs + KBCBC S -+- KwCwCs. (lO)


The concentration of vacant sites Cs is therefore:
Ct
Cs = (11)
(1 + KACA q- KpCp -[- KBC B Jr- KwCw) '

8.2. Controlling mechanism

Two important cases of the controlling mechanism are considered here


Case I: If the overall rate of reaction (ro) is controlled by Eq. (3) i.e. the reaction
between chemisorbed acetone (AS) and phenol (P) from the liquid phase,
according to the Eley-Rideal mechanism then:
-dCe
ro = rp -- d~ -- kR~CAsCp (12)
kRIKACACsCp
kR,KACACpCt
(1 + KACA -[- KpCp + KBCB -[- K w C w ) "

Case H: On the contrary if Eq. (6) is the rate controlling one, i.e. the reaction of
chemisorbed p-isopropenyl phenol (I2S) with phenol (P) from the liquid phase
48 G.D. Yadav,N. Kirthivasan/AppliedCatalysisA: General 154 (1997) 29-53

then, overall the rate of reaction is given by:


-dCp
ro = rp -- d~- -- kR2CIESCp (13)

-~ kR2KiCltsCp
= kR2KIkR, CAs (Cp) 2
kR2kRl K1KACA (Cp) 2Ct
(1 + KACA + KpCp + KBC. + KwCw)
There are several subcases possible depending on the relative rates of adsorption
of the various species. In the case of DTP/K-10, the following significant cases are
presented.

8.3. Adsorption of acetone alone is significant

This situation exists if


(1 + K A C A ) >> ( K p C p -~- KBCB "~ K w C w ) .

Since no poisoning of the catalyst was observed by phenol, water or bisphenol-A


during preliminary experiments, this assumption appeared to be valid.
Case I: If Eq. (3) is the rate determining step, then Eq. (12) becomes:
-dCp kR, KACACpCt
ro ---- rp - - - - -- (14)
dt (1 + KACA) "
The total sites concentration Ct can be written in terms of the catalyst loading w
(in mass/volume) because the number of sites are proportional to w.
Thus Eq. (14) can be rewritten as
-dCp k'RIwKACACp
ro = rp -- - - -- (15)
dt (1 +KACA) '
where UR1 incorporates the proportionality constant of Ct o~ w.
If the initial rates (roi) are measured, then Eq. (15) can be transformed into a very
convenient form as given below:
Cpo. w 1 1
-~ (16)
roi ' KA CAo
kRl k'R I "

A plot of t(Cpo, w ) / roi] against 1/CAo yields a slope of (k'R1 KA)-I and an intercept
of (MR1)- from which both the reaction rate constant Um and adsorption equili-
brium constant KA could be established.
Case II: On the contrary if Eq. (6) is the rate determining step, then
i ! 2
kR2kRIwKIKA C A Cp (17)
ro = rp = (I + KACA)
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 49

An equation of the same form as Eq. (16) can be written as follows:


C2o • w 1 1
(18)
roi kKCA kK~ '
where k--kRzk~Rl and K = KAK~. Further treatment to calculate k and K is
analogous.

8.4. Establishment of reaction kinetics

Eq. (17) suggests that the rate of reaction is a strong function of phenol
concentration and a slight increase in it over that of acetone, under otherwise
similar conditions, will lead to substantial increase in the rate of reaction. However,
the effect of mole ratio represented by Fig. 9 suggests that the rates do not increase
as greatly as expected.
A plot of (Cpo'W/ro) against (1/CAo) as suggested by Eq. (16) is shown in Fig. 12.
It is seen that the equation fits the data well thereby proving that step 3 i.e., the
reaction between chemisorbed acetone (AS) and phenol (P) from the liquid phase
is the rate determining step and the overall rate of reaction is given by Eq. (15).
From the intercept k~l is obtained as 0.18 (cm3/g s), and subsequently from the
slope/CA is obtained as 119.5 (cm3/gmol)
Since the system was found to obey the Eley-Rideal type of mechanism, it was
thought worthwhile to study the effect of temperature on the reaction rate and
selectivity. It is expected that/CA, the adsorption equilibrium constant will be lower

¢9

7"2
o

c.) :/

~ .*lJ~llhlltJJI,ihllrJJJidlrlJlllldl~t~Jiiih~llllrl~JInlEiJJl~ll~lldlJiiJiII dl~Jllll,i
1 2[ :3 4 5 6 7 B 9 10
(1/C,o)xlO -2, (gruel/eroS) -1

Fig. 12. Plot of Cpo.W/ro vs. 1/CAo. Catalyst loading: 2.68× 10 3 g/cm 2. Temperature: 135°C, speed: 1000 rpm.
50 G,D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

0.25

0.20

//
0.15

~'~0.10

0.05

0.00 .... E .... .... 4 . . . . . . .5. . 4 .... 7 E3


Time x 10 -~, s

Fig. 13. Plot of l n 4 ~ } vs. time. Catalyst loading: 2.6×10 -2 g/cm 3, Temperature: 150°C, speed: 1000 rpm,
mole ratio of PhOH to acetone=5 : 1.

at higher temperature. Should (1+ KACA)~ 1, then Eq. (14). would become

rp = k ~ I K A C A C p W (19)

which is a typical second order equation. On integration, it gives

1
In LM-(i --XA)] = C A o k l . w .t, (20)

where M = Cpo/CAo and ka =k~,KA'


Fig. 13 shows a plot of ln[(M - XA)/M(1 - XA)] against t at 150°C, which is a
straight line passing through the origin. The slope is equal to 0.34 cm6/gmol gm s
representing kl which is a product of kfR1KA. This value can be compared with that
obtained at 135°C to find that it is reasonable because k~l increases with
temperature whereas KA decreases. This is a further proof that step 3 is the rate
determining step. Thus, at lower temperatures I+KACA could be taken as
equivalent to KACA, thereby making Eq. (14) as:

ro = rp = k'R1 Cpw (21)


= Constant if Cpo >> CAo for the same catalyst loading.

At a mole ratio of M~5, the above equation suggests that the reaction rate will be
zero order in acetone concentration and the acetone conversion will be linear with
time. Inspection of Fig. 8 as well as the comments made earlier under the effect of
temperature reiterate these observations. The slope of this plot gives a value of
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 51

8.33×10 -6 s -1 from which using Eq. (21) k'Rx is obtained as 6.41×10 -3 cm6/
gmol g s which is 28 times less than that at 135°C.
Thus the kinetic model presented here shows very interesting behaviour
which fits the experimental data well under different reaction conditions.
The effect of temperature shows that the model could be tested quite
satisfactorily.

9. Conclusions

The reaction of phenol and acetone was studied over different catalysts. DTP/K-
10, Amberlyst-31 and Amberlyst XE-717p were found to be better catalysts.
DTP/K-10 is reusable and better as regards its use at higher temperatures. The
kinetics was studied with DTP/K-10 as catalyst where the rate determining step
is the formation of p-isopropenylphenol from chemisorbed acetone and phenol
from the liquid phase within pores according to Eley-Rideal mechanism. The
catalyst was characterised fully. By taking higher mole ratio of phenol to acetone, a
number of byproducts are avoided. The kinetic model was found to fit the data
satisfactorily.

10. Nomenclature

A reactant species A, acetone


AS chemisorbed A
B product species B, bisphenol-A
BS chemisorbed bisphenol
C concentration, gmol/cm 3 for liquid phase
CA concentration of acetone
CB concentration of bisphenol-A
C, total number of active sites,
gAS concentration of chemisorbed AS
Cs concentration of vacant sites
11 first intermediate
I~S chemisorbed first intermediate
I2 second intermediate
I2S chemisorbed second intermediate
ro overall rate of reaction based on liquid phase volume, gmol/cm3/s
rp rate of reaction of phenol, gmol/cm3/sec
k psuedo-constant k'R1KA given by Eq. (18).
kR1 surface reaction rate constant for forward reaction in Eq. (3).
modified surface reaction rate constant in Eq. (15).
kR2 surface reaction rate constant for forward reaction in Eq. (6)
52 G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53

KA adsorption equilibrium constant for A


KB adsorption equilibrium constant for B
Kp adsorption equilibrium constant for P
Kw adsorption equilibrium constant for W
Kn adsorption equilibrium constant for intermediate I1
KI2 adsorption equilibrium constant for intermediate I2
KI KI2/KII
M initial molar ratio of Cpo~CAo
S Vacant site
Xn fractional conversion of acetone,
W catalyst loading, g/cm3 liquid phase
W water

Acknowledgements

This work was supported under a research grant from CSIR, New Delhi,In-
dia,(Grant No. 2(346)/91-EMR-II, 1991 to Dr. G.D Yadav under which N.
Kirthivasan received SRF. GDY also gratefully acknowledges the Herdillia Che-
micals-UDCT Diamond Jubilee Distinguished Fellowship which enabled him to
spend time on creative pursuits.

References
[1] J.B. Mnffat, Stud. Surf. Sci. Catal., 31 (1987) 241.
[2] M. Misono, Appl. Catal. A, 49 (1995) 100.
[3] G.D. Yadav and N. Kirthivasan, J. Chem. Soc. Chem. Commun., (1995) 203.
[4] R.R. Wagner, (Hercules Powder ) USP 3,172,916, 1968.
[5] B.W. McNutt, B.B. Gammill, (Dow Chemicals) USP 3,394,089, 1968.
[6] T. Naito, T. Imai, Japanese Patent 7420565, 1975.
[7] A.P. Singh, Catal. Lett., 27 (1992) 431.
[8] D. Jean, E Pierre, Fr. Demonde (Rhone-Poulenc Chemie), 2,646,418, 1990.
[9] P. Laszlo (Ed.), Preparative Chemistry using Supported Reagents, Academic, San Diego, CA., 1987.
[10] K. Smith (Ed.), Solid Supports and Catalysts in Organic Synthesis, Ellis Horwood, Chichester, U.K., 1992.
[11] N. Kirthivasan, Heterogeneous Catalysis: Preparation, Characterisation and Appliations, Ph.D. Thesis,
University of Bombay, India, 1996.
[12] T.S. Thorat, Heterogeneous Catalytic Reactions: Preparation of Solid Acids and Superacids and their
Applications, Ph.D. Thesis, University of Bombay, India, 1994.
[13] M. Slaughter, I.H. Milne, in Earl Ingersen (Ed.), Clay and Clay Minerals, Monograph No. 5, Earth Science
Series, 114-124.
[14] R.H.S. Robertson and R.M. Ward, J. Pharm. Pharmcol., 5 (1951) 27.
[15] J.H. Clark, S.R. Cullen, S.J. Barlow and T.W. Bastok, Chem. Soc. J. Perkin Trans., 2 (1994) 1117-1130.
[16] I.V. Kozhevnikov, A. Sinnema, R.J.J. Jansen, K. Pamin and H. van Bekkum, Catal. Lett., 30 (1995)
241.
[17] N.E. Sharpless and J.S. Munday, Anal. Chem., 29 (1957) 1619.
[18] Z.-X. Su and T.-J. Wang, Reactive and Functional Polymers, 28 (1998) 97.
G.D. Yadav, N. Kirthivasan/Applied Catalysis A: General 154 (1997) 29-53 53

[19] Y. Izumi, K. Urabe, M. Onaka, Zeolites, Clays and Heteropoly Acids, VCH, Weinheim, 1992.
[20] Rohm and Haas Catalogue.
[21] Rohm and Haas, U.S Patent 5,233,096, 1993.
[22] J.I. Jong and F.H.D. Dethmers, Rec. Trav. Chim. Pays Bas, 84 (1965) 460.
[23] B.C. Gates, Catalytic Chemistry, Wiley, New York, 1992.
[24] R.A. Reinicker and B.C. Gates, AIChE J., 20 (1974) 933.

You might also like