+ All Categories
Home > Documents > Kinetics of Vapor-Phase Alkylation of Benzene with...

Kinetics of Vapor-Phase Alkylation of Benzene with...

Date post: 07-Feb-2018
Category:
Upload: truongbao
View: 230 times
Download: 2 times
Share this document with a friend
6
Kinetics of Vapor-Phase Alkylation of Benzene with Isopropyl Alcohol over Commercial H-Mordenite Catalyst K. Ganapathi S. Naidu a , Sunil K. Maity b , Narayan C. Pradhan c,* and Anand V. Patwardhan d Department of Chemical Engineering, IIT Kharagpur, Kharagpur-721302, India. Email: a [email protected], b [email protected], c [email protected], d [email protected] ABSTRACT Alkylation of benzene with isopropyl alcohol to produce cumene was carried out in the vapor phase over commercial H-mordenite catalyst in a fixed-bed down-flow laboratory scale reactor. The major by-products in this reaction are m-diisopropylbenzene (MDIPB) and p-diisopropylbenzene (PDIPB). The influences of various process parameters such as temperature, space-time, and benzene-to-isopropyl alcohol mole ratio on conversion and cumene selectivity were studied. The experiments were carried out to choose the zone in which the mass transfer effects were negligible. On the basis of product distribution and informations available in the literature, a mechanism of the cumene synthesis reaction over the catalyst was proposed to explain the course of the reaction. On the basis of product distribution pattern, a kinetic model for the cumene synthesis reaction was developed following Langmuir-Hinshelwood approach. The kinetic and adsorption parameters of the rate equation were determined by non-linear regression analysis. The apparent activation energy for the main reaction was evaluated. Keywords: Alkylation; Benzene; Isopropyl Alcohol; Cumene; H-Mordenite; Kinetics INTRODUCTION Commercial production of cumene by alkylation of benzene with propylene began in 1942 to supply high-performance fuel for military aircraft. A vapor-phase process with a phosphoric acid on kieselguhr catalyst and a liquid-phase alkylation in the presence of sulfuric acid was used [1]. Large scale production of cumene started for the production of phenol and acetone by oxidation of cumene by indirect three-step Hock process. The cumene route to phenol accounts for more than 90% of world phenol capacity. It is also an important chemical precursor in the production of detergents and polymers. It is well known to catalyze the alkylation of aromatics with a variety of Lewis or Bronsted acid catalysts. Typical commercial Friedel Crafts catalysts include phosphoric acid/kieselguhr, aluminum halides, boron trifluoride, antimony chloride, stannic chloride, zinc chloride, and hydrogen fluoride. These materials are highly corrosive to process equipment, create operational problems, and are often difficult to dispose of in an environmentally acceptable manner. Moreover, use of these acids may give rise to undesirable product composition. Thus, it would be preferable to use a safer, simpler catalyst, preferably in solid state like Zeolite catalyst. Several processes have been developed for the production of cumene by the alkylation of benzene with propylene using variety of zeolite catalysts like HZSM-5 zeolite [1,2], Zeolite EU-1 [2,3], metal loaded Zeolite Beta [4], mordenite [5], zeolite beta under partial liquid phase conditions [6], type-Y zeolite modified by treatment with a phosphorous compound [7], zeolite β, mordenite, ERB-1, USY, MTW under liquid phase conditions [8], Hβ zeolite [9] and various types of non- zeolite catalysts like Ni/γ-Al 2 O 3 catalyst [10], aluminium pillared bentonite exchanged with K + , La 3+ and Al 3+ [11], fluorinated alumina [12]. However in recent years, a lot of interest is being generated on the production of cumene with isopropyl alcohol (IPA) as the alkylating agent. This is due to some advantages of using IPA instead of propylene as the alkylation agent. The olefin is the major source of carbonaceous deposits on alkylation catalysts and, therefore, a long stable life of the catalysts is observed when alcohol rather than olefin is used as the alkylating agent. Besides, the propylene as alkylating agent is very active under alkylating conditions and will oligomerize and react to give a variety of undesired by-products. In order to minimize the formation of the undesired by-products, five to ten 1 Presented at CHEMCON, Ankleshwar, Gujarat, India, 2006.
Transcript
Page 1: Kinetics of Vapor-Phase Alkylation of Benzene with ...dspace.nitrkl.ac.in/dspace/bitstream/2080/467/1/Kinetics of Vapor... · Kinetics of Vapor-Phase Alkylation of Benzene with Isopropyl

Kinetics of Vapor-Phase Alkylation of Benzene with Isopropyl Alcohol over Commercial H-Mordenite Catalyst

K. Ganapathi S. Naidua, Sunil K. Maityb, Narayan C. Pradhanc,* and Anand V.

Patwardhand

Department of Chemical Engineering, IIT Kharagpur, Kharagpur-721302, India.

Email: a [email protected], b [email protected], c [email protected], d [email protected]

ABSTRACT Alkylation of benzene with isopropyl alcohol to produce cumene was carried out in the vapor phase over commercial H-mordenite catalyst in a fixed-bed down-flow laboratory scale reactor. The major by-products in this reaction are m-diisopropylbenzene (MDIPB) and p-diisopropylbenzene (PDIPB). The influences of various process parameters such as temperature, space-time, and benzene-to-isopropyl alcohol mole ratio on conversion and cumene selectivity were studied. The experiments were carried out to choose the zone in which the mass transfer effects were negligible. On the basis of product distribution and informations available in the literature, a mechanism of the cumene synthesis reaction over the catalyst was proposed to explain the course of the reaction. On the basis of product distribution pattern, a kinetic model for the cumene synthesis reaction was developed following Langmuir-Hinshelwood approach. The kinetic and adsorption parameters of the rate equation were determined by non-linear regression analysis. The apparent activation energy for the main reaction was evaluated. Keywords: Alkylation; Benzene; Isopropyl Alcohol; Cumene; H-Mordenite; Kinetics INTRODUCTION Commercial production of cumene by alkylation of benzene with propylene began in 1942 to supply high-performance fuel for military aircraft. A vapor-phase process with a phosphoric acid on kieselguhr catalyst and a liquid-phase alkylation in the presence of sulfuric acid was used [1]. Large scale production of cumene started for the production of phenol and acetone by oxidation of cumene by indirect three-step Hock process. The cumene route to phenol accounts for more than 90% of world phenol capacity. It is also an important chemical precursor in the production of detergents and polymers. It is well known to catalyze the alkylation of aromatics with a variety of Lewis or Bronsted acid catalysts. Typical commercial Friedel Crafts catalysts include phosphoric acid/kieselguhr, aluminum halides, boron trifluoride, antimony chloride, stannic chloride, zinc chloride, and hydrogen fluoride. These materials are highly corrosive to process equipment, create operational problems, and are often difficult to dispose of in an environmentally acceptable manner. Moreover, use of these acids may give rise to undesirable product composition. Thus, it would be preferable to use a safer, simpler catalyst, preferably in solid state like Zeolite catalyst. Several processes have been developed for the production of cumene by the alkylation of benzene with propylene using variety of zeolite catalysts like HZSM-5 zeolite [1,2], Zeolite EU-1 [2,3], metal loaded Zeolite Beta [4], mordenite [5], zeolite beta under partial liquid phase conditions [6], type-Y zeolite modified by treatment with a phosphorous compound [7], zeolite β, mordenite, ERB-1, USY, MTW under liquid phase conditions [8], Hβ zeolite [9] and various types of non-zeolite catalysts like Ni/γ-Al2O3 catalyst [10], aluminium pillared bentonite exchanged with K+, La3+ and Al3+ [11], fluorinated alumina [12]. However in recent years, a lot of interest is being generated on the production of cumene with isopropyl alcohol (IPA) as the alkylating agent. This is due to some advantages of using IPA instead of propylene as the alkylation agent. The olefin is the major source of carbonaceous deposits on alkylation catalysts and, therefore, a long stable life of the catalysts is observed when alcohol rather than olefin is used as the alkylating agent. Besides, the propylene as alkylating agent is very active under alkylating conditions and will oligomerize and react to give a variety of undesired by-products. In order to minimize the formation of the undesired by-products, five to ten

1

Presented atCHEMCON, Ankleshwar, Gujarat, India, 2006.

maity
Underline
maity
Highlight
maity
Underline
Administrator
Text Box
Sunil K. Maity is presently with NIT-Rourkela, India. [email protected] This paper is Archived in dspace@nitr http://dspace.nitrkl.ac.in/dspace
Page 2: Kinetics of Vapor-Phase Alkylation of Benzene with ...dspace.nitrkl.ac.in/dspace/bitstream/2080/467/1/Kinetics of Vapor... · Kinetics of Vapor-Phase Alkylation of Benzene with Isopropyl

molar excesses of the aromatic component are needed to promote the desired alkylation reaction. In the Present work, IPA was therefore used as the alkylating agent. Several researchers have also been reported the production of cumene by the alkylation of benzene with IPA as the alkylating agent using various types of zeolite catalysts like large pore zeolite designated as NCL-1[13], beta zeolite [14,15], ZSM-5 zeolite [14], MCM-41/ γ-Al2O3 catalyst [16], protonic forms of H/β zeolite [17,18] and SAPO-5 [18]. The minimization of isomer byproduct, n-propylbenzene (n-PB) is growing need for cumene manufacturing process. Kaeding and Holland [1] reported that during alkylation of benzene with propylene over HZSM-5 catalysts, n-PB becomes a major product (36% selectivity) at 3000C. However, at temperatures below 2500C, cumene isomerization is not a favorable reaction since there is a wide departure from the thermodynamic equilibrium ratios calculated for PBs. During the isopropylation of benzene with IPA over ZSM-5 and beta zeolite; significant amount of n-PB was formed even at 2250C especially with ZSM-5 zeolite [14]. Among variety of zeolites (H-(Al)ZSM-5, H-(Fe)ZSM-5, H-M, H-ZSM-12, H-Beta, H-Y) studied for the alkyalation of benzene with IPA; H-Mordenite was reported as the best catalyst in terms of n-PB formation [19]. The protonic form of mordenite was therefore used as the catalyst in the present study. Although a few studies on preparation of cumene by isopropylation of benzene were reported using H-mordenite [19, 20]; no detailed kinetic study was found in literature. Yet another objective of this study is to propose a suitable mechanism to explain the course of the reaction and to develop a kinetic model. EXPERIMENTAL Experimental Setup. The catalytic experiments were carried out in a fixed-bed, continuous down-flow cylindrical stainless steel (SS 316) tubular laboratory scale reactor (0.014m i.d. and 0.16m in length). The reactor was fitted with a preheater in the upstream and a condenser at its outlet. The reactor was heated electrically from outside and insulated to prevent heat loss. The temperature of the reactor and preheater was measured by a thermocouple placed in a thermowell placed at the centre of the respective parts of the setup. Procedure. In a typical run, about 0.003 kg of catalyst was loaded into the reactor and supported by inert beads on either side of the bed. The catalyst was activated ‘in situ’ for 2 h in an atmosphere of nitrogen before the experimental runs were started. The aromatic-alcohol mixture was introduced with the help of a metering pump and vaporized in the pre-heater before contacting the catalyst. The reactant vapors along with nitrogen entered the reactor from the top. The product vapors, along with un-reacted reactants, were condensed in the condenser and the liquid samples were collected and analyzed in a gas–liquid chromatography using a 2 m×3 mm stainless steel column packed with 10% OV-17 on Chromosorb W (80/100). The material balance was checked and it was >98%. RESULTS AND DISCUSSION Isopropylation of benzene to produce cumene was carried out over H-mordenite catalyst. The only by-products observed were m-DIPB and p-DIPB. However no o-DIPB was observed in the product mixture. The term ‘selectivity’ of the three products, cumene, m-DIPB, and p-DIPB, used afterward is defined as the fraction of IPA converted to a particular product divided by the total conversion of IPA. Reaction Mechanism. From the observed product distribution pattern; it seems that direct formation of o-DIPB is hindered either sterically or due to overcrowding at the ortho positions. The para isomer has smaller dimensions compared to ortho and meta isomers [8] and therefore, formation of para isomer is favored due to shape selective characteristic of the zeolite catalyst. This argument is supported by the fact that ortho/para ratio for a homogeneous electrophilic substitution of an alkylbenzene decreases with increasing dimension of the alkyl group [8]. It was also observed that highest ortho/para ratio was obtained with USY (=0.45), while it was very low for other zeolites (~0.03) (Zeolite β, MOR, ERB-1, and MTW) .This seems to be reasonable because of the presence of supercages (diameter 120A) in the Y structure, which should allow the formation of the bulky ortho isomer as well. This argument is further supported by the total strain energy of 4.26, 2.02, and 2.78 kcal/mol for the final energy minimized molecules of o-, m- and p-DIPB [13]. Since strain energy of o-DIPB is very high compared to the m- and p-DIPB; the formation of o-DIPB is very difficult.

2

CHEMCON, Ankleshwar, Gujarat, India, 2006.

Page 3: Kinetics of Vapor-Phase Alkylation of Benzene with ...dspace.nitrkl.ac.in/dspace/bitstream/2080/467/1/Kinetics of Vapor... · Kinetics of Vapor-Phase Alkylation of Benzene with Isopropyl

Table 1. Effect of External Diffusional Resistance on Conversion of IPAa

Conversion (%) of IPASpace-time (h. kg

catalyst/kmol) b c

8.23 2.40 2.43 9.60 3.21 3.2

14.40 4.75 4.72 a Conditions: pressure =1 atm; Temperature = 453 K; benzene-to-IPA mole ratio = 3.43; nitrogen-to-feed mole ratio = 0.5. b = 0.002 kg of catalyst. c= 0.003 kg of catalyst.

Table 2. Effect of Intraparticle Diffusion of Conversion of IPAa

Conversion (%) of IPA at space-time (h. kg catalyst/kmol) of

Partical size, dp

8.23 9.60 14.40 0.5 2.45 3.20 7.73 1.0 2.40 3.21 4.75 1.5 2.43 3.19 4.72

a Conditions: pressure =1 atm; Temperature = 453 K; benzene-toIPA mole ratio = 3.43; nitrogen-to-feed mole ratio = 0.5.

Since alkylation of aromatic compounds over acid catalysts is an electrophilic (hence ortho-para directing) reactions, m-DIPB is not expected to be formed as a primary product in the alkylation of benzene with IPA, in spite of its lowest strain energy among the three isomers. It was thought that p-DIPB formed by the alkylation of cumene with IPA isomerizes to energetically more stable form, m-DIPB. When an alcohol is used as an alkylating agent, the isopropylation reaction is proposed to take place by the reaction of activated alkene (formed by dehydration of alcohol) with benzene via carbonium ion mechanism. This mechanism was supported by the fact that alkylation of benzene with n-propanol gave isopropyl benzene (cumene) as the major product rather than n-PB [20]. This clearly shows that the n-propyl cation formed during this reaction rearranges rapidly to the most stable isopropyl (secondary carbonium ion) cation and reacts with benzene. The cumene formed by isopropylation of benzene can also undergo disproportination reaction producing benzene and DIPBs [21,22]. The DIPBs formed again can undergo transalkylation reaction with benzene forming cumene [3,4]. Therefore based on the product distribution following mechanism ([Scheme 1) is proposed to explain the course of the reaction. Time-on-stream behavior of the catalyst. The stability of commercial H-mordenite was tested for about 4 h time-on-streams at 523 K under atmospheric pressure. The conversion of benzene and IPA and the selectivity of cumene and DIPBs (m- and p-) remained almost constant during this period as shown in Fig. 1. The catalyst is therefore very stable under the reaction conditions and essentially colorless alkylated benzene was produced even after 4 h time-on-streams. For the alkylation of benzene with propylene; Lee et al. [5] also reported that essentially no deactivation of the catalyst (mordenite) was observed even after 500 hours of use. Effect of space-time on conversion and products selectivity. The effect of space-time on conversion of benzene and IPA and product distribution is shown in Fig. 2. As it is observed from the figure, the conversion of both benzene and IPA increases with increase in space-time as expected. However the selectivity of cumene was observed to decrease with increase in space-time whereas the selectivity of DIPBs increases with space-time. This is because with increase in space-time alkylation of cumene with IPA forming p-DIPB increases that result lower selectivity of cumene (and higher selectivity of DIPBs) at higher space-time. Table 3. Kinetic and Adsorption Constants of Equations 6 and 7

Temperature, K Parameters 433 453 473 493

k1, kmol/h.atm.kg

0.001 0.007 0.03 0.125

k2, kmol/h.atm.kg

0.94 1.66 4.79 17.2

k3, kmol/h.atm.kg

0.31 0.33 0.34 0.36

KA, atm-1 0.51 0.39 0.21 0.085

Alkylation of Benzene: )(+ )( ⎯→⎯+ WwaterCCumene (B) Benzene (A) IPA 1k

(1)

Alkylation of Cumene: )(+−⎯→⎯+ )( WWaterDIPB(D)pIPA(A)CCumene 2k

(2)

Isomerisation: (E) DIPB-mDIPB(D)p 3k⎯→⎯−

(3)

Transalkylation: 2Cumene(C)Benzene(B)DIPBs →+

(4)

Disproportination: Benzene(B)DIPBs2Cumene(C) +→

(5)

Scheme 1.

3

CHEMCON, Ankleshwar, Gujarat, India, 2006.

Page 4: Kinetics of Vapor-Phase Alkylation of Benzene with ...dspace.nitrkl.ac.in/dspace/bitstream/2080/467/1/Kinetics of Vapor... · Kinetics of Vapor-Phase Alkylation of Benzene with Isopropyl

0 30 60 90 120 150 180 210 2400

10

20

70

80

90

100

Cumene

IPA

Benzene

Time-on-stream (min)

Con

vers

ion

(%)

0

10

20

30

70

80

90

100

m-DIPB

p-DIPB

Selectivity (%)

Fig. 1. Effect of time-on-stream on conversion and products selectivity. Conditions: pressure = 1 atm; temperature = 523 K; benzene-to-IPA mole ratio = 4.87; space-time = 25.4 h. kg catalyst/kmol; nitrogen-to-feed mole ratio = 0.5.

8 10 12 14 16 18 20 22 24 265

10

15

70

80

90

100

p-DIPB

m-DIPB

Cumene

IPA

Benzene

Space-time (h. kg catalyst/kmol)

Con

vers

ion

(%)

0

10

20

70

80

90

selectivity (%)

Fig. 2. Effect of space-time on conversion and products selectivity. Conditions: pressure 1 = atm; temperature = 523 K; benzene-to-IPA mole ratio = 4.87; nitrogen-to-feed mole ratio = 0.5.

Effect of Temperature on Conversion and Products Selectivity. The effect of temperature on conversion of benzene and IPA and product distribution was studied in the temperature range of 443-543 K under otherwise identical experimental conditions as shown in the Fig. 3. As it is observed from the figure, the conversion of benzene and IPA increases with increase in temperature as expected. The selectivity of cumene was observed to increase with increase in temperature whereas opposite trend was observed for both m- and p-DIPB. At higher temperatures, transalkylation of DIPBs with benzene to cumene increases, thereby increasing the selectivity of cumene and decreasing the selectivity of DIPBs. Effect of benzene-to-IPA mole ratio on IPA conversion and products selectivity. The effect of benzene-to-IPA mole ratio on IPA conversion and products selectivity was studied by varying the benzene-to-IPA mole ratio from 1.57 to 7.75 as shown in Fig. 4. The conversion of IPA was found to increase with increase in benzene-to-IPA mole ratio. The selectivity of cumene was however found to increase with increase in benzene-to-IPA mole ratio. The increase in selectivity of cumene is due to the fact that the transalkylation of DIPBs with benzene is favorable at higher aromatic-to-alcohol mole ratios.

440 460 480 500 520 5400

10

20

30

40

50

60

70

80

90

100 Conversion

Benzene IPA

Temperature (K)

Con

vers

ion

(%)

0

5

10

15

75

80

85

90Selectivity

Cumene m-DIPB p-DIPB

Selectivity (%)

Fig. 3. Effect of temperature on conversion and products selectivity. Conditions: pressure = 1 atm; benzene-to-IPA mole ratio = 4.87; space-time = 25.4 h. kg catalyst/kmol; nitrogen-to-feed mole ratio =0.5.

1 2 3 4 5 6 7 80

10

20

30

40

50

60

70

80

90

DIPB

Cumene

Benzene-to-IPA mole ratio

Con

vers

ion

of IP

A (%

)

10

15

20

25

70

75

80

85

90

Selectivity (%)

Fig. 4. Effect of benzene-to-IPA mole ratio on IPA conversion and products selectivity. Conditions: pressure =1 atm; temperature = 483 K; space-time = 10.79 h. kg catalyst/kmol; nitrogen-to-feed mole ratio = 0.5.

4

CHEMCON, Ankleshwar, Gujarat, India, 2006.

Page 5: Kinetics of Vapor-Phase Alkylation of Benzene with ...dspace.nitrkl.ac.in/dspace/bitstream/2080/467/1/Kinetics of Vapor... · Kinetics of Vapor-Phase Alkylation of Benzene with Isopropyl

8 10 12 14 16 18 20 220

10

20

30

40

50

60

70

Con

vers

ion

of IP

A (%

)

Space-time (h. kg catalyst/kmol)

Temperature (K) 433 453 473 493

Fig. 5. Effect of space-time on conversion of IPA. Conditions: pressure =1 atm; benzene-to-IPA mole ratio = 3.43; nitrogen- to-feed mole ratio = 0.5.

0 10 20 30 40 50 60 700

10

20

30

40

50

60

70

Pred

icte

d IP

A c

onve

rsio

n (%

)

Experimental IPA conversion (%)

433 K 453 K 473 K 493 K

Fig. 6. Experimental vs predicted IPA conversions

Mass transfer considerations. For any kinetic study, it is important that the mass transfer resistances be negligible during the reaction. To estimate the external diffusional effects, experiments were carried out at constant space-time and catalyst size, but with varying feed rates. The results of Table 1 indicate that the conversion of IPA for both the series at constant W/F are independent of feed rate. Therefore, the external mass transfer resistance is negligible. Experiments were also conducted to test the intraparticle diffusional limitations by varying the catalyst particle size while keeping space-time constant. The experimental data obtained are presented in Table 2. The results showed that there was no change in conversion of IPA with catalyst size indicating negligible intraparticle mass transfer resistance in the particle size range studied. The particle sizes employed in the kinetic study were within the intraparticle diffusion free range. In zeolite-catalyzed reactions, two types of diffusion processes are involved: (i) micropore diffusion inside the zeolite crystal and (ii) macropore diffusion between the zeolite crystals within the catalyst pellets. The above experiments for mass transfer resistances confirms only the absence of diffusion in the macropores. The resistance due to micropore could not be evaluated, as it requires a modification of the synthesis conditions of the zeolite that affect the micropore size of the crystals, which would subsequently affect the diffusional characteristics. Hence, the kinetic parameters presented here include these diffusional effects, if any. KINETIC MODELING The kinetic runs were carried out at four different temperatures (namely, 433, 453, 473, and 493 K) and for each temperature; space-time was varied by changing the liquid feed rate. Fig. 5 presents the effect of space-time on IPA conversion at four different temperatures. It was observed that conversion of IPA increases with an increase in space-time at all temperatures. Various reaction rate models were formulated considering adsorption, desorption, and surface reaction controlling following Langmuir–Hinshelwood approach. All models, except the surface reaction controlling one, gave unrealistic values of various constants with improper trends. Hence, they were not considered. The following surface reaction-controlling model (IPA is only adsorbed on the active site of the catalyst) was found to fit the experimental data with proper trend of various constants. The rate of disappearance of IPA and rate of formation of m-DIPB is represented by following set of equations.

]/ZppKkppK[kDWDF)r( CAABAA

AA 21 +=−=− where AA pKZ += 1 (6)

DE

E pkDWDF)(r 3==

(7) The partial pressures (pi) of any species, i, as appeared in the above rate equations are related to its mole fraction (xi) and the total pressure (P) by the following expressions.

Pxp ii = (8)

5

CHEMCON, Ankleshwar, Gujarat, India, 2006.

Page 6: Kinetics of Vapor-Phase Alkylation of Benzene with ...dspace.nitrkl.ac.in/dspace/bitstream/2080/467/1/Kinetics of Vapor... · Kinetics of Vapor-Phase Alkylation of Benzene with Isopropyl

A nonlinear regression algorithm was used for parameter estimation. The optimum values of the parameters were estimated by minimizing the objective function (f) given by

2

1exp)( ]))((r)(r[f ipre.Ai

n

i.A −= ∑

=

(9)

The kinetic and adsorption constants of the model equations evaluated at various temperatures by nonlinear regression are tabulated in Table 3. Based on evaluated kinetic and adsorption constants of model Eq. 6; IPA conversions was calculated at four different temperatures and plotted against the experimentally obtained IPA conversions as shown in Fig. 6. It shows that the proposed reaction rate expression predicts the IPA conversion values comparable with the experimental ones. The kinetic constants evaluated and tabulated at various temperatures (Table 3) were used to determine the activation energy and frequency factor of the main reaction (Eq. 1) as 133.5 KJ/mol and 1.79×1013 respectively. CONCLUSIONS The commercial H-mordenite catalyst used for the synthesis of cumene by the alkylation of benzene with IPA was found to have good activity and selectivity towards cumene. No formation of n-PB was observed in this reaction. The major by-products in this reaction were m-DIPB and p-DIPB. The selectivity of cumene was found to increase with increase in temperature as well as benzene-to-IPA mole ratio. However the selectivity of cumene was found to decrease with space-time. Based on the product distribution, a suitable reaction mechanism was proposed to explain the course of the reaction. A kinetic model was developed to correlate the rate of disappearance of IPA. The kinetic and adsorption constants of the rate equations were estimated by non-linear regression algorithm. The activation energy for the cumene synthesis reaction was determined to be 133.5 kJ/mol. ACKNOWLEDGMENT Sunil K. Maity is thankful to the All India Council for Technical Education (AICTE), New Delhi, India, for the award of the National Doctoral Fellowship during the tenure of this work. REFERENCES

1. Kaeding, W. W.; Holland, R. E. Journal of Catalysis 109, 1988, 212-216. 2. Pradhan, A.R.; Kotasthane, A.N.; Rao, B.S. Applied Catalysis 72, 1991, 311-319. 3. Ramchandra, P. A.; Seshagiri, R. B. Patent number EP19910309531 19911016, 1993. 4. Seshagiri, R. B.; Ramchandra, P. A.; Paul, R. Patent number EP0538518, 1993. 5. Lee, G.J.; Garces, J.M.; Meima, G.R; Matheus, J. M. Patent number EP0433932, 1991. 6. Miller, S. J. Patent No. 20030204121, 2003. 7. Cavani, F.; Girotti, G.; Arrigoni, V.; Terzoni, G.; US Patent No. 5,650,547, 1997. 8. Perego, C.; Amarilli, S.; Millini, R.; Bellussi, G.; Girotti, G.; Terzoni, G. Microporous

Materials 6, 1996, 395-404. 9. Han, M.; Lin, S.; Roduner, E. Appl. Cata. A: Gen. 243, 2003, 175–184. 10. Panming, J.; Qiuying, W.; Chao, Z.; Yanhe, X. Appl. Cata. A: Gen. 91, 1992, 125-129. 11. Geatti, A.; Lenardo, M.; Storaro, L.; Ganzarla R.; Perissinotto, M. J. Mole. Cata. A: Chem.

121, 1997, 111-118. 12. Rodriguez, L.M.; Alcaraz, J.; Hernandez, M.; Taarit, Y. B.; Vrinat, M. Appl. Cata. A: Gen.

169, 1998, 15-27. 13. Sasidharan, M.; Reddy, K. R.; Kumar, R. Journal of Catalysis 154, 1995, 216-221. 14. Halgeri, A. B.; Das, J. Appl. Cata. A: Gen.181, 1999, 347-354. 15. Girotti, G.; Rivetti, F.; Ramello, S.; Carnelli, L. J. Mole. Cata. A: Chem. 204–205, 2003,

571–579. 16. Medma-Valuerra, J.; Zaldivar, O.; Sánchez, M. A.; Montoy, J.A.; Navarrete , J.; Reyes, J.

A. de los. Appl. Cata. A: Gen. 166, 1998, 387-392. 17. Kasture, M.W.; Niphadkar, P.S.; Sharanappa, N.; Mirajkar, S.P.; Bokade, V.V.; Joshi,

P.N. Journal of Catalysis 227, 2004, 375–383. 18. Sridevi, U.; Bokade, V.V.; Satyanarayana, C.V.V.; Rao, B.S.; Pradhan, N. C.; Rao, B.K.B.

J. Mole. Cata. A: Chem. 181, 2002, 257–262. 19. Wichterlová, B.; Čejka, J.; Zilková, N. Microporous Materials 6, 1996, 405-414. 20. Reddy, K.S.N.; Rao, B.S.; Shiralkar, V.P. Appl. Cata. A: Gen. 95, 1993, 53-63. 21. Tsai, T.; Ay, C.; Wang, I. Applied Catalysis 77, 1991, 199-207. 22. Tsai, T.; Chang, S.; Wang, I. Ind. Eng. Chem. Res. 42, 2003, 6053-6058.

6

CHEMCON, Ankleshwar, Gujarat, India, 2006.


Recommended