Você está na página 1de 8

Catalytic Distillation Modelling and Simulation using HYSYS.

Process Environment
Gheorghe BUMBAC1, Grigore BOZGA 1 , Valentin PLESU 1, Vasile BOLOGA 1 , Ilie. MUJA 2 and Corneliu Dan POPESCU 2
1

University POLITEHNICA of Bucharest, Department of Chemical Engineering, 1 Polizu Street, RO-78126, Bucharest, Romania, Tel/Fax:+40 (0)21 21.25.125, email: cttip@chim.upb.ro 2 S.N.P. PETROM, INCERP Ploiesti Subsidiary, B-dul Republicii Nr. 291 A, RO-2000, Jud. Prahova, Telefon +40 (0)244 135111, Fax +40 (0)244 198732, email: popescu@serv.incerp.ro
The catalytic-distillation process for the production of t-amyl-methyl-ether (TAME) from methanol and isoamylenes was simulated by developing the process model as a combination of unit operations from HYSYS operations palette. Geometrical characteristics of catalytic-distillation column are those of an industrial pilot plant and the results of simulation were compared with experimental data. The experimentally determined reactions kinetics was applied in the model. UNIQUAC-UNIFAC model equations were selected for the vapour-liquid equilibrium. The results show that fair agreements between the calculated and experimental data were obtained.

1. INTRODUCTION Methyl ethers replace lead compounds in gasoline. One of these ethers obtained by the etherification of an isoamylenes mixture (2-methyl-1-butene, 2M1B, and 2-methyl-2-butene) with methanol is t-amyl-methyl-ether (TAME). Catalytic distillation is a suitable technique for TAME synthesis due to the reversibility of the etherification reactions and the difference between the reactants and products volatilities. These particularities favours the enhancement of the reactant conversion and the increase of the interphase mass transfer potential. Despite of the important number of publications in the field of catalytic distillation, relatively few of them concern the industrial application of TAME synthesis. In this paper we focused on the relevance of the commercial software HYSYS for the simulation of catalytic distillation problems. To predict the behaviour of TAME synthesis reactor and reactive distillation column HYSYS.Process environment was used. The simulation results were compared with pilot plant experimental data. The pilot plant system consists mainly from a tubular, fixed bed pre-reactor (TFBR) and a reactive distillation (RD) packed column. The purpose of this study is: a) to develop a suitable simulation module for heterogeneous RD with HYSYS and b) to apply the model to an industrial applications. The TFBR is used to bring the reaction mixture near its equilibrium composition. The advantage of using a pre-reactor for TAME synthesis is based on the fact that the greatest part of reaction components can react before RD column and the throughput of reaction system increases. HYSYS provides many built-in modules for simulating various processes. Unfortunately the COLUMN subflowsheet environment allows simulation of RD with reactions taking places only in homogeneous phase. In the heterogeneous catalytic distillation process the solid catalyst particles are placed into many special packing envelopes, serving also as vapour-liquid contacting. The reactions occur inside the catalytic package where the liquid contacts the catalyst particles. Then the product flows out of the catalytic zone. Additional separation takes place on the packing placed below and above the reaction section of the column.
1

In a RD column the reaction and separation actually take place in the different locations of the column i.e. reaction on the catalyst pellets of the packaging and separation in the inert packing. Therefore, the parameters of the liquid residence time or the liquid hold-up on the trays or packings in the heterogeneous process can only be used in the separation calculation whereas the reaction calculation needs the parameters of contacting time of liquid with catalyst in the catalytic package. We underline that in the current version of HYSYS the built-in RD module is not directly suitable for the simulation of the heterogeneous catalytic distillation process. To overcome the above problem our study concentrated to develop a model for heterogeneous RD and implement the model in the HYSYS simulation environment. In this model, the catalyst space velocity appearing in the reaction system equations represents the contacting time of liquid with catalyst. 2. Reaction kinetics and thermodynamics Industrial processes for TAME synthesis are based on the reversible reactions of isoamylenes (2M1B and 2M2B) with methanol. The equilibrium conversion of isoamylenes to TAME, at 60C, is 56% if stoechiometric amounts of isoamylenes and methanol are used 6 and increases slightly with the increase of methanol/isoamylenes ratio. In Table 1 the equilibrium conversion as a function of temperature, for stoechiometric methanol/ isoamylenes molar ratio, is presented 6 . A typical industrial process for TAME synthesis involves at least 8 components: isoamylenes, n-pentane, i-pentane, methanol, 1-pentene and trans-2-pentene. Since methanol associates almost all hydrocarbon components into simple and complex azeotropic pairs, the system shows strong non-ideal properties. The property package used to calculate the liquid activities of the considered components is based on UNIQUAC-UNIFAC model. Table 1. Equilibrium isoamylenes conversion as a function of temperature Temperature Conversion o ( C) 50 0.618 60 0.560 70 0.491 80 0.446
Table 2. The feed composition.

The two reactive olefins (2M1B and 2M2B) are contained in the hydrocarbon mixture, resulted as a C5 fraction from Fluid Catalytic Cracking (FCC) unit. In Table 2 the composition of the feeding mixture used in the simulated process scheme is presented. Residual TAME is present in this stream from recycled stream. All components from the C5 fraction are producing azeotropes with methanol and the composition of these azeotropes is presented in Table 3. The methanol concentration in these azeotropes increases with pressure.

Table 3. Azeotrope compositions in TAME synthesis. p=2.5 bar p=4 bar o Component 1 Component 2 x1 t, C x1 t, o C methanol 2M1B 0.21 53.76 0.243 69.24 methanol 2M2B 0.28 58.69 0.31 73.78 methanol n-pentane 0.295 58.64 0.328 73.96 methanol i-pentane 0.21 51.22 0.252 66.61 methanol 1-pentene 0.22 53.64 0.267 69.06 methanol 2-pentene 0.265 56.73 0.301 72.29 methanol TAME 0.763 87.56 0.793 102.45

p=5.5 bar x1 t, o C 0.268 80.07 0.331 84.69 0.347 85.20 0.280 77.85 0.283 80.75 0.322 82.72 0.802 113.20

The synthesis of TAME from methanol and isoamylenes, catalysed by acid ion-exchange resin catalyst is a reversible process as shown in following reaction mechanism containing the main and secondary reactions (17):
r1
CH 2 = C CH 2 CH 3 (l ) + CH 3 OH ( l) | CH 3 2 Methyl 1 Butene ( 2M1B ) O CH 3 | CH 3 C CH 2 CH 3 (l ) | CH 3 TAME (1)

r2

r3
CH 3 C = CH CH 3 (l ) + CH 3 OH (l ) | CH 3 2 Methyl2 Butene (2 M 2 B )
CH 2 = C CH 2 CH 3 ( l ) | CH 3 2 M 1B

r4

O CH 3 | CH 3 C CH 2 CH 3 (l ) | CH 3 TAME
C = CH CH 3 ( l ) | CH 3

( 2)

r5 r6

CH 3

( 3)

2 M 2B

C 5 H 10 ( l ) + C 5 H 10 ( l ) 2 M 1B 2M 2B

C 10 H 20 ( l ) di iso amylene

(4)

C 5 H 10 ( l ) + C 5 H 10 ( l ) 2M 2B 2M 2B 2CH 3 OH ( l)
C5 H10 (l) + H 2O

C 10 H 20 ( l ) di iso amylene
+ H 2 O ( l) 3 ( l) di methyl ether C 5H12 O (l)
tert amyl alcohol

(5)

CH 3 O CH

( 6)
(7)

iso amylenes

There are three main reactions (reactions 13, one for etherification of 2M1B, one for etherification of 2M2B and an isomerisation reaction between 2M1B and 2M2B) and four secondary reactions (47). Both etherification reactions, (1) and (2), are exothermic, i.e. the equilibrium conversion of TAME decreases with temperature. The isomerisation reaction at operation temperature (between 60C and 120C) favours the 2M2B formation and this component will have the greatest concentration in the reaction mixture. From kinetic point of
3

view this situation is not advantageous because a faster reaction (1) is replaced by a slower one (2). High temperatures and low methanol concentrations are favourable conditions for isoamylenes oligomers formation. On the other hand, excess methanol produces higher dimethyl-ether concentration, whereas t-amyl-alcohol formation is very limited, in equilibrium conditions, due to a very small water concentration. Frequently used catalysts are sulphonic acid ion-exchange resins (Amberlyst 15 or 35, Levatit SPC 118). The kinetic mechanism is based on the consideration that methanol and TAME are stronger adsorbed on the catalysts surface, compared to isoamylenes. According to our knowledge kinetic studies for TAME synthesis were published by Muja et. al 7 , Randriamahefa et al 8 , Piccoli and Lovisi , Oost and Hoffmann 3 and Rihko et al. 2 etc. In this work the TAME synthesis kinetic model of Rihko et al. 3 on Amberlyst 16 was used. The 2M1B and 2M2B consumption rates have the expressions:

r1B

aT 1 k B1 aM a 2M 1B 1 K a a 1 a2 M 2 B 1 M 2 M 1B (8) = k B5 a2 M 1B 1 K3 a2 M 1B KT K aT + aM M
1 aT k B3 aM a 2M 2 B 1 K 2 a M a 2M 2B = KT K aT + a M M +k 1 a2 M 2 B a 2M 1B 1 K 3 a 2M 1B

r2 B

B5

(9)

The activation energies for the reactions (1), (2) and (3) are: EkB1 = 76 kJ / mol ; EkB3 = 94.1 kJ / mol ; E kB5 = 81.6 kJ / mol and mol mol mol ; k B3 ( 343K ) = 0.125 ; k B5 ( 343K ) = 0.107 g h gh g h are reaction rates constants at 343 K. k B1 ( 343 K ) = 0.286 The equilibrium constants in the above reaction rates as temperature functions are: K1 = exp ( 8.3881 + 4041.2 / T ) (10) K 2 = exp ( 8.2473 + 3225.3 / T ) (11) K 3 = exp ( 0.1880 + 833.3 / T) (12) KT = 0.1405 0.00061 (T 334) (13) KM The ion exchange capacity Amberlyst 15 is 5 mequiv/g 8,9. The reaction kinetic data have been verified using pilot plant synthesis of TAME in which the fixed bed reactor packing of the same catalyst was used. Molar ratio between methanol and isoamylenes in the feed stream was 1.256. 3. Process flowsheet In figure 1 TAME synthesis process flowsheet using catalytic distillation is presented. The C5 fraction is mixed with methanol and the resulting stream is fed to the preliminary

reactor (IV). In the Table 4 the composition of the mixture at the preliminary reactor exit is presented. The resulting product is mixed with a recycled methanol stream and is fed to a catalytic distillation column, with three zones, below the reaction zone. The stripping zone of the catalytic-distillation column is simulated as reboiled absorber, a standard operation in HYSYS. The second part is the reaction separation zone, represented in our model by a backflow cell model (BCM) with forward flow of the liquid and backward flow of the vapour in the reactive part of RD zone. The BCM consist of series of five continuous stirred tank reactors (CSTR) units with the same geometry and size of the individual unit. The third part is another pure mass transfer unit, representing the rectifying zone of the reactive distillation. This zone is simulated as refluxed absorber a HYSYS standard operation. Both stripping and the rectifying zones are represented as non-catalytic packed columns. From the computational point of view, each cell of the series was assumed to be at VL equilibrium, the increase of conversion being calculated as in a CSTR reactor. Here reactions take place in liquid-solid interface, following the kinetic law mentioned above. System characteristics: The main characteristics of the catalytic distillation column are: pre-reactor volume: 0.12 m3 , stripping zone: 6 theoretical stages, rectifying zone: 3 3 theoretical stages and in cell model there are five CSTRs, a CSTR vessel has 0.02 m of catalyst. Catalyst particles average diameters considered in the simulation were of 1 mm as in the pilot plant case. These characteristics are in agreement with those of the pilot plant. More explicit characteristics for stripping and rectifying zones of the RD column are presented in Tables 6 and 7. Table 4. Feed conditions for the RD column system
Table 5. Column heat exchangers

Table 6. Stripping zone characteristics

Table 7. Rectifying zone characteristics

4. Results and discussion: Several results from the simulation of main streams are presented in Table 8. The isoamylenes conversion in the reactive distillation column and pre-reactor is 80.76 %. Pilot plant scheme is presented in Figure 2. The characteristics of the experimental pilot plant are: pre-reactor volume 120 l; rectifying zone packing height 2 m; stripping zone packing height 3.5 m. Feeding condition and thermodynamic regime was the same as in the simulation. The simulation results with HYSYS for the TAME synthesis reactive distillation module set-up, presented in this work, allow drawing the following conclusions: - From the chemical transformation point of view it is profitable to place the reaction zone as close as possible to the top of the column. However, above the reaction zone a separation zone is needed to separate TAME from the distillate. - It is recommended to place the column feed bellow the reaction zone in order to ensure high concentration for the reactants in this zone (as there are more volatile compared with the reaction product). - The best structure for the RD column, obtained from this simulation study, involve 15 theoretical plates. As we denoted plates from top to bottom, the best position for the reaction zone are the theoretical plates 3 and 4, and the feed plate is the 5-th plate. The optimal reflux ratio is 2, as result of the trade-off between separation degree and energy saving. To describe the flow and fluid phase mixing in the reaction zone, a classical, multicellular, model was used, considering the back-flow of the vapour phase. In each cell the conversion increase was calculated considering uniform distribution of the catalyst and vapour-liquid equilibrium. The results obtained (over 90% conversion, much bigger than the equilibrium conversion) emphasises the advantage of catalytic distillation compared with the classical scheme, because the chemical transformation is not limited by the chemical equilibrium as result of continuous separation of the reaction product from the mixture. Table 8. Simulation results.

III

LN+1

n=N

VN VN-1 LN-1 n=3 L4 V2

II

LN VN-2

n=N-1

V3 L3 n=2

V1

n=1

L2

IV

I
L1 V0

a) process flowsheet

b) cells in series model

Figure 1. Simplified flowsheet for RD column.

Figure 2. TAME Pilot Plant with Catalytic Distillation Column. Iso-amylenes conversion data (Table 9) are obtained using the pilot plant presented in fig. 2, under the same experimental conditions.

Table 9. Experimental results for TAME synthesis.

5. Conclusions This paper presents a theoretical study for the modelling of reactive distillation column operation in TAME synthesis. The simulation procedure is based on a mathematical model considering chemical reaction kinetics for the main reactions and the vapour-liquid equilibrium. Phase contact in the reaction zone is described with the back-flow cell model. The problem statement in HYSYS.Process environment was made considering three zones for the catalytic distillation column (rectifying, reaction and stripping). Constructive and operational characteristics of the column are specified as a consequence of the parametric study: reaction zone position, feed position and reflux ratio, in order to obtain maximum yield for the transformation of C5 reactive olefins in the pilot plant. The simulation results are in good agreement with experimental data obtained in the experimental pilot plant at SNP PETROM, INCERP Ploiesti subsidiary. The quality of the results obtained in this paper is limited by the uncertainty introduced by the phase hydrodynamics in the reaction zone the phase equilibrium hypothesis. The authors foreseen additional studies in order to better describe phase hydrodynamics, to consider interphase mass transfer inside the catalyst pallets on process performances.

Nomenclature: ai activity of component i; Kj equilibrium constants in TAME synthesis reactions; kBm rate constants in reactions (8), (9);

ri reaction rate; p pressure; T absolute temperature, K; R ideal gas constant.

REFERENCES 1. L.K. Rihko, A.I.O. Krause, Ind. Eng. Chem. Res. 1995, 34, 1172. 2. L.K. Rihko, P.K. Kiviranta-Pkknen, A.I.O. Krause, Ind. Eng. Chem. Res., 1997, 36, 614. 3. C. Oost, U. Hoffmann, Chem. Eng. Sci. 1996, 51, 329. 4. W.B.Su, J.R. Chang, Ind. Eng. Chem. Res, 2000, 39, 4140. 5. A.P. Higler, R. Taylor, R. Krishna, Chem. Eng. Sci. 1999, 54, 1389. 6. K. Sundmacher, U. Hoffmann, Chem. Eng. Sci, 1994, 49, 4443. 7. L. Muja, et. al., Revista de Chimie, 1986, 37, 1047. 8. S. Randriamahefa, R. Gallo, J. Mol. Catal. 1988, 49, 85. 9. H. Subawalla, J.R. Fair, Ind. Eng. Chem. Res., 1999, 38, 3696. 10. J.L. DeGarmo, V.N. Parulekar, V. Pinjala, Chem. Eng. Progress, 1992, 88, 43. 11. J.C. Gonzales, H. Subawalla, J.R. Fair, Ind. Eng. Chem. Res., 1997, 36, 3845. 12. J.C. Gonzales, J.R. Fair, Ind. Eng. Chem. Res. 1997, 36, 3833. 13. S. Ung, M.F. Doherty, Chem. Eng. Sci., 1995, 50, 23.

Você também pode gostar