Sie sind auf Seite 1von 9

Ind. Eng. Chem. Res.

2006, 45, 1613-1621

1613

Multiple Product Solutions of tert-Butyl Alcohol Dehydration in Reactive Distillation


Zhiwen Qi and Kai Sundmacher*,,
Max-Planck-Institute for Dynamics of Complex Technical Systems, Sandtorstrasse 1, D-39106 Magdeburg, Germany, and Process Systems Engineering, Otto-Von-Guericke-UniVersity Magdeburg, UniVersitatsplatz 2, D-39106 Magdeburg, Germany

The feasibility of tert-butyl alcohol (TBA) dehydration to selectively produce the high purity products isobutene or diisobutene by using reactive distillation has been illustrated through process simulations. Correspondingly, different column configurations are proposed. The influence of the important operating parameters on the column performance is analyzed by continuation methods, and the operating windows of the parameters are suggested. Advantages of employing reactive distillation for TBA dehydration include the mild conditions required, high TBA conversion per pass, and high selectivity of the desired products. Moreover, the excellent coupling of TBA hydration and isobutene dimerization in one column leads this system to be a good example for studying the complex behaviors of reactive distillation integrating multiple chemical reactions.
1. Introduction tert-Butyl Alcohol (TBA) production is of interest due to its important use as a gasoline additive to meet part of the oxygenate requirement for reformulated gasoline. Since TBA is completely soluble in water, even small amounts originally present in gasoline can result in high levels in groundwater. As indicated by its organic carbon partition coefficient, TBA is somewhat more retarded in its movement through soils than methyl tert-butyl ether (MTBE), but after TBA reaches groundwater, it is even more difficult to remove by carbon filters or air stripping. TBA has been detected at concentrations as great as 18 000 g/L (much higher than the drinking water action level of 12 g/L) in groundwater near source areas of groundwater plumes in the USA.1 From a toxicological point of view, TBA shows evidence of carcinogenicity. Therefore, it seems that TBA is not a good replacement of MTBE as a gasoline additive and should be decomposed to manufacture other products. There has been an enormous technological interest in TBA dehydration during the past years, first, as a primary route to MTBE and, more recently, for the production of isobutene (IB) and polyisobutene. The valuable advantage of produced IB from TBA dehydration is that no other C4 compound exists together with IB, which is important for some IB downstream processing. Moreover, selectively produced dimerization products are of interest as a replacement for MTBE, which has attracted negative publicity because of its water solubility. The ban on the use of MTBE in California has spurred interest in new octaneenhancing components based on IB.2 Dimerization of IB produces diisobutenes (DIB, i.e., 2,2,4-trimethyl-4-pentene and 2,4,4-trimethyl-1-pentene), which can be used in gasoline as an additive or hydrogenated to isooctane. Another option is to etherify the DIBs to components such as 2-methoxy-2,4,4trimethylpentane and 2-ethoxy-2,4,4-trimethyl-pentane.3 A number of commercializable TBA dehydration processes have been developed, which typically involve either vapor phase
* To whom correspondence should be addressed. Phone: +49-391-6110351. Fax: +49-391-6110353. E-mail: sundmacher@ mpi-magdeburg.mpg.de. Max-Planck-Institute for Dynamics of Complex Technical Systems. Otto-von-Guericke-University Magdeburg.

reaction over a silica-aluminum catalyst at 260-370 C or liquid phase reaction utilizing homogeneous or solid acid catalysis.4-7 In such conventional reactors, the TBA conversion is not complete due to the chemical equilibrium limitation and the selectivity is not high due to the IB oligomerizations under the high temperature. The nature of the TBA dehydration (i.e., fast, water inhibited, and chemical equilibrium controlled) and the difference in the volatilities of the species involved (difference in the normal boiling points between IB and others is more than 100 K) leads to this reaction as a potential candidate for implementation in a reactive distillation column. The reactive distillation process of TBA dehydration has been demonstrated experimentally with reduced pressure8 and enhanced pressure.9 It is also a potential of reactive distillation technology for simultaneous dehydration of TBA and further dimerization of the produced IB by high reflux of IB inside the column. Both reactions are fast and use a common catalyst, for example, ionexchange resin10,11 can be applied. Though the DIB selectivity is favored by a polar component like TBA and water during the IB dimerization reaction, as suggested by Honkela and Krause,12 it is still the key concern in such a reactive distillation process that couples two reactions. This is because of the relatively high concentration of IB and the high reaction temperature in part of the reactive zone, which may promote the side reactions. In this contribution, TBA dehydration for the products IB and DIB is analyzed based on simulation studies. The influence of the most important parameters is investigated by continuation methods. Special attention is given to the conversion of TBA and selectivities toward the desired products. 2. Kinetic Models for Reactions Involved In this work, two main chemical reactions for producing IB and DIB are involved:

TBA S IB + H2O 2IB f DIB

(1) (2)

Beside the main reaction of IB dimerization, oligomerization of IB to higher oligomers may occur theoretically. As argued

10.1021/ie0511027 CCC: $33.50 2006 American Chemical Society Published on Web 01/19/2006

1614

Ind. Eng. Chem. Res., Vol. 45, No. 5, 2006

Table 1. Kinetics of the Reactions Involved reaction TBA S IB + H2O kinetics source 10

r1 )

kTBA(KaaTBA - aH2OaIB) aTBA + 1.5aH2O

kTBA ) 0.756 exp a

( 3111.9 K ) exp(7.6391 T )
kDIBaIB2

1 18000 1 R T 343

)]

(kmol/(h kgcat))

2IB f DIB

11

r2 )

[aIB + 7.0(aTBA + aH2O)]2

kDIB ) 0.82 exp DIB + IB f TIB

1 30000 1 R T 373.15

)]

(kmol/(h kgcat))
11

r3 )

kTIBaIBaDIB [aIB + 7.0(aTBA + aH2O)]3

kTIB ) 0.065 exp -

1800 1 1 R T 373.15

)]

(kmol/(h kgcat))

by Honkela and Krause,11 tetramers and higher oligomers need not to be considered because of their very low formation rate compared to the dimer and trimer. So, only one side reaction for triisobutene (TIB) formation is taken into account:

DIB + IB f TIB

(3)

The kinetics of TBA dehydration or IB hydration have been widely studied. The solid catalysts are mainly classified as inorganic ones such as zeolite13 and ion-exchange resins.10,14-16 Though they are reversible, the hydration of IB on ion-exchange resin is different from the dehydration of TBA because of a large amount of water on the catalyst that leads to the resin being fully swollen. Since the concentration profile of water is very different in reactive distillation, a model for TBA dehydration is needed for the column simulation. The recent presented kinetics of TBA dehydration and IB dimerization by Krause and co-workers are probably the most suitable ones when these two reactions are coupled in one process. In their works, they took the TBA dehydration under IB dimerization conditions using the same catalyst of an ionexchange resin. The kinetics were fitted using the same activity model. In this work, the kinetics and the parameters from Krauses works are adopted, as listed in Table 1. Accordingly, the temperature limitation of the catalyst ion-exchange resin should be always taken into account in simulations (i.e., <397 K). One should note that, in the kinetics, the adsorption effects were not treated consistently such that different adsorption expressions were adopted. Moreover, in the IB dimerization and trimerization kinetics, the water adsorption is now considered by modifying the original ones. The adsorption constant of water is assumed as the same as TBA. Such an assumption seems reasonable since the activity coefficients vary between 13 and 42 for water and between 2 and 8 for TBA so that the strong inhibition effect of water can be compensated by the water activity coefficient. 3. System Thermodynamics For the involved mixtures, liquid phase splitting may take place, which does not change the concentration of the vapor phase but also influences the reaction rate especially when the reaction takes place only in one liquid phase. To determine a

suitable phase equilibrium model for column simulations of the reaction systems, system thermodynamics is initially investigated. The liquid nonideality is described by the Dortmund modified UNIFAC model20 as used in column simulations. First, the phase equilibrium of the ternary system of IB, water, and TBA is studied. As indicated in Figure 1, at 1 atm, two liquid phases are predicted from the vapor-liquid-liquid equilibrium (VLLE) model, a critical point of x ) (0.0150, 0.7401, 0.2449) is located on the liquid-liquid envelope. When superposing the chemical equilibrium curve of eq 1 over the map, one can find that this curve is totally outside the liquidliquid envelope. Therefore, for the IB formation system, a homogeneous liquid mixture can be expected, which will be confirmed in the later simulation. For the DIB formation system, liquid phase splitting occurs in a wide range of mixture concentration. As example, Table 2 gives the VLLE prediction of typical mixtures in the DIB column simulation (see Figure 7a). Therefore, the column simulation for DIB formation should generally consider a heterogeneous liquid mixture (i.e., the VLLE model).

Figure 1. VLLE prediction and chemical equilibrium curve of a ternary mixture of IB, water, and TBA at 1 atm.

Ind. Eng. Chem. Res., Vol. 45, No. 5, 2006 1615


Table 2. VLLE Predictions of Typical Mixtures of IB, DIB, TIB, Water, and TBA stage 14 25 29 mole fractiona x ) (0.03099, 0.24317, 0.00105, 0.29889, xa ) (0.00065, 0.00019, 0.00000, 0.96357, 0.03560) xo ) (0.03294, 0.25878, 0.00112, 0.25618, 0.45098) x ) (0.00318, 0.26617, 0.00080, 0.44049, 0.28936) xa ) (0.00007, 0.00021, 0.00000, 0.96601, 0.03370) xo ) (0.00452, 0.38067, 0.00115, 0.21424, 0.39942) x ) (0.00015, 0.40505, 0.00091, 0.58031, 0.01358) xa ) (0.00000, 0.00011, 0.00000, 0.99707, 0.00282) xo ) (0.00032, 0.89332, 0.00201, 0.07781, 0.02655) 0.42590)b 0.06038 0.30095 0.54664 T (K) 379.65 388.34 396.40

Note: x ) mixture; xa ) aqueous phase; xo ) organic phase. b The sequence is IB, DIB, TIB, water, TBA. Table 3. Column Configurations and Performance IB column total stages reactive stages pressure (atm) reflux ratio reboil ratio heat duty (MW) ratio of condenser vapor to feed catalyst loading per stage (kg) feed flow rate (kmol/h) feed position TBA conversion (%) IB selectivity (%) DIB selectivity (%) IB DIB TIB Water TBA IB DIB TIB water TBA Configuration 15 3-13 1.0 5.0 1.61 0.9999 159.1 100 7 DIB column 30 2-28 4.13 infinite 6.1 2.12 0 54.5 100 14 99.59 97.49

4. Column Mathematical Model and Solution Method The studied reactive distillation column with three zones is shown in Figure 2. The reactions occur only in the reactive zone where the catalyst is present, and the feed is introduced to a reactive stage. For both IB and DIB formations, no distillate is withdrawn from the top, i.e., total liquid reflux. Instead, a vapor outlet from the partial condenser is set for IB formation; however, it is closed for DIB formation. Since the TBA dehydration studies indicate that a mass transport limitation does not exist when macroporous ionexchange resins are used as a catalyst,17 a dynamic equilibrium stage model considering mass and energy balances18,19 is applied to simulate the column. In the model, the solid catalyst is treated as quasi-homogeneous with the liquid phase, i.e., the same liquid composition exists at the catalytically active site as in the bulk liquid composition. The energy balance is considered with the mass balances. The vapor phase is assumed to be ideal at the investigated pressure, and the liquid activity coefficients are described by the Dortmund modified UNIFAC model20 as was done for the kinetics fitting. For the IB formation column, only the VLE is applied, while the model considers the possibility of liquid splitting on any stage of the column for the DIB formation column. Moreover, The pressure drop, heat losses across the column walls, and the effect of fluid dynamics are neglected. The above model is formulated as differential algebraic equations, which are solved using the DIVA simulation environment.21 DIVA contains robust solvers for dynamic and steady state simulations as well as continuation solvers for bifurcation analyses, which is especially suitable for large-scale systems such as reactive distillation processes. In the following sections, the presented model will be applied for process analyses. All the three reactions are always considered simultaneously. The effect of the important param-

Performance 99.69 99.43

Top Product Concentration (Vapor) 0.9959 <1 10-5 <1 10-6 0.0036 0.0004 Bottom Product Concentration <1 10-6 0.0029 0.1 10-4 0.9944 0.0027 <1 10-5 0.3249 0.0056 0.6668 0.0027

eters, i.e., catalyst loading and column pressure as well as the reboil ratio, is analyzed. To quantify the column performance, the conversion of TBA and the selectivities are focused on. Since the amount of DIB and TIB in the vapor phase of the condenser is always very low in simulations, the selectivities of the products IB and DIB are defined according to their amounts in the bottom as

SIB ) SDIB )

VtopyIB,top VtopyIB,top + B(2xDIB,bottom + 3xTIB,bottom) 2BxDIB,bottom VtopyIB,top + B(2xDIB,bottom + 3xTIB,bottom)

(4)

(5)

where B and Vtop are the flow rates of the bottom product and top product of vapor from the condenser, respectively. 5. Process for IB Formation 5.1. Steady State Behaviors. Due to the difference in boiling points between IB and other components, the concentration of IB in the vapor phase is always higher than that in the equilibrated liquid phase. To obtain high purity IB, the product withdrawn from the vapor phase of the condenser is designed

Figure 2. Schematic of reactive distillation column for TBA dehydration and DIB formation.

1616

Ind. Eng. Chem. Res., Vol. 45, No. 5, 2006

Figure 3. Steady state profiles in a TBA dehydration column: (a) concentration; (b) temperature.

and the liquid from condenser is totally refluxed back to the column, as shown in Figure 2. Since the duty for separating IB is light, both the rectifying and stripping sections contain only one nonreactive stage. Moreover, the ratio of the condenser vapor to the feed is close to unity (i.e., 0.9999) and the column operating pressure is 1 atm. In brief, the configurations of the reactive distillation column for IB formation are listed in Table 3. The steady state column profiles are presented in Figure 3, and the corresponding column performance is given in Table 3. As seen in Figure 3a, the IB concentration in the vapor phase (0.9959; product) is higher than that in the liquid phase (0.9623) of the condenser. TBA exists mainly inside the column (stages 2-11). In the bottom, very small amounts of byproducts are found, which hold a very low concentration in the reactive section (the TIB concentration in the whole column is very low such that one cannot distinguish in Figure 3a). From Table 3, both the TBA conversion and the IB selectivity are quite high. By checking the VLLE model, only in the liquid mixture of the condenser does phase splitting occur with a very low relative molar holdup of the aqueous phase ( ) 0.0034). Good column performance is favored by the coupled effects of reaction and distillation in one unit: (1) The coproduct water is immediately separated from the IB as it forms favoring the equilibrium of TBA dehydration (eq 1) to be shifted far to the right and, thus, high TBA conversion to be achieved under mild

Figure 4. Bifurcation diagram of the effect of the catalyst loading on IB column performance.

operating conditions. (2) Once IB is generated, it is quickly rectified to the upper part of the column, which leads to a very low IB concentration in the most reactive zone and, thus, reduces the rates of side reactions. (3) DIB is simultaneously removed from the reactive zone to the bottom, which further suppresses the formation of the trimer and higher oligomers. (4) Lower pressure results in a lower temperature profile in the column (Figure 3b), which is in favor of catalyst life. In this work, the catalyst loading per stage is 159.1 kg and the total catalyst loading in the column is 1750 kg for a feed flow rate of 100 kmol/h. Comparing these values to those of the literature with the same catalyst,8 i.e., a total of 0.05 kg catalyst and a feed flow rate of 8.928 10-4 kmol/h, the capacity of the catalyst (i.e., the ratio of the feed flow rate to the total catalyst loading) in this work is higher, i.e., 0.0571 kmol/(h kgcat) vs 0.0179 kmol/(h kgcat)). For process development, it is important to analyze the influence of the important parameters, which are the catalyst loading, the operating pressure, and the reboil ratio for this IB formation process. In the following, the influence is analyzed by changing the investigated parameter while keeping others the same as in this base case. The continuation methods in DIVA are applied. 5.2. Effect of Catalyst Loading. The catalyst loading is one of the most crucial design parameters for a reactive distillation

Ind. Eng. Chem. Res., Vol. 45, No. 5, 2006 1617

Figure 5. Bifurcation diagram of the effect of the operating pressure on IB column performance.

Figure 6. Bifurcation diagram of the effect of the reboil ratio on IB column performance.

process. By changing the catalyst loading, the bifurcation diagram of the influence of the catalyst loading per stage is generated, as shown in Figure 4. As can be seen, at lower catalyst loading (less than 100 kg/stage), though the selectivity of IB is close to unity (due to the very weak side reactions), the TBA conversion is low due to the slow reaction rate. Therefore, the unreacted TBA should be withdrawn from the bottom. At catalyst loading between 101 and 3000 kg/stage, the TBA is almost completely converted and the IB conversion is high. This wide range is suitable for column design. At higher catalyst loading, the reaction rates of side reactions are promoted such that both IB purity and selectivity are reduced significantly. 5.3. Effect of Operating Pressure. TBA dehydration to IB is an endothermic reaction (heat of reaction is about 27-38 kJ/mol),10,14,15 and the forward reaction is theoretically favored by a high operating temperature and low pressure, which can be easily achieved by conventional reactors. However, in a reactive distillation column, the temperature and pressure strongly interact with each other. Therefore, one should carefully determine a suitable operating pressure generally considering three effects: (1) reaction rate, (2) separation efficiency, and (3) limitation of catalyst like resins. For this analysis, a range of pressure between 0.5 and 8 atm is suggested and the bifurcation diagrams are presented in Figure 5. It is quite interesting that at reduced pressure (about 0.8 atm) the best column performance can be obtained with high purity

and high selectivity of IB. This is mainly attributed to two reasons: (1) the enhanced distillation effect for IB as it is more easiliy evaporated to the top of column at reduced pressure due to its very low boiling point, which leads to lower IB concentration in column and increases the TBA dehydration rate and (2) the low temperature at the reactive zone due to the reduced pressure. Both result in slower rates of the side reactions. Though a reduced pressure of 0.5-1 atm was also adopted in the literature experiments,8 it is not recommendable as the improvement is not so significant and more equipment is needed to maintain such reduced pressure. As the pressure increases, the temperature inside the column increases, which favors the side reactions and byproduct formation. As a result, both the IB purity and selectivity are decreasing. Unexpectedly, the reaction of TBA dehydration is not enhanced, though the high temperature is helpful, which implies that the effect of pressure enhancement on the volatility of IB (i.e., lower IB volatility) is stronger than the effect of temperature enhancement (due to the increasing pressure) on the reaction rate of TBA dehydration. 5.4. Effect of Reboil Ratio. The reboil ratio not only affects the column performance but also determines the heat duties of the column. As seen in Figure 6, the improvement of column performance, especially the TBA conversion (Figure 6b), is significant as the reboil ratio increases in the lower range (<3). However, at higher reboil ratios (3-8), its influence is very slight such that the column performance is stable. In other steady

1618

Ind. Eng. Chem. Res., Vol. 45, No. 5, 2006

Figure 7. Steady state profiles in a DIB formation column: (a) concentration; (b) temperature; (c) relative molar holdup of the aqueous phase; (d) reaction rates.

state studies with reboil ratios of 3-4, the column performance is much more sensitive to other parameters. Therefore, the range of the reboil ratio between 4 and 5 is suggested for IB formation. 6. Process for DIB Formation 6.1. Steady State Behaviors. Different from the IB formation column, to generate DIB, the IB formed by TBA dehydration should remain totally inside the column. Therefore, no product to be withdrawn from the top is designed for this process. Moreover, higher pressure than the IB column is adopted to increase the reaction rates. For the sake of better overall column performance, the feed is set at stage 14. The column configurations are listed in Table 3. Under such conditions, the column profiles of concentration, temperature, reaction rates, and liquid phase splitting information are illustrated in Figure 7 and the column performance is given in Table 3. As can be seen, both the TBA conversion and DIB selectivity are high (Table 3), and the temperature in the reactive zone is below the limitation of the catalyst (Figure 7b). The good column performance and the column profiles again demonstrate the advantages of reactive distillation technology for this process that couples the two reactions of TBA dehydration and DIB formation: (1) TBA dehydration is an endothermic reaction (27-38 kJ/mol), and DIB formation is a highly exothermic reaction (-82.84 kJ/mol).22 Since 2 mol of

IB form 1 mol of DIB, coupling these two reactions effectively makes use of the heats of reaction without losses and significantly reduces the heat exchanger costs. (2) The remarkable behavior is that the feed location plays an important role in the profiles. From Figure 7a, IB dominates the upper part of the column with very low concentrations of water and TBA at stages 2-7 (<0.001 for TBA and <3 10-5 for water) while the higher boiling components are condensed at the lower part. Correspondingly, the temperature below the feed stage is much higher than that at the upper part (Figure 7b). Moreover, the predicted phase splitting in most stages (Figure 7c) shows very different relative holdup of the aqueous phase, which varies from homogeneous between the condenser and stage 11 to 0.6413 in the reboiler. (3) From the reaction rate profile (Figure 7d), the TBA dehydration mainly takes place below the feed location, while the DIB formation takes places above the feed position. Due to the difference in boiling points between IB and other components, in the lower part of the column, the formed IB is immediately condensed to the upper part, leading to a very low IB concentration at the lower part. Moreover, the DIB formed at the upper part is simultaneously removed from the IBenriched region to the lower part. The formations of IB and DIB in different column regions and the condensation of IB and DIB in different directions result in a very low rate of TIB formation (Figure 7c, hard to distinguish), and thus, a high DIB selectivity is achieved.

Ind. Eng. Chem. Res., Vol. 45, No. 5, 2006 1619

Figure 8. Effect of the catalyst loading on DIB column performance.

Figure 9. Effect of the operating pressure on DIB column performance.

It is worth noting that, at stages 9-13, TBA hydration rather than TBA dehydration occurs due to the low concentration of TBA and much higher concentration of IB in this section, even though it is not recommended to feed TBA at the upper part, which can enhance the formation of TIB and, thus, reduce the DIB selectivity. It is also worth noting that, in this process, the desired catalyst loading (54.5 kg/stage and 1472 kg in total) is unexpectedly lower than that in the first studied IB formation column. The reason is that, in the upper part of the reactive distillation column for IB formation, the reaction rate is quite slow due to the relatively high concentrations of isobutene and water. Therefore, part of the catalyst does not work efficiently. If high conversion of TBA is not required, the stages above the feed position could be nonreactive such that the catalyst loading can be reduced. This aspect also illustrates the valuable advantage of reactive distillation gained by coupling multiple reactions, which has, to our best knowledge, not been studied so far. 6.2. Effect of Catalyst Loading. In the same way, the influence of catalyst loading on the column performance is analyzed by continuation methods. The bifurcation diagram is shown in Figure 8. At low catalyst loading (<50 kg/stage), both the reaction rates of TBA dehydration and DIB formation are slow, leading to low TBA conversion such that some TBA should be drawn out from the bottom product. On the other hand, the DIB selectivity is higher since the less formed IB is mostly converted to DIB. At higher catalyst loading (>50 kg/ stage), the TBA conversion is close to unity. However, the DIB selectivity is slightly lower (around 4%), which is because the

side reaction for TIB formation is enhanced at higher catalyst loading. With further increasing catalyst loading, the DIB selectivity is significantly reduced, which is not shown here. As a suggestion, the range of 50-100 kg/stage is recommended. 6.3. Effect of Pressure. In the base case for DIB formation, the pressure is 4.13 atm, which mainly considers the temperature limitation of catalyst. As illustrated in Figure 9, lower pressure leads to worse column performance due to the decreased temperature (and thus slow reaction rates) in the reactive zone. Though higher pressure (6-10 atm) has not much influence on the column performance, it is not recommendable. One reason is that more energy should be supplied as the reboiler temperature is increased. Another reason is that the temperature in the reactive zone exceeds the limitation of catalyst. 6.4. Effect of the Reboil Ratio. The bifurcation analysis of the influence of the reboil ratio on the column performance shows similar behavior (Figure 10). Since the liquid of the top is fully refluxed back to the column, the liquid and vapor flux inside the column is only determined by the reboil ratio. At low reboil ratios (<5.0), the flux inside the column is limited such that it is difficult to generate a better concentration profile. As a result, TBA will mainly remain at the lower part of the column, leading to lower conversion. Moreover, at lower reboil ratios, the potential for rectifying IB to the top becomes weak, such that the IB concentration at the lower part is higher than that in the base case; thus, the TIB formation rate is enhanced and the DIB selectivity is reduced. At higher reboil ratios (>6), TBA conversion and DIB selectivity are slightly improved. However, the improvement

1620

Ind. Eng. Chem. Res., Vol. 45, No. 5, 2006

Figure 11. Proposed alternative for DIB production.

Figure 10. Effect of the reboil ratio on DIB column performance.

becomes much easier (the difference of normal boiling points between water and isoostane is 17.6 K). As an alternative for the case of DIB as the final product, a process consisting of two reactive distillation columns is suggested (Figure 11). The dehydration of TBA (to IB) and dimerization of IB to DIB are carried out separately. Such an arrangement can avoid the separation problem of DIB from water. The concern is that the DIB selectivity might be lower to a certain extent according to the recent study on the dimerization of isobutene.23 Moreover, the reboiler heat demand will be significantly increased from 2.12 MW for one single column to 4.50 MW for two columns (the heat duty of the second column is referred to in the previous work23). The final decision of the flowsheet should be made based on the optimization considering the product quality and the overall cost. 8. Conclusions Reactive distillation favors the dehydration of TBA for selective formation of isobutene and diisobutene, and the feasibility of the two processes is illustrated through simulation studies. High purity isobutene and diisobutene with high TBA conversion can be obtained in different reactive distillations. Correspondingly, the column configurations are different for better column performance. The influence of the important parameters on column performance is analyzed by using continuation methods. On the basis of the bifurcation diagrams, suitable operating windows are suggested. The coupling of the two reactions of TBA dehydration and DIB formation in one single reactive distillation column explores new features, which have not been reported so far. This reaction system could be a good candidate for studying the complicity of reactive distillation coupling multiple reactions in one column. Acknowledgment The financial support for this work from the Volkswagen Foundation in Germany (Project Coupling of Chemical Reactions in a Reactive Distillation Process, AZ.: I/79515) is gratefully acknowledged. Nomenclature ai ) liquid activity of component i k ) reaction rate constant, mol/(h gcat) S ) selectivity

in column performance may not compensate the cost of energy required. One should balance the overall cost and the product quality based on a detailed cost estimation. 7. Discussion For process analysis, the kinetics model and activity coefficient method usually play crucial role, especially for systems studied in this work with the two strongly polar components TBA and water. Since no more suitable kinetics studies under the same conditions are valuable, further investigation of the effect of kinetics and activity models cannot be carried out. Even though, we believe that the general observations and conclusions should be the same, but the desired windows of the operating parameters might be shifted. Moreover, impurities in the crude TBA feed, most notably water, methanol, acetone, and heavies, are not considered in this work since they do not appear to significantly inhibit the dehydration process (eq 1), although the presence of methanol clearly leads to the formation of additional MTBE (either through etherification of the TBA feedstock or the isobutene coproduct). However, the desired windows for the operating parameters might be changed if they are taken into account. Furthermore, for DIB production, the proposed process leads to another problem of separating DIB from water due to their close boiling points (the difference of normal boiling points between water and two kinds of dimers is 1.5-5.0 K). However, if the DIB is further used for isooctane production by subsequent hydrogenation of DIB, the separation of isooctane from water

Ind. Eng. Chem. Res., Vol. 45, No. 5, 2006 1621

r ) rate of reaction, kmol/(h gcat) xi ) liquid molar fractions of component i yi ) vapor molar fractions of component i ) relative molar holdup of aqueous phase to overall liquid mixture AbbreViations DIB ) diisobutenes IB ) isobutene TBA ) tert-butyl alcohol TIB ) triisobutenes Literature Cited
(1) Linder, S. C. MTBE May Not Be the Only Gasoline Oxygenate You Should Be Worrying About. LUSTLine 2000, February (Bulletin 34), 18. (2) Executive Order D-52-02 by the Governor of the State of California, Mar 15, 2002. (3) Karinen, R. S.; Linnekoski, J. A.; Krause, A. O. I. Etherification of C5- and C8-Alkenes with C1- to C4-Alcohols. Catal. Lett. 2001, 76, 81. (4) Levine, R.; Olechowski, J. R. Dehydration of Tertiary Butyl Alcohol. U.S. Patent 4,165,343, Aug 21, 1979. (5) Inoue, K.; Sato, T.; Kohayashi, M. Process for Production Isobutylene. U.S. Patent 4,873,391, Oct 10, 1989. (6) Gupta, V. P. Liquid-Phase Dehydration of Tertiary Butyl Alcohol. U.S. Patent 5,625,109, Apr 29, 1997. (7) Gupta, V. P.; Pa, B. Isobutylene Recovery Process. U.S. Patent 5,436,382, Jun 25, 1995. (8) Abella, L. C.; Gaspillo, P. D.; Itoh, H.; Goto, S. Dehydration of tert-Butyl Alcohol in Reactive Distillation. J. Chem. Eng. Jpn. 1999, 32, 742. (9) Knifon, J. F.; Sanderson, J. R.; Stockton, M. E. Tert-Butanol Dehydration to Isobutylene via Reactive Distillation. Catal. Lett. 2001, 73 (1), 55. (10) Honkela, M. L.; Ouni, T.; Krause, A. O. I. Thermodynamics and Kinetics of the Dehydration of tert-Butyl Alcohol. Ind. Eng. Chem. Res. 2004, 43, 4060. (11) Honkela, M.; Krause, A. O. I. Kinetic Modeling of the Dimerization of Isobutene. Ind. Eng. Chem. Res. 2004, 43, 3251.

(12) Honkela, M.; Krause, A. O. I. Influence of Polar Components in the Dimerisation of Isobutene. Catal. Lett. 2003, 87 (Nos. 3-4), 113. (13) Knifton, J. F., Edwards J. C., Methyl tert-Butyl Ether Synthesis from tert-Butanol via Inorganic Solid Acid Catalysis. Appl. Catal., A 1999, 183, 1. (14) Velo, E.; Puigjaner, L.; Recasens, F. Inhibition by Product in the Liquid-Phase Hydration of Isobutene to tert-Butyl Alcohol: Kinetics and Equilibrium Studies. Ind. Eng. Chem. Res. 1988, 27, 2224. (15) Delion, A.; Torck, B.; Hellin, M. Equilibrium Constant for the Liquid-Phase Hydration of Isobutylene over Ion-Exchange Resin. Ind. Eng. Chem. Res. 1986, 25, 889. (16) Abella, L. C.; Gaspillo, P. D.; Maeda, M.; Goto, S. Kinetics Study on the Dehydration of tert-Butyl Alcohol Catalyzed by Ion-Exchange Resins. Int. J. Chem. Kinet. 1999, 31, 854. (17) Heath, H. W., Jr.; Gates, B. C. Mass Transport and Reaction in Sulfonic Acid Resin Catalyst: The Dehydration of tert-Butyl Alcohol. AIChE J. 1972, 18, 321. (18) Qi, Z.; Sundmacher, K.; Stein, E.; Kienle, A.; Kolah, A. Reactive Separation of Isobutene from C4 Crack Fractions by Catalytic Distillation Processes. Sep. Purif. Technol. 2002, 26, 147. (19) Qi, Z.; Kienle, A.; Stein, E.; Mohl, K. D.; Tuchlenski, A.; Sundmacher, K. MTBE Decomposition in a Reactive Distillation Column. Chem. Eng. Res. Des. 2004, 82 (A2), 185. (20) Gmehling, J.; Li, J.; Schiller, M. A Modified UNIFAC Model. 2. Present Parameter Matrix and Results for Different Thermodynamic Properties. Ind. Eng. Chem. Res. 1993, 32, 178. (21) Mangold, M.; Kienle, A.; Gilles, E. D.; Mohl, K. D. Nonlinear Computation in DIVAsMethod and Applications. Chem. Eng. Sci. 2000, 55, 441. (22) Marchionna, M.; Di Girolamo, M.; Patrini, R. Light Olefin Dimerization to High Quality Gasoline Components. Catal. Today 2001, 65, 397. (23) Kamath, R.; Qi, Z.; Sundmacher, K.; Aghalayam, P.; Mahajani, S. Process Analysis of Dimerization of Isobutene in a Reactive Distillation Column. Ind. Eng. Chem. Res., in press, 2005.

ReceiVed for reView October 3, 2005 ReVised manuscript receiVed December 14, 2005 Accepted December 21, 2005 IE0511027

Das könnte Ihnen auch gefallen