Vous êtes sur la page 1sur 8

Article pubs.acs.

org/IECR

Alkylation Kinetics of Isobutane by C4 Olens Using Sulfuric Acid as Catalyst


Weizhen Sun, Yi Shi, Jie Chen, Zhenhao Xi, and Ling Zhao*
State-Key Laboratory of Chemical Engineering, East China University of Science and Technology, Shanghai 200237, China
S Supporting Information *

ABSTRACT: The alkylation kinetics of isobutane with butene using sulfuric acid as catalyst was investigated by batch experiments in the conditions of industrial interest. More than 16 alkylates were identied and quantied by GC-MS. On the basis of the classic carbonium ion mechanism, the kinetic model was established, which can predict the concentration change of three groups of key alkylates including trimethylpentanes (TMPs), undesirable dimethylhexanes (DMHs), and heavy ends (HEs). The agreement between experimental and model calculated data was quite satisfactory. The rate constants were found to be constant with the varied temperatures (276.2 to 285.2 K) except those accounting for the addition of H+ to isobutene and its reversible reaction. An anti-Arrhenius behavior was observed for the addition reaction of H+ to isobutene, in which the corresponding rate constant falls with the increasing temperatures. The kinetic model was conrmed by the simulation of the industrial alkylation reactor. Hopefully, the kinetic model developed in this work will be useful to the design and optimization of novel alkylation reactors. isobutane with olens, much classical literature had been published,711 in which the formation pathway of the majority of key components and intermediates in alkylate were well formulated although dierent publications would have dierent descriptions regarding some specic steps. However, to the best of our knowledge, there was little literature focusing on the alkylation kinetics of isobutane with butene in sulfuric acid to address the formation of several key alkylates such as trimethylpentanes (TMPs), dimethylhexanes (DMHs), and heavy ends (HEs). Using uniform hydrocarbon drops and short contact time, the sulfuric acid catalyzed reaction of isobutane with 1-butene, as well as the oligomerization of 1-butene, was investigated, in which only two key reactions were considered,12 but other important reaction steps in view of the presence of several dozen isoparans in commercial produced alkylates are far from being understood.13 The two-step process for the alkylation of isobutane with C4 olens was investigated, but the rate constants were not given.10 In this work, the alkylation kinetics of isobutane with butene using sulfuric acid as catalyst was measured under the condition of industrial interest. The kinetic model was established on the basis of the carbonium ion mechanism, in which three families of key isoparans were considered including TMPs, DMHs, and HEs.

1. INTRODUCTION In early 1990s, the USA reners had to start changing their strategy on gasoline composition to meet the mandatory CAA (Clean Air Act) specications.1 Since that time, gasoline was forced to move in a more environmentally friendly direction, such as reducing volatility, limitations in the aromatic content, increased amount of oxygenates, reduction of olens and sulfur, and elimination of lead. As a desirable blending component, without olens or aromatics, alkylate exclusively contains isoalkanes with high octane number. Since its commercial production in the last century, alkylate had been the most ideal blending component for a typical renery gasoline pool. Nearly 70% of the worlds alkylate production is from North America, and over 20% is produced from Europe.2 It was believed that alkylate will continue to be a desirable blending component as long as cars are operated on high octane gasoline.3 The alkylation reaction producing alkylates combines isobutane with light C3C5 olens in the presence of a strong acid catalyst. Currently, the only processes of commercial importance use either sulfuric or hydrouoric acid as catalysts.4 Although the number of established alkylation units using sulfuric acid is almost as much as that using hydrouoric acid, more new alkylation plants built worldwide chose sulfuric acid as catalyst in the recent past years.5 One of the reasons is that hydrouoric acid is a highly toxic liquid, and released into the atmosphere, it forms aerosol, which drifts downwind for several kilometers.6 Actually, both acids suer from certain drawbacks, but it is not the intent of this work to review in detail the pros and cons of sulfuric acid versus hydrouoric acid but to address the alkylation kinetics using sulfuric acid as a catalyst. It is well-known that the study on reaction kinetics is of fundamental importance not only to the design and optimization of a ripe chemical reactor but also to the development of a novel one. Also, it is helpful in understanding the reaction mechanism. As to the alkylation mechanism of
2013 American Chemical Society

2. EXPERIMENTAL SECTION The experiments were carried out in a batch setup with a glass reactor with the volume of 1 L. To keep isobutane and olen in
Special Issue: NASCRE 3 Received: Revised: Accepted: Published:
15262

February 4, 2013 April 2, 2013 April 3, 2013 April 3, 2013


dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. 2013, 52, 1526215269

Industrial & Engineering Chemistry Research

Article

Figure 1. Schematic diagram of experimental setup: 1, cylinder for mixed C4; 2, cylinder for N2; 3, dryer; 4, lter; 5, mass ow meter for liquid; 6, refrigerator; 7, mass ow meter for gas; 8, low temperature salt water; 9, reactor; 10, baes; 11, sampling cylinder.

liquid phase, the operating pressure was set to be 0.5 MPa. One refrigerant system was employed to obtain cycling salt water, which was used to keep the entire reactor and the hydrocarbons into and out of the reactor at constant temperature ranging from 276.2 to 285.2 K. Inside, the reactor baes and stirring apparatus were installed, which can be adjusted within 3000 rpm. For a typical experiment, rst, the weighed sulfuric acid was put into the reactor, and then, the nitrogen (N2) was introduced to remove air in the reactor, after which the pressure was set to be 0.5 MPa. Next, the refrigerant system started to work to keep the material inside the reactor at the set temperature. When the temperature inside the reactor reached the set temperature, the weighed mixture of isobutane and olen with certain molar ratio was quickly introduced into the bottom of the reactor. Almost at the same time, the agitator started to work to ensure the good dispersing of hydrocarbon upon contacting the sulfuric acid. According to the alkylation reaction progress, the sampling interval was shorter initially and then longer in the later period. In all of the experiments, the temperature inside the reactor was recorded and controlled to ensure it was within a set range. The schematic diagram of the experimental setup was shown in Figure 1. The gas chromatographymass spectrometry (GC-MS) was adopted to identify and quantify more than 16 alkylate components. A typical chromatogram of alkylate components is shown in Figure 2. The area normalization method was employed in this work, since all of the correction factors of

dierent components are very close to 1.0 with respect to the same standard substance benzene on FID detector. The maximum relative analysis deviation was less than 5%, which shown in Table 1. More detailed descriptions about GC-MS can be found in the Supporting Information. Table 1. Reproducibility of Analysis Method
mass content, % number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 alkylate components isopentane 2,3dimethylbutane 3-methylpentane 2,4dimethylpentane 2,2,3trimethylbutane 2-methylhexane 2,3dimethylpentane 3-methylhexane isooctane 2,5dimethylhexane 2,2,3trimethylpentane 2,4dimethylhexane 2,3,4trimethylpentane 3,3dimethylhexane 2,3dimethylhexane 2,2,5trimethylhexane sample 1 sample 2 1.491 3.475 0.427 2.936 0.348 0.207 1.744 0.152 21.843 4.054 2.395 1.100 9.507 12.174 0.272 10.083 1.374 3.286 0.403 2.758 0.345 0.227 1.692 0.162 21.348 3.982 2.377 1.105 9.276 11.904 0.287 9.976 average, % 1.433 3.380 0.415 2.847 0.347 0.217 1.718 0.157 21.595 4.018 2.386 1.102 9.391 12.039 0.280 10.030 relative deviation 0.041 0.028 0.030 0.031 0.004 0.046 0.015 0.029 0.011 0.009 0.004 0.002 0.012 0.011 0.027 0.005

Figure 2. Chromatogram of all alkylate components.


15263

The intrinsic chemical reaction of alkylation is very fast, leading to a strong mass transfer limitation of olen into the reaction region when the mass transfer rate is not big enough.14 The experiments showed that when exceeding above 2500 rpm the increasing stirring speeds have no inuence on the alkylation rate, which demonstrates the elimination of mass
dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. 2013, 52, 1526215269

Industrial & Engineering Chemistry Research

Article

Figure 3. Diagram of the reaction pathways network.

transfer limitation. Thus, all of the investigations of alkylation kinetics with varied temperatures were performed with the stirring speeds of 2800 rpm.

TMP+ + iC4 = TMP + C=C(CH3)C+


C=C(CH3)C+ + iC4 = C=C(CH3)CC(CH3)C+C

(6)

3. KINETIC MODEL It is well accepted that the alkylation of isobutane with olens follows the classic carbonium ion mechanism. Initially the tertbutyl cation is produced by this reaction:
iC4 = + H+ iC4 +
k1

(7)

(1)

Through suitable hydrogen and proton transfer reactions, the unsaturated cation ion forms a DMH molecule. Another possible pathway to give DMHs is through the addition of 1butene to iC4+,11
iC4 + + 1butene DMHs+
k5

where is isobutene and represents tert-butyl cation. Actually, regardless of whether the feed butene is either 1butene or 2-butene, it tends to give iC4= through fast isomerization, which occurs especially in the presence of sulfuric acid.15 The transform between dierent butenes proceeds by the following equilibrium reaction,
1 C4 H8 2 C4 H8 iC4 =
K1 K2

iC4=

iC4+

(8)

Similar to TMPs+, the DMHs+ are saturated through the hydride abstraction from isobutane, giving an iC4+ meanwhile,
DMHs+ + iC4 DMHs + iC4 +
k6

(9)

(2)

Interestingly, it is found that even when isobutane alkylated with olens rather than butenes, the trimethylpentanes are still generated by so-called self-alkylation. The key step for selfalkylation is the formation of isobutene, which in fact is a reversible reaction of the reaction 1:1
iC4 + iC4 = + H+
k2

According to the experimental data in this work, it was veried that the formation of DMHs is most likely through the latter pathway, that is, the addition of 1-butene to iC4+. By reactions involving olens, heavy ends (HEs) are produced, which are primarily polymerization reactions.15 The polymerization of C4 olens proceeds by the addition of iC4= or 2-butene to TMPs+ or DMHs+,
TMPs+ (or DMHs+) + iC4 = (or 2butene) iC12+
(10)

(3)

The tert-butyl cation adds to an olen to produce the corresponding C8 carbonium cation. For example, when the tert-butyl cation reacts to iC4= or 2-butene, the trimethylpentanes will be formed,
iC4 + + 2butene (or iC4 =) TMPs+
k3

(4)

The iC12 would further add iC4 or 2-butene to give iC16+. For simplication, in this work, we regard iC12+ and iC16+ as the pseudo components iCm+. When these polymer cations iCm+ are contacted with isobutane, the HEs is formed, along with the production of iC4+,
iCm+ + iC4 iCm + iC4 +
k7

where TMPs is the abbreviation of trimethylpentanes. These C8 carbonium cations tend to isomerize via methyl shifts and hydride transfer reaction to form several isomer cations.4 For simplication, in this work, all these C8 cations are regarded as one pseudo component TMP+. By rapid hydride transfer with isobutane, each TMP+ changes to the corresponding TMP component, regenerating the iC4+ to keep on the chain propagation.
TMP+ + iC4 TMP + iC4 +
k4

(11)

As one of the ve groups in alkylated hydrocarbon products, the light ends (LEs) include C5C7 isoparans. The production of LEs is believed to come from the fragmentation of large isoalkyl cations.9 Due to the strong oxidizing capacity of concentrated sulfuric acid, the isoalkyl cation will be generated again from HEs as follows,
iCm + H+ iCm+
(12)

(5)

where iC4 stands for isobutane. The source of undesirable dimethylhexanes (DMHs) with low octane numbers has several possible pathways. One was believed from the addition of iC4= and allylic carbocation, which is the result of a hydride ion abstraction from the allylic position of an olen molecule.10
15264

These heavy isoalkyl cations are fragmentated following the -scission rule, leading to smaller isoalkyl cations and olens.1

iCm+ iCx + + iCy =


k9

k8

(13)

dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. 2013, 52, 1526215269

Industrial & Engineering Chemistry Research

Article

Figure 4. Concentration prole of key components in alkylation of isobutane with butene. Temperature: (a) 276.2 K, (b) 279.2 K, (c) 282.2 K, (d) 285.2 K. Symbols: experimental data; Line: calculated values by kinetic model.

Isoalkyl cations generated by fragmentation (iCx+) and protonation of olens (iCy=) are thought to be the major precursors of LEs. Through the hydride ion transfer, these isoalkyl cations change into the corresponding LEs. According to the experimental observation in this work, the degradation of DMH was found to be obvious. It may be ascribed to the formation of DMH+ from the oxidation of DMH, which is further fragmentated into small LE through scission reaction.16 To simplify this multistage reaction, one single reaction step is used in this work,
DMH LE
k10

c1C4H8 = *cC4H8
c 2C4H8 = *cC4H8
ciC4= = (1 )*cC4H8

(15) (16) (17)

where and represent 1/(K1 + K1*K2 + 1), K1/(K1 + K1*K2 + 1) respectively. Following the above reaction steps and assumptions, the kinetic model can be formulated readily as follows,
dc1 = (1 )k1c1 (1 )k3c1c3 k5c1c3 dt (1 )k7c1c 2c4 + k2c 3
(18)

(14)

The diagram of the whole reaction pathway network was shown in Figure 3. The alkylation products are composed of more than two dozen branchedchain paranic hydrocarbons. Both TMPs and DMHs have multiple isomers. HE, as well as LE, also refers to one family of isoparans with close molecular weight. Including those carbonium components, there will be more than three dozen species in this system. However, only three or four groups of components can be measured due to the limit of analysis method and the faster transform between various species. To avoid the over t or over parametrization, the number of involved species and pathway should be simplied as much as possible. So rst, some species with similar properties are treated as one pseudo component, such as TMPs (or DMHs) mentioned above. HE, as well as LE, is regarded as one single component, respectively, because of close molecular weight. Second, some steps are assumed to be instantaneous, such as reactions 10 and 12, and the reaction 2 is assumed to be under chemical equilibrium control. Accordingly, the concentration distribution for these three butenes can be calculated, respectively, on the basis of the total concentration of butene,
15265

dc 2 = k4c 2c4 k6c 2c5 (1 )k7c1c 2c4 dt


dc 3 = (1 )k1c1 + k4c 2c4 + k6c 2c5 k4c1c3 dt k5c1c3 k2c 3 + (1 )k 7c1c 2c4
dc4 = (1 )k3c1c3 k4c 2c4 (1 )k7c1c 2c4 dt dc5 = k5c1c3 k6c 2c5 dt

(19)

(20)

(21)

(22)

dc6 = k4c 2c4 dt


dc 7 = k6c 2c5 k10c 7 dt

(23)

(24)

dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. 2013, 52, 1526215269

Industrial & Engineering Chemistry Research


dc 8 = (1 )k7c1c 2c4 + k9c 9c10 k8c 8 dt
dc 9 dt = k8c 8 k9c 9c10
(25)

Article

(26)

dc10 = k8c 8 k9c 9c10 dt

(27)

These ordinary dierential equations (ODEs) are completed with specications of the following initial conditions at time t = 0,
c1 = c10 ; c6 = 0; c2 = c20; c 7 = 0; c3 = 0; c4 = 0; c5 = 0; c10 = 0
(28)

c8 = 0; c 9 = 0;

In eqs 18 to 28, the relationship between the numbers of 1 through 10 and the corresponding species is as follows: 1, butene; 2, iC4; 3, iC4+; 4, TMP+; 5, DMH+; 6, TMP; 7, DMH; 8, HE; 9, iCx+; 10, iCy=.

4. RESULTS AND DISCUSSION The concentration proles of three groups of key components in alkylation of isobutane with butene were plotted in Figure 4. It shows that all of three species have dramatic changes within 2 min. Initially, the HEs go up sharply and then come down shortly. This demonstrates that the alkylation of isobutane with butene is a very fast reaction. Almost after 5 min, each of the species reaches to a platform individually, although the DMHs still continue to go down slightly. To t the experiments and estimate k1 through k10 in eqs 1828, the nonlinear least-squares tting method was employed by the following object function,
m

obj =

(ciexp cical)2
i=1

(29)

where and represent the experimental and calculated values of ith component, respectively, and m is the number of experimental data. An lsnonlin function in Matlab was used for the regress of the set of k1 through k10. Equations 1828 were resolved by the RungeKutta method in each iteration. Before the tting was performed, the values of and should be determined rst; that is, the equilibrium constants in reaction 2 should be computed. According to thermodynamics, K1 and K2 are the functions of temperature and can be calculated on the basis of the Gibbs free energy variation of reaction 2. As we know, the ASPEN Plus software is a useful tool in chemical process modeling and has a reliable thermodynamic database. The RGibbs module in ASPEN Plus was used to obtain the average values of K1 and K2 at the temperatures between 276.2 and 285.2 K, which is 28.39 and 6.433, respectively. Accordingly, the values of and can be calculated to be 0.004718 and 0.1340, respectively. Preliminary tting results show that with the variation of temperature most of the rate constants change a little, except k1 and k2. It indicates that those reaction steps corresponding to the rate constants of k3 through k10 have small or even no activation energy. A similar nding has been reported by Langley et al. in the alkylation kinetics of isobutane with propylene.17 In this work, these rate constants including k3 through k10 are assumed to be constant with varied temperatures in the range from 276.2 to 285.2 K. One of the advantages of such an assumption is to reduce the adjustable
15266

cexp i

ccal i

parameters and make the estimated rate constants more denite. So, in the retting of experimental data, the values of k3 through k10 are kept xed with temperatures, and the only adjustable parameters are k1 and k2. The tting results are shown in Figure 4, compared with the experiments. It can be seen that the agreement between experimental and calculated data is quite satisfactory. In particular, the dramatic change of HEs in short time is captured successfully by the model. The estimated k1 through k10 are listed in Table 2 with 95% condence intervals. We can see that most of the condence intervals are one magnitude smaller than the corresponding parameters except k3 and k6, which conrms the reliability of the estimated rate constants. However, an anti-Arrhenius behavior can be found for k1; that is, the value of k1 falls as the temperature increases. In fact, this phenomenon is not unusual or abnormal for hydrocarbon reaction with the presence of free radical molecules.18 The observations of antiArrhenius kinetics are usually ascribed to an entropic contribution to the free energy of activation. This antiArrhenius behavior is believed to be caused by a decrease of the equilibrium constant of a multistep reaction.19 The Arrhenius relationships of k1 and k2 are obtained by plotting ln(ki) against 1/T shown in Figures 5 and 6, respectively, both of which have good linear relationships. The activation energies of k1 and k2 are computed to be 58.3 and 173.2 kJ/mol, and the corresponding pre-exponential factors are 5.23 1011 and 6.87 1031 min1 respectively. According to the fundamental principle of physical chemistry, in a reversible reaction, the dierence of activation energy between the forward and backward reaction should be equal to the heat of the reaction. So, in this sense, the heat of reaction for the addition of H+ to the iC4= can be calculated to be 231.5 kJ/mol, which indicates that this is a strong exothermic reaction. Using uniform hydrocarbon drops, L. Lee et al. studied the sulfuric acid catalyzed reaction of isobutane with 1-butene and the 1-butene oligomerization.12 Kinetic constants for a simplied model were calculated, in which only the primary reactions were considered. When using sulfuric acid of 98%, the rate constants for alkylation of isobutane with 1-butene and 1butene oligomerization at 298 K were calculated to be 2.25 108 and 9.5 108 cm3/mol/s, respectively. However, in this work, there are no such reaction steps exactly corresponding to those in the literature.12 This is because the kinetic model in this work was developed on the basis of the carbonium ion reaction mechanism which used elementary steps, while L. Lee et al. dealt with the alkylation and oligomerization reactions by the empirical method on the basis of power exponent. Even so, it is found that one elementary step in our model is similar to the alkylation reaction of isobutane with 1-butene, which is the addition of 1-butene to iC4+ (eq 8). The rate constant for this step was estimated to be 1.26 104 kg/mol/min, which equals 1.17 105 cm3/mol/s. It seems much smaller than the alkylation reaction rate constant of isobutane with 1-butene reported by L. Lee et al., but it should be noted that our experiments were carried out at the temperatures from 276.2 to 285.2 K, which are 13 to 22 K smaller than that reported by L. Lee et al. (298 K). According to L. Albright et al.s work,10 at the temperature of 259 K, the alkylation reaction rate slowed down dramatically, in which hour was used as the unit of time. Thus, taking into consideration the sensibility of alkylation reaction to temperature, the dierence of the rate
dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. 2013, 52, 1526215269

Industrial & Engineering Chemistry Research


2.61 0.85

Article

k10, 102 k9 k8

1 1 1

min1 kgmol min min

0.63 0.06

5.48 0.82

Figure 5. Arrhenius relationship between k1 and 1/T.

kg mol min
2

kgmol min kgmol min

1.26 0.42

4.36 2.48

k6, 103

88.46 6.78

k7

k5, 104

Figure 6. Arrhenius relationship between k2 and 1/T.

Table 2. Estimated Rate Constants with 95% Condence Intervals

k1

min

6.37 5.06 4.15 2.79

0.72 0.57 0.50 0.71

constant of isobutane alkylation reaction with 1-butene between L. Lee et al.s and ours should be reasonable. In L. Albright et al.s work,10 an S-shaped curve of alkylation yield versus time for a batch run can be noticed. The author believed that the increasing rates of alkylate formation during the initial stages were ascribed to the increasing acidity of the acid phase because acid was regenerated, but this S-shaped curve can be easily understood if we illustrate this phenomenon from the point of view of reaction mechanism. In this work, eq 1 is a chain initiation step. By the subsequent addition of iC4+ and olen, the precursor of alkylates are generated. Thus, in the initial stage, the alkylation rate increases due to the continuous accumulation of the concentration of iC4+. At the last period, the alkylation rate slows down with the decrease of olen concentration. This is similar to the free radical chain reaction of hydrocarbon, in which the existence of induction period is very common in the oxidation of hydrocarbon, especially at low temperature. To conrm the reliability of the kinetic model developed in this work, we simulated the alkylation reactor operated at industrial conditions. Table 3 listed some basic conditions of the typical industrial alkylation reactor. The industrial alkylation Table 3. Operating Conditions for Industrial Alkylation Reactor
A/H, m3/m3a I/O, mol/molb 8:1 temperature, K 279 pressure, MPa 0.45 1.1:1
a

kgmol min kgmol min min K

k2 temperature

276.2 279.2 282.2 285.2

0.15 0.45 1.00 1.59

0.06 0.20 0.38 0.56

8.33 1.86

k3, 102

1.03 0.22

k4

Ratio of sulfuric acid to hydrocarbon (isobutane + olen) based on volume. bRatio of isobutane to olen based on mole.
15267
dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. 2013, 52, 1526215269

Industrial & Engineering Chemistry Research reactor of isobutane with olen is a well mixed ow reactor.1 For this type of reactor, the material balance equation can be written as follows:
xi 0 xi = *( ri)
(30)

Article

where and xi stand for the inlet and outlet concentration of ith component in alkylates with the unit of mol/kg and and ri represent the mean residence time of sulfuric acid in reactor and formation or consumption rate of the ith component with the unit of min and mol/kg/min. In the eq 30, ri is determined by the kinetic model in this work and xi0 is calculated through the operating conditions listed in Table 3. Thus, xi can be solved only if is given. The conversion of olen and alkylate yield versus mean residence time is shown in Figures 7 and 8, respectively. It can

xi0

5. CONCLUSIONS The alkylation kinetics of isobutane with butene was measured in the condition of industrial interest. The kinetic model was established on the basis of the classic carbonium ion mechanism. The agreement between experiments and model tting is quite satisfactory. Except the addition of H+ to isobutene and its reversible reaction, other reaction steps rate constants were found not to change with the varied temperatures from 276.2 to 285.2 K. An anti-Arrhenius behavior was observed for the addition of H+ to isobutene. Both of them also can be found in other hydrocarbon reactions. By the kinetic model, the successful simulation of industrial alkylation reactor was achieved. We hope that the alkylation kinetics reported in this work will be useful for the reactor design and optimization of isobutane alkylation with butene.

ASSOCIATED CONTENT

S Supporting Information *

Mass spectrogram of key alkylates and retention time of alkylate products in GC. This information is available free of charge via the Internet at http://pubs.acs.org/.

AUTHOR INFORMATION

Corresponding Author

*Tel.: +86 21 64253175. Fax: +86 21 64253528. E-mail: zhaoling@ecust.edu.cn.


Notes
Figure 7. Conversion of olen versus mean residence time.

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS The authors gratefully acknowledge the nancial support from the Joint Funds of the National Natural Science Foundation of China (U1162202), the Fundamental Research Funds for the Central Universities, the China Postdoctoral Science Foundation (2012M512055), and the 111 Project (B08021). REFERENCES

Figure 8. Alkylate yield versus mean residence time.

be seen that the conversion of olen will be greater than 99.5% when the mean residence time () is above 10 min, but the yield of alkylate is under 200% even when reaches 20 min. Usually, the alkylate yield is required to be more than 200%, so the should be kept greater than about 21 min according to Figure 8. It also indicates that when exceeds 30 min the alkylate yield will approach its theoretical value (203.6%). It was reported that in the industrial unit the mean residence time ranges from 20 to 30 min.1,20 Clearly, the simulation results show that the kinetic model developed in this work can predict the mean residence time of the industrial alkylation reactor successfully. As we all know, the mean residence time determines the reactor size when the scale of production is known and vice versa. Thus, using the kinetic model developed in this work, we will be allowed to design and optimize the novel alkylation reactors.
15268

(1) Corma, A.; Martnez, A. Chemistry, catalysts, and processes for isoparaffinolefin alkylation: Actual situation and future trends. Catal. Rev.: Sci. Eng. 1993, 35 (4), 483570. (2) Shorey, S. W. In Motor-Fuel Alkylation with CDAlky Process Technology. 103rd NPRA Annual Meeting, San Francisco, CA, March 1315, 2005; pp 116. (3) Hommeltoft, S. I. Isobutane alkylation: Recent developments and future perspectives. Appl. Catal., A: Gen. 2001, 221 (1), 421428. (4) Kranz, K. Intro to Alkylation Chemistry: Mechanisms, operating variables, and olen interactions; DuPont STRATCO Clean Fuel Technology: Leawood, 2008; pp 129. (5) Albright, L. F. Alkylation of isobutane with C3C5 olefins: Feedstock consumption, acid usage, and alkylate quality for different processes. Ind. Eng. Chem. Res. 2002, 41 (23), 56275631. (6) Feller, A.; Zuazo, I.; Guzman, A.; Barth, J. O.; Lercher, J. A. Common mechanistic aspects of liquid and solid acid catalyzed alkylation of isobutane with n-butene. J. Catal. 2003, 216 (12), 313 323. (7) Schmerling, L. The mechanism of the alkylation of paraffins. II. Alkylation of isobutane with propene, 1-butene and 2-butene. J. Am. Chem. Soc. 1946, 68 (2), 275280. (8) Schmerling, L. Reactions of hydrocarbons: Ionic mechansims. Ind. Eng. Chem. 1953, 45 (7), 14471455. (9) Hofmann, J. E.; Schriesheim, A. Ionic reactions occurring during sulfuric acid catalyzed alkylation. II. Alkylation of isobutane with C14labeled butenes. J. Am. Chem. Soc. 1962, 84 (6), 957961.
dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. 2013, 52, 1526215269

Industrial & Engineering Chemistry Research


(10) Albright, L. F.; Spalding, M. A.; Faunce, J.; Eckert, R. E. Alkylation of isobutane with C4 olefins. 3. Two-step process using sulfuric acid as catalyst. Ind. Eng. Chem. Res. 1988, 27 (3), 391397. (11) Albright, L. F.; Li, K. W. Alkylation of isobutane with light olefins using sulfuric acid. Reaction mechanism and comparison with HF alkylation. Ind. Eng. Chem. Process Des. Dev. 1970, 9 (3), 447454. (12) Lee, L.-m.; Harriott, P. The kinetics of isobutane alkylation in sulfuric acid. Ind. Eng. Chem. Process Des. Dev. 1977, 16 (3), 282285. (13) Albright, L. F.; Wood, K. V. Alkylation of isobutane with C3-C4 olefins: Identification and chemistry of heavy-end production. Ind. Eng. Chem. Res. 1997, 36 (6), 21102120. (14) Aschauer, S. J.; Jess, A. Effective and intrinsic kinetics of the twophase alkylation of i-paraffins with olefins using chloroaluminate ionic liquids as catalyst. Ind. Eng. Chem. Res. 2012, 51 (50), 1628816298. (15) Albright, L. F. Mechanism for alkylation of isobutane with light olens. In Industrial and Laboratory Alkylations; Albright, L. F., Goldsby, A. R., Eds. ACS: Washington, DC, 1977; Vol. 55, pp 128 146. (16) Doshi, B.; Albright, L. F. Degradation and isomerization reactions occurring during alkylation of lsobutane with light olefins. Ind. Eng. Chem. Process Des. Dev. 1976, 15 (1), 5360. (17) Langley, J. R.; Pike, R. W. The kinetics of alkylation of isobutane with propylene. AIChE J. 1972, 18 (4), 698705. (18) Yucel, I. Atmospheric Model Applications, 1st ed.; InTech: Rijeka, 2012; p 306. (19) Lebedeva, N. V.; Nese, A.; Sun, F. C.; Matyjaszewski, K.; Sheiko, S. S. Anti-Arrhenius cleavage of covalent bonds in bottlebrush macromolecules on substrate. Proc. Natl. Acad. Sci. U.S.A. 2012, 109 (24), 927680. (20) Albright, L. F. Alkylation of isobutane with C3-C5 olefins to produce high-quality gasolines: Physicochemical sequence of events. Ind. Eng. Chem. Res. 2003, 42 (19), 42834289.

Article

15269

dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. 2013, 52, 1526215269

Vous aimerez peut-être aussi