Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Article

pubs.acs.org/IECR

Alkylation Kinetics of Isobutane by C4 Olefins 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 identified and quantified 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 confirmed 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.

1. INTRODUCTION isobutane with olefins, much classical literature had been


In early 1990s, the USA refiners had to start changing their published,7−11 in which the formation pathway of the majority
strategy on gasoline composition to meet the mandatory CAA of key components and intermediates in alkylate were well
(Clean Air Act) specifications.1 Since that time, gasoline was formulated although different publications would have different
forced to move in a more environmentally friendly direction, descriptions regarding some specific steps. However, to the best
such as reducing volatility, limitations in the aromatic content, of our knowledge, there was little literature focusing on the
increased amount of oxygenates, reduction of olefins and sulfur, alkylation kinetics of isobutane with butene in sulfuric acid to
and elimination of lead. As a desirable blending component, address the formation of several key alkylates such as
without olefins or aromatics, alkylate exclusively contains trimethylpentanes (TMPs), dimethylhexanes (DMHs), and
isoalkanes with high octane number. Since its commercial heavy ends (HEs). Using uniform hydrocarbon drops and short
production in the last century, alkylate had been the most ideal contact time, the sulfuric acid catalyzed reaction of isobutane
blending component for a typical refinery gasoline pool. Nearly with 1-butene, as well as the oligomerization of 1-butene, was
70% of the world’s alkylate production is from North America, investigated, in which only two key reactions were consid-
and over 20% is produced from Europe.2 It was believed that ered,12 but other important reaction steps in view of the
alkylate will continue to be a desirable blending component as presence of several dozen isoparaffins in commercial produced
long as cars are operated on high octane gasoline.3 alkylates are far from being understood.13 The two-step process
The alkylation reaction producing alkylates combines for the alkylation of isobutane with C4 olefins was investigated,
isobutane with light C3−C5 olefins in the presence of a strong but the rate constants were not given.10
acid catalyst. Currently, the only processes of commercial In this work, the alkylation kinetics of isobutane with butene
importance use either sulfuric or hydrofluoric acid as catalysts.4 using sulfuric acid as catalyst was measured under the condition
Although the number of established alkylation units using of industrial interest. The kinetic model was established on the
sulfuric acid is almost as much as that using hydrofluoric acid, basis of the carbonium ion mechanism, in which three families
more new alkylation plants built worldwide chose sulfuric acid of key isoparaffins were considered including TMPs, DMHs,
as catalyst in the recent past years.5 One of the reasons is that and HEs.
hydrofluoric acid is a highly toxic liquid, and released into the
atmosphere, it forms aerosol, which drifts downwind for several 2. EXPERIMENTAL SECTION
kilometers.6 Actually, both acids suffer from certain drawbacks, The experiments were carried out in a batch setup with a glass
but it is not the intent of this work to review in detail the pros reactor with the volume of 1 L. To keep isobutane and olefin in
and cons of sulfuric acid versus hydrofluoric acid but to address
the alkylation kinetics using sulfuric acid as a catalyst.
Special Issue: NASCRE 3
It is well-known that the study on reaction kinetics is of
fundamental importance not only to the design and Received: February 4, 2013
optimization of a ripe chemical reactor but also to the Revised: April 2, 2013
development of a novel one. Also, it is helpful in understanding Accepted: April 3, 2013
the reaction mechanism. As to the alkylation mechanism of

© XXXX American Chemical Society A dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

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

liquid phase, the operating pressure was set to be 0.5 MPa. One different components are very close to 1.0 with respect to the
refrigerant system was employed to obtain cycling salt water, same standard substance benzene on FID detector. The
which was used to keep the entire reactor and the hydrocarbons maximum relative analysis deviation was less than 5%, which
into and out of the reactor at constant temperature ranging shown in Table 1. More detailed descriptions about GC-MS
from 276.2 to 285.2 K. Inside, the reactor baffles and stirring can be found in the Supporting Information.
apparatus were installed, which can be adjusted within 3000
rpm. For a typical experiment, first, the weighed sulfuric acid Table 1. Reproducibility of Analysis Method
was put into the reactor, and then, the nitrogen (N2) was
mass content, %
introduced to remove air in the reactor, after which the
pressure was set to be 0.5 MPa. Next, the refrigerant system alkylate relative
number components sample 1 sample 2 average, % deviation
started to work to keep the material inside the reactor at the set
temperature. When the temperature inside the reactor reached 1 isopentane 1.491 1.374 1.433 0.041
the set temperature, the weighed mixture of isobutane and 2 2,3- 3.475 3.286 3.380 0.028
dimethylbutane
olefin with certain molar ratio was quickly introduced into the 3 3-methylpentane 0.427 0.403 0.415 0.030
bottom of the reactor. Almost at the same time, the agitator 4 2,4- 2.936 2.758 2.847 0.031
started to work to ensure the good dispersing of hydrocarbon dimethylpentane
upon contacting the sulfuric acid. According to the alkylation 5 2,2,3- 0.348 0.345 0.347 0.004
reaction progress, the sampling interval was shorter initially and trimethylbutane
then longer in the later period. In all of the experiments, the 6 2-methylhexane 0.207 0.227 0.217 −0.046
7 2,3- 1.744 1.692 1.718 0.015
temperature inside the reactor was recorded and controlled to dimethylpentane
ensure it was within a set range. The schematic diagram of the 8 3-methylhexane 0.152 0.162 0.157 −0.029
experimental setup was shown in Figure 1. 9 isooctane 21.843 21.348 21.595 0.011
The gas chromatography−mass spectrometry (GC-MS) was 10 2,5- 4.054 3.982 4.018 0.009
adopted to identify and quantify more than 16 alkylate dimethylhexane
components. A typical chromatogram of alkylate components 11 2,2,3- 2.395 2.377 2.386 0.004
trimethylpentane
is shown in Figure 2. The area normalization method was
12 2,4- 1.100 1.105 1.102 −0.002
employed in this work, since all of the correction factors of dimethylhexane
13 2,3,4- 9.507 9.276 9.391 0.012
trimethylpentane
14 3,3- 12.174 11.904 12.039 0.011
dimethylhexane
15 2,3- 0.272 0.287 0.280 −0.027
dimethylhexane
16 2,2,5- 10.083 9.976 10.030 0.005
trimethylhexane

The intrinsic chemical reaction of alkylation is very fast,


leading to a strong mass transfer limitation of olefin 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 influence on the
Figure 2. Chromatogram of all alkylate components. alkylation rate, which demonstrates the elimination of mass
B dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 3. Diagram of the reaction pathways network.

transfer limitation. Thus, all of the investigations of alkylation TMP+ + iC4 = → TMP + C=C(CH3)−C+ (6)
kinetics with varied temperatures were performed with the
stirring speeds of 2800 rpm.
C=C(CH3)−C+ + iC4 =
3. KINETIC MODEL → C=C(CH3)−C−C−(CH3)C+−C (7)
It is well accepted that the alkylation of isobutane with olefins
follows the classic carbonium ion mechanism. Initially the tert- Through suitable hydrogen and proton transfer reactions, the
butyl cation is produced by this reaction: unsaturated cation ion forms a DMH molecule. Another
possible pathway to give DMHs is through the addition of 1-
k1
iC4 = + H+ → iC4 + (1) butene to iC4+,11
k5
where iC4= is isobutene and iC4+ represents tert-butyl cation. iC4 + + 1‐butene → DMHs+ (8)
Actually, regardless of whether the feed butene is either 1-
butene or 2-butene, it tends to give iC4= through fast Similar to TMPs+, the DMHs+ are saturated through the
isomerization, which occurs especially in the presence of hydride abstraction from isobutane, giving an iC4+ meanwhile,
sulfuric acid.15 The transform between different butenes k6
proceeds by the following equilibrium reaction, DMHs+ + iC4 → DMHs + iC4 + (9)
K1 K2
1 − C4 H8 ⇄ 2 − C4 H8 ⇄ iC4 = According to the experimental data in this work, it was verified
(2)
that the formation of DMHs is most likely through the latter
Interestingly, it is found that even when isobutane alkylated pathway, that is, the addition of 1-butene to iC4+.
with olefins rather than butenes, the trimethylpentanes are still By reactions involving olefins, heavy ends (HEs) are
generated by so-called self-alkylation. The key step for self- produced, which are primarily polymerization reactions.15
alkylation is the formation of isobutene, which in fact is a The polymerization of C4 olefins proceeds by the addition of
reversible reaction of the reaction 1:1 iC4= or 2-butene to TMPs+ or DMHs+,
k2
iC4 + → iC4 = + H+ (3) TMPs+ (or DMHs+) + iC4 = (or 2‐butene) → iC12+
(10)
The tert-butyl cation adds to an olefin to produce the
corresponding C8 carbonium cation. For example, when the The iC12 would further add iC4 or 2-butene to give iC16+.
+ =

tert-butyl cation reacts to iC4= or 2-butene, the trimethylpen- For simplification, in this work, we regard iC12+ and iC16+ as the
tanes will be formed, pseudo components iCm+. When these polymer cations iCm+
are contacted with isobutane, the HEs is formed, along with the
k3
iC4 + + 2‐butene (or iC4 =) → TMPs+ (4) production of iC4+,
k7
where TMPs is the abbreviation of trimethylpentanes. These iCm+ + iC4 → iCm + iC4 + (11)
C8 carbonium cations tend to isomerize via methyl shifts and
hydride transfer reaction to form several isomer cations.4 For As one of the five groups in alkylated hydrocarbon products,
simplification, in this work, all these C8 cations are regarded as the light ends (LEs) include C5−C7 isoparaffins. The
one pseudo component TMP+. By rapid hydride transfer with production of LEs is believed to come from the fragmentation
isobutane, each TMP+ changes to the corresponding TMP of large isoalkyl cations.9 Due to the strong oxidizing capacity of
component, regenerating the iC4+ to keep on the chain concentrated sulfuric acid, the isoalkyl cation will be generated
propagation. again from HEs as follows,
k4
TMP+ + iC4 → TMP + iC4 + (5) iCm + H+ → iCm+ (12)

where iC4 stands for isobutane. The source of undesirable These heavy isoalkyl cations are fragmentated following the
dimethylhexanes (DMHs) with low octane numbers has several β-scission rule, leading to smaller isoalkyl cations and olefins.1
possible pathways. One was believed from the addition of iC4= k8
and allylic carbocation, which is the result of a hydride ion iCm+ ⇄ iCx + + iCy =
abstraction from the allylic position of an olefin molecule.10 k9 (13)

C dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX


Industrial & Engineering Chemistry Research Article

Figure 4. Concentration profile 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 c1‐C4H8 = α*cC4H8 (15)


protonation of olefins (iCy=) are thought to be the major
precursors of LEs. Through the hydride ion transfer, these c 2‐C4H8 = β*cC4H8 (16)
isoalkyl cations change into the corresponding LEs.
According to the experimental observation in this work, the ciC4= = (1 − α − β)*cC4H8 (17)
degradation of DMH was found to be obvious. It may be
ascribed to the formation of DMH+ from the oxidation of where α and β represent 1/(K1 + K1*K2 + 1), K1/(K1 + K1*K2
DMH, which is further fragmentated into small LE through + 1) respectively.
scission reaction.16 To simplify this multistage reaction, one Following the above reaction steps and assumptions, the
single reaction step is used in this work, kinetic model can be formulated readily as follows,
k10 dc1
DMH → LE (14) = −(1 − α − β)k1c1 − (1 − α)k3c1c3 − αk5c1c3
dt
The diagram of the whole reaction pathway network was shown − (1 − α)k7c1c 2c4 + k2c 3 (18)
in Figure 3.
The alkylation products are composed of more than two dc 2
dozen branched−chain paraffinic hydrocarbons. Both TMPs = −k4c 2c4 − k6c 2c5 − (1 − α)k7c1c 2c4
dt (19)
and DMHs have multiple isomers. HE, as well as LE, also refers
to one family of isoparaffins with close molecular weight. dc 3
Including those carbonium components, there will be more = (1 − α − β)k1c1 + k4c 2c4 + k6c 2c5 − k4c1c3
dt
than three dozen species in this system. However, only three or
four groups of components can be measured due to the limit of − αk5c1c3 − k2c 3 + (1 − α)k 7c1c 2c4 (20)
analysis method and the faster transform between various
species. To avoid the over fit or over parametrization, the dc4
= (1 − α)k3c1c3 − k4c 2c4 − (1 − α)k7c1c 2c4
number of involved species and pathway should be simplified as dt (21)
much as possible. So first, some species with similar properties
are treated as one pseudo component, such as TMPs (or dc5
= αk5c1c3 − k6c 2c5
DMHs) mentioned above. HE, as well as LE, is regarded as one dt (22)
single component, respectively, because of close molecular
weight. Second, some steps are assumed to be instantaneous, dc6
= k4c 2c4
such as reactions 10 and 12, and the reaction 2 is assumed to be dt (23)
under chemical equilibrium control. Accordingly, the concen-
tration distribution for these three butenes can be calculated, dc 7
= k6c 2c5 − k10c 7
respectively, on the basis of the total concentration of butene, dt (24)

D dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX


Industrial & Engineering Chemistry Research Article

dc 8 parameters and make the estimated rate constants more


= (1 − α)k7c1c 2c4 + k9c 9c10 − k8c 8 definite. So, in the refitting of experimental data, the values
dt (25)
of k3 through k10 are kept fixed with temperatures, and the only
dc 9
= k8c 8 − k9c 9c10 adjustable parameters are k1 and k2. The fitting results are
dt (26) shown in Figure 4, compared with the experiments. It can be
seen that the agreement between experimental and calculated
dc10 data is quite satisfactory. In particular, the dramatic change of
= k8c 8 − k9c 9c10
dt (27) HEs in short time is captured successfully by the model.
These ordinary differential equations (ODEs) are completed The estimated k1 through k10 are listed in Table 2 with 95%
with specifications of the following initial conditions at time t = confidence intervals. We can see that most of the confidence
0, intervals are one magnitude smaller than the corresponding
parameters except k3 and k6, which confirms the reliability of
c1 = c10 ; c2 = c20; c3 = 0; c4 = 0; c5 = 0; the estimated rate constants. However, an anti-Arrhenius
behavior can be found for k1; that is, the value of k1 falls as
c6 = 0; c 7 = 0; c8 = 0; c 9 = 0; c10 = 0 (28) the temperature increases. In fact, this phenomenon is not
In eqs 18 to 28, the relationship between the numbers of 1 unusual or abnormal for hydrocarbon reaction with the
through 10 and the corresponding species is as follows: 1, presence of free radical molecules.18 The observations of anti-
butene; 2, iC4; 3, iC4+; 4, TMP+; 5, DMH+; 6, TMP; 7, DMH; Arrhenius kinetics are usually ascribed to an entropic
8, HE; 9, iCx+; 10, iCy=. contribution to the free energy of activation. This anti-
Arrhenius behavior is believed to be caused by a decrease of
4. RESULTS AND DISCUSSION the equilibrium constant of a multistep reaction.19
The concentration profiles of three groups of key components The Arrhenius relationships of k1 and k2 are obtained by
in alkylation of isobutane with butene were plotted in Figure 4. plotting ln(ki) against 1/T shown in Figures 5 and 6,
It shows that all of three species have dramatic changes within 2 respectively, both of which have good linear relationships.
min. Initially, the HEs go up sharply and then come down The activation energies of k1 and k2 are computed to be −58.3
shortly. This demonstrates that the alkylation of isobutane with and 173.2 kJ/mol, and the corresponding pre-exponential
butene is a very fast reaction. Almost after 5 min, each of the factors are 5.23 × 10−11 and 6.87 × 1031 min−1 respectively.
species reaches to a platform individually, although the DMHs According to the fundamental principle of physical chemistry,
still continue to go down slightly. in a reversible reaction, the difference of activation energy
To fit the experiments and estimate k1 through k10 in eqs between the forward and backward reaction should be equal to
18−28, the nonlinear least-squares fitting method was the heat of the reaction. So, in this sense, the heat of reaction
employed by the following object function, for the addition of H+ to the iC4= can be calculated to be
m −231.5 kJ/mol, which indicates that this is a strong exothermic
obj = ∑ (ciexp − cical)2 reaction.
i=1 (29) Using uniform hydrocarbon drops, L. Lee et al. studied the
sulfuric acid catalyzed reaction of isobutane with 1-butene and
where cexp
and ccal
represent the experimental and calculated
i i the 1-butene oligomerization.12 Kinetic constants for a
values of ith component, respectively, and m is the number of
simplified model were calculated, in which only the primary
experimental data. An lsnonlin function in Matlab was used for
reactions were considered. When using sulfuric acid of 98%, the
the regress of the set of k1 through k10. Equations 18−28 were
rate constants for alkylation of isobutane with 1-butene and 1-
resolved by the Runge−Kutta method in each iteration. Before
the fitting was performed, the values of α and β should be butene oligomerization at 298 K were calculated to be 2.25 ×
determined first; that is, the equilibrium constants in reaction 2 108 and 9.5 × 108 cm3/mol/s, respectively. However, in this
should be computed. According to thermodynamics, K1 and K2 work, there are no such reaction steps exactly corresponding to
are the functions of temperature and can be calculated on the those in the literature.12 This is because the kinetic model in
basis of the Gibbs free energy variation of reaction 2. As we this work was developed on the basis of the carbonium ion
know, the ASPEN Plus software is a useful tool in chemical reaction mechanism which used elementary steps, while L. Lee
process modeling and has a reliable thermodynamic database. et al. dealt with the alkylation and oligomerization reactions by
The RGibbs module in ASPEN Plus was used to obtain the the empirical method on the basis of power exponent. Even so,
average values of K1 and K2 at the temperatures between 276.2 it is found that one elementary step in our model is similar to
and 285.2 K, which is 28.39 and 6.433, respectively. the alkylation reaction of isobutane with 1-butene, which is the
Accordingly, the values of α and β can be calculated to be addition of 1-butene to iC4+ (eq 8). The rate constant for this
0.004718 and 0.1340, respectively. step was estimated to be 1.26 × 104 kg/mol/min, which equals
Preliminary fitting results show that with the variation of 1.17 × 105 cm3/mol/s. It seems much smaller than the
temperature most of the rate constants change a little, except k1 alkylation reaction rate constant of isobutane with 1-butene
and k2. It indicates that those reaction steps corresponding to reported by L. Lee et al., but it should be noted that our
the rate constants of k3 through k10 have small or even no experiments were carried out at the temperatures from 276.2 to
activation energy. A similar finding has been reported by 285.2 K, which are 13 to 22 K smaller than that reported by L.
Langley et al. in the alkylation kinetics of isobutane with Lee et al. (298 K). According to L. Albright et al.’s work,10 at
propylene.17 In this work, these rate constants including k3 the temperature of 259 K, the alkylation reaction rate slowed
through k10 are assumed to be constant with varied temper- down dramatically, in which “hour” was used as the unit of
atures in the range from 276.2 to 285.2 K. One of the time. Thus, taking into consideration the sensibility of
advantages of such an assumption is to reduce the adjustable alkylation reaction to temperature, the difference of the rate
E dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

2.61 ± 0.85
k10, 10−2
min−1
kg·mol−1·min−1
5.48 ± 0.82
k9

0.63 ± 0.06
min−1
k8

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


kg2·mol−2·min−1
88.46 ± 6.78
k7
kg·mol−1·min−1
4.36 ± 2.48
k6, 103
kg·mol−1·min−1
1.26 ± 0.42

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


k5, 104

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
kg·mol−1·min−1
1.03 ± 0.22

believed that the increasing rates of alkylate formation during


the initial stages were ascribed to the increasing acidity of the
k4
Table 2. Estimated Rate Constants with 95% Confidence Intervals

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+
kg·mol−1·min−1
8.33 ± 1.86

and olefin, the precursor of alkylates are generated. Thus, in the


k3, 102

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 olefin
concentration. This is similar to the free radical chain reaction
of hydrocarbon, in which the existence of induction period is
0.06
0.20
0.38
0.56

very common in the oxidation of hydrocarbon, especially at low


min−1
k2

±
±
±
±

temperature.
0.15
0.45
1.00
1.59

To confirm 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
0.72
0.57
0.50
0.71

the typical industrial alkylation reactor. The industrial alkylation


min−1
k1

±
±
±
±
6.37
5.06
4.15
2.79

Table 3. Operating Conditions for Industrial Alkylation


Reactor
A/H, m3/m3a I/O, mol/molb temperature, K pressure, MPa
temperature

1.1:1 8:1 279 0.45


276.2
279.2
282.2
285.2
K

a
Ratio of sulfuric acid to hydrocarbon (isobutane + olefin) based on
volume. bRatio of isobutane to olefin based on mole.

F dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX


Industrial & Engineering Chemistry Research Article

reactor of isobutane with olefin is a well mixed flow reactor.1 5. CONCLUSIONS


For this type of reactor, the material balance equation can be The alkylation kinetics of isobutane with butene was measured
written as follows: in the condition of industrial interest. The kinetic model was
xi 0 − xi = τ *( −ri) (30) established on the basis of the classic carbonium ion
mechanism. The agreement between experiments and model
where xi0and xi stand for the inlet and outlet concentration of fitting is quite satisfactory. Except the addition of H+ to
ith component in alkylates with the unit of mol/kg and τ and isobutene and its reversible reaction, other reaction steps’ rate
−ri represent the mean residence time of sulfuric acid in reactor constants were found not to change with the varied
and formation or consumption rate of the ith component with temperatures from 276.2 to 285.2 K. An anti-Arrhenius
the unit of min and mol/kg/min. In the eq 30, −ri is behavior was observed for the addition of H+ to isobutene.
determined by the kinetic model in this work and xi0 is Both of them also can be found in other hydrocarbon reactions.
calculated through the operating conditions listed in Table 3. By the kinetic model, the successful simulation of industrial
Thus, xi can be solved only if τ is given. alkylation reactor was achieved. We hope that the alkylation
The conversion of olefin and alkylate yield versus mean kinetics reported in this work will be useful for the reactor
residence time is shown in Figures 7 and 8, respectively. It can 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
The authors declare no competing financial interest.


Figure 7. Conversion of olefin versus mean residence time.
ACKNOWLEDGMENTS
The authors gratefully acknowledge the financial 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 Founda-
tion (2012M512055), and the 111 Project (B08021).

■ REFERENCES
(1) Corma, A.; Martínez, A. Chemistry, catalysts, and processes for
isoparaffin−olefin alkylation: Actual situation and future trends. Catal.
Rev.: Sci. Eng. 1993, 35 (4), 483−570.
(2) Shorey, S. W. In Motor-Fuel Alkylation with CDAlky Process
Technology. 103rd NPRA Annual Meeting, San Francisco, CA, March
Figure 8. Alkylate yield versus mean residence time. 13−15, 2005; pp 1−16.
(3) Hommeltoft, S. I. Isobutane alkylation: Recent developments and
future perspectives. Appl. Catal., A: Gen. 2001, 221 (1), 421−428.
(4) Kranz, K. Intro to Alkylation Chemistry: Mechanisms, operating
be seen that the conversion of olefin will be greater than 99.5% variables, and olefin interactions; DuPont STRATCO Clean Fuel
when the mean residence time (τ) is above 10 min, but the Technology: Leawood, 2008; pp 1−29.
yield of alkylate is under 200% even when τ reaches 20 min. (5) Albright, L. F. Alkylation of isobutane with C3−C5 olefins:
Usually, the alkylate yield is required to be more than 200%, so Feedstock consumption, acid usage, and alkylate quality for different
the τ should be kept greater than about 21 min according to processes. Ind. Eng. Chem. Res. 2002, 41 (23), 5627−5631.
Figure 8. It also indicates that when τ exceeds 30 min the (6) Feller, A.; Zuazo, I.; Guzman, A.; Barth, J. O.; Lercher, J. A.
alkylate yield will approach its theoretical value (203.6%). It Common mechanistic aspects of liquid and solid acid catalyzed
was reported that in the industrial unit the mean residence time alkylation of isobutane with n-butene. J. Catal. 2003, 216 (1−2), 313−
ranges from 20 to 30 min.1,20 323.
Clearly, the simulation results show that the kinetic model (7) Schmerling, L. The mechanism of the alkylation of paraffins. II.
Alkylation of isobutane with propene, 1-butene and 2-butene. J. Am.
developed in this work can predict the mean residence time of Chem. Soc. 1946, 68 (2), 275−280.
the industrial alkylation reactor successfully. As we all know, the (8) Schmerling, L. Reactions of hydrocarbons: Ionic mechansims.
mean residence time determines the reactor size when the scale Ind. Eng. Chem. 1953, 45 (7), 1447−1455.
of production is known and vice versa. Thus, using the kinetic (9) Hofmann, J. E.; Schriesheim, A. Ionic reactions occurring during
model developed in this work, we will be allowed to design and sulfuric acid catalyzed alkylation. II. Alkylation of isobutane with C14-
optimize the novel alkylation reactors. labeled butenes. J. Am. Chem. Soc. 1962, 84 (6), 957−961.

G dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX


Industrial & Engineering Chemistry Research Article

(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), 391−397.
(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), 447−454.
(12) Lee, L.-m.; Harriott, P. The kinetics of isobutane alkylation in
sulfuric acid. Ind. Eng. Chem. Process Des. Dev. 1977, 16 (3), 282−285.
(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), 2110−2120.
(14) Aschauer, S. J.; Jess, A. Effective and intrinsic kinetics of the two-
phase alkylation of i-paraffins with olefins using chloroaluminate ionic
liquids as catalyst. Ind. Eng. Chem. Res. 2012, 51 (50), 16288−16298.
(15) Albright, L. F. Mechanism for alkylation of isobutane with light
olefins. 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), 53−60.
(17) Langley, J. R.; Pike, R. W. The kinetics of alkylation of isobutane
with propylene. AIChE J. 1972, 18 (4), 698−705.
(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), 9276−80.
(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), 4283−4289.

H dx.doi.org/10.1021/ie400415p | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

You might also like