Vous êtes sur la page 1sur 7

446 Ind. Eng. Chem. Res.

2009, 48, 446–452

Direct Preparation Kinetics of 1,3-Dichloro-2-propanol from Glycerol Using


Acetic Acid Catalyst
Zheng-Hong Luo,* Xiao-Zi You, and Hua-Rong Li
Department of Chemical and Biochemical Engineering, College of Chemistry and Chemical Engineering,
Xiamen UniVersity, Xiamen 361005, China

Direct preparation of 1,3-dichloro-2-propanol from glycerol is carried out in a batch reactor using acetic acid
catalyst at 363-393 K. The analytical technique, gas chromatography, is employed in order to follow the
time evolution of the reagents. The kinetic model of the process is developed. Furthermore, the model
parameters of the process are also determined by data fitting. The results show that direct preparation follows
the SN2 mechanism. A kinetic model corresponding to the mechanism is proposed in this work. The
experimental results show that the kinetic model agrees well with the experiments.

1. Introduction presented. Recently, Lee et al. studied direct preparation using


heteropolyacid (HPA) catalyst.6 In their work direct preparation
The synthesis of 1,3-dichloro-2-propanol (DCP) has been
of DCP was carried out in a liquid-phase batch reactor with an
increasingly important because DCP is an intermediate in the
aim of searching for highly active HPA catalyst in the reaction.
production of epichlorohydrin (ECH, 1-chloro-2,3-epoypropane).1,2
For this purpose, they examined a wide set of heteroatom- and
Furthermore, ECH is an intermediate in the production of
polyatom-substituted HPA catalysts but did not mention the
epoxide resins, glycerine, and several pharmaceutical products.2-4
process kinetics. On the basis of the above discussion it is clear
The commercial preparation method of DCP includes two
that past studies on direct preparation of DCP from GLY are
consecutive processes, consisting of preparing allyl chloride
limited to the unit operation design and catalyst exploitation of
through chlorination of propylene at a high temperature and
the process. Furthermore, available information is mainly
preparing DCP through subsequent chlorination of allyl chloride
restricted to the patent data and technical feasibility studies.1-19
under an excess amount of industrial water.1,2 It is not
There are few reports on studies of the process kinetics, which
environmental friendly due to the high consumption of raw
describes the time evolution of the reagents and product
materials and large amount of byproduct and water produced.4-6
properties. In practice, the process kinetics is always the
In addition, the commercial method greatly depends on the
important proportion of process engineering and still the basic
petroleum products. At present, the shortages of the petroleum
data of process design.
products are obvious. Therefore, a direct preparation method
In the present study, the direct preparation kinetics of DCP
of DCP with high yield of DCP and low consumption of raw
from GLY is studied. Direct preparation of DCP is carried out
materials must be explored.
in a batch reactor using HAC-catalyst at 363-393 K. In
On the other hand, biodiesel as a green fuel has been produced
addition, the kinetic model and its parameters of the process
on a large scale.6-10 However, one of the major problems in
are also obtained.
the production of biodiesel is formation of a considerable amount
of glycerol (GLY) (ca. 10 wt %) as a byproduct.6 This makes
2. Experimental Details
the oversupply of GLY to the industrial market and decreases
the commercial value of GLY. Therefore, direct conversion of 2.1. Direct Preparation of DCP from GLY. Direct prepa-
GLY to high-value chemicals has attracted much attention.6,11 ration of DCP from GLY was studied in the presence of HAC.
One of the promising methods to convert GLY to high-value All reactions were performed in a 500 mL liquid-phase bath
chemical is to produce DCP from GLY in a single step.6,11,12 reactor equipped with reagent inlet, magnetic stirrer, condensa-
The direct process is performed by reaction/chlorination of GLY tion pipe, and temperature control device.
with hydrochloric acid. Accordingly, DCP produced includes The reactor was first heated to a certain temperature. Certain
2,3-dichloro-1-propanol (2,3-DCP) and 1,3-dichloro-2-propanol quantities of HAC-catalyst, GLY, and aqueous hydrochloric acid
(1,3-DCP). Here, DCP represents 1,3-DCP. Scheme 1 shows
the commercial and direct processes for producing DCP. Scheme 1. Commercial and Direct Processes for Producing DCP
According to Scheme 1, it is easy to know that the direct process
is an environmental process and has many economical advan-
tages compared to the commercial process.6,13-15
To date, there have been some reports on the direct process
for producing DCP from GLY. Several processes utilizing
carboxylic acids as catalyst were developed for direct preparation
of DCP from GLY.16-18 In addition, the main commercial
processes used acetic acid (HAC-catalyst) as catalyst. Corre-
sponding reports16-19 concentrated on purification of GLY or
separation and reuse of the catalyst; no actual kinetic data were
* To whom correspondence should be addressed. Tel.: +86-592-
2187190. Fax: +86-592-2187231. E-mail: luozh@xmu.edu.cn.

10.1021/ie8011177 CCC: $40.75  2009 American Chemical Society


Published on Web 11/17/2008
Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009 447

Figure 1. Concentration histories during the direct preparation of DCP from


GLY (Experimental conditions: cA0 ) 2.36 mol · L-1; cB0 ) 0.79 mol · L-1; Figure 2. Time-dependent mass fraction of MCP without HAC (Experi-
cD0 ) 32.53 mol · L-1; cE0 ) 9.42 mol · L-1; T ) 393 K). mental conditions: cA0 ) 0.00 mol · L-1; cB0 ) 0.00 mol · L-1; cD0 ) 20.12
mol · L-1; cE0 ) 6.08 mol · L-1; cF0 ) 6.08 mol · L-1; T ) 393 K).

solution (HCl) (37 wt %, chlorination agent) were added into


the reactor. Simultaneously, the reaction took place and lasted
for a certain time. The stirring speed was 800 rpm, at which
the diffusion effect on the reaction was minimized and could
be ignored.4,20,21 The reacting solution was withdrawn at regular
intervals. Calcium carbonate (CaCO3) (20 wt %) was added into
the samples to remove HCl. In addition, the samples handled
were diluted with methanol. Accordingly, the samples were
analyzed with a gas chromatograph equipped with a FID.
2.2. Measurement. The reaction products were analyzed
with a gas chromatograph (HX GC-950) equipped with a FID.
KB-Wax columns (cross-linked polyethylene Glycol, 30 m ×
0.32 mm × 1.00 µm) were used. Chromatogram operational
conditions were determined: vaporization temperature 533 K;
detection temperature 553 K. Oven temperature program: 383
K (1 min), 10 K · min-1 to 483 K (2 min) and then 15 K min-1 Figure 3. Concentration histories during the reaction between MCP and
to 523 K (5 min). Glycol was used as internal standard, and HCl in the presence of HAC (Experimental conditions: cA0 ) 0.00 mol · L-1;
the results were calculated by the internal standard method. cB0 ) 0.87 mol · L-1; cD0 ) 19.13 mol · L-1; cE0 ) 5.65 mol · L-1; cF0 )
5.65 mol · L-1; T ) 393 K).
3. Results and Discussion
On the basis of the above descriptions, we find that the
3.1. Results. In order to obtain the direct preparation kinetics concentration of MCP increases quickly and the concentration
of DCP, the typical kinetic experiment must be performed and of GLY decreases quickly in the period of 0-100 min. In
the concentration histories in the reaction system must be addition, the concentration of MCP decreases and the concentra-
recorded. In addition, here, we must point out that all records tion of GLY remains unchanged after that. Simultaneously, the
by the gas chromatogram in this study reveal that the main concentration of DCP still increases. Accordingly, we deduce
product and byproduct of the reaction are DCP and 3-chloro- that MCP is the intermediate in the production of DCP from
1,2-propanediol (MCP), respectively. Furthermore, all records
GLY. In order to verify the result, additional experiments are
also show that the concentrations of 2,3-DCP and trichloropro-
performed, and the results are shown in Figures 2 and 3.
pane formed by further chlorination of DCP are much less than
From Figure 2 one can find that the change of the concentra-
that of DCP, and their concentrations are nearly unchanged with
tion of MCP demonstrates ‘wave’ shape with the increase of
the increase of reaction time. In order to illustrate the above,
several typical gas chromatogram figures are listed in Appendix reaction time and no law is obtained. As a whole it shows a
A. Accordingly, formation of 2,3-DCP and trichloropropane will parallel line to the time axis. To illustrate the result, a smooth
not be considered in the next study. Figure 1 demonstrates the line is fitted and also shown in Figure 2. The smooth line further
time dependent on the concentrations of GLY, DCP, MCP, and demonstrates that the concentration of MCP remains nearly
HAC in the system. constant, and DCP cannot be produced efficiently in the reaction
According to Figure 1, we can obtain that the concentration system without HAC in this work, namely, MCP can hardly
of GLY decreases to its minimum at about 100 min and remains react with HCl without HAC under the experimental conditions
nearly unchanged after that; the concentration of HAC used as provided in this work, although MCP can be converted into
catalyst remains nearly unchanged with the reaction proceeding. DCP in the noncatalytic reaction.
Figure 1 also shows that MCP can be produced immediately in Figure 3 shows that the concentration of MCP decreases and
the reaction system and its concentration increases to its the concentration of DCP increases with the increase of time
maximum at about 100 min and decreases after that. Further- in the presence of HAC. Therefore, for the direct preparation
more, we obtain that the concentration of DCP increases during system, accumulation of DCP depends on the consumption of
the reaction from Figure 1. MCP and MCP can be obtained from reaction of GLY with
448 Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009

Scheme 2. SN2 Mechanism

Scheme 3. Direct Preparation Mechanism of DCP from GLYa

Figure 5. Comparison between experimental and simulated data of


concentration of MCP vs time (Experimental conditions: cA0 ) 2.36
mol · L-1; cB0 ) 0.79 mol · L-1; cD0 ) 32.53 mol · L-1; cE0 ) 9.42 mol · L-1).

a
A represents GLY, B represents HAC, C represents glycerol-1-acetate
(GLYA), D represents water (H2O), E represents HCl, F represents MCP,
G represents 3-chloropropandeiol-1-acetate (MCPA), and H represents DCP.

Figure 6. Comparison between experimental and simulated data of


concentration of DCP vs time (Experimental conditions: cA0 ) 2.36
mol · L-1; cB0 ) 0.79 mol · L-1; cD0 ) 32.53 mol · L-1; cE0 ) 9.42 mol · L-1).

Table 1. Obtained Kinetic Model Parameters


kinetic computational activation
constants equation, energies
(mol · L-1 · s-1) k ) Ae(-E/RT) (kJ · mol-1)
k1 k1 ) 78.5e(-4427.7/T) 36.8
k1′ k1′ ) 4 × 10-11e(-3892.5/T) 32.4
Figure 4. Comparison between experimental and simulated data of k3 k3 ) 63.8e(-5283.5/T) 43.9
concentration of GLY vs time (Experimental conditions: cA0 ) 2.36 3.3. Kinetic Model. The kinetic model of the direct prepara-
mol · L-1; cB0 ) 0.79 mol · L-1; cD0 ) 32.53 mol · L-1; cE0 ) 9.42 mol · L-1).
tion of DCP from GLY established in this paper is constructed
based on the direct preparation mechanism and mass balance
HCl. Therefore, we confirm that MCP is the intermediate in equations for the chemical species in the reaction system.
the production of DCP from GLY. The following modeling assumptions are made in this
3.2. Direct Preparation Mechanism of DCP from GLY. paper:3,4,11,21 (1) the concentration of HAC is constant due to
Many researchers6,22-27 have proved that reactions between its catalytic function in the reaction system, (2) the concentration
alcohol and hydrogen halide belong to nucleophilic substitution of H2O is constant due to its surplus in the reaction system,
reactions; the reaction mechanism in the reaction of primary and (3) the quasi-stationary-state assumption is applied. There-
alcohol and hydrogen halide is a second-order nucleophilic fore, according to the reaction steps shown in Scheme 3, the
substitution mechanism (SN2), which is shown in Scheme 2.24 mass balance equations can be worked out.
From the structure formula of GLY it can be found that chloride For GLY, the following mass balance equation can be derived
ion can react with the terminal carbon easily. In addition, dcA
dichloropropanol and monchloropropanediol can be found in ) -k1cAcB + k′1cCcD (1)
dt
the chlorination of GLY. Thus, it can be deduced that the Similar equations can be derived for HAC, GLYA, MCP,
chlorination mechanism of GLY is also a SN2 mechanism. MCPA, and DCP
On the basis of the above analysis and our experimental
results, the simplified mechanism of the direct preparation of dcB
) -k1cAcB + k′1cCcD + k2cCcE - k3cFcB + k4cGcE (2)
DCP from GLY is suggested and shown in Scheme 3.6,24-27 dt
Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009 449
dcC dcH
) k1cAcB - k2cCcE - k′1cCcD (3) ) k3cFcB (13)
dt dt
dcF By integrating eq 11 we obtain
) k2cCcE - k3cFcB (4)
dt
dcG
 
cA ) Ce-(k1cB+k1cD)t + e-(k1cB+k1cD)t ∫k c c

1 D A0e
(k1cB+k1 cD)t
dt )
) k3cFcB - k4cEcG (5) k′1cDcA0
dt 
Ce-(k1cB+k1cD)t + (14)
dcH k1cB + k′1cD
) k4cEcG (6)
dt
According to the quasi-stationary-state assumption we obtain
dcC
) k1cAcB - k2cCcE - k′1cCcD ) 0 (7)
dt
dcG
) k3cFcB - k4cEcG ) 0 (8)
dt
Equations 9 and 10 can be obtained by solving eqs 7 and 8
k1cAcB
cC ) (9)
k2cE + k′1cD
k3cFcB
cG ) (10)
k4cE
Therefore, we obtain eqs 11-13 by substituting eqs 7-10 into
eqs 1, 4, and 6
Figure 9. Temperature-dependent reaction constant (k3).
dcA
) k′1cDcA0 - (k1cB + k′1cD)cA (11)
dt
dcF
) (k1cB + k′1cD)cA - k′1cDcA0-k3cBcF (12)
dt

Figure 10. Comparison between experimental and simulated data of


concentration of GLY vs time (Experimental conditions: cB0 ) 0.79
mol · L-1; nE0 ) 4 mol; T ) 393 K).

Figure 7. Temperature-dependent reaction constant (k1).

Figure 11. Comparison between experimental and simulated data of


concentration of MCP vs time (Experimental conditions: cB0 ) 0.79
Figure 8. Temperature-dependent reaction constant (k1′). mol · L-1; nE0 ) 4 mol; T ) 393 K).
450 Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009

Equations 15-17 will be used to describe the variation of


cA, cF, and cH with time and temperature. In addition, there are
three rate constants in eqs 15-17, namely, k1, k1′, and k3. The
temperature dependence of these rate constants can be repre-
sented with Arrhenius equation and shown in eqs 18-20.
Hereinafter, we will examine the applicability of eqs 15-20
by experimental data. The values of kinetic parameters can be
obtained by the least-squares method with the data obtained
experimentally

k1 ) A1e-E1⁄RT (18)

k′1 ) A′1e-E1⁄RT (19)

k3 ) A3e-E3⁄RT (20)

Figure 12. Comparison between experimental and simulated data of


3.4. Kinetic Constant Estimation and Model Verifica-
concentration of DCP vs time (Experimental conditions: cB0 ) 0.79 mol · L-1; tion. Relating the experimental data shown in Figures 4, 5, and
nE0 ) 4 mol; T ) 393 K). 6 with eqs 15-17 we obtain the model parameters by the least-
squares method. The obtained model parameters are shown in
where C is a constant number and can be obtained by the initial Table 1.
value of cA. cA ) cA0 at t ) 0, then C ) k1cB/k1cB + k1′cDcA0. Figures 4, 5, and 6 illustrate the comparisons between the
Therefore, cA can be described by eq 15 experimental concentration and the simulated concentration
cA0 
during the reaction at different temperature. The simulated
cA ) (k1cBe-(k1cB+k1cD)t + k′1cD) (15) results meet experimental data well. The correlation coefficients
k1cB + k′1cD of eqs 15-17 for corresponding experimental data all exceed
Similar equations can also be obtained by integrating eqs 12 0.990. In addition, the average deviations of eqs 15, 16, and 17
and 13 are 3.27%, 4.82%, and 4.37%, respectively. Furthermore, we
obtain the temperaturedependent model parameters obtained,
k1cBcA0  which are shown in Figures 7, 8, and 9. The linear correlation
cF ) [e-(k1cB+k1cD)t - e-k3cBt] (16)
k3cB - (k1cB + k′1cD) coefficients in Figures 7, 8, and 9 all exceed 0.990.

( )

In order to further verify the above kinetic model, other
k1cBk3cBcA0 e-k3cBt e-(k1cB+k1cD)t k1cBcA0 experiment data at different initial mole ratios of GLY and HCl in
cH ) ′
- ′
+
k c
k3cB - (k1cB + k1cD) 3 B k1cB + k1cD k1cB + k′1cD the reaction system are also obtained. Corresponding comparisons
between the experimental data and the simulated data are shown
(17)
in Figures 10, 11, and 12. Figures 10, 11, and 12 prove that the
where the values of cB and cD are constant according to the simulated results meet experimental data quite well. The average
above modeling assumptions. correlation coefficients of the model for the experimental data

Figure 13. Typical gas chromatogram records: (a) 2,3-DCP and (b) trichloropropane (Experimental conditions same as Figure 1).
Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009 451
exceed 0.990. In addition, the average deviations of eqs 15, 16, 1,3-DCP) in food products using isotope dilution GC-MS. Food Control.
and 17 are 3.36%, 4.73%, and 5.06%, respectively. 2007, 18, 81–90.
(2) Horsley, H. L. Encyclopedia of Chemical Technology, 3rd ed.; Wiley:
The above results indicate that the kinetic models provide a Hoboken, NJ, 1965; pp 595-612.
reasonable fit of the experiment data and can be used to describe (3) Ma, L.; Zhu, J. W.; Yuan, X. Q.; Yue, Q. Synthesis of epichloro-
the time-dependent concentrations of GLY, MCP, and DCP hydrin from dichloropropanols kinetic aspects of the process. Trans.
under different reaction conditions. IChemE, Part A, Chem. Eng. Res. Des. 2007, 85, 1580–1858.
From Table 1 it can be found that the activation energies of (4) Carrà, S.; Santacesarla, E.; Morbldlll, M. Synthesis of epichlorohydrin
by elimination of hydrogen chloride from chlorohydrins: 1. kinetic aspects
the esterifications (Scheme 2) in this work are 36.8 and 43.9 of the process. Ind. Eng. Chem. Process Des. DeV. 1979, 18, 424–427.
kJ · mol-1, respectively, which approach that of the typical (5) Wang, L. L.; Liu, Y. M.; Xie, W. Highly efficient and selective
esterification following the SN2 mechanism.27-30 It further production of epichlorohydrin through epoxidation of allyl chloride with
proves that direct preparation of DCP from GLY in this work hydrogen peroxide over Ti-MWW catalyst. J. Catal. 2007, 246, 205–214.
(6) Lee, S. H.; Park, D. R.; Kim, H.; Lee, J.; Jung, J. C.; Woo, S. Y.;
follows the SN2 mechanism. Song, W. S.; Kwon, M. S.; Song, I. K. Direct preparation of dichloropro-
panol (DCP) from glycerol using heteropolyacid (HPA) catalysts: A catalyst
4. Conclusions screen study Catal. Commun. 2008; doi:10.1016/j.catcom.2008.03.020.
(7) Rashid, U.; Anwar, F. Production of biodiesel through optimized
In this work, the direct preparation kinetics of DCP from GLY alkaline-catalyzed transesterification of rapeseed oil. Fuel 2008, 87 (3), 265–
using HAC-catalyst is investigated experimentally and theoreti- 273.
cally. The results show that direct preparation follows the SN2 (8) Xie, W.; Peng, H.; Chen, L. Transesterification of soybean oil
catalyzed by potassium loaded on alumina as a solid base catalyst. Appl.
mechanism. Kinetic models, eqs 15-17, are proposed to Catal. A: Gen. 2006, 300, 67–74.
describe the kinetic data of the direct preparation of DCP from (9) Macleod, C. S.; Harvey, A. P.; Lee, A. F.; Wilson, K. Evaluation of
GLY using HAC-catalyst. The results show that the simulated the activity and stability of alkali-doped metal oxide catalysts for application
data meet experimental data quite well. to an intensified method of biodiesel production. Chem. Eng. J. 2008, 135,
63–70.
Appendix (10) Gerpen, J. V. Biodiesel processing and production. Fuel Process.
Technol. 2005, 86, 1097–1107.
A. Typical Gas Chromatogram Figures. Several gas chro- (11) Seraphim, P.; Stylianos, F.; Michel, F.; Isabelle, C.; Maria, C. P.;
matogram records are listed in Figure 13. Figure 13 demonstrates Michael, K.; Ivan, M.; George, A. Biotechnological valorization of raw
glycerol discharged after bio-diesel manufacturing process: Production of
that the concentrations of 2,3-DCP and trichloropropane formed 1,3-propanediol, citric acid and single cell oil. Biomass Bioenergy 2008,
by further chlorination of DCP are much less than that of DCP 32, 60–71.
and that their concentrations are nearly unchanged with the (12) Jiang, J. X.; Zhang, P. P.; Yao, C. The development of the
increase of reaction time. Accordingly, formation of 2,3-DCP production of epichlorohydrin from glycerol. Mod. Chem. Ind. 2006, 26,
71–73 (in Chinese).
and trichloropropane will not be considered in the next study. (13) Krafft, P.; Gilbeau, P.; Balthasart, D. Crude glycerol-based product,
process for its purification and its use in the manufacture of dichloroproanol.
Acknowledgment
PCT Patent WO/ 2007/144335, 2007.
The authors thank Yantai Wanhua Polyurethanes Co., Ltd. (14) Zhao, X. J.; Bai, Z. L. Study on the process of producing
dichloropropanol from glycerol catalysted by organic acid. Chem. Intermed.
for financial support. They also thank the anonymous referees 2008, 15, 19–22 (in Chinese).
for comments on this manuscript. (15) Gilbeau, P.; Krafft, P. Producing chlorinated organic compounds
e.g. dichloropropanol involves using glycerol obtained from renewable raw
Nomenclature materials, as a starting product. FR Patent FR2868419, 2005.
A1, A1′, and A3 ) collision frequency factors (16) Krafft, P.; Gilbeau, P.; Gosselin, B.; Claessens, S. Process for
producing dichloropropanol from glycerol, the glycerol coming eventually
cA ) concentration of GLY in reaction system, mol · L-1 from the conversion of animal fats in the manufacture of biodiesel. PCT
cA0 ) initial concentration of GLY in reaction system, mol · L-1 Patent WO2005/054167 A1,2005.
cB ) concentration of HAC in reaction system, mol · L-1 (17) Kubicek, P.; Sladek, P.; Buricova, I. Method of preparing dichlo-
cB0 ) initial concentration of HAC in reaction system, mol · L-1 roproanols from glycerine. PCT Patent WO2005/021476 A1,2005.
(18) Schreck, D. J.; Kruper, W. J.; Varjian, R. D.; Jones, M. E.;
cC ) concentration of GLYA in reaction system, mol · L-1 Campbell, R. M.; Kearns, K.; Hook, B. D.; Briggs, J. R.; Hippler, J. G.
cD ) concentration of H2O in reaction system, mol · L-1 Conversion of a multihydroxylated-aliphatic hydrocarbon or ester thereof
cD0 ) initial concentration of H2O in reaction system, mol · L-1 to a chlorohydrin. PCT Patent WO2006/020234 A1, 2006.
cE ) concentration of HCl in reaction system, mol · L-1 (19) Krafft, P.; Franck, C.; Andolenko, I. D.; Veyrac, R. Process for
cE0 ) initial concentration of HCl in reaction system, mol · L-1 the manufacture of dichloropropanol by chlorination of glycerol. PCT Patent
WO/2007/054505, 2007.
cF ) concentration of MCP in reaction system, mol · L-1 (20) Chien, J. C. W. Criteria for diffusion limitation in coordination
cF0 ) initial concentration of MCP in reaction system, mol · L-1 polymerization. J. Polym. Sci. 1979, 17, 2555–2565.
cG ) concentration of MCPA in reaction system, mol · L-1 (21) Carrà, S.; Santacesarla, E.; Morbldlll, M. Synthesis of epichloro-
cH ) concentration of DCP in reaction system, mol · L-1 hydrin by elimination of hydrogen chloride from chlorohydrins: 1. simulation
of the reaction unit. Ind. Eng. Chem. Process Des. DeV. 1979, 18, 428–
E1, E1′, and E3 ) activation energies, kJ · mol-1
433.
k0 ) reaction rate constant, mol · L-1 · s-1 (22) Philip, B.; Bernard, T. G.; Gilles, L.; Hossein, L. K.; Reza, R. K.;
k1, k1′, k2, k3, and k4 ) reaction rate constants, mol · L-1 · s-1 Majid, M. S. Fluorination and chlorination of nitroalkyl groups. Tetrahedron
nA0 ) initial mole number of GLY in reaction system, mol 2007, 63, 11160–11166.
nE0 ) initial mole number of HCl in reaction system, mol (23) Chung, K. H.; Chang, D. R.; Park, B. G. Removal of free fatty
acid in waste frying oil by esterification with methanol on zeolite catalyst.
R ) ideal gas constant, 8.314 J · mol-1 · K-1 Bioresour. Technol. 2008, 99, 7438–7443.
t ) reaction time, s (24) Sharifov, G. S. Kinetics of chlorohydrination of allyl chloride. Kinet.
T ) absolute temperature, K Catal. 2001, 42 (5), 679–683.
T0 ) absolute temperature, 1K (25) Yadav, J. S.; Bhunia, D. C.; Krishna, K. V.; Srihari, P. Niobium(V)
pentachloride: an efficient catalyst for C-,N-,O-,and S-nucleophilic substitu-
Literature Cited tion reactions of benzylic alcohols. Tetrahedron Lett. 2007, 48, 8306–8310.
(26) Michael, C. D.; Huang, C. H. Aqueous chlorination of the
(1) Msameer, A. E. H.; Maciej, J. B.; Zuhoor, I.; Huda, H.; Mohammed, antibacterial agent trimethoprim: Reaction kinetic and pathways. Water Res.
A. T. Rapid and simple determination of chloropropanols (3-MCPD and 2007, 41, 647–655.
452 Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009

(27) Sakthivel, A.; Nakamura, R.; Komura, K.; Sugi, Y. Esterification (30) Päivi, M. A.; Tapio, S.; Erkki, P. Kinetics of the chlorination of
of glycerol by lauric acid over aluminium and zirconium containing propanoic acid in the presence of chlorosulphonic acid. Chem. Eng. Sci.
mesoporous molecular sieves in supercritical carbon dioxide medium. J. 1995, 50, 2275–2288.
Supercrit. Fluid. 2007, 42, 219–225.
(28) Calvar, N.; González, B.; Dominguez, A. Esterification of acetic
acid with ethanol: Reaction kinetics and operation in a packed bed reactive ReceiVed for reView July 19, 2008
distillation column. Chem. Eng. Process. 2007, 46, 1317–1323. ReVised manuscript receiVed October 14, 2008
(29) López, D. E.; Goodwin, J. G., Jr.; Bruce, D. A.; Furuta, S. Accepted October 16, 2008
Esterification and transesterification using modified-zirconia catalysts. Appl.
Catal. A: Gen. 2008, 339, 76–83. IE8011177

Vous aimerez peut-être aussi