Sunteți pe pagina 1din 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257373906

Esterification of acrylic acid with 2-ethylhexan-1-ol: Thermodynamic and


kinetic study

Article  in  Applied Catalysis A General · January 2013


DOI: 10.1016/j.apcata.2012.11.018

CITATIONS READS

17 772

4 authors, including:

Tomasz Komoń P. Oracz


Instytut Chemii Przemysłowej University of Warsaw
3 PUBLICATIONS   17 CITATIONS    71 PUBLICATIONS   654 CITATIONS   

SEE PROFILE SEE PROFILE

Małgorzata E. Jamróz
Instytut Chemii Przemysłowej
16 PUBLICATIONS   170 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Vapor-liquid equilibria of binary systems. View project

All content following this page was uploaded by Tomasz Komoń on 24 November 2017.

The user has requested enhancement of the downloaded file.


Applied Catalysis A: General 451 (2013) 127–136

Contents lists available at SciVerse ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Esterification of acrylic acid with 2-ethylhexan-1-ol: Thermodynamic


and kinetic study
Tomasz Komoń a , Piotr Niewiadomski a , Paweł Oracz b , Małgorzata E. Jamróz a,∗
a
Industrial Chemistry Research Institute, Rydygiera 8, 01-793 Warsaw, Poland
b
Faculty of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland

a r t i c l e i n f o a b s t r a c t

Article history: The activity and selectivity of polystyrene sulfonic acid resins (Amberlyst 39, Amberlyst 46, Amberlyst
Received 29 June 2012 70, and Amberlyst 131) were measured in the reaction of 2-ethylhexyl acrylate formation. The kinetics
Received in revised form 8 November 2012 of the esterification of acrylic acid with 2-ethylhexan-1-ol was studied using Amberlyst 70 ion exchange
Accepted 9 November 2012
resin (1–10 wt%) as a catalyst superior over a group of commercial catalysts. The 7:1, 5:1, 3:1, 1:1, 1:3,
Available online xxx
1:5, and 1:7 alcohol to acid molar ratio was used to carry out the reaction in an isothermal batch reactor
at temperatures from 353 to 393 K. The studied mixtures are highly non-ideal, therefore, the kinetic
Keywords:
equations and equilibrium parameters were expressed in terms of activities.
Acrylic acid
2-Ethylhexan-1-ol © 2012 Elsevier B.V. All rights reserved.
Esterification
Kinetic
Ion exchange resin

1. Introduction increases selectivity toward the ester; facilitates product separa-


tion, recovery, and recycling of the catalysts.
Likewise many other acrylates, 2-ethylhexyl acrylate is widely The importance of ester production over solid catalysts comes
used in the production of homopolymers. Also, it is used to pro- from both streamlining of industrial processes and better protec-
duce a multitude of different copolymers, such as those with acrylic tion of the environment. The solid catalysts in the esterification
acid and its salts, amides, methacrylates, acrylonitriles, vinyls (pre- reaction of acids with alcohols must be active in the presence
dominantly vinyl acetate and vinyl chloride), styrene and butadiene of water, which is formed during the course of the reac-
[1,2]. Both, homopolymers and copolymers are mainly processed in tion. However, generally, the solid acids are less or even not
aqueous emulsions and polymer dispersions to produce coatings, active in water, and only few materials have been found to
adhesives, printing inks, and binders for paints. They are also be water-tolerant. The following catalysts satisfy this important
applied in manufacturing of elastomers, super absorbent polymers, feature: organic functional polymers, H-ZSM-5, salts of het-
flocculants, as well as fibers and plastics [3,4]. eropolyacids, some sulfated oxides, and other solid oxides [6,7].
2-Ethylhexyl acrylate is clear, volatile liquid, slightly soluble in Among the heteropoly compounds, the water-insoluble acidic
water and completely soluble in alcohols, ethers, and most organic Cs2.5 H0.5 PW12 O40 salt, exhibits high activity. Okuhara et al. [7]
solvents. Commercially, this important ester is produced from 2- reported the following activity order in the liquid-phase esterifica-
ethylhexan-1-ol (EHOH) and acrylic acid (AA) by esterification in tion of acrylic acid with butan-1-ol: Nb2 O5 < H-ZSM-5 < Amberlyst
the presence of homogeneous acid catalysts such as sulfuric or 15 < Cs2.5 H0.5 PW12 O40 ∼ SO4 2− /ZrO2  Nafion. Indeed, H-ZSM-5
p-toluenesulfonic acid [5]. However, the use of strong acids in a was utilized in ASAHI commercial process for hydration of cyclo-
homogeneous system has serious negative consequences such as hexene in excess water. Yet, the industrial use of H-ZSM-5 in water
corrosion, pollution of the environment, and loss of catalysts. On the has not been widespread because of specific zeolite structure of
other hand, in the liquid phase organic synthesis, the use of acidic pores and weakened acidity of H-ZSM-5 in water.
heterogeneous catalysts is beneficial, since it eliminates corrosion; Use of organic cross-linked functional polymers as catalysts
has attracted considerable scientific interest [8–13]. The two main
classes of sulfonated ion exchange resins are widely applied: one
based on polystyrene/divinylbenzene matrix (Amberlyst, Dowex
∗ Corresponding author. Tel.: +48 22 568 20 21; fax: +48 22 568 22 33. type resins) and the other based on perfluorinated sulfonic acid
E-mail address: malgorzata.jamroz@ichp.pl (M.E. Jamróz). resins like Nafion and Aciplex. They are particularly relevant

0926-860X/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apcata.2012.11.018
128 T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136

reported very scarcely. Fomin et al. [20] investigated this very reac-
Nomenclature tion in the presence of macroporous acidic KU-23 and F-4SK ion
exchange resins. Their experimental data showed that the highest
AA acrylic acid activity of KU-23 was achieved due to the highest acid capac-
EHOH 2-ethylhexan-1-ol ity of this resin. Use of the ion exchange resins to production
EHA 2-ethylhexyl acrylate of 2-ethylhexyl acrylate has been described almost exclusively in
W water patents [21,22].
dpore pore diameter, nm Amberlysts are commercially available [23], reusable, and
dparticle particle diameter, mm non-hazardous heterogeneous catalysts consisting of styrene-
Vsp volume of the swollen polymer phase, cm3 g−1 divinylbenzene (PS/DVB) matrix with anchored sulfonic acid
Tmax maximum operating temperature, K groups as active sites [10]. They are characterized by relatively
T absolute temperature, K high concentration of acid sites but poor thermal stability. There-
Ka thermodynamic equilibrium constant of reaction fore, they can be applied in the low temperature processes (up
Kx apparent equilibrium constant of the reaction to 423–463 K). The strength and concentrations of acid sites of
G0 standard Gibbs energy of reaction, J mol−1 Amberlyst resins were characterized by Siril et al. [24]. For the halo-
H0 standard enthalpy of reaction, J mol−1 genated (Amberlyst 70) and over-sulfonated (Amberlyst 35 and 36)
R universal gas constant, 8.31 J mol−1 K−1 ion exchange resins, they observed the acid strength higher than
ai activity of the ith component that of the conventional sulfonated Amberlyst 15, yet lower than
i stoichiometric factor for the ith component that of Nafion. The concentration of resin acid sites depends on
xi mole fraction of the ith component the degree of sulfonation and varies from ca. 1 to 5 mmol H+ /g for
i activity coefficient of the ith component Amberlyst 46 and Amberlyst 35, respectively. The different struc-
xi0 apparent equilibrium constant of the reaction tural properties of the acid resins were tested by Casas et al. [25].
ci concentration of the ith component, mol dm−3 They presented data on the volume of the swollen polymer phase
ka rate constant calculated with activities, min−1 in water (Vsp ) obtained by inverse steric exclusion chromatography
kx rate constant calculated with mole fractions, min−1 ISEC, and polymer fraction density (PFD) which are the impor-
kc rate constant calculated with molar concentration, tant parameters of ion exchange resins, particularly if the reaction
dm3 mol−1 min−1 medium is highly polar.
k0,a so-called frequency factor calculated with activities, The kinetic studies of the esterification of acrylic acid with 2-
min−1 ethylhexan-1-ol have been rarely reported as well. Nowak [26]
k0,x so-called frequency factor calculated with mole and Grzesik et al. [27] studied the kinetics of liquid phase syn-
fractions, min−1 thesis of 2-ethylhexyl acrylate catalyzed by sulfuric acid. Based on
k0,c so-called frequency factor calculated with molar the amount of catalyst and the concentrations of acid and alco-
concentration, dm3 mol−1 min−1 hol, they proposed the second order kinetic equation. The authors
Ea reaction activation energy, kJ mol−1 did not take into account the non-ideality of the reaction mix-
ture. On the other hand, Fomin et al. [20] described the kinetics of
the same reaction catalyzed by a macroreticular acid ion exchange
for etherification, esterification, hydration/dehydration, alkylation, resins of diverse ion exchange capacity. It was stated that the activ-
and condensation reactions. Nafion-type resins are classified as ity of the resins used increased with their ion exchange capacity.
superacids. They have to be used in the presence of polar medium Moreover, the activation energy and the enthalpy of the studied
because the swelling of Nafion is essential for the catalyst activity. reaction were reported to be of the order of 72.8–77.8 kJ mol−1 and
Water and lower alcohols are almost the only solvents capable of 70.1–74.6 kJ mol−1 , respectively.
swelling of Nafion pellets. The knowledge of the chemical equilibrium constant, heat
The sulfonated resins have been found to be generally ade- of reaction, and kinetic parameters (activation energy, reaction
quate for esterification of the acrylic acid with C2 –C4 alcohols. rate constant) are necessary for technology development. So far,
The esterification of acrylic acid with butan-1-ol in the presence the kinetics of reaction over the ion-exchange resins has been
of many solid acid catalysts was studied by Chen et al. [14]. They described in terms of quasi-homogeneous [14,28–31] or quasi-
found the Amberlyst 15 and Nafion catalysts to be effective for heterogeneous systems [31–34]. The use of quasi-homogeneous
acrylate production. Esterification of acrylic acid with propylene model can be rationalized by the fact that acid ion exchange resins
glycol over Amberlyst 15, Amberlyst 36, and cesium salt of phos- swell in contact with water and polar reagents (alcohol), making
phorous tungstic acid was studied by Altiokka and Ödeş [15]. sulfonic groups of the catalyst readily accessible to the reactants.
They found that Amberlysts exhibited the highest catalytic activity Then, the mechanism of the reaction is very similar to that cat-
and, for Amberlyst 15, they determined the esterification kinetics alyzed by liquid acids and a pseudo-homogeneous solution model
[15]. can be applied.
Acrylic acid esterification with but-1-ene at 343 K was investi- The main objective of this study has been to develop the
gated in the presence of Mn- and Fe-promoted sulfated zirconia technology of catalytic synthesis of 2-ethylhexyl acrylate in the
[16]. It was concluded that, in comparison to Amberlyst resins, presence of solid heterogeneous catalyst. Thus, special atten-
the presence of Mn- and Fe-improved neither catalytic activity tion has been paid to the economic aspects of the catalyst
nor selectivity to sec-butyl acrylate, but it suppressed the catalyst selection. Therefore, the ion exchange resins based on the
deactivation. PS/DVB matrix, such as Amberlyst, were taken into account. In
The equilibrium and kinetics of the butyl acrylate synthesis, this work, two types of sulfonated resins, gelular and macro-
catalyzed by acid ion exchange resin (Lewitat K 2621), were inves- porous (called macroreticular) were used. Comparison of the
tigated by Schwarzer and Hoffmann and applied to the simulation catalyst activity and selectivity toward 2-ethylhexyl acrylate
of the reactive distillation [17]. Polystyrene sulfonic acid resins are enabled the selection of the best esterification catalyst. There-
recommended for many esterification processes [5,12], in particu- after, the thermodynamic and kinetic studies of esterification
lar, for processes with the reactive distillation [17–19]. However, was conducted in the presence of the best selected cata-
the esterification of acrylic acid with 2-ethylhexan-1-ol has been lyst.
T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136 129

2. Experimental introduced errors in the determination of equilibrium constants.


To prevent this, it was necessary to prepare the initial mixture
2.1. Catalysts and materials composition close to that of the equilibrium state. Therefore, after
preliminary kinetic experiments, samples of different reagents in
Acrylic acid (>99.8 wt%), 2-ethylhexan-1-ol (>99.5 wt%) and ratios close to the ratio predicted for the equilibrium state from
Nafion NR50 were purchased from Sigma–Aldrich and were used both sides, above and below, were prepared and measured. Thus,
as supplied. Phenothiazine (≥98 wt%, Fluka) and 4-methoxyphenol the equilibrium state was reached from the two sides allowing for
(≥98 wt%, Fluka) were used as free-radical polymerization error minimization and unequivocal determination of the equilib-
inhibitors. Undesirable free-radical polymerization of acrylic acid rium composition.
and its ester can occur in the reaction condition. Therefore, Equimolar amounts of acrylic acid and 2-ethylhexan-1-ol were
polymerization inhibitors must be added. The effect of the poly- taken to calculate the starting mixture for the above mentioned
merization inhibitors studied elsewhere [35] led to the conclusion experiments. The reagents (of 0.35 mol each), polymerization
that a mixture of the two inhibitors produced a synergistic inhibi- inhibitors: phenothiazine and 4-methoxyphenol (of 0.1 wt% each,
tion of the studied polymerization. The most profitable result was per acrylic acid and ester), and 10 wt% of catalyst were loaded in the
obtained for the same 0.1 wt% content of the two inhibitors. reactor. The equilibrium was determined in temperatures from 333
The Amberlyst catalysts (A39, A46, A70, and A131), used in this to 373 K, above which the mixture was boiling. Amberlyst 70 and
study, were kindly delivered by Rohm and Haas Co. (France). These Nafion NR50 were used as catalysts, and, additionally, the reaction
catalysts were first washed with deionized water as long as pH of without any catalyst was carried out near the equilibrium state. It
supernatant liquid became neutral, to eliminate free acids. Next, for was found that 6 hours was long enough to reach and stabilize the
48 hours, the catalysts were dried at 363 K at atmospheric pressure equilibrium state. Samples were collected every 1 hour and ana-
and then, were kept for 24 hours over P2 O5 and under vacuum. lyzed by GC. In the presence of catalyst, negligibly small amounts
of by-products were observed, whereas without catalyst no traces
2.2. Analysis of by-products were formed.

The reaction products were analyzed with Agilent GC Model


3.3. Kinetic measurements
6980 gas chromatograph equipped with a flame ionization detec-
tor. The desired chromatographic peak separation was achieved
All the experiments were performed in 3-necked typical batch
by using Innowax polar capillary column (60 m, Ø = 0.25 mm). The
reactor of 100 ml capacity with jacket, magnetically agitated and
analysis was carried out as follows: the initial temperature of col-
fitted with a reflux condenser, sampling device, and thermome-
umn was 433 K, kept for 8 min, and then increased by 20 K min−1 , up
ter. The temperature was maintained within ±0.1 K by circulating
to 513 K was reached. The injection and detection chamber temper-
ethylene glycol from thermostat into the jacket of the reactor. In
atures were set at 523 K and helium flow was set at 1.4 ml min−1 .
a typical procedure, a known amount of alcohol and the catalyst
The method of the addition of internal standard (butyl acrylate)
(usually 5 wt%) were loaded to the reactor and heated to the desired
was used to determine concentration of all the components. Water
temperature. Next, acrylic acid was introduced and the measure-
present in the liquid phase was measured by Karl Fischer titration
ment of the reaction time was started. To avoid polymerization
(Metrohm KF-784).
reaction, phenothiazine and 4-methoxyphenol (of 0.1 wt% each, per
Nitrogen porosimetry was determined by using Micrometrics
acrylic acid) were added to the reagents. It was determined that
ASAP 2020. Adsorption-desorption isotherms were recorded at
to eliminate external mass transfer limitations, it is sufficient to
77 K after degassing at 373 K for 6 h. The surface areas and mean
continuously stir the reaction mixture at the rate of 500 rpm. To
pore diameters were calculated by the BET and BJH method, respec-
perform the GC analysis of the reaction mixture and to avoid any
tively.
further reaction, the collected samples were cooled down rapidly
The number of sulfonic acid groups in all ion exchange resins
to 293 K in ice bath. The tests were carried out at temperatures from
was determined using an acid-base titration. For two hours, a sam-
the 353–393 K interval. The following initial molar ratios of acrylic
ple of each catalyst was stirred with 0.1 M water solution of NaCl
acid to 2-ethylhexan-1-ol were used: 1:7, 1:5, 1:3, 1:1, 3:1, 5:1, and
and was left to the next day, when it was titrated with NaOH
7:1.
(0.01 M). The analysis was repeated three-times.

3. Methods 4. Results and discussion

3.1. Catalyst selection 4.1. Catalyst characterization

The ion exchange resins, Amberlyst 39, Amberlyst 46, Amberlyst Properties of the used catalysts are summarized in Table 1. In
70 and Amberlyst 131 have been tested in the esterification of general, two types of ion exchange resins were tested: macro-
acrylic acid with 2-ethylhexan-1-ol at 373 K and reactants molar porous (Amberlyst 39, Amberlyst 46, Amberlyst 70) and gelular
ratio of 1:1 (of 0.3 mol each) with the same amount of catalyst one (Amberlyst 131). They differ by their porous structure and
(5 wt%) and inhibitors: phenothiazine and 4-methoxyphenol (of physicochemical properties. The macroporous resins have higher
0.1 wt% each, per acrylic acid). The esterification was carried out in cross-linking degree of divinylbenzene (≥8%). Therefore, they are
a magnetically agitated three-neck flask with condenser and oper- characterized by pore structure in the dry state and they have
ated in batch mode. Samples of reaction mixture were taken every much more polymer matrix density than the gel type resins which
1 hour and analyzed by GC and Karl Fischer method. The reaction are low cross-linked (4% of DVB) and have no permanent pores.
was performed for 6 hours. Amberlyst 46 is characterized by high cross-linked degree, thus, it
practically does not have any gel microspheres. Additionally, it is
3.2. Chemical equilibrium measurements only surface sulfonated in contrast to fully sulfonated (convention-
ally) Amberlyst 39, 46 and 70 (one sulfonic group per styrene/DVB
In order to reach the equilibrium state, a long time of reac- ring). In consequence, the acid capacity of Amberlyst 46 is remark-
tion was required. As a result, many by-products were formed and ably decreased. The ion exchange capacity of Amberlyst 39 and
130 T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136

Table 1
Properties and structural parameters of tested acid resins.

Catalyst Short name Typea DVBb % Acidity (meq H+ g−1 ) Dry state Swollen in water [25] Tmax (K)

SBET c (m2 g−1 ) dpore (nm) dparticle (mm) Vsp (cm3 g−1 ) PFDd (nm nm−3 )

Amberlyst 39 A39 M 8 4.50 01.07 08.0 0.6–0.7 1.45 0.8–1.5 403


5.00 [23] 00.09 [25] 17.6 [25]
32 [23] 23.0 [23]

Amberlyst 46 A46 M High 0.98 63 22.6 0.8–0.9 0.16 1.5–2.0 393


0.43 [23] 57.4 [25] 19.2 [25]
75 [23] 23.5 [23]

Amberlyst 70 A70 M 8 2.61 00.24 09.8 0.5 1.40 0.4–0.8 463


3.01 [25] 00.02 [25] 22.0 [23]
2.55 [23] 36 [23]

Amberlyst 131 A131 G 4 4.57 – – 0.4–0.5 – – 403


4.80 [23]
a
M – macroreticular; G – gel type.
b
Divinylbenzene content.
c
BET surface area.
d
Polymer fraction density.

Amberlyst 131 are similar (Table 1) whereas Amberlyst 70 exhibits olefins, ether) can occur, respectively [38]. Almost all of these
a lower acid capacity because of halogenation of the polymer by-products were detected but only the significant amounts
matrix. On the other hand, Amberlyst 70 and Amberlyst 39 have (exceeding 0.05 wt%) of Michael adducts and 2-ethylhexyl 3-
very similar surface area and porosity. It is important that their BET hydroxypropionate were determined in the reaction mixture.
surface area and pore volume, determined in this study, are consis- Comparison of the conversion of acrylic acid (after 6 h of the
tent with results obtained by the other researchers [25,36,37], but reaction) and selectivity toward 2-ethylhexyl acrylate indicates
they are much more lower than data specified by the manufacturer that both the highest conversion and the highest selectivity are
[23]. Comparison of acid strength of chlorinated (A70) and conven- achieved with Amberlyst 70 (Table 2). The conversion, selectivity,
tional sulfonated resin (A15) was done by Siril et al. [24]. Based on and yield are defined as follows:
the vicinity of the sulfonic groups in the resins, it can be assumed
mole of AA reacted
that the heats of NH3 adsorption by A39, A46, and A131 are simi- conversion % = × 100
lar to A15 (−111 ± 2 kJ mol−1 ), whereas that of A70, in which the Cl mole of AA initially
atoms are accompanying the sulfonic groups, is greater and equal to
−117 ± 2 kJ mol−1 . The presence of Cl atoms in the resin, increases mole of EHA
selectivity % = × 100
also its thermal resistance, so A70 resin can be used in temperature mole of the product sum
to 463 K.
Another important matter is the resin swelling. The macropo-
mole of EHA
rous resins have both the permanent pores in the dry state and yield % = × 100
mole of AA initially
microspheres shape in the gel form. However, the gel type resins
pores are formed only in the presence of a polar medium. There- The reaction mixture is highly polar due to the presence of alco-
fore, volume of the swollen polymer phase (Vsp ), which indicates hol and water formed in the reaction. This enables the applied resins
how large space is formed between the polymer chains, is a very to be swollen by the reaction medium and makes its active acidic
considerable parameter characterizing the accessibility of the acid sites to be accessible for the reagents. The used resins differ in
centers of the catalyst. The Vsp values of the applied catalysts were composition (content of DVB), structure, and properties such as
determined in Refs. [25] and [36,37]. ion exchange capacity, specific surface, and specific volume of the
swollen polymer phase Vsp . These differences are reflected in the
resins catalytic activity.
4.2. Selection of the catalyst

The ion exchange resins, Amberlyst 39, Amberlyst 46, Amberlyst 80


70, and Amberlyst 131 have been tested for their catalytic perfor-
70
mance in the esterification of acrylic acid with 2-ethylhexan-1-ol.
The time courses of the reaction over various Amberlyst-type ion 60
exchange resins are presented in Fig. 1. It is noteworthy that, out of
yield %

50
the studied resins, the highest yield of 2-ethylhexyl acrylate over
the whole time of the reaction was observed for Amberlyst 70. In the 40
presence of A131 the highest quantity of by-products were formed
30 A39
since yield of acrylate was decreased in the course of reaction time.
In the esterification conditions, the unreacted alcohol and 20 A46
acrylic acid can undergo side reactions: first and foremost the A70
10
Michael addition to the ester unsaturated C C double bond A131
but also the self-addition of acrylic acid. The adducts are 0
0 50 100 150 200 250 300 350 400
the following: 2-ethylhexyl 3-(2-ethylhexyloxy)propanoate, 2-
t, min
ethylhexyl 3-acryloxypropanoate, and 2-carboxyethyl acrylate (AA
dimer). Additionally, the hydration and dehydration reactions of Fig. 1. Time courses of the reaction over various catalysts (5 wt%) at 373 K, 0.2 wt%
ester (2-ethylhexyl 3-hydroxypropanoate) and alcohol (diverse inhibitors, AA (0.3 mol), and EHOH (0.3 mol).
T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136 131

Table 2
Conversion, selectivity and yield for esterification over various catalysts.

Catalyst Conversion % Yield % Selectivitya %

EHA EHHP EHAP EHEP AA dimer

A39 68.9 62.2 90.3 4.1 3.4 1.9 0.2


A46 64.5 58.4 90.5 2.5 3.4 3.3 0.3
A70 76.0 71.6 94.2 2.1 1.7 1.8 0.2
A131 62.2 55.1 88.6 4.8 3.7 2.7 0.2
a
EHHP – 2-ethylhexyl 3-hydroxypropanoate; EHAP – 2-ethylhexyl 3-acryloxypropanoate; EHEP – 2-ethylhexyl 3-(2-ethylhexyloxy)propanoate; AA dimer – 2-carboxyethyl
acrylate.
The reaction conditions: AA (0.3 mol), EHOH (0.3 mol), inhibitors = 0.2 wt%, catalyst loading = 5 wt%, reaction time = 360 min, 373 K.

Table 3
Mole fractions and the evaluated activity coefficients of components in the equilibrium state of the reaction at various temperatures.

T (K) Acrylic acid 2-Ethylhexan-1-ol 2-Ethylhexyl acrylate Water

Mole fraction Activity coefficient Mole fraction Activity coefficient Mole fraction Activity coefficient Mole fraction Activity coefficient

333.2 0.1690 0.76425 0.1690 1.19570 0.3310 1.54687 0.3310 3.60132


343.2 0.1505 0.75357 0.1505 1.20761 0.3495 1.52913 0.3495 3.59361
353.2 0.1345 0.74549 0.1345 1.21780 0.3655 1.51218 0.3655 3.57928
363.2 0.1185 0.73830 0.1185 1.22828 0.3815 1.49530 0.3815 3.56199
373.2 0.1015 0.73146 0.1015 1.23996 0.3985 1.47831 0.3985 3.54306

The reaction conditions: AA (0.35 mol), EHOH (0.35 mol), 0.2 wt% inhibitors, catalyst: A70 (10 wt%) or Nafion (10 wt%) or none.

A39 and A70 differ in the number and strength of the acid cen- suggests that the spaces formed between the chains of swollen
ters and also in the density of the swollen polymer phase but they polymer are not sufficiently wide for diffusion of the branched
have similar volume of the pores in the swollen state Vsp . The high- chain molecules. As a result, Amberlyst 70 of the highest activity
est activity of Amberlyst 70, with low concentration of the acid and the best selectivity was chosen for further investigations. The
sites, is probably due to the acidic strength of these sites addi- by-products generated during esterification are serious problem in
tionally enhanced by the chlorine atom in the matrix of the resin industry. Thus, the main factor taken into account in the catalyst
[24]. Moreover, A70 has lower polymer density and larger space selection was the highest selectivity of A70, nevertheless, the price
between polymer chains than A39, hence, it shows higher accessi- of the A70 is higher than that of the other catalysts of this type.
bility of acid centers for the reactants (Table 1). All gel microspheres In this way a significant reduction of side-product formation was
of A70 have polymer density about 0.4–0.8 mm−2 while in A39 a achieved.
considerable dispersion is observed. The main part of A39 gel phase
is characterized by polymer density in the range of 0.8–1.5 mm−2 4.3. Chemical equilibrium
which corresponds to small pore spaces. However, there are some
polymer fractions of smaller density (0.2–0.4 mm−2 ) in which the The esterification of acrylic acid with 2-ethylhexan-1-ol runs
pores formed between polymer chains are much wider. Summing according to the following scheme (1):
CH3
O + O
H
H2C
OH
+ H3C OH H2C
O CH3 + H2O

CH3
AA EHOH EHA W (1)

up, a less steric hindrance for diffusion of molecules inside the The thermodynamic equilibrium constant of reaction, Ka is given
pores of A70 accounts for the improvement of the reaction selec- by Eq. (2).
tivity. Similar behavior of these Amberlyst catalysts used in ether
   
G0
synthesis were reported by the other authors [25]. Ka = exp − = aivi = (xi i )vi (2)
RT
On the other hand, despite the fact that A46 is sulfonated only i
at the polymer surface and exhibits very low concentration of The apparent equilibrium constant of the reaction, Kx , expressed
acidic sites, while A39 has 5-fold higher concentration of the acidic in terms of mole fractions can be written as follows (3):
centers, the catalytic activity of the two catalysts is quite similar
(Table 2). This suggests that a number of sulfonic groups in A39 is
 xEHA xW 2
xEHA
Kx = xivi = = 0 0
(3)
inaccessible to the reagents. This is probably due to too close pack- xAA xEHOH (xAA − xEHA )(xEHOH − XEHA )
ing of the polymer chains in A39, and thus, in practice the number
of active sites of A46 and A39 is similar. In turn, for A131, which The equilibrium constant can be found experimentally from the
has only the gelular structure and has the ion exchange capac- concentrations of reactants and products determined at the equi-
ity similar to A39, conversion of acrylic acid was slightly lower. librium conditions. Correct determination of the thermodynamic
It is caused by the lack of the macroporous structure in A131 and equilibrium constant of real mixtures requires the activities of the
resulting in lower swelling of the resin in reaction medium. It addi- compounds to be determined instead of their concentrations. Thus,
tionally confirms the previously mentioned fact that in A39 not all the activity coefficients can be evaluated by using one of the local
gelular microspheres are swollen enough. Moreover, the decrease composition models, such as NRTL or UNIQUAC model, assum-
in selectivity of the reaction in presence of Amberlyst 131 also ing that for all the constituent binary subsystems the relevant
132 T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136

Table 4
The chemical mole fraction and activity based equilibrium constants (Kx and Ka , Eq. (7), respectively), standard enthalpies of reaction (Hr0 and their standard errors (Hr0 ),
and the degree of equilibrium conversion of acrylic acid (˛x ).

T (K) Kx Hr0 (kJ mol−1 ) (Hr0 ) (kJ mol−1 ) Ka Hr0 (kJ mol−1 ) (Hr0 ) (kJ mol−1 ) ˛x

333.2 3.8 25.8 (2.5) 23.4 25.1 (2.5) 0.662


343.2 5.4 30.6 (1.5) 32.6 29.6 (1.6) 0.699
353.2 7.4 35.6 (0.7) 44.0 34.2 (0.8) 0.731
363.2 10.4 40.7 (0.9) 60.9 38.9 (0.9) 0.763
373.2 15.4 46.1 (1.8) 89.0 43.8 (1.9) 0.797

interaction parameters are available (or the VLE data at the deviations are the following: (Kx ) = 0.13 and (Ka ) = 0.82. Varia-
pertinent temperature range are known). Unfortunately, careful tion of Kx and Ka with temperature indicates that the latter is more
inspection of the available databases shows significant gaps in data sensitive to the temperature changes (Fig. 2).
sets. In such a case, one among the group-composition method can The thermodynamic equilibrium constant is related to the
be used. Here, we have selected the UNIFAC method, version 6 by standard enthalpy of reaction, Hr0 , by the van’t Hoff equation (6):
Wittig et al. [39]. Applicability of this very version was checked
 d ln K  Hr0
using all the available literature data for the constituent binary sub- a
= (6)
systems and/or for similar systems. The obtained results appeared dT p RT 2
to be satisfactory enough to accept the UNIFAC method for fur-
ther use. The calculated activity coefficients and the experimental The enthalpy of reaction, calculated using Eqs. (4) and (6) for
mole fractions of reactants at the equilibrium state for each of the both, the apparent and thermodynamic equilibrium constants is
mixture components equilibrated in selected definite temperatures given by the formula (7):
from the range of 333–373 K are collected in Table 3. The apparent
and the thermodynamic equilibrium constants calculated based on Hr0 = −R(b2 − b3 T 2 ) (7)
Eqs. (2) and (3) are gathered in Table 4 and plotted against temper-
Table 4 reports values of the enthalpy of reaction calculated
ature in Fig. 2.
for discrete temperatures together with their estimated standard
The temperature dependence of the apparent and thermody-
errors (calculated using data reported in Table 5).
namic equilibrium constants can be described by the following
Notice, that because of the endothermicity of the reaction, the
function (4):
  equilibrium of the reaction is shifted to the products as the temper-
b2 ature is increased. The calculated enthalpy of the reaction at 373 K
K = exp b1 + + b3 T (4)
T is equal to 46.0 ± 1.81 kJ mol−1 (Kx ) and 43.8 ± 1.86 kJ mol−1 (Ka ).
where bi is the ith adjustable parameter. Different results were obtained by Fomin et al. [19], who estimated
The parameters in this equation can be fitted to the experimen- the enthalpy of reaction to be 70.1–72.3 kJ mol−1 , as independent
tal data using the least squares method with the sum of squared on temperature in the range of 363–383 K.
deviations between experimental and calculated values evaluated For the sake of comparison, the enthalpy of the reaction was
for all experimental points as the objective function. As a measure also estimated from the appropriate combinations of standard
of the quality of the fit, the standard deviation can be used (Eq. (5)): enthalpies of formation f Hi0 (8):


 N

(K exp tl − K calc )2 Hr0 = vi f Hi0 − vi f Hi0 (8)
(K) = i i
(5) prod. substr.
(N − m)
i=1
The enthalpies of formation of each reagent in the liquid state
where N is the number of experimental points and m is the number at 298.15 K are gathered in Table 6.
of adjusted parameters. Taking into account the temperature dependence of the
The fitted parameters of Eq. (4), their standard errors and enthalpy of reaction at 298.15 K estimated in this work (Eq. (7)),
the correlation coefficients between the corresponding pairs of the calculated enthalpy is 9.9 kJ mol−1 and 10.5 kJ mol−1 , for Kx and
parameters are gathered in Table 5. The corresponding standard Ka , respectively. The enthalpy of reaction calculated by Eq. (8) was
found to be equal to 14.8 kJ mol−1 at 298.15 K. These values are not
100 KKx KKa far from the enthalpy calculated from the heats of formation.
90
80
Table 5
70
The bi parameters (Eq. (5)), their standard errors (bi ), and the correlation
60 coefficients between ij parameters for the apparent and thermodynamic equilib-
rium constants.
K

50
40 i 1 2 3
30 Thermodynamic equilibrium constant
20 bi −40.6749 5793.48 0.079435
(bi ) 12.68 2262 0.0177
10
1i −0.9999 −0.9999
0 2i −0.9994
330 340 350 360 370 380
Apparent equilibrium constant
T, K
bi −46.7641 6469.36 0.086200
(bi ) 12.41 2215 0.0174
Fig. 2. The temperature dependence of the apparent (Kx ) and thermodynamic (Ka )
1i -0.9999 −0.9999
equilibrium constant of the acrylic acid (0.35 mol) esterification with 2-ethylhexan-
2i −0.9994
1-ol (0.35 mol), catalyst: A70 (10 wt%) or Nafion (10 wt%) or none, 0.2 wt% inhibitors.
T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136 133

Table 6 0.0
Enthalpy of formation of the selected reagents according to [40]. 363 K

Compound Enthalpy of formation (kJ mol −1


) - 0.5 373 K

383 K
Acrylic acid −383.8

ln (CAA /C0AA )
2-Ethylhexan-1-ol −432.8 - 1.0 393 K
2-Ethylhexyl acrylate −516.0
Water −285.8
- 1.5
298.15 K, liquid state.
- 2.0
1.0
- 2.5
0 50 100 150 200 250
0.9 t, min

Fig. 5. First-order plots for the esterification of acrylic acid in the excess of alcohol
αAA

(AA = 0.06 mol, EHOH = 0.42 mol, A70 = 5 wt%, 0.2 wt% inhibitors).
0.8

1:1 1:2
0.7 rate of the ester formation linearly depends on the catalyst con-
1:3 1:5 centration (Fig. 6). Thus, the first-order of the reaction with respect
1:7 to the resin concentration can be assumed for low and moderate
0.6
330 340 350 360 370 380 390 400 concentrations of Amberlyst 70.
T, K
4.5.3. Effect of temperature and initial molar ratio of reagents
Fig. 3. The temperature dependence of equilibrium conversion of acrylic acid (˛AA )
on the initial molar ratio of reactants (AA:EHOH) determined from experimental The time runs of the reaction over Amberlyst 70 were performed
equilibrium constant Kx . at 353, 363, 373, 383, and 393 K (Figs. 7 and 8) and for the initial
molar ratio of reagents equal to 1:7, 1:5, 1:3, 3:1, 5:1, and 7:1. It
is clearly evident, that the reaction rate is affected by the temper-
4.4. Equilibrium conversion of acrylic acid
ature. Conversion of acrylic acid increased with both: the increase
of temperature and the molar excess of one of the reagents.
The influence of the initial molar ratio of reagents on the equilib-
rium conversion of acrylic acid at various temperatures (Fig. 3) was
determined from experimental equilibrium constant Kx . Results 4.5.4. Kinetic model
show, that the equilibrium conversion of acrylic acid decreases The reaction rate expression of the acrylate formation depends
with increase of the initial molar ratios. In particular, it is clearly on the mechanism of reagent adsorption on heterogeneous catalyst
shown that the temperature sensitivity of the acrylic acid conver- and is generally described as (9):
sion decreases with the alcohol excess.
ka (aAA aEHOH − (1/Ka )aEHA aW )
rEHA = 2
(9)
4.5. Kinetic study (1 + ˙Ki ai )

4.5.1. Reaction order where Ki – adsorption constant of the ith component.


The reaction order was determined in two series of independent According to the theory of ion exchange resin activity, in the
experiments run in an excess of one of the reagents. The reaction presence of at least one highly polar reactant initiating the swelling
was found to be of the first-order in both: the excess of acid as well of the resin, the active acid centers are easily accessible to reagents
as the excess of alcohol (Figs. 4 and 5). Hence, the overall reaction [41]. Then, the adsorbed reactants can penetrate very small gel-
is of the second-order as was previously found in Ref. [4]. type in microspheres while being in the equilibrium with the bulk
solution. Therefore, a quasi-homogeneous kinetic model can be
4.5.2. Effect of catalyst concentration applied to describe the rate of reaction catalyzed by Amberlyst 70
The catalyst concentration was varied from 2 to 10 wt% of dry ion exchange resin.
Amberlyst 70. The reaction was conducted at the temperature of
353 K and stoichiometric equimolar reagent mixture. The initial
0.020

0.5 353 K
- rAA0 , mol·dm-3·min -1

363 K 0.015
0.0 373 K
383 K
ln(CAA /C0AA )

-0.5 393 K 0.010

-1.0
0.005
-1.5

-2.0 0.000
0 100 200 300 400 0 2 4 6 8 10
t, min catalyst loading, wt%

Fig. 4. First-order plots for the esterification of acrylic acid in the excess of acid Fig. 6. The initial rate of acrylate formation against the A70 catalyst concentration
(AA = 0.77 mol, EHOH = 0.11 mol, A70 = 5 wt%, 0.2 wt% inhibitors). at 353 K for the equimolar (0.3 mol) amounts of reagents and 0.2 wt% inhibitors.
134 T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136

0.25 0.30
1:3 3:1
0.20 0.25

0.20
0.15
x EHA

xEHA
0.15
0.10 363 K
353 K
373 K 0.10
0.05 383 K 363 K
393 K 0.05 373 K
0.00 383 K
0 50 100 150 200 250 300 350 400 0.00
0 50 100 150 200 250 300 350 400
t, min
t, min
0.18
1:5 0.18
0.15 5:1
0.15
0.12
0.12
x EHA

0.09
363 K

xEHA
0.06 0.09
373 K
353 K
0.03 383 K 0.06
363 K
393 K
0 0.03 373 K
0 50 100 150 200 250 300 350 400 383 K
t, min 0.00
0 50 100 150 200 250 300 350 400
0.14 t, min
1:7
0.12
0.14
0.10 7:1
0.12
0.08
x EHA

0.10
0.06 363 K
0.08
xEHA

0.04 373 K
383 K 0.06 353 K
0.02
393 K 0.04 363 K
0.00
0 50 100 150 200 250 300 350 400 373 K
0.02
t, min 383 K
0.00
0 50 100 150 200 250 300 350 400
Fig. 7. The experimental and calculated kinetic runs for 2-ethylhexyl acrylate for-
mation at 363–393 K and for the initial molar excess of alcohol (AA:EHOH = 1:3, t, min
1:5, and 1:7), xEHA stands for mole fraction, points denote the experimental results,
dotted line represents calculated Eq. (11), solid line represents calculated Eq. (12), Fig. 8. The experimental and calculated kinetic runs for 2-ethylhexyl acrylate for-
A70 = 5 wt% and 0.2 wt% inhibitors. mation at the temperatures of 353–383 K and for the initial molar excess of acid
(AA:EHOH = 3:1, 5:1, and 7:1), xEHA stands for mole fraction, points denote the
experimental results, dotted line represents calculated Eq. (11), solid line represents
At first, the ideal quasi-homogeneous model (IQH) was exam- calculated Eq. (12), A70 = 5 wt% and 0.2 wt% inhibitors.
ined. For the second-order equilibrium reaction the kinetic
equation of IQH can be written as (10):
 1
 and temperature. Therefore, the individual rate constants for for-
rEHA = kc cAA cEHOH − cEHA cW (10)
Kx ward reactions were fitted to the experimental compositions of
acrylate with the internal numerical integration of Eq. (12). For
After integration Eq. (10) can be expressed as follows (11):
the integration procedure the Runge–Kutta method of the fourth
 0 + x0
√ 
2(1 − 1/Kx )xEHA − (xAA EHOH
)−  order was used. The relevant thermodynamic equilibrium constant
ln √ was calculated using Eq. (4) with the parameters given in Table 5.
0 + x0
2(1 − 1/Kx )xEHA − (xAA )+ 
EHOH
Data on the relative small conversion were used and the conver-
 0 √ 

0
−(xAA + xEHOH )−  sion range was individually selected depending on particular data
− ln √ = kx t  (11) set (see the next subsection).
0 + x0
−(xAA )+ 
EHOH The application of the IQH kinetic model leads to quite good

where  = 1 − 4 1 − K1x xAA0 x0 agreement of the calculated kinetic curves with the experimental
EHOH
However, because the reaction mixture is highly non-ideal, the results for the reaction mixtures with acrylic acid excess only for the
kinetics should be expressed in terms of activities (non-ideal quasi- initial times of the reaction (Fig. 8). However, for the longer reac-
homogeneous model – NIQH). tion times the experimental points are deviated from the calculated
 1
 functions. This is probably connected with non-ideality of the sys-
rEHA = ka xAA AA xEHOH EHOH − xEHA EHA xW W (12) tem and also the by-products formation. The description was better
Ka when the activity coefficients were used, but still worse for the
The required activity coefficients were calculated from the longer time of reaction. In contrast, the description of experimen-
group-contribution UNIFAC model. Activities (or equivalently tal findings by the IQH model is significantly worse in the systems
activity coefficients) are implicit function of overall composition containing alcohol excess in the whole range of the reaction time
T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136 135

-2.5 5. Conclusions
ln(k x )
Serie1 ln(k
ln(ka)a)
-3
Activity of different acidic ion exchange resins in the esterifi-
-3.5 cation of acrylic acid with 2-ethylhexan-1-ol was established. The
ln(k, min-1)

-4 Amberlyst 39, Amberlyst 46, Amberlyst 70, and Amberlyst 131 ion
exchange resins were tested. Among them, a noticeable acid con-
-4.5
version and selectivity toward 2-ethylhexyl acrylate were observed
-5 for A70. The best performance of the A70 was assigned to the
structural resin properties and to the strength of its acidic sites
-5.5
higher than in the other used catalysts. The accessibility of acid
-6 centers placed in the spaces between polymer chains for branched
0.0025 0.0026 0.0027 0.0028 0.0029 molecules was also a very important factor.
1/(T, K) Both A70 and Nafion NR50 were used as catalyst in the
thermodynamic part of the study, to make sure that the
Fig. 9. The linear correlation between ln(k) (min−1 ) and reciprocal temperature. The
A70 catalyst content is equal to 5 wt% of the reagent mixture and 0.2 wt% inhibitors. proper equilibrium state was achieved. On the other hand,
in the kinetics tests the A70 macroporous acid resin was
employed. The equilibria for 2-ethylhexyl acrylate synthesis,
(Fig. 7). In that model the calculated curves are placed well below determined for equimolar initial composition of the reac-
the experimental points. In turn, the model based on activities tants at temperatures from 333 to 373 K, allowed to find
(NIQH) allows to get a significantly better description of the kinetic out the temperature dependence of the concentration and
behavior in the systems containing alcohol excess. It is worth to activity equilibrium constants. The apparent (Kx ) and thermody-
stress that in these systems, the by-products were formed in minute namic (Ka ) equilibrium constants, were expressed according to
quantities. In contrast, the kinetic curve fitting by the NIQH model the following equations: ln(Kx ) = –46.7641 + 6469.36/T + 0.008620 T
in the acid excess systems is comparable with the fitting by the IQH and ln(Ka ) = –40.6749 + 5793.48/T + 0.079435 T, respectively. The
model. enthalpy of the reaction at 373.15 K was determined to be equal to
46.0 ± 1.8 kJ mol−1 (Kx ) and to 43.8 ± 1.4 kJ mol−1 (Ka ). Special care
was paid to the experimental procedure minimizing the formation
4.5.5. Reaction rate constant and activation energy of by-products.
The forward reaction rate constants were determined from the The kinetic of the acrylic acid esterification with 2-ethylhexan-
experimental measurements of the initial reaction rate (for conver- 1-ol was found to be of the second-order: i.e., of the first-order with
sion not higher than 40 wt%) for different initial concentrations of respect to each of the reagents in excess of the other. Moreover,
reagents and at the temperature range of 353–393 K. The variation the first-order of the reaction with respect to the resin concentra-
of the rate constant with the temperature can be described by the tion can be assumed as well. A quasi-homogeneous kinetic model
Arrhenius equation (13): was used to describe the reaction catalyzed by Amberlyst 70. Both,
 E  mole fractions and activities (derived from UNIFAC method) were
a
k = k0 exp − (13) taken into account to describe non-ideality of the liquid phase.
RT
The estimated kinetic parameters allowed for the determination
where k0 is the frequency factor, EA is the reaction activation of the activation energy to 52.3 ± 1.9 kJ mol−1 (using the mole frac-
energy. tions) and 50.1 ± 3.1 kJ mol−1 (using the activities). Description of
The ideal (IQH) and non-ideal (NIQH) quasi-homogeneous mod- the kinetic results by non-ideal quasi-homogeneous model was
els were applied to estimate kinetic parameters. The temperature satisfactorily.
dependences of reaction rate constants calculated using mole frac-
tion and activities are presented in Fig. 9. The activation energy and Acknowledgments
the frequency factor based on mole fractions, concentrations and
activities were determined by fitting the experimental data using This work was supported by the European Regional Devel-
the least squares method and the standard deviation (k, min−1 ) opment Fund POIG 01.03.01-00-010/08. The authors thank
are presented in Table 7. Rohm&Haas, France, for providing the Amberlyst catalysts.

References
Table 7
Parameters in the Arrhenius equation (13) with their standard errors (bi ) and
correlation coefficient between  parameters calculated with different models. [1] G. Klein, V. Le Houérou, R. Muller, C. Gauthier, Y. Holl, Tribol. Int. 53 (2012)
142–149.
Calculated with activities, (ka ) (min−1 ) = 0.0019 [2] Y. Peykova, O.V. Lebedeva, A. Diethert, P. Müller-Buschbaum, N. Willenbacher,
k0,a (min−1 ) EA,a (kJ mol−1 ) Int. J. Adhes. Adhes. 34 (2012) 107–116.
bi 1.409 × 105 50.12 [3] W. Bauer Jr., in: J.I. Kroschwitz, M. Howe-Grant (Eds.), Kirk-Othmer Encyclope-
dia of Chemical Technology, vol. 1, fourth ed., Wiley-Interscience, New York,
(bi ) 1.07 × 105 3.07
1991, pp. 287–314.
(k0,a , EA,a ) −0.9997
˛
[4] S. Kudła, M. Kałedkowska, Przem. Chem. 77/3 (1998) 86–91.
Calculated with mole fractions, (kx ) (min−1 ) = 0.00046 [5] J. Otera, J. Nishikido, Esterification Methods, Reactions and Applications, second
k0,x (min−1 ) EA,x (kJ mol−1 ) ed., Wiley-vch verlag GmbH & Co., KGaA, Weinheim, 2010.
bi 2.111 × 105 52.31 [6] Y. Izumi, Catal. Today 33 (1997) 371–409.
[7] T. Okuhara, M. Kimura, T. Kawai, Z. Xu, T. Nakato, Catal. Today 45 (1998) 73–77.
(bi ) 1.25 × 105 1.91
[8] M.M. Sharma, React. Funct. Polym. 26 (1995) 3–23.
(k0,x , EA,x ) -0.9997
[9] B. Corain, M. Zecca, K. Jeřábek, J. Mol. Catal. A: Chem. 177 (2001) 3–20.
Calculated with molar concentrations, (kc ) (dm3 mol−1 min−1 ) = 0.000063 [10] M.A. Harmer, Q. Sun, Appl. Catal. A: Gen. 221 (2001) 45–62.
k0,c (dm3 mol−1 min−1 ) EA,c (kJ mol−1 ) [11] P. Barbaro, F. Liguori, Chem. Rev. 109 (2009) 515–529.
bi 1.831 × 104 51.56 [12] G. Gelbard, Ind. Eng. Chem. Res. 44 (2005) 8468–8498.
[13] A. Molnar, Curr. Org. Chem. 15 (2011) 3928–3960.
(bi ) 1.35 × 104 2.36
[14] X. Chen, Z. Xu, T. Okuhara, Appl. Catal. A: Gen. 180 (1999) 261–269.
(k0,c , EA,c ) −0.9997
[15] M.R. Altiokka, E. Ödeş, Appl. Catal. A: Gen. 362 (2009) 115–120.
136 T. Komoń et al. / Applied Catalysis A: General 451 (2013) 127–136

[16] N. Essayem, V. Martin, A. Riondel, J.C. Verdine, Appl. Catal. A: Gen. 326 (2007) [31] T. Pöpken, L. Götze, J. Gmehling, Ind. Eng. Chem. Res. 39 (2000) 2601–2611.
74–82. [32] Y.-T. Tsai, H. Lin, M.-J. Lee, Chem. Eng. J. 171 (2011) 1367–1372.
[17] S. Schwarzer, U. Hoffmann, Chem. Eng. Technol. 25 (2002) 975–980. [33] R. Tesser, L. Casale, D. Verde, M. Di Serio, E. Santacesaria, Chem. Eng. J. 157
[18] K.-L. Zeng, C.-L. Kuo, I.-L. Chien, Chem. Eng. Sci. 61 (2006) 4417–4431. (2010) 539–550.
[19] I-L. Chien, K. Chen, Ch-L. Kuo, J. Proc. Cont. 18 (2008) 215–231. [34] C.S.M. Pereira, S.P. Pinho, V.M.T.M. Silva, A.E. Rodrigues, Ind. Eng. Chem. Res. 47
[20] V.A. Fomin, I.V. Etlis, V.I. Kulemin, J. Appl. Chem. USSR 64 (1991) 1811–1815. (2008) 1453–1463.
[21] J.-M. Paul, P. Busca, EP 1 219 587 (2001). [35] M.E. Jamróz, T. Komoń, P. Niewiadomski, unpublished results.
[22] D. Ramprasad, J.R. Collin, US 2005/0027135 (2005). [36] J. Guilera, R. Bringué, E. Ramírez, M. Iborra, J. Tejero, Appl. Catal. A: Gen. 413–414
[23] Rohm&Hass web site www.amberlyst.com (2012) 21–29.
[24] P.F. Siril, H.E. Cross, D.R. Brown, J. Mol. Catal. A: Chem. 279 (2008) 63–68. [37] R. Bringué, M. Iborra, J. Tejero, J.F. Izquierdo, F. Cunill, C. Fite, V.J. Cruz, J. Catal.
[25] C. Casas, R. Bringué, E. Ramirez, M. Iborra, J. Tejero, Appl. Catal. A: Gen. 396 244 (2006) 33–42.
(2011) 129–139. [38] S. Ostrowski, M.E. Jamróz, J.Cz. Dobrowolski, Comput. Theor. Chem. 974 (2011)
[26] P. Nowak, React. Kinet. Catal. Lett. 66/2 (1999) 375–380. 100–108.
[27] M. Grzesik, J. Skrzypek, M. Witczak, Przem. Chem 79/8 (2000) 271–273. [39] R. Wittig, J. Lohmann, J. Gmehling, Ind. Eng. Chem. Res. 42 (2003) 183–188.
[28] M. Brehelin, F. Forner, D. Rouzineau, J.-U. Repke, X. Meyer, M. Meyer, G. Wozny, [40] T.E. Daubert, R.P. Danner, Physical, Thermodynamic Properties of Pure Chemi-
Chem. Eng. Res. Des. 85 (2007) 109–117. cals, Taylor&Francis, Washington, 1998.
[29] B. Schmidt, M. Döker, J. Gmehling, Ind. Eng. Chem. Res. 47 (2008) 698–703. [41] A. Chakrabarti, M.M. Sharma, React. Polym. 20 (1993) 1–45.
[30] B. Erdem, M. Cebe, Korean J. Chem. Eng. 23 (2006) 896–901.

View publication stats

S-ar putea să vă placă și