Sunteți pe pagina 1din 9

Computational and Theoretical Chemistry 974 (2011) 100108

Contents lists available at ScienceDirect

Computational and Theoretical Chemistry


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

Formation of heavy adducts in esterication of acrylic acid: A DFT study


Sawomir Ostrowski a, Magorzata E. Jamrz a,, Jan Cz. Dobrowolski a,b,
a b

Industrial Chemistry Research Institute, 8 Rydygiera Street, 01-793 Warsaw, Poland National Medicines Institute, 30/34 Chemska Street, 00-725 Warsaw, Poland

a r t i c l e

i n f o

a b s t r a c t
Esterication of acrylic acid with alcohols and the side reactions were studied at the B3LYP/6-31G level. The main model reaction is predicted to be endoergic which is in line with experimental ndings. It appeared that the activation barriers for the esterication reactions in vacuum are equal to ca. 68 3 kcal/mol, they decrease in presence of the H+ ion to ca. 47.5 3.5 kcal/mol, and are relatively insensitive to change of environment simulated by PCM method (a matter of 0.5 kcal/mol). Out of four side reactions studied, the lowest activation barrier (36.5 1.5 kcal/mol) is for addition of the acrylic acid molecule to double bond in acrylates. Next, relatively easily occurring side reactions are the additions of water and alcohols to acrylates (barriers of ca. 48 1 kcal/mol in presence of the H+ ion). Activation barrier for dimerisation of acrylic acid, i.e., addition of one molecule to the double bond of the other, in catalytic reaction is equal to 56.5 kcal/mol. Finally, the addition of alcohol to the acid dimer (leading to the same product as addition of acid to an acrylate) needs to overcome the 61 kcal/mol barrier. Based on the above results we discuss qualitatively our experimental ndings of technology using heterogeneous acid catalysts. 2011 Elsevier B.V. All rights reserved.

Article history: Received 6 April 2011 Received in revised form 15 July 2011 Accepted 15 July 2011 Available online 23 July 2011 Keywords: Acrylic acid Acrylates Activation energy DFT Esterication Side reactions

1. Introduction Acrylates are intermediate products of high commercial value used chiey in the manufacturing of polymers and copolymers. They are particularly important in formulation of water-based acrylic emulsions for production of coatings, polishes, carpet backing compounds, adhesives, and sealants. For acrylic based polymers, they are primarily applied in construction, paper and textiles industry as well as for curing polymeric materials, used for instance in dentistry [15]. In 20022007 the world demand for acrylic acid increased ca. 3.4% per year. Although, a stagnation at the acrylic acid market was observed for the next few years, now a 1.6% increase is anticipated. It is important that more than half of the world production of acrylic acid is used to synthesize various acrylic acid esters, chiey n-butyl, and 2-ethylhexyl acrylates. Acrylates are obtained by direct acid catalyst esterication of acrylic acid with alcohols at elevated temperature [511]. This is an equilibrium process. A disadvantage of such an esterication is that the subsequent reactions occur in these very conditions. The unconverted starting alcohol or acrylic acid reacts with double bond of the ester formed. As a result, the alkoxyesters and acryloiloxyesters (the oxyesters) are formed. Moreover, acrylic acid is also involved in
Corresponding authors. Address: Industrial Chemistry Research Institute, 8 Rydygiera Street, 01-793 Warsaw, Poland. Tel.: +48 22 568 2021 (M.E. Jamrz), 48 22 568 2421 (J.Cz. Dobrowolski). E-mail addresses: Malgorzata.Jamroz@ichp.pl (M.E. Jamrz), janek@il.waw.pl (J.Cz. Dobrowolski).
2210-271X/$ - see front matter 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.comptc.2011.07.016

dimerization to 3-acryloiloxypropionic acid. Furthermore, water formed in esterication is added to acrylic moiety and hydroxyesters or hydroxyacid are formed, too. The generation of these by-products can be seriously disadvantageous for the production costs. Esterication reaction is apparently well recognized as is explained in each fundamental organic chemistry textbook (e.g. [12]). However, although ab initio studies on esterication were undertaken just in late 1970s and early 1980s [13,14], now, this relatively simple reaction is very seldom taken into consideration by using modern computational techniques. Indeed, a very recent DFT study on mechanism of direct esterication of the p-nitrobenzoic acid with n-butanol has shown the diethyl chlorophosphate to decrease by as much as ca. 20 kcal/mol [15]. Moreover, the reaction can proceed via different channels: one of them, through a fourmembered ring, is being a one-step reaction, whereas the other proceeds through two steps. In another very recent study the main focus of the computational part of investigation of esterication kinetics was to tell more about adsorption of acetic acid on a catalyst [16]. The energetics and polyesterication mechanisms of succinic acid with ethylene glycol were recently investigated at the B3LYP/6-31G level followed by IRC analysis [17]. It was found that the polyesterication ran in self-catalyzed concerted and nonself-catalyzed stepwise mechanisms. The self catalysts include acid, alcohol and water. In conclusion, it was found that the optimum channel of the esterication is a stepwise mechanism and that the self acid catalyst and dimerization are the rate-limiting steps. Also, for large set of esters the enthalpies of formation were elucidated by theoretical methods [18,19].

S. Ostrowski et al. / Computational and Theoretical Chemistry 974 (2011) 100108

101

The experimental determination of the equilibrium constant, kinetic and thermodynamic parameters for the ethyl acrylate formation were accomplished in Refs. [2024]. It appeared that the reaction is endothermic and equals ca. 6.5 kcal/mol [20,21]. The estimated activation barriers practically did not vary with the catalyst used and was equal to ca. 15.5 kcal/mol [2022]. On the other hand, for butyl acrylate, the reaction enthalpy was found to be equal from ca. 7.5 kcal/mol [25] to ca. 16.5 kcal/mol [26], and the barrier varied from 14 to 20 kcal/mol depending on the catalyst used [2629]. The physico-chemical aspects of the acrylic acid esterication with 2-ethyl-1-hexyl alcohol were studied only rarely [30,31]. The estimated enthalpy of the reaction was determined to be 16.518.0 kcal/mol [30] and the reaction barrier to ca. 1718.5 kcal/mol [30] and 13.5 kcal/mol [31]. Recently, we have been developing technology of acrylic acid and acrylates production based on waste glycerol derived from FAME manufacturing. Many of acrylate producers improve process of esterication by recovering and recycling reactants from their higher boiling adducts formed during the processing [7]. Nevertheless, there is still little evidence to suggest which heavy adducts formation is favored in acrylate manufacturing. Therefore, it is important to recognize the physicochemical properties and energy barriers of the secondary reactions leading to the main by-products (Scheme 1). This was the main motivation for the computational study. 2. Calculations All DFT calculations were performed by using the B3LYP functional [32,33], for which the reliability in calculations of the ground state geometries has been widely assessed [34]. The standard 631G basis set was applied throughout the study. Stationary structures have been recognized as true minima after checking for the lack of imaginary harmonic frequencies. The relative abundances of the most stable conformations were estimated using the Gibbs free energies at 298.15 K, DG, and referred to the most stable

conformer. The inuence of the solvent was also studied and the solutesolvent interactions were calculated by using the Tomasis polarized continuum model with the integral equation formalism (IEFPCM) [35,36]. In this procedure, the solvent is mimicked by a dielectric continuum with dielectric constant e surrounding a cavity with shape and dimension adjusted to the real geometric structure of the solute molecule. The latter polarizes the solvent which, as a response, induces an electric eld (the reaction eld) which interacts with the solute. In the IEF PCM, the electrostatic part of such an interaction is represented in terms of an apparent charge density spread on the cavity surface. Three solvents were considered: chloroform, ethanol, and water. Because of the way, the PCM methods simulate the solvent surroundings, the solvents were chosen to reveal the effect of increasing permittivity of the solvent rather than the solvent possibility to be the H+ ions carrier. The dielectric constants for the solvent used were 4.71, 24.85, and 78.36 for chloroform, ethanol, and water, respectively. Full geometry optimizations were performed both in the gas phase and in solvents. All stationary points (minima and TSs) were conrmed by calculating the harmonic vibrational frequencies, using analytical second derivatives. All calculations were performed by using Gaussian 09 packages of programs [37]. 3. Results and discussion Our experimental study on esterication is focused on developing acrylates technology, in particular on improvement of manufacturing of ethyl, butyl, and 2-ethylhexyl acrylates [38]. A number of technologies currently used has still a severe shortcoming they are performed with a homogeneous acidic catalyst highly corrosive and generating heavy waste, difcult to separate from the product. A solution to these problems can be the use of an efcient heterogeneous catalysts. Therefore, we apply solid acid catalysts in the form of sulfonated polystyrene/divinylbenzene resin, which are well recognized to be cheap, active and durable

OH
2C

H O

OH

2C

H O

O
2C

H O O

OH

3-acryloiloxy propionic acid (dimer)

O
2C

H O O O

OH

+
2C

OH

H H
+

O
2C

H O O

O R

OH R

2C

H O

H O

3-acryloiloxy propionic ester

O
2C

H O

OH

R O O

O R

3-Alkoxy propionic ester O


2C

H O

OH

H O O

O R

3-Hydroxy propionic ester


Scheme 1. Main secondary reactions related to esterication process of acrylic acid.

102

S. Ostrowski et al. / Computational and Theoretical Chemistry 974 (2011) 100108

Fig. 1. The B3LYP/6-31(d,p) most stable conformers of acrylic acid (AA), butyl alcohol (BuOH), and butyl ester (BA) and their the most stable protonated conformers.

industrial catalysts. In particular, we tested activity and selectivity of the esterication over Amberlyst 39, 46, 70, and 131. The smallest amounts of byproducts and quite a high yield were obtained in presence of Amberlyst 70. The following reactions were performed experimentally: acrylic acid with (a) ethyl alcohol, (b) n-butyl alcohol, and (c) 2-ethyl-1hexyl alcohol. The reaction conditions depended on the alcohol used for the esterication. For the ethyl acrylate formation, the temperature was 355 K and atmospheric pressure. The butyl and 2-ethyl-1-hexyl acrylates were obtained at the temperature range of 393403 K under 45 bars. For all the acrylates, the main byproducts are formed in the subsequent reactions of the appropriate ester with alcohol, acrylic acid, and water (being a product of ester formation). The exemplary content of products is presented in Table 1SI of supplementary materials. The limitation of byproducts formation is one of the most challenging tasks of our research. To better understand the investigated processes, we performed the following computational studies. First, the main esterication reactions of the acid and alcohols in vaccuo were considered. Second, the reactions were calculated in presence of H+ ion to simulate the presence of acidic catalyst. Finally, the inuence of solvent surrounding was taken into account by applying the IEFPCM model of the water, methanol, and chloroform solvents. Then, the acid dimerization reaction, that is the acidic OH addition to the double bond of another acid molecule, was studied with and without presence of a proton. In the subsequent steps the addition of water, alcohol and acrylic acid to double bond of esters previously formed was studied in presence and absence of catalytic proton. It is important, that in the experimental kinetic studies on inuence of catalyst concentration on reaction productivity it appeared that the process can be described as quasi-homogeneous. This justies the theoretical model(s) applied in this study in which the itemized details of the catalytic surface have not been considered. Also, it is known that the presence of a proton decreases the barrier height in the gas phase and therefore the inuence of H+ ion on the reaction course can be accepted as a rst step in modeling an acid catalyzed reaction [3841]. 3.1. Esterication The modeling of the acrylic acid esterication reactions by methyl, ethyl, and butyl alcohol and the selected side reactions in absence and in presence of acidic catalyst simulated by H+ ion was performed by using B3LYP/6-31(d,p) calculations. In the introductory step, the conformational spaces of each molecule participating in the reaction (acrylic acid, methyl, ethyl, and butyl

Table 1 The B3LYP/6-31G(d,p) calculated Gibbs free energies and relative Gibbs free energies of reactants, transition states and products of acrylic acid esterication with methyl, ethyl, and butyl alcohols calculated with and without presence of H+ catalyst (for abbreviations see Fig. 2). System Without H catalyst AA1 + MeOH TSMA(H2O) MA(H2O) AA1 + EtOH TSEA(H2O) EA(H2O) AA1 + BuOH TSBA(H2O) BA(H2O) With H+ catalyst AA1(H+) + MeOH TSMA(H2O) (H+) MA(H2O)(H+) AA1(H+) + EtOH TSEA(H2O)(H+) EA(H2O)(H+) AA1(H+) + BuOH TSBA(H2O)(H+) BA(H2O)(H+)
+

DG (kcal/mol)
0.0 71.2 5.4 0.0 64.9 5.8 0.0 64.6 5.9

0.0 57.8 9.2 0.0 46.8 9.0 0.0 44.1 9.1

alcohols, as well as the appropriate esters) were determined (Fig. 1). The same was done for each of the protonated forms (Fig. 1). First, the reactions without catalyst were considered. The acrylic acid esterication reactions studied at the B3LYP/631G(d,p) were endoergic (Table 1). This is in agreement with the experimental knowledge on the esterication processes. The activation barriers for these reactions (Fig. 2) were found to be the following: 65, 65, and 71 kcal/mol for synthesis of butyl, ethyl and methyl, acrylates, respectively (Table 1). As expected, the activation barriers decrease signicantly when H+ ion is accompanying the reaction. Indeed, the barriers in the catalytic process are the following: 44, 47, and 58 kcal/mol for synthesis of butyl, ethyl and methyl, respectively (Table 1). We found that the catalytic reactions were even more endoergic than the non-catalytic processes (Table 1). Finally, we checked whether the solvent surrounding has an inuence on the barrier height or not. It appeared, that this inuence (simulated by using the IEFPCM solvent model) is practically meaningless (a matter of 0.5 kcal/mol, Table 2). This means, that the center of reaction remains practically impenetrable by the solvent molecules in the reaction course and

S. Ostrowski et al. / Computational and Theoretical Chemistry 974 (2011) 100108

103

Fig. 2. The B3LYP/6-31G(d,p) reactants (reactant complexes), transition states, and products of esterication of acrylic acid with methyl, ethyl, and butyl alcohols in absence and presence of an acidic catalyst (H+) (AA acrylic acid; MeOH methyl alcohol; EtOH ethyl alcohol; BuOH butyl alcohol; MA methyl acrylate; EA ethyl acrylate; BA butyl acrylate; TS transition state).

that the charge distribution around this center for reactants, transition states, and products remains quite stable within the frame of the IEFPCM model applied. Last but not least, let us comment on the applied methodology. The primary source of protons are the SO3H groups of a catalyst. Although in the very early step of reaction water is not yet present in the system in an important amount, the H3O+ cation is a more probable H+ carrier than the H+ ion itself. Therefore, we tried to model the reactions with presence of the hydronium cation. The proton jumps from the SO3H groups to the closest basicity center. It is likely, that in presence of alcohol it will be the alcohol hydroxyl group. When the water molecules concentration is increased during the reaction course, the H3O+ ion formation start to be populated as well. However, the calculated systems are very exible and we could not nd the TSs when the hydronium was used

instead of H+ cation. This probably could be found assuming some additional constrains, yet, the error which could result from such an assumption, would be difcult to estimate. In this context, use of chloroform looks bizarre because it can hardly be a proton carrier. This very solvent was used only for the purpose to estimate a role of changeable permittivity of the reaction environment. 3.2. Side reactions The acrylic acid esterication is accompanied by a series of side reactions. This is a result of the presence of double bond in both the reactant and the acrylates. The acidic media activate the OH groups in the acid and alcohols, but, simultaneously they activate the double bonds in the reactant and in the products. This is the main difculty in industrial operation with esterication. In our

104

S. Ostrowski et al. / Computational and Theoretical Chemistry 974 (2011) 100108 Table 3 The B3LYP/6-31G(d,p) calculated relative Gibbs free energies (kcal/mol) of reactants, transition states and products of acrylic acid dimerization reaction calculated with and without presence of H+ catalyst (for abbreviations see Fig. 3). System Without H+ catalyst AA1 + AA1 TSAoxPA AoxPA With H+ catalyst AA1(H+) + AA1 TSAoxPA(H+) AoxPA(H+)

Table 2 The B3LYP/6-31G(d,p) calculated relative Gibbs free energies (kcal/mol) of reactants, transition states and products of acrylic acid esterication with the methyl, ethyl, and butyl alcohols in vacuum and in solvents simulated by the IEFPCM method (for abbreviations see Fig. 2). System AA1(H+) + MeOH TSMA(H2O)(H+) MA(H2O)(H+) AA1(H+) + EtOH TSEA(H2O)(H+) EA(H2O)(H+) AA1(H+) + BuOH TSBA(H2O)(H+) BA(H2O)(H+) Gas phase DG298 0.0 57.8 9.2 0.0 46.8 9.0 0.0 44.1 9.1 Water DG298 0.0 57.6 9.5 0.0 46.7 9.3 0.0 44.7 9.4 Ethanol DG298 0.0 57.6 9.5 0.0 46.7 9.3 0.0 44.6 9.3 Chloroform DG298 0.0 57.5 9.4 0.0 46.7 9.2 0.0 44.6 9.3

DG298
0.0 68.5 4.1 0.0 56.3 5.7

experimental studies, we detected addition products of water, alcohols and acrylic acid to double bonds in acrylates. These reactions lead to quite a number of byproducts. First, we examined the socalled dimerization of acrylic acid, i.e., the carboxylic OH group addition to double bond of the other acrylic acid molecule leading to formation of 3-acryloiloxypropionic acid (Fig. 3, Table 3). Observe, that the non-catalytic process is endoergic whereas the catalytic one is exoergic (Table 3). The barrier of the non-catalytic reaction is 68.5 kcal/mol and is found to decrease by ca. 12 kcal/mol in presence of a catalyst (Table 3). Next, we studied addition of water and alcohols to double bonds of acrylates (Fig. 4). These reactions are exoergic: a matter of 35 kcal/mol (Tables 4 and 5). The barriers in the reactions with the water molecule and methanol do not depend on the ester type and are equal to ca. 48 (1) kcal/mol (Tables 4 and 5). In the case of the addition of the acrylic acid molecule to acrylates the reaction is slightly exoergic (ca. 1.5 kcal/mol) and the barriers are equal to ca. 36.5 (1.5) kcal/mol (Table 6). 3.3. Energetics Now, let us summarize the energetics of the studied main and side reactions. First, the acrylic acid esterication is endoergic in line with the experimental ndings. The presence of the H+ ions cut down the barrier heights. However, the endoergicity of the catalytic processes is larger than that of the non-catalytic reactions. Thus, use of a catalyst facilitates the reaction, but more energy must be supplied to run the process. According to our calculations, the additions of the acrylic acid molecule to acrylates are the easiest processes: they exhibit the smallest activation barriers. Surprisingly, apparently similar

addition reaction of two acid molecules exhibits the barrier higher than the above reaction by ca. 20 kcal/mol (Tables 3 and 6). There is however, a small but signicant difference: two protons are involved in the addition of acid to an acrylate catalyzed by H+ ion, whereas three protons are involved into the analogous acid dimerization. In the former case, one of the protons can be localized at the double bond of acrylate (Fig. 4), whereas in the latter case, the presence of three protons forces a specic conformation of the system blocking the double bond accessibility (Fig. 5). Note, that the addition of alcohol to the acid dimer (leading to the same product as addition of acid to an acrylate) needs the 61 kcal/mol barrier to be overcome (Table 7). This means that this is the most difcult reaction in the whole reaction system. Finally, the formation of alkoxy and hydroxy propionic acid esters requires ca. 48 kcal/mol barrier to be overcome regardless the ester and alcohol types (Tables 4 and 5). Thus, the most important side reactions are all additions to double bond moiety of acrylic acid esters. Last but not least, the proton afnities (PA) are listed in Table 8 and Table 2SI of supporting information. The PA values are calculated for the most stable forms of the neutral and protonated compounds. There is a regular tendency to increase the proton afnity from methyl, through ethyl, to butyl derivative regardless the type of the studied compounds. The proton afnity tends to increase with the number of oxygen atoms in the molecular structure. For the studied reactions, the number of O-atom in the products is greater than in the reactants therefore, the protonation shifts the reaction towards the products. 3.4. Computational vs. experimental ndings Presentation of the computational results always bears a question about quantitative agreement of the energetical values found

Fig. 3. The B3LYP/6-31(d,p) optimized reactants, transition states, and products of acrylic acid (AA1) dimerization in absence and presence of H+ ion (AA acrylic acid; AoxPA acryloiloxypropionic acid; TS transition state).

S. Ostrowski et al. / Computational and Theoretical Chemistry 974 (2011) 100108

105

Fig. 4. The B3LYP/6-31(d,p) reactants (reactant complexes), transition states, and products of water addition (a) and MeOH addition (b), and acrylic acid addition (c) reactions to the double bonds of methyl, ethyl and butyl acrylates in presence of an acidic catalyst (H+) (AA acrylic acid; MA methyl acrylate; EA ethyl acrylate; BA butyl acrylate; TS transition state; MP methyl propanoate; EP ethyl propanoate; BP butyl propanoate; OHMP hydroxy methyl propanoate; OHEP hydroxy ethyl propanoate; OHBP hydroxy butyl propanoate; MMP metoxy methyl propanoate; MEP metoxy ethyl propanoate; MBP metoxy butyl propanoate; AoxMP acryloiloxy methyl propanoate; AoxEP acryloiloxy ethyl propanoate; AoxBP(H+) acryloiloxy butyl propanoate).

with the experimental ones. This, however, is a very complex problem. First, computations require simplications and use of methods guaranteeing nite time of calculations. In the case of many side reactions of non-rigid molecules, this is a crucial factor. Indeed, this has been the purpose of use of not very sophisticated B3LYP/6-31G level in this study. Moreover, it is known, that the B3LYP functional does not yield a correct barrier height [39,40], which is a consequence of parametrizing it to correctly reproduce thermochemical not kinetic data [42,43]. Last but not least, the largest and unpredictable errors arise from not fully adequate mechanism of reaction assumed in calculations. First of all, this includes differences in surrounding of reacting system. The PCM-like models of the solvent may be of some help, yet, much better solution to this problem is to combine PCM with a supermolecule approach for the rst solvation sphere. For non-rigid systems this is again a very difcult requirement to be satised. As a result, the

computationally studied reaction is considered in an abstract vacuum whereas the real reaction may be, for instance, sensitive to traces of water facilitating proton jump from one to the other electron donor center. Similar can be said about a catalyst with dened structure at the atomic scale, where some positive and negative centers facilitate the reaction. For the reaction studied in this paper there are limited thermodynamic and kinetic data [2031]. They say, that by using various acid catalysts and alcohols, the barrier for esterication falls into the range from ca. 15 to 20 kcal/mol, whereas for the reaction model assumed in this paper the barriers are ca. three times larger. The experimental thermochemical data for the side reactions are unavailable. The discrepancy between the calculated and the experimental data is difcult to be unequivocally interpreted. It may indicate that the assumed mechanism is oversimplied. It may indicate an error in the experimental data larger than given

106

S. Ostrowski et al. / Computational and Theoretical Chemistry 974 (2011) 100108

Fig. 4 (continued)

Table 4 The B3LYP/6-31G(d,p) calculated relative Gibbs free energies (kcal/mol) of reactants, transition states and products of the water molecule addition to the double bond of the methyl, ethyl, and butyl acrylates in presence of the H+ ion (for abbreviations see Fig. 4). System MA(H ) + H2O TSOHMP(H+) OHMP(H+) EA(H+) + H2O TSOHEP(H+) OHEP(H+) BA(H+) + H2O TSOHBP(H+) OHBP(H+)
+

Table 5 The B3LYP/6-31G(d,p) calculated relative Gibbs free energies (kcal/mol) of reactants, transition states and products of the methanol molecule addition to the double bond of the methyl, ethyl, and butyl acrylates in presence of the H+ ion (for abbreviations see Fig. 4). System BA(H ) + MeOH TSMBP(H+) MBP(H+) MA(H+) + MeOH TSMMP(H+) MMP(H+) EA(H+) + MeOH TSMEP(H+) MEP(H+)
+

DG298
0.0 47.5 3.1 0.0 48.2 3.6 0.0 48.5 3.7

DG298
0.0 47.8 5.4 0.0 47.1 4.7 0.0 47.6 5.2

in the papers. It may indicate incompatibility(ties) of the reaction conditions between the two studies. However, at the present stage of study, one may believe, that a constant systematic error is associated with all the calculated data and that qualitative conclusions can be drawn and can be use to interpret the experimental ndings. Therefore, another important question arises: what is the significance of our computational ndings for interpretation of experimental systems. First, note that for practical reasons, a heterogenous catalysis is applied in industrial processes. However, a heterogenous catalysis is hardly reproducible by using a sole proton to simulate the catalytic act. So, the calculations are likely to simulate homogenous rather than heterogenous system. The surrounding is argued to have a minor inuence on the reaction courses, the assumption which is rather strong, yet, it is quite difcult to go beyond this condition. Let us now compare qualitatively the results of acrylic acid esterication in presence of heterogeneous acid catalysts with the results of our modeling. First, in the experiments the main byproducts are alkoxy and hydroxy propionic acid esters. Despite the fact that the addition of acid to esters is the easiest, in the experimental

Table 6 The B3LYP/6-31G(d,p) calculated relative Gibbs free energies (kcal/mol) of reactants, transition states and products of the acrylic acid molecule addition to the double bond of the methyl, ethyl, and butyl acrylates in presence of the H+ ion (for abbreviations see Fig. 4). System MA(H+) + AA1 TSAoxMP(H+) AoxMP(H+) EA(H+) + AA1 TSAoxEP(H+) AoxEP(H+) BA(H+) + AA1 TSAoxBP(H+) AoxBP(H+)

DG298
0.0 36.0 1.4 0.0 37.0 1.7 0.0 37.5 1.8

conditions, concentration of acid is much lower than that of the alcohol. So, in the experimental systems the alkoxy propionates are much more abundant than the acryloiloxypropionates. Second, the presence of hydroxy propionic acid esters may seem strange

S. Ostrowski et al. / Computational and Theoretical Chemistry 974 (2011) 100108

107

Fig. 5. The B3LYP/6-31(d,p) reactants (reactant complexes), transition states, and products of methanol esterication of acrylic acid dimer in absence and in presence of an acidic catalyst (H+) (MeOH methanol; AoxPA acryloiloxypropionic acid; TS AoxMP transition state for acryloiloxy methyl propanoate; AoxMP acryloiloxy methyl propanoate).

Table 7 The B3LYP/6-31G(d,p) calculated Gibbs free energies (kcal/mol) of reactants, transition states and products of the methyl alcohol esterication of the dimer of acrylic acid in absence and in presence of the H+ ion (for abbreviations see Fig. 5). System Without H+ catalyst AoxPA + MeOH TSAoxMP AoxMP + H2O With H+ catalyst AoxPA(H+) + MeOH TSAoxMP(H+) AoxMP(H+) + H2O

DG298
0.0 81.3 15.4 0.0 60.9 6.3

neous catalysts. Thus it is always accompanying to proton activating the double bonds. Therefore, locally, in close vicinity of the catalytic centers concentration of water is high. Finally, the acrylic acid dimers are detectable, but, in minor concentrations, due to the fact that the acid is reacted with high excess of alcohols. Interestingly, the book-mechanism of acrylic acid dimerization is the Michael addition in presence of anions, while in the experimental system the cations predominate in the system. Thus, either a competitive mechanism may occur or small amounts of water induce acid dissociation and possibility for nucleophilic attack to double bond of the other acrylic acid molecule.

4. Conclusions The esterication of acrylic acid molecule with alcohols is industrially a very important reaction leading to acrylates which are monomers for obtaining water-based polyacrylic emulsions for production of coatings, polishes, carpet backing compounds, adhesives, sealants and polymers used in construction, paper and textiles industry, as well as for curing polymeric materials, used in dentistry. Unexpectedly, neither acrylate formation nor the side reactions were studied by computational chemistry methods, so far. In this study, we present results of the B3LYP/6-31G calculations on energetic and barriers of esterication of the acrylic acid molecule with methyl, ethyl, and butyl alcohols. Because of double bond moiety present in acrylate molecule, it is useful for further polymerization, however, it is also the target of subsequent alcohol attack to perform alkoxy propionic acid esters, as well as some other side products connected with undesired additions to acrylates double bond. Deeper inspection into formation of the side products in light of computational methods, was the aim of this study. We limited the investigations to by-products observed by us in the experimental studies in signicant extents, i.e., acrylic acid dimer, alkoxy propionic acid esters, hydroxy propionic acid ester, and the product of addition of acrylic acid to acrylate in presence and in absence of the H+ cation. As a result, we established, that the latter reaction exhibits the lowest activation barrier (36.5 1.5 kcal/mol). Next, additions of water and alcohols to acrylates needs ca. 48 1 kcal/mol to overcome the reaction barrier. The most difcult reaction to run is the alcohol addition to the acid dimer with ca. 60 kcal/mol activation barrier.

Table 8 The B3LYP/6-31G calculated proton afnities (kcal/mol) of the selected systems based on Gibbs free enthalpies calculated for 298 K and 1 atm. System Methanola Ethanola Butanola Acrylic acidb Acrylic acidc Methyl acrylateb Ethyl acrylateb Butyl acrylateb Acryloiloxy propionic acidb Hydroxy methyl propanoateb Hydroxy ethyl propanoateb Hydroxy butyl propanoateb Acryloiloxy methyl propanoateb Acryloiloxy ethyl propanoateb Acryloiloxy butyl propanoateb Metoxy methyl propanoateb Metoxy ethyl propanoateb Metoxy butyl propanoateb
a b c

Proton afnity (kcal/mol) 184.6 190.5 193.0 195.8 159.6 200.3 203.5 205.0 217.6 211.9 214.2 215.1 212.4 214.6 215.6 216.8 219.2 220.2

Protonation at the hydroxyl O-atom. Protonation at the O atom of the C@O group. Protonation at the C@C double bond.

taking into account the fact that concentration of water in the product is even smaller than that of acid (a matter of ca. 5%). However, it is quite easy to understand that fact. Water, even if it is in minor quantities, is attached to acidic groups (usually SO3H) of heteroge-

108

S. Ostrowski et al. / Computational and Theoretical Chemistry 974 (2011) 100108 [21] I.-L. Chien, K. Chen, C.-L. Kuo, Overall control strategy of coupled reactor/ columns process for the production of ethyl acrylate, J. Process Control 18 (2008) 215231. [22] M. Witczak, M. Grzesik, Kinetics of the esterication of acrylic acid with lower aliphatic alcohols in the presence of dodecatungstophosphoric acid as a catalyst, Chem. Process Eng. 27 (2006) 14551467. [23] D.J. Schreck, Esterication of Carboxylic Acids with Alcohols, UK Pat. 2 063 261, Union Carbide Co., 1980. [24] V.R. Dhanuka, V.C. Malshe, S.E. Chandalia, Kinetics of the liquid phase esterication of acrylic acids with alcohols in the presence of acid catalysts: re-interpretation of published data, Chem. Eng. Sci. 32 (1977) 551556. [25] K.-L. Zeng, C.-L. Kuo, I.-L. Chien, Design and control of butyl acrylate reactive distillation column system, Chem. Eng. Sci. 61 (2006) 44174431. [26] S. Schwarzer, U. Hoffmann, Experimental reaction equilibrium and kinetics of the liquid-phase butyl acrylate synthesis applied to reactive distillation simulations, Chem. Eng. Technol. 25 (2002) 975980. [27] J. Skrzypek, T. Witczak, M. Grzesik, M. Witczak, Kinetics of the synthesis of propyl and butyl acrylates in the presence of some heteropolyacids as catalysts, Int. J. Chem. Kinet. 41 (2009) 1217. [28] X. Chen, Z. Xu, T. Okuhara, Liquid phase esterication of acrylic acid with 1butanol catalyzed by solid acid catalysts, Appl. Catal. A: General 180 (1999) 261269. [29] O. Darge, F.C. Thyrion, Kinetics of the liquid phase esterication of acrylic acid with butanol catalysed by cation exchange resin, J. Chem. Technnol. Biotechnol. 58 (1993) 351355. [30] V.A. Fomin, I.V. Etlis, V.I. Kulemin, Some aspects of esterication of acrylic acid with 2-ethylhexyl alcohol on sulfonic cation-exchangers, J. Appl. Chem. USSR 64 (1991) 18111815. [31] P. Nowak, Kinetics of the liquid phase esterication of acrylic acid with Noctanol and 2-ethylhexanol catalyzed by sulfuric acid, React. Kinet. Lett. 66 (1999) 375380. [32] A.D. Becke, Density-functional thermochemistry. III. The role of exact exchange, J. Chem. Phys. 98 (1993) 56485652. [33] K. Burke, J.P. Perdew, Y. Wang, in: J.F. Dobson, G. Vignale, M.P. Das (Eds.), Electronic Density Functional Theory: Recent Progress and New Directions, Plenum, 1998, pp. 155172. [34] R. Janoschek, Quantum chemical B3LYP/cc-pvqz computation of ground-state structures and properties of small molecules with atoms of Z 6 18 (hydrogen to argon), Pure Appl. Chem. 73 (2001) 5211553. [35] M.T. Cances, B. Mennucci, J. Tomasi, J. Chem. Phys. 107 (1997) 3032. [36] J. Tomasi, B. Mennucci, R. Cammi, Chem. Rev. 105 (2005) 2999. [37] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E. Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, . Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian 09, Revision A.1, Gaussian, Inc., Wallingford CT, 2009 [38] R.C.D.M. Filho, S.A. Alves de Sousa, F. da Silva Pereira, M.M.C. Ferreira, Theoretical study of acid-catalyzed hydrolysis of epoxides, J. Phys. Chem. A. 114 (2010) 51875194. [39] P.U. Civcir, Computational investigations of the gas phase reactions between hydrogen chloride and protonated alkyl chlorides, J. Mol. Struct. THEOCHEM 848 (2008) 128138. [40] L.M. Pratt, N.V. Nguyen, B. Ramachandran, Computational strategies for evaluating barrier heights for gas-phase reactions of lithium enolates, J. Org. Chem. 70 (2005) 42794283. , M.E. Jamrz, J. Kijen ski, Esterication of Acrylic Acid with 1-Butanol [41] T. Komon in Column Reactor; Ion-eschange Resins as Catalysts, Abstract Book of 3rd EuCheMS Chemistry Congress, Nrnberg, Germany, 2010. [42] K.E. Riley, B.T. Opt Holt, K.M. Merz Jr., Critical assessment of the performance of density functional methods for several atomic and molecular properties, J. Chem. Theory Comput. 3 (2007) 407433. [43] J. Zheng, Y. Zhao, D.G. Truhlar, Representative benchmark suites for barrier heights of diverse reaction types and assessment of electronic structure methods for thermochemical kinetics, J. Chem. Theory Comput. 3 (2007) 569 582.

Although, the studied processes need much more attention and higher computational level to conrm quantitative analysis of the reaction energetics, qualitatively they help us in understanding the results of our experiments on heterogeneous catalytic esterication of acrylic acid. Acknowledgments This work was supported by Ministry of Science and Higher Education in Poland Grant No. POIG 01.03.01-00-010/08. The computational Grant G19-4 from the Interdisciplinary Center of Mathematical and Computer Modeling (ICM) at the University of Warsaw is gratefully acknowledged. Appendix A. Supplementary material Supplementary data associated with this article can be found, in the online version, at doi:10.1016/j.comptc.2011.07.016. References
[1] ICIS Chemical Business, 2008, August 417, p. 40 <http://www.icis.com/home/ Acrylicacid>. [2] Acrylic Acid and Derivatives, in: Ullmanns Encyclopedia, fourth ed., vol. A1, Wiley, 1987, p. 16. [3] Acrylic Acid and Derivatives, KirkOthmer Encyclopedia of Chemical Technology, fourth ed., vol. 1, Wiley, 1987, p. 287. [4] X. Chen, Z. Xu, T. Okuhara, Liquid phase esterication of acrylic acid with 1butanol catalyzed by solid acid catalysts, Appl. Catal. A: General 180 (1999) 261. [5] A. Van Eikeren, M.T. Plaumann, VOCO GmbH, Dental Masking Product for Teeth and Gum, US Pat. 7789662, September 7, 2010. [6] R. Shastry, Y.R. Mirajkar, N. Dixit, R. Cameron, Q. Wang, L. Zeidel, S.K. Chopra, M. Prencipe, Colgate-palmolive Co., Tooth Whitening Composition Containing Cross Linked PolymerPeroxides, PCT/US2005/046240, July, 13th, 2006. [7] W. Bauer Jr., J.T. Chapman, M.G.L. Mirabelli, J.J. Venter, Rohm and Haas Co., Process for Producing Butyl Acrylate, US Pat. 6180819, January 30th, 2001. [8] S. Nakahara, T. Nishimura, M. Ueoka, Nippon Shokubai Co., Ltd. Nippon Shokubai Co., Method for Production of (meth)acrylic Acid and (meth)acrylic esters, US Pat. 6695928, February 24th, 2004. [9] A. Clymo, A. Diefenbacher, T. Friese, BASF AG, Process for Preparing Alkyl Esters of (meth)acrylic acid, US Pat. Appl. 2006/0205972, September 13th, 2006. [10] A. Riondel, J. Bessalem, ATOFINA, Process for Preparing Butyl Acrylate by Direct Esterication, US Pat. 6846948, January 25, 2005. [11] S. Nakahara, M. Ueoka, Nippon Shokubai Co. Ltd, Method for Preparing (meth)acrylic Acid Ester, US Pat. 6649787, November 18th, 2003. [12] J. March, Advanced Organic Chemistry, fourth ed., Wiley, NY, 1992. [13] J. Emsley, O.P.A. Hoyte, R.E. Overill, Ab initio calculations on the very strong hydrogen bond of the biformate anion and comparative esterication studies, J. Am. Chem. Soc. 100 (1978) 33033306. [14] A.P. Mazurek, W. Szeja, Theoretical study on alkylation and esterication of methyl 3,6-anhydro-t-galactopyranoside, J. Chem. Soc. Perkin Trans II (1985) 5758. [15] W. Zhang, Y. Zhu, D. Wei, C. Zhang, D. Sun, M. Tang, Direct esterication of pnitrobenzoic acid with n-butanol using diethyl chlorophosphate in pyridine: a DFT study, Comp. Theor. Chem. 963 (2011) 1317. [16] S. Miao, B.H. Shanks, Mechanism of acetic acid esterication over sulfonic acidfunctionalized mesoporous silica, J. Catal. 279 (2011) 136143. [17] T. Yu, H.-B. Chang, W.-P. Lai, X.-F. Chen, Computational study of esterication between succinic acid and ethylene glycol in the absence of foreign catalyst and solvent, Polym. Chem. 2 (2011) 892896. [18] M.H. Liu, S.-R. Cheng, Determination modied enthalpy of formation of straight alkyl-chained carboxylic acids and esters, J. Mol. Struct. THEOCHEM 763 (2006) 149154. [19] E.A. Castro, Theoretical calculation of heats of formation of carboxylic acids and esters, J. Mol. Struct. THEOCHEM 339 (1995) 239242. [20] M. Witczak, M. Grzesik, J. Skrzypek, Kinetyka estrykacji kwasu akrylowego _ nizszymi alkoholami, Inz. Chem. Procesowa 25 (2004) 331340.

S-ar putea să vă placă și