Next Article in Journal
ZnI2-Mediated cis-Glycosylations of Various Constrained Glycosyl Donors: Recent Advances in cis-Selective Glycosylations
Previous Article in Journal
Red Grape By-Products from the Demarcated Douro Region: Chemical Analysis, Antioxidant Potential and Antimicrobial Activity against Food-Borne Pathogens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reaction Thermodynamic and Kinetics for Esterification of 1-Methoxy-2-Propanol and Acetic Acid over Ion-Exchange Resin

by
Xinyu Liu
,
Shu Wang
,
Mingxia Wang
,
Lifang Chen
* and
Zhiwen Qi
*
State Key Laboratory of Chemical Engineering, School of Chemical Engineering, East China University of Science and Technology, 130 Meilong Road, Shanghai 200237, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(19), 4709; https://doi.org/10.3390/molecules29194709
Submission received: 3 September 2024 / Revised: 30 September 2024 / Accepted: 30 September 2024 / Published: 4 October 2024
(This article belongs to the Special Issue Applied Chemistry in Asia)

Abstract

:
The esterification of 1-methoxy-2-propanol (PM) and acetic acid (AA) is an important reaction for the production of 1-methoxy-2-propyl acetate (PMA). Herein, we used the macroporous ion-exchange resin Amberlyst-35 as a catalyst to explore the effects of reaction conditions on the reaction rate and equilibrium yield of PMA. Under the optimized conditions of a reaction temperature of 353 K, using the initial reactant PM/AA with a molar ratio of 1:3, and a catalyst loading of 10 wt%, the PMA equilibrium yield reached 78%, which is the highest equilibrium yield so far. The reaction equilibrium constants and activity coefficients were estimated to obtain reaction thermodynamic properties, indicating the exothermicity of the reaction. Furthermore, pseudo-homogeneous (PH), Eley–Rideal (ER), and Langmuir–Hinshelwood–Hougen–Watson (LHHW) kinetic models were fitted based on experimental reaction kinetic data. The results demonstrate that the LHHW model is the most consistent with experimental data, indicating a surface reaction-controlled process and exhibiting an apparent activation energy of 62.0 ± 0.2 kJ/mol. This work represents a valuable example of calculating reaction thermodynamics and kinetics, which are particularly essential for promising industrial reactor designs.

1. Introduction

Methoxy propyl acetate, also known as 1-methoxy-2-propyl acetate (PMA), is widely used in paints, inks, dyes, cleaning agents, and photoresistors [1,2]. Moreover, its high solvency, thermal stability, and low toxicity have led to it gradually replacing ethylene glycol-based products [3,4,5] that have toxic effects on human metabolism [6,7,8,9]. Currently, ester exchange reactions and direct esterification are the main methods considered for the synthesis of PMA. The catalyst for ester exchange between 1-methoxy-2-propanol (PM) and methyl ethanoate [1] (or ethyl ethanoate [10]) is sodium methoxide [11], which is difficult to separate and highly susceptible to decompose in water, resulting in a reduction in catalytic activity. The direct esterification of PM with acetic acid (AA) is industrially catalyzed by sulfuric acid, which has gradually been ruled out owing to its environmental unfriendliness and corrosiveness to equipment. Thus, some environmentally-friendly catalysts, such as 4-methylbenzenesulfonic acid, titanium sulfate, and cation exchange resins, have been developed for PMA synthesis [3,12,13,14]. In addition, direct esterification of PMA synthesis is reversible and limited by a thermodynamic equilibrium [12,14,15,16,17,18,19,20,21]. Oh et al. indicated that, in a batch reactor, the maximum conversion of PM is only 46% due to the limitation of the chemical equilibrium [13].
In comparison with homogeneous catalysts, solid acid catalysts offer a number of advantages, including low corrosiveness and facile separation and, therefore, are excellent alternatives to traditional liquid acid catalysts [22,23,24]. Currently, solid acids are widely used in a variety of chemical reactions, including esterification [25,26,27], alkane isomerization [28,29], aldol condensation [30,31,32], hydrogenation [33], and formylation [34,35]. For direct esterification to synthesize PMA, Huang et al. examined the solid acid SO42−/TiO2 with a pseudo-homogeneous kinetic model [3], obtaining an apparent activation energy (Ea+) of 65.7 kJ/mol. Gadekar-Shinde et al. utilized the ion-exchange resin Amberlyst-15 as a catalyst [12] and developed a concentration-based pseudo-homogeneous kinetic model with an Ea+ value of 66.5 kJ/mol. Wang et al. employed the ion-exchange resin NKC-9 as a catalyst to obtain a maximum PM conversion of 46% in a batch reaction [14], and the Ea+ was determined to be 60.5 kJ/mol based on a second-order reaction kinetic model. However, these studies tend to focus on pseudo-homogeneous (PH) kinetic models, which are unable to reveal the real reaction mechanisms related to the surface adsorption and reaction.
The Langmuir–Hinshelwood–Hougen–Watson (LHHW) and Eley–Rideal (ER) models have been extensively employed in kinetic studies of heterogeneous catalytic systems. The LHHW model describes the surface reaction between adsorbed reaction species and has been successfully employed to predict numerous reactions catalyzed by heterogeneous acid catalysts. As evidenced by previous reports, reaction kinetic data, such as esterification of benzoic acid with ethanol [36], oligomerization of 1-decene [37], liquid-phase hydrogenation of 1-indanone [38], esterification of lactic acid with ethanol [39], synthesis of tert-butyl methyl ether [40], and catalytic hydrogenation of d-lactose to lactitol [41], are well-fitted to the LHHW model. Conversely, the ER model postulates the potential for a reaction to occur between an adsorbed reactant molecule and another reactant molecule in the bulk phase. Catalytic reactions, such as cyclohexene hydration [42], the esterification of acetic acid with butanol [43], the dehydration of 1-octanol to di-n-octyl ether [44], and the oximation of cyclohexanone [45], are illustrated to follow the ER mechanism. It is noted that the model for the reaction catalyzed by a heterogeneous catalyst is rather ambiguous from case to case and should be dependent on the catalyst and reactants involved [42].
At present, the reaction kinetics for the esterification of PM with AA using a developed heterogeneous kinetic model is lacking. Herein, ion-exchange resin Amberlyst-35 was used as the catalyst, and kinetic data were obtained through the esterification of pure AA with PM in a stirred batch reactor. The effects of operating parameters, including temperature, reactant ratio, and catalyst loading, were investigated. Subsequently, this work evaluated reaction thermodynamic properties and identified several kinetic models to describe a wide range of operating conditions.

2. Results and Discussion

2.1. Effect of Mass Transfer

As indicated, the external and internal mass transfer limitations of resin catalysts in catalytic esterification reactions can be ruled out through varying stirring speeds and particle sizes of catalysts [3,13,14]. In order to avoid mass transfer influencing the reaction rate, the effect of stirring speed on PMA yield was carried out at a relatively high reaction temperature of 363 K [44]. As presented in Figure 1, it can be observed that varying stirring speeds within a range of 200 to 600 rpm has no significant impact on the yield of PMA, indicating that the influence of external diffusion can be disregarded. Nevertheless, elevated rotational speeds may result in increased wear between the rotor and catalyst particles, which is detrimental to the recovery and reuse of the catalyst. Consequently, a stirring speed of 300 rpm can be established as the optimal setting for subsequent reaction processes.
The particle diameter of the catalyst has a direct impact on internal diffusion. Previous studies have employed the method of varying catalyst particle diameters to investigate the effects of internal diffusion [13,14,46,47,48]. The particle size of the cation exchange resin Amberlyst-35 is not homogeneous, making it difficult to study the effects of internal diffusion by particle size screening. In theory, the Weisz–Prater number (Cwp) can be employed to assess the influence of internal diffusion on mass transfer in a reaction system catalyzed by a cation exchange resin [49]. When Cwp is much less than 1, the effect of internal diffusion on the mass transfer resistance of the reaction is negligible. Cwp is calculated by Equation (1),
C wp = r obs ρ c R c 2 C s D e
where −robs represents the apparent reaction rate using an initial reaction rate to minimize the influence of external factors; Cs is the concentration of reactants on the catalyst surface, assumed to be equal to that of the liquid phase due to negligible external diffusion effects; De represents the effective diffusion coefficient; ρc is the density of catalyst; and Rc is the radius of catalyst particles (according to the report of Pera-Titus et al. [50], ρc is 1.542 g/cm3 and Rc is 0.016 cm). De is calculated by Equation (2),
D e = ε c D A τ
where εc and τ represent the porosity and tortuosity factors of catalyst particles (where εc is assumed to be 0.35 and τ is 1/εc), respectively, and DA is the infinite dilution diffusion coefficient. DA can be calculated using the Wilke–Chang empirical correlation by Equation (3),
D A = 7.4 × 10 8 ( Φ 2 M 2 ) 0.5 T μ 2 V 1 0.6
where Φ2 is the association factor of PM, M2 is the molar mass of PM, μ2 is the viscosity of PM, and V1 is the molar volume of AA at a normal boiling point.
Table 1 exhibits the calculated results of the Weisz–Prater number at different temperatures. Within the temperature range of 333.15 to 363.15 K, the Cwp values are significantly smaller than 1, indicating that the influence of internal diffusion can be neglected.

2.2. The Effect of Reaction Conditions

The impacts of various reaction parameters, including catalyst loading, temperature, and reactant molar ratio on the esterification of PM and AA were investigated. The effect of the amount of catalyst on the PMA yield was studied by varying catalyst loading from 5 to 12 wt% while all other reaction parameters were kept identical. The results, as illustrated in Figure 2, demonstrate that when catalyst loading is below 10 wt%, an increase in the catalyst dosage leads to an acceleration in the reaction rate. The maximum equilibrium yield of PMA is 78% when catalyst loading is 10 wt%. This indicates that the catalyst amount exerts an influence on the reaction rate but not on the reaction equilibrium, which is consistent with previous report [51].
However, only a slight increase in the PMA yield is observed when the catalyst loading is further increased from 10 to 12%, which is probably caused by the existence of mass transfer resistance when excess catalyst is used under the same reaction conditions [42,52]. Furthermore, the utilization of an excess of catalyst increases the cost of the reaction system. Consequently, a catalyst loading of 10 wt% is deemed to be the optimal catalyst amount for the esterification reaction between AA and PM.
The effect of the initial molar ratio of PM to AA on the reaction was studied by varying the molar ratios of PM to AA from 1:1 to 1:4, as shown in Figure 3, with a catalyst loading of 10 wt% and reaction temperature of 353 K. The results revealed that the conversion of PM significantly improved as the initial molar ratio of PM to AA increased from 1:1 to 1:3. Similar results have been observed in previous studies of esterification of benzyl acetate [47]. Nevertheless, as the initial molar ratio of PM to AA continued to increase to 1:4, the increased trend of PM conversion became slower. This is due to the fact that, within a certain range, an increase in the amount of AA results in an acceleration of collision frequency between reactant molecules, thereby increasing the reaction rate. Furthermore, AA also acted as a solvent to dilute the concentration of PM, which in turn results in a decrease in PM conversion. In consideration of subsequent separation issues, an excessively high concentration of AA would render separation more challenging, and consequently, the optimal initial molar ratio of PM to AA is set at 1:3.
In order to investigate the effect of temperature on the reaction, the molar ratio of PM to AA was fixed to 1:3 and the catalyst loading was fixed to 10 wt%. Figure 4 illustrates the relationship between reaction rate and temperature within the studied temperature range. As the reaction temperature increases from 333 to 363 K, the reaction rate accelerates, while the equilibrium yield of PMA remains relatively constant at different reaction temperatures. The experimental results indicated that the temperature had a more significant impact on the initial reaction rate than on the final reaction rate, which is consistent with previous reports [13]. An increase in temperature facilitates the free movement of molecules, thereby enhancing the frequency of collisions between reactant molecules, which in turn leads to an increase in the reaction rate [52]. Nevertheless, an increase in temperature also results in elevated energy consumption and safety concerns, and 353 K is identified as the optimal reaction temperature.
In order to evaluate the catalytic performance of Amberlyst-35 for direct synthesis of PMA, the esterification of PM and AA based on various solid catalysts are presented in Table 2. Solid acid SO42−/TiO2 prepared using impregnation achieved a PMA yield of 73% under conditions of 383.15 K, 10 wt% catalyst loading, and an initial molar ratio of 1:3 [3]. As well, the NKC-9 cation exchange resin used as the catalyst exhibited a PMA yield of only 46% with 10 wt% catalyst loading and an initial molar ratio of 1:1 at 353.15 K [14]. In addition, a high PMA yield (78%) was obtained using Amberlyst-15 at 353.15 K with 10 wt% catalyst loading and an initial molar ratio of 1:3 [12]. Herein, Amberlyst-35 is comparable in catalytic performance under similar reaction conditions. This could be attributed to its higher acid capacity and larger pore size, which effectively promote reactant diffusion and provide sufficient active sites, thereby enhancing the efficiency of esterification reactions [53].
Figure 5 illustrates that the initial reaction rate that exhibits exponential growth with temperature. The calculation of the initial reaction rate is represented by Equation (4).
r 0 = ( d C PM d t ) t = 0
For every 10-degree increase in temperature, the initial reaction rate doubles. The substantial impact of temperature on reaction rate suggests that the reaction is regulated by either internal diffusion or surface reaction. The result has demonstrated that the esterification reaction between PM and AA is not limited by internal diffusion, as shown in Figure 1. Consequently, the surface reaction is considered to be the limiting step of the esterification reaction.

2.3. Chemical Reaction Thermodynamic Equilibrium

The PM-AA-PMA-H2O system exhibits a relatively high level of non-ideality [12,14]. In general, the non-ideality of the liquid phase mixture necessitates the use of an activity-based model. In accordance with Equation (5), the activity αi of component i is proportional to its mole fraction xi:
α i = γ i x i
where γi represents the activity coefficient of component i. The activity coefficient approach is applicable to liquid mixtures based on a conductor-like screening model for real solvents (COSMO-RS) model. This method enables the prediction of interaction energies and activity coefficients in complex liquid systems without the need for experimental data. The COSMO-RS model is capable of calculating the chemical potential of any solute in any pure or mixed solvent, thereby enabling the prediction of thermodynamic properties, such as activity coefficients and solubilities.
The activity coefficient is calculated by Equation (6).
ln   γ i = ( μ i sol μ i p ) / R T
where μisol represents the chemical potential of solute i in the solvent, and μip represents the chemical potential of solute i in the pure solute.
The chemical reaction for the esterification of AA and PM is represented by Equation (7).
AA + PM     PMA + H 2 O
This reaction is an acid-catalyzed esterification and is subject to thermodynamic equilibrium. The reaction equilibrium constant based on mole fractions (Kx) is given by Equation (8).
K x = x i ν i = x PMA x H 2 O x PM x AA
The reaction equilibrium constant based on activities (Kα) is predicted by Equation (9).
K α = α i ν i = ( x i γ i ) ν i = x PMA x H 2 O x PM x AA γ PMA γ H 2 O γ PM γ AA
The calculated activity coefficients for each component corresponding to the experimental mole fraction at equilibrium within the temperature range of 333 to 363 K are listed in Table 3. The calculated activity coefficients for each component corresponding to the experimental mole fraction at equilibrium within the temperature range of 333 to 363 K are listed in Table 2. The activity coefficient of PM ranges from 0.74 to 0.89, which is below 1, indicating strong attractive interactions between PM and other components. This could be attributed to the presence of oxygen atoms and polar hydroxyl groups in PM molecules, which facilitate hydrogen bonding or van der Waals forces with other polar molecules like AA or H2O. Conversely, the high activity coefficient of H2O ranges from 1.7 to 1.9, indicating strong repulsive interactions between water and the other components, which is consistent with the typical non-ideal behavior observed when water is mixed with non-polar components. The activity coefficients of AA and PMA are from 0.90 to 0.94, and from 1.1 to 1.2, respectively, and being close to 1 suggests that their behavior in the solution approaches an ideal mixture.
The values of Kx and Kα are calculated using Equations (11) and (12). According to the van’t Hoff equation, the relationship between the reaction equilibrium constant and temperature is given by Equation (10).
ln   K = r H θ R T + r S θ R
A linear fit was performed by the ln of the calculated equilibrium constants Kx and Kα versus the inverse of experimental temperature values. The fitting results are shown in Figure 6, and the calculated values of the standard enthalpy of the reaction (ΔrHθ) and the standard entropy of the reaction (ΔrSθ) are presented in Table 4:
The expressions relating Kx and Kα to temperature (T) are given by Equations (11) and (12).
ln   K x = 809.92 ± 43.58 T ( 1.99 ± 0.13 )
ln   K α = 1440.27 ± 34.67 T ( 2.76 ± 0.10 )
The standard Gibbs free energy of the reaction (ΔrGθ) can be calculated by Equation (13).
r G θ = r H θ T r S θ
The standard enthalpy of the reaction was determined to be −11.97 ± 0.29 kJ/mol (Kα) based on activity calculations, while the standard enthalpy based on mole fraction calculations was −6.73 ± 0.36 kJ/mol (Kx). The standard enthalpy of the reaction based on activity calculation considers the intricate interactions between molecules in the solution, thereby enhancing the accuracy of the resulting data. Previous studies have indicated that the relationship between the equilibrium constant of the reaction and temperature is not particularly strong, suggesting a relatively low value for ΔrH [13,54], which is consistent with our results. These results indicate that the influence of temperature on the initial reaction rate was more significant than its impact on the equilibrium conversion rate, and the reaction is exothermal. A negative reaction entropy value indicates a reduction in the degree of chaos within the system. The standard Gibbs free energy of the reaction is calculated to be −5.12 ± 0.38 kJ/mol using Equation (13), indicating that the reaction is spontaneous; however, the reaction could not take place owing to the very slow reaction rate at a standard state.

2.4. Reaction Kinetic Modelling

Both internal and external mass transfer resistances have been eliminated, as shown in Figure 1 and Table 1, and thus, the reaction rate is dependent on the adsorption of the reaction components on the heterogeneous catalyst. The pseudo-homogeneous (PH), Eley–Rideal (ER), and Langmuir–Hinshelwood–Hougen–Watson (LHHW) kinetic models are frequently employed to correlate kinetic data pertaining to esterification reactions.
The PH model is widely applied in esterification systems, where the adsorption and desorption of all components can be neglected [52,55,56,57,58,59]. The PH model assumes that the catalyst swells completely upon contact with a polar solvent and that the cation exchange resin is equivalent to a liquid acid center, treating the entire reaction system as a homogeneous phase. Both the LHHW and ER models are suitable and applicable for multiphase catalytic reactions when the surface reaction is the controlling step. The LHHW model is effective in describing surface reactions between adsorbed molecules, while the ER model is well-suited to describing surface reactions between an adsorbed substance and a free substance in the liquid phase.
The esterification reaction between AA and PM is reversible, and an excess amount of AA is added in order to enhance the conversion of PM. Consequently, the reaction rate is expressed as the consumption rate of PM. The formulations for the PH, ER, and LHHW models are presented in Equations (14)–(16):
r PM = d C PM d t = k + ( C PM C AA 1 K x C PMA C H 2 O )
r PM = d C PM d t = k + ( C PM C AA ( 1 / K x ) C PMA C H 2 O ) ( 1 + K PM C PM + K H 2 O C H 2 O )
r PM = d C PM d t = k + ( C PM C AA ( 1 / K x ) C PMA C H 2 O ) ( 1 + K AA C AA + K PM C PM + K PMA C PMA + K H 2 O C H 2 O ) 2
where CPM, CAA, CPMA, and CH2O are the molar concentrations of PM, AA, PMA, and H2O, respectively. Kx is the reaction equilibrium constant based on mole fractions, Ki represents the adsorption equilibrium constant of component i, and k+ is the rate constant of the forward reaction.
In consideration of the non-ideality of the liquid phase, the activity-based kinetic models are represented by Equations (17)–(19):
r PM = d C PM d t = k + ( α PM α AA 1 K α α PMA α H 2 O )
r PM = d C PM d t = k + ( α PM α AA ( 1 / K α ) α PMA α H 2 O ) ( 1 + K PM α PM + K H 2 O α H 2 O )
r PM = d C PM d t = k + ( α PM α AA ( 1 / K α ) α PMA α H 2 O ) ( 1 + K AA α AA + K PM α PM + K PMA α PMA + K H 2 O α H 2 O ) 2
where αPM, αAA, αPMA, and αH2O represent the activities of PM, AA, PMA, and H2O, respectively, and Kα is the reaction equilibrium constant based on activities.
The adsorption equilibrium amount (qe, mg/g) of each component on the solid acid catalyst is calculated according to Equation (20):
q e = V ( C 0 C e ) m
where C0 and Ce (g/L) represent the mass concentrations of the solution before adsorption and at equilibrium, respectively, V (L) is the volume of the solution, and m (g) is the mass of the catalyst.
The Langmuir adsorption isotherm model is employed with relevant expression, which is provided by Equation (21):
1 q e = 1 q m K i 1 C e + 1 q m
where Ki is the adsorption equilibrium constant, Ce (g/L) is the mass concentration of the solution at adsorption equilibrium, and qm (g/g) represents the theoretical maximum adsorption capacity of the solid acid catalyst.
According to the Arrhenius equation, the relationship between k+ and reaction temperature (T) is given by Equation (22):
k + = k 0 + exp ( E a + R T )
where k0+ is the pre-exponential factor of the reaction, Ea+ (kJ/mol) is the activation energy of the reaction, and R is the gas constant.
As both the ER model and LHHW model involve the adsorption of components from the liquid phase onto solid acid catalysts, adsorption experiments were conducted as a preliminary step. The Langmuir adsorption isotherm model was employed, and its expression is given by Equation (20). The qm of the solid acid catalyst and Ki obtained from the Langmuir model are presented in Table 5.
Table 5 presents the adsorption equilibrium the constants of PM, AA, PMA, and H2O on Amberlyst-35 at 303 K. The variation in the adsorption equilibrium constants with temperature follows the van’t Hoff rule. Generally, as the temperature increased by 10 K, the adsorption equilibrium constants decreased 10–30%. Since highly accurate adsorption equilibrium constants are not required for fitting kinetic models, the predicted adsorption equilibrium constants are within a temperature range of 333 to 363 K based on experimental values at 313 K and reports in the literature [53,60], and are presented in Table 6.
The obtained adsorption equilibrium constants exhibit considerable error, rendering them unsuitable for quantitative analysis. They can, however, be employed as qualitative references. Consequently, these parameter values were employed as initial values for the fitting calculations of the kinetic models, rather than being directly utilized as the adsorption equilibrium constant terms in the ER and LHHW models.
Python is a high-level programming language that is widely used in scientific computing and data analysis. Due to the non-ideality of the system, the kinetic models were fitted using the activity-based Equations (17)–(19), employing the Python 3.8 programming language. The kinetic parameters and error indicators are presented in Table 7. Although all models exhibit similar values for the root mean square error (RMSE) and coefficient of determination (R2), the LHHW model had the smallest mean absolute error (MAE). Consequently, among the three models employed, the LHHW model exhibited the most favorable correlation. The parity plot for the experimental and predicted rate of reaction is shown in Figure 7, indicating that the predicted concentration values of the three models are in good agreement with the experimental values.
According to the fitting results of the LHHW model, the apparent activation energy for the esterification reaction between AA and PM catalyzed by Amberlyst-35 is determined to be 62.0 ± 0.2 kJ/mol. This result is in close agreement with the reported value of 60.5 kJ/mol for the forward reaction activation energy by Wang B et al. [14]. The satisfactory correlation between the experimental data and LHHW model suggests that the reaction is governed by surface reactions, which is in accordance with the findings in Section 3.2 regarding the impact of temperature on the esterification reaction.
According to the LHHW model, PM and AA are independently adsorbed on the catalyst surface (σ), as illustrated in Equations (23) and (24), after which a surface reaction occurs to form PMA, as shown in Equation (25). Finally, produced PMA and H2O are desorbed and diffuse into the liquid phase, as shown in Equations (26) and (27).
PM + σ     PM σ
AA + σ     AA σ
PM σ + AA σ     PMA σ +   H 2 O σ
PMA σ     PMA   + σ
H 2 O σ     H 2 O   + σ

3. Experimental

3.1. Material

Acetic acid (AA, ≥99%), 1-methoxy-2-propanol (PM, ≥99%), and 1-methoxy-2-propyl acetate (PMA, ≥99%) were supplied by Aladdin Biochemical Technology Co., Ltd (Shanghai, China). The chemicals were directly used as received without further purification. The ion-exchange resin Amberlyst-35 was supplied by Shanghai Eastern Rohm & Haas Co., Ltd (Shanghai, China). Before use, the catalyst was placed in a muffle furnace at 110 °C for 8 h to remove moisture and impurities.

3.2. Esterification Reaction

The esterification reaction was conducted in a three-necked, round-bottom flask with a volume of 100 mL. A condenser was attached to the top of the flask to prevent the loss of components with anti-backflow suction. The solid acid catalyst and a fixed molar ratio of AA and PM were loaded into the flask. Once the seal was confirmed, the flask was placed in a preheated oil bath with a specific temperature under continuous stirring. The acid-to-ether ratio was within the range of 1:1 to 1:4, and the temperature was within the range of 333 to 363 K. The catalyst Amberlyst-35 was employed in quantities of 5 to 12% in the total mass of AA and PM. Subsequent to the initiation of the reaction, the sample was collected at regular intervals. A precise quantity of 0.5 mL of the liquid sample was taken using a syringe, which was immediately cooled in order to prevent further reaction, and filtered using a needle filter.
In order to determine the adsorption equilibrium constants of the solid acid catalyst with each component, a series of independent experiments were conducted. The catalyst Amberlyst-35 was added to test tubes containing AA, PM, PMA, and H₂O, which were then vigorously stirred at room temperature using a constant temperature oscillator (SD2-100, China) at 800 rpm for 8 h. After adsorption was complete, the supernatant was collected and prepared for analysis.

3.3. Product Analysis

The concentrations of AA, PM, and PMA were determined using an Agilent 7890A gas chromatograph (USA) equipped with a flame ionization detector (FID) and an HP-FFAP chromatographic column (30 m × 0.32 mm × 0.25 μm). The analysis was conducted under the following conditions: nitrogen gas with a purity of 0.999 was used as a carrier gas, and the injector and detector temperatures were maintained at 150 °C. The initial column temperature of 60 °C was increased to 90 °C at a rate of 10 °C/min, kept for 1 min, and then rapidly heated to 230 °C at a rate of 30 °C/min. Toluene was used as an internal standard for quantitative analysis.
The conversion of PM (XPM) and the yield of PMA (YPMA) were calculated using Equations (28) and (29), respectively:
X PM = Mole   of   PM   reacted Mole   of   PM   initially
Y PMA = Mole   of   PMA   produced Mole   of   PM   initially
In all experiments, mass balance was maintained within ±5%. The yield of PMA was accurate within a range of ±2–5%, while the conversion of PM was accurate within ±3–6%.
For each experiment, a data point represented the average of three independent measurements and an error bar indicated the standard deviation. In the direct esterification of PM with AA, the PMA yield versus time exhibited an exponential trend, so the trend line was fitted with the ExpAssoc function, which can clearly visualize the data trend.
The formation rate of PMA (rPMA) was calculated from the slope of the produced PMA (mol) vs. time (s) and dry catalyst mass (W, g) by means of Equation (30):
r PMA ( t ) = 1 W d n PMA d t t

4. Conclusions

The ion-exchange resin Amberlyst-35 was employed as a catalyst for the esterification reaction of PM and AA to form PMA. The impacts of varying reaction parameters, including temperature, catalyst dosage, the initial molar ratio of reactants, and stirring speed on the reaction were investigated. Under optimized reaction conditions with initial reactant PM to AA molar ratio of 1:3 and catalyst loading of 10 wt% for a period of 3 h at 353 K, the equilibrium yield of PMA could reach 78%. A reaction thermodynamic equilibrium model was considered to analyze the reaction’s enthalpy and entropy changes. Furthermore, reaction kinetic models, including PH, ER, and LHHW, were fitted, and the LHHW model demonstrated the most favorable correlation, indicating that the reaction was controlled by surface reactions. The reaction thermodynamic equilibrium and kinetic models serve as a promising foundation to support further efforts to develop economical and environmentally-friendly PMA industrial production routes through the application of ion-exchange resin catalysts.

Author Contributions

Conceptualization, L.C. and Z.Q.; methodology, L.C.; software, X.L.; validation, X.L. and S.W.; formal analysis, X.L. and M.W.; investigation, X.L. and S.W.; resources, L.C.; data curation, X.L. and M.W.; writing—original draft preparation, X.L.; writing—review and editing, L.C.; visualization, X.L. and L.C.; supervision, L.C.; project administration, Z.Q.; funding acquisition, L.C. and Z.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (Grant Nos. 22278134 and 22472055).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hussain, A.; Chaniago, Y.D.; Riaz, A.; Lee, M. Process design alternatives for producing ultra-high-purity electronic-grade propylene glycol monomethyl ether acetate. Ind. Eng. Chem. Res. 2019, 58, 2246. [Google Scholar] [CrossRef]
  2. Chaniago, Y.D.; Hussain, A.; Andika, R.; Lee, M. Reactive pressure-swing distillation toward sustainable process of novel continuous ultra-high-purity electronic-grade propylene glycol monomethyl ether acetate manufacture. ACS Sustain. Chem. Eng. 2019, 7, 18677. [Google Scholar] [CrossRef]
  3. Huang, Z.; Lin, Y.; Li, L.; Ye, C.; Qiu, T. Preparation and shaping of solid acid SO42−/TiO2 and its application for esterification of propylene glycol monomethyl ether and acetic acid. Chin. J. Chem. Eng. 2017, 25, 1207. [Google Scholar] [CrossRef]
  4. Ye, C.S.; Dong, X.L.; Zhu, W.J.; Cai, D.R.; Qiu, T. Isobaric vapor-liquid equilibria of the binary mixtures propylene glycol methyl ether plus propylene glycol methyl ether acetate, methyl acetate plus propylene glycol methyl ether and methanol plus propylene glycol methyl ether acetate at 101.3 kPa. Fluid Phase Equilib. 2014, 367, 45. [Google Scholar] [CrossRef]
  5. Chaniago, Y.D.; Lee, M. Distillation design and optimization of quaternary azeotropic mixtures for waste solvent recovery. J. Ind. Eng. Chem. 2018, 67, 255. [Google Scholar] [CrossRef]
  6. Jyothi, M.S.; Nagarajan, V.; Chandiramouli, R. Adsorption properties of glycol ethers on cubic germanane nanosheets: A first-principles study. J. Phys. Chem. Solids. 2024, 188, 111888. [Google Scholar] [CrossRef]
  7. Kelsey, J.R. Ethylene oxide derived glycol ethers: A review of the alkyl glycol ethers potential to cause endocrine disruption. Regul. Toxicol. Pharmacol. 2022, 129, 105113. [Google Scholar] [CrossRef]
  8. Kelsey, J.R.; Cnubben, N.H.P.; Bogaards, J.J.P.; Braakman, R.B.H.; van Stee, L.L.P.; Smet, K. The urinary metabolic profile of diethylene glycol methyl ether and triethylene glycol methyl ether in Sprague-Dawley rats and the role of the metabolite methoxyacetic acid in their toxicity. Regul. Toxicol. Pharmacol. 2020, 110, 104512. [Google Scholar] [CrossRef]
  9. Wang, Y.-W.; Li, X.-X.; Zhang, Z.-L.; Liu, Q.; Li, D.; Wang, X. Volumetric, viscosimetric and spectroscopic studies for aqueous solution of ethylene glycol monoethyl ether. J. Mol. Liq. 2014, 196, 192. [Google Scholar] [CrossRef]
  10. Oh, J.; Sreedhar, B.; Donaldson, M.E.; Frank, T.C.; Schultz, A.K.; Bommarius, A.S.; Kawajiri, Y. Transesterification of propylene glycol methyl ether in chromatographic reactors using anion exchange resin as a catalyst. J. Chromatogr. A 2016, 1466, 84. [Google Scholar] [CrossRef]
  11. Oh, J.; Sreedhar, B.; Donaldson, M.E.; Frank, T.C.; Schultz, A.K.; Bommarius, A.S.; Kawajiri, Y. Transesterification of propylene glycol methyl ether by reactive simulated moving bed chromatography using homogeneous catalyst. Adsorpt.-J. Int. Adsorpt. Soc. 2018, 24, 309. [Google Scholar] [CrossRef]
  12. Gadekar-Shinde, S.; Reddy, B.; Khan, M.; Chavan, S.; Saini, D.; Mahajani, S. Reactive distillation for the production of methoxy propyl acetate: Experiments and simulation. Ind. Eng. Chem. Res. 2017, 56, 832. [Google Scholar] [CrossRef]
  13. Oh, J.; Agrawal, G.; Sreedhar, B.; Donaldson, M.E.; Schultz, A.K.; Frank, T.C.; Bommarius, A.S.; Kawajiri, Y. Conversion improvement for catalytic synthesis of propylene glycol methyl ether acetate by reactive chromatography: Experiments and parameter estimation. Chem. Eng. J. 2015, 259, 397. [Google Scholar] [CrossRef]
  14. Wang, B.; Cai, X.; Lu, L.; Ma, J. A combination of low-temperature reactive and extractive distillation for methoxy-2-propyl acetate synthesis and separation process: Simulations and experiments. J. Chem. Technol. Biotechnol. 2019, 94, 3292. [Google Scholar] [CrossRef]
  15. Cho, M.; Han, M. Dynamics and control of entrainer enhanced reactive distillation using an extraneous entrainer for the production of butyl acetate. J. Process Control. 2018, 61, 58. [Google Scholar] [CrossRef]
  16. Kiss, A.A. Novel catalytic reactive distillation processes for a sustainable chemical industry. Top. Catal. 2019, 62, 1132. [Google Scholar] [CrossRef]
  17. Rodriguez, M.; Li, P.Z.; Diaz, I. A control strategy for extractive and reactive dividing wall columns. Chem. Eng. Process. 2017, 113, 14. [Google Scholar] [CrossRef]
  18. Simasatitkul, L.; Kaewwisetkul, P.; Wiyaratn, W.; Assabumrungrat, S.; Arpornwichanop, A. Optimal design and performance analyses of the glycerol ether production process using a reactive distillation column. J. Ind. Eng. Chem. 2016, 43, 93. [Google Scholar] [CrossRef]
  19. Wang, R.Z.; Chen, G.Z.; Qin, H.; Cheng, H.Y.; Chen, L.F.; Qi, Z.W. Systematic screening of bifunctional ionic liquid for intensifying esterification of methyl heptanoate in the reactive extraction process. Chem. Eng. Sci. 2021, 246, 116888. [Google Scholar] [CrossRef]
  20. Zeng, Q.; Hu, B.; Cheng, H.Y.; Chen, L.F.; Huang, J.M.; Qi, Z.W. Liquid-liquid equilibrium for the system of ionic liquid [BMIm] [HSO4] catalysed isobutyl isobutyrate formation. J. Chem. Thermodyn. 2018, 122, 162. [Google Scholar] [CrossRef]
  21. Zeng, Q.; Song, Z.; Qin, H.; Cheng, H.Y.; Chen, L.F.; Pan, M.; Heng, Y.; Qi, Z.W. Ionic liquid [BMIm] [HSO4] as dual catalyst-solvent for the esterification of hexanoic acid with n-butanol. Catal. Today 2020, 339, 113. [Google Scholar] [CrossRef]
  22. Zhang, C.; Liu, T.; Wang, H.-J.; Wang, F.; Pan, X.-Y. Synthesis of acetyl salicylic acid over WO3/ZrO2 solid superacid catalyst. Chem. Eng. J. 2011, 174, 236. [Google Scholar] [CrossRef]
  23. Gao, A.; Chen, H.; Tang, J.; Xie, K.; Hou, A. Efficient extraction of cellulose nanocrystals from waste Calotropis gigantea fiber by SO42−/TiO2 nano-solid superacid catalyst combined with ball milling exfoliation. Ind. Crops Prod. 2020, 152, 112524. [Google Scholar] [CrossRef]
  24. Hassan, M.A.; Wang, W.; Chang, Z.; Li, M.; Dong, B.; Ndonfack, K.I.A.; Li, W.; Sun, C. TiO2/SO42− solid superacid catalyst prepared by recovered TiO2 from waste SCR and its application in transesterification of ethyl acetate with n-butanol. Waste Biomass Valorization 2023, 14, 4035. [Google Scholar] [CrossRef]
  25. Kuwahara, Y.; Kaburagi, W.; Nemoto, K.; Fujitani, T. Esterification of levulinic acid with ethanol over sulfated Si-doped ZrO2 solid acid catalyst: Study of the structure-activity relationships. Appl. Catal. A 2014, 476, 186. [Google Scholar] [CrossRef]
  26. Lokman, I.M.; Goto, M.; Rashid, U.; Taufiq-Yap, Y.H. Sub- and supercritical esterification of palm fatty acid distillate with carbohydrate-derived solid acid catalyst. Chem. Eng. J. 2016, 284, 872. [Google Scholar] [CrossRef]
  27. Sheikh, R.; Choi, M.-S.; Im, J.-S.; Park, Y.-H. Study on the solid acid catalysts in biodiesel production from high acid value oil. J. Ind. Eng. Chem. 2013, 19, 1413. [Google Scholar] [CrossRef]
  28. Firouzjaee, M.H.; Taghizadeh, M. n-Butane isomerization over 0.5 Pt/HPA/MCM-41 catalyst: Optimization study using central composite design. Chem. Eng. Technol. 2024, 47, 887. [Google Scholar] [CrossRef]
  29. Li, L.; Liu, Y.; Wu, K.; Liu, C.; Tang, S.; Yue, H.; Lu, H.; Liang, B. Catalytic solvent regeneration of a CO2-loaded MEA solution using an acidic catalyst from industrial rough metatitanic acid. Greenh. Gases. 2020, 10, 449. [Google Scholar] [CrossRef]
  30. Yati, I.; Yeom, M.; Choi, J.-W.; Choo, H.; Suh, D.J.; Ha, J.-M. Water-promoted selective heterogeneous catalytic trimerization of xylose-derived 2-methylfuran to diesel precursors. Appl. Catal. A 2015, 495, 200. [Google Scholar] [CrossRef]
  31. Xu, W.; Ollevier, T.; Kleitz, F. Iron-modified mesoporous silica as an efficient solid lewis acid catalyst for the mukaiyama aldol reaction. ACS Catal. 2018, 8, 1932. [Google Scholar] [CrossRef]
  32. Orabona, F.; Capasso, S.; Perez-Sena, W.Y.; Taddeo, F.; Eränen, K.; Verdolotti, L.; Tesser, R.; Di Serio, M.; Murzin, D.; Russo, V.; et al. Solvent-free condensation of ethyl levulinate with phenol promoted by Amberlyst-15: Kinetics and modeling. Chem. Eng. J. 2024, 493, 152677. [Google Scholar] [CrossRef]
  33. Badia, J.H.; Soto, R.; Ramírez, E.; Bringué, R.; Fité, C.; Iborra, M.; Tejero, J. Role of ion-exchange resins in hydrogenation reactions. Catalysts 2023, 13, 624. [Google Scholar] [CrossRef]
  34. Jiang, T.; Zhao, Q.; Li, M.; Yin, H. Preparation of mesoporous titania solid superacid and its catalytic property. J. Hazard. Mater. 2008, 159, 204. [Google Scholar] [CrossRef] [PubMed]
  35. Huang, H.; Mu, J.; Liang, M.; Qi, R.; Wu, M.; Xu, L.; Xu, H.; Zhao, J.; Zhou, J.; Miao, Z. One-pot synthesis of MoO3-ZrO2 solid acid catalyst for solvent-free solketal production from glycerol. Mol. Catal. 2024, 552, 113682. [Google Scholar] [CrossRef]
  36. Waldron, C.; Pankajakshan, A.; Quaglio, M.; Cao, E.; Galvanin, F.; Gavriilidis, A. Closed-Loop Model-Based Design of Experiments for Kinetic Model Discrimination and Parameter Estimation: Benzoic Acid Esterification on a Heterogeneous Catalyst. Ind. Eng. Chem. Res. 2019, 58, 22165. [Google Scholar] [CrossRef]
  37. Echaroj, S.; Asavatesanupap, C.; Chavadej, S.; Santikunaporn, M. Kinetic study on microwave-assisted oligomerization of 1-decene over a HY catalyst. Catalysts 2021, 11, 1105. [Google Scholar] [CrossRef]
  38. Bertero, N.M.; Apesteguia, C.R.; Marchi, A.J. Selective synthesis of 1-indanol by 1-indanone liquid-phase hydrogenation over metal-based catalysts: A LHHW kinetic approach. Chem. Eng. Sci. 2022, 254, 117629. [Google Scholar] [CrossRef]
  39. Delgado, P.; Sanz, M.T.; Beltran, S. Kinetic study for esterification of lactic acid with ethanol and hydrolysis of ethyl lactate using an ion-exchange resin catalyst. Chem. Eng. J. 2007, 126, 111. [Google Scholar] [CrossRef]
  40. Mao, W.; Wang, X.; Wang, H.; Chang, H.; Zhang, X.; Han, J. Thermodynamic and kinetic study of tert-amyl methyl ether (TAME) synthesis. Chem. Eng. Process. 2008, 47, 761. [Google Scholar] [CrossRef]
  41. Kuusisto, J.; Mikkola, J.-P.; Sparv, M.; Waerna, J.; Karhu, H.; Salmi, T. Kinetics of the catalytic hydrogenation of D-lactose on a carbon supported ruthenium catalyst. Chem. Eng. J. 2008, 139, 69. [Google Scholar] [CrossRef]
  42. Shan, X.; Cheng, Z.; Yuan, P. Reaction kinetics and mechanism for hydration of cyclohexene over ion-exchange resin and H-ZSM-5. Chem. Eng. J. 2011, 175, 423. [Google Scholar] [CrossRef]
  43. Sert, E.; Atalay, F.S. Kinetic study of the esterification of acetic acid with butanol catalyzed by sulfated zirconia. React. Kinet. Mech. Catal. 2010, 99, 125. [Google Scholar] [CrossRef]
  44. Casas, C.; Bringué, R.; Fité, C.; Iborra, M.; Tejero, J. Kinetics of the liquid phase dehydration of 1-octanol to di-n-octyl ether on Amberlyst 70. AIChE J. 2017, 63, 3966. [Google Scholar] [CrossRef]
  45. Yip, A.C.K.; Hu, X. Catalytic Activity of Clay-Based Titanium Silicalite-1 Composite in Cyclohexanone Ammoximation. Ind. Eng. Chem. Res. 2009, 48, 8441. [Google Scholar] [CrossRef]
  46. Akyalcin, S.; Altiokka, M.R. Kinetics of esterification of acetic acid with 1-octanol in the presence of Amberlyst 36. Appl. Catal. A 2012, 429, 79. [Google Scholar] [CrossRef]
  47. Ali, S.H.; Merchant, S.Q. Kinetic study of Dowex 50 Wx8-catalyzed esterification and hydrolysis of benzyl acetate. Ind. Eng. Chem. Res. 2009, 48, 2519. [Google Scholar] [CrossRef]
  48. Patel, D.; Saha, B. Esterification of Acetic Acid with n-Hexanol in Batch and Continuous Chromatographic Reactors Using a Gelular Ion-Exchange Resin as a Catalyst. Ind. Eng. Chem. Res. 2012, 51, 11965. [Google Scholar] [CrossRef]
  49. Martinez, A.F.; Sanchez, C.A.; Orjuela, A.; Gil, I.D.; Rodriguez, G. Isobutyl acetate by reactive distillation. Part II. Kinetic study. Chem. Eng. Res. Des. 2020, 160, 447. [Google Scholar] [CrossRef]
  50. Pera-Titus, M.; Bausach, M.; Tejero, J.; Iborra, M.; Fite, C.; Cunill, F.; Izquierdo, J.F. Liquid-phase synthesis of isopropyl tert-butyl ether by addition of 2-propanol to isobutene on the oversulfonated ion-exchange resin Amberlyst-35. Appl. Catal. A 2007, 323, 38. [Google Scholar] [CrossRef]
  51. Son, Y.R.; Park, J.K.; Shin, E.W.; Moon, S.P.; Park, H.E. Synthesis of propylene glycol methyl ether acetate: Reaction kinetics and process simulation using heterogeneous catalyst. Processes 2024, 12, 865. [Google Scholar] [CrossRef]
  52. Chin, S.Y.; Ahmad, M.A.A.; Kamaruzaman, M.R.; Cheng, C.K. Kinetic studies of the esterification of pure and dilute acrylic acid with 2-ethyl hexanol catalysed by Amberlyst 15. Chem. Eng. Sci. 2015, 129, 116. [Google Scholar] [CrossRef]
  53. Soto, R.; Oktar, N.; Fité, C.; Ramírez, E.; Bringué, R.; Tejero, J. Experimental study on the liquid-phase adsorption equilibrium of n-Butanol over Amberlyst™15 and contribution of diffusion resistances. Chem. Eng. Technol. 2021, 44, 2210. [Google Scholar] [CrossRef]
  54. Ali, S.H. Kinetics of catalytic esterification of propionic acid with different alcohols over Amberlyst 15. Int. J. Chem. Kinet. 2009, 41, 432. [Google Scholar] [CrossRef]
  55. Li, K.; Xue, W.; Zeng, Z.; Shi, X. Kinetics of the reaction of ethanol and lauric acid catalyzed by deep eutectic solvent based on benzyltrimethylammonium chloride. Can. J. Chem. Eng. 2019, 97, 1144. [Google Scholar] [CrossRef]
  56. Qin, H.; Hu, X.; Wang, J.; Cheng, H.; Chen, L.; Qi, Z. Overview of acidic deep eutectic solvents on synthesis, properties and applications. Green Energy Environ. 2020, 5, 8. [Google Scholar] [CrossRef]
  57. Wang, R.; Qin, H.; Wang, J.; Cheng, H.; Chen, L.; Qi, Z. Reactive extraction for intensifying 2-ethylhexyl acrylate synthesis using deep eutectic solvent Im:2PTSA. Green Energy Environ. 2021, 6, 405. [Google Scholar] [CrossRef]
  58. Yang, J.; Zhou, L.; Guo, X.; Li, L.; Zhang, P.; Hong, R.; Qiu, T. Study on the esterification for ethylene glycol diacetate using supported ionic liquids as catalyst: Catalysts preparation, characterization, and reaction kinetics, process. Chem. Eng. J. 2015, 280, 147. [Google Scholar] [CrossRef]
  59. Wang, R.; Qin, H.; Song, Z.; Cheng, H.; Chen, L.; Qi, Z. Toward reactive extraction processes for synthesizing long-chain esters: A general approach by tuning bifunctional deep eutectic solvent. Chem. Eng. J. 2022, 445, 136664. [Google Scholar] [CrossRef]
  60. Agrawal, G.; Oh, J.; Sreedhar, B.; Tie, S.; Donaldson, M.E.; Frank, T.C.; Schultz, A.K.; Bommarius, A.S.; Kawajiri, Y. Optimization of reactive simulated moving bed systems with modulation of feed concentration for production of glycol ether ester. J. Chromatogr. A 2014, 1360, 196. [Google Scholar] [CrossRef]
Figure 1. Effect of stirring speed on yield of PMA. Reaction conditions: PM:AA = 1:3 (molar ratio) and 10 wt% catalyst loading of total reactant mass at 363 K.
Figure 1. Effect of stirring speed on yield of PMA. Reaction conditions: PM:AA = 1:3 (molar ratio) and 10 wt% catalyst loading of total reactant mass at 363 K.
Molecules 29 04709 g001
Figure 2. Effect of catalyst loading on yield of PMA (wt% of total reactant mass). Reaction conditions: PM:AA = 1:3 (molar ratio) at 353 K.
Figure 2. Effect of catalyst loading on yield of PMA (wt% of total reactant mass). Reaction conditions: PM:AA = 1:3 (molar ratio) at 353 K.
Molecules 29 04709 g002
Figure 3. Effect of initial molar ratio of PM to AA. Reaction conditions: 10 wt% catalyst loading of the total reactant mass at 353 K.
Figure 3. Effect of initial molar ratio of PM to AA. Reaction conditions: 10 wt% catalyst loading of the total reactant mass at 353 K.
Molecules 29 04709 g003
Figure 4. Effect of temperature on yield of PMA. Reaction conditions: PM:AA = 1:3 (molar ratio), 10 wt% catalyst loading.
Figure 4. Effect of temperature on yield of PMA. Reaction conditions: PM:AA = 1:3 (molar ratio), 10 wt% catalyst loading.
Molecules 29 04709 g004
Figure 5. Effect of reaction temperature on the initial reaction rate. Reaction conditions: PM:AA = 1:3 (molar ratio), 10 wt% catalyst loading.
Figure 5. Effect of reaction temperature on the initial reaction rate. Reaction conditions: PM:AA = 1:3 (molar ratio), 10 wt% catalyst loading.
Molecules 29 04709 g005
Figure 6. Relationship between lnKx/lnKα and 1000/T.
Figure 6. Relationship between lnKx/lnKα and 1000/T.
Molecules 29 04709 g006
Figure 7. Parity plots for the experimental and predicted rate of reaction for (a) PH, (b) ER, and (c) LHHW.
Figure 7. Parity plots for the experimental and predicted rate of reaction for (a) PH, (b) ER, and (c) LHHW.
Molecules 29 04709 g007aMolecules 29 04709 g007b
Table 1. Calculated infinite dilution diffusion coefficient (DA), effective diffusion coefficient (De), and Weisz–Prater number (Cwp) at different temperatures.
Table 1. Calculated infinite dilution diffusion coefficient (DA), effective diffusion coefficient (De), and Weisz–Prater number (Cwp) at different temperatures.
Temperature (K)Cs (mol/cm3)robs (mol/(s·g))DA (cm2/s)De (cm2/s)Cwp
333.151.484 × 10−22.525 × 10−62.307 × 10−52.423 × 10−60.03
343.151.483 × 10−24.973 × 10−62.576 × 10−52.705 × 10−60.05
353.151.482 × 10−28.726 × 10−62.841 × 10−52.983 × 10−60.08
363.151.482 × 10−21.589 × 10−53. 990 × 10−53.254 × 10−60.13
Table 2. Comparison of PMA yield for various catalysts reported in references.
Table 2. Comparison of PMA yield for various catalysts reported in references.
CatalystTemperature (K)Initial Molar Ratio (PM:AA)Catalyst Loading (wt%)Yield of PMA (%)Reference
SO42−/TiO2383.151:31073[3]
NKC-9353.151:11046[14]
Amberlyst-15353.151:31078[12]
Amberlyst-35353.151:31078This work
Amberlyst-35353.151:11055This work
Table 3. Mole fractions and evaluated activity coefficients of components in the equilibrium state of the reaction at various temperatures (initial composition of PM:AA = 1:3).
Table 3. Mole fractions and evaluated activity coefficients of components in the equilibrium state of the reaction at various temperatures (initial composition of PM:AA = 1:3).
Temperature (K)PMAAPMAH2O
xPMγPMαPMxAAγAAαAAxPMAγPMAαPMAxH2OγH2OαH2O
333.150.05470.88700.04850.55160.93330.51480.19691.1870.23370.19691.8160.3575
343.150.05300.84330.04470.55040.92390.50850.19831.1790.23390.19831.8120.3593
353.150.05020.79630.04000.54930.91350.50180.20021.1690.23400.20021.8000.3604
363.150.04700.74470.03500.55040.90330.49720.20131.1520.23190.20131.7770.3577
Table 4. Thermodynamic parameters of the reaction.
Table 4. Thermodynamic parameters of the reaction.
Equilibrium ConstantΔrHθ (kJ/mol)ΔrSθ (J/(mol·K))R2
Kx−6.73 ± 0.36−16.5 ± 1.00.9942
Kα−11.97 ± 0.29−23.0 ± 0.80.9988
Table 5. Fitting calculation results of Langmuir adsorption isotherm models at 303 K.
Table 5. Fitting calculation results of Langmuir adsorption isotherm models at 303 K.
Componentqm (g/g)Ki (L/g)R2
PM0.47680.02710.995
AA0.35820.02500.833
PMA2.19210.00010.926
H2O0.73570.00710.953
Table 6. Predicted values of adsorption equilibrium constants for each component at different temperatures.
Table 6. Predicted values of adsorption equilibrium constants for each component at different temperatures.
Temperature (K)KPM (L/g)KAA (L/g)KPMA (mL/g)KH2O (L/g)
333.150.02330.02030.05900.0059
343.150.01930.01700.05500.0046
353.150.01500.01400.05100.0035
363.150.01070.00990.04700.0022
Table 7. Kinetic parameters for different models.
Table 7. Kinetic parameters for different models.
Modelk0+Ea+ (kJ/mol)KPMKAAKPMAKH2OMAERMSER2
PH2.88 × 10663.2 ± 0.3////1.26 × 10−31.59 × 10−30.9964
ER4.25 × 10663.3 ± 0.20.12//0.011.22 × 10−31.57 × 10−30.9947
LHHW6.84 × 10662.0 ± 0.20.130.048.2 × 10−40.151.13 × 10−31.59 × 10−30.9946
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, X.; Wang, S.; Wang, M.; Chen, L.; Qi, Z. Reaction Thermodynamic and Kinetics for Esterification of 1-Methoxy-2-Propanol and Acetic Acid over Ion-Exchange Resin. Molecules 2024, 29, 4709. https://doi.org/10.3390/molecules29194709

AMA Style

Liu X, Wang S, Wang M, Chen L, Qi Z. Reaction Thermodynamic and Kinetics for Esterification of 1-Methoxy-2-Propanol and Acetic Acid over Ion-Exchange Resin. Molecules. 2024; 29(19):4709. https://doi.org/10.3390/molecules29194709

Chicago/Turabian Style

Liu, Xinyu, Shu Wang, Mingxia Wang, Lifang Chen, and Zhiwen Qi. 2024. "Reaction Thermodynamic and Kinetics for Esterification of 1-Methoxy-2-Propanol and Acetic Acid over Ion-Exchange Resin" Molecules 29, no. 19: 4709. https://doi.org/10.3390/molecules29194709

Article Metrics

Back to TopTop