The Nature of Acid-Catalyzed Acetalization Reaction of 1,2-Propylene Glycol and Acetaldehyde

Similar documents
Reversible Additions to carbonyls: Weak Nucleophiles Relative Reactivity of carbonyls: Hydration of Ketones and Aldehydes

Supplementary Information for Efficient catalytic conversion of fructose into hydroxymethylfurfural by a novel carbon based solid acid

Original Articles. 1. Introduction. You Zhou 1,, Liwei Xue 1, Kai Yi 1, Li Zhang 1, Seung Kon Ryu 2 and Ri Guang Jin 1

Synthesis of jet fuel range cycloalkanes with diacetone alcohol. from lignocellulose

Esterification of Acetic Acid with Butanol: Operation in a Packed Bed Reactive Distillation Column

Effects of Different Processing Parameters on Divinylbenzene (DVB) Production Rate

Experimental and Simulation Study on the Reactive Distillation Process for the Production of Ethyl Acetate

Synthesis and adsorption property of hypercross-linked sorbent

KINETICS OF ESTERIFICATION OF CITRIC ACID WITH BUTANOL USING SULPHURIC ACID

Mechanism and Kinetics of the Synthesis of 1,7-Dibromoheptane

The effect of phase transition of methanol on the reaction rate in the alkylation of hydroquinone

Adsorption Processes. Ali Ahmadpour Chemical Eng. Dept. Ferdowsi University of Mashhad

Prepared for Presentation at the 2004 Annual Meeting, Austin, TX, Nov. 7-12

The esterification of acetic acid with ethanol in a pervaporation membrane reactor

Carbon dioxide removal processes by alkanolamines in aqueous organic solvents Hamborg, Espen Steinseth

Abstract Process Economics Program Report 37B ACETIC ACID AND ACETIC ANHYDRIDE (November 1994)

University of Groningen. Levulinic acid from lignocellulosic biomass Girisuta, Buana

Separation and recovery of mangenese- and cobalt catalysts from wastewater of PTA unit

Simulation of Butyl Acetate and Methanol Production by Transesterification Reaction via Conventional Distillation Process

January 19, 2012, Workshop on Chemical Data Reporting (CDR) Rule Case Studies for Byproduct/Recycling Reporting

Chapter 9 Aldehydes and Ketones

Measurement and correlation of solubility of xylitol in binary water+ethanol solvent mixtures between K and K

Kinetics. 1. Consider the following reaction: 3 A 2 B How is the average rate of appearance of B related to the average rate of disappearance of A?

Synthesis of isobutyl propionate using Amberlyst 15 as a catalyst

Investigation of adiabatic batch reactor

Aldehydes and Ketones 2. Based on Organic Chemistry, J. G. Smith 3rde.

Lecture 25: Manufacture of Maleic Anhydride and DDT

Chapter 15 Reactions of Aromatic Compounds

Kinetic Isotope Effects

Chemistry 6A F2007. Dr. J.A. Mack 11/19/07. Chemical Kinetics measure the rate of appearance of products or the rate of disappearance of reactants.

REFERENCE LIST. 2. Alime, & Halit, L. (2007). Kinetics of Synthesis of Isobutyl Propionate over Amberlyst-15. Turk J Chem.

Structure of the chemical industry

LATEST TECHNOLOGY IN Safe handling & Recovery OF Solvents in Pharma Industry

12/27/2010. Chapter 15 Reactions of Aromatic Compounds

Effect of Recycle and Feeding Method on Batch Reactive Recovery System of Lactic Acid

Production of Renewable 1,3-Pentadiene from Xylitol via Formic. Acid-Mediated Deoxydehydration and Palladium-Catalyzed

Catalytic Esterification of Acetic Acid with Methanol; Comparison of Photocatalytic and Acid Catalysed Esterification

Lecture (9) Reactor Sizing. Figure (1). Information needed to predict what a reactor can do.

Comparison of vapor adsorption characteristics of acetone and toluene based on polarity in activated carbon fixed-bed reactor

Direct Preparation of Dichloropropanol from Glycerol Biodiesel Based Using HCl 37%: a Preliminary Study

Bioreactor Engineering Laboratory

less stable intermediate

Chemical Kinetics. Chapter 13. Copyright The McGraw-Hill Companies, Inc. Permission required for reproduction or display.

Dual Use of a Chemical Auxiliary: Molecularly Imprinted Polymers for the Selective Recovery of Products from Biocatalytic Reaction Mixtures

Heterogeneous Kinetic Study for Esterification of Acetic Acid with Ethanol

Absorption of carbon dioxide into non-aqueous solutions of N-methyldiethanolamine

Optimization of Batch Distillation Involving Hydrolysis System

Kinetic investigations on the esterification of phthalic anhydride with n-heptyl, n-nonyl or n-undecyl alcohol over sulfuric acid catalyst

SUPPORTING INFORMATION. Elucidation of the role of betaine hydrochloride in glycerol esterification: towards bio-based ionic building blocks

Selective O-Alkylation Reaction of Hydroquinone with Methanol over Cs Ion-Exchanged Zeolites

GCSE Chemistry. Module C7 Further Chemistry: What you should know. Name: Science Group: Teacher:

Protein separation and characterization

Chemical Reactions and Kinetics of the Carbon Monoxide Coupling in the Presence of Hydrogen

Improvement of separation process of synthesizing MIBK by the isopropanol one-step method

Experiment 2 Solvent-free Aldol Condensation between 3,4-dimethoxybenzaldehyde and 1-indanone

Energy Changes, Reaction Rates and Equilibrium. Thermodynamics: study of energy, work and heat. Kinetic energy: energy of motion

Chapter 23 Aldehydes and Ketones

Salting-out extraction of 1,3-propanediol from fermentation broth

Lecture 15. More Carbonyl Chemistry. Alcohols React with Aldehydes and Ketones in two steps first O R'OH, H + OR" 2R"OH R + H 2 O OR" 3/8/16

Solvent Extraction Research and Development, Japan, Vol. 23, No 2, (2016)

Initial carbon carbon bond formation during synthesis gas conversion to higher alcohols on K Cu Mg 5 CeO x catalysts

CHAPTER 1 INTRODUCTION

Methanolysis of Poly(ethylene terephthalate) in Supercritical Phase

GCSE CHEMISTRY REVISION LIST

Purification of 2-mercaptobenzothiazole by solvent extraction

Methylation of benzene with methanol over zeolite catalysts in a low pressure flow reactor

Greener and sustainable method for alkene epoxidations by polymer-supported Mo(VI) catalysts

Organic Chemistry Review: Topic 10 & Topic 20

Supporting Information

Alkaline Hydrolysis of Polyethylene Terephthalate at Lower Reaction Temperature

CHEMICAL KINETICS (RATES OF REACTION)

Theoretical Models for Chemical Kinetics

Simultaneous Removal of COS and H 2 S at Low Temperatures over Nanoparticle α-feooh Based Catalysts

Solvent Extraction Research and Development, Japan, Vol. 23, No 2, (2016)

A NEW SOLVENT FOR CO2 CAPTURE R.

Agilent J&W PoraBOND Q PT Analyzes Oxygenates in Mixed C4 Hydrocarbon Streams by GC/FID and GC/MSD

SELECTIVE OXIDATION OF TOLUENE TO BENZALDEHYDE USING Cu/Sn/Br CATALYST SYSTEM

Overview. Types of Solutions. Intermolecular forces in solution. Concentration terms. Colligative properties. Osmotic Pressure 2 / 46

A NOVEL PROCESS CONCEPT FOR THE PRODUCTION OF ETHYL LACTATE

Esterification of acrylic acid with ethanol using pervaporation membrane reactor

Lecture Presentation. Chapter 14. James F. Kirby Quinnipiac University Hamden, CT. Chemical Kinetics Pearson Education, Inc.

Application Note. Authors. Abstract. Energy & Chemicals, Biofuels & Alternative Energy

Hydrogenation of CO Over a Cobalt/Cerium Oxide Catalyst for Production of Lower Olefins

Supporting Information

CHEM 251 (4 credits): Description

Chapter 7 Adsorption thermodynamics and recovery of uranium

Chapter 16 Aldehydes and Ketones I. Nucleophilic Addition to the Carbonyl Group

Chapter 12. Chemical Kinetics

Chapter 11 Rate of Reaction

An alcohol is a compound obtained by substituting a hydoxyl group ( OH) for an H atom on a carbon atom of a hydrocarbon group.

Properties of Solutions and Kinetics. Unit 8 Chapters 4.5, 13 and 14

Supporting Information

Preparation and Characterization of Double Metal Cyanide Complex Catalysts

Chapter 15. Reactions of Aromatic Compounds. Electrophilic Aromatic Substitution on Arenes. The first step is the slow, rate-determining step

Chapter 10: Carboxylic Acids and Their Derivatives

Module: 7. Lecture: 36

Lesmahagow High School CfE Higher Chemistry. Chemical Changes & Structure Controlling the Rate

Separation Characteristics of Lactic Acid in Reactive Extraction and Stripping

But in organic terms: Oxidation: loss of H 2 ; addition of O or O 2 ; addition of X 2 (halogens).

ORGANIC - BROWN 8E CH ALDEHYDES AND KETONES.

Transcription:

Korean Chem. Eng. Res., 53(4), 463-467 (2015) http://dx.doi.org/10.9713/kcer.2015.53.4.463 PISSN 0304-128X, EISSN 2233-9558 The Nature of Acid-Catalyzed Acetalization Reaction of 1,2-Propylene Glycol and Acetaldehyde Chen Cheng*, Hui Chen*,, Xia Li*, Jianli Hu* and Baochen Liang* *School of Chemistry & Chemical Engineering, Tianjin University of Technology, Tianjin, China. (Received 11 September 2014; Received in revised form 29 December 2014; accepted 30 December 2014) Abstract We investigated catalytic activity of ion-exchange resins in acetalization of 1,2-propylene glycol with acetaldehyde. The impacts of reaction variables, such as temperature, reaction time, catalyst loading and feedstock composition, on the conversion of 1,2-propylene glycol were measured. The life of the catalyst was also studied. Furthermore, the reaction kinetics of 1,2-propylene glycol acetalization was studied. It was found that reaction rate followed the firstorder kinetics to acetaldehyde and 1,2-propylene glycol, respectively. Therefore, overall acetalization reaction should follow the second-order reaction kinetics, expressed as r kc na nb A C B 19.74e -------------- 6650 T C 1 1 = = A C B. Key words: 1,2-propylene Glycol, 2,4-dimethyl-1,3-dioxolane, Ion-exchange Resin, Polyhydroxy Compounds, Acetalization 1. Introduction Polyhydroxy compounds such as 1,2-propylene glycol (PG) and ethylene glycol are important chemical intermediates widely used in producing polyester, surfactant, lubricant, and adhesives [1-5]. PG is difficult and expensive to recover from aqueous solution because of its high boiling point and affinity with water. Reactive distillation was reported to be applied to PG purification [6]. In the reactive distillation column where an acid catalyst is present, PG reacts reversibly with acetaldehyde to form 2,4-dimethyl-1,3-dioxolane. The reaction is known as an acetalization reaction. The reaction pathway involves protonation of acetaldehyde followed by the addition of nucleophilic PG molecule to the carbonyl group to form an intermediate, which is hemiacetal. In the presence of an acid catalyst, etherification reaction occurs within hemiacetal molecule to form 2,4-dimethyl-1,3-dioxolane and water [7]. The reaction equation is shown in Scheme 1. The acidic function on solid acid catalyst is important in activating PG molecule to facilitate acetalization reaction. Ion-exchange resins are very effective in catalyzing acetal formation [8,9]. The use of heterogeneous ion-exchange resin catalysts offers many advantages over homogeneous acidic catalyst. Dhale et al. studied the recovery of Scheme 1. Reaction equation of PG and acetaldehyde. To whom correspondence should be addressed. E-mail: chenhui_410@163.com This is an Open-Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/bync/3.0) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited. 1,2-propylene glycol from water via acetal formation using ionexchange resins as catalysts in reactive distillation [6]. Acid-catalyzed acetalization is used to obtain intermediates or final compounds. In the past few decades, most of the studies reported in the literature for acetalization reactions focused on analyzing and comparing the activity of the solid catalysts [10]. However, in a recent trend, researchers directed their attention to the acidic macroporous resin to protect the environment. Acetaldehyde and 1,2-propylene glycol can be recovered and purified via the reverse hydrolysis reaction. High yield of 2,4-dimethyl-1,3-dioxolane can be obtained in the equilibrium limited reaction by removing the products [11]. Although there have been several studies describing the recovery of aqueous glycol solution via reactive extraction and reactive distillation [12-14], systematic experimental data, especially kinetic modeling work, have not been reported. This study presents a comprehensive investigation on the formation of acetals via reaction between PG and acetaldehyde using an ionexchange resin catalyst. The objective of this research was to measure the catalytic activity of ion-exchange resin catalysts in PG acetalization under various reaction conditions and to determine the optimal conditions for the formation of acetals. The specific objective was to elucidate equilibrium reaction and kinetics of the acetalization reaction [15]. The experimental data obtained from this research serves as a basis for the design of commercial reactive distillation column for acetalization reaction of polyhydroxy compounds [16]. 2. Experimental Section 2-1. Apparatus and Procedure Analytical grade PG and acetaldehyde (40 wt% solution) were used in all experiments. The ion-exchanged resin, D072, was purchased from the Chemical Plant of Nankai University, Tianjin, China. The physical properties of the resin D072 are listed in Table 1. The strong 463

464 Chen Cheng, Hui Chen, Xia Li, Jianli Hu and Baochen Liang Table 1. The basic properties of D072 Type D072 The structure of resin Styrene-DVB Functional group - -SO 3 Ion-exchange capacity (mmol/g) 4.2 Granularity (0.315~1.25 mm) 95 The highest temperature ( o C) Na 120 PH range 1~14 The factory form Na Foreign reference product Amberlyst-15; Diaion HPK-16 acidic ion-exchange resin possesses high ion exchange rate. The resin framework consists of an irregular, macromolecular and three-dimensional network of hydrocarbon chains. The matrix carries sulfonic acid -SO 3 H as a functional group on the surface. The reaction temperature, reaction time, catalyst loading and acetaldehyde/pg volume ratio were variables measured in catalytic performance test and kinetic study. After the reaction, the catalyst was separated by centrifugation. A typical experiment procedure was as follows: 100 ml acetaldehyde, 40 ml PG (acetaldehyde/pg volume ratio=2.5:1) and 3.5 g D072 (25 g/l) were used in a 250 ml glass reactor, respectively. Experiments were under ambient pressure in a constant temperature bath at 288 K with a stirring speed of 150 rpm. Small amount of sample (0.2 micron liter) was drawn at 30 minutes time interval for GC analysis. 2-2. Analytical Method The PG was analyzed by GC (Ling Hua 9890A gas chromatograph equipped with FID detector and an SE-54 capillary column, 50 m 0.32 mm 2.0 μm). Column temperature was programmed to maintain at 333 K for 5.5 min initially, then increased to 473 K at the rate of 298 K/min. The concentration of PG was determined by using the internal standard method. Tetrahydrofuran was chosen as an internal standard for GC calibration. The PG conversion was calculated by using the following formula: Ф = [(C 0 -C t ) / C 0 ] 100% where C 0 is the initial concentration of PG and C t is the concentration of PG at a reaction time of t. 3. Results and Discussion 3-1. The Impact of Reaction Variables on PG Acetalization PG acetalization is an equilibrium-limited reaction; therefore, some reaction variables affect the time and extent to reach equilibrium state. The dependence of PG conversion on key reaction variables was investigated. These reaction variables include temperature, reaction time, catalyst loading and the acetaldehyde/pg volume ratio. The effect of reaction temperature on PG conversion over Resin D072 catalyst is illustrated in Fig. 1(a). By raising temperature from 273 K to 288 K, PG conversion significantly increases, reaching maximum conversion of 90.3±0.3% at 288 K. PG conversion starts to decline when temperature is continually increased. Below 288 K, PG acetalization reaction is a kinetically controlled region, i.e., forward reaction rate is faster than reverse reaction rate and reaction is not in equilibrium state. When reaction temperature is increased to approach 288 K, the reverse reaction rate is increased at a faster pace to equal the forward reaction rate; then equilibrium state is reached. When reaction temperature is further increased beyond 288 K, the reverse reaction rate is greater than the forward reaction rate, PG conversion decreases, shifting equilibrium to the left; therefore, a downward PG conversion curve is observed. PG acetalization is an exothermic reaction. Thermodynamically, an increase in reaction temperature favors reverse reaction. Experimental results shown in Fig. 1(a) indicate that at a reaction temperature greater than 288 K, PG acetalization reaction is in equilibrium controlled region. The impact of reaction time on PG conversion is shown Fig. 1(b). Initially, PG conversion increases sharply, then starts to level off as the reaction proceeds. At 180 min, PG conversion reaches 90.3±0.3% and remains at this level as the reaction continues. This indicates that the reaction has reached equilibrium under the experimental conditions. The ion-exchange resin catalysts are essential for PG acetalization to yield 2,4-dimethyl-1,3-dioxolane. The effect of catalyst loading on PG conversion was studied. Resin D072 catalyst was used and each experiment was at 288 K, 3 hours reaction time with acetaldehyde/pg volume ratio being 2.5:1. As shown in Fig. 1(c), PG conversion increases as catalyst loading is increased from 10 g/l to 25 g/l. The increase in reaction rate can be attributed to the increase in the total number of acid sites available for acetalization. Nevertheless, further increase in the catalyst loading beyond 30 g/l does not show any significant enhancement in PG conversion. Therefore, the optimal catalyst loading is determined to be 25 g/l for the PG acetalization reaction. The increase in PG conversion with the increase in the amount of catalyst loading is associated with the number of available active sites for acetalization of PG. PG acetalization with acetaldehyde is an equilibrium limited reaction. To maximize PG conversion, an excess amount of acetaldehyde is necessary to shift equilibrium to the right. The effect of acetaldehyde/pg volume ratio on PG conversion was studied under the conditions of 288 K, catalyst loading of 25 g/l and 3 hours reaction time. The acetaldehyde/pg volume ratio was varied from 0.5:1 to 3.5:1. The relationship between acetaldehyde/pg volume ratio and PG conversion is shown in Fig. 1(d). When acetaldehyde/pg volume ratio is increased from 0.5:1 to 2.5:1, PG conversion increases and reaches a maximum of 90.3±0.3%, after which a steep decrease in PG conversion is observed. The upward PG conversion curve at acetaldehyde/pg volume ratio less than 2.5:1 can be explained by the shift of equilibrium to the direction that favors acetal formation because of the increase in acetaldehyde concentration. However, because acetaldehyde is introduced as 40% water solution, further increase in acetaldehyde/pg volume ratio actually leads to a decrease in PG concentration; the equilibrium can shift to the left that favors hydrolysis of 2,4-dimethyl- 1,3-dioxolane, so a downward PG conversion curve is observed.

The Nature of Acid-Catalyzed Acetalization Reaction of 1,2-Propylene Glycol and Acetaldehyde 465 Fig. 1. Effects of reaction variables on PG acetalization over Resin D072 catalyst. (a) effect of temperature (catalyst loading=25 g/l, reaction time=3 hours, acetaldehyde/pg volume ratio=2.5:1), (b) effect of reaction time (T=288 K, catalyst loading=25 g/l, acetaldehyde/pg volume ratio=2.5:1), (c) effect of catalyst loading (T=288 K, reaction time=3 hours, acetaldehyde/pg volume ratio=2.5:1), (d) effect of acetaldehyde/pg volume ratio (T=288 K, catalyst loading=25 g/l reaction time=3 hours). 3-2. Catalyst Life Test The ion-exchange resin catalysts deactivate over time in acetalization of polyhydroxy compounds. After a certain period of time-on-stream operation, catalytic performance may not satisfy economic requirements from the manufacturing point of view. There is a need to determine how many cycles the catalyst can be used in batch reaction. To establish the reusability guideline for commercial application, after each test, the catalyst was recovered and tested for the subsequent acetalization of PG. Before the start of the subsequent test, the reclaimed catalyst was washed with solvent for several times and dried in an oven at 328 K. Each testing cycle was under the same reaction conditions. As illustrated in Fig. 2, during the first six cycles of the test, catalytic activity measured by PG conversion decreased from 90.3±0.3% to 67.7±0.3%, probably due to the loss of the number of functional group. The deactivation rate appears to be linear at the beginning but not catastrophic. The deactivation rate slows down after four cycles of test. Although the deactivation mechanism and regeneration are not the focus of this study, we hypothesize that the deactivation is Fig. 2. Performance of ion-exchange Resin D072 catalyst in repeated cycle test (Resin D072 loading=25 g/l, T=288 K, reaction time= 3 hours, acetaldehyde/pg volume ratio=2.5:1). caused by forming strong bonding between the acid sites and polymeric intermediates. Moreover, the agitator in the batch reactor can

466 Chen Cheng, Hui Chen, Xia Li, Jianli Hu and Baochen Liang also cause mechanical damage of the resin [17,18]. At the same time, we speculate that some of certain substances in the reaction process occupied the active site of catalyst, leading to decreased catalytic efficiency. To prove this conjecture, we used a 4% aqueous hydrochloric acid washing the catalyst, which has been used ten times. When the washed catalyst was put into use again, we found that PG conversion rate could be raised to 80±0.3%. This phenomenon demonstrates our inference very well. However, the material to occupy the center position of the catalyst activity requires further studies. 3-3. Reaction Kinetics Determining the reaction orders of the kinetics is important in the design of a reactive distillation column for acetalization of polyhydroxy compounds. Guemez et al. [19,20] studied kinetics of acetalization reaction between glycerol and n-butyraldehyde. However, the kinetics for the acetalization of PG with acetaldehyde has not been published yet. In the kinetic study, the reaction mixture of PG and acetaldehyde was charged to a batch reactor. PG and acetaldehyde conversions were measured as reaction proceeded. Fig. 3 shows the conversion of acetaldehyde as a function of time. ExpDec1 equation in Origin 9 was fitted to the data?which can be expressed as: y = y 0 + A 1 e -x/t1 Among them, -1/t 1 is the first order reaction rate constant. Non-linear regression calculation was carried out for the kinetic modeling work. The results indicated that the reaction was first order for acetaldehyde. Similarly, the conversion of PG as a function of time was measured according to the method described above. The result indicated that reaction was first-order for PG as well. Therefore, overall reaction follows second-order kinetics, which can be expressed as: r=kc A 1 C B 1 where r is reaction rate and k is an Arrhenius factor defined as: Fig. 3. The conversion of acetaldehyde as a function of time (Resin D072 loading=25 g/l, T=288 K, acetaldehyde/pg volume ratio =2.5:1). Fig. 4. The conversion of PG as a function of time at different temperature. Table 2. The influence of reaction temperature on kinetic rate constants for PG acetalization T/K k/min -1 273 0.01075 278 0.01351 283 0.02222 288 0.04000 293 0.05000 k = A 0 e Ea ------ RT The values of preexponential factor A 0 and activation energy E a were determined experimentally. PG conversion was measured at different temperature varying from 273 K to 293 K. As illustrated in Fig. 4, the reaction rate increases with temperature. Non-linear regression calculation was conducted to obtain the reaction rate constant k at each temperature. The result listed in Table 2 shows the apparent rate constant k increases with reaction temperature. To the temperature of the reaction, five sets of different reaction temperature were used in the experiment. The results (Fig. 4) show that as the reaction temperature increases, the reaction time is reduced to reach equilibrium. This phenomenon followed the endothermic reaction theory. When the reaction time exceeds 300 min, all reactions have reached equilibrium. After equilibrium is reached, under a different temperature, the conversion rate of PG has no significant change. However, when the reaction temperature reaches equilibrium at 293 K, the conversion rate of PG is lower than 288 K. Therefore, we conclude that the best reaction temperature is 288 K. The Arrhenius plot is illustrated in Fig. 5 where the apparent rate constant is plotted against 1/RT. A 0 and E a are obtained from the intercept and slope shown in Fig. 5. Therefore, the reaction kinetics equation for PG aldolization can be expressed as: r kc na nb = A C B = 19.74e -------------- 6650 T C 1 1 A C B [21].

The Nature of Acid-Catalyzed Acetalization Reaction of 1,2-Propylene Glycol and Acetaldehyde 467 Fig. 5. Arrhenius plot for acetylation of PG with acetaldehyde. 4. Conclusions This study not merely demonstrates that the resin based catalyst is capable of catalyzing 1,2-Propylene Glycol acetalization. It goes beyond the catalytic performance study by exploring the reaction kinetics that occurred on the catalyst. These experimental results have helped us to gain insights into the mechanism on the formation of acetals via reverse acetalization reaction over ion-exchange resin catalysts. The experimental data can be used as a basis for the design of commercial reactive distillation column and for the simulation of acetalization of polyhydroxy compounds with aldehyde to produce acetals. Acknowledgment The author would like to thank the financial support by the Golden Eagle Technology Co., Ltd. Tianjin, China. References 1. Dietrieh, A. and Norbert, W., Method of Preparing 1,3-propanediol, US Patent 5015789(1991). 2. Allen, K. D., Preparation of 1,3-Propanediol Used in Polyester Production, US Patent 5777182(1998). 3. Texaco Chemical Company, Process for Cogeneration of Ethylene Glycol and Dimethyl Carbonate, US Patent 5214182(1993). 4. Arcnson, D. R. and King, C. J., Separation of low Molecular Weight Alcohols from Dilute Aqueous Solutions by Reversible Chemical Complexation, Ph.D. Thesis. LBL Report 24944(1989). 5. Ko, C. H., Park, S. H., Jeon, J. K., Suh, D. J., Jeong, K. E. and Park, Y. K., Upgrading of Biofuel by the Catalytic Deoxygenation of Biomass, Korean J. Chem. Eng., 29(12), 1657-1665(2012). 6. Dhale, A. D., Myrant, L. K., Chopade, S. P., Jackson, L. E. and Miller, D. J., Propylene Glycol and ethylene glycol Recovery from Aqueous Solution via Reactive Distillation, Chem. Eng. Sci., 59(14), 2881-2890(2004). 7. Mahajani, S. M., Kolah, A. K. and Sharma, M. M., Extractive Reactions with Cationic Exchange Resins as Catalysts, React. Funct. Polym., 28(1), 29-38(1995). 8. Mamontov, G. V., Knyazev, A. S., Paukshtis, E. A. and Vodyankina, O. V., Adsorption and Conversion of Ethylene Glycol on the Surface of Ag-Containing Catalyst Modified with Phosphate, Kinet. Catal., 54(6), 735-743(2013). 9. Zhang, T. S. and Ding, Q. W., Acetalization over Amberlyst D072 Type Cation Exchange Resin Catalyst, Journal of Taiyuan University of Science and Technology., 3, 251-253(2007). 10. Chopade, S. P. and Sharma, M. M., Reaction of Ethanol and Formaldehyde: Use of Versatile Cation-exchange Resins as Catalyst in Batch Reactors and Reactive Distillation Columns, React. Funct. Polym., 32(1) 53-65(1997). 11. Agirre, I., Guemez, M. B., Veen, V. M. V., Motelica, A., Vente, J. F. and Arias, P. L., Acetalization Reaction of Ethanol with Butyraldehyde Coupled with Pervaporation. Semi-batch Pervaporation Studies and Resistance of HybSi Membranes to Catalyst Impacts, J. Membr. Sci., 371(1-2), 179-188(2011). 12. Tink, R. R. and Neish, A. C., Extraction of Polyhydroxy Compounds from Dilute Aqueous Solutions by Cyclic Acetalic Cetal formation I. As investigation of the Scope of the Process, Canadian Journal of Technology., 29, 243(1951). 13. Broekhuis, R. R., Lynn, S., King, C. J., Recovery of Propylene Glycol from Dilute Aqueous Solutions via Reversible Reaction with Aldehydes, Ind. Eng. Chem. Res., 33(12), 3230-3237(1994). 14. Chopade, S. P. and Sharma, M. M., Acetalization of Ethylene glycol with Formaldehyde using Cation-Exchange Resins as Catalysts: Batch versus Reactive Distillation, React. Funct. Polym., 34(1), 37-45(1997). 15. Li, Y. J., Zhu, J. W., Wu, Y. Y. and Liu, J. X., Reactive-extraction of 2,3-butanediol from Fermentation Broth by Propionaldehyde: Equilibrium and Kinetic Study, Korean J. Chem. Eng., 30(1), 73-81(2013). 16. Sun, Q. T., Pan, C. J. and Yan, X. F., Simulation and Analysis of Dehydration Distillation Column Based on Distillation Mechanism Integrated with Neural Network, Korean J. Chem. Eng., 30(3), 518-527(2013). 17. Dixit, A. B. and Yadav, G. D., Deactivation of Ion-exchange Resin Catalysts. Part I: Alkylation of o-xylene with Styrene, React. Funct. Polym., 31(3), 237-250(1996). 18. Dixit, A. B. and Yadav, G. D., Deactivation of Ion-exchange Resin Catalysts. Part II: Simulation by Network Models, React. Funct. Polym., 31(3), 251-263(1996). 19. Guemez, M. B., Requies, J., Agirre, I., Arias, P. L., Barrio, V. L. and Cambra, J. F., Acetalization Reaction Between Glycerol and n-butyraldehyde Using An Acidic Ion Exchange Resin: Kinetic Modeling, Chem. Eng. J., 228(15), 300-307(2013). 20. Hong, X., McGiveron, O., Kolah, A. K., Orjuela, A., Peereboom, L., Lira, C. T. and Miller, D. J., Reaction Kinetics of Glycerol Acetal Formation via Transacetalization with 1,1-Diethoxyethane, Chem. Eng. J., 222(15), 374-381(2013). 21. Erdem, B. and Cebe, M., Kinetics of Esterification of Propionic Acid with n-amyl Alcohol in the Presence of Cation Exchange Resin, Korean J. Chem. Eng., 23(6), 896-901(2006).