Skip to main content
ACS Omega logoLink to ACS Omega
. 2020 Feb 3;5(5):2088–2096. doi: 10.1021/acsomega.9b02230

Zn- and Ti-Modified Hydrotalcites for Transesterification of Dimethyl Terephthalate with Ethylene Glycol: Effect of the Metal Oxide and Catalyst Synthesis Method

Amarsinh L Jadhav 1, Radhika S Malkar 1, Ganapati D Yadav 1,*
PMCID: PMC7016940  PMID: 32064369

Abstract

graphic file with name ao9b02230_0013.jpg

The activity and selectivity of hydrotalcites (HTs) can be suitably enhanced by the addition of different metal oxides. Zinc and titanium are prospective candidates for such a modification. Transesterification of dimethyl terephthalate (DMT) with ethylene glycol (EG) using basic catalysts is an industrially important process for the production of bis(2-hydroxyethyl)terephthalate (BHET). BHET is a precursor for polyethylene terephthalate (PET) which is used in production of films, fibers, and molding materials. As against use of polluting liquid bases, solid bases could be employed. In the current work, transesterification of DMT with EG was studied over modified HT base catalysts wherein the HT was activated with the addition of zinc and titanium. These catalysts were prepared by the combustion synthesis using different fuels. The modified HT using Zn and Ti were well characterized by scanning electron microscopy, energy-dispersive X-ray spectrometry, Brunauer–Emmett–Teller surface area analyzer, temperature-programmed desorption, and X-ray diffraction. Effects of several parameters on the rate of reaction and conversion of the limiting reagent were investigated. Zinc-modified HT using glycine as fuel (Zn–HT–glycine) was found to be the most selective, active, and reusable catalyst. The Langmuir–Hinshelwood–Hougen–Watson model was used to establish the reaction mechanism and kinetics. All species were weakly adsorbed leading to a second-order kinetics. Using a mole ratio of 1:2 of DMT to EG and 0.05 g/cm3 Zn–HT–glycine loading resulted in to 64.1% conversion of DMT and 96.1% selectivity to BHET in 4 h at 180 °C. The apparent activation energy was 9.64 kcal/mol. The catalyst was robust and reusable.

1. Introduction

The green chemistry principles basically aim at waste reduction using atom economical safer processes, catalysis, and renewable resources.1 Catalysis is pivotal in developing economical and energy efficient processes and many organic transformations could be revisited from this angle.2 Heterogeneous catalysis is inherently superior to homogeneous catalysis and offers process intensification, easier catalyst separation, and less impact on the environment, reduces waste, and improves process economics.3 Many traditional bulk and fine chemical manufacturing facilities involve use of homogeneous bases which ought to be replaced by solid bases. Among all, hydrotalcites (HTs) and their modified structures have made great inroads in base catalyzed condensation reactions which cover aldol, cross-aldol, Claisen–Schmidt, and Knoevenagel condensations and also other reactions such as isomerization, alkylation, and hydrolysis reactions.46

Transesterification between dimethyl terephthalate (DMT) with ethylene glycol (EG) is a relevant process used for the production of bis(2-hydroxyethyl)terephthalate (BHET) which is a monomer for making polyethylene terephthalate (PET).79 PET is the fastest growing thermoplastic polymer which is used in the form of films, bottles, molding and fibers and increasingly in blends, composites, and nanocomposites.10,11

Transesterification is commonly catalyzed by acid or base catalysts12 and also by enzymes.14 Many catalysts such as Zn, Mn, Mg, and lead acetates7 are commonly employed for the transesterification between DMT and EG.13 In this reaction, a considerable amount of byproducts such as methyl 2-hydroxy-ethylterephthalate and methanol are formed along with BHET.9,14,15 The catalysts are homogeneous and nonreusable bases. In general, the esterification and transesterification reactions of carboxylic acids or esters with alcohols have been carried out over homogeneous catalysts such as mineral acids, metal hydroxides, and metal alkoxides. The replacement of these catalysts by solid catalysts enjoys various advantages, as was stated before.8,15 The effect of different homogeneous catalysts such as acetates of Pb, Co, Mg, and Mn and of the mixtures of Mg, Mn, and Zn acetates has been also studied.7,15,16 The reaction requires higher concentration of raw materials and catalyst which subsequently necessitates neutralization of the reaction mass and effluent treatment. Synthesis of poly(ethylene terepthalate) (PET) by transesterification of DMT with EG was performed in the presence of a few well-known catalysts including various lanthanide compounds.7,9 Mihail et al.16 suggested the reaction order of transesterification of DMT and EG being fractional over zinc oxide in the concentration range of 0.14–0.28%. Sorokin and Chebotareva17 reported that the reaction was first order with respect to EG over Zn stearate. However, it is reported that the transesterification reaction was first order in each DMT, EG, and homogenous catalyst concentration, making it overall a third-order reaction.18 Besnoin et al.7 dealt with semi-batch melt transesterification of DMT with EG using zinc acetate as a catalyst. However, all above catalysts gave less selectivity toward BHET, requiring large catalyst/reactant ratio and nongreen solvents. Serio et al.19 have prepared a basic heterogeneous catalyst for transesterification of DMT with EG. Three different catalysts such as Al2O3, MgO, and calcined HT of Al–Mg with varying combination of both metals were investigated for transesterification and successive polycondensation to get poly(ethylene terephthalate) (PET) at 180 °C.

Transesterification reactions can be effectively conducted using HT which could be modified to render better activity and selectivity. Synthesis of HTs can be achieved using coprecipitation,20 sol–gel,21 template-assisted synthesis,22 decomposition of nitrates,23 and so forth. Such procedures intrinsically contain a number of steps to get the final catalyst making them lengthy and cumbersome, as well as they need a lot of process water. Therefore, different protocols have been suggested in the literature for synthesizing metal oxide(s) like HTs. Among them, combustion synthesis is a powerful alternative for making materials which not only render reproducible results but also involve reduced number of steps. Combustion synthesis gives fine nanoscale metal oxides depending on the source of fuel and ignition temperature. It can produce metal oxides with high surface area, pore volume, and mesoporosity, and therefore, it is one of the best methods to synthesize different catalysts.24,25 Combustion synthesis needs a fewer steps to make porous materials with controllable pore radii and particle size according to application.24 In the current work, the combustion synthesis method was adopted to prepare Ti- and Zn-modified HTs by using glycine and glycerol as fuel which is one of the novel aspects of the current work. Previously, Ti- and Zn-modified HTs were prepared by coprecipitation method and characterized by different high-end characterization techniques. Velu et al.26 have developed a novel series of HT like anionic clays containing Zr4+ by a simple coprecipitation method. Further rehydration behavior of Mg2+, Al3+ and Ti4+ containing layered double hydroxides (LDH) were studied by Das and Samal.27 Selective catalytic reduction of NO with ammonia was performed over Cu, CO, and Mn containing HTs.28 In all these reports, the method of preparation of modified HT is coprecipitation. The present work deals with a comparative study of Zn- and Ti-modified HTs synthesized by using the combustion method with two different fuels, namely, glycine and glycerol. Thus, the prepared catalysts were used for the first time to study the transesterification of DMT with EG in a solvent-free condition, which has resulted into better selectivity for BHET. A thorough investigation of catalyst synthesis and characterization was undertaken, along with the reaction mechanism and kinetic model. The current reaction is carried out with novel and cost-effective zinc-modified HT base catalyst to achieve good selectivity for the desired BHET with minimum quantities of starting materials employing solvent-free condition to make the process green.

2. Results and Discussion

2.1. Catalyst Characterization

Among all of the catalysts, Zn–HT–glycine was the best for the reaction, and hence its complete characterization was done. It is recently published elsewhere2931 (Supporting Information). Only a brief description is provided here.

2.1.1. Scanning Electron Microscopy and Energy-Dispersive X-ray Spectrometry

Scanning electron microscopy (SEM) images of Zn–HT–glycine at 1500, 2500, and 5000 amplification are shown in Figure 1a–c, respectively. The average particle size is in the range of 50–100 μm. It shows uneven particles, which are relatively uniform in size. It is generally seen for materials synthesized by the combustion method.

Figure 1.

Figure 1

SEM images of Zn–HT–glycine catalyst at different magnifications (a) 1500× (b) 2500× (c) 5000×.

Energy-dispersive X-ray spectrometry (EDXS) was used to determine the whole composition of titanium, zinc, aluminum, and magnesium for all catalysts (Table 1). The ratio of composition of aluminum and magnesium was kept constant as follows: (Al/Mg ratio of 1:3) in HT–glycerol, (Al/Mg ratio of 1:2) in Ti–HT with glycine and glycerol as fuels, (Al/Mg of 1:4) in Zn–HT with glycine and glycerol as fuels.

Table 1. Composition of Different Catalysts Using EDXS.
catalyst-fuel Zn–HT–glycine
Zn–HT–glycerol
Ti–HT–glycine
Ti–HT–glycerol
HT–glycerol
element mass wt % mol % mass wt % mol % mass wt % mol % mass wt % mol % mass wt % mol %
O 29.66   31.77   42.69   42.66      
Mg 19.24 49.60 20.28 52.10 18.54 50.06 18.55 50.06 76.44 75
Al 6.97 8.60 10.37 12.00 20.76 25.26 20.79 25.26 23.56 25
Zn 44.14 42.31 37.37 35.90            
Ti         18.1 24.68 18.00 24.68    
total 100 100 100 100 100 100 100 100 100 100

2.2. Catalyst Activity

Scheme 1 shows the reaction of DMT with EG producing BHET and methanol, and the efficacy of all catalysts is given in Table 2.

Scheme 1. Overall Reaction for Production of Bis(2-hydroxyethyl)terephthalate (BHET).

Scheme 1

Table 2. Performance of Different Catalysts for Transesterification Reaction of DMT with EGa.

catalyst-fuel surface area (m2/g) pore volume (cm3/g) acidity (mmol/g) basicity (mmol/g) conversion of DMT (%) selectivity to BHET (%)
Ti–HT–glycine 226.2 0.252 0.0145 1.244 59.1 91.6
Zn–HT–glycine 132.4 0.339 0.0135 0.540 64.1 96.1
Zn–HT–glycerol 148.6 0.446 0.0145 1.770 61.0 92.7
Ti–HT–glycerol 54.6 0.227 0.719 1.085 57.0 89.8
HT–glycerol 127.0 0.418 0.6603 1.371 58.0 90.1
a

Selectivity (%) = [area of desired product/sum of area of all products] × 100.

2.2.1. Efficacy of Different Catalysts

Zn–HT–glycine, Zn–HT–glycerol, Ti–HT–glycine, Ti–HT–glycerol, and HT–glycerol were used (Figure 2) for the production of BHET. It was observed that Zn–HT–glycine having mild basicity as well as low acidity in comparison with other studied catalysts showed the best results giving 64.6% conversion of DMT and 96.1% selectivity toward BHET (Table 2).

Figure 2.

Figure 2

Effect of different catalysts on the conversion of DMT. Reaction conditions: DMT 0.12 mol, EG 0.24 mol, total volume 38.2 cm3, agitation speed 1000 rpm, catalyst loading 0.05 g/cm3, temperature 180 °C.

The results obtained under current investigation are better than those reported in the published literature where the mole ratio of DMT to EG was 1:20.7,8,15 In the current case, it was 1:2 with DMT as the limiting reactant. A blank reaction was conducted without use of any catalyst which showed that the conversion was only 1.7% in 4 h.

For all above catalysts, the conversion and selectivity were confirmed by taking repeated runs to avoid errors and are averaged.

2.3. Effect of Speed of Agitation

Different runs were conducted to study the influence of external mass-transfer resistance on conversion at agitation speed from 800 to 1200 rpm (Figure 3). The conversion increased only marginally with speed and it was practically the same at 1000 and 12000 rpm. Thus, there were no significant effect of mass-transfer resistance on rate of reaction and conversion. A theoretical calculation was also done to find diffusion coefficients, solid–liquid mass-transfer coefficients, and mass-transfer rates in comparison with observed rate of reaction. The rate of mass transfer was very high by an order of magnitude, and hence, there was no mass-transfer resistance. The theoretical background and method of calculation are given elsewhere.32,33 Thus, the subsequent runs were carried out at 1000 rpm.

Figure 3.

Figure 3

Effect of the speed of agitation on conversion of DMT. Reaction conditions: catalyst Zn–HT–glycine, DMT 0.12 mol, EG 0.24 mol, total volume 38.2 cm3, catalyst loading 0.05 g/cm3, temperature 180 °C.

2.4. Proof of Absence of External Mass-Transfer Resistance

There are different controlling mechanisms as reported in the literature for better understanding of solid–liquid or heterogeneous catalytic reactions. The liquid phase diffusivity of DMT (A) at 180 °C was calculated using Wilke–Chang equation as 1.341 × 10–5 cm2/s.34 Afterward, considering the limiting value of Sherwood number (ShA = kSL-A × dp/DAB) as 2, the mass-transfer coefficient was evaluated as 3.831 × 10–3 cm/s. The particle surface per unit liquid volume was obtained as

2.4. 1

The observed initial rate of reaction for DMT was calculated as, 3.03 × 10–5 mol cm–3 s–1, while the mass-transfer rate for DMT was evaluated as 6.6 × 10–4 mol cm–3 s–1

As

2.4. 2

that is, 3.29 × 104 ≫ 1.5 × 103

Hence, it proved that there was no external mass-transfer resistance for the reaction. The only resistance could be due to intraparticle diffusion, surface reaction, chemisorption, or desorption. It will be discussed later.

2.5. Effect of Catalyst Loading

The catalyst quantity was varied from 0.05 to 0.15 g/cm3 (Figure 4). The conversion of DMT was found to increase with catalyst loading w (g/cm3). A plot was made for initial rate (dXA/dt) versus time which shows that the rate of reaction is directly proportional to the number of available active sites, that is, mass of the catalyst (Figure 5). The Weisz–Prater modulus was also calculated to confirm the absence of intraparticle diffusion limitation and the reaction was kinetically controlled.35

Figure 4.

Figure 4

Effect of catalyst loading on conversion of DMT. Reaction conditions: catalyst Zn–HT–glycine, DMT 0.12 mol, EG 0.24 mol, total volume 38.2 cm3, speed of agitation 1000 rpm, temperature 180 °C.

Figure 5.

Figure 5

Plot of initial rate versus catalyst loading. Reaction conditions: catalyst Zn–HT–glycine, DMT 0.12 mol, EG 0.24 mol, total volume 38.2 cm3, speed of agitation 1000 rpm, temperature 180 °C.

2.6. Proof of Absence of Intraparticle Resistance

Considering 70 μm as the average particle size of the catalyst Zn–HT–glycine, theoretical calculations were performed to calculate the Weisz–Prater criterion (Cwp).36 The details of these methods are reported in the literature.37,38 The value of Cwp thus obtained is 0.051 which is far less than unity, confirming that the reaction was free from intraparticle diffusion resistance. Thus, the reaction was kinetically controlled which was further confirmed from the evaluation of apparent activation energy as given in Section 2.10.

2.7. Effect of Mole Ratio

The molar concentration of the limiting (DMT) to the excess reactant (EG) was varied from 1:1 to 1:3 at the same catalyst loading per unit volume (Figure 6). The initial rate of reaction increased with concentration.

Figure 6.

Figure 6

Effect of mole ratio of DMT to EG. Reaction conditions: catalyst Zn–HT–glycine, agitation speed 1000 rpm, catalyst loading 0.05 g/cm3, temperature 180 °C.

2.8. Effect of Temperature

Experiments were conducted at different temperatures ranging from 150 to 190 °C (Figure 7). Conversion of DMT increased with temperature. This result indicates that the reaction is controlled by kinetic step only, and mass-transfer and intraparticle diffusion resistances do not play any role during the reaction.

Figure 7.

Figure 7

Effect of temperature on conversion of DMT. Reaction conditions: catalyst Zn–HT–glycine, DMT 0.12 mol, EG 0.24 mol, total volume 38.2 cm3, agitation speed 1000 rpm, catalyst loading 0.05 g/cm3.

Table 3 gives the comparison of published research on transesterification of DMT with EG for the synthesis of BHET including the current work. Both super base catalysts, viz. CsxO/γ-Al2O3 and Na/NaOH/γ-Al2O3 resulted in to 88 and 90% selectivity of BHET, respectively, for 0.25 mol DMT per g of catalyst. However, Zn–HT–glycine in the current work showed 96.1% selectivity toward BHET for 0.072 mol DMT per g of catalyst, which is 3.47 times less. Rest of the catalysts listed in Table 4 are homogeneous in nature. Thus, the current work gives more selectivity toward BHET with the use of cheap heterogeneous base catalyst, Zn–HT–glycine, with excellent selectivity at much lower DMT/EG molar ratio, less DMT/catalyst ratio with reusable catalyst, and in solvent-free condition to make the process green. The volume of reaction mass governs the reaction rate.

Table 3. Comparative Literature of Transesterification of DMT and EG for the Synthesis of BHET with Reference to the Current Worka.

DMT/EG mol ratio catalyst temp (°C) time (h) catalyst/reactant ratio conv. (%) sel. (%) refs
1:10 CsxO/γ-Al2O3 (hetero) 180 2.5 0.25 mol DMT/g catalyst 100 88 (8)
1:20 Na/NaOH/γ-Al2O3 (hetero) 180 2.5 0.25 mol DMT/g catalyst 100 90 (8)
2:1 Mg/Mn/Zn acetate (homo) 175 3 2.55 × 104 mol DMT/mol catalyst 88 94 (15)
2:1 zinc acetate (homo) 180 3 1.83 × 104 mol DMT/mol cat   20(BHET) (15)
2:1 zinc acetate (homo) 190 4 2.5 × 103 mol DMT/mol cat 76   (7)
1:2 Zn–HT–glycine (hetero) 180 4 0.072 mol DMT/g-cat 64.1 96.1 current work
a

Selectivity (%) = [moles of desired product/sum of moles of all products] × 100.

Table 4. Kinetic Parameters of the Reaction.

reaction temperature T (°C) KA (L mol–1) KB (L mol–1) KE (L mol–1) KF (L mol–1)
160 0.1 0.002 0.0004 0.0007
170 0.16 0.0031 0.00054 0.0009
180 0.31 0.042 0.0041 0.0046
190 0.45 0.075 0.008 0.0091

2.9. Effect of Reusability of Catalyst

After each experiment, the catalyst was separated by filtration and washed with 50 cm3 methanol to desorb any adsorbed material from the catalyst pores and then dried at 120 °C. Generally, there was a loss of ∼2 to 3% catalyst which was compensated with a fresh catalyst (Figure 8). It was noticed that the catalyst activity and selectivity reduced very marginally on repetitive use of the same catalyst. Further experiments were done where no make up for the loss of catalyst during filtration was done. However, the volume of the reaction mass was reduced to maintain the same catalyst loading in g/cm3 and the same mole of reactants/mass of the catalyst. The experimental results were within ±2% proving that the activity was preserved.

Figure 8.

Figure 8

Effect of reusability on conversion of DMT Reaction conditions: catalyst Zn–HT–glycine, DMT 0.12 mol, EG 0.24 mol, total volume 38.2 cm3, agitation speed 1000 rpm, catalyst loading 0.05 g/cm3, temperature 180 °C.

2.10. Reaction Mechanism and Kinetic Model

A kinetic model was developed for the transesterification reaction in the absence of both external mass-transfer and intraparticle diffusion resistances. Out of the many models tried, the following was observed to fit the data reasonably well. In this reaction, chemisorption of A (DMT) and B (EG) takes place on two nearby vacant sites S1 and S2, respectively, according to the Langmuir–Hinshelwood–Hougen–Watson (LWHW) mechanism to give E (bis(2-hydroxyethyl terephthalate)) and F (methanol) (Scheme 2).

Scheme 2. Catalytic Cycle for the Transesterification Reaction Over Zn–HT–glycine (S1—Acidic Sites, S2 Basic Sites).

Scheme 2

Adsorption of DMT A on vacant sites S1

2.10. 3

Similarly, on another vacant site S2, EG is adsorbed as

2.10. 4

In the next step, the surface reaction of complexes takes place giving ES1 and FS2 as follows

2.10. 5

Finally, desorption of complexes formed ES1 and FS2 is represented by following reversible reactions

2.10. 6
2.10. 7

It gives regeneration of active sites S1 and S2.

If the surface reaction is the rate controlling step, then the rate of reaction of A is given by

2.10. 8

When the reaction is far away from equilibrium

2.10. 9

Replacing with total site concentration

2.10. 10
2.10. 11

If w is the catalyst loading, then

2.10. 12

The adsorption constants K1, KA, KB, KE, and KF were calculated using Polymath 6 and eq 12 (Table 4). Initial concentration of DMT (CA0) and EG (CB0) were taken as 3.63 and 7.27 mol L–1, respectively.

The values of adsorption constants were observed to be very small and hence, eq 11 leads to

2.10. 13

or

2.10. 14

where

2.10. 15

Table 4 shows that the adsorption equilibrium constants are negligible, leading to a power law model.

If the initial molar ratio of EG and DMT is Inline graphic at time t = 0, then eq 14 reduces to a typical second-order equation which can be integrated in terms of fractional conversion XA of A as follows

2.10. 16

For a fixed catalyst loading w and initial concentration of A, the pseudo-constant can be written as

2.10. 17

Thus, plots of Inline graphic versus t were made at different temperatures at fixed w and CA0 (Figure 9) which are straight lines passing through origin. It confirms that reaction is second order and validates the mathematical model. The slopes of these lines at various temperatures were used to make the Arrhenius plot (Figure 10). The apparent activation energy was obtained to be 9.64 kcal/mol which established that the reaction rate was controlled by intrinsic kinetics.

Figure 9.

Figure 9

Kinetic plot for typical second-order reaction at different temperatures using Zn–HT–glycine. Reaction conditions: DMT 0.12 mol, EG 0.24 mol, total volume 38.2 cm3, agitation speed 1000 rpm, catalyst 0.05 g/cm3 reaction mixture.

Figure 10.

Figure 10

Arrhenius plot for transesterification of DMT with EG.

3. Conclusions

Different basic zinc- and titanium-loaded HT catalysts were prepared by using the combustion synthesis method with glycerol and glycine as fuels. The effect of combustion fuel was noticed in the activity of the catalysts which were fully characterized and used in transesterification reaction of DMT with EG to synthesize bis(2-hydroxyethyl)terephthalate. Zn–HT–glycine was observed to be the best catalyst which showed better stability, activity, and selectivity. A mechanism was proposed and kinetics were thoroughly deduced using LHHW model. It was observed that all species are weakly adsorbed leading to the second-order rate equation having activation energy 9.64 kcal/mol. The current reaction was carried out with a novel cheaper zinc-modified HT as the catalyst for the first time to give good selectivity without any use of solvent to make the process green and sustainable. Comparison with the earlier literature shows that the catalyst and process conditions are superior to the previously reported research.

4. Materials and Methods

4.1. Chemicals

Aluminum nitrate nonahydrate (Al (NO3)3·9H2O), magnesium nitrate hexahydrate (Mg (NO3)2·6H2O), glycerol, zinc nitrate, glycine, titanium tetra-isopropoxide, nitric acid (70%), DMT, and EG were purchased from S.D. Fine Chemicals, Mumbai, India. A.R. grade chemicals were employed without any further purification for the synthesis of catalysts and its application.

4.2. Synthesis of Catalysts

4.2.1. Hydrotalcite

The Mg–Al–O mixed oxide catalyst was made by dissolving Al (NO3)3·9H2O (0.016 mol) and Mg (NO3)2·6H2O (0.048 mol) in glycerol (0.051 mol) which was used as fuel in the least amount of water. The Al and Mg nitrate ratio was kept at 1:3. Surplus water was removed by heating the mixture at 80 °C in a crucible. It resulted in to a thick paste which was then placed into a preheated muffle furnace at 500 °C, leading to spontaneous combustion. The solid puffy material after spontaneous combustion was calcined at 650 °C for 3 h.24,25Table 5 gives the preparation of Zn- and Ti-loaded catalysts vis-à-vis HT.

Table 5. Preparation of Different Catalysts with Amount of Starting Materialsa.
no. catalyst-fuel catalyst abbreviation Al(NO3)3·9H2O (mol) Mg(NO3)2·6H2O (mol) titanium isopropoxide (mol) zinc nitrate (mol) fuel (mol)
1 HT–glycerol HT 0.016 0.048     0.051
2 TiMg4Al2O9 glycine Ti–HT–glycine 0.016 0.032 0.008   0.054
3 ZnMg4Al2O8 glycine Zn–HT–glycine 0.016 0.032   0.008 0.054
4 TiMg4Al2O9 glycerol Ti–HT–glycerol 0.016 0.032 0.008   0.042
5 ZnMg4Al2O8 glycerol Zn–HT–glycerol 0.016 0.032   0.008 0.042
a

Slurry thickening temperature 80 °C, combustion temperature 500 °C, calcination temperature 650 °C.

4.3. Reaction Setup and Procedure

The transesterification reaction was conducted in a standard cylindrical glass reactor of 100 mL volume with four baffles, overhead stirrer, and reflux condenser. The stirrer was a 45° pitched blade turbine impeller connected to a speed regulator. The reactor was kept in a thermostatic oil bath to maintain the desired temperature. In a typical run, 0.12 mol DMT and 0.24 mol EG were introduced in the reactor. The total volume was 38.2 cm3, No solvent was used. The reaction mixture was heated to 180 °C at 1000 rpm. The catalyst loading was 0.05 g/cm3 of total reaction volume. Sampling was done periodically. At the end, the catalyst was removed from the reaction mixture by centrifugation. High-performance liquid chromatography (HPLC) (Agilent technologies 1260 infinity) was used with C-18 mid-polar capillary column (0.25 mm ID, 30 m), and the reaction progress was monitored. Acetonitrile was used as a mobile phase at a flow rate of 1.0 mL/min. Gas chromatography–mass spectrometry (GC–MS) (PerkinElmer Clarus 500) with capillary column Elite −1 (30 m, 0.25 mm ID) was used for product confirmation. DMT (limiting reactant) conversion was determined.

4.4. Analytical Method

Sampling was done at periodic intervals by reducing the speed of agitation to zero and allowing the catalyst to settle at the bottom of the reactor. Clear liquid samples were prepared by centrifuging them for 5 min. The sample (20 μL) was diluted in a 10 mL standard volumetric flask using the mobile phase. Analysis of the samples was performed over HPLC (Agilent Technology 1260 infinity; autosampler); Hiplex-H column (300 × 7.7 mm, particle size 8 μm; 55 °C column oven temperature, UV–Vis detector at 210 nm, RID at 55 °C cell temperature). A mobile phase of acetonitrile: water (1:1 v/v with 0.1% orthophosphoric acid) solution was used at a flow rate of 1 cm3/min. Fifteen microliter injector volumes were used in the autosampler. Products were confirmed by using GC–MS. The rates and conversions were based on the disappearance of DMT. The conversion of DMT and selectivity to the main product BHET were calculated by HPLC analysis is as follows

4.4.

where A0 and Ai are the area of DMT at time t = 0 and t = i, respectively.

The selectivity was calculated as

4.4.

where Ad is the area of the desired BHET product and Ai is the total area of all the products formed in the reaction.

4.5. Catalyst Characterization

All five catalysts were characterized by EDXS (JOEL JSM 6308LA analytical scanning microscope) using 10 kV voltage at a counting rate 519 cps and energy range of 0–20 keV. SEM images were procured at different amplifications. The textural analysis of catalysts was achieved by nitrogen adsorption using Micromeritics ASAP 2020 instrument. 10% v/v CO2 in He and 10% v/v NH3 in He temperature programmed desorption (TPD) (micromeritics Autochem II 2920) using TCD detector were used to determine basic and acidic site densities of the catalyst, respectively. A Bruker AXS diffractometer D8 advance Cu Kα radiation (λ = 1.540562) was used to obtain the powder X-ray diffraction (XRD) pattern of catalysts.

Acknowledgments

A. L. Jadhav thanks the management of D.Y. Patil College of Engineering and Technology, Kolhapur, for permitting him to do this doctoral work as a teacher fellow. Thanks are also due to Dr. Godfree Fernandes for his help. Thanks to the University Grants Commission (UGC), New Delhi, for an award of SRF to R. S. Malkar under its BSR program. G.D.Y. acknowledges the support from R.T. Mody Distinguished Professor Endowment, Tata Chemicals Darbari Seth Distinguished Professor of Leadership and Innovation, and J.C. Bose National Fellowship of Department of Science and Technology, GOI.

Glossary

Nomenclature

A

reactant species A, dimethyl terephthalate

B

reactant species B, ethylene glycol

CA

concentration of A, dimethyl terephthalate (mol/L)

CA0

initial concentration of A in bulk liquid phase (mol/L)

CAS1

adsorption concentration of A on active site S1 (mol/g-cat)

CB

concentration of B ethylene glycol (mol/L)

CB0

initial concentration of B in bulk liquid phase (mol/L)

CBS2

concentration of B on active site S2 (mol/g-cat)

CE

concentration of E, product species (mol/L)

CES1

concentration of E on active sites of type S1 (mol/g-cat)

CFS2

concentration of F on active sites of type S2 (mol/g-cat)

CS1

concentration of vacant sites of type S1 (mol/g-cat)

CS2

concentration of vacant sites of type S2 (mol/g-cat)

E

product species E

BHET

bis(2-hydroxyethyl)terephthalate

F

Product species F, methanol

KA

adsorption equilibrium constant for A (g-cat/mol)

KB

adsorption equilibrium constant for B (g-cat/mol)

KE

adsorption equilibrium constant for E (g-cat/mol)

KF

adsorption equilibrium constant for F (g-cat/mol)

M

molar ratio of CB0/CA0

rA

rate of reaction (mol L–1 s–1)

t

time (s)

w

catalyst loading (g/cm3)

XA

fractional conversion of A

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.9b02230.

  • Surface area measurement by N2 adsorption and pore size analysis and TPD and XRD of Zn–HT–glycine (PDF)

This study was funded by the University Grants Commission (UGC), New Delhi.

The authors declare no competing financial interest.

Supplementary Material

ao9b02230_si_001.pdf (827.3KB, pdf)

References

  1. Anastas P. T.; Warner J. C.. Green Chemistry: Theory and Practice; Oxford University Press, 1998. [Google Scholar]
  2. Sheldon R. A.; Arends I. W. C. E.; Henefeld U.. Green Chemistry and Catalysis; Wiley-VCH Verlag, 2007. [Google Scholar]
  3. Cybulski A.; Sharma M. M.; Sheldon R. A.; Moulijn J. A.. Fine Chemicals Manufacture: Technology and Engineering; Elsevier Science: Amsterdam, 2001. [Google Scholar]
  4. Hideshi H.; Basic catalysts and fine chemicals. Studies in Surface Science And Catalysis; Elsevier, 1993; Vol. 78, pp 35–49. [Google Scholar]
  5. Hattori H. Heterogeneous basic catalysis. Chem. Rev. 1995, 95, 537–558. 10.1021/cr00035a005. [DOI] [Google Scholar]
  6. Ono Y. Solid base catalysts for the synthesis of fine chemicals. J. Catal. 2003, 216, 406. 10.1016/s0021-9517(02)00120-3. [DOI] [Google Scholar]
  7. Besnoin J.-M.; Lei G. D.; Choi K. Y. Melt transesterification of dimethyl terephthalate with ethylene glycol. AIChE J. 1989, 35, 1445–1456. 10.1002/aic.690350905. [DOI] [Google Scholar]
  8. Gorzawski H.; Hoelderich W. F. Transesterification of methyl benzoate and dimethyl terephthalate with ethylene glycol over superbases. Appl. Catal., A 1999, 179, 131–137. 10.1016/s0926-860x(98)00307-x. [DOI] [Google Scholar]
  9. Yamanis J.; Adelman M. Significance of oligomerization reactions in the transesterification of dimethyl terephthalate with ethylene glycol. J. Polym. Sci., Polym. Chem. Ed. 1976, 14, 1945–1959. 10.1002/pol.1976.170140812. [DOI] [Google Scholar]
  10. Tomita K.; Ida H. Studies on the formation of poly(ethylene terephthalate): 3. catalytic activity of metal compounds in transesterification of dimethyl terephthalate with ethylene glycol. Polymer 1973, 16, 185. 10.1016/0032-3861(75)90051-8. [DOI] [Google Scholar]
  11. Visakh P. M.; Liang M.. Poly(Ethylene Terephthalate) Based Blends, Composites and Nanocomposites; Elsevier: Amsterdam, 2015. [Google Scholar]
  12. Schuchardt U.; Sercheli R.; Matheus R. Transesterification of vegetable oils: a review. J. Braz. Chem. Soc. 1998, 9, 199–210. 10.1590/s0103-50531998000300002. [DOI] [Google Scholar]
  13. Ganesan D.; Rajendran A.; Thangavelu V. An overview on the recent advances in the transesterification of vegetable oils for biodiesel production using chemical and biocatalysts. Rev. Environ. Sci. Bio/Technol. 2009, 8, 367–394. 10.1007/s11157-009-9176-9. [DOI] [Google Scholar]
  14. Ravindranath K.; Mashelkar R. A. Polyethylene terephthalate—I. Chemistry, thermodynamics and transport properties. Chem. Eng. Sci. 1986, 41, 2197–2214. 10.1016/0009-2509(86)85070-9. [DOI] [Google Scholar]
  15. Di Serio M.; Tesser R.; Trulli F.; Santacesaria E. Kinetic and catalytic aspects in melt transesterification of dimethyl terephthalate with ethylene glycol in the presence of different catalytic systems. J. Appl. Polym. Sci. 1996, 62, 409–415. . [DOI] [Google Scholar]
  16. Mihail R.; Istratoiv R.; Lupu Al.; Gerogesur E. Transesterification of dimethyl terephthalate with ethylene glycol. Acad. Rep. Populare Romine, Studii Cercet M Chim. 1958, 6, 161. [Google Scholar]; Chem. Abstract 1959, 6145.
  17. Sorokin M. F.; Chebotareva N. A. Transesterification of the dimethyl esters of terephthalic acid by ethylene glycol. Tr. Mosk. Khim-Tekhnol. Inst. 1969, 61, 103. [Google Scholar]; Chem. Abstract 1959, 6145.
  18. Datye K. V.; Raje H. M. Kinetics of transesterification of dimethyl terephthalate with ethylene glycol. J. Appl. Polym. Sci. 1985, 30, 205. 10.1002/app.1985.070300117. [DOI] [Google Scholar]
  19. Serio M. D.; Tesser R.; Santacesaria E. Heterogeneous basic catalysts for the transesterification and the polycondensation reactions in PET production from DMT. J. Mol. Catal. A: Chem. 2004, 212, 251–257. 10.1016/j.molcata.2003.10.032. [DOI] [Google Scholar]
  20. Wiyantoko B.; Kurniawati P.; Purbaningtias T. E.; Fatimah I. Synthesis and characterization of hydrotalcite at different Mg/Al molar ratios. Procedia Chem. 2015, 17, 21–26. 10.1016/j.proche.2015.12.115. [DOI] [Google Scholar]
  21. Lopez T.; Bosch P.; Ramos E.; Gomez R.; Novaro O.; Acosta D.; Figueras F. Synthesis and characterization of sol-gel hydrotalcites. structure and texture. Langmuir 1996, 12, 189–192. 10.1021/la940703s. [DOI] [Google Scholar]
  22. Lakshmi B. B.; Patrissi C. J.; Martin C. R. Sol-gel template synthesis of semiconductor oxide micro- and nanostructures. Chem. Mater. 1997, 9, 2544–2550. 10.1021/cm970268y. [DOI] [Google Scholar]
  23. Xu Z. P.; Zeng H. C. Decomposition pathways of hydrotalcite-like compounds Mg1-x Alx(OH)2(NO3)x·nH2O as a continuous function of nitrate anions. Chem. Mater. 2001, 13, 4564–4572. 10.1021/cm010347g. [DOI] [Google Scholar]
  24. Rao C. N. R.Chemical Approaches to the Synthesis of Inorganic Materials; Wiley Eastern Limited: New Delhi, 1994. [Google Scholar]
  25. Patil K.; Aruna S. T.; Ekambaram S. Combustion synthesis. Curr. Opin. Solid State Mater. Sci. 1997, 2, 158. 10.1016/s1359-0286(97)80060-5. [DOI] [Google Scholar]
  26. Velu S.; Sabde D. P.; Shah N.; Sivasanker S. New hydrotalcite-like anionic clays containing Zr4+ in the layers: Synthesis and physicochemical properties. Chem. Mater. 1998, 10, 3451–3458. 10.1021/cm980185x. [DOI] [Google Scholar]
  27. Das N.; Samal A. Synthesis, characterisation and rehydration behaviour of titanium (IV) containing hydrotalcite like compounds. Microporous Mesoporous Mater. 2004, 72, 219–225. 10.1016/j.micromeso.2004.04.004. [DOI] [Google Scholar]
  28. Wierzbicki D.; Dębek R.; Szczurowski J.; Basąg S.; Włodarczyk M.; Motak M.; Baran R. Copper, cobalt and manganese: Modified hydrotalcite materials as catalysts for the selective catalytic reduction of NO with ammonia. The influence of manganese concentration. C. R. Chim. 2015, 18, 1074–1083. 10.1016/j.crci.2015.06.009. [DOI] [Google Scholar]
  29. Jadhav A. L.; Yadav G. D. Clean synthesis of benzylidenemalononitrile by Knoevenagel condensation of benzaldehyde and malononitrile: Effect of combustion fuel on activity and selelctivity of Ti-hydrotalcite and Zn-hydrotalcite catalyst. J. Chem. Sci. 2019, 131, 1–14. 10.1007/s12039-019-1641-6. [DOI] [Google Scholar]
  30. Jadhav A. L.; Yadav G. D. Green synthesis of 2,3-oxybutyl malononitrile via Michael reaction of methyl vinyl ketone with malononitrile over titania and zinc loaded hydrotalcite catalysts. Catal. Green Chem. Eng. 2019, 2, 43–54. 10.1615/catalgreenchemeng.2019030025. [DOI] [Google Scholar]
  31. Jadhav A. L.; Yadav G. D.. A Green process for selective hydrolysis of cinnamaldehyde in water to natural benzaldehyde by using Ti and Zn modified hydrotalcites as catalysts. Curr. Green Chem. 2019, 6242. 10.2174/2213346106666191021105244 [DOI] [Google Scholar]
  32. Fernandes G. P.; Yadav G. D. Selective glycerolysis of urea to glycerol carbonate using combustion synthesized magnesium oxide as catalyst. Catal. Today 2018, 309, 153–160. 10.1016/j.cattod.2017.08.021. [DOI] [Google Scholar]
  33. Deshpande K.; Mukasyan A.; Varma A. Direct synthesis of iron oxide nanopowders by the combustion approach: Reaction mechanism and properties. Chem. Mater. 2004, 16, 4896–4904. 10.1021/cm040061m. [DOI] [Google Scholar]
  34. Reid R. C.; Prausnitz J. M.; Sherwood T. K.. The properties of gases and liquids, 3rd ed.; McGraw-Hill: New York, 1977. [Google Scholar]
  35. Yadav G. D.; Nair J. J. Isomerization of citronellal to isopulegol using eclectically engineered sulfated zirconia– carbon molecular sieve composite catalysts, UDCaT-2. Langmuir 2000, 16, 4072–4079. 10.1021/la9911178. [DOI] [Google Scholar]
  36. Fogler H. S.Elements of Chemical Reaction Engineering, 2nd ed.; Prentice Hall: New Delhi, 1995. [Google Scholar]
  37. Yadav G. D.; Asthana N. S. Selective decomposition of cumene hydroperoxide into phenol and acetone by a novel cesium substituted heteropolyacid on clay. Appl. Catal., A 2003, 244, 341–357. 10.1016/s0926-860x(02)00605-1. [DOI] [Google Scholar]
  38. Malkar R. S.; Yadav G. D. Synthesis of cinnamyl benzoate over novel heteropoly acid encapsulated ZIF-8. Appl. Catal., A 2018, 560, 54–65. 10.1016/j.apcata.2018.04.038. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao9b02230_si_001.pdf (827.3KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES