Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconia-supported zinc oxide

Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconia-supported zinc oxide

ARTICLE IN PRESS JMRTEC 960 1–9 j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx Available online at www.sciencedirect.com www.jmrt.co...

2MB Sizes 0 Downloads 24 Views

ARTICLE IN PRESS

JMRTEC 960 1–9

j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

Available online at www.sciencedirect.com

www.jmrt.com.br

Original Article

1

Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconia-supported zinc oxide

2

3

4

5

Q2 Q1

Shady M. EL-Dafrawy ∗ , Shawky M. Hassan, Mervat Farag Chemistry Department, Faculty of Science, Mansoura University, Egypt

6 7

a r t i c l e

8

i n f o

a b s t r a c t

9 10

Article history:

An investigation of the kinetics of the Pechmann condensation reaction over ZnO supported

11

Received 22 March 2019

on sulphated zirconia (ZnO/SZ) catalyst has been performed. The ZnO/SZ catalyst contains

12

Accepted 23 September 2019

both Lewis (L) and Brønsted (B) acid sites. The maximum B/L ratio was obtained with a ZnO

13

Available online xxx

content of 6%. The reaction is first-order half in both acetoacetate and resorcinol, which supports the proposed reaction mechanism. The reaction temperature was found to have a

14

great effect on the kinetics of coumarin synthesis, and the activation energy was determined

15

Keywords:

16

Zinc oxide/sulphated zirconia

to be 42.29 kJ mol−1 . The experimental activation parameters for the reaction were calculated

17

Ethyl acetoacetate

to be H =/ = 39.20 kJ mol−1 , S =/ = −152.62 J mol−1 K−1 and G =/ = 99.18 kJ mol−1 at 393 K. The

18

Resorcinol

negative value of the activation entropy (S =/ ) indicates that the entropy decreases upon

19

Brønsted acid sites

forming the transition state, which proves that the transition state is more highly ordered

20

Pechmann reaction

21

Kinetics and activation parameters

than the reactants. © 2019 Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1.

Introduction

Solid-acid catalysts, especially metal oxide catalysts, are of great significance in a wide variety of catalytic processes important in the chemical industry [1]. Their favourable activities and selectivities have been attributed mostly to the presence of suitable amounts of Lewis and Brønsted acid sites. Sulphated metal oxides (SMOS ), including ZrO2 , SnO2 , TiO2 and Fe2 O3 , possess both Brønsted acid sites (BASS ) derived from the sulphates present on the surface of the catalyst and Lewis acid sites (LASS ) derived from the metal oxides.

22 23 24 25 26 27 28 29 30



Q2

Corresponding author. E-mail: [email protected] (S.M. EL-Dafrawy).

SMOS have been widely used as solid-acid catalysts in various organic transformations because of their high efficiency, nontoxicity, non-corrosiveness, low cost, high stability at elevated temperatures, and resistance to deactivation [2–5]. They can also be easily separated from reaction mixtures by simple filtration and reused after activation [6–8]. In the case of modified sulphated zirconia catalysts, doping with active metal species can also enhance their activity by improving crystallinity, acidity and particle size. Modified sulphated zirconia has been used to effectively catalyse condensation reactions for organic materials such as coumarins [9]. Coumarin, a benzoic acid derivative, is a natural compound commonly found in natural species, especially in green plants and some fungi [10,11]. Coumarin and its derivatives have been extensively used as laser dyes [12,13], fluores-

https://doi.org/10.1016/j.jmrt.2019.09.063 2238-7854/© 2019 Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/ licenses/by-nc-nd/4.0/). Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

31 32 33 34 35 36 37 38 39 40 41 42 43 44 45

JMRTEC 960 1–9

2

46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94

95

96 97 98 99 100

ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

cent indicators [14,15] and photosensitizers [16,17]. Recently, coumarins have attracted great interest in synthetic organic chemistry due to their noticeable biological activities such as antibacterial [18,19], anticancer [20–22], antitumour [23], anti-Alzheimer [24], anti-HIV [25], antimicrobial [26,27], antiinflammatory [28,29], and antioxidant activities [30–32]. The Perkin reaction, developed in 1863, was the first method for synthesizing coumarin derivatives [33]. In the 1990s, the Knoevenagel reaction provided a remarkable synthetic method for preparing coumarin derivatives with a carboxylic acid substituent at the 3-position [34,35]. Among all the procedures for coumarin synthesis, the Pechmann condensation is thought to be the most widely utilized technique as it proceeds with very simple starting reactants and produces several coumarin derivatives in good yields [36–38]. The Pechmann condensation to form 7-hydroxy-4-methyl coumarin from resorcinol and ethyl acetoacetate (EAA) was suggested to follow a three-step mechanism: transesterification, intramolecular hydroxyl alkylation, and dehydration, as shown in Scheme 1 [39]. The mechanism depends on proton transfer to the keto group of EAA from the acid sites of the catalyst by the interaction between EAA and the catalyst, and then an intermediate and ethanol are produced due to nucleophilic attack by a hydroxyl group of resorcinol. The intermediate rapidly cyclizes through an intramolecular condensation to yield 7-hydroxy-4-methyl coumarin [40]. The reaction mechanism has been studied via several theoretical methodologies. Among those, density functional theory (DFT) with a hybrid functional has been found to be effective in addition to offering sufficient precision at a suitable cost. The B3LYP functional was utilized to investigate the vibrational and chemical shifts of different coumarin derivatives [41–43]. The M05-2X functional was used to study the Pechmann reaction between esters and phenol [44]. PCM-TDDFT was used to examine the acid dissociation constants (pKa) of excited coumarin [45]. Recently, the M06-2X functional was efficiently used to determine the optical spectra of coumarins [46]. In this study, the catalytic performance of ZnO/SZ for the Pechmann condensation reaction was examined. The crystallinity and acid properties of the ZnO/SZ catalyst were investigated utilizing X-ray powder diffractometry (XRD), energy dispersive X-ray (EDX) and pyridine-FTIR techniques. The rates of the reaction at different reaction temperatures ranging from 353 K to 393 K were experimentally calculated and then fitted into the Arrhenius equation to measure the apparent activation energy. Additionally, different thermodynamic parameters were also determined through transition state theory.

2.

Experimental

2.1.

Materials

Zirconium nitrate (Zr(NO3 )4 ), zinc nitrate (Zn(NO3 )2 ·6H2 O, 98%), distilled water, sulphuric acid [AnalaR], resorcinol (C6 H6 O2 , 99%), and ethyl acetoacetate (C6 H10 O3 , 99%) were purchased from Sigma-Aldrich. Chemicals were used without further purification.

2.2.

Preparation of sulphated zirconia (SZ)

101

Sulphated zirconia, the SZ catalyst, was prepared as follows: 17.5 g of zirconium nitrate was dissolved in 400 mL of distilled water with stirring. Sulphuric acid (2 N) was added dropwise to the zirconium solution with vigorous and continuous stirring at room temperature until a white gel had formed. The obtained white gel was dried on a rotary evaporator at 90 ◦ C. Finally, the dried powder was calcined at different calcination temperatures (400, 500, 600, and 800 ◦ C) for 3 h with a heating rate of 20 ◦ C min−1 [47].

2.3. Preparation of zinc oxide-doped sulphated zirconia (ZnO/SZ)

102 103 104 105 106 107 108 109 110

111 112

Zinc-doped sulphated zirconia, denoted ZnO/SZ, was synthesized by a simple impregnation method [47]. SZ was impregnated in a zinc nitrate solution. The amounts of zinc nitrate were calculated to give 4, 6, 8, 10, and 20 wt% ZnO. The appropriate amounts of zinc nitrate were dissolved in 50 mL of distilled water and then added dropwise to the above white gel under continuous stirring. The mixture was stirred magnetically for 2 h and then dried on a rotary evaporator at 90 ◦ C. The synthesized samples were calcined at 400, 500, 600 and 800 ◦ C for 3 h at a heating rate of 20 ◦ C min−1 .

113 114 115 116 117 118 119 120 121 122

2.4.

Characterization

123

2.4.1.

XRD analysis

124

The powder X-ray diffraction (PXRD) patterns were recorded on an X-ray powder diffractometer (XRD) PW 150 (Philips) using Ni-filtered Cu K␣ radiation (␭ = 1.540 Å) at 40 kV and 30 mA over a scanning range 2 of 18◦ –80◦ . The tetragonal phase of SZ was evaluated by the following equation [48]. Q3



Tetragonal% =

Im(2=28.16+Im(2=31.44 2

128 129

130

D = K/ˇcos ␪ Where k is the crystallite shape constant (≈1), ␭ is the radiation wavelength (Å), ␤ is the line breadth (radians) and  is the Bragg angle.

EDX analysis

Energy dispersive X-ray (EDX) spectroscopy was applied to determine the chemical composition of the catalysts. EDX analyses were conducted using an energy dispersive X-ray spectrometer connected to a scanning electron microscope [47].

2.4.3.

127

)

The crystallite size (nm) was determined by the reflection of the SZ tetragonal phase at 2 h of 30.15 by the Scherrer equation [49,50]

2.4.2.

126



It (2 = 30.15) It(2 = 30.15) + (

125

FT-IR spectrometry with adsorbed pyridine

The Lewis and Brønsted acid sites on the catalyst surface were measured at room temperature by FT-IR analysis with adsorbed pyridine [51]. Before pyridine adsorption, each catalyst was activated at 120 ◦ C for approximately 1 h. Dry pyridine was then flashed inside the vacuum system, and the cata-

Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

131 132 133

134

135 136 137

138 139 140 141 142 143

144 145 146 147 148 149

JMRTEC 960 1–9

ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

3

Scheme 1 – Mechanism of the synthesis of 7-hydroxy-4-methyl coumarin.

150 151 152 153 154 155

156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181

182

183 184 185

lysts were maintained under these conditions for a sufficient amount of time. Then, the excess pyridine was removed by evaporation. The FT-IR spectra of the samples were recorded using a Mattson 5000 FT-IR spectrophotometer at room temperature by combining 0.05 g of the catalyst with 0.1 g of KBr to afford a self-supporting disc 30 mm in diameter [52].

2.4.4. Kinetics investigation of the Pechmann condensation reaction The investigated reaction was carried out under reflux conditions. A mixture of resorcinol (which is considered a highly activated phenol and allows the reaction to be performed under much milder conditions) and ethyl acetoacetate was refluxed in the presence of freshly calcined catalyst; the reflux time and the amount of catalyst were varied to optimize the yield of coumarin [53,54]. Freshly calcined ZnO/SZ catalyst (0.1 g) was added, the mixture was refluxed for 2 h, and then the contents of the flask was poured onto crushed ice and scratched until the solid product was obtained [52,55]. The solid product was separated by filtration, dried and then extracted with hot ethanol. 7-Hydroxy-4-methyl coumarin was recovered from the alcoholic solution by evaporating the solvent in a hot water bath. The yield was determined by FT-IR (KBr, ␯, cm−1 ): 3504 (O H), 3110 ( C H), 1672 (C O), 1600–1400 (C C), 1392 ( CH3 ) and 840–790 (tri-substituted alkene), which are characteristic of 7-hydroxy4-methyl coumarin. The order of the reaction was investigated by varying the reactant concentration in a systematic way. The activation energy can be experimentally calculated from the relation between the rate constant and different temperatures (353–393 K), which can be explained by the Arrhenius equation as given in Eq. (1): K = Ae−Ea/RT

(1)

where k is the rate constant, A is the Arrhenius constant, Ea is the activation energy, R is the gas constant, and T is the temperature.

Activation enthalpy, entropy and free energy can be determined from the Eyring-Polanyi equation, which depends on transition state theory, as shown in Eq. (2): K=

KB T −S =/ /R -H =/ /RT e .e h

(2)

where k is the rate constant, H =/ is activation enthalpy (kJ mol−1 ), S =/ is activation entropy (J mol−1 K−1 ), kB is the Boltzmann constant (J K−1 ), T is the temperature in kelvin (K), h is Plank’s constant (J S), and R is the universal gas constant (J/mol K).

3.

Results and discussion

3.1.

XRD analysis

The XRD patterns indicate the impact of various ZnO concentrations and calcination temperatures on the crystal phase and crystallite size of SZ, as revealed in Figs. 1 and 2. The peaks at 2 = 28.16◦ and 31.44◦ are characteristic of the monoclinic phase, while the peak at 2 = 30.15◦ refers to the tetragonal phase, which is the more catalytically active phase [56]. Fig. 1 shows that undoped SZ contains a major monoclinic phase in addition to the tetragonal phase. The tetragonal phase of SZ with 6% ZnO is more stable as a result of decreasing lattice spacing and zirconia particle size [57,58] in addition to the surface energy effects, which stabilize the tetragonal phase as the particle size decreases [59]. The small particles are stabilized in the tetragonal phase because their surface energy is much lower than that of the monoclinic phase, which slows transformation from the metastable phase to the monoclinic phase. Therefore, the stability of the tetragonal phase might be related to the introduction of phase stabilizers on the surface (such as the sulphate groups) or in the bulk (such as the cationic dopant) [60]. Here, tetragonal phase stabilization was caused by the introduction of 6% ZnO into the SZ. However, when the ZnO concentration was increased to more than 6%, the content of the tetragonal phase decreased while that of the monoclinic phase increased. This behaviour can be

Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

186 187 188

189

190 191 192 193 194

195

196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218

JMRTEC 960 1–9

4

ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

Table 1 – Chemical analysis of SZ catalysts with different ZnO contents. Catalysts

Wt.% of ZnO calculated from EDX

SZ 4%ZnO/SZ 6%ZnO/SZ 10%ZnO/SZ

— 4.00% 5.95% 8.82%

They observed that 0.5% Al-doped ZnO affords a minimum crystallite size of 103 nm. However, the crystallite size in the doped catalysts did not vary in any regular pattern with the Al dopant concentration [62].

3.2.

Fig. 1 – XRD of (a) SZ, (b) 6%ZnO/SZ and (c) 10%ZnO/SZ catalysts calcined at 600 ◦ C.

Fig. 3 reveals the chemical composition of the SZ catalysts doped with different concentrations of ZnO. The results corroborated that Zr, O, S and Zn are present in the catalyst structure. Furthermore, they show that the weight percent of ZnO in the nanocatalysts is in good agreement with the calculated weight percent from the synthesis of the catalysts, as observed in Table 1.

3.3.

Fig. 2 – XRD of 6%ZnO/SZ catalyst calcined at (a) 500 ◦ C, (b) 600 ◦ C and (c) 800 ◦ C.

219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236

attributed to the large concentration of ZnO, which increases randomness, disfavouring oxygen diffusion and destabilising the tetragonal phase [61]. Fig. 2 shows the 6% Zn/SZ catalysts calcined at 500, 600 and 800 ◦ C. As revealed in Fig. 2, the catalyst is mainly amorphous at 500 ◦ C, while the highest proportion of tetragonal phase is observed at 600 ◦ C. Calcination at 800 ◦ C causes a sharp decrease in the tetragonal phase accompanied by an increase in the particle size. The crystallite size calculations indicate that the crystallite size of undoped SZ decreases with the addition of 6% ZnO. The ionic radius of the Zn2+ ion (0.74 Å) is smaller than that of the Zr4+ ion (0.84 Å). Therefore, replacing a larger ion Zr4+ with a smaller one Zn2+ decreases the lattice constants and changes the crystallite size. Thus, it can be concluded that the lattice disorder that occurs at ZnO concentrations above 6% may be due to the difference in the ionic radii of Zn2+ and Zr4+ ions. Tewari et al. reached a similar conclusion for Al-doped ZnO.

EDX analysis

Adsorbed pyridine-FT-IR spectroscopy

Analysing the IR spectra with pyridine adsorbed as a base onto the surface of solid acids is a widely applied technique for investigating the Brønsted and Lewis acidity of catalysts. The pyridine molecules adsorbed on Brønsted and Lewis acid sites have different vibrational bands because of the difference in their bonding [52]. Figs. 4 and 5 show the IR spectra if the investigated catalysts with adsorbed pyridine over the 2000–500 cm−1 region. The ZnO/SZ catalysts show bands typical of adsorbed pyridine at approximately 1450 cm−1 , assigned to pyridine coordinated to electron-acceptor sites (i.e., Lewis acid sites), and 1609 cm−1 and 1626 cm−1 , assigned to pyridinium ions (pyH+ ) adsorbed on protonic acid sites (i.e., Brønsted acid sites). The band at approximately 1489 cm−1 arises from the interaction of pyridine species with both Brønsted and Lewis acid sites [63,64]. The SZ catalyst exhibits a typical band attributable to Brønsted acid sites. The percentages of Lewis and Brønsted acid sites were examined based on the integrated areas or the relative intensities of the characteristic bands of pyridine adsorbed on Lewis and/or Brønsted acid sites. The percentages of Lewis and Brønsted acid sites can be calculated using the following equations [65]:

237 238 239 240

241

242 243 244 245 246 247 248

249

250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271

Lewis% = (AL /AL +AB )×100

272

Brønsted% = (AL /AL +AB )×100

273

where AB is the intensity of the peak located in the range of 1609–1626 cm−1 , which is attributed to the pyridinium cation adsorbed on Brønsted acid sites, and AL is the intensity of the peak at 1450 cm−1 from pyridine coordinated to Lewis acid sites. The sample with 6%ZnO/SZ calcined at 600 ◦ C showed the optimum Brønsted to Lewis acid site (B/L) ratio, as shown in Fig. 6. Further increasing the ZnO loading decreased the B/L

Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

274 275 276 277 278 279 280 281

JMRTEC 960 1–9

ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

5

Fig. 3 – EDX pattern of (a) SZ, (b) 4%ZnO/SZ, (c) 6%ZnO/SZ, and (d) 10%ZnO/SZ catalysts calcined at 600 ◦ C.

Fig. 5 – FTIR spectra of pyridine adsorbed on 6%ZnO/SZ catalyst calcined at (a) 500 ◦ C, (b) 600 ◦ C and (c) 800 ◦ C. Fig. 4 – FTIR spectra of pyridine adsorbed on (a) SZ, (b) 6%ZnO/SZ, (c) 8%ZnO/SZ, (d) 10%ZnO/SZ and (e) 20%ZnO/SZ catalysts.

ratio, which may be attributable to the aggregation of ZnO decreasing the accessibility of Lewis and Brønsted acid sites to pyridine. Karim et al. found similar results with WO3 loaded on ZrO2 [66].

Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

282 283 284 285

JMRTEC 960 1–9

6

ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

Table 2 – Systematic variation of concentration of resorcinol and ethyl acetoacetate. Experiment Resorcinol [A] (mol) Ethyl acetoacetate [B] (mol) Product (mol)

1

2

3

0.01 0.01 4.63 × 10−3

0.01 0.02 6.31 × 10−3

0.02 0.01 3.26 × 10−3

and S OH species may serve as Brønsted acid sites according to the following scheme:

296

(" ∗ "referstoeitherSorZr)

297

3.4.

Kinetic study

3.4.1.

298 299 300

301

The Pechmann reaction has been intensely studied over different catalysts; however, the order, rate constants, and activation energy were not experimentally determined as discussed in this research. Based on density functional theory with the M06-2X functional, Pornsatitworakul et al. studied the preparation of 7-hydroxy-4-methyl coumarin. The authors claimed that theoretically the transesterification was the ratedetermining step, and it had an overall activation energy of 40.0 kJ mol−1 . They suggested the theoretical pathway of 7hydroxy-4-methyl coumarin synthesis as shown in Fig. 8; [AD (activated complex), IN (intermediate), TS (transition state), PR (product)] [39].

Fig. 7 – Effect of calcination temperature on B/L ratio of 6%ZnO/SZ catalyst.

295

*—O—H + pyr → *—O− + + H—pyr

However, S OH is a stronger Brønsted acid site. Therefore, losses of S OH decrease the number of Brønsted acid sites [64,65]. Fig. 6 – Effect of ZnO content on the B/L ratio of ZnO/SZ catalyst calcined at 600 ◦ C.

294

Order of the reaction

302 303 304 305 306 307 308 309 310 311 312 313

314

For a chemical reaction, the order can be determined using the initial rate method or by using a program for data analysis. However, in reactions involving multiple reactants, the order of the reaction can be determined by varying the concentration of the reactants in a systematic way, as observed in Table 2, so that the order can be measured with respect to concentration change of one of the reactants by using the following rate equation:

315 316 317 318 319 320 321 322

ˇ

k1 [A]˛1 [B]1 rate 1 = ˇ rate 2 k2 [B]˛2 [B]2

Scheme 2 – Surface structure of SZ with both Brønsted and Lewis acid sites proposed by Ward et al. [68].

286 287 288 289 290 291 292 293

The effect of the calcination temperature on the B/L ratio for the 6%ZnO/SZ catalyst is shown in Fig. 7. The B/L ratio increased with increasing calcination temperature until it reached a maximum at 600 ◦ C and then it decreased. This may be due to decomposition of the sulphate species and the loss of S O, which has a covalent bond (Scheme 2) and acts as electron-withdrawing species, and then decreases the number of Lewis acid sites [67,68]. Moreover, it is reported that Zr OH

(3)

Where 1 and 2 refer to the experiment number, k is the rate constant, A is the concentration of resorcinol, B is the concentration of ethyl acetoacetate, the exponent ␣ is the order of the reaction with respect to A and the exponent ␤ is the order of the reaction with respect to B. Then, the total reaction order can be given by the sum of exponents ␣ and ␤. Table 2 shows the systematic variations in the concentrations of resorcinol and ethyl acetoacetate, and the results showed first-order reaction kinetics.

3.4.2.

Apparent activation energy

The reaction was performed over temperatures ranging from 353 to 393 K as observed in Fig. 9. The increase in the conversion of reactants by increasing the reaction temperature

Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

323

324 325 326 327 328 329 330 331 332

333 334 335 336

JMRTEC 960 1–9

ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

7

Fig. 8 – Reaction energy pathway of Pechmann condensation reaction for 7-hydroxy-4-methyl coumarin proposed by Pornsatitworakul et al. [39].

3.4.3.

Activation enthalpy, entropy and free energy

349

The Eyring-Polanyi equation is used in chemical kinetics to describe variations in the rate of a chemical reaction with temperature. This equation is based on transition state theory. The linear form of the Eyring-Polanyi equation is given as follows: ln

−H =/ 1 S =/ k kB = . + ln + T R T h R

(5)

where k is the rate constant, T is the temperature in kelvin (K), H =/ is the activation enthalpy (kJ mol−1 ), S =/ is the activation entropy (J mol−1 K−1 ), kB is the Boltzmann constant (J K−1 ), h is Plank’s constant (J. S), and R is the universal gas constant (J/mol K). The activation enthalpy and the activation entropy were calculated from the slope and intercept of the straight line of Fig. 10. The values of S =/ and H =/ were used to calculate the free energy of activation using the following equation: G =/ = H =/ -TS =/ (6)

Fig. 9 – Relation between ln k and 1/T of the synthesis of Pechmann condensation reaction with ZnO/SZ catalyst.

337 338 339 340 341

indicates that the investigated reaction is kinetically controlled [69]. The rate constants were calculated at different temperatures and plotted as a function of 1/T as shown in Fig. 9 according to the Arrhenius equation:

Where G =/ is the free energy of activation (kJ mol−1 ). The results summarized in Table 3 show that the transesterification step had apparent activation energy of 42.29 kJ mol−1 , making this step endothermic by 39.20 kJ mol−1 . The negative value of S =/ reveals that the entropy decreases upon forming the transition state, which indicates that the transition state is more highly ordered than the reactants.

4. 342

343 344 345 346 347 348

lnK = lnA–Ea /R.1/T

(4)

Where k is the rate constant, Ea is the activation energy, R is the gas constant, and A is the preexponential constant. The value of Ea was calculated to be 42.29 kJ mol−1 . This value is comparable to the theoretical value calculated by Pornsatitworakul et al. (40.0 kJ mol−1 ). Thus, transesterification is the rate-determining step.

350 351 352 353 354

355

356 357 358 359 360 361 362 363 364

365

366 367 368 369 370 371 372

Conclusion

ZnO/SZ nanoparticles have been synthesized by a simple impregnation method. The results showed that SZ has a major monoclinic phase in addition to a tetragonal phase. The incorporation of ZnO increases the stability of the tetragonal phase as a result of decreasing the lattice spacing and particle size of zirconia. Increasing the ZnO concentration to more than 6 wt% led to decreasing the stability by decreasing the proportion of tetragonal phase with a concomitant increase in the mono-

Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

373 374 375 376 377 378 379 380

ARTICLE IN PRESS

JMRTEC 960 1–9

8

j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

Table 3 – Activation energy, activation enthalpy, entropy and free energy values for the investigated pechmann condensation reaction. Ea (kJ mol−1 )

42.29

H =/ (kJ mol−1 )

S =/ (J mol−1 K−1 )

39.20

−152.62

Fig. 10 – Relation between ln (k/T) and 1/T of the synthesis of Pechmann condensation reaction with ZnO/SZ catalyst.

381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400

clinic phase. The catalyst is mainly amorphous at 500 ◦ C, while the greatest proportion of tetragonal phase was observed at 600 ◦ C. Calcination at 800 ◦ C caused a sharp decrease in the tetragonal phase accompanied by an increase in the particle size. The crystal size of ZnO/SZ nanoparticles decreased with increasing ZnO content. The chemical composition confirmed that Zr, O, S and Zn are present in the catalyst structure. The addition of ZnO improved the acidic properties of the ZnO/SZ nanoparticles. The SZ and ZnO/SZ nanoparticles were applied in the catalytic synthesis of 7-hydroxy-4-methyl coumarin. The kinetic parameters of the investigated Pechmann condensation were studied experimentally. The apparent activation energy (42.29 kJ mol−1 ) was calculated using the Arrhenius equation. The kinetics of the Pechmann condensation show that the transesterification step had an apparent activation energy of 42.29 kJ mol−1 , making this step endothermic by 39.20 kJ mol−1 . The negative value of S =/ indicates that the entropy decreases upon forming the transition state, which indicates that the transition state is more highly ordered than the reactants.

Conflict of interest 401 Q4

The authors declare no conflicts of interest.

402

references 403

404 Q5

G =/ (kJ mol−1 ) 353 K

373 K

383 K

93.08

96.13

97.66

393 K 99.18

[2] Wu Y, Liao S. Front Chem Eng Chin 2009;3:330–43. [3] Banerjee B, Bhunia S, Bhaumik A. J Appl Catal A: Gen 2015;502:380–7. [4] Suzuki T, Yokoi T, Otomo R, Kondo JN, Tatsumi T. J Appl Catal A: Gen 2011;408:117–24. [5] Ivanov VK, Baranchikov AY, Kopitsa GP, Lermontov SA, Yurkova LL, Gubanova NN, et al. J Solid State Chem 2013;198:496–505. [6] Feng Q, Wen S, Deng J, Zhao W. J Appl Surf Sci 2017;425: 8–15. [7] Feng Q, Zha W, Wen S, Cao Q. J Separ Purif Techn 2017;178:193–9. [8] Feng Q, Zhao W, Wen S. J Appl Surf Sci 2018;436:823–31. [9] El-Dafrawy ShM, Farag M, Abd El Hakam S, Hassan ShM. Egy J Chem 2017;60:329–45. [10] Liu M, Chen A, Wang Y, Wang C, Wang B, Sun D. J Food Chem Nanotechnol 2015;168:270–5. [11] Konc JT, Hejchman E, Maciejewska D, Wolska I. J Mol Struct 2011;998:42–8. [12] Keskin SS, Aslan N, Bayrakceken F. J Spectrochim Acta Part A 2009;72:254–9. [13] Bakhtiari G, Moradi S, Soltanali S. J Arab J Chem 2014;7:972–5. [14] Ma J, Sheng R, Wu J, Liu W, Zhang H. J Sens Actuat B 2014;197:364–9. [15] Kaya EN, Yuksel F, Ozpınar GA, Bulut M, Durmus M. J Sens Actuat B 2014;194:377–88. [16] Camur M, Durmus M, Bulut M. J Polyhed 2012;41:92–103. [17] Kandavelu V, Huang H-S, Jian J-L, Yang TC-K, Wang K-L, Huang S-T. J Solar Ener 2009;83:574–81. [18] Wang S, Yin Y, Wu X, Qiao F, Sha S, Lv P, et al. J Bioorg Med Chem 2014;22:5727–37. [19] Kong Y, Fu Y, Zu Y, Chang F, Chen Y, Liu X, et al. J Food Chem Nanotechnol 2010;121:1150–5. [20] Thakur A, Singla R, Jaitak V. Eur J Med Chem 2015;101:476–95. [21] Rajabi M, Hossaini Z, Khalilzadeh MA, Datta S, Halder M, Mousa SA. J Photochem Photobiol B Biol 2015;148:66–72. [22] Li H, Wang X, Xu G, Zeng L, Cheng K, Gao P, et al. J Bioorg Med Chem Lett 2014;24:5274–8. [23] Zhang W, Li Z, Zhou M, Wu F, Hou X, Luo H, et al. J Bioorg Med Chem Lett 2014;24:799–807. [24] Huang M, Xie S, Jiang N, Lan J, Kong L, Wang X. J Bioorg Med Chem Lett 2015;25:508–13. [25] Matsuda K, Hattori S, Kariya R, Komizu Y, Kudo E, Goto H, et al. J Biochem Biophys Res Comm 2015;457:288–94. [26] Abd El-Wahab H, Abd El-Fattah M, Abd El-Khalik N, Nassar HS, Abdelall MM. J Prog Org Coat 2014;77:1506–11. [27] Sahoo J, Mekap SK, Kumar PS. J Taibah Univ Sci 2015;9:187–95. [28] Azelmat J, Fiorito S, Taddeo VA, Genovese S, Epifano F, Grenier D. J Phytochem Lett 2015;13:399–405. [29] Arora RK, Kaur N, Bansaln Y, Bansal G. J Acta Pharmac Sinica B 2014;4:368–75. [30] Roussaki M, Zelianaios K, Kavetsou E, Hamilakis S, Hadjipavlou-Litina D, Kontogiorgis C, et al. J Bioorg Med Chem 2014;22:6586–94. [31] Witaicenis A, Seito LN, Chagas AdS, Junior LDdA, Luchini AC, R.-Orsi P, et al. J Phytomed 2014;21:240–6. [32] Arsenyan P, Vasiljeva J, Shestakova I, Domracheva I, Jaschenko E, Romanchikova N, et al. J C R Chim 2015;18:399–409.

[1] Atghia SV, Beigbaghlou SS. J C R Chim 2014;17:1155–9. Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463

JMRTEC 960 1–9

ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx

464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492

[33] Augustine JK, Bombrun A, Ramappa B, Boodappa C. J Tetrah Lett 2012;53:4422–5. [34] Khoshkholgh MJ, Lotfi M, Balalaie S, Rominger F. J Tetrah 2009;65:4228–34. [35] Voskressensky LG, Festa AA, Varlamov AV. J Tetrah 2014;70:551–72. [36] Rahmatpour A, Mohammadian S. J C R Chim 2013;16:271–8. [37] Sinhamahapatra A, Sutradhar N, Pahari S, Bajaj HC, Panda AB. J Appl Catal A: Gen 2011;394:93–100. [38] Karami B, Kiani M. J Catal Comm 2011;14:62–7. [39] Pornsatitworakul S, Boekfa B, Maihom Th, Treesukol P, Namuangruk S, Jarussophon S, et al. J Monatsh Chem 2017;148:1245–50. [40] Heravi MM, Khaghaninejad S, Mostofi M. J Adv Hetero Chem 2014;112:1–50. [41] Singh AK, Singh RK. J Mol Struct 2016;1122:318–23. [42] Kurashige Y, Nakajima T, Kurashige S, Hirao K, Nishikitani Y. J Phys Chem A 2007;111:5544–8. [43] Zhao G-J, Han K-L. J Acc Chem Res 2012;45:404–13. [44] Daru J, Stirling A. J Org Chem 2011;76:8749–55. [45] Houari Y, Jacquemin D, Laurent AD. J Chem Phys Lett 2013;583:218–21. [46] Budzak S, Charaf-Eddin A, Medved M, Gryko DT, Jacquemin D. J Comput Theor Chem 2016;1076:57–64. [47] El-Dafrawy ShM, Farag M, Hassan ShM. J Res Chem Intermed 2017;43:6343–65. [48] Devassy BM, Lefebvre F, Bohringes W, Fletcher J, Halligudi SB. J Mol Catal A Chem 2005;236:162–7. [49] Patterson A. Can J Appl Physiol 1939;56:978–82. [50] Jenkins R, Snyder RL. Int J Chem Anal Sci 1996;138:335–9.

9

[51] Matsuhashi H, Motoi H, Arata K. J Catal Lett 1994;26:325–8. [52] EL-Hakam SA, Hassan ShM, Ahmed AI, ELDafrawy ShM. J Am Sci 2011;7:682–93. [53] Sethna S, Phadke SM. J Org Reac 1953;7:1–58. [54] Shockravib A, Valizadeha H. J Tetrahed Lett 2005;46:3501–3. [55] Ahmed AI, El-Hakam SA, Khder AS, Abo El-Yazeed WS. J Mol Catal A Chem 2013;366:99–108. [56] Sang X, Zhang L, Wang H, He D, Deng L, Huang S, et al. J Powd Technol 2014;253:590–5. [57] Rekha M, Manjunath HR, Nagaraju N. J Ind Eng Chem 2013;19:337–46. [58] Kato K, Saito T, Shibayama Sh, Sakashita M, Takeuchi W, Taoka N, et al. J Thin Solid Films 2014;557:192–6. [59] Kumar S, Bhunia S, Ojha AK. J Phys E 2015;66:74–80. [60] Gao S, Chen X, Wang H, Mo J, Wu Z, Liu Y, et al. J Coll Interf Sci 2013;394:515–21. [61] Rui R, Duan W, Xiaodong W, Jun F, Lei W, Xiaodi W. J. Rare Earths 2011;29:1053–9. [62] Tewari S, Bhattacharjee A. J Phys 2011;76:153–63. [63] Sani YM, Alaba PA, Raji-Yahyab AO, AbdulAziz AR, Daud WMAW. J Taiwan Inst Chem Eng 2016;60:247–57. [64] Stevens RW Jr, Chuang SSC, Davis BH. J Thermochim Acta 2003;407:61–71. [65] Stevens RW Jr, Chuang SSC, Davis BH. J Appl Catal A: Gen 2003;252:57–74. [66] Karima AH, Triwahyono S, Abdul Jalil A, Hattori H. J Appl Catal A: Gen 2012;433–434:49–57. [67] Arata K, Hino M. J Appl Catal 1990;59:197–204. [68] Ward DA, Ko EI. J Catal 1994;150:18–33. [69] Yadav GD, Ajgaonkar NP, Varma A. J Catal 2012;292:99–110.

Please cite this article in press as: EL-Dafrawy SM, et al. Kinetics and mechanism of Pechmann condensation reaction over sulphated zirconiaJMRTEC 960 1–9 supported zinc oxide. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.09.063

493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522