Hostname: page-component-788cddb947-rnj55 Total loading time: 0 Render date: 2024-10-14T06:22:31.226Z Has data issue: false hasContentIssue false

Effects of the Inclusion of Ce and Ni Species on Ti for Modification of K10-Clay by Sol-Gel and their Use as Catalysts in the Liquid-Phase Esterification Systems

Published online by Cambridge University Press:  01 January 2024

G. Rangel-Porras*
Affiliation:
Department of Chemistry, Division of Nature and Exact Sciences, University of Guanajuato, Noria Alta s/n, 36050 Guanajuato, Guanajuato, Mexico
A. Quiroga-Almaguer
Affiliation:
Department of Chemistry, Division of Nature and Exact Sciences, University of Guanajuato, Noria Alta s/n, 36050 Guanajuato, Guanajuato, Mexico
A. Ramírez-Hernández
Affiliation:
Institute of Applied Chemistry, University of Papaloapan, Circuíto Central No. 200, Parque Industrial, Tuxtepec, 68301 Oaxaca, Mexico
B. Bachiller-Baeza
Affiliation:
Group of Heterogeneous Catalysts for Selective Chemical Processes, Institute of Catalysis and Petrochemistry, C/Marie Curie 2, Cantoblanco, 28049, Madrid, Spain
H. Pfeiffer-Perea
Affiliation:
Department of Low Dimensional Materials, Institute of Materials Research, National Autonomous University of Mexico, Circuíto Exterior s/n, Ciudad Universitaria, Coyoacán, 04510 Mexico City, Mexico
P. Rangel-Rivera*
Affiliation:
Department of Chemistry, Division of Nature and Exact Sciences, University of Guanajuato, Noria Alta s/n, 36050 Guanajuato, Guanajuato, Mexico
Rights & Permissions [Opens in a new window]

Abstract

The modification of montmorillonite with metallic species affects directly its crystalline structure, texture, porosity, and surface. The interaction of the metallic molecules with the clay matrix, derived from the modification pathway and the characteristics of the adsorbate, modifies the physicochemical properties of montmorillonite, enabling the creation of materials with varied characteristics to be used both as catalysts and adsorbents. Small amounts of metallic species can confer various structural and physicochemical characteristics on the same montmorillonite matrix, depending on the metal incorporated. The objective of the present study was to create an acid-base catalyst based on montmorillonite K10 (K10 Mnt), modified with Ti, Ce, and Ni, for the catalytic esterification of acetic acid and penta-1-ol. K10-Mnt was modified using particles of Ti and of Ti modified with Ce and Ni. The effect of the inclusion of Ti and modified Ti species on the transformation of the physicochemical properties of the K10 Mnt and their contributions to the catalytic esterification syntheses were investigated. Samples were characterized by scanning electron microscopy coupled to an energy-dispersive X-ray spectroscopy system (SEM-EDS), powder X-ray diffraction (XRD), Fourier-transform infrared spectroscopy (FTIR), physisorption of N2 at 77 K (BET and BJH), and thermogravimetric analysis (TGA-DTGA). Finally, the original and modified K10 Mnt samples were tested for their catalytic esterification of acetic acid and penta-1-ol in the liquid phase.

Type
Original Paper
Copyright
Copyright © The Author(s), under exclusive licence to The Clay Minerals Society 2023

Introduction

Esterification is an important process in the chemical industry as esters are used in perfumes, cosmetics, flavorings, additives, emulsifiers, plasticizers, polymer precursors, intermediates, and agents in fine chemicals and drugs (Hamerski et al., Reference Hamerski, Dusi, Fernandes Dos Santos, da Silva, Pedersen Voll and Corazza2020; Khan et al., Reference Khan, Javed, Shamair, Hafeez, Fazal, Aslam, Zimmerman and Rehman2021). This process is performed typically with the reaction of carboxylic acid and alcohol, homogeneously catalyzed by Brønsted acid centers in strong mineral acids, but this has several operational and environmental drawbacks deriving from the toxic and corrosive environment of acids, as well as the difficulty of separating the products from the homogeneous catalysis system.

Studies, therefore, have been undertaken (Saravanan et al., Reference Saravanan, Tyagi and Bajaj2016; Vijayakumar et al., Reference Vijayakumar, Mahadevaiah, Nagendrappa and Prakash2012) to change to a non-toxic, non-corrosive, heterogeneous catalytic process in which the catalyst can be recovered easily from the reaction media in an environmentally friendly manner. For this purpose, a wide variety of solid acid catalysts have been used, e.g. graphene (Roy et al., Reference Roy, Poulose, Bakandritsos, Varma and Otyepka2021), modified metal oxides (Nsir et al., Reference Nsir, Younes, Rives and Ghorbel2017), zeolites (Alismaeel et al., Reference Alismaeel, Abbas, Albayati and Doyle2018), membranes (Zhang et al., Reference Zhang, Li, He, Song, Ji, Cui, Li and Younas2020), and clay minerals (Tekale & Yadav, Reference Tekale and Yadav2021). The clay minerals, such as Mnt, have been used as catalytic or adsorbent materials either unmodified or as supports for other reactive phases or chemically active centers (Amaya et al., Reference Amaya, Moreno and Molina2021; Barakan & Aghazadeh, Reference Barakan and Aghazadeh2021; Chmielarz et al., Reference Chmielarz, Kowalczyk, Skoczek, Rutkowska, Gil, Natkański, Radko, Motak, Dębek and Ryczkowski2018; Kashif et al., Reference Kashif, Halepoto, Memon, Su, Abduallah and Soomro2020; Khalil et al., Reference Khalil, Chaabene, Boujday, Blanchard and Bergaoui2015; Zhang et al., Reference Zhang, Wang, Zhao, Bai, Wen, Kang, Song, Song and Komarneni2021a) because they have a unique combination of physicochemical and structural characteristics (Bahmanpour et al., Reference Bahmanpour, Héroguel, Baranowski, Luterbacher and Kröcher2018; Parisi et al., Reference Parisi, Lazzara, Merli, Milioto, Princivalle and Sciascia2019; Wen et al., Reference Wen, Zhu, Chen, Ma, Liu, Zhu, Xi and He2019).

Acidic sites are located at the terminal O-H groups at the edges of the 2:1 Mnt layers and involve Si–O(H+), Al–O(H+), Fe–O(H+), or Mg–O(H+). Structural hydroxyls at the interfacial plane between the tetrahedral and octahedral sheets may also act as Brønsted acid sites (Bailey et al., Reference Bailey, Lekkerkerker and Maitland2015). The water present in the interlayer can be polarized and display significant potential for donating protons. Likewise, the undersaturated Al3+ at the edges of the octahedral sheets can act as electron acceptor or Lewis acid sites (Huang et al., Reference Huang, Liu, She, Zhong, Christidis and Zhou2021). This coexistence of two types of acid sites makes Mnt an ideal material for use as a heterogeneous catalyst in esterification and other acid-catalyzed processes. The surface and/or the matrix of Mnt may also be modified to enhance its capability as a catalyst. The fact that charged species can be removed from the interlamellar space means that space is a dynamic system that can be exploited to increase the surface area, change the porosity, and provide new active centers for adsorption and/or catalysis (Wang et al., Reference Wang, Zheng, Zhao, Huang, Zhang, Cao, Dai, Hua and Liu2018). A wide variety of inorganic species has been used for the modification/pillaring of Mnt, such as Cr (Georgescu et al., Reference Georgescu, Nardou, Zichil and Nistor2018), Al (Cardona et al., Reference Cardona, Korili and Gil2021), Ti (Butman et al., Reference Butman, Gushchin, Ovchinnikov, Gusev, Zinenko, Karamysheva and Krämer2020), Fe (Khankhasaeva & Badmaeva, Reference Khankhasaeva and Badmaeva2020), Zr (Rathinam et al., Reference Rathinam, Atchudan and Edison2021), La (Silva et al., Reference Silva, Cabral, Araujo, Caldeira, Coriolano, Fernandes, Pergher and Araujo2021), Ce (Lai et al., Reference Lai, Yan, Wang, Qu, Shen and Zhang2021), Ni (Asgari et al., Reference Asgari, Vitale and Sundararaj2021), Cu (Yang et al., Reference Yang, Spyrou, Thomou, Kumar, Cao, Stuart, Pei, Gournis and Rudolf2020), and others. Ti species incorporated on the surface or within the interlayer space of the Mnt can lead to greater interlayer spacing, significant thermal and hydrothermal stability, larger pore sizes, larger specific surface area, and greater acidity than found with other modified/pillared materials (Jin et al., Reference Jin, Chen, Liu, Liu, Qian, Wei and Zheng2019; Romero et al., Reference Romero, Dorado, Ascencio, García and Valverde2006; Zhang et al., Reference Zhang, Zhang, Liu, Wu, Zhou, Zhang, Zheng, Han, Xie and Liu2021b). In the case of Ti-modified Mnt, a variety of other metals can be added to enhance further the catalytic performance. Of particular interest in the present study was the incorporation of Ce and Ni species, which promotes oxidation-reduction catalysis, photocatalysis, and surface acidity (Natsir et al., Reference Natsir, Putri, Wibowo, Maulidiyah, Salim, Azis, Bijang, Mustapa, Irwan, Arham and Nurdin2021; Zhong et al., Reference Zhong, Cai, Yu and Zhong2015). These species are particularly useful in modulating the relative contributions of Lewis and Brønsted acid centers leading to weak acid behavior that sometimes benefits the various catalytic processes (Bernardon et al., Reference Bernardon, Osman, Laugel, Louis and Pale2020; Zhang et al., Reference Zhang, Qin, Ji, Chu, Gao, Zhang and Song2017).

The present study, therefore, aimed to determine the catalytic capacity of K10 Mnt modified with Ti, combined with the addition of Ce and Ni, toward the esterification of acetic acid and penta-1-ol, and to relate the physicochemical properties to the catalytic performance.

Materials and Methods

Ti/K10-Based Catalyst Synthesis

The clay mineral used was the commercially available montmorillonite K10 (Fluka Inc., Munich, Germany), with a chemical composition of SiO2 (73%), Al2O3 (14%), Fe2O3 (2.6%), CaO (0.3%), MgO (1.1%), Na2O (0.6%), and K2O (1.9%). The precursor of the Ti source was titanium(IV) isopropoxide (Sigma-Aldrich Inc., St. Louis, Missouri, USA). Cerium(IV) dioxide (CeO2, Merck Inc., Darmstadt, Germany) and nickel(II) chloride (NiCl2, Analit Inc., Mexico City, Mexico) were used as sources of Ce and Ni, respectively. The K10 was activated initially by an acid treatment: K10 was dispersed using 6 M hydrochloric acid, maintaining a relationship of 8 mg of K10 per 1 mL of hydrochloric solution. After that, the mixture was stirred constantly and vigorously for 24 h at 25°C, recovering the solid by centrifugation (1.88 xg), and washing with deionized water at least three times. Finally, the solid was dried at 70°C for 24 h.

Titanium and titanium combined with Ce and Ni species were obtained using a process similar to sol-gel: The Ti-based particle suspensions were prepared by adding titanium(IV) isopropoxide dropwise to an aqueous solution of 0.1 M HCl and stirring vigorously until a gel was formed that contained metallic species in the form of titanium oxide and titanium oxide with Ce and Ni species that can modify the K10 using the procedure described below.

Once the activated K10 was totally dried, it was modified with the metallic species: Each suspension of Ti, Ce/Ti, and Ni/Ti was sonicated for 1 h. At the same time, K10-activated powder was suspended in water and stirred for 1 h under room conditions. Based on the chemical formula for K10, the Si content was 34 wt.%. Enough of the Ti alkoxide was added to the K10 to obtain a Ti/Si weight ratio of 1.02. In the synthesis of Ti modified by Ce and Ni, the amounts of CeO2 and NiCl2 required were deposited in the alkoxide/HCl suspension to obtain weight ratios for Ce/Ti and Ni/Ti of 0.02 in both cases. Subsequently, each suspension of Ti, Ce/Ti, and Ni/Ti was mixed with the K10-activated suspension with vigorous stirring for 3 h. The resulting materials were separated by centrifugation (1.88 xg) and washed with deionized water at least three times and dried at 70°C for 24 h. Finally, the K10 modified with metallic species were heated at 200°C for 4 h. Four samples were obtained and studied: K10-activated (K10), K10 modified by the inclusion of Ti species (Ti/K10), K10 modified with Ti species and subsequently with Ce (Ce-Ti/K10), and K10 modified with Ti species and subsequently with Ni (Ni-Ti/K10).

SEM-EDS

The SEM images were collected using a JEOL JSM 6010/LA analytical scanning electron microscope coupled to a Quantax EDS detector device (Bruker Inc., Billerica, Massachusetts, USA). The sample was mounted on an electrically insulating material and then inserted into the high vacuum system before performing the SEM-EDS analysis.

Powder X-ray Diffraction

An Inel-Equinox diffractometer (Inel Inc., Artenay, France) operated at 25°C, equipped with CuKα radiation (1.54 Å), was used. Samples were prepared as random powders by hand grinding using a mortar and pestle until a homogeneous sample was obtained. The samples were packed from behind into metal holders with the help of a glass slide, and the entire uncovered surface was exposed to the incident X-rays. The XRD patterns were obtained over the range 2–80°2θ at 5°/min scanning velocity, 30 kV voltage, and at a current of 30 mA. Bragg´s equation (nλ = 2dsinθ) was used to calculate the interlamellar spaces or d spacings.

FTIR Spectroscopy

The FTIR spectra were collected using a Bruker FTIR Spectrometer Tensor 27 series instrument (Bruker Inc., Markham, Canada). The samples were heated at 80°C for 1 h prior to spectrum acquisition. Using the KBr pellet method, with typical grinding of a mixture of solid and salt in a ratio of 1/100 by weight and pressing, the spectra were acquired in the 4000–400 cm–1 range.

Physisorption of N2 at 77 K

A Micromeritics ASAP-2010 instrument (Micromeritics Inc., Norcross, Georgia, USA) was used for physisorption of N2. The BET equation and BJH (Barrett-Joyner-Halenda) method were used to calculate the BET surface area and pore-size distribution, respectively. A vacuum pressure of 500 μm Hg at 200°C was applied to all samples, with the aim of eliminating residual N2 gas molecules from their surfaces and pores.

TGA-DTA

A Thermal Analyzer SDT-2960 (TA Instruments Inc., New Castle, Delaware, USA) under an airflow with a heating rate of 10°C/min was used.

Esterification of Acetic Acid and Penta-1-ol

A stirred-batch reflux system equipped with a Dean-Stark trap to collect water was used to carry out the catalytic process. A ratio of 1:1 v/v of penta-1-ol and toluene was placed in a 50 mL round-bottomed flask. After the mixture stabilized on reaching 90°C, a molar ratio of acetic acid containing penta-1-ol of 0.2:1 and 0.1 g/10 mL of the reaction mixture of each catalyst were added to the system under vigorous and constant stirring. The conversion was quantified by an acid-base volumetric titration technique using 0.1 M normal KOH and phenolphthalein as an indicator (Li et al., Reference Li, Yan, Li, Yu and Ge2020; Rangel-Rivera et al., Reference Rangel-Rivera, Bachiller-Baeza, Galindo-Esquivel and Rangel-Porras2018).

Results

SEM-EDS

The SEM images (Fig. 1) showed the particle morphologies of each sample. The sample labeled K10 has a flaky morphology. For K10 modified with Ti species (Ti/K10), compact aggregates were seen without significant alteration with respect to the structure of the original clay. In EDS semiquantitative analysis of elements, for K10 modified by Ti with species of Ce and Ni, these metallic materials apparently were dispersed homogeneously on the surface of the clay. Semi-quantitative analysis (Table 1) showed the variation in the relative percentages of elements depending on the metallic species used as a modifier.

Fig. 1 SEM images of the clay samples

Table 1 Semi-quantitative analysis (wt.%) of chemical elements by EDS

N.D. Not detected by EDS

Powder XRD

The XRD patterns contained peaks characteristic of montmorillonite (Fig. 2). For K10, the XRD pattern showed a wide diffraction band between 3 and 8°2θ with a maximum centered at ~5.2°2θ, corresponding to a d 001 for K10 of 1.68 nm. When K10 was modified with Ti species, the formation of crystalline phases typical of Ti was not observed. No anatase (25.3, 38, 47.6, and 54.8°2θ) or rutile signals (21, 27, 32, and 43°2θ) were noted. Nevertheless, the range 2–14°2θ (Fig. 3) showed a broad peak, associated with a basal or interlamellar space of d 001 of (4–8°2θ, 2.21–1.10 nm). A broad peak between 3.5 and 7°2θ (2.52 and 1.26 nm, respectively) was observed for each clay sample. The wide peak was centered at ~5.5°2θ (1.61 nm).

Fig. 2 Powder XRD of the clay samples. A = anatase (TiO2), R = rutile (TiO2)

Fig. 3 Powder XRD of the clay samples in the range 2–14°2θ

FTIR Spectroscopy

The FTIR spectra showed bands typical of montmorillonite (Fig. 4). Peaks between 3700 and 3000 cm–1 were associated with hydroxyl group vibrations and the band at 3620 cm–1 is characteristic of smectites with a large quantity of Al species occupying the octahedral sheets. Likewise, the Si–O stretching vibration modes induce a strong absorption band in the spectral range 1100–1000 cm–1 (Fig. 5). On the other hand, the bands at 525 and 480 cm–1 are associated with Si–O–Al of octahedral-aluminum ions and Si–O–Si of bending vibrations, respectively, which do not undergo alterations in all cases.

Fig. 4 FTIR spectra of clay samples

Fig. 5 FTIR spectra of the clay samples in the range 1100–1000 cm–1

Physisorption of N2 at 77 K

Nitrogen adsorption on K10 at 77 K (Fig. 6) showed a well defined type IV isotherm with an inflection point that started at a P/P 0 value of ~0.4; as values approached 1.0, the tendency was for the material to continue adsorbing, as indicated by an asymptotic graph at that value, suggesting that larger mesopores may accompany slit-shaped pores. The same characteristics were observed in the isotherms of Ti/K10-based samples, and a type H2 (IUPAC) hysteresis-loop was observed in all samples, with some differences in the surface characteristics (Table 2). The BJH adsorption and desorption distribution curves for all samples are shown in Fig. 7; a well-defined peak was observed at 34 Å while a second, less well defined one occurred at 48 Å for K10. The same peaks were present in the curves of Ti/K10-based samples. Note that a second peak was very wide and poorly defined for Ce-Ti/K10 and Ni-Ti/K10, but was better defined in Ti/K10 centered at ~45 Å, due to the morphology of the pores, as well as their dimensions being different both at their entrance and within the matrix.

Fig. 6 N2 adsorption-desorption isotherms at 77 K for the clay samples

Table 2 Surface area and porosity of clay samples

Fig. 7 BJH pore-size distribution of clay samples

TGA-DTGA

The TGA profiles (Fig. 8) for K10 showed mass losses in several defined stages as the temperature increased: 30–100°C (2.3 wt.%), 100–400°C (1.8 wt.%), and 400–800°C (2.2 wt.%), with a total mass loss of 6.3 wt.%. Mass losses from Ti/K10 for the same temperature stages were: 30–100°C (2.3 wt.%), 100–400°C (1.8 wt.%), and 400–800°C (1.9 wt.%), for a total weight loss of 6.0%. For Ce-Ti/K10, the weight losses were: 2.9 wt.% at 30–100°C, 3.1 wt.% at 100–400°C, and 2.1% at 400–800°C, resulting in a total weight loss of 8.1%. For Ni-Ti/K10, the values were: 2.4 wt.% (30–100°C), 1.7 wt.% (100–400°C), and 1.8 wt.% (400–800°C), with a total weight loss of 5.9%. The DTGA analysis (Fig. 8) for K10 showed an inflection point at 434°C, while for Ti/K10 the maxima were centered at 425 and 526°C. For Ce-Ti/K10, the corresponding values were 425 and 521°C, and for Ni-Ti/K10 they were 425 and 528°C; the DTGA profiles were very similar to each other in the Ti/K10 samples. All solids with Ti/K10 as the support presented similar weight losses related to endothermic peaks due to the loss of adsorbed water (100–200°C) and dehydroxylation (600–750°C) (Tan, Reference Tan1998).

Fig. 8 a TGA and b DTGA analyses of the clay samples

Esterification of Acetic Acid and Penta-1-ol

The use of clay minerals as heterogeneous catalysts in the esterification of acetic acid and penta-1-ol (Fig. 9) indicated that for K10 the first and second hours of reaction showed little activity; this increased after 3 h until reaching 80% of conversion after 7 h of reaction. During the first hour of reaction, the Ti/K10 catalyst gave 76% conversion; Ce-Ti/K10, 75.0%; and Ni-Ti/K10, 60.0%. After 3 h of reaction, the maximum conversion for both Ti/K10 and Ce-Ti/K10 was 86.0%. For Ni-Ti/K10 the greatest conversion rate, achieved after 8 h of reaction, was 83.0%. The blank experiment (no catalyst) gave ~15.0% of conversion by autocatalytic reaction and the rate of conversion was constant thoughout the 8 h of reaction.

Fig. 9 Conversion of acetic acid and penta-1-ol during the esterification reaction using the clay samples

Discussion

SEM-EDS

The morphology of the K10 sample was flaky in aggregates, a typical and common structure of clays in lamellar arrangements. When K10 was modified with Ti, a heterostructure of the clay was observed, with the appearance of small aggregations on the lamellar surface. The flaky particles of the original K10 changed to a more spongy morphology as the Ti particles were dispersed on the surface, but without creating a ‘house-of-cards’ type structure; thus the clay layers conglomerated together to create a structure similar to that created by stacking playing cards on top of each other, in the shape of a pyramid. In the Ce-Ti- and Ni-Ti-modified K10 samples, the same behavior as for Ti/K10 was observed, but a more organized structure was noted than for with Ti/K10, i.e. a loose flaky aggregate but without forming a ‘house-of-cards’ structure, in agreement with results described by Belver et al. (Reference Belver, Bedia and Rodriguez2015). The EDS spectroscopy data indicated that the incorporation of Ti species was homogeneous, as well as the detection of other elements, especially in the decrease of percentage of Al and Fe. The homogeneous distribution of the Ti, Ce-Ti, and Ni-Ti species on the surface of the K10 layers was confirmed by these data.

Powder XRD

The XRD pattern of K10 showed a well defined peak associated with the d 001 basal spacing, indicating the swelling of the lamellar structure after the HCl acid treatment, with the subsequent increase in the basal spacing into flakes dispersed heterogeneously (Rangel-Porras et al., Reference Rangel-Porras, Rangel-Rivera, Pfeiffer-Perea and Gonzalez-Muñoz2014; Rangel-Rivera et al., Reference Rangel-Rivera, Rangel-Porras, Pfeiffer-Perea and Lima-Muñoz2014). The absence of peaks assigned to Ti in the XRD patterns of the Ti/K10-based samples indicated a very small amount of Ti, or the presence of non-crystalline Ti particles, or even the absence of Ti particles; the same effect was observed during modification by Ce-Ti and Ni-Ti species. In previous reports (Rangel-Rivera et al., Reference Rangel-Rivera, Rangel-Porras, Pfeiffer-Perea and Lima-Muñoz2014) the K10-activated clay showed a broad peak centered at 4.9°2θ (1.8 nm), which is similar to the broad peak observed in the present study at ~5.5°2θ (1.61 nm), both of which indicate layer stacking disorder. These interlamellar spaces are greater than those reported in the literature for TiO2-pillared montmorillonites (Binitha & Sugunan, Reference Binitha and Sugunan2006; Liu et al., Reference Liu, Dong, Zuo and Yu2009), leading to the conclusion that the clay minerals studied were not pillared by Ti species, but were only surface modified by Ti, Ce-Ti, and Ni-Ti species, even delaminating K10 to a certain degree.

FTIR Spectroscopy

Focusing on the –OH stretching region of Ti/K10, the band with a maximum at 3430 cm–1 corresponds to water molecules fixed to the surface by hydrogen bonds, resulting in overlapping symmetric-asymmetric stretching vibrations. The peak at 3615 cm–1 was associated with the structural hydroxyl groups of K10, which suggested that the inclusion of Ni-Ti and Ce-Ti in K10 modified the chemical environment of the hydroxyl groups in the K10 structure into a more homogeneously process than Ti species alone. The shape of the Si–O band is an indicator of montmorillonite intercalation (Cole, Reference Cole2008; Maier et al., Reference Maier, Beuntner and Thienel2021); nevertheless, no variation was observed in the region between 900 and 850 cm–1, indicating a similar chemical structure of the cations between samples. A slight change in shape, more defined and elongated, was noted in the Ni-Ti/K10 sample but not in the others, indicating that the interlayer space environment may have been altered. The unmodified bands, thus, at 1090, 525, and 480 cm–1 suggested that the incorporation of inorganic species in K10 does not modify substantially either the structure or the chemical environment of the interlamellar space, but the modification affects the surface of the material. These data corroborate the information obtained by SEM and XRD, thus reaffirming the hypothesis of modification by dispersion-deposition but not by pillar or by the formation of a ‘house-of-cards’ structure.

Physisorption of N2 at 77 K

A type IV curve is common in layered materials with the presence of mesopores and explains the tendency to follow the adsorption process at relative pressures (P/P 0) >1.0, relating to pore expansion (Gregg & Sing, Reference Gregg and Sing1995). Desorption hysteresis is an indicator of the shape of the pores and their capillary adsorption, indicating that this material with mesopores and H2 hysteresis is related to solids containing aggregates of particles in the form of parallel plates, with straight, cylindrical pores, which gives rise to pores in the form of slits, although in this case it is the greater heterogeneity of the pores that made up these clays. Furthermore, the presence of micropores was possible due to low adsorption at low relative pressures (P/P 0 < 0.03). In the BJH patterns, K10 showed the typical pore distribution of a layered material, and the trend was similar for all the Ti/K10-based samples that were analyzed, which indicated that the shape of the aggregates was practically identical among them. Following Xia and collaborators, the mesopores located at 30 Å correspond to pores inside the internal layers of the clay, and the pores >40 Å are assigned to mesopores outside the internal layers (Xia et al., Reference Xia, Jian, Li, Sun, Xue and Chen2009). The BJH pattern of K10 showed an incipient peak at 44 Å, which was undetectable in Ti/K10 samples. Deposition of the Ce and Ni species did not alter the structure or geometry of the pores of the original material. Ti/K10, as well as the formation of Ti species on the K10 surface, induced the presence of mesopores exclusively in internal layers of the structure. The differences in desorption hysteresis indicated the formation of these mesopores, where the presence of micropores made only a moderate contribution to the surface area. This type of architectural arrangement at the nanoscale is typical of exfoliated specimens or specimens in which the layers are disordered, losing their characteristic interlayer space while the metallic particles adhere to the surfaces of the lamellae (Serwicka, Reference Serwicka2021). The incorporation of Ti species does not produce pillaring or delamination of the K10; likewise, modification with Ce or Ni species does not produce these effects. In the present case, the incorporation of Ni and Ce species in the modified K10 with Ti prompted a decrease in pore size and BET surface area, as some Ce-Ti and Ni-Ti particles remained on the surfaces and edges of the layers. The greater increase in BET surface area of the Ce-Ti/K10 than that of Ti/K10 is an effect of the crystal sizes of the Ti species tending to shrink in the presence of Ce species, so the BET surface area increased. On the other hand, the presence of Ni species modified the pathways of incorporation of Ti species on the K10 matrix, occupying a minor BET surface area. Finally, two pathways of incorporation are proposed here: (1) the dispersion of species of Ti, Ce-Ti, and Ni-Ti which form, exclusively, the internal mesoporous structure arrangement; and (2) non-hydrolyzed species forming agglomerates which lead to micropores through the aggregation of metallic nanoparticle-clay layers, blocking the presence of mesopores outside the internal layers. These assertions are supported by SEM, XRD, and FTIR characterizations where the morphology of the materials, the absence of pillaring or ‘house-of-cards’ type structures, and the non-delamination of K10 after the inclusion of metallic species are all noted (Fig. 10).

Fig. 10 a Schematic representation of the structure of the clay samples, b esterification process of the clay samples

TGA-DTGA

When the samples were heated, the first stage was the elimination of all physically adsorbed water; the second stage, the removal of interlayer water and initiation of dehydroxylation of the K10. The third weight-loss stage, >400°C, was due to K10 dehydroxylation but also to the carbonization of the organic compounds from the organic systems of Ti species formed during the intercalation of Ti in the clay by a sol-gel process in all samples (Bineesh et al., Reference Bineesh, Kim, Kim and Park2011), which were the result of the hydrolysis (TiOH), alcoxolation-oxolation (Ti-O-Ti), and olation (Ti-OH-Ti) processes, respectively. On the other hand, a variety of compounds such as Ti{OCH(CH3)2}3OH, Ti{OCH(CH3)2}2(OH)OH, Ti{OCH(CH3)2}2(OH)2, or Ti{OCH(CH3)2}(OH)3 are derived from the aforementioned processes and present in the sol-gel matrix [Wright & Sommerdijk, Reference Wright and Sommerdijk2001]. One study reported that the latter could include the decomposition of two types of organic compounds, one adsorbed on the surface of Ti/K10 (>400°C) and the second intercalated in the interlamellar space (>600°C), both remnants of the intercalation process of Ti species via sol-gel (Kaneko et al., Reference Kaneko, Shimotsuma, Kajikawa, Hatamachi, Kodama and Kitayama2001). During the present study, however, no evidence of intercalation within the interlamellar space was obtained and, thus, the two types apparently were not present. Nevertheless, when the Ti/K10 was modified by Ce and Ni species, some slight differences were observed in the TGA profiles of Ti/K10 and Ni-Ti/K10, but Ce-Ti/K10 showed the greatest mass loss of all Ti/K10-based samples.

Due to Ce-Ti/K10 having the largest surface area of all samples, the incorporation of Ce into Ti/K10 may have exposed the metal, which could then become hydrated and account for the slight increase in adsorbed water content observed by TGA (Fig. 5).

Esterification of Acetic Acid and Penta-1-ol

The catalysis of esterification using K10 and Ti/K10-based catalysts was carried out with the aim of determining the effect of the inclusion of modified Ce and Ni species on Ti-modified K10. The esterification process is known to be a typical condensation reaction that is acid catalyzed. Note that montmorillonite (K10 included) is a system of colloidal particles which are very stable and highly swelling, leading to its ability to exfoliate into monolayers (Nicolosi et al., Reference Nicolosi, Chhowalla, Kanatzidis, Strano and Coleman2013). The phenomenon of delamination or exfoliation has attracted research into their use in heterogeneous catalysis (Roth et al. Reference Roth, Sasaki, Wolski, Song, Tang, Ebina, Ma, Grzybek, Kałahurska, Gil, Mazur, Zapotoczny and Cejka2020; Shawky et al., Reference Shawky, El-Sheikh, Rashed, Abdo and El-Dosoqy2019; Yi et al., Reference Yi, Zhao and Song2021), and has been especially focused in organic synthesis reactions in the liquid phase. The recovery of the catalyst in these cases is relatively simple, either by deposition of the K10 colloids or by mechanical means such as centrifugation (Chellapandi & Madhumitha, Reference Chellapandi and Madhumitha2022). Once the material is washed, settled, and dried, it can be used again, giving rise to an economic and ecofriendly catalyst (Mahanta et al., Reference Mahanta, Raul, Saikia, Bora and Thakur2017). The differences in the catalytic activity are due to changes in the total acidity of the K10 and Ti/K10-based catalysts. On the other hand, the esterification reaction is promoted with the presence of a Brønsted acid catalyst, which is responsible for providing protons (H+) to the reaction medium. Protons interact with one or both reactants, starting with the catalytic process. K10 showed a lower catalytic performance in the first 3 h of the reaction, reaching its maximum after 6 h. Using the Ti/K10 and Ce-Ti/K10 catalysts offered the best performances. This behavior is related to their greater BET surface area and ability to retain water molecules. The water molecules will be able to disperse in a greater proportion, as well as be retained in the metallic centers. This arrangement results in the dissociation of protons on a larger scale. Furthermore, it highlights the behavior of Ni-Ti/K10 which reached its maximum activity 5 h later than the other two catalysts. This phenomenon may be due to the tendency of nickel species to form agglomerates in inorganic matrices (Iwanschitz et al., Reference Iwanschitz, Holzer, Mai and Schütze2011; Lee & Kim, Reference Lee and Kim2021), which was observed in the decrease of surface characteristics such as BET surface area, pore size, and pore volume. When they are agglomerated after the incorporation process, the Ni particles reduce the textural properties in order of porosity and BET surface area, as well as the number of metallic centers where the Brønsted acid sites are in the Ti/K10 catalyst. The result of this behavior is an active acid material for esterification processes, but with a delayed catalytic activity relative to the other materials used as catalysts (Fig. 9). Finally, K10 required a longer time to obtain appreciable yields because its acid centers are not as strong or as dispersed as those of the modified materials, despite having greater mesoporosity. Comparing K10 and Ti/K10-based materials with other types of heterogeneous and homogeneous catalysts (Table 3) used for the conversion of pentyl acetate, the Ti/K10-based catalysts, except for Ni-Ti/K10, offer shorter reaction times and lower temperatures than most of the catalysts presented. In addition, the use of this type of catalyst is advantageous in terms of easy separation, environmental compatibility, and practical manipulation of catalytic process parameters.

Table 3 Comparison of various catalysts in the synthesis of pentyl acetate

Conclusions

The inclusion of Ti, Ce-Ti, and Ni-Ti species on K10 montmorillonite did not affect its crystalline structure inasmuch as the crystalline phases corresponding to the modified metallic species were not present. Nevertheless, the chemical environment of the catalyst changed depending on modifiers, such that Ti/K10 and Ce-Ti/K10 had surface hydration levels greater than that for Ni-Ti/K10, which generated a greater availability of active acid centers. The incorporation of metallic species did not provoke the formation of structures described in the literature such as pillars or the “house-of-cards” structure; nonetheless, the presence of two types of mesopores in K10 was affected by the incorporation of the metallic species, creating only internal mesopores, affecting the bimodal structure in terms of mesoporosity that was perceptible in K10. Specifically, the incorporation of Ce species caused an increase in the surface area, while modification with Ni species had the opposite effect. This behavior led to a greater amount of organic matter and water in Ce-Ti/K10 so that the incorporation of Ti and Ti modified with Ce species produced a greater dispersion of the reagents and increased the availability/accessibility of active acid centers. Finally, the inclusion of Ce species produced a structure with a larger porosity that induced a larger BET surface area than its counterparts, while the Ni-modified species led to the formation of agglomerated particles that decreased the BET surface area of the original material. These characteristics were decisive in the catalytic activity of the catalysts, such that Ni-Ti/K10 showed the least conversion over the longest reaction time.

Acknowledgments

This study was supported by the DAIP-Universidad de Guanajuato office. The authors acknowledge Cristina Daniela Moncada-Sánchez of the Laboratory for Research and Characterization of Minerals and Materials of the University of Guanajuato (LICAMM-UG-Mexico) for the support in the SEM-EDS analysis. Likewise, the authors are grateful to CONACYT-Mexico for the fellowships provided to P. Rangel-Rivera and A. Quiroga-Almaguer.

Author Contributions

P. Rangel-Rivera, G. Rangel-Porras, and B. Bachiller-Baeza contributed to the experimental development, analysis of results, writing, reviewing, and final editing of the manuscript. A. Ramírez Hernández and A. Quiroga-Almaguer contributed to the analysis of results, reviewing, and final editing of the manuscript. H. Pfeiffer-Perea contributed to the experimental development and reviewing of the manuscript.

Declarations

Conflict of Interest

The authors declare no conflicts of interest with respect to this article.

Footnotes

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

References

Alismaeel, Z. T., Abbas, A. S., Albayati, T. M., Doyle, A. M. Biodiesel from batch and continuous oleic acid esterification using zeolite catalysts Fuel 2018 234 170176 10.1016/j.fuel.2018.07.025CrossRefGoogle Scholar
Amaya, J., Moreno, S., Molina, R. Heteropolyacids supported on clay minerals as bifunctional catalysts for the hydroconversion of decane Applied Catalysis B: Environmental 2021 297 120464 10.1016/j.apcatb.2021.120464CrossRefGoogle Scholar
Asgari, M., Vitale, G., Sundararaj, U. Synthesis and characterization of a novel nickel pillared–clay catalyst: In-situ carbon nanotube–clay hybrid nanofiller from Ni-PILC Applied Clay Science 2021 205 106064 10.1016/j.clay.2021.106064CrossRefGoogle Scholar
Bahmanpour, A. M., Héroguel, F., Baranowski, C. J., Luterbacher, J. S., Kröcher, O. Selective synthesis of dimethyl ether on eco-friendly K10 montmorillonite clay Applied Catalysis A: General 2018 560 165170 10.1016/j.apcata.2018.05.006CrossRefGoogle Scholar
Bailey, L., Lekkerkerker, HNW, Maitland, G. C. Smectite clay – Inorganic nanoparticle mixed suspensions: Phase behaviour and rheology Soft Matter 2015 11 222236 10.1039/C4SM01717JCrossRefGoogle ScholarPubMed
Barakan, S., Aghazadeh, V. The advantages of clay mineral modification methods for enhancing adsorption efficiency in wastewater treatment: A review Environmental Science and Pollution Research 2021 28 25722599 10.1007/s11356-020-10985-9CrossRefGoogle ScholarPubMed
Belver, C., Bedia, J., Rodriguez, J. J. Titania–clay heterostructures with solar photocatalytic applications Applied Catalysis B: Environmental 2015 176 278287 10.1016/j.apcatb.2015.04.004CrossRefGoogle Scholar
Bernardon, C., Osman, M. B., Laugel, G., Louis, B., Pale, P. Acidity versus metal-induced Lewis acidity in zeolites for Friedel-crafts acylation Comptes Rendus Chimie 2020 20 2029 10.1016/j.crci.2016.03.008CrossRefGoogle Scholar
Bineesh, K. V., Kim, D. Y., Kim, M. I., Park, D. W. Selective catalytic oxidation of H2S over V2O5 supported on TiO2-pillared clay catalysts in the presence of water and ammonia Applied Clay Science 2011 53 204211 10.1016/j.clay.2010.12.022CrossRefGoogle Scholar
Binitha, N. N., Sugunan, S. Preparation, characterization and catalytic activity of titania pillared montmorillonite clays Microporous and Mesoporous Materials 2006 93 8289 10.1016/j.micromeso.2006.02.005CrossRefGoogle Scholar
Butman, M. F., Gushchin, A. A., Ovchinnikov, N. L., Gusev, G. I., Zinenko, N. V., Karamysheva, S. P., Krämer, K. W. Synergistic effect of dielectric barrier discharge plasma and TiO2-pillared montmorillonite on the degradation of rhodamine B in an aqueous solution Catalysts 2020 10 359376 10.3390/catal10040359CrossRefGoogle Scholar
Cardona, Y., Korili, S. A., Gil, A. A nonconventional aluminum source in the production of alumina-pillared clays for the removal of organic pollutants by adsorption Chemical Engineering Journal 2021 425 130708 10.1016/j.cej.2021.130708CrossRefGoogle Scholar
Chellapandi, T., Madhumitha, G. Montmorillonite clay-based heterogenous catalyst for the synthesis of nitrogen heterocycle organic moieties: A review Molecular Diversity 2022 26 23112339 10.1007/s11030-021-10322-3CrossRefGoogle ScholarPubMed
Chmielarz, L., Kowalczyk, A., Skoczek, M., Rutkowska, M., Gil, B., Natkański, P., Radko, M., Motak, M., Dębek, R., Ryczkowski, J. Porous clay heterostructures intercalated with multicomponent pillars as catalysts for dehydration of alcohols Applied Clay Science 2018 160 116125 10.1016/j.clay.2017.12.015CrossRefGoogle Scholar
Cole, K. C. Use of infrared spectroscopy to characterize clay intercalation and exfoliation in polymer nanocomposites Macromolecules 2008 41 834843 10.1021/ma0628329CrossRefGoogle Scholar
Curini, M., Epifano, F., Marcotullio, M. C., Rosati, O., Rossi, M. Heterogeneous catalysis in acetylation of alcohols and phenols promoted by zirconium sulfophenyl phosphonate Synthetic Communications 2000 30 13191329 10.1080/00397910008087154CrossRefGoogle Scholar
Gao, X., Ding, Q., Wu, Y., Jiao, Y., Zhang, J., Li, X., Li, H. Kinetic study of esterification over structured ZSM-5-coated catalysts based on fluid flow situations in macrocellular foam materials Reaction Chemistry & Engineering 2020 5 485494 10.1039/C9RE00445ACrossRefGoogle Scholar
Georgescu, A. M., Nardou, F., Zichil, V., Nistor, I. D. Adsorption of lead(II) ions from aqueous solutions onto Cr-pillared clays Applied Clay Science 2018 152 4450 10.1016/j.clay.2017.10.031CrossRefGoogle Scholar
Gregg, S. K., Sing, K. S. Adsorption surface area and porosity 1995 2 Academic Press Inc.Google Scholar
Hamerski, F., Dusi, F. G., Fernandes Dos Santos, J. T., da Silva, V. R., Pedersen Voll, F. A., Corazza, M. L. Esterification reaction kinetics of acetic acid and n-pentanol catalyzed by sulfated zirconia International Journal of Chemical Kinetics 2020 52 499512 10.1002/kin.21365CrossRefGoogle Scholar
Hassan, HMA, Betiha, M. A., Mohamed, S. K., El-Sharkawy, E. A., Ahmed, E. A. Stable and recyclable MIL-101(Cr)–ionic liquid based hybrid nanomaterials as heterogeneous catalyst Journal of Molecular Liquids 2017 236 385394 10.1016/j.molliq.2017.04.034CrossRefGoogle Scholar
Huang, W. J., Liu, J. H., She, Q. M., Zhong, J. Q., Christidis, G. E., Zhou, C. H. Recent advances in engineering montmorillonite into catalysts and related catalysis Catalysis Reviews 2021 63 157Google Scholar
Iwanschitz, B., Holzer, L., Mai, A., Schütze, M. Nickel agglomeration in solid oxide fuel cells: The influence of temperature Solid State Ionics 2011 211 6973 10.1016/j.ssi.2012.01.015CrossRefGoogle Scholar
Jin, J., Chen, B., Liu, L., Liu, R., Qian, G., Wei, H., Zheng, J. A study on modified bitumen with metal doped nano-TiO2 pillared montmorillonite Materials 2019 12 19101923 10.3390/ma12121910CrossRefGoogle Scholar
Kaneko, T., Shimotsuma, H., Kajikawa, M., Hatamachi, T., Kodama, T., Kitayama, Y. Synthesis and photocatalytic activity of titania pillared clays Journal of Porous Materials 2001 8 295301 10.1023/A:1013165014982CrossRefGoogle Scholar
Kashif, M., Halepoto, A., Memon, A., Su, Y., Abduallah, M., Soomro, M. Y. Gallium oxide impregnated on porous clay heterostructures material for selective catalytic reduction of nitrogen oxide with C3H6 Journal of Environmental Chemical Engineering 2020 8 103943 10.1016/j.jece.2020.103943CrossRefGoogle Scholar
Khalil, T. A., Chaabene, S. B., Boujday, S., Blanchard, J., Bergaoui, L. A new method for elaborating mesoporous SiO2/montmorillonite composite materials Journal of Sol-Gel Science and Technology 2015 75 436446 10.1007/s10971-015-3716-2CrossRefGoogle Scholar
Khan, Z., Javed, F., Shamair, Z., Hafeez, A., Fazal, T., Aslam, A., Zimmerman, W. B., Rehman, F. Current developments in esterification reaction: A review on process and parameters Journal of Industrial and Engineering Chemistry 2021 103 80101 10.1016/j.jiec.2021.07.018CrossRefGoogle Scholar
Khankhasaeva, S. T., Badmaeva, S. V. Removal of p-aminobenzene sulfanilamide from water solutions by catalytic photo-oxidation over Fe-pillared clay Water Research 2020 185 116212 10.1016/j.watres.2020.116212CrossRefGoogle Scholar
Lai, F., Yan, F., Wang, P., Qu, F., Shen, X., Zhang, Z. Efficient one-pot synthesis of ethyl levulinate from carbohydrates catalysed by Wells-Dawson heteropolyacid supported on Ce–Si pillared montmorillonite Journal of Cleaner Production 2021 324 129276 10.1016/j.jclepro.2021.129276CrossRefGoogle Scholar
Lee, J. R., Kim, Y. H. Agglomeration of nickel oxide particle during hydrogen reduction at high temperature in a fluidized bed reactor Chemical Engineering Research and Design 2021 168 193201 10.1016/j.cherd.2021.02.005CrossRefGoogle Scholar
Li, L., Yan, B., Li, H., Yu, S., Ge, X. Decreasing the acid value of pyrolysis oil via esterification using ZrO2/SBA-15 as a solid acid catalyst Renewable Energy 2020 146 643650 10.1016/j.renene.2019.07.015CrossRefGoogle Scholar
Liu, J., Dong, M., Zuo, S., Yu, Y. Solvothermal preparation of TiO2/montmorillonite and photocatalytic activity Applied Clay Science 2009 43 156159 10.1016/j.clay.2008.07.016CrossRefGoogle Scholar
Mahanta, A., Raul, P. K., Saikia, S., Bora, U., Thakur, A. J. Methanol aided synthesis of PdNPs decorated on montmorillonite K 10 and its implication in Suzuki Miyaura type cross coupling reaction under base free condition Applied Organometallic Chemistry 2017 32 e4192 10.1002/aoc.4192CrossRefGoogle Scholar
Maier, M., Beuntner, N., Thienel, K. C. Mineralogical characterization and reactivity test of common clays suitable as supplementary cementitious material Applied Clay Science 2021 202 105990 10.1016/j.clay.2021.105990CrossRefGoogle Scholar
Mirkhani, V., Tangestaninejad, S., Moghadam, M., Yadollahi, B., Alipanah, L. Cerium polyoxometalate as a reusable catalyst for acetylation and formylation of alcohols Chemical Monthly 2004 135 12571263 10.1007/s00706-004-0196-4CrossRefGoogle Scholar
Mirza-Aghayan, M., Boukherroub, R., Rahimifard, M. Graphite oxide as an efficient solid reagent for esterification reactions Turkish Journal of Chemistry 2014 38 859864 10.3906/kim-1401-81CrossRefGoogle Scholar
Natsir, M., Putri, Y. I., Wibowo, D., Maulidiyah, M., Salim, LOA, Azis, T., Bijang, C. M., Mustapa, F., Irwan, I., Arham, Z., Nurdin, M. Effects of Ni–TiO2 pillared clay–montmorillonite composites for photocatalytic enhancement against reactive orange under visible light Journal of Inorganic and Organometallic Polymers and Materials 2021 31 33783388 10.1007/s10904-021-01980-9CrossRefGoogle Scholar
Nicolosi, V., Chhowalla, M., Kanatzidis, M. G., Strano, M. S., Coleman, J. N. Liquid exfoliation of layered materials Science 2013 340 6139 12264191226419 10.1126/science.1226419CrossRefGoogle Scholar
Nsir, S. B., Younes, M. K., Rives, A., Ghorbel, A. Characterization and reactivity of zirconia-doped phosphate ion catalyst prepared by sol–gel route and mechanistic study of acetic acid esterification by ethanol Journal of Sol-Gel Science and Technology 2017 84 349360 10.1007/s10971-017-4509-6CrossRefGoogle Scholar
Parisi, F., Lazzara, G., Merli, M., Milioto, S., Princivalle, F., Sciascia, L. Simultaneous removal and recovery of metal ions and dyes from wastewater through montmorillonite clay mineral Nanomaterials 2019 9 16991714 10.3390/nano9121699CrossRefGoogle ScholarPubMed
Pushpaletha, P., Lalithambika, M. Modified attapulgite: An efficient solid acid catalyst for acetylation of alcohols using acetic acid Applied Clay Science 2011 51 424430 10.1016/j.clay.2010.12.033CrossRefGoogle Scholar
Rangel-Porras, G., Rangel-Rivera, P., Pfeiffer-Perea, H., Gonzalez-Muñoz, P. Changes in the characteristics of acid-treated clay after the inclusion of proteins Surface and Interface Analysis 2014 47 135141 10.1002/sia.5685CrossRefGoogle Scholar
Rangel-Rivera, P., Rangel-Porras, G., Pfeiffer-Perea, H., Lima-Muñoz, E. Thermoanalytical study of acid-treated clay containing amino acid immobilized on its surface Journal of Thermal Analysis and Calorimetry 2014 115 13591369 10.1007/s10973-013-3464-xCrossRefGoogle Scholar
Rangel-Rivera, P., Bachiller-Baeza, M. B., Galindo-Esquivel, I., Rangel-Porras, G. Inclusion of Ti and Zr species on clay surfaces and their effect on the interaction with organic molecules Applied Surface Science 2018 445 229241 10.1016/j.apsusc.2018.03.157CrossRefGoogle Scholar
Rathinam, K., Atchudan, R., Edison, TNJI Zirconium oxide intercalated sodium montmorillonite scaffold as an effective adsorbent for the elimination of phosphate and hexavalent chromium ions Journal of Environmental Chemical Engineering 2021 9 106053 10.1016/j.jece.2021.106053CrossRefGoogle Scholar
Romero, A., Dorado, F., Ascencio, I., García, P. B., Valverde, J. L. Ti-pillared clays: Synthesis and general characterization Clays and Clay Minerals 2006 54 737747 10.1346/CCMN.2006.0540608CrossRefGoogle Scholar
Roth, W. J., Sasaki, T., Wolski, K., Song, Y., Tang, D-M, Ebina, Y., Ma, R., Grzybek, J., Kałahurska, K., Gil, B., Mazur, M., Zapotoczny, S., Cejka, J. Liquid dispersions of zeolite monolayers with high catalytic activity prepared by soft-chemical exfoliation Science Advances 2020 6 16 10.1126/sciadv.aay8163CrossRefGoogle ScholarPubMed
Roy, A. S., Poulose, A. C., Bakandritsos, A., Varma, R. S., Otyepka, M. 2D graphene derivatives as heterogeneous catalysts to produce biofuels via esterification and trans-esterification reactions Applied Materials Today 2021 23 101053 10.1016/j.apmt.2021.101053CrossRefGoogle Scholar
Saravanan, K., Tyagi, B., Bajaj, H. C. Esterification of stearic acid with methanol over mesoporous ordered sulfated ZrO2–SiO2 mixed oxide aerogel catalyst Journal of Porous Materials 2016 23 937946 10.1007/s10934-016-0151-xCrossRefGoogle Scholar
Serwicka, E. W. Titania-clay mineral composites for environmental catalysis and photocatalysis Catalysts 2021 11 10871159 10.3390/catal11091087CrossRefGoogle Scholar
Shawky, A., El-Sheikh, S. M., Rashed, M. N., Abdo, S. M., El-Dosoqy, T. I. Exfoliated kaolinite nanolayers as an alternative photocatalyst with superb activity Journal of Environmental Chemical Engineering 2019 7 103174 10.1016/j.jece.2019.103174CrossRefGoogle Scholar
Silva, J. B., Cabral, G. G., Araujo, MDS, Caldeira, VPS, Coriolano, ACF, Fernandes, V. J. Jr, Pergher, SBC, Araujo, A. S. Catalytic pyrolysis of atmospheric residue of petroleum using pillared interlayed clay containing lanthanum and aluminum polyhydroxications (LaAl13-PILC) Petroleum Science and Technology 2021 39 704717Google Scholar
Tan, K. H. Principles of soil chemistry 1998 3 CRC PressGoogle Scholar
Tekale, D. P., Yadav, G. D. Esterification of propanoic acid with 1,2-propanediol: Catalysis by cesium exchanged heteropoly acid on K-10 clay and kinetic modelling Reaction Chemistry & Engineering 2021 6 313320 10.1039/D0RE00337ACrossRefGoogle Scholar
Vijayakumar, B., Mahadevaiah, N., Nagendrappa, G., Prakash, BSJ Esterification of stearic acid with p-cresol over modified Indian bentonite clay catalysts Journal of Porous Materials 2012 19 201210 10.1007/s10934-011-9461-1CrossRefGoogle Scholar
Wang, Y., Zheng, Z., Zhao, Y., Huang, J., Zhang, Z., Cao, X., Dai, Y., Hua, R., Liu, Y. Adsorption of U(VI) on montmorillonite pillared with hydroxyaluminum Journal of Radioanalytical and Nuclear Chemistry 2018 317 6980 10.1007/s10967-018-5913-2CrossRefGoogle Scholar
Wen, K., Zhu, J., Chen, H., Ma, L., Liu, H., Zhu, R., Xi, Y., He, H. Arrangement models of Keggin-Al30 and Keggin-Al13 in the interlayer of montmorillonite and the impacts of pillaring on surface acidity: A comparative study on catalytic oxidation of toluene Langmuir 2019 35 382390 10.1021/acs.langmuir.8b03447CrossRefGoogle Scholar
Wright, J. D., Sommerdijk, NAJM Sol-gel materials chemistry and applications 2001 CRC Press Taylor & Francis GroupGoogle Scholar
Xia, M., Jian, Y., Li, F., Sun, M., Xue, B., Chen, X. Preparation and characterization of bimodal mesoporous montmorillonite by using single template Colloids and Surfaces A 2009 338 16 10.1016/j.colsurfa.2008.12.043CrossRefGoogle Scholar
Yang, F., Spyrou, K., Thomou, E., Kumar, S., Cao, H., Stuart, MCA, Pei, Y., Gournis, D., Rudolf, P. Smectite clay pillared with copper complexed polyhedral oligosilsesquioxane for adsorption of chloridazon and its metabolites Environmental Science: Nano 2020 7 424436Google Scholar
Yi, H., Zhao, Y., Song, S. Development of superior stable two-dimensional montmorillonite nanosheet based working nanofluids for direct solar energy harvesting and utilization Applied Clay Science 2021 200 105886 10.1016/j.clay.2020.105886CrossRefGoogle Scholar
Zhang, L., Qin, Y., Ji, D., Chu, G., Gao, X., Zhang, X., Song, S. Effect of cerium ions initial distribution on the crystalline structure and catalytic performance of CeY zeolite Journal of Rare Earths 2017 35 791799 10.1016/S1002-0721(17)60978-5CrossRefGoogle Scholar
Zhang, J., Li, X., He, B., Song, Y., Ji, Y., Cui, Z., Li, J., Younas, M. Biodiesel production through heterogeneous catalysis using a novel poly(phenylene sulfide) catalytic membrane Energy & Fuels 2020 34 74227429 10.1021/acs.energyfuels.0c00522CrossRefGoogle Scholar
Zhang, T., Wang, W., Zhao, Y., Bai, H., Wen, T., Kang, S., Song, G., Song, S., Komarneni, S. Removal of heavy metals and dyes by clay-based adsorbents: From natural clays to 1D and 2D nano-composites Chemical Engineering Journal 2021 420 127574 10.1016/j.cej.2020.127574CrossRefGoogle Scholar
Zhang, Q., Zhang, Y., Liu, S., Wu, Y., Zhou, Q., Zhang, Y., Zheng, X., Han, Y., Xie, C., Liu, N. Adsorption of deoxynivalenol by pillared montmorillonite Food Chemistry 2021 343 128391 10.1016/j.foodchem.2020.128391CrossRefGoogle ScholarPubMed
Zhong, L., Cai, W., Yu, Y., Zhong, Q. Insights into synergistic effect of chromium oxides and ceria supported on Ti-PILC for NO oxidation and their surface species study Applied Surface Science 2015 325 5263 10.1016/j.apsusc.2014.11.024CrossRefGoogle Scholar
Figure 0

Fig. 1 SEM images of the clay samples

Figure 1

Table 1 Semi-quantitative analysis (wt.%) of chemical elements by EDS

Figure 2

Fig. 2 Powder XRD of the clay samples. A = anatase (TiO2), R = rutile (TiO2)

Figure 3

Fig. 3 Powder XRD of the clay samples in the range 2–14°2θ

Figure 4

Fig. 4 FTIR spectra of clay samples

Figure 5

Fig. 5 FTIR spectra of the clay samples in the range 1100–1000 cm–1

Figure 6

Fig. 6 N2 adsorption-desorption isotherms at 77 K for the clay samples

Figure 7

Table 2 Surface area and porosity of clay samples

Figure 8

Fig. 7 BJH pore-size distribution of clay samples

Figure 9

Fig. 8 a TGA and b DTGA analyses of the clay samples

Figure 10

Fig. 9 Conversion of acetic acid and penta-1-ol during the esterification reaction using the clay samples

Figure 11

Fig. 10 a Schematic representation of the structure of the clay samples, b esterification process of the clay samples

Figure 12

Table 3 Comparison of various catalysts in the synthesis of pentyl acetate