Next Article in Journal
Synthesis and Characterization of Supported Pd Catalysts for Potential Application in Glycerol Electro-Oxidation
Next Article in Special Issue
Methane Hydrate Formation in Hollow ZIF-8 Nanoparticles for Improved Methane Storage Capacity
Previous Article in Journal
Electrochemical Degradation of Nitrobenzene Wastewater: From Laboratory Experiments to Pilot-Scale Industrial Application
Previous Article in Special Issue
Modification of Silica Xerogels with Polydopamine for Lipase B from Candida antarctica Immobilization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress on Sulfated Nanozirconia as a Solid Acid Catalyst in the Hydrocracking Reaction

1
Department of Chemistry, Faculty of Mathematics and Natural Science, University of Pattimura, Ambon 97235, Indonesia
2
Department of Chemistry, Faculty of Mathematics and Natural Science, Universitas Gadjah Mada, Yogyakarta 55281, Indonesia
3
Department of Chemistry, Faculty of Mathematics and Natural Sciences, Universitas Islam Indonesia, Yogyakarta 55281, Indonesia
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(2), 191; https://doi.org/10.3390/catal12020191
Submission received: 17 December 2021 / Revised: 26 January 2022 / Accepted: 27 January 2022 / Published: 3 February 2022

Abstract

:
Zirconia has advantageous thermal stability and acid–base properties. The acidity character of ZrO2 can be enhanced through the sulfation process forming sulfated zirconia (ZrO2-SO4). An acidity test of the catalyst produced proved that the sulfate loading succeeded in increasing the acidity of ZrO2 as confirmed by the presence of characteristic absorptions of the sulfate group from the FTIR spectra of the catalyst. The ZrO2-SO4 catalyst can be further modified with transition metals, such as Platinum (Pt), Chromium (Cr), and Nickel (Ni) to increase catalytic activity and catalyst stability. It was observed that variations in the concentrations of Pt, Cr, and Ni produced a strong influence on the catalytic activity as the acidity and porosity of the catalyst increased with their addition. The activity, selectivity, and catalytic stability tests of Pt/ZrO2-SO4, Cr/ZrO2-SO4 and Ni/ZrO2-SO4 were carried out with their application in the hydrocracking reaction to produce liquid fuel. The percentage of liquid fractions produced using these catalysts were higher than the fraction produced using pure ZrO2 and ZrO2-SO4 catalyst.

1. Introduction

Zirconium dioxide, known as zirconia, is a crystalline oxide of zirconium that found form in the mineral baddeleyite. Zirconia is a white material that does not react with water or another solvent, and that has acid–base properties and excellent thermal dan chemical stabilization. ZrO2 materials are of wide interest and development in their application in various fields, such as heterogeneous catalyst, optics, electronics, magnetics, and ceramics owing to the high melting point (≥2700 °C), low thermal conductivity, corrosion resistance, and good thermal and mechanical strength [1,2,3].
Modification of zirconia to increase its catalytic activity has been developed. Many studies have showed that modification zirconia such as sulfated process or metal supported was effective during the chemical process [3,4,5]. Sulfated zirconia catalyst via hydrothermal treatment for hydrocracking of LDPE plastic waste into liquid fuels was examined by Utami et al. [6]. The total acidity of zirconia increased after the sulfation process, thus increasing the amount of liquid yield. However, the catalytic activity of sulfated zirconia catalyst during the hydrocracking reaction at high temperature decreases due to the deactivation catalyst. This process therefore requires an appropriate catalyst to increase the catalytic activity of sulfated zirconia, as well as the acidity and liquid yield [7,8].
Supported noble metals such as Platinum (Pt), Chromium (Cr), and Nickel (Ni) as a promoter have shown good catalytic activity in the hydrocracking reaction. The synthesis of Pt/sulfated zirconia [9], Cr/sulfated zirconia [10], and Ni/sulfated zirconia [11] catalyst in the hydrocracking reaction reported the enhance of acidity of sulfated zirconia and the liquid product of hydrocracking after the addition of promoter Pt, Cr, and Ni, respectively.

2. Zirconium Dioxide (ZrO2)

Zirconium dioxide (ZrO2) is a polymorphic material with three crystalline phases, namely the monoclinic, tetragonal, and cubic phases, as shown in Figure 1. The monoclinic and tetragonal phases are stable up to temperatures of 1170 and 2360 °C, respectively, while the cubic phase is stable at temperatures above 2680 °C. ZrO2 as a catalyst is normally used in its monoclinic and metastable tetragonal crystalline phases [12,13,14]. Transformation of the crystalline phases of ZrO2 is driven by changes in temperature. ZrO2 calcined at <800 °C forms monoclinic and metastable tetragonal phases [15,16].
The transformations of one ZrO2 phase to another are accompanied by changes in lattice parameters. The rate of phase transformation of ZrO2 is also influenced by the particle size of the ZrO2 precursor used. The larger the ZrO2 particle size, the faster the phase transformation occurs. Nano-sized particles have a high surface area, allowing for more of the atoms of the particles to interact and form bonds [15,17,18]. Based on surface properties and polymorphic form, ZrO2 is also often used as a catalyst or carrier material because it has acidic and basic properties [19,20]. In addition, ZrO2 has a structure with vacant sites on its surface that can allow cations to easily enter [21,22].

3. ZrO2-SO4 Catalyst

Zirconia is a catalyst containing high Lewis acid sites. Ore [15] mentioned that ZrO2 can be modified with acid or base to achieve, based on the intended application, the appropriate and desired strong acid or base characteristics. Sulfation is a method that can increase the strength of ZrO2 acid through the formation of Brønsted acids on the zirconia substrate that consequently increases its catalytic activity. ZrO2-SO4 can be prepared from sulfate precursors such as H2SO4, (NH4)2SO4, (NH4)2S2O3, and (NH4)2S) [24,25,26].
The formation of ZrO2-SO4 occurs through the chelation of the zirconium cations (Zr4+) with sulfate ions. The SO42− ions acts as ligands that donate their lone pair of electrons from the O atom, thus forming a coordination bond with two Zr4+ as the central atoms and causing the acid molecule to release two protons simultaneously. After relaxation, adsorptive complex molecules are produced from the coordination of SO42− ions onto the ZrO2 surface through the sharing of two O atoms [27,28,29]. Saravan et al. [30] illustrate the surface model of ZrO2-SO4 with the Lewis and Brønsted acid sites shown as shown in Figure 2.
Modification of zirconia to ZrO2-SO4 produces materials with high Lewis and Brønsted acid strengths. Wang et al. [29] presented an illustration of the ZrO2-SO4 surface as presented in Figure 3. The high acid strength of the Brønsted acid site is associated with the location of Zr4+ ions that are adjacent to the S=O bond which attracts electrons from the bisulfate to form the Brønsted acid site.

3.1. Functional Group Characterization for ZrO2-SO4

Research conducted by Utami et al. [31] reported the preparation of sulfated zirconia catalysts with various H2SO4 concentrations and calcination temperatures. Figure 4 shows the same absorption peak at wave number 424–741 cm−1, indicating the Zr–O–Zr bond [32]. Absorptions at 3426–3449 and 1636 cm−1 refer to the O-H stretching and bending vibrations of the water molecules adsorbed on the material [33]. In addition, according to Ore et al. [15], the broadband in the absorption region of 3400 cm−1 signifies the bridge between the hydroxyl group with two or three Zr atoms.
The presence of SO42− ions on the surface of ZrO2 can be confirmed by the formation of a new peak at 995-1404 cm−1 which is typical of the bidentate SO42− chelate ion covalently bonded to the Zr4+ cation [7,27]. The absorption peaks at 995–1003, 1049–1096, 1134–1157, and 1227 cm−1 are S–O symmetric vibrations, S–O asymmetric vibrations, S=O symmetry vibrations, and S=O asymmetric vibrations [34,35,36]. The peak with low intensity in the area of 1404 cm−1 is the stretching vibration of S=O, indicating the formation of SO3 species on the surface of ZrO2 [37]. The presence of characteristic bands of ZrO2-SO4 proves that the impregnation of H2SO4 on ZrO2 has been successfully carried out.
The use of too high a concentration of H2SO4 can lead to loss of several absorption bands of the SO42− ion which is covalently bonded to Zr4+ cations as the ZrO2 structure is degraded [9,37]. Such a case implies that an appropriate or optimum concentration of H2SO4 is needed for the ZrO2 activation process. The absorption intensity of ZrO2-SO4 increased at a temperature range of 500–600 °C but decreased at temperatures of 700–800 °C. Calcination treatment at 600 °C (SZ-0.8-600) was found to be optimum resulting in the highest SO42− dispersion. According to Ore [15], the maximum calcination temperature in the sulfation process is 650 °C. Temperatures above 650 °C cause the decomposition of SO42− ions, thereby reducing the acidity and reactivity of the catalyst.
The acidity test of the catalyst was carried out by the gravimetric method based on the amount of NH3 vapor absorbed by the catalyst. Table 1 shows the results of the catalyst acidity test. ZrO2 has a total acidity of 0.18 mmol/g. The acidity of ZrO2 comes from the Zr4+ cations that act as Lewis acid sites [7,8]. After modification with H2SO4, the acidity of ZrO2 increased. SZ-0.8-600 catalyst showed the highest acidity value of 1.06 mmol/g. Ore et al. [15] reported that the number of Brønsted and Lewis acid sites depends on the SO42− concentration present on the catalyst surface. Sulfation with optimum sulfate concentration can increase the catalytic activity of the catalyst, while the decrease in catalyst acidity with increasing temperature occurs due to dehydration of protonic sites and loss of SO42− groups on the surface of the catalyst [38,39,40].
Qualitatively, the number of Brønsted and Lewis acid sites can be observed absorption spectra intensity that denotes the interaction between the catalyst acid sites and NH3 [41]. Figure 5 presents the FTIR spectra of the catalyst after the acidity test. The absorption peak at 1119-1126 cm−1 indicates the presence of NH3 coordinated to the Lewis acid site. The peak at 1404 cm−1 confirmed the presence of NH4+ ions formed by proton transfer from the Brønsted acid site to NH3 [42,43]. The higher the intensity of the absorption band, the higher the number of Brønsted and Lewis acid sites. It was found that the SZ-0.8-600 catalyst had the highest acid site intensities. The absorption intensities of the acid sites decreased with the increase in calcination temperature. Calcination treatment at high temperature decreases the acidity of the catalyst due to the decrease in the number of acid sites on the surface of the catalyst [44,45].

3.2. ZrO2-SO4 Crystal Structure Characterization

Figure 6 presents the diffraction pattern of the SZ catalyst at various sulfate concentrations and calcination temperatures. The main diffraction peaks appear at 2θ = 28.34° (d-111) and 31.64° (d111), referring to the ZrO2 monoclinic crystalline phase [6,15]. In general, the diffraction pattern showed stable crystallinity even after the addition of acid and calcination treatment. However, the intensity of the ZrO2 monoclinic diffraction peak decreased after acid treatment. SZ-0.8-600 catalyst with the highest total acidity showed the lowest monoclinic peak intensity. The addition of a high concentration of H2SO4 causes a large number of SO42− ions to cover the surface of ZrO2, decreasing crystallinity [37,38]. The intensities of the monoclinic peaks at temperatures of 800 and 900 °C were higher than those at 600 and 700 °C. This occurred because the high calcination temperature caused SO42− ions to decompose from the catalyst surface, increasing the crystallinity of the catalyst.
ZrO2-SO4 calcined at a temperature of <800 °C can exist in a metastable tetragonal phase and a monoclinic phase. Similar results were reported by Ore et al. [15] which stated that ZrO2-SO4 calcined at a temperature of 600 °C consists of a mixture of metastable tetragonal and monoclinic phases. The ZrOCl2·8H2O amorphous precursor used made it possible for transformation to a metastable tetragonal structure to occur. However, the ZrO2 used in this study is commercial ZrO2, which contains the monoclinic structure of high stability and crystallinity, hindering it from undergoing phase transformation [46]. The catalyst diffraction pattern in the research by Utami et al. [6] demonstrated the ZrO2 catalyst and its modifications consisting only of monoclinic structure.

4. Platinum/Sulfated Zirconia (Pt/ZrO2-SO4) Catalyst

A heterogeneous catalyst is a catalyst material composed of two components, namely the doping and carrier components. The metal catalyst, when used in its pure form, has low thermal stability and tends to sinter that can leading to a decrease in surface area and deactivation [47,48,49]. Appropriate distribution of metal catalysts on the carrier material having acid–base sites and a large surface area is necessary to avoid sintering [50,51]. The ZrO2-SO4 material has many Brønsted and Lewis acid sites in which, despite its high acidity, this catalyst can be rapidly deactivated. The addition of Pt metal can increase the stability of the catalyst with the simultaneous presence of hydrogen gas (H2) [52,53,54].
The distribution mechanism of H2 through the Pt surface on the ZrO2-SO4 carrier is illustrated in Figure 7. The H2 molecule dissociates on the surface of the Pt particle homolytically to form two H radicals which then bind to the unpaired electrons in the 5d orbitals. The H+ ions released from Pt are distributed on the ZrO2-SO4 carrier and migrate to the electron-rich O atomic sites, forming Brønsted acid sites [55,56]. Figure 8 shows the physical appearance of ZrO2-SO4 catalyst which a white powder and after the impregnation with Pt metal as a Pt/ZrO2-SO4 catalyst, solid powder which is darker in color is formed due to the presence of the Pt metal that impregnated to the ZrO2-SO4 catalyst [6,9].

4.1. FTIR and Acidity Characterization of Pt/ZrO2-SO4 Catalyst

Utami et al. [6] researched the synthesis of Pt-promoted zirconia (Pt/SZ) catalyst and its application in hydrocracking LDPE plastic into liquid fuel. The FTIR spectra of the sulfated zirconia impregnated with Pt metal can be seen in Figure 9. Overall, the Pt1/SZ, Pt2/SZ, and Pt3/SZ spectra showed the same absorption peaks as ZrO2 and SZ. Pt metal impregnated catalysts had characteristic peaks of ZrO2-SO4 at wave number 1065–1126 cm−1. The addition of Pt metal caused some of the absorption peaks of ZrO2-SO4 to disappear. This is an early indication that Pt metal had been impregnated on nanoSZ [9].
Table 2 shows the total acidity of the catalysts of the present study. Pt metal impregnation on sulfated zirconia was proven to increase the total acidity of the catalyst significantly (from 1.06 to 10.75 mmol/g). Pt1/SZ, Pt2/SZ, and Pt3/SZ catalysts showed increasing acidity values with increasing concentrations of Pt metal. The increase in the acidity of the catalyst occurs because the Pt metal provides vacant orbitals that can act as electron-pair acceptors (Lewis acid sites) and the presence of unpaired electrons in the d orbitals that form Brønsted acid sites [57,58].
FTIR spectra interpretations of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ are shown in Figure 10. Increasing concentration of Pt metal produced increased adsorption of NH3 bound to the Brønsted and Lewis acid sites as shown at wavenumbers of 1396–1404 and 1119 cm−1, indicating that the higher the concentration of Pt metal, the higher the number of acid sites contained in the catalyst. The Pt3/SZ catalyst showed the highest intensity of Brønsted and Lewis acid absorptions. Based on the results of the acidity test of the catalyst, the Pt3/SZ catalyst was confirmed to have the highest acidity value.

4.2. XRD and GSA Characterizations of Pt/ZrO2-SO4 Catalyst

SZ catalysts are shown in Figure 11. Based on crystal identification, all samples showed the presence of the monoclinic phase. According to Ore [15], SO42− species can be thermally crystallized through the calcination process and undergo a crystalline phase transformation, further stabilizing the ZrO2 crystalline phase. The addition of Pt metal to SZ would not have caused changes in the crystal structure of the material. A decrease in the intensity of the diffraction peak after the addition of Pt metal was observed. This phenomenon indicated that Pt metal had been successfully impregnated on the SZ surface, where a higher concentration of the impregnated Pt metal would cause the intensity of the monoclinic peak to decrease [6,9].
Based on the catalyst diffraction pattern, Pt metal characteristic peaks were not detected. This is because the concentration of impregnated Pt metal was relatively low. This is by research conducted by Aboul-Gheit et al. [59] that showed similar results whereby the diffraction peak of Pt metal was not identified after the addition of 0.6% Pt metal to ZrO2-SO4. A relatively low concentration of Pt metal was used in the present study to prevent agglomeration of Pt particles on the ZrO2-SO4 surface, which could cause a decrease in catalytic activity.
Characterization carried out by GSA (Table 3) showed an increase in surface area and pore volume along with the increasing concentration of Pt metal impregnated onto SZ. The Pt1/SZ catalyst had a surface area and pore volume of 13.49 m2/g and 0.05 cm3/g, respectively, which saw an increase to 20.23 m2/g and 0.06 cm3/g for Pt2/SZ and 29.48 m2/g and 0.08 cm3/g for Pt3/SZ. Based on the data reported in the study by Utami et al. [9], the increase in the surface area and pore volume of the SZ catalyst can be attributed to the inhibition of the agglomeration process related to the high presence of Pt metal.
The increase in the surface area of the catalyst can also be attributed to the smaller crystal size dimensions. Crystal size results of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ (Table 4) presented a decrease in size after metal impregnation of Pt. This indicated that the Pt metal impregnated by the reflux technique was evenly dispersed on the surface and pores of SZ [9]. In contrast, Aboul-Gheit et al. [59] impregnated Pt metal on ZrO2-SO4 through the wet impregnation method and reported an increase in the concentration of Pt metal causing the surface area and pore volume to increase due to the Pt metal being not evenly distributed and covering most of the ZrO2-SO4 pores.

4.3. Elemental Composition Characterization Using EDXRF

Table 5 shows the concentrations of elements contained in the catalysts. The concentrations of Pt metals identified in the Pt1/SZ, Pt2/SZ, and Pt3/SZ samples were 0.35, 0.90, and 1.19%. The metal content of Pt in SZ was strongly influenced by the dispersion ability on the surface and pores of the carrier material. The reduction treatment with H2 gas flow was carried out after the calcination step which aimed to obtain Pt0 particles that would result in dispersion [14].
Based on the elemental compositions above, the concentrations of Pt that were observed were lower than the theoretical concentration of Pt metal by the addition of PtCl4 solution used in the impregnation process. This discrepancy occurred due to competition between impregnated Pt metals, causing the formation of multilayer stacking of the active metal in the pore mouth area of the carrier material [60,61]. The active Pt metal that sat at the top position would have weak interaction and would experience easier desorption [62]. Because of this, the amount of active metal in the pore area of the carrier material was observed to be less.

4.4. Pt metal Composition Identification Using XPS

Subsequent research related to the synthesis of platinum-loaded sulfated zirconia catalysts using the hydrothermal method was reported by Utami et al. [9]. XPS spectra were used to determine the composition of Pt in the samples based on a comparison of binding energy values. The peaks generated from XPS are not single peaks, so that deconvolution of the peaks was needed to identify the multiple peaks that made up each peak. Figure 12 presents the XPS spectra along with the deconvolution of Pt 4f peaks from Pt/nano ZrO2-SO4 consisting of Pt0 4f7/2, Pt0 4f5/2, and Pt2+ 4f7/2. The spectra indicated the interaction of electrons between the Pt particles and the nanoZS surface. Table 6 shows the relative area of the deconvoluted peaks of Pt0 4f7/2, Pt0 4f5/2, and Pt2+ 4f7/2 in the Pt/nano ZrO2-SO4 sample. The data obtained showed that the detected Pt0 composition was 81.82%. The results indicated that the reduction treatment with H2 gas flow at the metal impregnation stage succeeded in forming Pt0 particles.

4.5. Thermal Stability Characterization with TG/DTA

The TG/DTA curve provides information about changes in thermal conditions with mass changes in the sample. Figure 13 shows the TGA curves for nano ZrO2, nano ZrO2-SO4, and Pt/nano ZrO2-SO4 samples analyzed at temperatures of 30–900 °C. The TGA curve of nano ZrO2 did not indicate mass decrease indicating that nano Z had good thermal stability [63]. Mass decrease in nano ZrO2-SO4 and Pt/nano ZrO2-SO4 in the range of 50–200 °C by 2.41 and 1.24% were associated with the elimination of water molecules physically adsorbed on the material. At 500–700 °C, decreases in a mass of 5.69 and 2.48%, indicating the decomposition of SO42− ions bound to the ZrO2 surface. The decomposition of SO42− ions at 600–1000 °C affected the structural changes of ZrO2-SO4, causing a decrease in catalytic activity [64]. The decomposition of H2SO4 occurs through a two-stage endothermic process at high temperatures according to Equations (1) and (2).
H2SO4 (aq)   →   H2O (l) + SO3 (g)
SO3 (g)      →   SO2 (g) + ½ O2 (g)

4.6. Activity and Selectivity of Pt/ZrO2-SO4 Catalyst in LPDE Hydrocracking Application

Utami et al. [6] reported the activity and selectivity of the Pt/ZrO2-SO4 catalyst applied in LPDE hydrocracking. The hydrocracking liquid fraction of the LDPE plastic waste thermal cracking is shown in Figure 14. Physically, the hydrocracking liquid fraction has a yellow color and pungent odor, which indicate the success of the hydrocracking process. The percentages of liquid fractions produced using the Pt1/SZ, Pt2/SZ, and Pt3/SZ catalysts were higher than that of the SZ catalyst. The percentage of liquid fraction obtained with the SZ catalyst was 57.92%, while with Pt1/SZ, Pt2/SZ, and Pt3/SZ the liquid fractions were 70.58, 71.15, and 74.60%, respectively. Through MS data, it was found that the hydrocarbon compounds in the gasoline range (C5–C12) were more commonly found in the liquid fraction samples that had used Pt3/SZ (catalyst with the highest acidity). Table 7 shows the gasoline fraction percentages from the hydrocracking reaction using Pt1/SZ, Pt2/SZ, and Pt3/SZ catalysts, which were 48.76, 64.22, and 67.51 w/w%, respectively. This cased that besides affecting the amount of hydrocracking liquid fraction produced, the addition of Pt metal can also increase selectivity towards the gasoline fraction [65].
The percentage of hydrocarbon compounds in the C5–C12 range can be seen to be greater than that of C13–C20 hydrocarbons. This shows that the hydrocracking liquid fraction of LDPE plastic waste had a higher gasoline fraction than the diesel fraction and that the four types of catalysts used had good selectivity towards the hydrocracking reaction that produces liquid fuel fraction (gasoline fraction).
Figure 15 shows the proportion of hydrocarbon compounds contained in the hydrocracked liquid fraction in the gasoline range, namely olefins, paraffin, isoparaffins, and naphthenes, with a total composition of 56.36 and respective amounts of 20.07, 14.60, and 6.81% w/w with use of the Pt3/SZ catalyst. Aromatic compounds in small amounts were also produced with a composition of 0.70% w/w, and only 1.46% w/w was indicated to be non-hydrocarbon compounds. Overall, olefin (unsaturated/double-bonded compound) was dominantly produced from the hydrocracking. This is because LDPE, as the plastic feed used, consists of olefin in which, during the hydrocracking reaction, not all olefins react with the hydride to become paraffin (saturated/single bond compound) [6,9].

4.7. Stability Test of Pt/ZrO2-SO4

A stability test of the Pt/nano ZrO2-SO4 catalyst, along with nano ZrO2 and nano ZrO2-SO4 for comparison, was carried out by Utami et al. [9] through a hydrocracking reaction of LDPE plastic waste with a catalyst/feed ratio 1% w/w and a temperature of 250 °C for 60 min. The catalyst stability test was carried out for six cycles with the same reaction conditions. Figure 16 shows that ZrO2, SZ, and Pt3/SZ had good catalytic performances when first used. The hydrocracking reaction with the ZrO2 catalyst showed a significant decrease in the percentage of liquid fraction produced in the fourth cycle, while in SZ this significant reduction occurred in the second cycle.
The catalytic activity of Pt3/SZ showed very good stability up to the sixth cycle. Aboul-Gheit et al. [59] reported that the catalytic activity of ZrO2-SO4 in the n-pentane isomerization reaction decreased in the fourth cycle and became inactive in the eighth cycle. In contrast, the Pt/ZrO2-SO4 catalyst showed stable activity until the tenth cycle. The Pt/ZrO2-SO4 catalyst showed high resistance to the deactivation related to the removal of coke from the catalyst surface, thus increasing the stability of the catalyst [36,37].
Figure 16 illustrates the proposed origin of the catalytic stability of ZrO2-SO4 and Pt/nanoZS3-600 nanoscales. Based on the activity and catalytic selectivity data, the conversion of LDPE using nano ZrO2-SO4 produced the highest amount of coke. In addition, this material had a high initial activity but low resistance to the deactivation process as its catalytic properties decreased rapidly in its consequent cycle of use. The formation of a coke can causes the pores and the active sites of the catalyst to be closed and thus reduce activity [44], i.e., the deactivation process that occurred in the ZrO2-SO4 nanocatalyst would have been difficult to control. To restore the activity of nanoZS3-600, catalyst regeneration, i.e., coke removal, is indispensable, especially in large-scale industrial applications. Promisingly, the Pt/nanoZS3-600 catalyst showed good activity, selectivity, and stability even after repeated use.

5. Cr/ZrO2-SO4 Catalyst

The metals that are widely used for bifunctional catalysts are usually transition metals with have incomplete orbitals that function as Lewis acid sites. One of the said transition metals is chromium (Cr) [66]. Cr can be doped on a carrier to enhance the catalytic activity of the host [67,68]. The addition of Cr metal to sulfated zirconia can also increase the acidity of the catalyst as the metal would contribute to the presence of Lewis acid sites [69]. The presence of Cr metal on sulfated zirconia also affects increasing the surface area of the catalyst [70,71]. Figure 17 illustrates the reaction mechanism of sulfated zirconia impregnated with Cr metal when interacting with ammonia during an acidity test. Hauli et al. [71] stated that the addition of chromium metal to sulfated zirconia not only increases the surface area of the catalyst but also stabilizes pure zirconia at high calcination temperatures. Increased temperature can thus remove sulfate groups and damage the porosity of the structure, causing the catalyst to deactivate. The physical appearance of the ZrO2-SO4 and Cr/ZrO2-SO4 catalyst (Figure 18) shows solid particle color change after the impregnation of Cr metal to the ZrO2-SO4 catalyst. The gray solid particle is formed due to the presence of Cr metal.

5.1. FITR and Acidity Characterizations of Cr/ZrO2-SO4 Catalyst

Research on Cr/ZrO2-SO4 catalyst has been reported by Hauli et al. [72]. FTIR results of ZrO2, SZ 0.8-600 (SZ), 0.5% Cr/SZ (Cr1/SZ), 1.0% Cr/SZ (Cr2/SZ), and 1.5% Cr/SZ (Cr3/SZ) are shown in Figure 19. The spectra showed no specific differences between the spectra of Cr-embedded zirconia catalyst and sulfated zirconia. However, the specific absorption peak of the sulfate group at 1053–1224 cm−1, decreased with the presence of Cr. This is because the heating process involved in the loading of the metal had allowed sulfates to be released from the ZrO2 surface. The presence of metal produced not sulfate decomposition but rather the release of the sulfate groups from the surface of ZrO2 during the heating process [72]. Hanifah et al. [38] reported similar phenomena, namely that the presence of monometals and bimetals on ZrO2-SO4 release of sulfate group materialized whilst its decomposition did not.
The presence of empty orbitals in Cr metal allows electrons from other atoms to take their place in the orbital in what would then contribute to higher acidity of the Cr-containing material. Acidity test results of Cr-containing catalysts are presented in Table 8 confirming this. The acidity of the catalyst increased after the addition of Cr metal. The Cr1/SZ catalyst was the catalyst with the highest acidity value with 8.22 mmol/g. The low acidity value at high Cr-metal concentration was caused by the presence of metal particle aggregates that had covered the active metal sites on the carrier material. The more Cr metal that was embedded, the greater the amount of metal not accommodated in the pores of the carrier material, leading to the formation of aggregates.
The acidity test of the catalyst was carried out to determine the total acidity value of each catalyst. The FTIR spectra after the acidity test results for the ZrO2, SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ catalysts are shown in Figure 20. The results show the same absorption peaks at wave numbers 1119 and 1404 cm−1, confirming the presence of the Lewis and Brønsted acid sites on each catalyst. Cr-embedded sulfated zirconia catalyst showed a higher intensity of Lewis and Brønsted acid sites than before Cr metal was added. The spectra of the Cr/SZ catalyst had the highest acid site intensity, indicating that Cr2/SZ had a high acidity value. The results of the acidity test of the catalysts confirmed that Cr2/SZ had the highest acidity value of 8.22 mmol/g.

5.2. XRD Characterization of Cr/ZrO2-SO4 Catalyst

The diffraction patterns of ZrO2, SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ catalysts are presented in Figure 21. The bearing of Cr metal on sulfated zirconia provided no change in the crystal phase structure of ZrO2. The main peaks were found in the 2θ region around 28°, 31°, and 50° denoting plane distances of d-111 (3,2 Å), d111 (2,8 Å), d220 (1,8 Å) with the highest intensity in the 28° region. Sulfate and Cr impregnation on ZrO2 have a lower intensity peak than ZrO2. This is due to the presence of sulfate on ZrO2 and to Cr metal covering the surface of SZ, inhibiting the growth of ZrO2 crystals and thereby decreasing the crystallinity of the materials [73,74].

5.3. SAA Characterization of Cr/ZrO2-SO4 Catalyst

Characterization of pore characteristics including surface area, pore diameter, and pore volume was carried out by SAA analysis. The results of the SAA measurements of the catalysts are shown in Table 9. As can be seen, the addition of Cr metal to sulfated zirconia increased the surface area, pore diameter, and pore volume of the catalyst. This can be attributed to the uniform distribution of Cr metal on the catalyst surface [69,70]. The surface area of the catalyst, however, decreased upon the addition of higher concentrations of Cr metal, namely at Cr3/ZS, due to the entry of Cr metal into the catalyst pores causing agglomeration of metal atoms that covered the catalyst pores [74].
The adsorption and desorption isotherm patterns for ZrO2, SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ catalysts are shown in Figure 22. Based on the IUPAC classification, all catalysts showed a type IV isotherm pattern, which is characteristic of the isotherm pattern for mesoporous materials with pore diameter sizes of 2–50 nm. The adsorption–desorption isotherm patterns of the Cr1/SZ and Cr2/SZ catalysts demonstrated monolayer absorption of nitrogen gas on the surface when P/P0 < 0.2. At a relative pressure of 0.2 < P/P0 < 1, the isotherm curve experienced a sharp increase in volume representing a multilayer arrangement. In the Cr3/SZ catalyst, the absorption of a monolayer of nitrogen gas on the surface occurred when P/P0 < 0.4 and experienced a sharp increase in volume at 0.4 < P/P0 < 1. The ZrO2, Cr2/SZ, and Cr3/SZ catalyst had a type H4 hysteresis, while SZ and Cr1/SZ had type H3. Type H3 hysteresis showed no absorption limit at high P/P0. Type H4 hysteresis is associated with narrow slit pores [4,72]. These three types of catalysts exhibit the characteristics of porous materials. This form of porous material is composed of aggregates of particles such as plates that form pore gaps [10].

5.4. Activity and Selectivity Tests of Cr/ZrO2-SO4 Catalyst

A catalytic activity test of the Cr/ZrO2-SO4 catalyst in the hydrocracking reaction of LDPE into liquid fuel has been carried out by Hauli et al. [72]. The hydrocracking reaction was carried out at a temperature of 250 °C using ZrO2, SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ catalysts. The percentages of conversion yields of the hydrocracking of each catalyst are presented in Table 10. The presence of sulfate and Cr metal in ZrO2 can increase the yield of liquid products from the hydrocracking reaction. The liquid product increased along with the increase in the amount of Cr metal deposited on ZrO2. This can be attributed to the role of acid sites in the catalyst (from sulfate and metal Cr) which contribute to the active sites of the catalyst, thereby increasing its catalytic activity [50].
The highest conversion of liquid product was produced using the Cr2/SZ catalyst, which was 40.99%, with a lesser amount of 37.51% with the use of the Cr3/SZ catalyst. Moreover, the Cr2/SZ catalyst had a higher acidity value than Cr3/SZ. In addition, the surface area of the catalyst also affects the catalytic activity of the catalyst in the hydrocracking of LDPE plastics [75,76]. Accordingly, the Cr2/SZ had a larger surface area, resulting in greater conversion. The modification of Cr on sulfated zirconia demonstrated a reduction in coke formation, thereby increasing the feed interaction on the active sites of the catalysts.
The selectivity of the catalysts towards liquid products in the hydrocracking reaction are shown in Figure 23. All the catalysts showed higher selectivity in the gasoline fraction than the diesel fraction. The presence of sulfate and Cr in sulfated zirconia increased the selectivity towards the gasoline fraction (C5–C12) and decreased the selectivity towards the diesel fraction (C13–C20), as expected. The highest selectivity of the gasoline fraction was obtained from the use of the Cr2/SZ catalyst, which was at 93.42%.

6. Ni/ZrO2-SO4 Catalyst

6.1. FITR Characterization of Ni/ZrO2-SO4 Catalyst

Research on the ZrO2-SO4 catalyst was also carried out about its application in the hydrocracking of used cooking oil into liquid fuel. Modification of ZrO2-SO4 with Ni metal has been reported by Aziz et al. [73]. Figure 24 show the FTIR spectra of Ni-SZ 1, Ni-SZ 2, and Ni-SZ 3. Impregnated Cr metal to SZ caused the presence of a new peak in the area of 1103 and 1141 cm−1, confirming the S-O-S stretching of from SO4 ion from SZ coordinated with Ni metal [76,77]. However, the sulfate spectra at 1002–1219 cm−1 disappear due to the Ni metal coverage on the SZ surface and the higher calcination temperature.

6.2. Acidity and SAA Analysis of Ni/ZrO2-SO4 Catalyst

The total acidity and pore characteristic of Ni-SZ catalysts is shown in Table 11. It can be seen that the sulfation of ZrO2 (SZ catalyst) and Ni metal impregnated to SZ catalysts successfully increase the total acidity of the catalyst. Ni metal has a vacant p orbital that will accepting electron pair and acts as a Lewis acid site [78]. The Ni-SZ 3 catalyst was the catalyst with the highest total acidity of 4.235 mmol/g.
The pore characteristic of the catalyst shows the increase in surface area after the impregnation of Cr metal onto the SZ catalyst due to the highly dispersion of Ni metal on the pore and surface of the SZ catalyst [79]. However, the presence of Ni causing the form of pore-blocking that significantly decreases the pore diameter of Ni-SZ catalysts [72].
The selectivity of the Ni-ZrO2-SO4 catalyst is shown in Table 12. ZrO2 and SZ catalysts were observed to have lower activity and selectivity than Ni-SZ due to the lowest acidity and surface area that will produce considerable amounts of coke (block the active site) and decreases the amount of liquid product [80,81]. The addition of Ni metal (Ni-SZ 3 catalyst) increased the acidity (4.23 mmol/g) and surface area (11.68 m2/g) of ZrO2, thereby increasing its activity and selectivity in the hydrocracking process [82,83]. The largest amount of liquid product produced was gasoline with the highest selectivity was produced by the Ni-SZ 3 catalyst with the diesel fraction (C5–C12) reached 100%.

7. Conclusions

Zirconia and its modified heterogeneous catalyst forms hold great potential in hydrocracking reaction applications to convert LDPE waste into liquid fuels with their excellent activity, selectivity, and stability. The sulfation process on ZrO2 with various concentrations of sulfuric acid and calcination temperatures succeeded in increasing the acidity of ZrO2. The ZrO2-SO4 catalyst treated with Platinum (Pt) and Chrome (Cr) transition metals had significantly increased acidity. Characterization analyses confirmed that Pt and Cr metals had been successfully impregnated on the SZ surface. The percentage of liquid fraction obtained with the use of the Pt/ZrO2-SO4 catalyst proofed better activity, selectivity, and stability than Cr/ZrO2-SO4 and ZrO2-SO4. The optimum amount of liquid fraction produced from the hydrocracking reaction LDPE plastic waste with the Pt/ZrO2-SO4 catalyst was 67.515%, while that from the Cr/ZrO2-SO4 catalyst was 40.99%. The development of Ni on the ZrO2-SO4 catalyst also demonstrated an increase in selectivity in the hydrocracking reaction of used cooking oil into liquid fuel. The selectivity of the Ni-SZ catalyst was found to be better than that of SZ. The percentages of gasoline fractions produced were 100%.

8. Future Suggestion

Zirconia-based nanocatalysts have bright prospects for application in various industrial areas such as petroleum cracking, biofuel synthesis, pharmaceutical, and the synthesis of various organic materials. The potential application of sulfated nanozirconia catalysts is shown in the Figure 25. Due to characteristics such as being non-toxic and easy to regenerate, as well as having a large surface area and high thermal and structural resistance, nanozirconia catalysts also have the potential to be used in the pharmaceutical industry. Preliminary studies in our laboratory have shown that this catalyst has the potential to be used as a solid acid catalyst in the synthesis of nitrobenzene from benzene.

Author Contributions

Conceptualization, R.A.P., S.J.S., M.U., L.H. and A.K.A.; validation, R.A.P., M.U., L.H. and A.K.A.; investigation, R.A.P.; resources, R.A.P.; data curation, R.A.P., M.U., L.H. and A.K.A.; writing—original draft preparation, R.A.P. and S.J.S.; writing—review and editing, R.A.P.; visualization, K.W.; supervision, R.A.P. and K.W.; project administration, K.W.; funding acquisition, K.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Postdoctoral Research Grant Universitas Gadjah Mada (Contract Number: 6144/UN1/DITLIT/DIT-LIT/PT/2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, D.; Feng, W.; Chen, C.; Chen, S.; Fan, G.; Liao, S.; Wu, G.; Wang, Z. Transesterification of Litsea cubeba kernel oil to biodiesel over zinc supported on zirconia heterogeneous catalysts. Renew. Energy 2021, 177, 13–22. [Google Scholar] [CrossRef]
  2. Hanafi, M.F.; Sapawe, N. Reusability study of zirconia catalyst toward photocatalytic degradation of Remazol brilliant blue dye. Mater. Today Proc. 2020, 31, 266–2688. [Google Scholar] [CrossRef]
  3. Amin, A.K.; Trisunaryanti, W.; Wijaya, K. Effect of promoters and calcination temperature on surface and acidity of modified zirconia. J. Nano Res. 2018, 57, 31–39. [Google Scholar] [CrossRef]
  4. Hauli, L.; Wijaya, K.; Armunanto, R. Preparation and characterization of sulfated zirconia from a commercial zirconia nanopowder. Orient. J. Chem. 2018, 34, 1559–1564. [Google Scholar] [CrossRef]
  5. Yan, C.; Zheng, S.; Chen, N.; Yuan, S.; Chen, Y.; Li, B.; Zhang, Y. Sulfated zirconia catalysts supported on mesoporous Mg-SBA-15 with different morphologies for highly efficient conversion of fructose to 5-hydroxymethylfurfural. Micropor. Mesopor. Mater. 2021, 328, 111507. [Google Scholar] [CrossRef]
  6. Utami, M.; Wijaya, K.; Trisunaryanti, W. Pt-promoted Sulfated Zirconia as Catalyst for Hydrocracking of LDPE Plastic Waste into Liquid Fuels. Mater. Chem. Phys. 2018, 213, 548–555. [Google Scholar] [CrossRef]
  7. Yan, G.X.; Wang, A.; Wachs, I.E.; Baltrusaitis, J. Critical review on the active site structure of sulfated zirconia catalysts and prospects in fuel production. Appl. Catal. A. Gen. 2019, 572, 210–225. [Google Scholar] [CrossRef]
  8. Wijaya, K.; Kurniawan, M.A.; Saputri, W.D.; Trisunaryanti, W.; Mirzan, M.; Hariani, P.L.; Tikoalu, A.D. Synthesis of nickel catalyst supported on ZrO2/SO4 pillared bentonite and its application for conversion of coconut oil into gasoline via hydrocracking process. J. Environ. Chem. Eng. 2021, 9, 105399. [Google Scholar] [CrossRef]
  9. Utami, M.; Trisunaryanti, W.; Shida, K.; Tsushida, M.; Kawakita, H.; Ohto, K.; Wijaya, K.; Tominaga, M. Hydrothermal preparation of a platinum-loaded sulphated nanozirconia catalyst for the effective conversion of waste low density polyethylene into gasoline-range hydrocarbons. RSC Adv. 2019, 9, 41392–41401. [Google Scholar] [CrossRef] [Green Version]
  10. Hauli, L.; Wijaya, K.; Syoufian, A. Fuel production from LDPE-based plastic waste over chromium supported on sulfated zirconia. Indones. J. Chem. 2019, 20, 422–429. [Google Scholar] [CrossRef] [Green Version]
  11. Amin, A.K.; Trisunaryanti, W.; Wijaya, K. The Catalytic Performance of ZrO2-SO4 and Ni/ZrO2-SO4 Prepared from Commercial ZrO2 in Hydrocracking of LDPE Plastic Waste into Liquid Fuels. Orient. J. Chem. 2018, 34, 3070–3078. [Google Scholar] [CrossRef] [Green Version]
  12. Treccani, L.; Klein, T.Y.; Meder, F.; Pardun, K.; Rezwan, K. Functionalized ceramics for biomedical, biotechnological and environmental applications. Acta Biomater. 2013, 9, 7115–7150. [Google Scholar] [CrossRef] [PubMed]
  13. Skovgaard, M.; Gudik-Sorensen, M.; Almdal, K.; Ahniyas, A. Nanoporous zirconia microspheres prepared by salt-assisted spray drying. SN Appl. Sci. 2020, 2, 784. [Google Scholar] [CrossRef] [Green Version]
  14. Utami, M. Platinum/Sulfated Nanozirconia: Preparation, Characterization, and Their Application in Conversion of LDPE Plastic Waste into Liquid Fuels. Ph.D. Thesis, Department of Chemistry, Universitas Gadjah Mada, Yogyakarta, Indonesia, 2019. [Google Scholar]
  15. Le Ore, M.S.; Wijaya, K.; Trisunaryanti, W.; Saputri, W.D.; Heraldy, E.; Yuwana, M.W.; Hariani, P.L.; Budiman, A.; Sudiono, S. The Synthesis of SO4/ZrO2 and Zr/CaO catalysts via hydrothermal treatment and their application for conversion of low-grade coconut oil into biodiesel. J. Environ. Chem. Eng. 2020, 8, 104205. [Google Scholar] [CrossRef]
  16. Burger, W.; Kiefer, G. Alumina, zirconia and their composite ceramics with properties tailored for medical applications. J. Compos. Sci. 2021, 5, 306. [Google Scholar] [CrossRef]
  17. Siddiqui, M.R.H.; Al-Wassil, A.I.; Alotaibi, A.M.; Mahfouz, R. Effects of precursor on the morphology and size of ZrO2 nanoparticles, synthesized by sol-gel method in non-aqueous medium. Mater. Res. 2012, 15, 986–989. [Google Scholar] [CrossRef] [Green Version]
  18. Halter, W.; Eisele, R.; Rothenstein, D.; Bill, J.; Allgower, F. Moment dynamics of zirconia particle formation foe optimizing particle size distribution. Nanomaterials 2019, 9, 333. [Google Scholar] [CrossRef] [Green Version]
  19. Maleki, F.; Pacchiono, G. Characterization of acid and basic sites on zirconia surfaces and nanoparticles by adsorbed probe molecules: A theoretical study. Top. Catal. 2020, 63, 1717–1730. [Google Scholar] [CrossRef]
  20. Glorius, M.; Markovits, M.A.C.; Breitkopf, C. Design of specific acid-base-properties in CeO2-ZrO2-mixed oxides via templating and Au modification. Catalysts 2018, 8, 358. [Google Scholar] [CrossRef] [Green Version]
  21. De Souza, E.F.; Appel, L.G. Oxygen vacancy formation and their role in the CO2 activation on Ca doped ZrO2 surface: An an-initio DFT study. App. Surf. Sci. 2021, 553, 149589. [Google Scholar] [CrossRef]
  22. Song, L.; Cao, X.; Li, L. Engineering stable surface oxygen vacancies on ZrO2 by hydrogen-etching technology: An efficient support of gold catalysts for water-gas shift reaction. ACS Appl. Mater. Interfaces 2018, 109, 31249–31259. [Google Scholar] [CrossRef] [PubMed]
  23. Gautam, C.; Joyner, J.; Gautam, A.; Rao, J.; Vajtai, R. Zirconia Based Dental Ceramics: Structure, Mechanical Properties, Biocompatibility and Applications. Dalton Trans. 2016, 45, 19194–19215. [Google Scholar] [CrossRef] [PubMed]
  24. Hauli, L.; Wijaya, K.; Syoufian, A. Hydrocracking of LDPE plastic waste into liquid fuel over sulfated zirconia from a commercial zirconia nanopowder. Orient. J. Chem. 2019, 35, 128–133. [Google Scholar] [CrossRef]
  25. El-Desouki, D.S.; Ibrahim, A.H.; Abdelazim, S.M.; Aboul-Gheit, N.A.K.; Abdel-Hafizar, D.R. The optimum condition for methanol conversion to dimethyl ether over modified sulfated zirconia catalysts prepared by different methods. J. Fuel Chem. Technol. 2021, 49, 63–71. [Google Scholar] [CrossRef]
  26. Yu, S.; Wu, S.; Li, L.; Ge, X. Upgrading bio-oil from waste cooking oil by esterification using SO42−/ZrO2 as catalyst. Fuel 2020, 276, 118019. [Google Scholar] [CrossRef]
  27. Wang, P.; Yue, Y.; Wang, T.; Bao, X. Alkane isomerization over sulfated zirconia solid acid system. Int. J. Energ. Res. 2020, 44, 3270–3294. [Google Scholar] [CrossRef]
  28. Rabee, A.I.M.; Mekhemer, G.A.H.; Osatiashtiani, A.; Isaacs, M.A.; Lee, A.F.; Wilson, K.; Zaki, M.I. Acidity-reactivity relationships in catalytic esterification over ammonium sulfate-derived sulfated zirconia. Catalysts 2017, 7, 204. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, P.; Zhang, J.; Han, C.; Yang, C.; Li, C. Effect of Modification Methods on the Surface Properties and n-Butane Isomerization Performance of La/Ni-Promoted SO42−/ZrO2-Al2O3. Appl. Surf. Sci. 2016, 378, 489–495. [Google Scholar] [CrossRef]
  30. Saravanan, K.; Tyagi, B.; Bajaj, H.C. Esterifcation of caprylic acid with alcohol over nano-crystalline sulfated zirconia. J. Sol-Gel Sci. Technol. 2012, 62, 13–17. [Google Scholar] [CrossRef]
  31. Utami, M.; Wijaya, K.; Trisunaryanti, W. Effect of Sulfuric Acid Treatment and Calcination on Commercial Zirconia Nanopowder. Key Eng. Mater. 2017, 757, 131–137. [Google Scholar] [CrossRef]
  32. Yang, X.; Ma, X.; Yu, X.; Ge, M. Exploration of strong metal-support interaction in zirconia supported catalysts for toluene oxidation. App. Catal. B Environ. 2020, 263, 118355. [Google Scholar] [CrossRef]
  33. Mirzan, M.; Syoufian, A.; Wijaya, K. Physico-chemical properties of nano ZrO2-pillared bentonite with nickel as supporting metal. Nano Hybrids Compos. 2020, 30, 9–17. [Google Scholar] [CrossRef]
  34. Loveless, B.T.; Gyanani, A.; Muggli, D.S. Discrepancy between TPD- and FTIR- based measurements of Brønsted and Lewis acidity for sulfated zirconia. Appl. Catal. B Environ. 2008, 84, 591–597. [Google Scholar] [CrossRef]
  35. Li, L.; Yan, B.; Li, H.; Yu, S.; Liu, S.; Yu, H.; Ge, X. SO42−/ZrO2 as catalyst for upgrading of pyrolysis oil by esterification. Fuel 2018, 226, 190–194. [Google Scholar] [CrossRef]
  36. Pratika, R.A.; Wijaya, K.; Trisunaryanti, W. Hydrothermal treatment of SO4/TiO2 and TiO2/CaO as heterogeneous catalysts for the conversion of Jatropha oil into biodiesel. J. Environ. Chem. Eng. 2021, 9, 106547. [Google Scholar] [CrossRef]
  37. Utami, M.; Safitri, R.; Pradipta, M.F.; Wijaya, K.; Chang, S.W.; Ravindran, B.; Ovi, D.; Rajabathar, J.R.; Poudineh, N.; Gengan, R.M. Enhanced catalytic conversion of palm oil into biofuels by Cr-incorporated sulphated zirconia. Mater. Lett. 2022, 309, 131472. [Google Scholar] [CrossRef]
  38. Hanifah, A.; Nadia, A.; Saputri, W.D.; Syoufian, A.; Wijaya, K. Performance of Ni-Mo sulfated nanozirconia catalyst for conversion of waste cooking oil into biofuel via hydrocracking process. Mater. Sci. Forum 2021, 1045, 79–89. [Google Scholar] [CrossRef]
  39. Sari, E.P.; Wijaya, K.; Trisunaryanti, W.; Syoufian, A.; Hasanudin, H.; Saputri, W.D. The effective combination of zirconia superacid and zirconia-imregnated CaO in biodiesel manufacturing: Utilization of used coconut cooking oil (UCCO). Int. J. Energ. Environ. Eng. 2021, 1–12. [Google Scholar] [CrossRef]
  40. Wijaya, K.; Nadia, A.; Dinana, A.; Pratiwi, A.F.; Tikoalu, A.D.; Wibowo, A.C. Catalytic hydrocracking of fresh and waste frying oil over Ni- and Mo-based catalysts supported on sulfated silica for biogasoline production. Catalysts 2021, 11, 1150. [Google Scholar] [CrossRef]
  41. Fernandez-Morales, J.M.; Castillejos, E.; Asedegbega-Nieto, E.; Dongil, A.B.; Rodrigues-Ramos, I.; Guerrero-Ruiz, A. Comparative study of different acidic surface structures in solid catalysts applied for the isobutene dimerization reaction. Nanomaterial 2020, 10, 1235. [Google Scholar] [CrossRef]
  42. Purba, S.E.; Wijaya, K.; Trisunaryanti, W.; Pratika, R.A. Dealuminated and desilicated natural zeolite as a catalyst for hydrocracking of used cooking oil into biogasoline. Mediterr. J. Chem. 2020, 11, 74–83. [Google Scholar] [CrossRef]
  43. Wijaya, K.; Malau, M.L.L.; Utami, M.; Mulijani, S.; Patah, A.; Wibowo, A.C.; Chandrasekaran, M.; Rajabathar, J.R.; Al-Lohedan, H. Synthesis, characterizations and catalysis of sulfated silica and nickel modified silica catalysts for diethyl ether (DEE) production from ethanol towards renewable energy applications. Catalysts 2021, 11, 1511. [Google Scholar] [CrossRef]
  44. Serrano, D.P.; Garcia, R.A.; Linares, M.; Gil, B. Influence of the calcination treatment on the catalytic properties of hierarchical ZSM-5. Catal. Today 2012, 179, 91–101. [Google Scholar] [CrossRef]
  45. Swamidoss, C.M.A.; Sheraz, M.; Anus, A.; Jeong, S.; Park, Y.-K.; Kim, Y.-M.; Kim, S. Effect of Mg/Al2O3 and calcination temperature on the catalytic decomposition of HFC-134a. Catalysts 2019, 9, 270. [Google Scholar] [CrossRef] [Green Version]
  46. Abedin, M.A.; Kanitkar, S.; Bhattar, S.; Spivey, J.J. Methane dehydroaromatization using Mo supported on sulfated zirconia catalyst: Effect of promoters. Catal. Today 2021, 365, 71–79. [Google Scholar] [CrossRef]
  47. Susi, E.P.; Wijaya, K.; Wangsa; Pratika, R.A.; Hariani, P.L. Effect of nickel concentration in natural zeolite as catalyst in hydrocracking process of used cooking oil. Asian J. Chem. 2020, 32, 2773–2777. [Google Scholar] [CrossRef]
  48. Yin, P.; Hu, S.; Qian, K.; Wei, Z.; Zhang, L.-L.; Lin, Y.; Huang, W.; Xiong, H.; Li, W.X.; Liang, H.-W. Quantification of critical particle distance for mitigating catalyst sintering. Nat. Commun. 2021, 12, 4865. [Google Scholar] [CrossRef]
  49. Prašnikar, A.; Pavlišič, A.; Ruiz-Zepeda, F.; Kovač, J.; Likozar, B. Mechanism of Copper-based catalyst deactivation during CO2 reduction to methanol. Ind. Eng. Chem. Res. 2019, 58, 13021–13029. [Google Scholar] [CrossRef] [Green Version]
  50. Rahmati, M.; Safdari, M.S.; Fletcher, T.H.; Argyle, M.D.; Bartholomew, C.H. Chemical and thermal sintering of supported metals with emphasis on cobalt catalysts during Fischer–Tropsch synthesis. Chem. Rev. 2020, 120, 4455–4533. [Google Scholar] [CrossRef]
  51. Zhou, S.; Song, Y.; Zhao, J.; Zhou, X.; Chen, L. Study on the Mechanism of Water Poisoning Pt-Promoted Sulfated Zirconia Alumina in n-Hexane Isomerization. Energy Fuels 2021, 35, 14860–14867. [Google Scholar] [CrossRef]
  52. Peng, S.; Li, M.; Yang, X.; Li, P.; Liu, H.; Xiong, W.; Peng, X. Atomic layer deposition of Pt nanoparticles on ZrO2 based metal-organic frameworks for increased photocatalytic activity. Ceram. Int. 2019, 45, 18128–18134. [Google Scholar] [CrossRef]
  53. Bikmetova, L.I.; Kazantsev, K.V.; Zatolokina, L.V.; Smolikov, M.D.; Belyi, A.S. A study on the synthesis steps of alumina-supported Pt/SO4/ZrO2 catalysts for isomerization of n-hexane. AIP Conf. Proc. 2020, 2285, 020003. [Google Scholar] [CrossRef]
  54. Souza, I.C.A.; Manfro, R.L.; Souza, M.M.V.M. Hydrogen production from steam reforming of acetic acid over Pt-Ni bimetallic catalysts supported on ZrO2. Biomass Bioenergy 2022, 156, 106317. [Google Scholar] [CrossRef]
  55. Ullah, I.; Taha, T.A.; Alenad, A.M.; Uddin, I.; Hayat, A.; Hayat, A.; Sohail, M.; Irfan, A.; Khan, J.; Palamanit, A. Platinum-alumina modified SO42−ZrO2/Al2O3 based bifunctional catalyst for significantly improved n-butane isomerization performance. Surf. Interfaces 2021, 25, 101227. [Google Scholar] [CrossRef]
  56. Tamizhdurai, P.; Lavanya, M.; Meenakshisundaram, A.; Shanti, K.; Sivasanker, S. Isomerization of alaknes over Pt-sulphated zirconia supported on SBA-15. Adv. Porous Mater. 2017, 5, 169–174. [Google Scholar] [CrossRef]
  57. Cho, H.J.; Kim, D.; Li, S.; Su, D.; Ma, D.; Xu, B. Molecular-level proximity of metal and acid sites in zeolite-encapsulated Pt nanoparticles for selective multistep tandem catalysis. ACS Catal. 2020, 10, 3340–3348. [Google Scholar] [CrossRef]
  58. Gan, T.; Yang, J.; Morris, D.; Chu, X.; Zhang, P.; Zhang, W.; Zuo, Y.; Yan, W.; Wei, S.-H.; Liu, G. Electron donation of non-oxide supports boosts O2 activation on nano-platinum catalysts. Nat. Commun. 2021, 12, 2741. [Google Scholar] [CrossRef] [PubMed]
  59. Aboul-Gheit, A.K.; Gad, F.K.; Abdel-Aleem, G.M.; El-Desouki, D.S.; Abdel Hamid, S.M.; Ghoneim, S.S.; Ibrahim, A.H. Pt, Re and Pt-Re incorporation in sulfated zirconia as catalysts for n-pentane isomerization. Egypt. J. Pet. 2014, 23, 303–314. [Google Scholar] [CrossRef] [Green Version]
  60. Zhao, P.-P.; Chen, J.; Yu, H.-B.; Cen, B.-H.; Wang, W.-Y.; Luo, M.-F.; Lu, J.-Q. Insight into propane combustion over MoO3 promoted Pt/ZrO2 catalysts: The generation of Pt-MoO3 interface and its promotional role on catalytic activity. J. Catal. 2020, 391, 80–90. [Google Scholar] [CrossRef]
  61. Kondratowicz, T.; Drozdek, M.; Michalik, M.; Gac, W.; Gajewska, M.; Kustroski, P. Catalytic activity of Pt species variously dispersed on hollow ZrO2 spheres in combustion of volatile organic compounds. App. Surf. Sci. 2020, 513, 145788. [Google Scholar] [CrossRef]
  62. Verga, L.G.; Rusell, A.E.; Skylaris, C.K. Ethanol, O, and CO adsorption on Pt nanoparticles: Effect of nanoparticle size and graphene support. Phys. Chem. Chem. Phys. 2018, 20, 25918–25930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Wang, S.; Pu, J.; Wu, J.; Liu, H.; Xu, H.; Li, X.; Wang, H. SO42−/ZrO2 as a solid acid for the esterification of palmitic acid with methanol: Effects of the calcination time and recycle method. ACS Omega 2020, 5, 30139–30147. [Google Scholar] [CrossRef] [PubMed]
  64. Ren, K.; Kong, D.; Meng, X.; Wang, X.; Shi, L.; Liu, N. The effect of ammonium sulfate and sulfamic acid on the surface acidity of sulfated zirconia. J. Saudi Chem. Soc. 2019, 23, 198–204. [Google Scholar] [CrossRef]
  65. Vance, B.C.; Kots, P.A.; Wang, C.; Hinton, Z.R.; Quinn, C.M.; Epps, T.H., III; Korley, L.T.J.; Vlachos, D.G. Single pot catalyst strategy to branched products via adhesive isomerization and hydrocracking of polyethylene over platinum tungstated zirconia. App. Catal. B. Environ. 2021, 299, 120483. [Google Scholar] [CrossRef]
  66. Liu, S.; Lots, P.; Vance, B.C.; Danielson, A.; Vlachos, D.G. Plastic waste to fuels by hydrocracking at mild conditions. Sci. Adv. 2021, 7, eabf8283. [Google Scholar] [CrossRef] [PubMed]
  67. Chen, J.; Zhang, Y.; Chen, X.; Dai, S.; Bao, Z.; Yang, Q.; Ren, Q.; Zhang, Z. Cooperative interplay of Bronsted acid and Lewis acid sites in MIL-101 (Cr) for cross-dehydrogenative coupling of C–H. Bonds. ACS Appl. Mater. Interfaces 2021, 13, 10845–10854. [Google Scholar] [CrossRef] [PubMed]
  68. Hu, Z.-P.; Wang, Z.; Yuan, Z.-Y. Cr/Al2O3 catalysts with strong metal-support interactions for stable catalytic dehydrogenation of propane to propylene. Mol. Catal. 2020, 493, 111052. [Google Scholar] [CrossRef]
  69. Annuar, N.H.R.; Triwahyono, S.; Jalil, A.A.; Basar, N.; Abdullah, T.A.T.; Ahmad, A. Effect of Cr2O3 loading on the properties and cracking activity of Pt/Cr2O3-ZrO2. App. Catal. A Gen. 2017, 541, 77–86. [Google Scholar] [CrossRef]
  70. Wang, H.; Lin, N.; Xu, R.; Yu, Y.; Zhao, X. First principles studies of electronic, mechanical and optical properties of Cr-doped cubic ZrO2. Chem. Phys. 2020, 539, 110972. [Google Scholar] [CrossRef]
  71. Hauli, L. Chromium/Sulfated Nanozirconia: Preparation, Characterization, and Their Application in Conversion of LDPE Plastic Waste into Liquid Fuels. Ph.D. Thesis, Department of Chemistry, Universitas Gadjah Mada, Yogyakarta, Indonesia, 2019. [Google Scholar]
  72. Hauli, L.; Wijaya, K.; Armunanto, R. Preparation of Cr metal supported on sulfated zirconia catalyst. Mater. Sci. Forum 2019, 948, 221–227. [Google Scholar] [CrossRef]
  73. Aziz, I.T.A.; Saputri, W.D.; Trisunaryanti, W.; Sudiono, S.; Syoufian, A.; Budiman, A.; Wijaya, K. Synthesis of Nickel-loaded sulfated zirconia catalyst and its application for converting used palm cooking oil to gasoline via hydrocracking process. Period. Polytech. Chem. Eng. 2021, 66, 101–113. [Google Scholar] [CrossRef]
  74. Tao, Y.; Zhu, Y.; Liu, C.; Yue, H.; Ji, J.; Yuan, S.; Jiang, W.; Liang, B. A highly selective Cr/ZrO2 catalyst for the reverse water-gas shift reaction prepared from simulated Cr-containing wastewater by a photocatalytic deposition process with ZrO2. J. Environ. Chem. Eng. 2018, 6, 6761–6770. [Google Scholar] [CrossRef]
  75. Serrano, D.; Escola, J.; Briones, L.; Arroyo, M. Hydroprocessing of the LDPE thermal cracking oil into transportation fuels over Pd supported on hierarchical ZSM-5 catalyst. Fuel 2017, 206, 190–198. [Google Scholar] [CrossRef]
  76. Wong, S.L.; Ngadi, N.; Abdullah, T.A.T.; Inuwa, I.M. Conversion of low-density polyethylene (LDPE) over ZSM-5 zeolite to liquid fuel. Fuel 2017, 192, 71–82. [Google Scholar] [CrossRef]
  77. Lyu, Y.; Xu, R.; Williams, O.; Wang, Z.; Sievers, C. Reaction paths of methane activation and oxidation of surface intermediates over NiO on Ceria-Zirconia catalysts studied by In-situ FTIR spectroscopy. J. Catal. 2021, 404, 334–347. [Google Scholar] [CrossRef]
  78. Dahdah, E.; Estephane, J.; Gennequin, C.; Aboukais, A.; Abi-Aad, E.; Aouad, S. Zirconia supported nickel catalysts for glycerol steam reforming: Effect of zirconia structure on the catalytic performance. Int. J. Hydrogen Energy 2020, 45, 4457–4467. [Google Scholar] [CrossRef]
  79. Kou, J.; Yi, L.; Li, G.; Cheng, K.; Wang, R.; Zhnag, D.; Jin, H.; Guo, L. Structural effect of ZrO2 on supported Ni-based catalysts for supercritical water gasification of oil-containing water. Int. J. Hydrogen Energ. 2021, 46, 12874–12885. [Google Scholar] [CrossRef]
  80. Li, X.; Shao, Y.; Zhang, S.; Wang, Y.; Xiang, J.; Hu, S.; Xu, L.; Hu, X. Pore diameters of Ni/ZrO2 catalysts affect properties of the coke in steam reforming of acetic acid. Int. J. Hydrogen Energy 2021, 46, 23642–23657. [Google Scholar] [CrossRef]
  81. Kou, J.; Feng, H.; Wei, W.; Wang, G.; Sun, J.; Jin, H.; Guo, L. Study on the detailed reaction pathways and catalytic mechanism of a Ni/ZrO2 catalyst for supercritical water gasification of diesel oil. Fuel 2022, 312, 122849. [Google Scholar] [CrossRef]
  82. Papageridis, K.N.; Charisiou, N.D.; Douvartzides, S.L.; Sebastian, V.; Hinder, S.J.; Baker, M.A.; Alkhoori, S.; Polychronopoulou, K.; Goula, M.A. Effect of operating parameters on the selective catalytic deoxygenation of palm oil to produce renewable diesel over Ni supported on Al2O3, ZrO2, and SiO2 catalysts. Fuel Process. Technol. 2020, 209, 106547. [Google Scholar] [CrossRef]
  83. Munir, M.; Ahmad, M.; Saeed, M.; Waseem, A.; Nizami, A.-S.; Sultana, S.; Zafar, M.; Rehan, M.G.; Srinivasan, G.R.; Ali, A.M.; et al. Biodeisel production from novel non-edible caper (Capparis spinosa L.) seeds oil employing Cu–Ni doped ZrO2 catalyst. Renew. Sustain. Energy Rev. 2021, 138, 110558. [Google Scholar] [CrossRef]
Figure 1. Crystal structures of (a) monoclinic, (b) tetragonal, and (c) cubic ZrO2. Copyright Elsevier, reprinted from Ref. [23].
Figure 1. Crystal structures of (a) monoclinic, (b) tetragonal, and (c) cubic ZrO2. Copyright Elsevier, reprinted from Ref. [23].
Catalysts 12 00191 g001
Figure 2. Brønsted and Lewis acid sites on ZrO2-SO4. Copyright MDPI, reprinted from Ref. [17].
Figure 2. Brønsted and Lewis acid sites on ZrO2-SO4. Copyright MDPI, reprinted from Ref. [17].
Catalysts 12 00191 g002
Figure 3. ZrO2-SO4 surface model of bidentate chelate type. Reprinted from ref. [29].
Figure 3. ZrO2-SO4 surface model of bidentate chelate type. Reprinted from ref. [29].
Catalysts 12 00191 g003
Figure 4. FTIR spectra of SZ catalyst at various sulfate concentrations and calcination temperatures. Reprinted with permission from Dr. Utami, Ref. [31]. Copyright 2019 Trans Tech Publication.
Figure 4. FTIR spectra of SZ catalyst at various sulfate concentrations and calcination temperatures. Reprinted with permission from Dr. Utami, Ref. [31]. Copyright 2019 Trans Tech Publication.
Catalysts 12 00191 g004
Figure 5. FTIR spectra of SZ catalyst at various sulfate concentrations and calcination temperatures after acidity test. Reprinted with permission from Dr. Utami, Ref. [31]. Copyright 2019 Trans Tech Publication.
Figure 5. FTIR spectra of SZ catalyst at various sulfate concentrations and calcination temperatures after acidity test. Reprinted with permission from Dr. Utami, Ref. [31]. Copyright 2019 Trans Tech Publication.
Catalysts 12 00191 g005
Figure 6. Diffraction patterns of SZ catalyst at various sulfate concentrations and calcination temperatures. Reprinted with permission from Dr. Utami, Ref. [31]. Copyright 2019 Trans Tech Publication.
Figure 6. Diffraction patterns of SZ catalyst at various sulfate concentrations and calcination temperatures. Reprinted with permission from Dr. Utami, Ref. [31]. Copyright 2019 Trans Tech Publication.
Catalysts 12 00191 g006
Figure 7. Mechanism of H2 distribution through the Pt surface on ZrO2-SO4. Reprinted and modified with permission from Dr. Utami, Ref. [9]. Copyright 2019 The Royal Society of Chemistry.
Figure 7. Mechanism of H2 distribution through the Pt surface on ZrO2-SO4. Reprinted and modified with permission from Dr. Utami, Ref. [9]. Copyright 2019 The Royal Society of Chemistry.
Catalysts 12 00191 g007
Figure 8. ZrO2-SO4 (right) and Pt/ZrO2-SO4 (left) catalyst. Reprinted with permission from Dr. Utami, Ref. [42].
Figure 8. ZrO2-SO4 (right) and Pt/ZrO2-SO4 (left) catalyst. Reprinted with permission from Dr. Utami, Ref. [42].
Catalysts 12 00191 g008
Figure 9. FTIR spectra of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ. Reprinted with permission from Dr. Utami, Ref. [6]. Copyright 2019 Elsevier.
Figure 9. FTIR spectra of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ. Reprinted with permission from Dr. Utami, Ref. [6]. Copyright 2019 Elsevier.
Catalysts 12 00191 g009
Figure 10. FTIR spectra of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ after acidity test. Reprinted with permission from Dr. Utami, Ref. [6]. Copyright 2019 Elsevier.
Figure 10. FTIR spectra of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ after acidity test. Reprinted with permission from Dr. Utami, Ref. [6]. Copyright 2019 Elsevier.
Catalysts 12 00191 g010
Figure 11. Diffraction patterns of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ. Reprinted with permission from Dr. Utami, Ref. [6]. Copyright 2019 Elsevier.
Figure 11. Diffraction patterns of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ. Reprinted with permission from Dr. Utami, Ref. [6]. Copyright 2019 Elsevier.
Catalysts 12 00191 g011
Figure 12. XPS spectra of Pt/nano ZrO2-SO4 in the 4f region and the catalyst model. Reprinted and modified with permission from Dr. Utami, Ref. [9]. Copyright 2019 The Royal Society of Chemistry.
Figure 12. XPS spectra of Pt/nano ZrO2-SO4 in the 4f region and the catalyst model. Reprinted and modified with permission from Dr. Utami, Ref. [9]. Copyright 2019 The Royal Society of Chemistry.
Catalysts 12 00191 g012
Figure 13. TGA curves of (a) nano ZrO2, (b) nano ZrO2-SO4, (c) Pt/nano ZrO2-SO4. Reprinted and modified with permission from Dr. Utami, Ref. [9]. Copyright 2019 The Royal Society of Chemistry.
Figure 13. TGA curves of (a) nano ZrO2, (b) nano ZrO2-SO4, (c) Pt/nano ZrO2-SO4. Reprinted and modified with permission from Dr. Utami, Ref. [9]. Copyright 2019 The Royal Society of Chemistry.
Catalysts 12 00191 g013
Figure 14. The physical appearance of the hydrocracking liquid fraction (a) without catalyst, and with catalyst (b) ZrO2, (c) SZ, (d) Pt1/SZ, (e) Pt2/SZ3, (f) Pt3/SZ. Reprinted with permission from Dr. Utami, Ref. [14].
Figure 14. The physical appearance of the hydrocracking liquid fraction (a) without catalyst, and with catalyst (b) ZrO2, (c) SZ, (d) Pt1/SZ, (e) Pt2/SZ3, (f) Pt3/SZ. Reprinted with permission from Dr. Utami, Ref. [14].
Catalysts 12 00191 g014
Figure 15. The composition of liquid yield in the gasoline fraction from the hydrocracking reaction of LDPE plastic waste using Pt3/SZ at 250 °C. Reprinted with permission from Dr. Utami, Ref. [6]. Copyright 2019 Elsevier.
Figure 15. The composition of liquid yield in the gasoline fraction from the hydrocracking reaction of LDPE plastic waste using Pt3/SZ at 250 °C. Reprinted with permission from Dr. Utami, Ref. [6]. Copyright 2019 Elsevier.
Catalysts 12 00191 g015
Figure 16. Hydrocracking liquid fraction graphs of (a) nano ZrO2, (b) nano ZrO2-SO4, and (c) Pt/nano ZS3 Pt/nanoZS3-600. Reprinted and modified with permission from Dr. Utami, Ref. [9]. Copyright 2019 The Royal Society of Chemistry.
Figure 16. Hydrocracking liquid fraction graphs of (a) nano ZrO2, (b) nano ZrO2-SO4, and (c) Pt/nano ZS3 Pt/nanoZS3-600. Reprinted and modified with permission from Dr. Utami, Ref. [9]. Copyright 2019 The Royal Society of Chemistry.
Catalysts 12 00191 g016
Figure 17. Proposed model of ammonia interaction on metal-embedded sulfated zirconia. Reprinted from Ref. [68]. Copyright Elsevier.
Figure 17. Proposed model of ammonia interaction on metal-embedded sulfated zirconia. Reprinted from Ref. [68]. Copyright Elsevier.
Catalysts 12 00191 g017
Figure 18. ZrO2-SO4 (left) and Cr/ZrO2-SO4 (right) catalyst. Reprinted with permisission from Dr. Hauli, Ref. [71].
Figure 18. ZrO2-SO4 (left) and Cr/ZrO2-SO4 (right) catalyst. Reprinted with permisission from Dr. Hauli, Ref. [71].
Catalysts 12 00191 g018
Figure 19. FTIR spectra of ZrO2, SZ, Cr1/SZ, Cr2/SZ, Cr3/SZ. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Figure 19. FTIR spectra of ZrO2, SZ, Cr1/SZ, Cr2/SZ, Cr3/SZ. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Catalysts 12 00191 g019
Figure 20. FTIR spectra of SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ after acidity test. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Figure 20. FTIR spectra of SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ after acidity test. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Catalysts 12 00191 g020
Figure 21. Diffraction patterns of ZrO2, SZ, Cr1/SZ, Cr2/SZ, Cr3/SZ. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Figure 21. Diffraction patterns of ZrO2, SZ, Cr1/SZ, Cr2/SZ, Cr3/SZ. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Catalysts 12 00191 g021
Figure 22. Adsorption desorption isotherm patterns of ZrO2, SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Figure 22. Adsorption desorption isotherm patterns of ZrO2, SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Catalysts 12 00191 g022
Figure 23. Liquid product selectivity of ZrO2, SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ catalysts. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Figure 23. Liquid product selectivity of ZrO2, SZ, Cr1/SZ, Cr2/SZ, and Cr3/SZ catalysts. Reprinted with permission from Dr. Hauli, Ref. [72]. Copyright 2019 Trans Tech Publication.
Catalysts 12 00191 g023
Figure 24. FTIR spectra of SZ (a), Ni-SZ 1 (b), Ni-SZ 2 (c), and Ni-SZ 3 (d) catalysts. Reprinted with permission from Aziz, Ref. [73]. Copyright 2020 Budapest University.
Figure 24. FTIR spectra of SZ (a), Ni-SZ 1 (b), Ni-SZ 2 (c), and Ni-SZ 3 (d) catalysts. Reprinted with permission from Aziz, Ref. [73]. Copyright 2020 Budapest University.
Catalysts 12 00191 g024
Figure 25. Potential application of sulfated nanozirconia catalysts.
Figure 25. Potential application of sulfated nanozirconia catalysts.
Catalysts 12 00191 g025
Table 1. Acidity test for the SZ catalyst at various sulfate concentrations and calcination temperatures [31].
Table 1. Acidity test for the SZ catalyst at various sulfate concentrations and calcination temperatures [31].
CatalystAcidity (mmol/g)
600 °C700 °C800 °C900 °C
ZrO20.18---
SZ-0.20.390.320.290.26
SZ-0.50.830.330.300.28
SZ-0.81.060.560.530.33
Table 2. Acidity test results of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ [6].
Table 2. Acidity test results of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ [6].
SampleAcidity (mmol/g)
ZrO20.18
SZ1.06
Pt1/SZ10.75
Pt2/SZ11.05
Pt3/SZ11.14
Table 3. Textural characteristics of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ [6].
Table 3. Textural characteristics of ZrO2, SZ, Pt1/SZ, Pt2/SZ, and Pt3/SZ [6].
SampleSurface Area (m2/g)Pore Volume (cm3/g)Pore Diameter (nm)
ZrO222.050.083.49
SZ12.490.053.82
Pt1/SZ13.490.053.83
Pt2/SZ20.230.063.84
Pt3/SZ29.480.083.86
Table 4. Crystal sizes of ZrO2, SZ, Pt1/SZ, Pt3/SZ Pt2/SZ and Pt3/SZ [6].
Table 4. Crystal sizes of ZrO2, SZ, Pt1/SZ, Pt3/SZ Pt2/SZ and Pt3/SZ [6].
SampleCrystal Size (nm)
ZrO231.54
SZ34.06
Pt1/SZ32.80
Pt2/SZ31.54
Pt3/SZ31.53
Table 5. Elemental compositions of ZrO2, SZ, Pt1/SZ, Pt2/SZ dan Pt3/SZ [6].
Table 5. Elemental compositions of ZrO2, SZ, Pt1/SZ, Pt2/SZ dan Pt3/SZ [6].
SampleElemental Compositions (% w/w)
ZrPtOS
ZrO264.08-35.460.46
SZ60.51-37.422.07
Pt1/SZ75.240.3524.020.74
Pt2/SZ74.370.9025.020.61
Pt3/SZ75.341.1924.400.26
Table 6. Percentage of relative areas of Pt 4f deconvoluted peaks [4].
Table 6. Percentage of relative areas of Pt 4f deconvoluted peaks [4].
Pt 4f PeakPeak Position (eV)Relative Area (%)
Pt0 4f7/271.4536.36
Pt0 4f5/274.7545.46
Pt2+4f7/273.0818.18
Table 7. Distribution of hydrocracking products using different catalysts (T = 250 °C, t = 60 m, catalyst to feed ratio = 1% w/w) [6].
Table 7. Distribution of hydrocracking products using different catalysts (T = 250 °C, t = 60 m, catalyst to feed ratio = 1% w/w) [6].
SampleHydrocracking Product (% w/w)
LiquidSolidGas
C5–C12C13–C20
SZ42.9514.970.3541.73
Pt1/SZ48.7621.820.1529.27
Pt2/SZ64.226.930.1628.69
Pt3/SZ67.517.090.1525.25
Table 8. Acidity values of the catalysts [72].
Table 8. Acidity values of the catalysts [72].
SampleAcidity (mmol/g)
SZ3.81
Cr1/SZ6.24
Cr2/SZ8.22
Cr3/SZ6.75
Table 9. Pore characteristics of the catalysts [72].
Table 9. Pore characteristics of the catalysts [72].
SampleSurface Area (m2/g)Pore Diameter (nm)Pore Volume (cm3/g)
ZrO212.273.700.07
SZ7.793.680.08
Cr1/SZ12.623.700.12
Cr2/SZ14.563.700.08
Cr3/SZ11.919.040.04
Table 10. Distribution of LDPE plastic hydrocrack products on various catalysts [72].
Table 10. Distribution of LDPE plastic hydrocrack products on various catalysts [72].
CatalystProduct Conversion (% b/b)
LiquidCokeGas
ZrO217.390.3632.23
SZ28.720.3429.51
Cr1/SZ33.480.0125.78
Cr2/SZ40.990.0124.84
Cr3/SZ37.510.0126.77
Table 11. The acidity and pore characteristics of the Ni-SZ catalysts [73].
Table 11. The acidity and pore characteristics of the Ni-SZ catalysts [73].
SampleAciditySurface Area (m2/g)Pore Diameter
(nm)
Pore Volume (cm3/g)
ZrO20.3555.6330.720.04
SZ2.0044.0435.840.04
Ni-SZ 13.9057.093.820.06
Ni-SZ 24.0618.753.820.05
Ni-SZ 34.23511.683.820.05
Table 12. Selectivity of ZrO2, SZ, Ni-SZ 1, Ni-SZ 2, and Ni-SZ 3 catalysts in the hydrocracking reaction of used cooking oil into liquid fuel [73].
Table 12. Selectivity of ZrO2, SZ, Ni-SZ 1, Ni-SZ 2, and Ni-SZ 3 catalysts in the hydrocracking reaction of used cooking oil into liquid fuel [73].
CatalystSelectivity (wt%)
(C1–C4)(C5–C12)(C13–C20)Non-Hydrocarbon
ZrO21.8677.3119.261.57
SZ1.9367.8328.152.54
Ni-SZ 10.0051.5915.1633.25
Ni-SZ 20.0087.0112.990.00
Ni-SZ 30.00100.000.000.00
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sekewael, S.J.; Pratika, R.A.; Hauli, L.; Amin, A.K.; Utami, M.; Wijaya, K. Recent Progress on Sulfated Nanozirconia as a Solid Acid Catalyst in the Hydrocracking Reaction. Catalysts 2022, 12, 191. https://doi.org/10.3390/catal12020191

AMA Style

Sekewael SJ, Pratika RA, Hauli L, Amin AK, Utami M, Wijaya K. Recent Progress on Sulfated Nanozirconia as a Solid Acid Catalyst in the Hydrocracking Reaction. Catalysts. 2022; 12(2):191. https://doi.org/10.3390/catal12020191

Chicago/Turabian Style

Sekewael, Serly Jolanda, Remi Ayu Pratika, Latifah Hauli, Amalia Kurnia Amin, Maisari Utami, and Karna Wijaya. 2022. "Recent Progress on Sulfated Nanozirconia as a Solid Acid Catalyst in the Hydrocracking Reaction" Catalysts 12, no. 2: 191. https://doi.org/10.3390/catal12020191

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop