Next Article in Journal
Zn-Cr Layered Double Hydroxides for Photocatalytic Transformation of CO2 under Visible Light Irradiation: The Effect of the Metal Ratio and Interlayer Anion
Next Article in Special Issue
Promoter Effect on Carbon Nanosphere-Encapsulated Fe-Co Catalysts for Converting CO2 to Light Olefins
Previous Article in Journal
Study on the Hydrogenation of Ethyl Stearate to the Fatty Alcohol 1-Octadecanol over Ru on Tungstated Zirconia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hydroisomerization Catalysts for High-Quality Diesel Fuel Production

1
Physical and Colloidal Chemistry Department, Faculty of Chemical Technology and Ecology, Gubkin University, 119991 Moscow, Russia
2
Petro-Processes Engineering Department, Faculty of Petrochemical Engineering, Alfurat University, Deir ez-Zur 84C7+XHV, Syria
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(10), 1363; https://doi.org/10.3390/catal13101363
Submission received: 19 September 2023 / Revised: 5 October 2023 / Accepted: 9 October 2023 / Published: 11 October 2023
(This article belongs to the Special Issue Feature Papers in "Industrial Catalysis" Section)

Abstract

:
Upgrading the properties of diesel fractions is considered one of the crucial processes in the petrochemical industry; and for this purpose in laboratory-scale researching it is studied on the base of the hydroisomerization of n-hexadecane as a main model reaction. Recently, zeolite-based bifunctional catalysts have proven their efficiency due to their remarkable acidity, shape-selectivity and relative resistance to deactivation. In this review, different topological-type zeolite-based catalysts, the mechanism of their catalytic effect in n-C16 isomerization, and the principles of shape-selectivity are reviewed. A comparison of their structural-operational characteristics is made. The impact of some feedstock impurities on the catalyst’s performance and deactivation due to carbonaceous deposits as well as various modern eco-friendly cost-effective synthesis techniques are also discussed.

1. Introduction

Currently, a key focus in the global production of petroleum products is to raise the standards for motor fuels. The inadequate quality of such fuels is a leading factor contributing to environmental pollution. To address this issue, improving the chemical composition of the fuel is used to meet high quality requirements [1,2]. One of the most promising technologies is isodewaxing process. Instead of removing n-paraffins by solvent extraction, the is the odewaxing technology modifies the molecular structure of the wax through catalytic isomerization, producing isoparaffins with superior cold flow properties, high viscosity index, and excellent oxidation resistance.
The isodewaxing process comprises four reactions: (i) isomerization of the normal paraffins to obtain their isomers, (ii) hydrocracking of the long-chain paraffins into shorter-chain ones, (iii) hydrogenation of the unsaturated aromatic components, and (iv) hydrodecyclization with the ring-opening reaction. In this review we will focus on the first type of reaction, which usually occurs in the presence of hydrogen, so it is referred to as hydroisomerization. Usually, this reaction is accompanied by hydrocracking in which the extent varies according to the operational conditions of the process and the characteristics of the catalyst used. The rates of hydroisomerization and hydrocracking reactions define the composition of products.
Considerable research efforts have been dedicated towards the production of value-added products from low-grade waxy feedstocks (Figure 1). In particular, the focus has been on converting these feedstocks into branched products to improve the fuel quality. These researches have been centered on catalysts development, with a specific emphasis on micro/mesoporous catalysts to accelerate mass transfer and optimize the residence time of reagents on the active sites, thereby increasing their efficiency in the desired chemical reactions [3,4,5,6].
Hydroisomerization catalysts are generally categorized as bifunctional catalysts, composed of a metal, typically platinum, that is dispersed on an acidic support. The metal serves as a dehydrogenation/hydrogenation agent, while the acidic support aids in the skeletal isomerization reaction of the olefinic intermediates created over the metal sites and in cracking [7,8]. A well-balanced bifunctional catalyst promotes hydrogenation and dehydrogenation reactions, with rearrangements of the hydrocarbon intermediates over acid sites dictating reaction rates [9]. As a result, hydroisomerization and hydrocracking reactions occur in sequence, leading to mono- and multibranched isomers, and cracked paraffins [10,11]. One of the primary challenges in advancing the hydroisomerization technology is the need for highly effective and selective bifunctional catalysts that can minimize the formation of byproducts resulting from hydrocracking [12,13].
In this review, the latest findings in improving the isomerization processes of linear alkanes through the development of catalyst structures, methods of their preparation, and studies related to the operational conditions of the process will be discussed. A comprehensive analysis will be conducted to compare the performance of various catalysts during the hydroisomerization of n-C16. The operating conditions of the hydroisomerization of n-C16 will be discussed in details.

2. Straight-Chain Alkane Hydroisomerization Process

2.1. The Principle of the Process and Its Technical Significance

Hydroisomerization of diesel fraction is considered as a vital chemical process that includes the hydroconversion of normal paraffins into their branched isomers. This reaction comprises breaking and rearranging C-C bonds to obtain more favorable isomers providing the higher cetane number of diesel, thus reducing engine emissions and improving the cold flow characteristics of fuels, making them less prone to clogging at low temperatures.
As shown in Table 1, the presence of linear alkanes improves the cetane number of diesel, however, the melting point also increases. Given the importance of diesel quality, maintaining an adequate cetane number is crucial. According to the 2019 Worldwide Fuel Charter, the minimum cetane number equals 51.0 and up to 55.0 for some markets having high requirements for emission control [14].
It should be noted that branched-chain paraffins typically have lower cetane numbers than their linear counterparts. Among the mono-branched isomers with alkyl substituents or with methyl groups at the end of the hydrocarbon chain, the latter not only have lower melting points but also higher cetane numbers (Table 2). Therefore, the selective hydroisomerizing n-alkanes to these isomers can produce diesel with improved cold flow properties and a shorter ignition delay [15].
Table 1. Characteristics of n-alkanes C3-C18 as a fuel component (sourced from PubChem database [16]).
Table 1. Characteristics of n-alkanes C3-C18 as a fuel component (sourced from PubChem database [16]).
AlkaneChemical
Formula
Boiling Point (°C)Melting Point (°C)Flash Point (°C)Cetane Number
n-propaneC3H8−42.0−187.6−104.0−20.0
n-butaneC4H100.5−138.3−60.020.6
n-pentaneC5H1236.0−129.0−40.030.0
n-hexaneC6H1469.0−95.4−22.047.9
n-heptaneC7H1698.4−90.6−4.056.0
n-octaneC8H18125.6−56.713.064.4
n-nonaneC9H20150.5−53.531.060.9
n-decaneC10H22174.1−29.746.065.5
n-undecaneC11H24195.9−25.562.079.0
n-dodecaneC12H26216.3−9.683.072.9
n-tridecaneC13H28235.4−5.494.088.0
n-tetradecaneC14H30253.65.9100.095.0
n-pentadecaneC15H32270.610.0N/A98.0
n-hexadecaneC16H34286.918.2135.0100.0
n-heptadecaneC17H36303.022.0149.0105.0
n-octadecaneC18H38316.028.2166.0110.0
From a thermodynamic point of view, the formation of iso-paraffins is more favorable at lower temperatures; therefore, to achieve better selectivity towards isomers, catalysts are usually designed for low temperature exploitation [17,18].

2.2. Illuminating the Pathways of Molecular Transformation

Zeolites composed entirely of silica have a neutral charge, however, if Si4+ is replaced with Al3+, the framework becomes negatively charged. Protons can neutralize the framework charges, resulting in the development of acid sites that accelerate the main processes of oil refining: hydrocarbon isomerization and cracking. Based on the zeolite topology and acidity, these processes can yield desirable products by converting linear alkanes into diverse branching isomers and breaking down higher hydrocarbons into smaller ones [19]. Before digging into the influence of shape selectivity on these reactions, it is crucial to look at the case where shape selectivity is omitted.
Conventional hydroisomerization proceeds with the formation of an olefinic intermediate via dehydrogenation on the metal site. As a result of the highly endothermic nature of the dehydrogenation step, the concentration of olefins in the hydroisomerization reaction mixture is typically low. Thus, it is crucial to convert these olefinic compounds promptly to carbocations on the acid sites of the catalyst to ensure efficient conversion. Due to the weak basicity of the hydrocarbons, the equilibrium concentration of carbocations on the acid sites of the zeolite is relatively low [20,21].
According to Jacob et al. [22], there are two types of carbocations:
  • Tricoordinated carbenium ion;
  • Pentacoordinated carbonium ion (extremely unstable).
Carbenium ions can be formed via three distinct routes depending on feedstock properties and the catalyst acidity. These routes include (i) hydride elimination on Lewis acid sites (Figure 2a), which can occur through a variant pathway of the carbenium ion R+ adsorbed on active sites of the zeolite, known as “bimolecular hydride transfer” (Figure 2b); (ii) alkane protonation followed by hydrogen abstraction (Figure 2c); and (iii) olefin protonation, leading to the formation of an alkylcarbenium ion (Figure 2d). The last route comprises the protonation of olefinic intermediates instead of the energetically unfavorable direct protonation of alkanes. Thus, the isomerization via route (iii) accelerates as compared to the direct activation of the alkanes [22].
As per the classical bifunctional mechanism of the isomerization, the metal component of the catalyst facilitates the dehydrogenation of alkanes to form alkenes that are protonated to transitional alkylcarbenium ions on the Brønsted acid sites. These ions undergo rearrangement and scission reactions to obtain ultimately obtain, after hydrogenation on the metal sites, the saturated product—mono-substituted isoalkanes. In the case of availability of potent Brønsted acid sites, mono-substituted tertiary carbenium ions may undergo additional isomerization, finally leading to the formation of carbenium ions containing multi-branches [23].
Aluminosilicate zeolites that incorporate platinum or palladium metal typically operate based on the classical bifunctional mechanism (Figure 3), particularly at high conversions, elevated hydrogen pressure, and a sufficiently high carbon number of the feedstock. The pathway of converting hydrocarbons on metals often compete with the bifunctional mechanism, and the significance of the former increases in the case of more severe conditions and the weak acidity of zeolite. The presence of CH4 and C2H6 in the final reaction products is indicative of metal-catalyzed cracking, a process commonly referred to as hydrogenolysis.
Weiss demonstrated that alkenes displayed high reactivity in the isomerization on Pt-free aluminosilicate, whereas alkanes did not [24]. Several studies have reported an increase in the isomerization rate with rising Pt content, which can be attributed to high concentration of the olefinic intermediate [25,26].
Due to the formation of strong covalent bonds, both the enthalpy of protonation (∆Hp) and the activation energy of the isomerization step (Eiso) are expected to be high, as a result of the deformation of the C-O-Z bond (Z-acid site). The skeletal isomerization reaction is postulated to occur through the rearrangement of the carbenium ion via a cycloalkyl intermediate. The resulting isoparaffinic carbenium ion is subsequently transformed into an olefin via the loss of proton at the acid site. Finally, the isoolefin intermediate is rapidly hydrogenated to the isoalkane product [27].
The rate of isomerization has been observed to be heavily influenced by the length of the alkane chain. The longer chains result in the greater stability of bound carbenium ions and the higher rate of the isomerization reaction. However, high isomerization selectivity becomes increasingly challenging as the chain length increase, due to adverse reactions, namely carbocation intermediate hydrocracking on acid sites and hydrogenolysis reactions on metal sites. The balance between hydrogenation and acid function is a key factor that governs the prevalence of acid- or metal-catalyzed reactions, ultimately determining the composition of products [28,29].
Three mechanisms of alkane cracking have been established: (a) non-catalytic thermal cracking that proceeds at high temperatures as radical process, (b) proteolytic α-cracking including carbonium ion and (c) β-cracking via the carbenium ion. The commonly acknowledged mechanisms of catalytic cracking in hydroisomerization conditions can be described using two fundamental steps: the formation of carbenium ions and β-cracking.
In contrast to acid-catalyzed cracking, the product of thermal cracking of alkanes via free-radical intermediates possesses a the distinct composition (Table 3).
Typically, the formation of free-radical intermediates in the process is of kinetic significance at temperatures higher than 530 °C [30,31].
According to the traditional catalytic cracking mechanism (also known as the carbenium ion mechanism), an alkane is subjected to hydride abstraction by a carbenium ion, resulting in the formation of another carbenium ion, which then undergoes β-scission, or the cleavage of the C-C bond β-situated to the trivalent carbocation [32]. Carbenium ions are known to act as chain carriers, with the most stable ones, typically tertiary ions like the t-butyl cation, being the predominant carriers [33,34].

3. n-Hexadecane Hydroisomerization on Bifunctional Catalysts

3.1. Development of Hydroisomerization Catalysts

To improve the formation and isomerization of the carbocation intermediate in the hydroisomerization reaction, the catalyst must have a significant acidity [35,36]. Since the discovery of aluminum chloride in 1933 as a catalyst for hexane and heptane isomerization, four generations of hydroisomerization catalysts have been developed over the course of several decades (Table 4) [37,38,39]. Currently, only third and fourth generations are used in the industry, although the academic interest in these generations of catalysts has increased dramatically over the past 10 years (Figure 4), and this review focuses on the last generation. A critical challenge in the research and development of innovative hydroisomerization catalysts is to achieve high selectivity of isomerization and considerable n-paraffin conversion, constraining the formation of cracking byproducts. This parallel reaction reduces the yield of the intended products [40].
Typical zeolites-based catalysts are prone to poisoning and deactivation in the hydroisomerization process of linear alkanes due to the presence of sulfur-containing substances in the feed. Currently, the major developments in the design of catalysts for n-paraffin hydroisomerization are the high efficiency at temperatures 100–200 °C and tolerance to both sulfur and water.
Synthetic zeolites have been recognized as solid acid catalysts and are widely used in a variety of applications in industry, notably in petrochemical synthesis and oil refining [41]. The topology and morphology of the zeolites defines their suitability for the particular application process (Table 5), which is defined by interlinked channel structures or a hierarchical structure that integrates micro- and mesopores, in addition to acid-base properties [42,43,44,45,46].

3.2. Analysis of the Relationship between Catalyst Characteristics and Activity

3.2.1. Shape-Selectivity

The study of the structural effect of zeolites on their selectivity deals with the concept of “shape-selectivity” that is classified into two basic categories: configurational selectivity and transition state selectivity (Figure 5). Typically, zeolites acquire the requisite shape selectivity at the expense of restricted mass transfer inside micropores to and from the active sites. To overcome this issue, several approaches have been used to improve the accessibility of active sites inside zeolites. One successful way is to take advantage of molecular shape selectivity, which acts inside zeolite’ channel systems and has shown higher efficiency and selectivity in various chemical reactions.
Configurational selectivity arose from the insight that molecule cannot pass through the channels of the zeolite. In the case where the molecule will not be adsorbed it will keep staying intact in the resultant mixtures (reactant shape selectivity) (Figure 5a). In the case where the molecule will not be desorbed, it can exist as an adsorption transitional-state, and only molecules originating from this state will be included in the product via the following reactions (product shape selectivity) (Figure 5b). In addition, the effective pore diameter may strongly affect the formation of transition-state intermediates and direct them towards forming states that fit its topology or towards participating in secondary intermediate interactions (transitional-state shape selectivity) (Figure 5c).
In some cases, transition state intermediates approximating the topological framework of the zeolite are quickly formed (inverse shape selectivity) (Figure 6a), this phenomenon was caused by sorption shape selectivity instead of transition state shape selectivity, as previously thought [47,48].
Transition state selectivity can be influenced by restrictions that could occur either on the surface or within the pores of the framework of the zeolite. In the locations where the paths of the pores happen to intersect, this case can affect the diffusion of the reactant molecules—whether due to geometric reasons or the characteristics of molecular diffusion. As a result, molecules of the specific reactant may enter one of the pores and do not enter to the other one (molecular traffic control) (Figure 6b). Otherwise, the ends of the straight-chain alkanes may enter simultaneously into two (or more) adjacent pores (Pore mouth & key-lock selectivity) (Figure 6c) [49,50,51,52,53].
In addition to the previous two classifications, some effects related to the interactions on the outer surfaces of zeolites were studied and formulated by Chen et al. [54], including the window or cage effect (Figure 6d), to understand how the shape selectivity of erionite affects the product distribution of cracking of linear alkanes (C16H22–C36H74). This phenomenon attempts to explain the differences in diffusion rates of n-paraffin having various chain lengths inside a particular zeolite structure, which are referred to as “up and down” diffusivity variations. Generally, as the chain length of the n-paraffin increases, its diffusivity within the zeolite structure decreases due to the higher steric hindrance [55]. Derouane [56] also formulated the nest effect (Figure 7), which tried to explain the non-shape selectivity of zeolite at transition-state intermediates formation that occur within the cavities located at the ends of the pores.
Various approaches have been established to manipulate the morphological characteristics of zeolites based on using and developing new structure-directing agents in order to gain a better distribution of products of n-hexadecane hydroisomerization. Several of these approaches have led to acceleration of hydrocarbons’ transport within the crystals, and eventually to decreasing the residence time of intermediate alkylcarbenium in microspores and the rate of secondary cracking.
Romero D. et al. [57] observed the effect of member ring (MR) size of the zeolite framework on the selectivity of n-C16 hydroisomerization over 3D-10MR ZSM-5 (bulk and nanostructured), 1D-12MR ZSM-12, 1D-10MR ZSM-22 and extra-large intersecting pores (10MR + 21MR) EMM-23. Zeolite ZSM-5 having hierarchical structure and shorter intercrystalline length (higher diffusion rate) demonstrates an upgraded shape-selectivity for di-branched isomers (dimethyl) cracking owing to a snug fit of isomers at (straight-zigzag) channel intersections in the ZSM-5 framework. This shape of intersections is absent in the framework of ZSM-22; however, the cracking of di-branched isomers could be explained by the pore mouth type of shape-selectivity. The wide size of the 12MR pores provides an acceptable performance of ZSM-12, whilst in the case of EMM-23 the 10MR pores prevailed over the framework at the expense of 21MR, as a result of molecular defects of silanol groups.
Table 6 shows ZSM-12 prepared using various types of organic templates giving different morphological and acidic properties of zeolite. For instance, ZSM-12 with high specific area (>300 m2.g−1) and good total acidity (~200 µmol/g) with low Si/Al ratios (~40) is synthesized applying template comprising tetraethylamonium cations (like TEABr and TEAOH). In contrast, samples prepared with MTEACl and MTEABr as templates have high specific area with Si/Al ratios near to 70, to increase their acidity up to 300–700 µmol/g. As for ZSM-22, the majority of researchers use 1,6-diaminohexane as the template, giving zeolite with good crystallinity and optimal textural properties. With regard to ZSM-23, it may be produced using template comprising amide and amine functional groups, however, the last type has proven its superiority with Si/Al = 100–150 despites of the weak acidity of the zeolite produced.
The wide pores are not always preferred, where it could allow secondary reactions to happen. For instance, 12MR zeolites have the capability to accommodate mono-, di- and multi-branched hydrocarbons. However, multi-branched hydrocarbons are prone to undergoing cracking reactions [78,79].
An optimal hydroisomerization catalyst provides a high yield of isomerized products alongside the decreasing of cracking products formation (Figure 8). As a result, 10MR zeolites-based catalysts have been identified as the preferred catalyst because they allow unhindered diffusion of monobranched hydrocarbons while discouraging diffusion of multi-branched hydrocarbons via shape selection, promoting superior isomerization selectivity. Zhang M. et al. [80] studied different 10MR zeolites (ZSM-22, -23, -35 and -48)—corresponding topologies (TON, MTT, FER and MRE respectively)—as catalysts’ supports n-C16 hydroisomerization reaction. The products distribution of n-C16 hydroisomerization over 10MR zeolites was significantly influenced by tiny distinctions in the channel configuration.
The predominated hydroisomerization product was primarily 2-methylpentadecane. This likely occurred due to fast diffusion of 2-methylpentadecane. 4-Methylpentadecane and 5-methylpentadecane were not observed in products obtained over the ZSM-22-based catalysts, due to the smallest pore size in zeolite of this type.
The hydroisomerization mechanism might be explained by the pore mouth reaction pathway due to the lack of detection of multi-branched isomers. These results are supported by the study of Zhang et al. [79], in which three different types of zeolites were tested (ZSM-12, ZSM-22 and Beta) as supports of 0.5 wt. % Pt catalysts. 1D-10MR- ZSM-23 gave the highest concentration of mono-branched C16 isomers, whereas ZSM-12 and ZSM-48 zeolites gave the highest concentrations of multi-branched isomers. These could be described by pore channel sizes for ZSM-12 and ZSM-48 (0.57 × 0.61 mm and 0.53 × 0.56 mm respectively). Significant isomerization yields have been reported with non-interconnected mesopore zeolites and zeotypes, such as ZSM-12, IZM-2 and ZSM-23 [81,82].

3.2.2. Effect of the Active Metal Site

In the literature, much attention is given to the impact of the metal precursor and its loading on the catalyst selectivity (Table 7). The metal precursor utilized in synthesis can also contribute to heterogeneities in the metal structure obtained on both bulk and nanometer scales, which are referred to as precursor effects [83,84,85]. Distinct characteristics of each metal precursor, such as its decomposition temperature and interaction with the support, can significantly affect the distribution of metal particles across the support and the particle size distribution [86]. As a result, the metal to acid site ratio (nM/nA) within the sample varies. Ultimately, this can have a substantial impact on the performance of a catalyst [87,88,89,90]. In contrast, when considering the specific n-paraffin and operating conditions, the performance of bifunctional catalysts depends not only on the zeolite utilized but also on the ratio of metal/acid sites [25,91]
Table 7 shows that changing Si/Al—using [Pt(NH3)4](OH)2 as a metal precursor—does not effect the performance of ZSM-12 especially on the side of conversion, but on the other hand, the selectivity slightly improves with increasing ratio Si/Al. Also the use of different metal precursors greatly affects the selectivity of zeolite ZSM-22, as the case of using Pt(NO3)2, Pt(NH3)4Cl2 and H2PtCl6, it is clear that the latter offers the best selectivity. In general, it is obvious that the best metal loading for all zeolite-based catalysts is 0.50% achieving an optimum balance between metal and acidic functions along with obtaining and guarantee the high dispersion of metal on the support.

3.2.3. Effect of the Acidity

The zeolite framework contains exchangeable cations, which play an important role on the neutralization of the charge produced by isomorphous substitution (Figure 9). For example, the substitution of Si+4 with Al+3 leads to the formation of bridging [Si-OH-Al] hydroxyls, which acts as a Brønsted acid sites (BAS) providing the catalytic activity of zeolites [100]. The removal of the 4-coordinate Al associated with the BAS via dealumination results in the formation of extra-framework Al (EFAL) species within the pores, which often function as Lewis acid sites (LAS) [101,102,103]. These EFAL species have been found to modify the thermal stability of zeolites and enhance their catalytic activity. Notably, the hydrothermally dealuminated Y zeolite (USY) serves as a general catalyst in the fluid catalytic cracking process.
Moreover, extra-framework or framework-incorporated metals, such as Ga, B, Ti, and Sn, may form LAS. These metals frequently provide zeolites with bifunctionality characterized by both redox property and acidity, which extends their application beyond petrochemistry to biomass transformation [104,105,106].
Achieving a balance between the density and strength of acidic sites is critical for the reactivity and selectivity of hydroisomerization reactions, and catalysts with optimal hydrogenation capacity and suitable acidity are preferred [107,108]. Aluminosilicate zeolites with medium pores have significant activity and selectivity for the isomerization of long-chain n-alkanes. However, excessive acidity may lead to cracking reactions, and extensive efforts have been made to modify the acidity via improved synthetic methods or post-treatments, such as altering acid site distribution using mixed zeolitic catalysts, among others [109].
A catalyst with high acidity could cause both hydroisomerization and hydrocracking processes, resulting in lower isoparaffin yields and the breakdown of long-chain n-paraffins to obtain less valued and lighter products. To promote hydroisomerization and suppress hydrocracking, a catalyst with the strong hydrogenation activity and moderate acidity is ideal. Furthermore, the size of the pore aperture in zeolites may have a substantial impact on catalyst selectivity. In the case where pore sizes are small enough to prevent larger iso-paraffin formation at the acid sites, the catalyst will be good on converting n-paraffins [110].
Over the past few decades, researchers have emphasized the significant influence of the zeolite acidity of hydroisomerization catalysts on their activity and selectivity [111,112,113]. The general consensus is that a proper balance between hydrogenation and acid functions is essential to maximize isomer yields. However, the role of acid site density and strength remains a subject of debate due to variations in zeolite structure and reaction conditions [114]. For instance, in the case of hydrogen spillover highlights, the olefinic intermediates formed on acidic sites can be hydrogenated with H adsorbed over the far Pt cluster, and as a result, the physical mixture of Pt/SiO2 and zeolite can exhibit good selectivity. Batalha et al. [115] propose that catalyst selectivity is determined by the intimacy between metal and protonic sites, which is the number of acid sites that olefinic intermediates encounter during diffusion between two metallic sites.
The acidity of zeolites can be adjusted by altering the ratio of SiO2/Al2O3. Furthermore, some alkaline earth metals (Ba [116], Mg [117]) and transition metals (Fe [118]) can be included into the structure to decrease the acidity; the same effect exerts coating the zeolite with alumina [119].

4. Impact of Operating Parameters on Process Efficiency

The operational conditions of hydroisomerization of the diesel fraction are largely influenced by the composition of feedstocks and the desired quality of the final product. However, the choice of catalyst is equally critical. Initially, the hydroisomerization process was performed at high hydrogen pressures ranging from 10 to 25 MPa. Upon the development of more active catalysts, however, the pressure requirements have been reduced to 4–5 MPa.
The temperature of the process is determined by both the boiling range of the feedstock and the activity of the catalyst. Usually, higher boiling fractions necessitate higher process temperatures. The application of highly efficient zeolite-based catalysts has enabled a significant temperature reduction in recent years. For example, kerosene fractions are processed at temperatures ranging from 240 to 280 °C, as opposed to previous temperatures as high as 450 °C. Thus, the progress of the hydroisomerization process could be defined as a trend to significant reduction of temperature and pressure due to the use of more active catalysts. Despite numerous investigations in the field of knowledge, there is a relatively limited understanding of the interconnected influence of these operational parameters. Usually, experiments are performed by varying temperature within a predetermined range while keeping the remaining parameters fixed. Consequently, the predominant impact of temperature variation on catalyst performance can be observed. In general, with rise in temperature, conversion increases, but selectivity decreases (Figure 10).
Typically, recent studies of n-hexadecane hydroisomerization have performed at temperature range of 200–380 °C, operating pressure up to 6 MPa, the weight hourly space velocity (WHSV) between 1–6 h−1, and H2/n-C16 ratios ranging from ~10 up to 1000.
As Table 8 shows, the majority of catalysts operate at temperature of at least 270 °C and pressure exceeding 2 MPa. Analysis of the performance of catalysts above reveals that zeolites ZSM-12, ZSM-22, and SAPO-11 at temperatures above 310 °C and ZSM-23 at temperatures between 270 and 290 °C are superior supports for platinum-impregnated catalysts. In contrast, zeolites SAPO-11, SAPO-31, and SAPO-41 at temperatures exceeding 340 °C are preferable for palladium-impregnated catalysts, while catalysts supported with BETA, USY, and mordenite have the modest performance.
However, ZSM-5 supports demonstrated the worst selectivity (whatever type of active metal) even in the nanostructural form. As for the effect of pressure change, it depends on the type of zeolites. For example, in the case of ZSM-5, the increase in pressure from 2 to 6 MPa has affected remarkable effect on the conversion of n-alkanes. On the other hand, in the case of ZSM-22, the effect of pressure is somewhat proportionally limited on the reaction conversion and selectivity. However, the effect of WHSV and H2/n-C16 is still unclear considering that they have not ever been the main objectives of any investigation.
Parmar et al. [116] studied the impact of increasing WSHV on the performance of bifunctional catalysts comprising Pt/ZSM-22 modified with Group-II cations (Ca+2, Ba+2, Mg+2). Their findings revealed that higher WSHV and temperatures > 300 °C led to a decrease in the conversion and an increase in selectivity towards iso-paraffins C16. This observation can be due to the shortened reaction time, which has a negative impact on both the main (isomerization) and secondary (cracking) reactions. Subsequently, Zhang et al. [80] investigated the impact of WHSV at temperatures below 300 °C on the performance of different Pt-impregnated 10MR zeolites. These data confirmed an advantageous effect of increasing WHSV for both MTT and FER zeolite types (Pt loadings ~0.93 wt %), resulting in improved selectivity. This phenomenon can be attributed to the suppression of cracking iso-paraffins C16. In contrast, increasing WSHV led to a decrease in selectivity for both TON and MRE zeolite types (Pt loading ~0.91 wt %), due to the lesser content of Brønsted acid sites.
Upon the investigation of n-hexadecane hydroisomerization over Pd-Pt/β-zeolite catalyst, Bauer et al. [129] observed a decline in the conversion from 89.0% under pressure of 3 MPa to 53.0% at 7.5 MPa. This result can be attributed to the competition of olefinic intermediates with hydrogen adsorbed over active sites under high-pressure conditions. Although higher hydrogen pressures have a negative impact on the conversion of n-paraffins and provide minimal enhancements in selectivity and catalyst stability, it should be noted that industrial specifications are usually required to exploit catalysts within the pressure range of 3–5 MPa.

5. Tolerance of Bifunctional Catalysts to Poisons and Deactivation

5.1. The Effect of Impurities-Containing Feed

Commercial feedstock is frequently impaired with nitrogen- and sulfur-containing compounds, which could damage the metallic and acidic active sites of bifunctional catalyst [132]. Nitrogen poisoning is a serious issue in catalysis, especially with zeolite-based catalysts employed in n-hexadecane isomerization. Understanding the forms of nitrogen adsorption and the further blocking of active sites allows us to understand the mechanisms that govern catalyst deactivation and selectivity loss. The deactivation of bifunctional catalysts on hydroisomerization of n-hexadecane with sulfur and nitrogen compounds has got some attention, but the understanding of their impacts on catalytic isodewaxing is unclear [133,134].
According to Lee et al. [135] the polymetallic bifunctional Pt-Pd-Mg/ZSM-23 catalyst exhibited a high stability of product properties for more than 30 days in the presence of sulfur impurities (1–2 wt.%).
Pimerzin et al. [136] examined the stability of two type of SAPO-11-supported bifunctional catalysts (transition metal sulfides and noble metal) for the isomerization of hexadecane. Nitrogen-containing feedstock inhibits isomerization catalytic activity more strongly, and whether the quantity of sulfur in the feedstock is equivalent to or higher than 100 ppm, transition metal-based bifunctional catalysts surpass Pt-catalyst in n-hexadecane isomerization (Figure 11).
Later, the effect of nitrogen-containing feed was retraced by Vinogradov et al. [130] with different zeolite-supported catalysts (ZSM-5, Beta, and Y). It was observed that the nitrogen impurities (quinoline) could play a role as a selective-controlling agent, as a result of the predominant adsorption of quinoline on strong acid sites performing the cracking process.

5.2. The Impact of Water

The effect of water-containing feed on the process of hydroisomerization of n-alkanes was considered recently [137,138,139]. Controlled amounts of water and iso-pentanol promoted the performance of FAU-type structured zeolite (with rare-earth metal—REM) in the hydroconversion of n-hexadecane [140]. According to the results, the support was activated by creating new protonic active centers via the hydration of REM cations. However, due to competing adsorption and hydrating protons located in the zeolite, water decreases acidic strength and deactivates the catalyst.
The impact of competing water adsorption on the acid sites of Pt/MFI, Pd/MFI, and Pt/HY was investigated by Brosius et al. [141]. Authors discovered that water prevented n-hexadecane cracking products from the secondary isomerization. Moreover, in the case of using Pt(0.9 wt%)/HY products with a great extent of branching were obtained. In contrast, Pt(0.9 wt%)/MFI produced a wider extent of linear products (Figure 12).

5.3. Deactivation by Coke Deposition

Upon the hydroisomerization of n-alkanes over zeolite-based catalysts, carbonaceous deposits are obtained as a result of complicated chain of reactions, namely the oligomerization, cyclization, and condensation of alkylaromatics to generate polycyclic aromatics (Figure 13) [142,143,144,145]. Coke is composed of heavy components that are extremely deposited as a byproduct on the surface of the zeolite, poisoning the active sites and clogging the pores. It was observed that the interaction between intermediates of coking and other reactions intermediates impacts on the distribution and selectivity of the products. Coking occurs faster on strong acid sites than on weak acid sites, because the last are less sensitive to coke formation [146,147].
The coke can be formed in one of three ways: (a) synchronously with the desired reaction (competitive phase), (b) consecutively to the directed reaction (successive phase), or (c) both of (a + b) simultaneously [145]. Coke combustion can renew the catalyst, however, the intense regeneration conditions and the presence of water prevent full catalytic activity from being regained [148].
Bauer and Karge [149] proposed a classification system that differentiates between two forms of coke obtained. Condensation, oligomerization, and alkylation processes involving reactants and their intermediates produce “Coke-type I” at lower temperatures. “Coke-type II” is found at higher temperatures and comprises chemical stages requiring greater activation energies, such as dehydrocyclization, thermal cracking, and dehydrogenation.
The coke formation rate is influenced by the same parameters as the rate of the catalytic reaction itself. These parameters encompass the nature of the reactants [150,151], the pore structure of the catalyst [152], the density of the active sites [151,153], and the operating conditions [153].
Carbonaceous deposits are thought to disable the catalyst by poisoning the active sites or by pore blockage [154], but the latter causes more severe deactivation than active site poisoning [155], because one active site per coke molecule can be deactivated, but pore clogging with a single carbonaceous molecule could limit or prohibit reactant molecules from accessing all active sites inside the clogged pore.
The coke may accumulate via three ways: (a) uniform surface deposition [156], (b) pore-mouth blockage [157], and (c) bulk-phase blockage [156,158]. Each mode has a particular effect on the diffusivity of reactants and, as a result, on catalytic activity.
In the case of uniform surface deposition, it has been hypothesized that carbonaceous deposits are uniformly distributed throughout the micropore channels of catalyst. Nonetheless, this distribution is likely to shorten the micropore diameter decreasing the diffusion rate and catalytic activity. According to the pore-mouth blockage scenario, carbonaceous deposits tend to build at the pore entrances, limiting access for the reactant to the active sites within the pore. Although the pores are not totally blocked, this process reduces the diffusion rate. In this situation, carbonaceous deposits cover a tiny fraction of active sites, in which productivity will be proportional to the number of accessible active sites [157]. The bulk-phase blocking scenario incorporates both of the aforementioned carbonaceous deposition mechanisms. However, in this instance, a large quantity of coke is generated on the outer surface of catalyst, resulting in full clogging of the pores.
During the catalytic cracking of n-hexane and 1-hexene over ultrastable Y zeolite, Chen and Manos [159,160] differentiated coke into coke precursors and hard coke. Coke precursors are easily volatilized in inert nitrogen and removed off the catalyst sample, whereas hard coke stays on the catalyst even at high temperatures (600 °C) and is removed by burning. Bauer et al. [129] investigated the formation of coke depositions during the hydroisomerization of n-hexadecane on the Pt-Pd/H-beta bifunctional catalyst. It was revealed that both the volume and surface area of micropores were decreased by roughly 8% after 62 days on-stream as a result of pore occlusion with coke.
Bifunctional catalysts including mesopores have enhanced the lifetime. In the most cases, deactivation is caused by pore-mouth occlusion by carbonaceous deposition. Numerous studies have claimed that the favorable effect on the lifetime is due to an increased tolerance to the impacts of coke rather than a lower rate of the coke formation. Shortening diffusion lengths combined with increased pore openings in the mesoporous makes the pores more resistant to blocking, which may explain the reduced effects on coking. This simple reasoning, however, may not fully describe the situation. Holm et al. [161] reasoned that, due to mesoporosity, products obtained will have a shorter retention time in the micropores, potentially suppressing subsequent secondary coking.

6. Cost-Effective and Eco-Friendly Synthesis Approaches

Conventional methodologies for zeolites synthesis have been extensively documented in the scientific literature [162]. However, numerous challenges exist, including the high expenses associated with structure-directing agents and starting materials, energy consumption and, recycling wastes. Moreover, the multistage production comprising expensive procedures hinders commercialization [163,164,165]. As a result, the synthesis of hierarchical zeolites using an environmentally friendly approach has gained attention [166,167]. Reliable syntheses aim to accommodate the effectiveness of conventional methods with strategies that reduce environmental impact, such as utilizing alternative starting materials, carrying out solvent-free or template-free syntheses, and reusing water [168,169,170].
Recent studies have reported the successful synthesis of hierarchical zeolites using an eco-friendly approach [171,172]. These works have demonstrated the feasibility of zeolite production through the application of green chemistry principles. However, certain drawbacks have also been observed, including low crystallinity, non-uniform pore structures, and reduced process efficiency (Table 9).
Significant efforts have been devoted to establishing the connections between the characteristics and catalytic performance of zeolites, driven by their economic significance. In this regard, the development of innovative syntheses and alternative raw materials have a focal point. An example of a recent eco-friendly approach is template-free synthesis, which has been tested for the production of the MFI framework structure [173,174,175]. Notably, tubular clay nanomaterials, specifically Halloysite, have been utilized as triple-functional materials: the source of alumina and silica, and hard template [176,177,178,179,180]. It should be noted that bioresources, such as water reeds (Phragmites) [181] and rice husk ash [182,183,184,185,186,187], mineral sources including rectorite [188] and diatomite [189,190], and industrial wastes like coal fly ash [191,192], have a minor scale, but employed as sources of silica and alumina in the synthesis above. This advancement represents a notable step forward in green chemistry for zeolite production.
Occasionally, traditional synthesis approaches necessitate substantial quantities of organic or water-based solvents that are economic and ecological drawbacks, particularly evident in the context of silicoaluminophosphate (SAPO) materials [193,194].
Recently, Xiao et al. effectively synthesized a variety of zeolites without the need for extra solvents. The zeolites obtained comprised SAPO-11, SAPO-20, SAPO-34), aluminophosphates (AlPO-11), metal-containing aluminophosphate zeolites (Mg-AlPO-11, Co-AlPO-11, Mg-SAPO-46, and Co-SAPO-46) [195], as well as aluminosilicates (ZSM-5, silicate-1, ZSM-49, SOD, MOR, Beta, and FAU) [196]. This innovation in synthesis indicates that it has the potential to produce a wide range of zeolite frameworks without the need for extra solvents.
Most current studies have focused on the seed-assisted approach to crystal pattern modification as a simple and ecologically benign method [197,198,199]. This method includes the inducing and orientation of crystals growth by adding into the gel a minor amount of the same targeted zeolite structure as a seed. This method has been shown to be successful in syntheses of zeolite frameworks such as MFI [200,201,202,203], MTW [64,204], MEL [205], FAU [206], MOR [207], CHA [208], and MRE [209].
Nonetheless, despite its potential benefits, the seed-assisted approach has several limits, such as the broad particle size distribution, tendency to particle aggregation, and possibility of pore blockage [210,211]. To maximize the efficiency and performance of the seed-assisted synthesis technique, these constraints must be handled and minimized.
Unconventional methodologies for zeolite synthesis and the ability to modify surface properties through post-treatments are two effective strategies employed to improve the zeolite performance. One such approach is microwave-assisted hydrothermal synthesis (MAHyS), which significantly accelerates synthesis. For instance, Muraza et al. [212] successfully increased the rate of synthesis of the TON framework zeolite from 66–72 hours to 24 hours using the MAHyS technique.
Another promising technique is the ultrasound-assisted synthesis of zeolites, which offers several advantages, including increased surface area and shorter synthesis time [213,214]. These effects have been confirmed in the syntheses of MCM-22 [215], ZSM-5 [182,216], and SAPO-11 [217] zeolites, where the ultrasound-assisted method promotes the dissolution of reactants and accelerates the progression of crystallization.

7. Conclusions

The isomerization of n-hexadecane and its multiple isomers is critical to upgrading the low-temperature characteristics of diesel, such as cloud point, pour point and cold filter plugging point, also enhancing the cetane number, therefore, the effective catalysts and the optimal operating conditions have been reviewed.
The concept of main and secondary types of shape selectivity was analyzed in detail. On this basis the requirements for the preparation procedure of zeolites were formulated: mainly, mesopores should possess the product shape, transition-state, and pore mouth selectivity, but pores with the reactant shape, mutual intersections, and nest effect selectivity should be excluded.
The fundamentals of mechanism of n-hexadecane hydroisomerization, characterizing zeolite were highlighted. To improve the yield, the choice of structural and morphological characteristics of catalyst, combined with the optimal balance between metal and acidic sites (nM/nA) have the greatest impact on the process. Large external surface area and mesoporous structures minimize the diffusion limitations and unfavorable cracking. Diffusion on a high surface area enhances adsorption at pore-mouths and improves accessibility to acid sites, resulting in better conversion and selectivity.
The effects of nitrogen-, sulfur- and water-containing feedstock on the catalyst performance were also discussed. It was concluded that sulfur plays a lesser deactivation role than nitrogen impurities as a selectivity-control agent, poisoning the acid sites and deactivating the isomerization function. In contrast, water prevents the deactivation of catalysts and suppresses cracking reactions.
The recent trend in the synthesis zeolite-supported catalysts from natural sources—such as clay and natural minerals is considered very attractive from both economic and environmental standpoints. Although it is still limited to the ZSM-5 zeolite, it promises a lot of progress in developing new types of zeolites. The novel methods for synthesis of bifunctional catalysts, and their success were discussed in overcoming the problems that may face traditional methods, despite the presence of some challenges that still exist in environmental and economic areas, and the textural and structural properties of zeolites.
Furthermore, the use of alternative feedstocks, such as renewable sources, as well as the integration of hydroisomerization to other refining processes, such as hydrotreating and hydrocracking, could open up new avenues for producing high-quality diesel fuels.
The authors thank GZ (FSZE-2022-0002).

Author Contributions

Conceptualization, Y.A., V.S. and A.G.; writing—original draft preparation, Y.A. and M.R.; writing—review and editing, Y.A., V.S. and A.G.; supervision, A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kazakov, M.O.; Smirnova, M.Y.; Dubinin, M.E.; Bogomolova, T.S.; Dik, P.P.; Golubev, I.S.; Revyakin, M.E.; Klimov, O.V.; Noskov, A.S. Combining USY and ZSM-23 in Pt/zeolite hydrocracking catalyst to produce diesel and lube base oil with improved cold flow properties. Fuel 2023, 344, 128085. [Google Scholar] [CrossRef]
  2. Viswanadham, N.; Saxena, S.K.; Panwar, R.; Ray, A. A single-step catalytic process for the production of high-octane molecules from normal paraffins using zeolite supported bi-functional catalysts. Chem. Eng. Process. Process Intensif. 2022, 177, 108990. [Google Scholar] [CrossRef]
  3. Liu, S.; Zhang, L.; Zhang, L.; Zhang, H.; Ren, J. Function of well-established mesoporous layers of recrystallized ZSM-22 zeolites in the catalytic performance of n-alkane isomerization. New J. Chem. 2020, 44, 4744–4754. [Google Scholar] [CrossRef]
  4. Li, K.; Cui, T.; Zhao, J.; Liang, C.; Li, C. Regulate the mesopores and acidic structure of Pt/USY zeolite catalyst for improved performance in the hydroisomerization of Phenanthrene to Alkyl-adamantane. Chem. Eng. Sci. 2023, 279, 118957. [Google Scholar] [CrossRef]
  5. Chang, T.; Liu, S. Role of the mesoporous diameters of hierarchical ZSM-22/MCM-41 zeolite for n-alkane isomerization. Mol. Catal. 2021, 503, 111420. [Google Scholar] [CrossRef]
  6. Li, L.; Shen, K.; Huang, X.; Lin, Y.; Liu, Y. SAPO-11 with preferential growth along the a-direction as an improved active catalyst in long-alkane isomerization reaction. Microporous Mesoporous Mater. 2021, 313, 110827. [Google Scholar] [CrossRef]
  7. Mäki-Arvela, P.; Azkaar, M.; Vajglová, Z.; Aho, A.; Hemming, J.; Peurla, M.; Eränen, K.; Kumar, N.; Murzin, D.Y. Hexadecane hydrocracking for production of jet fuels from renewable diesel over proton and metal modified H-Beta zeolites. Mol. Catal. 2019, 476, 110515. [Google Scholar]
  8. Mäki-Arvela, P.; Kaka khel, T.A.; Azkaar, M.; Engblom, S.; Murzin, D.Y. Catalytic hydroisomerization of long-chain hydrocarbons for the production of fuels. Catalysts 2018, 8, 534. [Google Scholar] [CrossRef]
  9. Patrylak, L.; Krylova, M.; Pertko, O.; Voloshyna, Y.; Yakovenko, A. n-Hexane isomerization over nickel-containing mordenite zeolite. Chem. Chem. Technol. 2020, 14, 234–238. [Google Scholar] [CrossRef]
  10. Akhmedov, V.M.; Al-Khowaiter, S.H. Recent advances and future aspects in the selective isomerization of high n-Alkanes. Catal. Rev. 2007, 49, 33–139. [Google Scholar] [CrossRef]
  11. Misra, P.; Alvarez-Majmutov, A.; Chen, J. Isomerization catalysts and technologies for biorefining: Opportunities for producing sustainable aviation fuels. Fuel 2023, 351, 128994. [Google Scholar] [CrossRef]
  12. Echevskii, G.; Weixin, Q.; Kodenev, E.; Toktarev, A.; Chunmei, D. Study of Bifunctional Hydrocracking and Hydroisodeparaffinization Catalysts Based on Zeolites and Alumophosphates, Part 2: Effect of the Activity of Hydrogenating and Acidic Components on the Activity and Selectivity of Hydroisodeparaffinization Catalysts. Catal. Ind. 2020, 12, 81–87. [Google Scholar] [CrossRef]
  13. Liu, P.; Qiu, Z.; Shi, H.; Song, Y.; Zhao, D.; Wang, P.; Wang, T.; Bao, X. Reducing the diffusion barriers of Pt/Beta catalyzed n-hexane isomerization by SBA-15 addition and high-energy milling. Microporous Mesoporous Mater. 2023, 356, 112591. [Google Scholar] [CrossRef]
  14. Worldwide Fuel Charter. Gasoline and Diesel Fuel; Worldwide Fuel Charter: Brussels, Belgium, 2019. [Google Scholar]
  15. Do, P.T.; Crossley, S.; Santikunaporn, M.; Resasco, D.E. Catalytic strategies for improving specific fuel properties. Catalysis 2007, 20, 33–64. [Google Scholar] [CrossRef]
  16. Yanowitz, J.; Ratcliff, M.A.; McCormick, R.L.; Taylor, J.D.; Murphy, M.J. Compendium of Experimental Cetane Numbers; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2017. [Google Scholar]
  17. Ribeiro, F.; Marcilly, C.; Guisnet, M. Hydroisomerization of n-hexane on platinum zeolites: I. Kinetic study of the reaction on platinum/Y-zeolite catalysts: Influence of the platinum content. J. Catal. 1982, 78, 267–274. [Google Scholar] [CrossRef]
  18. Chica, A.; Corma, A. Hydroisomerization of pentane, hexane, and heptane for improving the octane number of gasoline. J. Catal. 1999, 187, 167–176. [Google Scholar] [CrossRef]
  19. Kosareva, O.A.; Gerasimov, D.N.; Maslov, I.A.; Pigoleva, I.V.; Zaglyadova, S.V.; Fadeev, V.V. Effect of the Zeolite Type on Catalytic Performance in Dewaxing of the Diesel Fraction under Sour Conditions. Energy Fuel 2021, 35, 16020–16034. [Google Scholar] [CrossRef]
  20. Hiraoka, K.; Kebarle, P. Stabilities and energetics of pentacoordinated carbonium ions. The isomeric protonated ethane ions and some higher analogs: Protonated propane and protonated butane. J. Am. Chem. Soc. 1976, 98, 6119–6125. [Google Scholar] [CrossRef]
  21. Liu, S.; Luo, C.; Deng, X.; Fang, Y. Toward rational design of narrowly-distributed mesopore on ZSM-22 zeolite for enhanced Pt dispersion and n-alkane isomerization performance. Fuel 2022, 328, 125282. [Google Scholar] [CrossRef]
  22. Jacobs, P.A.; Martens, J.A. Introduction to acid catalysis with zeolites in hydrocarbon reactions. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 1991; Volume 58, pp. 445–496. [Google Scholar]
  23. Weitkamp, J. Catalytic hydrocracking—Mechanisms and versatility of the process. ChemCatChem 2012, 4, 292–306. [Google Scholar] [CrossRef]
  24. Weisz, P.; Swegler, E. Stepwise reaction on separate catalytic centers: Isomerization of saturated hydrocarbons. Science 1957, 126, 31–32. [Google Scholar] [CrossRef] [PubMed]
  25. Alvarez, F.; Ribeiro, F.; Perot, G.; Thomazeau, Y.C.; Guisnet, M. Hydroisomerization and hydrocracking of alkanes: 7. influence of the balance between acid and hydrogenating functions on the transformation of n-decane on pthy catalysts. J. Catal. 1996, 162, 179–189. [Google Scholar] [CrossRef]
  26. Ravishankar, R.; Sivasanker, S. Hydroisomerization of n-hexane over Pt—H-MCM-22. Appl. Catal. A Gen. 1996, 142, 47–59. [Google Scholar] [CrossRef]
  27. Kazansky, V.; Frash, M.; Van Santen, R. Quantumchemical study of the isobutane cracking on zeolites. Appl. Catal. A Gen. 1996, 146, 225–247. [Google Scholar] [CrossRef]
  28. Martens, G.; Marin, G. Kinetics for hydrocracking based on structural classes: Model development and application. AICHE J. 2001, 47, 1607–1622. [Google Scholar] [CrossRef]
  29. Zschiesche, C.; Himsl, D.; Rakoczy, R.; Reitzmann, A.; Freiding, J.; Wilde, N.; Gläser, R. Hydroisomerization of Long-Chain n-Alkanes over Bifunctional Zeolites with 10-Membered-and 12-Membered-Ring Pores. Chem. Eng. Technol. 2018, 41, 199–204. [Google Scholar] [CrossRef]
  30. Kotrel, S.; Knözinger, H.; Gates, B. The Haag–Dessau mechanism of protolytic cracking of alkanes. Microporous Mesoporous Mater. 2000, 35, 11–20. [Google Scholar] [CrossRef]
  31. Shakor, Z.M.; Ramos, M.J.; AbdulRazak, A.A. A detailed reaction kinetic model of light naphtha isomerization on Pt/zeolite catalyst. J. King Saud Univ. Eng. Sci. 2022, 34, 303–308. [Google Scholar] [CrossRef]
  32. Noh, G.; Shi, Z.; Zones, S.I.; Iglesia, E. Isomerization and β-scission reactions of alkanes on bifunctional metal-acid catalysts: Consequences of confinement and diffusional constraints on reactivity and selectivity. J. Catal. 2018, 368, 389–410. [Google Scholar] [CrossRef]
  33. Qi, Y.; Cai, C.-F.; Sun, P.; Wang, D.-W.; Zhu, H.-J. Crude oil cracking in deep reservoirs: A review of the controlling factors and estimation methods. Pet. Sci. 2023, 20, 1978–1997. [Google Scholar] [CrossRef]
  34. Hu, W.; Noh, G.; Iglesia, E. Consequences of metal-acid site proximity for alkane isomerization and β-scission mediated by bifunctional catalytic cascades. J. Catal. 2023, 425, 125–142. [Google Scholar] [CrossRef]
  35. Wu, X.; Lei, X.; Wang, S.; Zhang, H.; Cui, Q.; Bao, X.; Wang, T.; Yuan, P. Tuning the acidity and pore structures of NiP based bifunctional catalyst for efficient n-hexane isomerization. Fuel 2022, 324, 124396. [Google Scholar] [CrossRef]
  36. Voloshyna, Y.; Pertko, O.; Patrylak, L.; Yakovenko, A. Peculiarities of palladium-containing MFI-type zeolites as catalysts of isomerization of linear alkanes. Vopr. Khimii I Khimicheskoi Tekhnol. 2020, 6, 26–32. [Google Scholar] [CrossRef]
  37. Leprince, P. Petroleum Refining. Vol. 3 Conversion Processes; Editions Technip: Paris, France, 2001; Volume 3. [Google Scholar]
  38. McKetta, J.J., Jr. Encyclopedia of Chemical Processing and Design; CRC Press: Boca Raton, FL, USA, 1997. [Google Scholar]
  39. Del Campo, P.; Martínez, C.; Corma, A. Activation and conversion of alkanes in the confined space of zeolite-type materials. Chem. Soc. Rev. 2021, 50, 8511–8595. [Google Scholar] [CrossRef] [PubMed]
  40. Gomes, L.C.; de Oliveira Rosas, D.; Chistone, R.C.; Zotin, F.M.Z.; de Araujo, L.R.R.; Zotin, J.L. Hydroisomerization of n-hexadecane using Pt/alumina-Beta zeolite catalysts for producing renewable diesel with low pour point. Fuel 2017, 209, 521–528. [Google Scholar] [CrossRef]
  41. Dhar, A.; Vekariya, R.L.; Bhadja, P. n-Alkane isomerization by catalysis—A method of industrial importance: An overview. Cogent Chem. 2018, 4, 1514686. [Google Scholar] [CrossRef]
  42. Morris, R.E.; Nachtigall, P. Zeolites in Catalysis: Properties and Applications; Royal Society of Chemistry: London, UK, 2017. [Google Scholar]
  43. Primo, A.; Garcia, H. Zeolites as catalysts in oil refining. Chem. Soc. Rev. 2014, 43, 7548–7561. [Google Scholar] [CrossRef]
  44. Gao, S.; Wu, Z. Advances in the One- Dimensional Pore Zeolite for Hydroisomerization. J. Petrochem. Univ. 2023, 36, 51. [Google Scholar] [CrossRef]
  45. Zhang, L.; Bai, X.; Fu, W.; Yang, X.; Sun, F.; He, L.; Tang, T. Superior activity and isomerization selectivity in n-dodecane hydroisomerization over Ni clusters on ZSM-22 zeolite with structural defects. Fuel 2023, 332, 126204. [Google Scholar] [CrossRef]
  46. Zhang, Y.; Liu, D.; Lou, B.; Yu, R.; Men, Z.; Li, M.; Li, Z. Hydroisomerization of n-decane over micro/mesoporous Pt-containing bifunctional catalysts: Effects of the MCM-41 incorporation with Y zeolite. Fuel 2018, 226, 204–212. [Google Scholar] [CrossRef]
  47. Csicsery, S.M. Shape-selective catalysis in zeolites. Zeolites 1984, 4, 202–213. [Google Scholar] [CrossRef]
  48. Marcilly, C.R. Where and how shape selectivity of molecular sieves operates in refining and petrochemistry catalytic processes. Top. Catal. 2000, 13, 357–366. [Google Scholar] [CrossRef]
  49. Degnan Jr, T.F. The implications of the fundamentals of shape selectivity for the development of catalysts for the petroleum and petrochemical industries. J. Catal. 2003, 216, 32–46. [Google Scholar] [CrossRef]
  50. Han, Y.; Yuan, J.; Xing, M.; Cao, J.; Chen, Z.; Zhang, L.; Tao, Z.; Liu, Z.; Zheng, A.; Wen, X. Shape selectivity of zeolite for hydroisomerization of long-chain alkanes. New J. Chem. 2023, 47, 1401–1412. [Google Scholar] [CrossRef]
  51. Wang, M.; Han, Y.; Liu, S.; Liu, Z.; An, D.; Zhang, Z.; Cheng, K.; Zhang, Q.; Wang, Y. Pore-mouth catalysis boosting the formation of iso-paraffins from syngas over bifunctional catalysts. Chin. J. Catal. 2021, 42, 2197–2205. [Google Scholar] [CrossRef]
  52. Brito, L.; Pirngruber, G.D.; Perez-Pellitero, J.; Guillon, E.; Albrieux, F.; Martens, J.A. Shape selectivity effects in the hydroconversion of perhydrophenanthrene over bifunctional catalysts. Catal. Sci. Technol. 2021, 11, 7667–7682. [Google Scholar] [CrossRef]
  53. Guo, Z.; Li, X.; Hu, S.; Ye, G.; Zhou, X.; Coppens, M.O. Understanding the role of internal diffusion barriers in Pt/Beta zeolite catalyzed isomerization of n-heptane. Angew. Chem. 2020, 132, 1564–1567. [Google Scholar] [CrossRef]
  54. Chen, N.; Lucki, S.; Mower, E. Cage effect on product distribution from cracking over crystalline aluminosilicate zeolites. J. Catal. 1969, 13, 329–332. [Google Scholar] [CrossRef]
  55. Kärger, J.; Ruthven, D.M.; Valiullin, R. Diffusion in nanopores: Inspecting the grounds. Adsorption 2021, 27, 267–281. [Google Scholar] [CrossRef]
  56. Derouane, E.G. Shape selectivity in catalysis by zeolites: The nest effect. J. Catal. 1986, 100, 541–544. [Google Scholar] [CrossRef]
  57. Romero, D.; Rohling, R.; Meng, L.; Rigutto, M.; Hensen, E.J. Shape selectivity in linear paraffins hydroconversion in 10-membered-ring pore zeolites. J. Catal. 2021, 394, 284–298. [Google Scholar] [CrossRef]
  58. Yu, Q.; Huang, Z.; Sun, H.; Li, L.; Zhu, X.; Ren, S.; Shen, B. Investigation on n-Alkane Hydroisomerization, a Comparison of IM-5 to ZSM-5 Zeolites. Ind. Eng. Chem. Res. 2018, 57, 14448–14459. [Google Scholar] [CrossRef]
  59. Mehla, S.; Krishnamurthy, K.; Viswanathan, B.; John, M.; Niwate, Y.; Kumar, K.; Pai, S.M.; Newalkar, B. n-Hexadecane hydroisomerization over Pt/ZSM-12: Role of Si/Al ratio on product distribution. J. Porous Mater. 2013, 20, 1023–1029. [Google Scholar] [CrossRef]
  60. Araujo, A.S.; Silva, A.O.; Souza, M.J.; Coutinho, A.C.; Aquino, J.M.; Moura, J.A.; Pedrosa, A.M. Crystallization of ZSM-12 zeolite with different Si/Al ratio. Adsorption 2005, 11, 159–165. [Google Scholar] [CrossRef]
  61. Li, S.; Liutkova, A.; Kosinov, N.; Hensen, E.J. Zeolite Nanocrystals (MOR, EU-1, and ZSM-12) Synthesized Using a Versatile Diquaternary Ammonium Template as Robust Catalysts. ACS Appl. Nano Mater. 2022, 5, 16862–16871. [Google Scholar] [CrossRef]
  62. Gopal, S.; Yoo, K.; Smirniotis, P.G. Synthesis of Al-rich ZSM-12 using TEAOH as template. Microporous Mesoporous Mater. 2001, 49, 149–156. [Google Scholar] [CrossRef]
  63. Mehla, S.; Krishnamurthy, K.; Viswanathan, B.; John, M.; Niwate, Y.; Kumar, S.K.; Pai, S.M.; Newalkar, B.L. n-Hexadecane hydroisomerization over BTMACl/TEABr/MTEABr templated ZSM-12. Microporous Mesoporous Mater. 2013, 177, 120–126. [Google Scholar] [CrossRef]
  64. Masoumifard, N.; Kaliaguine, S.; Kleitz, F. Synergy between structure direction and alkalinity toward fast crystallization, controlled morphology and high phase purity of ZSM-12 zeolite. Microporous Mesoporous Mater. 2016, 227, 258–271. [Google Scholar] [CrossRef]
  65. Li, S.; Mezari, B.; Wu, H.; Kosinov, N.; Hensen, E.J. ZSM-12 nanocrystals with tunable acidity directed by rigid diquats: Synthesis and catalytic applications. Fuel 2023, 333, 126363. [Google Scholar] [CrossRef]
  66. Silva, B.J.; de Sousa, L.V.; Quintela, P.H.; Júnior, N.R.A.; Alencar, S.L.; Maciel, P.A.; Santos, J.R.; Sarmento, L.R.; Meneghetti, S.M.; Silva, A. Preparation of ZSM-22 zeolite with hierarchical pore structure. Mater. Lett. 2018, 218, 119–122. [Google Scholar] [CrossRef]
  67. Verboekend, D.; Chabaneix, A.M.; Thomas, K.; Gilson, J.-P.; Pérez-Ramírez, J. Mesoporous ZSM-22 zeolite obtained by desilication: Peculiarities associated with crystal morphology and aluminium distribution. CrystEngComm 2011, 13, 3408–3416. [Google Scholar] [CrossRef]
  68. Niu, P.; Xi, H.; Ren, J.; Lin, M.; Wang, Q.; Jia, L.; Hou, B.; Li, D. High selectivity for n-dodecane hydroisomerization over highly siliceous ZSM-22 with low Pt loading. Catal. Sci. Technol. 2017, 7, 5055–5068. [Google Scholar] [CrossRef]
  69. Wang, Y.; Yu, J.; Yang, F.; Zhao, Y.; Zhuang, J.; Lu, X.; Zhu, X.; Zhu, K. Synthesis of ZSM-22 with Enhanced Acidity Using Alcoholamine as a Structure-Directing Agent to Promote the Hydroisomerization of n-Heptane. Ind. Eng. Chem. Res. 2023, 62, 10012–10023. [Google Scholar] [CrossRef]
  70. Wang, B.; Tian, Z.; Li, P.; Wang, L.; Xu, Y.; Qu, W.; He, Y.; Ma, H.; Xu, Z.; Lin, L. A novel approach to synthesize ZSM-23 zeolite involving N,N-dimethylformamide. Microporous Mesoporous Mater. 2010, 134, 203–209. [Google Scholar] [CrossRef]
  71. Molino, A.; Łukaszuk, K.; Rojo-Gama, D.; Lillerud, K.P.; Olsbye, U.; Bordiga, S.; Svelle, S.; Beato, P. Conversion of methanol to hydrocarbons over zeolite ZSM-23 (MTT): Exceptional effects of particle size on catalyst lifetime. Chem. Comm. 2017, 53, 6816–6819. [Google Scholar] [CrossRef]
  72. Chen, Y.; Li, C.; Chen, X.; Liu, Y.; Liang, C. Synthesis of ZSM-23 zeolite with dual structure directing agents for hydroisomerization of n-hexadecane. Microporous Mesoporous Mater. 2018, 268, 216–224. [Google Scholar] [CrossRef]
  73. Ahmed, M.H.; Muraza, O.; Yoshioka, M.; Yokoi, T. Effect of multi-step desilication and dealumination treatments on the performance of hierarchical EU-1 zeolite for converting methanol to olefins. Microporous Mesoporous Mater. 2017, 241, 79–88. [Google Scholar] [CrossRef]
  74. Du, Y.; Feng, B.; Jiang, Y.; Yuan, L.; Huang, K.; Li, J. Solvent-Free Synthesis and n-Hexadecane Hydroisomerization Performance of SAPO-11 Catalyst. Eur. J. Inorg. Chem. 2018, 2018, 2599–2606. [Google Scholar] [CrossRef]
  75. Zhang, Y.; Wang, W.; Jiang, X.; Su, X.; Kikhtyanin, O.; Wu, W. Hydroisomerization of n-hexadecane over a Pd–Ni2P/SAPO-31 bifunctional catalyst: Synergistic effects of bimetallic active sites. Catal. Sci. Technol. 2018, 8, 817–828. [Google Scholar] [CrossRef]
  76. Bai, X.; Wei, X.; Liu, Y.; Wu, W. Effect of dealumination of citric acid-modified Hβ molecular sieve on hydroisomerization performance of n-Hexadecane. In Proceedings of the IOP Conference Series: Materials Science and Engineering, The 6th Annual International Conference on Material Science and Environmental Engineering, Chongqing, China, 23–25 November 2018; p. 012040. [Google Scholar]
  77. Bai, X.; Wei, X.; Liu, Y.; Wu, W. Hydroisomerization of n-hexadecane over Hβ molecular sieve loading palladium bifunctional catalyst: Effect of SiO2/Al2O3 molar ratios. In Proceedings of the IOP Conference Series: Materials Science and Engineering, The 2nd International Workshop on Materials Science and Mechanical Engineering (IWMSME2018), Qingdao, China, 26–28 October 2018; p. 012042. [Google Scholar]
  78. Smirniotis, P.G.; Ruckenstein, E. Comparison of the Performance of ZSM-5,.beta. Zeolite, Y, USY, and Their Composites in the Catalytic Cracking of n-Octane, 2,2,4-Trimethylpentane, and 1-Octene. Ind. Eng. Chem. Res. 1994, 33, 800–813. [Google Scholar] [CrossRef]
  79. Zhang, J.; Maximov, A.; Bai, X.; Wang, W.; Xiao, L.; Lin, H.; Wu, W. Shape selectivity in hydroisomerization of n-Hexadecane over Pd supported on zeolites: ZSM-22, ZSM-12 and Beta. Russ. J. Appl. Chem. 2020, 93, 1427–1437. [Google Scholar] [CrossRef]
  80. Zhang, M.; Chen, Y.; Wang, L.; Zhang, Q.; Tsang, C.-W.; Liang, C. Shape selectivity in hydroisomerization of hexadecane over Pt supported on 10-ring zeolites: ZSM-22, ZSM-23, ZSM-35, and ZSM-48. Ind. Eng. Chem. Res. 2016, 55, 6069–6078. [Google Scholar] [CrossRef]
  81. Liu, Y.; Pinnavaia, T.J. Aluminosilicate nanoparticles for catalytic hydrocarbon cracking. J. Am. Chem. Soc. 2003, 125, 2376–2377. [Google Scholar] [CrossRef] [PubMed]
  82. Karlsson, A.; Stoecker, M.; Schmidt, R. Composites of micro-and mesoporous materials: Simultaneous syntheses of MFI/MCM-41 like phases by a mixed template approach. Microporous Mesoporous Mater. 1999, 27, 181–192. [Google Scholar] [CrossRef]
  83. Cheng, K.; van der Wal, L.I.; Yoshida, H.; Oenema, J.; Harmel, J.; Zhang, Z.; Sunley, G.; Zečević, J.; de Jong, K.P. Impact of the spatial organization of bifunctional metal–zeolite catalysts on the hydroisomerization of light alkanes. Angew. Chem. 2020, 132, 3620–3628. [Google Scholar] [CrossRef]
  84. Chen, Z.; Liu, S.; Wang, H.; Ning, Q.; Zhang, H.; Yun, Y.; Ren, J.; Li, Y.-W. Synthesis and characterization of bundle-shaped ZSM-22 zeolite via the oriented fusion of nanorods and its enhanced isomerization performance. J. Catal. 2018, 361, 177–185. [Google Scholar] [CrossRef]
  85. Morávková, J.; Pilař, R.; Bortnovsky, O.; Kaucký, D.; Vondrová, A.; Rathousky, J.; Sádovská, G.; Sazama, P. The effect of the nanoscale intimacy of platinum and acid centres on the hydroisomerization of short-chain alkanes. Appl. Catal. A Gen. 2022, 634, 118535. [Google Scholar] [CrossRef]
  86. Ye, G.; Sun, Y.; Guo, Z.; Zhu, K.; Liu, H.; Zhou, X.; Coppens, M.-O. Effects of zeolite particle size and internal grain boundaries on Pt/Beta catalyzed isomerization of n-pentane. J. Catal. 2018, 360, 152–159. [Google Scholar] [CrossRef]
  87. Wang, W.; Liu, C.-J.; Wu, W. Bifunctional catalysts for the hydroisomerization of n-alkanes: The effects of metal–acid balance and textural structure. Catal. Sci. Technol. 2019, 9, 4162–4187. [Google Scholar] [CrossRef]
  88. Wang, Y.; Tao, Z.; Wu, B.; Xu, J.; Huo, C.; Li, K.; Chen, H.; Yang, Y.; Li, Y. Effect of metal precursors on the performance of Pt/ZSM-22 catalysts for n-hexadecane hydroisomerization. J. Catal. 2015, 322, 1–13. [Google Scholar] [CrossRef]
  89. Geng, L.; Gong, J.; Qiao, G.; Ye, S.; Zheng, J.; Zhang, N.; Chen, B. Effect of metal precursors on the performance of Pt/SAPO-11 catalysts for n-dodecane hydroisomerization. ACS Omega 2019, 4, 12598–12605. [Google Scholar] [CrossRef] [PubMed]
  90. Chen, H.; Yi, F.; Ma, C.; Gao, X.; Liu, S.; Tao, Z.; Wu, B.; Xiang, H.; Yang, Y.; Li, Y.-W. Hydroisomerization of n-heptane on a new kind of bifunctional catalysts with palladium nanoparticles encapsulating inside zeolites. Fuel 2020, 268, 117241. [Google Scholar] [CrossRef]
  91. Guisnet, M. “Ideal” bifunctional catalysis over Pt-acid zeolites. Catal. Today 2013, 218, 123–134. [Google Scholar] [CrossRef]
  92. Wang, Y.; Tao, Z.; Wu, B.; Huimin, C.; Xu, J.; Yang, Y.; Li, Y. Shape-controlled synthesis of Pt particles and their catalytic performances in the n-hexadecane hydroconversion. Catal. Today 2016, 259, 331–339. [Google Scholar] [CrossRef]
  93. Roy, S.; Bouchy, C.; Pham-Huu, C.; Crouzet, C.; Ledoux, M.J. Slurry isomerization of n-hexadecane over MoO3-carbon-modified, Pt/β-zeolite, Pt/ZSM-22 and Pt/SAPO-11 catalysts at medium pressure. In Studies in Surface Science and Catalysis; Corma, A., Melo, F.V., Mendiorozm, S., Fierro, J.L.G., Eds.; Elsevier: Amsterdam, The Netherlands, 2000; Volume 130, pp. 2423–2428. [Google Scholar]
  94. Fedyna, M.; Żak, A.; Jaroszewska, K.; Mokrzycki, J.; Trawczyński, J. Composite of Pt/AlSBA-15+ zeolite catalyst for the hydroisomerization of n-hexadecane: The effect of platinum precursor. Microporous Mesoporous Mater. 2020, 305, 110366. [Google Scholar] [CrossRef]
  95. Chen, Y.; Meng, J.; Bai, D.; Li, C.; Chen, X.; Liang, C. Effect of Extra-Framework Fe Species in Pt/Fe/ZSM-23 Catalysts on Hydroisomerization Performance of n-Hexadecane. Ind. Eng. Chem. Res. 2021, 61, 279–286. [Google Scholar] [CrossRef]
  96. Song, C.Y.; Meng, J.P.; Chuang, L.I.; Zeyaodong, P.; Liang, C.H. Synthesis of ZSM-22/ZSM-23 intergrowth zeolite as the catalyst support for hydroisomerization of n-hexadecane. J. Fuel Chem. Technol. 2021, 49, 712–725. [Google Scholar] [CrossRef]
  97. Zhang, M.; Li, C.; Chen, X.; Chen, Y.; Liang, C. Hierarchical ZSM-48-supported nickel catalysts with enhanced hydroisomerization performance of hexadecane. Ind. Eng. Chem. Res. 2019, 58, 19855–19861. [Google Scholar] [CrossRef]
  98. Zhang, M.; Wang, L.; Chen, Y.; Zhang, Q.; Liang, C. Creating mesopores in ZSM-48 zeolite by alkali treatment: Enhanced catalyst for hydroisomerization of hexadecane. J. Energy Chem. 2016, 25, 539–544. [Google Scholar] [CrossRef]
  99. Pimerzin, A.A.; Roganov, A.A.; Verevkin, S.P.; Konnova, M.E.; Pilshchikov, V.A.; Pimerzin, A.A. Bifunctional catalysts with noble metals on composite Al2O3-SAPO-11 carrier and their comparison with CoMoS one in n-hexadecane hydroisomerization. Catal. Today 2019, 329, 71–81. [Google Scholar] [CrossRef]
  100. Zhao, X.; Xu, J.; Deng, F. Solid-state NMR for metal-containing zeolites: From active sites to reaction mechanism. Front. Chem. Sci. Eng. 2020, 14, 159–187. [Google Scholar] [CrossRef]
  101. Schulman, E.; Wu, W.; Liu, D. Two-dimensional zeolite materials: Structural and acidity properties. Materials 2020, 13, 1822. [Google Scholar] [CrossRef]
  102. Almutairi, S.M.; Mezari, B.; Filonenko, G.A.; Magusin, P.C.; Rigutto, M.S.; Pidko, E.A.; Hensen, E.J. Influence of extraframework aluminum on the brønsted acidity and catalytic reactivity of faujasite zeolite. ChemCatChem 2013, 5, 452–466. [Google Scholar] [CrossRef]
  103. Izan, S.; Triwahyono, S.; Jalil, A.; Majid, Z.; Fatah, N.; Hamid, M.; Ibrahim, M. Additional Lewis acid sites of protonated fibrous silica@BEA zeolite (HSi@BEA) improving the generation of protonic acid sites in the isomerization of C6 alkane and cycloalkanes. Appl. Catal. A Gen. 2019, 570, 228–237. [Google Scholar] [CrossRef]
  104. Vogt, E.T.; Weckhuysen, B.M. Fluid catalytic cracking: Recent developments on the grand old lady of zeolite catalysis. Chem. Soc. Rev. 2015, 44, 7342–7370. [Google Scholar] [CrossRef]
  105. Ennaert, T.; Van Aelst, J.; Dijkmans, J.; De Clercq, R.; Schutyser, W.; Dusselier, M.; Verboekend, D.; Sels, B.F. Potential and challenges of zeolite chemistry in the catalytic conversion of biomass. Chem. Soc. Rev. 2016, 45, 584–611. [Google Scholar] [CrossRef]
  106. Dzhikiya, O.; Smolikov, M.; Belopukhov, E.; Yablokova, S.; Doronin, A.; Gulyaeva, T.; Muromtsev, I.; Belyi, A. A study of fluorine-containing catalysts based on MOR and BEA zeolites in isomerization of n-hexane. In Proceedings of the AIP Conference Proceedings, Oil and Gas Engineering (OGE-2020), Omsk, Russia, 26–29 February 2020. [Google Scholar]
  107. Sun, J.; Mu, C.; Guo, D.; Zhao, Y.; Wang, S.; Ma, X. Effects of intimacy between acid and metal sites on the isomerization of n-C16 at the large/minor nanoscale and atomic scale. ACS Catal. 2022, 12, 4092–4102. [Google Scholar] [CrossRef]
  108. Harmel, J.; van der Wal, L.; Zečević, J.; de Jongh, P.; de Jong, K.P. Metal-Acid Bifunctional Catalysts for Hydro-Isomerization of Alkanes: Impact of Intimacy for Mesoporous Solid Acids. In Proceedings of the 2019 North American Catalysis Society Meeting, Chicago, IL, USA, 23–28 June 2019. [Google Scholar]
  109. Chen, Y.; Li, C.; Chen, X.; Liu, Y.; Tsang, C.-W.; Liang, C. Synthesis and characterization of iron-substituted ZSM-23 zeolite catalysts with highly selective hydroisomerization of n-hexadecane. Ind. Eng. Chem. Res. 2018, 57, 13721–13730. [Google Scholar] [CrossRef]
  110. Blasco, T.; Chica, A.; Corma, A.; Murphy, W.; Agúndez-Rodríguez, J.; Pérez-Pariente, J. Changing the Si distribution in SAPO-11 by synthesis with surfactants improves the hydroisomerization/dewaxing properties. J. Catal. 2006, 242, 153–161. [Google Scholar] [CrossRef]
  111. Gopal, S.; Smirniotis, P.G. Factors affecting isomer yield for n-heptane hydroisomerization over as-synthesized and dealuminated zeolite catalysts loaded with platinum. J. Catal. 2004, 225, 278–287. [Google Scholar] [CrossRef]
  112. Yang, X.; Zhao, W.; Liu, L.; Niu, X.; Wang, Q. Tuning the Structure and Acidity of Pt/Hierarchical SSZ-32 Catalysts to Boost the Selective Hydroisomerization of n-Hexadecane. Catalysts 2023, 13, 702. [Google Scholar] [CrossRef]
  113. Li, T.; Wang, W.; Feng, Z.; Bai, X.; Su, X.; Yang, L.; Jia, G.; Guo, C.; Wu, W. The hydroisomerization of n-hexane over highly selective Pd/ZSM-22 bifunctional catalysts: The improvements of metal-acid balance by room temperature electron reduction method. Fuel 2020, 272, 117717. [Google Scholar] [CrossRef]
  114. Wei, C.; Zhang, G.; Zhao, L.; Gao, J.; Xu, C. Effect of metal–acid balance and textual modifications on hydroisomerization catalysts for n-alkanes with different chain length: A mini-review. Fuel 2022, 315, 122809. [Google Scholar] [CrossRef]
  115. Batalha, N.; Pinard, L.; Bouchy, C.; Guillon, E.; Guisnet, M. n-Hexadecane hydroisomerization over Pt-HBEA catalysts. Quantification and effect of the intimacy between metal and protonic sites. J. Catal. 2013, 307, 122–131. [Google Scholar] [CrossRef]
  116. Parmar, S.; Pant, K.K.; John, M.; Kumar, K.; Pai, S.M.; Gupta, S.; Newalkar, B.L. Hydroisomerization of n-hexadecane over Brønsted acid site tailored Pt/ZSM-12. J. Porous Mater. 2014, 21, 849–857. [Google Scholar] [CrossRef]
  117. Lee, S.-W.; Ihm, S.-K. Characteristics of magnesium-promoted Pt/ZSM-23 catalyst for the hydroisomerization of n-hexadecane. Ind. Eng. Chem. Res. 2013, 52, 15359–15365. [Google Scholar] [CrossRef]
  118. Liu, S.; Ren, J.; Zhu, S.; Zhang, H.; Lv, E.; Xu, J.; Li, Y.-W. Synthesis and characterization of the Fe-substituted ZSM-22 zeolite catalyst with high n-dodecane isomerization performance. J. Catal. 2015, 330, 485–496. [Google Scholar] [CrossRef]
  119. Batalha, N.; Morisset, S.; Pinard, L.; Maupin, I.; Lemberton, J.; Lemos, F.; Pouilloux, Y. BEA zeolite nanocrystals dispersed over alumina for n-hexadecane hydroisomerization. Microporous Mesoporous Mater. 2013, 166, 161–166. [Google Scholar] [CrossRef]
  120. Parmar, S.; Pant, K.K.; John, M.; Kumar, K.; Pai, S.M.; Newalkar, B.L. Hydroisomerization of n-hexadecane over Pt/ZSM-22 framework: Effect of divalent cation exchange. J. Mol. Catal. A Chem. 2015, 404, 47–56. [Google Scholar] [CrossRef]
  121. Liu, L.; Zhang, M.; Wang, L.; Zhang, X.; Li, G. Modulating acid site distribution in MTT channels for controllable hydroisomerization of long-chain n-alkanes. Fuel Process. Technol. 2023, 241, 107605. [Google Scholar] [CrossRef]
  122. Zhang, M.; Li, C.; Chen, Y.; Tsang, C.-W.; Zhang, Q.; Liang, C. Hydroisomerization of hexadecane over platinum supported on EU-1/ZSM-48 intergrowth zeolite catalysts. Catal. Sci. Technol. 2016, 6, 8016–8023. [Google Scholar] [CrossRef]
  123. Wang, D.; Liu, J.; Cheng, X.; Kang, X.; Wu, A.; Tian, C.; Fu, H. Trace Pt Clusters Dispersed on SAPO-11 Promoting the Synergy of Metal Sites with Acid Sites for High-Effective Hydroisomerization of n-Alkanes. Small Methods 2019, 3, 1800510. [Google Scholar] [CrossRef]
  124. Guo, C.; Wang, W.; Zhang, Y.; Lin, H.; Jia, G.; Li, T.; Xin, Q.; Bai, X.; Wu, W. Influences of the metal-acid proximity of Pd-SAPO-31 bifunctional catalysts for n-hexadecane hydroisomerization. Fuel Process. Technol. 2021, 214, 106717. [Google Scholar] [CrossRef]
  125. Wei, X.; Kikhtyanin, O.; Parmon, V.; Wu, W.; Bai, X.; Zhang, J.; Xiao, L.; Su, X.; Zhang, Y. Synergetic effect between the metal and acid sites of Pd/SAPO-41 bifunctional catalysts in n-hexadecane hydroisomerization. J. Porous Mater. 2018, 25, 235–247. [Google Scholar] [CrossRef]
  126. Jia, G.; Guo, C.; Wang, W.; Bai, X.; Wei, X.; Su, X.; Li, T.; Xiao, L.; Wu, W. The synergic effects of highly selective bimetallic Pt-Pd/SAPO-41 catalysts for the n-hexadecane hydroisomerization. Front. Chem. Sci. Eng. 2021, 15, 1111–1124. [Google Scholar] [CrossRef]
  127. Wu, Q.; Jia, G.; Zhang, Y.; Liu, Z.; Li, X.; Wang, W.; Wu, W. Bifunctional catalysts based on hierarchical SAPO-41 nanosheet for Highly-efficient hydroisomerization of n-Hexadecane. Fuel 2023, 352, 129066. [Google Scholar] [CrossRef]
  128. Mendes, P.S.; Mota, F.M.; Silva, J.M.; Ribeiro, M.F.; Daudin, A.; Bouchy, C. A systematic study on mixtures of Pt/zeolite as hydroisomerization catalysts. Catal. Sci. Technol. 2017, 7, 1095–1107. [Google Scholar] [CrossRef]
  129. Bauer, F.; Ficht, K.; Bertmer, M.; Einicke, W.-D.; Kuchling, T.; Gläser, R. Hydroisomerization of long-chain paraffins over nano-sized bimetallic Pt–Pd/H-beta catalysts. Catal. Sci. Technol. 2014, 4, 4045–4054. [Google Scholar] [CrossRef]
  130. Vinogradov, N.; Rubtsova, M.; Glotov, A.; Tochilin, N.; Vinokurov, V.; Pimerzin, A. Hydroconversion of n-Hexadecane on Zeolite-Containing Sulfide-Based Catalysts: Influence of Nitrogen Impurity in the Feedstock on the Hydroisomerization Selectivity. Pet. Chem. 2021, 61, 739–747. [Google Scholar] [CrossRef]
  131. Jaroszewska, K.; Fedyna, M.; Trawczyński, J. Hydroisomerization of long-chain n-alkanes over Pt/AlSBA-15+ zeolite bimodal catalysts. Appl. Catal. B Environ. 2019, 255, 117756. [Google Scholar] [CrossRef]
  132. Bogomolova, T.; Smirnova, M.Y.; Klimov, O.; Noskov, A. Studying the Hydroisomerization of Diesel Fractions with Different Concentrations of Nitrogen-Containing Compounds on Bifunctional Catalysts Based on ZSM-23 and Non-Noble Metals. Catal. Ind. 2023, 15, 182–189. [Google Scholar] [CrossRef]
  133. Chen, N.; Degnan, T.; Lutner, J.; Pelrine, B. The deactivation of ZSM-5 in catalytic dewaxing. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 1991; Volume 68, pp. 773–782. [Google Scholar]
  134. Galperin, L. Hydroisomerization of N-decane in the presence of sulfur and nitrogen compounds. Appl. Catal. A Gen. 2001, 209, 257–268. [Google Scholar] [CrossRef]
  135. Lee, S.-W.; Ihm, S.-K. Hydroisomerization and hydrocracking over platinum loaded ZSM-23 catalysts in the presence of sulfur and nitrogen compounds for the dewaxing of diesel fuel. Fuel 2014, 134, 237–243. [Google Scholar] [CrossRef]
  136. Pimerzin, A.; Savinov, A.; Vutolkina, A.; Makova, A.; Glotov, A.; Vinokurov, V.; Pimerzin, A. Transition metal sulfides-and noble metal-based catalysts for N-hexadecane hydroisomerization: A study of poisons tolerance. Catalysts 2020, 10, 594. [Google Scholar] [CrossRef]
  137. Iskenderova, A. Mechanism of Catalytic Isomerization-Disproportionation Processing of Straight-Run Gasoline. Azerbaijan Chem. J. 2022, 2, 28–33. [Google Scholar] [CrossRef]
  138. Sun, J.; Yu, Y.; Mu, C.; Guo, X.; Zhao, Y.; Wang, S.; Ma, X. The gradient fabrication and optimization of metal-acid bifunctional catalysts for the hydrogenation of n-C16 to produce bio-jet fuel component. Fuel 2023, 353, 129229. [Google Scholar] [CrossRef]
  139. Pham, T.N.; Nguyen, V.; Wang, B.; White, J.L.; Crossley, S. Quantifying the influence of water on the mobility of aluminum species and their effects on alkane cracking in zeolites. ACS Catal. 2021, 11, 6982–6994. [Google Scholar] [CrossRef]
  140. Yan, T. The promotional effect of water in hydrocracking. J. Catal. 1972, 25, 204–211. [Google Scholar] [CrossRef]
  141. Brosius, R.; Kooyman, P.J.; Fletcher, J.C. Selective formation of linear alkanes from n-hexadecane primary hydrocracking in shape-selective MFI zeolites by competitive adsorption of water. ACS Catal. 2016, 6, 7710–7715. [Google Scholar] [CrossRef]
  142. Guisnet, M.; Magnoux, P. Coking and deactivation of zeolites: Influence of the pore structure. Appl. Catal. 1989, 54, 1–27. [Google Scholar] [CrossRef]
  143. Guisnet, M. “Coke” molecules trapped in the micropores of zeolites as active species in hydrocarbon transformations. J. Mol. Catal. A Chem. 2002, 182, 367–382. [Google Scholar] [CrossRef]
  144. Sun, Y.; Wei, L.; Zhang, Z.; Zhang, H.; Li, Y. Coke formation over zeolite catalysts in light alkanes aromatization and anti-carbon-deposition strategies and perspectives: A review. Energy Fuel 2023, 37, 1657–1677. [Google Scholar] [CrossRef]
  145. Guisnet, M.; Magnoux, P. Deactivation of zeolites by coking. Prevention of deactivation and regeneration. In Zeolite Microporous Solids: Synthesis, Structure, and Reactivity; Springer: Dordrecht, The Netherlands, 1992; pp. 457–474. [Google Scholar]
  146. Nazarova, G.Y.; Ivashkina, E.N.; Ivanchina, E.D.; Vosmerikov, A.V.; Vosmerikova, L.N.; Antonov, A.V. A model of catalytic cracking: Product distribution and catalyst deactivation depending on saturates, aromatics and resins content in feed. Catalysts 2021, 11, 701. [Google Scholar] [CrossRef]
  147. Muhammad, I.; Makwashi, N.; Ahmed, T.G.; Manos, G.; Zhao, D. A Mechanistic Model on Catalyst Deactivation by Coke Formation in a CSTR Reactor. Processes 2023, 11, 944. [Google Scholar] [CrossRef]
  148. Guisnet, M.; Magnoux, P. Organic chemistry of coke formation. Appl. Catal. A Gen. 2001, 212, 83–96. [Google Scholar] [CrossRef]
  149. Bauer, F.; Karge, H.G. Characterization of coke on zeolites. In Characterization II; Springer: Berlin/Heidelberg, Germany, 2006; pp. 249–364. [Google Scholar]
  150. Guisnet, M.; Magnoux, P. Deactivation by coking of zeolite catalysts. Prevention of deactivation. Optimal conditions for regeneration. Catal. Today 1997, 36, 477–483. [Google Scholar] [CrossRef]
  151. Guisnet, M.; Costa, L.; Ribeiro, F.R. Prevention of zeolite deactivation by coking. J. Mol. Catal. A Chem. 2009, 305, 69–83. [Google Scholar] [CrossRef]
  152. Shehab, A. The Role of Carbon in the Catalytic Isomerisation-Cracking of n-Alkanes. Ph.D. Thesis, University of Sheffield, Sheffield, UK, 2018. [Google Scholar]
  153. Brillis, A.A.; Manos, G. Coke formation during catalytic cracking of C8 aliphatic hydrocarbons over ultrastable Y zeolite. Ind. Eng. Chem. Res. 2003, 42, 2292–2298. [Google Scholar] [CrossRef]
  154. Forzatti, P.; Lietti, L. Catalyst deactivation. Catal. Today 1999, 52, 165–181. [Google Scholar] [CrossRef]
  155. Rombi, E.; Monaci, R.; Solinas, V. Kinetics of catalyst deactivation. An example: Methylnaphthalene transformation. Catal. Today 1999, 52, 321–330. [Google Scholar] [CrossRef]
  156. Richardson, S.M.; Nagaishi, H.; Gray, M.R. Initial coke deposition on a NiMo/γ-Al2O3 bitumen hydroprocessing catalyst. Ind. Eng. Chem. Res. 1996, 35, 3940–3950. [Google Scholar] [CrossRef]
  157. Muegge, B.; Massoth, F. Basic studies of deactivation of hydrotreating catalysts with anthracene. Fuel Process. Technol. 1991, 29, 19–30. [Google Scholar] [CrossRef]
  158. Zhang, X.; Chodakowski, M.; Shaw, J.M. Impact of multiphase behavior on coke deposition in a commercial hydrotreating catalyst under sedimentation conditions. Energy Fuel 2005, 19, 1405–1411. [Google Scholar] [CrossRef]
  159. Chen, S.; Manos, G. Study of coke and coke precursors during catalytic cracking of n-hexane and 1-hexene over ultrastable Y zeolite. Catal. Lett. 2004, 96, 195–200. [Google Scholar] [CrossRef]
  160. Chen, S.; Manos, G. In situ thermogravimetric study of coke formation during catalytic cracking of normal hexane and 1-hexene over ultrastable Y zeolite. J. Catal. 2004, 226, 343–350. [Google Scholar] [CrossRef]
  161. Holm, M.S.; Taarning, E.; Egeblad, K.; Christensen, C.H. Catalysis with hierarchical zeolites. Catal. Today 2011, 168, 3–16. [Google Scholar] [CrossRef]
  162. Da Rocha, J.D.G.; Macuvele, D.L.P.; de Andrade, C.J.; Riella, H.G.; Padoin, N.; Soares, C. Advances and Environmental Aspects on the Synthesis of Hierarchical Zeolites Revisited: A State-of-the-Art Description. J. Environ. Chem. Eng. 2023, 11, 109397. [Google Scholar] [CrossRef]
  163. Bornes, C.; Santos-Vieira, I.C.; Vieira, R.; Mafra, L.; Simões, M.M.; Rocha, J. Challenges and opportunities for zeolites in biomass upgrading: Impediments and future directions. Catal. Today 2023, 419, 114159. [Google Scholar] [CrossRef]
  164. Hu, G.; Yang, J.; Duan, X.; Farnood, R.; Yang, C.; Yang, J.; Liu, W.; Liu, Q. Recent developments and challenges in zeolite-based composite photocatalysts for environmental applications. Chem. Eng. J. 2021, 417, 129209. [Google Scholar] [CrossRef]
  165. Gosselink, R.W.; Sagala, S.L.; Meeldijk, J.D.; de Jongh, P.E.; de Jong, K.P. Alkaline treatment on commercially available aluminum rich mordenite. Appl. Catal. A Gen. 2010, 382, 65–72. [Google Scholar] [CrossRef]
  166. Sazama, P.; Pastvova, J.; Kaucky, D.; Moravkova, J.; Rathousky, J.; Jakubec, I.; Sadovska, G. Does hierarchical structure affect the shape selectivity of zeolites? Example of transformation of n-hexane in hydroisomerization. J. Catal. 2018, 364, 262–270. [Google Scholar] [CrossRef]
  167. Meng, J.; Cui, T.; Bai, D.; Li, C.; Chen, X.; Liang, C. Excellent catalytic performance over hierarchical ZSM-48 zeolite: Cooperative effects of enhanced mesoporosity and highly-accessible acidity. Fuel 2022, 324, 124589. [Google Scholar] [CrossRef]
  168. Gackowski, M.; Tarach, K.; Podobiński, J.; Jarczewski, S.; Kuśtrowski, P.; Datka, J. Hierarchical zeolites Y obtained by desilication: Porosity, acidity and catalytic properties. Microporous Mesoporous Mater. 2018, 263, 282–288. [Google Scholar] [CrossRef]
  169. Aumond, T.; Rousseau, J.; Pouilloux, Y.; Pinard, L.; Sachse, A. Synthesis of hierarchical zeolite templated carbons. Carbon Trends 2021, 2, 100014. [Google Scholar] [CrossRef]
  170. Hu, B.; Zhang, Z.X.; Xie, W.L.; Liu, J.; Li, Y.; Zhang, W.M.; Fu, H.; Lu, Q. Advances on the fast pyrolysis of biomass for the selective preparation of phenolic compounds. Fuel Process. Technol. 2022, 237, 107465. [Google Scholar] [CrossRef]
  171. Yue, Y.; Hu, Y.; Dong, P.; Li, X.; Liu, H.; Bao, J.; Wang, T.; Bi, X.; Zhu, H.; Yuan, P. Mesoscale depolymerization of natural rectorite mineral via a quasi-solid-phase approach for zeolite synthesis. Chem. Eng. Sci. 2020, 220, 115635. [Google Scholar] [CrossRef]
  172. Imbachi-Gamba, C.F.; Villa, A. Statistical analysis of the influence of synthesis conditions on the properties of hierarchical zeolite Y. Mater. Today Chem. 2021, 20, 100442. [Google Scholar] [CrossRef]
  173. Jermy, B.R.; Tanimu, A.; Siddiqui, M.A.; Qureshi, Z.S.; Aitani, A.; Akah, A.; Xu, Q.; AlHerz, M. Crude oil conversion to chemicals over green synthesized ZSM-5 zeolite. Fuel Process. Technol. 2023, 241, 107610. [Google Scholar] [CrossRef]
  174. Ndlovu, N.Z.; Ameh, A.E.; Petrik, L.F.; Ojumu, T.V. Synthesis and characterisation of pure phase ZSM-5 and sodalite zeolites from coal fly ash. Mater. Today Commun. 2023, 34, 105436. [Google Scholar] [CrossRef]
  175. Nguyen, D.K.; Dinh, V.P.; Nguyen, H.Q.; Hung, N.T. Zeolite ZSM-5 synthesized from natural silica sources and its applications: A critical review. J. Chem. Technol. Biotechnol. 2023, 98, 1339–1355. [Google Scholar] [CrossRef]
  176. Demikhova, N.; Rubtsova, M.; Kireev, G.; Cherednichenko, K.; Vinokurov, V.; Glotov, A. Micro-mesoporous catalysts based on ZSM-5 zeolite synthesized from natural clay nanotubes: Preparation and application in the isomerization of C-8 aromatic fraction. Chem. Eng. J. 2023, 453, 139581. [Google Scholar] [CrossRef]
  177. Stavitskaya, A.; Rubtsova, M.; Glotov, A.; Vinokurov, V.; Vutolkina, A.; Fakhrullin, R.; Lvov, Y. Architectural design of core–shell nanotube systems based on aluminosilicate clay. Nanoscale Adv. 2022, 4, 2823–2835. [Google Scholar] [CrossRef] [PubMed]
  178. Glotov, A.; Levshakov, N.; Stavitskaya, A.; Artemova, M.; Gushchin, P.; Ivanov, E.; Vinokurov, V.; Lvov, Y. Templated self-assembly of ordered mesoporous silica on clay nanotubes. Chem. Commun. 2019, 55, 5507–5510. [Google Scholar] [CrossRef] [PubMed]
  179. Glotov, A.; Vutolkina, A.; Pimerzin, A.; Nedolivko, V.; Zasypalov, G.; Stytsenko, V.; Karakhanov, E.; Vinokurov, V. Ruthenium catalysts templated on mesoporous MCM-41 type silica and natural clay nanotubes for hydrogenation of benzene to cyclohexane. Catalysts 2020, 10, 537. [Google Scholar] [CrossRef]
  180. Glotov, A.; Vutolkina, A.; Pimerzin, A.; Vinokurov, V.; Lvov, Y. Clay nanotube-metal core/shell catalysts for hydroprocesses. Chem. Soc. Rev. 2021, 50, 9240–9277. [Google Scholar] [CrossRef]
  181. Al-Jubouri, S.M.; Al-Batty, S.I.; Holmes, S.M. Using the ash of common water reeds as a silica source for producing high purity ZSM-5 zeolite microspheres. Microporous Mesoporous Mater. 2021, 316, 110953. [Google Scholar] [CrossRef]
  182. Khoshbin, R.; Karimzadeh, R. The beneficial use of ultrasound in free template synthesis of nanostructured ZSM-5 zeolite from rice husk ash used in catalytic cracking of light naphtha: Effect of irradiation power. Adv. Powder Technol. 2017, 28, 973–982. [Google Scholar] [CrossRef]
  183. Wang, Y.; Jia, H.; Chen, P.; Fang, X.; Du, T. Synthesis of La and Ce modified X zeolite from rice husk ash for carbon dioxide capture. J. Mater. Res. Technol. 2020, 9, 4368–4378. [Google Scholar] [CrossRef]
  184. Flores, C.G.; Schneider, H.; Dornelles, J.S.; Gomes, L.B.; Marcilio, N.R.; Melo, P.J. Synthesis of potassium zeolite from rice husk ash as a silicon source. Clean. Eng. Technol. 2021, 4, 100201. [Google Scholar] [CrossRef]
  185. Tran-Nguyen, P.L.; Ly, K.-P.; Thanh, L.H.V.; Angkawijaya, A.E.; Santoso, S.P.; Tsai, M.-L.; Ju, Y.-H. Facile synthesis of zeolite NaX using rice husk ash without pretreatment. J. Taiwan Inst. Chem. Eng. 2021, 123, 338–345. [Google Scholar] [CrossRef]
  186. Masoumi, F.; Safari, S.; Khoshbin, R.; Karimzadeh, R. Utilization of agricultural waste (rice husk) in synthesis of TS-1 zeolite as a support for NiMo nanocatalyst employed in hydrodesulfurization of heavy oil. Adv. Powder Technol. 2023, 34, 104134. [Google Scholar] [CrossRef]
  187. Lee, J.B.; Ahmed, I.; Lee, G.; Kim, T.-W.; Kim, C.-U.; Jhung, S.H. Synthesis of SSZ-13 zeolites using calcined rice husk as silica source for propylene production from ethylene and carbon dioxide adsorption. J. Ind. Eng. Chem. 2023. [Google Scholar] [CrossRef]
  188. Liu, H.; Yue, Y.; Shen, T.; Wang, W.; Li, T.; Bao, X. Transformation and crystallization behaviors of titanium species in synthesizing Ti-ZSM-5 zeolites from natural rectorite mineral. Ind. Eng. Chem. Res. 2019, 58, 11861–11870. [Google Scholar] [CrossRef]
  189. Yue, Y.; Kang, Y.; Bai, Y.; Gu, L.; Liu, H.; Bao, J.; Wang, T.; Yuan, P.; Zhu, H.; Bai, Z. Seed-assisted, template-free synthesis of ZSM-5 zeolite from natural aluminosilicate minerals. Appl. Clay Sci. 2018, 158, 177–185. [Google Scholar] [CrossRef]
  190. Servatan, M.; Ghadiri, M.; Yazdi, M.K.; Jouyandeh, M.; Mahmodi, G.; Samadi, A.; Zarrintaj, P.; Habibzadeh, S.; Ganjali, M.R.; Saeb, M.R. Synthesis of cost-effective hierarchical MFI-type mesoporous zeolite: Introducing diatomite as silica source. Silicon 2021, 13, 3461–3472. [Google Scholar] [CrossRef]
  191. Asl, S.M.H.; Masomi, M.; Tajbakhsh, M. Hybrid adaptive neuro-fuzzy inference systems for forecasting benzene, toluene & m-xylene removal from aqueous solutions by HZSM-5 nano-zeolite synthesized from coal fly ash. J. Clean. Prod. 2020, 258, 120688. [Google Scholar] [CrossRef]
  192. Rajakrishnamoorthy, P.; Karthikeyan, D.; Saravanan, C. Emission reduction technique applied in SI engines exhaust by using zsm5 zeolite as catalysts synthesized from coal fly ash. Mater. Today Proc. 2020, 22, 499–506. [Google Scholar] [CrossRef]
  193. Song, C.-M.; Feng, Y.; Ma, L.-L. Characterization and hydroisomerization performance of SAPO-11 molecular sieves synthesized by dry gel conversion. Microporous Mesoporous Mater. 2012, 147, 205–211. [Google Scholar] [CrossRef]
  194. Cai, R.; Sun, M.; Chen, Z.; Munoz, R.; O’Neill, C.; Beving, D.E.; Yan, Y. Ionothermal synthesis of oriented zeolite AEL films and their application as corrosion-resistant coatings. Angew. Chem. Int. Ed. 2008, 47, 525–528. [Google Scholar] [CrossRef] [PubMed]
  195. Jin, Y.; Chen, X.; Sun, Q.; Sheng, N.; Liu, Y.; Bian, C.; Chen, F.; Meng, X.; Xiao, F.S. Solvent-Free Syntheses of Hierarchically Porous Aluminophosphate-Based Zeolites with AEL and AFI Structures. Chem. A Eur. J. 2014, 20, 17616–17623. [Google Scholar] [CrossRef] [PubMed]
  196. Ren, L.; Wu, Q.; Yang, C.; Zhu, L.; Li, C.; Zhang, P.; Zhang, H.; Meng, X.; Xiao, F.-S. Solvent-free synthesis of zeolites from solid raw materials. J. Am. Chem. Soc. 2012, 134, 15173–15176. [Google Scholar] [CrossRef]
  197. Zhang, H.; Wang, B.; Yan, W. The structure-directing role of heterologous seeds in the synthesis of zeolite. Green Energy Environ. 2023. [Google Scholar] [CrossRef]
  198. Ren, B.; Wang, J.; Zhou, Z.; Du, P.; Zhang, X. Regulation of the composition of metakaolin-based geopolymer: Effect of zeolite crystal seeds. Case Stud. Constr. Mater. 2023, 19, e02421. [Google Scholar] [CrossRef]
  199. Kadja, G.T.; Azhari, N.J.; Mukti, R.R.; Khalil, M. A mechanistic investigation of sustainable solvent-free, seed-directed synthesis of ZSM-5 zeolites in the absence of an organic structure-directing agent. ACS Omega 2020, 6, 925–933. [Google Scholar] [CrossRef] [PubMed]
  200. Shestakova, D.O.; Babina, K.A.; Sladkovskiy, D.A.; Parkhomchuk, E.V. Seed-assisted synthesis of hierarchical zeolite ZSM-5 in the absence of organic templates. Mater. Chem. Phys. 2022, 288, 126432. [Google Scholar] [CrossRef]
  201. Bu, L.; Wang, Y.; Liu, W.; Chu, K.; Guo, N.; Huang, Y.; Qu, L.; Su, X.; Zhang, X.; Li, Y. Green synthesis of shaped nano-ZSM-5 toward more efficient methanol-to-hydrocarbon conversion. Appl. Catal. A Gen. 2023, 665, 119393. [Google Scholar] [CrossRef]
  202. Nguyen, D.-K.; Dinh, V.-P.; Dang, N.; Khan, D.T.; Hung, N.T.; Tran, N.H.T. Effects of aging and hydrothermal treatment on the crystallization of ZSM-5 zeolite synthesis from bentonite. RSC Adv. 2023, 13, 20565–20574. [Google Scholar] [CrossRef]
  203. Han, S.; Liu, Y.; Yin, C.; Jiang, N. Fast synthesis of submicron ZSM-5 zeolite from leached illite clay using a seed-assisted method. Microporous Mesoporous Mater. 2019, 275, 223–228. [Google Scholar] [CrossRef]
  204. Jain, R.; Rimer, J.D. Seed-Assisted zeolite synthesis: The impact of seeding conditions and interzeolite transformations on crystal structure and morphology. Microporous Mesoporous Mater. 2020, 300, 110174. [Google Scholar] [CrossRef]
  205. Chawla, A.; Jain, R.; Linares, N.; Martínez, J.G.; Rimer, J.D. Synthesis of Hierarchical Zeolite ZSM-11 Catalysts Via a Novel Organic-Free Route. In Proceedings of the 2019 North American Catalysis Society Meeting, Chicago, IL, USA, 23–28 June 2019. [Google Scholar]
  206. Zhu, D.; Wang, L.; Cui, W.; Tan, J.; Tian, P.; Liu, Z. High-silica zeolite Y: Seed-assisted synthesis, characterization and catalytic properties. Inorg. Chem. Front. 2022, 9, 2213–2220. [Google Scholar] [CrossRef]
  207. Velaga, B.; Peela, N.R. Seed-assisted and OSDA-free synthesis of H-mordenite zeolites for efficient production of 5-hydroxymethylfurfural from glucose. Microporous Mesoporous Mater. 2019, 279, 211–219. [Google Scholar] [CrossRef]
  208. Deng, X.; Zhou, P.; Yan, X.; Xiong, R.; Kou, H.; Luo, W. Green synthesis of low-silica CHA zeolite without organic structural directing agents, fluoride media and seeds. Microporous Mesoporous Mater. 2021, 310, 110618. [Google Scholar] [CrossRef]
  209. Meng, J.; Li, C.; Chen, X.; Song, C.; Liang, C. Seed-assisted synthesis of ZSM-48 zeolite with low SiO2/Al2O3 ratio for n-hexadecane hydroisomerization. Microporous Mesoporous Mater. 2020, 309, 110565. [Google Scholar] [CrossRef]
  210. Kamimura, Y.; Iyoki, K.; Elangovan, S.P.; Itabashi, K.; Shimojima, A.; Okubo, T. OSDA-free synthesis of MTW-type zeolite from sodium aluminosilicate gels with zeolite beta seeds. Microporous Mesoporous Mater. 2012, 163, 282–290. [Google Scholar] [CrossRef]
  211. Kamimura, Y.; Itabashi, K.; Okubo, T. Seed-assisted, OSDA-free synthesis of MTW-type zeolite and “Green MTW” from sodium aluminosilicate gel systems. Microporous Mesoporous Mater. 2012, 147, 149–156. [Google Scholar] [CrossRef]
  212. Muraza, O.; Abdul-lateef, A.; Tago, T.; Nandiyanto, A.B.; Konno, H.; Nakasaka, Y.; Yamani, Z.H.; Masuda, T. Microwave-assisted hydrothermal synthesis of submicron ZSM-22 zeolites and their applications in light olefin production. Microporous Mesoporous Mater. 2015, 206, 136–143. [Google Scholar] [CrossRef]
  213. Sadeghpour, P.; Haghighi, M.; Ebrahimi, A. Ultrasound-assisted rapid hydrothermal design of efficient nanostructured MFI-Type aluminosilicate catalyst for methanol to propylene reaction. Ultrason. Sonochem. 2021, 72, 105416. [Google Scholar] [CrossRef] [PubMed]
  214. Ramirez-Mendoza, H.; Valdez Lancinha Pereira, M.; Van Gerven, T.; Lutz, C.; Julian, I. Ultrasound-assisted preparation of Mo/ZSM-5 zeolite catalyst for non-oxidative methane dehydroaromatization. Catalysts 2021, 11, 313. [Google Scholar] [CrossRef]
  215. Wang, B.; Wu, J.; Yuan, Z.-Y.; Li, N.; Xiang, S. Synthesis of MCM-22 zeolite by an ultrasonic-assisted aging procedure. Ultrason. Sonochem. 2008, 15, 334–338. [Google Scholar] [CrossRef]
  216. Kumar, N.; Masloboischikova, O.; Kustov, L.; Heikkilä, T.; Salmi, T.; Murzin, D.Y. Synthesis of Pt modified ZSM-5 and beta zeolite catalysts: Influence of ultrasonic irradiation and preparation methods on physico-chemical and catalytic properties in pentane isomerization. Ultrason. Sonochem. 2007, 14, 122–130. [Google Scholar] [CrossRef]
  217. Palanisamy, S.; Palanisamy, D.; Gul, M.; Kandasamy, K.; Gevert, B.S. Hydro-treating and Hydro-isomerisation of Sunflower Oil using Pt/SAPO-11: Influence of Templates in Ultrasonic Assisted with Hydrothermal Synthesis. Bull. Chem. React. Eng. Catal. 2021, 16, 120–135. [Google Scholar] [CrossRef]
Figure 1. Distribution of published papers on isomerization of n-C16 paraffins by countries [sourced from Scopus database].
Figure 1. Distribution of published papers on isomerization of n-C16 paraffins by countries [sourced from Scopus database].
Catalysts 13 01363 g001
Figure 2. Routes of carbenium ions formation. (a) hydride elimination at Lewis type acid sites, (b) hydride elimination on adsorbed carbenium ion, (c) alkane protonation, and (d) olefin protonation.
Figure 2. Routes of carbenium ions formation. (a) hydride elimination at Lewis type acid sites, (b) hydride elimination on adsorbed carbenium ion, (c) alkane protonation, and (d) olefin protonation.
Catalysts 13 01363 g002
Figure 3. Mechanism of hydroisomerization of n-hexadecane.
Figure 3. Mechanism of hydroisomerization of n-hexadecane.
Catalysts 13 01363 g003
Figure 4. The study of n-alkanes isomerization on zeolite-based catalysts over the last decade [ScienceDirect database].
Figure 4. The study of n-alkanes isomerization on zeolite-based catalysts over the last decade [ScienceDirect database].
Catalysts 13 01363 g004
Figure 5. Representation of basic shape selectivity types in the zeolite: (a) reactant shape selectivity; (b) transitional-state shape selectivity; (c) product shape selectivity.
Figure 5. Representation of basic shape selectivity types in the zeolite: (a) reactant shape selectivity; (b) transitional-state shape selectivity; (c) product shape selectivity.
Catalysts 13 01363 g005
Figure 6. Representation of secondary shape selectivity types in the zeolite: (a) inverse shape selectivity; (b) molecular traffic shape selectivity; (c) pore mouth and key-lock selectivity; (d) the window effect.
Figure 6. Representation of secondary shape selectivity types in the zeolite: (a) inverse shape selectivity; (b) molecular traffic shape selectivity; (c) pore mouth and key-lock selectivity; (d) the window effect.
Catalysts 13 01363 g006
Figure 7. Representation of the nest effect in zeolite.
Figure 7. Representation of the nest effect in zeolite.
Catalysts 13 01363 g007
Figure 8. Product distribution of n-C16 hydroisomerization over 0.5% Pt/ZSM-5, -23, -35, 48 and 0.5% Pd/ZSM-12, -22 [58,79,80].
Figure 8. Product distribution of n-C16 hydroisomerization over 0.5% Pt/ZSM-5, -23, -35, 48 and 0.5% Pd/ZSM-12, -22 [58,79,80].
Catalysts 13 01363 g008
Figure 9. Developing the acid sites in the zeolite framework by cation substitutions.
Figure 9. Developing the acid sites in the zeolite framework by cation substitutions.
Catalysts 13 01363 g009
Figure 10. Performance of zeolite-based catalysts (~1 wt% Pd) for hydroisomerization of n-C16 at the operating condtions: temperature 220–340 °C, WHSV = 10 h−1, H2/n-C16 = 20, P = 6 MPa. Reprinted with permission from [57]. Copyright 2020 Elsevier Inc.
Figure 10. Performance of zeolite-based catalysts (~1 wt% Pd) for hydroisomerization of n-C16 at the operating condtions: temperature 220–340 °C, WHSV = 10 h−1, H2/n-C16 = 20, P = 6 MPa. Reprinted with permission from [57]. Copyright 2020 Elsevier Inc.
Catalysts 13 01363 g010
Figure 11. The stability of the conversion of bifunctional catalysts on processing feedstocks containing sulfur (a) and nitrogen (b). Adapted with permission from [136]. Copyright 2020 MDPI.
Figure 11. The stability of the conversion of bifunctional catalysts on processing feedstocks containing sulfur (a) and nitrogen (b). Adapted with permission from [136]. Copyright 2020 MDPI.
Catalysts 13 01363 g011
Figure 12. Distributions of n-hexadecane hydrocracking products (cracking yield 55%) with water over (a) 0.9% Pt/MFI and (b) 0.9% Pd/MFI and (c) (cracking yield 44%) over 0.9% Pt/H-Y. Reprinted with permission from [141]. Copyright 2016 American Chemical Society.
Figure 12. Distributions of n-hexadecane hydrocracking products (cracking yield 55%) with water over (a) 0.9% Pt/MFI and (b) 0.9% Pd/MFI and (c) (cracking yield 44%) over 0.9% Pt/H-Y. Reprinted with permission from [141]. Copyright 2016 American Chemical Society.
Catalysts 13 01363 g012
Figure 13. General illustration of the routes of the coke formation over bifunctional catalysts.
Figure 13. General illustration of the routes of the coke formation over bifunctional catalysts.
Catalysts 13 01363 g013
Table 2. Characteristics of hexadecane isomers as a fuel component [PubChem database]; [16].
Table 2. Characteristics of hexadecane isomers as a fuel component [PubChem database]; [16].
IsomerChemical StructureBoiling Point a, °CFlash Point a, °CMelting Point a, °C Cetane Number
Mono-isomers
2-methylpentadecane
Catalysts 13 01363 i001282.095.3−9.2100
8-methylpentadecaneCatalysts 13 01363 i002274.394.0−25.675
5-butyldodecaneCatalysts 13 01363 i003150.7 bN/A−47.445
Di-isomersCatalysts 13 01363 i004248.0N/A−37.0
(3R,4R)-dimethyltetradecane61
(3R,4S)-dimethyltetradecane64
(3S,4R)-dimethyltetradecane58
(3S,4S)-dimethyltetradecane62
(5R,6R)-dimethyltetradecaneCatalysts 13 01363 i005N/AN/AN/A51
(5R,6S)-dimethyltetradecane51
(5S,6R)-dimethyltetradecane48
(5S,6S)-dimethyltetradecane52
7,8-dimethyltetradecaneCatalysts 13 01363 i006270.0−86.2−86.240
Multi-branched isomers
2,2,4,4,6,8,8-heptamethylnonane
Catalysts 13 01363 i007247.0102.0N/A15
a Measured at 750 mmHg. b Measured at 20 mmHg.
Table 3. Comparison of thermal and catalytic cracking processes.
Table 3. Comparison of thermal and catalytic cracking processes.
HydrocarbonsThermal Cracking ProductsCatalytic Cracking Products
n-HexadecanePrimary products—alkanes C1-C3; significant quantity of n-olefins C4-C15; relatively low concentrations of branched aliphaticsPrimary products—alkanes C3-C6; minor amounts of n-olefins; high concentrations of branched aliphatics
AliphaticsMinor conversion to aromatics at 500 °CLarge conversion to aromatics at 500 °C
n-olefinsSlow sigmatropic rearrangement of double bond with a low isomerization Rapid sigmatropic rearrangement of double bond with a significant isomerization conversion
Branched olefinsNon-selective hydrogen transfer and lower rate of cracking as compared to that of the corresponding alkanesSelective hydrogen transfer and higher rate of cracking than that of the corresponding alkanes
NaphthenesThe rate of cracking is lower than that of the corresponding alkanes.The rate of cracking is comparable to that of the corresponding alkanes.
Alkyl aromaticsCracking happens within the side of alkyl functional groupsCracking happens within the side of aromatic ring itself
Alkyl aromatics with larger functional groupsThe rate of cracking is lower than that of the corresponding alkanesThe rate of cracking is higher than that of the corresponding alkanes
Table 4. Catalysts generations for hydroisomerization of n-alkanes.
Table 4. Catalysts generations for hydroisomerization of n-alkanes.
GenerationCatalysts
Composition
Operating Temperature, °CAdvantagesDisadvantages
1stFriedel-Crafts, AlCl3 80–100High activitySensitivity to trace water; fast deactivation
2ndMetal/support 350–500Easy to apply;
corrosion issues are mitigated; sensitivity problems are less severe
Thermodynamic limitations on conversion
3rdMetal/halogenated support120–160Enhanced acidity due to halogenation (Cl, F) of alumina supportSensitive to impurities;
continuous feeding chlorine to maintain catalyst activity; corrosion of equipment
4thBifunctional zeolite-based catalysts250–340Easy to use;
tolerant to poisons;
no feedstock pretreatment
Selectivity issues;
more expensive to produce
Table 5. The standard structural properties of zeolites [Database of Zeolite Structure].
Table 5. The standard structural properties of zeolites [Database of Zeolite Structure].
ZeoliteTopologyMember Ring (MR) Channel DimensionalityAccessible Volume, %
ZSM-5MFI10, 6, 5, 43D9.81
ZSM-12MTW12, 6, 5, 41D9.42
ZSM-22TON10, 6, 51D8.04
ZSM-23MTT10, 6, 51D7.98
ZSM-35FER10, 8, 6, 52D10.01
ZSM-48MRE10, 6, 5, 41D6.55
SAPO-11AEL10, 6, 41D6.77
BetaBEA12, 6, 5, 43D20.52
MordeniteMOR12, 8, 5, 42D12.27
EMM-23EWT21, 10, 6, 5, 43D21.95
Table 6. The morphological properties of zeolite-based catalysts of hydroisomerization prepared with using various organic structure-directing agents (OSDAs).
Table 6. The morphological properties of zeolite-based catalysts of hydroisomerization prepared with using various organic structure-directing agents (OSDAs).
Zeolite Support TypeOSDAsSi/AlSpecific Area a, m2·g−1Total Pore Volume b, cm3·g−1Total Acidity c, µmol/gRef.
ZSM-5Tetrapropylammonium hydroxide 323710.232238[58]
ZSM-12Tetraethylammonium bromide (TEABr)45
60
90
120
310
270
220
260
N/A195
92
80
54
[59]
ZSM-12Methyltriethylammonium chloride (MTEACl)25280N/A720[60]
40293530
50287430
75314300
100323270
ZSM-12453010.160601[61]
ZSM-12Tetraethylammonium hydroxide
(TEAOH)
40323N/AN/A[62]
50331
71300
100291
ZSM-12Benzyltrimethylammonium chlorid (BTMACl)
TEABr
Methyltriethylammonium bromide (MTEABr)
~90260
220
260
N/A107
80
115
[63]
ZSM-12TEAOH
MTEACl
61
94
258
380
0.120
0.230
N/A[64]
ZSM-12P-Phenylenedimethylene-bis(N-methylpyrrolidinium) dibromide
P-Phenylenedimethylene-bis(N-methylpiperidinium) dibromide
20
50
297
306
0.370
0.340
898
550
[65]
ZSM-221,6-diaminohexane
(+Organosilane)
(+Silanized silica beads)
76
72
72
258
257
251
0.210
0.190
0.230
N/A[66]
ZSM-221,6-diaminohexane (with acid treatment)462290.320N/A[67]
ZSM-221,6-diaminohexane40
60
100
240
255
244
N/A192
145
95
[68]
ZSM-226-amino-1-hexanol35
50
100
150
110
137
134
144
0.210
0.190
0.140
0.120
150.2
121.7
83.2
59.1
[69]
ZSM-23N,N-dimethylformamide40150N/AN/A[70,71]
ZSM-23Pyrrolidine (PY)100
150
305
228
0.280
0.230
68 *
52 *
[72]
ZSM-23Isopropylamine (IPA)2002060.26049 *
ZSM-23Dual-OSDA (PY + IPA)100
150
200
278
282
251
0.190
0.200
0.200
22 *
19 *
17 *
EU-1Hexamethonium bromide
253250.160847[61,73]
IM-51,5-bis(methylpyrrolidinium)-
pentane bromide
333380.257289[58]
IM-5
(micro-
pores)
213810.412252
SAPO-11Dipropylamine phosphate (DPA.H3PO4)0.242090.134245.3[74]
SAPO-31Di-n-butylamine0.62220.14457.9[75]
BETA(TEAOH)20
30
40
50
648
614
591
553
0.341
0.314
0.277
0.271
850
830
580
490
[76,77]
a Calculated by BET method. b Obtained using t-plot method at p/p0 = 0.99. c Obtained using FTIR data of pyridine absorption considering bands at 1545, 1490, and 1445 cm−1. * Obtained by NH3-TPD considering peaks centered at 441–445 °C, 313–348 °C, and 220–230 °C.
Table 7. The performance of zeolite-based catalysts of hydroisomerization prepared with using various metal precursors.
Table 7. The performance of zeolite-based catalysts of hydroisomerization prepared with using various metal precursors.
Bifunctional CatalystMetal
Loading, %
PrecursorSi/AlConversion, %Selectivity to i-C16, %Ref.
Pt/ZSM-120.50[Pt(NH3)4](OH)245
60
90
120
88
88
86
85
86
87
87
92
[59]
Pt/ZSM-120.50908287[63]
Pt/ZSM-220.50Pt(NO3)2
H2PtCl6
Pt(NH3)4Cl2
10033
32
30
52
72
45
[88]
Pt/ZSM-22
(Pt nanocrystals)
Octahedral
Spherical
Cubes
0.50H2PtCl6100
30
35
32

71
56
53
[92]
Pt/ZSM-220.46[Pt(NH3)4](NO3)2N/A6085[93]
Pt/BETA0.702676
Pt/SAPO-110.909075
Pt/SBA-15 + BETA0.50H2PtCl6
[Pt(NH3)4](NO3)2
Pt(NH3)4Cl2
SBA(7)
BETA(75)
80
80
80
70
85
60
[94]
Pt/Fe/ZSM-230.98:0.98
1.00:1.90
0.99:3.66
1.03:7.00
H2PtCl6
Fe(NO3)·9H2O
5057
70
64
62
80
82
65
64
[95]
Pt/ZSM-22/ZSM-23
(40:60)
0.50H2PtCl61005878[96]
Ni/ZSM-48N/ANi(NO3)2·6H2O2006050[97,98]
Pt/SAPO-110.50H2PtCl60.249658[74]
Pd-Ni2P/SAPO-310.05 Pd
4.00 Ni2P
Pd(NO3)2
(NH4)2HPO4 + Ni(NO3)2
0.68372[75]
Pt/(SAPO-11/Al2O3)
SAPO-11:Al2O3 =
0.43
0.67
1.00
1.00H2PtCl6
0.015
0.020
0.025

50
56
57

99
99
99
[99]
Table 8. The performance of zeolite-based catalysts of hydroisomerization at optimal operating conditions.
Table 8. The performance of zeolite-based catalysts of hydroisomerization at optimal operating conditions.
Bifunctional CatalystActive Metal Loading, %Si/AlOptimal ConditionsConversion, %Selectivity, %Ref.
T, °CP, MPaWHSV, h−1H2/n-C16
Pd/ZSM-51.00402806.010.020954[57]
Pd/ZSM-5 a1.00612905816
Pt/ZSM-50.50322402.06.010334[58]
Pt/ZSM-120.50903106.01.1108687[63]
Pd/ZSM-121.00623106.010.0208592[57]
Pt/ZSM-220.46N/A3204.0N/AN/A8085[93]
Pd/ZSM-221.00613306.010.0208556[57]
Pt/ZSM-220.45453056.00.7109476[120]
Pt/Ca-ZSM-220.423129182
Pt/Ba-ZSM-220.443121.19086
Pt/ZSM-220.901003004.0N/A6001533[81]
Pt/ZSM-230.912907080
Pt/ZSM-230.60602703.01.069182[117]
Pt/ZSM-230.50582903.02.010008585[121]
Pd/EMM-231.00602906.010208013[57]
Pt/ZSM-350.95403004.0N/A6009248[80]
Pt/ZSM-480.932002054
Pt/EU-1/ZSM-480.972003004.0N/A6008055[122]
Pt/SAPO-11 b0.150.63102.03.16509496[123]
Pt/SAPO-110.90N/A3504.0N/AN/A9075[93]
Pd/SAPO-310.150.33402.03.75009090[124]
Pd/SAPO-410.8089096[125]
Pt-Pd/SAPO-410.15:0.359694[126]
Pd/SAPO-410.100.33602.03.75008795[127]
Pd-Ni2P/SAPO-310.05/4.000.33802.03.75008587[75]
Pt/IM-50.50332402.06.010435[58]
Pt/BETA0.40142000.53.0135063[128]
Pt/BETA0.79-2804.0N/AN/A5070[93]
Pt-Pd/BETA0.40/0.40252000.13.07507893[129]
CoMo/Al2O3-Y3.50–13.002.53601.51.01501155[130]
Pt/BETA1.00142000.53.0135068[117]
Pt/USY0.701721561
Pt/USY+ Pt/BETA0.70/1.0014/1720565
Pt/Al-SBA-15 + BETA0.50753005.03.558471[131]
Pt/Al-SBA-15 + MOR454381
a Nanostructure. b Vacuum-assisted synthesis.
Table 9. Comparison of traditional and modern approaches of zeolite-based catalyst production.
Table 9. Comparison of traditional and modern approaches of zeolite-based catalyst production.
ApproachPrincipleInfluencing FactorsAdvantagesDisadvantages
Solvothermal methodUsing solventsTemperature, pressure, reactants sources and compositions, Si/Al, aging time, alkalinity, stirring conditions, seeding, solvents typeSimple; cheaperEnvironmentally harmful; energetic and time-consuming approach
Alkali-fusion methodFussing raw materials with alkali before hydrothermal treatmentSi/Al, concentration of the alkali medium, temperature of crystallizationUsing low-grade raw materials without purification; offers the zeolite with high purityHigh energy consumption; expensive; multistep;
time-consuming
Sol-gel methodFormation of an inorganic sol and its gelation in liquid phase gel to form a 3D structureHydrolysis rate, temperature, heating rate, PHHigh-quality product; homogeneity; does not require expensive equipmentCost of precursors
Microwave methodUsing microwave irradiationSi/Al ratios, alkalinity, wavelength, temperature, crystallization timeConcise time;
high purity product;
small particle size
Low crystal aspect ratio
Ultra-sound methodUsing ultrasound with frequency 20 kHz–2 MHzDuration and frequency of ultrasonicationVery simple; rapid reaction; high crystals growth rate; suitable particle size distribution and morphology;
control on nucleation process
The mechanism of ultrasonic effect remains unclear
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Aljajan, Y.; Stytsenko, V.; Rubtsova, M.; Glotov, A. Hydroisomerization Catalysts for High-Quality Diesel Fuel Production. Catalysts 2023, 13, 1363. https://doi.org/10.3390/catal13101363

AMA Style

Aljajan Y, Stytsenko V, Rubtsova M, Glotov A. Hydroisomerization Catalysts for High-Quality Diesel Fuel Production. Catalysts. 2023; 13(10):1363. https://doi.org/10.3390/catal13101363

Chicago/Turabian Style

Aljajan, Yamen, Valentin Stytsenko, Maria Rubtsova, and Aleksandr Glotov. 2023. "Hydroisomerization Catalysts for High-Quality Diesel Fuel Production" Catalysts 13, no. 10: 1363. https://doi.org/10.3390/catal13101363

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