Next Article in Journal
Metal-Doped HZSM-5 Zeolite Catalysts for Catalytic Cracking of Raw Bio-Oil: Exploring Activity toward Value-Added Products
Next Article in Special Issue
Mechanistic Insight into the Propane Oxidation Dehydrogenation by N2O over Cu-BEA Zeolite with Diverse Active Site Structures
Previous Article in Journal
Design of Highly Efficient Nickel-Cobalt-Manganese-Molybdenum (NCMM) Nano-Catalysts Supported on Activated Carbon for Desulfurization Process
Previous Article in Special Issue
Silver-Doped Zeolitic Imidazolate Framework (Ag@ZIF-8): An Efficient Electrocatalyst for CO2 Conversion to Syngas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research and Application Development of Catalytic Redox Technology for Zeolite-Based Catalysts

1
Department of Aircraft Maintenance Engineering, Guangzhou Civil Aviation College, Guangzhou 510403, China
2
State Key Laboratory of Silicate Materials for Architectures, Wuhan University of Technology, Wuhan 430070, China
3
Center for Materials Research and Analysis, Wuhan University of Technology, Wuhan 430070, China
4
College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 256600, China
5
Key Laboratory of Optoelectronic Chemical Materials and Devices, Ministry of Education, Jianghan University, Wuhan 430056, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(8), 1197; https://doi.org/10.3390/catal13081197
Submission received: 8 July 2023 / Revised: 7 August 2023 / Accepted: 7 August 2023 / Published: 10 August 2023

Abstract

:
Zeolites are porous materials with plentiful and adjustable pore structures, which are widely applied in various fields such as fossil fuel energy conversion, preparation of clean energy, chemical product conversion, CO2 capture, VOC treatment, and so on. Zeolites exhibited advantageous adsorption compared with traditional adsorbents such as activated carbon; in addition, they can also provide abundant reaction sites for various molecules. The chemical composition, structural acidity, and distribution of pore size can distinctly affect the efficiency of the reaction. The modification of zeolite structure, the development of novel and efficient preparation methods, as well as the improvement of reaction efficiency, have always been the focus of research for zeolites.

1. Introduction

Adsorption and catalytic oxidation technology have extensive uses in energy, chemical engineering, and environmental industries, such as the cracking of fossil fuels, conversion of biomass, synthesis of non-petroleum products, low-cost production of hydrogen, catalytic treatment of industrial exhaust gas, etc. [1,2]. The efficiency of catalytic reactions is strongly influenced by adsorbents and catalysts. Under normal conditions, adsorbents are porous materials with huge specific surface areas that can provide enough space for redox reactions. Activated carbons (ACs), porous alumina, metal-organic frameworks (MOFs), and molecular sieves are constantly used as adsorbents because of their pore structure. More specifically, considering the thermal stability and the modification of pore structure, molecular sieves exhibited better performance than activated carbons.
Zeolite has generally been utilized in various synthetic processes in industrial products since its development (as shown in Figure 1). For zeolites, the basic structural unit (TO4) is connected by bridging oxygen atoms to form a secondary building structure (SBU), which can be further combined to form composite building structures (CBUs), and multiple rings (n) are simultaneously formed [3]. Zeolites can be classified as different types according to the pore size and the number of rings: (1) small pore zeolites with n = 8, (2) medium pore zeolites with n = 10, (3) large pore zeolites with n = 12, (4) extra-large pore zeolites with n > 12.
Due to the different combinations of basic units, there are significant differences in the pore structure and size of different types of zeolite molecular sieves. So far, A type, X type, Y type, β-zeolite, and ZSM-5 zeolite have been the most widely used zeolites.
A type (Linde type A, LTA) zeolites have a cubic crystal system. It has 3-dimensional channels ranging from 3 Å to 5 Å and an 8-membered topological structure. β zeolite is composed of 12-membered rings with an average channel size of 6 Å − 8 Å. X type and Y type zeolites are classified as FAU zeolites. Both of them belong to cubic crystals and are composed of hexagonal column cages and β cages formed from 6-membered rings. The Si/Al ratio of the X type varies from 2.2 to 3, and the Y type is higher than 3. Due to the special channels, X type and Y type zeolites have a large adsorption capacity for organic pollutant molecules. However, the adsorption capacity of organic molecules is reduced in moist gases. ZSM-5 is a typical MFI zeolite with a “zigzag” shaped 10-membered ring channel and vertically oriented 10-membered ring channels. The medium pore size (~5 Å) and excellent hydrothermal stability facilitated its application in the petrochemical industry.
Additionally, catalysts are pretty active components in the reaction process, including three types of catalysts: noble metals, non-noble metal oxides, and composite metal oxides. In order to increase the efficiency of the catalytic oxidation process, the selection of adsorbents and catalysts and the regulation of their modification should be prioritized.
This article illustrates the main applications of zeolite catalysts in summary and analyzes new advances in zeolite catalytic redox chemistry. The effects of porous structure, acidity, and other parameters, including zeolite type, catalyst, and reaction temperature, on the activity, selectivity, stability, and deactivation of zeolites are also summarized. Finally, the challenges and strategies of zeolite catalytic technology are discussed further.

2. The Application of Zeolite Catalysts

2.1. Catalytic Cracking of Petroleum Resources

Fluid catalytic cracking (FCC) technology is an effective alkyl removal technology that is increasingly widely utilized in the biomass oil and petrochemical industry. FCC catalysts consist of various zeolites with multi-component active ingredients that can greatly improve the efficiency of catalytic cracking [4,5].
For crude oil, it is extremely important to ameliorate the selectivity of catalysts towards products and increase the yield of specific products. In addition, Y-type zeolite has gained widespread use as a molecular sieve. Introducing and increasing mesoporous surface area can effectively reduce heavy oil production. Liu et al. prepared ordered silica aluminate on an industrial scale using a pre-crystallization unit of Y zeolite precursor. However, when a single layer of zeolite was dispersed on the catalyst surface, the composite catalyst exhibited the best activity for heavy oil cracking [6]. Pre-cracking is the first process to be carried out in the base pores, followed by further cracking and selective generation of its different products. Although Y zeolite is commonly used in heavy oil catalytic cracking and hydrocracking, the mass transfer efficiency between active sites is lower than expected. Cui et al. obtained mesoporous materials with octahedra by synthesizing ultra-stable mesoporous high-silicon molecular sieves directly. Compared to industrial FCC catalysts, the conversion rate was increased by 7.64%, and the gasoline yield was increased by 16.37%, even after these catalysts were aged at 800 °C [7].
Due to the requirements of the sixth national standard for the olefin content of automotive fuels, it is of great necessity to reduce the olefin content in gasoline. Through hydrothermal treatment, the acidity of ZSM-5 loaded metal catalysts (La-Ni-Zn) was reduced, and the B/L ratio (B = Brønsted acid, L = Lewis acid) was increased, which finally increased the aromatic production by 5% and the isoalkane production by 16%, making it an industrial catalyst for the catalytic cracking of gasoline [8]. To enhance the steam stability of Y zeolite, Yu et al. prepared rare earth (RE)-exchanged Y zeolite. The substitution of Y3+ ions for counter-ions Na+ resulted in a noticeable shrinkage of the unit cell owing to the relatively small ionic radius and high charge density of Y3+ ions. Therefore, this kind of zeolite exhibited better performance in steam stability and was more favorable for the generation of liquefied petroleum gas and C5+ gasoline products in n-dodecane cracking [9]. Except for Y3+, Al-exchanged Y zeolite was also proved to be able to promote liquefied petroleum gas production as a result of its strong acidity [10].
Except for routine chemical products, Li et al. have conducted extensive research on the production of high value-added products through dealkylation of light fraction light circulating oil and found that mesoporous BEA zeolite catalysts with appropriate acidity and mass transfer ability can greatly raise the yield of value-added products [11]. It can be seen that, although FCC was regarded as an efficient technology of dealkylation, the modifying of micropores and B acidic sites of zeolite need more research to elevate mass transfer ability. Thus, the optimization of catalysts in FCC, such as product selectivity, thermal stability, and cycle life, should receive focused attention so as to increase production.

2.2. Conversion of Biomass

Biomass is a promising green and renewable organic carbon source, but its calorific value is sharply lower than that of fossil fuels due to its abundant aerobic molecules, making it difficult to utilize directly [12,13]. It is necessary to transform biomass into valuable bioproducts. Two critical processes are involved in this strategy. One is the conversion of biomass into platform molecules, and the other is the upgrading of platform molecules into valuable fuels. Usually, the biomass comprised 40–50% cellulose, 25–35% hemicellulose, 15–20% lignin, and others [14]. Levulinic acid (LeA) and 5-hydroxymethylfurfural (HMF) are two common biomass platform molecules due to the highly reactive functional groups such as carboxyls, aldehydes, hydroxymethyls, and furan rings. LeA can be obtained from cellulose via the C5 route or hemicellulose via the C6 route, and HMF is usually produced from cellulose via the C6 route.
(1)
LeA via C5 and C6 route
In the C6 route, cellulose was hydrolyzed to form C6 sugar, which was dehydrated to form HMF. LA was obtained after the hydrolysis of HMF. Due to the high separation cost of LA and formic acid, synthesizing LA from C6 sugar was mainly adopted in the laboratory. For the C5 route, intermediates such as furfuryl and furfuryl alcohol (FAL) were successively converted to LeA, which made the carbon utilization more efficient than that of the C6 route [15]. HY zeolite mixed with ionic liquid (ionic liquid: HY = 0.5) has more Lewis acid sites than Bronsted acid sites and exhibits a LeA yield of 62.2% [16]. ZSM-5 zeolite with mesopores was treated by tandem alkaline and acid washing, and ZSM-5-OH0.2-H zeolite was therefore obtained; the modification of pore structure and Al distribution promoted the hydrolytic efficiency of FAL to LeA, the LeA yield reached 64.5% after three cycles [17].
(2)
HMF via the C6 route
HMF from biomass can be utilized to synthesize various biochemicals. The conversion of HMF from biomass can be divided into three processes: glucose hydrolyzed from cellulose, the isomerization of glucose to fructose, and the dehydration of fructose [15]. Therefore, fructose was usually used as a model molecule in HMF preparation, and many zeolites such as ZSM, HY, H-USY, and MAPO were adopted. Scholars prepared Cu-Cr/ZSM-5 zeolite using the ion-exchange method; when the zeolite was 20 wt.% of glucose, 50.4% HMF was achieved from glucose at 140 °C for 4 h [18]. To efficiently convert carbohydrates into HMF, β zeolite doping with 0.4 wt.% Cr was prepared and exhibited a superior 72% HMF yield and 83% selectivity due to the moderate L/B (Lewis acid sites/Brønsted acid sites) [19]. Low L/B hindered the isomerization of glucose to fructose; high L/B excessively accelerated the dehydration of fructose and led to the degradation of fructose to humin.
As the most abundant part of biomass, cellulose can be hydrolyzed and further converted into various chemical substances, one of which is lactic acid. Lactic acid (LA) from biomass was also among the top 30 candidates for synthetic fossil product substitutes [20]. Many important industrial chemicals can be derived from LA molecules, such as polylactic acid (PLA) [21], acrylic acid (AA), and propanoic acid (PA) (as shown in Figure 2). While cellulose was hydrolyzed to produce lactic acid (LA), various products such as levulinic acid and 5-hydroxymethylfurfural (HMF) were also produced, which resulted in a lower LA yield (around 30%). Therefore, it is necessary to improve the LA selectivity. In recent years, catalysts with acidic sites, especially solid Lewis acid catalysts, have received extensive research [22,23]. It was found that yttrium-modified siliceous material β zeolite catalysts can effectively regulate the surface acidity of zeolite and inhibit the yield of dehydration products such as HMF and other derivatives. The results indicated that when Lewis acidity was increased, the yield of LA from cellulose reached 49.2% within 30 min, more efficient than previous research studies [24].
It was noteworthy that the composition of some organic compounds may have side effects on the catalytic performance of the catalyst. For example, phenolic substances derived from lignin not only contain O molecules but also have adverse impacts on catalysts. For this reason, the key to the catalytic cracking of bio-oil conversion is the optimization of efficient catalysts. Different catalysts have been adopted for different biomass and products. In addition, NiMo/ASA Al2O3, Co Mo, HY zeolite, and ASA Al2O3 were simulated and found to be suitable for producing high distillate diesel [25]. The theoretical calculation results also indicated that larger mesoporous volume and surface area accelerated the diffusion rate of biomass inside the zeolite. However, this effect was continually weakened with increasing reaction time [26]. Because of the damage of by-products to catalysts and zeolites, the mass transfer efficiency was consequently decreased. To restrict the destruction of zeolites and catalysts and decrease the coke formation, an SFCC catalyst was employed in the cracking of high acid-value waste edible oil to produce biofuels. Moreover, the catalytic efficiency was attributed to doping with rare earth metals. Nguyen-Phuc et al. found that propylene and liquefied petroleum gas production were obviously improved by increasing the content of rare earth elements in ZSM-5 zeolite. Specifically, the yields of diesel, gasoline, and liquefied petroleum gas reached 29 wt.%, 42 wt.%, and 18 wt.%, respectively [27]. Additionally, to promote the conversion of biomass to valuable energy, a novel catalytic system similar to the Dylison Cycle Riser (DCR) should be designed and developed, which can produce hydrocarbon fuel intermediates from biomass-derived pyrolysis steam, such as pine and oak [28].

2.3. Preparation of Propylene

Propylene has been widely applied in diverse chemical products that are badly needed. Developing alternative methods to produce propylene using economical raw materials has attracted considerable interest, such as the hydrogenation of methanol to propylene and the dehydrogenation of propane to propylene [29,30,31].

2.3.1. Methanol to Propylene (MTP)

Methanol to propylene (MTP) was considered a substitutable approach to propylene production because of the extensive sources of methanol available. Due to its selectivity for propylene and the high resistance of the zeolite, ZSM-5 was preferred in the process (Table 1).
Si/Al makes a great difference to the morphology and aluminum distribution of molecular sieves. By increasing the Si/Al ratio in HZSM-5 zeolite, the selectivity of zeolite for propylene can be improved [32]. In the ZSM-5 catalyst, when the channel intersection is rich in Al pairs, the selectivity for ethylene and aromatics is higher. While the channel intersection has multiple single Al sites and lower Al pairs, it exhibits higher propylene selectivity and lower aromatic hydrocarbon selectivity [33].
Ion doping can also scale up the selectivity of zeolites towards specific products. When boron was introduced into the ZSM-5 structure through impregnation and hydrothermal synthesis, the orthogonal crystal structure of the zeolite was characterized, which retained the preferred growth orientation and high crystallinity hierarchical structure and changed the surface acidity. The selectivity of propylene was increased to 67%, and the propylene/ethylene ratio reached 8 [34]. When Mn enters the framework structure of ZSM-5, weak Brønsted acid sites increase the Al in the channel of the framework. Consequently, the carbon deposition rate was lowered, and the selectivity for propylene was improved [35].
Preventing catalyst deactivation and coke deposition from clogging pores and improving the lifespan of catalysts are still facing great challenges. Usually, modification of structural properties and catalyst acidity can efficiently solve these problems. To reduce acidity and consequently inhibit the aromatic hydrocarbon cycle, fluoride-assisted low-temperature crystallization can be applied to prepare plate-like MFI molecular sieves with similar crystal morphology and controllable acidity, which made propylene selectivity reach 52%, and catalyst lifespan 252 h [36]. In the initial process of MTP, long chain alkenes not being cracked diffused out from MFI zeolite and were found in the products (as shown in Figure 3, blue frame). Simultaneously, the long-chain alkenes propagated from olefins or alkenes were converted to polymethyl-benzenes and dienylic carbenium ions. Subsequently, lower olefins were produced in the aromatic cycle (as shown in Figure 3, red frame). Moreover, lower-density plate-like MFI can effectively inhibit the circulation of aromatic hydrocarbon groups, promote the hydrocarbon cycle, and exhibit excellent stability [37].
Table 1. Effects of Si/Al and specific surface area on the selectivity of propylene in MTP.
Table 1. Effects of Si/Al and specific surface area on the selectivity of propylene in MTP.
CatalystsMethodSi/AlSBETSelectivity of
Propylene
P/E
HZSM-5two-stage crystallization130
130
124
139
36150.95
48.13
49.68
50.10
4.40
4.44
5.89
7.48
[32]
356
352
340
MFItemplate202428587.9[34]
B-MFI20540466.37.9
Cr-MFI20140859.82.7
[Al/Ga]MFIfluoride-assisted low-temperature
crystallization
70–73
65–275 *
415–42352%-[36]
411–441
ZSM-5quasi-solid-phase5035341.4%-[35]
* Si/Ga determined by ICP.

2.3.2. Propane to Propylene (PDH)

The dehydrogenation of propane (PDH) was a conventional way to produce propylene. Driven by the huge demand for propylene, Pt- or Cr- containing catalysts have drawn great attention and been applied to industrial manufacture. Although ZSM-5 is frequently used for PDH reactions due to its unique pore structure and good thermal stability, the high acidity of ZSM-5 can lead to a decrease in propylene selectivity. Through catalyst dealumination and additive regulation, the surface acidity of ZSM-5 can be reduced, and its propylene selectivity and catalytic stability in PDH can be improved. Some scholars have adopted the sequential impregnation method to introduce Ga and Mg to improve the Si/Al in HZSM-5, which can produce an effect on its stability and propylene selectivity (90.8%) [38].
Additionally, the separation of propylene and propane is another indispensable process in PDH, and it can be promoted by enhancing the propylene affinity of the adsorption site. What provided new ideas for the effective separation of propylene and propane is that scholars have prepared Ag exchanged Y zeolite (Ag-Y); this modified zeolite exhibited rapid adsorption kinetics and reversible propylene adsorption [39]. The results of molecular dynamics simulation research indicated that raising the temperature was beneficial for an increase in propane adsorption capacity, as the critical temperature of propane is higher than that of propylene. The smaller radius of non-skeletal cations in zeolite contributed to the higher efficiency of propylene adsorption [40]. After dealumination, vacant T-atom sites appeared in Si-BEA and silanol group forms. The introduced Co2+ can interact with silanol groups and form four coordinated structures in the zeolite (as shown in Figure 4), which consequently selectively activates the C–H bonds in propane, thereby improving the desorption efficiency after propylene generation and improving the long-term stability of the catalyst [41].

2.4. Selective Catalytic Reduction (SCR) of NOx

Nitrogen oxides (NOx) were one of the most destructive polluting gases that were mainly produced in the combustion of fossil fuels, such as car exhaust and coal combustion. In China, more than 90% of NOx comes from vehicle exhaust emissions. Selective catalytic technology (SCR) is one efficient method to solve the problem of nitrogen oxide emissions. The catalytic reduction reactions between NOx and NH3 were conducted on the surface of the catalyst, and NOx was therefore transformed into N2. Currently, the adopted catalysts are mainly divided into metal matrix catalysts and zeolite-based catalysts. V2O5-WO3/TiO2 was once supposed to be the most effective catalyst for NOx treatment. However, the biological toxicity, poor thermostability, and narrow working temperature range extremely restrained its application. In the 1980s, a ZSM-5 catalyst load with Cu2+ was discovered and exhibited excellent activity to NOx and anti-toxicity.
Adsorption is the key process in catalytic oxidation and reduction. Zeolite catalysts exhibit good adsorption performance on NOx owing to the large specific surface area, good thermal stability, adjustable acidity, excellent absorption, and ion exchange capabilities [42,43]. The overall efficiency of NOx reduction can be greatly improved by metal doping and modification of zeolite catalysts. Cu-doped and Fe-doped zeolite catalysts have received more attention, such as Cu/ZSM-5, Cu/BEA, Cu/SAPO-34, and Cu/SSZ-13, which have been reported for the NH3-SCR reaction of NOx. The Cu-based catalyst exhibits excellent anti-propylene poisoning ability in the NH3-SCR reaction. Researchers have found that moderately copper-doped zeolite, such as the Cu7.55-ZK-5 catalyst, can isolate a large amount of Cu2+, which is beneficial for the adsorption and activation of NOx [44].
Based on Cu/MOR, the addition of promoting the separation of Cu2+ can create more Brønsted acid sites and inhibit the high-temperature ammonia oxygen reactions. This catalyst exhibited a NOx conversion of about 88% and N2 selectivity above 99% at 350–560 °C [45].
Hydrothermal stability is another factor influencing the application of zeolite catalysts. Yttrium-doped Cu-based zeolite catalysts can significantly reduce the fracture of Si-O-Al, improve the dispersion of Cu2+ active components and promote the adsorption and conversion efficiency of NOx [46].
NOx can be adsorbed with five reactions on Cu-ZSM-5 zeolites (as shown in Figure 5). The adsorption can be enhanced by the coordination interaction between Cu and N. Cu can be precisely dopped into H-ZSM-5 using an improved method called initial wetting impregnation microwave drying (IM). Compared with H-ZSM-5 and Na-ZSM-5, the desorption energy was reduced and the NOx adsorption was significantly increased [47]. By designing novel pore structures, such as three dimensionally-ordered (3DOM) microporous zeolites, the contact area can be obviously extended. Some scholars have developed the steam seed-assisted colloid crystal template (SSAC) method to prepare 3DOM zeolites. After sonication, drying, and calcination, the PrxMn1−xOδ/3DOM ZSM-5 catalyst with different Pr/Mn ratios was obtained (as shown in Figure 6) and showed an NO conversion rate of 90% [48].
Except for these zeolite catalysts, scholars tried to combine metal-based catalysts (V2O5-WO3/TiO2) with zeolite catalysts (H-ZSM-5) to prepare the composite catalyst PM. The PM catalyst exhibited higher catalytic activity than the V2O5 WO3/TiO2 catalyst because the PM catalyst adsorbed more NO2 and N2O4 [49]. Given the complexity of the exhaust purification system in SCR, some researchers adopted an H2 selective catalytic reduction method using Pt/KFI molecular sieve catalysts from 150 °C to 250 °C, and the conversion rate of NOx reached 80% [50].

2.5. Hydrogen Preparation

Hydrogen energy has been viewed as a potential alternative energy in the future. However, producing hydrogen economically and environmentally is extremely challenging. Hydrogen can be prepared from natural gas (reforming of methane steam to produce hydrogen, SMR), coal gasification, electrolysis of water, and transformation of biomass. Various hydrogen production processes depend on catalysts. Due to the wide and renewable sources, the utilization of organic matter to produce hydrogen energy is gradually receiving attention. Relevant technical methods mainly include two methods. One is thermochemistry hydrogen production, which consists of high-temperature pyrolysis, liquefaction, and gasification, and the other is biomass hydrogen production which refers to the preparation of hydrogen from biomass after anaerobic fermentation, biological metabolism, and reforming. The catalysts used in hydrogen are mainly precious metals and transition metals, such as Rh, Ru, Pd, Pt, Cu, Co, Ni, etc. Molecular sieves are often used as carriers for these catalysts.
At present, ethanol steam reforming (ESR) is the most suitable industrial method for ethanol to produce hydrogen, but there are some vital problems with this technology: (1) high energy consumption; (2) expensive separation of CO2 and H2. Choosing appropriate catalyst carriers and catalysts can help improve catalytic activity and reduce costs. Furthermore, the Ni catalyst was relatively inexpensive and had good hydrogen generation activity. Zeolites have a large content of micropores, and the structure can be modified. However, carbon deposition during the catalytic process of Ni-loaded zeolite can easily lead to decreasing catalytic activity. A new type of quaternary ammonium cation cationic surfactant was introduced to synthesize the supported calcium-modified nickel-step classification β zeolite catalysts. The BET and pore volume of this modified zeolite were larger than that without modification. The introduction of Ca2+ changed the internal structure of the catalyst to some extent, which promoted the interaction between the carrier and Ni. Consequently, the yield of H2 was increased [51,52]. When Mg was introduced into Rh/β zeolite, Rh clusters from the atomic level to the sub-nanometer scale were formed in microporous channels. It was found that the C=O formed in the reaction was the key to increasing selectivity, which can be promoted by Mg addition [53]. The DFT calculation results indicated that the stability was obviously affected by the zeolite structure. Specifically, after dealumination, the binding energy of Rh to the vacancy defect was reduced by 0.2 eV. If the Mg was added to the zeolite structure, the binding energy was additionally reduced to −1.927 eV (as shown in Figure 7a). Due to the proximity of Rh and Mg in zeolite with Mg, the energy barrier of O-H was significantly reduced, which was beneficial to the dissociation of phenol to phenoxy (as shown in Figure 7b). ITQ-6 zeolite loaded with Ni and Co can also be used in ESR. Among them, the ITQ–6 zeolite loaded with Co exhibited higher hydrogen production on account of the smaller size of Co particles. Moreover, the carbon deposition effect of Co/ITQ-6 was weaker, and the deactivation was lower [54].
In order to reduce catalyst deactivation, some scholars have tried to design catalysts with small metal nanoparticles. An encapsulated ultra-small Ni catalyst with mesoporous and hollow structure was prepared using separated metal dispersions Si-1 zeolite with high Ni dispersion (Ni@Si-1), which facilitated local mass transfer and reduced carbon deposition [55]. The research on supercritical water dynamics indicated that the gasification of catalysts such as Ni can be promoted by improving hydrogen selectivity. The pore distribution of Ni particles can be divided into micropores (1–10 nm) and mesopores (20–60 nm). When the size of the mesopores was distributed reasonably in zeolite, the active sites increased, and the catalytic activity was higher [56].
In addition to ethanol, methanol can also be used as the raw material for hydrogen production, that is, methanol steam reforming (MSR). A catalyst suitable for the catalytic reaction of MSR was prepared using an attapulgite molecular sieve as a carrier and Cu-Zr active components. The addition of Zr significantly increased the distribution of active metal particles and reaction sites, as well as the H2 yield. At the same time, the sintering of active metals and coke were inhibited [57].
Photocatalysis is another important method for producing hydrogen gas, and zeolite loaded with TiO2 photocatalyst was supposed to be one of the potentially efficient catalysts [58,59]. To improve the transmission efficiency of photoelectrons, some scholars have designed Cu2O@TiO2@ZIF-8, where an internal electric field was formed due to the p-n junction, which promoted the transfer of electrons to the conduction band. The photoelectrons generated from TiO2 were transferred to the conduction band and further transferred through ZIF-8. The water can be therefore reduced by these photoelectrons (as shown in Figure 8) [52]. To provide more proton sources and active sites, Na+/K+/Ca2+ can be added into zeolite, which consequently promotes the hydrogen evolution rate [60].

2.6. VOCs Abatement

Volatile organic compounds (VOCs) refer to organic compounds with a boiling temperature in the range of 50–260 °C. They were usually emitted in the petrochemical industry and the rubber industry and had serious effects on human health and the natural environment. With the rapid development of industry and increasing consumption of fossil fuel energy in vehicles, the emission of VOCs has dramatically increased in recent years. According to their boiling points, molecular structures, and molecular polarities, VOCs can be classified into different types [61]. They were hard to eliminate using direct combustion due to the low concentration of VOCs. To reduce the pollution and damage of VOCs, multiple technologies have been developed and applied, such as adsorption, a photocatalytic degradation method, plasma degradation, catalytic combustion, etc. Whether adopting an adsorption method or catalytic oxidation method, it is no wonder that developing efficient adsorbent and porous materials are principal considerations. An excellent adsorbent should have massive adsorption sites, good hydrothermal stability, and weak resistance to molecular diffusion. Worldwide, research studies show that activated carbons [62], metal-organic frameworks, and zeolites are all potential adsorption materials that have exhibited excellent adsorption performance in many experiments. However, the combustion of activated carbons and the high cost of metal-organic frameworks restrained their application in VOC treatment.
The zeolite of the 255 framework type, approved by the International Zeolite Association (IZA), was considered a promising absorbent due to its abundant micropores, adjustive chemical stability, etc. The adsorption efficiency of different zeolite structures on different VOCs molecules and the competitive adsorption mechanism of various VOCs have been a puzzle to reveal. The absorption of VOCs can be influenced by the specific surface area, pore structure, and surface functional groups. Additionally, the kinetic diameter and polarity of the VOC molecule may also affect the absorption process.
The absorption efficiency was also proved to be related to humidity. Therefore, developing novel zeolites and improving the hydrophobicity of zeolites contributed to the adsorption efficiency of zeolites under humid conditions. The hydrophobicity of zeolites can be improved through ion doping and coating. A Mo-doped MEL zeolite was developed using a self-developed recrystallization method for the adsorption and reduction of non-methane hydrocarbons (NMHC) in cooking oil fumes. The affinity for VOC molecules is significantly enhanced due to its high atomic coordination level and lack of silanol, especially under humid conditions. This Mo MEL zeolite exhibited longer adsorption saturation time, larger adsorption capacity, and better hydrophobicity, which led to the more excellent adsorption efficiency of Mo MEL than that of Si MEL and Al MEL [63]. In addition to Mo, introducing Mn (Cu or Fe) to replace Al can also improve the hydrophobicity of zeolites [64]. The addition of Mn affected the crystallization behavior of zeolites, which in turn, changed the distribution of Mn in zeolites and the properties of zeolites. Y-type zeolite, with a high specific surface area and unique microporous structure, can have its hydrophobicity and diffusion efficiency improved by being coated with mesoporous SiO2 [65]. In general, after being coated with mesoporous hydrophobic materials, the diffusion activation energy of VOCs was decreased, and the diffusion coefficient was increased. The π-n electron pair interaction was weakened, while the dipole interaction was enhanced. The dispersion interaction plays a positive role in the adsorption of VOCs and gradually enhances with increasing relative humidity. However, the repeatability of core-shell composite materials needs to be verified. The evaporation-induced self-assembly (EISA) method was adopted to prepare NaY zeolite loaded with metal oxide nanoparticles. It strongly improved the adsorption capacity of VOCs. Li SHI et al. successfully prepared uniformly dispersed Y@MxOy nanoparticles, which significantly improved the adsorption effect of Y-zeolite on isopropanol and acetone in humid environments under RH = 50%. The metal oxides enhanced the adsorption capacity of VOCs, and greatly reduced the competitive adsorption between water molecules and VOCs molecules [66]. Multilevel mesoporous USY molecular sieves can be prepared using etching and surfactant template processes, during which the non-skeletal aluminum was effectively removed inside the zeolite, leading to a mesoporous size of approximately 4 nm [67]. Some scholars have also proposed a new strategy called citric acid sacrifice that can synthesize non adhesive monolithic zeolites. The mechanical strength and the water resistance were obviously improved. After the in situ dealuminization, the specific surface area and mesoporous volume were greatly improved, the adsorption capacities of toluene and acetone were also increased. When the relative humidity was 90%, the adsorption efficiency of acetone and toluene reached 90% of that attained in dry conditions [68].
Adjusting the pore structure of molecular sieves is also vital for improving the adsorption and reaction rates of VOCs. The hierarchical design and synthesis of zeolite structures are gradually receiving attention [69,70]. For example, the microporous size of ZSM-5 zeolite is generally less than 2 nm, and larger VOC molecules have a slower diffusion rate during the adsorption process, making it difficult to quickly enter the reaction active center and the generation rate of specific products is pretty low. When ZSM-5 is modified with organic functionalized silica and organic additives, the size and surface morphology of ZSM-5 nanocrystals can be greatly improved, forming a hierarchical mesoporous structure, and the adsorption capacity for toluene is greatly increased [71]. According to the hierarchical crystallization mechanism revealed in the corresponding research (as shown in Figure 9), organic silica fragments modified by organosilane PHAPTMS were combined and then gathered around MFI. After hydrolysis of -SiO3, PHAPTMS participated in the formation of nanocrystalline frameworks in a hydrothermal process. The inter-crystalline and intra-crystalline mesopores were finally formed after calcination.
The adsorption and catalytic efficiency of zeolite can be effectively improved by regulating the active sites. Effective regulation of active sites can reduce carbon deposition in pore structures and improve the durability of catalysts. Nano Pt is usually adopted as an active ingredient, and its high dispersibility facilitates the catalytic oxidation of aromatic hydrocarbons and alkanes. Synergistically, it can interact with surface acid sites that contribute to the removal of VOCs at low temperatures [72]. Generally, adjusting the Si/Al ratio in zeolites can help regulate the acidic sites. A controllable framework modulation strategy can be adopted to prepare high-silica zeolites. In this strategy, it was very important to control the matching degree of dealumination and Si-insertion. For Pt/β zeolite, Pt0 has a stronger activation ability for O2 and shows higher efficiency than Ptδ+. That makes it much easier to supply oxygen, which is more conducive to the ring opening reaction of the benzene ring, thus reducing the production of gaseous benzene on the high silicon catalyst. Furthermore, both high-silicon and low-silicon Pt/Beta catalysts show excellent durability and follow a similar toluene oxidation path to form intermediates, including alcohol oxide, carboxylate, and anhydride [73]. For Pt/ZSM-5, the main active oxidation site is Pt0. For the deep catalytic oxidation of oxygen/nitrogen-containing VOCs, Pt/ZSM-5 had abundant acidic sites on its surface. Different SiO2/Al2O3, Pt0 ratio, and Pt dispersion equilibrium in the catalyst may be changed, and the synergistic effect of acidic sites and oxidation sites were promoted, which led to higher reaction activity of Pt/ZSM-5 (25), with a T90% value of only 207 °C for acetonitrile and 175 °C for ethyl acetate. It is beneficial for low-temperature catalytic degradation of specific VOCs [74].
In industrial activities, multiple types of VOCs are often generated simultaneously. Ideally, a single adsorbent and catalyst can simultaneously handle multiple VOCs. However, due to the various molecular structure of VOCs, the different pore structures, and the complexity of the catalyst’s electronic structure, ensuring the treatment effect of VOCs is difficult. Therefore, it is particularly important to explore and understand the adsorption and oxidation mechanism of catalysts, which contributes to the development of novel catalysts and adsorbents aimed at VOC mixtures. The molecular polarity and volatility of VOCs play a crucial role in the adsorption process. For these VOC mixtures, the priority order of adsorption varies. For example, under static equilibrium conditions, the absorption order was determined as acetone > ethyl acetate > toluene. However, under dynamical conditions, molecular polarity and volatility played key roles. The absorption order was found to be ethyl acetate > toluene > acetone [75]. Complex interference effects were identified when this VOC mixture was adsorbed onto different zeolites. The adsorption effect of toluene on USY is significantly better than that on 13X. However, for acetone, the adsorption of 13X and ZSM-5 is better. ZSM-5 and β Zeolites can selectively adsorb ethyl acetate, as evidenced by DFT calculations. This competitive adsorption widely exists in the treatment of various VOCs, and the differential adsorption effect is one of the problems to be solved in the future [75]. To develop efficient catalysts, some scholars have attempted to explore the possibility of synergistic treatment of acetone and benzene by changing the amount of Mn added [76]. When Mn was introduced, the activity of the catalyst for acetone was significantly increased, while the activity for benzene was decreased. Because the introduction of Mn regulated the electronic structure of Pt, leading to the transfer of electrons to Pt, which enhanced the adsorption of acetone and gaseous oxygen and improved the performance of the acetone oxidation catalyst, and promoted the degradation of acetone. However, the high electron density of Pt inhibited the adsorption and degradation of benzene, reducing the production of formic acid products.
In addition, the treatment of VOCs in open spaces has gradually attracted the attention of researchers. For example, asphalt materials are often used in the process of road paving and continually release VOCs into the open air. Due to the limitations of open spaces, it is very difficult to collect and dispose of these VOCs. Steel slag and red mud were usually used as road materials, zeolites synthesized from these materials can reduce the volatility of asphalt and absorb VOCs in road structures [77]. Therefore, preparing efficient zeolites using in-road materials (such as steel slag) was the key approach to deal with VOC emissions in open spaces.

2.7. CO2 Capture

CO2 has been regarded as the main greenhouse gas resulting in global warming, which is one of the most challenging environmental issues. To prevent the damage caused by global warming, various technologies are being developed [78]. Among them, one technology called carbon capture, utilization, and storage (CCUS) has received great attention. In the CCUS system, CO2 capture is the key stage that matures this technology. According to the manufacturing methods of CO2, pre-combustion capture, post-combustion capture, and oxyfuel combustion were the main approaches. Regardless of different approaches, the CO2 separation process was the critical part of CO2 capture, such as absorption, membranes, adsorption, chemical looping combustion, and calcium looping, and adsorption has been considered one of the most promising ways to capture CO2.
Compared with other CO2 adsorbents, zeolite exhibited a stable cycle, huge surface area, and fast kinetics of CO2 adsorption. X-type zeolite, Y-type zeolite, and A-type zeolite were usually studied as CO2 capture adsorbents. Therefore, zeolite has received increasing attention in CO2 capture over the past decades. It has been found that there are several factors influencing the capture efficiency of CO2, such as Si/Al, distribution of pore sizes, exchangeable ions, and moisture effects.
As a kind of aluminosilicate, Al ions can replace Si ions in the TO4 structure, which breaks the charge balance of the original structure and leads to increasing basicity. It was reported that higher basicity contributes to higher CO2 adsorption [78]. Zeolites with a lower Si/Al ratio were more stable at high temperatures and less likely to exchange with other ions. To achieve higher CO2 capacity, the chemical composition of zeolites can be optimized. Scholars found that the Si/Al of GIS-type zeolite strongly affected the mechanism of CO2 adsorption. Specifically, for Na-GIS with Si/Al lower than 2.2, the CO2 adsorption was hindered, and CO2 uptake was negligible due to the large number of extra-framework cations near the 8-ring window [79].
The pore size of different zeolites was usually distributed in the range from 0.5 nm to 1.2 nm due to their crystalline nature. It means that only when the kinetic diameter of gas molecules is smaller than the pore size the gas can be adsorbed. To strengthen the adsorption efficiency of CO2, it is very promising to synthesize zeolites with mesopores. Due to the hierarchical structure, the obstruction of molecular diffusion is obviously reduced, which can consequently promote CO2 adsorption. Chen et al. prepared mesoporous LTA zeolite (Meso-LTA); it exhibited faster CO2 adsorption kinetics (1 bar) and higher CO2 adsorption capacities (>10 bar) at 298 K compared with microporous LTA zeolite [80].
CO2 adsorption is essentially the interaction between CO2 molecules and the electric field caused by ions in zeolites, which is similar to CO2 desorption. Therefore, the CO2 capture efficiency can be promoted by adjusting the chemical composition of zeolite, for instance, ion exchange. For instance, the simulation and experimental results of ion-exchanged 13X zeolite (FAU type) indicated that, compared with Na+ and K+, LiX-80 exhibited better separation of CO2/N2. Furthermore, the doping of Pd2+ and Ag+ changed the gradient of potential and strengthened the electrostatic potential (ESP), which increased the CO2 adsorption capacity (1.89%) and the CO2/N2 selectivity (85.97%) compared with LiX-80 [81].
Apart from the above factors, in flue gas, vapor is always accompanied by CO2. The water molecules may compete with CO2 molecules for the active sites in zeolites, which dramatically reduces the efficiency of CO2 adsorption. Some scholars proposed a strategy that fabricating a shell around the zeolite using sol-gel coating and a poly-ethylenimine impregnation process, and the diffusion of water molecules was hindered [82].
As a kind of promising CO2 solid adsorbent, zeolite has drawn great attention for the past decades and exhibited excellent performance. However, its application still faces a lot of challenges due to its structure characteristics and chemical composition. The modification of zeolite used for CO2 capture will still be a research focus in the future.

3. Conclusions and Future Perspectives

Zeolite catalyst is featured in energy conversion, selective separation, pollution prevention, and treatment. Generally, it exhibited better adsorption when the pore size of zeolite was equivalent to that of the adsorbed molecule. Therefore, it was very important to select the appropriate zeolite for different catalytic reactions. Since a state of long-term hydrothermal treatment at high temperatures is a necessary condition for its synthesis, numerous problems such as long synthesis time, high energy consumption, high equipment cost, and safety issues caused by the use of high-pressure hot-pressing tanks continuously emerged. There is no doubt that shortening synthesis time and lowering synthesis temperature are desperately needed. Reducing synthesis costs (energy consumption, autoclave productivity, etc.) and environmental footprint and safety issues has always been the primary issue.
Despite the special structure and excellent properties, the promotion of the application of zeolite catalysts also faces great challenges, such as the selectivity of specific products, temperature adaptability, and the lifetime of catalysts. It was also noticed that the retention of the reaction medium leads to the generation of by-products, resulting in micropore blockage and reduced catalytic activity. However, it can be noticed that upgrading the preparation method, regulating the pore structure, and increasing acid sites are three awesome measures to improve the properties of zeolite catalysts. Zeolites are expected to show their great advantage in energy-saving catalytic processes, environmentally friendly adsorption and separation, and energy storage.

Author Contributions

Conceptualization, W.Z., D.F. and Y.Z.; methodology, W.Z. and D.F.; software, W.Z.; validation, G.H., D.F. and Y.Z.; formal analysis, D.F. and Y.Z.; investigation, W.Z. and D.F.; resources, D.L. and Y.Z.; data curation, G.H., D.L. and Y.Z.; writing—original draft preparation, W.Z.; writing—review and editing, D.F. and G.H.; visualization, D.L. and Y.Z.; supervision, D.F. and Y.Z.; project administration, D.F.; funding acquisition, D.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the 2023 Teaching Case Program of Graduate School for the course named contemporary analytical techniques for materials characterization (Wuhan University of Technology).

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mishra, R.K.; Chistie, S.M.; Naika, S.U.; Mohanty, K. Catalytic pyrolysis of biomass over zeolites for bio-oil and chemical production: A review on their structure, porosity and acidity co-relation. Bioresour. Technol. 2022, 366, 128189. [Google Scholar] [CrossRef] [PubMed]
  2. Moliner, M.; Martinez, C.; Corma, A. Multipore Zeolites: Synthesis and Catalytic Applications. Angew. Chem. Int. Ed. 2015, 54, 3560–3579. [Google Scholar] [CrossRef] [Green Version]
  3. Li, Y.; Li, L.; Yu, J.H. Applications of Zeolites in Sustainable Chemistry. Chemistry 2017, 3, 928–949. [Google Scholar] [CrossRef] [Green Version]
  4. Martínez, C.; Vidal-Moya, A.; Yilmaz, B.; Kelkar, C.P.; Corma, A. Minimizing rare earth content of FCC catalysts: Understanding the fundamentals on combined P-La stabilization. Catal. Today 2023, 418, 114123. [Google Scholar] [CrossRef]
  5. Wang, T.; Le, T.; Ravindra, A.V.; Jue, H.; Zhang, L.; Wang, S. Enhanced regeneration of spent FCC catalyst by using oxalic acid-sulfuric acid mixture under ultrasonic irradiation. J. Mater. Res. Technol. 2021, 15, 7085–7099. [Google Scholar] [CrossRef]
  6. Zhang, L.; Hu, Q.; Qin, Y.; Liu, H.; Liu, H.; Cao, G.; Gao, X.; Song, L.; Zhaolin, S. Optimizing the accessibility of zeolite Y on FCC catalyst to improve heavy oil conversion capacity. Microporous Mesoporous Mater. 2023, 359, 112627. [Google Scholar] [CrossRef]
  7. Cui, W.; Zhu, D.; Tan, J.; Chen, N.; Fan, D.; Wang, J.; Han, J.; Wang, L.; Tian, P.; Liu, Z. Synthesis of mesoporous high-silica zeolite Y and their catalytic cracking performance. Chin. J. Catal. 2022, 43, 1945–1954. [Google Scholar] [CrossRef]
  8. Zhang, R.; Ju, Y.; Wu, P.; Chen, J.; Lv, Z.; Zhang, Y.; Song, S.; Zhang, Z.; Ma, C.; Zhang, R.; et al. Efficiently reducing olefin content of FCC gasoline over ZSM-5 zeolite based catalyst via hydro-upgrading. Catal. Today 2022, 405–406, 57–65. [Google Scholar] [CrossRef]
  9. Liu, P.; Cui, Y.; Wang, J.; Du, X.; Zhang, H.; Humphries, A.; Jia, M.; Yu, J. Structure stabilization of zeolite Y induced by yttrium and its role in promoting n-docosane conversion. Microporous Mesoporous Mater. 2021, 323, 111225. [Google Scholar] [CrossRef]
  10. Yamazaki, H.; Hasegawa, H.; Tanaka, C.; Takamiya, Y.; Mitsui, T.; Mizuno, T. Al ion-exchanged USY in FCC catalyst for high LPG yield. Catal. Commun. 2021, 159, 106354. [Google Scholar] [CrossRef]
  11. Miao, P.; Zhu, X.; Zhou, Z.; Feng, X.; Miao, J.; Hou, C.; Li, C. Combined dealkylation and transalkylation reaction in FCC condition for efficient conversion of light fraction light cycle oil into value-added products. Fuel 2021, 304, 121356. [Google Scholar] [CrossRef]
  12. Zhong, M.; Li, X.; Chu, X.; Gui, H.; Zuo, S.; Yao, C.; Li, Z.; Chen, Y. Solar driven catalytic conversion of cellulose biomass into lactic acid over copper reconstructed natural mineral. Appl. Catal. B Environ. 2022, 317, 121718. [Google Scholar] [CrossRef]
  13. Perego, C.; Bosetti, A. Biomass to fuels: The role of zeolite and mesoporous materials. Microporous Mesoporous Mater. 2011, 144, 28–39. [Google Scholar] [CrossRef]
  14. Ong, H.C.; Yu, K.L.; Chen, W.-H.; Pillejera, M.K.; Bi, X.; Tran, K.-Q.; Pétrissans, A.; Pétrissans, M. Variation of lignocellulosic biomass structure from torrefaction: A critical review. Renew. Sustain. Energy Rev. 2021, 152, 111698. [Google Scholar] [CrossRef]
  15. Yan, P.; Wang, H.; Liao, Y.; Wang, C. Zeolite catalysts for the valorization of biomass into platform compounds and biochemicals/biofuels: A review. Renew. Sustain. Energy Rev. 2023, 178, 113219. [Google Scholar] [CrossRef]
  16. Abu Zarin, M.A.; Zainol, M.M.; Ramli, N.A.S.; Amin, N.A.S. Zeolite immobilized ionic liquid as an effective catalyst for conversion of biomass derivatives to levulinic acid. Mol. Catal. 2022, 528, 112506. [Google Scholar] [CrossRef]
  17. Yan, P.; Wang, H.; Liao, Y.; Sun, P.; Wang, C. Introducing mesopore and regulating Al distribution for improving catalytic performances of ZSM-5 in furfuryl alcohol to levulinic acid. Fuel 2022, 329, 125213. [Google Scholar] [CrossRef]
  18. Chung, N.H.; Oanh, V.T.; Thoa, L.K.; Hoang, P.H. Catalytic Conversion of Glucose into 5-Hydroxymethyl Furfural Over Cu-Cr/ZSM-5 Zeolite. Catal. Lett. 2020, 150, 170–177. [Google Scholar] [CrossRef]
  19. Xu, S.Q.; Pan, D.H.; Hu, F.; Wu, Y.F.; Wang, H.Z.; Chen, Y.; Yuan, H.; Gao, L.J.; Xiao, G.M. Highly efficient Cr/beta zeolite catalyst for conversion of carbohydrates into 5-hydroxymethylfurfural: Characterization and performance. Fuel Process. Technol. 2019, 190, 38–46. [Google Scholar] [CrossRef]
  20. Werpy, T.; Petersen, G. Top Value Added Chemicals from Biomass: Volume I—Results of Screening for Potential Candidates from Sugars and Synthesis Gas; Medium: Boston, MA, USA, 2004; 76p. [Google Scholar]
  21. Venschott, M.; Hoelderich, W.F.; Eisenacher, M. 2nd generation PLA. Lactide formation directly from aqueous lactic acid. Catal. Commun. 2023, 177, 106636. [Google Scholar] [CrossRef]
  22. Ma, H.; Tingelstad, P.; Chen, D. Lactic acid production by catalytic conversion of glucose: An experimental and techno-economic evaluation. Catal. Today 2023, 408, 2–8. [Google Scholar] [CrossRef]
  23. Kim, J.; Bang, J.; Choi, J.-S.; Lim, D.-H.; Guk, D.; Jae, J. Selective conversion of lactic acid to renewable acrylic acid over SDA-free Na-ZSM-5: The critical role of basic sites of sodium oxide. J. Catal. 2023, 421, 271–284. [Google Scholar] [CrossRef]
  24. Ye, J.; Chen, C.; Zheng, Y.; Zhou, D.; Liu, Y.; Chen, D.; Ni, L.; Xu, G.; Wang, F. Efficient conversion of cellulose to lactic acid over yttrium modified siliceous Beta zeolites. Appl. Catal. A Gen. 2021, 619, 118133. [Google Scholar] [CrossRef]
  25. Koyunoğlu, C.; Gündüz, F.; Karaca, H.; Çınar, T.; Soyhan, G.G. Developing an adaptive catalyst for an FCC reactor using a CFD RSM, CFD DPM, and CFD DDPM–EM approach. Fuel 2023, 334, 126550. [Google Scholar] [CrossRef]
  26. Silva, J.M.; Ribeiro, M.F.; Graça, I.; Fernandes, A. Bio-oils/FCC co-processing: Insights into the adsorption of guaiacol on Y zeolites with distinct acidity and textural properties. Microporous Mesoporous Mater. 2021, 323, 111170. [Google Scholar] [CrossRef]
  27. Le-Phuc, N.; Tran, T.V.; Phan, T.T.; Ngo, P.T.; Ha, Q.L.M.; Luong, T.N.; Tran, T.H.; Phan, T.T. High-efficient production of biofuels using spent fluid catalytic cracking (FCC) catalysts and high acid value waste cooking oils. Renew. Energy 2021, 168, 57–63. [Google Scholar] [CrossRef]
  28. Magrini, K.; Olstad, J.; Peterson, B.; Jackson, R.; Parent, Y.; Mukarakate, C.; Iisa, K.; Christensen, E.; Seiser, R. Feedstock and catalyst impact on bio-oil production and FCC Co-processing to fuels. Biomass Bioenergy 2022, 163, 106502. [Google Scholar] [CrossRef]
  29. Zabihpour, A.; Ahmadpour, J.; Yaripour, F. Strategies to control reversible and irreversible deactivation of ZSM-5 zeolite during the conversion of methanol to propylene (MTP): A review. Chem. Eng. Sci. 2023, 273, 118639. [Google Scholar] [CrossRef]
  30. Qiu, B.; Lu, W.-D.; Gao, X.-Q.; Sheng, J.; Ji, M.; Wang, D.; Lu, A.-H. Boosting the propylene selectivity over embryonic borosilicate zeolite catalyst for oxidative dehydrogenation of propane. J. Catal. 2023, 417, 14–21. [Google Scholar] [CrossRef]
  31. Zhao, Y.-N.; Fan, S.-B.; Ma, Q.-X.; Zhang, J.-L.; Zhao, T.-S. Methanol converting to propylene on weakly acidic and hierarchical porous MFI zeolite. J. Fuel Chem. Technol. 2022, 50, 210–217. [Google Scholar] [CrossRef]
  32. Feng, R.; Zhou, P.; Liu, B.; Yan, X.; Hu, X.; Zhou, M. Direct synthesis of HZSM-5 zeolites with enhanced catalytic performance in the methanol-to-propylene reaction. Catal. Today 2022, 405–406, 299–308. [Google Scholar] [CrossRef]
  33. Feng, R.; Liu, B.; Zhou, P.; Yan, X.; Hu, X.; Zhou, M.; Yan, Z. Influence of framework Al distribution in HZSM-5 channels on catalytic performance in the methanol to propylene reaction. Appl. Catal. A Gen. 2022, 629, 118422. [Google Scholar] [CrossRef]
  34. Kalantari, N.; Bekheet, M.F.; Delir Kheyrollahi Nezhad, P.; Back, J.O.; Farzi, A.; Penner, S.; Delibaş, N.; Schwarz, S.; Bernardi, J.; Salari, D.; et al. Effect of chromium and boron incorporation methods on structural and catalytic properties of hierarchical ZSM-5 in the methanol-to-propylene process. J. Ind. Eng. Chem. 2022, 111, 168–182. [Google Scholar] [CrossRef]
  35. Tuo, J.; Lv, J.; Fan, S.; Li, H.; Yang, N.; Cheng, S.; Gao, X.; Zhao, T. One-pot synthesis of [Mn,H]ZSM-5 and the role of Mn in methanol-to-propylene reaction. Fuel 2022, 308, 121995. [Google Scholar] [CrossRef]
  36. Zhang, L.; Yang, L.; Liu, R.; Shao, X.; Dai, W.; Wu, G.; Guan, N.; Guo, Z.; Zhu, W.; Li, L. Design of plate-like H[Ga]MFI zeolite catalysts for high-performance methanol-to-propylene reaction. Microporous Mesoporous Mater. 2022, 333, 111767. [Google Scholar] [CrossRef]
  37. Dai, W.; Zhang, L.; Liu, R.; Huo, Z.; Dai, W.; Guan, N. Facile fabrication of a plate-like ZSM-5 zeolite as a highly efficient and stable catalyst for methanol to propylene conversion. Mater. Today Sustain. 2023, 22, 100364. [Google Scholar] [CrossRef]
  38. Yang, G.; Yan, X.; Chen, Y.; Guo, X.-J.; Lang, W.-Z.; Guo, Y.-J. Improved propylene selectivity and superior catalytic performance of Ga-xMg/ZSM-5 catalysts for propane dehydrogenation (PDH) reaction. Appl. Catal. A Gen. 2022, 643, 118778. [Google Scholar] [CrossRef]
  39. Xiong, Y.; Tian, T.; L’Hermitte, A.; Méndez, A.S.J.; Danaci, D.; Platero-Prats, A.E.; Petit, C. Using silver exchange to achieve high uptake and selectivity for propylene/propane separation in zeolite Y. Chem. Eng. J. 2022, 446, 137104. [Google Scholar] [CrossRef]
  40. Moradi, H.; Azizpour, H.; Mohammadi, M. Study of adsorption of propane and propylene on CHA zeolite in different Si/Al ratios using molecular dynamics simulation. Powder Technol. 2023, 419, 118329. [Google Scholar] [CrossRef]
  41. Wei, S.; Dai, H.; Long, J.; Lin, H.; Gu, J.; Zong, X.; Yang, D.; Tang, Y.; Yang, Y.; Dai, Y. Nonoxidative propane dehydrogenation by isolated Co2+ in BEA zeolite: Dealumination-determined key steps of propane CH activation and propylene desorption. Chem. Eng. J. 2023, 455, 140726. [Google Scholar] [CrossRef]
  42. Liu, Y.; Tao, H.; Yang, X.; Wu, X.; Li, J.; Zhang, C.; Yang, R.T.; Li, Z. Adsorptive purification of NOx by HZSM-5 zeolites: Effects of Si/Al ratio, temperature, humidity, and gas composition. Microporous Mesoporous Mater. 2023, 348, 112331. [Google Scholar] [CrossRef]
  43. Sunil Kumar, M.; Alphin, M.S.; Manigandan, S.; Vignesh, S.; Vigneshwaran, S.; Subash, T. A review of comparison between the traditional catalyst and zeolite catalyst for ammonia-selective catalytic reduction of NOx. Fuel 2023, 344, 128125. [Google Scholar] [CrossRef]
  44. Zang, Y.; Bi, Y.; Liu, C.; Zhang, Y.; Li, Q.; Wang, Y.; Zhang, M.; Liu, Q.; Zhang, Z. The study on Cu-ZK-5 catalyst of copper content regulation and anti-propylene poisoning mechanism in NH3-SCR reaction. Fuel 2023, 340, 127442. [Google Scholar] [CrossRef]
  45. Jin, P.; Yang, L.; Sheng, Z.; Chu, X.; Chen, D. Selective catalytic reduction of NOx with NH3 and tolerance to H2O & SO2 at high temperature over zeolite supported indium-copper bimetallic catalysts for gas turbine. J. Environ. Chem. Eng. 2023, 11, 109218. [Google Scholar] [CrossRef]
  46. Gao, L.; Gao, W.; Wang, H.; Xu, S.; Tian, X.; Cao, J.; Chen, J.; Zhang, Q.; Ning, P.; Hao, J. Boosting low-temperature and high-temperature hydrothermal stability of Cu/SAPO-34 for NOx removal via yttrium decoration. Chem. Eng. J. 2023, 455, 140520. [Google Scholar] [CrossRef]
  47. Liu, Y.; Wu, X.; Yang, X.; Tao, H.; Li, J.; Zhang, C.; Yang, R.T.; Li, Z. Enhancement of NOx adsorption performance on zeolite via a facile modification strategy. J. Hazard. Mater. 2023, 443, 130225. [Google Scholar] [CrossRef]
  48. Wang, L.; Ren, Y.; Yu, X.; Peng, C.; Yu, D.; Zhong, C.; Hou, J.; Yin, C.; Fan, X.; Zhao, Z.; et al. Novel preparation method, catalytic performance and reaction mechanisms of PrxMn1−xOδ/3DOM ZSM-5 catalysts for the simultaneous removal of soot and NOx. J. Catal. 2023, 417, 226–247. [Google Scholar] [CrossRef]
  49. Kang, T.H.; Kim, H.S.; Lee, H.; Kim, D.H. Synergistic effect of V2O5-WO3/TiO2 and H-ZSM-5 catalysts prepared by physical mixing on the selective catalytic reduction of NOx with NH3. Appl. Surf. Sci. 2023, 614, 156159. [Google Scholar] [CrossRef]
  50. Zhang, L.; Shan, Y.; Yan, Z.; Liu, Z.; Yu, Y.; He, H. Efficient Pt/KFI zeolite catalysts for the selective catalytic reduction of NOx by hydrogen. J. Environ. Sci. 2024, 138, 102–111. [Google Scholar] [CrossRef]
  51. Wang, S.; He, B.; Wang, Y.; Wu, X.; Duan, H.; Di, J.; Yu, Z.; Liu, Y.; Xin, Z.; Jia, L.; et al. Hydrogen production from the steam reforming of bioethanol over novel supported Ca/Ni-hierarchical Beta zeolite catalysts. Int. J. Hydrogen Energy 2021, 46, 36245–36256. [Google Scholar] [CrossRef]
  52. Husin, H.; Erdiwansyah, E.; Ahmadi, A.; Nasution, F.; Rinaldi, W.; Abnisa, F.; Mamat, R. Efficient hydrogen production by microwave-assisted catalysis for glycerol-water solutions via NiO/zeolite-CaO catalyst. South Afr. J. Chem. Eng. 2022, 41, 43–50. [Google Scholar] [CrossRef]
  53. Zhang, H.; Zhong, L.; Bin Samsudin, I.; Okumura, K.; Tan, H.-R.; Li, S.; Jaenicke, S.; Chuah, G.-K. Mg-stabilized subnanometer Rh particles in zeolite Beta as highly efficient catalysts for selective hydrogenation. J. Catal. 2022, 405, 489–498. [Google Scholar] [CrossRef]
  54. Da Costa-Serra, J.F.; Miralles-Martínez, A.; García-Muñoz, B.; Maestro-Cuadrado, S.; Chica, A. Ni and Co-based catalysts supported on ITQ-6 zeolite for hydrogen production by steam reforming of ethanol. Int. J. Hydrogen Energy 2022, 48, 26518–26525. [Google Scholar] [CrossRef]
  55. Ismaila, A.; Chen, H.; Fan, X. Nickel encapsulated in silicalite-1 zeolite catalysts for steam reforming of glycerol (SRG) towards renewable hydrogen production. Fuel Process. Technol. 2022, 233, 107306. [Google Scholar] [CrossRef]
  56. Yu, L.; Zhang, R.; Cao, C.; Liu, L.; Fang, J.; Jin, H. Hydrogen production from supercritical water gasification of lignin catalyzed by Ni supported on various zeolites. Fuel 2022, 319, 123744. [Google Scholar] [CrossRef]
  57. Chen, M.; Sun, G.; Wang, Y.; Liang, D.; Li, C.; Wang, J.; Liu, Q. Steam reforming of methanol for hydrogen production over attapulgite-based zeolite-supported Cu-Zr catalyst. Fuel 2022, 314, 122733. [Google Scholar] [CrossRef]
  58. Zhou, Y.; Wang, P.; Qin, L.; Kang, S.-Z.; Li, X. Double shell composite nanoarchitectonics of Cu2O core with TiO2/metal-organic frameworks for efficient hydrogen generation. Int. J. Hydrogen Energy 2023, 48, 629–639. [Google Scholar] [CrossRef]
  59. Saka, C. Highly active hydrogen generation from sodium borohydride methanolysis and ethylene glycolysis reactions using protonated chitosan-zeolite hybrid metal-free particles. Appl. Catal. B Environ. 2023, 325, 122335. [Google Scholar] [CrossRef]
  60. Sun, T.; Wei, J.; Zhou, C.; Wang, Y.; Shu, Z.; Zhou, J.; Chen, J. Facile preparation and enhanced photocatalytic hydrogen evolution of cation-exchanged zeolite LTA supported TiO2 photocatalysts. Int. J. Hydrogen Energy 2023, 48, 13851–13863. [Google Scholar] [CrossRef]
  61. Shen, Y. Biomass-derived porous carbons for sorption of Volatile organic compounds (VOCs). Fuel 2023, 336, 126801. [Google Scholar] [CrossRef]
  62. Yang, F.; Li, W.; Ou, R.; Lu, Y.; Dong, X.; Tu, W.; Zhu, W.; Wang, X.; Li, L.; Yuan, A.; et al. Superb VOCs capture engineering carbon adsorbent derived from shaddock peel owning uncompromising thermal-stability and adsorption property. Chin. J. Chem. Eng. 2022, 47, 120–133. [Google Scholar] [CrossRef]
  63. Zhang, Y.; Yu, Q.; Yuan, Y.; Tang, X.; Zhao, S.; Yi, H. Adsorption behavior of Mo-MEL zeolites for reducing VOCs from cooking oil fumes. Sep. Purif. Technol. 2023, 322, 124059. [Google Scholar] [CrossRef]
  64. Yu, Q.; Feng, Y.; Wei, J.; Tang, X.; Yi, H. Development of Mn-Si-MEL as a bi-functional adsorption-catalytic oxidation material for VOCs elimination. Chin. Chem. Lett. 2022, 33, 3087–3090. [Google Scholar] [CrossRef]
  65. Liu, H.; Wei, K.; Long, C. Enhancing adsorption capacities of low-concentration VOCs under humid conditions using NaY@meso-SiO2 core–shell composite. Chem. Eng. J. 2022, 442, 136108. [Google Scholar] [CrossRef]
  66. Yin, T.; Meng, X.; Wang, S.; Yao, X.; Liu, N.; Shi, L. Study on the adsorption of low-concentration VOCs on zeolite composites based on chemisorption of metal-oxides under dry and wet conditions. Sep. Purif. Technol. 2022, 280, 119634. [Google Scholar] [CrossRef]
  67. Feng, A.; Yu, Y.; Mi, L.; Cao, Y.; Yu, Y.; Song, L. Synthesis and VOCs adsorption performance of surfactant-templated USY zeolites with controllable mesopores. Chem. Phys. Lett. 2022, 798, 139578. [Google Scholar] [CrossRef]
  68. Lu, S.; Han, R.; Wang, H.; Song, C.; Ji, N.; Lu, X.; Ma, D.; Liu, Q. Three birds with one stone: Designing a novel binder-free monolithic zeolite pellet for wet VOC gas adsorption. Chem. Eng. J. 2022, 448, 137629. [Google Scholar] [CrossRef]
  69. Chen, L.H.; Sun, M.H.; Wang, Z.; Yang, W.M.; Xie, Z.K.; Su, B.L. Hierarchically Structured Zeolites: From Design to Application. Chem. Rev. 2020, 120, 11194–11294. [Google Scholar] [CrossRef]
  70. Ivanova, I.I.; Knyazeva, E.E. Micro-mesoporous materials obtained by zeolite recrystallization: Synthesis, characterization and catalytic applications. Chem. Soc. Rev. 2013, 42, 3671–3688. [Google Scholar] [CrossRef]
  71. Zhao, C.; Hu, X.; Liu, C.; Chen, D.; Yun, J.; Jiang, X.; Wei, N.; Li, M.; Chen, Z. Hierarchical architectures of ZSM-5 with controllable mesoporous and their particular adsorption/desorption performance for VOCs. J. Environ. Chem. Eng. 2022, 10, 106868. [Google Scholar] [CrossRef]
  72. Wang, J.; Shi, Y.; Kong, F.; Zhou, R. Low-temperature VOCs oxidation performance of Pt/zeolites catalysts with hierarchical pore structure. J. Environ. Sci. 2023, 124, 505–512. [Google Scholar] [CrossRef] [PubMed]
  73. Li, D.; Wang, L.; Lu, Y.; Deng, H.; Zhang, Z.; Wang, Y.; Ma, Y.; Pan, T.; Zhao, Q.; Shan, Y.; et al. New insights into the catalytic mechanism of VOCs abatement over Pt/Beta with active sites regulated by zeolite acidity. Appl. Catal. B Environ. 2023, 334, 122811. [Google Scholar] [CrossRef]
  74. Kong, F.; Li, G.; Wang, J.; Shi, Y.; Zhou, R. Promoting effect of acid sites in hierarchical porous Pt/ZSM-5 catalysts for low-temperature removal of VOCs. Appl. Surf. Sci. 2022, 606, 154888. [Google Scholar] [CrossRef]
  75. Yu, B.; Deng, H.; Lu, Y.; Pan, T.; Shan, W.; He, H. Adsorptive interaction between typical VOCs and various topological zeolites: Mixture effect and mechanism. J. Environ. Sci. 2024, 136, 626–636. [Google Scholar] [CrossRef]
  76. Zheng, Y.; Han, R.; Yang, L.; Yang, J.; Shan, C.; Liu, Q. Revealing opposite behaviors of catalyst for VOCs Oxidation: Modulating electronic structure of Pt nanoparticles by Mn doping. Chem. Eng. J. 2023, 465, 142807. [Google Scholar] [CrossRef]
  77. Wu, R.; Xiao, Y.; Zhang, P.; Lin, J.; Cheng, G.; Chen, Z.; Yu, R. Asphalt VOCs reduction of zeolite synthesized from solid wastes of red mud and steel slag. J. Clean. Prod. 2022, 345, 131078. [Google Scholar] [CrossRef]
  78. Yu, C.-H.; Huang, C.-H.; Tan, C.-S. A Review of CO2 Capture by Absorption and Adsorption. Aerosol Air Qual. Res. 2012, 12, 745–769. [Google Scholar] [CrossRef] [Green Version]
  79. Choi, H.J.; Hong, S.B. Effect of framework Si/Al ratio on the mechanism of CO2 adsorption on the small-pore zeolite gismondine. Chem. Eng. J. 2022, 433, 133800. [Google Scholar] [CrossRef]
  80. Chen, C.; Ahn, W.-S. CO2 adsorption on LTA zeolites: Effect of mesoporosity. Appl. Surf. Sci. 2014, 311, 107–109. [Google Scholar] [CrossRef]
  81. Chen, S.J.; Zhu, M.; Tang, Y.C.; Fu, Y.; Li, W.L.; Xiao, B. Molecular simulation and experimental investigation of CO2 capture in a polymetallic cation-exchanged 13X zeolite. J. Mater. Chem. A 2018, 6, 19570–19583. [Google Scholar] [CrossRef] [Green Version]
  82. Liu, X.; Gao, F.; Xu, J.; Zhou, L.; Liu, H.; Hu, J. Zeolite@Mesoporous silica-supported-amine hybrids for the capture of CO2 in the presence of water. Microporous Mesoporous Mater. 2016, 222, 113–119. [Google Scholar] [CrossRef]
Figure 1. Various applications of zeolites in sustainable energy.
Figure 1. Various applications of zeolites in sustainable energy.
Catalysts 13 01197 g001
Figure 2. Reaction pathway of bio-renewable lactic acid (reproduced from reference [23]).
Figure 2. Reaction pathway of bio-renewable lactic acid (reproduced from reference [23]).
Catalysts 13 01197 g002
Figure 3. Dual-cycle mechanism in the MTP conversion over the plate-like H[Ga]MFI catalysts [36].
Figure 3. Dual-cycle mechanism in the MTP conversion over the plate-like H[Ga]MFI catalysts [36].
Catalysts 13 01197 g003
Figure 4. Local structure evolution of BEA zeolite during dealumination and Co2+ introduction (reproduced from reference [41]).
Figure 4. Local structure evolution of BEA zeolite during dealumination and Co2+ introduction (reproduced from reference [41]).
Catalysts 13 01197 g004
Figure 5. Proposed schematic diagram of NOx adsorption process on Cu-HZSM5_IM [47].
Figure 5. Proposed schematic diagram of NOx adsorption process on Cu-HZSM5_IM [47].
Catalysts 13 01197 g005
Figure 6. Schematic diagram for the preparation of PrxMn1−xOd/3DOM ZSM-5 catalysts [48].
Figure 6. Schematic diagram for the preparation of PrxMn1−xOd/3DOM ZSM-5 catalysts [48].
Catalysts 13 01197 g006
Figure 7. (a) Structural configurations of Rh located on three different supports and corresponding binding energies and (b) relative energy levels for O–H bond dissociation of phenol in zeolite Beta with and without Mg [53].
Figure 7. (a) Structural configurations of Rh located on three different supports and corresponding binding energies and (b) relative energy levels for O–H bond dissociation of phenol in zeolite Beta with and without Mg [53].
Catalysts 13 01197 g007
Figure 8. Proposed mechanism in the Cu2O@TiO2@ZIF-8 photocatalyst [58].
Figure 8. Proposed mechanism in the Cu2O@TiO2@ZIF-8 photocatalyst [58].
Catalysts 13 01197 g008
Figure 9. The crystallization mechanism diagram for the ZSM-5 zeolites [71].
Figure 9. The crystallization mechanism diagram for the ZSM-5 zeolites [71].
Catalysts 13 01197 g009
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

Zhang, W.; Fang, D.; Huang, G.; Li, D.; Zheng, Y. Research and Application Development of Catalytic Redox Technology for Zeolite-Based Catalysts. Catalysts 2023, 13, 1197. https://doi.org/10.3390/catal13081197

AMA Style

Zhang W, Fang D, Huang G, Li D, Zheng Y. Research and Application Development of Catalytic Redox Technology for Zeolite-Based Catalysts. Catalysts. 2023; 13(8):1197. https://doi.org/10.3390/catal13081197

Chicago/Turabian Style

Zhang, Wentao, De Fang, Guanlin Huang, Da Li, and Yun Zheng. 2023. "Research and Application Development of Catalytic Redox Technology for Zeolite-Based Catalysts" Catalysts 13, no. 8: 1197. https://doi.org/10.3390/catal13081197

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