Next Article in Journal
Marine Mycosilver Nanoparticles: Screening, Evaluation of Their Antimicrobial Properties, and Synthesis Optimization
Previous Article in Journal
Adsorptive Removal of Hexavalent Chromium from Aqueous Solution Utilizing Activated Carbon Developed from Spathodea campanulata
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring the Potential of Zeolites for Sustainable Environmental Applications

by
Maura Mancinelli
and
Annalisa Martucci
*
Department of Physics and Earth Sciences, University of Ferrara, 44122 Ferrara, Italy
*
Author to whom correspondence should be addressed.
Sustain. Chem. 2025, 6(1), 9; https://doi.org/10.3390/suschem6010009
Submission received: 27 December 2024 / Revised: 18 February 2025 / Accepted: 10 March 2025 / Published: 17 March 2025

Abstract

:
Zeolites are amongst the most extensively explored crystalline microporous materials because of their variable chemical composition, framework geometry, pore dimensions, and tunability. Due to their high surface area, adsorption selectivity, mechanical, biological, chemical, and thermal stability, these molecular sieves are widely used in adsorption, catalysis, ion exchange, and separation technologies. This short review highlights the notable progress achieved in leveraging the properties of zeolite materials for multiple applications, including gas separation and storage, adsorption, catalysis, chemical sensing, and biomedical applications. The aim is to emphasize their capabilities by showcasing important achievements that have driven research in this field toward new and unforeseen areas of material chemistry.

Graphical Abstract

1. Introduction

Zeolites are among the most extensively studied crystalline porous materials due to their tunable framework geometry, variable chemical composition, and well-defined pore structures. Their high surface area, adsorption selectivity, and exceptional stability make them indispensable in catalysis, adsorption, ion exchange, and separation technologies [1,2,3].
A key feature of zeolites is their porosity, which governs molecular transport and reactivity, playing a critical role in applications such as gas separation, water purification, and catalysis.
The simplest way to classify zeolites is by the number of tetrahedra that form their pore openings (see Table 1). According to Flanigen et al. [1], zeolites with tiny pores have channels defined by eight-membered rings (8R) with diameters ranging from 3.5 to 4.5 Å (for example, gismondine). Medium-pore zeolites contain 10-membered rings (10R) with apertures between 4.5 and 6 Å (such as ferrierite). Large-pore zeolites are characterized by 12-membered rings (12R), with free diameters ranging from 6 to 8 Å (for instance, mordenite). Finally, zeolites with pore openings consisting of 14 or more membered rings (14 + R) fall into the category of extra-large pores (such as coverite) [2] (Figure 1).
The structural properties of the channel system—including orientation, geometry, and connectivity—determine their activity, selectivity, and diffusion processes. While increased channel connectivity enhances catalytic performance, lower-dimensional channels are prone to rapid deactivation. Shape selectivity is strongly dictated by pore architecture, as variations in channel geometry (e.g., sinusoidal or linear) influence molecular transport and reaction efficiency [3,4,5,6].
However, the intrinsic microporosity of zeolites often restricts the diffusion of reactants and products, limiting accessibility to active sites. To address these constraints, researchers have explored mesoporosity engineering through templating, post-synthetic modifications (e.g., dealumination, desilication), and hydrothermal steam treatments. The integration of mesopores within microporous frameworks significantly enhances catalytic efficiency by reducing intracrystalline diffusion barriers while preserving shape-selective properties [7,8,9].
This review highlights recent advancements in the application of zeolites across several fields, including gas separation, adsorption, catalysis, chemical sensing, and biomedical technologies. It emphasizes their emerging roles in CO2 capture, bioprocessing, and pollutant removal, particularly in wastewater treatment and energy storage. The novelty of this work lies in its comprehensive assessment of zeolites’ evolving functionalities, bridging traditional applications with cutting-edge environmental and technological solutions. Additionally, it critically examines challenges in pollutant degradation, underscoring the need for further research in zeolite-based environmental remediation.
This review synthesizes recent innovations and provides insights into the dynamic interplay between zeolite structure and functionality, offering new directions for future research and technological development.

2. Zeolites for Sustainable Environmental Applications

To effectively control and optimize the properties of zeolites, it is crucial to understand their framework types, structural connectivity of tetrahedrally coordinated atoms, and crystal chemistry. The Si/Al ratio (SAR) of zeolites is a critical parameter affecting properties such as the maximum ion-exchange capacity, thermal and hydrothermal stability, hydrophobicity, concentration, and strength of acid sites, along with catalytic activity and selectivity (Figure 2). Zeolites’ physical, chemical, and environmental remediation applications are significantly influenced by their Si/Al molar ratio. Because of their strong ion-exchange capacity, low silica zeolites (Si/Al ≤ 2) are excellent at removing heavy metals and ammonium. Intermediate silica zeolites (2 < Si/Al < 5) are also good at removing organic pollutants. Finally, high silica zeolites (Si/Al ≥ 5) are better at removing organic pollutants due to their increased hydrophobicity.
Zeolites with high SAR exhibit higher thermal stability, reaching temperatures as high as 1300 °C, and lower dehydration temperatures, due to their surfaces becoming more hydrophobic [10,11,12,13], because of the selective adsorption capability alongside the size and shape of the pores and channels through which host molecules diffuse. Consequently, zeolites can act like molecular sieves, meaning that, when a mixture containing two components differing in shape and size flows through the zeolite channels, the different components can be successfully separated.
Specifically, the possibility to modify both the type and location of extra-framework cations, as well as the incorporation of framework heteroatoms, allows for changes in the selectivity and catalytic activity of zeolites [14,15]. A wide range of metal species (M = Al, Ga, Ti, Sn, etc.) can be incorporated into these sites, each exerting a distinct influence on the catalytic behavior of the zeolite. Three different categories can be recognized: (1) metal-containing zeolites, which include framework and framework-associated metals; (2) metal-exchanged zeolites, containing exchangeable extra-framework metal cations; and (3) metal–zeolite composites, comprising (a) extra-framework metal single atoms, (b) subnanometric metal clusters (<1 nm) in confined sites, and (c) nanoparticles (>1 nm) requiring lattice disruption to traverse micropores. These extra-framework species are collectively termed metal–zeolites due to their transformation into clusters or nanoparticles via sintering or oxidation–reduction processes. Metals at zeolite framework sites are four-fold coordinated with framework oxygen, generating Si–O–M linkages. Framework-associated metals are partially displaced, exhibiting both framework Si–O–M linkages and terminal M–OH groups. Extra-framework metal species, fully dislodged from the framework, lack Si–O–M linkages and are hosted in channels, cavities, or secondary mesopores, interacting with the framework via electrostatic, van der Waals, or chemical interactions.
The synthesis methods are integral in determining the distribution and localization of metal species within the framework, framework-associated, or extra-framework positions, all of which have a significant impact on their catalytic properties. The distribution, location, and size of these metal species are critical for establishing structure–activity relationships, thereby advancing the design and optimization of highly efficient heterogeneous catalysts with enhanced activity and stability. Conventional hydrothermal synthesis is the primary method for preparing metal-containing zeolites by introducing metal sources into synthesis gels. The dry-gel conversion method effectively increases metal content, while post-synthesis treatments, which replace framework T atoms with other elements, offer an alternative when direct methods are insufficient. Interzeolite transformation can also be used to incorporate metal atoms into zeolite frameworks ([14] and references therein).
Zeolite-supported metal catalysts facilitate the formation of ultrasmall and remarkably stable metal species, contributing to their exceptional thermal and hydrothermal stability during catalytic reactions, particularly in challenging environments. Additionally, they exhibit unique shape-selectivity, which improves their performance in a wide range of catalytic activities. In heterogeneous catalysis, these materials favor essential reactions, such as hydrogenation of CO and CO2, dehydrogenation, oxidation, selective catalytic reduction of NOx using ammonia (NH3-SCR), and hydroisomerization. Nickel and cerium–zeolites (i.e., MFI, BEA, FAU, MOR, ITE) are efficient catalysts for converting CO2 to methane, because the presence of oxygen vacancies in CeO2 species enhances the activation of CO2 and increases the dispersion of Ni species [15,16,17]. Not only rare-earth- and noble-metal-supported zeolites (Pd, Ru, and Rh) but also alkali and alkaline metal ions (Li, Na, K, Cs, Ba, Ca, Mg) enhance CO2 adsorption and activation. Their unique properties (i.e., excellent selectivity, long-term stability, and recyclability) not only improve reaction outcomes, but also pave the way for more sustainable and efficient chemical processes.
Indeed, the presence of hydroxyl Si–OH–Al groups (Brønsted acid sites, BAS) along with the insertion of tri- or tetravalent heteroatoms within the framework sites confer enhanced catalytic properties. Similarly, extra-framework aluminum species have also been reported to act as a Lewis acid site (LAS). BAS and LAS in zeolites are extensively explored key roles in many biomass conversion processes. For example, Brønsted acidic zeolites with large pores facilitate the conversion of lactic acid (LA) into lactide, as well as catalyze condensation reactions. The zeolite H-Beta demonstrates remarkable selectivity in the synthesis of lactide, achieving a yield of approximately 79% at full conversion of LA. Additionally, zeolites with large and medium pores may also serve as catalysts for the oxidation of bioethanol to acetaldehyde or for the conversion of sugar-based 2,5-dimethylfuran to aromatic compounds. Indeed, the channels of high dimensions, together with the presence of additional supercages, render large-pore zeolites suitable for large-scale processes such as the fluid catalytic cracking of oil for gasoline production and transalkylation in the refinery processes. Furthermore, medium-pore zeolites are widely used as hydrocracking catalysts in processes that necessitate greater shape selectivity toward organic compounds, including olefin isomerization, dewaxing in refineries, and xylene isomerization ([18] and references therein). LAS are most extensively evaluated for alkene epoxidation, Baeyer−Villiger oxidation of ketones and aromatic aldehydes, alcohol dehydration, and conversion of sugars, furan, and acid derivatives. The combination of LAS with additional functional sites (for example, Brønsted acid sites, metal sites, and alkaline sites) could open new possibilities for producing high-value chemicals or fuels from biomass-derived oxygenates.

2.1. Zeolites for Inorganic Contaminant Removal

The SAR significantly impacts the ability to host water and extra-framework cations [19]. The high ion-exchange capability and the selectivity of zeolites are vital for removing ammonia, heavy metals, and radionuclides from natural waters and wastewater. According to Jin et al. [20], heavy metal adsorption initially involves the external surface of the particles, followed by the counter-diffusion of interchangeable cations, and, finally, the sorption in the microporosities of the zeolite. Zeolites such as X, Y, A, and P can achieve the removal of up to 96% of heavy metals, 90% of phosphoric compounds, 96% of dyes, 80% of nitrogen compounds (96% specifically for ammonium), and 89% of organic substances [21,22]. The sequestration of Hg, Cu, Cd, Zn, Cr, and Ni using natural zeolites like clinoptilolite and chabazite can be enhanced through iron coating or by utilizing nanoscale zero-valent iron combined with zeolite materials. Core–shell ZnO/Y particles are effective at enhancing lead uptake and have antibacterial properties [23]. Due to the similarity in size between metallic ions and the pores, the kinetics are fast, regardless of the type of zeolite or the heavy metal being removed.
Zeolites are radiologically stable and exhibit a strong affinity for radionuclides like 90Sr and 137Cs. Natural mordenite and clinoptilolite are effective in decontaminating nuclear power plant wastewater [24]. Furthermore, all-silica zeolites are used to capture iodine released during the dissolution of nuclear fuel rods [25,26].
Recently, Kwon et al. [27] investigated the efficiency of zeolites with both different framework topologies and Si/Al ratios to carefully evaluate their sequestration capabilities for Cs+ and Sr2+ in simulated groundwater and seawater. The strong attraction of the Cs+ ion to zeolites with high Si/Al ratios can be explained through dielectric theory. The geometry and dimensions of zeolite openings are also significant in their ion-exchange capabilities. Zeolite frameworks containing eight-membered rings (8MR), such as LTA, GIS, CHA, and MOR, demonstrate exceptional selectivity for Cs+ compared to frameworks without 8MR units. This occurs because Cs+ is likely to fit within the center of the 8MR units, and its ionic size (3.6 Å) is well matched to the size of these cavities which ranges from 3.6 to 4.1 Å [28]. On the other hand, zeolites with larger and more ellipsoidal cages, like chabazite, phillipsite, stilbite, and heulandite (with SAR, ranging from 2 to 3), tend to prefer larger cations, such as Cs over Na+ or Sr2+. Conversely, lower Si/Al ratios are crucial for enhancing the attraction to strontium ions, as this allows for the formation of Al-Al pairs that can effectively interact with strontium.
Moreover, zeolites with low Si/Al ratios can release extra-framework sodium and calcium ions from their structure to sequester ammonium and potassium from soils (Figure 3). This process reduces the leaching of nutrients from the soil and enhances the recovery of insoluble phosphorus from agricultural fields, promoting a more sustainable and healthier farming environment. Zeolites exhibiting a high Si/Al ratio possess fewer sites available for exchange, a lower electric field gradient, and a reduced level of hydration [2]. In these materials, the ion-exchange capacity is strongly related to defect sites (SiO) and the framework-terminating Si–OH groups.

2.2. Zeolites for the Removal of Organic Contaminants

Zeolites are pivotal in environmental remediation efforts, contributing to outdoor air quality monitoring, as well as in the purification of water and wastewater by efficiently removing cationic species, including ammonium, heavy metals, rare-earth cations, radioactive species, and various organic pollutants [29,30,31,32,33,34,35]. It has been noted that zeolites’ adsorption is strongly affected by their composition and structural features. Factors such as pore window size, internal pore volume, and steric hindrance are crucial in determining adsorption selectivity. Medium-pore zeolites, including ZSM-5 and ferrierite, exhibit exceptional efficiency in removing volatile organic compounds (VOCs) and organic molecules of moderate size from dilute aqueous solutions [36,37]. Conversely, zeolite Y demonstrates noteworthy adsorption capabilities for larger molecules, especially drugs [37,38,39] (Figure 3).
Recent studies focused on the effectiveness of 13X and ZSM-5 adsorbent materials have been undertaken in the industrial area of Tito Scalo, located in the Basilicata Region of Southern Italy, specifically focusing on removing heavy metals and VOCs from polluted water. The results reveal that ZSM-5 demonstrated remarkable efficacy in the removal of VOCs, specifically 1,2-dichloroethylene and trichloroethylene, achieving removal efficiencies exceeding 87%. On the other hand, the 13X zeolite exhibited remarkable selectivity for the in situ abatement of heavy metals, with efficiencies reaching up to 100%. These noteworthy findings suggest that a sequential filtration system using both ZSM-5 and 13X zeolites could function as a highly effective adsorption method for the remediation of groundwater contaminants, similar to the functionality of permeable reactive barriers [29,40].
Zeolites have emerged as a focal point in research aimed at reducing emerging contaminants in aqueous environments, including pesticides, pharmaceuticals, and perfluorinated compounds (PFAS) [41]. Previous investigations have shown that some monomers of humic acids possess the ability to compete with these contaminants for adsorption sites on zeolites [42]. This competition is made more complex by the potential for complex interactions between organic contaminants and naturally dissolved organic matter which can significantly influence the adsorption process [43]. Recently, an investigation has been conducted to assess the impact of humic monomers, specifically vanillin and caffeic acid, on the adsorption of sulfamethoxazole by the high-silica zeolite Y [44,45,46]. The results indicate that the adsorption of caffeic acid was minimal and did not significantly affect the adsorption of sulfamethoxazole within a pH range of 5 to 8. However, more recent studies have revealed that caffeic acid concentrations below 20 mg L−1 can enhance the adsorption of sulfonamide antibiotics on ZSM-5 at a pH of 9 [44].
It is also crucial to consider that these substances frequently coexist with a variety of other contaminants in both natural and wastewater systems, i.e., per- and polyfluoroalkyl substances (PFAS). PFAS are emerging contaminants that are gaining attention for their ubiquitous distribution, persistence, and toxicity in the environment and ecosystems. As regulatory frameworks become increasingly stringent regarding environmental and health standards for PFAS, adsorption is emerging as a promising technique to address these challenges effectively. Among the PFAS removal approaches, the adsorption from water and wastewater through porous materials has shown great effectiveness, albeit the PFOA and PFOS adsorption mechanisms are not yet completely understood. Zeolites differing in framework topology (including FAU, LTL, BEA, MOR, CHA, and KFI) and SiO2/Al2O3 have been evaluated for several types of PFAS (i.e., PFCAs, PFSA, FTSAs, and FOSA) [45,46,47,48,49,50]. Several factors, including the perfluoroalkyl chain length, the presence of functional groups, and the molecular size, influence the adsorption capacity of individual PFAS. Other key factors affecting the affinity of zeolite for PFAS include SAR, the nature of tetrahedral cations, the occurrence of silanol groups, the zeolite hydrophobicity, and any post-synthesis modifications like ion-exchange and ammonium fluoride treatments [46,47,48,49,50,51,52].
PFASs were recently removed from different polluted waters (untreated water from drinking water facilities, wastewater discharges from processing plants, leachate from landfills, and groundwater from waste sites) in Uppsala, Sweden. The results indicate that the zeolite Beta (SAR = 25) and mordenite (SAR = 240) demonstrated absorption capacities of 99.5% and 99.2%, respectively [47]. The ability of these materials to absorb individual PFASs was affected by several factors, including the length of the perfluoroalkyl chain, the presence of functional groups, and the molecular size. Furthermore, it has been suggested that modifying the most effective zeolites with silver could help prevent biofilm formation while improving antifouling and adsorption capabilities in wastewater treatment processes. Notably, silver-functionalized zeolites demonstrated superior PFAS uptake capacities, indicating that silver functionalization may facilitate catalytic reactions in the degradation of PFAS [49]. Ag-ion-exchanged zeolites show promise as photocatalysts, forming silver clusters within their micropores that exhibit photoluminescence. The energy transfer between silver clusters is notably efficient, enhancing their photocatalytic properties. Their utility spans critical environmental processes, including de-NOx and de-SOx treatments, CO2 photoreduction, water splitting, and other reactions that contribute to renewable energy development [53].
A key advantage of these materials lies in the structure of dehydrated zeolites, which contain empty, compartmentalized voids. These spaces serve as effective sites for positioning photosensitizers or organizing multicomponent systems without the need for covalent bonds—an arrangement reminiscent of the natural photosynthetic centers found in plants and algae [54]. For example, incorporating TiO2 nanoparticles into the zeolite framework significantly enhances the degradation of organic pollutants in water. In such composites, the rigid structure of the zeolite stabilizes the TiO2 particles and facilitates the assembly of complex systems akin to natural photosynthetic units. Furthermore, the TiO2 clusters embedded within the zeolite promote interfacial electron transfer, thereby increasing efficiency in decomposing nitric oxide (NO), removing ammonia and hydrogen sulfide, and reducing CO2 in the presence of water [55].
In addition to these applications, zeolite-based photocatalysts have demonstrated effectiveness in degrading organic contaminants in wastewater and in disinfecting swimming pools and aquariums [56]. Zeolite membranes can remove salts and oils from water, offering an alternative to reverse osmosis membranes. Natural zeolite membranes, like heulandite and clinoptilolite, show high rejection rates for metal cations and toluene in synthetic seawater.
Zeolite-coated mesh films, such as those made with silicalite-1, achieve efficient gravity-driven oil–water separation thanks to their super hydrophilicity and underwater superoleophobicity [57,58]. Composite zeolite–polymeric membranes (like polysulfone, polyethersulfone, polyacrylonitrile, polyvinylidene fluoride, alcohol/agar, and cellulose acetate) result in a powerful mix of mechanical toughness, chemical stability, flexibility, adsorption ability, selectivity, and surface area [59,60]. Recent studies highlight the great potential of zeolite/activated carbon composites in water treatment technologies. These composites are highly effective at trapping heavy metals and dyes from water because of the tiny pore sizes of zeolites, while the multilayer structure of activated carbon facilitates the absorption of larger molecules [61]. Increased amounts of graphene oxide greatly improve dye decolorization. Additionally, a new hybrid magnetic composite combining zeolite and graphene oxide has successfully adsorbed 97.346 mg/g of methylene blue dye [62].
Lastly, the development of a water purification membrane made of zeolite, graphene oxide, and polyvinylidene fluoride (PVDF) has demonstrated outstanding water permeability of 28.9 L/m2/h and a remarkable desalination efficiency of 98% [63]. These cutting-edge innovations are leading toward a cleaner and more sustainable future for water treatment.

2.3. Zeolites for CO2 Capture and Storage

Zeolites with FAU-type topology (i.e., 13X, 4A, and NaY) have been investigated for their ability to capture and retain CO2 concentrations from 400 ppm to 20–30% [64]. The trapping mechanism involves both reversible physical adsorption and irreversible chemical adsorption. Typically, zeolites display moderate adsorption capacities (1–7 mmol g−1) at low pressures (1 bar) and elevated temperatures. Surface modification of zeolites through acid treatment increases performance by raising the number of acid sites on their surfaces. Moreover, the rate of CO2 adsorption can be increased through amine functionalization (i.e., involving mono-ethanolamine, diethanolamine, and triethanolamine) due to the chemical reactions among the –NH2 functional group, CO2 molecules, and the active sites of the zeolites. Amine grafting is a chemical method that increases CO2 uptake by promoting interactions between CO2 and amine groups and improving the zeolites’ surface area and porosity. The CO2 adsorption performance can be further enhanced by ion exchange with Li+, H+, NH4+, Ba2+, Mg2+, Ca2+, and Fe3+ which modifies the surface properties and structure of the zeolite. For large- and medium-pore zeolites, the spatial constraints are usually negligible because the molecular dimensions of CO2 are considerably smaller than the window apertures. Here, the adsorption capacity and selectivity are mainly influenced by the interactions between CO2 and the cations. For example, the CO2 adsorption capacity of 13X is enhanced after Li loading, as the smaller atomic size and higher basicity of lithium lead to stronger ion-quadrupole interactions in comparison to larger cations like Na+, K+, Rb+, and Cs+.
This is clearly illustrated by the “molecular trapdoor/cation gating” and “gate breathing” effects observed in high-aluminum RHO-type, CHA-type, GIS-type, and MER-type zeolites [65,66]. These phenomena can lead to significantly enhanced selectivity for CO2/N2 and CO2/CH4, highlighting the complex relationship between zeolite structure and gas adsorption behavior. For example, zeolites like EDI, FER, and ISV show about 5% electrostatic interaction, while structures like GIS, MER, MOR, and RHO reveal interactions exceeding 15% [65]. Membranes based on small-pore zeolites (with eight-membered rings, or 8MR) are being actively studied for their capabilities in separating light gases because their pore sizes are ideally suited for accommodating various light gases. Key examples of 8MR zeolites implemented in catalysis, adsorption, and membrane separation include DDR (3.6 × 4.4 Å), LTA (4.1 × 4.1 Å), CHA (3.8 × 3.8 Å), ERI (3.6 × 5.1 Å), and MER (3.4 × 5.1 Å) [65]. Natural clinoptilolite, particularly when exchanged with alkali and alkaline earth cations (following order of effectiveness: Cs+ > Rb+ > K+ > Na+ > Li+ and Ba2+ > Sr2+ > Ca2+ > Mg2+), emerges as an up-and-coming candidate for CO2 uptake due to its adsorption capacity, widespread availability, and economic viability [66,67].

2.4. Zeolites in Bioprocessing Applications

In the last years, the potential of zeolites for recovering amino acids from fermentation broths, underscoring their practical relevance in bioprocessing, has been reported [68,69,70]. Zeolites used as sorbent media for α-amino acid adsorption usually belong to medium and large pores. The affinity between the sorbents and α-amino acids strongly depends on their net charge, which can vary from positive/neutral to negative in water solutions [71,72], along with occurring host–guest interactions. The uptake of amino acids is mainly influenced by the electrostatic interactions between the positively charged ammonium group of the amino acids and the negatively charged surface of the zeolite [73]. This fundamental interaction forms the basis for the binding process. Additionally, the hydrophobic interactions arising from the non-polar side chains of the adsorbed amino acid molecules further enhance the overall binding mechanism. These ionic and hydrophobic interactions synergistically aid adsorption, providing insights into how amino acids interact with zeolite surfaces [73,74,75]. The intricate interplay between these forces is crucial for comprehending the mechanisms that control adsorption, making it a significant topic in material sciences and biochemistry. Notably, Beltrami et al. [76] demonstrated that, when L-lysine is adsorbed onto L zeolite, it adopts an α-helical conformation, which is stabilized by strong hydrogen bonds formed between the terminal amino groups of lysine, co-adsorbed water, and the oxygen atoms in the zeolite framework.
Furthermore, investigations by Chen et al. [77,78] on achiral MFI zeolites have shown that these materials exhibit differential adsorption properties for the L- and D-Lys enantiomers in acidic solutions (pH < 2). This finding suggests that achiral zeolites may be applicable to chiral separations, thereby expanding their applications in biochemical separation processes. Furthermore, zeolites are increasingly recognized for their significance in the medical and biotechnological fields. From a medical perspective, zeolites enhance livestock’s nutritional status and immune response while serving as detoxifying agents for humans and animals [77,78,79,80]. This multifaceted utility underscores the vital role that zeolites play in advancing sustainable technologies, i.e., in detecting biomarkers, controlled gene delivery, tissue engineering, biomaterial coating, and separation of cells and biomolecules [81,82,83]. Their distinctive porous architecture and remarkable ion-exchange capabilities are efficiently utilized in contemporary drug delivery systems, wound healing applications, antibacterial coatings, contrast media, and detoxification methods, such as hemodialysis and the elimination of toxic ions [79]. A significant benefit of zeolites is their capacity to immobilize microorganism, thus acting as a support medium in anaerobic fluidized bed reactor (AFBR). This process represents a sustainable methodology for the conversion of complex organic substrates into methane and other biofuels, harnessing the synergistic actions of different microbial communities [84,85]. In this context, zeolites play a crucial role due to their unique properties, such as ion exchange, adsorption capabilities, and enhanced potential for microbial bioattachment, all of which significantly optimize anaerobic digestion efficiency [86]. For instance, the presence of metal ions such as Na+ and K+ within zeolites has been shown to positively influence microbial enrichment, resulting in improved performance in anaerobic digestion [86]. Additionally, modified zeolites have been utilized as a substrate for biofilm development, further advancing hydrogen production and other valuable byproducts. Fe-modified zeolites act as substitutes for traditional zeolites, enhancing the acetate fermentation pathway and boosting hydrogen production. This advancement has led to increased productivity over a continuous period of 72 days within our hybrid-Fe bioreactor. The synergistic effects of the Fe-modified zeolite supply essential micronutrients, aid in the synthesis of enzymes and co-enzymes, activate favorable metabolic pathways, promote microbial immobilization, and facilitate electron transfer among interconnected hydrogen-producing bacteria. Consequently, the hybrid-Fe bioreactor demonstrates significant potential as an effective and sustainable approach to biohydrogen production for practical applications [86].

3. Toward the Development of a Sustainable Zeolite-Based Roadmap for Environmental Remediation

To fully exploit the effectiveness of zeolites for extensive, long-lasting environmental cleanup, it is crucial to develop an all-encompassing plan that incorporates zeolite solutions into wider environmental management strategies. In environmental technologies, roadmapping should combine fundamental research with industrial applications, evaluating the progress of zeolite-based materials research. This method underscores new prospects in fields like carbon capture, water treatment, contaminant elimination, catalysis, and energy storage. It also outlines the technological advancements necessary to enhance the efficiency and applicability of zeolite-based solutions.
In particular, an exhaustive zeolite technology roadmap should integrate the following core components:
  • Advancements in technology and material developments: this includes monitoring the progress of zeolite structures, surface functionalization, and composite materials to improve efficiency in applications such as adsorption, ion exchange, and catalysis.
  • Market and industry alignment: it is essential to assess how zeolite-based materials meet the increasing global demand for sustainable environmental technologies, green chemistry, and renewable energy solutions.
  • Competitive and regulatory landscape: understanding the alignment of zeolite technologies with existing and emerging environmental regulations, sustainability policies, and industrial standards
  • R&D priorities and implementation challenges: identifying key innovation areas such as improving zeolite stability, scalability, and selectivity while addressing synthesis costs, resource availability, and large-scale applicability.
Figure 4 shows the roadmap outlining the key steps for leveraging zeolite-based technologies in addressing emerging contaminants, with a focus on research, scalable solutions, and global policy integration. By adopting a collaborative approach and continuous innovation, zeolite solutions will play a vital role in protecting ecosystems and human health, contributing to a sustainable future.

4. Conclusions

Zeolites are crystalline microporous materials extensively studied because of their distinctive features, including shape selectivity, adsorption, ion-exchange ability, and low manufacturing costs. Due to their extensive surface area and stability in mechanical, biological, chemical, and thermal processes, these molecular sieves are utilized not just in conventional processes (petrochemical, coal chemical industries, separation, adsorption), but also as catalysts for the conversion of greenhouse gases (CO2 and CH4) and biomass. Multifunctional zeolites, which combine Brønsted and Lewis acid sites, improve the efficacy of multistep reactions, thus offering new approaches for generating high-value chemicals from biomass. Moreover, zeolites are utilized to produce hydrogen and methanol for fuel cells, providing outstanding membrane characteristics and selectivity, deserving further investigation. Zeolites play a crucial role in generating hydrogen and methanol for fuel cells, functioning as membranes that inhibit methanol crossover while providing excellent selectivity and remarkable mechanical and thermal stability, justifying further investigation into this technology.
Zeolites can trap CO2 due to their powerful electric fields and transform it into fuels and chemicals, demonstrating excellent selectivity and stability.
Zeolite–water adsorption systems can store and release energy, making them useful for solar energy applications and industrial waste heat storage. By coating zeolites on a metallic bed, heat and mass transfer efficiency is enhanced, and the adsorption system is made more compact and lasts longer. Zeolite composites can be used efficiently to eliminate heavy metals and dyes in water purification, providing a cost-effective solution enhancing soil fertility and agricultural practices through their cation exchange properties.
Furthermore, zeolites decrease air pollutants by reducing NOx and NH3 emissions and turning volatile organic compounds into harmless byproducts, although there are still issues with emissions treatment.
Despite significant advancements, the application of zeolites in sustainable processes can still be further enhanced. Primary objectives include enhancing catalyst selectivity, reducing expenses, and evaluating the economic viability of novel zeolites. With nearly a century of substantial history, the future for zeolitic research and its applications appears promising.

Author Contributions

Conceptualization, M.M. and A.M.; validation, M.M. and A.M.; investigation, M.M. and A.M.; resources, A.M.; data curation, M.M. and A.M.; writing—original draft preparation, M.M. and A.M.; writing—review and editing, M.M. and A.M.; visualization, M.M. and A.M.; supervision, A.M.; funding acquisition, A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

The authors acknowledge the National Recovery and Resilience Plan (NRRP), Mission 04 Component 2 Investment 1.5—NextGenerationEU, call for tender n. 3277 dated 30 December 2021, Award Number: 0001052 dated 23 June 2022.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Li, J.; Gao, Z.R.; Lin, Q.F.; Liu, C.; Gao, F.; Lin, C.; Zhang, S.; Deng, H.; Mayoral, A.; Fan, W.; et al. A 3D extra-large-pore zeolite enabled by 1D-to-3D topotactic condensation of a chain silicate. Science 2023, 379, 283–287. [Google Scholar] [CrossRef] [PubMed]
  2. Casci, J.L. Zeolite molecular sieves: Preparation and scale-up. Microporous Mesoporous Mater. 2005, 82, 217–226. [Google Scholar] [CrossRef]
  3. Corma, A. State of the art and future challenges of zeolites as catalysts. J. Catal. 2003, 216, 298–312. [Google Scholar] [CrossRef]
  4. Serrano, D.P.; Centi, G.; Diddams, P.A.; Čejka, J. Outlooks for zeolite catalysts in a low-carbon scenario. Catal. Today 2024, 426, 114365. [Google Scholar] [CrossRef]
  5. Van Donk, S.; Janssen, A.H.; Bitter, J.H.; de Jong, K.P. Generation, characterization, and impact of mesopores in zeolite catalysts. Catal. Rev. 2003, 45, 297–319. [Google Scholar] [CrossRef]
  6. Mancinelli, M.; Precisvalle, N.; Ardit, M.; Beltrami, G.; Gigli, L.; Catizzone, E.; Migliori, M.; Giordano, G.; Martucci, A. Thermal stability of templated ZSM-5 zeolites: An in-situ synchrotron X-ray powder diffraction study. Microporous Mesoporous Mater. 2023, 362, 112777. [Google Scholar] [CrossRef]
  7. Čejka, J.; Millini, R.; Opanasenko, M.; Serrano, D.P.; Roth, W.J. Advances and challenges in zeolite synthesis and catalysis. Catal. Today 2020, 345, 2–13. [Google Scholar] [CrossRef]
  8. Velty, A.; Corma, A. Advanced zeolite and ordered mesoporous silica-based catalysts for the conversion of CO2 to chemicals and fuels. Chem. Soc. Rev. 2023, 52, 1773–1946. [Google Scholar] [CrossRef]
  9. Ducamp, M.; Coudert, F.X. Systematic study of the thermal properties of zeolitic frameworks. J. Phys. Chem. C 2021, 125, 15647–15658. [Google Scholar] [CrossRef]
  10. Krivovichev, S. Topology of microporous structures. Rev. Mineral. Geochem. 2005, 57, 17–68. [Google Scholar] [CrossRef]
  11. Cruciani, G. Zeolites upon heating: Factors governing their thermal stability and structural changes. J. Phys. Chem. Solids 2006, 67, 1973–1994. [Google Scholar] [CrossRef]
  12. Bish, D.L.; Carey, J.W. Thermal behavior of natural zeolites. Rev. Mineral. Geochem. 2001, 45, 403–452. [Google Scholar] [CrossRef]
  13. Confalonieri, G.; Grand, J.; Arletti, R.; Barrier, N.; Mintova, S. CO2 adsorption in nanosized RHO zeolites with different chemical compositions and crystallite sizes. Microporous Mesoporous Mater. 2020, 306, 110394. [Google Scholar] [CrossRef]
  14. Rahmah, W.; Kadja, G.T.M.; Mahyuddin, M.H.; Saputro, A.G.; Dipojono, H.K.; Wenten, I.G. Small-pore zeolite and zeotype membranes for CO2 capture and sequestration—A review. J. Environ. Chem. Eng. 2022, 10, 108707. [Google Scholar] [CrossRef]
  15. Colella, C. Natural zeolites in environmentally friendly processes and applications. Stud. Surf. Sci. Catal. 1999, 125, 641–655. [Google Scholar] [CrossRef]
  16. Boronat, M.; Corma, A. Factors controlling the acidity of zeolites. Catal. Lett. 2015, 145, 162–172. [Google Scholar] [CrossRef]
  17. Čejka, J.; Wichterlová, B. Acid-catalyzed synthesis of mono-and dialkyl benzenes over zeolites: Active sites, zeolite topology, and reaction mechanisms. Catal. Rev. 2002, 44, 375–421. [Google Scholar] [CrossRef]
  18. Pérez-Botella, E.; Valencia, S.; Rey, F. Zeolites in adsorption processes: State of the art and future prospects. Chem. Rev. 2022, 122, 17647–17695. [Google Scholar] [CrossRef]
  19. Jin, W.; Yang, Y.; Jin, J.; Xu, M.; Zhang, Z.; Dong, F.; Shao, M.; Wan, Y. Characterization of phosphate modified red mud-based composite materials and study on heavy metal adsorption. Environ. Sci. Pollut. Res. 2024, 31, 43687–43703. [Google Scholar] [CrossRef]
  20. Daer, S.; Kharraz, J.; Giwa, A.; Hasan, S.W. Recent applications of nanomaterials in water desalination: A critical review and future opportunities. Desalination 2015, 367, 37–48. [Google Scholar] [CrossRef]
  21. Swenson, P.; Tanchuk, B.; Bastida, E.; An, W.; Kuznicki, S.M. Water desalination and de-oiling with natural zeolite membranes—Potential application for purification of SAGD process water. Desalination 2012, 286, 442–446. [Google Scholar] [CrossRef]
  22. Li, Y.; Li, L.; Yu, J. Applications of zeolites in sustainable chemistry. Chem 2017, 3, 928–949. [Google Scholar] [CrossRef]
  23. Jiménez-Reyes, M.; Almazán-Sánchez, P.T.; Solache-Ríos, M. Radioactive waste treatments by using zeolites. A short review. J. Environ. Radioact. 2021, 233, 106610. [Google Scholar] [CrossRef]
  24. Riley, B.J.; Chong, S.; Schmid, J.; Marcial, J.; Nienhuis; Bera, M.K.; Lee, S.; Canfield, N.L.; Kim, S.; Derewinski, M.A.; et al. Role of zeolite structural properties toward iodine capture: A head-to-head evaluation of framework type and chemical composition. ACS Appl. Mater. Interfaces 2022, 14, 18439–18452. [Google Scholar] [CrossRef]
  25. Pham, T.C.T.; Docao, S.; Hwang, I.C.; Song, M.K.; Moon, D.; Oleynikov, P.; Yoon, K.B. Capture of iodine and organic iodides using silica zeolites and the semiconductor behavior of iodine in a silica zeolite. Energy Environ. Sci. 2016, 9, 1050–1062. [Google Scholar] [CrossRef]
  26. Kwon, S.; Kim, C.; Han, E.; Lee, H.; Cho, H.S.; Choi, M. Relationship between zeolite structure and capture capability for radioactive cesium and strontium. J. Hazard. Mater. 2021, 408, 124419. [Google Scholar] [CrossRef]
  27. Hijikata, T.; Koyama, T.; Aikyo, Y.; Shimura, S.; Kawanishi, M. Strontium adsorption characteristics of natural Zeolites for permeable reactive barrier in Fukushima Daiichi nuclear power station. J. Nucl. Sci. Technol. 2021, 58, 1079–1098. [Google Scholar] [CrossRef]
  28. Belviso, C. Special Issue “Sustainable Remediation Processes Based on Zeolites”. Processes 2021, 9, 2153. [Google Scholar] [CrossRef]
  29. de Gennaro, B.; Aprea, P.; Liguori, B.; Galzerano, B.; Peluso, A.; Caputo, D. Zeolite-rich composite materials for environmental remediation: Arsenic removal from water. Appl. Sci. 2020, 10, 6938. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Cao, B.; Yin, H.; Meng, L.; Jin, W.; Wang, F.; Xu, J.; Al-Tabbaa, A. Application of zeolites in permeable reactive barriers (PRBs) for in-situ groundwater remediation: A critical review. Chemosphere 2022, 308, 136290. [Google Scholar] [CrossRef]
  31. Eberle, S.; Börnick, H.; Stolte, S. Granular natural zeolites: Cost-effective adsorbents for the removal of ammonium from drinking water. Water 2022, 14, 939. [Google Scholar] [CrossRef]
  32. Mancinelli, M.; Arfè, A.; Martucci, A.; Pasti, L.; Chenet, T.; Sarti, E.; Vergine, G.; Belviso, C. Evaluation for the removal efficiency of VOCs and heavy metals by zeolites-based materials in the wastewater: A case study in the Tito Scalo industrial area. Processes 2020, 8, 1519. [Google Scholar] [CrossRef]
  33. Wilfong, W.C.; Ji, T.; Duan, Y.; Shi, F.; Wang, Q.; Gray, M.L. Critical review of functionalized silica sorbent strategies for selective extraction of rare earth elements from acid mine drainage. J. Hazard. Mater. 2022, 424, 127625. [Google Scholar] [CrossRef]
  34. Mijailović, N.R.; Nedić Vasiljević, B.; Ranković, M.; Milanović, V.; Uskoković-Marković, S. Environmental and pharmacokinetic aspects of zeolite/pharmaceuticals systems. Two facets of adsorption ability. Catalysts 2022, 12, 837. [Google Scholar] [CrossRef]
  35. Mangla, D.; Sharma, A.; Ikram, S. Critical review on adsorptive removal of antibiotics: Present situation, challenges and future perspective. J. Hazard. Mater. 2022, 425, 127946. [Google Scholar] [CrossRef]
  36. Song, J.Y.; Jhung, S.H. Adsorption of pharmaceuticals and personal care products over metal-organic frameworks functionalized with hydroxyl groups: Quantitative analyses of H-bonding in adsorption. J. Chem. Eng. 2017, 322, 366–374. [Google Scholar] [CrossRef]
  37. Sarti, E.; Chenet, T.; Stevanin, C.; Costa, V.; Cavazzini, A.; Catani, M.; Martucci, A.; Precisvalle, N.; Beltrami, G.; Pasti, L. High-silica zeolites as sorbent media for adsorption and pre-concentration of pharmaceuticals in aqueous solutions. Molecules 2020, 25, 3331. [Google Scholar] [CrossRef]
  38. Liu, Y.; Liu, X.; Lu, S.; Zhao, B.; Wang, Z.; Xi, B.; Guo, W. Adsorption and biodegradation of sulfamethoxazole and ofloxacin on zeolite: Influence of particle diameter and redox potential. Chem. Eng. J. 2020, 384, 123346. [Google Scholar] [CrossRef]
  39. Senila, M.; Cadar, O. Modification of natural zeolites and their applications for heavy metal removal from polluted environments: Challenges, recent advances, and perspectives. Heliyon 2024, 10, e25303. [Google Scholar] [CrossRef]
  40. Chenet, T.; Mancinelli, M.; Sarti, E.; Costa, V.; D’Anna, C.; Martucci, A.; Pasti, L. Competitive Adsorption of 4-Hydroxybenzaldehyde and Toluene onto High Silica Zeolites. Environ. Process. 2024, 11, 49. [Google Scholar] [CrossRef]
  41. Rocha, L.C.C.; Zuquette, L.V. Evaluation of zeolite as a potential reactive medium in a permeable reactive barrier (PRB): Batch and column studies. Geosciences 2020, 10, 59. [Google Scholar] [CrossRef]
  42. de Sousa, D.N.R.; Insa, S.; Mozeto, A.A.; Petrovic, M.; Chaves, T.F.; Fadini, P.S. Equilibrium and kinetic studies of the adsorption of antibiotics from aqueous solutions onto powdered zeolites. Chemosphere 2018, 205, 137–146. [Google Scholar] [CrossRef] [PubMed]
  43. Zuo, X.; Qian, C.; Ma, S.; Xiong, J. Sulfonamide antibiotics sorption by high silica ZSM-5: Effect of pH and humic monomers (vanillin and caffeic acid). Chemosphere 2020, 248, 126061. [Google Scholar] [CrossRef]
  44. Bal’zhinimaev, B.S.; Paukshtis, E.A.; Toktarev, A.V.; Kovalyov, E.V.; Yaranova, M.A.; Smirnov, A.E.; Stompel, S. Effect of water on toluene adsorption over high silica zeolites. Microporous Mesoporous Mater. 2019, 277, 70–77. [Google Scholar] [CrossRef]
  45. Mancinelli, M.; Ardit, M.; Chenet, T.; Pasti, L.; Martucci, A. Desorption of humic monomers from Y zeolite: A high-temperature X-ray diffraction and thermogravimetric study. Microporous Mesoporous Mater. 2024, 379, 113270. [Google Scholar] [CrossRef]
  46. Mancinelli, M.; Martucci, A.; Ahrens, L. Exploring the adsorption of short and long-chain per-and poly-fluoroalkyl substances (PFAS) to different zeolites using environmental samples. Environ. Sci. Water Res. Technol. 2023, 9, 2595–2604. [Google Scholar] [CrossRef]
  47. Ponge, C.A.; Sheehan, N.P.; Corbin, D.R.; Peltier, E.; Hutchison, J.M.; Shiflett, M.B. Zeolites for Sorption of PFAS from Water. Ind. Eng. Chem. Res. 2024, 63, 12102–12112. [Google Scholar] [CrossRef]
  48. Mancinelli, M.; Stevanin, C.; Ardit, M.; Chenet, T.; Pasti, L.; Martucci, A. PFAS as emerging pollutants in the environment: A challenge with FAU type and silver-FAU exchanged zeolites for their removal from water. J. Environ. Chem. Eng. 2022, 10, 108026. [Google Scholar] [CrossRef]
  49. Ponge, C.A.; Corbin, D.R.; Sabolay, C.M.; Shiflett, M.B. Designing zeolites for the removal of aqueous PFAS: A perspective. Ind. Chem. Mater. 2024, 2, 270–275. [Google Scholar] [CrossRef]
  50. Mancinelli, M.; Martucci, A.; Salani, G.M.; Bianchini, G.; Gigli, L.; Plaisier, J.R.; Colombo, F. High-temperature behavior of Ag-exchanged Y zeolites used for PFAS sequestration from water. Phys. Chem. Chem. Phys. 2023, 25, 20066–20075. [Google Scholar] [CrossRef]
  51. Omo-Okoro, P.N.; Curtis, C.J.; Marco, A.M.; Melymuk, L.; Okonkwo, J.O. Removal of per- and polyfluoroalkyl substances from aqueous media using synthesized silver nanocomposite-activated carbons. J. Environ. Health Sci. Eng. 2021, 19, 217–236. [Google Scholar] [CrossRef] [PubMed]
  52. Corma, A.; Garcia, H. Zeolite-based photocatalysts. Chem. Commun. 2004, 13, 1443–1459. [Google Scholar] [CrossRef] [PubMed]
  53. Hamd, A.; Shaban, M.; Al-Senani, G.M.; Alshabanat, M.N.; Al-Ghamdi, A.; Dryaz, A.R.; Ahmed, S.A.; El-Sayed, R.; Soliman, N.K. Comprehensive evaluation of zeolite/marine alga nanocomposite in the removal of waste dye from industrial wastewater. Sci. Rep. 2023, 13, 8082. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, H.; Liu, H.; Chen, Z.; Veksha, A.; Lisak, G.; You, C. Interaction between SO2 and NO in their adsorption and photocatalytic conversion on TiO2. Chemosphere 2020, 249, 126136. [Google Scholar] [CrossRef]
  55. Izadpanah, M.; Sarrafzadeh, M.H.; Rezaei-Dashtarzhandi, M.; Vojoudi, H. Aquatic center sewage reclamation and water reuse, using an integrated system combining adsorption, RO membrane system, and TiO2/Fe3O4 photocatalytic oxidation. J. Environ. Chem. Eng. 2021, 9, 104957. [Google Scholar] [CrossRef]
  56. Ghasemi, Z.; Sourinejad, I.; Kazemian, H.; Rohani, S. Application of zeolites in aquaculture industry: A review. Rev. Aquac. 2018, 10, 75–95. [Google Scholar] [CrossRef]
  57. Zhao, Q.; Liao, C.; Chen, G.; Liu, R.; Wang, Z.; Xu, A.; Ji, S.; Shih, K.; Zou, L.; Duan, T. In situ confined synthesis of a copper-encapsulated silicalite-1 zeolite for highly efficient iodine capture. Inorg. Chem. 2022, 61, 20133–20143. [Google Scholar] [CrossRef]
  58. Mambetova, M.; Dossumov, K.; Baikhamurova, M.; Yergaziyeva, G. Sorbents Based on Natural Zeolites for Carbon Dioxide Capture and Removal of Heavy Metals from Wastewater: Current Progress and Future Opportunities. Processes 2024, 12, 2071. [Google Scholar] [CrossRef]
  59. Velarde, L.; Nabavi, M.S.; Escalera, E.; Antti, M.L.; Akhtar, F. Adsorption of heavy metals on natural zeolites: A review. Chemosphere 2023, 328, 138508. [Google Scholar] [CrossRef]
  60. Tufail, M.K.; Ifrahim, M.; Rashid, M.; Haq, I.U.; Asghar, R.; Uthappa, U.T.; Selvaraj, M.; Kurkuri, M. Chemistry of zeolites and zeolite based composite membranes as a cutting-edge candidate for removal of organic dyes & heavy metal ions: Progress and future directions. Sep. Purif. Technol. 2024, 354, 128739. [Google Scholar] [CrossRef]
  61. Li, H.; Zheng, F.; Wang, J.; Zhou, J.; Huang, X.; Chen, L.; Bashir, S.; Liu, J.L. Facile preparation of zeolite-activated carbon composite from coal gangue with enhanced adsorption performance. Chem. Eng. J. 2020, 390, 124513. [Google Scholar] [CrossRef]
  62. Mahmoodi, N.M.; Saffar-Dastgerdi, M.H. Clean Laccase immobilized nanobiocatalysts (graphene oxide-zeolite nanocomposites): From production to detailed biocatalytic degradation of organic pollutant. Appl. Catal. B Environ. 2020, 268, 118443. [Google Scholar] [CrossRef]
  63. Hassan, F.; Mushtaq, R.; Saghar, S.; Younas, U.; Pervaiz, M.; Muteb Aljuwayid, A.; Habila, M.A.; Sillanpaa, M. Fabrication of graphene-oxide and zeolite loaded polyvinylidene fluoride reverse osmosis membrane for saltwater remediation. Chemosphere 2022, 307, 136012. [Google Scholar] [CrossRef] [PubMed]
  64. Shi, X.; Xiao, H.; Azarabadi, H.; Song, J.; Wu, X.; Chen, X.; Lackner, K.S. Sorbents for the direct capture of CO2 from ambient air. Angew. Chem. Int. Ed. 2020, 59, 6984–7006. [Google Scholar] [CrossRef]
  65. Fu, D.; Davis, M.E. Carbon dioxide capture with zeotype materials. Chem. Soc. Rev. 2022, 51, 9340–9370. [Google Scholar] [CrossRef]
  66. Zhang, Q.; Gao, S.; Yu, J. Metal sites in zeolites: Synthesis, characterization, and catalysis. Chem. Rev. 2022, 123, 6039–6106. [Google Scholar] [CrossRef]
  67. Zhang, H.; bin Samsudin, I.; Jaenicke, S.; Chuah, G.K. Zeolites in catalysis: Sustainable synthesis and its impact on properties and applications. Catal. Sci. Technol. 2022, 12, 6024–6039. [Google Scholar] [CrossRef]
  68. Sun, Q.; Wang, N.; Yu, J. Advances in catalytic applications of zeolite-supported metal catalysts. Adv. Mater. 2021, 33, 2104442. [Google Scholar] [CrossRef]
  69. Dusselier, M.; Van Wouwe, P.; Dewaele, A.; Jacobs, P.A.; Sels, B.F. Shape-selective zeolite catalysis for bioplastics production. Science 2015, 349, 78–80. [Google Scholar] [CrossRef]
  70. Faisal, A.; Holmlund, M.; Ginesy, M.; Holmgren, A.; Enman, J.; Hedlund, J.; Grahn, M. Recovery of L-arginine from model solutions and fermentation broth using zeolite-Y adsorbent. ACS Sustain. Chem. Eng. 2019, 7, 8900–8907. [Google Scholar] [CrossRef]
  71. Jatoi, A.S.; Baloch, H.A.; Mazari, S.A.; Mubarak, N.M.; Sabzoi, N.; Aziz, S.; Soomro, S.A.; Abro, R.; Shah, S.F. A review of extractive fermentation via ion exchange adsorption resins opportunities, challenges, and prospects. Biomass Convers. Biorefin. 2021, 13, 3543–3554. [Google Scholar] [CrossRef]
  72. Wang, Q.; Li, H.; Feng, K.; Liu, J. Oriented fermentation of food waste towards high-value products: A review. Energies 2020, 13, 5638. [Google Scholar] [CrossRef]
  73. Munsch, S.; Hartmann, M.; Ernst, S. Adsorption and separation of amino acids from aqueous solutions on zeolites. Chem. Comm. 2021, 19, 1978–1979. [Google Scholar] [CrossRef] [PubMed]
  74. Polisi, M.; Fabbiani, M.; Vezzalini, G.; Di Renzo, F.; Pastero, L.; Quartieri, S.; Arletti, R. Amino acid encapsulation in zeolite MOR: Effect of spatial confinement. Phys. Chem. Chem. Phys. 2021, 23, 20541–20552. [Google Scholar] [CrossRef]
  75. Beltrami, G.; Martucci, A.; Pasti, L.; Chenet, T.; Ardit, M.; Gigli, L.; Cescon, M.; Suard, E. L-Lysine Amino Acid Adsorption on Zeolite L: A Combined Synchrotron, X-Ray and Neutron Diffraction Study. ChemistryOpen 2020, 9, 978–982. [Google Scholar] [CrossRef]
  76. Chen, T.; Wun, C.K.T.; Day, S.J.; Tang, C.C.; Lo, T.W.B. Enantiospecificity in achiral zeolites for asymmetric catalysis. Phys. Chem. Chem. Phys. 2020, 22, 18757–18764. [Google Scholar] [CrossRef]
  77. Chen, T.; Huang, B.; Day, S.; Tang, C.C.; Tsang, S.C.E.; Wong, K.; Lo, T.W.B. Differential Adsorption of L-and D-Lysine on Achiral MFI Zeolites as Determined by Synchrotron X-Ray Powder Diffraction and Thermogravimetric Analysis. Angew. Chem. Int. Ed. 2020, 59, 1093–1097. [Google Scholar] [CrossRef]
  78. Serati-Nouri, H.; Jafari, A.; Roshangar, L.; Dadashpour, M.; Pilehvar-Soltanahmadi, Y.; Zarghami, N. Biomedical applications of zeolite-based materials: A review. Mater. Sci. Eng. C 2020, 116, 111225. [Google Scholar] [CrossRef]
  79. Elsherbeni, A.I.; Youssef, I.M.; Kamal, M.; Youssef, M.A.; El-Gendi, G.M.; El-Garhi, O.H.; Alfassam, H.E.; Rudayni, H.A.; Allam, A.A.; Moustafa, M.; et al. Impact of Adding Zeolite to broilers’ Diet and Litter on Growth, Blood Parameters, Immunity, and Ammonia Emission. Poult. Sci. 2024, 103, 103981. [Google Scholar] [CrossRef]
  80. Papaioannou, D.; Katsoulos, P.D.; Panousis, N.; Karatzias, H. The role of natural and synthetic zeolites as feed additives on the prevention and/or the treatment of certain farm animal diseases: A review. Microporous Mesoporous Mater. 2005, 84, 161–170. [Google Scholar] [CrossRef]
  81. Kucherenko, I.S.; Soldatkin, O.O.; Dzyadevych, S.V.; Soldatkin, A.P. Application of zeolites and zeolitic imidazolate frameworks in biosensor development. Biomater. Adv. 2022, 143, 213180. [Google Scholar] [CrossRef] [PubMed]
  82. Bacakova, L.; Vandrovcova, M.; Kopova, I.; Jirka, I. Applications of zeolites in biotechnology and medicine—A review. Biomater. Sci. 2018, 6, 974–989. [Google Scholar] [CrossRef] [PubMed]
  83. He, Z.W.; Zou, Z.S.; Ren, Y.X.; Tang, C.C.; Zhou, A.J.; Liu, W.; Wang, L.; Li, Z.; Wang, A. Roles of zero-valent iron in anaerobic digestion: Mechanisms, advances and perspectives. Sci. Total Environ. 2022, 852, 158420. [Google Scholar] [CrossRef]
  84. Jin, H.Y.; He, Z.W.; Ren, Y.X.; Tang, C.C.; Zhou, A.J.; Liu, W.; Liang, B.; Li, Z.H.; Wang, A. Current advances and challenges for direct interspecies electron transfer in anaerobic digestion of waste activated sludge. Chem. Eng. J. 2022, 450, 137973. [Google Scholar] [CrossRef]
  85. Montalvo, S.; Huiliñir, C.; Borja, R.; Sánchez, E.; Herrmann, C.J.B.T. Application of zeolites for biological treatment processes of solid wastes and wastewaters—A review. Bioresour. Technol. 2020, 301, 122808. [Google Scholar] [CrossRef]
  86. Zhao, C.; Sharma, A.; Ma, Q.; Zhu, Y.; Li, D.; Liu, Z.; Yang, Y. A developed hy-brid fixed-bed bioreactor with Fe-modified zeolite to enhance and sustain biohydrogen production. Sci. Total Env. 2021, 758, 143658. [Google Scholar] [CrossRef]
Figure 1. Zeolites with mono-, bi-, and tri-directional channels. In yellow, the atoms in tetrahedral coordination (Si, Al), and in red, the oxygens of the framework.
Figure 1. Zeolites with mono-, bi-, and tri-directional channels. In yellow, the atoms in tetrahedral coordination (Si, Al), and in red, the oxygens of the framework.
Suschem 06 00009 g001
Figure 2. Main zeolite properties.
Figure 2. Main zeolite properties.
Suschem 06 00009 g002
Figure 3. Cationic exchange in zeolites. In yellow, the atoms in tetrahedral coordination (Si, Al), and in red, the oxygens of the framework. The blue arrows represent species incorporated into the extra framework cavities, while the green arrows denote species expelled from the framework.
Figure 3. Cationic exchange in zeolites. In yellow, the atoms in tetrahedral coordination (Si, Al), and in red, the oxygens of the framework. The blue arrows represent species incorporated into the extra framework cavities, while the green arrows denote species expelled from the framework.
Suschem 06 00009 g003
Figure 4. Key steps in zeolite-based technologies, with a focus on research, scalable solutions, and global policy integration.
Figure 4. Key steps in zeolite-based technologies, with a focus on research, scalable solutions, and global policy integration.
Suschem 06 00009 g004
Table 1. Channel systems are categorized by decreasing tetrahedral atoms in the largest rings (bold). They include channel direction, bold T-atom counts in the ring controlling diffusion, and crystallographic channel diameters in Angstroms. Asterisks denote whether the system is one, two, or three-dimensional.
Table 1. Channel systems are categorized by decreasing tetrahedral atoms in the largest rings (bold). They include channel direction, bold T-atom counts in the ring controlling diffusion, and crystallographic channel diameters in Angstroms. Asterisks denote whether the system is one, two, or three-dimensional.
Extra-large pores
CLOCloverite<100> 20 4.0 × 13.2 *** | <100> 8 3.8 × 3.8 ***
VFIVPI-5[001] 18 12.7 × 12.7 *
AETAlPO-8[001] 14 7.9 × 8.7 *
CFICIT-5[010] 14 7.2 × 7.5 *
DONUTD-1F[010] 14 8.1 × 8.2 *
OSOOSB-1[001] 14 5.4 × 7.3 * Ʇ [001] 8 2.8 × 3.3 **
Large pores
AFIAlPO-5[001] 12 7.3 × 7.3 *
*BEABeta<100> 12 6.6 × 6.7 ** [001] 12 5.6 × 5.6 *
BOGBoggsite[100] 12 7.0 × 7.0 * [010] 10 5.5 × 5.8 *
CANCancrinite[001] 12 5.9 × 5.9 *
FAUFaujasite<111> 12 7.4 × 7.4 ***
GMEGmelinite[001] 12 7.0 × 7.0 * Ʇ [001] 8 3.6 × 3.9 **
ISVITQ-7<100> 12 6.1 × 6.5 ** [001] 12 5.9 × 6.6 *
LTLLinde Type L[001] 12 7.1 × 7.1 *
MAZMazzite[001] 12 7.4 × 7.4 * | [001] 8 3.1 × 3.1 ***
MORMordenite[001] 12 6.5 × 7.0 * {[010] 8 3.4 × 4.8 * [001] 8 2.6 × 5.7} **
MTWZSM-12[010] 12 5.6 × 6.0 *
OFFOffretite[001] 12 6.7 × 6.8 * Ʇ [001] 8 3.6 × 4.9 **
Medium pores
DACDachiardite[010] 10 3.4 × 5.3 * [001] 8 3.7 × 4.8 *
EPIEpistilbite[100] 10 3.4 × 5.6 * [001] 8 3.7 × 4.5 *
FERFerrierite[001] 10 4.2 × 5.4 * [010] 8 3.5 × 4.8 *
HEUHeulandite{[001] 10 3.1 × 7.5 * + 8 3.6 × 4.6 *} × [100] 8 2.8 × 4.7 *
MFIZSM-5{[100] 10 5.1 × 5.5 [010] 10 5.3 × 5.6} ***
STIStilbite[100] 10 4.7 × 5.0 * [001] 8 2.7 × 5.6 *
TERTerranovaite[100] 10 5.0 × 5.0 * [001] 10 4.1 × 7.0 *
NATNatrolite<100> 8 2.6 × 3.9 ** [001] 9 2.5 × 4.1 *
Small pores
ABWLi-A[001] 8 3.4 × 3.8 *
BREBikitaite[010] 8 2.8 × 3.7 *
CHAChabaziteꞱ [001] 8 3.8 × 3.8 ***
LTALinde Type A<100> 8 4.1 × 4.1 ***
YUGYugawaralite[100] 8 2.8 × 3.6 * [001] 8 3.1 × 5.0 *
VNIVPI-9{<110> 8 3.1 × 4.0 [001] 8 3.5 × 3.6} ***
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

Mancinelli, M.; Martucci, A. Exploring the Potential of Zeolites for Sustainable Environmental Applications. Sustain. Chem. 2025, 6, 9. https://doi.org/10.3390/suschem6010009

AMA Style

Mancinelli M, Martucci A. Exploring the Potential of Zeolites for Sustainable Environmental Applications. Sustainable Chemistry. 2025; 6(1):9. https://doi.org/10.3390/suschem6010009

Chicago/Turabian Style

Mancinelli, Maura, and Annalisa Martucci. 2025. "Exploring the Potential of Zeolites for Sustainable Environmental Applications" Sustainable Chemistry 6, no. 1: 9. https://doi.org/10.3390/suschem6010009

APA Style

Mancinelli, M., & Martucci, A. (2025). Exploring the Potential of Zeolites for Sustainable Environmental Applications. Sustainable Chemistry, 6(1), 9. https://doi.org/10.3390/suschem6010009

Article Metrics

Back to TopTop