Nanosheet zeolites: controlled synthesis, characterization, and advanced catalysis applications

Yankai Bian, Xiaoyang Zhang, Jun Yu* and Weili Dai*
School of Materials Science and Engineering, Nankai University, Tianjin 300350, China. E-mail: yujun@nankai.edu.cn; weilidai@nankai.edu.cn

Received 31st March 2025 , Accepted 5th June 2025

First published on 6th June 2025


Abstract

Zeolites are extensively utilized in over 40% of acid-catalyzed reactions owing to their well-ordered microporous structures, excellent hydrothermal stabilities, and tunable active sites. However, the exclusive presence of micropores coupled with extended diffusion pathways impedes molecular transport, leading to catalyst deactivation because of carbon deposition. Consequently, nanosheet zeolites with two-dimensional structures have emerged as promising candidates to mitigate the diffusion limitations, attracting considerable research interest in the past few decades. In this review, we comprehensively summarize the recent advances in the synthesis strategies and catalytic applications of nanosheet zeolites. Various synthesis approaches, including in situ hydrothermal synthesis and post-synthetic treatments are highlighted. Furthermore, we systematically analyze the physicochemical properties of nanosheet zeolites and their demonstrated effectiveness in diverse catalytic reactions, such as methanol conversion, cracking, isomerization, alkylation, carbonylation, and oxidation reactions. This review provides a foundational framework for the rational design of nanosheet zeolite catalysts and offers insights into their potential application in next-generation industrial catalysis.



Wider impact

Nanosheet zeolites with two-dimensional structures have emerged as promising candidates to mitigate diffusion limitations of zeolites, attracting considerable research interest in the past few decades. In this review, we provide a systematic overview of recent progress in the synthesis, structural characterization, and catalytic applications of nanosheet zeolites. By elucidating key advancements and outlining the future research directions, this review highlights the pivotal role of zeolite nanosheets as versatile functional materials with significant potential in advancing next generation catalytic technologies. In contrast to prior reviews, we present a more comprehensive discussion, encompassing the synthetic strategies for nanosheet zeolites, structure–property correlations governing their physicochemical characteristics and their catalytic performance towards diverse reaction systems. Additionally, we summarize the latest breakthroughs in nanosheet zeolite catalysts, offering critical insights into emerging opportunities in the field. Therefore, we believe that this review will be of great interest to researchers in the fields of nanosheet zeolites and industrial catalysis.

1. Introduction

The history of zeolites can be traced back to 1756, when Swedish mineralogist Cronstedt discovered a mineral in nature that continuously bubbled upon heating in water, resembling boiling. He named that mineral “zeolite”, meaning “boiling stone.” Following this, early researchers discovered several minerals with zeolite-like properties, such as stilbite, natrolite, and other low-silica zeolites. As the discovery of natural zeolites progressed, researchers began to explore their properties, leading to their application in adsorption, separation, and ion exchange.1–3 A pivotal moment in zeolite science occurred in 1948, when Barrer synthesized a zeolite that did not exist in nature by transforming known minerals in a high-temperature concentrated salt system through iterative experiments. This breakthrough marked the beginning of artificial zeolite synthesis. Building on Barrer's work, Milton and colleagues used the hydrothermal synthesis method to successfully synthesize various zeolites, including A-type and X-type zeolites. During this period, zeolite synthesis primarily focused on low-silica zeolites. However, the relatively high aluminum content of these zeolites resulted in poor hydrothermal stability, thereby constraining their practical applications. Through continued research and experience accumulation, significant progress has been made in the artificial synthesis of zeolites. A notable development was Barrer's introduction of a novel approach incorporating quaternary ammonium salts into the zeolite synthesis system in 1961. As synthesis techniques for zeolites have advanced, the diversity of zeolite structures has expanded considerably, with the number of known zeolite topologies reaching 255.

Traditionally, zeolites have been extensively utilized in diverse applications, including catalysis, separation processes, and environmental remediation.4–9 However, the relatively long microporous channels of traditional zeolites limit the mass transfer and diffusion of reactants and products between active sites, particularly in the case of large molecules. This often results in undesirable side reactions, carbon deposition, rapid catalyst deactivation, and other complications. Therefore, effectively reducing diffusion limitations has emerged as a research focus in the controlled synthesis of zeolites. In recent years, the emergence of zeolite nanosheets, characterized by a two-dimensional (2D) morphology, has garnered significant attention owing to their unique structural and functional advantages.10–18 These nanosheets, typically with a thickness in the order of a few nanometers, offer significantly increased surface areas, more accessible active sites, and improved mass transport properties compared with their bulk counterparts.19,20

To date, the synthesis of zeolite nanosheets has been the focus of numerous studies, employing methods such as in situ hydrothermal synthesis with organic structure-directing agents,21–30 additive-assisted synthesis,32–37 seed-induced synthesis,38–41 inter-growth synthesis,42–44 chemical etching,45,46 swelling, exfoliation, and pillaring47–53 by post-synthetic to control their structure and properties. The transition to reduced dimensionality not only alters the physical and chemical characteristics of zeolites including their morphology,54–63 structure,47,67 porosity,68,69 and acidity,70–77 but also imparts unique catalytic behavior, making zeolite nanosheets particularly attractive to realize challenging catalytic transformations, such as methanol conversion,78–84 cracking,85–91 isomerization,92–99 alkylation,100,101 carbonylation,102–104 and catalytic oxidation.105–112

Notably, recent comprehensive reviews have covered the developments in two-dimensional zeolites.113–115 However, their discussion mainly focused on the qualitative correlations between surface properties and performance, while detailed case studies comparing the catalytic behavior of different nanosheet zeolite structures in specific reaction systems are still lacking. Thus, in the present review, we specifically emphasize the controlled synthesis strategies, advanced structural characterization, and catalytic applications of nanosheet zeolites, providing a systematic summary of the most recent advancements in nanosheet zeolites. The discussion focuses on the impact of nanosheet morphology on catalytic performance and explores the potential of nanosheet zeolites in emerging applications, such as energy conversion and environmental catalysis. By highlighting critical developments and future directions in this field, this review underscores the significance of zeolite nanosheets as multifunctional materials with considerable potential for next-generation catalytic technologies.

2. Synthesis of nanosheet zeolites

Recently, strategies for the synthesis of nanosheet zeolites have been widely reported. Nanosheet zeolites with diverse frameworks, including MFI, MWW, FAU, FER, and MOR, have been successfully fabricated. The synthesis of 2D or nanosheet zeolites can be primarily categorized into two distinct methodologies, as follows: (I) in situ hydrothermal synthesis, which limits crystal growth in specific directions to form lamellar structures via the use of template agents or induced seeds. (II) Post-synthesis treatments, such as chemical etching, pillaring, and exfoliation. In the past few decades, nanosheet zeolites, particularly MFI, MWW, and FER, have achieved industrial-scale production and demonstrated exceptional catalytic performances in various practical applications. More details of their synthesis methods are elaborated in the following section.

2.1 In situ hydrothermal synthesis

2.1.1 Organic structure-directing agents with special functional groups. Basically, the design and synthesis of quaternary ammonium surfactants as organic structure-directing agents (OSDAs) have been used to optimize the interaction between the template agents and the precursor, thereby reducing phase separation during the zeolite crystallization process. Quaternary ammonium surfactants consist of two key components, a long-chain alkyl group (tail) and a quaternary ammonium group (head). The quaternary ammonium groups serve as SDAs, facilitating the formation of the zeolite framework. Meanwhile, the long-chain alkyl groups play a pivotal role in controlling the growth of zeolite crystals, resulting in the formation of nanosheet structures.

Ryoo et al. first reported the synthesis of ultrathin MFI nanosheets in 2009.21 They successfully synthesized the multilamellar ZSM-5 nanosheet zeolites by designing an organic structure-directing agent with a bifunctional group of C22H45–N+(CH3)2–C6H12–N+(CH3)2–C6H13(C22-6-6Br2). As shown in Fig. 1a, two quaternary ammonium groups are located at the channel intersections, one is inside the framework, and the other is at the pore mouth of the external surface. The dual ammonium ions in this structure-directing agent play a crucial role in templating the zeolite framework during crystallization. Simultaneously, the presence of long-chain organic functional groups attached to both sides of the ammonium ions induce pronounced hydrophobic interactions, which effectively suppress excessive crystal growth along the b-axis, enabling the formation of well-defined nanostructures.


image file: d5mh00579e-f1.tif
Fig. 1 Proposed structure model for a single (a) MFI nanosheet. Reprinted with permission from ref. 21. Copyright 2009, Springer Nature. (b) MOR nanosheet. Reprinted with permission from ref. 25. Copyright 2020, Wiley-VCH GmbH. (c) MWW nanosheet. Reprinted with permission from ref. 26. Copyright 2015, The Royal Society of Chemistry. (d) SAPO-34 nanosheet. Reprinted with permission from ref. 30. Copyright 2019, Springer Nature.

As a result, the obtained ZSM-5 nanosheets demonstrate exceptional structural characteristics, with a remarkably reduced thickness of only 2.5 nm and an impressive surface area reaching 720 m2 g−1, which significantly enhances their mass transfer properties. In addition, varying the Na+ concentration in the synthesis system can alter the structure of the nanosheets.22 When the concentration of Na+ is high, the ZSM-5 nanosheet zeolite with an ordered arrangement is obtained, whereas a low Na+ concentration results in a disordered arrangement. Due to the disordered arrangement, the ZSM-5 nanosheet exhibits an increased specific surface area, with the inhibited condensation of silica–hydroxyl groups between the lamellae during calcination. By varying the length of the organic long-chain and short-chain alkyl groups, it was found that the chain lengths of the alkyl groups affect not only the ZSM-5 nanosheet architecture but also the interlamellar spacing. Furthermore, researchers also designed an organic functional group with three quaternary ammonium ions of C18H37–N+(CH3)2–C6H12–N+(CH3)2–C6H12–N+(CH3)2–C18H37 by modifying the long-chain organic groups. Using this OSDA, nanosheets with a single-cell thickness of only 1.5 nm were successfully synthesized,23 which are also the thinnest MFI nanosheet zeolites synthesized to date via direct synthesis.

Building on Ryoo's pioneering work and inspired by this synthetic concept, other researchers designed similar organic structure-directing agents containing specific functional groups. The π–π interactions between aromatic groups have emerged as widespread forces in chemistry and biology, playing a pivotal role in molecular self-assembly, recognition, and structure determination. In research on the synthesis of zeolites, π–π interactions can effectively stabilize micelle structures and aromatic ring orientations during the self-assembly process, thereby influencing the zeolite growth in a selective manner, which is crucial for synthesizing mesoporous and nanosheet zeolites. Regarding the bifunctional organic structure-directing agents reported by Ryoo's group, Che et al. designed amphiphilic organic structure-directing agents containing aromatic groups (C6H5–C6H4–O–C10H20–N+(CH3)2–C6H13(Br)),24 leading to the successful synthesis of ZSM-5 zeolites with a thickness of ∼30 nm. Additionally, by modifying the aromatic-containing OSDA, a bis-quaternary ammonium organic structure-directing agent was further synthesized.

The application of the quaternary ammonium surfactants has extended beyond MFI-type zeolites, demonstrating the versatility in the synthesis of diverse nanosheet zeolites. Wu et al. used a well-designed bifunctional amphiphilic surfactant, C16H33–N+(CH3)2–C4H8–N+(CH3)2–benzylamine, as an organic structure-directing agent.25 The organic–inorganic interaction between the surfactant and the zeolite framework is crucial for forming the multilayer MOR nanosheet structure. The benzyl di-quaternary ammonium cation controls the formation of the MOR topology, while the long hydrophobic hexadecyl tail group effectively suppresses crystal growth along the b-axis. As a result, aluminosilicate MOR zeolite nanosheets with a thickness of 11 nm, highly exposed (010) planes, and 8-membered ring (8-MR) windows were successfully synthesized. Zhu et al. reported the synthesis of aluminosilicate beta zeolite nanosheets (beta-NS) using a polycationic organic structure-directing agent, [–N+(CH3)2–C5H10–N+(CH3)2–C6H12–]n[Br]2n(Me4–C5-6). The OSDA preferentially directed the aluminum atoms near the ammonium moieties within the straight channels, resulting in enhanced acidity and diffusion properties. The obtained nanosheets exhibited a “house-of-cards” architecture with interconnected space, which significantly improved their catalytic performance towards n-heptane cracking.26

MWW zeolites, recognized for their exceptional efficacy in the alkylation of benzene with ethylene, undergo a crystallization process to produce a multi-layered precursor known as MCM-22(P). In contrast, the material identified as MCM-56 is characterized by a distinct structure of singular nanosheets. However, due to the low mesoporous activity of the MCM-56 zeolite, the direct synthesis of ultrathin MWW zeolite nanosheets remains a significant challenge. Thus, to address this issue, Román et al. developed an innovative method for the one-pot synthesis of MIT-1, consisting of delaminated MWW zeolite nanosheets.27 MIT-1 was synthesized using a rationally designed OSDA integrating the features of the traditional OSDA used for the synthesis of the MWW layered zeolite precursor and the quaternary ammonium surfactant employed for swelling and delamination. The resulting MIT-1 exhibited superior structural characteristics, including high crystallinity, surface area exceeding 500 m2 g−1, significant mesoporosity, and high acid site density. Later, Corma et al. synthesized a highly delaminated MWW zeolite monolayer, referred to as DS-ITQ-2,28 using a single-step process involving a combination of hexamethyleneimine (HMI) and a bifunctional organic structure-directing agent (N-hexadecyl-N′-methyl-DABCO, C16DC1). HMI is responsible for the crystallization of the MWW framework, while the bifunctional quaternary ammonium surfactant inhibits crystal growth and stacking along the c-axis. The nanosheet MWW zeolite consisted of approximately 70% single and double layers, with a well-structured external surface area of ∼300 m2 g−1.

The FER zeolite is a high-silica material with excellent catalytic properties, but previous attempts to synthesize ultrathin FER nanosheets failed to maintain its crystallinity. Corma and colleagues synthesized nanosheet FER zeolites with high external surface areas and a well-preserved microporous structure.29 Their approach involved the use of a dual-template system consisting of piperidine and a modified surfactant (C16MPip) as OSDAs to prepare nano-ferrierite zeolites. The C16MPip surfactant played a dual role in the crystallization process of ferrierite zeolite, where its long carbon chain restricts crystal growth, while its head group adapts to the microporous channel system, thereby influencing the size and morphology of the crystals. As a result, nano-ferrierite crystals with a controlled size of 10–20 nm were obtained, particularly along the 10-ring channel direction [001]. In a significant advancement, Xiao's research team achieved the direct synthesis of ultrathin nanosheet FER zeolites (6–8 nm) using a small organic ammonium (N,N-diethyl-cis-2,6-dimethyl piperidinium, DMP) as the structure-directing agent.30 The DMP molecules preferentially adsorb on the FER {100} surface, effectively inhibiting crystal growth along the [100] direction. Additionally, these molecules function as a structure-directing agent during the crystallization of FER zeolite, further inhibiting zeolite growth along the [100] direction, which led to the formation of ultrathin FER zeolite nanosheets.

Quaternary ammonium surfactants have also been employed in the synthesis of SAPO-34 nanosheets. Chen et al. synthesized hierarchical SAPO-34 zeolites with a unique nanosheet-assembled morphology by employing the commercial organosilane surfactant [3-(trimethoxysilyl)propyl]octadecyl dimethyl-ammonium chloride (TPOAC) as a mesoscopic aggregation agent and tetraethylammonium hydroxide (TEAOH) as a micropore structure-directing agent.31 In this system, TPOAC acts as a surface stabilizer in the synthesis of the SAPO-34 zeolite by reducing the surface energy, which influences the morphology of the zeolite. Simultaneously, TPOAC assists in forming the mesoscopic agglomeration structure of the zeolite, promoting the formation of nanosheet SAPO-34 zeolite.

2.1.2 Additive-assisted synthesis. During the synthesis of zeolites, the introduction of additives regulates crystal growth by interacting with the molecular zeolite precursors through hydrogen bonds, van der Waals force, or hydrophobic interactions. Alternatively, additives can prevent the aggregation of the precursors by interacting with them, thus affecting the zeolite growth. As illustrated in Fig. 2, additives directionally adsorb on different crystal faces of the zeolite, clearly inducing differential growth rates of different crystal faces and effectively controlling the oriented growth of the crystal faces. Compared to costly designed templates, additives employed in the synthesis nanosheet zeolites are commercially available, making them attractive for large-scale industrial production. Recent advancements in this field have successfully applied this method to the synthesis of nanosheet zeolites.
image file: d5mh00579e-f2.tif
Fig. 2 Schematic of the modifications in the zeolite morphology after adding growth modifiers. (a) Reprinted with permission from ref. 35. Copyright 2021, the American Chemical Society. (b) Reprinted with permission from ref. 32. Copyright 2012, Wiley-VCH GmbH. (c) Reprinted with permission from ref. 36. Copyright 2024, Elsevier.

Rimer et al. employed spermine and D-arginine to significantly modify the morphology of the silicalite-1 crystal, including reducing the platelet thickness and directing the formation of new crystal facets.32 Xiao et al. introduced urea into the synthesis, resulting in the formation of TS-1 and ZSM-5 zeolites with a nanosheet morphology.33 Fluoride, as an important halide, acts as a mineralizing agent in zeolite synthesis systems, which also aids in the preparation of zeolite nanosheets. Xu et al. demonstrated that fluoride affected the synthesis process of the ITQ-13 zeolite.34 Experimental results revealed that the fluorine-to-silicon ratio in the product decreased with an increase in the H2O/Si ratio of the reactants, which hindered the formation of ITQ-13. Dai et al. combined preliminary aging with fluoride-assisted low-temperature crystallization to obtain micrometer-sized MFI zeolites along the a- and c-axes, while maintaining ultrathin dimensions (tens of nanometers) along the b-axis.35 The addition of NH4F disrupted the gel structure and generated nanoparticles, promoting the formation of plate-like crystals. Meanwhile, it reduced the pH value, changing the solubility of silicate species and leading to preferential growth in the a- and c-directions, thus forming ultra-thin crystals along the b-axis. More recently, they prepared plate-like TS-1 zeolites with a controllable thickness along the b-axis (80–1000 nm) using L-lysine as a growth modifier. Compared to the conventional TS-1 zeolite and fluoride-mediated plate-like TS-1, the obtained plate-like TS-1 showed improved catalytic performance in the epoxidation of 1-hexene.36

Recently, Shen et al. synthesized a plate-like MOR catalyst via a two-step crystallization process using triethylenediamine (TEDA) as the structure-directing agent (SDA) and γ-aminobutyric acid as the crystal growth medium.37 The as-obtained catalyst has a plate-like morphology with a thickness of 80–300 nm and a width of 2–3 μm, with the plates arranged in a cross-stacked “house of cards” pattern.

2.1.3 Seed-induced synthesis. In this synthesis method, seed crystals of the target zeolite are initially prepared through the conventional method with template agents. Subsequently, a small amount of the synthesized seed crystals is introduced into a template-free reactant gel to form the desired zeolite phase. The addition of zeolite seed crystals accelerates the nucleation and crystallization rates, thereby improving the quality of the desired zeolite phase by enhancing crystallization. This seed-induced crystallization of zeolites without organic templates has been successfully employed in the synthesis of several important zeolites. Tsapatsis et al. used a nanocrystal-seeded growth approach triggered by a single rotational intergrowth to produce 5 nm-thick MFI nanosheets. These nanosheets could be used to fabricate thin, defect-free coatings that effectively covered porous substrates, enabling the production of high-flux and ultra-selective MFI membranes.38 Fig. 3a illustrates the growth process of MFI zeolite nanosheets. Initially, the seed crystals exhibit a near-cylindrical morphology (∼140 nm in length), with the b-axis as the primary growth direction. At this stage, crystal growth predominantly proceeds via a surface epitaxial growth mechanism. However, after 20–40 h of hydrothermal treatment, the growth mode undergoes a distinctive transition, as nanosheets begin to rapidly extend from one corner of the seed crystals, eventually encapsulating them completely. The resulting nanosheets exhibited a uniform thickness corresponding to five pentagonal ring chains (5 nm). Moreover, crystallographic analysis revealed a 90° rotational intergrowth relationship between the nanosheets and the seed crystals, wherein they shared a common c-axis but displayed a 90° rotation between the a- and b-axes. This unique crystallographic alignment was proposed to drive the accelerated growth of the nanosheets. Notably, even with smaller and smoother seed crystals, an analogous growth process occurred, ultimately yielding 5 nm-thick MFI nanosheets. Thus, the transition from slow epitaxial growth to rapid nanosheet formation is governed by the orthogonal rotational intergrowth between the seed crystal and the nanosheet. Zhu et al. synthesized hierarchical ZSM-5 nanosheets containing intracrystalline mesopores with a honeycomb-like morphology.39 In this method, cetyltrimethylammonium bromide (CTAB) was employed as the secondary template under hydrothermal conditions. The proposed mechanism for the self-assembly between sub-nanocrystals and CTAB under hydrothermal conditions is illustrated in Fig. 3. The resulting hierarchical ZSM-5 nanosheets displayed an excellent catalytic performance towards the alkylation of toluene with methanol, primarily due to their increased abundant accessible acid sites and enhanced mass transfer properties.
image file: d5mh00579e-f3.tif
Fig. 3 (a) Growth stages of MFI nanosheet from a seed crystal. Reprinted with permission from ref. 38. Copyright 2017, Springer Nature. (b) Mechanism of the formation of hierarchical ZSM-5 nanosheets and conventional ZSM-5. Reprinted with permission from ref. 39. Copyright 2015, the Royal Society of Chemistry. (c) Idealized schematic of heterogeneous nucleation and growth of SPP zeolites on amorphous interfaces in the absence of organics. Reprinted with permission from ref. 41. Copyright 2021, Wiley.

Tatsuya et al. successfully synthesized MCM-22 (MWW-type) zeolite using an OSDA-free gel system.40 The in situ synthesis process necessitated the continuous presence of MCM-22 seeds throughout crystallization to provide a growth surface for the MWW-type zeolite, preventing the spontaneous nucleation of mordenite. Subsequently, Rimer et al. reported a seed-assisted method for the preparation of self-pillared pentasil zeolites without employing organic structure-directing agents.41 This shows that the use of MEL- or MFI-type zeolites as crystalline seeds can induce the formation of hierarchical zeolites with a large external surface area and large proportion of external acid sites. This is achieved through heterogeneous nucleation and the growth of branched nanosheets from inorganic precursors, thereby enhancing the performance and reducing the mass transfer limitations of the catalyst. In contrast to the use of expensive OSDAs, this method offers an environmentally benign and cost-effective alternative for producing industrially important zeolites.

2.1.4 Intergrowth synthesis. In the field of zeolite synthesis, heteroepitaxial growth is a common phenomenon observed in systems such as MFI/MEL, EMT/FAU, ETS-4/ETS-10, and CAN/SOD. The synthesis of nanosheet zeolites, particularly when the sheet thickness is reduced to a certain extent, often results in the condensation of the silicon hydroxyl groups during calcination or subsequent processing. This causes sheet agglomeration, significantly diminishing the advantages of nanosheet zeolites in applications. The heteroepitaxial growth mode employed in nanosheet zeolite synthesis effectively alleviates the agglomeration and condensation between lamellae. Additionally, the “house of cards” growth mode introduces mesopores, further enhancing the catalytic performance of the nanosheet zeolites. Tsapatsis et al. demonstrated the successful synthesis of lamellar zeolites with a “house of cards” morphology formed by MFI and MEL zeolites using TBPOH, achieving a lamellae thickness as low as 2 nm.42 They also synthesized intergrowth FAU/EMT zeolites, predominantly consisting of FAU-type zeolites.43 As illustrated in Fig. 4a, the nucleation of EMT zeolites occurred randomly in any part of the (111) crystallographic plane. However, when both EMT and FAU grow on the (111) plane, the two crystals rapidly extend along this plane, forming uncrossed lamellar intergrowth zeolites. If nucleation occurs at the edge of lamellar FAU, specifically at the (111) and (11[1 with combining macron]) facets, the growth of lamellar EMT at the FAU edge competes for the precursors, slowing FAU growth along the (111) plane. As EMT continues to grow, it gradually transforms into lamellar EMT, ultimately leading to the formation of an intersection between EMT and FAU. As shown in Fig. 4f, Zhu et al. synthesized beta-nanosheets with only straight micropore channels, which involved the intergrowth of polymorph B and polymorph C (BEC topology).26 Tsapatsis et al. used transmission electron microscopy (TEM) to identify the presence of one- to few-unit-cell-wide intergrowths of the zeolite MEL within 2D-MFI nanosheets. They found that a fraction of the nanosheets have a high (∼25% by volume) MEL content, while the majority are MEL-free, as shown in Fig. 4d. Moreover, they also found that confining the growth of MFI nanosheets in graphite can increase the MEL content.44
image file: d5mh00579e-f4.tif
Fig. 4 Illustration of the branching mechanism in T-H-FAU. (a) Nucleation of EMT close to the edge of sheet 1 on (111). (b) TEM image highlighting the triangular assembly of T-H-FAU sheets with 3-fold symmetry. (c) Structural model showing defect formation when the EMT domains on (111) and (11[1 with combining macron]) meet. Reprinted with permission from ref. 41. Copyright 2014, Wiley-VCH GmbH. (d) Filtered ADF-STEM image (left) and FFT (right) for an area with typical MFI structure (yellow) and with MEL (red) domains intergrown within the MFI framework. (e) Cross-correlated and averaged ADF-STEM section of MFI–MEL heterostructure from the dashed region. Reprinted with permission from ref. 44. Copyright 2020, Springer Nature. (f) HRTEM image taken from [110] direction of polymorph B and [100] direction of polymorph C. (g) Corresponding SAED and simulation ED patterns are shown in. (h) Proposed structural model of the epitaxial intergrowth. Reprinted with permission from ref. 26. Copyright 2025, Wiley.

2.2 Post-synthetic treatments

In contrast to the above-mentioned synthetic strategy, nanosheet zeolites can also be fabricated through the post-synthetic treatment of zeolite crystals, yielding a diverse range of zeolites with two-dimensional layered structures. The primary synthesis methods include chemical etching, swelling, exfoliation and pillaring by amorphous SiO2. Post-synthetic treatments are extensively utilized in industry and play a pivotal role in the commercialization and application of zeolites. These treatments enable the formation of zeolites with excellent catalytic and adsorptive properties, including high stability, tailored compositions, and desired acid-site distributions.
2.2.1 Chemical etching. Valtchev et al. synthesized layered ZSM-5 zeolite through a synthesis process involving fluoride etching with ammonium fluoride (NH4F).45 The purpose of NH4F etching in zeolite modification is to extract the framework cations in a selective manner, open the structure, and enhance the number of active sites, while maintaining the chemical composition or acidity. NH4F etching allowed precise tuning of the hierarchical structure of the zeolites, thereby improving their performance in catalytic reactions.

Liu et al. prepared single-crystalline hierarchical ZSM-5 nanosheets via alkaline etching with TPAOH.46 Alkaline etching increased the Brønsted acid site concentration and acid strength of the ZSM-5 zeolite by controlling the Si leaching, introducing additional Al species into the framework, and creating mesoporous structures. Additionally, alkaline etching preserved the morphological integrity of the ZSM-5 zeolite at low Si/Al ratios, improving its chemical and mechanical stability, which enhanced its catalytic performance. By optimizing the hydrothermal treatment conditions, uniform ZSM-5 nanosheets with smooth surfaces, a significantly reduced Si/Al ratio, and good crystallinity could be obtained.

2.2.2 Swelling, exfoliation and pillaring. The pillaring process involves the insertion of guest molecules between the layers of layered materials, which not only preserves the ordered structural integrity of the host material but also introduces additional chemical functionalities beyond its intrinsic properties. In the context of nanosheet-structured zeolites, a pillar support is predominantly achieved through two approaches, the intercalation of organic compounds or incorporation of thermally stable inorganic “pillars” into the interlayer spaces to ensure layer separation. These “pillars” are critical for maintaining the spatial separation between the layers, while the decomposition of adsorbed swelling agents generates interconnected pore networks within the voids created by their removal. This process enhances the structural stability and functional versatility of the material.

Lamellar zeolite precursors, a special type of 2-dimensional zeolites, have garnered considerable attention from researchers in this field. During the post-treatment process, multiple methods often need to interact synergistically to obtain lamellar zeolite precursors. The uncalcined MWW zeolite precursor (MWW-22P) is first subjected to swelling treatment using the long-chain organic surfactant cetyltrimethylammonium bromide (CTAB) as the swelling agent. The positively charged CTA+ interacts electrostatically with the negatively charged zeolite surface, causing it to swell. Additionally, during the swelling process, the acidity or alkalinity of the swelling system and the amount and type of swelling agent strongly affect the degree of swelling. Once swelling is achieved, SiO2 is introduced for pillaring to separate the dissolved lamellae and subsequent calcination produces the SiO2-pillared MCM-36. Moreover, Corma et al. further synthesized the ITQ-2 zeolite through a delamination method.47 As exhibited in Fig. 5a, the specific steps are as follows: firstly, the MCM-22(P) precursor consisting of inorganic layers and organic molecules (hexamethyleneimine) is prepared. Next, the MCM-22(P) precursor is suspended in an aqueous solution containing hexadecyltrimethylammonium bromide and tetra-propylammonium hydroxide, followed by refluxing at 353 K for 16 h to promote its expansion. The expanded solution is then placed in an ultrasonic bath for 1 h to further separate the layered structure. Finally, organic species are removed through acidification and calcination to obtain the ITQ-2 zeolite. This delamination process significantly increases the external surface area of ITQ-2, while preserving its internal structure, resulting in better catalytic activity compared to both its parent MCM-22 and the pillared MCM-36.


image file: d5mh00579e-f5.tif
Fig. 5 Schematic for the preparation of swelling, exfoliation and pillaring obtained via the (a) vapor-phase pillaring process of MFI nanosheet. Reprinted with permission from ref. 45. Copyright 2019, the American Chemical Society; (b) pillared MCM-22, Reprinted with permission from ref. 47. Copyright 2011, the American Chemical Society; (c) self-pillared MFI nanosheets. Reprinted with permission from ref. 46. Copyright 2021, the American Chemical Society; and (d) exfoliated SAPO-34 precursor. Reprinted with permission from ref. 51. Copyright 2022, Elsevier.

Liu et al. reported the preparation of a self-pillared two-dimensional layered MFI zeolite through a vapor-phase transformation process, in which TEOS was intercalated between layers of MFI zeolite nanosheets and converted into stable SiO2 pillars upon calcination.48 As shown in Fig. 5a, the intercalation efficiency of TEOS reached up to 100% with the minimal amount of water. The “pillaring” process endowed these MFI zeolite nanosheets with high stability, and even after calcination, the mesoporous voids created by the surfactant tails were preserved. Furthermore, the resulting mesoporous structures after “pillaring” provided additional sites for transition metal loading. The abundant Si–OH groups on the nanosheet surfaces facilitated further functionalization through silanization, enabling the modification of their surface properties or the grafting of catalytically active groups and metals.

More recently, Yu et al. used self-pillared MFI zeolite nanosheets as supports to immobilize ultrasmall monometallic (Rh and Ru) and bimetallic (Rh–Ru, Rh–Au, etc.) clusters.49 The metal clusters were uniformly dispersed within the zeolite framework and remained stable at high temperatures. Compared to conventional zeolite-supported catalysts, the zeolite nanosheet-supported samples offer distinct advantages of high surface area, abundant Si–OH groups, and improved mass transport. As displayed in Fig. 5c, the isolated Si–OH defects serve as effective anchoring sites for sub-nanometer metal clusters, and increasing the density of isolated Si–OH defects led to improved metal dispersion and smaller cluster sizes. During the aforementioned post-modification process, due to the entrapment of some gas molecules within the zeolite framework, the metal precursor solution is less likely to fill the entire micropores of the zeolite. This results in excessive metal deposition on the external surface during the reduction and reaction processes, which further aggregates to form larger nanoclusters or nanoparticles.

Notably, performing this process typically requires high surfactant and hydroxide (OH) concentrations under elevated temperatures, which often induce partial amorphization. Zones et al. synthesized UCB-1 by delaminating MCM-22(P) at a pH value of 9 using an aqueous solution containing cetyltrimethylammonium bromide, tetrabutylammonium fluoride, and tetrabutylammonium chloride at 353 K.50 This contrasts with the conventional high-pH delamination method that requires sonication. Unlike the previously reported delaminated zeolite precursor ITQ-2, UCB-1 maintained a high degree of structural integrity and did not undergo amorphization. Katz et al. treated the borosilicate layered zeolite precursor (ERB-1P) with a warm Al(NO3)3 solution.51 This treatment process caused the substitution of aluminum for boron in the zeolite framework, disrupting the hydrogen bonding interactions between the interlayers and transforming a regular layered structure in single layers and curved structures. Similarly, this single-step delamination method does not require corrosive pH conditions, high-energy ultrasonication, or organic surfactants, offering a simple and efficient route. Moreover, Roth et al. achieved the exfoliation of ZSM-55 and MCM-56 layers into monolayers using tetrabutylammonium hydroxide (TBAOH) solution.52,53

The exfoliation method was also reported for the synthesis of SAPO-34 nanosheets by Xing et al. SAPO-34 nanosheets can be prepared using the solvent-assisted freeze–thaw method, which involves the following steps:54 (I) 0.1 g of layered SAPO-34 crystals is dispersed in hexane (3 mL), followed by freezing in a liquid nitrogen bath, and sonication, repeating this process 20 times. (II) The treated suspension is mixed with ethanol and left to stand for 12 h, causing the unexfoliated larger crystals to settle at the bottom. (III) High-speed centrifugation (8000 rpm) separates the SAPO-34 nanosheets from the upper suspension. Notably, heat treatment, such as drying in an oven or muffle furnace, is avoided to prevent nanosheet aggregation. The obtained nanosheets had a thickness of ∼4 nm, length greater than 1 μm, and high aspect ratio ranging from 230 to 560.

3. Physicochemical properties of nanosheet zeolite

Nanosheet zeolites exhibit an ultrathin sheet morphology with a thickness ranging from 2 to 100 nm, displaying unique catalytic properties. Typically, surfactant molecules align along the channels of the MFI framework. Assemblies are formed along the b-axis to create multilayer structures, or adopt random arrangements to form monolayer configurations. These structural variations significantly influence the distribution of active sites. In contrast to the conventional three-dimensional (3D) zeolites with active sites predominantly situated within the micropores, two-dimensional (2D) zeolites predominantly feature active sites located on their external surface, resulting in unique application potential. Consequently, a comprehensive understanding of the structural and chemical characteristics of nanosheets is crucial for optimizing their catalytic performance.

3.1 Morphology analysis

In contrast to the typical hexagonal-prismatic or coffin-shaped morphology of the conventional ZSM-5, nanosheet MFI zeolites exhibit diverse morphologies, which are influenced by both the nanosheet architecture and the assembly process. Hu et al. reported the successful synthesis of a high-silica nanosheet MFI zeolite using low-cost raw materials and a bifunctional organic surfactant ([C18H37–N+(CH3)2–(CH2)6–N+(CH3)2–C6H13]Br2), confirming that the obtained nanosheet zeolite exhibits a lamellar stacking morphology, with three-dimensionally intergrown nanosheets.55 Ryoo et al. investigated the hierarchically structure-directing effect of multi-ammonium surfactants in the synthesis of MFI nanosheets.21–23 They synthesized a range of surfactants with a different number of ammonium groups, alkyl tail lengths, and spacer lengths between the ammonium groups. The surfactant molecule must contain at least two ammonium groups to effectively direct the formation of the zeolite structure, alongside the amphiphilicity and micelle packing ability of the surfactant. The –C3H6–, –C6H12–, and –C8H16– spacers (CiH2i) in C22-iN2 led to the formation of micrometer-sized bulk crystals, a multilamellar structure, and disordered nanosheets, respectively. Additionally, the thickness of the resulting zeolite nanosheets could be tuned by varying the number of ammonium groups in the surfactant. Using the C22-6N2 surfactant, each nanosheet consisted of three five-membered ring layers with a thickness of ∼2 nm, whereas the C22-6N3 and C22-6N4 surfactants induced the formation of thicker nanosheets with five and more than seven five-membered ring layers, respectively. This is due to the fact that the amino head groups of the surfactant act as structure-directing agents for the zeolite, and the increase in the number of amino groups promotes preferential growth of the zeolite along the b-axis, thereby increasing the nanosheet thickness. The morphology of the different types of nanosheet zeolites is exhibited in Fig. 6.
image file: d5mh00579e-f6.tif
Fig. 6 Morphology of nanosheet zeolites: (a)–(d) MFI nanosheet. Reprinted with permission from ref. 21. Copyright 2009, Springer Nature; reprinted with permission from ref. 35. Copyright 2021, the American Chemical Society; reprinted with permission from ref. 23. Copyright 2011, the American Chemical Society. (e) MWW nanosheet. Reprinted with permission from ref. 43. Copyright 2020, the American Chemical Society. (f) FER nanosheet. Reprinted with permission from ref. 62. Copyright 2022, the Royal Society of Chemistry. (g) FAU nanosheet. Reprinted with permission from ref. 58. Copyright 2016, Elsevier. (h) and (i) MOR nanosheet. Reprinted with permission from ref. 37. Copyright 2023, Elsevier; reprinted with permission from ref. 25. Copyright 2020, Wiley-VCH GmbH.

Fang et al. reported the direct synthesis of hierarchically structured MFI zeolite nanosheet assemblies via a seed-assisted strategy.56 The resulting assemblies showed a narrow mesopore distribution, and their morphology could be adjusted from nanosheet stacks to house-of-cards structures by varying the Si/Al ratio to 31. A lower Si/Al ratio led to irregular stacks, whereas a higher Si/Al ratio produced a nanosponge-like morphology. Additionally, the particle size of the obtained zeolite was strongly dependent on the seed content. Increasing the seed content in the gel from 5 to 30 wt% reduced the particle size from 0.8–1.2 μm to ∼500 nm. Zhan et al. synthesized ZSM-5 zeolite nanosheets with thicknesses ranging from ∼4 nm to 20 nm by adjusting the ratio of surfactant to tetraethyl orthosilicate (TEOS), which influenced their pore structure and catalytic performance.

Wattanakit et al. fabricated hierarchical FAU nanosheets via a hydrothermal process, using dimethyloctadecyl-(3-(trimethoxysilyl)propyl) ammonium chloride (TPOAC) as a hierarchical porogen.57 They found that the TPOAC content and crystallization temperature played significant roles in determining the morphologies and textural properties of the hierarchical FAU nanosheets. As the TPOAC content increased, the FAU nanosheets transitioned from smooth surfaces to rough spherical crystals, while excessive TPOAC inhibited the self-assembly of the nanosheets. Simultaneously, the microporous specific surface area increased with an increase in the crystallization temperature. Therefore, an appropriate TPOAC content (0.015–0.045 mole fraction) and a high crystallization temperature (85 °C) are crucial for achieving high-crystallinity and uniform FAU nanosheet morphologies. In a distinct approach, Ferdov et al. established the hierarchical assembly of FAU nanosheets by controlling the conditions near the boundary of FAU and EMT zeolite co-crystallization.58 FAU-type zeolite nanosheets with a ball-like morphology and a diameter of ∼2 μm were obtained without employing any organic or inorganic morphology-directing templates or agents. In a template-free system, Li et al. synthesized 50–100 nm thick H-mordenite with a nanosheet stack morphology by adjusting the water content.59 A decrease in alkalinity (higher H2O/SiO2 ratio) promoted the growth and assembly of H-MOR nanocrystals, forming a nanosheet structure.60 Dai et al. used γ-aminobutyric acid as a growth modifier to produce plate-like ZSM-5 with a b-axis thicknesses of ∼50 nm. Low pH values facilitated the formation of plate-like ZSM-5 zeolites.61 The addition of γ-aminobutyric acid as a growth regulator effectively suppressed growth in the [100] direction, while simultaneously promoting oriented growth along the c-axis, thereby yielding a thinner plate-like structure.

Yang et al. developed ultrathin FER nanosheets (designated as SCM-37) through a dual-template strategy using octyl-trimethylammonium chloride (OTMAC) and 4-dimethylaminopyridine (4-DMAP).62 The 13C MAS NMR results indicated that OTMAC, as an organic template agent, participates in the crystallization process of ultrathin FER zeolites. Its ammonium group is likely located within the pores and cages of the zeolite, synergizing with 4-DMAP to control the morphology and structure of SCM-37. J. D. Rimer et al. reported the synthesis of two-dimensional MWW zeolite nanosheets using a one-step approach.63 They employed the surfactant cetyltrimethylammonium (CTA) as a dual OSDA and exfoliating agent to produce MWW-type layers with an average thickness of 3.5 nm (∼1.5 unit cells). The use of CTA altered the Al sites without affecting the overall Si/Al ratio in the final product. This method offers a straightforward approach for the direct synthesis of 2D MWW-type materials, and is potentially applicable to other layered zeolites.

Ryoo et al. performed TEM, which revealed that each nanosheet is composed of alternating 2 nm-thick layers and 2.8 nm-thick surfactant layers, yielding an overall thickness of about 5 nm.21 The formation of both multilayer and single-layer MFI nanosheets was determined by the orientation and arrangement of surfactant molecules within the MFI framework. Qian et al. synthesized MFI zeolite nanosheets with a controllable morphology and interlayer spacing by adjusting the surfactant structure.64 TEM images revealed that surfactants with different chain lengths led to distinct nanosheet morphologies of MFI zeolite nanosheets. For instance, templates such as C6-12-diphe and C6-6-diphe produced multilayered lamellar structures, each composed of three pentasil zeolite layers separated by surfactant layers. In contrast, the C6-10-diphe template yielded nanosheets with an irregular and indistinct morphology, lacking clear layer stratification. Furthermore, an increase in the hydrophobic chain length of the surfactant resulted in a proportional increase in the interlayer spacing of the MFI zeolite nanosheets. This phenomenon arises from the geometric parameters of the surfactants, which dictate the formation of varying mesoporous structures. Fig. 7a presents the selected-area electron diffraction (SAED) pattern, exhibiting polycrystalline ring-shaped diffraction features consistent with the characteristic peaks of beta zeolite.26 The TEM (Fig. 7b) image revealed that the beta nanosheet is assembled from ∼16 nm-thick nanosheets, forming a three-dimensional ‘roof-tile’ architecture (∼800 nm). Moreover, high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) combined with electron tomography (Fig. 7c) further elucidated the internal ‘roof-tile’ morphology, consisting of interwoven nanosheets with mesoporous regions of low electron density at the core. Rimer et al. used cryogenic transmission electron microscopy (cryo-TEM) to observe the MWW nanosheet.63 Fig. 7f presents the bright-field high-resolution TEM image of randomly oriented MWW nanosheets in the d-MWW8.0 sample. These nanosheets predominantly exhibit an “edge-on” orientation (with the [001] perpendicular to the electron beam), appearing as narrow ribbon-like features. The enlarged view (Fig. 7g) showed two representative nanosheets, one with a unit-cell thickness and another with a bilayer configuration (two-unit-cell thickness). Fig. 7h reveals the high-resolution image of a planar region with the MWW zeolite crystal structure oriented along the (001) plane. For structural clarification, Fig. 7i superimposes the zeolite framework model onto a localized area of Fig. 7h, where the bright contrast regions are interpreted as the corresponding to zeolitic cage cavities. Fig. 7k and m display the STEM images of an MFI nanosheet8 and plate-like MOR,37 respectively. Wannapakdee et al. developed a facile and efficient approach for the synthesis of FAU zeolite nanosheets, utilizing an organosilane surfactant as a structure-directing agent.66 TEM analysis revealed that the FAU nanosheets exhibit a multi-layered structure, featuring narrow slits between the layers and surfactant-derived mesoporous cavities. Additionally, variations in the crystallization temperature and surfactant concentration enabled the fine-tuning of the FAU morphology, particle size, crystallite size, and porous structure characteristics.


image file: d5mh00579e-f7.tif
Fig. 7 TEM characterization of calcined beta-NS sample. (a) SAED pattern and (b) TEM image of a typical particle. (c) HAADF-STEM image of a representative particle. (d) Electron tomography reconstructed 3D volume of one particle. (e) Slices of the 3D reconstructed particle. Reprinted with permission from ref. 26. Copyright 2025, Wiley-VCH GmbH. HRTEM images of disordered MWW (gel Si/Al = 15, 8 wt% CTA). (f) Bright field image of a typical cluster of d-MWW8.0 nanosheets with a large number of sheets viewed edge-on and suitable for thickness measurements. (g) Magnified detail of panel A with two very thin sheets that are one-unit cell and two-unit cell thick. (h) High-resolution image of a sheet whose surface normal (001) is oriented parallel to the electron beam (a projection of the (001) plane). (i) Model of MWW structure superimposed on a portion of the image. (j) Defocus and information limit can be calculated from the Fourier transformation image. The rings coincide with the maxima of contrast transfer as a function of the reciprocal vector. Most of the diffraction spots are located close to these maxima, thereby enhancing image contrast. Reprinted with permission from ref. 63. Copyright 2020, the American Chemistry of Society. (k) SAED pattern and HRTEM (l) images of SPP viewed along the [010] and [001] zone axes. Reprinted with permission from ref. 8. Copyright 2022, the American Chemistry of Society. (m) STEM images of the cross-section of a plate viewed along the [001] direction of MOR crystal. Inset is the fast Fourier transform pattern; the orange spheres illustrate Si (or Al) atomic columns constructing the 12-MR and 8-MR pores with marked sizes. Reprinted with permission from ref. 37. Copyright 2023, Elsevier.

3.2 Structural analysis

Structural characterization of zeolite nanosheets, including their crystallinity, pore size, and surface area, is crucial for elucidating their physical properties. Powder X-ray diffraction (XRD) is widely used to characterize the structural properties of zeolites, particularly their crystallinity. The framework types of zeolites can be reliably distinguished based on their XRD patterns.

In contrast to 3D zeolites, nanosheet zeolites typically exhibit broad diffraction patterns, attributed to the absence of three-dimensional periodicity and disordered interlayer stacking. As shown in Fig. 8a, the 2D structure of the MWW family of zeolites can be identified by their characteristic peaks in the 2θ range of 6–10°, reflecting the interlayer spacing and interlayer orientation.67 As the size of the zeolite crystals decreases, their XRD patterns show broader interlayer (l00) diffraction peaks with only the in-plane (hk0) diffraction peaks retained. Meanwhile, a sharp reconfiguration of the 2D layer order in the (h0l) plane occurs, producing sharp peaks. Analogously, Fig. 8b shows the difference in the diffraction peaks between nanocrystal and plate-like ZSM-5. A sharp peak of h0l reflections can be observed, demonstrating that the thickness along the b-axis significantly decreased.


image file: d5mh00579e-f8.tif
Fig. 8 (a) XRD patterns of UTL zeolite and the products obtained after treatment in water. Reprinted with permission from ref. 67. Copyright 2011, the American Chemical Society. (b) XRD patterns of calcined plate-like HZSM-5 (top, Si/Al = 72) and commercial ZSM-5 nanocrystals (bottom, Si/Al = 74). Reprinted with permission from ref. 35. Copyright 2021, the American Chemical Society. (c) XRD patterns of MCM-22 (P) and its derivatives. Reprinted with permission from ref. 47. Copyright 1999, the American Chemical Society.

Corma and coworkers investigated the structural evolution of the layered MWW precursor during the post-synthesis process.47 MCM-36 was obtained by pillaring the swollen precursor materials, and the XRD pattern of MCM-36 (Fig. 8c) displayed the characteristic peaks of the MWW topological structure. The delamination of the layered precursor of the MCM-22 zeolite to form ITQ-2 produced an almost amorphous XRD pattern, indicating a significant reduction in long-range order. Compared to the MCM-22-type zeolite, the high-angle diffraction peaks of ITQ-2 are broader, consistent with its reduced crystallite size. In summary, MCM-36 retained the key features of the MWW topological structure, with an XRD pattern resembling that of MWW-type zeolites. In contrast, the XRD pattern of ITQ-2 was altered by the delamination of its layered structure.

3.3 Pore analysis

Generally, nanosheet zeolites exhibit a greater mesopore volume and larger external surface area compared to conventional zeolites. Nitrogen (N2) adsorption–desorption analysis is commonly employed as a reliable technique to determine the pore volume and surface area of zeolitic materials. For example, Qian et al. used the Barrett–Joyner–Halenda (BJH) model to analyze the pore size distribution of their samples.65 The synthesized MFI nanosheet samples exhibited an enhanced specific surface area, external surface area, and larger pore volume compared to the commercial ZSM-5 samples with a lower specific surface area but higher micropore area and volume. Furthermore, adjusting the structure of the surfactant, such as varying the hydrophobic chain length, could effectively control the pore size distribution of the layered MFI nanosheets. Similarly, Yan et al. reported that MFI nanosheets exhibited a large specific surface area and abundant mesoporous structure (366–665 m2 g−1), which is larger than that of conventional MFI zeolites (293 m2 g−1).68 In contrast, the traditional MFI zeolite does not exhibit a significant hysteresis loop, consistent with the absence of mesopores. Additionally, Liu et al. synthesized an MFI nanosheet zeolite via the pillaring method, which also exhibited a relatively higher mesopore volume than the conventional ZSM-5.69 Moreover, the pore size distribution of nanosheet zeolites varies significantly with their synthesis method, where the intercalation method is more efficient in generating mesoporous structures, whereas the dual-template approach produces more microporous structures.

Limtrakul et al. further found that the porous structural features of layered FAU nanosheets can be modulated by their synthesis conditions, including surfactant content and crystallization temperature. Specifically, as the surfactant content increases (TPOAC/Al2O3 ratio from 0.01 to 0.04), the specific surface area and pore volume of the samples gradually decreased (from 734 to 566 m2 g−1, and from 560 to 407 cm3 g−1, respectively).66 Corma et al. demonstrated that the nanosheet zeolite obtained by the delamination process (ITQ-2) exhibited higher total and mesoporous surface areas than both pillared MCM-36 and layered precursor MCM-22 samples.47

3.4 Acid analysis

The introduction of trivalent heteroatoms (e.g., Al3+, Fe3+, Ga3+, and B3+) into the zeolite framework generates a negative charge. When a proton balances the negative charge, a Brønsted acid site is formed within the zeolite framework. Dehydration of the Brønsted acid sites can convert them into Lewis acid sites. The acidic characteristics of zeolite catalysts, such as acid type, density, and strength, significantly impact the catalyst activity and product selectivity in various chemical reactions. Therefore, the effective characterization of zeolite acidity is crucial for understanding the structure–activity relationship of zeolite catalysts. To this end, numerous techniques have been developed to determine the acidity of zeolites both quantitatively and qualitatively.

Brønsted acidity in zeolites originates from their bridging Si–OH–Al groups, whereas Lewis acidity arises from their three-coordinated framework Al species, rather than tetrahedrally coordinated Al. The acidic properties of zeolites (e.g., strength and density) can be effectively characterized by ammonia temperature-programmed desorption (NH3-TPD), where weak acid sites correspond to desorption peaks at 373–673 K, while strong acid sites are assigned to the peaks in the range of 673–1073 K. Complementarily, the type of acidity can be analyzed by Fourier-transform infrared spectroscopy (FTIR). The Lewis acid sites (LAS) can be detected at 1454, 1487, and 1625 cm−1, and the Brønsted acid sites (BAS) can be observed at 1487, 1540, and 1634 cm−1.70 Furthermore, ammonia (NH3)-assisted 1H MAS NMR is employed to quantitatively analyze both the BAS and LAS concentrations. The coordination environment of aluminum (Al) within the zeolite framework can be investigated via 27Al MAS NMR. The peak at 55 ppm corresponds to tetrahedrally coordinated Al atoms integrated into the zeolite framework, while the peak at 0 ppm is associated with extra-framework Al species.71 The acidity properties can be enhanced by increasing the crystallinity of the zeolite (at the same Si/Al ratio), given that higher crystallinity results in more internal-framework Al sites and fewer external-framework Al sites. The internal-framework Al sites contribute more to the acid strength due to their better tetrahedral geometry.

Fig. 9a shows the NH3-TPD analysis of plate-like ZSM-5. All the samples exhibit peaks corresponding to weak and strong acid sites. The areas of these desorption peaks are consistent with the aluminum content, reflecting the acid site density. Pyridine adsorption infrared spectroscopy (Py-FTIR) identifies the acid types (Fig. 9b), revealing that all the samples contain Brønsted acid sites and a small amount of Lewis acid sites. Quantitative analysis indicates that the total acid amount of the samples is in the range of 361 to 388 μmol g−1, with the Brønsted acid sites being the dominant type.72,73 Wattanakit et al. investigated the effect of the synthesis precursors.74 The use of pure silica nanobeads as the precursor produced a significantly higher intensity at 0 ppm compared to aluminosilicate nanobeads. This suggests that hierarchical ZSM-5 nanospheres synthesized from aluminosilicate nanospheres incorporate more aluminum species into the zeolite framework, forming Brønsted acid sites, rather than remaining in the extra-framework state.


image file: d5mh00579e-f9.tif
Fig. 9 (a) NH3-TPD profiles of the zeolite samples. Reprinted with permission from ref. 72. Copyright 2023, Elsevier. (b) FTIR spectra of pyridine adsorption on different zeolite samples. Reprinted with permission from ref. 62. Copyright 2022, the American Chemical Society. (c) 31P MAS NMR spectra of bulk TMPO and TMPO-adsorbed H[Ga]MFI catalysts. Reprinted with permission from ref. 76. Copyright 2022, Elsevier. (d) 2D 27Al 3Q MAS NMR spectra, and the extracted 1D slices are shown as insets. Reprinted with permission from ref. 37. Copyright 2023, Elsevier.

In the case of heteroatom-substituted zeolites, their acid properties are determined by their heteroatom species. Yan et al. evaluated the acid strength of isomorphous MFI nanosheet zeolites (MFI (M), M = Al, Ga, and Fe).75 Compared to the ZSM-5 zeolite, the MFI-Ga zeolite exhibited a Brønsted-to-Lewis acid site ratio. However, the acidity and acid strength of the Fe-MFI zeolite were relatively low, resulting in an inferior catalytic performance towards the catalytic cracking of n-dodecane relative to ZSM-5. Dai et al. further investigated the acid strength and density of the plate-like H[Ga]MFI zeolite using TMPO-assisted 31P MAS NMR spectroscopy.76 As shown in Fig. 8c, the primary signal in the 31P MAS NMR spectrum at 64–83 ppm corresponds to the TMPO probe molecules bound to the Brønsted acid sites with varying acid strengths. Additionally, the weak signal at 47–49 ppm likely originates from the Lewis acid sites, extra-framework Ga species, or small GaOx clusters within the zeolite pores.

In contrast to the three-dimensional structure, the nanosheet structure is more likely to enhance the acid sites and improve their accessibility. Choi et al. employed 2,6-di-tert-butylpyridine as a probe to quantitatively measure the accessibility of the acid sites in TNU-9 and BEA zeolites via infrared spectroscopy.77 The extinction coefficient for the characteristic peak corresponding to the interaction between 2,6-di-tert-butylpyridine and Brønsted acid sites was determined. The results showed that the dealumination treatment markedly improved the accessibility of the external acid sites. Similar results were observed in other zeolite samples, such as MCM-22, ITQ-2, ZSM-5, and AS-824. Hensen et al. evaluated the total and external acidity of the MFI zeolite using FTIR with pyridine and 2,4,6-collidine as probe molecules.73 The Brønsted acid sites in the nanosheet zeolites exhibit comparable strength to the bulk H-ZSM-5, but the BAS concentration on the external (mesoporous) surface was higher. In addition, single-layer zeolites possess higher external BAS and silanol group concentrations than multi-layer zeolites.

The ultrathin morphology of nanosheet zeolites not only significantly increases their external surface area (reaching up to ∼50% of the total) but also shortens the diffusion paths. Consequently, a higher proportion of acid sites (particularly Lewis acid sites) becomes accessible compared to bulk zeolites. These features greatly promote the processing of bulky molecules. For instance, methanol conversion over ZSM-5 nanosheets achieved enhanced conversion along with significantly reduced coking. The introduction of mesoporosity through pillaring further improved the mass transport.78 Further details on the applications of nanosheet zeolites are discussed in the following section.

4. Catalytic applications of nanosheet zeolites

Nanosheet zeolites, with their crystal size in one dimension reaching the nanometer scale or even the thickness of a single or a few unit cells, dramatically shortening the diffusion paths. Furthermore, this morphological feature endows nanosheet zeolites with a more open pore structure and better accessibility to acid sites. Owing to these structural advantages, nanosheet zeolites exhibit excellent catalytic performances in diverse reactions, such as methanol conversion, cracking, carbonylation, isomerization, alkylation, and oxidation. Table 1 summarizes representative examples of their applications across various reactions.
Table 1 Applications of nanosheet zeolites in catalytic reactions
Catalysts Physico-chemical properties Catalytic reaction Conversion (%) Selectivity (%) Ref.
Si/M n(B/L) St (m2 g−1) Vt (cm3 g−1)
Note: Si/M (M = Al, Ga, Fe, Ti); n(B/L) = n(Brønsted[thin space (1/6-em)]acid[thin space (1/6-em)]sites)/n(Lewis[thin space (1/6-em)]acid[thin space (1/6-em)]sites) mmol g−1. St = surface areatotal(internal,[thin space (1/6-em)]external); Vt = volumetotal(mirco,[thin space (1/6-em)]meso).
NS ZSM-5 55 5.57 537 0.640 Methanol to propylene 98.1 78.1 (C=2–C=4) 78
B-CN ZSM-5 55 7.75 485 0.482 96.2 77.3 (C=2–C=4)
W-ZSM-5 236 4.3 654 0.453 Methanol to propylene 92 53.82 (propylene) 80
H[Ga]-MFI-50 65 441 0.22 Methanol to propylene >99 49 (propylene) 76
H[Ga]-MFI-100 142 411 0.23 >99 51 (propylene)
H[Ga]-MFI-150 224 413 0.29 >99 52 (propylene)
Bulky-ZSM-5 149 2.82 496 0.50 Methanol to propylene 99.4 5.4 (ethylene)/38.7 (propylene)/23.3 (butylenes) 81
Lam-ZSM-5 141 1.55 447 0.22   99.8 4.9 (ethylene)/40 (propylene)/23.8 (butylenes)  
NS ZSM-5 48 1.38 404 0.099 Cracking of n-decane 92 6.3 (ethylene)/9.0 (propylene)/5.9 (aromatics) 85
Dual-template NS ZSM-5 51 1.16 478 0.092   92 8.5 (ethylene)/16.9 (propylene)/12.4 (aromatics)  
MFI-Al 43 479 1.05 Cracking of n-dodecane   11.7 (ethylene)/14.2 (propylene)/1.2 (aromatics) 86
MFI-Fe 46 394 0.72     8.4 (ethylene)/10.3 (propylene)/0.9 (aromatics)  
MFI-Ga 48 452 0.87     9.1 (ethylene)/17.7 (propylene)/0.8 (aromatics)  
Al-ECNU-7 31 512 Cracking of TIPB 71 38 (DIPB)/37 (IPB)/17 (BZ) 90
MCM-22 28 452   23 63 (DIPB)/27 (IPB)/7 (BZ)  
Lam-ZSM-5 30 0.74 393 0.315 Isomerization of xylene 34 4.8 (p-xylene)/57.8 (m-xylene) 92
Pillared-ZSM-5 36 0.87 472 0.388   37 24.7 (p-xylene)/74.1 (m-xylene)  
De-pillared-ZSM-5 53 1.04 429 0.278   26 26.6 (p-xylene)/72.7 (m-xylene)  
Pillared-ZSM-5 56.3 0.843 405 0.30 Benzylation of mesitylene with benzyl alcohol ∼15 ∼70 (2-benzyl-1,3,5-trimethylbenzene) 93
ITQ-2 13.2 0.49 582 0.77 Aromatization of acetylene >50 ∼35 (BTX) 94
Ga-ZSM-5 38 1.96 508 1.03 Aromatization of n-pentane 40 45 (BTEX) 95
NS-mordenite 5.9 514 0.389 Carbonylation of dimethyl ether 42 98 (methyl acetate) 60
Fe/ZSM-5 41 475 0.95 Benzene oxidation to phenol 23.44 >99 (phenol) 108
C-TS-1 108 464 0.236 Phenol oxidation 44 8 (benzoquinone)/49 (catechol)/43 (hydroquinone) 105
M-TS-1 111 410 0.308   35 48 (benzoquinone)/28 (catechol)/24 (hydroquinone)  
P-TS-1 147 595 0.521   35 7 (benzoquinone)/57 (catechol)/36 (hydroquinone)  
NS-Ti-MFI 69 590 0.75 Epoxidation of cyclohexene 23.3 75.6 (epoxide, cyclohexene) 106
Epoxidation of cyclooctene 18 75.9 (epoxide, cyclooctene)
Plate-like TS-1-ly 58 443 0.21 Epoxidation of 1-hexene 23 95 (epoxide) 26
C-TS-1 54 394 0.18   17 91 (epoxide)  


4.1 Methanol conversion

Methanol can be converted into hydrocarbons to produce fuel via the methanol-to-hydrocarbons (MTH) process. Based on the product distribution, the MTH reaction can be categorized into three types including methanol-to-olefins (MTO), methanol-to-gasoline (MTG), and methanol-to-aromatics (MTA). Initially, methanol undergoes dehydration to form dimethyl ether (DME) over the Brønsted acid sites. The methanol and DME mixture reaches an equilibrium state, and subsequently transforms into hydrocarbons via further dehydration.79 Nanosheet zeolites, with their shortened diffusion paths and increased external acid site exposure, significantly enhance the reaction kinetics and suppress coke formation during methanol conversion. Their unique structure facilitates the faster access of reactants to the active sites, improving the selectivity towards light olefins and aromatics.80,81

The MTO reaction, producing ethylene and propylene, widely utilizes SAPO-34 with a CHA framework as a highly efficient catalyst owing to its high selectivity for light olefins (over 80% C2–C4 olefins). Nevertheless, the commercial SAPO-34 catalysts suffer from rapid deactivation due to coke accumulation within their micropores. This limitation can be mitigated by employing SAPO-34 nanosheets. Yu et al. found that nanosheet SAPO-34 with a thickness of 20 nm exhibited an extended catalyst lifetime along with reduced coke deposition rate in the MTO reaction.82 A decrease in crystal size improved the mass transfer, thereby reducing coke deposition. Moreover, SAPO-34 and SAPO-18 share similar double 6-membered ring layer structures, which facilitate the intergrowth of the two frameworks and enhance the catalyst performance. Wu et al. reported that an 11-nm-thick nanosheet MOR zeolite achieved markedly improved ethylene selectivity (42.1%) in the MTO reaction compared to conventional bulk MOR crystals (3.3%).25 The nanosheet mordenite catalyst featured highly exposed (010) crystal planes and 8-membered ring channels, not only enhancing the accessibility of methanol molecules to these channels, but also increasing the diffusion rate of ethylene molecules along the b-axis, leading to heightened ethylene selectivity.

Dou et al. demonstrated that the nanosheet MFI zeolite displayed an excellent catalytic performance in the MTP reaction, with high propylene selectivity (51.0%), high propylene/ethylene ratio (12.1), and long catalyst lifetime (240 h).55 Recently, Dai et al. synthesized plate-like MFI zeolites via a fluoride-assisted approach, achieving a remarkably long service life of 750 h in the MTP reaction.72 The significant improvement in stability originated from the unique morphological features of the zeolite. As mentioned above, the plate-like ZSM-5 possessed a higher crystal surface aspect ratio, resulting in a higher surface fraction of (010) than the conventional form. Therefore, methanol is more likely to reach the acidic sites of MFI nanosheets through the straight channels. In addition, the use of sheet-like ZSM-5 catalysts can shorten the reactant diffusion path, effectively suppressing coke formation and improving the stability and recyclability of the catalyst. Dai et al. designed plate-like H[Ga]MFI zeolites with controllable acidity, which exhibited a high performance in methanol-to-propylene (MTP) reactions.76 As illustrated in Fig. 9, the catalyst lifetime of the plate-like H[Ga]MFI zeolite initially increased, and then declined with a decrease in BAS density (Fig. 10b–f). Notably, the H[Ga]MFI-150 catalyst with an optimal BAS density of 0.10 mmol g−1 demonstrated the longest catalyst lifetime of 252 h (Fig. 10d). In contrast, a slight reduction in BAS density from 0.10 to 0.07 mmol g−1 in the H[Ga]MFI-200 catalyst led to a pronounced decrease in catalyst lifetime from 252 h to 43 h (Fig. 10e). This underscores the critical importance of precisely modulating the BAS density in plate-like H[Ga]MFI zeolites for optimizing their stability as catalysts in methanol-to-propylene conversion. In the case of H[Ga]MFI catalysts with comparable acid strength, the BAS density exerted a significant influence on the product distribution. As the BAS density decreased from H[Ga]MFI-50 to H[Ga]MFI-200, the selectivity towards ethylene and aromatics declined sharply (Fig. 10b–e and g), whereas the selectivity towards long-chain alkenes (e.g., C5–C8 alkenes) increased substantially. This phenomenon is likely attributed to the enhanced olefin-based cycle facilitated by H[Ga]MFI catalysts with a lower BAS density. Simultaneously, some long-chain alkenes may diffuse out of the MFI channels, contributing to the detected products. A direct comparison of the C4–C8 alkene selectivity (associated with the olefin-based cycle) and aromatic selectivity (linked to the aromatic-based cycle) across catalysts with varying BAS densities is presented in Fig. 10h. Apparently, the aromatics selectivity correlated positively with the BAS density, whereas the C4–C8 alkenes selectivity exhibited the inverse relationship.


image file: d5mh00579e-f10.tif
Fig. 10 Methanol conversion and product distribution over plate-like H[Ga]MFI-50-T (a), H[Ga]MFI-50 (b), H[Ga]MFI-100 (c), H[Ga]MFI-150 (d) and H[Ga]MFI-200 (e) in MTP conversion at 748 K with a WHSV of 3.0 h−1. (f) Propylene/ethylene ratios and catalyst lifetime over different H[Ga]MFI catalysts during the MTP conversion at 748 K with a TOS of 20 h. (g) Comparison of the product distribution over different plate-like H[Al]MFI and H[Ga]MFI zeolites in MTP conversion at 748 K with a TOS of 20 h. (h) Selectivity of C4–C8 alkenes and aromatics as a function of BAS density over different plate-like H[Al]MFI and H[Ga]MFI zeolites in MTP conversion at 748 K with a TOS of 20 h. Reprinted with permission from ref. 76. Copyright 2022 Elsevier.

Niaei et al. reported the synthesis of M-substituted (M: Mn, Ce, W) MFI zeolite nanosheets and their catalytic performance towards the methanol-to-propylene reaction.80 The results showed that the W-substituted MFI nanosheets exhibited the best performance, achieving complete methanol conversion with a propylene selectivity of 55.7%, total light olefin selectivity of 88% and catalyst lifetime of 81 h. Luo et al. explored the influence of two different types of ZSM-5 crystals as seeds on the morphology, pore structure, acidity, and defect sites of synthesized ZSM-5 zeolites, and investigated their catalytic performance in the MTP reaction.83 The result showed that the ZSM-5 samples derived from hydrothermally treated spherical seeds exhibited a longer lifetime and higher propylene selectivity compared to that from plate-like seeds. This is because the hydrothermally treated samples possess suitable acidity, accelerated diffusion rate, and reduced framework defects. Kim et al. investigated the influence of MFI nanosheet thickness and Si/Al ratios on the MTP reaction.84 The ultrathin ZSM-5 nanosheet (2.5 nm) exhibited better catalytic activity and propylene selectivity compared to the 7.5-nm-thick ZSM-5 nanosheet and commercial ZSM-5 zeolites. The optimal performance was observed at an Si/Al ratio of 500, which is attributed to the appropriate acid density and facile diffusion of molecules in the ZSM-5 nanosheet zeolites. Ryoo et al. found that the MFI nanosheets exhibited a longer catalytic lifetime in the methanol-to-gasoline reaction relative to traditional MFI zeolites.21 In the MTG reaction, the catalytic conversion rate curve of MFI nanosheets over time is significantly better than that of traditional MFI zeolites. This is because the coke on MFI nanosheets is primarily deposited on their external surface (mesopores) rather than in their internal micropores. The coke in the internal micropores is more effective at causing catalyst deactivation, while external coke has a relatively smaller obstruction to diffusion.

4.2 Catalytic cracking

The fluid catalytic cracking (FCC) process is pivotal refining technology for converting heavy feedstocks into lighter and more valuable products, such as gasoline, diesel, and petrochemical feedstocks. Aluminosilicate zeolites have been extensively employed in FCC units due to their unique structural properties, which are a key factor in achieving high activity and light olefin selectivity. However, the cracking of long-chain hydrocarbons inevitably leads to coke formation. The accumulation of coke can block the active sites of zeolites, thereby deactivating the catalyst, but nanosheet zeolites can efficiently mitigate this problem. In catalytic cracking, nanosheet zeolites mitigate secondary reactions and coke deposition owing to their ultrathin morphology, high external surface area, and improved mass transport properties. These structural advantages lead to a higher reaction rate and improved product selectivity, especially towards light olefins.85,86

In the catalytic cracking of long-chain alkanes, the crucial factors influencing the product selectivity are the textural properties and acidity of zeolites. Liu et al. reported the performance of pillared ZSM-5 nanosheet zeolites in n-decane catalytic cracking for the production of light olefins.87,88 The pillared structure stabilized the hierarchical porosity and preserved the interlayer mesopores, which facilitated product diffusion and restricted secondary reactions that consume light olefins. The pillared ZSM-5 nanosheets displayed a remarkable light olefin selectivity of up to 37.8% at 500 °C, with an n-decane conversion of ∼92%, nearly twice as high as that of the conventional ZSM-5. Moreover, the pillared structure exhibited exceptional anti-coking stability, with the deactivation rate of the conventional ZSM-5 zeolite reaching 68.24%, which was more than 11-times higher than that of the ZSM-5 nanosheets (5.81%). Recently, they synthesized pillared nanosheet H-ZSM-5 for the catalytic cracking of supercritical n-dodecane. Compared to the parent nanosheet H-ZSM-5 zeolite, the pillared nanosheet zeolite presented a higher surface area, mesoporous volume, and Brønsted acid site accessibility, which led to an improved catalytic performance with a conversion of 76.8%, TOF value of 130.92 s−1, and deactivation rate of 9.11%. Xiao et al. synthesized aluminosilicate ITH zeolite using a cationic oligomer as the organic template.89 In the vacuum gas oil cracking reaction, the Al-ITH zeolite enhanced the propylene and butene yields without increasing the yield of saturated propane and butanes, thereby improving the olefin/paraffin ratio in the C3 and C4 fractions.

The large external surface area and strong acidity of zeolites provide abundant active sites for the reaction of large-molecule compounds. Nanosheet zeolites also exhibit an excellent performance in cracking large molecules. Wu et al. developed ECNU-7P, an MWW-type zeolite with a layered porous structure, which was prepared via the self-assembly of the CTAB surfactant and zeolite precursors.90 As shown in Fig. 11a, compared to the traditional MCM-22, Al-ECNU-7 exhibited about threefold higher conversion in cracking 1,3,5-triisopropylbenzene (TIPB), together with a significantly slower deactivation rate during the reaction process. Moreover, Al-ECNU-7 also delivered a superior cracking performance in terms of product distribution. Dai et al. synthesized plate-like ZSM-5 zeolites with a reduced b-axis thickness. As shown in Fig. 11b, in the cracking reactions of large hydrocarbon molecules such as propyl-benzene and 1,3,5-triisopropylbenzene, the plate-like ZSM-5 zeolites showed obvious advantages, which are attributed to the improved accessibility to their acid sites and reduced diffusion limitations.91


image file: d5mh00579e-f11.tif
Fig. 11 (a) Catalytic cracking of TIPB over (left) Al-ECNU-7 (Si/Al = 15) and (right) MCM-22 (Si/Al = 16) catalysts. Reaction conditions: cat., 0.2 g; feed rate, 1.7 mL h−1; N2, 30 mL min−1; temp., 573 K. Reprinted with permission from ref. 90. Copyright 2016, the American Chemical Society. (b) Conversion of TIPB with time on stream (left), and the correlation between the initial conversion of TIPB and external acid sites over the zeolite samples (right). Reprinted with permission from ref. 91. Copyright 2022, the American Chemical Society.

4.3 Isomerization

Isomerization is a crucial catalytic process in the petrochemical and fine chemical industries, involving the rearrangement of molecular structures without altering their atomic composition. In industrial applications, isomerization enhances the octane number of gasoline, produces valuable branched alkanes, and converts linear alkenes into more reactive isomers. Traditional zeolite catalysts, such as ZSM-5 and beta, have been widely employed in isomerization reactions due to their shape-selectivity and acidity. However, their microporous nature often limits the diffusion of bulky molecules, leading to lower activity and rapid deactivation. Nanosheet zeolites provide superior diffusion and accessibility to confined acid sites, which is especially beneficial for the isomerization of bulky molecules. Their highly open structure effectively minimizes pore blockage and allows better control over reaction pathways, leading to improved catalytic efficiency.92,93

Choi et al. evaluated the performance of acetylene aromatization over ITQ-2 and the traditional microporous MCM-22 zeolite.94 Compared to MCM-22, ITQ-2 produced more C3–C8 aliphatic hydrocarbons than aromatic hydrocarbons. This arises from the larger external surface area of ITQ-2, which shortened the reactant residence time and promoted the formation of aliphatic hydrocarbons. Besides, ITQ-2 exhibited superior resistance to deactivation than MCM-22. Wattanakit et al. demonstrated that Ga-modified HZSM-5 nanosheets significantly boosted the catalytic performance in C5 hydrocarbon aromatization, with the n-pentane conversion and BTEX (benzene, toluene, ethylbenzene, and xylene) selectivity exceeding 40% and 45%, respectively.95 The layered structure facilitates the migration and dispersion of Ga species, promoting the formation of tetrahedral Ga species (up to 43%) and enhancing the catalytic activity. Ryoo et al. investigated the effect of the zeolite crystal thickness on the n-heptane hydro-isomerization over the Pt/MFI zeolite.96 The results showed that reducing the zeolite crystal thickness to the nanoscale markedly improved the selectivity for branched isomers, which is attributed to the shorter diffusion path lengths, allowing the branched products to escape before undergoing cracking.

FER zeolites are widely employed in the isomerization of n-butene to isobutene due to their sheet-like morphology. Xu et al. analyzed the catalytic performance of FER nanosheets with sizes in the range of 2 μm to 100 nm,97 revealing that nano-scale FER presented better stability than micro-scale FER during the reaction due to its larger external surface area and reduced diffusion limitations. Nano-scale FER exhibited a coking amount of 10.6% after 72-h reaction, whereas micro-scale FER showed only 6.6% after 48-h reaction, revealing the superior coke resistance of nano-scale FER. Similarly, adjusting the morphology of the FER nanosheets could enhance their performance. Dai et al. reported the synthesis of nanorod-like FER zeolites.98 The nanorod-like FER zeolite exhibited a significantly reduced diffusion path along the c-axis, resulting in a larger external surface area and better accessibility to acid sites compared to its micron-sized counterpart. These improved properties endowed the nanorod-like FER zeolite with high selectivity and low deactivation rates in n-butene isomerization. In situ FTIR spectroscopy (Fig. 12d) revealed that the conventional FER (FER-F-0) exhibited two additional characteristic absorption bands at 1650 and 1460 cm−1 compared to nanorod-like FER (FER-F-0.2). The peak at 1650 cm−1 is assigned to π-coordinated butene/oligomeric olefin species, while the band at 1460 cm−1 corresponds to the C–H vibrations of CH2/CH3 groups, indicating the formation of highly branched aromatic compounds on the FER-F-0 catalyst. These bulky aromatic species rapidly block the accessibility of the acid sites to 1-butene molecules, thereby leading to catalyst deactivation.


image file: d5mh00579e-f12.tif
Fig. 12 Catalytic performances of 1-butene skeletal isomerization on as-synthesized FER-F-0 and FER-F-0.2 samples: (a) 1-butene conversion; (b) isobutene selectivity; (c) isobutene yield; and (d) in situ FTIR spectra recorded during the 1-butene skeletal isomerization on FER zeolite samples up to TOS = 120 min at the reaction temperature of 400 °C. Reprinted with permission from ref. 98. Copyright 2021, The Royal Society of Chemistry. (e) Top: Reaction scheme for converting GLU into FRU. Bottom: FRU yield versus GLU conversion and FRU yield versus reaction time over different catalysts. (f) Top: Isomerization of LAC to LACTU in methanol. Bottom: LACTU yield versus LAC conversion and LACTU yield versus reaction time over different catalysts. Reprinted with permission from ref. 99. Copyright 2015, Wiley-VCH GmbH.

Isomerization is also crucial in biomass conversion to obtain the target products. Tsapatsis et al. reported that the single-unit-cell Sn-MFI zeolite nanosheet displayed a superior catalytic performance in glucose and lactose isomerization reactions,99 with Sn sites uniformly distributed and exclusively located at the framework sites. As illustrated in Fig. 12e, the self-pillared Sn-MFI nanosheet achieved ∼65% fructose (FRU) yield from glucose (GLU) in ethanol solvent, outperforming traditional USY and Al-BEA. In the conversion of lactose (LAC) to lactulose (LACTU), the self-pillared Sn-MFI nanosheet also demonstrated an excellent catalytic performance, achieving 31% LAC conversion and 97% LACTU selectivity in methanol solvent, significantly surpassing traditional Sn-BEA and Sn-MCM-41. These exceptional performance stems from the uniform distribution of framework Sn sites, which exhibit both Lewis and Brønsted acidity, as well as unique nanosheet structural features, providing efficient diffusion channels and accessibility to active sites for large molecular substrates.

4.4 Alkylation

Alkylation and acylation are vital organic synthesis reactions widely used in the petrochemical, fine chemical, and pharmaceutical industries. The choice of catalyst is critical for improving the catalytic efficiency, selectivity, and environmental impact of these processes. Therefore, a balance between the Brønsted/Lewis acid ratio must be achieved to obtain the optimal catalytic performance. The nanosheet architecture of zeolites enables acid site accessibility and minimized steric constraints, which are crucial for the alkylation of aromatic hydrocarbons. These structural advantages deliver superior activity and selectivity, while simultaneously improving the catalyst stability under industrial conditions.

Wattanakit et al. fabricated hierarchical nano-spherical ZSM-5 zeolites using uniform aluminosilicate (AS) nanobeads as the starting material.74 This approach enabled better control of the distribution of Al in the ZSM-5 framework. AS nanospheres provided both silicon and aluminum sources, ensuring that most of the Al species were uniformly embedded into the zeolite framework in tetrahedral coordination. The uniform Al distribution generated a higher density of Brønsted acid sites, which facilitated efficient ethylene transfer from ethanol dehydration to the adjacent acid sites, thereby promoting the alkylation reaction. Consequently, the AS-derived hierarchical ZSM-5 nanosheets presented an improved benzene conversion (60%) and ethylbenzene selectivity (62%) compared to the conventional ZSM-5.

Fang et al. synthesized hierarchically structured MFI nanosheet assemblies, which demonstrated an extraordinary catalytic performance in the benzylation of benzyl alcohol with mesitylene.100 As shown in Fig. 13b, compared to the conventional ZSM-5, the layered MFI zeolite nanosheet showed a higher benzyl alcohol conversion and excellent cyclic stability, maintaining a conversion of 70.1% even after 6 cycles. In contrast, the conventional ZSM-5 zeolite was nearly deactivated after 6 cycles. Notably, the selectivity towards 2-benzyl-1,3,5-trimethylbenzene increased with a decrease in the Si/Al ratio, highlighting the pivotal role of the zeolite acidity in governing the product selectivity. Moreover, the spatial distribution of Al species determined both the reaction activity and product selectivity.


image file: d5mh00579e-f13.tif
Fig. 13 (a) Scheme for allylation (kalk) and self-etherification (keth) over MFI zeolite nanosheet assemblies; (A) mesitylene, (B) benzyl alcohol, (C) 1,3,5-trimethyl-2-benzylbenzene and (E) dibenzyl ether. (b) Catalytic benzylation of mesitylene and benzyl alcohol over hierarchically structured MFI zeolite nanosheet assembly conversion of benzyl alcohol vs. reaction time and product selectivity over a reaction time of 20 h. Reprinted with permission from ref. 100. Copyright 2019, Elsevier. (c) Model reaction used to assess the performance of MWW-type catalysts. (d) Benzyl alcohol (BA) conversion (blue, left axis) at 80 °C (30 mg of catalyst per 8.86 g of reaction mixture; benzene/BA = 196 mol mol−1). Reprinted with permission from ref. 65. Copyright 2020, the American Chemical Society.

Rimer et al. employed the surfactant cetyltrimethylammonium (CTA) as both a dual OSDA and exfoliating agent to produce MWW-type layers with an average thickness of 3.5 nm, termed H-d-MWW. As shown in Fig. 13d, the obtained H-d-MWW exhibited enhanced catalytic activity compared to the three-dimensional MWW (MCM-22) and 2D layers prepared by post-synthesis exfoliation (ITQ-2).63 During the Friedel–Crafts alkylation reaction, the conversion of benzyl alcohol over H-d-MWW was significantly faster than that on other catalysts. The comparison of the turnover numbers indicated that d-MWW showed similar catalytic activity to MCM-22. Simultaneously, the post-synthesis exfoliation process used to prepare 2D nanosheets, such as ITQ-2, introduced substantial structural defects, including extra-framework aluminum and distorted aluminum sites, which significantly compromised their catalytic performance.

Limtrakul et al. developed a simple and efficient method to prepare self-pillared FAU zeolite nanosheets.66 In the benzylation of toluene with benzyl chloride, the conventional FAU zeolite catalyst exhibited limited initial activity (8.78% conversion within 6 h), whereas the FAU nanosheet catalyst achieved a 3–4-times enhancement in conversion under identical conditions. This is because the conventional FAU zeolite has microporous characteristics, which limit the diffusion path for reactions involving large molecules. Yang et al. reported a novel method for the direct synthesis of partially and fully delaminated MWW-type zeolites (SCM-1).101 In the liquid-phase alkylation of benzene with ethylene, the SCM-1 nanosheets displayed nearly a two-fold increase in ethylene conversion compared to the conventional MCM-22 zeolite. Remarkably, the conventional MCM-22 zeolite suffered significant activity loss after 64-h reaction, while SCM-1 maintained a stable ethylene conversion with less adsorbed organic matter and carbon deposition. This enhanced catalytic performance is attributed to the improved mass transfer properties provided by the delaminated nanosheet structure.

4.5 Carbonylation

The carbonylation reaction involves introducing carbon monoxide (CO) into organic molecules, replacing their substituents (e.g., alkyl, hydrogen, etc.) with carbonyl (C[double bond, length as m-dash]O) groups. Typically, dimethyl ether (DME) carbonylation to methyl acetate (MA), followed by hydrogenation is a novel route for producing ethanol. In carbonylation reactions, the diffusion limitations of conventional zeolites are alleviated by the open framework of nanosheet structures. This facilitates the accelerated migration of reactants and intermediates, thereby boosting the turnover frequency and selectivity for carbonylated products.

Yang et al. synthesized the hierarchical nanosheet H-ZSM-35 zeolite, which outperformed conventional ZSM-35 in DME carbonylation.102 Li et al. developed a templating agent-free strategy to synthesize H-mordenite (H-MOR) nanosheet assemblies,60 which exhibited high catalytic activity and stability in the carbonylation of DME to MA. They found that decreasing the basicity of the synthesis gel by dilution with water promoted the growth of H-MOR nanocrystals, which assembled into nanosheet bundles. The resulting expanded external surface area was beneficial for the mass transfer of the reactants and products, suppressing the formation of hard coke in the micropores. Shen et al. fabricated MOR zeolite nanosheets with a thickness of 20–40 nm.103 During the carbonylation of dimethyl ether, the DME conversion reached 36.7% after 4 h on the nanosheets, versus merely 14.6% after 6 h on the micro-sized sample. The marked enhancement is ascribed to the shortened diffusion distances in the 8-MR and 12-MR channels, which accelerated molecular transport and improved the accessibility of the acid sites to the reactant molecules. More recently, they also studied the identification of the active sites on a plate-like mordenite (MOR) zeolite for DME carbonylation. As shown in Fig. 14a, the results show that the T3 site in the 8-MR channel serves as the active center, exhibiting unprecedented activity.37


image file: d5mh00579e-f14.tif
Fig. 14 (a) Reaction pathway of DME carbonylation on zeolites, (b) DME carbonylation over the plate-MOR and conventional-MOR at 473 K and (c) 12-MR and 8-MR pores connected via the 8-MR side pocket, and schematic of the 8-MR side pocket viewed from the y/y orientations. Reprinted with permission from ref. 37. Copyright 2023, Elsevier. (d) Conversion of DME over different zeolites. (e) MA selectivity over different zeolites. (f) Selectivity of MeOH, CO2 and CH4 over different zeolites. Reprinted with permission from ref. 104. Copyright 2021, the American Chemical Society.

Tsubaki et al. reported the development of a carbonylation catalyst, Al-RUB-41 zeolite, featuring a unique nanosheet structure, which remarkably enhanced the DME carbonylation efficiency.104 As shown in Fig. 14b, the distinctive nanosheet structure of Al-RUB-41, with the 8-MR channels oriented perpendicular to the thin sheets, facilitated the efficient mass transfer of the reactants and products, achieving >95% MA selectivity, which outperformed the commercial H-MOR and H-ZSM-35 catalysts. The deactivation of the Al-RUB-41 catalyst is attributed to its external surface acid sites, where larger coke precursor species (e.g., alkylated naphthalene) preferentially accumulate, rather than within the 8-MR channels. Coating the Al-RUB-41 with a silica layer (Al-RUB-41@SiO2) effectively removed its external acid sites, thereby preventing the formation of hard coke precursors and enhancing the long-term stability and efficiency of the catalyst.

4.6 Oxidation

The application of zeolites in oxidation reactions is typically coupled with metal components. Although conventional microporous ferro-silicate and titano-silicate zeolites demonstrate high efficiency in oxidation processes, their microporous structure can induce diffusion limitations, resulting in rapid coke deposition. Nanosheet zeolites exhibit superior accessibility to redox-active sites and improved dispersion of metal species due to their high external surface area, which facilitates electron and mass transport during oxidation reactions, thereby demonstrating higher activity and suppressed deactivation.105,106

Gao et al. evaluated the catalytic performance of Cu–ZSM-5 nanosheets for the decomposition of N2O.107 The results showed that the Cu–ZSM-5 nanosheets were quite stable in a 50 h stream-online test, maintaining ∼80% conversion at 475 °C without deactivation. In contrast, the N2O conversion over conventional Cu–ZSM-5 decreased gradually from the initial 74% to 54% after 50 h. The CO-IR characterization showed that the Cu+ species in the spent Cu–ZSM-5 nanosheets remained almost unchanged compared to the fresh Cu–ZSM-5, while the amount of Cu+ species in the spent conventional Cu–ZSM-5 decreased obviously. This is likely due to the oxidation of a portion of Cu+ by O2 generated from N2O decomposition at 475 °C. The O2-TPD profiles revealed that oxygen desorption from the Cu sites on the Cu–ZSM-5 nanosheets occurred more easily than on the conventional Cu–ZSM-5, which further enhanced the N2O decomposition reaction. Hensen et al. synthesized Fe/ZSM-5 nanosheet zeolite catalysts for the oxidation of benzene to phenol using N2O as the oxidant.108 By decreasing the Fe content and b-direction crystal thickness of the MFI nanosheets, consecutive phenol reactions were suppressed, improving the catalyst stability. As illustrated in Fig. 15a, the Fe/ZSM-5 nanosheets, with 0.24 wt% Fe content and a crystal size of ∼3 nm in the b-direction, achieved a phenol yield of 185 mmol g−1 in 24 h, the highest observed yield. Compared to bulk Fe/ZSM-5, the nanosheet catalyst showed significantly reduced deactivation due to the absence of diffusion limitations and lower Fe agglomeration. For instance, the phenol yield of Fe/ZSM-5 nanosheets was 6.0 mmol g−1 h−1 after 24 h, while the bulk Fe/ZSM-5 only produced 0.6 mmol g−1 h−1 phenol yield.


image file: d5mh00579e-f15.tif
Fig. 15 (a) Scheme of the oxidation of benzene to phenol; phenol formation rate as a function of time on stream for bulk and nanosheet Fe/ZSM-5 zeolites synthesized using (b) C22-6-3·Br2, (c) C22-6-6-6-3·Br4, and (d) TPAOH. Reprinted with permission from ref. 108. Copyright 2013, the American Chemical Society. (e) and (f) Catalytic performances of 1-hexene epoxidation. Reaction conditions: 10 mmol 1-hexene, 0.05 g catalyst, 10 mL CH3OH, 10 mmol H2O2 (30 wt%), 60 °C, 2 h. Reprinted with permission from ref. 36. Copyright 2024, Elsevier. Dependence of CHE conversion (g), CHO selectivity (h), and H2O2 efficiency (i) with reaction time over Ti-beta-NS and Ti-beta-C. Reaction conditions: 333 K, catalyst, 50 mg; CHE, 10 mmol; H2O2 (30 wt%), 10 mmol; MeCN, 10 mL. Reprinted with permission from ref. 26. Copyright 2025, Wiley.

Zhang et al. developed a sandwich-structured Pt@ZSM-5 nanosheet catalyst, prepared by controllably intercalating Pt nanoparticles (∼4.3 nm) between ZSM-5 single-layer sheets.109 The Pt nanoparticles confined between the nanosheet layers displayed better dispersion and higher thermal stability compared to conventional Pt/ZSM-5 catalysts during toluene combustion. Although the Pt@ZSM-5 catalyst showed slightly higher coke formation (1.36 wt%) than Pt/ZSM-5 (1.13 wt%), its large external surface area and stabilized mesopores enhanced its anti-coking ability and coke tolerance.

Ma et al. reported the synthesis of hierarchical titanium silicalite-1 (TS-1) nanosheets,110 which exhibited a unique house-of-cards-like structure with stable interlayer mesopores and macropores, enhancing the accessibility of the active Ti sites for the epoxidation of bulky cyclic olefins (e.g., cyclohexene and cyclooctene). The catalytic performance of the TS-1 nanosheets surpassed that of the conventional microporous and mesoporous TS-1 catalysts due to their larger external surface area and improved mass transfer. Additionally, post-fluoride treatment further increased their hydrophobicity and catalytic activity. Ryoo et al. synthesized titanosilicate MFI nanosheets with a single-unit-cell thickness as efficient oxidation catalysts for bulky molecular epoxidation, using peroxides such as H2O2 or t-butyl hydroperoxide.111 The conversion of the bulk TS-1 catalyst for large-ring olefins was very low because these large molecules could not enter the 10-membered ring pores of TS-1. In contrast, the nanosheet-like TS-1 exhibited relatively high catalytic activity, given that its external Ti sites were sufficient to catalyze the epoxidation of these large-ring olefins. More recently, Dai et al. constructed plate-like TS-1 zeolites using L-lysine as a growth modifier.36 As shown in Fig. 15e, in comparison to the traditional TS-1 zeolite and plate-like TS-1 prepared in fluorine-containing media, the plate-like TS-1 synthesized with L-lysine displayed a superior catalytic performance towards the 1-hexene epoxidation reaction. The addition of L-lysine effectively inhibited the aggregation of the titanium species, increasing the Ti content within the framework, and thereby enhancing the catalytic activity. Moreover, the plate-like TS-1 zeolite synthesized with L-lysine maintained a stable catalytic performance over five reuse cycles, demonstrating excellent stability. Zhu et al. demonstrated that Ti-beta zeolite nanosheets (Ti-beta-NS) showed an improved catalytic performance in cyclohexene epoxidation.26 As illustrated in Fig. 15h, Ti-beta-NS exhibited higher activity, 1,2-cyclohexanediol selectivity (increased by 13.3–15.0%) and hydrogen peroxide utilization efficiency than the conventional Ti-beta catalyst. These enhancements originated from its unique nanosheet structure, which improved the internal diffusion efficiency and alleviated mass transfer limitations. The oriented distribution of strong acid sites was also favorable for the consecutive reaction steps in the reaction mechanism.

Wu et al. developed a novel titanosilicate catalyst with an MWW structure (Ti-MWW), and applied it to the liquid-phase ammoximation of cyclohexanone.112 The Ti-MWW achieved >99% cyclohexanone conversion with high oxime selectivity under the optimized conditions. Compared to TS-1 and Ti-MOR, Ti-MWW tended to oxidize hydroxylamine to NOx compounds, which affected hydroxylamine formation and the cyclohexanone oxime formation.

5. Summary and perspective

In conclusion, nanosheet zeolites have garnered remarkable progress in the past few decades. The synthesis strategies of nanosheet zeolites can be broadly classified into in situ hydrothermal synthesis and post treatment methods. The in situ hydrothermal can effectively and precisely control the layer thickness, interlayer spacing, and structural ordering of nanosheet zeolites. OSDA-based methods employ quaternary ammonium surfactants, which direct the formation of zeolite frameworks, while restricting crystal growth along specific directions to obtain nanosheets. Additive-assisted synthesis involves utilizing commercially available additives (e.g., urea) to regulate crystal growth, introduce mesoporosity, and reduce the crystal thickness, offering a cost-effective alternative to OSDA-based methods. Seed-induced synthesis accelerates crystallization by adding seed crystals to a template-free gel, promoting rapid growth and high-quality nanosheets. However, the layered precursor may undergo condensation during the calcination process. Several post-synthesis strategies, such as chemical etching, swelling, exfoliation, and pillaring, can not only preserve the fundamental structural units and layered framework of the nanosheets but also endow them with large pore sizes, high external surface areas, and enhanced stability.

These methods are widely used in industrial applications to optimize the zeolite properties, including morphology, pore structure, acid-site, and catalytic performance. Nanosheet zeolites with ultrathin morphologies (2–100 nm) possess unique structures and high external surface areas, enhancing the accessibility to their active sites. Their synthesis relies on surfactant-directed methods, in which the variations in surfactant composition, structure, and crystallization conditions significantly influence their morphology, interlayer spacing, and thickness. The structural properties of nanosheet zeolites can be analyzed through techniques such as XRD, which highlights their reduced crystallinity and unique diffraction patterns. N2 adsorption–desorption studies reveal that nanosheet zeolites present larger mesopore volumes and specific surface areas compared to traditional zeolites. Their acidic properties, which play a pivotal role in their catalytic performance, are assessed via NH3-TPD, FTIR, and solid-state NMR. Nanosheet structures enhance the accessibility to the Brønsted acid sites and silanol groups on their external surface, with the acid strength and type determined by the framework composition and crystallinity. Nanosheet zeolites also demonstrate tailored catalytic properties through heteroatom substitutions (e.g., Al, Ga, and Fe), optimizing their activity and selectivity in diverse reactions. Also, their ultrathin morphology with nanoscale confinement in one dimension endows them with superior catalytic efficiency in methanol conversion, cracking, isomerization, alkylation, carbonylation, and oxidation reactions, owing to the shortened diffusion paths and enhanced acid site accessibility.

However, despite the advantages of nanosheet zeolites over conventional zeolites, several challenges are encountered in their synthesis, including high cost, time-consuming processes, and the reliance on complex multi-step organic reactions for the preparation of SDAs. In this case, the seed-induced approach has emerged as a simpler and more cost-effective alternative, potentially gaining broader adoption. To advance this field, future studies should focus on optimizing the crystal growth parameters to precisely control the nanosheet thickness, lateral dimensions, and chemical composition, producing nanosheet zeolites with desirable properties. Moreover, small molecules (e.g., urea) can be used to inhibit the crystal growth in specific directions, promoting the formation of nanosheets. Furthermore, the fabrication of nanosheet zeolites utilizing cost-effective and sustainable precursors, such as silica derived from agricultural waste, presents a viable and environmentally benign approach. This strategy not only reduces production costs but also aligns with the principles of green chemistry, offering a promising alternative for the synthesis of advanced zeolitic materials. Additionally, the synthesis of zeolite nanosheets containing redox heteroatoms remains limited compared to aluminosilicate or pure silica zeolite nanosheets. Given the importance of metal-based catalysts in industrial processes, incorporating heteroatoms into nanosheet frameworks can significantly enhance the catalytic performance of metal-based catalysts.

Looking ahead, the design of affordable bifunctional templates and the use of renewable sources for nanosheet zeolite synthesis should be prioritized. In situ characterizations coupled with computational studies provide valuable insights into the nanosheet formation mechanisms and reaction pathways. From a technological perspective, applying nanosheet zeolites in catalytic membrane reactors can revolutionize the catalysis industry, offering breakthroughs in the energy and fine chemical industries. To achieve this, scaling up the industrial synthesis and applications of nanosheet zeolites are critical to realize their full potential. To date, nanosheet zeolite catalysts have demonstrated considerable potential at the laboratory scale; however, their large-scale industrial application remains in its infancy. Although some industrial-scale production methods have been reported for materials (e.g. using 20 L reactors),35 the majority of nanosheet catalysts are still under development. Thus, future research should focus on establishing commercially viable synthesis routes and verifying their long-term stability to facilitate their practical implementation.

Conflicts of interest

The authors declare no competing interests.

Data availability

We declare that the data supporting the findings of this review are available within the paper. Should any raw data files be needed in another format, they are available from the corresponding author upon reasonable request.

Acknowledgements

This work was partially supported by the National Key R&D Program of China (2023YFB4103204) and the Fundamental Research Funds for the Central Universities (023-63253174).

References

  1. C. S. Cundy and P. A. Cox, Chem. Rev., 2003, 103, 663–701 CrossRef CAS PubMed .
  2. S. T. Wilson, Stud. Surf. Sci. Catal., 1991, 58, 137–151 CrossRef CAS .
  3. J. S. Beck, J. C. Vartuli, W. J. Roth, M. E. Leonowicz, C. T. Kresge, K. D. Schmitt, C. T. W. Chu, D. H. Olson, E. W. Sheppard, S. B. McCullen, J. B. Higgins and J. L. Schlenker, J. Am. Chem. Soc., 1992, 114, 10834–10843 CrossRef CAS .
  4. D. Zhao, Y. Wan and W. Zhou, Ordered Mesoporous Mater., 2013, 27, 131–149 Search PubMed .
  5. A. Maghfirah, M. M. Ilmi, A. T. N. Fajar and G. T. M. Kadja, Mater. Today Chem., 2020, 100348 CrossRef .
  6. G. T. M. Kadja, M. D. Rukmana, R. R. Mukti, M. H. Mahyuddin, A. G. Saputro and T. D. K. Wungu, Mater. Lett., 2021, 290, 129501 CrossRef CAS .
  7. E. B. Clatworthy, M. Debost, N. Barrier, S. Gascoin, P. Boullay, A. Vicente, J. P. Gilson, J. P. Dath, N. Nesterenko and S. Mintova, ACS Appl. Nano Mater., 2021, 4, 24–28 CrossRef CAS .
  8. M. Pan, J. Zheng, Y. Liu, W. Ning, H. Tian and R. Li, J. Catal., 2019, 369, 72–85 CrossRef CAS .
  9. J. Pérez-Ramírez, C. H. Christensen, K. Egeblad, C. H. Christensen and J. C. Groen, Chem. Soc. Rev., 2008, 37, 2530–2542 RSC .
  10. W. J. Roth, B. Gil, W. Makowski, B. Marszalek and P. Eliášová, Chem. Soc. Rev., 2016, 45, 3400–3438 RSC .
  11. V. Nicolosi, M. Chhowalla, M. G. Kanatzidis, M. S. Strano and J. N. Coleman, Science, 2013, 340, 72–75 CrossRef .
  12. E. Verheyen, C. Jo, M. Kurttepeli, G. Vanbutsele, E. Gobechiya, T. I. Korányi, S. Bals, G. Van Tendeloo, R. Ryoo, C. E. A. Kirschhock and J. A. Martens, J. Catal., 2013, 300, 70–80 CrossRef CAS .
  13. Q. Gong, T. Fang, Y. Xie, R. Zhang, M. Liu, F. Barzagli, J. Li, Z. Hu and Z. Zhu, Ind. Eng. Chem. Res., 2021, 60, 1633–1641 CrossRef CAS .
  14. M. L. Firmansyah, A. A. Jalil, S. Triwahyono, H. Hamdan, M. M. Salleh, W. F. W. Ahmad and G. T. M. Kadja, Catal. Sci. Technol., 2016, 6, 5178–5182 RSC .
  15. Y. Tian, B. Zhang, H. Liang, X. Hou, L. Wang, X. Zhang and G. Liu, Appl. Catal., A, 2019, 572, 24–33 CrossRef CAS .
  16. G. T. M. Kadja, T. R. Suprianti, M. M. Ilmi, M. Khalil, R. R. Mukti and Subagjo, Microporous Mesoporous Mater., 2020, 308, 110550 CrossRef CAS .
  17. J. Hao, D. G. Cheng, F. Chen and X. Zhan, Microporous Mesoporous Mater., 2021, 310, 110647 CrossRef CAS .
  18. L. Meng, G. Vanbutsele, R. Pestman, A. Godin, D. E. Romero, A. J. F. van Hoof, L. Gao, T. F. Kimpel, J. Chai, J. A. Martens and E. J. M. Hensen, J. Catal., 2020, 389, 544–555 CrossRef CAS .
  19. Y. Guefrachi, G. Sharma, D. Xu, G. Kumar, K. P. Vinter, O. A. Abdelrahman, X. Li, S. Alhassan, P. J. Dauenhauer, A. Navrotsky, W. Zhang and M. Tsapatsis, Angew. Chem., Int. Ed., 2020, 132, 9666–9672 CrossRef .
  20. Y. Ma, X. Tang, J. Hu, Y. Ma, W. Chen, Z. Liu, S. Han, C. Xu, Q. Wu, A. Zheng, L. Zhu, X. Meng and F. S. Xiao, J. Am. Chem. Soc., 2022, 144, 6270–6277 CrossRef CAS PubMed .
  21. M. Choi, K. Na, J. Kim, Y. Sakamoto, O. Terasaki and R. Ryoo, Nature, 2009, 461, 246–249 CrossRef CAS PubMed .
  22. K. Na, M. Chol, W. Park, Y. Sakamoto, O. Terasakl and R. Ryoo, J. Am. Chem. Soc., 2010, 132, 4169–4177 CrossRef CAS PubMed .
  23. W. Park, D. Yu, K. Na, K. E. Jelfs, B. Slater, Y. Sakamoto and R. Ryoo, Chem. Mater., 2011, 23, 5131–5137 CrossRef CAS .
  24. D. Xu, Y. Ma, Z. Jing, L. Han, B. Singh, J. Feng, X. Shen, F. Cao, P. Oleynikov, H. Sun, O. Terasaki and S. Che, Nat. Commun., 2014, 5, 1–9 Search PubMed .
  25. K. Lu, J. Huang, L. Ren, C. Li, Y. Guan, B. Hu, H. Xu, J. Jiang, Y. Ma and P. Wu, Angew. Chem., Int. Ed., 2020, 59, 6258–6262 CrossRef CAS PubMed .
  26. G. Qie, R. Liu, Q. Deng, D. Deng, M. Zhai, W. Liu, W. Dai, L. Han and K. Zhu, Small, 2025, 2411639, 1–11 Search PubMed .
  27. H. Y. Luo, V. K. Michaelis, S. Hodges, R. G. Griffin and Y. Román-Leshkov, Chem. Sci., 2015, 6, 6320–6324 RSC .
  28. V. J. Margarit, M. E. Martínez-Armero, M. T. Navarro, C. Martínez and A. Corma, Angew. Chem., Int. Ed., 2015, 54, 13724–13728 CrossRef CAS PubMed .
  29. V. J. Margarit, M. R. Díaz-Rey, M. T. Navarro, C. Martínez and A. Corma, Angew. Chem., Int. Ed., 2018, 57, 3459–3463 CrossRef CAS PubMed .
  30. H. Xu, W. Chen, G. Zhang, P. Wei, Q. Wu, L. Zhu, X. Meng, X. Li, J. Fei, S. Han, Q. Zhu, A. Zheng, Y. Ma and F. S. Xiao, J. Mater. Chem. A, 2019, 7, 16671–16676 RSC .
  31. H. Chen, M. Wang, M. Yang, W. Shang, C. Yang, B. Liu, Q. Hao, J. Zhang and X. Ma, J. Mater. Sci., 2019, 54, 8202–8215 CrossRef CAS .
  32. A. I. Lupulescu and J. D. Rimer, Angew. Chem., 2012, 124, 3401–3405 CrossRef .
  33. Z. Shan, H. Wang, X. Meng, S. Liu, L. Wang, C. Wang, F. Li, J. P. Lewis and F. S. Xiao, Chem. Commun., 2011, 47, 1048–1050 RSC .
  34. G. Xu, X. Zhu, X. Li, S. Xie, S. Liu and L. Xu, Microporous Mesoporous Mater., 2010, 129, 278–284 CrossRef CAS .
  35. W. Dai, C. Kouvatas, W. Tai, G. Wu, N. Guan, L. Li and V. Valtchev, J. Am. Chem. Soc., 2021, 143, 1993–2004 CrossRef CAS PubMed .
  36. W. Dai, J. Zhao, Z. Liu, Y. Wang, J. Zheng and R. Li, Mater. Today Sustainability, 2024, 27, 100943 CrossRef .
  37. Z. Xiong, G. Qi, E. Zhan, Y. Chu, J. Xu, J. Wei, N. Ta, A. Hao, Y. Zhou, F. Deng and W. Shen, Chem, 2023, 9, 76–92 CAS .
  38. M. Y. Jeon, D. Kim, P. Kumar, P. S. Lee, N. Rangnekar, P. Bai, M. Shete, B. Elyassi, H. S. Lee, K. Narasimharao, S. N. Basahel, S. Al-Thabaiti, W. Xu, H. J. Cho, E. O. Fetisov, R. Thyagarajan, R. F. Dejaco, W. Fan, K. A. Mkhoyan, J. I. Siepmann and M. Tsapatsis, Nature, 2017, 543, 690–694 CrossRef CAS PubMed .
  39. M. Liu, J. Li, W. Jia, M. Qin, Y. Wang, K. Tong, H. Chen and Z. Zhu, RSC Adv., 2015, 5, 9237–9240 RSC .
  40. Y. Kamimura, K. Itabashi, Y. Kon, A. Endo and T. Okubo, Chem. – Asian J., 2017, 12, 530–542 CrossRef CAS PubMed .
  41. R. Jain, A. Chawla, N. Linares, J. García Martínez and J. D. Rimer, Adv. Mater., 2021, 33, 1–8 CrossRef PubMed .
  42. X. Zhang, D. Liu, D. Xu, S. Asahina, K. A. Cychosz, K. V. Agrawal, Y. Al Wahedi, A. Bhan, S. Al Hashimi, O. Terasaki, M. Thommes and M. Tsapatsis, Science, 2012, 336, 1684–1687 CrossRef CAS PubMed .
  43. M. Khaleel, A. J. Wagner, K. A. Mkhoyan and M. Tsapatsis, Angew. Chem., Int. Ed., 2014, 53, 9456–9461 CrossRef CAS PubMed .
  44. P. Kumar, D. W. Kim, N. Rangnekar, H. Xu, E. O. Fetisov, S. Ghosh, H. Zhang, Q. Xiao, M. Shete, J. I. Siepmann, T. Dumitrica, B. McCool, M. Tsapatsis and K. A. Mkhoyan, Nat. Mater., 2020, 19, 443–449 CrossRef CAS PubMed .
  45. J. Přech, K. N. Bozhilov, J. El Fallah, N. Barrier and V. Valtchev, Microporous Mesoporous Mater., 2019, 280, 297–305 CrossRef .
  46. T. Zhou, D. Zhang, Y. Liu, Y. Sun, T. Ji, S. Huang and Y. Liu, J. Energy Chem., 2022, 72, 516–521 CrossRef CAS .
  47. A. Corma, V. Fornés, J. Martínez-Triguero and S. B. Pergher, J. Catal., 1999, 186, 57–63 CrossRef CAS .
  48. L. Wei, K. Song, W. Wu, S. Holdren, G. Zhu, E. Shulman, W. Shang, H. Chen, M. R. Zachariah and D. Liu, J. Am. Chem. Soc., 2019, 141, 8712–8716 CrossRef CAS PubMed .
  49. N. Wang, Q. Sun, T. Zhang, A. Mayoral, L. Li, X. Zhou, J. Xu, P. Zhang and J. Yu, J. Am. Chem. Soc., 2021, 143, 6905–6914 CrossRef CAS PubMed .
  50. I. Ogino, M. M. Nigra, S. J. Hwang, J. M. Ha, T. Rea, S. I. Zones and A. Katz, J. Am. Chem. Soc., 2011, 133, 3288–3291 CrossRef CAS PubMed .
  51. X. Ouyang, S. J. Hwang, R. C. Runnebaum, D. Xie, Y. J. Wanglee, T. Rea, S. I. Zones and A. Katz, J. Am. Chem. Soc., 2014, 136, 1449–1461 CrossRef CAS PubMed .
  52. W. J. Roth, T. Sasaki, K. Wolski, Y. Song, D. M. Tang, Y. Ebina, R. Ma, J. Grzybek, K. Kałahurska, B. Gil, M. Mazur, S. Zapotoczny and J. Cejka, Sci. Adv., 2020, 6, 2–7 Search PubMed .
  53. W. J. Roth, T. Sasaki, K. Wolski, Y. Ebina, D. M. Tang, Y. Michiue, N. Sakai, R. Ma, O. Cretu, J. Kikkawa, K. Kimoto, K. Kalahurska, B. Gil, M. Mazur, S. Zapotoczny, J. Čejka, J. Grzybek and A. Kowalczyk, J. Am. Chem. Soc., 2021, 143, 11052–11062 CrossRef CAS PubMed .
  54. B. Wang, N. Wang, X. Li, R. Zhou and W. Xing, J. Membr. Sci., 2022, 645, 120177 CrossRef CAS .
  55. S. Hu, J. Shan, Q. Zhang, Y. Wang, Y. Liu, Y. Gong, Z. Wu and T. Dou, Appl. Catal., A, 2012, 445–446, 215–220 CrossRef CAS .
  56. B. Liu, Z. Chen, J. Huang, H. Chen and Y. Fang, Microporous Mesoporous Mater., 2019, 273, 235–242 CrossRef CAS .
  57. S. Salakhum, T. Yutthalekha, M. Chareonpanich, J. Limtrakul and C. Wattanakit, Microporous Mesoporous Mater., 2018, 258, 141–150 CrossRef CAS .
  58. S. Ferdov, Microporous Mesoporous Mater., 2017, 242, 59–62 CrossRef CAS .
  59. Y. Shang, W. Wang, Y. Zhai, Y. Song, X. Zhao, T. Ma, J. Wei and Y. Gong, Microporous Mesoporous Mater., 2019, 276, 173–182 CrossRef CAS .
  60. Y. Liu, N. Zhao, H. Xian, Q. Cheng, Y. Tan, N. Tsubaki and X. Li, ACS Appl. Mater. Interfaces, 2015, 7, 8398–8403 CrossRef CAS PubMed .
  61. W. Dai, Z. Liu, J. Zhao, Y. Wang, J. Zheng and R. Li, Appl. Catal., A, 2024, 684, 119912 CrossRef CAS .
  62. Y. Li, D. Ma, W. Fu, C. Liu, Y. Wang, Z. Wang and W. Yang, RSC Adv., 2022, 12, 14183–14189 RSC .
  63. Y. Zhou, Y. Mu, M. F. Hsieh, B. Kabius, C. Pacheco, C. Bator, R. M. Rioux and J. D. Rimer, J. Am. Chem. Soc., 2020, 142, 8211–8222 CrossRef CAS PubMed .
  64. B. Liu, Q. Duan, C. Li, Z. Zhu, H. Xi and Y. Qian, New J. Chem., 2014, 38, 4380–4387 RSC .
  65. L. Meng, G. Vanbutsele, R. Pestman, A. Godin, D. E. Romero, A. J. F. van Hoof, L. Gao, T. F. Kimpel, J. Chai, J. A. Martens and E. J. M. Hensen, J. Catal., 2020, 389, 544–555 CrossRef CAS .
  66. T. Yutthalekha, C. Wattanakit, C. Warakulwit, W. Wannapakdee, K. Rodponthukwaji, T. Witoon and J. Limtrakul, J. Cleaner Prod., 2017, 142, 1244–1251 CrossRef CAS .
  67. W. J. Roth, O. V. Shvets, M. Shamzhy, P. Chlubná, M. Kubu, P. Nachtigall and J. Ĉejka, J. Am. Chem. Soc., 2011, 133, 6130–6133 CrossRef CAS PubMed .
  68. Q. Zhang, H. Yang and W. Yan, RSC Adv., 2014, 4, 56938–56944 RSC .
  69. Y. Tian, B. Zhang, H. Liang, X. Hou, L. Wang, X. Zhang and G. Liu, Appl. Catal., A, 2019, 572, 24–33 CrossRef CAS .
  70. A. Chang, T. C. Yang, M. Y. Chen, H. M. Hsiao and C. M. Yang, J. Hazard. Mater., 2020, 400, 123241 CrossRef CAS PubMed .
  71. R. Liu, Y. Bian and W. Dai, Chin. J. Struct. Chem., 2024, 43, 100250 CAS .
  72. W. Dai, L. Zhang, R. Liu, Z. Huo and N. Guan, Mater. Today Sustainability, 2023, 22, 100364 CrossRef .
  73. L. Wu, P. C. M. M. Magusin, V. Degirmenci, M. Li, S. M. T. Almutairi, X. Zhu, B. Mezari and E. J. M. Hensen, Microporous Mesoporous Mater., 2014, 189, 144–157 CrossRef CAS .
  74. K. Saenluang, T. Imyen, W. Wannapakdee, D. Suttipat, P. Dugkhuntod, M. Ketkaew, A. Thivasasith and C. Wattanakit, ACS Appl. Nano Mater., 2020, 3, 3252–3263 CrossRef CAS .
  75. Y. Ji, B. Shi, H. Yang and W. Yan, Appl. Catal., A, 2017, 533, 90–98 CrossRef CAS .
  76. L. Zhang, L. Yang, R. Liu, X. Shao, W. Dai, G. Wu, N. Guan, Z. Guo, W. Zhu and L. Li, Microporous Mesoporous Mater., 2022, 333, 111767 CrossRef CAS .
  77. K. Góra-Marek, K. Tarach and M. Choi, J. Phys. Chem. C, 2014, 118, 12266–12274 CrossRef .
  78. Y. Shang, W. Wang, Y. Zhai, Y. Song, X. Zhao, T. Ma, J. Wei and Y. Gong, Microporous Mesoporous Mater., 2019, 276, 173–182 CrossRef CAS .
  79. R. Liu, X. Shao, C. Wang, W. Dai and N. Guan, Chin. J. Catal., 2023, 47, 67–92 CrossRef CAS .
  80. N. Hadi, R. Alizadeh and A. Niaei, J. Ind. Eng. Chem., 2017, 54, 82–97 CrossRef CAS .
  81. H. Chen, W. Shang, C. Yang, B. Liu, C. Dai, J. Zhang, Q. Hao, M. Sun and X. Ma, Ind. Eng. Chem. Res., 2019, 58, 1580–1589 CrossRef CAS .
  82. Q. Sun, Y. Ma, N. Wang, X. Li, D. Xi, J. Xu, F. Deng, K. B. Yoon, P. Oleynikov, O. Terasaki and J. Yu, J. Mater. Chem. A, 2014, 2, 17828–17839 RSC .
  83. A. Xing, N. Zhang, D. Yuan, H. Liu, Y. Sang, P. Miao, Q. Sun and M. Luo, Ind. Eng. Chem. Res., 2019, 58, 12506–12515 CrossRef CAS .
  84. Y. Kim, J. C. Kim, C. Jo, T. W. Kim, C. U. Kim, S. Y. Jeong and H. J. Chae, Microporous Mesoporous Mater., 2016, 222, 1–8 CrossRef CAS .
  85. Y. Tian, B. Zhang, H. Liang, X. Hou, L. Wang, X. Zhang and G. Liu, Appl. Catal., A, 2019, 572, 24–33 CrossRef CAS .
  86. Y. Ji, B. Shi, H. Yang and W. Yan, Appl. Catal., A, 2017, 533, 90–98 CrossRef CAS .
  87. L. Liu, H. Wang, R. Wang, C. Sun, S. Zeng, S. Jiang, D. Zhang, L. Zhu and Z. Zhang, RSC Adv., 2014, 4, 21301–21305 RSC .
  88. Y. Tian, B. Zhang, S. Gong, L. Wang, X. Zhang, C. Qiao and G. Liu, Microporous Mesoporous Mater., 2021, 310, 110598 CrossRef CAS .
  89. C. Lei, Z. Dong, C. Martínez, J. Martínez-Triguero, W. Chen, Q. Wu, X. Meng, A. N. Parvulescu, T. De Baerdemaeker, U. Müller, A. Zheng, Y. Ma, W. Zhang, T. Yokoi, B. Marler, D. E. De Vos, U. Kolb, A. Corma and F. S. Xiao, Angew. Chem., Int. Ed., 2020, 59, 15649–15655 CrossRef CAS PubMed .
  90. L. Xu, X. Ji, S. Li, Z. Zhou, X. Du, J. Sun, F. Deng, S. Che and P. Wu, Chem. Mater., 2016, 28, 4512–4521 CrossRef CAS .
  91. W. Dai, L. Zhang, R. Liu, G. Wu, N. Guan and L. Li, ACS Appl. Mater. Interfaces, 2022, 14, 11415–11424 CrossRef CAS PubMed .
  92. J. Přech, K. N. Bozhilov, J. El Fallah, N. Barrier and V. Valtchev, Microporous Mesoporous Mater., 2019, 280, 297–305 CrossRef .
  93. Z. Chen, J. Huang, Q. Mu, H. Chen, F. Xu, Y. Fang and B. Liu, Chin. J. Chem. Eng., 2019, 27, 1981–1987 CrossRef CAS .
  94. T. Lee, Y. Bae, J. Jeon, J. Kim, J. Choi and K. S. Ha, Catal. Today, 2020, 352, 183–191 CrossRef .
  95. W. Wannapakdee, D. Suttipat, P. Dugkhuntod, T. Yutthalekha, A. Thivasasith, P. Kidkhunthod, S. Nokbin, S. Pengpanich, J. Limtrakul and C. Wattanakit, Fuel, 2019, 236, 1243–1253 CrossRef CAS .
  96. J. Kim, W. Kim, Y. Seo, J. C. Kim and R. Ryoo, J. Catal., 2013, 301, 187–197 CrossRef CAS .
  97. Y. Wang, Y. Gao, W. Chu, D. Zhao, F. Chen, X. Zhu, X. Li, S. Liu, S. Xie and L. Xu, J. Mater. Chem. A, 2019, 7, 7573–7580 RSC .
  98. W. Dai, V. Ruaux, X. Deng, W. Tai, G. Wu, N. Guan, L. Li and V. Valtchev, J. Mater. Chem. A, 2021, 9, 24922–24931 RSC .
  99. L. Ren, Q. Guo, P. Kumar, M. Orazov, D. Xu, S. M. Alhassan, K. A. Mkhoyan, M. E. Davis and M. Tsapatsis, Angew. Chem., Int. Ed., 2015, 54, 10848–10851 CrossRef CAS PubMed .
  100. B. Liu, Z. Chen, J. Huang, H. Chen and Y. Fang, Microporous Mesoporous Mater., 2019, 273, 235–242 CrossRef CAS .
  101. Z. Wang, M. O. Cichocka, Y. Luo, B. Zhang, H. Sun, Y. Tang and W. Yang, Chin. J. Catal., 2020, 41, 1062–1066 CrossRef CAS .
  102. X. Feng, P. Zhang, Y. Fang, W. Charusiri, J. Yao, X. Gao, Q. Wei, P. Reubroycharoen, T. Vitidsant, Y. Yoneyama, G. Yang and N. Tsubaki, Catal. Today, 2020, 343, 206–214 CrossRef CAS .
  103. M. Ma, X. Huang, E. Zhan, Y. Zhou, H. Xue and W. Shen, J. Mater. Chem. A, 2017, 5, 8887–8891 RSC .
  104. J. Yao, Q. Wu, J. Fan, S. Komiyama, X. Yong, W. Zhang, T. Zhao, Z. Guo, G. Yang and N. Tsubaki, ACS Nano, 2021, 15, 13568–13578 CrossRef CAS PubMed .
  105. W. Wu, D. T. Tran, X. Wu, S. C. Oh, M. Wang, H. Chen, L. Emdadi, J. Zhang, E. Schulman and D. Liu, Microporous Mesoporous Mater., 2019, 278, 414–422 CrossRef CAS .
  106. N. Li, M. Wang, Q. You, C. Bi, H. Chen, B. Liu, M. Sun, Q. Hao, J. Zhang and X. Ma, Catal. Sci. Technol., 2020, 10, 1323–1335 RSC .
  107. W. Zou, P. Xie, W. Hua, Y. Wang, D. Kong, Y. Yue, Z. Ma, W. Yang and Z. Gao, J. Mol. Catal. A: Chem., 2014, 394, 83–88 CrossRef CAS .
  108. L. Meng, X. Zhu and E. J. M. Hensen, ACS Catal., 2017, 7, 2709–2719 CrossRef CAS PubMed .
  109. G. Liu, Y. Tian, B. Zhang, L. Wang and X. Zhang, J. Hazard. Mater., 2019, 367, 568–576 CrossRef CAS PubMed .
  110. N. Li, M. Wang, Q. You, C. Bi, H. Chen, B. Liu, M. Sun, Q. Hao, J. Zhang and X. Ma, Catal. Sci. Technol., 2020, 10, 1323–1335 RSC .
  111. K. Na, C. Jo, J. Kim, W. S. Ahn and R. Ryoo, ACS Catal., 2011, 1, 901–907 CrossRef CAS .
  112. F. Song, Y. Liu, H. Wu, M. He, P. Wu and T. Tatsumi, J. Catal., 2006, 237, 359–367 CrossRef CAS .
  113. X. Wang, Y. Ma, Q. Wu, Y. Wen and F. S. Xiao, Chem. Soc. Rev., 2022, 51, 2431–2443 RSC .
  114. G. T. M. Kadja, N. J. Azhari, S. Mardiana, N. T. U. Culsum and A. Maghfirah, Results Eng., 2023, 17, 100910 CrossRef CAS .
  115. E. Schulman, W. Wu and D. Liu, Materials, 2020, 13, 1822 CrossRef CAS PubMed .

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.