Abstract
Metal–organic frameworks (MOFs) have emerged as attractive chemical sensing materials due to their exceptionally high porosity and chemical diversity. Nevertheless, the utilization of MOFs in chemiresistive type sensors has been hindered by their inherent limitation in electrical conductivity. The recent emergence of two-dimensional conductive MOFs (2D c-MOFs) has addressed this limitation by offering enhanced electrical conductivity, while still retaining the advantageous properties of MOFs. In particular, c-MOFs have shown promising advantages for the fabrication of sensors capable of operating at room temperature. Thus, active research on gas sensors utilizing c-MOFs is currently underway, focusing on enhancing sensitivity and selectivity. To comprehend the potential of MOFs as chemiresistive sensors for future applications, it is crucial to understand not only the fundamental properties of conductive MOFs but also the state-of-the-art works that contribute to improving their performance. This comprehensive review delves into the distinctive characteristics of 2D c-MOFs as a new class of chemiresistors, providing in-depth insights into their unique sensing properties. Furthermore, we discuss the proposed sensing mechanisms associated with 2D c-MOFs and provide a concise summary of the strategies employed to enhance the sensing performance of 2D c-MOFs. These strategies encompass a range of approaches, including the design of metal nodes and linkers, morphology control, and the synergistic use of composite materials. In addition, the review thoroughly explores the prospects of 2D c-MOFs as chemiresistors and elucidates their remarkable potential for further advancements. The insights presented in this review shed light on future directions and offer valuable opportunities in the chemical sensing research field.
Keywords: metal−organic frameworks, chemiresistors, two-dimensional materials, sensors, electrical conductivity, sensing mechanism, selectivity, sensitivity, composites
1. Introduction
Metal–organic frameworks (MOFs) are a unique class of porous materials constructed from metal nodes and organic linkers. Due to their inherent permanent porosity and chemical tunability, MOFs have found widespread applications in diverse fields, including gas storage,1−3 catalysis,4−6 energy storage,7−9 and separation processes.10−12 Among these applications, chemical sensing emerges as an ideal and prominent research field for the utilization of MOFs.13,14 The remarkable porosity of MOFs facilitates the easy diffusion of analytes, while the tunability of these materials enables the incorporation of various binding sites, leading to specific interaction with target gases.15 These unique properties make MOFs highly suitable for chemical sensing applications. Among various types of sensors, chemiresistive sensors simultaneously offer high cost-effectiveness, portability, and sensitivity.16,17 However, conventional MOFs, lacking efficient charge transport pathways through their abundant pores, have been considered as electrically insulating materials, limiting their use in chemiresistors. Despite efforts to utilize MOFs with narrow band gaps in chemiresistors, their limited conductivities still require thermal activation, which poses challenges in terms of stability and device complexity.18−20 Alternatively, MOFs can be employed in conjunction with other conductive materials. However, in such cases, the interaction of MOFs may not be entirely transduced into the electrical signal, resulting in poor responses.21
However, the recent emergence of two-dimensional conductive MOFs (2D c-MOFs) has opened up new possibilities for the application of MOFs in various electronic devices as active components.22−26 2D c-MOFs are composed of π-conjugated ligands capable of effective charge delocalization. These ligands are linked to metal nodes through π–d hybridization, forming extended conjugation throughout the frameworks. As a result, 2D c-MOFs offer sufficiently high electrical conductivity measurable at room temperature, combined with the advantageous properties of MOFs. In 2D c-MOFs, the facile gas adsorption and interactions via various active sites can be directly transduced into an electrical signal, making them highly desirable materials for chemiresistive sensors.
Due to the unique properties described above, 2D c-MOFs exhibit several advantages over other sensing materials, such as metal oxides,27−29 graphene,30−33 and transition metal dichalcogenide.34−36 First, 2D c-MOFs inherently offer tunable porous structures, which are a key design element in sensing materials. Unlike other sensing materials that require multiple nanostructuring processes to achieve efficient porous structures,37−39 as-synthesized 2D c-MOFs provide well-defined porous structures with higher porosity and surface area. This feature simplifies the fabrication process of sensors, allowing for improved gas adsorption and interactions within the material, thereby enhancing sensing performance. Second, 2D c-MOFs possess highly tunable electrical properties. The high electrical conductivity and room temperature operation of 2D c-MOFs offer a significant advantage by addressing the issues associated with thermally activated sensors, which typically operate at elevated temperatures ranging from 200 to 400 °C. By eliminating the need for such high-temperature operation, 2D c-MOFs can overcome challenges related to device complexity, power consumption, and instability.40−42 Furthermore, the conductivity of 2D c-MOFs can be easily tuned across a wide range, enabling the utilization of sensors with optimal intrinsic conductivity depending on specific measurement conditions.43,44 This makes 2D c-MOFs more practical and efficient for room temperature chemiresistive sensing application. Third, the high diversity in chemical structures of 2D c-MOFs provides an advantage in selectivity tuning. Semiconducting metal oxides, which are the most well-established chemiresistive sensing materials, have achieved a sufficiently high level of sensitivity.45−48 However, the selective detection of target gases in real environments, where interfering gases are present, continues to be a significant challenge.17 This is because metal oxides primarily rely on similar surface reactions involving chemisorbed oxygen when interacting with different analytes.49 In contrast, the high diversity of binding sites in 2D c-MOFs enables unique interactions for specific targets, offering great potential to address these challenges associated with selective detection in the presence of interfering gases. In recent years, significant progress has been made in the development of 2D c-MOF-based sensors, demonstrating their potential as promising sensing materials. However, a dedicated review paper specifically focusing on the application of 2D c-MOFs in chemiresistors is currently lacking, while there are review papers discussing the applications of MOFs in sensors50−55 or 2D c-MOFs-based electronics in a more general context.22,56,57
In this review, our objective is to provide comprehensive insights into the application of 2D c-MOFs as chemiresistors, offering cutting-edge sensing strategies and a deeper understanding of their unique characteristics relevant to sensing performance (Figure 1). First, we discuss the fundamental principles of 2D c-MOFs within the context of gas sensor applications, encompassing their general properties, their proposed sensing mechanisms, and their integration into electrical devices. Subsequently, we provide a comprehensive summary of the reported 2D c-MOFs sensors, highlighting the various strategies employed to enhance their sensing performance. These strategies include the design of metal nodes and linkers, the control of morphology, and the utilization of composites based on 2D c-MOFs. We discuss how these approaches contribute to enhanced sensing characteristics of 2D c-MOFs sensors. Finally, we outline the prospects of 2D c-MOFs sensors and their potential for further advancement as highly effective sensing materials.
Figure 1.
Overview of basic principles of 2D c-MOFs in the context of chemiresistors, and strategies for improving the sensing performance with their key advantages.
2. Basic Principles in 2D c-MOFs Chemiresistors
2.1. Structural and Electrical Properties
Two key features of 2D c-MOFs are their well-defined porous frameworks and high electrical conductivities.58 The porous structures are directly related to the sensing properties of 2D c-MOFs as they affect the mass transport of gases.15,59 Meanwhile, high electrical conductivity is beneficial for operating chemiresistors with lower power consumption, particularly at room temperature.60 The direct correlation between electrical conductivity and sensing characteristics has not been clearly established. However, several studies have indicated that 2D c-MOFs with lower conductivity may offer advantages in improving response.61,62
Both the porous structures and electrical properties of 2D c-MOFs can be tailored by selecting different linkers. Figure 2a–d and Table 1 show the lattice layers and conductivities of representative 2D c-MOFs with various linkers.63−90 One notable difference is the pore sizes, which depend on the sizes of the linker molecules. For instance, triphenylene-based (HXTP) 2D c-MOFs exhibit larger honeycomb pores compared to benzene-based 2D c-MOFs (Figure 2a and c). In contrast, Cu3(BHT) exhibits a nonporous structure that may hinder the diffusion of gases within the frameworks (Figure 2b). However, 2D c-MOFs with shorter linkers can possess a higher density of active sites, which enables enhanced interactions with gases. In terms of electrical properties, describing the influence of linkers in a straightforward manner is challenging due to the complexity of the various factors involved. These factors include the length of pi conjugation, symmetry, density, and oxidation state, all of which can affect the conductivity of 2D c-MOFs.22 One potential contributing factor is the type of heteroatom in the M–X linkage, which results in different degrees of π–d orbital overlap with the metals. The high degree of orbital overlap can contribute to improving charge mobility and narrowing band gap, thereby enhancing the conductivity of 2D c-MOFs.58 For instance, Cu3(BHT) exhibits an ultrahigh conductivity of approximately 2500 S/cm with metallic behavior, owing to the high π–d orbital overlap of the M–S bond.82 Meanwhile, metal nodes have a lesser influence on the overall structures of 2D c-MOFs compared to linkers. However, they can still significantly affect the electronic structures through their different capabilities to form orbital hybridizations with linkers. For example, the M3(HITP)2 series exhibits different band gaps (0.291–0.762 eV) and conductivities (5.8 × 10–3–55 S/cm) depending on the choice of Cu, Co, and Ni nodes (Figure 2e).88
Figure 2.
Structure of lattice layers in (a) benzene-based 2D c-MOFs, (b) Cu3(BHT), (c) triphenylene-based 2D c-MOFs, and (d) phthalocyanine-based 2D c-MOFs. (e) Variation in bandgaps and conductivities of M3(HITP)2 with Co, Cu, and Ni nodes. Reprinted with permission from ref (88). Copyright 2020 American Chemical Society. (f) Possible stacking arrangements (eclipsed, serrated, and inclined) of 2D layers. Structures of (g) extended frameworks and (h) discrete complexes.
Table 1. Summary of Conductivities of Representative 2D c-MOFsa.
2D c-MOFs | linker core, heteroatom | sample type | measurement method | conductivity (S/cm) | ref |
---|---|---|---|---|---|
Cu3(THQ)2 | benzene, O | pellet | van der Pauw | 7.3 × 10–8 | (75) |
Cu3(HIB)2 | benzene, N | pellet | van der Pauw | 13 | (85) |
Ni3(HIB)2 | benzene, N | pellet | van der Pauw | 8 | (85) |
Co3(HIB)2 | benzene, N | pellet | 4-probe | 1.57 | (67) |
Ni3(BHT)2 | benzene, S | pellet | 4-probe | 2.8–160 | (90) |
Pd3(BHT)2 | benzene, S | film | 4-probe | 0.028 | (84) |
Cu3(BHT) | benzene, S | pellet | 4-probe | 48–280 | (83) |
film | 4-probe | 2500 | (82) | ||
Cu3(HHTP)2 | triphenylene, O | pellet | 4-probe | 0.02 | (86) |
film | 4-probe | 0.29 | (81) | ||
nanorod (scb) | 4-probe | 1.5 | (87) | ||
nanoflake (sc) | 2-probe | 0.5 | (87) | ||
Ni–HHTP (Ni9(HHTP)4) | triphenylene, O | pellet | 4-probe | 0.01 | (80) |
film | van der Pauw | 1.1 × 10–3 | (79) | ||
Co–HHTP (Co9(HHTP)4) | triphenylene, O | pellet | van der Pauw | 0.032 | (79) |
film | van der Pauw | 3.3 × 10–3 | (79) | ||
Cu3(HITP)2 | triphenylene, N | pellet | 4-probe | 0.75 | (88) |
Ni3(HITP)2 | triphenylene, N | pellet | 4-probe | 55.4 | (88) |
film | van der Pauw | 40 | (68) | ||
nanorod (sc) | 4-probe | 150 | (87) | ||
Co3(HITP)2 | triphenylene, N | pellet | 4-probe | 0.024 | (88) |
Fe3(HTTP)2 | triphenylene, S | film | van der Pauw | 0.2 | (78) |
Co3(HTTP)2 | triphenylene, S | pellet | van der Pauw | 1.4 × 10–3 | (77) |
film | van der Pauw | 0.032 | (77) | ||
Pt3(HTTP)2 | triphenylene, S | pellet | 2-probe | 3.86× 10–6 | (72) |
NiPc–Cu | phthalocyanine, O | pellet | 4-probe | 0.014 | (60) |
NiPc–Ni | phthalocyanine, O | pellet | 4-probe | 7.2 × 10–4 | (60) |
CuPc–Cu | phthalocyanine, O | pellet | 2-probe | 9.4 × 10–8–1.6 × 10–6 | (76) |
NiIPc–Ni | phthalocyanine, N | film | 4-probe | 0.2 | (69) |
CuIPc–Ni | phthalocyanine, N | pellet | van der Pauw | 1 × 10–4 | (71) |
This table presents selected examples of reported conductivities. Measured conductivity of the same 2D c-MOF can vary by 1–3 orders of magnitude across studies, depending on synthetic conditions, measurement methods, atmospheres, and other factors.
Conductivity of a single crystal.
In addition to the 2D structures of layers, the stacking arrangement of the layers also plays a significant role in determining their structural and electrical properties. While 2D c-MOFs have been designed based on in-plane charge transport through extended conjugation pathways, recent studies have highlighted the importance of the out-of-plane pathway involving interlayer π–π stacking interactions.87 This pathway has been found to make a significant contribution comparable to the in-plane pathway. Consequently, 2D c-MOFs with shorter interlayer spacing generally exhibit higher conductivity due to enhanced π–π overlap.91 Moreover, different stacking patterns can be observed within the same layer, which includes eclipsed, serrated, and inclined arrangements (Figure 2f).92 These distinct stacking patterns result in variations in porosity and electrical properties. For example, in 2D c-MOFs with highly staggered arrangements, the pores can be completely blocked by adjacent layers.93 Additionally, some 2D c-MOFs are composed of two distinct layers with alternative stacking. A notable example is Co–HHTP, which consists of discrete molecular complexes integrated within the extended framework (Figure 2g and h). These discrete complexes possess different chemical structures and oxidation states compared to the extended layer, which can influence their activity toward guest molecules. Thus, considering the three-dimensional arrangement of the stacked layers is essential for gaining a comprehensive understanding of the structural and functional characteristics of 2D c-MOFs. Meanwhile, there are other 2D c-MOFs based on different linkers which are not described in this review.56,58 Although their sensing characteristics have not been extensively investigated, they all hold great potential as promising candidates for implementing diverse features into gas sensors.
2.2. Gas Sensing Mechanisms
As 2D c-MOFs are relatively new sensing materials with a wide range of structures, their underlying sensing mechanisms have not been clearly established. However, identifying potential gas binding sites can serve as an initial step toward understanding the mechanism of these materials. In a general sense, metal nodes of 2D c-MOFs can be considered as highly active sites for gas binding. Rubio-Giménez et al. confirmed the coordination of NH3 gas on metal nodes through both density functional theory (DFT) calculations and infrared spectroscopy.94 The Mirica research group conducted an extensive study on the interaction of probe gases (NH3, CD3CN, pyridine) with metal complexes in M3(HITP)2 and M3(HHTP)2 (M = Cu and Ni) (Figure 3a).95 They proposed that the metal sites can function as Lewis acid sites by directly accepting lone pairs from gases. These metal sites also can act as Brønsted acid sites through the mediation of hydrated H2O ligands. Diffuse reflectance infrared Fourier transform spectroscopy (DRIFT) analysis revealed that the metal sites of all MOFs (MO4 and MN4) serve as Lewis acid sites to bind NH3. Upon exposure to less basic gases (CD3CN, pyridine), it was revealed that the Lewis acidity of HHTP-linked MO4 sites is higher than that of HITP-linked MN4 sites. Furthermore, the presence of NH4+ in the spectra of M3(HHTP)2 indicates that MO4 complexes can also act as Brønsted acid sites simultaneously. In addition, metal sites in 2D c-MOFs can also function as catalytic sites, facilitating the transformation of gases into other species. For example, Ni nodes can oxidize adsorbed H2S to sulfite or sulfate.96 Although irreversible transformations can facilitate strong interactions with analytes, they often lead to irreversible sensing behavior.
Figure 3.
(a) Possible interactions of metal nodes in M3(HHTP)2 and M3(HITP)2 (M = Cu and Ni) with basic molecular probes. (b) The trend in redox activities and acidities of M3(HHTP)2 and M3(HITP)2. Adapted with permission from ref (95). Copyright 2020 American Chemical Society. (c) DFT-optimized structures and density of states of Cu3(HHTP)2 interacting with N2, H2O, and NH3. Reprinted with permission from ref (94). Copyright 2019 Wiley-VCH.
In addition to gas binding on metal nodes, linkers in 2D c-MOFs can also serve as potential binding sites. Their various heteroatoms, such as O, N, and S, interact with a wide range of gases through hydrogen bonding or polar interaction. For example, Campbell et al. reported opposite sensing responses of Cu3(HITP)2 toward different amines (nBuNH2 and Et3N), which is unusual in other sensing materials.97 Considering the different hydrogen bonding capabilities of the secondary and tertiary amines, this implies that −NH moieties in the linker can interact with gases via hydrogen bonds.
Notably, the binding properties of metal nodes and linkers further vary depending on their location within a 2D c-MOF. For instance, the axial position of a metal node within the interior frameworks may not be accessible to gases due to the narrow interlayer spacing (∼0.3 nm) of 2D c-MOFs. On the other hand, metal nodes located at the basal plane and the terminal edge are more accessible to gases.95 The different coordination environments of the metal nodes lead to distinct chemical properties and binding behaviors. Therefore, understanding and controlling these distinct metal sites within 2D c-MOFs are crucial to optimize their sensing performance.
In chemiresistive response, the binding of analytes needs to be transduced into an electrical signal. One straightforward example is the change in charge density of the sensing materials caused by the exposure of the reducing or oxidizing gases. This change in charge density affects the electrical conductivity of the material, resulting in a measurable variation in its resistance. The charge transfer mechanism is also partially valid for 2D c-MOF sensors.97,98 However, unusual sensing behaviors have been observed in 2D c-MOFs, suggesting that charge transfer alone is insufficient to explain their sensing mechanisms. In a study by the Mirica group, the oxidation state change of 2D c-MOFs after binding NH3 was investigated using X-ray photoelectron spectroscopy (XPS) and electron paramagnetic resonance (EPR) analyses.95 The results showed that the redox-active Cu nodes underwent oxidation or reduction upon analyte binding, but the overall frameworks of the MOFs maintained charge neutrality. Instead, the linker molecules were found to be reduced or oxidized to counterbalance the oxidation state change of the metal nodes (Figure 3b). The change in the oxidation state of the redox-active linker in 2D c-MOFs can lead to conductivity change through the generation of radicals or alterations in π–d hybridization.22 On the other hand, despite the clear binding of NH3, such redox events do not occur with Ni nodes since the stable oxidation state of Ni2+ is maintained.
In addition, the structural distortion of 2D c-MOFs upon binding gases can influence their electrical properties. Rubio-Giménez et al. observed that the binding of guest molecules led to the disruption of metal complexes and an increase in the interlayer spacing, which is related to out-of-plane charge transport (Figure 3c).94 These changes in interlayer spacing were correlated with changes in conductivity induced by the guest molecules (N2, H2O, and NH3). It is important to note that these mechanisms and other factors can simultaneously affect the electrical properties of 2D c-MOFs. The complex nature of the 2D c-MOFs, combined with the diverse binding sites of MOFs, poses challenges in elucidating the precise sensing mechanisms and designing optimal sensors. However, the distinctive sensing behavior of 2D c-MOFs, which is not observed in other sensing materials, holds immense potential for advancements in sensor technology.
2.3. Sensing Parameters and Related Design Principles
The performance of sensors upon interaction with target analytes can be effectively accessed by evaluating several key sensing parameters. First, the response can be defined by calculating normalized resistance (R) change upon exposure of analytes according to eq 1, where R0 represents the baseline resistance observed under balance gas, and Rgas represents the resistance measured under gas exposure.
![]() |
1 |
In addition, the response also can be defined in terms of current (I) or conductance (G) from eq 2.
![]() |
2 |
The response value is indeed a crucial parameter for comparing the performance of sensors. However, it should be noted that the two definitions of response, conductivity and resistivity, are not equivalent due to the inverse relationship between them. Also, the type of balance gas, which affects R0 and the adsorption of analytes, should be considered when comparing the response. As a similar parameter, sensitivity (S) represents the ability of sensors to detect small changes in gas concentration. It is calculated from the slope (dy/dx) in the response (y) vs concentration (x) curve. The response and sensitivity of 2D c-MOFs can be influenced by factors such as surface area, number of binding sites, and gas accessibility. A higher surface area allows for enhanced surface interaction between the 2D c-MOFs and the analyte, promoting increased adsorption and improved sensor response. The number of binding sites available in the sensing material directly affects the capacity for interaction with the analyte, thereby impacting the response of the sensor. Furthermore, the accessibility of the gas molecules to the binding sites within the 2D c-MOFs can influence the response, as restricted or limited access may hinder effective sensing. Considering these factors is crucial for optimizing the performance of 2D c-MOFs as chemiresistive sensors. These factors can be further tailored in 2D c-MOFs through the selection of suitable linkers and control of morphology.
The response and recovery speeds are also crucial parameters for determining the performance of gas sensors. The response time (tres) refers to the duration required for the sensor’s response to dynamically reach 90% of the maximum response during exposure to the target gas. On the other hand, the recovery time (trec) is the time taken for the signal to return to 10% of the maximum response after the gas has been purged. In the case of 2D c-MOFs, sluggish response and recovery times are common issues due to their numerous binding sites and the operation of sensors at room temperature. However, these issues can be addressed through appropriate design strategies. One approach is to optimize the porous structure of the 2D c-MOFs to facilitate efficient gas diffusion and enhance the interaction with analyte molecules. Furthermore, the utilization of catalysts within the 2D c-MOFs can also accelerate the response and recovery speeds by promoting the desired chemical reactions involved in the sensing process. Another straightforward solution is to activate sensors using external stimuli. In conventional sensing materials like metal oxides, thermal activation through heating facilitates gas adsorption/desorption kinetics. However, this may not be suitable for MOFs-based sensors due to the relatively low thermal stability of MOFs. In this context, activation by light can be an excellent alternative to enhance the sensing speed of 2D c-MOFs sensors. Notably, 2D c-MOFs have narrower optical bandgaps (<1 eV) compared to insulating 3D MOFs, enabling efficient absorption of photon energy under visible light.88,94,99,100
Lastly, selectivity is a fundamental parameter that characterizes the ability of sensors to distinguish the target gas from other interfering gases. It is typically evaluated by comparing the response of the target gas to the responses of different analytes at a specific concentration.
As previously mentioned, one of the distinctive characteristics of 2D c-MOFs is the presence of diverse binding sites for target gases. This property provides opportunities to control and manipulate the interactions between the analytes and the MOF structure. Although the exact mechanism of the interaction is not well-understood, precise control of the composition of metal nodes and organic linkers can potentially lead to the tuning of selectivity across different MOFs.
2.4. Fabrication in Electronic Devices
Solvothermal synthesis is a versatile and essential approach for preparing a wide range of 2D c-MOFs specifically designed for sensor devices. By controlling the solvothermal conditions or postsynthetic treatment, the size and shape of the 2D c-MOFs can be modulated, which in turn affects their interaction with gases. Detailed discussions regarding the specific methods for controlling morphology and their impact on gas sensing performance will be discussed in a subsequent section. Through the utilization of ex situ coating methods, such as drop-casting, spray-coating, and mechanical abrasion, sensing devices can be fabricated from the as-synthesized 2D c-MOFs. Since the film fabrication process is independent of MOF synthesis conditions, these methods allow for a wide range of 2D c-MOFs to be employed. However, these methods often result in poor adhesion and contact with substrate and/or electrode. Moreover, there is a risk of sample damage during preparation, which can impede the reproducible sensing measurements.57 Alternatively, 2D c-MOFs can be directly grown on a suitable substrate under solvothermal synthetic conditions (Figure 4a).101,102 The direct growth approaches generally enhance contact of devices and minimize damage to the samples. Nonetheless, achieving precise control over the thickness or orientation of the film is challenging in solvothermal synthesis.
Figure 4.
(a) Direct growth of M3(HHTP)2 polymer film. Reprinted with permission from ref (101). Copyright 2016 American Chemical Society (b) Synthesis of Cu3(BHT) at the liquid–liquid interface. Reprinted with permission under a Creative Commons CC BY License from ref (24). Copyright 2015 Springer Nature. (c) Synthesis of Ni3(HTTP)2 at the air–water interface assisted by the Langmuir–Blodgett method. Reprinted with permission from ref (103). Copyright 2015 Wiley-VCH. (d) Schematic illustration for the LBL assembly technique combined with spray-coating. Reprinted with permission from ref (98). Copyright 2017 Wiley-VCH. (e) Schematic illustration for the microfluidic channel-embedded solution shearing method. Reprinted with permission under a Creative Commons CC BY License from ref (104). Copyright 2021 Springer Nature.
To achieve more precise control over film fabrication, the liquid interfacial synthesis method can be employed.89 This approach allows the synthesis of 2D c-MOFs within a confined space at the interface, resulting in films with controlled thickness and orientation. For instance, Huang et al. demonstrated the synthesis of Cu3(BHT) at the interface of dichloromethane and water, where the linker and Cu precursors were dissolved separately (Figure 4b).24 The resulting Cu3(BHT) film exhibited controlled thickness (10–140 nm), with layers arranged in a parallel orientation. Another method for obtaining thin films of 2D c-MOFs is the Langmuir–Blodgett-assisted technique, which enables the formation of monolayers at the air–liquid interface (Figure 4c).103 In this process, ligands are spread over the liquid surface and compressed to form a dense monolayer. By introducing metal precursors onto the surface, monolayers of 2D c-MOFs can be synthesized. The films prepared at the liquid interface can be easily transferred onto sensing substrates.
Alternatively, a layer-by-layer (LBL) approach allows for the direct deposition of thin films of 2D c-MOFs on functionalized surfaces (Figure 4d).98 By repeated coating of linker and ligand layers, MOFs can be grown layer by layer with precise control over thickness (<2 nm per cycle). Another technique, microfluidic-based solution shearing, enables the large-area synthesis of very thin and uniform films of 2D c-MOFs on substrates (Figure 4e).104 In this method, MOF precursors are rapidly mixed and reacted within a microfluidic-channel embedded blade. The shearing of the blade facilitates controlled growth of the MOF from the reaction mixture at the edge of the meniscus (liquid–air interface). The thickness of the resulting thin film can be precisely controlled up to tens of nanometers by adjusting the shearing speed of the blade. While these thin-film fabrication methods require rather sophisticated techniques or specialized equipment, their high level of controllability over the growth of 2D c-MOFs offers advantages in developing high-performance and reliable sensing devices.
3. Applications and Strategies in Chemiresistors
3.1. Design and Selection of Metal Modes and Linkers
As mentioned above, the high structural tunability of MOFs is one of their key advantages as sensing materials. The diverse chemical properties arising from numerous combinations of organic linkers and metal nodes determine the interaction of MOFs with analytes. Although the exact mechanisms underlying these interactions have not been clearly elucidated, selecting or designing nodes and linkers should be the primary strategy to achieve the desired sensing performance. In this section, we introduce how different nodes or linkers affect the sensing properties of 2D c-MOFs and related strategies to further improve sensing performance based on their properties. The sensing performances of 2D c-MOF sensors with different nodes and linkers are summarized in Table 2.
Table 2. Summary of the Sensing Properties of 2D c-MOF Sensors Based on the Design of Metal Nodes and Linkers.
sensing material | response type | response | target gas | operating condition | ref |
---|---|---|---|---|---|
Cu3(HITP)2 | ΔG/G0 | 2.6%@10 ppm | NH3 | N2, r.t. | (105) |
Cu3(HHTP)2 | ΔI/I0 | 0.7%@80 ppm | NH3 | N2, r.t. | (101) |
ΔI/I0 | –1.8%@80 ppm | NO | N2, r.t. | ||
ΔI/I0 | 0.5%@80 ppm | H2S | N2, r.t. | ||
Ni3(HHTP)2 | ΔI/I0 | –1.7%@80 ppm | NO | N2, r.t. | (101) |
ΔI/I0 | 4.2%@80 ppm | H2S | N2, r.t. | ||
Ni3(HHTP)2 on cotton | –ΔG/G0 | –49%@80 ppm | NO | N2, r.t. | (102) |
–ΔG/G0 | 98%@80 ppm | H2S | N2, r.t. | ||
Ni3(HITP)2 on cotton | –ΔG/G0 | 81%@80 ppm | NO | N2, r.t. | (102) |
–ΔG/G0 | 97%@80 ppm | H2S | N2, r.t. | ||
0.1 mol % HITP doped Cu-HHTP | ΔR/R0 | 240%@100 ppm | NH3 | air, r.t | (62) |
10 mol % HITP doped Cu-HHTP | ΔR/R0 | 90%@100 ppm | benzene | air, r.t | (62) |
Cu-HHTP-THQ | ΔI/I0 | 21.6%@100 ppm | NH3 | air, r.t. | (61) |
Cu-THQ | ΔI/I0 | 63.5%@100 ppm | NH3 | air, r.t | (61) |
Cu3(HIB)2 | S (ΔG/G0) | –0.62% (100 ppm)–1 | CO2 | 80 RH% air, r.t. | (107) |
CoPc–Cu | –ΔG/G0 | 27.4%@80 ppm | CO | N2, r.t | (108) |
NiPc–Cu | –ΔG/G0 | 18.9%@80 ppm | CO | N2, r.t | (108) |
NiPc–Ni | –ΔG/G0 | 43%@80 ppm | NH3 | N2, r.t. | (60) |
–ΔG/G0 | 64%@80 ppm | H2S | N2, r.t. | ||
–ΔG/G0 | –657%@1 ppm | NO | N2, r.t. | ||
NiPc–Cu | –ΔG/G0 | 45%@80 ppm | NH3 | N2, r.t. | (60) |
–ΔG/G0 | 98%@80 ppm | H2S | N2, r.t. | ||
–ΔG/G0 | –397%@1 ppm | NO | N2, r.t. | ||
2D Cu–salphen-MOF | ΔI/I0 | 766%@100 ppm | NO2 | Air, r.t | (109) |
3.1.1. Tuning Metal Nodes
Metal nodes of 2D c-MOFs are commonly regarded as the most likely binding sites for analytes. These metal nodes often have coordination sites that can interact with target gases, leading to selective sensing capabilities. Therefore, changing nodes with different metals significantly affects the sensing behavior of 2D c-MOFs. Campbell et al. first reported 2D c-MOF-based chemiresistors using Cu3(HITP)2 and Ni3(HITP)2, which are composed of the same linker but different metals.105 Cu3(HITP)2 exhibited increased conductance (ΔG/G0 = 2.6%) upon exposure to 10 ppm of NH3, while Ni3(HITP)2 showed no significant response toward NH3. This result is consistent with the study conducted by the Mirica group, which proposed that redox-active Cu nodes are responsible for the electronic interaction with NH3, while inactive Ni nodes do not exhibit such interaction.95 Campbell et al. further investigated the different sensing behaviors of these MOFs toward 16 other VOCs (aliphatic/aromatic hydrocarbons, alcohols, ketones, and amines).97 Cu3(HITP)2 and Ni3(HITP)2 exhibited opposite changes in conductance toward most of the targets (Figure 5a). Considering that the charge transfer mechanism plays a crucial role in chemiresistive sensors, these results imply that the different charge densities associated with metal nodes (e.g., d8 NiII vs d9 CuII) can have a substantial impact on the sensing response toward a wide range of targets. Similar behaviors are observed for 2D c-MOFs based on the HHTP linker, the same triphenylene core with −OH groups. While Cu3(HHTP)2 sensors showed a positive conductance change to 80 ppm of NH3, the Ni counterparts exhibited no significant response (Figure 5b).101 Moreover, it is worth noting that more than two different nodes can be simultaneously incorporated in isoreticular 2D c-MOFs. By mixing and controlling the compositions of Cu, Ni, and Co nodes in M3M′3–x(HITP)2, continuous variation of the electrical conductivity over 6 orders of magnitude can be achieved.88 Although the sensing properties of these bimetallic MOFs have not been studied, this approach offers an additional option for fine-tuning the sensing characteristics of 2D c-MOFs.
Figure 5.
(a) Sensing responses of the MOF array to representative examples from different categories of VOCs. Reprinted with permission from ref (97). Copyright 2015 American Chemical Society. (b) Sensing traces of Cu3(HHTP)2 (orange) and Ni3(HTTP)2 (blue) toward 80 ppm of NH3, NO, and H2S. (c) Sensing responses of Cu3(HHTP)2 and Ni3(HTTP)2 with different concentrations of NH3, NO, and H2S and their principal component analysis results. Adapted with permission from ref (101). Copyright 2016 American Chemical Society.
Besides, the 2D c-MOF sensors with altered cross-sensitivities depending on the choice of nodes can be combined into an array to more precisely discriminate multiple gases. For example, an array of Ni– and Cu–HHTP-based sensors can differentiate 10–80 ppm of NO, NH3, and H2S using principal component analysis (PCA), which is known as one of the pattern recognition approaches (Figure 5c).101 However, it should be noted that the selection of the node is not solely responsible for the sensing characteristics of 2D c-MOFs and may not always result in different sensitivity. For example, all Cu–, Co–, Ni–, and Fe–HHTP-based sensors exhibited similar responses toward NO.86
3.1.2. Controlling Organic Linkers
Linkers, as another main component of 2D c-MOFs, also play a crucial role in influencing sensing properties. Similar to the case of changing metal nodes, substituting the linker with different heteroatoms can lead to contrasting responses in gas sensing behavior. For example, Ni3(HITP)2 exhibits a negative conductance change (ΔG/G0 = −81%) in response to 80 ppm of NO, which is opposite to its HHTP counterpart (ΔG/G0 = 49%) (Figure 6a).102 The opposing behaviors may arise from linker-dependent variations in different charge carrier types (p or n) or ligand-specific interactions with analytes (−OH and −NH2). The different activity of 2D c-MOFs with different linkers, often combined with the “tuning nodes” approach, can further enhance gas discrimination capabilities through PCA. Additionally, the combination of isostructural ligands in 2D c-MOFs offers the ability to further tune sensor performance. Wu et al. demonstrated this interesting feature by creating a dual-ligand Cu-HHTP/HITP thin film with a doping level of approximately up to 10 mol % HITP.62 By controlling the doping concentration of HITP, cross-sensitivity between various gases (amines, benzenes, hydrogen, and ketones) was gradually modulated (Figure 6b). Notably, the 10 mol% HITP-doped Cu-HHTP exhibited a higher response to 100 ppm of benzene (ΔR/R0 = 90%) compared to NH3 (ΔR/R0 = 80%), which contrasts with the selectivity observed in the 0.1 mol % HITP-doped counterpart (ΔR/R0 = 75% to benzene and 240% to NH3). This reversed selectivity is intriguing since typical 2D c-MOF sensors are more sensitive to highly reducing NH3 compared to nonpolar VOCs. The authors proposed that the n-type doping effect of HITP reduces the sensitivity toward electron-donating reducing gases. They also suggested that the interaction of abundant defects formed by HITP doping have influenced the selectivity of the 2D c-MOF sensors.
Figure 6.
(a) Sensing traces of Ni3(HITP)2 (blue) and Ni3(HHTP)2 (red) toward NO and H2S. Reprinted with permission from ref (102). Copyright 2017 American Chemical Society. (b) The selectivity chart of Cu-HHTP with different doping concentrations of HITP. Reprinted with permission from ref (62). Copyright 2021 Springer Nature Group. (c) Structures of isostructural 2D c-MOFs with THQ, HHTP, and HITP organic ligands. (d) Comparison of conductivities and the response values of the isostructural 2D c-MOFs. Reprinted with permission from ref (61). Copyright 2021 Royal Society of Chemistry. (e) Schematic illustration of Cu3(HIB)2 for CO2 sensing. Reprinted with permission from ref (107). Copyright 2019 American Chemical Society.
Furthermore, two structurally distinct ligands also can be mixed to develop a new structure of 2D c-MOFs. The Kitagawa research group successfully synthesized Cu–HHTP–THQ by hybridizing HHTP (molecular size: 0.97 nm) and much smaller THQ (tetrahydroxy-1,4-quinone, molecular size: 0.55 nm).61,106 Cu–HHTP–THQ has a moderate pore size (1.71 nm) and electrical conductivity (2.53 × 10–5 S/cm), intermediate to that of Cu–HHTP and Cu–THQ. With high porosity and moderate conductivity, Cu–HHTP–THQ showed high sensitivity and linearity toward NH3 concentrations below 100 ppm. Furthermore, a comparative study on Cu–THQ, Cu–HHTP–THQ, Cu–HHTP, and Cu–HITP was conducted to investigate the influence of linker-dependent properties on sensing performance (Figure 6c).61 A negative relationship was observed between electrical conductivity of 2D c-MOF and their sensing response toward 100 ppm of NH3. This suggests that 2D c-MOF with lower intrinsic conductivity may exhibit higher sensitivity to charge transfer from low-concentration gases (Figure 6d). Additionally, the different porous structures of the MOFs influenced the sensing kinetics. Cu3(THQ)2, with the smallest 1D honeycomb pore size (1.31 nm), showed much slower reaction and recovery speeds compared to other MOFs with larger pores (1.71–2.18 nm), attributable to sluggish gas diffusion through the small 1D pores. However, sensing speeds of the 2D c-MOFs with larger pores were found to be influenced by the overall pore volume (including interparticle pores and other mesopores determined by morphologies) rather than solely relying on the size of the honeycomb pores.
Ligands with a benzene core, such as THQ, represent the smallest linker units in 2D c-MOFs. Their subnanoporous structure may hinder the mass transport of large analytes and limit the uptake of high gas concentrations. However, their relatively dense structures are advantageous as gas sensors in certain aspects. Stassen et al. reported CO2 sensors using Cu3(HIB)2 based on acid–base interaction between CO2 and amines in HIB (Figure 6e).107 Due to the small size of HIB, Cu3(HIB)2 can have a high density of amine groups serving as gas binding sites, while the pore size is large enough (0.8 nm) to allow CO2 molecules to access the pores. Accordingly, Cu3(HIB)2 exhibited a sensitivity of −0.62% (100 ppm)−1, which is higher than the analogous Cu3(HITP)2 with a lower density of amine groups and even comparable to commercial nondispersive infrared (NDIR) sensors. Furthermore, the nanopores of Cu3(HIB)2 can be fully hydrated even at very low humidity levels (∼10% RH). Owing to this unique property, Cu3(HIB)2-based sensors can effectively minimize the effect of humidity, maintaining an invariant CO2 response above 10% RH. These characteristics set Cu3(HIB)2-based sensors apart from other sensors that are unable to maintain such invariant CO2 response in humid conditions.
3.1.3. Metallolinker-Based 2D c-MOFs
Recently, metallophthalocyanines (MPc) have been utilized to synthesize high-performance chemiresistors in the form of 2D c-MOFs. One notable characteristic of MPc-based 2D c-MOFs is that the linkers themselves contain metal sites, which increases the diversity of 2D c-MOFs and provides additional gas binding sites for intriguing sensing properties. The Mirica research group demonstrated the utility of these additional metal sites in MPc-based 2D c-MOFs for gas sensing.108 They synthesized CoPc–Cu and NiPc–Cu by connecting the Cu node with Co- or Ni-centered MPc. CoPc–Cu exhibited an enhanced response toward CO (−ΔG/G0 = 27.4%) compared to that of analogous NiPc–Cu (−ΔG/G0 = 18.9%) (Figure 7a). Furthermore, CoPc–Cu had a much lower detection limit (0.53 ppm) compared to NiPc–Cu (3.0 ppm). DRIFTS analysis indicated that Cu nodes were primarily responsible for CO binding in both 2D c-MOFs. However, the DFT calculation revealed that the Co sites serve as stable binding sites for CO (ΔE = −6.5 kJ/mol), which is comparable to Cu sites. On the other hand, the Ni sites are found to be unfavorable for CO binding (ΔE = 16.0 kJ/mol) (Figure 7b). This suggests that the metal center in MPc can serve as a secondary interaction site to further modulate the sensing response. In addition to changing the metal center, substituting the node and linker in isoreticular MPc-based 2D c-MOFs also contributes to the tuning of sensing properties. Meng et al. investigated the sensing properties of analogous MOFs with the combinations of Cu/Ni nodes with NiPc- and Ni-centered naphthalocyanine (NiNPc) linkers (Figure 7c).60 Intriguingly, different combinations of nodes and linkers resulted in noticeable variations in sensitivity, selectivity, kinetics, and reversibility toward NO, H2S, and NH3 (Figure 7d), respectively. For instance, NiPc-based MOFs exhibited higher responses to NO compared to NiNPc, and the Cu node was more efficient in detecting H2S than Ni.
Figure 7.
(a) Sensing traces of CoPc–Cu and NiPc–Cu upon exposure to 80–10 ppm of CO. (b) DFT-optimized structure of CO-adsorbed CoPc–Cu and NiPc–Cu and their binding energies. Reprinted with permission from ref (108). Copyright 2022 Wiley-VCH. (c) Synthetic scheme for isoreticular phthalocyanine and naphthalocyanine-based MOFs, NiPc–M, and NiNPc–M. (d) Sensing responses of NiPc–M and NiNPc–M MOFs to 40 ppm of NH3, 40 ppm of H2S, and 1 ppm of NO in dry nitrogen and in the presence of 5000 ppm of H2O. Reprinted with permission from ref (60). Copyright 2019 American Chemical Society. (e) The structure of nonplanar salphen and Cu–salphen-MOF. (f) Response of Cu–salphen-MOF toward NO and other 11 types of gases (100 ppm). Reprinted with permission from ref (109). Copyright 2023 Wiley-VCH.
Metallosalphen is a new type of metallolinker that can be utilized to construct 2D c-MOFs (Figure 7e).109 While typical 2D c-MOFs are synthesized from planar ligands, salphen has a nonplanar structure. However, salphen can be converted into a planar configuration through the coordination of a metal ion within its N2O2 pocket. In addition, six ortho-substituted hydroxyl groups are coordinated simultaneously to metal ions, forming a unique structure of 2D c-MOFs. The Cu–salphen MOF, owing to the additional Cu coordination in salphen, possesses two distinct Cu sites (Cu–N2O2 and Cu–O4) with higher metal density compared to typical 2D c-MOFs. Cu–salphen MOF exhibited a remarkably high response (ΔI/I0 = 786% to 100 ppm of NO2), which can be attributable to the high density of Cu sites (Figure 7f). DFT and XPS analyses have revealed that the unique Cu–N2O4 sites are particularly responsible for the high activity toward NO2.
3.2. Morphology Control
2D c-MOFs possess diverse gas binding sites that exhibit varying activity toward target gases. Furthermore, their anisotropic crystal structure leads to different chemical/electrical properties and gas diffusivities depending on their orientations. Hence, manipulating the morphology of 2D c-MOFs can result in significant differences in sensitivity, selectivity, and reaction dynamics. In this section, we introduce several interesting approaches for controlling the morphology including crystal growth control and thin-film fabrication, with the aim of improving gas sensing characteristics. Table 3 provides a summary of various 2D c-MOF sensors associated with morphological manipulation techniques.
Table 3. Summary of the Sensing Properties of 2D c-MOF Sensors with the Morphology Control Strategy.
sensing material | response type | response | target gas | operating condition | ref |
---|---|---|---|---|---|
Cu3(HHTP)2 nanoflakes | ΔR/R0 | 89.4%@5 ppm | NO2 | air, r.t | (99) |
crystalline Ni3(HIB)2 | ΔI/I0 | 0.8@80%RH | H2O | N2, r.t | (111) |
amorphous Ni3(HIB)2 | ΔI/I0 | 45@80%RH | H2O | N2, r.t | (111) |
Cu3(BHT) film | ΔR/R0 | 14.9%@100 ppm | NH3 | air, r.t | (112) |
Cu3(HHTP)2 film with LBL process | ΔR/R0 | 129%@100 ppm | NH3 | air, r.t. | (98) |
Cu3(HHTP)2-3C (3 LBL cycles) | ΔI/I0 | 55.9%@100 ppm | NH3 | air, r.t | (113) |
ΔI/I0 | 9.1%@100 ppm | H2 | air, r.t | ||
Cu3(HHTP)2-5C | ΔI/I0 | 197.6%@100 ppm | NH3 | air, r.t | (113) |
ΔI/I0 | 65%@100 ppm | H2 | air, r.t | ||
Cu3(HHTP)2-20C | ΔI/I0 | 133.9%@100 ppm | NH3 | air, r.t | (113) |
ΔI/I0 | 66.3%@100 ppm | H2 | air, r.t | ||
Cu3(HHTP)2 3D thin film | ΔR/R0 | 161%@100 ppm | NH3 | air, r.t | (114) |
Ni3(HITP)2 thin film | ΔR/R0 | 2085%@5 ppm | H2S | N2, r.t | (115) |
3.2.1. Control of Crystal Shape and Crystallinity
Reducing the size of crystallites is a universal strategy for enhancing sensitivity in gas sensing. Smaller crystal size provides advantages such as increased gas accessibility and effective surface area. Jo et al. investigated the effect of the crystal size of Cu3(HHTP)2 on its response to NO2.99 The sizes of Cu3(HHTP)2 nanoflakes were controlled by collecting the supernatant after centrifugation at different speeds (50–5000 rpm). The response gradually increased from 61.1% to 89.4% toward 5 ppm of NO2 as the particle size increased from 158 to 415 nm (Figure 8a). Yao et al. reported similar observations for the response of Cu3(HHTP)2 to NH3.61 In this case, Cu3(HHTP)2 nanosheets were synthesized by controlling the reaction conditions such as the solvent and precursor concentrations. Cu3(HHTP)2 nanosheets exhibited higher response and faster responding speed compared to rod-shaped counterparts synthesized under typical solvothermal conditions. This difference is attributed to the shorter diffusion length in the c-axis of nanosheets. Another useful approach for controlling crystal shape is the addition of a coordination modulator during synthesis. For example, the addition of ammonia during synthesis suppresses the growth of Cu3(HHTP)2 in the c-axis, resulting in nanoplatelets with a thickness of approximately 10 nm.110 The Cu3(HHTP)2 nanoplatelets can be spray-coated on a substrate with parallel orientation, further increasing the accessibility of 1D pores. The sensing devices fabricated from the spray-coating of nanoplatelets exhibited an immediate resistance change (ΔR/R0 = 140%) in response to MeOH.
Figure 8.
(a) Sensing responses for bulk and nanoflakes of Cu3(HHTP)2. Reprinted with permission under a Creative Commons CC BY-NC-ND License from ref (99). Copyright 2021 American Chemical Society. (b) Structure of amorphous Ni3(HIB)2 with missing-linker sites. (c) Humidity sensing responses of missing-linker Ni3(HIB)2 (Ni–HAB in the figure) and crystalline Ni3(HIB)2. Reprinted with permission from ref (111). Copyright 2021 American Chemical Society. (d) Surface morphologies of Cu3(BHT) film with different reaction times for film preparation. (e) Reaction-time-dependent NH3 response (bar chart) and Cu2c percentage (red dotted line) of Cu3(BHT) films. Reprinted with permission from ref (112). Copyright 2020 American Chemical Society.
Moreover, reducing crystallinity and generating defects can also be advantageous in gas sensing as it creates binding sites with higher activity. Liu et al. synthesized Ni–HIB by partially oxidizing hexaaminobenzene linkers (HIB) to pentaaminobenzene or tetraaminobenzene.111 The resulting Ni–HIB synthesized from the oxidized linkers contained numerous missing-linker defects, leading to a loss of crystallinity (Figure 8b). The Ni–HIB with missing-linker defects showed a much higher response and fast responding speed (ΔI/I0 = 45, tres = 4.92 s) compared to crystalline Ni3(HIB)2 (ΔI/I0 = 0.8, tres = 36.77 s) toward H2O (80% relative humidity) (Figure 8c). This high activity is attributed to the hydroxyl group substituted at the missing-linker sites, which facilitates interaction with water molecules through hydrogen bonding. In addition, Chen et al. reported an increase in response of Cu3(BHT) with low crystallinity.112 During the synthesis of the Cu3(BHT) film, it was observed that reducing the reaction time led to the formation of less crystalline sample with an increased number of defective Cu sites on the surface of the crystals (Figure 8d). The authors observed a synergistic relationship between the NH3 response and the number of defective Cu sites (Figure 8e). Through DFT analysis, it was revealed that the defective Cu sites are responsible for the high responses by providing strong binding energy toward NH3.
3.2.2. Fabrication of Thin Film
Fabricating sensing materials in thin-film form is a straightforward approach that promotes gas diffusion and improves sensitivity. The Xu research group developed a method for constructing thin films of 2D c-MOFs using a LBL assembly technique combined with spray-coating.98 By employing LBL processes, Cu3(HHTP)2 thin films were successfully fabricated with high orientation, excellent crystallinity, and well-controlled thickness ranging from 20 to 100 nm. The Cu3(HHTP)2 thin film exhibited a resistance change of 129% toward 100 ppm of NH3, which is 5 times higher than that of the powder-based thick-film sensor (Figure 9a). Additionally, the thin-film sensors exhibited faster gas diffusion rate compared to a powder-based sensor, with response and recovery speeds increased by 54% and 10%, respectively. In a subsequent study, the Xu group further investigated the LBL growth process of thin films and its impact on sensing properties (Figure 9b).113 During the initial few growth cycles, in-plane growth of grains was self-limited to sizes below 40 nm. The low crystallinity of the films was self-repaired after 10 cycles. Consequently, the responses of the thin film toward H2 and NH3 significantly increased between 5 and 10 growth cycles, reflecting the formation of a charge transport pathway for sensing signal transduction. Beyond the 10th cycle, the response toward NH3 decreased due to reduced mass transport in thicker film. However, this effect had a lesser impact on H2 due to its smaller kinetic diameter. Furthermore, the thin film can also be grown on a 3D substrate with the LBL process to further enhance the exposed surface area, leading to improved mass transport.114 Lin et al. successfully fabricated a 3D Cu3(HHTP)2 thin film on a substrate consisting of TiO2 nanowire arrays (Figure 9c).114 The 3D thin-film-based sensor exhibited a 2.5 times higher response and a 2.3 times faster responding speed compared to the 2D thin-film sensor (Figure 9d).
Figure 9.
(a) NH3 sensing responses of Cu3(HHTP)2 film with different LBL growth cycles. Reprinted with permission from ref (98). Copyright 2017 Wiley-VCH. (b) Growth-cycle-dependent responses of the Cu3(HHTP)2 film toward 100 ppm of NH3 and H2. Reprinted with permission from ref (113). Copyright 2022 Wiley-VCH. (c) Schematic illustration for fabrication of oriented MOF 3D thin film using the LBL assembly technique. (d) Sensing performance for powder, 2D thin film, and 3D thin film of Cu3(HHTP)2. Reprinted with permission from ref (114). Copyright 2021 Wiley-VCH. (e) The schematic figure for fabrication of Ni3(HITP)2 thin-film processing using the microfluidic-based solution shearing (f) Dynamic sensing trace of bulk and thin-film Ni3(HITP)2 toward 5 ppm of H2S. Reprinted with permission from ref (115). Copyright 2022 Wiley-VCH.
Lee et al. demonstrated the fabrication of Ni3(HITP)2 thin-film sensors using microfluidic-assisted solution shearing and postsynthetic crystallization techniques (Figure 9e).115 Initially, thin films of precrystallized Ni3(HITP)2 were formed through solution shearing, with precise control over the film thickness ranging from 11 to 64 nm. Subsequent amine treatment was applied to induce deprotonation of the linker, resulting in rapid crystallization of ∼30 nm crystallites with parallel orientation in the thin film. The Ni3(HITP)2 thin film exhibited a significantly higher response to H2S, with a 30-fold increase compared to the powder-based sensor (Figure 9f). The enhanced sensitivity can be attributed to the thin-film structure, parallel orientation, and uniform nanocrystallites achieved during the fabrication process.
3.3. Composites of 2D c-MOFs
Combining sensing materials with other materials is a well-known strategy for tuning the sensing properties. This approach can be highly effective for 2D c-MOFs, especially considering the incomplete understanding of their sensing mechanisms. By incorporating 2D c-MOFs with other materials whose effects on gases are relatively predictable, a more rational option can be provided for tuning the sensing characteristics in a desired manner, rather than directly modifying the 2D c-MOFs themselves. In this section, we present several examples of foreign materials in 2D c-MOF composites, along with their effects on sensing performance. The sensing performances of 2D c-MOF composite sensors are summarized in Table 4.
Table 4. Summary of the Sensing Properties of 2D c-MOF-Based Composite Sensors.
sensing material | response type | response | target gas | operating condition | ref |
---|---|---|---|---|---|
Cu3(HHTP)2 | ΔR/R0 | –29.95%@5 ppm | NO2 | air, r.t. | (120) |
Pd@Cu3(HHTP)2 | ΔR/R0 | –62.11%@5 ppm | NO2 | air, r.t. | (120) |
Pt@Cu3(HHTP)2 | ΔR/R0 | –57.38%@5 ppm | NO2 | air, r.t. | (120) |
Co3(HITP)2 | Rgas/Rair | 4.25@100 ppm | H2S | air, r.t | (121) |
Au2–Co3(HITP)2 | Rgas/Rair | 5.10@100 ppm | H2S | air, r.t | (121) |
Pd2–Co3(HITP)2 | Rgas/Rair | 8.08@100 ppm | H2S | air, r.t | (121) |
Pt2–Co3(HITP)2 | Rgas/Rair | 6.38@100 ppm | H2S | air, r.t | (121) |
Cu3(HHTP)2 thin film | ΔR/R0 | –53.7%@3 ppm | NO2 | air, r.t. | (104) |
Pt@Cu3(HHTP)2 thin film | ΔR/R0 | –89.9%@3 ppm | NO2 | air, r.t. | (104) |
PtRu@Cu3(HHTP)2 | ΔR/R0 | –53.0%@2 ppm | NO2 | air, r.t | (122) |
Pt@Cu3(HHTP)2 | ΔR/R0 | –39.6%@2 ppm | NO2 | air, r.t | (122) |
Ru@Cu3(HHTP)2 | ΔR/R0 | –20.3%@2 ppm | NO2 | air, r.t | (122) |
Fe2O3–Cu3(HHTP)2 nanoflakes | ΔR/R0 | –64%@5 ppm | NO2 | air, light illumination | (99) |
Ni2[CuPc(NH)8]–OTMS | –ΔG/G0 | 3.22%@400 ppm | MeOH | N2, r.t | (124) |
Cu–TCPP-on-Cu–HHTP | ΔR/R0 | 153%@100 ppm | benzene | air, r.t. | (126) |
Ni–HHTP@UiO-66–NH2 | ΔR/R0 | 3.37@5 ppm | H2S | air, r.t | (129) |
3.3.1. Decoration of Nanoparticles
Various metal nanoparticles (NPs) have been employed as effective sensitizers in gas sensors, which activate sensing materials chemically or electronically.116 MOFs are especially advantageous for the strategy due to their cavities for stabilizing NPs.117−119 Koo et al. utilized this concept to combine NPs and Cu3(HHTP)2 for application in highly sensitive NO2 gas sensors.120 The well-defined porous structure of Cu3(HHTP)2 enables confined growth of ultrasmall Pt or Pd NPs (<2 nm) with high dispersity (Figure 10a). The responses of Pt@Cu3(HHTP)2 (ΔR/R0 = −57.38%) and Pd@Cu3(HHTP)2 (−62.11%) toward 5 ppm of NO2 are higher than that of pristine Cu3(HHTP)2 (−29.95%). In more detail, the Pt and Pd NPs affected the response in a distinct mechanism. Pt@Cu3(HHTP)2 exhibited a lower activation energy for the adsorption of NO2, implying that Pt acts as a chemical sensitizer through the spillover effect. Meanwhile, Pd acts as an electronic sensitizer in Pd@Cu3(HHTP)2. NO2 adsorption and charge transfer on Pd lower the Schottky barrier between NPs and Cu3(HHTP)2, resulting in an amplified signal (Figure 10b). Using similar approaches, Sun et al. reported the sensitization of Co3(HITP)2 by incorporating Pd, Au, and Pt NPs for highly sensitive H2S sensors.121 In addition, the NPs@2D-c-MOFs can be fabricated as thin films, resulting in further increased sensitivity. Kim et al. demonstrated the fabrication of Pt@Cu3(HHTP)2 thin films using in situ growth by microfluidic channel-embedded solution shearing.104 The Pt@Cu3(HHTP)2 thin film exhibited a higher response (ΔR/R0 = −89.9% to 3 ppm of NO2) compared to both the Cu3(HHTP)2 thin film and Cu3(HHTP)2 powder-based thick film (Figure 10c).
Figure 10.
(a) Schematic illustration for nanoparticle encapsulation of Cu3(HHTP)2 (b) Selective gas sensing performance of Cu3(HHTP)2, Pd@Cu3(HHTP)2, and Pt@Cu3(HHTP)2. Reprinted with permission under a Creative Commons CC BY License from ref (120). Copyright 2019 Wiley-VCH. (c) Sensing performance for Cu3(HHTP)2 powder, thin film, and thin film with Pt decoration using microfluidic channel-embedded solution shearing. Reprinted with permission under a Creative Commons CC BY License from ref (104) . Copyright 2021 Springer Nature. (d) Sensing responses of Pt, Ru, and PtRu NP- encapsulated Cu3(HHTP)2 for 2 ppm of NO2. Reprinted with permission from ref (122). Copyright 2021 Wiley-VCH. (e) Reversible NO2 sensing mechanism of Fe2O3–Cu3(HHTP)2 using visible light. (f) Repetitive sensing traces of Fe2O3–Cu3(HHTP)2 and Cu3(HHTP)2 toward 5 ppm of NO2 under blue light illumination. Reprinted with permission under a Creative Commons CC BY-NC-ND License from ref (99). Copyright 2021 American Chemical Society.
Furthermore, the use of bimetallic nanoparticles, which exhibit higher activity due to the synergistic effect between two different metal elements, can further enhance sensing performance. Dipolar interactions and site-specific growth of two different metals within the cavities of 2D c-MOFs allow the facile synthesis of ultrasmall bimetallic NPs, which is challenging in typical synthetic methods.122 Cu3(HHTP)2 with bimetallic PtRu NPs showed an enhanced response (ΔR/R0 = −53.0% to 2 ppm of NO2) compared to its monometallic Pt and Ru counterparts (Figure 10d). The improved response of PtRu@Cu3(HHTP)2 originates from the synergistic combination of the spillover effect by Pt and strong gas adsorption on Ru. With this approach, a diverse range of sensitization effects can be designed based on the combination of two metals, leading to further control of sensing properties in NPs@2D c-MOFs.
Additionally, Jo et al. utilized a mixture of Cu3(HHTP)2 and Fe2O3 nanoparticles for the NO2 sensor, focusing on reversibility.99 Pristine Cu3(HHTP)2 exhibits high sensitivity and selectivity toward NO2. However, the response toward NO2 is not reversible due to the strong adsorption of NO2, requiring additional activation for desorption. Light illumination on Cu3(HHTP)2 generates photogenerated holes to desorb NO2–(ads), enhancing reversibility at room temperature (Figure 10e). However, complete recovery cannot be achieved due to the fast recombination of photogenerated carriers before desorption. To address this issue, Fe2O3 nanoparticles (<50 nm) are decorated on Cu3(HHTP)2 by physical mixing, forming a heterojunction where photogenerated carriers can be separated with a longer lifetime. As a result, Fe2O3–Cu3(HHTP)2 exhibited completely reversible sensing behavior toward NO2 under light illumination (Figure 10f).
3.3.2. Hybridization of Hydrophobic/Hydrophilic Materials
Modulating the surface hydrophobicity of sensors represents a rational approach for discriminating gases based on their polarity.123 This strategy holds particular promise for 2D c-MOF-based sensors, which typically exhibit high sensitivity toward a limited range of highly polar molecules such as NH3, H2S, and NO2, while their ability to discriminate less-polar VOCs remains underdeveloped. A convenient method to manipulate the polarity of 2D c-MOFs involves incorporating other materials with varying degrees of hydrophobicity. Wang et al. conducted a study on the polarity-dependent sensing properties of phthalocyanine-based MOFs (Ni2[Cu(NHPc)]) by coating the surface of the MOFs with silanes possessing different functional groups (Figure 11a).124 Upon coating with octadecyltrimethoxysilane (OTMS), which features a long hydrophobic alkyl chain, the sensor exhibited a reduction in response (from ΔG/G0 = 4.7% to 1.7%) toward 600 ppm of H2O, along with an enhanced recovery speed. This can be attributed to the suppression of water adsorption by the hydrophobic OTMS layer, consequently enabling faster desorption. The grafting of propylamine, a relatively hydrophilic moiety, had minimal impact on the H2O sensing properties. Furthermore, OTMS-grafted 2D c-MOFs exhibited an enhanced response and recovery toward methanol compared to the pristine sample. They also observed a positive relationship between the response speeds and the polarity of the analytes (methanol > ethanol > isopropanol). Although the impact of OTMS on the sensing response to polar VOCs remains unclear and complex, their study offers additional opportunities for tuning the selectivity toward VOCs.
Figure 11.
(a) Schematic illustration for surface-modification of Ni2[CuPc(NH)8] 2D c-MOF film with APTMS, PTCS, and OTMS coating. Reprinted with permission from ref (124). Copyright 2021 Wiley-VCH. (b) Schematic illustration for the fabrication of MOF-on-MOF film. Reprinted with permission from ref (126). Copyright 2019 Wiley-VCH. (c) Schematic illustration for the synthesis of the heterostructured Ni–HHTP@UiO-66–NH2. Reprinted with permission from ref (129). Copyright 2022 Elsevier.
3.3.3. Integration with Other MOFs
As MOFs have various interactions with a wide range of gases, compositing 2D c-MOFs with other MOFs can offer intriguing sensing properties. One of the representative functions is the gas separation ability of MOFs which can be applied to modulate the selectivity of sensors.125 Xu’s group fabricated MOF-on-MOF film and demonstrated the selectivity tuning.126 As a 2D c-MOF layer, Cu3(HHTP)2 thin film was prepared by the LBL process (Figure 11b). The Cu3(HHTP)2 film showed considerable response to benzene (134% to 100 ppm), but the response was weaker compared to the response to polar NH3 (230% to 100 ppm). To reverse the selectivity, a Cu–TCPP layer was transferred onto the Cu3(HHTP)2 thin film. The Cu–TCPP layer limited the permeation of NH3 through strong binding at unsaturated Cu sites. Consequently, the Cu–TCPP-on-Cu3(HHTP)2 film showed a significantly reduced response to NH3 (from 230% to 94%), while maintaining a similar level of response to benzene (from 134% to 153%). It is noteworthy that the Cu–TCPP layer was synthesized independently and coated onto the Cu3(HHTP)2 film, overcoming their structural dissimilarity with a van der Waals gap. This approach enables the incorporation of a diverse range of MOFs known for their gas separation capabilities, thereby imparting selectivity toward different gases.127
Furthermore, MOFs can be integrated via coordination bonds to exert stronger effects on each other.128 Cho et al. synthesized a heterostructure of UiO-66–NH2 and Ni–HHTP by connecting them through coordination bonds (Figure 11c).129 The heterostructure was formed by initially synthesizing UiO-66–NH2 crystallites as the core, followed by connecting HHTP linkers to the Zr nodes of UiO-66–NH2. Subsequent growth of the Ni–HHTP shell resulted in well-connected core–shell heterostructures of the two MOFs. The heterostructured sensors exhibited a high surface area of 1071 m2/g, which was close to the predicted value based on their composition and surface areas (449 m2/g for Ni-HHTP and 1320 m2/g for UiO-66–NH2), indicating that the highly porous structures and numerous active sites of both MOFs were maintained in the heterostructures. Additionally, several unique properties were observed at the interface between the two MOFs. First, the interfacial coordination of HHTP at Zr nodes led to the formation of a high concentration of ligand-centered radicals, which are advantageous for redox reactions with analytes. Second, DFT calculations revealed that the highest occupied molecular orbital (HOMO) level was localized at the interfacial HHTP, enhancing the effectiveness of charge transfer between reducing gases. As a result, the Ni–HHTP@UiO-66–NH2 exhibited a higher response to 5 ppm of H2S (ΔR/R0 = 3.37) compared to pristine Ni–HHTP (ΔR/R0 = 0.71). Moreover, UiO-66–NH2 in the heterostructure provided secondary binding sites for H2S, facilitating its desorption more easily than in Ni–HHTP alone. Consequently, Ni–HHTP@UiO-66–NH2 showed improved reversibility toward H2S.
4. Summary and Perspectives
2D c-MOFs possess well-defined porous structures, high electrical conductivity, and numerous active sites. These features are highly tunable depending on the combination of metal nodes and linkers, making them unique materials in gas sensing applications. Their tunable pore sizes and densities of active sites impact the diffusion and adsorption of gases. Tailoring electrical properties of 2D c-MOFs can affect power consumption, signal noise, and potentially response of the sensors. In addition, the strategic combinations of nodes and linkers can give rise to various distinct binding characteristics, including Lewis or Brønsted acid, hydrogen bonding, and a range of other polar interactions. In addition, mechanisms underlying the conductivity changes are also diverse and complex. Along with charge transfer, unique conductivity-change mechanisms have been observed in 2D c-MOFs, such as the redox of linkers counterbalancing the charge shift in nodes and the disruption of interplanar interactions. Overall, the precise control and optimization of these unique properties of 2D c-MOFs hold significant potential for the development of high-performance gas sensors.
In this regard, various 2D c-MOFs-based devices have been demonstrated based on three approaches: (1) design and selection of linkers, (2) control of morphologies, (3) and integration of foreign materials with 2D c-MOFs. Among them, the design of metal nodes and linkers is the most crucial approach for determining the sensing properties of 2D c-MOFs. Controlling metal nodes significantly influences the binding affinity toward specific target gases as strong binding sites. They also affect the subsequent change in electrical signal through different redox activity or charge density. These differences lead to even contrasting responses toward certain targets. While linkers have similar influences, they determine the porous structure of 2D c-MOFs, affecting gas diffusivity and response speed. Meanwhile, metal-centered ligands further increase the chemical diversity by introducing additional metal sites. However, a large barrier in this strategy is the lack of a clear understanding of the relationship between chemical structures and sensing behaviors. To fully exploit the design potential of 2D c-MOFs, further investigations are needed to establish a comprehensive understanding of the intricate relationship between their structure and sensing properties. Alternatively, a more predictable approach involves the combination of two different nodes or linkers with finely tuned compositions. The mixed 2D c-MOFs may induce moderate properties that lie between those of the two pristine MOFs. However, the synthesis of such mixed 2D c-MOFs, especially those with linkers possessing different cores, poses a significant challenge. To overcome this challenge, innovative synthetic methods should be further developed.
The morphology control approach in 2D c-MOFs offers an additional pathway to achieve distinct properties without altering the nodes or linkers. By manipulating synthetic conditions or applying postsynthetic treatments, the shape and size of the crystals can be tuned. Smaller crystallites, particularly in the form of nanosheets, facilitate gas diffusion within the interior volume, making them advantageous for gas sensing. Reducing crystallinity and introducing defect sites can enhance the gas binding activity. However, these approaches sacrifice the electrical conductivities of the 2D c-MOFs, requiring careful optimization to balance the trade-off. Instead, through various techniques, the thickness of the 2D c-MOFs film can be reduced to <100 nm, significantly reducing the gas diffusion length. Some of these techniques (e.g., the LBL method) induce parallel orientation of the crystallites, further improving the accessibility of the pores and binding sites. The further development of the film fabrication techniques may enable the exploration of the orientation effect on other crystallographic planes (e.g., preferred exposure of edge sites) or different stacking patterns. Thus, thin-film 2D c-MOFs can provide additional functionality to the sensors. Additionally, efforts are being made to fabricate high-quality films with larger domains and fewer defects.56 Although sensor studies reveal that contrasting properties (small crystallites and numerous defects) tend to be advantageous for higher sensing response, high-quality films may allow the deeper study of intrinsic conductivity changes in 2D c-MOFs and provide valuable insights into sensing mechanisms.
Hybridization of 2D c-MOFs with other well-defined selectors or catalysts offers a promising approach for the rational design of 2D c-MOFs chemiresistors, providing more predictability. For instance, metal or metal-oxide nanoparticles can be integrated to utilize their well-known functions in gas sensors, such as chemical or electronic sensitizers. In addition, other materials with different polarities, surface charges, or acidities can be integrated to facilitate or inhibit gas adsorption. In particular, the decoration of metal nanoparticles has been widely employed in the field of gas sensors, and numerous prior studies have elucidated the sensing mechanisms of various metals. The well-defined pores of MOFs enable the incorporation of diverse metal nanoparticles. Despite these advantages and the extensive previous research, the decoration of metal nanoparticles in 2D c-MOF sensors has not been extensively studied, leaving significant room for exploring and enhancing the sensing characteristics of such composites. However, at the same time, the 1D pore channel of 2D c-MOFs is prone to pore blocking issues when nanoparticles are decorated on the surface. This approach inevitably compromises the porosity and surface reaction sites of 2D c-MOFs and requires careful optimization of the decoration method and loading amount of nanoparticles. To address this, other porous MOFs can be integrated as foreign materials, minimizing the pore blocking problem. These MOFs act as a sieving layer when arranged as a MOF-on-MOF structure, allowing the diffusion of target gases while effectively blocking interfering gases. Currently, the utilization of these MOF layers is primarily focused on filtering gases and controlling selectivity. However, they can also contribute to enhancing the sensitivity and response speed of sensors by harnessing other functionalities of MOFs, such as gas storage and preconcentration. In addition, MOFs can be integrated with stronger bonds, benefiting from the catalytic effects of adjacent MOFs and exploiting various intriguing interfacial properties. Relevant studies have shown that the sensitization effect of MOFs can be comparable to the decoration of catalytic nanoparticles. Although integrating structurally distinct MOFs and achieving strong coupling pose challenges, the construction of appropriate heterostructures can effectively overcome the surface blocking issue associated with conventional sensitizers in gas sensing applications.
Despite the existing challenges and the need for further advancements, the utilization of 2D c-MOFs holds immense promise in addressing current limitations in gas sensors. While there is still a significant gap in achieving the desired sensing properties, we believe that the rapid development of this emerging material class will play a pivotal role in the functionalization of next-generation conducting MOF-based chemiresistors in the near future.
Acknowledgments
This work was supported by Korea Technology and Information Promotion Agency for SMEs (Grant No. 00141845) and the National Research Foundation of Korea (NRF) grant funded by the Korea government, Ministry of Science and ICT (Development of Ultra-High Density Nanofibers Yarn Based Multi-Modal Gas Sensor Platform) (Grant No. 2020R1A2C301312712).
Author Contributions
All the authors take full responsibility for the content of this review and approved its submission. CRediT: Chungseong Park writing-original draft (lead), writing-review & editing (lead); Jong Won Baek writing-original draft (supporting), writing-review & editing (supporting); Euichul Shin writing-original draft (supporting), writing-review & editing (supporting).
The authors declare no competing financial interest.
References
- Daglar H.; Gulbalkan H. C.; Avci G.; Aksu G. O.; Altundal O. F.; Altintas C.; Erucar I.; Keskin S. Effect of metal-organic framework (MOF) database selection on the assessment of gas storage and separation potentials of MOFs. Angew. Chem., Int. Ed. 2021, 60, 7828–7837. 10.1002/anie.202015250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen Z.; Kirlikovali K. O.; Idrees K. B.; Wasson M. C.; Farha O. K. Porous materials for hydrogen storage. Chem. 2022, 8, 693–716. 10.1016/j.chempr.2022.01.012. [DOI] [Google Scholar]
- Sharma A.; Lim J.; Lah M. S. Strategies for designing metal-organic frameworks with superprotonic conductivity. Coord. Chem. Rev. 2023, 479, 214995. 10.1016/j.ccr.2022.214995. [DOI] [Google Scholar]
- Liu J.; Goetjen T. A.; Wang Q.; Knapp J. G.; Wasson M. C.; Yang Y.; Syed Z. H.; Delferro M.; Notestein J. M.; Farha O. K.; Hupp J. T. MOF-enabled confinement and related effects for chemical catalyst presentation and utilization. Chem. Soc. Rev. 2022, 51, 1045–1097. 10.1039/D1CS00968K. [DOI] [PubMed] [Google Scholar]
- Li Z.; Gao R.; Feng M.; Deng Y. P.; Xiao D.; Zheng Y.; Zhao Z.; Luo D.; Liu Y.; Zhang Z.; Wang D.; Li Q.; Li H.; Wang X.; Chen Z. Modulating metal-organic frameworks as advanced oxygen electrocatalysts. Adv. Energy Mater. 2021, 11, 2003291. 10.1002/aenm.202003291. [DOI] [Google Scholar]
- Yue K.; Liu J.; Zhu Y.; Xia C.; Wang P.; Zhang J.; Kong Y.; Wang X.; Yan Y.; Xia B. Y. In situ ion-exchange preparation and topological transformation of trimetal-organic frameworks for efficient electrocatalytic water oxidation. Energy Environ. Sci. 2021, 14, 6546–6553. 10.1039/D1EE02606B. [DOI] [Google Scholar]
- Zhou J.; Li R.; Fan X.; Chen Y.; Han R.; Li W.; Zheng J.; Wang B.; Li X. Rational design of a metal-organic framework host for sulfur storage in fast, long-cycle Li-S batteries. Energy Environ. Sci. 2014, 7, 2715–2724. 10.1039/C4EE01382D. [DOI] [Google Scholar]
- Thakur A. K.; Majumder M.; Patole S. P.; Zaghib K.; Reddy M. V. Metal-organic framework-based materials: advances, exploits, and challenges in promoting post Li-ion battery technologies. Mater. Adv. 2021, 2, 2457–2482. 10.1039/D0MA01019G. [DOI] [Google Scholar]
- Mehek R.; Iqbal N.; Noor T.; Amjad M. Z. B.; Ali G.; Vignarooban K.; Khan M. A. Metal-organic framework based electrode materials for lithium-ion batteries: a review. RSC Adv. 2021, 11, 29247–29266. 10.1039/D1RA05073G. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qian Q.; Asinger P. A.; Lee M. J.; Han G.; Mizrahi Rodriguez K.; Lin S.; Benedetti F. M.; Wu A. X.; Chi W. S.; Smith Z. P. MOF-based membranes for gas separations. Chem. Rev. 2020, 120, 8161–8266. 10.1021/acs.chemrev.0c00119. [DOI] [PubMed] [Google Scholar]
- Lin R.-B.; Xiang S.; Zhou W.; Chen B. Microporous metal-organic framework materials for gas separation. Chem. 2020, 6, 337–363. 10.1016/j.chempr.2019.10.012. [DOI] [Google Scholar]
- Koo W.-T.; Jang J.-S.; Qiao S.; Hwang W.; Jha G.; Penner R. M.; Kim I.-D. Hierarchical metal-organic framework-assembled membrane filter for efficient removal of particulate matter. ACS Appl. Mater. Interfaces 2018, 10, 19957–19963. 10.1021/acsami.8b02986. [DOI] [PubMed] [Google Scholar]
- Yuan H.; Li N.; Fan W.; Cai H.; Zhao D. Metal-Organic Framework Based Gas Sensors. Adv. Sci. 2022, 9, 2104374. 10.1002/advs.202104374. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kreno L. E.; Leong K.; Farha O. K.; Allendorf M.; Van Duyne R. P.; Hupp J. T. Metal-organic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105–1125. 10.1021/cr200324t. [DOI] [PubMed] [Google Scholar]
- Koo W.-T.; Jang J.-S.; Kim I.-D. Metal-organic frameworks for chemiresistive sensors. Chem. 2019, 5, 1938–1963. 10.1016/j.chempr.2019.04.013. [DOI] [Google Scholar]
- Koo W.-T.; Cho H.-J.; Kim D.-H.; Kim Y. H.; Shin H.; Penner R. M.; Kim I.-D. Chemiresistive hydrogen sensors: fundamentals, recent advances, and challenges. ACS Nano 2020, 14, 14284–14322. 10.1021/acsnano.0c05307. [DOI] [PubMed] [Google Scholar]
- Jeong S.-Y.; Kim J.-S.; Lee J.-H. Rational design of semiconductor-based chemiresistors and their libraries for next-generation artificial olfaction. Adv. Mater. 2020, 32, 2002075. 10.1002/adma.202002075. [DOI] [PubMed] [Google Scholar]
- Yang H.; He X.-W.; Wang F.; Kang Y.; Zhang J. Doping copper into ZIF-67 for enhancing gas uptake capacity and visible-light-driven photocatalytic degradation of organic dye. J. Mater. Chem. 2012, 22, 21849–21851. 10.1039/c2jm35602c. [DOI] [Google Scholar]
- Chen E.-X.; Fu H.-R.; Lin R.; Tan Y.-X.; Zhang J. Highly selective and sensitive trimethylamine gas sensor based on cobalt imidazolate framework material. ACS Appl. Mater. Interfaces 2014, 6, 22871–22875. 10.1021/am5071317. [DOI] [PubMed] [Google Scholar]
- DMello M. E.; Sundaram N. G.; Singh A.; Singh A. K.; Kalidindi S. B. An amine functionalized zirconium metal-organic framework as an effective chemiresistive sensor for acidic gases. Chem. Commun. 2019, 55, 349–352. 10.1039/C8CC06875E. [DOI] [PubMed] [Google Scholar]
- Travlou N. A.; Singh K.; Rodriguez-Castellon E.; Bandosz T. J. Cu-BTC MOF-graphene-based hybrid materials as low concentration ammonia sensors. J. Mater. Chem. A 2015, 3, 11417–11429. 10.1039/C5TA01738F. [DOI] [Google Scholar]
- Wang M.; Dong R.; Feng X. Two-dimensional conjugated metal-organic frameworks (2D c-MOFs): chemistry and function for MOFtronics. Chem. Soc. Rev. 2021, 50, 2764–2793. 10.1039/D0CS01160F. [DOI] [PubMed] [Google Scholar]
- Nam K. W.; Park S. S.; Dos Reis R.; Dravid V. P.; Kim H.; Mirkin C. A.; Stoddart J. F. Conductive 2D metal-organic framework for high-performance cathodes in aqueous rechargeable zinc batteries. Nat. Commun. 2019, 10, 4948. 10.1038/s41467-019-12857-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang X.; Sheng P.; Tu Z.; Zhang F.; Wang J.; Geng H.; Zou Y.; Di C.-a.; Yi Y.; Sun Y.; Xu W.; Zhu D. A two-dimensional π-d conjugated coordination polymer with extremely high electrical conductivity and ambipolar transport behaviour. Nat. Commun. 2015, 6, 7408. 10.1038/ncomms8408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu G.; Huang J.; Zang Y.; He J.; Xu G. Porous field-effect transistors based on a semiconductive metal-organic framework. J. Am. Chem. Soc. 2017, 139, 1360–1363. 10.1021/jacs.6b08511. [DOI] [PubMed] [Google Scholar]
- Yoon H.; Lee S.; Oh S.; Park H.; Choi S.; Oh M. Synthesis of bimetallic conductive 2D metal-organic framework (CoxNiy-CAT) and its mass production: enhanced electrochemical oxygen reduction activity. Small 2019, 15, 1805232. 10.1002/smll.201805232. [DOI] [PubMed] [Google Scholar]
- Li T.; Yin W.; Gao S.; Sun Y.; Xu P.; Wu S.; Kong H.; Yang G.; Wei G. The combination of two-dimensional nanomaterials with metal oxide nanoparticles for gas sensors: a review. Nanomaterials 2022, 12, 982. 10.3390/nano12060982. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim D.-H.; Kim J. K.; Oh D.; Park S.; Kim Y. B.; Ko J.; Jung W.; Kim I.-D. Ex-Solution Hybrids Functionalized on Oxide Nanofibers for Highly Active and Durable Catalytic Materials. ACS Nano 2023, 17, 5842–5851. 10.1021/acsnano.2c12580. [DOI] [PubMed] [Google Scholar]
- Song L.; Ahn J.; Xu L.; Baek J. W.; Shin E.; Kim I.-D. Facile Synthesis of Co3O4/CoMoO4 Heterostructure Nanosheets for Enhanced Acetone Detection. ACS Sens. 2022, 7, 3540–3550. 10.1021/acssensors.2c01969. [DOI] [PubMed] [Google Scholar]
- Kwon B.; Bae H.; Lee H.; Kim S.; Hwang J.; Lim H.; Lee J. H.; Cho K.; Ye J.; Lee S.; Lee W. H. Ultrasensitive N-channel graphene gas sensors by nondestructive molecular doping. ACS Nano 2022, 16, 2176–2187. 10.1021/acsnano.1c08186. [DOI] [PubMed] [Google Scholar]
- Choi J. H.; Lee J.; Byeon M.; Hong T. E.; Park H.; Lee C. Y. Graphene-based gas sensors with high sensitivity and minimal sensor-to-sensor variation. ACS Appl. Nano Mater. 2020, 3, 2257–2265. 10.1021/acsanm.9b02378. [DOI] [Google Scholar]
- Shekhirev M.; Lipatov A.; Torres A.; Vorobeva N. S.; Harkleroad A.; Lashkov A.; Sysoev V.; Sinitskii A. Highly selective gas sensors based on graphene nanoribbons grown by chemical vapor deposition. ACS Appl. Mater. Interfaces 2020, 12, 7392–7402. 10.1021/acsami.9b13946. [DOI] [PubMed] [Google Scholar]
- Kang J.-Y.; Koo W.-T.; Jang J.-S.; Kim D.-H.; Jeong Y. J.; Kim R.; Ahn J.; Choi S.-J.; Kim I.-D. 2D layer assembly of Pt-ZnO nanoparticles on reduced graphene oxide for flexible NO2 sensors. Sens. Actuators, B 2021, 331, 129371. 10.1016/j.snb.2020.129371. [DOI] [Google Scholar]
- Burman D.; Raha H.; Manna B.; Pramanik P.; Guha P. K. Substitutional doping of MoS2 for superior gas-sensing applications: a proof of concept. ACS Sens. 2021, 6, 3398–3408. 10.1021/acssensors.1c01258. [DOI] [PubMed] [Google Scholar]
- Pham T.; Li G.; Bekyarova E.; Itkis M. E.; Mulchandani A. MoS2-based optoelectronic gas sensor with sub-parts-per-billion limit of NO2 gas detection. ACS Nano 2019, 13, 3196–3205. 10.1021/acsnano.8b08778. [DOI] [PubMed] [Google Scholar]
- Koo W. T.; Cha J. H.; Jung J. W.; Choi S. J.; Jang J. S.; Kim D. H.; Kim I. D. Few-layered WS2 nanoplates confined in Co, N-doped hollow carbon nanocages: abundant WS2 edges for highly sensitive gas sensors. Adv. Funct. Mater. 2018, 28, 1802575. 10.1002/adfm.201802575. [DOI] [Google Scholar]
- Baek J. W.; Kim Y. H.; Ahn J.; Kim D.-H.; Shin H.; Ko J.; Park S.; Park C.; Shin E.; Jang J.-S.; Kim I. D. Galvanic replacement reaction in perovskite oxide for superior chemiresistors. J. Mater. Chem. A 2022, 10, 23282–23293. 10.1039/D2TA05338A. [DOI] [Google Scholar]
- Nascimento E. P.; Firmino H. C.; Neves G. A.; Menezes R. R. A review of recent developments in tin dioxide nanostructured materials for gas sensors. Ceram. Int. 2022, 48, 7405–7440. 10.1016/j.ceramint.2021.12.123. [DOI] [Google Scholar]
- Zhou T.; Zhang T. Recent progress of nanostructured sensing materials from 0D to 3D: overview of structure-property-application relationship for gas sensors. Small Methods 2021, 5, 2100515. 10.1002/smtd.202100515. [DOI] [PubMed] [Google Scholar]
- Burgués J.; Marco S. Low power operation of temperature-modulated metal oxide semiconductor gas sensors. Sensors 2018, 18, 339. 10.3390/s18020339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Potyrailo R. A.; Go S.; Sexton D.; Li X.; Alkadi N.; Kolmakov A.; Amm B.; St-Pierre R.; Scherer B.; Nayeri M.; et al. Extraordinary performance of semiconducting metal oxide gas sensors using dielectric excitation. Nat. Electron. 2020, 3, 280–289. 10.1038/s41928-020-0402-3. [DOI] [Google Scholar]
- Rasch F.; Postica V.; Schütt F.; Mishra Y. K.; Nia A. S.; Lohe M. R.; Feng X.; Adelung R.; Lupan O. Highly selective and ultra-low power consumption metal oxide based hydrogen gas sensor employing graphene oxide as molecular sieve. Sens. Actuators, B 2020, 320, 128363. 10.1016/j.snb.2020.128363. [DOI] [Google Scholar]
- Luo Y.; Abidian M. R.; Ahn J.-H.; Akinwande D.; Andrews A. M.; Antonietti M.; Bao Z.; Berggren M.; Berkey C. A.; Bettinger C. J.; et al. Technology roadmap for flexible sensors. ACS Nano 2023, 17, 5211–5295. 10.1021/acsnano.2c12606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang L.; Yi N.; Zhu J.; Cheng Z.; Yin X.; Zhang X.; Zhu H.; Cheng H. Novel gas sensing platform based on a stretchable laser-induced graphene pattern with self-heating capabilities. J. Mater. Chem. A 2020, 8, 6487–6500. 10.1039/C9TA07855J. [DOI] [Google Scholar]
- Kim D.-H.; Cha J.-H.; Shim G.; Kim Y. H.; Jang J.-S.; Shin H.; Ahn J.; Choi S.-Y.; Kim I.-D. Flash-thermochemical engineering of phase and surface activity on metal oxides. Chem. 2022, 8, 1014–1033. 10.1016/j.chempr.2021.12.003. [DOI] [Google Scholar]
- Yu T.; Cheng X.; Zhang X.; Sui L.; Xu Y.; Gao S.; Zhao H.; Huo L. Highly sensitive H2S detection sensors at low temperature based on hierarchically structured NiO porous nanowall arrays. J. Mater. Chem. A 2015, 3, 11991–11999. 10.1039/C5TA00811E. [DOI] [Google Scholar]
- Shin E.; Kim D.-H.; Cha J.-H.; Yun S.; Shin H.; Ahn J.; Jang J.-S.; Baek J. W.; Park C.; Ko J.; et al. Ultrafast Ambient-Air Exsolution on Metal Oxide via Momentary Photothermal Effect. ACS Nano 2022, 16, 18133–18142. 10.1021/acsnano.2c05128. [DOI] [PubMed] [Google Scholar]
- Xiangfeng C.; Siciliano P. CH3SH-sensing characteristics of LaFeO3 thick-film prepared by co-precipitation method. Sens. Actuators, B 2003, 94, 197–200. 10.1016/S0925-4005(03)00340-X. [DOI] [Google Scholar]
- Shin H.; Kim D.-H.; Jung W.; Jang J.-S.; Kim Y. H.; Lee Y.; Chang K.; Lee J.; Park J.; Namkoong K.; et al. Surface activity-tuned metal oxide chemiresistor: toward direct and quantitative halitosis diagnosis. ACS Nano 2021, 15, 14207–14217. 10.1021/acsnano.1c01350. [DOI] [PubMed] [Google Scholar]
- Huang X.; Gong Z.; Lv Y. Advances in metal-organic frameworks-based gas sensors for hazardous substances. TrAC, Trends Anal. Chem. 2022, 153, 116644. 10.1016/j.trac.2022.116644. [DOI] [Google Scholar]
- Jo Y. M.; Jo Y. K.; Lee J. H.; Jang H. W.; Hwang I. S.; Yoo D. J. MOF-Based Chemiresistive Gas Sensors: Toward New Functionalities. Adv. Mater. 2023, 2206842. 10.1002/adma.202206842. [DOI] [PubMed] [Google Scholar]
- Yao M.-S.; Li W.-H.; Xu G. Metal-organic frameworks and their derivatives for electrically-transduced gas sensors. Coord. Chem. Rev. 2021, 426, 213479. 10.1016/j.ccr.2020.213479. [DOI] [Google Scholar]
- Zhang L. T.; Zhou Y.; Han S. T. The role of metal-organic frameworks in electronic sensors. Angew. Chem. 2021, 133, 15320–15340. 10.1002/ange.202006402. [DOI] [PubMed] [Google Scholar]
- Li H.-Y.; Zhao S.-N.; Zang S.-Q.; Li J. Functional metal-organic frameworks as effective sensors of gases and volatile compounds. Chem. Soc. Rev. 2020, 49, 6364–6401. 10.1039/C9CS00778D. [DOI] [PubMed] [Google Scholar]
- Campbell M. G.; Dincǎ M. Metal-organic frameworks as active materials in electronic sensor devices. Sensors 2017, 17, 1108. 10.3390/s17051108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J.; Chen Y.; Feng X.; Dong R. Conductive 2D Conjugated Metal-Organic Framework Thin Films: Synthesis and Functions for (Opto-) electronics. Small Struct. 2022, 3, 2100210. 10.1002/sstr.202100210. [DOI] [Google Scholar]
- Ko M.; Mendecki L.; Mirica K. A. Conductive two-dimensional metal-organic frameworks as multifunctional materials. Chem. Commun. 2018, 54, 7873–7891. 10.1039/C8CC02871K. [DOI] [PubMed] [Google Scholar]
- Xie L. S.; Skorupskii G.; Dincă M. Electrically conductive metal-organic frameworks. Chem. Rev. 2020, 120, 8536–8580. 10.1021/acs.chemrev.9b00766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang R.; Lu L.; Chang Y.; Liu M. Gas sensing based on metal-organic frameworks: Concepts, functions, and developments. J. Hazard. Mater. 2022, 429, 128321. 10.1016/j.jhazmat.2022.128321. [DOI] [PubMed] [Google Scholar]
- Meng Z.; Aykanat A.; Mirica K. A. Welding metallophthalocyanines into bimetallic molecular meshes for ultrasensitive, low-power chemiresistive detection of gases. J. Am. Chem. Soc. 2019, 141, 2046–2053. 10.1021/jacs.8b11257. [DOI] [PubMed] [Google Scholar]
- Yao M.-S.; Wang P.; Gu Y.-F.; Koganezawa T.; Ashitani H.; Kubota Y.; Wang Z.-M.; Fan Z.-Y.; Otake K.-i.; Kitagawa S. A comparative study of honeycomb-like 2D π-conjugated metal-organic framework chemiresistors: conductivity and channels. Dalton Transactions 2021, 50, 13236–13245. 10.1039/D1DT02323C. [DOI] [PubMed] [Google Scholar]
- Wu A.-Q.; Wang W.-Q.; Zhan H.-B.; Cao L.-A.; Ye X.-L.; Zheng J.-J.; Kumar P. N.; Chiranjeevulu K.; Deng W.-H.; Wang G.-E.; et al. Layer-by-layer assembled dual-ligand conductive MOF nano-films with modulated chemiresistive sensitivity and selectivity. Nano Res. 2021, 14, 438–443. 10.1007/s12274-020-2823-8. [DOI] [Google Scholar]
- Jiang Q.; Xiong P.; Liu J.; Xie Z.; Wang Q.; Yang X. Q.; Hu E.; Cao Y.; Sun J.; Xu Y.; et al. A redox-active 2D metal-organic framework for efficient lithium storage with extraordinary high capacity. Angew. Chem., Int. Ed. 2020, 59, 5273–5277. 10.1002/anie.201914395. [DOI] [PubMed] [Google Scholar]
- Miner E. M.; Wang L.; Dincǎ M. Modular O2 electroreduction activity in triphenylene-based metal-organic frameworks. Chemical Science 2018, 9, 6286–6291. 10.1039/C8SC02049C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhong H.; Ly K. H.; Wang M.; Krupskaya Y.; Han X.; Zhang J.; Zhang J.; Kataev V.; Büchner B.; Weidinger I. M.; et al. A phthalocyanine-based layered two-dimensional conjugated metal-organic framework as a highly efficient electrocatalyst for the oxygen reduction reaction. Angew. Chem., Int. Ed. 2019, 58, 10677–10682. 10.1002/anie.201907002. [DOI] [PubMed] [Google Scholar]
- Feng D.; Lei T.; Lukatskaya M. R.; Park J.; Huang Z.; Lee M.; Shaw L.; Chen S.; Yakovenko A. A.; Kulkarni A.; et al. Robust and conductive two-dimensional metal-organic frameworks with exceptionally high volumetric and areal capacitance. Nat. Energy 2018, 3, 30–36. 10.1038/s41560-017-0044-5. [DOI] [Google Scholar]
- Park J.; Lee M.; Feng D.; Huang Z.; Hinckley A. C.; Yakovenko A.; Zou X.; Cui Y.; Bao Z. Stabilization of hexaaminobenzene in a 2D conductive metal-organic framework for high power sodium storage. J. Am. Chem. Soc. 2018, 140, 10315–10323. 10.1021/jacs.8b06020. [DOI] [PubMed] [Google Scholar]
- Sheberla D.; Sun L.; Blood-Forsythe M. A.; Er S. l.; Wade C. R.; Brozek C. K.; Aspuru-Guzik A.; Dincă M. High electrical conductivity in Ni3(2, 3, 6, 7, 10, 11-hexaiminotriphenylene)2, a semiconducting metal-organic graphene analogue. J. Am. Chem. Soc. 2014, 136, 8859–8862. 10.1021/ja502765n. [DOI] [PubMed] [Google Scholar]
- Jia H.; Yao Y.; Zhao J.; Gao Y.; Luo Z.; Du P. A novel two-dimensional nickel phthalocyanine-based metal-organic framework for highly efficient water oxidation catalysis. J. Mater. Chem. A 2018, 6, 1188–1195. 10.1039/C7TA07978H. [DOI] [Google Scholar]
- Lian Y.; Yang W.; Zhang C.; Sun H.; Deng Z.; Xu W.; Song L.; Ouyang Z.; Wang Z.; Guo J.; et al. Unpaired 3d electrons on atomically dispersed cobalt centres in coordination polymers regulate both oxygen reduction reaction (ORR) activity and selectivity for use in zinc-air batteries. Angew. Chem., Int. Ed. 2020, 59, 286–294. 10.1002/anie.201910879. [DOI] [PubMed] [Google Scholar]
- Wang M.; Shi H.; Zhang P.; Liao Z.; Wang M.; Zhong H.; Schwotzer F.; Nia A. S.; Zschech E.; Zhou S.; et al. Phthalocyanine-based 2D conjugated metal-organic framework nanosheets for high-performance micro-supercapacitors. Adv. Funct. Mater. 2020, 30, 2002664. 10.1002/adfm.202002664. [DOI] [Google Scholar]
- Cui J.; Xu Z. An electroactive porous network from covalent metal-dithiolene links. Chem. Commun. 2014, 50, 3986–3988. 10.1039/C4CC00408F. [DOI] [PubMed] [Google Scholar]
- Dong R.; Han P.; Arora H.; Ballabio M.; Karakus M.; Zhang Z.; Shekhar C.; Adler P.; Petkov P. S.; Erbe A.; et al. High-mobility band-like charge transport in a semiconducting two-dimensional metal-organic framework. Nat. Mater. 2018, 17, 1027–1032. 10.1038/s41563-018-0189-z. [DOI] [PubMed] [Google Scholar]
- Hmadeh M.; Lu Z.; Liu Z.; Gándara F.; Furukawa H.; Wan S.; Augustyn V.; Chang R.; Liao L.; Zhou F.; et al. New porous crystals of extended metal-catecholates. Chem. Mater. 2012, 24, 3511–3513. 10.1021/cm301194a. [DOI] [Google Scholar]
- Park J.; Hinckley A. C.; Huang Z.; Feng D.; Yakovenko A. A.; Lee M.; Chen S.; Zou X.; Bao Z. Synthetic routes for a 2D semiconductive copper hexahydroxybenzene metal-organic framework. J. Am. Chem. Soc. 2018, 140, 14533–14537. 10.1021/jacs.8b06666. [DOI] [PubMed] [Google Scholar]
- Nagatomi H.; Yanai N.; Yamada T.; Shiraishi K.; Kimizuka N. Synthesis and electric properties of a two-dimensional metal-organic framework based on phthalocyanine. Chem.—Eur. J. 2018, 24, 1806–1810. 10.1002/chem.201705530. [DOI] [PubMed] [Google Scholar]
- Clough A. J.; Skelton J. M.; Downes C. A.; De La Rosa A. A.; Yoo J. W.; Walsh A.; Melot B. C.; Marinescu S. C. Metallic conductivity in a two-dimensional cobalt dithiolene metal-organic framework. J. Am. Chem. Soc. 2017, 139, 10863–10867. 10.1021/jacs.7b05742. [DOI] [PubMed] [Google Scholar]
- Clough A. J.; Orchanian N. M.; Skelton J. M.; Neer A. J.; Howard S. A.; Downes C. A.; Piper L. F.; Walsh A.; Melot B. C.; Marinescu S. C. Room temperature metallic conductivity in a metal-organic framework induced by oxidation. J. Am. Chem. Soc. 2019, 141, 16323–16330. 10.1021/jacs.9b06898. [DOI] [PubMed] [Google Scholar]
- Mähringer A.; Jakowetz A. C.; Rotter J. M.; Bohn B. J.; Stolarczyk J. K.; Feldmann J.; Bein T.; Medina D. D. Oriented thin films of electroactive triphenylene catecholate-based two-dimensional metal-organic frameworks. ACS Nano 2019, 13, 6711–6719. 10.1021/acsnano.9b01137. [DOI] [PubMed] [Google Scholar]
- Wu H.; Zhang W.; Kandambeth S.; Shekhah O.; Eddaoudi M.; Alshareef H. N. Conductive metal-organic frameworks selectively grown on laser-scribed graphene for electrochemical microsupercapacitors. Adv. Energy Mater. 2019, 9, 1900482. 10.1002/aenm.201900482. [DOI] [Google Scholar]
- Song X.; Wang X.; Li Y.; Zheng C.; Zhang B.; Di C. a.; Li F.; Jin C.; Mi W.; Chen L.; et al. 2D semiconducting metal-organic framework thin films for organic spin valves. Angew. Chem., Int. Ed. 2020, 59, 1118–1123. 10.1002/anie.201911543. [DOI] [PubMed] [Google Scholar]
- Huang X.; Zhang S.; Liu L.; Yu L.; Chen G.; Xu W.; Zhu D. Superconductivity in a copper (II)-based coordination polymer with perfect Kagome structure. Angew. Chem. 2018, 130, 152–156. 10.1002/ange.201707568. [DOI] [PubMed] [Google Scholar]
- Huang X.; Yao H.; Cui Y.; Hao W.; Zhu J.; Xu W.; Zhu D. Conductive copper benzenehexathiol coordination polymer as a hydrogen evolution catalyst. ACS Appl. Mater. Interfaces 2017, 9, 40752–40759. 10.1021/acsami.7b14523. [DOI] [PubMed] [Google Scholar]
- Pal T.; Kambe T.; Kusamoto T.; Foo M. L.; Matsuoka R.; Sakamoto R.; Nishihara H. Interfacial synthesis of electrically conducting palladium bis(dithiolene) complex nanosheet. ChemPlusChem. 2015, 80, 1255–1258. 10.1002/cplu.201500206. [DOI] [PubMed] [Google Scholar]
- Dou J.-H.; Sun L.; Ge Y.; Li W.; Hendon C. H.; Li J.; Gul S.; Yano J.; Stach E. A.; Dincă M. Signature of metallic behavior in the metal-organic frameworks M3(hexaiminobenzene)2 (M = Ni, Cu). J. Am. Chem. Soc. 2017, 139, 13608–13611. 10.1021/jacs.7b07234. [DOI] [PubMed] [Google Scholar]
- Ko M.; Aykanat A.; Smith M. K.; Mirica K. A. Drawing sensors with ball-milled blends of metal-organic frameworks and graphite. Sensors 2017, 17, 2192. 10.3390/s17102192. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Day R. W.; Bediako D. K.; Rezaee M.; Parent L. R.; Skorupskii G.; Arguilla M. Q.; Hendon C. H.; Stassen I.; Gianneschi N. C.; Kim P.; Dincă M. Single crystals of electrically conductive two-dimensional metal-organic frameworks: Structural and electrical transport properties. ACS Cent. Sci. 2019, 5, 1959–1964. 10.1021/acscentsci.9b01006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen T.; Dou J.-H.; Yang L.; Sun C.; Libretto N. J.; Skorupskii G.; Miller J. T.; Dincă M. Continuous electrical conductivity variation in M3(hexaiminotriphenylene)2 (M = Co, Ni, Cu) MOF alloys. J. Am. Chem. Soc. 2020, 142, 12367–12373. 10.1021/jacs.0c04458. [DOI] [PubMed] [Google Scholar]
- Kambe T.; Sakamoto R.; Hoshiko K.; Takada K.; Miyachi M.; Ryu J.-H.; Sasaki S.; Kim J.; Nakazato K.; Takata M.; Nishihara H. π-Conjugated nickel bis(dithiolene) complex nanosheet. J. Am. Chem. Soc. 2013, 135, 2462–2465. 10.1021/ja312380b. [DOI] [PubMed] [Google Scholar]
- Kambe T.; Sakamoto R.; Kusamoto T.; Pal T.; Fukui N.; Hoshiko K.; Shimojima T.; Wang Z.; Hirahara T.; Ishizaka K.; Nishihara H.; et al. Redox control and high conductivity of nickel bis(dithiolene) complex π-nanosheet: a potential organic two-dimensional topological insulator. J. Am. Chem. Soc. 2014, 136, 14357–14360. 10.1021/ja507619d. [DOI] [PubMed] [Google Scholar]
- Skorupskii G.; Trump B. A.; Kasel T. W.; Brown C. M.; Hendon C. H.; Dincǎ M. Efficient and tunable one-dimensional charge transport in layered lanthanide metal-organic frameworks. Nat. Chem. 2020, 12, 131–136. 10.1038/s41557-019-0372-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lukose B.; Kuc A.; Heine T. The structure of layered covalent-organic frameworks. Chem.—Eur. J. 2011, 17, 2388–2392. 10.1002/chem.201001290. [DOI] [PubMed] [Google Scholar]
- Dou J.-H.; Arguilla M. Q.; Luo Y.; Li J.; Zhang W.; Sun L.; Mancuso J. L.; Yang L.; Chen T.; Parent L. R.; et al. Atomically precise single-crystal structures of electrically conducting 2D metal-organic frameworks. Nat. Mater. 2021, 20, 222–228. 10.1038/s41563-020-00847-7. [DOI] [PubMed] [Google Scholar]
- Rubio-Giménez V.; Almora-Barrios N.; Escorcia-Ariza G.; Galbiati M.; Sessolo M.; Tatay S.; Martí-Gastaldo C. Origin of the chemiresistive response of ultrathin films of conductive metal-organic frameworks. Angew. Chem. 2018, 130, 15306–15310. 10.1002/ange.201808242. [DOI] [PubMed] [Google Scholar]
- Stolz R. M.; Mahdavi-Shakib A.; Frederick B. G.; Mirica K. A. Host-Guest Interactions and Redox Activity in Layered Conductive Metal-Organic Frameworks. Chem. Mater. 2020, 32, 7639–7652. 10.1021/acs.chemmater.0c01007. [DOI] [Google Scholar]
- Meng Z.; Stolz R. M.; Mirica K. A. Two-dimensional chemiresistive covalent organic framework with high intrinsic conductivity. J. Am. Chem. Soc. 2019, 141, 11929–11937. 10.1021/jacs.9b03441. [DOI] [PubMed] [Google Scholar]
- Campbell M. G.; Liu S. F.; Swager T. M.; Dinca M. Chemiresistive sensor arrays from conductive 2D metal-organic frameworks. J. Am. Chem. Soc. 2015, 137, 13780–13783. 10.1021/jacs.5b09600. [DOI] [PubMed] [Google Scholar]
- Yao M. S.; Lv X. J.; Fu Z. H.; Li W. H.; Deng W. H.; Wu G. D.; Xu G. Layer-by-layer assembled conductive metal-organic framework nanofilms for room-temperature chemiresistive sensing. Angew. Chem., Int. Ed. 2017, 56, 16510–16514. 10.1002/anie.201709558. [DOI] [PubMed] [Google Scholar]
- Jo Y.-M.; Lim K.; Yoon J. W.; Jo Y. K.; Moon Y. K.; Jang H. W.; Lee J.-H. Visible-light-activated Type II heterojunction in Cu3(hexahydroxytriphenylene)2/Fe2O3 hybrids for reversible NO2 sensing: Critical role of π-π* transition. ACS Central Science 2021, 7, 1176–1182. 10.1021/acscentsci.1c00289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao J.; Geng S.; Chen Y.; Cheng P.; Zhang Z. Theoretical exploration and electronic applications of conductive two-dimensional metal-organic frameworks. Top. Curr. Chem. 2020, 378, 25. 10.1007/s41061-020-0288-6. [DOI] [PubMed] [Google Scholar]
- Smith M. K.; Jensen K. E.; Pivak P. A.; Mirica K. A. Direct self-assembly of conductive nanorods of metal-organic frameworks into chemiresistive devices on shrinkable polymer films. Chem. Mater. 2016, 28, 5264–5268. 10.1021/acs.chemmater.6b02528. [DOI] [Google Scholar]
- Smith M. K.; Mirica K. A. Self-organized frameworks on textiles (SOFT): conductive fabrics for simultaneous sensing, capture, and filtration of gases. J. Am. Chem. Soc. 2017, 139, 16759–16767. 10.1021/jacs.7b08840. [DOI] [PubMed] [Google Scholar]
- Dong R.; Pfeffermann M.; Liang H.; Zheng Z.; Zhu X.; Zhang J.; Feng X. Large-area, free-standing, two-dimensional supramolecular polymer single-layer sheets for highly efficient electrocatalytic hydrogen evolution. Angew. Chem., Int. Ed. 2015, 54, 12058–12063. 10.1002/anie.201506048. [DOI] [PubMed] [Google Scholar]
- Kim J.-O.; Koo W.-T.; Kim H.; Park C.; Lee T.; Hutomo C. A.; Choi S. Q.; Kim D. S.; Kim I.-D.; Park S. Large-area synthesis of nanoscopic catalyst-decorated conductive MOF film using microfluidic-based solution shearing. Nat. Commun. 2021, 12, 4294. 10.1038/s41467-021-24571-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Campbell M. G.; Sheberla D.; Liu S. F.; Swager T. M.; Dincǎ M. Cu3(hexaiminotriphenylene)2: an electrically conductive 2D metal-organic framework for chemiresistive sensing. Angew. Chem., Int. Ed. 2015, 54, 4349–4352. 10.1002/anie.201411854. [DOI] [PubMed] [Google Scholar]
- Yao M. S.; Zheng J. J.; Wu A. Q.; Xu G.; Nagarkar S. S.; Zhang G.; Tsujimoto M.; Sakaki S.; Horike S.; Otake K.; et al. A Dual-Ligand Porous Coordination Polymer Chemiresistor with Modulated Conductivity and Porosity. Angew. Chem., Int. Ed. 2020, 59, 172–176. 10.1002/anie.201909096. [DOI] [PubMed] [Google Scholar]
- Stassen I.; Dou J.-H.; Hendon C.; Dincǎ M. Chemiresistive sensing of ambient CO2 by an autogenously hydrated Cu3(hexaiminobenzene)2 framework. ACS Cent. Sci. 2019, 5, 1425–1431. 10.1021/acscentsci.9b00482. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aykanat A.; Meng Z.; Stolz R. M.; Morrell C. T.; Mirica K. A. Bimetallic Two-Dimensional Metal-Organic Frameworks for the Chemiresistive Detection of Carbon Monoxide. Angew. Chem., Int. Ed. 2022, 61, e202113665 10.1002/anie.202113665. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Su X.; Zhong Z.; Yan X.; Zhang T.; Wang C.; Wang Y.-X.; Xu G.; Chen L. Facile Synthesis of Metallosalphen-Based 2D Conductive Metal-Organic Frameworks for NO2 Sensing: Metal Coordination Induced Planarization. Angew. Chem., Int. Ed. 2023, 62, e202302645 10.1002/anie.202302645. [DOI] [PubMed] [Google Scholar]
- Hoppe B.; Hindricks K. D.; Warwas D. P.; Schulze H. A.; Mohmeyer A.; Pinkvos T. J.; Zailskas S.; Krey M. R.; Belke C.; König S.; et al. Graphene-like metal-organic frameworks: morphology control, optimization of thin film electrical conductivity and fast sensing applications. CrystEngComm 2018, 20, 6458–6471. 10.1039/C8CE01264D. [DOI] [Google Scholar]
- Liu C.; Gu Y.; Liu C.; Liu S.; Li X.; Ma J.; Ding M. Missing-linker 2D conductive metal organic frameworks for rapid gas detection. ACS sens. 2021, 6, 429–438. 10.1021/acssensors.0c01933. [DOI] [PubMed] [Google Scholar]
- Chen X.; Lu Y.; Dong J.; Ma L.; Yi Z.; Wang Y.; Wang L.; Wang S.; Zhao Y.; Huang J.; et al. Ultrafast In Situ Synthesis of Large-Area Conductive Metal-Organic Frameworks on Substrates for Flexible Chemiresistive Sensing. ACS Appl. Mater. Interfaces 2020, 12, 57235–57244. 10.1021/acsami.0c18422. [DOI] [PubMed] [Google Scholar]
- Zheng R.; Fu Z. H.; Deng W. H.; Wen Y.; Wu A. Q.; Ye X. L.; Xu G. The Growth Mechanism of a Conductive MOF Thin Film in Spray-based Layer-by-layer Liquid Phase Epitaxy. Angew. Chem., Int. Ed. 2022, 61, e202212797 10.1002/anie.202212797. [DOI] [PubMed] [Google Scholar]
- Lin Y.; Li W. H.; Wen Y.; Wang G. E.; Ye X. L.; Xu G. Layer-by-Layer growth of preferred-oriented MOF thin film on nanowire array for high-performance chemiresistive sensing. Angew. Chem., Int. Ed. 2021, 60, 25758–25761. 10.1002/anie.202111519. [DOI] [PubMed] [Google Scholar]
- Lee T.; Kim J. O.; Park C.; Kim H.; Kim M.; Park H.; Kim I.; Ko J.; Pak K.; Choi S. Q.; et al. Large-Area Synthesis of Ultrathin, Flexible, and Transparent Conductive Metal-Organic Framework Thin Films via a Microfluidic-Based Solution Shearing Process. Adv. Mater. 2022, 34, 2107696. 10.1002/adma.202107696. [DOI] [PubMed] [Google Scholar]
- Park S.; Oh D.; Ahn J.; Kim J. K.; Kim D. H.; Kim S.; Park C.; Jung W.; Kim I. D. Promoting Ex-Solution from Metal-Organic-Framework-Mediated Oxide Scaffolds for Highly Active and Robust Catalysts. Adv. Mater. 2022, 34, 2201109. 10.1002/adma.202201109. [DOI] [PubMed] [Google Scholar]
- Zhang W.; Lu G.; Cui C.; Liu Y.; Li S.; Yan W.; Xing C.; Chi Y. R.; Yang Y.; Huo F. A family of metal-organic frameworks exhibiting size-selective catalysis with encapsulated noble-metal nanoparticles. Adv. Mater. 2014, 26, 4056–4060. 10.1002/adma.201400620. [DOI] [PubMed] [Google Scholar]
- Koo W.-T.; Choi S.-J.; Kim S.-J.; Jang J.-S.; Tuller H. L.; Kim I.-D. Heterogeneous sensitization of metal-organic framework driven metal@metal oxide complex catalysts on an oxide nanofiber scaffold toward superior gas sensors. J. Am. Chem. Soc. 2016, 138, 13431–13437. 10.1021/jacs.6b09167. [DOI] [PubMed] [Google Scholar]
- Jang J.-S.; Koo W.-T.; Choi S.-J.; Kim I.-D. Metal organic framework-templated chemiresistor: sensing type transition from P-to-N using hollow metal oxide polyhedron via galvanic replacement. J. Am. Chem. Soc. 2017, 139, 11868–11876. 10.1021/jacs.7b05246. [DOI] [PubMed] [Google Scholar]
- Koo W. T.; Kim S. J.; Jang J. S.; Kim D. H.; Kim I. D. Catalytic metal nanoparticles embedded in conductive metal-organic frameworks for chemiresistors: highly active and conductive porous materials. Adv. Sci. 2019, 6, 1900250. 10.1002/advs.201900250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun Y.; Wang B.; Hou Y.; Suematsu K.; Zhao Z.; Zhang W.; Shimanoe K.; Hu J. Noble metal nanoparticles functionalized conductive Co3(hexaiminotriphenylene)2 chemiresistor for hydrogen sulfide detection at Room-Temperature. Chem. Eng. J. 2023, 465, 142818. 10.1016/j.cej.2023.142818. [DOI] [Google Scholar]
- Park C.; Koo W. T.; Chong S.; Shin H.; Kim Y. H.; Cho H. J.; Jang J. S.; Kim D. H.; Lee J.; Park S.; et al. Confinement of Ultrasmall Bimetallic Nanoparticles in Conductive Metal-Organic Frameworks via Site-Specific Nucleation. Adv. Mater. 2021, 33, 2101216. 10.1002/adma.202101216. [DOI] [PubMed] [Google Scholar]
- Choi H. J.; Chung J.-H.; Yoon J.-W.; Lee J.-H. Highly selective, sensitive, and rapidly responding acetone sensor using ferroelectric ε-WO3 spheres doped with Nb for monitoring ketogenic diet efficiency. Sens. Actuators, B 2021, 338, 129823. 10.1016/j.snb.2021.129823. [DOI] [Google Scholar]
- Wang M.; Zhang Z.; Zhong H.; Huang X.; Li W.; Hambsch M.; Zhang P.; Wang Z.; St. Petkov P.; Heine T.; et al. Surface-Modified Phthalocyanine-Based Two-Dimensional Conjugated Metal-Organic Framework Films for Polarity-Selective Chemiresistive Sensing. Angew. Chem. 2021, 133, 18814–18820. 10.1002/ange.202104461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim D. H.; Chong S.; Park C.; Ahn J.; Jang J. S.; Kim J.; Kim I. D. Oxide/ZIF-8 Hybrid Nanofiber Yarns: Heightened Surface Activity for Exceptional Chemiresistive Sensing. Adv. Mater. 2022, 34, 2105869. 10.1002/adma.202105869. [DOI] [PubMed] [Google Scholar]
- Yao M. S.; Xiu J. W.; Huang Q. Q.; Li W. H.; Wu W. W.; Wu A. Q.; Cao L. A.; Deng W. H.; Wang G. E.; Xu G. Van der Waals Heterostructured MOF-on-MOF Thin Films: Cascading Functionality to Realize Advanced Chemiresistive Sensing. Angew. Chem. 2019, 131, 15057–15061. 10.1002/ange.201907772. [DOI] [PubMed] [Google Scholar]
- Cho S. R.; Kim D. H.; Jeon M.; Rani P.; Gyeon M.; Kim Y.; Jo M. k.; Song S.; Park J. Y.; Kim J.; et al. Overlaying Monolayer Metal-Organic Framework on PtSe2-Based Gas Sensor for Tuning Selectivity. Adv. Funct. Mater. 2022, 32, 2207265. 10.1002/adfm.202207265. [DOI] [Google Scholar]
- Kwon O.; Kim J. Y.; Park S.; Lee J. H.; Ha J.; Park H.; Moon H. R.; Kim J. Computer-aided discovery of connected metal-organic frameworks. Nat. Commun. 2019, 10, 3620. 10.1038/s41467-019-11629-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cho S.; Park C.; Jeon M.; Lee J. H.; Kwon O.; Seong S.; Kim J.; Kim I.-D.; Moon H. R. Interface-sensitized chemiresistor: Integrated conductive and porous metal-organic frameworks. Chem. Eng. J. 2022, 449, 137780. 10.1016/j.cej.2022.137780. [DOI] [Google Scholar]