Kevin Dedecker*a,
Benjamin Paretb,
Lionel Presmanes*b,
Benjamin Duployerb,
Antoine Barnabé
b,
Philippe Menini
c,
David Farrusseng
d,
Mikhael Bechelany
ae,
Martin Drobek
*a and
Anne Julbe
a
aInstitut Européen des Membranes (IEM-UMR 5635) CNRS, ENSCM, Univ Montpellier, Place Eugène Bataillon, 34095 Montpellier Cedex 5, France. E-mail: kevin.dedecker@umontpellier.fr; martin.drobek@umontpellier.fr
bCentre Interuniversitaire de Recherche et d'Ingénierie des Matériaux (CIRIMAT-UMR 5085), Université de Toulouse, Université Toulouse 3 Paul Sabatier, Toulouse INP, CNRS, 118 Route de Narbonne, Cedex 9, 31062 Toulouse, France. E-mail: lionel.presmanes@univ-tlse3.fr
cLaboratoire d'Analyse et d'Architecture des Systèmes (LAAS-UPR 8001), Université de Toulouse, CNRS, UPS, 7 Avenue du Colonel Roche, Cedex 4, 31031 Toulouse, France
dInstitut de recherches sur la catalyse et l'environnement de Lyon (IRCELYON-UMR 5256), Université de Lyon, CNRS, 2 Avenue Albert Einstein, 69626 Villeurbanne Cedex, France
eFunctional Materials Group, Gulf University for Science and Technology, Mubarak Al-Abdullah, 32093, Kuwait
First published on 10th June 2025
Functionalizing ZnO-based chemiresistive sensor surfaces with a ZIF-8 layer, which acts as a permselective membrane, is a well-established strategy to enhance sensor selectivity. This study examines the key factors influencing the conversion process, including the physicochemical properties of solvents or solvent mixtures and the thermal pretreatment, of Ga-doped ZnO (ZnO:
Ga). We have evidenced that the polarity, viscosity, and interfacial tension of the solvent significantly affect the dissolution of ZnO
:
Ga and the crystallization of ZIF-8. Methanol–water mixtures were found to effectively control the conversion process, with a 3
:
1 MeOH/H2O ratio being optimal for producing continuous ZIF-8 membranes. Additionally, it has been demonstrated that annealing can greatly enhance the reactivity of the oxide. However, while it enhances the dissolution of ZnO
:
Ga, excessively high temperatures can lead to over-dissolution, which hinders ZIF-8 formation. These insights are crucial for optimizing ZIF-8 layers on ZnO
:
Ga, paving the way for the development of highly selective chemiresistive sensors.
Zinc oxide (ZnO) is a key material for gas sensors due to its non-toxicity, low cost, and high electron mobility. As an n-type semiconductor with a wide bandgap (3.37 eV) and excellent carrier mobility,4–6 ZnO shows great potential for sensing applications. However, its high resistivity often limits the sensing performance.7 To overcome this limitation, doping with gallium (Ga) effectively reduces resistivity, enhances oxidation resistance, and improves thermal stability.8–10 Ga-doped ZnO (ZnO:
Ga) is therefore considered as one of the most promising materials for fabricating sensitive sensors,11 as demonstrated by Presmanes and coworkers in their sub-ppm NO2 sensing studies.12 Despite these advances, ZnO
:
Ga sensors, like their other MOx counterparts, continue to face challenges with selectivity, especially in the presence of interfering molecules. This necessitates further optimization for practical sensing applications.
To address the issue of limited selectivity in chemiresistive sensors, a promising solution involves applying a selective membrane as an overlayer on the MOx-based sensing material. This strategy has proven effective in various systems, including nanowire-based sensors like SnO2 coated with Al2O3 (SnO2@Al2O3)13 and ZnO combined with boron nitride and palladium (ZnO@boron nitride/Pd).14 These selective membranes not only enhance the selectivity of the sensors but also improve their stability and reliability in complex environments. It is worth noting that metal–organic frameworks (MOFs) have also been studied as membrane overlayers to enhance the selectivity of chemiresistive sensors15,16 forming interfaces conceptually similar to heterojunctions observed in certain metal-oxide clusters.17
MOFs are hybrid crystalline materials composed of metal nodes connected by coordinating ligands.18 Their versatility has generated significant interest in various fields, including adsorption, separation, catalysis, and sensing. This interest stems from the ability to fine-tune their properties by selecting specific metals,19–21 modifying ligand structures,22,23 and introducing functional groups24–26 to meet the demands of particular applications.27,28 ZIF-8 (zeolitic imidazolate framework),29 a type of MOF, has attracted considerable interest due to its exceptional properties. Constructed from zinc cations linked by 2-methylimidazolate ligands, ZIF-8 stands out for its thermal and chemical stability,30 ease of synthesis, and microporous structure. Its pore aperture of 3.4 Å and cavity size of 10 Å make it highly effective for molecular sieving and storage applications.31–33
To enhance the selectivity of chemiresistive sensors, ZIF-8 is often studied as a molecular sieving membrane that covers a sensitive ZnO layer. The ZnO layer serves a dual function: acting as a reservoir for Zn2+ ions and enabling in situ formation of the MOF. Various methods to convert ZnO into ZIF-8 have been reported, including mechanochemistry,34 sonochemistry,35 and solution-based chemistry,36 each offering unique advantages in efficiency and scalability. The direct conversion of ZnO layers deposited by atomic layer deposition (ALD),37 chemical vapor deposition (CVD)36 or sputtering37 onto various supports (e.g. alumina,38 silicon,39 PAN,..40) has been widely studied over the past decade. This research has also been extended to form advanced composites, such as 3D ZIF-8/CNT structures, with enhanced adsorption properties.41 This method is also applied to coat ZnO-based chemiresistive sensors with various morphologies,42 such as nanorods, nanofibers, tetrapods or supported thin films. Among various selective membrane approaches, ZIF-8 offers distinct advantages for improving gas-sensing applications thanks to its molecular sieving properties. Previous studies by our research group have demonstrated that ZIF-8 layers significantly enhance the selectivity of ZnO@ZIF-8 nanowire-based sensors.43 The presence of ZIF-8 (pore aperture of 3.4 Å), can effectively separate H2 from larger molecules like benzene and toluene, resulting in virtually zero gas sensing response to these larger molecules due to the size exclusion effect. Furthermore, Weber et al.44 demonstrated further improvement of selectivity for H2 over toluene, acetone, ethanol, and benzene using Pd-loaded ZIF-8 membranes on ZnO nanowire sensors showing high stability up to 300 °C during repetitive temperature cycling. Their results show that the addition of the ZIF-8 membrane layer achieved >1000% improvement in selectivity by completely eliminating sensor response to all interfering gases (toluene, acetone, ethanol, and benzene), while maintaining strong sensitivity to H2, effectively transforming a multi-gas responsive sensor into one exclusively selective for hydrogen.
Overall, the molecular sieving strategy significantly improves selectivity of metal oxide sensors compared to conventional approaches, such as catalyst loading in metal oxides, which show less efficient molecular discrimination.45,46 It also surpasses many other selective materials used directly as sensors like metal oxides or dense polymers, which often rely on other non-sieving mechanisms such as catalytic effects or differences in solution/diffusion rates.47
ZnO surfaces or films can be converted to ZIF-8 through two main methods: reaction with vaporized 2-methylimidazole48,49 or direct solution-based conversion.50 The solution-based approach is often preferred because it provides better control over ZIF-8 formation and allows ligand recycling (up to 50 times).51 It involves five stages: ZnO dissolution, Zn2+ ion release and diffusion, ZIF-8 nucleation, crystal growth, and Ostwald ripening. Imbalances in these steps can lead to defects or incomplete formation of the ZIF-8 layer. Factors such as temperature,52 ZnO deposition method,37 crystallinity, crystal orientation,53 solvent type,52 and reagent ratio54 critically influence the quality of ZIF-8 membranes.
While existing research offers valuable insights into optimizing the ZnO-to-ZIF-8 conversion process, several critical gaps remain. Specifically, the influence of solvent physicochemical properties - such as polarity, viscosity, and surface tension - has not been thoroughly investigated, leaving an incomplete understanding in this area. Additionally, studies on ZnO doping with other trivalent cations, such as aluminum have shown that solubility limits significantly affect material properties.55 These findings suggest that similar effects may occur in Ga-doped systems.
Furthermore, the importance of achieving controlled and partial consumption of the ZnO layer, which is essential for preserving its electrical properties and sensitivity during ZIF-8 membrane formation, has often been overlooked. In addition, the role of annealing treatments of ZnO:
Ga, commonly applied to enhance sensor performance, has not been adequately studied as a factor impacting the conversion process. Lastly, although doping is widely recognized for significantly enhancing the sensitivity and functionality of sensors, most studies focus solely on pristine ZnO, providing limited information on the behavior of doped ZnO systems during the conversion to ZIF-8. Hence, addressing these gaps is crucial for advancing the development of high-performance, selective sensors.
To tackle this issue, the present study aims to identify the critical factors affecting the controlled conversion of a high-performance ZnO:
Ga layer into a uniform ZIF-8 membrane while preserving the ZnO
:
Ga layer to maintain its low electrical conductivity. Specifically, we have examined how solvent properties - such as polarity, surface tension, and viscosity-influence the ZnO-to-ZIF-8 conversion process. Additionally, we have explored the role of ZnO
:
Ga annealing in determining the uniformity and of the resulting synthetized ZIF-8 membrane. This work thus provides valuable insights into optimizing the surface conversion of ZnO
:
Ga to ZIF-8, which will be applied in our future works to improve the selectivity of ZnO
:
Ga-based gas sensing layers.
A Ga-Focused Ion Beam (FIB)-FESEM dual beam FEI Helios NanoLab600i apparatus, equipped with an Oxford Instruments Aztec Advanced EDS system featuring an X-Max 80 mm2 detector, was used for surface characterization and the preparation and analysis of film cross-sections. This system enabled high-resolution imaging and semi-quantitative chemical analysis of the thin films in both surface and cross-section views. To prevent damage to the samples during Ga-FIB milling, a platinum layer was deposited on the film surface. For surface imaging, a carbon layer was deposited to enhance conductivity and improve image clarity. Due to Ga milling, the analysis of Ga traces is not feasible on the cross-sections.
Solvent | Dielectric constant (F m−1) | Dipole moment (D) | Dynamic viscosity at 20 °C (mPa s) | Interfacial tension (mN m−1) |
---|---|---|---|---|
MeOH | 33.0 | 1.70 | 0.588 | 22.6 |
EtOH | 25.3 | 1.69 | 1.095 | 22.3 |
DMF | 36.7 | 3.82 | 0.92 | 36.4 |
H2O | 78.4 | 1.86 | 0.354 | 72.6 |
Application of ethanol leads to the lowest crystallinity among the solvents tested, mainly due to its relatively low polarity when compared to DMF. Indeed, polarity plays a critical role in dissolving ionic compounds and facilitating the nucleation process necessary for ZIF-8 formation. Due to its lower polarity, ethanol struggles to effectively deprotonate the 2-mim ligand, a crucial step in the generation of ZIF-8 nuclei. This limitation results in fewer nuclei forming, thereby reducing the overall crystallinity. Additionally, its higher viscosity hampers the diffusion of reactants, leading to their less frequent mutual collisions, thereby impeding crystal growth. The resulting ZIF-8 crystals are smaller in size and poorly defined. Moreover, the higher interfacial tension of ethanol compared to DMF imposes a significant surface barrier to crystallization, which further suppresses nucleation and limits crystal growth.
In contrast, DMF demonstrates better performance due to its higher polarity and lower viscosity. Its superior polarity enhances the deprotonation of the 2-mim ligand, promoting a higher level of nucleation. The lower viscosity of DMF allows for more rapid diffusion of reactants, increasing their frequency of collisions and thus accelerating crystal growth. However, despite these advantages, the interfacial tension and nucleation dynamics associated with DMF are less effective to achieve high crystallinity compared to methanol.
In fact, methanol appears as the most effective solvent for promoting high crystallinity in ZIF-8. The combination of its high polarity and low viscosity creates an optimal environment for efficient nucleation and robust crystal growth. Its high polarity ensures advanced deprotonation of the 2-mim ligand, resulting in the formation of abundant nuclei. Simultaneously, its low viscosity facilitates rapid diffusion of reactants, enabling frequent interactions that support sustained crystal growth. These synergistic properties of methanol produce larger, well-defined ZIF-8 crystals with the highest degree of crystallinity obtained among the solvents tested.
![]() | ||
Fig. 2 SEM observations (surfaces and cross-sections) of ZnO![]() ![]() |
For the ZnO:
Ga/ZIF-8 bilayer, the ZnO
:
Ga layer thickness remains constant at approximately 55 nm, while the ZIF-8 layer grows to around 100 nm, indicating the consumption of about 36 nm of ZnO
:
Ga during the transformation. The intercrystalline spaces, initially filled by the ZnO
:
Ga layer, appear to serve as reservoirs of metal ions, supplying Zn2+ for ZIF-8 formation in adjacent regions. This interpretation is supported by EDS spectra (Fig. S3 and S4†), which show that areas with intercrystalline defects have low zinc content compared to silicon and oxygen from the wafer. In contrast, neighboring regions exhibit higher concentrations of zinc, carbon, and nitrogen, consistent with contributions from the ZIF ligand.
Cross-sectional SEM micrographs show a well-defined ZnO/ZIF-8 stack (Fig. 2). The ZnO:
Ga layer thickness decreases to about 39 nm, while the ZIF-8 layer grows to approximately 48 nm, consistent with the partial consumption of ZnO
:
Ga during conversion. In some areas, only ZIF-8 is observed, as confirmed by EDS analysis (Fig. S5 and S6†), indicating that pinholes may result from the complete local conversion of ZnO
:
Ga into ZIF-8.
However, these observations raise additional questions. Theoretical calculations reported elsewhere57 predict that the ZIF-8 layer should be thicker than the original ZnO:
Ga layer after complete conversion (an expansion factor up to 17 was calculated for ZIF-8, corresponding to the ratio of metal ion densities in the crystalline structure of the oxide and the MOF). Yet, in the pinhole regions, the layer thickness remains around 50 nm, comparable to the original ZnO
:
Ga layer. This anomaly suggests that in these regions a complete conversion releases excess Zn2+ ions into the surrounding medium. These ions may accumulate at the interface, potentially driving additional ZIF-8 formation near the pinholes.
The ZIF-8 crystals formed during methanol conversion display specific morphological features, notably their faceted geometry. SEM micrographs reveal that most crystals adopt a rhombic dodecahedral shape with 12 exposed {110} facets (Fig. 2). According to the well-established crystallography of ZIF-8,53,58 these facets correspond to {110} planes. This highly crystalline morphology is attributed to the slow growth rate of {110} facets,59 which, according to Wulff's rule, determines the final crystal structure. A small fraction of the crystals exhibits a truncated cubic shape, typically associated with slower growth of {100} facets relative to {110}. This shape is often induced by capping agents like surfactants, and its presence, even in small quantities, suggests that during the conversion process, the growth of {100} facets might be selectively decelerated. While direct visualization of these crystallographic planes would require advanced techniques such as HR-TEM, the observed external morphologies align with the crystallographic features of ZIF-8 reported in the literature.53,58,60
The formation of such large, faceted ZIF-8 crystals can be explained by the interfacial conversion mechanism. In the early stages, the release of Zn2+ ions from the ZnO:
Ga layer enriches the interfacial medium, temporarily creating a low ligand-to-metal molar ratio and a supersaturation regime. Similar conditions have been reported in the literature, where EDS spectroscopy showed elevated zinc concentrations on substrate surfaces outside ZIF-8 domains, peaking at 5.3% after three hours and declining gradually as Zn2+ is fully consumed after ten hours.61
This supersaturation regime enables two potential growth pathways. Both begin with the nucleation of crystalline primary particles (CPP), followed by growth through successive addition of either monomers (metal ions and ligands)62 or secondary building units (SBUs).63 These pathways contribute to the formation of the well-faceted crystals observed in SEM micrographs (Fig. 2). As the Zn2+ concentration diminishes over time, crystal intergrowth becomes more prominent, transitioning from monomer- or SBU-driven growth to processes like polynuclear growth, internal reorganization, and aggregative growth. These mechanisms further enhance the observed crystal morphology.64
The results indicate that among the solvents studied, methanol is the most effective for promoting ZIF-8 crystallization, while water primarily facilitates ZnO:
Ga dissolution. This highlights the dual importance of solvent selection in determining both the dissolution and crystallization dynamics of the conversion process. Other polar aprotic solvents with low viscosity, such as acetonitrile (dielectric constant: 37.5, dynamic viscosity: 0.369 mPa s, interfacial tension: 29.3 mN m−1) and acetone (dielectric constant: 20.7, dynamic viscosity: 0.306 mPa s, interfacial tension: 23.7 mN m−1), may exhibit intermediate behavior in this conversion process. The high polarity and very low viscosity of acetonitrile could potentially accelerate ZIF-8 nucleation while maintaining moderate dissolution rates, possibly resulting in smaller, and more numerous crystals with higher overall crystallinity compared to DMF. Conversely, acetone, with its lower polarity but similarly low viscosity, might favor controlled ZnO dissolution while promoting moderate crystal growth rates, potentially leading to a more uniform particle size distribution. Based on the trends observed in this study, these solvents could offer additional pathways for fine-tuning ZnO conversion kinetics. Future investigations should consider these solvents to expand the toolkit for ZIF-8 membrane engineering with tailored morphological properties.
As discussed previously, DMF exhibits notable potential in ZIF-8 synthesis but is unsuitable due to health and environmental risks. Ethanol forms high-quality layers but has limited impact on oxide dissolution and MOF crystallization process. Water effectively dissolves the ZnO:
Ga layer, providing Zn2+ ions for ZIF-8 formation, while methanol aids in crystallization, forming large, well-faceted ZIF-8 crystals. Both water and methanol influence ZnO
:
Ga dissolution and ZIF-8 formation, though neither component on its own can achieve the desired balance for optimal layer performance.
A water–methanol mixture was thus proposed to combine the strengths of both solvents while reducing their limitations. By adjusting methanol-to-water ratio, the process was optimized to enhance Zn2+ ion release or promote ZIF-8 crystallization. The conversion was optimized by maintaining a temperature of 100 °C, while reducing the reaction time from 12 hours to 1.5 hours. Three methanol-to-water ratios (1:
3, 1
:
1, and 3
:
1) were tested, along with pure methanol for comparison. The structural and morphological characteristics of the corresponding samples synthetized in water–methanol solvent mixtures are reported and discussed in the following sections.
As the water content in the solvent mixture decreases in favor of methanol, the intensity of the ZIF-8 XRD peaks weakens, and the peaks become broader, while the relative intensity of the ZnO:
Ga peaks increases. Such result indicates a slower conversion rate and smaller ZIF-8 crystal sizes obtained in methanol solution with reduced water content. The infrared spectra (Fig. 3b) further confirm these observations, demonstrating that the MeOH/H2O solvent ratio can effectively regulate the conversion process. This allows for a precise control over the conversion rates, ranging from minimal to advanced, depending on the target reaction outcome.
![]() | ||
Fig. 4 SEM observations (surfaces) of ZnO![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
![]() | ||
Fig. 5 SEM observations (cross-sections) of ZnO![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
The ZIF-8 crystals feature an imperfect rhombic dodecahedral shape, coated with smaller spherical nanoparticles. This diminished crystal quality, compared to those formed in pure methanol, is attributed to faster dissolution kinetics and change in the physicochemical conditions at the crystal interface. Typically, a supersaturated environment with high Zn2+ concentration at the interface promotes the growth of well-defined, faceted crystals. However, in this scenario, the reduced viscosity allows Zn2+ ions to diffuse away from the interface more rapidly, disrupting the sustained high cation concentrations necessary for prolonged and precise crystal growth.
Two potential mechanisms could explain these observations:
1) Sequential growth and diffusion mechanism: water facilitates the rapid release of Zn2+ ions, creating a localized environment at the interface that favors the nucleation of crystalline primary particles (CPPs). These particles subsequently grow through the addition of monomers or secondary building units (SBUs). However, due to the low viscosity of water, Zn2+ ions quickly diffuse away from the interface, depleting the interfacial concentration. Such depletion prompts the formation of crystals via CPP nucleation in a metastable phase, followed by polynuclear growth, resulting in smaller particles with irregular crystal shapes.
2) Growth, erosion, and secondary nucleation mechanism: initially, the rapid release of Zn2+ ions supports the growth of large, well-defined rhombic dodecahedral crystals. However, the strong erosive properties of water accelerate Zn2+ depletion, partially dissolving the ZIF-8 crystals through hydrolysis of zinc-nitrogen bonds. This process releases Zn2+ ions near the larger crystals, triggering secondary nucleation and forming smaller spherical nanoparticles. This secondary nucleation compromises the morphological integrity of the larger crystals.
The latter hypothesis is supported by reports of ZIF-8 undergoing cycles of growth and dissolution under similar conditions.61 However, in the present study, these processes occur within 1.5 hours, hence significantly faster than the 10 hours reported elsewhere.61 Such accelerated kinetics is likely due to the high water content, which enhances the hydrolysis of zinc–ligand bonds and speeds up dissolution.
Cross-sectional micrographs (Fig. 5) reveal the advanced dissolution of the ZnO:
Ga layer caused by the high water content in the solvent mixture. Two distinct surface regions are evidenced: one where the ZnO
:
Ga layer has been completely replaced by a ZIF-8 layer, and another where the ZnO
:
Ga layer remains intact at its original thickness. Additionally, the resulting ZIF-8 layer exhibits significant surface irregularities. These observations suggest that excessive ZnO
:
Ga dissolution compromises sample quality, highlighting the need to reduce the water proportion in the solvent mixture.
Cross-sectional SEM micrographs (Fig. 5) reveal uniform conversion across the sample surface. The remaining ZnO:
Ga layer retains a consistent thickness, indicating its controlled dissolution. However, isolated ZIF-8 regions are still observed on the surface, separated from the bilayer areas. These regions could affect system performance by increasing overall resistivity.
Cross-sectional SEM micrographs (Fig. 5) confirm these improvements. The ZnO:
Ga layer retains a constant thickness along the sample, similar to what was observed with the 1
:
1 solvent mixture. The lower water content significantly minimizes ZnO
:
Ga consumption during the conversion process. Importantly, the bilayer remains defect-free and continuous. Such optimized conditions allow for precise control of the ZnO
:
Ga dissolution rate, promoting effective ZIF-8 crystallization. This results in a high-quality bilayer system and a compact ZIF-8 membrane layer with only a slight reduction in ZnO
:
Ga thickness.
Interestingly, while a high water content in solvent mixtures adversely affects membrane formation, the presence of a small amount of water is crucial for achieving optimal membrane quality. This observation is supported by SEM micrographs of samples converted in pure methanol under the same reaction conditions (Fig. S17 and S18†), which reveal suboptimal results in the absence of water.
![]() | ||
Fig. 7 SEM observations (surfaces) of annealed ZnO![]() ![]() ![]() ![]() |
When ZnO is annealed at 600 °C, several structural changes occur at the atomic level. The thermal energy causes oxygen atoms to migrate to the surface, where they can desorb as O2, creating oxygen vacancies within the metal oxide lattice.65 These vacancies generate localized regions of positive charge (defects) due to the missing O2− ions. During the subsequent conversion process, these defects serve as preferential dissolution sites, allowing water molecules to more readily weaken the Zn–O bonds and facilitate the release of Zn2+ ions into the solution. Additionally, the increased surface energy at defect sites promotes faster chemical reactions.
The higher concentration of Zn2+ ions at the interface then promotes the formation of larger ZIF-8 crystals. This finding is not surprising as for smaller ZIF-8 particles higher molar ratios of 2-methylimidazole to zinc ions is typically required. It should be mentioned that excessive 2-methylimidazole can coat ZIF-8 surfaces and slow the growth of crystal nuclei, but this effect is less significant in this case. The defects observed in the annealed sample are therefore caused by the accelerated ZnO:
Ga dissolution, which hinders the uniform crystallization of ZIF-8 at the interface.
Cross-sectional micrographs show that the regions with ZIF-8 islands consist of a bilayer of ZnO:
Ga and ZIF-8, while the grain boundary defects reveal the absence of matter, exposing the silicon wafer below (Fig. 8). This indicates that the Zn2+ ions necessary for ZIF-8 crystallization are initially supplied by the ZnO
:
Ga surrounding the islands.
Cross-sectional micrographs provide further evidence for this hypothesis (Fig. 8). The images reveal a single layer of residual ZnO:
Ga surrounded by voids, exposing the silicon wafer to the environment. This confirms that under such reaction conditions, ZnO
:
Ga dissolution dominates over ZIF-8 crystallization, resulting in an incomplete conversion process.
These findings highlight the critical role of thermal treatment in influencing the dissolution kinetics of ZnO:
Ga. They emphasize the importance of adjusting conversion conditions to slow down dissolution while enhancing ZIF-8 crystallization, ensuring a balanced dissolution/crystallization process and its effective control. The quality parameters of the ZIF-8 membranes optimized in this study should directly impact their molecular sieving performance in gas sensing applications, as demonstrated in our previous works.43,44 Furthermore, the literature extensively documents how ZIF-8 membrane characteristics influence sensing selectivity and response.16 The controlled membrane thickness (∼50–100 nm) achieved in our study represents an optimal balance between selectivity and response kinetics. Literature indicates that thicker ZIF-8 membranes (>200 nm) typically result in lower selectivity, preventing even H2 molecules from passing through.67 Conversely, an insufficient conversion rate may lead to cracks in the membrane or incomplete coverage, compromising selectivity.
The high crystallinity of the ZIF-8 layer optimized in this work, evidenced by sharp XRD peaks (Fig. 3a), ensures consistent 3.4 Å pore apertures across the membrane. This aperture size is ideal for selectively allowing H2 (kinetic diameter: 2.9 Å) to pass while blocking larger molecules such as acetone (4.6 Å), ethanol (4.5 Å), and benzene (5.9 Å).68 Finally, the proposed controlled conversion process preserves a significant portion of the underlying ZnO layer, which is critical for maintaining electrical conductivity for the sensing mechanism. The membrane quality parameters achieved in this study, particularly defect density, crystallinity, and thickness control, align well with those identified in the literature as determinants of superior molecular sieving performance.
This investigation thus demonstrates that solvent properties play a crucial role in ZIF-8 membrane engineering, while the characteristics of the metal oxide layer significantly influence the dissolution mechanism and, consequently, membrane quality. Our optimized synthesis conditions (3:
1 MeOH/H2O ratio, 100 °C, 1.5 hours) provide precise control over ZnO dissolution and promote effective ZIF-8 crystallization. This results in a high-quality, uniform bilayer system with consistent thickness across the substrate and nanoscale crystals forming a compact layer, potentially suitable as a selective membrane barrier in chemiresistive sensor applications. Further examination of the metal oxide surface chemistry, which depends notably on deposition conditions and thermal treatment, is essential for a complete mechanistic understanding. These findings lay the groundwork for developing precise conversion protocols to engineer selective chemiresistive sensors, which will be further refined by additional studies of metal oxide chemistry. Our experimental approach has provided mechanistic insights through systematic investigation of solvent effects and thermal treatments. Recent advanced studies in the literature further support our mechanistic interpretations. Tao et al. investigated the interplay between organic linkers and ZnO substrates during ZIF-8 crystallization using in situ atomic force microscopy combined with ab initio molecular dynamics and density functional theory calculations. Their work revealed that organic ligands selectively bind to ZnO surface steps, controlling dissolution kinetics and thereby regulating MOF overgrowth.53 This selective binding mechanism aligns with our observations of how different solvents influence ZnO dissolution rates. Additionally, Sinnwell et al. utilized in situ X-ray diffraction to reveal the time-resolved formation of ZIF-8 and intermediary ZnO@ZIF-8 composites, demonstrating temperature-dependent mechanisms throughout the reaction process.69,70 These findings parallel our observations regarding the critical role of temperature in balancing dissolution and crystallization kinetics.
While our study focuses on optimizing practical conversion conditions through systematic experimental investigation, these advanced mechanistic studies provide complementary theoretical foundations that support the proposed conversion mechanisms and the observed effects of solvent properties and ZnO thermal treatment on ZIF-8 growth.
The results emphasize the importance of solvent physicochemical properties, such as polarity, viscosity, and interfacial tension, in controlling the ZnO:
Ga conversion rate and ZIF-8 crystallinity. Methanol, while effectively dissolving, excels in promoting ZIF-8 crystallization. Attractively, by utilizing methanol–water mixtures, the dissolution and crystallization processes can be precisely regulated and optimal balanced. For example, a 3
:
1 CH3OH/H2O ratio minimizes defects formation and ensures a thin, uniform bilayer structure, making it ideal for the preparation of continuous ZIF-8 layers on ZnO
:
Ga thin films.
The effect of annealing on ZnO:
Ga conversion underscores the critical role of controlling dissolution kinetics. Annealing has been shown to enhance the reactivity of ZnO
:
Ga by increasing oxygen vacancies, which accelerates dissolution, and influences MOF crystal growth. However, higher temperatures can lead to excessive dissolution, hindering ZIF-8 crystallization and compromising the structural integrity of the ZIF-8 layer.
This study highlights the importance of optimizing solvent compositions, conversion parameters, and metal oxide thermal treatments to produce defect-free ZIF-8 membrane layers. In a broader context, these findings offer valuable insights for improving fabrication techniques in chemiresistive sensors, targeting superior sensitivity, selectivity, and environmental sustainability for practical applications. These insights are crucial for optimizing ZIF-8 selective films on metal oxide thin layers, paving the way for developing highly selective gas sensors based on sensitive semiconductors like ZnO. Our future research will focus on implementing these optimized ZIF-8 membranes in functional gas sensors based on ZnO semiconductor thin films and evaluating their selectivity enhancement in real-world conditions for specific target gases. Building on our previous work with ZnO@ZIF-8 systems, sensors fabricated using optimized conversion protocols are expected to demonstrate enhanced selectivity while maintaining high sensitivity.
Footnote |
† Electronic supplementary information (ESI) available: Additional material characterizations and computational details. See DOI: https://doi.org/10.1039/d5lf00054h |
This journal is © The Royal Society of Chemistry 2025 |