Kugalur Shanmugam
Ranjith
a,
Sonam
Sonwal
b,
Ali
Mohammadi
a,
Ganji
Seeta Rama Raju
a,
Mi-Hwa
Oh
*c,
Yun Suk
Huh
*b and
Young-Kyu
Han
*a
aDepartment of Energy and Material Engineering, Dongguk University-Seoul, Seoul 04620, Republic of Korea. E-mail: ykenergy@dongguk.edu
bDepartment of Biological Sciences and Bioengineering, Nano Bio High-Tech Materials Research Center, Inha University, 100 Inha-ro, Michuhol-gu, Incheon 22212, Republic of Korea. E-mail: yunsuk.huh@inha.ac.kr
cNational Institute of Animal Science, Rural Development Administration, Wanju 55365, Republic of Korea. E-mail: moh@korea.kr
First published on 29th August 2024
The sensitivity of chemiresistive sensors is inherently compromised by ambient humidity and trace level detection of toxic gases has potential challenges at room temperature. Herein, we designed a metal–organic framework (MOF) on a layered MXene hybrid by tagging a ZIF-67-based MOF on layered Ti3C2Tx MXene and following this with a surface ligand exchange process to design a highly sensitive, humidity tolerant chemiresistive sensor for ultra-low ppb level (200 ppb) NH3 sensing. The gas selectivity of MXenes was influenced by surface tagging with the MOF, which creates high surface-active features that promote the interaction and selectivity of NH3 on the MXene surface. In addition, a passive shell ligand exchange reaction provides not only a hydrophobic surface and environmental stability to the hybridized surface but also contributes to the sensing performances. The hybridized H-MOF/MXene-based sensor exhibited a superior NH3 sensing response (ΔR/Rg = 6.9, 1 ppm) at room temperature with high selectivity and reliability and a theoretical detection limit of 12.8 ppb. Passive ligand exchange had a significant effect on the sensing response at room temperature but improved humidity resistance and long-term durability. The H-MOF/MXene response to NH3 was only reduced by 0.22% and 0.27% at relative humidities of 76% and 93%, which represented 1.2 and 8.3-fold improvements in the sensing response versus MOF6/MXene and bare MXene at an NH3 concentration of 10 ppm. Furthermore, the sensing mechanism involved electronic interactions and charge transfer through a Schottky junction between the MOF and MXenes and the synergistic promotion of the sensing response on the hybridized H-MOF/MXene platform. This work provides a means of designing a surface functionalized MOF on MXene heterostructures that enables the production of sensors tailored to diverse environmental conditions.
The design of metal–organic frameworks (MOFs) on MXene surfaces has shown promise in terms of maximizing sensitivity by increasing adsorption capacity, surface area, electronic interactions, and substrate porosity.19,20 MOFs are metal-coordinated cationic polymer nanomaterials consisting of metal nodes and organic linkers, and in addition to the advantages mentioned above, MOFs have low environmental toxicities.21 However, the inferior conductivity and poor response recoverability of MOFs limit their sensitivities in chemiresistive gas sensors.22 MOF-derived metal compositions or tagged MOF heterostructures are generally preferred as chemiresistive sensors due to their high adsorption capacities and porous nature, which maximize active sites.23,24 Zeolitic imidazolate framework (ZIF) based metal–organic frameworks have a high selectivity towards NH3 adsorption, through the selective pore size and acid–base interaction within the zeolite groups.25 Meanwhile, surface tagging of a MOF on a 2D layered surface prevents self-aggregation, maximizes surface-active sites, and favors high sensitivity.26 Tagging MXenes with MOFs was found to provide a more effective platform than other semiconducting compositions and to serve as a heterostructure interface with a Schottky junction that promoted gas molecule-induced resistance changes and combined these with the advantages of high conductivity and effective carrier transportation.27 The heterostructure assembly of MOFs on MXene templates is made possible by hydrogen bonding or electrostatic interactions that create a Schottky junction at the metal–semiconductor interface.28 To address the challenges of humidity-independent NH3 detection at room temperature (RT), the MOF on MXene hybrids needs to have good electrical conductivity and humidity-tolerant surface functionality. Reducing water affinity, enhancing moisture stability, and designing highly porous hydrophobic MOFs have led to a unique class of applications.29 The strategies used to fabricate hydrophobic MOFs involve ligand design and post-synthetic hydrophobization.30 However, despite its improved stability, specific surface areas and surface-active sites were diminished and parent properties were somewhat impaired due to capping functionalities that enhanced the hydrophobic nature of the MOF.31 Coupling MOFs and MXenes with the excellent heterostructure feature has promoted widespread application in electrochemical sensors.32,33 Building a porphyrin-based MOF on MXenes achieved superior sensitivity for NO2 sensing with a high response (Ra/Rg = 2.0, 10 ppm) at RT, high reliability, and a low particle limit of detection (pLOD, 200 ppb).19 Furthermore, a Co–TCPP(Fe)/Ti3C2Tx Schottky junction promoted electron transfer to provide outstanding selectivity and Fe–N4 functionality. Naveen et al.25 reported the preparation of Ti3C2Tx with a mixed matrix membrane filter as a chemiresistive sensor for NH3 sensing at RT with a pLOD of 1 ppb. So far, dual enhancement towards the selectivity and anti-humidity properties on chemiresistive sensors based on MOF–MXene hybrids at room temperature has not been reported.
In this work, a hydrophobic MOF (H-MOF) on MXenes (Ti3C2Tx) was specifically produced using a two-step fabrication strategy. First, thin-layered 2D MXene was prepared using a LiF/HCl-based acid etching process, and this was followed by delamination to produce extremely thin few layered sheets. Second, ZIF-67 tagged over MXenes in situ by hydrogen bonding was subjected to the shell ligand-exchange reaction (SLER) process. Tagging the MOF on the MXene surface promoted high selectivity towards NH3 through the effective adsorption due to selective pore size and acid–base interaction within the zeolite groups. Furthermore, the outer shell of ZIF-67 was modified to promote humidity tolerance without sacrificing its morphology or sensing performance. The designed H-MOF/MXene-based sensor exhibited superior sensing performance (∼8.3 times that of pristine MXenes; ranging from 200 ppb to 50 ppm) towards NH3 at RT coupled with a low LOD (12.8 ppb), high response, selectivity, and humidity-independent NH3 detection. The sensing mechanism responsible for this superior NH3 response was attributed to (i) presence of a MOF on the MXene surface alters the electrical conductance upon uptake of analyte gas, and due to the molecular sieving effect, the MOF shows a high interaction towards NH3 that promotes the selectively of the sensing device; (ii) charge transfer from the MOF to the MXene hybrid was made possible by the formation of a Schottky junction, which decreased the band gap and Fermi level of ZIF-67 in the vicinity of MXenes; and (iii) surface passivation by organic functionality, which promoted hydrophobicity and minimized interface contact with OH functionalities, facilitating humidity-independent NH3 detection. The dramatic synergistic effects of gas adsorption and charge transportation on the MOF on MXenes provide clues regarding the potential sensing applications of MXene-based sensors.
![]() | (1) |
![]() | ||
Scheme 1 Schematic illustration of the preparation of a hydrophobic MOF on the MXene hybrid heterostructure. |
The selective etching of Al in MAX to form MXenes was verified by XRD (Fig. S1†). Effective removal of the (104) plane with the mild shift in the (002) plane representing increased d-spacing, confirmed the formation of MXenes. SEM images of Ti3C2Tx MXene are provided in Fig. 1a and S2.† EDAX spectra confirmed the presence of Ti, C, F, O, and Cl and a minimal amount of Al compared to the MAX phase. TEM (Fig. S2†) confirmed the fabrication of few-layered MXene and confirmed a lattice spacing of 0.23 nm representing the crystal plane of MXenes (0110).36 The detailed morphologies of H-ZIF-67 and H-MOF6/MXene (6 wt% of MOF:
MXene) are shown in Fig. 1. The SEM images in Fig. 1b and c show the polyhedral morphology of H-ZIF-67 nanostructures and a particle size of ∼900 nm. High-magnification images showed that the rough surface features of H-ZIF-67 may have been caused by DMBIM surface capping during the SLER process; these features were not observed in pristine ZIF-67 (Fig. S3†). No other significant change in the morphology or size of ZIF-67 was attributed to DMBIM. Fig. 1d and e show the distribution of polyhedral nanostructures on the layered MXene surface, indicating surface tagging of H-ZIF-67 on MXenes. The intimate contact between the MOF and MXene might have been due to the surface nucleation of Co ions on MXenes. In the absence of a nucleation process, ZIF-67/MXene was a composite structure composed of a random distribution of MXenes and micro-structured ZIF-67 (Fig. S4†). Initial seeding of Co on MXenes maximized the interaction of ZIF-67 nucleates on the layered MXene surface and promoted the growth of ZIF-67 on the MXene surface. Furthermore, this interaction may have promoted Schottky junction formation. More information on this structural interface was obtained by TEM and image mapping analysis (Fig. 1f and g). Fig. 1f shows the intimate contact between MXenes and H-ZIF-67, which was impregnated into the layered surface. TEM mapping and EDX spectra demonstrated that Ti, C, O, and F were uniformly distributed on the layered surface. Meanwhile, Co ions were distributed in certain areas, indicating the successful preparation of a MOF on the MXene hybrid. Increasing the Co/2-MIM concentration during ZIF growth increased the density of the MOF on MXenes (Fig. S5†), but high MOF loadings on MXenes resulted in an increase in the resistance of the hybrid surface and reduced the sensing ability of the hybrid layer surface (Fig. S6†). As MOF6/MXene exhibited excellent sensing response with stability and selectivity, it was considered the optimum sample and used for subsequent sensing-related experiments.
The structural and surface characteristics of the MXene hybrid were investigated by XRD, FTIR, Raman spectroscopy, and XPS (Fig. 2). The structures of ZIF-67, MOF6/MXene, and H-MOF6/MXene were analyzed by XRD (Fig. 2a). The XRD pattern of polyhedral ZIF-67 showed diffraction peaks at 7.3°, 10.3°, 12.6°, 16.3°, 17.9°, 22.1°, and 26.5°, which correspond to the (011), (002), (112), (013), (222), (114) and (134) planes, respectively, on the standard JCPDS card on ZIF-67 (67-1073).37 The XRD pattern of the delaminated MXene showed the down shift of the (002) peak from 9.6° to 7.9°, resulting in the formation of MXenes with extended interlayer spacing (Fig. S1†). Heterostructure samples produced the combined diffraction peaks of ZIF-67 and MXenes. The interaction of the MOF on the MXene surface further downshifted the (002) peak of MXenes towards 6.3°, leading to the extended d-spacing of MXenes. It is worth noting that there is no significant peak shift between MOF6/MXene, and H-MOF6/MXene, suggesting that the crystal structure was not altered during the functionalization process. DMBIM-capping of heterostructures did not alter diffraction peaks, and neither did DMBIM functionalization of ZIF-67 (Fig. S7†). FTIR results for MXene, H-ZIF-67, MOF6/MXene, and H-MOF6/MXene heterostructures are shown in Fig. 2b. Vibration peaks at 427 and 1576 cm−1 were ascribed to Co–N and CN stretching modes, respectively, and indicated the presence of metal–organic linker-based interactions. For MXenes and MXene heterostructures, vibrations of –OH, C–F, and Ti–O were attributed to peaks at 3425, 1066, and 559 cm−1, respectively.38 Peaks observed between 900 and 1400 cm−1 corresponded to the bending and stretching modes of C–H bonds in MOF-related samples, respectively. A comparison of the FTIR spectra of pristine and DMBIM-functionalized samples (Fig. 2b and S8†) showed the addition of peaks at 1233 and 854 cm−1, which corresponded to C–N stretch and C–H out-of-plane deformation vibrations of the phenyl rings of DMBIM, respectively.35 The chemical structure and bonds between MXenes and ZIF-67 were investigated using Raman spectra (Fig. 2c). MOF6/MXene produced the characteristic of MXenes in the range of 100 to 800 cm−1. Peaks at 153 and 390 cm−1 were assigned to plasmonic resonance, and in-plane vibrations of Ti–C (Eg) and O (Eg). Peaks at 265 and 617 cm−1 represented A1g vibrations derived from the out-of-plane shearing modes of Ti–C (A1g) of H and C, respectively.39 ZIF-67-based samples had peaks in the lower Raman shift region attributed to Co ions in ZIF-67.40 Peaks at 421 and 673 cm−1 were assigned to the Co–N and 2-MIM-based functional states.41 Specifically, the band at 519 cm−1 was assigned to the α-Co(OH)2 stretch on the surface of ZIF-67. This band was not observed for H-ZIF-67, which was probably due to surface functionality removal by ligand exchange. On the other hand, a new peak at 470 cm−1 was assigned to the in-plane deformation vibration of the fused rings of DMBIM on SLER-treated samples.
XPS was used to determine the electronic states and compositions of the as-prepared heterostructures. Survey spectra of MXenes and the heterostructures (Fig. S9†) provided C, Ti, O, F, N, and Co compositions. The high-resolution C 1s spectra of the MXene, H-ZIF-67, and H-MOF6/MXene are shown in Fig. 2d. Pristine MXene produced four peaks at 281.6, 284.6, 286.4, and 288.9 eV, which were attributed to C–Ti, C–C, C–N, C–O, and CO/C–F, respectively. After ZIF-67 tagging, new peaks were observed at 282.3 and 285.4 eV, which represented O–Ti–C and C–N, with a mild peak shift at 281.2 (Ti–C) for H-MOF6/MXene, confirming surface modification. As Fig. 2e shows, Ti 2p of pristine MXene produced peaks at 455.1, 456.1, 456.8, and 458.9 eV corresponding to Ti–C (Ti+), Ti–X (Ti2+), Ti–X (Ti3+), and Ti–O (Ti4+) at Ti 2p3/2, respectively. Ti–X was related to sub-stoichiometric titanium carbide and titanium oxycarbide functionalities. MOF tagging of MXenes resulted in Ti–O and 2-MIM-based interactions on the MXene and intense Ti–X and Ti–O-related peaks. The high-resolution O 1s spectra (Fig. 2f) of pristine MXene exhibited four peaks at 529.6, 530.4, 531.6, and 532.8 eV, which were attributed to Ti–O, C–Ti–Ox, C–Ti–OHx, and C–O, respectively. Tagging of MXenes with the MOF also resulted in an intense peak at 531.8 eV, representing surface-adsorbed oxygen. A slight shift in the O 1s spectra of H-MOF6/MXene heterostructures resulted from a strong interaction between the MOF and the MXene surface. F 1s spectra of pristine MXene were attributed to Ti–F and C–F peaks at 684.3 and 685.9 eV, which was consistent with previous reports (Fig. 2i).18 The F 1s spectrum of MOF6/MXene was downshifted to 685.3 eV, which was attributed to intimate contact between the –COOH functionality of MOF and F atoms on the MXene surface. N 1s spectra (Fig. S10†) confirmed the presence of pyridinic-N, pyrrolic-N, and graphitic-N functionalities on the MOF and H-MOF6/MXene samples. Due to the intimacy of the interaction between the MOF and MXene in H-MOF6/MXene, peak positions were shifted to lower binding energy with dense oxidation-N energy states. The Co 2p peaks of H-MOF and H-MOF6/MXene-based heterostructures are shown in Fig. 2j. Pristine ZIF-67 had two split peaks at 780.13/795.09 eV and 781.6/797.5 eV, which were ascribed to the Co3+ and Co2+ electronic states and corresponding satellite peaks at 785.8 and 802.9 eV. Compared to ZIF-67, the H-MOF6/MXene-based heterostructures showed a slight positive shift in binding energy, signifying electron transfer from the MOF to MXene. The surface area and the pore size distribution of MXenes and ZIF-67/MXene were obtained by BET analysis (Fig. 2k). The N2 absorption–desorption isotherm represented a type 1a isotherm for the H-MOF6/MXene hybrid with a BET surface area of 812.44 m2 g−1 and a micropore volume of 0.1217 cm3 g−1. The surface areas of ZIF-67 (1656.11 m2 g−1) and H-ZIF-67 (1488.96 m2 g−1) showed a minimal decrease in BET surface area and pore volume after the SLER process. The MOF6/MXene hybrid had a slightly higher specific surface area (755.95 m2 g−1) and pore volume (0.1346 cm3 g−1) than the H-MOF6/MXene hybrid because the DMBIM treatment reduced the porosity of the ZIF-67 surface. Nonetheless, the exposed pore volume of H-MOF6/MXene, which was greater than that of the pristine MXene nanosheets, was sufficient to facilitate rapid gas adsorption.
The gas sensing performances of MXenes, ZIF-67, MOF6/MXene, H-MOF6/MXene, and physically mixed MXene/H-ZIF-67 (10:
1) composites were investigated at RT using compressed air as a carrier gas and different target gases. Over the interdigitated Au electrode, the MXene-based sensor exhibited a thickness of around 35 μm within an area of 0.5 × 0.5 cm (Fig. S11†). Fig. 3a shows gas responses over different MXene-based heterostructures to 10 ppm of NH3. Pristine MXenes produced a response of 1.37% to 10 ppm of NH3. Pristine H-ZIF-67 and physically mixed MXene/H-ZIF-67 nanocomposite sensors were not sensitive at RT because of their large resistances or insulating characteristics. MOF6/MXene and H-MOF6/MXene sensors exhibited responses of 8.73% and 10.86%, respectively. Moreover, DMBIM functionalization slightly increased the sensing response of ZIF-67/MXene. The heterostructure of the MOF on the MXene hybrid had excellent charge transfer capacity through a Schottky junction and provided high sensing performances. H-MOF6/MXene sensors exhibited rapid response and recovery (25 and 192 s) to 1 ppm of NH3 at RT (Fig. S12†), and the H-MOF6/MXene sensor had a high recovery rate compared to layered MXene. The mass ratio of the MOF on MXenes crucially influenced the sensing ability of hybrid sensors. Fig. 3b shows the sensing abilities of ZIF-67/MXene with different surface mass loadings of the MOF. Surface conductivity significantly changed as a function of MOF tagging. Increasing the mass loading of ZIF-67 (6 wt%) in the H-MOFx/MXene hybrid increased sensor response. However, further increasing the ZIF-67 loading reduced the conductivity and resulted in poor sensing performance.
Fig. 3c shows the electrical resistance of the sensing materials when exposed to NH3 at different times. The resistance of the pristine MXene nanosheets increased, probably due to carrier reduction after NH3 adsorption and the effect of this on conductivity.42 However, the resistance of MOF/MXene decreased, revealing n-type semiconductor behavior.43 The interaction between NH3 and oxygen functionalities on the heterostructure surface increased the carrier concentration on ZIF-67/MXene and reduced sensor resistance. The MOF network on MXenes can act as a bridge for carrier transfer that increases the internal redox reaction rate and enhances gas-sensing ability. The dynamic response (Fig. 3d) of the H-MOF6/MXene hybrid was more sensitive to NH3, and its limit of detection (LOD) was lower than that of MXenes due to a greater number of adsorption sites on the heterostructure surface. MOF/MXene provided a robust, stable response even at low NH3 concentration (200 ppb) (Fig. 3e). In particular, H-MOF6/MXene exhibited a highly stable response to NH3 with good sensitivity and excellent linearity from 800 ppb to 50 ppm (Fig. 3f). The LODs of H-MOF6/MXene and MXene sensors were 12.8 and 59.4 ppb, respectively. Selective adsorption of NH3 through surface ZIF-67 enhanced selectivity due to an electronic interaction within the zeolite functionality.44 A comparison of sensing response and durability over five consecutive cycles showed that the H-MOF6/MXene sensor produced a strong, durable response (Fig. 3g) to 20 ppm NH3, which would be helpful for long-term applications. To study the stability of this sensor, an aging test of the MXene-based sensors was performed over 8 weeks at a relative humidity of 51% at room temperature. Response to 10 ppm NH3 was reduced by only 5.2%, which showcased excellent long-term stability (Fig. 3h). Response reductions are mainly due to surface oxidization of the MXene and result in increased conductivity. Increasing the DMBIM concentration by increasing diffusion time did not appreciably influence the sensing response (Fig. S13†). Due to the molecular sieve effect, DMDIM is too large to pass through ZIF-67, and hence ligand exchange only occurred on the outermost surface of ZIF-67.35Fig. 3i shows the temperature response and temperature resistance of MXene, MOF6/MXene, and H-MOF6/MXene sensors for 20 ppm NH3 within the temperature range from 20 °C to 70 °C. While increasing temperature, the resistance of the sensor decreases owing to thermal excitation, which further decreases the response value. Near room temperature, the sensors exhibited high response. On increasing the substrate temperature, the sensing response was reduced due to the poor adsorption of NH3 on the MXene-based sensing surface. The resistance of the MOF/MXene is increased on modifying H-MOF6/MXene sensors, which may be due to the surface functionalization that decreases the contact interface and impedes the electron transportation between MXene nanosheets. With respective to the temperature, H-MOF6/MXene exhibited a higher response compared to MOF6/MXene, which may be due to the broadened depletion layer created on H-MOF6/MXene that increases the ability of MXenes to accept free electrons from NH3.
Humidity presents a serious challenge because the adsorption of water molecules on active surface layers, especially on hydrophilic surfaces like MXenes, directly affects sensor responses. However, a hydrophobic active sensing surface effectively minimizes water interference (Fig. 4a). Fig. 4b and c show the typical dynamic gas sensing responses of the MOF6/MXene and H-MOF6/MXene sensors at different NH3 concentrations and RH values (10, 36, 51, 76, and 93%) to 10 ppm NH3. Notably, the current responses of MOF6/MXene and H-MOF6/MXene sensors were good in a broad range of NH3 concentrations (1–50 ppm). Notably, the MOF6/MXene sensor response increased with RH (increases of 1.45% and 1.33% versus response at RH values of 76% and 93%, respectively). This was likely due to competition between NH3 and H2O molecules for adsorption on the sensor surface. On the other hand, H-MOF6/MXene sensors exhibited selectivity for NH3 with RH tolerance. H-MOF6/MXene showed changes of 0.22% at 72% and 0.27% at 93% versus an RH of 10%. The selectivity coefficient (SCRH) of MOF6/MXene was 1.08 at 51% RH, while that of H-MOF6/MXene was 0.998 at the same RH. The contact angles of MXenes, MOF6/MXene, and H-MOF6/MXene were measured to be 61°, 83°, and 121°, respectively (Fig. 4b, c and S14†). The large contact angle of H-MOF6/MXene was attributed to low surface energy due to its hydrophobic nature. Thus, the passive layer of DMDIM on MOF/MXene reduced surface affinity for water and protected the layered sensing surface from humidity. The responses of MOF6/MXene and H-MOF6/MXene sensors were also tested at different RH values in the absence of NH3 (Fig. S15†). While the responses of the MOF6/MXene sensor increased with humidity, the H-MOF6/MXene sensor exhibited only a minimal increase when varying the RH% to address the anti-humidity performance of the sensor. For the H-MOF6/MXene sensors, DMDIM functionality acts as a shield that minimizes the adsorption of water molecules on the sensing surface due to the hydrophobic nature. Therefore, water molecules cannot react with chemisorbed oxygen ions (O−) and induce free electrons, so the response for the H-MOF6/MXene sensor at different humidity levels remains almost the same.
Designing the MOF on MXene heterostructure assembly provided a high sensitivity and selectivity towards NH3. The selectivity of the different sensors is presented in Fig. 4d–f in the presence of different volatile organics. The MXene surface responded to most analytes (Fig. 4d), whereas MOF6/MXene exhibited better selectivity for NH3 with a higher response (Ra/Rg = 13.37, 10 ppm) than acetone (Ra/Rg = 1.41, 10 ppm), ethanol (Ra/Rg = 1.24, 10 ppm), methanol (Ra/Rg = 1.45, 10 ppm), formaldehyde (Ra/Rg = 2.12, 10 ppm), CO2 (Ra/Rg = 0.61, 10 ppm), triethylamine (Ra/Rg = 2.14, 10 ppm), and toluene (Ra/Rg = 1.36, 10 ppm). This selectivity for NH3 was due to the selective adsorption of NH3 on the ZIF-67 surface and the Schottky junction between the ZIF-67 and MXene, which promoted effective carrier transportation. H-MOF/MXene had an NH3 sensing response nearly 8 times greater than those of VOCs (Fig. 4f), and this response was better than those in previous reports.19,25 The conductivity of the MXene and the high absorption characteristics of ZIF-67 increased the charge carrier transfer rate on MOF/MXene, indicating a gas sensing mechanism involving a Schottky junction. Thus, the structural defects endowed by the MOF on MXenes, sufficient active sites, and a p–n multi-phase heterointerface increased electron transport capacity. To further address the superiority of the MOF/MXene sensor, Table 1 summarizes and compares the gas sensing response with that of other reported MXene-based sensors. The designed MOF/MXene sensor demonstrates high sensitivity and offers the advantages of low pLOD and response/recovery time.
Material | Gas | Gas concentration | Response | T res/Trec (s) | pLOD | Ref. |
---|---|---|---|---|---|---|
a S = (Ra − Rg)/Ra × 100% or S = (Rg − Ra)/Rg × 100%. b S = Ra/Rg. | ||||||
H-MOF6/MXene | NH3 | 10 ppm | 13.37a | 25/192 | 12.8 ppb | This work |
Co–TCPP(Fe)/Ti3C2Tx | NO | 10 ppm | 2.2b | 95/15 | 200 ppb | 19 |
Ti3C2Tx–PVDF–ZIF-67 | NH3 | 25 ppm | 10.56a | 41/75 | 1 ppb | 25 |
MXene/In2O3 | NH3 | 20 ppm | 100.7a | 60/300 | 1000 ppb | 45 |
MXene/graphene | NH3 | 100 ppm | 25.0a | 26/148 | 56 ppb | 46 |
MXene/SnO2 | NH3 | 50 ppm | 40a | 36/44 | 4.29 ppb | 47 |
Ti3C2Tx/TiO2 | NH3 | 10 ppm | 3.1a | 60/750 | 500 ppb | 48 |
Ti3C2Tx/WSe2 | Ethanol | 40 ppm | 9.2a | 9.7/6.6 | — | 49 |
MoS2/Ti3C2Tx | NO2 | 20 ppm | 40.1a | 525/155 | — | 50 |
Ti3C2Tx–TiO2 | Hexanal | 10 ppm | ∼3.4b | 293/461 | 10 ppm | 51 |
α-Fe2O3/Ti3C2Tx | Acetone | 5 ppm | 16.6a | 5/5 | 0.56 ppm | 52 |
CoPM-24 | NOx | 100 ppm | 27.9b | 2/73 | 30 ppb | 53 |
MXene@TiO2/MoS2 | NH3 | 100 ppm | 163.3a | 117/88 | — | 54 |
MXene/WS2 | NO2 | 2 ppm | 17.4a | 407/336 | 10 ppb | 55 |
In2O3/ZnO/MXene | Ethanol | 100 ppm | 6.5b | 48/116 | — | 56 |
2H-MoS2/Ti3C2Tx | NO2 | 100 ppm | 65.6a | 6.2/5.1 | 1.15 ppm | 57 |
CuO/Ti3C2Tx | NO2 | 100 ppm | 38.54a | 2.1/32.4 | 0.003 ppm | 58 |
CPAM/Ti3C2Tx | NH3 | 150 ppm | 3.1b | 12.7/14.6 | — | 59 |
SnO/SnO2/Ti3C2Tx | Acetone | 100 ppm | 12.1b | 18/9 | — | 60 |
The I–V characteristics of MXene hybrids were investigated to obtain insights into sensitivities, contact interfaces, and mechanisms. Sensors were fabricated on Au interdigitated electrodes, and total sensor resistances were expressed as Rtotal(Rtotal = Rmaterial + Rcontact + Relectrodes). The I–V characteristics of MXenes on the Au electrode exhibited a symmetrical linear characteristic representing an ohmic contact (Fig. 5a). On the other hand, the MOF on the Au electrode had poor electrical response due to its low conductivity at RT. On the other hand, the I–V curve of MOF6/MXene on the Au electrode surface indicated a symmetrical Schottky contact interface, and that of H-MOF6/MXene a symmetrical contact with non-rectifying behavior. The reverse current was slightly less than the forward current for H-MOF6/MXene. Additionally, XPS analysis of the O 1s spectra of H-MOF6/MXene (Fig. 5b) reveals that after the adsorption of NH3, the ratio of surface chemisorbed oxygen/surface hydroxyl groups on H-MOF6/MXene was reduced to 27%, demonstrating the chemical interaction of ammonia on the sensing surface promoting electron transfer. The peak shifted towards the right by 0.4 eV after NH3 adsorption evidencing the metal–hydroxyl formation on the sensing surface.61 This result shows that surface-adsorbed O functionalities also participate in sensing. Furthermore, the shift shown by Co 2p to higher binding energy after exposure to NH3 resulted in charge transfer between MOF/MXene and adsorbed NH3 (Fig. S16†). The FTIR spectrum of H-MOF6/MXene after NH3 adsorption (Fig. 5c) showed N–H stretch at 648, 1356, and 3335 cm−1, which resulted in NH3-related products being adsorbed on H-MOF6/MXene as NH3−. On introducing ZIF-67 on the MXene surface, the selectivity for NH3 has been promoted through the selective adsorption properties towards the NH3 gases.62 The selective pore size and acid–base interaction within the zeolite groups induce the selectivity towards NH3.25 Meanwhile, MXenes have high sensitivity towards NH3 through –OH, and –O based surface functionalities due to their high adoption energy as compared with that of other analytes such as CO2, NO2, CH4, and CO.63 On the other hand, a building block of MOF on MXenes as a hybrid interface induced selectivity by blocking specific pores, enabling only the selectivity towards NH3, which minimizes the interaction of other gases and VOCs. The loading density of the MOF on MXenes plays a crucial role in the conductivity of MXenes that impacts the sensitive response of the hybrid surface. Furthermore, the rapid desorption of NH3 from the hybrid surface would aid sensor recovery (Fig. S17†). The work functions of MXenes and the MOF were obtained from previous reports,19,64 and the band gap was calculated using UV-DRS absorption spectra and a K–M plot (Fig. S18†). Band gaps for MXenes, ZIF-67, MOF6/MXene, and the H-MOF6/MXene hybrid were 1.61, 1.96, 1.86, and 1.90 eV, respectively. Band gaps are crucial parameters because narrow band gaps favor charge transfer. The highest occupied molecular orbital (HOMO) and the lowest occupied molecular orbital (LOMO) of ZIF-67–MXene before and after assembly are provided in Fig. 5d. The Ef of MXenes was close to the HOMO which is represented by the p-type semiconductor behavior. The ZIF-67/MXene heterostructure exhibited the n-type semiconductor behavior which shows that the Ef will be close to the LOMO.
According to predicted work functions, an external bias voltage causes electrons to migrate from the MOF to MXenes, which is favored by Schottky junction formation and in line with our XPS results. After forming the heterostructure, the Ef of MOF/MXene was close to the MOF LOMO level, resulting from n-type semiconductive behavior which has been confirmed through the Mott–Schottky plot of the H-MXe6/MXene heterostructure (Fig. S19†). The electrochemical characteristics of the charge transfer capabilities of the MXene, MOF, and MOF6/MXene hybrid were analyzed (Fig. S20†). The equivalent circuit diagram of the H-MOF6/MXene electrode was provided with the respective specific impedance values. H-MOF6/MXene had the smallest Nyquist circle next to the MXene with low internal charge transfer resistance and improved charge carrier transportation efficiency. The internal charge transfer resistance of H-MOF6/MXene was lower than that of the MOF, which showed that the Schottky interface of MXenes promotes charge transfer efficiency. In addition, MOF integrity on MXenes would lead to favorable energy band positions and a Fermi level that reduces the electrical conductivity of Ti3T2Tx MXene. The NH3 sensing mechanism of the MOF on MXenes is shown in Fig. 5e. The H-MOF6/MXene heterostructure exhibited intimate contact between H-ZIF-67 and MXenes due to hydrogen bond formation. The heterostructure interface created a Schottky junction with a narrow band function, and the Ef of ZIF-67 in the vicinity of MXenes enabled charge transfer from the MOF/MXene surface to NH3. The diffusion coefficient of the analyte also plays a significant role in the hybrid surface. Loading the MOF on MXenes created many active porous features, which allowed analytes to flow through the rough features of the layered surface. The diffusion coefficient is inversely related to the square root of the molecular weight (g g−1 mol−1) of the target gas. Among the gases examined, NH3 (∼17.03) is the lightest, followed by methanol (∼32.04), CO2 (∼44.01), ethanol (∼46.07), and acetone (∼58.08), which resulted in the high possible interaction of NH3 on the sensing surface.25 Furthermore, when the H-MOF/MXene sensor was exposed to air, O2 molecules captured electrons and converted them into chemisorbed O2− species at the heterostructure interface, and when such sensors were exposed to NH3, target gases interacted with species, such as OH and –O, and captured electrons from the MOF to form NH3− (NH3 + e− → NH3−). As a result, the MOF has reduced the resistance of the MXene surface, which generated gas-sensing signals. The surface-tagged MOF on MXenes would prevent aggregation and maximize gas adsorption, which, in turn, would maximize free electron density on the sensor surface. Furthermore, the adsorption of chemisorbed oxygen would enhance the electron-depleted layer and the potential barrier at the contact interface. The creation of a space charge region at the interface would maximize the generation of adsorbed oxygen ions on the surface and increase sensor resistance. On the other hand, less interaction with NH3 would release more electrons, reduce the width of the electron depletion layer, increase the conductivity of the MOF/MXene heterostructure, and thus, improve sensitivity for NH3 detection. In addition, our observations showed that excessive amounts of MOF on MXenes destroy the Schottky junction. The rapid carrier response of the MOF underlies its reversible response, and MXenes play a supportive role as a conductive layer that contributes to gas sensing performance. The work function is a significant parameter for constructing high-performance heterostructured sensing materials with high sensitivity and selectivity.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta04656k |
This journal is © The Royal Society of Chemistry 2024 |