Next Article in Journal
Mesomechanical Aspects of the Strain-Rate Sensitivity of Armco-Iron Pulled in Tension
Next Article in Special Issue
Advances in the Field of Graphene-Based Composites for Energy–Storage Applications
Previous Article in Journal
Simulation of the Void Shape Evolution of High-Temperature Annealed Silicon Structures by means of a Custom Level-Set Formulation
Previous Article in Special Issue
Novel Mixed-Phase α/γ-Fe2O3 Micro-Flower Assembled with Nanosheets for Enhancing Acetone Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Porous, Tremella-like NiFe2O4 with Ultrathin Nanosheets for ppb-Level Toluene Detection

1
School of Materials and Energy, Yunnan University, Kunming 650091, China
2
Research School of Chemistry, Australian National University, Canberra 2601, Australia
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(6), 865; https://doi.org/10.3390/cryst13060865
Submission received: 6 May 2023 / Revised: 21 May 2023 / Accepted: 23 May 2023 / Published: 25 May 2023

Abstract

:
As a typical spinel ferrite, NiFe2O4 is suitable for use in gas sensors. Herein, we report the fabrication of porous, tremella-like NiFe2O4 assembled using porous, ultrathin nanosheets via the coordination of Ni2+ and Fe2+ with 1,4-phenylenediboronic acid. The optical band gap of the NiFe2O4 is estimated to be about 1.7 eV. Furthermore, the NiFe2O4 sensor annealed at 400 °C exhibits a low detection limit of 50 ppb, a fast response/recovery time (11.6 s/41.9 s to 10 ppm toluene), good reproducibility, and long-term stability at 220 °C. The suitable sensing performances can be attributed to the good catalytic activity of NiFe2O4 to toluene oxidation. Moreover, the ultrathin nanosheets with porous structures provide a large number of active sites to significantly favor the diffusion and adsorption/desorption of toluene molecules. This current work provides an insight into fabricating NiFe2O4 using 1,4-phenylenediboronic acid, which is promising for ppb-level toluene detection.

1. Introduction

Toluene (C7H8) is a typical volatile organic compound (VOC) which is extensively used in medicine, leather, pesticides, laboratory solvents, pharmaceuticals, and so on [1,2,3]. However, toluene is harmful for both human health and the environment. Exposure to toluene could pose a risk to human health by damaging the central nervous system, liver, kidneys, heart, etc. [4,5,6]. It is reported that short-term exposure to 70 ppb of toluene may negatively affect the cognitive function of the brain [7]. Furthermore, toluene is reported to be a vital biomarker in the diagnosis of lung cancer. The concentrations of toluene in patients with lung cancer were reported to range from 80 to 100 ppb [8,9]. Hence, rapid toluene detection with excellent selectivity at an early stage is very promising for multiple applications in health care, environmental protection, and medical diagnoses. Among various techniques, resistive gas sensors fabricated with semiconductor metal oxides have attracted wide attention due to their advantages of easy fabrication and low cost [10,11]. However, it is still challenging to improve their sensing performance, especially for selective and sensitive ppb-level toluene detection.
Recently, spinel ferrite materials (AB2O4) have shown significant advantages in the detection of various gases due to their unique chemical compositions and crystal structures in which two cation sites are occupied by either transition or post-transition cations [12,13]. The combination of various transition and post-transition metal cations in spinel ferrites could be favorable for a highly sensitive and selective gas sensor. Thus far, sensitive materials such as NiFe2O4, ZnFe2O4, ZnCo2O4, NiCo2O4, and NiMn2O4 have been developed for gas sensors [14,15,16]. NiFe2O4 is one of the particularly important spinel ferrite materials which has recently been used in a gas sensor for the detection of toluene. Various synthesis methods have been developed to tune the morphology of NiFe2O4 to improve its sensing properties. For example, Lai et al. used KIT-6 as a template to successfully obtain mesoporous NiFe2O4, which exhibited an excellent toluene-sensing property [17]. Kim’s group synthesized 3D hierarchical NiFe2O4 porous microspheres with the assistance of urea and glucose. The NiFe2O4 sensor achieved a good toluene gas response at a low working temperature [18].Yang and his co-workers prepared NiFe2O4 nanomaterials using a co-precipitation method; these materials exhibited good sensing properties in their sensitivity and selectivity to toluene detection [19]. Paula et al. prepared NiFe2O4 nanoparticles via levitation-jet synthesis which were shown to be remarkably sensitive to toluene gas. In our previous work, Ni/Fe-MOF-derived 1D porous NiFe2O4 nanorods showed outstanding selectivity and sensitivity to toluene [20]. Even so, NiFe2O4-based gas sensors usually exhibit unsatisfied sensing properties in their low gas responses and high detection limits. Therefore, a novel synthesis method is highly desirable for fabricating NiFe2O4 with a unique microstructure to enhance its toluene detection to the ppb level. Recently, using an organic acid as both a coordinating reagent and a soft template has been an effective way of fabricating metal oxides [21,22]. Studies have shown that boric acid is capable of coupling with the metal ions of Pb2+, Cd2+, and Fe2+ to form nanosheet complexes [23,24,25]. Carboxylphenyl boronic acid was reported to complex with metal ions (such as Zn2+, Cd2+, Ni2+, abd Ln3+) to form 2D coordination polymer nanosheets [26,27]. These pioneered works provide us a clue for synthesizing NiFe2O4 nanosheets by taking 1,4-phenylenediboronic acid to coordinate with both Ni2+ and Fe2+ ions.
Herein, we successfully prepared porous, tremella-like NiFe2O4 with ultrathin nanosheets by using 1,4-phenylenediboronic acid to coordinate Fe2+ and Ni2+. Our prepared NiFe2O4 possesses good crystallinity, catalytic activity, a porous, ultra-thin nanosheet structure, and a large specific surface area (77.8 m2/g). The sensor based on NiFe2O4 annealed at 400 °C exhibits good toluene-gas-sensing properties, especially with a low detection limit of 50 ppb. The findings of this study demonstrate the possibility of using 1,4-phenylenediboronic acid to fabricate NiFe2O4 nanosheets which are then suitable for application in trace toluene detection.

2. Experimental Section

2.1. Fabrication of NiFe2O4

All the starting reagents were of analytical grade. Deionized water (18 MΩ cm at room temperature) was used for this work. The tremella-like NiFe2O4 was synthesized via a hydrothermal process followed a thermal treatment. Specifically, 0.5 mmol (0.1188 g) NiCl2·6H2O (AR 99.9%) and 1 mmol (0.1988 g) FeCl2·4H2O (AR 98%) were completely dissolved into the mixture of 50 mL N,N-dimethylformamide (DMF) (AR 99.5%) with 5 mL anhydrous ethanol and 5 mL deionized water. Then, 3 mmol (0.4973 g) of 1,4-phenylenediboronic acid (C6H8B2O4) (AR 95%) was further added under magnetic stirring for 30 min to form a homogenous solution. After a hydrothermal reaction at 130 °C for 12 h, the obtained product was separated via centrifugation and rinsed with deionized water and ethanol. Next, the precursor products were collected after being dried at 60 °C for 6 h. Finally, the precursor was annealed in air at 350 °C, 400 °C, and 450 °C for 2 h with a heating rate of 3 °C/min, and the corresponding NiFe2O4 samples were labeled S-350, S-400, and S-450, respectively.

2.2. Characterization

The crystal phase of the NiFe2O4 was analyzed on a Rigaku TTRIII X-ray diffraction (XRD) equipped with Cu Kα radiation (1.54056 Å) from a scanning angle of 10° to a scanning angle of 70°. A field emission scanning electron microscope (FESEM, FEI QUANTA 200, Hillsboro, OR, USA) was utilized to analyze the morphological features of all samples. Transmission electron microscopy (TEM) images were taken on JEOL JEM-2100 microscope at an accelerating voltage of 100 kV. The corresponding elemental mapping image was acquired by the EDS attachment on the FESEM. A thermogravimetric analysis (TGA-DSC) was performed to investigate the decomposition of the precursor at a heating rate of 10 °C/min (TA SDT-2960, New Castle, DE, USA). X-ray photoelectron spectra (XPS) were characterized to measure elemental valence states, and the electron binding energy was calibrated by referring to the C1s peak at 284.6 eV. The UV–visible absorption spectra were tested on a UV–VIS spectrophotometer (SolidSpec-3700). The Brunauer–Emmett–Teller (BET) surface area was derived from nitrogen isotherms measured on a surface adsorption instrument (ASAP 2020 Plus).

2.3. Sensor Fabrication and Testing Properties

The gas sensor was fabricated and tested using our general procedure [28,29]. The as-synthesized NiFe2O4 powders were dispersed into a certain amount of deionized water and sonicated into a floating solution. The above suspension solution was then coated onto the outer surface of the alumina tube and dried at 120 °C for 2 h. A pair of Au electrodes and corresponding Pt wires at both ends of the alumina tube were set. To ensure good contact between the gold electrode and the sensing material, the sensors were annealed in air at 300 °C for 2 h. The gas-sensing performances were tested using a WS-30 A system (Weisheng Instruments Co., Zhengzhou, China) with a volume of the closed chamber of 18 L. The desired gas concentration was obtained based on the static liquified gas distribution method [28]. The response (β) of the sensor is defined as:
β = Rg/Ra
where Rg and Ra stand for the resistance of the sensor in the target gas and air, respectively. In addition, the response and recovery time were described as the time taken for 90% of the entire resistance change to be achieved. During the test of the gas-sensing properties, the relative humidity (RH) was maintained around 30%.

3. Results and Discussion

The crystalline structures were examined via X-ray diffraction (XRD). The XRD curves of the NiFe2O4 samples are shown in Figure 1. In comparison with the standard card (JCPDS: 54-0964), a cubic NiFe2O4 structure (space-group: F d 3 ¯ m ) can be confirmed. Specifically, the distinct peaks are detected at 2θ of 30.3°, 35.7°, 37.3°, 43.3°, 57.3°, and 63°, corresponding to the Bragg planes of (220), (311), (222), (400), (511), and (440), respectively. The crystallite size of the NiFe2O4 is estimated by using Scherrer’s equation (Equation (2)):
D = k λ β cos   θ
where D, k, λ, β, and θ are the crystallite size, the Scherrer constant (0.89), the X-ray wavelength (1.54056 Å), the peak width at half, and Bragg’s diffraction angle, respectively. It can be calculated that the grain sizes of the S-400 NiFe2O4 and S-450 NiFe2O4 are 13.1 and 16.2 nm. Due to the poor crystallinity of the S-350, its grain size cannot be calculated. The X-ray diffraction peak of the materials has no peaks for other impurities, indicating that the NiFe2O4 is of a high quality. In Figure 1, the XRD peak becomes sharp as the annealing temperature increases, indicating strong crystallinity. The band gap is an important parameter for a metal oxide semiconductor. Therefore, the UV-vis absorption spectra of NiFe2O4 materials have been studied (Figure 2), and they are used to calculate the band gap using the following equation [29,30]:
α h v 2 = A ( h v E g )  
where α, h, ν, Eg, and A represent the absorption coefficient, Planck’s constant, light frequency, band gap energy, and a constant, respectively. The Tauc-plot of (αhv)2 versus the photo energy (hv) of the NiFe2O4 (S-350, S-400, and S-450) is as shown in Figure 2b–d. In general, the intercept between the linear portion of the (αhv)2 to hv plots on the hv axis represents the value of the band gap energy (Eg). From Figure 2b–d, it can be seen that the band gap is estimated to be 1.73 eV, 1.70 eV, and 1.72 eV for the S-350, S-400, and S-450 samples. Although the band gap depends strongly on the defect level and microstructure derived from the different synthetic methods, our values are comparable with some reported values of the ban gaps for pure NiFe2O4, as summarized in Table 1.
The morphology and microstructure of the NiFe2O4 were observed via SEM. As can be seen from Figure 3a,b, the NiFe2O4 precursor shows a nanosheet structure. After being annealed in air, the nanosheets became wrinkled into a tremella-like structure. Similar morphologies are observed after annealing at different temperature of 350 °C (Figure 3c,d), 400 °C (Figure 3e,f), and 450 °C (Figure 3g,h). The NiFe2O4 exhibits a tremella-like shape with a diameter of 3–4 μm (Figure 3e). In the highly magnified SEM image in Figure 3f, the NiFe2O4 consists of closely interlaced ultrathin nanosheets, resulting in a porous nanostructure.
The TEM characterization provides more details about the microstructures. It can be seen that the tremella-like S-400 NiFe2O4 is composed of subunits of ultrathin nanosheets (Figure 4a,b). The ultrathin nanosheet structure is the result of the complex formation of boric acid ligands with Fe2+ and Ni2+ [37]. The homogeneous distributions of the Ni, Fe, and O elements on the tremella-like NiFe2O4 are illustrated in Figure 4c. Interestingly, a large number of holes can be clearly observed from Figure 4d. These pores range in diameter from 5 nm to 40 nm. Some typical NiFe2O4 nanoparticles are from 9.7 nm to 14.2 nm in diameter, as shown in Figure 4e. The HRTEM image (Figure 4f reveals a lattice spacing of 0.25 nm, which is ascribed to the (311) plane of NiFe2O4, and the results are well matched with XRD analysis.
Taking the NiFe2O4 annealed at 400 °C as an example (S-400), the textural properties of the tremella-like NiFe2O4 nanosheets were further studied via a nitrogen adsorption–desorption test, as shown in Figure 5. Obviously, the NiFe2O4 adsorption isotherm belongs to type IV, according to the IUPAC classification [38]. Moreover, the material has an H3-type hysteresis loop, which may be due to the accumulation of several nanosheets forming mesoporous properties. As a result, the unique NiFe2O4 possesses a high specific surface area of 77.80 m2/g through the N2 adsorption–desorption measurement. In addition, the pore distribution can be estimated by the Barrett–Joyner–Halenda (BJH) method, which shows that the average pore size of the NiFe2O4 is centered at about 18.5 nm (inset in Figure 5). The TEM images in Figure 4d,e reveal that the NiFe2O4 comprises numerous nanoparticles, giving rise to the formation of pores. The specific area of the NiFe2O4 is strongly dependent on the synthesis methods and morphologies, as shown in Table 2. Notably, our synthesized tremella-like NiFe2O4 with nanosheets exhibits a larger specific surface area than common 0D (nanoparticles), 1D (nanofibers and fusiformis), and 3D (nanocubes, hexagonal bipyramids, nanospheres and microspheres) NiFe2O4 materials. Thus, we provide a facile hydrothermal method using 1,4-phenylenediboronic acid to fabricate porous NiFe2O4 with high specific surface area which will facilitate the molecules adsorption and desorption of gas molecules as well as oxidation on the surface, leading to an good gas-sensing performance [39]. Therefore, NiFe2O4 is expected to show good gas-sensing properties towards toluene detection.
1,4-phenylenediboronic acid is rich in carboxyl groups which will easily coordinate with the metal ions of Fe2+ and Ni2+ to form complexes [47]. In addition, 1,4-phenylenediboronic acid may serve as a capping agent or template for the formation of a nanosheet structure during the synthesis process. In an atmosphere in which DMF is used as a complex solvent, the borate anions in 1,4 p-phenylboronic acid are drawn together with Fe2+ and Ni2+ via electrostatic interaction and form complexes of specific two-dimensional structures [48]. With subsequent heat treatment, the pyrolysis process occurs in the complex. The NiFe2O4 was then synthesized after thermal treatment, which was studied via a thermogravimetric analysis (TGA), as shown in Figure 6. It can be seen that the weight loss of the sample is aroud 3.5% below 150 °C, which may originate from the loss of absorbed water in the precursor. As the temperature rises further to 350 °C, the decomposition of the precursor results in a weight loss of approximately 18.4%. The weight loss at this stage is ascribed to the thermal decomposition of the complexes formed from Ni2+, Fe2+ and 1,4-phenylenediboronic acid. There is an endothermic peak at a temperature of 257.08 °C, which may be the peak for the decomposition of organic matter to absorb energy. The decomposition of organic matter will release a large amount of carbon dioxide and nitrogen oxide gases [49]. The kinetics of this process are very intense, resulting in a large number of holes in most of the product. After that, a slight weight loss could orignate from the removal of the residual carbon and the further crystallization of the NiFe2O4.
To identify the elemental compositions and chemical bonding states, the NiFe2O4 nanosheets were subjected to an XPS analysis. As shown in Figure 7a, the coexistence of Ni, Fe, and O elements on the S-400 NiFe2O4 was confirmed. The Ni 2p spectrum of the S-400 NiFe2O4 (Figure 7b) exhibits two spin-orbit doublets characteristic of Ni 2p3/2 and Ni 2p1/2 located at 854.11 eV and 873.36 eV, respectively, as well as satellite peaks centered at 861.15 eV and 878.90 eV, corresponding to the characteristic of Ni2+ [50,51]. Meanwhile, the Ni 2p3/2 and Ni 2p1/2 peaks can be deconvoluted into two peaks at 856.06 eV and 872.69 eV, which are attributed to Ni3+ species. Similarly, in the Fe 2p spectrum of the S-400 NiFe2O4 (Figure 7c), peaks at 713.85 eV and 726.10 eV can be assigned to Fe3+. The binding energies of 710.53 eV and 723.40 eV correspond to Fe2+. The coexistence of Ni2+/Ni3+ and Fe2+/Fe3+ cations is essential for catalyzing toluene to enhance the sensor’s gas-sensing performance [52]. In particular, the formation of high-oxidation states of Ni3+ and Fe3+ could favor the transport of electrons and the formation of abundant oxygen ions to catalyze the reaction of the target gas molecules for a high degree of sensitivity [53]. The O 1s peak in XPS (Figure 7d–f) is asymmetric in shape and can be divided into two major peaks. Taking the S-400 sample as an example (Figure 7e), the low binding energy of 530.06 eV at the peak is characteristic of lattice oxygen (Olattice), which does not react with the target gas. Conversely, at relatively high binding energies (532.00 eV), a well-resolved peak is assigned to surface-adsorbed oxygen, which is thought to react with the target gas molecules to determine the gas-sensing characteristics [54]. Although a similar behavior of O 1s was observed, the ratios of adsorbed oxygen were estimated to be 13.12%, 23.79%, and 20.16% for S-350, S-400, and S-450 NiFe2O4 (Figure 7d–f), respectively. These results indicate that the S-400 NiFe2O4 could show a better sensing property because the adsorbed oxygen will evolve into active oxygen species to enhance the reaction with the target gas molecules [8].
The gas-sensing performance of NiFe2O4 is evaluated by the change in the conductance of the sensor in different concentrations of toluene. Figure 8a displays the responses of the sensor based on NiFe2O4 annealed at 400 °C at a temperature from 180 °C to 260 °C to toluene concentrations of 10 ppm and 30 ppm. Regardless of the toluene concentrations, the sensor shows an increase in its response to the maxmium value at 220 °C and then drops with the further increase in temperature. Due to the insufficient catalytic reaction of the toluene molecules on the NiFe2O4 surface, the response values are relatively lower at both lower and higher temperatures [55]. Therefore, the following toluene gas-sensing tests were performed at an optimal temperature of 220 °C. Figure 8b shows the dynamic responses of the NiFe2O4 annealed at 350 °C, 400 °C, and 450 °C to toluene at a concentration of 30 ppm at 220 °C. The S-400 based sensor exhibits a better response of 5.8 compared to the S-350 sensor (2.1) and the S-450 sensor (1.9). Therefore, the gas-sensing performance of the following study is mainly focused on the S-400. As can be seen from Figure 8c, the S-400 sensor shows a good dynamic response to a ppb level of toluene. The NiFe2O4 sensor (S-400) still shows a response of 1.1 to 50 ppb of toluene, suggesting a low limit of detection (LOD) at the ppb level. Additionally, the dynamic response of the NiFe2O4 sensor to various toluene concentrations (1–50 ppm) was examined, as shown in Figure 8d. In addition, the relationship between the responses and the toluene concentrations is shown in Figure 8e. We can see that the sensor shows a good dynamic response to toluene at low concentrations from 50 ppb to 5 ppm. A linear relationship between the response (β) and toluene concentration (C) is fitted by the following equation (Equation (4)), with a relative correlation coefficient of R2 = 0.9943:
β = 0.5287C + 1.1563
When the tolenene concentration exceeds 5 ppm, the S-400 NiFe2O4 sensor gradually becomes saturated. The NiFe2O4 sensor exhibits a response of 4.7 with a fast response (11.6 s) and recovery (41.9 s) time (Figure 8f) to 10 ppm of toluene. The fast response recovery time is attributed to the loose, porous structure of the ultra-thin nanosheets.
Figure 9a shows the response and recovery time of the NiFe2O4 gas sensor corresponding to the transient curve in the 0.05–50 ppm concentrations of toluene. The response time of the gas sensor decreases with the increase in toluene concentration. In addition, it shows a longer response time at lower concentrations. Kim et al. proposed a nonlinear reaction–diffusion model to explain the relationship between the response/recovery time and gas concentration [56], as shown in following equation:
τ = kx02C0r−1/D
where k, x0, C0, and D represent the reaction rate constant, film thickness, gas concentration, and diffusion coefficient, respectively. According to Equation (5), the gas sensor exhibits a fast response in a high concentration of toluene due to the rapid diffusion via large kinetics. Due to the slower kinetic process of adsorption–desorption and the ionization of surface oxygen during recovery [57], a longer recovery time is needed for a higher gas concentration. In brief, our NiFe2O4 sensor shows a comparable sensing performance for toluene detection compared with typical sensing materials in the literature, as summarized in Table 3. On the other hand, the responses toward other tested gases with concentrations of 10 ppm, such as xylene, acetone, ethanol, methanol, isopropanol, n-butyl alcohol, formaldehyde, and ammonia, have been examined, as shown in Figure 9b. Obviously, the response to toluene is about 2–3 times than that of the interfering gases due to the catalytic feature of p-type NiFe2O4 to toluene at the optimal operating temperature of 220 °C [58].To investigate the short-term repeatability, five consecutive tests of 10 ppm and 30 ppm of toluene were verified, as shown in Figure 9c, which shows a good repeatability. Finally, the long-term stability of one month was validated in Figure 9d. For instance, the responses to 30 ppm of toluene fluctuate around 6.1, suggesting a good stability over a long period of time. The sensor based on porous, tremella-like NiFe2O4 with nanosheets has the characteristics of a high response, a low detection limit, and fast response and recovery times, which are conductive to practical applications in toluene detection.
As a widely accepted sensing mechanism, the adsorbeed oxygen species (O, O2− and O2) play an crucial role in determining the sensing performance of a sensor based on semiconductor oxides. The NiFe2O4 is a p-type semiconductor whose conductivity carriers come from holes. A porous, tremella-like NiFe2O4 nanosheet reacts with adsorbed oxygen and its toluene target gas as follows [1,67]:
O 2 gas O 2 adsorbed
O 2 adsorbed   +   e O 2
O 2   +   e 2 O
O   +   e O 2 -
C 6 H 5 CH 3 + O C 6 H 5 CHO +   H 2 O
C 6 H 5 CH 3 + O 2 C 6 H 5 CHO +   H 2 O + e
C 6 H 5 CHO C 6 H 5 CHO + e
When exposed to air, the formation of oxygen species, as confirmed by the above XPS analysis, will capture free electrons from its conductance band. Upon exposure to toluene, the toluene molecules will react with these oxygen species, resulting in the release of the electrons. Accordingly, the change in conductance of the sensor occurs, achieving the detection of toluene. The true resistance change in the NiFe2O4 sensor at 1 ppm of toluene is shown in Figure 10a. The resistance of around 7.7 kΩ and 13.7 kΩ are tested in the air and in 1 ppm of toluene for the NiFe2O4 sensor, producing a response of 1.8. Figure 10b shows the reaction of the target gas toluene molecules in the porous nanosheets. First, oxygen is adsorbed onto the surface of the porous nanosheet, which then reacts with electrons in the conduction band, giving rise to active oxygen species (O, O2− and O2). After exposure to the target gas, the toluene molecule will react with active oxygen species to produce benzaldehyde and H2O molecules, as illustrated by the above equations (Equations (6)–(12)) [60,68].
The gas-sensing performance of this sensor is mainly ascribed to the following aspects. Fistly, the gas-sensing performance is derived from the unique spinel structure of the NiFe2O4, which possesses good catalytic activity. NiFe2O4 with a spinel structure shows a reversible redox reaction (Ni2+ + h+ ↔ Ni3+, Fe2+ + h+ ↔ Fe3+), resulting in the catalytic ability for the toluene to react on the surface of the NiFe2O4 at moderate temperatures [69]. Secondly, the unique, tremella-like structure assembled by the ultrathin nanosheets of NiFe2O4 favors a toluene-gas-sensitive performance. The unique morphology not only possesses a large specific surface area to provide abundant active sites but also provides abundant mesoporous structures as effective channels for gas diffusion. Finally, the porous, ultrathin NiFe2O4 nanosheet structure plays an important role in the highly sensitive detection of toluene, especially at low concentrations. The porous nanosheet structure not only provides a large number of active sites for gas reactions but also gas molecule transport channels, offering a preference for the detection of low concentrations of toluene.

4. Conclusions

In summary, the porous, tremella-like NiFe2O4 assembled with ultrathin nanosheets was successfully synthesized with the assistance of 1,4-phenylenediboronic acid. The band gap of the as-synthesized NiFe2O4 is around 1.7 eV. The NiFe2O4 annealed at 400 °C shows better performance at sensing toluene. It exhibits a good response of 4.7, fast response and recovery times of 11.6 s and 41.9 s to 10 ppm of toluene at 220 ℃, and a low detection limit of 50 ppb. Our good sensing performance originates from the original catalytic activity of NiFe2O4, the high surface area (77.80 m2/g), and the unique porous nanosheet structure. The as-prepared porous, tremella-like NiFe2O4 with ultrathin nanosheets using 1,4-benzenediboronic acid as a ligand provides a feasible strategy for the fabrication of NiFe2O4 to detect ppb-level toluene. Finally, it is also worth noting that the response time and selectivity of the NiFe2O4 sensor toward toluene could be further enhanced by doping, coupling with other sensing materials, or functionalizing noble metal nanoparticles for practical applications.

Author Contributions

Conceptualization, C.D.; Formal analysis, H.Q. and H.G.; Data curation, Y.Z.; Writing—original draft, Y.Z.; Writing—review & editing, C.D. and Z.Y.; Funding acquisition, C.G. and H.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data that support the findings of this study are available within this article.

Acknowledgments

This work was financially supported by the National Natural Science Foundation of China (22165032), the Application Basic Research Fund of Yunnan Province (2019FB129 and 202301AT070207), and the Major Science and Technology Project of Precious Metal Materials Genetic Engineering in Yunnan Province (2019ZE001-6). The authors acknowledge the Advanced Analysis and Measurement Center of Yunnan University for the sample-testing service.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mirzaei, A.; Kim, J.-H.; Kim, H.W.; Kim, S.S. Resistive-based gas sensors for detection of benzene, toluene and xylene (BTX) gases: A review. J. Mater. Chem. C 2018, 6, 4342–4370. [Google Scholar] [CrossRef]
  2. Yücel, M.; Takagi, M.; Walterfang, M.; Lubman, D.I. Toluene misuse and long-term harms: A systematic review of the neuropsychological and neuroimaging literature. Neurosci. Biobehav. Rev. 2008, 32, 910–926. [Google Scholar] [CrossRef] [PubMed]
  3. Ren, F.; Gao, L.; Yuan, Y.; Zhang, Y.; Alqrni, A.; Al-Dossary, O.M.; Xu, J. Enhanced BTEX gas-sensing performance of CuO/SnO2 composite. Sens. Actuators B 2016, 223, 914–920. [Google Scholar] [CrossRef]
  4. Donald, J.M.; Hooper, K.; Hopenhayn-Rich, C. Reproductive and developmental toxicity of toluene: A review. Environ. Health Perspect. 1991, 94, 237–244. [Google Scholar]
  5. Wheeler, A.J.; Wong, S.L.; Khoury, C.; Zhu, J. Predictors of indoor BTEX concentrations in Canadian residences. Health Rep. 2013, 24, 11–17. [Google Scholar]
  6. Hermawan, A.; Asakura, Y.; Inada, M.; Yin, S. A facile method for preparation of uniformly decorated-spherical SnO2 by CuO nanoparticles for highly responsive toluene detection at high temperature. J. Mater. Sci. Technol. 2020, 51, 119–129. [Google Scholar] [CrossRef]
  7. Zhan, C.; Liu, J.; Cui, M.; Wu, W.; Hao, F.; Wang, Z.; Liu, J.; Zhang, X.; Pan, L.; Deng, M.; et al. Alterations in functional connectivity of executive control network reveal influence of indoor toluene on cognition at low concentration. Build. Environ. 2023, 232, 110031. [Google Scholar] [CrossRef]
  8. Kim, N.-H.; Choi, S.-J.; Yang, D.-J.; Bae, J.; Park, J.; Kim, I.-D. Highly sensitive and selective hydrogen sulfide and toluene sensors using Pd functionalized WO3 nanofibers for potential diagnosis of halitosis and lung cancer. Sens. Actuators B 2014, 193, 574–581. [Google Scholar] [CrossRef]
  9. Seekaew, Y.; Wisitsoraat, A.; Phokharatkul, D.; Wongchoosuk, C.J.S.; Chemical, A.B. Room temperature toluene gas sensor based on TiO2 nanoparticles decorated 3D graphene-carbon nanotube nanostructures. Sens. Actuators B 2019, 279, 69–78. [Google Scholar] [CrossRef]
  10. Hunter, G.W.; Akbar, S.A.; Bhansali, S.; Daniele, M.A.; Erb, P.D.; Johnson, K.J.; Liu, C.C.; Miller, D.R.; Oralkan, O.; Hesketh, P.J.; et al. Editors’ choice-critical review-a critical review of solid state gas sensors. J. Electrochem. Soc. 2020, 167, 037570. [Google Scholar] [CrossRef]
  11. Yu, H.; Guo, C.; Zhang, X.; Xu, Y.; Cheng, X.-L.; Gao, S.; Huo, L.-H. Recent development of hierarchical netal oxides based gas sensors: From gas sensing performance to applications. Adv. Sustain. Syst. 2022, 6, 2100370. [Google Scholar] [CrossRef]
  12. Šutka, A.; Gross, K.A. Spinel ferrite oxide semiconductor gas sensors. Sens. Actuators B 2016, 222, 95–105. [Google Scholar] [CrossRef]
  13. Olowoyo, J.O.; Kriek, R.J. Recent Progress on Bimetallic-Based Spinels as Electrocatalysts for the Oxygen Evolution Reaction. Small 2022, 18, 2203125. [Google Scholar] [CrossRef] [PubMed]
  14. Kaneti, Y.V.; Septiani, N.L.W.; Saptiama, I.; Jiang, X.; Yuliarto, B.; Shiddiky, M.J.A.; Fukumitsu, N.; Kang, Y.-M.; Golberg, D.; Yamauchi, Y. Self-Sacrificial Templated Synthesis of a Three-Dimensional Hierarchical Macroporous Honeycomb-like ZnO/ZnCo2O4 Hybrid for Carbon Monoxide Sensing. J. Mater. Chem. A 2019, 7, 3415–3425. [Google Scholar] [CrossRef]
  15. Kim, S.-H.; Yun, K.-S. Room-temperature hydrogen gas sensor composed of palladium thin film deposited on NiCo2O4 nanoneedle forest. Sens. Actuators B 2023, 376, 132958. [Google Scholar] [CrossRef]
  16. Guan, Y.; Yin, C.; Cheng, X.; Liang, X.; Diao, Q.; Zhang, H.; Lu, G. Sub-ppm H2S sensor based on YSZ and hollow balls NiMn2O4 sensingelectrode. Sens. Actuators B 2014, 193, 501–508. [Google Scholar] [CrossRef]
  17. Lai, X.; Cao, K.; Shen, G.; Xue, P.; Wang, D.; Hu, F.; Zhang, J.; Yang, Q.; Wang, X. Ordered mesoporous NiFe2O4 with ultrathin framework for low-ppb toluene sensing. Sci. Bull. 2018, 63, 187–193. [Google Scholar] [CrossRef]
  18. Karuppasamy, K.; Sharma, B.; Vikraman, D.; Jo, E.-B.; Sivakumar, P.; Kim, H.-S. Switchable p-n gas response for 3D-hierarchical NiFe2O4 porous microspheres for highly selective and sensitive toluene gas sensors. J. Alloys Compd. 2021, 886, 161281. [Google Scholar] [CrossRef]
  19. Yang, L.; Xie, Y.; Zhao, H.; Wu, X.; Wang, Y. Preparation and gas-sensing properties of NiFe2O4 semiconductor materials. Solid-State Electron. 2005, 49, 1029–1033. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Jia, C.; Wang, Q.; Kong, Q.; Chen, G.; Guan, H.; Dong, C. Highly sensitive and selective toluene sensor of bimetallic Ni/Fe-MOFs derived porous NiFe2O4 nanorods. Ind. Eng. Chem. Res. 2019, 58, 9450–9457. [Google Scholar] [CrossRef]
  21. Cho, K.; Kang, C.W.; Ryu, S.H.; Jang, J.Y.; Son, S.U. The rise of morphology-engineered microporous organic polymers (ME-MOPs): Synthesis and benefits. J. Mater. Chem. A 2022, 10, 6950–6964. [Google Scholar] [CrossRef]
  22. Hu, M.-L.; Masoomi, M.Y.; Morsali, A. Template strategies with MOFs. Coord. Chem. Rev. 2019, 387, 415–435. [Google Scholar] [CrossRef]
  23. Wen, T.; Liu, M.; Chen, S.; Li, Q.; Du, Y.; Zhou, T.; Ritchie, C.; Zhang, J.S. 2D boron imidazolate framework nanosheets with electrocatalytic applications for oxygen evolution and carbon dioxide reduction reaction. Small 2020, 16, 1907669. [Google Scholar] [CrossRef] [PubMed]
  24. Shen, D.; Chen, X.; Luo, J.; Wang, Y.; Sun, Y.; Pan, J. Boronate affinity imprinted Janus nanosheets for macroscopic assemblies: From amphiphilic surfactants to porous sorbents for catechol adsorption. Sep. Purif. Technol. 2021, 256, 117837. [Google Scholar] [CrossRef]
  25. Zhang, Q.-P.; Sun, N.; Zheng, J.; Wang, S.-W.; Wu, Y.; Luo, D.-L.; Zhou, Y.-L. Fabrication of lead borate single crystal nanosheets for attenuating both neutron and gamma radiations. Adv. Eng. Mater. 2017, 19, 1600650. [Google Scholar] [CrossRef]
  26. Dimowa, L.; Nikolova, R.; Dyulgerov, V.; Kostov-Kytin, V.; Shivachev, B. Synthesis of MOFs from Zn,Cd,Ni-4-carboxyphenylboronic acid with DMF solvent. Acta Crystallogr. A 2014, 70, C1129. [Google Scholar] [CrossRef]
  27. Abdallah, A.; Daiguebonne, C.; Suffren, Y.; Rojo, A.; Demange, V.; Bernot, K.; Calvez, G.; Guillou, O. Microcrystalline core-shell lanthanide-based coordination polymers for unprecedented luminescent properties. Inorg. Chem. 2019, 58, 1317–1329. [Google Scholar] [CrossRef]
  28. Dong, C.; Tian, R.; Zhang, Y.; Liu, K.; Chen, G.; Guan, H.; Yin, Z. MOF-on-MOF nanoarchitecturing of Fe2O3@ZnFe2O4 radial-heterospindles towards multifaceted superiorities for acetone detection. Chem. Eng. J. 2022, 442, 136094. [Google Scholar] [CrossRef]
  29. Dong, C.; Tian, R.; Qu, H.; Tan, H.; Chen, G.; Guan, H.; Yin, Z. Anchoring Pt particles onto mesoporousized ZnO holey cubes for triethylamine detection with multifaceted superiorities. Small 2023, 19, 2300756. [Google Scholar] [CrossRef]
  30. Henderson, B.; Imbusch, G.F. Optical Spectroscopy of Inorganic Solids; Oxford University Press: New York, NY, USA, 1989. [Google Scholar]
  31. Tien, N.A.; Mittova, V.; Sladkopevtsev, B.; Mai, V.; Mittova, I.; Vuong, B. Structural, optical and magnetic properties of Y-doped NiFe2O4 nanoparticles prepared by simple co-precipitation method. Mater. Solid State Sci. 2023, 138, 107149. [Google Scholar] [CrossRef]
  32. Chiad, B.T.; Mohammed, R.A. Spectroscopy, optical properties of NiFe2O4 nanoparticles prepared by co-sputtering technique in different gasses mixture ratio. Int. J. New Technol. Res. 2017, 3, 56–58. [Google Scholar]
  33. Tong, S.-K.; Chi, P.-W.; Kung, S.-H.; Wei, D.-H. Tuning bandgap and surface wettability of NiFe2O4 driven by phase transition. Sci. Rep. 2018, 8, 1338. [Google Scholar] [CrossRef] [PubMed]
  34. Gupta, N.K.; Ghaffari, Y.; Kim, S.; Bae, J.; Kim, K.S.; Saifuddin, M. Photocatalytic degradation of organic pollutants over MFe2O4 (M=Co, Ni, Cu, Zn) nanoparticles at neutral pH. Sci. Rep. 2020, 10, 4942. [Google Scholar] [CrossRef]
  35. Dileep, K.; Louaya, B.; Pachauri, N.; Gupta, A.; Datta, R. Probing optical band gaps at the nanoscale in NiFe2O4 and CoFe2O4 epitaxial films by high resolution electron energy loss spectroscopy. J. Appl. Phys. 2014, 116, 103505. [Google Scholar] [CrossRef]
  36. Sun, Q.C.; Mazumdar, D.; Ma, J.X.; Holinsworth, B.S.; O’Neal, K.R.; Kim, G.; Butler, W.H.; Gupta, A.; Musfeldt, J.L. Optical band gap hierarchy in a magnetic oxide: Electronic structure of NiFe2O4. Phys. Rev. B 2012, 86, 205106. [Google Scholar] [CrossRef]
  37. Timmerman, M.A.; Xia, R.; Le, P.T.P.; Wang, Y.; ten Elshof, J.E. Metal oxide nanosheets as 2D building blocks for the design of novel materials. Chem. Eur. J. 2020, 26, 9084–9098. [Google Scholar] [CrossRef] [PubMed]
  38. Sing, K.S.W. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  39. Majhi, S.M.; Ali, A.; Rai, P.; Greish, Y.E.; Alzamly, A.; Surya, S.G.; Qamhieh, N.; Mahmoud, S.T. Metal-organic frameworks for advanced transducer based gas sensors: Review and perspectives. Nanoscale Adv. 2022, 4, 697–732. [Google Scholar] [CrossRef]
  40. Liu, N.; Wang, X.-F.; Zhang, G.; Liang, H.; Li, T.; Zhao, Y.; Zhang, T.; Tan, Z.; Song, X.-Z. Metal-organic framework-derived porous NiFe2O4 nanoboxes for ethyl acetate gas sensors. ACS Appl. Nano Mater. 2022, 5, 14320–14327. [Google Scholar] [CrossRef]
  41. Song, X.-Z.; Meng, Y.-L.; Chen, X.; Sun, K.-M.; Wang, X.-F. Hollow NiFe2O4 hexagonal biyramids for high-performance n-propanol sensing at low temperature. New J. Chem. 2018, 42, 14071–14074. [Google Scholar] [CrossRef]
  42. Zhang, Y.; Zhou, Y.; Li, Z.; Chen, G.; Mao, Y.; Guan, H.; Dong, C. MOFs-derived NiFe2O4 fusiformis with highly selective response to xylene. J. Alloys Compd. 2019, 784, 102–110. [Google Scholar] [CrossRef]
  43. Zhou, T.; Zhang, T.; Zeng, Y.; Zhang, R.; Lou, Z.; Deng, J.; Wang, L. Structure-driven efficient NiFe2O4 materials for ultra-fast response electronic sensing platform. Sens. Actuators B 2018, 255, 1436–1444. [Google Scholar] [CrossRef]
  44. Bameri, I.; Saffari, J.; Baniyaghoob, S.; Ekrami-Kakhki, M.-S. Synthesis of magnetic nano-NiFe2O4 with the assistance of ultrasound and its application for photocatalytic degradation of Titan Yellow: Kinetic and isotherm studies. Colloids Interface Sci. Commun. 2022, 48, 100610. [Google Scholar] [CrossRef]
  45. Zeynizadeh, B.; Mohammadzadeh, I.; Shokri, Z.; Ali Hosseini, S. Synthesis and characterization of NiFe2O4@Cu nanoparticles as a magnetically recoverable catalyst for reduction of nitroarenes to arylamines with NaBH4. J. Colloid Interface Sci. 2017, 500, 285–293. [Google Scholar] [CrossRef] [PubMed]
  46. Toghan, A.; Khairy, M.; Kamar, E.M.; Mousa, M.A. Effect of particle size and morphological structure on the physical properties of NiFe2O4 for supercapacitor application. J. Mater. Res. Technol. 2022, 19, 3521–3535. [Google Scholar] [CrossRef]
  47. Wang, X.-M.; Hu, Z.-J.; Guo, P.-F.; Chen, M.-L.; Wang, J.-H. Boron-modified defect-rich molybdenum disulfide nanosheets: Reducing nonspecific adsorption and promoting a high capacity for isolation of immunoglobulin G. ACS Appl. Mater. Interfaces 2020, 12, 43273–43280. [Google Scholar] [CrossRef]
  48. Ling, Z.; Wang, Z.; Zhang, M.; Yu, C.; Wang, G.; Dong, Y.; Liu, S.; Wang, Y.; Qiu, J. Sustainable synthesis and assembly of biomass-derived B/N co-doped carbon nanosheets with ultrahigh aspect ratio for high-performance supercapacitors. Adv. Funct. Mater. 2016, 26, 111–119. [Google Scholar] [CrossRef]
  49. Wei, H.; Zhang, H.; Song, B.; Yuan, K.; Xiao, H.; Cao, Y.; Cao, Q. Metal–Organic Framework (MOF) Derivatives as Promising Chemiresistive Gas Sensing Materials: A Review. Int. J. Environ. Res. Public Health 2023, 20, 4388. [Google Scholar] [CrossRef]
  50. Yang, H.; Liu, Y.; Luo, S.; Zhao, Z.; Wang, X.; Luo, Y.; Wang, Z.; Jin, J.; Ma, J. Lateral-size-mediated efficient oxygen evolution reaction: Insights into the atomically thin quantum dot structure of NiFe2O4. ACS Catal. 2017, 7, 5557–5567. [Google Scholar] [CrossRef]
  51. Wu, Z.; Zou, Z.; Huang, J.; Gao, F. NiFe2O4 nanoparticles/NiFe layered double-hydroxide nanosheet heterostructure array for efficient overall water splitting at large current densities. ACS Appl. Mater. Interfaces 2018, 10, 26283–26292. [Google Scholar] [CrossRef]
  52. Jia, Y.; Jian, B.; Wang, X.; Wang, C.; Sun, Y.; Zhang, H.; Shimanoe, K.; Lu, G. MOF-derived porous NiO/NiFe2O4 nanocubes for improving the acetone detection. Sens. Actuators B 2022, 366, 131985. [Google Scholar]
  53. Hussain, A.; Zhang, X.; Shi, Y.; Bushira, F.A.; Barkae, T.H.; Ji, K.; Guan, Y.; Chen, W.; Xu, G. Generation of oxygen vacancies in metal-organic framework-derived one-dimensional Ni0.4Fe2.6O4 nanorice heterojunctions for ppb-level diethylamine gas sensing. Anal.Chem. 2023, 95, 1747–1754. [Google Scholar] [CrossRef]
  54. Shuai, Y.; Peng, R.; He, Y.; Liu, X.; Wang, X.; Guo, W. NiO/BiVO4 p-n heterojunction microspheres for conductometric triethylamine gas sensors. Sens. Actuators B 2023, 384, 133625. [Google Scholar] [CrossRef]
  55. Wang, C.; Cheng, X.; Zhou, X.; Sun, P.; Hu, X.; Shimanoe, K.; Lu, G.; Yamazoe, N. Hierarchical α-Fe2O3/NiO composites with a hollow structure for a gas sensor. ACS Appl. Mater. Interfaces 2014, 6, 12031–12037. [Google Scholar] [CrossRef] [PubMed]
  56. Choi, S.-J.; Chattopadhyay, S.; Kim, J.J.; Kim, S.-J.; Tuller, H.L.; Rutledge, G.C.; Kim, I.-D. Coaxial electrospinning of WO3 nanotubes functionalized with bio-inspired Pd catalysts and their superior hydrogen sensing performance. Nanoscale 2016, 8, 9159–9166. [Google Scholar] [CrossRef]
  57. Wen, Z.; Tian-mo, L. Gas-sensing properties of SnO2-TiO2-based sensor for volatile organic compound gas and its sensing mechanism. Physica B Condens. Matter 2010, 405, 1345–1348. [Google Scholar] [CrossRef]
  58. Song, X.-Z.; Sun, F.-F.; Dai, S.-T.; Lin, X.; Sun, K.-M.; Wang, X.-F. Hollow NiFe2O4 microspindles derived from Ni/Fe bimetallic MOFs for highly sensitive acetone sensing at low operating temperatures. Inorg.Chem. Front. 2018, 5, 1107–1114. [Google Scholar] [CrossRef]
  59. Liu, H.; He, Y.; Nagashima, K.; Meng, G.; Dai, T.; Tong, B.; Deng, Z.; Wang, S.; Zhu, N.; Yanagida, T.; et al. Discrimination of VOCs molecules via extracting concealed features from a temperature-modulated p-type NiO sensor. Sens. Actuators B 2019, 293, 342–349. [Google Scholar] [CrossRef]
  60. Yuan, C.; Li, H.; Xie, L.; Wang, F.; Deng, H.; Chang, F.; Sun, Y. Flower-like NiO nanostructures synthesized by electrodeposition method for efficient detection of toluene gas. RSC Adv. 2015, 5, 92128–92133. [Google Scholar] [CrossRef]
  61. Chi, X.; Liu, C.; Liu, L.; Li, H.; He, Y.; Bo, X.; Liu, L. Toluene-sensing properties of In2O3 nanotubes synthesized by electrospinning. J. Semicond. 2014, 35, 064005. [Google Scholar] [CrossRef]
  62. Wang, X.; Chen, F.; Yang, M.; Guo, L.; Xie, N.; Kou, X.; Song, Y.; Wang, Q.; Sun, Y. Dispersed WO3 nanoparticles with porous nanostructure for ultrafast toluene sensing. Sens. Actuators B 2019, 289, 195–206. [Google Scholar] [CrossRef]
  63. Kang, Y.; Kim, K.; Cho, B.; Kwak, Y.; Kim, J. Highly Sensitive Detection of Benzene, Toluene, and Xylene Based on CoPP-Functionalized TiO2 Nanoparticles with Low Power Consumption. ACS Sens. 2020, 5, 754–763. [Google Scholar] [CrossRef] [PubMed]
  64. Wang, L.; Zhou, T.; Zhang, R.; Lou, Z.; Deng, J.; Zhang, T. Comparison of toluene sensing performances of zinc stannate with different morphology-based gas sensors. Sens. Actuators B 2016, 227, 448–455. [Google Scholar] [CrossRef]
  65. Hermawan, A.; Asakura, Y.; Inada, M.; Yin, S. One-step synthesis of micro-/mesoporous SnO2 spheres by solvothermal method for toluene gas sensor. Ceram. Int. 2019, 45, 15435–15444. [Google Scholar] [CrossRef]
  66. Zhang, Y.; Jia, C.; Wang, Q.; Kong, Q.; Chen, G.; Guan, H.; Dong, C. MOFs-derived porous NiFe2O4 nano-octahedrons with hollow interiors for an excellent toluene gas sensor. Nanomaterials 2019, 9, 1059. [Google Scholar] [CrossRef]
  67. Kim, H.-J.; Lee, J.-H. Highly sensitive and selective gas sensors using p-type oxidesemiconductors: Overview. Sens. Actuators B 2014, 192, 607–627. [Google Scholar] [CrossRef]
  68. Lyu, Y.; Li, C.; Du, X.; Zhu, Y.; Zhang, Y.; Li, S. Catalytic oxidation of toluene over MnO2 catalysts with different Mn (II) precursors and the study of reaction pathway. Fuel 2020, 262, 116610. [Google Scholar] [CrossRef]
  69. Zhou, T.; Zhang, R.; Wang, Y.; Zhang, T. MOF-derived 1 D α-Fe2O3/NiFe2O4 heterojunction as efficient sensing materials of acetone vapors. Sens. Actuators B 2019, 281, 885–892. [Google Scholar] [CrossRef]
Figure 1. XRD parttern of NiFe2O4 annealed at 350 °C, 400 °C, and 450 °C with JCPDS No.54-0964.
Figure 1. XRD parttern of NiFe2O4 annealed at 350 °C, 400 °C, and 450 °C with JCPDS No.54-0964.
Crystals 13 00865 g001
Figure 2. (a) UV–vis spectrum absorption spectra and band gap estimations of the tremella-like NiFe2O4 nanosheets for (b) S-350, (c) S-400, and (d) S-450.
Figure 2. (a) UV–vis spectrum absorption spectra and band gap estimations of the tremella-like NiFe2O4 nanosheets for (b) S-350, (c) S-400, and (d) S-450.
Crystals 13 00865 g002
Figure 3. SEM image of (a,b) precursor, and tremella-like NiFe2O4 annealed at (c,d) 350 °C (S-350), (e,f) 400 °C (S-400), and (g,h) 450 °C (S-450) in air.
Figure 3. SEM image of (a,b) precursor, and tremella-like NiFe2O4 annealed at (c,d) 350 °C (S-350), (e,f) 400 °C (S-400), and (g,h) 450 °C (S-450) in air.
Crystals 13 00865 g003
Figure 4. (a,b) TEM images of tremella-like S-400 NiFe2O4 nanosheets, (c) the EDS mapping of S-400 NiFe2O4, (d,e) high-magnification TEM, and (f) the HRTEM of S-400 nanosheets.
Figure 4. (a,b) TEM images of tremella-like S-400 NiFe2O4 nanosheets, (c) the EDS mapping of S-400 NiFe2O4, (d,e) high-magnification TEM, and (f) the HRTEM of S-400 nanosheets.
Crystals 13 00865 g004
Figure 5. Nitrogen adsorption–desorption isotherms and pore size distributions (inset) of tremella-like NiFe2O4 (S-400) nanosheets.
Figure 5. Nitrogen adsorption–desorption isotherms and pore size distributions (inset) of tremella-like NiFe2O4 (S-400) nanosheets.
Crystals 13 00865 g005
Figure 6. TGA-DSC curve of the formation of NiFe2O4 precursor.
Figure 6. TGA-DSC curve of the formation of NiFe2O4 precursor.
Crystals 13 00865 g006
Figure 7. (a) The full XPS survey for S-400 NiFe2O4 and the high-resolution XPS spectra of (b) Ni 2p for S-400 NiFe2O4, (c) Fe 2p for S-400 NiFe2O4, O 1s for (d) S-350 NiFe2O4, (e) S-400 NiFe2O4, and (f) S-450 NiFe2O4.
Figure 7. (a) The full XPS survey for S-400 NiFe2O4 and the high-resolution XPS spectra of (b) Ni 2p for S-400 NiFe2O4, (c) Fe 2p for S-400 NiFe2O4, O 1s for (d) S-350 NiFe2O4, (e) S-400 NiFe2O4, and (f) S-450 NiFe2O4.
Crystals 13 00865 g007
Figure 8. (a) Responses of the NiFe2O4 sensor at different operating temperature to 10 ppm and 30 ppm toluene based on the S-400 NiFe2O4 sensor; (b) responses of the S-350, S-400, and S-450 NiFe2O4 sensors toward 30 ppm toluene at 220 ℃; the dynamic response of the sensor toward (c) 50 ppb–700 ppb and (d) 1 ppm–50 ppm toluene based on S-400 NiFe2O4 sensor; (e) the relationship between response and toluene concentration (inset is the linear fit) based on S-400 NiFe2O4 sensor; (f) dynamic sensing transient of the NiFe2O4 to 10 ppm toluene based on S-400 NiFe2O4 sensor.
Figure 8. (a) Responses of the NiFe2O4 sensor at different operating temperature to 10 ppm and 30 ppm toluene based on the S-400 NiFe2O4 sensor; (b) responses of the S-350, S-400, and S-450 NiFe2O4 sensors toward 30 ppm toluene at 220 ℃; the dynamic response of the sensor toward (c) 50 ppb–700 ppb and (d) 1 ppm–50 ppm toluene based on S-400 NiFe2O4 sensor; (e) the relationship between response and toluene concentration (inset is the linear fit) based on S-400 NiFe2O4 sensor; (f) dynamic sensing transient of the NiFe2O4 to 10 ppm toluene based on S-400 NiFe2O4 sensor.
Crystals 13 00865 g008
Figure 9. (a) The relationship between the response–recovery time and the toluene concentration based on S-400 NiFe2O4 sensor, (b) the selectivity of the sensor to 10 ppm gas at 220 °C based on S-400 NiFe2O4 sensor, (c) the response of the NiFe2O4 sensor exposed into 10 ppm and 30 ppm toluene at 220 °C based on S-400 NiFe2O4 sensor, and (d) the long-term stability of the sensor to 10 ppm and 30 ppm toluene based on S-400 NiFe2O4 sensor.
Figure 9. (a) The relationship between the response–recovery time and the toluene concentration based on S-400 NiFe2O4 sensor, (b) the selectivity of the sensor to 10 ppm gas at 220 °C based on S-400 NiFe2O4 sensor, (c) the response of the NiFe2O4 sensor exposed into 10 ppm and 30 ppm toluene at 220 °C based on S-400 NiFe2O4 sensor, and (d) the long-term stability of the sensor to 10 ppm and 30 ppm toluene based on S-400 NiFe2O4 sensor.
Crystals 13 00865 g009
Figure 10. (a) The real resitance variation of NiFe2O4 sensor exposed to air and 1 ppm toluene atmosphere based on S-400 NiFe2O4 sensor; (b) the gas-sensing mechanism based on NiFe2O4 sensor for toluene detection.
Figure 10. (a) The real resitance variation of NiFe2O4 sensor exposed to air and 1 ppm toluene atmosphere based on S-400 NiFe2O4 sensor; (b) the gas-sensing mechanism based on NiFe2O4 sensor for toluene detection.
Crystals 13 00865 g010
Table 1. Comparision of band gaps of NiFe2O4 synthesized with different methods.
Table 1. Comparision of band gaps of NiFe2O4 synthesized with different methods.
MaterialsSynthesis MethodMorphologyAnnealing (°C)Band Gap
(eV)
Ref.
NiFe2O4Co-precipitationNanoparticles700 1.59 [31]
800 1.67
900 1.71
NiFe1.95Y0.05O4800 1.62
NiFe1.9Y0.1O41.59
NiFe1. 85Y0.15O41.59
NiFe1.8Y0.2O41.57
NiFe2O4Co-sputtering
(Ar:O2 = 40:60)
Nanoparticles 1.785 [32]
Co-sputtering
(Ar:O2 = 30:70)
1.78
Co-sputtering
(Ar:O2 = 20:80)
1.779
NiFe2O4Magnetron sputteringThin films 1.78 [33]
400 1.88
5002.46
7002.72
NiFe2O4Co-precipitationNanoparticles700 1.59 [34]
NiFe2O4CVDThin films 2.34
/2.78
[35]
NiFe2O4PLD Thin Films6902.36 [36]
NiFe2O4Hydrothermal synthesisNano-sheets350 1.73This work
400 1.70
450 1.72
Table 2. Comparision of specific surface areas and crystal sizes of NiFe2O4 synthesized with different synthesis methods and morphologies.
Table 2. Comparision of specific surface areas and crystal sizes of NiFe2O4 synthesized with different synthesis methods and morphologies.
MaterialsSynthesis MethodMorphologySurface Area(m2/g)Crystal Size (nm)Ref.
NiFe2O4MOFs templatesNanocubes33.72 [40]
NiFe2O4MOFs templatesHexagonal biyramids64.78 [41]
NiFe2O4MOFs templatesFusiformis61.50 [42]
NiFe2O4Carbon sphere templatesNanospheres86.916[43]
NiFe2O4Ultrasound assistanceNanoparticles17.17817[44]
NiFe2O4Co-precipitationNanoparticles5.32 [45]
NiFe2O4ElectrospinningNanofibers5420[46]
NiFe2O4HydrothermalMicrospheres46.44 46.4 [18]
NiFe2O4HydrothermalNanosheet77.8013.1This work
Table 3. Comparision of sensing performances to toluene based on typical sensing materials.
Table 3. Comparision of sensing performances to toluene based on typical sensing materials.
MaterialsMicrostructuresT (°C)ResponseLODτresrecov (s)Ref.
NiONanopaticle2501.5@300 ppm100 ppm-[59]
NiOFlower-like2501.73@0.5 ppm0.5 ppm13.6/24.5@2 ppm[60]
In2O3Nanotube34010.9@100 ppm1 ppm17/40@40 ppm[61]
WO3Nanopaticle225132 @100 ppm2 ppm2/6@100 ppm[62]
TiO2Nanopaticle3275.15@1 ppm5 ppb-[63]
Zn2SnO4Nanosheet2804.7@5 ppm5 ppm1/3.55@10 ppm[64]
SnO2Sphere40020.2@50 ppm-52/74@50 ppm[65]
NiFe2O4Nano-octahedron2606.41@100 ppm1 ppm25/40@100 ppm[66]
NiFe2O4Tremella-like2204.67@10 ppm50 ppb11.6/41.9@10 ppmThis work
T: temperature; LOD: limit of detection; Ref.: references.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Y.; Qu, H.; Gang, C.; Guan, H.; Dong, C.; Yin, Z. Porous, Tremella-like NiFe2O4 with Ultrathin Nanosheets for ppb-Level Toluene Detection. Crystals 2023, 13, 865. https://doi.org/10.3390/cryst13060865

AMA Style

Zhang Y, Qu H, Gang C, Guan H, Dong C, Yin Z. Porous, Tremella-like NiFe2O4 with Ultrathin Nanosheets for ppb-Level Toluene Detection. Crystals. 2023; 13(6):865. https://doi.org/10.3390/cryst13060865

Chicago/Turabian Style

Zhang, Yanlin, Honglong Qu, Cheng Gang, Hongtao Guan, Chengjun Dong, and Zongyou Yin. 2023. "Porous, Tremella-like NiFe2O4 with Ultrathin Nanosheets for ppb-Level Toluene Detection" Crystals 13, no. 6: 865. https://doi.org/10.3390/cryst13060865

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop