Next Article in Journal
IHIBE: A Hierarchical and Delegated Access Control Mechanism for IoT Environments
Previous Article in Journal
Cognitive Effort during Visuospatial Problem Solving in Physical Real World, on Computer Screen, and in Virtual Reality
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Ultrasensitive Room-Temperature H2 Sensor Based on a TiO2 Rutile–Anatase Homojunction

Ministry-of-Education Key Laboratory for the Green Preparation and Application of Functional Materials, Collaborative Innovation Center for Advanced Organic Chemical Materials Co-Constructed by the Province and Ministry, School of Materials Science & Engineering, Hubei University, Wuhan 430062, China
*
Author to whom correspondence should be addressed.
Sensors 2024, 24(3), 978; https://doi.org/10.3390/s24030978
Submission received: 22 December 2023 / Revised: 24 January 2024 / Accepted: 30 January 2024 / Published: 2 February 2024
(This article belongs to the Section Nanosensors)

Abstract

:
Metal oxide semiconductor hetero- and homojunctions are commonly constructed to improve the performance of hydrogen sensors at room temperature. In this study, a simple two-step hydrothermal method was employed to prepare TiO2 films with homojunctions of rutile and anatase phases (denoted as TiO2-R/A). Then, the microstructure of anatase-phase TiO2 was altered by controlling the amount of hydrochloric acid to realize a more favorable porous structure for charge transport and a larger surface area for contact with H2. The sensor used a Pt interdigital electrode. At an optimal HCl dosage (25 mL), anatase-phase TiO2 uniformly covered rutile-phase TiO2 nanorods, resulting in a greater response to H2 at 2500 ppm compared with that of a rutile TiO2 nanorod sensor by a factor of 1153. The response time was 21 s, mainly because the homojunction formed by the TiO2 rutile and anatase phases increased the synergistic effect of the charge transfer and potential barrier between the two phases, resulting in the formation of more superoxide (O2) free radicals on the surface. Furthermore, the porous structure increased the surface area for H2 adsorption. The TiO2-R/A-based sensor exhibited high selectivity, long-term stability, and a fast response. This study provides new insights into the design of commercially competitive hydrogen sensors.

1. Introduction

As a renewable source of energy, hydrogen is widely used in chemical and metallurgical industries owing to its relatively clean and pollution-free nature and its high calorific value. It can also be stored on a large scale [1,2,3,4]. However, gaseous hydrogen is prone to leaking and is extremely flammable when it contacts oxygen in the atmosphere. Moreover, its tendency to explode upon encountering open flame poses a grave threat to human life and infrastructure. Therefore, detecting hydrogen gas in real-time at low concentrations is crucial to ensure the safe usage of hydrogen as an energy source [5,6].
Among the various gas detection methods, gas chromatographs use chromatographic columns to separate the individual gas components in a mixture to identify each component. Mass spectrometers identify gas molecules on the basis of their characteristic deflections from a magnetic field. Optical sensors utilize the change of optical properties of certain materials when they interact with H2 to detect H2. These methods typically require complex instrumentation and relatively large, expensive, and high maintenance [7,8,9], and electrochemical sensors generally have poor selectivity and long-term stability and are subject to interference from reducing gases [10]. Therefore, hydrogen sensors based on chemical resistance have been widely explored owing to their low cost, good stability, and excellent application prospects [11]. Among the numerous sensing materials, metal oxide semiconductors (MOSs) (e.g., TiO2, SnO2, and ZnO) are popular owing to their low cost and good sensitivity [12,13]. Among these, TiO2, a typical n-type semiconductor, is non-toxic, harmless, and inexpensive. Moreover, its unique performance in gas sensing is well-documented [14,15]. However, the response of TiO2 to hydrogen is limited by the rapid recombination of electron–hole pairs. Therefore, to improve the response of TiO2 to hydrogen, it is generally necessary to regulate its band structure and suppress the recombination rate of electrons and holes to ensure that more electrons combine with O2 to generate O2. One effective method to suppress the recombination rate is to construct heterojunctions of TiO2 such as TiO2/MoS2 [16], Bi2MoO6/TiO2 [17], TiO2/Cu2O [18], and TiO2/SrTiO3 [19]. Owing to the advantages of the two types of metal oxide semiconductors, heterojunction composites exhibit significantly different physical and chemical properties and sensitivity to gases than their corresponding single-component counterparts. However, heterojunction composites contain many defects, which are not conducive to the performance of the heterojunction in terms of gas sensitivity [20,21].
A homojunction is formed between different crystal phases of the same semiconductor with different band structures. When carriers are transferred across a homojunction, their recombination rate would be much smaller than that at the heterojunction interface [22,23]. TiO2 has two stable phases, viz., rutile and anatase, with bandgaps of approximately 3.0 and 3.2 eV, respectively [21,24,25]. Therefore, in this study, we prepared composite materials with rutile–anatase TiO2 homojunctions for gas sensing. Biphasic TiO2 with homojunctions has a uniform composition and near-perfect lattice matching, which can ensure a reduced contact barrier, the regulation of the band structure, and effective charge transfer at the interface of the two phases (homojunction) [26,27].
Table 1 compares the performance of some previously reported MOS H2 sensors. It is seen that most MOS H2 sensors need to operate at high temperatures (>150 °C) [28,29,30,31,32,33]. Some reports indicate that modifying MOS sensors with precious metals can lower the operating temperature to room temperature [30]. However, there are few reports on MOS sensors that have a high response, a low detection limit, and a fast response time and operate at room temperature simultaneously. In this work, we designed and developed a growth-oriented material with rutile–anatase TiO2 homojunctions (denoted as TiO2-R/A) using a simple two-step hydrothermal method for H2 detection. In the second step of the synthesis, we controlled the amount of hydrochloric acid (HCl) to adjust the morphology of anatase-phase TiO2, promote charge transfer across the junction [34], and increase the contact area of the material surface with H2. Our results showed that the response of the bilayered mixed-phase TiO2 films with homojunctions to picomolar hydrogen was much greater than that of single-layer rutile-phase TiO2. The best performance in H2 sensing was achieved when anatase TiO2 was uniformly formed on the surface of the rutile TiO2 nanorods, at a hydrogen concentration of 2500 ppm, the response reached 1661. The characterization of the mixed-phase TiO2 material using X-ray diffraction (XRD), Raman spectroscopy, scanning electron microscopy (SEM), potentiometry, and X-ray photoelectron spectroscopy (XPS) is indicated, and the sensing mechanism is analyzed.

2. Materials and Methods

2.1. Fabrication of Sensing Layers

2.1.1. Preparation of Rutile-Phase TiO2 Nanorod Arrays

First, a self-assembled rutile TiO2 nanorod array was grown on an FTO substrate, as reported previously [35]. Briefly, 28 mL of deionized water, 2 mL of ethanol, 30 mL of hydrochloric acid, and 1 mL of tetrabutyl titanate (TBOT) were mixed in a beaker and stirred to obtain a homogeneous precursor solution. Then, the precursor solution and FTO substrates, which were cleaned sequentially with acetone, ethanol, and deionized water, were placed in a 200 mL reactor and reacted at 150 °C for 8 h. After the reaction, the reactor was cooled to room temperature, and the FTO substrate was soaked in deionized water for 3 h. The rutile-phase TiO2 nanorod array was annealed in a 400 °C tube furnace for 20 min, and the prepared sample is referred to as NR-TiO2.

2.1.2. Preparation of Rutile- and Anatase-Phase TiO2 Homojunctions

A known volume of hydrochloric acid (X = 10, 15, 20, 25, or 30 mL), 1 mL of concentrated sulfuric acid, and 1 mL of TBOT were added to a solution of ethanol (2 mL) and deionized water (28 mL) in a beaker and mixed evenly by stirring. Then, the NR-TiO2 sample was placed in a 200 mL reactor and reacted with the acidified TBOT solution at 150 °C for 8 h. After the reaction, the sample was rinsed several times with deionized water and then soaked in deionized water for 3 h. Finally, the sample was annealed in a tube furnace at 400 °C for 20 min. The samples thus prepared were denoted as TiO2(R/A-X), where R and A represent rutile and anatase phases and X represents the volume of HCl (10, 15, 20, 25, or 30 mL).

2.2. Characterization

XRD (D8-Advance, Bruker, Cu line as the X-ray source), field-emission SEM (Sigma 500, Zeiss, Oberkochen, Germany), Raman spectrometry (CLY19 Lab RAM HR Evolution), and XPS were performed to characterize the microstructure, morphology, and chemical composition of the samples.
Platinum interdigitated electrodes were deposited onto thin films of TiO2(R/A-X) through DC magnetron sputtering. The standard test gas, 75% of H2 mixed with 25% of Ar or 0.75% of NH3, SO2, NO, C2H6, or NO2 dry gas mixed with 99.25% of Ar (Wuhan Xiangyun Industry and Trade Co., Ltd., Wuhan, China) was introduced into a 15 L sealed testing chamber via a computer-controlled mass flow controller (D07-7C, Beijing Seven Star Huachuang Flow Co., Ltd., Beijing, China). The sensing evaluation of the target gas was conducted at ambient temperature (25 °C) and environmental humidity (44%) unless specified otherwise. A Keithley 2400 multimeter was used to measure the fluctuation in sensor resistance while maintaining a constant voltage of 1 V across the sensor. Equation (1) can be used to ascertain the response of the sensor to H2 (S).
S = Rair/Rgas,
In Formulas (1) and (2), the initial resistance of the sensor when exposed to air is denoted as Rair, while the initial resistance in a H2 atmosphere is denoted as Rgas. ΔR represents the recorded difference in resistance between the sensor in air and in an H2 atmosphere. The response time refers to the duration required for the sensor’s resistance to a decrease by 90% of ΔR from Rair, whereas recovery time refers to the duration required for the sensor’s resistance to increase by 90% of ΔR from Rgas.
R = RairRgas

3. Results and Discussion

3.1. Characterization of the TiO2(R/A-X) Samples

Figure 1a presents the XRD patterns of the FTO substrate and the NR-TiO2 and TiO2(R/A-X) samples grown on the FTO substrates. The main sample peaks are consistent with those of rutile and anatase TiO2 (JCPDS No. 21-1276 and No. 21-1272). The diffraction peaks observed at 2θ = 36.1° and 62.7° correspond to the (101) and (002) crystallographic planes of rutile TiO2 (JCPDS 21-1276), and the peak at 2θ = 37.82° occurred due to the (004) plane of the anatase phase. A comparison with the standard patterns showed that the strongest peaks of the rutile and anatase phases did not appear in the XRD patterns of the prepared materials. The anatase (004) peak replaced the (101) peak as the strongest diffraction peak of the film. This indicates a tendency for the preferential growth of the nanorod in the (001) direction. The peak of 2θ = 62.7° in Figure 1a corresponds to the (002) crystal plane of rutile TiO2, which is consistent with previously reported results for rutile nanorod arrays, which tend to grow in the direction of (001) [35]. Therefore, the nanorods with rutile and anatase phases in the synthesized TiO2(A/R-X) thin films tended to grow in the (001) direction. As the main peak of the FTO substrate (PDF# 46-1088) at 2θ = 37.76° is very close to the (004) peak of the anatase phase, this region of Figure 1a was magnified. The locally magnified pattern revealed that the peaks observed near 2θ = 37.76° for the NR-TiO2 sample arose owing to the underlying FTO substrate. A comparison of the XRD patterns revealed that the diffraction peak of rutile (101) became progressively weaker with a gradual reduction in the amount of hydrochloric acid used in the second step of the synthesis. Further, the (004) diffraction peak of the anatase phase gradually increased, which was accompanied by a proportional decrease in the peak intensity of the rutile phase. This occurred because the inhibitory effect of Cl on anatase-phase TiO2 was weakened with the decrease in the amount of hydrochloric acid, which is consistent with previous reports [36,37]. The above phenomena were further confirmed by Raman spectroscopy.
Figure 1b shows the Raman spectra of the TiO2(R/A-X) samples. This is consistent with previous reports [38], which show a Raman band at 144 cm−1 corresponding to the Eg vibration mode of the anatase phase. The peaks at 448 and 608 cm−1 correspond to the Eg and A1g vibrational modes of the rutile phase of TiO2. A comparison of the spectra revealed that as the amount of hydrochloric acid used in the second step increased, the main peak of the anatase-phase TiO2 (144 cm−1) increased gradually. Notably, the Raman spectrum of the TiO2(R/A-30 mL) sample did not show the anatase peak at 144 cm−1 because the content of the anatase-phase TiO2 in this sample was too low to be detected. Thus, the XRD and Raman spectroscopy results confirm the successful preparation of gas-sensitive films composed of anatase-phase TiO2 supported on the rutile-phase TiO2 nanorod array using a two-step hydrothermal method.
Figure 2 shows the surface morphologies of the NR-TiO2 and TiO2(R/A-X) samples. The surface of the NR-TiO2 sample shows the cauliflower-like morphology of the nanorod tips (Figure 2a) with diameters of approximately 100–200 nm. In this study, the formation of anatase-phase TiO2 on NR-TiO2 was controlled by varying the ratio of Cl to SO42−. Figure 2 shows that as the amount of hydrochloric acid was decreased, a larger amount of anatase-phase TiO2 was generated in the reaction. When the amount of hydrochloric acid was decreased from X = 30 mL to X = 25 mL, the gaps between the nanorods became smaller. When the amount of hydrochloric acid was decreased further, below X = 25 mL, the gaps between the nanorods were gradually filled with anatase-phase TiO2.
Figure 3 shows the cross-sectional morphologies of the NR-TiO2 and TiO2(R/A-X) samples. The cross-sectional SEM image of the NR-TiO2 sample exhibited smooth nanorods (Figure 3a) with a height of approximately 2.3 μm. Figure 3b–f show the cross-sectional images of the TiO2(R/A-X) samples, which reveal that the morphology of the nanorods changed owing to modification with anatase-phase TiO2. As the amount of hydrochloric acid was decreased from X = 30 mL to X = 25 mL, the spaces between the nanorods gradually decreased. When X = 20 mL, anatase-phase TiO2 was enriched at the top of the nanorod, resulting in a distinct double-layer structure. With the reduction in the amount of hydrochloric acid, the anatase-phase-TiO2-enriched layer gradually became thicker, exhibiting an increase in thickness from 1.1 to 2.63 μm, as shown in Figure 3d–f. To compare the effects of the hydrochloric acid dosage (X = 25 mL and X = 30 mL) on anatase-phase TiO2 more clearly, we compared the micromorphologies of the NR-TiO2, TiO2(R/A-25 mL), and TiO2(R/A-30 mL) samples at a higher magnification (50 kx) (Figure 3g–i). The surface of the nanorods became rough due to modification with anatase-phase TiO2, and the surface roughness of the nanorods gradually increased with the decrease in hydrochloric acid dosage. Thus, SEM analyses confirmed that hydrochloric acid could significantly inhibit the growth of anatase-phase TiO2, which is consistent with the XRD and Raman results shown in Figure 1.

3.2. Sensing Properties

The hydrogen-sensing curves of NR-TiO2 and TiO2(R/A-X) samples are shown in Figure 4a–f. All samples exhibited a consistent n-type response to hydrogen. Owing to the reducing nature of hydrogen, the resistance of the sensor decreased when hydrogen entered the ventilation chamber and increased to the initial value when hydrogen exited the ventilation chamber. As shown in Figure 4a, the response of NR-TiO2 to hydrogen at room temperature was relatively poor, and the response values at 2500 and 20,000 ppm were 1.44 and 2.7, respectively (Figure S1). As shown in Figure 4b–f, the resistance of the TiO2(R/A-X) sample first increased and then decreased with the reduction in the amount of hydrochloric acid (X) used in the second step of the synthesis. Notably, in the sample prepared with X = 30 mL, only a very small number of anatase particles adhered to the sides of the rutile nanorods because the excessive chloride ions inhibited the growth of the anatase phase. In this case, the homojunction area formed between the rutile and anatase phases was relatively small, and the gap between the nanorods was large, resulting in relatively low resistance.
The response values of the NR-TiO2 and TiO2(R/A-X) samples during hydrogen sensing are shown in Figure 5a. TiO2(R/A-X) samples with homojunctions exhibited significantly higher response values than NR-TiO2. In particular, the highest response (1661) of 2500 ppm hydrogen was observed for the TiO2(R/A-25 mL) sample prepared using 25 mL of HCl. Further, with a further increase in hydrogen concentration, the response value of this sensor continued to show an increasing trend. At 20,000 ppm hydrogen, the response value of the TiO2(R/A-25 mL) sensor reached 76,702, which is 28,408 times that of NR-TiO2. It is worth noting that for the sample prepared with X = 30 mL, the response to hydrogen was only slightly improved compared with that of NR-TiO2 because only a very small amount of anatase TiO2 was formed on the rutile nanorod surface. According to the results shown in Figure 5, when the dosage of hydrochloric acid did not exceed 25 mL, the TiO2(R/A-X) sample showed a highly sensitive response to hydrogen at room temperature. In order to further evaluate the detection limit (LOD) of TiO2(R/A-25 mL), the response curve at a low concentration of 25–125 ppm H2 was measured, and the LOD was theoretically evaluated using Equation (3) [39,40], as shown in the Figure 5b.
LOD = 3 R n o i s e L s l o p e
where R n o i s e is the measured noise of the sensor and L s l o p e is the slope of the linear fitting curve. The R n o i s e as 0.01 is calculated using the equation [39,40] below:
R n o i s e = Σ ( R i R ) 2 N
where R i is the experimental data (i.e., various responses of H2 concentration) and R is the fitting value based on Figure 5b. Calculation reveals that the LOD of the TiO2(R/A-25 mL) sensor is ~6.3 ppm.
As shown in Figure 5c,d, the TiO2(R/A-X) samples exhibited a slightly longer response time and a longer recovery time than NR-TiO2 (see Figure S2). This was the case because the increased response causes the resistance to fluctuate over a wider range, and the resistance of the system takes more time to reach the equilibrium value, especially in environments with high hydrogen concentrations, where this phenomenon is much more evident. In summary, the response values of NR-TiO2 at all tested hydrogen concentrations ranged from 1 to 2, whereas the response values of the TiO2(R/A-X) samples at the same tested hydrogen concentrations were significantly better.
After the surface of the rutile TiO2 nanorods was coated with anatase TiO2, TiO2(R/A-25 mL) exhibited exceptional sensing properties. It is crucial to comprehend its key determinants. The sensing performance of this sample can be attributed to the presence of type II alternating band alignment of ~0.4 eV between anatase and rutile phases and the higher electron affinity of anatase-phase TiO2. Therefore, electron flow occurs from the rutile phase to the surface anatase phase, as shown in Figure 6a,b. Because the rutile and anatase phases of TiO2 have different band gaps (Eg) and different valence band and conductive band energies (EV and EC, respectively), when the rutile-phase TiO2 nanorod contacts the anatase-phase TiO2 layer, an internal potential is established at the junction of the two phases. Owing to the continuous flow of electrons from the rutile to the anatase phase of TiO2, an electron depletion layer is formed at the rutile phase TiO2 nanorods, whereas an electron accumulation layer is formed at the anatase layer, as illustrated in Figure 6b.
To investigate this, XPS analyses were conducted on samples of NRs-TiO2, anatase phase TiO2, and TiO2(R/A-X), as depicted in Figure 6c,d. The O1s spectrum of the NRs-TiO2 sample (Figure 6c) exhibited two prominent peaks at 529.71 eV and 530.91 eV, corresponding to lattice oxygen (Ti-O-Ti) and hydroxide (Ti-OH), respectively. In the Ti2p spectrum (Figure 6d), two distinct nuclear level signals may be observed at 464.30 eV and 458.54 eV, which can be attributed to the Ti2p3/2 and Ti2p1/2 levels of Ti4+, respectively.
Consistent with previously reported findings [41], the O1s spectra of anatase-phase TiO2 exhibited higher binding energy peaks for lattice oxygen (Ti-O-Ti) and hydroxyl groups (Ti-OH). For both the sensor TiO2(R/A-10 mL) and anatase-phase TiO2 samples, no significant change in peak values was observed; this can be explained by the aforementioned microscopic morphology, in which a rutile-phase nanorod was present on the top surface covered with anatase-phase TiO2 with a thickness of approximately 2.6 μm. Consequently, the XPS analysis only detected signals from the anatase phase of TiO2.
In addition, the formation of a homojunction between rutile phase TiO2 and anatase-phase TiO2 in the TiO2(R/A-25 mL) sample results in negative shifts in the binding energies of O1s (−0.27 eV), Ti2p3/2 (−0.22 eV), and Ti2p1/2 (−0.24 eV) compared to pure anatase-phase TiO2. This peak shift further confirms electron flow from rutile-phase TiO2 to anatase-phase TiO2.
To explain the response mechanism of the sensor to H2 more clearly, one usually needs to calculate the Schottky barrier formed at the two-phase contact surface. Therefore, we tested the I-V change curve of TiO2(R/A-X) at room temperature to estimate the difference in barrier height between in air (Figure 7a) and H2 (2500 ppm) atmosphere (Figure 7b). The IV curves of the entire series of TiO2(R/A-X) samples changed significantly before and after exposure to H2 atmosphere. By observing the phenomenon of the nonlinear relationship between I-V, we can consider that there was a potential barrier [42]. The effective height of the barrier can be calculated by Equation (5) [43,44].
Φ B = k T q   ln ( A I s A * T 2 ) ,
where k represents the Boltzmann constant, T represents the test temperature, A represents the area where the diode is constructed, Is represents the saturation current, and A* represents the Richardson constant of rutile-phase TiO2. In this series of sensors, the value of A was equal to 2.5 cm × 2.5 cm = 6.25 cm2, T at room temperature usually takes a value of 300 K, and A* = (m*/m) A/cm2 K, for rutile-phase TiO2 m*/m = 20. The saturation current can be calculated by the following formulas (6) and (7):
I = I s   exp ( q V n k T ) ,
ln I = ln I s + q V n k T
In the formula given above, q represents the amount of charge carried by the electron and n represents the ideal factor. There are two main factors that affect the height of the effective barrier in this series of sensors. First, H2 reacts with the topmost Pt electrode to form PtHx, which decreases the height of the Schottky barrier [44,45,46]. Second, the change in the charge concentration in the anatase-phase TiO2 layer also changes the barrier height. Therefore, the potential barrier of the TiO2(R/A-X) sample in air showed a trend of first increasing and then decreasing with increasing anatase TiO2 content. The highest potential barrier was observed for TiO2(R/A-25 mL). The increase in the barrier heights of TiO2(R/A-30 mL) and TiO2(R/A-25 mL) was due to the gradual increase in the generation of anatase TiO2 and the resultant gradual increase in the interface area between the rutile and anatase phases. As the generation of anatase TiO2 was increased further by reducing the amount of HCl, lateral current leakage occurred due to the accumulation of anatase TiO2 on the surface of the rutile-phase TiO2 nanorods, which decreased the barrier height.
Figure 7c,d show the calculated effective barrier heights of different samples in air and 2500 ppm H2 environments and the correlation between the barrier heights of the samples and the corresponding response of the sensor. Under 2500 ppm H2, the barrier heights of all samples were less than those in air, and the barrier difference was consistent with the curve of the sensing response. The largest change in the barrier height of 0.30 eV was observed for the TiO2(R/A-25 mL) sample, which exhibited the highest response to H2.
Figure 8a,b show the changes in the morphology of the anatase-phase TiO2 layer with the reduction in the amount of hydrochloric acid used in the synthesis. As the amount of hydrochloric acid was gradually decreased, anatase-phase TiO2 was first coated on the surface of rutile nanorods (Figure 8a), As the amount of hydrochloric acid continued to decrease, the deposition of anatase-phase TiO2 on the surface of the rutile nanorods gradually increased, resulting in a double-layer structure (Figure 8b). In the schematic shown in Figure 8a, anatase-phase TiO2 completely covers the surface of rutile-phase TiO2 nanorods to form a porous structure. In this situation, only the longitudinal current I1 along the nanorods is present (Figure 8a). As the amount of hydrochloric acid was decreased, anatase-phase TiO2 generated by the reaction was not only deposited on the nanorod surface but also filled the gaps between the nanorods. This structure not only has longitudinal current I1 along the nanorods but also generates transverse current I2 (Figure 8b). This model explains the initial increase followed by the decrease in the resistance of the TiO2(R/A-X) samples, as shown in Figure 4. The schematic in Figure 8c,d shows the response mechanism of the sensor to H2. Electrons flow from rutile-phase TiO2 nanorods to the surface anatase-phase TiO2 layer and react with O2 adsorbed to the surface to generate a large amount of O2 species in air. At this stage, a large depletion layer is formed on the sides of the rutile-phase nanorods (Figure 8c), and the resistance of the system increases. When H2 enters the test chamber, H2 reacts with O2 on the surface of the sensor so that the released electrons return to the rutile-phase TiO2 nanorods, the depletion layer becomes smaller, and the resistance of the system decreases (Figure 8d). The associated reaction can be represented as follows [44,45,46].
O2 + e→O2,
2H2 + O2→2H2O + e.
According to the micromorphology of the TiO2(R/A-25 mL) sample in Figure 2 and Figure 3, rutile-phase TiO2 nanorods are tightly wrapped by anatase-phase TiO2, which has a porous structure that is more conducive to contact with H2.
The selectivity of the TiO2(R/A-25 mL) sample was tested at 2500 ppm concentrations of H2, NO, SO2, C2H6, NH3, and NO2. The dynamic curve of the resistance is shown in Figure 9a, and the response values of the sensor corresponding to different gases are shown in Figure 9b. The sensor exhibited an N-type response to reducing gases, and TiO2, as a typical N-type semiconductor, exhibited this phenomenon in line with the oxygen adsorption theory [44]. However, NO2, an oxidizing gas, also exhibited an N-type response, which contradicts the theory of oxygen adsorption. This phenomenon was a result of the catalytic action of the Pt electrode. The highly oxidizing NO2 gas reacts preferentially with Pt, and NO2 is broken down into NOx and O, thereby transferring electrons to the surface of TiO2 and reducing the resistance of the film. When the sensor was exposed to NH3 and C2H6 atmospheres, the initial resistance of the sensor did not change, reflecting that the sensor was not responsive to these two gases. Because these two gas molecules show high stability at room temperature, they are less likely to break the chemical bond and cause a change in the resistance of the TiO2 [45]. As shown in Figure 9b, at the same concentration of 2500 ppm, the response of the TiO2(R/A-25 mL) sensor to hydrogen was 109 times greater than that of NO, 230 times more than that of SO2, and 503 times greater than that of NO2, indicating that the TiO2(R/A-25 mL) sensor has excellent selectivity to H2.
Figure 9c shows the results of the cyclic testing of the TiO2(R/A-25 mL) sample with 25, 250, 1250, 2500, and 5000 ppm of H2. The samples were exposed to room-temperature air for 170 d and tested in four cycles. The response value was then compared with the initial value to calculate the attenuation in the response (Figure 9d). As shown in Figure 9d, when the hydrogen concentrations were 25, 250, 1250, 2500, and 5000 ppm, the decay rates of the homojunction were 7, 8.5, 11, 11.4, and 20%, respectively, indicating that the homojunction is very stable at room temperature. Figure 9e shows the response of the TiO2(R/A-25 mL) sensor to H2 under different humidities of 44, 62, and 82%, respectively. As the relative humidity was increased gradually from 44 to 82%, the initial resistance of the sample decreased gradually. Figure 9f presents the response value of the sample under different humidities. As the humidity increased, the response value of the sensor to hydrogen decreased significantly because of the competitive adsorption of water molecules on the TiO2 film surface, which attenuates the H2-sensing performance. Due to the increase in humidity, water molecules were adsorbed on the surface of the sensor, resulting in reduced oxygen adsorption at the active site on the surface of the TiO2. As a result, the reaction between oxygen and H2 molecules weakened rapidly, and the drop in the surface resistance of TiO2 was reduced. Under conditions of 44 and 62% humidity, the sensor response was still very good, reflecting the good moisture resistance of the sample, which essentially meets the requirement of hydrogen detection in daily life.

4. Conclusions

A TiO2 rutile–anatase dual-phase homojunction material was prepared for hydrogen sensing. The key sensing layers with different TiO2 phases were prepared by the in situ modification of rutile TiO2 nanorod films produced by a hydrothermal method. In the second step of the hydrothermal reaction, the thickness and morphology of the second anatase phase were controlled by changing the volume of HCl used to control the growth of this phase. The relationship between the barrier height and microstructure of the TiO2 rutile–anatase dual-phase material was investigated through various measurements. The experimental results show that the interface of the TiO2 nanocrystalline rutile phase and anatase phase forms a uniform homojunction, and a porous structure more conducive to carrier transport is formed, thus enhancing the response of the sensor to hydrogen. The TiO2 rutile–anatase heterojunction plays a crucial role in regulating the barrier height under H2 conditions. Among the different samples, the TiO2(R/A-25 mL) sample exhibited the best response to 2500 ppm hydrogen (as high as 1661) with good long-term stability and selectivity. The results of this study provide novel insights to support the design of commercially competitive hydrogen sensors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/s24030978/s1, Figure S1: Response values of NRs-TiO2 at different concentrations; Figure S2: Response time (a) and recovery time (b) of NRs-TiO2 at different concentrations.

Author Contributions

X.W.: Carried out laboratory research on TiO2/TiO2 homojunction film fabrication, Pt sputtering, and H2 sensing performance of the samples and wrote the draft of the manuscript. Y.Z.: Carried out laboratory research on the humidity and selectivity of the samples. M.Z.: Carried out the stability measurement of the samples. J.L.: Optimized the electrode fabrication of the homojunction devices. Y.B.: Carried out data processing and supervised the device fabrication. X.X.: Offered ideas for experimental details. K.H.: Offered ideas for experimental details and valuable discussions about paper organization. M.L.: Carried out electrode fabrication and device measurement. Y.G.: Offered ideas for optimizing the process of the interface regulation of the thin films and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

National Natural Science Foundation of China (Nos: 12174092 and U21A20500), Program for Key Research and Development of Science and Technology in Hubei Province (grant No. 2023BEB002), and Overseas Expertise Introduction Center for Discipline Innovation (D18025).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author. The data are not publicly available due to privacy and ethical restrictions.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Duan, P.; Duan, Q.; Peng, Q.; Jin, K.; Sun, J. Design of ultrasensitive gas sensor based on self-assembled Pd-SnO2/rGO porous ternary nanocomposites for ppb-level hydrogen. Sens. Actuators B Chem. 2022, 369, 132280. [Google Scholar] [CrossRef]
  2. Hübert, T.; Boon-Brett, L.; Black, G.; Banach, U. Hydrogen sensors—A review. Sens. Actuators B Chem. 2011, 157, 329–352. [Google Scholar] [CrossRef]
  3. Kovač, A.; Paranos, M.; Marciuš, D. Hydrogen in energy transition: A review. Int. J. Hydrogen Energy 2021, 46, 10016–10035. [Google Scholar] [CrossRef]
  4. Meng, X.; Bi, M.; Xiao, Q.; Gao, W. Ultra-fast response and highly selectivity hydrogen gas sensor based on Pd/SnO2 nanoparticles. Int. J. Hydrogen Energy 2022, 47, 3157–3169. [Google Scholar] [CrossRef]
  5. Kondalkar, V.V.; Park, J.; Lee, K. MEMS hydrogen gas sensor for in-situ monitoring of hydrogen gas in transformer oil. Sens. Actuators B Chem. 2021, 326, 128989. [Google Scholar] [CrossRef]
  6. Miller, D.R.; Akbar, S.A.; Morris, P.A. Nanoscale metal oxide-based heterojunctions for gas sensing: A review. Sens. Actuators B Chem. 2014, 204, 250–272. [Google Scholar] [CrossRef]
  7. Lee, J.H.; Park, M.S.; Jung, H.; Choe, Y.-S.; Kim, W.; Song, Y.G.; Kang, C.-Y.; Lee, H.-S.; Lee, W. Selective C2H2 detection with high sensitivity using SnO2 nanorod based gas sensors integrated with a gas chromatography. Sens. Actuators B Chem. 2020, 307, 127598. [Google Scholar] [CrossRef]
  8. Imonigie, J.A.; Walters, R.N.; Gribb, M.M. Rapid Isothermal Gas Chromatography-Mass Spectrometry Method for Validating a Small Ion Mobility Spectrometer Sensor. Instrum. Sci. Technol. 2006, 34, 677–695. [Google Scholar] [CrossRef]
  9. Wang, X.-d.; Wolfbeis, O.S. Fiber-Optic Chemical Sensors and Biosensors (2015–2019). Anal. Chem. 2019, 92, 397–430. [Google Scholar] [CrossRef]
  10. Bakker, E. Electrochemical Sensors. Anal. Chem. 2004, 76, 3285–3298. [Google Scholar] [CrossRef]
  11. Tonezzer, M.; Dang, L.T.T.; Tran, H.Q.; Iannotta, S. Multiselective visual gas sensor using nickel oxide nanowires as chemiresistor. Sens. Actuators B Chem. 2018, 255, 2785–2793. [Google Scholar] [CrossRef]
  12. Bao, Y.; Wei, P.; Xia, X.; Huang, Z.; Homewood, K.; Gao, Y. Remarkably enhanced H2 response and detection range in Nb doped rutile/anatase heterophase junction TiO2 thin film hydrogen sensors. Sens. Actuators B Chem. 2019, 301, 127143. [Google Scholar] [CrossRef]
  13. Xia, X.; Wu, W.; Wang, Z.; Bao, Y.; Huang, Z.; Gao, Y. A hydrogen sensor based on orientation aligned TiO2 thin films with low concentration detecting limit and short response time. Sens. Actuators B Chem. 2016, 234, 192–200. [Google Scholar] [CrossRef]
  14. Li, Z.; Haidry, A.A.; Gao, B.; Wang, T.; Yao, Z. The effect of Co-doping on the humidity sensing properties of ordered mesoporous TiO2. Appl. Surf. Sci. 2017, 412, 638–647. [Google Scholar] [CrossRef]
  15. Kwon, H.; Lee, Y.; Hwang, S.; Kim, J.K. Highly-sensitive H2 sensor operating at room temperature using Pt/TiO2 nanoscale Schottky contacts. Sens. Actuators B Chem. 2017, 241, 985–992. [Google Scholar] [CrossRef]
  16. Han, J.; Zhang, S.; Song, Q.; Yan, H.; Kang, J.; Guo, Y.; Liu, Z. The synergistic effect with S-vacancies and built-in electric field on a TiO2/MoS2 photoanode for enhanced photoelectrochemical performance. Sustain. Energy Fuels 2021, 5, 509–517. [Google Scholar] [CrossRef]
  17. Wang, L.; Wang, R.; Zhou, Y.; Shen, Q.; Ye, J.; Wu, C.; Zou, Z. Three-dimensional Bi2MoO6/TiO2 array heterojunction photoanode modified with cobalt phosphate cocatalyst for high-efficient photoelectrochemical water oxidation. Catal. Today 2019, 335, 262–268. [Google Scholar] [CrossRef]
  18. Zhang, S.; Liu, Z.; Yan, W.; Guo, Z.; Ruan, M. Decorating non-noble metal plasmonic Al on a TiO2/Cu2O photoanode to boost performance in photoelectrochemical water splitting. Chin. J. Catal. 2020, 41, 1884–1893. [Google Scholar] [CrossRef]
  19. He, Y.; Wang, P.; Zhu, J.; Yang, Y.; Liu, Y.; Chen, M.; Cao, D.; Yan, X. Synergistical Dual Strategies Based on in Situ-Converted Heterojunction and Reduction-Induced Surface Oxygen Vacancy for Enhanced Photoelectrochemical Performance of TiO2. ACS Appl. Mater. Interfaces 2019, 11, 37322–37329. [Google Scholar] [CrossRef]
  20. Huang, X.; Zhang, R.; Gao, X.; Yu, B.; Gao, Y.; Han, Z.-G. TiO2-rutile/anatase homojunction with enhanced charge separation for photoelectrochemical water splitting. Int. J. Hydrogen Energy 2021, 46, 26358–26366. [Google Scholar] [CrossRef]
  21. Tian, J.; Cao, G. Control of Nanostructures and Interfaces of Metal Oxide Semiconductors for Quantum-Dots-Sensitized Solar Cells. J. Phys. Chem. Lett. 2015, 6, 1859–1869. [Google Scholar] [CrossRef]
  22. Guo, Z.; Liu, Z. Synthesis and control strategies of nanomaterials for photoelectrochemical water splitting. Dalton Trans. 2021, 50, 1983–1989. [Google Scholar] [CrossRef]
  23. Li, Y.; Liu, Z.; Li, J.; Ruan, M.; Guo, Z. An effective strategy of constructing a multi-junction structure by integrating a heterojunction and a homojunction to promote the charge separation and transfer efficiency of WO3. J. Mater. Chem. A 2020, 8, 6256–6267. [Google Scholar] [CrossRef]
  24. Scanlon, D.O.; Dunnill, C.W.; Buckeridge, J.; Shevlin, S.A.; Logsdail, A.J.; Woodley, S.M.; Catlow, C.R.A.; Powell, M.J.; Palgrave, R.G.; Parkin, I.P.; et al. Band alignment of rutile and anatase TiO2. Nat. Mater. 2013, 12, 798–801. [Google Scholar] [CrossRef]
  25. Xu, J.; Chen, J.; Dong, Z.; Meyuhas, O.; Chen, J.-K. Phosphorylation of ribosomal protein S6 mediates compensatory renal hypertrophy. Kidney Int. 2015, 87, 543–556. [Google Scholar] [CrossRef]
  26. Wang, F.; Pei, K.; Li, Y.; Li, H.; Zhai, T. 2D Homojunctions for Electronics and Optoelectronics. Adv. Mater. 2021, 33, 2005303. [Google Scholar] [CrossRef]
  27. Deák, P.; Aradi, B.; Frauenheim, T. Band Lineup and Charge Carrier Separation in Mixed Rutile-Anatase Systems. J. Phys. Chem. C 2011, 115, 3443–3446. [Google Scholar] [CrossRef]
  28. Mondal, B.; Basumatari, B.; Das, J.; Roychaudhury, C.; Saha, H.; Mukherjee, N. ZnO–SnO2 based composite type gas sensor for selective hydrogen sensing. Sens. Actuators B Chem. 2014, 194, 389–396. [Google Scholar] [CrossRef]
  29. Yin, X.-T.; Li, J.; Dastan, D.; Zhou, W.-D.; Garmestani, H.; Alamgir, F.M. Ultra-high selectivity of H2 over CO with a p-n nanojunction based gas sensors and its mechanism. Sens. Actuators B Chem. 2020, 319, 128330. [Google Scholar] [CrossRef]
  30. Wu, C.-H.; Zhu, Z.; Chang, H.-M.; Jiang, Z.-X.; Hsieh, C.-Y.; Wu, R.-J. Pt@NiO core–shell nanostructure for a hydrogen gas sensor. J. Alloys Compd. 2020, 814, 151815. [Google Scholar] [CrossRef]
  31. Yin, X.-T.; Zhou, W.-D.; Li, J.; Wang, Q.; Wu, F.-Y.; Dastan, D.; Wang, D.; Garmestani, H.; Wang, X.-M.; Ţălu, Ş. A highly sensitivity and selectivity Pt-SnO2 nanoparticles for sensing applications at extremely low level hydrogen gas detection. J. Alloys Compd. 2019, 805, 229–236. [Google Scholar] [CrossRef]
  32. Meng, X.; Bi, M.; Xiao, Q.; Gao, W. Ultrasensitive gas sensor based on Pd/SnS2/SnO2 nanocomposites for rapid detection of H2. Sens. Actuators B Chem. 2022, 359, 131612. [Google Scholar] [CrossRef]
  33. Cai, L.; Zhu, S.; Wu, G.; Jiao, F.; Li, W.; Wang, X.; An, Y.; Hu, Y.; Sun, J.; Dong, X.; et al. Highly sensitive H2 sensor based on PdO-decorated WO3 nanospindle p-n heterostructure. Int. J. Hydrogen Energy 2020, 45, 31327–31340. [Google Scholar] [CrossRef]
  34. Li, K.; Liu, J.; Sheng, X.; Chen, L.; Xu, T.; Zhu, K.; Feng, X. 100-Fold Enhancement of Charge Transport in Uniaxially Oriented Mesoporous Anatase TiO2 Films. Chem.—A Eur. J. 2017, 24, 89–92. [Google Scholar] [CrossRef] [PubMed]
  35. Zhou, X.; Wang, Z.; Xia, X.; Shao, G.; Homewood, K.; Gao, Y. Synergistic Cooperation of Rutile TiO2 {002}, {101}, and {110} Facets for Hydrogen Sensing. ACS Appl. Mater. Interfaces 2018, 10, 28199–28209. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, B.; Aydil, E. Growth of Oriented Single-Crystalline Rutile TiO2 Nanorods on Transparent Conducting Substrates for Dye-Sensitized Solar Cells. J. Am. Chem. Soc. 2009, 131, 3985–3990. [Google Scholar] [CrossRef]
  37. Li, Y.; Guo, M.; Zhang, M.; Wang, X. Hydrothermal synthesis and characterization of TiO2 nanorod arrays on glass substrates. Mater. Res. Bull. 2009, 44, 1232–1237. [Google Scholar] [CrossRef]
  38. Sacco, A.; Mandrile, L.; Tay, L.-L.; Itoh, N.; Raj, A.; Moure, A.; Del Campo, A.; Fernandez, J.F.; Paton, K.R.; Wood, S.; et al. Quantification of titanium dioxide (TiO2) anatase and rutile polymorphs in binary mixtures by Raman spectroscopy: An interlaboratory comparison. Metrologia 2023, 60, 055011. [Google Scholar] [CrossRef]
  39. Shrivastava, A.; Gupta, V. Methods for the determination of limit of detection and limit of quantitation of the analytical methods. Chron. Young Sci. 2011, 2, 21–25. [Google Scholar] [CrossRef]
  40. Li, X.; He, H.; Tan, T.; Zou, Z.; Tian, Z.; Zhou, W.; Bao, Y.; Xia, X.; Gao, Y. Annealing effect on the methane sensing performance of Pt–SnO2/ZnO double layer sensor. Appl. Surf. Sci. 2023, 640, 158428. [Google Scholar] [CrossRef]
  41. Nasralla, N.H.S.; Yeganeh, M.; Astuti, Y.; Piticharoenphun, S.; Šiller, L. Systematic study of electronic properties of Fe-doped TiO2 nanoparticles by X-ray photoemission spectroscopy. J. Mater. Sci. Mater. Electron. 2018, 29, 17956–17966. [Google Scholar] [CrossRef]
  42. Liu, J.; He, L.; Dong, F.; Hudson-Edwards, K.A. The role of nano-sized manganese coatings on bone char in removing arsenic(V) from solution: Implications for permeable reactive barrier technologies. Chemosphere 2016, 153, 146–154. [Google Scholar] [CrossRef] [PubMed]
  43. Schierbaum, K.; Kirner, U.; Kirner, J.; Gdrel, W. Schottky-barrier and conductivity gas sensors based upon Pd/SnO2 and Pt/TiO2. Sens. Actuators B Chem. 1991, 4, 87–94. [Google Scholar] [CrossRef]
  44. Li, X.; Sun, Z.; Bao, Y.; Xia, X.; Tao, T.; Homewood, K.P.; Li, R.; Gao, Y. Comprehensively improved hydrogen sensing performance via constructing the facets homojunction in rutile TiO2 hierarchical structure. Sens. Actuators B Chem. 2022, 350, 130869. [Google Scholar] [CrossRef]
  45. Sun, Z.; Huang, L.; Zhang, Y.; Wu, X.; Zhang, M.; Liang, J.; Bao, Y.; Xia, X.; Gu, H.; Homewood, K.; et al. Homojunction TiO2 thin film-based room-temperature working H2 sensors with non-noble metal electrodes. Sens. Actuators B Chem. 2024, 398, 134675. [Google Scholar] [CrossRef]
  46. Katoch, A.; Abideen, Z.U.; Kim, H.W.; Kim, S.S. Grain-Size-Tuned Highly H2-Selective Chemiresistive Sensors Based on ZnO–SnO2 Composite Nanofibers. ACS Appl. Mater. Interfaces 2016, 8, 2486–2494. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD spectrum of the FTO substrate, NR-TiO2/FTO, and TiO2(R/A-X)/FTO. (b) Raman spectra of TiO2(R/A-X).
Figure 1. (a) XRD spectrum of the FTO substrate, NR-TiO2/FTO, and TiO2(R/A-X)/FTO. (b) Raman spectra of TiO2(R/A-X).
Sensors 24 00978 g001
Figure 2. SEM images showing the surface morphologies of (a) NR-TiO2, (b) TiO2(R/A-10 mL), (c) TiO2(R/A-15 mL), (d) TiO2(R/A-20 mL), (e) TiO2(R/A-25 mL), and (f) TiO2(R/A-30 mL).
Figure 2. SEM images showing the surface morphologies of (a) NR-TiO2, (b) TiO2(R/A-10 mL), (c) TiO2(R/A-15 mL), (d) TiO2(R/A-20 mL), (e) TiO2(R/A-25 mL), and (f) TiO2(R/A-30 mL).
Sensors 24 00978 g002
Figure 3. SEM images showing the cross-sectional morphologies of (a) NR-TiO2, (b) TiO2(R/A-30 mL), (c) TiO2(R/A-25 mL), (d) TiO2(R/A-20 mL), (e) TiO2(R/A-15 mL), (f) TiO2(R/A-10 mL), and the cross-sectional morphologies with higher magnification (50 kx) of (g) NR-TiO2, (h) TiO2(R/A-30 mL) and (i) TiO2(R/A-25 mL).
Figure 3. SEM images showing the cross-sectional morphologies of (a) NR-TiO2, (b) TiO2(R/A-30 mL), (c) TiO2(R/A-25 mL), (d) TiO2(R/A-20 mL), (e) TiO2(R/A-15 mL), (f) TiO2(R/A-10 mL), and the cross-sectional morphologies with higher magnification (50 kx) of (g) NR-TiO2, (h) TiO2(R/A-30 mL) and (i) TiO2(R/A-25 mL).
Sensors 24 00978 g003
Figure 4. Hydrogen-sensing characteristics of the TiO2(R/A-X) sensors at 25 °C. Change in resistance with time at different hydrogen concentrations for (a) NR-TiO2, (b) TiO2(R/A-30 mL), (c) TiO2(R/A-25 mL), (d) TiO2(R/A-20 mL), (e) TiO2(R/A-15 mL), and (f) TiO2(R/A-10 mL).
Figure 4. Hydrogen-sensing characteristics of the TiO2(R/A-X) sensors at 25 °C. Change in resistance with time at different hydrogen concentrations for (a) NR-TiO2, (b) TiO2(R/A-30 mL), (c) TiO2(R/A-25 mL), (d) TiO2(R/A-20 mL), (e) TiO2(R/A-15 mL), and (f) TiO2(R/A-10 mL).
Sensors 24 00978 g004
Figure 5. Hydrogen-sensing performances of the TiO2(R/A-X) sensors. (a) Response values at different hydrogen concentration from 25 to 30,000 ppm and (b) LOD of TiO2(R/A-25 mL) at low H2 concentration. (c,d) Response and recovery times of the sensors at different hydrogen concentration from 25 to 30,000 ppm.
Figure 5. Hydrogen-sensing performances of the TiO2(R/A-X) sensors. (a) Response values at different hydrogen concentration from 25 to 30,000 ppm and (b) LOD of TiO2(R/A-25 mL) at low H2 concentration. (c,d) Response and recovery times of the sensors at different hydrogen concentration from 25 to 30,000 ppm.
Sensors 24 00978 g005
Figure 6. (a,b) Bandgaps (Eg) and the energies of the valence band (Ev) and conduction band (Ec) for the rutile and anatase phases of titania. (c) O1s and (d) Ti2p XPS profiles of the NR-TiO2 and TiO2(R/A-25 mL) samples.
Figure 6. (a,b) Bandgaps (Eg) and the energies of the valence band (Ev) and conduction band (Ec) for the rutile and anatase phases of titania. (c) O1s and (d) Ti2p XPS profiles of the NR-TiO2 and TiO2(R/A-25 mL) samples.
Sensors 24 00978 g006
Figure 7. (a,b) Representation of I-V characteristics of the TiO2(R/A-X) sensor in air and H2 (2500 ppm), respectively. (c) Changes in the Schottky barrier height of the TiO2(R/A-X) sensor before and after exposure to H2 at 2500 ppm. (d) Relationship between the barrier height of the TiO2(R/A-X) sensor and response value.
Figure 7. (a,b) Representation of I-V characteristics of the TiO2(R/A-X) sensor in air and H2 (2500 ppm), respectively. (c) Changes in the Schottky barrier height of the TiO2(R/A-X) sensor before and after exposure to H2 at 2500 ppm. (d) Relationship between the barrier height of the TiO2(R/A-X) sensor and response value.
Sensors 24 00978 g007
Figure 8. (a,b) Schematic showing the change in the morphology of TiO2 nanorods during the second step of the reaction. (c,d) Schematic illustration of the mechanism of the sensor response.
Figure 8. (a,b) Schematic showing the change in the morphology of TiO2 nanorods during the second step of the reaction. (c,d) Schematic illustration of the mechanism of the sensor response.
Sensors 24 00978 g008
Figure 9. Sensing performance of TiO2(R/A-25 mL) for H2, NO, SO2, NH3, C2H6, and NO2 at 2500 ppm: (a) resistance change with time and (b) sensor response to different gases. (c) Cycling stability of TiO2(R/A-25 mL) at 25, 250, 1250, 2500, and 5000 ppm hydrogen after being placed in air for 170 d. (d) Comparison of the sensor response after 170 d with the initial response and percentage decline in the response. (e) Resistance changes of the TiO2(R/A-25 mL) sensor in response to H2 under different humidities and (f) responses of the sample under different humidities.
Figure 9. Sensing performance of TiO2(R/A-25 mL) for H2, NO, SO2, NH3, C2H6, and NO2 at 2500 ppm: (a) resistance change with time and (b) sensor response to different gases. (c) Cycling stability of TiO2(R/A-25 mL) at 25, 250, 1250, 2500, and 5000 ppm hydrogen after being placed in air for 170 d. (d) Comparison of the sensor response after 170 d with the initial response and percentage decline in the response. (e) Resistance changes of the TiO2(R/A-25 mL) sensor in response to H2 under different humidities and (f) responses of the sample under different humidities.
Sensors 24 00978 g009
Table 1. Comparison of sensing performances of H2 sensors.
Table 1. Comparison of sensing performances of H2 sensors.
MaterialsT (°C)Concentration (ppm)Response (Ra/Rg)Ref.
ZnO-SnO215010,00052[28]
SnO2-Co3O45003504.5[29]
Pt@NiORT50004.25[30]
1 at.% Pt/SnO235010056.5[31]
1.0 at% Pd/SnS2/SnO230050095[32]
3 at% PdO/WO31505077[33]
TiO2(R/A-25 mL)RT25001661This work
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

Wu, X.; Zhang, Y.; Zhang, M.; Liang, J.; Bao, Y.; Xia, X.; Homewood, K.; Lourenco, M.; Gao, Y. An Ultrasensitive Room-Temperature H2 Sensor Based on a TiO2 Rutile–Anatase Homojunction. Sensors 2024, 24, 978. https://doi.org/10.3390/s24030978

AMA Style

Wu X, Zhang Y, Zhang M, Liang J, Bao Y, Xia X, Homewood K, Lourenco M, Gao Y. An Ultrasensitive Room-Temperature H2 Sensor Based on a TiO2 Rutile–Anatase Homojunction. Sensors. 2024; 24(3):978. https://doi.org/10.3390/s24030978

Chicago/Turabian Style

Wu, Xuefeng, Ya Zhang, Menghan Zhang, Jianhu Liang, Yuwen Bao, Xiaohong Xia, Kevin Homewood, Manon Lourenco, and Yun Gao. 2024. "An Ultrasensitive Room-Temperature H2 Sensor Based on a TiO2 Rutile–Anatase Homojunction" Sensors 24, no. 3: 978. https://doi.org/10.3390/s24030978

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