Next Article in Journal
Metal Oxide Semiconductor Gas Sensors for Lung Cancer Diagnosis
Next Article in Special Issue
MXene/NiO Composites for Chemiresistive-Type Room Temperature Formaldehyde Sensor
Previous Article in Journal
Pyranine Immobilized on Aminopropyl-Modified Mesoporous Silica Film for Paraquat Detection
Previous Article in Special Issue
Facet-Dependent Gas Adsorption Selectivity on ZnO: A DFT Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Potentiometric H2 Sensing of Pr0.1Ce0.9O2−δ Using Fe2O3 Surface Modification

State Key Laboratory of Fire Science, Department of Safety Science and Engineering, University of Science and Technology of China, Hefei 230026, China
*
Author to whom correspondence should be addressed.
Chemosensors 2023, 11(4), 250; https://doi.org/10.3390/chemosensors11040250
Submission received: 27 March 2023 / Revised: 10 April 2023 / Accepted: 11 April 2023 / Published: 17 April 2023

Abstract

:
Monitoring the concentration of hydrogen is very important as it is a flammable and explosive gas. Non-Nernstian potentiometric hydrogen sensors hold promising potentials for the sensitive detection of hydrogen. This paper reports the improved H2-sensing performance of a mixed oxide ion-electron conducting (MIEC) Pr0.1Ce0.9O2−δ (PCO) electrode using Fe2O3 surface modification. The Fe2O3-modified PCO exhibited a high response of −184.29 mV to 1000 ppm H2 at 450 °C. The response values exhibited a linear or logarithmic dependence on the H2 concentration for below or above 20 ppm, respectively. A sensitivity of −74.9 mV/decade in the concentration range of 20–1000 ppm was achieved, and the theoretical limit of detection was calculated to be 343 ppb. Moreover, a power-law relationship between the response time and the concentration value was also found. Electrochemical impedance analyses revealed that the excellent H2-sensing performance may be attributed to the large ratio of the electrochemical activity of the hydrogen oxidation reaction (HOR) over the oxygen exchange reaction (OER). In addition, the distribution of relaxation time (DRT) results reveal that the enhanced electrochemical kinetics caused by H2 presence in air is mainly related to acceleration of the electrode surface processes.

1. Introduction

Hydrogen (H2) is a very promising clean and renewable fuel with abundant reserves and high energy density which is widely utilized in many fields such as spaceships and many industrial sectors [1,2,3]. However, H2 is one of the most important conventional fire detection marker gases due to its many unusual properties such as its lack of color or odor, its ultra-low ignition energy of only 0.02 mJ and its very large flammable range (4–75%) [4,5]. During the storage and usage of H2, the hydrogen embrittlement phenomenon can lead to the failure of the protective layer and cause hydrogen spillage or even leakage, which can lead to fire and explosion accidents and pose a serious threat to the safety of human life and property [6]. Therefore, it is important for the popularization and development of the use of hydrogen energy to research low-concentration hydrogen sensors that can detect ppm or even ppb levels, in order to warn about hydrogen leaks as soon as possible [7]. In addition, low-concentration hydrogen detection has important potential value in the fields of earthquake early warning and respiratory gas disease monitoring [8,9].
Currently, the most widely used sensors are metal oxide semiconductors, catalytic combustion and electrochemical gas sensors. Semiconductor sensors can detect hydrogen at the ppm level, but with poor selectivity. Catalytic sensors are mainly used to detect the hydrogen in air at pre-explosive concentrations of 0.1–2% and have the advantages of low operating temperature and high selectivity [10,11]. Electrochemical non-Nernstian potentiometric gas sensors with simple configurations, wider operating temperatures, high sensitivity and low cost have attracted considerable attention from researchers. Solid electrolyte, sensitive electrode (SE) and reference electrode (RE) are generally the three important components of a non-Nernstian-type potentiometric gas sensor [12]. This type of sensor responds to the measured gas through changes in electrode potential caused by the competition between various electrochemical reactions proceeding over the three-phase boundary (TPB, which is the boundary between the electrolyte, electrode and gas), which do not follow the Nernst equation [13,14,15]. Three different mechanisms, namely the electrochemical, the electrical and the surface electrostatic mechanism, have been proposed for non-Nernstian sensing behavior, among which the electrochemical one has been very often adopted, according to mixed-potential theory [16,17,18]. Over the past decades, researchers have done a great deal to develop high-performance non-Nernstian gas sensors. It was found that SE materials play a crucial role in determining the sensor response. For example, Miura et al. developed a ZrSiO4(+Au) sensitive electrode material for detecting H2, which had a high sensitivity and good selectivity for H2 [15]. Zhang et al. reported a novel La0.8Sr0.2Cr0.5Fe0.5O3−δ sensitive material, which had a fast response, high sensitivity and selectivity to H2 [14]. Lu et al. fabricated a NH3 sensor based on a metal-oxide (NiO, SnO2, WO3)-modified FeVO4 sensitive electrode material, which had the advantages of a short response time, short recovery time and good oxygen resistance [19]. While a great amount of sensitive electrode materials has been investigated, the optimal potential sensing performance has not yet been observed.
Depending on the widely adopted mixed-potential principle, the ratio of electrochemical activity of the SE material to the interfacial electrochemical reactions proceeding over the TPB can serve as a valid indicator of the sensor response. A large activity ratio of hydrogen oxidation over oxygen reduction favors a high response to hydrogen [20,21]. In this sense, regulating the electrochemical activities of the SE material could tune the sensor response, which provides a strategy to select sensitive electrode materials rationally. For instance, Zhang et al. investigated the relationship between the mixed-potential hydrogen response and the electrochemical activity of perovskite oxides, which showed that the activity ratio can be used as a response descriptor potentially applicable to different materials and target gases [16]. In addition, Nicollet et al. recently found that the oxygen surface exchange (oxygen reduction/evolution) kinetics of MIEC oxides is highly dependent upon surface acidity, and infiltrating strongly acidic oxides onto the material surface may dramatically decrease the oxygen exchange kinetics [22]. This finding suggests that there is a possibility to boost the potentiometric gas sensing performance of MIEC electrodes using strongly acidic oxide surface modification. Although oxide surface modification has been used to prepare composite electrode materials [12], its effect on non-Nernstian gas sensing properties and the underlying mechanism remains unclear.
This work studied the influence of strongly acidic Fe2O3 surface modification on the potentiometric gas sensing performance of a model MIEC oxide, Pr0.1Ce0.9O2−δ (PCO). Results show that compared with pure PCO, the Fe2O3-modified sample exhibited a markedly enhanced H2 sensing performance. The excellent H2 sensing was ascribed to a large electrochemical activity ratio, and the enhanced electrochemical performance was associated with the acceleration of electrode surface processes.

2. Materials and Methods

2.1. Chemicals

Praseodymium nitrate (Pr(NO3)3·6H2O) was purchased from Shanghai Aladdin Biochemical Technology. Iron nitrate (Fe(NO3)3·9H2O), cerium nitrate (Ce(NO3)3·6H2O), ethylene glycol (EG) and citric acid (CA) were purchased from Sinopharm Chemical Reagent, China. No further purification of these reagents was required, and they all reached analytical purity.

2.2. Materials Synthesis and Sensors Preparation

An adjusted Pechini-type sol-gel approach was used to obtain the PCO powder. Pr(NO3)3·6H2O, Ce(NO3)3·6H2O, EG and CA were mixed according to the molar ratio of 1:1:2:2 to form the precursor solution. This solution was stirred and heated to 80 °C until the formation of a yellow sol, which was dried overnight at 110 °C, and finally sintered in a muffle furnace for 5 h under 700 °C to obtain a reddish-brown powder.
All the solid electrolytes in this experiment were square YSZ plates (yttria-stabilized zirconia) with a size of 13 mm × 13 mm × 0.25 mm, which were purchased from Hefei Kejing, China. Pt paste (PE-Pt-7840, Sino-Platinum Metals, Kunming, China) was painted on both sides of the YSZ as a current collector and RE, respectively, and it was strip-shaped with a size of 1 mm × 3 mm. Then, Pt wires were adhered to the Pt strips by using high-temperature Pt paste and finally calcined for 30 min at 1000 °C. We mixed the organic solvent (α-terpineol and ethyl cellulose) with the prepared PCO powder at a mass ratio of 1:1 and ground it for 2 h to obtain the SE pastes, which were subsequently painted on the side of the electrolyte opposite to the RE, and finally sintered for 2 h at 900 °C. The diameter of all SE was 3 mm and the distance between them was ~5 mm. To obtain the Fe2O3-modified electrodes, which was named PCO-Fe2O3, 0.4 M precursor iron nitrate solution was prepared. Then, 5 μL of the solution was dropped onto the surface of the PCO electrode using a pipette. After the surface modification was completed, samples were dried for 0.5 h at 100 °C, and finally sintered for 1h in an atmosphere at 600 °C.
Using similar procedures, symmetrical electrodes for the electrochemical tests were also prepared by painting identical SE pastes symmetrically on different sides of the YSZ substrate, followed by sintering.

2.3. Characterization and Gas Sensing Measurements

The phase structure and crystal composition of PCO and PCO-Fe2O3 electrodes sintered on YSZ were obtained through the X-ray powder diffraction (XRD) technique with a TTR-III diffractometer (Rigaku Corporation, Tokyo, Japan), using Cu Kα radiation and a 2θ collection range of 20°–90°. The microstructure of the electrodes was observed using a cold field emission scanning electron microscope (SEM, SU8220) equipped with an Oxford Aztec series X-ray spectrometer. The elemental composition of the electrodes was observed using X-ray photoelectron spectroscopy (XPS, ESCLAB 250) which used Al Kα as the excitation source and used the C1s of binding energy at 284.8 eV for energy calibration.
A home-made two-chamber device was used to obtain the pristine response values for these sensitive electrodes, the details of which can be found in the previous work of our group [18]. In this method, high-temperature silver paste (DAD-87, Shanghai Research Institute of Synthetic Resins, Shanghai, China) was used to glue the YSZ substrate to the quartz tube, thereby forming two separate chambers. Since the SE and RE were located at opposing sides of the YSZ, the atmosphere in the SE and RE can be regulated separately. During the test, the chamber in which the SE was located was alternately introduced into the mixture of air/test gas and air, while the atmosphere in the chamber where the RE was located was always air. In order to obtain the required test gas concentrations for the experiment, we mixed dry air with standard gas (H2, NO2, CO, NH3, C7H8, H2S and CH4), using N2 as the equilibrium gas. All gases were purchased from Nanjing Special Gases. The flow rate of each gas path was controlled using a mass flow meter (MFC, CS200, Beijing Seven-star Electronics, Beijing, China) and the total flow rate for the chambers of SE and RE was 200 mL/min.
The SE and RE were linked to the positive and negative terminals of an Agilent 34972A data acquisition unit to measure the response signal of the sensor, i.e., the open circuit voltage (OCV) between the SE and RE. In this paper, ΔV = Vgas − Vair is defined as the response value of the sensor, where Vgas and Vair represent the open circuit voltage between the SE and RE when the SE is exposed to the mixture of test gas/air and to the air, respectively. The linear fit slope of the dependence function of the response value on the test gas concentration or logarithm concentration is taken as the sensitivity (S). Response time is defined as the time taken from when SE is exposed to the test gas to when the change in OCV reaches 90% of the response value, while recovery time is defined as the time taken between when the SE is exposed to air again and when the change in OCV reaches 10% of the response value. The lower detection limit (LOD) is calculated by the following formula:
LOD = 3 rms/S
where rms is the standard deviation of the base OCV and S represents sensitivity [23]. Humid gas was obtained by bubbling the gas through deionized water or a saturated salt solution (MgCl2/NaBr saturated solution) and the relative humidity of the gas was tested using a hygrometer (Testo 605-H1).

2.4. Electrochemical Analysis Methods

Electrochemical impedance spectroscopy (EIS) was conducted on samples with symmetrical electrodes of the SE and a Pt RE with an electrochemical workstation (VersaSTAT 3). During the EIS test, the frequency range and amplitude were set to 0.01 Hz–1 MHz and 20 mV, respectively. Both SEs were exposed to the same atmosphere (air or H2/air). ZSimpwin software was used to fit the data obtained from the impedance test, wherein an equivalent circuit composed of three R-CPE elements connected in series was used. The distribution of relaxation time (DRT) is a powerful technique that allows the separation of different key electrochemical reaction steps [18]. The impedance data obtained from measurements in air or H2/air were transformed into a function of time constants τ for DRT analysis. For the DRT analysis, the value of the regularization parameter λ was set to 0.001.

3. Results

3.1. Characterization Results of Materials

The diffraction peaks of PCO are in perfect agreement with those in the standard card (JCPDS No.:43-1002), indicating a single-phase fluorite crystal structure, as shown in the XRD pattern of Figure 1. The XRD patterns of the PCO-Fe2O3 samples are similar to that of PCO, and the presence of the oxide modifier was not detected, probably due to the relatively low loading or small particle size.
XPS was used for the purpose of analyzing whether PCO-Fe2O3 contains other impurity elements and the oxidation state of the modified metal ions, which is shown in Figure 2. Figure 2a shows that PCO-Fe2O3 was composed of only Ce, Pr, Fe and O, without the presence of any other elements except C. Figure 2b shows Fe 2p3/2, Fe 2p1/2 and a satellite peak composed of the Fe 2p spectrum. The spacing between the Fe 2p3/2 and Fe 2p1/2 doublets is 13.6 eV, and the satellite peak is located at 718 eV. This satellite peak has been observed in other studies and can be used as an indication of the formation of Fe2O3 [24,25,26].
The SEM images presented in Figure 3a,b show that the electrode sintered on YSZ was mainly composed of particles with a size ranging from several tens of nanometers to a few micrometers. Some pores can be observed among the particles, which facilitate gas diffusion. Figure 3c–f shows the EDS analysis results of PCO-Fe2O3. The modified Fe elements are uniformly distributed on the SE surface, confirming a well-dispersed composite synthesis. In addition, Figure S1a–d also gives the cross-sectional EDS mapping of PCO-Fe2O3, indicating that Fe2O3 is present evenly not only on the surface of PCO but also inside PCO. These results reveal that Fe2O3 was successfully formed on the PCO surface but likely had a very fine particle size.

3.2. Sensing Performance

The pristine responses of the PCO-Fe2O3 and PCO sensors were measured in a two-chamber mode by exposing the SE to base air or test gas/air mixture and the Pt RE to an air reference. The dynamic variation of the response values of the PCO and PCO-Fe2O3 sensors with the increasing H2 concentration at 450 °C and 500 °C are clearly observed in Figure 4. When SEs were exposed to only air, the base OCV for all the sensors remained at ~0. The OCV shifted negatively and approached a steady value when exposing SEs to H2. While the SEs atmosphere was again switched back to base air, the OCV increased rapidly and returned to the initial value. Clearly, all the sensors exhibited negative responses to H2. The OCV variation became more pronounced and the response increased with the increasing H2 concentration. Under the tested conditions, the variation of OCV for PCO-Fe2O3 was significantly larger than that for PCO, indicating that the Fe2O3 modification successfully improved the response value. Note that the increase in response with the decrease in temperature has similarly been observed for other non-Nernstian gas sensors [12,14,27,28,29]. The temperature dependence of the response is a result of the joint effects of many affecting factors, including the electrochemical reaction mechanism and heterogeneous catalytic activities, which are also a function of temperature [29].
The response values of the PCO-Fe2O3 and PCO sensors as a function of H2 concentration or logarithm H2 concentration are shown in Figure 5. It can be directly observed that the response values of the PCO-Fe2O3 sensor to H2 are much greater than that of the PCO one, regardless of whether the H2 concentrations were low or high. The lowest H2 concentration in the experiment was measured at 1 ppm, and the response of PCO-Fe2O3 to 1 ppm H2 was −3.9 mV, while PCO had almost no response to 1 ppm H2. The response values of the two sensors showed a different dependence on the H2 concentration in the region of low and high H2 concentrations, respectively, i.e., the response values showed a linear relationship with the H2 concentration in the region of 1–20 ppm, while the response values showed a linear relationship with the logarithm H2 concentration in the region of 20–1000 ppm. For PCO-Fe2O3, sensitivities of −2.97 mV/ppm and −74.9 mV/decade were obtained in the low and high H2 concentration ranges, respectively, which were 7.4 and 1.3 times greater than those for PCO in the same concentration range. In addition, the LOD for PCO-Fe2O3 was calculated to be 343 ppb, which is around 3.5 times lower than that of PCO (1.2 ppm). Similar concentration dependence has often been observed in other systems, and this is in accordance with the mixed-potential sensing mechanism.
The relationship between response time and recovery time with the concentrations of H2 for the PCO-Fe2O3 sensor was also studied, as shown in Figure 6. It can be observed that the response time reduces quickly with the increase in H2 concentration in the range of 20–300 ppm, while in a higher concentration range, the response time reduces slowly with the increase in H2 concentration and remains almost constant (Figure 6a). The response time declined with the rising temperature, with a value of 32 s and 22 s for 1000 ppm H2 at 450 °C and 500 °C, respectively. A power-law relationship between response time and H2 concentration on the log-log scale. Such behavior has also been observed in previous studies, agreeing with a theoretical simulation according to a mixed-potential sensing mechanism [27]. For recovery time, the variation with H2 concentration was similar to that of the response time at 450 °C, while at 500 °C, the recovery time reduced slowly with the increase in H2 concentration in the tested concentration range. The recovery time also decreased with an increasing temperature, with a value of 322 s and 122 s for 1000 ppm H2 at 450 °C and 500 °C, respectively, which is much longer than the response time.
Figure 7 compares the response of PCO-Fe2O3 towards different gases at 500 °C. PCO-Fe2O3 responded most sensitively to H2 and moderately to CO and C7H8 but exhibited a rather small response to the other gases (NO2, NH3, H2S and CH4). Apparently, the PCO-Fe2O3 sensor has excellent selectivity for H2, with particular regard to CH4, H2S, NH4 and NO2. Humidity tolerance is also one important performance indicator of gas sensors. As shown in Figure 8a,b, the basic OCV of the PCO-Fe2O3 sensor remained almost constant as the relative humidity (RH) of the base gas varied in the range of 2.8–99%. The H2 response value decreased slightly with the increase in RH, and only an ~13.4% reduction was observed as RH increased from 2.8% to 99%. Repeatability is of significance for the practical application of gas sensors. As shown in Figure 8c, by alternately switching the test gas between air and the 300 ppm H2/air mixture, the response transients can be reproduced very well. The OCV varied by a similar magnitude in each cycle of the H2 response and can be restored to the baseline upon H2’s removal from air, indicating the excellent repeatability of the measurements. Figure 8d presents a 15-day test of the PCO-Fe2O3 sensor at 500 °C for 300 ppm H2. The response fluctuated between −97.32 mV and −107.88 mV, with a maximal variation of only 9.7%, revealing the good stability of the sample within the experimental period.

4. Discussion

For the widely adopted electrochemical (mixed-potential) mechanism, sensors usually exhibit a negative response to reducing gases such as H2 [16]. When the chamber in which the SE is located contains only air, just one oxygen exchange reaction (OER) will occur on the TPB:
O 2 + 4 e 2 O 2
Upon the introduction of the analyte (H2 in this case), a second electrochemical reaction takes place concurrently along with the OER, which is the hydrogen oxidation reaction (HOR):
H 2 + O 2 H 2 O ( g ) + 2 e
The HOR reaction results in a change in electric charges in the double layer, which behaves similarly to a capacitor [27], thereby altering the electrode potential. As electrons are generated in the forward HOR, a negative contribution from the electrochemical mechanism to ΔE (response) is yielded. Furthermore, the linear relationship between the response value and H2 concentration or its logarithm and the power-law concentration dependence of response time for PCO-Fe2O3 have also been observed for other electrode materials [14,28], in line with theoretical analyses based on mixed-potential theory [27,29]. Therefore, the non-Nernstian gas-sensing behavior of PCO-Fe2O3 most probably follows the electrochemical mixed-potential mechanism.
Based on the mixed-potential mechanism, response values of sensors (absolute values) under Tafel kinetics can be expressed as [16,29]
S = b c , 1 b c , 1 + b a , 2 Δ E e + b c , 1 b a , 2 b c , 1 + b a , 2 l g i 2 0 i 1 0
where ΔEe = Ee1Ee2 is the thermodynamic equilibrium voltage (the difference in the equilibrium potentials between the OER and HOR), i 1 0 and i 2 0 are the exchange current for the OER and HOR, respectively, and bc,1 and ba,2 are the logarithmic cathodic tafel slope of the OER and the logarithmic anodic tafel slope or the HOR, respectively.
Taking into account the inverse relationship between exchange current density and polarization impedance (Rc,1, Ra,2), Equation (1) can be deformed as [30]
S = b c , 1 b c , 1 + b a , 2 Δ E e + b c , 1 b a , 2 b c , 1 + b a , 2 l g 2 R c , 1 R a , 2
where Rc,1 and Ra,2 represent the indicator of electroactivity for the OER and HOR of the sensing electrode material. Equations (2) and (3) indicate that materials of large i 2 0 / i 1 0 (or Rc,1/Ra,2) are desirable for achieving a high response performance.
To measure the polarization resistances, symmetrical electrodes of PCO-Fe2O3 were studied with impedance spectroscopy. The impedance spectra obtained typically consisted of two semicircles. The semicircle at high frequency corresponds to ohmic resistance, while the width of the semicircle at low frequency gives the polarization resistance. A large polarization impedance represents slow electrochemical kinetics. Rc,1 and Ra,2 were obtained from the low-frequency semicircle width of the impedance spectra tested in base air and oxygen-free H2/N2, respectively [16]. As shown in Figure 9, Rc,1 is obviously much larger than Ra,2, indicating that the HOR is much more active than the OER. The H2 response increased almost linearly with the increasing logarithmic Rc,1/Ra,2, in line with Equation (2). The variation of response with the activity ratio has been interpreted in terms of an electrochemical mixed-potential mechanism. For PCO-Fe2O3, the HOR activity is much larger than the OER one. Hence, the HOR competed strongly with the OER while the SE was exposed to an atmosphere containing H2, resulting in the significant cathodic polarization of the OER and thus a large response [29]. Furthermore, Ra,2 decreased with the increase in H2 concentration, suggesting the higher the H2 concentration, the greater the HOR reaction rate will be. Thus, the response increased with increasing H2 concentration.
The distribution of relaxation time (DRT) was applied to analyze, in detail, the impedance spectra of PCO-Fe2O3 tested in air and H2/air to gain insight into changes in the electrochemical processes occurring at the electrode [31]. As shown in Figure 10a, the DRT plots obtained from the tests in each atmosphere can be divided into low-frequency (LF) peaks, intermediate-frequency (IF) peaks and high-frequency (HF) peaks, according to the frequency magnitude, which is consistent with the impedance fitting equivalent circuit plots. The literature suggests that these peaks correspond to different sub-processes or steps of OER for an MIEC electrode, i.e., the HF peak is associated with ionic transmission through the electrode/electrolyte intersection; the IF peak is related to charge transport on the electrode surface; and the LF peak is related to oxygen adsorption and diffusion on the electrode surface [32]. The magnitude of the peak area represents the resistance for the corresponding process, and a large peak corresponds to slow kinetics. As shown in Figure 10a, the LF peak is obviously larger than IF and HF ones, which indicates that the oxygen adsorption and diffusion on the electrode surface associated with the LF peak play a determining role in the overall electrochemical reaction rate. In addition to this, we also observed that the intensity of the LF peak decreases when the SE is exposed to an atmosphere containing H2. The LF peak significantly decreased in 60 ppm H2, which became more pronounced as the H2 concentration increased. The IF peak also decreased in 60 ppm H2/air, but the peak area remained almost constant afterwards, at higher H2 concentrations. Concerning the HF peak, no significant effect of H2 presence was observed. These results reveal that the enhanced electrochemical kinetics caused by H2’s presence in air is mainly related to acceleration of the electrode surface processes.
When compared with reported non-Nernstian hydrogen sensors [14,15,16,17,28,33,34,35,36,37,38], PCO-Fe2O3 exhibited comparable H2 response characteristics (Table S1). In addition, for practical applications, hydrogen sensors need to meet a number of technical requirements. According to the US DOE (Department of Energy), stationary hydrogen sensors should be capable of detecting down to 1000 ppm of hydrogen with a response/recovery time of 30 s [39]. Although PCO-Fe2O3 meets the requirement for the detection limit and response time (at 500 °C), its recovery speed is insufficient. It is also noteworthy that hydrogen sensors operable at lower temperaturs are most desirable, which can increase safety and reduce the power consumption of sensors. Further enhancement of the sensor performance is needed to meet these practical requirements, which may be achieved by optimizing the materials, including the electrode materials and morphology, oxide modifiers and solid electrolyte.

5. Conclusions

In this paper, a composite sensitive PCO-Fe2O3 electrode was fabricated using surface modification of a mixed-conducting PCO electrode with strongly acidic Fe2O3, and the potential response performance and electrochemical activity of PCO-Fe2O3 were investigated. Results showed that modification with strongly acidic Fe2O3 led to a remarkable improvement in the H2-sensing characteristics, including the response, sensitivity, LOD and response speed. The sensing behavior is in good agreement with the electrochemical mixed-potential mechanism. This enhanced H2 sensing by Fe2O3 was primarily due to an increase in the ratio of electrochemical activity of hydrogen oxidation over oxygen reduction. In addition, the enhanced electrochemical kinetics caused by H2 presence in air is mainly related to acceleration of electrode surface processes. These findings show that acidic oxide surface modification may be an effective way to develop high-performance sensitive electrode materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemosensors11040250/s1. Figure S1: (a–d) Cross-section EDS mapping for PCO-Fe2O3; (e) Surface EDS quantification results of PCO-Fe2O3. Table S1: Comparison of H2 sensing characteristics of PCO-Fe2O3 with literature data.

Author Contributions

Conceptualization, J.Y.; investigation, L.W.; validation, L.W.; writing—original draft, L.W.; writing—review & editing, J.Y. and L.W.; funding acquisition, J.Y.; project administration, J.Y.; supervision, J.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (Grant 2021YFC3000300), the National Natural Science Foundation of China (Grant 62271462, 61871359), and Fundamental Research Funds for the Central Universities (Grant WK2320000053).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available from authors upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hassan, K.; Tung, T.T.; Yap, P.L.; Nine, M.J.; Kim, H.C.; Losic, D. Fast response hydrogen gas sensor based on Pd/Cr nanogaps fabricated by a single-step bending deformation. Anal. Chim. Acta 2020, 1138, 49–58. [Google Scholar] [CrossRef] [PubMed]
  2. Buttner, W.J.; Post, M.B.; Burgess, R.; Rivkin, C. An overview of hydrogen safety sensors and requirements. Int. J. Hydrog. Energy 2011, 36, 2462–2470. [Google Scholar] [CrossRef]
  3. Pak, Y.; Lim, N.; Kumaresan, Y.; Lee, R.; Kim, K.; Kim, T.H.; Kim, S.-M.; Kim, J.T.; Lee, H.; Ham, M.-H.; et al. Palladium Nanoribbon Array for Fast Hydrogen Gas Sensing with Ultrahigh Sensitivity. Adv. Mater. 2015, 27, 6945–6952. [Google Scholar] [CrossRef] [PubMed]
  4. Gutmacher, D.; Hoefer, U.; Wöllenstein, J. Gas sensor technologies for fire detection. Sens. Actuators B Chem. 2012, 175, 40–45. [Google Scholar] [CrossRef]
  5. Hübert, T.; Boon-Brett, L.; Black, G.; Banach, U. Hydrogen sensors—A review. Sens. Actuators B Chem. 2011, 157, 329–352. [Google Scholar] [CrossRef]
  6. Toribio, J.; Lorenzo, M.; Aguado, L. Innovative Design of Residual Stress and Strain Distributions for Analyzing the Hydrogen Embrittlement Phenomenon in Metallic Materials. Materials 2022, 15, 9063. [Google Scholar] [CrossRef]
  7. Chauhan, P.S.; Bhattacharya, S. Hydrogen gas sensing methods, materials, and approach to achieve parts per billion level detection: A review. Int. J. Hydrogen Energy 2019, 44, 26076–26099. [Google Scholar] [CrossRef]
  8. Shen, B.; Luo, J.; Xie, Y.; Zhang, D.; Fan, P.; Zhong, A. Hydrogen gas ppb-level detection based on AlGaN/GaN high electron mobility transistor with 2.0 nm thick Pt gate layer. Appl. Phys. Lett. 2019, 115, 254104. [Google Scholar] [CrossRef]
  9. Podlepetsky, B.; Samotaev, N.; Kovalenko, A. Responses’ parameters of hydrogen sensors based on MISFET with Pd(Ag)-Ta2O5-SiO2-Si structure. Sens. Actuators B Chem. 2019, 290, 698–705. [Google Scholar] [CrossRef]
  10. Ivanov, I.I.; Baranov, A.M.; Talipov, V.A.; Mironov, S.M.; Akbari, S.; Kolesnik, I.V.; Orlova, E.D.; Napolskii, K.S. Investigation of catalytic hydrogen sensors with platinum group catalysts. Sens. Actuators B Chem. 2021, 346, 130515. [Google Scholar] [CrossRef]
  11. Wu, R.-J.; Tian, X.-M.; Hua, Z.-Q.; Lu, N.; Wang, P. A low temperature catalytic combustible gas sensor based on Ru supported zeolite catalyst films. Chin. J. Anal. Chem. 2021, 49, 63–68. [Google Scholar] [CrossRef]
  12. Bhardwaj, A.; Kumar, A.; Bae, H.; Park, C.J.; Song, S.J. Surface decorated spinel-oxide electrodes for mixed-potential ammonia sensor: Performance and DRT analysis. J Hazard Mater. 2020, 396, 122601. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, B.; Liu, F.; Yang, X.; Guan, Y.; Ma, C.; Hao, X.; Liang, X.; Liu, F.; Sun, P.; Zhang, T.; et al. Fabrication of Well-Ordered Three-Phase Boundary with Nanostructure Pore Array for Mixed Potential-Type Zirconia-Based NO2 Sensor. ACS Appl. Mater. Interfaces 2016, 8, 16752–16760. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, H.; Yi, J.; Jiang, X. Fast Response, Highly Sensitive and Selective Mixed-Potential H2 Sensor Based on (La, Sr)(Cr, Fe)O3-δ Perovskite Sensing Electrode. ACS Appl. Mater. Interfaces 2017, 9, 17218–17225. [Google Scholar] [CrossRef] [PubMed]
  15. Anggraini, S.A.; Ikeda, H.; Miura, N. Tuning H2 Sensing Performance of Zirconia-based Sensor using ZrSiO4(+Au) Sensing-electrode. Electrochim. Acta 2015, 171, 7–12. [Google Scholar] [CrossRef]
  16. Zhang, H.; Yi, J.; Zhang, Z.; Zhang, H. The relation between mixed-potential hydrogen response and electrochemical activities for perovskite oxides. Sens. Actuators B Chem. 2022, 352, 130988. [Google Scholar] [CrossRef]
  17. Miura, N.; Sato, T.; Anggraini, S.A.; Ikeda, H.; Zhuiykov, S. A review of mixed-potential type zirconia-based gas sensors. Ionics 2014, 20, 901–925. [Google Scholar] [CrossRef]
  18. Zhang, Z.; Yi, J.; Han, H.; Meng, Y.; Zhang, H.; Jiang, Y. Electrochemical Response of Mixed Conducting Perovskite Enables Low-Cost High-Efficiency Hydrogen Sensing. ACS Appl. Mater. Interfaces 2022, 14, 33580–33588. [Google Scholar] [CrossRef]
  19. Lu, Q.; Huang, L.; Hao, X.; Li, W.; Wang, B.; Wang, T.; Liang, X.; Liu, F.; Wang, C.; Lu, G. Mixed potential type NH3 sensor based on YSZ solid electrolyte and metal oxides (NiO, SnO2, WO3) modified FeVO4 sensing electrodes. Sens. Actuators B Chem. 2021, 343, 130043. [Google Scholar] [CrossRef]
  20. Lu, G.; Miura, N.; Yamazoe, N. High-temperature hydrogen sensor based on stabilized zirconia and a metal oxide electrode. Sens. Actuators B Chem. 1996, 35, 130–135. [Google Scholar] [CrossRef]
  21. Liu, T.; Zhang, X.; Yuan, L.; Yu, J. A review of high-temperature electrochemical sensors based on stabilized zirconia. Solid State Ion. 2015, 283, 91–102. [Google Scholar] [CrossRef]
  22. Nicollet, C.; Toparli, C.; Harrington, G.F.; Defferriere, T.; Yildiz, B.; Tuller, H.L. Acidity of surface-infiltrated binary oxides as a sensitive descriptor of oxygen exchange kinetics in mixed conducting oxides. Nat. Catal. 2020, 3, 913–920. [Google Scholar] [CrossRef]
  23. Li, J.; Lu, Y.; Ye, Q.; Cinke, M.; Han, J.; Meyyappan, M. Carbon Nanotube Sensors for Gas and Organic Vapor Detection. Nano Lett. 2003, 3, 929–933. [Google Scholar] [CrossRef]
  24. Balamurugan, S.; Naresh, N.; Prakash, I.; Satyanarayana, N. Capacity fading mechanism of Li2O loaded NiFe2O4/SiO2 aerogel anode for lithium-ion battery: Ex-situ XPS analysis. Appl. Surf. Sci. 2021, 535, 147677. [Google Scholar] [CrossRef]
  25. Fan, Y.; Wang, W.; Zhang, J.; Lu, Y.; Liu, C.; Adimi, S.; Zhou, J.; Ruan, S.; Chen, Y. Construction of p-n heterojunctions by modifying MOF-derived α-Fe2O3 with partially covered cobalt tungstate for high-performance ethyl acetate detection. Sens. Actuators B Chem. 2021, 344, 130129. [Google Scholar] [CrossRef]
  26. Yan, W.; Fan, H.; Zhai, Y.; Yang, C.; Ren, P.; Huang, L. Low temperature solution-based synthesis of porous flower-like α-Fe2O3 superstructures and their excellent gas-sensing properties. Sens. Actuators B Chem. 2011, 160, 1372–1379. [Google Scholar] [CrossRef]
  27. Zhang, J.; Zhang, H.; Teng, Y.; Yi, J. Simulation of dynamic response of mixed-potential hydrogen sensors. Int. J. Hydrog. Energy 2022, 47, 33891–33902. [Google Scholar] [CrossRef]
  28. Yi, J.; Zhang, H.; Zhang, Z.; Chen, D. Hierarchical porous hollow SnO2 nanofiber sensing electrode for high performance potentiometric H2 sensor. Sens. Actuators B Chem. 2018, 268, 456–464. [Google Scholar] [CrossRef]
  29. Yi, J.; Han, H. Analysis of factors affecting response for mixed potential gas sensors. Electrochim. Acta 2021, 379, 138129. [Google Scholar] [CrossRef]
  30. Liu, Q.; Dong, X.; Xiao, G.; Zhao, F.; Chen, F. A Novel Electrode Material for Symmetrical SOFCs. Adv. Mater. 2010, 22, 5478–5482. [Google Scholar] [CrossRef]
  31. Lu, M.Y.; Scipioni, R.; Park, B.-K.; Yang, T.; Chart, Y.A.; Barnett, S.A. Mechanisms of PrOx performance enhancement of oxygen electrodes for low and intermediate temperature solid oxide fuel cells. Mater. Today Energy 2019, 14, 100362. [Google Scholar] [CrossRef]
  32. Adler, S.B. Factors Governing Oxygen Reduction in Solid Oxide Fuel Cell Cathodes. Chem. Rev. 2004, 104, 4791–4844. [Google Scholar] [CrossRef] [PubMed]
  33. Sekhar, P.K.; Brosha, E.L.; Mukundan, R.; Nelson, M.A.; Williamson, T.L.; Garzon, F.H. Development and testing of a miniaturized hydrogen safety sensor prototype. Sens. Actuators B Chem. 2010, 148, 469–477. [Google Scholar] [CrossRef]
  34. Anggraini, S.A.; Breedon, M.; Miura, N. Sensing characteristics of aged zirconia-based hydrogen sensor utilizing Zn–Ta-based oxide sensing-electrode. Electrochem. Commun. 2013, 31, 133–136. [Google Scholar] [CrossRef]
  35. Tang, Z.; Li, X.; Yang, J.; Yu, J.; Wang, J.; Tang, Z. Mixed potential hydrogen sensor using ZnWO4 sensing electrode. Sens. Actuators B Chem. 2014, 195, 520–525. [Google Scholar] [CrossRef]
  36. Li, Y.; Li, X.; Tang, Z.; Tang, Z.; Yu, J.; Wang, J. Hydrogen sensing of the mixed-potential-type MnWO4/YSZ/Pt sensor. Sens. Actuators B Chem. 2015, 206, 176–180. [Google Scholar] [CrossRef]
  37. Li, Y.; Li, X.; Tang, Z.; Wang, J.; Yu, J.; Tang, Z. Potentiometric hydrogen sensors based on yttria-stabilized zirconia electrolyte (YSZ) and CdWO4 interface. Sens. Actuators B Chem. 2016, 223, 365–371. [Google Scholar] [CrossRef]
  38. Li, Y.; Li, X.; Wang, J.; Jun, Y.; Tang, Z. A cobalt tungstate compound sensing electrode for hydrogen detection based upon mixed-potential type sensors. RSC Adv. 2017, 7, 2919–2925. [Google Scholar] [CrossRef]
  39. Darmadi, I.; Nugroho, F.A.A.; Langhammer, C. High-Performance Nanostructured Palladium-Based Hydrogen Sensors—Current Limitations and Strategies for Their Mitigation. ACS Sens. 2020, 5, 3306–3327. [Google Scholar] [CrossRef]
Figure 1. XRD patterns for PCO and PCO-Fe2O3.
Figure 1. XRD patterns for PCO and PCO-Fe2O3.
Chemosensors 11 00250 g001
Figure 2. XPS spectra for PCO-Fe2O3. (a) Survey spectrum; (b) Fe 2p for PCO-Fe2O3.
Figure 2. XPS spectra for PCO-Fe2O3. (a) Survey spectrum; (b) Fe 2p for PCO-Fe2O3.
Chemosensors 11 00250 g002
Figure 3. Surface SEM pictures for (a) PCO and (b) PCO-Fe2O3, (cf) EDS spectrum of PCO-Fe2O3.
Figure 3. Surface SEM pictures for (a) PCO and (b) PCO-Fe2O3, (cf) EDS spectrum of PCO-Fe2O3.
Chemosensors 11 00250 g003
Figure 4. Dynamic response of sensors to H2 at (a) 450 °C and (b) 500 °C.
Figure 4. Dynamic response of sensors to H2 at (a) 450 °C and (b) 500 °C.
Chemosensors 11 00250 g004
Figure 5. The function between the H2 concentration and the response value of PCO-Fe2O3 and PCO sensors at (a) 450 °C and (b) 500 °C. The insets display the function between the logarithmic H2 concentration and the response value of sensors.
Figure 5. The function between the H2 concentration and the response value of PCO-Fe2O3 and PCO sensors at (a) 450 °C and (b) 500 °C. The insets display the function between the logarithmic H2 concentration and the response value of sensors.
Chemosensors 11 00250 g005
Figure 6. The relationship between (a) response time and (b) recovery time with the concentrations of H2 for a PCO-Fe2O3 sensor at 450 °C and 500 °C. The inset in (a) illustrates the response time as a function of the H2 concentration on the log-log scale.
Figure 6. The relationship between (a) response time and (b) recovery time with the concentrations of H2 for a PCO-Fe2O3 sensor at 450 °C and 500 °C. The inset in (a) illustrates the response time as a function of the H2 concentration on the log-log scale.
Chemosensors 11 00250 g006
Figure 7. Comparison of response values of PCO-Fe2O3 sensors for different gases. Concentration: 100 ppm; temperature: 500 °C.
Figure 7. Comparison of response values of PCO-Fe2O3 sensors for different gases. Concentration: 100 ppm; temperature: 500 °C.
Chemosensors 11 00250 g007
Figure 8. (a,b) Effect of humidity on (a) baseline and (b) response; (c) continuous dynamic response curves; and (d) stability test for PCO-Fe2O3 sensor to 300 ppm H2 at 500 °C.
Figure 8. (a,b) Effect of humidity on (a) baseline and (b) response; (c) continuous dynamic response curves; and (d) stability test for PCO-Fe2O3 sensor to 300 ppm H2 at 500 °C.
Chemosensors 11 00250 g008
Figure 9. (a) Impedance spectra tested in different atmospheres for symmetrical PCO-Fe2O3 at 500 °C. (b) Relationship between electrochemical activity ratio (Rc,1/Ra,2) and response value.
Figure 9. (a) Impedance spectra tested in different atmospheres for symmetrical PCO-Fe2O3 at 500 °C. (b) Relationship between electrochemical activity ratio (Rc,1/Ra,2) and response value.
Chemosensors 11 00250 g009
Figure 10. DRT analysis of PCO-Fe2O3 at 500 °C. (a) DRT plots; (b) variation of DRT peak area with H2 concentration, where the inset shows zoom-in of the HF curve.
Figure 10. DRT analysis of PCO-Fe2O3 at 500 °C. (a) DRT plots; (b) variation of DRT peak area with H2 concentration, where the inset shows zoom-in of the HF curve.
Chemosensors 11 00250 g010
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

Wang, L.; Yi, J. Enhancing the Potentiometric H2 Sensing of Pr0.1Ce0.9O2−δ Using Fe2O3 Surface Modification. Chemosensors 2023, 11, 250. https://doi.org/10.3390/chemosensors11040250

AMA Style

Wang L, Yi J. Enhancing the Potentiometric H2 Sensing of Pr0.1Ce0.9O2−δ Using Fe2O3 Surface Modification. Chemosensors. 2023; 11(4):250. https://doi.org/10.3390/chemosensors11040250

Chicago/Turabian Style

Wang, Liang, and Jianxin Yi. 2023. "Enhancing the Potentiometric H2 Sensing of Pr0.1Ce0.9O2−δ Using Fe2O3 Surface Modification" Chemosensors 11, no. 4: 250. https://doi.org/10.3390/chemosensors11040250

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