Next Article in Journal
Molecular Recognition and Shape Studies of 3- and 4-Substituted Diarylamide Quasiracemates
Next Article in Special Issue
Optimal Polyethyleneimine Molecular Weight and Arrangement for Modification of γ-Cyclodextrin Metal Organic Frameworks (γ-CD-MOFs) for Post-Combustion CO2 Capture
Previous Article in Journal
Structural, Optical, Electrical and Antibacterial Properties of Fe-Doped CeO2 Nanoparticles
Previous Article in Special Issue
A Scalable Prototype by In Situ Polymerization of Biodegradables, Cross-Linked Molecular Mode of Vapor Transport, and Metal Ion Rejection for Solar-Driven Seawater Desalination
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanocrystalline Surface Layer of WO3 for Enhanced Proton Transport during Fuel Cell Operation

1
School of Electronic Engineering, Nanjing Xiaozhuang University, Nanjing 211171, China
2
Zhijiang College, Zhejiang University of Technology, Shaoxing 312030, China
3
Huadong Engineering Corporation Limited, Hangzhou 311122, China
4
Jiangsu Provincial Key Laboratory of Solar Energy Science and Technology, Energy Storage Joint Research Center, School of Energy and Environment, Southeast University, Nanjing 210096, China
5
School of Materials Science and Engineering, Hubei University, Wuhan 430062, China
6
New Energy Technologies Group, Department of Applied Physics, Aalto University School of Science, Aalto, FI-00076 Espoo, Finland
7
Hubei Collaborative Innovation Center for Advanced Organic Chemical Materials, Faculty of Physics and Electronic Science, Hubei University, Wuhan 430062, China
*
Authors to whom correspondence should be addressed.
Submission received: 27 November 2021 / Revised: 15 December 2021 / Accepted: 18 December 2021 / Published: 20 December 2021

Abstract

:
High ionic conductivity in low-cost semiconductor oxides is essential to develop electrochemical energy devices for practical applications. These materials exhibit fast protonic or oxygen-ion transport in oxide materials by structural doping, but their application to solid oxide fuel cells (SOFCs) has remained a significant challenge. In this work, we have successfully synthesized nanostructured monoclinic WO3 through three steps: co-precipitation, hydrothermal, and dry freezing methods. The resulting WO3 exhibited good ionic conductivity of 6.12 × 10−2 S cm−1 and reached an excellent power density of 418 mW cm−2 at 550 °C using as an electrolyte in SOFC. To achieve such a high ionic conductivity and fuel cell performance without any doping contents was surprising, as there should not be any possibility of oxygen vacancies through the bulk structure for the ionic transport. Therefore, laterally we found that the surface layer of WO3 is reduced to oxygen-deficient when exposed to a reducing atmosphere and form WO3−δ/WO3 heterostructure, which reveals a unique ionic transport mechanism. Different microscopic and spectroscopic methods such as HR-TEM, SEM, EIS, Raman, UV-visible, XPS, and ESR spectroscopy were applied to investigate the structural, morphological, and electrochemical properties of WO3 electrolyte. The structural stability of the WO3 is explained by less dispersion between the valence and conduction bands of WO3−δ/WO3, which in turn could prevent current leakage in the fuel cell that is essential to reach high performance. This work provides some new insights for designing high-ion conducting electrolyte materials for energy storage and conversion devices.

1. Introduction

Fuel cells (FCs) provide a clean and efficient way to generate electricity from H2 and hydrocarbon fuels. Due to high operating temperatures, solid oxide fuel cells (SOFCs) can also be used in combined heat and power applications. One of the main challenges for SOFCs is developing high oxygen ion (O2−) conductivity of electrolyte materials. Structural doping has been remained a general methodology for developing high ionic conductivities [1,2]. In this methodology, the host cations could often be replaced by a lower valence state, which produces an oxygen-deficient structure to conduct O2− (For example, Zr4+ or Ce4+ are replaced with Y3+ and Sm3+) [3,4]. However, this approach does not significantly enhance fuel cell performance at low operating temperatures due to limited ionic conductivity [4].
Protonic conducting fuel cells (PCFCs), a sub-class of SOFCs, have emerged with great potential for lowering the operating temperature to the range of 400–700 °C. PCFCs can achieve higher fuel utilization by producing the water on the cathode side and avoiding fuel dilution effectively [5]. Another advantage of PCFCs is that many proton-conducting electrolytes such as BaZr0.8Y0.2O3−δ, BaZr0.1Ce0.7Y0.2O3−δ, and Yb-doped BaZr0.1Ce0.7Y0.2O3−δ offer high protonic conductivity at an intermediate operating temperature of 600–800 °C [6,7,8]. Further reduction of the operating temperature, (e.g., 400–600 °C) reduces the power output sharply. If PCFCs could operate in the range of 400–600 °C, it would help prolong the lifetime. The main critical obstacle to achieving high-performance low-temperature PCFCs is the limited proton conductivity of the electrolyte and cathode materials for oxygen reduction reaction (ORR), both of them perform poorly at low temperatures [9,10,11,12,13,14].
However, obtaining the benchmark in proton conductivity (H+) for electrolytes and their chemical stability is a big challenge. The best proton-conducting electrolytes, such as doped barium cerates (BaCeO3−δ), lead to poor CO2 tolerance and rapid decomposing due to forming carbonates at a temperature higher than 600 °C, which is a significant limitation for their use in the device [15]. Recently, next-generation proton-conducting electrolytes such as MTO4, where M = lanthanide/alkali metals and T = Nb, W, Mo, Mn, shows high CO2 tolerance [16,17]. Moreover, the semiconductor-based electrolyte has been reported to deliver the best performance at a low operating temperature [10,11,12,13,14]. The doped BaCeO3−δ only exhibits dominant H+ conductivity under wet conditions or reduced atmosphere and shows significant p-type conductors under oxidizing conditions [15]. WO4/WO3 based materials almost remain pure H+ conductors in both wet and oxidizing conditions, although they exhibit moderate H+ conductivity. However, WO3 to be a direct bandgap semiconductor with less disseminative valence and conduction bands has shown considerable interest for multiple applications in energy devices. WO4/WO3− type oxides exhibit a high oxide ion conduction, e.g., Pb0.9Sm0.1WO4+δ shows a conductivity of ∼2 × 10−2 S cm−1 at 800 °C, which is comparable to that of YSZ (3.6 × 10−2 S cm−1 at 800 °C) [18].
Controlling grain boundary conduction (GBs) into nanocrystalline material is an emerging field. Therefore, we have synthesized nanocrystalline WO3 in three steps to obtain the fine morphology to effectively modulate its electrical properties. We have synthesized nano-structure WO3 by combining three steps following one by one, such as (i) co-precipitation, (ii) hydrothermal, and (iii) dry freezing method. The structural, chemical and morphological analysis of prepared WO3 powders is analyzed. However, in situ formations of WO3−δ/WO3 heterostructure when exposed to H2- side in the fuel cell that could help migrate ions accompanied with the creation of the internal cavity (charge accumulation and depletion region) could be explained by the “surface locking” effect. The prepared WO3 spontaneously facilitated ionic transport exhibiting high ionic conductivity of 6.12 × 10−2 S cm−1 and excellent power density of 478 mW cm−2 at 550 °C. The results demonstrate that the approach proposed is helpful for developing new materials with unique functionalities for advanced PCFCs/SOFCs.

2. Materials and Methods

2.1. Material Preparation Methods

The powders of WO3 were prepared in three steps. Initially, an appropriate amount of ammonium tungstate with a chemical formula of (NH4)10H2(W2O7)6 was bought from Shanghai Macklin Biochemical Co., Ltd., Shanghai, China (purity 98.5%) was dissolved into deionized water to prepare the 0.5 mol/L solutions. In parallel, the 1 mol/L of Na2CO3 was prepared into 100 mL water; afterward, the Na2CO3 solution was added dropwise to the above solution under continuous stirring resulted in white milky precipitates. The obtained precipitates were moved to a Teflon autoclave bottle for hydrothermal treatment at 180°C for 6 h at 180 in a vacuum oven. Afterward, the precipitates were collected to wash and filter numerous times with distilled water and absolute ethanol to eliminate surface-adsorbed water. The resultant precursors were dry freezes at −40 °C for 4 h. Furthermore, it was followed by a vacuum at 1.0 P to cool down to room temperature to remove the absorbed water. Moreover, obtained powders were calcined at 750 °C for 6 h to obtain WO3 nanoparticles. Furthermore, tungsten trioxide (WO3) purchased from Shanghai Macklin Biochemical Co., Ltd., Shanghai, China (99.8%) was used for comparative study to synthesize WO3.

2.2. Material Characterizations and Electrochemical Measurements

The Bruker D8 advanced X-ray diffractometer (Bruker, Hanau, Germany, with Cu Kα radiation, λ = 1.5418 Å) was employed to study the crystalline structure of the synthesized WO3 powders in 2θ ranges from 10–70°. Furthermore, crystal structure, microstructure, and chemical composition of prepared WO3 were studied via FEI Tecnai GI F30 and JEOL JSM7100F (resolution transmission electron microscopy; HR-TEM and field emission scanning electron microscope). X-ray photoelectron spectroscopy (XPS, Physical Electronics Quantum 2000, Al Kα X-ray source) was engaged for chemical oxidation states where the raw data was analyzed through CASA XPS software. Raman spectra were carried out on NT-MDT (Russia) Raman spectrometer at room temperature, with 532 nm solid-state laser as the excitation source and laser power of 20 mW. The UV-Vis 3600 spectrophotometer was used to measure the absorbance spectrum. The electron spin resonance (ESR) measurements were performed using JES X320 (JEOL). Keithley 2400 source meter was used to measure dc-four probed method conductivity.

2.3. Fuel Cells Fabrication and Electrochemical Measurements

We utilized the dry-pressing method to fabricate the SOFC cell with dense electrolyte and porous electrodes. The WO3 powder was pressed between the symmetrical electrodes of Ni0.8Co0.15Al0.05LiO2−δ (NCAL) to form the electrolyte membranes as thin as 285 µm using low filling density powders. In detail used as an electrode was purchased from Bamo Sci. & Tech. Joint Stock Company Ltd., Tianjin, China. In detail, a slurry of NCAL was prepared and painted on porous Ni-foam followed by drying at 120 °C for 2 h. The synthesized and commercially purchased WO3 powders were pressed between the prepared NCAL electrodes under 240 MPa to produce the solid cells. The designed fuel cell pellet’s diameter, thickness, and active area were about 13 mm, 1.5 mm, and 0.64 cm2, respectively. DC electronic load instrument (ITECH8511, ITECH Electrical Co., Ltd., New Taipei, Taiwan) was employed to determine the fuel cell performance of fabricated cells under the H2 as fuel and air as an oxidant with a flow rate of 100–110 mL/min and 100 mL/min, respectively. The I-V (current density-voltage) and I-P (current density–power density) were recorded to present the electrochemical characteristics of the prepared FC devices. Electrochemical impedance spectroscopy (EIS) was measured by Gamry Reference 3000, USA workstation under the fuel cell open-circuit voltage (OCV) conditions in the frequency range of 0.1 to 106 Hz with 10 mV of dc signal. ZSIMPWIN software was used to fit the model circuit with obtained EIS data.

3. Results & Discussion

3.1. Structural and Compositional Study

Figure 1a shows the XRD pattern of WO3, whose main diffraction peaks are located at 2θ° of the 22.8, 23, 24, 26, 28, 33, 33.5, 34.5, 41, 50.5, 56, which are corresponding to (002), (020), (200), (120), (112), (202), (022), (220), (222), (232) and (114) planes of monoclinic crystal structure of WO3 (JCPDS no 43–1035), with space group Pi (C~), and lattice of a = 7.309, b = 7.522, c = 7.678, α = 88.81, β = 90.92, γ = 90.93. There are no extra peaks observed in the patterns, eliminating the possibility of additional phase formation except the monoclinic WO3 phase. To verify crystallography of synthesized WO3, Nano-STAR (Bruker-AXS was used to measure the wide-angle scattering. Figure 1b,c shows a wide-angle plot in the 2θ range from 2°–45° and its corresponding 3-D mapping image, respectively. The wide-angle confirms the high purity monoclinic crystal structure of WO3 measured by XRD. Moreover, the crystallographic structure of WO3 is confirmed by HR-TEM, as shown in Figure 1d,e. The d-spacing values calculated using a digital micrograph are 0.236 and 0.213 nm, which can be nominated to (200) and plane (020) planes, respectively. These planes could also be confirmed by a selected area electron diffraction (SAED) pattern, as shown in Figure 1f [19,20].
Figure 2 shows nano-scaled HR-TEM and s STEM images for synthesized WO3, where sophisticated and fine nano-particles with a particle size of <50 nm can be seen clearly. The restrain of agglomerates and the growth of particles in WO3 can be attributed to the purification of the synthesis method we have used. Moreover, a line scan using high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) was used to confirm the elemental distribution at the particle level. Figure 2c shows a HAADF-STEM image with elements mapping of W and O, where a homogenous chemical distribution can be observed. Figure 2d shows the chemical distribution along with the line scan compared to the HAADF image in the 0–100 nm range. Figure 3f shows the energy dispersive spectroscopy (EDS) image measured using the HAADF-STEM image. The results mentioned above manifest that the particles are at nano-level and are well connected, which leads to enhancement of the surface area with more active sties to enhance the overall performance of fuel cell devices [21,22].

3.2. Electrochemical Performance Measurements

The electrochemical performance of the prepared WO3 in comparison to commercially purchased WO3 in a fuel cell as an ion-conducting electrolyte was evaluated. Figure 3a displays typical current-voltage (I–V) and corresponding power density (I-P) characteristics of fuel cells using commercial WO3 and our synthesized WO3 material as an electrolyte at 450–550 °C. The OCV of 1.03 V and the maximum power density (Pmax) of 418 ± 2% mW cm−2 were achieved for our prepared WO3 electrolyte, while commercially purchased WO3 exhibited (Pmax) of only 260 mW cm−2. The OCV (e.g., 1.03 V at 550 °C) agrees with the Nernst theoretical potential at this temperature, showing that the ionic transfer number is near unity and WO3 acts electronically insulating [23,24,25]. Moreover, our synthesized WO3 electrolyte fuel cell displayed good electrochemical performance even at low temperatures, as shown in Figure 3b. It exhibited peak power densities of 360 and 310 mW cm−2 at 500 °C and 450 °C, respectively. The enhanced electrochemical performance of synthesized WO3 over commercially purchased WO3 suggests the vital role of synthesis method and nanostructured morphology, where the surface layer of WO3 can easily be reduced and produce the strong WO3−δ/WO3 heterostructure for creating the surface path for easy transport of H+ when it’s exposed to H2 in fuel cell conditions [20]. During the online electrochemical process, the prepared WO3 material could facilitate the better intercalation of protons because of the high surface-active area and easy reduction of the surface layer. Therefore, obtained results suggested that the nanostructured stoichiometry synthesized by three steps synthesis could be helpful to control the surface structural properties. In this way, the WO3 surface kinetically favors the heterolytic dissociation of H2 to form W–H and O–H species. The resulting W–H further evolves to the thermodynamically more stable O–H species, accompanied by the reduction of W6+ to W5+ into the surface layer.
Additionally, Figure 3c shows a typical cross-sectional scanning electron microscopy (SEM) image of the fuel cell with WO3 acquired after online sintering and test. The SEM image indicates that the WO3 electrolyte layer appears fully dense, without noticeable connected pores, and appears well-adhered to the anode substrate, without any cracking or delamination after fuel cell testing as compared with before testing (Figure 3d,f). Such adhesive and dense structure of WO3 electrolyte guarantee better fuel cell performance [23,26].

3.3. Electrochemical Impedance and Electrical Conductivity

Further, EIS characterizations were performed for the cell with synthesized WO3 electrolyte in both air and H2/air atmosphere at 450–550 °C under OCV conditions. Figure 3a,b shows Nyquist curve of the measured EIS spectra and Fitted data using ZSIMPWIN software with equivalent circuit modeling of Ro-(R1-CPE1) − (R2-CPE2), where Ro, is the ohmic resistance from the electrolyte, R1 and R2 belong to charge transfer and mass transfer losses of the fuel cell with WO3 electrolyte, respectively. It can be seen from the EIS spectra of Figure 4a,b that Ro of the fuel cell with WO3 dramatically reduced in H2/air (fuel cell operating conditions) as compared to in the air. For example, fuel cell with WO3 electrolyte exhibited Ro of only 1.09 Ω cm2 in H2/air at 550 °C, while in the air it shows 12.85 Ω cm2 Similarly, a clear difference in Ro values at a low operating temperature of 500 and 450 °C also can be visualized. These results describe that when WO3 is exposed to the H2 atmosphere, its surface could be reduced to W5+ to form WO3−δ/WO3, and protons can easily be transported through this layer as reported for CeO2. Afterward, it breaks the accumulation layer at the interface of electrode/electrolyte and hence reduces the charge transfer resistance (R1), mass transport resistance (R2) of fuel cells, as shown in Figure 4a,b. The decrease in charge transfer resistance at the interface of electrolyte/electrode for each cell may be followed by value capacitance for that cell (i.e., R1∼C1, R2∼C2), where capacitance can be determined by C i = R i Q i 1 n R i , where R is corresponding resistance and n to Frequency power [0 ≥ n ≤1] of the Q’s values [11]. The cell’s charge transfer resistance in H2/air compared to EIS spectra in air decreases from 10.5 to 0.17 Ω cm2, followed by a decline in space charge capacitance from 1.231 × 10−3 F to 1.713 × 10−5 F at 550 °C [24,25,27].
The ionic conductivity of synthesized WO3 in both air and H2/air was calculated using Ro values of the fitted data. As shown in Figure 4c, the ionic conductivity of synthesized WO3 in H2/air is much higher than only in air atmospheres. The synthesized WO3 exhibits the ionic conductivity of 6.12 × 10−2 S cm−1 in H2, while in the air, only 1.21 × 10−3 S cm−1. The conductivity results support our findings that the WO3 surface could be reduced to deficient WO3−δ and support fast protonic transport. Moreover, cell ionic conductivity was measured over 24 h to check its behavior, and it shows the stable ionic conductivity as can be seen in Figure 4d. Furthermore, the proton conductivity of WO3 was measured using Ag current collectors in H2 without using the NCAL electrode. In these conditions, WO3 shows a little lower proton conductivity as compared to fuel cell conditions (Figure 4c). However, these results indicate that WO3 synthesized by three-step methods could be a good candidate for electrolyte application in advanced SOFCs [20].

3.4. Spectroscopic Analysis

Moreover, different spectroscopic techniques, such as Raman, UV-visible, and X-ray photoelectron spectroscopies, were employed to study further structural properties of WO3 powders before and after fuel cell measurements. Figure 5a displays specific Raman shifts bands for synthesized WO3 centered at 266, 322, 709, and 803 cm−1 links to the stretching vibrations of (O-W-O), (O-W-O), (W2O6 & W3O8), and (W-O-W) from 200–2400 Raman shifts/cm−1. After performing fuel cell measurements, a small redshift of 0.5 cm−1 in the Raman shift band of WO3 was observed. However, overall, WO3 shows good structure stability after fuel cell measurements. Moreover, UV-Visible absorbance spectra of WO3 for before and after fuel cell performance measurements are presented in Figure 5b. There is just a little bit of difference in absorbance spectra that can be observed. The difference in absorbance spectra is an obvious indicator for lower down in the energy band gap of WO3, which only could be due to the reducing the surface and producing oxygen vacancies. The low energy gap will help to improve the ionic conductivity and, hence, better fuel cell performance [21,24,25,28].
High-resolution XPS spectra of as-synthesized WO3 powders are shown in Figure 6a,b. After subtracting Shirley’s background, high-resolution XPS spectra were fitted by the mixture function of Lorentzian and Gaussian. Our focus was to observe the chemical and electronic state configuration changes of W-4f and O1s spectra before and after electrochemical performance measurements. Figure 6a shows the XPS spectra of 4f-W6+ (5/2, 7/2) and 4-f W5+ (5/2, 7/2) that appear at 35.32/37.52 and 35.7/38.05 eV, whereas in after fuel cell performance measurements at 35.12/37.22 and 35.9/37.85 eV. These downshifts in B.E of W6+ 4f (5/2, 7/2) and W5 + 4f (5/2, 7/2) manifest a reduction in WO3 after fuel cell measurements. Moreover, O1s spectra of the material also influence the ionic conductivity of a material [29,30]. The O1s spectra after fuel cell measurements contain lattice oxygen (lattice O2−) and oxygen vacancy peaks, as shown in Figure 6b. The O1s spectra of WO3 after fuel cell measurements display two partially superimposed peaks (Figure 6b). There are two significant excitations: the first includes O1s of lattice oxygen bands and the second, WO3−δ band with binding energy (BE) ranging from 528 to 533.5 eV [28,29]. The low BE peak at 529.2 can be ascribed to the lattice oxygen (O Lattice), higher at 531.4- to extra oxygen vacancies. The increased area percentage ratio of Olat/Ovac of WO3 after fuel cell measurements indicates high oxygen vacancies concentration, which plays an essential role in high fuel cell performance [21,28]. However, these results provide clear evidence that an in situ surface reduction produces oxygen vacancies for the fast protons transports, as shown in Figure 6c. Therefore, our developed strategy could help create high-performance LT-SOFCs electrolyte materials in a new way [31]. The XPS data shows that W4f and O1s percentage is different (W4f−18.37; O1s 56.65%) as compared to as synthesized sample (W4f−20.5; O1s 54.26). It means the after the fuel cell performance measurements the WO3 has more oxygen vacancies than as pristine WO3. Furthermore, electron spin resistance (ESR) configures the change in structural properties after fuel cell measurements. Figure 6c shows the full spectrum of proton unirradiated WO3 and proton irradiated WO3 phases. Where proton irradiation-induced defect phase lines and the well-resolved group of the ESR spectrum can be seen after fuel cell measurements. However, ESR results confirm our proposed mechanism that the surface layer of WO3 is reduced during the in-situ operation of the fuel cell, and it facilitates proton transport.
The schematic diagram of the process involved for pristine WO3 to deficient WO3−δ layer formation for migration of oxygen and proton ions is shown in Figure 7. The proton transport is accompanied by a surface layer of WO3/WO3−δ as shown in the last step of Figure 7.

4. Conclusions

In summary, we have successfully synthesized and characterized WO3 nanostructured material by combining three synthesis methods. Furthermore, the synthesized nanostructured WO3 with a monoclinic structure demonstrated excellent proton conductivity during fuel cell operation. The ionic conductivity reached 6.12 × 10−2 S cm−1 in an H2 atmosphere as compared to 1.21 × 10−3 S cm−1 in the air. The fabricated fuel cell using prepared WO3 as elect4rolyte exhibited a high-power density of 418 mW cm−2 at 550 °C. Furthermore, we used different microscopic and spectroscopic analyses to study the mechanism behind the drastic increase in ionic conductivity of WO3 during fuel cell operation. We found that synthesized nanostructured WO3 surfaces can easily be reduced to form an oxygen-deficient layer and facilitate protons transport effectively when exposed to a reduction in the atmosphere. The ex-situ spectroscopies that included Raman, UV-visible, XPS, and ESR clearly described our findings and the structural change properties of WO3 during the fuel cell operation. In conclusion, this method could form the basis of interest to develop new WO3 based proton-conducting electrolytes, which could be useful for all energy devices and material systems.

Author Contributions

N.M. provided the concept of this study; X.S. performed the experiments and wrote the first draft; W.G. and Y.G. contributed to the data curation; M.A.K.Y.S. and M.S.I. conducted the formal analysis of the results and the manuscript; P.D.L. and M.I.A. helped to revise the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

School of Electronic Engineering, Nanjing Xiaozhuang University, 211171 Nanjing, China for providing experimental facilities. Further, this work was supported by National Natural Science Foundation of China (NSFC) under the grant #51772080 and 11604088 and Southeast University (SEU PROJET # 3203002003A1). Dr. Asghar thanks the Hubei Talent 100 program and Academy of Finland (grant no. 13329016 and 13322738) for their support.

Acknowledgments

Authors Acknowledged School of Electronic Engineering, Nanjing Xiaozhuang University, 211171 Nanjing, China for providing experimental facilities and publication charges.

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, C.; Kee, R.J.; Zhu, H.; Karakaya, C.; Chen, Y.; Ricote, S.; Jarry, A.; Crumlin, E.J.; Hook, D.; Braun, R.; et al. Highly durable, coking and sulfur tolerant, fuel-flexible protonic ceramic fuel cells. Nature 2018, 557, 217–222. [Google Scholar] [CrossRef] [PubMed]
  2. Choi, S.; Kucharczyk, C.J.; Liang, Y.; Zhang, X.; Takeuchi, I.; Ji, H.-I.; Haile, S.M. Exceptional power density and stability at intermediate temperatures in protonic ceramic fuel cells. Nat. Energy 2018, 3, 202–210. [Google Scholar] [CrossRef] [Green Version]
  3. Malavasi, L.; Fisher, C.A.J.; Islam, M.S. Oxide-ion and proton conducting electrolyte materials for clean energy applications: Structural and mechanistic features. Chem. Soc. Rev. 2010, 39, 4370–4387. [Google Scholar] [CrossRef]
  4. Goodenough, J.B. Oxide-Ion Conductors by Design. Nature 1999, 404, 821–823. [Google Scholar] [CrossRef]
  5. Zhang, L.; Chae, S.-R.; Hendren, Z.; Park, J.-S.; Wiesner, M.R. Recent advances in proton exchange membranes for fuel cell applications. Chem. Eng. J. 2012, 204–206, 87–97. [Google Scholar] [CrossRef]
  6. Bi, L.; Da’As, E.H.; Shafi, S.P. Proton-conducting solid oxide fuel cell (SOFC) with Y-doped BaZrO3 electrolyte. Electrochem. Commun. 2017, 80, 20–23. [Google Scholar] [CrossRef]
  7. Hakim, M.; Yoo, C.-Y.; Joo, J.H.; Yu, J.H. Enhanced durability of a proton conducting oxide fuel cell with a purified yttrium-doped barium zirconate-cerate electrolyte. J. Power Sources 2015, 278, 320–324. [Google Scholar] [CrossRef]
  8. Muccillo, R.; Muccillo, E.N. Synthesis and Properties of BaZr0.1Ce0.7Y02-xMxO3-δ (x = 0, 0.1; M = Dy, Yb) Compounds. ECS Trans. 2011, 35, 1251. [Google Scholar] [CrossRef]
  9. Liu, Z.; Zhou, M.; Chen, M.; Cao, D.; Shao, J.; Liu, M.; Liu, J. A high-performance intermediate-to-low temperature protonic ceramic fuel cell with in-situ exsolved nickel nanoparticles in the anode. Ceram. Int. 2020, 46, 19952–19959. [Google Scholar] [CrossRef]
  10. Shah, M.Y.; Mushtaq, N.; Rauf, S.; Akbar, N.; Xing, Y.; Wu, Y.; Wang, B.; Zhu, B. Advanced fuel cell based on semiconductor perovskite La–BaZrYO3-δ as an electrolyte material operating at low temperature 550 °C. Int. J. Hydrog. Energy 2020, 45, 27501–27509. [Google Scholar] [CrossRef]
  11. Shah, M.A.K.Y.; Rauf, S.; Mushtaq, N.; Zhu, B.; Tayyab, Z.; Yousaf, M.; Hanif, M.B.; Lund, P.D.; Lu, Y.; Asghar, M.I. Novel Perovskite Semiconductor Based on Co/Fe-Codoped LBZY (La0.5Ba0.5Co0.2Fe0.2Zr0.3Y0.3O3−δ) as an Electrolyte in Ceramic Fuel Cells. ACS Appl. Energy Mater. 2021, 4, 5798–5808. [Google Scholar] [CrossRef]
  12. Lu, Y.; Mi, Y.; Li, J.; Qi, F.; Yan, S.; Dong, W. Recent Progress in Semiconductor-Ionic Conductor Nanomaterial as a Membrane for Low-Temperature Solid Oxide Fuel Cells. Nanomaterials 2021, 11, 2290. [Google Scholar] [CrossRef]
  13. Liu, Y.; Xia, C.; Wang, B.; Tang, Y. Layered LiCoO2–LiFeO2 Heterostructure Composite for Semiconductor-Based Fuel Cells. Nanomaterials 2021, 11, 1224. [Google Scholar] [CrossRef]
  14. Xu, D.; Yan, A.; Xu, S.; Zhou, Y.; Yang, S.; Zhang, R.; Yang, X.; Lu, Y. Self-Assembled Triple (H+/O2/e) Conducting Nanocomposite of Ba-Co-Ce-YO into an Electrolyte for Semiconductor Ionic Fuel Cells. Nanomaterials. 2021, 11, 2365. [Google Scholar] [CrossRef] [PubMed]
  15. Sailaja, J.M.; Murali, N.; Margarette, S.; Jyothi, N.K.; Rajkumar, K.; Veeraiah, V. Chemically stable proton conducting doped BaCeO3 by citrate-EDTA complexing sol-gel process for solid oxide fuel cell. S. Afr. J. Chem. Eng. 2018, 26, 61–69. [Google Scholar]
  16. Fabbri, E.; Bi, L.; Pergolesi, D.; Traversa, E. Towards the next generation of solid oxide fuel cells operating below 600 °C with chemically stable proton-conducting electrolytes. Adv. Mater. 2012, 24, 195–208. [Google Scholar] [CrossRef] [PubMed]
  17. Stevens, J.; Wieczorek, W.; Raducha, D.; Jeffrey, K. Proton conducting gel/H3PO4 electrolytes. Solid State Ion. 1997, 97, 347–358. [Google Scholar] [CrossRef]
  18. Yashima, M.; Tsujiguchi, T.; Sakuda, Y.; Yasui, Y.; Zhou, Y.; Fujii, K.; Torii, S.; Kamiyama, T.; Skinner, S.J. High Oxide-Ion Conductivity through the Interstitial Oxygen Site in Ba7Nb4MoO20-Based Hexagonal Perovskite Related Oxides. Nat. Commun. 2021, 12, 556. [Google Scholar] [CrossRef] [PubMed]
  19. Wang, B.B.; Zhong, X.X.; He, C.L.; Zhang, B.; Cvelbar, U.; Ostrikov, K. Solvent-dependent structures and photoluminescence of WO3-x nanomaterials grown in nonaqueous solutions. J. Alloy. Compd. 2021, 854, 157249. [Google Scholar] [CrossRef]
  20. Wang, Z.; Fan, X.; Li, C.; Men, G.; Han, D.; Gu, F. Humidity-sensing performance of 3DOM WO3 with controllable structural modification. ACS Appl. Mater. Interfaces 2018, 10, 3776–3783. [Google Scholar] [CrossRef]
  21. Shah, M.Y.; Rauf, S.; Zhu, B.; Mushtaq, N.; Yousaf, M.; Lund, P.D.; Xia, C.; Asghar, M.I. Semiconductor Nb-Doped SrTiO3 − δ Perovskite Electrolyte for a Ceramic Fuel Cell. ACS Appl. Energy Mater. 2021, 4, 365–375. [Google Scholar] [CrossRef]
  22. Mushtaq, N.; Xia, C.; Dong, W.; Wang, B.; Raza, R.; Ali, A.; Afzal, M.; Zhu, B. Tuning the energy band structure at interfaces of the SrFe0.75Ti0.25O3−δ–Sm0.25Ce0.75O2−δ heterostructure for fast ionic transport. ACS Appl. Mater. Interfaces 2019, 11, 38737–38745. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, B.; Cai, Y.; Xia, C.; Kim, J.S.; Liu, Y.; Dong, W.; Wang, H.; Afzal, M.; Li, J.; Raza, R.; et al. Semiconductor-ionic membrane of LaSrCoFe-oxide-doped ceria solid oxide fuel cells. Electrochim. Acta 2017, 248, 496–504. [Google Scholar] [CrossRef] [Green Version]
  24. Shah, M.A.K.Y.; Zhu, B.; Rauf, S.; Mushtaq, N.; Yousaf, M.; Ali, N.; Tayyab, Z.; Akbar, N.; Yang, C.P.; Wang, B. Electrochemical properties of a co-doped SrSnO3−δ-based semiconductor as an electrolyte for solid oxide fuel cells. ACS Appl. Energy Mater. 2020, 3, 6323–6333. [Google Scholar] [CrossRef]
  25. Mushtaq, N.; Lu, Y.; Xia, C.; Dong, W.; Wang, B.; Shah, M.Y.; Rauf, S.; Akbar, M.; Hu, E.; Raza, R.; et al. Promoted electrocatalytic activity and ionic transport simultaneously in dual functional Ba0.5Sr0.5Fe0.8Sb0.2O3-δ-Sm0.2Ce0. 8O2-δ heterostructure. Appl. Catal. B: Environ. 2021, 298, 120503. [Google Scholar] [CrossRef]
  26. Shah, M.A.K.Y.; Mushtaq, N.; Rauf, S.; Xia, C.; Zhu, B. The semiconductor SrFe0.2Ti0.8O3-δ-ZnO heterostructure electrolyte fuel cells. Int. J. Hydrog. Energy 2019, 44, 30319–30327. [Google Scholar] [CrossRef]
  27. Shah, M.Y.; Tayyab, Z.; Rauf, S.; Yousaf, M.; Mushtaq, N.; Imran, M.A.; Lund, P.D.; Asghar, M.I.; Zhu, B. Interface engineering of bi-layer semiconductor SrCoSnO3-δ-CeO2-δ heterojunction electrolyte for boosting the electrochemical performance of low-temperature ceramic fuel cell. Int. J. Hydrog. Energy 2021, 46, 33969–33977. [Google Scholar] [CrossRef]
  28. Chen, G.; Liu, H.; He, Y.; Zhang, L.; Asghar, M.I.; Geng, S.; Lund, P.D. Electrochemical mechanisms of an advanced low-temperature fuel cell with a SrTiO3 electrolyte. J. Mater. Chem. A 2019, 7, 9638–9645. [Google Scholar] [CrossRef] [Green Version]
  29. Katrib, A.; Hemming, F.; Wehrer, P.; Hilaire, L.; Maire, G. The multi-surface structure and catalytic properties of partially reduced WO3, WO2 and WC + O2 or W + O2 as characterized by XPS. J. Electron Spectrosc. Relat. Phenom. 1995, 76, 195–200. [Google Scholar] [CrossRef]
  30. Efkere, H.I.; Gümrükçü, A.E.; Özen, Y.; Kınacı, B.; Aydın, S. Şebnem; Ates, H.; Özçelik, S. Investigation of the effect of annealing on the structural, morphological and optical properties of RF sputtered WO3 nanostructure. Phys. B Condens. Matter. 2021, 622, 413350. [Google Scholar] [CrossRef]
  31. Wang, B.; Zhu, B.; Yun, S.; Zhang, W.; Xia, C.; Afzal, M.; Cai, Y.; Liu, Y.; Wang, Y.; Wang, H. Fast ionic conduction in semiconductor CeO2−δ electrolyte fuel cells. NPG Asia Mater. 2019, 11, 1–12. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structure characterization (a) X-ray diffraction pattern of synthesized WO3; (b,c) wide-angle X-ray diffraction pattern and its corresponding mapping for the lattice fringes image of WO3; (d,e) crystal structure and d-spacing of WO3 measured by HR-TEM and (f) selected area electron diffraction (SAED) pattern of the synthesized WO3.
Figure 1. Structure characterization (a) X-ray diffraction pattern of synthesized WO3; (b,c) wide-angle X-ray diffraction pattern and its corresponding mapping for the lattice fringes image of WO3; (d,e) crystal structure and d-spacing of WO3 measured by HR-TEM and (f) selected area electron diffraction (SAED) pattern of the synthesized WO3.
Crystals 11 01595 g001
Figure 2. Morphology and compositional characterization (a) HR-TEM image of WO3; (b) STEM image of synthesized WO3; (c) STEM-HAADF elements mapping for W and O; (d) line scan distribution of WO3 corresponding to HAADF; (e) EDS image of WO3 for actual chemical composition study.
Figure 2. Morphology and compositional characterization (a) HR-TEM image of WO3; (b) STEM image of synthesized WO3; (c) STEM-HAADF elements mapping for W and O; (d) line scan distribution of WO3 corresponding to HAADF; (e) EDS image of WO3 for actual chemical composition study.
Crystals 11 01595 g002
Figure 3. Electrochemical performance characterizations: (a) typical current (I)–voltage (V) characteristics curves of using our synthesized material (WO3) and commercially purchased WO3 as an electrolyte in fuel cell operating at 550 °C; (b) fuel cell using our synthesized WO3 as an electrolyte at different operating temperatures of 450–550 °C; (c) tri-layer cross-sectional SEM images of anode supported symmetrical fuel cell with our synthesized WO3 electrolyte after performing the electrochemical test; (d,e) enlarged SEM image of WO3 electrolyte layer, before and after fuel cell testing, respectively.
Figure 3. Electrochemical performance characterizations: (a) typical current (I)–voltage (V) characteristics curves of using our synthesized material (WO3) and commercially purchased WO3 as an electrolyte in fuel cell operating at 550 °C; (b) fuel cell using our synthesized WO3 as an electrolyte at different operating temperatures of 450–550 °C; (c) tri-layer cross-sectional SEM images of anode supported symmetrical fuel cell with our synthesized WO3 electrolyte after performing the electrochemical test; (d,e) enlarged SEM image of WO3 electrolyte layer, before and after fuel cell testing, respectively.
Crystals 11 01595 g003
Figure 4. (a) Impedance spectra for the fuel cell with synthesized WO3 electrolyte layer measured in air at 450–550 °C and the corresponding fitting data; (b) impedance spectra for the fuel cell with synthesized WO3 electrolyte layer measured in H2/air at 450–550 °C; (c) comparison of the ionic conductivity of WO3 measured in air and H2 atmosphere with NCAL and electrodes and (d) stability of ionic conductivity measured over 24 h.
Figure 4. (a) Impedance spectra for the fuel cell with synthesized WO3 electrolyte layer measured in air at 450–550 °C and the corresponding fitting data; (b) impedance spectra for the fuel cell with synthesized WO3 electrolyte layer measured in H2/air at 450–550 °C; (c) comparison of the ionic conductivity of WO3 measured in air and H2 atmosphere with NCAL and electrodes and (d) stability of ionic conductivity measured over 24 h.
Crystals 11 01595 g004
Figure 5. (a,b) Raman and UV-visible spectra of as-synthesized WO3 powders and for WO3 after fuel cell performance measurements.
Figure 5. (a,b) Raman and UV-visible spectra of as-synthesized WO3 powders and for WO3 after fuel cell performance measurements.
Crystals 11 01595 g005
Figure 6. (a) X-ray photoelectron spectra of (a,b) W-4f and O1s spectra of WO3 for as-synthesized WO3 powders and after fuel cell performance measurements; (c) electron spinning resonance (ESR) study of as-synthesized WO3 and after fuel cell measurements.
Figure 6. (a) X-ray photoelectron spectra of (a,b) W-4f and O1s spectra of WO3 for as-synthesized WO3 powders and after fuel cell performance measurements; (c) electron spinning resonance (ESR) study of as-synthesized WO3 and after fuel cell measurements.
Crystals 11 01595 g006
Figure 7. The different processes involved in the phase transition of WO3 to WO3−δ and schematic of our proposed proton transport mechanism in WO3 during fuel cell operation.
Figure 7. The different processes involved in the phase transition of WO3 to WO3−δ and schematic of our proposed proton transport mechanism in WO3 during fuel cell operation.
Crystals 11 01595 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, X.; Guo, W.; Guo, Y.; Mushtaq, N.; Shah, M.A.K.Y.; Irshad, M.S.; Lund, P.D.; Asghar, M.I. Nanocrystalline Surface Layer of WO3 for Enhanced Proton Transport during Fuel Cell Operation. Crystals 2021, 11, 1595. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11121595

AMA Style

Song X, Guo W, Guo Y, Mushtaq N, Shah MAKY, Irshad MS, Lund PD, Asghar MI. Nanocrystalline Surface Layer of WO3 for Enhanced Proton Transport during Fuel Cell Operation. Crystals. 2021; 11(12):1595. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11121595

Chicago/Turabian Style

Song, Xiang, Weiqing Guo, Yuhong Guo, Naveed Mushtaq, M. A. K. Yousaf Shah, Muhammad Sultan Irshad, Peter D. Lund, and Muhammad Imran Asghar. 2021. "Nanocrystalline Surface Layer of WO3 for Enhanced Proton Transport during Fuel Cell Operation" Crystals 11, no. 12: 1595. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11121595

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