Next Article in Journal / Special Issue
Multi-Dimensional Analysis of Load Characteristics of Electrical Vehicles Based on Power Supply Side Data and Unsupervised Learning Method
Previous Article in Journal / Special Issue
Online Parameters Identification and State of Charge Estimation for Lithium-Ion Battery Using Adaptive Cubature Kalman Filter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Constructing Supports–Network with N–TiO2 Nanofibres for Highly Efficient Hydrogen–Production of PEM Electrolyzer

1
Clean Energy Automotive Engineering Center, School of Automotive Studies, Tongji University, Shanghai 201814, China
2
Department of Electrical Engineering, Tongji University, Shanghai 201814, China
*
Author to whom correspondence should be addressed.
World Electr. Veh. J. 2021, 12(3), 124; https://doi.org/10.3390/wevj12030124
Submission received: 13 July 2021 / Revised: 13 August 2021 / Accepted: 16 August 2021 / Published: 17 August 2021

Abstract

:
Hydrogen production with a proton exchange membrane (PEM)electrolyzer utilized with renewable energy power is considered to be an efficient and clean green technique, but the poor oxygen evolution performance results in high energy consumption and low efficiency. In this work, a strategy is reported for the construction of a support network of the anodic catalyst layer to simultaneously ameliorate its sluggish reaction kinetics and mass transport in order to realize highly efficient hydrogen production of the PEM electrolyzer. After in situ synthesis of IrO2 nanoparticles on N–doped TiO2 nanofibers, the as–prepared IrO2/N–TiO2 electrode shows substantially enhanced Ir utilization and accelerated mass transport, consequently decreasing the corresponding cell potential of 107 mV relative to pure IrO2 at 2 A cm−2. The enhanced activity of IrO2/N–TiO2 could be due to the fact that the N–TiO2 nanofiber support can form a porous network, endowing IrO2/N–TiO2 with a large reactive contact interface and favorable mass transfer characters. The strategy in this work supplies a pathway to develop high–efficiency interfacial reaction materials for diverse applications.

1. Introduction

PEM electrolyzer technology is a promising alternative to splitting water into oxygen and clean hydrogen, especially as the intrinsic characteristics of a fast response and fluctuation adaptability of PEM electrolyzer are closely integrated with intermittent energy sources [1]. Unfortunately, the insufficient performance due to sluggish oxygen evolution reaction (OER) kinetics together with depressed mass transport gives rise to excessive consumption of electrical power and a high cost of hydrogen production [2,3]; consequently, research efforts devoted to the highly efficient anodic catalysts of PEM electrolyzer have gained attention. To date, numerous research works have been devoted to enhancing electrocatalytic performance, thus minimizing the overall voltage loss [4,5]. Among the modified methods, an attractive alternative that can be carried out to improve the efficiency of hydrogen production is the retention of Ir–based materials in nanoparticles using a support network; not only could this maximize Ir utilization to increase the number of active sites, but it could also realize fast mass transport via the construction of a mass diffusion pathway [6,7].
One–dimensional (1D) nanostructure materials are known for their unique physical and chemical properties, which have been widely applied in fuel cells, lithium–ion batteries, supercapacitors, etc. [8,9]. When employed as supports for a catalyst, 1D nanostructured materials can offer a large specific surface area and expedite electron (ion) transport, contributing to the high electrocatalytic performance of the catalyst–support system [10,11]. On the other hand, the 1D nanostructure could form a support network via the fabrication of the catalyst layer, bringing about the appropriate reaction and diffusion pathways for reactant intermediates and reduction products [12]. Inspired by the above discussion, a well–designed support network is crucial for a catalyst to display robust reaction kinetics and rapid mass transfer. However, little research has paid attention to developing a 1D nanostructured support network for efficient PEM electrolyzer performance. Moreover, the relationship between the catalyst–support system, mass transport, and reaction kinetics of OER is rarely involved.
In this work, a N–doped TiO2 (N–TiO2) composite nanofiber is developed to construct a support network for IrO2 nanoparticles in order to accelerate the OER kinetics and improve mass transport. IrO2/N–TiO2 is successfully prepared via the solvothermal method followed by heat treatment. SEM and TEM technologies have demonstrated that sub–2 nm IrO2 nanoparticles are well dispersed on N–TiO2 nanofiber support, and the support network of the anodic catalytic layer is effectively constructed. The IrO2/N–TiO2 electrode shows optimized OER activity and low cell potential of 107 mV at 2 A cm−2 when compared to that of pure IrO2. Electrochemical measurements confirm the increased active sites, accelerated OER kinetics, and favorable mass transfer. This could be assigned to the accelerated kinetics and rapid mass transport via the construction of the mass diffusion pathway.

2. Experiment

2.1. Preparation of N–TiO2, IrO2, and IrO2/N–TiO2

The hydrothermal method was employed to synthesize N–TiO2 nanofibers. As shown in Scheme 1, 0.3 gTiO2 (P25) and 0.078 g urea were dispersed into a 70 mL NaOH solution (10 M) and ultrasonicated for 30 min. Then, the suspension liquid was transferred to a hydrothermal reactor processed at 200 ℃ for 18 h. The white precipitate (Na–Ti–N–O) was washed to neutral with deionized water and added to a 500 mL HCl solution (0.1 M) for 24 h. After that, the product was washed to neutral with deionized water again and dried at 60 ℃ to obtain N–TiO2.
IrO2/N–TiO2 was obtained with heat treatment after polyol reduction (Scheme 1). Briefly, 0.256 g of N–TiO2 and 0.47 g of H2IrCl6 6H2O were added to a 100 mL solution of isopropyl alcohol and water (vol% 1:1). The mixed solution was then heated at 80 ℃ for 6 h. After cooling down, the solution was filtrated and dried 60 ℃. IrO2/N–TiO2 was finally achieved after calcining at 500 ℃ for 30 min. IrO2 supported on the N–TiO2 support was controlled at 80 wt.%. The IrO2 sample was obtained in a similar manner to that for IrO2/N–TiO2 but without the addition of N–TiO2.

2.2. Preparation of MEA

Nafion 117 membranes acted as the electrolyte and were pre–processed continuously in H2O2 solution, H2SO4 solution, and deionized water at 90 °C for 1 h. Pt/C and IrO2/N–TiO2, acting as the cathode and anode catalysts, respectively, were sprayed on each side of the N117 membranes to prepare membrane electrode assembly (MEA). The loading mass of Pt and IrO2 were controlled with 0.4 and 1.5 mg/cm2, respectively. Regarding the catalyst ink, a certain amount of the catalyst and ionomer (Nafion solution) was added to the solution of isopropyl alcohol and water (vol% 1:1). The amount of Nafion was 25 wt% for the cathode and anode catalysts. The active area for the prepared MEA was 4 cm2.

2.3. Physicochemical Characterization

An X–ray diffractometer (XRD, Bruker D8 Advance X–ray diffractometer) was used to detect the phase and crystal structure. A scanning electron microscope (SEM, FEI QUANTA250FEG*) and transmission electron microscope (TEM, JEOL JEM 2100F microscope) equipped with energy–dispersive X–ray (EDX) spectroscopy and high–angle annular dark–field imaging (HAADF) were used to record the microtopography features of the samples. A Kruss DSA100 system was employed to record the static contact angles of the prepared anodic catalyst layer by the drop shape method using water as the wetting liquid.

2.4. Electrochemical Analysis

Before electrochemical analysis, a homemade PEM single cell assembled with the prepared MEA was inputted into high–purity water for 24 h at a flow rate of 20 mL/min. Cyclic voltammetry (CV) curves of the prepared MEA were measured from the potential from 0 to 1.23 V with a scan rate of 50 mV/s. Polarization curves of the MEAs were performed on an incremental current density from 0.01 to 2 A cm−2 with a DC supply. Electrochemical impedance spectroscopy (EIS) was tested at 1.5 V at a frequency range of 0.01 to 10 K Hz with an amplitude of 5 mV using a connection of a Solartron 1287 potentiostat and a 1260 frequency response analyzer. The high–frequency resistance values (HFR) of the PEM single cell were taken at the frequency of 10 kHz. All of the electrochemical tests of the PEM single cell were conducted at the conditions of 80 °C and flow rate of 20 mL/min.

3. Results

3.1. Materials Characterization

XRD patterns (Figure 1) were obtained to analyze the crystal structure of the prepared samples. For IrO2, the peaks located at 28.05°, 34.71°, 40.06°, 54.02°, 57.94°, and 66.05° match well with the (110), (101), (200), (211), (220), and (112) planes, respectively, of the rutile phase IrO2 (JCPDS 15–0870). The diffraction peaks of N–TiO2 at about 25.28°, 37.80°, 48.05°, 53.89°, and 55.06° are attributed to the (101), (004), (200), (105), and (211) planes, respectively, of tetragonal anatase TiO2 (JCPDS 21–1272). Regarding IrO2, N–TiO2, and IrO2/N–TiO2, it can be seen that no impurity phase was detected.
The SEM image (Figure 2a) confirms IrO2/N–TiO2 in the form of 1D nanofibers with an average diameter of 100 nm. TEM (Figure 2b and HRTEM). The images in Figure 2c confirm the well–dispersed IrO2 nanoparticles on the surface of the N–TiO2 nanofibers, and the particle size of IrO2 is mainly less than 2 nm (obtained from the particle distribution statistics). Figure 2d shows the lattice fringe values of 0.352 and 0.258 nm that could be assigned to the (101) crystal plane of N–TiO2 and the (101) crystal plane of IrO2, respectively, which is consistent with the XRD results. IrO2 nanoparticles uniformly supported on the surface of N–TiO2 nanofibers were further verified by the HAADF image and the corresponding EDX mapping of Ir, Ti, O, and N elements (Figure 2e). The formed nanofiber is expected to supply more nucleation sites for Ir ions, realizing higher Ir utilization, which is favorable to the increase in active sites and enhancement of the mass activity of IrO2.
Figure 3a,b reveal the SEM cross–sections of IrO2 and IrO2/N–TiO2 catalyst layers, respectively. It can be seen that the average thickness of the IrO2/N–TiO2 catalyst layers is 4.85 μm, which is larger than that of IrO2 (3.02 μm). IrO2 catalyst layers show compact microtopography features (Figure 3c), while the IrO2/N–TiO2 catalyst layers reveal a porous network with more porosity (Figure 3d). Although the IrO2/N–TiO2 catalyst layers are thicker than those of IrO2, it can be assumed that the formed porous networks endowed IrO2/N–TiO2 with a larger reactive contact interface and favorable mass transfer characteristics [13].

3.2. Electrocatalytic Performance Analysis

To explore the support effect of N–TiO2 nanofibers on the activity of IrO2, the electrocatalytic performance of the electrode was examined in a homemade PEM single cell (Figure 4a). Carbon paper and Ti mesh were used as the cathode and anode porous transport layers, respectively. Ti plates with flow fields were employed as bipolar plates. Figure 4b shows the measured polarization curves of pure IrO2 and IrO2/N–TiO2 with potential from 0.01 to 2 A cm−2 at 80 °C. The polarization curve could be divided into three regions: the kinetic–dominated region at the current density of less than 0.15 Acm−2; the ohmic resistance–dominated region at the current density of 0.15–1.8 Acm−2; and the region dominated by ohmic resistance and mass transport at the current density above 1.8 Acm−2. Notably, IrO2 shows a clear increased cell potential that is affected by mass transport when the current density is above 1.8 Acm−2. The obtained cell voltage values of pure IrO2 are 1.858 V at 1 A cm−2 and 2.224 V 2 A cm−2, while the cell voltage values of IrO2/N–TiO2 are 1.797 V at 1 A cm−2 and 2.117 V 2 A cm−2. Note that IrO2/N–TiO2 has superior electrocatalytic performance to that of IrO2. Moreover, the electrocatalytic activity of the IrO2/N–TiO2 electrode is comparable to those in the recent literature. For example, Lim et al. [14] reported a novel IrO2 nanoneedle prepared using the molten salt method in which the cell potential of IrO2 nanoneedles reached >1.8 V at 1 A cm−2 and 2.15 V 2 A cm−2. IrOx/Nb–SnO2 synthesized by Ohno et al. using the flame pyrolysis method showed a high cell potential of 1.91 V at 1 A cm−2 [15]. Recently, Ramani’s group [16] prepared bifunctional Pt–IrO2/RuTiO2 using the co–deposition method; Pt–IrO2/RuTiO2 showed a cell potential of >2 V at 1.5 A cm−2.

3.3. Origin of the Efficient Electrocatalytic Performances

To gain insight into the origin of the superior electrocatalytic performance of IrO2/N–TiO2, CV, Tafel slope, and EIS analyses were further carried out. It has been reported that the amount of electrochemically active sites is generally identified by CV curves. Note that IrO2/N–TiO2 shows a current density that is higher than that of IrO2 (Figure 5a), indicating that the increase active sites due to the nanofiber could provide numerous nucleation sites for Ir ions. The ohmic resistance values (Figure 5b) of the PEM single cell at different densities were recorded at the frequency of 10 kHz. In terms of the poor conductivity of N–TiO2, the ohmic resistance values of IrO2/N–TiO2 are higher than those of IrO2. A decreasing trend of ohmic resistance values with the increasing current density could be ascribed to the occurring heat production at high current densities [18]. To explore the effect of support on the reaction kinetics of IrO2, Tafel slope values were further calculated using Tafel’s equation (Equation (1)).
η = a + b log j
where η is the overpotential and j is the current density. Figure 5c shows the calculated Tafel slope, and the slope values are 73 mV/dec and 48 mV/dec for IrO2 and IrO2/N–TiO2, respectively, suggesting that substantially improved kinetics of IrO2 were achieved after the use of supports. This is also verified by the EIS results. As shown in the EIS plots (Figure 5), the charge transfer resistance values for IrO2 and IrO2/N–TiO2 are 71 and 54 mΩ, respectively. Thus, the increased active sites, smaller Tafel slope value, and charge transfer resistance of IrO2/N–TiO2 indicate that nanofiber supports could efficiently ameliorate the sluggish reaction kinetics of IrO2, which is a benefit of enhancing the OER activity of IrO2.
Figure 6a,b show the contact angle tests of IrO2 and IrO2/N–TiO2 samples using deionized water as the wetting liquid. In addition, the contact angles of IrO2 significantly decreased from 66° to 40° after N–TiO2 introduction, which could be ascribed to fact that the hydrophilic TiO2 nanofibers endow the catalyst–support system with a more hydrophilic surface. The more hydrophilic surface of IrO2/N–TiO2 indicates superior electrolyte penetration and gas adsorption–desorption. To verify the superior kinetics and mass transport of IrO2/N–TiO2, the voltage losses of kinetic losses, ohmic losses, and mass transport losses were estimated [19]. For an electrolysis cell, the cell voltage is the sum of thermodynamic voltage E0(p, T), ohmic losses (R j), kinetics losses, and mass transport losses (Equation (2)).
E = E 0 p , T + R   j + b l o g j + η m t x
In Equation (2), the thermodynamic cell voltage E0(p, T) is calculated from the Nernst equation (Equation (3)):
E 0   p , T   = E 0 T   + R T 2 F + l n a H 2     a O 2 a H 2 O
The thermodynamic cell voltage value is calculated as 1.168 V at 80 °C and atmospheric pressure. Then, the three main potential losses are achieved, as shown in Figure 6c,d. Notably, the mass transport losses and kinetics losses at 2 A cm−2 for IrO2 are 12.1% and 71.4%, respectively, while for IrO2/N–TiO2, the mass transport losses and kinetics losses for IrO2 are 9.8% and 70.6%, respectively. Thus, summarizing the above discussion, the increased electrocatalytic activity of IrO2/N–TiO2 could be assigned to the fact that the IrO2/N–TiO2 nanofiber support system can increase reaction kinetics and accelerate mass transport.

4. Conclusions

In summary, an IrO2/N–TiO2 anodic catalyst was prepared with sub–2 nm IrO2 nanoparticles well dispersed on N–TiO2 nanofiber support, and a porous network structure was constructed in an IrO2/N–TiO2 catalytic layer. The IrO2/N–TiO2 electrode showed cell potential of 2.117 V at 2 A cm−2, which is lower than that of the IrO2 electrode (2.224 V at 2 A cm−2), suggesting the enhanced OER activity of IrO2/N–TiO2. In addition, IrO2 the with N–TiO2 support electrode had more active sites, a smaller Tafel slope value, and higher charge transfer resistance compared with IrO2 without a N–TiO2 support electrode. Moreover, a more hydrophilic surface was observed in IrO2/N–TiO2 after N–TiO2 introduction, which indicates superior electrolyte penetration and gas adsorption–desorption. Therefore, the enhanced electrocatalytic performance of IrO2/N–TiO2 could be ascribed to the increased reaction kinetics and rapid mass transport via the construction of the mass diffusion pathway, as verified by the electrochemical measurements. Constructing support networks of the anodic catalyst layer was demonstrated as a simple and effective method to improve reaction kinetics and favorable mass transport for the highly efficient electrocatalytic performance of the PEM electrolyzer.

Author Contributions

Conceptualization and software, X.S. and W.J.; methodology, H.L. and S.W.; validation, H.L., X.S., and C.Z.; formal analysis, Y.S.; investigation, S.W.; resources, S.W.; data curation, S.W.; writing—original draft preparation, S.W.; writing—review and editing, H.L.; visualization, W.J.; supervision, C.Z.; project administration, H.L.; funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key R&D Program of China (No. 2018YFB1503100).

Acknowledgments

Thanks to the organizers of the 34th World Electric Vehicle Congress for their commendable work. We would like to thank the editors and reviewers for their valuable comments in improving this article.

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. Kang, Z.; Yang, G.; Mo, J.; Li, Y.; Yu, S.; Cullen, D.; Retterer, S.T.; Toops, T.J.; Bender, G.; Pivovar, B.S.; et al. Novel thin/tunable gas diffusion electrodes with ultra–low catalyst loading for hydrogen evolution reactions in proton exchange membrane electrolyzer cells. Nano Energy 2018, 47, 434–441. [Google Scholar] [CrossRef]
  2. Bele, M.; Stojanovski, K.; Jovanovič, P.; Moriau, L.; Podboršek, G.K.; Moškon, J.; Umek, P.; Sluban, M.; Dražič, G.; Hodnik, N.; et al. Towards Stable and Conductive Titanium Oxynitride High-Surface-Area Support for Iridium Nanoparticles as Oxygen Evolution Reaction Electrocatalyst. ChemCatChem 2019, 11, 5038–5044. [Google Scholar] [CrossRef]
  3. Hartig–Weiss, A.; Miller, M.; Beyer, H.; Schmitt, A.; Siebel, A.; Freiberg, A.T.S.; Gasteiger, H.A.; El–Sayed, H.A. Iridium Oxide Catalyst Supported on Antimony–Doped Tin Oxide for High Oxygen Evolution Reaction Activity in Acidic Media. ACS Appl. Nano Mater. 2020, 3, 2185–2196. [Google Scholar] [CrossRef]
  4. Yang, G.; Yu, S.; Kang, Z.; Li, Y.; Bender, G.; Pivovar, B.S.; Green, J.B., Jr.; Cullen, D.A.; Zhang, F.Y. Building electron/proton nanohighways for full utilization of water splitting catalysts. Adv. Energy Mater. 2020, 10, 1903871. [Google Scholar] [CrossRef]
  5. Wang, S.; Lv, H.; Tang, F.; Sun, Y.; Ji, W.; Zhou, W.; Shen, X.; Zhang, C. Defect engineering assisted support effect: IrO2/N defective g–C3N4 composite as highly efficient anode catalyst in PEM water electrolysis. Chem. Eng. J. 2021, 419, 129455. [Google Scholar] [CrossRef]
  6. Lv, H.; Wang, S.; Hao, C.; Zhou, W.; Li, J.; Xue, M.; Zhang, C. Oxygen-Deficient Ti0.9Nb0.1O2-x as an Efficient Anodic Catalyst Support for PEM Water Electrolyzer. ChemCatChem 2019, 11, 2511–2519. [Google Scholar] [CrossRef]
  7. Yu, H.; Bonville, L.; Jankovic, J.; Maric, R. Microscopic insights on the degradation of a PEM water electrolyzer with ultra–low catalyst loading. Appl. Catal. B Environ. 2020, 260, 118194. [Google Scholar] [CrossRef]
  8. Phadke, S.; Lee, J.–Y.; West, J.; Peumans, P.; Salleo, A. Using Alignment and 2D Network Simulations to Study Charge Transport Through Doped ZnO Nanowire Thin Film Electrodes. Adv. Funct. Mater. 2011, 21, 4691–4697. [Google Scholar] [CrossRef]
  9. Jackson, E.D.; Prieto, A.L. Copper Antimonide Nanowire Array Lithium Ion Anodes Stabilized by Electrolyte Additives. ACS Appl. Mater. Interfaces 2016, 8, 30379–30386. [Google Scholar] [CrossRef] [PubMed]
  10. SSaha, M.S.; Li, R.; Cai, M.; Sun, X. Nanowire–based three–dimensional hierarchical core/shell heterostructured electrodes for high performance proton exchange membrane fuel cells. J. Power Sources 2008, 185, 1079–1085. [Google Scholar] [CrossRef]
  11. Pan, C.; Wu, H.; Wang, C.; Wang, B.; Zhang, L.; Cheng, Z.; Hu, P.; Pan, W.; Zhou, Z.; Yang, X.; et al. Nanowire–Based High–Performance “Micro Fuel Cells”: One Nanowire, One Fuel Cell. Adv. Mater. 2008, 20, 1644–1648. [Google Scholar] [CrossRef]
  12. Hong, W.; Shang, C.; Wang, J.; Wang, E. Bimetallic PdPt nanowire networks with enhanced electrocatalytic activity for ethylene glycol and glycerol oxidation. Energy Environ. Sci. 2015, 8, 2910–2915. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Gao, F.; Song, T.; Wang, C.; Chen, C.; Du, Y. Novel networked wicker–like PtFe nanowires with branch–rich exteriors for efficient electrocatalysis. Nanoscale 2019, 11, 15561–15566. [Google Scholar] [CrossRef] [PubMed]
  14. Lim, J.; Park, D.; Jeon, S.S.; Roh, C.–W.; Choi, J.; Yoon, D.; Park, M.; Jung, H.; Lee, H. Ultrathin IrO2 Nanoneedles for Electrochemical Water Oxidation. Adv. Funct. Mater. 2018, 28, 170479. [Google Scholar] [CrossRef]
  15. Gayen, P.; Liu, X.; He, C.; Saha, S.; Ramani, V.K. Bidirectional energy & fuel production using RTO–supported–Pt–IrO2 loaded fixed polarity unitized regenerative fuel cells. Sustain. Energy Fuels 2021, 5, 2734–2746. [Google Scholar]
  16. Ohno, H.; Nohara, S.; Kakinuma, K.; Uchida, M.; Uchida, H. Effect of Electronic Conductivities of Iridium Oxide/Doped SnO2 Oxygen–Evolving Catalysts on the Polarization Properties in Proton Exchange Membrane Water Electrolysis. Catalysts 2019, 9, 74. [Google Scholar] [CrossRef] [Green Version]
  17. Lv, H.; Wang, S.; Li, J.; Shao, C.; Zhou, W.; Shen, X.; Xue, M.; Zhang, C. Self–assembled RuO2@IrOx core–shell nanocomposite as high efficient anode catalyst for PEM water electrolyzer. Appl. Surf. Sci. 2020, 514, 145943. [Google Scholar] [CrossRef]
  18. Suermann, M.; Schmidt, T.J.; Büchi, F.N. Investigation of Mass Transport Losses in Polymer Electrolyte Electrolysis Cells. ECS Trans. 2015, 69, 1141–1148. [Google Scholar] [CrossRef]
  19. Lee, J.K.; Lee, C.; Fahy, K.F.; Kim, P.J.; Krause, K.; Lamanna, J.M.; Baltic, E.; Jacobson, D.L.; Hussey, D.S.; Bazylak, A. Accelerating Bubble Detachment in Porous Transport Layers with Patterned Through–Pores. ACS Appl. Energy Mater. 2020, 3, 9676–9684. [Google Scholar] [CrossRef]
Scheme 1. Schematics of the preparation method of N–TiO2 and IrO2/N–TiO2.
Scheme 1. Schematics of the preparation method of N–TiO2 and IrO2/N–TiO2.
Wevj 12 00124 sch001
Figure 1. XRD patterns of the prepared samples.
Figure 1. XRD patterns of the prepared samples.
Wevj 12 00124 g001
Figure 2. (a) SEM image; (b) TEM; (c,d) HR–TEM; (e) HAADF image; and the corresponding EDX mapping of Ir, Ti, O, and N elements of IrO2/N–TiO2.
Figure 2. (a) SEM image; (b) TEM; (c,d) HR–TEM; (e) HAADF image; and the corresponding EDX mapping of Ir, Ti, O, and N elements of IrO2/N–TiO2.
Wevj 12 00124 g002
Figure 3. (a,c) SEM cross–sections and (b,d) SEM images of the IrO2 and IrO2/N–TiO2 catalyst layers, respectively.
Figure 3. (a,c) SEM cross–sections and (b,d) SEM images of the IrO2 and IrO2/N–TiO2 catalyst layers, respectively.
Wevj 12 00124 g003
Figure 4. (a) Schematics of the home–made PEM single cell; reproduced with permission from Lv, H.; Wang, S.; Li, J.; Shao, C.; Zhou, W.; Shen, X.; Xue, M.; Zhang, C, Appl. Surf. Sci.; published by Elsevier, 2020 [17]; (b) polarization curves of IrO2 and IrO2/N–TiO2 with the potential from 0.01 to 2 A cm−2 at 80 °C.
Figure 4. (a) Schematics of the home–made PEM single cell; reproduced with permission from Lv, H.; Wang, S.; Li, J.; Shao, C.; Zhou, W.; Shen, X.; Xue, M.; Zhang, C, Appl. Surf. Sci.; published by Elsevier, 2020 [17]; (b) polarization curves of IrO2 and IrO2/N–TiO2 with the potential from 0.01 to 2 A cm−2 at 80 °C.
Wevj 12 00124 g004
Figure 5. (a) CV curves, (b) HFR resistance, (c)Tafel slopes, and (d) EIS of IrO2 and IrO2/N–TiO2.
Figure 5. (a) CV curves, (b) HFR resistance, (c)Tafel slopes, and (d) EIS of IrO2 and IrO2/N–TiO2.
Wevj 12 00124 g005
Figure 6. (a,b) Contact angles and (c,d) voltage losses of IrO2 and IrO2/N–TiO2.
Figure 6. (a,b) Contact angles and (c,d) voltage losses of IrO2 and IrO2/N–TiO2.
Wevj 12 00124 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, S.; Lv, H.; Sun, Y.; Ji, W.; Shen, X.; Zhang, C. Constructing Supports–Network with N–TiO2 Nanofibres for Highly Efficient Hydrogen–Production of PEM Electrolyzer. World Electr. Veh. J. 2021, 12, 124. https://doi.org/10.3390/wevj12030124

AMA Style

Wang S, Lv H, Sun Y, Ji W, Shen X, Zhang C. Constructing Supports–Network with N–TiO2 Nanofibres for Highly Efficient Hydrogen–Production of PEM Electrolyzer. World Electric Vehicle Journal. 2021; 12(3):124. https://doi.org/10.3390/wevj12030124

Chicago/Turabian Style

Wang, Sen, Hong Lv, Yongwen Sun, Wenxuan Ji, Xiaojun Shen, and Cunman Zhang. 2021. "Constructing Supports–Network with N–TiO2 Nanofibres for Highly Efficient Hydrogen–Production of PEM Electrolyzer" World Electric Vehicle Journal 12, no. 3: 124. https://doi.org/10.3390/wevj12030124

Article Metrics

Back to TopTop