Next Article in Journal
Delivering Combination Chemotherapies and Targeting Oncogenic Pathways via Polymeric Drug Delivery Systems
Next Article in Special Issue
A Bioinspired Functionalization of Polypropylene Separator for Lithium-Sulfur Battery
Previous Article in Journal
Functional Nanostructured Oligochitosan–Silica/Carboxymethyl Cellulose Hybrid Materials: Synthesis and Investigation of Their Antifungal Abilities
Previous Article in Special Issue
A Comparative Study on Physiochemical, Thermomechanical, and Electrochemical Properties of Sulfonated Poly(Ether Ether Ketone) Block Copolymer Membranes with and without Fe3O4 Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of 3-DOM IrO2 Electrocatalysts Templated by PMMA for Oxygen Evolution Reaction

State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
*
Authors to whom correspondence should be addressed.
Polymers 2019, 11(4), 629; https://doi.org/10.3390/polym11040629
Submission received: 4 March 2019 / Revised: 30 March 2019 / Accepted: 1 April 2019 / Published: 4 April 2019

Abstract

:
Three-dimensional ordered macroporous (3-DOM) IrO2 material was prepared using PMMA as a template and ammonia as a chelator. These 3-DOM IrO2 honeycomb arrays showed a large surface area and ordered macropores (155 nm in diameter) cross-linked by secondary mesopores. Internal structures of 3-DOM IrO2 material were observed microscopically through these secondary pores. According to the X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS) spectra, 3-DOM IrO2 has a rutile crystal structure and is mainly composed of iridium dioxide. In acidic electrolytes, the overpotential of 3-DOM IrO2 material at 0.5 mV cm−2 was only 0.22 V. Accelerated durability tests demonstrated excellent durability of 3-DOM IrO2 as an oxygen evolution reaction (OER) catalyst.

Graphical Abstract

1. Introduction

Hydrogen, as an environmentally friendly energy carrier with high thermal value, has many advantages, which solves the energy problems and environmental issues associated with other energy resources [1,2,3]. The preparation of hydrogen is mainly based on fossil fuel cracking and water electrolysis. Solid polymer electrolyte (SPE) water electrolysis is considered to be one of the most promising methods for hydrogen production because of its high energy conversion efficiency and power density. Hydrogen production systems are safe and clean. Products of this process are just oxygen and hydrogen, both of which possess high purity [4]. However, some problems associated with SPE water electrolysis still exist, including high anodic overpotential and the high cost of the system [4,5].
Despite numerous research studies in this area, high anodic overpotential remains a great challenge for oxygen evolution reaction (OER). Additionally, the stability of electrocatalysts is another factor restricting widespread SPE applications. The majority of metals and their oxides cannot be stabilized under high anode potential and strong acidic conditions; thus, choices for anode catalysts are very limited. Platinum metals have good catalytic performance because of their excellent adsorption affinity to hydrogen and excellent catalytic properties. Thus, Pt has been widely used in SPE water electrolysis [6,7], fuel cells [8,9,10,11], and hydrogen storage [12,13]. However, Pt-based catalysts are not suitable for OER because of the high overpotential of Pt as it forms platinum oxides, which result in the poor conductivity of Pt [14]. Fortunately, Ir and Ru have good catalytic OER activity. Their oxides also have good conductivity and are stable in acidic environments. Interfaces of Ir and Ru oxides are also very reactive.
To increase the stability of the catalyst at the anode during SPE water electrolysis, a lot of work has been performed especially involving dimensionally stable anodes (DSA). Durable material to load on IrO2 and RuO2 were selected, most of which were oxides of transition metals. Although these materials have excellent chemical stability, their conductivity is poor, which makes OER electron transfer dependent on IrO2 and RuO2. The result does not reduce the amount of precious metals, and only when the content of IrO2 in the IrO2-ZrO2 solid solution reaches 80% does the catalyst demonstrate good performance [15]. Other solid oxides forming catalysts with high levels of IrO2 and/or RuO2 are TiO2 [16,17,18,19], SiC [20], SnO2 [21,22,23,24], Ta2O5 [25,26], and Nb2O5 [27], etc. Research for a variety of other support materials is still underway.
To increase the catalytic activity during SPE water electrolysis, the activity of precious metal catalysts needs to be improved, for example by studying and analyzing its surface area and micropore structure [28,29,30,31]. Recently, Li et al. [30] synthesized a nanoporous IrO2 catalyst based on a template-free ammonia complex. Unlike the Adams fusion method, ammonia was used as a chelating agent to replace Cl in H2IrCl4. This IrO2 catalyst demonstrated the highest specific surface area ever reported (363.3 m2 g−1) as well as outstanding performance. Ortel et al. [31] synthesized ordered mesoporous IrO2 film templated by PEO–PB–PEO block copolymers. The films possess nanocrystalline walls and ordered pores, which enhance the durability and activity of OER.
Three-dimensional ordered macroporous (3-DOM) materials are novel molecular sieves with uniform and ordered pores and well-controlled pore structures. The ordered porosity of the 3-DOM material is very conducive to an even distribution of the electrolyte between the electrode, which improves the mass transfer of ions to the electrode surface [32]. This can increase the number of reactive sites and improve the catalytic activity of materials. The interconnected structure of 3-DOM materials can increase conductivity [33], which makes them suitable as electrochemical catalysts. 3-DOM materials were successfully used in a lithium-ion battery and fuel cell [32,34]. However, few 3-DOM oxide materials have been reported for OER electrocatalysts. Hu et al. first synthesized 3-DOM IrO2 using a SiO2 microsphere template [35,36], in which 3-DOM IrO2 exhibited an ordered honeycomb array of macropores. The overpotential of this OER catalyst was only 0.22 V at 0.5 mA cm−2. However, organic templates have not been reported for the preparation of 3-DOM IrO2. Herein, we synthesized 3-DOM IrO2 using a PMMA template. Compared to the SiO2 template, PMMA templates can be removed by annealing, which makes the preparation process very simple, thereby reducing its cost. However, the decomposition temperature of the PMMA template is under 300 °C, which is too low for iridium chloride to form iridium oxide. Because the formation of 3-DOM structures is difficult under such conditions, we used ammonia as a chelator, which reacts with iridium chloride to form ammonium chloride and iridium hydroxide. Ammonium chloride is volatile and easily decomposed. Iridium hydroxide is easily dehydrated to produce iridium oxide under 300 °C, which can protect the 3-DOM structure against collapse. 3-DOM IrO2 was characterized by SEM, Brunauer-Emmett-Teller (BET), X-ray diffraction (XRD), and X-ray photoelectron spectroscopy (XPS). Ordered macropores were clearly seen with a significant decrease in overpotential.

2. Experimental

2.1. Preparation of PMMA Templates

Conventional emulsifier-free emulsion polymerization was used to prepare polymethyl methacrylate (PMMA) templates [37]. First, 260 mL of deionized water was placed into a 500 mL round-bottomed flask and heated to 70 °C inside a heat-collecting thermostatic magnetic stirrer. After nitrogen was passed through the flask for 15 min, potassium persulfate (KPS, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) and methyl methacrylate (MMA, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) were added in a ratio of 1:118. The whole mixture was stirred for 1 h under nitrogen. After that, the PMMA emulsion was centrifuged and washed three times with deionized water, following which PMMA microspheres were dispersed ultrasonically in deionized water. A certain amount of PMMA emulsion was placed into a 25 mL beaker, which was then covered by tin foil with holes. PMMA microspheres were settled down and dried at room temperature. Bright fluorescence on the surface indicated that the PMMA microspheres were hexagonally close-packed.

2.2. Preparation of 3-DOM IrO2

IrCl4·2H2O (Shanxi Kaida Chemical Engineering Co., Ltd., Shanxi, China) was added to H2O and isopropyl alcohol (IPA, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) (H2O: IPA = 3:1). Then, 4 mL of IrCl4·2H2O 0.03 mol L−1 was added to PMMA templates until it was immersed. After 30 min, the excess solution was removed by a dropper. PMMA templates were put into the oven at 70 °C. After drying, the above steps were repeated three times. Finally, the templates were heated to 450 °C in air at a rate of 5 °C/min in a tube furnace for 4 h. The final product was a black powder (marked as non-ammonia IrO2). Ammonia (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) (IrCl4·nH2O: NH4·H2O = 1:4) was then added to the above solution that contains isopropyl alcohol, H2O, and IrCl4·2H2O. 3-DOM IrO2 was prepared by repeated preparation of non-ammonia IrO2. Colloidal IrO2 was prepared by the colloidal method [35].

2.3. Physical Characterization

Scanning electron microscopy (SEM) was performed using a Hitachi S-4800 microscope operated at 8 kV. X-ray photoelectron spectroscopy (XPS) spectra were recorded by an ESCALAB250Xi spectrometer using Mg Kα as an X-ray source. X-ray diffraction (XRD) spectra were recorded by Bruker D8 Advance using Cu–Kα radiation. Particle sizes were measured by the dynamic light scattering (DLS) technique. Specific surface areas and pore distribution were obtained by implementing Brunauer-Emmett-Teller (BET) and Barrett-Joyner-Halenda (BJH) methods, respectively, using N2 gas absorption/desorption data collected by Micromeritics ASAP 2020 (BELSORP-max).

2.4. Electrochemical Characterization

Electrochemical properties were tested using a three-electrode system at room temperature. The electrolyte was 0.5 mol L−1 H2SO4 solution. The counter electrode was platinum foil, and a saturated calomel electrode (SCE, 0.242 V) was a reference electrode. The working electrode was a glassy carbon (GC) 5 mm in diameter. The catalyst ink was prepared using 5 mg of testing material, 980 µL of isopropanol, and 20 µL of 5% Nafion. Both catalysts had a load of 106 µg cm−2. Cyclic voltammograms (CV) were measured between 0 V and 1.4 V at a 50 mV s−1 scan rate. Steady-state polarization curves for OER were obtained at a 5 mV s−1 scan rate. Before all tests, the electrode was activated by performing 20 cycles of the CV curve at the scanning rate (500 mV s−1). All electrochemical tests were performed using the CHI660 workstation.

3. Results and Discussion

3.1. Physical Characterization

Since PMMA microspheres serve as a synthesis template, their particle size distribution is an important physical property. Analysis of the particle size distribution of PMMA by DLS demonstrates that the average diameter of PMMA microspheres is ~220 nm (Figure 1a). Polydispersity of PMMA microspheres is 1.7%, which is indicative of good uniformity. A single particle size of PMMA microspheres is an important reason for forming a regular colloidal crystal template. Also, sizes of PMMA microspheres are controlled by the quantity of the initiator and MMA. The PMMA template consists of hexagonally close-packed PMMA microspheres (Figure 1b), which have a uniform distribution of particle size (the diameter of the particles is 220 nm). An obvious layered structure can be clearly observed in Figure 1b. All layers have the same close-packed hexagonal arrangement of the microspheres. Several point defects are observed as well as some microsphere misalignments (markings in Figure 1b). These defects have no impact on the template ordered macroporous material formed because of this ordered regular template. The PMMA template is covered by the precursor, as observed in Figure 1d. After immersion, the PMMA template retains the same close-packed hexagonal arrangement of the microspheres. The PMMA colloidal crystal template exhibits fluorescence (Figure 1c) under natural sunlight, which is a typical Bragg behavior of ordered microspheres.
3-DOM IrO2 exhibits a three-dimensional ordered honeycomb macroporous structure (Figure 2). Most of the macropores are 155 nm in diameter, which is ~30% smaller than the diameters of PMMA microspheres in the template, probably because of the carbonization and dehydration of the organic templates during annealing. PMMA contains many hydrogen and oxygen elements, and dehydration can greatly reduce the microsphere size as well as carbonize the PMMA template. The honeycomb pore structure can be seen in the high-magnification SEM image. In addition, small pores in the walls of the large pores, which can connect to the surrounding large pore structures, can also be clearly observed (Figure 2b). These pore structures are formed because when accumulated hexagonally close-packed PMMA microspheres are in contact with each other, the precursor solution cannot pass through the points of contact of the PMMA spheres during impregnation. After the removal of the PMMA template, these contact points form small pores. The solid structure of the 3-DOM IrO2 can be seen through these small pores. These pores are favorable for the even distribution of the electrolyte between the electrodes, which will improve the mass transfer of ions to the surface of the electrodes. This can also increase reactive sites and improve the catalytic activity of the material. Additionally, such interconnected structures can increase the overall conductivity. A few pores can be observed in non-ammonia IrO2 (Figure 2c). These pores are broken and disordered probably because the decomposition temperature of IrCl4 is higher than that of the organic template. Before IrCl4 decomposes, the decomposition of the organic template causes the pore structure to collapse (Figure 2d).
The specific surface area of the OER catalyst is an important physical property. Because of its microporous structure, it has a large specific surface area. Figure 3a exhibits N2 adsorption/desorption isotherms for 3-DOM IrO2 and non-ammonia IrO2. According to the International Union of Pure and Applied Chemistry (IUPAC) classification [38], the curve of 3-DOM IrO2 belongs to the type IV with H3-type hysteresis, which is usually associated with highly consistent mesopore structures. Non-ammonia IrO2 also shows the type IV with H3-type hysteresis. However, the hysteresis loop is smaller than that in 3-DOM IrO2, showing that 3-DOM IrO2 has more mesopores. According to the BJH pore-size distribution analysis (Figure 3b), the pore sizes of both materials are concentrated below 20 nm. Diameters of 3-DOM IrO2 mesopores are ~1.2 nm, while diameters of non-ammonia IrO2 mesopores are ~2.1 nm. The BJH surface areas of the mesopores of 3-DOM IrO2 and non-ammonia IrO2 are 34.5 and 23.8 m2 g−1, respectively, indicating that 3-DOM IrO2 has more mesopores. The BET surface areas of 3-DOM IrO2 and non-ammonia IrO2 are 67.3 and 37.1 m2 g−1, respectively. The BET surface area of 3-DOM IrO2 is close to the mesoporous IrO2 structure reported by Ortel et al. [31].
Figure 4 shows the XRD patterns of two kinds of materials. XRD patterns of 3-DOM IrO2 showing diffraction peaks corresponding to the rutile IrO2 peak at 2θ angles equal to 28°, 34°, 40°, 54°, and 58°, which correspond to the (110), (101), (200), (211), and (220) planes, can be easily seen.
Non-ammonia IrO2 pattern does not show rutile IrO2 peaks. However, characteristic peaks of the elemental Ir at 2θ of ~40°, 47°, and 69° (which correspond to (111), (200), and (220) planes) could be clearly seen. This indicates that 3-DOM IrO2 contained metallic Ir formed because of template carbonization; after PMMA carbonization, internal Ir4+ had no access to air. Thus, a small amount of IrO2 was reduced to Ir. However, Ir4+ in the precursor that has no ammonia is IrCl4, and since Cl is less electronegative than O, Ir4+ was reduced almost entirely.
XPS analysis of 3-DOM suggests that IrO2 is present in the catalyst (Figure 5). No other impurities are observed on the surface (Figure 5a). Characteristic peaks of Ir are clearly seen as well. High-resolution spectra of the Ir4f demonstrates characteristic peaks at 61.5 eV (Figure 5b). This binding energy value agrees with values obtained from nanoporous IrO2 synthesized by the template-free ammonia complex [30].

3.2. Electrocatalytic Characterization

The electrocatalytic performances of two materials are shown in Figure 6. The peaks observed in the CV curve of colloidal IrO2 (Figure 6a) clearly indicate Ir transformation (IrIII/IrIV and IrIV/IrIV), accompanied by protonation/deprotonation. The CV curve of 3-DOM IrO2 demonstrated a distinct redox peak at a low potential, which was related to the elemental Ir in 3-DOM IrO2. Ir, similar to Pt and Ru, can absorb hydrogen in the hydrogen underpotential deposition region [39,40]. 3-DOM IrO2 showed significantly larger current density in comparison to colloidal IrO2. Current responses consist of two effects, the redox pseudo-capacitance and the double-layer capacitance, both of which are related to the electroactive surface areas (ESAs).
The linear sweep voltammetry (LSV) of two catalysts demonstrating their OER activities are shown in Figure 6b. 3-DOM IrO2 demonstrates superior OER activity compared to colloidal IrO2. Compared to the thermodynamic minimum potential of OER, the overpotential of 3-DOM IrO2 is 0.22 V at 0.5 mA cm−2, which outperforms colloidal IrO2 (0.25 V) and is the same as the 3-DOM IrO2 material synthesized by the silica colloidal crystal template [35] and the mesoporous IrO2 templated by PEO–PB–PEO copolymers [31]. However, the preparation of PMMA microspheres was easy as 3-DOM IrO2 prepared by the PMMA template did not require acid solution treatment, thereby reducing the cost of preparation. Figure 6b suggests that 3-DOM IrO2 (106 µg cm−2) has the same catalytic activity as colloidal IrO2 (265 µg cm−2), while the OER activity of 3-DOM IrO2 is outstanding compared to that of colloidal IrO2, because of their different ESAs (Figure 6a). The rutile IrO2 shows lower O2-evolving overpotential than the metal Ir [41]; therefore, the metal Ir is not the main reason for the increase of catalytic activity. Catalytic data were normalized relative to the corresponding CV charges (inset of Figure 6c). The curves represent intrinsic activities of the catalysts, which are almost identical (Figure 6c). Thus, intrinsic OER activities are identical, and the difference in porosity does not affect catalytic performance. However, 3-DOM IrO2 shows superior catalytic performance under the same mass condition. The slope of the Tafel curve for 3-DOM IrO2 is 57.9 mV dec−1 (Figure 6d), which is lower than that of colloidal IrO2 (which is equal to 62.2 mV dec−1). The Tafel slope is similar to that of nanoporous IrO2 and within the normal range of conventional IrO2 [30]. This also confirms the robust catalytic activity of 3-DOM IrO2, indicating a faster rate for the OER.
Figure 7 shows current change as a function of time. From these curves, the OER stabilities of both catalysts in acidic environments can be obtained. Current density dropped initially for both catalysts. After 3000 s, the decline of the current density slowed down. However, 3-DOM IrO2 still showed higher current density than that of colloidal IrO2. Slopes of the fitting curves in the range of 15,000–20,000 s (insert in Figure 7) are −2.09 × 10−4 and −2.37 × 10−4, respectively. Thus, the current change of colloidal IrO2 deceases faster and 3-DOM IrO2 exhibits superior stability than that of colloidal IrO2.

4. Conclusions

Synthesis of 3-DOM IrO2 using a PMMA template and ammonia sol has been performed. The issue of the organic template having difficulty in maintaining the 3-DOM IrO2 structure has been solved by the ammonia sol. The sol consists of iridium chloride, isopropanol, and ammonia. IrCl4 and ammonia form a stable complex solution, in which NH4+ and Cl form ammonium chloride, and ammonium chloride can easily be removed. The prepared 3-DOM IrO2 material has a honeycomb array macroporous structure with an average pore diameter of 155 nm and a large surface area. Many mesoscale pores are observed on the wall of the macropores. 3-DOM IrO2 demonstrates superior OER activity compared to colloidal IrO2, and the overpotential of 3-DOM IrO2 is lower than that of colloidal IrO2 by about 30 mV at 0.5 mA cm−2. In the same catalytic activity, the loading of 3-DOM IrO2 is much lower than that of colloidal IrO2. Our novel method to synthesize the 3-DOM IrO2 catalyst can reduce the anode overpotential of water electrolysis. The results suggest that the 3-DOM IrO2 catalyst is a promising material for OER.

Author Contributions

Conceptualization—F.L., J.Y., and H.T.; Data curation—F.L., X.S., X.C., and C.L.; Investigation—F.L., J.Y., and H.T.; Resources—X.S., X.C., and C.L.; Supervision—J.Y. and H.T.; Writing—original draft—F.L.; Writing—review and editing—F.L., J.Y., and H.T.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Bockris, J.O.M. The hydrogen economy: Its history. Int. J. Hydrogen Energy 2013, 38, 2579–2588. [Google Scholar]
  2. Singh, S.; Jain, S.; Ps, V.; Tiwari, A.K.; Nouni, M.R.; Pandey, J.K.; Goel, S. Hydrogen: A sustainable fuel for future of the transport sector. Renew. Sustain. Energy Rev. 2015, 51, 623–633. [Google Scholar] [CrossRef]
  3. Cheng, J.; Zhang, H.; Ma, H.; Zhong, H.; Zou, Y. Study of carbon-supported IrO2 and RuO2 for use in the hydrogen evolution reaction in a solid polymer electrolyte electrolyzer. Electrochim. Acta 2010, 55, 1855–1861. [Google Scholar] [CrossRef]
  4. Sharma, S.; Ghoshal, S.K. Hydrogen the future transportation fuel: From production to applications. Renew. Sustain. Energy Rev. 2015, 43, 1151–1158. [Google Scholar] [CrossRef]
  5. Schmidt, O.; Gambhir, A.; Staffell, I.; Hawkes, A.; Nelson, J.; Few, S. Future cost and performance of water electrolysis: An expert elicitation study. Int. J. Hydrogen Energy 2017, 42, 30470–30492. [Google Scholar] [CrossRef]
  6. Sapountzi, F.M.; Gracia, J.M.; Weststrate, C.J.; Fredriksson, H.O.A.; Niemantsverdriet, J.W. Electrocatalysts for the generation of hydrogen, oxygen and synthesis gas. Prog. Energy Combust. Sci. 2017, 58, 1–35. [Google Scholar] [CrossRef]
  7. Eftekhari, A. Electrocatalysts for hydrogen evolution reaction. Int. J. Hydrogen Energy 2017, 42, 11053–11077. [Google Scholar] [CrossRef]
  8. Yang, H.N.; Park, S.H.; Lee, D.C.; Yi, S.C.; Kim, W.J. Electrochemical properties of hybrid typed electrocatalyst using Pt/carbon molecular sieve synthesized by zeolite template and Pt carbon black. Microporous Mesoporous Mater. 2013, 172, 161–166. [Google Scholar] [CrossRef]
  9. Kaewsai, D.; Yeamdee, S.; Supajaroon, S.; Hunsom, M. Orr activity and stability of PtCr/C catalysts in a low temperature/pressure PEM fuel cell: Effect of heat treatment temperature. Int. J. Hydrogen Energy 2018, 43, S0360319918301769. [Google Scholar] [CrossRef]
  10. Cantane, D.A.; Oliveira, F.E.R.; Santos, S.F.; Lima, F.H.B. Synthesis of Pt-based hollow nanoparticles using carbon-supported Co@Pt and Ni@Pt core–shell structures as templates: Electrocatalytic activity for the oxygen reduction reaction. Appl. Catal. B Environ. 2013, 136–137, 351–360. [Google Scholar] [CrossRef]
  11. Yun, Y.H.; Kim, S.D.; Park, S.W.; Lee, J.Y.; Yi, S.C.; Kim, W.J. Cell performance of MEA fabricated with Pt-ZSM-5-carbon electrode for PEMFC. Microporous Mesoporous Mater. 2010, 131, 122–127. [Google Scholar] [CrossRef]
  12. Kishor, R.; Singh, S.B.; Ghoshal, A.K. Role of metal type on mesoporous KIT-6 for hydrogen storage. Int. J. Hydrogen Energy 2018, S0360319918312643. [Google Scholar] [CrossRef]
  13. Alam, N.; Mokaya, R. Characterisation and hydrogen storage of Pt-doped carbons templated by Pt-exchanged zeolite y. Microporous Mesoporous Mater. 2011, 142, 716–724. [Google Scholar] [CrossRef]
  14. Miles, M.H.; Klaus, E.A.; Gunn, B.P.; Locker, J.R.; Serafin, W.E.; Srinivasan, S. The oxygen evolution reaction on platinum, iridium, ruthenium and their alloys at 80 °C in acid solutions. Electrochim. Acta 1978, 23, 521–526. [Google Scholar] [CrossRef]
  15. Benedetti, A.; Riello, P.; Battaglin, G.; Battisti, A.D.; Barbieri, A. Physicochemical properties of thermally prepared Ti-supported IrO2 + ZrO2 electrocatalysts. J. Electroanal. Chem. 1994, 376, 195–202. [Google Scholar] [CrossRef]
  16. Lu, Z.-X.; Shi, Y.; Yan, C.-F.; Guo, C.-Q.; Wang, Z.-D. Investigation on IrO2 supported on hydrogenated TiO2 nanotube array as OER electro-catalyst for water electrolysis. Int. J. Hydrogen Energy 2017, 42, 3572–3578. [Google Scholar] [CrossRef]
  17. Nikolić, B.Ž.; Panić, V.V.; Dekanski, A.B. Intrinsic potential-dependent performances of a Sol–Gel-Prepared electrocatalytic IrO2–TiO2 coating of dimensionally stable anodes. Electrocatalysis 2012, 3, 360–368. [Google Scholar] [CrossRef]
  18. Osman, J.R.; Crayston, J.A.; Pratt, A.; Richens, D.T. Sol–gel processing of IrO2–TiO2 mixed metal oxides based on a hexachloroiridate precursor. J. Sol-Gel Sci. Technol. 2007, 44, 219–225. [Google Scholar] [CrossRef]
  19. Oakton, E.; Lebedev, D.; Povia, M.; Abbott, D.F.; Fabbri, E.; Fedorov, A.; Nachtegaal, M.; Copéret, C.; Schmidt, T.J. IrO2-TiO2: A high-surface-area, active, and stable electrocatalyst for the oxygen evolution reaction. ACS Catal. 2017, 7, 2346–2352. [Google Scholar] [CrossRef]
  20. Nikiforov, A.V.; García, A.L.T.; Petrushina, I.M.; Christensen, E.; Bjerrum, N.J. Preparation and study of IrO2/SiC-Si supported anode catalyst for high temperature pem steam electrolysers. Int. J. Hydrogen Energy 2011, 36, 5797–5805. [Google Scholar] [CrossRef]
  21. Xu, W.; Tayal, J.; Basu, S.; Scott, K. Nano-crystalline RuxSn1-xO2 powder catalysts for oxygen evolution reaction in proton exchange membrane water electrolysers. Int. J. Hydrogen Energy 2011, 36, 14796–14804. [Google Scholar]
  22. Ferro, S.; Rosestolato, D.; Martínez-Huitle, C.A.; Battisti, A.D. On the oxygen evolution reaction at IrO2-SnO2 mixed-oxide electrodes. Electrochim. Acta 2014, 146, 257–261. [Google Scholar] [CrossRef]
  23. Xu, J.; Liu, G.; Li, J.; Wang, X. The electrocatalytic properties of an IrO2/SnO2 catalyst using SnO2 as a support and an assisting reagent for the oxygen evolution reaction. Electrochim. Acta 2012, 59, 105–112. [Google Scholar] [CrossRef]
  24. Jiang, G.; Yu, H.; Hao, J.; Chi, J.; Fan, Z.; Yao, D.; Qin, B.; Shao, Z. An effective oxygen electrode based on Ir0.6Sn0.4O2 for pem water electrolyzers. J. Energy Chem. 2019, 39, 23–28. [Google Scholar] [CrossRef]
  25. Yan, Z.; Zhang, H.; Feng, Z.; Tang, M.; Yuan, X.; Tan, Z. Promotion of in situ tin x interlayer on morphology and electrochemical properties of titanium based IrO2-Ta2O5 coated anode. J. Alloys Compd. 2017, 708, 1081–1088. [Google Scholar] [CrossRef]
  26. Cuevas, O.; Herrada, R.A.; Corona, J.L.; Olvera, M.G.; Sepúlveda-Guzmán, S.; Sirés, I.; Bustos, E. Assessment of IrO2-Ta2O5|Ti electrodes for the electrokinetic treatment of hydrocarbon-contaminated soil using different electrode arrays. Electrochim. Acta 2016, 208, 282–287. [Google Scholar] [CrossRef]
  27. Kadakia, K.S.; Jampani, P.H.; Velikokhatnyi, O.I.; Datta, M.K.; Patel, P.; Chung, S.J.; Park, S.K.; Poston, J.A.; Manivannan, A.; Kumta, P.N. Study of fluorine doped (Nb, Ir)O2 solid solution electro-catalyst powders for proton exchange membrane based oxygen evolution reaction. Mater. Sci. Eng. B 2016, 212, 101–108. [Google Scholar] [CrossRef]
  28. El Sawy, E.N.; Birss, V.I. Nano-porous iridium and iridium oxide thin films formed by high efficiency electrodeposition. J. Mater. Chem. 2009, 19, 8244–8252. [Google Scholar] [CrossRef]
  29. Cherevko, S.; Reier, T.; Zeradjanin, A.R.; Pawolek, Z.; Strasser, P.; Mayrhofer, K.J.J. Stability of nanostructured iridium oxide electrocatalysts during oxygen evolution reaction in acidic environment. Electrochem. Commun. 2014, 48, 81–85. [Google Scholar] [CrossRef]
  30. Li, G.; Li, S.; Xiao, M.; Ge, J.; Liu, C.; Xing, W. Nanoporous IrO2 catalyst with enhanced activity and durability for water oxidation owing to its micro/mesoporous structure. Nanoscale 2017, 9, 9291. [Google Scholar] [CrossRef]
  31. Ortel, E.; Reier, T.; Strasser, P.; Kraehnert, R. Mesoporous IrO2 films templated by PEO-PB-PEO Block-Copolymers: Self-Assembly, Crystallization Behavior, and Electrocatalytic performance. Chem. Mater. 2011, 23, 3201–3209. [Google Scholar] [CrossRef]
  32. Lei, W.; Wang, F.; Zhu, J.; Xin, Z.; Yi, T.; Xing, W. Synthesis and electrochemical performance of three-dimensional ordered hierarchically porous Li4Ti5O12 for high performance lithium ion batteries. Ceram. Int. 2017, 44, S0272884217317418. [Google Scholar]
  33. Ergang, N.S.; Lytle, J.C.; Lee, K.T.; Oh, S.M.; Smyrl, W.H.; Stein, A. Photonic crystal structures as a basis for a Three-Dimensionally interpenetrating electrochemical-cell system. Adv. Mater. 2010, 18, 1750–1753. [Google Scholar] [CrossRef]
  34. Chunzhen, Y.; Ming, Z.; Qian, X. Three-dimensional ordered macroporous MnO2/carbon nanocomposites as high-performance electrodes for asymmetric supercapacitors. Phys. Chem. Chem. Phys. 2013, 15, 19730–19740. [Google Scholar]
  35. Hu, W.; Wang, Y.; Hu, X.; Zhou, Y.; Chen, S. Three-dimensional ordered macroporous IrO2 as electrocatalyst for oxygen evolution reaction in acidic medium. J. Mater. Chem. 2012, 22, 6010. [Google Scholar] [CrossRef]
  36. Hu, W.; Zhou, P.; Xu, S.; Chen, S.; Xia, Q. Template synthesis of 3-DOM IrO2 powder catalysts: Temperature-dependent pore structure and electrocatalytic performance. J. Mater. Sci. 2015, 50, 2984–2992. [Google Scholar] [CrossRef]
  37. Haddadine, N.; Agoudjil, K.; Abouzeid, K.; Castano, C.E.; Bouslah, N.; Benaboura, A.; Samy El-Shall, M. Optical and physical properties of iridescent photonic crystals obtained by self-assembled polymethyl methacrylate nanospheres within graphene oxide nanoplatelets. Polym. Adv. Technol. 2018, 29, 244–253. [Google Scholar] [CrossRef]
  38. Zhao, D.; Yang, P.; Chmelka, B.F.; Stucky, G.D. Multiphase assembly of Mesoporous-Macroporous Membranes. Chem. Mater. 1999, 11, 1174–1178. [Google Scholar] [CrossRef]
  39. Woods, R. Hydrogen adsorption on platinum, iridium and rhodium electrodes at reduced temperatures and the determination of real surface area. J. Electroanal. Chem. Interfacial Electrochem. 1974, 49, 217–226. [Google Scholar] [CrossRef]
  40. Papaderakis, A.; Tsiplakides, D.; Balomenou, S.; Sotiropoulos, S. Probing the hydrogen adsorption affinity of Pt and Ir by surface interrogation scanning electrochemical microscopy (si-secm). Electrochem. Commun. 2017, 83, 77–80. [Google Scholar] [CrossRef]
  41. Rossmeisl, J.; Qu, Z.W.; Zhu, H.; Kroes, G.J.; Nørskov, J.K. Electrolysis of water on oxide surfaces. J. Electroanal. Chem. 2007, 607, 83–89. [Google Scholar] [CrossRef]
Figure 1. (a) Dynamic light scattering (DLS) particle size distribution of PMMA emulsion; (b) SEM surface image of PMMA template; (c) natural fluorescence of PMMA emulsion under natural sunlight; (d) SEM surface image of PMMA template covered by the precursor.
Figure 1. (a) Dynamic light scattering (DLS) particle size distribution of PMMA emulsion; (b) SEM surface image of PMMA template; (c) natural fluorescence of PMMA emulsion under natural sunlight; (d) SEM surface image of PMMA template covered by the precursor.
Polymers 11 00629 g001
Figure 2. (a) High- and (b) low-magnification SEM images of three-dimensional ordered macroporous (3-DOM) IrO2 honeycomb structures; (c) high- and (d) low-magnification SEM images of non-ammonia IrO2 structures.
Figure 2. (a) High- and (b) low-magnification SEM images of three-dimensional ordered macroporous (3-DOM) IrO2 honeycomb structures; (c) high- and (d) low-magnification SEM images of non-ammonia IrO2 structures.
Polymers 11 00629 g002
Figure 3. (a) N2 adsorption/desorption isotherms; (b) BJH pore-size distribution for 3-DOM IrO2 and non-ammonia IrO2.
Figure 3. (a) N2 adsorption/desorption isotherms; (b) BJH pore-size distribution for 3-DOM IrO2 and non-ammonia IrO2.
Polymers 11 00629 g003
Figure 4. XRD patterns of 3-DOM IrO2 and colloidal IrO2.
Figure 4. XRD patterns of 3-DOM IrO2 and colloidal IrO2.
Polymers 11 00629 g004
Figure 5. (a) X-ray photoelectron spectroscopy (XPS) spectra of 3-DOM IrO2; (b) high-resolution spectra of the Ir4f peak.
Figure 5. (a) X-ray photoelectron spectroscopy (XPS) spectra of 3-DOM IrO2; (b) high-resolution spectra of the Ir4f peak.
Polymers 11 00629 g005
Figure 6. (a) Cyclic voltammetry (CV) curves; (b) steady-state polarization curves with 106 µg cm−2 loadings; (c) the current density is normalized to CV; (d) Tafel curves for 3-DOM IrO2 and colloidal IrO2.
Figure 6. (a) Cyclic voltammetry (CV) curves; (b) steady-state polarization curves with 106 µg cm−2 loadings; (c) the current density is normalized to CV; (d) Tafel curves for 3-DOM IrO2 and colloidal IrO2.
Polymers 11 00629 g006
Figure 7. Time evolution currents for two IrO2 materials under a constant potential (1.6 V) in 0.5 mol L−1 H2SO4 solution.
Figure 7. Time evolution currents for two IrO2 materials under a constant potential (1.6 V) in 0.5 mol L−1 H2SO4 solution.
Polymers 11 00629 g007

Share and Cite

MDPI and ACS Style

Liu, F.; Sun, X.; Chen, X.; Li, C.; Yu, J.; Tang, H. Synthesis and Characterization of 3-DOM IrO2 Electrocatalysts Templated by PMMA for Oxygen Evolution Reaction. Polymers 2019, 11, 629. https://doi.org/10.3390/polym11040629

AMA Style

Liu F, Sun X, Chen X, Li C, Yu J, Tang H. Synthesis and Characterization of 3-DOM IrO2 Electrocatalysts Templated by PMMA for Oxygen Evolution Reaction. Polymers. 2019; 11(4):629. https://doi.org/10.3390/polym11040629

Chicago/Turabian Style

Liu, Fei, Xuechu Sun, Xiu Chen, Cuicui Li, Jun Yu, and Haolin Tang. 2019. "Synthesis and Characterization of 3-DOM IrO2 Electrocatalysts Templated by PMMA for Oxygen Evolution Reaction" Polymers 11, no. 4: 629. https://doi.org/10.3390/polym11040629

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