Fabrication of spinel ferrite based alkaline anion exchange membrane water electrolysers for hydrogen production

T. Pandiarajan, L. John Berchmans and S. Ravichandran*
Electro Inorganic Chemicals Division, CSIR-Central Electrochemical Research Institute, Karaikudi-630003, India. E-mail: sravi@cecri.res.in

Received 20th January 2015 , Accepted 23rd February 2015

First published on 23rd February 2015


Abstract

Alkaline anion exchange membrane water electrolysis (AEMWE) is considered to be an alternative to proton exchange membrane water electrolysis (PEMWE), owing to the use of non-noble meta/metal oxides in AEMWE. Here, we report a highly durable and low-cost AEM-based electrolysis cell with active spinel ferrite catalysts for hydrogen production. Ce-substituted MnFe2O4 was synthesized by a combustion method and investigated as the electro catalyst for oxygen evolution reaction (OER). Substitution of Ce in the cubic lattice of MnFe2O4 increases the conductivity of CexMnFe(2−x)O4, which results in a negative shift in the OER onset potential. At 25 °C, the single cell with Ce0.2MnFe1.8O4 exhibited a current density of 300 mA cm−2 at 1.8 V. Notably, Ce0.2MnFe1.8O4 demonstrates a durability of >100 hours in continuous electrolysis.


1. Introduction

The gradual progression of the world toward a serious energy crisis has motivated intense studies on alternative energy conversion and storage systems.1–5 To determine the ultimate renewable energy source for the future, a number of energy conversion and storage techniques have been developed, such as water splitting, fuel cells, and batteries, etc.6–8 One of the most promising energy conversion methods, electrochemical water splitting, has received much attention due to its commercial efficiency and minimal impact on the environment.3,9,10 Alkaline water electrolysis (AWE) is a simple and cheap method for producing hydrogen, which is recognized as an efficient fuel for energy storage systems. Moreover, AWE has dominated large-scale hydrogen production technology for several decades, because of cost-effective electrolyser fabrication and the simple control of corrosion.11–15 However, conventional liquid alkaline electrolysers face a set of challenges, such as limited current density, low operating pressure, and low partial load range.11,16 These challenges are due to the use of an electrolyte of 30% KOH solution in liquid alkaline water electrolysers for the electro catalytic water splitting.

The problems associated with the use of a concentrated alkali are largely related to the formation of carbonate ions, which can decrease the electrolyser lifetime and purity of hydrogen.17–19 As a result, water electrolysis via solid polymer electrolyte (SPE), viz. proton exchange membranes (PEMs), and anion exchange membranes (AEMs) have recently received greater attention. Moreover, an alkaline polymer electrolyte possesses the advantages of both a solid polymer electrolyte and an alkaline pH.

While PEM water electrolysis (PEMWE) offers high current densities, compact electrolyser designs, and high hydrogen production rates, the non-platinum group metal lacks durability in an acidic environment.9,16 Furthermore, the fabrication of the membrane electrode assembly (MEA) for PEMWE is expensive, resulting in resources being diverted to the development of alkaline anion exchange membrane water electrolysis (AEMWE).20,21

The efficiency of electrolysers depends on the MEA, a key component of water electrolysis that hinges on the sluggish kinetic reactions of oxygen evolution (OER) and oxygen reduction reaction (ORR). Although the OER activity of noble metal oxides (Ir, Ru, and Pt) was outstanding in acidic medium, stability problems under alkaline conditions prevent the utilization of noble metal oxides as a practical electro catalyst for OER.22–24

The exhaustive search for non-noble catalysts, conducted over many years, has identified the mixed metal oxides,25,26 specifically the AB2O4 spinel,27–31 spinel ferrites,32–34 organic compounds,35 and perovskites,6,36–38 as potential OER catalysts capable of replacing expensive noble metal-based compounds. A detailed discussion of the electrocatalytic activities of non-noble metal oxides can be found in the cited ref. 37. Moreover, several studies have reported Ni and Co-based materials as excellent catalysts for water oxidation,37 but only in alkaline solution; changing the electrolyte to deionized water significantly decreases the efficiency of electrolysers. To date, only a few studies have addressed in AAEMWE in deionized water. For example, Xu Wu et al. have reported a current density of 100 mA cm−2 at 1.8 V (25 °C) for MEA with Cu0.7Co2.3O4.39 In addition, a current density of 300 mA cm−2 at 2.2–2.5 V with a Li0.21Co2.79O4 anode has been reported.40 More recently, Lin Zhuang and co-workers41 demonstrated a current density of 400 mA cm−2 at 1.85 V with a Ni–Mo-based AEM electrolyser using only deionized water. Similarly, Chao-yang Wang et al.18 report a life-time of >500 h with AEM water electrolysis. However, the Ni–Mo-based AEM electrolyser exhibits a durability of only 8 h, and the latter requires noble metal oxides for long-term stability. Moreover, each of these electrolysers required an operating temperature of 50 to 75 °C.

Therefore, the challenge of developing a non-noble metal catalysts for AEMWE that exhibits a current density of j > 0.5 A cm−2 at an over potential of ηO2 < 0.3 V with long-term stability, at room temperature remains to be achieved.

Among non-noble metal catalysts, spinel ferrites32 are of particular interest because they possess low-cost, high catalytic activity, and durability at high pH; such attributes make spinel ferrites relevant for use in catalysis, sensors, and magnetic devices. Moreover, the spinel ferrite based catalysts demonstrate a great potential for OER33,34,42–44 and ORR,45 which is comparable to commercial Pt. However, in practice, the low electrical conductivity restricts the application of spinel ferrite as a catalyst. Here, we report the performance of a laboratory scale AAEMWE with spinel ferrite catalysts for pure hydrogen production.

2. Experimental section

2.1. Membrane pre-treatment

The Alkaline Anion Exchange Membrane (Fumasep® FAA-3-PK-130) was purchased from the Fumatech company (Germany), delivered in bromide form and dry state. The membrane thickness was 130 μm and the specific conductivity was 7.6 mS cm−1 at 25 °C. For alkaline applications, the membrane must be converted into a hydroxyl form. To achieve this, the membrane was kept in an aqueous solution of 1 M KOH for 24 h at 20–30 °C, and then transferred to deionized water for 1 h. To avoid CO2 contamination, the pre-treated membrane was stored under humidified and CO2 free conditions.

2.2. Ionomer preparation

Instead of using commercial ionomer solution, we used the AAEM dissolved in dimethyl sulfoxide (DMSO) solvent. To prepare AAEM ionomer solution 0.1 g of AAEM was added to 3 ml of DMSO solvent. However, the substrate used as a support for membrane not dissolved completely in the DMSO and settle down as precipitate. So we discard the precipitate and used the solution as ionomer for catalyst ink preparation.

2.3. Catalyst ink preparation

To prepare the catalyst ink, 3.5 mg of the catalyst powder dispersed, in the mixture of 60μ ml AAEM solution, 2 ml of 2-propanol and 30μ ml deionized water. After 30 minutes of sonication, the solution became homogeneous and was used for the MEA preparation.

2.4. MEA preparation

The MEA was fabricated by brush coating the catalyst mixture onto each side of the membrane, which was allowed to dry and hot pressed at 70 °C. The catalyst loadings on the anode and cathode were 3.5 mg cm−2 of Ce0.2MnFe1.8O4 and Ni powder (Sigma-Aldrich) respectively. This was followed by the cell assembly, utilizing a pair of platinum-coated titanium mesh as current collectors. The current collectors were placed on each side of the MEA to ensure proper contact between electrodes and the current carrier plate. Finally, a stainless steel end plate was placed for mechanical support. Fig. 7 shows the schematic diagram of the single cell adopted for this work. The working area of the single cell was 2 cm × 2 cm. The single-cell experiments were carried out using deionized water. Polarisation curves of the single cell with CexMnFe2−xO4 were measured with a scan rate of 1 mV s−1.

3. Results & discussions

3.1. Half-cell studies

In our previous work,46 we performed the initial screening of catalysts in a standard three-electrode system using various concentrations of KOH electrolytes and showed that 0.2 M cerium substitution enhances the electrocatalytic activity of MnFe2O4 for OER in alkaline solution.

The Working electrode for electrochemical studies was prepared by sonicating 3 mg of the catalyst, in 60 μl isopropyl alcohol for 15 minutes. 20 μl of the suspension was pipette out and dried on a Pt metal substrate (1 cm × 1 cm) by heating, in a pre-heated oven at 80 °C. Subsequently the substrate was sintered, at 1200 °C in presence of air for 2 h.

The X-ray diffractograms of the as prepared MnFe2O4 and Ce0.2 Fe1.8O4 are shown in Fig. 1. The diffraction patterns are distinct and sharp, clearly delineating the crystalline nature of the Cex Fe2−xO4. Furthermore, the observed diffraction pattern of MnFe2O4 are good agreement with the spinel-type MnFe2O4 (JCPDS: 073-1964), with a cubic lattice spinel structure. Due to the large ionic radius the incorporation of Ce3+ in the spinel lattice increases the lattice constant value of MnFe2O4[thin space (1/6-em)]47 from a = 8.515 Å to 8.614 Å. The substitution of Ce3+ on MnFe2O4 was confirmed by X-ray photoelectron spectroscopy (XPS).


image file: c5ra01123j-f1.tif
Fig. 1 XRD patterns of CexMnFe2−xO4 with x = 0.0, 0.2, 0.4, 0.6, 0.8.

Fig. 2 illustrates the XPS spectrum of the Ce0.2 Fe1.8O4 after sintering at 1200 °C, in ambient air for 2 hours. The survey spectrum presented in Fig. 2A, shows the compositional elements of the Ce0.2 Fe1.8O4. Moreover, the elemental composition of Ce3+, Mn2+, and Fe3+ were further analyzed with the EDAX spectrum (Fig. 2B).


image file: c5ra01123j-f2.tif
Fig. 2 (a). XPS spectra of the Ce0.2MnFe1.8O4 after sintering (b). EDAX spectra of Ce0.2MnFe1.8O4 after sintering. Reproduced from ref. 46.

Fig. 3A shows the XPS spectra for Mn 2p which provides clear evidence for the presence of Mn2+ in the CexMnFe2−xO4. It can be seen from the spectrum that the peaks at 641.5 and 652.3 eV are caused by Mn 2p3/2 and Mn 2p1/2 respectively, with a satellite peak at 41.3 eV.43,48 In Fig. 3C, the peaks at 711.0 and 724.6 eV were assigned to Fe 2p3/2 and Fe 2p1/2.43 Moreover, two main peaks at 718.5 and 732.5 eV[thin space (1/6-em)]49–51 specifies the presence of Fe3+. Similarly, the characterized peaks with binding energies 845 and 905 eV[thin space (1/6-em)]52–54 in the XPS spectra (Fig. 3B) occurs due to the presence of Ce3+ on MnFe2O4. Thus, the above XPS data clearly suggest that Ce3+ incorporated on MnFe2O4 and sintering the catalysts at 1200 °C, resulted in the reduction of Fe3+ on the ferrite system.


image file: c5ra01123j-f3.tif
Fig. 3 XPS spectra of the catalysts. Insert XPS spectra of (a) Mn2+ (b) Ce3+ (c) Fe2+ & Fe3+ on as prepared Ce0.2MnFe1.8O4 and (d) Fe2+ & Fe3 on Ce0.2MnFe1.8O4 after sintering. Reproduced from ref. 46.

Fig. 4 shows the morphology of the prepared Cex Fe2−xO4 (x = 0.0, 0.2) were analyzed using a scanning electron microscope (SEM). In all cases, we observed the collapsed microstructures of ferrites due to the high-temperature treatment. Further, sintering the samples at 1200 °C increased the agglomeration of ferrites; this resulted in the formation of a thin-film layer.34 Subsequently, it insulates the Pt surface from the electrolyte solution. Therefore, no redox peaks appeared for Pt on cyclic voltammetry studies of samples in KOH solution, which ruled out the synergetic effect or diffusion of Pt on the catalysts layers.42 Fig. 4C & D shows SEM recorded on samples coated on Pt surface and sintered in air for 2 hours at 1200 °C.


image file: c5ra01123j-f4.tif
Fig. 4 SEM images of CexMnFe2−xO4 (x = 0.0, 0.2). Insert (A & B) before sintering and (C & D) after sintering. Reproduced from ref. 46

The cyclic voltammograms (CV) of catalysts were recorded in 1 M KOH solution with a scan rate of 10 mV s−1 at 25 °C. Fig. 5 shows the CV curves recorded between 1.2–1.8 V vs. RHE). No redox peaks were seen in the CV,42 indicating that the ferrite does not undergo any oxidation process in the reaction conditions. However, the OER onset of Ce-substituted manganese ferrite shifts towards more negative than the manganese ferrite.


image file: c5ra01123j-f5.tif
Fig. 5 Cyclic voltammograms of Pt/CexMnFe2−xO4 (x = 0.0 ≤ 0.8) at 25 °C, with scan rate 10 mV s−1 in 1 M KOH solution. Adapted from ref. 46.

The electrochemical impedance spectrum (EIS) of the catalysts recorded during the OER is shown in Fig. 6. The Ce-substituted MnFe2O4 shows smaller charge transfer resistance, compared to MnFe2O4. Such a drastic change in electrochemical behavior is due to the increase in the electronic conductivity of Ce0.2MnFe1.8O4.


image file: c5ra01123j-f6.tif
Fig. 6 Electrochemical impedance spectra of catalysts as a function of Ce substitution, measured at 1.6 V during oxygen evolution. Reproduced from ref. 46.

In general, the electrical conductivity of ferrite is explained by Verwey–de Boer mechanism and the polaron effect. However, the electrical conduction in ferrites originates between Fe2+ and Fe3+ ions at site B, conduction is mostly influenced by the Fe2+ concentration in the ferrite system. Due to the high temperature sintering, some lattice oxygen escapes from the spinel lattice, causing an oxygen deficiency in the spinel ferrites. Therefore, to balance the electrical charge created in the lattice, Fe3+ is reduced to Fe2+,55,56 which facilitates the formation of excess Fe2+ ions in the system (as shown in Fig. 3D). Hence, the hopping rate of electrons in ferrites increases which is the cause of their higher electronic conductivity.55,57

The X-ray photoelectron spectroscopy (XPS) spectrum, shown in Fig. 3C & D, confirms the co-existence of Fe2+ ions with Fe3+ in the spinel lattice. In addition, the replacement of metal ions with rare earth metal ions forms grain boundaries in ferrites but using high concentration (≥0.4) of such metal ions lead to the formation of secondary phases (i.e. ABO3) on the ferrite system. These secondary phases hinder the mobility of charge carriers and increase the electrical resistivity. However, in the case of Ce0.2MnFe1.8O4, the electronic conductivity was further enhanced by the cooperative effect58 between the Ce3+ and Mn2+. The interaction between the electron-rich Mn2+ and electron-deficient Ce3+ ion enhances the Lewis acid properties of the mixed valence centers, as a result, the electrocatalytic activity of Ce0.2MnFe1.8O4 increases.

3.2. Single cell studies

Based on results of half-cell studies, we selected Ce0.2MnFe1.8O4 as a catalyst for single-cell experiments, since Ce0.2MnFe1.8O4 demonstrates an excellent electro-catalytic activity and stability for OER, in alkaline solutions.

In general, the addition of the ionomer (during catalyst ink preparation) to the catalyst layer shows a direct effect on the efficiency of the electrolysers.59,60 This is because the ionomer addition promotes ion (H+ or OH) transport from the bulk of the catalyst layer to the membrane, which consequently enhances the electrolyser efficiency by reducing the interfacial resistance between the membrane and electrode, or the ionomer and catalyst. Conversely, the electrical conductivity of catalysts is decreased by the addition of electron resistant ionomer solutions. Hence, it is important to optimize the ionomer loading for the catalyst ink and MEA preparation.

Furthermore, we modified the current collectors by platinum-coating on Ti mesh for single cell studies. As commercial AEMs possess a lower conductivity than Nafion membranes, single cells based on AEMs are essentially sensitive to the requirement of a good current collector; which should be corrosion resistant, have strong electrical conductivity, and provide mechanical support to the membrane.16,61 Such a current collector should effectively expel the gases and allow the water to reach the catalytic sites in its counter flow. Hence, Pt-coated Ti meshes are better choice, as they display good electrical properties and do not passivate over time. As a result, the performance and durability of the MEA may increase in electrochemical studies.

The electro catalytic performances of single cell with Ce0.2MnFe1.8O4 and MnFe2O4 anodes were first evaluated by recording a Linear Sweep Voltammograms (LSV) at room temperature. Fig. 8 shows the LSV curves of the MEA with Ce0.2MnFe1.8O4 and MnFe2O4 anodes. At 25 °C, the Ce0.2MnFe1.8O4 exhibits an over potential of 320 mV with a current density of 10 mA cm−2, which is 80 mV lower than the MnFe2O4 by passing deionised water. When the Ce0.2MnFe1.8O4 initiates the water splitting at 1.48 V, the electrolyser starts with a current density of 3 mA. Moreover, gradual increase the applied potential increases the current density to a maximum of 300 mA cm−2 at 1.8 V. As shown in Fig. 9, the potential above 1.55 V the charge transfer resistance values of MEA using Ce0.2MnFe1.8O4 shifts significantly to smaller values and it accelerates the water splitting to a greater extent. To validate the synergistic effect or catalytic effect of Pt coated Ti meshes, we measured the polarization curve without catalysts on AAEM (Fig. 8). However, the single cell did not show any significant catalytic activity on water electrolysis, which ruled out the synergistic effect of Pt.


image file: c5ra01123j-f7.tif
Fig. 7 Schematic diagram of the single cell.

image file: c5ra01123j-f8.tif
Fig. 8 Polarisation curves of the MEA with Ce0.2MnFe1.8O4, MnFe2O4 and without catalysts in deionised water at 25 °C.

image file: c5ra01123j-f9.tif
Fig. 9 Nyquist plots for MEA with Ce0.2MnFe1.8O4 as a function of potential from 1.50 to 1.75 V during oxygen evolution reaction.

In contrast, the PEMWES using noble metal oxides will probably yield a maximum current density of >1 A cm−2 under similar voltage, due to the high conductivity of per fluorinated membranes (Nafion). However, our results are superior to the electrolysers with Co-based catalysts such as Cu0.7Co2.3O4[thin space (1/6-em)]39 and Li0.21Co2.79O4[thin space (1/6-em)]40 anodes. Though, the MEA with the above catalysts exhibit promising activity in KOH solution, they show only a small current density performance with AAEM. Most significantly, the electro catalytic activity of Ce0.2MnFe1.8O4 was comparable to the AAE water electrolysers using Ni–Fe as the anode.41 Previously, Hickner and Wang[thin space (1/6-em)]18 reported a maximum current density of 399 mA cm−2 at 1.8 V (50 °C) with IrO2 anode and Pt cathode. However, recently, Zhuang et al. demonstrated the same current density with a slightly higher over potential and operating temperature of 75 °C in AAEMWE using Ni–Fe anode working only with pure water. Although our experiments were carried out at room temperature, we achieved a performance comparable to that of the Ni–Fe anode performance at higher temperature. The electrocatalytic activities of various electrodes using AAEM are given in Table 1.

Table 1 OER activity of various anode materials recorded in deionized water
S. no Anode material Operating potential (V) Temperature (°C) Current density (mA cm−2) Reference
(1) Ce0.2MnFe1.8O4 1.8 25 300 This work
(2) Li0.21Co2.79O4 2.2–2.05 20–45 300 39
(3) IrO2 1.8 50 399 17
(4) Cu0.7Co2.3O4 2.0 25 200 38
(5) Ni–Fe 1.85 70 400 40


The efficiency of MEA depends on the effective utilization of catalysts for electrochemical reactions viz., OER and HER, which are ultimately governed by the quantity of intermediates that reach the catalyst surface. Here, we used an AEM ionomer solution prepared by dissolving pre-treated membrane in DMSO. As a result, which enhances the hydroxyl ions transportation from the bulk catalyst layer to the membrane, and vice versa. Hence, the ionic conduction in the catalyst layer is significantly enhanced due to AAEM ionomer, and improves the electrolyser efficiency. Similarly, the Pt coated Ti meshes improve the durability of the electrolysers by reducing the corrosion of anode components.40,61 Together with the ionomer effect and current collectors, the reduction of Fe3+ at high temperature also contributes for the overall efficiency of the single cell.

To evaluate the kinetic parameters (Table 2), we fit the polarization curves achieved with Ce0.2MnFe1.8O4 and MnFe2O4 to the following equation.

 
η = a + b[thin space (1/6-em)]log[thin space (1/6-em)]j (1)
where η is the over potential and j is the current density.

Table 2 Electrode Kinetic parameters for OER in deionized water
S. no Catalysts Lattice parameter (Å) ηO2 (mV) at current density (j) Tafel slope value (b) mV per decade
10 mA cm−2 100 mA cm−2 300 mA cm−2
1 Ce0.2MnFe1.8O4 8.614 320 460 570 64
2 MnFe2O4 8.515 390 690 910 95


Fig. 10 depicts the Tafel slope plots of MEA with CexMnFe2−xO4 for OER measured in deionized water. The Tafel slope of Ce0.2MnFe1.8O4 is 64 mV per decade, lower than 45 mV per decade obtained on the MEA using MnFe2O4. The observation of lower Tafel slope value for Ce0.2MnFe1.8O4 compared to that of MnFe2O4 indicates the increased electron transfer for OER. A common mechanism for OER listed in eqn (2)–(4) has been used to account for the observed the kinetic parameters.

 
S + OH ↔ S–OH + e (2)
 
S–OH + OH ↔ S–O + H2O + e (3)
 
2S–O ↔ 2S + O2 (4)
where “S” is an active species.


image file: c5ra01123j-f10.tif
Fig. 10 Tafel plots of OER current in Fig. 8.

The Tafel slope (b) value of Ce0.2MnFe1.8O4 indicates that the eqn (3) is the rate determining step for OER. The literature studies39,42 show that if eqn (3) is the rate determining step then it gives a second order reaction in OH and a typical b value of 60 mV per decade. Again, it can be noticed from the discussions that, Ce0.2MnFe1.8O4 was the most active catalyst for the OER among all cerium substituted manganese ferrite.

3.3. Stability studies

The short term stability of the MEA estimated under an electrolysis current of 200 mA cm−2 at room temperature. During this study the cell voltage steadily increased with time (Fig. 11). However, the commercial utilization of any catalyst requires long-term stability toward OER. As shown in Fig. 12, the stability of the Ce0.2MnFe1.8O4 and MnFe2O4 was investigated by running chronoamperometric responses at 2 V (vs. RHE) by circulating deionized water. Current density of >390 mA cm−2 remained constant. There was no apparent drop in the performance of Ce0.2MnFe1.8O4 electrodes after 100 h of operation. Moreover, we analyzed the electrolyte every 10 h to ensure the stability of the catalysts during continuous electrolysis. Atomic absorption spectroscopy (AAS) analysis shows the absence of metal ions, further supporting that these catalysts are stable during AEM water electrolysis.
image file: c5ra01123j-f11.tif
Fig. 11 Chronopotentiometry curve of single with Ce0.2MnFe1.8O4 under current density of 200 mA cm−2.

image file: c5ra01123j-f12.tif
Fig. 12 Current density profile for constant electrolysis of MEA using MnFe2O4 and Ce0.2MnFe1.8O4 with applied potential of 2 V at 25 °C.

4. Conclusion

In summary, we concluded that the enhanced electrocatalytic activity of Ce0.2MnFe1.8O4 results from the incorporation of Ce3+ into the MnFe2O4 spinel lattice. The single cell fabricated with Ce0.2MnFe1.8O4 catalysts exhibit a current density of ∼300 mA cm−2 at 1.8 V with a lifetime of >100 h on constant electrolysis. Moreover, our studies reveal that using spinel ferrite as a catalyst, we can fabricate cost-effective, highly active AEM-based electrolysers for water oxidation. Moreover, studies are on-going to investigate the effect of temperature on water electrolysis and to improve the efficiency of the cell.

Acknowledgements

T. Pandiarajan thanks UGC, India for providing Junior Research Fellowship. We also thank Mr R. Ram Kumar and M. Baskar, CSIR-CECRI, Karaikudi – for the fabrication of single cell. We are greatly indebted to Dr Vijayamohanan K. Pillai, Director, CSIR-CECRI for his keen interest and encouragement.

References

  1. C. A. Rodriguez, M. A. Modestino, D. Psaltis and C. Moser, Energy Environ. Sci., 2014, 7, 3828–3835 CAS.
  2. N. S. Lewis and D. G. Nocera, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 15729–15735 CrossRef CAS PubMed.
  3. D. L. Stojić, M. P. Marčeta, S. P. Sovilj and Š. S. Miljanić, J. Power Sources, 2003, 118, 315–319 CrossRef.
  4. H. B. Gray, Nat. Chem., 2009, 1, 7 CrossRef CAS PubMed.
  5. M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. Mi, E. A. Santori and N. S. Lewis, Chem. Rev., 2010, 110, 6446–6473 CrossRef CAS PubMed.
  6. J. Suntivich, H. A. Gasteiger, N. Yabuuchi, H. Nakanishi, J. B. Goodenough and Y. Shao-Horn, Nat. Chem., 2011, 3, 546–550 CrossRef CAS PubMed.
  7. R. D. Smith, M. S. Prevot, R. D. Fagan, Z. Zhang, P. A. Sedach, M. K. Siu, S. Trudel and C. P. Berlinguette, Science, 2013, 340, 60–63 CrossRef CAS PubMed.
  8. M. W. Kanan and D. G. Nocera, Science, 2008, 321, 1072–1075 CrossRef CAS PubMed.
  9. K. Zeng and D. Zhang, Prog. Energy Combust. Sci., 2010, 36, 307–326 CrossRef CAS PubMed.
  10. A. Marshall, B. Børresen, G. Hagen, M. Tsypkin and R. Tunold, Energy, 2007, 32, 431–436 CrossRef CAS PubMed.
  11. E. Zoulias, E. Varkaraki, N. Lymberopoulos, C. N. Christodoulou and G. N. Karagiorgis, A Review on Water Electrolysis, TCJST, 2004, 4(2), 41–71 Search PubMed.
  12. V. M. Nikolic, G. S. Tasic, A. D. Maksic, D. P. Saponjic, S. M. Miulovic and M. P. Marceta Kaninski, Int. J. Hydrogen Energy, 2010, 35, 12369–12373 CrossRef CAS PubMed.
  13. S. Marini, P. Salvi, P. Nelli, R. Pesenti, M. Villa, M. Berrettoni, G. Zangari and Y. Kiros, Electrochim. Acta, 2012, 82, 384–391 CrossRef CAS PubMed.
  14. A. Manabe, M. Kashiwase, T. Hashimoto, T. Hayashida, A. Kato, K. Hirao, I. Shimomura and I. Nagashima, Electrochim. Acta, 2013, 100, 249–256 CrossRef CAS PubMed.
  15. M. Bodner, A. Hofer and V. Hacker, Wiley Interdiscip. Rev.: Energy Environ., 2014 Search PubMed.
  16. M. Carmo, D. L. Fritz, J. Mergel and D. Stolten, Int. J. Hydrogen Energy, 2013, 38, 4901–4934 CrossRef CAS PubMed.
  17. J. R. Varcoe, M. Beillard, D. M. Halepoto, J. P. Kizewski, S. Poynton and R. C. T. Slade, ECS Trans., 2008, 16, 1819–1834 CAS.
  18. Y. Leng, G. Chen, A. J. Mendoza, T. B. Tighe, M. A. Hickner and C. Y. Wang, J. Am. Chem. Soc., 2012, 134, 9054–9057 CrossRef CAS PubMed.
  19. J. R. Varcoe, P. Atanassov, D. R. Dekel, A. M. Herring, M. A. Hickner, P. A. Kohl, A. R. Kucernak, W. E. Mustain, K. Nijmeijer, K. Scott, T. Xu and L. Zhuang, Energy Environ. Sci., 2014, 7, 3135–3191 CAS.
  20. M. K. Datta, K. Kadakia, O. I. Velikokhatnyi, P. H. Jampani, S. J. Chung, J. A. Poston, A. Manivannan and P. N. Kumta, J. Mater. Chem. A, 2013, 1, 4026 CAS.
  21. R. Subbaraman, N. Danilovic, P. P. Lopes, D. Tripkovic, D. Strmcnik, V. R. Stamenkovic and N. M. Markovic, J. Mater. Chem. C, 2012, 116, 22231–22237 CAS.
  22. T. Reier, M. Oezaslan and P. Strasser, ACS Catal., 2012, 2, 1765–1772 CrossRef CAS.
  23. Y.-H. Fang and Z.-P. Liu, J. Am. Chem. Soc., 2010, 132, 18214–18222 CrossRef CAS PubMed.
  24. Y. Lee, J. Suntivich, K. J. May, E. E. Perry and Y. Shao-Horn, J. Phys. Chem. Lett., 2012, 399–404 CrossRef.
  25. R. L. Doyle, I. J. Godwin, M. P. Brandon and M. E. Lyons, Phys. Chem. Chem. Phys., 2013, 15, 13737–13783 RSC.
  26. R. Subbaraman, D. Tripkovic, K. C. Chang, D. Strmcnik, A. P. Paulikas, P. Hirunsit, M. Chan, J. Greeley, V. Stamenkovic and N. M. Markovic, Nat. Mater., 2012, 11, 550–557 CrossRef CAS PubMed.
  27. Y. Liang, Y. Li, H. Wang, J. Zhou, J. Wang, T. Regier and H. Dai, Nat. Mater., 2011, 10, 780–786 CrossRef CAS PubMed.
  28. M. R. Gao, Y. F. Xu, J. Jiang, Y. R. Zheng and S. H. Yu, J. Am. Chem. Soc., 2012, 134, 2930–2933 CrossRef CAS PubMed.
  29. D. U. Lee, B. J. Kim and Z. Chen, J. Mater. Chem. A, 2013, 1, 4754 CAS.
  30. D. Chanda, J. Hnát, M. Paidar and K. Bouzek, Int. J. Hydrogen Energy, 2014, 39, 5713–5722 CrossRef CAS PubMed.
  31. J. P. Singh and R. N. Singh, J. New Mater. Electrochem. Syst., 2000, 3, 131–139 Search PubMed.
  32. F. Cheng, J. Shen, B. Peng, Y. Pan, Z. Tao and J. Chen, Nat. Chem., 2011, 3, 79–84 CrossRef CAS PubMed.
  33. M. Raghasudha, D. Ravinder and P. Veerasomaiah, Adv. Mater. Phys. Chem., 2013, 03, 89–96 CrossRef.
  34. S. Ida, K. Yamada, T. Matsunaga, H. Hagiwara, Y. Matsumoto and T. Ishihara, J. Am. Chem. Soc., 2010, 132, 17343–17345 CrossRef CAS PubMed.
  35. E. Mirzakulova, R. Khatmullin, J. Walpita, T. Corrigan, N. M. Vargas-Barbosa, S. Vyas, S. Oottikkal, S. F. Manzer, C. M. Hadad and K. D. Glusac, Nat. Chem., 2012, 4, 794–801 CrossRef CAS PubMed.
  36. T. Takeguchi, T. Yamanaka, H. Takahashi, H. Watanabe, T. Kuroki, H. Nakanishi, Y. Orikasa, Y. Uchimoto, H. Takano, N. Ohguri, M. Matsuda, T. Murota, K. Uosaki and W. Ueda, J. Am. Chem. Soc., 2013, 135, 11125–11130 CrossRef CAS PubMed.
  37. D. E. Hall, J. Electrochem. Soc., 1985, 132, 41C–48C CrossRef CAS PubMed.
  38. F. Manzer, C. M. Hadad and K. D. Glusac, Nat. Chem., 2012, 4, 794–801 CrossRef PubMed.
  39. X. Wu and K. Scott, J. Mater. Chem., 2011, 21, 12344 RSC.
  40. X. Wu and K. Scott, Int. J. Hydrogen Energy, 2013, 38, 3123–3129 CrossRef CAS PubMed.
  41. L. Xiao, S. Zhang, J. Pan, C. Yang, M. He, L. Zhuang and J. Lu, Energy Environ. Sci., 2012, 5, 7869 CAS.
  42. N. K. Singh, S. K. Tiwari, K. L. Anitha and R. N. Singh, J. Chem. Soc., Faraday Trans., 1996, 92, 2397 RSC.
  43. Y. Fu, P. Xiong, H. Chen, X. Sun and X. Wang, Ind. Eng. Chem. Res., 2012, 51, 725–731 CrossRef.
  44. C. Iwakura, M. Nishioka and H. Tamura, Nippon Kagaku Kaishi, 1982, 7, 136 Search PubMed.
  45. H. Zhu, S. Zhang, Y. X. Huang, L. Wu and S. Sun, Nano Lett., 2013, 13, 2947–2951 CrossRef CAS PubMed.
  46. T. Pandiarajan, S. Ravichandran and L. J. Berchmans, RSC Adv., 2014, 4, 64364–64370 RSC.
  47. L. J. Berchmans, M. P. I. Devi and K. Amalajyothi, Int. J. Self-Propag. High-Temp. Synth., 2009, 18, 11–14 CrossRef CAS.
  48. Y. Gorlin, B. Lassalle-Kaiser, J. D. Benck, S. Gul, S. M. Webb, V. K. Yachandra, J. Yano and T. F. Jaramillo, J. Am. Chem. Soc., 2013, 135, 8525–8534 CrossRef CAS PubMed.
  49. M. Gong, Y. Li, H. Wang, Y. Liang, J. Z. Wu, J. Zhou, J. Wang, T. Regier, F. Wei and H. Dai, J. Am. Chem. Soc., 2013, 135, 8452–8455 CrossRef CAS PubMed.
  50. L. Zhang and Y. Wu, J. Nanomater., 2013, 2013, 1–6 Search PubMed.
  51. A. Mukherjee, S. Basu, P. K. Manna, S. M. Yusuf and M. Pal, J. Mater. Chem. C, 2014, 2, 5885–5891 RSC.
  52. M. Jean-Baptiste, J. Phys. D: Appl. Phys., 2013, 46, 143001 CrossRef.
  53. J. Hou, W. Jiang, Y. Fang and F. Huang, J. Mater. Chem. C, 2013, 1, 5892 RSC.
  54. D. Channei, B. Inceesungvorn, N. Wetchakun, S. Ukritnukun, A. Nattestad, J. Chen and S. Phanichphant, Sci. Rep., 2014, 4 Search PubMed.
  55. Y. Fukuda, S. Nagata and K. Echizenya, J. Magn. Magn. Mater., 2004, 279, 325–330 CrossRef CAS PubMed.
  56. A. Tawfiks, Y. Atassi and I. S. Daewish, Latv. J. Phys. Tech. Sci., 2005, N3, 68–74 Search PubMed.
  57. B. Parvatheeswara Rao and K. H. RAO, J. Mater. Sci., 1997, 32, 6049–6054 CrossRef.
  58. D. W. Stephan, Coord. Chem. Rev., 1989, 95, 41 CrossRef CAS.
  59. M. Mamlouk and K. Scott, J. Power Sources, 2012, 211, 140–146 CrossRef CAS PubMed.
  60. Y. Zhang, C. Wang, N. Wan, Z. Liu and Z. Mao, Electrochem. Commun., 2007, 9, 667–670 CrossRef CAS PubMed.
  61. H. Ito, T. Maeda, A. Nakano, A. Kato and T. Yoshida, Electrochim. Acta, 2013, 100, 242–248 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra01123j

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.