Next Article in Journal
Combined Cross-Linked Enzyme Aggregates as Biocatalysts
Next Article in Special Issue
Earth-Abundant Electrocatalysts in Proton Exchange Membrane Electrolyzers
Previous Article in Journal
Selective Catalytic Reduction of NOx
Previous Article in Special Issue
Hydrogen Oxidation on Ni-Based Electrocatalysts: The Effect of Metal Doping
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxygen Reduction Reaction Electrocatalysis in Alkaline Electrolyte on Glassy-Carbon-Supported Nanostructured Pr6O11 Thin-Films

by
Rakesh K. Sharma
1,
Verónica Müller
1,
Marian Chatenet
1,2,* and
Elisabeth Djurado
1,*
1
LEPMI, Université Grenoble Alpes, CNRS, Grenoble-INP (Institute of Engineering Univ. Grenoble Alpes), LEPMI, F-38000 Grenoble, France
2
Institut Universitaire de France, IUF, 75231 Paris Cedex 05, France
*
Authors to whom correspondence should be addressed.
Submission received: 18 September 2018 / Revised: 10 October 2018 / Accepted: 12 October 2018 / Published: 17 October 2018
(This article belongs to the Special Issue Catalysts for Polymer Membrane Fuel Cells)

Abstract

:
In this work, hierarchical nanostructured Pr6O11 thin-films of brain-like morphology were successfully prepared by electrostatic spray deposition (ESD) on glassy-carbon substrates. These surfaces were used as working electrodes in the rotating disk electrode (RDE) setup and characterized in alkaline electrolyte (0.1 M NaOH at 25 ± 2 °C) for the hydrogen evolution reaction (HER), the oxygen evolution reaction (OER), and the oxygen reduction reaction (ORR) for their potential application in alkaline electrolyzers or in alkaline fuel cells. The electrochemical performances of these electrodes were investigated as a function of their crystallized state (amorphous versus crystalline). Although none of the materials display spectacular HER and OER activity, the results show interesting performances of the crystallized sample towards the ORR with regards to this class of non-Pt group metal (non-PGM) electrocatalysts, the activity being, however, still far from a benchmark Pt/C electrocatalyst.

1. Introduction

With the development of renewable electricity—a necessity to face the present fossil energy crisis and limit the harmful emissions of related greenhouse gases—the means to store this energy upon production peaks and to release it upon demand peaks is mandatory [1]. One very efficient manner to do so lies in the so-called hydrogen economy, where H2 (and O2) is produced in water electrolyzers (acidic [2], alkaline [3], or, if necessary, with advanced physical procedures to boost the reactions [4]), stored, and then converted back to water and electrical energy in fuel cells. The so-called unitized regenerative fuel cells (URFC) [5,6], as well as photocatalytic water-splitting devices [7,8] are also convenient means to do so. In these systems, the reactions of oxygen reduction (ORR) and oxygen evolution (OER) limit the operation in fuel cell and electrolyzer mode, respectively, whatever the pH of operation. Moreover, the electrocatalysts that can sufficiently and durably accelerate these reactions in acidic (proton-exchange membrane) conditions are based on platinum group metals (PGM) [9,10], which conveys dramatic issues in terms of availability and cost of the electrodes. As a result, the sustained hydrogen cycle might only be viable for alkaline systems for which the potential choice of more available and less costly non-noble electrocatalysts is greater (and simply possible) [11]. There are extremely vast numbers of potential non-PGM electrocatalysts for the ORR [12,13,14,15,16,17,18] and OER [19,20,21,22,23,24,25,26,27,28] (these examples being by-no-means comprehensive), and metal oxides in the perovskite structure play a large role in this hot area, owing to their non-negligible activity for the oxygen reduction and evolution reactions, and the extremely large variety of components available [6,11,29,30,31,32]. Perovskite oxides, usually based on rare-earth component(s), are usually not sufficiently electron-conductive on their own to exhibit sufficient performances, and they are, therefore, generally used in composite electrodes while including some carbon additive [32,33,34,35,36]. Addition of carbon is usually very positive for the ORR [37,38,39,40,41] and results from a combination of effects like (i) an increase of electronic conductivity of the active layer, (ii) a bifunctional mechanism in which carbon enhances hydrogen peroxide formation, the latter being decomposed/reduced on the perovskite oxide, and (iii) a change of the perovskite electronic and/or crystalline structure(s) promoted by carbon [32,35]. On the contrary, adding carbon is detrimental for the OER as a result of its poor durability in operation [32,37,38]. Nevertheless, very interesting performances have been reached to date, both for the ORR and OER on this class of materials [31,42], but it is unclear from these studies what the intrinsic activity of the perovskite component is. More recently, carbon-free active layers based on perovskite oxides have been described and tested, which shed light on the intrinsic electrocatalytic properties of the electrocatalysts, and highlighted the role of their electronic conductivity and lattice strain, two parameters which need a certain balance to reach high ORR and OER activity [43]. The importance of the crystalline nature of the perovskites has been ascertained, with smaller crystals yielding larger ORR activity [44]. The influence of the facets has also been put forth [45] as well as the crystalline versus amorphous nature of the oxide [46,47], but in these cases, composite electrodes using carbon additives were used.
Rare-earth oxides are also popular in the field of high-temperature solid-state electrochemistry and in particular for solid oxide fuel cells (SOFCs). For example, Pr6O11 has been reported to enhance electrochemical ORR performance when used as an additive in composite oxygen electrodes: Pr6O11-infiltrated LaNi0.6Fe0.4O3-δ (LNF) electrodes exhibited improved performance over LNF (LaNi0.6Fe0.4O3-δ) electrodes, as reported by Chiba et al. [48] and Ding et al. [49]. More recently, polarization resistances as low as 0.026 Ω cm2 at 600 °C were reported for a pure Pr6O11 columnar-type cathode prepared by electrostatic spray deposition (ESD) [50]. The ESD technique is a film-coating method based on the principle of electrostatic atomization [51]. First, a positive high voltage is applied to the stainless steel nozzle from which positively-charged droplets of a solution of precursors are generated. Then, they are directed to the grounded substrate thanks to the electrical field. Finally, the size of the droplets at the moment of impact on the heated substrate and the combined effects of relative spreading and drying rates will determine film morphology. A schematic drawing of the ESD system has been presented in Reference [52]. This innovative process is a low-cost and versatile technique to elaborate a wide variety of oxide materials with large reproducibility and controlled microstructures [50,53,54,55,56,57].
In the present contribution, Pr-based oxides (Pr6O11) have been prepared for the first time as thin-films on glassy-carbon substrates by ESD; these electrodes, prepared without any addition of carbon black additive, were tested for their activity of hydrogen and oxygen evolution reactions (HER and OER), and oxygen reduction reactions (ORR) in dilute alkaline electrolyte. The oxide crystallinity, in particular, has been evaluated by comparing the fate of amorphous and crystalline Pr6O11.

2. Results and Discussion

2.1. Microstructural Characterization and Elemental Analysis

The microstructures of the films deposited with a flow rate of Q = 1.5 mL h−1, a substrate temperature of Ts = 230 °C, a nozzle-to-substrate distance of dns = 20 mm, and a deposition time of t = 3 h, are shown in Figure 1A–F. In these conditions, both as-prepared (amorphous) and sintered (crystalline) Pr6O11 samples showed similar brain-like type microstructures (Figure 1). The average grain size was too small to be resolved by Scanning Electron Microscopy, SEM (Figure 1C,F), and was therefore determined using Transmission Electron Microscopy, TEM, and X-Ray Diffraction, XRD, analyses (see Section 2.2.). Such microstructure was found reproducible via five replicate coatings.
With the ESD deposition technique, the microstructure of the film is controlled via the size of the droplets that impact the substrate, which depends on the solution properties (solvent density ρ , surface tension γ, boiling point b.p., and electrical conductivity K) and the deposition parameters (flow rate Q, substrate temperature T, nozzle-to-substrate distance dns, and deposition time t) as per following the Gañan-Calvo’s relationship (Equation (1)) [58]
d droplet ~ ( ε 0 ρ Q 3 γ K ) 1 / 6  
where ddroplet is the droplet diameter, ε0 is the vacuum permeability and Q is the flow rate.
The physicochemical properties of the precursor solution also play a crucial role on evaporation of the solvents during the transport and spreading of the droplets when they impact the substrate, resulting in different microstructures [51,56,57]. In these ESD conditions, the boiling point of the solution of precursors, around T = 200 °C, is close to the deposition temperature (i.e., that of the substrate, Ts = 230 °C). Large droplets full of liquid are therefore expected, according to Equation (1), especially when a high flow rate of Q = 1.5 mL h−1 is used. In addition, at a short nozzle-to-substrate distance of dns = 20 mm, these droplets do not have time to evaporate, and instead spread on the heated substrate upon impact. Since the drying step is occurring simultaneously with the boiling step, mechanical stress is generated by drying and contraction of the film on the heated substrate. This drives the film to severe cracks and a “brain-like” type morphology could be observed (Figure 1A,B,E,F). The average thickness of the amorphous (Figure 1D) and crystalline (Figure 1H) samples was found near similar at around 16 ± 2 µm. The mass of the samples was measured by simply weighing the RDE tips before and after deposition. The values obtained were close to the sensitivity of the scale used, so only estimates of the samples weight could be given: 0.7 ± 0.3 mg was obtained for both the amorphous and crystalline samples. These values led to an apparent density of the deposits of ca. 2.23 g cm−3, a value ca. three times smaller than the expected density of Pr6O11 (6.55 g cm−3), in agreement with the existence of macro-porosity in the thin-films (Figure 1A,B,E,F), with porosity on the smaller scale being monitored by TEM (see Section 2.2). The authors point out that an essential understanding of the brain-like type microstructure observed in Figure 1A–H would require more in-depth experimental and theoretical work, which is beyond the scope of the present contribution. The X-ray energy dispersive (X-EDS) spectra in Figure 2 confirmed the presence of Pr and O elements in both films, the amorphous one and the crystalline one, as expected. No noticeable differences in film composition were observed upon crystallization.

2.2. Structural Characterization

X-ray diffraction patterns of both the as-prepared film (amorphous) and the one sintered in air for 2 h at T = 550 °C (crystalline) are shown in Figure 3. As shown in Figure 3A, the as-prepared film was indeed amorphous, whereas all diffraction peaks of the sintered one (Figure 3B) were indexed with the fluorite cubic structure of Pr6O11 (JCPDS 00-042-1121) crystallizing in the Fm 3 ¯ m space group (N° 225), and of the glassy-carbon support. No secondary phases were observed within the detection limits of the powder XRD. The unit-cell parameter of the fluorite cubic structure Pr6O11 was a = 5.4678(0) Å. The average crystallite size calculated by using Scherrer equation for (111), (200), (220) and (311) diffraction peaks was about 16 nm for the films sintered at T = 550 °C.
The microstructure was further investigated by TEM, high-resolution transmission electron microscopy (HRTEM), and selected area electron diffraction (SAED) as shown in Figure 4. For this, the powder was scratched from the brain-like type film sintered at T = 550 °C (crystalline). The brain-like morphology of the crystalline Pr6O11 consists of numerous interconnected nanoparticles of ca. 10–20 nm on average (Figure 4A), in agreement with the XRD data. The HRTEM image in Figure 4B shows distinct lattice fringes which revealed the high crystallinity of the Pr6O11 coating. In fact, the SAED pattern (Figure 4C) shows the diffraction rings for the (111), (200), (220), (311) and (222) crystal planes of fluorite cubic structure of polycrystalline Pr6O11 phase, in agreement with the XRD results.

2.3. Electrochemical Properties

2.3.1. Pr6O11 As-Prepared (Amorphous)

The behavior of the amorphous (as-prepared) Pr6O11/GC sample (no annealing for crystallization, hereafter simply denoted amorphous Pr6O11) in supporting electrolyte and under inert atmosphere is presented on Figure 5 together with the voltamperograms showing incursions to low (HER) and high (OER) potential values. Obviously, the alkaline HER and OER activity of this sample was minor versus the state-of-the-art for non-noble electrocatalysts in the literature [59,60,61,62,63,64]. For this reason, focus on the ORR activity of the amorphous Pr6O11 electrocatalyst will be made hereafter (Figure 6). The onset potential of the ORR is ca. Eonset = 0.835 V vs. RHE, in line with the performances reached by Poux et al. [33,36,65] for similar types of ORR catalytic materials in absence of carbon electron-conductive additives. However, the reaction kinetics was rather sluggish, and the limiting plateau was ill-defined (Figure 6); the latter appeared at smaller current densities compared to the study of Poux et al., which can be connected to the small loading of amorphous Pr6O11 electrocatalyst immobilized at the RDE tip in the present case (ca. 700 µg, see above). To be more specific, the limiting current densities monitored were far below (ca. a factor of 15) the ones expected if the ORR was complete and involving 4 electrons per O2 molecule (in this case, the limiting current densities of ORR should be ca. 3.1, 4.6, 6.2, and 7.7 mA cm−2geometric at RDE revolution rates of Ω = 400, 900, 1600, and 2500 rpm, respectively, these values being calculated using the Levich equation (Equation (2)):
jlim = 0.62 n F D2/3 C v−1/6 Ω1/2
with n = 4, F = 96,500 C mol−1, D = 1.9 × 10−5 cm2 s−1 [66], C = 1.3 µmol cm−3 [66] and v = 10−2 cm2 s−1, for a 0.1 M NaOH solution at T = 25 °C).
This lower-than-expected limiting current density could be ascribed to an insufficient loading of amorphous Pr6O11 (though the authors think it is not likely given the coverage of the glassy-carbon disks being near-complete (see Figure 1)), but also to the fact that (i) the ORR predominantly proceeds according to a 2-electron pathway generating HO2 instead of OH, or (ii) that the electronic conductivity of the amorphous Pr6O11 layer was insufficient to enable its complete operation (in other words, it is likely that not all the catalytic sites could operate owing to their “electrical disconnection” from the GC current collector, the effect being emphasized here because no carbon black additive was used). This will not be further investigated here, as the results of ORR on amorphous Pr6O11 were not sufficiently convincing for an application in alkaline fuel cell cathodes.
Besides, the amorphous Pr6O11 did not show adequate tolerance to NaBH4 (Figure 7), as non-negligible NaBH4 oxidation current was witnessed in nearly all ORR potential regions. This result is disappointing, as non-noble metals (oxides) usually show remarkable tolerance to strong reducers (fuels), e.g., rare-earth perovskites [67], monoatomic metal complexes [68,69,70,71], manganese oxides [16,69,72,73], or cobalt oxides [74].
As a result, this material does not reach sufficient performances for application in practical alkaline fuel cell systems. Therefore, the focus will be made on samples heat-treated at a sufficiently high temperature to enable crystallization of the Pr6O11 phase.

2.3.2. Pr6O11 Sintered (Crystalline)

The behavior of Pr6O11 heat-treated for 2 h at T = 550 °C (sintering/annealing that enabled crystallization of the Pr6O11 phase, see previous sections, herein denoted crystalline Pr6O11) in supporting electrolyte and under an inert atmosphere is presented on Figure 8. As for the as-prepared (amorphous) sample, the HER and OER activity were too small to render the material competitive versus the present state-of-the-art, even though the performances were slightly better compared to the amorphous sample. Therefore, the focus will be made on the ORR activity of the material.
The ORR activity of the crystalline Pr6O11 sample was better than that of the as-prepared one (Figure 9A), as clearly revealed when directly superposing the two sets of data on the same axes (Figure 9B). While the onset potential value (Eonset = 0.835 V vs RHE, in the average for such class of materials [46,75,76,77,78,79,80,81], although in these papers, the materials were tested with carbon as electron-conductive additive) was barely changed compared to the amorphous sample, the limiting current plateaus were better defined. However, they still do not match the expected values for a 4-electron ORR, by a factor ca. 4, a behavior also noted by Poux et al. in the absence of electron-conductive additives in their inks [33,36] and by Fabbri et al. [35]. These authors linked this surprising behavior to multiple (positive) effects in the presence of carbon: (i) an enhanced electronic conductivity; (ii) a bifunctional effect between carbon (that generates HO2) and the perovskite oxide (that converts it to OH); and also (iii) a ligand effect that positively modulates the electronic structure of the metal oxides. Whatever the positive influence of crystallization noted here, one can nevertheless affirm that it is still not approaching the activity of the best rare-earth oxide-based electrodes for alkaline ORR [33,36,65], and is even less than that of a state-of-the-art Pt/C electrocatalyst (Figure 9C). In addition, as was for its amorphous counterpart, the tolerance of the crystalline Pr6O11 to strong reducers was not evident, as NaBH4 oxidation currents were monitored on the whole ORR potential range (Figure 10); this finding is in opposition to the results of Yang et al. who used a LaNi0.8Co0.2O3 ORR electrocatalyst [67].
These results were reproduced for at least two different electrodes in each case: they show that crystallization of the Pr6O11 phase by 2 h annealing at T = 550 °C is beneficial to the ORR activity of the sample. The so-obtained sample, nevertheless, did not approach the performances of state-of-the-art rare-earth oxide-based ORR electrocatalysts reported in the literature [33,36,42,59,65,67], but performed in full-absence of carbon additive to enhance the electronic conduction of the electrode. Although this last property is clearly a step forward in the practical application of such materials in alkaline fuel cell cathodes, more work is needed to obtain an electrocatalyst that approaches Pt-group metal-based ORR electrocatalysts. One possibility would be that the present deposits were too thick to fully benefit from the high electronic conductivity of the glassy-carbon substrate and the outer surface of the Pr6O11 sample; this shall be tested in the future for thinner (but well-covering) films at the GC substrate.

3. Materials and Methods

3.1. Pr6O11 Film Preparation

Pr6O11 films have been deposited on glassy-carbon substrates (0.5 mm in diameter, Sgeometric = 0.196 cm²) by Electrostatic Spray Deposition (ESD), a technique based on the principle of electrostatic atomization [51,53,54,55,56,57,82,83,84,85,86]. The ESD technique is basically a simple three-step process, as described hereafter. The first step consists of the creation of an aerosol from the precursor solution at the needle tip by applying a high dc voltage between the tip and the substrate. The second step is the transportation of the aerosol from the needle to the surface of the substrate in the electric field. The third and last step is the formation of the coating from the impact of the droplets on the heated substrate. The last two steps play a major role in the microstructure of the coating since the droplet size has to be well-controlled versus the ESD process parameters. For this, a praseodymium nitrate hexahydrate (Pr(NO3)3·6H2O, 99.9%, Aldrich) based precursor solution of 0.02 M concentration was prepared in a mixture of ethanol (EtOH, CH3CH2OH, >99.9%, VWR Chemicals, referred as EtOH) and butyl carbitol (diethylene glycol monobutyl ether CH3(CH2)3(OC2H4)2OH, Acros Organics, 99+%, referred to as BC) at an EtOH:BC ratio of 1:2, under vigorous stirring. In this contribution, deposition has been carried out at a nozzle-to-substrate distance (dns), flow rate (Q), substrate temperature (Ts), and deposition time (t) of, 20 mm, 1.5 mL h−1, 230 °C and 3 h, respectively. During the ESD, the voltage was fixed to approximately 5–7 kV. The as-prepared films (noted Pr6O11/GC) were found to be amorphous and a subsequent sintering at T = 550 °C in air for 2 h led to the crystallization of the films. The area of the deposited Pr6O11 was 0.196 cm2, corresponding to a full-coverage of the glassy-carbon tip of the rotating disk electrode (RDE). It is worth adding that the Pr6O11/GC films prepared herein did not contain electron-conductive additives (e.g., carbon black), which is often encountered in the literature and known to modulate the apparent activity of such rare-earth-based samples [32,33,34,35,36].

3.2. Microstructural and Electrochemical Characterization

X-ray diffraction (XRD, Philips X’Pert-MPD system, Cu Kα radiation, λ = 1.54056 Å) in the Bragg–Brentano configuration was used to determine the crystal structure and purity of the prepared films. Phase identification was carried out using the International Centre for Diffraction Data (ICDD), the lattice parameters being determined from the refined X-ray pattern by profile matching using the Fullprof software [87]. The surface-morphology of each film was analyzed by scanning electron microscopy (ZEISS Ultra 55 instrument, Jena, Germany) with a field emission gun (FEG) and X-ray energy-dispersive spectroscopy (XEDS) for composition analysis. High-resolution transmission electron microscopy (HRTEM) and selective area electron diffraction (SAED) were also performed on a JEOL 2010 (JEOL, Tokyo, Japan) equipped with a LaB6 filament to evaluate the particle size and crystal structure. For the TEM analysis, the samples were prepared by scratching the surface of amorphous or crystalline Pr6O11 using the diamond tip, and immobilizing the scratched Pr6O11 on a copper (Cu) holey carbon grid.
Electrochemical measurements of the activity of both amorphous and crystalline Pr6O11 electrodes were performed towards usual reactions of electrocatalysis (namely the hydrogen evolution reaction (HER), oxygen evolution reaction (OER), and the oxygen reduction reaction (ORR), the latter one performed in absence or presence of a strong reducer in solution (NaBH4) for potential application in direct borohydride fuel cell cathodes [16,72,88,89,90]). They were performed in a Pyrex®-based four-electrode cell, with glassy-carbon as the counter electrode (to avoid any possible contamination of the working electrode (WE) with the dissolution of the, usually Pt-based, counter electrode [91], a very likely occurrence in base electrolytes [92,93,94,95]), Pt as the auxiliary electrode, a hydrogen reference electrode (all the potential values are expressed on the RHE scale hereafter) and Pr6O11/GC (either bare (amorphous) or heat-treated (crystalline)) as the WE. The cell is equipped with an internal PTFE beaker to limit contamination by the possible dissolution of the Pyrex in strong base [96]. The potentiostat used was a multichannel VMP3 from Bio-Logic® piloted by the EC-lab software. In any case, the cell and glassware were soaked in Caro’s acid overnight and thoroughly rinsed using ultrapure water (18.2 MΩ cm, <3 ppb total organic carbon, Millipore®, Billerica, MA, USA) prior to each measurement. All the electrochemical experiments were performed using a dynamic correction of the Ohmic-drop, the high-frequency resistance being measured prior to each experiment at 10,000 Hz and automatically corrected at 85%.
An aqueous solution of 0.1 M NaOH at T = 25 ± 2 °C for ORR, OER, and HER study, and aqueous solution of 0.1 M NaOH + 10−2 M NaBH4 aqueous solution (T = 25 ± 2 °C) for borohydride oxidation reaction (BOR) and ORR in the presence of BH4 for tolerance study, have been used. The aqueous solutions were saturated by bubbling Ar or O2 in solution (4N purity, Messer, Bad Soden, Germany), depending on the test. All the salts used to prepare the electrolytes were purchased in Suprapur® quality (Merck, Darmstadt, Germany) and used without further purification. The electrolytes were prepared using ultrapure water.

4. Conclusions

In this work, we have successfully prepared brain-like type hierarchical nanostructured Pr6O11 by electrostatic spray deposition (ESD) on glassy-carbon tips of rotating disk electrodes (RDE). These materials have been tested in liquid alkaline electrolyte (0.1 M NaOH) in view of their potential application in alkaline electrolyzers or alkaline fuel cells. The as-prepared Pr6O11 film is amorphous and shows minimal HER, OER, and ORR activities in 0.1 M NaOH electrolyte. The amorphous nature of the film disappears upon 2 h sintering in air at T = 550 °C, leading to crystalline Fm 3 ¯ m fluorite cubic structure. In that case, the brain-like type microstructure consists of agglomerated crystalline nanoparticles, with sizes in the range of ca. 10–20 nm on average. The crystalline Pr6O11 exhibits better ORR activity than its amorphous counterpart but still fails to direct the reaction towards a complete 4-electron pathway and to reach the activity of the best non-PGM electrocatalysts reported in the literature (and even less than that of PGM-based electrocatalysts). In addition, the material does not exhibit significant HER and OER activities and does not seem tolerant to BH4 fuel. Nevertheless, the significant ORR performances reached were obtained in the absence of any electron-conductive additive, which grants the material some interest for potential use in alkaline fuel cell cathodes and leaves hope for improvements in activity if the electronic conductivity of the oxide layer is improved and/or if its thickness is decreased.

Author Contributions

R.S. performed the materials syntheses, some of their materials’ characterizations, some of the data analysis, and participated in the writing. V.M. performed some of the materials’ characterizations. M.C. performed the electrochemical analyses and made the corresponding data analysis. E.D. and M.C. equally performed some of the materials’ characterizations, some of the data analysis, and participated in the writing of the contribution.

Funding

This research received no external funding.

Acknowledgments

This work was performed within the framework of the Centre of Excellence of Multifunctional Architectured Materials “CEMAM” n° AN-10-LABX-44-01funded by the “Investments for the Future” Program. The authors would like to thank S. Coindeau and T. Encinas for XRD and R. Martin for SEM and X-EDS analyses in CMTC (Grenoble INP, France). MC thanks the French IUF for its support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ghoniem, A.F. Needs, resources and climate change: Clean and efficient conversion technologies. Prog. Energ. Comb. Sci. 2011, 37, 15–51. [Google Scholar] [CrossRef]
  2. Carmo, M.; Fritz, D.L.; Mergel, J.; Stolten, D. A comprehensive review on PEM water electrolysis. Int. J. Hydrogen Energy 2013, 38, 4901–4934. [Google Scholar] [CrossRef]
  3. Marini, S.; Salvi, P.; Nelli, P.; Pesenti, R.; Villa, M.; Berrettoni, M.; Zangari, G.; Kiros, Y. Advanced alkaline water electrolysis. Electrochim. Acta 2012, 82, 384–391. [Google Scholar] [CrossRef]
  4. Niether, C.; Faure, S.; Bordet, A.; Deseure, J.; Chatenet, M.; Carrey, J.; Chaudret, B.; Rouet, A. Improved water electrolysis using magnetic heating of FeC–Ni core–shell nanoparticles. Nat. Energy 2018, 3, 476–483. [Google Scholar] [CrossRef]
  5. Chen, G.; Bare, S.R.; Mallouk, T.E. Development of Supported Bifunctional Electrocatalysts for Unitized Regenerative Fuel Cells. J. Electrochem. Soc. 2002, 149, A1092–A1099. [Google Scholar] [CrossRef]
  6. Narayan, S.R.; Manohar, A.K.; Mukerjee, S. Bi-functional oxygen electrodes challenges and prospects. Electrochem. Soc. Interface 2015, 24, 65–69. [Google Scholar] [CrossRef]
  7. Liao, P.; Carter, E.A. New concepts and modeling strategies to design and evaluate photo-electro-catalysts based on transition metal oxides. Chem. Soc. Rev. 2013, 42, 2401–2422. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, X.; Bieberle-Hütter, A. Modeling and Simulations in Photoelectrochemical Water Oxidation: From Single Level to Multiscale Modeling. ChemSusChem 2016, 9, 1223–1242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Gasteiger, H.A.; Vielstich, W.; Yokokawa, H. Handbook of Fuel Cells; John Wiley & Sons Ltd.: Chichester, UK, 2009. [Google Scholar]
  10. Vielstich, W.; Lamm, A.; Gasteiger, H.A. Handbook of Fuel Cells; Wiley: Chichester, UK, 2003. [Google Scholar]
  11. Jörissen, L. Bifunctional oxygen/air electrodes. J. Power Sour. 2006, 155, 23–32. [Google Scholar] [CrossRef]
  12. Vidakovic, T.R.; Avramov-Ivic, M.L.; Ivanovic, V.Z.; Zecevic, S.K.; Nikolic, B.Z. Electrochemical study of metal-tetrasulfonated phthalocyanines adsorbed on gold, platinum and iron electrodes and their role in electrochemical oxygen reduction. J. Serbian Chem. Soc. 1998, 63, 41–51. [Google Scholar]
  13. Chatenet, M.; Génies-Bultel, L.; Aurousseau, M.; Durand, R.; Andolfatto, F. Oxygen reduction on silver catalysts in solutions containing various concentrations of sodium hydroxide—Comparison with platinum. J. Appl. Electrochem. 2002, 32, 1131–1140. [Google Scholar] [CrossRef]
  14. Roche, I.; Chainet, E.; Chatenet, M.; Vondrak, J. Carbon-supported manganese oxide nanoparticles as electrocatalysts for the Oxygen Reduction Reaction (ORR) in alkaline medium: Physical characterizations and ORR mechanism. J. Phys. Chem. C 2007, 111, 1434–1443. [Google Scholar] [CrossRef]
  15. Piana, M.; Catanorchi, S.; Gasteiger, H.A. Kinetics of Non-Platinum Group Metal Catalysts for the Oxygen Reduction Reaction in Alkaline Medium. ECS Trans. 2008, 16, 2045–2055. [Google Scholar]
  16. Garcia, A.C.; Lima, F.H.B.; Ticianelli, E.A.; Chatenet, M. Carbon-supported nickel-doped manganese oxides as electrocatalysts for the oxygen reduction reaction in the presence of sodium borohydride. J. Power Sources 2013, 222, 305–312. [Google Scholar] [CrossRef]
  17. Benhangi, P.H.; Alfantazi, A.; Gyenge, E. Manganese Dioxide-based bifunctional oxygen reduction/evolution electrocatalysts: Effect of perovskite doping and potassium ion insertion. Electrochim. Acta 2014, 123, 42–50. [Google Scholar] [CrossRef]
  18. Tarasevich, M.R.; Davydova, E.S. Nonplatinum cathodic catalysts for fuel cells with alkaline electrolyte (Review). Rus. J. Electrochem. 2016, 52, 193–219. [Google Scholar] [CrossRef]
  19. Singh, R.N.; Madhu; Awasthi, R.; Tiwari, S.K. Iron molybdates as electrocatalysts for O2 evolution reaction in alkaline solutions. Int. J. Hydrogen Energy 2009, 34, 4693–4700. [Google Scholar] [CrossRef]
  20. Moureaux, F.; Stevens, P.; Toussaint, G.; Chatenet, M. Development of an oxygen-evolution electrode from 316L stainless steel: Application to the oxygen evolution reaction in aqueous lithium–air batteries. J. Power Sources 2013, 229, 123–132. [Google Scholar] [CrossRef]
  21. Suntivich, J.; Perry, E.; Gasteiger, H.; Shao-Horn, Y. The Influence of the Cation on the Oxygen Reduction and Evolution Activities of Oxide Surfaces in Alkaline Electrolyte. Electrocatalysis 2013, 4, 49–55. [Google Scholar] [CrossRef]
  22. Gerken, J.B.; Shaner, S.E.; Massé, R.C.; Porubsky, N.J.; Stahl, S.S. A survey of diverse earth abundant oxygen evolution electrocatalysts showing enhanced activity from Ni-Fe oxides containing a third metal. Energy Environ. Sci. 2014, 7, 2376–2382. [Google Scholar] [CrossRef]
  23. Bates, M.K.; Jia, Q.; Doan, H.; Liang, W.; Mukerjee, S. Charge-Transfer Effects in Ni-Fe & Ni-Fe-Co Mixed-Metal-Oxides for the Alkaline Oxygen Evolution Reaction. ACS Catal. 2015, 6, 155–161. [Google Scholar]
  24. Li, N.; Xia, W.Y.; Wang, J.; Liu, Z.L.; Li, Q.Y.; Chen, S.Z.; Xu, C.W.; Lu, X.H. Manganese oxides supported on hydrogenated TiO2 nanowire array catalysts for the electrochemical oxygen evolution reaction in water electrolysis. J. Mater. Chem. A 2015, 3, 21308–21313. [Google Scholar] [CrossRef]
  25. Schäfer, H.; Chatenet, M. Steel: The Resurrection of a Forgotten Water-Splitting Catalyst. ACS Energy Lett. 2018, 3, 574–591. [Google Scholar] [CrossRef]
  26. Davodi, F.; Mühlhausen, E.; Tavakkoli, M.; Sainio, J.; Jiang, H.; Gökce, B.; Marzun, G.; Kallio, T. Catalyst Support Effect on the Activity and Durability of Magnetic Nanoparticles: Toward Design of Advanced Electrocatalyst for Full Water Splitting. ACS Appl. Mater. Interfaces 2018, 10, 31300–31311. [Google Scholar] [CrossRef] [PubMed]
  27. Luo, Z.; Martí-Sànchez, S.; Nafria, R.; Joshua, G.; de la Mata, M.; Guardia, P.; Flox, C.; Martínez-Boubeta, C.; Simeonidis, K.; Llorca, J.; et al. Fe3O4@NiFexOy Nanoparticles with Enhanced Electrocatalytic Properties for Oxygen Evolution in Carbonate Electrolyte. ACS Appl. Mater. Interfaces 2016, 8, 29461–29469. [Google Scholar] [CrossRef] [PubMed]
  28. Davodi, F.; Tavakkoli, M.; Lahtinen, J.; Kallio, T. Straightforward synthesis of nitrogen-doped carbon nanotubes as highly active bifunctional electrocatalysts for full water splitting. J. Catal. 2017, 353, 19–27. [Google Scholar] [CrossRef] [Green Version]
  29. Neburchilov, V.; Wang, H.; Martin, J.J.; Qu, W. A review on air cathodes for zinc-air fuel cells. J. Power Sources 2010, 195, 1271–1291. [Google Scholar] [CrossRef]
  30. Ge, X.; Sumboja, A.; Wuu, D.; An, T.; Li, B.; Goh, F.W.T.; Hor, T.S.A.; Zong, Y.; Liu, Z. Oxygen Reduction in Alkaline Media: From Mechanisms to Recent Advances of Catalysts. ACS Catal. 2015, 5, 4643–4667. [Google Scholar] [CrossRef]
  31. Risch, M. Perovskite electrocatalysts for the Oxygen reduction reaction in alkaline media. Catalysts 2017, 7, 154. [Google Scholar] [CrossRef]
  32. Zhu, Y.; Zhou, W.; Shao, Z. Perovskite/Carbon Composites: Applications in Oxygen Electrocatalysis. Small 2017, 13. [Google Scholar] [CrossRef] [PubMed]
  33. Poux, T.; Napolskiy, F.S.; Dintzer, T.; Kéranguéven, G.; Istomin, S.Y.; Tsirlina, G.A.; Antipov, E.V.; Savinova, E.R. Dual role of carbon in the catalytic layers of perovskite/carbon composites for the electrocatalytic oxygen reduction reaction. Catal. Today 2012, 189, 83–92. [Google Scholar] [CrossRef]
  34. Sunarso, J.; Torriero, A.A.J.; Zhou, W.; Howlett, P.C.; Forsyth, M. Oxygen reduction reaction activity of La-based perovskite oxides in alkaline medium: A thin-film rotating ring-disk electrode study. J. Phys. Chem. C 2012, 116, 5827–5834. [Google Scholar] [CrossRef]
  35. Fabbri, E.; Mohamed, R.; Levecque, P.; Conrad, O.; Kötz, R.; Schmidt, T.J. Composite electrode boosts the activity of Ba0.5Sr0.5Co0.8Fe0.2O3−δ perovskite and carbon toward oxygen reduction in alkaline media. ACS Catal. 2014, 4, 1061–1070. [Google Scholar] [CrossRef]
  36. Poux, T.; Bonnefont, A.; Kéranguéven, G.; Tsirlina, G.A.; Savinova, E.R. Electrocatalytic oxygen reduction reaction on perovskite oxides: Series versus direct pathway. ChemPhysChem 2014, 15, 2108–2120. [Google Scholar] [CrossRef] [PubMed]
  37. Miyahara, Y.; Miyazaki, K.; Fukutsuka, T.; Abe, T. Catalytic roles of perovskite oxides in electrochemical oxygen reactions in alkaline media. J. Electrochem. Soc. 2014, 161, F694–F697. [Google Scholar] [CrossRef]
  38. Miyahara, Y.; Miyazaki, K.; Fukutsuka, T.; Abe, T. Influence of Surface Orientation on the Catalytic Activities of La0.8Sr0.2CoO3 Crystal Electrodes for Oxygen Reduction and Evolution Reactions. ChemElectroChem 2016, 3, 214–217. [Google Scholar] [CrossRef]
  39. Celorrio, V.; Dann, E.; Calvillo, L.; Morgan, D.J.; Hall, S.R.; Fermin, D.J. Oxygen Reduction at Carbon-Supported Lanthanides: TheRole of the B-Site. ChemElectroChem 2016, 3, 283–291. [Google Scholar] [CrossRef]
  40. Celorrio, V.; Calvillo, L.; Granozzi, G.; Russell, A.E.; Fermin, D.J. AMnO3 (A = Sr, La, Ca, Y) Perovskite Oxides as Oxygen Reduction Electrocatalysts. Top. Catal. 2018, 61, 154–161. [Google Scholar] [CrossRef]
  41. Gobaille-Shaw, G.P.A.; Celorrio, V.; Calvillo, L.; Morris, L.J.; Granozzi, G.; Fermín, D.J. Effect of Ba Content on the Activity of La1−xBaxMnO3 Towards the Oxygen Reduction Reaction. ChemElectroChem 2018, 5, 1922–1927. [Google Scholar] [CrossRef] [PubMed]
  42. Jung, J.I.; Risch, M.; Park, S.; Kim, M.G.; Nam, G.; Jeong, H.Y.; Shao-Horn, Y.; Cho, J. Optimizing nanoparticle perovskite for bifunctional oxygen electrocatalysis. Energy Environ. Sci. 2016, 9, 176–183. [Google Scholar] [CrossRef] [Green Version]
  43. Stoerzinger, K.A.; Choi, W.S.; Jeen, H.; Lee, H.N.; Shao-Horn, Y. Role of strain and conductivity in oxygen electrocatalysis on LaCoO3 thin films. J. Phys. Chem. Lett. 2015, 6, 487–492. [Google Scholar] [CrossRef] [PubMed]
  44. Hwang, C.; Gwon, O.; Jo, H.; Ok, K.M.; Kim, G. Major Role of Surface Area in Perovskite Electrocatalysts for Alkaline Systems. ChemElectroChem 2017, 4, 468–471. [Google Scholar] [CrossRef]
  45. Huang, K.; Yuan, L.; Jiang, Y.; Zhang, J.; Geng, Z.; Luo, L.; Feng, S. Hydrothermal shape controllable synthesis of La0.5Sr0.5MnO3 crystals and facet effect on electron transfer of oxygen reduction. Inorg. Chem. Front. 2018, 5, 732–738. [Google Scholar] [CrossRef]
  46. Yu, J.; Chen, D.; Saccoccio, M.; Lam, K.; Ciucci, F. Promotion of Oxygen Reduction with Both Amorphous and Crystalline MnOx through the Surface Engineering of La0.8Sr0.2MnO3−δ Perovskite. ChemElectroChem 2018, 5, 1105–1112. [Google Scholar] [CrossRef]
  47. Hua, B.; Zhang, Y.Q.; Yan, N.; Li, M.; Sun, Y.F.; Chen, J.; Li, J.; Luo, J.L. The Excellence of Both Worlds: Developing Effective Double Perovskite Oxide Catalyst of Oxygen Reduction Reaction for Room and Elevated Temperature Applications. Adv. Funct. Mater. 2016, 26, 4106–4112. [Google Scholar] [CrossRef]
  48. Chiba, R.; Aono, H.; Kato, K. An SOFC cathode infiltrated with Pr6O11. ECS Trans. 2013, 57, 1831–1840. [Google Scholar] [CrossRef]
  49. Ding, X.; Zhu, W.; Hua, G.; Li, J.; Wu, Z. Enhanced oxygen reduction activity on surface-decorated perovskite La0.6Ni0.4FeO3 cathode for solid oxide fuel cells. Electrochim. Acta 2015, 163, 204–212. [Google Scholar] [CrossRef]
  50. Sharma, R.K.; Djurado, E. An efficient hierarchical nanostructured Pr6O11 electrode for solid oxide fuel cells. J. Mater. Chem. A 2018, 6, 10787–10802. [Google Scholar] [CrossRef]
  51. Chen, C.H.; Emond, M.H.J.; Kelder, E.M.; Meester, B.; Schoonman, J. Electrostatic sol-spray deposition of nanostructured ceramic thin films. J. Aerosol Sci. 1999, 30, 959–967. [Google Scholar] [CrossRef]
  52. Jaworek, A. Electrospray droplet sources for thin film deposition. J. Mater. Sci. 2007, 42, 266–297. [Google Scholar] [CrossRef]
  53. Marinha, D.; Rossignol, C.; Djurado, E. Influence of electrospraying parameters on the microstructure of La0.6Sr0.4Co0.2F0.8O3−δ films for SOFCs. J. Solid State Chem. 2009, 182, 1742–1748. [Google Scholar] [CrossRef]
  54. Marinha, D.; Hayd, J.; Dessemond, L.; Ivers-Tiffée, E.; Djurado, E. Performance of (La,Sr)(Co,Fe)O3−x double-layer cathode films for intermediate temperature solid oxide fuel cell. J. Power Sources 2011, 196, 5084–5090. [Google Scholar] [CrossRef]
  55. Sharma, R.K.; Burriel, M.; Djurado, E. La4Ni3O10−δ as an efficient solid oxide fuel cell cathode: Electrochemical properties versus microstructure. J. Mater. Chem. A 2015, 3, 23833–23843. [Google Scholar] [CrossRef]
  56. Sharma, R.K.; Burriel, M.; Dessemond, L.; Martin, V.; Bassat, J.M.; Djurado, E. An innovative architectural design to enhance the electrochemical performance of La2NiO4+δ cathodes for solid oxide fuel cell applications. J. Power Sources 2016, 316, 17–28. [Google Scholar] [CrossRef]
  57. Marinha, D.; Dessemond, L.; Djurado, E. Comprehensive Review of Current Developments in IT-SOFCs. Curr. Inorg. Chem. 2013, 3, 2–22. [Google Scholar] [CrossRef]
  58. Gañán-Calvo, A.M.; Dávila, J.; Barrero, A. Current and droplet size in the electrospraying of liquids. Scaling laws. J. Aerosol Sci. 1997, 28, 249–275. [Google Scholar] [CrossRef]
  59. Jung, K.N.; Jung, J.H.; Im, W.B.; Yoon, S.; Shin, K.H.; Lee, J.W. Doped lanthanum nickelates with a layered perovskite structure as bifunctional cathode catalysts for rechargeable metal-air batteries. ACS Appl. Mater. Interfaces 2013, 5, 9902–9907. [Google Scholar] [CrossRef] [PubMed]
  60. Jung, S.; McCrory, C.C.L.; Ferrer, I.M.; Peters, J.C.; Jaramillo, T.F. Benchmarking nanoparticulate metal oxide electrocatalysts for the alkaline water oxidation reaction. J. Mater. Chem. A 2016, 4, 3068–3076. [Google Scholar] [CrossRef] [Green Version]
  61. Menezes, P.W.; Indra, A.; Bergmann, A.; Chernev, P.; Walter, C.; Dau, H.; Strasser, P.; Driess, M. Uncovering the prominent role of metal ions in octahedral: Versus tetrahedral sites of cobalt-zinc oxide catalysts for efficient oxidation of water. J. Mater. Chem. A 2016, 4, 10014–10022. [Google Scholar] [CrossRef]
  62. Menezes, P.W.; Indra, A.; González-Flores, D.; Sahraie, N.R.; Zaharieva, I.; Schwarze, M.; Strasser, P.; Dau, H.; Driess, M. High-Performance Oxygen Redox Catalysis with Multifunctional Cobalt Oxide Nanochains: Morphology-Dependent Activity. ACS Catal. 2015, 5, 2017–2027. [Google Scholar] [CrossRef]
  63. Menezes, P.W.; Indra, A.; Sahraie, N.R.; Bergmann, A.; Strasser, P.; Driess, M. Cobalt–Manganese-Based Spinels as Multifunctional Materials that Unify Catalytic Water Oxidation and Oxygen Reduction Reactions. ChemSusChem 2015, 8, 164–171. [Google Scholar] [CrossRef] [PubMed]
  64. Su, H.-Y.; Gorlin, Y.; Man, I.C.; Calle-Vallejo, F.; Norskov, J.K.; Jaramillo, T.F.; Rossmeisl, J. Identifying active surface phases for metal oxide electrocatalysts: A study of manganese oxide bi-functional catalysts for oxygen reduction and water oxidation catalysis. Phys. Chem. Chem. Phys. 2012, 14, 14010–14022. [Google Scholar] [CrossRef] [PubMed]
  65. Poux, T.; Bonnefont, A.; Ryabova, A.; Kéranguéven, G.; Tsirlina, G.A.; Savinova, E.R. Electrocatalysis of hydrogen peroxide reactions on perovskite oxides: Experiment versus kinetic modeling. Phys. Chem. Chem. Phys. 2014, 16, 13595–13600. [Google Scholar] [CrossRef] [PubMed]
  66. Case, B. The diffusivity of oxygen in dilute alkaline solution from 0° to 65 °C. Electrochim. Acta 1973, 18, 293–299. [Google Scholar] [CrossRef]
  67. Yang, X.; Li, S.; Liiu, Y.; Wei, X.; Liu, Y. LaNi0.8Co0.2O3 as a cathode catalyst for a direct borohydride fuel cell. J. Power Sources 2011, 196, 4992–4995. [Google Scholar] [CrossRef]
  68. Serov, A.; Aziznia, A.; Benhangi, P.H.; Artyushkova, K.; Atanassov, P.; Gyenge, E. Borohydride-tolerant oxygen electroreduction catalyst for mixed-reactant Swiss-roll direct borohydride fuel cells. J. Mater. Chem. A 2013, 1, 14384–14391. [Google Scholar] [CrossRef]
  69. Liu, B.H.; Li, Z.P. Current status and progress of direct borohydride fuel cell technology development. J. Power Sources 2009, 187, 291–297. [Google Scholar] [CrossRef]
  70. Cheng, H.; Scott, K. Investigation of non-platinum cathode catalysts for direct borohydride fuel cells. J. Electroanal. Chem. 2006, 596, 117–123. [Google Scholar] [CrossRef]
  71. Baranton, S.; Coutanceau, C.; Roux, C.; Hahn, F.; Léger, J.M. Oxygen reduction reaction in acid medium at iron phthalocyanine dispersed on high surface area carbon substrate: Tolerance to methanol, stability and kinetics. J. Electroanal. Chem. 2005, 577, 223–234. [Google Scholar] [CrossRef]
  72. Chatenet, M.; Micoud, F.; Roche, I.; Chainet, E.; Vondrak, J. Kinetics of sodium borohydride direct oxidation and oxygen reduction in sodium hydroxide electrolyte—Part II. O2 reduction. Electrochim. Acta 2006, 51, 5452–5458. [Google Scholar] [CrossRef]
  73. Garcia, A.; Linares, J.; Chatenet, M.; Ticianelli, E. NiMnOx/C: A Non-noble Ethanol-Tolerant Catalyst for Oxygen Reduction in Alkaline Exchange Membrane DEFC. Electrocatalysis 2014, 5, 41–49. [Google Scholar] [CrossRef]
  74. Hernández-Rodríguez, M.A.; Goya, M.C.; Arévalo, M.C.; Rodríguez, J.L.; Pastor, E. Carbon supported Ag and Ag–Co catalysts tolerant to methanol and ethanol for the oxygen reduction reaction in alkaline media. Int. J. Hydrogen Energy 2016, 41, 19789–19798. [Google Scholar] [CrossRef]
  75. Zhu, Y.; Zhou, W.; Yu, J.; Chen, Y.; Liu, M.; Shao, Z. Enhancing Electrocatalytic Activity of Perovskite Oxides by Tuning Cation Deficiency for Oxygen Reduction and Evolution Reactions. Chem. Mater. 2016, 28, 1691–1697. [Google Scholar] [CrossRef]
  76. Li, Z.; Lv, L.; Wang, J.; Ao, X.; Ruan, Y.; Zha, D.; Hong, G.; Wu, Q.; Lan, Y.; Wang, C.; et al. Engineering phosphorus-doped LaFeO3−δ perovskite oxide as robust bifunctional oxygen electrocatalysts in alkaline solutions. Nano Energy 2018, 47, 199–209. [Google Scholar] [CrossRef]
  77. Liu, X.; Gong, H.; Wang, T.; Guo, H.; Song, L.; Xia, W.; Gao, B.; Jiang, Z.; Feng, L.; He, J. Cobalt-Doped Perovskite-Type Oxide LaMnO3 as Bifunctional Oxygen Catalysts for Hybrid Lithium–Oxygen Batteries. Chem. Asian J. 2018, 13, 528–535. [Google Scholar] [CrossRef] [PubMed]
  78. Afzal, R.A.; Park, K.Y.; Cho, S.H.; Kim, N.I.; Choi, S.R.; Kim, J.H.; Lim, H.T.; Park, J.Y. Oxygen electrode reactions of doped BiFeO3 materials for low and elevated temperature fuel cell applications. RSC Adv. 2017, 7, 47643–47653. [Google Scholar] [CrossRef]
  79. Dzara, M.J.; Christ, J.M.; Joghee, P.; Ngo, C.; Cadigan, C.A.; Bender, G.; Richards, R.M.; O’Hayre, R.; Pylypenko, S. La and Al co-doped CaMnO3 perovskite oxides: From interplay of surface properties to anion exchange membrane fuel cell performance. J. Power Sources 2018, 375, 265–276. [Google Scholar] [CrossRef]
  80. Hua, B.; Sun, Y.F.; Li, M.; Yan, N.; Chen, J.; Zhang, Y.Q.; Zeng, Y.; Shalchi Amirkhiz, B.; Luo, J.L. Stabilizing Double Perovskite for Effective Bifunctional Oxygen Electrocatalysis in Alkaline Conditions. Chem. Mater. 2017, 29, 6228–6237. [Google Scholar] [CrossRef]
  81. Chen, C.F.; King, G.; Dickerson, R.M.; Papin, P.A.; Gupta, S.; Kellogg, W.R.; Wu, G. Oxygen-deficient BaTiO3−x perovskite as an efficient bifunctional oxygen electrocatalyst. Nano Energy 2015, 13, 423–432. [Google Scholar] [CrossRef]
  82. Sharma, R.K.; Burriel, M.; Dessemond, L.; Bassat, J.M.; Djurado, E. Lan+1NinO3n+1 (n = 2 and 3) phases and composites for solid oxide fuel cell cathodes: Facile synthesis and electrochemical properties. J. Power Sources 2016, 325, 337–345. [Google Scholar] [CrossRef]
  83. Sharma, R.K.; Burriel, M.; Dessemond, L.; Bassat, J.M.; Djurado, E. Design of interfaces in efficient Ln2NiO4+δ (Ln = La, Pr) cathodes for SOFC applications. J. Mater. Chem. A 2016, 4, 12451–12462. [Google Scholar] [CrossRef]
  84. Djurado, E.; Salaün, A.; Mignardi, G.; Rolle, A.; Burriel, M.; Daviero-Minaud, S.; Vannier, R.N. Electrostatic spray deposition of Ca3Co4O9 + δ layers to be used as cathode materials for IT-SOFC. Solid State Ionics 2016, 286, 102–110. [Google Scholar] [CrossRef]
  85. Çelikbilek, Ӧ.; Jauffrès, D.; Siebert, E.; Dessemond, L.; Burriel, M.; Martin, C.L.; Djurado, E. Rational design of hierarchically nanostructured electrodes for solid oxide fuel cells. J. Power Sources 2016, 333, 72–82. [Google Scholar] [CrossRef]
  86. Constantin, G.; Rossignol, C.; Barnes, J.P.; Djurado, E. Interface stability of thin, dense CGO film coating on YSZ for solid oxide fuel cells. Solid State Ionics 2013, 235, 36–41. [Google Scholar] [CrossRef]
  87. Rodríguez-Carvajal, J. Recent advances in magnetic structure determination by neutron powder diffraction. Phys. B Phys. Condens. Matter 1993, 192, 55–69. [Google Scholar] [CrossRef]
  88. Chatenet, M.; Concha, B.M.; Parrour, G.; Diard, J.-P.; Lima, F.H.B.; Ticianelli, E.A. Potential and limitation of the Direct Borohydride Fuel Cell. Special emphasis on the Borohydride Oxidation Reaction (BOR) mechanism and kinetics on gold electrocatalysts. In Boron Hydrides, High Potential Hydrogen Storage Materials; Demirci, U.B., Miele, P., Eds.; Nova Science: New York, NY, USA, 2011; pp. 103–135. [Google Scholar]
  89. Olu, P.-Y.; Deschamps, F.; Caldarella, G.; Chatenet, M.; Job, N. Investigation of platinum and palladium as potential anodic catalysts for direct borohydride and ammonia borane fuel cells. J. Power Sources 2015, 297, 492–503. [Google Scholar] [CrossRef]
  90. Olu, P.-Y.; Job, N.; Chatenet, M. Evaluation of anode (electro)catalytic materials for the direct borohydride fuel cell: Methods and benchmarks. J. Power Sources 2016, 327, 235–257. [Google Scholar] [CrossRef]
  91. Chen, J.G.; Jones, C.W.; Linic, S.; Stamenkovic, V.R. Best Practices in Pursuit of Topics in Heterogeneous Electrocatalysis. ACS Catal. 2017, 7, 6392–6393. [Google Scholar] [CrossRef]
  92. Zadick, A.; Dubau, L.; Sergent, N.; Berthomé, G.; Chatenet, M. Huge Instability of Pt/C Catalysts in Alkaline Medium. ACS Catal. 2015, 5, 4819–4824. [Google Scholar] [CrossRef]
  93. Lafforgue, C.; Zadick, A.; Dubau, L.; Maillard, F.; Chatenet, M. Selected review of the degradation of Pt and Pd-based carbonsupported electrocatalysts for alkaline fuel cells: Towards mechanisms of degradation. Fuel Cells 2018, 18, 229–238. [Google Scholar] [CrossRef]
  94. Cherevko, S.; Zeradjanin, A.R.; Keeley, G.P.; Mayrhofer, K.J.J. A comparative study on gold and platinum dissolution in acidic and alkaline media. J. Electrochem. Soc. 2014, 161, H822–H830. [Google Scholar] [CrossRef]
  95. Schalenbach, M.; Kasian, O.; Ledendecker, M.; Speck, F.D.; Mingers, A.M.; Mayrhofer, K.J.J.; Cherevko, S. The Electrochemical Dissolution of Noble Metals in Alkaline Media. Electrocatalysis 2018, 9, 153–161. [Google Scholar] [CrossRef]
  96. Mayrhofer, K.J.J.; Wiberg, G.K.H.; Arenz, M. Impact of Glass Corrosion on the Electrocatalysis on Pt Electrodes in Alkaline Electrolyte. J. Electrochem. Soc. 2008, 155, P1–P5. [Google Scholar] [CrossRef]
Figure 1. SEM micrographs of Pr6O11/GC ESD films obtained with a EtOH:BC (1:2) solution with a concentration of 0.02 M at Ts = 230 °C for a nozzle to substrate distance of dns = 20 mm, a flow rate of Q = 1.5 mL h−1, and a deposition time of t = 3 h. (AC) surface view and (D) cross-section of amorphous (as-prepared) Pr6O11/GC; (EG) surface view; (H) cross-section of crystalline (sintered at T = 550 °C for 2 h in air) Pr6O11/GC.
Figure 1. SEM micrographs of Pr6O11/GC ESD films obtained with a EtOH:BC (1:2) solution with a concentration of 0.02 M at Ts = 230 °C for a nozzle to substrate distance of dns = 20 mm, a flow rate of Q = 1.5 mL h−1, and a deposition time of t = 3 h. (AC) surface view and (D) cross-section of amorphous (as-prepared) Pr6O11/GC; (EG) surface view; (H) cross-section of crystalline (sintered at T = 550 °C for 2 h in air) Pr6O11/GC.
Catalysts 08 00461 g001
Figure 2. Typical X-EDS spectra of the Pr6O11/GC film deposited on the glassy-carbon substrate: (A) amorphous; (B) crystalline, sintered for 2 h in air at T = 550 °C.
Figure 2. Typical X-EDS spectra of the Pr6O11/GC film deposited on the glassy-carbon substrate: (A) amorphous; (B) crystalline, sintered for 2 h in air at T = 550 °C.
Catalysts 08 00461 g002
Figure 3. X-ray diffraction patterns of Pr6O11 ESD film deposited on vitreous carbon substrate: (A) as-deposited (amorphous); (B) sintered for 2 h in air at T = 550 °C (crystalline).
Figure 3. X-ray diffraction patterns of Pr6O11 ESD film deposited on vitreous carbon substrate: (A) as-deposited (amorphous); (B) sintered for 2 h in air at T = 550 °C (crystalline).
Catalysts 08 00461 g003
Figure 4. TEM, HRTEM, and SAED pattern of powder scratched from the brain-like type Pr6O11 ESD film sintered in air for 2 h at T = 550 °C (crystalline Pr6O11): (A) TEM, (B) HRTEM, (C) SAED pattern.
Figure 4. TEM, HRTEM, and SAED pattern of powder scratched from the brain-like type Pr6O11 ESD film sintered in air for 2 h at T = 550 °C (crystalline Pr6O11): (A) TEM, (B) HRTEM, (C) SAED pattern.
Catalysts 08 00461 g004
Figure 5. Voltamperogram in 0.1 M NaOH-supporting electrolyte of amorphous Pr6O11 (Ar atmosphere), with incursions in the HER and OER regions of Pr6O11; T = 25 ± 2 °C, v = 20 mV s−1.
Figure 5. Voltamperogram in 0.1 M NaOH-supporting electrolyte of amorphous Pr6O11 (Ar atmosphere), with incursions in the HER and OER regions of Pr6O11; T = 25 ± 2 °C, v = 20 mV s−1.
Catalysts 08 00461 g005
Figure 6. Oxygen reduction reaction voltamperogram in 0.1 M NaOH electrolyte of amorphous Pr6O11 (O2 atmosphere); T = 25 ± 2 °C, v = 5 mV s−1.
Figure 6. Oxygen reduction reaction voltamperogram in 0.1 M NaOH electrolyte of amorphous Pr6O11 (O2 atmosphere); T = 25 ± 2 °C, v = 5 mV s−1.
Catalysts 08 00461 g006
Figure 7. Oxygen reduction reaction voltamperogram in 0.1 M NaOH + 0.01 M NaBH4 electrolyte of amorphous Pr6O11 (O2 atmosphere); T = 25 ± 2 °C, v = 5 mV s−1.
Figure 7. Oxygen reduction reaction voltamperogram in 0.1 M NaOH + 0.01 M NaBH4 electrolyte of amorphous Pr6O11 (O2 atmosphere); T = 25 ± 2 °C, v = 5 mV s−1.
Catalysts 08 00461 g007
Figure 8. Voltamperogram in 0.1 M NaOH-supporting electrolyte of crystalline Pr6O11 annealed at T = 550 °C (Ar atmosphere), with incursions in the HER and OER regions; T = 25 ± 2 °C, v = 20 mV s−1.
Figure 8. Voltamperogram in 0.1 M NaOH-supporting electrolyte of crystalline Pr6O11 annealed at T = 550 °C (Ar atmosphere), with incursions in the HER and OER regions; T = 25 ± 2 °C, v = 20 mV s−1.
Catalysts 08 00461 g008
Figure 9. (A) Oxygen reduction reaction voltamperogram in 0.1 M NaOH electrolyte of crystalline Pr6O11 annealed at T = 550 °C (O2 atmosphere), (B) comparison with amorphous Pr6O11 (not-annealed, data set in dashed lines), and (C) comparison of amorphous and crystalline Pr6O11 with a state-of-the-art Pt/C electrocatalyst at Ω = 900 rpm; T = 25 ± 2 °C, v = 5 mV s−1.
Figure 9. (A) Oxygen reduction reaction voltamperogram in 0.1 M NaOH electrolyte of crystalline Pr6O11 annealed at T = 550 °C (O2 atmosphere), (B) comparison with amorphous Pr6O11 (not-annealed, data set in dashed lines), and (C) comparison of amorphous and crystalline Pr6O11 with a state-of-the-art Pt/C electrocatalyst at Ω = 900 rpm; T = 25 ± 2 °C, v = 5 mV s−1.
Catalysts 08 00461 g009
Figure 10. Oxygen reduction reaction voltamperogram in 0.1 M NaOH + 0.01 M NaBH4 electrolyte of crystalline Pr6O11 annealed at T = 550 °C (O2 atmosphere); T = 25 ± 2 °C, v = 5 mV s−1.
Figure 10. Oxygen reduction reaction voltamperogram in 0.1 M NaOH + 0.01 M NaBH4 electrolyte of crystalline Pr6O11 annealed at T = 550 °C (O2 atmosphere); T = 25 ± 2 °C, v = 5 mV s−1.
Catalysts 08 00461 g010

Share and Cite

MDPI and ACS Style

Sharma, R.K.; Müller, V.; Chatenet, M.; Djurado, E. Oxygen Reduction Reaction Electrocatalysis in Alkaline Electrolyte on Glassy-Carbon-Supported Nanostructured Pr6O11 Thin-Films. Catalysts 2018, 8, 461. https://0-doi-org.brum.beds.ac.uk/10.3390/catal8100461

AMA Style

Sharma RK, Müller V, Chatenet M, Djurado E. Oxygen Reduction Reaction Electrocatalysis in Alkaline Electrolyte on Glassy-Carbon-Supported Nanostructured Pr6O11 Thin-Films. Catalysts. 2018; 8(10):461. https://0-doi-org.brum.beds.ac.uk/10.3390/catal8100461

Chicago/Turabian Style

Sharma, Rakesh K., Verónica Müller, Marian Chatenet, and Elisabeth Djurado. 2018. "Oxygen Reduction Reaction Electrocatalysis in Alkaline Electrolyte on Glassy-Carbon-Supported Nanostructured Pr6O11 Thin-Films" Catalysts 8, no. 10: 461. https://0-doi-org.brum.beds.ac.uk/10.3390/catal8100461

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