Next Article in Journal
Electrochemical Analysis of Corrosion Resistance of Manganese-Coated Annealed Steel: Chronoamperometric and Voltammetric Study
Previous Article in Journal
Group Contribution Revisited: The Enthalpy of Formation of Organic Compounds with “Chemical Accuracy” Part VI
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Electrochemical Characterization of Ru-Modified Iridium Oxide Catalysts for PEM Electrolysis

by
Stanford Chidziva
1,*,
Dorcas Zide
1,2,*,
Joshua John Bambo
1,
Anele Sinto
1,2,*,
Sivakumar Pasupathi
1 and
Bernard J. Bladergroen
1
1
South African Institute for Advance Materials Chemistry (SAIAMC), University of the Western Cape, Robert Sobukwe Road, Bellville, Cape Town 7535, South Africa
2
Department of Chemistry, Cape Peninsula University of Technology, Symphony Way, Bellville, Cape Town 7535, South Africa
*
Authors to whom correspondence should be addressed.
AppliedChem 2024, 4(4), 353-366; https://doi.org/10.3390/appliedchem4040022
Submission received: 18 July 2024 / Revised: 13 September 2024 / Accepted: 26 September 2024 / Published: 24 October 2024

Abstract

:
In the search of sustainable energy solutions, proton exchange membrane water electrolyzers (PEMWEs) have emerged as a promising alternative for sustainable clean hydrogen production. This study focuses on synthesis and characterization of Ruthenium (Ru)-modified iridium oxide (IrO2) catalysts. The anode is the principal reason for the high overpotential of PEMWEs and it also greatly increases the cost of the electrolyzers. IrO2 is highly stable and corrosion-resistant, particularly in acidic environments, making it a durable catalyst for the oxygen evolution reaction (OER) in PEMWEs, though it suffers from a relatively high overpotential. Ruthenium oxide (RuO2), on the other hand, is more catalytically active with a lower overpotential, but is less stable under the same conditions. In this study, the goal was to improve the catalytic activity and stability of the anode catalyst, IrO2, through the controlled incorporation of Ru and to reduce overall catalyst cost due to the reduced iridium content. This synergistic combination allows for better performance in terms of conductivity, efficiency, and durability, making Ru-modified IrO2 an ideal catalyst for OER in PEMWE applications. The Adams fusion method was adapted and used to synthesize the catalysts. The modified catalysts were characterized using analytical instruments. These analyses provided insights into the structural, morphological, and electrochemical properties of the Ru-modified IrO2 catalysts.

1. Introduction

The increase in the global population is causing a global energy crisis and a demand for alternative energy sources, such as renewable energy sources [1,2,3]. Consumption of fossil fuels increases the risk of global warming, raising concern among the public [1,3] Fossil fuels are limited energy resources compared to renewable energy sources such as solar energy, which is used to generate green hydrogen [2]. Globally, climate change disproportionately affects the most disadvantaged and vulnerable people [4]. Sustainable objectives that set the stage for climate change-resistant and low-carbon development were established by the UN Sustainable Development Goals (SDGs) and the Paris Climate Agreement (PCA) in the year 2015 [4]. The SDGs must be achieved soon to stop climate change and mitigate its effects [4]. Energy consumption in the end-use sector, namely transport, buildings, and industry, will also have to be considered to achieve time targets for decarbonization because increasing the options available will offer the best opportunity for doing so [4].
Using fossil fuels as an energy source causes significant pollution. The burning of fossil fuels produces SO2, which is the cause of acid rain [3]. Large amounts of greenhouse gasses being released into the atmosphere have caused concern globally, which has led to an international agreement on the reduction of CO2 emissions known as the Kyoto Protocol [1,3].
Hydrogen has become a widespread energy source in recent years, and this is due to it being greener compared to fossil fuels [3]. Hydrogen is produced by an electrochemical process called electrolysis, where water is split into oxygen and hydrogen [2]. Hydrogen as an energy carrier has benefits over hydrocarbon fuels, including its high specific energy density, and does not emit pollutants such as CO2 [2]. Global hydrogen production for chemical industries, fertilizer production, iron refineries and oil refineries, is estimated to be around 95 MtH2 per year in 2022 [5]. The Institute for Energy Economics and Financial Analysis (IEEFA) conservatively estimates that around 2.9 million tonnes of hydrogen per year could be produced from sustainable energy sources by 2030 [6]. The demand for hydrogen will continue to increase yearly. Alternative cheap electrochemical catalysts used in PEMWE for hydrogen production need to be developed. The high cost associated with components such as precious metal catalysts and the proton conducting membrane is the primary challenge facing the Proton Exchange Membrane (PEM) water electrolyzer [7]. Significant efforts by various researchers aiming to reduce the cost of the PEM water electrolyzer electrocatalysts and improve their specific performance and durability using the precious metal loading requirement have been conducted [7,8,9].
At the anode, the electrocatalyst plays a critical role as it undergoes the oxygen evolution reaction (OER), exhibiting the highest overpotential in standard operating conditions [7,9]. Similar to the oxygen reduction reaction (ORR) in Proton Exchange Membrane Fuel Cells (PEMFC), the OER is kinetically sluggish, since it is thermodynamically and kinetically unfavorable to remove four electrons for oxygen–oxygen bond formation [9].
Among metal oxides, IrO2 exhibits good durability with the second-best activity, while RuO2 exhibits the highest activity with poor durability [10]. Several methods have been studied for synthesizing nano-sized metal oxides, including Adams fusion, the molten salt method, metal–organic chemical vapor deposition method, sulphite complex route method, sol–gel method, modified polyol method, and hydrothermal method [7]. However, in the context of the scaling-up process, some of these methods involve steps requiring complex equipment and present technical and economic challenges [7].
In this study, research was carried out to synthesize nanostructured Ru-modified IrO2 electrocatalysts to improve the catalytic activity and the corrosion stability of the modified metal oxide compared to pure noble metal oxide electrocatalysts and to reduce the cost of the catalysts. Modifying IrO2 catalysts with ruthenium (Ru) presents a promising avenue to overcome the challenges associated with pure IrO2 and unlock improved performance in PEM electrolyzers. Ruthenium, characterized by lower cost and relatively more abundant supply than iridium, addresses the economic barriers associated with IrO2 catalysts. Furthermore, incorporating Ru can significantly enhance the electrochemical activity of the catalyst, reducing the overpotential required for the OER. This translates to improved energy efficiency and reduced electrolysis costs. Moreover, Ru modification can potentially mitigate the stability issues of IrO2 by providing enhanced structural stability and resistance to degradation under harsh operating conditions. The synergistic effects of the IrO2-Ru hybrid catalyst hold promise for achieving high catalytic performance, increased durability, and cost-effectiveness, thereby accelerating the adoption of PEM electrolyzers for sustainable hydrogen production.

2. Electrolysers

2.1. Alkaline Water Electrolyser

Alkaline water electrolysis uses an aqueous electrolyte solution consisting of approximately 20–30% KOH or NaOH and operates at relatively low temperatures (60–80 °C) [11]. Nickel-based materials and separators such as zirfon and asbestos fabricate the anode and cathode electrodes. In the alkaline process, two moles of water are reduced to one mole of H2 and two OH ions at the cathode [11]. The resulting H2 is released from the cathodic surface. The electric potential between the anode and the cathode force the hydroxile ions through a porous separator to the anode side where they are oxidized at the anode to produce half a molecule of O2 and one molecule of H2O [11].
The limitations to alkaline water electrolysis include limited current density, reduced partial load, and low operating pressure, which can lead to reduced energy efficiency [12]. Consequently, there is a promise within the field of ALK water electrolysis technologies for assessing anionic conductive polymer-based AEMs with alternative asbestos diaphragms [11].

2.2. Anion Exchange Membrane(AEM) Water Electrolyser

AEM electrolysis uses distilled water, or a low concentration alkaline solution compared to water electrolysis, where a concentrated ALK solution is usually used as electrolyte [13]. AEM electrolysis technologies are currently under the development stage to the kW scale and require further investigation. Anode and cathode catalytic electrodes primarily utilize Ni and NiFeCo-based alloy materials. Standard AEMs include quaternary ammonium ion exchange membranes, such as Fumasep® FAA3, Tokuyama A201, Ionomr Aemion™, Dioxide materials Sustainion® [14]. The gas diffusion layers are Ni-foam, porous Ni-mesh, or C-cloths for both the anode and cathode. Ni-coated stainless steel and stainless-steel separator plates are used as bipolar plates and end plates [13]. Challenges the AEM water electrolyzers face include limited stability and high hydrogen production cost, and these hinder the commercialization of AEM water electrolyzer systems.

2.3. Solid Oxide Electrolyser

In solid oxide electrolysis, steam at the anode combines with electrons from the external circuit to form hydrogen gas and negatively charged oxygen ions. The oxygen ions pass through the solid ceramic membrane and gets reduced at the cathode to form oxygen gas and generate electrons for the external circuit.
The solid oxide electrolyzer cell has two capabilities that stand out: its ability to perform high-temperature electrolysis and smooth integration with industrial processes, thus enabling improvement in overall efficiency in the hydrogen production process. Conventional SOE makes use of O2− conductors made from nickel/yttria stabilized zirconia; however, now, the use of ceramic proton conducting materials exhibits superior efficiency at high operating temperatures [15]. Comparing the efficiency of the SOE to AKL and AEM, each is ~80%, ~78%, and ~59%, respectively [16]. However, SOE technology suffers from lack of stability, which hinders the commercialization of it [15].

2.4. Proton Exchange Membrane Water Electrolyser

Recently much attention has been paid to the production of H2 from H2O using a proton exchange membrane (PEM) water electrolyzer [17]. Grubbs in the 1950s introduced the concept of PEM electrolysis [18,19], which was advanced by General Electric Co. in 1966 [15,20]. The PEM fuel cell technology is similar to PEM water electrolysis technology, as both technologies utilize a solid polysulfonated membrane (Nafion®) which acts as an electrolyte [11]. The membrane exhibits many advantages such as low gas permeability, high proton conductivity and high-pressure operational capacity [11]. Normal operation of the PEM water electrolysis is between 20–80 °C with high current densities (<2 A/cm2) and produces high purity (99.999%) of H2 and O2 gasses [11].
PEM technology has received attention by many water electrolyzer manufacturers for industrial and transportation applications since it has a smaller carbon footprint and an absence of hazard electrolytes, as well as being safer compared to ALK water electrolysis [21].

3. Anode Catalysts for the Proton Exchange Membrane Water Electrolyzer

The use of catalysts plays an important role in effectively facilitating the oxygen evolution reaction (OER) to produce O2 from the oxidation of H2O at the anode of the water electrolyzer. Therefore, it is important to develop an electrocatalyst which is highly conductive and has high stability. The acidic environment of the PEM water electrolyzer destabilizes transition metals such as Ni, Co, and Mn, thus these metals undergo corrosion which increases the risk of poisoning the membrane.

3.1. Iridium Oxide

One of the commonly used electrocatalysts for the oxygen evolution reaction (OER) due to its high stability, it is considered ‘the-state-of-the-art’ OER catalyst. According to [7], IrO2 is one of the catalysts that exhibit the required stability and activity for use in acidic solid electrolyte environments. Applications of IrO2 other than in electrocatalysts include field emission, sensing, electrical properties, and Li-air batteries [22,23]. Arico et al. demonstrated the effect of reaction conditions on the morphology of the IrO2 [24].

3.2. Ruthenium Oxide

RuO2 exhibits higher OER activity in a low overpotential range but with lower stability [24]. Combining iridium and ruthenium to form an oxide is known as a binary metal oxide (Irx−1RuxO2.). According to Adamson et al. [25], binary metal oxides exhibit the following properties: enhanced electrocatalytic activity [26], favorable morphologic for ion transfer [27], and the creation of unique crystal interfaces [28]. Wang et al. [29] report on the superior activity and enhanced cell efficiency shown by the Ir0.7Ru0.3O2 catalyst. When Wang et al. [29] studied the effect of different preparation conditions on the Ir0.7Ru0.3O2 catalyst, it was found that leaching Ru in Ir0.7Ru0.3O2 (EC) showed 13-fold higher OER activity than Ir0.7Ru0.3O2 (TT). It was noted that, in the initial catalysis stage, the Ru component in Ir0.7Ru0.3O2 was unstable [29].
Ir mitigates the degradation of RuO2 during prolonged electrochemical operations, leading to a longer-lasting electrode. This synergy between Ir and Ru enhances the overall performance and stability of the anode, making it more effective for applications such as the OER in water electrolysis. The presence of RuO2 alongside IrO2 provides additional active sites for the OER, thereby improving the rate of oxygen production and lowering the activation energy required for the reaction.

4. Material and Methodology

4.1. Chemicals and Apparatus

The chemicals used for the synthesis of the Ru-modified IrO2 catalyst and to prepare and test the working electrode (WE) were H2IrCl6; RuCl3; NaNO3; Isopropanol; IrO2 commercial; and RuO2 commercial, respectively. The chemicals used for the preparation of the working electrode (WE) were Nafion® Solution 5 wt%; Isopropanol; and Ultrapure H2O with a resistance of 18.2 MΩ.cm.
The apparatuses used during synthesis and electrochemical testing were a magnetic stirrer and hotplate; water bath; drying oven; ultrasonic bath; muffle furnace; and a PGSTAT302N Potentiostat/Galvanostat with working electrode (WE), counter electrode (CE), and reference electrode (RE).

4.2. Synthesis of Ir0.8Ru0.2O2 Catalyst

4.2.1. Modified Adams Fusion Method

A predetermined amount of 0.36 g of total precursor salt (amounting to a ratio of 0.8:0.2 of iridium precursor to ruthenium precursor) was dissolved in 10 mL isopropanol and magnetically stirred for 30 min at 250–300 rotation per minute (rpm). An amount of finely grounded NaNO3 was added to the mixture and stirred for another 30 min. The mixture was then placed on a hot plate set at 80 °C to evaporate excess isopropanol and then further dried in an oven for 10 min at 80 °C. The dried mixture was transferred to a porcelain crucible. The crucible with the dried mixture was placed into a preheated furnace at 350 °C for 2 h. After the synthesis duration of 2 h, the mixture was taken out of the furnace and left to cool overnight. The obtained metal oxide was filtered using approximately 2 L of ultrapure H2O to ensure that unreacted salt was removed. A solution of 0.1 M AgNO3 was used to the filtrate to ensure there was no chlorine present. Finally, the Ir0.8Ru0.2O2 was dried in a preheated oven at 80 °C overnight.
The Ir:Ru ratio and synthesis temperature were determined from the literature and the optimized ratio was used in this research.

4.2.2. Preparation of Working Electrode

For all electrochemical measurements, a glassy carbon (GC) working electrode (WE) was used. To clean the GC, a 0.05 μm alumina paste was used to polish the GC. Thereafter, the GC was placed in an ultrasonic H2O bath to further remove surface particles. Figure 1 shows how the GC was polished and cleaned [30].
The catalyst ink was prepared by dispersing 8 mg of electrocatalyst, 50 μL Nafion solution (5 wt%), and 1950 μL ultrapure H2O using an ultrasonic homogenizer for 15 min. A total of 30 μL of electrocatalyst was dropped onto the GC WE with a micropipette. The ink was dried at room temperature overnight and the GC and WE were covered by a beaker to prevent them from falling onto the wet ink. Before electrochemical measurements were taken, the GC and WE were rotated for 15 min to ensure the ink was dry.

4.2.3. Electrochemical Characterization

In this study, a three-electrode electrochemical cell was used for the electrochemical measurements. All tests were performed at 25 °C using a warm bath and 1 atm using the Autolab potentiostat PGSTAT302N. The cell setup was in electrolyte of 0.5 M H2SO4(aq) with the 3 electrodes, the GC with catalyst coat, a 3 M Ag/AgCl, and a 1 Pt sheet which are the WC, RE, and CE, respectively. All potentials in this experiment were calibrated from the 3 M Ag/AgCl electrode to the RHE, since the Ag/AgCl had a positive potential shift of +0.21 V compared to the RHE [8] The cell was purged using N2 for 15 min, and thereafter, an electrode activation was performed using CV at a scan rate of 0.02 V/s for 50 cycles to remove any impurities on the surface of the WE. The cell setup used in this experiment can be seen in Figure 2 below. The parameters for the cyclic voltammetry are shown in Table 1, Table 2 shows the parameters for linear sweep voltammetry analysis and Table 3 shows the parameters for CP (Chronopotentiometry) analysis.

5. Results and Discussion

Modified catalysts: Physical and electrochemical characterization of in-house Ir0.8Ru0.2O2 catalyst.
There have been extensive studies of Ir-Ru mixed oxides for many years. According to the literature, the addition of Ru to IrO2 improves the electrocatalytic activity of IrO2 [32,33]. In this study, IrO2 was modified by adding Ru, the composition of Ru to Ir was determined by in-house (IH) literature and, according [33] to the best performing Ru-modified IrO2 catalyst, had a composition ratio of Ir to Ru of 0.8:0.2 respectively.
The EDS results of IrO2 commercial, RuO2 commercial, and Ir0.8Ru0.2O2-IH are represented by 3. The data from EDS shows trace amounts of impurities in the IrO2, which could be due to the synthesis procedure, or the trace amounts were detected from the tray the analyte was placed in. However, both the RuO2 and Ir0.8Ru0.2O2-IH samples were found to be pure the ratio of Ir to Ru was found to be 0.8:0.2, respectively. Table 4 shows the elemental composition of commercial and IH samples.
Figure 3 shows the XRD spectra of the IH-Ir0.8Ru0.2O2 catalyst synthesized at 350 °C as well as both commercial IrO2 and RuO2 catalysts. The IrO2 and RuO2 standard peaks were assigned using the data available from the JCPDS database. Both commercial catalysts IrO2 and RuO2 represent a similar presence of the crystalline phase, which is confirmed when their similar characteristic phases are cross-referenced with the JCPDS database, JCP2-43-1019 for IrO2 and JCP2-40-1290 for RuO2 [34]. Comparing the peaks of the IH-Ru-modified IrO2 to the commercial catalysts (IrO2 and RuO2), the commercial catalysts have narrower peaks compared to IH-Ru-modified IrO2 catalyst. This indicates that the IH-Ru-modified IrO2 is more amorphous in nature and the commercial catalysts being more crystalline in nature. According to literature, the amorphous phase usually has smaller particle size than the crystalline phase [35]. For both commercial samples, the (110) facet is there the main diffraction peaks however, the (101) facet was the main diffraction peak for the IH sample, and a similar result was achieved by [7] by using a synthesis temperature of 350 °C. The crystallite sizes at each facet were calculated using the Scherrer formula below.
d p = K λ β   c o s θ
The crystallite sizesw for the (110), (101) and (211) facets were calculated using Equation (1) and reported in Table 5 along with the average crystallite sizes of both the commercial and IH samples.
Figure 4 shows the surface morphology of the IH Ru-modified IrO2 and the commercial IrO2 and RuO2 catalyst at a scale of 100 nm. There is a drastic change in the surface morphology of the IH sample compared to the commercial samples. XRD and BET results correlate to the change in surface morphology, where the IH sample has a smoother surface compared to the commercial samples.
Figure 5 represents the TEM images of the IH Ru-modified IrO2 and, commercial IrO2 and RuO2. The TEM image of the IH samples shows small particles that are extremely agglomerated compared to the commercial samples, thus making it difficult to particle size. The commercial samples in Figure 5a,c show cubic particles. The particle size of the IH Ir0.8Ru0.2O2 and commercial IrO2 and RuO2 are reported below.
Figure 6 represents the N2 adsorption–desorption analysis for commercial IrO2 and RuO2 and Ir0.8Ru0.2O2-IH. Both the Ir0.8Ru0.2O2-IH and commercial RuO2 show distinct hysteresis loops; however, the IrO2 does not have the hysteresis loop. The BET surface area reported in Table 6 shows that Ir0.8Ru0.2O2-IH has the largest surface area compared to the commercial RuO2 catalysts, with IrO2 having the smallest surface area. The significant increase in surface area for the Ir0.8Ru0.2O2-IH sample compared to the commercial IrO2 and RuO2 samples is attributed to the nature of synthesized Ir0.8Ru0.2O2, which is amorphous. Amorphous materials usually have higher surface areas due to their irregular structure and the random arrangement of atoms, which creates more surface irregularities and pores.
There is a direct correlation between surface area and pore size, with the IH sample having the smallest pore size and commercial IrO2 having a larger pore size. The BET results correlate with the results found in TEM, SEM and XRD.
Figure 7 represents the CV analysis of Ir0.8Ru0.2O2-IH and commercial IrO2 and RuO2. The addition of Ru to IrO2 caused the CV of Ir0.8Ru0.2O2-IH to resemble that of RuO2 more than IrO2 with a distinct negative tail at the cathodic potential scan at 0 V vs. RHE, which is attributed to hydrogen adsorption (Hads) [7,32]. The Ir(III)/Ir(IV) and Ir(IV)/Ir(V) redox couples are reported in Table 7. An extra peak can be seen for the commercial IrO2 which was reported by [33] and suggested it could be the presence of active Ir(III) sites.
Figure 8 represents the LSV analysis of Ir0.8Ru0.2O2-IH and commercial IrO2 and RuO2. LSV results show an increase in catalytic activity for OER with the addition of Ru to IrO2, which is expected since RuO2 is known to have higher catalytic activity than IrO2. However, as seen in Figure 8, the catalytic activity of RuO2 decreases with increased potential. This result is due to the harsh acidic environment which causes corrosion and catalyst degradation, making the RuO2 catalyst less stable than IrO2. Table 8 represents the current densities of Ir0.8Ru0.2O2-IH and commercial IrO2 and RuO2 at the operational potential of 1.7 V as this is the typical potential a PEM electrolyzer operates. From the results reported in Table 8, it is seen that Ir0.8Ru0.2O2-IH is 3 times more active than both commercial IrO2 and RuO2.
CP studies the OER stability of the catalyst at a fixed current, Figure 9 shows the CP results of the Ir0.8Ru0.2O2-IH and commercial IrO2 and RuO2 that were obtained at the current density of 10 mA.cm−2. Table 9 shows the summary of the OER stability of the same catalysts. The Ir0.8Ru0.2O2-IH sample had excellent stability at a constant potential of ±1.5 V over 24 h. This result is significant compared to the commercial sample, with RuO2 being the least stable, shown in both Figure 9 and Table 9.

6. Conclusions

This study focused on a bimetallic anode of Ir0.8Ru0.2O2-IH, compared with single-metal IrO2 and RuO2 electrodes to assess performance in oxygen evolution reactions (OER). By benchmarking RuO2 and using it to support IrO2, the study demonstrated the enhanced catalytic activity and stability of the bimetallic anode over single metal oxides, showcasing the benefits of metal combinations in improving electrochemical performance. The ratio of Ir/Ru was optimized based on the best-performing mixed metal oxide, synthesized via a modified Adams fusion method. The resulting Ru-modified IrO2 catalyst showed superior electrocatalytic activity and stability compared to commercial IrO2 and RuO2, with an extended operational stability of around 24 h. XRD and TEM analyses revealed that adding Ru to IrO2 produced amorphous particles at 350 °C, while BET analysis indicated a larger surface area and smaller pore size in the synthesized catalyst. These improvements suggest that the modified catalyst is highly suitable for water electrolysis applications.

Author Contributions

Conceptualization, S.P. and D.Z.; methodology, D.Z., J.J.B. and A.S.; software, D.Z., S.C., S.P. and B.J.B.; validation, D.Z.; formal analysis, J.J.B. and A.S.; resources, D.Z., S.C., S.P. and B.J.B.; writing—original draft preparation, S.C.; writing—review and editing, D.Z. and B.J.B.; funding acquisition, S.P. and B.J.B.; supervision, D.Z., S.C., S.P. and B.J.B.; project administration, D.Z. All authors have read and agreed to the published version of the manuscript.

Funding

Funding was provided by the Department of Trade, Industry and Competition of the Republic of South Africa as part of the Technology and Human Resources for Industry Program (THRIP/19/31/08/2018). Some additional funding was provided by the Department of Science and Innovation through the financial and strategic support received from the South African National Energy Development Institute (SANEDI) and the Energy & Water Sector Education Training Authority (EWSETA) in South Africa.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Additional data is available from the authors upon request.

Acknowledgments

We would like to express our gratitude to the Department of Science and Innovation through the financial and strategic support received from the South African National Energy Development Institute (SANEDI) and the Energy & Water Sector Education Training Authority (EWSETA) in South Africa is greatly appreciated. Also, the funding from Technology and Human Resources for Industry Program (THRIP/19/31/08/2018) under the Department of Trade, Industry and Competition of the Republic of South Africa is greatly appreciated. Our special thanks go to students and staff of South Africa Institute of Advanced Material Chemistry at the University of the Western Cape and the Chemistry Department at Cape Peninsula University of Technology for their insightful comments and suggestions, which greatly contributed to improving the quality of this paper. Finally, we are grateful to the anonymous reviewers for their constructive feedback and thoughtful suggestions, which have enhanced the clarity and impact of this manuscript.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Akella, A.; Saini, R.; Sharma, M. Social, economical and environmental impacts of renewable energy systems. Renew. Energy 2009, 34, 390–396. [Google Scholar] [CrossRef]
  2. Rasten, E.; Hagen, G.; Tunold, R. Electrocatalysis in water electrolysis with solid polymer electrolyte. Electrochim. Acta 2003, 48, 3945–3952. [Google Scholar] [CrossRef]
  3. Saxena, R.C.; Adhikari, D.K.; Goyal, H.B. Biomass-based energy fuel through biochemical routes: A review. Renew. Sustain. Energy Rev. 2009, 13, 167–178. [Google Scholar] [CrossRef]
  4. Bhagwat, S.; Olczak, M. Green Hydrogen: Bridging the Energy Transition in Africa and Europe; European University Institute: Fiesole, Italy, 2020. [Google Scholar]
  5. International Renewable Energy Agency (IRENA). Shaping Sustainable International Hydrogen Value Chains; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates, 2024. [Google Scholar]
  6. Zhiznin, S.; Shvets, N.; Timokhov, V.; Gusev, A. Economics of hydrogen energy of green transition in the world and Russia. Part I. Int. J. Hydrogen Energy 2023, 48, 21544–21567. [Google Scholar] [CrossRef]
  7. Felix, C.; Bladergroen, B.J.; Linkov, V.; Pollet, B.G.; Pasupathi, S. Ex-Situ Electrochemical Characterization of IrO2 Synthesized by a Modified Adams Fusion Method for the Oxygen Evolution Reaction. Catalysts 2019, 9, 318. [Google Scholar] [CrossRef]
  8. Marshall, A.; Sunde, S.; Tsypkin, M.; Tunold, R. Performance of a PEM water electrolysis cell using IrxRuyTazO2 electrocatalysts for the oxygen evolution electrode. Int. J. Hydrogen Energy 2007, 32, 2320–2324. [Google Scholar] [CrossRef]
  9. Cheng, Y.; Jiang, S.P. Advances in electrocatalysts for oxygen evolution reaction of water electrolysis-from metal oxides to carbon nanotubes. Prog. Nat. Sci. 2015, 25, 545–553. [Google Scholar] [CrossRef]
  10. Hrbek, T.; Kúš, P.; Košutová, T.; Veltruská, K.; Dinhová, T.N.; Dopita, M.; Matolín, V.; Matolínová, I. Sputtered Ir–Ru based catalysts for oxygen evolution reaction: Study of iridium effect on stability. Int. J. Hydrogen Energy 2022, 47, 21033–21043. [Google Scholar] [CrossRef]
  11. Kumar, S.S.; Himabindu, V. Hydrogen production by PEM water electrolysis—A review. Mater. Sci. Energy Technol. 2019, 2, 442–454. [Google Scholar] [CrossRef]
  12. Khan, M.A.; Zhao, H.; Zou, W.; Chen, Z.; Cao, W.; Fang, J.; Xu, J.; Zhang, L.; Zhang, J. Recent progresses in electrocatalysts for water electrolysis. Electrochem. Energy Rev. 2018, 1, 483–530. [Google Scholar] [CrossRef]
  13. Kumar, S.S.; Lim, H. An overview of water electrolysis technologies for green hydrogen production. Energy Rep. 2022, 8, 13793–13813. [Google Scholar] [CrossRef]
  14. Henkensmeier, D.; Najibah, M.; Harms, C.; Žitka, J.; Hnát, J.; Bouzek, K. Overview: State-of-the Art Commercial Membranes for Anion Exchange Membrane Water Electrolysis. J. Electrochem. Energy Convers. Storage 2021, 18, 024001. [Google Scholar] [CrossRef]
  15. Russell, J.H.; Nuttall, L.J.; Fickett, A.P. Hydrogen generation by solid polymer electrolyte water electrolysis. Am. Chem. Soc. Div. Fuel Chem. Prepr. 1973, 18, 24–40. [Google Scholar]
  16. Bianchi, F.R.; Bosio, B. Operating principles, performance and technology readiness level of reversible solid oxide cells. Sustainability 2021, 13, 4777. [Google Scholar] [CrossRef]
  17. van der Merwe, J.; Uren, K.; van Schoor, G.; Bessarabov, D. Characterisation tools development for PEM electrolysers. Int. J. Hydrogen Energy 2014, 39, 14212–14221. [Google Scholar] [CrossRef]
  18. Grubb, W.T.; Niedrach, L.W. Batteries with solid ion-exchange membrane electrolytes: II. Low-temperature hydrogen-oxygen fuel cells. J. Electrochem. Soc. 1960, 107, 131–135. [Google Scholar] [CrossRef]
  19. Grubb, W.T. Batteries with solid ion exchange electrolytes: I. secondary cells employing metal electrodes. J. Electrochem. Soc. 1959, 106, 275. [Google Scholar] [CrossRef]
  20. Sun, X.; Xu, K.; Fleischer, C.; Liu, X.; Grandcolas, M.; Strandbakke, R.; Bjørheim, T.S.; Norby, T.; Chatzitakis, A. Earth-abundant electrocatalysts in proton exchange membrane electrolyzers. Catalysts 2018, 8, 657. [Google Scholar] [CrossRef]
  21. Nasser, M.; Megahed, T.F.; Ookawara, S.; Hassan, H. A review of water electrolysis–based systems for hydrogen production using hybrid/solar/wind energy systems. Environ. Sci. Pollut. Res. 2022, 29, 86994–87018. [Google Scholar] [CrossRef]
  22. Huang, K.; Li, Y.; Xing, Y. Increasing round trip efficiency of hybrid Li–air battery with bifunctional catalysts. Electrochim. Acta 2013, 103, 44–49. [Google Scholar] [CrossRef]
  23. Ahmed, J.; Mao, Y. Ultrafine iridium oxide nanorods synthesized by molten salt method toward electrocatalytic oxygen and hydrogen evolution reactions. Electrochim. Acta 2016, 212, 686–693. [Google Scholar] [CrossRef]
  24. Aricò, A.S.; Siracusano, S.; Briguglio, N.; Baglio, V.; Di Blasi, A.; Antonucci, V. Polymer electrolyte membrane water electrolysis: Status of technologies and potential applications in combination with renewable power sources. J. Appl. Electrochem. 2013, 43, 107–118. [Google Scholar] [CrossRef]
  25. Adamson, W.; Bo, X.; Li, Y.; Suryanto, B.H.; Chen, X.; Zhao, C. Co-Fe binary metal oxide electrocatalyst with synergistic interface structures for efficient overall water splitting. Catal. Today 2020, 351, 44–49. [Google Scholar] [CrossRef]
  26. Cheng, H.; Su, C.-Y.; Tan, Z.-Y.; Tai, S.-Z.; Liu, Z.-Q. Interacting ZnCo2O4 and Au nanodots on carbon nanotubes as highly efficient water oxidation electrocatalyst. J. Power Sources 2017, 357, 1–10. [Google Scholar] [CrossRef]
  27. Dong, C.; Yuan, X.; Wang, X.; Liu, X.; Dong, W.; Wang, R.; Duan, Y.; Huang, F. Rational design of cobalt–chromium layered double hydroxide as a highly efficient electrocatalyst for water oxidation. J. Mater. Chem. A 2016, 4, 11292–11298. [Google Scholar] [CrossRef]
  28. Liu, P.F.; Yang, S.; Zhang, B.; Yang, H.G. Defect-Rich Ultrathin Cobalt–Iron Layered Double Hydroxide for Electrochemical Overall Water Splitting. ACS Appl. Mater. Interfaces 2016, 8, 34474–34481. [Google Scholar] [CrossRef]
  29. Wang, L.; Saveleva, V.A.; Zafeiratos, S.; Savinova, E.R.; Lettenmeier, P.; Gazdzicki, P.; Gago, A.S.; Friedrich, K.A. Highly active anode electrocatalysts derived from electrochemical leaching of Ru from metallic Ir0.7Ru0.3 for proton exchange membrane electrolyzers. Nano Energy 2017, 34, 385–391. [Google Scholar] [CrossRef]
  30. Elgrishi, N.; Rountree, K.J.; McCarthy, B.D.; Rountree, E.S.; Eisenhart, T.T.; Dempsey, J.L. A practical beginner’s guide to cyclic voltammetry. J. Chem. Educ. 2017, 95, 197–206. [Google Scholar] [CrossRef]
  31. Karels, S.; Felix, C.; Pasupathi, S. Development of unsupported IrO2 nano-catalysts for polymer electrolyte membrane water electrolyser applications. S. Afr. J. Sci. 2024, 120. [Google Scholar] [CrossRef]
  32. 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]
  33. Cheng, J.; Zhang, H.; Chen, G.; Zhang, Y. Study of IrxRu1−xO2 oxides as anodic electrocatalysts for solid polymer electrolyte water electrolysis. Electrochim. Acta 2009, 54, 6250–6256. [Google Scholar] [CrossRef]
  34. Audichon, T.; Mayousse, E.; Morisset, S.; Morais, C.; Comminges, C.; Napporn, T.W.; Kokoh, K.B. Electroactivity of RuO2eIrO2 mixed nanocatalysts toward the oxygen evolution reaction in a water electrolyzer supplied by a solar profile. Int. J. Hydrogen Energy 2014, 39, 16785–16796. [Google Scholar] [CrossRef]
  35. Siracusano, S.; Baglio, V.; Di Blasi, A.; Briguglio, N.; Stassi, A.; Ornelas, R.; Trifoni, E.; Antonucci, V.; Aricò, A. Electrochemical characterization of single cell and short stack PEM electrolyzers based on a nanosized IrO2 anode electrocatalyst. Int. J. Hydrogen Energy 2010, 35, 5558–5568. [Google Scholar] [CrossRef]
Figure 1. Illustration showing how the GC WE were cleaned.
Figure 1. Illustration showing how the GC WE were cleaned.
Appliedchem 04 00022 g001
Figure 2. Illustration showing the setup of the electrochemical cell.
Figure 2. Illustration showing the setup of the electrochemical cell.
Appliedchem 04 00022 g002
Figure 3. XRD spectra of IH Ru-modified IrO2 catalyst and commercial IrO2 and RuO2.
Figure 3. XRD spectra of IH Ru-modified IrO2 catalyst and commercial IrO2 and RuO2.
Appliedchem 04 00022 g003
Figure 4. SEM images of (a) IrO2 commercial, (b) Ir0.8Ru0.2O2-IH and (c) RuO2 commercial at scale of 100 nm.
Figure 4. SEM images of (a) IrO2 commercial, (b) Ir0.8Ru0.2O2-IH and (c) RuO2 commercial at scale of 100 nm.
Appliedchem 04 00022 g004
Figure 5. TEM image of (a) RuO2 commercial, (b) Ir0.8Ru0.2O2-IH and (c) IrO2 commercial at scale of 20 nm.
Figure 5. TEM image of (a) RuO2 commercial, (b) Ir0.8Ru0.2O2-IH and (c) IrO2 commercial at scale of 20 nm.
Appliedchem 04 00022 g005
Figure 6. Nitrogen adsorption and desorption of the (a) IrO2 commercial, (b) RuO2 commercial and (c) Ir0.8Ru0.2O2-IH.
Figure 6. Nitrogen adsorption and desorption of the (a) IrO2 commercial, (b) RuO2 commercial and (c) Ir0.8Ru0.2O2-IH.
Appliedchem 04 00022 g006
Figure 7. CV analysis of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH in 0.5 M of H2SO4 electrolyte.
Figure 7. CV analysis of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH in 0.5 M of H2SO4 electrolyte.
Appliedchem 04 00022 g007
Figure 8. LSV analysis of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH in 0.5 M of H2SO4 electrolyte.
Figure 8. LSV analysis of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH in 0.5 M of H2SO4 electrolyte.
Appliedchem 04 00022 g008
Figure 9. CP analysis of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH in 0.5 M of H2SO4 electrolyte at 10 mA·cm−2.
Figure 9. CP analysis of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH in 0.5 M of H2SO4 electrolyte at 10 mA·cm−2.
Appliedchem 04 00022 g009
Table 1. The parameters of the CV (Cyclic voltammetry) analysis [31].
Table 1. The parameters of the CV (Cyclic voltammetry) analysis [31].
Start potential0.02 V
Upper vertex potential1.2 V
Lower vortex potential−0.2 V
Stop potential0.02 V
Number of scans3
Scan rate0.02 V/s
Step0.00244 V
Interval time0.12207 s
Table 2. The parameters of the LSV (Linear sweep voltammetry) analysis [31].
Table 2. The parameters of the LSV (Linear sweep voltammetry) analysis [31].
Start potential0.8 V vs. RHE
Stop potential1.8 V vs. RHE
Scan rate0.002 V/s
Step0.00244 V
Interval time1.2207 s
Current range1 mA–100 nA
Rotation speed1600 rpm
Table 3. The parameters of CP (Chronopotentiometry) analysis [31].
Table 3. The parameters of CP (Chronopotentiometry) analysis [31].
Applied current0.02 A/cm2
Maximum cutoff voltage1.8 V vs. RHE
Rotation speed1600 rpm
Table 4. Elemental composition of commercial and IH samples.
Table 4. Elemental composition of commercial and IH samples.
Weight %IrO2 CommerczialRuO2 CommercialIrRuO2 IH
O17.3030.5325.90
Ru0.0069.4716.73
Ir69.430.0057.37
K3.720.000.00
Ti0.110.000.00
Cr1.800.000.00
Fe7.160.000.00
Na0.030.000.00
Ir/Ru0.000.000.8/0.2
Table 5. XRD d-spacing and average crystallite size of commercial catalyst and IH Ru-modified IrO2 calculated using Scherrer equation.
Table 5. XRD d-spacing and average crystallite size of commercial catalyst and IH Ru-modified IrO2 calculated using Scherrer equation.
Sample Named(110) (nm)d(101) (nm)d(211) (nm)Average Size (nm)
IrO2 commercial14.8011.8712.3313.00
RuO2 commercial20.9320.9526.2622.71
Ir0.8Ru0.2O2 IH1.612.082.211.97
Table 6. BET surface area and pore size of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH.
Table 6. BET surface area and pore size of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH.
Sample NameBET Surface Area (m2/g)Pore Size (nm)
IrO2 commercial5.7519.91
RuO2 commercial8.4013.95
Ir0.8Ru0.2O2-IH205.642.52
Table 7. Comparison of Commercial and Synthesized IrO2.
Table 7. Comparison of Commercial and Synthesized IrO2.
Sample NameIr(III)/Ir(IV) Redox Couple (V vs. RHE)Ir(IV)/Ir(V) Redox Couple (V vs. RHE)
IrO2 commercial0.741.26
Ir0.8Ru0.2O2-IH0.651.22
Table 8. Current densities of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH at operational voltage of 1.7 V.
Table 8. Current densities of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH at operational voltage of 1.7 V.
IrO2 CommercialRuO2 CommercialIr0.8Ru0.2O2-IH
Operational voltage (V)1.71.71.7
Current density (Acm−2)0.0990.0780.26
Table 9. OER stability of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH in 0.5 M of H2SO4 electrolyte [31].
Table 9. OER stability of commercial IrO2 and RuO2, and Ir0.8Ru0.2O2-IH in 0.5 M of H2SO4 electrolyte [31].
Sample NameOER Stability (h)
IrO2 commercial5
RuO2 commercial3
Ir0.8Ru0.2O2-IH24
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chidziva, S.; Zide, D.; Bambo, J.J.; Sinto, A.; Pasupathi, S.; Bladergroen, B.J. Synthesis and Electrochemical Characterization of Ru-Modified Iridium Oxide Catalysts for PEM Electrolysis. AppliedChem 2024, 4, 353-366. https://doi.org/10.3390/appliedchem4040022

AMA Style

Chidziva S, Zide D, Bambo JJ, Sinto A, Pasupathi S, Bladergroen BJ. Synthesis and Electrochemical Characterization of Ru-Modified Iridium Oxide Catalysts for PEM Electrolysis. AppliedChem. 2024; 4(4):353-366. https://doi.org/10.3390/appliedchem4040022

Chicago/Turabian Style

Chidziva, Stanford, Dorcas Zide, Joshua John Bambo, Anele Sinto, Sivakumar Pasupathi, and Bernard J. Bladergroen. 2024. "Synthesis and Electrochemical Characterization of Ru-Modified Iridium Oxide Catalysts for PEM Electrolysis" AppliedChem 4, no. 4: 353-366. https://doi.org/10.3390/appliedchem4040022

APA Style

Chidziva, S., Zide, D., Bambo, J. J., Sinto, A., Pasupathi, S., & Bladergroen, B. J. (2024). Synthesis and Electrochemical Characterization of Ru-Modified Iridium Oxide Catalysts for PEM Electrolysis. AppliedChem, 4(4), 353-366. https://doi.org/10.3390/appliedchem4040022

Article Metrics

Back to TopTop