Next Article in Journal
Effect of Auxins on the Accumulation of Alkaloids in Ungrafted Annona emarginata (Schltdl.) H. Rainer and Annona emarginata (Schltdl.) H. Rainer Grafted with Annona atemoya Mabb.
Previous Article in Journal
Synthesis and Fluorescence Mechanism of Nitrogen-Doped Carbon Dots Utilizing Biopolymer and Urea
Previous Article in Special Issue
Photo(electro)catalytic Water Splitting for Hydrogen Production: Mechanism, Design, Optimization, and Economy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reduced Graphene Oxide-Coated Iridium Oxide as a Catalyst for the Oxygen Evolution Reaction in Alkaline Water Electrolysis

by
Shengyin Luo
1,†,
Ziqing Zuo
2,† and
Hongbin Sun
1,*
1
College of Sciences, Northeastern University, Shenyang 110819, China
2
Kang Chiao International School, East China Campus, Kunshan 215332, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2025, 30(9), 2069; https://doi.org/10.3390/molecules30092069
Submission received: 9 March 2025 / Revised: 4 May 2025 / Accepted: 5 May 2025 / Published: 7 May 2025
(This article belongs to the Special Issue Design and Mechanisms of Photo(electro)catalysts for Water Splitting)

Abstract

:
Producing hydrogen by water electrolysis has attracted significant attention as a potential renewable energy solution. In this work, a catalyst with reduced graphene oxide (rGO) loaded on IrO2/TiO2 (called rGO/IrO2/TiO2) was designed for the catalytic oxygen evolution reaction (OER). The catalyst was synthesized by coating graphene oxide onto a pretreated IrO2/TiO2 precursor, followed by thermal treatment at 450 °C to achieve reduction and the adhesion of graphene to the substrate. The graphene support retained its intact sp2 carbon framework with minor oxygen-containing functional groups, which enhanced electrical conductivity and hydrophilicity. Benefiting from the synergistic effect of an rGO, IrO2, and TiO2 matrix, the rGO/IrO2/TiO2 catalyst only needed overpotentials of 240 mV and 320 mV to reach 10 mA cm−2 and 100 mA cm−2 in the OER, along with excellent stability over 50 h. Its morphology and crystalline structure were characterized by SEM and XRD spectroscopy, and its electrochemical performance was tested by LSV analysis, EIS impedance spectrum, and double-layer capacitance (Cdl) measurements. This work introduces an innovative and eco-friendly strategy for constructing a high-performance, functionalized Ir-based catalyst.

Graphical Abstract

1. Introduction

Currently, fossil fuels are the primary sources of hydrogen production due to their efficiency and cost-effectiveness [1]. However, with the rising demand for hydrogen as a purely clean energy source and growing concerns about the environmental impact of fossil fuels, such as global warming, it has become imperative to develop sustainable and economical methods for large-scale hydrogen production [2,3]. Currently, hydrogen produced by electrocatalytic water splitting using sustainable electricity is considered an ideal alternative energy source because it is pollution-free and carbon-free in both its production and utilization stages [4,5].
As shown in Equation (1), water electrolysis, or water splitting, is an endothermic reaction and has a ΔG of 237.1 kJ mol−1 under standard conditions (25 °C, 101.325 kPa), indicating a non-spontaneous process. Water electrolysis involves two half-reactions: the hydrogen evolution reaction (HER) at the cathode and the oxygen evolution reaction (OER) at the anode. Among the different electrolysis conditions, alkaline conditions are particularly advantageous for an OER due to their ability to enhance catalyst stability and performance [6]. Alkaline media also offer higher tolerance to corrosion and degradation, making them suitable for long-term industrial applications [7]. In these conditions, water molecules at the cathode gain electrons to produce hydrogen, and the remaining hydroxide anions transport to the anode to lose electrons and produce oxygen, as shown in Equations (2) and (3).
Water splitting: 2H2O (l)→2H2 (g) + O2 (g)
In alkaline electrolyte:
HER: 2H2O (l) + 4e→2H2 (g) + 4OH
OER: 4OH→O2 (g) + 2H2O (l) + 4e
The HER entails a two-electron transfer, while the OER necessitates a four-electron transfer, which requires the reaction to surmount a higher energy barrier. Consequently, individuals are dedicated to discovering exceptional OER electrocatalysts to lower the energy barrier of water electrolysis [8]. Under standard conditions, the theoretical potential to run the water electrolysis reaction is 1.23 V (0 V for HER and 1.23 V for OER, vs. RHE) [9]. However, the actual voltage required for water electrolysis is much higher than the theoretical value due to several factors, including energy barriers related to electrochemical kinetics, mass transport limitations, and resistance arising from the electrode–electrolyte interface [10]. These extra voltage demands are also known as overpotential. Therefore, the actually applied potential (voltage) for water electrolysis (Eap) can be determined using Equation (4), where ηHER and ηOER are overpotentials and iR is the voltage drop due to system resistances.
Eap = 1.23 V + ηHER + ηOER + iR
To reduce the overpotential of an OER, many chemists have attempted to develop high-efficiency OER catalysts, such as noble metals Pt [11], Ru [12], Ir [13], and Pd [14], and their oxides, as well as some transition metal-based alloys such as Ni-Fe [15], Ni-Cu [16], Ni-Co [17], and Ni-Ir [18]. Currently, Ir-based catalysts are considered promising due to their excellent OER performance and stability [19]. However, there is still room to further enhance their electrocatalytic performance. For improvement, Zhang et al. synthesized a Ru@Ir-O catalyst by introducing a core–shell nanostructure with oxygen bound to the Ir shell, leading to an increase in activity and resulting in an extremely low OER overpotential of 238 mV (at 10 mA cm−2) [20]. Zhu et al. prepared the Ir-Co3O4 catalyst by doping dispersed Ir atoms into spinel Ir-Co3O4. The introduction of Ir single atoms increased electronic conductivity and decreased the adsorption energy barrier, resulting in an overpotential of 236 mV (at 10 mA cm−2) [21].
In addition to enhancing the intrinsic activity of electrocatalysts, increasing the number of active sites—regions on the surface of a catalyst that adsorb and desorb reactants to catalyze the reaction—can also improve OER performance, as demonstrated by measuring the electrochemical surface area (ECSA) [22]. Reduced graphene oxide (rGO), a material with an exceptional surface area and electrical conductivity, can provide more catalytically active sites [23]. Moreover, despite the nitrogen gas reduction process that eliminates oxygen-containing functional groups from graphene oxide (GO) and creates large defect regions that enhance electron transfer, carbonyl groups (C=O) at the edges of GO nanosheets may still remain and serve as active sites by adsorbing intermediate species for water splitting [24]. These properties make rGO a promising material for further enhancing OER catalysts. Huang et al. developed an rGO-coated Ni3Se2 catalyst on nickel foam, denoted as rGO/Ni3Se2/NF, which provided abundant active sites due to rGO integration and demonstrated excellent OER and HER performances with low overpotentials of 292.61 mV and 251.01 mV, respectively, at ±10 mA cm−2 [25]. Bhosale et al. developed a vanadium oxide-reduced graphene oxide–nickel oxide (VrG/NiO) electrocatalyst that utilizes active sites at the interface of V2O5, rGO, and NiO to achieve an efficient oxygen evolution reaction (OER) performance with a low overpotential of 155.47 mV at 10 mA cm−2 [26].
Even though they have achieved good performance, the potential benefits of integrating rGO and IrO2 have yet to be fully realized.
Herein, a reduced graphene oxide (rGO)-coated IrO2 catalyst on TiO2 (denoted as rGO/IrO2/TiO2) was designed by solution coating and then calcination at 450 °C for a catalytic oxygen evolution reaction (OER). Through the synergetic effect of rGO and IrO2, the overpotential decreased to 240 mV and 320 mV at 10 mA cm−2 and 100 mA cm−2 in the OER, respectively. Accordingly, it exhibits a low Tafel slope and large double-layer capacitance (Cdl) values. This article highlights the impact of rGO on IrO2 for OER performance, providing innovative insight for developing highly efficient Ir-based catalysts.

2. Experimental Methods

2.1. Chemicals

Every material was commercially obtainable: N-propanol (CH3CH2CH2OH, 99.5%) and oxalic acid (H2C2O4, 99.5%) were procured from Macklin (Shanghai, China), and KOH (99.5%) was purchased from Sinopharm (Beijing, China). H2IrCl6·6H2O (99.5%) was procured from Johnson Matthey (London, UK). Chemicals were utilized as provided, without an additional purification process.

2.2. Characterization Methods

The structural analysis and crystalline phase purity of all samples were examined by X-ray diffraction (XRD) on a Panalytical Empyrean diffractometer (Almelo, The Netherland) utilizing Cu Kα radiation spectra (λ = 1.54 Å, 300 W) in the 2θ range of 20–80°. Scanning electron microscopy (SEM) and energy dispersive spectroscopy (EDS) mapping were performed before and after 50 h, testing for structural morphologies and elemental analysis at 5 kV using a Tescan MAIA3 XMH (15 kV) (Kohoutovice, Czech Republic). Furthermore, Raman spectroscopy was conducted using a Renishaw (Gloucestershire, UK) inVia Raman microscope to analyze vibrational modes and confirm the presence of carbon-based structures. X-ray photoelectron spectroscopy (XPS) was performed on a Shimadzu AXIS SUPRA+ instrument (Kyoto, Japan). The charge was corrected by the C1s = 284.8 eV binding energy standard, and elemental binding energy was analyzed based on the confirmation of C1s at 284.8 eV.

2.3. Electrochemical Measurements

The electrochemical performance of the proposed catalysts was evaluated using the electrochemical workstation CH Instrument Ins (CHI 660) (Austin, TX, USA). A conventional three-electrode setup was employed, consisting of a Pt foil as the counter electrode, an Hg/HgO reference electrode (0.098 V vs. SHE), and rGO/IrO2/TiO2 and IrO2/TiO2 as the working electrodes for all electrochemical measurements. Catalytic activity was assessed in a 1 M KOH electrolyte (pH = 14) at room temperature (25 °C). The working potential of samples was converted from E (V vs. SHE) to E (V vs. RHE) using the Nernst equation—Equation (5):
E V   v s .   R H E = E H g / H g O θ + E H g / H g O + 0.059 × p H i R
where E H g / H g O θ is 0.098 V, E H g / H g O is the measured potential, and the pH is 14. In addition, the current (i/A) was directly converted into current density (j/mA cm−2), as the electrodes had a reacting surface area of 1 cm2. All polarization curves were corrected by iR compensation, where the potential values were corrected for the IR drop by subtracting the voltage loss due to the ohmic resistance of the electrolyte.

2.4. Synthesis of IrO2/TiO2

First, a solution of 1-propanol and chloroiridinic acid with a mass ratio of 14:1 was prepared and stirred thoroughly by a magnetic stirrer (15 min). Using a pipette and an analytical balance, 0.3 g of this solution was added to a 10 × 10 × 0.1 cm3 pure Ti plate that had been acid-etched with 10% HCl (30 min), and was spread evenly on the plate using a brush that had been dipped in the same solution. After the solution dried, the plate was calcined in the atmosphere furnace at 450 °C for 2 h. This process was repeated another 4 times on the same Ti plate, as this could allow the IrO2 to be more evenly distributed and result in fewer cracks between IrO2 particles on the plate. Additionally, the surface of the Ti plate produced titanium dioxide during calcination. As a result, an IrO2 catalyst with approximately 1.0 mg/cm2 of Ir was synthesized. A detailed schematic illustration of the synthesis procedure is depicted as Step 1 in Figure 1.

2.5. Synthesis of rGO/IrO2/TiO2

Another solution of 1-propanol and graphene oxide powder with a mass ratio of 11:1 was prepared and stirred thoroughly by a magnetic stirrer (15 min). Next, 0.3 g of this solution was added to the IrO2/TiO2 and was applied evenly using a brush that had been dipped in the same solution. After the solution dried, the plate was calcined in nitrogen at 450 °C for 2 h, ensuring that the GO reduced successfully. After that, the synthesis of rGO/IrO2/TiO2 with approximately 0.2 mg/cm2 of graphene was finished (Step 2 in Figure 1), and both the IrO2/Ti and the rGO/IrO2/TiO2 were sliced into 1 × 1.5 × 0.1 cm3 samples to prepare for the experiment (Figure S1).

3. Results and Discussion

3.1. Material Characterizations

As shown in Figure 2a, the phase structure and crystallinity of catalysts IrO2/TiO2 and rGO/IrO2/TiO2 were investigated through XRD. XRD spectra of the catalyst IrO2/TiO2 demonstrated characteristic peaks at 28.04°, 34.79°, 40.19°, and 53.99°, matching the (110), (101), (200), and (211) planes of the IrO2 (PDF#00-043-1019), respectively. Likewise, XRD spectra of the catalyst rGO/IrO2/TiO2 demonstrated characteristic peaks at 27.64°, 35.06°, 40.14°, and 54.11°, also matching the (110), (101), (200), and (211) planes of the IrO2 (PDF#00-043-1019), respectively. This indicates that both catalysts contained IrO2. XRD spectra of the catalyst IrO2/TiO2 also demonstrated characteristic peaks at 27.62, 35.98, 41.20, 54.28, and 56.63, matching the (110), (101), (111), (211), and (220) of the rutile TiO2 (PDF#00-001-1292), respectively. Similarly, XRD spectra of the catalyst rGO/IrO2/TiO2 demonstrated characteristic peaks at 27.31, 36.09, 41.24, 54.37, and 56.65, matching the (110), (101), (111), (211), and (220) of the rutile TiO2 (PDF#00-001-1292), respectively. This suggests that the surfaces of the Ti plates oxidized to TiO2. However, only the XRD spectra of the catalyst rGO/IrO2/TiO2 demonstrated little characteristic peaks at 26.66, 42.15, 44.60, 50.65, and 54.73, corresponding to the (002), (100), (101), (102), and (004) of the graphite (PDF#00-043-1019), respectively. This justified the coated rGO in the catalyst.
According to Figure 2b, Raman spectra of bare IrO2/TiO2 and rGO/IrO2/TiO2 samples provide clear evidence of successful graphene deposition and offer insight into the structural quality of the rGO layer. The IrO2/TiO2 control exhibits a featureless baseline in the 1000–3000 cm⁻1 region, confirming the absence of graphitic carbon. Raman peaks at 554.49 cm−1 and 733.47 cm−1 are attributed to the characteristic vibrational modes of iridium oxide (IrO2). In contrast, the rGO/IrO2/TiO2 composite displays the three hallmark bands of graphene-based materials: the D band at 1371.50 cm−1, the G band at ~1590.09 cm−1, and the 2D band near 2916.70 cm−1. The D band, arising from the breathing modes of sp2 carbon rings activated by defects and edge sites, and the G band, corresponding to the in-plane E2g vibration of sp2 carbon networks, are both broadened relative to those of pristine graphene, reflecting residual oxygen functionalities and interfacial interactions with the IrO2/TiO2 substrate [27]. The intensity ratio ID/IG of approximately 0.73 indicates a moderate defect density—typical for reduced graphene oxide—that is advantageous for catalysis because defect sites can serve as additional active centers for electron transfer. Meanwhile, the high I2D/IG ratio of 1.91 is characteristic of few-layer graphene, rather than bulk graphite, and confirms that thermal reduction preserved a relatively thin graphene coating [28]. A slight upshift in the G peak further suggests a charge transfer between the rGO and the underlying oxide, consistent with the enhanced electrical conductivity and accelerated OER kinetics observed in electrochemical measurements. Together, these Raman features validate the presence of a defect-rich, few-layer rGO film whose structural and electronic characteristics underpin the superior electrocatalytic performance of the rGO/IrO2/TiO2 composite.
The surface morphologies of IrO2/TiO2 and rGO/IrO2/TiO2 were examined using a scanning electron microscope (SEM) (Figure 3 and Figure S2). Figure 3a,b present SEM images of rGO/IrO2/TiO2 at different magnifications for comparison. The surface morphology of rGO/IrO2/TiO2 displays a cracked mud-like structure comparable to that of IrO2/TiO2, yet with discernible distinctions. The edges of the polygonal platelets exhibit a slight degree of smoothness and definition, indicative of a thin graphene layer coating the IrO2 structure. At higher magnification (Figure 3b), the surface of the platelets displays a more textured appearance relative to that of IrO2/TiO2, which can be attributed to the graphene coating.
As illustrated in Figure 3c, EDS mappings of rGO/IrO2/TiO2 provide insight into its elemental composition following the graphene coating process. In addition to Ir and O, there was a notable increase in the carbon signal, which was discernible and distributed uniformly across the surface. This result corroborates the successful deposition of the graphene layer onto IrO2/TiO2. Consequently, the distribution of Ir, O, and C appeared to be uniform and followed the pattern of the cracked electrode surface.
Scanning electron microscopy (SEM) imaging revealed a distinct cracked mud-like structure for both catalysts, with rGO/IrO2/TiO2 exhibiting a smoother surface texture, indicative of the graphene coating. Energy-dispersive X-ray spectroscopy (EDS) mapping further confirmed the successful deposition of the graphene layer on the IrO2/TiO2 surface. While the overall cracked mud-like structure was maintained, the graphene coating introduced subtle changes to the surface texture and significantly increased the carbon content of the electrode. These structural modifications likely played a crucial role in the enhanced electrocatalytic properties observed for rGO/IrO2/TiO2, as discussed in subsequent sections.
Further X-ray photoelectron spectroscopy (XPS) analysis was conducted to determine the elemental composition and chemical states of the samples. The XPS survey spectrum (Figure 4 and Figure S3) confirmed the presence of C, O, Ir, and Ti, indicating the successful synthesis of rGO/ IrO2/TiO2. The C 1s XPS spectrum (Figure 4a) exhibited five characteristic peaks at 295.53, 292.81, 289.52, 284.08, and 285.89 eV, corresponding to the Auger peak of Ir, the conjugated structure (π→π*) of graphene, sp2-hybridized graphitic carbon, and sp3-hybridized carbon, respectively. Compared with that of IrO2/TiO2, the C 1s spectrum of rGO/IrO2/TiO2 displayed distinct graphene-related peaks, along with a weakened C=O peak, confirming the presence of reduced graphene oxide (rGO) while maintaining the integrity of the graphene framework. The deconvoluted O 1s XPS spectrum (Figure 4b) revealed two peaks of rGO/IrO2/TiO2 at 531.34 and 529.3 eV, and two peaks of IrO2/TiO2 at 532.49 and 530.96 eV. The incorporation of reduced graphene oxide (rGO) induced metal–oxygen bond reconstruction, while interfacial electron transfer promoted oxygen vacancy formation. As shown in Figure 4c, the Ir 4f XPS spectrum exhibited two characteristic peaks corresponding to 4f₇/2 and 4f5/2 [29]. Notably, rGO/ IrO2/TiO2 displayed only a single oxidation state (4f5/2), unlike IrO2/TiO2. The Ti 2p XPS spectrum (Figure 4d) showed two peaks assigned to 2p1/2 and 2p3/2, confirming the structural integrity and stability of the TiO2 substrate before and after rGO loading. These findings demonstrate that loading rGO onto IrO2/TiO2 not only introduced graphene-like catalytic properties, but also served as a structural support layer, significantly improving the electrode’s electronic conductivity. Moreover, IrO2 and TiO2 remained uniformly distributed without forming low-valence states or alloys.

3.2. Electrocatalytic Properties

A conventional three-electrode electrochemical system was employed to assess the electrocatalytic oxygen evolution reaction (OER) activity of the resulting catalyst in an alkaline solution to ascertain the electrocatalytic performance. Figure 5a shows the polarization curves of each sample. It can be observed that the OER catalytic activity was significantly enhanced after the graphene layer was coated, where its overpotentials at 10 and 100 mA cm−2 were only 240 mV and 320 mV, respectively, compared to 320 mV and 400 mV for the original IrO2 catalyst at the same current density.
Figure 5e shows the Tafel slope of each tested electrode. It demonstrates that the rGO/IrO2/TiO2 had the smallest value of 53.81 mV dec−1, which was less than IrO2/TiO2’s 82.93 mV dec−1. The lower Tafel slope indicated a more rapid reaction kinetics and a more favorable OER mechanism, which was likely associated with the efficient adsorption and desorption of oxygen intermediates, facilitating the four-electron transfer pathway commonly observed for iridium-based catalysts.
Electrochemical impedance spectroscopy (EIS) measurements coupled with equivalent circuit modeling were employed to more accurately elucidate the kinetics of the oxygen evolution reaction (OER) process. Figure 5c shows that the diameter of the semicircle for the rGO/IrO2/TiO2 electrode was notably smaller than that of pristine IrO2/TiO2, indicating that the rGO coating lowered the overall impedance and accelerated interfacial charge transfer. This reduced impedance corroborates the enhanced electron transport pathways provided by the conductive graphene network on the catalyst surface.
Figure 5e shows the capacitive current densities as a function of the scanning rate for each catalyst, which was extracted from the CV curves tested (Figure 5d and Figure S4). These results illustrate that the Cdl of the rGO/IrO2/TiO2 was 54.69 mF·cm−2, higher than the IrO2/TiO2’s 52.08 mF·cm−2, proving that graphene coating on electrode materials can increase their electrochemically active sites, which is a core measurement of the electrocatalytic property.
In addition, to evaluate the stability of the rGO/IrO2/TiO2 catalyst, a chronoamperometry test under the condition of 1.1 V (vs. Hg/HgO) was performed. As shown in Figure 5f, the change was very subtle within 50 h of continuous operation, indicating the excellent electrochemical stability of the catalyst.
Figure 6a compares the LSV polarization curves of the rGO/IrO2/TiO2 electrode recorded initially and after 50 h of continuous OER testing in 1.0 M KOH. The overpotential required to reach 10 mA cm−2 increased only from 240 mV to 251 mV, and that for 100 mA cm−2 rose from 320 mV to 332 mV—corresponding to modest shifts of ~11 mV and ~12 mV, respectively. Such small potential increases demonstrate that the catalyst retained most of its activity under prolonged operation. SEM imaging after 50 h (Figure 6b) shows that the characteristic cracked “mud-like” morphology of the catalyst surface remained intact, and EDS maps (Figure 6c) confirm a uniform distribution of C, O, and Ir. Together, these results indicate that the rGO coating effectively preserved the structural integrity and catalytic performance of the IrO2/TiO2 electrode during extended alkaline electrolysis. For comparative evaluation of electrochemical performance, overpotentials at 10 and 100 mA cm−2 were benchmarked against similar catalysts in Table 1.
Electrochemical measurements showed that the rGO/IrO2/TiO2 catalyst exhibited superior performance. The enhanced performance of the rGO/IrO2/TiO2 catalyst can be ascribed to several contributing factors. The incorporation of rGO is believed to have augmented the overall electrical conductivity of the catalyst, thereby facilitating accelerated electron transfer during the OER process. The interaction between rGO and IrO2 may have created beneficial synergistic effects, potentially altering the electronic structure of the active sites and optimizing the adsorption energies of reaction intermediates. The elevated surface area of rGO furnished supplementary active sites for the OER, as substantiated by the augmented Cdl value. Moreover, the graphene coating may have safeguarded the underlying IrO2 from degradation, thereby enhancing the catalyst’s long-term stability.

4. Conclusions

In summary, this study demonstrated a promising strategy for enhancing the performance of Ir-based OER catalysts through the integration of reduced graphene oxide. The surface morphology of rGO/IrO2/TiO2 exhibited a comparable cracked mud-like structure, with a uniform dispersion of graphene across the surface. XPS analysis revealed that Ir existed in the +4 oxidation state (IrO2), while Ti was present as +4 (TiO2). The graphene support retained its intact sp2 carbon framework (C 1s at 284.8 eV) with minor oxygen-containing functional groups (C=O at 289.52 eV), which enhanced electrical conductivity and hydrophilicity. Raman spectroscopy (I2D/IG = 1.91) further verified the defective structure of graphene, which facilitated metal oxide anchoring and charge transfer. The rGO/IrO2/TiO2 catalyst demonstrated exceptional catalytic activity. It only needed an overpotential of 240 mV to reach 10 mA cm−2 and 320 mV to reach 100 mA cm−2 in the OER, and had the smallest Tafel slope of 53.81 mV dec−1. Its double-layer capacitance (Cdl) was 54.69 mF·cm−2, which further illustrates its large electrochemically active surface area. These findings contribute to the ongoing efforts to develop high-performance catalysts for water electrolysis and provide valuable insights for future research in this field.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30092069/s1, Figure S1. Photographs of (a) rGO/IrO2/TiO2; (b) IrO2/TiO2. Figure S2. Morphology and structure characterizations. (a,b) SEM images of IrO2/TiO2 at different magnifications; (c) EDS mappings of IrO2/TiO2. Figure S3. Summary XPS spectra of rGO/IrO2/TiO2 and IrO2/TiO2. Figure S4. CV curves at 10 to 50 mV s−1 scan rates of IrO2/TiO2. Tabel S1. XPS test data related to rGO/IrO2/TiO2. Tabel S2. XPS test data related to IrO2/TiO2.

Author Contributions

Writing—original draft, S.L. and Z.Z.; Writing—review & editing, H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 21872020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jamadar, A.S.; Sutar, R.; Patil, S.; Khandekar, R.; Yadav, J.B. Progress in metal oxide-based electrocatalysts for sustainable water splitting. Mater. Rep. Energy 2024, 4, 100283. [Google Scholar] [CrossRef]
  2. Akpasi, S.O.; Smarte Anekwe, I.M.; Tetteh, E.K.; Amune, U.O.; Mustapha, S.I.; Kiambi, S.L. Hydrogen as a clean energy carrier: Advancements, challenges, and its role in a sustainable energy future. Clean Energy 2025, 9, 52–88. [Google Scholar] [CrossRef]
  3. Yu, M.; Budiyanto, E.; Tüysüz, H. Principles of Water Electrolysis and Recent Progress in Cobalt-, Nickel-, and Iron-Based Oxides for the Oxygen Evolution Reaction. Angew. Chem. Int. Ed. 2022, 61, e202103824. [Google Scholar] [CrossRef]
  4. Liu, Y.; Vijayakumar, P.; Liu, Q.; Sakthivel, T.; Chen, F.; Dai, Z. Shining Light on Anion-Mixed Nanocatalysts for Efficient Water Electrolysis: Fundamentals, Progress, and Perspectives. Nano-Micro Lett. 2022, 14, 43. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, X.; Wang, Y.; Zou, J.; Zhao, S.; Hou, H.; Yao, W.; Wang, H. Unraveling the oxygen evolution activity of biomass-derived porous carbon plate as self-supported metal-free electrocatalyst for water splitting. Prog. Nat. Sci. Mater. Int. 2024, 34, 967–976. [Google Scholar] [CrossRef]
  6. Fan, R.; Lu, S.; Wang, F.; Zhang, Y.; Hojamberdiev, M.; Chai, Y.; Dong, B.; Zhang, B. Enhancing catalytic durability in alkaline oxygen evolution reaction through squaric acid anion intercalation. Nat. Commun. 2025, 16, 3407. [Google Scholar] [CrossRef] [PubMed]
  7. Allard, C. Efficient catalyst for alkaline water. Nat. Rev. Mater. 2025, 10, 84. [Google Scholar] [CrossRef]
  8. Parra-Puerto, A.; Ng, K.L.; Fahy, K.; Goode, A.E.; Ryan, M.P.; Kucernak, A. Supported Transition Metal Phosphides: Activity Survey for HER, ORR, OER, and Corrosion Resistance in Acid and Alkaline Electrolytes. ACS Catal. 2019, 9, 11515–11529. [Google Scholar] [CrossRef]
  9. Miao, L.; Jia, W.; Cao, X.; Jiao, L. Computational chemistry for water-splitting electrocatalysis. Chem. Soc. Rev. 2024, 53, 2771–2807. [Google Scholar] [CrossRef]
  10. Li, Z.; Wang, D.; Kang, H.; Shi, Z.; Hu, X.; Sun, H.; Xu, J. Atomic cation and anion co-vacancy defects boosted the oxide path mechanism of the oxygen evolution reaction on NiFeAl-layered double hydroxide. J. Mater. Chem. A 2025, 13, 587–594. [Google Scholar] [CrossRef]
  11. Zysler, M.; Shokhen, V.; Hardisty, S.S.; Muzikansky, A.; Zitoun, D. Bifunctional Pt–Ni Electrocatalyst Synthesis with Ultralow Platinum Seeds for Oxygen Evolution and Reduction in Alkaline Medium. ACS Appl. Energy Mater. 2022, 5, 4212–4220. [Google Scholar] [CrossRef]
  12. Bai, J.; Deng, Y.; Lian, Y.; Zhou, Q.; Zhang, C.; Su, Y. WCx-Supported RuNi Single Atoms for Electrocatalytic Oxygen Evolution. Molecules 2023, 28, 7040. [Google Scholar] [CrossRef] [PubMed]
  13. Xing, Y.; Ku, J.; Fu, W.; Wang, L.; Chen, H. Inductive effect between atomically dispersed iridium and transition-metal hydroxide nanosheets enables highly efficient oxygen evolution reaction. Chem. Eng. J. 2020, 395, 125149. [Google Scholar] [CrossRef]
  14. Hussain, A.; Asim, M.; Samanci, M.; Kausar Janjua, N.; Bayrakçeken, A. Oxygen evolution reaction activity of carbon aerogel supported Pd–Ni–Al catalysts synthesized by microwave irradiation method. Int. J. Hydrogen Energy 2024, 81, 93–109. [Google Scholar] [CrossRef]
  15. Huang, T.; Liu, Y.; Zhao, Z.; Liu, Y.; Ye, R.; Hu, J. Amorphous and outstandingly stable Ni(OH)2.0.75H2O@Ni(OH)2/FeOOH heterojunction nanosheets for efficient oxygen evolution performance. Chem. Commun. 2025, 61, 4010–4013. [Google Scholar] [CrossRef]
  16. Gautam, R.P.; Pan, H.; Chalyavi, F.; Tucker, M.J.; Barile, C.J. Nanostructured Ni–Cu electrocatalysts for the oxygen evolution reaction. Catal. Sci. Technol. 2020, 10, 4960–4967. [Google Scholar] [CrossRef]
  17. Cheng, Y.; Guo, X.; Ma, Z.; Dong, K.; Miao, L.; Du, S. Highly Efficient and Stable Mn-Co1.29Ni1.71O4 Electrocatalysts for Alkaline Water Electrolysis: Atomic Doping Strategy for Enhanced OER and HER Performance. Molecules 2025, 30, 1162. [Google Scholar] [CrossRef]
  18. Xu, J.; Cao, S.; Zhong, M.; Chen, X.; Li, W.; Yan, S.; Wang, C.; Wang, Z.; Lu, X.; Lu, X. Iridium incorporated cobalt-based hydroxide on nickel-contained carbon nanofibers renders highly efficient oxygen evolution reaction. Sep. Purif. Technol. 2023, 324, 124638. [Google Scholar] [CrossRef]
  19. Qin, R.; Chen, G.; Feng, X.; Weng, J.; Han, Y. Ru/Ir-Based Electrocatalysts for Oxygen Evolution Reaction in Acidic Conditions: From Mechanisms, Optimizations to Challenges. Adv. Sci. 2024, 11, 2309364. [Google Scholar] [CrossRef]
  20. Zhang, J.; Fu, X.; Xia, F.; Zhang, W.; Ma, D.; Zhou, Y.; Peng, H.; Wu, J.; Gong, X.; Wang, D.; et al. Core-Shell Nanostructured Ru@Ir–O Electrocatalysts for Superb Oxygen Evolution in Acid. Small 2022, 18, 2108031. [Google Scholar] [CrossRef]
  21. Zhu, Y.; Wang, J.; Koketsu, T.; Kroschel, M.; Chen, J.-M.; Hsu, S.-Y.; Henkelman, G.; Hu, Z.; Strasser, P.; Ma, J. Iridium single atoms incorporated in Co3O4 efficiently catalyze the oxygen evolution in acidic conditions. Nat. Commun. 2022, 13, 7754. [Google Scholar] [CrossRef] [PubMed]
  22. Huang, C.-J.; Xu, H.-M.; Shuai, T.-Y.; Zhan, Q.-N.; Zhang, Z.-J.; Li, G.-R. A review of modulation strategies for improving catalytic performance of transition metal phosphides for oxygen evolution reaction. Appl. Catal. B Environ. 2023, 325, 122313. [Google Scholar] [CrossRef]
  23. Yang, P.; Yang, X.; Liu, W.; Guo, R.; Yao, Z. Graphene-based electrocatalysts for advanced energy conversion. Green Energy Environ. 2023, 8, 1265–1278. [Google Scholar] [CrossRef]
  24. Reddy, D.A.; Choi, J.; Lee, S.; Ma, R.; Kim, T.K. Self-assembled macro porous ZnS–graphene aerogels for photocatalytic degradation of contaminants in water. RSC Adv. 2015, 5, 18342–18351. [Google Scholar] [CrossRef]
  25. Huang, S.-J.; Balu, S.; Barveen, N.R.; Sankar, R. Surface engineering of reduced graphene oxide onto the nanoforest-like nickel selenide as a high performance electrocatalyst for OER and HER. Colloids Surf. A Physicochem. Eng. Asp. 2022, 654, 130024. [Google Scholar] [CrossRef]
  26. Bhosale, M.; Thangarasu, S.; Magdum, S.S.; Jeong, C.; Oh, T.-H. Enhancing the electrocatalytic performance of vanadium oxide by interface interaction with rGO and NiO nanostructures for electrochemical water oxidation. Int. J. Hydrogen Energy 2024, 54, 1449–1460. [Google Scholar] [CrossRef]
  27. Jing, F.; Zhang, S.; Shao, H.; Zhang, S.; Shi, P.; Sun, Z. Construction of Co-Ni3B/GDY heterostructured electrocatalyst for boosting oxygen evolution in alkaline media. J. Alloys Compd. 2025, 1010, 177401. [Google Scholar] [CrossRef]
  28. Guo, J.; Pan, L.; Sun, J.; Wei, D.; Dai, Q.; Xu, J.; Li, Q.; Han, M.; Wei, L.; Zhao, T. Metal-free Fabrication of Nitrogen-doped Vertical Graphene on Graphite Felt Electrodes with Enhanced Reaction Kinetics and Mass Transport for High-performance Redox Flow Batteries. Adv. Energy Mater. 2023, 14, 2302521. [Google Scholar] [CrossRef]
  29. Ma, M.; Zhu, W.; Liao, F.; Yin, K.; Huang, H.; Feng, K.; Gao, D.; Chen, J.; Li, Z.; Zhong, J.; et al. Sulfonated carbon dots modified IrO2 nanosheet as durable and high-efficient electrocatalyst for boosting acidic oxygen evolution reaction. Nano Res. 2024, 17, 8017–8024. [Google Scholar] [CrossRef]
  30. Boter-Carbonell, J.; Calabrés-Casellas, C.; Sarret, M.; Andreu, T.; Cabot, P.L. IrOx Supported on Submicron-Sized Anatase TiO2 as a Catalyst for the Oxygen Evolution Reaction. Catalysts 2025, 15, 79. [Google Scholar] [CrossRef]
  31. Zhong, W.; Lin, Z.; Feng, S.; Wang, D.; Shen, S.; Zhang, Q.; Gu, L.; Wang, Z.; Fang, B. Improved oxygen evolution activity of IrO2 by in situ engineering of an ultra-small Ir sphere shell utilizing a pulsed laser. Nanoscale 2019, 11, 4407–4413. [Google Scholar] [CrossRef] [PubMed]
  32. Li, G.; Li, S.; Xiao, M.; Ge, J.; Liu, C.; Xing, W. Nanoporous IrO2 catalyst with enhanced activity and durability for water oxidation owing to its micro/mesoporous structure. Nanoscale 2017, 9, 9291–9298. [Google Scholar] [CrossRef] [PubMed]
  33. Deng, Q.; Li, B.; Wang, J. IrO2-Based Monolithic Electrodes for Efficient and Stable Oxygen Evolution Reaction in Acid. Key Eng. Mater. 2020, 861, 401–406. [Google Scholar] [CrossRef]
Figure 1. Schematic of rGO/IrO2/TiO2 synthesis procedure.
Figure 1. Schematic of rGO/IrO2/TiO2 synthesis procedure.
Molecules 30 02069 g001
Figure 2. Crystal and structural characterization of catalysts IrO2/TiO2 and rGO/IrO2/TiO2. (a) XRD analysis and (b) Raman spectra.
Figure 2. Crystal and structural characterization of catalysts IrO2/TiO2 and rGO/IrO2/TiO2. (a) XRD analysis and (b) Raman spectra.
Molecules 30 02069 g002
Figure 3. Morphology and structure characterizations. (a,b) SEM images of rGO/IrO2/TiO2 at different magnifications and (c) EDS mappings of rGO/IrO2/TiO2.
Figure 3. Morphology and structure characterizations. (a,b) SEM images of rGO/IrO2/TiO2 at different magnifications and (c) EDS mappings of rGO/IrO2/TiO2.
Molecules 30 02069 g003
Figure 4. High-resolution XPS spectra of rGO/IrO2/TiO2 and IrO2/TiO2 for (a) C 1s, (b) O 1s, (c) Ir 4f, and (d) Ti 2p.
Figure 4. High-resolution XPS spectra of rGO/IrO2/TiO2 and IrO2/TiO2 for (a) C 1s, (b) O 1s, (c) Ir 4f, and (d) Ti 2p.
Molecules 30 02069 g004
Figure 5. (a) Polarization curves; (b) corresponding Tafel plots in 1.0 M KOH electrolyte of IrO2/TiO2 and rGO/IrO2/TiO2; (c) EIS Nyquist plots; (d) CV curves at 10 to 50 mV s−1 scan rates; (e) double-layer capacitance (Cdl) values; and (f) stability test for rGO/IrO2/TiO2.
Figure 5. (a) Polarization curves; (b) corresponding Tafel plots in 1.0 M KOH electrolyte of IrO2/TiO2 and rGO/IrO2/TiO2; (c) EIS Nyquist plots; (d) CV curves at 10 to 50 mV s−1 scan rates; (e) double-layer capacitance (Cdl) values; and (f) stability test for rGO/IrO2/TiO2.
Molecules 30 02069 g005
Figure 6. (a) Polarization curves of the rGO/IrO2/TiO2 catalyst after 50 h of continuous electrolysis in 1.0 M KOH; (b) SEM image of the electrode surface after the 50 h stability test (scale bar 10 µm); (c) corresponding EDS elemental maps of C (green), O (red) and Ir (yellow) after 50 h of electrolysis (scale bar 10 µm).
Figure 6. (a) Polarization curves of the rGO/IrO2/TiO2 catalyst after 50 h of continuous electrolysis in 1.0 M KOH; (b) SEM image of the electrode surface after the 50 h stability test (scale bar 10 µm); (c) corresponding EDS elemental maps of C (green), O (red) and Ir (yellow) after 50 h of electrolysis (scale bar 10 µm).
Molecules 30 02069 g006
Table 1. Electrochemical properties.
Table 1. Electrochemical properties.
CatalystElectrolyteOverpotential at 10 mA cm−2 (mV)Overpotential at 100 mA cm−2 (mV)Reference
rGO/IrO2/TiO2 1 M KOH240320This work
IrOx/TiO2 (10:90) composite1 M KOH300Josep Boter-Carbonell et al. [30]
Core–shell IrO2@Ir1 M KOH255Wenwu Zhong et al. [31]
Porous IrO2 (1:100) at 450 °C0.5 M H2SO4276Guoqiang Li et al. [32]
IrO2/Ti foil annealed at 400 °C for 60 h0.5 M H2SO4282391Deng, Qian et al. [33]
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

Luo, S.; Zuo, Z.; Sun, H. Reduced Graphene Oxide-Coated Iridium Oxide as a Catalyst for the Oxygen Evolution Reaction in Alkaline Water Electrolysis. Molecules 2025, 30, 2069. https://doi.org/10.3390/molecules30092069

AMA Style

Luo S, Zuo Z, Sun H. Reduced Graphene Oxide-Coated Iridium Oxide as a Catalyst for the Oxygen Evolution Reaction in Alkaline Water Electrolysis. Molecules. 2025; 30(9):2069. https://doi.org/10.3390/molecules30092069

Chicago/Turabian Style

Luo, Shengyin, Ziqing Zuo, and Hongbin Sun. 2025. "Reduced Graphene Oxide-Coated Iridium Oxide as a Catalyst for the Oxygen Evolution Reaction in Alkaline Water Electrolysis" Molecules 30, no. 9: 2069. https://doi.org/10.3390/molecules30092069

APA Style

Luo, S., Zuo, Z., & Sun, H. (2025). Reduced Graphene Oxide-Coated Iridium Oxide as a Catalyst for the Oxygen Evolution Reaction in Alkaline Water Electrolysis. Molecules, 30(9), 2069. https://doi.org/10.3390/molecules30092069

Article Metrics

Back to TopTop