Next Article in Journal
Construction of Fire Safe Thermoplastic Polyurethane/Reduced Graphene Oxide Hierarchical Composites with Electromagnetic Interference Shielding
Next Article in Special Issue
Assembly of Homochiral Magneto-Optical Dy6 Triangular Clusters by Fixing Carbon Dioxide in the Air
Previous Article in Journal
Heteronuclear Complexes of Hg(II) and Zn(II) with Sodium Monensinate as a Ligand
Previous Article in Special Issue
Ab Initio Electronic, Magnetic, and Optical Properties of Fe Phthalocyanine on Cr2O3(0001)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrocatalytic Properties of Quasi-2D Oxides LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, and Zn) for Hydrogen and Oxygen Evolution Reactions

by
Kinithi M. K. Wickramaratne
and
Farshid Ramezanipour
*
Department of Chemistry, University of Louisville, Louisville, KY 40292, USA
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(13), 3107; https://doi.org/10.3390/molecules29133107
Submission received: 2 May 2024 / Revised: 18 June 2024 / Accepted: 27 June 2024 / Published: 29 June 2024
(This article belongs to the Special Issue Exclusive Feature Papers in Inorganic Chemistry, 2nd Edition)

Abstract

:
Designing cost-effective and highly efficient electrocatalysts for water splitting is a significant challenge. We have systematically investigated a series of quasi-2D oxides, LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, Zn), to enhance the electrocatalytic properties of the two half-reactions of water-splitting, namely oxygen and hydrogen evolution reactions (OER and HER). The four materials are isostructural, as confirmed by Rietveld refinements with X-ray diffraction. The oxygen contents and metal valence states were determined by iodometric titrations and X-ray photoelectron spectroscopy. Electrical conductivity measurements in a wide range of temperatures revealed semiconducting behavior for all four materials. Electrocatalytic properties were studied for both half-reactions of water-splitting, namely, oxygen-evolution and hydrogen-evolution reactions (OER and HER). For the four materials, the trends in both OER and HER were the same, which also matched the trend in electrical conductivities. Among them, LaSrMn0.5Co0.5O4 showed the best bifunctional electrocatalytic activity for both OER and HER, which may be attributed to its higher electrical conductivity and favorable electron configuration.

1. Introduction

The development of green and sustainable energy solutions is crucial to addressing the energy challenges of the future. The persistent use of carbon-rich fossil fuels leads to carbon dioxide emissions, worsening the greenhouse effect. Additionally, the harmful gases produced during the burning of those fuels cause environmental problems. These challenges have inspired the scientific community to explore renewable and sustainable energy sources. Hydrogen is widely considered an attractive alternative to conventional fuels due to its high energy density and absence of carbon emissions when used as a fuel. Among different methods of hydrogen production, electrochemical water splitting is highly promising. It is an excellent approach for converting and storing renewable energy if cost-effective and efficient catalysts are used to facilitate large-scale hydrogen production. Electrochemical water splitting involves two half-cell reactions, the hydrogen-evolution reaction (HER) and the oxygen-evolution reaction (OER). The OER also occurs in other energy storage and conversion devices, including rechargeable metal–air batteries (MxO2 → Mx + O2) [1]. The OER in water electrolysis involves a four-electron-transfer mechanism (H2O → O2 + 4H+ + 4e) with a more complex reaction pathway and slower kinetics compared to the HER, which follows a two-electron-transfer mechanism (2H+ + 2e → H2). There are efficiency challenges associated with water electrolysis due to the sluggish kinetics of the two half-reactions, particularly the OER. This leads to substantial overpotentials beyond the theoretical thermodynamic potential of 1.23 V versus the reversible hydrogen electrode (RHE) [2]. Although catalysts based on noble metals, such as Pt, Ru, and Ir, often exhibit very high performance for water-splitting, their high cost and scarcity limit their wide-scale application, making it necessary to explore cost-effective and earth-abundant alternatives [3,4].
Perovskite oxides have received growing interest as non-precious metal-based catalysts due to their versatility in composition, tunable electronic structures, affordability, and strong inherent catalytic performance [5,6,7,8]. Partial replacement or doping with different metal cations leads to changes in their performance for different applications driven by the resulting modifications to their physical, chemical, and electronic properties [9,10]. Several perovskite oxides, such as the benchmark catalyst Ba0.5Sr0.5Co0.8Fe0.2O3−δ (BSCF), have demonstrated efficient OER catalysis, gaining prominence for their OER performance in alkaline environments, on par with IrO2 [5]. Studies have also explored the HER characteristics of perovskite oxides such as Pr0.5(Ba0.5Sr0.5)0.5Co0.8Fe0.2O3-δ [11] and SrNb0.1Co0.7Fe0.2O3-δ [12]. A variety of other structural arrangements are derived from perovskites, such as double perovskite, brownmillerite, Ruddlesden-popper, and the so-called K2NiF4-type [13] structure. The latter is the topic of the present work and has a quasi-2D structure (Figure 1) with the general formula A2BO4, where A is usually a rare-earth or alkaline-earth metal, and B is frequently a transition metal. Some materials with this structure type have been reported to show electrocatalytic activity for OER or HER, such as SrLaFeO4, SrLaCo0.5Fe0.5O4, and SrLaCoO4-δ [13]. Nevertheless, oxide catalysts that demonstrate efficient catalytic activity for both OER and HER are not commonly found.
In the present work, we investigated a series of quasi-2D oxides LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, Zn), with the aim of improving their electrocatalytic performance for HER and OER through the incorporation of different transition metals. Previous reports on these materials have been limited to their synthesis and some structural characterizations [14,15,16,17,18]. Their electrocatalytic properties have not been previously studied. Our comprehensive studies involved the investigation of their electrical and electrocatalytic properties, as well as detailed X-ray diffraction, X-ray photoelectron spectroscopy (XPS), and scanning electron microscopy (SEM) studies. Our findings showed semiconducting behavior in all four compounds, with electrical conductivity increasing as a function of temperature. The trend in electrical conductivity correlated with that of the electrocatalytic activity for both HER and OER. Importantly, the bifunctional electrocatalytic properties found in these materials are remarkable, particularly for LaSrMn0.5Co0.5O4, which exhibited the highest electrical charge-transport and superior electrocatalytic performance.

2. Results and Discussion

2.1. Crystal Structure

Powder X-ray diffraction (XRD) was employed to assess the crystal structures of LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, Zn) using Rietveld refinements (Figure 1). Table 1 shows a representative example of refined structural parameters for LaSrMn0.5Co0.5O4. The structural parameters for the other three materials are listed in Tables S1–S3. The data revealed that all samples consisted of a single phase, indicating the absence of impurities. The space group and structures of these materials were consistent with previous reports [14,15,16,17,18]. All four materials are isostructural with a body-centered tetragonal space group I4/mmm. The crystal structure is shown in Figure 2. All four materials can be described by the general formula A2BO4, where the A-site is occupied by a mix of La and Sr, while the B-site contains a mix of Mn with Co, Ni, Cu, or Zn. Their structures comprise BO6 octahedra that share corners to form a 2D layer. The layers are separated from each other, and the gap between them is where the A-site cations reside. The A-site cations have 9 oxygens in their coordination environment. The morphologies of sintered pellets were probed using scanning electron microscopy (SEM), and the results are shown in Figure 3, indicating similar microstructures for the four samples. The EDS mapping showed the surface elemental compositions, as shown in Figure S1, which illustrates representative data for LaSrMn0.5Co0.5O4, matching the expected atomic percentages.

2.2. Oxidation State and Oxygen Content

The synthesis conditions and elevated temperatures necessary for synthesizing these compounds can result in different valence states, which can affect the oxygen contents [19,20]. The oxygen contents of these materials were determined by iodometric titrations, which were repeated at least three times with excellent reproducibility, giving oxygen stoichiometries of 4.0173 ± 0.0016 for LaSrMn0.5Co0.5O4, 3.8530 ± 0.0073 for LaSrMn0.5Cu0.5O4-δ, 4.0039 ± 0.0105 for LaSrMn0.5Ni0.5O4, and 4.0120 ± 0.0164 for LaSrMn0.5Zn0.5O4. Therefore, iodometric titrations indicate four oxygens per formula unit for LaSrMn0.5Co0.5O4, LaSrMn0.5Ni0.5O4, and LaSrMn0.5Zn0.5O4, matching the expected stoichiometries. However, LaSrMn0.5Cu0.5O4-δ shows oxygen deficiency, with an oxygen stoichiometry of ~3.85 per formula unit (δ ≈ 0.15).
We also investigated all four materials by X-ray photoelectron spectroscopy (XPS) to determine the oxidation states of transition metals. Given the close overlap of binding energies of Mn3+ and Mn4+ peaks, we also conducted additional measurements of standard Mn2O3 and MnO2 samples (Figure 4). The manganese spectra for three materials, LaSrMn0.5Co0.5O4, LaSrMn0.5Ni0.5O4, and LaSrMn0.5Cu0.5O4-δ, are very similar to each other and encompass binding energies corresponding to both trivalent and tetravalent manganese, as evident from comparison to Mn2O3 and MnO2 standards in Figure 4 and previously reported manganese binding energies in other oxide materials [21,22,23]. Therefore, these three materials contain manganese in a mixed +3/+4 state. On the other hand, the manganese spectrum of LaSrMn0.5Zn0.5O4 shows peaks at a higher binding energy compared to the other three compounds, indicating a significantly greater tetravalent character. This observation is consistent with iodometric titration results that indicated four oxygens per formula unit. This oxygen content requires manganese in a tetravalent state, given the divalent state of zinc in LaSrMn0.5Zn0.5O4.
For LaSrMn0.5Co0.5O4, the cobalt spectrum was obtained (Figure 5a) and showed binding energies consistent with a mix of +2/+3 oxidation states. Divalent cobalt is expected to give a 2p3/2 peak just below 780 eV [24], while the peak for trivalent cobalt is expected to be just above 780 eV [24]. The Co 2p3/2 peak for our material is centered around 780 eV, with a satellite peak at about 786 eV, indicative of Co2+ [25], and a satellite at about 789 eV, indicative of Co3+ [24]. We ruled out the presence of tetravalent cobalt, as we did not observe any pronounced peak at ~281.5 eV [24]. These analyses are consistent with iodometric titration results, indicating four oxygens per formula unit for LaSrMn0.5Co0.5O4, and the manganese XPS spectrum, indicating manganese in mixed +3/+4 state, which would require cobalt to be in mixed +2/+3 state for the charge neutrality of the material.
For LaSrMn0.5Ni0.5O4, the nickel spectrum was obtained, as shown in Figure 5b. In this spectrum, the first peak, just above 850 eV, corresponds to La3+ 3d3/2 [26]. The 2p3/2 peak for Ni2+ is expected to appear close to 854 eV [26,27,28], but higher binding energies, such as ~855 eV, have also been reported for divalent nickel [29]. On the other hand, the Ni3+ 2p3/2 peak is expected to be close to 856–857 eV [26,27,29]. In our nickel spectra, the 2p3/2 peak is centered around ~254.6 eV, consistent with the predominant presence of Ni2+, and a small high-energy shoulder at ~856.5 eV, which may indicate the presence of some Ni3+.
For LaSrMn0.5Cu0.5O4-δ, the copper spectrum was obtained, as illustrated in Figure 5c. The Cu2+ 2p3/2 peak is expected to appear at ~933.5–934.6 eV [26,30,31,32], while the peak for Cu1+ should be close to ~932.4 eV [30,31,32]. In our copper spectra, the 2p3/2 peak is centered around ~933.7 eV, with strong satellite peaks at about 7–10 eV higher than the 2p3/2 peak, which is considered the signature of Cu2+ [31,33]. These observations are consistent with the presence of copper in a predominantly divalent state. They also match the observation of a Mn3+/Mn4+ mixture in the manganese spectrum and the oxygen vacancies determined by iodometric titrations, which will be required for charge neutrality.
For LaSrMn0.5Zn0.5O4, the zinc spectrum was obtained, as illustrated in Figure 5d. As expected, zinc is in divalent state, evident from the 2p3/2 peak at ~1021.3 eV [26,34]. As stated in the discussion of the manganese spectrum, the presence of Zn2+ and Mn4+ are consistent with iodometric titration results indicating four oxygens per formula unit.

2.3. Electrical Conductivity

We conducted a study on the electrical conductivity of all four compounds over a range of temperatures, from 25 to 800 °C (298 to 1073 K). The resistance is determined by applying Ohm’s law to the current response obtained from DC measurements, and this resistance value is subsequently converted into electrical conductivity, denoted by σ (Equation (1)) [35].
σ = l R A = I V .   l A
In this equation, l indicates the thickness of the measured pellet, and A represents the cross-sectional area of the pellet through which the current is applied. Table 2 lists the conductivity values at room temperature, and Figure 6a shows the plot of conductivity at different temperatures.
As the temperature rises, the conductivities of all four materials increase, a behavior that is commonly observed in semiconducting materials. As temperature rises, charge carriers become more mobile, resulting in increased conductivity described by Equation (2).
σ = n e μ
In this equation, σ represents electrical conductivity, n denotes the concentration of charge carriers, e stands for charge of an electron, and μ represents the mobility of charge carriers [36,37]. LaSrMn0.5Co0.5O4 shows higher conductivity than the other three materials in the whole temperature range. The electrical conductivity in oxides typically occurs by electron/hole hopping through metal–oxygen–metal (M–O–M) bonds [38]. For this process to happen, metals with more than one stable oxidation state, such as Mn3+/Mn4+ or Co3+/Co4+, are preferred. The transition metal 3d orbitals overlap with oxygen 2p orbitals, allowing the electron hopping through the M–O–M pathway. The activation energy (Ea) for the temperature-dependent increase in the conductivity of these materials can be calculated using the Arrhenius equation (Equation (3)) [35,39].
σ T = σ 0 e E a K T
The activation energy (Ea) for the change in conductivity with temperature can be determined by analyzing the slope of the line of best fit in the log σT versus the 1000/T plot. In this equation, σ° is a pre-exponential factor, while Ea, k, and T represent activation energy, the Boltzmann constant, and absolute temperature, respectively. Figure 6b shows the Arrhenius plots for the four materials. The activation energies, determined from the slopes, are presented in Table 2. It is important to emphasize that the activation energy signifies the energy barrier associated with the temperature-dependent increase in conductivity. The Ea values are consistent with the electrical conductivity trend, where LaSrMn0.5Co0.5O4, which exhibits higher conductivity, also has a lower activation energy.

2.4. Hydrogen Evolution Reaction

The hydrogen evolution reaction is the cathodic half-reaction of water electrolysis and involves a two-electron transfer process [6,40]. To overcome the sluggish reaction kinetics and accelerate the rate of the HER, it is necessary to apply an overpotential. Designing efficient and stable electrocatalysts is essential to improving the reaction kinetics and minimizing the overpotential. The mechanism of the HER involves a series of steps in both acidic and alkaline media [41]. Each condition has advantages and disadvantages. The efficiency of HER in acidic media is expected to be better than that of alkaline environments due to faster kinetics [42]. The source of the hydrogen adsorbed on the catalyst in acidic media is a proton, whereas in alkaline HER, the adsorbed hydrogen comes from breaking the bond in the H2O molecule [43], as shown below.
  • Volmer reaction (acidic): H3O+ + M + e ⇌ M − H * + H2O
  • Volmer reaction (alkaline): H2O + M + e ⇌ M − H * + OH
  • Heyrovsky reaction (acidic): M − H * + H3O+ + e ⇌ M + H2 + H2O
  • Heyrovsky reaction (alkaline): M − H * + H2O + e ⇌ M + H2 + OH
  • Tafel reaction (acidic and alkaline): 2M − H * ⇌ 2M + H2
Alkaline electrolysis may be favored in an industrial setting [44], due to some limitations of acidic conditions, such as problems with the stability of metal oxides in acids [13]. In this work, the HER experiments were performed in both alkaline (1 M KOH) and acidic (0.5 M H2SO4) conditions. In an acidic medium (Figure 7a), among the four materials, LaSrMn0.5Co0.5O4 shows the lowest overpotential of η10 = 427 mV at 10 mA/cm2, which is a current density used by convention and corresponds to a 12.3% solar-to-hydrogen efficiency. It is followed by LaSrMn0.5Cu0.5O4-δ showing η10 = 535 mV, LaSrMn0.5Ni0.5O4 with η10 = 647 mV, and LaSrMn0.5Zn0.5O4 with η10 = 724 mV. Therefore, the HER activities show the following order: LaSrMn0.5Zn0.5O4 < LaSrMn0.5Ni0.5O4 < LaSrMn0.5Cu0.5O4-δ < LaSrMn0.5Co0.5O4. The acidic HER overpotential of LaSrMn0.5Co0.5O4 is not as low as those observed in some other catalysts such as MoO3-y nanofilms (η10 = 0.201 V) [45], WO3 nanoplates (η10 = 0.117 V) [46], and LaCa2Fe3O8 (η10 = 0.400 V) [47]. However, it is lower than those reported for some other oxides in acidic media, such as perovskite LaFeO3 (η10 = 0.490 V) [48], Ruddlesden-Popper oxide Sr2LaCoMnO7 (η10 = 0.612 V) [49], bilayered brownmillerite Ca3Fe2MnO8 (η10 = 0.507 V) [21], and K2NiF4-type oxide SrLaCoO4-δ (η10 = −0.547 V) [13].
To explore the reaction kinetics for the four materials, the Tafel equation was used (Equation (4)) [50,51].
η = a + b   l o g j
where η is the overpotential and j is the current density. The slope of η vs. logj is indicative of the reaction kinetics. A smaller slope suggests a faster reaction, as it indicates that a small change in overpotential is associated with a large change in current density [51,52]. As shown in Figure 7b, the Tafel slopes of the four materials show the same trend as their overpotentials. LaSrMn0.5Co0.5O4 shows the smallest Tafel slope among the four compounds, indicating faster reaction kinetics and consistent with its greater electrocatalytic activity. Chronopotentiometry experiments performed on the most active material, LaSrMn0.5Co0.5O4, exhibit a stable response for over 12 h, as shown in the inset of Figure 7a. Furthermore, X-ray diffraction data, before and after 100 cycles of HER (Figure 7c), indicates little change, confirming that the material’s structural integrity remains unaltered. We also calculated the mass activity to evaluate the catalysts’ inherent qualities. In Figure 7d, we have compared the mass activities of the four compounds at various overpotentials, showing that LaSrMn0.5Co0.5O4 has a consistently higher mass activity.
We also determined the double-layer capacitance, Cdl, (Figure 8) by analyzing the average current densities (javg) at various scan rates (υ), as obtained from cyclic voltammograms in a non-faradaic region (Figure S2) [53]. The importance of Cdl is that it is proportional to the electrochemically active surface area [54]. The slope of the plot depicting javg against ν provides the double-layer capacitance, calculated according to Equation (5) [6].
javg = Cdl × ν
Notably, the Cdl values align with the trend observed in electrocatalytic activity, with the most active catalyst, LaSrMn0.5Co0.5O4, also displaying the highest Cdl value.
The HER activities of the four materials in basic conditions, 1 M KOH, show a similar trend to that in acidic conditions, where LaSrMn0.5Co0.5O4 has the best activity, followed by LaSrMn0.5Cu0.5O4-δ, LaSrMn0.5Ni0.5O4, and LaSrMn0.5Zn0.5O4 (Figure 9a). The overpotential values (η10) at 10 mA/cm2 are 0.482 V, 0.680 V, 0.740 V, and 0.805 V, respectively. Similarly, the same trend is observed for the HER kinetics, as evident from the Tafel slopes shown in Figure 9b. XRD data, before and after 100 HER cycles (Figure 9c), indicate that the structure remains unchanged, confirming the stability. In addition, the comparison of mass activity at various overpotentials in the basic medium again shows consistently higher mass activity for LaSrMn0.5Co0.5O4 (Figure 9d).
We also determined the double-layer capacitance (Cdl) in 1 M KOH (Figure 10) using cyclic voltammograms in the non-faradaic region (Figure S2). Among the four materials, LaSrMn0.5Co0.5O4 displays the highest Cdl, followed by LaSrMn0.5Cu0.5O4-δ, LaSrMn0.5Ni0.5O4, and LaSrMn0.5Zn0.5O4, the same trend as the electrocatalytic activity.

2.5. Oxygen Evolution Reaction

The OER electrocatalytic activities of the four materials were studied in both acidic and alkaline media. These materials showed little OER activity in acidic conditions, but the alkaline OER activity, especially for two of the materials, was substantial. Therefore, alkaline OER is discussed here. As demonstrated by the polarization curves in Figure 11a, the electrocatalytic performance for OER follows a similar trend to that observed in the HER. Among the four compounds, LaSrMn0.5Ni0.5O4, and LaSrMn0.5Zn0.5O4 did not produce sufficient current to reach 10 mA cm−2 within the experimental potential range. On the other hand, LaSrMn0.5Cu0.5O4-δ shows an overpotential of η10 = 550 mV, beyond the thermodynamic potential of 1.23 V at 10 mA cm−2, while LaSrMn0.5Co0.5O4 exhibits a lower overpotential of η10 = 450 mV. The OER overpotential of LaSrMn0.5Co0.5O410 = 450) is not as low as those observed for some oxides such as IrO2 (400 mV) and RuO210 ≈ 420 mV) [49]. Nevertheless, LaSrMn0.5Co0.5O4 shows a better activity than several previously reported oxides, including well-known catalyst BSCF (η10 ≈ 500 mV) [5], La0.5Sr0.5Co0.8Fe0.2O310 = 600 mV) [55], and La0.6Sr0.4CoO3−δ10 = 590 mV) [56].
The kinetics of the OER were analyzed using Tafel plots (Figure 11b), indicating a smaller Tafel slope, i.e., faster reaction kinetics, for LaSrMn0.5Co0.5O4 compared to the other three materials and the same trend as the OER activity. The stability of the best catalyst in the series, LaSrMn0.5Co0.5O4, was evaluated by chronopotentiometry, showing a steady response for at least 12 h, as shown in the inset of Figure 11a. Also, X-ray diffraction data after 100 cycles of OER confirmed the material’s structural integrity was preserved (Figure 11c). The mass activities for OER at various overpotentials are shown in Figure 11d, highlighting the significantly higher activity of LaSrMn0.5Co0.5O4. Figure 12 shows the TEM images before and after the chronopotentiometry experiment under OER conditions for LaSrMn0.5Co0.5O4.
Some comments on the observed electrocatalytic properties are in order. The trend in electrocatalytic activity matches the trend in electrical conductivity. It is important to note that both OER and HER involve the transfer of electrons and are affected by the conductivity of catalysts. Some previous studies have shown that perovskite oxides with better electronic conductivity exhibit improved electrocatalytic performance [6,36].
There are also other parameters to consider. Some researchers have suggested that the filling of eg-level orbitals can be a descriptor for the OER activity [5]. They proposed that high OER activity is expected for oxides, containing metal cations that have eg occupancy close to one, due to the optimum binding with reaction intermediates. In the best catalyst in our series, LaSrMn0.5Co0.5O4, a mixture of Co2+ and Co3+ oxidation states, is observed. Previous studies have indicated that the intermediate spin state is prominent for Co3+ in some perovskite oxides [5,57]. This intermediate spin state would give an electron configuration of t2g5 eg1 for Co3+. Also, the mixed oxidation states of transition metals can accelerate OER activity [1,43], an advantage offered by the presence of cobalt. Regarding the next best material in the series, LaSrMn0.5Cu0.5O4-δ, it is possible that the observed activity is influenced by the presence of oxygen vacancies in this material, as confirmed by iodometric titrations and XPS, which might facilitate the adsorption of reaction intermediates [6,47,58]. In addition, some studies on HER catalysts containing both copper and nickel have shown the better activity of copper compared to nickel. For example, a study of NiCu showed that surface Cu exhibits a d-band center closer to that of the highly active catalyst Pt [59]. The Ni site of NiCu was found to play a less substantial role in HER, primarily due to its d-band being too high, while Cu emerged as the more active catalytic site [59,60]. Finally, the least active material in the series, LaSrMn0.5Zn0.5O4, has the lowest electrical conductivity and contains zinc, which is only in a divalent state.

3. Experimental Methods

Synthesis and Characterization. Polycrystalline samples of LaSrMn0.5Co0.5O4, LaSrMn0.5Ni0.5O4, LaSrMn0.5Cu0.5O4-δ, and LaSrMn0.5Zn0.5O4 were synthesized under an argon atmosphere by solid-state synthesis method. The stoichiometric proportions of precursors La2O3, SrCO3, MnO4, CoO, NiO, CuO, and ZnO were thoroughly mixed in agate mortar and pestle. For example, for the synthesis of LaSrMn0.5Co0.5O4, a mixture of La2O3 (0.1500 g), SrCO3 (0.1359 g), CoO (0.0343 g), and MnO2 (0.0400 g) was used. The powder mixtures were pressed into pellets and were heated at 1200 °C for 24 h followed by slow cooling. The heating and cooling rates of the furnace for all samples were set at 100 °C/h. The phase purity and structure of polycrystalline samples were determined by powder X-ray diffraction (XRD) at room temperature using Cu Kα1 radiation (λ = 1.54056 Å) on an X-ray diffractometer equipped with a monochromator. Rietveld refinements were carried out using GSAS [61] program and EXPIGU interface [62]. The oxygen content was determined using iodometric titrations [63], where excess potassium iodide (2 g) and 50 mg of the sample were dissolved in 100 mL of argon-purged 1 M HCl, allowing the mixture to react overnight. Then, 5 mL of the reacted mixture, containing the generated iodine, was titrated against 0.025 M Na2S2O3 (I2 + 2S2O32− → 2I + S4O62−) using starch indicator (0.6 mL), which was added near the titration endpoint. Excess KI reduced the oxide samples to form oxides with the lowest stable oxidation states of transition metals. The amount of Na2S2O3 needed for titration indicated the amount of oxygen lost during metal ion reduction. Measurements were repeated three times for error analysis. The morphologies of the samples were studied by Scanning Electron Microscopy (SEM) on sintered pellets using a Thermo Fisher Apreo C LoVac Field Emission SEM at a magnification of 5000×, accompanied by an energy-dispersive X-ray spectroscopy (EDS) detector. X-ray photoelectron spectroscopy (XPS) measurements were performed using a ThermoFisher Scientific K-Alpha instrument. The instrument employed an aluminum monochromatic X-ray source with an energy of hν = 1486.69 eV and incorporated an electron flood gun for effective charge neutralization. Wide survey scans were carried out at a pass energy of 160 eV, while high-resolution scans were performed with a pass energy of 20 eV.
Electrical Conductivity. Electrical conductivity measurements were conducted at variable temperatures from 25 °C to 800 °C. These measurements were performed using a two-probe direct current (DC) method [64] on pellets that had been sintered at 1250 °C. Prior to conducting the measurements, a layer of gold paste was applied to both sides of the pellet and then dried by heating for 3 h at 800 °C. Gold wires were attached to gold foils and used as electrodes, ensuring contact with the two gold-painted surfaces on each side of the pellet. A voltage of 0.01 V was applied for the measurements, which were conducted at approximately 100 °C intervals. Equilibrium conductivity was achieved at each measurement temperature after about 30 min, as indicated by a stable plateau in the DC conductivity data. Heating and cooling rates during the conductivity measurements were maintained at 3 °C/min.
Electrochemical Measurements. The catalyst ink for electrochemical measurements was prepared using 35 mg of the catalyst material, 7 mg of carbon black powder, 40 μL of Nafion D-521 solution (5% w/w in water and 1-propanol), and 7 mL of Tetrahydrofuran (THF). The mixture was ultrasonically dispersed in water for 30 min. The drop-casting was carried out by placing two coats of 10 μL of the mixture onto the surface of a glassy carbon electrode (GCE) with a diameter of 5 mm and an area of 0.196 cm2, with a mass loading of 1.02 mg/cm2, followed by overnight air-drying. This catalyst-loaded electrode was used as the working electrode. Electrochemical measurements were conducted using a standard three-electrode electrochemical cell connected to a rotating disk electrode at 1600 rpm. The OER and HER experiments were carried out in 1 M KOH solution by using a Hg/HgO reference electrode (1 M NaOH). The HER experiments were also conducted in 0.5 M H2SO4 solution using Ag/AgCl (4 M KCl) as the reference electrode. A carbon electrode was used as the counter electrode for all experiments. All potentials were iR-corrected and converted to potential vs. reversible hydrogen electrode (RHE) using the Nernst equation EvsRHE = Evs Reference electrode + 0.059 pH + E0 Reference electrode, where E0Ag/AgCl = 0.197 V for 4 M KCl and E0Hg/HgO = 0.098 for 1 M NaOH [36,65]. For each material, the electrocatalytic measurements were repeated at least three times, using at least two different batches synthesized independently. Chronopotentiometry was utilized to investigate the catalyst stability under both HER and OER conditions, employing the same three-electrode setup and a constant current of 10 mA/cm2.

4. Conclusions

A range of parameters can influence electrocatalytic properties, such as electrical conductivity, the presence of metals with more than one stable oxidation state, oxygen-vacancies, and electron configuration. These parameters were studied in this work through the investigation of a series of quasi-2D oxides to reveal their structural and electrocatalytic properties. X-ray diffraction, X-ray photoelectron spectroscopy, iodometric titrations, electrical conductivity studies, and OER and HER experiments were used for the investigation of the four materials LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, and Zn). The electrical conductivity measurements, as well as electrocatalytic HER and OER, all showed the same trend, LaSrMn0.5Zn0.5O4 < LaSrMn0.5Ni0.5O4 < LaSrMn0.5Cu0.5O4-δ < LaSrMn0.5Co0.5O4. The best-performing compound in this series, LaSrMn0.5Co0.5O4, can act as a bifunctional catalytic material, facilitating both OER and HER. Its lower overpotential, faster kinetics, and greater mass activity were demonstrated and correlated with several descriptors that can be responsible for its higher electrocatalytic performance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29133107/s1, Figure S1: SEM images and corresponding EDS; Figures S2 and S3: Cyclic voltammetry data in non-faradaic region in 0.5 M H2SO4 and 1 M KOH, respectively; Tables S1–S3: Refined structural parameters.

Author Contributions

Conceptualization, F.R.; Methodology, K.M.K.W. and F.R.; Validation, K.M.K.W.; Formal Analysis, K.M.K.W.; Investigation, K.M.K.W.; Resources, F.R.; Writing—Original Draft Preparation, K.M.K.W.; Writing—Review & Editing, F.R.; Visualization, K.M.K.W.; Supervision, F.R.; Project Administration, F.R.; Funding Acquisition, F.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Foundation (NSF) under grant no. DMR-1943085.

Data Availability Statement

Data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lee, J.G.; Hwang, J.; Hwang, H.J.; Jeon, O.S.; Jang, J.; Kwon, O.; Lee, Y.; Han, B.; Shul, Y.-G. A new family of perovskite catalysts for oxygen-evolution reaction in alkaline media: BaNiO3 and BaNi0. 83O2. 5. J. Am. Chem. Soc. 2016, 138, 3541–3547. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, J.; Hu, L.; Zhao, P.; Lee, L.Y.S.; Wong, K.-Y. Recent advances in electrocatalytic hydrogen evolution using nanoparticles. Chem. Rev. 2019, 120, 851–918. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, Y.; Suntivich, J.; May, K.J.; Perry, E.E.; Shao-Horn, Y. Synthesis and activities of rutile IrO2 and RuO2 nanoparticles for oxygen evolution in acid and alkaline solutions. J. Phys. Chem. Lett. 2012, 3, 399–404. [Google Scholar] [CrossRef] [PubMed]
  4. Tian, J.; Wu, W.; Tang, Z.; Wu, Y.; Burns, R.; Tichnell, B.; Liu, Z.; Chen, S. Oxygen reduction reaction and hydrogen evolution reaction catalyzed by Pd–Ru nanoparticles encapsulated in porous carbon nanosheets. Catalysts 2018, 8, 329. [Google Scholar] [CrossRef]
  5. Suntivich, J.; May, K.J.; Gasteiger, H.A.; Goodenough, J.B.; Shao-Horn, Y. A perovskite oxide optimized for oxygen evolution catalysis from molecular orbital principles. Science 2011, 334, 1383–1385. [Google Scholar] [CrossRef] [PubMed]
  6. Wickramaratne, K.M.; Ramezanipour, F. Impact of oxygen-vacancies on electrical conductivity and electrocatalytic activity of La3-xCaxFe2GaO9-δ (x= 0, 2; δ= 0, 1). Solid State Sci. 2023, 141, 107208. [Google Scholar] [CrossRef]
  7. Karki, S.B.; Hona, R.K.; Yu, M.; Ramezanipour, F. Enhancement of Electrocatalytic Activity as a Function of Structural Order in Perovskite Oxides. ACS Catal. 2022, 12, 10333–10337. [Google Scholar] [CrossRef]
  8. Hona, R.K.; Karki, S.B.; Cao, T.; Mishra, R.; Sterbinsky, G.E.; Ramezanipour, F. Sustainable oxide electrocatalyst for hydrogen-and oxygen-evolution reactions. ACS Catal. 2021, 11, 14605–14614. [Google Scholar] [CrossRef]
  9. Choi, M.-J.; Kim, T.L.; Kim, J.K.; Lee, T.H.; Lee, S.A.; Kim, C.; Hong, K.; Bark, C.W.; Ko, K.-T.; Jang, H.W. Enhanced Oxygen Evolution Electrocatalysis in Strained A-Site Cation Deficient LaNiO3 Perovskite Thin Films. Nano Lett. 2020, 20, 8040–8045. [Google Scholar] [CrossRef]
  10. Wang, J.; Gao, Y.; Chen, D.; Liu, J.; Zhang, Z.; Shao, Z.; Ciucci, F. Water splitting with an enhanced bifunctional double perovskite. ACS Catal. 2018, 8, 364–371. [Google Scholar] [CrossRef]
  11. Tian, S.; He, J.; Huang, H.; Song, T.-S.; Wu, X.; Xie, J.; Zhou, W. Perovskite-based multifunctional cathode with simultaneous supplementation of substrates and electrons for enhanced microbial electrosynthesis of organics. ACS Appl. Mater. Interfaces 2020, 12, 30449–30456. [Google Scholar] [CrossRef] [PubMed]
  12. Zhu, Y.; Zhou, W.; Zhong, Y.; Bu, Y.; Chen, X.; Zhong, Q.; Liu, M.; Shao, Z. A perovskite nanorod as bifunctional electrocatalyst for overall water splitting. Adv. Energy Mater. 2017, 7, 1602122. [Google Scholar] [CrossRef]
  13. Alom, M.S.; Ramezanipour, F. Layered Oxides SrLaFe1-xCoxO4-δ (x = 0–1) as Bifunctional Electrocatalysts for Water-Splitting. ChemCatChem 2021, 13, 3510–3516. [Google Scholar] [CrossRef]
  14. El Shinawi, H.; Greaves, C. Structural and Magnetic Characterisation of La1+xSr1–xCo0. 5M0. 5O4±δ (M = Cr, Mn). Z. Anorg. Allg. Chem. 2009, 635, 1856–1862. [Google Scholar] [CrossRef]
  15. El Shinawi, H.; Marco, J.; Berry, F.; Greaves, C. Synthesis and characterization of La0. 8Sr1. 2Co0. 5M0. 5O4−δ (M = Fe, Mn). J. Solid State Chem. 2009, 182, 2261–2268. [Google Scholar] [CrossRef]
  16. McCabe, E.; Greaves, C. Synthesis and Structural and Magnetic Characterization of Mixed Manganese−Copper n = 1 Ruddlesden−Popper Phases. Chem. Mater. 2006, 18, 5774–5781. [Google Scholar] [CrossRef]
  17. Burley, J.C.; Battle, P.D.; Gaskell, P.J.; Rosseinsky, M.J. Structural and magnetic chemistry of La2Sr2BMnO8 (B = Mg, Zn). J. Solid State Chem. 2002, 168, 202–207. [Google Scholar] [CrossRef]
  18. El Shinawi, H.; Greaves, C. Synthesis and characterization of La1. 5+xSr0.5−xCo0.5Ni0.5O4±δ (x = 0, 0.2). J. Mater. Chem. 2010, 20, 504–511. [Google Scholar] [CrossRef]
  19. Yin, W.-J.; Weng, B.; Ge, J.; Sun, Q.; Li, Z.; Yan, Y. Oxide perovskites, double perovskites and derivatives for electrocatalysis, photocatalysis, and photovoltaics. Energy Environ. Sci. 2019, 12, 442–462. [Google Scholar] [CrossRef]
  20. Zhao, Y.-N.; Liu, C.; Xu, S.; Min, S.; Wang, W.; Mitsuzaki, N.; Chen, Z. A/B-Site Management Strategy to Boost Electrocatalytic Overall Water Splitting on Perovskite Oxides in an Alkaline Medium. Inorg. Chem. 2023, 62, 12590–12599. [Google Scholar] [CrossRef]
  21. Wickramaratne, K.M.K.; Karki, S.B.; Ramezanipour, F. Electrocatalytic Properties of Oxygen-Deficient Perovskites Ca3Fe3-xMnxO8 (x = 1–2) for Hydrogen Evolution Reaction. Inorg. Chem. 2023, 62, 20961–20969. [Google Scholar] [CrossRef] [PubMed]
  22. Bhowmick, S.; Mohanta, M.K.; Qureshi, M. Transcription methodology for rationally designed morphological complex metal oxides: A versatile strategy for improved electrocatalysis. Sustain. Energy Fuels. 2021, 5, 6392–6405. [Google Scholar] [CrossRef]
  23. Bhowmick, S.; Dhankhar, A.; Sahu, T.K.; Jena, R.; Gogoi, D.; Peela, N.R.; Ardo, S.; Qureshi, M. Low overpotential and stable electrocatalytic oxygen evolution reaction utilizing doped perovskite oxide, La0.7Sr0.3MnO3, modified by cobalt phosphate. ACS Appl. Energy Mater. 2020, 3, 1279–1285. [Google Scholar] [CrossRef]
  24. Dupin, J.-C.; Gonbeau, D.; Benqlilou, H.; Vinatier, P.; Levasseur, A. XPS Analysis of New Lithium Cobalt Oxide Thin-films Before and After Lithium Deintercalation. Thin Solid Films 2001, 384, 23–32. [Google Scholar] [CrossRef]
  25. Davison, N.; McWhinnie, W.R.; Hooper, A. X-Ray Photoelectron Spectroscopic Study of Cobalt(II) and Nickel(II) Sorbed on Hectorite and Montmorillonite. Clays Clay Miner. 1991, 39, 22–27. [Google Scholar] [CrossRef]
  26. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X-ray Photoelectron Spectroscopy: A Reference Book of Standard Spectra for Identification and Interpretation of XPS Data; Chastain, J., Ed.; Perkin-Elmer: Eden Prairie, MN, USA, 1992. [Google Scholar]
  27. Zhao, Q.; Fang, C.; Tie, F.; Luo, W.; Peng, Y.; Huang, F.; Ku, Z.; Cheng, Y.-B. Regulating the Ni3+/Ni2+ ratio of NiOx by plasma treatment for fully vacuum-deposited perovskite solar cells. Mater. Sci. Semicond. Process. 2022, 148, 106839. [Google Scholar] [CrossRef]
  28. Fan, L.; Liu, P.F.; Yan, X.; Gu, L.; Yang, Z.Z.; Yang, H.G.; Qiu, S.; Yao, X. Atomically isolated nickel species anchored on graphitized carbon for efficient hydrogen evolution electrocatalysis. Nat. Commun. 2016, 7, 10667. [Google Scholar] [CrossRef]
  29. Cheng, M.; Fan, H.; Song, Y.; Cui, Y.; Wang, R. Interconnected hierarchical NiCo2O4 microspheres as high-performance electrode materials for supercapacitors. Dalton Trans. 2017, 46, 9201–9209. [Google Scholar] [CrossRef]
  30. Wang, P.; Liu, Z.; Han, C.; Ma, X.; Tong, Z.; Tan, B. Cu2O/CuO heterojunction formed by thermal oxidation and decorated with Pt co-catalyst as an efficient photocathode for photoelectrochemical water splitting. J. Nanoparticle Res. 2021, 23, 268. [Google Scholar] [CrossRef]
  31. Dan, Z.; Yang, Y.; Qin, F.; Wang, H.; Chang, H. Facile Fabrication of Cu2O Nanobelts in Ethanol on Nanoporous Cu and Their Photodegradation of Methyl Orange. Materials 2018, 11, 446. [Google Scholar] [CrossRef]
  32. Akgul, F.A.; Akgul, G.; Yildirim, N.; Unalan, H.E.; Turan, R. Influence of thermal annealing on microstructural, morphological, optical properties and surface electronic structure of copper oxide thin films. Mater. Chem. Phys. 2014, 147, 987–995. [Google Scholar] [CrossRef]
  33. Jin, Z.; Liu, C.; Qi, K.; Cui, X. Photo-reduced Cu/CuO nanoclusters on TiO2 nanotube arrays as highly efficient and reusable catalyst. Sci. Rep. 2017, 7, 39695. [Google Scholar] [CrossRef] [PubMed]
  34. Xu, D.; Fan, D.; Shen, W. Catalyst-free direct vapor-phase growth of Zn1−xCuxO micro-cross structures and their optical properties. Nanoscale Res. Lett. 2013, 8, 46. [Google Scholar] [CrossRef] [PubMed]
  35. Andoulsi, R.; Horchani-Naifer, K.; Ferid, M. Electrical conductivity of La1−xCaxFeO3−δ solid solutions. Ceram. Int. 2013, 39, 6527–6531. [Google Scholar] [CrossRef]
  36. Karki, S.B.; Ramezanipour, F. Pseudocapacitive Energy Storage and Electrocatalytic Hydrogen-Evolution Activity of Defect-Ordered Perovskites SrxCa3–xGaMn2O8 (x = 0 and 1). ACS Appl. Energy Mater. 2020, 3, 10983–10992. [Google Scholar] [CrossRef]
  37. Asenath-Smith, E.; Lokuhewa, I.N.; Misture, S.T.; Edwards, D.D. p-Type thermoelectric properties of the oxygen-deficient perovskite Ca2Fe2O5 in the brownmillerite structure. J. Solid State Chem. 2010, 183, 1670–1677. [Google Scholar] [CrossRef]
  38. Kozhevnikov, V.; Leonidov, I.; Mitberg, E.; Patrakeev, M.; Petrov, A.; Poeppelmeier, K. Conductivity and carrier traps in La1− xSrxCo1−zMnzO3−δ (x = 0.3; z = 0 and 0.25). J. Solid State Chem. 2003, 172, 296–304. [Google Scholar] [CrossRef]
  39. Sudha, L.; Sukumar, R.; Uma Rao, K. Evaluation of activation energy (Ea) profiles of nanostructured alumina polycarbonate composite insulation materials. Int. J. Mater. Mech. Manuf. 2014, 2, 96–100. [Google Scholar]
  40. Alom, M.S.; Kananke-Gamage, C.C.; Ramezanipour, F. Perovskite Oxides as Electrocatalysts for Hydrogen Evolution Reaction. ACS Omega 2022, 7, 7444–7451. [Google Scholar] [CrossRef]
  41. Si, C.; Zhang, W.; Lu, Q.; Guo, E.; Yang, Z.; Chen, J.; He, X.; Luo, J. Recent Advances in Perovskite Catalysts for Efficient Overall Water Splitting. Catalysts 2022, 12, 601. [Google Scholar] [CrossRef]
  42. Strmcnik, D.; Uchimura, M.; Wang, C.; Subbaraman, R.; Danilovic, N.; Van Der Vliet, D.; Paulikas, A.P.; Stamenkovic, V.R.; Markovic, N.M. Improving the hydrogen oxidation reaction rate by promotion of hydroxyl adsorption. Nat. Chem. 2013, 5, 300–306. [Google Scholar] [CrossRef] [PubMed]
  43. Khan, R.; Mehran, M.T.; Naqvi, S.R.; Khoja, A.H.; Mahmood, K.; Shahzad, F.; Hussain, S. Role of perovskites as a bi-functional catalyst for electrochemical water splitting: A review. Int. J. Energy Res. 2020, 44, 9714–9747. [Google Scholar] [CrossRef]
  44. Wang, S.; Lu, A.; Zhong, C.-J. Hydrogen production from water electrolysis: Role of catalysts. Nano Converg. 2021, 8, 4. [Google Scholar] [CrossRef]
  45. Wang, H.; Zhou, H.; Zhang, W.; Yao, S. Urea-assisted synthesis of amorphous molybdenum sulfide on P-doped carbon nanotubes for enhanced hydrogen evolution. J. Mater. Sci. Mater. 2018, 53, 8951–8962. [Google Scholar] [CrossRef]
  46. Han, H.; Nayak, A.K.; Choi, H.; Ali, G.; Kwon, J.; Choi, S.; Paik, U.; Song, T. Partial dehydration in hydrated tungsten oxide nanoplates leads to excellent and robust bifunctional oxygen reduction and hydrogen evolution reactions in acidic media. ACS Sustain. Chem. Eng. 2020, 8, 9507–9518. [Google Scholar] [CrossRef]
  47. Karki, S.B.; Andriotis, A.N.; Menon, M.; Ramezanipour, F. Bifunctional Water-Splitting Electrocatalysis Achieved by Defect Order in LaA2Fe3O8 (A = Ca, Sr). ACS Appl. Energy Mater. 2021, 4, 12063–12066. [Google Scholar] [CrossRef]
  48. Galal, A.; Hassan, H.K.; Atta, N.F.; Jacob, T. An Efficient and Durable Electrocatalyst for Hydrogen Production Based on Earth-Abundant Oxide-Graphene Composite. ChemistrySelect 2017, 2, 10261–10270. [Google Scholar] [CrossRef]
  49. Kananke-Gamage, C.C.; Ramezanipour, F. Variation of the electrocatalytic activity of isostructural oxides Sr2LaFeMnO7 and Sr2LaCoMnO7 for hydrogen and oxygen-evolution reactions. Dalton Trans. 2021, 50, 14196–14206. [Google Scholar] [CrossRef]
  50. Murthy, A.P.; Theerthagiri, J.; Madhavan, J. Insights on Tafel constant in the analysis of hydrogen evolution reaction. J. Phys. Chem. C 2018, 122, 23943–23949. [Google Scholar] [CrossRef]
  51. Shinagawa, T.; Garcia-Esparza, A.T.; Takanabe, K. Insight on Tafel slopes from a microkinetic analysis of aqueous electrocatalysis for energy conversion. Sci. Rep. 2015, 5, 13801. [Google Scholar] [CrossRef]
  52. Ling, C.; Ouyang, Y.; Shi, L.; Yuan, S.; Chen, Q.; Wang, J. Template-grown MoS2 nanowires catalyze the hydrogen evolution reaction: Ultralow kinetic barriers with high active site density. ACS Catal. 2017, 7, 5097–5102. [Google Scholar] [CrossRef]
  53. Hona, R.K.; Ramezanipour, F. Remarkable Oxygen-Evolution Activity of a Perovskite Oxide from the Ca2−xSrxFe2O6−δ Series. Angew. Chem. 2019, 131, 2082–2085. [Google Scholar] [CrossRef]
  54. Pan, Y.; Chen, Y.; Li, X.; Liu, Y.; Liu, C. Nanostructured nickel sulfides: Phase evolution, characterization and electrocatalytic properties for the hydrogen evolution reaction. RSC Adv. 2015, 5, 104740–104749. [Google Scholar] [CrossRef]
  55. Park, H.W.; Lee, D.U.; Park, M.G.; Ahmed, R.; Seo, M.H.; Nazar, L.F.; Chen, Z. Perovskite–nitrogen-doped carbon nanotube composite as bifunctional catalysts for rechargeable lithium–air batteries. ChemSusChem 2015, 8, 1058–1065. [Google Scholar] [CrossRef] [PubMed]
  56. Oh, M.Y.; Jeon, J.S.; Lee, J.J.; Kim, P.; Nahm, K.S. The bifunctional electrocatalytic activity of perovskite La0.6Sr0.4CoO3−δ for oxygen reduction and evolution reactions. RSC Adv. 2015, 5, 19190–19198. [Google Scholar] [CrossRef]
  57. Moritomo, Y.; Higashi, K.; Matsuda, K.; Nakamura, A. Spin-state transition in layered perovskite cobalt oxides: La2−xSrxCoO4 (0.4≤ x ≤ 1.0). Phys. Rev. B 1997, 55, R14725. [Google Scholar] [CrossRef]
  58. Alom, M.S.; Ramezanipour, F. Vacancy effect on the electrocatalytic activity of LaMn1/2Co1/2O3−δ for hydrogen and oxygen evolution reactions. Chem. Commun. 2023, 59, 5870–5873. [Google Scholar] [CrossRef] [PubMed]
  59. Wei, C.; Sun, Y.; Scherer, G.N.G.; Fisher, A.C.; Sherburne, M.; Ager, J.W.; Xu, Z.J. Surface composition dependent ligand effect in tuning the activity of nickel–copper bimetallic electrocatalysts toward hydrogen evolution in alkaline. J. Am. Chem. Soc. 2020, 142, 7765–7775. [Google Scholar] [CrossRef] [PubMed]
  60. Rajput, A.; Kundu, A.; Chakraborty, B. Recent Progress on Copper-Based Electrode Materials for Overall Water-Splitting. ChemElectroChem 2021, 8, 1698–1722. [Google Scholar] [CrossRef]
  61. Larson, A.C.; Von Dreele, R.B. General Structure Analysis System (GSAS); Los Alamos National Laboratory Report LAUR; Los Alamos National Laboratory: Los Alamos, NM, USA, 2004; pp. 86–748.
  62. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  63. Hona, R.K.; Karki, S.B.; Ramezanipour, F. Oxide Electrocatalysts Based on Earth-Abundant Metals for Both Hydrogen-and Oxygen-Evolution Reactions. ACS Sustain. Chem. Eng. 2020, 8, 11549–11557. [Google Scholar] [CrossRef]
  64. Li, Q.; Thangadurai, V. A comparative 2 and 4-probe DC and 2-probe AC electrical conductivity of novel co-doped Ce0.9−xRExMo0.1O2.1–0.5x (RE = Y, Sm, Gd; x = 0.2, 0.3). J. Mater. Chem. 2010, 20, 7970–7983. [Google Scholar] [CrossRef]
  65. Niu, S.; Li, S.; Du, Y.; Han, X.; Xu, P. How to reliably report the overpotential of an electrocatalyst. ACS Energy Lett. 2020, 5, 1083–1087. [Google Scholar] [CrossRef]
Figure 1. Rietveld refinement profiles using powder X-ray diffraction data of (a) LaSrMn0.5Co0.5O4, (b) LaSrMn0.5Ni0.5O4, (c) LaSrMn0.5Cu0.5O4-δ, and (d) LaSrMn0.5Zn0.5O4. The cross symbols, solid red line, olive vertical tick marks, and lower magenta line correspond to experimental data, the calculated pattern for the structural model, Bragg peak positions, and the difference plot, respectively.
Figure 1. Rietveld refinement profiles using powder X-ray diffraction data of (a) LaSrMn0.5Co0.5O4, (b) LaSrMn0.5Ni0.5O4, (c) LaSrMn0.5Cu0.5O4-δ, and (d) LaSrMn0.5Zn0.5O4. The cross symbols, solid red line, olive vertical tick marks, and lower magenta line correspond to experimental data, the calculated pattern for the structural model, Bragg peak positions, and the difference plot, respectively.
Molecules 29 03107 g001
Figure 2. Crystal structure of LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, Zn). Pink spheres denote La/Sr, small red spheres indicate oxygen, and green spheres represent Mn/M.
Figure 2. Crystal structure of LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, Zn). Pink spheres denote La/Sr, small red spheres indicate oxygen, and green spheres represent Mn/M.
Molecules 29 03107 g002
Figure 3. Scanning electron microscopy (SEM) images of (a) LaSrMn0.5Co0.5O4, (b) LaSrMn0.5Ni0.5O4, (c) LaSrMn0.5Cu0.5O4-δ, and (d) LaSrMn0.5Zn0.5O4.
Figure 3. Scanning electron microscopy (SEM) images of (a) LaSrMn0.5Co0.5O4, (b) LaSrMn0.5Ni0.5O4, (c) LaSrMn0.5Cu0.5O4-δ, and (d) LaSrMn0.5Zn0.5O4.
Molecules 29 03107 g003
Figure 4. XPS data showing the manganese spectra for all four materials. The purple dashed line is drawn to show that the binding energy for the 2p3/2 peak of LaSrMn0.5Zn0.5O4 is shifted to higher energy compared to those of the other three materials.
Figure 4. XPS data showing the manganese spectra for all four materials. The purple dashed line is drawn to show that the binding energy for the 2p3/2 peak of LaSrMn0.5Zn0.5O4 is shifted to higher energy compared to those of the other three materials.
Molecules 29 03107 g004
Figure 5. XPS data showing the spectra for (a) cobalt in LaSrMn0.5Co0.5O4, (b) nickel in LaSrMn0.5Ni0.5O4, (c) copper in LaSrMn0.5Cu0.5O4–δ, and (d) zinc in LaSrMn0.5Zn0.5O4. The peak marked by * is the lanthanum 3d3/2 peak, which appears very close to the nickel 2p3/2.
Figure 5. XPS data showing the spectra for (a) cobalt in LaSrMn0.5Co0.5O4, (b) nickel in LaSrMn0.5Ni0.5O4, (c) copper in LaSrMn0.5Cu0.5O4–δ, and (d) zinc in LaSrMn0.5Zn0.5O4. The peak marked by * is the lanthanum 3d3/2 peak, which appears very close to the nickel 2p3/2.
Molecules 29 03107 g005
Figure 6. (a) Electrical conductivities as a function of temperature. (b) Arrhenius plots to determine the activation energies (Ea) for the temperature-activated increase in conductivity.
Figure 6. (a) Electrical conductivities as a function of temperature. (b) Arrhenius plots to determine the activation energies (Ea) for the temperature-activated increase in conductivity.
Molecules 29 03107 g006
Figure 7. (a) Polarization curves for HER in 0.5 M H2SO4. The inset shows chronopotentiometry data for the best-performing material, LaSrMn0.5Co0.5O4. (b) Tafel plots and slopes. (c) X-ray diffraction data for LaSrMn0.5Co0.5O4 before and after 100 cycles of HER in 0.5 M H2SO4. (d) HER mass activities at different overpotentials in 0.5 M H2SO4.
Figure 7. (a) Polarization curves for HER in 0.5 M H2SO4. The inset shows chronopotentiometry data for the best-performing material, LaSrMn0.5Co0.5O4. (b) Tafel plots and slopes. (c) X-ray diffraction data for LaSrMn0.5Co0.5O4 before and after 100 cycles of HER in 0.5 M H2SO4. (d) HER mass activities at different overpotentials in 0.5 M H2SO4.
Molecules 29 03107 g007
Figure 8. Plots of javg against ν in 0.5 M H2SO4. The slopes indicate double-layer capacitance, Cdl.
Figure 8. Plots of javg against ν in 0.5 M H2SO4. The slopes indicate double-layer capacitance, Cdl.
Molecules 29 03107 g008
Figure 9. (a) HER polarization curves in 1 M KOH. The inset shows chronopotentiometry data for the best-performing material LaSrMn0.5Co0.5O4. (b) Tafel plots and slopes. (c) X-ray diffraction data for LaSrMn0.5Co0.5O4 before and after 100 cycles of HER in 1 M KOH. (d) HER mass activities at different overpotentials in 1 M KOH.
Figure 9. (a) HER polarization curves in 1 M KOH. The inset shows chronopotentiometry data for the best-performing material LaSrMn0.5Co0.5O4. (b) Tafel plots and slopes. (c) X-ray diffraction data for LaSrMn0.5Co0.5O4 before and after 100 cycles of HER in 1 M KOH. (d) HER mass activities at different overpotentials in 1 M KOH.
Molecules 29 03107 g009
Figure 10. Plots of javg against ν in 1 M KOH. The slopes indicate double-layer capacitance, Cdl.
Figure 10. Plots of javg against ν in 1 M KOH. The slopes indicate double-layer capacitance, Cdl.
Molecules 29 03107 g010
Figure 11. (a) Polarization curves for OER in 1 M KOH. The inset shows chronopotentiometry data for the best-performing material, LaSrMn0.5Co0.5O4. (b) Tafel plots and slopes. (c) X-ray diffraction data for LaSrMn0.5Co0.5O4 before and after 100 cycles of OER in 1 M KOH. (d) OER mass activities at different overpotentials in 1 M KOH.
Figure 11. (a) Polarization curves for OER in 1 M KOH. The inset shows chronopotentiometry data for the best-performing material, LaSrMn0.5Co0.5O4. (b) Tafel plots and slopes. (c) X-ray diffraction data for LaSrMn0.5Co0.5O4 before and after 100 cycles of OER in 1 M KOH. (d) OER mass activities at different overpotentials in 1 M KOH.
Molecules 29 03107 g011
Figure 12. The TEM data (a) before and (b) after chronopotentiometry experiment under OER conditions.
Figure 12. The TEM data (a) before and (b) after chronopotentiometry experiment under OER conditions.
Molecules 29 03107 g012
Table 1. Refined structural parameters for LaSrMn0.5Co0.5O4 at room temperature using powder X-ray diffraction data. Space group: I4/mmm, a = 3.8435 (1) Å, c = 12.5933 (2) Å, Rp = 0.0392, wRp = 0.0504.
Table 1. Refined structural parameters for LaSrMn0.5Co0.5O4 at room temperature using powder X-ray diffraction data. Space group: I4/mmm, a = 3.8435 (1) Å, c = 12.5933 (2) Å, Rp = 0.0392, wRp = 0.0504.
AtomxyzOccupancyMultiplicityUiso2)
La000.35985 (8)0.540.0192 (5)
Sr000.35985 (8)0.540.0192 (5)
Mn0000.520.0131 (9)
Co0000.520.0131 (9)
O10.500140.031 (2)
O2000.1642 (6)140.038 (2)
Table 2. Room temperature electrical conductivity and activation energies.
Table 2. Room temperature electrical conductivity and activation energies.
Electrical Conductivity (S/cm)Activation Energy (eV)
LaSrMn0.5Co0.5O40.00260.2673 (2)
LaSrMn0.5Ni0.5O40.00140.2746 (2)
LaSrMn0.5Cu0.5O4-δ8.5961 × 10−40.2676 (3)
LaSrMn0.5Zn0.5O47.0755 × 10−40.4727 (7)
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

Wickramaratne, K.M.K.; Ramezanipour, F. Electrocatalytic Properties of Quasi-2D Oxides LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, and Zn) for Hydrogen and Oxygen Evolution Reactions. Molecules 2024, 29, 3107. https://doi.org/10.3390/molecules29133107

AMA Style

Wickramaratne KMK, Ramezanipour F. Electrocatalytic Properties of Quasi-2D Oxides LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, and Zn) for Hydrogen and Oxygen Evolution Reactions. Molecules. 2024; 29(13):3107. https://doi.org/10.3390/molecules29133107

Chicago/Turabian Style

Wickramaratne, Kinithi M. K., and Farshid Ramezanipour. 2024. "Electrocatalytic Properties of Quasi-2D Oxides LaSrMn0.5M0.5O4 (M = Co, Ni, Cu, and Zn) for Hydrogen and Oxygen Evolution Reactions" Molecules 29, no. 13: 3107. https://doi.org/10.3390/molecules29133107

Article Metrics

Back to TopTop