The development of water electrolysis catalysts that accelerate the oxygen evolution reaction (OER) is a crucial challenge. Ni-based oxides are promising OER catalysts; however, quantitative studies of Ni-based oxides remain unexplored. In this study, we quantitatively evaluated the OER activity of LiNi0.5Mn1.5O4 as a thin-film electrode catalyst. The LiNi0.5Mn1.5O4 thin film fabricated using a sputtering method exhibited a current density of 6.6 and ∼2.6 mAcm−2 for geometric and estimated areas, respectively, at 1.78 V vs. a reversible hydrogen electrode. X-ray photoelectron spectroscopy indicated the presence of Ni3+ in the as-grown and post-OER LiNi0.5Mn1.5O4 thin films. These results suggest that Ni3+ plays a key role in the OER of LiNi0.5Mn1.5O4.
I. INTRODUCTION
Water electrolysis converts water into hydrogen and oxygen.1–4 Ir and Ru oxides have attracted considerable attention as catalysts for oxygen evolution reactions (OER) because of their high catalytic activity and durability.5–9 Because these noble metals are rare and expensive, it is highly desirable to develop alternatives composed of abundant 3d transition metals, such as Ni, Fe, and Co.10–13 However, 3d transition metal oxide catalysts exhibit lower catalytic activity, that is, higher overpotentials, than Ir and Ru oxide catalysts.14,15 Consequently, elucidating the OER mechanism is necessary to lower the overpotentials of 3d transition metal oxide catalysts.
Oxides containing Ni3+ exhibit high OER catalytic activity, as demonstrated in NiOOH,16 NixFe1−xOOH,17–19 Ni3−xCoxO4,20,21 and LaNiO3.22–24 In contrast, oxides containing Ni2+, such as Ni(OH)2, NiO, and NiFe2O4, exhibit lower OER activity than Ni3+-containing oxides. In Ni2+-containing oxides, Ni3+ is suggested to contribute during the catalytic cycle.25 Therefore, understanding the chemical state of Ni is critical for developing Ni-oxide catalysts with high OER activity.
LiNi0.5Mn1.5O4 (LNMO) is an intriguing compound that exhibits OER catalytic activity despite the formal charge of 2+ for Ni26 and 4+ for Mn.27,28 The oxidation state of Ni can be tuned from 2+ to 4+ by the insertion/removal of Li, indicating the potential for enhanced OER activity. To date, the OER catalytic activity of LNMO has only been studied in powder form, leading to less accurate current density values. The difficulties in quantitative studies using powders are as follows: (1) determining the separate contributions of various additives to the reaction is difficult, (2) the potential might be underestimated by the voltage drop owing to the low electrical conductivity of LNMO (∼10−6 Scm−1)29 and the grain boundaries, and (3) measuring the surface area of the powder is difficult. Thin-film electrodes are suitable platforms for quantitative evaluations. LNMO thin films fabricated on a flat current collector have the following advantages: (1) only the catalyst surface is exposed, (2) the voltage drop is minimized by uniformly applying a potential from the current collector, and (3) the surface area can be defined if the surface roughness is small. However, there are no reports on the quantitative evaluation of OER activity using LNMO thin-film electrodes.
In this paper, we report the quantitative evaluation of the OER using LNMO thin-film electrodes. LNMO thin films deposited on an Au current collector exhibit a current density of 6.6 and ∼2.6 mA cm−2 for geometric and roughness-modeled areas, respectively, at 1.78 V vs. a reversible hydrogen electrode (VRHE). The current density using the estimated area is comparable to that of the previously reported value for the powder form.27 X-ray photoelectron spectroscopy (XPS) revealed the presence of Ni3+ and Mn2+ in both the as-grown and post-OER LNMO thin films. These results suggest the importance of Ni3+ originating from the electron transfer between Ni2+ and Mn4+.
II. EXPERIMENTAL
The LNMO thin-film electrode catalyst was fabricated using the following process. A direct-current magnetron sputtering method was employed to deposit a current collector (30 nm Ti layer and 100 nm Au layer) on an Al2O3(0001) substrate at room temperature. Using a radio-frequency (RF) magnetron sputtering method, a 120-nm thick LNMO thin film was deposited on a current collector using radio-frequency magnetron sputtering. A target with an excess Li composition of Li1.2Ni0.5Mn1.5O4 (diameter: 48 mm, Toshima Manufacturing Co., Ltd.) was used to compensate for the Li deficiency.30,31 During sputtering, the substrate temperature, RF power, and deposition time were set at 500 °C, 100 W, and 60 min, respectively. The total pressure and total flow rate during RF sputtering were fixed at 1.0 Pa and 20.0 SCCM, respectively. Ar and O2 gases were introduced at concentrations of 98% and 2.0%, respectively.
The structural and compositional characterization of the fabricated films was performed using x-ray diffraction [XRD, Bruker, D8 DISCOVER, Cu Kα rays (wavelength: 0.15406 nm)], Raman spectroscopy (HORIBA, XploRA PLUS, wavelength: 532 nm), scanning electron microscopy (SEM, Hitachi High-Tech, S-5500), and energy-dispersive x-ray analysis (EDX, Bruker AXS, XFlash® 5010). The electronic states of the thin-film surfaces were investigated using XPS (ULVAC-PHI, VersaProbeIII PHI5000). The water droplets on the LNMO thin-film electrode retrieved from the electrolyte were removed using an air duster to perform XPS. The LNMO thin film was then dried overnight (∼8 h) in a vacuum desiccator. The samples were subsequently transferred to the XPS apparatus in ambient air.
The electrochemical measurements were performed using a three-electrode electrochemical cell. The electrolyte used was a 0.1M KOH aqueous solution prepared by adding KOH (Guaranteed Reagent, FUJIFILM Wako Pure Chemical Corp.) to ultrapure water (>19.0 MΩ cm, Merck Millipore, Direct-Q UV3). Fe impurity removal was not performed. Fe was not detected by XPS on the surface after OER (Fig. S1 in the supplementary material). We believe that the effect of impurities below a detection sensitivity (∼1%) is negligible. The LNMO thin films on a current collector, Pt coil, and Ag/AgCl were used as the working, counter, and reference electrodes, respectively. Linear sweep voltammetry (LSV), cyclic voltammetry, chronopotentiometry, and electrochemical impedance spectroscopy (EIS) were performed. The geometric surface area of the working electrode was determined through image processing using ImageJ software.32 The quantification of oxygen evolution through chronopotentiometry was conducted using gas chromatography (Shimadzu Corp., GC-8A). Tafel analysis was performed based on the results of the LSV. EIS measurements were performed with a sinusoidal voltage variation of 20 mV over the frequency range of 100 mHz to 10 MHz at 1.70, 1.75, 1.80, and 2.00 VRHE. The potential drop during measurement was compensated by an automatic function in the instrument (Biologic SP-50) during EIS measurements. A sinusoidal voltage of 100 kHz and 10 mV amplitude was used to compensate for a feedback ratio of 85%. The results were analyzed using pyZwx software.33 Initially, the two-pole logarithmic Hilbert transform method34 was used to exclude measurement points that did not meet the causality and stability criteria. Subsequently, fitting was performed using the equivalent circuit models, as discussed later. Additionally, the uncompensated resistance of the electrochemical cell was corrected by the EIS results.
III. RESULTS AND DISCUSSION
First, we characterize the structures of the fabricated films [Fig. 1(a)]. The out-of-plane XRD patterns in Fig. 1(b) show the diffraction from the spinel-type LNMO and the diffraction from the Au current collector. The two-dimensional XRD pattern of LNMO is elongated in the ψ direction, indicating that the film is a weakly (111)-oriented LNMO polycrystalline thin film. The Raman spectrum also confirms the formation of LNMO [Fig. 1(c)], where the peaks at ∼494 and ∼610 cm−1 correspond to the F2g and A1g modes of LNMO, respectively.35 The significant background observed in the Raman spectrum was due to the reflection of the excited laser to the Au current collector underneath. These results confirm the formation of a weakly oriented polycrystalline LNMO thin film on the Au current collector.
Characterization of LiNi0.5Mn1.5O4 (LNMO) thin films. (a) Cross-sectional schematic of thin-film electrode. (b) X-ray diffraction patterns. Upper: two-dimensional (2D) XRD patterns, lower: 2θ plot in out-of-plane direction after integration from 2D-XRD pattern. The Ti peaks are not visible in the figure because they overlap with Au. (c) Raman spectrum of LNMO thin film. Purple dotted curve is the Raman spectrum of bulk reported in Ref. 35. (d) Scanning electron microscopy (SEM) image (upper left) and energy-dispersive x-ray (EDX) mapping of LNMO thin film. O (blue, upper right), Mn (red, lower left), and Ni (green, lower right) distribute uniformly in the film.
Characterization of LiNi0.5Mn1.5O4 (LNMO) thin films. (a) Cross-sectional schematic of thin-film electrode. (b) X-ray diffraction patterns. Upper: two-dimensional (2D) XRD patterns, lower: 2θ plot in out-of-plane direction after integration from 2D-XRD pattern. The Ti peaks are not visible in the figure because they overlap with Au. (c) Raman spectrum of LNMO thin film. Purple dotted curve is the Raman spectrum of bulk reported in Ref. 35. (d) Scanning electron microscopy (SEM) image (upper left) and energy-dispersive x-ray (EDX) mapping of LNMO thin film. O (blue, upper right), Mn (red, lower left), and Ni (green, lower right) distribute uniformly in the film.
SEM and EDX measurements confirm the uniformity of the crystal grain structure and composition of the thin film [Fig. 1(d)]. The SEM image shows crystal grains approximately 50–100 nm in size. Additionally, the EDX mapping of each element shows a homogeneous distribution of Ni, Mn, and O. Table S1 in the supplementary material presents the results of elemental composition measurements using EDX. The Mn/Ni ratio shows a value close to stoichiometry, which is approximately 3.5. The signal of Al originating from the substrate is also detected. This detection makes it challenging to quantify the O content within the LNMO films. We note that Li is not detectable by EDX.
The LSV curve of the device structure [Fig. 2(a)] shows the increase in the current density at ∼1.6 VRHE, which reaches 6.60 mA cm−2 at 1.78 VRHE; here, the reaction area is a geometric area estimated from a photograph [Fig. 2(a)]. Considering the increase in real area due to surface roughness, the current density may decrease to 0.4 times. We establish two types of surface three-dimensional models to estimate the real area: a rectangular prism and a square-pyramid-shaped model [see Figs. S2(a) and S2(b) in the supplementary material]. The arithmetic mean roughness (Ra) was used as the evaluation scale for surface roughness. The Ra of the LNMO thin film was 35 nm through atomic force microscopy measurements. Hence, the roughness heights were set to match Ra. Compared to the directly measured simple geometric surface areas in this study, the areas in Figs. S1(a) and S1(b) in the supplementary material are 2.4 and 2.7 times larger, respectively. As a result, the estimated current density is ∼2.6 mA cm−2. In such a scenario, the oxygen evolution current density of the LNMO thin film electrode is comparable to that of the octahedral LNMO powders reported previously, 2.9227 and 0.1428 mAcm−2 at 1.78 VRHE. The turnover frequency (TOF) was calculated to be 2.7 × 10 s−1. From the electrical current in the vicinity of 1.3 VRHE (corresponding to the change from Ni2+ to Ni3+), the number of electrons transferred is ∼1.6 × 1013, which corresponds to the number of active sites.36 From the current density (6.6 mA cm-2) at 1.78 VRHE and electrode area (0.10 cm2), we obtain TOF = 2.7 × 10 s−1. We note that the use of Ag/AgCl in 0.1M KOH may induce an electrode liquid junction potential of approximately 30 mV.37,38
Electrochemical oxygen evolution reaction (OER) measurements of polycrystalline LiNi0.5Mn1.5O4 (LNMO) thin films. (a) Left: schematic of electrode structure used in this study. Right: photograph of electrode for OER setup. (b) Linear sweep voltammogram. Blue and green dashed curves are reproduced with permission from Maiyalagan et al., ACS Catal. 4, 421–425 (2014). Copyright 2014 American Chemical Society and Ren et al., ACS Appl. Energy Mater. 4, 10731–10738 (2021). Copyright 2021 American Chemical Society, respectively. Inset shows the Tafel plot for LNMO thin film electrode. (c) Chronopotentiometry for LNMO thin film electrode for 90 h.
Electrochemical oxygen evolution reaction (OER) measurements of polycrystalline LiNi0.5Mn1.5O4 (LNMO) thin films. (a) Left: schematic of electrode structure used in this study. Right: photograph of electrode for OER setup. (b) Linear sweep voltammogram. Blue and green dashed curves are reproduced with permission from Maiyalagan et al., ACS Catal. 4, 421–425 (2014). Copyright 2014 American Chemical Society and Ren et al., ACS Appl. Energy Mater. 4, 10731–10738 (2021). Copyright 2021 American Chemical Society, respectively. Inset shows the Tafel plot for LNMO thin film electrode. (c) Chronopotentiometry for LNMO thin film electrode for 90 h.
Subsequently, we investigated the rate-determining step of the OER from the Tafel slope39 determined from the current densities ranging from 0.5 to 10 mA cm−2. We obtained a Tafel slope of 69.8 mV dec−1 [Fig. 2(b), inset], which is in good agreement with the previously reported value for octahedral LNMO powders comprising (111) facets (70 mV dec−1).27 Theoretically, the Tafel slope of ∼70 mV dec−1 suggests that the conversion of adsorbed OH species to O2− is the rate-limiting step of OER.39 Accordingly, OH species adsorbed on the LNMO thin film may slowly convert to O2−.39 It should be noted that when the current values are large >100 mAcm−2, steady-state measurements are essential. However, since the current values in this study are small (0.5–5 mA cm−2), we believe that the error of Tafel analysis noted in the literature40 is negligible. As for the potential used to calculate the Tafel gradient, it exceeds the equilibrium potential by more than 120 mV.41 In the present study, we conclude that the error is negligible because it does not fall into the critical region.
Chronopotentiometry was performed to evaluate the durability of the LNMO catalysts [Fig. 2(c)]. The observed potential increase over 90 h was 0.18 V, indicating sufficient durability of the catalyst. Furthermore, the stable evolution of oxygen during chronopotentiometry is confirmed by gas chromatography (Fig. S3 in the supplementary material) and Faraday efficiency is ∼90%.
The crystal structure of LNMO remains unchanged before and after the OER. The changes in the peak structure of the XRD and Raman spectroscopy [Figs. 3(a) and 3(b), respectively] are absent before and after the OER, showing that LNMO maintains its spinel-type structure during and after the OER with no degradation or dissolution. We speculate that the decrease in the 111 peak intensity is due to Li desorption from LNMO in the OER process. The Li 1s XPS shows the decrease in Li intensity after OER (Fig. S4 in the supplementary material), indicating that Li desorbed from the surface. We believe that the crystallinity has decreased and the diffraction intensity has also decreased because of the Li desorption. SEM images [Fig. 3(c)] also indicate no change in the LNMO electrode surface before and after the OER. Therefore, the structure of LNMO films remains intact during the OER, and the electrochemical measurements in this study reflect the intrinsic activity of LNMO.
Characterization of LiNi0.5Mn1.5O4 (LNMO) thin-film structures before and after the oxygen evolution reaction (OER) measurements. (a) X-ray diffraction patterns, (b) Raman spectra, and (c) scanning electron microscopy images of LNMO films. Note: the purple dotted line in (b) is reproduced with permission from Wang et al., Solid State Ionics 193, 32–38 (2011). Copyright 2011 Elsevier B. V.
Characterization of LiNi0.5Mn1.5O4 (LNMO) thin-film structures before and after the oxygen evolution reaction (OER) measurements. (a) X-ray diffraction patterns, (b) Raman spectra, and (c) scanning electron microscopy images of LNMO films. Note: the purple dotted line in (b) is reproduced with permission from Wang et al., Solid State Ionics 193, 32–38 (2011). Copyright 2011 Elsevier B. V.
To investigate the surface electronic states of the OER, we performed XPS (Fig. 4). The O 1s spectrum consists of two peaks at 529.4 and 531.0 eV [Fig. 4(a)], attributed to lattice oxygen and chemisorbed water on the surface, respectively.42 After the OER, an increase in the peak corresponding to chemisorbed water suggests an increase in the amount of adsorbed water molecules on the LNMO surface after the immersion of the samples in the KOH aqueous solution.
X-ray photoelectron spectra of LiNi0.5Mn1.5O4 (LNMO) thin films before and after the oxygen evolution reaction (OER) measurements. (a) O 1s, (b) Mn 2p, and (c) Ni 2p 3/2.
X-ray photoelectron spectra of LiNi0.5Mn1.5O4 (LNMO) thin films before and after the oxygen evolution reaction (OER) measurements. (a) O 1s, (b) Mn 2p, and (c) Ni 2p 3/2.
The Mn 2p spectra [Fig. 4(b)] indicate negligible changes before and after the OER measurements. The XPS spectra exhibit two peaks at 653.5 and 641.8 eV, corresponding to Mn 2p 1/2 and Mn 2p 3/2, respectively,43 along with a broad tail centered at ∼636.4 eV. In particular, the broad tail at ∼636 eV is lower than the binding energy observed for Mn3+,44 suggesting a lower valence state of Mn2+.45
The Ni 2p 3/2 spectra [Fig. 4(c)] comprise three peaks at 861.2, 856.0, and 854.8 eV, assigned to the satellites from the shake-up process, Ni3+, and Ni2+, respectively.26 Surprisingly, we confirm the presence of a formal charge of Ni3+ even in the as-grown LNMO. Similar to the XPS spectra of Mn, no changes were observed in the peak positions or relative peak areas of Ni before and after the OER measurements. Therefore, the presence of Ni3+ on the LNMO surface may have contributed to the observed OER activity. The origins of Ni3+ states remain unclear. It is essential to determine the valence states within thin films. Given that XPS is surface-sensitive, a future challenge lies in evaluating the interior using hard XPS or x-ray absorption spectroscopy, which have larger probing depths.
We estimated the proportions of Mn2+ and Ni3+ using the ratio of the peak areas (I) in the XPS spectra. Here, we denote the fraction of Mn2+ as x (the fraction of Mn4+ corresponds to 1 − x). The value of x/(1 − x) is approximately equal to IMn2+/IMn4+ ∼ 0.06, yielding x ∼ 0.057. Here, IMn2+ and IMn4+ are the areas of the fitted Gaussian functions at around 636.4 and 641.8 eV, respectively. This indicates that ∼5.7% of the original Mn4+ was converted to Mn2+ and that 0.16 electrons were accepted in Mn per chemical formula unit of LNMO. The fraction of Ni3+ (y) was similarly evaluated. The value of y/(1 − y) was approximately equal to INi3+/INi2+ ∼ 0.56, yielding y ∼ 0.36. Here, INi3+ and INi2+ are the areas of the fitted Gaussian functions at around 855.5 and 854.4 eV, respectively. These evaluations suggest that 0.17 electrons are donated from Ni per chemical formula unit of LNMO. The value of donated electrons from Mn (0.16 e−) is in good agreement with the accepted electrons in Ni (0.17 e−), suggesting an electron transfer from Mn and Ni in the LNMO thin films.
The binding energy shift observed in LNMO is about 0.3 eV. Adsorption of K is suspected as the reason for this shift. Indeed, when LNMO is immersed in an electrolyte solution, K in the electrolyte is detected on its surface (Fig. S5 in the supplementary material). Electron doping by K adsorption or Li/K exchange near the surface may induce a shift in the binding energies.
To examine the OER mechanism on LNMO, we conduct EIS at 1.70, 1.75, 1.80, and 2.00 VRHE. Figures 5(a) and 5(b) show the Nyquist plots, indicating three distinct resistance components: the uncompensated resistance corresponding to the real-axis intercept on the high-frequency side and two resistance components from the capacitive semicircle. The uncompensated resistance primarily consists of the wiring and electrolyte resistances.
Electrochemical impedance spectrometry (EIS) measurements for structure shown in Fig. 1(a). (a) and (b) Nyquist plots and fitted results at (a) 1.70 and 1.75 VRHE and (b) 1.80 and 2.00 VRHE. The inset shows equivalent model circuit applied for fitting. (c) and (d) Fitted values of (c) resistance and (d) constant-phase element from simulated EIS curves. The inset shows equivalent model circuit applied for fitting.
Electrochemical impedance spectrometry (EIS) measurements for structure shown in Fig. 1(a). (a) and (b) Nyquist plots and fitted results at (a) 1.70 and 1.75 VRHE and (b) 1.80 and 2.00 VRHE. The inset shows equivalent model circuit applied for fitting. (c) and (d) Fitted values of (c) resistance and (d) constant-phase element from simulated EIS curves. The inset shows equivalent model circuit applied for fitting.
The Nyquist plots were fitted to three candidate equivalent circuit models (X, Y, Z; see Table S2 in the supplementary material).46 Fitting was conducted using the Levenberg–Marquardt method within a confidence region (weight: absolute value of impedance; power factor: 2.0), resulting in calculated values within a certain range and within the confidence region. Models Y and Z exhibited relative errors exceeding 100% (Table S2 in the supplementary material). In contrast, model X exhibited errors of approximately 30% or less, except for 2.0 VRHE, where the error remains within 60%. Considering that a single model should be applicable at any potential in electrode reactions, model X was determined to be the most appropriate in this study.
The curves of the parameters obtained by fitting model X (Table I) are depicted as solid lines in Figs. 5(a) and 5(b). The equivalent circuit Model X [Fig. 5(a), inset] consisted of three resistance components (Rs, Ra, Rc) and two pseudocapacitance components (CPEd and CPEa). Note that CPE ≡ A(jω)−α is defined, where A is the CPE constant, α is the fractal parameter, ω is the frequency, and j is the imaginary unit. When α ∼ 1, the CPE value is equivalent to the capacitance, whereas when α ∼ 0, the CPE value is equivalent to the resistance. According to previous reports,47 the resistance components (Rs, Ra, and Rc) and two pseudocapacitance components (CPEd and CPEa) correspond to the uncompensated resistance (Rs), resistance derived from the adsorbed OH intermediates (Ra), charge transfer resistance at the thin film and electrolyte interface (Rc), pseudocapacitance of the electric double layer (CPEd), and pseudocapacitance derived from the adsorbed OH intermediates (CPEa).
Parameters of electrochemical impedance spectrometry (EIS) for model X. Rs: uncompensated resistance; Ra and Rc: resistances for the equivalent circuit of model X; CPEd: pseudo capacitance of the electric double layer; αd: fractal parameter of CPEd; CPEa: pseudo capacitance of OH adsorption; αa: fractal parameter of CPEa.
V/VRHE . | Rs (Ω cm2) . | Ra (Ω cm2) . | Rc (Ω cm2) . | CPEd/μFsα−1 . | αd . | CPEa/μFsα−1 . | αa . |
---|---|---|---|---|---|---|---|
1.70 | 3.60 ± 0.02 | 97.1 ± 2.3 | 16.8 ± 2.0 | 86.5 ± 7.7 | 1.00 ± 0.03 | 362 ± 23 | 0.670 ± 0.007 |
1.75 | 3.52 ± 0.02 | 10.9 ± 2.9 | 21.4 ± 2.7 | 344 ± 118 | 0.641 ± 0.030 | 61.2 ± 4.5 | 1.00 ± 0.05 |
1.80 | 3.41 ± 0.03 | 6.83 ± 1.06 | 5.09 ± 1.03 | 841 ± 453 | 0.527 ± 0.056 | 40.0 ± 3.5 | 0.975 ± 0.037 |
2.00 | 3.41 ± 0.09 | 1.33 ± 0.42 | 1.64 ± 0.65 | 1990 ± 1480 | 0.399 ± 0.089 | 30.0 ± 12.2 | 1.00 ± 0.09 |
V/VRHE . | Rs (Ω cm2) . | Ra (Ω cm2) . | Rc (Ω cm2) . | CPEd/μFsα−1 . | αd . | CPEa/μFsα−1 . | αa . |
---|---|---|---|---|---|---|---|
1.70 | 3.60 ± 0.02 | 97.1 ± 2.3 | 16.8 ± 2.0 | 86.5 ± 7.7 | 1.00 ± 0.03 | 362 ± 23 | 0.670 ± 0.007 |
1.75 | 3.52 ± 0.02 | 10.9 ± 2.9 | 21.4 ± 2.7 | 344 ± 118 | 0.641 ± 0.030 | 61.2 ± 4.5 | 1.00 ± 0.05 |
1.80 | 3.41 ± 0.03 | 6.83 ± 1.06 | 5.09 ± 1.03 | 841 ± 453 | 0.527 ± 0.056 | 40.0 ± 3.5 | 0.975 ± 0.037 |
2.00 | 3.41 ± 0.09 | 1.33 ± 0.42 | 1.64 ± 0.65 | 1990 ± 1480 | 0.399 ± 0.089 | 30.0 ± 12.2 | 1.00 ± 0.09 |
The phase vs. frequency plot (Fig. S6 in the supplementary material) exhibits a peak at 10 Hz when 1.70 VRHE. The phase angle approaches 0° with increasing potential. These trends are similar to those observed in previous impedance measurements for alkaline electrolysis using Ni electrodes.48 Therefore, we discuss the impedance results based on the interpretations in Ref. 47 as follows.
First, Rs remained nearly constant and independent of the applied potential. Considering that the setup of the water electrolysis cell was uninfluenced by the applied potential, this is reasonable. In contrast, the values of Ra decrease significantly with increasing potential and Rc decreases beyond 1.75 VRHE [Figs. 5(c), and 5(d), respectively]. This trend is analogous to that reported in Ref. 46. According to this literature, not only the charge transfer resistance at the thin film and electrolyte interface (Rc) but also the contribution of the charge transfer resistance to adsorbed OH intermediates (Ra) is highlighted. Therefore, for the LNMO thin-film electrode, a similar suggestion was made regarding the contribution of the charge-transfer resistance to the adsorbed OH species.46 This agrees with the Tafel plot results, indicating that the conversion of the adsorbed OH intermediates is the rate-determining step [Fig. 2(b)]. Consequently, reducing the charge-transfer resistance to the adsorbed OH intermediates at lower applied potentials may lead to enhanced OER activity.
Finally, we discuss the constant-phase elements. With an increase in potential, the capacitance of the electric double layer (CPEd) significantly increases, while that of the constant-phase element associated with adsorption (CPEa) decreases (Table I). In particular, the reduction in CPEa was similarly observed in the alkaline water electrolysis of Fe oxides.48 According to this literature, CPEa appears owing to the adsorption of OH intermediates, and at high potentials, these intermediates almost entirely cover the surface (i.e., a coverage rate of 1). Therefore, for the LNMO thin-film electrode, we assume that at potentials above 1.75 VRHE, the adsorbed OH intermediates almost completely cover the surface, ensuring a sufficient supply of reactants to the reaction sites. Consequently, we estimate that the adsorbed OH intermediates are adequately supplied to the reaction active sites at potentials above 1.75 VRHE, leading to reduced Ra. In contrast, at 1.70 VRHE, the significant value of CPEa suggests that OH adsorption intermediates are not adequately supplied, resulting in a reduced OER current.
IV. CONCLUSION
We quantitatively evaluated the OER using LNMO polycrystalline thin films as electrode catalysts. XPS measurements indicate the presence of Ni3+ and Mn2+ on both the as-grown and post-OER LNMO surfaces, suggesting the formation of Ni3+ due to the charge transfer between the Ni and Mn ions. In addition, we discussed the reaction mechanism through electrochemical impedance spectrometry measurements. In future studies, higher OER activity is anticipated in Li1−xNi0.5Mn1.5O4 (x = 0.00–0.50) with high-valence Ni electronic states by the Li desertion.
SUPPLEMENTARY MATERIAL
See the supplementary material for details on the equivalent circuit model and a conceptual diagram of the surface model. The supplementary material also includes the results of EDX, XPS, and oxygen evolution reactions.
ACKNOWLEDGMENTS
This study was supported by JSPS Kakenhi (Grant Nos. 18H03876, 21K18892, 21H00141, 22H01953, 17H05216, 19H02596, 19H04689, 18H03876, and 18H05514), the JST-PRESTO (Grant No. JPMJPR17N6), JST-CREST (Grant No. JPMJCR1523) programs, and DX-GEM (Grant No. JPMXP1122712807). K.H. acknowledges funding from the Advanced Human Resource Development Fellowship for Doctoral Students (Tokyo Institute of Technology). XPS measurements of this work were supported by “Advanced Research Infrastructure for Materials and Nanotechnology in Japan (ARIM)” of the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Grant No. JPMXP1223UT0202. We are also grateful to the platform of Analytical Instruments for Chemistry and Materials Science (Tokyo Institute of Technology) for providing the equipment for the Raman and SEM measurements. We are thankful to Professor Takanabe and Assistant Professor Obata for the support of gas chromatography.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Kentaro Hatagami: Conceptualization (lead); Investigation (lead); Writing – original draft (lead); Writing – review & editing (equal). Kazunori Nishio: Writing – review & editing (supporting). Ryota Shimizu: Conceptualization (supporting); Writing – review & editing (equal). Taro Hitosugi: Funding acquisition (lead); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.