ACS Publications. Most Trusted. Most Cited. Most Read
Light-Mediated Electrochemical Synthesis of Manganese Oxide Enhances Its Stability for Water Oxidation
My Activity
  • Open Access
Article

Light-Mediated Electrochemical Synthesis of Manganese Oxide Enhances Its Stability for Water Oxidation
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

ACS Nanoscience Au

Cite this: ACS Nanosci. Au 2023, 3, 4, 310–322
Click to copy citationCitation copied!
https://doi.org/10.1021/acsnanoscienceau.3c00002
Published April 23, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

New methods are needed to increase the activity and stability of earth-abundant catalysts for electrochemical water splitting to produce hydrogen fuel. Electrodeposition has been previously used to synthesize manganese oxide films with a high degree of disorder and a mixture of oxidation states for Mn, which has led to electrocatalysts with high activity but low stability for the oxygen evolution reaction (OER) at high current densities. In this study, we show that multipotential electrodeposition of manganese oxide under illumination produces nanostructured films with significantly higher stability for the OER compared to films grown under otherwise identical conditions in the dark. Manganese oxide films grown by multipotential deposition under illumination sustain a current density of 10 mA/cm2 at 2.2 V versus reversible hydrogen electrode for 18 h (pH 13). Illumination does not enhance the activity or stability of manganese oxide films grown using a constant potential, and films grown by multipotential deposition in the dark undergo a complete loss of activity within 1 h of electrolysis. Electrochemical and structural characterization indicate that photoexcitation of the films during growth reduces Mn ions and changes the content and structure of intercalated potassium ions and water molecules in between the disordered layers of birnessite-like sheets of MnOx, which stabilizes the nanostructured film during electrocatalysis. These results demonstrate that combining multiple external stimuli (i.e., light and an external potential) can induce structural changes not attainable by either stimulus alone to make earth-abundant catalysts more active and stable for important chemical transformations such as water oxidation.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2023 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Hydrogen production through the splitting of water has the potential to provide a clean fuel source to replace petroleum for the transportation industry. In electrochemical water splitting to generate H2 and O2, the oxidative half-reaction, known as the oxygen evolution reaction (OER) is kinetically more challenging to achieve at low overpotentials as it involves the transfer of four protons and four electrons and the formation of an oxygen–oxygen double bond. (1) While oxides of precious metals like Ir and Ru are currently the best-performing electrocatalysts for the OER, (2) their rarity makes them prohibitively expensive in addressing the world’s current and future energy needs. (3,4) First-row transition metal oxides including the oxides of Mn, Fe, Ni, and Co are highly earth-abundant, (3) so they could be incorporated into systems for photo- and electrochemical water splitting to produce hydrogen fuel on a terawatt scale. (2,5−11) However, methods to improve the activity and long-term stability of such first-row transition metal oxides for the OER are needed.
Manganese oxides exhibit a wide variety of different stoichiometries and structures and have provided a testbed for elucidating the structure–activity relationships in transition metal oxide catalysts for water oxidation. (12−36) For example, Mn2O3 with a bixbyite structure exhibits higher activity for photochemical water oxidation relative to other stoichiometries (e.g., Mn3O4 and different phases of MnO2). (17,18) Melder and coworkers have recently compiled a comparison of the activity and stability of different manganese oxides for electrochemical water oxidation. (36) Bixbyite Mn2O3 and disordered birnessite MnOx (containing a mixture of Mn3+ and Mn4+ ions) have demonstrated some of the lowest overpotentials of manganese oxides for electrochemical water oxidation, with both operating at a current density of 10 mA/cm2 in 1 M KOH with overpotentials between 330 and 340 mV. (23,35) Nanoparticles of γ-MnO2 (a mixture of the pyrolusite and ramsdellite phases of MnO2) have exhibited high stability, performing the OER in acid (pH 2) at a current density of 10 mA/cm2 for 8000 h (24) While different methods of measuring the activity for water oxidation (e.g., chemical, photochemical, or electrochemical) and different experimental conditions have led to the development of various structure–activity relationships, two important trends have emerged for electrochemical water oxidation: (1) the presence of manganese in a mixture of oxidation states (e.g., Mn3+ and Mn4+) at potentials where the OER occurs is important for the redox cycling of active sites that takes place during water oxidation (13,17,20,25−27,29−34,37) and (2) relative to highly crystalline forms of manganese oxide, disordered structures can exhibit higher activity. (13,14,22,27,31,32,37)
The electrochemical growth of manganese oxide provides a method to both synthesize films with a mixture of oxidation states for Mn and introduce disorder into the resulting films. Multipotential deposition through alternating or cycling the applied potential to both oxidize Mn2+ ions in solution and reduce Mn4+ ions in the growing film increases the relative amount of Mn3+ incorporated into the films (relative to deposition at a constant potential). MnOx films prepared through multipotential electrodeposition (in which the Mn ions are in a mixture of +3 and +4 oxidation states) possess a structure that resembles birnessite MnO2 but with a high amount of disorder between layers of edge-sharing MnO6 octahedra; these disordered films exhibit higher activity than ordered birnessite MnO2 films electrodeposited at a constant potential. (27,29,31,32) While electrodeposited MnOx films exhibit promising activity for electrochemical water oxidation, they are typically stable for only a few hours at current densities between 0.1 and 1 mA/cm2. (27,31,36) Current densities near 10 mA/cm2 are needed to match the incident flux of sunlight and convert solar energy into hydrogen either through photoelectrochemical water splitting or by coupling a photovoltaic device with an electrolyzer. Thus, electrodeposited MnOx films need to be more stable at higher current densities before they can be incorporated into such systems.
Illumination during the electrochemical growth of metal oxide (e.g., Cu2O) and chalcogenide (e.g., PbSe and Se–Te) semiconductors has been used to alter their composition, structure, and nanoscale morphology. (38−44) Photoexcitation during electrodeposition both enhances the concentration of mobile charge carriers and alters the spatial distribution of charge transfer at the interface between the deposition solution and the growing semiconductor film. (38,39,43) Birnessite MnO2 is a light-absorbing semiconductor, but its efficiency for the photoelectrochemical conversion of incident photons into chemical products has so far been relatively low. (45−47) On the other hand, the photoreduction of Mn4+ ions in birnessite MnO2 by sunlight with the concomitant oxidation of organic matter plays an important role in the biogeochemical cycle of aquatic ecosystems. (13,48−51) Thus, we hypothesized that the photoreduction of Mn4+ may be a way to alter the structure of MnOx films during electrochemical growth. (27)
We show that manganese oxide films synthesized by multipotential electrodeposition under illumination with a 405 nm light-emitting diode (LED) exhibit higher current densities for the OER and significantly better stability compared to films synthesized under otherwise identical conditions in the dark. MnOx films grown under illumination have a Tafel slope of 64.1 mV/decade in alkaline solution (pH 13), and they maintain a stable overpotential (∼2.2 V vs RHE) for 18 h while performing the OER at a current density of 10 mA/cm2. On the other hand, MnOx films grown in the dark lose nearly all their activity for the OER after 1 h of electrolysis at this current density. Importantly, the enhanced stability is only observed when combining multipotential electrodeposition with illumination during growth. Electrochemical and structural characterization indicate that the enhanced catalytic activity and stability for the OER observed in MnOx films deposited under illumination arises from light-induced structural and compositional changes during growth. MnOx films grown by multipotential deposition contain Mn ions in a mixture of the +2, +3, and +4 oxidation states and possess a disordered layered structure. We propose that the interlayer structure of intercalated potassium ions and water molecules in MnOx films grown under illumination enhances their stability during the redox cycling and the transport of interlayer species that occurs during electrochemical water oxidation.

Results and Discussion

Click to copy section linkSection link copied!

MnOx films were electrodeposited based on modifications to a procedure developed by Huynh and coworkers. (31) A square-wave potential waveform is applied to the working electrode (i.e., glass coated with fluorine-doped tin oxide) in a solution containing MnCl2 and KNO3 (see the Experimental Section for details). Electrodeposition of manganese oxide while cycling or alternating the potential has been shown to incorporate a higher fraction of Mn3+ ions as well as induce disorder into the growing manganese oxide film relative to deposition at constant potential (see the Supporting Information for additional discussion of this reaction). (27,29,31,32) We thus refer to the manganese oxide films as MnOx as they contain Mn in different oxidation states (i.e., Mn2+, Mn3+, and Mn4+ as discussed further below). We varied the values of the two potentials used in the square wave and their duration (see Supporting Discussion and Figure S1a,b in the Supporting Information). Although we modified the procedure by increasing the overall deposition time and illuminating the films during growth, we found that the square-wave potential originally developed by Huynh and coworkers (31) produced MnOx films with the highest activity for the OER. This sequence consists of an oxidizing potential of 0.9 V versus Ag/AgCl for 3 s (used to oxidize Mn2+ ions in solution to Mn4+ and induce the crystallization of MnO2) and a reducing potential of −0.6 V Ag/AgCl for 2 s [used to increase the local pH through the reduction of NO3 and induce the formation of Mn(OH)2]. We varied the total deposition time using this pulse sequence (Figure S1c) and found that deposition times between 40 and 60 min gave the best combination of activity and stability when the MnOx films were used as catalysts for the OER.
We next compared the activity and stability of MnOx films grown through multipotential electrodeposition under illumination with those of films grown in the dark. A high-power LED with a peak wavelength of 405 nm was used as the light source to grow the illuminated MnOx films (see the Experimental Section for further details). All subsequent electrochemical testing of the films was performed in the dark. Cyclic voltammetry (CV) was performed in a solution of 0.1 M KOH and 0.9 M KNO3 to characterize the as-synthesized, MnOx films grown in the dark and under illumination, as shown in Figure 1a. The initial current densities of both the illuminated and dark films at potentials where the OER can occur were similar during the first CV scan. The onset potential of the as-synthesized, illuminated MnOx film for the OER was 1.57 V versus the reversible hydrogen electrode (RHE), and the film produced a current density of 9.9 mA/cm2 at 2.15 V versus RHE. Furthermore, the film grown under illumination retained the same high activity during the second CV scan. The onset potential of the as-synthesized MnOx film grown in the dark was 1.65 V versus RHE, and the film produced a current density of 8.8 mA/cm2 at 2.15 V versus RHE during the first CV scan. However, the current density dropped to 7.8 mA/cm2 at 2.15 V versus RHE during the second CV scan. CV scans of the illuminated and dark MnOx films after 1 h of constant-current electrolysis at a current density of 10 mA/cm2 are shown in Figure 1b. The onset potential of the illuminated sample remained at 1.57 V versus RHE, and the current density at 2.15 V remained at 9.9 mA/cm2 after electrolysis. However, the current density of the dark sample decreased significantly to 0.8 mA/cm2 at 2.15 V versus RHE. The onset potential for the dark film after electrolysis was difficult to measure because it was too close to background current density produced by a bare fluorine-doped tin oxide (FTO) substrate. Potential sweeps were also performed for films after electrolysis at a current density of 3 mA/cm2. At a lower current density used for performing electrolysis, MnOx films grown under illumination were also more stable relative to films grown in the dark and retained their activity toward OER (Figure S2). These results indicate that MnOx films grown under illumination exhibit enhanced stability for water oxidation compared to films grown under otherwise identical conditions in the dark.

Figure 1

Figure 1. (a,b) CV scans (using IUPAC convention) of MnOx films measured in 0.1 M KOH and 0.9 M KNO3. (a) CV scans of the as-synthesized films grown under illumination (red trace) and in the dark (blue trace). (b) CV scans of the same illuminated film (orange trace) and dark film (light-blue trace) after 1 h of constant-current electrolysis at a current density of 10 mA/cm2. The arrows indicate the onset potentials for the OER. (c) Applied potential vs time for an illuminated MnOx film (red trace) and a dark film (blue trace) during constant-current electrolysis at a current density of 10 mA/cm2 to test the stability of the films during the OER.

The stability of the MnOx films during the OER was further quantified by performing electrolysis at a constant current density of 10 mA/cm2 in a solution containing 0.1 M KOH and 0.9 M KNO3, as shown in Figure 1c. The stability of each electrode was evaluated based on the change in the potential needed to maintain a constant current; an active and stable electrocatalyst maintains a high current density at a lower overpotential (i.e., a less positive potential for oxidation). During the 40 min test, the MnOx film grown under illumination operated at a relatively stable potential window near 2.2 V versus RHE (Figure 1c). However, for the dark film, the potential required to maintain 10 mA/cm2 continuously increased over the 40 min period and eventually reached a maximum value of 3.25 V versus RHE. In a control experiment, a bare FTO substrate required a similar potential (∼3.45 V vs RHE) to produce a current density of 10 mA/cm2, indicating a near-complete loss of activity for the MnOx film grown in the dark. Based on these electrochemical tests, MnOx films grown under illumination exhibit better performance than the dark films with higher current density and stability for the OER. A MnOx film grown under illumination was tested at a current density of 10 mA/cm2 for a period of 18 h (Figure S3), and the potential only increased by 0.15 V over the testing period.
Scanning electron microscopy (SEM) images of MnOx films before and after using them as electrocatalysts for the OER are shown in Figure 2. The initial films grown under illumination and in the dark exhibit similar morphologies consisting of aggregates of nanoscale platelets that cover the FTO substrate. After electrolysis for 30 min at a current density of 3 mA/cm2, the MnOx film grown under illumination maintained the nanoplate morphology. However, the individual nanoplates became less distinct for the MnOx film grown in the dark. The electrocatalytic activity of the MnOx film grown under illumination was stable, while the sample grown in the dark dramatically decreased in activity; these observations indicate that the nanoplate morphology is important in determining the surface area that is catalytically active.

Figure 2

Figure 2. SEM images of (a) as-synthesized MnOx film grown under illumination with a 405 nm LED, (b) MnOx film synthesized under illumination after 30 min of electrolysis at 3 mA/cm2, (c) as-synthesized MnOx film grown in the dark, and (d) MnOx film synthesized in the dark after 30 min of electrolysis at 3 mA/cm2. The scale bar of 1 μm applies to all images.

The electrochemically active surface areas (ECSAs) of different MnOx films were compared by measuring their electrochemical double-layer capacitance, CDL, which is directly proportional to the ECSA. The double-layer capacitance of each film was calculated using CV scans that were performed at different scan rates over a potential window where there is no Faradaic current (see Figure S4). (52) The as-synthesized MnOx film grown in the dark had a CDL of 1.6 ± 0.1 mF/cm2, but this value decreased to 0.25 ± 0.02 mF/cm2 after constant-current electrolysis for 40 min at 5 mA/cm2. The as-synthesized MnOx film grown under illumination had a CDL of 2.8 ± 0.1 mF/cm2, and this value decreased to 2.0 ± 0.2 mF/cm2 after electrolysis. These results indicate that the initial ECSA of the illuminated MnOx film was greater than that of the dark film. The difference in active surface area is consistent with the higher initial current density for OER observed for the illuminated MnOx film compared to the dark film (Figure 1a). Assuming a specific capacitance of 0.04 mF/cm2, a value that is recommended for metal oxide electrodes in alkaline solution when the true specific capacitance is not known, (22,52) the ESCA of the MnOx electrode grown under illumination is 70 cm2 for a geometric area of 1 cm2. Thus, the nanoscale morphology of the electrodeposited MnOx films provides a high surface area for electrocatalysis. Furthermore, the significant decrease in the ECSA of the dark film after electrolysis helps to explain the loss of activity for the OER from a morphological perspective.
Tafel plots for the OER using MnOx films synthesized in the dark and under illumination before and after electrolysis are shown in Figure 3a. The Tafel slope reflects how much the applied potential needs to be increased to increase the current density by a factor of 10. A low Tafel slope is desirable as it means the overpotential needed to drive the reaction is small. When the electrode reaction is not limited by mass transport, the Tafel slope reflects the kinetics of the OER. The potential applied to each MnOx film was scanned from 2.0 to 1.9 V versus RHE under stirring at 500 rpm to minimize mass-transport effects. The Tafel slopes were obtained by linear fitting, and the errors are provided after each Tafel slope. The as-synthesized, MnOx film grown under illumination exhibited a Tafel slope of 64.1 ± 0.4 mV/decade, while the as-synthesized, dark film had a slope of 72.7 ± 0.7 mV/decade. A Tafel slope near 60 mV/decade is consistent with a rapid one-electron, charge-transfer step that is in quasi-equilibrium and followed by a rate-determining chemical process. For the OER using MnOx films in alkaline solution, proton-coupled electron transfer at a surface Mn3+ site followed by the release of O2 from the electrode surface as the rate-determining step has been proposed to account for this Tafel slope. (30,53) Next, both MnOx electrodes were subjected to constant-current electrolysis at 5 mA/cm2 for 40 min in an electrolyte with 0.1 M KOH and 0.9 M KNO3. The Tafel slope for the illuminated MnOx film after electrolysis was 71.4 ± 0.8 mV/decade, while the slope for dark film after electrolysis was 138 ± 2 mV/decade. These results show that the MnOx film grown under illumination was able to maintain a relatively high activity during the OER as its Tafel slope only increased by 7.3 mV/decade. However, the Tafel slope of the dark film increased significantly after OER by 65.9 mV/decade, which indicates that it was not stable during oxygen evolution. The changes in Tafel slopes support characterization by CV and constant-current electrolysis, which indicates that MnOx films synthesized under illumination are more stable and active during the OER.

Figure 3

Figure 3. Electrochemical characterization of electrodeposited MnOx films. (a) Tafel plots of MnOx films synthesized under illumination before (red triangles) and after (orange circles) electrolysis at 5 mA/cm2 for 40 min and films synthesized in the dark before (blue triangles) and after (light-blue squares) electrolysis. The symbols represent the actual data, and the lines connecting the symbols represent linear fits of the data used to calculate Tafel slopes. The errors were determined by linear fitting, and the R2 values of the linear fits were greater than 0.995. (b) Mott–Schottky plots of MnOx films synthesized under illumination before (red triangles) and after (orange circles) electrolysis at 5 mA/cm2 for 1 h, and films synthesized in the dark before (blue triangles) and after (light-blue squares) electrolysis. The lines represent linear fits of the raw data used to calculate the Mott–Schottky slopes. (c) Nyquist plots of MnOx films synthesized under illumination before (red) and after (orange) electrolysis at 3 mA/cm2 for 1.5 h, and films synthesized in the dark before (blue) and after (light blue) electrolysis. The inset shows the equivalent circuit model used to fit the Nyquist plots. R1 represents the uncompensated resistance from the solution, FTO substrate, and MnOx film, C2 represents the double-layer capacitance, R2 represents the polarization resistance, and W2 represents a Warburg impedance.

We used electrochemical impedance spectroscopy (EIS) to correlate the changes in the activity of the MnOx films for the OER with changes in their resistivity and ECSA. EIS was carried out in an aqueous solution containing 0.1 M KOH and 0.9 M KNO3 in the dark at different applied potentials. Two types of plots were derived from the EIS data: Mott–Schottky plots and Nyquist plots. The Mott–Schottky equation relates the capacitance (C) of a semiconductor electrode to its mobile carrier concentration (Nd) and its flat-band potential (VFB), as shown in eq 1
1C2=(2eεε0A2Nd)[VVFBkBTe]
(1)
where e is the charge of an electron, ε0 is the permittivity of vacuum, ε is the relative permittivity of the semiconductor, A is the ESCA, V is the applied potential, kB is the Boltzmann’s constant, and T is the temperature.
Mott–Schottky plots (i.e., 1/C2 versus the applied potential, V vs RHE) measured at a frequency of 10 kHz for both the as-synthesized MnOx films and films after 1 h of electrolysis at a current density of 5 mA/cm2 are shown in Figure 3b. Figure S5 shows a magnified view of the Mott–Schottky plots at 10 kHz near the abscissa to show that the slopes are positive for all four samples, which indicates the MnOx films are n-type semiconductors, similar to previous reports. (46) Mott–Schottky plots made at other frequencies gave similar results and are shown in Figure S6. The slope of the Mott–Schottky plot for the MnOx film grown under illumination was similar before and after electrolysis. On the other hand, the slope of the dark film after electrolysis was much greater compared to the dark film before electrolysis and films grown under illumination both before and after electrolysis. As seen from eq 1, decreases in the surface area, carrier concentration, and relative permittivity of the semiconductor will all increase the Mott–Schottky slope. Based on measurements of the double-layer capacitance (Table S1), the ESCA for the MnOx film grown in the dark decreased by a factor of 6.4 after electrolysis. However, the Mott–Schottky slope increased by a factor of 300 for the dark MnOx film after electrolysis, indicating that the effective carrier concentration and relative permittivity also change during the OER for films grown in the dark. These results show that in addition to the morphological changes observed by SEM after electrolysis, the structural changes that occur during the OER also lead to a much lower conductivity for MnOx films grown in the dark.
Nyquist plots were used to quantify changes in the resistance of the MnOx films after electrolysis (Figure 3c). The x- and y-axes of the Nyquist plots show the real and imaginary components of the impedance, respectively. The inset in Figure 3c shows the equivalent circuit model used to fit each of the Nyquist plots. The resistor, R1, accounts for uncompensated resistance in the electrochemical cell including that from the solution, the FTO substrate, and the MnOx film. While the Nyquist plots were obtained at a potential where the OER does not occur (i.e., 0.2 V vs Ag/AgCl), the second resistor element, R2, accounts for any polarization resistance away from the open-circuit potential. Based on fitting the impedance data to this circuit model, the values of R1 and R2 for the as-synthesized MnOx film grown under illumination were 38 ± 1 Ω and 63 ± 3 Ω, respectively. After electrolysis at 3 mA/cm2 for 1.5 h, the values of R1 and R2 increased to 43 ± 2 Ω and 70 ± 3 Ω. The resistance of the MnOx film grown in the dark increased much more substantially after electrolysis. The values of R1 and R2 for the as-synthesized film grown in the dark were 28 ± 1 Ω and 54 ± 3 Ω, respectively, while they increased to 159 ± 2 Ω and 161 ± 3 Ω after electrolysis at 3 mA/cm2 for 1.5 h. The higher resistance for the dark film after OER as measured from the fit to the Nyquist plot agrees with the large increase in slope seen in the Mott–Schottky plot. Thus, the loss in activity for MnOx films grown in the dark is due to a combination of structural and morphological changes that decrease both the film’s active surface area and conductivity. The cumulative results from electrochemical characterization indicate that illumination of the MnOx films during their growth alters their structure to stabilize them against such changes while they are undergoing the OER.
Cycling the applied potential during electrochemical growth of manganese oxide films has been previously shown to produce disordered MnOx films in which the Mn ions possess a mixture of oxidation states. (27,29,31,32) Separately, illumination has been shown to reduce Mn in both synthetic and naturally occurring manganese oxides. (13,48−51) However, the combined effect of these two stimuli on the distribution of the resulting oxidation states for Mn has not been studied. We used X-ray photoelectron spectroscopy (XPS) to compare the differences in MnOx films grown in the dark and under illumination (see Figure 4 and Table 1). For each sample, we fit the photoelectron peak in the binding energy region for Mn 2p3/2 electrons based on a procedure developed by Nesbitt and Banerjee (see the Experimental Section for details of the fitting procedure). (54) Each oxidation state of Mn was decomposed into five different peaks due to coupling between unpaired valence electrons and unpaired core electrons (once photoionization has taken place). (55−57) This procedure has been applied to manganese oxide powders including MnO, MnOOH, Mn2O3, MnO2, and KMnO4. (20,54,57) More recently, it has been used to characterize the distribution of oxidation states in disordered, birnessite-like MnOx particles (similar in structure to the MnOx films studied here). (37) We first applied this fitting procedure to a series of manganese oxide powders (see Tables 1, S2, and Figure S7). The 2p3/2 peak of a bixbyite Mn2O3 powder could be fit to predominantly Mn3+. The 2p3/2 peak of a hausmannite Mn3O4 powder could be fit to a mixture of Mn3+ and Mn2+ but with an excess of Mn3+ relative to the 2:1 ratio that would be expected for this compound (possibly due to surface oxidation). Birnessite MnO2 synthesized by a hydrothermal method showed a mixture of Mn4+ and Mn3+ as expected based on its X-ray diffraction pattern. An MnOx film electrodeposited at a constant potential with the birnessite structure (based on Raman spectroscopy described below) also showed a mixture of Mn4+ and Mn3+. However, the electrodeposited film exhibited a higher percentage of Mn2+ relative to the birnessite powder prepared by hydrothermal synthesis. Although there is significant overlap in the binding energies of the Mn 2p3/2 peaks in different oxidation states, results obtained from manganese oxide powders show that the fitting procedure can be used to compare the differences in the distribution of the oxidation states for the electrodeposited films.

Figure 4

Figure 4. XPS of electrodeposited MnOx films in the binding energy region of Mn 2p3/2 electrons. (a,b) MnOx films synthesized under illumination before (a) before and (b) after electrolysis at 5 mA/cm2 for 1 h (c,d) MnOx films synthesized in the dark (c) before and (d) after electrolysis. For each spectrum, the shaded regions show the deconvolution of the peak into contributions from Mn4+ (purple), Mn3+ (green), and Mn2+ (light red). The gray circles show the raw data, and the black line shows the sum of fitting the peak to the different oxidation states of Mn. The dashed line shows the residual of the fit. The peak maxima for each peak in the deconvolution are provided in Table S3 of the Supporting Information.

Table 1. Relative Peak Areas for Mn in Different Oxidation States After Deconvoluting the Mn 2p3/2 Peak in the Photoelectron Spectra of Different Manganese Oxide Samplesa
componentratio of peak areas (%)
sampleilluminated MnOxilluminated MnOx–OERdark MnOxdark MnOx–OERMnOx–constant potentialbirnessite MnO2bixbyite Mn2O3hausmannite Mn3O4
Mn4+13.629.923.025.862.857.64.85.0
Mn3+67.162.770.467.820.140.495.271.8
Mn2+19.37.46.66.417.12.00.023.2
a

Illuminated MnOx = film grown by multipotential deposition under illumination, dark MnOx = film grown by multipotential deposition in the dark. OER indicates that the film underwent electrolysis at a current density of 5 mA/cm2 for 1 h. MnOx – constant potential = film grown at a constant potential of 0.9 V vs Ag/AgCl.

XPS reveals that illumination during multipotential electrodeposition reduces the average oxidation state of Mn relative to MnOx films grown in the dark (Figure 4 and Table 1). MnOx films grown under illumination exhibit a similar percentage of Mn3+ compared with films grown in the dark, but they possess a higher percentage of Mn2+ (and correspondingly a lower percentage of Mn4+). After constant-current electrolysis at 5 mA/cm2 for 1 h, the fraction of Mn4+ increases for both films grown under illumination and in the dark, indicating that the OER also leads to partial oxidation of the MnOx films. Notably, the relative amounts of Mn in different oxidation states were similar after the OER for both films grown under illumination and in the dark. Characterization of the samples by X-ray absorption near-edge spectroscopy (XANES) supports the conclusions drawn from XPS (see Figures S8 and S9). The MnOx film grown under illumination has a lower average oxidation state compared to the film grown in the dark, and both films become more oxidized after the OER.
The structures of the MnOx films were characterized by Raman spectroscopy (Figure 5a). For comparison, the Raman spectrum of a film grown at a constant potential of 0.9 V versus Ag/AgCl is shown in Figure S10. The peaks of this film at 493, 567, and 651 cm–1 are indicative of birnessite MnO2. (58) Birnessite is composed of layers of edge-sharing MnO6 octahedra (Figure 5b). Different types of birnessite (e.g., hexagonal vs triclinic) exhibit different stacking sequences of the layers. While Mn ions are primarily in the +4 oxidation state, a portion of the Mn ions can be in the +3 oxidation state. Metal cations (e.g., Na+ or K+) along with H2O can intercalate between the layers of MnO2 to compensate for Mn ions in a reduced oxidation state and to screen charge between layers. Thus, the interlayer spacing and peak positions in the Raman spectra of birnessite and birnessite-like films are sensitive to the amount and structure of intercalated species. Because the electrolyte solution used for both electrochemical deposition and electrolysis in our studies contained KOH and KNO3, we assume K+ is the primary ion intercalated into our MnOx films, which is supported by XPS results showing the presence of K+ (Figure S11). Typically, the peak at 651 cm–1 (which can vary between ∼620 and 651 cm–1) is assigned to the symmetric stretching of the Mn–O bonds in MnO6 octahedra, while the peak at 567 cm–1 (which can vary between ∼570 and 590 cm–1) is attributed to the stretching of the Mn–O bonds parallel to the basal plane of the layers of MnO2. (58−60) A recent report by Scheitenberger and coworkers, which used density-functional theory to calculate the Raman spectra of birnessite films with different structures of interlayer H2O, has called these traditional assignments into question and provides evidence that the peak at 651 cm–1 arises from the vibrations of interlayer H2O rather than Mn–O bonds. (60)

Figure 5

Figure 5. (a) Raman spectra of electrodeposited MnOx films. Red trace: as-synthesized MnOx film grown under illumination; orange trace: illuminated film after electrolysis at 5 mA/cm2 for 1 h; blue trace: as-synthesized MnOx film grown in the dark; light-blue trace: dark film after electrolysis at 5 mA/cm2 for 1 h. (b) Crystal structure of potassium-intercalated birnessite hydrate, K0.27(Mn0.98O2)(H2O)0.51. Mn atoms are pink, oxygen atoms are gray, and potassium atoms are purple. The oxygen atoms within the layer represent intercalated H2O (hydrogen atoms not shown). The structure was made using collection code no 55407 from the Inorganic Crystal Structure Database.

The square-wave potential (i.e., alternating between 0.9 V versus Ag/AgCl for 3 s and −0.6 V for 2 s) leads to significant disorder in the resulting MnOx films (both for deposition in the dark and under illumination). Thus, the Raman spectra shown in Figure 5a, exhibit reduced peak intensities compared to the film grown at constant potential. Raman spectra of disordered MnOx particles and films have been previously interpreted based on their deviations from well-ordered birnessite MnO2. (14,32,37) While all peaks observed in the Raman spectra (probing the bulk structure) of our samples are indicative of disordered birnessite, XPS indicates a large fraction of Mn3+ ions and the presence of Mn2+ at the surface of the films. Hausmannite Mn3O4 is composed of Mn3+ and Mn2+ ions, and electrochemical cycling of manganese oxide films have been previously shown to produce a disordered mixture of birnessite MnO2 and hausmannite Mn3O4. (31,32,61,62) As the most intense peak in the Raman spectrum of Mn3O4 occurs at 657 cm–1 (i.e., at a similar frequency of 651 cm–1 for birnessite), we cannot exclude the presence of Mn3O4 in our films.
The as-synthesized MnOx films possess a peak at 651 cm–1 for the film grown under illumination and at 652 cm–1 for the film grown in the dark. After electrolysis at 5 mA/cm2 for 1 h, the intensity of this peak decreases significantly for both films (Figure 5a). SEM images show little change in the nanoplate morphology of the MnOx film grown under illumination after the OER (Figure 2). We attribute the loss of the peak at 651 cm–1 along with the appearance of a peak at 567 cm–1 (described further below) that occurs after electrolysis using the MnOx film grown under illumination to reflect a change in the content and structure of intercalated species (i.e., K+ and H2O) rather than a change in the structure of the MnOx layers. After electrolysis, the Raman spectrum of the MnOx film grown in the dark only shows a broad hump with low intensity between 500 and 700 cm–1; the loss of long-range order observed by Raman spectroscopy agrees with the changes in morphology observed by SEM.
The MnOx film grown under illumination also shows a peak at 611 cm–1 before electrolysis, which shifts to 615 cm–1 after electrolysis (Figure 5a). This peak is not observed for MnOx films grown in the dark. A peak between 606 and 611 cm–1 is occasionally observed in birnessite films, (32,37,59,63,64) but this peak is typically left unassigned. Along with the peak at 651 cm–1, Scheitenberger and coworkers also assigned this peak to vibrations of interlayer H2O in birnessite. (60) Finally, a peak appears at 567 cm–1 after electrolysis for the MnOx film grown under illumination. XPS shows that the amount of K+ in the film increased significantly after electrolysis only for the MnOx film grown under illumination (Figure S11). Chen and coworkers previously saw an increase in the intensity of this peak with an increase in the amount of K+ intercalated into birnessite MnO2 films. (59) Thus, we assign the appearance of the peak at 567 cm–1 to a change in the interlayer spacing induced by the intercalation of K+ during the OER. Together, the Raman spectra of these samples indicate that the content and structure of species intercalated between the layers of MnOx differ significantly between films grown in the dark and under illumination.
The presence of Mn3+ in manganese oxides has been correlated with higher activity for water oxidation. (17,20,25−27,29−34) Different coordination geometries for catalytically active Mn3+ ions have been proposed. Bixbyite Mn2O3, which consists of edge-sharing, distorted MnO6 octahedra (as the Mn3+ ions possess a Jahn–Teller distortion), exhibits high activity for both photochemical and electrochemical water oxidation. (17,18,20,23) Corner-sharing Mn3+O6 octahedra and tetrahedrally coordinated Mn3+ ions have also been proposed to be active sites for water oxidation. (14,20,32) Our initial hypothesis was that photochemical reduction could enhance the amount of Mn3+ incorporated into electrodeposited MnOx films; this hypothesis was based on work by Marafatto and coworkers, who showed that photoexcitation of birnessite MnO2 particles with 400 nm illumination led to the photoreduction of Mn4+ to Mn3+. (51) However, XPS of our samples indicates that illumination during multipotential electrodeposition leads to a higher fraction of Mn2+ incorporation rather than Mn3+. The photoreduction of Mn4+ in biogenic manganese oxides by sunlight normally leads to the dissolution of Mn2+ ions. (48−50) Hocking and coworkers showed that disordered, birnessite-like nanoparticles (containing a mixture of Mn4+ and Mn3+ ions) embedded in Nafion could be reversibly photoreduced to Mn2+ and then electrochemically oxidized over multiple cycles. (13) In a similar manner, photoreduction during multipotential deposition could enhance the concentration of Mn2+ ions trapped in electrodeposited MnOx films. Electrochemical oxidation of Mn during the OER (performed without photoexcitation) then brings the concentration of Mn2+ to a level similar to those of films grown by multipotential deposition in the dark. Interestingly, we found that illumination during constant potential deposition did not improve the activity or stability of the films for the OER. Films grown at constant potential under illumination were inactive for the OER, as has been previously observed for films grown at constant potential in the dark. (27,31,32) Thus, photoexcitation appears to work in concert with the disorder induced by multipotential deposition to incorporate a higher fraction of Mn2+ ions and change the interlayer structure of the MnOx films.
While both XPS and XANES indicate that the photoreduction of MnOx films grown by multipotential electrodeposition leads to films with a lower average oxidation state for Mn, these techniques do not provide information on the coordination geometry of Mn ions. Following electrolysis, the relative amounts of Mn4+, Mn3+, and Mn2+ at the surface of films grown in the dark and under illumination are similar (based on XPS). However, the films grown under illumination remain active for the OER, while films grown in the dark have become inactive. SEM imaging, Raman spectroscopy, EIS, Tafel plots, and measurements of the active surface area all indicate that the OER causes different degrees of structural and morphological changes in these samples. The high amount of disorder for MnOx films synthesized by multipotential electrodeposition makes it difficult to quantify the differences in structure that enhance the stability of films grown under illumination. The films do not show any peaks by X-ray diffraction. While there are differences in the extended X-ray absorption fine structure spectra for MnOx films grown in the dark and under illumination, we have not yet been able to fit a satisfactory model to explain these differences (likely because the Mn ions exhibit a range of different local coordination geometries in the disordered films). While the peaks are relatively weak, the Raman spectra indicate that the major differences in the samples are the spacing of the MnOx layers and the structure of intercalated species between these layers.
A mechanism consistent with the sum of our electrochemical and structural data and with prior literature is that photoreduction of Mn4+ ions during growth of the MnOx films followed by their partial oxidation during the OER induces a different coordination geometry for catalytically active Mn sites compared with films grown in the absence of illumination. The multipotential method we used for electrodepositing MnOx films is based on a procedure developed by Nocera and coworkers. (31,32) They proposed that Mn3+ ions introduced by the comproportionation between MnO2 and Mn(OH)2 were trapped in tetrahedral coordination sites that stabilized them against disproportionation during the OER. On the other hand, Marafatto and coworkers used time-resolved X-ray absorption spectroscopy to follow the photoreduction of Mn4+ ions in birnessite MnO2 particles to Mn3+. (51) Following photoreduction, the Jahn–Teller distortion of Mn3+ ions causes them to move from within the layer of edge-sharing MnO6 octahedra to sit above the layer where they share corner oxygen atoms with three MnO6 octahedra within the layer. Finally, Deibert and coworkers studied the structural changes that occur in a metal–organic framework (MOF) containing Mn2+ ions during photochemical water oxidation. (37) In the initial stage of photoexcitation, Mn2+ ions are oxidized, and the MOF transforms into a disordered birnessite composed of Mn4+, Mn3+, and Mn2+ ions with intercalated H2O and Na+ ions. They proposed that Mn3+ ions occupy sites both within the layers of birnessite and in between them.
The concentration and distribution of Mn ions in a reduced oxidation state will affect the spacing between layers in birnessite-like MnOx; Mn3+ and Mn2+ ions can be charge compensated through the intercalation of cations such as Na+ and K+. The additional peaks at 611 and 615 cm–1 (before and after electrolysis) in the Raman spectra of MnOx films grown under illumination, the appearance of a peak at 567 cm–1 for the illuminated film after electrolysis (Figure 5a), and the increase in the amount of intercalated K+ after electrolysis (Figure S11) support the hypothesis that the interlayer structure is different for these films compared to films grown in the dark. Previous reports have shown that the identity of intercalated cations (e.g., H+, Li+, Na+, K+, Rb+, Cs+, Mg2+, Ca2+, or Sr2+) in layered manganese oxides affects their activity for water oxidation. (15,16,22) Furthermore, there is evidence that water oxidation can occur in between the layers of birnessite MnO2. (16,21) In MnO2 electrodes used for Li+-ion and Na+-ion batteries, the interlayer spacing and structure of interlayer water strongly affect the ability to reversibly cycle between different oxidation states of Mn while intercalating/de-intercalating Li+ or Na+ ions. (65−67) Our results (Figure S11) and previous reports (68,69) both indicate that potassium ion intercalation/de-intercalation occurs during the OER. While we have not yet elucidated the exact structure of MnOx films grown under illumination, it appears that photoreduction of Mn ions during growth and their subsequent oxidation during the OER changes the interlayer structure (possibly through incorporating more corner-sharing, interlayer Mn3+ ions that are active sites for the OER). The resulting structure (not accessible to films grown in the dark or at constant potential) stabilizes the illuminated MnOx films during the OER by facilitating the transport of intercalated species (e.g., H+, H2O, and K+), while interlayer Mn sites catalyze water oxidation in between disordered layers of MnOx.

Conclusions

Click to copy section linkSection link copied!

We demonstrate how the combination of multiple external stimuli (i.e., light and an external potential) can be used to direct the nanoscale structure of manganese oxide films. MnOx films synthesized by multipotential electrodeposition under illumination with a 405 nm LED exhibit higher activity and stability toward OER relative to samples synthesized by multipotential deposition in the dark or at a constant potential under illumination. The nanoplate morphology of the MnOx films only degrades during the OER for the films grown in the dark, which decreases both their active surface area and conductivity. Changes in the interlayer structure of disordered MnOx films induced by illumination during growth appear to facilitate the intercalation and de-intercalation of species (e.g., H+, K+, and H2O) during the OER, which provides a stable electrocatalyst. The interlayer structure could be further modified through introducing other metal cations during synthesis (e.g., Rb+ or Cs+) to change the interlayer spacing. (22) This strategy of photoexcitation during electrochemical cycling could also be applied to other semiconductor metal oxides that are active for water oxidation, such as nickel oxide.

Experimental Section

Click to copy section linkSection link copied!

Materials

All chemicals were used as received. Manganese chloride tetrahydrate (MnCl2·4H2O, ≥99%), manganese(IV) oxide (MnO2, ≥99%), manganese(II, III) oxide (Mn3O4, ≥97%), manganese(III) oxide (Mn2O3, ≥99%), manganese(II) oxide (MnO, ≥99%), sodium hydroxide (NaOH, ≥97%), potassium hydroxide (KOH, ≥85%), potassium nitrate (KNO3, ≥99%), potassium permanganate (KMnO4, ≥99%), manganese carbonate (MnCO3, ≥99.9%, trace metal basis), and manganese nitrate tetrahydrate (Mn(NO3)2·4H2O, ≥97.0%) were purchased from Sigma-Aldrich/Millipore Sigma. Platinum gauze (Pt, 100 mesh, 99.9% metals basis), Pt wire (0.5 mm diameter, 99.95% metals basis), and d-(+)-glucose (anhydrous, ≥99%) were purchased from Alfa Aesar. All aqueous solutions were prepared using purified water from a GenPure Pro water purification system with a resistivity of 18.2 MΩ·cm.

Electrochemical Synthesis of MnOx Films

Preparation of Electrodes

The counter electrode for all electrochemical experiments consisted of a Pt gauze attached to a Pt wire. The Pt wire was soldered to a tinned copper wire, and the soldered joint was sealed in a glass tube. A silver/silver chloride (Ag/AgCl) electrode in 3 M NaCl was used as the reference electrode. Glass slides coated with FTO from MSE Supplies with a resistivity of 15 Ω/sq were cut into smaller pieces to be used as working electrodes with typical lateral dimensions of 1 cm × 2 cm. Each FTO slide was rinsed with purified water and then dried using a stream of nitrogen gas before use. The FTO electrodes were connected to the potentiostat with a copper clip.

Electrodeposition of MnOx Films

A custom cell was used for all electrochemical experiments. The cell was made from borosilicate glass with 14/20 ground-glass joints at the top to insert the electrodes and a flat window on the side for illumination. A BioLogic VSP-300 potentiostat/galvanostat operated using EC-Lab software (V11.33) was used to perform all electrochemical experiments. A photo scanner was used to scan images of the working electrodes, and their geometric areas were measured using ImageJ software. The MnOx films were electrodeposited from an aqueous solution containing 0.5 mM MnCl2 and 0.9 M KNO3 based on a previous report with modifications. (31) The pH of the deposition solution was adjusted to 8 using a diluted solution of KOH before use. The MnOx films were prepared by multipotential deposition, where the working electrode (i.e., FTO-coated glass) was held at 0.9 V versus Ag/AgCl for 3 s, followed by −0.6 V versus Ag/AgCl for 2 s; this cycle was repeated for either 40 min or 60 min. Variations of this deposition protocol are described in the Supporting Information and Figure S1. The electrodeposition was performed either in the dark or under illumination with a high-power, LED with a peak wavelength of 405 nm and a full width at half maximum of ∼30 nm. An aspheric condenser lens (Thor, ACL3026) was used to collimate the light. The FTO electrode faced the window of the electrochemical cell where it was irradiated by the 405 nm LED. The irradiance was measured using a calibrated Si photodiode from ThorLabs (FDS-100 CAL), which was placed at the same position as the working electrode prior to deposition. A typical current reading was 2 mA, which based on the responsivity and area (3.6 × 3.6 mm) of the photodiode gives an irradiance of 173 mW/cm2. For comparison, a MnO2 film was also deposited at a constant potential of 0.9 V versus Ag/AgCl for 20 min.

Electrochemical Measurements

Cyclic Voltammetry

CV was performed to characterize the onset potential and current density for oxygen evolution when using MnOx films as the working electrode. All CV scans started at the open-circuit potential. For each scan, the potential was first swept in the positive direction to 1.2 V versus Ag/AgCl and then swept in the negative direction to 0 V versus Ag/AgCl. The scan rate was 20 mV/s, and two consecutive CV scans were performed for each electrode to observe the change in current density. CV scans are plotted using IUPAC convention. The onset potential for the OER using different MnOx films was calculated by drawing tangent lines in the non-Faradaic and Faradaic regions of the CV scan. The abscissa of the point where the two lines intersect indicates the onset potential. In this work, the CV scans were referenced to the Ag/AgCl half-cell, which can be converted to the RHE scale by the following relation: ERHE = EAg/AgCl + 0.197 V + 0.059pH. The overpotential, η, can be calculated as η = ERHE – 1.23 V.

Electrode Stability

The stabilities of the MnOx films for the OER were evaluated by constant-current electrolysis. The MnOx films were tested at various current densities (e.g., 3, 5, and 10 mA/cm2) in a solution containing 0.1 M KOH and 0.9 M KNO3 as the supporting electrolyte. The solutions were stirred at ∼600 rpm throughout the experiment using a magnetic stir plate and stir bar. The potential was measured every 0.5 s, and the electrolysis times ranged from 40 min to 18 h. The electrode stability was analyzed by tracking the change of potential over time. A CV scan was taken after each stability test.

Electrochemical Impedance Spectroscopy

EIS was performed to determine the impedance and capacitance of the MnOx films before and after constant-current electrolysis. Mott–Schottky plots (i.e., 1/C2 vs the applied potential, where C is the space-charge capacitance) and Nyquist plots were extracted from EIS data and were fit using EC-Lab software (V11.33). Impedance measurements were conducted in a solution containing 0.1 M KOH and 0.9 M KNO3. For Mott–Schottky plots, the AC potential oscillation was 10 mV, and the DC potential range was from 1.0 to 0.2 V versus Ag/AgCl. The frequencies scanned ranged from 1 to 10 kHz. The data obtained at 10 kHz are shown in Figure 3b, and data obtained at other frequencies are shown in Figure S6 in the Supporting Information. No iR compensation was applied during these measurements. For Nyquist plots, the frequencies scanned ranged from 30 Hz to 3 MHz, the AC potential oscillation was 10 mV, and the DC potential range was 0.1 to 0.4 V versus Ag/AgCl. Only the data at a potential of 0.2 V are shown in Figure 3. A modified Randles circuit was used to model the electrochemical cell based on the fitting of Nyquist plots at different applied potentials. This circuit model has been previously used to describe electrodeposited MnOx films. (31) It consists of two resistors (R1 and R2), a capacitor (C2), and a Warburg impedance (W2) element: [R1 + C2/(R2 + W2)], as shown in Figure 3c.

Tafel Plots

To compare the activities of different MnOx films for the OER, we quantified the relationship between the steady-state current density and the applied potential during oxygen evolution by measuring their Tafel slopes. Tafel plots were tested in an aqueous solution containing 0.1 M KOH and 0.9 M KNO3 as the supporting electrolyte following previously published protocols. (30,31) The solutions were stirred at ∼500 rpm with a magnetic stir bar and stir plate to minimize the effect of mass-transport during the measurements. To limit the pseudocapacitance, each electrode was first held at 1.05 V versus Ag/AgCl for 200 s. A series of decreasing potentials, ranging from 1.05 to 0.95 V versus Ag/AgCl with an interval of 0.01 V, was then applied with each potential lasting for 30 s. The potentials were automatically iR compensated during the measurements by the EC-Lab software for the solution and substrate resistances. The applied potential versus the log of the current density was used to calculate the Tafel slope for each sample. Bare FTO slides exhibited Tafel slopes that were over 500 mV/decade and current densities that were at least 2 orders of magnitude lower than those of the MnOx films deposited on FTO, indicating that the substrate by itself is a poor catalyst for the OER.

Electrochemically Active Surface Area

Differences in the ECSAs of the MnOx films were determined from their electrochemical double-layer capacitance. (52) CV scans were measured over a potential window in which there was no Faradaic current in an aqueous solution containing 0.1 M KOH and 0.9 M KNO3. The potential window used was ±0.05 V with respect to the open-circuit potential of the cell using the MnOx film as the working electrode. CV scans were conducted by sweeping the potential within this window at eight different scan rates: 5, 25, 50, 75, 100, 125, 150, and 175 mV/s. At each vertex potential, the working electrode was held for 10 s before reversing the sweep direction.

Structural Characterization

Scanning Electron Microscopy

The morphologies of the MnOx films were characterized by using a ThermoFisher Quattro S environmental SEM operated at an acceleration voltage of 10 kV. The conductive top surface of each substrate was connected to the SEM sample holder with copper tape to avoid charge buildup. To compare films before and after constant-current electrolysis, the as-synthesized MnOx film on FTO-coated glass was cut into two smaller pieces of approximately 1 cm × 1 cm. One piece of the sample was imaged by SEM and the other underwent constant-current electrolysis as described in the manuscript before imaging.

X-ray Absorption Near-Edge Spectroscopy

X-ray absorption spectra at the Mn K-edge were collected in fluorescence mode using a four-element Vortex detector at beamline 10-BM-A,B of the Advanced Photon Source at Argonne National Laboratory. The XANES data were analyzed using Athena software. The first derivative of each spectrum was calculated to determine the edge energy of the peak. Manganese powders purchased from Millipore Sigma (MnO, Mn2O3, and MnO2) were used as references for the edge energies of Mn in different oxidation states. A linear fit was made to the edge energies and oxidation states of the reference manganese oxide powders. The average oxidation states of Mn in the electrodeposited MnOx films were determined through comparison of their edge energies to this linear fit.

X-ray Photoelectron Spectroscopy

XPS was performed by using a Physical Electronics 5000 VersaProbe II Scanning ESCA Microprobe system with a base pressure below 1 × 10–9 Torr. XPS data were acquired using the 1486.6 eV line from a monochromatic Al Kα source at 150 W with a spherical capacitor analyzer set to a pass energy of 23.5 eV for the high-resolution scans. The step size was 0.05 eV. High-resolution spectra in the binding regions for Mn 2p, Mn 3s, Mn 3p, C 1s, and K 2p electrons were measured for the MnOx films before and after constant-current electrolysis. Manganese powders prepared by hydrothermal synthesis or calcination (MnO2, Mn2O3, Mn3O4) were used as references for the peak positions of Mn in different oxidation states.
To monitor changes in the average Mn oxidation state of the MnOx films, we fit the Mn 2p3/2 peak using the method proposed by Nesbitt and Banerjee. (54) The binding energies of the MnOx films were first corrected based on the C 1s peak. A Shirley background was used in the binding energy region for Mn 2p3/2 electrons. Each oxidation state of Mn was decomposed into a multiplet of five different peaks based on calculations by Gupta and Sen, (55,56) who accounted for coupling between unpaired valence electrons in high-spin manganese and unpaired core electrons (once photoionization has taken place). The peak positions and the ratios of peak areas were constrained based on the fitting parameters used in prior work. (20,37,54,57) Each peak was fit by using a 70% Gaussian and 30% Lorentzian peak shape. For the reference manganese oxide powders, the full width at half maximum (fwhm) for each peak in the multiplet was optimized to reduce the residual standard deviation (see Table S2). The fwhm for peaks in the electrodeposited films was set to 1.25 eV to compare the distribution of oxidation states. For each oxidation state of Mn, we allowed the peak position to vary by up to 0.3 eV from the reference value. For example, when fitting the five peaks for Mn3+ multiplet, the binding energy of the peak with the lowest binding energy was 640.7 ± 0.3 eV, and the binding energies for the other four peaks were fixed based on their relative position to the peak with the lowest binding energy. The binding energies of the peaks with the lowest binding energy among the multiplets for Mn2+ and Mn4+ were 640.0 ± 0.3 eV and 641.9 ± 0.3 eV, respectively. The area for the peak with the lowest binding energy was allowed to vary freely to obtain the lowest residual standard deviation. The areas for the other four peaks were constrained based on their relative intensity to the peak with the lowest binding energy.

Raman Spectroscopy

Raman spectroscopy was used to characterize the structure of the MnOx films with an InVia Raman microscope (Renishaw, UK). The excitation source was a 514 nm laser (∼4 mW), and a grating of 1800 lines/mm was used. A 20× Leica objective and a reduced power of 50% were utilized. The integration time for each measurement was 10 s, and the step size was 1.2 cm–1. Raman spectra of MnOx films grown on FTO substrates in the dark and under illumination were measured before and after constant-current electrolysis.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsnanoscienceau.3c00002.

  • Procedures used to synthesize manganese oxide powders and obtain their diffraction patterns, variations of the square-wave potential used to electrodeposit MnOx films, procedure for obtaining the double-layer capacitances and ECSAs, and characterization of electrodeposited MnOx films and reference powders by XANES; summary of electrochemical characterization of MnOx films and the parameters used to fit the X-ray photoelectron spectra of different MnOx films and reference powders; additional characterization by linear sweep voltammetry and constant-current electrolysis to measure the stability of MnOx films, CV used to measure the double-layer capacitances of the MnOx films, Mott–Schottky plots of MnOx films at different frequencies, X-ray absorption near edge spectroscopy and linear fitting of XANES data, Raman spectrum of a MnO2 film grown at constant potential, and XPS of electrodeposited MnOx films in the region for K 2p electrons (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Bryce Sadtler - Department of Chemistry, Washington University, St. Louis, Missouri 63130, United StatesInstitute of Materials Science & Engineering, Washington University, St. Louis, Missouri 63130, United StatesOrcidhttps://orcid.org/0000-0003-4860-501X Email: [email protected]
  • Authors
  • Author Contributions

    CRediT: Chu Qin conceptualization (supporting), formal analysis (lead), investigation (lead), writing-original draft (equal); Jiang Luo formal analysis (supporting), investigation (supporting), writing-original draft (supporting); Dongyan Zhang formal analysis (supporting), investigation (supporting); Logan Brennan investigation (supporting); shijun Tian formal analysis (supporting); Ashlynn Berry investigation (supporting); Brandon M. Campbell investigation (supporting); Bryce Sadtler conceptualization (lead), formal analysis (supporting), supervision (lead), writing-original draft (equal).

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This material is based upon work supported by the National Science Foundation (NSF) under grant no. CHE-1753344 to B.S. This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science user facility operated for the DOE Office of Science by Argonne National Laboratory under Contract no DE-AC02-06CH11357. B.M.C. acknowledges support from the MARC U-STAR program at Washington University. Electron microscopy and XPS were performed at the Institute of Materials Science & Engineering at Washington University. X-ray diffraction was performed in the Department of Earth and Planetary Sciences at Washington University. The authors thank S. Singamaneni for use of his Raman spectrometer and M. Warren for assistance in acquiring and interpreting the X-ray absorption spectra.

References

Click to copy section linkSection link copied!

This article references 69 other publications.

  1. 1
    Lewis, N. S.; Nocera, D. G. Powering the Planet: Chemical Challenges in Solar Energy Utilization. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 1572915735,  DOI: 10.1073/pnas.0603395103
  2. 2
    Tahir, M.; Pan, L.; Idrees, F.; Zhang, X.; Wang, L.; Zou, J.-J.; Wang, Z. L. Electrocatalytic Oxygen Evolution Reaction for Energy Conversion and Storage: A Comprehensive Review. Nano Energy 2017, 37, 136157,  DOI: 10.1016/j.nanoen.2017.05.022
  3. 3
    Vesborg, P. C. K.; Jaramillo, T. F. Addressing the Terawatt Challenge: Scalability in the Supply of Chemical Elements for Renewable Energy. RSC Adv. 2012, 2, 79337947,  DOI: 10.1039/c2ra20839c
  4. 4
    Gaultois, M. W.; Sparks, T. D.; Borg, C. K. H.; Seshadri, R.; Bonificio, W. D.; Clarke, D. R. Data-Driven Review of Thermoelectric Materials: Performance and Resource Considerations. Chem. Mater. 2013, 25, 29112920,  DOI: 10.1021/cm400893e
  5. 5
    Pinaud, B. A.; Benck, J. D.; Seitz, L. C.; Forman, A. J.; Chen, Z.; Deutsch, T. G.; James, B. D.; Baum, K. N.; Baum, G. N.; Ardo, S.; Wang, H.; Miller, E.; Jaramillo, T. F. Technical and Economic Feasibility of Centralized Facilities for Solar Hydrogen Production Via Photocatalysis and Photoelectrochemistry. Energy Environ. Sci. 2013, 6, 19832002,  DOI: 10.1039/c3ee40831k
  6. 6
    Trotochaud, L.; Ranney, J. K.; Williams, K. N.; Boettcher, S. W. Solution-Cast Metal Oxide Thin Film Electrocatalysts for Oxygen Evolution. J. Am. Chem. Soc. 2012, 134, 1725317261,  DOI: 10.1021/ja307507a
  7. 7
    Sun, K.; Park, N.; Sun, Z.; Zhou, J.; Wang, J.; Pang, X.; Shen, S.; Noh, S. Y.; Jing, Y.; Jin, S.; Yu, P. K. L.; Wang, D. Nickel Oxide Functionalized Silicon for Efficient Photo-Oxidation of Water. Energy Environ. Sci. 2012, 5, 78727877,  DOI: 10.1039/c2ee21708b
  8. 8
    Kenney, M. J.; Gong, M.; Li, Y.; Wu, J. Z.; Feng, J.; Lanza, M.; Dai, H. High-Performance Silicon Photoanodes Passivated with Ultrathin Nickel Films for Water Oxidation. Science 2013, 342, 836840,  DOI: 10.1126/science.1241327
  9. 9
    Zou, S.; Burke, M. S.; Kast, M. G.; Fan, J.; Danilovic, N.; Boettcher, S. W. Fe (Oxy)Hydroxide Oxygen Evolution Reaction Electrocatalysis: Intrinsic Activity and the Roles of Electrical Conductivity, Substrate, and Dissolution. Chem. Mater. 2015, 27, 80118020,  DOI: 10.1021/acs.chemmater.5b03404
  10. 10
    Wang, H.; Lee, H.-W.; Deng, Y.; Lu, Z.; Hsu, P.-C.; Liu, Y.; Lin, D.; Cui, Y. Bifunctional Non-Noble Metal Oxide Nanoparticle Electrocatalysts through Lithium-Induced Conversion for Overall Water Splitting. Nat. Commun. 2015, 6, 7261,  DOI: 10.1038/ncomms8261
  11. 11
    Hill, J. C.; Landers, A. T.; Switzer, J. A. An Electrodeposited Inhomogeneous Metal–Insulator–Semiconductor Junction for Efficient Photoelectrochemical Water Oxidation. Nat. Mater. 2015, 14, 11501155,  DOI: 10.1038/nmat4408
  12. 12
    Jiao, F.; Frei, H. Nanostructured Cobalt and Manganese Oxide Clusters as Efficient Water Oxidation Catalysts. Energy Environ. Sci. 2010, 3, 10181027,  DOI: 10.1039/c002074e
  13. 13
    Hocking, R. K.; Brimblecombe, R.; Chang, L.-Y.; Singh, A.; Cheah, M. H.; Glover, C.; Casey, W. H.; Spiccia, L. Water-Oxidation Catalysis by Manganese in a Geochemical-Like Cycle. Nat. Chem. 2011, 3, 461466,  DOI: 10.1038/nchem.1049
  14. 14
    Iyer, A.; Del-Pilar, J.; King’ondu, C. K.; Kissel, E.; Garces, H. F.; Huang, H.; El-Sawy, A. M.; Dutta, P. K.; Suib, S. L. Water Oxidation Catalysis Using Amorphous Manganese Oxides, Octahedral Molecular Sieves (OMS-2), and Octahedral Layered (OL-1) Manganese Oxide Structures. J. Phys. Chem. C 2012, 116, 64746483,  DOI: 10.1021/jp2120737
  15. 15
    Wiechen, M.; Zaharieva, I.; Dau, H.; Kurz, P. Layered Manganese Oxides for Water-Oxidation: Alkaline Earth Cations Influence Catalytic Activity in a Photosystem II-Like Fashion. Chem. Sci. 2012, 3, 23302339,  DOI: 10.1039/c2sc20226c
  16. 16
    Boppana, V. B. R.; Yusuf, S.; Hutchings, G. S.; Jiao, F. Nanostructured Alkaline-Cation-Containing δ-MnO2 for Photocatalytic Water Oxidation. Adv. Funct. Mater. 2013, 23, 878884,  DOI: 10.1002/adfm.201202141
  17. 17
    Robinson, D. M.; Go, Y. B.; Mui, M.; Gardner, G.; Zhang, Z.; Mastrogiovanni, D.; Garfunkel, E.; Li, J.; Greenblatt, M.; Dismukes, G. C. Photochemical Water Oxidation by Crystalline Polymorphs of Manganese Oxides: Structural Requirements for Catalysis. J. Am. Chem. Soc. 2013, 135, 34943501,  DOI: 10.1021/ja310286h
  18. 18
    Pokhrel, R.; Goetz, M. K.; Shaner, S. E.; Wu, X.; Stahl, S. S. The “Best Catalyst” for Water Oxidation Depends on the Oxidation Method Employed: A Case Study of Manganese Oxides. J. Am. Chem. Soc. 2015, 137, 83848387,  DOI: 10.1021/jacs.5b05093
  19. 19
    Frey, C. E.; Kurz, P. Water Oxidation Catalysis by Synthetic Manganese Oxides with Different Structural Motifs: A Comparative Study. Chem.─Eur. J. 2015, 21, 1495814968,  DOI: 10.1002/chem.201501367
  20. 20
    Smith, P. F.; Deibert, B. J.; Kaushik, S.; Gardner, G.; Hwang, S.; Wang, H.; Al-Sharab, J. F.; Garfunkel, E.; Fabris, L.; Li, J.; Dismukes, G. C. Coordination Geometry and Oxidation State Requirements of Corner-Sharing MnO6 Octahedra for Water Oxidation Catalysis: An Investigation of Manganite (γ-MnOOH). ACS Catal. 2016, 6, 20892099,  DOI: 10.1021/acscatal.6b00099
  21. 21
    Thenuwara, A. C.; Cerkez, E. B.; Shumlas, S. L.; Attanayake, N. H.; McKendry, I. G.; Frazer, L.; Borguet, E.; Kang, Q.; Remsing, R. C.; Klein, M. L.; Zdilla, M. J.; Strongin, D. R. Nickel Confined in the Interlayer Region of Birnessite: An Active Electrocatalyst for Water Oxidation. Angew. Chem., Int. Ed. 2016, 55, 1038110385,  DOI: 10.1002/anie.201601935
  22. 22
    Kang, Q.; Vernisse, L.; Remsing, R. C.; Thenuwara, A. C.; Shumlas, S. L.; McKendry, I. G.; Klein, M. L.; Borguet, E.; Zdilla, M. J.; Strongin, D. R. Effect of Interlayer Spacing on the Activity of Layered Manganese Oxide Bilayer Catalysts for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2017, 139, 18631870,  DOI: 10.1021/jacs.6b09184
  23. 23
    Kölbach, M.; Fiechter, S.; van de Krol, R.; Bogdanoff, P. Evaluation of Electrodeposited α-Mn2O3 as a Catalyst for the Oxygen Evolution Reaction. Catal. Today 2017, 290, 29,  DOI: 10.1016/j.cattod.2017.03.030
  24. 24
    Li, A.; Ooka, H.; Bonnet, N.; Hayashi, T.; Sun, Y.; Jiang, Q.; Li, C.; Han, H.; Nakamura, R. Stable Potential Windows for Long-Term Electrocatalysis by Manganese Oxides under Acidic Conditions. Angew. Chem., Int. Ed. 2019, 58, 50545058,  DOI: 10.1002/anie.201813361
  25. 25
    Takashima, T.; Hashimoto, K.; Nakamura, R. Mechanisms of pH-Dependent Activity for Water Oxidation to Molecular Oxygen by MnO2 Electrocatalysts. J. Am. Chem. Soc. 2012, 134, 15191527,  DOI: 10.1021/ja206511w
  26. 26
    Takashima, T.; Hashimoto, K.; Nakamura, R. Inhibition of Charge Disproportionation of MnO2 Electrocatalysts for Efficient Water Oxidation under Neutral Conditions. J. Am. Chem. Soc. 2012, 134, 1815318156,  DOI: 10.1021/ja306499n
  27. 27
    Zaharieva, I.; Chernev, P.; Risch, M.; Klingan, K.; Kohlhoff, M.; Fischer, A.; Dau, H. Electrosynthesis, Functional, and Structural Characterization of a Water-Oxidizing Manganese Oxide. Energy Environ. Sci. 2012, 5, 70817089,  DOI: 10.1039/c2ee21191b
  28. 28
    Ramírez, A.; Hillebrand, P.; Stellmach, D.; May, M. M.; Bogdanoff, P.; Fiechter, S. Evaluation of MnOx, Mn2O3, and Mn3O4 Electrodeposited Films for the Oxygen Evolution Reaction of Water. J. Phys. Chem. C 2014, 118, 1407314081,  DOI: 10.1021/jp500939d
  29. 29
    Gorlin, Y.; Lassalle-Kaiser, B.; Benck, J. D.; Gul, S.; Webb, S. M.; Yachandra, V. K.; Yano, J.; Jaramillo, T. F. In Situ X-Ray Absorption Spectroscopy Investigation of a Bifunctional Manganese Oxide Catalyst with High Activity for Electrochemical Water Oxidation and Oxygen Reduction. J. Am. Chem. Soc. 2013, 135, 85258534,  DOI: 10.1021/ja3104632
  30. 30
    Huynh, M.; Bediako, D. K.; Nocera, D. G. A Functionally Stable Manganese Oxide Oxygen Evolution Catalyst in Acid. J. Am. Chem. Soc. 2014, 136, 60026010,  DOI: 10.1021/ja413147e
  31. 31
    Huynh, M.; Shi, C.; Billinge, S. J. L.; Nocera, D. G. Nature of Activated Manganese Oxide for Oxygen Evolution. J. Am. Chem. Soc. 2015, 137, 1488714904,  DOI: 10.1021/jacs.5b06382
  32. 32
    Morgan Chan, Z.; Kitchaev, D. A.; Nelson Weker, J.; Schnedermann, C.; Lim, K.; Ceder, G.; Tumas, W.; Toney, M. F.; Nocera, D. G. Electrochemical Trapping of Metastable Mn3+ Ions for Activation of MnO2 Oxygen Evolution Catalysts. Proc. Natl. Acad. Sci. U.S.A. 2018, 115, E5261E5268,  DOI: 10.1073/pnas.1722235115
  33. 33
    Chinnadurai, D.; Nallal, M.; Kim, H.-J.; Li, O. L.; Park, K. H.; Prabakar, K. Mn3+ Active Surface Site Enriched Manganese Phosphate Nano-Polyhedrons for Enhanced Bifunctional Oxygen Electrocatalyst. ChemCatChem 2020, 12, 23482355,  DOI: 10.1002/cctc.202000164
  34. 34
    Shao, C.; Yin, K.; Liao, F.; Zhu, W.; Shi, H.; Shao, M. Rod-Shaped α-MnO2 Electrocatalysts with High Mn3+ Content for Oxygen Reduction Reaction and Zn-Air Battery. J. Alloys Compd. 2021, 860, 158427,  DOI: 10.1016/j.jallcom.2020.158427
  35. 35
    Melder, J.; Mebs, S.; Heizmann, P. A.; Lang, R.; Dau, H.; Kurz, P. Carbon Fibre Paper Coated by a Layered Manganese Oxide: A Nano-Structured Electrocatalyst for Water-Oxidation with High Activity over a Very Wide pH Range. J. Mater. Chem. A 2019, 7, 2533325346,  DOI: 10.1039/c9ta08804k
  36. 36
    Melder, J.; Bogdanoff, P.; Zaharieva, I.; Fiechter, S.; Dau, H.; Kurz, P. Water-Oxidation Electrocatalysis by Manganese Oxides: Syntheses, Electrode Preparations, Electrolytes and Two Fundamental Questions. Z. Phys. Chem. 2020, 234, 925978,  DOI: 10.1515/zpch-2019-1491
  37. 37
    Deibert, B. J.; Zhang, J.; Smith, P. F.; Chapman, K. W.; Rangan, S.; Banerjee, D.; Tan, K.; Wang, H.; Pasquale, N.; Chen, F.; Lee, K.-B.; Dismukes, G. C.; Chabal, Y. J.; Li, J. Surface and Structural Investigation of a MnOx Birnessite-Type Water Oxidation Catalyst Formed under Photocatalytic Conditions. Chem.─Eur. J. 2015, 21, 1421814228,  DOI: 10.1002/chem.201501930
  38. 38
    Sadtler, B.; Burgos, S. P.; Batara, N. A.; Beardslee, J. A.; Atwater, H. A.; Lewis, N. S. Phototropic Growth Control of Nanoscale Pattern Formation in Photoelectrodeposited Se–Te Films. Proc. Natl. Acad. Sci. U.S.A. 2013, 110, 1970719712,  DOI: 10.1073/pnas.1315539110
  39. 39
    Carim, A. I.; Batara, N. A.; Premkumar, A.; Atwater, H. A.; Lewis, N. S. Self-Optimizing Photoelectrochemical Growth of Nanopatterned Se–Te Films in Response to the Spectral Distribution of Incident Illumination. Nano Lett. 2015, 15, 70717076,  DOI: 10.1021/acs.nanolett.5b03137
  40. 40
    Lowe, J. M.; Yan, Q.; Benamara, M.; Coridan, R. H. Direct Photolithographic Patterning of Cuprous Oxide Thin Films Via Photoelectrodeposition. J. Mater. Chem. A 2017, 5, 2176521772,  DOI: 10.1039/c7ta05321e
  41. 41
    Tan, C.; Qin, C.; Sadtler, B. Light-Directed Growth of Metal and Semiconductor Nanostructures. J. Mater. Chem. C 2017, 5, 56285642,  DOI: 10.1039/c7tc00379j
  42. 42
    Carim, A. I.; Hamann, K. R.; Batara, N. A.; Thompson, J. R.; Atwater, H. A.; Lewis, N. S. Template-Free Synthesis of Periodic Three-Dimensional PbSe Nanostructures via Photoelectrodeposition. J. Am. Chem. Soc. 2018, 140, 65366539,  DOI: 10.1021/jacs.8b02931
  43. 43
    Meier, M. C.; Cheng, W.-H.; Atwater, H. A.; Lewis, N. S.; Carim, A. I. Inorganic Phototropism in Electrodeposition of Se–Te. J. Am. Chem. Soc. 2019, 141, 1865818661,  DOI: 10.1021/jacs.9b10579
  44. 44
    Qin, C.; Campbell, B. M.; Shen, M.; Zhao, T.; Sadtler, B. Light-Driven, Facet-Selective Transformation of Cuprous Oxide Microcrystals to Hollow Copper Nanoshells. Chem. Mater. 2019, 31, 80008011,  DOI: 10.1021/acs.chemmater.9b02240
  45. 45
    Sakai, N.; Ebina, Y.; Takada, K.; Sasaki, T. Photocurrent Generation from Semiconducting Manganese Oxide Nanosheets in Response to Visible Light. J. Phys. Chem. B 2005, 109, 96519655,  DOI: 10.1021/jp0500485
  46. 46
    Pinaud, B. A.; Chen, Z.; Abram, D. N.; Jaramillo, T. F. Thin Films of Sodium Birnessite-Type MnO2: Optical Properties, Electronic Band Structure, and Solar Photoelectrochemistry. J. Phys. Chem. C 2011, 115, 1183011838,  DOI: 10.1021/jp200015p
  47. 47
    Hsu, Y.-K.; Chen, Y.-C.; Lin, Y.-G.; Chen, L.-C.; Chen, K.-H. Birnessite-Type Manganese Oxides Nanosheets with Hole Acceptor Assisted Photoelectrochemical Activity in Response to Visible Light. J. Mater. Chem. 2012, 22, 27332739,  DOI: 10.1039/c1jm14355g
  48. 48
    Sunda, W. G.; Huntsman, S. A. Photoreduction of Manganese Oxides in Seawater. Mar. Chem. 1994, 46, 133152,  DOI: 10.1016/0304-4203(94)90051-5
  49. 49
    Sherman, D. M. Electronic Structures of Iron(III) and Manganese(IV) (Hydr)Oxide Minerals: Thermodynamics of Photochemical Reductive Dissolution in Aquatic Environments. Geochim. Cosmochim. Acta 2005, 69, 32493255,  DOI: 10.1016/j.gca.2005.01.023
  50. 50
    Kwon, K. D.; Refson, K.; Sposito, G. On the Role of Mn(IV) Vacancies in the Photoreductive Dissolution of Hexagonal Birnessite. Geochim. Cosmochim. Acta 2009, 73, 41424150,  DOI: 10.1016/j.gca.2009.04.031
  51. 51
    Marafatto, F. F.; Strader, M. L.; Gonzalez-Holguera, J.; Schwartzberg, A.; Gilbert, B.; Peña, J. Rate and Mechanism of the Photoreduction of Birnessite (MnO2) Nanosheets. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 46004605,  DOI: 10.1073/pnas.1421018112
  52. 52
    McCrory, C. C. L.; Jung, S.; Peters, J. C.; Jaramillo, T. F. Benchmarking Heterogeneous Electrocatalysts for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 1697716987,  DOI: 10.1021/ja407115p
  53. 53
    Huynh, M.; Bediako, D. K.; Liu, Y.; Nocera, D. G. Nucleation and Growth Mechanisms of an Electrodeposited Manganese Oxide Oxygen Evolution Catalyst. J. Phys. Chem. C 2014, 118, 1714217152,  DOI: 10.1021/jp501768n
  54. 54
    Nesbitt, H. W.; Banerjee, D. Interpretation of XPS Mn(2p) Spectra of Mn Oxyhydroxides and Constraints on the Mechanism of MnO2 Precipitation. Am. Mineral. 1998, 83, 305315,  DOI: 10.2138/am-1998-3-414
  55. 55
    Gupta, R. P.; Sen, S. K. Calculation of Multiplet Structure of Core p-Vacancy Levels. Phys. Rev. B: Solid State 1974, 10, 7177,  DOI: 10.1103/physrevb.10.71
  56. 56
    Gupta, R. P.; Sen, S. K. Calculation of Multiplet Structure of Core p-Vacancy Levels. II. Phys. Rev. B: Solid State 1975, 12, 1519,  DOI: 10.1103/physrevb.12.15
  57. 57
    Biesinger, M. C.; Payne, B. P.; Grosvenor, A. P.; Lau, L. W. M.; Gerson, A. R.; Smart, R. S. C. Resolving Surface Chemical States in XPS Analysis of First Row Transition Metals, Oxides and Hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 27172730,  DOI: 10.1016/j.apsusc.2010.10.051
  58. 58
    Julien, C.; Massot, M.; Baddour-Hadjean, R.; Franger, S.; Bach, S.; Pereira-Ramos, J. P. Raman Spectra of Birnessite Manganese Dioxides. Solid State Ionics 2003, 159, 345356,  DOI: 10.1016/s0167-2738(03)00035-3
  59. 59
    Chen, D.; Ding, D.; Li, X.; Waller, G. H.; Xiong, X.; El-Sayed, M. A.; Liu, M. Probing the Charge Storage Mechanism of a Pseudocapacitive MnO2 Electrode Using in Operando Raman Spectroscopy. Chem. Mater. 2015, 27, 66086619,  DOI: 10.1021/acs.chemmater.5b03118
  60. 60
    Scheitenberger, P.; Euchner, H.; Lindén, M. The Hidden Impact of Structural Water – How Interlayer Water Largely Controls the Raman Spectroscopic Response of Birnessite-Type Manganese Oxide. J. Mater. Chem. A 2021, 9, 1846618476,  DOI: 10.1039/d1ta05357d
  61. 61
    Wu, T.-H.; Hesp, D.; Dhanak, V.; Collins, C.; Braga, F.; Hardwick, L. J.; Hu, C.-C. Charge Storage Mechanism of Activated Manganese Oxide Composites for Pseudocapacitors. J. Mater. Chem. A 2015, 3, 1278612795,  DOI: 10.1039/c5ta03334a
  62. 62
    Yang, L.; Cheng, S.; Ji, X.; Jiang, Y.; Zhou, J.; Liu, M. Investigations into the Origin of Pseudocapacitive Behavior of Mn3O4 Electrodes Using in Operando Raman Spectroscopy. J. Mater. Chem. A 2015, 3, 73387344,  DOI: 10.1039/c5ta00223k
  63. 63
    Boumaiza, H.; Renard, A.; Rakotomalala Robinson, M.; Kervern, G.; Vidal, L.; Ruby, C.; Bergaoui, L.; Coustel, R. A Multi-Technique Approach for Studying Na Triclinic and Hexagonal Birnessites. J. Solid State Chem. 2019, 272, 234243,  DOI: 10.1016/j.jssc.2019.02.017
  64. 64
    Scheitenberger, P.; Brimaud, S.; Lindén, M. XRD/Raman Spectroscopy Studies of the Mechanism of (De)Intercalation of Na+ from/into Highly Crystalline Birnessite. Mater. Adv. 2021, 2, 39403953,  DOI: 10.1039/d1ma00161b
  65. 65
    Nam, K. W.; Kim, S.; Yang, E.; Jung, Y.; Levi, E.; Aurbach, D.; Choi, J. W. Critical Role of Crystal Water for a Layered Cathode Material in Sodium Ion Batteries. Chem. Mater. 2015, 27, 37213725,  DOI: 10.1021/acs.chemmater.5b00869
  66. 66
    Lu, K.; Hu, Z.; Xiang, Z.; Ma, J.; Song, B.; Zhang, J.; Ma, H. Cation Intercalation in Manganese Oxide Nanosheets: Effects on Lithium and Sodium Storage. Angew. Chem., Int. Ed. 2016, 55, 1044810452,  DOI: 10.1002/anie.201605102
  67. 67
    Shan, X.; Guo, F.; Charles, D. S.; Lebens-Higgins, Z.; Abdel Razek, S.; Wu, J.; Xu, W.; Yang, W.; Page, K. L.; Neuefeind, J. C.; Feygenson, M.; Piper, L. F. J.; Teng, X. Structural Water and Disordered Structure Promote Aqueous Sodium-Ion Energy Storage in Sodium-Birnessite. Nat. Commun. 2019, 10, 4975,  DOI: 10.1038/s41467-019-12939-3
  68. 68
    Chigane, M.; Ishikawa, M. XRD and XPS Characterization of Electrochromic Nickel Oxide Thin Films Prepared by Electrolysis–Chemical Deposition. J. Chem. Soc., Faraday Trans. 1998, 94, 36653670,  DOI: 10.1039/a806459h
  69. 69
    Chigane, M.; Ishikawa, M. Manganese Oxide Thin Film Preparation by Potentiostatic Electrolyses and Electrochromism. J. Electrochem. Soc. 2000, 147, 2246,  DOI: 10.1149/1.1393515

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 4 publications.

  1. Yang Liu, Shiqing Zhang, Shaokai Ma, Xinyu Sun, Ying Wang, Fang Liu, Ying Li, Yuanhui Ma, Xuewen Xu, Yanming Xue, Chengchun Tang, Jun Zhang. Electronic Structure Modification of MnO2 Nanosheet Arrays with Enhanced Water Oxidation Activity and Stability by Nitrogen Plasma. ACS Applied Materials & Interfaces 2024, 16 (28) , 36498-36508. https://doi.org/10.1021/acsami.4c07973
  2. Masaharu Nakayama, Wataru Yoshida. Electrodeposited Manganese Dioxides and Their Composites as Electrocatalysts for Energy Conversion Reactions. ChemSusChem 2025, 18 (5) https://doi.org/10.1002/cssc.202401907
  3. Chu Qin, Shijun Tian, Jialong Wu, Junliang Mou, Lan Feng, Zhongqing Jiang. Plasma‐Assisted Surface Engineering to Stabilize Mn 3+ in Electrodeposited Manganese Oxide Films for Water Oxidation. ChemCatChem 2024, 16 (22) https://doi.org/10.1002/cctc.202401033
  4. Elisa Morales, Lauren Formanski, Shaner Sarah, Stone Kari. Characterizing Biogenic MnOx Produced by Pseudomonas putida MnB1 and Its Catalytic Activity towards Water Oxidation. Life 2024, 14 (2) , 171. https://doi.org/10.3390/life14020171

ACS Nanoscience Au

Cite this: ACS Nanosci. Au 2023, 3, 4, 310–322
Click to copy citationCitation copied!
https://doi.org/10.1021/acsnanoscienceau.3c00002
Published April 23, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

1738

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (a,b) CV scans (using IUPAC convention) of MnOx films measured in 0.1 M KOH and 0.9 M KNO3. (a) CV scans of the as-synthesized films grown under illumination (red trace) and in the dark (blue trace). (b) CV scans of the same illuminated film (orange trace) and dark film (light-blue trace) after 1 h of constant-current electrolysis at a current density of 10 mA/cm2. The arrows indicate the onset potentials for the OER. (c) Applied potential vs time for an illuminated MnOx film (red trace) and a dark film (blue trace) during constant-current electrolysis at a current density of 10 mA/cm2 to test the stability of the films during the OER.

    Figure 2

    Figure 2. SEM images of (a) as-synthesized MnOx film grown under illumination with a 405 nm LED, (b) MnOx film synthesized under illumination after 30 min of electrolysis at 3 mA/cm2, (c) as-synthesized MnOx film grown in the dark, and (d) MnOx film synthesized in the dark after 30 min of electrolysis at 3 mA/cm2. The scale bar of 1 μm applies to all images.

    Figure 3

    Figure 3. Electrochemical characterization of electrodeposited MnOx films. (a) Tafel plots of MnOx films synthesized under illumination before (red triangles) and after (orange circles) electrolysis at 5 mA/cm2 for 40 min and films synthesized in the dark before (blue triangles) and after (light-blue squares) electrolysis. The symbols represent the actual data, and the lines connecting the symbols represent linear fits of the data used to calculate Tafel slopes. The errors were determined by linear fitting, and the R2 values of the linear fits were greater than 0.995. (b) Mott–Schottky plots of MnOx films synthesized under illumination before (red triangles) and after (orange circles) electrolysis at 5 mA/cm2 for 1 h, and films synthesized in the dark before (blue triangles) and after (light-blue squares) electrolysis. The lines represent linear fits of the raw data used to calculate the Mott–Schottky slopes. (c) Nyquist plots of MnOx films synthesized under illumination before (red) and after (orange) electrolysis at 3 mA/cm2 for 1.5 h, and films synthesized in the dark before (blue) and after (light blue) electrolysis. The inset shows the equivalent circuit model used to fit the Nyquist plots. R1 represents the uncompensated resistance from the solution, FTO substrate, and MnOx film, C2 represents the double-layer capacitance, R2 represents the polarization resistance, and W2 represents a Warburg impedance.

    Figure 4

    Figure 4. XPS of electrodeposited MnOx films in the binding energy region of Mn 2p3/2 electrons. (a,b) MnOx films synthesized under illumination before (a) before and (b) after electrolysis at 5 mA/cm2 for 1 h (c,d) MnOx films synthesized in the dark (c) before and (d) after electrolysis. For each spectrum, the shaded regions show the deconvolution of the peak into contributions from Mn4+ (purple), Mn3+ (green), and Mn2+ (light red). The gray circles show the raw data, and the black line shows the sum of fitting the peak to the different oxidation states of Mn. The dashed line shows the residual of the fit. The peak maxima for each peak in the deconvolution are provided in Table S3 of the Supporting Information.

    Figure 5

    Figure 5. (a) Raman spectra of electrodeposited MnOx films. Red trace: as-synthesized MnOx film grown under illumination; orange trace: illuminated film after electrolysis at 5 mA/cm2 for 1 h; blue trace: as-synthesized MnOx film grown in the dark; light-blue trace: dark film after electrolysis at 5 mA/cm2 for 1 h. (b) Crystal structure of potassium-intercalated birnessite hydrate, K0.27(Mn0.98O2)(H2O)0.51. Mn atoms are pink, oxygen atoms are gray, and potassium atoms are purple. The oxygen atoms within the layer represent intercalated H2O (hydrogen atoms not shown). The structure was made using collection code no 55407 from the Inorganic Crystal Structure Database.

  • References


    This article references 69 other publications.

    1. 1
      Lewis, N. S.; Nocera, D. G. Powering the Planet: Chemical Challenges in Solar Energy Utilization. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 1572915735,  DOI: 10.1073/pnas.0603395103
    2. 2
      Tahir, M.; Pan, L.; Idrees, F.; Zhang, X.; Wang, L.; Zou, J.-J.; Wang, Z. L. Electrocatalytic Oxygen Evolution Reaction for Energy Conversion and Storage: A Comprehensive Review. Nano Energy 2017, 37, 136157,  DOI: 10.1016/j.nanoen.2017.05.022
    3. 3
      Vesborg, P. C. K.; Jaramillo, T. F. Addressing the Terawatt Challenge: Scalability in the Supply of Chemical Elements for Renewable Energy. RSC Adv. 2012, 2, 79337947,  DOI: 10.1039/c2ra20839c
    4. 4
      Gaultois, M. W.; Sparks, T. D.; Borg, C. K. H.; Seshadri, R.; Bonificio, W. D.; Clarke, D. R. Data-Driven Review of Thermoelectric Materials: Performance and Resource Considerations. Chem. Mater. 2013, 25, 29112920,  DOI: 10.1021/cm400893e
    5. 5
      Pinaud, B. A.; Benck, J. D.; Seitz, L. C.; Forman, A. J.; Chen, Z.; Deutsch, T. G.; James, B. D.; Baum, K. N.; Baum, G. N.; Ardo, S.; Wang, H.; Miller, E.; Jaramillo, T. F. Technical and Economic Feasibility of Centralized Facilities for Solar Hydrogen Production Via Photocatalysis and Photoelectrochemistry. Energy Environ. Sci. 2013, 6, 19832002,  DOI: 10.1039/c3ee40831k
    6. 6
      Trotochaud, L.; Ranney, J. K.; Williams, K. N.; Boettcher, S. W. Solution-Cast Metal Oxide Thin Film Electrocatalysts for Oxygen Evolution. J. Am. Chem. Soc. 2012, 134, 1725317261,  DOI: 10.1021/ja307507a
    7. 7
      Sun, K.; Park, N.; Sun, Z.; Zhou, J.; Wang, J.; Pang, X.; Shen, S.; Noh, S. Y.; Jing, Y.; Jin, S.; Yu, P. K. L.; Wang, D. Nickel Oxide Functionalized Silicon for Efficient Photo-Oxidation of Water. Energy Environ. Sci. 2012, 5, 78727877,  DOI: 10.1039/c2ee21708b
    8. 8
      Kenney, M. J.; Gong, M.; Li, Y.; Wu, J. Z.; Feng, J.; Lanza, M.; Dai, H. High-Performance Silicon Photoanodes Passivated with Ultrathin Nickel Films for Water Oxidation. Science 2013, 342, 836840,  DOI: 10.1126/science.1241327
    9. 9
      Zou, S.; Burke, M. S.; Kast, M. G.; Fan, J.; Danilovic, N.; Boettcher, S. W. Fe (Oxy)Hydroxide Oxygen Evolution Reaction Electrocatalysis: Intrinsic Activity and the Roles of Electrical Conductivity, Substrate, and Dissolution. Chem. Mater. 2015, 27, 80118020,  DOI: 10.1021/acs.chemmater.5b03404
    10. 10
      Wang, H.; Lee, H.-W.; Deng, Y.; Lu, Z.; Hsu, P.-C.; Liu, Y.; Lin, D.; Cui, Y. Bifunctional Non-Noble Metal Oxide Nanoparticle Electrocatalysts through Lithium-Induced Conversion for Overall Water Splitting. Nat. Commun. 2015, 6, 7261,  DOI: 10.1038/ncomms8261
    11. 11
      Hill, J. C.; Landers, A. T.; Switzer, J. A. An Electrodeposited Inhomogeneous Metal–Insulator–Semiconductor Junction for Efficient Photoelectrochemical Water Oxidation. Nat. Mater. 2015, 14, 11501155,  DOI: 10.1038/nmat4408
    12. 12
      Jiao, F.; Frei, H. Nanostructured Cobalt and Manganese Oxide Clusters as Efficient Water Oxidation Catalysts. Energy Environ. Sci. 2010, 3, 10181027,  DOI: 10.1039/c002074e
    13. 13
      Hocking, R. K.; Brimblecombe, R.; Chang, L.-Y.; Singh, A.; Cheah, M. H.; Glover, C.; Casey, W. H.; Spiccia, L. Water-Oxidation Catalysis by Manganese in a Geochemical-Like Cycle. Nat. Chem. 2011, 3, 461466,  DOI: 10.1038/nchem.1049
    14. 14
      Iyer, A.; Del-Pilar, J.; King’ondu, C. K.; Kissel, E.; Garces, H. F.; Huang, H.; El-Sawy, A. M.; Dutta, P. K.; Suib, S. L. Water Oxidation Catalysis Using Amorphous Manganese Oxides, Octahedral Molecular Sieves (OMS-2), and Octahedral Layered (OL-1) Manganese Oxide Structures. J. Phys. Chem. C 2012, 116, 64746483,  DOI: 10.1021/jp2120737
    15. 15
      Wiechen, M.; Zaharieva, I.; Dau, H.; Kurz, P. Layered Manganese Oxides for Water-Oxidation: Alkaline Earth Cations Influence Catalytic Activity in a Photosystem II-Like Fashion. Chem. Sci. 2012, 3, 23302339,  DOI: 10.1039/c2sc20226c
    16. 16
      Boppana, V. B. R.; Yusuf, S.; Hutchings, G. S.; Jiao, F. Nanostructured Alkaline-Cation-Containing δ-MnO2 for Photocatalytic Water Oxidation. Adv. Funct. Mater. 2013, 23, 878884,  DOI: 10.1002/adfm.201202141
    17. 17
      Robinson, D. M.; Go, Y. B.; Mui, M.; Gardner, G.; Zhang, Z.; Mastrogiovanni, D.; Garfunkel, E.; Li, J.; Greenblatt, M.; Dismukes, G. C. Photochemical Water Oxidation by Crystalline Polymorphs of Manganese Oxides: Structural Requirements for Catalysis. J. Am. Chem. Soc. 2013, 135, 34943501,  DOI: 10.1021/ja310286h
    18. 18
      Pokhrel, R.; Goetz, M. K.; Shaner, S. E.; Wu, X.; Stahl, S. S. The “Best Catalyst” for Water Oxidation Depends on the Oxidation Method Employed: A Case Study of Manganese Oxides. J. Am. Chem. Soc. 2015, 137, 83848387,  DOI: 10.1021/jacs.5b05093
    19. 19
      Frey, C. E.; Kurz, P. Water Oxidation Catalysis by Synthetic Manganese Oxides with Different Structural Motifs: A Comparative Study. Chem.─Eur. J. 2015, 21, 1495814968,  DOI: 10.1002/chem.201501367
    20. 20
      Smith, P. F.; Deibert, B. J.; Kaushik, S.; Gardner, G.; Hwang, S.; Wang, H.; Al-Sharab, J. F.; Garfunkel, E.; Fabris, L.; Li, J.; Dismukes, G. C. Coordination Geometry and Oxidation State Requirements of Corner-Sharing MnO6 Octahedra for Water Oxidation Catalysis: An Investigation of Manganite (γ-MnOOH). ACS Catal. 2016, 6, 20892099,  DOI: 10.1021/acscatal.6b00099
    21. 21
      Thenuwara, A. C.; Cerkez, E. B.; Shumlas, S. L.; Attanayake, N. H.; McKendry, I. G.; Frazer, L.; Borguet, E.; Kang, Q.; Remsing, R. C.; Klein, M. L.; Zdilla, M. J.; Strongin, D. R. Nickel Confined in the Interlayer Region of Birnessite: An Active Electrocatalyst for Water Oxidation. Angew. Chem., Int. Ed. 2016, 55, 1038110385,  DOI: 10.1002/anie.201601935
    22. 22
      Kang, Q.; Vernisse, L.; Remsing, R. C.; Thenuwara, A. C.; Shumlas, S. L.; McKendry, I. G.; Klein, M. L.; Borguet, E.; Zdilla, M. J.; Strongin, D. R. Effect of Interlayer Spacing on the Activity of Layered Manganese Oxide Bilayer Catalysts for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2017, 139, 18631870,  DOI: 10.1021/jacs.6b09184
    23. 23
      Kölbach, M.; Fiechter, S.; van de Krol, R.; Bogdanoff, P. Evaluation of Electrodeposited α-Mn2O3 as a Catalyst for the Oxygen Evolution Reaction. Catal. Today 2017, 290, 29,  DOI: 10.1016/j.cattod.2017.03.030
    24. 24
      Li, A.; Ooka, H.; Bonnet, N.; Hayashi, T.; Sun, Y.; Jiang, Q.; Li, C.; Han, H.; Nakamura, R. Stable Potential Windows for Long-Term Electrocatalysis by Manganese Oxides under Acidic Conditions. Angew. Chem., Int. Ed. 2019, 58, 50545058,  DOI: 10.1002/anie.201813361
    25. 25
      Takashima, T.; Hashimoto, K.; Nakamura, R. Mechanisms of pH-Dependent Activity for Water Oxidation to Molecular Oxygen by MnO2 Electrocatalysts. J. Am. Chem. Soc. 2012, 134, 15191527,  DOI: 10.1021/ja206511w
    26. 26
      Takashima, T.; Hashimoto, K.; Nakamura, R. Inhibition of Charge Disproportionation of MnO2 Electrocatalysts for Efficient Water Oxidation under Neutral Conditions. J. Am. Chem. Soc. 2012, 134, 1815318156,  DOI: 10.1021/ja306499n
    27. 27
      Zaharieva, I.; Chernev, P.; Risch, M.; Klingan, K.; Kohlhoff, M.; Fischer, A.; Dau, H. Electrosynthesis, Functional, and Structural Characterization of a Water-Oxidizing Manganese Oxide. Energy Environ. Sci. 2012, 5, 70817089,  DOI: 10.1039/c2ee21191b
    28. 28
      Ramírez, A.; Hillebrand, P.; Stellmach, D.; May, M. M.; Bogdanoff, P.; Fiechter, S. Evaluation of MnOx, Mn2O3, and Mn3O4 Electrodeposited Films for the Oxygen Evolution Reaction of Water. J. Phys. Chem. C 2014, 118, 1407314081,  DOI: 10.1021/jp500939d
    29. 29
      Gorlin, Y.; Lassalle-Kaiser, B.; Benck, J. D.; Gul, S.; Webb, S. M.; Yachandra, V. K.; Yano, J.; Jaramillo, T. F. In Situ X-Ray Absorption Spectroscopy Investigation of a Bifunctional Manganese Oxide Catalyst with High Activity for Electrochemical Water Oxidation and Oxygen Reduction. J. Am. Chem. Soc. 2013, 135, 85258534,  DOI: 10.1021/ja3104632
    30. 30
      Huynh, M.; Bediako, D. K.; Nocera, D. G. A Functionally Stable Manganese Oxide Oxygen Evolution Catalyst in Acid. J. Am. Chem. Soc. 2014, 136, 60026010,  DOI: 10.1021/ja413147e
    31. 31
      Huynh, M.; Shi, C.; Billinge, S. J. L.; Nocera, D. G. Nature of Activated Manganese Oxide for Oxygen Evolution. J. Am. Chem. Soc. 2015, 137, 1488714904,  DOI: 10.1021/jacs.5b06382
    32. 32
      Morgan Chan, Z.; Kitchaev, D. A.; Nelson Weker, J.; Schnedermann, C.; Lim, K.; Ceder, G.; Tumas, W.; Toney, M. F.; Nocera, D. G. Electrochemical Trapping of Metastable Mn3+ Ions for Activation of MnO2 Oxygen Evolution Catalysts. Proc. Natl. Acad. Sci. U.S.A. 2018, 115, E5261E5268,  DOI: 10.1073/pnas.1722235115
    33. 33
      Chinnadurai, D.; Nallal, M.; Kim, H.-J.; Li, O. L.; Park, K. H.; Prabakar, K. Mn3+ Active Surface Site Enriched Manganese Phosphate Nano-Polyhedrons for Enhanced Bifunctional Oxygen Electrocatalyst. ChemCatChem 2020, 12, 23482355,  DOI: 10.1002/cctc.202000164
    34. 34
      Shao, C.; Yin, K.; Liao, F.; Zhu, W.; Shi, H.; Shao, M. Rod-Shaped α-MnO2 Electrocatalysts with High Mn3+ Content for Oxygen Reduction Reaction and Zn-Air Battery. J. Alloys Compd. 2021, 860, 158427,  DOI: 10.1016/j.jallcom.2020.158427
    35. 35
      Melder, J.; Mebs, S.; Heizmann, P. A.; Lang, R.; Dau, H.; Kurz, P. Carbon Fibre Paper Coated by a Layered Manganese Oxide: A Nano-Structured Electrocatalyst for Water-Oxidation with High Activity over a Very Wide pH Range. J. Mater. Chem. A 2019, 7, 2533325346,  DOI: 10.1039/c9ta08804k
    36. 36
      Melder, J.; Bogdanoff, P.; Zaharieva, I.; Fiechter, S.; Dau, H.; Kurz, P. Water-Oxidation Electrocatalysis by Manganese Oxides: Syntheses, Electrode Preparations, Electrolytes and Two Fundamental Questions. Z. Phys. Chem. 2020, 234, 925978,  DOI: 10.1515/zpch-2019-1491
    37. 37
      Deibert, B. J.; Zhang, J.; Smith, P. F.; Chapman, K. W.; Rangan, S.; Banerjee, D.; Tan, K.; Wang, H.; Pasquale, N.; Chen, F.; Lee, K.-B.; Dismukes, G. C.; Chabal, Y. J.; Li, J. Surface and Structural Investigation of a MnOx Birnessite-Type Water Oxidation Catalyst Formed under Photocatalytic Conditions. Chem.─Eur. J. 2015, 21, 1421814228,  DOI: 10.1002/chem.201501930
    38. 38
      Sadtler, B.; Burgos, S. P.; Batara, N. A.; Beardslee, J. A.; Atwater, H. A.; Lewis, N. S. Phototropic Growth Control of Nanoscale Pattern Formation in Photoelectrodeposited Se–Te Films. Proc. Natl. Acad. Sci. U.S.A. 2013, 110, 1970719712,  DOI: 10.1073/pnas.1315539110
    39. 39
      Carim, A. I.; Batara, N. A.; Premkumar, A.; Atwater, H. A.; Lewis, N. S. Self-Optimizing Photoelectrochemical Growth of Nanopatterned Se–Te Films in Response to the Spectral Distribution of Incident Illumination. Nano Lett. 2015, 15, 70717076,  DOI: 10.1021/acs.nanolett.5b03137
    40. 40
      Lowe, J. M.; Yan, Q.; Benamara, M.; Coridan, R. H. Direct Photolithographic Patterning of Cuprous Oxide Thin Films Via Photoelectrodeposition. J. Mater. Chem. A 2017, 5, 2176521772,  DOI: 10.1039/c7ta05321e
    41. 41
      Tan, C.; Qin, C.; Sadtler, B. Light-Directed Growth of Metal and Semiconductor Nanostructures. J. Mater. Chem. C 2017, 5, 56285642,  DOI: 10.1039/c7tc00379j
    42. 42
      Carim, A. I.; Hamann, K. R.; Batara, N. A.; Thompson, J. R.; Atwater, H. A.; Lewis, N. S. Template-Free Synthesis of Periodic Three-Dimensional PbSe Nanostructures via Photoelectrodeposition. J. Am. Chem. Soc. 2018, 140, 65366539,  DOI: 10.1021/jacs.8b02931
    43. 43
      Meier, M. C.; Cheng, W.-H.; Atwater, H. A.; Lewis, N. S.; Carim, A. I. Inorganic Phototropism in Electrodeposition of Se–Te. J. Am. Chem. Soc. 2019, 141, 1865818661,  DOI: 10.1021/jacs.9b10579
    44. 44
      Qin, C.; Campbell, B. M.; Shen, M.; Zhao, T.; Sadtler, B. Light-Driven, Facet-Selective Transformation of Cuprous Oxide Microcrystals to Hollow Copper Nanoshells. Chem. Mater. 2019, 31, 80008011,  DOI: 10.1021/acs.chemmater.9b02240
    45. 45
      Sakai, N.; Ebina, Y.; Takada, K.; Sasaki, T. Photocurrent Generation from Semiconducting Manganese Oxide Nanosheets in Response to Visible Light. J. Phys. Chem. B 2005, 109, 96519655,  DOI: 10.1021/jp0500485
    46. 46
      Pinaud, B. A.; Chen, Z.; Abram, D. N.; Jaramillo, T. F. Thin Films of Sodium Birnessite-Type MnO2: Optical Properties, Electronic Band Structure, and Solar Photoelectrochemistry. J. Phys. Chem. C 2011, 115, 1183011838,  DOI: 10.1021/jp200015p
    47. 47
      Hsu, Y.-K.; Chen, Y.-C.; Lin, Y.-G.; Chen, L.-C.; Chen, K.-H. Birnessite-Type Manganese Oxides Nanosheets with Hole Acceptor Assisted Photoelectrochemical Activity in Response to Visible Light. J. Mater. Chem. 2012, 22, 27332739,  DOI: 10.1039/c1jm14355g
    48. 48
      Sunda, W. G.; Huntsman, S. A. Photoreduction of Manganese Oxides in Seawater. Mar. Chem. 1994, 46, 133152,  DOI: 10.1016/0304-4203(94)90051-5
    49. 49
      Sherman, D. M. Electronic Structures of Iron(III) and Manganese(IV) (Hydr)Oxide Minerals: Thermodynamics of Photochemical Reductive Dissolution in Aquatic Environments. Geochim. Cosmochim. Acta 2005, 69, 32493255,  DOI: 10.1016/j.gca.2005.01.023
    50. 50
      Kwon, K. D.; Refson, K.; Sposito, G. On the Role of Mn(IV) Vacancies in the Photoreductive Dissolution of Hexagonal Birnessite. Geochim. Cosmochim. Acta 2009, 73, 41424150,  DOI: 10.1016/j.gca.2009.04.031
    51. 51
      Marafatto, F. F.; Strader, M. L.; Gonzalez-Holguera, J.; Schwartzberg, A.; Gilbert, B.; Peña, J. Rate and Mechanism of the Photoreduction of Birnessite (MnO2) Nanosheets. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 46004605,  DOI: 10.1073/pnas.1421018112
    52. 52
      McCrory, C. C. L.; Jung, S.; Peters, J. C.; Jaramillo, T. F. Benchmarking Heterogeneous Electrocatalysts for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 1697716987,  DOI: 10.1021/ja407115p
    53. 53
      Huynh, M.; Bediako, D. K.; Liu, Y.; Nocera, D. G. Nucleation and Growth Mechanisms of an Electrodeposited Manganese Oxide Oxygen Evolution Catalyst. J. Phys. Chem. C 2014, 118, 1714217152,  DOI: 10.1021/jp501768n
    54. 54
      Nesbitt, H. W.; Banerjee, D. Interpretation of XPS Mn(2p) Spectra of Mn Oxyhydroxides and Constraints on the Mechanism of MnO2 Precipitation. Am. Mineral. 1998, 83, 305315,  DOI: 10.2138/am-1998-3-414
    55. 55
      Gupta, R. P.; Sen, S. K. Calculation of Multiplet Structure of Core p-Vacancy Levels. Phys. Rev. B: Solid State 1974, 10, 7177,  DOI: 10.1103/physrevb.10.71
    56. 56
      Gupta, R. P.; Sen, S. K. Calculation of Multiplet Structure of Core p-Vacancy Levels. II. Phys. Rev. B: Solid State 1975, 12, 1519,  DOI: 10.1103/physrevb.12.15
    57. 57
      Biesinger, M. C.; Payne, B. P.; Grosvenor, A. P.; Lau, L. W. M.; Gerson, A. R.; Smart, R. S. C. Resolving Surface Chemical States in XPS Analysis of First Row Transition Metals, Oxides and Hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 27172730,  DOI: 10.1016/j.apsusc.2010.10.051
    58. 58
      Julien, C.; Massot, M.; Baddour-Hadjean, R.; Franger, S.; Bach, S.; Pereira-Ramos, J. P. Raman Spectra of Birnessite Manganese Dioxides. Solid State Ionics 2003, 159, 345356,  DOI: 10.1016/s0167-2738(03)00035-3
    59. 59
      Chen, D.; Ding, D.; Li, X.; Waller, G. H.; Xiong, X.; El-Sayed, M. A.; Liu, M. Probing the Charge Storage Mechanism of a Pseudocapacitive MnO2 Electrode Using in Operando Raman Spectroscopy. Chem. Mater. 2015, 27, 66086619,  DOI: 10.1021/acs.chemmater.5b03118
    60. 60
      Scheitenberger, P.; Euchner, H.; Lindén, M. The Hidden Impact of Structural Water – How Interlayer Water Largely Controls the Raman Spectroscopic Response of Birnessite-Type Manganese Oxide. J. Mater. Chem. A 2021, 9, 1846618476,  DOI: 10.1039/d1ta05357d
    61. 61
      Wu, T.-H.; Hesp, D.; Dhanak, V.; Collins, C.; Braga, F.; Hardwick, L. J.; Hu, C.-C. Charge Storage Mechanism of Activated Manganese Oxide Composites for Pseudocapacitors. J. Mater. Chem. A 2015, 3, 1278612795,  DOI: 10.1039/c5ta03334a
    62. 62
      Yang, L.; Cheng, S.; Ji, X.; Jiang, Y.; Zhou, J.; Liu, M. Investigations into the Origin of Pseudocapacitive Behavior of Mn3O4 Electrodes Using in Operando Raman Spectroscopy. J. Mater. Chem. A 2015, 3, 73387344,  DOI: 10.1039/c5ta00223k
    63. 63
      Boumaiza, H.; Renard, A.; Rakotomalala Robinson, M.; Kervern, G.; Vidal, L.; Ruby, C.; Bergaoui, L.; Coustel, R. A Multi-Technique Approach for Studying Na Triclinic and Hexagonal Birnessites. J. Solid State Chem. 2019, 272, 234243,  DOI: 10.1016/j.jssc.2019.02.017
    64. 64
      Scheitenberger, P.; Brimaud, S.; Lindén, M. XRD/Raman Spectroscopy Studies of the Mechanism of (De)Intercalation of Na+ from/into Highly Crystalline Birnessite. Mater. Adv. 2021, 2, 39403953,  DOI: 10.1039/d1ma00161b
    65. 65
      Nam, K. W.; Kim, S.; Yang, E.; Jung, Y.; Levi, E.; Aurbach, D.; Choi, J. W. Critical Role of Crystal Water for a Layered Cathode Material in Sodium Ion Batteries. Chem. Mater. 2015, 27, 37213725,  DOI: 10.1021/acs.chemmater.5b00869
    66. 66
      Lu, K.; Hu, Z.; Xiang, Z.; Ma, J.; Song, B.; Zhang, J.; Ma, H. Cation Intercalation in Manganese Oxide Nanosheets: Effects on Lithium and Sodium Storage. Angew. Chem., Int. Ed. 2016, 55, 1044810452,  DOI: 10.1002/anie.201605102
    67. 67
      Shan, X.; Guo, F.; Charles, D. S.; Lebens-Higgins, Z.; Abdel Razek, S.; Wu, J.; Xu, W.; Yang, W.; Page, K. L.; Neuefeind, J. C.; Feygenson, M.; Piper, L. F. J.; Teng, X. Structural Water and Disordered Structure Promote Aqueous Sodium-Ion Energy Storage in Sodium-Birnessite. Nat. Commun. 2019, 10, 4975,  DOI: 10.1038/s41467-019-12939-3
    68. 68
      Chigane, M.; Ishikawa, M. XRD and XPS Characterization of Electrochromic Nickel Oxide Thin Films Prepared by Electrolysis–Chemical Deposition. J. Chem. Soc., Faraday Trans. 1998, 94, 36653670,  DOI: 10.1039/a806459h
    69. 69
      Chigane, M.; Ishikawa, M. Manganese Oxide Thin Film Preparation by Potentiostatic Electrolyses and Electrochromism. J. Electrochem. Soc. 2000, 147, 2246,  DOI: 10.1149/1.1393515
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsnanoscienceau.3c00002.

    • Procedures used to synthesize manganese oxide powders and obtain their diffraction patterns, variations of the square-wave potential used to electrodeposit MnOx films, procedure for obtaining the double-layer capacitances and ECSAs, and characterization of electrodeposited MnOx films and reference powders by XANES; summary of electrochemical characterization of MnOx films and the parameters used to fit the X-ray photoelectron spectra of different MnOx films and reference powders; additional characterization by linear sweep voltammetry and constant-current electrolysis to measure the stability of MnOx films, CV used to measure the double-layer capacitances of the MnOx films, Mott–Schottky plots of MnOx films at different frequencies, X-ray absorption near edge spectroscopy and linear fitting of XANES data, Raman spectrum of a MnO2 film grown at constant potential, and XPS of electrodeposited MnOx films in the region for K 2p electrons (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.