Kristina
Gočanin
a,
Yasemin
Aykut
b,
Dušan
Mladenović
*a,
Diogo M. F.
Santos
c,
Ayşe
Bayrakçeken
b,
Gulin S. P.
Soylu
d and
Biljana
Šljukić
ac
aUniversity of Belgrade, Faculty of Physical Chemistry, Studentski trg 12-16, 11158 Belgrade, Serbia. E-mail: dusan.mladenovic@ffh.bg.ac.rs
bDepartment of Chemical Engineering, Atatürk University, Erzurum, 25240, Turkey
cCenter of Physics and Engineering of Advanced Materials, Laboratory of Physics for Materials and Emerging Technologies, Chemical Engineering Department, Instituto Superior Técnico, Universidade de Lisboa, 1049-001 Lisbon, Portugal
dFaculty of Engineering, Chemical Engineering Department, Istanbul University-Cerrahpasa, 34320, Avcilar, Istanbul, Turkey
First published on 1st September 2025
Developing efficient, low-cost catalysts for oxygen reduction and evolution reactions (ORR and OER) is key to advancing metal–air batteries and regenerative fuel cells. In this study, nitrogen-doped binary metal (Mn and Ni) oxides (N–BMOs) and Pt-decorated N–BMOs were synthesised using three methods and tested as ORR and OER catalysts in alkaline media. Their physicochemical properties were characterised by XRD, N2-sorption, TEM, and XPS, while their electrochemical performance was evaluated using voltammetry and impedance spectroscopy. Among all tested materials, the best bifunctional catalyst proved to be Pt/N–Mn2O3–NiO (1:
1) (S3) with the highest achieved diffusion limited current density (−4.98 mA cm−2 at 1800 rpm), the highest kinetic current density (−15.3 mA cm−2), low Tafel slope (75 mV dec−1) in ORR potential region, and overpotential of 0.56 V to reach benchmark current value of 10 mA cm−2 during OER. The ΔE was calculated to be 0.95 V, comparable to or even better than that of similar materials reported in the literature. Pt/N–Mn2O3–NiO (1
:
1) (S3) demonstrated striking stability during long-term operation with preserved morphology and catalytic activity.
Transition metals and transition metal oxides (TMOs) have garnered significant attention from researchers due to their high natural abundance and low cost,8 along with excellent performance in oxygen reduction and oxygen evolution reactions.9–11 Moreover, these materials have proven to be effective carriers (supports) of active particles. Zhao and coworkers synthesised MOF-derived nitrogen-doped carbon nanotubes encapsulated with bimetallic oxide, FeNiO@NCNT, which reached a diffusion-limited current density, jd, of −2.5 mA cm−2 at 1600 rpm with ΔE of 0.82 V.12 Morales et al. synthesised trimetallic Mn–Fe–Ni oxides on multi-walled carbon nanotubes (MWCNTs) by first growing and oxygen-functionalising MWCNTs to obtain MWCNTs-Ox with an average outer diameter of 8 to 10 nm. MnOx was then incorporated into Fe–Ni oxide previously supported on MWCNTs-Ox.13 Mn0.5(Fe0.3Ni0.7)0.5Ox/MWCNTs-Ox showed low values of ORR and OER overpotentials with the ΔE of 0.73 V and jd in the ORR potential region of −4.2 mA cm−2 at 1600 rpm. Shao et al. synthesised carbon dots that bridge NiO and Mn2O3, NiO–Mn2O3–CDs, which showed an almost four-electron oxygen reduction (n = 3.85) and ΔE 0.72 V.14 Furthermore, the material achieved jd of −6 mA cm−2 in ORR mode and 10 mA cm−2 in OER mode at an overpotential of 0.298 V. In the authors’ previous work, non-doped bimetallic Pt/Mn2O3–NiO15 and nitrogen-doped Pt/Mn2O3–NiO–N16 were synthesised and tested under the same conditions as bifunctional URFC catalysts. The undoped Pt/Mn2O3–NiO showed remarkable activity, achieving −4.475 mA cm−2 at 1800 rpm and ΔE of 1.03 V, while the Pt/Mn2O3–NiO–N showed ΔE values of 0.88–0.99 V and jd of −4.43 to −4.81 mA cm−2, depending on the BMO to N ratio. Though Pt might outperform TMOs, it should be kept in mind that Pt is known to undergo oxidation at higher potentials,17 which effectively increases the ORR overpotential17,18 and reduces the OER activity of the catalyst, unlike transition metal oxides that cannot be so easily oxidised when increasing applied potentials.10,19,20
This work continued the investigation of nitrogen-doped N–Mn2O3–NiO and Pt/N–Mn2O3–NiO (with 20 wt% Pt) with two different ratios of BMO to N by slightly modifying the synthesis procedure to improve the catalytic performance of the materials. Three modified procedures were used to synthesise the materials designated as (S1), (S2), and (S3), which were then tested in alkaline media as bifunctional oxygen electrode catalysts. This design is based on the synergy between the excellent catalytic activity and electrical conductivity of Pt with the tuneable surface chemistry, defect structure, and stability of N-doped BMOs. Nitrogen-doping introduces electronic modifications and surface defects that can enhance Pt anchoring, dispersion, and charge transfer kinetics. By engineering the metal–support interface and optimising particle distribution, the composite materials aim to simultaneously boost electrochemically active surface area (ECSA), reduce overpotentials for ORR and OER, and improve long-term catalytic durability. Electrochemical testing revealed significant improvements in both OER and ORR activity compared to the undoped BMO, demonstrating the effectiveness of the combined doping-deposition approach and highlighting the potential of these hybrid materials as bifunctional electrocatalysts for energy conversion systems such as metal–air batteries and regenerative fuel cells.
Mn2O3 was prepared by co-precipitation. An appropriate amount of metal nitrate (Mn(NO3)2·4H2O) was dissolved in deionised warm water, and the resulting solution was heated to 65 °C. This mixture was precipitated by gradually adding NH3 solution (25 wt%) until the pH value reached 10. The resultant solution was processed following the same steps as in the case of NiO.
The binary metal oxides (Mn2O3 + NiO) were prepared by the solid-state dispersion (SSD) method. Mn2O3 and NiO were thoroughly mixed in a 1:
1 weight ratio using ethanol in an agate pestle and mortar; the solvent was then removed by evaporation during the mixing process. Samples prepared by this method were dried at 110 °C for 90 min and calcined at 450 °C for 6 h to obtain binary oxide catalysts. The resultant binary oxide was ground at a constant vibration rate of 300 rpm for 15 min in a Retsch MM 200 vibratory ball mill with 12 mm ZrO2 milling balls in a ZrO2 milling container.
Electrochemical studies were conducted using cyclic voltammetry, linear scan voltammetry with a rotating disc electrode and electrochemical impedance spectroscopy (EIS). Experimental details of electrode preparation and electrochemical measurements are given in the SI.
Surface area and pore structure are defined as essential characteristics of catalyst support materials, and they have a significant impact on the catalyst's activity. Fig. S1 shows the N2-adsorption/desorption isotherms and pore size distributions for the N-doped oxide structures. These structures display a Type IV isotherm according to IUPAC classification, indicating the presence of both mesopores and micropores. Capillary condensation occurs within mesopores, where gas molecules are adsorbed, leading to the formation of a hysteresis loop.21 This hysteresis is observed in the relative pressure (P/P0) range of 0.8–1.0, Fig. S1(a–c). However, in the case of N–Mn2O3–NiO (1:
2) (S2) structure, hysteresis extends up to a relative pressure (P/P0) of 0.4, resembling Type-3 hysteresis loops. This type of hysteresis loop shows the presence of sheet-like pores within the structure. Fig. S1 presents the pore size distribution curves, which reveal differences in textural properties as suggested by the N2 isotherms. The inner graphs show variations in pore size distributions from 0 to 30 nm. The data indicate that most pores in the N–Mn2O3–NiO structures are between 0 and 5 nm wide. These structures, designed as catalyst support materials, exhibit mesoporous characteristics within this size range. Changes in pore size distribution also affect the BET surface area of the N-doped structures. Additionally, the effect of the nitrogen doping ratio (1
:
1 and 1
:
2) on the surface morphology of the composites was evaluated based on the BET surface area and pore structure properties determined by the BJH method (Table 1). The results reveal that the amount of nitrogen doping is a significant determinant of both the specific surface area and pore structure. Lower surface area values were obtained for a nitrogen doping level of 1
:
1. In samples S1 and S2, nitrogen doping with a 1
:
1 melamine-to-support ratio resulted in lower surface area values. In the S3 groups, significantly lower surface areas were achieved at both doping ratios. This suggests that synthesis under less basic conditions limits porosity formation, regardless of nitrogen content, and that pH plays a crucial role in determining the structure of the resulting material. Furthermore, if the nitrogen content exceeds a certain level, the porous structure may become partially blocked, micropores may convert into mesopores, or pore walls may collapse during doping, resulting in a lower surface area. This assessment is also supported by data on pore volume and average pore diameter. Therefore, not only nitrogen content but also the preservation of pore structure and the homogeneous distribution of nitrogen throughout the structure are critical for improving surface properties. All findings reveal that the combined evaluation of the melamine-to-support ratio and synthesis pH is the key factor in controlling the structural properties of the composites.22
Sample | BET surface area (m2 g−1) | BJH adsorption cumulative pore volume (cm3 g−1) | BJH desorption cumulative pore volume (cm3 g−1) | BJH adsorption average pore width (nm) | BJH desorption average pore width (nm) |
---|---|---|---|---|---|
N–Mn2O3–NiO (1![]() ![]() |
20.31 | 0.226 | 0.227 | 40.86 | 38.33 |
N–Mn2O3–NiO (1![]() ![]() |
28.13 | 0.241 | 0.241 | 30.56 | 29.04 |
N–Mn2O3–NiO (1![]() ![]() |
24.62 | 0.087 | 0.087 | 12.33 | 11.44 |
N–Mn2O3–NiO (1![]() ![]() |
37.71 | 0.118 | 0.119 | 10.94 | 9.79 |
N–Mn2O3–NiO (1![]() ![]() |
2.77 | 0.012 | 0.012 | 16.18 | 15.45 |
N–Mn2O3–NiO (1![]() ![]() |
3.05 | 0.020 | 0.019 | 21.41 | 20.82 |
To identify the crystalline phases in the synthesised structures, the samples were characterised by X-ray diffraction (Fig. 1). The peaks observed at ∼32.9°, 38.2°, and 55.1° in all samples belong to the (222), (400), and (440) planes of the Mn2O3 structure. The sharp peaks at 37.2°, 43.2°, 62.8°, 75.2°, and 79.3° are attributed to the (111), (200), (220), (311), and (222) planes of NiO, respectively. Again, characteristic diffraction peaks were observed at 18.5°, 30.5°, and 57.2°, corresponding to the (111), (200), and (511) planes of NiMn2O4 in all catalyst structures. The characteristic peaks of Pt face cubic centred (fcc) at 40.1°, 46.5°, 68.0°, and 81.9° correspond to the (111), (200), (220), and (311) planes, respectively. Furthermore, an increase in the nitrogen content in the samples has resulted in a relative widening of the peaks.16,23
Fig. 2 presents the XPS survey spectra of the synthesised samples. Distinct peaks corresponding to C 1s, N 1s, O 1s, Mn 2p, and Ni 2p are observed at binding energies of 284.7 eV, 400.7 eV, 532.9 eV, 642.6 eV, and 855.0 eV, respectively. The prominent signals for Mn, Ni, and O confirm the presence of the constituent elements of the support materials. Notably, the N 1s peak becomes increasingly sharp with higher nitrogen content, suggesting enhanced nitrogen incorporation. The C 1s peak is attributed to carbon atoms derived from melamine, which was used as the nitrogen source during synthesis.
The successful N-doping of the support material as targeted (Table S1), with a high intensity of the N 1s peak, enables the decomposition of the N 1s peak into various nitrogen species. The high-resolution XPS spectra of the nitrogen-doped support materials in the N 1s region are displayed in Fig. 3. When the samples’ N 1s spectra were decomposed into their sub-peaks, pyridinic-N, pyrrolic-N, graphitic-N, and pyridinic N-oxide showed their peaks at binding energies of 398.07 ± 0.5 eV, 398.65 ± 0.5 eV, 399.25 ± 0.5 eV, and 400.10 ± 0.5 eV, respectively. It is well established that the electrochemical activity is linked more to these types of nitrogen species than to the overall nitrogen content.24 Nitrogen in the pyridine structure contributes one or two π electrons to the aromatic π system and bonds with two nearby carbon atoms in six-sided or five-sided carbon rings. Graphite-type nitrogen lets three surrounding carbon atoms connect with a carbon atom entering the sp2 hybridisation state. Bonds developed by two carbon and one oxygen atom define the pyridinic N-oxide structure. Deconvolution analysis shows that the primary configurations in the synthesised metal oxide-based structures were pyrrolic-N and pyridinic-N.25 The data for the nitrogen element obtained by XPS confirm the elemental analysis findings presented in Table S1. The peaks in the N 1s XPS spectra of the sample groups doped with a (1:
1) ratio are significantly higher, and the peaks belonging to low-binding-energy pyridinic and pyrrolic nitrogen compounds are particularly high compared to other peaks. This indicates that nitrogen doping promotes the formation of more active nitrogen centres at edge and defect sites on the support material structures. In contrast, in samples with a (1
:
2) ratio, the intensity of the N 1s peaks decreased due to the decrease in total nitrogen content, resulting in a narrower area. The contribution of pyridinic and pyrrolic nitrogen species decreased, while the relative proportion of the graphitic nitrogen component increased. This finding suggests that the limited amount of nitrogen is primarily localised to more graphitic locations within the support material lattice and that active nitrogen formation at edge/defect sites is limited. While increasing the graphitic N ratio offers the potential to improve the material's electrical conductivity, it can negatively affect catalytic performance by limiting reactive centre formation due to the limited active surface area.26 The percentages of different N-species were calculated by dividing the area of individual peaks in the N 1s spectrum by the total area of the four peaks (Table 2). N–Mn2O3–NiO (1
:
1) (S3) sample contains a higher proportion of pyridinic (66.46%) and pyrrolic (21.09%) nitrogen compared to the other samples. Nitrogen doping plays a critical role in electrocatalytic processes by regulating the electronic properties of materials and promoting the formation of active centres. Pyridinic and pyrrolic nitrogen configurations, in particular, constitute active centres for reactions such as HER, ORR, and OER, due to their capacity to alter charge distribution and electron density. Qu et al.27 reported that high onset potentials and current densities are achieved in catalysts rich in pyridinic nitrogen, as these regions serve as suitable centres for the activation of the O2 molecule. Similarly, Gong et al.28 reported that the synergistic effect of pyridinic and graphitic nitrogen significantly increased catalytic efficiency by promoting the four-electron ORR pathway. It was also stated that the electron delocalisation provided by pyridinic nitrogen accelerated charge transfer and supported the stability of active sites. Zhang et al.29 determined that N-doped carbon-supported Pt electrocatalysts with high pyrolic and pyridinic nitrogen contents and no other nitrogen species exhibited approximately three times higher ORR activity compared to Pt/XC electrocatalysts without nitrogen doping. A study by Ning et al.30 reported that the sample with the highest pyridinic-N/graphitic-N ratio exhibited the highest ORR activity, a result consistent with the view that pyridinic nitrogen provides the most active sites for ORR. The obtained data suggest that samples with a 1
:
1 doping ratio offer greater advantages in terms of potential surface reactivity and catalytic activity due to their higher pyridinic/pyrrolic nitrogen content, whereas samples with a (1
:
2) doping ratio may exhibit more limited performance in this regard due to their lower active nitrogen content.
Sample | Pyridinic-N (%) | Pyrrolic-N (%) | Graphitic-N (%) | Pyridinic N-oxide (%) |
---|---|---|---|---|
N–Mn2O3–NiO (1![]() ![]() |
54.22 | 26.99 | 10.47 | 8.298 |
N–Mn2O3–NiO (1![]() ![]() |
51.54 | 22.86 | 20.54 | 5.045 |
N–Mn2O3–NiO (1![]() ![]() |
55.42 | 13.78 | 18.52 | 12.26 |
N–Mn2O3–NiO (1![]() ![]() |
43.10 | 22.51 | 23.41 | 10.95 |
N–Mn2O3–NiO (1![]() ![]() |
66.46 | 21.09 | 10.12 | 2.316 |
N–Mn2O3–NiO (1![]() ![]() |
58.75 | 19.77 | 17.97 | 3.483 |
TEM and HR-TEM analysis were employed to examine the structure and size of the particles in the Pt/N–Mn2O3–NiO catalysts in greater detail (Fig. 4). Some aggregation of nanoparticles on the support material was observed, possibly due to the insufficient mixing of the metal precursor and the support material. Additionally, the surface area of the support material is another parameter that affects the homogeneous distribution of Pt particles. In this direction, it can be seen that agglomeration is more pronounced in the S3 structured support material, which has a lower surface area. The TEM images of the catalysts reveal that the particles are nearly spherical. The mean particle size values derived from the TEM analysis images were calculated using ImageJ software, yielding average sizes within a narrow range of 3 to 4 nm. Specifically, Pt particles were of ca. 3.61 nm, 3.73 nm, 3.43 nm, 3.21 nm, 4.02 nm, and 3.82 nm size for the Pt/N–Mn2O3–NiO (1:
1) (S1), Pt/N–Mn2O3–NiO (1
:
2) (S1), Pt/N–Mn2O3–NiO (1
:
1) (S2), Pt/N–Mn2O3–NiO (1
:
2) (S2), Pt/N–Mn2O3–NiO (1
:
1) (S3), and Pt/N–Mn2O3–NiO (1
:
2) (S3) catalysts, respectively. However, despite Pt particle size remaining constant over different synthesis methods, factors such as the chemical composition and electronic characteristics of the BMO support, Pt–BMO interface interactions, and nitrogen doping directly affect the surface electronic properties and charge density of Pt nanoparticles, thus playing a decisive role in electrochemical activity.
![]() | ||
Fig. 4 TEM (50 nm scale bar) and HR-TEM (5 nm scale bar) images of catalysts and Pt particle size distributions (left to right: TEM – HRTEM – Pt particle size histograms). |
The recorded cyclic voltammograms of the second set of synthesised catalysts are shown in Fig. S3. The highest Cdl value was obtained for Pt/N–Mn2O3–NiO (1:
1) (S2) (2.53 mF cm−2), corresponding to an ECSA of 63.3 cm2. In contrast, the calculated ECSA values for N–Mn2O3–NiO (1
:
1) (S2), N–Mn2O3–NiO (1
:
2) (S2), and Pt/N–Mn2O3–NiO (1
:
2) (S2) were significantly lower, amounting to only 12.3 cm2, 6.25 cm2, and 20.3 cm2, respectively.
The recorded cyclic voltammograms of the third set of synthesised catalysts are shown in Fig. S4. As in the case of the previous two syntheses, the highest value of Cdl, 2.38 mF cm−2, was obtained for Pt/N–Mn2O3–NiO (1:
1) (S3), corresponding to an ECSA of 59.5 cm2. A slightly lower ECSA value of 46.3 cm2 was calculated Pt/N–Mn2O3–NiO (1
:
2) (S3). The pure N-doped binary oxides, N–Mn2O3–NiO (1
:
1) (S3) and N–Mn2O3–NiO (1
:
2) (S3) showed similar ECSA values of 8.25 cm2 and 8.75 cm2 per 1 cm2 of the geometric surface area of the electrode.
A comparison of all the obtained Cdl and ECSA values for the investigated catalysts is presented in Table S3. The Pt-decorated materials exhibited significantly higher ECSA values than the pure N-doped binary oxides. Moreover, Pt-decorated materials with an N to Mn2O3–NiO ratio of 1:
1 were characterised by a higher ECSA compared to those with an N to Mn2O3–NiO ratio of 1
:
2.
The incorporation of Pt into N-doped transition metal oxides has been shown to significantly enhance capacitive currents, primarily due to synergistic improvements in electronic conductivity, surface characteristics, and interfacial charge dynamics.32 Pt facilitates rapid electron transport within the electrode, effectively reducing internal resistance and enabling more efficient charging and discharging of the electric double layer. Additionally, the incorporation of Pt nanoparticles can increase the ECSA by introducing surface roughness and porosity, thereby providing more sites for double-layer formation.33 At the interface, strong electronic interactions between Pt and the N-doped TMO matrix can lead to beneficial charge redistribution, further increasing Cdl. Moreover, Pt promotes faster ion adsorption and desorption kinetics (especially in acidic electrolytes), which supports the rapid accumulation of surface charges characteristic of capacitive behaviour.34 Finally, Pt can improve charge separation and suppress recombination processes in systems where pseudocapacitive contributions are present, further boosting the observed current response. The synergy between Pt addition and nitrogen doping thus led to a pronounced increase in capacitive currents for the decorated TMO materials.35
Although TEM analysis revealed relatively similar Pt nanoparticle sizes across the S1, S2, and S3 samples, the observed differences in ECSA augmentation (2.56× for S1 (1:
1), 5.16× for S2 (1
:
1), and 7.12× for S3 (1
:
1) compared to their non-Pt decorated analogues) suggest that additional factors beyond apparent particle size influence the electrochemical behaviour. Notably, intrinsic differences in the physicochemical properties of the deposited Pt, including crystallinity, surface energy, and electronic structure, may vary subtly between samples depending on the synthesis environment and precursor–substrate interactions, thereby affecting the density of active sites.36 Moreover, variations in the synthesis protocols of the N-doped TMO substrates themselves may have impacted Pt anchoring, dispersion stability, and interfacial charge transfer. Post-synthesis factors such as catalyst ink formulation, degree of dispersion, and electrode fabrication also play a critical role. Poor ink dispersion during film preparation may have led to localised Pt clustering and uneven ionomer distribution, reducing surface accessibility and limiting double-layer formation, while improved ink homogeneity and film morphology may have facilitated more uniform Pt distribution and higher surface utilisation.37 These findings emphasise that both the synthesis of the material and the post-synthesis handling, particularly ink formulation and deposition, are key to optimising the electrochemical performance of Pt-modified N-doped TMO materials.
Material | ORR parameters | OER parameters | Source | ||||||
---|---|---|---|---|---|---|---|---|---|
j d/mA cm−2 | j k/mA cm−2 | E 1/2/V | b/mV dec−1 | n | η 10 mA cm−2/V | b/mV dec−1 | ΔE/V | ||
N–Mn2O3–NiO (1![]() ![]() |
−2.66 | −1.25 | 0.66 | 96 | 3.04 | — | 267 | — | This work |
N–Mn2O3–NiO (1![]() ![]() |
−1.89 | −1.04 | 0.65 | 115 | 2.04 | — | 246 | — | This work |
Pt/N–Mn2O3–NiO (1![]() ![]() |
−4.86 | −8.05 | 0.86 | 65 | 3.67 | 0.63 | 272 | 1.00 | This work |
Pt/N–Mn2O3–NiO (1![]() ![]() |
−1.64 | −0.52 | 0.64 | 102 | 2.93 | — | 255 | — | This work |
N–Mn2O3–NiO (1![]() ![]() |
−3.17 | −1.88 | 0.67 | 93 | 3.15 | 0.67 | 177 | 1.23 | This work |
N–Mn2O3–NiO (1![]() ![]() |
−2.76 | −0.14 | 0.70 | 82 | 3.07 | — | 225 | — | This work |
Pt/N–Mn2O3–NiO (1![]() ![]() |
−3.60 | −1.72 | 0.89 | 93 | 3.98 | 0.56 | 230 | 0.90 | This work |
Pt/N–Mn2O3–NiO (1![]() ![]() |
−3.90 | −0.34 | 0.71 | 90 | 3.37 | 0.57 | 205 | 1.09 | This work |
N–Mn2O3–NiO (1![]() ![]() |
−2.71 | −1.18 | 0.70 | 91 | 3.21 | — | 195 | — | This work |
N–Mn2O3–NiO (1![]() ![]() |
−2.93 | −2.44 | 0.63 | 93 | 2.54 | — | 189 | — | This work |
Pt/N–Mn2O3–NiO (1![]() ![]() |
−4.98 | −15.3 | 0.84 | 75 | 3.56 | 0.56 | 128 | 0.95 | This work |
Pt/N–Mn2O3–NiO (1![]() ![]() |
−4.53 | −4.56 | 0.86 | 86 | 3.90 | 0.62 | 197 | 0.99 | This work |
NiO/NiCo2O4 | — | — | 0.37 | 85.4 | — | — | 130 | — | 41 |
MnO/Co/PGC | — | — | 0.78 | 69 | — | — | 77 | — | 42 |
MnO2–C(ultrathin amorphous) | −5.81 | — | 0.81 | — | 4.00 | — | — | — | 43 |
MnOx/CNTs | −4.90 | — | 0.77 | — | 3.60 | — | — | — | 44 |
Pt/Mn2O3–NiO | −4.48 | −4.34 | 0.79 | 62 and 109 | 3.73 | 0.54 | 154 | 0.98 | 15 |
PtNi/Mn2O3–NiO | −4.32 | −3.10 | 0.79 | 63 and 103 | 3.99 | 0.53 | 140 | 0.97 | 15 |
Mn2O3–NiO | −1.93 | −0.67 | 0.66 | 151 | 2.90 | 0.57 | 155 | 1.14 | 16 |
Mn2O3–NiO–N (1![]() ![]() |
−2.75 | −1.65 | 0.75 | 82 | 2.73 | — | 230 | — | 16 |
Pt/Mn2O3–NiO–N (1![]() ![]() |
−4.69 | −2.97 | 0.87 | 90 | 3.88 | 0.63 | 249 | 0.99 | 16 |
IrO2 | — | — | — | — | — | 0.36 | 84 | — | 45 |
RuO2 | — | — | — | — | — | 0.40 | — | — | 46 |
Pt/C (40 wt%) | −6.44 | 14.9 | 0.86 | 79 and 60 | 3.97 | 0.58 | 198 | 0.95 | 15 |
The LSV curves at 1800 rpm were subjected to Tafel analysis (Fig. S7(a)). The lowest Tafel slope (b) value was obtained for Pt/N–Mn2O3–NiO (1:
1) (S1) (65 mV dec−1), indicating faster ORR kinetics for this material compared to the other three materials obtained by this synthesis method. For N–Mn2O3–NiO (1
:
1) (S1), N–Mn2O3–NiO (1
:
2) (S1), and Pt/N–Mn2O3–NiO (1
:
2) (S1), the calculated b values amount to 96 mV dec−1, 115 mV dec−1, and 102 mV dec−1, respectively.
To determine the number of exchanged electrons (n) during an electrochemical reaction, Koutecký–Levich (K–L) analysis is performed, an electrochemical method used to separate the kinetic current density (jk) from the diffusion-limited current density (jd) in rotating disc electrode experiments. The number of exchanged electrons, n, during the elementary step of the reaction is determined using the K–L equation (eqn (1)),
![]() | (1) |
B = 0.62nFAD2/3ν−1/6C | (2) |
Upon comparison of all analysed ORR parameters, Pt/N–Mn2O3–NiO (1:
1) (S1) emerged as the most effective catalyst synthesised using this method, achieving the highest diffusion-limited current density and half-wave potential, the lowest Tafel slope, and the highest number of transferred electrons.
Following the analysis of the first set of materials, the same procedure (cyclic voltammetry and LSVs at varying scan rates) was applied to the second set of materials, as shown in Fig. S8 and S9. The oxygen reduction maximum is observed at the cyclic voltammograms in 0.1 M KOH saturated with O2 at 0.91 V and 0.80 V for Pt/N–Mn2O3–NiO (1:
1) (S2) and Pt/N–Mn2O3–NiO (1
:
2) (S2), respectively. Higher jd were delivered by the Pt/N–Mn2O3–NiO (1
:
2) (S2) catalyst (−3.90 mA cm−2) and Pt/N–Mn2O3–NiO (1
:
1) (S2) (−3.60 mA cm−2) compared to the undecorated oxides. N–Mn2O3–NiO (1
:
1) (S2), and N–Mn2O3–NiO (1
:
2) (S2) delivered jd values of −3.17 mA cm−2 and −2.76 mA cm−2, respectively. The values of the half-wave potential were determined to be 0.67 V, 0.70 V, 0.89 V, and 0.71 V for N–Mn2O3–NiO (1
:
1) (S2), N–Mn2O3–NiO (1
:
2) (S2), Pt/N–Mn2O3–NiO (1
:
1) (S2), and Pt/N–Mn2O3–NiO (1
:
2) (S2), respectively.
Tafel slopes (Fig. S10(a)) were of similar values for Pt-decorated and non-decorated N-doped materials within a 82–93 mV dec−1 range (Table 3).
The Koutecký–Levich analysis (Fig. S10(b)) revealed that oxygen reduction on Pt/N–Mn2O3–NiO (1:
1) (S2) proceeds mainly via the four-electron reduction mechanism, while on Pt/N–Mn2O3–NiO (1
:
2) (S2) and non-decorated BMOs, ORR proceeds via mixed kinetics. Namely, the number of exchanged electrons was calculated to be 3.98 and 3.37 for Pt/N–Mn2O3–NiO (1
:
1) (S2) and Pt/N–Mn2O3–NiO (1
:
2) (S2), respectively. For N–Mn2O3–NiO (1
:
1) (S2) and N–Mn2O3–NiO (1
:
2) (S2), the n was found to be 3.15 and 3.07, respectively.
As in the case of the materials obtained in the previous synthesis, the Pt-decorated catalyst with the BMO to N ratio of 1:
1 (Pt/N–Mn2O3–NiO (1
:
1) (S2)) showed the overall best performance for ORR in terms of the highest number of exchanged electrons in the elementary step of the reaction, the highest value of E1/2 and the second highest obtained jd.
Fig. S11(a–d) show the results of the CV experiments obtained with the third set of materials. Clear oxygen reduction maxima can be observed at 0.70 V, 0.67 V, 0.88 V, and 0.90 V for N–Mn2O3–NiO (1:
1) (S3), N–Mn2O3–NiO (1
:
2) (S3), Pt/N–Mn2O3–NiO (1
:
1) (S3), and Pt/N–Mn2O3–NiO (1
:
2) (S3), respectively. As expected, the Pt-decorated materials showed higher diffusion-limited current densities and higher half-wave potentials compared to the undecorated N-doped materials. The highest jd value was reached by Pt/N–Mn2O3–NiO (1
:
1) (S3) (−4.98 mA cm−2), which also delivered the highest kinetic current density (−15.27 mA cm−2). Pt/N–Mn2O3–NiO (1
:
2) (S3) reached jd of −4.53 mA cm−2 and jk of −4.56 mA cm−2, while N–Mn2O3–NiO (1
:
1) (S3) and N–Mn2O3–NiO (1
:
2) (S3) show significantly lower values. The half-wave potentials for these materials were determined to be 0.70 V, 0.63 V, 0.84 V, and 0.86 V for N–Mn2O3–NiO (1
:
1) (S3), N–Mn2O3–NiO (1
:
2) (S3), Pt/N–Mn2O3–NiO (1
:
1) (S3), and Pt/N–Mn2O3–NiO (1
:
2) (S3), respectively. Within this set of materials, Pt/N–Mn2O3–NiO (1
:
1) (S3) exhibited the lowest Tafel slope value of 75 mV dec−1, while N–Mn2O3–NiO (1
:
1) (S3), N–Mn2O3–NiO (1
:
2) (S3) and Pt/N–Mn2O3–NiO (1
:
2) (S3) showed slightly higher values of 91 mV dec−1, 93 mV dec−1, and 86 mV dec−1, respectively (Fig. S13(a)). The number of exchanged electrons was calculated to be between 2.54 and 3.90, Fig. S13(b). The highest value of n was calculated for Pt/N–Mn2O3–NiO (1
:
2) (S3) (n = 3.90), while the lowest value was calculated for N–Mn2O3–NiO (1
:
2) (S3) (n = 2.54). For N–Mn2O3–NiO (1
:
1) (S3) and Pt/N–Mn2O3–NiO (1
:
1) (S3), it was calculated that n = 3.21 and n = 3.56, respectively.
Comparing the obtained results, both Pt-decorated materials from the third set showed similar ORR performance. On both materials, ORR proceeds mainly via the four-electron mechanism, and the jd values reached with these two were among the highest of all tested materials. Furthermore, E1/2 values for Pt/N–Mn2O3–NiO (1:
1) (S3) and Pt/N–Mn2O3–NiO (1
:
2) (S3) were almost identical, indicating that with this synthesis method, the BMO to N ratio has a lower or no impact on catalytic performance than with the previous two syntheses.
Fig. S14(b) shows the results of testing the second set of synthesised catalysts. Similar to the first synthesis, the catalyst Pt/N–Mn2O3–NiO (1:
1) (S2) exhibited the lowest onset potential for the reaction, with oxygen evolution starting at 1.14 V. The current density of 10 mA cm−2 with this catalyst was reached at an η10 of 0.56 V. The same current density with N–Mn2O3–NiO (1
:
1) (S2) and Pt/N–Mn2O3–NiO (1
:
2) (S2) was achieved at an overpotential of 0.67 V and 0.57 V, respectively, while with N–Mn2O3–NiO (1
:
2) (S2), the current density of 10 mA cm−2 was not reached. The OER on Pt/N–Mn2O3–NiO (1
:
2) (S2) starts at 1.44 V, while on N–Mn2O3–NiO (1
:
1) (S2) and N–Mn2O3–NiO (1
:
2) (S2), it starts at notably higher potentials of 1.70 V and 1.76 V, respectively. The highest Tafel slope values of 230 and 225 mV dec−1 were obtained for Pt/N–Mn2O3–NiO (1
:
1) (S2) and N–Mn2O3–NiO (1
:
2) (S2), respectively, while for Pt/N–Mn2O3–NiO (1
:
2) (S2) a value of 205 mV dec−1 was obtained. A lower value of b was obtained only with N–Mn2O3–NiO (1
:
1) (S2) (177 mV dec−1).
Finally, the catalytic performance of the catalysts synthesised in the third synthesis was examined, in Fig. S14(c). It can be seen that the reaction starts the earliest on the Pt/N–Mn2O3–NiO (1:
2) (S3) catalyst (onset potential of 1.61 V), followed by Pt/N–Mn2O3–NiO (1
:
1) (S3) (1.64 V). Higher onset potentials of 1.84 V and 1.73 V were observed for N–Mn2O3–NiO (1
:
1) (S3) and N–Mn2O3–NiO (1
:
2) (S3), respectively. Pt/N–Mn2O3–NiO (1
:
1) (S3) reaches 10 mA cm−2 current density earliest (minimum overpotential value of 0.56 V), while Pt/N–Mn2O3–NiO (1
:
2) (S3) reaches 10 mA cm−2 at an overpotential of 0.62 V. The current density of 10 mA cm−2 was not reached with the remaining two catalysts. As in the previous two sets investigations, the lowest Tafel slope value of 128 mV dec−1 was obtained for Pt/N–Mn2O3–NiO (1
:
1) (S3), while values of 195 mV dec−1, 189 mV dec−1, and 197 mV dec−1 were obtained for N–Mn2O3–NiO (1
:
1) (S3), N–Mn2O3–NiO (1
:
2) (S3), and Pt/N–Mn2O3–NiO (1
:
2) (S3), respectively.
EIS is typically used to gain a better insight into the origin of resistance in an electrochemical system. Nyquist and Bode plots of catalysts from all three syntheses in 0.1 M KOH at 1.8 V are shown in Fig. 6. All EIS data were fitted using the equivalent circuit shown in Fig. 6(a), and the comparison of EIS fit parameters is shown in Table S4. Notably, the solution and wiring resistance (Rs) values vary only slightly, from 46 Ω to 76 Ω, indicating the good experimental setup and the consistency of the performed experiments. On the other hand, charge-transfer resistance (Rct) values range from 47.27 Ω for Pt/N–Mn2O3–NiO (1:
1) (S3) to 605.9 Ω for N–Mn2O3–NiO (1
:
2) (S1), supporting the better performance of the Pt-decorated materials compared to the non-decorated BMOs. The two best-performing electrocatalysts, Pt/N–Mn2O3–NiO (1
:
1) (S3) and Pt/N–Mn2O3–NiO (1
:
1) (S2), showed the smallest values of Rct, 47.27 Ω and 64.43 Ω, respectively. Nyquist plots of all investigated materials showed slight flattening of the semicircles, indicating dispersive capacitance,38 which led to using the CPE (constant phase element) for fitting the data instead of the ideal capacitance. CPE is an empirical element that accounts for the capacitive element's non-ideal behaviour due to the surface's inhomogeneities. The effective capacitance C can be obtained from CPE using eqn (3),39,40
C = CPE-T1/CPE-PR(1-CPE-P)/CPE-P | (3) |
Of pure bimetallic oxides tested for the OER, only N–Mn2O3–NiO (1:
1) (S2) reached a current density of 10 mA cm−2, suggesting limited activity of pure oxides toward this reaction. In contrast, Pt-decorated catalysts, particularly Pt/N–Mn2O3–NiO (1
:
1) (S3) and Pt/N–Mn2O3–NiO (1
:
1) (S2), showed excellent performance, all achieving a benchmark current density of 10 mA cm−2 with the latter two requiring an overpotential of 0.56 V. Similar to the ORR results, higher nitrogen doping consistently enhanced catalytic performance for oxygen evolution. This suggests that the nitrogen content remains one of the key factors, enhancing the ECSA and overall material activity, regardless of minor adjustments to the synthesis process.
Investigation of the bifunctional performance of the materials provided better insight into the catalyst's ability to alternate between the two mentioned operating modes in the unitised regenerative fuel cells. As a measurement of the bifunctional performance, ΔE was used, representing the difference between the potential at which the material achieves a benchmark current density of 10 mA cm−2, E10, in the OER potential region and half-wave potential, E1/2, in the ORR potential region. As can be seen from Fig. S14 and Fig. 5, not all materials reached 10 mA cm−2 within the investigated OER potential window, and therefore the determination of the ΔE parameter was not possible for these materials.
The determined ΔE values are presented in Table 3 and in Fig. 5(f). The smallest ΔE was calculated for the Pt/N–Mn2O3–NiO (1:
1) (S2) of 0.90 V, indicating the best bifunctional performance. A somewhat higher value of this parameter was calculated for Pt/N–Mn2O3–NiO (1
:
1) (S3), 0.95 V. Both of these values are comparable with the values that were calculated for the commercial 40 wt% Pt/C catalyst in the authors’ previous work (ΔE = 0.94 V), even though herein synthesised materials contained only 20 wt% of Pt.15 The values were further comparable to that of undoped PtNi/Mn2O3–NiO (0.97 V) tested in the same work.15 However, it should be kept in mind that when evaluating the bifunctional performance, all relevant parameters for both investigated reactions should be taken into account. For example, although Pt/N–Mn2O3–NiO (1
:
1) (S2) showed the smallest value of ΔE, Pt/N–Mn2O3–NiO (1
:
1) (S3) showed a significantly higher value of ORR jd, −4.98 mA cm−2 compared with −3.60 mA cm−2 for Pt/N–Mn2O3–NiO (1
:
1) (S2), and both materials have achieved 10 mA cm−2 in OER mode at the same overpotential of 0.56 V. As authors suggested in their previous work,10 the E1/2 values and, consequently, calculated ΔE values should be taken with some reserve. Therefore, it might be better to use a potential at which a defined value of current density is achieved (e.g., 3 mA cm−3) instead of the E1/2 for the determination of ΔE.
For the rest of the synthesised materials, significantly higher ΔE values were calculated, ranging from 0.99 V in the case of Pt/N–Mn2O3–NiO (1:
2) (S3) up to 1.23 V for N–Mn2O3–NiO (1
:
1) (S2). It should also be mentioned that N–Mn2O3–NiO (1
:
1) (S2) is the only Pt-undecorated material that achieved 10 mA cm−2 in OER mode and the only Pt-undecorated material for which we were able to determine ΔE.
![]() | ||
Fig. 7 Results of the (a) accelerated stress test (AST), (b) switch test, and SEM images (c) before and (d) after the AST of the best performing material, Pt/N–Mn2O3–NiO (1![]() ![]() |
The results of the switch test performed with the Pt/N–Mn2O3–NiO (1:
1) (S3) are presented in Fig. 7(b). This test simulates real-life conditions in which a bifunctional material operates under constant switching between reduction and oxidation potentials. The switch test also demonstrated the remarkable stability of the synthesised material, where the ORR and OER currents, after an initial drop, remained almost unchanged during the ten-hour experiment.
This study demonstrated the successful development of Pt/N–Mn2O3–NiO (1:
1) (S3) bifunctional ORR/OER electrocatalyst by integrating Pt with N-doped bimetallic oxide. The design leverages the synergistic effects between Pt's superior catalytic activity and electrical conductivity, and the tuneable surface chemistry, defect structure, and inherent stability of N-doped BMOs. Mn2O3 contributed mainly through redox mediation and peroxide decomposition, improving ORR selectivity and supporting moderate OER activity. NiO contributed to the enhancement of OER through electronic modification and conductivity, by forming NiOOH, a highly active species for OER. Nitrogen doping in this case led to a decrease in specific surface area; however, this doping introduced surface defects and electronic structure modifications that enhanced Pt anchoring, promoted uniform dispersion, and facilitated charge transfer at the catalyst/support interface. These led to a significantly increased ECSA and reduced overpotentials for both oxygen electrode reactions. Among different identified nitrogen species, the pyridinic-N and graphitic-N particularly decrease the ORR onset potential and lead to higher diffusion-limited current density, respectively. Pyridinic and quaternary nitrogen can also act as active sites for the OER. Nitrogen doping modifies the electronic structure of the Mn and Ni oxides, creating specific Mn–N or Ni–N bonds and active sites, which promote the absorption of intermediates critical for both oxygen electrode reactions. Electrochemical measurements confirmed marked improvements in ORR and OER performance compared to the undoped BMO counterpart, validating the effectiveness of the combined N doping and Pt deposition strategy. Pt/N–Mn2O3–NiO (1
:
1) (S3) also demonstrated excellent long-term catalytic durability due to stabilisation of Pt nanoparticles by BMO and prevention of their agglomeration, detachment, and Ostwald ripening; this underscores high structural and electrochemical stability of Pt/N–Mn2O3–NiO (1
:
1) (S3), a critical requirement for practical deployment in energy conversion systems such as metal–air batteries and regenerative fuel cells.
This journal is © The Royal Society of Chemistry 2025 |