Md Yeasin Pabel,
Muhammed Shah Miran
and
Md. Mominul Islam
*
Department of Chemistry, University of Dhaka, Dhaka, 1000, Bangladesh. E-mail: mominul@du.ac.bd
First published on 28th August 2025
The selective two-electron oxygen reduction reaction (2e− ORR) producing H2O2 through an electrocatalytic approach is an attractive alternative to the industrial anthraquinone oxidation method, enabling decentralized H2O2 production. This study explored the potential use of an inverse nanocomposite comprising polyaniline (PAni)-embedded α-MnO2 nanorods for a selective 2e− ORR in acidic medium. α-MnO2, PAni, and inverse composites of α-MnO2/PAni with different compositions were synthesized using hydrothermal, chemical oxidative polymerization, and solution sonication methods, respectively. Transmission electron microscopy, Fourier-transform infrared spectroscopy, X-ray diffraction, and X-ray photoelectron spectroscopy were employed to clarify various aspects, such as morphology, functional groups, microphase, crystallinity, etc., of the prepared materials. The electrocatalytic activity of the prepared inverse nanocomposites containing minute amounts of PAni dispersed phase towards the ORR in an acidic solution was evaluated with voltammetry at a rotating ring disk electrode, while the rate of H2O2 production was determined through chronoamperometry and the iodometric titration method. The rod-like, inverse α-MnO2/PAni nanocomposite catalyst performs commendably for the 2e− ORR with, on average, 80% selectivity toward H2O2 at a production rate of 433 mmol gcat−1 h−1. A plausible mechanism for the ORR on the α-MnO2/PAni nanocomposite in an acidic medium is discussed. These findings highlight the promise of inverse α-MnO2/PAni nanocomposites as affordable and efficient electrocatalysts for on-site production of H2O2 through the ORR.
Researchers have mainly focused on commercially feasible catalyst-based transition metal oxides, which offer excellent catalytic activity towards the oxygen reduction reaction (ORR).10,19–24 Among them, manganese oxides (MnOx) have garnered significant interest due to their notable benefits, including abundance, cost-effectiveness, environmental sustainability, and substantial catalytic efficacy in facilitating the electrochemical ORR.21–23,25–27 However, most research has focused on studying the ORR on MnOx under alkaline conditions, indicating that MnOx typically supports the 4e− ORR.21–23,28–30 Ohsaka et al. have reported that the ORR on MnOOH encompasses an electrochemical process, succeeded by the subsequent disproportionation of electrochemical reduction intermediates, specifically O2− and HO2−, resulting in a quasi 4e− pathway.23 Tang et al. have showed that MnOx characterized by higher Mn valence states, like γ-MnOOH and Mn(OH)4 composites, exhibits enhanced catalytic efficiency in the ORR when contrasted with counterparts possessing lower Mn valences.29 The catalytic activities of MnO2 are notably affected by its crystallographic structures in the order of α > β > γ.21,31 Moreover, α-MnO2 nanowires have been reported to demonstrate size-dependent functionality, where their nano-form shows a superior performance in comparison to microparticles.30 However, the activity and selectivity of the ORR on MnOx depend on the valence states of Mn and its crystallographic phases.21,23,25,28–30 The mechanism of the ORR either to be 2e− or 4e− also depends on adsorption orientation and the binding energy of O2 to the active sites of the catalysts,18,20 regulating the selectivity of the ORR (Scheme 1). There are three adsorption modes for O2 on the catalyst surface as illustrated in Scheme 1A–C. In the Griffiths mode, two oxygen atoms bind on a single site, usually leading to dissociative adsorption and water formation. The Pauling mode involves one oxygen atom binding with partial charge transfer, resulting in peroxide formation. The bridge mode is similar to the Griffiths model but requires two adsorption sites. The adsorption mode depends on the nature of the catalyst surface, while the ultimate product is influenced by both binding energy and modes of adsorption of O2. However, there are several challenges for efficient electrochemical production of H2O2 using an MnOx modified electrode in acidic medium. One challenge is the insufficient electrical conductivity of the MnOx modified electrode because of the insulating nature of MnOx.25,32 Another challenge is poor adhesion of the MnOx catalysts to the electrode surface.25 Additionally, the concentration of O2 is limited due to its low solubility in aqueous solution (ca. 1 mM) at the electrode surface.14,15,33,34 These issues can be addressed by integrating MnO2 with conducting polymeric materials such as polypyrrole (PPy), polythiophene, polyaniline (PAni), or poly(vinyl alcohol).22,26,27,35–37 The electrochemically prepared MnO2/PAni and MnO2/PPy were studied for the ORR in Na2SO4 and H2SO4 solutions, respectively.22,36 These studies used only the cyclic voltammetric technique and proposed a probable mechanism based solely on the observed activity. However, the specific selectivity of the ORR on MnO2/PAni, whether it involves a 2e− or 4e− reduction process (Scheme 1E), remains ambiguous.
The composites of MnO2 using a minute amount of PAni, called inverse composites,38,39 wherein MnO2 and PAni are the matrix phase and dispersed phase, respectively, would offer outstanding properties required for selective and efficient catalysis of the ORR for enhanced production of H2O2. In fact, the concept of formation of an inverse composite, an intriguing class of materials,38 is relatively new, and it provides extreme adhesion promotion of the metal-based matrix phase, i.e., MnO2 required for developing a highly stable, robust electrode. In addition, inverse nanocomposites containing a higher MnO2 content compared to PAni may be designed with an appropriate adsorption orientation of molecular O2 for the selective ORR (Scheme 1D). In our previous study, the enhancement and a way of tuning the solubility of O2 have been achieved by creating a hydrophobic micelle core that actually serves as a nest for O2 in aqueous solutions, e.g., solubility enhanced by 35-times using 24 mM sodium dodecyl sulfate surfactant.33 Therefore, the hydrophobic polymer moiety-embedded in the inverse composites with the desired metal oxides are thus thought to potentially boost the local concentration of O2, regulate the adsorption modes of O2 on the catalyst surface required for selective ORR, and address the mentioned limitations of using MnO2 as an electrode material. These features are obviously lucrative for ORR catalyst design and hence motivated us to comprehensively perform ORR on the inverse MnO2/PAni nanocomposites.
Herein, we report, for the first time to the best of our knowledge, the synthesis and investigation of an inverse MnO2/PAni nanocomposite for efficient and selective H2O2 production. This work reveals changes in the electronic structure of Mn upon incorporation of PAni, as evidenced by atomic-level characterization using XPS. The ORR activity was evaluated under static conditions using cyclic voltammetry and under hydrodynamic conditions using a rotating ring disk electrode. Additional electrochemical and kinetic insights were obtained through analysis of Tafel plots and Koutecky–Levich plots, etc. The nanocomposite demonstrates ∼80% selectivity toward the 2e− ORR pathway in acidic media, with excellent operational stability. The incorporation of PAni with MnO2 is proposed to facilitate selective 2e− ORR. These findings highlight MnO2/PAni nanocomposites as promising electrocatalysts for on-site H2O2 generation. Furthermore, this approach can be extended to other oxide-polymer systems, offering new possibilities for catalyst development in the ORR and other catalytic applications.
The simplicity of the FTIR spectrum of MnO2 (Fig. 2A-a) suggests a higher symmetry in its structure. In the far-infrared region (400–1000 cm−1), the bands result from vibrations of octahedral MnO6 units within the α-MnO2 lattice.41 The band at 750 cm−1 is specifically linked to the asymmetric stretching mode of the MnO6 octahedral unit.44 However, the bands at 1642 and 3450 cm−1 are attributed to the presence of minute amounts of water in the prepared α-MnO2. On the other hand, the FTIR spectrum of PAni possesses distinctive bands at 1620 and 1430 cm−1, as shown in Fig. 2A-b and c, corresponding to the CC stretching vibrations of the benzenoid and quinonoid rings in PAni.26,37 The band around 1142 cm−1, attributed to C–N stretching, serves as a characteristic indicator of the conductivity of PAni, particularly in its emeraldine salt form.37 Furthermore, the observed bands at 2912 and 2852 cm−1 are signatures of the formation of hydrogen bonding inside the PAni phase.37 The spectrum of α-MnO2/PAni (Fig. 2A-c) reveals the presence of all the characteristic bands associated with both α-MnO2 and PAni, affirming the successful preparation of the composite. The band at 520 cm−1 for MnO2 in MnO2/PAni shifted to a higher frequency region. Moreover, a new band emerges at 952 cm−1 in the spectrum of the MnO2/PAni composite, as shown in Fig. 2A-c. The development of a new band, along with the observed shifting of the original bands of the constituting components, is considered as the marker of the formation of the composite.37
Fig. 2B shows the XRD patterns of α-MnO2 and α-MnO2/PAni. The peaks at 2θ = 12.5°, 17.8°, 25.5°, 28.6°, 37.4°, 41.9°, 49.7°, 55.9°, 60.1°, 65.3°, 69.3°, 72.9° and 77.4° observed for the MnO2 nanorods (Fig. 2B-a) are for the crystal planes of (110), (200), (220), (310), (211), (301), (411), (600), (512), (002), (541), (312) and (402), respectively, of the tetragonal phase of MnO2.40–42 The absence of distinctive peaks associated with other phases of manganese oxides or impurities indicates the high performance of the hydrothermal method in synthesizing pristine tetragonal MnO2. The average crystallite size (d) of α-MnO2, calculated from the width of the (211) peak using the Scherrer formula d = 0.9λ/β1/2cos
θ, where λ is the X-ray wavelength, β1/2 is the corrected width of the main diffraction peak at half height, and θ is the diffraction angle, was ca. 37 nm. In the diffraction patterns of the MnO2/PAni composite shown in Fig. 2B-b, a slight change of the (220) peak of pure MnO2 occurred due to the overlapping of the (200) peak of PAni. In addition, a low-intensity peak positioned at 2θ = 20° associated with the (020) plane of PAni was observed.
So far, the synthesis of inverse MnO2/PAni nanocomposites has been successful as confirmed by the results of TEM, FTIR spectra and XRD analyses. While low-valent states of transition metals in catalyst families such as mixed oxides, spinels, bronzes, polyoxometalates, and metal–organic frameworks have often been associated with enhanced catalytic activity, such enhancement is typically attributed to the formation of electron-rich centers that facilitate charge transfer, generation of vacancies, electronegative doping, or synergistic effects in composite systems that enable back-donation of electrons.25,29,45–48 However, in the case of MnO2 for the ORR, particularly under acidic conditions, higher oxidation states of Mn are generally more favorable.21,23,28–30 These oxidized states (e.g., Mn4+) provide greater structural stability, suppress Mn dissolution, and support sustained redox cycling,21,23 whereas reduced Mn species (Mn2+ and Mn3+) are often unstable and prone to leaching, as discussed later. In this study, the observed development of a new band and shifting of bands of the FTIR spectrum of the inverse composites direct the presence of additional electronic interaction that must be specified and quantified further. In this regard, XPS is a powerful tool to identify the elemental information that was adopted for characterization of the materials studied.49
Fig. 3A depicts the XPS survey spectra of MnO2 and the inverse MnO2/PAni nanocomposites, revealing peaks attributed to the expected elements, i.e., Mn, O, C, and N. In the high-resolution XPS spectrum of Mn 2p present in pure MnO2, two distinguishable peaks at 642.2 and 653.8 eV are observed as shown in Fig. 3B, corresponding to Mn 2p3/2 and Mn 2p1/2 of the α-MnO2 nanorods, respectively, exhibiting a spin energy separation of 11.6 eV.50,51 This observation aligns well with both the SAED (Fig. 1D) and XRD patterns (Fig. 2B-b), indicating the successful synthesis of α-MnO2 nanorods. Previously, Pereira et al. have obtained analogous peaks at 642.2–642.25 eV and 653.8–654.05 eV for Mn 2p3/2 and Mn 2p1/2, respectively, with a spin energy separation of ca. 11.60 eV for α-MnO2 nanowires.50 On the other hand, remarkable shifting of these peaks towards higher values was observed in the case of the inverse MnO2/PAni nanocomposites, where the peaks corresponding to Mn 2p3/2 and Mn 2p1/2 are observed at 649.8 and 661.7 eV, respectively, with a spin energy separation of 11.9 eV (Fig. 3C). These shifts in peak position towards higher energy levels, along with the increased spin energy separation, are likely attributable to the phenomenon of back donation of the electron from MnO2 to PAni.52 Such an electron transfer prompts a redistribution of electron density around the metal ions, leading to a stronger electrostatic attraction between the Mn 2p electrons and the nucleus of Mn, consequently shifting the binding energy of the electrons to higher energy levels.49 It is also plausible that the interaction with PAni may not be uniform across all the orbitals of Mn in the composite, resulting in a larger apparent spin energy separation. Fig. 3D shows the high resolution XPS spectrum of N 1s present in PAni of the MnO2/PAni nanocomposite exhibiting three peaks at 399.5, 400.8 and 402.2 eV, corresponding to the quinoid imine, benzenoid imine and positively charged imine, respectively.53,54 The presence of different electronic states of N was also supported by the FTIR spectrum (Fig. 2A). These observations suggest that PAni is in a protonated state. The presence of PAni in MnO2 could serve a dual purpose: firstly, aiding in the adsorption of hydrophobic O2 on the surface of the catalyst, and secondly, enhancing the activity and selectivity of the 2e− ORR.
The CV measurements for the ORR at all electrodes were further conducted at various potential scan rates (ν). Fig. 4B shows the plots of cathodic peak current densities (jp) vs. square root of ν (ν1/2). These plots are linear, passing through the origin with the regression coefficients of 0.97, 0.99, 0.94 and 0.99 at MnO2, MnO2/PAni-10, MnO2/PAni-15, and MnO2/PAni-20 modified GC electrodes, respectively. Thus, the ORR studied using these inverse MnO2/PAni composites is an irreversible, diffusion-controlled process as generally observed in aqueous solutions at metal, non-metal or modified electrodes.14,15 However, the slopes of the straight lines (Fig. 4B) that indicate the number of electrons involved (n) in the ORR, diffusion coefficient (DO2) and saturation concentration of O2 (CO2) varied according to the well-known Randles–Sevcik equation (eqn (1)) as follows:
jp = 2.69 × 105n3/2CO2(DO2ν)1/2 | (1) |
In general, the slopes obtained for the inverse composites are higher than that obtained with MnO2, and its magnitude increased as the amount of PAni was increased. With the help of eqn (1), one may determine the product of the DO2 and CO2 from the obtained slope if the number of electrons involved is known. By considering the average number of electrons, n, involved in the ORR as 2, the values of the product of CO2 and were determined as 8.1 × 10−9, 10.2 × 10−9, 11.4 × 10−9, and 12 × 10−8 mol cm−2 s−1/2 for MnO2, MnO2/PAni-10, MnO2/PAni-15, and MnO2/PAni-20 modified GC electrodes, respectively, indicating enhanced solubility of O2 with increasing PAni.33,34 Nissim et al. have developed a simple and generally applicable electrochemical approach to ascertain the solubility of O2 in a micellar system and found that CO2 increased as the concentration of surfactant increases in the micellar system.34 In our previous study, we employed cyclic voltammetry to evaluate the characteristics of the ORR in micellar environments.33 A notable increase of CO2 approximately 15-times higher compared to an H2SO4 solution, was achieved by using SDS in H2SO4 solution with a 4.5-times decreased DO2. This increased solubility of O2 was attributed to the hydrophobic nature of the core of the micelles, potentially acting as a high-density O2 reservoir.33
The cyclic voltammetric study reveals that the ORR can be carried out on the inverse MnO2/PAni nanocomposites, and the polymeric micro-domain enhanced the solubility of O2. To evaluate detailed catalytic features and the selectivity towards the ORR of the materials under consideration, the RRDE system was employed to quantify the number of electrons involved, Tafel slope, percentage of H2O2 production, etc., for the ORR at the inverse α-MnO2/PAni nanocomposites. Fig. 5A shows the characteristic linear sweep voltammograms (LSVs) for the ORR measured at different catalyst-modified GC disks of the RRDE, where the lower set of curves refers to the oxygen reduction current and the upper set refers to the generated H2O2 oxidation current on a Pt ring electrode. The measured LSVs at the modified disk shown in Fig. 5A do not exhibit an oxygen reduction current reaching a diffusion-limited current plateau as observed in the corresponding CVs (Fig. 4A). This phenomenon, commonly observed in various studies under acidic conditions, suggests merging the ORR with the HER process.14 On the other hand, the LSVs measured at the Pt ring exhibit the expected ideal shape with the current plateau occurring for the oxidation of the product generated at the disk, i.e., peroxide species. However, for comparison, the current densities at 2000 rpm are shown in Fig. S1, illustrating a decreasing trend in the order of MnO2/PAni-20 > MnO2/PAni-15 > MnO2/PAni-10 > MnO2. The ring current corresponding to the MnO2/PAni-20 disk electrode (Fig. 5A-c′) decreases with an increasing disk potential after −0.7 V. This could be attributed to the pronounced generation of hydrogen at this disk electrode, which flushes out the produced H2O2 beyond this potential, as indicated in the CV (Fig. 4A). The low onset potentials and reduction currents at the MnO2/PAni modified electrodes as well as the corresponding ring current indicate a 2e− pathway ORR on the catalysts, which will be further clarified later.
To evaluate the ORR kinetics at the MnO2/PAni catalysts, the rotation-rate dependent LSVs were recorded at 800 to 2000 rpm (Fig. S2–S4). The typical Koutecký–Levich (K–L) plot of j−1 vs. ω−1/2 derived from the measured LSVs is presented in Fig. 5B. The good linearity and parallelism of these plots indicate the first-order kinetics with respect to molecular oxygen.24,55 The decreasing trend of the slopes of the Tafel plots is as follows (Fig. 5C): −277 mV dec−1 for MnO2/PAni-10 > −204 mV dec−1 for MnO2/PAni-15 > −158 mV dec−1 for MnO2/PAni-20. The values of the transfer coefficient (α) are also determined. Moreover, from the intercept of the Tafel plots, exchange current density (j0) was estimated for evaluating the values of the standard heterogamous rate constant (k0). The values of onset potential, E, α, j0 and k0 obtained for the ORR at three different electrode materials are compared in Table 1.
The value of k0 varies when we change the ratio of metal oxide to polymer in the inverse composite, along with shifts in the onset potential (Table 1). These fluctuations of k0 imply the existence of distinct reaction mechanisms occurring at the surface of different catalysts. The reliance of k0 on both the onset potential and metal oxide loading again underscores the intricate nature of the ORR occurring at the MnO2/PAni interface. At the potential region at which the ORR was studied on the composites, no potential-dependent redox reaction of PAni takes place, and only the emeraldine salt (ES) phase of PAni can persist.56 While the structure of MnO2 may be influenced by the applied potential in acidic environments.57 The ES phase of PAni and the potential dependent structure of the MnO2 rods in the composite may modulate the adsorption mode of O2 (Scheme 1) for a selective and faster electrode reaction, producing H2O2 as evidenced by the enhanced ring current of the LSVs shown in Fig. 5A. Consequently, the cumulative effects of these factors, either individually or in combination, may contribute to the observed variations in k0. The mechanism of this catalytic performance of the composites towards the ORR and the selectivity towards high percent production of H2O2 were revealed by analysis of the LSVs using eqn (2) and (3).
![]() | (2) |
![]() | (3) |
MnO2 + H2O + e− ↔ MnOOH + OH− | (4) |
MnOOH + O2 ↔ MnOO⋯O2(ads.) | (5a) |
2MnOOH + O2 ↔ 2(MnOOH–O(ads.)) | (5b) |
MnOOH⋯O2(ads.) + e− ↔ MnO2 + HO2− | (6a) |
2(MnOOH⋯O(ads.)) + 2e− ↔ 2MnO2 + 2OH− | (6b) |
The overall reaction of eqn (4), (5a) and (6a) results in a 2e− reduction process, while the combination of eqn (4), (5b) and (6b) results in an overall 4e− transfer. These two schemes differ in the manner of O2 adsorption. Cheng et al. have reported that both processes could occur on MnO2, depending on the structure of MnO2.21 Moreover, the surface structure of PAni changes with pH, and under acidic conditions it mainly exists in the protonated ES form, which is more conductive and active.59 This helps improve ORR performance (reaction (7)), but the exact surface changes during the reaction are still not fully understood and need further investigation. The ES phase of PAni reacts with MnO2 to form MnOOH and pernigraniline (PN) as shown in eqn (7).22 Consequently, at the MnO2/PAni surface, reaction (4) could be replaced by the following reaction (7). Furthermore, Khomenko et al. have suggested that during the ORR on a PPy-modified graphite electrode, O2 molecules can interact with electron-donor N atoms in PPy as well as with MnOOH (reaction (8)).23,35,58 In the acidic medium, HO2− is stable and desorbs from the catalyst surface.23
([C6H4NH]2[C6H4N]2)n(PAni-ES) + MnO2 ↔ MnOOH + (C6H4N)n(PAni-PN) | (7) |
PAni⋯O![]() | (8) |
The percentage of H2O2 production at different electrodes for different potentials is shown in Fig. 5E. It has been observed that the percentage of H2O2 production is also dependent on the applied voltage.5 The number of electrons involved and the selectivity of H2O2 exhibit a continuous increase at increasingly negative potentials (Fig. 5D and E), suggesting a gradual reduction of the generated H2O2 to H2O at the electrode surfaces. However, the average values of selectivity to H2O2 production in the potential range of −0.5 to −0.75 V for MnO2/PAni-10, MnO2/PAni-15, and MnO2/PAni-20 are equal to 82, 71, and 70%, respectively, demonstrating a high-percentage production of H2O2. Table S1 compares the variations of these characteristic parameters of the ORR studied with different catalysts in both acidic and basic solutions. A notable trade-off is observed between the ORR kinetics and H2O2 selectivity across the MnO2/PAni composites. While increasing the PAni content enhances the apparent electron transfer rate (Fig. 5A and Table 1), it does not lead to a proportional increase in H2O2 selectivity (Fig. 5E). This suggests that the conductive polymer, beyond a certain threshold, may alter the local electronic structure or adsorption environment at the MnO2 interface, affecting the binding modes of oxygen intermediates and shifting the reaction pathway. These results highlight the necessity of optimizing the composite architecture to simultaneously maintain efficient electron transfer and high selectivity toward the 2e− ORR pathway. Efficient H2O2 production was also confirmed by chronoamperometric measurements in a two-compartment cell (Fig. S5), and the produced H2O2 was determined by iodometric measurements. The calculated result shows that the production rate of H2O2 is 433 mmol gcat−1 h−1 at the MnO2/PAni-15 modified GC electrode in O2-saturated 0.05 M H2SO4 solution.
The catalytic stability is a key factor in catalytic performance. In acidic media, MnO2 is first reduced to MnOOH (Mn3+) via proton insertion and electron uptake (reaction (4)),21 and this unstable Mn3+ species then either reacts with O2 (reactions (5) and (6))23,58 or undergoes disproportionation, yielding soluble Mn2+ and regenerating MnO2, leading to dissolution.60 However, in low-concentration acid (0.05 M H2SO2), proton activity is reduced, slowing the dissolution pathways (reaction (4)). Additionally, PAni acts as a redox mediator (reaction (7)), stabilizing Mn valence states and preventing Mn3+ buildup, while its protonated ES form buffers local acidity, together effectively suppressing Mn2+ leaching during cycling.27 Fig. 5F illustrates the initial (a) and final (b) CVs of the MnO2/PAni-15 modified GC electrodes after undergoing 1000 cycles (inset of Fig. 5F) in an O2-saturated 0.05 M H2SO4 solution. The cyclic voltammetric studies were performed at a scan rate of 10 mV/s and a rotation speed of 1000 rpm. The absence of a discernible peak in the CV can be attributed to the experimental conditions, specifically the dynamic environment and the utilization of a large electrode surface. During the measurement, the half-wave potential exhibits minimal change, with only a small reduction in diffusion-limiting currents. This property can be ascribed to the interaction between the catalyst and the electrode as well as the protective nature of PAni in acidic solutions.27 Zhang et al. also reported the stability of PAni/MnO2/MWCNTs in acidic solutions, attributing this stability to the protective coating of PAni.27
In this study, XRD measurements were carried out to reveal the basic structural information and the determination indexing of the crystal planes of the materials studied. For this, the experimental X-ray diffraction patterns given in this study were matched with appropriate entries from the Powder Diffraction File database provided by the International Centre for Diffraction Data. No structural refinement was, however, conducted that would necessitate to generate a Crystallographic Information File. Hence, no data is available for deposition in a repository system. Other data will be made available on request.
This journal is © The Royal Society of Chemistry 2025 |