Abstract

Electrocatalytic water splitting is a cost-effective and environment-friendly technique for producing oxygen and hydrogen through the oxygen/hydrogen evolution reaction (OER/HER). Developing the highly active and stable electrocatalyst, particularly for bifunctional water electrolysis (i.e., both OER and HER), is still a formidable challenge. Herein, we demonstrated the enhanced bifunctional water splitting activities by utilizing the molybdenum trioxide-anchored activated carbon (MoO3/AC) nanocomposites. The MoO3/AC samples were fabricated by the ultrasonication method using sol-gel synthesized MoO3 and biomass-derived AC, and they displayed a nanostreusel-like morphology with spherical MoO3 nanoparticle-decorated AC nanosheets. For the water electrolysis test, the MoO3/AC nanocomposites exhibited the excellent bifunctional electrocatalytic OER and HER performances with low overpotential and small Tafel slope values. Through analyzing the material characteristics and the electrochemical properties of MoO3/AC, it was found that the superb bifunctional OER-HER activities were attributed to the synergistic effects from the hybridization of highly conductive AC and electrochemically active α-MoO3. The results pronounce that the MoO3/AC nanocomposites possess an aptitude as a superb bifunctional OER/HER electrocatalyst for high-performance water electrolysis.

1. Introduction

Due to its zero-carbon emission as well as high energy density, hydrogen has emerged as a promising green energy source that can alternate the fossil fuels and release the environmental crisis [13]. Electrocatalytic water splitting is a fascinating technique for producing the renewable hydrogen energy source from water because the method possesses many advantages such as ecofriendliness, recyclability, and sustainability [48]. Water electrolysis consists of two basic half-cell reactions at the anode and the cathode; i.e., the oxygen evolution reaction (OER) takes place at the anode, and in turn, the hydrogen evolution reaction (HER) occurs at the cathode [911]. For OER and HER, Ir/Ru- and Pt-based materials were typically used as electrode materials because of their superior electrocatalytic performances [1215]. However, the low natural abundance and the high cost of both novel metal and rare earth material would somewhat restrict their practical applications [1619]. Meanwhile, it is well known that OER is less effective than HER because of its sluggish reaction kinetics and high energy consumption [2023]. To perform the high electrocatalytic water splitting activities of both OER and HER, therefore, developing a highly efficient bifunctional electrocatalyst is vital. In practical application point of view, particularly, finding a highly durable and low-cost electrode material is of great essence.

Recently, transition metal oxides (TMOs) have garnered ample attention as a proficient catalyst because of their high electrochemical activity and cost-effectiveness [2428]. Among various TMOs, molybdenum trioxide (MoO3) has emerged as an efficient electrocatalyst because of its double-layer planar structure, nontoxicity, high stability, and intrinsic characteristics [2933]. Crystallographically, MoO3 possesses three different types of crystalline phases (i.e., α, β, and h) [3436]. Among them, orthorhombic α-MoO3 consists of the octahedral MoO6 double-layer structure that performs high electrochemical reactivity and good thermodynamic stability [37, 38]. These play a key role for promoting the oxygen-associated reaction during the electrocatalytic water splitting process [33, 36]. Accordingly, many research groups have attempted to fabricate high-quality α-MoO3 by utilizing several experimental methods (e.g., microwave-assisted chemical synthesis [39], ultrasonication [40], water-based exfoliation [41], hydrothermal growth [38, 42], wet chemical process [43, 44], sol-gel [4547], and ball milling [48]). Despite such substantial efforts, the electrocatalytic performances of α-MoO3 are still unsatisfiable because of its small electrochemically active site, poor ionic conductivity, poor electronic conductivity, and sluggish kinetics [33, 49]. Aiming at improving the electrocatalytic activity of α-MoO3, therefore, it was intensively tried to materialize the α-MoO3-based nanocomposites by hybridizing with various carbonaceous nanostructures (e.g., carbon nanotube [50], graphene [51], graphene oxide [52], and activated carbon (AC) [53]). Among those, AC is of good use for fabricating the MoO3/AC nanocomposites as a high-performance OER and HER catalyst [53]. This is because high conductivity and large surface area of AC could help enhance the electronic charge transfer and the ionic motion, respectively. Generally, AC could be synthesized by two different approaches, i.e., physical (thermal) and chemical activation processes. In physical activation, initially, the raw material was carbonized and subsequently gasified at higher temperatures (500–1000°C) in a stream of oxidizing gas (e.g., air, steam, and CO2) [54, 55]. For chemical activation, the precursor material is mixed with the chemical agents (e.g., KOH, ZnCl2, NaOH, H2SO4, LiOH, and H3PO4) and followed by annealing at moderate temperatures (400–700°C) [5658]. In both cases, the textural characteristics and the pore size distribution of AC can also be modulated by changing the precursors, ambient gases, and activation agents [59]. When using biomass-derived AC, furthermore, there are additional advantages such as a cost-effectiveness, an ecofriendliness, a massive abundance, and a fast regeneration [6063]. For instance, Hoang et al. [64] prepared the hybrid Ni/NiO nanocomposite with N-doped biomass AC and exhibited an overpotential of 346 mV (for OER) and 180 mV (for HER) in 0.1 M KOH. Sangeetha et al. [53] prepared the biomass AC-decorated h-MoO3 nanocomposite via the hydrothermal method and showed an overpotential of 282 mV and Tafel slope of 169 mV/dec for HER in 0.5 sulphuric acid. Kou et al. [65] synthesized the NiO-based nanocomposites decorated with nitrogen-doped biomass AC dots and demonstrated an overpotential of 380 mV at 10 mA/cm2 for OER. Yaseen et al. [66] fabricated the hierarchical Co/MoO2@N-doped biomass AC nanocomposites via the simple annealing process and exhibited the excellent OER and HER activity with a low overpotential of 272 mV and 130 mV in 1 M KOH. In spite of all the above benefits from both MoO3 and biomass AC, the electrocatalytic activity of the MoO3/AC nanocomposites has rarely been investigated [53].

All these prompt us to study on the fabrication of robust MoO3/AC that could act as an excellent bifunctional OER/HER electrocatalyst. In this study, we synthesized and characterized the MoO3/AC nanocomposites that were devised through facile ultrasonication by utilizing the sol-gel synthesized MoO3 nanoparticles and the biomass human hair-derived AC nanosheets. From the fabricated MoO3/AC nanocomposites, we found that the excellent bifunctional electrocatalytic water splitting activities could be achievable. For example, low overpotential values (OER: 280 mV at 10 mA/cm2 and HER: 353 mV at 10 mA/cm2) were recorded from MoO3/AC. Herein, the material characteristics and the electrocatalytic performances are thoroughly assessed and discussed in detail.

2. Experimental Details

2.1. Growth of MoO3 Nanoparticles

Figure 1 represents a schematic diagram of the fabrication procedure for obtaining high-quality MoO3/AC. Firstly, the α-MoO3 nanoparticles were synthesized through the sol-gel process by using ammonium heptamolybdate ((NH4)6Mo7O24) and nitric acid (HNO3). As a primary task, (NH4)6Mo7O24 (6 g) was added into deionized (DI) water (30 mL). After continuous stirring for 10 min, HNO3 (15 mL) was injected drop-by-drop into the above aqueous ammonium heptamolybdate solution and subsequently stirred for 4 h at 100°C in a sand bath. After cooling down to 27°C, the mixture solution was filtered and washed five times in DI water and ethanol. Then, the suspension was dried for 10 h at 80°C. Finally, the α-MoO3 nanoparticles were obtained by performing postgrowth thermal annealing at 700°C for 3 h. Here, we note that no additional purification was performed for all the used materials that were purchased from Sigma-Aldrich (St. Louis, MO, USA).

2.2. Derivation of AC Nanosheets

The AC nanosheets were derived from the biomass resource of human hair (HH) through thermal carbonization and KOH activation. Initially, the collected biomass HH bundles were cleaned and rinsed with DI water (three times for each) and dried in air for 24 h. After that, the carbonization of HH was carried out in air at 300°C for 60 min. Then, the mixture of carbonized HH ashes (5 g) and KOH (20 g) was prepared by utilizing a mortar. After subsequent annealing at 600°C for 120 min in a muffle furnace, the KOH-activated HH ashes were rinsed for 12 h in DI water to extract potassium and its related precipitates. Finally, the pure HH-derived AC nanosheets were collected via filtrating, rinsing, and drying (120°C, 12 h). The experimental procedures in more detail for obtaining high-quality HH-derived AC were reported elsewhere in our previous study [18].

2.3. Fabrication of MoO3/AC Nanocomposites

The nanocomposites of MoO3/AC were fabricated by the facile ultrasonication method. Initially, the MoO3 nanoparticles (1 g) were blended and stirred in DI water (100 mL) for 15 min. Next, the AC nanosheets (0.5 g) were added into the above MoO3-blended aqueous solution and stirred for 30 min. Subsequently, the MoO3-AC mixture solution was sonicated for 1 h (, ). During the ultrasonication step, the high-power ultrasound supplies an excessive energy enough to cleave and break the big clusters into the tiny species [6769]. Then, the dispersity of the nanoparticles could be increased in the entire composite system. After sonication, the powder-type MoO3/AC nanocomposites were obtained via sieving, cleaning, rinsing, and drying (120°C, 8 h) of the prepared colloidal suspension.

2.4. Material Characterization

The structural and the vibrational properties of the prepared materials (i.e., MoO3, AC, and MoO3/AC) were examined by X-ray diffractometry (XRD) using an Ultima IV system (Rigaku, Tokyo, Japan) and Raman scattering spectroscopy using a LabRAM HR-800 system (Jobin Yvon, Longjumeau, France), respectively. In addition, the morphological structure and the chemical composition of MoO3 and MoO3/AC were monitored by field emission scanning electron microscopy (FE-SEM) and in situ energy-dispersive X-ray spectroscopy (EDX), respectively, using an Inspect F50 system (FEI, Mahwah, NJ, USA). Furthermore, the microstructural properties of the samples were characterized by transmission electron microscopy (TEM) using a JEM 2100F system (JEOL, Tokyo, Japan).

2.5. Electrocatalytic Measurements

The water electrolysis characteristics of MoO3 and MoO3/AC were assessed by using a VersaSTAT 3 workstation (Ametek Scientific Instruments, Berwyn, PA, USA). To examine the electrocatalytic OER and HER performances, as a primary task, the conventional three-electrode system was prepared in a KOH solution (1 M). Firstly, the working electrodes were devised by using the bare MoO3 nanoparticles and the MoO3/AC nanocomposites. For this step, initially, each catalyst (i.e., either MoO3 or MoO3/AC) was blended with N-methyl-2-pyrrolidinone and coated onto the stainless steel substrate (). Thereafter, the prepared substrates were dried for 8 h at 180°C to cure the active materials. To configure the three-electrode system, the counter electrode of platinum mesh (Nilaco Co., Tokyo, Japan) and the reference electrode of saturated calomel (Sanxin, Shenyang, China) were also prepared. Then, the cyclic voltammetry (CV) measurements were conducted at the potential window of 0.0–0.5 V under various scan rates of 10–100 mV/s. Additionally, the linear sweep voltammetry (LSV) characteristics were measured at potential windows of -0.1–1 V (for OER) and -1.8–0 V (for HER) under the scan rate of 1.0 mV/s. Furthermore, the chronopotentiometric (CP) characteristics of the prepared materials were investigated upon varying the applied current density (i.e., “10 → 20 → 30 → 40 → 50 → 100 mA/cm2 for OER” and “-10 → -20 → -30 → -40 → -50 → -100 mA/cm2 for HER”). Here, the working electrodes were polarized at every current densities, and each polarization state at each current density was maintained for 10 min. To examine the electrical properties of MoO3 and MoO3/AC, furthermore, the electrochemical impedance spectroscopy (EIS) characteristics were also evaluated at 1 Hz–10 kHz.

3. Results and Discussion

Figure 2(a) shows the crystallographic properties of MoO3 and MoO3/AC. In both samples, the polycrystalline XRD patterns were detected at 12.78, 23.42, 25.74, 27.33, 33.71, 35.50, 39.02, 45.67, 46.34, 49.26, 58.83, 64.47, and 67.53°, which were sprouted from the lattice planes of (020), (110), (040), (021), (111), (041), (060), (200), (210), (002), (081), (062), and (0100) of orthorhombic α-MoO3 (JCPDS card no.: 05-0508 [7073]), respectively. No secondary phases in both MoO3 and MoO3/AC indicate the high crystal quality of the synthesized samples. Here, we notice that the absence of carbon-related peaks in MoO3/AC is thought as resulting from the small volume of the AC nanosheets in the entire MoO3/AC composites. The average crystallite distribution sizes were calculated to be 45 and 32 nm for MoO3 and MoO3/AC, respectively, by using Scherrer’s formula, as follows [74, 75]: where , , , and are the Scherrer constant, Bragg angle, X-ray wavelength, and full width at half-maximum in radian, respectively. Such a result is thought as attributing to the sonochemical exfoliation of the MoO3 particles (i.e., interconnected nMoO3 → exfoliated MoO3(n)), as discussed later.

The vibrational properties of MoO3 and MoO3/AC were further characterized by the Raman scattering spectroscopy. As can be seen from Figure 2(b), both MoO3 and MoO3/AC exhibited the Raman peaks at 122, 150, 190, 214, 239, 282, 334, 374, 466, 661, 814, and 992 cm-1, arising from the local vibrations in orthorhombic α-MoO3 [76]. The Raman scattering mode at 992 cm-1 originates from the stretching vibration of terminal oxygen (Mo6+=O), which is indicative of the layered α-MoO3 structure [77]. The strong Raman peak at 814 cm-1 comes from the stretching vibration of bicoordinated oxygen (Mo-O-Mo), which is corner-shared in two MoO6 octahedra [78]. The peak at 661 cm-1 corresponds to the stretching vibration mode of triply coordinated oxygen (Mo3-O), where the edge-shared oxygen atom mutually exists with three octahedra [79]. Asymmetric O-Mo-O bending and stretching are observed at 466 cm-1. The Raman scattering modes at 374 and 334 cm-1 are associated with the O-Mo-O scissoring and bending vibrations, respectively [80]. The vibrational modes at 282, 239, 214,190, 150, and 122 cm-1 are credited to the wagging and twisting modes of O=Mo=O as well as MoO4 rotational rigid chains [76, 81]. In the case of MoO3/AC, one can see that there are two additional Raman scattering features at 1339 and 1596 cm-1. Through double-checking with the bare AC nanosheets (Figure 2(b)), the former and the latter were confirmed to originate from the D and G bands of graphitized 2D AC, respectively [48, 63]. The D band arises from the disordered graphitic AC, and the G band belongs to the E2g vibrational Raman scattering mode at sp2-hybridized carbon lattices [82, 83]. The large intensity ratio of IG/ID (≈1.02) implies that the AC nanosheets in MoO3/AC were highly graphitized with small numbers of stacked layers [8492]. Thus, one can conjecture the present MoO3/AC nanocomposites to comprise highly conductive AC that may improve the electrical conductivity of the entire MoO3/AC nanocomposite system.

Figure 3 displays the morphological properties of the synthesized materials. In the case of MoO3, the sample exhibited the nanoparticle-interconnected cluster-like morphology (Figures 3(a) and 3(b)). After sonicating the MoO3 nanoparticles with the biomass HH-derived AC nanosheets (see also Figure S1 for the AC morphology), the sample exhibited the MoO3/AC nanocomposite structure where the MoO3 nanoparticles were aggregated with the cleaved AC nanosheets (Figures 3(c) and 3(d)). Here, it should be noticed that the MoO3 particle size was much decreased after ultrasonication (see also Figure 4). Namely, the average particle size was reduced from 600 nm (bare MoO3) to 200 nm (MoO3/AC composite). This can be explained by the following ultrasonication mechanism. During ultrasonication in the aqueous solution, H2O renders the two important radicals that can promote the sonochemical reaction; i.e., one is hydrogen (H) and the other is hydroxyl (OH). These two radicals play key roles as the reductants during the sonication process [9397]. Namely, when applying high ultrasonic power in H2O, the bigger MoO3 particles (i.e., interconnected nMoO3) would be cleaved and collapsed so that the smaller MoO3 particles (i.e., exfoliated MoO3(n)) could be formed via the following reactions:

From the in situ EDX measurements (Figures 3(e) and 3(f)), it was confirmed that both MoO3 and MoO3/AC were composed of their own elements of Mo, O, and C. Here, it should be noted that the Pt peaks in the EDX spectra were arisen from the ultrathin Pt layer, which was coated onto the sample surface before the FE-SEM measurements to minimize the electron charging effect.

To monitor the microstructural insights into both MoO3 and MoO3/AC, TEM measurements were carried out. The MoO3 sample exhibited a typical topography of the colloidal nanoparticles where the spherical MoO3 nanoparticles were interconnected with each other (Figures 4(a) and 4(b)). As depicted in Figure 4(c), one can confirm the interplanar spacing of MoO3 to be 0.38 nm (i.e., direction of ), which coincides with that of orthorhombic (110) α-MoO3 [98, 99]. Furthermore, the SAED pattern clearly showed the circularly dispersed spots, presenting a polycrystalline nature of orthorhombic α-MoO3 [78, 100] (Figure 4(d)). Different from bare MoO3, the MoO3/AC nanocomposites clearly exhibited an intertwingled texture of the AC nanosheet-encapsulated spherical MoO3 nanoparticles (Figures 4(e) and 4(f)). Moreover, it can be seen from the high-resolution TEM image that the crystallized spherical MoO3 nanoparticles were aggregated with the amorphous AC nanosheets (Figure 4(g)). The coexistence of both ring patterns (i.e., amorphous AC [56, 101]) and circularly dispersed spots (i.e., polycrystalline MoO3 [78, 100]) further verifies the effective formation of the MoO3/AC composites (Figure 4(h)).

To investigate the effects of the AC incorporation on the electrochemical characteristics of MoO3/AC, we examined and compared the CV characteristics of MoO3 and MoO3/AC. Figures 5(a) and 5(b) present the CV curves of MoO3 and MoO3/AC, respectively. Both samples exhibited the typical oxidation and reduction peaks, representing the insertion and the desertion of electrolyte ions through the cathodic and the anodic reactions, respectively. The feature of the observed redox reaction belongs to the pseudocapacitive behavior, which would strongly affect the OER and HER activities. As the scan rate increased, the current density also increased because of the low diffusion resistance of the active catalyst material. Compared to bare MoO3, MoO3/AC displayed a larger CV window and a higher current density. This depicts that MoO3/AC possesses a larger number of active sites than that of MoO3. We believe such an improved electrochemical activity of MoO3/AC to arise from two possible reasons. One is the increased electrical conductivity [17, 53], and the other is the increased active surface area [24, 102, 103]. The former will be discussed in detail later at the EIS part, and the latter is to be verified as follows.

To clarify the improved electrocatalytic performance of MoO3/AC, firstly, we calculated the electrochemically active surface area (ECSA) by using the following equations [17, 24, 104, 105]: where is the double-layer charging current, is the non-Faradic capacitance, is the scan rate, is the electrode area, and is the electrolyte capacitance (0.04 mF/cm2 for KOH). From the non-Faradic CV region, the magnitudes of ECSA were determined to be 124 and 153 cm2 for MoO3 and MoO3/AC, respectively (Figures 5(c) and 5(d)). This depicts that MoO3/AC has a larger number of the electrochemically active sites than that of MoO3. One can ascribe such a result to the sonochemical exfoliation of the MoO3 crystallites, as confirmed from FE-SEM, XRD, and TEM. In other words, since the smaller crystallites provide the larger surface-to-volume ratio, it can be inferred that the electrochemically active surface area was increased in MoO3/AC via the sonochemical exfoliation of the MoO3 crystallites.

Next, the electrochemical resistive behaviors were examined by the EIS measurements. As shown in the Nyquist plots (Figures 5(e) and 5(f)), both MoO3 and MoO3/AC exhibited the straight lines and the parabolic curves at the low and the high frequency regions, respectively. The former is relevant to the series resistance () of the working electrodes, relating to the charge transfer resistance of the catalyst materials [18, 106]. The latter is attributed to the dispersion of electrolytes inside the working electrodes [24, 51]. Using the equivalent circuit model (insets of Figures 5(e) and 5(f)), the magnitudes of were determined to be 4.02 and 1.06 Ω for MoO3 and MoO3/AC, respectively. Thus, one can surmise that MoO3/AC has a smaller charge transfer resistance than that of bare MoO3.

The increases in both the electrical conductivity and the active surface area help promote the fast electronic charge transport and the swift ionic diffusion, and it may in turn lead to the enhanced OER/HER activities. To assess the electrocatalytic OER/HER performances, therefore, we measured and compared the LSV characteristics for MoO3 and MoO3/AC. Figure 6(a) shows the -corrected LSV curves at 1 mV/s of MoO3 and MoO3/AC. From the measured LSV data at 10 mA/cm2, the overpotential () values of MoO3 and MoO3/AC were determined to be 350 and 280 mV, respectively, by using the following equations: where and are the standard potentials of the reversible hydrogen electrode and the reference electrode, respectively. The values obtained from the present MoO3 and MoO3/AC catalysts are comparable to and even lower than the literature values (Table S1). Particularly, MoO3/AC recorded the lowest among the well-known metal oxide-based electrocatalysts. For better understanding, we note that the turnover frequencies of MoO3 and MoO3/AC were estimated to be 0.019 and 0.712 s-1 at the potential of 1.51 V, respectively (Figure S4(a)). This implies the MoO3/AC nanocomposites to hold the improved intrinsic reaction kinetics, which could enhance the OER performance in the alkaline electrolyte [17, 24, 38, 43]. The improved OER reaction kinetics can be further authenticated through evaluating the Tafel slope (), which can be calculated by using the following equation: where and are the current density applied to the electrodes and the fitting parameter, respectively. From the measured Tafel curves (Figure 6(b)), the small values were determined for both MoO3 (47 mV/dec) and MoO3/AC (35 mV/dec). Particularly, MoO3/AC exhibited to have a smaller value, compared to other metal oxide-based electrocatalysts (Table S1). Namely, as discussed above, MoO3/AC showed the improved intrinsic reaction kinetics because of its larger ECSA and smaller .

The intrinsic reaction kinetics may also affect the chronopotentiometric responses. As shown in voltage-step profiles (Figure 6(c)), MoO3/AC showed the smaller overpotential values at every current density step than those of bare MoO3. This validates that sonochemical hybridization of MoO3/AC could help enhance the ion storage performance as well as the catalytic activity. In addition, MoO3/AC showed the long-term stability, compared to MoO3 (Figure 6(d)). Initially, both MoO3 and MoO3/AC revealed an enhanced potential activity because of the preliminary activation of the catalyst. After a few hours, however, the potential dramatically reduced and maintained constantly because of the catalyst stabilization. Furthermore, the sample showed the nearly identical LSV characteristics even after 10 h of the stability test (Figure S2). These suggest that MoO3/AC could act as a superior and stable electrocatalytic OER catalyst.

Finally, we examined the HER performances to assess the bifunctional water electrolysis activities. Figure 7(a) displays the -corrected LSV characteristic curves, which were measured in the KOH electrolyte at 1 mV/s. From LSV data at -10 mA/cm2, the values were calculated to be 387 and 353 mV for MoO3 and MoO3/AC, respectively. In addition, the low values of 146 and 124 mV/dec were obtained from MoO3 and MoO3/AC, respectively, as can be seen from the Tafel curves (Figure 7(b)). Compared to bare MoO3, the values of and of MoO3/AC are comparable to and even smaller than those of other metal oxide-based electrocatalysts (Table S2). Furthermore, the turnover frequencies of the prepared MoO3 and MoO3/AC catalysts were calculated to be 0.00111 and 0.00184 s-1 at the potential of 0.35 V, respectively (Figure S4(b)). For the chronopotentiometric HER performance test (Figure 7(c)) and the stability test (Figure 7(d) and Figure S3), MoO3/AC showed the better electrocatalytic activities than those of MoO3. From all the above result, therefore, it can be concluded that sonochemical hybridization of orthorhombic α-MoO3 and highly graphitized AC is of good use for improving the bifunctional OER/HER activities. For further improvements of both bifunctionality and stability, using the highly conductive and highly porous substrates (e.g., Ni foam, C clothes, and C sheets) can also be the next step toward the practical OER/HER application of the MoO3/AC catalyst.

4. Conclusions

The high-performance bifunctional OER/HER electrocatalysts of MoO3/AC were fabricated through the facile ultrasonication by using the biomass-derived AC nanosheets and the sol-gel grown α-MoO3 nanoparticles. MoO3/AC exhibited the excellent electrocatalytic water splitting activities in both OER and HER. Namely, MoO3/AC not only showed the superb OER performances (i.e., low of 280 mV and low of 35 mV/dec) but also displayed the remarkable HER performances (i.e., low of 353 mV and low of 124 mV/dec). Such excellent bifunctional OER/HER activities were confirmed to arise from the large ECSA (153 cm2) and the (1.06 Ω) of MoO3/AC, attributing to sonochemical hybridization of the highly conductive nanoconstituent (i.e., graphitized AC nanosheets) and the electrochemically active nanoconstituent (i.e., α-MoO3 nanoparticles).

Data Availability

Data will be made available upon request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Authors’ Contributions

Sankar Sekar carried out the investigation, formal analysis, methodology, and writing of the original draft. Ji-Seop Yun assisted in the investigation and formal analysis. Seoyeon Park carried out the investigation and formal analysis. Deuk Young Kim was responsible for the supervision and validation. Youngmin Lee made great contribution to data curation, methodology, and validation. Sejoon Lee assisted in the conceptualization, supervision, and funding acquisition and wrote, reviewed, and edited the manuscript.

Acknowledgments

This research was supported by the National Research Foundation (NRF) of Korea through the basic science research programs (2016R1A6A1A03012877 and 2023R1A2C1005421) funded by the Korean Government.

Supplementary Materials

Supplementary data includes the morphological and the structural properties of the AC nanosheets, the electrocatalytic properties of MoO3 and MoO3/AC, and the comparison of the electrocatalytic OER and HER activities for various materials. Figure S1: FE-SEM image of the AC nanosheets. Figure S2: LSV curves of MoO3 and MoO3/AC for OER before and after the stability test. Figure S3: LSV curves of MoO3 and MoO3/AC for HER before and after the stability test. Figure S4: turnover frequency plots of MoO3 and MoO3/AC for OER and HER. Table S1: comparison of the OER performances of MoO3 and MoO3/AC with previously reported metal oxide-based electrocatalysts. Table S2: comparison of the HER performances of MoO3 and MoO3/AC with previously reported metal oxide-based electrocatalysts. (Supplementary Materials)