Next Article in Journal
Transport Property of Wrinkled Graphene Nanoribbon Tuned by Spin-Polarized Gate Made of Vanadium-Benzene Nanowire
Previous Article in Journal
Intensity-Dependent Optical Response of 2D LTMDs Suspensions: From Thermal to Electronic Nonlinearities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Controlled Fabrication of Hierarchically Structured MnO2@NiCo-LDH Nanoarrays for Efficient Electrocatalytic Urea Oxidization

1
Shenzhen Key Laboratory of Special Functional Materials, Guangdong Research Centre for Interfacial Engineering of Functional Materials, College of Materials Science and Engineering, Shenzhen University, Shenzhen 518060, China
2
Fujian Key Laboratory of Functional Marine Sensing Materials, College of Materials and Chemical Engineering, Minjiang University, Fuzhou 350108, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(15), 2268; https://doi.org/10.3390/nano13152268
Submission received: 1 July 2023 / Revised: 1 August 2023 / Accepted: 2 August 2023 / Published: 7 August 2023
(This article belongs to the Section Energy and Catalysis)

Abstract

:
Urea, a prevalent component found in wastewater, shows great promise as a substrate for energy-efficient hydrogen production by electrolysis. However, the slow kinetics of the anodic urea oxidation reaction (UOR) significantly hamper the overall reaction rate. This study presents the design and controlled fabrication of hierarchically structured nanomaterials as potential catalysts for UOR. The prepared MnO2@NiCo-LDH hybrid catalyst demonstrates remarkable improvements in reaction kinetics, benefiting from synergistic enhancements in charge transfer and efficient mass transport facilitated by its unique hierarchical architecture. Notably, the catalyst exhibits an exceptionally low onset potential of 1.228 V and requires only 1.326 V to achieve an impressive current density of 100 mA cm−2, representing a state-of-the-art performance in UORs. These findings highlight the tremendous potential of this innovative material designing strategy to drive advancements in electrocatalytic processes.

1. Introduction

As a promising and environmentally-friendly energy carrier, hydrogen (H2) has garnered significant attention as a potential substitute for fossil fuels in future energy systems [1,2]. The electrochemical water splitting reaction (2H2O → 2H2 + O2) provides an eco-friendly and sustainable pathway for hydrogen production, enabling the efficient conversion of surplus or low-quality electricity into chemical energy stored in the form of hydrogen gas [3,4,5,6]. Regrettably, the efficiency of water electrolysis is significantly impeded due to the high theoretical reaction potential of 1.23 V and the sluggish kinetics of the oxygen evolution reaction (OER) occurring at the anode side [7]. In order to address this limitation, several novel reaction schemes have been proposed as alternatives to the OER. The urea oxidation reaction (UOR, CO(NH2)2 + 6OH → N2 + 5H2O + CO2 + 6e), with a favorable thermodynamic potential of only 0.37 V, has therefore attracted considerable research interest [8]. Urea, which occurs in abundance in human urine and/or industrial wastewater, serves as an resource for this reaction. Despite its great promise, the UOR still encounters significant kinetic barriers that hinder its efficiency [9,10]. Therefore, the development of active catalysts is imperative to overcome these challenges. Numerous transition metal compounds have been extensively exploited as potential catalysts for the UOR [10,11,12,13,14,15,16,17,18]. Among them, Ni–Co binary hydroxides stand out as particularly attractive options due to their abundant availability in nature and remarkable catalytic performance for UORs [17,19,20]. Despite great promise, there remains a considerable requirement for further enhancing performance to meet the demands of practical applications, calling for the implementation of innovative material design strategies. Catalytic reactions take place at the catalyst’s surface, which underscores the importance of reducing its size to maximize the availability of active sites. However, size reduction often results in increased boundaries that impede electron conduction. To strike a balance, the implementation of hierarchical 3D nanostructures arises as the superior choice, as they can offer exceptionally large surface areas while preserving well-connected networks for effective charge carrier conduction.
In this article, we present the design and fabrication of hierarchically structured MnO2@NiCo-LDH nanosheet arrays supported on nickel foam (NF) and their applications as an efficient UOR electrocatalyst. Ultrathin MnO2 nanosheet arrays were first formed on NF via hydrothermal growth, which served as a conductive substrate for the electrodeposition of a (Ni, Co)-based layered double hydroxide (NiCo-LDH) nanosheet. The NiCo-LDH loading density was readily controlled by tuning the deposition current density and time. The formation of the hierarchical nanostructure was confirmed by microscopic investigations, and remarkable enhancement in electrocatalytic performance for UOR was validated by leveraging the advantages of its unique architecture.

2. Materials and Methods

2.1. Preparation of MnO2@NiCo-LDH Nanosheet Arrays

Commercial nickel foam (NF, 1 mm in thickness) was cut into pieces measuring 1 cm × 3 cm for subsequent use. These nickel foam pieces underwent an initial treatment in a 1 M hydrochloric acid solution for 15 min to remove the naturally formed oxide layer at the surface.
The cleaned NF pieces were utilized as substrates for the growth of ultrathin MnO2 nanosheet arrays via hydrothermal synthesis by adopting our previously reported method [21]. Typically, a piece of cleaned NF was placed in a 50 mL polytetrafluoroethylene (PTFE)-lined stainless steel reaction vessel that was loaded with 30 mL of 50 mM aqueous solution of KMnO4, and the reaction was then held at 180 °C for 2.5 h in an electric oven. Electrodeposition was employed to form NiCo-LDH nanosheets over the primary MnO2 nanosheet arrays from aqueous solutions of the metal nitrate salts, with the MnO2-coated nickel foam as the working electrode, a piece of platinum (Pt) foil as the counter electrode, and a saturated calomel electrode (SCE) as the reference electrode. The electrolyte was an aqueous solution that contained 0.1 M Ni(NO3)2·6H2O and 0.1 M Co(NO3)2·6H2O. Typically, the deposition was carried out at a constant current density of −20 mA cm−2, and the NiCo-LDH loading amount was controlled by tuning the deposition time (50 s, 100 s, 150 s, 200 s).

2.2. Characterizations

Scanning electron microscopy (SEM) was acquired in a secondary electron collection mode with an accelerating voltage of 5 KeV by using an APREO-S field emission scanning electron microscope (Thermo Fisher Scientific Inc., Waltham, MA, USA). Transmission electron microscopy (TEM) investigations were conducted on an accelerating voltage of 200 KeV on the JEM-F200 scanning transmission electron microscope (JEOL Ltd., Akishima, Tokyo, Japan). A ESCALAB X-ray photoelectron spectrometer (Thermo Fisher Scientific Inc., Waltham, MA, USA) was used for collecting X-ray photoelectron spectroscopy (XPS) data and all spectra were calibrated by aligning the C 1s peak at 284.8 eV. X-ray diffraction (XRD) patterns were obtained using a RIGAKU Miniflex X-ray diffractometer (Applied Rigaku Technologies, Inc., Akishima, Tokyo, Japan), with Cu-Kα radiation, at a scan rate of 5° per minute. Inductively coupled plasma-atomic emission spectroscopy (ICP-AES) analysis was performed on an iCAP 7000 Series spectrometer (Thermo Fisher Scientific Inc., Waltham, MA, USA) to assess the possible dissolution of the catalyst component after the prolonged electrochemical stability test. For the analysis, 2 mL of the electrolyte was taken and then acidified using 4 mL of 1 M HCl. The solution was further filtered through a 0.22 μm nylon filter before being sent for the ICP-AES test.

2.3. Electrochemical Measurements

All electrochemical measurements were performed by using a Interface 1010E potentiostat (Gamry Instruments, Warminster, PA, USA). The OER tests were carried out in a standard 1 M KOH (pH = 13.6) solution. For the UOR tests, 0.5 M urea was added to the alkaline solution. Linear sweep voltammetry (LSV) curves were recorded with a scan rate of 5 mV/s, and iR-correction was applied using the current interrupt (CI) method. Electrochemical impedance spectroscopy (EIS) measurements were conducted in the potential static mode, using frequencies ranging from 1 to 105 Hz and an oscillation amplitude of 10 mV at an overpotential of 450 mV vs. Hg/HgO (balanced with 1 M KOH). The stability of the electrocatalyst was tested using chronopotentiometry at a constant current density of 100 mA cm−2 for a duration of 20 hours. To account for the pH effects, the potentials were converted to the reversible hydrogen electrode (RHE) scale using the Nernst equation (E vs. RHE = E vs. ref + 0.059 × pH + 0.114 V). The potential of the Hg/HgO electrode was also calibrated against a Pt|H2 (g) electrode in the 1 M KOH electrolyte using a multimeter [21], which showed a potential of 0.916 V (Figure S1), consistent with the calculation mentioned above.

3. Results and Discussion

Ultrathin MnO2 nanosheet arrays supported on NF were utilized as a substrate for the deposition of NiCo-LDH. These MnO2 nanosheet arrays were selected because of their half-metallic electronic properties, easy synthesis, and large accessible surface areas [22,23]. SEM images (Figure S2) confirm the formation of vertically aligned MnO2 nanosheets on the surface of NF, which was consistent with our previous report [21]. The bare NF and NF-supported MnO2 nanosheet arrays were employed as substrates for the electrodeposition of the NiCo-LDH active layer from solutions containing nickel and cobalt nitrate salts. As depicted in Equations (1) and (2), the deposition was triggered through an increase in local pH at the electrode’s surface as a result of the electrochemical reduction in nitrate ions [24].
NO 3 + 7   H 2 O + 8   e NH 4 + + 10   OH
x   Ni 2 + + y   Co 2 + + 2 x + y   OH Ni x Co y ( OH ) 2
The formation of NiCo-LDH over the MnO2 was investigated using SEM. As shown in Figure 1, when the deposition was carried out at the current density of −20 mA cm−2 and kept for a short time of 50 s, a uniform coating of LDH nanosheets was observed on the surfaces of individual MnO2 nanosheets. Importantly, there were still ample free spaces present among the MnO2@NiCo-LDH hybrid nanosheets. However, as the deposition time increased, a continuous overcoating of LDH occurred. For example, after deposition for 150 s, nearly all the available spaces between the primary MnO2 nanosheets were occupied with LDH, and concurrently, cracks started to form within the overall deposition layer. These cracks became more pronounced as the deposition time was further extended to 200 s. As a control, NiCo-LDH was also deposited on bare NF, and the SEM image reveals that the LDH nanosheets were organized into close-packed microspheres (Figure S3).
With a hierarchical structure, the MnO2@NiCo-LDH fabricated with a deposition time of 50 s was further investigated in TEM. The presence of vertically aligned NiCo-LDH over the MnO2 sheets was clearly demonstrated by the emergence of conspicuous black strips observed atop the bulk nanosheets in the TEM image (Figure 2a). Upon conducting further analysis using high-resolution transmission electron microscopy (HRTEM), distinct lattice fringes were observed, with a measured spacing of 0.46 nm (Figure 2b). These fringes are indicative of the interlayer spacing of the (001) atomic planes of the hydroxide, as documented in the hydroxide’s crystallographic database entry (PDF#73-1520). Notably, the selected area electron diffraction (SAED) pattern (Figure 2c) exhibited exclusive diffraction rings that matched the δ-MnO2 phase (PDF#80-1098), with no indication of NiCo-LDH signals observed, suggesting the low crystallinity of the hydroxide deposits. Moreover, the high-angle annular dark field (HAADF) scanning transmission electron (STEM) image, combined with the energy-dispersive X-ray spectroscopy (EDS) mapping images (Figure 2d–g), provided compelling evidence of the homogeneous distribution of the Mn, Ni, and Co components throughout the sample. The general atomic contents of these elements were determined to be 16.41% (Mn), 7.13% (Ni), and 4.81% (Co), according to the EDS spectrum (Figure S4).
The performance of MnO2@NiCo-LDH for urea oxidation reactions (UOR) was assessed in the alkaline solution added with urea (0.5 M). NiCo-LDH/NF, MnO2/NF, and bare NF were used as reference catalysts. The LSV curves and potential statistics required to achieve specified current densities are presented in Figure 3a,b, respectively. It is evident that MnO2@NiCo-LDH possesses a significantly improved performance for UORs compared to the reference catalysts. Notably, by enlarging the low current density range of the LSV curves (Figure S5), the onset potential of MnO2@NiCo-LDH was measured as 1.228 V vs. RHE. This value is below both the thermodynamic redox potential of the OER (1.230 V) and the onset potentials of NiCo-LDH/NF (1.267 V), MnO2/NF (1.311 V), and NF (1.377 V). Significantly, when subjected to an applied potential of 1.5 V vs. RHE, the current density of MnO2@NiCo-LDH impressively reaches 905 mA cm−2. This value surpasses the current densities of NiCo-LDH/NF (398 mA cm−2) and MnO2/NF (337 mA cm−2) by more than two-fold. Likewise, to attain current densities of 10 and 100 mA cm−2, the MnO2@NiCo-LDH necessitates potentials of 1.261 and 1.326 V vs. RHE, respectively. These values represent a considerable reduction compared to the potentials required by the reference catalysts. Furthermore, they are on par with the most superior UOR catalysts reported in the literature [20] (Table S1), underscoring the exceptional electrochemical capability of MnO2@NiCo-LDH. In addition, Tafel plots (Figure 3c) were generated based on the LSV curves for analyzing the kinetics of the UOR on the different catalysts. MnO2/NF displayed a Tafel slope of 45.8 mV dec−1, indicating a rapid UOR rate. However, its overall performance was limited due to a high onset potential. On the other hand, NiCo-LDH/NF exhibited a Tafel slope of only 42.2 mV dec−1 in the low potential region, but this slope significantly increased to 111.5 mV dec−1 at higher applied potentials. This discontinuity is likely attributed to an increase in mass transport resistance caused by an insufficient exposed surface area at high potentials. In comparison to the aforementioned catalysts, MnO2@NiCo-LDH demonstrated a larger Tafel slope of 62.3 mV dec−1, but this value remained consistent across the displayed potential range. These findings corroborate our hypothesis that the combination of high conductivity in MnO2 nanosheets and the low onset potential of NiCo-LDH work synergistically to enhance the activity of the hybrid catalyst. The electrochemical surface areas (ECSA) of the catalysts were estimated by measuring their double-layer capacitance (Cdl) at the non-faradaic region through CV scanning at different rates. The definitive Cdl values were calculated as half the slopes of the plots that depict the difference in current densities (Δj) between the anodic and cathodic scans against the scanning rate (Figure 3d and Figure S6). The Cdl of MnO2@NiCo-LDH reaches 4.03 mF cm−2, which is almost twice that of MnO2/NF (2.16 mF cm−2) and 2.8 times that of NiCo-LDH (1.44 mF cm−2). This significant increase in Cdl of MnO2@NiCo-LDH reflects a much greater accessible surface area for the electrocatalytic reaction. EIS was utilized to further explore the charge transport kinetics of the UOR. As shown in Figure 3e, all the samples exhibit typical semicircular curves in their EIS, where the diameters of the semicircles reflect the charge transfer resistances (Rct). MnO2@NiCo-LDH has the smallest semicircle diameter, indicating a lowered Rct at the catalyst/electrolyte interface. In earlier investigations, Botte et al. [25,26] suggested an indirect “electrochemical-chemical” (E-C) mechanism for the UOR process on nickel-based catalysts. Based on this theory, electrochemically generated Ni3+ ions act as the active species responsible for the oxidation of adsorbed urea molecules. Therefore, the potential for the regeneration of Ni2+ to Ni3+ plays a critical role in the catalytic process. Figure 3f illustrates the cyclic voltammetry (CV) curves of MnO2@NiCo-LDH and bare NF, revealing that the redox potential of Ni2+/Ni3+ in MnO2@NiCo-LDH is shifted to a lower value compared to the pure nickel-based sample. This shift indicates that urea oxidation on MnO2@NiCo-LDH can occur at a lower onset potential, consistent with the results obtained from the LSV curves of UOR.
XPS was utilized to investigate any potential alteration in the chemical states of the Ni and Co components during the electrolysis. Figure 4 shows the full XPS survey spectra and the high-resolution spectra of the Ni 2p, Co 2p, and O 1s of the sample before and after the UOR test, all of which were calibrated by aligning the C-C binding energy to the standard value of 284.8 eV in the high-resolution spectra of C 1s (Figure S7). The full survey spectra (Figure 4a) showed no significant change observed in the composition. In the high resolution spectra of Ni 2p (Figure 4b), the initial sample exhibited typical features of 2p3/2 and 2p1/2, and two satellites at binding energies of 855.4, 873.3 eV, 861.5, and 879.3 eV, signifying that the Ni had a valence state of +2 [27]. Interestingly, after the UOR test, the Ni 2p spectra exhibited no noticeable position shift, indicating little change of the Ni valence. In the spectra of Co 2p (Figure 4c) of the original sample, the photoemission peak of the 2p3/2 components was located at 780.8 eV with noticeable satellite peaks emerging at around 785.5 eV, which has been typically observed in Co(OH)2 [28]; however, after the UOR test, the 2p3/2 peak shifted to a lower binding energy of 780.3 eV, and meanwhile, the satellite peaks were depleted, indicating the formation of Co3+ species by the UOR test [29,30]. The O 1s spectrum (Figure 4d) of MnO2@NiCo-LDH demonstrated a dominate peak at 510.0 eV, which was ascribed to the oxygen in the hydroxide group (OH) [31]. The peak shifted slightly to the binding energy centered at 530.6 eV after the UOR test, which was likely associated with the absorbed carbonate anions at the surface due to the generation of CO2 during UOR. During UOR tests at the high potential region, Ni2+ and Co2+ ions undergo partial oxidation, forming Ni3+ and Co3+ species, respectively. However, the Ni 2p spectrum of the sample after UOR revealed the absence of Ni3+, indicating the rapid backward reduction of these high-valency species to Ni2+. The results obtained provide compelling evidence in favor of the indirect electrochemical (E-C) mechanism of UOR, where the electrochemically generated Ni3+ species undergo a simultaneous reaction with urea molecules, leading to their reduction back to Ni2+. The regenerated Ni2+ species can actively partake in subsequent cycles of the electrocatalysis process.
The catalysts obtained with extended deposition times (100 s, 150 s, 200 s) were also evaluated for their performances in the urea oxidation reactions (UOR). However, their performances were found to be subpar compared to the sample deposited for 50 s, particularly in the high-potential regions (Figure S8). This outcome is to be expected due to the presence of additional NiCo-LDH over-coatings, which exhibit poor conductivity and consequently impede the efficient transfer of electrons from the supporting electrode to the catalytic sites. Furthermore, these over-coatings obstruct the available space between the nanosheets, thereby impeding the mass-transport processes at the surfaces of the electrodes.
The above analyses strongly support the idea that the strategic integration of MnO2 nanosheets and NiCo-LDH within a hierarchical architecture not only amplifies the exposure and accessibility of active sites from the NiCo-LDH components but also promotes effective electron and reactant transfer during the electrochemical process, and this remarkable synergy leads to a state-of-the-art performance for UOR.
To showcase the advantages of UOR as an alternative anodic reaction to OER, a comparison was made by measuring the LSV curve of MnO2@NiCo-LDH (deposited for 50 s) in a pure KOH solution (1 M). As illustrated in Figure S9, the LSV curves of UOR emerge at a much lower potential region, showing a negative shift (ΔE) of 274 mV, compared to that of OER (1.60 V vs. RHE) for achieving a current density of 100 mA cm−2, which corresponds to a 17.1% reduction in energy consumption by replacing the OER with UOR. It is noteworthy that the shift is much larger than that of NiCoGe oxyhydroxide (260 mV), the best reported catalyst of UOR [20]. Considering that the best reported OER catalysts typically require an overpotential of about 250 mV (or a potential of 1.480 V vs. RHE) to achieve the same current level [32], the energy saved still amounts to 10.4%. This finding suggests great promise in adopting UOR as a replacement anodic reaction in energy-saving hydrogen production.
A constant-current chronopotentiometry measurement was carried out to evaluate the stability of the catalyst. Figure 5 illustrates the potential recorded over a time course of 20 h at a current density of 100 mA cm−2. It is evident that the potential experienced only a minor change during the time course, demonstrating the excellent catalytic stability of MnO2@NiCo-LDH for UOR. Additionally, an ICP-AES was performed to analyze the possible presence of dissolved metal cations resulting from the effect of urea complexation. The concentrations of dissolved Co, Ni, and Mn cations were found to be 0.180 ppm, 0.045 ppm, and 0.001 ppm, respectively. These remarkably low values provide further evidence of the outstanding chemical stability of the electrocatalysts under the operational conditions.

4. Conclusions

In summary, this study focused on the development of a precisely engineered hierarchical structure of MnO2@NiCo-LDH nanoarrays for effective electrocatalytic urea oxidation. The synthesized hybrid catalyst showcased improved reaction kinetics, which can be attributed to the combined advantages of the high electrical conductivity of MnO2 nanosheets and the intrinsic activity of the NiCo-LDH deposits. Additionally, the three-dimensional hierarchical architecture provided ample space for efficient mass transport during the reaction. Remarkably, the resulting catalyst exhibited an onset potential of only 1.228 V for UOR and required only 1.326 V to achieve a current density of 100 mA cm−2. Overall, the controlled fabrication of hierarchically structured nanoarrays offers a potential pathway for efficient electrocatalytic urea oxidation and holds great promise for application in energy conversion and storage systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13152268/s1, Figure S1: Potential calibration of the SCE and Hg/HgO against a Pt|H2(g) electrode in 1 M KOH; Figure S2: SEM images of MnO2 nanosheet arrays grown on nickel foam (NF); Figure S3: SEM image of CoNi-LDH deposited on NF; Figure S4: EDS spectrum of the mapping area; Figure S5: Enlarged view of the UOR LSV curves of different catalysts; Figure S6: CV curves of (a) MnO2@NiCo-LDH, (b) NiCo-LDH/NF, (c) MnO2/NF, and (d) NF at different potential scan rates at the non-Faradaic region for ECSA estimations; Figure S7: High resolution calibrated C 1s XPS spectra of MnO2@NiCo-LDH before and after the UOR test; Figure S8: LSV curves of UOR by MnO2@NiCo-LDH with different deposition times; Figure S9: LSV curves of MnO2@NiCo-LDH deposited for 50 s for UOR and OER. Table S1: Performance comparison of UOR catalysts reported recently.

Author Contributions

Conceptualization, W.X., G.D. and M.F.; Data curation, W.L. and G.D.; Formal analysis, W.L. and M.F.; Investigation, W.L. and W.X.; Methodology, W.X., G.D. and M.F.; Resources, G.D. and M.F.; Supervision, M.F.; Validation, W.X.; Writing—original draft, M.F.; Writing—review and editing, W.L. and M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Shenzhen City, grant number JCYJ20190808141015383; the Natural Science Foundation of Fujian Province, grant number 2022I0043; and the Science and Technology Project of Fuzhou, Grant No. 021-S-237.

Data Availability Statement

Data are available on request from the authors.

Acknowledgments

We thank the Instrumental Analysis Center of Shenzhen University for their help in TEM imaging.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yue, M.; Lambert, H.; Pahon, E.; Roche, R.; Jemei, S.; Hissel, D. Hydrogen Energy Systems: A Critical Review of Technologies, Applications, Trends and Challenges. Renew. Sustain. Energy Rev. 2021, 146, 111180. [Google Scholar] [CrossRef]
  2. Yusaf, T.; Laimon, M.; Alrefae, W.; Kadirgama, K.; Dhahad, H.A.; Ramasamy, D.; Kamarulzaman, M.K.; Yousif, B. Hydrogen Energy Demand Growth Prediction and Assessment (2021–2050) Using a System Thinking and System Dynamics Approach. Appl. Sci. 2022, 12, 781. [Google Scholar] [CrossRef]
  3. Shih, A.J.; Monteiro, M.C.O.; Dattila, F.; Pavesi, D.; Philips, M.; da Silva, A.H.M.; Vos, R.E.; Ojha, K.; Park, S.; van der Heijden, O.; et al. Water Electrolysis. Nat. Rev. Methods Primer 2022, 2, 84. [Google Scholar] [CrossRef]
  4. Yu, Z.-Y.; Duan, Y.; Feng, X.-Y.; Yu, X.; Gao, M.-R.; Yu, S.-H. Clean and Affordable Hydrogen Fuel from Alkaline Water Splitting: Past, Recent Progress, and Future Prospects. Adv. Mater. 2021, 33, 2007100. [Google Scholar] [CrossRef]
  5. Miller, H.A.; Bouzek, K.; Hnat, J.; Loos, S.; Bernäcker, C.I.; Weißgärber, T.; Röntzsch, L.; Meier-Haack, J. Green Hydrogen from Anion Exchange Membrane Water Electrolysis: A Review of Recent Developments in Critical Materials and Operating Conditions. Sustain. Energy Fuels 2020, 4, 2114–2133. [Google Scholar] [CrossRef]
  6. Qi, J.; Zhang, W.; Cao, R. Solar-to-Hydrogen Energy Conversion Based on Water Splitting. Adv. Energy Mater. 2018, 8, 1701620. [Google Scholar] [CrossRef]
  7. Chatenet, M.; Pollet, B.G.; Dekel, D.R.; Dionigi, F.; Deseure, J.; Millet, P.; Braatz, R.D.; Bazant, M.Z.; Eikerling, M.; Staffell, I.; et al. Water Electrolysis: From Textbook Knowledge to the Latest Scientific Strategies and Industrial Developments. Chem. Soc. Rev. 2022, 51, 4583–4762. [Google Scholar] [CrossRef]
  8. Xiao, Z.; Qian, Y.; Tan, T.; Lu, H.; Liu, C.; Wang, B.; Zhang, Q.; Sarwar, M.T.; Gao, R.; Tang, A.; et al. Energy-Saving Hydrogen Production by Water Splitting Coupling Urea Decomposition and Oxidation Reactions. J. Mater. Chem. A 2022, 11, 259–267. [Google Scholar] [CrossRef]
  9. Geng, S.-K.; Zheng, Y.; Li, S.-Q.; Su, H.; Zhao, X.; Hu, J.; Shu, H.-B.; Jaroniec, M.; Chen, P.; Liu, Q.-H.; et al. Nickel Ferrocyanide as a High-Performance Urea Oxidation Electrocatalyst. Nat. Energy 2021, 6, 904–912. [Google Scholar] [CrossRef]
  10. Qin, H.; Ye, Y.; Li, J.; Jia, W.; Zheng, S.; Cao, X.; Lin, G.; Jiao, L. Synergistic Engineering of Doping and Vacancy in Ni(OH)2 to Boost Urea Electrooxidation. Adv. Funct. Mater. 2022, 33, 2209698. [Google Scholar] [CrossRef]
  11. Ji, Z.; Song, Y.; Zhao, S.; Li, Y.; Liu, J.; Hu, W. Pathway Manipulation via Ni, Co, and V Ternary Synergism to Realize High Efficiency for Urea Electrocatalytic Oxidation. ACS Catal. 2022, 12, 569–579. [Google Scholar] [CrossRef]
  12. Cai, M.; Zhu, Q.; Wang, X.; Shao, Z.; Yao, L.; Zeng, H.; Wu, X.; Chen, J.; Huang, K.; Feng, S. Formation and Stabilization of NiOOH by Introducing α-FeOOH in LDH: Composite Electrocatalyst for Oxygen Evolution and Urea Oxidation Reactions. Adv. Mater. 2023, 35, 2209338. [Google Scholar] [CrossRef]
  13. Sun, H.; Liu, J.; Kim, H.; Song, S.; Fei, L.; Hu, Z.; Lin, H.-J.; Chen, C.-T.; Ciucci, F.; Jung, W. Ni-Doped CuO Nanoarrays Activate Urea Adsorption and Stabilizes Reaction Intermediates to Achieve High-Performance Urea Oxidation Catalysts. Adv. Sci. 2022, 9, 2204800. [Google Scholar] [CrossRef]
  14. Fang, M.; Xu, W.-B.; Han, S.; Cao, P.; Xu, W.; Zhu, D.; Lu, Y.; Liu, W. Enhanced Urea Oxidization Electrocatalysis on Spinel Cobalt Oxide Nanowires via on-Site Electrochemical Defect Engineering. Mater. Chem. Front. 2021, 5, 3717–3724. [Google Scholar] [CrossRef]
  15. Fang, M.; Xu, W.-B.; Shen, Y.; Cao, P.; Han, S.; Xu, W.; Zhu, D.; Lu, Y.; Liu, W. Electrochemical Tuning of Nickel Molybdate Nanorod Arrays towards Promoted Electrocatalytic Urea Oxidization. Appl. Catal. Gen. 2021, 622, 118220. [Google Scholar] [CrossRef]
  16. Tong, Y.; Chen, P.; Zhang, M.; Zhou, T.; Zhang, L.; Chu, W.; Wu, C.; Xie, Y. Oxygen Vacancies Confined in Nickel Molybdenum Oxide Porous Nanosheets for Promoted Electrocatalytic Urea Oxidation. ACS Catal. 2018, 8, 1–7. [Google Scholar] [CrossRef]
  17. Wang, X.-H.; Hong, Q.-L.; Zhang, Z.-N.; Ge, Z.-X.; Zhai, Q.-G.; Jiang, Y.-C.; Chen, Y.; Li, S.-N. Two-Dimensional Nickel–Cobalt Bimetallic Hydroxides towards Urea Electrooxidation. Appl. Surf. Sci. 2022, 604, 154484. [Google Scholar] [CrossRef]
  18. Li, D.; Zhou, X.; Liu, L.; Ruan, Q.; Zhang, X.; Wang, B.; Xiong, F.; Huang, C.; Chu, P.K. Reduced Anodic Energy Depletion in Electrolysis by Urea and Water Co-Oxidization on NiFe-LDH: Activity Origin and Plasma Functionalized Strategy. Appl. Catal. B Environ. 2023, 324, 122240. [Google Scholar] [CrossRef]
  19. Song, W.; Xu, M.; Teng, X.; Niu, Y.; Gong, S.; Liu, X.; He, X.; Chen, Z. Construction of Self-Supporting, Hierarchically Structured Caterpillar-like NiCo2S4 Arrays as an Efficient Trifunctional Electrocatalyst for Water and Urea Electrolysis. Nanoscale 2021, 13, 1680–1688. [Google Scholar] [CrossRef]
  20. Wang, P.; Bai, X.; Jin, H.; Gao, X.; Davey, K.; Zheng, Y.; Jiao, Y.; Qiao, S.-Z. Directed Urea-to-Nitrite Electrooxidation via Tuning Intermediate Adsorption on Co, Ge Co-Doped Ni Sites. Adv. Funct. Mater. 2023, 33, 2300687. [Google Scholar] [CrossRef]
  21. Fang, M.; Han, D.; Xu, W.-B.; Shen, Y.; Lu, Y.; Cao, P.; Han, S.; Xu, W.; Zhu, D.; Liu, W.; et al. Surface-Guided Formation of Amorphous Mixed-Metal Oxyhydroxides on Ultrathin MnO2 Nanosheet Arrays for Efficient Electrocatalytic Oxygen Evolution. Adv. Energy Mater. 2020, 10, 2001059. [Google Scholar] [CrossRef]
  22. Wang, H.; Zhang, J.; Hang, X.; Zhang, X.; Xie, J.; Pan, B.; Xie, Y. Half-Metallicity in Single-Layered Manganese Dioxide Nanosheets by Defect Engineering. Angew. Chem. Int. Ed. 2015, 54, 1195–1199. [Google Scholar] [CrossRef]
  23. Zhao, Y.; Chang, C.; Teng, F.; Zhao, Y.; Chen, G.; Shi, R.; Waterhouse, G.I.N.; Huang, W.; Zhang, T. Defect-Engineered Ultrathin δ-MnO2 Nanosheet Arrays as Bifunctional Electrodes for Efficient Overall Water Splitting. Adv. Energy Mater. 2017, 7, 1700005. [Google Scholar] [CrossRef]
  24. Li, Z.; Shao, M.; An, H.; Wang, Z.; Xu, S.; Wei, M.; Evans, D.G.; Duan, X. Fast Electrosynthesis of Fe-Containing Layered Double Hydroxide Arrays toward Highly Efficient Electrocatalytic Oxidation Reactions. Chem. Sci. 2015, 6, 6624–6631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Vedharathinam, V.; Botte, G.G. Direct Evidence of the Mechanism for the Electro-Oxidation of Urea on Ni(OH)2 Catalyst in Alkaline Medium. Electrochim. Acta 2013, 108, 660–665. [Google Scholar] [CrossRef]
  26. Vedharathinam, V.; Botte, G.G. Understanding the Electro-Catalytic Oxidation Mechanism of Urea on Nickel Electrodes in Alkaline Medium. Electrochim. Acta 2012, 81, 292–300. [Google Scholar] [CrossRef]
  27. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving Surface Chemical States in XPS Analysis of First Row Transition Metals, Oxides and Hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
  28. Xu, Y.; Liu, Z.; Chen, D.; Song, Y.; Wang, R. Synthesis and Electrochemical Properties of Porous α-Co(OH)2 and Co3O4 Microspheres. Prog. Nat. Sci. Mater. Int. 2017, 27, 197–202. [Google Scholar] [CrossRef]
  29. Liardet, L.; Hu, X. Amorphous Cobalt Vanadium Oxide as a Highly Active Electrocatalyst for Oxygen Evolution. ACS Catal. 2018, 8, 644–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Ye, S.; Wang, J.; Hu, J.; Chen, Z.; Zheng, L.; Fu, Y.; Lei, Y.; Ren, X.; He, C.; Zhang, Q.; et al. Electrochemical Construction of Low-Crystalline CoOOH Nanosheets with Short-Range Ordered Grains to Improve Oxygen Evolution Activity. ACS Catal. 2021, 11, 6104–6112. [Google Scholar] [CrossRef]
  31. Zhang, H.; Tian, W.; Guo, X.; Zhou, L.; Sun, H.; Tadé, M.O.; Wang, S. Flower-like Cobalt Hydroxide/Oxide on Graphitic Carbon Nitride for Visible-Light-Driven Water Oxidation. ACS Appl. Mater. Interfaces 2016, 8, 35203–35212. [Google Scholar] [CrossRef] [PubMed]
  32. Chowdhury, P.R.; Medhi, H.; Bhattacharyya, K.G.; Hussain, C.M. Recent Progress in the Design and Functionalization Strategies of Transition Metal-Based Layered Double Hydroxides for Enhanced Oxygen Evolution Reaction: A Critical Review. Coord. Chem. Rev. 2023, 483, 215083. [Google Scholar] [CrossRef]
Figure 1. SEM images of MnO2@NiCo-LDH with electrodeposition times of (a) 50 s, (b) 100 s, (c) 150 s, and (d) 200 s.
Figure 1. SEM images of MnO2@NiCo-LDH with electrodeposition times of (a) 50 s, (b) 100 s, (c) 150 s, and (d) 200 s.
Nanomaterials 13 02268 g001
Figure 2. (a) TEM, (b) HRTEM, (c) SAED pattern, and (dg) HAADF STEM images and the corresponding EDS elemental mapping images of MnO2@NiCo-LDH.
Figure 2. (a) TEM, (b) HRTEM, (c) SAED pattern, and (dg) HAADF STEM images and the corresponding EDS elemental mapping images of MnO2@NiCo-LDH.
Nanomaterials 13 02268 g002
Figure 3. (a) LSV curves, (b) performance statistics, (c) Tafel curves, (d) double layer capacitance calculation, (e) EIS spectra of MnO2@NiCo-LDH and the references of NiCo-LDH/NF, MnO2/NF, and NF tested in a 1 M KOH solution added with 0.5 M urea. The inset shows the equivalent circuit of the EIS data. Note: the R1, Rct, and CPE represent the series resistance, charge transfer resistance, and constant phase element, respectively. (f) CV curves of MnO2@NiCo-LDH and NF tested in 1 M KOH.
Figure 3. (a) LSV curves, (b) performance statistics, (c) Tafel curves, (d) double layer capacitance calculation, (e) EIS spectra of MnO2@NiCo-LDH and the references of NiCo-LDH/NF, MnO2/NF, and NF tested in a 1 M KOH solution added with 0.5 M urea. The inset shows the equivalent circuit of the EIS data. Note: the R1, Rct, and CPE represent the series resistance, charge transfer resistance, and constant phase element, respectively. (f) CV curves of MnO2@NiCo-LDH and NF tested in 1 M KOH.
Nanomaterials 13 02268 g003
Figure 4. (a) Full XPS survey spectra, (bd) high-resolution XPS spectra of Ni 2p, Co 2p and O 1s of the MnO2@NiCo-LDH before and after UOR tests.
Figure 4. (a) Full XPS survey spectra, (bd) high-resolution XPS spectra of Ni 2p, Co 2p and O 1s of the MnO2@NiCo-LDH before and after UOR tests.
Nanomaterials 13 02268 g004
Figure 5. Chronopotential curve of MnO2@NiCo-LDH at 100 mA cm−2.
Figure 5. Chronopotential curve of MnO2@NiCo-LDH at 100 mA cm−2.
Nanomaterials 13 02268 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, W.; Xu, W.; Dong, G.; Fang, M. Controlled Fabrication of Hierarchically Structured MnO2@NiCo-LDH Nanoarrays for Efficient Electrocatalytic Urea Oxidization. Nanomaterials 2023, 13, 2268. https://doi.org/10.3390/nano13152268

AMA Style

Liu W, Xu W, Dong G, Fang M. Controlled Fabrication of Hierarchically Structured MnO2@NiCo-LDH Nanoarrays for Efficient Electrocatalytic Urea Oxidization. Nanomaterials. 2023; 13(15):2268. https://doi.org/10.3390/nano13152268

Chicago/Turabian Style

Liu, Wenjun, Wenbo Xu, Guofa Dong, and Ming Fang. 2023. "Controlled Fabrication of Hierarchically Structured MnO2@NiCo-LDH Nanoarrays for Efficient Electrocatalytic Urea Oxidization" Nanomaterials 13, no. 15: 2268. https://doi.org/10.3390/nano13152268

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop