Next Article in Journal
An Electrochemical Sensor Based on a Nitrogen-Doped Carbon Material and PEI Composites for Sensitive Detection of 4-Nitrophenol
Previous Article in Journal
Recent Developments and Advancements in Graphene-Based Technologies for Oil Spill Cleanup and Oil–Water Separation Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrothermal Synthesis of MnWO4@GO Composite as Non-Precious Electrocatalyst for Urea Oxidation

by
Patnamsetty Chidanandha Nagajyothi
1,
Kisoo Yoo
1,
Rajavaram Ramaraghavulu
2,* and
Jaesool Shim
1,*
1
School of Mechanical Engineering, Yeungnam University, Gyeongsan 38541, Korea
2
Department of Humanities and Sciences, Annamacharya Institute of Technology and Sciences, Rajampet, Kadapa 516126, India
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(1), 85; https://doi.org/10.3390/nano12010085
Submission received: 19 November 2021 / Revised: 22 December 2021 / Accepted: 24 December 2021 / Published: 29 December 2021

Abstract

:
In this study, manganese tungstate (MW) and MW/graphene oxide (GO) composites were prepared by a facile hydrothermal synthesis at pH values of 7 and 12. X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS), and Raman spectroscopy were used for the structural, compositional, and morphological characterization of the nanoparticles (NPs). The XRD analysis revealed that the formation of monoclinic MnWO4 did not have impurities. The SEM and TEM analyses showed that the synthesized NPs were rod-shaped and well-distributed on the GO. The as-synthesized samples can be used as electrocatalysts for the urea oxidation reaction (UOR). The MW@GO-12 electrocatalyst exhibited higher current density values compared to other electrocatalysts. This study provides a new platform for synthesizing inexpensive nanocomposites as promising electrocatalysts for energy storage and conversion applications.

1. Introduction

Urea (CO(NH2)2) is an abundant and easily available substance, which is found in human and animal urine and is widely used in the chemical industries and agriculture [1]. Urea-rich wastewater causes serious health issues like throat and lung irritation and environmental damages like eutrophication and acid rain [2,3]. However, urea is an efficient hydrogen (H2) carrier with non-flammable, non-toxic, and colorless properties [4,5,6]. Various traditional methods are used to treat urea-rich wastewater such as hydrolysis [7], adsorption [8], chemical oxidation [9], and biodegradation [10]; however, these methods are inefficient and expensive. Hence, it is important to develop an efficient, cost-effective, and convenient method to degrade urea-rich wastewater. The electro-oxidation of urea is essential in wastewater treatment and hydrogen production, solving both energy and environmental problems [11]. Noble metal-based catalysts are efficient catalysts for the urea oxidation reaction (UOR), but they are expensive, scarce, and limited to large-scale applications [12]. Hence, developing active, cost-efficient catalysts, which are easily available, is essential.
Manganese tungstate (MnWO4) is an inexpensive and eco-friendly material, and it has recently attracted attention owing to its high electronic conductivity, effective redox chemistry, and electrochemical, multiferroic, and ionic properties [13,14,15,16]. It is used in photocatalysis [17], electrocatalysis [18], gas sensors [19], and supercapacitors [20]. The use of transition metal oxides combined with carbon-based materials is a suitable way to improve the electrochemical properties [21,22,23]. Carbon-based materials play an important role in energy storage devices, transistors, sensors, etc., as they possess high physical, chemical, electrical, and thermal properties [24]. Among carbon materials, GO, the oxidized form of graphene, has a large surface area and displays strong hydrophilicity owing to the abundant oxygen-containing groups on its edges [25]. Furthermore, these functional groups (carboxylic, hydroxyl, and carbonyl) have been actively used to build novel composites. Mallick et al. synthesized cotton-fabric-derived mesoporous carbon-supported MnWO4 nanostructures for Zn-ion batteries [26]. Sardar et al. synthesized a MnWO4/amorphous CNT hybrid material for supercapacitor applications [27]. Xu et al. synthesized a GO/MnWO4 composite for magnetic resonance/photoacoustic dual-modal imaging and tumor photothermo-chemotherapy [28]. Jianhua et al. reported a layered MnWO4/rGO nanocomposite for supercapacitor applications [29].
In this study, we synthesized MW@GO at pH values of 7 and 12 by simple hydrothermal synthesis, and it had a nanorod-like morphology. It is extremely active in electrocatalytic urea oxidation under alkaline conditions. The XRD, TEM, Raman, and XPS results revealed that no impurity was observed between MnWO4 and GO. The hydrothermal synthesis is simple, rapid, and cost-effective, and the as-synthesized catalysts are inexpensive alternatives to precious metal catalysts for electrocatalytic UORs.

2. Materials and Methods

2.1. Synthesis of Electrocatalysts

All the chemicals used were of analytical grade and were obtained from Sigma-Aldrich. They were used directly without additional purification. Initially, GO (0.1 mg/mL) was sonicated for 1 h in a 50:50 v/v ratio of ethanol and DI water, and later Mn (NO3)2 (1.0 g) was added and stirred for 30 min to completely dissolve the salt. A second solution was prepared (50:50 v/v ratio of ethanol and DI water) and Na2WO4 (1.2 g) was also added and stirred for 30 min. The second solution (Na2WO4) was added dropwise to the first solution (GO-Mn (NO3)2) and stirred for 30 min. The ammonia solution was added until the pH reached 12, the reaction was shifted to a hydrothermal setup, and the temperature was maintained at 180 °C for 12 h. The mixture was allowed to cool naturally, after which the solid electrocatalyst was collected by centrifugation and dried at 70 °C. It was labeled as MW@GO-12 and left overnight. Other electrocatalysts were synthesized by the same method. The sample synthesized under neutral condition (pH 7) with GO was labeled MW@GO-7. The samples synthesized without GO at pH 7 and 12 were labeled MW-7 and MW-12, respectively. The characterization is presented in the Supplementary Information.

2.2. Electrode Preparation

An electrocatalyst slurry was prepared by mixing the electrocatalyst, acetylene carbon, and the binder (PVDF) at a ratio of 80:10:10 (wt. %) in 500 µL of NMP and sonicated for 30 min. The slurry was then drop-coated on the nickel foam and dried in a vacuum oven at 70 °C. Pt and Hg/HgO were used as counter and reference electrodes, respectively, and a whole electrochemical urea oxidation reaction was carried out in a 1.0 M KOH aqueous electrolyte with and without urea, at room temperature.

3. Results and Discussion

3.1. Characterization of Electrocatalysts

Pure MnWO4 and MnWO4@GO were obtained at pH values of 7 and 12. At pH 7, the XRD peaks of the synthesized MnWO4 increased compared with MnWO4 at pH 12. Figure 1 shows the diffraction peaks of the MW-7 and MW-12 at 15.27° (010), 18.35° (100), 23.53° (011), 24.14° (110), 29.71° (−111), 30.18° (111), 30.99° (020), 35.83° (021), 37.23° (200), 40.12° (210) 40.79° (102), 43.28° (−112), 43.95° (112), 44.90° (211), 47.31° (030), 48.99° (022), 49.26° (220), 51.08° (130), 52.15° (122), 60.54° (032), 61.35° (113), 62.30° (−311), 63.83° (−132), 64.31° (040), and 67.47° (041). These values agree with the monoclinic crystal structure; space group: P2/c; space group number: 13 (JCPDS card no: 010-0477). Compared to pure MW-7 and MW-12 samples, there were no significant changes in MW@GO-7 and MW@GO-12 samples, and no noticeable diffraction peak of the GO was observed in these samples, as the regular stack of graphene oxide was destroyed by the intercalation of NPs [25].
Figure 2 shows the Raman spectra of the GO, MW-7, MW-12, MW@GO-7, and MW@GO-12. Pure GO exhibited the D and G bands at 1355 cm−1 and 1587 cm−1, respectively [30,31,32]. Pure MW-7 and MW-12 showed weak peaks at 548, 701, and 778 cm−1, which corresponds to the tensile vibration of Mn-O and the symmetric and asymmetric tensile vibration modes of W-O-W bonds, respectively [33,34,35]. The strong peak located at 889 cm−1 is related to the strong symmetrical stretching of the WO2 group in MnWO4 [35]. Both GO- and MnWO4-related bands were present in MW@GO-7 and MW@GO-12 samples. However, a slight shift in the D and G bands was observed in the composites. The Raman spectra results revealed the absence of the impurity phase in any of the synthesized catalysts. The UV-Vis spectra results are presented in the Supplementary Information (Figure S1).
For an in-depth elemental analysis, the chemical nature was examined by the XPS as shown in Figure 3. Figure 3a shows the XPS survey spectrum of the hydrothermally synthesized MW-12 and MW@GO-12 composite, which showed the presence of Mn, W, O, and C. The synthesized MW-12 and MW@GO-12 were analyzed by a high-resolution XPS, and the Mn 2p spectrum was deconvoluted into four peaks (Figure 3b) with binding energies (BEs) of approximately 640.2, 645.3, 652.7, and 656.9 eV corresponding to Mn 2p3/2 and Mn 2p1/2, confirming the +2 oxidation state of Mn [26,36]. The W 4f spectra showed two peaks at 34.9 and 37.1 eV (Figure 3c), corresponding to W 4f7/2 and W 4f5/2, respectively, suggesting a +6 oxidation state of W [18]. The XPS spectra of as-synthesized MW-12 in the O 1s region showed three peaks at 529.4, 530.8, and 532. With 3 eV, however, the MW@GO-12 showed four peaks at 529.5, 531, 532.6, and 534.5 eV (Figure 3d). In both samples, the most dominant peak was observed at approximately 529 eV, associated with the Mn-O-W bond in MnWO4 and the remaining peaks related to O−2 present in the lattice of MnWO4, O=C-OH, and C-OH bonds, respectively [26,27,36,37]. The C1s spectra show four peaks at 284.2, 285.7, 287.6, and 288.9 eV (Figure 3e), corresponding to C=C/C-C, C-OH, C-O, and O=C-OH, respectively [26,38].
Figure 4 shows the SEM images of MW-7, MW-12, MW@GO-7, and MW@GO-12 synthesized under different pH conditions. At pH 7, the synthesized MnWO4 was shaped as a rod (Figure 4(a1,a2)). As the pH increased from 7 to 12, the MnWO4 nanorods aggregated to a flower shape (Figure 4(c1,c2)). The SEM images of MW@GO-7 and MW@GO 12 are shown in Figure 4 b1 and b2 and Figure 4 d1 and d2, respectively. They did not show obvious changes in morphology when compared to pure MW-7 and MW-12 samples. The HR-TEM images of MW-12 and MW@GO-12 clearly show that the synthesized NPs had a nanorod shape and were randomly distributed on the graphene oxide nanosheets (Figure 5(a1,b2)). The d-spacing was measured using the Gatan software, and the results were shown in Figure 5(a2). The d-spacing of ~ 0.48 nm, which corresponds to the(100) plane. The SAED pattern reveals bright diffraction rings corresponding to the lattice planes of (110), (010), (−132), and (030), indicating that the synthesized NPs were crystalline in nature, which is well in line with the XRD results (Figure 5(a3,b3). The EDX analysis and elemental mapping results confirmed the presence of Mn, W, O, and C in the synthesized samples (Figure S2).

3.2. Urea Oxidation Reaction

The electrochemical oxidation reaction of urea was investigated using a conventional three-electrode system with the CV and CA techniques. The CV curves of the synthesized electrocatalysts are shown in Figure 6; they were recorded at various scan rates (5–80 mV/s) in a fixed potential range (0.0–0.7 V) for Figure 6a–d, corresponding to MW-7, MW-12, MW@GO-7, and MW@GO-12, respectively. The CV curves showed clear redox peaks; there were no observable changes in peak shift with an increase in pH, although the peak sharpness increased (Figure 6a,b). Similarly, after the addition of GO, there were no observable changes in peak shift, although an enhancement in the current density was noticeable, which indicated that the addition of GO increased the current density of the electrocatalysts (Figure 6b,d). The surface coverage (Γ*) was calculated for the electrocatalysts [39], and the corresponding plot is shown in Figure S3a. The Γ*values of electrocatalysts MW-7, MW-12, MW@GO-7, and MW@GO-12 are 1.47 × 10−7, 1.94 × 10−7, 4.53 × 10−7, and 8.19 × 10−7, respectively. The Γ* value increased rapidly with the addition of GO, and moderately with an increase in pH. The linear fit analysis was carried out for redox peak current densities vs. the square root of the scan rates for all electrocatalysts (Figure S3b). The linear fit plots suggest that the diffusion-control processes dominated the electrochemical studies. The electrochemical active surface area (ECSA) is a crucial parameter to identify the electrochemical performance. The ECSA was measured by the previously reported method [39], and the corresponding CV curves are shown in Figure S4. The double-layer capacitance (Cdl) of electrocatalysts were 3, 5, 9, and 30 mF g−1 for MW-7, MW-12, MW@GO-7, and MW@GO-12, respectively (Figure 7a,b). From these values, MW@GO-12 electrocatalysts displayed superior ECSA. The addition of carbon will increase the catalytical active surface area. Sourav et al., reported that on introducing carbon into the electrocatalyst (MnWO4) the ECSA increased [26]; therefore, the ORR activity was also increased. Askari et al. developed ZnFe2O4-rGO nanohybrids, and, on adding GO, the ECSA was increased [40]. These reports suggest that the introduction of GO into the electrocatalysts will increase the ECSA.
The electro-oxidation of urea was performed using the electrocatalysts, and the results are shown in Figure 8a–d, labeled MW-7, MW-12, MW@GO-7, and MW@GO-12, respectively. UOR was carried out at various scan rates (5–80 mV/s) and a fixed potential range (0.0–0.7 V). The current densities of MW-12 slightly increased compared to MW-7 owing to the smaller size of the MnWO4 nanorods, whereas, on adding GO, the current density increased rapidly due to the higher electron transfer between GO and the electrocatalyst. The MW@GO-7 shows higher current densities compared to MW-7 and MW-12 due to the addition of GO. The MW@GO-12 electrocatalyst exhibited higher current density values compared to other electrocatalysts owing to its smaller size and the presence of GO [27,29]. Tang et al. synthesized layered MnWO4/RGOcomposite. They studied the supercapacitor application and noted a higher capacitance and stability compared to pristine MnWO4 [29]. To obtain more information on the kinetics reaction of UOR, the linear fit analysis was carried out for the anodic peak current densities of the electrocatalysts; its results are shown in Figure S5. The results suggest that the current densities increased linearly with an increase in the scan rate and that the UOR reaction process of all electrocatalysts was controlled by the diffusion-control process [41,42]. Electrochemical impedance spectroscopy (EIS) plays a vital role in understanding kinetic processes in electrochemical studies. Figure S6 shows the EIS plot along with the circuit and corresponding fit values of the electrocatalysts; it reveals that the MW@GO-12 electrocatalyst exhibited a moderate solution resistance, a lower charge-transfer resistance, and Warburg impedance compared to other electrocatalysts. These results also show that the MW@GO-12 electrocatalyst exhibited a higher catalytic performance.
The stability of the electrocatalysts in the UOR reaction plays a vital role, and, in this study, the CA technique was used to identify the stability of the electrocatalysts. Figure S7 shows the CA analysis of the electrocatalysts in the presence and absence of urea. The CA analysis was performed up to 9000 s after adding urea to the electrolyte, and the current density of the electrocatalysts increased compared to the pure electrolytes. The MW-7 electrocatalyst exhibited a lower current density but a higher stability compared to other electrocatalysts. The thermal stability of the synthesized MW-12 and MW/GO-12 catalysts was ascertained using TGA under the air atmosphere with a temperature ramp of 10 °C min−1 (Figure S8). The weight losses were observed above 700 °C, and the MW-12 and MW@GO-12 showed a 2.9% and 4.4% weight loss, respectively.

4. Conclusions

The facile hydrothermal-synthesized non-precious MnWO4/GO composites have been introduced as electrocatalysts for urea oxidation under alkaline conditions. The MW@GO-12 electrocatalyst exhibited higher current density values compared to the other electrocatalysts, i.e., MW-7, MW-12, and MW@GO-7. The enhanced electrocatalytic urea oxidation using MnWO4@GO composites revealed a high potential for future applications in wastewater remediation, fuel cells, and hydrogen production.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano12010085/s1. Figure S1: UV-Visible spectra of the electrocatalysts. Figure S2: EDS spectrum of MW@GO-12 electrocatalyst, and corresponding elemental mapping analysis. Figure S3: Current density vs. scan rate plots (a) and current density vs. the square root of scan rate plots of the electrocatalysts (b). Figure S4: CV curves of the electrocatalysts at different scan rates for electrochemical active surface area tests; MW-7 (a), MW-12(b), MW@GO-7 (c), and MW@GO-12 (d). Figure S5: Linear fit analysis of UOR for electrocatalysts, and the plots of peak current density vs. scan rate. Figure S6: EIS analysis of the electrocatalysts; inset shows the equivalent fitted circuit along with experimental data. Figure S7: Chronoamperometry analysis of the electrocatalysts in the presence and absence of the urea in 1.0 M KOH electrolyte. Figure S8. TGA analysis of pristine MW-12 and MW@GO-12 measured in air atmosphere.

Author Contributions

P.C.N.: Investigation, writing–original draft, K.Y.: Conceptualization, R.R.: Data curation and formal analysis, J.S.: Supervision, project administration, and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grants funded by the Korean government (2020R1A4A1019227 and 2020R1A2C1012439).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data can be available upon request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, F.; Ye, K.; Cheng, K.; Wang, G.; Cao, D. Preparation of nickel nanowire arrays electrode for urea electrooxidation in alkaline medium. J. Power Sources 2015, 278, 562–568. [Google Scholar] [CrossRef]
  2. Alotaibi, N.; Hammud, H.H.; Al Otaibi, N.; Prakasam, T. Electrocatalytic properties of 3D hierarchical graphitic carbon–cobalt nanoparticles for urea oxidation. ACS Omega 2020, 5, 26038–26048. [Google Scholar] [CrossRef]
  3. Wang, S.; Xu, P.; Tian, J.; Liu, Z.; Feng, L. Phase structure tuning of graphene supported Ni-NiO nanoparticles for enhanced urea oxidation performance. Electrochim. Acta 2021, 370, 137755. [Google Scholar] [CrossRef]
  4. Ding, R.; Li, X.; Shi, W.; Xu, Q.; Wang, L.; Jiang, H.; Yang, Z.; Liu, E. Mesoporous Ni-P nanocatalysts for alkaline urea electrooxidation. Electrochim. Acta 2016, 222, 455–462. [Google Scholar] [CrossRef]
  5. Rollinson, A.N.; Jones, J.; Dupont, V.; Twigg, M.V. Urea as a hydrogen carrier: A perspective on its potential for safe, sustainable and long-term energy supply. Energy Environ. Sci. 2014, 4, 1216–1224. [Google Scholar] [CrossRef] [Green Version]
  6. Wang, D.; Wang, Z.; Ren, S.Y.; Xu, J.; Wang, C.; Hu, P.; Fu, J.J. Molecular engineering of a colorless, extremely tough, superiorly self-recoverable, and healable poly(urethane-urea) elastomer for impact-resistant applications. Mater. Horiz. 2021, 8, 2238–2250. [Google Scholar] [CrossRef]
  7. Liu, Z.; Zhao, Q.; Wang, K.; Lee, D.; Qiu, W.; Wang, J. Urea hydrolysis and recovery of nitrogen and phosphorus as MAP from stale human urine. J. Environ. Sci. 2008, 20, 1018–1024. [Google Scholar] [CrossRef]
  8. Kameda, T.; Horikoshi, K.; Kumagai, S.; Saito, Y.; Yoshioka, T. Adsorption of urea, creatinine, and uric acid from three solution types using spherical activated carbon and its recyclability. Chin. J. Chem. Eng. 2020, 28, 2993–3001. [Google Scholar] [CrossRef]
  9. Zaher, A.; Shehata, N. Recent advances and challenges in management of urea wastewater: A mini review. IOP Conf. Series Mater. Sci. Eng. 2021, 1046, 012021. [Google Scholar] [CrossRef]
  10. Rittstieg, K.; Robra, K.H.; Somitsch, W. Aerobic treatment of a concentrated urea wastewater with simultaneous stripping of ammonia. Appl. Microbiol. Biotechnol. 2001, 56, 820–825. [Google Scholar] [CrossRef]
  11. Zequine, C.; Wang, F.; Li, X.; Guragain, D.; Mishra, S.R.; Siam, K.; Kahol, P.K.; Gupta, R.K. Nanosheets of CuCo2O4 as a high-performance electrocatalyst in urea oxidation. Appli. Sci. 2019, 9, 793. [Google Scholar] [CrossRef] [Green Version]
  12. Song, M.; Zhang, Z.; Li, Q.; Jin, W.; Wu, Z.; Fu, G.; Liu, X. Ni-foam supported Co(OH)F and Co-P nanoarrays for energy-efficient hydrogen production via urea electrolysis. J. Mater. Chem. A 2019, 7, 3697–3703. [Google Scholar] [CrossRef]
  13. Muthamizh, S.; Suresh, R.; Giribabu, K.; Manigandan, R.; Praveen Kumar, S.; Munusamy, S.; Narayanan, V. MnWO4 nanocapsules: Synthesis, characterization and electrochemical sensing property. J. Alloys Compd. 2015, 619, 601–609. [Google Scholar] [CrossRef]
  14. Wu, K.; Fu, P.; Ruan, B.; Wu, M.; Wu, M.; Wu, R. Performances of MnWO4@AC mixed oxide composite materials as Pt-free counter electrodes for high efficiency dye sensitized solar cells. New J. Chem. 2021, 45, 1686–1694. [Google Scholar] [CrossRef]
  15. Ramasamy Raja, V.; Karthika, A.; Suganthi, A.; Rajarajan, M. Facile synthesis of MnWO4/BiOI nanocomposites and their efficient photocatalytic and photoelectrochemical activities under the visible-light irradiation. J. Sci. Adv. Mater. Devices 2018, 3, 331–341. [Google Scholar] [CrossRef]
  16. Mostafa, K.; Davar, M.B.; Mojtaba, A. Synthesis, structural characterization and reactivity of manganese tungstate nanoparticles in the oxidative degradation of methylene blue. Compts Rendus Chim. 2015, 18, 199–203. [Google Scholar]
  17. Wu, W.; Qin, W.; He, Y.; Wu, Y.; Wu, T. The effect of pH value on the synthesis and photocatalytic performance of MnWO4 nanostructure by hydrothermal method. J. Exp. NanoSci. 2012, 7, 390–398. [Google Scholar] [CrossRef]
  18. Tiwari, A.; Singh, V.; Nagaiah, T.C. Tuning the MnWO4 morphology and its electrocatalytic activity towards oxygen reduction reaction. J. Mater. Chem. A 2018, 6, 2681–2692. [Google Scholar] [CrossRef]
  19. Trung, D.D.; Cuong, N.D.; Trung, K.Q.; Nguyen, T.D.; Toan, N.V.; Hung, C.M.; Hieu, N.V. Controlled synthesis of manganese tungstate nanorods for highly selective NH3 gas sensor. J. Alloys Compd. 2018, 735, 787–794. [Google Scholar] [CrossRef]
  20. Naik, K.K.; Gangan, A.S.; Pathak, A.; Chakraborty, B.; Nayak, S.K.; Rout, C.S. Facile hydrothermal synthesis of MnWO4 nanorods for non-enzymatic glucose sensing and supercapacitor properties with insights from density functional theory simulations. ChemistrySelect 2017, 2, 5707–5715. [Google Scholar] [CrossRef]
  21. Kang, R.; Zheng, L.; Tong, W.; Zhuangjun, F. Recent developments of transition metal compounds-carbon hybrid electrodes for high energy/power supercapacitors. Nano Micro Lett. 2021, 13, 129. [Google Scholar]
  22. Malkhandi, S.; Trinh, P.; Aswin, K.M.; Jayachandrababu, K.C.; Kindler, A.; Surya Prakash, G.K.; Narayanan, S.R. Electrocatalytic activity of transition metal oxide-carbon composites for oxygen reduction in alkaline batteries and fuel cells. J. Electrochem. Soc. 2013, 160, F943. [Google Scholar] [CrossRef]
  23. Hyun-Woo, S.; Ah-Hyeon, L.; Jae-Chen, K.; Gwang-Hee, L.; Dong-Wan, K. Hydrothermal realization of a hierarchical, flowerlike MnWO4@MWCNTs nanocomposites with enhanced reversible Li storage as a new anode material. Chem. Asian J. 2013, 8, 2851–2858. [Google Scholar]
  24. Anusha, V.; Eberechukwu, V.M.; Yingduo, C.; Chris, P. Carbon nanotube assembly and integration for applications. Nanoscale Res. Lett. 2019, 14, 220. [Google Scholar]
  25. Huang, L.; Yang, H.; Zhang, Y.; Xiao, W. Study on synthesis and antibacterial properties of AgNPs/GO nanocomposites. J. Nanomater. 2016, 2016, 5685967. [Google Scholar] [CrossRef] [Green Version]
  26. Mallick, S.; Samanta, A.; Retna Raj, C. Mesoporous carbon-supported manganese tungstate nanostructures for the development of zinc-air battery powered long lifespan asymmetric supercapacitor. Sustain. Energy Fuels 2020, 4, 4008–4017. [Google Scholar] [CrossRef]
  27. Sardar, K.; Thakur, S.; Maiti, S.; Besra, N.; Bairi, P.; Chandra, K.; Majumdhar, G.; Chattopadhyay, K.K. Amalgamation of MnWO4 nanorods with amorphous carbon nanotubes for highly stabilized energy efficient supercapacitor electrodes. Dalton Trans. 2021, 50, 5327–5347. [Google Scholar] [CrossRef] [PubMed]
  28. Xu, C.; Yixue, Z.; Puqun, X.; Mengqing, Z.; Huixia, W.; Shiping, Y. Graphene oxide/MnWO4 nanocomposite for magnetic resonance/photoacoustic dual-model imaging and tumor photothermochemotherapy. Carbon 2018, 138, 397–409. [Google Scholar]
  29. Jianhua, T.; Jianfeng, S.; Na, L.; Mingxin, Y. Facile synthesis of layered MnWO4/ reduced graphene oxide for supercapacitor applications. J. Alloys Compd. 2016, 666, 15–22. [Google Scholar]
  30. Dubale, A.A.; Su, W.N.; Tamirat, A.G.; Pan, C.J.; Aragaw, B.A.; Chen, H.M.; Chen, C.H.; Hwang, B.J. The synergetic effect of graphene on Cu2O nanowire arrays as a highly efficient hydrogen evolution photocathode in water splitting. J. Mater. Chem. A 2014, 2, 18383–18397. [Google Scholar] [CrossRef]
  31. Ahn, S. Changing the sp2 carbon clusters in graphene oxide during exfoliation. Trans. Electr. Electron. Mat. 2015, 16, 49–52. [Google Scholar] [CrossRef] [Green Version]
  32. Dinh, T.T.; Van, N.N. rGO/persulfate metal-free catalytic system for the degradation of tetracycline: Effect of reaction parameters. Mater. Res. Express 2020, 7, 075501. [Google Scholar]
  33. Dai, R.C.; Ding, X.; Wang, Z.P.; Zhang, Z.M. Pressure and temperature dependence of Raman scattering of MnWO4. Chem. Phys. Lett. 2013, 586, 76–80. [Google Scholar] [CrossRef]
  34. Iliev, M.N.; Gospodinov, M.M.; Litvinchuk, A.P. Raman spectroscopy of MnWO4. Phys. Rev. B 2009, 80, 212302. [Google Scholar] [CrossRef]
  35. Jung, J.Y.; Yi, S.S.; Hwang, D.H.; Son, C.S. Structure, luminescence, and magnetic properties of crystalline manganese tungstate doped with rare earth ion. Materials 2021, 14, 3717. [Google Scholar] [CrossRef] [PubMed]
  36. Sundaresan, R.; Mariyappan, V.; Chen, S.M.; Keerthi, M.; Ramachandran, R. Electrochemical sensor for detection of tryptophan in the milk sample based on MnWO4 nanoplates encapsulated RGO nanocomposite. Colloid Surf. A 2021, 625, 126889. [Google Scholar] [CrossRef]
  37. Teeparthi, S.R.; Awin, E.W.; Kumar, R. Dominating role of crystal structure over defect chemistry in black and white zirconia on visible light photocatalytic activity. Sci. Rep. 2018, 8, 5541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Shulga, Y.M.; Baskakov, S.A.; Knerelman, E.I.; Davidova, G.I.; Badamshina, E.R.; Shulga, N.Y.; Skryleva, E.A.; Agapov, A.L.; Voylov, D.N.; Sokolov, A.P.; et al. Carbon nanomaterial produced by microwave exfoliation of graphite oxide: New insights. RSC Adv. 2014, 4, 587–592. [Google Scholar] [CrossRef]
  39. Sreekanth, T.V.M.; Tamilselvan, M.; Yoo, K.; Kim, J. Microwave-assisted in situ growth of VO2 nanoribbons on Ni foam as inexpensive bifunctional electrocatalyst for the methanol oxidation and oxygen evolution reactions. Appl. Surf. Sci. 2021, 570, 151119. [Google Scholar] [CrossRef]
  40. Askari, M.B.; Salarizadeh, P.; Seifi, M.; Zadeh, M.H.R.; Di Bartolomeo, A. ZnFe2O4 nanorods on reduced graphene oxide as advanced supercapacitor electrodes. J. Alloys Compd. 2021, 860, 158497. [Google Scholar] [CrossRef]
  41. Sreekanth, T.V.M.; Dillip, G.R.; Nagajyothi, P.C.; Yoo, K.; Kim, J. Integration of Marigold 3D flower-like Ni-MOF self-assembled on MWCNTs via microwave irradiation for high-performance electrocatalytic alcohol oxidation and oxygen evolution reactions. Appl. Catal. B Environ. 2021, 285, 119793. [Google Scholar] [CrossRef]
  42. Imanzadeh, H.; Habibi, B. Electrodeposition of ternary CuNiPt alloy nanoparticles on graphenized pencil lead electrode as a new electrocatalyst for electro-oxidation of ethanol. Solid State Sci. 2020, 105, 106239. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the electrocatalysts, MW-7 (a), MW-12 (b), MW@GO-7 (c) and [email protected] (d).
Figure 1. XRD patterns of the electrocatalysts, MW-7 (a), MW-12 (b), MW@GO-7 (c) and [email protected] (d).
Nanomaterials 12 00085 g001
Figure 2. Raman analysis of the electrocatalysts, GO (a), MW-7 (b), MW-12 (c), MW@GO-7 (d), and MW@GO-12 (e).
Figure 2. Raman analysis of the electrocatalysts, GO (a), MW-7 (b), MW-12 (c), MW@GO-7 (d), and MW@GO-12 (e).
Nanomaterials 12 00085 g002
Figure 3. XPS analysis of MW-12 (above) and MW@GO-12 (below), survey scan spectrum (a), Mn-2p of MW-12 (above) and MW@GO-12 (below) (b), W 4f of MW-12 (above) and MW@GO-12 (below) (c), O 1s of MW-12 (above) and MW@GO-12 (below) (d), and C 1s of MW-12 (above) and MW@GO-12 (below) (e).
Figure 3. XPS analysis of MW-12 (above) and MW@GO-12 (below), survey scan spectrum (a), Mn-2p of MW-12 (above) and MW@GO-12 (below) (b), W 4f of MW-12 (above) and MW@GO-12 (below) (c), O 1s of MW-12 (above) and MW@GO-12 (below) (d), and C 1s of MW-12 (above) and MW@GO-12 (below) (e).
Nanomaterials 12 00085 g003
Figure 4. SEM image of hydrothermally synthesized electrocatalysts. MW-7 (a1,a2); MW@GO-7 (b1,b2); MW-12 (c1,c2); and MW@GO-12 (d1,d2).
Figure 4. SEM image of hydrothermally synthesized electrocatalysts. MW-7 (a1,a2); MW@GO-7 (b1,b2); MW-12 (c1,c2); and MW@GO-12 (d1,d2).
Nanomaterials 12 00085 g004
Figure 5. TEM analysis of electrocatalysts, low resolution MW-12 (a1), MW@GO-12 (b1), high resolution MW-12 (a2) MW@GO-12 (b2) and the corresponding SAED pattern MW-12 (a3), MW@GO-12 (b3).
Figure 5. TEM analysis of electrocatalysts, low resolution MW-12 (a1), MW@GO-12 (b1), high resolution MW-12 (a2) MW@GO-12 (b2) and the corresponding SAED pattern MW-12 (a3), MW@GO-12 (b3).
Nanomaterials 12 00085 g005
Figure 6. CV curves of the electrocatalysts at various scan rates; MW-7 (a), MW-12 (b), MW@GO-7 (c), and MW@GO-12 (d).
Figure 6. CV curves of the electrocatalysts at various scan rates; MW-7 (a), MW-12 (b), MW@GO-7 (c), and MW@GO-12 (d).
Nanomaterials 12 00085 g006
Figure 7. Comparative CV curves of the electrocatalysts for the ECSA analysis (a) and the corresponding electrochemical double-layer capacity linear fit plots of the electrocatalysts (b).
Figure 7. Comparative CV curves of the electrocatalysts for the ECSA analysis (a) and the corresponding electrochemical double-layer capacity linear fit plots of the electrocatalysts (b).
Nanomaterials 12 00085 g007
Figure 8. CV curves of the electrocatalysts of UOR analysis at various scan rates; MW-7 (a), MW-12 (b), MW@GO-7 (c), and MW@GO-12 (d).
Figure 8. CV curves of the electrocatalysts of UOR analysis at various scan rates; MW-7 (a), MW-12 (b), MW@GO-7 (c), and MW@GO-12 (d).
Nanomaterials 12 00085 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nagajyothi, P.C.; Yoo, K.; Ramaraghavulu, R.; Shim, J. Hydrothermal Synthesis of MnWO4@GO Composite as Non-Precious Electrocatalyst for Urea Oxidation. Nanomaterials 2022, 12, 85. https://doi.org/10.3390/nano12010085

AMA Style

Nagajyothi PC, Yoo K, Ramaraghavulu R, Shim J. Hydrothermal Synthesis of MnWO4@GO Composite as Non-Precious Electrocatalyst for Urea Oxidation. Nanomaterials. 2022; 12(1):85. https://doi.org/10.3390/nano12010085

Chicago/Turabian Style

Nagajyothi, Patnamsetty Chidanandha, Kisoo Yoo, Rajavaram Ramaraghavulu, and Jaesool Shim. 2022. "Hydrothermal Synthesis of MnWO4@GO Composite as Non-Precious Electrocatalyst for Urea Oxidation" Nanomaterials 12, no. 1: 85. https://doi.org/10.3390/nano12010085

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