Next Article in Journal
Terahertz Broadband Polarization Conversion for Transmitted Waves Based on Graphene Plasmon Resonances
Previous Article in Journal
Mechanistic Insight of Sensing Hydrogen Phosphate in Aqueous Medium by Using Lanthanide(III)-Based Luminescent Probes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of NiCo2O4 Nanostructures and Their Electrochemial Properties for Glucose Detection

1
School of Materials Science and Engineering, Inha University, 25 Younghyun-Dong, Incheon 22201, Korea
2
Department of Materials Science and Engineering, Hanyang University, 222 Wangsimni-ro, Seoul 04763, Korea
3
Korea Institute of Industrial Technology, 137-41 Gwahakdanji-ro, Gangneung 25440, Korea
4
Korea Institute of Industrial Technology, 156 Gaetbeol-ro, Incheon 21999, Korea
5
Daegu Mechatronics & Materials Institute, Seongseogongdan-r0 11-gil, Dalseo-gu, Daegu 42714, Korea
6
Department of Energy Engineering, Konkuk University, 120 Neungdong-ro, Seoul 05029, Korea
7
Department of Advanced Materials Engineering, Kyonggi University, 154-42 Gwanggyosan-ro, Suwon 16227, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2021, 11(1), 55; https://doi.org/10.3390/nano11010055
Submission received: 30 November 2020 / Revised: 22 December 2020 / Accepted: 24 December 2020 / Published: 28 December 2020

Abstract

:
In this work, we prepared spinel-type NiCo2O4 (NCO) nanopowders as a low-cost and sensitive electrochemical sensor for nonenzymatic glucose detection. A facile and simple chemical bath method to synthesize the NCO nanopowders is demonstrated. The effect of pH and annealing temperature on the formation mechanism of NCO nanoparticles was systematically investigated. Our studies show that different pHs of the precursor solution during synthesis result in different intermediate phases and relating chemical reactions for the formation of NCO nanoparticles. Different morphologies of the NCO depending on pHs are also discussed based on the mechanism of growth. Electrochemical performance of the prepared NCO was characterized towards glucose, which reveals that sensitivity and selectivity of the NCO are significantly related with the final microstructure combined with constituent species with multiple oxidation states in the spinel structure.

1. Introduction

Diabetes is a disease that impairs the human body’s ability to control glucose or sugar in the blood [1,2,3]. High glucose levels can cause serious health problems including heart disease, stroke, and kidney disease, and thus, it is important to maintain the glucose levels in blood via continuous monitoring with accurate detection of glucose [4,5,6]. Since the oxygen electrode was developed, different types of glucose sensors including an optical sensor and electrochemical sensor have been introduced [7,8]. Especially, the electrochemical glucose sensor has been actively studied to increase the sensitivity and to reduce the detection time for its excellent applicability to the real-time monitoring system.
Electrochemical glucose sensors can be divided into enzymatic and nonenzymatic sensors. The enzymatic glucose sensor exhibits high sensitivity and selectivity through direct immobilization of glucose oxidase. However, its natural limitations include a short lifetime due to poor chemical and thermal stability, and expensive processing costs hindering further advancement of the enzymatic biosensor for industrial applications [9,10,11]. Thanks to their several advantages such as long-term stability and reproducibility, and effective processing cost, different types of nonenzymatic electrocatalysts have been developed including noble metals, transition metals and alloys, and metal oxides [12,13,14]. Among them, transition-metal-based oxides have drawn great interest due to their configurational flexibility of transition metals, which promotes glucose oxidation with excellent sensitivity and selectivity [15,16,17,18].
Different types of synthesis methods for transition-metal-based oxides have been developed including hydroxide decomposition [19], nanocasting [20], electrodeposition [21], coprecipitation [22], and hydrothermal synthesis [23]. However, their complex and energy-consuming processing hinders the industrial application to glucose sensors. Therefore, exploring fast, environmentally friendly, and energy-efficient synthetic methods for transition-metal-based oxides is urgent. Recently, it has been reported that chemical bath synthesis has proven to be capable of controlling the size and morphology of materials by control of reaction parameters, such as temperature, pH, and solvent concentration [23,24,25,26]. It is expected that the transition-metal-based oxides prepared by chemical bath synthesis can be applied as a glucose sensor. However, there is limited information available on the effect of the processing parameters of chemical bath synthesis on the electrochemical performance for glucose detection.
In this work, spinel-type NiCo2O4 (NCO) are successfully synthesized by a simple and facile chemical bath method. The morphology of the NCOs depending on pH during synthesis was investigated, which is closely related to the growth mechanism. Also, the formation of the NCOs was investigated to show excellent electrochemical performance for glucose detection including sensitivity, selectivity, and low detection limits.

2. Materials and Methods

2.1. Materials and Reagents

Nickel nitrate hexahydrate (Ni(NO3)2·6H2O, 99.99%), cobalt nitrate hexahydrate Co(NO3)2·6H2O, 99.9%), ammonia solution (NH4OH), sodium hydroxide (NaOH), D-(+)-glucose, uric acid (UA), dopamine (DA), L-ascorbic acid (LA), and acetic acid (AA) were purchased from Sigma-Aldrich (Seoul, Korea). All of the reagents used were of analytical grade and used as received without further purification.

2.2. Chemical Bath Synthesis of NiCo-Layered Double Hydroxide and NiCo2O4

An aqueous solution of Ni(NO3)2·6H2O (0.005 M) and Co(NO3)2·6H2O (0.01 M) was prepared by dissolving the salts in 100 mL of deionized water (DI water) with vigorous stirring for 60 min. Ammonia solution was added to the precursor solution until each scheduled pH value (11, 12, 13, and 14) was reached, followed by heating at 80 °C on a hotplate for 14 h, resulting in a thick, viscous, dark greenish fluid. It is noted that the pH of the precursor solution without the addition of ammonia solution was 8. Obtained products prepared at pH 8, 11, 12, 13, 14 are denoted as NCO8B, NCO11B, NCO12B, NCO13B, and NCO14B, respectively. The fluid was filtered through filter paper several times with DI water and ethanol. Subsequently, the filtered materials were dried in air for 24 h followed by annealing in air for 4 h at different temperatures at 450 °C with a heating rate of 10 °C/min, which turned the material black in color. Final products prepared at pH 8, 11, 12, 13, 14 are denoted as NCO8, NCO11, NCO12, NCO13, and NCO14, respectively. Also, crystalline NiCo2O4 with spinel structure is abbreviated as NCO. For clarity, detailed information is presented in Table S1 in the Supporting Information.

2.3. Material Characterizations and Electrochemical Measurements

The morphologies of NCO were investigated by scanning electron microscopy (SEM, Nova NanoSEM 450, FEI, Portland, OR, USA). The X-ray diffraction (XRD) patterns were collected using a PANalytical X-ray diffractometer (Empyrean, PANanalytical, Almelo, Nederland) with Cu-Kα radiation (λ = 0.1548 nm).
All the electrochemical measurements including cyclic voltammetry (CV) and chronoamperometry (CA) were performed on an IVIUMSTAT electrochemical analyzer (IVIUMSTAT, Ivium Technologies, Eindhoven, Netherland) using a three-electrode system in a 0.1 M aqueous NaOH solution at room temperature. A glassy carbon electrode (GCE), an Ag/AgCl electrode, and a platinum plate were used as the working electrode, reference electrode, and counter electrode, respectively. The samples (10 mg), ethanol (0.5 mL), and Nafion solution (30 μL) were mixed for the preparation of the working electrode. Subsequently, drop-casting of the dispersion on the GC electrode was performed followed by drying under ambient conditions overnight.
The CV response was recorded between 0 and 0.6 V at different scanning rates of 5–100 mV/s. To get the optimal potential of the CA response of the sample, glucose was added to the 0.1 M NaOH solution at various potentials from +0.4 to +0.6 V, as shown in Figure S1. The optimal potential of +0.5 V was chosen, which was highly responsive and stable as the working potential for glucose detection. The CA response of the samples to the glucose was carried out at an applied potential of 0.5 V under stirred conditions. For sensing performance evaluation, 0.01–6 mM glucose solutions were used, with LA, DA, AA, and UA detection performed at concentrations of 0.1 mM in 0.1 M NaOH alkaline electrolyte.

3. Results

The influence of pH during synthesis on the crystallization of the NCO was investigated as shown in Figure 1. As shown in Figure 1a, Ni2(NO3)2(OH)2 · 2H2O and Co(NO3)2· 6H2O were observed from NCO8B, while NiCo-layered double hydroxide (NiCo-LDH) was observed from NCO11B to 14B [27]. As annealing temperature increased, different phase transformations were observed depending on pH as shown in Figure 1b, c. Phase transformation of the NCO8B to spinel-type NCO (NCO8) occurred from 150 °C, and NCO single phase with improved crystallinity was observed at 350 °C, which implies the chemical reaction of the intermediates (Ni3(NO3)2(OH)4 and Co(NO3)2· 6H2O) with oxygen for the formation of spinel-type NCO8. However, NCO11B–14B with LDH structure directly transformed into spinel-type NCOs (NCO11—14) without any chemical reaction between intermediates with increasing temperature. Regardless of pH during synthesis, spinel-type NCO single phase was observed after annealing at 450 °C (Figure 1d) [28]. In the chemical bath process, an increase of OH ions in a precursor solution containing Ni2+, Co2+, and NO3 occurred by applying NH4OH, which turned into an alkaline condition (pH = 11~14) in the chemical bath. Under the condition, the chemical reaction among different ions such as Ni2+, Co2+, and OH ions led to the formation of the NiCo-LDH as expressed by reaction (1) [29,30]. Simultaneously, H2O molecules and NO3− ions were intercalated into NiCo-LDH interlayer to retain the LDH structure through a hydrogen bond. Subsequent annealing of the as-synthesized NiCo-LDH caused structural transformation into spinel-type NiCo2O4 (NCO), as described by the reaction (2) [29,30,31,32].
Ni2+ + 2Co2+ + 6OH → NiCo2(OH)6 (NiCo-LDH)
2NiCo2(OH)6 + O2 → 2NiCo2O4 + 6H2O
The pH depending on OH ions in the precursor solution also determines the morphology of the NCO during synthesis. As depicted in Figure 2a, a high concentration of OH ions for the reaction environment reveals the flower-like morphology as observed from NCO11B–13B, which is originated from anisotropic grain growth of LDH. [33,34] However, the unique flower-like morphology disappears at pH 14 (NCO14B) due to further grain growth of LDHs. After annealing, the morphological transformation of the NCOB prepared at different pH values was observed as shown in Figure 2b. Spherical nanoparticles were observed from NCO8, derived by the chemical reaction between nitrates and oxygen. However, the transformation from NiCo-LDH prepared at high pH (NCO11–13) to NCO maintains the sheet-like morphology. It is noted that the nanosheets consist of the assembly of spherical nanoparticles after transformation from LDH to NCO. Also, an increase of pH from 8 to 13 during synthesis results in smaller particle sizes after annealing. As expected, the morphology of the NCO14 is transferred from NCO14B. Regardless of different pH, homogenous distribution of the constituent elements (Ni, Co, and O) was observed from NCOBs and NCOs as shown in Figure S2, which implies that the synthesis route using the chemical bath method is applicable to prepare the spinel-type NiCo2O4 nanostructures at comparatively low temperature (450 °C).
Different microstructures of the NCOs can be determined by different processing conditions including pH, thus showing different electrochemical properties. The dependence of the cyclic voltammetric (CV) curves for NCOs on the pH during chemical bath synthesis was measured to investigate the electrochemical behavior of the NCOs under alkaline conditions (0.1 M NaOH) at various scan rates as shown in Figure 3. There is a negligible effect of pH on the redox peak potentials for the NCO electrodes. Regardless of pH, redox peak currents of the NCOs were increased with increasing CV scan rate. Oxidation peaks of the NCOs correspond to Ni2+/Ni3+, Co2+/Co3+, and Co3+/Co4+ due to oxidation of Ni2+, Co2+ and Co3+ to Ni3+, Co3+, and Co4+, respectively. It is noted that the redox peak potential of Co3+/Co4+ is close to that of Ni2+/Ni3+ and Co2+/Co3+, which shows overlapped redox peaks in the CV curve [35,36,37]. In addition, the reduction peaks of NCOs are attributed to the Ni3+, Co4+, and Co3+ to Ni2+, Co3+, and Co2+, respectively. The redox peak currents for the NCOs at the square root of the scan rates are presented in Figure 3f. All NCOs synthesized at different pH show a linear proportionality relationship between the redox peak currents and the square root of the scan rates, suggesting that NCOs undergo diffusion-controlled electrochemical behavior [38,39]. Also, CV responses of NCOs were synthesized at different pH in response to 5 mM glucose under alkaline conditions (0.1 M NaOH) at a scan rate of 50 mVs−1, as shown in Figure S3. Regardless of the pH, all NCOs oxidize glucose (C6H12O6) to gluconolactone (C6H10O6), which implies that NCOs synthesized at different pH can be applied to electrochemical glucose sensors [40,41].
The electrochemical performance of NCOs on glucose oxidation was investigated as shown in Figure 4. Chronoamperometry (CA) responses of NCOs were measured by stepwise changes in glucose concentrations in 0.1 M NaOH at 60 s intervals under an applied potential of 0.50 V. With increasing pH during synthesis, the sensitivity of NCO shows in the range between 48.71–146.26 μA/mM (cm2) with 0.995–0.998 (R2), which shows a linear detection limit in the range between 0.01–6 mM. The limits of detection (LOD) of NCOs are in the range between 0.0475–0.393 µM, as shown in Figure S4. Based on the results from the CA test, NCO13 shows superior electrochemical performance for glucose detection, supported by excellent linear sensitivity (146.24 μA/mM (cm2)) in a wide detection range. The excellent sensitivity of the NCOs is strongly associated with the redox reaction of active sites. As expected, Co2+, Co3+, Ni2+, and Ni3+ as active sites in NCO were investigated in XPS results, as shown in Figure S5. In the XPS spectra of Co2p and Ni2p, Co2+/3+ and Ni2+/3+ were observed on the NCO. It is believed that multi-valence states of Ni and Co cations play an important role as oxidizing agents for glucose detection [17,42,43]. Thus, the reversible conversion of Ni2+/Ni3+ and Co2+/Co3+ in NCOs enables repetitive glucose detection [44,45,46].
The selectivity of NCOs is also an important factor for accurate glucose detection: current response to other reagents can detract from the determination of glucose [47,48,49]. The selectivity of the NCOs depending on different pH was investigated as shown in Figure 5. The current response of all NCOs to glucose is obvious. However, there is no change in the current response to uric acid (UA), dopamine (DA), L-ascorbic acid (LA), and acetic acid (AA) at the same glucose concentration of 1 and 2 mM, which implies that NCOs have excellent selectivity for glucose [41,50,51]. Based on the results above, it is suggested that the sensitivity of the NCOs for glucose detection is strongly dependent on morphology, however, selectivity for glucose detection is significantly determined by the redox reaction of the chemical components. The electrochemical performance of the NCOs prepared in this work was summarized in Table 1. The sensitivity of the NiCo2O4/rGO shows the highest value when compared to the other materials including NCOs, induced by the excellent electrochemical performance of the NCOs combined with fast electron transfer from supportive rGO [52,53,54,55]. However, NCOs show higher sensitivity with a lower detection limit in response to glucose compared to the rest of the materials in Table 1 [56,57,58]. Therefore, the NCOs as pure oxides can be expected to be practical for the application of glucose sensors.

4. Conclusions

Spinel-type NiCo2O4 (NCO) nanostructure was synthesized by a simple chemical bath method for electrochemical glucose sensors. Although different chemical reactions and formation of intermediates depending on pHs occur during synthesis, only spinel-type NCO was prepared after annealing at 450 °C. However, the morphology and particle size of the NCOs are strongly influenced by pH value, which emphasizes the importance of synthetic routes for the formation of the NCO. In our study, NCO13 with flower-like morphology assembled with small nanoparticles shows superior glucose detection including a sensitivity of 143 μA/mM (cm2) up to 6 mM with good linearity. It is revealed that the different morphology and particle size of the NCO determine the sensitivity for glucose detection. Also, the selectivity of the NCO is determined by the unique spinel structure and redox reaction of Ni and Co ions.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/11/1/55/s1, Figure S1: CA response of NCO13 electrode upon addition of 1 mM glucose in 1m M NaOH solution at different applied potentials. Figure S2: SEM-elemental mapping images of (a) NCOBs (8B, 11B, 12B, 13B, and 14B), and (b) NCOs (8, 11, 12, 13, and 14). Figure S3: CV curves of (a) NCO8, (b) NCO11, (c) NCO12, (d) NCO13, and (e) NCO14 electrodes in the absence of glucose and with 5 mM concentration of glucose at a scan rate 50 mVs−1. Figure S4: CA response of (a) NCO8, (b) NCO11, (c) NCO12, (d) NCO13, and (e) NCO14 electrodes with the addition of 10 μM glucose in 0.1 M NaOH solution at 0.50 V. Figure S5: The XPS spectra of Ni2p and Co2p (NCO13). Table S1: Sample notations of As-prepared and after annealing samples.

Author Contributions

K.-b.J.: data curation, formal analysis, investigation, and writing-original draft. K.R.P.: data curation, formal analysis, investigation, and writing-original draft. K.M.K.: investigation, and methodology. S.-k.H.: data curation, and investigation. J.-e.J.: data curation, investigation, and methodology. Y.S.S.: data curation, and investigation. S.-k.P.: formal analysis, and funding acquisition. K.-i.M.: formal analysis, and funding acquisition. C.A.: data curation, investigation, and resources. S.-c.L.: formal analysis, and project administration. J.L.: investigation, and methodology. J.C.K.: writing-review and editing. H.H.: writing-review and editing. S.M.: conceptualization, formal analysis, writing-review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This study has been conducted with the support of the Korea Institute of Industrial Technology as “Development of intelligent root technology with add-on modules (kitech EO200017)”. This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. NRF-2020R1A2C1101466).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cho, N.; Shaw, J.E.; Karuranga, S.; Huang, Y.; da Rocha Fernandes, J.D.; Ohlrogge, A.W.; Malanda, B. IDF Diabetes Atlas: Global estimates of diabetes prevalence for 2017 and projections for 2045. Diabetes Res. Clin. Pract. 2018, 138, 271–281. [Google Scholar] [CrossRef]
  2. Song, S.O.; Song, Y.D.; Nam, J.Y.; Park, K.H.; Yoon, J.H.; Son, K.M.; Lim, D.H. Epidemiology of type 1 diabetes mellitus in Korea through an investigation of the national registration project of type 1 diabetes for the reimbursement of glucometer strips with additional analyses using claims data. Diabetes. Metab. J. 2016, 40, 35–45. [Google Scholar] [CrossRef] [Green Version]
  3. Abellán-Llobregat, A.; Jeerapan, I.; Bandodkar, A.; Vidal, L.; Canals, A.; Wang, J.; Morallon, E. A stretchable and screen-printed electrochemical sensor for glucose determination in human perspiration. Biosens. Bioelectron. 2017, 91, 885–891. [Google Scholar] [CrossRef] [Green Version]
  4. Ensafi, A.A.; Zandi-Atashbar, N.; Rezaei, B.; Ghiaci, M.; Taghizadeh, M. Silver nanoparticles decorated carboxylate functionalized SiO2, new nanocomposites for non-enzymatic detection of glucose and hydrogen peroxide. Electrochim. Acta. 2016, 214, 208–216. [Google Scholar] [CrossRef]
  5. Bandodkar, A.J.; Wang, J. Non-invasive wearable electrochemical sensors: A review. Trends Biotechnol. 2014, 32, 363–371. [Google Scholar] [CrossRef]
  6. Yao, H.; Liao, Y.; Lingley, A.R.; Afanasiev, A.; Lähdesmäki, I.; Otis, B.P.; Parviz, B.A. A contact lens with integrated telecommunication circuit and sensors for wireless and continuous tear glucose monitoring. J. Micromech. Microeng. 2012, 22, 075007. [Google Scholar] [CrossRef]
  7. Kim, D.M.; Moon, J.M.; Lee, W.C.; Yoon, J.H.; Choi, C.S.; Shim, Y.B. A potentiometric non-enzymatic glucose sensor using a molecularly imprinted layer bonded on a conducting polymer. Biosens. Bioelectron. 2017, 91, 276–283. [Google Scholar] [CrossRef] [PubMed]
  8. Chung, J.S.; Hur, S.H. A highly sensitive enzyme-free glucose sensor based on Co3O4 nanoflowers and 3D graphene oxide hydrogel fabricated via hydrothermal synthesis. Sens. Actuators B Chem. 2016, 223, 76–82. [Google Scholar]
  9. Wang, J. Glucose biosensors: 40 years of advances and challenges. Electroanalysis. 2013, 13, 983–988. [Google Scholar] [CrossRef]
  10. Jia, W.Z.; Wang, K.; Zhu, Z.J.; Song, H.T.; Xia, X.H. One-step immobilization of glucose oxidase in a silica matrix on a Pt electrode by an electrochemically induced sol−gel process. Langmuir 2007, 23, 11896–11900. [Google Scholar] [CrossRef] [PubMed]
  11. Bo, X.; Ndamanisha, J.C.; Bai, J.; Guo, L. Nonenzymatic amperometric sensor of hydrogen peroxide and glucose based on Pt nanoparticles/ordered mesoporous carbon nanocomposite. Talanta. 2010, 82, 85–91. [Google Scholar] [CrossRef]
  12. Yuan, J.H.; Wang, K.; Xia, X.H. Highly ordered platinum-nanotubule arrays for amperometric glucose sensing. Adv Funct Mater. 2005, 15, 803–809. [Google Scholar] [CrossRef]
  13. Martins, A.; Ferreira, V.; Queirós, A.; Aroso, I.; Silva, F.; Feliu, J. Enantiomeric electro-oxidation of d-and l-glucose on chiral gold single crystal surfaces. Electrochem Commun. 2003, 5, 741–746. [Google Scholar] [CrossRef]
  14. Niu, X.; Lan, M.; Chen, C.; Zhao, H. Nonenzymatic electrochemical glucose sensor based on novel Pt–Pd nanoflakes. Talanta. 2012, 99, 1062–1067. [Google Scholar] [CrossRef]
  15. Yuan, C.; Wu, H.B.; Xie, Y.; Lou, X.W. Mixed transition-metal oxides: Design, synthesis, and energy-related applications. Angew. Chem. Int. Ed. 2014, 53, 1488–1504. [Google Scholar] [CrossRef] [PubMed]
  16. Wu, Z.; Zhu, Y.; Ji, X. NiCo2O4-based materials for electrochemical supercapacitors. J. Mater. Chem. A 2014, 2, 14759–14772. [Google Scholar] [CrossRef]
  17. Naik, K.K.; Kumar, S.; Rout, C.S. Electrodeposited spinel NiCo2O4 nanosheet arrays for glucose sensing application. RSC Adv. 2015, 5, 74585–74591. [Google Scholar] [CrossRef]
  18. Li, G.; Huo, H.; Xu, C. Ni0.31Co0.69S2 nanoparticles uniformly anchored on a porous reduced graphene oxide framework for a high-performance non-enzymatic glucose sensor. J. Mater. Chem. A 2015, 3, 4922–4930. [Google Scholar] [CrossRef]
  19. Marco, J.F.; Gancedo, J.R.; Gracia, M.; Gautier, J.L.; Rios, E.; Berry, F.J. Characterization of the nickel cobaltite, NiCo2O4, prepared by several methods: An XRD, XANES, EXAFS, and XPS study. J. Solid State Chem. 2000, 153, 74–81. [Google Scholar] [CrossRef]
  20. Cabo, M.; Pellicer, E.; Rossinyol, E.; Estrader, M.; López-Ortega, A.; Nogués, J.; Baró, M.D. Synthesis of compositionally graded nanocast NiO/NiCo2O4/Co3O4 mesoporous composites with tunable magnetic properties. J. Mater. Chem. 2010, 20, 7021–7028. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, N.; Sun, B.; Zhao, P.; Yao, M.; Hu, W.; Komarneni, S. Electrodeposition preparation of NiCo2O4 mesoporous film on ultrafine nickel wire for flexible asymmetric supercapacitors. Chem. Eng. J. 2018, 345, 31–38. [Google Scholar] [CrossRef]
  22. Zhang, J.; Sun, Y.; Li, X.; Xu, J. Fabrication of NiCo2O4 nanobelt by a chemical co-precipitation method for non-enzymatic glucose electrochemical sensor application. J. Alloys Compd. 2020, 154796. [Google Scholar] [CrossRef]
  23. Wang, J.; Zhang, Y.; Ye, J.; Wei, H.; Hao, J.; Mu, J.; Hussain, S. Facile synthesis of three-dimensional NiCo2O4 with different morphology for supercapacitors. RSC Adv. 2016, 6, 70077–70084. [Google Scholar] [CrossRef]
  24. Chi, B.; Li, J.; Han, Y.; Chen, Y. Effect of temperature on the preparation and electrocatalytic properties of a spinel NiCo2O4/Ni electrode. Int. J. Hydrog. Energy. 2004, 29, 605–610. [Google Scholar] [CrossRef]
  25. Guragain, D.; Zequine, C.; Poudel, T.; Neupane, D.; Gupta, R.K.; Mishra, S.R. Influence of Urea on the Synthesis of NiCo2O4 Nanostructure: Morphological and Electrochemical Studies. J. Nanosci. Nanotechnol. 2020, 20, 2526–2537. [Google Scholar] [CrossRef]
  26. Cui, B.; Lin, H.; Li, J.B.; Li, X.; Yang, J.; Tao, J. Core–ring structured NiCo2O4 nanoplatelets: Synthesis, characterization, and electrocatalytic applications. Adv. Funct. Mater. 2008, 18, 1440–1447. [Google Scholar] [CrossRef]
  27. Valdez, R.; Grotjahn, D.B.; Smith, D.K.; Quintana, J.M.; Olivas, A. Nanosheets of Co-(Ni and Fe) layered double hydroxides for electrocatalytic water oxidation reaction. Int. J. Electrochem. Sci. 2015, 10, 909–918. [Google Scholar]
  28. Hu, L.; Wu, L.; Liao, M.; Fang, X. High-Performance NiCo2O4 Nanofilm Photodetectors Fabricated by an Interfacial Self-Assembly Strategy. Adv. Mater. 2011, 23, 1988–1992. [Google Scholar] [CrossRef]
  29. Cai, L.; Li, Y.; Xiao, X.; Wang, Y. The electrochemical performances of NiCo2O4 nanoparticles synthesized by one-step solvothermal method. Ionics (Kiel) 2017, 23, 2457–2463. [Google Scholar] [CrossRef]
  30. Chen, X.; Li, H.; Xu, J.; Jaber, F.; Musharavati, F.; Zalezhad, E.; Liu, J. Synthesis and Characterization of a NiCo2O4@NiCo2O4 Hierarchical Mesoporous Nanoflake Electrode for Supercapacitor Applications. Nanomaterials. 2020, 10, 1292. [Google Scholar] [CrossRef]
  31. Uke, S.J.; Chaudhari, G.N.; Bodade, A.B.; Mardikar, S.P. Morphology dependant electrochemical performance of hydrothermally synthesized NiCo2O4 nanomorphs. Mater. Sci. Technol. 2020, 3, 289–298. [Google Scholar] [CrossRef]
  32. Liu, Y.; Jiang, J.; Yuan, Y.; Jiang, Q.; Yan, C. Vertically Aligned NiCo2O4 Nanosheet-Encapsulated Carbon Fibers as a Self-Supported Electrode for Superior Li+ Storage Performance. Nanomaterials. 2019, 9, 1336. [Google Scholar] [CrossRef] [Green Version]
  33. Yan, T.; Li, R.; Li, Z. Nickel–cobalt layered double hydroxide ultrathin nanoflakes decorated on graphene sheets with a 3D nanonetwork structure as supercapacitive materials. Mater. Res. Bull. 2014, 51, 97–104. [Google Scholar] [CrossRef]
  34. Song, Q.; Tang, Z.; Guo, H.; Chan, S.L.I. Structural characteristics of nickel hydroxide synthesized by a chemical precipitation route under different pH values. J. Power Sources. 2002, 112, 428–434. [Google Scholar] [CrossRef]
  35. Guo, Q.; Zeng, W.; Liu, S.; Li, Y. In situ formation of Co3O4 hollow nanocubes on carbon cloth-supported NiCo2O4 nanowires and their enhanced performance in non-enzymatic glucose sensing. Nanotechnology. 2020, 31, 265501. [Google Scholar] [CrossRef]
  36. Wang, J.; Qiu, T.; Chen, X.; Lu, Y.; Yang, W. Hierarchical hollow urchin-like NiCo2O4 nanomaterial as electrocatalyst for oxygen evolution reaction in alkaline medium. Power Sources. 2014, 268, 341–348. [Google Scholar] [CrossRef]
  37. Yu, Z.; Li, H.; Zhang, X.; Liu, N.; Tan, W.; Zhang, X.; Zhang, L. Facile synthesis of NiCo2O4@Polyaniline core–shell nanocomposite for sensitive determination of glucose. Biosens. Bioelectron. 2016, 75, 161–165. [Google Scholar] [CrossRef]
  38. Hassanpoor, S.; Aghely, F. Hierarchically self-assembled NiCo2O4 nanopins as a high-performance supercapacitor cathodic material: A morphology controlled study. RSC Adv. 2020, 10, 35235–35244. [Google Scholar] [CrossRef]
  39. Zhan, J.; Cai, M.; Zhang, C.; Wang, C. Synthesis of mesoporous NiCo2O4 fibers and their electrocatalytic activity on direct oxidation of ethanol in alkaline media. Electrochim. Acta. 2015, 154, 70–76. [Google Scholar] [CrossRef]
  40. Yu, H.; Jin, J.; Jian, X.; Wang, Y.; Qi, G.C. Preparation of cobalt oxide nanoclusters/overoxidized polypyrrole composite film modified electrode and its application in nonenzymatic glucose sensing. Electroanalysis 2013, 25, 1665–1674. [Google Scholar] [CrossRef]
  41. Pasta, M.; La Mantia, F.; Cui, Y. Mechanism of glucose electrochemical oxidation on gold surface. Electrochim. Acta. 2010, 55, 5561–5568. [Google Scholar] [CrossRef]
  42. Saraf, M.; Natarajan, K.; Mobin, S.M. Multifunctional porous NiCo2O4 nanorods: Sensitive enzymeless glucose detection and supercapacitor properties with impedance spectroscopic investigations. New J. Chem. 2017, 41, 9299–9313. [Google Scholar] [CrossRef]
  43. Amin, B.G.; Masud, J.; Nath, M. A non-enzymatic glucose sensor based on a CoNi2Se4/rGO nanocomposite with ultrahigh sensitivity at low working potential. J. Mater. Chem. B. 2019, 7, 2338–2348. [Google Scholar] [CrossRef]
  44. Li, H.; Zhang, L.; Mao, Y.; Wen, C.; Zhao, P. A simple electrochemical route to access amorphous Co-Ni hydroxide for non-enzymatic glucose sensing. Nanoscale Res. Lett. 2019, 14, 1–12. [Google Scholar] [CrossRef] [Green Version]
  45. Cui, S.; Zhang, J.; Ding, Y.; Gu, S.; Hu, P.; Hu, Z. Rectangular flake-like mesoporous NiCo2O4 as enzyme mimic for glucose biosensing and biofuel cell. Sci. China Mater. 2017, 60, 766–776. [Google Scholar] [CrossRef] [Green Version]
  46. Huang, W.; Cao, Y.; Chen, Y.; Peng, J.; Lai, X.; Tu, J. Fast synthesis of porous NiCo2O4 hollow nanospheres for a high-sensitivity non-enzymatic glucose sensor. Appl. Surf. Sci. 2017, 396, 804–811. [Google Scholar] [CrossRef]
  47. Wang, C.; Du, J.; Wang, H.; Zou, C.E.; Jiang, F.; Yang, P.; Du, Y. A facile electrochemical sensor based on reduced graphene oxide and Au nanoplates modified glassy carbon electrode for simultaneous detection of ascorbic acid, dopamine and uric acid. Sens. Actuators B Chem. 2014, 204, 302–309. [Google Scholar] [CrossRef]
  48. Jiménez-Fiérrez, F.; González-Sánchez, M.I.; Jiménez-Pérez, R.; Iniesta, J.; Valero, E. Glucose Biosensor Based on Disposable Activated Carbon Electrodes Modified with Platinum Nanoparticles Electrodeposited on Poly (Azure A). Sensors. 2020, 20, 4489. [Google Scholar] [CrossRef]
  49. Gong, X.; Gu, Y.; Zhang, F.; Liu, Z.; Li, Y.; Chen, G.; Wang, B. High-performance non-enzymatic glucose sensors based on CoNiCu alloy nanotubes arrays prepared by electrodeposition. Front. Mater. 2019, 6, 3. [Google Scholar] [CrossRef] [Green Version]
  50. Mondal, S.; Madhuri, R.; Sharma, P.K. Probing the shape-specific electrochemical properties of cobalt oxide nanostructures for their application as selective and sensitive non-enzymatic glucose sensors. J. Mater. Chem. C. 2017, 5, 6497–6505. [Google Scholar] [CrossRef]
  51. Ding, Y.; Wang, Y.; Su, L.; Bellagamba, M.; Zhang, H.; Lei, Y. Electrospun Co3O4 nanofibers for sensitive and selective glucose detection. Biosens. Bioelectron. 2010, 26, 542–548. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, L.; Lu, X.; Ye, Y.; Sun, L.; Song, Y. Nickel-cobalt nanostructures coated reduced graphene oxide nanocomposite electrode for nonenzymatic glucose biosensing. Electrochim. Acta. 2016, 114, 484–493. [Google Scholar] [CrossRef]
  53. Ma, G.; Yang, M.; Li, C.; Tan, H.; Deng, L.; Xie, S.; Song, Y. Preparation of spinel nickel-cobalt oxide nanowrinkles/reduced graphene oxide hybrid for nonenzymatic glucose detection at physiological level. Electrochim. Acta. 2016, 220, 545–553. [Google Scholar] [CrossRef]
  54. Liu, Y.; Liu, X.; Guo, Z.; Hu, Z.; Xue, Z.; Lu, X. Horseradish peroxidase supported on porous graphene as a novel sensing platform for detection of hydrogen peroxide in living cells sensitively. Biosens. Bioelectron. 2017, 87, 101–107. [Google Scholar] [CrossRef]
  55. Zhang, C.; Zhang, Y.; Du, X.; Chen, Y.; Dong, W.; Han, B.; Chen, Q. Facile fabrication of Pt-Ag bimetallic nanoparticles decorated reduced graphene oxide for highly sensitive non-enzymatic hydrogen peroxide sensing. Talanta 2016, 159, 280–286. [Google Scholar] [CrossRef]
  56. Kannan, P.K.; Hu, C.; Morgan, H.; Rout, C.S. One-Step Electrodeposition of NiCo2S4 Nanosheets on Patterned Platinum Electrodes for Non-Enzymatic Glucose Sensing. Chem. Asian. J. 2016, 11, 1837–1841. [Google Scholar] [CrossRef] [Green Version]
  57. Liu, S.; Hui, K.S.; Hui, K.N. Flower-like copper cobaltite nanosheets on graphite paper as high-performance supercapacitor electrodes and enzymeless glucose sensors. ACS Appl. Mater. Interfaces. 2016, 8, 3258–3267. [Google Scholar] [CrossRef]
  58. Ding, Y.; Wang, Y.; Zhang, L.; Zhang, H.; Lei, Y. Preparation, characterization and application of novel conductive NiO–CdO nanofibers with dislocation feature. J. Mater. Chem. 2012, 22, 980–986. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) NCOBs (8B, 11B, 12B, 13B, and 14B), (b) NCO8 by different annealing temperatures (150, 250, and 350 °C), (c) NCO13 by different annealing temperatures (150, 250, and 350 °C), and (d) NCOs (8, 11, 12, 13, and 14) after annealing at 450 °C.
Figure 1. XRD patterns of (a) NCOBs (8B, 11B, 12B, 13B, and 14B), (b) NCO8 by different annealing temperatures (150, 250, and 350 °C), (c) NCO13 by different annealing temperatures (150, 250, and 350 °C), and (d) NCOs (8, 11, 12, 13, and 14) after annealing at 450 °C.
Nanomaterials 11 00055 g001
Figure 2. FE-SEM images of (a) NCOBs (8B, 11B, 12B, 13B, and 14B), and (b) NCOs (8, 11, 12, 13, and 14).
Figure 2. FE-SEM images of (a) NCOBs (8B, 11B, 12B, 13B, and 14B), and (b) NCOs (8, 11, 12, 13, and 14).
Nanomaterials 11 00055 g002
Figure 3. CV curves of (a) NCO8, (b) NCO11, (c) NCO12, (d) NCO13, and (e) NCO14 electrodes at different scan rates in 0.1 M NaOH solution. (f) Respective Randles–Sevcik plots of NCO electrodes.
Figure 3. CV curves of (a) NCO8, (b) NCO11, (c) NCO12, (d) NCO13, and (e) NCO14 electrodes at different scan rates in 0.1 M NaOH solution. (f) Respective Randles–Sevcik plots of NCO electrodes.
Nanomaterials 11 00055 g003
Figure 4. CA responses of (a) NCO8, (b) NCO11, (c) NCO12, (d) NCO13, and (e) NCO14 electrodes with the addition of glucose to 0.1 M NaOH solution at 0.50 V. Their respective calibration curves of current response versus glucose concentration plots inset in the figure.
Figure 4. CA responses of (a) NCO8, (b) NCO11, (c) NCO12, (d) NCO13, and (e) NCO14 electrodes with the addition of glucose to 0.1 M NaOH solution at 0.50 V. Their respective calibration curves of current response versus glucose concentration plots inset in the figure.
Nanomaterials 11 00055 g004
Figure 5. CA responses (CA) of (a) NCO8, (b) NCO11, (c) NCO12, (d) NCO13, and (e) NCO14 electrodes to addition of 1 mM glucose and 0.1 mM interfering species (LA, DA, AA, and UA) in a 0.1 M NaOH solution at 0.5 V.
Figure 5. CA responses (CA) of (a) NCO8, (b) NCO11, (c) NCO12, (d) NCO13, and (e) NCO14 electrodes to addition of 1 mM glucose and 0.1 mM interfering species (LA, DA, AA, and UA) in a 0.1 M NaOH solution at 0.5 V.
Nanomaterials 11 00055 g005
Table 1. Comparison of our work to the nonenzymatic glucose sensor.
Table 1. Comparison of our work to the nonenzymatic glucose sensor.
ElectrodeMethodSensitivity
(μA/mM (cm2))
Linear Range
(mM)
Correlation
Coefficient (R2)
Detection Limit
(μM)
Refs.
NCO8CA48.710.01–60.9920.539This work
NCO11CA90.5490.01–60.9950.393This work
NCO12CA94.900.01–60.9980.0503This work
NCO13CA142.140.01–60.9970.0433This work
NCO14CA73.100.01–60.9960.0475This work
NiCo2O4/CNTCA66.150.02–12.120.995[52]
NiCo2O4/rGOCA548.90.005–8.560.992[53]
CuCo2O4CA3.625Up to 0.32-5[56]
NiOCA32.91Up to 1.94 1.28[57]
Co3O4CA36.25Up to 2.04 0.97[58]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jang, K.-b.; Park, K.R.; Kim, K.M.; Hyun, S.-k.; Jeon, J.-e.; Song, Y.S.; Park, S.-k.; Moon, K.-i.; Ahn, C.; Lim, S.-c.; et al. Synthesis of NiCo2O4 Nanostructures and Their Electrochemial Properties for Glucose Detection. Nanomaterials 2021, 11, 55. https://doi.org/10.3390/nano11010055

AMA Style

Jang K-b, Park KR, Kim KM, Hyun S-k, Jeon J-e, Song YS, Park S-k, Moon K-i, Ahn C, Lim S-c, et al. Synthesis of NiCo2O4 Nanostructures and Their Electrochemial Properties for Glucose Detection. Nanomaterials. 2021; 11(1):55. https://doi.org/10.3390/nano11010055

Chicago/Turabian Style

Jang, Kyu-bong, Kyoung Ryeol Park, Kang Min Kim, Soong-keun Hyun, Jae-eun Jeon, Young Sik Song, Soo-keun Park, Kyoung-il Moon, Chisung Ahn, Sung-chul Lim, and et al. 2021. "Synthesis of NiCo2O4 Nanostructures and Their Electrochemial Properties for Glucose Detection" Nanomaterials 11, no. 1: 55. https://doi.org/10.3390/nano11010055

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