Open Access Article
R.
Manigandan
a,
T.
Dhanasekaran
a,
A.
Padmanaban
a,
K.
Giribabu
b,
R.
Suresh
c and
V.
Narayanan
*a
aDepartment of Inorganic Chemistry, University of Madras, Guindy Campus, Chennai, India. E-mail: vnnara@yahoo.co.in
bElectrodics and Electrocatalysis Division, CSIR-CECRI, Karaikudi, India
cDepartment of Analytical and Inorganic Chemistry, University of Concepcion, Chile
First published on 31st January 2019
Ni0/NiO (nickel/nickel oxide) core–shell nanostructures were synthesized through a facile combustible redox reaction. Remarkably, the hetero-phase boundary with different crystalline orientations offered dual properties, which helped in bifunctional catalysis. Presence of a metallic Ni phase changed physicochemical properties and some emerging applications (magnetic properties, optical conductivity, electrochemical sensitivity, catalytic behaviour) could be foreseen. Moreover, formation of a NiO layer on metal surface prevented magnetism-induced aggregation, arrested further oxidation by hindering oxygen diffusion, and acted as a good sorbent to enhance the surface adsorption of the analyte. Hexagonal Ni/NiO nanostructures manifested well-defined ferromagnetic behavior and the catalyst could be collected easily at the end of the catalytic reduction. Ni/NiO core–shell catalysts at the nanoscale had outstanding catalytic performance (reduction of 4-nitrophenol to 4-aminophenol) compared with pure NiO catalysts beyond a reaction time of ∼9 min. The estimated sensitivity, limit of detection and limit of quantification towards the electrochemical sensing of serotonin were 0.185, 0.43 and 1.47 μM μA−1, respectively. These results suggest that a bifunctional Ni/NiO nanostructure could be a suitable catalyst for electrochemical detection of serotonin and reduction of 4-nitrophenol.
Key challenges in the synthesis of core–shell materials are scalability and multi-step processes. Modification of a core material with a shell by problematic complex procedures or addition of a reactive material can affect the phase boundary of core structures. It is necessary to have a general method to prepare nanoparticles with controlled size to attain unique properties. Core/shell nanostructured materials have been prepared by several methods: template-mediated synthesis, low-pressure spray pyrolysis and polymer-matrix assisted synthesis.17,18 Few of the methods stated above are suitable for large-scale synthesis; they suffer from size-heterogeneity and expense. To overcome these issues, we chose a rapid, single-step procedure for large-scale material synthesis with facile methodology. Hence, a combustible redox reaction would be preferred to obtain highly crystalline material, and rapid accomplishment could arrest the post-oxidation of metallic nickel.19
Addition of 1–10% of a non-ionic kosmotrope (e.g., glycerol, propylene glycol and ethylene glycol (EG)) in an aqueous system makes it highly viscous due to the formation of strong water–water hydrogen bonds, which affect the size of nanoparticles without the assistance of surfactants or templates.20,21 Herein, we used ethylene glycol (EG) as a multipurpose solvent to attain a high boiling point and high viscous nature and maintained a reducing atmosphere to control nucleation.
Most nitro-aromatic compounds have been found to be environmentally poisonous materials. 4-Nitrophenol (4-NP) is one of the most refractory pollutants often present in industrial effluents, but is important in the preparation of several analgesic and antipyretic drugs such as paracetamol and phenacetin.22–24 In addition, 5-hydroxytryptamine (5-HT, serotonin) is a monoamine neurotransmitter involved in a large number of physiological processes in the digestive system, mood, sleep, emesis, sexuality, and appetite. Changes in serotonin levels within the digestive system can cause neurological disorders and several diseases.25 Therefore, detection of the 5-HT level is a crucial requirement for understanding the mechanisms of gastrointestinal functions and monitoring pathophysiological conditions. Hence, it is worthy to conduct the reduction of 4-NP to 4-aminophenol and electrochemical detection of 5-HT using an earth-abundant catalyst. The initial focus of this work was to understand materials properties through optimizing nanostructured Ni/NiO for potential applications (electrochemical detection of 5-HT and catalytic reduction of 4-AP). We tried to gain better understanding of the effect of the metallic-nickel phase on physicochemical properties (structural, magnetic and optical) through metal/metal oxide hybrid nanostructure approaches. The relatively simple method of preparation makes them an attractive class of hybrid catalysts.
We utilized the auto combustible redox reaction of Ni/NiO to further improve magnetic and catalytic properties. Briefly, a detailed investigation and discussion on structural analysis, elemental analysis, electronic transition, surface morphology, and magnetic properties of hexagonal magnetic Ni/NiO nanostructures were reported. Additionally, to study the bifunctional catalytic ability of magnetic Ni/NiO, we undertook electrochemical sensing of 5-HT and catalytic reduction of 4-NP.
:
1 ratio. Excessive addition of a hydrazine
:
ammonium carbonate (hydrazinocarbonic acid) dissolved the intermediate precursor to a dark-blue solution. This suspension was stirred for 20 min at 60 °C with a magnetic stirrer. Furthermore, the reaction mixture was transferred directly to a silica boat and then evaporated on a hot plate for a combustible redox reaction. The temperature was increased until the sample mixture was ignited. Once complete combustion was over, samples were removed immediately from the hot plate to arrest complete oxidation. This reaction mixture is too vigorous and explosive when we used nickel nitrate instead of nickel acetate. The resulting black spongy powder (as-combusted or Ni–C) was heated in a muffle furnace at 300 °C, 400 °C, and 500 °C for 10 min to improve its crystalline nature.
m (225) space group. The diffraction peaks appearing at 2θ = 44.4°, and 51.5° were indexed as the (111), and (200) planes of the cubic phase (a = 3.535 Å) of metallic Ni (JCPDS no. 065-0380), with an Fm
m (225) space group. Moreover, the XRD characteristic peaks were broader with high intensity, which indicated a small crystallite size. Upon calcination, the intensity of metallic peaks at 45° and 52° started to vanish, which could have been due to oxidation of Ni0 to its oxide material by higher thermal energy. However, little signal intensity was observed up to 400 °C, in accordance with the diffraction of the metallic phase. For (Ni-500), after calcination at 500 °C, no diffraction peaks correspond to the metallic phase. Diffraction bands disappeared completely and NiO characteristic peaks were narrow with high intensity. Hence, complete oxidation of the metallic phase to the pure NiO phase as evidenced by increasing crystallite size and disappearance of Ni0 bands, and in good agreement with the literature,27 was observed. The average crystallite size of Ni and NiO were calculated by the deconvolution of highly intense diffraction peaks, based on the Scherrer formula:D = 0.9λ/β cos θ | (1) |
| Name | Phase | XRD | XPS | DRS | VSM | |||||
|---|---|---|---|---|---|---|---|---|---|---|
| 2θ (°) | d (hkl) (Å) | FWHM (°) | Crystallite D (nm) | ΔEB (eV) | E g (eV) | M s (emu g−1) | M r (emu g−1) | H c (Oe) | ||
| Ni–C | Ni0 | 44.56 | 2.03(111) | 0.478 | ≈18.77 | 18.2 | 2.62 | ≈16.87 | ≈2.418 | ≈115 |
| NiO | 43.27 | 2.09(200) | 1.284 | ≈6.95 | ||||||
| Ni-300 | Ni0 | 44.45 | 2.03(111) | 0.410 | ≈21.87 | — | 2.71 | ≈12.86 | ≈2.620 | ≈141 |
| NiO | 43.26 | 2.09(200) | 2.118 | ≈4.22 | ||||||
| Ni-400 | Ni0 | 44.44 | 2.03(111) | 0.550 | ≈16.30 | — | 3.08 | ≈00.91 | ≈0.212 | ≈152 |
| NiO | 43.20 | 2.09(200) | 0.926 | ≈9.64 | ||||||
| Ni-500 | NiO | 43.28 | 2.09(200) | 0.448 | ≈19.93 | 17.9 | 3.23 | ≈00.02 | ≈0.002 | ≈72 |
Fig. 2A and B show the XPS survey, Ni (2p), and C (1s) core level spectrum of Ni/NiO hybrid materials calcined at Ni-300 °C, and Ni-500 °C, respectively. The binding energies (EB) of obtained spectra were charge-corrected by referencing the C 1s peak to 284.6 eV. The deconvoluted C 1s spectrum in the range 281–290 eV is shown in Fig. 2i. The Ni (2p) core-level (Fig. 2ii) emission was split into two significant peaks at EB of 857.6 and 875.3 eV with two satellite peaks corresponding to the spin–orbit components of 2p3/2 and 2p1/2, respectively. The spin–orbit separation ΔEB = 17.9 (NiO) & ΔEB = 18.2 eV (Ni/NiO) and the peak 2p3/2 looked stronger than 2p1/2. The spin–orbit sublevels 2p3/2 and 2p1/2 were superimposed multiplet splitting, which was due purely to the electrostatic interaction between 2p and 3d shells. Apart from that, an additional small shoulder peak appeared at lower energy than Ni2+ at 853.9 eV that could be assigned to metallic Ni in the Ni/NiO hybrid structure (Fig. 2B), which was not observed in NiO-500 sample (Fig. 2A) and which was in good agreement with the literature.17,28 The O (1s) core-level spectra of binding energy vs. intensity are shown in Fig. S1.† The two peaks with binding energies of 528–535.7 eV were due to a hybrid metal oxide (Ni–O) and surface hydroxyl group, respectively. The presence of a surface hydroxyl group was due to the hygroscopic character of the metal oxide.29,30 The obtained XPS elemental analyses provided good information about the presence of Ni0, and Ni2+ species in Ni/NiO, and were in accordance with the results discussed above (Fig. 1).
![]() | ||
| Fig. 2 XPS survey, C (1s), Ni (2p) core-level spectrum of Ni/NiO hybrid materials obtained at (A) Ni-300, and (B) Ni-500. | ||
![]() | ||
| Fig. 3 (a) FESEM image, (b, d & e) HRTEM images and (c) SAED pattern of a Ni/NiO nanocomposite (Ni–C). | ||
:
N2H4 mixture along with ammonium carbonate acted as a redox agent to create an in situ mildly basic condition and inert atmosphere upon combustion. According to experimental results, the possible formation mechanism of the Ni/NiO chain-like core–shell nanoarchitecture could be explained in two main steps. First, the self-assembly of EG
:
N2H4 results in micelle formation, which reduces flocculation and stimulates the formation of a well-defined structure by creating a steric atmosphere on the surface of nuclei due to the inherent van der Waals attraction of the surfactant.31 Second, the combustible intermediates (N, O–centred bidentate hydrazidocarbonic anion derivatives (N2H3COO−)) could be formed by a metathesis reaction between hydrazine hydrate and ammonium carbonate (N2H4
:
(NH4)2CO3 mixture), which would result in nickel hydrazinocarboxylate salts (Ni(N2H3COO)2/Ni(N2H3COO)2(N2H4)2) by coordination with nickel cations.32 Herein, the EG
:
N2H4
:
(NH4)2CO3 mixture acted as an agent as well as a fuel. Briefly, addition of hydrazine nucleated metallic Ni0 partially along with combustible nickel hydrazinocarboxylate (Ni(N2H3COO)2/(N2H4)2Ni(N2H3COO)2) on the surface of Ni0 grains due to the formation of the N2H3COO− intermediate, which also controlled the size of particles by arresting further agglomeration. Formation of an N–N-based highly energetic precursor (nickel hydrazinocarboxylate) was evidenced by characteristic Fourier transform infrared (FTIR) peaks (Fig. S2†) and the thermal behavior of the precursor was also evaluated by thermal gravimetry-differential thermal gravimetry (TG-DTG) (Fig. S3†), which were in good agreement with the literature.33 To achieve a crystalline NiO shell, the resulting precursor was combusted at above ≈250 °C. At a calcination temperature of 500 °C, the core metallic nickel was oxidized to pure NiO. Furthermore, at this stage of crystal growth (Fig. 7), formation of a porous nature would be rational due to the removal of N2, NO2 and a large amount of CO2 in the presence of atmospheric oxygen. Moreover, the metallic nickel was covered by porous nickel oxide, which allowed contact with reactants on core and shell surfaces, and the nickel oxide layer arrested further oxidation of Ni0. Based on the results stated above, a solution-solid-solid growth mechanism was proposed.
![]() | ||
| Fig. 5 (A) UV-VIS DRS absorption spectra, (B) Tauc plot of Ni/NiO (a) Ni–C, (b) Ni-300, (c) Ni-400, and (d) Ni-500. Inset: plot of estimated band gap against temperature. | ||
The band gap can be calculated from eqn (2):
| (αhν) = A(hν − Eg)2 | (2) |
Band gaps deduced from a Tauc plot of (αhν)2vs. hν are shown in Fig. 5B. The extrapolated value of hν up to zero absorption elicits the valence-to-conduction-band direct-allowed transition or optical band gap energy (Eg) of various materials. It is well known that a semiconductor with nanoscale size shows a blue shift in its spectrum due to quantum confinement effects.34 The synthesized Ni/NiO resulted in a lower Eg value than that of pure NiO. This effect was likely due to the chemical defects present in the inter-granular regions, which generate new energy levels to decrease the distance between the valence to conduction band.
000 Oe to 15
000 Oe. The magnetization versus magnetic field plots (measurements of M–H magnetic hysteresis loops) are shown in Fig. 6 carried out at room temperature. The resulting graph provided evidence for ferromagnetic behavior, which could be concluded based on the obtained coercivity (Hc) and retentivity of the hysteresis loops. Ni/NiO hybrid nanostructures manifested well-defined ferromagnetic behavior at room temperature and had coercivity, which was due to the higher thermal energy than the anisotropy energy of the nanoparticles (Table 1). The saturation magnetization (Ms), remanent magnetization (Mr), and Hc were determined to be 16.17 emu g−1, 2.41 emu g−1, and 115 Oe, respectively, for Ni/NiO as a combusted sample. The Ms of Ni/NiO compared with that reported for pure Ni0 nanoparticles exhibited a large difference, which could be attributed to weak magnetic interfaces (nonmagnetic residual carbon and antiferromagnetic nickel oxide layer covered on the surface of metallic Ni nanoparticles). Accordingly, the magnetization of nanoparticles is usually smaller than that of the corresponding bulk materials.35
![]() | ||
| Fig. 6 Plot of magnetization versus magnetic field for Ni/NiO hybrid nanostructures inset: magnified view. | ||
The reduction reactions followed pseudo first-order kinetics with respect to 4-NP, which was defensible from a linear correlation between ln(At/A0) with time. Catalytic performances depended strongly on the morphology and environment of the catalyst. The reaction proceeded rapidly, with conversion completer at a reaction time of ∼9 min (Fig. 8). Moreover, Ni/NiO core-shell-structured catalysts with nanoscale facilitated the outstanding catalytic performance compared with pure NiO catalysts. Finally, the fact that the reduction efficiency decreased with the increasing calcination temperature of catalysts could have been due to the lattice defect, phase transformation and increasing particle size. Those defects in Ni/NiO lattices lead to changes in the band gap (electronic effect), loss of synergetic behaviour (adsorption of substrates) and decreased surface-to-volume ratio.
Fig. 9b depicts the electrochemical oxidation of 5-HT on the surface of bare GCE (9a) and Ni/NiO modified GCE (Fig. 9b–d) as the working electrode (GCE/Ni/NiO/5-HT) in PBS containing 5 × 10−5 M 5-HT. Bare and modified electrodes showed an anodic peak with slight potential differences and more current variations in the potential region of 0.27–0.33 V. Ni/NiO-modified GCE toward 5-HT showed higher electrocatalytic activity than that of bare GCE. The Ni/NiO phase shifted the oxidation potential to less positive potential with greater electrocatalytic activity compared with that of pure NiO. The surface pH effect on the modified electrodes could be the reason for the less positive potential shift (a synergetic effect between zero-valent Ni sites and acid–base sites in Ni–Ni–O catalysts improves the catalytic performance by shifting the potential towards a less positive region27) because pH altered the anodic oxidation peak. Moreover, Ni/NiO samples showed an increase in conductivity due to the electronic effect between the core and shell. Electron transfer within phases altered the electronegativity and it had two electronic environments in each grain boundary, which enhanced the current density.
The effect of scan rate on the electrochemical response of 5 × 10−5 M 5-HT at the surface of GCE/Ni/NiO was studied by cyclic voltammetry to investigate electrode reaction kinetics (Fig. 10a). When the scan rate was increased, the oxidation peak current moved positively and the reduction peak current moved negatively. The inset of Fig. 10a shows a plot of peak current (Ip) vs. square root of scan rate (ν1/2). Fig. 10b shows that the oxidation peak currents were linearly proportional to the scan rate from 50 to 500 mV s−1. The linear regression equation for the anodic process was Ia = 0.518ν0.5 − 1.608 (R2 = 0.997). The number of electrons involved in the overall reaction could be obtained by eqn (3) (n = number of electrons, Ip = peak current, R = gas constant, T = 298 K, F = Faraday constant, Q = charge, ν = scan rate). The total number of electrons involved in the redox process of 5-HT was ∼2, which is consistent with the literature. These data suggest that the electrode reaction corresponded to a diffusion-controlled process.40–42
| n = 4IRT/FQν | (3) |
![]() | ||
| Fig. 10 (a and b) Cyclic voltammograms of GCE/Ni/NiO/5-HT at different scan rates in PBS containing 5 × 10−5 M 5-HT. (cand d) Differential pulse voltammograms of 5-HT at Ni/NiO/GCE. | ||
The DPV response of 5-HT at Ni/NiO/GCE is displayed in Fig. 10c. The oxidation peak current of 5-HT increased gradually at 0.25 × 10−6 to 3 × 10−6 M. The calibration graph for 5-HT determination by Ni/NiO/GCE is shown as the inset in Fig. 10d. The linear regression equation was found to be Ia (μA) = 0.185[5-HT] (μM) + 1.234 (R2 = 0.991). This corresponded to a sensitivity of 0.185 μM μA−1, showing that Ni/NiO/GCE was highly sensitive towards 5-HT. The small drop in current upon increasing the concentration found in the linear plot might have been due to the kinetic limitation of the electrode. The calculated limit of detection (3σ/s) was 0.437 μM μA−1 and limit of quantification (10σ/s) was 1.470 μM μA−1 (S/N = 3). The obtained results of Ni/NiO/GCE/5-HT were comparable with those of other electrochemical sensors. Furthermore, the stability and reproducibility of Ni/NiO/GCE/5-HT was investigated by measuring electrode sensitivity after 30 day storage in air under ambient conditions: repetitive measurements revealed acceptable catalytic-current signal responses. These data suggested that Ni/NiO/GCE (Ni–C) possessed high sensitivity, stability and reproducibility towards 5-HT oxidation.
We believe that our method and choice of material has several advantages: simple reaction conditions, large-scale synthesis, cost-effective earth-abundant material, and easy recoverable magnetism. The ferromagnetic Ni/NiO prepared by our method showed high sensitivity towards 5-HT and better catalytic reduction toward 4-NP. These findings are important for enabling catalytic studies and providing a sensor platform. Our method can be utilized to fabricate storage devices due to their outstanding optoelectronic properties.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c8na00342d |
| This journal is © The Royal Society of Chemistry 2019 |