Next Article in Journal
Imbalanced Fault Diagnosis of Rotating Machinery Based on Deep Generative Adversarial Networks with Gradient Penalty
Next Article in Special Issue
Effect of Mn and Cu Substitution on the SrFeO3 Perovskite for Potential Thermochemical Energy Storage Applications
Previous Article in Journal
Interactive Mechanism of Potential Inhibitors with Glycosyl for SARS-CoV-2 by Molecular Dynamics Simulation
Previous Article in Special Issue
Intelligent and Data-Driven Fault Detection of Photovoltaic Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Performance of Aluminum Doped Ni1−xAlxCo2O4 Hierarchical Nanostructure: Experimental and Theoretical Study

1
Department of Physics and Materials Science, The University of Memphis, Memphis, TN 38152, USA
2
Department of Chemistry, Pittsburg State University, Pittsburg, KS 66762, USA
*
Author to whom correspondence should be addressed.
Submission received: 3 August 2021 / Revised: 13 September 2021 / Accepted: 22 September 2021 / Published: 30 September 2021
(This article belongs to the Special Issue Redesign Processes in the Age of the Fourth Industrial Revolution)

Abstract

:
For electrochemical supercapacitors, nickel cobaltite (NiCo2O4) has emerged as a new energy storage material. The electrocapacitive performance of metal oxides is significantly influenced by their morphology and electrical characteristics. The synthesis route can modulate the morphological structure, while their energy band gaps and defects can vary the electrical properties. In addition to modifying the energy band gap, doping can improve crystal stability and refine grain size, providing much-needed surface area for high specific capacitance. This study evaluates the electrochemical performance of aluminum-doped Ni1−xAlxCo2O4 (0 ≤ x ≤ 0.8) compounds. The Ni1−xAlxCo2O4 samples were synthesized through a hydrothermal method by varying the Al to Ni molar ratio. The physical, morphological, and electrochemical properties of Ni1−xAlxCo2O4 are observed to vary with Al3+ content. A morphological change from urchin-like spheres to nanoplate-like structures with a concomitant increase in the surface area, reaching up to 189 m2/g for x = 0.8, was observed with increasing Al3+ content in Ni1−xAlxCo2O4. The electrochemical performance of Ni1−xAlxCo2O4 as an electrode was assessed in a 3M KOH solution. The high specific capacitance of 512 F/g at a 2 mV/s scan rate, 268 F/g at a current density of 0.5 A/g, and energy density of 12.4 Wh/kg was observed for the x = 0.0 sample, which was reduced upon further Al3+ substitution. The as-synthesized Ni1−xAlxCo2O4 electrode exhibited a maximum energy density of 12.4 W h kg−1 with an outstanding high-power density of approximately 6316.6 W h kg−1 for x = 0.0 and an energy density of 8.7 W h kg−1 with an outstanding high-power density of approximately 6670.9 W h kg−1 for x = 0.6. The capacitance retention of 97% and 108.52% and the Coulombic efficiency of 100% and 99.24% were observed for x = 0.0 and x = 0.8, respectively. First-principles density functional theory (DFT) calculations show that the band-gap energy of Ni1−xAlxCo2O4 remained largely invariant with the Al3+ substitution for low Al3+ content. Although the capacitance performance is reduced upon Al3+ doping, overall, the Al3+ doped Ni1−xAlxCo2O4 displayed good energy, powder density, and retention performance. Thus, Al3+ could be a cost-effective alternative in replacing Ni with the performance trade off.

1. Introduction

Researchers are actively engineering an efficient, clean, and cost-effective electrode for better electrochemical performance [1]. Conventional capacitors and batteries are broadly exploited by energy storage devices [2]. Previously, secondary batteries were suitable electrical energy storage devices, but they were not very satisfactory because of their smaller power density, presence of toxic waste substances, and shorter lifetime [3]. On the other hand, conventional capacitors display high power density compared to a battery, but because of their lower energy density, researchers are interested in advanced energy storing devices that mitigate the demerits of both battery and capacitor. The energy gap between batteries and capacitors can be bridged by a supercapacitor (SC) [4]. SC electrode materials must have a large specific surface area, controlled porosity, high electronic conductivity, suitable electroactive sites, high thermal stability, low raw material costs, and ease of manufacture [5]. Supercapacitors have recently attracted attention due to their quick charging/discharging, high power density, wide range of working temperatures, high efficiency, and extended lifespans [6]. SCs are of three types: (i) electrical double layer capacitance (EDLCs) examples are highly porous carbon materials, (ii) pseudocapacitor (PCs) examples are transition metal oxides and conducting polymers, and (iii) hybrid capacitor examples are battery type hybrid devices such as the lithium-ion capacitor [7]. EDLCs store electrical energy in the electric double layer, which occurs at the interface between the electrode’s active material and the electrolyte by an electrostatic accumulation process. In EDLCs, energy density becomes lower (between 3 and 6 Wh/Kg), and specific capacitance is restricted because of the specific surface area of the active electrode material [8]. The ionic double layer’s fast diffusion provides a very rapid discharge profile and hence high-power densities. PCs, on the other hand, are caused by reversible Faradaic processes at the electrode surface, which cause charge storage [9]. Moreover, in PCs, the electrochemical process occurs near the surface of the active electrode material. Thus, it can give specific capacitance 10–100 times greater than that provided by EDLCs [10]. Moreover, hybrid capacitors are a combination of both EDLCs and PCs. Usually, hybrid capacitor electrodes are made from (1) both EDLCs materials and PCs materials; symmetric geometry (2) with one EDLCs electrode and other PCs or battery type electrode; and asymmetric geometry (3) with an asymmetric structure with one PC electrode and another rechargeable battery type electrode [7].
In a pseudocapacitor, a redox reaction takes place between the active electrode material and the electrolyte. Various kinds of transition metal oxides having higher theoretical specific capacitance, such as RuO2 [11], MnO2 [12], NiO [13], Ni(OH)2 [14], CuO [15], Co3O4 [16], and Co(OH)2 [17], have been investigated as an active electrode material. However, their poor electronic conductivity [18,19], lower cyclic performance [20,21], and substantial volume change during the charge/discharge process [22] demand the development of nanostructured materials to meet the increasing and urgent demands for energy storing devices. Recently, mixed-transition-metal-oxides (MTMOs), AxB3−xO4 materials have emerged with promising applications for electrodes. The MTMOs with A and B metal ions can provide many oxidation states for efficient redox charge reactions [23]. Nickel cobaltite (NiCo2O4) is one of the most valuable electrode materials studied because of its high theoretical specific capacitance, lower cost, low toxic nature, good redox activity, higher electrical conductivity, and environmentally benign nature [24]. NiCo2O4 offers high specific capacitance due to the presence of Ni and Co ions together, which are responsible for producing higher electrochemical activity and electrical conductivity [25]. NiCo2O4 is a spinel-type, AB2O4, binary metal oxide where nickel cations occupy octahedral sites and cobalt cations are randomly distributed on both tetrahedral and octahedral sites [26].
Doped NiCo2O4 with various transition metals, e.g., Ni1−xRuxCo2O4 [27], Ni1−xMnxCo2O4 [28], Ni1−xCrxCo2O4 [29], Ni1−xZnxCo2O4 [30], Ni1−xFexCo2O4 [31], and Ni1−xCaxCo2O4 [32], have been studied to understand the effect on morphology and their electrochemical performance. This paper describes a simple hydrothermal method for making Al3+ doped NiCo2O4, i.e., Ni1−xAlxCo2O4. The study systematically assesses the effect of Al3+ doping on the crystal structure, morphology, chemical composition, and electrochemical performance of Ni1−xAlxCo2O4. Aluminum ion substitution is favored because of its inexpensive cost, large abundance, non-toxic nature, and superior corrosion resistance. First-principles density functional theory (DFT) calculations were employed to assess the energy band-gap of the doped Ni1−xAlxCo2O4.

2. Experimental

2.1. Synthesis

The required high purity nitrate salts, viz. Al(NO3)2.6H2O, Co(NO3)2.6H2O, and Ni(NO3)2.6H2O, were purchased from Sigma Aldrich (St. Louis, MO, USA). Table 1 lists the required amount of precursor salts needed to prepare Ni1−xAlxCo2O4. The nitrate salts were dissolved in 35 mL of water with a fixed amount of Urea (CO(NH2)2) and stirred for 30 min. The solution was then transferred to a 45 mL Teflon-lined autoclave and maintained at 190 °C for 12 h, followed by a natural cooling to room temperature. The precipitates were washed with distilled water and acetone several times by centrifugation and then dried for 12 h at 80 °C. After this, the as derived metal hydroxide composite particles were calcined at 350 °C in the air for 3.5 h.

2.2. Characterization

X-ray diffraction (XRD) was used to examine the crystalline structure of calcined powder using a Bruker D8 Advance X-ray diffractometer (Bruker, Billerica, MA, USA). The surface morphology of prepared samples was studied by a scanning electron microscope (SEM) (at 10 keV using Phenom, Oak Park, CA, USA). Autosorb (Quantachrome, Boynton Beach, FL 33426, USA, model No. AS1MP) was used to quantify specific surface area and pore volume at 77K using nitrogen as the adsorbing gas. The Brunauer–Emmett–Teller (BET) method was used to measure the specific surface area of as-prepared samples. Also, surface area, average pore volume, and pore radius were obtained from the Barret–Joyner–Halenda (BJH) model. Between 400 and 1000 cm−1, FTIR spectra were acquired using an FTIR spectrometer (Nicolet iS10, Thermo Fisher Scientific, Waltham, MA, USA). The x-ray photoelectron spectrometer (Thermo Scientific K-alpha, XPS) equipped with a monochromatic X-ray source at 1486.6 eV, corresponding to the Al Kα line, was employed to collect XPS data (for the brevity XPS spectra not shown). The atomic weight percent of elements was determined using XPS.
The working electrode was prepared using nickel foam. Nickel foam was first cleaned with acetone and 3 M HCl solution for 10 min, which helped remove the oxide layer from the nickel foam’s surface. After washing and drying the nickel foam, the working electrode was made by combining 80 wt.% of the synthesized powder, 10 wt. % of acetylene black, and 10 wt. % of polyvinylidene difluoride (PVDF) with N-methyl pyrrolidinone (NMP). The slurry paste was obtained after thoroughly mixing the components and pasted onto the nickel foam. The prepared electrode on nickel foam was placed in a vacuum oven at a temperature of 60 °C for 10 h. The actual sample loading on nickel foam was determined by weighing the nickel foam before and after loading the nickel foam sample using an analytical balance (MS105DU, Mettler Toledo, max. 120 g, 0.01 mg of resolution). A platinum wire counter electrode, a saturated calomel electrode (SCE) as a reference electrode, and pasted samples on nickel foam as a working electrode were utilized in the electrochemical cell. The electrochemical performance of the electrode was assessed in 3 M KOH as an electrolyte. The electrochemical performance of the as-prepared electrode was evaluated using cyclic voltammetry (CV), galvanostatic charge-discharge techniques, and electrochemical impedance spectroscopy (EIS) by using a Versastat4-500 electrochemical workstation (Princeton Applied Research, Oak Ridge, TN, USA).

2.3. DFT Study

First-principles density functional theory (DFT) calculations were performed using the Perdew–Burke–Ernzerhof (PBE) [33] exchange-correlation functional for self-consistent calculations and geometry optimization along with the advanced hybrid functional HSE06 method to find the electronic density of states of the AlxNi1−xCo2O4 structure. The variable x ranges from 0 to 1 with a step of 0.25. Calculations were performed using VASP [34] (Vienna ab initio simulation package) with the projected augmented wave (PAW) [35] type pseudo-potential and the plane-wave basis set. The exchange and correlation part for the HSE calculations was described by hybrid functional [36] containing 40% Hartree–Fock exchange. We used a 4 × 4 × 2 k-points mesh centered at the Gamma point with a plane wave cutoff of 400 eV, which is enough for the self-consistent calculation. The global break condition for the electronic and ionic self-consistent (SC) loop was set to be 10−4 eV and 10−3 eV, respectively.

3. Results and Discussion

The room temperature powder XRD patterns of Ni1−xAlxCo2O4 samples are shown in Figure 1. The XRD patterns of the as-synthesized doped Ni1−xAlxCo2O4 are almost identical and can be assigned to NiCo2O4 (ICDD card no.02-0770). The diffraction peaks are located at 18.9°, 31.3°, 36.8°, 38.4°, 44.5°, 55.4°, 59.4°, and 65.2° can be assigned to the (111), (220), (311), (222), (400), (422), (511), and (440) agreeing with the standard crystallographic pattern of spinel NiCo2O4 phase respectively [37]. The absence of impurities in the XRD patterns confirms the complete dissolution of Al3+ into the Ni1−xAlxCo2O4 compound. The lattice parameter of Ni1−xAlxCo2O4 is listed in Table 2. As expected, the lattice parameter of Ni1−xAlxCo2O4 decreases with the content x due to Al3+ (rion ~0. 535 nm) replacing a bigger Ni ion (rionic ~0.6 nm). The average crystallite size of the NiCo2O4 calculated using Scherrer’s formula [38] is listed in Table 2. The average crystallite size of Ni1−xAlxCo2O4 falls within the range of 12.28 nm for x = 0.8 to 16.52 nm for x = 0.2. Moreover, sharp diffraction peaks indicated the crystalline nature of the calcined Ni1−xAlxCo2O4 nanostructures. The aluminum content in Ni1−xAlxCo2O4 as determined via XPS (XPS study is not reported here) is listed in Table 2 and is within the accepted range of stoichiometry of the compound.
Figure 2 shows the SEM images of Ni1−xAlxCo2O4 as a function of Al3+ content. Figure 2 shows the gradual evolution of Ni1−xAlxCo2O4 nanostructures from an urchin-like architecture for x = 0.0 to nanoplates for x = 0.8. It is evident that the initial urchin whiskers slowly merge and morph into a plate-like structure with increasing Al content. The morphology of Ni1−xAlxCo2O4 is dependent on the Al3+ content in the sample. Usually, the morphology of hierarchal nanostructures depends on the synthesis techniques and urea concentration [4]. As a dispersing agent, urea helps avoid agglomeration by finely separating nanoparticles and plays a vital role in controlling the particles’ morphology [39,40,41].
Furthermore, NH3 reacts with water to form NH4+ and OH, and CO2 hydrolyzes to form Co32− ions [42]. Tiny crystalline nuclei are created during the oxide crystal development phase, and nanoparticles of this oxide are precipitated by a rise in pH caused by NH4+ ions generated from NH3 owing to urea breakdown at higher temperatures [43,44]. Urea’s hydrolysis leads to a rise in the pH due to the increased release of NH4+ in the solution [45]. The urea hydrolysis progresses slowly at milder circumstances and at a certain urea level, and the basic solution undergoes supersaturation of the metal-hydroxide species [46].
Figure 3a shows N2 adsorption/desorption measurements were performed at 77 K between relative pressures of P/Po ~ 0.029 to 0.99. From these curves, BET-specific surface areas and corresponding pore sizes were calculated by the BJH method for all Ni1−xAlxCo2O4 (0 ≤ x ≤ 0.8). Figure 2c inset shows the pore distribution. These curves show the largest number of pores distribution at around 3 nm for all Ni1−xAlxCo2O4. This pore distribution lies between ~2.0 and 6.0 nm, a range best suited to electrode materials [46]. The larger pore volume helps boost the diffusion kinetics inside the electrode material [40,46,47]. Table 2 lists the measured BJH pore radius, volume, and BET surface area of Ni1−xAlxCo2O4 samples. The high surface area of samples, in the range of 81.74–189 m2/g, indicates that the hydrated electrolyte ions could have an increased contact area at the electrolyte/electrode surface for the Faradaic redox reaction [48,49].
The FTIR spectrum Figure 3b shows two different bands arising from the stretching vibrations of the metal-oxygen bonds at 534.2 (ν1) and 628.7 (ν2) cm−1 [1]. For the crystal structure of NiCo2O4, the ν1 band is attributed to M-O (M = Ni) bond vibrations in octahedral coordination, whereas the ν2 band is attributed to M-O (M-Co) bond vibrations in tetrahedral coordination. The stretching vibration of Co3+-O2− in the octahedral sites is shown by the FTIR spectrum at 534.2 cm−1, and the vibration of Ni2+-O at the tetrahedral sites is indicated by the FTIR spectrum at 628.7 cm−1 for NiCo2O4, respectively [50]. The presence of these vibration bands indicates that pure phase spinel NiCo2O4 nanostructures have formed. Table 3 lists the band positions for the tetrahedral and octahedral sites of the compound. Aluminum doping brings the shift of the stretching peak towards the right with an increase in Al3+ content. As the vibrational frequency is inversely proportional to the mass, the increase in vibrational frequency with Al3+ is expected due to the lower atomic weight of Al3+ replacing the Ni ion.
Cyclic voltammetry (CV) and galvanostatic charge-discharge (GCD) measurements give the electrochemical performance of Ni1−xAlxCo2O4 nanostructures. Figure 4a–e shows the cyclic voltammograms obtained at different scan rates (2–300 mV/s) in a voltage window of 0–0.6 V (vs. SCE). The Faradaic reaction for NiCo2O4, x = 0 is given by the equation below [51,52].
MO + OH ↔ MOOH + e
NiCo2O4 + OH + H2O ↔ NiOOH + 2CoOOH + e
In both the cathodic and anodic scans, there are noticeable oxidation and reduction peaks [53]. Obtained non-rectangular and unsymmetric CV curves indicate the pseudocapacitive nature of electrodes [54]. The CV curves of Ni1−xAlxCo2O4 show no extra redox peaks, implying that Ni1−xAlxCo2O4 redox processes are very similar to NiCo2O4 redox processes. The peaks are due to the redox reaction related to M-O/M-O-OH, where M stands for Ni or Co ions [55]. A positive shift in oxidation peak potential and a negative shift in reduction peak potential are found as scan rate increases [56]. This PC characteristic is derived from the Faradaic redox reaction related to the reversible reaction of Ni2+/Ni3+ and Co3+/Co4+ transitions. Figure 5a–e shows the cyclic stability curves of Ni1−xAlxCo2O4 measured up to 2000 cycles without any change in the position of redox peaks.
The Randles–Sevcik plots of the Ni1−xAlxCo2O4 samples are shown in Figure 5f. When the scan rate is increased from 2 to 300 mV s−1, the anodic peak current (Ipa) and the cathodic peak current (Ipc) both increase. More precisely, the Randles–Sevcik equation shows that both Ipa and Ipc change linearly with the square root of the scan rate (ν1/2) [57,58]. These results suggest that the electrode–electrolyte interface’s redox events are rapid, quasi-reversible, and only limited by electrolyte diffusion [59]. The positive shift of oxidation peak and negative shift of reduction peak potential indicate relatively low electrode materials resistance and good electrochemical reversibility [1].
The calculation of specific capacitance was done from CV curves using the equation below [60]:
C sp = V 1 V 2 I V d V m v ( V 2 V 1 )
where I (A) is current, m (g) is the mass of the active material, v (mV/s) is the scan rate, V1 and V2 (V) are the start and end voltage of the CV scan.
Figure 4f shows the Csp of Ni1−xAlxCo2O4 samples as a function of scan rates. The obtained values of Csp for Ni1−xAlxCo2O4 nanostructures are 512, 368, 371, 380, and 356 F/g at a scan rate of 2 mV/s for Ni1−xAlxCo2O4 x = 0.0, 0.2, 0.4, 0.6, and 0.8, respectively. Compared to all Ni1−xAlxCo2O4 nanostructures, NiCo2O4 x = 0.0 exhibited the highest specific capacitance at all scan rates with a maximum value of 512 F/g at 2 mV/s. The higher electrochemical utilization and high electroactive surface area of the synthesized nanostructure occur for Ni1−xAlxCo2O4, x = 0. The observed decrease in Csp with Al3+ substitution suggests that in the presence of Al3+, the Ni and Co in Ni1−xAlxCo2O4 may not behave synergistically as an effective Faradaic charging site as compared with the Ni and Co sites in NiCo2O4. Moreover, it can be deduced that, with the increase in Al3+ content, overall effective redox sites have decreased, which leads to a decrease in the specific capacitance.
The total charge storage process is affected by factors such as the Faradaic contribution from the insertion process of electrolyte ions, the Faradaic contribution from the charge-transfer process, and due to the high surface area, the contribution from both pseudocapacitance and the non-Faradaic contribution from the double layer effects [61]. CV data are analyzed at different scan rates and can characterize capacitive effects by equation [62,63].
I = avb
where the peak current is I (A), the scan rate is v (mV/s), and a and b are adjustable parameters. The value of b specifies the charge storage mechanism. If b = 1, it indicates a surface-controlled reaction that the capacitive surface mechanism is dominant, and if b = 0.5, it indicates a diffusion-controlled Faradaic reaction during the charge storage mechanism [64]. Figure 5f shows the peak current vs. square root of the scan rate curve. These curves are fitted with the Equation (2) which gives values of b ~ 0.6239, 0.6398, 0.6218, 0.6134, and 0.5894 for x = 0.0, 0.2, 0.4, 0.6, and 0.8, respectively. That b > 0.5 for all electrodes indicates both diffusion-controlled and surface-controlled reactions [54]. Moreover, in Figure 5f, the lower value of the scan rate region has a linear dependence relation with the square root of the scan rate; this indicates a quasi-reversible electrochemical reaction [65]. Figure 5f also indicates that, at a higher scan rate > 100 mV/s, the ion diffusion is limited to the surface of the active material of the electrode, i.e., EDLC dominates the pseudocapacitor, and diffusion of OH ions can adhere only to the outer layer of the nanostructure, which contributes less to the electrochemical capacitive behavior [66]. On the other hand, the Faradaic redox reaction dominates at scan rates < 100 mV/s due to more effective usage of the working electrode’s active material [67]. Moreover, the diffusion of OH- ions can easily penetrate deep into the nanostructure’s interlayer, which leads to adsorption of more ions and hence ends up with higher specific capacitance [68].
The capacitive and diffusion processes are the two mechanisms for current response at a fixed potential [69,70]. The equation below gives their contributions [71].
Csp = k1 + k2−1/2
k1 and k2 are determined from the Csp vs.V1/2 linear plot, where k2 is the slope and k1 is the intercept. k1 indicates diffusion and k2 capacitance contribution to the total specific capacitance for a given voltage. For the calculation, the specific capacitance is plotted against the slow scan rate up to a value of 20 mV/s and performed a regression fit using Equation (3). The obtained value for k1 and k2 was used to determine the fractional contribution in terms of diffusion and capacitance from total specific capacitance [72], as given in Figure 5g. Figure 5g shows that the contribution to the current response at a fixed potential is more diffusive than capacitive and decreases with the Al content.
To further estimate the potential application of as-prepared electrodes for supercapacitors, the galvanostatic charge-discharge (GCD) measurement was performed. GCD was measured within the voltage window of 0.0 to 0.6 at different current densities between 1 A/g to 30 A/g in 3 M KOH solution are given in Figure 6a–e. These figures indicate that prepared electrodes of Ni1−xAlxCo2O4 can charge and discharge rapidly with good electrochemical reversibility at different constant current densities. As in CV curves, the non-linear relationship between the potential and time in both the charge-discharge cycle indicates the capacitance of studied nanomaterials is not constant in between the potential range and reflects typical PCs behavior [73]. The non-linearity of the GCD curve is a consequence of the Co3+/Co4+, and Ni2+/Ni3+ ions redox reactions with OH. The discharge process occurs in three distinct steps. First, there is a rapid potential drop that occurs owing to the internal resistance. The second is a slow potential decay at an intermediate time, which is due to the Faradaic redox reaction. The third is a fast potential decay which is created by electric double layer capacitance [74]. The specific capacitance values for electrodes of Ni1−xAlxCo2O4 were calculated by using Equation (4) [75].
C sp = I t m Δ V
where Csp is the specific capacitance (F/g), I (A) is the charge-discharge current, ΔV (V) is the potential range, m (g) is the mass of the electroactive materials and t (s) is the discharging time.
The pseudocapacitance behavior of electrodes with respect to their discharging time is confirmed by the presence of voltage plateau on GCD curves [76]. The specific capacitance values of Ni1−xAlxCo2O4 were calculated using Equation (4) at 0.5 A/g are 268, 194, 192, 176, and 167 F/g for x = 0.0, 0.2, 0.4, 0.6, and 0.8, respectively, and are listed in Table 4. Figure 6f shows the dependence of current density with specific capacitance. A decrease in the specific capacitance with the increase of the discharge current is caused by the insufficient time available for the diffusion of the electrolyte ions into the inner electrode surface and the increase of the potential drop towards higher discharge currents [77].
The cyclic stability of Ni1−xAlxCo2O4, x = 0.0, and x = 0.8, electrodes were evaluated by the repeated charge-discharge measurements up to 5000 cycles at a constant current density of 10 A/g and 5 A/g in the potential range between 0.0 and 0.6 V in 3 M KOH, shown in Figure 7a,b respectively. The percentage retention in specific capacitance was calculated using Equation (5) [78].
% retention in specific capacitance = (C#/C1) × 100
where C# and C1 are specific capacitance at various cycles and the 1st cycle, respectively. Hence, the % retention in specific capacitance after 5000 cycles for Ni1−xAlxCo2O4 x = 0.0 and x = 0.8 is 97% and 108.5% respectively. This shows that even after 5000 cycles, the electrode shows outstanding cyclic stability, and hence there is not much high specific capacitance fade from its initial values in x = 0.0. Similarly, in the case of x = 0.8, the specific capacitance improves to 108.5% of its initial value, and this suggests an improvement in the surface moistened by electrolyte ions, access of electrolyte ions up to the pore of electrode materials, and activation of electrodes by electrolyte ions upon cycle repetition [79]. Furthermore, the Coulombic efficiency for x = 0.0 and x = 0.8 is 100% and 99.2%, as shown in Figure 7a,b, respectively.
Ragone plots of synthesized Ni1−xAlxCo2O4 are given in Figure 8a. The energy (E) and power (P) densities are determined using the equations below [32],
E = (1/2)CV2
P = E/t
where C is the electrodes’ specific capacitance, V is the operating voltage, and the letter t denotes the discharge time in seconds. An excellent-performance supercapacitor must exhibit a high value of both energy density and power density together. Figure 8a shows the superior performance of electrodes with an energy density of up to 12.4 Wh/kg for x = 0.0, with up to a peak power density of 6670.9 W/kg for x = 0.6. Furthermore, Table 5 compares the electrodes’ specific capacitances derived from the doping of transition elements on NiCo2O4 using the hydrothermal method.
The superior electrochemical performance of as-synthesized Ni1−xAlxCo2O4 nanostructure electrode materials was further confirmed by the electrochemical impedance spectroscopy (EIS) measurement held at 10 mV AC perturbation in the frequency range from 10 kHz to 0.05 Hz. Figure 8b shows the tendency of the real part of the impedance (Z′) to decrease with increasing frequency for all samples. When x = 0.0, it has the lowest real part of the impedance. The magnitude of the real part of the impedance (Z’) increases as Al3+ content increases, which affects the rise of grains, grain boundaries, and electrode interface resistance. Figure 8c shows the Nyquist plots for all of the Ni1−xAlxCo2O4 nanostructures. In this plot, the NiCo2O4 nanostructure electrode exhibits a small semicircle that suggests low internal resistance and charge transfer resistance, and hence this resistance is smaller than Al3+ doped NiCo2O4. Towards higher frequencies, the real part (Zreal) of the impedance represents a combined resistance of contact resistance at active material/current collector interface, ionic resistance of the electrolyte, and intrinsic resistance of the nanomaterials used in the electrode [80]. The semicircle at the high-frequency range is due to the Faradaic charge transfer resistance (Rct) of the redox reaction between the electrode and electrolyte. The decrease in the diameter of the semicircle indicates lower charge transfer resistance [81]. These variations of the EIS curve from semicircle towards higher frequencies and linearity towards lower frequencies relate to reversible Faradaic redox reaction and access of OH ions into the pore of the electrode material. This behavior of the EIS curve describes the PCs properties of the prepared Ni1−xAlxCo2O4 nanostructure electrode [82]. The slope of the 45° portion of the curve is called Warburg resistance (Zw) and results from the frequency dependence of ion diffusion/transport in the electrolyte to the electrode surface [83,84]. The Nyquist plot is a vertical line for an ideal electrode material, and more is the vertical line represents better electrolyte diffusion and capacitive behavior [85]. Figure 8c shows that Rct is smaller for NiCo2O4 and has a more vertical line, which offers better rate capability.
First-principles density functional theory (DFT) calculations were performed using VASP2 [34] (Vienna ab initio simulation package) with the projected augmented wave (PAW) [35] type pseudo-potential and the plane-wave basis set. The exchange and correlation part for the HSE calculations was described by hybrid functional [36] containing a 40% Hartree-Fock exchange. Here, we used a 4 × 4 × 2 k-points mesh centered at the Gamma point with a plane wave cutoff of 400 eV, which is enough for the self-consistent calculation. The global break condition for the electronic and ionic self-consistent (SC) loop was set to be 10−4 eV and 10−3 eV, respectively. In our calculation, we dealt with a cell containing 28 atoms as a total. The value of gaussian smearing was adjusted to be 0.20. Even though the ratio of x is 20% in the experiment, we chose 25% in the calculations because changing it to 20% would require a much larger supercell, which eventually increases the computing time and memory excessively. Figure 9a shows the side view along the x-axis of the crystal structure of Ni1−xAlxCo2O4, while Figure 9b shows the band-gap variation from 2.42 eV to 4.82 eV for the different values of x. For the system without Al atom, which is Ni1Co2O4, the band gap was found at 2.42 eV, which agrees with the previously reported experimental values of 2.06 eV [86] and 2.64 eV [87]. This value is the minimum among all the cases of adding Al to the system. Adding an Al atom above 0.8, the band-gap reached its maximum value of 4.82 eV for the case where Al completely replaces Ni, as shown in Figure 9a. These results suggest that the energy band-gap of Ni1−xAlxCo2O4 remains largely unaltered (x < 0.8) with the Al3+ substitution, which translates to the fact that the electrical conductivity of Ni1−xAlxCo2O4 could remain unaltered with the Al3+ substitution below x = 0.8. In the absence of the contribution of electrical conductivity to the electrochemical performance of Ni1−xAlxCo2O4, at least for x < 0.8, it can be concluded that the electrochemical performance of Ni1−xAlxCo2O4 could be dictated by other factors, such as morphology, the oxidation potential of participating ions, active metal sites, and the internal resistance of the compound.

4. Conclusions

Finally, a simple and cost-effective technique was developed to report the effect of Al3+ doping on the structural and electrochemical performance of the NiCo2O4 nanostructure. The experimental results revealed the dependence of surface area, specific capacitance, and electrochemical performance of NiCo2O4 on Al3+ content. The doped compound’s overall electrochemical performance, viz. energy and power density, was observed to be significantly superior to other doped NiCo2O4 compounds. The DFT calculation suggests that the energy band-gap of Ni1−xAlxCo2O4 (<0.8) remains largely invariant, and hence the electrical conductivity. Thus, the observed reduction in the specific capacitance of doped Ni1−xAlxCo2O4 could be attributed to the decrease in active redox sites with Al3+ substitution. The study indicates that Al3+ can be an excellent cost-effective substitution for expensive cobalt in NiCo2O4 to provide the desired electrochemical performance. Hence, this work highlights the development of cheaper and more promising electrode doped materials without sacrificing the performance of NiCo2O4.

Author Contributions

Conceptualization and supervision, S.R.M.; methodology, D.G., data collection and validation. J.C., W.L. and F.A.P.; software, R.B. and X.S.; formal analysis, S.R.M. and R.K.G.; data curation, D.G.; writing—D.G. and X.S.; writing—review and editing, S.R.M. and R.K.G.; supervision, S.R.M.; project administration, S.R.M.; funding acquisition, S.R.M. All authors have read and agreed to the published version of the manuscript.

Funding

Dr. Mishra extends thanks to the FRG grant provided by the University of Memphis, Memphis, TN, USA. Dr. Ram K. Gupta expresses his sincere acknowledgment of the Polymer Chemistry Initiative at Pittsburg State University, Pittsburg, KS, USA, for providing financial and research support for this project.

Institutional Review Board Statement

None.

Informed Consent Statement

Not applicable.

Acknowledgments

Authors acknowledge the Integrated Microscopy Center (IMC) at the University of Memphis for helping with XPS study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhu, Y.; Pu, X.; Song, W.; Wu, Z.; Zhou, Z.; He, X.; Lu, F.; Jing, M.; Tang, B.; Ji, X. High capacity NiCo2O4 nanorods as electrode materials for supercapacitor. J. Alloy. Compd. 2014, 617, 988–993. [Google Scholar] [CrossRef]
  2. Zequine, C.; Ranaweera, C.K.; Wang, Z.; Singh, S.; Tripathi, P.; Srivastava, O.N.; Gupta, B.K.; Ramasamy, K.; Kahol, P.K.; Dvornic, P.R.; et al. High performance and flexible supercapacitors based on carbonized bamboo fibers for wide temperature applications. Sci. Rep. 2016, 6, 31704. [Google Scholar] [CrossRef]
  3. Yong, S. Fabrication and Characterization of Fabric Supercapacitor. Ph.D. Thesis, University of Southampton, Southampton, UK, 2016. [Google Scholar]
  4. Dubal, D.P.; Gomez-Romero, P.; Sankapal, B.R.; Holze, R. Nickel cobaltite as an emerging material for supercapacitors: An overview. Nano Energy 2015, 11, 377–399. [Google Scholar] [CrossRef]
  5. Zhi, M.; Xiang, C.; Li, J.; Li, M.; Wu, N. Nanostructured carbon–metal oxide composite electrodes for supercapacitors: A review. Nanoscale 2013, 5, 72–88. [Google Scholar] [CrossRef]
  6. Han, C.; Xu, X.; Mu, H.; Tian, Q.; Li, Q.; Liu, Y.; Zhang, X.; Zhao, Z.; Su, X. Construction of hierarchical sea urchin-like manganese substituted nickel cobaltite@tricobalt tetraoxide core-shell microspheres on nickel foam as binder-free electrodes for high performance supercapacitors. J. Colloid Interface Sci. 2021, 596, 89–99. [Google Scholar] [CrossRef]
  7. Zhong, C.; Deng, Y.; Hu, W.; Qiao, J.; Zhang, L.; Zhang, J. A review of electrolyte materials and compositions for electrochemical supercapacitors. Chem. Soc. Rev. 2015, 44, 7484–7539. [Google Scholar] [CrossRef]
  8. Sarangapani, S.; Tilak, B.V.; Chen, C. Materials for Electrochemical Capacitors: Theoretical and Experimental Constraints. J. Electrochem. Soc. 1996, 143, 3791. [Google Scholar] [CrossRef]
  9. Huang, L.; Chen, D.; Ding, Y.; Feng, S.; Wang, Z.L.; Liu, M. Nickel–Cobalt Hydroxide Nanosheets Coated on NiCo2O4 Nanowires Grown on Carbon Fiber Paper for High-Performance Pseudocapacitors. Nano Lett. 2013, 13, 3135–3139. [Google Scholar] [CrossRef]
  10. Conway, B.E. Electrochemical Supercapacitors: Scientific Fundamentals and Technological Applications; Springer Science and Business Media: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  11. Liu, T.; Pell, W.G.; Conway, B.E. Self-discharge and potential recovery phenomena at thermally and electrochemically prepared RuO2 supercapacitor electrodes. Electrochim. Acta 1997, 42, 3541–3552. [Google Scholar] [CrossRef]
  12. Cheng, Q.; Tang, J.; Ma, J.; Zhang, H.; Shinya, N.; Qin, L.-C. Graphene and nanostructured MnO2 composite electrodes for supercapacitors. Carbon 2011, 49, 2917–2925. [Google Scholar] [CrossRef]
  13. Yuan, C.; Zhang, X.; Su, L.; Gao, B.; Shen, L. Facile synthesis and self-assembly of hierarchical porous NiO nano/micro spherical superstructures for high performance supercapacitors. J. Mater. Chem. 2009, 19, 5772–5777. [Google Scholar] [CrossRef]
  14. Ji, J.; Zhang, L.; Ji, H.; Li, Y.; Zhao, X.; Bai, X.; Fan, X.; Zhang, F.; Ruoff, R.S. Nanoporous Ni(OH)2 Thin Film on 3D Ultrathin-Graphite Foam for Asymmetric Supercapacitor. ACS Nano 2013, 7, 6237–6243. [Google Scholar] [CrossRef]
  15. Moosavifard, S.E.; El-Kady, M.F.; Rahmanifar, M.S.; Kaner, R.B.; Mousavi, M.F. Designing 3D highly ordered nanoporous CuO electrodes for high-performance asymmetric supercapacitors. ACS Appl. Mater. Interfaces 2015, 7, 4851–4860. [Google Scholar] [CrossRef]
  16. Meher, S.K.; Rao, G.R. Ultralayered Co3O4 for High-Performance Supercapacitor Applications. J. Phys. Chem. C 2011, 115, 15646–15654. [Google Scholar] [CrossRef]
  17. Wang, R.; Yan, X.; Lang, J.; Zheng, Z.; Zhang, P. A hybrid supercapacitor based on flower-like Co(OH)2 and urchin-like VN electrode materials. J. Mater. Chem. A 2014, 2, 12724–12732. [Google Scholar] [CrossRef]
  18. Dubal, D.P.; Gund, G.; Lokhande, C.D.; Holze, R. Decoration of Spongelike Ni(OH)2 Nanoparticles onto MWCNTs Using an Easily Manipulated Chemical Protocol for Supercapacitors. ACS Appl. Mater. Interfaces 2013, 5, 2446–2454. [Google Scholar] [CrossRef]
  19. Dubal, D.; Holze, R. A successive ionic layer adsorption and reaction (SILAR) method to induce Mn3O4 nanospots on CNTs for supercapacitors. New J. Chem. 2013, 37, 403–408. [Google Scholar] [CrossRef]
  20. Aricò, A.S.; Bruce, P.; Scrosati, B.; Tarascon, J.-M.; VAN Schalkwijk, W. Nanostructured materials for advanced energy conversion and storage devices. Mater. Sustain. Energy 2010, 148–159. [Google Scholar] [CrossRef]
  21. Bruce, P.G.; Scrosati, B.; Tarascon, J.-M. Nanomaterials for Rechargeable Lithium Batteries. Angew. Chem. Int. Ed. 2008, 47, 2930–2946. [Google Scholar] [CrossRef]
  22. Wang, G.; Zhang, L.; Zhang, J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41, 797–828. [Google Scholar] [CrossRef] [Green Version]
  23. Zhang, Y.; Li, L.; Su, H.; Huang, W.; Dong, X. Binary metal oxide: Advanced energy storage materials in supercapacitors. J. Mater. Chem. A 2015, 3, 43–59. [Google Scholar] [CrossRef]
  24. Wang, Q.; Wang, X.; Xu, J.; Ouyang, X.; Hou, X.; Chen, D.; Wang, R.; Shen, G. Flexible coaxial-type fiber supercapacitor based on NiCo2O4 nanosheets electrodes. Nano Energy 2014, 8, 44–51. [Google Scholar] [CrossRef]
  25. Shen, L.; Yu, L.; Yu, X.-Y.; Zhang, X.; Lou, X.W. (David) Self-Templated Formation of Uniform NiCo2O4Hollow Spheres with Complex Interior Structures for Lithium-Ion Batteries and Supercapacitors. Angew. Chem. Int. Ed. 2015, 54, 1868–1872. [Google Scholar] [CrossRef]
  26. Chakrabarty, S.; Mukherjee, A.; Basu, S. RGO-MoS2 supported NiCo2O4 catalyst toward solar water splitting and dye degradation. ACS Sustain. Chem. Eng. 2018, 6, 5238–5247. [Google Scholar] [CrossRef]
  27. Yang, S.; Yu, J.; Jiang, T.; Zhu, L.; Xu, X. High performance symmetric solid state supercapacitor based on electrode of RuxNi1−xCo2O4 grown on nickel foam. J. Alloy. Compd. 2018, 764, 767–775. [Google Scholar] [CrossRef]
  28. Mu, H.; Su, X.; Zhao, Z.; Han, C.; Wang, Z.; Zhao, P. Facile synthesis of Ni0.5Mn0.5Co2O4 nanoflowers as high-performance electrode material for supercapacitors. J. Am. Ceram. Soc. 2019, 102, 6893–6903. [Google Scholar]
  29. Uke, S.J.; Akhare, V.P.; Meshram-Mardikar, S.P.; Bodade, A.B.; Chaudhari, G.N. PEG Assisted Hydrothermal Fabrication of Undoped and Cr Doped NiCo2O4 Nanorods and Their Electrochemical Performance for Supercapacitor Application. Adv. Sci. Eng. Med. 2019, 11, 357–366. [Google Scholar] [CrossRef]
  30. Gao, K.; Li, S.D. Hollow fibrous NiCo2O4 electrodes with controllable Zn substitution sites for supercapacitors. J. Alloy. Compd. 2020, 832, 154927. [Google Scholar] [CrossRef]
  31. Liu, L.; Zhang, H.; Fang, L.; Mu, Y.; Wang, Y. Facile preparation of novel dandelion-like Fe-doped NiCo2O4 microspheres@nanomeshes for excellent capacitive property in asymmetric supercapacitors. J. Power Sources 2016, 327, 135–144. [Google Scholar] [CrossRef]
  32. Guragain, D.; Zequine, C.; Bhattarai, R.; Choi, J.; Gupta, R.K.; Shen, X.; Mishra, S.R. Effect of dopant on the morphology and electrochemical performance of Ni1−xCaxCo2O4 (0≤ x≤ 0.8) oxide hierarchical structures. MRS Adv. 2020, 5, 2487–2494. [Google Scholar] [CrossRef]
  33. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  34. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef] [PubMed]
  35. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  36. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
  37. Chaudhari, S.; Bhattacharjya, D.; Yu, J.-S. Facile Synthesis of Hexagonal NiCo2O4Nanoplates as High-Performance Anode Material for Li-Ion Batteries. Bull. Korean Chem. Soc. 2015, 36, 2330–2336. [Google Scholar] [CrossRef]
  38. Sumadiyasa, M.; Manuaba, I. Determining Crystallite Size Using Scherrer Formula, Williamson-Hull Plot, and Particle Size with SEM. Bul. Fis. 2018, 19, 28–34. [Google Scholar] [CrossRef] [Green Version]
  39. Gonzalo-Juan, I.; Ferrari, B.; Colomer, M. Influence of the urea content on the YSZ hydrothermal synthesis under dilute conditions and its role as dispersant agent in the post-reaction medium. J. Eur. Ceram. Soc. 2009, 29, 3185–3195. [Google Scholar] [CrossRef]
  40. Hadke, S.; Kalimila, M.T.; Rathkanthiwar, S.; Gour, S.; Sonkusare, R.; Ballal, A. Role of fuel and fuel-to-oxidizer ratio in combustion synthesis of nano-crystalline nickel oxide powders. Ceram. Int. 2015, 41, 14949–14957. [Google Scholar] [CrossRef]
  41. Adhikari, H.; Ghimire, M.; Ranaweera, C.K.; Bhoyate, S.; Gupta, R.K.; Alam, J.; Mishra, S.R. Synthesis and electrochemical performance of hydrothermally synthesized Co3O4 nanostructured particles in presence of urea. J. Alloy. Compd. 2017, 708, 628–638. [Google Scholar] [CrossRef]
  42. Pan, J.H.; Zhang, X.; Du, A.J.; Bai, H.; Ng, J.; Sun, D. A hierarchically assembled mesoporous ZnO hemisphere array and hollow microspheres for photocatalytic membrane water filtration. Phys. Chem. Chem. Phys. 2012, 14, 7481–7489. [Google Scholar] [CrossRef]
  43. Zhu, Y.; Ji, X.; Yin, R.; Hu, Z.; Qiu, X.; Wu, Z.; Liu, Y. Nanorod-assembled NiCo2O4 hollow microspheres assisted by an ionic liquid as advanced electrode materials for supercapacitors. RSC Adv. 2017, 7, 11123–11128. [Google Scholar] [CrossRef] [Green Version]
  44. Jadhav, A.; Bandal, H.A.; Kim, H. NiCo2O4 hollow sphere as an efficient catalyst for hydrogen generation by NaBH 4 hydrolysis. Mater. Lett. 2017, 198, 50–53. [Google Scholar] [CrossRef]
  45. Burt, C.D.; Cabrera, M.L.; Rothrock, M.J.; Kissel, D.E. Urea Hydrolysis and Calcium Carbonate Precipitation in Gypsum-Amended Broiler Litter. J. Environ. Qual. 2018, 47, 162–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Zhi, J.; Wang, Y.; Deng, S.; Hu, A. Study on the relation between pore size and supercapacitance in mesoporous carbon electrodes with silica-supported carbon nanomembranes. RSC Adv. 2014, 4, 40296–40300. [Google Scholar] [CrossRef]
  47. Kondrat, S.; Perez, C.; Presser, V.; Gogotsi, Y.; Kornyshev, A.A. Effect of pore size and its dispersity on the energy storage in nanoporous supercapacitors. Energy Environ. Sci. 2012, 5, 6474–6479. [Google Scholar] [CrossRef]
  48. Largeot, C.; Portet, C.; Chmiola, J.; Taberna, P.-L.; Gogotsi, Y.; Simon, P. Relation between the Ion Size and Pore Size for an Electric Double-Layer Capacitor. J. Am. Chem. Soc. 2008, 130, 2730–2731. [Google Scholar] [CrossRef]
  49. Guragain, D.; Zequine, C.; Poudel, T.; Neupane, D.; Gupta, R.K.; Mishra, S.R. Facile Synthesis of Bio-Templated Tubular Co3O4 Microstructure and Its Electrochemical Performance in Aqueous Electrolytes. J. Nanosci. Nanotechnol. 2020, 20, 3182–3194. [Google Scholar] [CrossRef]
  50. Khairy, M.; Mousa, M. Synthesis of Ternary and Quaternary metal oxides based on Ni, Mn, Cu, and Co for high-performance Supercapacitor. J. Ovonic Res. 2019, 15, 181–198. [Google Scholar]
  51. Liu, X.; Zhang, Y.; Xia, X.; Shi, S.; Lu, Y.; Wang, X.; Gu, C.; Tu, J. Self-assembled porous NiCo2O4 hetero-structure array for electrochemical capacitor. J. Power Sources 2013, 239, 157–163. [Google Scholar] [CrossRef]
  52. Wang, X.; Han, X.; Lim, M.; Singh, N.; Gan, C.L.; Jan, M.; Lee, P.S. Nickel Cobalt Oxide-Single Wall Carbon Nanotube Composite Material for Superior Cycling Stability and High-Performance Supercapacitor Application. J. Phys. Chem. C 2012, 116, 12448–12454. [Google Scholar] [CrossRef]
  53. Pendashteh, A.; Rahmanifar, M.; Kaner, R.B.; Mousavi, M.F. Facile synthesis of nanostructured CuCo2O4 as a novel electrode material for high-rate supercapacitors. Chem. Commun. 2014, 50, 1972–1975. [Google Scholar] [CrossRef]
  54. Kirubasankar, B.; Vijayan, S.; Angaiah, S. Sonochemical synthesis of a 2D–2D MoSe2/graphene nanohybrid electrode material for asymmetric supercapacitors. Sustain. Energy Fuels 2019, 3, 467–477. [Google Scholar] [CrossRef]
  55. Zou, R.; Xu, K.; Wang, T.; He, G.; Liu, Q.; Liu, X.; Zhang, Z.; Hu, J. Chain-like NiCo2O4 nanowires with different exposed reactive planes for high-performance supercapacitors. J. Mater. Chem. A 2013, 1, 8560–8566. [Google Scholar] [CrossRef]
  56. Liu, X.; Long, Q.; Jiang, C.; Zhan, B.; Li, C.; Liu, S.; Zhao, Q.; Huang, W.; Dong, X. Facile and green synthesis of mesoporous Co3O4 nanocubes and their applications for supercapacitors. Nanoscale 2013, 5, 6525–6529. [Google Scholar] [CrossRef] [PubMed]
  57. Ren, J.; Bai, W.; Guan, G.; Zhang, Y.; Peng, H. Flexible and Weaveable Capacitor Wire Based on a Carbon Nanocomposite Fiber. Adv. Mater. 2013, 25, 5965–5970. [Google Scholar] [CrossRef]
  58. Wei, H.; Zhu, J.; Wu, S.; Wei, S.; Guo, Z. Electrochromic polyaniline/graphite oxide nanocomposites with endured electrochemical energy storage. Polymer 2013, 54, 1820–1831. [Google Scholar] [CrossRef]
  59. Xu, P.; Gu, T.; Cao, Z.; Wei, B.; Yu, J.; Li, F.; Byun, J.-H.; Lu, W.; Li, Q.; Chou, T.-W. Carbon Nanotube Fiber Based Stretchable Wire-Shaped Supercapacitors. Adv. Energy Mater. 2014, 4, 1300759. [Google Scholar] [CrossRef]
  60. Ghosh, D.; Giri, S.; Das, C.K. Hydrothermal synthesis of platelet β Co(OH)2 and Co3O4: Smart electrode material for energy storage application. Environ. Prog. Sustain. Energy 2014, 33, 1059–1064. [Google Scholar] [CrossRef]
  61. 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]
  62. Augustyn, V.; Come, J.; Lowe, M.A.; Kim, J.W.; Taberna, P.-L.; Tolbert, S.H.; Abruña, H.D.; Simon, P.; Dunn, B. High-rate electrochemical energy storage through Li+ intercalation pseudocapacitance. Nat. Mater. 2013, 12, 518–522. [Google Scholar] [CrossRef]
  63. Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S.-E. Li+Ion Insertion in TiO2 (Anatase). 1. Chronoamperometry on CVD Films and Nanoporous Films. J. Phys. Chem. B 1997, 101, 7710–7716. [Google Scholar] [CrossRef]
  64. Zhai, S.; Karahan, H.E.; Wang, C.; Pei, Z.; Wei, L.; Chen, Y. 1D Supercapacitors for Emerging Electronics: Current Status and Future Directions. Adv. Mater. 2020, 32, e1902387. [Google Scholar] [CrossRef]
  65. Muhammad, H.; Tahiri, I.A.; Muhammad, M.; Masood, Z.; Versiani, M.A.; Khaliq, O.; Latif, M.; Hanif, M. A comprehensive heterogeneous electron transfer rate constant evaluation of dissolved oxygen in DMSO at glassy carbon electrode measured by different electrochemical methods. J. Electroanal. Chem. 2016, 775, 157–162. [Google Scholar] [CrossRef]
  66. Ali, G.; Fouad, O.A.; Makhlouf, S.; Yusoff, M.; Chong, K.F. Co3O4/SiO2 nanocomposites for supercapacitor application. J. Solid State Electrochem. 2014, 18, 2505–2512. [Google Scholar] [CrossRef] [Green Version]
  67. Guragain, D.; Zequine, C.; Gupta, R.K.; Mishra, S.R. Facile Synthesis of Bio-Template Tubular MCo2O4 (M = Cr, Mn, Ni) Microstructure and Its Electrochemical Performance in Aqueous Electrolyte. Processes 2020, 8, 343. [Google Scholar] [CrossRef] [Green Version]
  68. Ali, G.A.; Tan, L.L.; Jose, R.; Yusoff, M.M.; Chong, K.F. Electrochemical performance studies of MnO2 nanoflowers recovered from spent battery. Mater. Res. Bull. 2014, 60, 5–9. [Google Scholar] [CrossRef] [Green Version]
  69. Liu, T.C.; Pell, W.G.; Conway, B.E.; Roberson, S.L. Behavior of molybdenum nitrides as materials for electrochemical capacitors comparison with ruthenium oxide. J. Electrochem. Soc. 1998, 145, 1882–1888. [Google Scholar] [CrossRef]
  70. Wang, J.; Polleux, J.; Lim, A.J.; Dunn, B. Pseudocapacitive Contributions to Electrochemical Energy Storage in TiO2(Anatase) Nanoparticles. J. Phys. Chem. C 2007, 111, 14925–14931. [Google Scholar] [CrossRef]
  71. Jiao, X.; Hao, Q.; Liu, P.; Xia, X.; Lei, W. Facile synthesis of T-Nb2O5 nanosheets/nitrogen and sulfur co-doped graphene for high performance lithium-ion hybrid supercapacitors. Sci. China Mater. 2017, 61, 273–284. [Google Scholar] [CrossRef] [Green Version]
  72. Wang, H.; Yang, Y.; Li, Q.; Lu, W.; Ning, J.; Zhong, Y.; Zhang, Z.; Hu, Y. Molecule-assisted modulation of the high-valence Co3+ in 3D honeycomb-like CoxSy networks for high-performance solid-state asymmetric supercapacitors. Sci. China Mater. 2021, 64, 840–851. [Google Scholar] [CrossRef]
  73. Fu, H.; Liu, Y.; Chen, L.; Shi, Y.; Kong, W.; Hou, J.; Yu, F.; Wei, T.; Wang, H.; Guo, X. Designed formation of NiCo2O4 with different morphologies self-assembled from nanoparticles for asymmetric supercapacitors and electrocatalysts for oxygen evolution reaction. Electrochim. Acta 2019, 296, 719–729. [Google Scholar] [CrossRef]
  74. Hao, C.; Zhou, S.; Wang, J.; Wang, X.; Gao, H.; Ge, C. Preparation of Hierarchical Spinel NiCo2O4 Nanowires for High-Performance Supercapacitors. Ind. Eng. Chem. Res. 2018, 57, 2517–2525. [Google Scholar] [CrossRef]
  75. Yu, J.; Fu, N.; Zhao, J.; Liu, R.; Li, F.; Du, Y.; Yang, Z. High Specific Capacitance Electrode Material for Supercapacitors Based on Resin-Derived Nitrogen-Doped Porous Carbons. ACS Omega 2019, 4, 15904–15911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Guo, J.; Yin, Z.; Zang, X.; Dai, Z.; Zhang, Y.; Huang, W.; Dong, X. Facile one-pot synthesis of NiCo2O4 hollow spheres with controllable number of shells for high-performance supercapacitors. Nano Res. 2017, 10, 405–414. [Google Scholar] [CrossRef]
  77. Qing, X.; Liu, S.; Huang, K.; Lv, K.; Yang, Y.; Lu, Z.-G.; Fang, D.; Liang, X. Facile synthesis of Co3O4 nanoflowers grown on Ni foam with superior electrochemical performance. Electrochim. Acta 2011, 56, 4985–4991. [Google Scholar] [CrossRef]
  78. Mishra, S.; Adhikari, H.; Kunwar, D.; Ranaweera, C.; Sapkota, B.; Ghimire, M.; Gupta, R.; Alam, J. Facile Hydrothermal Synthesis of Hollow Fe3O4 Nanospheres: Effect of Hydrolyzing Agents and Electrolytes on Electrocapacitive Performance of Advanced Electrodes. Int. J. Met. Met. Phys. 2017, 2, 1–15. [Google Scholar] [CrossRef] [Green Version]
  79. Sennu, P.; Aravindan, V.; Lee, Y.-S. High energy asymmetric supercapacitor with 1D@2D structured NiCo2O4@Co3O4 and jackfruit derived high surface area porous carbon. J. Power Sources 2016, 306, 248–257. [Google Scholar] [CrossRef]
  80. Gamby, J.; Taberna, P.; Simon, P.; Fauvarque, J.; Chesneau, M. Studies and characterisations of various activated carbons used for carbon/carbon supercapacitors. J. Power Sources 2001, 101, 109–116. [Google Scholar] [CrossRef]
  81. Zequine, C.; Wang, F.; Li, X.; Guragain, D.; Mishra, S.; Siam, K.; Kahol, P.K.; Gupta, R.K. Nanosheets of CuCo2O4 as a High-Performance Electrocatalyst in Urea Oxidation. Appl. Sci. 2019, 9, 793. [Google Scholar] [CrossRef] [Green Version]
  82. Jokar, E.; Zad, A.I.; Shahrokhian, S. Synthesis and characterization of NiCo2O4 nanorods for preparation of supercapacitor electrodes. J. Solid State Electrochem. 2014, 19, 269–274. [Google Scholar] [CrossRef]
  83. Di Fabio, A.; Giorgi, A.; Mastragostino, M.; Soavi, F. Carbon-poly (3-methylthiophene) hybrid supercapacitors. J. Electrochem. Soc. 2001, 148, a845–a850. [Google Scholar] [CrossRef]
  84. Wu, M.-S.; Huang, C.-Y.; Lin, K.-H. Electrophoretic deposition of nickel oxide electrode for high-rate electrochemical capacitors. J. Power Sources 2009, 186, 557–564. [Google Scholar] [CrossRef]
  85. Stoller, M.D.; Park, S.; Zhu, Y.; An, J.; Ruoff, R.S. Graphene-Based Ultracapacitors. Nano Lett. 2008, 8, 3498–3502. [Google Scholar] [CrossRef] [PubMed]
  86. Bitla, Y.; Chin, Y.-Y.; Lin, J.-C.; Van, C.N.; Liu, R.; Zhu, Y.; Liu, H.-J.; Zhan, Q.; Lin, H.-J.; Chen, C.-T.; et al. Origin of metallic behavior in NiCo2O4 ferrimagnet. Sci. Rep. 2015, 5, 15201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Shi, X.; Bernasek, S.L.; Selloni, A. Formation, Electronic Structure, and Defects of Ni Substituted Spinel Cobalt Oxide: A DFT+U Study. J. Phys. Chem. C 2016, 120, 14892–14898. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction pattern for Ni1−xAlxCo2O4.
Figure 1. X-ray diffraction pattern for Ni1−xAlxCo2O4.
Processes 09 01750 g001
Figure 2. SEM images for Ni1−xAlxCo2O4 samples (a) x = 0, (b) x = 0.2, (c) x = 0.4, (d) x = 0.6, and (e) x = 0.8.
Figure 2. SEM images for Ni1−xAlxCo2O4 samples (a) x = 0, (b) x = 0.2, (c) x = 0.4, (d) x = 0.6, and (e) x = 0.8.
Processes 09 01750 g002
Figure 3. (a) Adsorption-desorption curves and inset pore volume distribution for Ni1−xAlxCo2O4, and (b) FTIR of Ni1−xAlxCo2O4.
Figure 3. (a) Adsorption-desorption curves and inset pore volume distribution for Ni1−xAlxCo2O4, and (b) FTIR of Ni1−xAlxCo2O4.
Processes 09 01750 g003
Figure 4. (ae) are cyclic voltammetry curves of Ni1−xAlxCo2O4, electrode obtained in the scan range of 5 mV/s to 300 mV/s measured in 3M KOH electrolytes. Figure (f) specific capacitance vs. scan rate.
Figure 4. (ae) are cyclic voltammetry curves of Ni1−xAlxCo2O4, electrode obtained in the scan range of 5 mV/s to 300 mV/s measured in 3M KOH electrolytes. Figure (f) specific capacitance vs. scan rate.
Processes 09 01750 g004
Figure 5. (ae) are cyclic stability curves of Ni1−xAlxCo2O4 measured up to 2000 cycles in 3M KOH electrolyte, (f) peak current vs. (scan rate)1/2, and (g) diffusion and capacitive contribution to the specific capacitance.
Figure 5. (ae) are cyclic stability curves of Ni1−xAlxCo2O4 measured up to 2000 cycles in 3M KOH electrolyte, (f) peak current vs. (scan rate)1/2, and (g) diffusion and capacitive contribution to the specific capacitance.
Processes 09 01750 g005
Figure 6. (ae) Charge-discharge curves of Ni1−xAlxCo2O4, electrode measured in the current density window of 0.75 to 30 A/g under 3 M KOH electrolytes. Figure (f) Comparison of specific capacitance as a function of current density.
Figure 6. (ae) Charge-discharge curves of Ni1−xAlxCo2O4, electrode measured in the current density window of 0.75 to 30 A/g under 3 M KOH electrolytes. Figure (f) Comparison of specific capacitance as a function of current density.
Processes 09 01750 g006
Figure 7. (a,b) Cyclic stability and Coulombic efficiency tested at 10 A/g and 5 A/g current density up to 5000 cycles in 3 M KOH electrolytes for Ni1−xAlxCo2O4, x = 0, and x = 0.8.
Figure 7. (a,b) Cyclic stability and Coulombic efficiency tested at 10 A/g and 5 A/g current density up to 5000 cycles in 3 M KOH electrolytes for Ni1−xAlxCo2O4, x = 0, and x = 0.8.
Processes 09 01750 g007
Figure 8. (a) Ragone plot of power density vs. energy density, (b) frequency-dependent real impedance measured, and (c) Nyquist plot for Ni1−xAlxCo2O4 at open circuit potential obtained in 3M KOH electrolytes.
Figure 8. (a) Ragone plot of power density vs. energy density, (b) frequency-dependent real impedance measured, and (c) Nyquist plot for Ni1−xAlxCo2O4 at open circuit potential obtained in 3M KOH electrolytes.
Processes 09 01750 g008
Figure 9. (a) Side view, along the x-axis of the crystal structure and (b) energy band-gap of Ni1−xAlxCo2O4 obtained from DFT calculations as a function of content, x.
Figure 9. (a) Side view, along the x-axis of the crystal structure and (b) energy band-gap of Ni1−xAlxCo2O4 obtained from DFT calculations as a function of content, x.
Processes 09 01750 g009
Table 1. Stoichiometry of chemicals used in the synthesis of Ni1−xAlxCo2O4.
Table 1. Stoichiometry of chemicals used in the synthesis of Ni1−xAlxCo2O4.
Ni1−xAlxCo2O4, xCo(NO3)2·6H2O
(gm)
Ni(NO3)2·6H2O (gm)Al(NO3)2·6H2O
(gm)
Urea (gm)
0.02.4201.2090.0001.497
0.22.4850.9930.3201.497
0.42.5500.7650.6581.497
0.62.6200.5251.0161.497
0.83.8500.3841.9851.497
Table 2. Surface area parameters derived from the adsorption-desorption curve, crystallite size obtained using Scherrer’s formula, lattice parameter obtained from XRD refinement, and Al3+ content was obtained from the XPS analysis of Ni1−xAlxCo2O4 compound.
Table 2. Surface area parameters derived from the adsorption-desorption curve, crystallite size obtained using Scherrer’s formula, lattice parameter obtained from XRD refinement, and Al3+ content was obtained from the XPS analysis of Ni1−xAlxCo2O4 compound.
Ni1−xAlxCo2O4, xAl Content Determined via XPSBET Surface Area (m2/g)BJH Surface Area (m2/g)BJH Avg. Pore Radius (nm)BJH Avg. Pore Volume (cc/g)Lattice Parameter, a (Å)Crystallite Size (nm)
0.00.0081.7489.091.4190.2398.098313.212
0.20.3783.16126.131.4230.1838.080416.525
0.40.57107.19120.261.5220.1928.066613.804
0.60.79106.29128.812.1440.1948.058514.776
0.80.89189.00354.831.5930.2768.044912.286
Table 3. Two distinct band positions from FTIR spectrum for Ni1−xAlxCo2O4 samples.
Table 3. Two distinct band positions from FTIR spectrum for Ni1−xAlxCo2O4 samples.
Ni1−xAlxCo2O4, xν1 (cm−1)ν2 (cm−1)
0.0534.2628.7
0.2544.0645.2
0.4551.0652.7
0.6553.6653.9
0.8556.1665.1
Table 4. Data of specific capacitance obtained from cyclic voltammetry, specific capacitance obtained from charge-discharge curves, energy density, and power density for Ni1−xAlxCo2O4.
Table 4. Data of specific capacitance obtained from cyclic voltammetry, specific capacitance obtained from charge-discharge curves, energy density, and power density for Ni1−xAlxCo2O4.
x, ContentSpecific Capacitance at 2 mv/s (F/g)Specific Capacitance at 0.5 A/g (F/g)Energy Density
(Wh/Kg)
Power Density
(W/Kg)
0.051226812.46316.6
0.23681948.96289.1
0.43711929.15281.2
0.63801768.76670.9
0.83561678.24660.4
Table 5. Comparison of specific capacitance of electrodes derived from doping of transition elements on NiCo2O4 using the hydrothermal method.
Table 5. Comparison of specific capacitance of electrodes derived from doping of transition elements on NiCo2O4 using the hydrothermal method.
Electrode MaterialScan Rate
(mV/s)
Specific Cap. (Csp)Current Density
(A/g)
Specific Cap.
(Csp)
Reference
Ni1−xRuxCo2O4
(x = 0.00, 0.02, 0.05, 0.10, 0.20)
---------1831 F/g, 913 F/g, 1004 F/g, 1527 F/g, 1483 F/g[27]
Ni1−xMnxCo2O4
(x =1.0, 0.7, 0.5, 0.3, 0.0)
---------1243 F/g (x = 0.3), 366 F/g (x = 0.5, max), 162 F/g (x = 0.7)[28]
Ni1−xZnxCo2O4
(x = 0.5)
-------21228 F/g[30]
Ni1−xFexCo2O4
(x = 1)
--------12237 F/g at 1 h,2208 F/g at 3 h, 1976 F/g at 5 h.[31]
Ni1−xCaxCo2O4 (x = 0.0, 0.2, 0.4, 0.6, 0.8)2506 F/g,508 F/g, 355 F/g, 934 F/g, 571 F/g0.5284 F/g, 365 F/g, 317 F/g, 558 F/g, 253 F/g[32]
Ni1−xAlxCo2O4 (x = 0.0, 0.2, 0.4, 0.6, 0.8)2512 F/g,368 F/g, 371 F/g, 380 F/g, 356 F/g0.5268 F/g, 194 F/g, 192 F/g, 176 F/g, 167 F/g[This study]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guragain, D.; Bhattarai, R.; Choi, J.; Lin, W.; Gupta, R.K.; Shen, X.; Perez, F.A.; Mishra, S.R. Electrochemical Performance of Aluminum Doped Ni1−xAlxCo2O4 Hierarchical Nanostructure: Experimental and Theoretical Study. Processes 2021, 9, 1750. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9101750

AMA Style

Guragain D, Bhattarai R, Choi J, Lin W, Gupta RK, Shen X, Perez FA, Mishra SR. Electrochemical Performance of Aluminum Doped Ni1−xAlxCo2O4 Hierarchical Nanostructure: Experimental and Theoretical Study. Processes. 2021; 9(10):1750. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9101750

Chicago/Turabian Style

Guragain, Deepa, Romakanta Bhattarai, Jonghyun Choi, Wang Lin, Ram Krishna Gupta, Xiao Shen, Felio A. Perez, and Sanjay R. Mishra. 2021. "Electrochemical Performance of Aluminum Doped Ni1−xAlxCo2O4 Hierarchical Nanostructure: Experimental and Theoretical Study" Processes 9, no. 10: 1750. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9101750

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