| Home | E-Submission | Sitemap | Contact Us |  
Environ Eng Res > Volume 30(3); 2025 > Article
Yudha, Putro, Haryadi, Paramitha, Paramitha, and Jumari: Nanostructured NiO derived from spent nickel catalyst as an efficient photocatalyst: Characterization, performance, and kinetic study

Abstract

This study aimed to recover and regenerate nanostructured nickel oxide (NiO) from a spent nickel catalyst (SNC) using the simple and economical hydrometallurgical method. The experiment was carried out by exploring Ni leaching kinetic models, such as the shrinking-core model (SCM), logarithmic rate law, and the Avrami model. HCl was used as a lixiviant, achieving a high 99% Ni leaching efficiency (LE) at a solid/liquid ratio of 100 g/L and temperature of 70°C The suitable mechanism for the Ni leaching from SNC using HCl solution is the Avrami model. The nanostructured NiO was successfully obtained by precipitation and sintering. Furthermore, XRD, SEM-EDX, TEM, Raman spectroscopy, and surface area analysis confirmed the formation of SNC-derived nanostructured-NiO (NiO-Cat) with needle-like nanosized particle size and a surface area of 77.2 m2/g. The NiO-Cat was examined for UV-irradiated photo-Fenton-like degradation activity on synthetic cationic and anionic dyes, including rhodamine B (RhB), methylene blue (MB), acid orange 7 (AO7), and methyl orange (MO), which showed degradation efficiency of 96%, 98%, 93%, and 83%, respectively, after 120 min of reaction. The overall process is economically and environmentally attractive. Meanwhile, the leaching and photocatalysis data can be further developed in the industry for equipment design.

Graphical Abstract

/upload/thumbnails/eer-2024-433f7.gif

1. Introduction

Nickel-based catalyst is widely used in the production of valuable chemicals in hydrocracking, hydrogenation, reforming, hydrotreating, and alkylation processes of petroleum and petrochemical industries. This widespread application has a significant effect on the final products of the overall production cost and economic value. Catalyst is often discarded in landfills when deactivated after prolonged use due to irreversible poisoning phenomena. Most countries consider spent nickel catalyst (SNC) to be hazardous materials, showing the need for rapid and careful handling to reduce the pollution risk. Moreover, some environmental policies state that the processing of SNC is mandatory due to the carcinogenic and leachable nature. This is because the potential harm to humans and other living organisms will significantly increase with exposure to high dosages of Ni species such as Ni2+ ions. SNC processing and recovery of valuable metals offer several advantages [1,2], including reduction of environmental issues and threats, as well as possible secondary sources of nickel and nickel-based products. In addition to catalyst regeneration, value-added products such as battery materials, capacitors, and catalyst can also be obtained.
Generally, the processing of SNC can be performed by pyrometallurgy, hydrometallurgy, or bio-metallurgy, as the most recent and eco-friendly method. Based on previous studies, hydrometallurgical method is the most feasible due to the effectiveness, scalability, and lower carbon footprints [3,4]. In hydrometallurgical processing, SNC is broken down and soluble by leaching agents [5,6], which can extract the targeted minerals, such as nickel, under mild conditions requiring no sophisticated apparatus. Although both alkaline and acid-leaching agents have been successfully used to extract nickel, acid-leaching is preferred due to the fast-leaching mechanism and low energy consumption. The leached metals are processed further to obtain new products, such as nickel salts and Li-ion battery material precursors [79].
The presence of synthetic dyes in wastewater streams of textile dyeing industries has become a serious environmental problem. Dyes such as rhodamine B (RhB), methylene blue (MB), MO, and acid orange 7 (AO7) are hazardous with carcinogenic, mutagenic, and toxic properties at some level [10,11]. These dyes have a high persistence toward natural degradation and can potentially risk the aquatic ecosystem and humans. To overcome the potential risk, several methods have been identified, including adsorption [12], biosorption [13,14], reverse osmosis/membrane technology [15], and photocatalyzed degradation [16]. Adsorption and reverse osmosis are the most preferred options due to the dye removal efficiency of approximately 99% through a simple separation mechanism [17,18]. However, the post-treatment of pollutants at higher concentrations requires extensive effort, leading to the generation of secondary pollution.
Biodegradation is another method that includes the use of living organisms to decompose dyes. However, the process requires heavy control of the operating conditions and a significant amount of time [19]. Advanced oxidation processes, such as photocatalytic degradation, are promising methods to rapidly treat synthetic organic pollutants in water. The heterogenous photocatalytic process has multiple advantages, such as an abundant catalyst source and the ability to ‘destroy’ the dye molecule into simple compounds such as carbon dioxide and water under mild operating conditions in a short amount of time. This method is sustainable due to its high recoverability and good cycle ability, which improve the feasibility of the waste treatment process [2022]. The photocatalysts can be selected from various metal oxides, including ZnO, CuO, WO3, TiO2, and NiO [2325], or mesoporous silica-based catalysts, including SBA-15, MCM-41, and MCM-48 [2628]. Nickel oxide or NiO has good stability (Ni2+), high availability, and efficiency, as well as the ability to generate hydroxyl and per hydroxyl radicals under low-intensity UV light irradiation. These radicals effectively break down the dye pollutant, eliminating the potential waste hazard. Nanostructured NiO can be obtained through several methods, including precipitation, flame spray pyrolysis, sol-gel, and combustion. Among these methods, precipitation is considered the most economically feasible for obtaining NiO at an industrial scale using water-soluble nickel salts as the precursor [29,30]. These nickel salts can be effectively obtained from SNC.
Based on the description, this study will focus on the kinetics of Ni leaching in detail to find a kinetic model that best describes the leaching process, determine the activation energy required, and describe the order of reaction applicable in the leaching of nickel from SNC catalysts. Kinetic data are crucial for designing reactors at pilot-plant or industrial scale. This work further emphasizes the preparation of nanostructured nickel oxide (NiO), starting from the leach solution, followed by its thorough characterization and evaluation as a heterogenous Fenton-like photocatalyst. The performance was analyzed through the decoloration/degradation of MB, RhB, MO, and AO7.SNC was treated by hydrometallurgy at a high pulp density, ensuring efficiency and less specific water consumption of the overall process. The properties of NiO were extensively investigated, such as crystal structure, morphology, chemical composition, photocatalytic, and optical performance. A conversion of nickel-based catalyst to photocatalyst is a facile method of waste handling and use, generating a product that can benefit the environment.

2. Materials and Method

2.1. Processing of SNC

In this study, SNC waste was obtained from an Indonesian petrochemical industry, specifically at the hydro-refining unit. The acid used was HCl (Hydrochloric acid, Merck, Germany) with a starting concentration of 37% v/v. The SNC was in extrudate form with a large grain size of > 250 μm. Before the destruction, the pre-dried SNC powders were ground and sieved using a 100-mesh screen (max. grain size = 149 μm). The 10 g of sieved SNC was leached in a flask for 120 minutes. The effect of the concentration of the lixiviant (1.25–6 M), the pulp density or solid-to-liquid (S/L) ratio (25, 50, 100, 150, and 200 gsnc/L), and the leaching temperature (50, 60, and 70°C) were examined at a rotation speed of 200 rpm. The filtrates were collected and analyzed using Atomic Absorption Spectroscopy (AAS) (Shimadzu, Japan) to understand SNC’s leaching behaviors in an HCl solution. The LE was calculated using the following equations (Eq. (1) and Eq. (2)):
(1)
LE=CCNi×100%
(2)
LE=MW×C×Vm×xt×100%
where C, CNi, MW, V, m, and xt represent the concentration of nickel, the total concentration of nickel, the molecular weight of nickel, the total HCl volume, the total solid mass, and nickel mass fraction in the solid. The leaching efficiency tests were performed three times.

2.2. Nickel Recovery and Synthesis of NiO

Each leachate was separated from the residue by gravity filtration and the pH was gradually increased to 9 using 10% of NH4OH (Merck, Germany). Subsequently, 4 M NaOH was added to the mixture to obtain a light green precipitate until the pH reached 12. The stirring speed and temperature of the reaction were kept at 200 rpm and 60°Cfor 2 hours. After aging, a green precipitate was filtered using standard gravity filtration. The collected nickel hydroxide (Ni(OH)2) precipitate was washed using demineralized water to a neutral pH and dried in a vacuum oven at 80°C. The Ni(OH)2 was labeled as Ni(OH)2-Cat. Ni(OH)2 synthesized using NiCl2·6H2O (Zenith, Brazil) were used as a comparison and labeled as Ni(OH)2-NiCl2. The dried Ni(OH)2 samples were fired in a muffle furnace under an air atmosphere for 3 hours at 500°C. The as-prepared nanostructured NiO samples were labeled as NiO-Cat and NiO-NiCl2 and the samples were stored in a dry environment for further analysis.

2.3. Material Characterization

2.3.1. Examination of SNC

Nickel content and the topological feature of SNC were investigated using Scanning Electron Microscope-Energy Dispersive X-ray spectroscopy (SEM-EDX)(JCM-7000, JEOL, Japan), meanwhile the X-ray diffraction (XRD) (D2-Phaser, Bruker, Germany) was used to investigate the structural and crystal phases of SNC.

2.3.2. Characterization of the as-prepared precursors and NiO powders

Both precursors and nanostructured NiO powders were also characterized using XRD, Fourier Transformed Infra-Red Spectroscopy (FTIR), SEM, and SEM-EDX. The surface area and pore analysis were conducted using the N2 isothermal adsorption-desorption method, performed at 77 K using Belsorp Mini X (Microtrac, Germany). Thermogravimetric Analysis/Differential Thermal Analysis (TG/DTA) was performed using Thermogravimetric Analyzers 60, Shimadzu, Japan. UV-visible spectra of nanostructured NiO powders were analyzed using Specord 200 Plus, Analytik Jena, Germany. The UV-Vis spectra were used to develop the Tauc-Plot. The Raman spectrum of nanostructured NiO powders were recorded by Raman Spectrometer (Horiba Scientific LabRam HR, Japan). The TEM imaging of nanostructured NiO samples were performed using Transmission Electron Microscopy (Thermofisher Scientific Fei Tecnai G2 Supertwin, USA).

2.4. Photocatalytic Experiments in Batch Operation

Photocatalytic activity test of nanostructured NiO samples was conducted by the photodegradation process of various pollutants such as RhB, MB, MO, and AO7 with an initial concentration of 20 mg/L. The UV irradiation source was obtained from a 2×8 W UV-C lamp, where the distance between the solution and the lamp was 12 cm. A total of 0.1–1.6 mg/mL of nanostructured NiO was dispersed in magnetically stirred pollutant solutions. The photodegradation processes were performed in a closed box. A 10 mL sample was collected and centrifuged at 3000 rpm for 10 minutes at various times to investigate the degradation process. The absorbance/UV-Vis spectra of samples were determined using UV-visible spectroscopy (Spectronic Genesys 150 Spectrophotometer, Thermofisher, USA) at the wavelength of 200–900 nm and a scan rate of 2 nm. The efficiency of pollutant decolorization due to the photodegradation process, or namely decolorization efficiency (DE), was calculated using the following Eq. (3):
(3)
DE=C0-CtC0×100%
where C0 and Ct are the initial and final dye pollutant concentrations, respectively. The photocatalyst performance tests were repeated and recorded three times.

3. Results and Discussion

3.1. Metal Leaching from SNC

3.1.1. Effect of HCl concentration

The characterization of SNC can be seen in Fig. S1, proving the presence of NiO [31,32]. Due to the acid strength of HCl, the leaching reaction of NiO is explained in Eq. (4):
(4)
NiO(s)+2HCl(aq)NiCl2(aq)+H2O(l)
The effect of HCl concentration was examined by varying the concentration of HCl in the leaching process at 60 minutes, 60°C, and the S/L value of 100 g/Llixiviant, with the result shown in Fig. 1a. The LEs increased along with HCl concentration, although the improvement at concentration above 3.75 M was not significant (from 89.02% to 90.78% and 92.9%). This showed that HCl quickly reacted with NiO, the stoichiometric value of HCl should at least be achieved, particularly when the S/L or pulp density value was high. Based on the Fig. 2a., 3.75 M HCl concentration was selected.

3.1.2. Effect of S/L ratio

High pulp density or S/L significantly affected the product throughput and attracted industries’ attention as a technological feasible process. In this study, the effect of S/L ratio during the leaching process was examined by maintaining the HCl concentration at 3.75 M and a temperature of 60°C for 60 minutes. Based on Fig. 1b., LE was decreased with the increasing S/L ratio. Generally, the acid consumption of pure NiO is theoretically two molHCl/molNiO or approximately 139 gNiO/L. At the S/L value above 100 g/L, LE dropped significantly due to acid depletion and the increasing slurry viscosity, leading to poor mass transfer. Therefore, the optimal S/L ratio to maintain the high product throughput was 100 g/L [3335].

3.1.3. Effect of temperature

The temperature effect during the SNC leaching was investigated using 3.75 M HCl with an S/L ratio of 100 g/L and a leaching time of 60 minutes. As shown in Fig. 1c., the LE increased along with the leaching temperature, with the highest value of 98.8% obtained at 70°C. The increasing LE followed the Arrhenius law, where the reaction rate increased along with leaching. Furthermore, the high leaching temperature would lead to a decrease in solution viscosity and better mass transfer. Moreover, the required threshold is limited to 70°C, where a significant increase will pose a challenge to parameter control [36].

3.1.4. Leaching behavior

The leaching kinetics were initially investigated by measuring LE at 0–120 minutes and multiple temperatures of 30, 50, 60, and 70°Cat the pre-determined S/L and HCl concentrations. Based on the trend shown in Fig. 1d., high reaction rates were achieved in the first 25 minutes. Subsequently, the reaction tended to slow down based on the insignificant increase of LE over time, showing an initially high HCl reactivity. Continuous consumption of HCl caused the reduction of [H+], lower concentration gradient, and increase in viscosity. The increase in leaching efficiency after 120 minutes is negligible due to the reduced concentration gradient, which slows down the reaction. Therefore, a leaching time of 60 minutes (1 hour) was selected. A shorter reaction time also suggests a lower residence time. On a pilot or larger scale, this reduction in residence time would decrease the reactor volume, leading to lower capital costs. A heterogeneous reaction phenomenon was analyzed using the SCM, logarithmic rate law, and Avrami model to understand the mechanism of Ni leaching from SNC and determine the suitable leaching kinetic equation [37,38].

3.2. Kinetics Models of Ni Leaching from SNC

The heterogeneous reaction between solids (SNC) and liquids (HCl) occurs during the process. This shows that the reaction occurred on the surface of the solid particles. As the solid Ni is converted to Ni ions, the size of the solid shrinks and leaves the unreacted part of the SNC to stay in the solid form. Therefore, the leaching mechanism can be expected to follow the unreacted SCM. The infamous SCM considers a 5-step mechanism during the leaching process. These include, first, the diffusion of the acid molecules from the bulk solution to the liquid-liquid interface. Second, the molecules diffuse from the interface to the solid surface. Third, a reaction occurs due to the reactivity between the compound in the solid surface and the acid molecules, producing soluble metal ions. Fourth, the as-produced metal ions penetrate the diffusion layer through the liquid-liquid interface. Fifth, the metal ions diffuse to the solution. These steps can be summarized into three control mechanism equations, namely(i) mass transfer of liquid reactants, (ii) surface chemical reaction, and (iii) diffusion process of the products through the ash/residue layer, which the formula is expressed in Eq. (5), Eq. (6), and Eq. (7) [37,39,40], respectively.
(5)
LE=k1t
(6)
-(1-LE)13=k2t
(7)
1-23LE-(1-LE)23=k3t
Previous studies have used several models to explain the leaching process of metal oxides using logarithmic rate law and the Avrami model, which are expressed in Eq. (8) and Eq. (9) [36,37], respectively:
(8)
(-ln(1-LE))2=k4t
(9)
ln(-ln(1-LE))=ln k5+n ln t
where LE is the LE of Ni (%), k is the reaction rate constant (min−1), t is the leaching time (min), and n is the Avrami parameter, with each equation plot, i.e. SCM-Mass transfer control, SCM-Surface reaction control, SCM-Diffusion control, logarithmic model, and Avrami model, presented in Fig. 2(a–e). The fitting parameter (R2) can be used to determine which model is suitable for explaining the leaching parameter, The closer the R2 value is to 1, the better the equation fits the data. The values of k and R2 of each plot used in this study are shown in Table 1. In the SCM, product diffusion through leaching residue appears to be the control mechanism with an R2 value of >0.95. The leaching residue morphology and quantitative analysis are presented in Fig. S2. and Table S1. The results show that Al and Si are present as the unreacted reactant, causing resistance in product diffusion. However, Avrami equation has the highest fitting parameters of >0.98, making Avrami model suitable for explaining the leaching mechanism of Ni from the SNC. This model describes the leaching mechanism as a dissolution process without forming a solid product, which is considered an analog of a reversed crystallization process. In this case, the solid nickel catalyst continuously reacts with HCl solution without forming solid or crystal nickel chloride salt [36]. The n value in the range of 0.5–1.0 shows that the reaction rate is initially high. Meanwhile, the rate decreases over time due to the reduction in concentration gradient and nickel from the SNC.
The activation energy can be evaluated by the Arrhenius equation expressed in Eq. (10) and Eq. (11):
(10)
k=Ae-EaRT
(11)
ln k=ln A-EaRT
where A is the Arrhenius constant (min−1), Ea is the activation energy (kJ/mol), R is the ideal gas constant (J/mol K), and T is the leaching process (K). Based on the Arrhenius plot in Fig. 2f., the A and Ea values for Avrami model are 28.8757 min−1 and 14.05 kJ/mol, as shown in Fig. S3., respectively. The Ea value less than 40 kJ/mol shows a fast chemical reaction, suggesting that the resistance can be neglected, in line with the previous leaching kinetic study [1,41]. Based on these results, the apparent nickel leaching rate law can be expressed in the following Eq. (1215):
(12)
1-LE=exp (k5tn)
(13)
1-LE=exp (A exp(-EaRT)tn)
(14)
LE=1-exp (A exp(-EaRT)tn)
(15)
LE=1-exp(28.8757exp(-1690.3/T)tn

3.3. Regeneration of Nanostructured NiO: Characterization

3.3.1. Characterization of Ni(OH)2

X-ray diffraction method was used to investigate the crystal structure of the as-prepared powders. Fig. 3a. shows the X-ray diffractogram of the Ni(OH)2 samples obtained from the spent catalyst leaching solution and the commercial nickel chloride through ammoniacal precipitation. Based on the results, both samples showed a crystalline phase due to the sharp peaks that were well indexed to the JCPDS No. 14-0117 or PDF card number 14–117 with no impurities detected [42,43]. This showed that the leaching process of SNC with aluminum supported only selectively precipitates nickel at pH 11–12 [44]. Furthermore, the Al ions remained in the solution due to the formation of soluble aluminate ions at elevated pH. The peaks (100) and (110) did not show line broadening, indicating the anisotropic property of β-Ni(OH)2 [42]. Using the Debye-Scherer equation, the crystallite size of Ni(OH)2-Cat and Ni(OH)2-NiCl2 at (011) was 6.4 nm and 9.9 nm, respectively. The smaller crystallite size of the Ni(OH)2-Cat compared to the Ni(OH)2-NiCl2 sample was due to the presence of Al ions. This prevented the formation of particle growth during the hydroxide precipitation through the disturbance of the Ostwald ripening mechanism [45]. Fig. 3(b-c). shows the morphology and elemental analysis of the as-prepared samples. The morphology of Ni(OH)2-Cat in Fig. 3b. (inset) and Fig. 3c. showed that the Ni(OH)2 particle had a quasi-spherical secondary particle with a diameter of 12 μm. The particle consisted of nano-sheets, similar to the previous study. The EDX mapping showed the homogenous distribution of Ni, Al, and O atoms, with Al atoms originating from the residue of Al2O3 support. Meanwhile, Fig. 3d. confirms the presence of a dense secondary particle and the absence of Al atoms. Fig. 3e. confirms the characteristic of the β-phase of precipitates through the infra-red spectra analysis with the presence of peaks at wavenumbers of 3600, 3400, and 1600 cm−1, showing the presence of stretching mode of disturbed and free hydroxyl groups. The peaks at the wavenumber of 1300 and 1000 cm−1 corresponded to the presence of carbonate species on the surface of the powders due to a side reaction between the hydroxyl group and CO2 gas from the humid open atmosphere [46]. Furthermore, Ni(OH)2-Cat showed a prominent broadening peak at 3425 cm−1 due to trapped water molecules. Bands detected at <600 cm−1 were correlated to the Ni-O stretching vibration mode, showing the presence of nickel hydroxide [47], while a band in <500 cm−1 was related to the Al-O mode [48]. The TG/DTA curve of the Ni(OH)2-Cat and Ni(OH)2-NiCl2 presented in Fig. 3f., showed two endothermic peaks attributed to the desorption of trapped water molecules at the temperature of below 200°C and the calcination temperature of 288°C which produced nanostructured NiO powder [41]. The results of qualitative and quantitative analysis in Table S2 are consistent with the XRD and FTIR.

3.3.2. Characterization of NiO

The nanostructured NiO powders obtained from the sintering of Ni(OH)2 samples were characterized Fig. 4 shows the characterization of NiO-Cat in comparison with NiO-NiCl2 by XRD, FTIR, Raman spectroscopy, Tauc-Plot from UV-Visible DRS, surface area analysis using Brunauer-Emmet-Teller (BET) method, SEM-EDX, and TEM. Meanwhile, Fig. 4a shows the X-ray diffraction patterns of the as-prepared and commercial samples. NiO-Cat and NiO-NiCl2 show a face-centered cubic structure of NiO, which is well indexed to the commercial NiO and the reference JCPDS or PDF No. 04-0835 [30,49]. Compared to commercial powder, NiO-Cat and NiO-NiCl2 have line broadening, which can be attributed to the small crystallite size. In addition, the comparison of the as-prepared NiO-Cat and NiO-NiCl2 provides strong evidence for product commercialization. Table S3. shows the structural parameters of the samples. The as-prepared nanostructured NiO powder crystallite sizes are significantly lower than in the commercial, showing the formation of nanostructure during hydroxide precipitation [50]. However, the lattice parameter does not show a significant difference between each sample.
Fig. 4b shows the infra-red spectra of the as-prepared NiO samples with Ni-O bond present in the wavenumber of <500 cm−1 and surface impurities. These include carbonate and hydroxide groups due to the adsorption of water and CO2 molecules from the atmosphere. NiO-Cat has a significant -OH group detected at 3420 cm−1 due to a higher amount of H2O adsorption than NiO-NiCl2 sample, showing a high surface area [51]. The Raman spectra in Fig. 4c. show slight pattern similarity between NiO-Cat and NiO-NiCl2, except with the peak broadening. A total of 4 modes can be identified from the nanosized NiO consisting of the first order of phonon modes at 300–700 cm−1 (LO and TO) and the second order at 700–1200 cm−1 (2LO and 2TO) [5254]. The Raman spectra of NiO-Cat have higher second-order phonon modes than NiO-NiCl2. UV-visible spectroscopy analysis of the as-prepared samples was conducted to obtain the Tauc-Plot presented in Fig. 4d. and Fig. 4e., respectively. Based on the calculation, the energy gap of NiO-Cat and NiO-NiCl2 were 3.50 and 3.91 eV, respectively, showing the presence of p-type semiconductor materials and their possible application for UV irradiated photocatalytic [55,56]. The result of the surface area and pore analysis of NiO-Cat are shown in Fig. 4f. The N2 adsorption and desorption isotherms behavior were related to the mesoporous characteristic of NiO-Cat with the BET surface area of 77.2 m2/g and the average pore radius of 3.18 nm, based on the Barret-Joyner-Halenda (BJH) method. Meanwhile, in Fig. 4g. NiO-NiCl2 has a BET surface area of 38.747 m2/g and an average pore radius of 3.342 nm. The large surface area of NiO-Cat is beneficial for catalyst application due to the presence of large active sites and promoting reaction rates, especially in heterogenous photocatalysis [57,58]. The BET surface area analysis result of NiO-Commercial can be seen in Fig. S4.
The morphology and elemental analysis of NiO samples were examined using SEM, as shown in Fig. 4(g–k). In comparison, the morphology feature of NiO-Cat in Fig. 4a. and Fig. 4b. are different from the Ni(OH)2 precursors in Fig. 3b. and Fig. 3d.. The decomposition of Ni(OH)2 into NiO lead to the change of nanosheet particle shape into a needle-like nanoparticle. The elemental analysis by EDX shows the presence of evenly distributed Ni, Al, and O elements. Fig. 4i. exhibits the HRTEM analysis of NiO-Cat sample and confirms nanosized particles with diameters 2–12 nm and a dspacing of 0.24 nm, as presented in Table S3. Moreover, the morphology and elemental analysis of NiO-NiCl2 in Fig. 4j. show a sphere-like shape of particles with high agglomeration levels but no Al elements. These results suggest that the as-prepared NiO-Cat has comparable characteristics to NiO made using fresh raw materials, namely NiO-NiCl2 [9]. Therefore, NiO-Cat is a promising material originating from SNC and can be used as a new catalyst for the photodegradation of pollutants.

3.4. Photodegradation of Dyes in a Photo-Fenton-like Process

The performance of nanostructured NiO-Cat was evaluated by photodegradation of organic pollutants, namely synthetic dyes comprised of cationic and anionic. RhB (C28H31ClN2O3) and MB (C16H18ClN3S) represented the cationic dyes. Meanwhile, AO7 (C16H11N2NaO4S) and MO (C14H14N3NaO3S) represented the anionic dyes.

3.4.1. Decolorization of cationic dyes

The photocatalysis parameters were first investigated on 20 mg/L RhB solution. The effect of H2O2, UV-C irradiation, and catalyst loading can be seen in Fig. 5(a–b). Fig. 5a. shows that UV-C itself can decolorize the RhB through photolysis. The addition of H2O2 increased the decolorization rate of RhB under UV-C irradiation. However, using NiO-Cat photocatalyst, the decolorization rate increased significantly, which is indicated by the increasing slope. In Fig. 5b. the catalyst loading increased when the catalyst loading was 1.6 g/L as a result of increasing active sites and photogenerated reactive compounds. However, at a higher value, the DE significantly dropped due to light scattering and shading, as we can see that the catalyst can cause turbidity, thus reducing the intensity of the light [59,60].
Fig. 5(c–f). shows the photocatalytic result of 20 mg/L RhB and MB using 1.6 mg/mL NiO-Cat and 10 mM H2O2 under UV-C irradiation. Based on the results, the absorbance peaks of both dyes are decreasing over time. This shows the successful decomposition of organic synthetic dyes using the as-prepared photocatalyst with RhB and MB DE of 81% and 88% after 70 minutes of photocatalytic process under UV-C, respectively, and DE above 98% after 120 minutes.

3.4.2. Decolorization of anionic dyes

Fig. 5(g–j). shows the photocatalytic result of 20 mg/L AO7 and MO using similar parameters, showing that absorbance peaks of both dyes are decreasing over time. This shows the successful decomposition of organic synthetic dyes using the as-prepared sample with AO7 and MO DE of 76% and 39% after 70 minutes of photocatalytic under UV-C, respectively. After 120 minutes, the DE for AO7 and MO is 92 and 83%, respectively.

3.4.3. Photocatalysis mechanism and kinetic

NiO-Cat can behave as a photocatalyst due to the gap energy, serving as a semiconducting material. Moreover, the nanostructured morphology and large surface area exhibited by NiO-Cat also contribute to the photocatalytic performance. Generally, electrons are capable of exciting at the edge of the conduction band (c) and recombining with the holes at the edge of the valence band (v). When holes are generated, their reaction with water molecules produces oxidizing radicals of hydroxyl (•OH). Subsequently, the radicals attack dye molecules, leading to decomposition into simpler molecules while losing color in the process. The formation of hydroxyl radicals can be accelerated by additional reducing agents such as H2O2 [25,59].
In this study, radical trapping analysis was carried out by adding several radical scavengers during the photodegradation process of 20 mg/L RhB with 1.6 g NiO/LRhBsolution and 10 mM H2O2 Generally, the active species for organic pollutant degradation consisted of photo-generated holes (h+), superoxide radicals (•O2), and hydroxyl radicals (•OH). However, isopropyl alcohol (IPA), disodium ethylenediaminetetraacetic acid (EDTA-2Na), and benzoquinone (BQ) were used during the RhB photodegradation process as the radical avengers for h+, •O2, and •OH, respectively. The results of radical scavenger addition with each concentration 10 mM were shown in Fig. 6a., suggesting a severe inhibition effect due to •OH scavenging with a 76% reduction of DE. The suppressed photocatalyst activity was also detected in the presence of EDTA-2Na and IPA. Therefore, the decolorization was highly affected by hydroxyl radicals [10], as expressed in Eq. (1620) and visualized in Fig. 6b.
(16)
NiO+hvNiO (h+V+e-C)
(17)
H2O+NiO(h+V)OH+H++NiO
(18)
H2O2+NiO(e-C)NiO+OH-+OH
(19)
H2O2+hv2OH
(20)
OH+dyes (complex molecules)simple molecules+H2O
The dye decomposition kinetic curves in Fig. 5(d,f,h,j). were constructed based on the PFO model (Eq. (21)) and half-time (Eq.(22)). The curves showed R2 above 90%, indicating a proper model to study the phenomena [10]. The pseudo-second-order (PSO) model (Eq. (2324)) result was shown in Fig. S5, while the degradation constant, R2, and half-time were presented in Table 2. The PFO and PSO models were applied due to the simplicity of the reaction mechanism, which is only affected by the reaction between pollutant molecules with the photogenerated reactive compounds or radicals [59].
(21)
ln (CtC0)=-kpfo×t
(22)
t12pfo=0.693kpfo
(23)
1Ct-1C0=kpso×t
(24)
t12pso=1C0×kpso
Ct is the dye concentration at various times, C0 is the initial dye concentration, kpfo and kPSO is the pseudo-first-order and pseudo-second-order dye degradation constant (min−1 and L/mg/min), and t1/2pfo and t1/2pso is the pseudo-first-order and pseudo-second-order reaction half-time (min). Based on the kinetic data in Table 2, the cationic dyes show large k and short t1/2 due to hydroxyl radicals and hydroxyl ions. Cationic dyes degrade quickly under alkaline conditions, while anionic dye degrades more slowly under the same conditions[11]. The R2 coefficient shows that the PFO kinetic model is suitable for dye degradation. The cycle performance of the photocatalyst shown in Fig. S6. provides a promising reusability of NiO-Cat.
Since radicals are heavily included in the process, the treated water was analyzed using atomic absorption spectroscopy to detect the presence of leached Ni in the sample. This was attributed to the harmful effects of Ni ions on the environment, particularly humans. Based on the analysis, the concentration of Ni was 4.2 ppb, which was significantly lower than the standardized water quality parameters determined by the Indonesian Ministry of Health (Kemenkes) of less than 70 μg/L or 70 ppb [61]. To further minimize the risk of Ni leaching, the addition of stable and eco-friendly coatings was suggested. Carbonaceous materials including graphene and reduced graphene oxide (rGO) have also shown promising potential to be applied as a co-doping or binary photocatalyst for future investigation. Based on the overall process, this study served as a step toward a new direction for sustainable technology. The effectiveness of the method used was comparable with the previous studies, as shown in Table S4. Moreover, the presence of H2O2 was compensated with the low-powered UV light used in the process [62].
The results showed the characterization and photocatalytic performance of the as-prepared NiO. Based on the overall process, the method showed promising potential to be considered suitable to simultaneously handle the chemical and textile industry waste, specifically SNC and dye-containing wastewater. The NiO-Cat is also promising to be applied during the natural organic matters (NOMs) content reduction in a wastewater treatment plant. On the other hand, the effect of free cations in the wastewater and the poisoning mechanism of the NiO-Cat still need to be further developed in the future. The economic aspect of the process was attractive for commercial development due to the inclusion of commercially available and cheap chemical reagents such as HCl, ammonium hydroxide, and sodium hydroxide. Based on the environmental impact, the conversion process generated less hazardous waste, such as catalyst support containing aluminosilicates and wastewater comprising ammonium and sodium chloride. The ammonium chloride in wastewater could be used as a fertilizer or converted into ammonium hydroxide solution through NaOH treatment.
A simple gate-to-gate analysis was performed, and the estimated specific energy consumption of the SNC conversion to NiO photocatalyst was 146 kWh/kgproduct. The emission factor of a coal power plant in Indonesia was found to be approximately 0.75–0.85 kg CO2e/kWh, leading to a carbon footprint of 106–120 kg CO2e/kgproduct. This value was significantly lower compared to the freshly prepared commercial NiO of ~369 kg CO2e/kg [63]. Moreover, the complete cradle-to-grave life cycle analysis of the method used in this study was an interesting topic that could be investigated in the future.

4. Conclusions

In conclusion, this study successfully explored the leaching process of SNC, regenerated NiO powders, and repurposed the powder as a photocatalyst. Based on the kinetics study, the Avrami model was the most suitable, as shown by the R2 value. The nanostructured NiO derived from the SNC was successfully synthesized and confirmed by XRD, FTIR, SEM-EDX, TEM, Raman, and surface area analysis. The optical band-gap energy of the as-prepared sample was 3.50 eV, which was considered a good p-type semiconductor photocatalyst. Furthermore, the analysis of photocatalytic activity was conducted by the photo-Fenton-like degradation of cationic and anionic dyes, namely RhB, MB, AO7, and MO. The results of DE of each dye showed that the photocatalytic of cationic dyes was more favorable compared to anionic dyes. The photodegradation of dyes followed the PFO kinetic model and was highly affected by the generation of hydroxyl radicals. This method was effective, efficient, and promising for handling SNC as well as organic pollutant processing applications. Moreover, data kinetics can be further used in equipment designs, especially for leaching reactors and photocatalysis reactors at the industrial level.

Supplementary Information

Acknowledgments

The authors are grateful for the financial support received from Lembaga Penelitian dan Pengabdian Masyarakat, Universitas Sebelas Maret, through Penelitian Unggulan Terapan Research Grant with contract number: 194.2/UN27.22/PT.01.03/2024. Furthermore, the authors are grateful to Prof. Hendri Widiyandari and Prof. Agus Purwanto for the fruitful discussion during this study.

Notes

Conflict of Interest Statement

The authors declare that they have no conflict of interest.

Author Contribution

C.S.Y. (Postgraduate) developed the methodology, performed experiments, data validation and investigation, resources management, and wrote the original draft. F.A.P. (Postgraduate) Developed the investigation and conceptualization, performed experiments, and wrote the final manuscript. H.H. (Associate Professor) performed supervision and data validation. T.P. (Ph.D. student) developed methodology and data visualization. T.P. (Postgraduate) conducted a review and editing of the manuscript and performed data visualization. A.J. (Associate Professor) performed supervision, reviewed the manuscript, and data validation.

References

1. Arslanoglu H, Yaras A. Recovery of precious metals from spent Mo-Co-Ni/Al2O3 catalyst in organic acid medium: Process optimization and kinetic studies. Pet Sci Technol. 2019;37:2081–2093. https://doi.org/10.1080/10916466.2019.1618867
crossref

2. Miceli M, Frontera P, Macario A, Malara A. Recovery/reuse of heterogeneous supported spent catalysts. Catalysts. 2021;1:1–17. https://doi.org/10.3390/catal11050591
crossref

3. Han Y, Chen J, Deng Y, et al. A leaching solvent extraction, stripping, precipitation, and calcination process for the recovery of MoO3 and NiO from spent hydrofining catalysts. Hydrometallurgy. 2023;218. https://doi.org/10.1016/j.hydromet.2023.106046
crossref

4. Jiang S, Hua H, Zhang L, et al. Environmental impacts of hydrometallurgical recycling and reusing for manufacturing of lithium-ion traction batteries in China. Sci Total Environ. 2022;811. https://doi.org/10.1016/j.scitotenv.2021.152224
crossref pmid

5. Le-Phuc N, Luong TN, Van Tran T, et al. Recycling nickel from spent catalyst of Phu My fertilizer plant as a precursor for exhaust gas treatment catalysts preparation. J. Mater. Cycles Waste Manag. 2017;19:1085–1092. https://doi.org/10.1007/10163-017-0609-1
crossref

6. Huang H, Tan M, Wang X, et al. Synthesis of Mesoporous ã-Alumina-Supported Co-Based Catalysts and Their Catalytic Performance for Chemoselective Reduction of Nitroarenes. ACS Appl. Mater. Interfaces. 2018;10:5413–5428. https://doi.org/10.1021/csami.7b14513
crossref pmid

7. Muzayanha SU, Yudha CS, Nur A, et al. A fast metals recovery method for the synthesis of lithium nickel cobalt aluminum oxide material from cathode waste. Metals. 2019;9: https://doi.org/10.3390/met9050615
crossref

8. Al-Mansi NM, Monem NMA. Recovery of nickel oxide from spent catalyst. Waste Manag. 2001;22:85–90. https://doi.org/0.1016/S0956-053X(01)00024-1
crossref

9. Nisa SS, Nurohmah AR, Yudha CS, et al. Preliminary Investigation of NiO Anode for NCA/NiO Battery from Spent Catalyst Recovery. IOP Conf. Ser. Mater. Sci. Eng. 2021;1096:12140. https://doi.org/10.1088/1757-899x/1096/1/012140
crossref

10. Widiyandari H, Prilita O, Al Ja’farawy MS, et al. Nitrogen-doped carbon quantum dots supported zinc oxide (ZnO/N-CQD) nanoflower photocatalyst for methylene blue photodegradation. Results Eng. 2023;17:100814. https://doi.org/10.1016/j.rineng.022.100814
crossref

11. Joulaee S, Mirzaei M, Hassanpour A, et al. Efficient removal of anionic and cationic dyes from waste water using green ZnO/NiO/graphene quantum dots nano photocatalyst. Optik. 2023;290:171324. https://doi.org/10.1016/j.ijleo.2023.171324
crossref

12. Chafidz A, Astuti W, Augustia V, et al. Removal of methyl violet dye via adsorption using activated carbon prepared from Randu sawdust (Ceiba pentandra). IOP Conf Ser Earth Environ Sci. 2018;167: https://doi.org/10.1088/1755-1315/167/1/012013
crossref

13. Gao Y, Chang H, Liang D, et al. Harnessing Prunella vulgaris L. residues for enhanced biosorption of dyes in wastewater. Chem. Eng. Res. Des. 2023;200:469–480. https://doi.org/10.1016/.cherd.2023.10.053
crossref

14. Safardastgerdi M, Doulati Ardejani F, Mahmoodi NM. Lignocellulosic biomass functionalized with EDTA dianhydride for removing Cu (II) and dye from wastewater: Batch and fixed-bed column adsorption. Miner. Eng. 2023;204:108423. https://doi.org/10.1016/j.mineng.2023.108423
crossref

15. Shah IA, Bilal M, Almanassra IW, et al. A comprehensive review of graphene oxide-based membranes for efficient dye removal from water sources. Sep. Purif. Technol. 2024;330:25277. https://doi.org/10.1016/j.seppur.2023.125277
crossref

16. Minisha S, Johnson J, Mohammad Wabaidur S, et al. Synthesis and Characterizations of Fe-Doped NiO Nanoparticles and Their Potential Photocatalytic Dye Degradation Activities. Sustainability. 2023;15. https://doi.org/10.3390/su151914552
crossref

17. Nizam NUM, Hanafiah MM, Mahmoudi E, et al. The removal of anionic and cationic dyes from an aqueous solution using biomass-based activated carbon. Sci Rep. 2021;11. https://doi.org/10.1038/s41598-021-88084-z
crossref

18. Abdelhamid AE, Elsayed AE, Naguib M, Ali EAB. Effective Dye Removal by Acrylic-Based Membrane Constructed from Textile Fibers Waste. Fibers Polym. 2023;24:2391–2399. https://oi.org/10.1007/s12221-023-00247-z
crossref pdf

19. Varjani S, Rakholiya P, Ng HY, et al. Microbial degradation of dyes: An overview. Bioresour Technol. 2020;314. https://doi.org/10.1016/j.biortech.2020.123728
crossref pmid

20. Singh A, Mishra BK. Removal of chlorhexidine digluconate from aqueous solution by heterogenous photocatalysis using sunlight-driven Ni-doped TiO2 material. Environ Eng Res. 2023;28. https://doi.org/10.4491/eer.2021.546
crossref

21. Luu HQ, Pham MT, Wang YF, You SJ. Enhancing NO degradation in visible light through plasma-treated photocatalytic substrates featuring TiO2@g-C3N4 Z-scheme structure. Environ Eng Res. 2024;29. https://doi.org/10.4491/eer.2023.560
crossref

22. Stiadi Y, Wendari TP, Zilfa , et al. Tuning the structural magnetic and optical properties of ZnO/NiFe2O4 heterojunction photocatalyst for simultaneous photodegradation of Rhodamine B and Methylene Blue under natural sunlight. Environ Eng Res. 2023; https://doi.org/10.4491/eer.2022.074
crossref

23. Haryadi , Purnama MRW, Wibowo A. C dots derived from waste of biomass and their photocatalytic activities. Indones. J. Chem. 2018;18:594–599. https://doi.org/10.22146/ijc.26652
crossref

24. Ahmaruzzaman M. Metal oxides (ZnO CuO and NiO)-based nanostructured materials for photocatalytic remediation of organic contaminants. Nanotechnol. Environ. Eng. 2023;8:219–235. https://doi.org/10.1007/s41204-022-00284-8
crossref

25. Dong C, Xing M, Zhang J. Recent Progress of Photocatalytic Fenton-Like Process for Environmental Remediation. Front. Environ. Chem. 2020;1:1–21. https://doi.org/10.4491/eer.2022.074
crossref

26. Castanheira B, Otubo L, Oliveira CLP, et al. Functionalized mesoporous silicas SBA-15 for heterogeneous photocatalysis towards CECs removal from secondary urban wastewater. Chemosphere. 2022;287. https://doi.org/10.1016/j.chemosphere.021.132023
crossref

27. Woszuk A, Panek R, Madej J, et al. Mesoporous silica material MCM-41: Novel additive for warm mix asphalts. Constr Build Mater. 2018;183:270–274. https://doi.org/10.1016/j.conbuildmat.2018.06.177
crossref

28. Ali NS, Salih IK, Harharah HN, et al. Utilization of Loaded Cobalt onto MCM-48 Mesoporous Catalyst as a Heterogeneous Reaction in a Fixed Bed Membrane Reactor to Produce Isomerization Product from n-Heptane. Catalysts. 2023;13. https://doi.org/10.3390/catal13071138
crossref

29. Ahmad W, Chandra Bhatt S, Verma M, et al. A review on current trends in the green synthesis of nickel oxide nanoparticles, characterizations, and their applications. Environ Nanotechnol Monit Manag. 2022; https://doi.org/10.1016/j.enmm.022.100674
crossref

30. Olajire AA, Mohammed AA. Green synthesis of nickel oxide nanoparticles and studies of their photocatalytic activity in degradation of polyethylene films. Adv. Powder Technol. 2020;31:211–218. https://doi.org/10.1016/j.apt.2019.10.012
crossref

31. Leyva C, Rana MS, Ancheyta J. Surface characterization of Al2O3-SiO2 supported NiMo catalysts: An effect of support composition. Catal. Today. 2008;130:345–353. https://doi.org/10.1016/j.cattod.2007.10.113
crossref

32. Campos-Molina MJ, Corral-Perez JJ, Mariscal R, Granados ML. Silica-poly(styrenesulphonic acid) nanocomposites as promising acid catalysts. Catal. Today. 2017;279:155–163. http://x.doi.org/10.1016/j.cattod.2016.06.042
crossref

33. Li L, Zhai L, Zhang X, et al. Recovery of valuable metals from spent lithium-ion batteries by ultrasonic-assisted leaching process. J. Power Sources. 2014;262:380–385. https://doi.org/1.1016/j.jpowsour.2014.04.013
crossref

34. Gupita LT, Purwanto A, Muzayanha SU, et al. Green hydrometallurgical route for recycling process of NCA cathode scrap. AIP Conf. Proc. 2020;2197:2–7. https://doi.org/10.1063/1.5140914
crossref

35. Chen WS, Ho HJ. Recovery of Valuable Metals from Lithium-Ion Batteries NMC Cathode Waste Materials by Hydrometallurgical Methods. Metals. 2018; https://doi.org/10.3390/met8050321
crossref

36. Zhuang L, Sun C, Zhou T, et al. Recovery of valuable metals from LiNi0.5Co0.2Mn0.3O2 cathode materials of spent Li-ion batteries using mild mixed acid as leachant. Waste Manag. 2019;85:175–185. https://doi.org/10.1016/j.wasman.2018.12.034
crossref pmid

37. Muzayanha SU, Yudha CS, Hasanah LM, et al. Comparative study of various kinetic models on leaching of nca cathode material. Indones. J. Chem. 2020;20:1291–1300. https://doi.org/0.22146/ijc.49412
crossref

38. Habbache N, Alane N, Djerad S, Tifouti L. Leaching of copper oxide with different acid solutions. Chem. Eng. J. 2009;152:503–508. https://doi.org/10.1016/j.cej.2009.05.020
crossref

39. He L, Sun S, Mu Y, et al. Recovery of Lithium Nickel Cobalt and Manganese from Spent Lithium-Ion Batteries Using l-Tartaric Acid as a Leachant. ACS Sustain Chem Eng. 2017;5–12. https://doi.org/10.1021/acssuschemeng.6b02056
crossref

40. Li L, Bian Y, Zhang X, et al. Process for recycling mixed-cathode materials from spent lithium-ion batteries and kinetics of leaching. Waste Manag. 2018;71:362–371. https://doi.org/10.1016/.wasman.2017.10.028
crossref pmid

41. Yao L, Yao H, Xi G, Feng Y. Recycling and synthesis of LiNi1/3Co1/3Mn1/3O2 from waste lithium ion batteries using d, l-malic acid. RSC Adv. 2016;6:17947–17954. https://doi.org/0.1039/c5ra25079j
crossref

42. Chu Q, Su Y, Xu Q, et al. Anisotropic strain induced fracture of Ni(OH)2 nanosheet electrodes for supercapacitors through a charge-ordered mixed-valent intermediate. Mater Sci Eng B. 2023;296. https://doi.org/10.1016/j.mseb.2023.116671
crossref

43. Li C, Xue Z, Qin J, et al. Synthesis of nickel hydroxide/delaminated-Ti3C2 MXene nanosheets as promising anode material for high performance lithium ion battery. J. Alloys Compd. 2020;842:155812. https://doi.org/10.1016/j.jallcom.2020.155812
crossref

44. Li Y, Pan G, Xu W, et al. Effect of Al substitution on the microstructure and lithium storage performance of nickel hydroxide. J Power Sources. 2016;307:114–121. https://doi.org/0.1016/j.jpowsour.2015.12.129
crossref

45. Li Y, Li X, Wang Z, et al. An Ostwald ripening route towards Ni-rich layered cathode material with cobalt-rich surface for lithium ion battery. Sci. China Mater. 2018;61:719–727. https://doi.org/10.1007/s40843-017-9162-3
crossref

46. Yudha CS, Muzayanha SU, Rahmawati M, et al. Fast production of high performance LiNi0.815Co0.15Al0.035O2 cathode material via urea-assisted flame spray pyrolysis. Energies. 2020;13. https://doi.org/10.3390/en13112757
crossref

47. Martinez-Sanchez C, Regmi C, Lee SW, Rodriguez-Gonzalez V. Effects of Ce Doping on the Photocatalytic and Electrochemical Performance of Nickel Hydroxide Nanostructures. Top. Catal. 2021;64:73–83. https://doi.org/10.1007/s11244-020-01295-y
crossref

48. Gonzalez-Gomez MA, Belderbos S, Yanez-Vilar S, et al. Development of superparamagnetic nanoparticles coated with polyacrylic acid and aluminum hydroxide as an efficient contrast agent for multimodal imaging. Nanomaterials. 2019;9: https://doi.org/10.3390/nano9111626
crossref pmid

49. Mohamed RM, Harraz FA. Photoreduction coupling of NiO/SiO2 nanocomposite with palladium and yttria nanoparticles: Visible-light-driven photocatalysts. Mater. Res. Bull. 2020;131:110965. https://doi.org/10.1016/j.materresbull.2020.10965
crossref

50. He K, Ruan Z, Teng X, Zhu Y. Facile synthesis and electrochemical properties of spherical LiNi0.85–xCo0.15AlxO2 with sodium aluminate via co-precipitation. Mater. Res. Bull. 2017;90:131–137. http://dx.doi.org/10.1016/j.materresbull.2017.1.039
crossref

51. Saputra RE, Astuti Y, Darmawan A. Hydrophobicity of silica thin films: The deconvolution and interpretation by Fouriertransform infrared spectroscopy. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018;199:12–20. https://doi.org/10.1016/j.saa.018.03.037
crossref pmid

52. Vivek P, Sivakumar R, Selva Esakki E, Deivanayaki S. Fabrication of NiO/RGO nanocomposite for enhancing photocatalytic performance through degradation of RhB. J. Phys. Chem. Solids. 2023;176:111255. https://doi.org/10.1016/j.jpcs.023.111255
crossref

53. Chandekar KV, Palanivel B, Alkallas FH, et al. Photocatalytic activities of Mg doped NiO NPs for degradation of Methylene blue dye for harmful contaminants: A kinetics mechanism and recyclability. J. Phys. Chem. Solids. 2023;178:111345. https://doi.org/10.1016/j.jpcs.2023.111345
crossref

54. Sheokand S, Kumar P, Jabeen S, Samra KS. 3D highly porous microspherical morphology of NiO nanoparticles for supercapacitor application. J. Solid State Electrochem. 2023;27:727–738. https://doi.org/10.1007/s10008-022-05366-w
crossref

55. Ezeh CV, Egbo KO, Musah JD, Yu KM. Wide gap p-type NiO-Ga2O3 alloy via electronic band engineering. J. Alloys Compd. 2023;932:167275. https://doi.org/10.1016/j.jallcom.2022.67275
crossref

56. Tong B, Ichimura M. Electrochemical Deposition of Transparent p-Type Semiconductor NiO. Electron Commun Jpn. 2018;101:5–50. https://doi.org/10.1002/ecj.12043
crossref

57. Navada KM, Nagaraja GK, D’Souza JN, et al. Phyto assisted synthesis and characterization of Scoparia dulsis L. leaf extract mediated porous nano CuO photocatalysts and its anticancer behavior. Appl. Nanosci. 2020;10:4221–4240. https://doi.org/10.1007/s13204-020-01536-2
crossref

58. Li G, Lian Z, Wan Z, et al. Ni(OH)2 clusters decorated NiTiO3 Z-scheme heterojunctions with strong O2 activation and robust interfacial built-in electric field for efficient NO photo-oxidation. Chem. Eng. J. 2023;451:3–12. https://doi.org/10.1016/j.cej.2022.138625
crossref

59. Betim FS, Marins AAL, Coelho ELD, et al. Evaluation of photocatalytic properties of zinc and cobalt mixed oxide recycled from spent Li-ion and Zn-MnO2 batteries in photo-Fenton-like process. Mater. Res. Bull. 2023;162:112179. https://doi.org/10.1016/j.materresbull.2023.112179
crossref

60. Darvishi Cheshmeh Soltani R, Khataee A, Mashayekhi M, Safari M. Photocatalysis of formaldehyde in the aqueous phase over ZnO/diatomite nanocomposite. Turk. J. Chem. 2016;40:402–411. https://doi.org/10.3906/kim-1507-54
crossref

61. Hartono B, Utomo SW, Kusnoputranto H, et al. Health Risk Analysis of Nickel and Lead Exposure in Drinking Water at Kawasi Village Obi Island South Halmahera District 2015. Asian J. Appl. Sci. 2018;6:1–7. https://www.ajouronline.com/index.php/AJAS/article/view/5199


62. Balarak D, Mostafapour FK. Photocatalytic degradation of amoxicillin using UV/Synthesized NiO from pharmaceutical wastewater. Indones. J. Chem. 2019;19:211–218. https://doi.org/10.22146/ijc.33837
crossref

63. Wei W, Samuelsson PB, Tilliander A, et al. Energy consumption and greenhouse gas emissions of nickel products. Energies. 2020;13. https://doi.org/10.3390/en13215664
crossref

Fig. 1
Leaching behavior of SNC using HCl at various (a) HCl concentrations, (b) S/L ratio, (c) leaching temperature, and (d) leaching time.
/upload/thumbnails/eer-2024-433f1.gif
Fig. 2
Leaching kinetic curve using (a) SCM-Mass transfer control, (b) SCM-Surface reaction control, (c) SCM-Diffusion control, (d) logarithmic model, (e) Avrami model, and (f) the Arrhenius plot of the Avrami model.
/upload/thumbnails/eer-2024-433f2.gif
Fig. 3
(a) X-ray diffractogram of Ni(OH)2-Cat and Ni(OH)-NiCl2, (b) SEM image of Ni(OH)2-Cat, (c) EDX-Mapping of Ni(OH)2 -Cat, (d) SEM-EDX Mapping of Ni(OH)2-NiCl2, (e) FTIR spectra of Ni(OH)2-Cat and Ni(OH)-NiCl2 and TG/DTA curve of (f) Ni(OH)2-Cat and (g) Ni(OH)-NiCl2.
/upload/thumbnails/eer-2024-433f3.gif
Fig. 4
Characterization of NiO catalysts derived from spent nickel catalyst (NiO-Cat) compared to NiO-NiCl2 and commercial NiO. (a) XRD patterns. (b) FTIR spectra. (c) Raman spectra. (d) UV-Vis DRS absorption spectra. (e) Tauc plots. (f,g) N2 adsorption-desorption isotherms for NiO-Cat and NiO-NiCl2, respectively. (h) SEM image of NiO-Cat (inset: higher magnification). (i) SEM-EDX elemental mapping of NiO-Cat for Ni, O, and Al. (j) TEM images of NiO-Cat at different magnifications, with particle size distribution histogram. (k) SEM image of NiO-NiCl2 with corresponding EDX elemental mapping for Ni and O.
/upload/thumbnails/eer-2024-433f4.gif
Fig. 5
Photocatalytic degradation of various dyes using NiO derived from spent nickel catalyst. (a) Degradation efficiency (DE%) towards RhB over time for different parameters. (b) Effect of photocatalyst mass loading on degradation efficiency. (c–j) Degradation kinetics and absorption spectra for individual dyes: (c,d) RhB, (e,f) MB, (g,h) AO7, and (MO).
/upload/thumbnails/eer-2024-433f5.gif
Fig. 6
(a) Photocatalytic decoloration of 20 mg/L RhB in the presence of multiple radical scavengers and (b) The photocatalytic mechanism of cationic/anionic dyes using NiO-Cat.
/upload/thumbnails/eer-2024-433f6.gif
Table 1
Kinetic model plot results.
Leaching Temperature (°C) Plot result

k1 (min−1) k2 (min−1) k3 (min−1) k4 (min−1) k5 (min−1) n
30 0.016
R2=0.877
0.0075
R2=0.923
0.0018
R2=0.9881
0.0226
R2=0.9288
0.1080
R2=0.9838
0.6378
50 0.019
R2=0.856
0.0097
R2=0.922
0.0028
R2=0.9869
0.0684
R2=0.9861
0.1542
R2=0.9927
0.6263
60 0.020
R2=0.840
0.0108
R2=0.915
0.0033
R2=0.9801
0.1007
R2=0.9656
0.1812
R2=0.9911
0.6204
70 0.022
R2=0.825
0.0108
R2=0.917
0.0041
R2=0.9795
0.1492
R2=0.9700
0.2154
R2=0.9954
0.6249
Table 2
Kinetic data evaluation based on a PFO and PSO model
Pollutant Dye Degradation Kinetics

Kpfo(min−1) t1/2pfo (min) kpso(L.(mg.min)−1) t1/2pso (min)
RhB 0.0257
R2=0.9932
11.713 0.007
R2=0.766
7.143
MB 0.0331
R2=0.9946
9.094 0.0128
R2=0.719
3.906
AO7 0.0223
R2=0.9744
13.499 0.0042
R2=0.7877
11.905
MO 0.0114
R2=0.0114
26.402 0.0013
R2=0.7060
38.462
Editorial Office
464 Cheongpa-ro, #726, Jung-gu, Seoul 04510, Republic of Korea
FAX : +82-2-383-9654   E-mail : eer@kosenv.or.kr

Copyright© Korean Society of Environmental Engineers.        Developed in M2PI
About |  Browse Articles |  Current Issue |  For Authors and Reviewers