Next Article in Journal
Unveiling the Role of In Situ Sulfidation and H2O Excess on H2S Decomposition to Carbon-Free H2 over Cobalt/Ceria Catalysts
Previous Article in Journal
Recent Progress in the Production of Cyanide-Converting Nitrilases—Comparison with Nitrile-Hydrolyzing Enzymes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Green Synthesis and Photocatalytic Dye Degradation Activity of CuO Nanoparticles

1
Institute of Chemistry, The Islamia University of Bahawalpur, Bahawalpur 63100, Pakistan
2
LAQV-REQUIMTE, Department of Chemistry, NOVA School of Science and Technology, Universidade NOVA de Lisboa, 2829-516 Caparica, Portugal
3
Department of Chemistry, Quaid-e-Azam University, Islamabad 44000, Pakistan
4
Department of Chemistry, University of Sahiwal, Sahiwal 57000, Pakistan
5
Department of Chemistry, College of Science, University of Jeddah, Jeddah 23218, Saudi Arabia
6
Department of Biochemistry, College of Science, University of Jeddah, Jeddah 23218, Saudi Arabia
7
State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
8
Department of Chemical Engineering, Faculty of Engineering, Khon Kaen University, Khon Kaen 40002, Thailand
9
Department of Zoology, College of Science, King Saud University, Riyadh 11362, Saudi Arabia
10
Department of Biological and Chemical Engineering, Aarhus University, Nørrebrogade 44, 8000 Aarhus C, Denmark
11
Department of Chemical Engineering, School of Mining, Metallurgy and Chemical Engineering, University of Johannesburg, P.O. Box 17011, Doornfontein 2028, South Africa
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(3), 502; https://doi.org/10.3390/catal13030502
Submission received: 20 January 2023 / Revised: 17 February 2023 / Accepted: 20 February 2023 / Published: 28 February 2023
(This article belongs to the Section Photocatalysis)

Abstract

:
The degradation of dyes is a difficult task due to their persistent and stable nature; therefore, developing materials with desirable properties to degrade dyes is an important area of research. In the present study, we propose a simple, one-pot mechanochemical approach to synthesize CuO nanoparticles (NPs) using the leaf extract of Seriphidium oliverianum, as a reducing and stabilizing agent. The CuO NPs were characterized via X-ray diffraction (XRD), scanning electron microscopy (SEM), photoluminescence (PL) and Fourier-transform infrared spectroscopy (FTIR). The photocatalytic activity of CuO NPs was monitored using ultraviolet-visible (UV-Vis) spectroscopy. The CuO NPs exhibited high potential for the degradation of water-soluble industrial dyes. The degradation rates for methyl green (MG) and methyl orange (MO) were 65.231% ± 0.242 and 65.078% ± 0.392, respectively. Bio-mechanochemically synthesized CuO NPs proved to be good candidates for efficiently removing dyes from water.

1. Introduction

The development of industrialization has increased the risk of environmental pollution [1,2]. Several types of waste can seriously threaten water bodies [3,4]. Organic dyes are significant pollutants produced by different industries, such as food, pharmaceutical, leather, textile, inks, cosmetics, etc. [5,6,7]. Every year, tons of complex dyes are formed and discharged into water bodies, which cause harmful effects on aquatic life [8,9,10]. Developing methods for degrading dyes is a challenge for researchers [11,12,13,14,15]. Some of these methods are ion exchange, chemical oxidation, ozonolysis, photocatalytic degradation, and coagulation-based techniques [16,17,18,19,20,21]. Adsorption is often used to remove several water pollutants [22,23,24,25,26,27,28,29,30,31,32,33,34]. However, photocatalytic degradation using a semiconductor photocatalyst is considered a cost-effective and green approach to successfully degrade dyes [35,36,37,38,39].
Light active materials with large surface areas are needed to degrade organic dyes [40,41]. Several metal oxides or semiconductors are active photocatalysts under sunlight, due to their small band gap [42,43]. CuO is a p-type semiconductor having a narrow band gap of 2.1 to 2.71 eV [44]. CuO nanoparticles (NPs) show a proper response towards optical, mechanical, and photolytic applications [42,45,46,47]. Different methods are used to synthesize CuO NPs, including sol-gel, solvothermal, microwave irradiation, hydrothermal, arc discharge, etc. [48,49].
The synthesis of CuO NPs using plant extracts is considered the most practical and green approach [50,51]. Different biological methods for synthesizing CuO NPs involve the use of bacteria, fungi, algae, and plants as bioactive materials by using leaves, flowers, or stem extracts, including Solanum americanum, Solanum nigram, Camellia japonica, Pterospermum acerifolium, Gum karaya, Soya bean, etc. [38,51,52,53].
Several efforts were made to understand the biological mechanism and phytochemicals involved in the green synthesis of CuO NPs and their characteristics [54]. Different biomolecules responsible for stabilizing and reducing nanoparticles can be amino acids, proteins/enzymes, alkaloids, polysaccharides, vitamins, and alcoholic compounds [55,56]. The prepared CuO NPs depend on the reduction power of ions and the reduction capability of plants having biochemicals such as polyphenols, enzymes, and other chelating agents [57,58].
Seriphidium oliverianum belongs to the family Asteraceae, widely used in folk medicines [59,60]. It has many active biomolecules, such as cardenolides, anthraquinones, tannins, alkaloids, flavonoids, terpenoids, phenolic acids, and carbohydrates [61]. The biological CuO NPs can be effectively reduced by biomolecules [62].
Catalysis-based processes play a crucial role in producing high-value goods such as fuel, chemicals, pharmaceuticals, etc., from cheap raw materials. Catalysts are considered the engines behind these processes [63]. It is projected that the catalysis-based sector may produce commodities worth several trillion euros annually, with a total sales value of catalytic materials of about 20 billion euros, highlighting the importance of catalysis to our community [64]. The chemical industry mainly uses heterogeneous catalysis for several reasons, such as simple catalyst separation, durability, and suitability for continuous operation. However, the designing of catalyst materials is not a straightforward process due to their complex architecture and the poor understanding of active centers [65]. This idea becomes exceptionally crucial for the consistent mass production of solid catalysts. Despite the advanced level of technology in this sector, catalyst synthesis is often seen as more of an art than a science. Therefore, a significant amount of research focuses on the creation of catalysts, which is an indication of the tremendous effort put into understanding the rational synthesis of active, selective, and stable catalysts. This idea is crucial for the consistent mass production of solid catalysts. Consequently, despite the advanced level of technology in this sector, catalyst synthesis is frequently seen as more of an art than a science. Therefore, it is not surprising that there is much research focusing on the creation of catalysts, as they attest to the tremendous effort put into understanding the logical synthesis of active, selective, and stable catalysts [66,67].
Precipitation, deposition–precipitation, the hydrothermal approach, and impregnation are the main pathways for synthesizing industrial-scale catalysts that are currently developed to ensure a reasonable level of control over the catalyst properties and performance [68]. Other methods such as solid-state reactions and fusing can also be used. However, solution-based procedures always generate a significant amount of solvent waste due to their inherent nature. In addition, nitrate or chloride metal salts are commonly used as precursors, which may result in the production of poisonous gases during subsequent calcination stages. Necessary measures may be required to prevent these gases from escaping into the atmosphere [69]. Wet chemistry procedures are also frequently viewed as tedious and challenging to scale up for a particular formulation of the catalytic material. Furthermore, due to their energy requirements and possible role as producers of hazardous waste, solution processes and additional treatment stages, carried out at high temperatures, frequently fail to fulfill current environmental standards. Thus, there is a large interest in creating alternative synthetic procedures that are less harmful to the environment, easier, more economical, more productive, and scalable.
Superior features of the generated materials and more advantageous economic or environmental factors of the processes can stimulate the creation of innovative synthetic techniques. Due to the growing urgency of environmental issues and energy depletion resources, environmentally friendly production techniques of catalyst synthesis are particularly advantageous [70]. Reactive extrusion and ball milling are two prominent, quick, and efficient mechanochemical processes to create catalytic materials.
In the last decades, the mechanochemical approach was developed as a sustainable method for the large-scale production of various nanomaterials [71]. This procedure can generate well-dispersed metal oxide nanoparticles to be used in wide-ranging applications, including environmental monitoring, energy storage, conversion, or biomedical uses. The mechanochemical synthesis is relatively simple, easy to scale-up and create a uniform reaction [71]. The motivation of work is that chemical reactions can proceed in the absence of excess solvents or heating, making this a key reason for the recent interest in green chemistry.
To the best of our knowledge, the mechanical synthesis of CuO NPs using Seriphidium oliverianum extract was not yet reported in the literature. In this study, we develop a straightforward bio-mechanochemical approach using an electric mortar grinder mill to synthesize CuO NPs in the presence of Seriphidium oliverianum leaf extract. We also evaluated the degradation efficiency of water-soluble dyes, namely methyl green (MG) and methyl orange (MO).

2. Results and Discussion

2.1. X-ray Diffraction (XRD)

XRD was employed to study and explore the crystalline nature of the CuO nanostructured material. The average grain size of the material was obtained using the Debye–Scherrer’s formula:
D = k λ / β cos θ
where “D” is the crystallite size (nm), “k” is Scherrer’s constant, equal to 0.98, “β” is full width at half maximum (FWHM), and “θ” is the angle of diffraction.
The calculated average crystallite size of the NPs is 12.44 nm. The PXRD diffractogram (Figure 1) displays several characteristics peaks of the monoclinic structure for CuO NPs (standard JCPDS data card no. 00-001-1117 [27]).

2.2. Scanning Electron Microscopy (SEM)

The accumulation of fine CuO nanoparticles originated aggregates. High surface area to volume ratio of nanoparticles provides very high surface energy. To minimize its surface energy, the nanoparticles tend to agglomerate. Uncontrolled agglomeration may occur due to attractive van der Waals forces between particles. The average grain size obtained for CuO NPs was 1.48 µm. Figure 2 shows that the green synthesis of CuO NPs produces small, aggregated particles.

2.3. Fourier Transform Infrared (FTIR)

FTIR also allowed us to examine the composition and functional groups of bio-mechanochemically synthesized CuO NPs, from 400 to 4000 cm−1. The strong vibrational bands found in the FTIR spectrum of CuO NPs (Figure 3) may be due to the biochemicals found in Seriphidium oliverianum extract (Figure S1), which capped the CuO NPs. Figure 3 shows a broad band at 3358 cm−1, which matches the hydroxyl functional group of alcoholic or phenolic compounds found on the NPs surface. Another FTIR band at 1616 cm−1 corresponds to the aromatic bending vibrational frequency of the alkene group (C=C). It may be due to the bio components of leaves, which play a role in the reduction and stabilization of NPs. The sharp band at 1352 cm−1 can be ascribed to frequencies of the C–H group of alkanes. The influential stretching band of C–O of the plant extract bio element alcoholic group is found at 1085 cm−1. The bending vibration band of the aromatic group appears at 834 cm−1. FTIR vibrational frequency ranges from 400 to 600 cm−1, being attributed to Cu–O linkage, which confirms the formation of CuO NPs [72]. The functional groups associated with phytochemicals of Seriphidium oliverianum include glycoalkaloid, tropane alkaloid, and atropine, assigned to hydroxyl, aromatic, phenolic, and amino groups, which confirm the role of the plant extract as a reducing agent in the CuO NP synthesis.

2.4. UV-Visible Spectroscopy

Figure 4 shows the UV-Vis spectrum of CuO NPs, from 300 nm to 550 nm. It displays the expected absorption band at 324 nm for CuO NPs, and the band at 334 nm, due to the interband transition of Cu metal core electrons [73]. Bio components found in plant extracts play a role in the synthesis of stable CuO NPs [74]. Some aspects such as reaction time, temperature, concentration of precursor salt and aqueous leaf extract, and morphology of nanoparticles have an impact on the location of the absorption band in the UV-Vis spectrum. The sharp band also shows a high concentration of nanoparticles.

2.5. Photoluminescence Spectroscopy (PL)

PL allows us to reveal more details about exterior interstices, oxygen vacancies, surface flaws, optical emission facts, and photochemical characteristics of photocatalytic CuO NPs. The fluorescent process can also study the separation and transportation of electrons, the recombination process, and their effects on photocatalysis. The suggested mechanism of photoluminescence includes the movement of an electron from the valence to the conduction band after energy absorption via the generation of a hole. The recombination process occurs by shifting back the electron to the valence band with a simultaneous emission of energy. Furthermore, the small-sized particles are attributed to excellent facet fault and oxygen vacancies, resulting in a sharp luminescent peak [75].
The PL spectra of CuO NPs carried at 300 nm and 350 nm are shown in Figure 5. Two well-defined peaks are situated at 421 nm and 597 nm, for a wavelength of 300 nm (Figure 5a). The band at 421 nm is assigned to band edge-free excitons, and the band at 597 nm is attributed to bound excitons. The strong peak of PL spectra may be allocated to the small particle size and exterior defects. The intense band is also related to a high recombination rate. In Figure 5b, two distinct bands located at 450 nm and 699 nm are obtained by using a wavelength of 350 nm. These bands differ from those obtained at 300 nm (Figure 5a) because the excitation takes place at different wavelengths. Electron transformation occurs on different energy levels by absorbing different radiant energies and recombining them back to the valence band, with different conditions. So, the PL spectra covers different ranges. The intensity of band edge-free and bound excitons was found to be higher at a wavelength of 350 nm than at 300 nm. This confirms the existence of a UV adsorption band in the range of 324–334 nm, as previously mentioned in Section 3.4.

2.6. A Plausible Mechanism for Biogenic CuO NPs

The bio reduction of the precursor salt starts instantly, and the formation of CuO NPs is demonstrated by the solution color changing from blue to dark brown. The biochemicals found in the plant extracts play a main role in the stabilization of CuO NPs [38].
It is expected that several functional groups, found in flavonoids, can be used as reductants and contribute to NP formation [76]. Moreover, the release of H atoms during the conversion of enol flavonoids into keto flavonoids reduces Cu ions into metal Cu NPs. Nevertheless, the precise mechanism for the synthesis of CuO NPs mediated by plant extracts is still unknown. It is believed that the depth of the nanoparticles’ color, in an open environment, after one hour, can be attributed to oxidation, which is responsible for the formation of CuO NPs. Many factors might be involved, for e.g., it is possible that oxidation occurs due to environmental oxygen or biochemicals binding reduced metal ions before stabilization. Given electrostatic attraction, the ions of the metal oxide bind together forming NPs that are stabilized to prevent cluster formation. Despite the lack of a clear understanding of the mechanism, the use of plant extracts to synthesize nanoparticles is a promising approach due to its safe, environmentally friendly, and cost-effective nature. Figure 6 shows the proposed mechanism for CuO NP synthesis.

2.7. Evaluation of Photocatalytic Activity

The photocatalytic degradation of MG and MO, in the presence of CuO NPs, is depicted in Figure 7. All parameters, namely irradiation time, light source, concentration of dyes, and catalyst were identical for all reactions. The degradation of dyes was evaluated using natural sunlight as the light source. The confined bandgap and high surface area significantly influenced the degradation activity. Absorption spectra were measured at regular intervals using a UV-Vis spectrometer for all experiments. It was found that the intensity of the absorption band decreased as the illumination time under sunlight increased.
The main factors responsible for the decolorization of dyes were hydroxyl and oxy radicals, which degrade toxic contaminants formed when a hole–electron pair was created. Furthermore, the color of dyes simultaneously became lighter with time. Degradation efficiency for MG and MO was 65% and 65%, respectively, after 60 min of exposure to sunlight in the presence of a photocatalyst, as shown in Figure 8, Table 1. The rate constant is different for the two dyes, as they have different compositions and react differently. The higher the rate constant, the faster the reaction rate and vice versa.

2.7.1. Kinetic Studies

In order to determine the degradation rate of organic compounds under optimal conditions, kinetic studies were conducted using the following relationship:
ln A o / A t = kt
In the equation, “Ao” is the absorbance of dyes at the time t = 0, “At” is the absorbance of dyes at time t, and “k” is the rate constant. Figure 9a shows the plot of Ao/At versus time, and Figure 9b shows ln Ao/At versus time. The slope of the graph represents the order of reaction (pseudo-first-order kinetics, which are attributable to the degradation of dyes) [77].

2.7.2. Mechanism

The primary species involved in detoxifying dyes under sunlight irradiation were identified via an analysis of the process. Figure 10 depicts the proposed mechanism for the photocatalytic degradation of dyes. Nanomaterials with a small bandgap promote the creation of hole–electron pairs, as low absorption energy is necessary for electrons to move between the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO). When a catalyst absorbs, the absorption intensity is equal to the bad gap energy. The electron movement occurs from the ground state to the excited state, creating a gap, valence band hole ( h VB + ), a free electron, and a conduction band electron ( e CB ). The positive hole is a promising candidate to accept an electron from the pollutants in order to degrade them. Highly oxidizing species convert the water molecules into hydroxyl radicals (OH˙), degrading the organic contaminants. Molecular oxygen combines with an electron and converts into a superoxide radical (O2˙) [78]. The process involved in the reduction of pollutants is given as follows:
Photocatalyst + h ν     e ( CB ) +   h ( VB ) +
It is believed that hollow spaces and hydroxyl radicals are the main reactive species during photocatalytic degradation of water-soluble dyes.

2.7.3. Parameters Affecting Photocatalytic Degradation

In order to achieve a remarkable photocatalytic activity, some features must be considered, such as the concentration of dyes and photocatalyst, the nature and structural morphology of the catalyst, irradiation time, and light source. The effect of the concentration of the dyes and catalyst loading on the degradation are given in Figures S2 and S3 of Supporting Information. The thermodynamic parameters such as ΔG°, ΔS°, and ΔH° are given in Figure S4 and Tables S1 and S2. The standard calculation deviation and R2 value are also given in Tables S3 and S4.
An inhibited reaction is observed when the concentration of dyes is increased, as there is no interaction with the active sites of the catalyst. This is due to the lack of absorption of light intensity by the dyes and the difficulty of electrons being able to migrate to the photocatalyst, resulting in insufficient hydroxyl radicals for dye degradation, and thus poor results are obtained [79]. MG and MO dye solution (with a 10-ppm concentration) showed much better results, compared to solutions with 15 or 20 ppm.
Upon increasing the photocatalyst amount, more interaction sites become available, leading to an enhanced production of hole–electron couples and hydroxyl radicals for the efficient detoxification of organic pollutants. The design, morphology, and particle size of the catalyst are important aspects in photocatalytic degradation. Smaller-sized particles provide more active sites for the absorption of dyes, increasing the efficiency of the process [80]. In our case, a good result was obtained using 10 mg CuO NPs. The removal rate of 65% was achieved after 60 min of illumination. 100% efficiency could be obtained if the exposure time was increased.
Although there is no report in the literature on the degradation of mechanochemical synthesized CuO NPs (used for the first time in the present work), other processes were used by several authors. Table 2 provides a detailed comparison between our synthesized photocatalyst and other CuO-based materials reported in the literature.
Concerning MO, UV light is more effective, as degradation rates varying from 45.23% [85] up to 96.4% [82] are obtained, using CuO NPs prepared via green synthesis. Using sunlight, the efficiencies vary from 31.95% [85] to 95% [86] for CuO NPs prepared via green synthesis, but are much smaller for CuO nanorods prepared via a hydrothermal method (22%) [87]. Our materials were the first prepared through mechanochemical synthesis and achieved a 65% degradation of MO using sunlight. This value is within the values reported in the literature for other preparation methods, but our method is simpler.
Concerning MG, no studies were found in the literature using CuO materials. A comparison is given for CuO-based materials used for the degradation of other dyes, showing the potential of this metal oxide.

2.7.4. Recyclability of Photocatalyst

The reusability of CuO NPs for the degradation of dyes was also analyzed. After complete degradation, the photocatalyst was removed and washed with deionized water. Then the photocatalyst was sonicated for a half an hour in 50 mL of deionized water and dried in air for 24 h. The CuO photocatalyst was then used in several consecutive degradation reactions. Excellent results were obtained for up to five runs, revealing the stability of CuO NPs (Figure S5). After the fifth run, the photocatalytic activity of the material decreased, possibly due to the formation of intermediates during the degradation of dyes, as suggested by other authors [100]. Additionally, the particle surface might decay during the degradation process, leading to a reduction in the overall activity after repeated use. However, the reusability of photocatalysts is a crucial factor in practical applications, as it allows for the effective and sustainable removal of dyes from wastewater.

3. Material and Methods

All chemical and precursor materials were purchased from Sigma-Aldrich (99.99%) and handled as acquired without any further treatment. Cu(NO3)2.3H2O was used as precursor material. All other chemicals were of analytical grade and used without any additional purification. Deionized water was utilized for the preparation of standard solutions.

3.1. Preparation of Seriphidium Oliverianum Leaf Extract

The leaves of Seriphidium oliverianum were washed with distilled water and dried in air for a few days. Dry leaves were mashed to form a powder, using a mortar and a pestle. 5 g of powder was dispersed in 50 mL of deionized water. After that, the dispersion was kept for 24 h and then heated at 70 °C for 30 min under continuous stirring, followed by filtration with filter paper (Whatman No. 1) twice to remove the suspended particles altogether. The obtained leaf extract was stored for further experiments. We used dry leaves because morphology, size, and shape may vary for fresh leaves [101]. Additionally, the amounts and types of flavonoid groups change, depending on thermal stability during leaf drying and extract preparation. Moreover, when fresh leaves are used, the UV spectra may not give a clear absorption band, compared to dry leaves.

3.2. Synthesis of CuO NPs

The required amount of copper nitrate was crushed into a fine powder, and 40 mL of plant extract was added ([Cu] = 0.1 M), followed by continuous grinding for 3 h in an electric mortar grinder mill (Model 911MPEMG100) at 70 rpm speed. The formation of CuO NPs was noticed by a color change of the solution (from blue to dull, dark brown). After that, the solution was placed in an open-air atmosphere for one hour, intensifying the color from light to dark brown. The mixture was centrifuged at room temperature for 30 min at 4000 rpm and washed with deionized water to eliminate the excesses of leaf extract or precursor salt. The obtained CuO NPs were collected in a Petri dish and air-dried. The complete process is depicted in Figure 11.

3.3. Photocatalytic Experiments

The photocatalytic activity of photocatalytic CuO NPs was tested in the degradation of different dyes, namely, methyl green (MG) and methyl orange (MO), used as reference models (Table S5). The used sunlight came from a visible light source (average solar flux = 500 km h−1m−2).
The reaction was initiated by adding 10 mg of CuO NPs to a 10-ppm solution (10 mgL−1) of each dye. The mixture was stirred for 30 min in the dark in order to establish the adsorption–desorption equilibrium. A well-established spectrum of UV-vis absorption was seen in all experiments. Different bands were analyzed in the UV-Vis spectra for MG and MO, at 632 nm and 462 nm, respectively. The solution was stirred under sunlight irradiation, and 2 mL of suspension was withdrawn every 10 min, up to 60 min, to observe the absorption peak, which was considered the absorption of dyes at the time “t”, and analyzed with UV-Vis.
The following equation was used to measure the dye degradation [29].
Degradation   ( % ) = ( A 0 A t A 0 ) × 100
where A0 is the absorbance at time = 0 and At is the absorbance at time = t.

3.4. Characterization

The optical characteristics of synthesized CuO NPs were analyzed via UV-vis spectroscopy (Cecil 7500 UV-Vis Spectrometer), from 295 to 550 nm. The structural and chemical composition of CuO NPs were characterized using a Fourier-transform infrared spectrophotometer (FTIR, Tensor 27) that had a vibrational frequency ranging from 400 to 4000 cm−1. A powder X-ray Diffractometer (PXRD) (Bruker D8 Advance PXRD) with Cu-Kα radiation source and wavelength λ equal to 1.540598 Å was employed to determine the crystallite size, nature, and phase description of CuO NPs [27]. In addition, CuO NPs morphology was investigated with scanning electron microscopy (SEM) using a MIRA-III TESCON apparatus. Surface deformity, photochemical, optical, and structural analysis of the obtained products were characterized using a Photoluminescence (PL) spectrometer at wavelengths ranging from 300 nm to 350 nm (Cary Eclipse Agilent technology) [28].

4. Conclusions

This study successfully reported an unprecedented environment-friendly bio mechanochemical approach for the synthesis of CuO NPs, using an aqueous extract from the Seriphidium oliverianum leaves. The bio components present in leaves were used as stabilizing and reducing agents. PXRD analysis identified a monoclinic CuO phase with a crystallite size of 12.44 nm. PL spectra identified two separate bands at 421 nm and 597 nm, indicating the presence of oxygen vacancies within the CuO NPs, which improved the photocatalytic activity. The synthesized nanosized material effectively demonstrated catalytic activity under sunlight illumination to degrade MG and MO dyes. The photocatalytic reduction followed a pseudo-first-order kinetics, with a removal rate of 65% for both MG and MO dyes. These promising results offered a new means for researchers to produce cost-effective and environmentally friendly photocatalysts to efficiently remove dyes from water.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13030502/s1, Figure S1. FTIR spectrum of Seriphidium oliverianum; Figure S2: Effect of different concentrations of dyes on the degradation process; Figure S3. Effect of different concentrations of photocatalyst on the degradation process; Figure S4. Thermodynamic study of dyes degradation; Figure S5. Dyes degradation upon recycling of the photocatalyst; Table S1. Thermodynamic parameters for MG degradation on CuO NPs; Table S2. Thermodynamic parameters for MO degradation on CuO NPs; Table S3. RMSE calculations for degradation of MG; Table S4. RMSE calculations for degradation of MO. Table S5. Dyes used in this work. Refs. [102,103,104,105] are cited in supplementary materials.

Author Contributions

Conceptualization: M.B.T.; investigation: W.A., Mika Sillanpaa, S.A.C.C.; software: R.Y.; formal analysis: A.R. and T.J.; writing—original draft: S.A.; writing—review and editing: F.V., M.S., K.K., S.A.-F. and S.A.C.C., supervision: M.B.T., S.A.C.C. and I.B.; resources: I.B.; project administration: M.B.T. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by Researchers Supporting Project Number (RSP-2023R7) King Saud University, Riyadh, Saudi Arabia. S.A.C.C. acknowledges support from FCT/MCTES (Fundação para a Ciência e Tecnologia and Ministério da Ciência, Tecnologia e Ensino Superior) through projects UIDB/50006/2020 and UIDP/50006/2020 and for the Scientific Employment Stimulus—Institutional Call (CEECINST/00102/2018). The authors thank The Islamia University of Bahawalpur for providing basic facilities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are present within the manuscript body.

Conflicts of Interest

There are no conflict of interest between the authors.

References

  1. Zafar, A.; Ullah, S.; Majeed, M.T.; Yasmeen, R. Environmental pollution in Asian economies: Does the industrialisation matter? OPEC Energy Rev. 2020, 44, 227–248. [Google Scholar] [CrossRef]
  2. Ukaogo, P.O.; Ewuzie, U.; Onwuka, C.V. Environmental pollution: Causes, effects, and the remedies. In Microorganisms for Sustainable Environment and Health; Elsevier: Amsterdam, The Netherlands, 2020; pp. 419–429. [Google Scholar]
  3. Hasnat, G.T.; Kabir, M.A.; Hossain, M.A. Major environmental issues and problems of South Asia, particularly Bangladesh. Handb. Environ. Mater. Manag. 2018, 2, 1–40. [Google Scholar]
  4. Reid, A.J.; Carlson, A.K.; Creed, I.F.; Eliason, E.J.; Gell, P.A.; Johnson, P.T.; Kidd, K.A.; MacCormack, T.J.; Olden, J.D.; Ormerod, S.J. Emerging threats and persistent conservation challenges for freshwater biodiversity. Biol. Rev. 2019, 94, 849–873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Hanafi, M.F.; Sapawe, N. A review on the water problem associate with organic pollutants derived from phenol, methyl orange, and remazol brilliant blue dyes. Mater. Today Proc. 2020, 31, A141–A150. [Google Scholar] [CrossRef]
  6. Pavithra, K.G.; Jaikumar, V. Removal of colorants from wastewater: A review on sources and treatment strategies. J. Ind. Eng. Chem. 2019, 75, 1–19. [Google Scholar] [CrossRef]
  7. Ismail, M.; Akhtar, K.; Khan, M.; Kamal, T.; Khan, M.A.; M Asiri, A.; Seo, J.; Khan, S.B. Pollution, toxicity and carcinogenicity of organic dyes and their catalytic bio-remediation. Curr. Pharm. Des. 2019, 25, 3645–3663. [Google Scholar] [CrossRef] [PubMed]
  8. Choudhary, M.; Peter, C.; Shukla, S.K.; Govender, P.P.; Joshi, G.M.; Wang, R. Environmental issues: A challenge for wastewater treatment. In Green Materials for Wastewater Treatment; Springer: Berlin/Heidelberg, Germany, 2020; pp. 1–12. [Google Scholar]
  9. Mehra, S.; Singh, M.; Chadha, P. Adverse impact of textile dyes on the aquatic environment as well as on human beings. Toxicol. Int. 2021, 28, 165–176. [Google Scholar]
  10. Maheshwari, K.; Agrawal, M.; Gupta, A. Dye Pollution in Water and Wastewater. In Novel Materials for Dye-Containing Wastewater Treatment; Springer: Berlin/Heidelberg, Germany, 2021; pp. 1–25. [Google Scholar]
  11. Nemiwal, M.; Zhang, T.C.; Kumar, D. Recent progress in g-C3N4, TiO2 and ZnO based photocatalysts for dye degradation: Strategies to improve photocatalytic activity. Sci. Total Environ. 2021, 767, 144896. [Google Scholar] [CrossRef]
  12. Reddy, C.V.; Reddy, K.R.; Harish, V.a.; Shim, J.; Shankar, M.; Shetti, N.P.; Aminabhavi, T.M. Metal-organic frameworks (MOFs)-based efficient heterogeneous photocatalysts: Synthesis, properties and its applications in photocatalytic hydrogen generation, CO2 reduction and photodegradation of organic dyes. Int. J. Hydrog. Energy 2020, 45, 7656–7679. [Google Scholar] [CrossRef]
  13. Ihsanullah, I.; Jamal, A.; Ilyas, M.; Zubair, M.; Khan, G.; Atieh, M.A. Bioremediation of dyes: Current status and prospects. J. Water Process Eng. 2020, 38, 101680. [Google Scholar] [CrossRef]
  14. Rashid, R.; Shafiq, I.; Akhter, P.; Iqbal, M.J.; Hussain, M. A state-of-the-art review on wastewater treatment techniques: The effectiveness of adsorption method. Environ. Sci. Pollut. Res. 2021, 28, 9050–9066. [Google Scholar] [CrossRef] [PubMed]
  15. Khan, S.; Bhardwaj, U.; Iqbal, H.M.; Joshi, N. Synergistic role of bacterial consortium to biodegrade toxic dyes containing wastewater and its simultaneous reuse as an added value. Chemosphere 2021, 284, 131273. [Google Scholar] [CrossRef] [PubMed]
  16. Ahmad, A.; Mohd-Setapar, S.H.; Chuong, C.S.; Khatoon, A.; Wani, W.A.; Kumar, R.; Rafatullah, M. Recent advances in new generation dye removal technologies: Novel search for approaches to reprocess wastewater. RSC Adv. 2015, 5, 30801–30818. [Google Scholar] [CrossRef]
  17. Piaskowski, K.; Świderska-Dąbrowska, R.; Zarzycki, P.K. Dye removal from water and wastewater using various physical, chemical, and biological processes. J. AOAC Int. 2018, 101, 1371–1384. [Google Scholar] [CrossRef] [PubMed]
  18. Kumar, P.S.; Joshiba, G.J.; Femina, C.C.; Varshini, P.; Priyadharshini, S.; Karthick, M.; Jothirani, R. A critical review on recent developments in the low-cost adsorption of dyes from wastewater. Desalin. Water Treat 2019, 172, 395–416. [Google Scholar] [CrossRef]
  19. Adane, T.; Adugna, A.T.; Alemayehu, E. Textile industry effluent treatment techniques. J. Chem. 2021, 2021, 5314404. [Google Scholar] [CrossRef]
  20. Ejraei, A.; Aroon, M.A.; Saravani, A.Z. Wastewater treatment using a hybrid system combining adsorption, photocatalytic degradation and membrane filtration processes. J. Water Process Eng. 2019, 28, 45–53. [Google Scholar] [CrossRef]
  21. Rao, S.; AS, S.; Jayaprakash, G.K.; Swamy, M.M.; K, S.; Kumar, D. Plant seed extract assisted, eco-synthesized C-ZnO nanoparticles: Characterization, chromium(VI) ion adsorption and kinetic studies. Luminescence, 2022; in press. [Google Scholar] [CrossRef]
  22. Carabineiro, S.A.C.; Thavorn-Amornsri, T.; Pereira, M.F.R.; Figueiredo, J.L. Adsorption of ciprofloxacin on surface-modified carbon materials. Water Res. 2011, 45, 4583–4591. [Google Scholar] [CrossRef]
  23. Carabineiro, S.A.C.; Thavorn-amornsri, T.; Pereira, M.F.R.; Serp, P.; Figueiredo, J.L. Comparison between activated carbon, carbon xerogel and carbon nanotubes for the adsorption of the antibiotic ciprofloxacin. Catal. Today 2012, 186, 29–34. [Google Scholar] [CrossRef]
  24. Silva, A.R.; Martins, P.M.; Teixeira, S.; Carabineiro, S.A.C.; Kuehn, K.; Cuniberti, G.; Alves, M.M.; Lanceros-Mendez, S.; Pereira, L. Ciprofloxacin wastewater treated by UVA photocatalysis: Contribution of irradiated TiO2 and ZnO nanoparticles on the final toxicity as assessed by Vibrio fischeri. Rsc Adv. 2016, 6, 95494–95503. [Google Scholar] [CrossRef]
  25. Chakraborty, R.; Asthana, A.; Singh, A.K.; Yadav, S.; Susan, M.A.; Carabineiro, S.A.C. Intensified elimination of aqueous heavy metal ions using chicken feathers chemically modified by a batch method. J. Mol. Liq. 2020, 312, 113475. [Google Scholar] [CrossRef]
  26. Yadav, S.; Asthana, A.; Chakraborty, R.; Jain, B.; Singh, A.K.; Carabineiro, S.A.C.; Susan, M.A. Cationic Dye Removal Using Novel Magnetic/Activated Charcoal/beta-Cyclodextrin/Alginate Polymer Nanocomposite. Nanomaterials 2020, 10, 170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Yadav, S.; Asthana, A.; Singh, A.K.; Chakraborty, R.; Vidya, S.S.; Susan, M.A.; Carabineiro, S.A.C. Adsorption of cationic dyes, drugs and metal from aqueous solutions using a polymer composite of magnetic/beta-cyclodextrin/activated charcoal/Na alginate: Isotherm, kinetics and regeneration studies. J. Hazard. Mater. 2021, 409, 124840. [Google Scholar] [CrossRef] [PubMed]
  28. Yadav, S.; Asthana, A.; Singh, A.K.; Chakraborty, R.; Vidya, S.S.; Singh, A.; Carabineiro, S.A.C. Methionine-Functionalized Graphene Oxide/Sodium Alginate Bio-Polymer Nanocomposite Hydrogel Beads: Synthesis, Isotherm and Kinetic Studies for an Adsorptive Removal of Fluoroquinolone Antibiotics. Nanomaterials 2021, 11, 568. [Google Scholar] [CrossRef] [PubMed]
  29. Lilhare, S.; Mathew, S.B.; Singh, A.K.; Carabineiro, S.A.C. Calcium Alginate Beads with Entrapped Iron Oxide Magnetic Nanoparticles Functionalized with Methionine-A Versatile Adsorbent for Arsenic Removal. Nanomaterials 2021, 11, 1345. [Google Scholar] [CrossRef] [PubMed]
  30. Martins, P.M.; Santos, B.; Salazar, H.; Carabineiro, S.A.C.; Botelho, G.; Tavares, C.J.; Lanceros-Mendez, S. Multifunctional hybrid membranes for photocatalytic and adsorptive removal of water contaminants of emerging concern. Chemosphere 2022, 293, 133548. [Google Scholar] [CrossRef]
  31. Lilhare, S.; Mathew, S.B.; Singh, A.K.; Carabineiro, S.A.C. Aloe Vera Functionalized Magnetic Nanoparticles Entrapped Ca Alginate Beads as Novel Adsorbents for Cu(II) Removal from Aqueous Solutions. Nanomaterials 2022, 12, 2947. [Google Scholar] [CrossRef]
  32. Chakraborty, R.; Asthana, A.; Singh, A.K.; Verma, R.; Sankarasubramanian, S.; Yadav, S.; Carabineiro, S.A.C.; Susan, M.A. Chicken feathers derived materials for the removal of chromium from aqueous solutions: Kinetics, isotherms, thermodynamics and regeneration studies. J. Dispers. Sci. Technol. 2022, 43, 446–460. [Google Scholar] [CrossRef]
  33. Arif, M.; Shahid, M.; Irfan, A.; Nisar, J.; Wang, X.; Batool, N.; Ali, M.; Farooqi, Z.H.; Begum, R. Extraction of copper ions from aqueous medium by microgel particles for in-situ fabrication of copper nanoparticles to degrade toxic dyes. Z. Für Phys. Chem. 2022, 236, 1219–1241. [Google Scholar] [CrossRef]
  34. Arif, M. Extraction of iron (III) ions by core-shell microgel for in situ formation of iron nanoparticles to reduce harmful pollutants from water. J. Environ. Chem. Eng. 2023, 11, 109270. [Google Scholar] [CrossRef]
  35. Hajiani, M.; Sayadi, M.H.; Mozafarjalali, M.; Ahmadpour, N. Green Synthesis of Recyclable, Cost-Effective, Chemically Stable, and Environmentally Friendly CuS@ Fe3O4 Nanoparticles for the Photocatalytic Degradation of Dye. J. Clust. Sci. 2022, 33, 1–13. [Google Scholar] [CrossRef]
  36. Ahmed, A.; Usman, M.; Yu, B.; Ding, X.; Peng, Q.; Shen, Y.; Cong, H. Efficient photocatalytic degradation of toxic Alizarin yellow R dye from industrial wastewater using biosynthesized Fe nanoparticle and study of factors affecting the degradation rate. J. Photochem. Photobiol. B Biol. 2020, 202, 111682. [Google Scholar] [CrossRef]
  37. Safajou, H.; Ghanbari, M.; Amiri, O.; Khojasteh, H.; Namvar, F.; Zinatloo-Ajabshir, S.; Salavati-Niasari, M. Green synthesis and characterization of RGO/Cu nanocomposites as photocatalytic degradation of organic pollutants in waste-water. Int. J. Hydrog. Energy 2021, 46, 20534–20546. [Google Scholar] [CrossRef]
  38. Aroob, S.; Taj, M.B.; Shabbir, S.; Imran, M.; Ahmad, R.H.; Habib, S.; Raheel, A.; Akhtar, M.N.; Ashfaq, M.; Sillanpää, M. In situ biogenic synthesis of CuO nanoparticles over graphene oxide: A potential nanohybrid for water treatment. J. Environ. Chem. Eng. 2021, 9, 105590. [Google Scholar] [CrossRef]
  39. Muhammad, T.; Alkahtani, M.; Raheel, A.; Shabbir, S.; Fatima, R.; Aroob, S.; Noor, S.; Ahmad, R.; Alshater, H. Bioconjugate Synthesis of NiFe2O4 Using Juglans Regia Leaves Extract: Phytochemical Analysis, Optical Activity, Removal of Ciprofloxacin and Congo Red From Water. 2020; Preprint from Research Square. [Google Scholar] [CrossRef]
  40. Mitra, M.; Ahamed, S.T.; Ghosh, A.; Mondal, A.; Kargupta, K.; Ganguly, S.; Banerjee, D. Polyaniline/reduced graphene oxide composite-enhanced visible-light-driven photocatalytic activity for the degradation of organic dyes. ACS Omega 2019, 4, 1623–1635. [Google Scholar] [CrossRef] [Green Version]
  41. Feng, S.; Li, F. Photocatalytic dyes degradation on suspended and cement paste immobilized TiO2/g-C3N4 under simulated solar light. J. Environ. Chem. Eng. 2021, 9, 105488. [Google Scholar] [CrossRef]
  42. Karthikeyan, C.; Arunachalam, P.; Ramachandran, K.; Al-Mayouf, A.M.; Karuppuchamy, S. Recent advances in semiconductor metal oxides with enhanced methods for solar photocatalytic applications. J. Alloy. Compd. 2020, 828, 154281. [Google Scholar] [CrossRef]
  43. Yadav, P.; Dwivedi, P.K.; Tonda, S.; Boukherroub, R.; Shelke, M.V. Metal and non-metal doped metal oxides and sulfides. In Green Photocatalysts; Springer: Berlin/Heidelberg, Germany, 2020; pp. 89–132. [Google Scholar]
  44. TAŞDEMİRCİ, T.Ç. Synthesis of copper-doped nickel oxide thin films: Structural and optical studies. Chem. Phys. Lett. 2020, 738, 136884. [Google Scholar] [CrossRef]
  45. Naseem, T.; Durrani, T. The role of some important metal oxide nanoparticles for wastewater and antibacterial applications: A review. Environ. Chem. Ecotoxicol. 2021, 3, 59–75. [Google Scholar] [CrossRef]
  46. Pourmoslemi, S.; Bayati, N.; Mahjub, R. Application of Box–Behnken design to optimize a sol-gel synthesis method for Ag and Zn doped CuO nanoparticles with antibacterial and photocatalytic activity. J. Sol-Gel Sci. Technol. 2022, 104, 319–329. [Google Scholar] [CrossRef]
  47. Elysabeth, T.; Mulia, K.; Ibadurrohman, M.; Dewi, E.L. A comparative study of CuO deposition methods on titania nanotube arrays for photoelectrocatalytic ammonia degradation and hydrogen production. Int. J. Hydrog. Energy 2021, 46, 26873–26885. [Google Scholar] [CrossRef]
  48. Forouzandeh, P.; Ganguly, P.; Dahiya, R.; Pillai, S.C. Supercapacitor electrode fabrication through chemical and physical routes. J. Power Sources 2022, 519, 230744. [Google Scholar] [CrossRef]
  49. Khairy, M.; El-Shaarawy, M.; Mousa, M. Characterization and super-capacitive properties of nanocrystalline copper ferrite prepared via green and chemical methods. Mater. Sci. Eng. B 2021, 263, 114812. [Google Scholar] [CrossRef]
  50. Gebremedhn, K.; Kahsay, M.H.; Aklilu, M. Green synthesis of CuO nanoparticles using leaf extract of catha edulis and its antibacterial activity. J. Pharm. Pharmacol. 2019, 7, 327–342. [Google Scholar] [CrossRef] [Green Version]
  51. Galan, C.R.; Silva, M.F.; Mantovani, D.; Bergamasco, R.; Vieira, M.F. Green synthesis of copper oxide nanoparticles impregnated on activated carbon using Moringa oleifera leaves extract for the removal of nitrates from water. Can. J. Chem. Eng. 2018, 96, 2378–2386. [Google Scholar] [CrossRef]
  52. El-Batal, A.I.; El-Sayyad, G.S.; Mosallam, F.M.; Fathy, R.M. Penicillium chrysogenum-mediated mycogenic synthesis of copper oxide nanoparticles using gamma rays for in vitro antimicrobial activity against some plant pathogens. J. Clust. Sci. 2020, 31, 79–90. [Google Scholar] [CrossRef]
  53. Bukhari, S.I.; Hamed, M.M.; Al-Agamy, M.H.; Gazwi, H.S.; Radwan, H.H.; Youssif, A.M. Biosynthesis of copper oxide nanoparticles using Streptomyces MHM38 and its biological applications. J. Nanomater. 2021, 2021, 6693302. [Google Scholar] [CrossRef]
  54. Kumar, J.A.; Krithiga, T.; Manigandan, S.; Sathish, S.; Renita, A.A.; Prakash, P.; Prasad, B.N.; Kumar, T.P.; Rajasimman, M.; Hosseini-Bandegharaei, A. A focus to green synthesis of metal/metal based oxide nanoparticles: Various mechanisms and applications towards ecological approach. J. Clean. Prod. 2021, 324, 129198. [Google Scholar] [CrossRef]
  55. Mustapha, T.; Misni, N.; Ithnin, N.R.; Daskum, A.M.; Unyah, N.Z. A Review on Plants and Microorganisms Mediated Synthesis of Silver Nanoparticles, Role of Plants Metabolites and Applications. Int. J. Environ. Res. Public Health 2022, 19, 674. [Google Scholar] [CrossRef] [PubMed]
  56. Waris, A.; Din, M.; Ali, A.; Ali, M.; Afridi, S.; Baset, A.; Khan, A.U. A comprehensive review of green synthesis of copper oxide nanoparticles and their diverse biomedical applications. Inorg. Chem. Commun. 2021, 123, 108369. [Google Scholar] [CrossRef]
  57. El Shafey, A.M. Green synthesis of metal and metal oxide nanoparticles from plant leaf extracts and their applications: A review. Green Process. Synth. 2020, 9, 304–339. [Google Scholar] [CrossRef]
  58. Siddiqi, K.S.; Husen, A. Current status of plant metabolite-based fabrication of copper/copper oxide nanoparticles and their applications: A review. Biomater. Res. 2020, 24, 11. [Google Scholar] [CrossRef] [PubMed]
  59. Rehman, A.; Saeed, S.; Ahmed, A. Genetic diversity and population structure of Seriphidium Sub-genus of Artemisia from different terrains of Balochistan, Pakistan. Biodiversitas J. Biol. Divers. 2021, 22, 1826–1834. [Google Scholar] [CrossRef]
  60. Shafiq, N.; Shafiq, S.; Rafiq, N.; Parveen, S.; Javed, I.; Majeed, H.N.; Mahmood, A.; Noor, N.; Anjum, A. Phytochemicals of the Seriphidium, economically and pharmaceutically important genus of Asteraceae family. Mini-Rev. Org. Chem. 2020, 17, 158–168. [Google Scholar] [CrossRef]
  61. Abbas, A.; Naqvi, S.A.R.; Rasool, M.H.; Noureen, A.; Mubarik, M.S.; Tareen, R.B. Phytochemical analysis, antioxidant and antimicrobial screening of seriphidium oliverianum plant extracts. Dose-Response 2021, 19, 15593258211004739. [Google Scholar] [CrossRef]
  62. Hano, C.; Abbasi, B.H. Plant-based green synthesis of nanoparticles: Production, characterization and applications. Biomolecules 2021, 12, 31. [Google Scholar] [CrossRef]
  63. Fechete, I.; Wang, Y.; Védrine, J.C. The past, present and future of heterogeneous catalysis. Catal. Today 2012, 189, 2–27. [Google Scholar] [CrossRef]
  64. Munnik, P.; De Jongh, P.E.; De Jong, K.P. Recent developments in the synthesis of supported catalysts. Chem. Rev. 2015, 115, 6687–6718. [Google Scholar] [CrossRef]
  65. Copéret, C.; Allouche, F.; Chan, K.W.; Conley, M.P.; Delley, M.F.; Fedorov, A.; Moroz, I.B.; Mougel, V.; Pucino, M.; Searles, K. Bridging the Gap between Industrial and Well-Defined Supported Catalysts. Angew. Chem. Int. Ed. 2018, 57, 6398–6440. [Google Scholar] [CrossRef]
  66. Schüth, F. Control of solid catalysts down to the atomic scale: Where is the limit? Angew. Chem. Int. Ed. 2014, 53, 8599–8604. [Google Scholar] [CrossRef] [PubMed]
  67. Schüth, F. Strukturierung fester Katalysatoren bis hinab zur atomaren Skala: Wo liegen die Grenzen? Angew. Chem. 2014, 126, 8741–8747. [Google Scholar] [CrossRef]
  68. Gallei, M.; Schwab, E. Handbook of Heterogenous Catalysis. Ertl. G 2008, 4, 57–66. [Google Scholar]
  69. Amrute, A.P.; De Bellis, J.; Felderhoff, M.; Schüth, F. Mechanochemical synthesis of catalytic materials. Chem. A Eur. J. 2021, 27, 6819–6847. [Google Scholar] [CrossRef]
  70. Karge, H.G. Handbook of Heterogeneous Catalysis; Ertl, G., Knozinger, H., Schuth, F., Eds.; VCH: Weinheim, Germany, 2008. [Google Scholar]
  71. Tsuzuki, T. Mechanochemical synthesis of metal oxide nanoparticles. Commun. Chem. 2021, 4, 143. [Google Scholar] [CrossRef]
  72. Bonnia, N.N.; Fairuzi, A.A.; Akhir, R.M.; Yahya, S.M.; Rani, M.A.A.; Ratim, S.; Rahman, N.A.; Akil, H.M. Comparison study on biosynthesis of silver nanoparticles using fresh and hot air oven dried IMPERATA CYLINDRICA leaf. IOP Conf. Ser. Mater. Sci. Eng. 2017, 290, 012002. [Google Scholar] [CrossRef]
  73. Usha, V.; Kalyanaraman, S.; Thangavel, R.; Vettumperumal, R. Effect of catalysts on the synthesis of CuO nanoparticles: Structural and optical properties by sol–gel method. Superlattices Microstruct. 2015, 86, 203–210. [Google Scholar] [CrossRef]
  74. Sharma, J.K.; Akhtar, M.S.; Ameen, S.; Srivastava, P.; Singh, G. Green synthesis of CuO nanoparticles with leaf extract of Calotropis gigantea and its dye-sensitized solar cells applications. J. Alloy. Compd. 2015, 632, 321–325. [Google Scholar] [CrossRef]
  75. Saif, S.; Tahir, A.; Asim, T.; Chen, Y. Plant mediated green synthesis of CuO nanoparticles: Comparison of toxicity of engineered and plant mediated CuO nanoparticles towards Daphnia magna. Nanomaterials 2016, 6, 205. [Google Scholar] [CrossRef] [Green Version]
  76. Bhardwaj, R.; Bharti, A.; Singh, J.P.; Chae, K.H.; Goyal, N. Influence of Cu doping on the local electronic and magnetic properties of ZnO nanostructures. Nanoscale Adv. 2020, 2, 4450–4463. [Google Scholar] [CrossRef]
  77. Shreyash, N.; Bajpai, S.; Khan, M.A.; Vijay, Y.; Tiwary, S.K.; Sonker, M. Green synthesis of nanoparticles and their biomedical applications: A review. ACS Appl. Nano Mater. 2021, 4, 11428–11457. [Google Scholar] [CrossRef]
  78. Abebe, B.; Murthy, H.; Amare, E. Summary on Adsorption and Photocatalysis for Pollutant Remediation: Mini Review. J. Encapsulation Adsorpt. Sci. 2018, 8, 225–255. [Google Scholar] [CrossRef] [Green Version]
  79. Zhao, J.; Chen, C.; Ma, W. Photocatalytic degradation of organic pollutants under visible light irradiation. Top. Catal. 2005, 35, 269–278. [Google Scholar] [CrossRef]
  80. Kumar, A.P.; Bilehal, D.; Tadesse, A.; Kumar, D. Photocatalytic degradation of organic dyes: Pd-γ-Al2O3 and PdO-γ-Al2O3 as potential photocatalysts. RSC Adv. 2021, 11, 6396–6406. [Google Scholar] [CrossRef] [PubMed]
  81. Sharma, S.; Kumar, K. Aloe-vera leaf extract as a green agent for the synthesis of CuO nanoparticles inactivating bacterial pathogens and dye. J. Dispers. Sci. Technol. 2021, 42, 1950–1962. [Google Scholar] [CrossRef]
  82. Sharma, S.; Kumar, K.; Thakur, N.; Chauhan, S.; Chauhan, M.S. Eco-friendly Ocimum tenuiflorum green route synthesis of CuO nanoparticles: Characterizations on photocatalytic and antibacterial activities. J. Environ. Chem. Eng. 2021, 9, 105395. [Google Scholar] [CrossRef]
  83. Pourmortazavi, S.M.; Rahimi-Nasrabadi, M.; Ahmadi, F.; Ganjali, M.R. CuCO3 and CuO nanoparticles; facile preparation and evaluation as photocatalysts. J. Mater. Sci. Mater. Electron. 2018, 29, 9442–9451. [Google Scholar] [CrossRef]
  84. Mageshwari, K.; Sathyamoorthy, R.; Park, J. Photocatalytic activity of hierarchical CuO microspheres synthesized by facile reflux condensation method. Powder Technol. 2015, 278, 150–156. [Google Scholar] [CrossRef]
  85. Ikram, A.; Jamil, S.; Fasehullah, M. Green Synthesis of Copper Oxide Nanoparticles from Papaya/Lemon tea Extract and its Application in Degradation of Methyl Orange. Mater. Innov. 2022, 2, 115–122. [Google Scholar] [CrossRef]
  86. Kayalvizhi, S.; Sengottaiyan, A.; Selvankumar, T.; Senthilkumar, B.; Sudhakar, C.; Selvam, K. Eco-friendly cost-effective approach for synthesis of copper oxide nanoparticles for enhanced photocatalytic performance. Optik 2020, 202, 163507. [Google Scholar] [CrossRef]
  87. Sali, R.K.; Pujar, M.S.; Nadgir, A.; Sidarai, A.H. Photocatalytic activities of CuO nanorods over methyl orange: Hydrothermal method. In AIP Conference Proceedings, Coimbatore, India, 17–18 July 2020; AIP Publishing LLC: Baltimore, MD, USA, 2020. [Google Scholar]
  88. Zeid, E.F.A.; Ibrahem, I.A.; Mohamed, W.A.A.; Ali, A.M. Study the influence of silver and cobalt on the photocatalytic activity of copper oxide nanoparticles for the degradation of methyl orange and real wastewater dyes. Mater. Res. Express 2020, 7, 026201. [Google Scholar] [CrossRef]
  89. Sreeju, N.; Rufus, A.; Philip, D. Studies on catalytic degradation of organic pollutants and anti-bacterial property using biosynthesized CuO nanostructures. J. Mol. Liq. 2017, 242, 690–700. [Google Scholar] [CrossRef]
  90. Singh, J.; Kumar, V.; Kim, K.-H.; Rawat, M. Biogenic synthesis of copper oxide nanoparticles using plant extract and its prodigious potential for photocatalytic degradation of dyes. Environ. Res. 2019, 177, 108569. [Google Scholar] [CrossRef]
  91. Sonia, S.; Poongodi, S.; Kumar, P.S.; Mangalaraj, D.; Ponpandian, N.; Viswanathan, C. Hydrothermal synthesis of highly stable CuO nanostructures for efficient photocatalytic degradation of organic dyes. Mater. Sci. Semicond. Process. 2015, 30, 585–591. [Google Scholar] [CrossRef]
  92. Akter, J.; Sapkota, K.P.; Hanif, M.A.; Islam, M.A.; Abbas, H.G.; Hahn, J.R. Kinetically controlled selective synthesis of Cu2O and CuO nanoparticles toward enhanced degradation of methylene blue using ultraviolet and sun light. Mater. Sci. Semicond. Process. 2021, 123, 105570. [Google Scholar] [CrossRef]
  93. Manikandan, D.B.; Arumugam, M.; Veeran, S.; Sridhar, A.; Krishnasamy Sekar, R.; Perumalsamy, B.; Ramasamy, T. Biofabrication of ecofriendly copper oxide nanoparticles using Ocimum americanum aqueous leaf extract: Analysis of in vitro antibacterial, anticancer, and photocatalytic activities. Environ. Sci. Pollut. Res. 2021, 28, 33927–33941. [Google Scholar] [CrossRef]
  94. Benhadria, N.; Hachemaoui, M.; Zaoui, F.; Mokhtar, A.; Boukreris, S.; Attar, T.; Belarbi, L.; Boukoussa, B. Catalytic reduction of methylene blue dye by copper oxide nanoparticles. J. Clust. Sci. 2022, 33, 249–260. [Google Scholar] [CrossRef]
  95. Nazim, M.; Khan, A.A.P.; Asiri, A.M.; Kim, J.H. Exploring rapid photocatalytic degradation of organic pollutants with porous CuO nanosheets: Synthesis, dye removal, and kinetic studies at room temperature. ACS Omega 2021, 6, 2601–2612. [Google Scholar] [CrossRef]
  96. Rafique, M.; Shafiq, F.; Gillani, S.S.A.; Shakil, M.; Tahir, M.B.; Sadaf, I. Eco-friendly green and biosynthesis of copper oxide nanoparticles using Citrofortunella microcarpa leaves extract for efficient photocatalytic degradation of Rhodamin B dye form textile wastewater. Optik 2020, 208, 164053. [Google Scholar] [CrossRef]
  97. Kannan, D.K.; Radhika, S.V.; Sadasivuni, K.K.; Ojiaku, A.A. Urvashi Verma, Facile fabrication of CuO nanoparticles via microwave-assisted method: Photocatalytic, antimicrobial and anticancer enhancing performance. Int. J. Environ. Anal. Chem. 2020, 102, 1095–1108. [Google Scholar] [CrossRef]
  98. Shayegan Mehr, E.; Sorbiun, M.; Ramazani, A.; Taghavi Fardood, S. Plant-mediated synthesis of zinc oxide and copper oxide nanoparticles by using ferulago angulata (schlecht) boiss extract and comparison of their photocatalytic degradation of Rhodamine B (RhB) under visible light irradiation. J. Mater. Sci. Mater. Electron. 2018, 29, 1333–1340. [Google Scholar] [CrossRef]
  99. Katwal, R.; Kaur, H.; Sharma, G.; Naushad, M.; Pathania, D. Electrochemical synthesized copper oxide nanoparticles for enhanced photocatalytic and antimicrobial activity. J. Ind. Eng. Chem. 2015, 31, 173–184. [Google Scholar] [CrossRef]
  100. Pimol, P.; Khanidtha, M.; Prasert, P. Influence of particle size and salinity on adsorption of basic dyes by agricultural waste: Dried Seagrape (Caulerpa lentillifera). J. Environ. Sci. 2008, 20, 760–768. [Google Scholar] [CrossRef]
  101. El-Berry, M.F.; Sadeek, S.A.; Abdalla, A.M.; Nassar, M.Y. Microwave-assisted fabrication of copper nanoparticles utilizing different counter ions: An efficient photocatalyst for photocatalytic degradation of safranin dye from aqueous media. Mater. Res. Bull. 2021, 133, 111048. [Google Scholar] [CrossRef]
  102. Daneshvar, N.; Salari, D.; Khataee, A.R. Photocatalytic degradation of azo dye acid red 14 in water: Investigation of the effect of operational parameters. J. Photochem. Photobiol. A Chem. 2003, 157, 111–116. [Google Scholar]
  103. Dhananasekaran, S.; Palanivel, R.; Pappu, S. Adsorption of methylene blue, bromophenol blue, and coomassie brilliant blue by α-chitin nanoparticles. J. Adv. Res. 2016, 7, 113–124. [Google Scholar] [CrossRef] [Green Version]
  104. Kumar, A.; Pandey, G. A review on the factors affecting the photocatalytic degradation of hazardous materials. Mater. Sci. Eng. Int. J 2011, 1, 106. [Google Scholar] [CrossRef] [Green Version]
  105. Moradi, S.E.; Haji Shabani, A.M.; Dadfarnia, S.; Emami, S. Effective removal of ciprofloxacin from aqueous solutions using magnetic metal–organic framework sorbents: Mechanisms, isotherms and kinetics. Mater. Sci. Eng. Int. J 2016, 13, 1617–1627. [Google Scholar] [CrossRef]
Figure 1. XRD diffractogram of biogenic CuO NPs (monoclinic CuO phase structure).
Figure 1. XRD diffractogram of biogenic CuO NPs (monoclinic CuO phase structure).
Catalysts 13 00502 g001
Figure 2. SEM image (a) and size distribution (b) of biogenic CuO NPs.
Figure 2. SEM image (a) and size distribution (b) of biogenic CuO NPs.
Catalysts 13 00502 g002
Figure 3. FTIR spectrum of biogenic CuO NPs.
Figure 3. FTIR spectrum of biogenic CuO NPs.
Catalysts 13 00502 g003
Figure 4. UV-vis spectrum of biogenic CuO NPs.
Figure 4. UV-vis spectrum of biogenic CuO NPs.
Catalysts 13 00502 g004
Figure 5. PL spectra of biogenic CuO NPs at: (a) 300 nm, (b) 350 nm.
Figure 5. PL spectra of biogenic CuO NPs at: (a) 300 nm, (b) 350 nm.
Catalysts 13 00502 g005
Figure 6. Mechanism of biogenic synthesis of CuO NPs.
Figure 6. Mechanism of biogenic synthesis of CuO NPs.
Catalysts 13 00502 g006
Figure 7. Absorption spectra of the dyes at different time intervals: (a) MG, (b) MO.
Figure 7. Absorption spectra of the dyes at different time intervals: (a) MG, (b) MO.
Catalysts 13 00502 g007
Figure 8. Degradation rate of (a) MG and (b) MO dyes.
Figure 8. Degradation rate of (a) MG and (b) MO dyes.
Catalysts 13 00502 g008
Figure 9. Kinetic results for dye degradation: (a) A/Ao vs. time; (b) ln (A/Ao) vs. time.
Figure 9. Kinetic results for dye degradation: (a) A/Ao vs. time; (b) ln (A/Ao) vs. time.
Catalysts 13 00502 g009
Figure 10. Schematic representation of photocatalytic degradation of dyes.
Figure 10. Schematic representation of photocatalytic degradation of dyes.
Catalysts 13 00502 g010
Figure 11. Mechanochemical synthesis of biogenic CuO NPs.
Figure 11. Mechanochemical synthesis of biogenic CuO NPs.
Catalysts 13 00502 g011
Table 1. Degradation rates (%) and rate constants for photocatalytic degradation of dyes using biogenic CuO NPs.
Table 1. Degradation rates (%) and rate constants for photocatalytic degradation of dyes using biogenic CuO NPs.
DyesDegradation Rate%Rate Constant (min−1)
Methyl green (MG)65.2310.0175285
Methyl orange (MO)65.0780.0175348
Table 2. Comparative study of photocatalytic degradation of dyes using CuO based materials.
Table 2. Comparative study of photocatalytic degradation of dyes using CuO based materials.
PhotocatalystSynthesis MethodTime (min)Light SourceDyeDegradation RateRef.
CuO NPsGreen Synthesis24UV lightMO96%[81]
CuO NPsGreen Synthesis24UV lightMO96.4%[82]
CuO NPsChemical Precipitation Method120UV lightMO90%[83]
CuO MicrospheresReflux Condensation Method93
130
UV lightMO
MB
89.39%
92%
[84]
CuO NPsGreen Synthesis60
60
UV light
Sunlight
MO
MO
45.23%
31.95%
[85]
CuO NPsGreen Synthesis60SunlightMO95%[86]
CuO NanorodHydrothermal Method90SunlightMO22%[87]
CuO NPsCo-precipitation Method120Xenon lampMO39%[88]
CuO NPsGreen Synthesis4
12
4
4
Visible lightMO
MB
MR
EY
80%
91%
89%
97%
[89]
CuO NPsGreen Synthesis120Solar lightNB
RY 160
93%
81%
[90]
CuO
nano leaves
Hydrothermal Synthesis180UV lightMB
MV
89%
96%
[91]
CuO NPsSelective Method60
120
UV light
Sunlight
MB81%
63%
[92]
CuO NPsGreen Synthesis200SunlightEY
Rh123
MB
75.69%
34.12%
71.06%
[93]
CuO NPsPrecipitation Method15Visible sourceMB74%[94]
CuO NanosheetsRT Synthesis6SunlightAR96.99%[95]
CuO NPsGreen Synthesis90UV light sourceRB98%[96]
CuO NPsMicrowave-assisted method90SunlightMB99%[97]
CuO NPsGreen Synthesis150Visible lightRB84%[98]
CuO NPsElectrochemical Method120SunlightMB
MR
CR
93%
90%
85%
[99]
CuO NPsMechanochemical synthesis60SunlightMO65%This work
CuO NPsMechanochemical synthesis60SunlightMG65%This work
Abbreviations: NB—Nile blue; RY160—Reactive yellow; CBB—Coomassie brilliant blue; EY—Eosin yellow; Rh123—rhodamine 123; MR—Methyl red; CR—Congo red; AR—Allura red; AB—Acid black 210; RT—room temperature.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Aroob, S.; Carabineiro, S.A.C.; Taj, M.B.; Bibi, I.; Raheel, A.; Javed, T.; Yahya, R.; Alelwani, W.; Verpoort, F.; Kamwilaisak, K.; et al. Green Synthesis and Photocatalytic Dye Degradation Activity of CuO Nanoparticles. Catalysts 2023, 13, 502. https://doi.org/10.3390/catal13030502

AMA Style

Aroob S, Carabineiro SAC, Taj MB, Bibi I, Raheel A, Javed T, Yahya R, Alelwani W, Verpoort F, Kamwilaisak K, et al. Green Synthesis and Photocatalytic Dye Degradation Activity of CuO Nanoparticles. Catalysts. 2023; 13(3):502. https://doi.org/10.3390/catal13030502

Chicago/Turabian Style

Aroob, Sadia, Sónia A. C. Carabineiro, Muhammad Babar Taj, Ismat Bibi, Ahmad Raheel, Tariq Javed, Rana Yahya, Walla Alelwani, Francis Verpoort, Khanita Kamwilaisak, and et al. 2023. "Green Synthesis and Photocatalytic Dye Degradation Activity of CuO Nanoparticles" Catalysts 13, no. 3: 502. https://doi.org/10.3390/catal13030502

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