Home Physical Sciences Photodegradation of methyl orange under solar irradiation on Fe-doped ZnO nanoparticles synthesized using wild olive leaf extract
Article Open Access

Photodegradation of methyl orange under solar irradiation on Fe-doped ZnO nanoparticles synthesized using wild olive leaf extract

  • Tahani Saad Algarni , Naaser A. Y. Abduh , Ahmed Aouissi EMAIL logo and Abdullah Al Kahtani
Published/Copyright: September 16, 2022
Become an author with De Gruyter Brill

Abstract

Fe-doped ZnO nanoparticles (NPs) with different Fe contents (0.1–5.0 wt%) were prepared using extract of wild olive leaves growing in Saudi Arabia (region of Abha). The biosynthesized NPs were characterized by Fourier transform infrared spectroscopy, X-ray diffraction, Brunauer–Emmett–Teller, scanning electron microscopy, transmission electron microscopy, and photoluminescence (PL). Characterization results showed that undoped ZnO and Fe-doped ZnO powders were crystallized in the wurtzite structure with a small shift for the doped samples. Neither Fe3O4 nor another iron oxide phase was observed in the samples, which proves the incorporation of Fe into the ZnO lattice. Doping has a pronounced effect on the physical and optical properties. Indeed, the size of the crystallites, the energy of the bandgap as well as the intensity of the PL emission decreased with the Fe content. Photocatalytic tests revealed that the doped samples degraded methyl orange (MO) more efficiently than pure ZnO and pure Fe3O4. Moreover, the photocatalytic activity improved with increasing Fe content. The best photocatalyst of the series (Fe–ZnO-5) was found degrading MO by 92.1%, in 90 min in a pseudo-first order reaction.

Graphical abstract

1 Introduction

Several industries including textiles, paper, plastics, and leather tanneries use synthetic dyes in large quantities [1,2]. Among these dyes, the azo compounds are the most abundant compounds (60–70%) in textile waste, which contributes significantly to water pollution [3]. Because of their toxicity, they generally pose a great threat to living aquatic organisms [4,5]. That is why research for their disposal of wastewater has been intensified. Methyl orange (MO), one of the azo dyes often used is non-biodegradable and therefore it can cause several pollution problems by generating toxic carcinogenic compounds in water [6]. Its degradation is therefore necessary because of its properties harmful to humanity [7]. Various methods physical [8,9], chemical [10,11], and biological [12] for removing azo dyes in wastewater have been explored. Unfortunately, in most cases these methods do not completely destroy the dye molecules [13,14,15]. Thus, it is necessary to develop an efficient method for their complete degradation. Advanced oxidation processes, whose principle is based on the production of hydroxyl radicals (·OH), are effective methods for the complete degradation of dyes. These methods using materials capable of collecting solar energy (photocatalysis) and mechanical energy (piezocatalysis) have found applications for wastewater treatment due to their efficiency and environmental friendliness [16,17,18]. It has proven to be the most efficient and environmentally friendly method for wastewater treatment. Thanks to this alternative, the pollutants are effectively mineralized into H2O and CO2. The degradation is carried out in the presence of a photocatalyst, according to an oxidation-reduction process without causing secondary pollution [19]. In the case of piezocatalysis, the mechanical vibrations, by increasing the number of free electrons and holes on the surface of the piezocatalyst, produce free radicals which decompose organic pollutants by oxidation [20].

On the other hand, in the case of photocatalysis the organic pollutants are decomposed by oxidation using solar energy. The oxidation is triggered by photons whose energy matches or exceeds that of the bandgap of a given semiconductor [21]. In this process, where an electron is excited from the valence band (VB) to the conduction band (CB), a positive hole in the VB is formed due to the action of the OH˙ radical with the OH ion in the water, which can then be used for oxidation. Meanwhile, the excited electron can reduce the oxygen in the CB, thereby acting as an oxidant [21]. Due to their instability in the excited state, the electrons generated can recombine in their respective holes. This causes dissipation of the input light energy and results in low efficiency of the photocatalyst. Photocatalysis is a cost-effective treatment method to get rid of toxic pollutants present in industrial wastes and has the advantage of degrading various chemical compounds under the action of sunlight.

In order to develop an effective photocatalyst, several metallic semiconductor materials have been explored as a photocatalyst [22]. Among these semiconductor materials, ZnO nanoparticles (NPs) have often been used because of their interesting properties compared to other nanomaterials. In addition, they are much more available and less expensive for wastewater treatment [23]. ZnO is a semiconductor that has a wide bandgap (about 3.37 eV) and a high excitation binding energy (60 mV). This allows the creation of electron–hole pairs under light irradiation. Unfortunately, its wide band limits its use in the UV region of sunlight [24,25]. An efficient photocatalyst must absorb visible radiation in addition to ultraviolet radiation because solar radiation is composed of a large proportion of visible radiation and a small proportion of ultraviolet radiation [26]. It is therefore essential to extend the optical absorption of ZnO from the UV region to the visible region for photocatalytic applications. Various approaches, including controlling the morphology and the size, coupling with other semiconductors, and doping with several elements have been explored to overcome this constraint. In the case of doping, several dopants were used including non-metals and transition metals [27]. Each dopant can induce a specific characteristic that can influence the properties of the photocatalyst. It is therefore possible by inserting transition metal ion dopants to enhance the photocatalytic properties of ZnO. Various photocatalytic semiconductors have been studied but their application in photocatalysis still remains on a laboratory scale. The difficulty of their industrial application is generally attributed to the low photocatalytic activity under solar lighting [28]. On the other hand, considerable research work has been done to improve the photocatalytic properties. Iron-based oxides have the advantage of having a narrow bandgap, making it easier to absorb photons in the visible region, thus increasing the visible light absorption of ZnO by narrowing the bandgap. By doping zinc oxide having a large surface area with iron oxide having a smaller bandgap, an efficient solar photocatalyst can be developed. This work aims to develop an environmentally friendly photocatalyst for wastewater treatment, operating under conditions similar to those encountered in the natural environment. In order to reach this goal, ZnO and Fe/ZnO nanomaterials were synthesized by a green procedure using wild olive leaf extract and were fully characterized via X-ray diffraction (XRD), transmission electron microscopy (TEM), scanning electron microscope (SEM), Brunauer–Emmett–Teller (BET) surface analysis, and UV-visible (UV-Vis) spectrophotometer. To our knowledge, the synthesis of Fe-doped ZnO nanomaterials by a procedure using an extract of wild olive leaves growing in Saudi Arabia (Abha region) is a work that has not been studied before. The efficiency of the photocatalysts was evaluated in the degradation of MO under direct sunlight irradiation.

2 Materials and methods

2.1 Materials

Wild olive leaves were collected from Abha region in southern Kingdom of Saudi Arabia. Zn (NO3)2‧6H2O powder (>99%), NaOH (98%), and C14H14N3NaO3S (MO, 99.8%) were purchased from British Drug Houses Chemicals Ltd, Poole, England. All of these chemicals were used as received.

2.2 Preparation of leaf extract

Undamaged wild olive leaves were washed in deionized water and dried in the shade for 2–3 weeks. After cutting them into small pieces, 10 g of these pieces are dissolved in 100 mL of deionized water, boiled at 60°C for 15 min to kill the pathogens in the solution. Then, this aqueous leaf extract solution was cooled, filtered through filter paper, and stored at 4°C for later use.

2.3 Synthesis of ZnO NPs

The preparation of the ZnO NPs was carried out as follows: 5 g zinc nitrate were added to the 30 mL of extract and stirred for 2 h. 16 mL of NaOH (3.5 M) were added dropwise, the resulting solution was heated to 75°C with stirring in a closed container for 5 h. The precipitate thus formed was separated from the solution by filtration and washed several times with an aqueous solution of ethanol with a water/ethanol ratio of 1/3. The nanomaterial thus obtained was dried in an oven at 60°C overnight and then calcined in a furnace at 500°C for 3 h.

2.4 Synthesis of Fe-dopped ZnO NPs

In order to prepare Fe-dopped ZnO NPs, desired amount of zinc nitrate and iron(iii) chloride was added into 30 mL of the extract and stirred for 2 h. 18 mL of NaOH (3.5 M) were added dropwise, the resulting mixture was heated to 75°C with stirring in a closed vessel for 5 h. The precipitate formed is separated and washed several times with an aqueous solution of ethanol in a water/ethanol proportion of 1/3. Then, it was put in the oven at 60°C overnight. The resulting nanomaterial was then calcined at 500°C for 3 h. A series of Fe-doped ZnO with different Fe contents (Fe: 1%, 2%, 3%, 4%, 5%, and 7%) were prepared. They are denoted as Fe–ZnO-x where x is the mass percentage of Fe in the Fe–ZnO photocatalyst: Fe–ZnO-1, Fe–ZnO-2, Fe–ZnO-3, Fe–ZnO-4, Fe–ZnO-5, and Fe–ZnO-7.

2.5 Photocatalytic experiment

The photo degradation of MO was carried out under solar irradiation at atmospheric pressure. Typically, 50 mg NPs catalyst were added into a beaker containing MO dye solution (10 mL, 10 mg·L−1). Prior to exposure to sunlight, the suspension was magnetically stirred in the dark for 30 min to reach the adsorption–desorption equilibrium. The experiment was carried out on a sunny day (outside temperature 40°C, sunlight intensity in the range of 15–25 mW·cm−2). A control solution was prepared and kept under similar conditions to see if there was any color change in the MO dye solution. At given time intervals, 10 mL of suspension was sampled and centrifuged (7,000 rpm) for 10 min, the supernatant was collected for absorption analysis on a UV-Vis spectrophotometer. The photocatalytic degradation of MO was followed by the gradual decrease in absorption intensity at λ max = 464 nm and a wavelength scan ranging from 600 to 350 nm.

The photodegradation of MO was conducted under solar irradiation at atmospheric pressure. Typically, to a beaker containing 10 mL of MO aqueous solution, 50 mg NPs catalyst are added gradually. Before exposure to sunlight, the suspension is stirred magnetically in the dark for 30 min to reach adsorption–desorption equilibrium. The experiments were carried out on sunny days at a temperature of about 40°C and a sunlight intensity in the range of 15–25 mW‧cm−2. A control solution was prepared and exposed under similar conditions to see if there was a color change in the MO solution. At given periods of time, 10 mL of suspension are taken and centrifuged (7,000 rpm) for 10–15 min. The supernatant is then separated and analyzed using a UV-Vis spectrophotometer. The photocatalytic degradation of MO was followed by the gradual decrease in absorption intensity at λ max = 464 nm and a wavelength sweep ranging from 600 to 350 nm.

The MO photodegradation rate was determined using Eq. 1:

(1) Degradation % = A o A t A 0

where A 0 represents the initial absorbance of MO and A t the absorbance of the solution at time t after solar irradiation.

2.6 Characterization of the photocatalysts

The optical and structural properties of the prepared undoped ZnO and Fe-doped ZnO NPs were characterized by various technics. Functional groups of the compounds were examined by using Fourier transform infrared spectroscopy (FTIR). Infrared spectra were recorded with an infrared spectrometer Genesis II FTIR (4,000–350 cm−1) using the KBr Pellet technique. The crystal structure of the nanomaterials was examined by an XRD, using an Ultima IV X-ray Rigaku diffractometer using Cu-Kα radiation. The data were collected over a 2θ range of 20–80°. TEM micrographs were performed on a JEOL JEM-2100F field emission electron microscope (JEOL, Japan) with an acceleration voltage of 110 kV. Surface morphology of the catalysts were determined by means of a Scanning Electron Microscopy (SEM) on a JEOL JSM-7600F) (JEOL, Japan) with an accelerating voltage of 100 V and a beam current of 30A. UV-Vis analysis was performed by using a double beam UV-Vis. The specific surface area (BET) was measured by a Micromeritics Tristar II 3020 surface area and porosity analyzer. The fluorescence spectra were measured using Shimadzu RF-6000 fluorescence spectrophotometer. Particle aqueous suspension was sonicated for 30 min. The excitation wavelength was 320 nm and the fluorescence spectra were recorded in the wavelength range from 300 to 600 nm. A pass width of 10 nm was used for the measurement of the emission and excitation spectra.

3 Results and discussion

3.1 Catalysts characterization

3.1.1 FTIR

An FTIR study was performed to determine the major functional groups present in wild olive leaf extract (WOLE) to suggest biomolecules responsible for reducing and capping Zn2+ leading to the formation of ZnO NPs. The WOLE spectrum shows the presence of a multitude of absorption peaks, indicating its complex nature. The spectrum is shown in Figure 1. The IR bands at 3,405 and 1,710 cm−1 in WOLE are due to the O‒H and C…O stretching modes of the OH group and the C…O group, respectively [25,26]. The strong band observed at 2,931 cm−1 seems to be due to the –CH stretching vibrations of the methyl and methylene groups. The peak at 1,635 cm−1 is characteristic of the CO, C‒O, and O‒H groups. The band at 1,635 cm−1, due to the C‒O stretch, indicates the probable presence of carboxyl coupled to the amide bond in amide I and the band at 1,517 cm−1, due to NH stretching, indicates the presence of amide II. The band at 1,405 cm−1 is attributed to methylene vibrations. The intense band at 1,075 cm−1 can be attributed to C‒N stretching vibrations of aliphatic amines [29] or C–OH vibrations in the WOLE [30]. Bands at 930 and 765 cm−1 may be due to C–N stretching amine groups [31]. The 3 small peaks at 618, 542, and 765 cm−1 can be assigned to the C–H bend [32]. The results of the present study suggest that the amine (N–H), hydroxyl (–OH), and carboxyl (–C…O) groups evidenced by the FTIR could be responsible for the reduction and stabilization of the ZnO NPs.

Figure 1 
                     FTIR spectrum of WOLE.
Figure 1

FTIR spectrum of WOLE.

Figure 2 depicts the FTIR spectrum of ZnO and Fe-doped ZnO NPs. FTIR bands of ZnO NPs appear between 441 and 665 cm−1 [33]. The strong peaks at 442 cm−1 can be ascribed to the vibrational modes of the Zn–O bonds as reported in the literature [34]. The strong peak at 442 and 565 cm−1 can be attributed to the vibration of Zn–O bond and confirms the formation of product [35]. It has been reported in the literature that these bands can be ascribed to the stretching vibration of Zn–O bond [34,36]. The band at 665 cm−1 is assigned to the Fe–O stretching. Based on similar FTIR results [37], the insertion of iron at the position of zinc in the ZnO fraction can be confirmed. It is important to note the shift of the peaks of the FTIR spectrum of the NPs compared to the leaf extract such as for example from 3,455 to 3,405, from 1,635 to 1,630, from 1,075 to 1,040, and from 930 to 910 cm−1. This indicates the possible involvement of alcohols, amines, carboxylic acids, ketones, and polyols in the bioreduction of metal ions [38]. In fact, it has been reported that similar functional groups of heterocyclic compounds present in the leaf extract are used as capping ligands [39].

Figure 2 
                     FTIR spectra of: (a) Fe3O4, (b) ZnO, (c) Fe–ZnO-1, (d) Fe–ZnO-2, (e) Fe–ZnO-3, (f) Fe–ZnO-4, (g) Fe–ZnO-5, and (h) Fe–ZnO-7.
Figure 2

FTIR spectra of: (a) Fe3O4, (b) ZnO, (c) Fe–ZnO-1, (d) Fe–ZnO-2, (e) Fe–ZnO-3, (f) Fe–ZnO-4, (g) Fe–ZnO-5, and (h) Fe–ZnO-7.

3.1.2 XRD analysis

The XRD patterns for ZnO, Fe3O4, and Fe-doped ZnO samples calcined at 500°C are shown in Figure 3. The result shows the presence of crystallite peaks at 2θ = 31.79°, 34.45°, 36.29°, 47.58°, 56.63°, 62.89°, 66.37°, 67.98°, 69.10°, and 77.05° corresponding to the reflections of planes (100), (002), (101), (102), (110), (103), (200), (112), (201), and (202), designating the polycrystalline hexagonal wurtzite structure [JCPDS file no. 36–1451]. The peak which appears at 45.5° corresponds to Zn(OH)2 [40,41]. It is due to the absorption of moisture induced by the high surface area. As for the Fe-doped Zinc oxide samples, a small shift in the position of the peaks was observed. For example, for Fe–ZnO-5, the peaks were observed at 2θ = 31.72°, 34.42°, 36.22°, 47.53°, 56.56°, 62.84°, 66.30°, 67.93°, 69.04°, and 76.94°. This result is in agreement with that of ref. [42]. Neither the Fe3O4 phase nor another iron oxide phase was observed. This proves that Fe has indeed been inserted into the lattice of ZnO. The decrease in crystallite size may be due to the smaller ionic radius of Fe3+ (0.064 nm) compared to Zn2+ (0.074 nm).

Figure 3 
                     XRD pattern of: (a) ZnO, (b) Fe–ZnO-1, (c) Fe–ZnO-2, (d) Fe–ZnO-3, (e) Fe–ZnO-4, Fe–ZnO-5, and (g) Fe–ZnO-7.
Figure 3

XRD pattern of: (a) ZnO, (b) Fe–ZnO-1, (c) Fe–ZnO-2, (d) Fe–ZnO-3, (e) Fe–ZnO-4, Fe–ZnO-5, and (g) Fe–ZnO-7.

3.1.3 TEM

Figure 4 shows the TEM micrographs of the biosynthesized ZnO and ZnO doped with Fe. The images of the ZnO NPs (Figure 4a) confirm the nanometric scale of the biosynthesized nanomaterials and show that the NPs have a uniform distribution with a spherical shape and with very little aggregation. In addition, TEM images show that ZnO NPs have smooth surfaces and that the size of the NPs varied from 86 to 170 nm. TEM images of Fe-doped ZnO (Figure 4b) showed that the material obtained is heterogeneous and consists of a mixture of NPs of spherical and rod shape with a length varying between 123 and 170 nm. Fe-doped ZnO exhibits more aggregation than pure ZnO. This result indicates that the shape and the size depend strongly on the doping.

Figure 4 
                     TEM micrographs of: (a) undoped ZnO and (b) ZnO-Fe-5.
Figure 4

TEM micrographs of: (a) undoped ZnO and (b) ZnO-Fe-5.

3.1.4 SEM with analysis

SEM micrographs of pure ZnO and iron-doped ZnO are depicted in Figure 5. From the micrograph (Figure 5a), it appears that the morphology of pure ZnO NPs tends toward the spherical shape. ZnO particles are well distributed with very low aggregation. As noted, ZnO particles do not have a uniform size distribution. As for Fe–ZnO-5 (Figure 5b), the image shows an overall porous and rough shape with some aggregations but with a large number of particles smaller than that of pure ZnO.

Figure 5 
                     SEM images of: (a) undoped ZnO and (b) ZnO-Fe-5.
Figure 5

SEM images of: (a) undoped ZnO and (b) ZnO-Fe-5.

3.1.5 Bandgap determination

The optical bandgap (E g) is one of the properties of semiconductor materials which can be greatly influenced by the doping and the size of the NPs. To estimate its value, the UV-Vis spectroscopy is the most used technique. This method was used to follow the evolution of the bandgap value after doping the pure ZnO by iron cation Fe3+. The bandgap value of the undoped and doped ZnO samples were evaluated using the Tauc equation:

(2) ( α h v ) 1 / n = A ( h v E g )

where h, ν, and α account for Planck’s constant, frequency, and absorption coefficient, respectively. A is the proportionality constant and n indicates the type of electronic transition (for permitted transitions, n = 1/2).

The results (Figure 6 and Table 1) show that the bandgap value of iron-doped ZnO is equal to 3.18 eV. This value is within the range of reported values for pure ZnO NPs [43]. The energy bandgap of the Fe-doped ZnO is slightly lower than that of undoped ZnO. As reported in the literature [44], the substitution of Zn2+ by Fe3+ increases the number of oxygen vacancies and energy levels, which decreases the energy gap.

Figure 6 
                      Optical band gab of undoped ZnO and Fe–doped ZnO nanoparticles with different Fe contents (0.1–5.0 wt%).
Figure 6

Optical band gab of undoped ZnO and Fe–doped ZnO nanoparticles with different Fe contents (0.1–5.0 wt%).

Table 1

Crystallite size and bandgap value of the undoped ZnO and Fe-doped ZnO metal oxide NPs

Photocatalyst Physical properties
Crystallite size* (nm) Bandgap (eV)
ZnO 22.50 3.18
ZnO-Fe-1 22.14 3.15
ZnO-Fe-2 21.46 3.14
ZnO-Fe-3 21.00 3.12
ZnO-Fe-4 20.25 3.09
ZnO-Fe-5 21.06 3.05
ZnO-Fe-7 15.49 3.14

*The values of crystallite size have been calculated by the Scherrer’s equation.

3.1.6 Photoluminescence (PL) measurements

The influence of Fe doping on the physical properties of ZnO NPs was investigated by PL analysis. It is well known that the absorbance is related to different parameters such as the size of the NPs, the lack of oxygen, the bandgap, and the defects present in the structure [45]. Figure 7 shows the PL spectra of pure samples of ZnO, Fe–ZnO-3, and Fe–ZnO-5 obtained using ultraviolet light with a wavelength of 320 nm as an excitation source. These spectra show 2 peaks appearing at 363 and 409 nm. As it can be seen, the position of the peaks remains unchanged for each of the samples; the pure ZnO and the Fe-doped ZnO. The first peak observed at 363 nm is assigned to the near band edge [46,47] resulting in the band-to-band transition. The second peak (a shoulder) appearing around 409 nm is attributed to the recombination of an electron–hole pair in the Zn vacancy [48]. It is worth noting that the intensity of the peaks decreases after doping and it decreases further when the doping amount is increased. This can be explained by the fact that when Fe3+ is inserted in the ZnO lattice it induces defects, especially increases oxygen vacancies which ultimately increases the energy levels localized near the VB.

Figure 7 
                     PL spectra of: (a) undoped ZnO, (b) Fe–ZnO-3, and (c) Fe–ZnO-5.
Figure 7

PL spectra of: (a) undoped ZnO, (b) Fe–ZnO-3, and (c) Fe–ZnO-5.

3.2 Photocatalytic activity of ZnO and Fe3O4

To investigate the influence of Fe dopant content on the catalytic activity of undoped ZnO, first the photoactivity of both pure Fe3O4 and pure ZnO was evaluated. The degradation of MO by photolysis and by photocatalysis in the presence of pure ZnO and Fe3O4 were followed by the discoloration of the MO dye solution under sunlight for 45 min. The results obtained (Figure 8) show that in the absence of photocatalyst, the degradation of the MO dye is negligible. This indicated that the effect of photolysis on the MO dye is insignificant. In the dark, only a slight decrease in MO absorbance was observed when the experiment was conducted in the presence of pure ZnO and Fe3O4 separately. This could be due to the adsorption on the surface of the NPs. Under the sunlight, the results showed that undoped ZnO is about 10 times more efficient than Fe3O4. In fact, ZnO degraded 40.7% of MO in 45 min, while Fe3O4 only degraded 4.7% in the same irradiation time. One of the causes of the low photoactivity of Fe3O4 compared to ZnO, is the rapid recombination of the electron–hole pairs of Fe3O4.

Figure 8 
                  UV-visible absorbance spectra of degraded MO-dye under sunlight irradiation for 45 min over pure ZnO and Fe3O4.
Figure 8

UV-visible absorbance spectra of degraded MO-dye under sunlight irradiation for 45 min over pure ZnO and Fe3O4.

3.3 Photocatalytic activity of Fe-doped ZnO

The effect of Fe content on the photocatalytic degradation of aqueous MO solution under solar irradiation is depicted in Figure 9. As can be seen from the figure, all Fe-doped ZnO samples show an activity higher than that of pure ZnO and that of pure Fe3O4. Regarding the amount of Fe dopant on ZnO, it can clearly be seen that the photocatalytic activity increases with increasing amount of Fe up to a Fe content equal to 5%. Beyond this content, a decay was observed. So, the ZnO oxide doped with 5% iron is the most efficient photocatalyst of the series. The low photocatalytic activity of ZnO under the action of sunlight is due to its wide bandgap, only allowing the use of sunlight at the region of UV light. As for the low activity of Fe3O4, it is due to the rapid recombination of photogenerated electrons and holes. When Fe ions are used as dopants (in small amounts), they were inserted into the ZnO matrix and can therefore serve as trap centers for the capture of electrons or holes [49,50]. This result is corroborated by UV and XRD characterizations (Table 1) where it was observed that iron-doped ZnO solids exhibit a lower bandgap and crystallite size than pure ZnO. Similar results have been reported in the literature. Huang et al. [51] reported that doping with Fe3+ ions improve the separation of electron–hole pairs due to the decrease in the bandgap, resulting in a shift of the absorption towards the visible light spectrum and it also decreases the particle size. In their study on arrays of Fe-doped ZnO nanorods with different doping amounts, Xiao et al. [52] observed better photocatalytic activity for 1.0% Fe-doped ZnO nanorods. Ba-Abbad et al. [53] also explained the enhancement of photocatalytic activity by the reduction in the particle size by adding small amounts of Fe dopant. It should be emphasized that other factors can also make a contribution, but little, to the performance of photocatalytic activities. Indeed, it has been reported that the photocatalytic activity is related to the bandgaps, to the nanosized structures, and to the surface areas [54,55]. Based on the above discussion, the photocatalytic activity enhancement of Fe-doped samples may also be due to the high specific surface area in addition to the decrease in the bandgap. In fact, Fe-doped ZnO (ZnO-Fe-5) has a BET surface area of 16.23 m·g−2, while pure ZnO has only a surface area of 3.68 m·g−2. It could then be said that Fe doping improves the activity by decreasing the bandgap energy and crystallite size and also increasing the surface area.

Figure 9 
                  Effect of Fe content on bandgap, crystallite size, and MO degradation. The photodegradation was carried out under solar irradiation for 45 min.
Figure 9

Effect of Fe content on bandgap, crystallite size, and MO degradation. The photodegradation was carried out under solar irradiation for 45 min.

3.3.1 Effect of irradiation time

Figure 10 represents the effect of irradiation time on MO degradation over the best photocatalyst of the series, namely, Fe–ZnO-5. The result shows that the degradation rate of MO dye increased when the irradiation time increased from 0 to 90 min, indicating that the reduction/decomposition process of MO dye is carried out by Fe–ZnO-5 NPs. After 90 min of solar radiation, the degradation of the MO dye reaches 92.1%. This clearly shows the efficiency of Fe–ZnO-5 photocatalyst for MO dye photodegradation.

Figure 10 
                     Degradation as a function of irradiation time of MO dye under solar irradiation on Fe–ZnO-5.
Figure 10

Degradation as a function of irradiation time of MO dye under solar irradiation on Fe–ZnO-5.

3.3.2 Kinetic study

The degradation of MO catalyzed by Fe–ZnO-5 proceeds in two steps. In the first one, the MO molecule is adsorbed on Fe–ZnO-5 and in the second step, the degradation reaction is triggered. Information on the adsorption mechanism of a heterogeneous surface reaction can be obtained from the kinetics of the adsorption process described by the Langmuir–Hinshelwood mechanism for surface catalyzed reactions [56,57]. This process can be detailed by two basic mechanisms:

  1. successive attacks of the coloring matter by the hydroxyl radical (HO˙) leading to its oxidation.

  2. reaction of the dye with generated holes.

When MO molecules and one of the oxygen-containing species are adsorbed on neighboring sites, an oxygen in its molecular form on the surface of the Fe-doped ZnO can play the role of an electron acceptor, thus leading via the superoxide radical O 2 to the hydrodioxyl radical (HO2˙) formation known as a precursor of HO˙ radical [58].

In the case where the degradation of the MO dye in aqueous solution over Fe–ZnO oxide catalyst obeys the kinetics of a pseudo-first order reaction, the description according to Langmuir–Hinshelwood mechanism is as follows:

(3) r = d c d t = k 1 KC 1 + KC

(4) r = K app C 1 + KC

where k accounts for the intrinsic rate constant, K is the Langmuir–Hinshelwood adsorption equilibrium constant, K app (k.K) is the apparent rate constant, and C is the concentration of MO at time t.

After approximation (KC ≪ 1) and integration, the above rate expression gives the pseudo-first-order equation:

(5) ln C o C t = k app t

The curve ln(C o/C t ) representing the photodegradation with time of MO by Fe–ZnO-5 photocatalyst under sunlight irradiation is represented in Figure 11. The curve obtained is a straight line with a correlation constant R 2 = 0.9832. As the correlation constant for a straight line must be R 2 > 0.95, the kinetics of the MO degradation reaction in the presence of Fe–ZnO-5 is of pseudo first-order. The value of the rate constant determined from the slope of the curve is 0.0254 min−1. This value further indicates that the Fe–ZnO-5 photocatalyst has good photocatalytic reactivity which corroborates the MO degradation efficiency.

Figure 11 
                     Kinetics of MO-dye photodegradation over Fe-ZnO-5 under sunlight irradiation.
Figure 11

Kinetics of MO-dye photodegradation over Fe-ZnO-5 under sunlight irradiation.

3.3.3 Effect of catalyst loading

To investigate the effect of the amount of catalyst on MO photodegradation, experiments were performed by varying the Fe–ZnO-5 photocatalyst loading from 0.25 to 2.00 g·L−1. The dye concentration was kept to 0.03 mM and the duration of irradiation was for 40 min. The results showed that the percentage of MO degradation increased from 58.9% to 95.3% when the catalyst loading increased from 0.25–2.00 g·L−1 (Figure 12). The increase in the degradation rate with the loading of Fe–ZnO-5 is due to the increase in the number of active sites. In fact, the increase in the latter leads to an increase in the hydroxyl radicals formed, thus leading to an increase in the number of degraded dye molecules. It should be mentioned that the use of a large amount of catalyst increases the turbidity of the solution which limits the penetration of UV light into the reaction mixture and thus leads to a decrease in the reaction rate. This is why the experiments were carried out with small quantities in order to avoid the turbidity of the solution.

Figure 12 
                     Effect of Fe–ZnO-5 photocatalyst loading on MO dye degradation under solar irradiation.
Figure 12

Effect of Fe–ZnO-5 photocatalyst loading on MO dye degradation under solar irradiation.

3.3.4 Reusability

Recycling of photocatalysts is one of the important processes as it contributes to economic and environmental development. Thus, the photocatalytic stability of a photocatalyst for its reuse is a critical factor, especially for large-scale applications. In order to study the stability of the biosynthesized Fe–ZnO-5 photocatalyst, experiments for its recycling for MO dye degradation were conducted under the same conditions. After each experiment, Fe–ZnO-5 is recovered from the aqueous MO solution by centrifugation, washed three times with an ethanol–acetone solution (1:1 by volume), and then placed in an oven at 80°C overnight to be reused. The recovered Fe–ZnO was reused twice, and the results (Figure 13) obtained show that the fresh Fe–ZnO-5 sample degraded the MO by 92.1%, while in the case of its second and third use, it has degraded the MO by 87.3% and 75.4%, respectively. This slight decrease in the rate of degradation was expected, indeed, many studies have shown that the photocatalytic activity decreases after repeated use [59,60]. These results therefore confirm that Fe–ZnO maintained an acceptable performance allowing it to be used successfully for the removal of MO dye in wastewater.

Figure 13 
                     Recycling efficiency of Fe–ZnO-5 photocatalyst on MO dye degradation under solar irradiation.
Figure 13

Recycling efficiency of Fe–ZnO-5 photocatalyst on MO dye degradation under solar irradiation.

4 Conclusion

ZnO and iron-doped ZnO NPs were biosynthesized by a simple method using WOLEs. They have been characterized and tested for MO degradation using sunlight. The effects of Fe doping content on physical and photocatalytic properties were investigated.

The shift evidenced by the FTIR analysis for the obtained peaks suggests that the amine (NH), hydroxyl (–OH), and carboxyl (–C…O) groups could be involved in the reduction and stabilization of the ZnO NPs.

Fe doping leads to a reduction in the size of ZnO crystallites without modifying the crystal structure of ZnO. The absence of Fe3O4 or any Fe(iii)-oxide phase in the XRD pattern indicates that Fe3 + has been inserted into the ZnO.

The TEM image reveals that the ZnO NPs have a spherical shape and do not show aggregation, while the Fe-doped ZnO NPs consist of a mixture of spherical and rod shape showing aggregation. This result clearly indicates that the size and shape of the NPs strongly depend on the doping.

Fe doping decreased the bandgap energy of ZnO and as the Fe content increased, the bandgap decreased. This may suggest that the substitution of Zn2+ by Fe3+ causes the formation of oxygen vacancies resulting in additional energy levels, thus reducing the bandgap.

The enhanced photocatalytic performance of doped ZnO NPs is the result of Fe doping which reduced the bandgap energy and crystal size and increased surface area.

  1. Funding information: The authors extended their appreciation to the Deanship of Scientific Research at King Saud University: Grant Number RG-1441-507.

  2. Author contributions: Naaser A.Y. Abduh carried out the photocatalytic tests and the chromatographic analyses. Tahani Al-Garni collected the olive leaves and carried out the characterizations of the prepared photocatalysts. Abdullah Al Kahtani as head of laboratory provided the equipment and the necessary funding for the realization of this work, Ahmed Aouissi planned the realization of this work and wrote the article. All authors have read and accepted the published version of the manuscript.

  3. Conflict of interest: Authors state no conflict of interest.

References

[1] Ito T, Adachi Y, Yamanashi Y, Shimada Y. Long–term natural remediation process in textile dye-polluted river sediment driven by bacterial community changes. Water Res. 2016;100:458–65.10.1016/j.watres.2016.05.050Search in Google Scholar PubMed

[2] Liu J, Peng G, Jing X, Yi Z. Treatment of methyl orange by the catalytic wet peroxide oxidation process in batch and continuous fixed bed reactors using Fe-impregnated 13X as catalyst. Water Sci Technol. 2018;78:936–46.10.2166/wst.2018.372Search in Google Scholar PubMed

[3] Aziztyana AP, Wardhani S, Prananto YP, Purwonugroho D. IOP conference series. Mater Sci Eng. IOP Publishing; 2019. p. 042047.10.1088/1757-899X/546/4/042047Search in Google Scholar

[4] Ozdemir O, Armagan B, Turan M, Celik MS. Comparison of the adsorption characteristics of azo-reactive dyes on mezoporous minerals. Dye Pigment. 2004;62:49–60.10.1016/j.dyepig.2003.11.007Search in Google Scholar

[5] Wang S, Li H, Xie S, Liu S, Xu L. Physical and chemical regeneration of zeolitic adsorbents for dye removal in wastewater treatment. Chemosphere. 2006;65:82–7.10.1016/j.chemosphere.2006.02.043Search in Google Scholar PubMed

[6] Khatri J, Nidheesh PV, Singh TA, Kumar MS. Advanced oxidation processes based on zero-valent aluminium for treating textile wastewater. Chem Eng J. 2018;348:67–73.10.1016/j.cej.2018.04.074Search in Google Scholar

[7] Zou H, Wang Y. Functional collaboration of biofilm-cathode electrode and microbial fuel cell for biodegradation of methyl orange and simultaneous bioelectricity generation. Env Sci Pollut R. 2019;26:23061–9.10.1007/s11356-019-05617-wSearch in Google Scholar PubMed

[8] Garrido-Cardenas JA, Esteban-García B, Agüera A, Sánchez-Pérez JA, Manzano-Agugliaro F. Wastewater treatment by advanced oxidation process and their worldwide research trends. Int J Env Res Pub He. 2020;17:170.10.3390/ijerph17010170Search in Google Scholar PubMed PubMed Central

[9] Al-Baldawi IA, Abdullah SRS, Almansoory AF, Ismail NI, Hasan HA, Anuar N. Role of Salvinia molesta in biodecolorization of methyl orange dye from water. Sci Rep-Uk. 2020;10:1–9.10.1038/s41598-020-70740-5Search in Google Scholar PubMed PubMed Central

[10] Sen SK, Raut S, Bandyopadhyay P, Raut S. Fungal decolouration and degradation of azo dyes: a review. Fungal Biol Rev. 2016;30:112–33.10.1016/j.fbr.2016.06.003Search in Google Scholar

[11] Wang S, Yun J-H, Luo B, Butburee T, Peerakiatkhajohn P, Thaweesak S, et al. Recent progress on visible light responsive heterojunctions for photocatalytic applications. J Mater Sci Technol. 2017;33:1–22.10.1016/j.jmst.2016.11.017Search in Google Scholar

[12] Guo Q, Zhou C, Ma Z, Yang X. Fundamentals of TiO2 photocatalysis: concepts, mechanisms, and challenges. Adv Mater. 2019;31:1901997.10.1002/adma.201901997Search in Google Scholar PubMed

[13] Muhd Julkapli N, Bagheri S, Bee Abd Hamid S. Recent advances in heterogeneous photocatalytic decolorization of synthetic dyes. Sci World J. 2014;2014.10.1155/2014/692307Search in Google Scholar PubMed PubMed Central

[14] Minas F, Chandravanshi BS, Leta S. Chemical precipitation method for chromium removal and its recovery from tannery wastewater in Ethiopia. Chem Int. 2017;3:291–305.Search in Google Scholar

[15] Jafarinejad S. Activated sludge combined with powdered activated carbon (PACT process) for the petroleum industry wastewater treatment: a review. Chem Int. 2017;3:368.Search in Google Scholar

[16] Ismail M, Wu Z, Zhang L, Ma J, Jia Y, Hu Y, et al. High-efficient synergy of piezocatalysis and photocatalysis in bismuth oxychloride nanomaterial for dye decomposition. Chemosphere. 2019;228:212–8.10.1016/j.chemosphere.2019.04.121Search in Google Scholar PubMed

[17] You H, Wu Z, Jia Y, Xu X, Xia Y, Han Z, et al. High-efficiency and mechano-/photo-bi-catalysis of piezoelectric-ZnO@ photoelectric-TiO2 core-shell nanofibers for dye decomposition. Chemosphere. 2017;183:528–35.10.1016/j.chemosphere.2017.05.130Search in Google Scholar PubMed

[18] Xu X, Wu Z, Xiao L, Jia Y, Ma J, Wang F, et al. Strong piezo-electro-chemical effect of piezoelectric BaTiO3 nanofibers for vibration-catalysis. J Alloy Compd. 2018;762:915–21.10.1016/j.jallcom.2018.05.279Search in Google Scholar

[19] El Nemr A, Hassaan MA, Madkour FF. Advanced oxidation process (AOP) for detoxification of acid red 17 dye solution and degradation mechanism. Env Process. 2018;5:95–113.10.1007/s40710-018-0284-9Search in Google Scholar

[20] Tu S, Guo Y, Zhang Y, Hu C, Zhang T, Ma T, et al. Piezocatalysis and piezo‐photocatalysis: catalysts classification and modification strategy, reaction mechanism, and practical application. Adv Funct Mater. 2020;48:2005158.10.1002/adfm.202005158Search in Google Scholar

[21] Saleh TA, Gupta VK. Photo-catalyzed degradation of hazardous dye methyl orange by use of a composite catalyst consisting of multi-walled carbon nanotubes and titanium dioxide. J Colloid Interf Sci. 2012;371:101–6.10.1016/j.jcis.2011.12.038Search in Google Scholar PubMed

[22] Hernández-Alonso MD, Fresno F, Suárez S, Coronado JM. Development of alternative photocatalysts to TiO2: challenges and opportunities. Energ Env Sci. 2009;2:1231–57.10.1039/b907933eSearch in Google Scholar

[23] Ray PZ, Shipley HJ. Inorganic nano-adsorbents for the removal of heavy metals and arsenic: a review. RSC Adv. 2015;5:29885–907.10.1039/C5RA02714DSearch in Google Scholar

[24] Kong D, Zheng Y, Kobielusz M, Wang Y, Bai Z, Macyk W, et al. Recent advances in visible light-driven water oxidation and reduction in suspension systems. Mater Today. 2018;21:897–924.10.1016/j.mattod.2018.04.009Search in Google Scholar

[25] Shu H-Y, Chang M-C, Tseng T-H. Solar and visible light illumination on immobilized nano zinc oxide for the degradation and mineralization of orange G in wastewater. Catalysts. 2017;7:164.10.3390/catal7050164Search in Google Scholar

[26] Nair MG, Nirmala M, Rekha K, Anukaliani A. Structural, optical, photo catalytic and antibacterial activity of ZnO and Co doped ZnO nanoparticles. Mater Lett. 2011;65:1797–800.10.1016/j.matlet.2011.03.079Search in Google Scholar

[27] Yi S, Cui J, Li S, Zhang L, Wang D, Lin Y. Enhanced visible-light photocatalytic activity of Fe/ZnO for rhodamine B degradation and its photogenerated charge transfer properties. Appl Surf Sci. 2014;319:230–6.10.1016/j.apsusc.2014.06.151Search in Google Scholar

[28] Low J, Yu J, Jaroniec M, Wageh S, Al‐Ghamdi AA. Heterojunction photocatalysts. Adv Mater. 2017;29:1601694.10.1002/adma.201601694Search in Google Scholar PubMed

[29] Shameli K, Ahmad MB, Jazayeri SD, Shabanzadeh P, Sangpour P, Jahangirian H, et al. Investigation of antibacterial properties silver nanoparticles prepared via green method. Chem Cent J. 2012;6:1–10.10.1186/1752-153X-6-73Search in Google Scholar PubMed PubMed Central

[30] Khalil MM, Ismail EH, El-Baghdady KZ, Mohamed D. Green synthesis of silver nanoparticles using olive leaf extract and its antibacterial activity. Arab J Chem. 2014;7:1131–9.10.1016/j.arabjc.2013.04.007Search in Google Scholar

[31] Gupta M, Tomar RS, Kaushik S, Mishra RK, Sharma D. Effective antimicrobial activity of green ZnO nano particles of Catharanthus roseus. Front Microbiol. 2018;2030.10.3389/fmicb.2018.02030Search in Google Scholar PubMed PubMed Central

[32] Md Salim R, Asik J, Sarjadi MS. Chemical functional groups of extractives, cellulose and lignin extracted from native Leucaena leucocephala bark. Wood Sci Technol. 2021;55(2):295–313.10.1007/s00226-020-01258-2Search in Google Scholar

[33] Suresh J, Pradheesh G, Alexramani V, Sundrarajan M, Hong SI. Green synthesis and characterization of zinc oxide nanoparticle using insulin plant (Costus pictus D. Don) and investigation of its antimicrobial as well as anticancer activities. Adv Nat Sci-Nanosci. 2018;9:015008.10.1088/2043-6254/aaa6f1Search in Google Scholar

[34] Rodrıguez-Paéz J, Caballero A, Villegas M, Moure C, Duran P, Fernandez J. Controlled precipitation methods: formation mechanism of ZnO nanoparticles. J Eur Ceram Soc. 2001;21:925–30.10.1016/S0955-2219(00)00283-1Search in Google Scholar

[35] PP V. In vitro biocompatibility and antimicrobial activities of zinc oxide nanoparticles (ZnO NPs) prepared by chemical and green synthetic route – A comparative study. BionanoScience. 2020;10:112–21.10.1007/s12668-019-00698-wSearch in Google Scholar

[36] Páez CA, Liquet DY, Calberg C, Lambert SD, Willems I, Germeau A, et al. Study of photocatalytic decomposition of hydrogen peroxide over ramsdellite-MnO2 by O2-pressure monitoring. Catal Commun. 2011;15:132–6.10.1016/j.catcom.2011.08.025Search in Google Scholar

[37] Ferin Fathima A, Jothi Mani R, Sakthipandi K, Manimala K, Hossain A. Enhanced antifungal activity of pure and iron-doped ZnO nanoparticles prepared in the absence of reducing agents. J Inorg Organomet P. 2020;30:2397–405.10.1007/s10904-019-01400-zSearch in Google Scholar

[38] Bala N, Saha S, Chakraborty M, Maiti M, Das S, Basu R, et al. Green synthesis of zinc oxide nanoparticles using Hibiscus subdariffa leaf extract: effect of temperature on synthesis, anti-bacterial activity and anti-diabetic activity. RSC Adv. 2015;5:4993–5003.10.1039/C4RA12784FSearch in Google Scholar

[39] Peletiri C, Matur B, Ihongbe J, Okoye M. The effect of Azadirachta indica (Neem Tree) on human plasmodiasis: the laboratory perspective. Res J Med Sci. 2012;2:13–7.Search in Google Scholar

[40] Wang M, Jiang L, Kim EJ, Hahn SH. Electronic structure and optical properties of Zn (OH) 2: LDA + U calculations and intense yellow luminescence. RSC Adv. 2015;5:87496–87503.10.1039/C5RA17024ASearch in Google Scholar

[41] Shankar JS, Ashok Kumar S, Periyasamy BK, Nayak SK. Studies on optical characteristics of multicolor emitting MEH-PPV/ZnO hybrid nanocomposite. Polym-Plast Technol. 2019;58:148–57.10.1080/03602559.2018.1466171Search in Google Scholar

[42] Johnson EM, Ojwang JO, Szekely A, Wallace TL, Warnock DW. Comparison of in vitro antifungal activities of free and liposome-encapsulated nystatin with those of four amphotericin B formulations. Antimicrob Agents Ch. 1998;42:1412–6.10.1128/AAC.42.6.1412Search in Google Scholar PubMed PubMed Central

[43] Rai A, Singh A, Ahmad A, Sastry M. Role of halide ions and temperature on the morphology of biologically synthesized gold nanotriangles. Langmuir. 2006;22:736–41.10.1021/la052055qSearch in Google Scholar PubMed

[44] Dong S, Xu K, Liu J, Cui H. Photocatalytic performance of ZnO: Fe array films under sunlight irradiation. Phys B. 2011;406:3609–12.10.1016/j.physb.2011.06.053Search in Google Scholar

[45] Raja K, Ramesh P, Geetha D. Structural, FTIR and photoluminescence studies of Fe doped ZnO nanopowder by co-precipitation method. Spectrochim Acta Part A. 2014;131:183–8.10.1016/j.saa.2014.03.047Search in Google Scholar

[46] Huang MH, Wu Y, Feick H, Tran N, Weber E, Yang P. Catalytic growth of zinc oxide nanowires by vapor transport. Adv Mater. 2001;13(2):113–6.10.1002/1521-4095(200101)13:2<113::AID-ADMA113>3.0.CO;2-HSearch in Google Scholar

[47] Williams G, Kamat PV. Graphene-semiconductor nanocomposites: excited-state interactions between ZnO nanoparticles and graphene oxide. Langmuir. 2009;25(24):13869–73.10.1021/la900905hSearch in Google Scholar

[48] Karthika K, Ravichandran K. Tuning the microstructural and magnetic properties of ZnO nanopowders through the simultaneous doping of Mn and Ni for biomedical applications. J Mate Sci Technol. 2015;31:1111–7.10.1016/j.jmst.2015.09.001Search in Google Scholar

[49] Azimirad R, Khayatian A, Safa S, Kashi MA. Enhancing photoresponsivity of ultra violet photodetectors based on Fe doped ZnO/ZnO shell/core nanorods. J Alloy Compd. 2014;615:227–33.10.1016/j.jallcom.2014.06.157Search in Google Scholar

[50] Afifah N, Djaja NF, Saleh R. Materials science forum. Trans Tech Publ. 2015;19–24.10.4028/www.scientific.net/MSF.827.19Search in Google Scholar

[51] Huang W, Tang X, Felner I, Koltypin Y, Gedanken A. Preparation and characterization of FexOy–TiO2 via sonochemical synthesis. Mater Res bull. 2002;37:1721–35.10.1016/S0025-5408(02)00821-8Search in Google Scholar

[52] Xiao S, Zhao L, Lian J. Enhanced photocatalytic performance of supported Fe doped ZnO nanorod arrays prepared by wet chemical method. Catal Lett. 2014;144:347–54.10.1007/s10562-013-1106-3Search in Google Scholar

[53] Ba-Abbad MM, Kadhum AAH, Mohamad AB, Takriff MS, Sopian K. Visible light photocatalytic activity of Fe3+-doped ZnO nanoparticle prepared via sol–gel technique. Chemosphere. 2013;91:1604–11.10.1016/j.chemosphere.2012.12.055Search in Google Scholar

[54] Sun T, Qiu J, Liang C. Controllable fabrication and photocatalytic activity of ZnO nanobelt arrays. J Physl Chem C. 2008;112:715–21.10.1021/jp710071fSearch in Google Scholar

[55] Jang ES, Won JH, Hwang SJ, Choy JH. Fine tuning of the face orientation of ZnO crystals to optimize their photocatalytic activity. Adv Mater. 2006;18:3309–12.10.1002/adma.200601455Search in Google Scholar

[56] Machado AEH, de Miranda JA, de Freitas RF, Duarte ETF, Ferreira LF, Albuquerque YD, et al. Destruction of the organic matter present in effluent from a cellulose and paper industry using photocatalysis. J Photoch Photobio A. 2003;155:231–41.10.1016/S1010-6030(02)00393-3Search in Google Scholar

[57] Rauf M, Meetani M, Hisaindee S. An overview on the photocatalytic degradation of azo dyes in the presence of TiO2 doped with selective transition metals. Desalination. 2011;276:13–27.10.1016/j.desal.2011.03.071Search in Google Scholar

[58] Matos J, Laine J, Herrmann J-M. Effect of the type of activated carbons on the photocatalytic degradation of aqueous organic pollutants by UV-irradiated titania. J Catal. 2001;200:10–20.10.1006/jcat.2001.3191Search in Google Scholar

[59] Eskizeybek V, Sarı F, Gülce H, Gülce A, Avcı A. Preparation of the new polyaniline/ZnO nanocomposite and its photocatalytic activity for degradation of methylene blue and malachite green dyes under UV and natural sun lights irradiations. Appl Catal B-Environ. 2012;119:197–206.10.1016/j.apcatb.2012.02.034Search in Google Scholar

[60] Kitture R, Koppikar SJ, Kaul-Ghanekar R, Kale S. Catalyst efficiency, photostability and reusability study of ZnO nanoparticles in visible light for dye degradation. J Phys Chem Solids. 2011;72:60–6.10.1016/j.jpcs.2010.10.090Search in Google Scholar

Received: 2022-03-22
Revised: 2022-08-06
Accepted: 2022-08-22
Published Online: 2022-09-16

© 2022 Tahani Saad Algarni et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Kinetic study on the reaction between Incoloy 825 alloy and low-fluoride slag for electroslag remelting
  3. Black pepper (Piper nigrum) fruit-based gold nanoparticles (BP-AuNPs): Synthesis, characterization, biological activities, and catalytic applications – A green approach
  4. Protective role of foliar application of green-synthesized silver nanoparticles against wheat stripe rust disease caused by Puccinia striiformis
  5. Effects of nitrogen and phosphorus on Microcystis aeruginosa growth and microcystin production
  6. Efficient degradation of methyl orange and methylene blue in aqueous solution using a novel Fenton-like catalyst of CuCo-ZIFs
  7. Synthesis of biological base oils by a green process
  8. Efficient pilot-scale synthesis of the key cefonicid intermediate at room temperature
  9. Synthesis and characterization of noble metal/metal oxide nanoparticles and their potential antidiabetic effect on biochemical parameters and wound healing
  10. Regioselectivity in the reaction of 5-amino-3-anilino-1H-pyrazole-4-carbonitrile with cinnamonitriles and enaminones: Synthesis of functionally substituted pyrazolo[1,5-a]pyrimidine derivatives
  11. A numerical study on the in-nozzle cavitating flow and near-field atomization of cylindrical, V-type, and Y-type intersecting hole nozzles using the LES-VOF method
  12. Synthesis and characterization of Ce-doped TiO2 nanoparticles and their enhanced anticancer activity in Y79 retinoblastoma cancer cells
  13. Aspects of the physiochemical properties of SARS-CoV-2 to prevent S-protein receptor binding using Arabic gum
  14. Sonochemical synthesis of protein microcapsules loaded with traditional Chinese herb extracts
  15. MW-assisted hydrolysis of phosphinates in the presence of PTSA as the catalyst, and as a MW absorber
  16. Fabrication of silicotungstic acid immobilized on Ce-based MOF and embedded in Zr-based MOF matrix for green fatty acid esterification
  17. Superior photocatalytic degradation performance for gaseous toluene by 3D g-C3N4-reduced graphene oxide gels
  18. Catalytic performance of Na/Ca-based fluxes for coal char gasification
  19. Slow pyrolysis of waste navel orange peels with metal oxide catalysts to produce high-grade bio-oil
  20. Development and butyrylcholinesterase/monoamine oxidase inhibition potential of PVA-Berberis lycium nanofibers
  21. Influence of biosynthesized silver nanoparticles using red alga Corallina elongata on broiler chicks’ performance
  22. Green synthesis, characterization, cytotoxicity, and antimicrobial activity of iron oxide nanoparticles using Nigella sativa seed extract
  23. Vitamin supplements enhance Spirulina platensis biomass and phytochemical contents
  24. Malachite green dye removal using ceramsite-supported nanoscale zero-valent iron in a fixed-bed reactor
  25. Green synthesis of manganese-doped superparamagnetic iron oxide nanoparticles for the effective removal of Pb(ii) from aqueous solutions
  26. Desalination technology for energy-efficient and low-cost water production: A bibliometric analysis
  27. Biological fabrication of zinc oxide nanoparticles from Nepeta cataria potentially produces apoptosis through inhibition of proliferative markers in ovarian cancer
  28. Effect of stabilizers on Mn ZnSe quantum dots synthesized by using green method
  29. Calcium oxide addition and ultrasonic pretreatment-assisted hydrothermal carbonization of granatum for adsorption of lead
  30. Fe3O4@SiO2 nanoflakes synthesized using biogenic silica from Salacca zalacca leaf ash and the mechanistic insight into adsorption and photocatalytic wet peroxidation of dye
  31. Facile route of synthesis of silver nanoparticles templated bacterial cellulose, characterization, and its antibacterial application
  32. Synergistic in vitro anticancer actions of decorated selenium nanoparticles with fucoidan/Reishi extract against colorectal adenocarcinoma cells
  33. Preparation of the micro-size flake silver powders by using a micro-jet reactor
  34. Effect of direct coal liquefaction residue on the properties of fine blue-coke-based activated coke
  35. Integration of microwave co-torrefaction with helical lift for pellet fuel production
  36. Cytotoxicity of green-synthesized silver nanoparticles by Adansonia digitata fruit extract against HTC116 and SW480 human colon cancer cell lines
  37. Optimization of biochar preparation process and carbon sequestration effect of pruned wolfberry branches
  38. Anticancer potential of biogenic silver nanoparticles using the stem extract of Commiphora gileadensis against human colon cancer cells
  39. Fabrication and characterization of lysine hydrochloride Cu(ii) complexes and their potential for bombing bacterial resistance
  40. First report of biocellulose production by an indigenous yeast, Pichia kudriavzevii USM-YBP2
  41. Biosynthesis and characterization of silver nanoparticles prepared using seeds of Sisymbrium irio and evaluation of their antifungal and cytotoxic activities
  42. Synthesis, characterization, and photocatalysis of a rare-earth cerium/silver/zinc oxide inorganic nanocomposite
  43. Developing a plastic cycle toward circular economy practice
  44. Fabrication of CsPb1−xMnxBr3−2xCl2x (x = 0–0.5) quantum dots for near UV photodetector application
  45. Anti-colon cancer activities of green-synthesized Moringa oleifera–AgNPs against human colon cancer cells
  46. Phosphorus removal from aqueous solution by adsorption using wetland-based biochar: Batch experiment
  47. A low-cost and eco-friendly fabrication of an MCDI-utilized PVA/SSA/GA cation exchange membrane
  48. Synthesis, microstructure, and phase transition characteristics of Gd/Nd-doped nano VO2 powders
  49. Biomediated synthesis of ZnO quantum dots decorated attapulgite nanocomposites for improved antibacterial properties
  50. Preparation of metal–organic frameworks by microwave-assisted ball milling for the removal of CR from wastewater
  51. A green approach in the biological base oil process
  52. A cost-effective and eco-friendly biosorption technology for complete removal of nickel ions from an aqueous solution: Optimization of process variables
  53. Protective role of Spirulina platensis liquid extract against salinity stress effects on Triticum aestivum L.
  54. Comprehensive physical and chemical characterization highlights the uniqueness of enzymatic gelatin in terms of surface properties
  55. Effectiveness of different accelerated green synthesis methods in zinc oxide nanoparticles using red pepper extract: Synthesis and characterization
  56. Blueprinting morpho-anatomical episodes via green silver nanoparticles foliation
  57. A numerical study on the effects of bowl and nozzle geometry on performances of an engine fueled with diesel or bio-diesel fuels
  58. Liquid-phase hydrogenation of carbon tetrachloride catalyzed by three-dimensional graphene-supported palladium catalyst
  59. The catalytic performance of acid-modified Hβ molecular sieves for environmentally friendly acylation of 2-methylnaphthalene
  60. A study of the precipitation of cerium oxide synthesized from rare earth sources used as the catalyst for biodiesel production
  61. Larvicidal potential of Cipadessa baccifera leaf extract-synthesized zinc nanoparticles against three major mosquito vectors
  62. Fabrication of green nanoinsecticides from agri-waste of corn silk and its larvicidal and antibiofilm properties
  63. Palladium-mediated base-free and solvent-free synthesis of aromatic azo compounds from anilines catalyzed by copper acetate
  64. Study on the functionalization of activated carbon and the effect of binder toward capacitive deionization application
  65. Co-chlorination of low-density polyethylene in paraffin: An intensified green process alternative to conventional solvent-based chlorination
  66. Antioxidant and photocatalytic properties of zinc oxide nanoparticles phyto-fabricated using the aqueous leaf extract of Sida acuta
  67. Recovery of cobalt from spent lithium-ion battery cathode materials by using choline chloride-based deep eutectic solvent
  68. Synthesis of insoluble sulfur and development of green technology based on Aspen Plus simulation
  69. Photodegradation of methyl orange under solar irradiation on Fe-doped ZnO nanoparticles synthesized using wild olive leaf extract
  70. A facile and universal method to purify silica from natural sand
  71. Green synthesis of silver nanoparticles using Atalantia monophylla: A potential eco-friendly agent for controlling blood-sucking vectors
  72. Endophytic bacterial strain, Brevibacillus brevis-mediated green synthesis of copper oxide nanoparticles, characterization, antifungal, in vitro cytotoxicity, and larvicidal activity
  73. Off-gas detection and treatment for green air-plasma process
  74. Ultrasonic-assisted food grade nanoemulsion preparation from clove bud essential oil and evaluation of its antioxidant and antibacterial activity
  75. Construction of mercury ion fluorescence system in water samples and art materials and fluorescence detection method for rhodamine B derivatives
  76. Hydroxyapatite/TPU/PLA nanocomposites: Morphological, dynamic-mechanical, and thermal study
  77. Potential of anaerobic co-digestion of acidic fruit processing waste and waste-activated sludge for biogas production
  78. Synthesis and characterization of ZnO–TiO2–chitosan–escin metallic nanocomposites: Evaluation of their antimicrobial and anticancer activities
  79. Nitrogen removal characteristics of wet–dry alternative constructed wetlands
  80. Structural properties and reactivity variations of wheat straw char catalysts in volatile reforming
  81. Microfluidic plasma: Novel process intensification strategy
  82. Antibacterial and photocatalytic activity of visible-light-induced synthesized gold nanoparticles by using Lantana camara flower extract
  83. Antimicrobial edible materials via nano-modifications for food safety applications
  84. Biosynthesis of nano-curcumin/nano-selenium composite and their potentialities as bactericides against fish-borne pathogens
  85. Exploring the effect of silver nanoparticles on gene expression in colon cancer cell line HCT116
  86. Chemical synthesis, characterization, and dose optimization of chitosan-based nanoparticles of clodinofop propargyl and fenoxaprop-p-ethyl for management of Phalaris minor (little seed canary grass): First report
  87. Double [3 + 2] cycloadditions for diastereoselective synthesis of spirooxindole pyrrolizidines
  88. Green synthesis of silver nanoparticles and their antibacterial activities
  89. Review Articles
  90. A comprehensive review on green synthesis of titanium dioxide nanoparticles and their diverse biomedical applications
  91. Applications of polyaniline-impregnated silica gel-based nanocomposites in wastewater treatment as an efficient adsorbent of some important organic dyes
  92. Green synthesis of nano-propolis and nanoparticles (Se and Ag) from ethanolic extract of propolis, their biochemical characterization: A review
  93. Advances in novel activation methods to perform green organic synthesis using recyclable heteropolyacid catalysis
  94. Limitations of nanomaterials insights in green chemistry sustainable route: Review on novel applications
  95. Special Issue: Use of magnetic resonance in profiling bioactive metabolites and its applications (Guest Editors: Plalanoivel Velmurugan et al.)
  96. Stomach-affecting intestinal parasites as a precursor model of Pheretima posthuma treated with anthelmintic drug from Dodonaea viscosa Linn.
  97. Anti-asthmatic activity of Saudi herbal composites from plants Bacopa monnieri and Euphorbia hirta on Guinea pigs
  98. Embedding green synthesized zinc oxide nanoparticles in cotton fabrics and assessment of their antibacterial wound healing and cytotoxic properties: An eco-friendly approach
  99. Synthetic pathway of 2-fluoro-N,N-diphenylbenzamide with opto-electrical properties: NMR, FT-IR, UV-Vis spectroscopic, and DFT computational studies of the first-order nonlinear optical organic single crystal
Downloaded on 6.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/gps-2022-0077/html
Scroll to top button