Paper The following article is Open access

GAD plasma-assisted synthesis of ZnO nanoparticles and their photocatalytic activity

, , , , and

Published 10 January 2024 © 2024 The Author(s). Published by IOP Publishing Ltd
, , Citation Ridha Messai et al 2024 Mater. Res. Express 11 015006 DOI 10.1088/2053-1591/ad1a82

2053-1591/11/1/015006

Abstract

In this study we present an efficient method for synthesizing highly pure ZnO nanoparticles using a Gliding Arc Discharge (GAD) plasma system as a non-thermal plasma source. This approach offers distinct advantages over conventional techniques, including simplicity, a short synthesis time, utilization of readily available air as the source gas, and potential scalability, rendering it a promising alternative for sustainable ZnO nanoparticle production. The synthesized nanoparticles physicochemical properties were characterized using various techniques, such as scanning electron microscopy (SEM), energy dispersive x-ray analysis (EDAX), UV-visible spectroscopy (UV–vis), Fourier-transform infrared spectroscopy (FTIR), x-ray diffraction (XRD), Thermogravimetric Analysis (TGA), and Differential Scanning Calorimetry (DSC). Furthermore, we evaluated the effectiveness of the synthesized ZnO nanoparticles for wastewater treatment by assessing their photocatalytic activity against methylene blue (MB), Brilliant Cresyl Blue (BCB), and Congo Red (CR) under UV light irradiation for 2 h and 30 min. The results confirmed the successful synthesis of highly pure ZnO nano-powder with an average size of 27.18 nm and a band gap energy of 3.28 eV in an exceptionally brief duration and through straightforward steps. Additionally, GAD plasma-assisted ZnO nanoparticles exhibited a significant dye removal rate, showcasing their potential as highly effective materials for photocatalytic wastewater treatment. This study contributes new insights into the application of GAD plasma for nanoparticle synthesis.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Over the last two decades, nanotechnology has revolutionized our modern life in countless ways, from improving the efficacy of medical treatments to enhancing the performance of electronics and renewable energy technologies. These tiny structures, typically less than 100 nanometers, have unique properties due to their small size and high surface area-to-volume ratio, making them attractive for a wide range of applications. By manipulating and engineering nanomaterials at the atomic and molecular level, scientists are able to create materials with specific unique optical, electrical, magnetic, or catalytic properties that were previously impossible to achieve [1].

Therefore, exploiting nanoparticles has opened up a world of possibilities for researchers and innovators seeking to develop new materials with desirable properties. Zinc oxide stands out as one of the most remarkable nanomaterials in terms of its properties and potential applications. These nanoparticles exhibit a unique combination of attributes, including high transparency and a wide band gap [2], making them excellent candidates for optoelectronic devices such as solar cells, sensors, and LEDs. Additionally, their outstanding catalytic properties make them useful in chemical synthesis applications, particularly in environmental remediation, where they have been reported to remove heavy metals, organic compounds, and even microorganisms through adsorption or photocatalytic degradation [35]. Furthermore, zinc oxide nanoparticles have demonstrated electrochemical sensing and antimicrobial properties, opening up possibilities for their use in medical applications such as drug delivery and wound healing [5, 6].

Several biological, physical, and chemical methods have been employed for synthesizing zinc oxide nanoparticle. These methods include vapor deposition [7], where zinc precursor compounds are vaporized and condensed to form nanoparticles. Laser ablation [8] involves using laser pulses to vaporize zinc and collect the resulting nanoparticles. Sol gel utilizes hydrolysis and condensation reactions to transform a precursor solution into a gel, which is then further processed [9]. Spray pyrolysis atomizes a precursor solution into droplets that are rapidly heated, resulting in the formation of solid nanoparticles [10]. Hydrothermal/solvothermal synthesis involves the controlled reaction between a precursor solution and a solvent under high temperature and pressure conditions [6, 11]. Lastly, biosynthesis methods utilize plant extracts or bacteria to reduce and stabilize zinc ions, facilitating the formation of nanoparticles with desired properties [1214].

These limitations and concerns can be efficiently overcome by utilizing non-thermal plasma (NTP) sources to generate reactive species for oxidizing metallic ions, thus producing safe nanoparticles without hazardous chemicals or solvents. The plasma method enables the production of nanoparticles with a high degree of purity and can be carried out at relatively low temperatures, making it suitable for temperature-sensitive materials. This method is distinguished by its short synthesis time compared to the other methods, which makes it highly flexible and scalable for industrial applications [15].

Previous studies have explored various techniques, such as PLD, sputtering, radio frequency (RF) plasma, microwave plasma, and atmospheric pressure plasma, for synthesizing ZnO nanoparticles [16, 17]. These methods require high power and elevated temperatures, making them unsuitable for temperature-sensitive materials. Additionally, it necessitates a cooling system, involves complex equipment, operates with a mixed gas at low pressure, and demands effective equipment sealing. Despite the advancements, there is still a research gap in the realm of plasma-based synthesis of ZnO nanoparticles. To further illustrate the advantages of GAD plasma, table 1 provides a comprehensive comparison of key parameters between GAD plasma, PLD, sputtering, and other plasma technologies used for ZnO nanoparticle synthesis [1722]. This table illustrates the efficiency of the GAD plasma method, showcasing lower operating temperatures that are conducive to temperature-sensitive materials and contribute to enhanced sustainability. The GAD plasma method provides a high degree of controllability and tunability across various parameters, facilitating precise manipulation of ZnO nanoparticle characteristics tailored for specific applications.

Table 1. Comparison of different physical process used for synthesizing ZnO nanoparticles.

MethodZn Salt PrecursorDischarge ParametersAverage Particle Size nmZnO MorphologyApplicationReferences
  Gas PressurePower kWSynth time minSynth temp KYield    
Cold Plasma GADZn(CH3COO)2 H2OAir0.960Room temp90%27.18pseudo-spheres, and non-uniform hexagonal particlesPhoto-degradation of organic dyesThis work
  Pr = 1 bar        
PLD MethodO2  30–45150–180 [17]
  Pr = 10–2 mbar        
RF plasma torchAr/ O2 P = 30 kW f = 3.5 MHz300 K with water cooling100–200Spherical and hexagonal ZnO wurtzite structure [18]
  Pr = 0.2 bar        
Micro-wave plasmaZn(CH3)2 80% Ar 20% O2 P = 60 W90>1800 K04–08spherical [19]
  Pr = 30 mbar        
RF inductively coupled thermal plasmaZn(CH3)2 Ar asP = 28–31 kW f = 1–3 MHz15–901300 K60–70 at Pr1 30–50 at Pr2 200–500 at Pr3polygonal at Pr1 round at Pr2 flower-like with hexagonal nanorods at Pr3[20]
  Pr = 0.47, 0.87 and 0.93 Bar        
AC (pulse) plasma jetZn(NO3)2 ZnCl2 ZnSO4· 7 H2OAirP = 0.063 kW f = 1–3 MHz60310 KWeak400–1600 (ZN)rods, pillar and, flower-like for ZNAnti-bacterial activity[21]
  Pr = 1 bar    150–1000 (ZC)flower-like and rods for ZC  
       500–1000 (ZS)layered and aggregate structures for ZS  
AC sinusoidal DBD plasmaZn(CH3COO)2 H2OH2 U = 15KV f = 38 MHz05–151000 K90–150spherical[22]
  Pr = 1 bar        
RF Sputtering MethodAr/ O2 P = 180 W150–180[17]
  Pr = 10–3 mbar        

Notably, existing studies lack exploration of the GAD plasma method for ZnO nanoparticle synthesis. We present the first work utilizing this promising approach, which offers several advantages: (i) utilizing air as the source gas, promoting environmentally friendly synthesis; (ii) generating a high concentration of reactive species like OH radicals, potentially enhancing particle growth and crystallinity; (iii) operating at atmospheric pressure, simplifying process implementation; and (iv) eliminating the need for chemical products, leading to greener production. Given its unique benefits, this GAD plasma method stands as a powerful tool for producing ZnO nanoparticles with desirable properties. These nanoparticles hold significant potential for diverse applications, including photocatalytic degradation of pollutants, next-generation electronics, and targeted drug delivery in biomedicine.

Therefore, this study aims to investigate the potential of the GAD plasma method for synthesizing ZnO nanoparticles and to characterize their physicochemical properties using techniques such as SEM, EDAX, UV–vis, FTIR, TGA/DSC, and XRD. Furthermore, the photocatalytic activity of the synthesized ZnO nanoparticles was tested against BCB, MB, and CR to evaluate their effectiveness in wastewater treatment.

2. Preparation techniques and materials

2.1. Chemicals

All of the chemicals used, including zinc acetate Zn(CH3COO)2 H2O, Congo Red (CR) (C32H22N6Na2O6S2) (696,663 g mol−1), Brilliant Cresyl Blue (BCB) (C17H20ClN3O)2 (385.96 g mol−1), and methylene bleu (MB) (C16H18ClN3S) (319,852 g mol−1) were purchased from Sigma-Aldrich and used exactly as specified. Demineralized water was used to prepare all of the solutions for the zinc oxide NP formation and the photodegradation studies.

2.2. Sample preparation

The materials for the experiment were prepared with high precision. To start, 20 grams of Zn(CH3COO)2·2H2O was dissolved in 1000 milliliters of demineralized water, using a magnetic stirrer, to create a zinc acetate solution. This solution was stirred for 10 min at room temperature, ensuring the mixture was well combined. The resulting solution was designed to yield approximately 6.6 grams of pure ZnO nanoparticles, which was required for the experiment. Once the zinc acetate solution was prepared, individual solutions of CR, MB, and BCB dyes were methodically formulated. Each solution was created to have a concentration of 7 ppm and was rigorously agitated for 30 min to ensure the dyes were completely dissolved.

2.3. Gliding arc discharge (GAD) apparatus

In this work, ZnO nanoparticle synthesis utilizes a custom-designed gliding arc discharge (GAD) reactor (figure 1), drawing inspiration from various configurations reported in the literature for treating polluted liquids [2332]. This setup operates on the principle of generating a non-thermal plasma plume at atmospheric pressure.

Figure 1.

Figure 1. GAD- reactor.

Standard image High-resolution image

The key components and their roles:

  • A compressed air saturator: ensures adequate water content in the air stream entering the reactor.
  • Inlet nozzle (1 mm diameter): controls the plasma gas flow and its interaction with the electrodes.
  • Diverging electrodes with a 3 mm gap: Generate the electric arc and define the plasma plume volume.
  • A high-voltage transformer from Aupem Sefli, with a voltage setting of 220 V/9 kV: provides the necessary voltage (9 kV) to initiate and sustain the arc.
  • A plasma plume: sweeps along the electrode gap, generating active species.
  • The target solution: uses zinc acetate as a source of Zn2+ ions. Optimization considerations: GAD performance depends on several factors, which were optimized in previous studies and maintained constant in this work [2532]:
  • Nozzle diameter: Affects plasma gas flow and species diffusion towards the target.
  • Inter-electrode distance: Determines plasma plume size and stability.
  • Plasma gas: Air saturated with water ensures cost-effectiveness and the generation of desired radicals.
  • Gas flow rate: Set at 800 l/h for optimal plasma generation and interaction with the solution.
  • Power input: Maintained at 0.9 KW based on established power estimation methods [23, 24].

2.4. Synthesis of ZnO nanoparticle using gliding arc discharge

In our experiment, we started by passing compressed air through a bubbler to ensure that the air became saturated with water. Then, the airflow passed through an inlet nozzle with a diameter of 1 mm between two diverging electrodes. These electrodes, with a minimum gap of 3 mm, were connected to a high-voltage. This setup generated an alternating voltage of 9 kV and a current of 100 mA. The whole setup consumed 0.9 kW of power, which was impressively efficient. When we turned on the high voltage, an electric arc formed between the electrodes. This happened because of the voltage difference between the electrodes. The arc did not stay in one place; instead, it moved away from where it started due to the flowing air. It traveled along the space between the electrodes, creating a larger area filled with a special kind of glowing air called plasma. This process happened repeatedly, with each new arc following the same steps as the previous one.

The result of this process is the creation of a non-thermal plasma under atmospheric pressure. This plasma generates various species, including OH, O, H, NO, and HNOOH. In the subsequent phase of the experimental system, a flask containing zinc acetate is positioned 5 cm below the upper part of the plasma column. This arrangement is maintained for a duration of 60 min, resulting in the development of an opaque solution. This opaqueness serves as an indicator that the chemical constituents have participated in a redox mechanism, effectively converting ${{Zn}}^{2+}$ into ${{Zn}({OH})}_{2}.$

Following this treatment, the sample undergoes evaporation and drying at 100 °C for a span of 10 min to eliminate water content. Subsequently, a critical annealing process is undertaken at 600 °C, spanning a duration of 3 h. This annealing procedure serves multiple purposes, including the removal of organic components, the initiation of crystallization processes, and the production of zinc oxide nanoparticles.

Research indicates that temperatures at 600 °C play a crucial role in facilitating the complete decomposition of the precursor, resulting in the formation of well-defined ZnO crystals. Moreover, as the temperature increases, the crystalline shape transforms into a hexagonal structure [33]. The annealing process at 600 °C is instrumental in achieving optimal degradation of the organic group attached to the precursor, simultaneously fostering favorable crystallization, a conclusion corroborated by additional studies [34].

These processes were iterated ten times, with each iteration involving the utilization of 100 ml of solution. The aim was to achieve a cumulative quantity of approximately 6.6 g of pure ZnO nanoparticles.

2.5. Characterization

Powder XRD data were obtained using a Proto AXRD-2 diffractometer with sample spinner and a Dectris Mythen 1 K (3.22° active length*) 1D-detector in Bragg–Brentano geometry with a Copper Line Focus x-ray tube and Ni kβ absorber. The Thermo Scientific TM Quattro SEM combines all-around imaging and analytics performance with an innovative environmental mode (ESEM) that allows materials to be examined in their natural condition. The powder samples' optical characteristics were measured using a Shimadzu 3101PC double-beam spectrophotometer. The spectrophotometer (Shimadzu model IR Iraffinity-1) measured FTIR absorption in the spectral region 400–4000 cm−1. To discover the photocatalytic activity of the NPs, the Shimadzu Spectrophotometer UV-1800 was employed. METTLER TOLEDO's Simultaneous Thermal Analysis Device (TGA/DSC): TGA/DSC 3+ 1600 °C has been used to track changes in mass and thermal flow as a function of temperature. It allows users to evaluate a wide range of samples. The ability to conduct analyses in a controlled environment at temperatures ranging from 25 to 1600 °C (air, nitrogen or argon). Heater rate: 0.1 to 100 °C min−1. Balance: 5 g measurement range with 1 ug creams available: aluminum and platinum creams (70 ul) and huge aluminum creams (600 and 900 ul). LUXMETER 400,000 LUX - USB used for measuring sunlight intensity.

2.6. Photocatalytic activity

Before the photocatalytic process, the dispersion was stirred constantly in the dark room for 30 min to allow the dye particles and the surface of the catalyst to reach a state of equilibrium between adsorption and desorption. After the dispersion was stirred for 30 min, it was exposed to UV light for 150 min. Every 15 min, a beaker was removed, and 4 ml was centrifuged to separate the nanoparticulate suspensions (NPs) of ZnO from the solution. The irradiation of a 365 nm, 30 W UV lamp was assured by placing it 15 cm above the solution at room temperature. Ten samples (25 ml) of each aqueous solution dye (CR, MB, and BCB) with initial concentrations of 7 ppm were mixed with 25 mg of prepared ZnO in an open glass (50 ml) used as a photo-reactor. The concentration of the dye was determined using a UV-visible spectrophotometer with wavelengths of 498 nm and 625 nm, respectively. The samples were taken at the prescribed time, centrifuged, and the dye concentration was measured at 498 nm and 625 nm. Using the following formula (equation (1)), the rate of degradation was determined.

Equation (1)

Where C0 represents the initial concentration (mg/l) and Ct represents the concentration of the dye at time t.

3. Results and discussion

3.1. Chemical interactions of gliding arc plasma with zinc acetate solution

A gliding arc discharge is a type of ionized gas that is created by the passage of gas through an electric current. The gas is typically air, and it is often humidified to increase its conductivity. The arc moves along the surface of the electrodes, and it produces a variety of chemical species, including radicals, ions, and molecules. The chemical species that are produced in a gliding arc discharge depend on the gas composition, the humidity, and the operating parameters. In humid air (N2, CO2 , O2, H2O), the primary chemical species are OH, O, H, NO, O3, ${{CO}}_{3}^{2-},$ and CH3COOH [3537]. These species are formed as a result of electron impact, dissociation, addition, transformation, collision events that occur in accordance with the gas vector, and chemical reactions.

The OH and NO radicals are of particular interest because they are very reactive. They can disperse in the treated target, and inducing both acidifying and oxidizing effects in aqueous solutions, react with other molecules to form new compounds, or they can initiate chemical reactions. This makes gliding arc discharge a promising technology for a variety of applications, such as water treatment, waste treatment, and material synthesis.

As a result of electron impact within the arc, the primary formation of NO and OH radicals is explained through (equations (2–4)) [37].

Equation (2)

Equation (3)

Equation (4)

These species are capable of diffusing in aqueous solutions and interacting with molecules. The hydroxyl radical is the principal species responsible for the production of metal oxide nanoparticles [38], whose standard potential is E°(OH/H2O) = 2.85 V N−1H−1E [39]. The hydroxyl radical OH is a very reactive species that can react with other molecules. It can also react with an electron received from ions and molecules dissolved in water as impurities, as shown in (equation (5)). This reaction produces a hydroxide ion (OH-) [40].

Equation (5)

The dissolution of zinc acetate in water is expressed by the following, (equation (6)) and (equation (7)) [41].

Equation (6)

Equation (7)

The zinc ion reacts with hydroxide ion and then zinc hydroxide is produced as (equation (8))

Equation (8)

Zinc oxide tends to coagulate and deposit in water, so we perform the vaporization process as follows: (equation (9))

Equation (9)

The reactions mentioned from (equations (29)) demonstrate the fundamental process of ZnO NPs creation by gliding arc discharge with the air as vector gas.

3.2. Characterization of synthesized ZnO

3.2.1. UV-visible

ZnO is a transition oxide with a hexagonal structure and a direct bandgap of 3.28 eV for n-type semiconductor materials. For example, the direct band gap of ZnO NPs is between 3.10 and 3.37 eV. Many parameters, including crystallinity, grain size, particle size, particle shape, and dislocation, affect the band gap of ZnO NPs. The UV–vis spectrum in figure 2 shows the absorption of ZnO nanoparticles produced using the GAD reactor. The high absorption peak at 362 nm is due to the surface plasmon resonance of ZnO NPs [42, 43].

Figure 2.

Figure 2. Absorbance spectra of ZnO NPs.

Standard image High-resolution image

The following equation (equation (10)) calculates the nanoparticles' direct transition optical band gap energy [44]:

Equation (10)

where:

α is the absorption coefficient,

Eg is the optical band gap of the NPs,

hν is the photon energy,

A is a constant independent on hν.

As a result, Tauc's plot is used in order to compute the optical band gap. This is accomplished by plotting (hv)2 versus (hv) and then extrapolating the linear section in order to determine the energy axis intercept figure 3(a). Synthesized ZnO NPs have an optical band gap of 3.28 eV, compared to bulk ZnO. According to the literature review [45], the band gap varied according to several parameters, such as the ZnO crystal defect, particle shape, and particle size.

Figure 3.

Figure 3. (a) Plot of (αhν)2 versus photon energy of synthesized ZnO. (b) Inset shows the Urbach energy plot of ln(α) versus(hν).

Standard image High-resolution image

The optical, electrical, and catalytic properties of ZnO nanoparticles are influenced by the anisotropy of their band gap. The band gap refers to the energy difference between the valence and conduction bands of electrons. The variation in the ZnO nanoparticle band gap directly affects electrical conductivity, with ZnO having a wide band gap. By reducing the band gap, electrical conductivity can be improved, which is crucial for applications such as transparent conductive films, solar cells, and electronics. Furthermore, the band gap difference plays a significant role in determining ZnO's redox and catalytic properties, as the surfaces of ZnO nanoparticles catalyze various chemical reactions. The presence of oxygen vacancies enhances catalytic activity. Overall, ZnO nanoparticles possess the ability to clean, convert energy, and degrade pollutants due to their unique band gap properties.

Band tailing in the band gap commonly results from semiconductor impurity incorporation. Since optical transitions take place from occupied states in the valence band tail to unoccupied ones at the conduction band edge, the absorption coefficient α at the band edge is exponentially dependent on the photon energy. The empirical Urbach (equation (11)) describes local defects' band tail energy (E00) [46]:

Equation (11)

Where αo is a constant and E00 is the gap area tail width of localized states. To calculate αo and Ε00, plot the logarithmic scale of the absorption coefficient as a function of photon energy hν. The slope of log (α) versus incident photon energy hν determines the Urbach tail [47]. From the inverse slope of the linear plot of ln(α) versus hν in figure 3(b), ZnO NPs' Urbach energy is 0.430 eV. Possible explanations for this result include structural and thermal disturbances.

3.2.2. FTIR spectroscopy

In order to identify the functional groups for ZnO NPs, FTIR analysis is used. The spectra of the treated zinc acetate solution by non-thermal plasma (GAD) before and after annealing at 600 °C for 3 h are shown in figure 4. The wavenumbers of the peaks that appeared in the FTIR and their functional origins are shown in table 2. As a result of plasma treatment of zinc acetate solution, FTIR spectra before annealing treatment revealed a number of absorption bands (697, 954, 1037, 1045, 1386, 1451, and 1542 cm−1) corresponding to the functional groups of the organic molecules [48]. The broad and strong bands at 1451, and 1542 cm−1 attributed to nitrogen-bonded (O=N, and C=N) respectively. The bands at 1037 and 1045 cm−1 can be attributed to the (C=O) stretching and C-O-C stretching asymmetric vibrations and the (C-N) stretching vibrations of the aromatic amine.

Figure 4.

Figure 4. FTIR spectra of ZnO NPs and inset shows the enlarged spectra in the range <600 cm−1.

Standard image High-resolution image

Table 2. IR peaks and their assignments for the NPs of ZnO system.

AssignmentsWavenumber (cm−1)Wave number (cm−1)
 Before annealingAfter annealing
Zn-O Stretching< 600< 600
N-O Stretching697Absent
C-N954Absent
C=O1037Absent
C-O-C/C-N1045Absent
O-H1386Absent
O=N1451Absent
C=N1542Absent
O≡N20102009
C≡N21582156

After annealing, the FITR spectrum of the ZnO NPs shows a sharp peak at 448 cm−1 associated with Zn-O vibration and a broad peak at 3430 cm−1 related to the O-H bonding due to moisture absorption.

In wurtzite hexagonal type ZnO crystals, the Zn-O stretching vibration of the oxygen sublattice (${E}_{2}H$) is responsible for the 426 and 565 cm−1 absorption bands, while the 1386 cm−1 oxygen vacancies are characterized as hydrogen-related defects on the surface of ZnO [49].

3.2.3. X-ray diffraction

An x-ray diffractometer was used to study the structure and elements of ZnO nanoparticles, and the results are displayed in figure 5. As can be observed, there are many peaks that correspond to the planes (100), (002), (101), (102), (110), and (103) for ZnO that is hexagonal in shape. The treated sample was made up of tiny particles, according to the XRD data.

Figure 5.

Figure 5. X-ray diffraction pattern of ZnO NPs.

Standard image High-resolution image

Figure 5 displays the XRD pattern of the ZnO NPs sample after preparation. The ZnO single crystal phase is verified by the diffraction pattern. The crystallographic plane (101) exhibits a hexagonal crystal structure, and all the diffraction peaks observed in the patterns (100), (002), (101), (102), (110), (103), (200), (112), (201), (004), and (202) indicate a polycrystalline nature. These values closely correspond to those documented in the standard card (JCPDS number: 01-089-0551) [50].

Table 3 shows the ZnO NPs' crystallite size of 27.18 nm, calculated using Scherrer's formula (equation (12)) [51].

Equation (12)

The wavelength for a hexagonal crystal structure is λ =1.54 Å, the full-width at half-maximum (FWHM) is β, k = 0.91, and θ is the measured angle at the peak (101).

Table 3. Estimation of structural parameters of ZnO NPs.

Sample name2θ (deg)FWHM (deg)D (nm)Lattice parameter ($\hat{{\bf{A}}}$) ε × 10–4 σ(GPa) ξ × 1014 (line/m2)
ZnO pure36.3180.295227.18a = bc1.921−0.8611.35
    3.24945.2051   

In addition, the peak location deviation from the reference data of bulk ZnO reveals the presence of stress in the NPs. These equations (equations (13 and14)) were used to calculate an approximation of the stress level:

Equation (13)

Equation (14)

Where, σ is strain, C33(210 GPa) , C11(210 GPa), C12(120 GPa), and C13 (105 GPa) are the elastic stiffness constants, ε is the uniform strain, ZnO NPs have a lattice parameter denoted by the letter c, which may be determined from the plane labeled (101), and the lattice parameter for the bulk material is denoted by c0 [52]. Table 3 provides approximated values for 2θ, lattice parameter, FWHM, crystallite size, stress, strain, and dislocation. The negative signs of the stress suggest the existence of compressive stress [53]. This compressive stress is usually caused by distortions and defects in the lattice.

3.2.4. Scanning electron microscopy (SEM) and EDAX

Scanning electron microscopy (SEM) was employed to investigate the formation, size, shape, and morphology of the prepared sample's particles. Figures 6(a), and (b) show low-magnification SEM images revealing the typical microstructure of ZnO NPs, which is mainly of different sizes and agglomerated grains of nanopowder which consist of pseudo-spheres and non-uniform hexagonal particles [54], and show particle aggregation related to the self-assembly of the ZnO nanoparticles [55]. Figure 6(c) depicts the ZnO NPs size distribution histogram; the averaged grain size is 52 ± 1.26 nm, which is larger than the measurements reported by XRD analyses. While XRD solely reveals data on the crystalline components and not the amorphous ones, FESEM analysis delivers insights into the overall shape of the grains [54]. The energy-dispersive analysis x-ray spectra (EDAX) reveal the chemical composition and atomic percentage of nanoparticles (NPs). Figure 6(d) shows the EDAX spectra obtained for a sample of ZnO NPs. The elements identified are zinc and oxygen; however, the peak of carbon is not considered as it is absorbed by the superficial area of the ZnO NPs.

Figure 6.

Figure 6. SEM image of the ZnO NPs plasma synthesized with a magnification of (a) 35.00 KX; (b) 100.00 KX; (c) their corresponding particle size distribution and (d) Atomic concentration and Typical EDAX spectra.

Standard image High-resolution image

3.2.5. Thermogravimetric analysis TGA / differential scanning calorimetry (DSC)

The DSC/TGA curve in the (figure 7) shows the product's behavior when subjected to a 10 °C/min heating rate from 25 °C to 700 °C. The TGA profile indicates a weight reduction interrupted by three distinct shifts. Meanwhile, the DSC curve shows three endothermic peaks, which correspond to different stages of decomposition: the release of physically adsorbed water in the initial stage and the release of chemically adsorbed water and carbonic matter in the second and third stages, respectively [12, 56, 57]. These findings are consistent with the peaks observed in the FTIR analysis of these groups. It is showing negligible mass loss of about 0.42% as a function of temperature, indicating thermal stability between 25 °C and 700 °C. Hence, we can conclude that the sample is pure.

Figure 7.

Figure 7. TGA/DSC thermogram identifies the transformation of the synthesized ZnO.

Standard image High-resolution image

3.3. ZnO NPs photocatalytic activity

The effectiveness of synthesized ZnO NPs in removing organic pollutants from water was investigated using the cationic and anionic dyes BCB, MB, and CR (figure 8). The photo-degradation of these dyes was observed under UV light, and the results were analyzed in terms of contact time and the presence of ZnO NPs.

Figure 8.

Figure 8. (a) Absorbance spectra of dye solution under UV light irradiation, (b) photocatalytic degradation of dye at different times: (1) BCB, (2) CR and (3) MB.

Standard image High-resolution image

The findings of the study demonstrate that after 2 h and 30 min of exposure to UV light, the rates of degradation were 92.9%, 71%, and 88% for BCB, MB and CR dyes, respectively. It was also observed that the photo-degradation efficiency of BCB was higher than that of MB and CR. The kinetics of the photodegradation reaction of the dyes were determined using the pseudo-first order equation (equation (15)) [58]:

Equation (15)

Where Ct and C0 are the concentrations of dye at time t = 0 min and t, respectively, and k is the pseudo-first-order rate constant.

Figure 9 shows the relationship between ln (${C}_{0}$/${C}_{t}$) and the irradiation period for BCB, MB, and CR dyes. As seen in figure 9, a linear relationship was found between ln (${C}_{0}$/${C}_{t}$) and time, confirming that the photocatalytic degradation followed the first order mechanism. The value of the degradation rate constant (k) was estimated to be 0.0190 min−1, 0.0081 min−1, and 0.0086 min−1 for BCB, MB and CR, respectively, based on the slope of the linear equation utilizing the above equation.

Figure 9.

Figure 9. The plot of ln(C0/Ct) versus time for (a) BCB, (b) CR (c) and MB dyes.

Standard image High-resolution image

Table 4 presents a comparison of the photocatalytic activities of various nanoparticles in several previous studies and ZnO synthesized by GAD plasma, for the degradation of BCB, MB, and CR. The results of this comparison indicate that ZnO nanoparticles produced through Gliding Arc Discharge plasma exhibited higher photocatalytic activity compared to other synthesis methods [13, 5980].

Table 4. Comparative study of the degradation of CR, BCB and MB by photocatalysis of different nanoparticles and ZnO synthetizes by plasma.

Sample nameSynthesis methodSize nmGap Energy eVDyeExperimental conditionsCatalysis dose g/lDeg efficiency / TimeReferences
ZnO purehydrothermal method3.3CRV = 50 ml, 20 pmm,0.447%/ 1.16 h 41%/ 1.16 h[59]
NiO pure 3.0 pH = 30.4  
     400 W- Hg Lamp   
  CRV = 300 ml, 50 ppm, pH = 70.1697%[60]
ZnOCommercial   Sun-light 2 h 
  212.88CRV = 20 ml /120 ppm20099.9%[61]
ZnOBiosynthesis   T° =30 °C, 1.2 h 
     (λ > 254 nm- 6 W)   
TiO2-ZnOhydrothermal method17.4−4.18CRV = 100 ml, 1.0 mM,0.564.5%/ 1.17 h[62]
TiO2-ZnO/Ag 34.82.24 under visible light irradiation TiO2-ZnO 
     (150 W) 87.5% TiO2-ZnO/Ag 
TiO2, ZnOhydrothermal method44.13.16CRV = 100 ml, 1.0 mM,0.580%/1.17 h[62]
TiO2-ZnO 743.07 UV light (120 W)   
TiO2-ZnO/Ag 17.4−4.18     
  34.82.24     
ZnOCommercial7–30CRV = 5000 ml, 20 ppm, pH = 70.365%/ 8 h[63]
     UV light   
     (λ= 253.7 nm- 18 W)   
     Dark Room 30 min   
CuOhydrothermal method CRV = 100 ml, 20 ppm,0.5 [64]
nano-rods 16  UV light 67%/ 3.5 h 
nano-leaves 14  (λ= 256 nm- 18 W) 48%/ 3.5 h 
nano-sheets 18  Dark Room 30 min 12%/ 3.5 h 
Co@C-60126CRV = 50 ml, 100 ppm, pH = 3115.6% / 3 h UV + Co@C-60[65]
     UV light 98.1%/3 h UV + H2O2 + Co@C-60 
     (λ= 254 nm- 8 W)   
ZnOSol gel23.6CRV = 100 ml, 50 ppm, pH = 4194%/ 1 h[66]
     solar light   
     Dark Room 60 min   
ZnOgreen synthesis10–153.42CRV = 250 ml, 20 ppm, pH = 90.2490%/ 1 h[13]
     UV Light (λ = 254nm-30 W)   
     Dark Room 60 min   
TiO2Micro wave method11.923.4CRV = 50 ml, 100 ppm,0.190%/2 h/ UV light[67]
     UV Light (15 W) 99%/2 h/ Vis light 
     Visible light (400 W)   
     Dark Room 30 min   
ZnO pureCold Plasma27.183.28CRV = 25 ml, 7 ppm, T° = 20 °C, pH = 7188%This work
 GAD   UV Light / 2.5 h 
     (λ= 365 nm—30 W)   
     Dark Room 30 min   
TiO2CommercialBCBV = 100 ml, 3 ppm,2.674%[68]
     UV Light (200 W) /2 h 
     Dark Room 20 min   
Al2O3CommercialBCBV = 100 ml, 50 ppm, pH = 101.792.87%/1 h Al2O3+ UV + 10 cm3 /min air bubble[69]
     (8.44 mW cm−2 light intensity)   
     Dark Room 30 min   
g-C3N4/ ZnOCo-precipitation40BCBV = 50 ml, 25 ppm, pH = 100.599.51%/1 h[70]
 method   Visible Light   
     (250 w)   
     Dark Room 30 min   
Co3O4/Fe2O3 202.12BCBV = 100 ml, 38.6 ppm, pH = 10197% 3 h[71]
     Sun-light   
ZnO/CuOPrecipitation18 to 402.88BCBV = 100 ml, 15 ppm,0.1597.30% 1.66 h[72]
 method   Visible Light   
     (250 w)   
     Dark Room 30 min   
Al2O3 doped Mn3O4low-temperature stirring method5.30CBCV = 100 ml, 38.6 ppm, pH = 101.250%–65% / 5 h[73]
     Sun-light   
     Dark Room 30 min   
TiO2CommercialCBCV = 100 ml, 19.3 ppm, pH = 111.596% / 8 h[74]
     Visible light (500 W)   
     Dark Room 10 min   
ZnO pureCold Plasma27.183.28BCBV = 25 ml, 7 ppm, T° = 20 °C, pH = 7192.92%This work
 GAD   UV Light / 2.5 h 
     (λ= 365 nm—30 W)   
     Dark Room 30 min   
Ag/ZnOsol–gel method3.10MBV = 100 ml, 6.4 ppm,1100%[75]
     Visible lamp / 2.5 h 
     (250 W)   
     Dark Room 15 min   
ZnOPrecipitation method50–200MBV = 20 ml, 5 ppm,595%[76]
     UV Light (12 W) Dark Room 30 min / 1.5 h 
ZnOSol–Gel20–1002.69MB10 ppm,0.8695%[77]
     solar Light / 0.75 h 
     (300 W)   
     Dark Room 45 min   
ZnO pureCo-precipitation26.7 27.923.05MBV = 20 ml, 20 ppm,193%/4 h[78]
Ag1.7%/ZnOAg3.4%/ZnO Ag4.4%/ZnOmethod.27.40 25.603.02 UV-C lamp for UV 98%/4 h 
   2.98   93%/4 h 
   2.94   92%/4 h 
ZnO pureCo-precipitation26.7 27.923.05MBV = 20 ml, 20 ppm,143%/4 h[78]
Ag1.7%/ZnOAg3.4%/ZnO Ag4.4%/ZnOmethod.27.40 25.603.02 three Philips TLD 18 W/54–765 lamp for visible light 48%/4 h 
   2.98   36%/4 h 
   2.94   35%/4 h 
ZnOSol–Gel80–1003.54MBV = 20 ml, 10 ppm,0.2586%/1 h[79]
     UV lamp (365 nm −4 W)   
     Dark Room 60 min   
ZnAl2O4: Dy3+ Green183.4MBV = 100 ml, 10 ppm, pH = 9195%/3 h[80]
 combustion   UV light   
ZnO pureCold Plasma27.183.28MBV = 25 ml, 7 ppm, T° = 14 °C, pH = 7171%This work
 GAD   UV Light / 2.5 h 
     (λ= 365 nm—30 W)   
     Dark Room 30 min   

The findings of this research could suggest that the method of synthesizing ZnO NPs can significantly affect their photocatalytic activity, which in turn affects their efficiency in degrading organic pollutants in water. Therefore, the use of ZnO NPs generated by the plasma GAD of air may be a more effective approach for water treatment applications.

4. Conclusion

In this paper, we have developed a novel and effective method for synthesizing ZnO NPs using plasma GAD and air as the vector gas. The zinc acetate aqueous solution was subjected to gliding arc discharge plasma, which generated hydroxyl radicals. These radicals reacted with the zinc ions in the solution to form zinc hydroxide, which was then vaporized to form ZnO NPs.

The physical and chemical properties of the synthesized ZnO NPs were characterized using various analytical techniques, including UV–vis, FTIR, XRD, TGA, DSC, and SEM. The UV–vis analysis confirmed the band gap energy of ZnO NPs to be 3.3 eV. The FTIR analysis revealed the presence of absorption bands at 426 and 565 cm−1, which are characteristic of the hexagonal structure of ZnO NPs. The XRD analysis confirmed the wurtzite hexagonal structure of the ZnO NPs, with an estimated diameter of 27.18 nm. The TGA and DSC exhibited excellent thermal stability of the synthesized ZnO NPs.

The SEM images showed that the ZnO NPs had a porous structure with grains of varying sizes that were agglomerated and had a regular shape. The synthesized ZnO NPs were found to be highly effective in degrading organic pollutants in water when exposed to UV light. The photodegradation rates of BCB, MB, and CR dyes reached 92.9%, 71%, and 88%, respectively, which is a promising result for water treatment applications.

The simplicity and effectiveness of the plasma GAD to synthesize ZnO NPs make it a promising approach for the development of new and improved photocatalysts for the degradation of organic pollutants in water. Future work should investigate the synthesis of different nanostructures to further enhance the photocatalytic activity of ZnO NPs.

Acknowledgments

This work was supported by a grant from the Directorate General for Scientific Research and Technological Development (DG-SRTD) of the Ministry of Higher Education and Scientific Research of Algeria. This work was supported by the Deanship of Scientific Research, Vice Presidency for Graduate Studies and Scientific Research, King Faisal University, Saudi Arabia [Project No: GRANT5242].

Data availability statement

No new data were created or analysed in this study.

Declarations

Ethical approval

This study did not involve the use of human or animal subjects. As such, no ethical approval was required for this research.

Conflict of interest

The authors declare that they have no conflict of interest.

Authors' contributions

M, W, A: R M writing the manuscript, S S: M F F and M S A revising and editing, B B data collection, N A and D B result analysis, M R G supervision and editing, all authors reviewed the manuscript.

Please wait… references are loading.