Skip to main content
  • Original Article
  • Open access
  • Published:

Three different methods for ZnO-RGO nanocomposite synthesis and its adsorption capacity for methylene blue dye removal in a comparative study

Abstract

Water is one of the vital needs of life. However, due to rapid industrialization, urbanization and lack of awareness, the world population now facing the threat of water shortage. To ensure that future living conditions are preserved, it is crucial to reduce water pollution and protect the ecosystem. Zinc oxide- reduced graphene oxide (ZnO-RGO) nanocomposite is used in this study as an adsorbent for the adsorption of methylene blue (MB) dye from an aqueous solution. An easy strategy was used for the synthesis of reduced graphene oxide nanoparticles (RGO), Zinc oxide nanoparticles (ZnO) and ZnO-RGO nanocomposite. The synthesis of reduced graphene oxide (RGO) was accomplished through the exothermic reaction of a modified Hummer's method. In a novel approach, zinc oxide nanoparticles (ZnO NPs) were synthesized using the green Leidenfrost technique. This study presents a comparative investigation of ZnO-RGO nanocomposite synthesis employing both green and chemical methods. Three distinct approaches were utilized to prepare the ZnO-RGO nanocomposite: (1) the innovative Leidenfrost green method for composite A1, (2) a chemical precipitation method for composite A2, and (3) a physical mixing sonication method for composite A3. This research marks the first application of the Leidenfrost technique in the synthesis of ZnO-RGO nanocomposites, contributing to the growing body of knowledge in this field. X-ray diffraction (XRD), Burnauer-Emmett-Teller (BET), Fourier transform infrared (FTIR), Zeta potential, transmittance electron microscope (TEM) and scanning electron microscope (SEM) analyses are conducted for synthesized sample characterization. Comparing the XRD patterns of the three synthesis methods, it is notable that the intensity peaks of composite A3 were the highest when ZnO was synthesized using a green method, indicating a higher degree of crystallinity. FTIR analysis approves that combining ZnO with RGO affects the functional groups of the three nanocomposite surfaces. The SEM analysis shows ZnO NPs and RGO sheets are incorporated together. In the case of A1 composite sharp angles make a flower shape was observed due to the unique synthesizing method. The surface area for A2 composite is the highest (7.29 m2/g) compared with A1 (2.91 m2/g) and A3(1.90 m2/g). A comparison study is made among the three nanocomposites for MB dye removal. The effect of adsorbent dose, pH, contact time and initial dye concentration on dye adsorption has been studied. The results show that A1 and A2 nanocomposites removed 85.5 and 87.5% of MB at the optimum adsorbent dose of 0.15 g/100 ml at pH8 and A3 removed 95% of MB at the optimum dose of 0.1 g/100 ml at pH 2. All three composites exhibited adherence to the Langmuir isotherm model, with correlation coefficients (R2) of 0.9858, 0.9904, and 0.9959 for A1, A2, and A3, respectively. Kinetic study results demonstrated that the pseudo-second-order model best described the adsorption process for all three composites, yielding R2 values of 0.9998, 0.9988, and 1.0000 for A1, A2, and A3, respectively. The A3 nanocomposite shows the highest adsorption capacity (104.5 mg/g) compared to the other composites (87.7 and 97.5 mg/g for A1 and A2, respectively). Desorption experiments revealed that the dye removal percentages varied with the ratio of the ethanol–water mixture used. Absolute ethanol achieved a 90% removal compared with 1:1 and 1:2 aqueous ethanol solutions (87.5% and 80%, respectively).

Peer Review reports

Introduction

Industrial pollutants often make water sources unsuitable for irrigation and life demands. The universe is full of cancer-causing substances because of increasing water contamination at a steady rate. Fast action is required to develop a potential and effective water management technology. Nowadays, everyone is concerned about water management due to the rising freshwater demand. Water sources are contaminated because of the discharge of various items such as colored pigments, pesticides, herbicides, diseases and heavy metals [1]. It is known that dye pollution affects human health, including the appearance of skin diseases and digestive problems that may increase the risk of cancer [2, 3]. One of the largest challenges is dye removal from the manufacturing and textile industry effluent [4]. These colored impurities are classified as anionic, cationic, or nonionic [1]. Methylene blue (MB) is one of these cationic-colored and most popular dyes. MB is frequently employed as a coloring and disinfectant in dyestuffs, rubbers, medicines, insecticides and varnishes [5]. In addition, it is used in coloring paper, wool, silk and hair coloring [6, 7]. MB dye is a highly toxic dye that is considered a source of several health problems such as respiratory disease, severe headaches, nausea, vomiting, stomach pain, chest pain, mental confusion, as well as a burning feeling if it is breathed or ingested [8, 9].

To effectively remove pollutants from wastewater, various techniques have been established. OF these, coagulation [10], sedimentation [11], filtration [12], chemical purification [13], reverse osmosis [14], photocatalytic degradation [15, 16], ion exchange [17], biological treatments [18] and other techniques were noted for water remediation [1]. The selection of water treatment techniques depends on flexibility and the lowest economic cost, so adsorption is considered an extremely widely used technique because it is a simple and inexpensive method [19]. Adsorption is a physicochemical surface interaction between the adsorbate and the adsorbent [20]. The adsorption method is affected by several factors, including temperature, the forces that interact between the adsorbent and adsorbate, the pH, contact time and concentration [21].

There are various adsorbents studied for MB dye removals such as crushed brick and cedar sawdust [22], garlic peel [23], bentonite [24], bamboo activated carbon [25], unburned carbon [26] and carbon nanotube [27]. Numerous studies have examined the effectiveness of graphene, graphene oxide (GO), reduced graphene oxide (RGO) and their composites due to their adsorption, oxidation and catalytic properties [28]. These compounds have significant pore volume, high conductivity and rich surface chemistry which make them favorable for the adsorption and catalysis of organic pollutants [29]. They have many applications for the adsorption of pollutants from an aqueous solution like the removal of heavy metal ions [30], oil spill [31], phenols [32], chlorophenols [33], pesticides [34], endocrine-disrupting compounds [35], photocatalysts [36], emerging pollutants [37] and microplastics [38].

To combine the properties of graphene and graphene oxide, RGO is prepared by thermal reduction of GO, where GO can regain graphene-like properties by reduction treatments that transform GO into RGO [39]. RGO is considered an intermediate of graphene and GO [40]. The region of the sufficient conjugated structure of graphene is considered an important point for RGO to adsorb organic dyes, particularly the aromatic structure. RGO possesses oxygen functional groups and surface defects (vacancies places), so RGO is better than GO and graphene in adsorption properties [41]. RGO gives favorable adsorption of MB due to the electrostatic interactions, π– π binding, and hydrophobic association [41,42,43].

Hummer’s method is considered one of the most popular ways for GO production. This technique has become extensively utilized for GO synthesis. It is a very quick and safe method where it was replaced sodium nitrate (NaNO3) and potassium permanganate (KMnO4) instead of nitric acid (HNO3) and potassium perchlorate (KClO4). It doesn’t produce harmful gases such as ClO2 (chlorine dioxide) or acidic fog. Now Hummer’s method is modified where it no longer uses NaNO3, so modified Hummer’s technique becomes more environmentally friendly [44].

ZnO is a favorable material due to its low cost, wide availability, eco-friendly nature, electrochemical activity and relatively broadband gap (3.37 eV) [45]. ZnO NPs have many applications as catalysis [46], photochemical capability [47], medical effects [48], fungicidal [49], antibacterial [50] UV filtration [51] and wastewater treatment [52]. Various techniques were used to synthesize ZnO NPs, in this study Leidenfrost method was used. Leidenfrost is considered a green nanotechnology that focuses on developing simple, novel and environmentally friendly ways to make nanoparticles [53]. The Leidenfrost effect occurs when a liquid comes into contact with a heated plate that has a temperature substantially higher than the liquid's boiling point creating an insulating vapor layer that prevents the liquid from rapidly boiling [53, 54]. The metal oxide forms when a drop of water is scattered on a hot surface at a surface temperature above the water boiling point, where the water molecules droplet ionized into H+ and OH¯. There is a negative charge inside the droplet because of the large number of hydroxyl ions, while there is a positive charge outside the droplet owing to the hydronium ion formation. Metal ion reacts with hydroxide ion forming a metal hydroxide which ends up into metal oxide [55].

"ZnO-RGO nanocomposite possesses remarkable properties; it combines the properties of both RGO and ZnO nanoparticles. It enhanced photocatalytic performance [56], energy storage capacity [57], sensing capability [58] and optoelectronic properties [59]. These attributes make them highly effective as adsorbents for the thorough removal of toxic metals and dyes from aqueous solutions. The increased surface area and diverse functional groups in graphene-based adsorbents contribute to their excellent adsorption potential for toxic metals and dyes. ZnO-RGO composite was synthesized using a new green method (Leidenfrost) that is different from traditional methods and is considered environmentally friendly which achieves the principle of sustainable development.

In this study, ZnO-RGO nanocomposites synthesized via a green method and chemical methods are compared for the first time. We examine the effect of the synthesis method on the resulting nanopowder properties and evaluate the efficacy of each composite in removing methylene blue (MB) dye from aqueous solutions, offering a comprehensive comparative analysis."

Experimental

Materials

In this study, all chemicals (graphite powder, Zinc nitrate-hexahydrate, potassium permanganate, Sulfuric acid (99%), phosphoric acid (85%), hydrochloric acid (99%) hydrogen peroxide (30%), methylene blue dye, ammonium hydroxide and potassium hydroxide) were of analytical grades used without additional purification.

Synthesis of reduced graphene oxide

Using the modified Hummers method, pure graphite powder was used to prepare reduced graphene oxide (RGO). During this process, 27 ml of sulfuric acid and 3 ml of phosphoric acid (9:1 volume ratio) were mixed and stirred for a few minutes. While stirring, 0.225 g of graphite powder was added to the above solution. 1.32 g Potassium permanganate (KMnO4) was slowly added to the mixture. The obtained mixed solution was stirred for 6 h until the solution became dark green. To remove the excess KMnO4, 0.675 ml of hydrogen peroxide (H2O2) was added gradually and stirred for 10 min. This reaction is considered an exothermic reaction. The above mixture was centrifuged with 10 ml of concentrated hydrochloric acid and 30 ml of distilled water for 7 min. The residual was washed several times using distilled water. The result RGO was dried in an oven at 90 °C for 24 h. Functional groups in the GO are eliminated by a quick temperature change and GO is exfoliated to form RGO. Thermal shock is caused by an abrupt temperature change, which removes function groups from the GO lattice as water vapor, CO, and CO2. The primary cause of exfoliation is the pressure created by the evolution of gases between two stacked GO layers [60].

Synthesis of ZnO

ZnO is prepared using the Leidenfrost method according to our previous work [61]. To a hot beaker (300 °C), a 50 ml of 10 mM solution of Zinc nitrate hexahydrate was added from a burette drop by drop to simulate the Leidenfrost effect by forming a vapor layer. Rapid vaporization processes take place in a very short time to create ZnO nanoparticles by heat convection, providing the required activation energy for the nucleation of nanoparticles.

Synthesis of ZnO-RGO nanocomposite

First method (A1 nanocomposite)

For the first time, the ZnO-RGO nanocomposite is prepared using the Leidenfrost method. 50 ml of (0.01 M) Zn(NO3)2.6H2O was mixed with 50 ml (0.1 g) RGO. This mixture was sonicated for 10 min, then the mixture was added drop by drop (Leidenfrost effect) to a hot beaker (300 °C) to obtain the ZnO-RGO nanocomposite.

Second method (A2 nanocomposite)

50 ml of 0.01 M Zn(NO3)2.6H2O was mixed for an hour in an ice water bath using a magnetic bar. RGO solution was prepared by adding 0.0904 g of RGO to 100 ml of distilled water. The RGO solution was sonicated for one hour. Zn (NO3)2 solution was added to the RGO solution and mixed for 10 min, then 5.6 g of potassium hydroxide was added to the above mixture. The resulting solution was left for the night. The solution has become alkaline, so to make the solution pH 7, the precipitate was washed with distilled water during the filtration process. The obtained powder was dried in a 200°C oven for an hour.

Third method (A3 nanocomposite)

0.1g of RGO (50 ml) was sonicated for an hour. 0.4M zinc oxide nanoparticles solution was added slowly with stirring to the RGO solution. After stirring this mixture for 10 min, the suspension was allowed to age for 12 h before filtration. The obtained powder was dried for 3 h at 80°C.

Characterization techniques

The synthesized samples were analyzed using an X-ray Diffraction (XRD) system from Germany, employing Cu Kα radiation at wavelength 0.154 nm. Field Emission Scanning Electron Microscopy (FE-SEM QUANTAFEG 250, Netherlands) was utilized at a voltage of 20 kV. Transmittance Electron Microscope (TEM), JEOL (JEM-HR-2100 ELECTRON MICROSCOPE, USA) was used for further examination. Fourier Transform Infrared analysis (FTIR) was conducted using a JASCO 3600 (Tokyo, Japan). These spectral measurements covered a range from 400 to 4000 cm−1 with a resolution of 2 cm−1, aiming to identify and analyze the functional groups present in the synthesized composites. The particle size distribution and charge characteristics of the synthesized composites were assessed using Zeta sizer Ver. 7.03 (temperature 25 °C, measurement position 2 mm). Additionally, (Burner–Emmett- Teller (BET) surface area analysis (Quanta CHROME NOVA2000 Series, UK) and a UV-1800 TOMOS spectrophotometer from China were employed at room temperature.

Batch experiments

MB powder was dissolved in distilled water to get a stock solution of 500 mg/L. The standard working solutions were prepared daily by dilution of the stock solution. Batch experiments were carried out in 250 mL conical flasks containing 100 mL of standard solution. After adsorption equilibrium, the resultant solution was filtered using Whatman filter paper No. (1001 125), then the obtained solution was measured by the UV–visible spectrophotometer at λ(max) = 664 nm.

The percentage removal (% removal) was calculated using the following equation:

$$\%\, Removal=\frac{{C}_{o}-{C}_{e}}{{C}_{o}}\times 100$$
(1)

where: Co is the initial concentration; Ce is the final concentration in the solution.

The effect of pH on the adsorption of MB was studied in a pH range 2–10 for a 100 mL solution of MB (10, 20, 30, 50 and 100 mg/L). The pH of the solution was adapted to the required pH value using a 1M solution of HCl and NH4OH. The effect of the composite dose on the adsorption of the dye was examined with different dosages (0.02–0.2g). The contact time effect was studied in the range of 15–180 min at the optimum condition of pH, dye concentration and composites dose.

Results and discussion

Characterization

X-ray diffraction analysis (XRD)

Figure 1a, b shows XRD analysis of RGO and ZnO nanoparticles. From the XRD pattern of the RGO, Fig. 1, two distinguished peaks can be observed the first which is the most intense peak at 2θ = 24.25º and the second, less intense at 2θ = 42.79º. The crystallite size of RGO was determined instrumentally to be 2.42 nm. For ZnO NPs, Fig. (1: b), multi-crystal with high crystallinity and pure phase of ZnO NPs was observed [61].

Fig. 1
figure 1

a XRD Pattern of Prepared RGO, b XRD Pattern of Prepared ZnO, c XRD Pattern of A1, A2 and A3

Figure 1c displays the XRD patterns of ZnO-RGO for the three prepared composites (A1, A2 and A3) which showed the combination peaks of ZnO and RGO. This Figure represents seven peaks were shifted to 2θ = 31.7°, 34.3°, 36.2°, 47.4°, 56.5°, 62.8° and 67.9° of ZnO for the three composites. As well, there is a low-intensity peak at 2θ = 28.8°, 28.7° and 24.2° for A1, A2 and A3, respectively which related to RGO [62]. The intensity of RGO peaks is low due to (1) its relatively low diffraction intensity compared to ZnO [63, 64] and (2) the dispersion of ZnO on the outer side of the RGO [65]. Notably, the intensity peaks of composite A3 were the highest where ZnO is prepared by a green method (Leidenfrost method), this indicates a higher degree of crystallinity. The crystallite size was determined to be 16.7, 33.8 and 65.2 nm for A1, A2 and A3, respectively.

Fourier transform infrared spectrometer (FTIR)

The FTIR analysis was studied for the RGO, ZnO and ZnO-RGO composites with a frequency range of 400 to 4000 cm−1. Figures 2a, b show an FT-IR of RGO and ZnO. For RGO, Fig. 2a, a typical broadband attributed to the − OH group was found at 3363 cm−1 which resulted from the adsorbed moisture [66]. The peak at 1715 cm−1 is assigned to the stretching C=O. Moreover, the peak at 1056 cm−1 refers to C–O stretching [67].

Fig. 2
figure 2

a FT-IR Spectrum of RGO, b FT-IR Spectrum of ZnO, c FT-IR Spectrum of A1, A2, A3

For ZnO, Fig. 2b, the peak at 452 cm−1 represented the stretching vibration of the Zn–O, in addition, a peak at 3414 cm−1 is related to the O–H stretching [61, 68,69,70].

Figure 2c shows the FTIR spectrum of the ZnO-RGO nanocomposites. A broad peak at 3232, 3381, 3376 and 1414 cm−1, which attributed to the O–H stretching vibration of the C–OH groups and water [71, 72]. The peaks at 1099.71, 1220.15 and 1119.67 cm−1 correspond to C–O bonds. The nanocomposite spectrum shows a strong transmittance band at 691, 552 and 631 cm−1 which is due to the vibrations of the ZnO bonds [70, 73] this confirms the ZnO-RGO Nanocomposite formation.

Comparing the peaks of RGO and the three nanocomposites, it was found that RGO has four main function groups that appeared at 3363, 1715, 1583 and 1056 cm−1 (OH, C=O of a carboxyl group, C=O of the carbonyl group and C-O, respectively). The same peaks with small shifts formed for A3 nanocomposite (3238, 1720, 1577 and 1119 cm−1) as well as the metal oxide peak at 631 cm−1. Alternatively, for A1 and A2 nanocomposites, only three peaks of RGO are noticed which are related to OH, C–O and C=O of the carbonyl group while the peak at 1715 cm−1 related to the carboxyl group is disappearing. The disappearance of the C=O of the carboxyl group in A1 and A2 nanocomposites means more reduction of RGO takes place [74, 75]. According to previous literature as the reduction process of RGO increases, the composite surface becomes more positive [74]. In contrast, the A3 nanocomposite surface accepted more negative charge compared with the RGO nanoparticle. This is because the A3 nanocomposite shows more broad and intensifying peaks for C=O of both carboxyl and carbonyl groups compared with RGO. In addition, four peaks of the OH group appeared at 3760, 3693, 3238 and 1414 cm−1 compared with RGO which gives two peaks only at 3757 and 3363 cm−1. The variation in the results between the three nanocomposites is related to the variation in their synthesis methods. For A1 and A2 composites, the ZnO NPs grew on the RGO, whereas for the A3 nanocomposite, the ZnO NPs mixed with RGO which accumulated on the surface of the RGO.

Zeta potential analysis

The stability of the synthesized ZnO-RGO nanocomposites was assessed using zeta potential measurements, where a higher positive or negative value indicates greater nanoparticle stability [76]. The negatively charged surface of ZnO-RGO reflects the presence of oxygenated surface groups on RGO. The zeta potential values for A1, A2, and A3 were − 0.013, − 0.03, and − 0.012 mV, respectively, as shown in Fig. 3a–c, likely due to a reduction in oxygen functional groups [77]. Figure 3 further illustrates the consistent zeta potential distributions of A1, A2, and A3."

Fig. 3
figure 3

Zeta potential analysis (a) A1, (b) A2, (c) A3

Scanning electron microscope (SEM)

Figure 4a, b presents SEM images of RGO which appeared as a random aggregation of parallel thin sheets with different edges, wrinkled surfaces and folding [78]. Figure 4c, d represents the morphology of ZnO. They have six folded facets and are similar to the closely packed hexagonal bipyramid morphology [79]. Figure 4e, f shows the SEM images of the A1 composite. It exhibits a cross-linked flaky assembly as well as excellent dispersibility without cluster [80]. The structures randomly overlap one another and form sharp angles making a flower shape this is due to the unique method in which the composite was prepared [80]. Figure 4g–j show the SEM images of A2 and A3 composites, respectively. It shows ZnO nanoparticles and the RGO sheets are incorporated together [81].

Fig. 4
figure 4figure 4

Scanning Electron Microscope (SEM) (a & b) for RGO, (c& d) for ZnO, (e& f) for A1composite, (g& h) for A2 composite, (i& j) forA3 composite

Transmittance electron microscope (TEM)

The structure of synthesized samples was studied using the TEM technique. Figure 5a, b displays the TEM image of RGO which represents typical sheet-like silky waves and clumped structure, this Figure confirms the nano size of RGO in the range of (1.4–3.04 nm) [82]. The selected area electron diffraction (SAED) pattern of RGO displays well-defined diffraction rings and bright spots, indicating the crystalline nature of RGO [83,84,85]. For ZnO nanoparticles, Fig. 5c, d exhibits the formation of spherical-shaped nanoparticles which are agglomerated to form a cluster-like structure in the sample with particle size in the range of (5.8–11.47 nm) [86]. The SAED pattern of ZnO nanoparticles exhibits a characteristic ring diffraction pattern with some brighter, distinct spots, suggesting the presence of larger crystallites randomly oriented, though still in the nanometer range [87]. Figure 6a–c refers to the TEM images of A1, A2 and A3 composites, respectively. The ZnO nanoparticles are distributed randomly onto the surfaces of the RGO sheets. Additionally, the image shows that ZnO and RGO are in close contact. The particle size of A1 was in the order of (2.24–3.97 nm), for A2 (1.96–4.96 nm) and A3 (0.52–1.47 nm).

Fig. 5
figure 5

(a) TEM of RGO and (b) SAED pattern, (c) TEM for ZnO and (d) SAED pattern of ZnO

Fig. 6
figure 6

TEM images a A1 composite, b A2 composite, c A3 composite

The surface area (BET) analysis

The Burnauer-Emmett-Teller (BET) analysis is the most common one for measuring surface area by measuring the adsorption–desorption of nitrogen gas on the surface of the material [88]. The isotherm results of the RGO, ZnO and ZnO-RGO nanocomposites are presented in Fig. 7 and summarized in Table 1. It was observed that all nanocomposites (A1, A2 and A3) exhibited a type IV mesoporous isotherm with an H-1 hysteresis loop, according to IUPAC classification [89, 90]. It is noticeable that the BET-specific surface area of A2 is the highest among RGO, ZnO, A1 and A3 with a surface area of 7.29 m2/g. Conversely, the A3 composite gives the lowest surface area (1.90 m2/g) compared with the other composites. The A1 and A2 composites show the highest surface area which means the growth of ZnO NPs on RGO is effective and improves the properties of RGO and ZnO NPs individually. For the A3 composite, the surface area is lower than the RGO alone because the RGO pore is blocked with ZnO NPs because of the high concentration of ZnO that exists in this composite [91,92,93].

Fig. 7
figure 7

Nitrogen adsorption/desorption isotherm of a for RGO, b for ZnO, c for A1, d for A2, e for A3

Table 1 Surface Areas, Pore Sizes and Pore Volumes for RGO, ZnO, A1, A2 and A3

Batch experiment

Effect of pH

Generally, pH value plays an important role in many adsorption processes. For example, the change in pH affects the way functional groups form on the adsorbent surface, which affects how adsorbent and adsorbate interact [5].

The adsorption results for the removal of MB by the three composites adsorbents (A1, A2 and A3) at various pH values (2 to 10) under the experimental conditions (adsorbent dose 0.02 g /100mL, contact time of 60min and initial concentration 10 mg/L) were presented in Fig. 8. For A1 and A2 composites, the removal percentage increases from pH 2 (25% for and 27.5% for A2) to be maximum at pH 8 with 52.5 and 63% for A1 and A2, respectively. In contrast, the A3 composite shows high percentage removal at pH 2 with 87.5%, which decreases gradually to be minimized at pH 10 with 27.5%.

Fig. 8
figure 8

Effect of pH on MB adsorption (10 mg/L initial concentration, 0.02 g dose, 60 min and 380 rpm) using A1, A2, A3

A1 and A2 composites achieve high removal at weak basic medium (pH 8) which is explained by the fact that the pH affects the adsorbent surface as follows: (a) below pH 4, the solution is intense with H+ ions which adsorbed together with the cationic dye molecules. (b) The second region (pH above 6) is practically neutral and only cationic dye molecules interact with the surface of the composites. (c) The third region (pH 10–11) has excess oxygen, OH− ions, which interact with the cationic dye molecules, keeping them suspended in the solution [94]. Furthermore, the growth of ZnO on RGO leads to a decrease in ZnO PZC (pHPZC = 9.07) to be neutral (pHPZC = 7.2–7.7) for the ZnO-RGO composite [91, 95,96,97,98,99]. This means above the PZC the surface has a negative charge resulting in an electrostatic attraction force between the A1 and A2 composites and the positively charged MB cationic dye at pH 8.

In contrast, the optimum value for A3 at pH 2 (very acidic medium) where the percentage removal was 87.5%. This is attributed to two reasons: First, from BET analysis, the A3 composite surface pore is blocked by ZnO nanoparticles and in a very acidic medium (pH = 3) ZnO dissolved and transferred to the ionic state Zn2+ [91, 100]. This means at pH 2 an activation process takes place for the surface by dissociation of ZnO leading to more empty pores for MB dye adsorption. Increasing pH results in decreasing the adsorption percentage because of the presence of ZnO NPs on the RGO surface which hinders the adsorption of the dye. Secondly, from the FTIR analysis, Fig. 2c, the A3 composite contains more oxygenated function groups with higher intensity and broader than RGO making the composite surface has a negative charge and become more acidic. Increasing these groups (C–O, C=O and OH) leads to a negative value of zeta potential and reduces the PZC of the surface to be more acidic, where these groups are responsible for the acidity properties of the surface [74, 101]. This means the A3 composite PZC is below the PZC of RGO (RGO PZC = 3.5 [95]) so at pH lower than 3.5 the A3 surface has a negative charge facilitating the electrostatic attraction with cationic MB dye.

Comparing the three composites, A3 gives the highest removal percent of the MB dye, while A2 shows the lowest removal.

The results of A1 and A2 composites are in agreement with Mittal, et al., [102, 103], Mijinyawa, et al., [104], Ning, et al.,[105], Sarmah and Karak, [106], Sarkar, et al., [107], Yu, et al., [108] and Zhao, et al., [109] whose found that MB dye removed at pH 8. Otherwise, other studies found that the high adsorption capacity of MB dye takes place in an acidic medium [110,111,112].

Effect of adsorbent dose

The impact of MB dye removal using the three adsorbents at different dosages 0.02, 0.05, 0.1, 0.15 and 0.2 g/100 mL under the experimental conditions was studied. From Fig. 9, the general trend for the three composites is identical. Increasing the dose amount results in increasing the percent removal of MB dye to be maximum at 0.15 g for both A1 and A2 (72.5% and 80%, respectively) and 0.1 g for A3 (90%). Instead, by increasing the dose above 0.15g for A1 and A2 composites and 0.1g for A3 composites, the removal percentage decreases gradually. This is explained by the fact that increasing the adsorbent dose increased the number of active sites on the surface [113]. After equilibrium, the removal percentage decreased because of the saturation of the adsorbent surfaces, where all active sites are fully exposed and used. Also, at high doses, the adsorbent particles collide frequently, resulting in fewer active sites exposed and occupied by MB dye [114].

Fig. 9
figure 9

Effect of dose toward MB adsorption (10 mg/L initial concentration, 60 min and 380 rpm) using A1and A2 at pH = 8, A3 at pH = 2

Effect of shaking time

Determining the contact time is essential for determining the time required for the adsorption process to reach equilibrium. The studied interval time was in the range of 15, 30, 45, 60, 120 and 180 min. The results are represented in Fig. 10. The graph reveals a rapid adsorption equilibrium time takes place for the three composites. Composites A1 and A2 show nearly similarity in results where, the removal increased from 15 to be maximum at 30 min with 85% and 85.9% for A1 and A2, respectively. For the A3 composite, the removal was raised from 15 to 45 min with 92.5% removal. For the three composites after equilibrium, the adsorption percentage goes down with increasing interval time to be the lowest at 180 min. The findings show that adsorption is a fast process and the nanocomposites have a high rate of adsorption due to their large exterior surface area and potential for rapid site occupancies, as well as having single-atom layers that are advantageous for quickly attracting dye molecules [98].

Fig. 10
figure 10

Effect of time toward MB adsorption (10 mg/L initial concentration, 0.15 g and pH = 8 for A1 and A2, 0.1 g and pH = 2 for A3 and 380 rpm)

Effect of concentration

The impact of MB dye solution concentration on the adsorption capacity was evaluated in the range of 10- 100 mg/L. Figure 11 illustrates that there is a reverse relation between concentration and removal percentage; where the increase in concentration from 10 to 100 mg/L results in a decrease in the adsorption of the MB dye for the three composites. The highest removal for MB was obtained at 10 mg/L with removal percentages 85.5%, 87.5% and 95% for A1, A2 and A3, respectively. This may be explained by the fact that at lower dye concentrations the active sites are sufficient to adsorb the dye molecules but at a high dye concentration these active sites become saturated with dye molecules [115]. Finally, A3 gives the highest removal percentage for all studied concentration ranges compared with A1 and A2 composites.

Fig. 11
figure 11

Effect of concentration toward MB adsorption (0.15 g, pH = 8, 30 min for A1 and A2, 0.1 g, pH = 2, 45 min for A3 at 380 rpm)

Figure 12 illustrates the proposed adsorption mechanism of the ZnO-RGO composite for MB. The ZnO-RGO composite facilitates favorable adsorption of MB through hydrogen bonding, π-π interactions, and electrostatic attraction. In an aqueous solution, MB exists as positively charged ions, while ZnO-RGO tends to acquire a negative charge in water, enhancing the electrostatic attraction between the dye and the adsorbent surface.

Fig. 12
figure 12

The mechanism of the adsorption process

Adsorption isotherm

Several mathematical models have been applied to describe equilibrium studies for the adsorption process. Langmuir and Freundlich models are the most applicable models for adsorption systems. Freundlich model can describe adsorption on heterogeneous rough surfaces and the interactions within the adsorbed molecules [41]. The Freundlich isotherm is introduced as an empirical Model (Eq. 2):

$$Log\, {q}_{m}=Log\, {K}_{f}+\frac{1}{n}Log\,{ C}_{e}$$
(2)

where qe the equilibrium capacity of dye on adsorbent (mg/g), Ce is the equilibrium concentration of heavy metals or dye solution (mg/L), Kf is the Freundlich isotherm constant and (n) is the adsorption intensity. Kf and 1/n can be calculated from the intercept and slope of the linear plot of Log qe against Log Ce [116].

According to the Langmuir adsorption isotherm, adsorption only occurs on homogenous surfaces [41]. The linearized form of the Langmuir equation is described by equation (3):

$$\frac{{C}_{e}}{{q}_{m}}=\frac{{C}_{e}}{{q}_{m}}+\frac{1}{{q}_{m}{K}_{L}}$$
(3)

where Ce is the equilibrium concentration of dye solution (mg/L), qm is a monolayer adsorption capacity of the adsorbent (mg/g), KL is the Langmuir constant related to the energy and affinity of binding sites of adsorption (L/mg). The qm and KL can be calculated from the intercept and slope of the linear plot of Ce/qe against Ce.

The linearized form of the Temkin equation is expressed as:

$${q}_{e}=B\, lnK+B\, ln{C}_{e}$$
(4)

where B = RT/b, b is the Temkin constant related to the heat of sorption. Qe (mg/g) and Ce (mg/L) are the amount of adsorbed dye per unit weight of adsorbent and un-adsorbed dye concentration at equilibrium, respectively. Therefore, a plot of qe versus ln Ce is used to determine the constant B which is the constant related to the heat of sorption (J/mol) and K, is the Temkin isotherm equilibrium binding constant [117].

The linearized D-R equation can be written as:

$$Ln\, {q}_{e}=\text{ln}\,{q}_{m}-\beta {E}^{2}$$
(5)

where β, is a constant related to the adsorption energy (mol2/kJ2), qm is a constant that indicates the sorption degree characterizing the sorbent (mg/g) and E is the Polanyi potential, which can be obtained by following equation [118]:

$$E=RT\,\text{ln}(1+\frac{1}{{C}_{e}})$$
(6)

where R, is the ideal gas constant (R = 8.314 J/mol) and T is absolute temperature (K). By plotting ln qe vs E2 it is possible to determine the value of β from the slope and the value of qm from the intercept.

The mean free energy E (mol/kJ) of sorption can be estimated by using B values as expressed in the following equation:

$$E = \frac{1}{{(2\beta )^{1/2} }}$$
(7)

The magnitude of E may characterize the type of adsorption as chemical ion exchange (E = 8–16 kJ/mol) or physical adsorption (E < 8 kJ/mol) [119].

Figure 13a–c represents the Freundlich isotherm of A1, A2 and A3 composites. According to Eq. (2), the kf is determined to be 4.70, 5.27 and 13.21 for A1, A2 and A3, correspondingly. The values of 1/n are summarized in Table 2. The value of 1/n between 0 and 1 indicates the adsorption intensity of a normal isotherm [120]. The R2 were 0.9858, 0.9896 and 0.9682 for A1, A2 and A3, individually.

Fig. 13
figure 13

Freundlich isotherm model a A1, b A2, c A3

Table 2 The Summary of Freundlich, Langmuir Isotherms, Temkin and Dubinin-Radushkevich Isotherm Models

The Langmuir isotherms, (Eq. 3), were represented in Fig. 14a–c and Table 2. The adsorption capacities (qm) were determined to be 87.7, 97.5 and 104.5 for A1, A2 and A3, respectively. The correlation coefficients (R2) were 0.9760, 0.9904 and 0.9959 for A1, A2 and A3, correspondingly. These findings suggest that Freundlich isotherm applies to the three composites but compared with the R2 of Langmuir, the Langmuir model is more convenient for the three composites. Furthermore, from Table 2, it is noticed that the kL < 1 which confirms the adsorption of MB on A1, A2 and A3 favors the Langmuir isotherm [120]. Comparing the qm values of the three composites indicated that the A3 composite has high-capacity adsorption for MB dye compared with A1 and A2 composites.

Fig. 14
figure 14

Langmuir Isotherm Model a A1, b A2, c A3

Figure 15a–c shows the Temkin isotherm plots for the A1, A2, and A3 composites, with the values of B and K provided in Table 2. The B values were 15.21, 16.07, and 20.41, respectively, while the K values were 0.85, 0.85, and 2.35, correspondingly. Figure 16a–c presents the Dubinin-Radushkevich isotherm for A1, A2, and A3, with Table 2 indicating that the mean free energy (E) values are 0.79, 0.79, and 1.58 kJ/mol. These values suggest that MB adsorption on A1, A2, and A3 occurs via physical adsorption. Therefore, it can be concluded that physical adsorption plays a dominant role in the MB adsorption process on these three composites.

Fig. 15
figure 15

Temkin Isotherm Model a A1, b A2, c A3

Fig. 16
figure 16

Dubinin-Radushkevich Isotherm Model a A1, b A2, (c)

Adsorption kinetic studies

The adsorption kinetic study provides facts about the adsorption rate, adsorbent performance and mass transfer mechanisms. There are three steps in the kinetics of adsorption mass transference. The first step is external diffusion, in which the adsorbate moves across the liquid film surrounding the adsorbent. The external diffusion is driven by the concentration dissimilarity between the bulk solution and the adsorbent surface. The second step is internal diffusion. Internal diffusion refers to the movement of adsorbate through the pores of an adsorbent. The third step is the adsorption of the adsorbate in the active sites of the adsorbent. The models of the pseudo-first-order and the pseudo-second-order are the most used [121]. The pseudo-first-order kinetic model is described by Eq. (8):

$$Log\left({q}_{e}-{q}_{t}\right)=Log{q}_{e}-\left(\frac{{K}_{1}}{2.303}\right)t$$
(8)

where, qe and qt are the adsorption capacities at equilibrium and time (mg/g), respectively. K1 is the rate constant of the pseudo-first-order [122]. A plot of log (qe-qt) vs t. Figure 17a–c gives a linear relationship with the slope k1 and intercept log qe according to Eq. (8). The results in Table 3 show that k1 (min−1) for A1, A2 and A3 are 0.012, 0.02 and 0.027, respectively. The correlation coefficients (R2) are 0.4261, 0.3505 and 0.9641 for A1, A2 and A3, respectively.

Fig. 17
figure 17

Pseudo-first order model a A1, b A2, c A3

Table 3 Summary of pseudo-first order, pseudo-second order, intra-particle diffusion and Elovich models

The linearized form of the pseudo-second-order kinetic model is represented as:

$$\frac{t}{{q}_{t}}=\frac{1}{{k}_{2}{q}_{e}^{2}}+\frac{1}{{q}_{e}}t$$
(9)

where, qe and qt are the adsorption capacities at equilibrium and time (mg/g), respectively. A plot of t/qt vs. t. Figure 18a–c shows a linear relationship. The data in Table 3 shows the equilibrium adsorption capacity (qe) for A1, A2 and A3 are 5.65, 5.62 and 9.38, respectively. The correlation coefficients (R2) for A1, A2 and A3 are 0.9998, 0.9988 and 1, respectively. All the correlation coefficients (R2) of pseudo-second-order reaction kinetics are higher than the first-order reaction kinetics, indicating that the removal of MB using A1, A2 and A3 is well described by pseudo-second-order kinetics.

Fig. 18
figure 18

Pseudo-second order model: a A1, b A2, c A3

The intra-particle diffusion model is represented as [123]:

$$q_{t} = K_{i } t^{1/2} + C$$
(10)

where Ki (mg g−1 min−1/2) is the rate constant of the intra-particle diffusion model and C (mg/g) reflects the boundary effect. The intra-particle diffusion plots of qt vs t1/2 are shown in Fig. 19a–c and the values are given in Table 3. The data in Table 3 shows that the C values for A1, A2 and A3 are 0.035, 0.085 and 0.12 mg/g, respectively.

Fig. 19
figure 19

Intra-particle diffusion model: a A1, b A2, c A3

The Elovich equation was:

$${q}_{t }=\frac{1}{\beta }\text{ln}\left(\alpha \beta \right)+\frac{1}{\beta }\text{ln}t$$
(11)

where, α and β are the initial adsorption rate and the adsorption coefficient, respectively. The constants αand β can be calculated when qt is plotted vs ln t [124]. The results are shown in Fig. 20a–c. The data in Table 3 shows the initial adsorption rate (α) for A1, A2 and A3 are 1.6*1024, 1.4*105 and 1.6*1012, respectively. The adsorption constant (β) for A1, A2 and A3 are 11.29, 4.71 and 3.59, respectively.

Fig. 20
figure 20

Elovich model: a A1, b A2, c A3

Table 4 lists the batch experimental optimum condition and qm of this study in comparison to the results of other studies. It is noticeable that the pH values were between 5 and 10, which is nearly neutral, only A3 prefer the acidic medium. Comparing the three adsorbents, the ability of the adsorbent surfaces to remove MB is 87.7, 97.5, and 104 mg/g for A1, A2, and A3, respectively. Compared to A1 and A2, the qm for A3 is very high.

Table 4 The maximum removal percent and capacity of Methylene Blue at different conditions using various adsorbents

Desorption of ZnO-RGO composite

The desorption process using volatile organic solvents was explored as an effective approach for dye removal [138]. The choice of solvent is crucial for efficient dye desorption; it should be affordable, environmentally friendly, and possess high dye solubility [139]. Ethanol was selected to regenerate the ZnO-RGO composite. For the desorption experiment, ZnO-RGO composite was immersed in 50 ml of absolute ethanol, 1:1 aqueous ethanol, and 1:2 aqueous ethanol for 10 min to facilitate MB dye desorption. A 10 ml sample of 10 mg/L MB dye was filtered, and the filtrate was analyzed using a spectrophotometer. The results, shown in Table 5, indicate that the removal percentages of the dye after regeneration with ZnO-RGO composite were 90.0%, 87.5%, and 80.0% for absolute ethanol, 1:1 aqueous ethanol, and 1:2 aqueous ethanol, respectively, with absolute ethanol achieving the highest removal percentage of 90.0%. This efficiency may be attributed to the hydrogen bonding strength of MB with alcohol molecules, which enhances MB solubility [140]. Overall, the results demonstrated that the removal percentage remained high after desorption, confirming the composite’s effectiveness in MB removal."

Table 5 Desorption of ZnO-RGO Composite

Application

The three composites are used for adsorption of MB from a real sample (industrial drainage) under the optimum conditions (pH8, dose 0.15 g/100ml and 30 min for A1 and A2), while for A3 pH2, dose 0.1g/100 and 45 min. A standard addition method was used because the concentrations of all the assayed species were lower than the detection limit. 10 ppm MB was added to the solution, and then the residual concentrations of MB were calculated by the difference. As illustrated in Table 6, A3 nanocomposite achieves the highest removal percentage (97%) compared with A1 and A2 nanocomposites.

Table 6 Removal Percentage of MB Adsorption Using A1, A2 or A3

Conclusion

The three nanocomposites were successfully synthesized using three distinct methods. The first, a novel Leidenfrost green approach, was used to prepare the ZnO-RGO nanocomposite (A1), followed by the precipitation chemical method for the second composite (A2) and the physical mixing sonication method for the third (A3). Composites A1 and A2 underwent a reduction process due to the growth of ZnO nanoparticles on the RGO surface, while the A3 composite exhibited surface oxidation due to the mixing process between ZnO nanoparticles and RGO. A comparison of XRD results revealed that composite A3 showed the highest intensity peaks, indicating a higher degree of crystallinity compared to A1 and A2. The crystallite sizes were determined to be 16.7, 33.8, and 65.2 nm for A1, A2, and A3, respectively. TEM analysis showed ZnO nanoparticles distributed randomly on the RGO surface sheets, with ZnO in close contact with RGO. The surface areas for A1, A2, and A3 were 2.91, 7.29, and 1.90 m2/g, respectively.

For MB dye adsorption, A1 and A2 showed 85.5% and 87.5% effectiveness at pH 8, a dose of 0.15 g/100 mL, and a contact time of 30 min, while A3 showed 95% effectiveness at pH 2, a dose of 0.1 g/100 mL, and a contact time of 45 min. The isotherm study indicated that all three composites followed the Langmuir isotherm, with A1 showing the highest adsorption capacity of 104.5 mg/g, compared to A2 and A3, with 87.7 mg/g and 97.5 mg/g, respectively. The kinetic study showed that A3 had an R2 value greater than 0.93 across four study models, with the best fit to the pseudo-second-order model (R2 nearly 1). In contrast, A1 and A2 had R2 values less than 0.9 for all models except for pseudo-second-order, with values of 0.9998 for A1 and 0.9988 for A2."

Data availability

The authors declare that the data supporting the findings of this study are available within the paper.

Abbreviations

SEM:

Scanning electron microscopy

TEM:

Transmission electron microscopy

XRD:

X-ray diffraction analysis

FTIR:

Fourier transform infrared spectroscopy

BET:

Brunauer–Emmett–Teller

MB:

Methylene blue

GO:

Graphene oxide

RGO:

Reduced graphene oxide

ZnO:

Zinc oxide

A1:

Composite 1

A2:

Composite 2

A3:

Composite 3

PZC:

Point of zero charge

References

  1. Anand KV, et al. A review on the removal of dye, pesticide and pathogens from waste water using quantum dots. Eur J Adv Chem Res. 2020;1:5.

    Article  Google Scholar 

  2. Yagub MT, et al. Dye and its removal from aqueous solution by adsorption: a review. Adv Coll Interface Sci. 2014;209:172–84.

    Article  CAS  Google Scholar 

  3. Vakili M, et al. Application of chitosan and its derivatives as adsorbents for dye removal from water and wastewater: a review. Carbohyd Polym. 2014;113:115–30.

    Article  CAS  Google Scholar 

  4. Lima A, et al. Reduced graphene oxide filtration membranes for dye removal—production and characterization. Mater Proc. 2020;4(1):29.

    Google Scholar 

  5. Zhang F, et al. Adsorption behavior and mechanism of methyl blue on zinc oxide nanoparticles. J Nanopart Res. 2013;15(11):1–10.

    Article  Google Scholar 

  6. Yang Z-Y. Kinetics and mechanism of the adsorption of methylene blue onto ACFs. J China Univ Min Technol. 2008;18(3):437–40.

    Article  CAS  Google Scholar 

  7. Rafatullah M, et al. Adsorption of methylene blue on low-cost adsorbents: a review. J Hazard Mater. 2010;177(1–3):70–80.

    Article  CAS  PubMed  Google Scholar 

  8. Ghosh D, Bhattacharyya KG. Adsorption of methylene blue on kaolinite. Appl Clay Sci. 2002;20(6):295–300.

    Article  CAS  Google Scholar 

  9. Yousefzadeh Y, et al. UiO-66-NH2/guanidine-functionalized chitosan: a new bio-based reusable bifunctional adsorbent for removal of methylene blue from aqueous media. Int J Biol Macromol. 2024;254: 127391.

    Article  CAS  PubMed  Google Scholar 

  10. Cui H, et al. Application progress of enhanced coagulation in water treatment. RSC Adv. 2020;10(34):20231–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Nieto LM, et al. Flocculation–sedimentation combined with chemical oxidation process. Clean Soil Air Water. 2011;39(10):949–55.

    Article  CAS  Google Scholar 

  12. Hube S, et al. Direct membrane filtration for wastewater treatment and resource recovery: a review. Sci Total Environ. 2020;710: 136375.

    Article  CAS  PubMed  Google Scholar 

  13. Gupta VK, et al. Chemical treatment technologies for waste-water recycling—an overview. RSC Adv. 2012;2(16):6380–8.

    Article  CAS  Google Scholar 

  14. Malaeb L, Ayoub GM. Reverse osmosis technology for water treatment: State of the art review. Desalination. 2011;267(1):1–8.

    Article  CAS  Google Scholar 

  15. Akter S, et al. UV/TiO2 photodegradation of metronidazole, ciprofloxacin and sulfamethoxazole in aqueous solution: an optimization and kinetic study. Arab J Chem. 2022;15(7): 103900.

    Article  CAS  Google Scholar 

  16. Kader S, et al. Enhanced photodegradation of methyl orange dye under UV irradiation using MoO3 and Ag doped TiO2 photocatalysts. Environ Technol Innov. 2022;27: 102476.

    Article  CAS  Google Scholar 

  17. Raghu S, Basha CA. Chemical or electrochemical techniques, followed by ion exchange, for recycle of textile dye wastewater. J Hazard Mater. 2007;149(2):324–30.

    Article  CAS  PubMed  Google Scholar 

  18. Bhatia D, et al. Biological methods for textile dye removal from wastewater: a review. Crit Rev Environ Sci Technol. 2017;47(19):1836–76.

    Article  CAS  Google Scholar 

  19. Abu-Nada A, Abdala A, McKay G. Removal of phenols and dyes from aqueous solutions using graphene and graphene composite adsorption: a review. J Environ Chem Eng. 2021;9(5): 105858.

    Article  CAS  Google Scholar 

  20. Bernal V, Giraldo L, Moreno-Piraján JC. Physicochemical properties of activated carbon: their effect on the adsorption of pharmaceutical compounds and adsorbate–adsorbent interactions. C. 2018;4(4):62.

    CAS  Google Scholar 

  21. Gusain R, et al. Adsorptive removal and photocatalytic degradation of organic pollutants using metal oxides and their composites: a comprehensive review. Adv Coll Interface Sci. 2019;272: 102009.

    Article  CAS  Google Scholar 

  22. Hamdaoui O. Batch study of liquid-phase adsorption of methylene blue using cedar sawdust and crushed brick. J Hazard Mater. 2006;135(1–3):264–73.

    Article  CAS  PubMed  Google Scholar 

  23. Hameed BH, Ahmad AA. Batch adsorption of methylene blue from aqueous solution by garlic peel, an agricultural waste biomass. J Hazard Mater. 2009;164(2–3):870–5.

    Article  CAS  PubMed  Google Scholar 

  24. Hong S, et al. Adsorption thermodynamics of methylene blue onto bentonite. J Hazard Mater. 2009;167(1–3):630–3.

    Article  CAS  PubMed  Google Scholar 

  25. Hameed BH, Din ATM, Ahmad AL. Adsorption of methylene blue onto bamboo-based activated carbon: kinetics and equilibrium studies. J Hazard Mater. 2007;141(3):819–25.

    Article  CAS  PubMed  Google Scholar 

  26. Wang S, et al. Unburned carbon as a low-cost adsorbent for treatment of methylene blue-containing wastewater. J Colloid Interface Sci. 2005;292(2):336–43.

    Article  CAS  PubMed  Google Scholar 

  27. Yao Y, et al. Adsorption behavior of methylene blue on carbon nanotubes. Biores Technol. 2010;101(9):3040–6.

    Article  CAS  Google Scholar 

  28. Bai S, et al. The influence of wrinkling in reduced graphene oxide on their adsorption and catalytic properties. Carbon. 2013;60:157–68.

    Article  CAS  Google Scholar 

  29. Thakur K, Kandasubramanian B. Graphene and graphene oxide-based composites for removal of organic pollutants: a review. J Chem Eng Data. 2019;64(3):833–67.

    Article  CAS  Google Scholar 

  30. Ahmad SZN, et al. Adsorptive removal of heavy metal ions using graphene-based nanomaterials: toxicity, roles of functional groups and mechanisms. Chemosphere. 2020;248: 126008.

    Article  CAS  PubMed  Google Scholar 

  31. Elhenawy S, et al. Recent developments and advancements in graphene-based technologies for oil spill cleanup and oil–water separation processes. Nanomaterials. 2021;12(1):87.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Al-Ghouti MA, et al. Effective removal of phenol from wastewater using a hybrid process of graphene oxide adsorption and UV-irradiation. Environ Technol Innov. 2022;27: 102525.

    Article  CAS  Google Scholar 

  33. Wei D, et al. Sorption mechanism and dynamic behavior of graphene oxide as an effective adsorbent for the removal of chlorophenol based environmental-hormones: a DFT and MD simulation study. Chem Eng J. 2019;375: 121964.

    Article  CAS  Google Scholar 

  34. Wang H, et al. Emerging role of graphene oxide as sorbent for pesticides adsorption: experimental observations analyzed by molecular modeling. J Mater Sci Technol. 2021;63:192–202.

    Article  CAS  Google Scholar 

  35. Liao Z, et al. Recent advances in the synthesis, characterization, and application of carbon nanomaterials for the removal of endocrine-disrupting chemicals: a review. Int J Mol Sci. 2022;23(21):13148.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Sodeinde KO, et al. Enhanced adsorptional-photocatalytic degradation of chloramphenicol by reduced graphene oxide-zinc oxide nanocomposite. Sci Rep. 2022;12(1):17054.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Zhu S, et al. Adsorption of emerging contaminant metformin using graphene oxide. Chemosphere. 2017;179:20–8.

    Article  PubMed  Google Scholar 

  38. Dey TK, Jamal M, Uddin ME, Fabrication and performance analysis of graphene oxide-based composite membrane to separate microplastics from synthetic wastewater. 2022.

  39. Ali I, et al. Graphene based adsorbents for remediation of noxious pollutants from wastewater. Environ Int. 2019;127:160–80.

    Article  CAS  PubMed  Google Scholar 

  40. Wang X, et al. Correlation between the adsorption ability and reduction degree of graphene oxide and tuning of adsorption of phenolic compounds. Carbon. 2014;69:101–12.

    Article  CAS  Google Scholar 

  41. Minitha CR, et al. Adsorption behaviour of reduced graphene oxide towards cationic and anionic dyes: co-action of electrostatic and π–π interactions. Mater Chem Phys. 2017;194:243–52.

    Article  CAS  Google Scholar 

  42. Rahman DA, et al. Adsorption of dyes using graphene oxide-based nano-adsorbent: a review. J Presipitasi Media Komunikasi dan Pengembangan Teknik Lingkungan. 2022;19(2):384–97.

    Article  Google Scholar 

  43. Vassileva P, et al. Study on the dye removal from aqueous solutions by graphene-based adsorbents. Materials. 2023;16(17):5754.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Singh A, et al. Graphene and its derivatives: synthesis and application in the electrochemical detection of analytes in sweat. Biosensors. 2022;12(10):910.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Ebnalwaled AA, El-Fadl AA, Tuhamy MA. Synthesis and characterization of reduced graphene oxide/zinc oxide nanocomposites. Synthesis. 2019;48(2):45–69.

    Google Scholar 

  46. Sun Y, et al. The applications of morphology controlled ZnO in catalysis. Catalysts. 2016;6(12):188.

    Article  Google Scholar 

  47. Das PP, et al. Photochemical performance of ZnO nanostructures in dye sensitized solar cells. Solid State Sci. 2015;48:237–43.

    Article  CAS  Google Scholar 

  48. Bisht G, Rayamajhi S. ZnO nanoparticles: a promising anticancer agent. Nanobiomedicine. 2016. https://doiorg.publicaciones.saludcastillayleon.es/10.5772/63437.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Ali M, et al. Antifungal activity of zinc nitrate derived nano Zno fungicide synthesized from Trachyspermum ammi to control fruit rot disease of grapefruit. Ecotoxicol Environ Saf. 2022;233: 113311.

    Article  CAS  PubMed  Google Scholar 

  50. Qi K, et al. Review on the improvement of the photocatalytic and antibacterial activities of ZnO. J Alloy Compd. 2017;727:792–820.

    Article  CAS  Google Scholar 

  51. Kim K-B, et al. Risk assessment of zinc oxide, a cosmetic ingredient used as a UV filter of sunscreens. J Toxicol Environ Health Part B. 2017;20(3):155–82.

    Article  CAS  Google Scholar 

  52. Rambabu K, et al. Green synthesis of zinc oxide nanoparticles using Phoenix dactylifera waste as bioreductant for effective dye degradation and antibacterial performance in wastewater treatment. J Hazard Mater. 2021;402: 123560.

    Article  CAS  PubMed  Google Scholar 

  53. Abdelaziz R, et al. Green chemistry and nanofabrication in a levitated Leidenfrost drop. Nat Commun. 2013;4(1):1–10.

    Article  Google Scholar 

  54. Sobac B, et al. Leidenfrost effect: accurate drop shape modeling and refined scaling laws. Phys Rev E. 2014;90(5): 053011.

    Article  CAS  Google Scholar 

  55. Moghazy MA. Leidenfrost green synthesis method for MoO3 and WO3 nanorods preparation: characterization and methylene blue adsorption ability. BMC Chem. 2023;17(1):1–11.

    Article  Google Scholar 

  56. Haghshenas SSP, et al. Photocatalytic and photoluminescence properties of ZnO/graphene quasi core-shell nanoparticles. Ceram Int. 2019;45(7):8945–61.

    Article  CAS  Google Scholar 

  57. Ramadoss A, Kim SJ. Facile preparation and electrochemical characterization of graphene/ZnO nanocomposite for supercapacitor applications. Mater Chem Phys. 2013;140(1):405–11.

    Article  CAS  Google Scholar 

  58. Singh AK. Synthesis, characterization, electrical and sensing properties of ZnO nanoparticles. Adv Powder Technol. 2010;21(6):609–13.

    Article  CAS  Google Scholar 

  59. Panda D, Tseng T-Y. One-dimensional ZnO nanostructures: fabrication, optoelectronic properties, and device applications. J Mater Sci. 2013;48(20):6849–77.

    Article  CAS  Google Scholar 

  60. Sengupta I, et al. Thermal reduction of graphene oxide: How temperature influences purity. J Mater Res. 2018;33(23):4113–22.

    Article  CAS  Google Scholar 

  61. Hussein SA, Taha GM, Moghazy MA. Comparative study on synthesis ZnO nanoparticles using green and chemical methods and its effect on crystallite size and optical properties. Aswan Univ J Environ Stud. 2023;4(4):205–18.

    Article  Google Scholar 

  62. Li J, Liu X, Sun J. One step solvothermal synthesis of urchin-like ZnO nanorods/graphene hollow spheres and their NO2 gas sensing properties. Ceram Int. 2016;42(1):2085–90.

    Article  CAS  Google Scholar 

  63. Moussa H, et al. ZnO rods/reduced graphene oxide composites prepared via a solvothermal reaction for efficient sunlight-driven photocatalysis. Appl Catal B. 2016;185:11–21.

    Article  CAS  Google Scholar 

  64. Wen MYS, Abdullah AH, Ngee LH. Synthesis of ZnO/rGO Nanohybrid for improved photocatalytic activity. Malay J Anal Sci. 2017;21(4):889–900.

    Google Scholar 

  65. Alharthi FA, et al. Synthesis and characterization of rGO@ ZnO nanocomposites for esterification of acetic acid. ACS Omega. 2022;7(3):2786–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Sharma N, et al. Synthesis and characterization of graphene oxide (GO) and reduced graphene oxide (rGO) for gas sensing application. Hoboken: Wiley Online Library; 2017.

    Book  Google Scholar 

  67. Thakur S, Karak N. Green reduction of graphene oxide by aqueous phytoextracts. Carbon. 2012;50(14):5331–9.

    Article  CAS  Google Scholar 

  68. Ashok A, Kennedy LJ, Vijaya JJ. Structural, optical and magnetic properties of Zn1-xMnxFe2O4 (0≤ x≤ 0.5) spinel nano particles for transesterification of used cooking oil. J Alloys Comp. 2019;780:816–28.

    Article  CAS  Google Scholar 

  69. Haghighatzadeh A, et al. Hollow ZnO microspheres self-assembled from rod-like nanostructures: morphology-dependent linear and Kerr-type nonlinear optical properties. J Mater Sci: Mater Electron. 2021;32(18):23385–98.

    CAS  Google Scholar 

  70. Al-Rawashdeh NAF, Allabadi O, Aljarrah MT. Photocatalytic activity of graphene oxide/zinc oxide nanocomposites with embedded metal nanoparticles for the degradation of organic dyes. ACS Omega. 2020;5(43):28046–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Thema FT, et al. Synthesis and characterization of graphene thin films by chemical reduction of exfoliated and intercalated graphite oxide. J Chem. 2013. https://doiorg.publicaciones.saludcastillayleon.es/10.1155/2013/150536.

    Article  Google Scholar 

  72. Ahmad SZN, et al. Pb(II) removal and its adsorption from aqueous solution using zinc oxide/graphene oxide composite. Chem Eng Commun. 2020;208(5):646–60.

    Article  Google Scholar 

  73. Geetha MS, Nagabhushana H, Shivananjaiah HN. Green mediated synthesis and characterization of ZnO nanoparticles using Euphorbia Jatropa latex as reducing agent. J Sci Adv Mater Dev. 2016;1(3):301–10.

    Google Scholar 

  74. Toral-Sánchez E, et al. Role of the intrinsic properties of partially reduced graphene oxides on the chemical transformation of iopromide. Carbon. 2016;99:456–65.

    Article  Google Scholar 

  75. Bhattacharya G, et al. Aloe vera assisted facile green synthesis of reduced graphene oxide for electrochemical and dye removal applications. RSC Adv. 2017;7(43):26680–8.

    Article  CAS  Google Scholar 

  76. Shah R, et al. Optimisation and stability assessment of solid lipid nanoparticles using particle size and zeta potential. J Phys Sci. 2014;25:1.

    Google Scholar 

  77. Tran DNH, Kabiri S, Losic D. A green approach for the reduction of graphene oxide nanosheets using non-aromatic amino acids. Carbon. 2014;76:193–202.

    Article  CAS  Google Scholar 

  78. Shalaby A, et al. Structural analysis of reduced graphene oxide by transmission electron microscopy. Bul Chem Commun. 2015;47(1):291–5.

    Google Scholar 

  79. Priyadharshini SS, et al. Photocatalytic degradation of methylene blue and metanil yellow dyes using green synthesized zinc oxide (ZnO) nanocrystals. Crystals. 2021;12(1):22.

    Article  Google Scholar 

  80. Lin Y, et al. Green synthesis of ZnO-GO composites for the photocatalytic degradation of methylene blue. J Nanomater. 2020. https://doiorg.publicaciones.saludcastillayleon.es/10.1155/2020/4147357.

    Article  Google Scholar 

  81. Manjunatha C, et al. One pot green synthesis of novel rGO@ ZnO nanocomposite and fabrication of electrochemical sensor for ascorbic acid using screen-printed electrode. J Nanostruct. 2020;10(3):531–9.

    CAS  Google Scholar 

  82. Sengunthar P, et al. Physical properties and enhanced photocatalytic activity of ZnO-rGO nanocomposites. Appl Phys A. 2020;126(7):1–9.

    Article  Google Scholar 

  83. George G, et al. Green and facile approach to prepare polypropylene/in situ reduced graphene oxide nanocomposites with excellent electromagnetic interference shielding properties. RSC Adv. 2018;8(53):30412–28.

    Article  CAS  Google Scholar 

  84. Jose PPA, et al. Reduced graphene oxide/silver nanohybrid as a multifunctional material for antibacterial, anticancer, and SERS applications. Appl Phys A. 2020;126(1):58.

    Article  CAS  Google Scholar 

  85. Kumar P, Jain P, Sarkar A. Binary solvent-assisted graphene synthesis by direct exfoliation of graphite powder and its cytotoxicity on SiHa cell line. Ind J Pharm Educ Res. 2022;56(3):732–9.

    Article  CAS  Google Scholar 

  86. Lingaraju K, et al. Biogenic synthesis of zinc oxide nanoparticles using Ruta graveolens (L.) and their antibacterial and antioxidant activities. Appl Nanosci. 2016;6(5):703–10.

    Article  CAS  Google Scholar 

  87. Emil E, et al. Tuning the morphology of ZnO nanostructures with the ultrasonic spray pyrolysis process. Metals. 2018;8(8):569.

    Article  Google Scholar 

  88. Walton KS, Snurr RQ. Applicability of the BET method for determining surface areas of microporous metal—organic frameworks. J Am Chem Soc. 2007;129(27):8552–6.

    Article  CAS  PubMed  Google Scholar 

  89. Morsy M, et al. Highly efficient photocatalysts for methylene blue degradation based on a platform of deposited GO-ZnO nanoparticles on polyurethane foam. Molecules. 2023;28(1):108.

    Article  CAS  Google Scholar 

  90. Sing KSW. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations 1984). Pure Appl Chem. 1985;57(4):603–19.

    Article  CAS  Google Scholar 

  91. Dehghan S, et al. Preparation and photocatalytic performance of reduced graphene oxide/ ZnO nanocatalyst for degradation of metalaxyl from aqueous solution: effect of operational parameters, mineralisation and toxicity bioassay. Int J Environ Anal Chem. 2022;102(18):7112–34.

    Article  CAS  Google Scholar 

  92. Peng W, et al. A review on heavy metal ions adsorption from water by graphene oxide and its composites. J Mol Liq. 2017;230:496–504.

    Article  CAS  Google Scholar 

  93. Babitha KB, et al. Catalytically engineered reduced graphene oxide/ZnO hybrid nanocomposites for the adsorption, photoactivity and selective oil pick-up from aqueous media. RSC Adv. 2015;5(62):50223–33.

    Article  CAS  Google Scholar 

  94. Arias Arias F, et al. The adsorption of methylene blue on eco-friendly reduced graphene oxide. Nanomaterials. 2020;10(4):681.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Landström A, et al. Reduced graphene oxide-ZnO hybrid composites as photocatalysts: the role of nature of the molecular target in catalytic performance. Ceram Int. 2021;47(14):19346–55.

    Article  Google Scholar 

  96. Tatykayev B, et al. Synthesis of core/shell ZnO/rGO nanoparticles by calcination of ZIF-8/rGO composites and their photocatalytic activity. ACS Omega. 2017;2(8):4946–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Sabouni R, Gomaa H. Photocatalytic degradation of pharmaceutical micro-pollutants using ZnO. Environ Sci Pollut Res Int. 2019;26(6):5372–80.

    Article  CAS  PubMed  Google Scholar 

  98. Li Y, et al. Comparative study of methylene blue dye adsorption onto activated carbon, graphene oxide, and carbon nanotubes. Chem Eng Res Des. 2013;91(2):361–8.

    Article  CAS  Google Scholar 

  99. Rout DR, Jena HM. Polyethylene glycol functionalized reduced graphene oxide coupled with zinc oxide composite adsorbent for removal of phenolic wastewater. Environ Res. 2022;214(Pt 3): 114044.

    Article  CAS  PubMed  Google Scholar 

  100. Le AT, et al. Synergistic effect of pH solution and photocorrosion of ZnO particles on the photocatalytic degradation of Rhodamine B. Bull Mater Sci. 2021. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s12034-020-02281-6.

    Article  Google Scholar 

  101. Krishnamoorthy K, et al. The chemical and structural analysis of graphene oxide with different degrees of oxidation. Carbon. 2013;53:38–49.

    Article  CAS  Google Scholar 

  102. Mittal H, Al Alili A, Alhassan SM. High efficiency removal of methylene blue dye using κ-carrageenan-poly(acrylamide-co-methacrylic acid)/AQSOA-Z05 zeolite hydrogel composites. Cellulose. 2020;27(14):8269–85.

    Article  CAS  Google Scholar 

  103. Mittal H, et al. GO crosslinked hydrogel nanocomposites of chitosan/carboxymethyl cellulose—a versatile adsorbent for the treatment of dyes contaminated wastewater. Int J Biol Macromol. 2021;167:1248–61.

    Article  CAS  PubMed  Google Scholar 

  104. Mijinyawa AH, Durga G, Mishra A. A sustainable process for adsorptive removal of methylene blue onto a food grade mucilage: kinetics, thermodynamics, and equilibrium evaluation. Int J Phytorem. 2019;21(11):1122–9.

    Article  CAS  Google Scholar 

  105. Ning F, et al. Hydroxyethyl cellulose hydrogel modified with tannic acid as methylene blue adsorbent. J Appl Polym Sci. 2020;138:8.

    Google Scholar 

  106. Sarmah D, Karak N. Double network hydrophobic starch based amphoteric hydrogel as an effective adsorbent for both cationic and anionic dyes. Carbohydr Polym. 2020;242: 116320.

    Article  CAS  PubMed  Google Scholar 

  107. Sarkar N, Sahoo G, Swain SK. Reduced graphene oxide decorated superporous polyacrylamide based interpenetrating network hydrogel as dye adsorbent. Mater Chem Phys. 2020;250:20.

    Article  Google Scholar 

  108. Yu H, et al. Preparation of porous microspherical adsorbent via pine pollen stabilized O1/W/O2 double emulsion for high-efficient removal of cationic dyes. Colloid Surf A Physicochem Eng Aspects. 2020;601:104.

    Article  Google Scholar 

  109. Zhao J, et al. Preparation of chitosan and carboxymethylcellulose-based polyelectrolyte complex hydrogel via SD-A-SGT method and its adsorption of anionic and cationic dye. J Appl Polym Sci. 2020;137:34.

    Article  Google Scholar 

  110. Mokhtar A, et al. Preparation and characterization of anionic composite hydrogel for dyes adsorption and filtration: non-linear isotherm and kinetics modeling. J Polym Environ. 2020;28(6):1710–23.

    Article  CAS  Google Scholar 

  111. Kamel S, et al. Carboxymethyl cellulose-hydrogel embedded with modified magnetite nanoparticles and porous carbon: effective environmental adsorbent. Carbohydr Polym. 2020;242: 116402.

    Article  CAS  PubMed  Google Scholar 

  112. Li B, et al. Adsorption of methylene blue and Cd(II) onto maleylated modified hydrochar from water. Environ Pollut. 2019;254(Pt B): 113014.

    Article  CAS  PubMed  Google Scholar 

  113. El Qada EN, Allen SJ, Walker GM. Adsorption of basic dyes onto activated carbon using microcolumns. Ind Eng Chem Res. 2006;45(17):6044–9.

    Article  Google Scholar 

  114. Liu T, et al. Adsorption of methylene blue from aqueous solution by graphene. Colloids Surf, B. 2012;90:197–203.

    Article  CAS  Google Scholar 

  115. Nizam NUM, et al. The removal of anionic and cationic dyes from an aqueous solution using biomass-based activated carbon. Sci Rep. 2021;11(1):1–17.

    Article  Google Scholar 

  116. Desta MB. Batch sorption experiments: Langmuir and Freundlich isotherm studies for the adsorption of textile metal ions onto teff straw (Eragrostis tef) agricultural waste. J Thermodyn. 2013;2013:1.

    Article  Google Scholar 

  117. Johnson RD, Arnold FH. The Temkin isotherm describes heterogeneous protein adsorption. Biochim Biophys Acta BBA-Protein Struct Mol Enzymol. 1995;124:293–7.

    Google Scholar 

  118. Hutson ND, Yang RT. Theoretical basis for the Dubinin-Radushkevitch (DR) adsorption isotherm equation. Adsorption. 1997;3:189–95.

    Article  CAS  Google Scholar 

  119. Inglezakis VJ, Zorpas AA. Heat of adsorption, adsorption energy and activation energy in adsorption and ion exchange systems. Desalin Water Treat. 2012;39(1–3):149–57.

    Article  CAS  Google Scholar 

  120. Khurana I, et al. Removal of dyes using graphene-based composites: a review. Water Air Soil Pollut. 2017;228(5):1–17.

    Article  CAS  Google Scholar 

  121. Wang J, Guo X. Adsorption kinetic models: physical meanings, applications, and solving methods. J Hazard Mater. 2020;390: 122156.

    Article  CAS  PubMed  Google Scholar 

  122. Moussout H, et al. Critical of linear and nonlinear equations of pseudo-first order and pseudo-second order kinetic models. Karbala Int J Modern Sci. 2018;4(2):244–54.

    Article  Google Scholar 

  123. Graaf GH, et al. Intra-particle diffusion limitations in low-pressure methanol synthesis. Chem Eng Sci. 1990;45(4):773–83.

    Article  CAS  Google Scholar 

  124. Aharoni C, Tompkins FC. Kinetics of adsorption and desorption and the Elovich equation. In: Aharoni C, editor. Advances in catalysis. Amsterdam: Elsevier; 1970. p. 1–49.

    Google Scholar 

  125. Salem AI, et al. A comparative study for adsorption of methylene blue dye from wastewater on to three different types of rice ash. J Pharm Appl Chem. 2018;4(1):30–9.

    Google Scholar 

  126. Al-Mahmoud SM. Kinetic and thermodynamic studies for the efficient removal of methylene blue using hordeum murinum as a new biosorbent. Egypt J Chem. 2020;63(9):3381–90.

    Google Scholar 

  127. Al-Asadi ST, et al. Adsorption of methylene blue dye from aqueous solution using low-cost adsorbent: kinetic, isotherm adsorption, and thermodynamic studies. Environ Monit Assess. 2023;195(6):676.

    Article  CAS  PubMed  Google Scholar 

  128. Mehdizadeh A, et al. Preparation of a new magnetic nanocomposite for the removal of dye pollutions from aqueous solutions: synthesis and characterization. Mater Chem Horizon. 2022;1(1):23–34.

    Google Scholar 

  129. Myneni VR, Kanidarapu NR, Vangalapati M. Methylene blue adsorption by magnesium oxide nanoparticles immobilized with chitosan (CS-MgONP): response surface methodology, isotherm, kinetics and thermodynamic studies. Ir J Chem Chem Eng IJCCE. 2020;39(6):29–42.

    CAS  Google Scholar 

  130. Noori M, Tahmasebpoor M, Foroutan R. Enhanced adsorption capacity of low-cost magnetic clinoptilolite powders/beads for the effective removal of methylene blue: adsorption and desorption studies. Mater Chem Phys. 2022;278: 125655.

    Article  CAS  Google Scholar 

  131. Lagergren SK. About the theory of so-called adsorption of soluble substances. Sven Vetenskapsakad Handingarl. 1898;24:1–39.

    Google Scholar 

  132. Ahmad N, et al., Adsorption of methylene blue in aqueous solution by Musa paradisiaca stem powder. ARPN J Eng Appl Sci. 2016; 11.

  133. Tran HD, Nguyen DQ. Study on methylene blue adsorption using cashew nut shell-based activated carbon. Chimica Techno Acta. 2023;10(4):202310401.

    Article  CAS  Google Scholar 

  134. Mouni L, et al. Removal of methylene blue from aqueous solutions by adsorption on kaolin: kinetic and equilibrium studies. Appl Clay Sci. 2018;153:38–45.

    Article  CAS  Google Scholar 

  135. Mekuria D, et al. Adsorptive removal of methylene blue dye using biowaste materials: barley bran and enset midrib leaf. J Chem. 2022;2022:1–13.

    Article  Google Scholar 

  136. Lelifajri L, et al. Study on methylene blue dye adsorption in aqueous solution by heat-treated Gnetum gnemon shell waste particles as low-cost adsorbent. Melville: AIP Publishing; 2020.

    Book  Google Scholar 

  137. Kerrou M, et al. A comparative study of the kinetics and isotherm of adsorption of a cationic dye by different natural wastes. EDP Sci. 2021;234:59.

    CAS  Google Scholar 

  138. Kanwal A, et al. Phytoremediative adsorption methodologies to decontaminate water from dyes and organic pollutants. RSC Adv. 2023;13(38):26455–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Salimi A, Roosta A. Experimental solubility and thermodynamic aspects of methylene blue in different solvents. Thermochim Acta. 2019;675:134–9.

    Article  CAS  Google Scholar 

  140. Samiey B, Ashoori F. Adsorptive removal of methylene blue by agar: effects of NaCl and ethanol. Chem Cent J. 2012;6(1):1–13.

    Article  Google Scholar 

Download references

Acknowledgements

I want to express my gratitude to Professor Fatma S. M. Hassan, a professor of inorganic chemistry, for her support and valuable manuscript revision.

Funding

Open access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB).

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed and helped in the completion of this study. [Safaa A. Hussein] performed the experiments, analyzed the data and wrote the first draft of the manuscript. [Gharib M. Taha] suggested the study idea, revised the results and supervised the project. [F. A. Adam] supervised the project. [Marwa A. Moghazy] assisted in the suggestion of the idea of the study, supervised the practical part and data analysis and revised the results and final writing of this manuscript.

Corresponding author

Correspondence to Marwa A. Moghazy.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hussein, S.A., Taha, G.M., Adam, F.A. et al. Three different methods for ZnO-RGO nanocomposite synthesis and its adsorption capacity for methylene blue dye removal in a comparative study. BMC Chemistry 19, 18 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13065-025-01381-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13065-025-01381-w

Keywords