Next Article in Journal
Phase-Field Simulation of the Microstructure Evolution in the Eutectic Alloy NiAl-31Cr-3Mo
Next Article in Special Issue
Recent Advances in the Raman Investigation of Structural and Optical Properties of Graphene and Other Two-Dimensional Materials
Previous Article in Journal
Effects of Annealing on Surface Residual Impurities and Intrinsic Defects of β-Ga2O3
Previous Article in Special Issue
Modelling the Structure and Optical Properties of Reduced Graphene Oxide Produced by Laser Ablation: Insights from XPS and Time-Dependent DFT
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unleashing the Power of Graphene-Based Nanomaterials for Chromium(VI) Ion Elimination from Water

by
Keloth Paduvilan Jibin
1,*,†,
Silpa Augustine
2,†,
Prajitha Velayudhan
1,
Jesiya Susan George
1,
Sisanth Krishnageham Sidharthan
3,
Sylas Variyattel Paulose
4,* and
Sabu Thomas
1,3,*
1
School of Chemical Sciences, Mahatma Gandhi University, Kottayam 686560, India
2
Department of Chemistry, Baselius College, Kottayam 686560, India
3
International and Inter-University Centre for Nano Science and Nanotechnology, Mahatma Gandhi University, Kottayam 686560, India
4
School of Environmental Sciences, Mahatma Gandhi University, Kottayam 686560, India
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this manuscript.
Crystals 2023, 13(7), 1047; https://doi.org/10.3390/cryst13071047
Submission received: 1 June 2023 / Revised: 27 June 2023 / Accepted: 28 June 2023 / Published: 1 July 2023
(This article belongs to the Special Issue Recent Advances in Graphene and Other Two-Dimensional Materials)

Abstract

:
Graphene-oxide-based nanomaterials have gained attention in recent years in the field of heavy metal removal. In this work, graphene oxide (GO) and graphene-oxide-coated silica nanoparticles (GO/SiO2) were synthesized for the efficient removal of Cr(VI) ions from water. Synthesized nanosorbents were characterized by FTIR, Raman spectroscopy, and Transmission Electron Microscopy (TEM). The effects of the pH and the concentration of Cr(VI) ions in adsorption, using GO and GO/SiO2, was studied using the batch process. The results of the study indicated that the maximum removal percentage was shown at pH 3 for both adsorbents. Comparatively, GO/SiO2 showed a higher removal percentage (92.28%) than GO (86.15%) for Cr(VI) at a concentration of 50 ppm. The results validate that the removal of Cr(VI) ions is highly concentration-dependent and pH-dependent. This study shows that GO and GO/SiO2 are efficient adsorbents and that GO/SiO2 has great potential over GO for the removal of Cr(VI) ions from water.

1. Introduction

The availability of clean drinking water is critical for maintaining a healthy life. Freshwater resources have been steadily declining in recent decades due to increased human consumption, pollution, and climate change. Among the toxic pollutants found in water bodies, heavy metals deserve special attention. Heavy metals are generally referred to as those metals which possess a specific density of more than 5 g/cm3 and adversely affect the environment and living organisms [1]. They are very toxic even at low doses, very persistent in the environment and living tissues, and easily migrate upstream of the food chain. The toxic effects of these metals are harmful to the human body and its proper functioning. Moreover, monitoring and removing them are costly procedures. The most commonly found heavy metals in wastewater include arsenic, cadmium, chromium, copper, lead, nickel, mercury, and zinc, all of which badly affect human health and the environment [2]. Heavy metals enter into the surrounding environment by natural means and as a result of human activities. Various sources of heavy metals include soil erosion, natural weathering of the earth’s crust, mining, industrial discharge, urban runoff, sewage discharge, pest- or disease-control agents applied to crops, and many others [3].
Chromium (Cr) is the seventh most abundant element on earth [4]. It occurs in several oxidation states in the environment, ranging from Cr2+ to Cr6+.The most commonly occurring forms of Cr are trivalent- Cr3+ and hexavalent- Cr6+, and both states are toxic to animals, humans, and plants [5]. Hexavalent chromium is thought to have 100-fold extra toxicity than trivalent chromium for each acute and persistent exposure, due to its excessive water solubility and mobility and easy reduction. The respiratory tract is the important goal organ for hexavalent chromium after inhalation exposure in humans. Chronic human exposure to high concentrations of hexavalent chromium through inhalation or oral exposure may produce effects on the liver, kidney, gastrointestinal, and immune systems, and possibly the blood. Dermal exposure to hexavalent chromium can result in contact dermatitis, sensitivity, and ulceration of the skin [6].
Heavy metal pollution has become a serious problem in both developed and developing countries [7,8,9,10]. In addition to the detection of heavy metal ions, their elimination from water is quite a crucial concern. Many strategies can be employed for the removal process, and common methods for removing heavy metals from water sources include precipitation, flocculation, membrane separation, ion exchange, and evaporation [11,12,13]. Among the many methods, adsorption is often the most effective technique of choice, due to its low cost, simple design, strong operability, and especially its high elimination performance from dilute solutions [14,15]. Some widely used adsorbents for the adsorption of metal ions include activated carbon, clay minerals, biomaterials, industrial solid wastes, and zeolites [16]. The development of nanotechnology has provided various nanomaterials as efficient adsorbents for removing metal ions from water [17,18,19,20]. Heavy metal removal using graphene and graphene-based nanomaterials has grown significantly during these years. Development of various functionalized nanomaterials, such as graphene oxide (GO), enhances the removal efficiency toward metal ion uptake. The presence of oxygen-containing functional groups (hydroxyl, carboxyl, epoxy, etc.) for metal ion complexation, large specific surface area, fast adsorption speed, affinity for a large number of metal ions, and low preparation cost make GO a promising adsorbent for heavy metal removal [21,22]. Rajesh et al. reported a novel and straightforward approach for effectively removing toxic chromium by using the intriguing interaction between exfoliated graphene oxide (EGO), trioctylamine (TOA), and Cr(VI) [23]. Recently, numerous research works have studied the development of graphene-oxide-based nanomaterials and their environmental applications [24]. For instance, Deng et al. synthesized magnetic graphene oxide and used it as an adsorbent for the removal of Cd2+ from wastewater [25]. Likewise, a synthesized magnetic graphene oxide (mGO) grafted with amine groups (mGO-PAMAM) was found to be effective in the removal of Cd2+, Pb2+, and Cu2+ from wastewater and to have excellent regeneration capability for future use [26]. Prepared magnetic chitosan and graphene nanocomposites functionalized with EDTA (EDTA-MCS/GO) were used for the removal of Cr3+ from water [27]. In this study, we synthesized graphene oxide (GO),amine-modified silica using APTES and graphene-oxide-coated silica nanoparticles (GO/SiO2), for the removal of Cr(VI) ions from water. Generally, the interaction between metal ions and the surface oxygen-containing groups plays an important role in the adsorption capacity of heavy metals on CNTs and graphene-based nanomaterials [28]. Since silica nanoparticles and their derivatives have been found to possess excellent qualities, such as a definite pore size, high surface area, strong selectivity, and adsorption capability, a lot of studies have been conducted employing them as adsorbents [29]. Further modification of nano silica and synthesized GO is possible and thereby overcomes the self-aggregation and also increases the removal efficiency. When APTES-modified silica is coated with GO, the adsorption rate is found to be increased due to its enhanced properties.
In the present work, we successfully prepared an amine-modified silica-GO self-assembled novel high-efficiency adsorbent for the successful removal of Cr(VI) from the solution and compared the result with GO. The combination of amine-modified silica and GO creates a synergistic effect that improves Cr(VI) adsorption. Graphene oxide possesses a large surface area and abundant oxygen-containing functional groups, such as hydroxyl and carboxyl groups, which can form coordination complexes with metal ions. The amine-modified silica, with its amino groups, can interact with the functional groups on GO, creating a favorable environment for Cr(VI) adsorption. The synergistic effect between the amine-modified silica and GO enhances the overall adsorption capacity for Cr(VI). The properties of the prepared adsorbent were studied using various spectroscopic and microscopic methods. We achieved a high-percentage removal of Cr(VI) in the silica-GO hybrid system.

2. Materials and Methods

2.1. Materials

Potassium dichromate (K2Cr2O7), graphite (20-micron, synthetic CAT#282863), potassium permanganate (KMnO), Conc.H2SO4, sodium nitrate (NaNO3), hydrochloric acid (HCl), sodium carbonate (Na2CO3), Nano silica (10 nm diameter, spec. surface area 175–225 m2/g (BET), 99.8% trace metals basis. CAT#718483), 3-Aminopropyltriethoxysilane (APTES), HCl, ammonia, ethanol, and 30% H2O2 were purchased from Sigma Aldrich (St. Louis, MI, USA). All the chemicals were used as received, unless otherwise specified.

2.2. Methods

2.2.1. Synthesis of Graphene Oxide

Graphene oxide was synthesized by modified Hummer’s method [30]. Then, 5 g of graphite, 115 mL of sulfuric acid, and 5 g of NaNO3 were taken in a beaker and stirred for 30 min in an ice bath. After that, 15 g of KMnO4 was added very slowly and stirred for 40 min. We then slowly raised the temperature to 35 °C for 12 h, and to this, 230 mL of distilled water was added slowly. The temperature was further raised to 90 °C for 1 h, and to this, 400 mL of distilled water and 50 mL of 30% H2O2 were added. Then, we removed the supernatant and washed it with 5% HCl several times to remove sulphate salt. The solution was then made alkaline by treating it with 5% Na2CO3. Afterward, it was centrifuged and dried it in an oven.

2.2.2. Synthesis of Amine Modified Silica

The APTES and silica nanoparticles were used to create amine modified silica nanoparticles. This was accomplished by dissolving 1 mL of APTES in 200 mL of ethanol in a beaker. This solution was then mixed with 10 g of SiO2. The resulting mixture was sonicated and stirred vigorously for one hour. The solution was then given 150 mL of water, followed by 30 min of stirring and 15 min of sonication. The mixture was centrifuged for 20 min at 14,000 rpm, the solid residue recovered and repeatedly rinsed with ethanol. The amine modified silica nanoparticles collected by gently heating the gel-like substance for three hours at 50 °C and 0.2 mbar of reduced pressure.

2.2.3. Synthesis of Self-Assembled Graphene-Oxide-Coated Silica Nanoparticle

Synthesized graphene oxide was dispersed in distilled water. Modification of amine on nano silica was accomplished by using 3-aminopropyl triethoxysilane in ethanol/water mixture (9:1). Nano silica purchase from sigma Aldrich is modified by using 3-Aminopropyl triethoxysilane (APTES) in ethanol medium (Figure 1) and the amine modified silica was also dispersed in distilled water and prepared the self-assembled structures of graphene oxide coated silica nanoparticles by mixing along with stirring and the core shell precipitates out and washed and dried. Then,10 g of amine-modified silica was dispersed in 500 mL distilled water via sonication and stirring. In another beaker, 0.1 g of graphene oxide was dispersed in 500 mL of distilled water. The two 500 mL solutions were mixed thoroughly and stirred for another 6 h. Centrifuge the precipitate and dried in vacuum oven at 80 °C.

2.2.4. Preparation of Stock Solution

Stock solution of chromium was prepared by dissolving 100 mg of potassium dichromate (K2Cr2O7) in a 100 mL standard flask. From this stock solution of 1000 ppm, different concentrations of standard solutions (1, 10, 30, 50, 100, and 200) were prepared.

2.2.5. Optimization of pH

A series of standard solutions, 1 ppm, 10 ppm, 30 ppm, 50 ppm, 100 ppm, and 200 ppm, were prepared in a 100 mL standard flask. The pH was adjusted to 3, 5, 7, and 9 for a 30 ppm solution, and 20 mg of GO was added to each, sonicated for 20 min, and allowed to settle down for 24 h. The solutions were decanted using Whatman No. 41 filter paper, and the samples to be analyzed were filtered using a syringe filter. Then ICP-MS analysis was performed, and the results were used for determining the pH at which the maximum adsorption takes place. The same procedure was carried out using silica-coated GO for a 30 ppm standard solution, and the optimum pH was determined.

2.2.6. Batch Adsorption Test on GO and GO-Coated Silica Nanoparticles

To investigate the adsorption of heavy metal ions (Cr (VI)) on GO and GO-coated silica nanocomposites, a batch adsorption test was performed. First, 20 mg of GO was added to 20 mL of different concentrations of standard solutions (1 ppm, 10 ppm, 30 ppm, 50 ppm, 100 ppm, and 200 ppm). Then, 20 mg of GO-coated silica nanoparticles was added to the next set of 20 mL standard solutions. The pH was optimized, then the sample was sonicated for 20 min, allowed to cool at room temperature, and it was left at room temperature for 24 h for equilibration. It was then decanted using Whatman No.1 filter paper, and the sample to be analyzed was filtered using a syringe filter. Then ICP-MS analysis was carried out for the filtrate that was obtained after sonication, filtration, and syringe filtration. The quantity of adsorbed heavy metal ions was calculated as the percentage of the difference between concentrations of metal ions before and after adsorption as follows [31,32]:
P e r c e n t a g e   A d s o r p t i o n   ( R e m o v a l ) = C i C e C e × 100
where Ci is the initial heavy metal concentration (ppm), and Ce is the heavy metal concentration at final equilibrium (ppm).

2.3. Characterizations

Characterization of graphite and graphene oxide was performed using FTIR spectrometer (FTIR spectrometer Perkin Elmer Spectrum 2). The Raman spectra were obtained using WITEC Alpha300 RA—Confocal Raman Microscope with AFM. Morphological studies of GO and GO-coated silica nanoparticle were performed with a transmission electron microscope (TEM—JEOL 2100, high resolution, Tokyo, Japan). Amounts of heavy metal removal from standard solutions were determined by using ICP-MS (iCAP Q-model-Thermo).

3. Results and Discussion

3.1. Characterization of GO and GO/SiO2

FTIR measurement was employed to investigate the bonding interactions in graphite before and after the oxidation process. Figure 2 shows the FTIR spectra of graphite, graphene oxide, and GO/SiO2.
Evidence of graphite oxidation could be obtained from the analysis of spectral data. The different bands present in the spectra indicate different functional groups present on GO. Figure 2 shows that the synthesized GO has a peak between 1000 and 1100 cm−1 that is attributed to the epoxy (-O-) group [33]. The peak between 1700 cm−1 and 1720 cm−1 is associated with the stretching of C=O bonds of the carboxyl group (-COOH) [34] or carbonyl group [26,35]. The peak located between 3500 and 3400 cm−1 is ascribed to the -OH stretching mode hydroxyl group [26]. The FTIR spectrum shows oxygen-containing functional groups such as carbonyl (C=O), carboxyl—(COOH), epoxy (C-O-C), and hydroxyl (O-H) groups present on GO. The FTIR spectra of graphite do not show such properties, thereby confirming the successful incorporation of oxygen-containing groups into the graphite. From the FTIR spectra of GO-SiO2, we can confirm the successful incorporation of SiO2 into the GO sheets. The characteristic peaks at 1199 cm−1, 1060 cm−1, 970 cm−1, and 800 cm−1 are attributed to the Si-O-C asymmetric stretching vibration Si-O-Si asymmetric stretching vibration, Si-OH stretching vibration, and Si-O-Si symmetrical stretching vibration, respectively. Moreover, the peaks were due to the Si-O-Si groups having a higher intensity than the Si-O-C and Si-OH groups. Furthermore, the peak at 1721 cm−1 on GO (C=O) completely disappeared, suggesting the conversion of carbonyl groups into Si-O-C bonds.
The FTIR analysis shown in Figure 3 proves the existence of modified silica nanoparticles. The bands at 810 and 475 cm−1 found in the FTIR spectrum of silica nanoparticles illustrated in Figure 3 can be attributed, respectively, to the stretching and bending of Si-O-Si bonds. Furthermore, siloxane vibrations from (SiO)n groups are implicated according to the absorption bands at 1100 cm−1. The O-H stretching band of the surface silanol groups and the remaining adsorbed water molecules are responsible for the peaks at 3450 and 1630 cm−1, respectively, which developed after the APTES modification of the silica nanoparticles (Figure 3). The bands around 2800 cm−1 can be connected to the asymmetrical and symmetrical stretching vibrations of CH2 groups after silica nanoparticles were modified with APTES (Figure 3). The bending vibration of the N-H group causes the appearance of a new band at 1560 cm−1 in response to the addition of an amine group. Furthermore, the stretching vibration of the N-H group can be linked to the peak around 3500 cm−1. These combined results show that amine-functionalized silica nanoparticles were successfully prepared.
Raman spectroscopy is a standard non-destructive technique employed for the characterization of carbon-based materials [36]. The oxidation of graphite to GO can be further confirmed by Raman spectroscopy. Characteristic peaks of graphitic carbon-based materials are the G and D bands and their overtones [37]. As shown in Figure 4, characteristic bands for graphite and GO appear at around 1300–1350 cm−1 (D band) and between 1550 and 1600 cm−1 (G band) [26,38]. Their intensity ratio (ID/IG) indicates the degree of the disorder and defects [38]. A universal observation is that the higher the disorder of graphite, the broader the G and D bands, which have a higher relative intensity compared to the G band [39]. As shown in Figure 3, it is observed that there is a small shift for GO, and, also, the bands are broader than they are in graphite. The G peak and D peak of GO are shifted to higher frequencies with respect to that of graphite. The increase in intensity and shifting of the D peak to a higher wavelength corresponds to the presence of defects and the distortion of the sp2 crystal structure of graphite [37]. The 2D band appears at 2660 cm−1 [40]. The intensity of the 2D band is smaller after oxidation because of the breaking of the stacking order and incorporation of functional groups due to the oxidation reaction [41]. Thus, the intense D band in GO confirms the oxygen functionalization of graphite. The ratio of ID and IG portrays the defect window regime in GO and GO/SiO2. In general, the oxidation of graphite is accomplished by the increase of ID/IG. Hence, the higher the value of this ID/IG ratio, the more there will be structural defects or disorders in the system. Table 1 shows the ID/IG values of graphite, GO, and GO/SiO2.
The ID/IG ratios of GO and GO/SiO2 were 0.05 and 0.60, respectively. Thus, the increased value of the ID/IG ratio of GO/SiO2 indicates an increased degree of disorder and improved lamellar spacing, which would assist in the better dispersion of GO-SiO2 nanoparticles, and this result is in line with the TEM images.
Transmission electron microscopy (TEM) is considered as an effective method for the morphological studies of GO. With increases in the oxidation level, GO becomes highly transparent, since it possesses a high amount of the oxygenated functional group, which makes them suitable for exfoliation into monolayers or just a few layers of GO after ultra-sonication. As shown in Figure 5, the TEM images of GO sheets are coupled with numerable wrinkles and ripples on them. The transparency level of the image indicates the presence of more layers and functionalization with oxygen-containing groups [42]. Moreover, the GO-coated silica shows a rough external surface, with visible irregular cavities (Figure 5b). This must be due to the nanoparticles embedded in the silica backbone. The TEM images containing the inner core are outer shell shows that the GO was successfully coated on the silica particles.

3.2. Effect of pH on Adsorption of Cr(VI) Ions

The pH is the key factor that affects the removal efficiency of metal ions. It affects the surface charge of the adsorbent, the degree of ionization, and the speciation of the adsorbate [43]. To study the effect of the pH on adsorption, experiments were conducted under the same experimental conditions, at a concentration of 30 ppm and a pH in the range of 3.0–9.0. The experimental results are shown in Table 2.
From Table 2, it is clear that the optimum pH for the efficient removal of chromium(VI) ions is pH 3. Basic medium, i.e., pH 9, leads to the least removal of Cr(VI) for both of the adsorbents (Figure 6). As reported by Zhao et al. [44] and Shirkhanloo et al. [45], at low pH values, the surface charge of GO is positive due to the protonation reaction. At lower pH values, the predominant Cr(VI) species mainly exists in the monovalent HCrO4 form, which is then gradually changed to the divalent CrO42− and Cr2O72− as the pH increases [46,47]. Hence, at a lower pH, the surface of GO and GO-SiO2 becomes protonated. Then, the positively charged surface of the adsorbents strongly attracts oppositely charged Cr(VI) species through electrostatic interaction, and the adsorption rate increases [48]. Moreover, the adsorption free energy of HCrO4 ion is lower than that of CrO42−, and, hence, HCrO4 is more favorably adsorbed than CrO42− at the same concentration [49,50]. As the pH increases, the availability of the OH ions’ concentration increases, thus causing them to compete with the Cr(VI) ions, leading to the reduction in percentage of adsorption [32]. Thus, this study reveals that the adsorption quantities of Cr(VI) ions at a lower pH are larger than those of at a higher pH.

3.3. Evaluation of the Adsorption Efficiency of GO and Si/GO at Different Concentrations of Solutions

Adsorption experiments were conducted with different concentrations of heavy metal solutions (Ci, 1–200 ppm) at pH 3. Table 3 shows the effect of initial concentration of the solution under the optimum pH.
At a concentration of 50 ppm, it was noticed that the percentages of adsorption of Cr(VI) by GO and GO-SiO2 were at the maximum, i.e., at optimum concentration; the removal percentage was 86.15% and 92.28% for GO and GO-SiO2, respectively. As the initial concentration increases, the adsorption percentage of Cr(VI) increases and then decreases for both the adsorbents. At lower concentrations, the number of Cr(VI) ions present in the solution is comparatively lower compared to the number of available sites on the GO and GO-SiO2 nanoparticles [51]. The highest heavy metal concentration (200 ppm) leads to the lowest removal percentage [52]. Imtithal et al. [31] reported that, at a lower concentration of metal ions, vacant sites are available for the adsorption of metal ions, and, hence, there is a higher adsorption rate. At higher concentrations, the capacity of the adsorbent becomes exhausted due to the non-availability of the surface sites; i.e., as the concentration increases, the surface sites of the adsorbents become saturated, leading to less or no more adsorption (Figure 7). Moreover, it is clear that the adsorption efficiency of GO-SiO2 is higher than GO at all concentrations. Although the functional groups, such as epoxy (C-O-C), hydroxyl (-OH), carboxyl (-COOH), and carbonyl (C=O), on the surface of GO promote the adsorption of metal ions, the aggregation of its layers due to strong interplanar interactions leads to a reduced metal ion adsorption capacity [53]. GO and amine-modified silica are good adsorbents, and their combination leads to the formation of an excellent adsorbent, i.e., GO-SiO2. The increases in the specific surface area and the number of binding sites of the GO-SiO2 composite contribute to its excellent removal efficiency [54]. Both amine-modified silica and graphene oxide (GO) are good adsorbents for removing heavy metals from wastewater. Due to the abundance of functional groups such as -OH, -COOH, and -C=O, GO has a wide surface area and high adsorption capacity. These groups have high interactions with heavy metal ions through coordination and electrostatic interactions, making GO an efficient adsorbent. On the other hand, amino functional groups in amine-modified silica show potent metal-ion-chelating capabilities. It is a great adsorbent for metal ions because the amino groups can form complexes with metal ions through coordination interactions. The GO-SiO2 composite that results from the combination of these two substances has improved the adsorption abilities for the removal of Cr(VI) from wastewater. The amine-modified silica provides chelating sites for Cr(VI) ions, whereas GO provides a high surface area for the adsorption of metal ions. A better adsorption effectiveness can also result from the presence of amine functional groups on the silica surface, which can encourage the dispersion of GO and prevent its aggregation. The interaction between protonated amine and hydro chromate anion exhibits a strong affinity for graphene oxide, primarily through cation-pi and electrostatic interactions [23,55,56]. Chromium(VI) is easily adsorbed onto the surface of GO due to hydrogen bonding and the electrostatic interaction of CrO42− with the oxygen functionalities in GO. Figure 8 illustrates the interaction between the amine and silica, graphene oxide with amine-modified silica, and Cr(VI).

3.4. Comparison of Removal Efficiency of Cr(VI) Ions Using Different Adsorbents

In order to compare the efficiency of synthesized nanoparticles for the removal of Cr(VI) ions from water with the ones reported in the literature, it is appropriate to compare the removal percentage or adsorbance capacity of graphene oxide (GO) and graphene-oxide-coated silica nanoparticles (GO/SiO2). In this paper, a comparison of the removal percentage of different adsorbents with synthesized adsorbents is presented Table 4.
It can be observed that an acidic pH in the range from 2.0 to 3.0 was found to be optimum in most of the studies. An acidic pH is often preferred for the removal of Cr(VI) from wastewater because it helps to convert the highly toxic and mobile Cr(VI) into its less toxic and less mobile Cr(III) form. This is because, at lower pH levels, the reduction of Cr(VI) to Cr(III) is more favorable and occurs more quickly. The reduction of Cr(VI) to Cr(III) occurs through various mechanisms, such as chemical reduction, microbial reduction, or electrochemical reduction. However, the efficiency of these mechanisms is affected by the pH of the solution. At a low pH, the chemical reduction of Cr(VI) is favored, while at higher pH values, microbial reduction is more favorable. Moreover, at lower pH values, the solubility of Cr(III) is significantly reduced, which leads to its precipitation and removal from the solution. This precipitation can occur in the form of various Cr(III) compounds, such as hydroxides, oxides, or salts. These precipitates are less soluble and less mobile than Cr(VI), reducing the risk of groundwater contamination and other adverse environmental impacts.
Compared with other adsorbents, a high absorbance percentage was recorded for synthesized GO and GO/SiO2 at pH 3. Other adsorbents that are reported in the literature with a high removal percentage showed maximum adsorption at a pH < 3. A high surface area and the presence of oxygen-containing functional groups on the surface of GO facilitate the removal efficiency by complexation with Cr(VI) ions. The higher removal efficiency of GO/SiO2 can be justified by its enhanced properties.

4. Conclusions

In this study, graphene oxide (GO) and graphene-oxide-coated silica nanoparticles were successfully synthesized and characterized by FTIR and Raman spectroscopy and TEM. The synthesized nanoparticles were used for the removal of chromium (VI) ions from water. The pH was optimized to 3, and the maximum adsorption percentage of Cr(VI) was observed at a concentration of 50 ppm for both of the adsorbents. A difference of 6.13% adsorption was observed between GO and GO-coated silica nanoparticles at optimum conditions. The results of the present study showed that both GO and GO-coated silica nanoparticles are effective adsorbents for the Cr(VI) removal from water. The adsorption of Cr(VI) on these synthesized adsorbents is strongly dependent on the pH values and initial concentration of the solution. GO-coated silica nanoparticles are excellent adsorbents compared to GO alone, and we highly recommend that they can be used in water treatment for removing Cr(VI). Further optimization studies are essential for the various applications of these nanoparticles.

Author Contributions

Conceptualization, K.P.J., S.T. and S.V.P.; methodology, K.P.J., S.A., S.K.S., S.T. and S.V.P.;validation, K.P.J., S.T., S.V.P. and S.A.; formal analysis, K.P.J., S.A., S.K.S., P.V., J.S.G. and S.V.P.; investigation, K.P.J., S.A., S.K.S., P.V., J.S.G. and S.V.P.; resources, K.P.J., S.T. and S.V.P.; data curation, K.P.J., S.T. and S.V.P.; writing—original draft preparation, K.P.J., S.A., S.K.S., P.V. and J.S.G.; writing—review and editing, K.P.J., S.T. and S.V.P.; supervision, S.T. and S.V.P.; project administration, S.T.; funding acquisition, S.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the DST Nanomission project (SR/NM/NT-1054/2015) by the Department of Science and Technology, Government of India, Defence Research and Development Organization (DRDO), Naval Research Board (NRB-429/MAT/18-19), RUSA scheme under Ministry of Education, Government of India, Mahatma Gandhi University for providing University JRF (4446/AcA6/2022/MGU) and Department of Science and Technology (DST) through Innovation of Science Pursuit for Inspire Research programme (INSPIRE-IF190284).

Data Availability Statement

Data are available upon request.

Acknowledgments

We would like to acknowledge Anu A.S. (TEM Engineer) at the International and Inter University Centre for Nanoscience and Nanotechnology, Mahatma Gandhi University, Kottayam, for performing the TEM Analysis.

Conflicts of Interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

References

  1. Järup, L. Medical bulletin and undefined 2003: Hazards of heavy metal contamination. Br. Med. Bull. 2003, 68, 167–182. [Google Scholar] [CrossRef] [PubMed]
  2. Jaishankar, M.; Tseten, T.; Anbalagan, N.; Mathew, B.B.; Beeregowda, K.N. Toxicity, mechanism and health effects of some heavy metals. Interdiscip. Toxicol. 2014, 7, 60. [Google Scholar] [CrossRef]
  3. Brevik, E.C.; Burgess, L.C. Soils and Human Health; CRC Press: Boca Raton, FL, USA, 2012. [Google Scholar]
  4. Sperling, M. Chromium. Encycl. Anal. Sci. Second Ed. 2004, 113. [Google Scholar] [CrossRef]
  5. Rodríguez, M.C.; Barsanti, L.; Passarelli, V.; Evangelista, V.; Conforti, V.; Gualtieri, P. Effects of chromium on photosynthetic and photoreceptive apparatus of the alga Chlamydomonas reinhardtii. Environ. Res. 2007, 105, 234. [Google Scholar] [CrossRef]
  6. Saha, R.; Nandi, R.; Saha, B. Chemistry, and undefined 2011: Sources and toxicity of hexavalent chromium. J. Coord. Chem. 2011, 64, 1782–1806. [Google Scholar] [CrossRef]
  7. Li, Y.; Zhou, Q.; Ren, B.; Luo, J.; Yuan, J.; Ding, X.; Bian, H.; Yao, X. Trends and Health Risks of Dissolved Heavy Metal Pollution in Global River and Lake Water from 1970 to 2017. Rev. Environ. Contam. Toxicol. 2020, 251, 1–24. [Google Scholar]
  8. Tumolo, M.; Ancona, V.; De Paola, D.; Losacco, D.; Campanale, C.; Massarelli, C.; Uricchio, V.F. Chromium pollution in European water, sources, health risk, and remediation strategies: An overview. Int. J. Environ. Res. Public Health 2020, 17, 5438. [Google Scholar] [CrossRef]
  9. Vardhan, K.H.; Kumar, P.S.; Panda, R.C. A review on heavy metal pollution, toxicity and remedial measures: Current trends and future perspectives. J. Mol. Liq. 2019, 290, 111197. [Google Scholar] [CrossRef]
  10. Lisiak-Zielińska, M.; Borowiak, K.; Budka, A.; Kanclerz, J.; Janicka, E.; Kaczor, A.; Żyromski, A.; Biniak-Pieróg, M.; Podawca, K.; Mleczek, M.; et al. How polluted are cities in central Europe?—Heavy metal contamination in Taraxacum officinale and soils collected from different land use areas of three representative cities. Chemosphere 2021, 266, 129113. [Google Scholar] [CrossRef]
  11. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2021, 92, 407. [Google Scholar] [CrossRef]
  12. Singh, K.; Renu, N.A.; Agarwal, M. Methodologies for removal of heavy metal ions from wastewater: An overview. Interdiscip. Environ. Rev. 2017, 18, 124. [Google Scholar] [CrossRef]
  13. Zhao, D.; Gao, X.; Chen, S.; Xie, F.; Feng, S.; Alsaedi, A.; Hayat, T.; Chen, C. Interaction between U(VI) with sulfhydryl groups functionalized graphene oxides investigated by batch and spectroscopic techniques. J. Colloid Interface Sci. 2018, 524, 129. [Google Scholar] [CrossRef] [PubMed]
  14. Malik, L.A.; Bashir, A.; Qureashi, A.; Pandith, A.H. Detection and removal of heavy metal ions: A review. Environ. Chem. Lett. 2019, 17, 1495. [Google Scholar] [CrossRef]
  15. Vakili, M.; Deng, S.; Cagnetta, G.; Wang, W.; Meng, P.; Liu, D.; Yu, G. Regeneration of chitosan-based adsorbents used in heavy metal adsorption: A review. Sep. Purif. Technol. 2019, 224, 373. [Google Scholar] [CrossRef]
  16. Arora, R. Adsorption of heavy metals—A review. Mater. Today Proc. 2019, 18, 4745. [Google Scholar] [CrossRef]
  17. Prachi; Gautam, P.; Madathil, D.; Nair, A.N.B. Nanotechnology in waste water treatment: A review. Int. J. ChemTech Res. 2013, 5, 2303. [Google Scholar]
  18. Gangadhar, G.; Maheshwari, U.; Gupta, S. Application of nanomaterials for the removal of pollutants from effluent streams. Nanosci. Nanotechnol.-Asia 2012, 2, 140–150. [Google Scholar] [CrossRef]
  19. Potts, J.R.; Dreyer, D.R.; Bielawski, C.W.; Ruoff, R.S. Graphene-based polymer nanocomposites. Polymer 2011, 52, 5–25. [Google Scholar] [CrossRef]
  20. Sadegh, H.; Ali, G.A.; Gupta, V.K.; Makhlouf, A.S.H.; Shahryari-Ghoshekandi, R.; Nadagouda, M.N.; Sillanpää, M.; Megiel, E. The role of nanomaterials as effective adsorbents and their applications in wastewater treatment. J. Nanostructure Chem. 2017, 7, 1–14. [Google Scholar] [CrossRef]
  21. Liu, X.; Ma, R.; Wang, X.; Ma, Y.; Yang, Y.; Zhuang, L.; Zhang, S.; Jehan, R.; Chen, J.; Wang, X. Graphene oxide-based materials for efficient removal of heavy metal ions from aqueous solution: A review. Environ. Pollut. 2019, 252, 62. [Google Scholar] [CrossRef]
  22. Li, L.; Zhao, L.; Ma, J.; Tian, Y.; Tian, Y. Preparation of graphene oxide/chitosan complex and its adsorption properties for heavy metal ions. Green Process. Synth. 2020, 9, 294. [Google Scholar] [CrossRef]
  23. Kumar, A.S.K.; Kakan, S.S.; Rajesh, N. A novel amine impregnated graphene oxide adsorbent for the removal of hexavalent chromium. Chem. Eng. J. 2013, 230, 328. [Google Scholar] [CrossRef]
  24. Xu, L.; Wang, J. The application of graphene-based materials for the removal of heavy metals and radionuclides from water and wastewater. Crit. Rev. Environ. Sci. Technol. 2017, 47, 1042. [Google Scholar] [CrossRef]
  25. Deng, J.H.; Zhang, X.R.; Zeng, G.M.; Gong, J.L.; Niu, Q.Y.; Liang, J. Simultaneous removal of Cd (II) and ionic dyes from aqueous solution using magnetic graphene oxide nanocomposite as an adsorbent. Chem. Eng. J. 2013, 226, 189–200. [Google Scholar] [CrossRef]
  26. Peer, F.E.; Bahramifar, N.; Younesi, H. Removal of Cd (II), Pb (II) and Cu (II) ions from aqueous solution by polyamidoamine dendrimer grafted magnetic graphene oxide nanosheets. J. Taiwan Inst. Chem. Eng. 2018, 87, 225–240. [Google Scholar] [CrossRef]
  27. Shahzad, A.; Miran, W.; Rasool, K.; Nawaz, M.; Jang, J.; Lim, S.R.; Lee, D.S. Heavy metals removal by EDTA-functionalized chitosan graphene oxide nanocomposites. RSC Adv. 2017, 7, 9764. [Google Scholar] [CrossRef]
  28. Xu, J.; Cao, Z.; Zhang, Y.; Yuan, Z.; Lou, Z.; Xu, X.; Wang, X. A review of functionalized carbon nanotubes and graphene for heavy metal adsorption from water: Preparation, application, and mechanism. Chemosphere 2018, 195, 351–364. [Google Scholar] [CrossRef] [PubMed]
  29. Manyangadze, M.; Chikuruwo, N.M.H.; Narsaiah, T.B.; Chakra, C.S.; Charis, G.; Danha, G.; Mamvura, T.A. Adsorption of lead ions from wastewater using nano silica spheres synthesized on calcium carbonate templates. Heliyon 2020, 6, e05309. [Google Scholar] [CrossRef]
  30. Hummers, W.S., Jr.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  31. Sheet, I.; Kabbani, A.; Holail, H. Removal of heavy metals using nanostructured graphite oxide, silica nanoparticles and silica/graphite oxide composite. Energy Procedia 2014, 50, 130. [Google Scholar] [CrossRef]
  32. Mondal, N.K.; Chakraborty, S. Adsorption of Cr(VI) from aqueous solution on graphene oxide (GO) prepared from graphite: Equilibrium, kinetic and thermodynamic studies. Appl. Water Sci. 2020, 10, 1–10. [Google Scholar] [CrossRef]
  33. Lee, D.W.; De Los Santos V, L.; Seo, J.W.; Felix, L.L.; Bustamante D, A.; Cole, J.M.; Barnes, C.H.W. The structure of graphite oxide: Investigation of its surface chemical groups. J. Phys. Chem. B 2010, 114, 5723–5728. [Google Scholar]
  34. Li, D.; Müller, M.B.; Gilje, S.; Kaner, R.B.; Wallace, G.G. Processable aqueous dispersions of graphene nanosheets. Nat. Nanotechnol. 2008, 3, 101. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, W.; Langmuir, H.C. Silica-graphene oxide hybrid composite particles and their electroresponsive characteristics. ACS Publ. 2012, 28, 7055. [Google Scholar] [CrossRef] [PubMed]
  36. Björkman, Å. Thermische Klärschlammbehandlung. Schweiz. Z. Hydrol. 1969, 31, 632. [Google Scholar]
  37. Eda, G.; Chhowalla, M. Chemically derived graphene oxide: Towards large-area thin-film electronics and optoelectronics. Adv. Mater. 2010, 22, 2392. [Google Scholar] [CrossRef]
  38. Sun, L.; Yu, H.; Fugetsu, B. Graphene oxide adsorption enhanced by in situ reduction with sodium hydrosulfite to remove acridine orange from aqueous solution. J. Hazard. Mater. 2012, 203–204, 101. [Google Scholar] [CrossRef]
  39. Kudin, K.N.; Ozbas, B.; Schniepp, H.C.; Prud, R.K.; Aksay, I.A.; Car, R. Raman Spectra of Graphite Oxide and Functionalized Graphene Sheets. Nano Lett. 2008, 8, 36–41. [Google Scholar]
  40. Gu, J.; She, J.; Yue, Y. Micro/nanoscale thermal characterization based on spectroscopy techniques. ES Energy Environ. 2020, 9, 15–27. [Google Scholar] [CrossRef]
  41. Khalili, D. Graphene oxide: A promising carbocatalyst for the regioselective thiocyanation of aromatic amines, phenols, anisols and enolizable ketones by hydrogen peroxide/KSCN in water. N. J. Chem. 2016, 40, 2547. [Google Scholar] [CrossRef]
  42. Krishnamoorthy, K.; Veerapandian, M.; Yun, K.; Kim, S.J. The chemical and structural analysis of graphene oxide with different degrees of oxidation. Carbon N. Y. 2013, 53, 38. [Google Scholar] [CrossRef]
  43. Mishima, K.; Du, X.; Sekiguchi, S.; Kano, N. Experimental and theoretical studies on the adsorption and desorption mechanisms of chromate ions on cross-linked chitosan. J. Funct. Biomater. 2017, 8, 3. [Google Scholar] [CrossRef]
  44. Zhao, G.; Li, J.; Ren, X.; Chen, C.; Wang, X. Few-layered graphene oxide nanosheets as superior sorbents for heavy metal ion pollution management. Environ. Sci. Technol. 2011, 45, 10454. [Google Scholar] [CrossRef]
  45. Shirkhanloo, H.; Khaligh, A.; Mousavi, H.Z.; Rashidi, A. Graphene oxide-packed micro-column solid-phase extraction combined with flame atomic absorption spectrometry for determination of lead (II) and nickel (II) in water. J. Environ. Anal. Chem. 2015, 95, 16–32. [Google Scholar] [CrossRef]
  46. Zhang, L.-H.; Sun, Q.; Yang, C.; Lu, A.-H. Synthesis of magnetic hollow carbon nanospheres with superior microporosity for efficient adsorption of hexavalent chromium ions. Sci. China Mater. 2015, 58, 611. [Google Scholar] [CrossRef]
  47. Hu, J.; Chen, C.; Zhu, X.; Wang, X. Removal of chromium from aqueous solution by using oxidized multiwalled carbon nanotubes. J. Hazard. Mater. 2009, 162, 1542. [Google Scholar] [CrossRef] [PubMed]
  48. Matei, E.; Predescu, A.M.; Râpă, M.; Tarcea, C.; Pantilimon, C.M.; Favier, L.; Berbecaru, A.C.; Sohaciu, M.; Predescu, C. Removal of Chromium(VI) from Aqueous Solution Using a Novel Green Magnetic Nanoparticle–Chitosan Adsorbent. Anal. Lett. 2019, 52, 2416. [Google Scholar] [CrossRef]
  49. Hu, J.; Lo, I.M.C.; Chen, G. Removal of Cr(VI) by magnetite nanoparticle. Water Sci. Technol. 2004, 50, 139. [Google Scholar] [CrossRef]
  50. Liu, W.; Yang, L.; Xu, S.; Chen, Y.; Liu, B.; Li, Z.; Jiang, C. RSC Advances Efficient removal of hexavalent chromium from water by an adsorption-reduction mechanism with sandwiched nanocomposites. RSC Adv. 2018, 8, 15087–15093. [Google Scholar] [CrossRef]
  51. Samuel, M.S.; Bhattacharya, J.; Raj, S.; Santhanam, N.; Singh, H.; Singh, N.D.P. Efficient removal of Chromium(VI) from aqueous solution using chitosan grafted graphene oxide (CS-GO) nanocomposite. Int. J. Biol. Macromol. 2019, 121, 285. [Google Scholar] [CrossRef] [PubMed]
  52. Konicki, W.; Aleksandrzak, M.; Mijowska, E. Equilibrium and kinetics studies for the adsorption of Ni2+ and Fe3+ ions from aqueous solution by graphene oxide. Polish J. Chem. Technol. 2017, 19, 120. [Google Scholar] [CrossRef]
  53. Donga, C.; Mishra, S.; Aziz, A.; Ndlovu, L.; Kuvarega, A.; Mishra, A.K. 3-Aminopropyl) Triethoxysilane (APTES) Functionalized Magnetic Nanosilica Graphene Oxide (MGO) Nanocomposite for the Comparative Adsorption of the Heavy Metal [Pb(II), Cd(II) and Ni(II)] Ions from Aqueous Solution. J. Inorg. Organomet. Polym. Mater. 2022, 32, 2235–2248. [Google Scholar] [CrossRef]
  54. Ma, M.; Li, H.; Xiong, Y.; Dong, F. Rational design, synthesis, and application of silica/graphene-based nanocomposite: A review. Mater. Des. 2021, 198, 109367. [Google Scholar] [CrossRef]
  55. Kalidhasan, S.; Kumar, A.S.K.; Rajesh, V.; Rajesh, N. Enhanced adsorption of hexavalent chromium arising out of an admirable interaction between a synthetic polymer and an ionic liquid. Chem. Eng. J. 2013, 222, 454. [Google Scholar] [CrossRef]
  56. Kumar, A.S.K.; Rajesh, N.; Kalidhasan, S.; Rajesh, V. An enhanced adsorption methodology for the detoxification of chromium using n-octylamine impregnated Amberlite XAD-4 polymeric sorbent. J. Environ. Sci. Health—Part A Toxic/Hazard. Subst. Environ. Eng. 2011, 46, 1598. [Google Scholar] [CrossRef]
  57. Zhang, L.; Song, F.; Wang, S.; Wang, H.; Yang, W.; Li, Y. Efficient removal of hexavalent chromium and congo red by graphene oxide/silica nanosheets with multistage pores. ACS Publ. 2020, 65, 4368. [Google Scholar] [CrossRef]
  58. Hu, J.; Lo, I.M.C.; Chen, G. Fast removal and recovery of Cr(VI) using surface-modified jacobsite (MnFe2O4) nanoparticles. Langmuir 2005, 21, 11173. [Google Scholar] [CrossRef] [PubMed]
  59. Hu, J.; Chen, G.; Lo, I.M.C. Removal and recovery of Cr(VI) from wastewater by maghemite nanoparticles. Water Res. 2005, 39, 4528. [Google Scholar] [CrossRef]
  60. Najafabadi, H.H.; Irani, M.; Rad, L.R.; Haratameh, A.H.; Haririan, I. Removal of Cu2+, Pb2+ and Cr6+ from aqueous solutions using a chitosan/graphene oxide composite nanofibrous adsorbent. RSC Adv. 2015, 5, 16532. [Google Scholar] [CrossRef]
  61. Liu, Y.; Shan, H.; Zeng, C.; Zhan, H.; Pang, Y. Removal of Cr(VI) from Wastewater Using Graphene Oxide Chitosan Microspheres Modified with α–FeO(OH). Materials 2022, 15, 4909. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Synthesis of amine modified silica and GO/SiO2.
Figure 1. Synthesis of amine modified silica and GO/SiO2.
Crystals 13 01047 g001
Figure 2. FTIR spectra of graphite, GO, and GO/SiO2.
Figure 2. FTIR spectra of graphite, GO, and GO/SiO2.
Crystals 13 01047 g002
Figure 3. FTIR spectra of SiO2 and amine-modified SiO2.
Figure 3. FTIR spectra of SiO2 and amine-modified SiO2.
Crystals 13 01047 g003
Figure 4. Raman spectra of GO and graphite.
Figure 4. Raman spectra of GO and graphite.
Crystals 13 01047 g004
Figure 5. (a) Transmission electron microscopy (TEM) images of synthesized graphene oxide (GO) at different magnifications. (b) TEM images of graphene-oxide-coated silica in different magnifications (100 nm, 20 nm and 5 nm).
Figure 5. (a) Transmission electron microscopy (TEM) images of synthesized graphene oxide (GO) at different magnifications. (b) TEM images of graphene-oxide-coated silica in different magnifications (100 nm, 20 nm and 5 nm).
Crystals 13 01047 g005
Figure 6. Effect of pH on the adsorption of Cr(VI) ion.
Figure 6. Effect of pH on the adsorption of Cr(VI) ion.
Crystals 13 01047 g006
Figure 7. Adsorption of Cr(VI) using GO and Si/GO at different concentrations.
Figure 7. Adsorption of Cr(VI) using GO and Si/GO at different concentrations.
Crystals 13 01047 g007
Figure 8. Schematic representation of the Cr(VI) adsorption mechanism.
Figure 8. Schematic representation of the Cr(VI) adsorption mechanism.
Crystals 13 01047 g008
Table 1. ID/IG values of graphite, GO, and GO/SiO2.
Table 1. ID/IG values of graphite, GO, and GO/SiO2.
Sample CodeID/IG
Graphite0.05
GO0.60
GO/SiO20.94
Table 2. Percentage of adsorption of Cr(VI) ions using GO and GO-coated-silica nanoparticles at 30 ppm.
Table 2. Percentage of adsorption of Cr(VI) ions using GO and GO-coated-silica nanoparticles at 30 ppm.
pH of the
Sample
Initial Concentration
(Ci)
(ppm)
Final Concentration (Ce) (ppm)Adsorbance
Percentage (%)
GOSi/GOGOSi/GO
3308.083315.7724073.06 ± 0.680.76 ± 0.125
53010.0987110.9756766.34 ± 0.363.41 ± 0.312
73010.0040010.6862366.65 ± 0.02564.38 ± 0.421
93010.1146511.0076966.28 ± 0.53260.97 ± 0.25
Table 3. Percentage of adsorption of Cr(VI) ions from different concentrations of samples, using GO and GO-coated silica nanoparticles.
Table 3. Percentage of adsorption of Cr(VI) ions from different concentrations of samples, using GO and GO-coated silica nanoparticles.
Initial Concentration of the Sample (Ci)
(ppm)
Concentration after Adsorption (Ce)
(ppm)
Percentage of Adsorption (%)
GOSi/GOGOSi/GO
10.615040.2976038.49 ± 0.2570.24 ± 0.65
105.946742.4663040.53 ± 012675.34 ± 35
308.083315.7724073.06 ± 0.680.76 ± 0.69
506.923333.8582586.15 ± 0.892.28 ± 0.45
10065.3400952.3145334.66 ± 0.5847.69 ± 0.52
200132.2878118.822633.86 ± 0.8740.59 ± 0.85
Table 4. Summary of removal percentage of Cr(VI) using various adsorbents.
Table 4. Summary of removal percentage of Cr(VI) using various adsorbents.
No.AdsorbentOptimum pHRemoval % (a)/Adsorption Capacity (mg/g) (b)Reference
1Graphene oxide/silica nanosheets with multistage pores290 (a)[57]
2Surface-modified Jacobsite (MnFe2O4) nanoparticles231.55 (b)[58]
3GO prepared from Graphite492.8 (a)[32]
4Maghemite nanoparticles2.597.3 (a)[59]
5Magnetite nanoparticle2.599.8 (a)[49]
6Nanostructured graphite oxide363 (a)[31]
7Silica/graphite oxide composite360.24 (b)[31]
8Chitosan-grafted graphene oxide (CS-GO) nanocomposite296 (a)[51]
9Chitosan GO3310.4 (b)[60]
10Graphene oxide chitosan microspheres modified with α–FeO(OH)397.69 (a)[61]
11Trioctylamine-exfoliated graphene oxide (TOA–EGO)2.5–396.3[23]
12Graphene oxide (GO)386.15 ± 0.8 (a)Present study
13GO-SiO2392.28 ± 0.45 (a)Present study
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jibin, K.P.; Augustine, S.; Velayudhan, P.; George, J.S.; Krishnageham Sidharthan, S.; Paulose, S.V.; Thomas, S. Unleashing the Power of Graphene-Based Nanomaterials for Chromium(VI) Ion Elimination from Water. Crystals 2023, 13, 1047. https://doi.org/10.3390/cryst13071047

AMA Style

Jibin KP, Augustine S, Velayudhan P, George JS, Krishnageham Sidharthan S, Paulose SV, Thomas S. Unleashing the Power of Graphene-Based Nanomaterials for Chromium(VI) Ion Elimination from Water. Crystals. 2023; 13(7):1047. https://doi.org/10.3390/cryst13071047

Chicago/Turabian Style

Jibin, Keloth Paduvilan, Silpa Augustine, Prajitha Velayudhan, Jesiya Susan George, Sisanth Krishnageham Sidharthan, Sylas Variyattel Paulose, and Sabu Thomas. 2023. "Unleashing the Power of Graphene-Based Nanomaterials for Chromium(VI) Ion Elimination from Water" Crystals 13, no. 7: 1047. https://doi.org/10.3390/cryst13071047

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop