Next Article in Journal
Elite Sport and Sustainable Psychological Well-Being
Next Article in Special Issue
Radiation Shielding Enhancement of Polyester Adding Artificial Marble Materials and WO3 Nanoparticles
Previous Article in Journal
Web GIS for Sustainable Education: Towards Natural Disaster Education for High School Students
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sustainable Remedy Waste to Generate SiO2 Functionalized on Graphene Oxide for Removal of U(VI) Ions

1
Nuclear Materials Authority, El Maadi, Cairo P.O. Box 530, Egypt
2
Department of Chemistry, College of Science, Princess Nourah bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
3
Institute of Physics and Technology, Ural Federal University, St. Mira, 19, 620002 Yekaterinburg, Russia
4
Department of Nuclear Medicine Research, Institute for Research and Medical Consultations (IRMC), Imam Abdulrahman Bin Faisal University (IAU), P.O. Box 1982, Dammam 31441, Saudi Arabia
5
Department of Physics, Faculty of Science, Isra University, Amman 11622, Jordan
*
Authors to whom correspondence should be addressed.
Sustainability 2022, 14(5), 2699; https://0-doi-org.brum.beds.ac.uk/10.3390/su14052699
Submission received: 13 January 2022 / Revised: 1 February 2022 / Accepted: 11 February 2022 / Published: 25 February 2022
(This article belongs to the Special Issue Sustainable Polymer Composites and Nanocomposites)

Abstract

:
The Hummer process is applied to generate graphene oxide from carbon stocks’ discharged Zn-C batteries waste. SiO2 is produced from rice husks through the wet process. Subsequently, SiO2 reacted with graphene oxide to form silica/graphene oxide (SiO2/GO) as a sorbent material. XRD, BET, SEM, EDX, and FTIR were employed to characterize SiO2/GO. Factors affecting U(VI) sorption on SiO2/GO, including pH, sorption time, a dosage of SiO2/GO, U(VI) ions’ concentration, and temperature, were considered. The experimental data consequences indicated that the uptake capacity of SiO2/GO towards U(VI) is 145.0 mg/g at a pH value of 4.0. The kinetic calculations match the pseudo second-order model quite well. Moreover, the sorption isotherm is consistent with the Langmuir model. The sorption procedures occur spontaneously and randomly, as well as exothermically. Moreover, SiO2/GO has essentially regenerated with a 0.8 M H2SO4 and 1:50 S:L phase ratio after 60 min of agitation time. Lastly, the sorption and elution were employed in seven cycles to check the persistent usage of SiO2/GO.

1. Introduction

Uranium ions may develop into the environment through applications, industry, and mining, posing risks to health of humans and biological situations because of radiation and toxicity [1,2]. Nevertheless, by the widespread use of atomistic energy and further applications, U(VI) is exceptionally loose in running solutions; it threatens environmental safety and biological balance [3,4]. Thus, U(VI) elimination and deposition from liquid origins are important for nucleate energy engendering and peripheral protection [5,6]. To acquire U(VI) outputting from radioactive wastewater that is financially workable, the extracting materials must maintain tremendous extracting ability, selectivity, departure rate, and reusability applications. Until recently, most studies have been directed towards the application of extracting materials for U(VI) adsorption due to its high performance, capacity, low cost, and broad versatility [7,8].
Recently, numerous investigations were conducted for the removal and extraction of U(VI), comprising ion exchange [9,10,11,12], precipitation [13], liquid extraction [14], film separation [14,15], and sorption [16,17,18,19,20]. Adsorption methods were commonly examined because they remain faithful and are low risk and low cost; furthermore, the separation power could be enormously promoted by originating relevant sorption materials [21,22,23,24,25,26,27]. Graphene-based sorbents are commonly criticized as subjective sorbents due to their distinctive qualities, such as low mass, chemical force, and fundamental variability.
Recently, advances in current nano-technology, science, and standardized hybrid permeable substances have gained considerable attention for sorption and separation purposes, owing to their orderly, uniform, and interpenetrating mesoporous, tunable hole sizes and extended surface areas. Numerous organic-inorganic porous hybrids evolved for excluding heavy metals (Ba(II), Hg(II), Cd(II), Cu(II), and Pb(II)) from wastewater [28,29,30]. Graphene-based mesoporous silica was utilized for Pb(II) ions’ elimination [31]. Reduced graphene oxide was combined with heavy metal oxides to enhance the efficiency of sonocatalytic degradation and photocatalytic hydrogen production. Multifunctional WO3 and tungsten trioxide/reduced graphene oxide (WO3/rGO) composite photocatalysts were used to investigate the sonocatalytic degradation of toxic Congo red azo dye as well as photocatalytic hydrogen production. Furthermore, Platinum has been used as a co-catalyst with WO3/rGO to boost hydrogen production [32]. Reduced graphene oxide/SnO2 (rGO/SnO2) nanocomposite used ultrasound-generated cavitational events to improve the uniform deposition of SnO2 on rGO nanotubes. These rGO/SnO2 nanocomposites were explored for their potential gas sensing and electrochemical implementations [33]. The spongy ball-like CuO and CuO/rGO photocatalysts achieved optical results, which proved that the bandgap is within the visible range of the solar spectrum. In addition, the CuO/rGO composite produce more photocatalytic hydrogen compared to CuO [34].
Furthermore, graphene oxide-related materials have been fabricated widely and exploited in several applications. Nevertheless, several convergent surveys have been performed, particularly inorganic/GO for uranium removal. It was assumed that the recent advancement of inorganic/GO for adsorption of U(VI) promotes the refinement of inorganic/GO materials for uranium disposal and recovery purposes. GO/MnO2 and magnetite nanoparticle/GO composites have been utilized for U(VI) removal [35,36]. Fe-Ni/GO composite was manufactured to monitor the exclusion of U(VI) from liquid solution in environmental conditions [37]. Magnetic Fe3O4/SiO2 composite particles were applied for swift removal of U(VI) from liquid solutions. The maximum uranium adsorption capacity onto magnetic Fe3O4/SiO2 composite particles was estimated to be about 52 mg/g at 25 °C [38]. A functional surface layer based on quaternary ammonium is produced in situ on the surface of silica-coated magnetic nanoparticles for uranium removal, with an uptake efficiency of 87 mg/g at 120 min of equilibrium time [39]. The graphene oxide/bentonite composite was synthesized to adsorb uranium ions at room temperature with 234.19 mg/g sorption capacity [40].
Here, the SiO2/GO sorbent was produced from the obtained graphene oxide from spent carbon rods batteries, and silica was established from the rice husk to remove U(VI) ions from the aqueous solution. The experimental parameters are pH, initial U(VI) concentration, contact time, SiO2/GO dose, and temperature, with which we examined U(VI) adsorption by batch experiments. In addition, adsorption isotherms and thermodynamic considerations were evaluated.

2. Materials and Methods

2.1. Chemicals and Techniques

A UV/Vis 6405 spectrophotometer was utilized for analyzing U(VI) at 650.0 nm with Arsenazo III (as a chromophore) [41]. Furthermore, SEM, EDX, BET, and XRD were exploited to designate the morphologies of GO, SiO2, SiO2/GO, and U/SiO2/GO. FTIR (IRPrestige-21, Shimadzu, Kyoto, Japan) analysis with IR steadfastness software was utilized to expose the function groups. A FT-Raman spectrometer RAMII (Bruker, Billerica, MA, USA) equipped with a 500 mW Nd: YAG laser was used to collect Raman spectra at 1064 nm. The experimental conditions lead to a spectral resolution of 5 cm−1.
The chemicals and reagents can be used without purification. Arsenazo III, UO2(CH3COO)2·2H2O, phosphoric acid, FeSO4.7H2O, sodium hydroxide, and NaNO3 were obtained from Merch. HCl, hydrogen peroxide, methanol, ethyl acetate, and sulfuric acid were obtained from Sigma Aldrich. The chosen source for rice husk adopted here was a local rice harvester at Mina El-Qamh, Qalyubia Governorate, Egypt. The spent carbon rods were taken from Egyptian Zn-C battery solid wastes.

2.2. Fabrication of Silica

Silica is fabricated from low-cost, agricultural sources such as rice husk (RH) [42,43]. The collected RH (200 g) was washed by tap water to eliminate dust and further contaminations and then dried at 100 °C. Subsequently, it was ignited under suction; the black RH was refluxed with 1.0 M HCl at 80 °C and stirring constantly for 5 h. After cooling, the treated RH was filtered and washed with Milli-Q water until it became acid-free (pH 7.0) and then dried at 100 °C overnight; the attained product was calcinated at 750 °C for 8 h. A white silica product was acquired after calcination.
A 20 g white product was dissolved completely in 200 mL of 2.5 M NaOH at 70 °C under reflux. The produced sticky solution was filtered before being washed with 200 mL of boiled Mili-Q water, and then permitted to cool. The mixture was stirred for 2 h, and subsequently, 50.0% H2SO4 was added dropwise to the mixture, stirring constantly, till white silica formed at pH ~8.0. The white silica was rendered alkali-free after washing with hot Milli-Q H2O and then desiccation for 12 h at 70 °C. The desiccated silica was dried at 112 °C for 11 h (Figure 1a).

2.3. Graphene Oxide Preparation

The spent Zn-C batteries (dry cell battery wastes) were broken to attain waste carbon rods, which were milled to obtain a fine powder. Hot water was used to wash the carbon powder three times, and this was followed by washing with 30.0% sulfuric acid to eliminate inorganic impurities such as MnO2 and NH4Cl. The Hummer method was employed using treated carbon powder (graphite) to obtain graphene oxide [44,45]. Briefly, treated graphite powder (30 g) and NaNO3 (20 g) were added into a 3.0 L container. An acid mixture of 500 mL (98.0%) H2SO4 and 130 mL (85.0%) H3PO4 was combined inside a container with stirring. The compounds were frozen to 0–4 °C in an ice bath.
Thereafter, KMnO4 (36 g) was regularly added after 125 min with severe stirring and refrigeration to keep the mixture’s temperature beneath 15 °C for 33–125 min. Subsequently, the mixture warmed to 32 °C within a water bath for 11 h, where it converted to a sickly brownish-gray color. Then, 600 mL purified water was gradually added; moreover, the mixtures were constantly stirred for 24 min below 95 °C. Vast volumes of H2O (1.8 L) and 95 mL H2O2 (29%) were added into the earlier mix over 32 h, mixing to complete the reaction, which resulted in a bright yellow color. Subsequently, the mixture was treated with 85 mL (17%) HCl to eliminate any residual ions, centrifuged, and washed multiple times to eliminate HCl. Subsequently, a 32 min centrifugation (3100 rpm) was undertaken to eliminate any impurities. The sample was filtering under a vacuum, rinsed with deionized H2O many times, and placed to dry at 85 °C. Eventually, the gained GO was desiccated at 55 °C for 24 h (Figure 1b).

2.4. Preparation of Silica/GO

The GO was functionalized with a silanol unit of silica to gain silica/graphene oxide (GO/SiO2) by in situ hydrolysis [46]. For this purpose, 12 g of GO was dispersed in 85 mL water and 66 mL ethanol with 60 min sonication. Then, a 3.5 g SiO2 dissolving in 90 mL ammonia was added to the solution with constant stirring for 11 h. Afterwards, acetic acid and ammonia was used to adjust the pH of the mixture to 8.5; the mixture was kept for 3 h at 65 °C for complete condensation of the silica on GO. The SiO2/GO precipitate was washed many times by Milli-Q H2O. The SiO2/GO was heated to 82 °C in a furnace for 12 h (Figure 1c).

2.5. Adsorption Measures

A series of trials were performed to obtain the optimal parameters of the related conditions governing a U(VI) sorption method, such as pH of solution, equilibrium time, SiO2/GO quantity, U(VI) ions concentration, and reaction temperature. Group tests were completed in replicate. The sorption experiments were developed by incorporating 50 mL of various U(VI) concentrations, (10–100 mg) SiO2/GO dose at 220 rpm mechanical shaker, agitation time changing from 5–120 min, and (25–55 °C) temperature. The pH was varied from 1–6 and was adjusted by 1.0 M NaOH or 1.0 M H2SO4. The sorption experiments studied the sorption kinetics, equilibrium, and thermodynamic isotherms. In each sorption trial, U(VI) were put in SiO2/GO, and the uptake qe (mg/g) and distribution constant (Kd) were estimated using the subsequent equations:
q e = ( U o U e ) × V L m g
K d = ( U o U e U e ) × v m L m g
where Uo and Ue = initial and equilibrium U(VI) ions concentration (mg/L), respectively, VL= volume per L, vmL = volume per mL, and mg= SiO2/GO weight (g).

3. Results and Discussion

3.1. Descriptions

3.1.1. XRD Scrutiny

XRD analyses of SiO2, GO, SiO2/GO, and U/SiO2/GO are displayed in Figure 2. The SiO2 exhibited major broad peaks at 2θ = 19°, 22°, 24°, 28°, 29°, 32°, 34°, 36°, 39°, and 49°, which correspond to the database of Bruker software COD 9014256 (Figure 2a). The data of the SiO2 pattern revealed that the SiO2 was composed mainly of cristobalite [47]. As shown in Figure 2b, a broad peak at 2θ = 14° and two peaks with low intensity at 2θ = 20° and 42° were characterized of graphene oxide (GO), which correspond to the database of Bruker software COD 2000183 and 2002929.
XRD spectra of SiO2/GO and U/SiO2/GO are displayed in Figure 2c,d. In the XRD of SiO2 and GO (Figure 2a,b), a new peak appears at 2θ = 9°; the peak position and peak shape of GO with high intensity at 2θ = 14° were not altered, whereas the broad peaks of SiO2 overlap with the GO pattern. A new composite (SiO2/GO) was formed from these data due to the surface chemical and electrostatic interaction between SiO2 and GO. It has a broad peak at 2θ = 20°, as well as small two peaks, which appear at 2θ= 27° and 28°, according to the corresponding database of Bruker software COD 4025951 and 4124041 (Figure 2). In the XRD pattern of U/SiO2/GO (Figure 2d), some new peaks were observed after adsorption according to the database of Bruker software COD 8103695, 9000080, and 9009686. In contrast, the intensity of peaks was slightly changed, signifying that the SiO2/GO crystallinity did not imply alteration after U(VI) adsorption.

3.1.2. SEM Analysis

This analysis was established to check the reconstruction of exterior and natural configurations of SiO2, GO, SiO2/GO, and U/SiO2/GO (Figure 3). SEM of SiO2 and GO demonstrate definite shapes and include distinct irregular compositions. The particle sizes of SiO2 and GO may be 18 nm and 55 nm. However, the particle size of SiO2/GO may be 40 nm due to SiO2 being fixed on GO. The SiO2 and GO skeletons were established of small irregular bits with minute diameters, and they were not related to each other.
In contrast, the photograph of SiO2/GO shows smooth surfaces with several cavities due to the collection and affixation of SiO2 on the surface of GO. Additionally, the photograph of SiO2/GO was constructed of aggregate particles with bigger interstitial holes. After U(VI) sorption on SiO2/GO, the photograph shows that the holes were packed by U(VI). The SiO2/GO surface was irregular and agglomerated by U(VI), and the particle size of U/SiO2/GO may be 88 nm, which is larger than the other sizes. Additionally, the SEM results reveal that the SiO2/GO structure gave a more widespread surface area for U(VI) adsorption, which was better than that of GO or SiO2.

3.1.3. EDX Analysis

EDX spectra were used to conduct semi-quantitative analyses of SiO2, GO, and SiO2/GO, besides U/SiO2/GO (Figure 4). From the outcomes, Si and O bands were conferred in the SiO2 scale, and no extra bands were disclosed (Figure 4a), but the GO scale has C and O peaks (Figure 4b). The EDX analysis of SiO2/GO contained C, O, and Si bands (Figure 4c). These results confirmed that the Si, C, and O bands happened due to SiO2/GO formation. Ultimate U(VI) sorption above SiO2/GO only discerned plain U(VI) bands that admitted and validated U(VI) sorption above SiO2/GO (Figure 4d).

3.1.4. BET Analysis

The BET is operated to assess solid materials’ surface area and pore size. Figure 5a–d demonstrated the N2 sorption–desorption isotherms of SiO2, GO, SiO2/GO, and U/SiO2/GO. The obtained data in Table 1 showed that the surface areas of SiO2, GO, SiO2/GO, and U/SiO2/GO were 25.89, 37.37, 35.45, and 32.15 m2/g, respectively. The surface area (SBET) and pore volume, besides the pore size of SiO2/GO, were measured between the consistent SiO2 and GO values, and it might be considered that SiO2 decorated GO pores. The SiO2 surface area and porosity improved the GO surface to pick up uranium ions. From the data, the surface area, pore size, along with pore volume of SiO2/GO were reduced after U(VI) sorption due to pore-blocking with ions. The results showed that U(VI) ions were powerfully captured on SiO2/GO due to more active sites.

3.1.5. FTIR Analysis

The FTIR spectra of SiO2, GO, SiO2/GO, and U/SiO2/GO were determined at 4000–400 cm1, as characterized in Figure 6. Regarding the SiO2 in Figure 6a, the extensive band at 3450 cm−1 offered O–H of Si–OH, but a distinctive band at 1646 cm−1 of reacted H2O on SiO2 was distinguished, which was not wholly displaced via drying [48]. The influential bands at 1216 and 1088 cm−1 were determined through a siloxane group (Si–O–Si) [47,49,50]. The attendance peak at 956 cm−1 was also accepted in the silanol group (Si–OH). The absorption peak at 795 cm−1 was documented in the Si–O–Si group, and the detected peak was 466 cm−1 in the Si–O–Si group [51,52].
Moreover, Figure 6b of the prepared GO demonstrates a deep band at 3432 cm−1, consistent with the O–H group. The vibrational band at 1646 cm−1 expressed C=C or OH of H2O sorbed upon GO. An assignment at 1704 cm−1 was declared to C=O of COOH groups. The slight vibrations at 1183 and 957 cm−1 correspond to epoxy groups (C–O), and the 1102 cm−1 is in agreement with C–O of COOH groups [53].
The spectrum of SiO2/GO (Figure 6c) showed an abroad peak at 3398 cm−1 that belonged to –OH groups of SiO2 and GO. The assignments at 3112 cm−1 match –CH aromatic, and features at 1676 and 1593 cm−1 are consistent with C=O and C=C of GO. Additionally, the peaks at 1246, 1145, and 1093 cm−1 were concerned with the siloxane groups. Moreover, a Si–OH peak at 943 cm−1 was documented, and the vibrational band at 788 cm−1 corresponds to the Si–O–Si group. From these data, it was confirmed that SiO2/GO was formed.
After U(VI) sorption in the U/SiO2/GO spectrum (Figure 6d), the –OH, epoxy (–O–), and C=O stretching vibration bands were condensed and relocated to redshift with 4–11 cm−1, probably because of U(VI) adsorbed on SiO2/GO. Moreover, fresh peaks consistent with of O=U=O are found near 974 and 894 cm−1 [54]. Moreover, two weak peaks appeared near ≈465 and ≈425 cm−1 due to the U–O group [55]. Accordingly, it realized that the SiO2/GO acted more friendly to U(VI) sorption.

3.1.6. Raman Analysis

Raman spectra of SiO2, GO, and SiO2/GO are illustrated in Figure 7(a–c). The spectrum of the prepared SiO2 from rice husk waste is given in Figure 2a, the distinctive bands of the SiO2 network, D1 and D2, can be observed, which are associated with the silica material, besides the regular silica band observed at 985 cm−1, which is assigned to the SiOH groups [56]. The Raman spectrum of the prepared graphene oxide from rod batteries wastes is depicted in Figure 2b. The G band is noticed at 1585 cm−1, whereas the D peak is also observed at 1355 cm−1. The G band is standard for all sp2 carbon forms, and it appears from the C-C bond stretch. This band is assembled from first-order Raman scattering [57,58]. The G band in GO is due to the presence of oxygen atoms in the GO structure, forming from sp2 ordered crystalline graphite-like configurations. The D band in GO is broadened due to a disordered sp3 carbon structure.
The Raman spectrum of the prepared SiO2/GO is shown in Figure 7c. It detected the characteristic bands of SiO2 and GO. It was observed that the bands of the SiO2 network wrre D1, D2, and SiOH groups for SiO2, whereas the D and G bands were attributed to GO. All bands of SiO2/GO are minor when compared with the bands of its constituents (SiO2 and GO). The relative intensity of the D and G peaks offers crucial information concerning the chemical environment. The intensity of both D and G bands display a significant alteration without changing the position or shape when GO reacts with silica particles that acquire covalent attached to the GO [59,60,61]. However, in the SiO2/GO spectrum (Figure 7c), the D and G bands’ intensity ratio was small, demonstrating a noncovalent interaction between silica and GO [62,63]. Hence, it is quite evident that silica and GO accumulated via the random packing procedure according to the van der Waal potential.

3.2. Sorption Data

U(VI) sorption was intended on SiO2/GO from the uranium sulfate complex solution. Some testing experiments were directed on U(VI) sorption to study the impact of pH, SiO2/GO dose, contact time, initial U(VI) concentration, and sorption temperature.

3.2.1. pH Influence

pH of solution is an important factor affecting U(VI) extraction on the SiO2/GO sorbent. The pH result at SiO2/GO was intended with several pH values from 1.0–6.0 at permanent settings of 50 mg SiO2/GO and the 50 mL assaying 150 mg/L U(VI), 40 min sorption time, and 25 °C (Figure 8a). The U(VI) uptake increased to 115 mg/g through an increasing pH of 4.0. Successively, as pH increased to 6.0, U(VI) uptake dropped to 40 mg/g. As a consequence, the maximum uptake was attained at pH 4.0.
The graphene oxide active sites are –OH, –COOH, and –O–; however, SiO2/GO active sites are increased to improve uptake. The active sites increased in the SiO2/GO due to existing –OH, –COOH, –O–, and silanol (Si–OH) groups on the SiO2/GO surface. At high acidity (pH < 4.0), the sorption sites of SiO2/GO take positive charges. However, the electrostatic attraction between the positive active sites of SiO2/GO and anionic species [UO2(SO4)3]4− and [UO2(SO4)2]2− were carried out due to a great volume of HSO4 that easily reacted with positive active sites. Some neutral complexes of uranium ions were formed by reacting some H+ ions with uranium anionic species. Additionally, HSO4 ions competed with uranium anionic species during sorption. Hence, U(VI) sorption was decreased. At pH 4.0, U(VI) was observed in a hydrolyzed appearance, and the resulting cationic varieties recognized UO22+, [(UO2)2(OH)2]2+ dimmer, along with trimmer [(UO2)3(OH)5]+ [64]. Oxygen atoms of SiO2/GO deprotonated. Multiple uranium hydrolysis outputs formed were exceedingly adsorbed by electrostatic magnetism and complexation operations. Hence, the uranium uptake realized maximum sorption at pH 4.0. Alternatively, at pH ˃ 4.0, uptake progressively decreased, owing to the precipitation of uranyl hydroxide species [65]. For that reason, it was not necessary to examine the effect of pH 6.0.

3.2.2. Dose Influence

The effect of SiO2/GO dose on uranium sorption was measured using a variety of doses from 10 to 100 mg, at 4.0 pH, 150 mg/L of U(VI), and 40 min sorption time (Figure 8b). The results informed that U(VI) sorption improved by increasing the SiO2/GO dose. Due to increasing the active sites, more active sites were available for U(VI) sorption at higher dosages. Henceforth, U(VI) sorption improved to 96.66% by increasing the SiO2/GO doses to 50 mg. Nevertheless, the sorption percent prevailed constantly until the 100 mg dose. Hence, the 50 mg SiO2/GO dose was chosen for sorption tests.

3.2.3. Sorption Time and Kinetics

The role of contact time on U(VI) uptake was applied on SiO2/GO at an altered contact time (5–120 min) at pH 4.0, 150 mg/L of U(VI), 50 mg of SiO2/GO dose (Figure 9a). The data showed that the uranium uptake was increased from 62–145 mg/g by increasing contact time from 5–50 min. Instantaneously varying sorption time to 120 min should have no noticeable impact on U(VI) uptake. Henceforth, a contact time of 50 min was an appropriate supplementary condition of uptake.
Pseudo-first-order as well as pseudo-second-order model was maintained for the good data to estimate U(VI) sorption kinetics on SiO2/GO sorbent. The nonlinear equation of the two forms is as follows [66]:
q t = q e ( 1 e k 1 t )
q t = k 2 q e 2 t 1 + k 2 q e t
where qe and qt (mg/g) are U(VI) amounts sorbed at equilibrium and at sorption time t (min), and k1 (1/min) refers to the first-order constant, while k2 (g/mg.min) refers to the second-order constant. q1, k1, and R2 were 124 mg/g, 0.115 1/min, and 0.89, respectively, based on the nonlinear regression of first-order rate in Figure 9b, as well as the reduced Chi-square was 67.72. Nonetheless, for the nonlinear regression of second-order, q2, k2, and R2 were 150 mg/g, 9.67 g/mg.min, and 0.96, respectively, with a reduced Chi-square of 18.88. From the data in Figure 9b, the R2 for the second-order model was close to unity, and q2 = 150 mg/g was closer to experimental uptake, qe (145 mg/g). Therefore, it suggested that the second-order model described U(VI) sorption kinetics upon SiO2/GO. Hence, U(VI) sorption by SiO2/GO signified that it was most likely controlled by the chemisorption process [67].

3.2.4. Initial U(VI) Concentration and Equilibrium Isotherms

Many experiments were conducted to determine the effect of U(VI) concentration and its uptake on SiO2/GO. These tests were performed by stirring 50 mL of various U(VI) concentrations vary between 25–600 mg/L at pH 4.0, 50 mg of SiO2/GO dose, and 50 min sorption time. Figure 10a shows that as initial U(VI) concentration increased, the U(VI) uptake augmented to reach a maximum uptake of 145 mg/g at 150 mg/L; after that, the uptake remained constant.
Adsorption isotherms are critical for understanding the specifics of adsorption conduct, SiO2/GO sorbent surface properties, and further factors that affect sorption procedures. The sorption experiments were carried out at the optimum conditions to determine the respectable well-matched isotherm. Nonlinear models of Langmuir and Freundlich engaged in regulating the convenient model and competently related the practical information. Nonlinear most minor square optimization performance was employed, and isotherm factors were estimated with the assistance of nonlinear fit software. The reduced Chi-square was operated as a condition for the fitting advantage. The squared alteration between the experimental and calculated data was divided into the calculated data gained from the Langmuir alongside Freundlich models [68]. The reduced Chi-square was signified by:
x 2 = ( q e ( exp ) q e ( m o d ) ) 2 q e ( m o d )
where qe(exp) and qe(mod) (mg/g) are the uptakes of experimental and model calculation. x2 has reduced Chi-square. The assessment of x2 was utilized to assess the fit of the best isotherm to experimental data, where the smaller the reduced Chi-square, the better the isotherm fit. Permitting to Langmuir model sorption employed at homogeneous sites inside the SiO2/GO adsorbent and formerly U(VI) occupied at sites. The nonlinear Langmuir isotherm was determined using the following equation:
q e = q m a x k L C e 1 + k L C e
where qe (mg/g) is the sorbed U(VI) amount on SiO2/GO adsorbent, Ce (mg/L) is the equilibrium U(VI) concentration, qmax (mg/g) is the maximum sorbed U(VI) amount on SiO2/GO adsorbent, and kL (L/mg) is defined as the Langmuir sorption constant.
The empirical nonlinear Freundlich model according to sorption via heterogeneous surface was set as:
q e = k f C e 1 / n
where kf and n symbolize for the uptake and sorption intensity, respectively. From 10b shows that the Langmuir model’s sorption uptake (148.5 mg/g) is relevant to the experimental uptake (145.0 mg/g), and its R2 (0.962) is close to unity when compared to the Freundlich model’s kf (79.71 mg/g) and R2 (0.691). Additionally, the Langmuir’s Chi-square is smaller than Freundlich’s Chi-square. As a result, Langmuir’s model is the most accurate for the experimental data.

3.2.5. Temperature and Thermodynamic Investigations

The temperature impact on U(VI) uptake on SiO2/GO sorbent was measured at 25–55 °C at pH 4.0, 50 g of SiO2/GO dose, 50 min contact time, and 150 mg/L of U(VI). The data in Figure 11a show that as the sorption temperature rose to 55 °C, the uptake decreased to 142.20 mg/g. As a result, it was probably the increase in temperature that led to the decomposition of the SiO2/GO sorbent. Additionally, it may be the decrease in SiO2/GO active sites that started the diminishing uptake.
The sorption mechanism was assessed through thermodynamic calculations for U(VI) on SiO2/GO sorbent, and they were assessed from the Van’t Hoff calculations [69,70]:
l o g K d = Δ S ° 2.303 R Δ H ° 2.303 R T
Δ G ° = Δ H ° T Δ S °
where Kd (L/g) refers to the distribution constant, ΔG° (kJ/mol) symbolizes for the sorption free energy, ΔH° (kJ/mol) stands for the enthalpy sorption changes, ΔS° represents the entropy sorption changes (J/mol.K), R is 8.314 J/mol.K, and T expresses the Kelvin temperature (K). Thermodynamic considerations of SiO2/GO sorbent are shown in Figure 11b and Table 2. The negative ΔG° values stated that the adsorption of (VI) on SiO2/GO was achievable spontaneously. The sorption progression also formed U(VI) electrostatic interaction with SiO2/GO. The negative Δ value suggested that the sorption method was exothermic. Correspondingly, the negative Δ value evidenced the feasibility and randomness of the sorption.

3.3. Desorption Investigation

3.3.1. Eluting Type

Eluting HNO3, HCl, H2SO4, NaCl, NaNO3, and Na2SO4 agents were performed to acquire the U(VI) desorption from U/SiO2/GO. Still, the other desorption factors were kept constant at a 1.0 M eluting concentration, 60 min eluting time, and 1:30 (S:L) phase ratio. As presented in Figure 12a, U(VI) elution from U/SiO2/GO using 1.0 M H2SO4 was extended to the maximum elution (88.0%). Consequently, H2SO4 is the best eluting agent.

3.3.2. Acid Concentration

U(VI) desorption from U/SiO2/GO was studied. Many concentrations of H2SO4 ranging from 0.2–1.2 M were employed to detect the best acid concentration for desorption, and other parameters were constant at 1:30 S:L ratio, 60 min desorption time, and 25 °C. The acquired data in Figure 12b appeared to improve U(VI) desorption by increasing the eluting concentration until realizing the highest value (85.0%) at 0.8 M H2SO4. Subsequently, 0.8 M of H2SO4 was suitable for eluting of U(VI).

3.3.3. Solid: Liquid Ratio

The S:L ratio’s outcome was inspected by a difference of the S:L phase range from 1:10–1:70, while other parameters were settled at 0.8 M acid for 60 min contact time (Figure 12c). The comprehensive data indicate that U(VI) desorption was continuously improved through increasing the S:L phase ratio until 89.0% at 1:50 S:L and constantly persisted from 1:50 to 1:70 S:L phase ratio. Hence, the S:L phase ratio of 1:50 optimized U(VI) desorption.

3.3.4. Eluting Time

Eluting time affects U(VI) desorption or elution from U/SiO2/GO, and it was examined many times from 10 to 90 min. The other parameters constantly persist at 0.8 M H2SO4, 1:50 S:L phase ratio. In the data in Figure 12d, desorption was improved from 75.0 to 99.0% by increasing the agitation time from 10 to 60 min, but the desorption constantly persisted after 60–90 min. Therefore, 60 min was the best eluting time.

3.4. Regeneration

Regeneration remains the most significant advantage in the re-reusability of SiO2/GO adsorbent. The SiO2/GO adsorbent was refreshed using 0.8 M H2SO4, 1:50 ratio of S:L, and 60 min for recycling (Figure 13). The sorption–desorption behaviors were recurrent until desorption dropped to 80.0% after the seventh cycle for SiO2/GO. Hence, this validated the use of SiO2/GO for U(VI) sorption.

4. Conclusions

Silica/GO was prepared from the environmental solid wastes and employed for U(VI) sorption. SiO2/GO adsorbent was produced from graphene oxide obtained from spent carbon rod batteries and silica was obtained from rice husk to eliminate U(VI) ions from their solutions. The best sorption parameters were pH of 4.0, 150 mg/L U(VI), 50 mg SiO2/GO dosage, and 50 min of equilibrium time. The recognized superior uptake was at 145.0 mg/g. In addition, kinetic results were stated to suit the second-order kinetic model.
Furthermore, the Langmuir isotherm model was well-matched for labeling adsorption developments. The negative ∆S° and negative ∆H° values suggested that the thermodynamic calculations recognized U(VI) sorption randomness and exothermic nature. Moreover, the negative ∆G° values pointed out to spontaneous sorption. The U(VI) ions were also stripped from U/SiO2/GO by 0.8 M H2SO4, a 1:50 ratio of S:L, and 60 min of stripping time. The sorption–desorption developments were frequently iterated till desorption was reduced to 80.0% seven recurring times. Finally, SiO2/GO adsorbent permitted a good U(VI) adsorption from aqueous sources.

Author Contributions

Conceptualization, M.A.H., S.H.N. and A.K.S.; methodology, M.A.H., M.A.Y. and A.K.S.; software, A.K.S., M.Y.H. and M.F.C.; validation, M.I.S. and M.F.C.; formal analysis, A.K.S. and M.F.C.; investigation, A.K.S. and M.F.C.; resources, A.K.S. and M.F.C.; data curation, A.K.S., M.Y.H. and M.F.C.; writing—original draft preparation, A.K.S. and M.F.C.; writing—review and editing, A.K.S., M.F.C. and M.Y.H.; visualization, A.K.S., M.F.C. and T.F.M.; supervision, H.I.M., T.F.M., M.F.C. and M.I.S.; project administration, H.I.M., T.F.M. and M.F.C.; funding acquisition, J.S.A.-O. All authors have read and agreed to the published version of the manuscript.

Funding

The authors express their gratitude to Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R13), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors express their gratitude to Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R13), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cheira, M.F.; Atia, B.M.; Kouraim, M.N. Uranium(VI) recovery from acidic leach liquor by Ambersep 920U SO4 resin: Kinetic, equilibrium and thermodynamic studies. J. Radiat. Res. Appl. Sci. 2017, 10, 307–319. [Google Scholar] [CrossRef] [Green Version]
  2. Chen, C.; Wang, J. Uranium removal by novel graphene oxide-immobilized Saccharomyces cerevisiae gel beads. J. Environ. Radioact. 2016, 162–163, 134–145. [Google Scholar] [CrossRef] [PubMed]
  3. Atia, B.M.; Gado, M.A.; Cheira, M.F. Kinetics of uranium and iron dissolution by sulfuric acid from Abu Zeneima ferruginous siltstone, Southwestern Sinai, Egypt. Euro-Mediterr. J. Environ. Integr. 2018, 3, 39. [Google Scholar] [CrossRef]
  4. Sakr, A.; Mohamed, S.; Mira, H.; Cheira, M. Successive leaching of uranium and rare earth elements from El Sela mineralization. J. Sci. Eng. Res. 2018, 5, 95–111. [Google Scholar]
  5. Yao, W.; Wang, X.; Liang, Y.; Yu, S.; Gu, P.; Sun, Y.; Xu, C.; Chen, J.; Hayat, T.; Alsaedi, A.; et al. Synthesis of novel flower-like layered double oxides/carbon dots nanocomposites for U(VI) and 241Am(III) efficient removal: Batch and EXAFS studies. Chem. Eng. J. 2018, 332, 775–786. [Google Scholar] [CrossRef]
  6. Liu, X.; Xu, X.; Sun, J.; Alsaedi, A.; Hayat, T.; Li, J.; Wang, X. Insight into the impact of interaction between attapulgite and graphene oxide on the adsorption of U(VI). Chem. Eng. J. 2018, 343, 217–224. [Google Scholar] [CrossRef]
  7. Dai, Z.; Zhang, H.; Sui, Y.; Ding, D.; Li, L. Preparation of polyamidoxime/magnetic graphene oxide composite and Its application for efficient extraction of uranium(VI) from aqueous solutions in an ultrasonic field. J. Chem. Eng. Data 2018, 63, 4215–4225. [Google Scholar] [CrossRef]
  8. Cheira, M.F.; El-Didamony, A.M.; Mahmoud, K.F.; Atia, B.M. Equilibrium and kinetic characteristics of uranium recovery by the strong base Ambersep 920U Cl resin. IOSR J. Appl. Chem 2014, 7, 32–40. [Google Scholar] [CrossRef]
  9. Zidan, I.H.; Cheira, M.F.; Bakry, A.R.; Atia, B.M. Potentiality of uranium recovery from G.Gattar leach liquor using Duolite ES-467 chelating resin: Kinetic, thermodynamic and isotherm features. Int. J. Environ. Anal. Chem. 2020, 100, 1–23. [Google Scholar] [CrossRef]
  10. Cheira, M.F. Synthesis of pyridylazo resorcinol–functionalized Amberlite XAD-16 and its characteristics for uranium recovery. J. Environ. Chem. Eng. 2015, 3, 642–652. [Google Scholar] [CrossRef]
  11. Gado, M.A.; Atia, B.M.; Cheira, M.F.; Abdou, A.A. Thorium ions adsorption from aqueous solution by amino naphthol sulphonate coupled chitosan. Int. J. Environ. Anal. Chem. 2021, 101, 1419–1436. [Google Scholar] [CrossRef]
  12. Atia, B.M.; Gado, M.A.; Cheira, M.F.; El-Gendy, H.S.; Yousef, M.A.; Hashem, M.D. Direct synthesis of a chelating carboxamide derivative and its application for thorium extraction from Abu Rusheid ore sample, South Eastern Desert, Egypt. Int. J. Environ. Anal. Chem. 2021, 101, 1–24. [Google Scholar] [CrossRef]
  13. Jo, Y.; Lee, J.-Y.; Yun, J.-I. Adsorption of uranyl tricarbonate and calcium uranyl carbonate onto γ-alumina. Appl. Geochem. 2018, 94, 28–34. [Google Scholar] [CrossRef]
  14. Cheira, M.F. Solvent extraction of uranium and vanadium from carbonate leach solutions of ferruginous siltstone using cetylpyridinium carbonate in kerosene. Chem. Pap. 2020, 74, 2247–2266. [Google Scholar] [CrossRef]
  15. Zhou, C.; Ontiveros-Valencia, A.; de Cornette Saint Cyr, L.; Zevin, A.S.; Carey, S.E.; Krajmalnik-Brown, R.; Rittmann, B.E. Uranium removal and microbial community in a H2-based membrane biofilm reactor. Water Res. 2014, 64, 255–264. [Google Scholar] [CrossRef] [PubMed]
  16. Cheira, M.F. Characteristics of uranium recovery from phosphoric acid by an aminophosphonic resin and application to wet process phosphoric acid. Eur. J. Chem. 2015, 6, 48–56. [Google Scholar] [CrossRef] [Green Version]
  17. Abdien, H.G.; Cheira, M.; Abd-Elraheem, M.; El-Naser, T.A.S.; Zidan, I.H. Extraction and pre-concentration of uranium using activated carbon impregnated trioctyl phosphine oxide. Elix Appl. Chem. 2016, 100, 3462–43469. [Google Scholar]
  18. Cheira, M.F. Synthesis of aminophosphonate-functionalised ZnO/polystyrene-butadiene nanocomposite and its characteristics for uranium adsorption from phosphoric acid. Int. J. Environ. Anal. Chem. 2021, 101, 1710–1734. [Google Scholar] [CrossRef]
  19. Jiao, X.; Zhang, L.; Qiu, Y.; Guan, J. Comparison of the adsorption of cationic blue onto graphene oxides prepared from natural graphites with different graphitization degrees. Colloids Surf. A Physicochem. Eng. Asp. 2017, 529, 292–301. [Google Scholar] [CrossRef]
  20. Cheira, M.F.; Kouraim, M.N.; Zidan, I.H.; Mohamed, W.S.; Hassanein, T.F. Adsorption of U(VI) from sulfate solution using montmorillonite/polyamide and nano-titanium oxide/polyamide nanocomposites. J. Environ. Chem. Eng. 2020, 8, 104427. [Google Scholar] [CrossRef]
  21. Atia, B.M.; Khawassek, Y.M.; Hussein, G.M.; Gado, M.A.; El-Sheify, M.A.; Cheira, M.F. One-pot synthesis of pyridine dicarboxamide derivative and its application for uranium separation from acidic medium. J. Environ. Chem. Eng. 2021, 9, 105726. [Google Scholar] [CrossRef]
  22. Cheira, M.F.; Rashed, M.N.; Mohamed, A.E.; Hussein, G.M.; Awadallah, M.A. Removal of some harmful metal ions from wet-process phosphoric acid using murexide-reinforced activated bentonite. Mater. Today Chem. 2019, 14, 100176. [Google Scholar] [CrossRef]
  23. Zong, P.; Wang, S.; Zhao, Y.; Wang, H.; Pan, H.; He, C. Synthesis and application of magnetic graphene/iron oxides composite for the removal of U(VI) from aqueous solutions. Chem. Eng. J. 2013, 220, 45–52. [Google Scholar] [CrossRef]
  24. Hassanin, M.A.; El-Gendy, H.S.; Cheira, M.F.; Atia, B.M. Uranium ions extraction from the waste solution by thiosemicarbazide anchored cellulose acetate. Int. J. Environ. Anal. Chem. 2021, 101, 351–369. [Google Scholar] [CrossRef]
  25. Gomaa, H.; Cheira, M.F.; Abd-Elraheem, M.A.; Seaf El-Naser, T.A.; Zidan, I.H. Removal of uranium from acidic solution using activated carbon impregnated with tri butyl phosphate. Biol. Chem. Res. 2016, 2016, 313–340. Available online: http://www.ss-pub.org/wp-content/uploads/2016/11/BCR2016081701.pdf (accessed on 12 January 2022).
  26. Li, Z.-J.; Wang, L.; Yuan, L.-Y.; Xiao, C.-L.; Mei, L.; Zheng, L.-R.; Zhang, J.; Yang, J.-H.; Zhao, Y.-L.; Zhu, Z.-T.; et al. Efficient removal of uranium from aqueous solution by zero-valent iron nanoparticle and its graphene composite. J. Hazard. Mater. 2015, 290, 26–33. [Google Scholar] [CrossRef]
  27. Allam, E.M.; Lashen, T.A.; Abou El-Enein, S.A.; Hassanin, M.A.; Sakr, A.K.; Cheira, M.F.; Almuqrin, A.; Hanfi, M.Y.; Sayyed, M.I. Rare Earth Group Separation after Extraction Using Sodium Diethyldithiocarbamate/Polyvinyl Chloride from Lamprophyre Dykes Leachate. Materials 2022, 15, 1211. [Google Scholar] [CrossRef]
  28. Salazar-Hernández, M.M.; Salazar-Hernández, C.; Elorza-Rodríguez, E.; Juárez Ríos, H. The use of mesoporous silica in the removal of Cu(I) from the cyanidation process. J. Mater. Sci. 2015, 50, 439–446. [Google Scholar] [CrossRef]
  29. Gomaa, H.; Shenashen, M.A.; Elbaz, A.; Kawada, S.; Seaf El-Nasr, T.A.; Cheira, M.F.; Eid, A.I.; El-Safty, S.A. Inorganic-organic mesoporous hybrid segregators for selective and sensitive extraction of precious elements from urban mining. J. Colloid Interface Sci. 2021, 604, 61–79. [Google Scholar] [CrossRef]
  30. Verma, S.; Kim, K.-H. Graphene-based materials for the adsorptive removal of uranium in aqueous solutions. Environ. Int. 2022, 158, 106944. [Google Scholar] [CrossRef]
  31. Li, X.; Wang, Z.; Li, Q.; Ma, J.; Zhu, M. Preparation, characterization, and application of mesoporous silica-grafted graphene oxide for highly selective lead adsorption. Chem. Eng. J. 2015, 273, 630–637. [Google Scholar] [CrossRef]
  32. Yadav, A.A.; Hunge, Y.M.; Kang, S.-W. Porous nanoplate-like tungsten trioxide/reduced graphene oxide catalyst for sonocatalytic degradation and photocatalytic hydrogen production. Surf. Interfaces 2021, 24, 101075. [Google Scholar] [CrossRef]
  33. Choudhari, A.; Bhanvase, B.A.; Saharan, V.K.; Salame, P.H.; Hunge, Y. Sonochemical preparation and characterization of rGO/SnO2 nanocomposite: Electrochemical and gas sensing performance. Ceram. Int. 2020, 46, 11290–11296. [Google Scholar] [CrossRef]
  34. Yadav, A.A.; Hunge, Y.M.; Kang, S.-W. Spongy ball-like copper oxide nanostructure modified by reduced graphene oxide for enhanced photocatalytic hydrogen production. Mater. Res. Bull. 2021, 133, 111026. [Google Scholar] [CrossRef]
  35. Yang, A.; Zhu, Y.; Huang, C.P. Facile preparation and adsorption performance of graphene oxide-manganese oxide composite for uranium. Sci. Rep. 2018, 8, 9058. [Google Scholar] [CrossRef] [Green Version]
  36. Ding, C.; Cheng, W.; Nie, X.; Yi, F. Synergistic mechanism of U(VI) sequestration by magnetite-graphene oxide composites: Evidence from spectroscopic and theoretical calculation. Chem. Eng. J. 2017, 324, 113–121. [Google Scholar] [CrossRef]
  37. Zhang, Q.; Zhao, D.; Ding, Y.; Chen, Y.; Li, F.; Alsaedi, A.; Hayat, T.; Chen, C. Synthesis of Fe–Ni/graphene oxide composite and its highly efficient removal of uranium(VI) from aqueous solution. J. Clean. Prod. 2019, 230, 1305–1315. [Google Scholar] [CrossRef]
  38. Fan, F.-L.; Qin, Z.; Bai, J.; Rong, W.-D.; Fan, F.-Y.; Tian, W.; Wu, X.-L.; Wang, Y.; Zhao, L. Rapid removal of uranium from aqueous solutions using magnetic Fe3O4@SiO2 composite particles. J. Environ. Radioact. 2012, 106, 40–46. [Google Scholar] [CrossRef]
  39. Aljarrah, M.T.; Al-Harahsheh, M.S.; Mayyas, M.; Alrebaki, M. In situ synthesis of quaternary ammonium on silica-coated magnetic nanoparticles and it’s application for the removal of uranium (VI) from aqueous media. J. Environ. Chem. Eng. 2018, 6, 5662–5669. [Google Scholar] [CrossRef]
  40. Liu, H.; Xie, S.; Liao, J.; Yan, T.; Liu, Y.; Tang, X. Novel graphene oxide/bentonite composite for uranium(VI) adsorption from aqueous solution. J. Radioanal. Nucl. Chem. 2018, 317, 1349–1360. [Google Scholar] [CrossRef]
  41. Hamza, M.F.; Fouda, A.; Elwakeel, K.Z.; Wei, Y.; Guibal, E.; Hamad, N.A. Phosphorylation of Guar Gum/Magnetite/Chitosan Nanocomposites for Uranium (VI) Sorption and Antibacterial Applications. Molecules 2021, 26, 1920. [Google Scholar] [CrossRef] [PubMed]
  42. Setyawan, N.; Yuliani, S. Synthesis of silica from rice husk by sol-gel method. IOP Conf. Ser. Earth Environ. Sci. 2021, 733, 012149. [Google Scholar] [CrossRef]
  43. Le, V.H.; Thuc, C.N.H.; Thuc, H.H. Synthesis of silica nanoparticles from Vietnamese rice husk by sol–gel method. Nanoscale Res. Lett. 2013, 8, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zaaba, N.I.; Foo, K.L.; Hashim, U.; Tan, S.J.; Liu, W.-W.; Voon, C.H. Synthesis of Graphene Oxide using Modified Hummers Method: Solvent Influence. Procedia Eng. 2017, 184, 469–477. [Google Scholar] [CrossRef]
  45. Abd El-Rahem, K.A.; Taher, M.A.; Cheira, M.F. Uranium ions removal from aqueous media using graphene oxide derived from black carbon. Int. J. Adv. Res. Rev. 2021, 6, 1–17. [Google Scholar]
  46. Dehghani, Z.; Sedghi-Asl, M.; Ghaedi, M.; Sabzehmeidani, M.M.; Adhami, E. Ultrasound-assisted adsorption of paraquat herbicide from aqueous solution by graphene oxide/ mesoporous silica. J. Environ. Chem. Eng. 2021, 9, 105043. [Google Scholar] [CrossRef]
  47. Gholami, T.; Salavati-Niasari, M.; Bazarganipour, M.; Noori, E. Synthesis and characterization of spherical silica nanoparticles by modified Stöber process assisted by organic ligand. Superlattices Microstruct. 2013, 61, 33–41. [Google Scholar] [CrossRef]
  48. Cheira, M.F.; Mira, H.I.; Sakr, A.K.; Mohamed, S.A. Adsorption of U(VI) from acid solution on a low-cost sorbent: Equilibrium, kinetic, and thermodynamic assessments. Nucl. Sci. Tech. 2019, 30, 156. [Google Scholar] [CrossRef]
  49. Sankar, S.; Sharma, S.K.; Kaur, N.; Lee, B.; Kim, D.Y.; Lee, S.; Jung, H. Biogenerated silica nanoparticles synthesized from sticky, red, and brown rice husk ashes by a chemical method. Ceram. Int. 2016, 42, 4875–4885. [Google Scholar] [CrossRef]
  50. Cheira, M.F. Performance of poly sulfonamide/nano-silica composite for adsorption of thorium ions from sulfate solution. SN Appl. Sci. 2020, 2, 398. [Google Scholar] [CrossRef] [Green Version]
  51. Sakr, A.K.; Cheira, M.F.; Hassanin, M.A.; Mira, H.I.; Mohamed, S.A.; Khandaker, M.U.; Osman, H.; Eed, E.M.; Sayyed, M.I.; Hanfi, M.Y. Adsorption of Yttrium Ions on 3-Amino-5-Hydroxypyrazole Impregnated Bleaching Clay, a Novel Sorbent Material. Appl. Sci. 2021, 11, 10320. [Google Scholar] [CrossRef]
  52. Lu, P.; Hsieh, Y.-L. Highly pure amorphous silica nano-disks from rice straw. Powder Technol. 2012, 225, 149–155. [Google Scholar] [CrossRef]
  53. Faniyi, I.O.; Fasakin, O.; Olofinjana, B.; Adekunle, A.S.; Oluwasusi, T.V.; Eleruja, M.A.; Ajayi, E.O.B. The comparative analyses of reduced graphene oxide (RGO) prepared via green, mild and chemical approaches. SN Appl. Sci. 2019, 1, 1181. [Google Scholar] [CrossRef] [Green Version]
  54. Sureshkumar, M.K.; Das, D.; Mallia, M.B.; Gupta, P.C. Adsorption of uranium from aqueous solution using chitosan-tripolyphosphate (CTPP) beads. J. Hazard. Mater. 2010, 184, 65–72. [Google Scholar] [CrossRef]
  55. Wang, G.; Liu, J.; Wang, X.; Xie, Z.; Deng, N. Adsorption of uranium (VI) from aqueous solution onto cross-linked chitosan. J. Hazard. Mater. 2009, 168, 1053–1058. [Google Scholar] [CrossRef]
  56. Alessi, A.; Agnello, S.; Buscarino, G.; Gelardi, F.M. Raman and IR investigation of silica nanoparticles structure. J. Non-Cryst. Solids 2013, 362, 20–24. [Google Scholar] [CrossRef]
  57. Perumbilavil, S.; Sankar, P.; Priya Rose, T.; Philip, R. White light Z-scan measurements of ultrafast optical nonlinearity in reduced graphene oxide nanosheets in the 400–700 nm region. Appl. Phys. Lett. 2015, 107, 051104. [Google Scholar] [CrossRef]
  58. Bahadur, J.; Das, A.; Prakash, J.; Singh, P.; Khan, A.; Sen, D. Role of trapped water on electroresponsive characteristic of silica-graphene oxide composite microspheres. J. Appl. Phys. 2019, 126, 204301. [Google Scholar] [CrossRef]
  59. Kou, L.; Gao, C. Making silica nanoparticle-covered graphene oxide nanohybrids as general building blocks for large-area superhydrophilic coatings. Nanoscale 2011, 3, 519–528. [Google Scholar] [CrossRef]
  60. Innocenzi, P.; Malfatti, L.; Lasio, B.; Pinna, A.; Loche, D.; Casula, M.F.; Alzari, V.; Mariani, A. Sol–gel chemistry for graphene–silica nanocomposite films. New J. Chem. 2014, 38, 3777–3782. [Google Scholar] [CrossRef]
  61. Park, K.-W.; Kwon, O.-Y. Preparation of Novel Mesoporous Silica Using a Self-Assembled Graphene Oxide Template. Sci. Rep. 2020, 10, 6173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Carboni, D.; Lasio, B.; Alzari, V.; Mariani, A.; Loche, D.; Casula, M.F.; Malfatti, L.; Innocenzi, P. Graphene-mediated surface enhanced Raman scattering in silica mesoporous nanocomposite films. Phys. Chem. Chem. Phys. 2014, 16, 25809–25818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Mo, J.; Ma, W.; Qiu, G.; Shi, Y. Mesoporous silica coated graphene oxide: Fabrication, characterization and effects on the dielectric properties of its organosilicon hybrid films. J. Mater. Sci. Mater. Electron. 2019, 30, 130–146. [Google Scholar] [CrossRef]
  64. Gado, M.A.; Atia, B.M.; Cheira, M.F.; Elawady, M.E.; Demerdash, M. Highly efficient adsorption of uranyl ions using hydroxamic acid-functionalized graphene oxide. Radiochim. Acta 2021, 109, 743–757. [Google Scholar] [CrossRef]
  65. Abd El-Rahem, K.A.; Taher, M.A.; Cheira, M.F. Peculiarities of u(vi) adsorption from acidic solution using chitin-derived chitosan as a low-priced bio-adsorbent. Al-Azhar Bull. Sci. 2021, 32, 57–70. [Google Scholar] [CrossRef]
  66. Salem, A.R.; El-Maghrabi, H.H. Preparation and characterization of modified anion exchange resin for uranium adsorption: Estimation of nonlinear optimum isotherm, kinetic model parameters, error function analysis and thermodynamic studies. J. Dispers. Sci. Technol. 2021, 42, 1–15. [Google Scholar] [CrossRef]
  67. Haggag, E.S.A.; Abdelsamad, A.A.; Masoud, A.M. Potentiality of uranium extraction from acidic leach liquor by polyacrylamide-acrylic acid titanium silicate composite adsorbent. Int. J. Environ. Anal. Chem. 2020, 100, 204–224. [Google Scholar] [CrossRef]
  68. Meng, J.; Lin, X.; Li, H.; Zhang, Y.; Zhou, J.; Chen, Y.; Shang, R.; Luo, X. Adsorption capacity of kelp-like electrospun nanofibers immobilized with bayberry tannin for uranium(vi) extraction from seawater. RSC Adv. 2019, 9, 8091–8103. [Google Scholar] [CrossRef] [Green Version]
  69. Mahmoud, N.S.; Atwa, S.T.; Sakr, A.K.; Abdel Geleel, M. Kinetic and thermodynamic study of the adsorption of Ni (II) using Spent Activated clay Mineral. N. Y. Sci. J. 2012, 5, 62–68. [Google Scholar]
  70. Abdel Geleel, M.; Atwa, S.T.; Sakr, A.K. Removal of Cr (III) from aqueous waste using Spent Activated Clay. J. Am. Sci. 2013, 9, 256–262. [Google Scholar]
Figure 1. A schematic diagram of (a) fabrication of silica from rice husk, (b) preparation of GO from spent zinc–carbon batteries, (c) preparation of SiO2/GO, (d) U(VI) sorption at SiO2/GO.
Figure 1. A schematic diagram of (a) fabrication of silica from rice husk, (b) preparation of GO from spent zinc–carbon batteries, (c) preparation of SiO2/GO, (d) U(VI) sorption at SiO2/GO.
Sustainability 14 02699 g001
Figure 2. XRD spectra of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Figure 2. XRD spectra of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Sustainability 14 02699 g002
Figure 3. SEM of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Figure 3. SEM of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Sustainability 14 02699 g003
Figure 4. EDX of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Figure 4. EDX of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Sustainability 14 02699 g004
Figure 5. N2 sorption/desorption isotherm of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Figure 5. N2 sorption/desorption isotherm of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Sustainability 14 02699 g005
Figure 6. FTIR of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Figure 6. FTIR of (a) SiO2, (b) GO, (c) SiO2/GO, and (d) U/SiO2/GO.
Sustainability 14 02699 g006
Figure 7. Raman of (a) SiO2, (b) GO, (c) and SiO2/GO.
Figure 7. Raman of (a) SiO2, (b) GO, (c) and SiO2/GO.
Sustainability 14 02699 g007
Figure 8. (a) Effect of pH on U(VI) uptake; (b) effect of SiO2/GO on U(VI) sorption.
Figure 8. (a) Effect of pH on U(VI) uptake; (b) effect of SiO2/GO on U(VI) sorption.
Sustainability 14 02699 g008
Figure 9. (a) Effect of sorption time on U(VI) uptake; (b) pseudo’s first-order and second-order nonlinear models for U(VI) uptake on SiO2/GO.
Figure 9. (a) Effect of sorption time on U(VI) uptake; (b) pseudo’s first-order and second-order nonlinear models for U(VI) uptake on SiO2/GO.
Sustainability 14 02699 g009
Figure 10. (a) Influence of initial U(VI) concentration on U(VI) uptake at SiO2/GO, (b) the Langmuir and Freundlich nonlinear models for U(VI) adsorption.
Figure 10. (a) Influence of initial U(VI) concentration on U(VI) uptake at SiO2/GO, (b) the Langmuir and Freundlich nonlinear models for U(VI) adsorption.
Sustainability 14 02699 g010
Figure 11. (a) Effect of temperature on U(VI) uptake at SiO2/GO sorbent; (b) log Kd vs. 1/T relation for uptake.
Figure 11. (a) Effect of temperature on U(VI) uptake at SiO2/GO sorbent; (b) log Kd vs. 1/T relation for uptake.
Sustainability 14 02699 g011
Figure 12. Effect of (a) stripping agents; (b) H2SO4 concentration; (c) S:L ratio; (d) eluting time upon U(VI) desorption from U/SiO2/GO.
Figure 12. Effect of (a) stripping agents; (b) H2SO4 concentration; (c) S:L ratio; (d) eluting time upon U(VI) desorption from U/SiO2/GO.
Sustainability 14 02699 g012
Figure 13. Sorption–desorption of U(VI) according to relation for sorption–desorption cycle.
Figure 13. Sorption–desorption of U(VI) according to relation for sorption–desorption cycle.
Sustainability 14 02699 g013
Table 1. Surface area, pore size, and pore volume of SiO2, GO, SiO2/GO, and U/SiO2/GO.
Table 1. Surface area, pore size, and pore volume of SiO2, GO, SiO2/GO, and U/SiO2/GO.
MaterialsSBET, m2/gPore-Volume, cc/gPore-Size, nm
SiO225.890.0372.87
GO37.370.0523.19
SiO2/GO35.450.0452.99
U/SiO2/GO32.150.0392.85
Table 2. Thermodynamic parameters of U(VI) sorption upon SiO2/GO sorbent.
Table 2. Thermodynamic parameters of U(VI) sorption upon SiO2/GO sorbent.
Temp, KG°, kJ/molH°, kJ/molS°, kJ/(mol.K)
298−8.2980−12.33−1.35 × 10−2
303−8.230
308−8.163
313−8.095
318−8.027
323−7.959
328−7.892
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hassanin, M.A.; Negm, S.H.; Youssef, M.A.; Sakr, A.K.; Mira, H.I.; Mohammaden, T.F.; Al-Otaibi, J.S.; Hanfi, M.Y.; Sayyed, M.I.; Cheira, M.F. Sustainable Remedy Waste to Generate SiO2 Functionalized on Graphene Oxide for Removal of U(VI) Ions. Sustainability 2022, 14, 2699. https://0-doi-org.brum.beds.ac.uk/10.3390/su14052699

AMA Style

Hassanin MA, Negm SH, Youssef MA, Sakr AK, Mira HI, Mohammaden TF, Al-Otaibi JS, Hanfi MY, Sayyed MI, Cheira MF. Sustainable Remedy Waste to Generate SiO2 Functionalized on Graphene Oxide for Removal of U(VI) Ions. Sustainability. 2022; 14(5):2699. https://0-doi-org.brum.beds.ac.uk/10.3390/su14052699

Chicago/Turabian Style

Hassanin, Mohamed A., Sameh H. Negm, Mohamed A. Youssef, Ahmed K. Sakr, Hamed I. Mira, Tarek F. Mohammaden, Jamelah S. Al-Otaibi, Mohamed Y. Hanfi, M. I. Sayyed, and Mohamed F. Cheira. 2022. "Sustainable Remedy Waste to Generate SiO2 Functionalized on Graphene Oxide for Removal of U(VI) Ions" Sustainability 14, no. 5: 2699. https://0-doi-org.brum.beds.ac.uk/10.3390/su14052699

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