Next Article in Journal
Correction: Ren et al. Injectable and Antioxidative HT/QGA Hydrogel for Potential Application in Wound Healing. Gels 2021, 7, 204
Next Article in Special Issue
Investigation of Gelation Techniques for the Fabrication of Cellulose Aerogels
Previous Article in Journal
Creation of Polymer Hydrogelator/Poly(Vinyl Alcohol) Composite Molecular Hydrogel Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption Ability of Graphene Aerogel and Reduced Graphene Aerogel toward 2,4-D Herbicide and Salicylic Acid

by
Alexandra Yu. Kurmysheva
1,*,
Oleg Yanushevich
2,
Natella Krikheli
2,
Olga Kramar
2,
Marina D. Vedenyapina
3,
Pavel Podrabinnik
1,
Nestor Washington Solís Pinargote
1,
Anton Smirnov
1,
Ekaterina Kuznetsova
1,
Vladislav V. Malyavin
4,
Pavel Peretyagin
1,2 and
Sergey N. Grigoriev
1
1
Laboratory of Electric Current Assisted Sintering Technologies, Moscow State University of Technology “STANKIN”, Vadkovsky per. 1, 127055 Moscow, Russia
2
Scientific Department, A.I. Yevdokimov Moscow State University of Medicine and Dentistry, Delegatskaya St., 20, p. 1, 127473 Moscow, Russia
3
N. D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, Leninsky Prospect 47, 119991 Moscow, Russia
4
Laboratory of Petroleum Chemistry and Petrochemical Synthesis, Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences, Leninsky Prospect 29, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Gels 2023, 9(9), 680; https://doi.org/10.3390/gels9090680
Submission received: 7 August 2023 / Revised: 19 August 2023 / Accepted: 22 August 2023 / Published: 23 August 2023
(This article belongs to the Special Issue Characterization and Applications of Aerogels)

Abstract

:
Within this work, new aerogels based on graphene oxide are proposed to adsorb salicylic acid (SA) and herbicide 2,4-Dichlorophenoxyacetic acid (2,4-D) from aqueous media. Graphene oxide aerogel (GOA) and reduced graphene oxide aerogel (rGOA) were obtained by freeze-drying processes and then studied by Raman spectroscopy, Fourier-transform infrared spectroscopy (FT-IR), and Brunauer–Emmett–Teller (BET) analysis. The influence of contact time and the concentration of the adsorbates were also assessed. It was found that equilibrium for high adsorption is reached in 150 min. In a single system, the pseudo-first-order, pseudo-second-order kinetic models, Intraparticle diffusion, and Elovich models were used to discuss the detail of the aerogel adsorbing pollutant. Moreover, the Langmuir, Freundlich, and Temkin adsorption models were applied to describe the equilibrium isotherms and calculate the isotherm constants.

1. Introduction

In recent years, the number of hazardous organic compounds entering natural water systems has increased dramatically with urban and industrial development. Among the organic pollutants, biologically active organic compounds (BAOCs) pose a significant threat to the normal functioning of natural aquatic ecosystems. For example, 2,4-Dichlorophenoxyacetic acid (2,4-D) is a popular agricultural herbicide and also is a common environmental pollutant [1]. Due to poor biodegradability, low soil absorption coefficient, and high water solubility, 2,4-D, after application, leaches into the soil and then into the surface and groundwater [2]. In addition, 2,4-D affects not only weed plants but also may curb growth rates, cause problems with the reproductive system, and provoke deviations in behavior or death of organisms [1]. Furthermore, 2,4-D is toxic to the human central nervous system, kidneys, and liver [3].
Another group of BAOCs includes medicines and, particularly, salicylic acid (SA), which is one of the most widespread compounds for analgesic and antipyretic agents [4]. After use, SA is excreted from the body as the parent drug and its metabolites and eventually ends up in sewers and wastewater treatment plants [5]. Because SA is not completely removed by conventional wastewater treatment processes; it is discharged into the receiving environment along with effluent and sludge [6]. SA is a highly toxic BAOC and causes serious environmental pollution [4].
To extract SA and 2,4-D from aquatic environments, different materials such as polymers [7,8], biochar [9,10], metal–organic frameworks [11,12], and activated carbon [13,14,15,16] were used. Among the various adsorbing materials, graphene oxide (GO) is a promising compound, as it is reported that oxygen-containing functional groups on the surface of GO can act as active adsorption sites for various pollutants [17]. However, 2-dimensional GO has poor structural stability in aquatic environments making it difficult to retrieve and reuse [18]. Three-dimensional structures (sponges, aerogels) can preserve GO features while allowing multiple reuses while in service [19]. Advanced GO-based aerogels (GOAs) are applied in many fields, including medicine [20,21,22].
Both reduced and non-reduced GOAs are well-studied and known as powerful adsorbents for extracting different oils [23,24], organic colorants [25,26,27,28], and metals [29,30,31,32] from aquatic environments. Therefore, it is of interest to study the adsorption capacity of GOA for SA and 2,4-D.
The work aims to study the adsorption capacity of GOAs in reduced and non-reduced forms for SA and 2,4-D. For this purpose, samples of graphene oxide aerogel and reduced graphene oxide aerogel (rGOA) were produced. The specific surface area was measured, and Raman and FT-IR studies were carried out. Adsorption rates were calculated by pseudo-first and pseudo-second-order kinetic models, and adsorption isotherms were described by Langmuir, Freundlich, and Temkin models.

2. Results and Discussion

2.1. Characterization of Adsorbents

Figure 1 illustrates N2 adsorption–desorption isotherms that were conducted on graphene oxide aerogels to determine their specific surface areas and pore sizes. The adsorption/desorption behavior of nitrogen on GOA and rGOA was typical for type IV with an isothermal H3 hysteresis loop indicating the presence of a mesoporous structure in the samples. The specific surface area measured by the BET method (SBET) for GOA and rGOA was 80.1 and 11.9 m2/g, respectively, while the total pore volume was 0.67 and 0.06 cm3/g. For the GOA sample, following the pore size distribution diagram (Figure 1c), pore sizes ranged from 5 nm to 16 nm, proving that GOA had a mesoporous structure. The average pore size of the GOA sample was 11.6 nm. The average pore diameter of the rGOA sample was impossible to calculate due to the low total pore volume.
The specific surface area of the rGOA was much smaller than that of GOA, which may be due to the adhesion of graphene during the reduction of GO with hydrazine [33] and the agglomeration of the sample during the BET study [34].
Raman spectroscopy was used to examine structural changes in the GOA and rGOA samples after reduction (Figure 2). The obtained results show that the GOA sample has two characteristic peaks at 1350 cm−1 and 1582 cm−1, indicating D and G bands, respectively. Normally, the D band is disordered owing to structural defects, edge effects, and dangling sp2 carbon bonds, while the G band is due to the in-plane bond stretching of the sp2 carbons [35]. For the GOA sample, the G band is usually found at 1582 cm−1. After reduction, it was shifted to a lower wavenumber of 1580 cm−1, suggesting overlapping or sticking of several layers of graphene on top of each other after reduction with hydrazine hydrate [36,37,38]. This opinion is also supported by the results of specific surface area measurements performed by BET. Importantly, the intensity ratio of D and G bands for non-reduced forms was 1.08 against 1.86 for the rGOA, which is 1.7 times less. Thus, it can be concluded that newly formed sp2-hybrid domains in the reduced GOA are smaller in size than in the non-reduced one but more widespread [39].
The obtained compounds were also studied by FT-IR spectroscopy by identifying the molecular groups’ structures before and after the reduction process. The FT-IR spectrum of the aerogels is shown in Figure 3.
Since the samples were thoroughly dried, no significant signals belonging to water molecules encapsulated within the structure and stretching vibration of hydroxyl groups, theoretically located around 3500–3000 cm−1, were almost completely removed [40]. A weak peak observed on the GOA spectrum at 2360 cm−1 can be attributed to stretching vibrations of the C–H bond [41]. Peaks at 1718 cm−1 of the GOA sample correspond to the stretching vibrations of carbonyl groups [42], while peaks at 1619 cm−1 could be associated with the C=O stretching vibrations from carbonyl or carboxyl groups [43]. The peaks at 1361 cm−1 and 1222 cm−1 correspond to the stretching vibrations of the C=O and C–O bond, respectively [44]. The peak at 1050 cm−1 can be attributed to the C–O–C stretching vibration of the alkoxy group [45]. According to [46], a peak at 981 cm−1 is assigned to the ring out-of-plane deformation. As seen in Figure 3, the peak intensities of the rGOA oxygen-containing functional groups decreased compared to GOA, and some are gone completely. Thus, the reduction of GO by hydrazine hydrate was successful. Along with that, a new peak at 850 cm−1, identified as the C–O–C epoxy group, was found in the rGOA spectrum. [47]. As reported in [48], the presence of surface functional groups is critical for adsorbents removing 2,4-D and SA from aqueous solutions.
The morphology and the structure of the GOA and rGOA samples studied by scanning electron microscopy (SEM) are observed in Figure 4. Both samples demonstrate a wrinkled sheet structure [32]. However, compared to non-reduced aerogel, rGOA is characterized by smaller and more compacted sheets less distanced from each other. It suggests that such changes in morphology are attributed to the weakening of electrostatic repulsion between graphene nanosheets due to the removal of oxygen-containing groups during the reduction process [39].

2.2. Adsorption Kinetics

Defining the adsorption kinetics helps to understand the mass exchange mechanism during adsorption and specify the rate-limiting stage [24]. Adsorption kinetics were studied using the pseudo-first-order (PFO) [49], the pseudo-second-order (PSO) [50], the Elovich model [51], and the intraparticle diffusion model (IPD) [52]. The model parameters and correlation coefficients (R2) were obtained by nonlinear regression using Origin Pro 2016 software.
The pseudo-first-order model suggests that the rate of change in the absorption of solute materials over time is directly proportional to the difference in saturation concentration and amount of solid material absorbed over time. This assumption is usually applicable at the beginning of the process. Normally, adsorption kinetics follow this model if the adsorption process proceeds through diffusion through the phase boundary. The pseudo-first-order model can be described by the following equation:
q t = q e · 1 e x p k 1 · t ,
where k 1 (min−1) is the rate constant, and q e and q t (mg/g) are the number of adsorbed ions of 2,4-D or SA at equilibrium and at time t , respectively.
On the other hand, the pseudo-second-order model assumes that chemisorption is a rate-limiting stage of the adsorption process and predicts its behavior over the entire adsorption range. In this case, the adsorption rate depends on the adsorption capacity of the sorbent and not on the concentration of the adsorbate [50]. One of the main advantages of this model compared to the PFO model is that the PSO model can be used to calculate the equilibrium adsorption capacity at any time. The pseudo-second-order model equation can be represented as follows:
q t = q e 2 · k 2 · t 1 + q e 2 · k 2 · t ,
where k 2 (g/(mg∙min)) is the rate constant in the PSO model.
The Elovich model is based on the hypothesis that the adsorption rate decreases exponentially as the amount of adsorbed solute material grows. It is expressed in the following way [51]:
q t = 1 β ln α β t + 1 ,
where α (mg/g min)–initial adsorption rate; β (g/mg)–desorption constant during each experiment.
Figure 5 and Figure 6 present the experimental results on the 2,4-D and SA adsorption on the aerogel samples described by three models: PFO, PSO, and Elovich model. As seen from the figures, the adsorption equilibrium is achieved in less than 150 min. The parameters calculated according to the models are shown in Table 1 and Table 2. The PSO model gives the best fit with higher regression coefficients (R2) (>98%) for both 2,4-D and SA for both aerogel samples, signalizing that chemisorption is the adsorption rate-limiting step [53].
The kinetic process is directed by various mechanisms and stages of the adsorption phenomena. Usually, there are four main rate-limiting stages [54]: (1) volumetric diffusion, in which the adsorbate is transferred from the volume of the solution to the liquid film surrounding the solid adsorbent; (2) mass transfer of a solute from the boundary film to the surface of the adsorbent, also called external diffusion; (3) intraparticulate diffusion of solute material in the pores of the adsorbent. The porosity of the adsorbent is a key factor for the stage; (4) interaction with surface areas through chemisorption. Both the first and second stages can be attributed to film diffusion. The third stage is the stage of resistance to the internal diffusion of particles. The IPD model serves to evaluate the influence of the third adsorption step on the rate of the entire adsorption process and is studied by applying the Weber and Morris model as shown in Equation (4):
q t = k p i · t 0.5 + C
where k p i (mg/g min0.5) is the IPD rate constant; t0.5 is the square root of time; C (mg/g) is an intercept. According to Equation (5), if the kinetic process is dominated by the intraparticle diffusion, the plot of qt versus t0.5 should be a straight line. In addition, if the intercept of the plot (C) is equal to 0, then the intraparticle diffusion is the only rate-limiting step [55].
As can be seen from Figure 7, the plots of qt versus t0.5 do not take a linear form during the entire adsorption time. This shows that the adsorption of 2,4-D and SA on GOA and rGOA does not follow the intraparticle diffusion model during the whole adsorption process. However, polylinearity is observed on the plots indicating the presence of two or more stages involved in the adsorption process. Moreover, each linear section of these graphs that deviates from the origin also shows that intraparticle diffusion is not the only rate-limiting stage.
According to [56], the kpi values can reflect the degree of contribution of initial sorption and intraparticle diffusion to the entire sorption process. At the first stage, in all cases, the best linearization with the corresponding first-rate constants IPD model kp1 (Table 3 and Table 4) was obtained, indicating the fastest process that can be associated with film diffusion. As a result, the hydrodynamic boundary layer was probably formed due to the diffusion of 2,4-D and SA molecules from the solution onto the outer surfaces of GOA and rGOA. In contrast, the second and last stages showed a decrease in the second- and third-rate constant of the IPD model (kp2 and kp3) with corresponding lower values of R2. The decrease in the order of the adsorption rate (kpi) in the sequence kp1 > kp2 > kp3 may be due to the resistance to a mass transfer occurring in the second phase of adsorption [56]. The second stage denotes a process in which the movement of adsorbate molecules can migrate from the outer surfaces of the adsorbent to the inner regions of the pores. The rate-limiting stage may be determined by a non-zero intersection calculated from the intersection (C) on the second area of the graph. In related studies, the calculated intersections (C) on the graphs were positive (>0), suggesting that the process showed slow diffusion within the particles, mainly due to the small concentration of adsorbate molecules remaining in the aqueous medium and the saturation of the remaining pores with molecules 2,4-D and SA [55].

2.3. Adsorption Isotherms

Three commonly used isotherm models, Langmuir [57], Freundlich [58], and Temkin [59] were adopted to depict the solid-liquid adsorption system.
The nonlinear form of the Langmuir isotherm could be described as follows:
q e = q m b l c e / 1 + b L c e ,
where q m —maximum capacity of the monolayer (mg/g), and b l —the adsorption coefficient (L/g).
Mathematically, Freundlich isotherm can be presented as in Equation (4):
q e = K F c e 1 / n ,
where K F —the coefficient of distribution or adsorption coefficient (L/g).
Temkin isotherm is shown in Equation (5):
q e = R T / b T ln A c e ,
where b T —the adsorption coefficient (J/mol); R—the universal gas constant of 8.314 J/(mol∙K); A—the constant, L/g; T—the absolute temperature (K).
Three kinds of isotherms have their own assumptions [24]: the Langmuir model assumes that the active sites from homogeneous surface share equal relations with the adsorbate. The hypothesis of the Freundlich model describes heterogeneous adsorption surfaces with different affinities of adsorption sites. The Temkin model assumes that the adsorption heat of adsorbate decreases linearly with the increase of coverage.
Figure 7a,b and Figure 8a,b show the results of applying the isotherm models to the experimental data of the adsorption of 2,4-D and SA ions on GOA and rGOA. Experimental and calculated parameters for each isotherm model are given in Table 5.
As can be seen from Figure 8 and Figure 9, the capacity of the adsorbents with respect to SA is a bit higher than that of 2,4-D, which can be attributed to a smaller size of SA molecules.
It can be concluded from Table 5 that, according to the correlation coefficient (R2), the isotherm can be fitted with the Langmuir and Freundlich isotherms with high precision. However, the Freundlich fitting model appeared more representative than Langmuir one, comparing the R2 values. The adsorption process, according to the Langmuir model, must meet certain conditions, such as a homogeneous adsorbent surface, monolayer adsorption of the adsorbate on the adsorbent surface, and the absence of interaction between adsorbed particles [55]. However, some adsorption interactions do not always follow this rule. It may be a reason for the slight deviation of the 2,4-D and SA adsorption isotherms on aerogel samples from the Langmuir model in this work. The Freundlich isotherm is an empiric model, which is not limited to covering the surface of a sorbent by adsorbate molecules, but also describes multilayer adsorption and considers interactions between the adsorbed molecules. The Freundlich model constant KF reflects how favorable the adsorption is. The constant of the 1/n in the Freundlich model is related to the adsorption intensity, which varies with the heterogeneity of materials. Adsorption is favorable when the 1/n ratio is between 0 and 1 [55]. As seen from Table 5, the value of the 1/n ratio is between 0 and 1, indicating that the adsorption behavior of 2,4-D and SA on aerogels is favorable.
The maximum adsorption capacity is higher for GOA for both adsorbates than for rGOA, which may be due to the absence of oxygen-containing groups on the surface of rGOA. The maximum adsorption capacity and the rate of achieving the adsorption equilibrium of GOA in comparison with the results obtained in other works are presented in Table 6. On that background, the obtained results may be found relevant.

3. Conclusions

Within this work, both non-reduced and reduced with hydrazine hydrate GO-based aerogels were studied. Raman spectroscopy results confirmed the successful reduction process. The adsorption capacity of GOA for 2,4-D and SA was measured to be 42.63 mg/g and 57.61 mg/g, respectively, while rGOA showed only 10.92 mg/g and 16.01 mg/g. The difference in adsorption capacity may be attributed to the absence of oxygen-containing functional groups on the surface of rGOA, which are essential for the successful adsorption of the 2,4-D and SA molecules. The adsorption kinetics for both aerogel samples are consistent with the PSO model, indicating that the rate-limiting step may be chemical adsorption by sharing or exchanging electrons between the adsorbate and adsorbent.
Thus, the sorption ability of the herbicide 2,4-D and salicylic acid of aerogels based on graphene oxide was successfully confirmed in the work.

4. Materials and Methods

4.1. Materials and Reagents

All the chemicals in this study were of analytical grade and were used as received without any additional purification. Graphene oxide (GO) suspension (50 mg/mL) was provided by Graphenox (Moscow, Russia). The element composition of the GO suspension was provided by the manufacturer and included C–46%, O–49%, H–2.5%, and S < 1.5%. The hydrazine hydrate solution (100%), 2,4-Dichlorophenoxyacetic acid, and salicylic acid were purchased from Merck (Darmstadt, Germany). Hydrochloric acid HCl was provided by Himprom (Kemerovo, Russia).

4.2. Preparation of GOA and rGOA

GO was dispersed into water by ultrasonication with a concentration of 3 mg/mL. The GO suspension was frozen at −70 °C and sublimated for 30 h by a freeze-dryer FreeZone 2.5 (Labconco, Kansas, MO, USA) to obtain the product. Further, the GOA samples were subjected to a chemical reduction in hydrazine vapor for 24 h to obtain rGOA. The obtained aerogels are presented in Figure 10.

4.3. Characterization of GOA and rGOA

Fourier-transform infrared spectroscopy (FTIR) spectra were obtained on a Vertex 70 infrared Fourier spectrometer (Bruker Optik GmbH, Ettlingen, Germany). The Raman spectra of graphene oxide and reduced graphene oxide aerogels were obtained on a DXRTM 2 Raman Micro-scope instrument (Thermo Fisher Scientific, Waltham, MA, USA) equipped with a 780 nm laser with a power of 15 mW. The laser beam was focused through the lens of an optical microscope with a magnification of 50× into a 0.8 μm spot on an area under investigation. The integration time for each spectrum was approximately 6 s. Scanning electron microscopy (SEM) images were obtained using SEM JEOL JSM-7001F (JEOL Ltd., Tokyo, Japan). The aerogel samples were preliminarily coated with a thin electric conductive gold film by an ion sputter coater Quorum Q150T Emscope sputter coater (Quorum Technologies Ltd., Newhaven, UK).
Nitrogen adsorption isotherms were recorded by using a BELSORP-miniX instrument (MicrotracBEL Japan, Inc. BELSORP-mini, Osaka, Japan). The preliminary preparation stage included thermal degasification of samples at 200 °C and 10 Pa pressure for 8 h. At the end of the preparation, the cuvette with the sample was cooled and weighed again to determine the weight loss of the sample after processing. After that, the cuvette was placed in the analytical port of the device, and the adsorption capacity was analyzed with the sequential supply of nitrogen until an equilibrium pressure in the system was reached at a temperature of 77 K in the relative pressure range p/p0 = 0–0.993. The total specific surface area was estimated according to the Brunauer–Emmett–Teller method (BET). The total pore volumes were estimated from a single point adsorption at p/p0 = 0.99.

4.4. Adsorption Experiments

The adsorption of SA и 2,4-D on graphene aerogel samples from aqueous solutions was carried out in a thermostated cell with a reflux condenser with nonstop mixing with a magnetic stirrer (150 rpm). The initial pH value of the solutions was set to 3 by adding the HCl solution dropwise. This pH value is optimal for better adsorption of 2,4-D and SA molecules in an acidic medium [1,2,3,4].
Adsorption kinetics were studied by tracking changes in the concentration of each adsorbate in time and the initial concentrations of 25–125 mg/L. The concentration of the SA and 2,4-D during the adsorption was determined by the collected specimens via UV spectroscopy from adsorption at 295 nm and 283 nm, respectively, on a Hitachi U-1900 instrument (Tokyo, Japan). The following Equation (8) was used to estimate the number of ions of SA and 2,4-D adsorbed per 1 g of graphene aerogel (mg/g):
q e = C 0 C e · V m ,
where C 0 and C e —initial and equilibrium concentration of each adsorbate ion, respectively; mg/L; V —volume of the solution, L; m —the mass of the adsorbent, g.

Author Contributions

Conceptualization, A.Y.K., N.W.S.P. and M.D.V.; methodology, A.Y.K. and M.D.V.; software, A.Y.K. and V.V.M.; validation, A.Y.K. and M.D.V.; formal analysis, A.Y.K., N.W.S.P. and E.K.; investigation, A.Y.K., N.W.S.P., M.D.V. and A.S.; resources, S.N.G. and M.D.V.; data curation, S.N.G. and M.D.V.; writing—original draft preparation, A.Y.K., P.P. (Pavel Podrabinnik) and E.K.; writing—review and editing, N.W.S.P. and M.D.V.; visualization, A.Y.K. and E.K.; supervision, S.N.G. and M.D.V.; project administration, S.N.G.; funding acquisition, S.N.G., O.Y., N.K., O.K. and P.P. (Pavel Peretyagin). All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Health of the Russian Federation under project 056-00041-23-00.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The study was carried out on the equipment of the Center of Collective Use “State Engineering Center” of the MSUT “STANKIN” (project 075-15-2021-695, unique id RF----2296.61321X0013).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vinayagam, R.; Ganga, S.; Murugesan, G.; Rangasamy, G.; Bhole, R.; Goveas, L.C.; Varadavenkatesan, T.; Dave, N.; Samanth, A.; Devi, V.R.; et al. 2,4-Dichlorophenoxyacetic acid (2,4-D) adsorptive removal by algal magnetic activated carbon, nanocomposite. Chemosphere 2023, 310, 136883. [Google Scholar] [CrossRef] [PubMed]
  2. Salomón, Y.L.O.; Georgin, J.; Franco, D.S.P.; Netto, M.S.; Piccilli, D.G.A.; Foletto, E.L.; Oliveira, L.F.S.; Dotto, G.L. High-performance removal of 2,4-dichlorophenoxyacetic acid herbicide in water using activated carbon derived from Queen palm fruit, endocarp (Syagrus romanzoffiana). J. Environ. Chem. Eng. 2021, 9, 104911. [Google Scholar] [CrossRef]
  3. Wu, G.; Ma, J.; Li, S.; Wang, S.; Jiang, B.; Luo, S.; Li, J.; Wang, X.; Guan, Y.; Chen, L. Cationic metal-organic frameworks as an efficient adsorbent for the removal of 2,4-dichlorophenoxyacetic acid from aqueous solutions. Environ. Res. 2020, 186, 10954. [Google Scholar] [CrossRef] [PubMed]
  4. Sun, Y.; Bai, L.; Wang, T.; Cao, S.; Han, C.; Sun, X. Effective scavenging and selective adsorption of salicylic acid from wastewater using a novel deep eutectic solvents-based chitosan-acrylamide surface molecularly imprinted hydrogel. Appl. Surf. Sci. 2023, 608, 155102. [Google Scholar] [CrossRef]
  5. Lozano, I.; Pérez-Guzmán, C.J.; Mora, A.; Mahlknecht, J.; Aguilar, C.L.; Cervantes-Avilés, P. Pharmaceuticals and personal care products in water streams: Occurrence, detection, and removal by electrochemical advanced oxidation processes. Sci. Total Environ. 2022, 827, 154348. [Google Scholar] [CrossRef]
  6. Couto, C.F.; Lange, L.C.; Amaral, M.C.S. Occurrence, fate and removal of pharmaceutically active compounds (PhACs) in water and wastewater treatment plants—A review. J. Water Process. Eng. 2019, 32, 100927. [Google Scholar] [CrossRef]
  7. Fernández-García, M.; Ciarlantini, C.; Francolini, I.; Girelli, A.; Piozzi, A. Molecularly Imprinted Polymers Based on Chitosan for 2,4-Dichlorophenoxyacetic Acid Removal. Int. J. Mol. Sci. 2022, 23, 13192. [Google Scholar] [CrossRef]
  8. Huang, J.; Yang, L.; Wang, X.; Li, H.; Chen, L.; Liu, Y.-N. A novel post-cross-linked polystyrene/polyacryldiethylenetriamine (PST_pc/PADETA) interpenetrating polymer networks (IPNs) and its adsorption towards salicylic acid from aqueous solutions. Chem. Eng. J. 2014, 248, 216–222. [Google Scholar] [CrossRef]
  9. Ćwieląg-Piasecka, I.; Jamroz, E.; Medy’ nska-Juraszek, A.; Bednik, M.; Kosyk, B.; Polláková, N. Deashed Wheat-Straw Biochar as a Potential Superabsorbent for Pesticides. Materials 2023, 16, 2185. [Google Scholar] [CrossRef]
  10. Ahmed, M.J.; Hameed, B.H. Adsorption behavior of salicylic acid on biochar as derived from the thermal pyrolysis of barley straws. J. Clean. Prod. 2018, 195, 1162–1169. [Google Scholar] [CrossRef]
  11. Wang, J.; Teng, Y.; Jia, S.; Li, W.; Yang, T.; Cheng, Y.; Zhang, H.; Li, X.; Li, L.; Wang, C. Highly efficient removal of salicylic acid from pharmaceutical wastewater using a flexible composite nanofiber membrane modified with UiO-66(Hf) MOFs. Appl. Surf. Sci. 2023, 625, 157183. [Google Scholar] [CrossRef]
  12. Lazarou, S.; Antonoglou, O.; Mourdikoudis, S.; Serra, M.; Sofer, Z.; Dendrinou-Samara, C. Magnetic Nanocomposites of Coated Ferrites/MOF as Pesticide Adsorbents. Molecules 2023, 28, 39. [Google Scholar] [CrossRef] [PubMed]
  13. Taoufik, N.; Elmchaouri, A.; Mahmoudi, S.E.; Korili, S.A.; Gil, A. Comparative analysis study by response surface methodology and artificial neural network on salicylic acid adsorption optimization using activated carbon. Environ. Nanotechnol. Monit. Manag. 2021, 15, 100448. [Google Scholar] [CrossRef]
  14. Bernal, V.; Giraldo, L.; Moreno-Piraján, J.C. Thermodynamic analysis of acetaminophen and salicylic acid adsorption onto granular activated carbon: Importance of chemical surface and effect of ionic strength. Thermochim. Acta 2020, 683, 178467. [Google Scholar] [CrossRef]
  15. Hazrin, H.M.M.N.; Lim, A.; Li, C.; Chew, J.J.; Sunarso, J. Adsorption of 2,4-dichlorophenoxyacetic acid onto oil palm trunk-derived activated carbon: Isotherm and kinetic studies at acidic, ambient condition. Mater. Today Proc. 2022, 64, 1557–1562. [Google Scholar] [CrossRef]
  16. Vieira, Y.; Schnorr, C.; Piazzi, A.C.; Netto, M.S.; Piccini, W.M.; Franco, D.S.P.; Mallmann, E.S.; Georgin, J.; Silva, L.F.O.; Dotto, G.L. An advanced combination of density functional theory simulations and statistical physics modeling in the unveiling, and prediction of adsorption mechanisms of 2,4-D pesticide to activated carbon. J. Mol. Liq. 2022, 361, 119639. [Google Scholar] [CrossRef]
  17. Liu, L.; Zhang, B.; Zhang, Y.; He, Y.; Huang, L.; Tan, S.; Cai, X. Simultaneous Removal of Cationic and Anionic Dyes from Environmental Water Using Montmorillonite-Pillared Graphene Oxide. J. Chem. Eng. Data 2015, 60, 1270–1278. [Google Scholar] [CrossRef]
  18. Shen, Y.; Fang, Q.; Chen, B. Environmental Applications of Three-Dimensional Graphene-Based Macrostructures: Adsorption, Transformation, and Detection. Environ. Sci. Technol. 2015, 49, 67–84. [Google Scholar] [CrossRef]
  19. Sun, Z.; Fang, S.; Hu, Y.H. 3D Graphene Materials: From Understanding to Design and Synthesis Control. Chem. Rev. 2020, 120, 10336–10453. [Google Scholar] [CrossRef]
  20. Trembecka-Wójciga, K.; Sobczak, J.J.; Sobczak, N. A comprehensive review of graphene-based aerogels for biomedical applications. The impact of synthesis parameters onto material microstructure and porosity. Arch. Civ. Mech. Eng. 2023, 23, 133. [Google Scholar] [CrossRef]
  21. Qiao, F.; Wang, X.; Han, Y.; Kang, Y.; Yan, H. Preparation of poly (methacrylic acid)/graphene oxide aerogel as solid-phase extraction adsorbent for extraction and determination of dopamine and tyrosine in urine of patients with depression. Anal. Chim. Acta 2023, 1269, 341404. [Google Scholar] [CrossRef]
  22. Lim, H.; Kim, J.Y.; Yoon, M.G.; Kang, Y.-M.; Park, Y.M.; Lee, H.-N.; Moon, S.-H.; Koh, W.-G.; Kim, H.-J. Facile control of porous structure of graphene aerogel via two-step drying process and its effect on drug release. J. Porous Mater. 2023. [Google Scholar] [CrossRef]
  23. Diao, S.; Liu, H.; Chen, S.; Xu, W.; Yu, A. Oil adsorption performance of graphene aerogels. J. Mater. Sci. 2020, 55, 4578–4591. [Google Scholar] [CrossRef]
  24. Huang, J.; Yan, Z. Adsorption Mechanism of Oil by Resilient Graphene Aerogels from Oil–Water Emulsion. Langmuir 2018, 34, 1890–1898. [Google Scholar] [CrossRef] [PubMed]
  25. Nguyen, V.T.; Ha, L.Q.; Nguyen, T.D.L.; Ly, P.H.; Nguyen, D.M.; Hoang, D. Nanocellulose and Graphene Oxide Aerogels for Adsorption and Removal Methylene Blue from an Aqueous Environment. ACS Omega 2022, 7, 1003–1013. [Google Scholar] [CrossRef] [PubMed]
  26. Mkrtchyan, E.S.; Neskoromnaya, E.A.; Burakova, I.V.; Ananyeva, O.A.; Revyakina, N.A.; Babkin, A.V.; Kuznetsova, T.S.; Kurnosov, D.A.; Burakov, A.E. Comparative Analysis of the Adsorption Kinetics of the Methylene Blue Dye on Graphene Aerogel and Activated, Coconut Carbon. Adv. Mater. Technol. 2020, 4, 21–28. [Google Scholar] [CrossRef]
  27. Jiang, L.; Hu, X.; Yang, B.; Yang, Z.; Lu, C. Preparation of porous diethylene triamine reduced graphene oxide aerogel for efficient pollutant dye adsorption. J. Porous Mater. 2023. [Google Scholar] [CrossRef]
  28. Nguyen, T.N.; Nguyen, T.T.; Tran, H.T.; Nguyen, M.D.; Nguyen, H.H. Synthesis and application of graphene aerogel as an adsorbent for water treatment. Vietnam J. Sci. Technol. Eng. 2022, 61, 23–28. [Google Scholar] [CrossRef]
  29. Chen, L.; Diao, H.; Shu, Q.; Yang, T. A Graphene-Based Aerogel Was Prepared as Solid Adsorbent for the Enrichment of Platinum (IV) at Trace Concentration. Adv. Mater. Phys. Chem. 2023, 13, 17–29. [Google Scholar] [CrossRef]
  30. Mi, X.; Huang, G.; Xie, W.; Wang, W.; Liu, Y.; Gao, J. Preparation of graphene oxide aerogel and its adsorption for Cu2+ ions. Carbon 2012, 50, 4856–4864. [Google Scholar] [CrossRef]
  31. Kondratowicz, I.; Żelechowska, K.; Nadolska, M.; Jażdżewska, A.; Gazda, M. Comprehensive study on graphene hydrogels and aerogels synthesis and their ability of gold nanoparticles adsorption. Colloids Surf. A-Physicochem. Eng. Asp. 2017, 528, 65–73. [Google Scholar] [CrossRef]
  32. Huo, J.B.; Yu, G. Poly(vinyl) Alcohol-Assisted Fabrication of Magnetic Reduced Graphene Oxide Aerogels and their Adsorption Performance for Cd(II) and Pb(II). Water Air Soil Pollut. 2023, 234, 231. [Google Scholar] [CrossRef]
  33. Korepanov, V.I.; Kabachkov, E.N.; Baskakov, S.A.; Shul’ga, Y.M. Raman Spectra of Composite Aerogels of Polytetrafluoroethylene and Graphene Oxide. Russ. J. Phys. Chem. 2020, 94, 2250–2254. [Google Scholar] [CrossRef]
  34. Zhang, S.; Wang, H.; Liu, J.; Bao, C. Measuring the specific surface area of monolayer graphene oxide in water. Mater. Lett. 2020, 261, 127098. [Google Scholar] [CrossRef]
  35. Grigoriev, S.; Smirnov, A.; Pinargote, N.W.S.; Yanushevich, O.; Kriheli, N.; Kramar, O.; Pristinskiy, Y.; Peretyagin, P. Evaluation of Mechanical and Electrical Performance of Aging Resistance ZTA Composites Reinforced with Graphene Oxide Consolidated by SPS. Materials 2022, 15, 2419. [Google Scholar] [CrossRef] [PubMed]
  36. McAllister, M.J.; Li, J.-L.; Adamson, D.H.; Schniepp, H.C.; Abdala, A.A.; Liu, J.; Herrera-Alonso, M.; Milius, D.L.; Car, R.; Prud’homme, R.K.; et al. Single Sheet Functionalized Graphene by Oxidation and Thermal Expansion of Graphite. Chem. Mater. 2007, 19, 4396–4404. [Google Scholar] [CrossRef]
  37. Graf, D.; Molitor, F.; Ensslin, K.; Stampfer, C.; Jungen, A.; Hierold, C.; Wirtz, L. Spatially Resolved Raman Spectroscopy of Single- and Few-Layer Graphene. Nano Lett. 2007, 7, 238–242. [Google Scholar] [CrossRef]
  38. Dato, A.; Radmilovic, V.; Lee, Z.; Phillips, J.; Frenklach, M. Substrate-Free Gas-Phase Synthesis of Graphene Sheets. Nano Lett. 2008, 8, 2012–2016. [Google Scholar] [CrossRef]
  39. Shen, Y.; Zhu, X.; Chen, B. Size effects of graphene oxide nanosheets on the construction of three-dimensional graphene-based macrostructures as adsorbents. J. Mater. Chem. A 2016, 4, 12106–12118. [Google Scholar] [CrossRef]
  40. Ciszewski, M.; Szatkowska, E.; Koszorek, A.; Majka, M. Carbon aerogels modified with graphene oxide, graphene and CNT as symetric supercapacitor electrodes. J. Mater. Sci. Mater. Electron. 2017, 28, 4897–4903. [Google Scholar] [CrossRef]
  41. Alruwais, R.S.; Adeosun, W.A.; Marwani, H.M.; Jawaid, M.; Asiri, A.M.; Khan, A. Novel Aminosilane (APTES)-Grafted Polyaniline@Graphene Oxide (PANI-GO) Nanocomposite for Electrochemical Sensor. Polymers 2021, 13, 2562. [Google Scholar] [CrossRef] [PubMed]
  42. Aliyev, E.; Filiz, V.; Khan, M.M.; Lee, Y.J.; Abetz, C.; Abetz, V. Structural Characterization of Graphene Oxide: Surface Functional Groups and Fractionated Oxidative Debris. Nanomaterials 2019, 9, 1180. [Google Scholar] [CrossRef] [PubMed]
  43. Cham Sa-Ard, W.; Fawcett, D.; Fung, C.C.; Chapman, P.; Rattan, S.; Poinern, G.E.J. Synthesis, characterisation and thermo-physical properties of highly stable graphene oxide-based aqueous nanofluids for potential low-temperature direct absorption solar ap, plications. Sci. Rep. 2021, 11, 16549. [Google Scholar] [CrossRef] [PubMed]
  44. Balaji, A.; Yang, S.; Wang, J.; Zhang, J. Graphene oxide-based nanostructured DNA sensor. Biosensors 2019, 9, 74. [Google Scholar] [CrossRef]
  45. Hanifah, M.F.; Jaafar, J.; Aziz, M.; Ismail, A.F.; Rahman, M.A.; Othman, M.H. Synthesis of graphene oxide nanosheets via modified hummers’ method and its physicochemical properties. J. Teknol. 2015, 74, 195–198. [Google Scholar] [CrossRef]
  46. Stroe, M.; Cristea, M.; Matei, E.; Galatanu, A.; Cotet, L.C.; Pop, L.C.; Baia, M.; Danciu, V.; Anghel, I.; Baia, L.; et al. Optical Properties of Composites Based on Graphene Oxide and Polystyrene. Molecules 2020, 25, 2419. [Google Scholar] [CrossRef] [PubMed]
  47. García-Bordejé, E.; Víctor-Román, S.; Sanahuja-Parejo, O.; Benito, A.M.; Maser, W.K. Control of the microstructure and surface chemistry of graphene aerogels via pH and time manipulation by a hydrothermal method. Nanoscale 2018, 10, 3526–3539. [Google Scholar] [CrossRef] [PubMed]
  48. Isaeva, V.I.; Vedenyapina, M.D.; Kurmysheva, A.Y.; Weichgrebe, D.; Nair, R.R.; Nguyen, N.P.T.; Kustov, L.M. Modern Carbon–Based Materials for Adsorptive Removal of Organic and Inorganic Pollutants from Water and Wastewater. Molecules 2021, 26, 6628. [Google Scholar] [CrossRef]
  49. Lagergren, S. About the theory of so-called adsorption of soluble substances. Sven. Vetenskapsakad. Handingarl 1898, 24, 1–39. [Google Scholar]
  50. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Proc. Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  51. Zeldowitsch, J. The catalytic oxidation of carbon monoxide on manganese dioxide. Acta Physicochim. URSS 1934, 1, 364–449. [Google Scholar]
  52. Weber, W.J.; Morriss, J.C. Kinetics of adsorption on carbon from solution. J. Sanit Eng. Div. Am. Soc. Civ. Eng. 1963, 89, 31–60. [Google Scholar] [CrossRef]
  53. Touihri, M.; Guesmi, F.; Hannachi, C.; Hamrouni, B.; Sellaoui, L.; Badawi, M.; Poch, J.; Fiol, N. Single and simultaneous adsorption of Cr(VI) and Cu (II) on a novel Fe3O4/pine cones gel beads nanocomposite: Experiments, characterization and isotherms, modeling. Chem. Eng. J. 2021, 416, 129101. [Google Scholar] [CrossRef]
  54. Toor, M.; Jin, B. Adsorption Characteristics, Isotherm, Kinetics, and Diffusion of Modified Natural Bentonite for Removing Diazo Dye. Chem. Eng. J. 2012, 187, 79. [Google Scholar] [CrossRef]
  55. Ding, L.; Lu, X.; Deng, H.; Zhang, X. Adsorptive Removal of 2,4-Dichlorophenoxyacetic Acid (2,4-D) from Aqueous Solutions Using MIEX Resin. Ind. Eng. Chem. Res. 2012, 51, 11226–11235. [Google Scholar] [CrossRef]
  56. George, L.Y.; Ma, L.; Zhang, W.; Yao, G. Parametric modelling and analysis to optimize adsorption of Atrazine by MgO/Fe3O4-synthesized porous carbons in water environment. Environ. Sci. Eur. 2023, 35, 21. [Google Scholar] [CrossRef]
  57. Langmuir, I. The constitution and fundamental properties of solids and liquids. PART I.SOLIDS. J. Frankl. Inst. 1916, 184, 102–105. [Google Scholar]
  58. Gamal, R.; Rizk, S.E.; El-Hefny, N.E. The adsorptive removal of Mo(VI) from aqueous solution by a synthetic magnetic chromium ferrite nanocomposite using a nonionic surfactant. J. Alloys Comp. 2021, 853, 157039. [Google Scholar] [CrossRef]
  59. Srivastava, V.; Sharma, Y.C.; Sillanpää, M. Application of nano-magnesso ferrite (n-MgFe2O4) for the removal of Co2+ ions from synthetic wastewater: Kinetic, equilibrium and thermodynamic studies. Appl. Surf. Sci. 2015, 338, 42–54. [Google Scholar] [CrossRef]
  60. Liu, W.; Yang, Q.; Yang, Z.; Wang, W. Adsorption of 2,4-D on magnetic graphene and mechanism study. Colloids Surf. A-Physicochem. Eng. Asp. 2016, 509, 367–375. [Google Scholar] [CrossRef]
  61. Yılmaz, Ş.; Zengin, A.; Şahan, T.; Gubbuk, I.H. Efficient Removal of 2,4-Dichlorophenoxyacetic Acid from Aqueous Medium Using Polydopamine/Polyacrylamide Co-deposited Magnetic Sporopollenin and Optimization with Response Surface Methodology Approach. J. Polym. Environ. 2023, 31, 36–49. [Google Scholar] [CrossRef]
  62. Deokar, S.K.; Jadhav, A.R.; Pathak, P.D.; Mandavgane, S.A. Biochar from microwave pyrolysis of banana peel: Characterization and utilization for removal of benzoic and salicylic acid from aqueous solutions. Biomass Convers. Biorefin. 2022. [Google Scholar] [CrossRef]
  63. Farghal, H.H.; Nebsen, M.; El-Sayed, M.M.H. Multifunctional Chitosan/Xylan-Coated Magnetite Nanoparticles for the Simultaneous Adsorption of the Emerging Contaminants Pb(II), Salicylic Acid, and Congo Red Dye. Water 2023, 15, 829. [Google Scholar] [CrossRef]
Figure 1. The nitrogen adsorption-desorption isotherms of GOA (a), rGOA (b), and pore size distribution of GOA (c).
Figure 1. The nitrogen adsorption-desorption isotherms of GOA (a), rGOA (b), and pore size distribution of GOA (c).
Gels 09 00680 g001
Figure 2. Raman spectra of GOA and rGOA.
Figure 2. Raman spectra of GOA and rGOA.
Gels 09 00680 g002
Figure 3. FT-IR spectra of GOA and rGOA.
Figure 3. FT-IR spectra of GOA and rGOA.
Gels 09 00680 g003
Figure 4. SEM images of GOA (a) and rGOA (b).
Figure 4. SEM images of GOA (a) and rGOA (b).
Gels 09 00680 g004
Figure 5. Effect of contact time on 2,4-D (a) and SA (b) adsorption on GOA. Dots are experimentally obtained data for various C0; lines are the results of model calculations.
Figure 5. Effect of contact time on 2,4-D (a) and SA (b) adsorption on GOA. Dots are experimentally obtained data for various C0; lines are the results of model calculations.
Gels 09 00680 g005aGels 09 00680 g005b
Figure 6. Effect of contact time on 2,4-D (a) and SA (b) adsorption on rGOA. Dots are experimentally obtained data for various C0; lines are the results of model calculations.
Figure 6. Effect of contact time on 2,4-D (a) and SA (b) adsorption on rGOA. Dots are experimentally obtained data for various C0; lines are the results of model calculations.
Gels 09 00680 g006aGels 09 00680 g006b
Figure 7. Intraparticle diffusion model of 2,4-D on GOA (a) and rGOA (b); SA on GOA (c) and rGOA (d).
Figure 7. Intraparticle diffusion model of 2,4-D on GOA (a) and rGOA (b); SA on GOA (c) and rGOA (d).
Gels 09 00680 g007aGels 09 00680 g007b
Figure 8. Plots of nonlinear isotherm models for the adsorption of 2,4-D (a) and SA (b) adsorption on GOA.
Figure 8. Plots of nonlinear isotherm models for the adsorption of 2,4-D (a) and SA (b) adsorption on GOA.
Gels 09 00680 g008
Figure 9. Plots of nonlinear isotherm models for the adsorption of 2,4-D (a) and SA (b) adsorption on rGOA.
Figure 9. Plots of nonlinear isotherm models for the adsorption of 2,4-D (a) and SA (b) adsorption on rGOA.
Gels 09 00680 g009
Figure 10. GOA (a) and rGOA (b) samples.
Figure 10. GOA (a) and rGOA (b) samples.
Gels 09 00680 g010
Table 1. PFO and PSO kinetic models for the removal of 2,4-D and SA using GOA and rGOA.
Table 1. PFO and PSO kinetic models for the removal of 2,4-D and SA using GOA and rGOA.
C0, mg/LPseudo-First OrderPseudo-Second Order
2,4-DSA2,4-DSA
k1qe1, mg/gR2k1qe1, mg/gR2k2qe2, mg/gR2k2qe2, mg/gR2
GOA
250.059.620.9850.059.950.9880.00810.40.990.00710.790.988
500.0416.450.980.0518.770.9910.00417.930.9950.00420.370.993
750.0422.690.9850.0427.40.990.00324.810.9970.00229.930.992
1000.0827.270.9790.0435.450.9860.00628.130.9340.00238.680.994
1250.0533.190.9880.0642.770.9920.00235.930.9960.00245.730.993
rGOA
250.042.830.9860.053.780.9840.0163.130.9930.0174.130.991
500.033.880.9740.045.280.9740.0114.290.990.0145.780.996
750.044.930.9840.056.890.9780.0125.390.9910.0127.410.994
1000.055.920.9630.088.370.9850.0136.380.9920.0168.840.996
1250.076.880.9860.19.620.9810.0177.320.9960.0210.060.998
Table 2. Elovich kinetic model for the removal of 2,4-D and SA using GOA and rGOA.
Table 2. Elovich kinetic model for the removal of 2,4-D and SA using GOA and rGOA.
C0, mg/LElovich Equation
2,4-DSA
α, mg/(g⋅min)β, g/mgR2α, mg/(g⋅min)β, g/mgR2
GOA
259.3750.7470.9319.1890.7110.931
509.4140.4010.94114.9990.370.933
7510.8280.2810.95412.7280.2330.937
10024.8640.3610.94719.4110.1840.943
12534.9250.2170.944132.350.1970.939
rGOA
250.7542.0290.9442.1811.7390.939
500.9311.4540.9562.3951.1970.963
752.5631.310.93911.3721.1250.951
1008.0141.2710.95618.4751.2610.958
12551.8891.3650.95328.9911.3850.976
Table 3. Intraparticle diffusion constants of adsorption SA using GOA and rGOA.
Table 3. Intraparticle diffusion constants of adsorption SA using GOA and rGOA.
C0 SA, mg/LIPD
kp1, mg/g min1/2C1, mg/gR2kp2, mg/g min1/2C2, mg/gR2kp3, mg/g min1/2C3, mg/gR2
GOA
251.475−0.1590.9890.4845.3330.8910.069.1220.75
502.574−0.0360.9991.0368.9670.9390.06517.9090.867
753.531−0.0040.9981.28913.8710.9490.03527.1290.536
1004.869−0.2460.9991.72817.2840.9330.05535.1810.683
1256.3990.5690.9961.58628.1010.9190.07641.9860.653
rGOA
250.531−0.0450.9880.1791.8880.8740.0123.6490.392
500.7240.0070.9990.3052.0750.9680.0254.9870.588
751.0490.0520.9960.3893.2060.980.0116.8270.237
1001.4820.0980.9920.3685.1730.9870.0128.3070.242
1251.8370.2320.9840.2717.1880.8270.0119.6310.385
Table 4. Intraparticle diffusion constants of adsorption 2,4-D using GOA and rGOA.
Table 4. Intraparticle diffusion constants of adsorption 2,4-D using GOA and rGOA.
C0 2,4-D, mg/LIPD
kp1, mg/g min1/2C1, mg/gR2kp2, mg/g min1/2C2, mg/gR2kp3, mg/g min1/2C3, mg/gR2
GOA
251.396−0.0020.9990.7812.8210.9940.0399.0780.252
502.215−0.0270.9990.758.6680.9740.04715.9130.432
752.9730.0770.9991.4368.6270.9960.13420.8470.853
1004.5530.6340.9891.11316.9490.9110.17725.1610.867
1254.824−0.0420.99951.11721.2230.9740.04833.090.331
rGOA
250.35574−0.0430.990.1261.3880.9110.0112.6880.635
500.451770.0250.9910.1691.8730.9590.0043.9090.599
750.67995−0.0610.9930.172.9960.968−0.0035.1020.119
1000.89790.0820.9860.2862.9630.9780.0036.0350.698
1251.128170.0950.9880.2674.4320.9980.0126.8040.466
Table 5. Isotherm parameters for 2,4-D and SA adsorption on GOA and rGOA.
Table 5. Isotherm parameters for 2,4-D and SA adsorption on GOA and rGOA.
Model ParametersGOArGOA
2,4-DSA2,4-DSA
Langmuirqm, mg/g42.6357.6110.9216.01
bl, L/g0.080.170.0150.015
R20.9820.980.9870.99
FreundlichKF, mg/g6.8611.90.550.73
1/n0.430.470.540.56
R20.9860.9920.9910.995
TemkinbT, J/mol333.43241.211058.05733.7
A, L/g1.643.10.160.16
R20.9630.9660.9830.984
Table 6. Comparison of capacity values for 2,4-D and SA ions sorbed by different adsorbents.
Table 6. Comparison of capacity values for 2,4-D and SA ions sorbed by different adsorbents.
AdsorbentAdsorbateqe, mg/gEquilibrium Time, minRef.
Magnetic Fe3O4@graphene nanocomposite2,4-D32.31720[60]
Polydopamine/polyacrylamide co-deposited magnetic
sporopollenin
62.2-[61]
Algal magnetic activated carbon nanocomposite60.61-[1]
GOA42.63150This work
Chitosan-acrylamide surface molecularly imprinted hydrogelSA44.8720[4]
Biochar36.39720[62]
Chitosan/xylan-coated magnetite nanoparticles13.49120[63]
GOA57.61100This work
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

Kurmysheva, A.Y.; Yanushevich, O.; Krikheli, N.; Kramar, O.; Vedenyapina, M.D.; Podrabinnik, P.; Solís Pinargote, N.W.; Smirnov, A.; Kuznetsova, E.; Malyavin, V.V.; et al. Adsorption Ability of Graphene Aerogel and Reduced Graphene Aerogel toward 2,4-D Herbicide and Salicylic Acid. Gels 2023, 9, 680. https://doi.org/10.3390/gels9090680

AMA Style

Kurmysheva AY, Yanushevich O, Krikheli N, Kramar O, Vedenyapina MD, Podrabinnik P, Solís Pinargote NW, Smirnov A, Kuznetsova E, Malyavin VV, et al. Adsorption Ability of Graphene Aerogel and Reduced Graphene Aerogel toward 2,4-D Herbicide and Salicylic Acid. Gels. 2023; 9(9):680. https://doi.org/10.3390/gels9090680

Chicago/Turabian Style

Kurmysheva, Alexandra Yu., Oleg Yanushevich, Natella Krikheli, Olga Kramar, Marina D. Vedenyapina, Pavel Podrabinnik, Nestor Washington Solís Pinargote, Anton Smirnov, Ekaterina Kuznetsova, Vladislav V. Malyavin, and et al. 2023. "Adsorption Ability of Graphene Aerogel and Reduced Graphene Aerogel toward 2,4-D Herbicide and Salicylic Acid" Gels 9, no. 9: 680. https://doi.org/10.3390/gels9090680

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