Next Article in Journal
Characterization of Caseinate–Carboxymethyl Chitosan-Based Edible Films Formulated with and without Transglutaminase Enzyme
Next Article in Special Issue
Stability Enhancement of Laser-Scribed Reduced Graphene Oxide Electrodes Functionalized by Iron Oxide/Reduced Graphene Oxide Nanocomposites for Nitrite Sensors
Previous Article in Journal
An Efficient Method to Determine the Thermal Behavior of Composite Material with Loading High Thermal Conductivity Fillers
Previous Article in Special Issue
Green Reduction of Graphene Oxide Involving Extracts of Plants from Different Taxonomy Groups
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanocomposites of Copper Trimesinate and Graphene Oxide as Sorbents for the Solid-Phase Extraction of Organic Dyes

by
Igor E. Uflyand
*,
Victoria N. Naumkina
and
Vladimir A. Zhinzhilo
Department of Chemistry, Southern Federal University, 344090 Rostov-on-Don, Russia
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2022, 6(7), 215; https://doi.org/10.3390/jcs6070215
Submission received: 28 June 2022 / Revised: 12 July 2022 / Accepted: 17 July 2022 / Published: 20 July 2022
(This article belongs to the Special Issue Graphene Oxide Composites)

Abstract

:
A nanocomposite based on graphene oxide and copper trimesinate was obtained by the in situ method. The samples have permanent porosity and a microporous structure with a large surface area corresponding to the adsorption type I. A study of the adsorption properties of the obtained composites with respect to organic dyes (malachite green, indigo carmine, brilliant green, Rose Bengal, crystal violet) showed that adsorption largely depends on the content of graphene oxide in the composites. The complex is an effective sorbent for the extraction of cationic and neutral organic dyes when the content of graphene oxide in the nanocomposite is 20% of the calculated copper trimesinate due to electrostatic forces of interaction. For anionic dyes, the maximum adsorption is achieved when using a composite containing 5% graphene oxide due to the predominance of physical sorption. Experimental results show that the obtained sorbent can be used for extraction in a wide pH range, illustrating the excellent pH window offered by this adsorbent. Kinetics data were properly fitted with the pseudo-second-order model. Equilibrium data were best correlated with the Freundlich model. The process was endothermic and spontaneous in nature. The composite makes it possible to achieve a maximum sorption of 393 mg/g, which is a sufficiently high value for the absorption of dyes.

1. Introduction

Organic dyes, widely used in the printing, textile, paper, paint and varnish industries and in the plastics industry, are recognized manufactured pollutants that can slow down and change the chemistry of the photosynthesis cycle. In addition, they contribute to the destabilization of metabolic reactions in plant organisms and can cause mutagenic and carcinogenic diseases in humans and animals [1,2,3,4]. Among the developed methods for the removal of dyes from wastewater, ion-exchange filtration [5], membrane filtration [6], chemical oxidation [7], photocatalysis [8,9,10], catalytic degradation [11,12], electrochemical oxidation [13,14], biodegradation [15,16] and coagulation/flocculation [17,18,19] should be noted. The main disadvantages of most traditional methods are a low selectivity, a low rate of decomposition of dyes and a high energy consumption [20,21]. The adsorption is the most efficient, cheap, flexible, applicable, compatible, eco-friendly, operable, rapid, reusable and widely used technique compared to other methods for removing organic dyes [22]. Therefore, researchers have devoted considerable attention to the search for new adsorbents as advanced materials for the effective removal of dyes.
In recent decades, the attention of scientists has been focused on a class of polyfunctional materials called metal-organic frameworks (MOFs), which can be used as solid-state adsorbents for organic dyes [23]. They are characterized by a high chemical resistance and a significant specific internal surface hundreds of times greater than carbon materials (up to 10,000 m2/g). Among the various MOFs, MOFs such as Cu3(BTC)2(H2O)3·xH2O (where H3BTC = 1,3,5-benzenetricarboxylic or trimesic acid) containing the binuclear Cu(II) paddlewheel structures are of particular interest for testing the adsorption capacity. This MOF was first synthesized in 1999 [24] and designated as HKUST-1 (Hong Kong University of Science and Technology) or MOF-199. It has a high specific surface area and an interconnected 3D pore system with pore sizes of 9 Å × 9 Å [25,26]. Various Cu–BTC complexes were studied, including Cu3(BTC)2·3H2O [27,28], Cu2(OH)(BTC)(H2O)2·nH2O [29] and Cu(BTC–H2)2(H2O)2·3H2O [30].
However, MOFs have some limitations associated with high selectivity, easily lose their adsorption activity in a humid atmosphere and are difficult to post-synthetic functionalize. To expand the practical application of MOFs, a promising approach is to create their composites with other materials, which makes it possible to obtain polyfunctional materials with improved technological properties, which, in addition to the physical sorption provided by the network structure of MOFs, retain adsorbate molecules using chemisorption processes. Such materials are graphene-based compounds, since separated and functionalized graphene sheets will increase dispersion interactions, while MOFs will increase the pore space in which adsorbates can be stored [31,32].
The purpose of the work was to obtain a nanocomposite material based on a copper(II) trimesinate and graphene oxide (GO) and to study the possibilities of such a material for the solid-phase extraction of organic dyes. The objectives of the study were the synthesis of a hybrid composite material based on copper trimesinate and GO, the selection of MOF:GO ratios to obtain the most effective material, the study of the ability of the obtained material regarding the solid-phase extraction of model pollutants and the study of the influence of various variables (pH, initial dye concentration, temperature and contact time) on the process of solid-phase extraction.

2. Materials and Methods

2.1. Materials

Methanol (MeOH, 99.9%), ethanol (EtOH, 98%) and methylene chloride were supplied by Sigma-Aldrich (Moscow, Russia); the purity of the solvents was determined by selective gas chromatography. Copper(II) sulfate pentahydrate (CuSO4·5H2O, ≥99.0%), 1,3,5-benzenetricarboxylic acid (H3BTC, 95.0%), NaOH, H2SO4, NaNO3, KMnO4 and H2O2 were purchased from Sigma-Aldrich (Moscow, Russia) and used without further purification. As a starting material, we used S-1 flaky graphite (technical grade, Lenreaktiv, Russia) without additional purification. The graphite particle size was 125–1000 µm (mean size 200 µm).

2.2. Characterization Techniques

The Fourier transform IR (FTIR) spectra were taken with a Perkin Elmer Spectrum 100 FTIR spectrometer (Perkin Elmer) from KBr pellets, using Softspectra software for the data analysis (Shelton, CT). Scanning electron microscopic (SEM) images were taken with a ZEISS Crossbeam 340 device (Carl Zeiss) at an accelerating voltage of 3 kV. Secondary electrons were detected with an Everhart-Thornley detector (SE2). The distribution of chemical elements on the surface of the samples was determined by X-ray energy dispersive microanalysis (EDX) on an Oxford X-max 80 microanalyzer with an electron probe energy of ≤10 keV. The nitrogen adsorption/desorption experiments were performed at 77 K (liquid N2) using the AUTOSORB-1 system (Quantachrome, Boynton Beach, FL, USA) by the static volumetric method; before the analysis, the samples were degassed by heating at 150 °C for 12 h in vacuum. The Brunauer–Emmett–Teller surface area (SBET) was obtained from the amount of N2 physically sorbed at various relative pressures (P/P0) based on the linear part of the six-point adsorption data at P/P0 = 0.02–0.10. The total pore volume (Vtotal) was calculated by the Horvath Kawazoe method at P/P0 = 0.99. The micropore volume (Vmicro) was obtained by the Barrett–Joyner–Halenda adsorption and the t-plot methods, respectively. To assess the adsorption capacity, ultrahigh purity gases (99.995%) were used.

2.3. Synthesis of the Sorbent

The composite was synthesized in situ in two stages: obtaining GO by the modified Hummers method, and the subsequent introduction of copper trimesinate into the synthesis procedure.

2.3.1. Synthesis of Graphene Oxide

GO was prepared by the Hummers method [33] in accordance with the previously developed protocol [34]. A total of 15 g of finely ground sodium nitrate powder was added to 115 mL of concentrated sulfuric acid at a temperature not exceeding 0 °C. The mixture was stirred until completely dissolved. A total of 5 g of crushed graphite was added to the mixture, stirred until homogeneous and then carefully oxidized, introducing 15 g of potassium permanganate in small portions so that the temperature does not rise above 20 °C. Then, 230 mL of distilled water was added in small portions, making sure that the temperature of the system does not exceed 98 °C. The rest of the potassium permanganate is reduced with a concentrated solution of hydrogen peroxide. As a result, a thin suspension of GO in water was obtained, stable for a month (Figure 1). GO was purified by dialysis.

2.3.2. Synthesis of Nanocomposite

First, a soluble form of 1,3,5-benzenetricarboxylic acid is prepared as the sodium salt. Then, 3.48 g (0.0166 mol) of 1,3,5-benzenetricarboxylic acid was added to an aqueous solution containing 1.98 g of sodium hydroxide (in 200 mL of water) while heating to 80 °C. Then, 0.25 g of GO and an aqueous solution of 6.22 g (0.0250 mol) of copper sulfate are added to the resulting solution of sodium trimesinate. The precipitate that formed was sonicated in distilled water (60 W, 40 kHz) for 24 h, filtered off and dried in air. At the final stage, it is sequentially kept in absolute alcohol and then in dry ethyl acetate and dichloromethane. The exposure in each solvent was 8 h. At the end of the procedure, the precipitate was centrifuged, washed with dry dichloromethane and dried for 10 h in a dynamic vacuum at 120 °C. GO was introduced into the composite in an amount of 5 wt.% of the calculated copper trimesinate. The yield was 4.56 g (91.23% in terms of copper trimesinate). Similarly, the synthesis of the composite was conducted in the amounts of 10 and 20 wt.% GO of the calculated copper trimesinate. The yield was 4.85 g and 4.74 g (96.90% and 94.71% in terms of copper trimesinate, respectively).

2.4. Dye Adsorption Study

In this work, serial adsorption experiments were conducted to remove dyes from aqueous solutions by using the synthesized sorbents. The following dyes were studied in the work: malachite green (MG), indigo carmine (IC), brilliant green (BG), Rose Bengal (RB) and crystal violet (CV) (Table 1). The dye stock solutions were prepared by dissolving precisely weighed dye powders in distilled water to a concentration of 40 mg/L. Experimental solutions were obtained by diluting the initial dye solution in exact proportions to the required initial concentrations.

2.4.1. Study of the Dependence of Sorption on the Initial Dye Concentration

Solutions of dyes with a volume of 40 mL and with concentrations of 20, 10, 5, 2.5 and 1.25 mg/L were placed in a 100 mL beaker and thermostated at 278, 293 and 308 K on a magnetic stirrer, adjusting the rotation speed so that the mixing was effective but air was not drawn into the liquid phase. When the solutions reached a predetermined temperature, a 10 mg sorbent was introduced. Upon reaching 3 h, 10 mL of the sorbent suspension was taken from each dye solution and subjected to centrifugation for 5 min. The residual dye concentration was determined in the filtrate by using a spectrophotometer at specific wavelengths for each dye.
The adsorption capacities (q (mg/g)) were evaluated by the following equations:
qt = (C0Ct)V/m
qe = (C0Ce)V/m
where qt and qe are the dye amount (mg/g) adsorbed on the sorbent at time t and in equilibrium, respectively; C0, Ct and Ce are the dye concentrations in the solution (mg/L) at the initial stage at time t and in equilibrium, respectively; m (g) and V (L) are the sorbent dosage and the volume of the dye solution, accordingly.
The degree of adsorption R (%) (extraction ratio of the sorbate) was calculated by the formula:
R (%) = (C0Ce)/C0 × 100%
The data obtained during the experiment were analyzed using the Langmuir and Freundlich models.

2.4.2. Study of the Dependence of Sorption on Contact Time

Solutions of dyes with a volume of 40 mL and with a concentration of 20 mg/L were placed in a 100 mL beaker and thermostated at 278, 293 and 308 K on a magnetic stirrer, adjusting the rotation speed so that the mixing was effective but air was not drawn into the liquid phase. When the solutions reached a predetermined temperature, a 20 mg sorbent was introduced, and a timer was started. Every 5, 10, 15, 30, 45 and 60 min, 10 mL of the sorbent suspension was taken from each dye solution and subjected to centrifugation for 5 min. The residual dye concentration was determined in the filtrate by using a spectrophotometer at specific wavelengths for each dye.
The data obtained during the study of the adsorption process were analyzed using kinetic models of pseudo-first and pseudo-second order reactions [41,42]. The pseudo-first order rate equation is:
ln (qeqt) = ln qek1t,
where k1 (min−1) is the adsorption rate constant of the pseudo first order model.
The differential form of the classical pseudo-second order rate equation is:
d q t d t = k 2 q e q t 2
where k2 is the adsorption rate constant of the pseudo-second order model (g/mmol min).
Gibbs free energy (ΔG), standard enthalpy (ΔH) and entropy (ΔS) were calculated from adsorption isotherms at various temperatures. The thermodynamic characteristics were calculated analytically using the equations [42]:
Δ G 0 = R T l n K c
where Kc is the thermodynamic equilibrium constant, which can be determined from the equation:
K c = C t C e
The Gibbs equation can also be expressed as follows:
ΔG0 = ΔH0 − TΔS0

2.4.3. Study of the Dependence of Sorption on the pH

Dye solutions with a volume of 40 mL and a concentration of 20 mg/L were placed in a 100 mL beaker and thermostated at 278, 293 and 308 K on a magnetic stirrer. The pH of the initial solution was adjusted by adding 0.1 N NaOH or HCl solution dropwise. Next, adsorption studies were conducted, as mentioned earlier.

2.4.4. Determining the Number of Work Cycles

A certain volume of a dye solution with a standard concentration is placed in a beaker and thermostated at a given temperature. The calculated mass of the sorbent is added and kept at a given temperature until the equivalence point is reached (at least 3 h). The sorbent is separated by centrifugation, washed with cold distilled water and treated with an eluent (ethanol) at a temperature of 40–50 °C to desorb the dye; the procedure is repeated several times until the color of the eluent is gone. The sorbent is washed repeatedly with water and conditioned. The procedure is repeated 4–5 times.

3. Results

3.1. Synthesis and Characterization of the Sorbent

In this work, we considered the removal of dyes from an aqueous medium using a MOF-GO nanocomposite material obtained by the in situ method. A characteristic feature of this synthesis is the introduction of GO at the stage of MOF synthesis to generate MOF crystals on the surface of GO with a size of up to 100 nm, which leads to the formation of a nanocomposite material.
The initial GO sample was analyzed by elemental analysis, using EDX analysis and conventional combustion to determine the content of the elements. Found (%): C–66.5; O–31.8 (EDX analysis, Figure 2); C–66.4; H–1.72 (burning method).
We have obtained the IR spectra of the key compounds involved in the synthesis of the composite by pressing in a tablet with potassium bromide in the frequency range from 4000 to 400 cm−1 (Figure 3). The IR spectrum of GO (Figure 3a) contains a peak in the region of 3400 cm−1, assignable to stretching vibrations of hydroxy groups of adsorbed water. There are absorption bands in the region of 2927 and 2849 cm−1, characteristic of symmetric and asymmetric vibrations of CH2 groups, an intense band in the region of 1720 cm−1, corresponding to vibrations of the carboxyl group, and a low-intensity band in the region of 1619 cm−1, corresponding to vibrations of C-C bonds which are probably due to a slight admixture of unoxidized graphite [43]. These data agree well with the results of studying laboratory and commercial samples of graphene oxide [44,45]. In the IR spectrum of copper trimesinate (Figure 3c), one can see the formation of a strong absorption peak in the region of 760 cm−1, which can be interpreted as the formation of a Me-O bond. This proves that deprotonation of carboxyl groups occurred with the formation of a Cu–O bond. However, in the composite, such a peak is shifted by 20 cm−1 to the long wavelength region. Characteristically, in copper trimesinate, the intense band of pure trimesic acid in the region of 1720 cm−1, corresponding to the carboxyl group in the non-ionized form, completely disappears. This allows us to assert that there are no free carboxyl groups in the obtained complex; all three carboxyl groups are deprotonated and form bonds with copper atoms. In the IR spectrum of copper trimesinate, new bands appear at 1375 and 1560 cm−1, which are characteristic of symmetric and asymmetric vibrations of carboxy groups. Since Δν = 185 cm−1, it can be argued that the carboxylate ion can exhibit the character of both a monodentate and a bidentate ligand. In the composite, these peaks are preserved. At the same time, GO introduces into the composite a wide strong band with a peak in the region of 3400 cm−1, which can be attributed to stretching vibrations of adsorbed water hydroxo groups. The intense band at 1740 cm−1 corresponds to vibrations of carboxyl groups and epoxy carbon atoms, which are also contributed by GO. The absorption band at 1625 cm−1 refers to bending vibrations of hydroxo groups. Intense absorption bands at 1200, 1100 and 850 cm−1 can also characterize the formation of epoxy groups. The IR spectrum of the resulting composite includes a set of spectra of GO and copper trimesinate, regardless of the amount of GO (Figure 3d).
Since porosity is an important parameter of porous materials as dye adsorbents, the specific surface area, total pore volume and pore size distribution of MOF and MOF-GO composites were analyzed using N2 adsorption-desorption isotherms at 77 K. All samples show typical type I isotherms according to the IUPAC classification, suggesting the predominance of micropores existing in the materials. A high ratio of micropore volume to total pore volume (Vmicro/Vtotal) above 70% can also confirm this. As can be seen from Table 2, MOF has a large surface area of 1562 m2/g and a pore volume of 0.85 cm3/g. GO is not porous when nitrogen is used as a molecular probe [46], and a low content of GO in composites leads to a slight decrease in the surface area of MOFs [31]. In our work, a significant increase in the GO content leads to a sharp decrease in the surface area and the average pore radius in all three composites due to the structure of the coating. At the same time, Vtotal and Vmicro practically do not change, which indicates the same microporous nature of MOFs and composites.
The composite does not form as dense of a structure as the initial MOF does, but it has a layered structure irrespective of the amount of the introduced GO (Figure 4). GO sheets are directly adjacent to copper trimesinate crystals, forming a common system. The diffraction patterns of the composites have a similar shape which does not depend on the GO content and coincides with the previously published data [47].
Based on the data we obtained and the available literature data, the following structure can be assigned to the composites (Figure 5).

3.2. Adsorption of Dyes on the Sorbent

The adsorption isotherm plays a crucial role in analyzing the adsorption capacity of the materials and providing information about the solution–surface interaction. The adsorption capacity of the sorbent with the increase in the initial dye concentration is shown in Figure 6. The adsorbed amounts of the dyes increased with the increase in equilibrium dye concentration. It is evident that a high concentration in the solution implicates a high dye molecule fixed at the surface of the adsorbent. When studying the dependence of the solid-phase extraction on the initial concentration, it is noticeable that, with an increase in the initial concentration of the sorbate, the adsorption increases monotonically.
An analysis of the parameters of the Langmuir adsorption isotherm model (Figure 7) allows us to conclude that this Langmuir model describes the adsorption process well. The correlation coefficients are quite high, and the maximum adsorption calculated by the Langmuir model satisfactorily coincides with the experimental one (Table 3).
An analysis of the Freundlich adsorption isotherms shows that the obtained experimental data satisfactorily agree with this model, since the correlation coefficient is quite high (R2 is more than 0.91) (Figure 8). The values of the Freundlich constants within the framework of the description of this model characterize the resulting composites as sorbents with satisfactory adsorption properties. The values of the coefficient 1/n range from 0 to 1, which indicates favorable conditions for adsorption and is in good agreement with the other experimental data (Table 3).
Contact time is an important design parameter on which the adsorption process depends, as it provides information about the dynamic of the reaction in terms of the order and the rate constant. As shown in Figure 9, the degree of extraction of dyes on the composite increases rapidly during the initial contact time and then slows down. Rapid diffusion to the outer surface is accompanied by diffusion into the pores of the matrix at a significant rate, which ultimately leads to a rapid achievement of equilibrium.
The linear plots of ln(qeqt) versus t for a pseudo-first order reaction and of t/qt versus t for a pseudo-second order reaction for dye adsorption were evaluated to obtain the rate parameters. Kinetic parameters under different conditions were calculated from the graphs (Figure 10). The applied pseudo-first-order equation satisfactorily describes the patterns of adsorption at the initial stages (up to 60 min in the experiment) of the adsorption process, when the phenomenon of film diffusion has a significant effect on the process. Apparently, an increase in the concentration of sorbate molecules at the surface of the adsorbent at the initial moments of the process and at low degrees of the filling of the adsorption space stimulates the movement of sorbate molecules into the pores of the adsorbent under the influence of the concentration gradient, but, subsequently, this process slows down, which in turn potentiates other mechanisms. The calculated values of the determination coefficient make it possible to judge that the pseudo-first-order kinetic adsorption model describes the process less accurately, and, consequently, the pseudo-second-order kinetic model is more preferable (Table 4).
From the data given in Table 5, we can conclude that positive adsorption enthalpies characterize the process as endothermic, except for the processes of solid-phase extraction with a nanocomposite containing 5% GO of the dyes MG and IR. For these dyes, the enthalpy value has negative values, which characterize the processes as exothermic. In order for the organic dye particles to reach the active centers of the adsorbent, it is necessary that they be freed from the hydration shell, or at least part of it, for which energy must be expended. As a result, the calculated value of the enthalpy is the result of two processes—the dehydration of organic dye ions (an endothermic process) and the adsorption of the latter on the active centers of the adsorbent, which proceeds with the release of energy (the heat of adsorption). A positive entropy value indicates an increase in the number of degrees of freedom at the solid–liquid interface during the dye adsorption process and reflects the sorbate affinity for adsorbent molecules. The activation energies are small values, which indicate the feasibility of the adsorption process under these conditions. In addition, since the values are below 40 kJ/mol, it can be assumed with a reasonable degree of confidence that the limiting process is physical adsorption.
The study of the dependence of solid-phase extraction on the pH of the medium allows us to conclude that the obtained sorbent can be used for extraction in a wide pH range, illustrating the excellent pH window offered by this adsorbent (Table 6).
The possibility of the repeated use of the sorbent for the extraction of dyes is an important parameter. Studies have been conducted on the regeneration of the sorbent after use in the adsorption of dyes. The data obtained (Table 7) show that the sorbent can be used for five cycles without a significant loss of sorption capacity.
The sorbent makes it possible to achieve a maximum sorption of 393 mg/g, which is a sufficiently high value for the absorption of dyes (Table 8). The complex is an effective sorbent for the extraction of cationic and neutral organic dyes when the content of GO in the nanocomposite is 20 wt.% of the calculated copper trimesinate. With an increase in the content of GO, which consists of many carboxyl groups, in the composite in addition to physical sorption on the composite, cationic and neutral dyes are also bound chemically on GO. In the case of anionic dyes, the maximum adsorption is achieved when using a composite containing 5% GO due to the predominance of physical sorption. A comparison of the maximum adsorption values for the studied dyes leads to the conclusion that cationic dyes are adsorbed best. From the obtained metrological characteristics of the adsorption of dyes, presented in Table 8, we can conclude that the accuracy score does not exceed 4%, and the accuracy rate does not exceed 5%.

4. Conclusions

The developed one-stage method for the in situ preparation of composites based on a copper trimesinate and graphene oxide is economical and can be used for large-scale production. The IR spectrum of the resulting composite includes a set of spectra of graphene oxide and copper trimesinate, regardless of the amount of graphene oxide. The diffraction patterns of the composites have similar shapes, which do not also depend on the graphene oxide content. Nanocomposites showed a good ability to adsorb organic dyes, such as malachite green, indigo carmine, brilliant green, Rose Bengal and crystal violet. The complex is an effective sorbent for the extraction of cationic and neutral organic dyes when the content of GO in the nanocomposite is 20 wt.% of the calculated copper trimesinate. In the case of anionic dyes, the maximum adsorption is achieved when using a composite containing 5% GO. The accuracy score does not exceed 4%, and the accuracy rate does not exceed 5%. The sorbent can be used for five cycles without a significant loss of sorption capacity.

Author Contributions

Conceptualization, I.E.U. and V.A.Z.; methodology, I.E.U., V.N.N. and V.A.Z.; writing—original draft, I.E.U. and V.A.Z.; writing—review and editing, I.E.U., V.N.N. and V.A.Z.; supervision, I.E.U. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abe, F.R.; Mendonça, J.N.; Moraes, L.A.; de Oliveira, G.A.; Gravato, C.; Soares, A.M.; de Oliveira, D.P. Toxicological and behavioral responses as a tool to assess the effects of natural and synthetic dyes on zebrafish early life. Chemosphere 2017, 178, 282–290. [Google Scholar] [CrossRef] [PubMed]
  2. Li, J.; Wang, L.; Liu, Y.; Song, Y.; Zeng, P.; Zhang, Y. The research trends of metal-organic frameworks in environmental science: A review based on bibliometric analysis. Environ. Sci. Pollut. Res. 2020, 27, 19265–19284. [Google Scholar] [CrossRef] [PubMed]
  3. Pavithra, K.G.; Kumar, P.S.; Jaikumar, V.; Rajan, P.S. Removal of colorants from wastewater: A review on sources and treatment strategies. J. Ind. Eng. Chem. 2019, 75, 1–19. [Google Scholar] [CrossRef]
  4. Gorbunova, M.O.; Bayan, E.M.; Shevchenko, A.V.; Kulyaginova, M.S. Digital colorimetric determination of chlorides in water using gas extraction and methyl orange. Anal. Kontrol 2017, 21, 274–280. [Google Scholar] [CrossRef] [Green Version]
  5. Khan, M.A.; Khan, M.I.; Zafar, S. Removal of different anionic dyes from aqueous solution by anion exchange membrane. Membr. Water Treat. 2017, 8, 259–277. [Google Scholar] [CrossRef]
  6. Zwane, S.; Masheane, M.L.; Kuvarega, A.T.; Vilakati, G.D.; Mamba, B.B.; Nyoni, H.; Mhlanga, S.D.; Dlamini, D.S. Polyethersulfone/Chromolaena odorata (PES/CO) adsorptive membranes for removal of Congo red from water. J. Water Process. Eng. 2019, 30, 100498. [Google Scholar] [CrossRef]
  7. Xiao, X.Z.; Dai, T.T.; Guo, J.; Wu, J.H. Flowerlike Brochantite Nanoplate Superstructures for Catalytic Wet Peroxide Oxidation of Congo Red. ACS Appl. Nano Mater. 2019, 2, 4159–4168. [Google Scholar] [CrossRef]
  8. Khan, J.; Sayed, M.; Ali, F.; Khan, H.M. Removal of Acid Yellow 17 Dye by Fenton Oxidation Process. Z. Phys. Chem. 2018, 232, 507–525. [Google Scholar] [CrossRef]
  9. Rashed, M.N.; Eltaher, M.A.; Abdou, A.N.A. Adsorption and photocatalysis for methyl orange and Cd removal from wastewater using TiO2/sewage sludge-based activated carbon nanocomposites. R. Soc. Open Sci. 2017, 4, 170834. [Google Scholar] [CrossRef] [Green Version]
  10. Xu, R.; Su, M.; Liu, Y.; Chen, Z.; Ji, C.; Yang, M.; Chang, X.; Chen, D. Comparative study on the removal of different-type organic pollutants on hierarchical tetragonal bismutite microspheres: Adsorption, degradation and mechanism. J. Clean. Prod. 2020, 242, 118366. [Google Scholar] [CrossRef]
  11. Tran, V.A.; Vu, K.B.; Vo, T.-T.-T.; Le, V.T.; Do, H.H.; Bach, L.G.; Lee, S.-W. Experimental and computational investigation on interaction mechanism of Rhodamine B adsorption and photodegradation by zeolite imidazole frameworks-8. Appl. Surf. Sci. 2021, 538, 148065. [Google Scholar] [CrossRef]
  12. Abdellah, M.H.; Nosier, S.A.; El-Shazly, A.H.; Mubarak, A.A. Photocatalytic decolorization of methylene blue using TiO2/UV system enhanced by air sparging. Alex. Eng. J. 2018, 57, 3727. [Google Scholar] [CrossRef]
  13. Garcia-Segura, S.; Ocon, J.D.; Chong, M.N. Electrochemical oxidation remediation of real wastewater effluents—A review. Process Saf. Environ. 2018, 113, 48–67. [Google Scholar] [CrossRef]
  14. Pacheco-Alvarez, M.O.A.; Picos, A.; Perez-Segura, T.; Peralta-Hernandez, J.M. Proposal for highly efficient electrochemical discoloration and degradation of azo dyes with parallel arrangement electrodes. J. Electroanal. Chem. 2019, 838, 195–203. [Google Scholar] [CrossRef]
  15. Gomaa, O.M.; Fapetu, S.; Kyazze, G.; Keshavarz, T. Probing the mechanism of simultaneous bioenergy production and biodegradation process of Congo red in microbial fuel cells. J. Chem. Technol. Biotechnol. 2019, 94, 2092–2097. [Google Scholar] [CrossRef]
  16. Shanmugam, S.; Ulaganathan, P.; Swaminathan, K.; Sadhasivam, S.; Wu, Y.R. Enhanced biodegradation and detoxification of malachite green by Trichoderma asperellum laccase: Degradation pathway and product analysis. Int. Biodeter. Biodegr. 2017, 125, 258–268. [Google Scholar] [CrossRef]
  17. Castañeda-Díaz, J.; Pavón-Silva, T.; Gutiérrez-Segura, E.; Colín-Cruz, A. Electrocoagulation-Adsorption to Remove Anionic and Cationic Dyes from Aqueous Solution by PV-Energy. J. Chem. 2017, 2017, 5184590. [Google Scholar] [CrossRef] [Green Version]
  18. Kristianto, H.; Rahman, H.; Prasetyo, S.; Sugih, A.K. Removal of Congo red aqueous solution using Leucaena leucocephala seed’s extract as natural coagulant. Appl. Water Sci. 2019, 9, 88. [Google Scholar] [CrossRef] [Green Version]
  19. Gadekar, R.; Ahammed, M.M. Coagulation/flocculation process for dye removal using water treatment residuals: Modelling through artificial neural networks. Desalin. Water Treat. 2016, 57, 26392–26400. [Google Scholar] [CrossRef]
  20. Lu, L.; Wu, J.; Wang, J.; Liu, J.Q.; Li, B.H.; Singh, A.; Kumar, A.; Batten, S.R. An uncommon 3D 3,3,4,8-c Cd(II) metal-organic framework for highly efficient luminescent sensing and organic dye adsorption: Experimental and theoretical insight. CrystEngComm 2017, 19, 7057–7067. [Google Scholar] [CrossRef]
  21. Wang, C.K.; Xing, F.F.; Bai, Y.L.; Zhao, Y.M.; Li, M.X.; Zhu, S.R. Synthesis, structure of semi-rigid tetracarboxylate Cu(II) porous coordination polymers and their versatile high efficiency catalytic dye degradation in neutral aqueous solution. Cryst. Growth Des. 2016, 16, 2277–2288. [Google Scholar] [CrossRef]
  22. Wang, X.; Ye, N. Recent advances in metal-organic frameworks and covalent organic frameworks for sample preparation and chromatographic analysis. Electrophoresis 2017, 38, 3059–3078. [Google Scholar] [CrossRef] [PubMed]
  23. Uflyand, I.E.; Zhinzhilo, V.A.; Nikolaevskaya, V.O.; Kharisov, B.I.; Oliva González, C.M.; Kharissova, O.V. Recent strategies to improve MOF performance in solid phase extraction of organic dyes. Microchem. J. 2021, 168, 106387. [Google Scholar] [CrossRef]
  24. Chui, S.S.-Y.; Lo, S.M.-F.; Charmant, J.P.H.; Guy Orpen, A.; Williams, I.D. A Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n. Science 1999, 283, 1148–1150. [Google Scholar] [CrossRef]
  25. Gascon, J.; Aguado, S.; Kapteijn, F. Manufacture of dense coatings of Cu3(BTC)2 (HKUST-1) on α-alumina. Microporous Mesoporous Mater. 2008, 113, 132–138. [Google Scholar] [CrossRef]
  26. Mustafa, D.; Breynaert, E.; Bajpe, S.R.; Martens, J.A.; Kirschhock, C.E.A. Stability improvement of Cu3(BTC)2 metal–organic frameworks under steaming conditions by encapsulation of a Keggin polyoxometalate. Chem. Commun. 2011, 47, 8037–8039. [Google Scholar] [CrossRef] [Green Version]
  27. Seo, Y.K.; Hundal, G.; Jang, I.T.; Hwang, Y.K.; Jun, C.H.; Chang, J.-S. Microwave synthesis of hybrid inorganic–organic materials including porous Cu3(BTC)2 from Cu(II)-trimesate mixture. Microporous Mesoporous Mater. 2009, 119, 331–337. [Google Scholar] [CrossRef]
  28. Chen, J.; Yu, T.; Xiao, H.; Zhou, G.; Weng, L.; Tu, B.; Zhao, D. Synthesis and Structure of a New 3D Porous Cu(II)–Benzene-1,3,5-tricarboxylate Coordination Polymer, [Cu2(OH)(BTC)(H2O)]n·2nH2O. Chem. Lett. 2003, 32, 590–591. [Google Scholar] [CrossRef]
  29. Mao, Y.; Shi, L.; Huang, H.; Yu, Q.; Ye, Z.; Peng, X. Mesoporous separation membranes of {[Cu(BTC–H2)2·(H2O)2]·3H2O} nanobelts synthesized by ultrasonication at room temperature. CrystEngComm 2013, 15, 265–270. [Google Scholar] [CrossRef]
  30. Zornoza, B.; Seoane, B.; Zamaro, J.M.; Téllez, C.; Coronas, J. Combination of MOFs and zeolites for mixed-matrix membranes. ChemPhysChem 2011, 12, 2781–2785. [Google Scholar] [CrossRef]
  31. Uflyand, I.E.; Naumkina, V.N.; Zhinzhilo, V.A. Nanocomposites of Graphene Oxide and Metal-Organic Frameworks. Russ. J. Appl. Chem. 2021, 94, 1453–1468. [Google Scholar] [CrossRef]
  32. Zheng, M.; Xu, L.; Chen, C.; Labiadh, L.; Yuan, B.; Fu, M.-L. MOFs and GO-based composites as deliberated materials for the adsorption of various water contaminants. Separ. Purif. Technol. 2022, 294, 121187. [Google Scholar] [CrossRef]
  33. Hummers, W.S., Jr.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  34. Novikova, A.A.; Burlakova, V.E.; Varavka, V.N.; Uflyand, I.E.; Drogan, E.G.; Irkha, V.A. Influence of glycerol dispersions of graphene oxide on the friction of rough steel surfaces. J. Mol. Liq. 2019, 284, 1–11. [Google Scholar] [CrossRef]
  35. Rauf, M.A. Solvatochromic behavior on the absorption and fluorescence spectra of Rose Bengal dye in various solvents. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2009, 72, 133–137. [Google Scholar] [CrossRef]
  36. Naser, J.A. Adsorption kinetic of malachite green dye from aqueous solutions by electrospun nanofiber Mat. Orient. J. Chem. 2017, 33, 3121–3129. [Google Scholar] [CrossRef]
  37. Latif, I.K.; Abdullah, H.M.; Saleem, M.H. Magnetic conductive hydrogel nanocomposites as drug carrier. Nanosci. Nanotechnol. 2016, 6, 48–58. [Google Scholar]
  38. Baidya, K.S.; Kumar, U. Adsorption of brilliant green dye from aqueous solution onto chemically modified areca nut husk. S. Afr. J. Chem. Eng. 2021, 35, 33–43. [Google Scholar]
  39. Rehman, R.; Muhammad, S.J.; Arshad, M. Brilliant green and acid orange 74 dyes removal from water by Pinus roxburghii leaves in naturally benign way: An application of green chemistry. J. Chem. 2019, 2019, 3573704. [Google Scholar] [CrossRef] [Green Version]
  40. Jaspal, D.; Malhotra, S.; Malviya, A. Column Studies for the Adsorption of Brilliant Green, Fast Green FCF and Phenol Red Dyes on De-Oiled Soya and Bottom Ash. Asian J. Chem. 2012, 24, 5082–5086. [Google Scholar]
  41. Simonin, J.-P. On the comparison of pseudo-first order and pseudo-second order rate laws in the modeling of adsorption kinetics. Chem. Eng. J. 2016, 300, 254–263. [Google Scholar] [CrossRef] [Green Version]
  42. Pehlivan, E.; Arslan, G. Comparison of adsorption capacity of young brown coals and humic acids prepared from different coal mines in Anatolia. J. Hazard. Mater. B 2006, 138, 401–408. [Google Scholar] [CrossRef] [PubMed]
  43. Shahriary, L.; Athawale, A.A. Graphene oxide synthesized by using modified Hummers approach. Int. J. Renew. Energy Environ. Eng. 2014, 2, 58–63. [Google Scholar]
  44. Lian, P.; Zhu, X.; Liang, S.; Li, Z.; Yang, W.; Wang, H. Large reversible capacity of high quality graphene sheets as an anode material for lithium-ion batteries. Electrochim. Acta 2010, 55, 3909–3914. [Google Scholar] [CrossRef]
  45. Manoratne, C.H.; Rosa, S.R.D.; Kottegoda, I.R.M. XRD-HTA, UV Visible, FTIR and SEM interpretation of reduced graphene oxide synthesized from high purity vein graphite. Mater. Sci. Res. India 2017, 14, 19–30. [Google Scholar] [CrossRef]
  46. Seredych, M.; Petit, C.; Tamashausky, A.V.; Bandosz, T.J. Role of graphite precursor in the performance of graphite oxides as ammonia adsorbents. Carbon 2009, 47, 445–456. [Google Scholar] [CrossRef]
  47. Domán, A.; Klébert, S.; Madarász, J.; Sáfrán, G.; Wang, Y.; László, K. Graphene Oxide Protected Copper Benzene-1,3,5-Tricarboxylate for Clean Energy Gas Adsorption. Nanomaterials 2020, 10, 1182. [Google Scholar] [CrossRef]
Figure 1. Suspension of GO in water.
Figure 1. Suspension of GO in water.
Jcs 06 00215 g001
Figure 2. Photomicrograph of a section of the surface of a GO sheet and elemental analysis at this point.
Figure 2. Photomicrograph of a section of the surface of a GO sheet and elemental analysis at this point.
Jcs 06 00215 g002
Figure 3. IR spectra of GO (a), copper trimesinate (b), trimesic acid (c) and nanocomposite (d).
Figure 3. IR spectra of GO (a), copper trimesinate (b), trimesic acid (c) and nanocomposite (d).
Jcs 06 00215 g003
Figure 4. SEM images of the composites of copper trimesinate with 5, 10 and 20% GO, respectively.
Figure 4. SEM images of the composites of copper trimesinate with 5, 10 and 20% GO, respectively.
Jcs 06 00215 g004
Figure 5. Scheme of the location of copper trimesinate between graphene layers in the composite. The Cu atoms are shown in blue; O atoms, red; C atoms, brown; H atoms, pink.
Figure 5. Scheme of the location of copper trimesinate between graphene layers in the composite. The Cu atoms are shown in blue; O atoms, red; C atoms, brown; H atoms, pink.
Jcs 06 00215 g005
Figure 6. Dependence of the extraction on the initial concentration of the sorbate for IC dye.
Figure 6. Dependence of the extraction on the initial concentration of the sorbate for IC dye.
Jcs 06 00215 g006
Figure 7. Langmuir isotherms of the solid-phase extraction of IC dye in non-linearized (a) and linearized (b) form.
Figure 7. Langmuir isotherms of the solid-phase extraction of IC dye in non-linearized (a) and linearized (b) form.
Jcs 06 00215 g007
Figure 8. Freundlich isotherms of the solid-phase extraction of IC dye in non-linearized (a) and linearized (b) form.
Figure 8. Freundlich isotherms of the solid-phase extraction of IC dye in non-linearized (a) and linearized (b) form.
Jcs 06 00215 g008
Figure 9. Degree of extraction (a) and kinetic curves (b) of solid-phase extraction versus contact time for IC dye.
Figure 9. Degree of extraction (a) and kinetic curves (b) of solid-phase extraction versus contact time for IC dye.
Jcs 06 00215 g009
Figure 10. Pseudo-first order (a) and pseudo-second order (b) models of the solid-phase extraction rate for IC dye.
Figure 10. Pseudo-first order (a) and pseudo-second order (b) models of the solid-phase extraction rate for IC dye.
Jcs 06 00215 g010
Table 1. Studied dyes.
Table 1. Studied dyes.
NameMolecular Structureλmax (nm)Ref.
RB Jcs 06 00215 i001546[35]
MG Jcs 06 00215 i002617[36,37]
IC Jcs 06 00215 i003610 (600)[37]
BG Jcs 06 00215 i004625[38,39]
CV Jcs 06 00215 i005590 (594)[40]
Table 2. Characteristics of MOFs, GO and nanocomposites.
Table 2. Characteristics of MOFs, GO and nanocomposites.
SampleSBET (m2/g)Vtotal, (cm3/g)Vmicro (cm3/g)Average Pore Radius (Å)
MOF1562.30.850.6110.9
GO562--6.2
Nanocomposite (5% GO)7620.850.66.5
Nanocomposite (10% GO)7130.850.66.6
Nanocomposite (20% GO)6890.850.66.8
Table 3. Isotherm model constants of dyes by the nanocomposite.
Table 3. Isotherm model constants of dyes by the nanocomposite.
ModelSorbateConstantsT (K)
283293308
Nanocomposite with 5% GO
LangmuirCVqmax (mg/g)90.974.353.1
KL (L/mg)0.0360.0130.010
R20.9920.9800.883
RBqmax (mg/g)29.021.419.6
KL (L/mg)3.312.202.02
R20.9590.7650.855
MGqmax (mg/g)37.0029.418.8
KL (L/mg)0.030.110.53
R20.8900.9830.964
BGqmax (mg/g)35.529.227.1
KL (L/mg)1.291.110.76
R20.9380.9490.918
ICqmax (mg/g)27.617.512.4
KL (L/mg)0.780.660.60
R20.8140.7560.737
FreundlichCV1/n0.9970.730.63
KF (mg1−1/n/g·L−1/n)0.110.320.35
R20.9510.9130.965
RB1/n0.710.750.78
KF (mg1−1/n/g·L−1/n)0.100.070.08
R20.9720.9690.966
MG1/n0.530.640.60
KF (mg1−1/n/g·L−1/n)0.850.200.10
R20.9450.9600.930
BG1/n0.840.650.58
KF (mg1−1/n/g·L−1/n)0.160.220.24
R20.9560.9390.998
IC1/n0.900.530.32
KF (mg1−1/n/g·L−1/n)0.100.200.40
R20.9110.9790.919
Nanocomposite with 10% GO
LangmuirCVqmax (mg/g)27.2722.4621.45
KL (L/mg)0.130.180.17
R20.8060.7950.768
RBqmax (mg/g)34.8832.4130.54
KL (L/mg)0.130.140.12
R20.9730.9340.924
MGqmax (mg/g)9.038.678.01
KL (L/mg)0.190.180.17
R20.8760.8470.879
BGqmax (mg/g)32.7129.0420.50
KL (L/mg)0.060.050.07
R20.9420.9170.913
ICqmax (mg/g)18.0515.4015.10
KL (L/mg)0.040.020.015
R20.8690.9100.923
FreundlichCV1/n0.600.620.63
KF (mg1−1/n/g·L−1/n)2.512.462.34
R20.9900.9910.989
RB1/n0.580.550.53
KF (mg1−1/n/g·L−1/n)1.161.121.08
R20.8440.7970.776
MG1/n0.550.540.53
KF (mg1−1/n/g·L−1/n)1.991.791.67
R20.9360.9390.940
BG1/n0.690.650.60
KF (mg1−1/n/g·L−1/n)1.271.131.04
R20.9610.9390.910
IC1/n0.610.510.49
KF (mg1−1/n/g·L−1/n)1.851.351.08
R20.9810.9940.998
Nanocomposite with 20% GO
LangmuirCVqmax (mg/g)30.7724.7523.56
KL (L/mg)0.180.210.14
R20.9980.9570.967
RBqmax (mg/g)30.7821.4418.04
KL (L/mg)0.210.280.32
R20.9220.8230.851
MGqmax (mg/g)24.7516.0013.76
KL (L/mg)0.060.170.20
R20.9450.9460.961
BGqmax (mg/g)27.4718.8116.64
KL (L/mg)0.130.160.17
R20.9770.9530.975
ICqmax (mg/g)14.8813.0412.05
KL (L/mg)0.120.110.10
R20.8190.8860.970
FreundlichCV1/n0.750.700.69
KF (mg1−1/n/g·L−1/n)3.1423.0502.268
R20.9840.9930.985
RB1/n0.530.550.56
KF (mg1−1/n/g·L−1/n)3.3702.8382.532
R20.9980.9960.994
MG1/n0.810.800.82
KF (mg1−1/n/g·L−1/n)1.3901.3331.290
R20.9990.9980.999
BG1/n0.730.600.58
KF (mg1−1/n/g·L−1/n)3.0052.9502.803
R20.9470.9790.973
IC1/n0.550.600.75
KF (mg1−1/n/g·L−1/n)1.8531.5011.030
R20.9840.9960.992
Table 4. Kinetic parameters of the process of adsorption by a nanocomposite.
Table 4. Kinetic parameters of the process of adsorption by a nanocomposite.
DyeT (K)Ce (mg/g)Time to Reach Equilibrium (min)R2k1
(min−1)
k2 × 10−3 
(g/mg min)
Pseudo-First OrderPseudo-Second Order
Nanocomposite with 5% GO
IC27814.05900.9500.9900.600.36
29314.63900.9850.9820.650.34
30818.01900.9550.9510.811.66
BG2788.88900.9880.9880.682.03
2939.56900.9820.9720.811.59
30811.28900.9900.9380.551.45
CV2786.64900.9870.9950.681.19
2937.94900.9760.9980.601.52
3089.86900.9600.9620.651.76
RB2785.10900.9660.8890.611.02
2935.18900.9740.8950.510.98
3085.65900.9960.9950.340.86
MG27814.50900.9230.9630.452.24
29315.38900.9960.9980.605.69
30817.94900.9580.9940.6219.61
Nanocomposite with 10% GO
IC27814.88900.9520.9550.551.35
29316.10900.98009250.731.26
30816.54900.9010.9721.1113.40
BG27810.36900.9950.9220.401.33
29310.96900.9940.9760.551.65
30811.16900.9880.9460.651.94
CV2786.30900.9900.9410.670.77
2936.57900.9880.9290.700.76
3086.72900.9860.9480.730.75
RB2784.76900.9380.9950.602.80
2935.08900.9510.9960.612.77
3085.20900.9570.9920.632.74
MG27815.54900.9920.9680.603.81
29315.90900.9610.9770.935.58
30816.70900.9770.9170.414.55
Nanocomposite with 20% GO
IC27812.92900.9550.9280.651.18
29312.98900.9360.9670.611.05
30813.04900.9380.9720.511.01
BG2788.80900.9640.9640.731.49
29310.26900.9000.9750.601.70
30812.00900.9510.8550.701.10
CV2786.00900.9530.9530.630.69
2936.70900.9380.9550.680.76
3087.70900.9240.9620.700.82
RB2784.42900.9880.9630.870.80
2934.70900.9780.9770.610.71
3084.95900.9670.9840.530.65
MG27810.50900.8500.9500.311.23
29310.70900.8730.9710.351.25
30811.20900.9650.8890.471.36
Table 5. Thermodynamic parameters of dye adsorption by the nanocomposite.
Table 5. Thermodynamic parameters of dye adsorption by the nanocomposite.
SorbateT, KKcΔG0 (kJ/mol)ΔH0 (kJ/mol)ΔS0 (J/mol K)
Nanocomposite with 5% GO
CV2781.420−2.5130.1212.1
2931.512−2.189
3081.705−2.103
RB2781.592−4.0890.671.57
2931.680−4.203
3081.769−4.115
MG2781.873−5.835−0.931.86
2932.315−6.112
3082.562−4.768
BG2780.425−2.1230.840.42
2930.769−1.912
3080.913−1.105
IC2782.440−5.636−1.2810.6
2932.211−5.356
3080.280−0.7
Nanocomposite with 10% GO
CV2781.470−3.3980.8108.94
2931.408−3.430
3081.374−3.518
RB2781.857−4.2920.72712.24
2931.771−4.314
3081.739−4.453
MG2781.928−4.4560.7158.11
2931.662−4.049
3081.555−3.982
BG2780.621−1.4350.6946.53
2930.501−1.220
3080.460−1.178
IC2781.071−2.4760.7026.74
2930.510−1.242
3080.374−0.958
Nanocomposite with 20% GO
CV2781.541−3.5620.7328.98
2931.379−3.359
3081.162−2.976
RB2781.953−4.5140.57717.53
2931.873−4.560
3081.805−4.622
MG2781.717−3.9680.72516.09
2931.638−3.990
3081.571−4.023
BG2780.934−2.1590.7547.90
2930.641−1.561
3080.289−0.740
IC2780.092−0.2130.7812.01
2930.079−0.192
3080.065−0.166
Table 6. Dependence of the degree of extraction of the dyes by the composite (5% GO) on the pH.
Table 6. Dependence of the degree of extraction of the dyes by the composite (5% GO) on the pH.
BGpH2345678
R (%)51.751.951.951.95252.252.2
CVpH2345678910
R (%)48.350.457.260.360.360.350.743.230.1
RBpH45678910
R (%)60.368.274.174.174.174.174.073.873.6
MGpH45678910
R (%)232323.123.123.123.123.1
ICpH45678910
R (%)2525.325.425.625.625.524
Table 7. Dependence of the degree of extraction on the number of working cycles.
Table 7. Dependence of the degree of extraction on the number of working cycles.
CVCycle 12345
5% GOR, %60.355.751.146.239.4
CVCycle 12345
10% GOR, %67.667.467.166.465.2
CVCycle12345
20% GOR, %76.576.175.875.273.1
BGCycle 12345
5% GOR, %52.248.644.238.732.1
BGCycle 12345
10% GOR, %45.445.244.640.138.7
BGCycle 12345
20% GOR, %42.742.141.540.138.6
RBCycle 12345
5% GOR, %74.272.370.468.160.2
RBCycle 12345
10% GOR, %74.774.273.878.272.6
RBCycle 12345
20% GOR, %76.676.175.470.368.6
MGCycle 12345
5% GOR, %23.122.421.620.118.1
MGCycle 12345
10% GOR, %20.620.520.320.119.8
MGCycle 12345
20% GOR, %46.646.245.241.438.2
ICCycle 12345
5% GOR, %25.725.625.525.325.1
ICCycle 12345
10% GOCycle35.235.034.834.133.3
Table 8. Values of maximum adsorption and degree of extraction for studied dyes (n = 3, p = 0.95).
Table 8. Values of maximum adsorption and degree of extraction for studied dyes (n = 3, p = 0.95).
DyeGO Content in the Composite (%)Dye Natureqmax (mg/g)R (%)Limit of Detection (mg/L)Limit of Quantification (mg/L)Accuracy Score (Systematic Error)
±Δc (%)
Accuracy Rate (Error Characteristic) ±Δ (%)
IC5%Anionic196.025.60.6250.625–103.1 ± 2%4.2 ± 2%
10%Anionic105.635.40.30.3–12.43.1 ± 2%4.2 ± 2%
20%Anionic187.535.40.6250.625–7.53.1 ± 2%4.2 ± 2%
BG5%Cationic74.855.60.30.3–203.4 ± 3%3.7 ± 3%
10%Cationic12848.20.30.3–203.4 ± 3%3.7 ± 3%
20%Cationic13056.00.30.3–203.4 ± 3%3.7 ± 3%
MG5%Cationic89.327.50.6250.625–15.25.1 ± 2%2.9 ± 2%
10%Cationic51.522.31.251.25–205.1 ± 2%2.9 ± 2%
20%Cationic233.647.50.6250.625–205.1 ± 2%2.9 ± 2%
RB5%Neutral84.074.50.30.3–202.9 ± 4%2.1 ± 3%
10%Neutral304.076.20.30.3–202.9 ± 4%2.1 ± 3%
20%Neutral345.677.90.30.3–202.9 ± 4%2.1 ± 3%
CV5%Cationic93.068.50.30.3–203.7 ± 2%3.5 ± 2%
10%Cationic380.468.50.30.3–203.7 ± 2%3.5 ± 2%
20%Cationic393.070.00.30.3–203.7 ± 2%3.5 ± 2%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Uflyand, I.E.; Naumkina, V.N.; Zhinzhilo, V.A. Nanocomposites of Copper Trimesinate and Graphene Oxide as Sorbents for the Solid-Phase Extraction of Organic Dyes. J. Compos. Sci. 2022, 6, 215. https://doi.org/10.3390/jcs6070215

AMA Style

Uflyand IE, Naumkina VN, Zhinzhilo VA. Nanocomposites of Copper Trimesinate and Graphene Oxide as Sorbents for the Solid-Phase Extraction of Organic Dyes. Journal of Composites Science. 2022; 6(7):215. https://doi.org/10.3390/jcs6070215

Chicago/Turabian Style

Uflyand, Igor E., Victoria N. Naumkina, and Vladimir A. Zhinzhilo. 2022. "Nanocomposites of Copper Trimesinate and Graphene Oxide as Sorbents for the Solid-Phase Extraction of Organic Dyes" Journal of Composites Science 6, no. 7: 215. https://doi.org/10.3390/jcs6070215

Article Metrics

Back to TopTop