Next Article in Journal
Capillary Electrophoresis-Laser Induced Fluorescence Method Development and Validation for Quantification of Nine Gangliosides—Application to Analysis of Cell Lines of CNS Origin
Previous Article in Journal
Exploring the Mutated Kinases for Chemoenzymatic Synthesis of N4-Modified Cytidine Monophosphates
Previous Article in Special Issue
Preparation of 4-Amino-3-hydrazino-1,2,4-triazol-5-thiol-Modified Graphene Oxide and Its Greatly Enhanced Selective Adsorption of Gallium in Aqueous Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Graphene Oxide Covalently Functionalized with 5-Methyl-1,3,4-thiadiazol-2-amine for pH-Sensitive Ga3+ Recovery in Aqueous Solutions

College of Chemistry, Sichuan University, Chengdu 610065, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(16), 3768; https://doi.org/10.3390/molecules29163768
Submission received: 12 June 2024 / Revised: 19 July 2024 / Accepted: 31 July 2024 / Published: 9 August 2024
(This article belongs to the Special Issue Design and Application Based on Versatile Nano-Composites)

Abstract

:
A novel graphene-based composite, 5-methyl-1,3,4-thiadiazol-2-amine (MTA) covalently functionalized graphene oxide (GO-MTA), was rationally developed and used for the selective sorption of Ga3+ from aqueous solutions, showing a higher adsorption capacity (48.20 mg g−1) toward Ga3+ than In3+ (15.41 mg g−1) and Sc3+ (~0 mg g−1). The adsorption experiment’s parameters, such as the contact time, temperature, initial Ga3+ concentration, solution pH, and desorption solvent, were investigated. Under optimized conditions, the GO-MTA composite displayed the highest adsorption capacity of 55.6 mg g−1 toward Ga3+. Moreover, a possible adsorption mechanism was proposed using various characterization methods, including scanning electron microscopy (SEM) equipped with X-ray energy-dispersive spectroscopy (EDS), elemental mapping analysis, Fourier transform infrared (FT-IR) spectroscopy, and X-ray photoelectron spectroscopy (XPS). Ga3+ adsorption with the GO-MTA composite could be better described by the linear pseudo-second-order kinetic model (R2 = 0.962), suggesting that the rate-limiting step may be chemical sorption or chemisorption through the sharing or exchange of electrons between the adsorbent and the adsorbate. Importantly, the calculated qe value (55.066 mg g−1) is closer to the experimental result (55.60 mg g−1). The well-fitted linear Langmuir isothermal model (R2 = 0.972~0.997) confirmed that an interfacial monolayer and cooperative adsorption occur on a heterogeneous surface. The results showed that the GO-MTA composite might be a potential adsorbent for the enrichment and/or separation of Ga3+ at low or ultra-low concentrations in aqueous solutions.

1. Introduction

The extraction of vanadium from vanadium titano-magnetite ores by salt roasting is one of the most representative routes. After sodium vanadate is leached out, several rare, dispersed elements, such as indium (In), gallium (Ga), and scandium (Sc) at low concentration levels, remain in the vanadium slag-processing residue (VSPR), causing undesirable environmental risks and resource waste [1], receiving considerable attention from the Ministry of Ecology and Environment. Ga is one of the most important strategic materials for high-tech applications, playing an irreplaceable role in the electronic and aerospace industries. However, Ga-containing resources are unevenly distributed globally, and their presence cannot be guaranteed [2]. In addition, the low Ga content in minerals makes it extremely difficult to guarantee its economically feasible recovery. To recover Ga, In, and Sc at low concentration levels in aqueous solutions, various technologies have been developed. Fortunately, the enrichment of Ga seems possible through adsorption treatment technologies [3]. Thus, the development of effective adsorbents for the highly efficient and selective recovery of Ga is urgent and worthy of investigation.
Adsorption is thought to be one of the most efficient and convenient strategies, but its mechanism is still unclear [2,4]. The adsorption of Ga3+ and In3+ by γ-Al2O3 suggests that bidentate Ga3+ and In3+ and monodentate GaOH2+/InOH2+ might be formed on the surface, benefiting efficient adsorption [5]. The adsorption of Ga3+ from aqueous solutions by oxidized coir (OC) was implemented in another study, and a high adsorption rate of 70.53% and an adsorption capacity of 19.42 mg g−1 could be obtained at pH = 3 in the pH range of 1~3 due to the ion exchange interactions [6]. The pH and the adsorbent dosage were found to have a positive effect on the adsorption of Ga3+ by bentonite, indicating that the former might be an important and significant factor [7]. The cationic or anionic contaminants adsorbed on the surfaces of an adsorbent material are usually dominated by a specific mode of adsorption—chemisorption [8]. That is, ion exchange, electrostatic interactions, and an electro-donor/acceptor (EDA) might be involved in the adsorption process. The solution’s pH values greatly affect the adsorbent’s surface charges and the hydrolysis of the adsorbate; therefore, optimal adsorption is usually favored at a specific pH value, which is close to the acid dissociation constant (pKa). Additionally, the intrinsic surface complexation constant (KC,intr) between the positive/negative ions and the adsorbent is also very important [9]. To develop a highly efficient and selective adsorbent, the rational design of its structure and the grafted functional groups are important, as these are undoubtedly dominant in influencing its adsorption performance.
Graphene oxide (GO)-based composites have recently attracted great attention in the field of adsorption and molecular separation [10]. The modification of GO with different organic molecules could endow it with specific adsorption properties, such as high efficiency, good selectivity, and excellent reusability [11,12,13]. Specifically, the adsorbability of GO-based composites toward heavy metal ions, rare earth elementals (REEs), or organic contaminants (including dyes, phenols, antibiotics, and pesticides) could be regulated [14,15,16]. Previous research has suggested that nitrogen (N)-containing heterocyclic compounds could be used to modify GO to selectively adsorb specific metal ions in aqueous solutions [17,18]. Evidently, the existence of C=N and -NH2 groups in 2-aminobenzothiazole (ABT) would benefit the selective adsorption of REEs [17], while N-enriched lignosulfonate could selectively interact with hexavalent chromium (Cr6+) [18]. Previous research has also suggested that 2-mercapto-5-methyl-1,3,4-thiadiazole and 2,5-bis(methylmercapto)-1,3,4-thiadiazole could be combined with Ni2+, Cu2+, Zn2+, Cd2+, and Hg2+ [19,20], revealing that the different bonding modes of N-containing groups produce quite different adsorption characteristics or affinities.
In another study, the highly effective adsorption of Ga3+ from waste Bayer solution was successfully carried out using polyacrylonitrile nanofiber membranes (PAN NFs) containing amidoxime groups [21], suggesting that certain N-containing groups could selectively interact with Ga3+. There have been few reports so far on the evaluation of N-containing molecules with modified GO for Ga3+ adsorption, and it is of great significance to investigate novel GO-based composites for the highly efficient and selective enrichment of Ga3+ in aqueous solutions. The construction of Ga3+-imprinted polyacrylic acid on GO (IIP-GO/PAA) using controllable polymerization and surface-imprinting technology has previously been implemented; this exhibited a better adsorption capacity (221.56 mg g−1) for Ga3+ and higher selectivity in comparison with Al3+, Fe3+, Mg2+, and Ca2+ in the acid-leaching process of fly ash [22]. However, some shortcomings, such as the complicated and uncontrollable operations, were also apparent. The selective adsorption of Ga3+ in the presence of Sc3+ and In3+ seems to be more difficult because their properties are very similar. Previously, a 4-amino-3-hydrazino-1,2,4-triazol-5-thiol (AHTZT)-modified GO (GO-AHTZT) composite, possessing a relatively lower adsorption capacity and an unremarkable selectivity for Ga3+, was developed; this process was preferentially protonated due to the existence of free amino and hydrazine groups, causing inevitable electrostatic repulsions with Ga3+ [23].
In this work, to improve the adsorption capacity and selectivity of GO-based materials, an interesting organic ligand, 5-methyl-1,3,4-thiadiazol-2-amine (MTA), was covalently coupled with GO to produce a novel composite through nucleophilic addition. We proposed that it might possess adsorption selectivity for Ga3+ due to the heterocyclic -C=N-N=C- groups. Because of the basic properties of the heterocyclic -C=N-N=C- and -NH- groups, the GO-MTA composite might also be pH-sensitive. Batch adsorption experiments were performed to evaluate the adsorbability of a GO-MTA composite for Ga3+, and the adsorption mechanism was proposed by fitting the experimental data with different adsorption kinetic and isotherm models.

2. Results and Discussion

2.1. Characterization Results

2.1.1. SEM, EDS, and Elemental Mapping Analyses

The surface morphologies of the fabricated samples were recorded via SEM (Figure 1). Lamellar GO structures could be clearly observed, indicating that the chemical exfoliation of the flake graphite had been successfully performed (Figure 1A). Many wrinkles near the edges of the GO sheets could also be found, suggesting that single- or multi-layered lamellar structures can provide abundant active sites for further modification [24,25]. After the chemical attachment of MTA, the prepared GO-MTA composite exhibited a thickened multi-layered structure with a rougher surface due to the improved π-π stacking and hydrogen bonding interactions (Figure 1B). After Ga3+ adsorption, the morphology of the GO-MTA composite remained almost unchanged, confirming the structural stability of the developed GO-based adsorbent (Figure 1C).
The elemental distribution of the GO-MTA composite post Ga3+ adsorption was further elucidated (Figure 2A–F). Noticeably, N was uniformly distributed, along with C and O, on the surface of the GO-MTA composite (Figure 2B–D), while S was distributed almost evenly over the surface (Figure 2E). The distribution of Ga3+ was consistent with that of N, indicating that it had been successfully adsorbed onto the GO-MTA through strong cation–lone pair interactions (Figure 2C,F) [26]. According to the EDS analysis, 79.62 wt.% of C, 5.37 wt.% of N, 13.13 wt.% of O, and 4.62 wt.% of S were measured (Figure 2G).

2.1.2. FT-IR Spectra

The FT-IR spectra of the GO (Figure S1), MTA, and the GO-MTA composite were also recorded (Figure 3). For MTA (Figure 3A), the broad peaks observed at the wavenumbers of 3414.07, 3255.87, and 3101.23 cm−1 clearly confirmed the existence of hydrogen-bonding interactions and stretching vibrations of N-H [27]. The peaks at 2971.5 and 2786.91 cm−1 corresponded to the symmetric- and asymmetric-stretching vibrations of saturated C-H, respectively. A weak characteristic peak corresponding to S-H could be observed at 2617.8 cm−1, while two weak characteristic peaks corresponding to C=N could be detected at the wavenumbers of 1641.05 and 1531.67 cm−1. The peak at 686.51 cm−1 could be assigned to the ν(C-S)-stretching mode [28], and the characteristic peak at 1385.59 cm−1 confirmed the existence of -CH3.
In comparison, the GO-MTA composite possessed characteristic peaks corresponding to N-H (3417.6~3792.8 cm−1), -CH3 (2910.81, 2822.16, 1384.26 cm−1), C=N (1634 cm−1), and the skeleton vibrations of the benzene ring (1619.61 cm−1), confirming the successful attachment of MTA onto GO (Figure 3B). Due to the specific functional groups on the surface, adsorption might involve two processes: the release of protons from the surface -COOH groups, followed by Ga3+ coordination to the de-pronated sites, and the coordination interactions between Ga3+ and the N-N component (Scheme 1) [29]. The intrinsic Ga3+ adsorption mechanism of the GO-MTA composite was further determined using XPS and adsorption kinetic/isothermal models.

2.1.3. XPS

To reveal its composition and chemical state, XPS surface characterization of the GO-MTA composite pre and post Ga3+ adsorption was performed. The survey spectra of the GO-MTA composite indicated the presence of C, N, S, O, and Ga, while the XPS of the GO-MTA-Ga3+ revealed a new peak corresponding to the Ga (Figure 4A). Deconvolution of the C 1s region led to six peaks, located at 290.47, 288.68, 286.55, 284.80, 283.35, and 282.39 eV, corresponding to O=C-OH, O=C-H, C-O, C=C, C-S, and C=N bonds, respectively. After Ga3+ adsorption, peaks at 291.12, 288.89, 286.58, 284.80, 283.26, and 282.69 eV could be observed. The clear shifts in the binding energy of the O=C-OH and C=N bonds indicate that these two groups were involved in the adsorbent–adsorbate interactions (Figure 4B) [30].
Further deconvolution of the N 1s region showed two peaks located at 400.39 and 399.32 eV, which could be attributed to the C-N and -C=N bonds, respectively. After Ga3+ adsorption, the peaks of the GO-MTA-Ga3+ shifted to 399.52 and 397.59 eV, respectively, again confirming that the -C=N and C-N bonds might coordinate with Ga3+ due to EDA interactions (Figure 4C, Scheme 1) [31].
Previous studies have confirmed that the O-containing functional groups of oxidized carbon nanomaterials, such as the carboxyl (-COOH) and hydroxyl (-OH) groups, may participate in the adsorption process of positive metal cations [32]. In our study, the XPS peak-differentiation-imitating analyses of the O 1s region displayed three peaks located at 532.89, 530.57, and 530.44 eV, which could be assigned to the C-O-C, -O-C=O, and O-H bonds, respectively (Figure 4D). For GO-MTA-Ga3+, the abovementioned peaks shifted to 534.62, 532.61, and 531.34 eV, respectively, also guaranteeing the EDA interactions between the O-containing functional groups and Ga3+ (Scheme 1).
To evaluate the possible interactions between the S-containing group and Ga3+, deconvolution of the S 2p region was performed (Figure 4E). Three peaks were located at 168.43, 165.08 and 164.00 eV, corresponding to the characteristic peaks of S 2p3/2 and S 2p1/2, respectively, shifting to 164.48 and 160.96 eV after Ga3+ adsorption [33]. These results suggested that the S-containing group might also be involved in the adsorption process (Scheme 1). As shown in Scheme 1, the simultaneous interaction of Ga3+ with the C-N, -OH, and C-S groups appeared to be infeasible or difficult. We therefore propose that the -C=N-N=C- groups might have dominated in the adsorption process, guaranteeing the comparatively high adsorption selectivity of the GO-MTA composite for Ga3+. The lone pairs of the two N atoms in heterocyclic -C=N-N=C- were located in the sp2 hybrid orbitals, making them relatively better electron donors [34].

2.2. Adsorption Experiments

2.2.1. Adsorption Selectivity

Different metal ion solutions including Ga3+, Sc3+, and In3+ were used to evaluate the adsorption properties of the GO-MTA composite. A total of 5.0 mg of the GO-MTA composite was added to a 50.0 mL conical bottle containing 20.0 mL of metal ion solution (50.0 mg L−1) and stirred at 25 °C for 180 min. The adsorption capacity of the GO-MTA composite for Ga3+, Sc3+, and In3+ was calculated to be 48.20, 15.41, and ~0 mg g−1, respectively (Figure 5), indicating that it possessed a relatively higher adsorption selectivity toward Ga3+ than the other two metal ions. We therefore propose that the N-N group was able to more strongly interact with Ga3+ than In3+ and Sc3+. As a result, Ga3+ was selected as a model ion to investigate the effects of conditional parameters on the adsorption process in the following experiments.

2.2.2. Effect of Contact Time and Adsorption Kinetics

The effect of the contact time on Ga3+ adsorption onto the GO-MTA composite was investigated (Figure 6A). We found that 120 min was enough to achieve adsorption equilibrium under the experimental conditions (adsorbent dosage = 5.0 mg, C0 = 50.0 mg L−1, and T = 298 K). Meanwhile, fast adsorption occurred in 15 min in the initial stage, followed by a gradually increased adsorption capacity. To achieve adsorption equilibrium, a contact time of 120 min was applied in the subsequent adsorption experiments.
The experimental data of the GO-MTA composite for Ga3+ at 298 K were fitted using different kinetic models (linear/nonlinear pseudo-first- and second-order models) (Equations (S1)–(S4)). As shown in Figure 6B–E and Table 1, the experimental data were distorted by the nonlinear kinetic forms, indicating that the linear ones should be primarily used to calculate the adsorption parameters [35]. The adsorption data could be well described by the linear pseudo-second-order kinetic model (R2 = 0.962), suggesting that the rate-limiting step may be chemical sorption or chemisorption through the sharing or exchange of electrons between the adsorbent and the adsorbate [36]. Importantly, the calculated qe value (55.066 mg g−1) from the plots of t/qt versus t was closer to the experimental result (55.60 mg g−1).

2.2.3. Effect of Solution pH

The initial pH value of a solution is one of the most important parameters able to affect metal cation adsorption. Figure 7A reflects the relationship between the adsorption capacity of the GO-MTA composite and the initial pH value of the solution (t = 120 min, adsorbent dosage = 5.0 mg, C0 = 50.0 mg L−1, and T = 298 K). It is clear that its adsorption capacity increases with an increase in the initial pH in the range of 2.0–3.0. The active sites of the N-containing groups on the surface of the GO-MTA composite are protonated under strongly acidic conditions, negatively influencing adsorption due to Ga3+ also being positively charged [37]. That is, the active sites are readily occupied by H+ rather than Ga3+. It is well known that Ga3+ is precipitated at higher pH values [38], whereby sedimentation rather than adsorption occurs, resulting in decreased adsorption capacities. The zeta potentials of the GO-MTA composite in the pH range of 2.0–6.0 were measured (Figure 7B). A large number of N-N groups were present on the surface of the GO-MTA composite. Noticeably, its zeta potentials in the tested pH range decreased gradually and then turned negative at pH = 5.0. The maximum adsorption capacity at pH = 3.0 indicated that the O-H, C-N, and S-C groups were positively charged at low pH values and that the electrostatic repulsions between the adsorbent and the adsorbate had a negative effect on adsorption. The complexation interactions between the N-N groups of the GO-MTA composite and Ga3+ mainly contributed to the effective adsorption, and the Ga3+ adsorption capacity of the composite gradually increased as the solution’s pH increased.

2.2.4. Adsorption Isotherms

Adsorption isotherms can be established when dynamic adsorption equilibrium is achieved. The adsorption progress is usually affected by experimental factors, including the solution’s pH, the contact temperature, the initial adsorbate concentration, and especially, the nature of the adsorbent. Optimizing adsorption processes by rationally changing the conditional parameters is of great significance. Most often, the adsorption isotherm could be obtained by changing the initial adsorbate concentration and contact temperature whilst keeping the other conditions unchanged. In this work, different types of isothermal models (linear/nonlinear Langmuir and Freundlich isothermal models) were employed to diagnose the nature of Ga3+ adsorption by the GO-MTA composite (Equations (S5)–(S8)).
Noticeably, the experimental data of the GO-MTA composite for Ga3+ could be better described by the linear Langmuir isothermal model. As can be seen from Figure 8a–e and Table 2, the linear regression of Ce/qe vs. Ce provides the best fit for the equilibrium data (R2 = 0.972–0.998). The estimated adsorption capacity qm is 55.556 mg g−1 at 308 K, which is much closer to the experimental data. In addition, the values of 0 < kL < 1 suggest that adsorption is an endothermic process, which is also consistent with the experimental data (Figure 8a) [39]. The interfacial monolayer’s capture efficiency for the adsorbate was also confirmed by the linear Langmuir isothermal model. However, cooperative adsorption caused primarily by attractive interactions between the monolayer and the dissolved adsorbate on a heterogeneous surface might occur.

2.2.5. Adsorption Thermodynamics

Figure S2 shows the experimental data and the fitted curve of Inkd versus 1/T calculated from the Van’t Hoff plots of the GO-MTA composite for Ga3+ adsorption. We can observe that the adsorption process is temperature-dependent and that an increase in the contact temperature results in a corresponding increase in the Ga3+ adsorption capacity of the GO-MTA composite. Noticeably, the results are in accordance with the adsorption isotherms fitted using the linear Langmuir isothermal model.
The standard free-energy change, ΔGo, was calculated. The average standard enthalpy change (ΔHo) was calculated from the Van’t Hoff equation. The thermodynamic parameters are listed in Table S1. Noticeably, the ΔGo values at all the tested temperatures and initial Ga3+ concentrations are negative, suggesting that Ga3+ adsorption by the GO-MTA composite is spontaneous. The ΔGo values become more negative as the contact temperature increases, indicating that higher temperatures benefit adsorption. Interestingly, the ΔGo values become less negative with increasing initial Ga3+ concentrations, implying that the presence of a specific external surface dominates physical adsorption. As the initial Ga3+ concentration increases, chemical adsorption due to the coordination effects between the adsorbate and the N-, O-, and S-containing functional groups of the GO-MTA composite dominates the process, as further verified by the positive ΔHo values calculated. That is, Ga3+ adsorption by the GO-MTA composite is an endothermic process, as supported by the increased Ga3+ adsorption with an increase in the contact temperature (Table S1).
Finally, the positive ΔSo values could be attributed to the transfer of the Ga3+ ion in the solutions caused by the interactions between Ga3+ and the functional groups of the GO-MTA composite [40].

2.3. Comparison of Ga3+ Adsorption with Previously Reported Adsorbents

The Ga3+ adsorption capacity of the newly developed GO-MTA composite was compared with that of various adsorbents previously reported in the literature (Table 3). Noticeably, in some cases, the experimental data from this investigation exhibited some advantages over previous data. The adsorbents’ adsorption capacity varied to a certain extent, depending on the individual physicochemical characteristics, the substrate, and the surface functional groups, as well as the initial adsorbate concentration. Importantly, the GO-MTA composite was relatively easier to fabricate and possessed a higher adsorption selectivity for Ga3+. To rationally design and develop a practically applicable adsorbent with specific adsorption selectivity toward Ga3+, the following experiments were carried out. These experiments could be used as a useful reference for follow-up studies.

2.4. Desorption and Reutilization

From the perspective of efficient resource use for sustainable development, the reutilization of a newly developed adsorbent should be evaluated. In this study, diluted nitric acid (HNO3, 1.0 mol L−1) was used as the eluent during the desorption process to reveal the GO-MTA composite’s reusability. After Ga3+ adsorption at room temperature for 6 h, the adsorbent was stirred, obtaining an elution rate of ~90%. The GO-MTA composite’s adsorption efficiency during 10 adsorption–desorption cycles was investigated, showing that the removal rate remained above ~80% even after all the cycles (Figure 9). Generally, the GO-MTA composite possesses a relatively higher stability and may be used for practical applications.

2.5. Adsorption Properties of the GO-MTA Composite in a Mixed Solution

The GO-MTA composite (50.0 mg) was applied to a mixed solution (20.0 mL) containing Sc3+ (10.0 mg L−1) and Ga3+ (10.0 mg L−1) in a conical flask (50.0 mL) at 35 °C for 180 min. Relatively lower adsorption capacities of 0.45 mg mL−1 and 15.76, respectively, were determined. Noticeably, the GO-MTA composite exhibited a relatively higher adsorption selectivity toward Ga3+ even in a mixed solution, suggesting its potential applications in real samples.

3. Experimental Details

3.1. Fabrication of the GO Nanosheets

The GO suspension was easily prepared according to the modified Hummer method [47,48]. In brief, flake graphite (0.60 g) was evenly mixed with KMnO4 powder (3.00 g), then transferred into a round-bottom bottle (500.0 mL) placed in an ice-water bath. Under constant stirring, a mixture of H3PO4 (85%, 8.0 mL) and H2SO4 (98%, 72.0 mL) was slowly added to the reaction. Once the dispersion had gradually turned a bottle-green color, it was heated to 50 °C and reacted for 12 h until it turned purple–red. Once the mixture had naturally cooled to room temperature, excess H2O2 was added until it turned bright yellow, with no bubbles. The mixture was washed repeatedly with a certain amount of HCl (1.0 mol L−1) and ultrapure water, and it was separated by centrifugation until the supernatant’s pH was close to being neutral. Finally, the black solid residual was carefully collected, dispersed in a certain amount of ultrapure water, and then freeze-dried at −50 °C for 48 h to obtain GO nanosheets.

3.2. Fabrication of the GO-MTA Composite

GO (500.0 mg) was dispersed in 50.0 mL of ultrapure water in a 100.0 mL round-bottom flask and then ultrasonically treated to obtain an evenly mixed GO dispersion. Afterwards, 1.0 mL of CH3COOH (99.5 wt.%) and 2.0 g of MTA were added to the reaction, followed by heating the mixture to 110 °C in an oil bath for 8 h. After the reaction, the mixture was naturally cooled to room temperature, diluted with a certain amount of ultrapure water, and then filtered and washed 3 times with ethanol and ultrapure water until the filtrate had turned neutral. The residual solid on the filter paper was collected, then dispersed in a certain amount of ultrapure water, and finally, freeze-dried at −50 °C for 36 h to obtain the GO-MTA composite.

3.3. Batch Adsorption Experiments

The effects of the contact time, initial pH of the solution, initial Ga3+ concentration, and contact temperature on the GO-MTA composite’s Ga3+ adsorption performance were investigated. Ga3+ solutions were prepared by dissolving a certain amount of Ga(NO3)3∙H2O in ultrapure water and then transferred into a volumetric flask and diluted to a certain volume. A total of 5.0 mg of the GO-MTA composite was mixed with 20.0 mL of Ga(NO3)3 solution in a 50.0 mL conical flask and continuously stirred at a certain temperature (15 °C, 20 °C, 25 °C, 30 °C, and 35 °C) for a certain time (0–180 min). Three parallel experiments were performed, and the adsorption data were calculated using Equation (1) to obtain the equilibrium adsorption capacity (qe; mg g−1), while the removal rate (R, %) was calculated using Equation (2):
q e = ( C 0 C e ) V m
R = C 0 C e C 0 × 100 %
where C0 (mg L−1) represents the initial Ga3+ concentration; Ce (mg L−1) represents the equilibrium Ga3+ concentration; and V (L) and m (g) represent the volume of the Ga3+ solution and the mass of the GO-MTA composite, respectively.

3.4. Sample Characterization

The morphology and elemental distributions of the GO-MTA composite pre and post Ga3+ adsorption were verified using a field emission scanning electron microscope (FE-SEM; JSM-7900F, JEOL Corp.; Tokyo, Japan) equipped with an energy-dispersive X-ray spectroscope (EDS) at an accelerating voltage of 10 kV. The Fourier transform infrared (FTIR) spectra were recorded on a Shimadzu FTIR (IR Prestige-21; Shimadzu, Ltd., Tokyo, Japan) spectrophotometer in the wavenumber range of 4000–400 cm−1 and at a resolution of 4 cm−1 to detect the surface functional groups of the GO, MTA, and GO-MTA composite. The composition and chemical states of the GO-MTA composite pre and post Ga3+ adsorption were characterized using an X-ray photoelectron spectroscope (XPS) on a Perkin Elmer PHI 5000 C ESCA instrument (Perkin Elmer Co.; Eden Prairie, MN, USA) using Al Kα radiation (1486.6 eV) with a detection angle of 54° and a cathode voltage of 14 kV, operated at 250 W. The metal ion concentration was determined via an inductively coupled plasma–optical emission spectrometer (ICP-OES; IRIS Intrepid II XSP; Thermo Electron Corporation, Waltham, MA, USA).

4. Conclusions

In this study, a highly selective GO-MTA composite with a specific N-containing functional group was fabricated through the nucleophilic addition of MTA and GO. This GO-MTA composite exhibits a stacked lamellar structure. Due to the surface N-, O-, and S-containing functional groups, the GO-MTA composite exhibits a relatively higher adsorption capacity and selectivity for Ga3+ compared to Sc3+ and In3+. In our experiments, at an initial concentration of 50.0 mg L−1, a contact temperature of 35 °C, a contact time of 120 min, and an adsorbent dosage of 5.0 mg, a high Ga3+ adsorption capacity of 55.6 mg g−1 could be observed, and the adsorption process could be best described by the linear Langmuir isothermal model (R2 = 0.972–0.997) and the linear pseudo-second-order kinetic model (R2 = 0.962). Ga3+ adsorption by the GO-MTA composite under acidic conditions was mainly dominated by the coordination interactions of the adsorbate with the adsorbent’s -C=N-N=C- functional group. Elution rates of ~90% could be obtained using HNO3, and a removal rate above ~80% could be found even after 10 adsorption–desorption cycles. Application of the GO-MTA composite in a mixed solution containing Sc3+ and Ga3+ confirmed its excellent adsorption selectivity. These results suggest that the GO-MTA composite could be implemented as a potential adsorbent for Ga3+ separation/enrichment processes in aqueous solutions.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules29163768/s1: Figure S1: FT-IR spectra of GO; Figure S2: Experimental data and the fitted curve of Inkd versus 1/T calculated from the Van’t Hoff plots of the GO-MTA composite for Ga3+ with different concentrations—(A) 10 mg L−1; (B) 20 mg L−1; (C) 30 mg L−1; (D) 40 mg L−1; and(E) 50 mg L−1; and Table S1: Adsorption thermodynamic parameters of the GO-MTA composite for Ga3+.

Author Contributions

Conceptualization, X.Z. and Y.G.; methodology, X.Z. and Y.G.; validation, X.Z. and Y.G.; formal analysis, X.Z. and Y.G.; investigation, X.Z.; resources, X.Z. and Y.G.; data curation, X.Z. and Y.G.; writing—original draft preparation, X.Z. and B.Z.; writing—review and editing, X.Z., Y.G. and B.Z.; visualization, X.Z.; supervision, Y.G.; and project administration, Y.G. 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 original contributions presented in this study are included in this article/the Supplementary Materials. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Syu, C.H.; Chen, L.Y.; Lee, D.Y. The growth and uptake of gallium (Ga) and indium (In) of wheat seedlings in Ga- and In-contaminated soils. Sci. Total Environ. 2021, 759, 143943. [Google Scholar] [CrossRef]
  2. Long, J.; Zhang, S.; Luo, K. Discovery of anomalous gallium enriched in stone coal: Significance, provenance and recommendations. Geosci. Front. 2023, 14, 101538. [Google Scholar] [CrossRef]
  3. Qin, S.; Sun, Y.; Li, Y.; Wang, J.; Zhao, C.; Gao, K. Coal deposits as promising alternative sources for gallium. Earth-Sci. Rev. 2015, 150, 95–101. [Google Scholar] [CrossRef]
  4. Qi, H.; Gong, N.; Zhang, S.Q.; Li, J.; Yuan, G.L.; Liu, X.F. Research progress on the enrichment of gallium in bauxite. Ore Geol. Rev. 2023, 160, 105609. [Google Scholar] [CrossRef]
  5. Lin, C.F.; Chang, K.S.; Tsay, C.W.; Lee, D.Y.; Lo, S.L.; Yasunaga, T. Adsorption mechanism of gallium(III) and indium(III) onto γ-Al2O3. J. Colloid Interface Sci. 1997, 188, 201–208. [Google Scholar] [CrossRef]
  6. Suryavanshi, U.S.; Shukla, S.R. Adsorption of Ga(III) on Oxidized Coir. Ind. Eng. Chem. Res. 2009, 48, 870–876. [Google Scholar] [CrossRef]
  7. Chegrouche, S.; Bensmaili, A. Removal of Ga(III) from aqueous solution by adsorption on activated bentonite using a factorial design. Water Res. 2002, 36, 2898–2904. [Google Scholar] [CrossRef] [PubMed]
  8. Hua, T.; Haynes, R.J.; Zhou, Y.F.; Boullemant, A.; Chandrawana, I. Potential for use of industrial waste materials as filter media for removal of Al, Mo, As, V and Ga from alkaline drainage in constructed wetlands–Adsorption studies. Water Res. 2015, 71, 32–41. [Google Scholar] [CrossRef] [PubMed]
  9. Bi, X.; Westerhoff, P. Adsorption of III/V ions (In(III), Ga(III) and As(V)) onto SiO2, CeO2 and Al2O3 nanoparticles used in the semiconductor industry. Environ. Sci. Nano 2016, 3, 1014–1026. [Google Scholar] [CrossRef]
  10. Cui, Y.X.; Duan, M.Y.; Liu, Y.P.; Hu, W.G.; Jiang, X.Y.; Yu, J.G. Fabrication of novel 4-methyl-5-thiazoleethanol covalently-linked graphene oxide composite with adsorption selectivity for Cu2+ from aqueous solutions. J. Taiwan Inst. Chem. Eng. 2024, 156, 105369. [Google Scholar] [CrossRef]
  11. Ren, H.; Cao, Z.F.; Chen, Y.Y.; Jiang, X.Y.; Yu, J.G. Graphene oxide-Bicine composite as a novel adsorbent for removal of various contaminants from aqueous solutions. J. Environ. Chem. Eng. 2021, 9, 106769. [Google Scholar] [CrossRef]
  12. Chen, Y.Y.; Lan, X.W.; Ren, H.; Li, W.-J.; Chen, J.; Jiang, X.Y.; Yu, J.G. Three-dimensional hybrid nitrogen/oxygen-containing components modified graphene oxide as a recyclable adsorbent for rapid adsorption of REEs. J. Environ. Chem. Eng. 2021, 9, 106500. [Google Scholar] [CrossRef]
  13. Velusamy, S.; Roy, A.; Sundaram, S.; Kumar Mallick, T. A review on heavy metal ions and containing dyes removal through graphene oxide-based adsorption strategies for textile wastewater treatment. Chem. Rec. 2021, 21, 1570–1610. [Google Scholar] [CrossRef]
  14. Sherlala, A.I.A.; Raman, A.A.A.; Bello, M.M.; Asghar, A. A review of the applications of organo-functionalized magnetic graphene oxide nanocomposites for heavy metal adsorption. Chemosphere 2018, 193, 1004–1017. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, Y.; Pan, C.; Chu, W.; Vipin, A.K.; Sun, L. Environmental remediation applications of carbon nanotubes and graphene oxide: Adsorption and catalysis. Nanomaterials 2019, 9, 439. [Google Scholar] [CrossRef] [PubMed]
  16. Peng, W.; Li, H.; Liu, Y.; Song, S.J. A review on heavy metal ions adsorption from water by graphene oxide and its composites. J. Mol. Liq. 2017, 230, 496–504. [Google Scholar] [CrossRef]
  17. Liao, C.; Liu, Y.P.; Ren, H.; Jiang, X.Y.; Yu, J.G.; Chen, X.Q. Rational assembly of GO-based heterocyclic sulfur- and nitrogen-containing aerogels and their adsorption properties toward rare earth elementals. J. Hazard. Mater. 2021, 419, 126484. [Google Scholar] [CrossRef] [PubMed]
  18. Geng, J.; Liang, Q.; Yu, W.; Chen, W.; Lu, G.; Luo, H. Enhanced removal of Cr(VI) from aqueous solutions by polymer-mediated nitrogen-rich reduced graphene oxide. J. Hazard. Mater. 2022, 436, 129184. [Google Scholar] [CrossRef] [PubMed]
  19. Fabretti, A.C.; Franchini, G.C.; Peyronel, G. Cobalt(II), nickel(II), copper(II) and copper(I) complexes of 2-mercapto-5-methyl-1,3,4-thiadiazole and 2,5-bis(methylmercapto)-1,3,4-thiadiazole. Transit. Met. Chem. 1982, 7, 105–108. [Google Scholar] [CrossRef]
  20. Bharati, P.; Bharti, A.; Bharty, M.K.; Kashyap, S.; Singh, U.P.; Singh, N.K. Synthesis, spectral and structural characterization of Ni(II), Cu(II), Zn(II), Cd(II) and Hg(II) complexes with 2-mercapto-5-methyl-1,3,4-thiadiazole: A Zn(II) complex acting as a new sensitive and selective fluorescent probe for the detection of Hg2+ in H2O–MeOH medium. Polyhedron 2013, 63, 222–231. [Google Scholar]
  21. Qin, Z.; Liao, Y.; Wang, Z.; Wang, S.; Song, L.; Ma, K.; Luo, D.; Yue, H. Innovative amidoxime nanofiber membranes for highly effective adsorption of Ga(III) from waste Bayer solution. Ind. Eng. Chem. Res. 2023, 62, 11140–11150. [Google Scholar] [CrossRef]
  22. Zhang, M.; Zhang, X.; Liang, C.; Qin, L.; Fu, D.; Wang, M.; Bai, Y.; Liu, W.; Liu, X. Preparation of graphene-based surface ion-imprinted adsorbent for Ga(Ⅲ) selective extraction from acid leaching of fly ash. Sep. Purif. Technol. 2023, 325, 124681. [Google Scholar] [CrossRef]
  23. Zhu, X.; Guo, Y.; Zheng, B. Preparation of 4-Amino-3-hydrazino-1,2,4-triazol-5-thiol-Modified graphene oxide and its greatly enhanced selective adsorption of gallium in aqueous solution. Molecules 2024, 29, 2778. [Google Scholar] [CrossRef] [PubMed]
  24. Shen, K.; Li, P.; Lin, J.; Wang, Z.; Cai, G.; Ming, X.; Liu, Y.; Gao, C.; Xu, Z. Intercalated oligomer doubles plasticity for strong and conductive graphene papers and composites. Carbon 2023, 208, 160–169. [Google Scholar] [CrossRef]
  25. Zeng, X.; Zhu, B.B.; Qiu, W.; Li, W.L.; Zheng, X.H.; Xu, B. A review of the preparation and applications of wrinkled graphene oxide. New Carbon Mater. 2022, 37, 290–302. [Google Scholar] [CrossRef]
  26. Laurita, G.; Seshadri, R. Chemistry, structure, and function of lone pairs in extended solids. Acc. Chem. Res. 2022, 55, 1004–1014. [Google Scholar] [CrossRef] [PubMed]
  27. Hipler, F.; Winter, M.; Fischer, R.A. N–H⋯S hydrogen bonding in 2-mercapto-5-methyl-1,3,4-thiadiazole. Synthesis and crystal structures of mercapto functionalised 1,3,4-thiadiazoles. J. Mol. Struct. 2003, 658, 179–191. [Google Scholar] [CrossRef]
  28. Silva, A.L.R.; Gonçalves, J.M.; Morais, V.M.F.; Ribeiro da Silva, M.D.M.C. Thermodynamic properties of 2-mercapto-, 2,5-dimethyl- and 2-mercapto-5-methyl-1,3,4-thiadiazole. J. Chem. Thermodyn. 2022, 165, 106644. [Google Scholar] [CrossRef]
  29. Qin, Z.; Wang, S.; Zhang, S.; Xie, J.; Zhou, C.-a.; Wang, C.; Song, L.; Ma, K.; Luo, D.; Yue, H. Cross-linked amidoxime porous resin for selective gallium separation in Bayer solutions: Reaction mechanism and kinetic study. Chem. Eng. J. 2024, 481, 148340. [Google Scholar] [CrossRef]
  30. Vinothkumar, K.; Rajkumar, G.; Sethuraman, M.G. Enhancement of protection of copper through electropolymerised poly-2-amino-1,3,4-thiadiazole and its composite films. Mater. Chem. Phys. 2021, 259, 123987. [Google Scholar] [CrossRef]
  31. Zhang, X.; Liu, J.; Zhang, H.; Zhang, Q.; Shen, J.; Wei, Y.; Wang, C. Facile construction of a stable core-shell spherically magnetic polyimide covalent organic framework for efficient extraction of phenylurea herbicides. Talanta 2024, 275, 126184. [Google Scholar] [CrossRef] [PubMed]
  32. Kong, Q.; Preis, S.; Li, L.; Luo, P.; Wei, C.; Li, Z.; Hu, Y.; Wei, C. Relations between metal ion characteristics and adsorption performance of graphene oxide: A comprehensive experimental and theoretical study. Sep. Purif. Technol. 2020, 232, 115956. [Google Scholar] [CrossRef]
  33. Lang, P.; Nogues, C. Self-assembled alkanethiol monolayers on a Zn substrate: Interface studied by XPS. Surf. Sci. 2008, 602, 2137–2147. [Google Scholar] [CrossRef]
  34. Del Valle, M.; Gacitúa, M.; Hernández, F.; Luengo, M.; Hernández, L. Nanostructured conducting polymers and their applications in energy storage devices. Polymers 2023, 15, 1450. [Google Scholar] [CrossRef] [PubMed]
  35. Khan, A.A.; Naqvi, S.R.; Ali, I.; Arshad, M.; AlMohamadi, H.; Sikandar, U. Algal-derived biochar as an efficient adsorbent for removal of Cr(VI) in textile industry wastewater: Non-linear isotherm, kinetics and ANN studies. Chemosphere 2023, 316, 137826. [Google Scholar] [CrossRef] [PubMed]
  36. Khader, E.H.; Khudhur, R.H.; Abbood, N.S.; Albayati, T.M. Decolourisation of anionic azo dye in industrial wastewater using adsorption process: Investigating operating parameters. Environ. Process. 2023, 10, 34. [Google Scholar] [CrossRef]
  37. Ji, Q.; Hu, C.; Liu, H.; Qu, J. Development of nitrogen-doped carbon for selective metal ion capture. Chem. Eng. J. 2018, 350, 608–615. [Google Scholar] [CrossRef]
  38. Kubicek, V.; Havlickova, J.; Kotek, J.; Tircsó, G.; Hermann, P.; Tóth, É.; Lukes, I. Gallium (III) complexes of DOTA and DOTA−monoamide: Kinetic and thermodynamic studies. Inorg. Chem. 2010, 49, 10960–10969. [Google Scholar] [CrossRef] [PubMed]
  39. Guo, X.; Wang, J. Comparison of linearization methods for modeling the Langmuir adsorption isotherm. J. Mol. Liq. 2019, 296, 111850. [Google Scholar] [CrossRef]
  40. Alrefaee, S.H.; Aljohani, M.; Alkhamis, K.; Shaaban, F.; El-Desouky, M.G.; El-Bindary, A.A.; El-Bindary, M.A. Adsorption and effective removal of organophosphorus pesticides from aqueous solution via novel metal-organic framework: Adsorption isotherms, kinetics, and optimization via Box-Behnken design. J. Mol. Liq. 2023, 384, 122206. [Google Scholar] [CrossRef]
  41. Li, W.; Zhou, C.; Li, C.; Zhu, W.; Shi, J.; Liu, G. Synthesis of UiO-66 series metal–organic framework composites and the adsorption effect on gallium. Chem. Eng. J. 2023, 455, 140881. [Google Scholar] [CrossRef]
  42. Li, Y.; Chen, X.; Guo, B.; Dai, Z.; Kong, Z.; Li, F.; Ou, J. Synthesis of polyacrylate-divinylbenzene hydroxamic resins and its gallium adsorption performance in sulfuric acid solution. J. Water Process Eng. 2024, 60, 105191. [Google Scholar] [CrossRef]
  43. Li, S.; Fan, J.; Gao, L. Conductive biomass carbon aerogel with high adsorption performance for gallium in alkaline solution. Miner. Eng. 2023, 204, 108396. [Google Scholar] [CrossRef]
  44. Li, W.; Zhou, C.; Li, C.; Zhu, W.; Shi, J.; Liu, G. Synthesis of CNT/UiO-66-NH2 adsorbent and the selective adsorption of gallium in solution. Sep. Purif. Technol. 2023, 323, 124464. [Google Scholar] [CrossRef]
  45. Raj, P.; Patel, M.; Karamalidis, A.K. Chemically modified polymeric resins with catechol derivatives for adsorption, separation and recovery of gallium from acidic solutions. J. Environ. Chem. Eng. 2023, 11, 110790. [Google Scholar] [CrossRef]
  46. Fan, S.; Liu, Y.; Zhang, Z.; Huang, M.; Wang, Y.; Gao, J.; Xiong, Y. A green synthesis method of a mussel-inspired polyphenol-functionalized silica-based material and its highly efficient adsorption of gallium. Sep. Purif. Technol. 2024, 349, 127670. [Google Scholar] [CrossRef]
  47. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved synthesis of graphene oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  48. Dimiev, A.M.; Tour, J.M. Mechanism of graphene oxide formation. ACS Nano 2014, 8, 3060–3068. [Google Scholar] [CrossRef]
Figure 1. SEM images of the samples: (A) GO; (B) GO-MTA composite; and (C) GO-MTA composite post Ga3+ adsorption.
Figure 1. SEM images of the samples: (A) GO; (B) GO-MTA composite; and (C) GO-MTA composite post Ga3+ adsorption.
Molecules 29 03768 g001
Figure 2. Elemental-mapping images of the GO-MTA composite post Ga3+ adsorption—(A) area tested, (B) C, (C) N, (D) O, (E) S, and (F) Ga3+—and (G) EDS of the composite.
Figure 2. Elemental-mapping images of the GO-MTA composite post Ga3+ adsorption—(A) area tested, (B) C, (C) N, (D) O, (E) S, and (F) Ga3+—and (G) EDS of the composite.
Molecules 29 03768 g002
Figure 3. FT-IR spectra of the samples: (A) MTA and (B) GO-MTA composite.
Figure 3. FT-IR spectra of the samples: (A) MTA and (B) GO-MTA composite.
Molecules 29 03768 g003
Scheme 1. A schematic diagram of the functionalization of the GO-MTA composite and its Ga3+ adsorption mechanism.
Scheme 1. A schematic diagram of the functionalization of the GO-MTA composite and its Ga3+ adsorption mechanism.
Molecules 29 03768 sch001
Figure 4. XPS spectra of the GO-MTA composite pre and post Ga3+ adsorption: (A) survey spectra and XPS peak-differentiation-imitating analyses of (B) C 1s, (C) N 1s, (D) O 1s, and (E) S 2p.
Figure 4. XPS spectra of the GO-MTA composite pre and post Ga3+ adsorption: (A) survey spectra and XPS peak-differentiation-imitating analyses of (B) C 1s, (C) N 1s, (D) O 1s, and (E) S 2p.
Molecules 29 03768 g004
Figure 5. Adsorption properties of the GO-MTA composite for Ga3+, Sc3+, and In3+ (dosage of adsorbent = 5.0 mg, V = 20.0 mL, C0 = 50.0 mg L−1, T = 25 °C, and t = 180 min).
Figure 5. Adsorption properties of the GO-MTA composite for Ga3+, Sc3+, and In3+ (dosage of adsorbent = 5.0 mg, V = 20.0 mL, C0 = 50.0 mg L−1, T = 25 °C, and t = 180 min).
Molecules 29 03768 g005
Figure 6. The adsorption capacity in MTA: (A) effect of the contact time (adsorbent dosage = 5.0 mg, C0 = 50.0 mg L−1, T = 298 K, and V = 20.0 mL; RSD = 0.14%~3.45%); (B) fitted using the linear pseudo-first-order kinetic model; (C) fitted using the nonlinear pseudo-first-order kinetic model; (D) fitted using the linear pseudo-second-order kinetic model; and (E) fitted using the nonlinear pseudo-second-order kinetic model.
Figure 6. The adsorption capacity in MTA: (A) effect of the contact time (adsorbent dosage = 5.0 mg, C0 = 50.0 mg L−1, T = 298 K, and V = 20.0 mL; RSD = 0.14%~3.45%); (B) fitted using the linear pseudo-first-order kinetic model; (C) fitted using the nonlinear pseudo-first-order kinetic model; (D) fitted using the linear pseudo-second-order kinetic model; and (E) fitted using the nonlinear pseudo-second-order kinetic model.
Molecules 29 03768 g006
Figure 7. (A) Effects of the pH on the Ga3+ adsorption capacity of the GO-MTA composite (adsorbent dosage = 5.0 mg, C0 = 50.0 mg L−1, T = 298 K, V = 20.0 mL, and t = 120 min; RSD = 0.25%~3.04%). (B) Zeta potential of the GO-MTA composite in the pH range of 2.0–6.0.
Figure 7. (A) Effects of the pH on the Ga3+ adsorption capacity of the GO-MTA composite (adsorbent dosage = 5.0 mg, C0 = 50.0 mg L−1, T = 298 K, V = 20.0 mL, and t = 120 min; RSD = 0.25%~3.04%). (B) Zeta potential of the GO-MTA composite in the pH range of 2.0–6.0.
Molecules 29 03768 g007
Figure 8. Effects of the contact time on the GO-MTA composite’s Ga3+ adsorption capacity: (a) effect of the initial concentration and temperature on the adsorption capacity of the GO-MTA (adsorbent dosage = 5.0 mg, C0 = 50.0 mg L−1, T = 298 K, V = 20.0 mL, t = 120 min, and pH = 3.0; RSD = 0.32%~2.12%); (b) fitted curve of the linear Langmuir isothermal model; (c) fitted curve of the nonlinear Langmuir isothermal model; (d) fitted curve of the linear Freundlich isothermal model; and (e) fitted curve of the nonlinear Freundlich isothermal model.
Figure 8. Effects of the contact time on the GO-MTA composite’s Ga3+ adsorption capacity: (a) effect of the initial concentration and temperature on the adsorption capacity of the GO-MTA (adsorbent dosage = 5.0 mg, C0 = 50.0 mg L−1, T = 298 K, V = 20.0 mL, t = 120 min, and pH = 3.0; RSD = 0.32%~2.12%); (b) fitted curve of the linear Langmuir isothermal model; (c) fitted curve of the nonlinear Langmuir isothermal model; (d) fitted curve of the linear Freundlich isothermal model; and (e) fitted curve of the nonlinear Freundlich isothermal model.
Molecules 29 03768 g008
Figure 9. The reusability of the GO-MTA composite.
Figure 9. The reusability of the GO-MTA composite.
Molecules 29 03768 g009
Table 1. Adsorption kinetic parameters for adsorption by MTA.
Table 1. Adsorption kinetic parameters for adsorption by MTA.
Adsorption Kinetic ModelR2k1 (min−1)k2 (min−1)qe (mg g−1)
Linear pseudo-first-order 0.9710.013-9.622
Nonlinear pseudo-first-order 0.6210.139-43.470
Linear pseudo-second-order0.962-0.001555.066
Nonlinear pseudo-second-order0.769-0.003648.156
Table 2. Isothermal parameters for adsorption onto the GO-MTA composite.
Table 2. Isothermal parameters for adsorption onto the GO-MTA composite.
T (K)Langmuir ModelFreundlich Model
R2kLqm (mg g−1)R2kFn
Linear2880.9720.13437.0370.9565.7172.646
2930.9930.58347.6190.9228.5795.039
2980.9970.70347.8470.9318.4684.965
3030.9900.57652.6310.9428.7715.062
3080.9780.47455.5560.9318.8244.727
Nonlinear 2880.9160.13736.5670.9318.7312.803
2930.5210.37746.5600.91324.8775.706
2980.4060.34346.8930.92425.7305.734
3030.7250.55350.1600.93426.5105.656
3080.7150.50755.2600.91426.9245.188
Notes: The units of kL and kF are L mg−1 and mg L1/n g−1 mg−1/n, respectively.
Table 3. Ga3+ adsorption comparison with previously reported adsorbents.
Table 3. Ga3+ adsorption comparison with previously reported adsorbents.
Adsorbentqe (mg g−1)C0 (mg L−1)T (°C)Reference
γ-Al2O3Not provided21025[5]
PAN NFs39.8020030[21]
GO-AHTZT34.725035[23]
A-PSD resin14.6720025[29]
GO/UiO-66-NH2661.6740040[41]
RCONHOH27.6550060[42]
NHPA120.48300RT[43]
CNT/UiO-66-NH2925.4440040[44]
Catechol-modified resin28.7160RT[45]
KIT-6@CA-PEI88.652040[46]
GO-MTA55.65035This work
Notes: A-PSD: amidoxime-grafted polyacrylonitrile-styrene–divinylbenzene; PAN NFs: polyacrylonitrile nanofiber membranes; RCONHOH: polyacrylate-divinylbenzene hydroxamic resins; NHPA: nitrogen-rich functional group biomass carbon aerogel; CNT: carbon nanotube; RT: room temperature; and KIT-6@CA-PEI: mesoporous silica@x-catechol/polyethyleneimine.
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

Zhu, X.; Guo, Y.; Zheng, B. Graphene Oxide Covalently Functionalized with 5-Methyl-1,3,4-thiadiazol-2-amine for pH-Sensitive Ga3+ Recovery in Aqueous Solutions. Molecules 2024, 29, 3768. https://doi.org/10.3390/molecules29163768

AMA Style

Zhu X, Guo Y, Zheng B. Graphene Oxide Covalently Functionalized with 5-Methyl-1,3,4-thiadiazol-2-amine for pH-Sensitive Ga3+ Recovery in Aqueous Solutions. Molecules. 2024; 29(16):3768. https://doi.org/10.3390/molecules29163768

Chicago/Turabian Style

Zhu, Xi, Yong Guo, and Baozhan Zheng. 2024. "Graphene Oxide Covalently Functionalized with 5-Methyl-1,3,4-thiadiazol-2-amine for pH-Sensitive Ga3+ Recovery in Aqueous Solutions" Molecules 29, no. 16: 3768. https://doi.org/10.3390/molecules29163768

Article Metrics

Back to TopTop