Next Article in Journal
NiFe Layered Double Hydroxide Electrocatalyst Prepared via an Electrochemical Deposition Method for the Oxygen Evolution Reaction
Next Article in Special Issue
Synthesis of High-Energy Faceted TiO2 Nanocrystals with Enhanced Photocatalytic Performance for the Removal of Methyl Orange
Previous Article in Journal
Quinoline Derivatives with Different Functional Groups: Evaluation of Their Catecholase Activity
Previous Article in Special Issue
Enhanced Fe(III)/Fe(II) Redox Cycle for Persulfate Activation by Reducing Sulfur Species
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Dithiocarbamate Chitosan Modified SBA-15 for Catalytic Reductive Removal of Vanadium(V)

1
School of Resource and Environmental Sciences, Wuhan University, Wuhan 430072, China
2
School of Environmental Science and Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
3
Key Laboratory of Material Chemistry and Service Failure, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1469; https://doi.org/10.3390/catal12111469
Submission received: 27 October 2022 / Revised: 11 November 2022 / Accepted: 15 November 2022 / Published: 18 November 2022
(This article belongs to the Special Issue Advanced Catalytic Material for Water Treatment)

Abstract

:
We have successfully synthesized dithiocarbamate chitosan modified SBA-15 (CS2C@SBA) composites, with promise in vanadium (V(V)) elimination. Among the three composites using different mass ratios of dithiocarbamate chitosan to SBA-15, CS2C@SBA−3, which had the highest CS2 substitution, showed the best performance on V(V) removal of which the maximum adsorption capacity could achieve 218.00 mg/g at pH 3.0. The adsorption kinetics were best fitted with a pseudo−second order reaction model, suggesting a chemisorption mechanism. Meanwhile, the Langmuir model fitted better with the adsorption isotherm, revealing a monolayer adsorption behavior. Through FTIR and XPS analysis, the functional group −SH was identified as dominating reduction sites on this composite, which reduced 73.1% of V(V) into V(IV) and V(III). The functional group −NH− was the main adsorption site for vanadium species. This reaction followed a catalytic reduction coupled adsorption mechanism reducing most of V(V) into less toxic vanadium species. Furthermore, CS2C@SBA−3 showed great selectivity towards V(V) in the presence of various co−existing ions in synthetic wastewater and real water samples. Moreover, CS2C@SBA−3 could retain a removal efficiency over 90% after five adsorption−desorption cycles. Based on the aforementioned results, we can conclude that CS2C@SBA−3 has great potential to be applied in efficient remediation of vanadium water−pollution.

1. Introduction

Vanadium pollution has become a worldwide problem with the extending process of industrialization, such as steelmaking, battery manufacturing, and medicinal processing [1,2,3]. Vanadium is a potential toxic contaminant that may cause pulmonary tumors and oxidative cell damage if it is over ingested [4]. There exist various valence states of vanadium including +2, +3, +4 and +5 [5,6]. The higher the valence state of vanadium is, the more toxicity and stability it has [7]. Over the past, a variety of methods including ion exchange, surface complexation and electrostatic interaction have been applied for the elimination of V(V). Anion exchange resins such as D201, D314 and 711 were employed for the removal of V(V), achieving a removal efficiency up to 99.0% [8]. Lactococcus raffinolactis was proved to have promising prospects in the elimination of V(V), offering microbial resources for the bioremediation of a V(V)−polluted environment [9]. Among aforementioned traditional technologies for the remediation of V(V), adsorption was the most widely used, due to its high removal efficiency and low cost. For example, surface modified pine bark was investigated with the maximum adsorption capacity of 35.00 mg/g with quaternary nitrogen groups on its surface as the dominant adsorption sites [10]. Kong et al. synthesized a double hydroxide−supported nanoscale zero−valent iron (nZVI@LDH) for the elimination of vanadium, which achieved the maximum adsorption capacity of 93.70 mg/g at initial pH of 3.0 [11].
Chitosan is a natural and linear copolymer comprised of β−(1,4)−glucopyranose units containing randomly distributed N−acetyl−glucosamine and glucosamine residues [12,13]. The application of pristine chitosan to the removal of heavy metals is limited by its low surface area and restricted accessibility of surface functional groups [14,15]. Thus, various physical and chemical synthetic methods have been adopted to modify chitosan into more application−worthy composites with improved adsorption properties [16,17,18]. Due to the presence of abundant heteroatoms and polar functional groups serving as potential complexation sites with metal ions, derivatives of chitosan can be considered as potential remediation agents for the elimination of heavy metals [19]. A dithiol functional terminal may be introduced to chitosan (CS2−chitosan) [20]. The successfully grafted thiol functional groups (−SH) may then serve as reactive sites to reduce vanadium at high valence states into lower valence states. It is well known that the −SH/−SS− redox couple displays a standard oxidation potential of 0.420 V [21,22], while the standard reduction potential of a V(V/IV) couple is 1.420 V [23]. Thus, thiol groups (−SH) are a promising reducing site theoretically to donate electrons to V(V). Furthermore, the abundance of −NH2 and −OH may promote the adsorption process due to their coordination with V(V). Thus, CS2−chitosan may preclude the desorption of vanadium species back into the water matrix due to its folder configuration. Applications of chitosan alone may be limited in neutral or alkaline conditions by the difficulty of convoluted separation procedures, since it is hydrophobic in nature [24].Thus, coating CS2−chitosan on a stable support may be feasible to synthesize a promising heterogeneous adsorbent for practical applications to the remediation of vanadium contaminated wastewater. As is acknowledged, Santa Barbara Amorphous type 15 material (SBA-15) is one of the most popular support materials for the development of heterogeneous composites in which the 2D hexagonal array of uniformly distributed mesopores plays a vital role in the easy diffusion of substrate molecules [25]. Due to the distinctive properties, high surface area and abundance of hydroxyl groups on its surface [26], SBA-15 can compensate for the disadvantages of CS2−chitosan, facilitating the rafting of polymers.
Inspired by the aforementioned facts, we proposed a dithiocarbamate chitosan modified SBA-15 (CS2C@SBA) composite as a heterogeneous remediation agent for the catalytic reduction coupled adsorption of V(V) in this work. During this research, we investigated the impacts of various parameters including pH value, the dosage of CS2C@SBA, contact time, initial concentration of V(V) and co−existing anions on the elimination of V(V). Afterwards, recycling and regeneration studies were carried out. As the mechanism of catalytic reduction of V(V) by −SH functional groups is not clear, a series of characterization techniques such as Fourier transform infrared spectroscopy (FTIR), scanning electron microscope combined with an electron dispersive X-ray (SEM−EDS) and X-ray photoelectron spectroscopy (XPS) were employed to analyze the change of the physicochemical properties of CS2C@SBA before and after the reaction. Three CS2C@SBA materials with distinct ratios of SBA-15 and modified dithiocarbamate chitosan were tested to disclose the structure−performance relationships. Finally, the performance was tested in real waters to further evaluate the potential of CS2C@SBA in future practical applications. By all the aforementioned efforts, we aim to shed light on the adsorption and catalytic reduction mechanism of V(V) by CS2C@SBA, and to confirm the potential of its future applications.

2. Result and Discussion

2.1. Structural Characterizations of Modified CS2C@SBA−1,2,3

Three CS2C@SBA materials which were synthesized with the mass ratios of CS2−chitosan to SBA-15 of 2:1, 1:1 and 1:2 were named as CS2C@SBA−1, CS2C@SBA−2 and CS2C@SBA−3, respectively. The elemental compositions of the synthesized materials, which were confirmed by elemental analyzer, are listed in Table 1. The elemental contents of N (weight percentage) of CS2C@SBA−1, CS2C@SBA−2 and CS2C@SBA−3 were 4.50%, 3.52% and 2.62%, while the contents of S were 1.97%, 1.73% and 1.36%, respectively (Table 1). To calculate the CS2 substitution, the corresponding results of CS2C@SBA−1, CS2C@SBA−2 and CS2C@SBA−3 were 19.19%, 21.52% and 22.72%, respectively (Table 1). Among the three CS2C@SBA, the CS2 substitution of CS2C@SBA−3 was the highest, which indicated that more −CS2 functional groups (22.76%) are grafted with CS2C@SBA−3 composites (Table 1). FTIR spectra also suggested that the −CS2 functional groups were fabricated on the surface successfully. The obvious bands at 1084 and 783 cm−1 were the asymmetric and symmetric stretching of Si−O−Si, respectively, reflecting the successful hybridization of CS2−chitosan and SBA-15 (Figure 1A) [26]. The bands stretching at 3454, 2920, 2873, 1629 and 1465 cm−1 represented the existence of −OH/−NH2/−SH, −CH3, −CH2, −CNH, and −C=S (Figure 1A) [20,27,28,29], respectively, which were same as the stretching bands in the FTIR spectra of CS2−chitosan. Additionally, the intensities of those bands increased along with the increase of the amount of CS2 substitution. The band intensities of asymmetric and symmetric stretching of Si−O−Si increased along with the increase of mass ratio of SBA-15. The pHzpc of SBA-15 and CS2−chitosan were determined as 7.7 and 6.0, respectively (Figure 1B). With the hybrid of SBA-15 and CS2chitosan, the pHzpc of CS2C@SBA−3 changed into 6.9 (Figure 1B). XPS survey spectra (Figure 1C) of CS2C@SBA−3 before the reaction were consistent with the results of elemental analysis (Table 1). All the signals of O, N, C, S and Si were found. XPS spectra of CS2C@SBA−3 confirmed the existence of sulfur signals at the binding energies of 161.7 and 162.8 eV, which indicated the presence of −C=S and −CS− functional groups, respectively (Figure 1D) [30]. According to these characterization results, we can conclude that the CS2chitosan was successfully fabricated on the surface of SBA-15.
The physical properties of SBA-15 and CS2C@SBA−3 are shown in Table S1 and Figure S1. As the successful hybrid of CS2−chitosan, the Brunauer–Emmett–Teller (BET) surface areas of SBA-15 and CS2C@SBA−1, CS2C@SBA−2 and CS2C@SBA−3 were calculated as 632.6, 121.4, 223.4 and 309.2 m2/g (Table S1), respectively. The BET surface area increased with the increasing mass ratio of CS2−chitosan to SBA-15. Furthermore, the pore width increased from 2.04 to 2.28 nm while the pore volume increased from 0.1637 to 0.1762 cm3/g, respectively (Table S1 and Figure S1). CS2C@SBA was mainly micro−mesopore structure (>2 nm). The X-ray diffraction (XRD) patterns revealed the ordered mesoporous structure of SBA-15 and CS2C@SBA−3 (Figure S2). The peak at 23° was assigned for the amorphous SiO2. Typical peaks at 0.8° (100), 1.6° (110) and 1.8° (200) indicated the mesopore structure of SBA-15 [31,32,33,34]. Furthermore, SEM images (Figure 2A–F) clearly showed the surface morphology of SBA-15, CS2−chitosan and CS2C@SBA−3 [35,36]. In the SEM image of CS2C@SBA−3, it can be found that SBA-15 was uniformly wrapped by CS2−chitosan (Figure 2E,F). To sum up, all the analyses above from the different characterizations proved the coating of CS2−chitosan on the surface of SBA-15 [37]. TEM images (Figure S3) of SBA-15 exhibited cylindrical silicate layer morphologies [21,35]. After reaction with CS2−chitosan, TEM images of CS2C@SBA−3 (Figure S3) exhibited a core−shell structure which indicated the successful hybrid of CS2−chitosan and SBA-15 [21]. Besides, after the reaction with V(V), CS2−chitosan was still coated on the surface of SBA-15 (Figure S4). The phenomena suggested a robust nature of CS2C@SBA−3 in the adsorption coupled catalytic reduction processes.

2.2. Adsorption Properties on Vanadium (V) Elimination

2.2.1. Effects of pH

The solution pH is one of the most important influence factors in the removal of heavy metal since it can change not only the surface charge properties of adsorbents but also the distribution of metal ions species [38,39]. As is acknowledged, the distribution of V(V) oxyanions exists in different forms at different solution pH values [40,41]. Taking those facts into consideration, we decided to determine the removal of V(V) by CS2C@SBA−1, CS2C@SBA−2 and CS2C@SBA−3 at a specific pH ranging from 2.0 to 10.0.
Figure 3 showed the effects of pH value on the removal of V(V). CS2C@SBA−1, CS2C@SBA−2 and CS2C@SBA−3 achieved V(V) removal of 62.2%, 72.1% and 98.2% at pH 3.0, respectively (Figure 3). CS2C@SBA−3 demonstrated the highest efficiency in the removal of V(V). Furthermore, the removal of V(V) can be strongly influenced by the mass ratio of SBA-15/CS2−chitosan. With the increase of SBA-15/CS2−chitosan mass ratio, the removal efficiency was remarkably enhanced since there were more functional groups grafted on the surface of the composites, which acted as reactive sites to remove V(V). However, the removal efficiency of V(V) by CS2C@SBA−3 decreased from 98.2% to 4.6% with the increase of solution pH from 2.0 to 10.0, while it decreased from 62.2% and 72.1% to 4.5% and 4.6% for CS2C@SBA−1 and CS2C@SBA−2, respectively (Figure 3). Hence, the optimum pH value for the removal of V(V) was determined as 3.0. The pHzpc of CS2C@SBA−3 is 6.9 (Figure 1B) which suggested a negative surface charge at pH over 6.9 and a positive surface charge at pH under 6.9. Vanadium anions were reported to mainly exist as H2V10O284−, H2VO4− and H3V10O283− at pH 3.0 [2]. Therefore, positively charged CS2C@SBA−3 should have an electrostatic attraction with V(V) oxyanions at solution pH 3.0, favoring the adsorption of V(V) oxyanions. The analysis above suggests a strong co−relationship between pH value and the removal of V(V). An acidic environment favored the immobilization of V(V) by CS2C@SBA−3.

2.2.2. Effects of the Dosage of CS2C@SBA−3

To determine the effects of the dosage of CS2C@SBA−3, a series of different dosages were examined in a 100 ppm V(V) solution. As the dosage increased from 0.1 to 2.0 g/L, the removal efficiency of V(V) increased from 31.2% to 99.7% (Figure 4). This might be attributed to the increase of the density of reactive sites on the surface. However, the removal efficiency of V(V) increased only slightly with increases in dosage from 1.0 to 1.5 and 2.0 g/L. The phenomena suggested that the removal of V(V) attained an equilibrium with the dosage of CS2C@SBA−3 higher than 1.0 g/L. Thus, the adsorption capacity decreased from 312.44 to 49.85 mg/g along with the increase of the dosage of CS2C@SBA−3 (Figure 4). The moles ratio of V(V) to CS2C@SBA-15 is inadequate at higher adsorbent dosages, indicating that the reactive sites on the surface of CS2C@SBA−3 did not reach saturation point in those cases.

2.3. Adsorption Kinetics

Figure 5A–C and Table 2 show the kinetics of the elimination of V(V). The effects of reaction time were investigated in a 100 ppm V(V) solution. Within the initial 5 min, 48.9% of V(V) was removed by CS2C@SBA−3 (Figure 5A). The elimination rate was fast in the first 10 min. These results might be attributed to the abundance of reactive sites on the surface of CS2C@SBA−3 which were quickly enriched with V(V). After 360 min of reaction, an equilibrium was reached, with 98.2% removal of V(V) due to the saturation of reactive sites (Figure 5A). Based on the data, the maximum adsorption capacity was calculated as 98.24 mg/g. The kinetics data were carefully fitted by both pseudo−first−order model and pseudo−second−order models. The regression coefficient (R2) values of pseudo−second−order model were higher (0.9999) compared with pseudo−first−order model (R2 = 0.9851) and Elovich model (R2 = 0.8788) (Figure 5B–D). Hence, the pseudo−second−order kinetic model fitted better than the pseudo−first−order kinetic model and Elovich model did. The phenomena indicated that the elimination of V(V) by CS2C@SBA−3 was primarily directed by the chemisorption [37,40,42,43].

2.4. Adsorption Isotherm

To identify the adsorption interaction and the type of adsorption, adsorption isotherms were investigated at a series of initial V(V) concentrations (10–500 ppm). Figure 6 and Table 2 show the isotherm results of V(V) removal. The V(V) removal efficiency decreased from 98. 6% to 43.6% while the adsorption capacity increased from 8.64 to 218.00 mg/g along with the increase of V(V) concentration from 10 to 500 ppm (Figure 6A). Two different isotherm models including Langmuir and Freundlich ones were used to fit the data [44,45]. From the parameters listed in Table 2, the Langmuir model fitted better than Freundlich models, since their correlation coefficient (R2) were 0.9982 and 0.9218, respectively (Figure 6B and Table 2). From the calculation results of the Langmuir model, Qm of CS2C@SBA−3 is 221.19 mg/g which is appropriately equal to the measured experimental value (218.00 mg/g) (Table 2). The phenomena suggested that the adsorption of V(V) by CS2C@SBA−3 was based on the assumption of monolayer distribution of V(V) on the surface of CS2C@SBA−3 and no lateral interaction between V(V) and CS2C@SBA−3 [46,47]. As long as a molecule occupies the reactive site on CS2C@SBA−3, there is no adsorption that can take place at this site any more.

2.5. Effects of Competing Anions

In real wastewater or groundwater, there are abundant co−existing anions which may have negative impacts on the removal of V (V). The co−existing anions may compete with V(V) for the reactive sites on the surface of composites. Thus, to simulate the real groundwater, the effects of co−existing anions (60 ppm PO43−, 600 ppm SO42−, 400 ppm NO3−, 400 ppm HCO3−) and 60 ppm Cu2+ and ionic strength (0.01, 0.1, and 0.3 M) on the removal of V(V) were studied [11,48].
Figure 7A,B indicate the impacts of various of co−existing anions and ionic strength on the elimination of V(V) by CS2C@SBA−3. CS2C@SBA−3 showed a predominant selectivity for the uptake of V(V). Despite of the presence of the competing anions, over 98.0% of V(V) could be removed from the solution except for two cases in the presence of 600 ppm SO42− and 400 ppm HCO3−. The removal efficiency rates of V(V) were 95.9% and 97.2% in the presence of SO42− and HCO3−, respectively (Figure 7A). This could be attributed to slight competition on reactive sites between V(V) and the two anions. Furthermore, the ionic strength demonstrated only slight impacts on the elimination of V(V) by CS2C@SBA−3 because V(V) removal efficiency could achieve over 98.0% in a wide range of NaCl concentrations from 10 to 300 ppm (Figure 7B). All the aforementioned results suggest a promising performance of CS2C@SBA−3 in the presence of common anions in wastewater or groundwater.

2.6. The Elimination of V(V) in Real Waters

In order to evaluate the practical potential of CS2C@SBA−3, the elimination of V(V) was tested in three types of real water samples including lake, pond and tap waters. At pH 3.0, the removal efficiency of V(V) from lake, pond and tap waters was 99.5%, 99.8% and 99.3% respectively (Figure 7C). The results indicated that CS2C@SBA−3 was highly efficient in real water bodies. However, V(V) removal efficiency by CS2C@SBA−3 decreased to 77.5%, 75.2% and 75.4% in lake, pond and tap waters, respectively, at a local pH of 8.0 (Figure 7C). Hence, we can conclude that pH value has a significant impact on the removal of V(V), which was consistent with our previous results in Section 2.2.1. The dramatic decrease of V(V) removal may be attributed to the electrical repulsion between negatively charged CS2C@SBA−3 and V(V) oxyanions at pH 8.0. The results suggest that CS2C@SBA−3 can be successfully employed for the remediation of V(V) in real wastewater with appropriate adjustment of pH value.

2.7. Mechanism for V(V) Removal by CS2C@SBA−3

SEM−EDS analysis was conducted to investigate the morphology and elemental composition of CS2C@SBA−3 before and after the remediation of V(V). Figure 8A–E and Figure S4 show the structure and morphology of CS2C@SBA−3 after the reaction. Comparing with Figure 2E,F, SBA-15 was still coated with CS2−chitosan uniformly and there was nearly no obvious change on the morphology after treatment with 100 ppm V(V) for 24 h. Table 3, Figures S5 and S6 show the results of EDS elemental mapping of CS2C@SBA−3 before and after the reaction. The results are consistent with the results of elemental analysis which can be considered as solid evidence supporting the successful complexation of the composites (Table 1 and Table 3). Before the reaction, no signal of V was detected (Figure 8F). After treatment with large amounts of vanadium ions, CS2C@SBA-15 was enriched with V which accounted for 12.2 wt% of the composite (Figure 8G and Table 3). The results once again indicated that large amounts of V(V) were adsorbed on the surface of CS2C@SBA−3.
FTIR and XPS measurement was further conducted to uncover the deeper mechanisms at work. Figure 9A–F reveals the results of FTIR spectra and XPS spectra of CS2C@SBA−3 before and after the reaction with V(V). The bands stretching at 3454, 2924, 2854, 1632 and 1465 cm−1 represents the presence of −OH/−NH2/−SH, −CH3, −CH2, −CNH, and −C=S, respectively, on CS2C@SBA−3 after the reaction (Figure 9A) [49,50,51]. Figure 9B, Figures S7 and S8 show the XPS survey spectra of CS2C@SBA−3 before and after the reaction. Comparing these two XPS results, there was a distinct signal of V 2p at 516.8 eV which could prove the adsorption of V(V) on the surface of CS2C@SBA−3 (Figure 9B) [39]. To illustrate the mechanism of the reduction of V(V) by CS2C@SBA−3, XPS spectra of V 2p, V 2p1/2 and V 2p2/3 are provided in Figure 9C, D and E, respectively. There were three peaks at binding energies of 524.0, 516.8, and 513.3 eV, which occupied 26.9%, 66.9% and 6.6% of peak area, respectively (Figure 9C). The three peaks in XPS spectra of V 2p were characteristic of V2O5, V(IV), and V(III) respectively [52,53]. The results indicated that 73.1% of V(V) was reduced to V(IV) and V(III) on the surface of CS2C@SBA−3 (Figure 9C–E).
The successful synthesis of CS2C@SBA−3 introduced organic functional groups including −SH, −OH, −NH2 to the composite which can enable the adsorption coupled reduction of V(V). Figure 1D and Figure 9F demonstrate the results of the XPS spectra of S 2p before and after the elimination of V(V), respectively. Before the reaction with V(V), S 2p XPS peaks at 163.1 and 161.8 eV could be attributed to the sulphur species of CS2−chitosan including −C=S and −CS− respectively (Figure 1D). Their distribution molar percentages were 51.2% and 48.8%. After the reaction with V(V), a new S 2p XPS peak appeared at 163.2 eV (46.3%) which was attributed to the oxidation of −SH to disulfide bonds (−S−S−) (Figure 9F) [21]. The weak peak at 540 cm−1 in FTIR spectra also confirmed the formation of disulfide bonds (Figure 9A) [49,54]. The results suggested that thiol functional groups on the surface of CS2C@SBA−3 donated electrons to V(V) for its reduction. Meanwhile, a new peak appeared at 162.2 eV (38.5%) accounting for the protonated −C=SH+ or −SH2+ due to the doping of H+ on dithiocarbamate (Figure 9F). The formation of protonated sulphur species −C=SH+ or −SH2+ would foster the electrostatic attraction between the residual V(V) anions and CS2C@SBA−3, facilitating the nearly total elimination of V(V).
To identify other factors which might support the elimination of V(V), N 1s and O 1s XPS spectra were de−convoluted. Figure 10A–D show the XPS spectra of N 1s and O 1s before and after the elimination of V(V). Before the reaction with V(V), N 1s XPS peaks appearing at binding energies of 402.4, 400.2 and 399.3 eV could be associated to N−SiO2, −NH2 and −NH−, respectively (Figure 10A) [55]. After the reaction with V(V), the corresponding binding energies of N 1s peaks changed to 401.5, 399.4 and 398.7 eV, respectively (Figure 10B). The relative amounts of N−SiO2 and −NH2 increased from 4.4% and 15.0% to 41.5% and 55.3%, respectively. In contrast to the two functional groups, the relative amount of −NH− dramatically decreased from 80.6% to 3.2% after the elimination of V(V). The results suggest that large amounts of V(V) occupied −NH−. We could conclude that −NH− constituted the dominant adsorption site for V(V) among the nitrogen functional groups. Furthermore, O 1s XPS spectra located at binding energies of 533.4, 532.6 and 531.8 eV corresponded to C−O−C, −CH2OH and −C−O−Si, respectively (Figure 10C,D) [56]. The new peak at 529.8 eV indicated the presence of V2O5, as illustrated in Figure 9D [53]. After the elimination of V(V), the relative density of −OH decreased from 57.6% to 51.5%. The decrease of the density of –OH was most likely attributable to the formation of hydrogen bonds between the V species and hydroxyl groups. As discussed above, the reaction between the CS2C@SBA−3 and V(V) species is a catalytic reduction process in which most of V(V) (73.1%) is reduced into V(IV) and V(III) species which are less toxic. A small amount of V(V) was transformed into V(V) oxide, V2O5. Meanwhile, the abundance of functional groups −NH− enhanced the adsorption between vanadium species and CS2C@SBA−3. The functional group −SH is the dominant reactive site for reducing V(V) and was oxidized into disulfide bonds (−S−S−) in this catalytic reduction process.

2.8. Recycling Application

As recyclability is one of the most important factors to judge the practical potential of these findings, a series of recycling application experiments were conducted. Figure 11 shows the removal efficiency of V(V) in five consecutive adsorption–desorption cycles. The removal efficiency of V(V) by CS2C@SAB−3 decreased from 98.2% to 90.8% from the first to the fifth cycle of application. The results suggested that a V(V) removal up to 90.0% could be achieved after five adsorption−desorption cycles. Thus, CS2C@SAB−3 can be regarded as a promising composite for the elimination of V(V) via catalytic reduction coupled adsorption process with satisfying reusability and good practical potential.

3. Materials and Methods

3.1. Materials

Deionized water (DI water, 18.2 MΩ∙cm−1) was obtained from Lichen water purifier UPTA. Chitosan was purchased from Meryer (Shanghai, China) without any purification. NaOH (≥96.0%), HCl (38.0%), NH3·H2O (28%), Na3PO4·12H2O (98%), NaNO3 (99.0%), NaHCO3 (≥99.8%), Na2S2O3 (99%) and CS2 (99.9%) were purchased from Hushi (Shanghai, China). Glutaraldehyde (50% v/v) and sodium metavanadate (NaVO3, 99.9%) were obtained from Aladdin. All the chemicals mentioned above were used at analytical grade without further purification.

3.2. Preparation of SBA-15, CS2−Chitosan and CS2C@SBA

SBA-15 and modified dithiocarbamate chitosan (CS2-Chitosan) was synthesized based on the methods in previous literature [49]. Details can be found in Texts S1 and S2 respectively. CS2C@SBA was synthesized by the following method. Initially, 0.60 g CS2−chitosan was dissolved in 100 mL acetic acid solution (1.2% v/v). Then, different dosages of SBA-15 were added (mass ratios SBA-15: CS2−chitosan = 1:2, 1:1 and 2:1). According to these mass ratios, CS2C@SBA composites are named as CS2C@SBA−1, CS2C@SBA−2 and [57,58] CS2C@SBA−3, respectively. Afterwards, the above mixture was stirred for 30 min at 313 K, followed by the addition of 0.50 mL glutaraldehyde (50.0% v/v) and further stirred for 2 h. After that, the pH of the mixture was adjusted to 9.0 by adding 0.1M NaOH, and stirred for 30 min. Finally, the solid suspension was filtered, washed several times with distilled water and dried at 333 K overnight.

3.3. Characterizations

Fourier Transform Infrared Spectra (FTIR) was measured by Nicolet 5700 spectrometer through the KBr pellet method over a wavelength range from 400 to 4000 cm−1 under a 4 cm−1 resolution. The Scanning Electron Microscopy (SEM) was performed on SEM Virion 200 combined with an electron dispersive X-ray (EDS) analyzer. The Transmission Electron Microscopy (TEM) was conducted on JEM−2100. The elemental composition (C, H, N and S wt%) of the materials was analyzed by Vario MICRO cube elemental analyzer (German). The Brunauer–Emmett–Teller (BET) pore properties of materials were measured via Belsorp−minill. Before the N2 adsorption−desorption measurement, the composite was dried and degassed at 373 K under vacuum for 24 h. The X-ray diffraction (XRD, MiniFlex600, Japan) was conducted with radiation of Cu Kα and scanning rate of 0.2°/min from 0.6° to 2.5° and 5°/min from 10° to 80°. Zeta potentials of the materials were measured by the Zetasizer Nano series at different pH (2.0–9.0). Firstly, 150 mg of composite was added into a plastic conical flask containing 65 mL DI water. Then the flask was mechanically shaken for 24 h. The resultant suspension was collected for zeta potentials determination in 0.01 mM NaCl solution. The pH was adjusted with 0.01 mM NaOH and HCl. The point of zero charge (pHZPC) was calculated by the zeta potentials under different pHs. The X-ray photoelectron spectroscopy (XPS) equipped with Al/Mg Kα X-ray source with 30 eV pass energy in 0.5 eV step over an area of 650 mm × 650 mm to the samples was carried out (ESCALAB 250Xi) at vacuum of 10−9 Torr, and the results were corrected by a reference of the C 1s peak from adventitious carbon at 284.8 eV and fitted with the Shirley method of background subtraction on XPSpeak4.1 Software.

3.4. Adsorption Experiments

The stock solution of V(V) (1000 ppm) was prepared by dissolving NaVO3 with DI water. The effects of pH (2.0–10.0) on vanadium removal was conducted by adding 30 mg CS2C@SBA composites into 30 mL of V(V) solution (100 ppm) under stirring (200 rpm) at normal temperature for 24 h. The pH of solutions was adjusted by 0.1 M HCl and NaOH. The effect of adsorbent dosage was conducted in 25 mL V(V) solution (100 ppm) with varying adsorbent dosages (0.1, 0.3, 0.5, 1.0, 1.5 and 2 g/L). The kinetics experiment was conducted by adding 30 mg CS2C@SBA composites into 30 mL of V(V) solution (100 ppm) at pH 3.0 under stirring (200 rpm) at 298 K. The contact time was studied from 5 min to 12 h. The adsorption isotherm experiment was conducted by adjusting the concentration of V(V) from 10 to 500 ppm adding 1 g/L CS2C@SBA−3. In all experiments, the residual concentration of total V was measured by Inductively Coupled Plasma − Atomic Emission Spectroscopy (ICP−AES, Agilent 5110). Adsorption capacity (qt) (mg/g) and percentage removal (%) were calculated by the following equations [41,59,60].
q t = C 0 C t V m
% R e m o v a l = C 0 C t C 0 × 100
where C0 and Ct are, respectively, the initial concentration and equilibrium concentration of total V (ppm); V is the solution volume (mL) while m is the adsorbent amount (g).
The results of the adsorption kinetic data were fitted by two models, pseudo−first−order, pseudo−second−order and Elovich models, as indicated below [10,61,62].
ln q e q t = ln q e k 1 t
t q t = 1 k 2 q e 2 + t q e
q t = 1 β ln 1 + β t
where qe and qt are, respectively, the adsorbed capacity of V(V) on CS2C@SBA−3 at equilibrium time and time t. k1 (1/min) and k2 (g/mg·min) are, respectively, the adsorption rate constants for pseudo−first−order and pseudo−second−order models. α (mg·min/g) is the initial rate constant and β (mg/g) is the desorption constant for Elovich model;
To determine the adsorption isotherm, Langmuir and Freundlich models are fitted to calculate the adsorption data, as indicated as the following equations [63]:
C e q e = 1 q m K L + C e q m
ln q e = ln K f + 1 n ln C e
where qe and qm are, respectively, the adsorption capacity at equilibrium and the maximum adsorption capacity of V(V) by CS2C@SBA−3. Ce (ppm) is the concentration of V(V) at equilibrium. KL (L/mg) and Kf ((mg/g)/(mg/L)1/n) are, respectively, the Langmuir and Freundlich constants.

3.5. Regeneration and Recycle Experiments

To explore the recyclability of the material, the desorption was performed by using 25 mL NaOH solution (0.50 M) as the eluent at 298 K with 1 g/L CS2C@SBA−3, and then using 0.01 M Na2S2O3 solution to regenerate the oxidized disulfide bonds. The regenerated CS2C@SBA−3 was employed to eliminate V(V) again, followed by the regeneration procedures. The experiment was repeated for five cycles. During the recycle experiments, the removal of V(V) was measured.

3.6. Effects of Groundwater Impurities

To investigate the effects of co−existing anions (60 ppm PO43−, 600 ppm SO42−, 400 ppm NO3, 400 ppm HCO3) and 60 ppm Cu2+, based on the typical concentration of common ions in ground water, 1 g/L CS2C@SBA−3 was tested in 100 ppm V(V) solution at pH 3.0 for 24 h at 298 K. The effects of ionic strength were studied in 100 ppm V(V) solution at pH 3.0 for 24 h at 298 K with varying NaCl concentrations (10, 100 and 300 mM)

3.7. V(V) Elimination Assays in Spiked Real Water

The applicability of the adsorbent was also tested in spiked real water samples (10 ppm V(V)) including lake water, pond water and tap water collected from Wuhan at pH 3.0 and at their local pH 8.0. The physical and chemical properties of these samples are listed in Table S2.

4. Conclusions

In this study, dithiocarbamate chitosan modified SBA-15 (CS2C@SBA) was successfully synthesized via a facile and robust procedure. Among the three CS2C@SBA composites, CS2C@SBA−3 which embraced the highest mass ratio of dithiocarbamate chitosan showed the most efficient performance in the elimination of V(V). At 298 K, a V(V) removal efficiency of 98.2% was achieved by 1 g/L CS2C@SBA−3 in a solution containing 100 ppm of V(V). In that case, the adsorption capacity by CS2C@SBA−3 was 218.00 mg/g. The adsorption kinetics fitted better with pseudo−second−order model (R2 = 0.9999), indicating that the elimination of V(V) by CS2C@SBA−3 was primarily triggered by chemisorption. The adsorption isotherm followed the Langmuir model (R2 = 0.9981), revealing the monolayer adsorption on the surface of CS2C@SBA−3. FTIR, SEM−EDS and XPS characterization results demonstrated that the functional group −SH played a vital role, acting as a proton−coupled electron donor to reduce 73.5% V(V) into V(IV) and V(III) and transformed residual V(V) into V(V) oxide V2O5. Via the reduction of high−valent vanadium, its toxicity was reduced significantly. Furthermore, CS2C@SBA−3 displayed high selectivity towards V(V) elimination in the presence of various co−existing ions in synthetic and spiked real water samples which could be attributed to the abundant −NH− on its surface. Hence, we can conclude that CS2C@SBA−3 is a promising composite to eliminate V(V) via an adsorption coupled catalytic reduction mechanism, with good practical potentials. CS2C@SBA−3 can be applied in the reductive elimination of high−valent heavy metals in efficient remediation of water pollution.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12111469/s1, Text S1: Details of synthesis procedures of SBA-15; Text S2: Details of synthesis procedures of CS2−chitosan; Figure S1: Nitrogen sorption isotherms of SBA-15 and CS2C@SBA−3; Figure S2: The XRD patterns of preparation composite; Figure S3: TEM images of SBA-15 and CS2C@SBA−3; Figure S4: SEM images of CS2C@SBA−3 after reaction with V(V); Figure S5: EDS elemental mapping images of CS2C@SBA−3 before reaction with V(V); Figure S6: EDS elemental mapping images of CS2C@SBA−3 after reaction with V(V); Figure S7: The XPS survey spectra of CS2C@SBA−3 before reaction; Figure S8: The XPS survey spectra of CS2C@SBA−3 after reaction; Table S1: Porous properties of SBA-15 and CS2C@SBA−3; Table S2: Physico-chemical properties of water samples used in this study.

Author Contributions

Y.H.: Data curation, Methodology, Validation, Investigation, Writing—original draft. J.W.: Methodology, Supervision, Writing—review and editing. M.L.: Methodology. Z.Y.: Funding acquisition, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 21573163) and the Natural Science Foundation of Hubei Province (No. 2015CFA017).

Acknowledgments

We thank the Analytical and Testing Centre of Wuhan University for help with the characterization.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Abbreviations

Santa Barbara Amorphous type 15 material (SBA-15); Dithiocarbamate chitosan (CS2−chitosan); Dithiocarbamate chitosan modified SBA-15 (CS2C@SBA); Vanadium(V) [V(V)]; Fourier Transform Infrared Spectra (FTIR); Scanning Electron Microscopy with an electron dispersive X-ray (SEM-EDS); Brunauer-Emmett-Teller (BET); X-ray diffraction (XRD); Transmission Electron Microscopy (TEM); X-ray photoelectron spectroscopy (XPS).

References

  1. Wang, Z.L.; Zhang, B.G.; He, C.; Shi, J.X.; Wu, M.X.; Guo, J.H. Sulfur−based Mixotrophic Vanadium (V) Bio−reduction towards Lower Organic Requirement and Sulfate Accumulation. Water Res. 2021, 189, 116655. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, P.; Hu, J.; Wang, Y.D.; Liu, T.Y. Enhanced elimination of V5+ in wastewater using zero−valent iron activated by ball milling: The overlooked crucial roles of energy input and sodium chloride. J. Hazard. Mater. 2022, 435, 129050. [Google Scholar] [CrossRef]
  3. Sun, F.J.; Liu, M.; Yuan, B.; He, J.; Wu, P.; Liu, C.J.; Jiang, W. Separation of vanadium and chromium by selective adsorption by titanium−based microspheres. Chem. Eng. J. 2022, 450, 138039. [Google Scholar] [CrossRef]
  4. Aregay, G.G.; Ali, J.; Shahzad, A.; Ifthikar, J.; Oyekunle, D.T.; Chen, Z. Application of layered double hydroxide enriched with electron rich sulfide moieties (S2O4(2−)) for efficient and selective removal of vanadium (V) from diverse aqueous medium. Sci. Total Environ. 2021, 792, 148543. [Google Scholar] [CrossRef] [PubMed]
  5. Schiffer, S.; Liber, K. Toxicity of aqueous vanadium to zooplankton and phytoplankton species of relevance to the athabasca oil sands region. Ecotoxicol. Environ. Saf. 2017, 137, 1–11. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, H.S.; Chen, N.; Feng, C.P.; Deng, Y. Synchronous microbial V(V) reduction and denitrification using corn straw as the sole carbon source. Sci. Total Environ. 2022, 839, 156343. [Google Scholar] [CrossRef]
  7. Chen, L.; Liu, J.R.; Hu, W.F.; Gao, J.; Yang, J.Y. Vanadium in soil−plant system: Source, fate, toxicity, and bioremediation. J. Hazard. Mater. 2021, 405, 124200. [Google Scholar] [CrossRef]
  8. Zhu, X.B.; Li, W.; Zhang, Q.; Zhang, C.X.; Chen, L.J. Separation characteristics of vanadium from leach liquor of red mud by ion exchange with different resins. Hydrometallurgy 2018, 176, 42–48. [Google Scholar] [CrossRef]
  9. Zhang, B.G.; Li, Y.N.; Fei, Y.M.; Cheng, Y.T. Novel Pathway for Vanadium(V) Bio−Detoxification by Gram−Positive Lactococcus raffinolactis. Environ. Sci. Technol. 2021, 55, 2121–2131. [Google Scholar] [CrossRef] [PubMed]
  10. Zhang, R.C.; Leiviska, T. Surface modification of pine bark with quaternary ammonium groups and its use for vanadium removal. Chem. Eng. J. 2020, 385, 123967. [Google Scholar] [CrossRef]
  11. Kong, X.; Chen, J.; Tang, Y.; Lv, Y.; Chen, T.; Wang, H. Enhanced removal of vanadium(V) from groundwater by layered double hydroxide−supported nanoscale zerovalent iron. J. Hazard. Mater. 2020, 392, 122392. [Google Scholar] [CrossRef] [PubMed]
  12. Kong, D.; Foley, S.R.; Wilson, L.D. An Overview of Modified Chitosan Adsorbents for the Removal of Precious Metals Species from Aqueous Media. Molecules 2022, 27, 978. [Google Scholar] [CrossRef] [PubMed]
  13. Kazachenko, A.S.; Akman, F.; Malyar, Y.N.; Issaoui, N.; Vasilieva, N.Y.; Karacharov, A.A. Synthesis optimization, DFT and physicochemical study of chitosan sulfates. J. Mol. Struc. 2021, 1245, 131083. [Google Scholar] [CrossRef]
  14. Crini, G.; Lichtfouse, E.; Wilson, L.D.; Morin−Crini, N. Conventional and non−conventional adsorbents for wastewater treatment. Environ. Chem. Lett. 2019, 17, 195–213. [Google Scholar]
  15. Steiger, B.G.K.; Wilson, L.D. Modular Chitosan−Based Adsorbents for Tunable Uptake of Sulfate from Water. Int. J. Mol. Sci. 2020, 21, 7130. [Google Scholar] [CrossRef]
  16. Nuryono, N.; Miswanda, D.; Sakti, S.C.W.; Rusdiarso, B.; Krisbiantoro, P.A.; Utami, N.; Otomo, R.; Kamiya, Y. Chitosan−functionalized natural magnetic particle@silica modified with (3−chloropropyl)trimethoxysilane as a highly stable magnetic adsorbent for gold(III) ion. Mater. Chem. Phys. 2020, 255, 123507. [Google Scholar] [CrossRef]
  17. Zhuang, S.T.; Zhu, K.K.; Wang, J.L. Fibrous chitosan cellulose composite as an efficient adsorbent for Co(II) removal. J. Clean. Prod. 2021, 285, 124911. [Google Scholar] [CrossRef]
  18. Bui, T.H.; Lee, W.; Jeon, S.B.; Kim, K.W.; Lee, Y. Enhanced Gold(III) adsorption using glutaraldehyde−crosslinked chitosan beads: Effect of crosslinking degree on adsorption selectivity, capacity, and mechanism. Sep. Purif. Technol. 2020, 248, 116989. [Google Scholar]
  19. Bilican, I.; Pekdemir, S.; Onses, M.S.; Akyuz, L.; Altuner, E.M.; Koc−Bilican, B.; Zang, L.S.; Mujtaba, M.; Mulercikas, P.; Kaya, M. Chitosan Loses Innate Beneficial Properties after Being Dissolved in Acetic Acid: Supported by Detailed Molecular Modeling. ACS Sustain. Chem. Eng. 2020, 8, 18083–18093. [Google Scholar] [CrossRef]
  20. Yin, Z.; Qiu, D.; Zhang, M.Y. Molecular level study of cadmium adsorption on dithiocarbamate modified chitosan. Environ. Pollut. 2021, 271, 116322. [Google Scholar] [CrossRef] [PubMed]
  21. Ifthikar, J.; Chen, Z.; Chen, Z.; Jawad, A. A self−gating proton−coupled electron transfer reduction of hexavalent chromium by core−shell SBA−Dithiocarbamate chitosan composite. J. Hazard. Mater. 2020, 384, 121257. [Google Scholar] [CrossRef]
  22. Ifthikar, J.; Zhao, M.; Shahzad, A.; Shahib, I.I.; Wang, J.; Wang, H.B.; Sellaoui, L.; Chen, Z.Q.; Chen, Z.L. Recyclable process modeling study of hexavalent chromium elimination by thiol−based electron donor: Implications for practical applicability. J. Environ. Chem. Eng. 2021, 9, 105645. [Google Scholar] [CrossRef]
  23. Roncevic, S.; Nemet, I.; Ferri, T.Z.; Matkovic−Calogovic, D. Characterization of nZVI nanoparticles functionalized by EDTA and dipicolinic acid: A comparative study of metal ion removal from aqueous solutions. RSC Adv. 2019, 9, 31043–31051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Ifthikar, J.; Jiao, X.; Ngambia, A.; Wang, T.; Khan, A.; Jawad, A.; Xue, Q.; Liu, L.; Chen, Z. Facile One−Pot Synthesis of Sustainable Carboxymethyl Chitosan − Sewage Sludge Biochar for Effective Heavy Metal Chelation and Regeneration. Bioresour. Technol. 2018, 262, 22–31. [Google Scholar] [CrossRef]
  25. Verma, P.; Kuwahara, Y.; Mori, K.; Raja, R.; Yamashita, H. Functionalized mesoporous SBA-15 silica: Recent trends and catalytic applications. Nanoscale 2020, 12, 11333–11363. [Google Scholar] [CrossRef] [PubMed]
  26. Qiao, Y.; Said, N.; Rauser, M.; Yan, K.; Qin, F.; Theyssen, N.; Leitner, W. Preparation of SBA-15 supported Pt/Pd bimetallic catalysts using supercritical fluid reactive deposition: How do solvent effects during material synthesis affect catalytic properties? Green Chem. 2017, 19, 977–986. [Google Scholar] [CrossRef]
  27. Kazachenko, A.S.; Akman, F.; Vasilieva, N.Y.; Issaoui, N.; Malyar, Y.N.; Kondrasenko, A.A.; Borovkova, V.S.; Miroshnikova, A.V.; Kazachenko, A.S.; Al−Dossary, O.; et al. Catalytic Sulfation of Betulin with Sulfamic Acid: Experiment and DFT Calculation. Int. J. Mol. Sci. 2022, 23, 1602. [Google Scholar] [CrossRef] [PubMed]
  28. Kazachenko, A.S.; Akman, F.; Vasilieva, N.Y.; Malyar, Y.N.; Fetisova, O.Y.; Lutoshkin, M.A.; Berezhnaya, Y.D.; Miroshnikova, A.V.; Issaoui, N.; Xiang, Z.Y. Sulfation of Wheat Straw Soda Lignin with Sulfamic Acid over Solid Catalysts. Polymers 2022, 14, 3000. [Google Scholar] [CrossRef] [PubMed]
  29. Kazachenko, A.; Akman, F.; Medimagh, M.; Issaoui, N.; Vasilieva, N.; Malyar, Y.N.; Sudakova, I.G.; Karacharov, A.; Miroshnikova, A.; Al−Dossary, O.M. Sulfation of Diethylaminoethyl−Cellulose: QTAIM Topological Analysis and Experimental and DFT Studies of the Properties. Acs Omega 2021, 6, 22603–22615. [Google Scholar] [CrossRef]
  30. Zhang, L.; Guo, D.; Tantai, X.; Jiang, B.; Sun, Y.; Yang, N. Synthesis of Three−Dimensional Hierarchical Flower−Like Mg–Al Layered Double Hydroxides with Excellent Adsorption Performance for Organic Anionic Dyes. Trans. Tianjin Univ. 2020, 27, 394–408. [Google Scholar] [CrossRef]
  31. Hongmanorom, P.; Ashok, J.; Zhang, G.H.; Bian, Z.F.; Wai, M.H.; Zeng, Y.Q.; Xi, S.B.; Borgna, A.; Kawi, S. Enhanced performance and selectivity of CO2 methanation over phyllosilicate structure derived Ni-Mg/SBA-15 catalysts. Appl. Catal. B-Environ. 2021, 282, 119564. [Google Scholar] [CrossRef]
  32. Alkafajy, A.M.; Albayati, T.M. High performance of magnetic mesoporous modification for loading and release of meloxicam in drug delivery implementation. Mater. Today Commun. 2020, 23, 100890. [Google Scholar] [CrossRef]
  33. Betiha, M.A.; Moustafa, Y.M.; El−Shahat, M.F.; Rafik, E. Polyvinylpyrrolidone−Aminopropyl−SBA-15 schiff Base hybrid for efficient removal of divalent heavy metal cations from wastewater. J. Hazard. Mater. 2020, 397, 122675. [Google Scholar] [CrossRef] [PubMed]
  34. Daoura, O.; Fornasieri, G.; Boutros, M.; El Hassan, N.; Beaunier, P.; Thomas, C.; Selmane, M.; Miche, A.; Sassoye, C.; Ersen, O.; et al. One-pot prepared mesoporous silica SBA-15-like monoliths with embedded Ni particles as selective and stable catalysts for methane dry reforming. Appl. Catal. B-Environ. 2021, 280, 119417. [Google Scholar] [CrossRef]
  35. Shimon, D.; Chen, C.H.; Lee, J.J.; Didas, S.A.; Sievers, C.; Jones, C.W.; Hayes, S.E. (15)N Solid State NMR Spectroscopic Study of Surface Amine Groups for Carbon Capture: 3−Aminopropylsilyl Grafted to SBA-15 Mesoporous Silica. Environ. Sci. Technol. 2018, 52, 1488–1495. [Google Scholar] [CrossRef]
  36. Cao, T.; Li, Z.; Xiong, Y.; Yang, Y.; Xu, S.; Bisson, T.; Gupta, R.; Xu, Z. Silica−Silver Nanocomposites as Regenerable Sorbents for Hg(0) Removal from Flue Gases. Environ. Sci. Technol. 2017, 51, 11909–11917. [Google Scholar] [CrossRef]
  37. Dong, W.C.; Liu, J.; Hao, J.M.; Zeng, Y. Adsorption of DTC−CTS on sphalerite (110) and Cu−activated sphalerite (110) surfaces: A DFT study. Appl. Surf. Sci. 2021, 551, 149466. [Google Scholar] [CrossRef]
  38. Ifthikar, J.; Shahib, I.I.; Sellaoui, L.; Jawad, A.; Zhao, M.M.; Chen, Z.Q.; Chen, Z.L. pH tunable anionic and cationic heavy metal reduction coupled adsorption by thiol cross−linked composite: Physicochemical interpretations and fixed−bed column mathematical model study. Chem. Eng. J. 2020, 401, 126041. [Google Scholar] [CrossRef]
  39. Ifthikar, J.; Wang, J.; Wang, Q.L.; Wang, T.; Wang, H.B.; Khan, A.; Jawad, A.; Sun, T.T.; Jiao, X.; Chen, Z.Q. Highly Efficient Lead Distribution by Magnetic Sewage Sludge Biochar: Sorption Mechanisms and Bench Applications. Bioresour. Technol. 2017, 238, 399–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Dong, Y.B.; Lin, H.; Zhao, Y.M.; Menzembere, E. Remediation of vanadium−contaminated soils by the combination of natural clay mineral and humic acid. J. Clean. Prod. 2021, 279, 123874. [Google Scholar] [CrossRef]
  41. Gogoi, H.; Zhang, R.C.; Matusik, J.; Leiviska, T.; Ramo, J.; Tanskanen, J. Vanadium removal by cationized sawdust produced through iodomethane quaternization of triethanolamine grafted raw material. Chemosphere 2021, 278, 130445. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, L.; Zhu, Y.Y.; Luo, H.Q.; Yang, J.Y. Characteristic of adsorption, desorption, and co−transport of vanadium on humic acid colloid. Ecotox. Environ. Safe. 2020, 190, 110087. [Google Scholar] [CrossRef]
  43. Zhu, X.B.; Li, W.; Zhang, C.X. Extraction and removal of vanadium by adsorption with resin 201*7 from vanadium waste liquid. Environ. Res. 2020, 180, 108865. [Google Scholar] [CrossRef] [PubMed]
  44. Gan, C.D.; Liu, M.; Lu, J.; Yang, J.Y. Adsorption and Desorption Characteristics of Vanadium (V) on Silica. Water Air Soil Poll. 2020, 231, 10. [Google Scholar] [CrossRef]
  45. He, C.; Zhang, B.G.; Lu, J.P.; Qiu, R. A newly discovered function of nitrate reductase in chemoautotrophic vanadate transformation by natural mackinawite in aquifer. Water Res. 2021, 189, 116664. [Google Scholar] [CrossRef]
  46. Li, M.; Zhang, B.G.; Zou, S.Q.; Liu, Q.S.; Yang, M. Highly selective adsorption of vanadium (V) by nano−hydrous zirconium oxide−modified anion exchange resin. J. Hazard. Mater. 2020, 384, 121386. [Google Scholar] [CrossRef]
  47. Zhang, B.G.; Jiang, Y.F.; Zuo, K.C.; He, C.; Dai, Y.R.; Ren, Z.J. Microbial vanadate and nitrate reductions coupled with anaerobic methane oxidation in groundwater. J. Hazard. Mater. 2020, 382, 121228. [Google Scholar] [CrossRef]
  48. Aregay, G.G.; Ali, J.; Du, Y.S.; Shahzad, A.; Chen, Z.Q. Efficient and selective removal of chromium (VI) by sulfide assembled hydrotalcite compounds through concurrent reduction and adsorption processes. J. Mol. Liq. 2019, 294, 111532. [Google Scholar] [CrossRef]
  49. Yong, S.K.; Skinner, W.M.; Bolan, N.S.; Lombi, E.; Kunhikrishnan, A.; Ok, Y.S. Sulfur crosslinks from thermal degradation of chitosan dithiocarbamate derivatives and thermodynamic study for sorption of copper and cadmium from aqueous system. Environ. Sci. Pollut. Res. 2016, 23, 1050–1059. [Google Scholar] [CrossRef]
  50. Kazachenko, A.S.; Vasilieva, N.Y.; Fetisova, O.Y.; Sychev, V.V.; Elsuf’ev, E.V.; Malyar, Y.N.; Issaoui, N.; Miroshnikova, A.V.; Borovkova, V.S.; Kazachenko, A.S.; et al. New reactions of betulin with sulfamic acid and ammonium sulfamate in the presence of solid catalysts. Biomass Convers. Biorefinery 2022. [Google Scholar] [CrossRef]
  51. Kazachenko, A.S.; Akman, F.; Abdelmoulahi, H.; Issaoui, N.; Malyar, Y.N.; Al−Dossary, O.; Wojcik, M.J. Intermolecular hydrogen bonds interactions in water clusters of ammonium sulfamate: FTIR, X-ray diffraction, AIM, DFT, RDG, ELF, NBO analysis. J. Mol. Liq. 2021, 342, 117475. [Google Scholar] [CrossRef]
  52. Fan, K.; Chen, H.; Ji, Y.; Huang, H.; Claesson, P.M.; Daniel, Q.; Philippe, B.; Rensmo, H.; Li, F.; Luo, Y.; et al. Nickel−vanadium monolayer double hydroxide for efficient electrochemical water oxidation. Nat. Commun. 2016, 7, 11981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Silversmit, G.; Depla, D.; Poelman, H.; Marin, G.B.; De Gryse, R. Determination of the V2p XPS binding energies for different vanadium oxidation states (V5+ to V0+). J. Electron Spectrosc. Relat. Phenom. 2004, 135, 167–175. [Google Scholar] [CrossRef]
  54. Brandt, N.N.; Chikishev, A.Y.; Kruzhilin, V.N. Raman study of the cleavage of disulphide bonds in albumin, chymotrypsin, and thrombin. Vib. Spectrosc. 2017, 89, 75–80. [Google Scholar] [CrossRef]
  55. Das, S.; Chakraborty, P.; Ghosh, R.; Paul, S.; Mondal, S.; Panja, A.; Nandi, A.K. Folic Acid−Polyaniline Hybrid Hydrogel for Adsorption/Reduction of Chromium(VI) and Selective Adsorption of Anionic Dye from Water. ACS Sustain. Chem. Eng. 2017, 5, 9325–9337. [Google Scholar] [CrossRef]
  56. Veerakumar, P.; Thanasekaran, P.; Lin, K.−C.; Liu, S.−B. Biomass Derived Sheet−like Carbon/Palladium Nanocomposite: An Excellent Opportunity for Reduction of Toxic Hexavalent Chromium. ACS Sustain. Chem. Eng. 2017, 5, 5302–5312. [Google Scholar] [CrossRef]
  57. Kazachenko, A.S.; Medimagh, M.; Issaoui, N.; Al−Dossary, O.; Wojcik, M.J.; Kazachenko, A.S.; Miroshnokova, A.V.; Malyar, Y.N. Sulfamic acid/water complexes (SAA−H2O((1−8))) intermolecular hydrogen bond interactions: FTIR,X-ray, DFT and AIM analysis. J. Mol. Struct. 2022, 1265, 133394. [Google Scholar] [CrossRef]
  58. Levdansky, A.V.; Vasilyeva, N.Y.; Malyar, Y.N.; Kondrasenko, A.A.; Fetisova, O.Y.; Kazachenko, A.S.; Levdansky, V.A.; Kuznetsov, B.N. An Efficient Method of Birch Ethanol Lignin Sulfation with a Sulfaic Acid−Urea Mixture. Molecules 2022, 27, 6356. [Google Scholar] [CrossRef]
  59. He, W.Y.; Liao, W.; Yang, J.Y.; Jeyakumar, P.; Anderson, C. Removal of vanadium from aquatic environment using phosphoric acid modified rice straw. Bioremediation J. 2020, 24, 80–89. [Google Scholar] [CrossRef]
  60. Yang, J.Y.; Luo, H.Q.; Zhu, Y.Y.; Yu, Y.Q.; He, W.Y.; Wu, Z.Z.; Wang, B. Adsorption-desorption and co-migration of vanadium on colloidal kaolinite. Environ. Sci. Pollut. R. 2020, 27, 17910–17922. [Google Scholar] [CrossRef]
  61. Yap, P.L.; Auyoong, Y.L.; Hassan, K.; Farivar, F.; Tran, D.N.H.; Ma, J.; Losic, D. Multithiol functionalized graphene bio−sponge via photoinitiated thiol−ene click chemistry for efficient heavy metal ions adsorption. Chem. Eng. J. 2020, 395, 124965. [Google Scholar] [CrossRef]
  62. Abukhadra, M.R.; Bakry, B.M.; Adlii, A.; Yakout, S.M.; El−Zaidy, M.E. Facile conversion of kaolinite into clay nanotubes (KNTs) of enhanced adsorption properties for toxic heavy metals (Zn2+, Cd2+, Pb2+, and Cr6+) from water. J. Hazard. Mater. 2019, 374, 296–308. [Google Scholar] [CrossRef] [PubMed]
  63. Shahat, A.; Hassan, H.M.A.; Azzazy, H.M.E.; El−Sharkawy, E.A.; Abdou, H.M.; Awual, M.R. Novel hierarchical composite adsorbent for selective lead(II) ions capturing from wastewater samples. Chem. Eng. J. 2018, 332, 377–386. [Google Scholar] [CrossRef]
Figure 1. FTIR spectra of synthesized composites (A); Zeta potential plots of SBA-15, CS2chitosan and CS2C@SBA−3 (B). The XPS survey spectra of CS2C@SBA−3 before reaction (C). High resolution XPS spectra S 2p of CS2C@SBA−3 (D).
Figure 1. FTIR spectra of synthesized composites (A); Zeta potential plots of SBA-15, CS2chitosan and CS2C@SBA−3 (B). The XPS survey spectra of CS2C@SBA−3 before reaction (C). High resolution XPS spectra S 2p of CS2C@SBA−3 (D).
Catalysts 12 01469 g001
Figure 2. SEM images of SBA-15 (A,B); CS2−chitosan (C,D); CS2C@SBA−3 (E,F).
Figure 2. SEM images of SBA-15 (A,B); CS2−chitosan (C,D); CS2C@SBA−3 (E,F).
Catalysts 12 01469 g002
Figure 3. Effects of pH on V(V) elimination. Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA dosage = 1.0 g/L, experimental temperature = 298 K, contact time = 24 h.
Figure 3. Effects of pH on V(V) elimination. Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA dosage = 1.0 g/L, experimental temperature = 298 K, contact time = 24 h.
Catalysts 12 01469 g003
Figure 4. Effects of the dosage of CS2C@SBA−3 on the removal of V(V). Conditions: V(V) initial concentration = 100 ppm, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Figure 4. Effects of the dosage of CS2C@SBA−3 on the removal of V(V). Conditions: V(V) initial concentration = 100 ppm, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Catalysts 12 01469 g004
Figure 5. Kinetics studies. (A) V(V) removal percentage versus time; (B) fitted results for linearized pseudo−first order; (C) pseudo−second order kinetic models; and (D) Elovich model. Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K.
Figure 5. Kinetics studies. (A) V(V) removal percentage versus time; (B) fitted results for linearized pseudo−first order; (C) pseudo−second order kinetic models; and (D) Elovich model. Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K.
Catalysts 12 01469 g005
Figure 6. Isotherm studies. (A) removal efficiency versus V(V) initial concentration, (B) adsorption capacity versus equilibrium concentration. Conditions: CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Figure 6. Isotherm studies. (A) removal efficiency versus V(V) initial concentration, (B) adsorption capacity versus equilibrium concentration. Conditions: CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Catalysts 12 01469 g006
Figure 7. Effects of co−existing ions (A), effects of ionic strength (B) and the elimination of V(V) in real water samples (C). Conditions (A,B): V(V) initial concentration = 100 ppm V(V), and 60 ppm PO43−, 600 ppm SO42−, 400 ppm NO3−, 400 ppm HCO3 and 60 ppm Cu2+, CS2C@SBA−3 dosage = 1.0 g/L, experimental temperature = 298 K, contact time = 24 h; Condition(C): V(V) initial concentration = 10 ppm, CS2C@SBA−3 dosage = 1.0 g/L, experimental temperature = 298 K, contact time = 24 h.
Figure 7. Effects of co−existing ions (A), effects of ionic strength (B) and the elimination of V(V) in real water samples (C). Conditions (A,B): V(V) initial concentration = 100 ppm V(V), and 60 ppm PO43−, 600 ppm SO42−, 400 ppm NO3−, 400 ppm HCO3 and 60 ppm Cu2+, CS2C@SBA−3 dosage = 1.0 g/L, experimental temperature = 298 K, contact time = 24 h; Condition(C): V(V) initial concentration = 10 ppm, CS2C@SBA−3 dosage = 1.0 g/L, experimental temperature = 298 K, contact time = 24 h.
Catalysts 12 01469 g007
Figure 8. SEM−EDS images of CS2C@SBA−3 before adsorption with V(V) (A) and EDS elemental mapping images (B,F), SEM−EDS images of CS2C@SBA−3 after adsorption with V(V) (C) and EDS elemental mapping images (D,E,G). Conditions: Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Figure 8. SEM−EDS images of CS2C@SBA−3 before adsorption with V(V) (A) and EDS elemental mapping images (B,F), SEM−EDS images of CS2C@SBA−3 after adsorption with V(V) (C) and EDS elemental mapping images (D,E,G). Conditions: Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Catalysts 12 01469 g008
Figure 9. FTIR spectra of CS2C@SBA−3 before and after adsorption (A). The XPS survey spectra of CS2C@SBA−3 before and after adsorption (B). V 2p (C), V 2p1/2 (D), and V 2p2/3 (E). The XPS high−resolution of S 2p (F).
Figure 9. FTIR spectra of CS2C@SBA−3 before and after adsorption (A). The XPS survey spectra of CS2C@SBA−3 before and after adsorption (B). V 2p (C), V 2p1/2 (D), and V 2p2/3 (E). The XPS high−resolution of S 2p (F).
Catalysts 12 01469 g009
Figure 10. The XPS high−resolution of N 1S of CS2C@SBA−3 before adsorption (A), and after adsorption (B). The XPS high−resolution of O 1S of CS2C@SBA−3 before adsorption (C), and after adsorption (D). Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Figure 10. The XPS high−resolution of N 1S of CS2C@SBA−3 before adsorption (A), and after adsorption (B). The XPS high−resolution of O 1S of CS2C@SBA−3 before adsorption (C), and after adsorption (D). Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Catalysts 12 01469 g010
Figure 11. Effects of recyclability of CS2C@SBA−3 on V(V) removal. Condition: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Figure 11. Effects of recyclability of CS2C@SBA−3 on V(V) removal. Condition: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Catalysts 12 01469 g011
Table 1. Elemental composition of synthesized materials.
Table 1. Elemental composition of synthesized materials.
AdsorbentElemental Composition a (wt%)Degree of CS2 Substitution (%)
CHNS
SBA-150.091.170.030.01NA
CS2Chitosan42.547.427.923.9220.96
CS2C@SBA−132.836.214.501.9719.19
CS2C@SBA−225.405.353.521.7321.52
CS2C@SBA−318.994.212.621.3622.72
a Determined through elemental analysis. NA: Not applicable.
Table 2. The parameters of kinetic and thermodynamics models for the adsorption of V(V) on CS2C@SBA−3.
Table 2. The parameters of kinetic and thermodynamics models for the adsorption of V(V) on CS2C@SBA−3.
ModelsParameters
Pseudo−first−order modelK1 (min−1)Qe (mg∙g−1)R2
0.01439.8130.9851
Pseudo−second−order modelK2 (mg∙g−1∙min−1)Qe (mg∙g−1)R2
0.00198.0450.9999
Elovich modelα (mg∙min∙g−1) β (mg∙g−1)R2
41.3419.0570.8788
Langmuir modelQm (mg∙g−1)KL (L∙mg−1)R2
221.1930.2600.9982
Freundlich modelnKf (mg1−n∙Ln∙g−1)R2
4.25265.6630.9218
Table 3. Elemental composition of CS2C@SBA−3 before adsorption and after adsorption.
Table 3. Elemental composition of CS2C@SBA−3 before adsorption and after adsorption.
AdsorbentElemental Composition a (wt %)
CSiONSV
CS2C@SBA−3 before adsorption
CS2C@SBA−3 after adsorption b
55.24.532.75.51.80.0
46.35.832.62.70.412.2
a determined by EDS. b Conditions: V(V) initial concentration = 100 ppm, CS2C@SBA−3 dosage = 1.0 g/L, pH value = 3.0, experimental temperature = 298 K, contact time = 24 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, Y.; Wang, J.; Li, M.; You, Z. Application of Dithiocarbamate Chitosan Modified SBA-15 for Catalytic Reductive Removal of Vanadium(V). Catalysts 2022, 12, 1469. https://doi.org/10.3390/catal12111469

AMA Style

Huang Y, Wang J, Li M, You Z. Application of Dithiocarbamate Chitosan Modified SBA-15 for Catalytic Reductive Removal of Vanadium(V). Catalysts. 2022; 12(11):1469. https://doi.org/10.3390/catal12111469

Chicago/Turabian Style

Huang, Yilin, Jia Wang, Mengwei Li, and Zhixiong You. 2022. "Application of Dithiocarbamate Chitosan Modified SBA-15 for Catalytic Reductive Removal of Vanadium(V)" Catalysts 12, no. 11: 1469. https://doi.org/10.3390/catal12111469

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

Article Metrics

Back to TopTop