Next Article in Journal
A Targeted Mass Spectrometric Approach to Evaluate the Anti-Inflammatory Activity of the Major Metabolites of Foeniculum vulgare Mill. Waste in Human Bronchial Epithelium
Previous Article in Journal
Mid-Level Data Fusion Techniques of LC-MS and HS-GC-MS for Distinguishing Green and Ripe Forsythiae Fructus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Remediation of Hg-Contaminated Groundwater via Adsorption on Supramolecular Polymers in Batch Process and Column Test

1
Department of Environment Engineering, Yellow River Conservancy Technical Institute, Kaifeng Engineering Research Center for Municipal Wastewater Treatment, Kaifeng 475004, China
2
Product Quality Inspection and Testing Center of Kaifeng, Kaifeng 475004, China
3
MOE Key Laboratory of Yellow River and Huai River Water Environmental and Pollution Control, School of Environment, Henan Normal University, Xinxiang 453007, China
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(7), 1406; https://doi.org/10.3390/molecules30071406
Submission received: 23 December 2024 / Revised: 17 March 2025 / Accepted: 18 March 2025 / Published: 21 March 2025

Abstract

Mercury contamination in groundwater seriously affects human health and ecosystem security. The remediation of Hg-contaminated groundwater remains a challenging task. The applicability of an as-synthesized supramolecular polymer (SP) for low-concentration mercury in a high-salinity groundwater matrix has been verified through a batch process and column test. The remediation of mercury-contaminated groundwater, particularly in complex high-salinity environments, represents a significant and enduring challenge in environmental science. The batch test study demonstrated that the SP can efficiently adsorb Hg from groundwater with superior selectivity and a high uptake capacity (up to 926.1 ± 165.3 mg g−1). Increasing the pH and dissolved organic matter (DOM) and reducing the ionic strength can facilitate Hg adsorption; the coexistence of heavy metal ions slightly weakens the removal. In terms of its performance as a permeable reactive barrier, the SP can intercept Hg in flowing groundwater with a capacity of up to 3187 mg g−1. A low influent mercury concentration, low pore velocity, and high SP dosage can effectively extend the breakthrough time in column tests. Additionally, the Yan model (R2 = 0.960−0.989) can accurately depict the whole dynamic interception process (150 PVs) of SPs in a fixed column, and the Adams–Bohart model (R2 = 0.916−0.964) describes the initial stage (≤35 PVs) well. Considering the functional group in the SP and the Hg species in groundwater, complexation, electrostatic attraction, ion exchange, and precipitation/co-precipitation are the plausible mechanisms for mercury removal based on the characterization results of scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), and Fourier transform infrared spectrometer (FT-IR). These impressive features render the SP a promising candidate for the remediation of trace Hg in saline groundwater using permeable reactive barrier (PRB) technology.

1. Introduction

Mercury (Hg) is one of the most toxic persistent heavy metals, and its presence poses a serious threat to the ecological environment and human health. The reduction of mercury in the environment and health has become a global management and technology focus [1,2]. Among the various valence states of mercury, Hg2+ is the most abundant. Incidents of mercury contamination in groundwater have been reported worldwide, with Hg concentrations in groundwater ranging from 66 μg L−1 to 2005 μg L−1 [3,4,5,6]. Furthermore, the established maximum contaminant level (MCL) established toward Hg in drinking water is 2 μg L−1 (USEPA, 2010), and the World Health Organization (WHO) guideline for Hg in drinking water is 1 μg L−1. Groundwater is an important source of drinking water and irrigation water for agriculture. The guideline value of Hg with respect to groundwater for drinking in China is 1 μg L−1 [7]. Therefore, it is ecologically and strategically imperative to explore cost-effective technologies to address this global dilemma.
Groundwater is rich in minerals, and the major species among them ( Na + , Ca 2 + , Mg 2 + , K + , HCO 3 , SO 4 2 - , Cl , and NO 3 ) always account for >99% of the solute content [8], leading to high ionic strength and high salinity. The presence of multiple ions affects the form of mercury [9], which is related to the fluidity and toxicity of mercury. In addition, the water quality and hydrological conditions of groundwater, such as the pH, mercury concentration, dissolved organic matter (DOM), and pore velocity, also have a significant influence on the presence and migration of mercury [10,11,12]. The difficulty of removing mercury from groundwater is compounded by the complex substrate [2,9]. Among the various methods of removing contaminants from flowing groundwater, permeable reactive barriers (PRBs) stand out for their simple operation, effective in situ interception and ease of scale-up from laboratory to industrial scale [13,14]. Reactive sorbents with high efficiency and low cost, filled in PRBs to trap the migrating contaminants in groundwater, have become the bottleneck in promoting the applications of PRBs. To date, persistent efforts have been made to fabricate reactive engineered materials for the in situ remediation of heavy metal contaminated groundwater, such as activated carbon, resins, zeolites, clay, zero-valent iron (ZVI), graphene oxide (GO), solid waste residue, chitosan, chelating ligands, metal organic frameworks (MOFs), titanate nanotubes, etc. [13,15,16,17,18,19,20,21,22,23,24,25]. Limited by the low sorption capacity of activated carbon, resins, zeolites, chitosan, and clay for mercury; the chemical instability of ZVI in groundwater; the potential decrease in permeability of GO and solid waste residue at a given flow rate; and the environmentally unfriendly and high cost of chelating ligands and MOFs, a reliable sorbent with high selectivity and capacity is highly desirable for the remediation of Hg-contaminated groundwater in PRBs.
As an environment-friendly material, polymeric material has become one of the most promising candidates, benefiting from the excellent controllable chemical functionality, dimensional stability, and extraordinary adsorption property [26,27]. The development of an innovative synthesis method and the screening of suitable raw materials for the production of highly efficient polymeric materials is, therefore, the crucial link. Trithiocyanuric acid (TTCA), the main precursor of the commercial adsorbent TMT15, which is limited in practice by aggregation, pH limitation, and difficult recovery, has a strong complexing ability of sulfur ligands with many soft Lewis acid metal ions (Pb2+, Hg2+, etc.) [27,28,29,30,31,32]. The organodisulfide polymer derived from TTCA has the potential to be an excellent heavy metal scavenger but suffers from the use of toxic reagents, complex synthesis steps, and limited yield, which prevent large-scale application [33,34,35,36]. Recent work has shown that melamine (MA)-based dendrimers also exhibit high scavenging capacity due to the complexation between –NH2 and heavy metals [37,38]. It is still essential to find a way to combine the advantages of TTCA and MA while avoiding their disadvantages. Few attempts have been made to enrich precious metals from aqueous substances by the sequential addition of TTCA and MA [39,40]. Furthermore, a chemically stable supramolecular polymer (SP), through non-covalent combination, has been synthesized through the high hydrogen bonding of N–H···S and N–H···N via the reorganization of MA and TTCA at the molecular level [41,42,43,44,45,46]. A magnetic supramolecular polymer fabricated via hydrogen bonding showed excellent capture ability for Pb and Cr in wastewater based on our previous study [47,48]. However, to the best of our knowledge, the investigation of the SP as a reactive material in PRBs for Hg2+ removal in groundwater has not yet been reported.
Motivated by the above research gaps, we systematically investigated the batch and column performance of the SP to assess its applicability in removing Hg2+ from simulated groundwater. The objectives of this study were (1) to explore the application of the SP to mitigate mercury concentrations through batch experiments by investigating the effects of the pH, dosage, dissolved organic matter (DOM), ionic strength, and coexisting ions on the adsorption performance of Hg2+ in groundwater; (2) to describe the breakthrough curves of the SP toward Hg2+ using typical models (Adams–Bohart, Thomas, Yoon–Nelson, and Yan) and to investigate the effects of operation conditions (including the influent Hg2+ concentration, pore velocity, and sorbent dosage) for the SP in PRBs via a fixed-bed column test; and (3) to reveal the plausible mechanism of Hg2+ adsorption on the SP based on the results of routine and column experiments coupled with component characterization.

2. Results and Discussion

2.1. Material Characterization

The SP obviously exhibited a distinct rod-like structure and smooth surface (Figure 1 and Figure S1), and the uniformly distributed presence of C, N, and S atoms on the surface was observed. In the FT-IR spectra (Figure 2a), the N–H stretching vibration that correspond to hydrogen bonding (N–H···S and N–H···N) are centered at 3420 cm−1. The signal at 880 cm−1 is attributed to the bending vibration of heterocycles in the polymer structure of the SP. The peaks around 1633 cm−1, 1142 cm−1, and 780 cm−1 correspond to the thione, the C=S stretching vibration, and the thiazine ring vibration, respectively [28,47]. The peak at 3420 cm⁻1 in the SP provides clear evidence of hydrogen bonding interactions.
The SP also has high crystallinity (Figure 2b), confirming the formation of a supramolecular polymer. Compared with the XRD patterns of MA and TTCA, there are three signals at 18.57°, 13.19°, and 12.25°, which can be attributed to in-planar packing, and the peak at 24.76° is indexed to the graphite-like structure of SP sheets [42,49,50]. The crystallinity indices (CrI) from XRD for TTCA, SP, and MA are 0.773, 0.579, and 0.797, respectively. As shown in Figure 2c, three signals centered at 288.1, 399.6, and 162.0 eV represent the coexistence of C, N, and S elements in the SP, corresponding to C1s, N1s, and S2p, respectively [51,52]. In the high−resolution XPS spectra of C1s (Figure 2d), there are three forms, including the thione (33.68%, 284.9 eV), the thiazine ring (29.76%, 287.6 eV), and the aromatic trithiol (38.36%, 288.6 eV) in the SP. For the N1s XPS spectra (Figure 2e), signals at 399.3 eV (69.99%), 400.0 eV (10.20%), and 400.5 eV (10.20%) represent the aromatic trithiol, the –NH–, and the thione, respectively [28,51,53,54]. The peaks situated at 162.1 and 162.3 eV fit the spectra of S2p3/2 (61.45%) and S2p1/2 (6.40%) in the aromatic trithiol and that of 163.2 eV indicates the S2p3/2 in the thione (Figure 2f) [34,55,56].

2.2. Interfacial Processes of Hg in Groundwater

To understand the interfacial processes of Hg adsorption in groundwater, the pseudo-first- and pseudo-second-order kinetic models were used [52,57,58,59]. The sorption kinetics was analyzed using both the pseudo-first- and pseudo-second-order kinetic models to provide a comprehensive understanding of the interfacial processes (Text S1). The SP exhibited a higher Hg removal efficiency (96.39%), superior adsorption capacity (38.63 ± 0.03 mg g−1), and shorter equilibration time (5 h) (Figure S2a). The uniform dispersion of N and S in the SP significantly facilitated the adsorption to Hg (Figure 1c,d). In addition, the fitting was conducted through the plots of ln (qeqt) vs. t, t/qt vs. t (Figure 3a,b). The R2 (0.9996) of pseudo-second-order equation for Hg2+ removal was better than that of pseudo-first model (0.9262). In addition, the equilibrium adsorption capacity qe (39.22 ± 0.31 mg g−1) based on the fitted results of the pseudo-second-order kinetics is approximately equal to the experimental result (Table S1). The pseudo-second-order model is based on the assumption that the rate-limiting step may be chemical adsorption or chemisorption, involving electron sharing or electron transfer between the sorbate and the sorbent. The better fit of the pseudo-second-order kinetic model suggests that the rate-limiting step in the adsorption process may be chemisorption through valency forces involving electron sharing between mercury and the SP on heterogeneous surfaces [52,60].
To further understand the effect of the concentration on adsorption, varying initial concentrations C0 (0.01–1.50 mg L−1) were studied. An increase in the adsorption capacity qe was accompanied by an increase in the C0, which can be attributed to the stronger ability of overcoming mass-transfer resistance resulting from the concentration gradient. The qt increases rapidly with an increasing time before 300 min, and it starts to stabilize after 300 min (Figure S2a). By contrast, the removal efficiency by the SP decreased significantly (Figure S2b). Langmuir isotherm and Freundlich isotherm models (Text S2) were used to depict the adsorption equilibrium [61,62]. The nonlinear fitting of qe vs. ce is shown in Figure 3c. The Langmuir isotherm (R2 = 0.9803) matched the adsorption data better than the Freundlich isotherm (R2 = 0.9780) (Table S2). The Freundlich model predicts nonideal sorption on heterogeneous or multilayer surfaces, but the Langmuir model can be applied to a constant monolayer-sorption capacity. This indicates that Hg was mainly adsorbed via monolayer adsorption on heterogeneous surfaces of the SP during the interfacial process [52,62]. Furthermore, the qm of the SP is up to 926.1 ± 165.3 mg g−1 due to the rich –SH and –NH2 in the supramolecular polymer (SP), which is higher than that of the most recently reported capacities (Table 1).
To better understand the temperature effect on mercury adsorption, thermodynamic studies were performed. It can be seen that a higher temperature is favorable for Hg removal (Figure S2b), indicating that the adsorption process is endothermic. Furthermore, the negative free energy changes (ΔG) and the decreasing trend with an increasing temperature indicate the spontaneity of the adsorption (Table S3). The positive enthalpy change (ΔH) shows that the adsorption process is endothermic, and the positive entropy change (ΔS) suggests the increasing randomness at the interface between Hg and the SP during the adsorption. This also indicates that the adsorption is likely to be chemisorption, which is consistent with the kinetic analysis.
For the selected simulated groundwater, Hg(OH)2, HgClOH (aq), HgCl2 (aq), and HgOHCO 3 (aq) were the predominant Hg species, with proportions of 45.9%, 40.4%, 7.1%, and 5.7%, respectively [2]. Ca 2 + , CO 3 2 , and SO 4 2 - were the major anions and cations, which may lead to the presence of CaCO3 precipitation. Precipitation/co-precipitation of Hg2+ and Ca 2 + could occur. The surface charge of SP is negative (Figure 3d), and the positive (Hg2+, HgCl+, and HgOH+) or uncharged charge (Hg(OH)2, HgClOH, and HgCl2) could be adsorbed on the SP via electrostatic attraction. Therefore, the SP could selectively adsorb mercury species before calcium species, resulting in an increase in both the zeta potential (from −59.6 mV to −33.5 mV) and hydraulic diameter (from 1.26 μm to 1.49 μm) (Figure 3d,e). The pH change throughout the process is almost negligible (Figure 3f).

2.3. Effects of pH, Ionic Strength, DOM, and Coexisting Ions

The efficiency of the mercury removal by the SP increased from 11.66% to 92.94% in a pH range from 2.0 to 6.0 (Figure 4a). The pH can both vary the surface charge of the SP and affect the mercury species in simulated groundwater [65,66]. At pH < 4.0, H+ occupies partial active sites on the SP. This can lead to a weakening of the negative charges, weakening the electrostatic attraction of the SP to mercury and causing competition (e.g., cation–π interaction and ligand exchange) between H+ and Hg2+. With an increasing pH, the decrease in H+, the increase in negative charge, and the enhancement of mercury precipitation/co-precipitation result in a significant increase in the Hg removal rate and a small increase in precipitate removal (from 1.75% to 5.03%). Higher ionic strength hindered the Hg2+ removal (Figure 4b). Reasons for this include competition from other cations (Ca2+ and Na+) for adsorption sites on the SP, shielding caused by a high salt content with a neutral electrical charge (e.g., NaCl, Na2SO4, CaCl2, and NaHCO3) between Hg and the SP, the strong inhibition of precipitation on interfaces, and the limitations imposed by C1 due to metal chlorides [2,65,67,68,69,70]. Regarding the role of the groundwater matrix (e.g., Ca2+, Na+), it can be seen that a certain amount of Ca (0.01%) and Na (0.43%) is adsorbed onto the SP in the groundwater without mercury (Figure 5a,b). However, in the Hg-contaminated groundwater, perhaps due to competition or subsequent ion exchange, the Na content decrease to 0.14% (2.45% for Hg), indicating that the presence of Na+ has a negative but negligible effect (Figure 5c). However, as the Hg content increases to 2.5%, Ca also shows an increase (from 0.01% to 0.57%), which may be due to precipitation or co-precipitation, indicating that the presence of Ca2+ can promote mercury removal. This shows that the SP has good applicability to real-world conditions.
As a typical macromolecular DOM, humic acid (HA) has a strong affinity for Hg due to the functional groups in HA (e.g., –OH, –NH2, –COOH, and –SH) [67,71]. Therefore, mercury in groundwater was scavenged by both the SP and HA through complexation. At low HA concentrations, HA in solution will very readily complex Hg in groundwater. This leads directly to a decrease in the removal of Hg by the SP. With an increasing HA concentration, some HA complexed with mercury in groundwater was adsorbed on the SP, which resulted in a significant enhancement of Hg removal from the aqueous solution to the SP (from 74.19% to 95.87%) (Figure 4c). Increasing the concentration of glucose (GLu) had little effect on interfacial adsorption. For both HA and GLu, Hg precipitation (1.19%−1.69%) showed robust resistance to changes in the concentration. In the presence of coexisting heavy metal ions (Cd, Cu, Zn, Pb, Cr, and Ni), the selectivity of the SP for mercury removal in simulated groundwater was investigated (Figure 4d). It can be seen that SP removal slightly decreased but remained relatively high (94.63%) compared to the six other metal species (1.17−11.11%). This suggests that despite the competition for active sites between seven heavy metals, the SP still has a higher affinity for Hg due to the strong soft acid–soft base interactions by the form of Hg···N and Hg···S bonds [47].

2.4. Regeneration and Reusability

To minimize economic costs and reduce secondary pollution, the regeneration and reusability of the SP is also crucial for practical applications. Here, 0.05 M EDTA-2Na was used to elute the adsorbed Hg on the SP in the regeneration experiment. It can be seen that there is only a slight decrease in the removal (from 96.34% to 91.99%) and the uptake capacity (from 38.20 mg g−1 to 36.48 mg g−1) for Hg after five cycles (Figure S3). Therefore, the resulting SP has easy regeneration and good reproducibility.

2.5. Mathematical Modeling of Breakthrough Curves

The breakthrough curves of the fixed column show that Ct/C0 generally increases with increasing pore volumes (PVs) (Figure 6). This may be due to the subsequent reduction of available sites on the SP as more groundwater flows through the column. After 150 PVs (the total groundwater influent volume is 1194 mL), the Ct/C0 for the SP reached 9.50−25.37%. Control tests (CK) showed that quartz sand and column materials had minimal impact on Hg interception (<5%). The fitted results of four mathematical models are shown in Table 2 and Text S3. For 150 PVs, the Yan model, with higher R2 values (0.960–0.989) and lower error (SS and ARE), can describe the process more accurately under various operating conditions. Based on the fitting of the Yan model, the maximum adsorption capacity qY can be up to 3187 mg g−1, the tb (95% breakthrough time) can reach 69.4 years, and the pτ (pore volume of 50% breakthrough) ranges from 3646 to 11,531 PVs, indicating that the SP has an ultra-high uptake for mercury in groundwater. In addition, the Adams–Bohart model (R2 = 0.916−0.964) can better characterize the dynamic process of mercury interception during the initial phase (≤35 PVs).
In general, the operating variables in the column tests, including the initial concentration C0, influent pore velocity νp, and adsorbent dosage m, significantly affected the Hg interception by the SP. As shown in Figure 6a–c and Table 2, a shorter breakthrough time was reached with a higher C0. When the C0 increased from 0.024 to 0.080 mg L−1, the tb decreased from 42.7 years to 11.8 years, and the qe increased from 50.0 to 144.0 mg g−1 (for 150 PVs). Mercury can be more easily transported from the aqueous solution to the SP surface due to the higher mass-transfer driving force caused by the high Hg concentration [64,72]. Similarly, the tb decreased from 69.4 years to 7.16 years with an increasing νp from 0.080 to 0.253 cm min−1 (qe dropped from 115.6 to 99.1mg g−1) (Figure 6d–f and Table 2). A high νp facilitates a reduction in the thickness of the mass-transfer film, which shortens the residence time for Hg from the external to the internal sites on the SP, resulting in low occupiable active sites on the SP [64,73]. With respect to the dosage m, a high m increased the tb but inhibited the qe (Figure 6g–i and Table 2). A high dosage provided more adsorption sites, leading to a relative surplus of available active sites on the SP [14,74].

2.6. Plausible Mechanism for Hg(II) Adsorption

To obtain insights into the mechanism of adsorption, the characterization methods SEM, XPS, and FT-IR were used. The highly consistent distribution and surface percentages of N, S atoms confirm the formation of the SP molecule (Figure 1, Figure 2c and Figure 5a). Compared to the SP, the presence of Hg (2.45%) and the consistent distribution of Hg, N, and S atoms demonstrate the complexation between Hg species (e.g., (Hg(OH)2, HgClOH, HgCl2, and HgOHCO 3 ) in groundwater and the N, S composition in the SP, attributable to the strong soft acid–soft base interactions by the form of Hg···N and Hg···S bonds (Figure 5c). Signals centered at 288.13 eV, 532.29 eV, 162.03 eV, 399.63 eV, 199.91 eV, 347.87 eV, 1071.97 eV, and 101.08 eV indicate the presence of C1s, O1s, S2p, N1s, Cl2p, Ca2p, Nals, and Hg4f (Figure 2c). The doublet 4f peaks for Hg(II) suggest that mercury was successfully uploaded onto the SP surfaces (Figure 7c). For N1s high-resolution XPS spectra, compared to the SP, the lower content (from 9.50% to 3.02%) for the thione, the higher shift (from 400.0 eV to 400.2 eV) for –NH–, and the higher content (from 69.59% to 75.28%) for the aromatic trithiol also suggest the formation of Hg···N coordination bonds in the SP-G-Hg (Figure 7a). For S2p high-resolution XPS spectra, compared to the SP, the lower content for the thione (from 9.50% to 3.02%), the higher shift for –NH– (from 400.0 eV to 400.2 eV), and the higher content for the aromatic trithiol (from 69.59% to 75.28%) also suggest the formation of Hg···N coordination bonds in the SP-G-Hg (Figure 7b). For S2p high-resolution XPS spectra, compared to the SP, the lower content (from 45.73% to 15.93%) and higher shift (from 162.8 eV to 163.3 eV) for the thione S2p3/2, the higher content (from 2.60% to 16.92%) and the higher shift (from 162.2 eV to 162.8 eV) for aromatic trithiol S2p1/2, and the higher content (from 51.67% to 67.15%) and the higher shift (from 161.8 eV to 161.9 eV) for the aromatic trithiol S2p3/2 also suggest the formation of Hg···S coordination bonds in the SP-G-Hg (Figure 7b).
In the FTIR spectra (Figure 2a), the two undermining peaks around 1633 cm−1 and 1140 cm−1 indicate the formation of coordination bonds (metal–nitrogen and metal–sulfur) in the SP-G-Hg. The red shifts of aromatic trithiol (from 1482, 1240 and 780 cm−1 to 1420, 1235 and 774 cm−1) suggest that the coordinate bonds between mercury and ligands could stabilize the vibration of C3N3S3 rings in the SP-G-Hg. In addition, the SP can adsorb a small number of cations (Na+ and Ca2+) in groundwater without mercury duo to the electrostatic attraction driven by the electronegativity of the SP (Figure 5b). But under conditions of Hg-contaminated groundwater, mercury can replace these ions via ion exchange (Figure 5c), leading to a decrease in Na from 0.43% to 0.14%. For the formation of precipitates/co-precipitates, the Ca content shows an increasing trend. The adsorption mechanism can be described using a schematic diagram (Figure S4).

3. Materials and Methods

3.1. Materials and Characterization

All reagents used were of analytical grade or better without purification. Detailed information on the chemicals and characterizations is presented in the Supplementary Materials (Text S4).

3.2. Synthesis of SP

The supramolecular polymer (SP) was prepared according to our previous work, with some modifications [47]. Firstly, 1.773 g of trithiocyanuric acid was dispersed in 40 mL of water and heated at 70 °C via sonication for 1 h to form a homogeneous solution. Secondly, 1.261 g of melamine was mixed thoroughly with 30 mL of water accompanied by sonication for 0.5 h at 70 °C. Thirdly, the above two solutions were mixed together, transferred into a Teflon-lined autoclave, and kept at 100 °C for a further 2 h. The resulting light-yellow solid was then collected via filtration, purified with deionized water, and freeze dried (–60 °C, 48 h), which is referred to as the SP. Figure S5 shows photographs of the raw materials and the as-prepared SP, together with the proposed synthesis route of the SP.

3.3. Batch Experiments

For batch sorption procedures of mercury in groundwater, 10 mg of the SP was dispersed into 30 mL of the desired Hg2+ simulated groundwater (containing 234.3 mg L−1 C1, 96.0 mg L−1  SO 4 2 - , 183.0 mg L−1  HCO 3 , 230.0 mg L−1 Na+, and 32.0 mg L−1 Ca2+) in sealed 50 mL PTFE vials. The initial pH of the simulated groundwater was 8.08 ± 0.2. HCl and NaOH were used to adjust the pH. The suspension was shaken under 200 rpm at 25 °C and then filtered using 0.22 μm PTFE. The concentration of Hg in the filtrate was determined using atomic fluorescence spectrometry (AFS), and that of the other metal ions was detected using inductively coupled plasma optical emission spectrometry (ICP-OES). For sorption kinetics, the initial Hg2+ concentration was 0.08 mg L−1, the sorbent dosage was 2.0 mg L−1, and the experiment was conducted under pH 6.0. In sorption isotherms, the Hg2+ concentrations were 0.01–1.50 mg L−1, and the other experimental conditions were the same as above. The effects of environmental factors, including the pH, DOM, ionic strength, and coexisting ions (Cd, Zn, Pb, Cr, Cu, and Ni), were also determined. All experiments were carried out in triplicate. Control tests without the SP (called CK) showed that the Hg loss during the process was <5%. Calculations of the removal efficiency R (%) and the adsorption capacity qe (mg g−1) are shown in Text S5.

3.4. Column Tests

Continuous fixed-bed column tests were performed to investigate the remediation performance of the SP in permeable reactive barriers (PRBs) for mercury removal in simulated groundwater. The column experiments were conducted in a glass column (006EZ-10-25-AF, Omnifit, Amersham, UK). Under the driving force of the pump, the simulated groundwater was continuously injected into the column and flowed through layers of quartz sands, SP sorbents, quartz sands, and glass wool in sequence from top to bottom (Figure S6). The effluent was periodically collected by the automated sampler for subsequent analysis. The analysis method for metal ion concentration was the same as the batch experiment. The control column (quartz sands/glass wool) test was conducted under identical conditions. Several operating conditions of the column test, such as the influent Hg2+ concentration C0, pore velocity νp, and sorbent dosage m, were studied to investigate the effects. The breakthrough time (tb) and the total Hg sorbed in the column (qt, mg g−1) were adopted and are presented in Text S6 [14].

3.5. Mathematical Models for Breakthrough Curves

Four mathematical models were used to simulate the breakthrough curves in order to describe the dynamic Hg concentration change of the effluent, including the Adams–Bohart [75], Thomas [76], Yan [77], and Yoon–Nelson models [78]. The models can be expressed as Equations (1)–(4), respectively.
C t C 0 = exp K AB C 0 V 0 Q n K AB N 0 H ν p
C t C 0 = 1 1 + exp 1000 K Th q Th m V 0 K Th C 0 n 1000 Q
C t C 0 = 1 1 1 + C 0 V 0 n 1000 q Y m
C t C 0 = 1 1 + exp K YN V 0 p τ K YN V 0 n Q
Detailed information on Ct/C0, KAB, C0, V0, H, Q, n, νp, KTh, qTh, m, qY, pτ, and KYN in the above models is given in Text S3.

4. Conclusions

Our study demonstrated that the as-synthesized SP with excellent applicability can effectively remove low-concentration mercury in high-salinity groundwater despite unfavorable conditions such as organic molecules, coexisting metal ions, and high ionic strength. The adsorption performance fitted well with the pseudo-second-order kinetic model (R2 = 0.9996) and the Langmuir isotherm model (R2 = 0.9803), signifying that the process belongs to chemisorption and monolayer sorption. The maximum adsorption capacity qm can reach 926.1 ± 165.3 mg g−1 due to the rich –SH and –NH2 in the SP. Furthermore, the SP can serve as an excellent PRB for intercepting mercury in groundwater, and the Yan model and Adams–Bohart model can better describe the breakthrough curves. Further mechanistic studies suggest that complexation, electrostatic attraction, ion exchange, and precipitation play a major role. These practical performances, including excellent selectivity, high uptake capacity, superior anti-interference ability, and longer breakthrough time, support that the SP will be a good candidate for the practical PRB remediation of mercury-contaminated groundwater.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30071406/s1: Text S1. The pseudo-first-order and pseudo-second-order models. Text S2. The Langmuir isotherm and Freundlich isotherm models. Text S3. The four mathematical models for breakthrough curves. Text S4. Materials and instrumentation. Text S5. The adsorption capacity and removal efficiency. Text S6. The time of 95% breakthrough. Figure S1. SEM images of SP. Figure S2. Mercury sorption kinetics in simulated groundwater (m = 2.0 mg L−1 adsorbent, C0 = 0.08 mg L−1; reaction time was 5 h) and adsorption removal efficiencies (R) at multiple initial Hg2+ concentrations after mercury uptake from simulated groundwater (m = 2.0 mg L−1, C0 = 0.01–1.50 mg L−1; reaction time was 5 h). Fitting of the isotherms (T = 308 K) and (d) (T = 318 K) in the simulated groundwater for mercury adsorption onto the SP (m = 2.0 mg L−1; C0 = 0.01–1.50 mg L−1). Figure S3. Recycling performance of mercury onto SP (m = 2.0 mg L−1; T = 298 K; C0 = 0.08 mg L−1). Figure S4. Underlying adsorption mechanisms of SP for Hg species in simulated groundwater. Figure S5. The proposed route of synthesis and the chemical formula of the SP; the photographs of raw materials and the as-prepared SP. Figure S6. Scheme of the fixed-bed column test system. Table S1. Pseudo-first-order and pseudo-second-order kinetic models used for simulating Hg sorption kinetic data and the resulting fitting parameters. Table S2. Fitting parameters of Langmuir and Freundlich isotherm models for Hg sorption. Table S3. Thermodynamic parameters for mercury adsorption onto SP at different temperatures (m = 2.0 mg L−1; C0 = 0.08 mg L−1). Table S4. Fitting parameters of Thomas and Yoon–Nelson models for Hg adsorption by SP in column test.

Author Contributions

Conceptualization, Z.W. and W.L.; investigation, X.S., Y.Y. and J.J.; data curation and funding acquisition, Z.W. and J.S.; project administration and supervision, Z.W. and Q.Z.; writing—original draft, W.L. and Q.Z.; writing—review and editing, Z.W. and J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Henan Provincial Science and Technology Research and Development Joint Fund (232103810037), the Science and Technology Research Project of Kaifeng (2201002), the Science and Technology Development Plan of Henan Province (242102310336), the Major Science and Technology Projects of Henan Province (221100320200), and the College Students’ Nursery Camp Innovation Project (2023CX035).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lourenço, M.A.O.; Figueira, P.; Pereira, E.; Gomes, J.R.B.; Lopes, C.B.; Ferreira, P. Simple, mono and bifunctional periodic mesoporous organosilicas for removal of priority hazardous substances from water: The case of mercury(II). Chem. Eng. J. 2017, 322, 263–274. [Google Scholar] [CrossRef]
  2. Wang, Z.; Jia, J.; Liu, W.; Huang, S.; Chen, X.; Zhang, N.; Huang, Y. Mercury speciation transformation mediated by thiolated biochar in high salinity groundwater: Interfacial processes, influencing factors, and mechanisms. Chem. Eng. J. 2024, 484, 149443. [Google Scholar] [CrossRef]
  3. Wang, M.; Li, Y.; Zhao, D.; Zhuang, L.; Yang, G.; Gong, Y. Immobilization of mercury by iron sulfide nanoparticles alters mercury speciation and microbial methylation in contaminated groundwater. Chem. Eng. J. 2020, 381, 122664. [Google Scholar] [CrossRef]
  4. Bollen, A.; Wenke, A.; Biester, H. Mercury speciation analyses in HgCl2-contaminated soils and groundwater-Implications for risk assessment and remediation strategies. Water Res. 2008, 42, 91–100. [Google Scholar] [CrossRef]
  5. Chaudhuri, S.; Sigmund, G.; Bone, S.E.; Kumar, N.; Hofmann, T. Mercury removal from contaminated water by wood-based biochar depends on natural organic matter and ionic composition. Environ. Sci. Technol. 2022, 56, 11354–11362. [Google Scholar] [CrossRef]
  6. Aleku, D.L.; Lazareva, O.; Pichler, T. Mercury in groundwater—Source, transport and remediation. Appl. Geochem. 2024, 170, 106060. [Google Scholar] [CrossRef]
  7. GB/T 14848-2017; Standard for Groundwater Quality. General Administration of Quality Supervision, Inspection and Quarantine of the People’s Republic of China: Beijing, China, 2017. Available online: https://www.chinesestandard.net/PDF.aspx/GBT14848-2017 (accessed on 22 December 2024).
  8. Edmunds, W.M.; Smedley, P.L. Groundwater geochemistry and health: An overview. Geol. Soc. Lond. Spec. Publ. 1996, 113, 91–105. [Google Scholar] [CrossRef]
  9. Wang, M.; Han, Q.; Shu, Y.; Wang, K.; Wang, L.; Liu, B.; Zucker, I.; Wang, Z. Matrix effects on the performance and mechanism of Hg removal from groundwater by MoS2 nanosheet. Environ. Sci. Adv. 2022, 1, 59–69. [Google Scholar] [CrossRef]
  10. Chen, M.; Price, R.M.; Yamashita, Y.; Jaffé, R. Comparative study of dissolved organic matter from groundwater and surface water in the Florida coastal Everglades using multi-dimensional spectrofluorometry combined with multivariate statistics. Appl. Geochem. 2010, 25, 872–880. [Google Scholar] [CrossRef]
  11. Richard, J.-H.; Bischoff, C.; Ahrens, C.G.M.; Biester, H. Mercury (II) reduction and co-precipitation of metallic mercury on hydrous ferric oxide in contaminated groundwater. Sci. Total Environ. 2016, 539, 36–44. [Google Scholar] [CrossRef]
  12. Wang, M.; Zhang, M.; Han, Q.; Shu, Y.; Liu, X.; Chen, B.; Chen, Y.; Liu, B.; Wang, Z. Effect of natural organic matter (NOM) on the removal efficiency of Hg(ii) by MoS2: Dependence on the Hg/MoS2 ratio and NOM properties. Environ. Sci. Nano 2024, 11, 1129–1141. [Google Scholar] [CrossRef]
  13. Weisener, C.G.; Sale, K.S.; Smyth, D.J.A.; Blowes, D.W. Field column study using zerovalent iron for mercury removal from contaminated groundwater. Environ. Sci. Technol. 2005, 391, 6306–6312. [Google Scholar] [CrossRef]
  14. Zhang, Y.; Jin, F.; Shen, Z.; Wang, F.; Lynch, R.; Al-Tabba, A. Adsorption of methyl tertbutyl ether (MTBE) onto ZSM-5 zeolite: Fixed-bed column tests, breakthrough curve modelling and regeneration. Chemosphere 2019, 220, 422–431. [Google Scholar] [CrossRef] [PubMed]
  15. Vilensky, M.Y.; Berkowitz, B.; Warshawsky, A. In situ remediation of groundwater contaminated by heavy- and transition-metal ions by selective ion-exchange methods. Environ. Sci. Technol. 2002, 36, 1851–1855. [Google Scholar] [CrossRef]
  16. Zhang, Y.; Wang, F.; Cao, B.; Yin, H.; Al-Tabbaa, A. Simultaneous removal of Pb and MTBE by mixed zeolites in fixed-bed column tests. J. Environ. Sci. 2022, 34, 41–49. [Google Scholar] [CrossRef]
  17. Bleiman, N.; Mishael, Y.G. Selenium removal from drinking water by adsorption to chitosan–clay composites and oxides: Batch and columns tests. J. Hazard. Mater. 2010, 183, 590–595. [Google Scholar] [CrossRef]
  18. Biterna, M.; Antonoglou, L.; Lazou, E.; Voutsa, D. Arsenite removal from waters by zero valent iron: Batch and column tests. Chemosphere 2010, 78, 7–12. [Google Scholar] [CrossRef]
  19. Abbasi, M.; Safari, E.; Baghdadi, M.; Janmohammadi, M. Enhanced adsorption of heavy metals in groundwater using sand columns enriched with graphene oxide: Lab-scale experiments and process modeling. J. Water Process Eng. 2021, 40, 101961. [Google Scholar] [CrossRef]
  20. Blue, L.Y.; Van Aelstyn, M.A.; Matlock, M.; Atwood, D.A. Low-level mercury removal from groundwater using a synthetic chelating ligand. Water Res. 2008, 42, 2025–2028. [Google Scholar] [CrossRef]
  21. Rahmi; Fathurrahmi; Lelifajri; PurnamaWati, F. Preparation of magnetic chitosan using local iron sand for mercury removal. Heliyon 2019, 5, e01731. [Google Scholar] [CrossRef]
  22. Yong, S.K.; Bolan, N.S.; Lombi, E.; Skinner, W.; Guibal, E. Sulfur-containing chitin and chitosan derivatives as trace metal adsorbents: A review. Crit. Rev. Environ. Sci. Technol. 2013, 43, 1741–1794. [Google Scholar] [CrossRef]
  23. Yoon, S.Y.; Jang, S.B.; Wong, K.T.; Kim, H.; Kim, M.J.; Choong, C.E.; Yang, J.-K.; Chang, Y.-Y.; Oh, S.-E.; Yoon, Y.; et al. Sulfur-anchored palm shell waste-based activated carbon for ultrahigh sorption of Hg(II) for in-situ groundwater treatment. J. Hazard. Mater. 2021, 417, 125995. [Google Scholar] [CrossRef] [PubMed]
  24. Biswas, A.; Chandra, B.P.; Prathibha, C. Highly efficient and simultaneous remediation of heavy metal ions (Pb(II), Hg(II), As(V), As(III) and Cr(VI)) from water using Ce intercalated and ceria decorated titanate nanotubes. Appl. Surf. Sci. 2023, 612, 155841. [Google Scholar] [CrossRef]
  25. Wang, Z.; Xia, Y.; Wei, Y.; Tang, H.; Zhao, J.; Zhang, D.; Yu, L.; Tan, M.; Liu, X.; Shi, J.; et al. Deep and consecutive removal of Hg2+ from water using a novel dopamine-bridged metal organic framework-carbon sponge composite. Sep. Purif. Technol. 2024, 346, 127512. [Google Scholar] [CrossRef]
  26. Venkateswarlu, S.; Yoon, M. Core-shell ferromagnetic nanorod based on amine polymer composite (Fe3O4@DAPF) for fast removal of Pb(II) from aqueous solutions. ACS Appl. Mater. Interfaces 2015, 7, 25362–25372. [Google Scholar] [CrossRef]
  27. Huang, X.; Yang, J.; Wang, J.; Bi, J.; Xie, C.; Hao, H. Design and synthesis of core-shell Fe3O4@PTMT composite magnetic microspheres for adsorption of heavy metals from high salinity wastewater. Chemosphere 2018, 206, 513–521. [Google Scholar] [CrossRef]
  28. Fu, W.; Wang, X.; Huang, Z. Remarkable reusability of magnetic Fe3O4-encapsulated C3N3S3 polymer/reduced graphene oxide composite: A highly effective adsorbent for Pb and Hg ions. Sci. Total Environ. 2019, 659, 895–904. [Google Scholar] [CrossRef]
  29. Xue, J.; Qu, Y.; Chen, Y.; Zhang, C.; Bu, X. Effective sulfide flotation of cerussite by using trithiocyanuric acid as a novel sulfurizing reagent. Miner. Eng. 2023, 198, 108087. [Google Scholar] [CrossRef]
  30. Wang, W.; Qi, L.; Han, S.; Yuan, H. Synthesis, characterization of chitosan-trithiocyanuric and its removal mechanism of Cr(VI) and Pb(II) from wastewater. Chem. Eng. Sci. 2023, 281, 119144. [Google Scholar] [CrossRef]
  31. Huang, S.-A.; Teng, H.-J.; Su, Y.-T.; Liu, X.-M.; Li, B. Trithiocyanurate-functionalized hydrochar for effetively removing methylene blue and Pb (II) cationic pollutants. Environ. Pollut. 2023, 337, 122585. [Google Scholar] [CrossRef]
  32. Gao, X.; Li, Z.; Wen, H.; Zhao, J.; Zhou, M.; Yang, S.; Liu, J. Instant oxidase- and peroxidase-mimic activities of in-situ reductive coordinated Cu(I)-polytrithiocyanuric acid for H2S colorimetric detection and antibacterial. J. Hazard. Mater. 2025, 483, 136722. [Google Scholar] [CrossRef] [PubMed]
  33. Ko, D.; Lee, J.S.; Patel, H.A.; Jakobsen, M.H.; Hwang, Y.; Yavuz, C.T.; Hansen, H.C.B.; Andersen, H.R. Selective removal of heavy metal ions by disulfide linked polymer networks. J. Hazard. Mater. 2017, 332, 140–148. [Google Scholar] [CrossRef] [PubMed]
  34. Ko, D.; Mines, P.D.; Jakobsen, M.H.; Yavuz, C.T.; Hansen, H.C.B.; Andersen, H.R. Disulfide polymer grafted porous carbon composites for heavy metal removal from stormwater runoff. Chem. Eng. J. 2018, 348, 685–692. [Google Scholar] [CrossRef]
  35. Lin, G.; Wang, S.; Zhang, L.; Hu, T.; Cheng, S.; Fu, L.; Xiong, C. Enhanced and selective adsorption of Hg2+ to a trace level using trithiocyanuric acid-functionalized corn bract. Environ. Pollut. 2019, 244, 938–946. [Google Scholar] [CrossRef]
  36. Li, Q.; Ruan, B.; Yu, Y.; Ye, L.; Dai, A.; You, S.; Zhao, B.; Ren, L. Green and Mild Fabrication of Magnetic Poly(trithiocyanuric acid) Polymers for Rapid and Selective Separation of Mercury(II) Ions in Aqueous Samples. Polymers 2024, 16, 3067. [Google Scholar] [CrossRef]
  37. Sharahi, F.J.; Shahbazi, A. Melamine-based dendrimer amine-modified magnetic nanoparticles as an efficient Pb(II) adsorbent for wastewater treatment: Adsorption optimization by response surface methodology. Chemosphere 2017, 189, 291–300. [Google Scholar] [CrossRef]
  38. Al Hamouz, O.C.S.; Adelabu, I.O.; Saleh, T.A. Novel cross-linked melamine based polyamine/CNT composites for lead ions removal. J. Environ. Manag. 2017, 192, 163–170. [Google Scholar] [CrossRef]
  39. Nagai, D.; Kuribayashi, T.; Tanaka, H.; Morinaga, H.; Uehara, H.; Yamanobea, T. A facile, selective, high recovery system for precious metals based on complexation between melamine and cyanuric acid. RSC Adv. 2015, 5, 30133–30139. [Google Scholar] [CrossRef]
  40. Nagai, D.; Nagashima, A.; Mori, M. A facile and high-recovery system for palladium(II) ion based on complexation between trithiocyanuric acid and melamine. Chem. Lett. 2016, 45, 1165–1167. [Google Scholar] [CrossRef]
  41. Ranganathan, A.; Pedireddi, V.R.; Rao, C.N.R. Hydrothermal synthesis of organic channel structures: 1:1 hydrogen-bonded adducts of melamine with cyanuric and trithiocyanuric acids. J. Am. Chem. Soc. 1999, 121, 1752–1753. [Google Scholar] [CrossRef]
  42. Fan, Q.; Liu, J.; Yu, Y.; Zuo, S.; Li, B. A simple fabrication for sulfur doped graphitic carbon nitride porous rods with excellent photocatalytic activity degrading RhB dye. Appl. Surf. Sci. 2017, 391, 360–368. [Google Scholar] [CrossRef]
  43. Whitesides, G.M.; Simanek, E.E.; Mathias, J.P.; Seto, C.T.; Chin, D.; Mammen, M.; Gordon, D.M. Noncovalent synthesis:Using physical-organic chemistry to make aggregates. Acc. Chem. Res. 1995, 28, 37–44. [Google Scholar] [CrossRef]
  44. Liu, Y.; Ni, S.; Wang, W.; Rong, M.; Cai, H.; Xing, H.; Yang, L. Functionalized hydrogen-bonded organic superstructures via molecular self-assembly for enhanced uranium extraction. J. Hazard. Mater. 2024, 464, 133002. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, A.; Guo, Y.; Xie, H.; Zhang, Y.; Fu, Y.; Ye, C.; Du, Y.; Zhu, M. Green and Controllable synthesis of kelp-like carbon nitride nanosheets via an ultrasound-mediated self-assembly strategy. J. Colloid Interface Sci. 2022, 628, 397–408. [Google Scholar] [CrossRef]
  46. Li, H.; Zuo, P.; Qu, S.; Qin, F.; Li, N.; Shen, W. Hybrid hierarchically porous carbon micron tubes for trace cadmium and lead ions electrochemical detection. Appl. Surf. Sci. 2023, 615, 156426. [Google Scholar] [CrossRef]
  47. Wang, Z.; Zhang, J.; Wu, Q.; Han, X.; Zhang, M.; Liu, W.; Yao, X.; Feng, J.; Dong, S.; Sun, J. Magnetic supramolecular polymer: Ultrahigh and highly selective Pb(II) capture from aqueous solution and battery wastewater. Chemosphere 2020, 248, 126042. [Google Scholar] [CrossRef]
  48. Huang, Y.; Huang, Y.; Fang, L.; Zhao, B.; Zhang, Y.; Zhu, Y.; Wang, Z.; Wang, Q.; Li, F. Interfacial chemistry of mercury on thiol-modified biochar and its implication for adsorbent engineering. Chem. Eng. J. 2023, 454, 140310. [Google Scholar] [CrossRef]
  49. Shalom, M.; Inal, S.; Fettkenhauer, C.; Neher, D.; Antonietti, M. Improving carbon nitride photocatalysis by supramolecular preorganization of monomers. J. Am. Chem. Soc. 2013, 135, 7118–7121. [Google Scholar] [CrossRef]
  50. Jun, Y.-S.; Lee, E.Z.; Wang, X.; Hong, W.H.; Stucky, G.D.; Thomas, A. From melamine-cyanuric acid supramolecular aggregates to carbon nitride hollow spheres. Adv. Funct. Mater. 2013, 23, 3661–3667. [Google Scholar] [CrossRef]
  51. Fu, W.; Huang, Z. One-pot synthesis of a two-dimensional porous Fe3O4/poly(C3N3S3) network nanocomposite for the selective removal of Pb(II) and Hg(II) from synthetic wastewater. ACS Sustain. Chem. Eng. 2018, 6, 14785–14794. [Google Scholar] [CrossRef]
  52. Huang, Y.; Xia, S.; Lyu, J.; Tang, J. Highly efficient removal of aqueous Hg2+ and CH3Hg+ by selective modification of biochar with 3-mercaptopropyltrimethoxysilane. Chem. Eng. J. 2019, 360, 1646–1655. [Google Scholar] [CrossRef]
  53. Dinda, D.; Saha, S.K. Sulfuric acid doped poly diaminopyridine/graphene composite to remove high concentration of toxic Cr(VI). J. Hazard. Mater. 2015, 291, 93–101. [Google Scholar] [CrossRef] [PubMed]
  54. Kucharski, M.; Chmiel-Szukiewicz, E. Reactions of trithiocyanuric acid with oxiranes. I. Synthesis of Polyetherols. J. Appl. Polym. Sci. 2000, 76, 439–445. [Google Scholar] [CrossRef]
  55. Kim, H.; Lee, J.; Ahn, H.; Kim, O.; Park, M.J. Synthesis of three-dimensionally interconnected sulfur-rich polymers for cathode materials of high-rate lithium-sulfur batteries. Nat. Commun. 2015, 6, 7278. [Google Scholar] [CrossRef]
  56. Wang, Q.; Dong, S.; Zhang, D.; Yu, C.; Lu, J.; Wang, D.; Sun, J. Magnetically recyclable visible-light-responsive MoS2@Fe3O4 photocatalysts targeting efficient wastewater treatment. J. Mater. Sci. 2017, 53, 1135–1147. [Google Scholar] [CrossRef]
  57. Lagergren, S. About the theory of so-called adsorption of soluble substances. K. Sven. Vetenskapsakademiens 1898, 24, 1–39. [Google Scholar]
  58. Ho, Y.S.; McKay, G. The kinetics of sorption of divalent metal ions onto sphagnum moss peat. Water Res. 2000, 34, 735–742. [Google Scholar] [CrossRef]
  59. Ghodbane, I.; Hamdaoui, O. Removal of mercury(II) from aqueous media using eucalyptus bark: Kinetic and equilibrium studies. J. Hazard. Mater. 2008, 160, 301–309. [Google Scholar] [CrossRef]
  60. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  61. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef]
  62. Ho, Y.S.; Porter, J.F.; McKay, G. Equilibrium isotherm studies for the sorption of divalent metal ions onto peat: Copper, Nickel and Lead single component systems. Water Air Soil Pollut. 2002, 141, 1–33. [Google Scholar] [CrossRef]
  63. Wang, Z.; Zhang, J.; Wen, T.; Liu, X.; Wang, Y.; Yang, H.; Sun, J.; Feng, J.; Dong, S.; Sun, J. Highly effective remediation of Pb(II) and Hg(II) contaminated wastewater and soil by flower-like magnetic MoS2 nanohybrid. Sci. Total Environ. 2020, 699, 134341. [Google Scholar] [CrossRef] [PubMed]
  64. Huang, Y.; Wang, M.; Gong, Y.; Zeng, E.Y. Efficient removal of mercury from simulated groundwater using thiol-modified graphene oxide/Fe–Mn composite in fixed-bed columns: Experimental performance and mathematical modeling. Sci. Total Environ. 2020, 714, 136636. [Google Scholar] [CrossRef] [PubMed]
  65. Sánchez-Polo, M.; Rivera-Utrilla, J. Adsorbent-adsorbate interactions in the adsorption of Cd(II) and Hg(II) on ozonized activated carbons. Environ. Sci. Technol. 2002, 36, 3850–3854. [Google Scholar] [CrossRef]
  66. Chandra, V.; Kim, K.S. Highly selective adsorption of Hg2+ by a polypyrrole-reduced graphene oxide composite. Chem. Commun. 2011, 47, 3942–3944. [Google Scholar] [CrossRef]
  67. Huang, Y.; Gong, Y.; Tang, J.; Xia, S. Effective removal of inorganic mercury and methylmercury from aqueous solution using novel thiol-functionalized graphene oxide/Fe-Mn composite. J. Hazard. Mater. 2019, 366, 130–139. [Google Scholar] [CrossRef]
  68. El-Bayaa, A.A.; Badawy, N.A.; AlKhalik, E.A. Effect of ionic strength on the adsorption of copper and chromium ions by vermiculite pure clay mineral. J. Hazard. Mater. 2009, 170, 1204–1209. [Google Scholar] [CrossRef]
  69. Wang, J.; Deng, B.; Chen, H.; Wang, X.; Zheng, J. Removal of aqueous Hg(II) by polyaniline: Sorption characteristics and mechanisms. Environ. Sci. Technol. 2009, 43, 5223–5228. [Google Scholar] [CrossRef]
  70. Cui, H.; Qian, Y.; Li, Q.; Zhang, Q.; Zhai, J. Adsorption of aqueous Hg(II) by a polyaniline/attapulgite composite. Chem. Eng. J. 2012, 211–212, 216–223. [Google Scholar] [CrossRef]
  71. Nascimento, F.H.D.; Masini, J.C. Influence of humic acid on adsorption of Hg(II) by vermiculite. J. Environ. Manag. 2014, 143, 1–7. [Google Scholar] [CrossRef]
  72. Goel, J.; Kadirvelu, K.; Rajagopal, C.; Garg, V.K. Removal of lead(II) by adsorption using treated granular activated carbon: Batch and column studies. J. Hazard. Mater. 2005, 125, 211–220. [Google Scholar] [CrossRef] [PubMed]
  73. Hussein, F.B.; Mayer, B.K. Fixed-bed column study of phosphate adsorption using immobilized phosphate-binding protein. Chemosphere 2022, 295, 133908. [Google Scholar] [CrossRef] [PubMed]
  74. Tang, N.; Niu, C.-G.; Li, X.-T.; Liang, C.; Guo, H.; Lin, L.-S.; Zheng, C.-W.; Zeng, G.-M. Efficient removal of Cd2+ and Pb2+ from aqueous solution with amino- and thiol-functionalized activated carbon: Isotherm and kinetics modeling. Sci. Total Environ. 2018, 635, 1331–1344. [Google Scholar] [CrossRef] [PubMed]
  75. Bohart, G.S.; Adams, E.Q. Some aspects of the behavior of charcoal with respect to chlorine. J. Frankl. Inst. Eng. Appl. Math. 1920, 189, 669. [Google Scholar] [CrossRef]
  76. Thomas, H.C. Heterogeneous ion exchange in a flowing system. J. Am. Chem. Soc. 1944, 66, 1664–1666. [Google Scholar] [CrossRef]
  77. Yoon, Y.H.; Nelson, J.H. Application of gas adsorption kinetics. I. A theoretical model for respirator cartridge service life. Am. Ind. Hyg. Assoc. J. 1984, 45, 509–516. [Google Scholar] [CrossRef]
  78. Yan, G.; Viraraghavan, T.; Chen, M. A new model for heavy metal removal in a biosorption column. Adsorpt. Sci. Technol. 2001, 19, 25–43. [Google Scholar] [CrossRef]
Figure 1. SEM images of SP (a); SEM mapping of elements C (b), N (c), and S (d) for SP.
Figure 1. SEM images of SP (a); SEM mapping of elements C (b), N (c), and S (d) for SP.
Molecules 30 01406 g001
Figure 2. FT−IR (a), XRD (b), and XPS (c) patterns of the as-prepared samples. SP-G (SP after adsorption in the simulated groundwater without Hg) and SP-G-Hg (SP after adsorption in the simulated groundwater with Hg); high-resolution XPS spectra of (d) C1s, (e) N1s, and (f) S2p for the SP.
Figure 2. FT−IR (a), XRD (b), and XPS (c) patterns of the as-prepared samples. SP-G (SP after adsorption in the simulated groundwater without Hg) and SP-G-Hg (SP after adsorption in the simulated groundwater with Hg); high-resolution XPS spectra of (d) C1s, (e) N1s, and (f) S2p for the SP.
Molecules 30 01406 g002
Figure 3. Fitting of the pseudo-second-order kinetic model (a), the pseudo-first-order model (b), and isotherms (c) in the simulated groundwater for mercury adsorption onto the SP (m = 2.0 mg L−1; C0 = 0.01–1.50 mg L−1; T = 298 K); changes in the (d) zeta potential, (e) hydrodynamic diameter, and (f) pH (m = 2.0 mg L−1; C0 = 0.08 mg L−1; T = 298 K).
Figure 3. Fitting of the pseudo-second-order kinetic model (a), the pseudo-first-order model (b), and isotherms (c) in the simulated groundwater for mercury adsorption onto the SP (m = 2.0 mg L−1; C0 = 0.01–1.50 mg L−1; T = 298 K); changes in the (d) zeta potential, (e) hydrodynamic diameter, and (f) pH (m = 2.0 mg L−1; C0 = 0.08 mg L−1; T = 298 K).
Molecules 30 01406 g003
Figure 4. Effects of the (a) pH, (b) ionic strength, (c) DOM, and (d) coexisting heavy metal ions for Hg removal performance from a mixed solution of heavy metal ions. Experimental conditions for (d): 2.0 mg L−1 for the SP, C0 = 4.0 × 10−4 mmol L−1 for Hg and each coexisting heavy metal, and a mass concentration of 0.08 mg L−1 for Hg2+.
Figure 4. Effects of the (a) pH, (b) ionic strength, (c) DOM, and (d) coexisting heavy metal ions for Hg removal performance from a mixed solution of heavy metal ions. Experimental conditions for (d): 2.0 mg L−1 for the SP, C0 = 4.0 × 10−4 mmol L−1 for Hg and each coexisting heavy metal, and a mass concentration of 0.08 mg L−1 for Hg2+.
Molecules 30 01406 g004
Figure 5. SEM-mapping images of individual elements C, N, and S in the SP (a), C, N, S, Ca, Na, O, and Cl in the SP-G (b), and C, N, S, Ca, Na, O, Cl, and Hg in the SP-G-Hg (c). The numbers in the mapping represent atomic percentages.
Figure 5. SEM-mapping images of individual elements C, N, and S in the SP (a), C, N, S, Ca, Na, O, and Cl in the SP-G (b), and C, N, S, Ca, Na, O, Cl, and Hg in the SP-G-Hg (c). The numbers in the mapping represent atomic percentages.
Molecules 30 01406 g005
Figure 6. Effects of initial Hg concentration C0, influent pore velocity νp, and adsorbent dosage m in groundwater on breakthrough curve of Hg adsorption on SP in fixed-bed column (25 ± 1 °C). Experimental conditions for (ac): C0 = 0.024–0.080 mg L−1, νp = 0.139 cm min−1, and dosage m = 0.5 mg. Experimental conditions for (df): νp = 0.080–0.253 cm min−1, C0 = 0.056 mg L−1, and m = 0.5 mg. Experimental conditions for (gi): m = 0.5–1.5 mg, C0 = 0. 080 mg L−1, and νp = 0.139 cm min−1.
Figure 6. Effects of initial Hg concentration C0, influent pore velocity νp, and adsorbent dosage m in groundwater on breakthrough curve of Hg adsorption on SP in fixed-bed column (25 ± 1 °C). Experimental conditions for (ac): C0 = 0.024–0.080 mg L−1, νp = 0.139 cm min−1, and dosage m = 0.5 mg. Experimental conditions for (df): νp = 0.080–0.253 cm min−1, C0 = 0.056 mg L−1, and m = 0.5 mg. Experimental conditions for (gi): m = 0.5–1.5 mg, C0 = 0. 080 mg L−1, and νp = 0.139 cm min−1.
Molecules 30 01406 g006
Figure 7. High-resolution XPS spectra of N1s (a), S2p (b), and Hg4f (c) for SP-G, SP-P-Hg (SP after adsorption in DI water with mercury), and SP-G-Hg.
Figure 7. High-resolution XPS spectra of N1s (a), S2p (b), and Hg4f (c) for SP-G, SP-P-Hg (SP after adsorption in DI water with mercury), and SP-G-Hg.
Molecules 30 01406 g007
Table 1. Comparison of mercury uptake using different adsorbents.
Table 1. Comparison of mercury uptake using different adsorbents.
Adsorbentsqmax (mg g−1)Ref.
rGO-p(C3N3S3)/Fe3O4400.0[28]
MoS2/Fe3O4425.5[63]
SGO/Fe–Mn112.03[64]
PSAC-S136[23]
CCTNT199.80[24]
BCS4.30[2]
MPTAPs)211[36]
Zr-MSA/DCS312.4[25]
SP926 ± 165Present work
Table 2. Fitting parameters of Adams–Bohart and Yan models for Hg adsorption by SP in column test.
Table 2. Fitting parameters of Adams–Bohart and Yan models for Hg adsorption by SP in column test.
VariablesYan ModelAdams–Bohart Model
KYqYR2SSAREtbpτKAB(×10−3)N0R2SSARE
(mg g−1) (×10−5)(%)(Year) (L min−1 mg−1)(mg L−1) (×10−5)(%)
C0 (mg L−1) a
0.0800.48531870.96211.88.2911.836467.502.750.9575.3813.8
0.0560.49823440.9833.1714.824.5614210.62.180.9166.0362.6
0.0240.52018700.9821.426.0842.7835022.81.110.9261.8716.5
νp (cm min−1) b
0.2530.38317640.9894.037.947.16392717.92.220.9225.6717.4
0.1390.49823440.9833.1714.824.5614210.62.180.9166.0362.6
0.0800.60329840.9777.6229.069.411,5316.012.220.9204.7623.3
m (mg) c
0.50.48531870.96211.88.2911.836467.502.750.9575.3813.8
1.00.50424740.9608.4656.426.3580627.332.820.9631.4413.32
1.50.57617590.9671.0316.947.9297057.372.800.9645.5613.29
Note: The results of the Yan model and Adams–Bohart model were fitted for 150 PVs and 35 PVs, respectively. C0 (mg L−1) is the influent concentration of Hg2+; vp (cm min−1) is the pore velocity of groundwater, i.e., the distance flowed per minute; m (mg) is the amount of material used; KY and qY (mg g−1) are the constant and the maximum adsorption capacity for the Yan model, respectively; tb (h) is the time of 95% breakthrough (Ct/C0 = 0.95); pτ is the number of pore volumes at 50% adsorbate breakthrough (Ct/C0 = 0.5) for the Yan model; KAB (L min−1 mg−1) and N0 (mg L−1) are the rate constant and saturated adsorption capacity per column volume for the Adams–Bohart model, respectively; R2, SS, and ARE are the coefficients of determination, least sum of squares, and average relative error, respectively. Detailed information is shown in Text S3. Experimental conditions: a: νp = 0.139 cm min−1, m = 0.5 mg, and 25 ± 1 °C; b: C0 = 0.056 mg L−1, m = 0.5 mg, and 25 ± 1 °C; c: C0 = 0.080 mg L−1, νp = 0.139 cm min−1, and 25 ± 1 °C.
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

Wang, Z.; Liu, W.; Sun, X.; Zhang, Q.; Ji, J.; Yan, Y.; Sun, J. Remediation of Hg-Contaminated Groundwater via Adsorption on Supramolecular Polymers in Batch Process and Column Test. Molecules 2025, 30, 1406. https://doi.org/10.3390/molecules30071406

AMA Style

Wang Z, Liu W, Sun X, Zhang Q, Ji J, Yan Y, Sun J. Remediation of Hg-Contaminated Groundwater via Adsorption on Supramolecular Polymers in Batch Process and Column Test. Molecules. 2025; 30(7):1406. https://doi.org/10.3390/molecules30071406

Chicago/Turabian Style

Wang, Zongwu, Wei Liu, Xiaoyan Sun, Qing Zhang, Jiapu Ji, Yimeng Yan, and Jianhui Sun. 2025. "Remediation of Hg-Contaminated Groundwater via Adsorption on Supramolecular Polymers in Batch Process and Column Test" Molecules 30, no. 7: 1406. https://doi.org/10.3390/molecules30071406

APA Style

Wang, Z., Liu, W., Sun, X., Zhang, Q., Ji, J., Yan, Y., & Sun, J. (2025). Remediation of Hg-Contaminated Groundwater via Adsorption on Supramolecular Polymers in Batch Process and Column Test. Molecules, 30(7), 1406. https://doi.org/10.3390/molecules30071406

Article Metrics

Back to TopTop