Next Article in Journal
A Comparative Study of Drying Technologies for Apple and Ginger Pomace: Kinetic Modeling and Antioxidant Properties
Previous Article in Journal
Feasibility Verification of Casing Drilling in Shallow Marine Formations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal of Cd2+ and Pb2+ from an Aqueous Solution Using Modified Coal Gangue: Characterization, Performance, and Mechanisms

1
State Key Laboratory of Water Resource Protection and Utilization in Coal Mining, Beijing 102211, China
2
School of Chemical and Environmental Engineering, China University of Mining and Technology (Beijing), Beijing 100083, China
3
National Institute of Clean and Low-Carbon Energy (NICE), Beijing 102211, China
*
Author to whom correspondence should be addressed.
Processes 2024, 12(10), 2095; https://doi.org/10.3390/pr12102095 (registering DOI)
Submission received: 6 September 2024 / Revised: 23 September 2024 / Accepted: 24 September 2024 / Published: 26 September 2024
(This article belongs to the Section Separation Processes)

Abstract

:
The impact of various modification methods on enhancing the adsorption performance of coal gangue (CG) for hazardous heavy metals has not been thoroughly investigated. In this study, three CG samples were first modified by calcination, followed by acid washing, alkali washing, and hydrothermal treatment, to obtain modified CG samples. The adsorption performance was assessed based on the adsorption capacities for Cd2⁺ and Pb2⁺ (i.e., qe,Cd and qe,Pb), and the kinetics of the adsorption processes were analyzed using kinetic equations. XRD, SEM-EDX, FTIR, and N2 adsorption–desorption isotherms were used to elucidate the adsorption mechanisms. Results indicated that qe,Cd and qe,Pb of raw CG samples were approximately 10 and 25 mg/g, respectively, with only slight changes observed after calcination, acid washing, and alkali washing. In contrast, hydrothermal treatment yielded NaP and NaA zeolites, which significantly enhanced qe,Cd and qe,Pb to values of 48.5–72.7 and 214.9–247.5 mg/g, respectively. The hydrothermally treated CG samples primarily adsorbed Cd2⁺ and Pb2⁺ through ion exchange with Na⁺ within the zeolite structure, facilitating the entry of these ions into the zeolite’s pore channels. The adsorption processes were effectively described by the pseudo-second-order kinetic model. By optimizing the conditions of hydrothermal modification, the adsorption performance of CG samples is anticipated to further improve due to the creation of additional adsorption sites.

1. Introduction

Heavy metal pollution from industrial wastewater has become a global concern. Due to their tendency to bioaccumulate in living organisms, heavy metals in aquatic environments pose significant threats to both ecological systems and public health [1,2]. These metals also adversely affect soil and plants, leading to issues, such as stunted growth, altered enzyme activities, and disrupted photosynthesis. Cadmium (Cd) and lead (Pb), as prominent examples of heavy metals, have become widely distributed in the environment due to their extensive use in industrial processes and commercial applications [3,4]. For example, excessive Pb levels in the human body can damage the nervous system and kidneys and affect the cardiovascular, skeletal, reproductive, and immune systems. Current treatment technologies for heavy-metal-containing wastewater include ion exchange, membrane separation, electrolysis, chemical precipitation, and adsorption. Among these, adsorption is widely promoted due to its low cost, simplicity, and lack of secondary pollution [5,6]. The core of adsorption technology lies in developing adsorbents with high adsorption capacity and low cost. Researchers have investigated a variety of adsorbents, mainly including zeolite materials [7], carbon-based materials [8], and clay minerals [9].
Coal gangue (CG) is a solid waste generated during coal mining, accounting for 10% to 20% of the raw coal yield. The primary mineral constituents of CG include clay minerals, such as kaolinite, montmorillonite, and illite, along with quartz, feldspar, mica, and pyrite. Modifying CG can enhance its porosity and ion exchange capacity, facilitating the development of modified CG adsorbents [10]. High-temperature calcination is a widely used method for CG modification [11,12]. For instance, Qiu et al. [11] incorporated sodium tetraborate (Na2B4O7·10H2O) into the calcination process, resulting in a more than seven-fold increase in the Mn2⁺ adsorption capacity of the modified CG compared to raw CG. They also introduced a hydrothermal modification technique that produces zeolite from CG, which significantly boosts its adsorption capacity for Cd2⁺ to 183.7 mg/g [13]. Additionally, acid washing and alkali washing following calcination have been shown to improve the adsorption performance of CG samples [14,15,16]. Researchers have also developed various methods to create adsorption sites on CG samples [17,18]. For example, Shang et al. [17] devised a technique to introduce mercapto groups (–SH) onto CG, enhancing its adsorption capacity for Pb2⁺, Cd2⁺, and Hg2⁺ to 332.8, 110.4, and 179.2 mg/g, respectively.
Coal gangue is rich in SiO2 and Al2O3, with concentrations ranging from 60% to 90%, making it a natural raw material for the production of zeolites. Zeolites exhibit excellent adsorption properties. However, traditional production methods typically use chemical raw materials, such as Al(OH)3, NaOH, and Na2SiO3·H2O, which are relatively costly. Using coal gangue as a synthesis source can significantly reduce production costs. Alkaline-fusion-assisted hydrothermal crystallization is the most common method used to synthesize zeolite [7]. NaX zeolite, NaY zeolite, and NaA zeolite were prepared, and their adsorption behaviors for heavy metal ions were explored [5,6,19]. For instance, Ge et al. [5] synthesized NaX zeolite from coal gangue, and the maximum Pb2+ adsorption capacity reached was 457 mg/g. Jin et al. [6] demonstrated a low-cost method of preparing zeolite NaA from coal gangue, which involves using a low alkali concentration (1.8 mol/L) and a longer crystallization time (10 h). In order to improve adsorption performance, Bu et al. [20] developed a sequential alkaline and ultrasonic post-treatment method that facilitates the creation of hierarchical mesopores, thereby enhancing the adsorption performance of mesoporous zeolites.
Due to economic considerations, the modification of CG presents a more feasible approach for producing absorbents compared to the synthesis of zeolite, which requires extensive amounts of sodium hydroxide and prolonged high-temperature calcination. Researchers have explored various modification techniques, including calcination, acid washing, alkali washing, hydrothermal treatment, and intercalation. However, there is limited research comparing the effectiveness of these methods in enhancing the adsorption performance of CG for heavy metal ions. This study addressed this gap by modifying three CG samples with different SiO2/Al2O3 molar ratios through calcination, followed by acid washing, alkali washing, and hydrothermal treatment, to produce modified CG samples. The adsorption performance was evaluated based on the capacities to adsorb Cd2⁺ and Pb2⁺, and the kinetics of the adsorption processes were analyzed using appropriate kinetic models. Adsorption mechanisms were elucidated using X-ray diffraction (XRD), scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM-EDX), Fourier-transform infrared spectroscopy (FTIR), and nitrogen adsorption–desorption isotherms. This study aimed to provide a comprehensive understanding of how different modification methods affect the compositional properties and adsorption capabilities of CG samples, offering guidance for future efforts to develop high-capacity, cost-effective heavy metal adsorbents.

2. Experimental Section

2.1. Materials

In this study, CG samples were collected from three different locations, the Fumin mine in Inner Mongolia, the Yangshita mine in Inner Mongolia, and the Sihou mine in Shanxi Province, which were denoted as FM, YST, and SH, respectively. The chemical compositions of these samples were analyzed using X-ray fluorescence spectroscopy (XRF-1800, Shimadzu, Kyoto, Japan), with the results summarized in Table 1. The molar ratios of SiO2 to Al2O3 were 5.49 for FM, 5.13 for YST, and 2.86 for SH, indicating that SH has a higher aluminum content compared to FM and YST.
The proximate analysis of CG samples was conducted following the Chinese national standard GB/T 212-2008 [21], and the results are shown in Table 2. The lowest ash content in the YST sample indicated that this sample contained the highest amount of organic matter among all the samples. The raw CG samples were initially crushed, screened to particles smaller than 180 μm, and then dried at 105 °C for 24 h. The dried CG powder was subsequently used for modification. Additional chemicals used in this study, including HCl, NaOH, hexadecyl trimethyl ammonium bromide (HTAB), CdCl2·2.5H2O, and Pb(NO3)2, were purchased from Sinopharm Chemical Reagent Co., Ltd. Solutions of Cd2⁺ and Pb2⁺ at a concentration of 1000 mg/L were prepared by dissolving precise amounts of CdCl2·2.5H2O and Pb(NO3)2 in distilled water.

2.2. Preparation of Modified Coal Gangue

The raw CG samples underwent various modification treatments, including high-temperature calcination, acid washing, alkali washing, and hydrothermal treatment, as illustrated in Figure 1. The procedures for each treatment are detailed next:
(1) Calcination modification: FM, YST, and SH were placed in a crucible and calcined at 600 °C for 5 h in a muffle furnace (DC-B08/12, Dotrust, Beijing, China). The resulting samples were designated as FM-C, YST-C, and SH-C, respectively.
(2) Acid washing treatment: In total, 10 g of the calcined CG was stirred with 100 mL of 1 M HCl solution at ambient temperature for 24 h. Afterward, the solution was removed by filtration under reduced pressure, and the insoluble residue was washed with distilled water until the pH of the filtrate approached 7.0. The filtered residue was then dried at 105 °C for 12 h, yielding the acid-modified samples FM-C-AC, YST-C-AC, and SH-C-AC.
(3) Alkali washing treatment: This treatment followed a procedure similar to that of acid washing, except that the calcined CG was stirred with 100 mL of 1 M NaOH solution at ambient temperature for 24 h. The resulting alkali-modified samples were designated as FM-C-AL, YST-C-AL, and SH-C-AL.
(4) Hydrothermal treatment: Hydrothermal modification was conducted, as described by Qiu et al. [13]. Briefly, 10 g of calcined CG, 100 mL of 2 M NaOH solution, and 1 g of hexadecyl trimethyl ammonium bromide (HTAB) were added to a 200 mL three-necked round-bottom flask equipped with a reflux condenser. The mixture was continuously stirred at 100 °C for 24 h. The resulting colloidal suspension was then filtered under vacuum and washed with distilled water, and the residue was dried at 105 °C for 12 h, resulting in the hydrothermally modified samples FM-C-HY, YST-C-HY, and SH-C-HY.

2.3. Adsorption of Cd2+ and Pb2+ in Aqueous Solution

2.3.1. Adsorption Capacity Test

For the adsorption capacity experiment, 50 mL each of 100 mg/L Cd2⁺ and Pb2⁺ solutions was mixed in a conical flask to simulate wastewater. These solutions were prepared by diluting a 1000 mg/L stock solution with distilled water, adjusting the final pH to approximately 7.0. To investigate adsorption, 0.1 g of the modified coal gangue was introduced into the wastewater. The mixture was incubated at room temperature for 1800 min on a horizontal shaker (HY-2, Supo, Shaoxing, China). After the adsorption period, the samples were filtered through a 0.45 μm membrane filter. The concentrations of Cd2⁺ and Pb2⁺ in the filtrate were determined using a UV–visible spectrophotometer (UV-1100, Mapada, Shanghai, China) [22]. In the tests involving hydrothermally modified CG samples, it was observed that both Cd2⁺ and Pb2⁺ ions were fully adsorbed by a 0.1 g sample. Consequently, the dosages were adjusted to 0.04 g for Cd2⁺ and 0.01 g for Pb2⁺. The adsorption capacity was calculated using Equation (1) [17]:
q e = c 0 c e × V m
where qe represents the adsorption capacity (mg/g) and c0 and ce denote the initial and equilibrium concentrations of the heavy metal ions, respectively (mg/L). The volume of the wastewater is represented by V (L), and m is the mass of the solid adsorbent added to the flask (g). Each adsorption test was performed in triplicate, and the average values were used for the final data analysis. Specifically, qe,Cd and qe,Pb refer to the adsorption capacities for Cd2⁺ and Pb2⁺, respectively.

2.3.2. Adsorption Kinetic Test

For adsorption kinetic tests, 1.6 g of hydrothermally modified CG samples was combined with 2000 mL of Cd2⁺ solution, while 0.4 g of the same samples was mixed with 2000 mL of Pb2⁺ solution. Both solutions had an initial concentration of 100 mg/L. The pH of the solutions was approximately 7.0 and was not adjusted further. The mixtures were incubated at room temperature on a horizontal shaker (HY-2, Supo, Shaoxing, China). At specified time intervals (0, 2, 5, 10, 20, 30, 60, 90, 120, 180, 360, 540, 1440, 1560, 1680, and 1800 min), 1 mL of the samples was withdrawn from the conical flasks. The total volume of the samples taken did not exceed 2% of the initial solution volume. The extracted samples were filtered through 0.45 μm membrane filters, and the ion concentrations were measured using a UV–visible spectrophotometer (UV-1100, Mapada, Shanghai, China) [22]. The amount of ions adsorbed at time t (qt, mg/g) was calculated using Equation (2) [13]:
q t = c 0 c t × V m
where qt represents the adsorption amount (mg/g) and ct denotes the concentration of heavy metal ions at time t (mg/L). Each adsorption test was performed in triplicate, with the average values used for final data analysis. Specifically, qt,Cd and qt,Pb refer to the adsorption amounts for Cd2⁺ and Pb2⁺, respectively. Following the adsorption tests, the CG samples were filtered, washed, and dried using the previously described procedures. The samples were then labeled as follows: FM-C-HY (Cd), YST-C-HY (Cd), and SH-C-HY (Cd) for the hydrothermally modified CG samples after Cd2⁺ adsorption and FM-C-HY (Pb), YST-C-HY (Pb), and SH-C-HY (Pb) for the hydrothermally modified CG samples after Pb2⁺ adsorption.

2.4. Modified CG Characterization

The crystal phase of the modified CG samples was analyzed using XRD with a Bruker D8 diffractometer, which used Cu Kα radiation (λ = 0.15418 nm) and operated at 40 kV and 100 mA. XRD scans were conducted over a 2θ range of 5–60° at a scanning rate of 10°/min. The surface morphology of the modified CG samples was examined with a scanning electron microscope (Gemini SEM 300, Zeiss, Oberkochen, Germany).
To explore the adsorption mechanism, EDX, FTIR and N2 adsorption–desorption isotherms were performed on hydrothermally modified CG samples both before and after adsorption. The distribution of heavy metals in the modified CG samples was analyzed using an energy-dispersive spectrometer (XFlash Detector 5010, Bruker, Ettlingen, Germany). FTIR spectra were recorded in the mid-infrared range from 400 to 4000 cm⁻1 with a Fourier transform infrared spectrometer (Nicolet 6700, ThermoFisher, Waltham, MA, USA). N2 adsorption–desorption measurements were carried out using an automatic surface and porosity analyzer (ASAP 2460, Micromeritics, Atlanta, GA, USA).

3. Results and Discussion

3.1. Properties of Modified CG Samples

Figure 2 displays the XRD patterns for both raw and modified CG samples. The primary mineral phases identified in the raw CG samples included quartz, kaolinite, and muscovite. Additionally, nontronite and illite were observed in the FM and SH samples, respectively. The diffraction peaks corresponding to kaolinite and nontronite were notably diminished following calcination at 600 °C. This reduction is attributed to the decomposition of hydroxyl groups within these clay minerals, which leads to the release of structural water and the subsequent collapse of the interlayer structure [15]. It has been proposed that kaolinite transforms into disordered metakaolinite upon calcination at this temperature [11]. In comparison to the calcined CG samples, both acid and alkali washing procedures enhanced the peak intensities of the mineral crystal phases. Acid treatment effectively dissolves metal oxides and carbonate minerals, such as Fe2O3, CaO, and CaCO3, while alkali treatment removes some amorphous silicates by forming soluble sodium silicate. The resulting increase in the purity and concentration of quartz and kaolinite thus improved the overall crystallinity of the coal gangue components [14].
Figure 2 also illustrates that the hydrothermal modification of the calcined CG samples resulted in the formation of different zeolites. Specifically, NaP zeolite was synthesized in the FM-C-HY and YST-C-HY samples, whereas NaA zeolite was produced in the SH-C-HY sample. Qiu et al. [13] also observed the formation of zeolite materials, such as Na6Al6Si10O32 and Na12Al12Si12O48, following the hydrothermal modification of combusted coal gangue. It is hypothesized that calcination and hydrothermal treatment partially disrupted the structures of the CG samples, generating AlO4 and SiO4 tetrahedral units. Under the structural guidance of HTAB, these active units condensed into relatively ordered aluminosilicate gels, which subsequently crystallized into zeolite structures [7]. The SiO2/Al2O3 ratio in the aluminosilicate gel plays a crucial role in determining the type of zeolite synthesized. Higher SiO2/Al2O3 ratios generally lead to the formation of NaP, NaX, and NaY zeolites, whereas lower ratios favor the production of NaA zeolite [23,24]. In this study, NaP zeolite was obtained from the FM and YST samples, which had high SiO2/Al2O3 ratios of 5.49 and 5.13, respectively, while NaA zeolite was synthesized from the SH sample, which had a lower SiO2/Al2O3 ratio of 2.86. These zeolite crystals were thought to improve the adsorption performance of modified CG samples.
Figure 3, Figure 4 and Figure 5 present SEM images of both raw and modified CG samples. As depicted in Figure 3a, Figure 4a and Figure 5a, the raw CG samples exhibited irregular shapes with a dense texture and were characterized by fine particles and layers adhering to their surfaces. Generally, raw CG samples exhibit poor adsorption performance due to their low pore volume and specific surface area [25]. Despite undergoing calcination, acid washing, and alkali washing, no significant changes in the morphological characteristics of the CG samples were observed. During calcination at 600 °C, the combustion of inherent organic matter and the decomposition of minerals, such as kaolinite, are expected to contribute to changes in the CG morphology. For instance, Jablonska et al. [10] reported that thermal modification at 600 °C increased the mesopore and macropore volumes of a Polish CG sample, while decreasing the total specific surface area. Similarly, Qiu et al. [11] found that the addition of Na2B4O7·10H2O during CG calcination weakened and fractured CG bonds, resulting in an increase in the BET surface area from 9.29 to 20.05 m2/g. Furthermore, Gao et al. [14] observed that treatment with HCl and KOH solutions reduced the particle size and generated irregular layered structures, enhancing the surface area. The effectiveness of calcination, acid washing, and alkali treatment in developing the pore structure of CG is likely influenced by various factors, including the organic matter content, mineral composition, concentration of the acid/alkali solution, and processing temperature. In this study, however, no significant changes in the surface morphology of CG samples were detected from SEM images.
Figure 3e, Figure 4e and Figure 5e display the SEM images of FM-C-HY, YST-C-HY, and SH-C-HY, respectively. Compared to the raw and other modified CG samples, these hydrothermally modified CG samples exhibited aggregates of fine particles with a loose and porous texture. Additionally, cubic particles, highlighted by red circles in Figure 4e and Figure 5e, were observed. These aggregates and cubic particles were likely zeolite precursors and crystals, respectively. To achieve high-crystallinity zeolites from CG, alkali fusion is commonly used to activate the CG, promoting the conversion of CG into soluble aluminosilicate gel [7]. It is presumed that the 2 M NaOH solution used in this study effectively dissolved only the amorphous silicate and aluminate minerals present in the CG samples. Due to the resulting solution’s low Al–Si supersaturation, the growth of zeolite nuclei during the crystallization phase was inhibited. Consequently, fewer zeolite crystals were formed, and the majority of the zeolite phase exhibited low crystallinity.

3.2. Adsorption of Cd2+ and Pb2+ by Modified CG Samples

3.2.1. Adsorption Capacity

Figure 6 illustrates the adsorption capacities of raw and modified CG samples for Cd2⁺ and Pb2⁺ in aqueous solutions. The raw CG samples exhibited relatively poor adsorption performance, with qe,Cd and qe,Pb values of approximately 10 and 25 mg/g, respectively. In contrast, the FM sample demonstrated the highest adsorption capacity for both Cd2⁺ and Pb2⁺, likely due to its higher clay mineral content. Clay minerals, such as kaolinite and nontronite, can adsorb metal ions through ion exchange [9]. After thermal modification at 600 °C, YST-C showed a greater ion adsorption capacity compared to FM-C and SH-C, though the differences were minimal. This enhancement is likely attributed to the significant amount of intrinsic organic matter in YST, which resulted in a more developed porous structure in YST-C following the combustion of the organic matter. Acid and alkali washing resulted in only minor changes in qe,Cd and qe,Pb, making it difficult to draw a definitive conclusion due to the complexity of the adsorption process. Gao et al. [14] reported that KOH washing increased the adsorption capacity of CG for U(VI) from 44 to 140 mg/g, while HCl washing had negligible effects. They noted that lower-binding-energy Al has a higher affinity for U(VI), and the increased Al/Si ratio after alkali washing contributes to better adsorption performance. Based on these findings, calcination, acid washing, and alkali washing appear to be ineffective methods for improving adsorption performance.
Figure 6 also shows significant enhancement in the adsorption performance of CG samples following hydrothermal treatment. The equilibrium adsorption capacities for Cd2+ (qe,Cd) of FM-C-HY, YST-C-HY, and SH-C-HY were 48.5, 72.7, and 65.6 mg/g, respectively. The qe,Pb values of FM-C-HY, YST-C-HY, and SH-C-HY were 214.9, 247.5, and 242.3 mg/g, respectively. This improvement in adsorption capacity can be attributed to the formation of zeolite, which offers a high surface area, a stable crystal structure, and a substantial cavity volume. The zeolite facilitates the accommodation of Cd2⁺ and Pb2⁺ ions through physical adsorption and ion exchange [6,26]. Notably, the adsorption capacity for Pb2⁺ was higher than that for Cd2⁺ in the hydrothermally modified samples. Shang et al. [17] observed similar results, finding that the maximum adsorption capacities of mercapto-modified CG for Cd2⁺ and Pb2⁺ were 110.4 and 332.8 mg/g, respectively. Wingenfelder et al. [27] noted that the hydrated forms of Cd2⁺ and Pb2⁺ slightly exceed the width of zeolite pore channels, indicating that only dehydrated cations can penetrate these channels. The lower hydration energy of Pb2⁺ accounts for its preferential sorption over Cd2⁺. The series YST-C-HY > SH-C-HY > FM-C-HY showed a decrease in both qe,Cd and qe,Pb despite YST-C-HY and FM-C-HY containing the same type of zeolite (NaP). SEM analysis suggested that the lower adsorption capacity of FM-C-HY may be due to the lower crystallinity of the NaP zeolite it contains. Well-crystallized zeolites possess a more developed pore channel system, which enhances their ability to retain more Cd2⁺ and Pb2⁺.

3.2.2. Adsorption Kinetics

Given that hydrothermal treatment significantly enhanced the adsorption performance for Cd2⁺ and Pb2⁺, our primary focus was on investigating the adsorption kinetics of the hydrothermally modified CG samples. Figure 7 illustrates the adsorption profiles of heavy metal ions over varying time periods. It is evident that both Cd2⁺ and Pb2⁺ ions exhibited a significant increase in adsorption within the initial 180 min, followed by a gradual approach to equilibrium by 540 min. This initial rapid uptake can be attributed to the high availability of vacant adsorption sites on the modified CG samples and the steep concentration gradient of metal ions between the solution and the adsorbent [26]. As these sites become occupied, the rate of adsorption slows down [13]. Additionally, a slight decrease in qt,Cd and qt,Pb toward the end of the adsorption period suggests a possible desorption of metal ions.
Adsorption kinetic models are valuable tools for elucidating the mechanism of adsorption and assessing the effectiveness of adsorbents in pollutant removal. Among these models, the pseudo-first-order and pseudo-second-order models are commonly used across various adsorption systems. The linear representation of the pseudo-first-order kinetics is given by Equation (3) [13,17]:
ln q e q t = ln q e k 1 t
where k1 represents the rate constant for the pseudo-first-order model (1/min). The values of qe and k1 are obtained from the slope and intercept of the linear plot of ln(qeqt) versus t, respectively. The linear form of the pseudo-second-order model is expressed by Equation (4) [13,17]:
t q t = 1 k 2 q e 2 + t q e
where k2 (g/(mg·min)) denotes the rate constant for the pseudo-second-order model. The values of qe and k2 are derived from the slope and intercept of the plot of t/qt versus t, respectively. Figure 8 presents the kinetic fitting for the adsorption of Cd2⁺ and Pb2⁺ onto the modified CG samples.
As presented in Table 3, the high correlation coefficients and the close agreement between the calculated qe values and the experimental qe values suggest that the pseudo-second-order model is highly suitable for describing the adsorption kinetics of Cd2⁺ and Pb2⁺ on the modified CG samples. These findings align with those reported by other researchers [11,13,28]. The pseudo-second-order model typically accounts for both the diffusion process and the surface reaction process. This implies that the rate-limiting step in the adsorption process is surface adsorption, which is influenced by the availability of adsorption sites. Some researchers have also highlighted that chemisorption is the predominant mechanism during the adsorption process [1].

3.3. Adsorption Mechanism

To gain a deeper understanding of the adsorption mechanisms of modified CG samples, FM-C-HY, YST-C-HY, and SH-C-HY were characterized both before and after the adsorption of Cd2⁺ and Pb2⁺. The characterization techniques used included XRD, SEM-EDX, FTIR, and nitrogen adsorption–desorption isotherms.

3.3.1. XRD

Figure 9 illustrates that the XRD patterns of the modified CG samples before and after the adsorption of Cd2⁺ and Pb2⁺ were largely similar, suggesting that the overall zeolite structure is not significantly altered by the adsorption processes. However, minor variations in the XRD patterns were observed post adsorption. Specifically, Figure 9a,b demonstrates that the peak at 2θ = 21.5° (1 1 2) shifted to a higher diffraction angle with Cd2⁺ adsorption, while it shifted to a lower diffraction angle with Pb2⁺ adsorption. Similar shifts in diffraction peaks were reported by Pu et al. [29], who observed a shift of the peak at 2θ = 4.67° (NaP zeolite) to a higher diffraction angle (2θ = 4.83°). They attributed this shift to the replacement of Na⁺ (r = 1.02 Å) in the NaP zeolite lattice by the smaller Cu2⁺ (r = 0.73 Å), which resulted in a reduction in unit cell dimensions and an increase in the diffraction angle. Considering that the ionic radii of Cd2⁺ and Pb2⁺ are 0.95 Å and 1.19 Å, respectively, it is hypothesized that Cd2⁺ adsorption leads to a decrease in unit cell dimensions, while Pb2⁺ adsorption results in an increase.
Figure 9 also demonstrates that the peaks corresponding to the (1 0 1) lattice plane of NaP zeolite, as well as the (2 2 0) and (2 2 2) lattice planes of NaA zeolite, nearly vanished following the adsorption of Cd2⁺ and Pb2⁺. Similar observations were reported by Steinike et al. [30], who noted that bivalent heavy metal cations, such as Zn2⁺, Cd2⁺, and Pb2⁺, could occupy the Na⁺ positions within the α-cages and β-cages of the NaA zeolite. The disappearance of these diffraction peaks indicates the disruption of the corresponding ordered stacks of unit cells. This effect may be attributed to the low crystallinity of the derived NaP and NaA zeolites, which were partially decomposed during the adsorption process in aqueous solutions. It is suggested that cation exchange plays a significant role in the removal of Cd2⁺ and Pb2⁺ by hydrothermally modified CG samples.

3.3.2. SEM-EDX

Figure 10 presents the SEM images and EDX spectra of modified CG samples after the adsorption of Cd2⁺ and Pb2⁺. These images confirmed the presence of Cd2⁺ and Pb2⁺ on the surface of the modified CG samples, indicating that adsorption has occurred. Additionally, the content of Na significantly decreased or was nearly undetectable following adsorption. This reduction is attributed to the ion exchange process involving NaP and NaA zeolites in the modified CG samples, where Na⁺ ions are replaced by Cd2⁺ and Pb2⁺ ions. Similar findings have been reported by other researchers [31,32]. For instance, Lv et al. [31] observed that after the adsorption of Cd2⁺ and Pb2⁺ by NaA zeolite, the distribution of Na decreases relative to other elements.

3.3.3. N2 Adsorption–Desorption

Table 4 summarizes the pore structure parameters of the modified CG samples before and after adsorption. For both FM-C-HY and YST-C-HY, the adsorption of Cd2+ or Pb2+ resulted in a decrease in the total surface area (SBET) and the total pore volume (Vtotal), indicating that the heavy metal ions occupied some pore channels. Notably, FM-C-HY has a relatively larger surface area, but it does not exhibit better adsorption performance, indicating that the adsorption process is not solely physical. Since ion exchange was identified as an important adsorption mechanism, exchangeable Na+ within the pore channels of zeolites were thought to be active adsorption sites. FM-C-HY should contain more non-active surface area formed by the stacking of inherent CG minerals, such as quartz and muscovite. By contrast, SH-C-HY showed an increase in both SBET and the total pore volume, Vtotal, after the adsorption process. The possible reason is that Cd2+ or Pb2+ adsorption leads to an enlargement of the pore system of NaA zeolite.

3.3.4. FTIR

Figure 11 displays the FTIR spectra of modified CG samples before and after the adsorption of Cd2⁺ and Pb2⁺ ions. The peak at approximately 3400 cm⁻1 corresponded to the stretching vibration of O–H bonds, while the peak at around 1640 cm⁻1 was attributed to the bending vibration of O–H groups. The band near 960 cm⁻1 was assigned to the asymmetric stretching of Si–O–Si or Si–O–Al bridges [6]. Peaks observed at about 2920 and 2850 cm⁻1 were attributed to the asymmetrical and symmetrical stretching of alkyl C−H, indicating the introduction of hexadecyltrimethylammonium in the CG samples post-modification [33]. Importantly, the characteristic peaks of the CG samples after adsorption remained largely unchanged, although the peak near 960 cm⁻1 shifted to a higher wavenumber. The presence of exchangeable cations can influence the IR spectra of zeolite frameworks due to factors such as mass, charge, ion size, and the cation’s environment [34]. This shift is primarily due to the ion exchange between Na⁺ and Cd2⁺ or Na⁺ and Pb2⁺, causing a slight alteration in the zeolite structure. Some researchers have also noted shifts in the O–H peaks around 3400 and 1600 cm⁻1, indicating the involvement of hydroxyl groups in the removal of heavy metal ions by NaA and NaP zeolites [29,31,35].

3.4. Limitations and Implications of This Study

From the perspective of adsorption capacities for Cd2⁺ and Pb2⁺, hydrothermal treatment emerged as the preferred method for modifying CG. In this study, the hydrothermal modification of CG was conducted under fixed experimental conditions. Further research is needed to explore how variables such as the NaOH concentration, HTAB dosage, and processing time, as well as their interactions, affect adsorption performance. Additionally, ion exchange was identified as the principal mechanism for Cd2⁺ and Pb2⁺ adsorption by hydrothermally modified CG samples. Given that cation exchange sites are predominantly associated with aluminum–oxygen tetrahedra, enhanced adsorption performance could potentially be achieved by introducing an additional aluminum source to increase the number of adsorption sites. This approach will be a central focus of our future research. These findings are significant for the development of high-performance, cost-effective adsorbents derived from CG through hydrothermal treatment.

4. Conclusions

This study investigated three CG samples with different SiO2/Al2O3 molar ratios, subjected to four distinct modification methods, to assess and compare their compositional properties and adsorption capabilities. Additionally, the adsorption kinetics and mechanisms of the hydrothermally modified CG samples were examined. The principal findings are as follows:
(1)
The initial adsorption capacities of the unmodified CG samples for Cd2⁺ and Pb2⁺ were approximately 10 and 25 mg/g, respectively. These values remain largely unchanged following calcination, acid washing, and alkali washing. In contrast, hydrothermal treatment produces NaP and NaA zeolites, leading to substantial increases in the adsorption capacities for Cd2⁺ and Pb2⁺, reaching values of 48.5–72.7 and 214.9–247.5 mg/g, respectively.
(2)
Hydrothermally modified CG samples primarily remove Cd2⁺ and Pb2⁺ through ion exchange with Na⁺ within the zeolite framework, facilitating the ingress of these ions into the zeolite’s pore channels. The adsorption data are well described by the pseudo-second-order kinetic model, indicating that chemisorption is the dominant mechanism.
(3)
The study highlights the effective removal of Cd2⁺ and Pb2⁺ by hydrothermally modified CG samples, establishing hydrothermal treatment as an efficient and cost-effective modification strategy. Future research should aim at optimizing the hydrothermal process and exploring methods to increase the number of adsorption sites on the CG samples.

Author Contributions

Conceptualization, writing—original draft preparation, Z.C. and C.L.; methodology, investigation, C.L. and Y.Y.; validation, writing—review and editing, Z.C. and L.B.; formal analysis, funding acquisition, Z.C.; resources, supervision, project administration, L.B., N.G. and Z.X.; data curation, L.B.; visualization, C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Open Fund of the State Key Laboratory of Water Resource Protection and Utilization in Coal Mining (GJNY-21-41-19), the CHN Energy Investment Group (GJNY-22-92), the Fundamental Research Funds for the Central Universities (2024ZKPYHH04), and the China Scholarship Council (202306430017).

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mahmoud, E.R.I.; Aly, H.M.; Hassan, N.A.; Aljabri, A.; Khan, A.L.; El-Labban, H.F. Biochar from date palm waste via two-step pyrolysis: A modified approach for Cu (II) removal from aqueous solutions. Processes 2024, 12, 1189. [Google Scholar] [CrossRef]
  2. Zuo, J.; Ren, J.; Jiang, L.; Tan, C.; Li, J.; Xia, Z.; Wang, W. Preparation of PVA/SA-FMB microspheres and their adsorption of Cr(VI) in aqueous solution. Processes 2024, 12, 443. [Google Scholar] [CrossRef]
  3. Zheng, L.; Peng, D.; Meng, P. Promotion effects of nitrogenous and oxygenic functional groups on cadmium (II) removal by carboxylated corn stalk. J. Clean Prod. 2018, 201, 609–623. [Google Scholar] [CrossRef]
  4. Liu, L.; Huang, Y.; Zhang, S.; Gong, Y.; Su, Y.; Cao, J.; Hu, H. Adsorption characteristics and mechanism of Pb(II) by agricultural waste-derived biochars produced from a pilot-scale pyrolysis system. Waste Manag. 2019, 100, 287–295. [Google Scholar] [CrossRef]
  5. Ge, Q.; Moeen, M.; Tian, Q.; Xu, J.; Feng, K. Highly effective removal of Pb2+ in aqueous solution by Na-X zeolite derived from coal gangue. Environ. Sci. Pollut. Res. 2020, 27, 7398–7408. [Google Scholar] [CrossRef]
  6. Bu, N.; Liu, X.; Song, S.; Liu, J.; Yang, Q.; Li, R.; Zheng, F.; Yan, L.; Zhen, Q.; Zhang, J. Synthesis of NaY zeolite from coal gangue and its characterization for lead removal from aqueous solution. Adv. Powder Technol. 2020, 31, 2699–2710. [Google Scholar] [CrossRef]
  7. Zhang, X.; Li, C.; Zheng, S.; Di, Y.; Sun, Z. A review of the synthesis and application of zeolites from coal-based solid wastes. Int. J. Miner. Metall. Mater. 2022, 29, 1–21. [Google Scholar] [CrossRef]
  8. Shakoor, M.B.; Shafaqat, A.; Muhammad, R.; Farhat, A.; Irshad, B.; Muhammad, R.; Usman, K.; Nabeel, K.N.; Jörg, R. A review of biochar-based sorbents for separation of heavy metals from water. Int. J. Phytoremediat. 2019, 22, 111–126. [Google Scholar] [CrossRef]
  9. Otunola, B.O.; Ololade, O.O. A review on the application of clay minerals as heavy metal adsorbents for remediation purposes. Env. Technol. Innov. 2020, 18, 100692. [Google Scholar] [CrossRef]
  10. Jabłońska, B.; Kityk, A.V.; Busch, M.; Huber, P. The structural and surface properties of natural and modified coal gangue. J. Env. Manag. 2017, 190, 80–90. [Google Scholar] [CrossRef]
  11. Qiu, R.; Cheng, F. Modification of waste coal gangue and its application in the removal of Mn2+ from aqueous solution. Water Sci. Technol. 2016, 74, 524–534. [Google Scholar] [CrossRef] [PubMed]
  12. Yang, K.; Hu, L.; Zhu, J.; Luo, Z.; Zhang, Z.; He, J.; Chen, X. Adsorption performance of calcined coal gangue for ethyl mercaptan. Energy Sources Part A Recovery Util. Environ. Eff. 2023, 45, 4050–4063. [Google Scholar] [CrossRef]
  13. Qiu, R.; Cheng, F.; Huang, H. Removal of Cd2+ from aqueous solution using hydrothermally modified circulating fluidized bed fly ash resulting from coal gangue power plant. J. Clean Prod. 2018, 172, 1918–1927. [Google Scholar] [CrossRef]
  14. Gao, Y.; Huang, J.; Li, M.; Dai, Z.; Jiang, R.; Zhang, J. Chemical modification of combusted coal gangue for U(Vi) adsorption: Towards a waste control by waste strategy. Sustainability 2021, 13, 8421. [Google Scholar] [CrossRef]
  15. Peng, L.; Wang, R.; Cheng, H.; Zhang, L.; He, Y.; Yin, C.; Zhang, X. Investigation on the adsorption performance of modified coal gangues to p-hydroxybenzenesulfonic acid. Korean J. Chem. Eng. 2023, 40, 1767–1774. [Google Scholar] [CrossRef]
  16. Li, Z.; Wu, L.; Sun, S.; Gao, J.; Zhang, H.; Zhang, Z.; Wang, Z. Disinfection and removal performance for escherichia coli, toxic heavy metals and arsenic by wood vinegar-modified zeolite. Ecotoxicol. Env. Saf. 2019, 174, 129–136. [Google Scholar] [CrossRef]
  17. Shang, Z.; Zhang, L.W.; Zhao, X.; Liu, S.; Li, D. Removal of Pb(II), Cd(II) and Hg(II) from aqueous solution by mercapto-modified coal gangue. J. Env. Manag. 2019, 231, 391–396. [Google Scholar] [CrossRef]
  18. Zhang, X.; Li, M.; Su, Y.; Du, C. A novel and green strategy for efficient removing Cr(VI) by modified kaolinite-rich coal gangue. Appl. Clay Sci. 2021, 211, 106208. [Google Scholar] [CrossRef]
  19. Jin, Y.; Li, L.; Liu, Z.; Zhu, S.; Wang, D. Synthesis and characterization of low-cost zeolite NaA from coal gangue by hydrothermal method. Adv. Powder Technol. 2021, 32, 791–801. [Google Scholar] [CrossRef]
  20. Zhou, J.; Zheng, F.; Li, H.; Wang, J.; Bu, N.; Hu, P.; Gao, J.; Zhen, Q.; Bashir, S.; Louise Liu, J. Optimization of post-treatment variables to produce hierarchical porous zeolites from coal gangue to enhance adsorption performance. Chem. Eng. J. 2020, 381, 122698. [Google Scholar] [CrossRef]
  21. GB/T 212-2008; General Administration of Quality Supervision, Inspection and Quarantine of the People’s Republic of China, Proximate Analysis of Coal. National Standards of the People’s Republic of China: Beijing, China, 2008.
  22. Watanabe, H.; Ohmori, H. Dual-wavelength spectrophotometric determination of Cadmium with cadion. Talanta 1979, 26, 959–961. [Google Scholar] [CrossRef] [PubMed]
  23. Inada, M.; Eguchi, Y.; Enomoto, N.; Hojo, J. Synthesis of zeolite from coal fly ashes with different silica-alumina composition. Fuel 2005, 84, 299–304. [Google Scholar] [CrossRef]
  24. Zhang, X.; Tang, D.; Zhang, M.; Yang, R. Synthesis of NaX zeolite: Influence of crystallization time, temperature and batch molar ratio SiO2/Al2O3 on the particulate properties of zeolite crystals. Powder Technol. 2013, 235, 322–328. [Google Scholar] [CrossRef]
  25. Jin, Y.; Liu, Z.; Han, L.; Zhang, Y.; Li, L.; Zhu, S.; Li, Z.P.J.; Wang, D. Synthesis of coal-analcime composite from coal gangue and its adsorption performance on heavy metal ions. J. Hazard Mater. 2022, 423, 127027. [Google Scholar] [CrossRef]
  26. Liang, Z.; Gao, Q.; Wu, Z.; Gao, H. Removal and kinetics of cadmium and copper ion adsorption in aqueous solution by zeolite NaX synthesized from coal gangue. Environ. Sci. Pollut. Res. 2022, 29, 84651–84660. [Google Scholar] [CrossRef] [PubMed]
  27. Wingenfelder, U.; Nowack, B.; Furrer, G.; Schulin, R. Adsorption of Pb and Cd by amine-modified zeolite. Water Res. 2005, 39, 3287–3297. [Google Scholar] [CrossRef] [PubMed]
  28. Marhoon, A.A.; Hasbullah, S.A.; Asikin-Mijan, N.; Mokhtar, W.N.A.W. Hydrothermal synthesis of high-purity zeolite X from coal fly ash for heavy metal removal: Kinetic and isotherm analysis. Adv. Powder Technol. 2023, 34, 104242. [Google Scholar] [CrossRef]
  29. Pu, X.; Yao, L.; Yang, L.; Jiang, W.; Jiang, X. Utilization of industrial waste lithium-silicon-powder for the fabrication of novel nap zeolite for aqueous Cu(II) removal. J. Clean Prod. 2020, 265, 121822. [Google Scholar] [CrossRef]
  30. Steinike, U.; Jancke, K.; Lutz, W.; Schreier, E.; Walther, G. NaA-zeolites modified by Me2+-cations. Mater. Sci. Forum 1996, 228, 669–676. [Google Scholar] [CrossRef]
  31. Lv, Y.; Ma, B.; Liu, Y.; Wang, C.; Chen, Y. Adsorption behavior and mechanism of mixed heavy metal ions by zeolite adsorbent prepared from lithium leach residue. Microporous Mesoporous Mater. 2022, 329, 111553. [Google Scholar] [CrossRef]
  32. Cui, W.L.; Tang, K.; Chen, Y.; Chen, Z.; Lan, Y.; Hong, Y.B.; Lan, W.G. Regulating the particle sizes of NaA molecular sieves toward enhanced heavy metal ion adsorption. New J. Chem. 2024, 48, 7863–7874. [Google Scholar] [CrossRef]
  33. Król, M.; Mozgawa, W.; Jastrzbski, W.; Barczyk, K. Application of IR spectra in the studies of zeolites from D4R and D6R structural groups. Microporous Mesoporous Mater. 2012, 156, 181–188. [Google Scholar] [CrossRef]
  34. Wen, J.; Yan, C.; Xing, L.; Wang, Q.; Yuan, L.; Hu, X.H. Simultaneous immobilization of As and Cd in a mining site soil using HDTMA-modified zeolite. Env. Sci. Pollut. Res. 2021, 28, 9935–9945. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, W.; Xu, F.; Wang, Y.; Luo, M.; Wang, D. Facile control of zeolite NaA dispersion into xanthan gum-alginate binary biopolymer network in improving hybrid composites for adsorptive removal of Co2+ and Ni2+. Chem. Eng. J. 2014, 255, 316–326. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the CG modification procedure.
Figure 1. Schematic diagram of the CG modification procedure.
Processes 12 02095 g001
Figure 2. XRD patterns of raw and modified CG samples: (a) FM CG samples, (b) YST CG samples, and (c) SH CG samples.
Figure 2. XRD patterns of raw and modified CG samples: (a) FM CG samples, (b) YST CG samples, and (c) SH CG samples.
Processes 12 02095 g002
Figure 3. SEM images of raw and modified FM samples: (a) FM, (b) FM-C, (c) FM-C-AC, (d) FM-C-AL, and (e) FM-C- HY.
Figure 3. SEM images of raw and modified FM samples: (a) FM, (b) FM-C, (c) FM-C-AC, (d) FM-C-AL, and (e) FM-C- HY.
Processes 12 02095 g003aProcesses 12 02095 g003b
Figure 4. SEM images of raw and modified YST samples: (a) YST, (b) YST-C, (c) YST-C-AC, (d) YST-C-AL, and (e) YST-C- HY.
Figure 4. SEM images of raw and modified YST samples: (a) YST, (b) YST-C, (c) YST-C-AC, (d) YST-C-AL, and (e) YST-C- HY.
Processes 12 02095 g004
Figure 5. SEM images of raw and modified SH samples: (a) SH, (b) SH-C, (c) SH-C-AC, (d) SH-C-AL, and (e) SH-C- HY.
Figure 5. SEM images of raw and modified SH samples: (a) SH, (b) SH-C, (c) SH-C-AC, (d) SH-C-AL, and (e) SH-C- HY.
Processes 12 02095 g005aProcesses 12 02095 g005b
Figure 6. Adsorption capacity of raw and modified CG samples for Cd2+ and Pb2+: (a) qe,Cd And (b) qe,Pb.
Figure 6. Adsorption capacity of raw and modified CG samples for Cd2+ and Pb2+: (a) qe,Cd And (b) qe,Pb.
Processes 12 02095 g006
Figure 7. The effect of adsorption time on the adsorption performance of modified CG samples: (a) Cd2+ and (b) Pb2+.
Figure 7. The effect of adsorption time on the adsorption performance of modified CG samples: (a) Cd2+ and (b) Pb2+.
Processes 12 02095 g007
Figure 8. Kinetic fitting curves of modified CG samples: (a) pseudo-first-order model fitting for Cd2+ adsorption, (b) pseudo-second-order model fitting for Cd2+ adsorption, (c) pseudo-first-order model fitting for Pb2+ adsorption and (d) pseudo-second-order model fitting for Pb2+ adsorption.
Figure 8. Kinetic fitting curves of modified CG samples: (a) pseudo-first-order model fitting for Cd2+ adsorption, (b) pseudo-second-order model fitting for Cd2+ adsorption, (c) pseudo-first-order model fitting for Pb2+ adsorption and (d) pseudo-second-order model fitting for Pb2+ adsorption.
Processes 12 02095 g008
Figure 9. XRD patterns of hydrothermally modified CG samples before and after adsorption: (a) FM-C-HY, (b) YST-C-HY, and (c) SH-C-HY.
Figure 9. XRD patterns of hydrothermally modified CG samples before and after adsorption: (a) FM-C-HY, (b) YST-C-HY, and (c) SH-C-HY.
Processes 12 02095 g009
Figure 10. SEM images and EDX spectra of modified CG samples following the adsorption of Cd2⁺ and Pb2⁺: (a) FM-C-HY (Cd), (b) FM-C-HY (Pb), (c) YST-C-HY (Cd), (d) YST-C-HY (Pb), (e) SH-C-HY (Cd), and (f) SH-C-HY (Pb).
Figure 10. SEM images and EDX spectra of modified CG samples following the adsorption of Cd2⁺ and Pb2⁺: (a) FM-C-HY (Cd), (b) FM-C-HY (Pb), (c) YST-C-HY (Cd), (d) YST-C-HY (Pb), (e) SH-C-HY (Cd), and (f) SH-C-HY (Pb).
Processes 12 02095 g010aProcesses 12 02095 g010bProcesses 12 02095 g010c
Figure 11. FTIR spectra of hydrothermally modified CG samples before and after adsorption: (a) FM-C-HY, (b) YST-C-HY, and (c) SH-C-HY.
Figure 11. FTIR spectra of hydrothermally modified CG samples before and after adsorption: (a) FM-C-HY, (b) YST-C-HY, and (c) SH-C-HY.
Processes 12 02095 g011
Table 1. Main chemical composition of the CG samples.
Table 1. Main chemical composition of the CG samples.
SampleSiO2Al2O3Fe2O3K2ONa2OCaOMgOSO3P2O5OthersSiO2/Al2O3
FM66.5920.644.102.131.251.971.990.120.250.965.49
YST64.6421.433.563.590.872.510.920.300.182.005.13
SH56.3434.711.054.151.190.240.550.220.051.502.86
Table 2. Proximate analysis of the CG samples.
Table 2. Proximate analysis of the CG samples.
SampleMadAdVdFCd
FM7.0694.055.780.17
YST2.3986.188.425.40
SH0.9787.917.614.47
Table 3. Kinetics parameters for Cd2+ and Pb2+ adsorption by modified CG samples.
Table 3. Kinetics parameters for Cd2+ and Pb2+ adsorption by modified CG samples.
AdsorptionExperimental qe (mg/g)Pseudo-First-Order ModelPseudo-Second-Order Model
k1R2Calculated qe (mg/g)k2
(×10−4)
R2Calculated qe (mg/g)
FM-C-HY (Cd2+)48.60.00240.87930.93.480.98748.8
YST-C-HY (Cd2+)72.80.00660.94240.76.030.99973.9
SH-C-HY (Cd2+)65.60.00330.87746.32.270.98867.0
FM-C-HY (Pb2+)215.00.00240.931108.41.120.997216.9
YST-C-HY (Pb2+)247.50.00300.83195.51.610.999250.0
SH-C-HY (Pb2+)242.40.00160.836122.80.920.995239.2
Table 4. Pore structure parameters of modified CG samples before and after adsorption.
Table 4. Pore structure parameters of modified CG samples before and after adsorption.
SampleSBET
(m2·g–1) a
Smicro
(m2·g–1) b
Sext
(m2·g–1) c
Vtotal
(cm3·g–1) d
Vmicro
(cm3·g–1) b
Dpore (nm) e
FM-C-HY23.27023.270.07603.399
FM-C-HY (Cd)13.84013.840.06703.403
FM-C-HY (Pb)20.57020.570.07703.400
YST-C-HY20.22020.220.08103.402
YST-C-HY (Cd)12.11012.110.07203.405
YST-C-HY (Pb)9.1109.110.05403.403
SH-C-HY13.01013.010.07303.404
SH-C-HY (Cd)20.312.0418.270.0880.0013.402
SH-C-HY (Pb)14.41014.410.05803.402
a Determined by the multipoint BET method. b Measured by the t-plot method. c Calculated by the difference. d Calculated from the absorbed volume of N2 at a relative pressure P/P0 of 0.99. e Determined by the BJH method using the adsorption branches of N2 isotherms.
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

Chang, Z.; Lu, C.; Bai, L.; Guo, N.; Xing, Z.; Yan, Y. Removal of Cd2+ and Pb2+ from an Aqueous Solution Using Modified Coal Gangue: Characterization, Performance, and Mechanisms. Processes 2024, 12, 2095. https://doi.org/10.3390/pr12102095

AMA Style

Chang Z, Lu C, Bai L, Guo N, Xing Z, Yan Y. Removal of Cd2+ and Pb2+ from an Aqueous Solution Using Modified Coal Gangue: Characterization, Performance, and Mechanisms. Processes. 2024; 12(10):2095. https://doi.org/10.3390/pr12102095

Chicago/Turabian Style

Chang, Zhibing, Chunwei Lu, Lu Bai, Nan Guo, Zhenguo Xing, and Yinuo Yan. 2024. "Removal of Cd2+ and Pb2+ from an Aqueous Solution Using Modified Coal Gangue: Characterization, Performance, and Mechanisms" Processes 12, no. 10: 2095. https://doi.org/10.3390/pr12102095

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