Next Article in Journal
Changes in Water Quality and Soil Property in the Rice–Freshwater Animal Co-Culturing System
Previous Article in Journal
The Spatial and Temporal Dynamics of Soil Conservation and Its Influencing Factors in the Ten Tributaries of the Upper Yellow River, China
Previous Article in Special Issue
Cobalt-Based MOF Material Activates Persulfate to Degrade Residual Ciprofloxacin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanism of Enhanced Fluoride Adsorption Using Amino-Functionalized Aluminum-Based Metal–Organic Frameworks

by
Yiting Luo
1,†,
Zhao Liu
2,†,
Mingqiang Ye
3,
Yihui Zhou
3,
Rongkui Su
4,*,
Shunhong Huang
4,
Yonghua Chen
4 and
Xiangrong Dai
5
1
Hunan First Normal University, Changsha 410114, China
2
Radiation Environmental Supervision Station of Xinjiang Uygur Autonomous Region, Urumqi 830010, China
3
Aerospace Kaitian Environmental Technology Co., Ltd., Changsha 410100, China
4
College of Life and Environmental Science, Central South University of Forestry and Technology, Changsha 410004, China
5
PowerChina Zhongnan Engineering Corporation Limited, Changsha 410004, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work as co-first author.
Water 2024, 16(20), 2889; https://doi.org/10.3390/w16202889
Submission received: 6 September 2024 / Revised: 2 October 2024 / Accepted: 9 October 2024 / Published: 11 October 2024

Abstract

:
Due to the increasing fluoride concentrations in water bodies, significant environmental concerns have arisen. This study focuses on aluminum-based materials with a high affinity for fluorine, specifically enhancing metal–organic frameworks (MOFs) with amino groups to improve their adsorption and defluorination performance. We systematically investigate the factors influencing and mechanisms governing the adsorption and defluorination behavior of amino-functionalized aluminum-based MOF materials in aqueous environments. An SEM, XRD, and FT-IR characterization confirms the successful preparation of NH2-MIL-101 (Al). In a 10 mg/L fluoride ion solution at pH 7.0, fluoride ion removal efficiency increases with the dosage of NH2-MIL-101 (Al), although the marginal improvement decreases beyond 0.015 g/L. Under identical conditions, the fluoride adsorption capacity of NH2-MIL-101 (Al) is seven times greater than that of NH2-MIL-101 (Fe). NH2-MIL-101 (Al) demonstrates effective fluoride ion adsorption across a broad pH range, with superior fluoride uptake in acidic conditions. At a fluoride ion concentration of 7 mg/L, with 0.015 g of NH2-MIL-101 (Al) at pH 3.0, adsorption equilibrium is achieved within 60 min, with a capacity of 31.2 mg/g. An analysis using adsorption isotherm models reveals that the fluoride ion adsorption on NH2-MIL-101 (Al) follows a monolayer adsorption model, while kinetic studies indicate that the predominant adsorption mechanism is chemical adsorption. This research provides a scientific basis for the advanced treatment of fluoride-containing wastewater, offering significant theoretical and practical contributions.

1. Introduction

Fluoride is ubiquitously present in nature, acting both as a toxicological marker for water bodies and as an essential trace element for human health and development [1]. In industrial processes, such as those in the solar cell, phosphate fertilizer, cement production, and thermal power generation industries, the byproducts often include fluoride-containing “three wastes”, which contribute to high concentrations of fluoride in wastewater [2]. Additionally, significant amounts of fluoride are released into the atmosphere during volcanic eruptions, where they subsequently settle into water bodies [3,4]. Fluorinated minerals also release fluoride through surface runoff erosion and natural weathering, leading to its accumulation in soil and water bodies [5,6,7]. Research indicates that the primary route for fluoride entry into the human body is through drinking water [8]. According to China’s drinking water hygiene standards (GB5749-2006) [9], fluoride levels in drinking water should be between 0.5 and 1.0 mg/L. This range is critical because insufficient fluoride intake can lead to dental caries, while excessive fluoride can cause fluorosis and skeletal fluorosis, adversely affecting health [10,11]. Concentrations exceeding 10 mg/L can severely impact the environment and human health, disrupting animal and plant reproduction and potentially causing neurological, immune, and reproductive system issues [7,12]. Long-term excessive fluoride intake can lead to memory loss, cognitive decline, and a weakened immune system, increasing susceptibility to diseases [13,14,15,16,17]. Addressing water pollution [18,19,20,21], particularly fluoride contamination, is a pressing global concern. Given the rising fluoride levels in water bodies, it is crucial to develop effective fluoride removal strategies to protect the ecological balance essential for human survival [22].
Common methods for fluoride removal include chemical precipitation, ion exchange, coagulation sedimentation, and adsorption [23,24,25]. Chemical precipitation involves solid–liquid separation by generating precipitates through chemical reactions. These precipitates often require further treatment; if not managed properly, the pollutants in the precipitates may be released back into the environment, causing secondary pollution [26,27]. Moreover, the efficiency of chemical precipitation is influenced by various operating conditions, such as pH, temperature, and stirring speed. Minor variations in these parameters can significantly affect the formation and separation of precipitates [28]. The ion exchange method is a separation technique based on reversible chemical reactions between active groups on ion exchange materials (e.g., ion exchange resins) and ions in the solution. This method effectively removes impurity ions from water and isolates valuable compounds [29,30,31]. However, it is costly, and its regeneration process typically requires substantial amounts of salts to supply exchange ions, leading to high salt consumption [32]. This consumption becomes particularly pronounced in wastewater with high ion concentrations, thereby increasing treatment costs [33]. Coagulation sedimentation is a widely used combined physical–chemical treatment technology in wastewater management. It operates on the principle that colloids and fine suspended particles aggregate into larger flocs under the influence of coagulants. These flocs then settle under gravity, achieving solid–liquid separation and removing harmful substances. This method is effective and versatile, finding applications in domestic, industrial, and rural sewage treatment. However, there is a risk of secondary pollution if coagulants are improperly selected or excessively used, potentially increasing certain wastewater parameters (e.g., pH, color), which poses a risk of secondary pollution. The adsorption method for fluoride removal utilizes specific adsorbents, such as activated alumina, polyaluminum salts, and lignite adsorbents. These materials possess porous structures and large specific surface areas, enabling effective fluoride ion adsorption [34]. When fluoride containing wastewater passes through these adsorbents, fluoride ions are adsorbed on the surface of the adsorbents to promote their removal [35]. The key to improving this method is identifying an adsorbent that offers high effectiveness at low cost [36,37]. An ideal fluoride adsorbent should feature rapid adsorption kinetics, cost-effectiveness, recyclability, high adsorption capacity, and minimal reactivity with other substances [38]. High-performance adsorbents typically have significant porosity and a specific surface area or possess chemical groups that readily react with pollutants [39,40].
The adsorption method distinguishes itself among various water purification technologies due to its superior defluorination performance, low cost, and straightforward manufacturing process, making it the prevalent choice for defluorination at present [41,42]. This method is not only efficient but also cost-effective and highly practical [43,44,45]. Metal–organic frameworks (MOFs) are a novel class of nanomaterials, typically formed through the self-assembly of metal ions coordinated with organic ligands [46,47,48,49]. MOFs are characterized by their high porosity and robust chemical stability, rendering them suitable for applications in catalysis [50,51,52], adsorption and separation [53], optical materials [54], and magnetic materials. These materials exhibit a significant adsorption capacity for specific organic pollutants in aqueous environments, thus offering promising advancements in water purification [55,56]. As a new type of functional molecular material, MOFs have the following advantages compared to activated carbon and zeolite materials: a large specific surface area, adjustable pore size, an ordered microporous structure, diverse pore size and skeleton structure, modifiable functional groups on the pore surface, and unsaturated metal coordination sites [57,58,59]. MOFs can be categorized into various types, such as Zeolitic Imidazolate Frameworks (ZIFs), Materials of Institut Lavoisier (MILs), and University of Oslo (UIO) series, based on their metal ions and organic ligands [60]. Current synthesis methods for MOFs include microwave-assisted synthesis, ultrasonic methods, liquid-phase diffusion, solvothermal synthesis, and mechanical stirring [61]. To enhance the porosity and performance of MOFs, several activation techniques are employed, including vacuum heat treatment, solvent exchange, supercritical carbon dioxide treatment (scCO2), freeze-drying, ultrasonic treatment, and chemical activation [62]. These activation methods effectively improve MOF porosity, thereby optimizing their performance in catalysis and adsorption applications [53].
This study successfully synthesized NH2-MIL-101 (Al) material using a solvothermal method and confirmed its synthesis quality through comprehensive material characterization. The fluoride ion removal efficiency and mechanism of this material were assessed using adsorption experiments. A series of controlled comparative experiments were conducted to explore the impacts of varying dosages, initial fluoride ion concentrations, pH levels, reaction times, and temperatures on the fluoride ion removal efficiency of NH2-MIL-101 (Al). The optimal dosage and conditions for fluoride ion adsorption by NH2-MIL-101 (Al) were determined, providing a robust theoretical and experimental foundation for its practical applications.

2. Materials and Methods

2.1. Instruments and Experimental Reagents

Sodium fluoride (NaF, analytical grade), 2-aminoterephthalic acid (C8H7NO4, 98%), and acetic acid (CH3COOH, 99.5%) were obtained from Shanghai McLean Biochemical Technology Co., Ltd. (Shanghai, China). N,N-Dimethylformamide (C3H7NO, analytical grade) was sourced from Taicang Hujian Reagent Co., Ltd. (Taicang, China). Sodium chloride (NaCl, superior grade), aluminum chloride hexahydrate (AlCl3·6H2O, analytical grade), methanol (CH4O, analytical grade), sodium hydroxide (NaOH, analytical grade), trisodium citrate (C6H5Na3O7, analytical grade), and hydrochloric acid (HCl, analytical grade) were purchased from China National Pharmaceutical Group Chemical Reagent Co., Ltd. (Beijing, China).
A digital pH meter (PHS-3E, Shanghai Yidian Scientific Instrument Co., Ltd., Shanghai, China), high-speed desktop centrifuge (TG16-WS, Hunan Xiangyi Centrifuge Instrument Co., Ltd., Changsha, China), electronic balance (FA2004B, Shanghai Youke Instrument Co., Ltd., Shanghai, China), constant-temperature drying oven (202 type, Beijing Yongguangming Medical Instrument Co., Ltd., Beijing, China), rotary vane vacuum pump (2XZ type, Jiaojiang Shuang’e Vacuum Equipment Factory, Taizhou, China), ultrasonic cleaning machine (820HT, Changsha Mingjie Instrument Co., Ltd., Changsha, China), constant-temperature oscillator (XMTE-205, Changzhou Guowang Instrument Co., Ltd., Changzhou, China), and deionized water preparation unit (LAN-19-H, Lide Water Treatment Equipment Co., Ltd., Chongqing, China) were obtained for this study.

2.2. Preparation of NH2-MIL-101 (Al)

NH2-MIL-101 (Al) was synthesized using a solvothermal method [63,64]. Precisely weigh 0.51 g of AlCl3·6H2O and 0.56 g of 2-aminoterephthalic acid using an analytical balance. Dissolve these reagents in 30 mL of N,N-dimethylformamide. Subject the resulting solution to ultrasonic agitation for 30 min to ensure thorough mixing. Transfer the homogenized solution into a 50 mL reaction vessel, seal it tightly, and perform the reaction at 130 °C for 72 h. Allow the reaction mixture to cool to room temperature naturally. After cooling, separate the product by centrifugation at 11,000 rpm to obtain a yellow precipitate. Wash the precipitate with N,N-dimethylformamide to remove the supernatant, then perform three additional washes with anhydrous methanol. Finally, place the washed precipitate in a vacuum drying oven at 60 °C and dry overnight to obtain NH2-MIL-101 (Al) as a light-yellow powder. The preparation method of NH2-MIL-101 (Fe) is based on our previous report [65].

2.3. Characterization Methods for NH2-MIL-101 (Al)

(1)
X-ray Diffraction Analysis (XRD)
To investigate the crystal structure and phase composition of NH2-MIL-101 (Al), an X-ray diffraction analysis was performed using a MiniFlex600 X-ray powder diffractometer (Rigaku Co., Ltd., Tokyo, Japan). Copper Kα radiation was utilized as the X-ray source to ensure high precision in the measurements. The diffraction measurements were conducted at a current of 40 mA and a voltage of 40 kV to maintain a stable instrument operation. The data were collected with a scanning speed of 10°/min over a 2θ range of 5° to 90° to capture comprehensive diffraction information of the material.
(2)
Fourier Transform Infrared Spectroscopy (FT-IR)
Fourier Transform Infrared Spectroscopy (FT-IR) was employed to analyze the chemical composition, functional groups, and intermolecular interactions of NH2-MIL-101 (Al). This technique utilizes the absorption and transmission characteristics of infrared light to generate spectra that reveal detailed information about chemical bonds and functional groups within the material. The FT-IR spectra were obtained using a Nicolet iS50 spectrometer (Thermo Fisher Scientific, Inc., Waltham, MA, USA), with samples prepared using the KBr pellet method, covering a wavelength range of 400 to 4000 cm−1. These spectral data provide valuable insights into the chemical properties of NH2-MIL-101 (Al).
(3)
Scanning Electron Microscopy (SEM)
To examine the surface microstructure, elemental distribution, and particle size of NH2-MIL-101 (Al), a JSM-7610FPlus scanning electron microscope (JEOL Co., Ltd., Tokyo, Japan) was employed. This high-resolution microscopy technique allowed for detailed observation of the fine surface structure and precise measurement of particle size.
(4)
Specific Surface Area and Pore Size Analysis (BET)
The specific surface area and pore structure of NH2-MIL-101 (Al) were analyzed using the Brunauer–Emmett–Teller (BET) method. A JWGB JY-BK 112 specific surface area analyzer (Beijing JWGB Sci.& Tech. Co., Ltd., Beijing, China) was utilized to determine key parameters such as the specific surface area, pore volume, and pore size of the sample. This analysis not only enhances the accuracy of the research but also provides a deeper understanding of the microstructural characteristics of NH2-MIL-101 (Al).

2.4. Adsorption Experiment and Analysis

2.4.1. Adsorption Experiment

Fluoride ion concentrations were determined using the standard curve method, employing a pH meter equipped with a fluoride ion selective electrode [66]. The electromotive force (E), measured in millivolts by the pH meter, decreases as the fluoride ion concentration in the solution increases. Before each experimental batch, it is essential to verify the electrode’s actual slope and re-construct the standard curve. To assess the impact of the material dosage and initial fluoride ion concentration on the fluoride ion removal efficiency of NH2-MIL-101 (Al), the following experiments were conducted:
Material Dosage Effect: Add 0.005 g, 0.010 g, 0.015 g, and 0.020 g NH2-MIL-101 (Al) to 150 mL of a 10 mg/L fluoride ion solution, adjusting the pH to 7.0–7.5. Then, keep the reaction solution in a constant temperature stirrer at 25 °C for 120 min. After the reaction, filter the solution using a 0.22 µm organic filter, measure the equilibrium concentration, and calculate the equilibrium adsorption capacity ( q e ).
Initial Concentration Effect: Prepare fluoride ion solutions at concentrations of 1, 2, 3, 4, 5, 6, 7, 8, 9, and 10 mg/L. Add 150 mL of each fluoride ion solution to separate beakers and introduce 0.015 g of NH2-MIL-101 (Al). React for 120 min, then filter the solutions using a 0.22 µm organic filter. Measure the equilibrium concentrations and determine the equilibrium adsorption capacity ( q e ).
To evaluate the influence of solution pH and reaction time on the fluoride ion removal performance of NH2-MIL-101 (Al), the following experiments were performed:
Reaction Time Effect: Prepare 150 mL of a 7 mg/L fluoride ion solution with the pH adjusted to 7, and add 0.015 g of NH2-MIL-101 (Al). React at 25 °C in a constant temperature mixer, and filter the sample solution every 10 min. Measure the equilibrium concentration after each time point and calculate the adsorption capacity ( q e ).
pH Effect: Prepare 150 mL of a 7 mg/L fluoride ion solution and sequentially adjust the pH to 1, 3, 5, 7, and 9. Add 0.015 g of NH2-MIL-101 (Al) to each solution and react at 25 °C for 60 min in a constant temperature stirrer. After the reaction, filter the solution using a 0.22 µm organic filter, measure the equilibrium concentrations, and calculate the equilibrium adsorption capacity ( q e ).

2.4.2. Adsorption Kinetics

Two kinetic models—the pseudo-first-order kinetic model and the pseudo-second-order kinetic model—were utilized to analyze and describe the interaction between the adsorbent and the adsorbate under specific temperature conditions [67,68].
(1)
Pseudo-First-Order Kinetic Model
The pseudo-first-order rate equation, based on solid adsorption capacity, is frequently employed. If kinetic data fit the pseudo-first-order model, it indicates that diffusion is the predominant mechanism in the adsorption process. The linear equation for this model is given by the following:
ln ( q e q t ) = ln q e k 1 t
where t is the adsorption time (minutes), qt is the adsorption capacity at time t (mg/g), qe is the adsorption capacity at equilibrium (mg/g), and k1 is the pseudo-first-order rate constant (min−1) [41].
(2)
Pseudo-Second-Order Kinetic Model
Assuming that the adsorption rate is governed primarily by a chemical adsorption mechanism, the pseudo-second-order kinetic model can be applied. This model suggests that the adsorption involves electron transfer between the adsorbent and the adsorbate. If the kinetic data align well with the pseudo-second-order model, it implies that the adsorption process is mainly controlled by chemical interactions. The linear equation is expressed as the following:
t q t = 1 k 2 q e 2 + t q e
where t is the adsorption time (min), qt is the adsorption capacity at time t (mg/g), qe is the adsorption capacity at equilibrium (mg/g), and k2 is the pseudo-second-order rate constant (g/(mg·min)).

2.4.3. Adsorption Isotherms

Two widely accepted adsorption models—the Langmuir adsorption isotherm model and the Freundlich adsorption isotherm model—are employed for detailed simulation and analysis [69,70].
(1)
Langmuir Adsorption Isotherm Model
The Langmuir adsorption isotherm model presumes a constant number of adsorption sites on the adsorbent, each with identical adsorption potential. Once a solute molecule occupies an adsorption site, it remains fixed and does not migrate to other sites, with no mutual influence between adsorbed molecules. This model provides a theoretical foundation for understanding the isothermal adsorption process. The linear form of this model is as follows:
1 q e = 1 K L q m a x 1 C e + 1 q m a x
where qe is the equilibrium adsorption capacity (mg/g), KL is the adsorption constant, qmax is the theoretical maximum adsorption capacity (mg/g), and Ce is the equilibrium concentration (mg/L).
The adsorption constant (KL) reflects the adsorption difficulty: a KL value of 0 indicates an irreversible reaction; KL between 0 and 1 signifies a feasible reaction; KL equal to 1 represents linear adsorption; and KL greater than 1 suggests a more challenging reaction.
(2)
Freundlich Adsorption Isotherm Model
The Freundlich adsorption isotherm model describes adsorption as a non-uniform distribution across multiple molecular layers. This model indicates that adsorption is not evenly distributed but occurs in a multi-layer structure with varying adsorption capacities at different sites. The model suggests that as the adsorbate concentration increases, so does the adsorption, highlighting the site-specific adsorption capacities and the complexity of multi-layer adsorption. The linear form is the following:
l n q e = 1 n l n C e + l n K F
where qe is the equilibrium adsorption capacity (mg/g), and KF is the adsorption constant.
For plotting isothermal adsorption data, lnqe is typically plotted against lnCe. A straight line is fitted to the experimental data, from which the Freundlich parameters KF and n are derived. KF is directly related to adsorption capacity, a higher KF indicates stronger adsorption. The n value represents the adsorption intensity; n > 1 usually indicates that the adsorption process is relatively favorable.

3. Results and Discussion

3.1. Characterization Analysis of NH2-MIL-101 (Al)

3.1.1. X-ray Diffraction Analysis (XRD)

Figure 1 presents the XRD pattern of NH2-MIL-101 (Al). The data reveal that the diffraction peaks exhibit high intensity. A comparison with the literature data indicate that the characteristic diffraction peaks of the material exhibit high diffraction intensity and align with the peaks documented in the literature [71,72]. The absence of impurity peaks confirms the successful synthesis of NH2-MIL-101 (Al). These results suggest that the synthesized NH2-MIL-101 (Al) material possesses a crystal phase structure consistent with that described in the literature, thereby validating its identity as NH2-MIL-101 (Al).

3.1.2. Fourier Transform Infrared Spectroscopy (FT-IR)

The chemical bonds in NH2-MIL-101 (Al) were analyzed using Fourier Transform Infrared Spectroscopy (FT-IR), as illustrated in Figure 2. According to reference [43], the absorption peak at 3491.9 cm−1 corresponds to the asymmetric and symmetric stretching vibrations of the -NH2 group in the organic ligand. The characteristic peaks at 1569.1 cm−1 and 1442.1 cm−1 are attributed to the N-H bending vibration and C-H stretching vibrations in aromatic rings, respectively. These results confirm the successful synthesis of the NH2-MIL-101 (Al) material.

3.1.3. Scanning Electron Microscopy (SEM)

Figure 3 reveals that the NH2-MIL-101 (Al) material exhibits a tightly packed octahedral grain arrangement, forming clusters with a rough surface, which enhances adsorption capacity. This observation is consistent with the findings reported in the literature [73].

3.1.4. Specific Surface Area and Pore Size Analysis (BET)

The specific surface area and pore size distribution of NH2-MIL-101 (Al) are presented in Figure 4. Figure 4a shows the N2 adsorption isotherm at 77 K. The low-pressure region indicates a low adsorption capacity for N2, while the capacity significantly increases as the relative pressure approaches 1.0, exhibiting a Class II adsorption isotherm with an overall S-shape [42]. Figure 4b illustrates the pore size distribution of NH2-MIL-101 (Al), with a concentration around 20 nm and an average pore size of 17.19 nm. The material is categorized as mesoporous, with an average pore volume of 0.28 cm3/g and a BET specific surface area of 676.94 m2/g.

3.2. Effect of Material Dosage on Adsorption Performance

Figure 5 illustrates the impact of different dosages on fluoride ion removal by NH2-MIL-101 (Al). Increasing the dosage enhances the adsorption capacity of NH2-MIL-101 (Al) for fluoride ions. At dosages of 0.005 g and 0.010 g, the equilibrium adsorption capacities are 25 mg/g and 30 mg/g, respectively. Compared to 0.005 g and 0.010 g, he adsorption capacity increases to 36.6 mg/g at 0.015 g, the adsorption capacity increases to 36.6 mg/g, and the removal rate rising from 34% and 40% to 48%, respectively. At a dosage of 0.020 g, the equilibrium adsorption capacity reaches 38.4 mg/g, and the removal rate improves from 48% to 51%. This change indicates that beyond 0.015 g, further increases in dosage do not significantly enhance adsorption performance. With the increase in adsorbent dosage (under the condition of constant fluoride solution concentration and volume), the removal rate increases, which may be due to the increase in the number of active sites on the MOF and the increase in surface area in contact with the solution. However, as the amount of MOFs increases, the increase in removal rate is not significant, because the number of remaining adsorption sites increases during the removal process [63]. Subsequent experiments were therefore conducted with 0.015 g of NH2-MIL-101 (Al).
Under identical conditions, a 120 min reaction time achieved equilibrium adsorption of fluoride. Figure 5 shows that NH2-MIL-101 (Al) exhibits a markedly superior adsorption capacity for fluoride compared to NH2-MIL-101 (Fe). Specifically, NH2-MIL-101 (Al) achieves an adsorption capacity of 36.6 mg/g, whereas NH2-MIL-101 (Fe) only reaches 5.9 mg/g. According to reference [74], this disparity is attributed to NH2-MIL-101 (Al)’s mesoporous structure, larger specific surface area, and abundant functional groups, which confer a substantial advantage in fluoride ion adsorption over NH2-MIL-101 (Fe) and enhance the interaction between NH2-MIL-101 (Al) and fluoride ions.

3.3. Effect of Initial Fluoride Solution Concentration on the Defluorination Efficiency of NH2-MIL-101 (Al)

Figure 6 depicts the relationship between the equilibrium adsorption capacity (qe) of NH2-MIL-101 (Al) and the initial concentration (C0) of fluoride ion solution. The data demonstrate that as the fluoride ion concentration increases, the adsorption capacity of NH2-MIL-101 (Al) also rises until it reaches an equilibrium state where further increases in concentration no longer affect the adsorption capacity.
The underlying mechanism can be explained as follows [75]: at lower fluoride ion concentrations, the number of available adsorption sites on NH2-MIL-101 (Al) significantly exceeds the number of free fluoride ions in the solution, resulting in incomplete adsorption equilibrium. As the concentration of fluoride ions increases, the number of free fluoride ions also increases, providing more ions for adsorption onto the NH2-MIL-101 (Al) crystals. This increase enhances the adsorption capacity by facilitating the movement of fluoride ions to the available adsorption sites. However, when the initial fluoride ion concentration reaches 7.0 mg/L, the adsorption capacity begins to stabilize. This stabilization is likely due to the saturation of adsorption sites on NH2-MIL-101 (Al), where additional fluoride ions can no longer be absorbed effectively.
In summary, when the fluoride ion concentration is below 7.0 mg/L, increasing the initial concentration enhances the adsorption performance of NH2-MIL-101 (Al). Once the concentration reaches 7.0 mg/L, the adsorption capacity remains relatively constant. As shown in Figure 6, at a fluoride ion concentration of 7.0 mg/L, the equilibrium adsorption capacity is 35.1 mg/g and does not increase further with higher initial fluoride ion concentrations. Therefore, subsequent experiments will utilize a fluoride ion solution with a concentration of 7 mg/L.

3.4. Effect of Solution pH on Fluorination Removal of NH2-MIL-101 (Al)

From Figure 7, it can be seen that the initial solution pH has little effect on the defluorination of the NH2-MIL-101 (Al) adsorbent, indicating that NH2-MIL-101 (Al) still has an excellent defluorination performance in a larger pH range, and its adsorption performance is even better under acidic conditions. Under acidic and neutral conditions, the adsorption capacity of NH2-MIL-101 (Al) can reach over 25 mg/g, and when the pH reaches 9.0, the adsorption capacity begins to decrease to below 20 mg/g. The possible mechanism is the consumption of hydroxyl ions by hydrogen ions in water, further promoting the interaction between fluoride ions and NH2-MIL-101 (Al) under acidic conditions [73]. Similarly, there is a strong competitive relationship between fluoride ions and hydroxyl ions under alkaline conditions, which leads to the inability of NH2-MIL-101 (Al) to fully adsorb free fluoride ions in water, resulting in weaker adsorption performance. In order to make the adsorption effect more obvious in subsequent experiments, the pH was adjusted to three before conducting further experiments.

3.5. Effect of Reaction Time on Defluorination Using NH2-MIL-101 (Al)

Determining the optimal adsorption time is essential for effective wastewater treatment using NH2-MIL-101 (Al). A slower adsorption rate extends the time required to achieve equilibrium, which may be suboptimal for practical applications in wastewater treatment. Figure 8 illustrates the relationship between adsorption time (t) and the amount of fluoride ions adsorbed (qt) by NH2-MIL-101 (Al). The data indicate that qt increases progressively with an adsorption time up to 60 min. Beyond this period, the adsorption capacity stabilizes with minimal variation, suggesting that NH2-MIL-101 (Al) has reached its adsorption saturation point for fluoride ions. Based on repeated experiments, this saturation level is considered the equilibrium adsorption capacity, with an average qt of 31.2 mg/g. Compared with previous studies on metal–organic frameworks, the MOF synthesized in this study has a higher qe and a shorter time required to reach equilibrium [76,77,78,79]. Compared to other adsorbents investigated for F removal, NH2-MIL-101(Al) adsorbs F in a shorter period of time, confirming the appropriate nature of the studied MOF for the purification of F-polluted water [63,79,80].

3.6. Adsorption Kinetics Analysis

To further elucidate the adsorption kinetics of fluoride ions on NH2-MIL-101 (Al), both pseudo-first-order and pseudo-second-order kinetic models were applied to analyze the adsorption data.

3.6.1. Pseudo-First-Order Kinetic Model

The pseudo-first-order kinetic model was evaluated by plotting the adsorption time (t) on the x-axis against ln(qeqt) on the y-axis, as shown in Table 1. The resulting linear equation enabled the determination of the model parameters. By calculating the slope and intercept, the equilibrium adsorption capacity (qe) and rate constant (k1) were obtained. An analysis of Figure 9 and Table 1 reveals that the regression coefficient (R2) for the pseudo-first-order model was 0.95278.

3.6.2. Pseudo-Second-Order Kinetic Model

For the pseudo-second-order kinetic model, the adsorption time (t) was plotted against t/qt, as presented in Table 2. An analysis of Figure 10 and Table 2 shows that the experimental data aligned closely with the fitted line, yielding a high regression coefficient (R2) of 0.99727.
Considering the R2 values of the kinetic equations and comparing the calculated and experimental values of equilibrium capacities, the pseudo-second-order kinetic model is more consistent with the obtained experimental results. Therefore, NH2-MIL-101 (Al) exhibits strong binding chemical adsorption towards fluoride ions [81]. Studies showed that the adsorption of fluoride ions by Al-MOF is achieved by replacing Al−OH on Al atoms to form a stable Al−F structure [82].

3.7. Adsorption Isotherm Analysis

3.7.1. Langmuir Model

To analyze the Langmuir model adsorption isotherm, plot “1/qe” on the horizontal axis and “1/Ce” on the vertical axis for three different temperatures, as illustrated in Figure 11. Linear regression of the data yields the fitting equations and parameters listed in Table 3.
The data presented in Table 3 indicate that the fluoride ion adsorption by NH2-MIL-101 (Al) is significantly influenced by temperature. As temperature increases, the adsorption capacity for fluoride ions also rises. Furthermore, the R2 values remain above 0.98 across different temperatures, suggesting that the adsorption process conforms well to the Langmuir model. This trend implies that the underlying mechanism for fluoride ion adsorption by NH2-MIL-101 (Al) is primarily single-molecule adsorption.

3.7.2. Freundlich Model

For the Freundlich model adsorption isotherms, plot “ln(Ce)” on the horizontal axis and “ln(qe)” on the vertical axis for three different temperatures, as shown in Figure 12. The fitting results and corresponding parameters are summarized in Table 4.
Figure 12 shows that the Freundlich model fits the NH2-MIL-101 (Al) fluoride ion adsorption data less effectively compared to the Langmuir model. The n values, which are all below 0.7, suggest that the adsorption process is not easily accomplished. Additionally, the linear fitting coefficients are consistently below 0.98, indicating a poorer fit compared to the Langmuir model.
Comparing the results from the two models, it is evident that the Langmuir model provides a superior fit for fluoride ion adsorption by NH2-MIL-101 (Al) compared to the Freundlich model. This result confirms that single-molecule adsorption is the dominant mechanism in this process [83,84].

4. Conclusions

This study successfully synthesized NH2-MIL-101 (Al) material using a solvothermal method and systematically investigated its effectiveness in removing fluoride ions from the solution, along with the underlying mechanisms and influencing factors. Characterization via SEM, XRD, and FT-IR confirms the successful preparation of the NH2-MIL-101 (Al) material. In a solution containing 10 mg/L fluoride ions at pH 7, fluoride removal efficiency improves with an increasing NH2-MIL-101 (Al) dosage; however, the marginal benefit decreases when the dosage exceeds 0.015 mg/g. Under identical conditions, the fluoride adsorption capacity of NH2-MIL-101 (Al) is seven times greater than that of NH2-MIL-101 (Fe). NH2-MIL-101 (Al) demonstrates effective fluoride ion adsorption across a broad pH range, with enhanced adsorption in acidic conditions, outperforming NH2-MIL-101 (Fe). When the fluoride ion concentration is 7 mg/L, and 0.015 g of NH2-MIL-101 (Al) is used at pH 3.00, the material achieves adsorption equilibrium within 60 min, with an equilibrium capacity of 31.2 mg/g. The analysis using adsorption isotherm models indicates that fluoride ion adsorption onto NH2-MIL-101 (Al) follows a monolayer adsorption pattern. The kinetic analysis reveals that the predominant mechanism of fluoride ion adsorption is chemical adsorption. The synthesis of NH2-MIL-101 (Al) has good scalability in industrial scale water treatment applications. However, cost and reuse are the main challenges and considerations for scaling up the process. This research provides a scientific basis for the advanced treatment of fluoride-containing wastewater, offering significant theoretical and practical implications.

Author Contributions

Conceptualization, Y.L. and R.S.; methodology, Y.L. and R.S.; formal analysis, Z.L., Y.Z. and R.S.; investigation, M.Y. and R.S.; resources, R.S.; data curation, S.H. and R.S.; writing—original draft, Y.L., Z.L. and R.S.; writing—review and editing, M.Y., X.D., Y.C., Y.Z. and S.H.; supervision, X.D. and Y.C.; project administration, R.S.; funding acquisition, R.S. and Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Major Program Natural Science Foundation of Hunan Province of China (2021JC0001), Hunan Provincial Natural Science Foundation of China (2023JJ31010, 2024JJ7647, 2024JJ7094), Key Project of Scientific Research Project of Hunan Provincial Department of Education (23A0225).

Data Availability Statement

The data are contained within this article.

Acknowledgments

The authors thank all the participants who devoted their free time to participate in this study. This research was also funded by National Nature Science Foundation of China (52000183), Hunan Province Environmental Protection Research Project (HBKYXM-2023038), Research project funded by the Education Department of Hunan Province (22B0883), National Local Joint Engineering Research Center of Heavy Metals Pollutants Control and Resource Utilization Open Fund (ES202380056), Scientific Research Foundation for Talented Scholars of CSUFT (2020YJ010).

Conflicts of Interest

Authors Mingqiang Ye and Yihui Zhou were employed by the Aerospace KaiTian Environmental Technology Co., Ltd. Author Xiangrong Dai was employed by the company PowerChina Zhongnan Engineering Corporation Limited. The remaining authors declare that this research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Kabir, H.; Gupta, A.K.; Tripathy, S. Fluoride and human health: Systematic appraisal of sources, exposures, metabolism, and toxicity. Crit. Rev. Environ. Sci. Technol. 2020, 50, 1116–1193. [Google Scholar] [CrossRef]
  2. Ghosh, A.; Mukherjee, K.; Ghosh, S.K.; Saha, B. Sources and toxicity of fluoride in the environment. Res. Chem. Intermed. 2013, 39, 2881–2915. [Google Scholar] [CrossRef]
  3. Liang, J.; Xu, B.; Zhang, Y.; Li, K.; Du, J. Removal Performance of Fluoride Ions from Drinking Water Using Metal Salt-Modified Activated Magnesium Oxide. J. Environ. Eng. 2023, 17, 2181–2191. [Google Scholar] [CrossRef]
  4. Liu, X.; Yang, Y.; Wang, P.; Liu, T. Advances in Functional Fluoride Removal Adsorbents for Drinking Water and Fluoride-Containing Wastewater. Environ. Chem. 2024, 43, 1–14. [Google Scholar] [CrossRef]
  5. Chang, H.Y.H.; Kuo, Y.L.; Liu, J.C. Fluoride at waste oyster shell surfaces—Role of magnesium. Sci. Total Environ. 2019, 652, 1331–1338. [Google Scholar] [CrossRef]
  6. Wambu, E.W.; Ambusso, W.O.; Onindo, C.; Muthakia, G.K. Review of fluoride removal from water by adsorption using soil adsorbents—An evaluation of the status. J. Water Reuse Desalination 2016, 6, 1–29. [Google Scholar] [CrossRef]
  7. Singh, K.; Lataye, D.H.; Wasewar, K.L. Removal of fluoride from aqueous solution by using bael (Aegle marmelos) shell activated carbon: Kinetic, equilibrium and thermodynamic study. J. Fluor. Chem. 2017, 194, 23–32. [Google Scholar] [CrossRef]
  8. Kheradpisheh, Z.; Mirzaei, M.; Mahvi, A.H.; Mokhtari, M.; Azizi, R.; Fallahzadeh, H.; Ehrampoush, M.H. Impact of drinking water fluoride on human thyroid hormones: A case-control study. Sci. Rep. 2018, 8, 2674. [Google Scholar] [CrossRef]
  9. GB 5749-2006; Sanitary Standards for Drinking Water. Ministry of Health of the People’s Republic of China and National Standardization Administration of China: Beijing, China, 2006.
  10. Srivastava, S.; Flora, S. Fluoride in drinking water and skeletal fluorosis: A review of the global impact. Curr. Environ. Health Rep. 2020, 7, 140–146. [Google Scholar] [CrossRef]
  11. Bhasin, C.P.; Pathan, A.; Patel, R.V. An Evaluation of Carbon Nanotube-based and Activated Carbon-based Nanocomposites for Fluoride and Other Pollutant Removal from Water: A Review. Curr. Mol. Med. 2024, 9, 16–40. [Google Scholar] [CrossRef]
  12. Tiwari, K.K.; Raghav, R.; Pandey, R. Recent advancements in fluoride impact on human health: A critical review. Environ. Sustain. Indic. 2023, 20, 100305. [Google Scholar] [CrossRef]
  13. Hussain, J.; Hussain, I.; Sharma, K. Fluoride and health hazards: Community perception in a fluorotic area of central Rajasthan (India): An arid environment. Environ. Monit. Assess. 2010, 162, 1–14. [Google Scholar] [CrossRef] [PubMed]
  14. Ali, S.; Thakur, S.K.; Sarkar, A.; Shekhar, S. Worldwide contamination of water by fluoride. Environ. Chem. Lett. 2016, 14, 291–315. [Google Scholar] [CrossRef]
  15. Ghosh, G.; Mukhopadhyay, D.K. Human health hazards due to arsenic and fluoride contamination in drinking water and food chain. In Groundwater Development and Management; Springer: Cham, Switzerland, 2019; pp. 351–369. [Google Scholar]
  16. Sunkari, E.D.; Adams, S.J.; Okyere, M.B.; Bhattacharya, P. Groundwater fluoride contamination in Ghana and the associated human health risks: Any sustainable mitigation measures to curtail the long term hazards? Groundw. Sustain. Dev. 2022, 16, 100715. [Google Scholar] [CrossRef]
  17. Mutai, I.K.; Kiriamiti, H.K.; M’Arimi, M.; Tewo, R.K. Fluoride removal from water using Al(OH)3-surface modified diatomite mixed with brick: Optimization, isotherm and kinetic studies. Aims Environ. Sci. 2024, 11. [Google Scholar] [CrossRef]
  18. Su, R.; Chai, L.; Tang, C.; Li, B.; Yang, Z. Comparison of the degradation of molecular and ionic ibuprofen in a UV/H2O2 system. Water Sci. Technol. 2018, 77, 2174–2183. [Google Scholar] [CrossRef] [PubMed]
  19. Luo, Y.; Su, R.; Yao, H.; Zhang, A.; Xiang, S.; Huang, L. Degradation of trimethoprim by sulfate radical-based advanced oxidation processes: Kinetics, mechanisms, and effects of natural water matrices. Environ. Sci. Pollut. Res. 2021, 28, 62572–62582. [Google Scholar] [CrossRef] [PubMed]
  20. Su, R.; Dai, X.; Wang, H.; Wang, Z.; Li, Z.; Chen, Y.; Luo, Y.; Ouyang, D. Metronidazole degradation by UV and UV/H2O2 advanced oxidation processes: Kinetics, mechanisms, and effects of natural water matrices. Int. J. Environ. Res. Public Health 2022, 19, 12354. [Google Scholar] [CrossRef]
  21. Luo, Y.; Su, R. Environmental impact of waste treatment and synchronous hydrogen production: Based on life cycle assessment method. Toxics 2024, 12, 652. [Google Scholar] [CrossRef]
  22. Podgorski, J.; Berg, M. Global analysis and prediction of fluoride in groundwater. Nat. Commun. 2022, 13, 4232. [Google Scholar] [CrossRef]
  23. Arab, N.; Derakhshani, R.; Sayadi, M.H. Approaches for the Efficient Removal of Fluoride from Groundwater: A Comprehensive Review. Toxics 2024, 12, 306. [Google Scholar] [CrossRef] [PubMed]
  24. Sun, Y.J.; Zhang, C.; Ma, J.Q.; Sun, W.Q.; Shah, K.J. Review of fluoride removal technology from wastewater environment. Desalination Water Treat. 2023, 299, 90–101. [Google Scholar] [CrossRef]
  25. Su, R.; Xue, R.; Ma, X.; Zeng, Z.; Li, L.; Wang, S. Targeted improvement of narrow micropores in porous carbon for enhancing trace benzene vapor removal: Revealing the adsorption mechanism via experimental and molecular simulation. J. Colloid. Interf. Sci. 2024, 671, 770–778. [Google Scholar] [CrossRef]
  26. Wang, T.Y.; Zhang, Y.; Qi, J.; Hu, C.Z.; Qu, J.H. Sulfate Doping Promotes Agglomeration of Calcium Fluoride Crystals. Environ. Sci. Technol. 2024, 58, 4450–4458. [Google Scholar] [CrossRef] [PubMed]
  27. Su, R.; Zhang, H.; Chen, F.; Wang, Z.; Huang, L. Applications of single atom catalysts for environmental management. Int. J. Environ. Res. Public Health 2022, 19, 11155. [Google Scholar] [CrossRef] [PubMed]
  28. Feng, S.Y.; Zhang, W.J.; Che, J.Y.; Wang, C.Y.; Chen, Y.Q. Controllable and selective fluoride precipitation from phosphate-rich wastewater. Sci. Total Environ. 2024, 951, 175507. [Google Scholar] [CrossRef] [PubMed]
  29. Jia, B.L.; Li, Y.; Guo, J.P.; Zhou, F.; Guo, Y.C.; Yang, Y.H. Nanometer lanthanum hydroxide modified anionic polymer for deep fluoride removal: Adsorption properties and mechanism study. Prog. Nat. Sci. Mater. Int. 2023, 33, 833–842. [Google Scholar] [CrossRef]
  30. Tipplook, M.; Tanaka, H.; Sudare, T.; Hagio, T.; Saito, N.; Teshima, K. Nanoarchitectonics Solution Plasma Polymerization of Amino-Rich Carbon Nanosorbents for Use in Enhanced Fluoride Removal. ACS Appl. Mater. Interfaces 2024, 16, 7038–7046. [Google Scholar] [CrossRef]
  31. Rathi, B.S.; Kumar, P.S.; Rangasamy, G.; Badawi, M.; Aminabhavi, T.M. Membrane-based removal of fluoride from groundwater. Chem. Eng. J. 2024, 488, 150880. [Google Scholar] [CrossRef]
  32. Wang, S.; Wang, Z.Y.; Zhang, Y.F.; Li, C.J.; Chen, W.L.; Fang, H.P.; Huang, F.; Zhang, Y.L.; Pan, L.; Zheng, Y.C.; et al. Ionic liquid-functionalized transition-metal oxides as highly recyclable and efficient nano-adsorbent for fluoride removal. Sep. Purif. Technol. 2024, 351, 128003. [Google Scholar] [CrossRef]
  33. Zhu, A.Q.; Wang, Z.R.; Li, P.C.; Zhou, X.J.; Shen, S.S. Preparation of a positive porous polyvinylidene fluoride membrane and its removal of mercury containing wastewater. Desalination Water Treat. 2023, 287, 59–66. [Google Scholar] [CrossRef]
  34. Lobo, C.C.; Lerner, J.E.C.; Bertola, N.C.; Zaritzky, N.E. Synthesis and characterization of functional calcium-phosphate-chitosan adsorbents for fluoride removal from water. Int. J. Biol. Macromol. 2024, 264, 130553. [Google Scholar] [CrossRef] [PubMed]
  35. Li, Y.X.; Xiao, Y.J.; Lu, T.X.; He, G.C.; Ding, Z.X.; Wang, D.L.; Zhang, P.; Hu, Y.L. MgO modified sucrose-derived porous carbon composite for fluoride adsorption. Desalination Water Treat. 2024, 319, 100529. [Google Scholar] [CrossRef]
  36. Costa, L.R.D.; Jurado-Davila, I.V.; De Oliveira, J.T.; Nunes, K.G.P.; Estumano, D.C.; de Oliveira, R.A.; Carissimi, E.; Féris, L.A. Exploring Key Parameters in Adsorption for Effective Fluoride Removal: A Comprehensive Review and Engineering Implications. Appl. Sci. 2024, 14, 2161. [Google Scholar] [CrossRef]
  37. Zhao, M.X.; Zhou, X.M.; Li, J.F.; Li, F.; Li, X.D.; Yu, J.X.; Guo, L.; Song, G.P.; Xiao, C.Q.; Zhou, F.; et al. Efficient removal of phosphate and fluoride from phosphogypsum leachate by lanthanum-modified zeolite: Synchronous adsorption behavior and mechanism. J. Environ. Chem. Eng. 2024, 12, 113294. [Google Scholar] [CrossRef]
  38. Deng, Y.L.; Wang, S.Y.; Shi, K.; Xiong, H.X. Adsorption removal of fluoride from polluted drinking waters using Mn-Al-La oxide. Environ. Sci. Pollut. Res. 2024, 31, 7122–7137. [Google Scholar] [CrossRef]
  39. Yang, H.R.; Zhang, H.; Lu, J.X.; Cui, Y.G.; Wang, Y.Q.; Wang, X.Y.; Xue, J.; Cao, H. Advanced magnetic adsorbents for enhanced phosphorus and fluoride removal from wastewater: Mechanistic insights and applications. Sep. Purif. Technol. 2025, 353, 128195. [Google Scholar] [CrossRef]
  40. Ma, X.; Qi, T.; Chen, R.; Su, R.; Zeng, Z.; Li, L.; Wang, S. Experimental and theoretical calculations insight into acetone adsorption by porous carbon at different pressures: Effects of pore structure and oxygen groups. J. Colloid. Interf. Sci. 2023, 646, 67–77. [Google Scholar] [CrossRef]
  41. Sugita, H.; Morimoto, K.; Saito, T.; Hara, J. Simultaneous removal of arsenate and fluoride using magnesium-based adsorbents. Sustainability 2024, 16, 1774. [Google Scholar] [CrossRef]
  42. El Messaoudi, N.; Franco, D.S.P.; Gubernat, S.; Georgin, J.; Senol, Z.M.; Cigeroglu, Z.; Allouss, D.; El Hajam, M. Advances and future perspectives of water defluoridation by adsorption technology: A review. Environ. Res. 2024, 252, 118857. [Google Scholar] [CrossRef]
  43. Park, G.; Gabbaï, F.P. Phosphonium boranes for the selective transport of fluoride anions across artificial phospholipid membranes. Angew. Chem. Int. Ed. 2020, 59, 5298–5302. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, X.Z.; Pan, S.L.; Zhang, M.; Qi, J.W.; Sun, X.Y.; Gu, C.; Wang, L.J.; Li, J.S. Modified hydrous zirconium oxide/PAN nanofibers for efficient defluoridation from groundwater. Sci. Total Environ. 2019, 685, 401–409. [Google Scholar] [CrossRef] [PubMed]
  45. Dhanasekaran, P.; Sai, P.M.S.; Gnanasekar, K.I. Fixed bed adsorption of fluoride by Artocarpus hirsutus based adsorbent. J. Fluor. Chem. 2017, 195, 37–46. [Google Scholar] [CrossRef]
  46. Kuang, L.Y.; Liu, Y.Y.; Fu, D.D.; Zhao, Y.P. FeOOH-graphene oxide nanocomposites for fluoride removal from water: Acetate mediated nano FeOOH growth and adsorption mechanism. J. Colloid. Interf. Sci. 2017, 490, 259–269. [Google Scholar] [CrossRef]
  47. Zhou, H.C.; Kitagawa, S. Metal-Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415–5418. [Google Scholar] [CrossRef]
  48. Çiftlik, A.; Semerci, T.G.; Sahin, O.; Semerci, F. Two-dimensional metal-organic framework nanostructures based on 4,4′-sulfonyldibenzoate for photocatalytic degradation of organic dyes. Cryst. Growth Des. 2021, 21, 3364–3374. [Google Scholar] [CrossRef]
  49. Su, R.; Xie, C.; Alhassan, S.I.; Huang, S.; Chen, R.; Xiang, S.; Wang, Z.; Huang, L. Oxygen reduction reaction in the field of water environment for application of nanomaterials. Nanomaterials 2020, 10, 1719. [Google Scholar] [CrossRef]
  50. Wang, F.; Tian, F.K.; Deng, Y.X.; Yang, L.L.; Zhang, H.H.; Zhao, D.S.; Li, B.; Zhang, X.T.; Fan, L.M. Cluster-based multifunctional copper(ii) organic framework as a photocatalyst in the degradation of organic dye and as an electrocatalyst for overall water splitting. Cryst. Growth Des. 2021, 21, 4242–4248. [Google Scholar] [CrossRef]
  51. Su, R.K.; Li, Z.S.; Cheng, F.H.; Dai, X.R.; Wang, H.Q.; Luo, Y.T.; Huang, L. Advances in the degradation of emerging contaminants by persulfate oxidation technology. Water Air Soil Pollut. 2023, 234, 754. [Google Scholar] [CrossRef]
  52. Luo, Y.; Su, R. Cobalt-based mof material activates persulfate to degrade residual ciprofloxacin. Water 2024, 16, 2299. [Google Scholar] [CrossRef]
  53. Tolkou, A.K.; Zouboulis, A.I. Fluoride Removal from Water Sources by Adsorption on MOFs. Separations 2023, 10, 467. [Google Scholar] [CrossRef]
  54. Zhou, C.Y.; Huang, J.; Wang, Q.; Tang, H.; Hu, Y.C.; Wang, W.L. MOFs-Based Photoelectrochemical Sensing Interface and Its Applications. Prog. Chem. 2024, 36, 893–903. [Google Scholar] [CrossRef]
  55. Liu, R.L.; Song, J.Y.; Zhao, J.Y.; Wang, Z.; Xu, J.; Yang, W.S.; Hu, J.P. Novel MOF(Zr)-on-MOF(Ce/La) adsorbent for efficient fluoride and phosphate removal. Chem. Eng. J. 2024, 497, 154780. [Google Scholar] [CrossRef]
  56. Li, L.X.; Han, J.Z.; Huang, X.H.; Qiu, S.; Liu, X.H.; Liu, L.L.; Zhao, M.J.; Qu, J.W.; Zou, J.L.; Zhang, J. Organic pollutants removal from aqueous solutions using metal-organic frameworks (MOFs) as adsorbents: A review. J. Environ. Chem. Eng. 2023, 11, 111217. [Google Scholar] [CrossRef]
  57. Ma, X.; Xu, W.; Su, R.; Shao, L.; Zeng, Z.; Li, L.; Wang, H. Insights into CO2 capture in porous carbons from machine learning, experiments and molecular simulation. Sep. Purif. Technol. 2023, 306, 122521. [Google Scholar] [CrossRef]
  58. Li, D.; Su, R.; Ma, X.; Zeng, Z.; Li, L.; Wang, H. Porous carbon for oxygenated and aromatic VOCs adsorption by molecular simulation and experimental study: Effect pore structure and functional groups. Appl. Surf. Sci. 2022, 605, 154708. [Google Scholar] [CrossRef]
  59. Luo, Y.; Liu, Z.; Ye, M.; Zhou, Y.; Su, R.; Huang, S.; Chen, Y.; Dai, X. Synergistic enhancement of oxytetracycline hydrochloride removal by UV/ZIF-67 (Co)-activated peroxymonosulfate. Water 2024, 16, 2586. [Google Scholar] [CrossRef]
  60. Chen, L.; Zhang, X.; Cheng, X.; Xie, Z.; Kuang, Q.; Zheng, L. The function of metal–organic frameworks in the application of MOF-based composites. Nanoscale Adv. 2020, 2, 2628–2647. [Google Scholar] [CrossRef]
  61. Almáši, M. A review on state of art and perspectives of Metal-Organic frameworks (MOFs) in the fight against coronavirus SARS-CoV-2. J. Coord. Chem. 2021, 74, 2111–2127. [Google Scholar] [CrossRef]
  62. Pramanik, B.; Sahoo, R.; Das, M.C. pH-stable MOFs: Design principles and applications. Coord. Chem. Rev. 2023, 493, 215301. [Google Scholar] [CrossRef]
  63. Beheshti, H.; Bagherian, G.; Forghani Tehrani, G. Application of NH2-MIL-101(Al) as a new adsorbent for defluoridation of aqueous solutions: Isotherms, kinetics, and thermodynamics. Int. J. Enivron. Sci. Technol. 2024. [Google Scholar] [CrossRef]
  64. Serra-Crespo, P.; Ramos-Fernandez, E.V.; Gascon, J.; Kapteijn, F. Synthesis and characterization of an amino functionalized MIL-101 (Al): Separation and catalytic properties. Chem. Mater. 2011, 23, 2565–2572. [Google Scholar] [CrossRef]
  65. Luo, Y.; Su, R. Preparation of NH2-MIL-101(Fe) metal organic framework and its performance in adsorbing and removing tetracycline. Int. J. Mol. Sci. 2024, 25, 9855. [Google Scholar] [CrossRef]
  66. Marzilli, J.R.; Collaborators. Determination of Fluoride in Fluoride Tablets and Solutions by Ion-Selective Electrode: Collaborative Study. J. Assoc. Off. Anal. Chem. 1984, 67, 682–684. [Google Scholar] [CrossRef]
  67. Wang, J.; Guo, X. Adsorption kinetic models: Physical meanings, applications, and solving methods. J. Hazard. Mater. 2020, 390, 122156. [Google Scholar] [CrossRef]
  68. Wongphat, A.; Wongcharee, S.; Chaiduangsri, N.; Suwannahong, K.; Kreetachat, T.; Imman, S.; Suriyachai, N.; Hongthong, S.; Phadee, P.; Thanarat, P.; et al. Using excel solver’s parameter function in predicting and interpretation for kinetic adsorption model via batch sorption: Selection and statistical analysis for basic dye removal onto a novel magnetic nanosorbent. Chemengineering 2024, 8, 58. [Google Scholar] [CrossRef]
  69. Suwannahong, K.; Wongcharee, S.; Kreetachat, T.; Imman, S.; Suriyachai, N.; Hongthong, S.; Rioyo, J.; Dechapanya, W.; Noiwimol, P. Comprehensive cost-benefit and statistical analysis of isotherm and kinetic models for heavy metal removal in acidic solutions using weakly base polymeric chelating resin as adsorbent. Water 2024, 16, 2384. [Google Scholar] [CrossRef]
  70. Wang, J.; Guo, X. Adsorption isotherm models: Classification, physical meaning, application and solving method. Chemosphere 2020, 258, 127279. [Google Scholar] [CrossRef]
  71. Ahmadipouya, S.; Ahmadijokani, F.; Molavi, H.; Rezakazemi, M.; Arjmand, M. CO2/CH4 separation by mixed-matrix membranes holding functionalized NH2-MIL-101(Al) nanoparticles: Effect of amino-silane functionalization. Chem. Eng. Res. Des. 2021, 176, 49–59. [Google Scholar] [CrossRef]
  72. Isaeva, V.I.; Tarasov, A.L.; Starannikova, L.E.; Yampol’skii, Y.P.; Alent’ev, A.Y.; Kustov, L.M. Microwave-assisted synthesis of mesoporous metal-organic framework NH2-MIL-101(Al). Russ. Chem. Bull. 2015, 64, 2791–2795. [Google Scholar] [CrossRef]
  73. Liu, R.T.; Chi, L.N.; Wang, X.Z.; Wang, Y.; Sui, Y.M.; Xie, T.T.; Arandiyan, H. Effective and selective adsorption of phosphate from aqueous solution via trivalent-metals-based amino-MIL-101 MOFs. Chem. Eng. J. 2019, 357, 159–168. [Google Scholar] [CrossRef]
  74. Zhang, P.; Jia, X.; Hou, X.; Xia, R.; Guo, Y.; Zhang, X. Characterization of magnetic aluminum-based metal-organic frameworks and their adsorption performance for fluoride in water. Environ. Sci. Res. 2021, 34, 1139–1147. [Google Scholar] [CrossRef]
  75. Guo, Y. Study on the Adsorption of Phosphorus and Fluorine in Water by ZIF-8 and NH2-MIL-101 (Al). Ph.D. Dissertation, Yunnan University, Kunming, China, 2015. [Google Scholar]
  76. Kamarehie, B.; Noraee, Z.; Jafari, A.; Ghaderpoori, M.; Karami, M.A.; Ghaderpoury, A. Data on the fluoride adsorption from aqueous solutions by metal-organic frameworks (ZIF-8 and Uio-66). Data Brief 2018, 20, 799–804. [Google Scholar] [CrossRef] [PubMed]
  77. Tan, T.L.; Krusnamurthy, P.A.P.; Nakajima, H.; Rashid, S.A. Adsorptive, kinetics and regeneration studies of fluoride removal from water using zirconium-based metal organic frameworks. RSC Adv. 2020, 10, 18740–18752. [Google Scholar] [CrossRef]
  78. Jeyaseelan, A.; Naushad, M.; Ahamad, T.; Viswanathan, N. Design and development of amine functionalized iron based metal organic frameworks for selective fluoride removal from water environment. J. Environ. Chem. Eng. 2021, 9, 104563. [Google Scholar] [CrossRef]
  79. Li, W.F.; Zhang, T.; Lv, L.; Chen, Y.X.; Tang, W.X.; Tang, S.W. Room-temperature synthesis of MIL-100(Fe) and its adsorption performance for fluoride removal from water. Colloid. Surf. A 2021, 624, 126791. [Google Scholar] [CrossRef]
  80. Wang, X.G.; Zhu, H.; Sun, T.S.; Liu, Y.B.; Han, T.; Lu, J.X.; Dai, H.L.; Zhai, L.Z. Synthesis and study of an efficient metal-organic framework adsorbent (MIL-96(Al)) for fluoride removal from water. J. Nanomater. 2019, 2019, 3128179. [Google Scholar] [CrossRef]
  81. Cui, X.; Li, H.; Wang, Y.; Hu, Y.; Hua, L.; Li, H.; Han, X.; Liu, Q.; Yang, F.; He, L.; et al. Room-temperature methane conversion by graphene-confined single iron atoms. Chem 2018, 4, 1902–1910. [Google Scholar] [CrossRef]
  82. Nehra, S.; Raghav, S.; Kumar, D. Biomaterial functionalized cerium nanocomposite for removal of fluoride using central composite design optimization study. Environ. Pollut. 2020, 258, 113773. [Google Scholar] [CrossRef] [PubMed]
  83. Kumari, U.; Behera, S.K.; Meikap, B.C. A novel acid modified alumina adsorbent with enhanced defluoridation property: Kinetics, isotherm study and applicability on industrial wastewater. J. Hazard. Mater. 2019, 365, 868–882. [Google Scholar] [CrossRef]
  84. Millar, G.J.; CouperthWaite, S.J.; Dawes, L.A.; Thompson, S.; Spencer, J. Activated alumina for the removal of fluoride ions from high alkalinity groundwater: New insights from equilibrium and column studies with multicomponent solutions. Sep. Purif. Technol. 2017, 187, 14–24. [Google Scholar] [CrossRef]
Figure 1. XRD of NH2-MIL-101 (Al).
Figure 1. XRD of NH2-MIL-101 (Al).
Water 16 02889 g001
Figure 2. FT-IR Spectrum of NH2-MIL-101 (Al).
Figure 2. FT-IR Spectrum of NH2-MIL-101 (Al).
Water 16 02889 g002
Figure 3. SEM image of NH2-MIL-101 (Al) (Left: 100 nm, Right: 5 µm).
Figure 3. SEM image of NH2-MIL-101 (Al) (Left: 100 nm, Right: 5 µm).
Water 16 02889 g003
Figure 4. N2 adsorption–desorption isotherm (a) and pore size analysis (b) of NH2-MIL-101(Fe).
Figure 4. N2 adsorption–desorption isotherm (a) and pore size analysis (b) of NH2-MIL-101(Fe).
Water 16 02889 g004
Figure 5. Effect of material dosage on adsorption performance.
Figure 5. Effect of material dosage on adsorption performance.
Water 16 02889 g005
Figure 6. Effect of initial fluoride concentration on adsorption performance.
Figure 6. Effect of initial fluoride concentration on adsorption performance.
Water 16 02889 g006
Figure 7. Effect of solution pH on adsorption performance.
Figure 7. Effect of solution pH on adsorption performance.
Water 16 02889 g007
Figure 8. Adsorption equilibrium curve of NH2-MIL-101 (Al).
Figure 8. Adsorption equilibrium curve of NH2-MIL-101 (Al).
Water 16 02889 g008
Figure 9. Pseudo-first-order kinetic model.
Figure 9. Pseudo-first-order kinetic model.
Water 16 02889 g009
Figure 10. Pseudo-second-order kinetic model.
Figure 10. Pseudo-second-order kinetic model.
Water 16 02889 g010
Figure 11. Langmuir model.
Figure 11. Langmuir model.
Water 16 02889 g011
Figure 12. Freundlich model.
Figure 12. Freundlich model.
Water 16 02889 g012
Table 1. Parameters of the pseudo-first-order kinetic model.
Table 1. Parameters of the pseudo-first-order kinetic model.
q e (mg·g−1) k 1 (min−1) R 2
30.4−0.023460.95278
Table 2. Parameters of the pseudo-second-order kinetic model.
Table 2. Parameters of the pseudo-second-order kinetic model.
qe (mg·g−1)K2 (g/(mg·min))R2
37.00.0003730.99727
Table 3. Langmuir adsorption isotherm parameters.
Table 3. Langmuir adsorption isotherm parameters.
Temperature/k SlopeInterceptqmKLR2
2980.015010.05606670.26770.9942
3080.012650.0501790.25250.9801
3180.0160.02593630.61700.9879
Table 4. Freundlich adsorption isotherm parameters.
Table 4. Freundlich adsorption isotherm parameters.
Temperature/k SlopeIntercept1/nKFR2
2981.12220.67310.673113.25200.9761
3081.19540.61990.619915.68300.9547
3181.33200.51990.519921.48220.9417
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

Luo, Y.; Liu, Z.; Ye, M.; Zhou, Y.; Su, R.; Huang, S.; Chen, Y.; Dai, X. Mechanism of Enhanced Fluoride Adsorption Using Amino-Functionalized Aluminum-Based Metal–Organic Frameworks. Water 2024, 16, 2889. https://doi.org/10.3390/w16202889

AMA Style

Luo Y, Liu Z, Ye M, Zhou Y, Su R, Huang S, Chen Y, Dai X. Mechanism of Enhanced Fluoride Adsorption Using Amino-Functionalized Aluminum-Based Metal–Organic Frameworks. Water. 2024; 16(20):2889. https://doi.org/10.3390/w16202889

Chicago/Turabian Style

Luo, Yiting, Zhao Liu, Mingqiang Ye, Yihui Zhou, Rongkui Su, Shunhong Huang, Yonghua Chen, and Xiangrong Dai. 2024. "Mechanism of Enhanced Fluoride Adsorption Using Amino-Functionalized Aluminum-Based Metal–Organic Frameworks" Water 16, no. 20: 2889. https://doi.org/10.3390/w16202889

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop