Next Article in Journal
Explosive Spalling Mechanism and Modeling of Concrete Lining Exposed to Fire
Previous Article in Journal
The Use of Synthesized CoO/Co3O4 Nanoparticles as A Corrosion Inhibitor of Low-Carbon Steel in 1 M HCl
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal of Fe2+ and Mn2+ from Polluted Groundwater by Insoluble Humic Acid/Tourmaline Composite Particles

1
School of Civil Engineering, Liaoning Technical University, Fuxin 123000, China
2
Information Industry Electronics Eleventh Design and Research Institute Technology Engineering Co., Ltd., Dalian Branch, Dalian 116000, China
3
Water Services Association of Malawi, Tikwere House, City Center, Private Bag 390, Lilongwe 207213, Malawi
*
Author to whom correspondence should be addressed.
Materials 2022, 15(9), 3130; https://doi.org/10.3390/ma15093130
Submission received: 24 March 2022 / Revised: 21 April 2022 / Accepted: 24 April 2022 / Published: 26 April 2022

Abstract

:
Insoluble humic acid/tourmaline composite particles (IHA/TM) were prepared by combining inorganic tourmaline (TM) with the natural organic polymer humic acid (HA) and carbonizing them at 330 °C to study the removal characteristics and mechanism of Fe2+ and Mn2+. The results showed that the optimal ratio of TM to IHA is 2:3. When the temperature of the IHA/TM composite particles was 35 °C and the pH was 6, the adsorption of Fe2+ and Mn2+ by IHA/TM reached equilibrium at 240 min. The optimum dose of the adsorbent was 10 g/L, and the equilibrium adsorption capacities of Fe2+ and Mn2+ were 5.645 mg/g and 3.574 mg/g, respectively. The process of IHA/TM adsorption of Fe2+ and Mn2+ in water was spontaneous, endothermic and sustainable, and cooling was not conducive to adsorption. The pseudo-second order kinetic equation can well reflect the adsorption mechanism of IHA/TM on Fe2+ and Mn2+, and the Langmuir adsorption model better describes the isothermal adsorption behaviour. The material characterisation and adsorption experiments indicate that surface coordination and chemical precipitation are the main mechanisms of Fe2+ and Mn2+ removal by IHA/TM.

1. Introduction

Iron and manganese are the fourth and second most abundant metal elements in the Earth’s crust and are usually present in groundwater in the form of divalent ions (Fe2+ and Mn2+) [1]. There are two main sources of these metals: geological sources and anthropogenic sources. Geological sources include minerals containing Fe and Mn, which dissolve when water penetrates soil and rocks and preserves them in an aqueous solution. Anthropogenic sources include industrial wastewater, landfill leakage, acid mine wastewater, etc., all of which result in high concentrations of Fe2+ and Mn2+ in groundwater [2]. There are serious iron and manganese pollution problems in groundwater in mining areas of northeast, northwest and north China due to mining and disorderly discharge of industrial wastewater. The concentrations of Fe2+ and Mn2+ in polluted groundwater in Shuangyashan mining area of Heilongjiang and Jilin Tonghua mining area in Northeast China can reach up to 60.00 mg/L and 14.17 mg/L [3]; the Fe2+ content in the polluted water of the Malan coal mine in North China reaches 102.90 mg/L, the Mn2+ content is 8.10 mg/L, and the pH is 4.83 [4]; the content of Mn2+ in the weakly acidic mine drainage of the Enhong coal mine in Northwest China is as high as 32 mg/L, and the iron ion content is as high as several hundred mg/L [5]. High concentrations of Fe2+ and Mn2+ exist in groundwater, which will not only cause serious color problems, but also damage the local ecological environment. Long-term excessive intake of iron and manganese-contaminated water caused by water shortage in mining areas may cause chronic poisoning and affect health [6]. How to remove the pollution of high concentrations of Fe2+ and Mn2+ in groundwater has always been the concern of the majority of environmental workers [7].
Domestic and foreign treatment methods for removing Fe2+ and Mn2+ from groundwater mainly include filtration, oxidation, precipitation, biological treatment, adsorption method and membrane technology [8,9]. Among them, the adsorption method has been proved to be a technology with low cost, good effect, high efficiency, and is easy to operate [10]. Selecting and preparing the adsorption material for Fe and Mn removal are key to research and application. Geremias et al. used coal mine tailings as adsorbents and successfully adsorbed iron and manganese ions in acidic mine drainage [11]. Apkar’yan et al. prepared an environmentally friendly glass-ceramic particle material based on broken glass, red clay and organic additives, which can effectively remove iron and manganese ions from groundwater [12]. Aziz et al. discovered the great potential of zeolite as a good adsorbent for heavy metal ions by exploring the adsorption capacity of activated clinoptilolite for iron and manganese ions in acid wastewater from palm oil mills [13]. In addition, some low-cost and easily available waste materials such as rice husk ash, citrus peel, and natural minerals such as hydroxyapatite have also become the current application hotspots [14,15]. However, the existing adsorbents for removing iron and manganese ions are still insufficient in terms of cost and adsorption effect, especially for groundwater contaminated with high concentrations of iron and manganese ions. The ion adsorbent has become the key to research and application, which has also become the focus of our research.
Tourmaline (TM) is a complex cyclic silicate mineral with high mechanical and chemical stability. It can continuously release negative ions, has spontaneous polarity, and does not pollute the environment. TM is a relatively environmentally friendly material with spontaneous electrode characteristics. In 1989, Kubo et al. found that tourmaline has the characteristics of a spontaneous electrode and first proposed the existence of an electrostatic field on the surface of tourmaline [16]. Subsequently, many scholars have conducted research on the adsorption ability and removal of heavy metal ions by tourmaline [17]. However, TM has the disadvantages of high agglomeration and difficult removal from water due to its large surface energy [18]. Currently, the use of surface modification or preparation of composite materials to reduce agglomeration and improve adsorption performance has become a common research topic in the application of TM [19]. Chen et al. synthesized bentonite and tourmaline montmorillonite composites (TMMs) by a vacuum sintering method that can better adsorb Pb (II) in water [20]. Liao et al. reported that silver-loaded polyacrylic acid-modified tourmaline composites could efficiently remove methylene blue and Cu (II) ions from water by catalytic degradation [21]. Studies have shown that the compounds used for surface modification and composite adsorbents have an important characteristic in common: strong metal complexing abilities. Humic acid (HA) is such a material. HA, a polymeric organic material that widely exists in nature, contains carboxyl, hydroxyl, carbonyl and other active groups that can adsorb onto heavy metals and undergo complexation and redox reactions with heavy metals in the environment [22,23]. However, as an adsorbent, humic acid has low mechanical strength, is difficult to separate from water and has poor biological stability, so it is still not directly used as an adsorbent; thus, the development of insoluble humic acid (IHA) has become an important research topic. Zhao et al. made insoluble humic acid and used it to remove Mn2+ from aqueous solution. At pH = 5.6 and at 25 °C, the rate of Mn2+ removal by insoluble humic acid was 92% [24]. Wei et al. prepared chitosan-crosslinked insoluble humic acid (CS @ IHA) by a dropping ball method, and this material had a high adsorption capacity for Pb (II) in water [25]. The combination of IHA and TM can not only change the surface properties and environmental behaviours of TM but also affect the adsorption effect of pollutants.
Based on this behaviour, a new composite adsorbent, insoluble humic acid/tourmaline (IHA/TM), was prepared by making full use of the respective advantages of HA and TM. The thermodynamic and kinetic behaviours of Fe2+ and Mn2+ were analysed, the key parameters of adsorption and adsorption capacity were determined, and the microscopic removal mechanism was discussed to provide a scientific basis and technical background for the practical application of IHA/TM in the treatment of groundwater containing Fe2+ and Mn2+.

2. Materials and Methods

2.1. Preparation of IHA/TM Composite Particles and Determination of Their Optimum Ratio

TM is mainly composed of Schorl. It had a particle size of 300 mesh, and its chemical formula was NaFe3Al6(BO3)3SiO6O18(OH)4. The chemical composition of tourmaline was determined by X-ray fluorescence (XRF) as follows: SiO2 36.28%, Al2O3 22.86%, Fe2O3 19.41%, B2O3 7.81%, MgO 4.66%, CO2 3.23%, CaO 2.30%, Na2O 1.77% and other 1.68%.
Preparation of IHA: HA was dissolved in 0.1 mol/L NaOH solution and undissolved impurities were discarded. The pH of the HA-containing solution was adjusted to 1.2 ± 0.2 with 6.0 mol/L HCl and placed in a water bath for heating at 65 ± 5 °C for 2 h. The precipitated HA was dried in an oven at 45 ± 5 °C to produce IHA.
Preparation of IHA/TM composite particles occurred as follows. The pretreated IHA was dissolved in 0.1 mol/L NaOH solution, and then controlled TM was mixed with IHA at an optimal mixing ratio. The pH was adjusted to 6.0, the mixture was shaken (25 °C, 150 r/min) for 12 h and centrifuged. The precipitate was washed three times with deionized water and then dried in an oven at 65 ± 5 °C for 24 h [26]. After grinding and screening, the solid was mixed with etherified starch (10/1, (v/v)) which was dextrinated at 95 °C to make adsorbent pellets with 3–5 mm particles, which were calcined in a muffle furnace at 300 °C for 1 h to form IHA/TM composite particles.
The optimum ratio of IHA and TM was determined. Based on the characteristics of typical groundwater high in iron and manganese in Fuxin, Liaoning Province, China, the Fe2+ content of the water sample was set at 25 mg/L, and the Mn2+ content was set at 10 mg/L. The mixing ratios of IHA and TM were controlled at 1:3, 2:3, 1:1, 3:2 and 3:1 in the preparation of the composite particles, and 1.2 g of composite particles was placed in a 100 mL water sample. After shaking in a shaker (25 °C, 150 r/min) for different times, the concentration of Fe2+ and Mn2+ and the pH value of the solution were measured.

2.2. Material Characterisation

Changes in surface morphology, particle size and agglomeration were analysed using a JEOL JSM-7500F scanning electron microscope (SEM, Tokyo, Japan). Elemental and content analyses of the materials were performed with an FYFS-2002E energy dispersive spectrometer (EDS). An XRD-6100 X-ray diffractometer (XRD, Shimadzu, Japan) was used for sample phase analysis. FTIR studies were performed with an IRPrestige-21 Fourier transform infrared spectrometer to characterise molecular structures, chemical bonding and functional group changes in adsorption materials.

2.3. Static Adsorption Test

2.3.1. Effect of IHA/TM Dose on Adsorption

The experimental study simulated the Xinqiu mining area in Liaoning, China. In this area, due to mining and industrial wastewater discharge, the groundwater is seriously polluted by iron and manganese ions. The maximum concentrations of Fe2+ and Mn2+ were 20.5–25.0 mg/L and 5.9–9.8 mg/L, respectively, and the pH was 5.8–6.1. A 100 mL of water sample with Fe2+ content 25 mg/L and Mn2+ content 10 mg/L was taken, pH adjusted to 6.0 with 0.1mol/L HCl and NaOH solution, and dosage of IHA/TM controlled to be 0.4–1.4 g. IHA/TM was weighed accurately and added to the water sample, shaken at 150 r/min for a fixed time at 25 °C, and after centrifugation, the concentrations of Fe2+ and Mn2+ were measured.
The amount of Fe2+ and Mn2+ adsorbed on IHA/TM per unit mass and removal rate R (%) were calculated using Equations (1) and (2), respectively [27].
q e =   ( C 0 C e )   V m
R = C 0 C e C 0 × 100 %
where qe is the adsorption capacity of IHA/TM composite particles at equilibrium, mg/g; C0 is the initial concentration of groundwater containing Fe2+ and Mn2, mg/L; Ce is the solution concentration when the adsorption reaches equilibrium, mg/L; V is the solution volume, L; and m is the adsorbent mass, g.

2.3.2. Effect of Initial pH on Adsorption

To determine the effect of the initial pH on the adsorption performance of IHA/TM, 1.0 g of IHA/TM composite particles was added to 100 mL of water with an Fe2+ content of 25 mg/L and Mn2+ content of 10 mg/L, and the pH was adjusted to 2.0–9.0. The samples were shaken at 150 r/min for 240 min at 25 °C, and the concentrations of Fe2+ and Mn2+ were measured after centrifugation.

2.3.3. Adsorption Isotherm Experiment

Different initial concentrations of Fe2+ and Mn2+ solutions were separated into 100 mL aliquots in 250 mL conical flasks, and the IHA/TM composite particles were stored in 1.0 g samples at pH 6. The filtrate was shaken at 150 r/min for 240 min at different temperatures to determine the Fe2+ and Mn2+ concentrations.

2.3.4. Adsorption Kinetics

Fe2+ and Mn2+ solutions (100 mL) with different initial concentrations were placed in 250 mL conical flasks. The dosage of IHA/TM composite particles was kept at 1.0 g at a pH of 6.0 and temperature of 25 °C. The concentrations of Fe2+ and Mn2+ in the filtrate were measured after different reaction times.

2.3.5. Regeneration Test

To evaluate the reusability of the IHA/TM composite particles, the IHA/TM composite particles after adsorption saturation were centrifugally filtered, the unabsorbed Fe2+ and Mn2+ ions were washed away, mixed with 0.1 mol/L HNO3 solution, and shaken at 25 °C for 240 min washed with deionized water until neutral, and dried at 105 °C. The adsorption and desorption experiments were repeated 5 times using the same batch of IHA/TM composite particles, and the removal rates of Fe2+ and Mn2+ by IHA/TM after each desorption were calculated.

3. Results and Discussion

3.1. Optimal Ratio of IHA and TM

As shown in Figure 1, the mixing ratio of IHA and TM greatly influences the rate of Fe2+ and Mn2+ removal. IHA can improve the dispersion performance of TM and reduce its agglomeration, and its surface has many active groups (such as -COO-, -COOH, and -OH) that can transfer electrons and improve the adsorption capacity [28]. However, these IHA functional groups also occupy the active sites of TM, reducing the adsorption performance of the adsorbent [29]. Therefore, there is an optimal TM:IHA ratio. When TM/IHA = 2:3, the rates of Fe2+ and Mn2+ removal were 99.02% and 97.67%, respectively. In this study, a composite material with TM/IHA = 2:3 was selected as the experimental adsorbent to explore the adsorption characteristics of the adsorbent for Fe2+ and Mn2+.

3.2. Effect of IHA/TM Dose on the Adsorption of Fe2+ and Mn2+

Figure 2 shows that with increasing IHA/TM composite particle dose, the rates of Fe2+ and Mn2+ removal first increased and then stabilised. This stabilisation is due to the increase in IHA/TM, which provides many surface active sites, increases the contact probability of IHA/TM with Fe2+ and Mn2+, and increases the rate of Fe2+ and Mn2+ removal. When the dose was 10 g/L, the rates of Fe2+ and Mn2+ removal were 99.75% and 99.45%, respectively. The rates of Fe2+ and Mn2+ removal remained mostly stable as the dose was continually increased. The dose of IHA/TM was selected as 10 g/L by comprehensive comparison.

3.3. Effect of Initial pH on Adsorption of Fe2+ and Mn2+

Studies have shown that the pH of the solution greatly influences the complexation reaction of TM, IHA and metal, and their complexing ability and stability increase with increasing pH [30]. Figure 3 shows that when the pH was 2–6, the rates of Fe2+ and Mn2+ removal increased significantly with increasing pH. When pH > 6, the rate of Fe2+ and Mn2+ removal tended to stabilise because the pH of the solution system not only affects the surface charge, ionisation degree and type of adsorbent, but also determines the forms of Fe2+ and Mn2+ present in the solution, resulting in a large difference in the performance of the adsorbent at different pH values [31].
When pH < 4, the high concentration of H+ competitively adsorbed onto the surface, blocking Fe2+ and Mn2+, and it more easily reacted with the active groups on the surface of the IHA/TM so that metal ions easily dissociated from the complex. When the pH was 4–6, the rate of Fe2+ and Mn2+ removal continued to increase, but the growth rate slowed slightly. The adsorption of H+ on the adsorbent was reduced, while H+ dissociated from functional groups such as carboxyl and hydroxyl groups as they were deprotonated. Therefore, the negative charge on the IHA/TM surface increased, and the adsorption sites increased [32]. Under acidic conditions, some of the Fe2+ combined with OH to form Fe(OH)+, the surface charge decreased, the electrostatic repulsion between the surface protonated adsorbent weakened, and Fe2+ was more easily adsorbed. In addition, the active functional groups of IHA tended to form complexes with iron and manganese ions at a high pH, which is more conducive to the formation of humic acid-iron-manganese complexes, thereby reducing Fe2+ and Mn2+ in the solution [33]. However, under neutral and basic conditions, Fe2+ and Mn2+ were precipitated as Fe(OH)2 and Mn(OH)2, nearly completing their removal.
Due to the good buffer performance of the TM in the adsorbent, the pH of the solution tended to be neutral after the reaction reached equilibrium. This phenomenon was due to the spontaneous polarity of TM (chemical Equations (3)–(5)). This electrode allowed the surrounding air to become weakly electrolyzed, H+ obtained electrons, and OH combined with H2O to generate negative ions. The performance was also related to the hydroxyl (-OH) crystal structure and the numerous atomic bonds on the surface. TM may be a very good adsorbent for the remediation of Fe2+ and Mn2+ contaminated water because it can be applied over a wide pH range [34].
H 2 O H + + OH
2 H + + 2 e H 2
OH + nH 2 O OH ( H 2 O ) n
The removal of Fe2+ and Mn2+ by adsorption onto IHA/TM is more successful under weakly acidic conditions. In this study, an initial pH of 6 was chosen; at this pH, Fe2+ and Mn2+ did not precipitate, and the rates of Fe2+ and Mn2+ removal were 99.84% and 99.20%, respectively.

3.4. Isothermal Adsorption Analysis

Figure 4a shows that at 35 °C, the qe values of TM for Fe2+ and Mn2+ adsorption were 2.615 mg/g and 1.937 mg/g, respectively, while the equilibrium adsorption capacities of IHA/TM for Fe2+ and Mn2+ were 5.645 mg/g and 3.574 mg/g, respectively. The capacity of IHA/TM for iron and manganese removal was 2.16 and 1.85 times higher than that of TM, indicating that the removal capacity of IHA/TM for Fe2+ and Mn2+ was better than that of TM.
To explore the adsorption behaviour of Fe2+ and Mn2+ on IHA/TM, the Langmuir, Freundlich and Temkin models were used to fit the isothermal adsorption data (Figure 4b–d).
Langmuir isothermal adsorption equation [35]:
C e q e = 1 ( K L q m ) + C e q m
Freundlich isothermal adsorption equation:
Ln q e = Ln K F + 1 n Ln C e
Temkin isothermal adsorption equation:
q e = B ln A + B ln C e
where Ce is the adsorption equilibrium solution concentration, mg/L; qe and qm are adsorption capacity at equilibrium and at saturation, respectively, mg/g; and KL, KF, n, A and B are all adsorption constants.
Linear fitting was performed using the Langmuir, Freundlich and Temkin models, and the relevant parameters are shown in Table 1. Based on R2 (>0.99), as shown in Table 1, the adsorption of Fe2+ and Mn2+ onto IHA/TM was in good agreement with the Langmuir isothermal adsorption equation. The results show that the adsorption of both Fe2+ and Mn2+ onto IHA/TM is monolayer adsorption, and the saturated adsorption capacity of IHA/TM for Fe2+ and Mn2+ increased with increasing temperature. For the dimensionless parameter separation factor RL, the expression RL = 1/(1 + KL C0) can be used to represent the adsorption properties of materials: the RL value indicates that the adsorption type is unfavourable (RL > 1), favourable (0 < RL < 1), linear (RL = 1) or irreversible (RL = 0) [36]. The RL of this experiment was 0–1, and the adsorption process of Fe2+ and Mn2+ onto IHA/TM was favourable.

3.5. Thermodynamic Analysis

To further determine the thermodynamic effects, chemical Equations (9) and (10) were used to estimate the thermodynamic parameters of the IHA/TM composite particles for the Fe2+ and Mn2+ adsorption processes. These parameters included the enthalpy change (ΔH), entropy change (ΔS) and Gibbs free energy change (ΔG).
Δ G = R T ln K d
ln K d = Δ S R Δ H R T
K d = q e C e
where R is the universal constant (8.314 J/(mol·K)), T is the absolute temperature (K), and Kd is the equilibrium constant calculated from Equation (11).
The enthalpy and entropy changes can be linearly fitted to the data based on the variation in the adsorption constants Kd with temperature. The results are given in Table 2. The negative values of ΔG for Fe2+ and Mn2+ indicate that the adsorption of IHA/TM is spontaneous and increases with increasing temperature. The positive values of ΔH for Fe2+ and Mn2+ indicate that the adsorption of IHA/TM is endothermic, and a higher temperature is more conducive to the reaction, which again verifies the conclusion of the Langmuir isothermal adsorption model. In addition, the positive value of ΔS indicates that the adsorption of Fe2+ and Mn2+ by the IHA/TM composite is accompanied by chemical reactions in which different types of ions are released into the liquid phase, indicating that the adsorption process is sustainable. Overall, the process of Fe2+ and Mn2+ adsorption onto IHA/TM is spontaneous, heat-absorbing and sustainable.

3.6. Adsorption Kinetic Analysis

Figure 5 shows the variation in the adsorption capacity versus adsorption time for different initial concentrations of Fe2+ and Mn2+ on IHA/TM at a temperature of 25 °C, pH of 6, and a dose of 10 g/L. Two different removal stages can be observed; in the first stage, the adsorption capacity for Fe2+ and Mn2+ increases rapidly with increasing adsorption time, and then in the second stage, the adsorption capacity increases slowly until it remains almost constant and the reaction reaches dynamic equilibrium.
The Fe2+ and Mn2+ adsorption kinetics were fitted to quasi-first order (Equation (12)) and quasi-second order (Equation (13)) models to further elucidate the kinetic mechanism of Fe2+ and Mn2+ removal by IHA/TM [37] (Figure 6a–d).
lg ( q e q t ) = lg q e K 1 2.203 t
t q t = 1 K 2 q e 2 + t q e
where qe and qt are the adsorption amount at equilibrium and t min, respectively, mg/g; K1 and K2 are the pseudo-first and second order kinetic rate constants, 1/min, mg/g/min.
The kinetic parameters of Fe2+ and Mn2+ adsorption by IHA/TM at different initial concentrations are given in Table 3. It can be seen from Table 3 that the K1 and K2 vary with the initial concentration regardless of Fe2+ or Mn2+, and the dispersion degree of kinetic constant K2 is larger than that of K1. With the increase in the initial concentration of Fe2+ and Mn2+, the kinetic rate constant K2 gradually decreased, which is consistent with the experimental results obtained by Nekouei et al. when the composite adsorbent CS-EDTA-mGO effectively adsorbed and removed Rhodamine B [38]. They believed that in the presence of a chemisorption reaction, the chemisorption rate was affected by the adsorbate concentration in the solution. Therefore, for the IHA/TM composite adsorbent, the complex chemical action occurs during the adsorption process, and the kinetic rate constant K is not independent of the initial concentrations of Fe2+ and Mn2+. This is different from the fact that the kinetic constant K, which is dominated by the physical adsorption process, is independent of the initial conditions [39]. It is very difficult to judge which kinetic model fits from the discreteness of the K value. Therefore, for the reaction process dominated by chemisorption, scholars still only judge the fitting kinetics from R2 and qe. The R2 values show that the fitting results of the adsorption kinetics of Fe2+ and Mn2+ onto IHA/TM are more consistent with the pseudo-second order kinetic equation (R2 > 0.99). The formation of chemical bonds is the main factor affecting pseudo-second order kinetics, indicating that adsorption mainly proceeds by chemisorption [40]. Surface adsorption, external liquid film diffusion and intraparticle diffusion explain the entire process of IHA/TM adsorption. Of course, when the adsorption reaction is chemical adsorption, it is worthwhile to further study the relationship between the adsorption kinetic rate constant K, the equilibrium adsorption capacity qe and the initial concentration of pollutants, as well as the effect on the fitted kinetics.
To determine the diffusion mechanism, the Weber-Morris equation (chemical Equation (14)) was used for piecewise linear fitting of the experimental data.
q t = k p t 0.5 + C
where qt is the adsorption amount at t min, mg/g; kp is the diffusion rate constant, mg/g·min1/2; and C is a constant related to the boundary layer thickness, mg/g.
In Figure 7, the fitted curves of qt and t0.5 are divided into two stages for different concentrations of Fe2+ and Mn2+. In the initial stage, the slope of the straight line is greater, the adsorption rate is faster, and many vacant active sites on the IHA/TM are available for Fe2+ and Mn2+ adsorption. The adsorption reaction is mainly controlled by the ion exchange between Fe2+ and Mn2+, the functional groups on the surface of the IHA/TM adsorbent and the diffusion of the outer liquid film. In the second stage, the adsorption rate is lower, and as the binding sites are gradually occupied, a concentration gradient forms between the surface and the interior of IHA/TM, promoting the diffusion of Fe2+ and Mn2+ inside IHA/TM, and the adsorption rate is controlled by the diffusion rate inside the particles [41]. Based on the different slopes of the first and second stages, adsorption occurs in gradual stages, surface adsorption is controlled by the thickness of the boundary layer, and the diffusion rate constant K2d is smaller than K1d (Table 4), which indicates that the internal diffusion rate of the particles is slower. In addition, the curve of qt versus t1/2 is bilinear over the entire time range, and the fitted curve does not cross the origin, indicating that the process of IHA/TM adsorption of Fe2+ and Mn2+ is not only controlled by intraparticle diffusion, but also affected by other adsorption stages [42,43].

3.7. Desorption and Reusability of IHA/TM

The results of the desorption experiment are given in Figure 8. Figure 8 shows that after 5 cycles, the removal rates of Fe2+ and Mn2+ decreased from 99.85% and 99.51% in the zeroth cycle to 89.4% and 87.07% in the fifth cycle (loss rates were 10.45% and 12.44%, respectively). The results showed that IHA/TM composite particles had good reusability and could remove Fe2+ and Mn2+ ions in actual groundwater.

3.8. Comparison of Saturated Adsorption Capacity of Materials

Table 5 shows that compared with similar adsorbents reported in the previous literature, the IHA/TM composite particles had certain advantages, including a relatively high saturated adsorption capacity for Fe2+ and Mn2; therefore, they are considered to be highly efficient adsorbents.

3.9. Microscopic Characterisation of IHA/TM Composite Particles

3.9.1. XRD Analysis

The crystal structure of the IHA/TM composite particles was studied by XRD, as shown in Figure 9a. TM is composed of NaFe3Al6(BO3)3SiO6O18(OH)4, which has several main characteristic peaks at 18.76°, 20.84°, 22.16°, 26.60°, 29.88°, 34.56°, 44.04°, 54.88° and 63.96°. All the characteristic peaks are consistent with the standard peaks of tourmaline reported in the literature [20]. By comparing the diffraction peaks of IHA/TM and raw materials, it was found that IHA/TM has diffraction peaks similar to those of TM, indicating that the sintering process does not destroy the crystal structures of the raw materials. After adsorption of Fe2+ and Mn2+, IHA/TM exhibited hydroxide precipitation containing manganese and iron, which reflected the electrode nature of TM. The adsorbent could react with Fe2+ and Mn2+ by chemical precipitation. Additionally, complex products containing manganese and iron also appeared, indicating that Fe2+ and Mn2+ underwent surface coordination reactions with the Al, Si, hydroxyl and carboxyl groups on the surface of the adsorbent.

3.9.2. FTIR Analysis

Figure 9b shows the FTIR spectra of TM, IHA and IHA/TM. The absorption band at 472.56 cm−1 corresponds to Si-O, while the peaks at 700.16 cm−1 and 775.38 cm−1 are assigned to M-O (M=Fe or Al). The peaks at 1653 cm−1,1031.92 cm−1, 3500–3300 cm−1, 2950–2850 cm−1, 1720–1700 cm−1 and 1650–1580 cm−1 correspond to H-O-H, O-Si-O stretching, N-H stretching vibration peak and O-H stretching vibration, C-H aliphatic stretching, C=O stretching of the carboxyl group, and C=C stretching of the aromatic group, respectively [46]. It can be concluded from the above data that IHA contains numerous carboxyl and phenolic hydroxyl functional groups, which can be used as surface adsorption sites, so it has a good ability to adsorb heavy metals. Compared with those of IHA, the vibrations of the C=O and O-H stretching vibration peaks in the IHA/TM spectra were weaker, indicating that IHA may undergo dehydration and decarboxylation as the composite particles are heated during preparation, resulting in the loss of some acidic groups. However, it can be inferred from the spectra that the prepared composite particles retain the active functional groups of TM and IHA, so the preparation process does not affect the adsorption of metal ions.
After adsorption of Fe2+ and Mn2+ by IHA/TM, the functional groups changed significantly, and the original absorption band at 3566.38 cm−1 corresponds to the M-OH in TM [47]. After the reaction, the absorption band at 3566.38 cm−1 was significantly weakened, indicating that the hydroxyl M-OH on the TM surface combined with Fe2+ and Mn2+ in the solution to form surface complexes. The original O-H absorption band, N-H stretching vibration peak at 3500~3300 cm−1 and C=O absorption band at 1720–1700 cm−1 were significantly weakened and migrated, and the C-N absorption peak at 1380 cm−1 was shifted upwards. M. Adhikari also observed similar phenomena, which he believed were caused by the formation of complexes between metal ions and humic acid [48]. The change in functional groups indicated that N-H, C-N, carboxyl and hydroxyl groups were involved in the adsorption of Fe2+ and Mn2+ by IHA/TM.

3.9.3. SEM Analysis

To further elucidate the morphological characteristics of IHA/TM, SEM images of TM, IHA and IHA/TM, before and after the reaction, were selected, as shown in Figure 10. As seen in Figure 10a, the surface of TM is in the form of a mutually stacked lamellar structure with a rough surface and cluster-like aggregation and adhesion properties, which is beneficial to the preparation of composite materials. As seen in Figure 10b, the surface of IHA is an inhomogeneous layered structure with an uneven surface that increases the specific surface area of the adsorbent and helps to adsorb heavy metal ions. Figure 10c shows that small tourmaline particles are attached to large humic acids, and there are flocculent or irregular flakes between the IHA surface and TM to combine the two, which may be because IHA reacts with Al, Si, Mg and Fe on TM to form IHA-A complexes (A represents Al, Si, Mg and Fe). After compounding IHA and TM, the voids and cavities on the surface of the insoluble IHA/TM composite particles increased, the adsorption sites increased, the specific surface area increased, the interparticle dispersion was more uniform, the particle shape was dense and irregular, the surface was much rougher and layered than the original TM, the polarity was reduced, the dispersion of TM was improved and the agglomeration problem of TM was solved. Figure 10d shows that after the adsorption of Fe2+ and Mn2+ by the IHA/TM composite particles, the surface structure morphology changed significantly, losing the complete cluster structure formed by large particles, forming small particles, appearing as a cover layer of particulate matter and producing a multitude of small particles and fragments because the surface of the IHA/TM adsorbent provides the necessary channels and sufficient adsorption space for the adsorption of Fe2+ and Mn2+ that facilitates the adsorption of Fe2+ and Mn2+. It is speculated that surface complexation may have occurred between the Fe2+ and Mn2+ ions and the IHA/TM composite adsorbent, and physical adsorption may have occurred at the surface. The soluble Fe2+, Mn2+, FeOH+, MnOH+, Fe(OH)2(aq) and Mn(OH)2(aq) were converted to Fe(OH)2(s) and Mn(OH)2(s).

3.9.4. EDS Analysis

The EDS spectra of the water samples before and after IHA/TM treatment are shown in Figure 11. Figure 11a shows that the main elements in IHA/TM composite particles were C, O, N, Al, Si, B and Fe, among which C, O, H and N were the main elements in IHA. After the adsorption of Fe2+ and Mn2+ by IHA/TM, the EDS (Figure 11b) spectrum showed higher levels of Fe than before the reaction, and a signal corresponding to Mn appeared at the same time, indicating that IHA/TM successfully adsorbed Fe2+ and Mn2+ ions in aqueous solution; in addition, the content of O, Al, Mg, Si and other elements decreased. It is speculated that the surface coordination reactions of Al, Si, hydroxyl and carboxyl groups with Fe2+ and Mn2+ ions may form surface complexes of Fe and Mn, resulting in a decrease in the O, Al and Si contents.

4. Mechanistic Analysis of Iron and Manganese Removal

The IHA/TM composite particles have good adsorption properties for Fe2+ and Mn2+, which is determined by the adsorption mechanism of TM and IHA. The structure of TM is compact and metal ions do not easily enter its crystal structure, so the adsorption of TM is mainly surface adsorption. TM is a cyclic silicate crystal mineral composed of elements such as Si, Al, Na, Ca, Mg, B, Fe, O and H. Its crystal structure can be regarded as composed of a [Si6O18] complex trigonal ring, [BO3] triangle and [Mg(Fe)-O5(OH)] triple octahedron with a common edge and a common vertex. There are many Si and Al elements on the surface of the mineral. It easily reacts with water in aqueous solution to form a hydroxylated surface (≡MeOH) (chemical Equations (15) and (16)). Chemical Equation (15) shows that the Si-O bond within the Si6O18 six-membered ring breaks and interacts with water molecules, with the Si bond bonding to OH and the O bonding to H, resulting in a bonded hydroxyl group on the surface of TM. Chemical Equation (16) shows that the intraoctahedral bond of AlO6 is also broken, and then the hydroxyl group is bonded, resulting in the exposure of a large number of metal cations. Then it forms surface complexes with Fe2+ and Mn2+ in water (chemical Equations (17) and (18)). Since Fe2+ and Mn2+ have empty orbitals, they tend to form ligands [49]. At the same time, the molecular formula of IHA is C9H9NO6, the basic structural unit is an aromatic ring, and the alkyl chain contains a variety of functional groups such as -NH2, -COOH, -OH and -C=O. These groups can be closely combined with metals, giving IHA a high adsorption capacity for metals. Therefore, Fe2+ and Mn2+ can be coordinated and adsorbed on the surface of IHA/TM to form complexes and generate chemical bond forces, which also confirms the conclusion of the adsorption kinetics. Additionally, the variation in surface functional groups in FTIR indicates that functional groups such as M-OH, NH, CN, carboxyl and hydroxyl groups are involved in the surface complexation reaction of IHA/TM on Fe and Mn ions. This result is consistent with the complexation products of AlFeO3, FeSiO3, MnAl2O4 and MnSiO3 obtained by XRD.
  -Si + OH = -SiOH ;   -SiO + H + = -SiOH ;
-Al + OH = -AlOH ;   -AlO + H + = -AlOH ;
MeOH + Fe 2 + MeOFe + + H +
MeOH + Mn 2 + MeOMn + + H +
In addition, based on the electrode nature of TM in IHA/TM, under the action of an electric field, the negative electrode of TM particles adsorbs Fe2+ and Mn2+ to its surface through electrostatic gravity, resulting in an increase in the concentration of Fe2+ and Mn2+ around the negative electrode, which combines with OH in water and reaches a certain concentration to produce precipitation (chemical Equations (19) and (20)), thus decreasing the solution of Fe2+ and Mn2+ concentration. The positive electrode of TM particles can adsorb OH in water, leading to an increase in the concentration of OH around it, again favouring the aggregation of ions in solution and the formation of insoluble precipitates [50]. This is consistent with the detection of Fe(OH)2 and Mn(OH)2 in the XRD analysis. The removal mechanism of IHA/TM is shown in Figure 12.
2 OH + Fe 2 + Fe ( OH ) 2
2 OH + Mn 2 + Mn ( OH ) 2

5. Conclusions

(1) Using the respective advantages of IHA and TM, IHA/TM composite particles were made to study the effect of adsorbing the heavy metals iron and manganese ions from polluted groundwater. The results showed that IHA/TM composite particles can effectively remove iron and manganese ions from groundwater. The optimal adsorption conditions of IHA/TM for Fe2+ and Mn2+ were as follows: TM:IHA was 2:3, the dosage of IHA/TM was 10 g/L, the pH was 6, the reaction time was 240 min, and the temperature was 35 °C. The equilibrium adsorption capacities of Fe2+ and Mn2+ were 5.645 mg/g and 3.574 mg/g, respectively. After five cycles of regeneration, the rates of Fe2+ and Mn2+ removal decreased by 10.45% and 12.44%, respectively. IHA/TM has good reusability and great potential for removing metals from water.
(2) The process of Fe2+ and Mn2+ adsorption onto IHA/TM conformed to the pseudosecond order kinetic model and Langmuir equation (R2 > 0.99). The experimental thermodynamic results were ΔG < 0, ΔH > 0 and ΔS > 0, indicating that the adsorption of Fe2+ and Mn2+ onto TM tended to be monolayer adsorption, and the adsorption process was spontaneous, endothermic and sustainable.
(3) Through XRD, FTIR, SEM and EDS analysis, it was confirmed that surface coordination and chemical precipitation are the main mechanisms of Fe2+ and Mn2+ removal by IHA/TM.
(4) The study is a laboratory simulation of high-concentration Fe2+ and Mn2+ polluted water, and no actual water samples are used for research. In the future, actual water samples in the field should be used to further study the influence of other coexisting ions in the polluted water on the adsorption of Fe2+ and Mn2+ by IHA/TM composite particles, and dynamic experimental research carried out in order to create conditions for the application of IHA/TM adsorbents in practical engineering.

Author Contributions

Conceptualization, L.L. and T.Z.; methodology, X.Y.; software, L.L., T.Z. and V.M.; validation, X.Y. and T.Z.; formal analysis, L.L. and X.L.; investigation, X.Y. and V.M.; resources, X.L. and L.L.; data curation, L.L. and X.Y.; writing—original draft preparation, L.L. and T.Z.; writing—reviewing and editing, X.L.; visualization, J.M.; supervision, L.L. and X.L.; project administration, L.L.; funding acquisition, L.L. and X.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the support of National Key R&D Program of China (No. 2017YFC1503105), Scientific Research Project of Educational Department of Liaoning Province of China (No. LJKZ0344) and Liaoning Provincial Natural Science Foundation of China ((No. 2019-ZD-0037).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data can be obtained from corresponding authors upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chitaranjan, D.; Ramakar, J.; Desai, V. Rice Husk and Sugarcane Baggase Based Activated Carbon for Iron and Manganese Removal. Aquat. Proc. 2015, 4, 1126–1133. [Google Scholar] [CrossRef]
  2. Akbar, N.; Abdul, A.; Adlan, M. Iron and Manganese Removal from Groundwater Using High Quality Limestone. Appl. Mech. Mater. 2015, 802, 460–465. [Google Scholar] [CrossRef]
  3. Adeyeye, O.; Xiao, C.; Zhang, Z.; Liang, X. State, source and triggering mechanism of iron and manganese pollution in groundwater of Changchun, Northeastern China. Environ. Monit. Assess. 2020, 192, 619–634. [Google Scholar] [CrossRef]
  4. Zhao, F.; Sun, H.; Li, W. Migration of hazardous lements in acid coal mine drainage. J. China Coal Soc. 2007, 3, 261–266. [Google Scholar]
  5. Ren, H.; Zhu, S.; Wang, X.; Liu, Y.; Cao, L. Study on Issues and Countermeasures in Coal Measures Mine Water Resources Exploitation and Utilization. Coal Geol. China 2020, 32, 9–20. [Google Scholar] [CrossRef]
  6. Ahmad, J.; Cheng, W.; Low, W.; Ali, N.; Megat, M. Study on the removal of iron and manganese in groundwater by granular activated carbon. Desalination 2005, 182, 347–353. [Google Scholar] [CrossRef]
  7. Shavandi, M.; Haddadian, Z.; Ismail, M.; Abdullah, N.; Abidin, Z. Removal of Fe(III), Mn(II) and Zn(II) from palm oil mill effluent (POME) by natural zeolite. J. Taiwan Inst. Chem. Eng. 2012, 43, 50–759. [Google Scholar] [CrossRef]
  8. Zhu, C.; Wang, S.; Hu, K.; Wang, W.; Cai, A.; Chang, W.; Li, B. Study on Fluoride, Iron and Manganese Removal from Aqueous Solutions by a Novel Composite Adsorbent. Adv. Mater. Res. 2013, 821–822, 1085–1092. [Google Scholar] [CrossRef]
  9. Li, X.; Yu, X.; Liu, L.; Yang, J.; Liu, S.; Zhang, T. Preparation, characterization serpentine-loaded hydroxyapatite and its simultaneous removal performance for fluoride, iron and manganese. RSC Adv. 2021, 11, 16201–16215. [Google Scholar] [CrossRef]
  10. Tahir, S.S.; Rauf, N. Removal of Fe(II) from the wastewater of a galvanized pipe manufacturing industry by adsorption onto bentonite clay. J. Environ. Manag. 2004, 73, 285–292. [Google Scholar] [CrossRef]
  11. Geremias, R.; Laus, R.; de Fávere, V.T.; Pedrosa, R.C. Rejeito de mineracao de carvao como adsorvente para remocao da acidz, Fe (III), Al (III) E Mn (II) EM drenagen acida. Quím. Nova 2009, 33, 1677–1681. [Google Scholar] [CrossRef]
  12. Apkar’yan, A.S.; Gubaidulina, T.A.; Kaminskaya, O.V. Foam-Glass Ceramic Based Filtering Material for Removing Iron and Manganese from Drinking Water. Glass Ceram. 2015, 71, 413–417. [Google Scholar] [CrossRef]
  13. Aziz, R.A.; Li, C.; Salleh, M. Removal of Iron and Manganese from Palm Oil Mill Effluent (POME) using Activated Clinoptilolite Zeolite. IOP Conf. Ser. Earth Environ. Sci. 2021, 765, 012029. [Google Scholar] [CrossRef]
  14. Dhakal, R.; Ghimire, K.; Inoue, K. Adsorptive separation of heavy metals from an aquatic environment using orange waste. Hydrometallurgy 2005, 79, 182–190. [Google Scholar] [CrossRef]
  15. Wang, M.; Zhang, K.; Wu, M.; Wu, Q.; Liu, J.; Yang, J.; Zhang, J. Unexpectedly High Adsorption Capacity of Esterified Hydroxyapatite for Heavy Metal Removal. Langmuir 2019, 35, 16111–16119. [Google Scholar] [CrossRef] [PubMed]
  16. Kubo, T.; Terutaro, N. Tourmaline group crystals reaction with water. Ferroelectrics 1992, 137, 13–31. [Google Scholar] [CrossRef]
  17. Wang, C.; Liu, J.; Zhang, Z.; Wang, B.; Sun, H. Adsorption of Cd(II), Ni(II), and Zn(II) by Tourmaline at Acidic Conditions: Kinetics, Thermodynamics, and Mechanisms. Ind. Eng. Chem. Res. 2012, 51, 4397–4406. [Google Scholar] [CrossRef]
  18. Leonard, D.; Michael, T.G.R.; Altangerel, A.; Hem, R.P.; Park, C.; Dong, W.K.; Cheol, S.K. Antibacterial and superhydrophilic electrospun polyurethane nanocomposite fibers containing tourmaline nanoparticles. Chem. Eng. J. 2012, 197, 41–48. [Google Scholar] [CrossRef]
  19. Hu, Y.; Yang, X. The surface organic modification of tourmaline powder by span-60 and its composite. Appl. Surf. Sci. 2012, 258, 7540–7545. [Google Scholar] [CrossRef]
  20. Chen, Y.; Wang, S.; Li, Y.; Liu, Y.; Chen, Y.; Wu, Y.; Zhang, J.; Li, H.; Peng, Z.; Xu, R.; et al. Adsorption of Pb(II) by tourmaline-montmorillonite composite in aqueous phase. J. Colloid Interface Sci. 2020, 575, 367–376. [Google Scholar] [CrossRef]
  21. Liao, G.; Zhao, W.; Li, Q.; Pang, Q.; Xu, Z. Novel Poly(acrylic acid)-modified Tourmaline/Silver Composites for Adsorption Removal of Cu(II) ions and Catalytic Reduction of Methylene Blue in Water. Chem. Lett. 2017, 46, 1631–1634. [Google Scholar] [CrossRef]
  22. Molinari, R.; Argurio, P.; Romeo, L. Studies on interactions between membranes (RO and NF) and pollutants (SiO2, NO3, Mn++ and humic acid) in water. Desalination 2001, 138, 271–281. [Google Scholar] [CrossRef]
  23. Wang, H.; Zhu, K.; Wei, X.; Liang, Y.; Lu, X. Desorption enhancement of diesel by humic sodium and surfactants in loess soil. J. Saf. Environ. 2004, 4, 522–551. [Google Scholar]
  24. Zhao, W.; Ren, B.; Andrew, H.; Jiang, F. The adsorption of Mn(II) by insolubilized humic acid. Water Sci. Technol. 2020, 82, 384–396. [Google Scholar] [CrossRef]
  25. Wei, Y.; Ma, M.; Li, S.; Zhao, H.; Shang, Q. Preparation and adsorption performance of chitosan cross linked insolublized humic acid beads. Ion. Exch. Adsorpt. 2016, 32, 43–53. [Google Scholar] [CrossRef]
  26. Jin, X.; Zheng, M.; Sarkar, B.; Naidu, R.; Chen, Z. Characterization of bentonite modified with humic acid for the removal of Cu (II) and 2,4-dichlorophenol from aqueous solution. Appl. Clay Sci. 2016, 134, 89–94. [Google Scholar] [CrossRef]
  27. Laabd, M.; Chafai, H.; Essekri, A.; Elamine, M.; Al-Muhtaseb, S.A.; Lakhmiri, R.; Albourine, A. Single and multi-component adsorption of aromatic acids using an eco-friendly polyaniline-based biocomposite. Sustain. Mater. Technol. 2017, 12, 35–43. [Google Scholar] [CrossRef]
  28. Li, X.; Shang, C. Role of Humic Acid and Quinone Model Compounds in Bromate Reduction by Zerovalent Iron. Environ. Sci. Technol. 2005, 39, 1092–1100. [Google Scholar] [CrossRef]
  29. Gao, B.; Li, P.; Yang, R.; Li, A.; Yang, H. Investigation of multiple adsorption mechanisms for efficient removal of ofloxacin from water using lignin-based adsorbents. Sci. Rep. 2019, 9, 637–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Jiang, K.; Sun, T.; Sun, L.; Li, H. Adsorption characteristics of copper, lead, zinc and cadmium ions by tourmaline. J. Environ. Sci. 2006, 18, 1221–1225. [Google Scholar] [CrossRef]
  31. Elmorsi, R.R.; El-Wakeel, S.T.; El-Dein, W.A.S.; Lotfy, H.R.; Rashwan, W.E.; Nagah, M.; Shaaban, S.A.; Sayed Ahmed, S.A.; El-Sherif, I.Y.; Abou-El-Sherbini, K.S. Adsorption of Methylene Blue and Pb2+ by using acid-activated Posidonia oceanica waste. Sci. Rep. 2019, 9, 3356–3368. [Google Scholar] [CrossRef] [PubMed]
  32. Ding, H.; Lei, T.; Nie, Y.; Ji, H. Characteristics and interactions of heavy metals with humic acid in gold mining area soil at a upstream of a metropolitan drinking water source. J. Geochem. Explor. 2018, 200, 266–275. [Google Scholar] [CrossRef]
  33. Samuel, B.; Zhang, J.; Wang, H.; Yin, H.; Chen, H. The influence of pH, co-existing ions, ionic strength, and temperature on the adsorption and reduction of hexavalent chromium by undissolved humic acid. Chemosphere 2018, 212, 209–218. [Google Scholar] [CrossRef]
  34. Xia, M.; Hu, C.; Zhang, H. Effects of tourmaline addition on the dehydrogenase activity of Rhodopseudomonas palustris. Process Biochem. 2006, 41, 221–225. [Google Scholar] [CrossRef]
  35. Choi, H.; Yu, S. Application of novel hybrid bioadsorbent, tannin/chitosan/sericite, for the removal of Pb(II) toxic ion from aqueous solution. Korean J. Chem. Eng. 2018, 35, 2198–2206. [Google Scholar] [CrossRef]
  36. Bagbi, Y.; Sarswat, A.; Mohan, D.; Pandey, A.; Solanki, P. Lead and Chromium Adsorption from Water using L-Cysteine Functionalized Magnetite (Fe3O4) Nanoparticles. Sci. Rep. 2017, 7, 7672–7687. [Google Scholar] [CrossRef]
  37. Wang, H.; Zhang, P.; Liu, J. Triethylene Tetramine Functionalized Magnetic Graphene Oxide Chitosan Composite with Superior Capacity for the Removal of Phosphate. J. Chem. Eng. Data 2017, 62, 3341–3352. [Google Scholar] [CrossRef]
  38. Nekouei Marnani, N.; Shahbazi, A. A novel environmental-friendly nanobiocomposite synthesis by EDTA and chitosan functionalized magnetic graphene oxide for high removal of Rhodamine B: Adsorption mechanism and separation property. Chemosphere 2019, 218, 715–725. [Google Scholar] [CrossRef]
  39. Duan, C.; Ma, T.; Wang, J.; Zhou, Y. Removal of heavy metals from aqueous solution using carbon-based adsorbents: A review. J. Water Process Eng. 2020, 37, 101339. [Google Scholar] [CrossRef]
  40. Chwastowski, J.; Bradło, D.; Zukowski, W. Adsorption of Cadmium, Manganese and Lead Ions from Aqueous Solutions Using Spent Coffee Grounds and Biochar Produced by Its Pyrolysis in the Fluidized Bed Reactor. Materials 2020, 13, 2782. [Google Scholar] [CrossRef]
  41. Zhang, M.; Yin, Q.; Ji, X.; Wang, F.; Gao, X.; Zhao, M. High and fast adsorption of Cd(II) and Pb(II) ions from aqueous solutions by a waste biomass based hydrogel. Sci. Rep. 2020, 10, 3285–3298. [Google Scholar] [CrossRef]
  42. Anbalagan, K.; Kumar, P.S.; Gayatri, K.S.; Hameed, S.S.; Sindhuja, M.; Prabhakaran, C.; Karthikeyan, R. Removal and recovery of Ni(II) ions from synthetic wastewater using surface modified Strychnos potatorum seeds: Experimental optimization and mechanism. Desalin. Water Treat. 2015, 53, 171–182. [Google Scholar] [CrossRef]
  43. Li, X.; Qi, Y.; Li, Y.; Zhang, Y.; He, X.; Wang, Y. Novel magnetic beads based on sodium alginate gel crosslinked by zirconium(IV) and their effective removal for Pb2+ in aqueous solutions by using a batch and continuous systems. Bioresour. Technol. 2013, 142, 611–619. [Google Scholar] [CrossRef] [PubMed]
  44. Esfandiar, N.; Bahram, N.; Ebadi, T. Removal of Mn(II) from groundwater by sugarcane bagasse and activated carbon (a comparative study): Application of response surface methodology (RSM). J. Ind. Eng. Chem. 2014, 20, 3726–3736. [Google Scholar] [CrossRef]
  45. Fan, X.; Lin, S.; He, S. Study on Adsorption Characteristics of Fe2+ by Attapulgite. Environ. Sci. Manag. 2010, 35, 77–79. [Google Scholar]
  46. Fu, H.; Xie, Q. Complexes of fulvic acid on the surface of hematite, goethite, and akaganeite: FTIR observation. Chemosphere 2006, 63, 403–410. [Google Scholar] [CrossRef]
  47. Wang, D.; Xu, H.; Ma, J.; Lu, X.; Qi, J.; Song, S. Strong promoted catalytic ozonation of atrazine at low temperature using tourmaline as catalyst: Influencing factors, reaction mechanisms and pathways. Chem. Eng. J. 2018, 354, 113–125. [Google Scholar] [CrossRef]
  48. Adhikari, M.; Hazer, G.C. Humus-metal complex-spectral studies. Indian Chem. Soe. 1976, 53, 513–515. [Google Scholar]
  49. Sun, N.; Tian, W.; Zhang, Y.; Tian, J. Adsorption properties of Fe2+ and Mn2+ from groundwater in cold regions by carbonated rice husk. Trans. Chin. Soc. Agric. Eng. 2016, 32, 197–205. [Google Scholar] [CrossRef]
  50. Sangi, M.; Shahmoradi, A.; Zolgharnein, J.; Azimi, G.; Ghorbandoost, M. Removal and recovery of heavy metals from aqueous solution using Ulmus carpinifolia and Fraxinus excelsior tree leaves. J. Hazard. Mater. 2008, 155, 513–522. [Google Scholar] [CrossRef]
Figure 1. Effect of different mixing ratios on the removal of Fe2+ (a) and Mn2+ (b).
Figure 1. Effect of different mixing ratios on the removal of Fe2+ (a) and Mn2+ (b).
Materials 15 03130 g001
Figure 2. Effect of different doses on the removal of Fe2+ (a) and Mn2+ (b).
Figure 2. Effect of different doses on the removal of Fe2+ (a) and Mn2+ (b).
Materials 15 03130 g002
Figure 3. Effect of pH on the removal efficiency.
Figure 3. Effect of pH on the removal efficiency.
Materials 15 03130 g003
Figure 4. Isothermal adsorption of Fe2+ and Mn2+ onto TM and IHA/TM (a); equation fitting curve for the Langmuir model (b), Freundlich model (c) and Temkin model (d).
Figure 4. Isothermal adsorption of Fe2+ and Mn2+ onto TM and IHA/TM (a); equation fitting curve for the Langmuir model (b), Freundlich model (c) and Temkin model (d).
Materials 15 03130 g004
Figure 5. Effects of initial concentrations and contact time on the Fe2+ (a) and Mn2+ (b) adsorption amount of IHA/TM at 25 °C.
Figure 5. Effects of initial concentrations and contact time on the Fe2+ (a) and Mn2+ (b) adsorption amount of IHA/TM at 25 °C.
Materials 15 03130 g005
Figure 6. Quasi-first order kinetic equation fitting curve of IHA/TM adsorption of Fe2+ (a) and Mn2+ (b); quasi-second order kinetic equation fitting curve of IHA/TM adsorption of Fe2+ (c) and Mn2+ (d).
Figure 6. Quasi-first order kinetic equation fitting curve of IHA/TM adsorption of Fe2+ (a) and Mn2+ (b); quasi-second order kinetic equation fitting curve of IHA/TM adsorption of Fe2+ (c) and Mn2+ (d).
Materials 15 03130 g006
Figure 7. Fitting curve of the internal diffusion equation of Fe2+ (a) and Mn2+ (b).
Figure 7. Fitting curve of the internal diffusion equation of Fe2+ (a) and Mn2+ (b).
Materials 15 03130 g007
Figure 8. The influence of the number of adsorption-desorption cycles on the removal effect.
Figure 8. The influence of the number of adsorption-desorption cycles on the removal effect.
Materials 15 03130 g008
Figure 9. XRD pattern of IHA/TM before and after adsorption (a); FTIR spectra of IHA/TM before and after adsorption (b).
Figure 9. XRD pattern of IHA/TM before and after adsorption (a); FTIR spectra of IHA/TM before and after adsorption (b).
Materials 15 03130 g009
Figure 10. SEM images of IHA/TM before and after adsorption of Fe2+ and Mn2+: (a) TM; (b) IHA; (c) IHA/TM; (d) IHA/TM after adsorption of Fe2+ and Mn2+.
Figure 10. SEM images of IHA/TM before and after adsorption of Fe2+ and Mn2+: (a) TM; (b) IHA; (c) IHA/TM; (d) IHA/TM after adsorption of Fe2+ and Mn2+.
Materials 15 03130 g010
Figure 11. EDS surface scan images of IHA/TM composite particles before (a) and after (b) adsorption of Fe2+ and Mn2+.
Figure 11. EDS surface scan images of IHA/TM composite particles before (a) and after (b) adsorption of Fe2+ and Mn2+.
Materials 15 03130 g011
Figure 12. Schematic diagram of the adsorption mechanism of iron and manganese by IHA/TM composite particles.
Figure 12. Schematic diagram of the adsorption mechanism of iron and manganese by IHA/TM composite particles.
Materials 15 03130 g012
Table 1. Fitting results of the isothermal model for IHA/TM adsorption of iron and manganese.
Table 1. Fitting results of the isothermal model for IHA/TM adsorption of iron and manganese.
Ions TypeTemperature (°C)LangmuirFreundlichTemkin
qm mg/gKLR2KF1/nR2lnABR2
Fe2+255.0100.3990.9922.4020.1890.9823.1450.6840.978
355.1750.4590.9912.5970.1720.9833.6350.6630.970
455.3190.5250.9912.8280.1570.9884.3740.6210.963
Mn2+253.4650.5340.9921.5560.2240.9003.7720.4490.964
353.5960.6400.9901.8610.1810.9515.5940.3700.972
453.6590.6850.9902.0540.1550.9537.1690.3180.960
Table 2. Thermodynamic parameters of IHA/TM adsorption of Fe2+ and Mn2+.
Table 2. Thermodynamic parameters of IHA/TM adsorption of Fe2+ and Mn2+.
Ion TypeT (°C)Ce (mg/L)qe (mg/g)ΔG (kJ/mol)ΔH (kJ/mol)ΔS (J/(mol k))
Fe2+250.542.46−3.7433.64138.24
350.472.47−4.12
450.272.58−5.95
Mn2+250.390.97−2.2522.1586.58
350.240.99−3.61
450.231.00−3.89
Table 3. Pseudo-first order and pseudo-second order kinetic fitting results.
Table 3. Pseudo-first order and pseudo-second order kinetic fitting results.
Ions TypeConcentration mg/LQuasi-First Order DynamicsQuasi-Second Order Dynamics
qe mg/gK1 1/minR2qe mg/gK2 mg/g/minR2
Fe2+252.1190.02030.8892.8370.01310.996
402.9730.03030.9652.4620.01320.993
605.1590.02430.9424.9390.00510.997
806.9120.03110.9216.0030.00300.995
1007.8530.02640.9217.4120.00210.991
Mn2+100.5410.01780.9591.0520.06020.999
200.9240.01200.9952.0350.02550.996
301.5430.01250.9912.7210.01310.994
402.4450.01220.9753.2050.00850.995
503.1880.01400.9863.4510.00420.999
Table 4. Internal diffusion model fitting results.
Table 4. Internal diffusion model fitting results.
Ion TypeConcentration mg/LK1d mg/g·min−1/2R12K2d mg/g·min−1/2R22
Fe2+250.0880.7340.0760.952
400.1960.9780.0180.637
600.3170.9250.0380.916
800.3720.9740.0660.674
1000.4830.9700.0390.664
Mn2+100.0320.9910.0240.950
200.0730.9620.0510.972
300.1170.9140.0790.972
400.1350.8150.1260.983
500.2700.9910.1350.997
Table 5. Comparison of the IHA/TM composite adsorbent with previously reported adsorbents for the removal of iron and manganese in terms of uptake capacity.
Table 5. Comparison of the IHA/TM composite adsorbent with previously reported adsorbents for the removal of iron and manganese in terms of uptake capacity.
AdsorbentpHqm of Fe2+(mg/g)qm of Mn2+(mg/g)Reference
Sugarcane bagasse4.5 0.676[44]
Graptolite60.352 [45]
Limestone6.2–6.70.030.007[2]
Granular activated carbon73.6012.545[6]
Slovakian natural zeolite71.1570.075[7]
Natural shells7.0–9.04.003.50[8]
IHA/TM composite granules6.05.6453.574This study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, L.; Zhang, T.; Yu, X.; Mkandawire, V.; Ma, J.; Li, X. Removal of Fe2+ and Mn2+ from Polluted Groundwater by Insoluble Humic Acid/Tourmaline Composite Particles. Materials 2022, 15, 3130. https://doi.org/10.3390/ma15093130

AMA Style

Liu L, Zhang T, Yu X, Mkandawire V, Ma J, Li X. Removal of Fe2+ and Mn2+ from Polluted Groundwater by Insoluble Humic Acid/Tourmaline Composite Particles. Materials. 2022; 15(9):3130. https://doi.org/10.3390/ma15093130

Chicago/Turabian Style

Liu, Ling, Tianyi Zhang, Xiaowan Yu, Vitumbiko Mkandawire, Jiadi Ma, and Xilin Li. 2022. "Removal of Fe2+ and Mn2+ from Polluted Groundwater by Insoluble Humic Acid/Tourmaline Composite Particles" Materials 15, no. 9: 3130. https://doi.org/10.3390/ma15093130

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