Next Article in Journal
Insights into the Contribution of Oxidation-Reduction Pretreatment for Mn0.2Zr0.8O2−δ Catalyst of CO Oxidation Reaction
Previous Article in Journal
Nonlinear Finite Element Model for Bending Analysis of Functionally-Graded Porous Circular/Annular Micro-Plates under Thermomechanical Loads Using Quasi-3D Reddy Third-Order Plate Theory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of MCS from Low-Grade Bauxite Desilication Lye and Adsorption of Heavy Metals

1
College of Materials and Metallurgy, Guizhou University, Guiyang 550025, China
2
Guizhou Province Key Laboratory of Metallurgical Engineering and Process Energy Saving, Guiyang 550025, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(9), 3506; https://doi.org/10.3390/ma16093506
Submission received: 26 March 2023 / Revised: 23 April 2023 / Accepted: 24 April 2023 / Published: 2 May 2023

Abstract

:
By utilizing low-grade bauxite desilication solution as raw material and adding lime after thermal reaction, adsorbent MCS was synthesized. X-ray diffraction, Brunauer–Emmett–Teller, Fourier transform infrared spectroscopy, and scanning electron microscopy were used to characterize the MCS, MCS-Pb, and MCS-Cu. The Freundlich model was found to be more suitable for isothermal adsorption, suggesting that the adsorption of Cu2+ and Pb2+ by MCS is not limited to monolayer adsorption. According to the results of the experiment, the maximum adsorption capacities of lead ion and copper ion were found to be Pb2+ (1921.506 mg/g) > Cu2+ (561.885 mg/g), and the adsorption was controlled by chemical reactions following pseudo-second-order kinetics. Electrolyte study results indicated that the presence of background electrolyte did not affect the adsorption of Cu2+ and Pb2+ by MCS.

1. Introduction

The main sources of heavy metal ions in industrial wastewater are smelting engineering, electroplating, and electrolytic processes. Water containing heavy metals is difficult to biodegrade, easily accumulates, and can cause serious toxicity reactions [1]. Therefore, the effective removal of heavy metal ions from water has become an urgent issue in the present day. Meanwhile, researchers have conducted extensive studies on the adsorption of heavy metal ions, resulting in an increasing interest in determining how to solve this problem efficiently and economically.
There are many traditional methods for treating heavy metal wastewater, including chemical precipitation, chelate flocculation, ferrite, ion exchange, membrane separation, electrochemistry, flotation, and adsorption. Due to its low cost and environmental friendliness, adsorption is one of the most used methods of removing heavy metals from solutions. Abukhadra [2] prepared kaolinite nanotubes for heavy metal adsorption using kaolinite as a precursor and demonstrated that the process not only involved chemical exchange and ion exchange, but also multilayer adsorption. Furthermore, the adsorbent is reusability-friendly and can be used in the future. Sharma [3] synthesized mesoporous zinc oxide and titanium dioxide using a nanoporous casting technique. These adsorbents are effective at adsorbing Pb2+ and Cd2+ from the solution. Ghiloufi [4] also obtained similar results. According to the Italian wastewater regulations, Di [5] used silica-supported hydrophilic carbon nanoparticles to adsorb heavy metals Cb2+, Ni2+, and Pb2+ in the solution with high adsorption rates. Li [6] prepared a polymer composite by utilizing a variety of solid wastes for the adsorption of heavy metals Pb2+ and Cd2+. During the synthesis of the composite, calcium silicate was formed and contributed to improved mechanical properties. Czech et al. [7] used calcium silicate for the adsorption of Cd2+ heavy metal, which provided advantages such as good adsorption capacity, a broad range of raw materials, and economic feasibility. A significant improvement in the phosphorus removal rate was observed when free Ca2+ on the surface of calcium silicate was washed away by deionized water in a study by Okano et al. [8]. Salem et al. [9] used nanoscale ZnO as an adsorbent for heavy metal ions, which was later used to adsorb acidic dye Acid Blue 92 (AB92) in aqueous solutions. Shao et al. [10] synthesized calcium silicate for heavy metal adsorption, which was subsequently used in the degradation of MB as a photocatalyst. As a result of this experiment path, solid waste adsorbents are utilized for heavy metal adsorption, followed by their use as catalysts. This provides a new reference for the treatment of adsorbents.
As a summary, synthesis of mesoporous calcium silicate (MCS) from low-quality bauxite ore using desilication liquid can effectively and economically remove silica and regenerate the alkali solution used in desilication. The MCS has excellent performance in adsorbing metal ions from water and is widely available, easily synthesized, inexpensive, and has many potential applications.

2. Experimental

2.1. Materials

Low-grade bauxite and limestone were obtained from an alumina plant in Guizhou. The limestone was calcined at 1000 °C for 3 h prior to use to increase its reactivity (effective CaO content was approximately 80%). Nitric acid (HNO3, purity ≥ 99 %), copper nitrate trihydrate (Cu(NO3)2 3H2O, purity ≥ 99%), lead nitrate hexahydrate (Pb(NO3)2·6H2O, purity ≥ 99%), sodium nitrate (NaNO3, purity ≥ 98%), and sodium hydroxide (NaOH, purity ≥ 98%) were purchased from Aladdin Chemistry Co. Ltd. (Shanghai, China). Deionized water was prepared in-house at Guizhou University, and its conductivity was less than 1 μS/cm.

2.2. Synthesis of MCS

Initially, low-grade bauxite was reacted with 110 g/L caustic soda solution at a liquid-to-solid ratio of 10 under 95 °C for 30 min, and a silicon-containing alkali solution was obtained. After that, MCS was prepared by hydrothermal reaction of silicon-containing alkali solution with lime-to-calcium-silicon ratio of 2 at 95 °C for 120 min, followed by filtration, washing, and drying.
The XRD and SEM patterns of the prepared MCS are shown in Figures 2 and 4a. It can be seen from Figure 2 that the peak of MCS was narrow, indicating that its grain size was significant, and it mainly consisted of calcium silicate gel, calcium hydroxide Ca(OH)2, and other phases. The porous structure of MCS is shown in Figure 4a, which can provide more adsorption sites for the adsorption process.

2.3. Analytical Method

The overall surface morphology was investigated using a BrukerAXS D8 Advance X-ray diffractometer, a scanning electron microscope, and energy dispersive X-ray spectroscopy (TESCAN MIRA LMS, Brno, Czech Republic). Brunauer–Emmett–Teller (BET) surface analysis was carried out using a Micromeritics Tristar II 3020 instrument (Mac ASAP2460, MICROMERITICS, Atlanta, GA, USA), while changes in the chemical bonds of MCS were identified using Fourier transform infrared spectroscopy (FT-IR) (Nicolet 670, Nicolite, Madison, WI, USA). A spectrophotometer (PERSEE A3G, Beijing, China) was used to detect heavy metal ions in the solution, and a pH meter (PHS = 3E, Shanghai, China) was used to monitor pH changes.

2.4. Adsorption Experiment

MCS was used to adsorb copper or lead ions. The experimental process was as follows: different NaOH and HNO3 solution concentrations were used to adjust the pH of Cu2+ and Pb2+ solutions with 200 mg/L concentrations. Then, 0.0040 g, 0.0080 g, 0.0120 g, 0.0160 g, 0.0240 g of MCS and 20 mL of Cu2+ and Pb2+ solutions were added at a concentration of 200 mg/L, pH = 2.00~6.00, respectively, to a 100 mL centrifuge tube. After 2 min of ultrasound with a high-power CNC ultrasonic cleaner, the oscillations were carried out in a water bath thermostatic oscillator at 303 K for 24 h. As a result of the shaking, the samples were collected and then filtered through a 0.22 m pinhole filter using a 5 mL syringe. Through the use of a stoppered colorimetric tube, the diluted samples were volumetrically determined using atomic absorption spectrometry, and then the Cu2+ and Pb2+ concentrations were determined using atomic absorption spectrometry to determine the removal rates and adsorption capacities of Cu2+ and Pb2+.
The determination results were substituted into (1) and (2) to calculate the adsorption capacity and removal rate. The adsorption capacity (q, mg∙g−1) and removal efficiency (R, %) of the adsorbent were calculated by the following formula:
q = C 0 C e v m
R = C 0 C e C 0 × 100   %
C0 and Ce represent the concentration of adsorbed substance under initial and equilibrium states, respectively, mg/L; m represents the mass of the adsorbent, g; v is the volume of the adsorbent solution, L.

2.5. Isothermal Adsorption Model

The adsorption isotherm is a graph plotting the amount of adsorbent removed from the system solution against the amount remaining in the solution at isothermal conditions. The adsorption isotherm allows for the study of the proportion of adsorbent molecules distributed between the solid and liquid phases at equilibrium [11,12].

2.5.1. Langmuir Model

According to the law of mass action and the characteristic of equal adsorption and desorption rates at adsorption equilibrium, the isothermal equation for adsorption is derived as [11,12]:
C e q e = 1 q m K L + C e q m
where Ce: liquid phase concentration at adsorption equilibrium, mg/L; C0: initial concentration of liquid phase, mg/L; qe: adsorption capacity at equilibrium, mg/g; KL: Langmuir adsorption isothermal equation coefficient, L/mg; qm: maximum adsorption capacity of adsorbent per unit mass, mg/g; RL: Langmuir adsorption equilibrium constant.
Using Ce/qe as a graph of Ce, the adsorption capacity of qmax and KL is generally related to the specific surface area and porosity of the adsorbent based on the slope and intercept of the line. When the adsorbent has a high specific surface area and pore, it generally has a high adsorption capacity [13]. Langmuir’s adsorption isothermal model is suitable for uniform adsorption, in which each molecule on the surface of the adsorbent has the same adsorption activation energy [14].
Adsorption that conforms to the Langmuir isotherm is chemisorption. It is generally believed that the activation energy of chemisorption is in the range of 40–400 kJ/mol and that, except in special circumstances, a spontaneous chemisorption process, which should be exothermic, will decrease in saturation adsorption with increasing isotherm [15]. KL is the equilibrium constant for adsorption, also known as the adsorption coefficient. Its magnitude is related to the adsorbent, the nature of the adsorbent, and the temperature; the larger the KL value, the stronger the adsorption capacity, and KL has the magnitude of the inverse of the concentration [16,17].

2.5.2. Freundlich Model

The Freundlich adsorption isotherm describes the heterogeneous system, namely reversible adsorption. This isotherm is not limited to single-layer adsorption, and its equation is [18,19]:
qe = KFCe1/n
where qe: the mass of adsorbent adsorbed by a unit adsorbent at equilibrium, mg/g; Ce: adsorption mass equilibrium concentration, mg/L; KF: Freundlich adsorption isothermal equation equilibrium constant, L/g; 1/n: Freundlich adsorption isothermal equation equilibrium constant.
It is generally believed that the value of 1/n is between 0 and 1, and its value indicates the strength of the influence of concentration on adsorption capacity. The smaller the 1/n, the better the adsorption performance [20]. Therefore, 1/n in 0.1~0.5, that is, n between 2~10, it is easy to adsorb; when 1/n is greater than 2, that is, n is between 0 and 2, it is difficult to adsorb [21].

2.6. Adsorption Kinetics

Models of adsorption kinetics are used in order to study the mechanism of diffusion between the adsorbent and adsorbate, the rapid control steps in adsorption, and the factors influencing the adsorption rate, which are closely related to the reaction time. The rate constants in adsorption kinetic models are frequently used to describe the adsorption rate. Frequently used adsorption kinetic models include pseudo primary adsorption kinetic models and pseudo secondary adsorption kinetic models [22].
(1)
Pseudo-first-order model adsorption kinetics model.
The differential form of the pseudo-first-order model dynamic model equation is as follows [23]:
dq t dt = k 1 q e q t
The boundary conditions of the integral are: when t = 0, qe = 0; when t = t, qt = qt.
Where qe refers to the adsorption amount when the adsorption equilibrium is reached; qt refers to the adsorption amount at any time during the adsorption process; t is the adsorption time, min; k1 is the adsorption rate constant of the pseudo-first-order model adsorption kinetic model [24,25].
(2)
Modified pseudo-first-order model dynamics: The modified pseudo-first-order model kinetic model is obtained by transforming the rate constant in the Equation (6). K1(min−1) is the rate constant of the modified equation of the pseudo-first-order model dynamics model [26].
When t ∈ (0,t), integrate the above equation, and in this range, the adsorption concentration increases from 0 to qt, and the following equation can be obtained:
q t q e + ln ( q e q t ) = ln ( q e ) k 1 t
If the adsorption process follows the modified first-order kinetic model equation, then the rate constant K1 and adsorption quantity qe can be obtained by drawing a line with (qt − qe) + ln(qe − qt) to t [27].
(3)
Pseudo-second-order kinetic model [28,29].
The equation of the pseudo-second-order dynamics model is expressed as follows:
t q t = 1 k 2 q e 2 + 1 q e t
In the above equation, qe is the equilibrium adsorption amount, and k2 is the equation constant of the pseudo-second-order kinetic model. Using t/q as a line against t, qe and k2 can be found from the slope and intercept of the line on the graph, and {τ = k2qe2} can be found [30,31].

3. Results

3.1. Effect of Operation Parameters

As shown in Figure 1, the effect of MCS addition and initial pH of the solution on the adsorption of Cu2+ and Pb2+ was investigated. The experimental procedure was as follows.
In order to adjust the pH of the 200 mg/L Cu2+ and Pb2+ solutions to 2.00, 3.00, 4.00, 5.00, and 6.00, different concentrations of NaOH and HNO3 solutions were used, respectively. First, 0.0040 g, 0.0080 g, 0.0120 g, 0.0160 g, and 0.0240 g of porous calcium silicate adsorbent was added to each of the five groups of six 100 mL centrifuge tubes (30 in total), and 0.0200 g, 0.0240 g of MCS adsorbent and 20 mL of Cu2+ solution with a concentration of 200 mg/L, pH = 2.00~6.00 to were added each group, respectively. After being sonicated for two minutes using a high-power CNC ultrasonic cleaner, the samples were shaken for 24 h at 303 K in a water bath. In this experiment, samples were collected after the shaking time had elapsed, the pH of each sample was measured, and the solution was then taken using a 5 mL syringe and filtered through a 0.22 μm pinhole filter. Filtered samples were diluted and volumetrically determined using a stoppered colorimetric tube. By atomic absorption spectrometry, the concentration of Cu2+ in the samples was determined in order to determine the removal rate and adsorption capacity of the heavy metal ions Cu2+.
It can be seen from Figure 1a that the removal rate of heavy metal ions Cu2+ increased with the addition of MCS at pH = 2. The effect of increasing the removal rate was slow when the addition amount was between 0.2 and 1.0 g/L. When the pH = 3, 4, 5, and 6, the removal rate of heavy metal ions Cu2+ increased with the addition of 0.2~0.6 g/L. After the addition of 0.6 g/L, the removal rate tended to be stable. As can be seen from Figure 1b, the removal rate of heavy metal ions Pb2+ did not change significantly with the addition of porous calcium silicate at pH = 2; at pH = 3, the removal rate of Pb2+ increased with the addition of porous calcium silicate up to 0.3 g/L, and then stabilized, and the trend was most obvious between 0.15 and 0.25 g/L. At pH = 4, 5, and 6, the removal rate of Pb2+ increased with the addition of porous calcium silicate up to 0.15 g/L, and then stabilized. At pH = 4, 5, and 6, the removal rate of Pb2+ increased with the addition of porous calcium silicate until 0.15 g/L and then stabilized. When the pH was greater than 2, the best adsorbents of MCS for Cu2+ and Pb2+ were 0.6 g/L and 0.15 g/L, respectively, at which time Cu2+ and Pb2+ in solution were almost completely adsorbed by MCS. It can also be seen that the affinity of MCS for Pb2+ was greater than that for Cu2+.
In Figure 1c, when pH = 2, the adsorption capacity of heavy metal ion Cu2+ increased with the increase of MCS addition, and the adsorption capacity increased slowly when the addition amount was between 0.2 and 1.0 g/L. However, after 1.0, the increase effect had an obvious change; when pH = 3, the adsorption capacity increased with the increase of porous calcium silicate addition amount until 0.6. At pH = 4, 5 and 6, the adsorption capacity increased with the addition of porous calcium silicate up to 0.4 g/L, but decreased with the addition of MCS after 0.4 g/L. In Figure 1d, the adsorption capacity of heavy metal ion Pb2+ did not change significantly with the increase of MCS addition at pH = 2; at pH = 3, the adsorption capacity increased with the increase of MCS addition up to 0.2 g/L, and the trend of increase was most obvious between 0.15 and 0.2 g/L, but decreased with the increase of MCS addition after 0.2 g/L. At pH = 3, the adsorption capacity increased with the increase of MCS addition up to 0.2 g/L, but decreased with the increase of MCS addition after 0.2 g/L. At pH = 4 and 5, the adsorption capacity increased with the addition of porous calcium silicate up to 0.15 g/L, but decreased with the addition of MCS after 0.15 g/L. At pH = 6, the adsorption capacity of Pb2+ decreased gradually with the addition of MCS. It can be seen from the graph that the adsorption capacity of MCS for Pb2+ was greater than that for Cu2+.
We conclude that H+ in solution inhibits the adsorption of Cu2+ by MCS. Despite the fact that the equilibrium pH in solution increased after adsorption equilibrium in Figure 1e,f, this was likely due to the exchange of -OH with heavy metal ions on the surface of MCS during the adsorption process, which allowed -OH to enter the solution and increase the pH value. It also explains why when the pH was low, the H+ in solution reacted preferentially with -OH, resulting in a lower removal of heavy metal ions.

3.2. Characterization of MCS

Figure 2 shows the XRD patterns of MCS. As seen in Figure 1a, the peaks of MCS were narrow and mainly consisted of Ca3SiO5, Ca(OH)2, and CaCO3 phases. CaCO3 was formed by the carbonation of Ca(OH)2 that absorbed CO2 from the air. The Figure 2b illustrates the XRD pattern of MCS after the adsorption of copper ions. We observed only the peak of Ca3SiO5, and the peak intensity has decreased. This can be attributed to the chemical adsorption of Cu2+ on MCS, which reduced the crystallinity of calcium silicate. The calcium silicate did not completely react and adapted to adsorb on the surface of the MCS, preventing any further chemical adsorption reactions.
As shown in Figure 2c, MCS after the adsorption of lead ions exhibited an XRD pattern. Based on the phase analysis in the Figure 2c, the main components were 2PbCO3·Pb(OH)2, Pb3(CO2)2 (OH)2, and Ca3SiO5, while the peak of calcium silicate almost disappeared. This indicates that chemical adsorption between MCS and Pb2+ was more intense. The peak of calcium silicate at 2θ = 29.43° had a stronger intensity for (a) > (b) > (c), and the adsorption capacity of calcium silicate for Cu2+ and Pb2+ was the same.
Figure 3a shows the sample exhibits characteristic absorption peaks at approximately 3440, 1630, 1440, 966, 664, and 471 cm−1. These peaks corresponded to functional groups as follows: the peak at approximately 3440 cm−1 corresponded to the asymmetric stretching vibration of water molecules and surface -OH groups. A peak at around 1630 cm−1 was caused by the vibration of adsorbed water molecules and surface -OH, indicating the presence of abundant reaction groups -OH on the sample surface. The peak around 1440 cm−1 was attributed to the calcium silicate peak, while 966 cm−1 was attributed to the symmetric stretching vibration of Si-O-Si. The characteristic peaks at 664 cm−1 and 471 cm−1 were caused by the deformation of Si-O-Si. From Figure 3b,c, it can be observed that the peaks of MCS FT-IR at 966 cm−1, 664 cm−1, and 471 cm−1 disappeared after the adsorption of copper ions and lead ions. Chemical adsorption resulted in the fracture of Si-O-Si and the disappearance of the characteristic peak due to chemical adsorption [32].
We examined the morphology of MCS before and after the adsorption of copper and lead heavy metal ions using scanning electron microscopy. The adsorbed MCS was analyzed using a scanning electron microscope in combination with an energy-dispersive X-ray spectrometer, as shown in Figure 4. The SEM image of MCS is shown in Figure 4a, from which it can be seen that MCS was fluffy and porous, which provided a large number of adsorption sites for the adsorption of heavy metals. Figure 4b shows MCS after the adsorption of Cu2+. It can be seen that the surface of MCS was covered by a layer of particles, and the thickness of MCS increased, indicating that copper ions were adsorbed on the surface and pores of MCS, making the porous calcium silicate no longer fluffy. This is confirmed by the energy-spectrum distribution map in Figure 4f. Figure 4c shows the morphology of MCS after adsorption of lead ions. The morphology of MCS changed from a fluffy porous state to a dense flaky state, indicating that the adsorption of Pb2+ is not simply a physical adsorption process [33].
In Figure 5, the nitrogen adsorption-desorption isotherms and pore size distribution curves are shown before and after the adsorption of copper and lead heavy metal ions by MCS. The specific surface area and the pore size of the samples were significantly reduced as a result of the adsorption of copper and lead ions by MCS. The specific surface area and pore size of MCS were calculated by the BET method and Barrett–Joyner–Halend method, which were 69.92 m2/g and 38.6 nm, respectively, while the specific surface area and average pore size of MCS after adsorption of copper and lead ions were 42.45 m2/g, 30.4 nm, 20.49 m2/g, and 28.8 nm, respectively. The adsorption of older heavy gold ions on the surface and in the pores of MCS, as shown in Figure 3, resulted in a reduction of both the specific surface area and pore size.

3.3. Isothermal Adsorption Mode

The adsorption isotherms of Cu2+ and Pb2+ (293 K, 303 K and 313 K) at different temperatures are shown in Figure 6. It was found that the adsorption capacity of MCS for Cu2+ increased continuously with the increase of initial concentration, and that it increased rapidly when the initial concentration of Cu2+ was below 200 mg/L. The Cu2+ concentration continued to increase, but the qe increased slowly. When the initial concentration of Pb2+ was lower than 20 mg/L, the adsorption capacity of MCS increased rapidly with the increase of the initial concentration of Pb2+, and when the concentration of Pb2+ was greater than 200 mg/L, the adsorption capacity almost ceased to change. Meanwhile, it is evident from Figure 5 that the adsorption capacity qe of MCS for heavy metal ions Cu2+ and Pb2+ increased with the increase of temperature. Therefore, the increase in temperature is beneficial for the adsorption of heavy metal ions Cu2+ and Pb2+ on MCS.
Table 1 and Table 2 show the fitted parameters of Langmuir and Freundlich for MCS to heavy metal ions Cu2+ and Pb2+, respectively. The correlation coefficient (R2) fitted by the Freundlich model was higher than that of Langmuir. This result indicates that the description of the adsorption of heavy metal ions by MCS is more consistent with the Freundlich model, suggesting that the adsorption process is not limited to monolayer adsorption.
Based on the Freundlich model, the maximum adsorption capacity qe of MCS for Cu2+ and Pb2+ was 561.885 mg/g and 1921.506 mg/g, respectively. This indicates that the affinity of MCS for lead ions was much greater than that for copper ions.

3.4. Adsorption Kinetics Study

The effect of mesoporous calcium polyacid on the adsorption capacity at different adsorption times for heavy metal ions with initial concentrations of 200 mg/L, 400 mg/L and 600 mg/L; the variation curves with contact time; the fitted first-order and second-order fitted lines; and the fitted parameters of the obtained copper and lead adsorption capacities are shown in Figure 7. From Figure 7a, it can be seen that the adsorption times for MCS to reach adsorption equilibrium for Cu2+ initial concentrations of 200 mg/L, 400 mg/L, and 600 mg/L were 30 min, 90 min, and 360 min, respectively.
As shown in Figure 7b, the adsorption time for MCS to reach the adsorption equilibrium for Pb2+ initial concentrations of 200 mg/L, 400 mg/L, and 600 mg/L, respectively, was 360 min. There was a linear relationship between the adsorption capacities of Cu2+ and Pb2+ in the same period of time: 600 mg/L > 400 mg/L > 200 mg/L. This indicates that the adsorption capacity of MCS for Cu2+ and Pb2+ at the same adsorption time is related to the concentration of Cu2+ and Pb2+ in the original solution. Pb2+ is more easily adsorbed by MCS than Cu2+, and this is due to the high surface area of the adsorbent and the distinct adsorption sites with Cu2+ as opposed to Pb2+, which form complexes more easily.
The correlation coefficients for the kinetics of Pb2+ and Cu2+ adsorption by MCS are shown in Table 3, and the fit is better when the heavy metal ion concentration is 200 mg/L. Therefore, the corresponding parameters are shown in Table 3. The fit coefficients for the pseudo first-order and pseudo second-order kinetic models indicate that the pseudo second-order model is a better fit and is the best model for describing the kinetics of Cu and Pb ion adsorption. They also indicate that the mesoporous calcium silicate for Cu2+ and Pb2+ adsorption is mainly chemisorption. During the adsorption process, heavy metal ions form chemical bonds by sharing or exchanging electrons and attach to the MCS surface, i.e., surface complexation.

3.5. Effect of Ionic Strength on Adsorbent

To study the effect of solution ionic strength on the adsorption process, this study used sodium nitrate as the background electrolyte to investigate the adsorption performance of MCS adsorbent on Pb2+ and Cu2+ in different concentrations of NaNO3 solution. In this experiment, NaNO3 was used at concentrations of 0.001 mol/L, 0.01 mol/L, 0.05 mol/L, 0.2 mol/L, and 0.2 mol/L, respectively. Then, the MCS adsorbent was added to study the effect on the removal rate of Pb2+ and Cu2+.
As shown in Figure 8, Na+ has no effect on the adsorption of Pb2+ and Cu2+ by MCS; thus, it is clear that the use of sodium nitrate as a background electrolyte does not affect the adsorption performance of MCS. The fact that MCS does not specifically adsorb sodium and nitrate ions does not affect the results of the experiment, i.e., sodium nitrate is an irrelevant electrolyte.

3.6. Elution, Recycling, and Recovery of Adsorbed Heavy Metals

Table 4 shows the elution and recovery rates of adsorbed Cu2+ and Pb2+ on MCS. Using 0.1 mol hydrochloric acid solution as the eluent, the experimental results are shown in Table 4. The elution rates of MCS-Cu and MCS-Pb using 0.1 mol hydrochloric acid as the eluent were both high and almost completely removed. This is because the lower pH conditions resulted in the dissolution of the MCS, allowing the heavy metals adsorbed on the MCS to dissolve in solution. Meanwhile, the heavy metal ions in solution were recovered almost completely using sodium hydroxide as a precipitant.

4. Discussion

Although China has abundant bauxite reserves, it is mainly bauxite of diaspore and kaolinite type, which is generally characterized by high aluminum, high silicon, a low aluminum-silicon ratio, high impurities, and miscellaneous components. Moreover, the annual consumption of resources is also enormous. With the increasing demand for bauxite and the depletion of domestic high-grade bauxite resources, the key to solving the strategic demand for bauxite is to improve the Al-Si ratio of low-grade bauxite and make it suitable for Bayer process production by pre-desilication. However, a large amount of desilication lye will be produced in the desilication process, and a large amount of slag will be produced during the desilication process, which will increase the production of red mud, which not only increases the cost but also causes a significant burden on the environment.
According to the above experimental conclusions, the synthesis of MCS from low-grade bauxite desilication liquid has an excellent adsorption effect on copper and lead ions. Using low-grade bauxite desilication liquid to prepare MCS can reduce the increase of solid waste produced during the mother liquor cycle, which is helpful to increase the economic benefit of bauxite. MCS is used to adsorb heavy metals in waste liquid to achieve the effect of abolishing, which provides a new idea for improving the process of the bauxite industry. Therefore, preparing MCS using low-grade bauxite desilication liquid has a good application prospect, which is more green and offers better environmental protection.

5. Conclusions

In this experiment, MCS was prepared by the hydrothermal reaction of added lime with a low-grade bauxite desilication solution and used as an adsorbent to adsorb copper and lead ions from wastewater. The main findings are as follows.
  • The MCS had high adsorption capacity for Cu2+ and Pb2+ when the adsorption solution was pH = 4~6, with Pb2+ (1921.506 mg/g) > Cu2+ (561.885 mg/g). It was found that the lower pH inhibited the adsorption of heavy metal ions from the solution by the adsorbent. The final adsorption conditions for Cu2+ and Pb2+ were pH = 4, with an adsorption temperature of 303 K. The adsorption rates of Cu2+ and Pb2+ after 24 h were 99.51% and 99.74%, respectively.
  • The kinetic adsorption of Cu and Pb ions by MCS was calculated to be more in line with the quasi-secondary kinetic model. The adsorption capacity of Cu2+ and Pb2+ was related to the initial solution concentration, which was 600 mg/L > 400 mg/L > 200 mg/L. The results showed that the adsorption of Cu2+ and Pb2+ by MCS was mainly chemisorption. During the adsorption process, heavy metal ions form chemical bonds by sharing or exchanging electrons and attach to the MCS surface.
  • Based on the calculations of the isothermal adsorption models for Cu2+ and Pb2+ by MCS, it was proposed that the adsorption of Cu2+ and Pb2+ by MCS was not limited to monolayer adsorption. It has been demonstrated that an increase in the adsorption temperature promotes the adsorption of Cu2+ and Pb2+ from MCS with a maximum adsorption capacity of qm(Pb) > qm(Cu). According to the Freundlich isotherm, the n value for Pb is higher than that for Cu, indicating that the adsorption of Pb2+ by MCS is greater than that of Cu2+.
  • By adjusting the background electrolyte NaNO3 concentration, it was found that the background electrolyte had no effect on the over-sorption process during the adsorption of Cu2+ and Pb2+ by MCS.

Author Contributions

Conceptualization, C.C. (Chaoyi Chen) and J.L.; methodology, formal analysis, C.C. (Cheng Chen) and X.L.; validation, G.W.; writing original draft preparation, D.N. and C.C. (Cheng Chen); investigation, G.W.; resources, project administration, funding acquisition, C.C. (Chaoyi Chen) and J.L; writing—review and editing, C.C. (Cheng Chen) and D.N.; supervision, data curation, G.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China (52164017, 52074096, 52274260), Guizhou Graduate Research Foundation YJSKYJJ [2021]001, Guizhou Science and Technology Cooperation Platform [2021]086.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data supporting this study’s findings are included within the article.

Acknowledgments

The authors gratefully acknowledge projects funded by the National Natural Science Foundation of China (52164017, 52074096, 52274260), the Guizhou Provincial Postgraduate Research Fund (YJSKYJJ[2021]001) and the Guizhou Provincial Science and Technology Cooperation Platform ([2021]086).

Conflicts of Interest

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

References

  1. Elsherbiny, A.S.; El-Hefnawy, M.E.; Gemeay, A.H. Adsorption Efficiency of Polyaspartate-Montmorillonite Composite towards the Removal of Pb(II) and Cd(II) from Aqueous Solution. J. Polym. Environ. 2017, 26, 411–422. [Google Scholar] [CrossRef]
  2. Abukhadra, M.R.; Bakry, B.M.; Adlii, A.; Yakout, S.M.; El-Zaidy, M.E. Facile conversion of kaolinite into clay nanotubes (KNTs) of enhanced adsorption properties for toxic heavy metals (Zn2+, Cd2+, Pb2+, and Cr6+) from water. J. Hazard. Mater. 2019, 374, 296–308. [Google Scholar] [CrossRef] [PubMed]
  3. Sharma, M.; Singh, J.; Hazra, S.; Basu, S. Adsorption of heavy metal ions by mesoporous ZnO and TiO2@ZnO monoliths: Adsorption and kinetic studies. Microchem. J. 2019, 145, 105–112. [Google Scholar] [CrossRef]
  4. Ghiloufi, I.; Ghoul, J.E.; Modwi, A.; Mir, L.E. Ga-doped ZnO for adsorption of heavy metals from aqueous solution. Mat. Sci. Semicond. Proc. 2016, 42, 102–106. [Google Scholar] [CrossRef]
  5. Di Natale, F.; Gargiulo, V.; Alfe, M. Adsorption of heavy metals on silica-supported hydrophilic carbonaceous nanoparticles (SHNPs). J. Hazard. Mater. 2020, 393, 122374–122386. [Google Scholar] [CrossRef]
  6. Li, J.; Li, J.X.; Wei, H.; Yang, X.D.; Benoit, G.; Jiao, X.K. Alkaline-thermal activated electrolytic manganese residue-based geopolymers for efficient immobilization of heavy metals. Constr. Build. Mater. 2021, 298, 123853–123862. [Google Scholar] [CrossRef]
  7. Czech, B.; Hojamberdiev, M.; Bogusz, A. Impact of thermal treatment of calcium silicate-rich slag on the removal of cadmium from aqueous solution. J. Clean. Prod. 2018, 200, 369–379. [Google Scholar] [CrossRef]
  8. Okano, K.; Uemoto, M.; Kagami, J.; Miura, K.; Aketo, T.; Toda, M.; Honda, K.; Ohtake, H. Novel technique for phosphorus recovery from aqueous solutions using amorphous calcium silicate hydrates (A-CSHs). Water Res. 2013, 47, 2251–2259. [Google Scholar] [CrossRef]
  9. Salem, I.A.; Salem, M.A.; El-Ghobashy, M.A. The dual role of ZnO nanoparticles for efficient capture of heavy metals and Acid blue 92 from water. J. Mol. Liq. 2017, 248, 527–538. [Google Scholar] [CrossRef]
  10. Shao, N.N.; Tang, S.Q.; Liu, Z.; Li, L.; Yan, F.; Liu, F.; Li, S.; Zhang, Z.T. Hierarchically Structured Calcium Silicate Hydrate-Based Nanocomposites Derived from Steel Slag for Highly Efficient Heavy Metal Removal from Wastewater. ACS Sustain. Chem. Eng. 2018, 6, 14926–14935. [Google Scholar] [CrossRef]
  11. Song, X.R.; Chai, Z.H.; Zhu, Y.; Li, C.L.; Liang, X.Q. Preparation and characterization of magnetic chitosan-modified diatomite for the removal of gallic acid and caffeic acid from sugar solution. Carbohydr. Polym. 2019, 219, 316–327. [Google Scholar] [CrossRef]
  12. Yang, W.J.; Ding, P.; Lei, Z.; Yu, J.G.; Chen, X.Q.; Jiao, F.P. Preparation of diamine modified mesoporous silica on multi-walled carbon nanotubes for the adsorption of heavy metals in aqueous solution. Appl. Surf. Sci. 2013, 282, 38–45. [Google Scholar] [CrossRef]
  13. Liu, L.H.; Zhao, L.; Liu, J.Y.; Yang, Z.C.; Su, G.; Song, H.S.; Xue, J.R.; Tang, A.P. Preparation of magnetic Fe3O4@SiO2@CaSiO3 composite for removal of Ag+ from aqueous solution. J. Mol. Liq. 2020, 299, 112222–112255. [Google Scholar] [CrossRef]
  14. Tao, P.; Shao, M.H.; Song, C.W.; Wu, S.H.; Cheng, M.R.; Cui, Z. Preparation of porous and hollow Mn2O3 microspheres and their adsorption studies on heavy metal ions from aqueous solutions. J. Ind. Eng. Chem. 2014, 20, 3128–3133. [Google Scholar] [CrossRef]
  15. Xia, P.; Wang, X.J.; Wang, X.; Song, J.K.; Wang, H.; Zhang, J.; Zhao, J.F. Struvite crystallization combined adsorption of phosphate and ammonium from aqueous solutions by mesoporous MgO-loaded diatomite. Colloid Surf. A 2016, 506, 220–227. [Google Scholar] [CrossRef]
  16. Li, X.L.; Guo, F.Q.; Jiang, X.C.; Zhao, X.M.; Peng, K.Y.; Guo, C.L. 2018 Study of low-cost and high-performance biomass activated carbon for phenol removal from wastewater: Kinetics, isotherms, and thermodynamics. Asia Pac. J. Chem. Eng. 2018, 13, 2240–2255. [Google Scholar] [CrossRef]
  17. Zou, J.J.; Guo, C.B.; Jiang, Y.S.; Wei, C.D.; Li, F.F. Structure, morphology and mechanism research on synthesizing xonotlite fiber from acid-extracting residues of coal fly ash and carbide slag. Mater. Chem. Phys. 2016, 172, 121–128. [Google Scholar] [CrossRef]
  18. Liu, L.H.; Liu, S.Y.; Peng, H.L.; Yang, Z.C.; Zhao, L.; Tang, A.P. Surface charge of MCS and its adsorption characteristics for heavy metal ions. Solid State Sci. 2020, 99, 106072–106084. [Google Scholar] [CrossRef]
  19. Shen, C.; Chen, C.; Wen, T.; Zhao, Z.; Wang, X.; Xu, A. Superior adsorption capacity of g-C3N4 for heavy metal ions from aqueous solutions. J. Colloid Interface Sci. 2015, 456, 7–14. [Google Scholar] [CrossRef]
  20. Yu, J.; Zheng, J.D.; Lu, Q.F.; Yang, S.X.; Zhang, X.M.; Wang, X.; Yang, W. Selective adsorption and reusability behavior for Pb2+ and Cd2+ on chitosan/poly(ethylene glycol)/poly(acrylic acid) adsorbent prepared by glow-discharge electrolysis plasma. Colloid Polym. Sci. 2016, 294, 1585–1598. [Google Scholar] [CrossRef]
  21. Wang, X.B.; Cai, W.P.; Liu, S.W.; Wang, G.Z.; Wu, Z.K.; Zhao, H.J. ZnO hollow microspheres with exposed porous nanosheets surface: Structurally enhanced adsorption towards heavy metal ions. Colloid Surf. A 2013, 422, 199–205. [Google Scholar] [CrossRef]
  22. Malakar, A.; Das, B.; Sengupta, S.; Acharya, S.; Ray, S. ZnS nanorod as an efficient heavy metal ion extractor from water. J. Water Process 2014, 3, 74–81. [Google Scholar] [CrossRef]
  23. Chowdhury, S.; Mazumder, M.A.J.; Al-Attas, O.; Husain, T. Heavy metals in drinking water: Occurrences, implications, and future needs in developing countries. Sci. Total Environ. 2016, 569, 476–488. [Google Scholar] [CrossRef]
  24. Hayati, B.; Maleki, A.; Najafi, F.; Daraei, H.; Gharibi, F.; McKay, G. Synthesis and characterization of PAMAM/CNT nanocomposite as a super-capacity adsorbent for heavy metal (Ni2+, Zn2+, As3+, Co2+) removal from wastewater. J. Mol. Liq. 2016, 224, 1032–1040. [Google Scholar] [CrossRef]
  25. Singh, J.; Sharma, M.; Basu, S. Heavy metal ions adsorption and photodegradation of remazol black XP by iron oxide/silica monoliths: Kinetic and equilibrium modelling. Adv. Powder Technol. 2018, 29, 2268–2279. [Google Scholar] [CrossRef]
  26. Wang, X.H.; Zhao, K.Y.; Yang, B.X.; Chen, T.; Li, D.Y.; Wu, H.; Wei, J.F.; Wu, X.Q. Adsorption of dibutyl phthalate in aqueous solution by MCS grafted non-woven polypropylene. Chem. Eng. J. 2016, 306, 452–459. [Google Scholar] [CrossRef]
  27. Huang, Z.; Wu, Q.; Liu, S.; Liu, T.; Zhang, B. A novel biodegradable beta-cyclodextrin-based hydrogel for the removal of heavy metal ions. Carbohydr. Polym. 2013, 97, 496–501. [Google Scholar] [CrossRef]
  28. Teoh, Y.P.; Khan, M.A.; Choong, T.S.Y. Kinetic and isotherm studies for lead adsorption from aqueous phase on carbon coated monolith. Chem. Eng. J. 2013, 217, 248–255. [Google Scholar] [CrossRef]
  29. Németh, G.; Mlinárik, L.; Török, Á. Adsorption and chemical precipitation of lead and zinc from contaminated solutions in porous rocks: Possible application in environmental protection. J. Afr. Earth Sci. 2016, 122, 98–106. [Google Scholar] [CrossRef]
  30. Kanagaraj, P.; Nagendran, A.; Rana, D.; Matsuura, T.; Neelakandan, S.; Karthikkumar, T.; Muthumeenal, A. Influence of N-phthaloyl chitosan on poly (ether imide) ultrafiltration membranes and its application in biomolecules and toxic heavy metal ion separation and their antifouling properties. Appl. Surf. Sci. 2015, 329, 165–173. [Google Scholar] [CrossRef]
  31. Liu, L.H.; Wu, J.; Li, X.; Ling, Y.L. Synthesis of poly(dimethyldiallylammonium chloride-co-acrylamide)-graft-triethylenetetramine–dithiocarbamate and its removal performance and mechanism of action towards heavy metal ions. Sep. Purif. 2013, 103, 92–100. [Google Scholar] [CrossRef]
  32. Salmani Abyaneh, A.; Fazaelipoor, M.H. Evaluation of rhamnolipid (RL) as a biosurfactant for the removal of chromium from aqueous solutions by precipitate flotation. J. Environ. Manag. 2016, 165, 184–187. [Google Scholar] [CrossRef] [PubMed]
  33. Segundo, J.E.D.V.; Salazar-Banda, G.R.; Feitoza, A.C.O.; Vilar, E.O.; Cavalcanti, E.B. Cadmium and lead removal from aqueous synthetic wastes utilizing Chemelec electrochemical reactor: Study of the operating conditions. Sep. Purif. 2012, 88, 107–115. [Google Scholar] [CrossRef]
Figure 1. Effect of initial solution pH and adsorbent addition amount on adsorption of Cu2+, Pb2+: Cu2+ removal rate (a), Cu2+ adsorption capacity (c); Cu2+ equilibrate pH (e); Pb2+ removal rate (b), Pb2+ adsorption capacity (d); Pb2+ equilibrate pH (f).
Figure 1. Effect of initial solution pH and adsorbent addition amount on adsorption of Cu2+, Pb2+: Cu2+ removal rate (a), Cu2+ adsorption capacity (c); Cu2+ equilibrate pH (e); Pb2+ removal rate (b), Pb2+ adsorption capacity (d); Pb2+ equilibrate pH (f).
Materials 16 03506 g001
Figure 2. X-ray diffraction patterns of MCS (a) and MCS loaded with Cu2+ (b), Ni2+ (c).
Figure 2. X-ray diffraction patterns of MCS (a) and MCS loaded with Cu2+ (b), Ni2+ (c).
Materials 16 03506 g002
Figure 3. Fourier transform infrared spectroscopy spectra of MCS (a) and MCS loaded with Pb2+ (MCS−Pb) (b), Cu2+ (MCS−Cu) (c).
Figure 3. Fourier transform infrared spectroscopy spectra of MCS (a) and MCS loaded with Pb2+ (MCS−Pb) (b), Cu2+ (MCS−Cu) (c).
Materials 16 03506 g003
Figure 4. Scanning electron microscopy images of MCS (a) and MCS loaded with Cu2+ (b), Pb2+ (c). Energy spectrum analysis diagram (d) of MCS: MCS−Pb (e), MCS−Cu (f).
Figure 4. Scanning electron microscopy images of MCS (a) and MCS loaded with Cu2+ (b), Pb2+ (c). Energy spectrum analysis diagram (d) of MCS: MCS−Pb (e), MCS−Cu (f).
Materials 16 03506 g004
Figure 5. Nitrogen adsorption-desorption isotherms and pore size distribution curves of MCS (a) and MCS loaded with Cu2+ (MCS−Cu) (b), Pb2+ (MCS−Pb) (c).
Figure 5. Nitrogen adsorption-desorption isotherms and pore size distribution curves of MCS (a) and MCS loaded with Cu2+ (MCS−Cu) (b), Pb2+ (MCS−Pb) (c).
Materials 16 03506 g005
Figure 6. Adsorption isotherms of MCS for Cu2+ (a), Pb2+ (b).
Figure 6. Adsorption isotherms of MCS for Cu2+ (a), Pb2+ (b).
Materials 16 03506 g006
Figure 7. Change in adsorption amount (qt) with time (t) at different Initial concentrations: Cu2+ (a), Pb2+ (b).
Figure 7. Change in adsorption amount (qt) with time (t) at different Initial concentrations: Cu2+ (a), Pb2+ (b).
Materials 16 03506 g007
Figure 8. Effect of ionic strength on adsorption of heavy metal ions by MCS: Pb2+ (a), Cu2+ (b).
Figure 8. Effect of ionic strength on adsorption of heavy metal ions by MCS: Pb2+ (a), Cu2+ (b).
Materials 16 03506 g008
Table 1. Parameters of the Cu2+ isotherm model for MCS adsorption.
Table 1. Parameters of the Cu2+ isotherm model for MCS adsorption.
Langmuir Equation
NameT (K)qm (mg/g)KL (L/mg)R2
MCS-Cu293487.6221.4070.975
303525.9471.2780.982
313561.8851.4850.987
Freundlich equation
NameT (K)KF(mg/g)(L/mg)1/nnR2
MCS−Cu293321.23413.7740.997
303322.93811.4070.999
313332.60611.1231.000
Table 2. Parameters of the Pb2+ isotherm model for MCS adsorption.
Table 2. Parameters of the Pb2+ isotherm model for MCS adsorption.
Langmuir Equation
NameT (K)qm (mg/g)KL (L/mg)R2
MCS−Pb2931827.0131.0710.617
3031879.0314.7230.657
3131921.5066.6430.676
Freundlich equation
NameT (K)KL (mg/g)(L/mg)1/nnR2
MCS−Pb2931194.35412.9180.993
3031269.23613.6120.981
3131293.00813.3670.978
Table 3. Kinetic parameters of adsorption of Cu2+ and Pb2+.
Table 3. Kinetic parameters of adsorption of Cu2+ and Pb2+.
ModelParametersCu2+Pb2+
200 mg/L400 mg/L600 mg/L200 mg/L400 mg/L600 mg/L
Pseudo-first-order modelR20.9560.7310.9350.7660.6400.721
k1169 × 10−337.8 × 10−348.8 × 10−335.6 × 10−355.1 × 10−369.9 × 10−3
qe (cal)329.725463.319528.2381269.231775.1721907.849
Pseudo-second-order modelR20.9970.9510.9740.9320.9380.968
k20.858 × 10−30.148 × 10−30.216 × 10−30.046 × 10−30.071 × 10−30.112 × 10−3
qe (cal)336.386481.484541.3521325.5381822.3371940.063
Table 4. Elution and recovery of Cu2+ and Pb2+ from the adsorbed MCS.
Table 4. Elution and recovery of Cu2+ and Pb2+ from the adsorbed MCS.
ItemsCu2+Pb2+
Elution ration (%)99.8799.95
Recovery ratio (%)99.6499.78
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

Chen, C.; Chen, C.; Li, J.; Wang, G.; Lin, X.; Ning, D. Preparation of MCS from Low-Grade Bauxite Desilication Lye and Adsorption of Heavy Metals. Materials 2023, 16, 3506. https://doi.org/10.3390/ma16093506

AMA Style

Chen C, Chen C, Li J, Wang G, Lin X, Ning D. Preparation of MCS from Low-Grade Bauxite Desilication Lye and Adsorption of Heavy Metals. Materials. 2023; 16(9):3506. https://doi.org/10.3390/ma16093506

Chicago/Turabian Style

Chen, Cheng, Chaoyi Chen, Junqi Li, Gangan Wang, Xin Lin, and Deyang Ning. 2023. "Preparation of MCS from Low-Grade Bauxite Desilication Lye and Adsorption of Heavy Metals" Materials 16, no. 9: 3506. https://doi.org/10.3390/ma16093506

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