Translate this page into:
Adsorption of Cd2+ and Pb2+ by biofuel ash-based geopolymer synthesized by one-step hydrothermal method
⁎Corresponding authors. fawangzhang1965@sina.com (Fawang Zhang), hanzhantao@tcare-mee.cn (Zhantao Han)
-
Received: ,
Accepted: ,
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.
Abstract
A hydrothermal alkali modification method was used to produce geopolymer (BFA-GP) from biofuel ash (BFA). The product was characterized by scanning electron microscopy with energy dispersive spectrometry (SEM-EDS), transmission electron microscopy (TEM), X-ray diffraction (XRD), Fourier transform infrared reflection (FTIR), and X-ray photoelectron spectroscopy (XPS). Significant amount of geopolymer and gismondine was produced by this modification. The surface area increased from 20.41 m2 g−1 to 56.63 m2 g−1. The adsorption capacity of BFA-GP can reach 29.92 and 137.49 mg g−1 for Cd2+ and Pb2+, respectively, in a pure solution at 298 K, which is about triple of that by the original BFA. Competitive adsorption of Cd2+ and Pb2+ showed that the binding affinity is Pb2+> Cd2+. Chemical sorption including electrostatic attraction, chelate reaction, and ion-exchange is the dominant mechanism of heavy metal adsorption on BFA-GP. And the adsorption thermodynamics indicates that adsorption reaction of heavy metal ions by BFA-GP is spontaneous and endothermic. This modification provided a cost-effective and environmentally friendly way to change BFA to adsorbent for heavy metals with promising application prospects.
Keywords
Biofuel ash
Hydrothermal method
Geopolymer
Heavy metal adsorption
1 Introduction
Wastewater contains high concentrations of soluble heavy metal ions discharged from many industries such as mining, metallurgy, and battery manufacturing (Xu et al., 2020, Shao et al. 2020a), which has led to contamination of groundwater and soil (Lin et al. 2020a). Heavy metals are one of the most hazardous materials due to their high toxicity and persistence (Qiao et al. 2020, Lin et al. 2020b).
Techniques used to remove heavy metals from water include chemical precipitation (Duarte-Nass et al. 2020), ion exchange (Nekouei et al. 2019), membrane filtration (Bai, Wang and Zhu 2020), and adsorption (Di Natale, Gargiulo and Alfe 2020). Adsorption has been proposed due to the advantages of easy operation, cost-effectiveness, and extensively available adsorbent materials (Xu et al. 2020). These are widely used in the treatment of heavy metal-contaminated water (Wieszczycka et al. 2020, Fakhre and Ibrahim 2018). Adsorbents include activated carbon (Shahrokhi-Shahraki et al. 2020), natural zeolites (Hong et al. 2019), fly ash(Alinnor 2007), and other industrial and agricultural wastes (Shao et al. 2020b) and organic materials such as resins and alginate hydrogels (Jiang et al. 2020). Among them, industrial waste-derived materials are one the most promising adsorbents for metal ions adsorption due to their high adsorption capacity and low cost. (Di Natale et al. 2020).
In recent years, geopolymers have attracted much attention as a new heavy metal adsorbent (Siyal et al. 2018, Chen et al. 2019b, Rasaki et al. 2019). Geopolymers are inorganic polymeric materials with a three-dimensional porous structure synthesized from industrial wastes or mineral materials containing aluminosilicate through alkaline activation (Khale and Chaudhary 2007). Geopolymers have high adsorption ability due to their large specific surface area, porous structure, and abundant surface functional groups (Tan et al. 2020). They also have good mechanical properties and thermal/chemical stability versus organic adsorbents (Chen et al. 2019a). Geopolymers have been widely used in heavy metals (Darmayanti et al. 2019) and dye (Barbosa et al. 2018) removal from liquid as well as VOCs (Tang et al. 2020) adsorption from waste gas. But most of the method to produce geopolymer needs long curing time (>10 days) (Andrejkovičová et al., 2016; Yousef et al., 2009; Huang and Han 2011), which is very costly in practice.
Precursor materials for the preparation of geopolymers include fly ash (Onutai et al. 2019, Wang et al. 2020), metakaolin (Cheng et al. 2012, Kara, Yilmazer and Akar 2017, Tian, Guo and Sasaki 2020), red mud (Chen et al. 2019b) and other industrial byproducts containing aluminosilicate (Chen et al. 2019a, Sarkar, Basu and Samanta 2019, Sha, Pan and Sun 2020). In the past years, large amount biomass power plants have been built up in China as a new way to produce renewable energy (Qin et al. 2018). It is estimated that 480 million tons of BFA will be produced by 2050 worldwide (Stanislav V. Vassilev et al. 2013). Thus, the disposal and utilization of BFA has become a challenging task worldwide. The main elements in BFA are silica, aluminum, oxygen, calcium, potassium, and other metals (Shi et al. 2017). The content of silica and aluminum oxide in BFA can reach more than 60%, making BFA a suitable precursor material for preparing geopolymer. Pérez-Villarejo et al. (2018) prepared geopolymers using biomass fly ash and aluminum industrial slag activated by a mixture of sodium hydroxide and sodium silicate. Novais et al. (2019) synthesized geopolymer from biomass fly ash by injecting FA slurry into a polyethylene glycol medium (85 °C) and curing at room temperature for 27 days. Based on these studies, we designed a fast hydrothermal alkali modification method to produce geopolymer by BFA, and the produced geopolymer was testified has superior adsorption capacity for Cd2+ and Pb2+ which are very popular heavy metal contaminants in China. It’s expected the BFA-GP could be widely used in the heavy metal wastewater treatment and heavy metal contaminated soil remediation.
2 Materials and methods
2.1 Materials
BFA used in this study obtained from the bottom of a boiler that produces electricity by burning several kinds of biomass as fuel in JinZhou biomass power plant (Shijiazhuang, China). XRF test results are listed in Table 1.
Composition
Content(%)
SiO2
Al2O3
CaO
Fe2O3
MgO
Na2O
P2O5
K2O
BFA
47.26
12.67
19.19
7.49
3.86
1.19
1.07
4.20
BFA-GP
51.9
15.4
11.79
5.60
3.12
3.3
0.579
2.08
Analytical reagent (AR) grade sodium hydroxide (NaOH), hydrochloric acid (HCl), cadmium nitrate (Cd(NO3)2·2H2O), lead nitrate (Pb(NO3)2) were purchased from Tianjin, China. All the water used in the experiment was ultrapure water.
2.2 Modification of BFA
BFA was ground to ≤ 74 μm and subsequently washed by ultrapure water until the conductivity was constant. It was then oven dried at 105 °C overnight.
Geopolymer was synthetized by one-step hydrothermal alkali modification method as shown in Fig. 1. A fixed amount of BFA was mixed with 2 M NaOH at solid-liquid ratio of 1:15. The resulting mixture was fed into a round flask and heated in a 90 °C water bath with stirring for 12 h. The material was cooled naturally followed by aging for 2 h at ambient temperature and separated by centrifugation. The solid products were washed three time by deionized water, and freeze dried for 48 h. The solid was sieved to the particle size of ≤ 74 μm, and the final sample was called BFA -GP.Schematic map showing BFA-GP synthetization.
2.3 Characterization
Scanning Electron Microscope with Energy Dispersive Spectrometer (SEM- EDS, ProX, Phenom, Holland) was utilized to observe the microscopic morphology and chemical elements of BFA and BFA-GP. Transmission electron microscopy (TEM) characterization was carried out by a TEM/STEM FEI Tecnai F30 microscope. The surface area, total pore volume and pore size distribution were measured through N2 adsorption–desorption method at approximately 77.3 K using Automatic Surface Area and Porosity Analyzer (ASAP 2460, Micromeritics, USA). The specific surface areas of BFA and BFA-GP were determined for the relative pressure range of 0.0–1.0 and calculated by the Brunauere-Emmette-Teller (BET) method. The total pore volume was determined for a single point P/P0 value of 0.990. The pore size distributions of the samples were inferred from the adsorption isotherms by means of the Barrett-Joyner-Halenda (BJH) method. Before testing, the sample was degassed at 200 °C for 6 h. Mineral composition of BFA and BFA-GP was obtained by X-ray diffraction (XRD, D8 Advance, Bruker, Germany) using Cu-Kα radiation with a wavelength of 1.5406 Å, tube voltage and current are 40 KV and 40 mA, respectively. The range of XRD patterns was from 2θ of 10° to 80° with a scanning rate at 1.5°2θ/min. The functional groups in BFA-GP were probed using Fourier-transform infrared spectroscopy (FTIR, iS10, Nicolet, USA) by KBr pelletized technique. Each spectrum in the 4000–400 cm−1 range was accumulated from 32 individual scans. The chemical components of samples were characterized by X-ray photoelectron spectroscopy (XPS) using a Thermo escalab 250XI spectrometer with monochromatic Al Kα radiation. The zeta potential of the geopolymer was measured at different pH (1–9) using Zeta-sizer instrument (MALVERN, UK) to calculate the point of zero charge (pHZPC).
2.4 Adsorption experiments
The solutions contained 320 mg L−1 of Cd2+ and 1600 mg L−1 of Pb2+ were prepared by dissolving the corresponding nitrate salts Cd(NO3)2·2H2O and Pb(NO3)2 in deionized water, and diluted to the desired concentration in subsequent experiments.
Batch experiments were carried out to get the equilibrium adsorption data of Cd2+ or Pb2+ or their mixture onto BFA-GP. The experimental procedures are described as following: 0.1 g BFA-GP was put into a 50-mL Teflon centrifuge tube, 40 mL solution containing Pb2+ or Cd2+ or their mixture with different concentrations was then added. The pH of mixture solution was adjusted to 6.5 ± 0.2 with 0.1 M HCl or NaOH and the tubes were sealed with screw caps. All samples were prepared in duplicate. The Teflon centrifuge tubes were shaken at 298 K and 150 rpm for 24 h. After shaking, the mixtures were centrifuged at 3000 rpm for 10 min and then filtered through a 13-mm syringe filter with a 0.45-μm membrane. The supernatants were collected and the metal ion concentrations were measured by atomic absorption spectrophotometry (AA-7000, Shimadazu, Japan).
Kinetic adsorption were performed by adding 40 mL 400 mg L−1 Pb2+ or 80 mg L−1 Cd2+ solutions, or their mixture into eighteen 50-mL Teflon centrifuge tubes. The pH of mixture solution was adjusted to 6.5 ± 0.2 with 0.1 M HCl or NaOH and sealed with screw caps. The samples were shaken at 298 K and 150 rpm and at determined time intervals, 2 samples were collected, centrifuged and the supernatant were filtered and the metal ion concentrations were measured by the same method as in equilibrium adsorption.
Another set of adsorption experiments were performed to investigate the adsorption thermodynamics. Three pare of samples were prepared as in previous steps: 40 mL Pb2+ and Cd2+ in the range of 20–800 mg L−1 and 2–160 mg/L and 0.1 g BFA-GP were put into 50-mL Teflon centrifuge tubes, the pH of sample solution was adjusted to 6.5 ± 0.2 with 0.1 M HCl or NaOH and the tubes were sealed with screw caps, but were shaken under different temperatures (298 K, 308 K, and 318 K) for 24 h before centrifugation and measure the Pb2+ and Cd2+ concentrations in supernatants.
The adsorption at equilibrium (
, mg g−1) were determined according to Eq. (1):
3 Results and discussion
3.1 Characterization of BFA-GP
3.1.1 Mineral composition
X-ray diffraction (XRD) was used to detect the mineral composition of BFA and BFA-GP. The XRD spectrograms of BFA and BFA-GP are shown in Fig. 2. The main minerals in BFA is quartz (JCPDS No: 46-1045) and margarite (JCPDS No: 18-0276) with some CaCO3. The amorphous halo between 2θ 20° and 35° observed in the BFA XRD pattern is the characteristic of amorphous nature of biofuel ash (Siyal et al. 2016, Hui-Teng et al. 2021). After modification, the intensity of the diffraction peak of quartz at 2θ = 20.8° decreased, and the peak of margarite at 2θ = 27.9° disappeared, indicating that some of the quartz and all the margarite in the BFA dissolved during the modification. The broad peak around 2θ between 20° and 35° in the XRD pattern of BFA-GP may be attributed to the presence of amorphous geopolymer (Singhal, Gangwar and Gayathry 2017). The diffraction peak at 2θ = 26.6° which is belongs to both quartz and gismondine (JCPDS No: 20-0452) intensified, and a new peak appear at 2θ = 36.5° which is belongs to gismondine. The formation of gismondine is very favorable for heavy metal cations adsorption due to the precense of the ion exchange centers in zeolite where Al and Si tetrahedra are connected.XRD patterns of BFA and BFA-GP.
3.1.2 Microstructural characteristics and surface area
Fig. 3 presents SEM images of BFA and BFA-GP, and their corresponding energy dispersive spectrometry. By comparing Fig. 2(a) and (b), many cracks and opening pores and some small particles were produced by the modification. The change in the morphology of the product likely causes a significant increase in the specific surface area as shown in Fig. 2(b).SEM micrographs of BFA (a), BFA-GP (b) and EDS analysis of BFA (c), BFA-GP (d).
As demonstrated in Fig. 2(c, d), the main elements of BFA and BFA-GP were O, Si and Al. But a Na peak appeared in the spectrum of BFA-GP which come from NaOH solution, indicating that alkali activator participated in the modification and Na ions were fixed into BFA-GP. Similar results have been reported in previous literature (Sha et al. 2020).
The microstructure of BFA and BFA-GP were shown in high-resolution TEM images (Fig. 3). Many spherical particles of around 100–800 nm were observed in BFA as shown in Fig. 4(a)–(b). But in the images of BFA-GP, many fiber-like gismondine crystals with diameter of around 5 nm can be clearly observed (Fig. 4(d)–(f)). The morphology of geopolymer prepared by biofuel ash is different from that by coal bottom ash (Santa et al. 2021). This may be due to the varied Si/Al ratio of raw material and different polymerization conditions(Rożek, Król and Mozgawa 2019). Potassium salt and sodium salt in raw biofuel ash were dissolved during washing.TEM images of BFA (a, b, c) and BFA-GP (d, e, f) at different magnifications.
The N2 adsorption-desorption isotherms of BFA and BFA-GP are displayed in Fig. 5. The specific surface area (SBET), total open pore volume at P/Po = 0.99(VT), t-Plot micropore volume (Vmic), and BJH desorption average pore width of BFA and BFA-GP were determined by N2 adsorption–desorption test. The surface area of the BFA increased from 20.41 m2 g−1 to 56.63 m2 g−1 after modification; meanwhile, the total open pore volume also increased from 0.0337 cm3 g−1 to 0.1600 cm3 g−1, and the pore size increased from 12.5 nm to 13.6 nm. The increase in specific surface area and pore volume may be due to the dissolution and rearrangement of aluminosilicate, which resulted in new pores and small particles as seen in the SEM and TEM images (Fig. 3 and Fig. 4). The specific surface area is generally considered to be the most important property determining the adsorption capacity (Liu et al. 2016).N2 adsorption–desorption isotherm curves of BFA and BFA-GP.
Usually, the characteristic pores in the material can be judged by the shape of the adsorption curve and the hysteresis loop. The adsorption curves of BFA and BFA-GP were type IV adsorption curves and were mostly produced by mesoporous substances. Moreover, according to the classification of international union of pure and applied chemistry (IUPAC), the hysteresis loop of BFA-GP was Type H3 type indicating that there were slit-like pores in the material. The pore size distributions of BFA and BFA-GP were displayed in the insert graph of Fig. 5. There were two main differences in pore size distribution between BFA-GP and BFA. One was the increase in pore width less than 2 nm, which may result from the formation of gismondine. Another significant difference was the pore width of 3–50 nm in the BFA-GP increased drastically. This change could be attributed to the formation of geopolymer identified as a mesoporous material with a characteristic pore size between 2 and 50 nm (Siyal et al. 2018).
3.1.3 FTIR and XPS analysis
FTIR spectra of BFA, BFA-GP, and BFA-GP after adsorption of Cd (BFA-GP-Cd) are shown in Fig. 6 (a). The comparison of BFA-GP with BFA shows some changes in the adsorption band. The broad absorption band around 3440 cm−1 and the peak near 1630 cm−1 represent asymmetric stretching vibration and deformational vibrations of H—O—H and —OH bond (Kara et al. 2017, Siyal et al. 2019). The two peaks intensified in the spectra of BFA-GP indicated that hydroxyls on the surface of geopolymer increased (Siyal et al. 2019). The peak at approximately 1420 cm−1 and the small one at 874 cm−1 in spectrum of BFA are ascribed to asymmetric stretching and out-of-plane bending vibrations of the O—C—O bonds of CO3−2 (Singhal et al. 2017). The decomposition of calcium carbonate in the hydrothermal process may explain the decrease of absorption intensity at these bands. Bands near 1030 cm−1 are associated with asymmetric stretching vibrations of Si—O—T (T = Si or Al) (Barbosa et al. 2018). This is the principal band used to recognize the synthesis of geopolymers (Siyal et al. 2019) and is well characterized in the literature (Siyal et al. 2019, Wang et al. 2020, Tian, Nakama and Sasaki 2019). This peak deepened obviously after modification, indicating there are significant amount of geopolymer in BFA-GP. Here, Al—O is longer and weaker than Si—O, and the shift of the peak to lower wavenumbers (from 1033 cm−1 for BFA to 1000 cm−1 for BFA-GP) indicates that the degree of silicon substitution by aluminum increased. The bands at 777 cm−1 corresponded to the cyclosilicate vibrations which is a kind of silica tetrahedron oligomer (Acisli, Acar and Khataee 2020, Lecomte et al. 2006) this is other evidence of rearrangement of aluminosilicate (He et al. 2020, Huang and Han 2011).FTIR spectrum of the BFA, BFA-GP and BFA-GP-Cd (a) and XPS spectra of BFA and BFA-GP: (b) wide scan, (c) Si 2p spectra, (d) Al 2p spectra, (e) O 1 s high resolution XPS spectra of BFA, (f) O 1 s high resolution XPS spectra of BFA-GP.
The spectrum of BFA-GP-Cd in Fig. 6 (a) show that the absorption of –OH at 3440 cm−1 decreased compared to BFA-GP suggesting that coordination reaction likely occurred between Cd2+ and surface hydroxyl group. Meanwhile, the peak of Si—O—T (T = Si or Al) shift to higher wavenumbers after adsorption, indicating Si—O—T (T = Si or Al) group also adsorbed significant amount of Cd2+.
The XPS spectra provided molecular structure information of the BFA and BFA-GP as shown in Fig. 6. The intensity of peaks corresponding to Si, O, Al, and Ca elements changed indicating the transformation of chemical bond during the modification. The intensification of O, Si, and Al peaks mainly due to the polymerization of aluminosilicate. Formation of gismondine (CaAl2Si2O8·4H2O) may explain the peak intensification corresponding to calcium. There are three types of primary chemical states of oxygen in geopolymer including Si—O—Al, Si—O—Si, and Si—O—H with the main absorption signal positions at 531, 532, and 533 eV according to the literature (Simonsen et al. 2009, Jhang, Boscoboinik and Altman 2020). The fitted parameters of the O1 s XPS spectra are shown in Fig. 4(e) and (f) for BFA and BFA-GP, respectively. There are more Si—O—Al and Si—O—H in BFA-GP, which is in accordance with the FTIR results. These results reconfirmed that new composition such as geopolymer and gismondine were formed.
3.1.4 Zeta potential and zero charge point
The zeta potential graphs (Fig. 7) at solution pH range of 1–9 reveal that the point of zero charge (pHPZC) of BFA-GP is at 1.3, where the adsorbent surface is having electron neutrality (Gao et al. 2015). The result clearly suggest that BFA-GP exhibited negatively charge under pH > 1.3, and Zeta potential is lower than −23 mV when the pH of solution higher than 3. This property is important for geopolymer to generate electrostatic attraction for the metal cations.Zeta potentials of BFA-GP sample under different pH conditions.
3.2 Comparison of Cd2+ adsorption on BFA-GP and BFA
Fig. 8 shows the adsorption amount of Cd2+ and Pb2+ by BFA and BFA-GP. The maximum adsorption amount of Cd2+ and Pb2+ increased from 7.94 to 29.92 mg g−1 and 32.87 to 137.49 mg g−1 respectively after modification, indicating BFA-GP synthetized is a high efficiency adsorbent for heavy metal ions.Adsorption isotherm of Cd2+ (a) and Pb2+ (b) on BFA-GP and BFA.
3.3 Effect of pH
The pH value of the aqueous solution affects both the surface charges of adsorbent and the degree of ionization of heavy metal ion, so it is an important variable to adsorption effect (Bao et al. 2013). The Fig. 9 illustrated the adsorption amounts of metal ions at different pH conditions. The sorption ability of BFA-GP for heavy metals increased rapidly with the increasing of solution pH. The adsorption amounts of Pb2+ increased at pH between 3 and 6, and amount adsorbed was near constant after pH 6. With regard to Cd2+, the adsorption amounts were very low when pH was below 5, and increased quickly between pH 5 and 6, after pH 6 adsorption amounts increased slowly and reached equilibrium gradually. This results may be related to the surface charge of BFA-GP. At low pH, higher concentration of H+ present in the reaction solutions, which can compete with heavy metal ions for the adsorption sites, and the functional groups of adsorbent were protonated, significant electrostatic repulsion exists between positively charge surface and heavy metal ions. With the increasing of solution pH, the concentration of the H+ decreases and deprotonation would occur on the surface of BFA-GP, the surface negative charge increased, which enhanced the adsorption of positively charged metal cations by electrostatic attraction. Similar theories have also been proposed by other researchers for the metal adsorption (Liu and Zhang 2009, Duan and Su 2014, Huang et al. 2020b, Liu et al. 2016). Additionally, adsorption amounts of Cd2+ and Pb2+ reach constant at different pH maybe due to the different hydrolysis precipitation property of two ions (Xu 2008). Considering the large amount of precipitation of metal ions could affect the adsorption under high pH, selecting pH = 6.5 as the optimum condition for subsequent experiments.Effect of initial pH on Cd2+ (a) and Pb2+ (b) adsorption.
3.4 Characterization of adsorption kinetics
Fig. 10 shows the kinetic adsorption curves of Cd2+ and Pb2+ on BFA-GP in pure and mixed solutions. The adsorption process of Cd2+ and Pb2+ by BFA-GP shows two different stages. The first stage is fast and should be adsorption on the surface and the second stage is slow adsorption and due to the diffusion to narrow cracks and pores(Qiu, Cheng and Huang 2018). At initial stage of the reaction, there are a large number of vacant adsorption sites on the adsorbent surface, and the heavy metal ions diffusing to the adsorbent surface can be captured by the adsorption sites quickly; on the other hand, the abundant Cd2+ and Pb2+ in the solution enhance mass transfer driving force in the liquid phase, which is beneficial to the diffusion of metal ions to the adsorbent surface. The adsorption rate slowed down after 240 min indicating that the adsorption sites on the adsorbent surface were saturated, the Cd2+ and Pb2+ began to diffuse to narrow pores in the BA-GP. At adsorption equilibrium, the adsorption quantity of Cd2+ and Pb2+ in pure solution was 26.84 and 143.73 mg g−1, respectively. The adsorption capacity in a mixed Cd2+ and Pb2+ solution was 8.80 and 135.88 mg g−1, this is 67.2% and 5.4% lower than their adsorption in pure solution.Adsorption kinetics of Cd2+ (a) and Pb2+ (b) onto BFA-GP in the pure and mixed metal system.
To further investigation mechanisms of adsorption, pseudo-first-order kinetics model, pseudo-second-order kinetics model and intra-particle diffusion model have been used to evaluate the kinetics data both in pure and mixed solution.
Pseudo-first-order kinetics model often express the adsorption process in which the reaction rate is controlled by physical diffusion (Eq. (2)). Pseudo-second-order kinetics model is used to describe chemisorption exists between the adsorbate and the adsorbents (Eq. (3)). The intra-particle diffusion model is employed to decide if intra particle diffusion is the only rate limiting step in the adsorption process (Eq. (4)) (Zhu et al. 2017).
The fitting results of pseudo-first-order and pseudo-second-order kinetics model are illustrated in Fig. 10, and the kinetic parameters are presented in Table 2. The correlation coefficient of pseudo-second-order kinetics model is higher than that of the pseudo-first-order model, the equilibrium adsorption capacity calculated by the pseudo-second-order kinetic model is closer to the experimental values in all the cases. The results show that the adsorption process of Cd2+ and Pb2+ onto BFA-GP in pure and mixed solutions is more consistent with the pseudo-second-order kinetics model, indicating chemisorption is the predominant step of the adsorption process (Gao et al. 2015). This was also found in other studies (Wang, Terdkiatburana and Tadé 2008). qee is the experimental value, qe is the equilibrium adsorption capacity calculated by Pseudo-first order model and Pseudo-second order model.
Solution
qee
Pseudo-first order model
Pseudo-second order model
(mg g−1)
qe (mg g−1)
k1 (min−1)
R2
qe (mg g−1)
k2 (g mg−1 min−1) (×104)
R2
Pure-Cd
26.84
22.23
0.0117
0.9979
29.88
3.19
0.9999
Mixed-Cd
8.8
8.22
0.0076
0.9988
9.91
6.71
0.9996
Pure-Pb
143.73
132.85
0.0118
0.9349
147.36
0.96
0.9840
Mixed-Pb
136.93
118.48
0.0136
0.9608
135.04
1.10
0.9695
It is generally believed that the adsorption process can be divided into three steps: i) external diffusion—heavy metal ions diffuse onto the external surface of the adsorbent, ii) internal diffusion—adsorbates diffuse into the pore channels of the adsorbent, and iii) the reaction of adsorbates and adsorption sites external or internal of the adsorbent. The rate of reaction is usually fast, and thus the first two processes are the factors controlling adsorption rate. The intra-particle diffusion model is suitable to describe the kinetics in which the diffusion inside the particle is dominated. If the curve of qt versus t0.5 is a straight line and passes through the origin, then this indicates that the intra-particle diffusion process is the only rate-limiting step to control the adsorption rate (Kong et al. 2016).
Fig. 11 shows the fitting line segments of the intraparticle diffusion model for the experimental data and the fitting parameters are laid out in Table 3. The curves did not pass through origin, and the experimental data showed multiple linear relationships indicating that the adsorption of Cd2+ and Pb2+ onto BFA-GP is controlled by two or more factors. The experimental data plot can be fitted as two linear stages: i) 0–240 min with a fast reaction rate that is the external diffusion process; ii) a slow adsorption stage controlled by the internal diffusion process of Cd2+ and Pb2+.Intra-particle diffusion model plots of Cd2+ (a) and Pb2+ (b) in pure and mixed solution.
Solution
Stage1
Stage2
(mg g−1 min−0.5)
(mg g−1 min−0.5)
Pure-Cd
1.493
0.9921
0.253
0.8545
Mixed-Cd
0.412
0.9950
0.088
0.9372
Pure-Pb
7.327
0.9933
1.229
0.9858
Mixed-Pb
5.851
0.9863
1.278
0.9589
3.5 Adsorption isotherms of Cd2+ and Pb2+ on BFA-GP
3.5.1 Adsorption isotherms in pure solution
Both Langmuir model (Eq. (5)) and Freundlich model (Eq. (6)) was used to fit the adsorption isotherms:
Nonlinear fitted curves of experimental data at different temperatures in pure solution by two adsorption isotherms are displayed in Fig. 12, the related constants are listed in Table 4. The correlation coefficients R2 show that the Freundlich isotherms can better describe the adsorption of Cd2+ on BFA-GP indicating the heterogeneity of the adsorbent surface and multi-layer adsorption may occur (Chen et al. 2017). The Pb2+ adsorption on BFA-GP is more consistent with the Langmuir adsorption isotherm indicating that the adsorption process of Pb2+ was monolayer adsorption containing chemical reactions such as ion exchange and electron sharing (Qiu et al. 2018). Based on the data in Table 4, all the kf values increase with increasing temperature, which indicates that high temperature benefits adsorption of Cd2+ and Pb2+ by BFA-GP. The values of n are always greater than 2 at different temperatures for both Cd2+ and Pb2+. This shows the high adsorption intensity between adsorbate and adsorbent. qee is the experimental value.Adsorption isotherms of Cd2+ (a) and Pb2+ (b) at different temperature in pure systems.
Metal
T/K
qee
Langmuir
Freundlich
Cd
298
29.92
1.0281
25.89
0.8794
11.5502
4.8099
0.9721
308
31.01
5.5461
28.10
0.9175
16.9278
7.1505
0.9691
318
32.34
5.3649
30.66
0.7001
22.2795
11.3308
0.9131
Pb
298
137.49
0.0359
149.01
0.8846
19.9510
3.0096
0.789
308
165.20
0.2441
179.30
0.7799
53.1437
4.7288
0.6112
318
185.62
0.3936
190.39
0.8053
59.1397
4.7893
0.6582
Comparison of the maximum adsorption capacities of BFA-GP for Cd2+ and Pb2+ with other similar adsorbents reported in literatures are listed in Table 5. Clearly, the BFA-GP produced in this study shown a higher or equivalent adsorption capacity for Cd2+ and Pb2+ compare with most of the similar adsorbents.
Adsorbents
Adsorptive capacity (mg g−1)
References
Cd
Pb
BFA and slags-based geopolymers
10.58
50
(Pérez-Villarejo et al. 2018)
Alkali-activated fly ash
26.47
–
(Krol et al. 2018)
Zeolite-based geopolymer
26.25
–
(Javadian, 2013)
Geopolymer-alginate-chitosan composites
–
142.67
(Yan et al. 2019)
Biofuel ash based geopolymer
29.92
137.49
This study
3.5.2 Competitive adsorption
The competitive adsorption of heavy metals by BFA-GP was shown in Fig. 13 (a) and (b). When the concentrations of Cd2+ and Pb2+ in the solution are below 20 and 100 mg L−1, respectively, the adsorption of Cd2+ and Pb2+ in mixed solution is consistent with adsorption in pure solution indicating that competitive adsorption is not significant at low concentrations. This is because BFA-GP can provide sufficient adsorption sites in low concentration solutions. If the concentration of the two ions in the solution is higher than 20 and 100 mg L−1, then the adsorption amount of two ions in mixed solution was lower than that in pure solutions.Comparison of adsorption quantity of Cd2+ (a) and Pb2+ (b) in pure and mixed solution and Langmuir competitive adsorption model for Cd2+ and Pb2+ adsorption on BFA-GP (c).
The adsorption amount of Cd2+ in mixed solution is obviously lower than in pure solution, but the adsorption amount of Pb2+ has no significant changes indicating that the presence of Pb2+ seriously limits the adsorption of Cd2+. Similar results were shown in previous literature (Ma et al. 2015, Cheng et al. 2012, Huang et al. 2020a). Smaller hydration ion radius, smaller surface free energy and stronger adsorption affinity lead to easier adsorption (Chen et al. 2019b). The hydrated ionic radius of lead (4.01 Å) is smaller than that of cadmium (4.26 Å), and thus Pb2+ enters the micro-pore of the adsorbent more easily. The surface free energy of Cd2+ and Pb2+ hydrated ions is 4.26 and 4.01 kcal mol−1, respectively. A larger free energy keeps Cd2+ in solution. The affinity depends on their hydrolysis constant K, and the ions with a large hydrolysis constant are more prone to forming metal hydroxy groups (e.g., Pb(OH)− and Cd(OH)−) in the solution. The pK values of Pb2+ and Cd2+ are 7.7 and 10.1, respectively. In other words, the hydrolysis constant of Pb2+ is about 250-fold that of Cd2+, which determines that the adsorption affinity order is Pb2+>Cd2+.
Langmuir competitive adsorption model was introduced to fit the adsorption data obtained in mixed solution. Langmuir competitive adsorption model can be expressed as follow (Huang et al. 2020b):
The Langmuir model fitting for the mixed system is shown in Fig. 13 (c). According to the competitive model fitting results, the correlation coefficients of Cd2+ and Pb2+ were 0.9995 and 0.9947, respectively, indicating that adsorption behaviors could be nicely fitted with a Langmuir competitive model. qmax of Cd2+ and Pb2+ calculated by Langmuir model was 2.58 and 126.10 mg g−1. The fitting results are very consistent with the experimental data. This result indicates that adsorption of heavy metal ions onto BFA-GP will be restricted by other coexisting ions.
3.6 Adsorption thermodynamics of Cd2+ and Pb2+ on BFA-GP
The effect of temperature on adsorption of Cd2+ and Pb2+ onto BFA-GP is shown in Fig. 14. The experiments were conducted at 298, 308, and 318 K. The distribution coefficient values (K) increased with increasing temperature indicating the endothermic nature of the adsorption.Plot of ln K against 1/T obtained for the adsorption of Cd2+ and Pb2+ onto BFA-GP.
Thermodynamic parameter, free energy change (
), enthalpy change (
) and entropy change (
) were calculated using Van’t Hoff equation (Gao et al. 2015).
Thermodynamic parameters of adsorption Cd2+ and Pb2+ on BFA-GP are listed in Table 6.
values are positive confirming that BFA-GP adsorption for Cd2+ and Pb2+ is endothermic(Chatterjee, Basu and Jana 2019), means a high temperature is favorable for the reaction. This is consistent with the increase of heavy metal ion adsorbing onto the BFA-GP with the increasing of temperature, as shown in Fig. 9. Although there is no clear criterion defining the adsorption type on the basis of
, the adsorption process with
ranges from 20.9 to 418.4 kJ mol−1 and is recognized as a chemisorption process (Nuri and Ersoz, 2019). The enthalpy change of Cd2+ and Pb2+ is 8.262 and 23.912 kJ mol−1, respectively. This confirmed that chemisorption is involved in the adsorption process of Pb2+. The ΔG values are negative indicating that the adsorption reaction is spontaneous. The entropy change (
) is negative, which indicates that the confusion degree of the system is reduced after the adsorption reaction, and the metal ions in the solution are fixed on the surface of the adsorbent.
Metal
ΔG (
)
ΔH (
)
ΔS (
)
298
308
318
Cd
−14.379
−15.185
−15.898
8.262
−0.07603
Pb
−14.142
−15.509
−16.694
23.912
−0.12779
3.7 Mechanism of Cd2+ and Pb2+ adsorption on BFA-GP
Fig. 15 displays the SEM images and elemental mapping images of BFA-GP after adsorption Cd2+ and Pb2+. The adsorption of heavy metal ions on BFA-GP was confirmed by EDS analysis. The characteristic signals of cadmium and lead elements are observed in Fig. 12 (c) and (f), respectively. In Fig. 15(a), amorphous geopolymer can be seen as white cloud-like particles on the surface of BFA-GP, and Cd can be observed predominantly adsorb on the surface of those geopolymer particles. In Fig. 15(d), needle like gismondine particles can be clearly observed, and Pb also predominantly adsorbed by gismondine particles as shown in Fig. 15(e). This clearly indicated the newly produced geopolymer and gismondine promoted the adsorption for Pb and Cd by BFA.SEM-EDS analysis of BFA-GP loaded with Cd2+ (a, b, c) and Pb2+ (d, e, f).
Fig. 16 shows the bond energy absorption analysis of BFA-GP before and after adsorption of Cd2+ ions and Pb2+ ions. The adsorption of Cd2+ and Pb2+ onto BFA-GP were also observed by the XPS spectrum. The characteristic signals of cadmium or lead elements are not observed in the spectra of BFA-GP, but appeared in the spectra of BFA-GP samples after adsorption. The bond energy absorption positions for Cd 3d are 405.58 and 412.28 eV corresponding to Cd 3d5/2 and Cd 3d3/2, respectively. The bond energy absorption positions for Pb 4f are 138.68 and 143.68 eV representing Pb 4f7/2 and Pb 4f5/2, respectively (Huang et al. 2020a). Moreover, Fig. 16 (d,e,f) showed that the binding energy peaks of O1s move from 531.88 to 531.68, 531.78, and 531.48 eV after adsorption of Cd, Pb, and Cd + Pb, respectively, it may be due to that the chelate reaction between heavy metal ions and oxygen containing functional groups including Si—O—Si, Si—O—Al, and Si—O—H changed the bonding energy of O1s (Kaewmee et al. 2020). The reduction in the bond strengths of Ca 2p after adsorption clearly indicated that ion-exchange occurred between heavy metal ions and Ca2+ in Ca—A—-Si—H gel (Qiu et al. 2018). Hence, the mechanisms of the adsorption process should include electrostatic attraction, chelate reaction, and ion-exchange.XPS spectra of BFA-GP before and after loading Cd2+ and Pb2+ in pure and mixed solution: (a) survey spectra, (b) Cd 3d spectra, (c) Pb 4f spectra, (d) O1s spectra of BFA-GP-Cd, (e) O1s spectra of BFA-GP-Pb, (f) O1s spectra of BFA-GP-Cd + Pb.
4 Conclusion
A hydrothermal alkali modification method to produce geopolymer from biofuel ash at low temperature was proposed. This process partly change quartz, margarite and CaCO3 to geopolymer and Gismondine. Specific surface area and pore volume of biofuel ash was increased from 20.41 m2 g−1 and 0.0337 cm3 g−1 to 56.63 m2 g−1 and 0.1600 cm3 g−1, respectively after the modification. The adsorption capacity of the modified biofuel ash for Cd2+ and Pb2+ can reach 29.92 and 143.73 mg g−1, respectively, which is better than most similar adsorbents. Meanwhile, the binding affinity of two cations on BFA-GP followed the order of Pb2+> Cd2+. Kinetic experimental data are more consistent with pseudo-second-order kinetic models. Both FTIR, XPS and adsorption thermodynamics indicated the adsorption of Cd2+ and Pb2+ on BFA-GP is mainly chemical sorption, including electrostatic attraction, chelate reaction, and ion-exchange happened during the adsorption. The produced geopolymer could be used as low-cost adsorbent for heavy metal removal from wastewater, and provided a new method for the use of biomass power plant ash.
Funding
This study was supported by Key Research & Development Project of Hebei Province, China (Grant No.18274232), National Key Research and Development Program of China (Grant No.2019YFC1805300) and Key Research & Development Project of Guangxi Province, China (Grant No.AB17195036), Basic scientific research operating expenses of universities of Hebei Province (Grant No. QN202119). We also thank LetPub (www.letpub.com) for its linguistic assistance during the preparation of this manuscript.
Declaration of Competing 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
- Preparation of a fly ash-based geopolymer for removal of a cationic dye: Isothermal, kinetic and thermodynamic studies. J. Ind. Eng. Chem.. 2020;83:53-63.
- [Google Scholar]
- Adsorption of heavy metal ions from aqueous solution by fly ash. Fuel. 2007;86:853-857.
- [Google Scholar]
- Adsorption of Cr(III) and Pb(II) by graphene oxide/alginate hydrogel membrane: Characterization, adsorption kinetics, isotherm and thermodynamics studies. Int. J. Biol. Macromol.. 2020;147:898-910.
- [Google Scholar]
- Adsorption of heavy metal ions from aqueous solutions by zeolite based on oil shale ash: Kinetic and equilibrium studies. Chem. Res. Chin. Univ.. 2013;29:126-131.
- [Google Scholar]
- Barbosa, T.R., Foletto, E.L., Do tto, G.L., Jahn, S.L., 2018. Preparation of mesoporous geopolymer using metakaolin and rice husk ash as synthesis precursors and its use as potential adsorbent to remove organic dye from aqueous solutions. Ceram. Int. 44, 416–423.
- Alumina-silica nano-sorbent from plant fly ash and scrap aluminium foil in removing nickel through adsorption. Powder Technol.. 2019;354:792-803.
- [Google Scholar]
- Chen, F., Wang, K., Shao, L., Muhammad, Y., Wei, Y., Gao, F., Wang, X., Cui, X., 2019a. Synthesis of Fe2O3-modified porous geopolymer microspheres for highly selective adsorption and solidification of F− from waste-water. Compos. Part B: Eng., 178.
- Removal of Cd(II) and Pb(II) ions from aqueous solutions by synthetic mineral adsorbent: Performance and mechanisms. Appl. Surf. Sci.. 2017;409:296-305.
- [Google Scholar]
- Utilization of red mud in geopolymer-based pervious concrete with function of adsorption of heavy metal ions. J. Cleaner Prod.. 2019;207:789-800.
- [Google Scholar]
- The heavy metal adsorption characteristics on metakaolin-based geopolymer. Appl. Clay Sci.. 2012;56:90-96.
- [Google Scholar]
- Structural alteration within fly ash-based geopolymers governing the adsorption of Cu(2+) from aqueous environment: Effect of alkali activation. J. Hazard Mater.. 2019;377:305-314.
- [Google Scholar]
- Adsorption of heavy metals on silica-supported hydrophilic carbonaceous nanoparticles (SHNPs) J. Hazard Mater.. 2020;393:122374
- [Google Scholar]
- Removal characteristics of Cd(II) from acidic aqueous solution by modified steel-making slag. Chem. Eng. J.. 2014;246:160-167.
- [Google Scholar]
- Application of microbe-induced carbonate precipitation for copper removal from copper-enriched waters: Challenges to future industrial application. J. Environ. Manage.. 2020;256:109938
- [Google Scholar]
- The use of new chemically modified cellulose for heavy metal ion adsorption. J. Hazard Mater.. 2018;343:324-331.
- [Google Scholar]
- Combined modification of fly ash with Ca(OH)2/Na2FeO4 and its adsorption of Methyl orange. Appl. Surf. Sci.. 2015;359:323-330.
- [Google Scholar]
- Low-cost and facile synthesis of geopolymer-zeolite composite membrane for chromium(VI) separation from aqueous solution. J. Hazard Mater.. 2020;392:122359
- [Google Scholar]
- Study of the adsorption of Cd (II) from aqueous solution using zeolite-based geopolymer, synthesized from Coal Fly Ash, Kinetic, isotherm and thermodynamic studies. Arabian J. Chem. 2013
- [CrossRef] [Google Scholar]
- Heavy metal adsorption with zeolites: The role of hierarchical pore architecture. Chem. Eng. J.. 2019;359:363-372.
- [Google Scholar]
- Competitive heavy metal adsorption onto new and aged polyethylene under various drinking water conditions. J. Hazard Mater.. 2020;385:121585
- [Google Scholar]
- Potential of removing Cd(II) and Pb(II) from contaminated water using a newly modified fly ash. Chemosphere. 2020;242:125148
- [Google Scholar]
- The influence of alpha-Al2O3 addition on microstructure, mechanical and formaldehyde adsorption properties of fly ash-based geopolymer products. J. Hazard Mater.. 2011;193:90-94.
- [Google Scholar]
- Formulation, mechanical properties and phase analysis of fly ash geopolymer with ladle furnace slag replacement. J. Mater. Res. Technol.. 2021;12:1212-1226.
- [Google Scholar]
- Ambient pressure x-ray photoelectron spectroscopy study of water formation and adsorption under two-dimensional silica and aluminosilicate layers on Pd(111) J. Chem. Phys.. 2020;152:084705
- [Google Scholar]
- Preparation of a novel bio-adsorbent of sodium alginate grafted polyacrylamide/graphene oxide hydrogel for the adsorption of heavy metal ion. Sci. Total Environ.. 2020;744:140653
- [Google Scholar]
- Porous and reusable potassium-activated geopolymer adsorbent with high compressive strength fabricated from coal fly ash wastes. J. Cleaner Prod.. 2020;272
- [Google Scholar]
- Metakaolin based geopolymer as an effective adsorbent for adsorption of zinc(II) and nickel(II) ions from aqueous solutions. Appl. Clay Sci.. 2017;139:54-63.
- [Google Scholar]
- Mechanism of geopolymerization and factors influencing its development: a review. J. Mater. Sci.. 2007;42:729-746.
- [Google Scholar]
- Synthesis of zeolite-supported microscale zero-valent iron for the removal of Cr(6+) and Cd(2+) from aqueous solution. J. Environ. Manage.. 2016;169:84-90.
- [Google Scholar]
- Influence of alkali metal cations/type of activator on the structure of alkali-activated fly ash - ATR-FTIR studies. Spectrochim. Acta A Mol. Biomol. Spectrosc.. 2018;198:33-37.
- [Google Scholar]
- (Micro)-structural comparison between geopolymers, alkali-activated slag cement and Portland cement. J. Eur. Ceram. Soc.. 2006;26:3789-3797.
- [Google Scholar]
- Application of algae for heavy metal adsorption: A 20-year meta-analysis. Ecotoxicol. Environ. Saf.. 2020;190:110089
- [Google Scholar]
- A comparative study on fly ash, geopolymer and faujasite block for Pb removal from aqueous solution. Fuel. 2016;185:181-189.
- [Google Scholar]
- Removal of lead from water using biochars prepared from hydrothermal liquefaction of biomass. J. Hazard Mater.. 2009;167:933-939.
- [Google Scholar]
- Competitive adsorption of heavy metal ions on carbon nanotubes and the desorption in simulated biofluids. J. Colloid Interface Sci.. 2015;448:347-355.
- [Google Scholar]
- Selective isolation of heavy metals from spent electronic waste solution by macroporous ion-exchange resins. J. Hazard Mater.. 2019;371:389-396.
- [Google Scholar]
- Synthesis of porous biomass fly ash-based geopolymer spheres for efficient removal of methylene blue from wastewaters. J. Cleaner Prod.. 2019;207:350-362.
- [Google Scholar]
- Nuri, Unlu, Ersoz, Mustafa, 2019. Adsorption characteristics of heavy metal ions onto a low cost biopolymeric sorbent from aqueous solutions. J. Hazard. Mater. B136, 272–280.
- Porous fly ash-based geopolymer composite fiber as an adsorbent for removal of heavy metal ions from wastewater. Mater. Lett.. 2019;236:30-33.
- [Google Scholar]
- Biomass fly ash and aluminium industry slags-based geopolymers. Mater. Lett.. 2018;229:6-12.
- [Google Scholar]
- Effect of coal fly ash's particle size on U adsorption in water samples and thermodynamic study on adsorption. Environ. Chem. Ecotoxicol.. 2020;2:32-38.
- [Google Scholar]
- Fabrication of superporous cellulose beads via enhanced inner cross-linked linkages for high efficient adsorption of heavy metal ions. J. Cleaner Prod.. 2020;253
- [Google Scholar]
- Biomass and biofuels in China: Toward bioenergy resource potentials and their impacts on the environment. Renew. Sustain. Energy Rev.. 2018;82:2387-2400.
- [Google Scholar]
- Removal of Cd2+ from aqueous solution using hydrothermally modified circulating fluidized bed fly ash resulting from coal gangue power plant. J. Cleaner Prod.. 2018;172:1918-1927.
- [Google Scholar]
- Geopolymer for use in heavy metals adsorption, and advanced oxidative processes: A critical review. J. Cleaner Prod.. 2019;213:42-58.
- [Google Scholar]
- Microstructural characteristics of geopolymer materials with twenty eight days of curing and after eight years stored at room temperature. J. Cleaner Prod.. 2021;278
- [Google Scholar]
- Experimental and kinetic study of fluoride adsorption by Ni and Zn modified LD slag based geopolymer. Chem. Eng. Res. Des.. 2019;142:165-175.
- [Google Scholar]
- The effect of natural zeolite on microstructure, mechanical and heavy metals adsorption properties of metakaolin based geopolymers. Appl. Clay Sci.. 2016;126:141-152.
- [Google Scholar]
- A novel raw material for geopolymers: Coal-based synthetic natural gas slag. J. Cleaner Prod.. 2020;262
- [Google Scholar]
- High efficiency removal of heavy metals using tire-derived activated carbon vs commercial activated carbon: Insights into the adsorption mechanisms. Chemosphere. 2020;264:128455
- [Google Scholar]
- An all-in-one strategy for the adsorption of heavy metal ions and photodegradation of organic pollutants using steel slag-derived calcium silicate hydrate. J. Hazard. Mater.. 2020;382
- [Google Scholar]
- Characteristics of biomass ashes from different materials and their ameliorative effects on acid soils. J. Environ. Sci. (China). 2017;55:294-302.
- [Google Scholar]
- XPS and FT-IR investigation of silicate polymers. J. Mater. Sci.. 2009;44:2079-2088.
- [Google Scholar]
- CTAB modified large surface area nanoporous geopolymer with high adsorption capacity for copper ion removal. Appl. Clay Sci.. 2017;150:106-114.
- [Google Scholar]
- Geopolymerization kinetics of fly ash based geopolymers using JMAK model. Ceram. Int.. 2016;42:15575-15584.
- [Google Scholar]
- A review on geopolymers as emerging materials for the adsorption of heavy metals and dyes. J. Environ. Manage.. 2018;224:327-339.
- [Google Scholar]
- Fly ash based geopolymer for the adsorption of anionic surfactant from aqueous solution. J. Cleaner Prod.. 2019;229:232-243.
- [Google Scholar]
- An overview of the composition and application of biomass ash. Part 1. Phase–mineral and chemical composition and classification. Fuel. 2013;105:40-76.
- [Google Scholar]
- Current development of geopolymer as alternative adsorbent for heavy metal removal. Environ. Technol. Innovation. 2020;18
- [Google Scholar]
- Reduce VOCs and PM emissions of warm-mix asphalt using geopolymer additives. Constr. Build. Mater.. 2020;244
- [Google Scholar]
- Immobilization mechanism of Se oxyanions in geopolymer: Effects of alkaline activators and calcined hydrotalcite additive. J. Hazard. Mater.. 2020;387:121994
- [Google Scholar]
- Immobilization of cesium in fly ash-silica fume based geopolymers with different Si/Al molar ratios. Sci. Total Environ.. 2019;687:1127-1137.
- [Google Scholar]
- Single and co-adsorption of heavy metals and humic acid on fly ash. Sep. Purif. Technol.. 2008;58:353-358.
- [Google Scholar]
- Effects of Si/Al ratio on the efflorescence and properties of fly ash based geopolymer. J. Cleaner Prod.. 2020;244
- [Google Scholar]
- Novel ionic liquid-modified polymers for highly effective adsorption of heavy metals ions. Sep. Purification Technol.. 2020;236
- [Google Scholar]
- Electrospun SiO2-MgO hybrid fibers for heavy metal removal: Characterization and adsorption study of Pb(II) and Cu(II) J. Hazard. Mater.. 2020;381:120974
- [Google Scholar]
- Adsorption of Pb(II) from aqueous solution to MX-80 bentonite: Effect of pH, ionic strength, foreign ions and temperature. Appl. Clay Sci.. 2008;41:37-46.
- [Google Scholar]
- Novel composites based on geopolymer for removal of Pb(II) Mater. Lett.. 2019;239:192-195.
- [Google Scholar]
- The influence of using Jordanian natural zeolite on the adsorption, physical, and mechanical properties of geopolymers products. J. Hazard. Mater.. 2009;165:379-387.
- [Google Scholar]
- Adsorption of Cd(II) and Pb(II) by in situ oxidized Fe3O4 membrane grafted on 316L porous stainless steel filter tube and its potential application for drinking water treatment. J. Environ. Manage.. 2017;196:127-136.
- [Google Scholar]