An invaluable utilization approach for industrial wastes is to employ them as effective adsorbents for environmental pollutants. This study aimed to investigate the phosphorus (P) adsorption behavior of coal wastes and zeolite in three forms of pristine powder (CP and ZP), nanoparticles (CNP and ZNP), and Fe (III)-modified nanoparticles (MCNP and MZNP). The adsorbents were characterized using X-ray diffraction (XRD), Fourier transform infrared (FTIR), scanning electron microscopy (SEM), and energy-dispersive spectroscopy (EDS) analyses. The effects of pH, initial P concentration, and contact time were studied under batch mode. Results showed an optimum pH range of 2–6 for the P adsorption process. The pseudo-second-order kinetic model and the Langmuir isotherm described the P adsorption data well. The P adsorption capacity of the studied adsorbents was enhanced after modifications. However, the coal-based modified adsorbents represented higher P adsorption performances rather than the zeolite ones. The maximum P adsorption capacity (Qmax) values were obtained as 0.36, 3.23, and 30.48 mg g−1 for CP, CNP, and MCNP, and 0.80, 2.84, and 6.99 mg g−1 for ZP, ZNP, and MZNP, respectively. The surface complexation, ligand exchange, and electrostatic attraction processes were identified as the main P adsorption mechanisms by the studied adsorbents.

  • Phosphorus nano-adsorbents were derived from coal solid wastes.

  • The FeCl3-modified adsorbents effectively removed the aquatic phosphorus.

  • Adsorption process was pH-dependent and dominated by chemisorption.

  • A sustainable approach in waste management and environmental protection was suggested.

Graphical Abstract

Graphical Abstract

Eutrophication is a global environmental concern that seriously threats aquatic ecosystems, and results in loss of biodiversity and large economic dissipations (Zhou et al. 2022). It occurs as a consequence of massive phosphorus (P) discharge into natural water resources through various anthropogenic activities including excessive use of agricultural compounds such as fertilizers and pesticides, paint industries, detergents, and various municipal wastes (Goscianska et al. 2018; Park et al. 2021). A maximum permissible range of 0.5–1 mg P L−1 has been established by the Environmental Protection Agency (EPA) in wastewater (Park et al. 2021). Hence, in the past few decades, efforts have been significantly made to control this essential element regarding water quality standards.

Several physicochemical and biological techniques have been used to remove and recover P from wastewater and industrial effluents including electrochemical precipitation, filtration, ion exchange, and microbial degradation processes (Di Capua et al. 2022). Among them, adsorption has been developed as a promising water treatment strategy following its major advantages such as low cost and low energy requirements, simplicity, non-toxicity, and reversibility (Abdellaoui et al. 2021). Consequently, a large number of synthetic materials have been used for P adsorption including carbonized sludge (Zhang et al. 2018), calcium meta-silicate minerals (Obradović et al. 2017), iron-impregnated biochar (Lee et al. 2018), and layered double hydroxide-loaded biochar (Bolbol et al. 2019).

Recently, a new comprehensive approach has been established based on the circular economy concept on the valorization of agricultural and industrial waste by-products by employing them as novel composite materials for environmental decontamination, particularly water resources restoration (Xu et al. 2022). In this context, various solid waste materials have been examined for P adsorption such as different biomass-derived biochars (Jung et al. 2015; Ngatia et al. 2017), zirconium-loaded orange waste (Biswas et al. 2008), and concrete powder (Liu et al. 2020). Most of them have represented considerable P removal performances regarding their favorable physicochemical and structural characteristics including high porosity, surface area, and mineral composition, as well as their cost-effectiveness and environmental friendship.

Coal fly ash is one of the most important industrial by-products worldwide, which originates from the coal combustion process and is excreted in large quantities in the environment as a waste material, causing many environmental problems such as soil and water contamination, as well as waste management issues (Usman et al. 2022). From the sustainable management perspective, several utilizing plans have been suggested for these waste materials such as using them in cement industry (Singh et al. 2019), rubber production (Ren & Sancaktar 2019), and other engineering or agricultural applications (Ahmaruzzaman 2010). In addition, effective recapture of P from wastewater can be considered an important alternative approach to utilize coal fly ash following its great potential to act as a P adsorbent.

Coal solid wastes have recently received much attention by environmental researchers to be used in P removal reactions from aquatic environments. In this context, some efforts have been made to enhance their P removal efficiencies through different modification procedures such as pre-treatment with different metal ions, and impregnation in acidic or alkaline solutions (Wang et al. 2016a; Xu et al. 2022). However, limited research data regarding enhancement of their adsorption capacity through simultaneous application of physical and chemical modification methods are available. Since the adsorption functionality of a material depends largely on its size, shape, and surface morphology, its smaller size results in higher chemical reactivity. It can be achieved through size reduction of the coal wastes to nanoscales, creating new waste-based nano-adsorbents. On the other hand, chemical surface modification of these nano-adsorbents through embedding high P affinity metals such as Fe can increase their active surface adsorption sites and improve their P removal performance. Reusing coal solid wastes to remove phosphorus from water would be a novel scheme that synergistically supports waste management and environmental remediation issues. Hence, the present study was conducted to investigate the P removal efficiency of the coal solid wastes as an industrial adsorbent in the raw, nano-size, and Fe-modified forms in comparison with similar forms of zeolite as a well-known natural P adsorbent. The adsorbents were prepared, modified, and characterized using SEM-EDS, FTIR, and XRD instrumental analyses and their P adsorption performances were evaluated under batch conditions as affected by solution P concentration, pH, and contact time. Furthermore, the plausible mechanisms involved in the P adsorption processes were explored and discussed in detail regarding the obtained data.

Materials

The coal solid wastes used in this study were collected from the Zarand Coal Washing Plant in the Kerman province, southeast Iran, and the natural zeolite particles were obtained from the Sartakht Natural Zeolite Mine located at the Semnan region, north-central of Iran. All chemical reagents used in this study were of analytical grade (purity > 99%) provided from Merck, Germany, including potassium dihydrogen phosphate (KH2PO4), potassium antimony tartrate (K(SbO)·C4H4O6·0.5H2O), ammonium heptamolybdate tetrahydrate ((NH4)6Mo7O24), hydrochloric acid (HCl), sulfuric acid (H2SO4), sodium chloride (NaCl), sodium hydroxide (NaOH), iron chloride hexahydrate (FeCl3·6H2O), and ascorbic acid. Distilled water (DW) was used in the preparation of all chemical solutions except for P stock solution (1,000 mg L−1) which was prepared by dissolving an appropriate amount of KH2PO4 in 0.01 M NaCl solution as background electrolyte. Desired P concentrations were then diluted from the prepared stock solution.

Preparation of the adsorbents

The adsorbents used in this study included the raw coal waste particles (CPs), coal waste nanoparticles (CNPs), FeCl3-modified coal waste nanoparticles (MCNPs), raw natural zeolite particles (ZPs), zeolite nanoparticles (ZNPs), and FeCl3-modified zeolite nanoparticles (MZNPs). The collected raw materials were powdered to pass through a 270-mesh size sieve, washed with DW several times to remove their impurities, and oven-dried at 70 °C for 24 h. The CNP and ZNP were prepared through physical modification (nano-size reduction) of the powdered particles using a Ball-mill for 10 h. The chemical surface modification of the prepared nanoparticles was carried out following Wang et al. (2016b). First, the CNP and ZNP particles were impregnated with 2 M NaOH and 1 M HCl, respectively, for 24 h, and after drying, they were added to 500 mL DW and the suspension pH was adjusted around 13. Then, a 0.5 M FeCl3 solution was added drop-wise into the stirring suspension until the pH value decreased to 5, and the suspension was allowed to precipitate for 24 h. Finally, the suspension was centrifuged and the remaining residues were oven-dried at 105 °C for 3 h, ground and sieved by a 270-mesh size sieve.

Characterization of the adsorbents

A MIRA3-TESCAN scanning electron microscope (SEM) equipped with an energy-dispersive spectroscopy detector (EDS) was used to investigate the surface morphology and elemental composition of the prepared adsorbents, respectively. The crystallographic structure of the adsorbents was analyzed over 2θ range of 10–60° using a Philips X'pert Pro MPD model X-ray diffractometer (XRD) with Cu Kα radiation, running at 40 kV and 30 mA. The surface functional groups and chemical bonds were identified through infrared spectra (400–4,000 cm−1) using a Bruker Tensor 27 Fourier transform infrared spectrometer (FTIR).

Determination of point of zero charge (pHpzc)

The pHpzc of the prepared adsorbents was determined following the method described by Feizi & Jalali (2016). Briefly, 10 mL of 0.01 M NaCl was placed into a series of 50 mL centrifuge tubes and their initial pH values were adjusted in the range of 2–12 using 0.1 M HCl and NaOH solutions. Then, about 0.1 g of each adsorbent was added to the prepared solutions and the obtained suspensions were shaken for 48 h at room temperature. Finally, the equilibrium pH of each suspension was recorded and plotted across the initial pH value. The pHpzc for each adsorbent was obtained from the intersection of the plotted curve with the blank sample line.

Batch adsorption experiments

The P adsorption characteristics of the prepared adsorbents including the effect of pH (2–10), contact time (0–1,440 min), and initial P concentration (0–500 mg L−1) were investigated under batch mode at room temperature. According to the pre-treatments, the initial P concentration and adsorbent dosage were selected as 300 mg L−1 and 10 g L−1, respectively. The suspensions were shaken at 250 rpm for 24 h in three replicates. The solid and liquid phases were separated by centrifugation at 4,000 rpm for 10 min, and the supernatants were filtered using Whatman 42 filter papers. The equilibrium P concentration in the supernatants was measured following Eaton et al. (2005) using a UV/VIS spectrophotometer at 880 nm wavelength. The P removal efficiency and equilibrium P adsorbed by the adsorbents were calculated using the following equations, respectively:
(1)
(2)
where R refers to the P removal efficiency (%), Ci and Ce represent the initial and equilibrium P concentrations, respectively (mg L−1), Q refers to the amount of P adsorbed (mg g−1) per unit mass of adsorbent, V is the volume of the aqueous solution (L), and m is the adsorbent dry mass (g).

Adsorption kinetics

The kinetic experiments were carried out on the 300 mg L−1 initial P concentration between 0 and 1,440 min time periods at room temperature, and the amount of P adsorbed at each time (Qt) was obtained by the following equation:
(3)
where Ci and Ct (mg L−1) are the P concentrations at initial and t time, respectively. The three most commonly used kinetic models including pseudo-first-order (Lagergren 1898), pseudo-second-order (Ho & McKay 1999), and intra-particle diffusion (Weber & Morris 1963) models were used to describe the P adsorption kinetic data using non-linear regression as follows:
(4)
(5)
(6)

In these relations, Qt and Qmax (mg g−1) refer to the P adsorption capacity at times t and equilibrium, respectively, k1 (min−1) and k2 (mg g−1 min−1) represent the rate constant of pseudo-first-order and pseudo-second-order models, respectively, and kp (mg g−1 min−0.5) and C (mg g−1) refer to the rate constant and the intercept of the intra-particle diffusion model from the origin, respectively.

Adsorption isotherms

The isothermal experiments were conducted on the P concentration ranges between 0 and 500 mg P L−1, and the experimental data were fitted to the Langmuir (Langmuir 1916), Freundlich (Freundlich 1906), and Dubinin–Radushkevich (D-R) (McEnaney 1987) models according to their non-linear mathematical equations as follows:
(7)
(8)
(9)

In these relations, Q (mg g−1) and Ce (mg L−1) represent the amount of adsorbed P per unit mass of adsorbent and the equilibrium P concentration in the solution, respectively. The Qmax (mg g−1) and Kl (L mg−1) are the Langmuir constants referring to maximum sorption capacity and the affinity of P ions to the sorption sites, respectively, and the KF (mg g−1), (L mg−1)1/n, and n (g L−1) are the Freundlich constants referring to the sorption capacity and intensity, respectively. β is the activity coefficient related to the mean adsorption energy per mole, R is the universal gas constant (8.314 J mol−1 K−1), and T is the absolute temperature (K).

The correlation coefficients (R2) and standard errors of estimate (SEE) values were used to determine the conformity between the experimental data and model-predicted values in both kinetic and isothermal studies.

Characterization of the adsorbents

The XRD patterns of the studied adsorbents are presented in Figure 1(a). As can be seen, the CNP is dominantly composed of Si and Al oxides, and Quartz (SiO2) is the main mineral in its structure. The appearance of sharp and intense peaks around 2θ values of 21 and 27° confirmed the significant presence of Quartz in the CNP structure. Moreover, Mullite (3Al2O3·2SiO2) and Calcite (CaCO3) minerals were also identified in the CNP structure through the appearance of relatively broad and low intense peaks between 30–40°, and around 45°, respectively (Kobayashi et al. 2020). In the case of the MCNP pattern, the Quartz peaks have been removed or attenuated and instead, a Hematite (Fe2O3) mineral peak has emerged around 2θ = 32°, which can be attributed to the successful loading of Fe atoms on the surfaces of the CNP. The Calcite peak has been also removed after modification of CNP with FeCl3, which can be explained by the dissolution of CaCO3 as a result of FeCl3 acidity (Wang et al. 2016a). The MZNP showed a similar XRD pattern to that of ZNP, indicating that the crystal structure of the natural zeolite was not destroyed after the reaction with FeCl3 (Kragović et al. 2012; Bahabadi et al. 2017).
Figure 1

XRD patterns (a) and FTIR spectra (b) of the studied adsorbents.

Figure 1

XRD patterns (a) and FTIR spectra (b) of the studied adsorbents.

Close modal

Figure 1(b) illustrates the FTIR spectra of the studied adsorbents. The broad absorption bands between 3,400 and 3,500 cm−1 in all adsorbent's spectra are signatures of –OH stretching vibration, demonstrating the presence of water in the adsorbents structure or between their pores. The absorption band around 2,925 cm−1 in the MCNP spectrum refers to the asymmetric stretching vibration of a methylene group (–C–H), which appeared after modification of CNP with FeCl3. Also, the 1,447 cm−1 peak which has intensified after FeCl3 treatment was attributed to the bending vibration of the methylene group. In addition, the 1,614, 1,037, and 797 cm−1 absorption bands were found in the adsorbents FTIR spectra, referring to the O–H, Si–O, and Si–O–Si groups, respectively (Fazlzadeh et al. 2017). The Al(Si)–O–(Si)Al asymmetric stretching vibration which has intensified in the MCNP spectrum, indicated its zeolite-like surface chemistry as a consequence of alkaline conditions (NaOH treatment) before FeCl3 modification (Kobayashi et al. 2020). Moreover, the 1,632 cm−1 band in the MZNP spectrum was related to the presence of the Fe–OH group in the zeolite surface after treatment with FeCl3 (Tandon et al. 2013).

The changes in surface morphologies of the adsorbents after modification with FeCl3 are displayed in Figure 2. It can clearly be seen that the CNP surface is comprised of irregular flakes, which are homogeneously covered by new rough particles after modification with FeCl3. These particles have originated from the FeCl3 hydrolysis and led the development of a porous aggregated morphology on the MCNP surface (Wang et al. 2016a). In addition, the results of EDS analyses showed an increase in the Fe peak in the MCNP spectrum compared with the CNP one, demonstrating the successful loading of Fe atoms on the MCNP surface. In the case of zeolite-based adsorbents, the tabular surface morphology of ZNP has been changed after modification with FeCl3, creating irregular clusters on the MZNP surface. The successful embedding of Fe on the MZNP surface was confirmed by EDS analyses through the appearance of the Fe peak in the MZNP spectrum rather than that of ZNP.
Figure 2

SEM micrographs and EDS spectrums of the studied adsorbents.

Figure 2

SEM micrographs and EDS spectrums of the studied adsorbents.

Close modal

Phosphorus adsorption characteristics

Effect of solution pH

The adsorption process is significantly affected by the solution pH because it simultaneously impacts the adsorbent through changing its surface charges, active sites, and functional groups, and also on the adsorptive through controlling its diffusion rate and chemical speciation in the solution (Kragović et al. 2012). Figure 3(a) and 3(b) represents the effects of solution pH on the adsorption capacities of the studied adsorbents toward phosphorus in solution with an initial concentration of 300 mg L−1, an adsorbent dosage of 10 g L−1, and a contact time of 24 h. The studied adsorbents showed a highly pH-dependent P adsorption, and optimum pH ranges obtained around 2–4 for ZP, and 2–6 for CP, CNP, MCNP, ZNP, and MZNP. The adsorption capacities of the studied adsorbents did not significantly change in these pH ranges, while they decreased sharply with the increase in solution pH out of these ranges. For instance, the amounts of P adsorbed by MCNP and MZNP in the pH range of 2–6 were about 27 and 19 mg g−1, respectively, and they decreased to 12 and 12.7 mg g−1 with an increase in pH toward 10. This could be explained by the electrostatic mechanism involved in the P adsorption process. At low pH ranges, the protonated surfaces of the adsorbents cause an increase in their affinities for P ions in the solution, and in contrast, at high pH ranges, the P ions will be electrostatically repulsed from the adsorbents surfaces as a result of their negatively charged surfaces. Moreover, at high pH values, the OH ions in solution will compete with the P anions for the adsorbents surface sites, resulting in a decrease in P adsorption capacity. Similar results have been also reported for P adsorption by different adsorbents (Feizi & Jalali 2016; Wang et al. 2016b; Bolbol et al. 2019).
Figure 3

Effect of solution pH on the P adsorption capacity of coal-based adsorbents (a), the zeolite-based adsorbents (b), pHpzc of the coal-based adsorbents (c), the zeolite-based adsorbents (d) (adsorbent dosage: 10 g L−1, initial P concentration: 300 mg L−1, contact time: 24 h).

Figure 3

Effect of solution pH on the P adsorption capacity of coal-based adsorbents (a), the zeolite-based adsorbents (b), pHpzc of the coal-based adsorbents (c), the zeolite-based adsorbents (d) (adsorbent dosage: 10 g L−1, initial P concentration: 300 mg L−1, contact time: 24 h).

Close modal

Studying the P speciation in solution as affected by pH provides useful information regarding the mechanisms involved in the P adsorption process. In the pH ranges between 2 and 6, the anion is the predominant P species in the solution, which mainly participates in the P adsorption process through OH exchange mechanism, resulting in an inner-sphere complex formation on the adsorbent surface (Krishnan & Haridas 2008; Feizi & Jalali 2016). When increasing the pH toward 10, the secondary orthophosphate () prevails in the solution and its higher adsorption free energy rather than the , leads to a decrease in P adsorption (Chubar et al. 2005). The pHpzc refers to the pH value that the net surface charge of an adsorbent is zero, and the anion exchange capacity equals the cation exchange capacity (Sparks 2003). This means that in pH ranges lower than the pHpzc, the adsorbent has a net positive surface charge and it has more affinity to adsorb anions from the solution electrostatically. Accordingly, in pH ranges higher than the pHpzc, the adsorbent prefers the cations to adsorb. The higher pHpzc values imply that the charge balance occurs at higher pH values, and therefore, the adsorbent has a positive surface charge over a wider pH range (Feizi & Jalali 2016). Figure 3(c) and 3(d) displays the pHpzc determination curves for the studied adsorbents. As can be seen, the pHpzc of the CP, CNP, and MCNP are approximately coincided with each other around pH = 8, indicating that the physical and chemical modifications of coal waste particles did not affect their pHpzc. However, in the case of ZP, ZNP, and MZNP, the pHpzc points have been increased after modifications, so their pHpzc were obtained as 3.8, 5.8, and 6.5, respectively. Overall, the optimum pH range for P adsorption by the studied adsorbents was lower than their pHpzc and a decreasing trend was found in their P adsorption capacity toward pHpzc. This suggests that the P adsorption process by the studied adsorbents has an electrostatic nature.

Effect of contact time

Figure 4 exhibits the P removal efficiency of the studied adsorbents as affected by contact time. The P removal processes by the CP and ZP showed a slow mono-phase behavior, so only about 4.6 and 4.2% of the aqueous P was removed by them in 24 h, respectively. However, the modified adsorbents (CNP, MCNP, ZNP, and MZNP) revealed a faster P removal process which occurred in two phases. The initial phase was rapid, but the removal rate decreased over time and reached equilibrium after 120 min. The P removal efficiencies were different for modified adsorbents in this time period and obtained as 18, 79, 16, and 58% for CNP, MCNP, ZNP, and MZNP, respectively. These results indicated that the physical (nano-scaling) and chemical (FeCl3) modifications have improved the P removal efficiencies of the adsorbents rather than their initial raw forms. Moreover, the higher P removal percent obtained for MCNP and MZNP was related to the presence of Fe particles on their matrix and their role in P adsorption enhancement (Xu et al. 2010). The initial rapid phase in the P removal process could be attributed to the single-layer adsorption of P on the surfaces of the studied adsorbents as a consequence of the electrostatic attraction of P anions in solution with the positively charged surfaces of the adsorbents. As time increases, the surface adsorption sites will be saturated with P anions, resulting in a decrease in the adsorption rate and cause the adsorption process to reach equilibrium gradually (et al. 2013).
Figure 4

Effect of contact time on the P removal efficiency of the coal-based adsorbents (a) and the zeolite-based adsorbents (b) (adsorbent dosage: 10 g L−1, pH: 4, initial P concentration: 300 mg L−1).

Figure 4

Effect of contact time on the P removal efficiency of the coal-based adsorbents (a) and the zeolite-based adsorbents (b) (adsorbent dosage: 10 g L−1, pH: 4, initial P concentration: 300 mg L−1).

Close modal

Effect of initial P concentration

The P removal efficiency of the studied adsorbents as a function of initial P concentration in solution is shown in Figure 5. With increasing the initial P concentration, all of the studied adsorbents presented a descending tendency in their P removal. As the solution P concentration increased from 50 to 500 mg L−1, the P removal efficiencies decreased from 100 to 39, 37 to 7, and 5 to 0.7 for the MCNP, CNP, and CP adsorbents, respectively. Similarly, the P removal efficiencies for the MZNP, ZNP, and ZP reduced from 80 to 20, 26 to 5, and 6 to 1.4, respectively. At all initial P concentrations, the removal efficiencies of the modified adsorbents were higher than those of the raw materials. However, the MCNP represented a higher performance to remove aqueous P rather than the MZNP, so in 50 mg P L−1 concentration, their removal efficiencies were obtained as 100 and 80%, respectively. The descending trend in the P removal process with the P concentration in solution could be explained by the ratio of the available active sites on the adsorbent's surfaces to the number of adsorptive P anions in the solution. In lower P concentrations, this ratio has higher values and the P species in solution have the most interaction with the surface functional groups of the adsorbents, resulting in higher P removal percentages. With increasing the P concentration in solution, the P removal rate will be decreased as a consequence of the surface adsorption sites saturation (Wang et al. 2012).
Figure 5

Effect of initial P concentration on the P removal efficiency of the coal-based adsorbents (a) and the zeolite-based adsorbents (b) (adsorbent dosage: 10 g L−1, pH: 4, contact time: 24 h).

Figure 5

Effect of initial P concentration on the P removal efficiency of the coal-based adsorbents (a) and the zeolite-based adsorbents (b) (adsorbent dosage: 10 g L−1, pH: 4, contact time: 24 h).

Close modal

Adsorption isotherms

The results of the isothermal studies on the P adsorption by studied adsorbents are displayed in Figure 6. The experimental data were described using Langmuir, Freundlich, and Dubinin–Radushkevich isotherm models and their parameters were extracted and presented in Table 1. As the given graphs illustrate, the Langmuir and Freundlich models have described the P adsorption process better than the Dubinin–Radushkevich one. However, the R2 and SE values show that the Langmuir model is better fitted to the P adsorption data. Considering the Langmuir model assumptions, these results could be explained by the monolayer P adsorption on the surface adsorption sites with similar levels of adsorption energy. The maximum P adsorption capacities predicted with this model were obtained as 0.36 and 0.80 mg g−1 for CP and ZP adsorbents, respectively. While the modified adsorbents showed higher P adsorption capacities as 3.23 and 30.48 mg g−1 for the CNP and MCNP, and 2.84 and 6.99 mg g−1 for ZNP and MZNP, respectively. Moreover, the values of the Langmuir constant (Kl) demonstrating the bonding energy in the P adsorption process were higher for the modified adsorbents compared with the raw materials, implying their higher affinities for P adsorption, resulting in longer retention of P by the modified adsorbents and decreasing their P desorption rate (Jung et al. 2015; Bolbol et al. 2019). Moreover, the 1/n values obtained for the studied adsorbents were all lower than 1, demonstrating the non-linearity in the Freundlich isotherm which commonly occurs for adsorbents with a limited adsorption capacity.
Table 1

Isotherm parameters of P adsorption by the studied adsorbents

Isotherm modelAdsorbent
CPCNPMCNPZPZNPMZNP
Freundlich 
KF 0.13 1.73 4.55 0.11 1.49 3.80 
 1/n 0.19 0.14 0.15 0.36 0.12 0.13 
R2 0.99 0.94 0.96 0.96 0.92 0.96 
 SE 0.01 0.34 2.40 0.06 0.33 0.60 
Langmuir 
Qmax (mg g−10.36 3.23 30.48 0.80 2.84 6.99 
Kl (L·mg−10.03 0.16 1.78 0.03 0.67 1.74 
R2 0.94 0.99 0.99 0.99 0.98 0.95 
 SE 0.03 0.03 0.99 0.04 0.16 0.61 
Dubinin–Radushkevich 
β 0.03 0.03 1.13 0.21 0.01 0.04 
qm (mg g−10.34 3.37 30.89 0.69 2.91 7.11 
R2 0.89 0.74 0.77 0.081 0.69 0.21 
 SE 0.05 0.73 5.82 0.13 0.67 2.61 
Isotherm modelAdsorbent
CPCNPMCNPZPZNPMZNP
Freundlich 
KF 0.13 1.73 4.55 0.11 1.49 3.80 
 1/n 0.19 0.14 0.15 0.36 0.12 0.13 
R2 0.99 0.94 0.96 0.96 0.92 0.96 
 SE 0.01 0.34 2.40 0.06 0.33 0.60 
Langmuir 
Qmax (mg g−10.36 3.23 30.48 0.80 2.84 6.99 
Kl (L·mg−10.03 0.16 1.78 0.03 0.67 1.74 
R2 0.94 0.99 0.99 0.99 0.98 0.95 
 SE 0.03 0.03 0.99 0.04 0.16 0.61 
Dubinin–Radushkevich 
β 0.03 0.03 1.13 0.21 0.01 0.04 
qm (mg g−10.34 3.37 30.89 0.69 2.91 7.11 
R2 0.89 0.74 0.77 0.081 0.69 0.21 
 SE 0.05 0.73 5.82 0.13 0.67 2.61 
Figure 6

Adsorption isotherms of P on the studied adsorbents.

Figure 6

Adsorption isotherms of P on the studied adsorbents.

Close modal

The FeCl3 modification showed higher performance in the case of coal-based adsorbents rather than the natural zeolite nanoparticles. As can be deduced from Table 1, the modification of CNP with FeCl3 has enhanced its P adsorption capacity 10-fold (from 3.23 to 30.48 mg g−1), while in the case of ZNP, a 2.5-fold increase is observed (from 2.84 to 6.99 mg g−1). Krishnan & Haridas (2008) reported a five to six times increase in P adsorption capacity (Qmax = 70.92 mg g−1) by the Fe-modified coir pith rather than the native one, and related it to the cation bridge effect of Fe cations on the adsorbent surface. In another study on the native and FeCl3-modified plant residues, Feizi & Jalali (2016) reported an approximately 2.5-fold increase in P adsorption capacity after modification of the adsorbents with Fe. They obtained Qmax values of 2.8, 4.3, 4, and 3.7 mg g−1 for native sunflower, potato, canola, and walnut shell residues, while these values were increased to 6.6, 9, 8.4, and 9 mg g−1, after Fe modification, respectively. Wang et al. (2012) also studied the P removal from an aqueous solution using two series of activated carbons modified by Fe (II) and Fe (III) and reported 14.12 and 8.73 mg g−1Qmax values, respectively. Overall, it can be concluded from the findings of the present study that the P adsorption capacity found for MCNP was comparable to other adsorbents used in literature, making this industrial waste capable, cost-effective, and eco-friendly P adsorbent (Table 2).

Table 2

Comparison of the P adsorption capacities of the studied adsorbents with various adsorbents

AdsorbentQmax (mg g−1)Reference
Fe (II)-modified activated carbon 14.12 Wang et al. (2012)  
Fe (III)-modified activated carbon 8.73 Wang et al. (2012)  
Fly ash 10.70 Wang et al. (2016b)  
Fe-modified walnut shell Feizi & Jalali (2016)  
La-modified zeolite 44 Goscianska et al. (2018)  
Layered double hydroxide loaded biochar 17.46 Bolbol et al. (2019)  
Coal thermal power plant fly ash 4.1 Park et al. (2021)  
La-modified coal fly ash 10.75 Xu et al. (2022)  
MCNP 30.48 Present study 
MZNP 6.99 Present study 
AdsorbentQmax (mg g−1)Reference
Fe (II)-modified activated carbon 14.12 Wang et al. (2012)  
Fe (III)-modified activated carbon 8.73 Wang et al. (2012)  
Fly ash 10.70 Wang et al. (2016b)  
Fe-modified walnut shell Feizi & Jalali (2016)  
La-modified zeolite 44 Goscianska et al. (2018)  
Layered double hydroxide loaded biochar 17.46 Bolbol et al. (2019)  
Coal thermal power plant fly ash 4.1 Park et al. (2021)  
La-modified coal fly ash 10.75 Xu et al. (2022)  
MCNP 30.48 Present study 
MZNP 6.99 Present study 

The thermodynamic nature of the P adsorption process was investigated using the standard free energy parameter calculated from the Langmuir constant (Kl) as follows (Ghosal & Gupta 2017):
(10)
where ΔG° refers to the Gibbs standard free energy (kJ·mol−1), R is the universal gas constant (8.314 J·mol−1 K−1), T is the absolute temperature (K), and Kl is the Langmuir parameter (L mol−1). The ΔG° values were obtained as −16.94, −21.09, −27.06, −16.94, −24.64, and −27 kJ·mol−1 for CP, CNP, MCNP, ZP, ZNP, and MZNP, respectively. The negative ΔG° values indicated the spontaneous nature of the P adsorption process by the studied adsorbents.

Adsorption kinetics

Adsorption kinetics provide useful information regarding the reaction direction and the associated mechanisms, as well as the adsorbent performance. Three commonly used empirical kinetic models including pseudo-first-order, pseudo-second-order, and intra-particle diffusion equations were applied in the present study to explain the P adsorption kinetic data by the studied adsorbents (Figure 7). The pseudo-first-order model has been established based on the assumption of linear reduction in the adsorption rate as a consequence of adsorption capacity increment, and usually describes a physical adsorption mechanism, while the pseudo-second-order model considers the interaction of the particles as the rate-limiting factor and is commonly used to explain a chemisorption mechanism (Wang et al. 2012). The kinetic models were fitted to the experimental kinetic data and their related parameters are given in Table 3. The pseudo-first-order and pseudo-second-order models presented acceptable results on the P adsorption kinetics, suggesting the simultaneous effects of the physical and chemical mechanisms in the P adsorption processes. However, based on the average values, the pseudo-second-order model represented relatively higher R2 and lower SE values rather than the pseudo-first-order one, indicating the chemisorption mechanism as the rate-limiting factor in the P adsorption process. Similar results on the well-fitting of P adsorption kinetics data with the pseudo-second-order model have been reported (et al. 2013; Jung et al. 2015; Bolbol et al. 2019). Besides, the Qmax values predicted by the pseudo-second-order model were higher for the modified adsorbents rather than the raw materials, which could be related to the size reduction (nano) and iron loading on the surfaces of the studied adsorbents. The higher K2 values in the pseudo-second-order model which refer to the rate constant also confirmed the effect of physical and chemical modifications on the P adsorption enhancement (Table 3).
Table 3

Kinetic parameters of P adsorption by the studied adsorbents

Kinetic modelAdsorbent
CPCNPMCNPZPZNPMZNP
qe(exp) (mg g−10.47 2.24 26.9 0.43 2.17 6.28 
Pseudo-first-order 
qmax (mg g−10.46 2.20 25.52 0.43 2.12 6.18 
K1 (min−10.01 0.02 0.04 0.01 0.02 0.03 
R2 0.99 0.98 0.98 0.96 0.97 0.98 
 SE 0.01 0.11 1.33 0.03 0.15 0.28 
Pseudo-second-order 
qmax (mg g−10.50 2.41 28.44 0.48 2.39 6.46 
K2 (mg g−1min−10.03 0.07 0.14 0.02 0.07 0.11 
R2 0.98 0.99 0.98 0.98 0.98 0.99 
 SE 0.02 0.06 1.25 0.02 0.11 0.13 
Intra-particle diffusion 
Kp (mg g−1 min−0.50.018 0.014 0.251 0.003 0.007 0.011 
C (mg g−10.098 1.764 17.88 0.032 1.900 5.882 
R2 0.94 0.82 0.96 0.75 0.97 0.88 
Kinetic modelAdsorbent
CPCNPMCNPZPZNPMZNP
qe(exp) (mg g−10.47 2.24 26.9 0.43 2.17 6.28 
Pseudo-first-order 
qmax (mg g−10.46 2.20 25.52 0.43 2.12 6.18 
K1 (min−10.01 0.02 0.04 0.01 0.02 0.03 
R2 0.99 0.98 0.98 0.96 0.97 0.98 
 SE 0.01 0.11 1.33 0.03 0.15 0.28 
Pseudo-second-order 
qmax (mg g−10.50 2.41 28.44 0.48 2.39 6.46 
K2 (mg g−1min−10.03 0.07 0.14 0.02 0.07 0.11 
R2 0.98 0.99 0.98 0.98 0.98 0.99 
 SE 0.02 0.06 1.25 0.02 0.11 0.13 
Intra-particle diffusion 
Kp (mg g−1 min−0.50.018 0.014 0.251 0.003 0.007 0.011 
C (mg g−10.098 1.764 17.88 0.032 1.900 5.882 
R2 0.94 0.82 0.96 0.75 0.97 0.88 
Figure 7

Adsorption kinetic models of P on the studied adsorbents.

Figure 7

Adsorption kinetic models of P on the studied adsorbents.

Close modal

The basis of the intra-particle diffusion kinetic model is to determine the diffusion rate of the adsorbate toward the adsorbent at the liquid–solid interface, which can be investigated by plotting the Qt values against the square root of the contact time (t0.5) (Weber & Morris 1963). In this context, a linear diagram crossing the origin of the coordinates implies the intra-particle diffusion mechanism as the main rate-limiting stage in the adsorption process. The P adsorption process generally occurs in three fundamental steps including film diffusion, intra-particle diffusion, and adsorption by the active surface sites. First, the P anions in bulk solution move to the film surrounding the adsorbent following the diffusion gradient and diffuse along it toward the adsorbent surface. Then the adsorption process gradually continues through an intra-particle mass transfer of P anions into the macro pores of the adsorbent, and finally, the physical or chemical adsorption reaction occurs depending on the surface reactivity (Wang et al. 2012). The final equilibrium occurs as a result of reduction in P concentration in solution, diffusion of P into micro-pores of the adsorbent, and electrostatic repulsion of P from the adsorbent surface.

Modeling the P adsorption kinetic data using the intra-particle diffusion model showed a two-phase diagram that did not cross the origin, demonstrating two or more rate-limiting steps in the P adsorption process by all of the studied adsorbents (Figure 8). The Kp and C constants representing the intra-particle diffusion rate and the boundary layer thickness were calculated from the slope and intersection of the second linear phases of the obtained diagrams, respectively (Table 3). It was observed that higher Kp values for the FeCl3-modified adsorbents compared with the raw materials, suggesting the increasing effects of surface-loaded Fe on the intra-particle diffusion mechanism in the P adsorption process. Additionally, the C-value increased from 0.098 to 17.88 in raw and modified coal waste particles, and from 0.032 to 5.882 in natural and modified zeolite, respectively. These results could be attributed to the greater role of external diffusion film in the rate-limiting step after modification of the adsorbents with Fe, resulting in a thicker boundary layer (Olgun et al. 2013).
Figure 8

The intra-particle diffusion plots for P adsorption on the studied adsorbents.

Figure 8

The intra-particle diffusion plots for P adsorption on the studied adsorbents.

Close modal

The feasibility to improve phosphorus adsorption capacity of coal solid wastes was investigated through physical (nano-size reduction) and chemical (FeCl3) modification methods. Characterization analyses including XRD, SEM-EDS, and FTIR confirmed the successful embedding of Fe (III) particles on the surfaces of the prepared adsorbents. The P adsorption capacities of the coal-based adsorbents were enhanced after both modification methods by approximately 9-fold for the physical (0.36–3.23 mg g−1) and 85-fold for the chemical (0.36–30.48 mg g−1) treatments, respectively. Whereas, the zeolite adsorbents showed a 3.5-fold (0.80–2.84 mg g−1) and 9-fold (0.80–6.99 mg g−1) increase in their adsorption capacity at the same modification conditions. The Langmuir equation described the isothermal data well, with a 4-fold maximum P adsorption capacity for MCNP (30.48 mg g−1) rather than the MZNP (6.99 mg g−1). These values demonstrated the higher potential of the coal waste materials to be used as P adsorbents rather than the zeolite ones. The kinetic data were explained by the pseudo-second-order model, implying the chemisorption mechanism involved in the P adsorption process. Besides the surface complexation, ligand exchange and electrostatic attraction were also found as effective mechanisms contributed to the P adsorption process. It can be concluded from the findings of the present study that coal waste materials are great choices to be used as invaluable P adsorbents regarding their desirable adsorption properties, suggesting a win-win paradigm supporting the sustainable management of waste materials as well as environmental protection disciplines.

The authors declare that no funds, grants, or other support were received during the preparation of this manuscript.

All the co-authors read and approved the final manuscript. S.H. contributed to the materials preparation, chemical analysis, and data collection. M.H.-M. contributed to the conceptualization, design, methodology, data analysis, and writing the first draft. H.H. and M.H.F. contributed to the review and editing.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Abdellaoui
Y.
,
Abou Oualid
H.
,
Hsini
A.
,
El Ibrahimi
B.
,
Laabd
M.
,
El Ouardi
M.
,
Giácoman-Vallejos
G.
&
Gamero-Melo
P.
2021
Synthesis of zirconium-modified Merlinoite from fly ash for enhanced removal of phosphate in aqueous medium: experimental studies supported by Monte Carlo/SA simulations
.
Chem. Eng. J.
404
,
126600
.
https://doi.org/10.1016/j.cej.2020.126600
.
Ahmaruzzaman
M.
2010
A review on the utilization of fly ash
.
Prog. Energy Combust. Sci.
36
,
327
363
.
https://doi.org/10.1016/j.pecs.2009.11.003
.
Bahabadi
F. N.
,
Farpoor
M. H.
&
Mehrizi
M. H.
2017
Removal of Cd, Cu and Zn ions from aqueous solutions using natural and Fe modified sepiolite, zeolite and palygorskite clay minerals
.
Water. Sci. Technol.
75
,
340
349
.
https://doi.org/10.2166/wst.2016.522
.
Biswas
B. K.
,
Inoue
K.
,
Ghimire
K. N.
,
Harada
H.
,
Ohto
K.
&
Kawakita
H.
2008
Removal and recovery of phosphorus from water by means of adsorption onto orange waste gel loaded with zirconium
.
Bioresour. Tech.
99
,
8685
8690
.
https://doi.org/10.1016/j.biortech.2008.04.015
.
Bolbol
H.
,
Fekri
M.
&
Hejazi-Mehrizi
M.
2019
Layered double hydroxide-loaded biochar as a sorbent for the removal of aquatic phosphorus: behavior and mechanism insights
.
Arab. J. Geosci.
12
,
1
11
.
https://doi.org/10.1007/s12517-019-4694-4
.
Chubar
N.
,
Kanibolotskyy
V.
,
Strelko
V.
,
Gallios
G.
,
Samanidou
V.
,
Shaposhnikova
T.
,
Milgrandt
V.
&
Zhuravlev
I.
2005
Adsorption of phosphate ions on novel inorganic ion exchangers
.
Colloids Surf. A: Physicochem. Eng. Aspects
255
,
55
63
.
https://doi.org/10.1016/j.colsurfa.2004.12.015
.
Di Capua
F.
,
De Sario
S.
,
Ferraro
A.
,
Petrella
A.
,
Race
M.
,
Pirozzi
F.
,
Fratino
U.
&
Spasiano
D.
2022
Phosphorous removal and recovery from urban wastewater: current practices and new directions
.
Sci. Total. Environ.
153750
.
https://doi.org/10.1016/j.scitotenv.2022.153750
.
Eaton
A.
,
Clesceri
L. S.
,
Rice
E. W.
,
Greenberg
A. E.
&
Franson
M.
2005
APHA: Standard Methods for the Examination of Water and Wastewater
, Centennial edn.
APHA, AWWA, WEF
,
Washington, DC
.
Fazlzadeh
M.
,
Rahmani
K.
,
Zarei
A.
,
Abdoallahzadeh
H.
,
Nasiri
F.
&
Khosravi
R.
2017
A novel green synthesis of zero valent iron nanoparticles (NZVI) using three plant extracts and their efficient application for removal of Cr (VI) from aqueous solutions
.
Adv. Powder Technol.
28
,
122
130
.
https://doi.org/10.1016/j.apt.2016.09.003
.
Feizi
M.
&
Jalali
M.
2016
Sorption of aquatic phosphorus onto native and chemically-modified plant residues: modeling the isotherm and kinetics of sorption process
.
Desalin. Water. Treat.
57
,
3085
3097
.
https://doi.org/10.1080/19443994.2014.981226
.
Freundlich
H. M. F.
1906
Over the adsorption in solution
.
J. Phys. Chem.
57
(
385471
),
1100
1107
.
Ghosal
P. S.
&
Gupta
A. K.
2017
Determination of thermodynamic parameters from Langmuir isotherm constant-revisited
.
J. Mol. Liq.
225
,
137
146
.
https://doi.org/10.1016/j.molliq.2016.11.058
.
Goscianska
J.
,
Ptaszkowska-Koniarz
M.
,
Frankowski
M.
,
Franus
M.
,
Panek
R.
&
Franus
W.
2018
Removal of phosphate from water by lanthanum-modified zeolites obtained from fly ash
.
J. Colloid. Interface Sci.
513
,
72
78
.
https://doi.org/10.1016/j.jcis.2017.11.003
.
Ho
Y. S.
&
McKay
G.
1999
Pseudo-second order model for sorption processes
.
Process Biochem.
34
(
5
),
451
465
.
https://doi.org/10.1016/S0032-9592(98)00112-5
.
Jung
K.-W.
,
Hwang
M.-J.
,
Ahn
K.-H.
&
Ok
Y.-S.
2015
Kinetic study on phosphate removal from aqueous solution by biochar derived from peanut shell as renewable adsorptive media
.
Int. J. Environ. Sci. Technol.
12
,
3363
3372
.
https://doi.org/10.1007/s13762-015-0766-5
.
Kobayashi
Y.
,
Ogata
F.
,
Nakamura
T.
&
Kawasaki
N.
2020
Synthesis of novel zeolites produced from fly ash by hydrothermal treatment in alkaline solution and its evaluation as an adsorbent for heavy metal removal
.
J. Environ. Chem. Eng.
8
,
103687
.
https://doi.org/10.1016/j.jece.2020.103687
.
Kragović
M.
,
Daković
A.
,
Sekulić
Ž
,
Trgo
M.
,
Ugrina
M.
,
Perić
J.
&
Gatta
G. D.
2012
Removal of lead from aqueous solutions by using the natural and Fe (III)-modified zeolite
.
Appl. Surf. Sci.
258
,
3667
3673
.
https://doi.org/10.1016/j.apsusc.2011.12.002
.
Krishnan
K. A.
&
Haridas
A.
2008
Removal of phosphate from aqueous solutions and sewage using natural and surface modified coir pith
.
J. Hazard. Mater.
152
,
527
535
.
https://doi.org/10.1016/j.jhazmat.2007.07.015
.
Lagergren
S. K.
1898
About the theory of so-called adsorption of soluble substances
.
Sven. Vetenskapsakad. Handingarl.
24
,
1
39
.
Langmuir
I.
1916
The constitution and fundamental properties of solids and liquids. Part I. Solids
.
J. Am. Chem. Soc.
38
(
11
),
2221
2295
.
https://doi.org/10.1021/ja02268a002
.
Lee
M. E.
,
Jeon
P.
,
Kim
J.-G.
&
Baek
K.
2018
Adsorption characteristics of arsenic and phosphate onto iron impregnated biochar derived from anaerobic granular sludge
.
Korean J. Chem. Eng.
35
,
1409
1413
.
https://doi.org/10.1007/s11814-018-0057-1
.
Liu
D.
,
Quan
X.
,
Zhu
H.
,
Huang
Q.
&
Zhou
L.
2020
Evaluation of modified waste concrete powder used as a novel phosphorus remover
.
J. Clean. Prod.
257
,
120646
.
https://doi.org/10.1016/j.jclepro.2020.120646
.
J.
,
Liu
H.
,
Liu
R.
,
Zhao
X.
,
Sun
L.
&
Qu
J.
2013
Adsorptive removal of phosphate by a nanostructured Fe–Al–Mn trimetal oxide adsorbent
.
Powder Technol.
233
,
146
154
.
https://doi.org/10.1016/j.powtec.2012.08.024
.
Ngatia
L.
,
Hsieh
Y.
,
Nemours
D.
,
Fu
R.
&
Taylor
R.
2017
Potential phosphorus eutrophication mitigation strategy: biochar carbon composition, thermal stability and pH influence phosphorus sorption
.
Chemosphere
180
,
201
211
.
https://doi.org/10.1016/j.chemosphere.2017.04.012
.
Obradović
N.
,
Filipović
S.
,
Marković
S.
,
Mitrić
M.
,
Rusmirović
J.
,
Marinković
A.
,
Antić
V.
&
Pavlović
V.
2017
Influence of different pore-forming agents on wollastonite microstructures and adsorption capacities
.
Ceram. Int.
43
,
7461
7468
.
https://doi.org/10.1016/j.ceramint.2017.03.021
.
Olgun
A.
,
Atar
N.
&
Wang
S.
2013
Batch and column studies of phosphate and nitrate adsorption on waste solids containing boron impurity
.
Chem. Eng. J.
222
,
108
119
.
https://doi.org/10.1016/j.cej.2013.02.029
.
Park
J.-H.
,
Hwang
S.-W.
,
Lee
S.-L.
,
Lee
J.-H.
&
Seo
D.-C.
2021
Sorption behavior of phosphate by fly ash discharged from biomass thermal power plant
.
Appl. Biol. Chem.
64
,
1
9
.
https://doi.org/10.1186/s13765-021-00614-5
.
Ren
X.
&
Sancaktar
E.
2019
Use of fly ash as eco-friendly filler in synthetic rubber for tire applications
.
J. Clean. Prod.
206
,
374
382
.
https://doi.org/10.1016/j.jclepro.2018.09.202
.
Singh
N.
,
Mithulraj
M.
&
Arya
S.
2019
Utilization of coal bottom ash in recycled concrete aggregates based self compacting concrete blended with metakaolin
.
Resour. Conserv. Recycl.
144
,
240
251
.
https://doi.org/10.1016/j.resconrec.2019.01.044
.
Sparks
D. L.
2003
Environmental Soil Chemistry
.
Academic Press, CA, USA
.
Tandon
P. K.
,
Shukla
R. C.
&
Singh
S. B.
2013
Removal of arsenic (III) from water with clay-supported zerovalent iron nanoparticles synthesized with the help of tea liquor
.
Ind. Eng. Chem. Res.
52
,
10052
10058
.
https://doi.org/10.1021/ie400702k
.
Usman
M.
,
Anastopoulos
I.
,
Hamid
Y.
&
Wakeel
A.
2022
Recent trends in the use of fly ash for the adsorption of pollutants in contaminated wastewater and soils: effects on soil quality and plant growth
.
Environ. Sci. Pollut. Res.
1
20
.
https://doi.org/10.1007/s11356-022-19192-0
.
Wang
Z.
,
Nie
E.
,
Li
J.
,
Yang
M.
,
Zhao
Y.
,
Luo
X.
&
Zheng
Z.
2012
Equilibrium and kinetics of adsorption of phosphate onto iron-doped activated carbon
.
Environ. Sci. Pollut. Res.
19
,
2908
2917
.
https://doi.org/10.1007/s11356-012-0799-y
.
Wang
W.
,
Ye
Z.
&
Li
F.
2016a
Removal of oil from simulated oilfield wastewater using modified coal fly ashes
.
Desalin. Water. Treat.
57
,
9644
9650
.
https://doi.org/10.1080/19443994.2015.1033470
.
Wang
Y.
,
Yu
Y.
,
Li
H.
&
Shen
C.
2016b
Comparison study of phosphorus adsorption on different waste solids: fly ash, red mud and ferric–alum water treatment residues
.
J. Environ. Sci.
50
,
79
86
.
https://doi.org/10.1016/j.jes.2016.04.025
.
Weber
W. J.
&
Morris
J. C.
1963
Kinetics of adsorption on carbon from solution
.
J. Sanitary Eng. Dev.
89
,
31
59
.
https://doi.org/10.1061/JSEDAI.0000430
.
Xu
Y.
,
Dai
Y.
,
Zhou
J.
,
Xu
Z. P.
,
Qian
G.
&
Lu
G. M.
2010
Removal efficiency of arsenate and phosphate from aqueous solution using layered double hydroxide materials: intercalation vs. precipitation
.
J. Mater. Chem.
20
,
4684
4691
.
https://doi.org/10.1039/B926239C
.
Xu
R.
,
Lyu
T.
,
Wang
L.
,
Yuan
Y.
,
Zhang
M.
,
Cooper
M.
,
Mortimer
R. J.
,
Yang
Q.
&
Pan
G.
2022
Utilization of coal fly ash waste for effective recapture of phosphorus from waters
.
Chemosphere
287
,
132431
.
https://doi.org/10.1016/j.chemosphere.2021.132431
.
Zhang
L.
,
Liu
J.
&
Guo
X.
2018
Investigation on mechanism of phosphate removal on carbonized sludge adsorbent
.
J. Environ. Sci.
64
,
335
344
.
https://doi.org/10.1016/j.jes.2017.06.034
.
Zhou
H.
,
Margenot
A. J.
,
Li
Y.
,
Si
B.
,
Wang
T.
,
Zhang
Y.
,
Li
S.
&
Bhattarai
R.
2022
Phosphorus pollution control using waste-based adsorbents: material synthesis, modification, and sustainability
.
Crit. Rev. Environ. Sci. Technol.
52
,
2023
2059
.
https://doi.org/10.1080/10643389.2020.1866414
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).