Nano particles of ZrO2 of average size 10.91 nm are successfully synthesized via green routes from a solvent blend of water and ethylene glycol (4:1 v/v). Bio-extract of seeds of Sapindus plant is employed as stabilizing and/or capping agent and homogeneous method of precipitation is adopted to generate the precipitating agent. The nZrO2 particles are immobilized in aluminum alginate beads (nZrO2-Al- alig). Nano-ZrO2 and beads are investigated as adsorbents for the extraction of phosphate from water. The controlling physicochemical parameters are studied for the maximum phosphate removal using simulate water. The optimum conditions are: pH: 7; sorbent dosage: 0.1 g/100 mL for nZrO2 and 0.08 g/100 mL for beads; equilibration time: 30 min.for nZrO2 and 35 min for beads; initial phosphate concentration: 50 mg/L; temperature: 30 ± 1 °C; 300 rpm. The adsorption capacities are: 126.2 mg/g for nZrO2 and 173.0 mg/g for ‘nZrO2-Al- alig’ and they are higher than many reported in literature. The beads, besides facilitating the easy filtration, are exhibiting enhanced cumulative phosphate-adsorption nature of nanoZrO2 and Al-alginate. X-ray diffraction (XRD), Fourier transform infrared (FTIR), field emission scanning electron microscopy (FESEM) and energy-dispersive X-ray (EDX) investigations are employed in characterizing the adsorbents. Of the various isotherm models analyzed to assess the nature of adsorption, Freundlich model provides the best correlation (R2 = 0.99 for nZrO2 and R2 = 0.99 for ‘nZrO2-Al-alig’), indicating the heterogeneous and multi-layered adsorption process. Thermodynamic studies reveal the endothermic and spontaneous nature of sorption. Pseudo-second-order model of kinetics describes the adsorption well. Spent adsorbents can be regenerated with marginal loss of adsorption capacity until five cycles. The sorbents are successfully applied to remove phosphate from polluted lake water samples.

  • nZrO2 is synthesized via green routes using bio-extracts of seeds of Sapindus plant.

  • nZrO2 and nZrO2 doped aluminium alginate beads are investigated as adsorbents.

  • Phosphate sorption capacities are high:126.2 mg/g for nZrO2 and 173.0 mg/g for beads sorbents are characterized using XRD, FTIR, FESEM and EDX.

  • Adsorption isotherms, thermodynamic and kinetic natures are analysed.

Graphical Abstract

Graphical Abstract
Graphical Abstract

Phosphate contamination of water bodies and its hazardous effects have been globally recognised and, in fact, some countries issued phosphate acts to control or ban the usage of phosphate in detergents and allied materials (Fadiran et al. 2008). The potential sources of phosphate pollution are: excessive application of fertilizers in agricultural fields; inadequately treated effluents from phosphate-based industries viz., detergents, drinks, food materials, etc. (Karageorgiou et al. 2007).

The presence of phosphate in water bodies causes eutrophication. This phenomenon of wild growth of aquatic plants results in depletion of dissolved oxygen (DO) content in water, generates offensive gases due to fermentation of decayed plant parts and increases acidity of water. Further, the sunlight will not reach to the needed depths in waters and thereby, the photosynthesis process is obstructed (Ames & Dean 1970; Batchelor & Dennis 1987). This causes stress to aqua plants and organisms and thereby, loss in eco-cycles. As per WHO, 0.1 ppm is the maximum limit of phosphate for discharging waters (Fadiran et al. 2008).

Removal of phosphate using conventional methods based on precipitation, biological degradation, ion-exchange and electrochemical process (Korngold 1973; PCA 2006; Lacasa et al. 2011; Yamashita & Yamamoto-Ikemoto 2014) are non-economical, involve complicating procedures and need expert supervision and are not adoptable in poor countries like India. Methods based on adsorption are attracting the researchers and in this aspect, many bio-materials are developed as effective adsorbents for the removal of phosphates (Jyothi et al. 2012; Rao & Ravindhranath 2015; Mohan et al. 2017; Sujitha & Ravindhranath 2017).

Developing adsorbents based on nano particles for water remediation, is one of the recent trend in the purification waters (Ozmen et al. 2010; Dehghani et al. 2015; Ghafar A et al. 2015; Ravindhranath & Mylavarapu 2017a, 2017b; Ravulapalli & Ravindhranath 2019). The nano particles are proving to be effective adsorbents in view of increase in surface area, quantum confinements and other morphological changes attributed to the nano particles.

One of the main disadvantages of nano-materials as adsorbents are percolation of water through them is low due to micro or nano-sized voids. This can be avoided by immobilizing the nano particle in beads. In fact, our research groups have successfully developed some composite adsorbents based on beads formed by cross linking of Sodium alginate with Ca2+ and Zn2+ (Mohan et al. 2017; Naga Babu et al. 2017; Sujitha & Ravindhranath 2017). The developing of adsorbents based on nano particles to have maximum sorption capacity and at the same to make easy the filtration process, is one of the potential aspects of water pollution control research (Ravindhranath & Mylavarapu 2017a, 2017b). Further, the production of nano particles via conventional methods using various materials as capping, reducing and stabilizing agents, are less encouraging as most of the chemicals used in these methods are toxic and not eco-friendly. Hence, green methods of synthesis are attractive. Using the green synthesized nano particles in the composite materials for being used as adsorbents, is one of the vital aspect of water remediation techniques.

In the present investigation, nano ZrO2 particles are tried to be synthesized using green routes. Non-toxic and bio-degradable extract of seeds of Sapindus plant is used as capping and/or stabilizing agent. Reagent (OH) is generated by urea hydrolysis method to prevent super-saturation. The mother liquid is a solvent blend of water and ethylene glycol (4:1). At these experimental conditions of high viscosity of media and slow generation of reagent, the nucleation and growth of the particle is slow and when the particle is reached to nano size, the capping agent prevents the further growth and stabilizes. This methodology is found to be successful in this investigation. Thus obtained green synthesized nZrO2 particles are immobilized in the beads produced by cross linking sodium alginate with Al3+. The nZrO2 and beads are investigated as adsorbents for the removal of phosphate from water. The cross linking with Al3+ against the conventional divalent Ca2+ has been done with an endeavour to increase the cumulative sorption nature of nano-ZrO2 and Al-alginate in view of enhanced positive charge of the cross linking metal ion (Al3+) and binding capacity of the functional groups of the beads. The controlling parameters namely, pH, sorbent dosage, equilibration time, initial phosphate concentration and temperature are optimized for the maximum removal of phosphate from simulated water. Natures of sorption, kinetic and thermodynamic parameters have been investigated. The developed procedures are applied to the samples collected at affected lakes in Prakasham District of Andhra Pradesh, India.

Chemicals and reagents

Analytical grade chemicals were used in this work. Solutions were prepared with double distilled water.

Green routes of synthesis of nano ZrO2

The conventional capping and stabilizing agents used in the synthesis of nano ZrO2 are toxic and costly. So investigations are being made to identify the plant extracts as substitutes. In this process, we identified that extract of Sapindus plant seed possesses good surfacting nature and it is employed in the present work. Further, the success of limiting the growth of particles to nano size depends upon particle growth conditions, mainly low super-saturation, viscosity of the mother liquor and entrapping ability of the capping agent (Vogel 1961). As per the classical theory of precipitations, the growth of particles comprises three important steps: super-saturation, nucleation and aggregation. When the growth of particle is at nano dimensions, the growth is to be arrested by employing a capping agent. Super-saturation influences particle size (Vogel 1961). The rate of nuclei formation is proportional to Q-S/S where Q = concentration of solute and S = equilibrium solubility. The speed of aggregation of molecules is to be retarded by increasing the viscosity of the medium. At low super-saturation and high viscosity of medium, the aggregation of particles is less and before the particles cross the nano size, the growth is to be stopped by capping agents (Vogel 1961; Sun et al. 2007; Pandian et al. 2015; Ravindhranath & Mylavarapu 2017a, 2017b).

In the present investigation, to achieve the low super saturation, we employed classical methods of homogenous precipitations (Vogel 1961). The precursor was zirconyl chloride and the precipitating agent was OH generated by the hydrolysis of urea. The viscosity of the mother liquid was enhanced by using ethylene glycol. With low super-saturation, high viscosity and by using an effective bio-surfactant derived from Sapindus plant seeds as capping and stabilizing agent, nano ZrO2 were successfully synthesized in this work as is evident from the following narration.

Preparation of bio-extract

Seeds of Sapindus plant were collected, washed and dried. The dried seeds were skinned out and the leathery bio-parts were dried at 90 °C. 5.0 g of the bio-material was extracted repeatedly with hot water. The extractions were combined, filtered and the filtrate was diluted to 500 mL. Thus obtained extract was preserved at 4 °C in a refrigerator.

Nano ZrO2 synthesis

Requisite amount of ZrOCl2.8H2O was dissolved in a solvent blend of water and ethylene glycol (4:1 V/V) to obtain 0.05 N solutions. To 100 mL of this solution taken in a 250 mL beaker, 20 mL of bio-extract obtained from the seeds of Sapindus plant was added and stirred for 30 min using a magnetic stirrer at 600 rpm. Then 10 g A.R. urea was added slowly and the resulting solution was heated to 80–90 °C with constant stirring until the pH was increased to pH 9. Then the heating was stopped and stirring was continued for 15 h. The solution was centrifuged and the particles were washed with distilled water and dried at 90 °C. Then the obtained material was calcinated at 500 °C for 6 h in a muffle furnace.

The experimental conditions adopted in this work for generating slowly the nuclei of metal hydroxide, provides sufficient time for capping agent to limit the growth of the particle to nano size. The particles were characterized with X-ray diffraction (XRD) and by using Scherrer formula (Dorofeeva et al. 2012), the average crystalline size for ZrO2 sample was 10.91 nm.

Aluminium alginate beads embedded with nano ZrO2 (nZrO2-Al-alig)

In the present work, the nano ZrO2 was immobilised in beads synthesized by cross linking sodium alginate with Al3+. The conventional beading with Ca2+ was replaced with Al3+ in this investigation with an aim to derive cumulative advantages of nano size of ZrO2 and residual positive charges of Al3+ in enhancing the adsorption ability of the composite towards negatively charged phosphate ions. In this investigation, good beads of Al-alginate embedded with nano ZrO2, were successively synthesized.

2.0% w/v sodium alginate solution in distilled water was heated slowly to reach 80 °C to obtain a homogenous gel. To this gel, 1.5 g of nano ZrO2 was added with constant stirring. The resulting solution was stirred (600 rpm) until a homogeneous solution was obtained. Then the solution was cooled to room temperature and was added to 3% acidic solution of aluminium chloride in HCl (2N) at 5 °C taken in a beaker, drop wise using a dropper having a good capillary delivery tube. The moment the drops were touched to Al3+ solution, excellent beads were developed. The formed beads were allowed to be in contact with the mother liquid for an overnight for the complete digestion to occur. Then beads were filtered, dried at 70 °C for 5 h. The beads obtained were named as nZrO2-Al-alig(nano zirconium oxide embedded aluminium alginate). The synthesis of the beads is graphically summarized in Figure 1.

Figure 1

Methodology for preparation of nZrO2 and nZrO2 embedded with aluminum alginate beads (ZrO2-Al-alig).

Figure 1

Methodology for preparation of nZrO2 and nZrO2 embedded with aluminum alginate beads (ZrO2-Al-alig).

Close modal

Surface characterization

XRD, Fourier transform infrared (FTIR), field emission scanning electron microscopy (FESEM) and energy-dispersive X-ray (EDX) investigations were made to study the nature of adsorption of the composite material: nZrO2 and ‘nZrO2-Al-alig’ towards phosphate.

X-ray powder diffract meter with CuKα (λ= 0.154 nm) radiation over the 2θ range of 20–80 °C was used to record XRD patterns of nZrO2. FESEM images of nZrO2 and nZrO2-Al-alig were taken at 500 to 10,000× magnifications at the acceleration voltage of 15.0 kV. FTIR and EDX spectra of nZrO2 and nZrO2-Al-alig before and after adsorption of phosphate were recorded. The results are presented in Figures 25.

Figure 2

XRD patterns of zirconium oxide nano particles (nZrO2).

Figure 2

XRD patterns of zirconium oxide nano particles (nZrO2).

Close modal
Figure 3

FTIR spectra: (a) nZrO2 and (b) nZrO2-Al-alig before and after adsorption of phosphate ion.

Figure 3

FTIR spectra: (a) nZrO2 and (b) nZrO2-Al-alig before and after adsorption of phosphate ion.

Close modal
Figure 4

SEM images: nZrO2 (a) before and (b) after; ZrO2-Al-Alig (c) before and (d) after – phosphate adsorption.

Figure 4

SEM images: nZrO2 (a) before and (b) after; ZrO2-Al-Alig (c) before and (d) after – phosphate adsorption.

Close modal
Figure 5

EDX images: nZrO2 (a) before and (b) after adsorption of phosphate; nZrO2-Al-alig (c) before and (d) after adsorption of phosphate.

Figure 5

EDX images: nZrO2 (a) before and (b) after adsorption of phosphate; nZrO2-Al-alig (c) before and (d) after adsorption of phosphate.

Close modal

Adsorption experiment

Batch adsorption experiments were employed as described in literature (A.R.K. 1995; Metcalf Eddy 2003). 100 mL of phosphate solutions of concentration: 50 mg/L, were taken into 250 mL conical flasks and to them, different quantities of nZrO2 and nZrO2-Al-alig (0.025 to 0.2 g) were added. Then, initial pHs of solutions were adjusted from 2 to 12 by using 0.1 M HCl/0.1 M NaOH. Then the flasks were agitated in orbital shaker for desired time at 300 rpm at room temperature (30 °C). After a definite time of agitation, the solutions were centrifuged/filtered. The residual phosphate in the filtrates was analysed spectrophotometrically adopting ‘Molybdenum Blue’ method (APHA (American Public Health Association) 1985) using UV-Visible-159 model spectrophotometer (ELICO). The percentage removal and adsorbed amount of phosphate (adsorbent capacity) were calculated by the equations: % removal = and qe = where m= mass of adsorbent (g), V= volume of the solution (L), C0 and Ci are the initial and final concentrations (mg/L) of phosphate ions (Dehghani et al. 2018a). The influence of pH, sorbent dosage, initial concentration of phosphate ions, co-ions and temperature on the % removal of phosphate were assessed. The results are presented in Figures 69.

Figure 6

(a) Evaluation of pHzpc of nZrO2-Al-alig; (b) evaluation of pHzpc of nZrO2; (c) effect of pH; (d) effect of sorbent dosage; (e) effect of contact time; (f) effect of initial phosphate concentration; (g) and (h) effect of temperature.

Figure 6

(a) Evaluation of pHzpc of nZrO2-Al-alig; (b) evaluation of pHzpc of nZrO2; (c) effect of pH; (d) effect of sorbent dosage; (e) effect of contact time; (f) effect of initial phosphate concentration; (g) and (h) effect of temperature.

Close modal
Figure 7

Adsorption isotherm models: (a) Freundlich; (b) Langmuir; (c) Temkin; and (d) Dubinin-Radushkevich.

Figure 7

Adsorption isotherm models: (a) Freundlich; (b) Langmuir; (c) Temkin; and (d) Dubinin-Radushkevich.

Close modal
Figure 8

Kinetic models: (a) pseudo-first-order; (b) pseudo-second-order; (c) Elovich; and (d) Bangham's kinetics.

Figure 8

Kinetic models: (a) pseudo-first-order; (b) pseudo-second-order; (c) Elovich; and (d) Bangham's kinetics.

Close modal
Figure 9

Effect of (a) interfering on the % removal of phosphate ions and (b) no. of regenerations vs % removal phosphate.

Figure 9

Effect of (a) interfering on the % removal of phosphate ions and (b) no. of regenerations vs % removal phosphate.

Close modal

Characterization analysis

X-ray diffraction (XRD)

XRD pattern of nano ZrO2 is depicted in Figure 2. The XRD pattern shows cubic phase as per JCPDS card No. 27-0997. The broadening of XRD peaks indicates fine size for ZrO2 particles. The average size of ZrO2 particles was calculated using Scherrer formula (Dorofeeva et al. 2012). The average size of ZrO2 particles is: 10.91 nm. This reflects the successful adoption of the present green route of synthesis and the classical homogenous method of precipitations in obtaining nanoZrO2.

Fourier transform infrared spectroscopy (FTIR)

The FTIR spectra for nZrO2 and ‘nZrO2-Al-alig’ are presented in Figure 3(a) and 3(b). IR spectrum of nZrO2 (prepared from green routes) resembles that of previously reported in literature (Bi et al. 2017). It is interesting to note the difference between before and after spectra. The frequencies at 745, 930, 1,072.6 and 1,146 cm−1of nZrO2 are shifted to 751, 937, 1,075, 1,153.6 cm−1 with emphatic increase in intensities after adsorption of phosphate. A new peak in the after-adsorption spectrum at 2,352.2 cm−1 is due to phosphate.

In the case of nZrO2-Al-alig, there is clear evidence that phosphate is adsorbed. The spectrum before adsorption has a broad peak between 3,036 to 3,275 cm−1 with an apex at 3,150 cm−1 pertains to –OH stretching. The broadness indicates hydrogen bonding. In addition to this peak, additional sharp peaks of different intensities are observed at 3,415 cm−1, 3,490 cm−1 and 3,556 cm−1 in after-IR adsorption spectrum. These are due to interactions of phosphate with the adsorbent. Further, a new additional peak at 2,049.8 cm−1 (after adsorption) is due to the surface structural changes caused by the adsorption of phosphate. Small peaks before adsorption pertaining to –C-O stretching at 1,038 cm−1 and 1,102 cm−1 are developed into a sharp merged peak with improving intensity at 1,143 cm−1. This is due to the adoption of HPO42−. A sharp small peak at 778 cm−1 before adsorption pertains to Zr-O (stretching), has been reduced to a very small peak due to the formation of ‘Zr-PO43−’. Consequently a new sharp peak with medium intensity is appeared at 630.7 cm−1. It is due to asymmetric stretching of ‘Zr/Al- PO43−. All these findings emphatically prove the adsorption of phosphate.

Field emission scanning electron microscopy (FESEM)

FESEM images of nZrO2 and ‘nZrO2-Al-alig (beads)’ are presented in Figure 4. The image before adsorption has cavities, small pores, and irregular surface. These have been converted to smooth surfaces after adsorption. These morphological changes reveal phosphate adsorption ‘onto’ nZrO2 and nZrO2-Al-alig.

Energy-dispersive X-ray spectroscopy (EDX)

The EDX spectra of nZrO2 and nZrO2-Al-alig before and after phosphate adsorption are depicted in Figure 5. It is clearly seen that phosphate peaks are noticed in the EDX spectra taken after adsorption (Figure 5(b) and 5(d)) and they are missing in the EDX spectra taken before adsorption (Figure 5(a) and 5(c)). Hence, it may conclude that the phosphate is adsorbed ‘onto’ the surface of both the adsorbents.

Effect of adsorption parameters on phosphate removal

Various experimental conditions were optimized for the maximum extraction of phosphate. Investigations were carried out by varying one parameter while maintaining the other experimental conditions constant. Investigations were made by varying the pH from 2 to 12, agitation time: 5 to 60 min, adsorbent dosage from 0.025 to 0.2 g/100 mL, at initial concentration: 50 mg/L and rpm: 300 and temperature 30 °C for ascertaining the optimum extraction conditions. The effect of initial concentration of phosphate and temperature on the adsorption efficiency of the two adsorbents was also investigated by maintaining all other parameters at optimum levels while varying only the parameter under study.

pH effect

The effect of pH on the adsorption of phosphate onto the surface of nZrO2 and nZrO2-Al-alig are depicted in Figure 6(c). The maximum removal of 93.0% for nZrO2 and 99.5% for nZrO2-Al-alig are observed at pH 7 and above or below, the removal is less favoured. H3PO4 is a tri-basic acid and it has pKa values: 2.14, 7.2, 12.4. So below pH 3, the main species is H3PO4, between pH 3–9, H2PO4 and HPO42− co-exist and between pH 9–12, HPO42− and PO43− co-exist and above pH 12, main species is PO43− (Vogel 1961). At pH 7, the H2PO4 and HPO42− are presented in almost equivalent concentrations and in fact, these ions form a buffer and the buffering capacity is maximum at pH ∼7. pH 0 values for ZrO2 and ZrO2-Al-alig are at pH 6.9 and 7.1, respectively (Figure 6(a) and 6(b)). So, above these pH values, the surface of the adsorbents possesses negative charge. As the charge of phosphate species H2PO4 and HPO42− is also negative, the repulsion is caused and so, there is less adsorption. When pH of solution is below 7, the surface hydroxyl groups of both the adsorbents are protonated and attain positive charge. So they attract the negative species of phosphate ions. At pH 7, both the negative species are in equimolar proportions and hence, there is maximum interaction of the species towards the surface. So, the extraction is at its maximum at pH ∼7. With the decrease in pH of the solution below 7, doubly charged HPO42− species are converted to mono charged negative H2PO4. This results in less interaction of phosphate species with the positively charged surface when compared to the species at pH 7. So, adsorption decreases as pH decreases from 7. For maximum adsorption, the surface of the adsorbents must be charged positively while the phosphate species possess high cumulative negative charge. This condition is acquired by the two adsorbents at pH 7 and so, maximum adsorption is observed.

Effect of adsorbent concentration

Investigations were made to optimize the adsorbent concentration for maximum extraction of phosphate. The concentrations were varied from 0.025 to 0.2 g/100 mL while maintaining other extraction conditions at optimum levels. Observations are presented in Figure 6(d). Adsorption is almost linearly increased up to 0.1 mg/100 mL for nZrO2 and 0.08 mg/100 mL for nZrO2-Al-alig. After that dosage, a steady state is resulted. The maximum removal is found to be 93.1% for nZrO2 and 99.5% for nZrO2-Al-alig. The linear increase in the adsorption with progressive raise in concentration of adsorbents is owing to the increase in sorption sites. But at high concentrations, active sites prevailing on surface are blocked due to agglomeration and hence, steady state.

Effect of agitation time

After equilibration of the phosphate solutions at different time intervals with the two adsorbents while maintaining constant the other optimum conditions of extractions, the samples were assayed for the residual phosphate. The findings are depicted in Figure 6(e). 99.5% extraction was observed after 35 min of equilibration for nZrO2-Al-alig and 93.1% after 30 min for nZrO2.

As is viewed from the Figure 6(e), the adsorption is progressively increased with an increase in time. After 30 min for nZrO2 and 35 min for nZrO2-Al-alig, steady states are reached. Initially many sorption sites are available for the adsorption of phosphate. So, the adsorption is maximum at the beginning. But with time, the active sites are progressively used up. As the amount of adsorbent is fixed, the sites available are also limited. Hence, with time, the rate of adsorption is decreased. After certain time, all sites are engaged in the equilibrium state of adsorption/desorption and hence steady state is resulted (Krishna Mohan et al. 2017).

Initial concentration effect

The initial concentration of phosphate solution has influence on the adsorption of phosphate ions. It was investigated by varying the initial concentrations from 60 to 150 mg/L while keeping constant (optimum) all other extraction conditions. The results are depicted Figure 6(f). With the raise in phosphate concentration from 60 to 150 mg/L, the adsorption falls from 92.37 to 72.24% for nZrO2 and 97.5 to 79.6% for nZrO2-Al-alig(bead).

As the adsorbent concentration is fixed (0.1 g/100 mL for nZrO2 and 0.08 g/100 mL for nZrO2-Al-alig), the active sites are limited. At low initial concentration of phosphate (adsorbate) active sites available per one molecule of phosphate are many and probabilities of interactions are more and so, adsorption is more. As the phosphate concentrations are increased, the sites offered per molecule of phosphate for adsorption process to occur is decreased and hence, % removal is less.

Effect of temperature

Temperature influences the adsorption process profoundly. Its effect was studied at temperatures 303, 313, 323 and 333 K on the adsorption of phosphate onto nZrO2 and nZrO2-Al-alig by equilibrating 100 mg/L of phosphate solution at other optimum conditions of extraction. The results are depicted in Figure 6(g). As the temperature increases from 303 to 333 K, the percentage removal of phosphate is increased from 86.3 to 92.2% for nZrO2 and 88.2 to 97.3% for nZrO2-Al-alig.

The rise in percentage removal with rise in temperature indicates the adsorption process is favourable at high temperatures. At elevated temperatures, the pores in the adsorbents are enlarged and, moreover, some new sites may be formed. These changes in surface morphology create additional paths for the phosphate molecules to reach the inner laying active sites. Thereby, hidden active sites in the folds of layers of the adsorbents are open for the adsorption process.

Elevated temperatures help phosphate ions to overcome activation energy barrier at the interface between the solution and the adsorbent. This enhances the mobility or diffusion of phosphate into the surface layers of the adsorbents. Due to these effects, the adsorption is more with raise in temperature (Mansooreh et al. 2014).

Thermodynamic studies

To assess the nature of adsorption process, various thermodynamic parameters namely, ΔH, ΔS and ΔG were evaluated as per the procedures in literature (Fan & Zhang 2018). The pertaining equations are: , and where, qe = amount of phosphate adsorbed at equilibrium, Kd = distribution coefficient, Ce = equilibrium concentration of phosphate, R = gas constant and T = temperature (Kelvin), ln KD vs. (1/T) plots, Figure 6(h), were used to evaluate ΔH and ΔS. Values obtained for all the parameters are presented in Table 1.

Table 1

Thermodynamics parameters for nZrO2 and nZrO2- Al-alig

AdsorbentΔH (kJ/mol)ΔS (J/mol)ΔG (kJ/mol)
R2
303313323333
nZrO2 16.77 70.28 −4.53 −5.23 −5.93 −6.63 0.997 
nZrO2- Al-alig 33.305 127.32 −5.272 −6.55 −7.82 −9.1 0.998 
AdsorbentΔH (kJ/mol)ΔS (J/mol)ΔG (kJ/mol)
R2
303313323333
nZrO2 16.77 70.28 −4.53 −5.23 −5.93 −6.63 0.997 
nZrO2- Al-alig 33.305 127.32 −5.272 −6.55 −7.82 −9.1 0.998 

Negative values for ΔG signify the spontaneity of adsorption process for both the adsorbents under investigation towards adsorbate (phosphate). Further, the increase in negative values of ΔG with the rise in temperature reflects the more spontaneity of the adsorption process at elevated temperatures.

ΔH values for both the adsorbents are positive and the value for nZrO2 (16.77) kJ/mol is less than nZrO2-Al-alig (33.305 kJ/mol). The positive values suggest the endothermic nature (Ngah & Hanafiah 2008). At high temperatures, the conditions are conducive for adsorbate (phosphate) to overcome the resistance at the solution–adsorbent interface and thereby, accelerating the adsorption process. Further, the magnitude of ΔH values indicates the nature of sorption process: physical adsorption 2.1–20.9 kJ/mol; chemical adsorption: 20.9–418.4 kJ/mol (Sun et al. 2012). As the values of ΔH are 16.77 kJ/mol and 33.305 kJ/mol for nZrO2 and nZrO2-Al-alig, respectively, the adsorption of phosphate is more inclined towards chemical in nature.

ΔS values are positive. This signifies the more disorder at the solid–liquid interface. Further higher values for nZrO2-Al-alig than nZrO2, reflect more disorderliness in the former than the latter. If the disorder is more, the randomness in the movement of phosphates (adsorbate) at the surface is more. This enables the phosphate molecules to penetrate deep into the layers of adsorbents. So, the adsorption is expected to be more in the case of nZrO2-Al-alig than nZrO2 (Borah et al. 2015).

Adsorption isotherms

The sorption nature was assessed by applying various adsorption isotherm models as has been described elsewhere (Freundlich 1906; Langmuir 1918; Temkin & Pyzhev 1940; Dubinin 1947). The results are presented in Table 2. The R2 values fall in the order: Freundlich (0.99) > Temkin (0.985) > Langmuir (0.97) > Dubinin-Radushkevich (0.78) for nZrO2; Freundlich (0.997) > Langmuir (0.965) > Temkin (0.963) > Dubinin-Radushkevich (0.56) for nZrO2-Al-alig. The values indicate that Freundlich model (Figure 7(a)) is fit to understand the adsorption nature of both the adsorbents and, further, it indicates the heterogeneous and multilayer of adsorption (Dehghani et al. 2016). Kf and n values were noted from the plots. As 1/n for nZrO2 (0.303) and nZrO2-Al-alig (0.248) are in the range: 0.1 < 1/n < 1, the adsorption of phosphate onto the surface of the said adsorbents is favourable.

Table 2

Parameters of adsorption isotherms

AdsorbentFreundlich isothermLangmuir isothermTemkin isothermDubinin-Radushkevich isotherm
nZrO2 Slope 0.303 0.0077 24.14 −2.6E-6 
Intercept 3.56 0.0282 17.32 4.55 
R2 0.997 0.97 0.985 0.78 
RL/1/n/B/E 1/n = 0.303 RL = 0.058 B = 24.14 E = 0.455 
nZrO2-Al-alig Slope 0.248 0.018 22.81 −3.63E-7 
Intercept 3.92 0.0568 43.695 4.58 
R2 0.997 0.965 0.963 0.56 
RL/1/n/B/E 1/n = 0.248 RL = 0.0121 B = 22.81 E = 1.17 
AdsorbentFreundlich isothermLangmuir isothermTemkin isothermDubinin-Radushkevich isotherm
nZrO2 Slope 0.303 0.0077 24.14 −2.6E-6 
Intercept 3.56 0.0282 17.32 4.55 
R2 0.997 0.97 0.985 0.78 
RL/1/n/B/E 1/n = 0.303 RL = 0.058 B = 24.14 E = 0.455 
nZrO2-Al-alig Slope 0.248 0.018 22.81 −3.63E-7 
Intercept 3.92 0.0568 43.695 4.58 
R2 0.997 0.965 0.963 0.56 
RL/1/n/B/E 1/n = 0.248 RL = 0.0121 B = 22.81 E = 1.17 

From the Langmuir isotherms (Figure 7(b)), RL (dimensionless separation factor) values were calculated by using R L= 1 (1 + aLC0). The RL values are 0.058 for nZrO2 and 0.0121 for nZrO2-Al-alig. As the RL values are falling in the range: 0 < RL< 1, the adsorption is favorable (Dehghani et al. 2018b).

Further, linear form of Temkin equation and Dubinin-Radushkevich equations were applied (Figure 7(c) and 7(d)). Isothermal constants along with the correlation coefficient values were presented in Table 2. The mean free energy (E) and heat of sorption (B) were calculated. As B value for the adsorbents: 24.14 for nZrO2 and 22.81 for nZrO2-Al-alig are more than 20 kJ/mol, the adsorption is more oriented towards chemical interactions than physical interactions.

Kinetics of adsorption

The kinetics of adsorption were evaluated using various conventional models like pseudo-first-order (Equation (1)) (Corbett 1972), pseudo-second-order (Equation (2)) (Ho & McKay 1999), Bangham's pore diffusion model (Equation (3)) (Ho et al. 2000), and Elovich equation (Equation (4)) (Lagergren 1898) as described in the literature. The concerned equations employed are:
(1)
(2)
(3)
(4)

The results are depicted in Figure 8(a)–8(d) and Table 3. On perusal of the R2 values, pseudo-second-order is a better fit to explain the kinetics of adsorption. This model is followed by Elovich model, then by Bangham's pore diffusion model and lastly by pseudo-first-order mode in the case of both the adsorbents.

Table 3

Kinetics of adsorption

AdsorbentsPseudo-first-orderPseudo-second-orderElovichBangham's pore diffusion
nZrO2 Slope −0.032 0.0178 7.62 0.311 
Intercept 1.502 0.0927 19.05 0.825 
R2 0.94 0.998 0.956 0.942 
nZrO2- Al-alig Slope −0.042 0.018 7.85 0.323 
Intercept 1.51 0.015 20.77 −0.916 
R2 0.87 0.995 0.96 0.934 
AdsorbentsPseudo-first-orderPseudo-second-orderElovichBangham's pore diffusion
nZrO2 Slope −0.032 0.0178 7.62 0.311 
Intercept 1.502 0.0927 19.05 0.825 
R2 0.94 0.998 0.956 0.942 
nZrO2- Al-alig Slope −0.042 0.018 7.85 0.323 
Intercept 1.51 0.015 20.77 −0.916 
R2 0.87 0.995 0.96 0.934 

Effect of interfering anions

The interferences caused by commonly present co-anions (Cl, F, HCO3 NO3 and SO42−) in water were assessed. For this, five-fold excess of the said ions was added to the known amount of phosphate solutions. Thus prepared simulated solutions were subjected to extraction under the optimum conditions established in this investigation with both the adsorbents. Percent removals were calculated. The results are presented in Figure 9(a). It can be seen from the figure that chlorides, fluorides, nitrate and bicarbonate have marginal effect; sulphate has some interference.

Regeneration of sorbents

To achieve cost effectiveness, the spent adsorbents are to be regenerated. The conditions to activate the spent adsorbents are to be optimized. For this object, many eluents at various concentrations were tried. 0.1 N Na2SO4 was a successful regenerating agent. The spent adsorbents were soaked in 0.1 N Na2SO4 solutions overnight. The adsorbent is filtered, washed with distilled water and dried at 105 °C. Thus regenerated adsorbents were employed for investigations. This cycle of regeneration and reuse was continued a number of times. The results are presented in Figure 9(b).

The efficiency of adsorption is marginally affected up to five cycles for both the adsorbents. For nZrO2, percent removal has been decreased from 92.4 (first) to 80.9 after the fifth cycle of regeneration. In the case of nZrO2-Al-alig, percent removal has decreased from 99.5 (first) to 90.3 (fifth). It may be concluded that by repetitive use of the adsorbents, complete removal of phosphate from water can be achieved.

Applications

The adoptability of nZrO2 and nZrO2- Al-alig as adsorbents was assessed with respect to the samples collected from different lakes in Prakasham District, A.P., India. The phosphate pollution problem is severe in these localities. So, the present developed new adsorbents were applied to different water samples collected in various localities in the district. The results are presented in Table 4. From the data, it may be concluded that both the adsorbents have effectively reduced the concentrations. It is more than 97.0% with nZrO2 and complete removal is observed with nZrO2- Al-alig.

Table 4

Applications: adoptability of nZrO2 and ‘nZrO2- Al-alig’ as adsorbents for real water samples

AdsorbentsSamplesCi (mg/lit)aCf (mg/lit)aPercentage removal
nZrO2 Sample 1 29.30 0.82 97.2 
Sample 2 22.5 100 
Sample 3 18.7 100 
Sample 4 26.2 0.36 98.6 
Sample 5 16.8 100 
nZrO2-Al-alig Sample 1 29.8 0.06 99.79 
Sample 2 18.4 100 
Sample 3 24.7 100 
Sample 4 22.5 100 
Sample 5 27.9 100 
AdsorbentsSamplesCi (mg/lit)aCf (mg/lit)aPercentage removal
nZrO2 Sample 1 29.30 0.82 97.2 
Sample 2 22.5 100 
Sample 3 18.7 100 
Sample 4 26.2 0.36 98.6 
Sample 5 16.8 100 
nZrO2-Al-alig Sample 1 29.8 0.06 99.79 
Sample 2 18.4 100 
Sample 3 24.7 100 
Sample 4 22.5 100 
Sample 5 27.9 100 

aAverage of five determinations; standard deviation: ±0.055%.

Comparative study

The adsorbents developed in this investigation are compared with nano adsorbents reported in literature for the removal of phosphate from water. The data are summarized in Table 5.

Table 5

Comparison of phosphate uptake capacity with other adsorbents from literature

S no.Adsorbent (nanomaterials)Optimum pHInitial conc. of phosphate ions (mg/L)Adsorption capacity (mg/g)Reference
Fe-Mg-La 6.4 100 48.3 Yu & Paul Chen (2015)  
Fe–Ti 6.8 50 35.4 Lu et al. (2015)  
CNT 5.6 200 14.5 Mahdavi & Akhzari (2016)  
Zr(IV)-Chitosan 10 1,000 149 Sowmya & Meenakshi (2014)  
GO-NZVI – 20 14.71 Liu et al. (2014)  
10 GO-ZrO2 20 16.45 Zong et al. (2013)  
11 Ag/TAC 30 13.61 Nguyen et al. (2020)  
12 La–Zr/Peel 6.21 200 40.21 Muhammad et al. (2020)  
13 Zr/Peel 6.21 200 30.2 
14 Fe3O4-ASC 7.2 150 128 Jiang et al. (2017)  
15 EL-MNP-Zeolite 6.35 100 38.91  Xu et al. (2019
16 nZrO2 200 126.2 Present work 
17 nZrO2-Al-alig 200 173 Present work 
S no.Adsorbent (nanomaterials)Optimum pHInitial conc. of phosphate ions (mg/L)Adsorption capacity (mg/g)Reference
Fe-Mg-La 6.4 100 48.3 Yu & Paul Chen (2015)  
Fe–Ti 6.8 50 35.4 Lu et al. (2015)  
CNT 5.6 200 14.5 Mahdavi & Akhzari (2016)  
Zr(IV)-Chitosan 10 1,000 149 Sowmya & Meenakshi (2014)  
GO-NZVI – 20 14.71 Liu et al. (2014)  
10 GO-ZrO2 20 16.45 Zong et al. (2013)  
11 Ag/TAC 30 13.61 Nguyen et al. (2020)  
12 La–Zr/Peel 6.21 200 40.21 Muhammad et al. (2020)  
13 Zr/Peel 6.21 200 30.2 
14 Fe3O4-ASC 7.2 150 128 Jiang et al. (2017)  
15 EL-MNP-Zeolite 6.35 100 38.91  Xu et al. (2019
16 nZrO2 200 126.2 Present work 
17 nZrO2-Al-alig 200 173 Present work 

It may be inferred from the table that the sorbents developed in this investigation, namely nZrO2 and nZrO2-Al-alig, have higher adsorption capacities than hitherto reported in literature.

In the present investigation, nano particles of zirconium oxide of average size, 10.91 nm, are successfully synthesized via green routes. In this synthesis, bio-extracts of seeds of Sapindus plant are employed as stabilizing and capping agents. Water: ethylene glycol (4:1 v/v) media is used to slow down the nucleation of ZrO2 particles. Further, a homogenous method of precipitation is adopted while generating slowly the reagent, OH by urea hydrolysis. At these experimental conditions of low super-saturation and high viscosity of medium, the aggregation of particles is less and before the particles acquire nano size, the growth is prevented by the capping agent.

Two adsorbents are prepared: nZrO2 and nZrO2-Al- alig. These adsorbents are investigated for their phosphate absorptivity by varying extraction conditions using batch methods. The optimum conditions established for the maximum removal of phosphate are: pH 7; sorbent dosage: 0.1 g/100 mL (nZrO2) and 0.08 g/100 mL (nZrO2-Al- alig); equilibration time: 30 min (nZrO2) and 35 min (nZrO2-Al-alig); initial concentration of phosphate: 50 mg/L; temperature: 30 ± 1 °C; 300 rpm.

The adoption capacity for nZrO2 is 126.2 mg/g and for nZrO2-Al- alig is 173 mg/g. These are higher than hitherto reported investigations. The beads, besides facilitating easy filtration, enhance the adsorption capacity. The increase in the adsorption capacity may be due to the residual positive charges of Al3+ which is used as crossing agent and also due to the binding capacity of functional groups pertaining to the Al-alginate beads.

The sorption of phosphate ‘onto’ the surface of nZrO2 and nZrO2-Al- alig are characterized by various techniques including XRD, FTIR, FESEM and EDX. The sorption mechanism is investigated using various isotherm models. Freundlich model fits well for both the adsorbents. Thermodynamic studies reveal the endothermic and spontaneous nature of sorption. Pseudo-second-order model of kinetics fits well for both the adsorbents.

Spent adsorbents can be regenerated. Even after five cycles of regeneration of the adsorbents, a significant amount of phosphate is removed to an extent of 80.9% for nZrO2 and 90.3% for nZrO2-Al-alig. The beads' structure is so strong that even after five regenerations, they are not deteriorated.

One of the major advantages of these two adsorbents nZrO2 and nZrO2-Al- alig, is that they show substantial phosphate adsorptivity in neutral pH (∼7). Generally, the polluted lake waters or natural waters have pH 6–8 and, hence, the present developed adsorbents can be applied directly without adjusting the pH of water samples and, thereby, avoid one step in the purification of water. The adsorbents nZrO2 and nZrO2-Al- alig, are successfully applied to remove phosphate from water samples of polluted lakes.

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

Ames
L. L.
Jr.
Dean
R. B.
1970
Phosphorus removal from effluents in alumina columns
.
Journal (Water Pollution Control Federation)
,
R161
R172
.
APHA (American Public Health Association)
1985
Standard Methods for the Examination of Water and Wastewater
.
APHA
,
Washington, DC, USA
.
A.R.K.
1995
Trivedy, Pollution Management in Industries
, 2nd edn.
Environmental Publications
,
Karad
,
India
.
Batchelor
B.
Dennis
R.
1987
A surface complex model for adsorption of trace components from wastewater
.
Journal (Water Pollution Control Federation)
,
1059
1068
.
Bi
C.
Li
J.
Peng
L.
Zhang
J.
2017
Bio fabrication of Zinc oxide nanoparticles and their in-vitro cytotoxicity towards gastric cancer (MGC803) cell
.
Biomedical Research
28
(
5
),
2065
2069
.
Borah
I.
Goswami
M.
Phukan
P.
2015
Adsorption of methylene blue and eosin yellow usingporous carbon prepared from tea waste: adsorption equilibrium, kinetics and thermodynamics study
.
Journal of Environmental Chemical Engineering
3
(
2
),
1018
1028
.
https://doi.org/10.1016/j.jece.2015.02.013
.
Corbett
J. F.
1972
Pseudo-first-order kinetics
.
Journal of Chemical Education
49
(
10
),
663
.
https://doi.org/10.1021/ed049p663
.
Dehghani
M. H.
Taher
M. M.
Bajpai
A. K.
Heibati
B.
Tyagi
I.
Asif
M.
Agarwal
S.
Gupta
V. K.
2015
Removal of noxious Cr(VI) ions using single-walled carbon nanotubes and multiwalled carbon nanotubes
.
Chemical Engineering Journal
279
,
344
352
.
https://doi.org/10.1016/j.cej.2015.04.151
.
Dehghani
M. H.
Faraji
M.
Mohammadi
A.
Kamani
H.
2016
Optimization of fluoride adsorption onto natural and modified pumice using response surface methodology: isotherm, kinetic and thermodynamic studies
.
Korean Journal of Chemical Engineering
34
(
2
),
454
462
.
doi:10.1007/s11814-016-0274-4
.
Dehghani
M. H.
Tajik
S.
Panahi
A.
Khezri
M.
Zarei
A.
Heidarinejad
Z.
Yousefi
M.
2018a
Adsorptive removal of noxious cadmium from aqueous solutions using poly urea-formaldehyde: a novel polymer adsorbent
.
MethodsX
5
,
1148
1155
.
doi:10.1016/j.mex.2018.09.010
.
Dehghani
M. H.
Farhang
M.
Alimohammadi
M.
Afsharnia
M.
McKay
G.
2018b
Adsorptive removal of fluoride from water by activated carbon derived from CaCl2-modified Crocus sativus leaves: equilibrium adsorption isotherms, optimization, and influence of anions
.
Chemical Engineering Communications
205
(
7
),
955
965
.
doi:10.1080/00986445.2018.1423969
.
Dorofeeva
G. A.
Streletskiib
A. N.
Povstugara
I. V.
Protasova
A. V.
Elsukova
E. P.
2012
Determination of nanoparticle sizes
.
74
(
6
),
710
720
.
https://doi.org/10.1134/S1061933X12060051
.
Dubinin
M. M.
1947
The equation of the characteristic curve of activated charcoal
.
Doklady Akademii Nauk SSSR
55
,
327
329
.
Fadiran
A. O.
Dlamini
S. C.
Mavuso
A.
2008
A comparative study of the phosphate levels in some surface and ground water bodies of Swaziland
.
Bulletin of the Chemical Society of Ethiopia
22
(
2
).
doi:10.4314/bcse.v22i2.61286
.
Fan
C.
Zhang
Y.
2018
Adsorption isotherms, kinetics and thermodynamics of nitrate and phosphate in binary systems on a novel adsorbent derived from corn stalks
.
Journal of Geochemical Exploration
188
,
95
100
.
https://doi.org/10.1016/j.gexplo.2018.01.020.
Freundlich
H. M. F.
1906
Over the adsorption in solution
.
Journal of Physical Chemistry
57
,
85
471
.
Ghafar A
H. H.
Ali
G. A. M.
Fouad
O. A.
Makhlouf
S. A.
2015
Enhancement of adsorption efficiency of methylene blue on Co3O4/SiO2 nanocomposite
.
Desalination and Water Treatment
53
(
11
),
2980
2989
.
https://doi.org/10.1080/19443994.2013.871343
.
Ho
Y. S.
McKay
G.
1999
Pseudo-second-order model for sorption processes
.
Process Biochemistry
34
(
5
),
451
465
.
https://doi.org/10.1016/S0032-9592(98)00112-5
.
Ho
Y. S.
Ng
J. C. Y.
McKay
G.
2000
Kinetics of pollutant sorption by biosorbents
.
Separation and Purification Methods
29
(
2
),
189
232
.
https://doi.org/10.1081/SPM-100100009
.
Jiang
D.
Amano
Y.
Machida
M.
2017
Removal and recovery of phosphate from water by a magnetic Fe3O4@ ASC adsorbent
.
Journal of Environmental Chemical Engineering
5
(
5
),
4229
4238
.
https://doi.org/10.1016/j.jece.2017.08.007
.
Jyothi
M. D.
Kiran
K. R.
Ravindhranath
K.
2012
New bio-sorbents in the control of phosphate pollution in wastewaters
.
International Journal of Applied Environmental Sciences
7
(
2
),
127
141
.
Karageorgiou
K.
Paschalis
M.
Anastassakis
G. N.
2007
Removal of phosphate species from solution by adsorption onto calcite used as natural adsorbent
.
Journal of Hazardous Materials
139
(
3
),
447
452
.
https://doi.org/10.1016/j.jhazmat.2006.02.038
.
Korngold
E.
1973
Removal of nitrates from potable water by ion exchange
.
Water, Air, and Soil Pollution
2
(
1
),
15
22
.
https://doi.org/10.1007/BF00572386
.
Krishna Mohan
G. V.
Naga Babu
A.
Kalpana
K.
Ravindhranath
K.
2017
Removal of chromium (VI) from water using adsorbent derived from spent coffee grounds
.
International Journal of Environmental Science and Technology
16
,
101
112
.
doi:10.1007/s13762-017-1593-7
.
Lacasa
E.
Cañizares
P.
Sáez
C.
Fernández
F. J.
Rodrigo
M. A.
2011
Electrochemical phosphates removal using iron and aluminium electrodes
.
Chemical Engineering Journal
172
(
1
),
137
143
.
https://doi.org/10.1016/j.cej.2011.05.080
.
Lagergren
S.
1898
The theory of so-called adsorption of soluble substances, adsorption gel
.
Sven, K., Vetensk. Akad. Handl
24
,
1
39
.
Langmuir
I.
1918
The adsorption of gases on plane surfaces of glass, mica and platinum
.
Journal of the American Chemical Society
40
(
9
),
1361
1403
.
https://doi.org/10.1021/ja02242a004
.
Liu
F.
Yang
J.
Zuo
J.
Ma
D.
Gan
L.
Xie
B.
Yang
B.
2014
Graphene-supported nanoscale zero-valent iron: removal of phosphorus from aqueous solution and mechanistic study
.
Journal of Environmental Sciences
26
(
8
),
1751
1762
.
https://doi.org/10.1007/s13762-017-1520-y
.
Lu
J.
Liu
D.
Hao
J.
Zhang
G.
Lu
B.
2015
Phosphate removal from aqueous solutions by a nano-structured Fe–Ti bimetal oxide sorbent
.
Chemical Engineering Research and Design
93
,
652
661
.
https://doi.org/10.1016/j.cherd.2014.05.001
.
Mahdavi
S.
Akhzari
D.
2016
The removal of phosphate from aqueous solutions using two nano-structures: copper oxide and carbon tubes
.
Clean Technologies and Environmental Policy
18
,
817
827
.
https://doi:org/10.1007/s10098-015-1058-y
.
Mansooreh
D.
Simin
N.
Mojtaba
K.
2014
Removal of 2,4-Dichlorophenolyxacetic acid (2,4-D) herbicide in the aqueous phase using modified granular activated carbon
.
Journal of Environmental Health Science and Engineering
12
,
28
.
https://doi.org/10.1186/2052-336X-12-28
.
Metcalf, Eddy
2003
Wastewater Engineering: Treatment of Reuse
, 4th edn.
McGraw Hill Co.
,
New York, NY, USA
.
Mohan
G. K.
Babu
A. N.
Kalpana
K.
Ravindhranath
K.
2017
Zirconium-treated fine Red Mud impregnated in Zn-alginate beads as adsorbent in removal of phosphate from water
.
Asian Journal of Chemistry
29
(
11
),
2549
2558
.
https://doi.org/10.1155/2017/4650594
.
Muhammad
A.
Xing
X.
Baoyu
G.
Qinyan
Y.
Shang
Y.
Rizwan
K.
Muhammad
A. I.
2020
Adsorptive removal of phosphate by the bimetallic hydroxide nanocomposites embedded in pomegranate peel
.
Journal of Environmental Science
91
,
189
198
.
https://doi.org/10.1016/j.jes.2020.02.005
.
Naga Babu
A.
Krishna Mohan
G. V.
Kalpana
K.
Ravindhranath
K.
2017
Removal of lead from water using calcium alginate beads doped with hydrazine sulphate-activated red mud as adsorbent
.
Journal of Analytical Methods in Chemistry
,
Article ID 4650594. doi:10.1155/2017/4650594
.
Ngah
W. W.
Hanafiah
M. A. K. M.
2008
Adsorption of copper on rubber (hevea brasiliensis) leaf powder: kinetic, equilibrium and thermodynamic studies
.
Biochemical Engineering Journal
39
(
3
),
521
530
.
https://doi.org/10.1016/j.bej.2007.11.006
.
Nguyen
T. M. P.
Van
H. T.
Nguyen
T. V.
Ha
L. T.
Vu
X. H.
Pham
T. T.
Nguyen
X. C.
2020
Phosphate adsorption by silver nanoparticles-loaded activated carbon derived from tea residue
.
Scientific Reports
10
(
1
),
1
13
.
https://doi.org/10.1038/s41598-020-60542-0
.
Ozmen
M.
Can
K.
Arslan
G.
Tor
A.
Cengeloglu
Y.
Ersoz
M.
2010
Adsorption of Cu (II) from aqueous solution by using modified Fe3O4 magnetic nanoparticles
.
Desalination
254
(
1–3
),
162
169
.
https://doi.org/10.1016/j.desal.2009.11.043
.
Pandian
C. J.
Palanivel
R.
Dhananasekaran
S.
2015
Green synthesis of nickel nanoparticles using Ocimum sanctum and their application in dye and pollutant adsorption
.
Chinese Journal of Chemical Engineering
23
(
8
),
1307
1315
.
https://doi.org/10.1016/j.cjche.2015.05.012
.
PCA
2006
Minnesota Pollution Control Agency
.
Phosphorus treatment and removal technologies
.
Rao
Y. H.
Ravindhranath
K.
2015
New methodologies in phosphate pollution control: using bio-sorbents derived from Terminalia arjuna and Madhuca indica plants
.
International Journal of ChemTech Research
8
(
12
),
784
796
.
Ravindhranath
K.
Mylavarapu
R.
2017a
Nano aluminum oxides as adsorbents in water remediation methods: a review
.
Rasayan Journal of Chemistry
10
(
3
),
716
722
.
http://dx.doi.org/10.7324/RJC.2017.1031762
.
Ravindhranath
K.
Mylavarapu
R.
2017b
Nickel based nano particles as adsorbents in water purification methods – a review
.
Oriental Journal of Chemistry
33
(
4
),
1603
16013
.
http://dx.doi.org/10.13005/ojc/330403
.
Ravulapalli
S.
Ravindhranath
K.
2019
Novel adsorbents possessing cumulative sorption nature evoked from Al2O3 nanoflakes, C. urens seeds active carbon and calcium alginate beads for defluoridation studies
.
Journal of the Taiwan Institute of Chemical Engineers
101
,
50
63
.
https://doi.org/10.1016/j.jtice.2019.04.034
.
Sowmya
A.
Meenakshi
S.
2014
Zr (IV) loaded cross-linked chitosan beads with enhanced surface area for the removal of nitrate and phosphate
.
International Journal of Biological Macromolecules
69
,
336
343
.
https://doi.org/10.1016/j.ijbiomac.2014.05.043
.
Sujitha
R.
Ravindhranath
K.
2017
Extraction of phosphate from polluted waters using calcium alginate beads doped with active carbon derived from Achyranthes aspera plant as adsorbent
.
Journal of Analytical Methods in Chemistry
,
Article ID 3610878. https://doi.org/10.1155/2017/3610878
.
Sun
Y. P.
Li
X. Q.
Zhang
W. X.
Wang
H. P.
2007
A method for the preparation of stable dispersion of zero-valent iron nanoparticles
.
Colloids and Surfaces A: Physicochemical and Engineering Aspects
308
(
1–3
),
60
66
.
https://doi.org/10.1016/j.colsurfa.2007.05.029
.
Sun
C.
Li
C.
Wang
C.
Qu
R.
Niu
Y.
Geng
H.
2012
Comparison studies of adsorption properties for Hg(II) and Au(III) on polystyrene-supported bis-8-oxyquinoline-terminated open-chain crown ether
.
Chemical Engineering Journal
200
(
202
),
291
299
.
https://doi.org/10.1016/j.cej.2012.06.007
.
Temkin
M. J.
Pyzhev
V.
1940
Recent Modifications to Langmuir Isotherms
.
Vogel
A. I.
1961
A Textbook of Quantitative Inorganic Analysis, Including Elementary Instrumental Analysis
, 3rd edn.
John Wiley and Sons, Inc.
,
New York, NY
,
USA
.
Xu
Q.
Li
W.
Ma
L.
Cao
D.
Owens
G.
Chen
Z.
2019
Simultaneous removal of ammonia and phosphate using green synthesized iron oxide nanoparticles dispersed onto zeolite
.
Science of The Total Environment
135002
.
https://doi.org/10.1016/j.scitotenv.2019.135002
.
Yamashita
T.
Yamamoto-Ikemoto
R.
2014
Nitrogen and phosphorus removal from wastewater treatment plant effluent via bacterial sulfate reduction in an anoxic bioreactor packed with wood and iron
.
International Journal of Environmental Research and Public Health
11
(
9
),
9835
99853
.
https://doi.org/10.3390/ijerph110909835
.
Yu
Y.
Paul Chen
J.
2015
Key factors for optimum performance in phosphate removal from contaminated water by a Fe–Mg–La trimetal composite sorbent
.
Journal of Colloid and Interface Science
445
,
303
311
.
https://doi.org/10.1016/j.jcis.2014.12.056
.
Zong
E.
Wei
D.
Wan
H.
2013
Adsorptive removal of phosphate ions from aqueous solution using zirconia-functionalized graphite oxide
.
Chemical Engineering Journal
221
,
193
203
.
https://doi.org/10.1016/j.cej.2013.01.088
.