10.8
CiteScore
 
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
10.8
CiteScore
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original article
10 (
1_suppl
); S461-S471
doi:
10.1016/j.arabjc.2012.10.006

Potential of nano crystalline calcium hydroxyapatite for Tin(II) removal from aqueous solutions: Equilibria and kinetic processes

Department of Engineering, Islamic Azad University, Maybod Branch, Maybod, P.O. Box 89516/175, Iran
Ceramics Department, Materials and Energy Research Center, Karaj, P.O. Box 31787-316, Iran
Department of Ceramics, Islamic Azad University, Shahrood Branch, Shahrood, Iran
Islamic Azad University, Science and Research Branch, Tehran, Iran

⁎Corresponding author. Tel.: +98 (261)6204131; fax: +98 2616201888. Iman.Mobasherpour@gmail.com (Iman Mobasherpour), I_Mobasherpour@merc.ac.ir (Iman Mobasherpour),

Disclaimer:
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.

Peer review under responsibility of King Saud University .

Abstract

The potential of the synthesized nano hydroxyapatite to remove Sn2+ from aqueous solutions was investigated in a batch reactor under different experimental conditions. The study also investigates the effects of process parameters such as initial concentration of Sn2+ ions, temperature, and adsorbent mass. Various thermodynamic parameters, such as ΔGo, ΔHo and ΔSo have been calculated. The thermodynamics of an Sn2+ ion onto nano HAp system indicates a spontaneous and endothermic nature of the process. Tin uptake was quantitatively evaluated using the Langmuir, Freundlich and Dubinin–Kaganer–Radushkevich models. The adsorption data follow the adsorption equilibrium described well by the Langmuir isotherm model with maximum adsorption capacity of 2500 mg/g of Sn2+ ions on nano HAp. Using the second-order kinetic constants, the activation energy of adsorption (Ea) was determined as 4.125 kJ mol−1 according to the Arrhenius equation.

Keywords

Adsorption
Nano crystalline hydroxyapatite
Sn2+ ion
Isotherms
Thermodynamics
Kinetic
1

1 Introduction

Over the past few decades, rapid industrialization has led to a tremendous increase in the use of heavy metals, which has inevitably resulted in an increased flux of metallic substances in the aquatic environment. Tin is one of the toxic metals found in most wastewaters. Many industries produce large quantities of waste streams containing low concentrations of tin, along with other metals from processes such as tin electroplating, aluminum anodizing, printed circuit board manufacturing, and metal pickling. Recycling metals from such solutions is attractive for environmental reasons and for the value of metals (Wasewar et al., 2009).

A level of tin (250 mg/kg) in canned food is generally accepted as a maximum tolerance level for humans. If a large amount of tin in canned food is taken daily over a long period, acute effects such as stomachache, anemia, and liver and kidney problems occur. The Occupational Safety and Health Administration (OSHA) has designated a workplace exposure limit of 0.1 mg/m3 for organic tin compounds, and 2 mg/m3 for inorganic tin compounds, except oxide. There are various treatment processes for the removal and/or recovery of heavy metal ions, including precipitation, oxidation, ultrafiltration, electrodialysis, solvent extraction, ion exchange, adsorption, etc. (Aksu and Tunc, 2005). For high strength and low volumes of wastewater, heavy metal removal by adsorption is a good proposition.

Hydroxyapatite is an ideal material for long-term containment of contaminants because of its high sorption capacity for actinides and heavy metals, low water solubility, high stability under reducing and oxidizing conditions, availability, and low cost (Krestou et al., 2004). It was used in stabilization of a wide variety of metals (e.g., Cr, Co, Cu, Cd, Zn, Ni, Pu, Pb, As, Sb, U, and V) by many investigators (Chen et al., 1997; Czerniczyniec et al., 2003; Vega et al., 1999; Reichert and Binner, 1996; Leyva et al., 2001; Fuller et al., 2002; McGrellis et al., 2001; Mobasherpour et al., 2011). They reported the sorption to take place through ionic exchange reaction, surface complexation with phosphate, calcium and hydroxyl groups and/or co-precipitation of new partially soluble phases.

Calcium hydroxyapatite (CaHAp), Ca10(PO4)6(OH)2, is used for the removal of heavy metals from contaminated soils, wastewater and fly ashes (Laperche et al., 1996; Ma et al., 1993; Ma et al., 1994; Mavropoulos, 2002; Nzihou and Sharrock, 2002; Takeuchi and Arai, 1990). Calcium hydroxyapatite (CaHAp) is a principal component of animal hard tissues and has been of interest in industrial and medical fields. Its synthetic particles find many applications in bioceramics, chromatographic adsorbents to separate protein and enzyme, catalysts for dehydration and dehydrogenation of alcohols, methane oxidation, and powders for artificial teeth and bones paste germicides (Elliott, 1994). These properties relate to various surface characteristics of HAp, e.g., surface functional groups, acidity and basicity, surface charge, hydrophilicity, and porosity. It has been found that HAp surface possesses 2.6 groups nm−2 of P–OH groups acting as sorption sites (Tanaka et al., 2005). The sorption properties of HAp are of great importance for both environmental processes and industrial purposes.

The objective of this study was to investigate the possible use of nano crystalline hydroxyapatite as an alternative adsorbent material for the removal of Sn2+ ions from aqueous solutions. The Langmuir, Freundlich and D–K–R models were used to fit the equilibrium isotherm. The dynamic behavior of the adsorption was investigated on the effect of initial metal ion concentration, temperature, and adsorbent mass of solution. The thermodynamic parameters were also evaluated from the adsorption measurements.

2

2 Material and methods

2.1

2.1 Preparation of nano crystalline hydroxyapatite sorbents

All chemicals used in this work were of analytical grade and the aqueous solutions were prepared using double distilled water. Nanocrystalline hydroxyapatite compounds were prepared by a solution-precipitation method (Mobasherpour et al., 2007) using (NH4)2HPO4 (Merck No. 1205) and Ca(NO3)2·4H2O (Analar No. 10305) as starting materials and ammonia solution as agents for pH adjustment. A suspension of Ca(NO3)2·4H2O was vigorously stirred and its temperature was maintained at 25 °C. A solution of (NH4)2HPO4 was slowly added dropwise to the Ca(NO3)2·4H2O solution. In all experiments the pH of Ca(NO3)2·4H2O solution by adding ammonia solution was 11. The precipitate HAP was removed from the solution by the centrifuge method at a rotation speed of 3000 rpm. The resulting powder was dried at 100 °C. The particles thus synthesized were characterized by the following methods. Transmission electron microscopy (TEM) was used to characterize the synthesized particles of HAp. For this purpose, particles were deposited onto Cu grids, which support a ‘‘holey'' carbon film. The particles were deposited onto the support grids by deposition from a dilute suspension in acetone or ethanol. The crystalline shapes and sizes were characterized by diffraction (amplitude) contrast and, for crystalline materials, by high resolution (phase contrast) imaging. The specific surface area was determined from N2 adsorption isotherm by the BET method using a Micromeritics surface area analyzer model ASAP 2010. The concentrations of Ca2+ and PO 4 3 - in the synthesized particles were determined by inductively coupled plasma atomic emission spectrometer (ICP-AES) by first dissolving the particles in HCl solution. The Fourier transform infrared spectra (FT-IR) of nano hydroxyapatite were recorded using a Perkin–Elmer 2000 FTIR spectrometer calibrated with a deuterated triglycine sulfate (DTGS) detector covering the frequency range of 500–4000 cm−1. The sample cell was purged with nitrogen gas throughout data collection to exclude carbon dioxide and water vapor. Ten milligrams of the dried samples was dispersed in 200 mg of spectroscopic grade KBr to record the spectra.

The crystal phase was identified by powder X-ray diffraction (XRD) using Siemens (30 kV and 25 mA) X-ray diffractometer with Cu Kα radiation ( λ = 1.5404 A ˚ ) and XPERT software.

2.2

2.2 Sorption study

All sorption experiments were carried out without any pre equilibrium processes that were imposed during the performance of any experiments. In order to determine the sorption capacity of nano HAp for Sn2+ cations, as well as the influence of the initial concentration of Sn2+ ion, adsorbent dosage and temperature, sorption experiments were performed by the batch equilibration technique.

Aqueous solutions containing Sn2+ ions of concentration (80, 100,120 and 140 mg/L) were prepared from Sn2+ chloride (SnCl2·2H2O, Merck No. 7815). 0.01 g of nano HAp was introduced in a stirred tank reactor containing 500 mL of the prepared solution. The stirring speed of the agitator was 300 rpm. The temperature of the suspension was maintained at 25 ± 1  C. The initial pH of the solution was adjusted to the value 6.5 by adding NH3 and HCl. Samples were taken after mixing the adsorbent and Sn2+ ion bearing solution at pre determined time intervals (5, 10, 20, 30, 60, and 120 min) for the measurement of residual metal ion concentration in the solution and to ensure equilibrium was reached. After any specified time the sorbents were separated from the solution by centrifuge and filtration through the filter paper (Whatman grade6). The exact concentration of metal ions was determined by AAS (GBC 932 Plus atomic absorption spectrophotometer). All experiments were carried out twice. The mass balance of tin is given by:

(1)
mq = V ( C 0 - C ) where m, q, V, C0, and C are the mass of nano-HAp (g), amount of tin removed by unit weight of HAp (Uptake capacity: mg Sn/g HAp), volume of tin solution (L), initial tin concentration in the solution (mg Sn/L), and the concentration of tin at the time t of adsorption (mg Sn/L). After 120 min, C and q will reach equilibrium values Ce and qe.

The percent removal (%) was calculated using the following equations:

(2)
% Removal = ( c 0 - c f ) c 0 × 100 where C0 and Cf are the concentrations of the metal ion in initial and final solutions (after 120 min).

3

3 Results and discussion

3.1

3.1 Characteristics of adsorbent

TEM micrograph of the HAp powders after drying is seen in Fig. 1a. The microstructure of the HAp crystalline after drying is observed to be almost like needle shape, with size in the range of 20–30 nm. The crystal structure analysis of HAp particles was performed, using X-ray diffraction, and the obtained diffractograms are represented in Fig. 1b. The produced reflection patterns match the ICDD standards (JCPDS) for HAp. The patterns show only the peaks characteristic of the synthesized HAp with no obvious evidence on the presence of other additional phases. The broad patterns around (2 1 1) and (0 0 2) indicate that the crystallites are very tiny in nature with much atomic oscillations. The sample was indexed in the hexagonal system with space group P63/m and Calcium Phosphate Hydroxide ICSD name (Ref. Code.01-074-0565 in XPERT software). The unit cell parameters of nano Hydroxyapatite synthesized were a = b = 0.9424 nm and c = 0.6879 nm. The analysis of the HAp sample has confirmed a low-crystalline product, with the specific surface area of 94.9 m2/g. The Ca/P molar ratio was found to be 1.65 ± 0.05 instead of 1.67 that represents the characteristic stoichiometric ratio. This cation deficiency is common for the sample obtained by the wet method.

TEM micrograph (a) and XRD pattern (b) of calcium nanocrystalline hydroxyapatite after drying at 100 °C.
Figure 1
TEM micrograph (a) and XRD pattern (b) of calcium nanocrystalline hydroxyapatite after drying at 100 °C.

The structure of the powder was analyzed using FT-IR spectroscopy after drying at 100 °C, as shown in Fig. 2. In the FT-IR analysis, mainly the peaks for PO 4 3 - and OH groups in the hydroxyapatite can be identified in Fig. 2. Peaks at 560–610 and 1000–1100 cm−1 must be due to PO 4 3 - . For the OH group the peak positions are 636 and 3572 cm−1.

FTIR Spectroscopy analysis of the nano hydroxyapatite powder after drying at 100 °C.
Figure 2
FTIR Spectroscopy analysis of the nano hydroxyapatite powder after drying at 100 °C.

3.2

3.2 Effect of initial Sn2+ concentration and adsorbent dosage

The sorption of Sn2+ ions was carried out at different initial Sn2+ ion concentrations ranging from 80 to 140 mg/L, at pH 6.5, at 300 rpm with 120 min of contact time using nano-HAp. A rapid kinetic reaction of Sn removal by sorbent occurred within the first 10 min (Fig. 3a). The aqueous Sn concentration at 10 min decreased to 42, 62, 80 and 98 mg/L by nano-HAp for 80, 100, 120 and 140 mg/L initial concentration respectively. Uptake of the Sn2+ also increased with increasing the initial metal concentration tending to saturation at higher metal concentrations. As shown in Fig. 3b, when the initial Sn2+ concentration increased from 80 to 140 mg/L, the uptake capacity of nano HAp increased from 1985 to 2400 mg/g. A higher initial concentration provides an important driving force to overcome all mass transfer resistances of the pollutant between the aqueous and solid phases, thus increasing the uptake (Aksu and Tezer, 2005).

Time dependent concentration (a) and effect of initial concentration on removal of aqueous Sn2+ by nano hydroxyapatite sorbents (pH 6.5, adsorbent dosage = 0.01 g/L, 300 rpm agitating rate).
Figure 3
Time dependent concentration (a) and effect of initial concentration on removal of aqueous Sn2+ by nano hydroxyapatite sorbents (pH 6.5, adsorbent dosage = 0.01 g/L, 300 rpm agitating rate).

The effect of nano-HAp dosage is presented in Fig. 4. It is evident that adsorption increases with the increase in the mass of sorbent and the uptake capacity of Sn2+ decreased from 2400 mg/g (34.28% removal) to 1233.33 mg/g (52.86% removal) with the increasing nano-HAp concentration from 0.01 to 0.03 g/L. This is because at the higher dosage of sorbent due to increased surface area, more adsorption sites are available causing higher removal of Sn2+.

Effect of adsorbent dosage on removal of Sn2+ by nano hydroxyapatite (pH 6.5, initial metal concentration = 140 mg/L, 300 rpm agitating rate).
Figure 4
Effect of adsorbent dosage on removal of Sn2+ by nano hydroxyapatite (pH 6.5, initial metal concentration = 140 mg/L, 300 rpm agitating rate).

3.3

3.3 Effect of temperature

To study the effect of this parameter on the uptake of Sn2+ ions by nano-HAp, we selected the following temperatures: 25, 45 and 65  C. Fig. 5, illustrates the relationship between temperature and the amount of Sn2+ ions adsorbed onto nano-HAp at equilibrium time (120 min). As seen in Fig. 5, the adsorption of Sn2+ on nano-HAp increased from 2400 mg/g (34.28% removal) to 2750 mg/g (39.28% removal) when temperature was increased from 25 to 65 °C at an initial concentration of 140 mg/L. The increase in the equilibrium sorption of Sn2+ with temperature indicates that Sn2+ ions removal by adsorption on nano-HAp favors a high temperature. This may be a result of increase in the mobility of the Sn2+ ion with temperature. An increasing number of molecules may also acquire sufficient energy to undergo an interaction with active sites at the surface. Furthermore, increasing temperature may produce a swelling effect within the internal structure of nano-HAp enabling metal ions to penetrate further (Dogan and Alkan, 2003).

Uptake capacity of Sn2+ ions at different temperatures (pH 6.5, initial metal concentration = 140 mg/L, adsorbent dosage = 0.01 g/L, 300 rpm agitating rate).
Figure 5
Uptake capacity of Sn2+ ions at different temperatures (pH 6.5, initial metal concentration = 140 mg/L, adsorbent dosage = 0.01 g/L, 300 rpm agitating rate).

3.4

3.4 Determination of thermodynamic parameters

Thermodynamic parameters such as free energy change (ΔGoHo), and entropy change (ΔSo) can be estimated using equilibrium constants changing with temperature. The free energy change of the sorption reaction is given by the following equation:

(3)
Δ G o = - RT Ln K d where ΔGo is the standard free energy change (J); R is the universal gas constant, 8.314 J/mol K and T is the absolute temperature (K).The distribution ratio (Kd) was calculated using Eq. (2):
(4)
K d = amount of metal in adsorbent amount of metal in solution = q e C e
The distribution ratio (Kd) values increased with temperature, indicating the endothermic nature of adsorption. A plot of Gibbs free energy changes, ΔGo, versus temperature, T (K); was found to be linear. The values of ΔHo and ΔSo) were determined from the slope and intercept of the plots according Eq. (5):
(5)
Δ G o = Δ H o - T Δ S o
The thermodynamic parameters Gibbs free energy change, ΔGo, are shown in Table 1. The enthalpy change, ΔHo, and the entropy change, ΔSo, for the sorption processes are calculated to be 4524 J/mol and 99.67 J/mol K, respectively. The negative values of ΔGo at various temperatures indicate the spontaneous nature of the adsorption process. The positive value of ΔSo indicates that there is an increase in the randomness in the system solid/solution interface during the adsorption process. In addition, the positive value of ΔHo indicates that the adsorption is endothermic. The positive value of ΔSo reflects the affinity of nano-HAp for Sn2+ ions and suggests some structural changes in tin and nano-HAp (Ho, 2003).
Table 1 Thermodynamic parameters for the adsorption of Sn2+ onto nano hydroxyapatite.
T (K) Kd ΔGo (J/mol) ΔHo (J/mol) ΔSo (J/mol k)
298 26086.96 −25194.90
318 28651.69 −27133.30 4524 99.67
338 32352.94 −2931.70

3.5

3.5 Adsorption isotherms

Analysis of the equilibrium data is important to develop an equation which accurately represents the results and which could be used for design purposes (Aksu, 2002). Several isotherm equations have been used for the equilibrium modeling of adsorption systems.

The sorption data have been subjected to different sorption isotherms, namely, Langmuir, Freundlich, and Dubinin–Kaganer–Radushkevich.

The equilibrium data for metal cations over the concentration range from 80 to 140 mg/L at 25  C have been correlated with the Langmuir isotherm (Langmuir, 1918):

(6)
C e q e = 1 Q 0 K + C e Q 0 where Ce is the equilibrium concentration of metal in the solution (mg/L), qe is the amount absorbed at equilibrium onto nano-HAp (mg/g), Q0 and K are Langmuir constants related to sorption capacity and sorption energy, respectively. Maximum sorption capacity (Q0) represents the monolayer coverage of the sorbent with the sorbate and K represents enthalpy of sorption and should vary with temperature. A linear plot is obtained when Ce/qe is plotted against Ce over the entire concentration range of metal ions investigated.

The linearized Langmuir adsorption isotherms of Sn2+ ions are given in Fig. 6a. An adsorption isotherm is characterized by certain constants whose values express the surface properties and affinity of the sorbent and can also be used to find the sorption capacity of the sorbent.

Linear fits of experimental data obtained using Langmuir (a), Freundlich (b) and D–K–R (c) sorption isotherms for the adsorption of Sn2+ onto nano Hydroxyapatite (pH 6.5, initial metal concentration = 80, 100, 120 and 140 mg/L, adsorbent dosage = 0.01 gr/L, 300 rpm agitating rate).
Figure 6
Linear fits of experimental data obtained using Langmuir (a), Freundlich (b) and D–K–R (c) sorption isotherms for the adsorption of Sn2+ onto nano Hydroxyapatite (pH 6.5, initial metal concentration = 80, 100, 120 and 140 mg/L, adsorbent dosage = 0.01 gr/L, 300 rpm agitating rate).

The Freundlich sorption isotherm, one of the most widely used mathematical descriptions, usually fits the experimental data over a wide range of concentrations. This isotherm gives an expression encompassing the surface heterogeneity and the exponential distribution of active sites and their energies. The Freundlich adsorption isotherms were also applied to the removal of Sn2+ on nano-HAp (Fig. 6b).

(7)
Ln q e = Ln k f + 1 n Ln C e where qe is the amount of metal ion sorbed at equilibrium per gram of adsorbent (mg/g), Ce the equilibrium concentration of metal ion in the solution (mg/L), kf, and n the Freundlich model constants (Malkoc and Nuhoglu, 2003; Kadirvelu et al., 2001). Freundlich parameters, kf and n, were determined by plotting Ln qe versus Ln Ce. The numerical value of 1/n < 1 indicates that adsorption capacity is only slightly suppressed at lower equilibrium concentrations. This isotherm does not predict any saturation of the sorbent by the sorbate; thus infinite surface coverage is predicted mathematically, indicating multilayer adsorption on the surface (Hasany et al., 2002).

The Dubinin–Kaganer–Radushkevich (D–K–R) equation has been used to describe the sorption of metal ions on clays. The D–K–R equation has the form:

(8)
Ln C ads = Ln X m - β ε 2 where Cads is the number of metal ions adsorbed per unit weight of adsorbent (mol/g), Xm (mol/g) is the maximum sorption capacity, β (mol2/J2) is the activity coefficient related to mean sorption energy, and ε is the Polanyi potential, which is equal to:
(9)
= RT ln ( 1 + 1 / Ce )
where R is the gas constant (kJ/mol K) and T is the temperature (K). The saturation limit Xm represents the total specific micropore volume of the sorbent. The sorption potential is independent of the temperature but varies according to the nature of the sorbent and the sorbate (Khan et al., 1995). The slope of the plot of ln Cads versus ε gives β (mol2/J2) and the intercept yields the sorption capacity, Xm (mol/g). The sorption space in the vicinity of a solid surface is characterized by a series of equipotential surfaces having the same sorption potential. This sorption potential is independent of the temperature but varies according to the nature of the sorbent and the sorbate. The sorption energy can also be worked out using the following relationship:
(10)
E = 1 / - 2 β .
It is known that the magnitude of apparent adsorption energy E is useful for estimating the type of adsorption and if this value is below 8 kJ/mol the adsorption type can be explained by physical adsorption, between 8 and 16 kJ/mol the adsorption type can be explained by ion exchange, and over 16 kJ/mol the adsorption type can be explained by a stronger chemical adsorption than ion exchange; (Lin and Juang, 2002; Wang, 2004; Krishna et al., 2000). The plot of Ln Cads against ε2 for metal ion sorption on nano-HAp is shown in Fig. 6c. The Langmuir, Freundlich and D–K–R adsorption constants from the isotherms and their correlation coefficients are also presented in Table 2.
Table 2 Langmuir, Freundlish and D–K–R constants for adsorption of Sn2+ onto nano hydroxyapatite.
Langmuir adsorption isotherm constants
Qo (mg/g) K (L/mg) R2
2500 0.07 0.997
Freundlich adsorption isotherm constants
kf (mg/g) n R2
899.64 4.58 0.929
DKR adsorption isotherm constants
Xm (mg/g) β (mol2/J2) R2
5092.06 −2 × 10−9 0.935

The correlation factors R (0.997, 0.929 and 0.935 for Langmuir, Freundlich and D–K–R models, respectively) confirm good agreement between the two theoretical models and our experimental results. The maximum sorption capacity, Q0, calculated from the Langmuir equation is 2500 mg/g, while Langmuir constant K is 0.07 L/mg. The values obtained for Sn2+ from the Freundlich model showed a maximum adsorption capacity (Kf) of 899.64 mg/g with an affinity value (n) equal to 4.58. The values of sorption constants, derived from the D–K–R model are: 5092.06 mg/g (42.89 mmol/g) for Xm, −2 × 10−9 mol2/J2 for β and 15.81 kJ/mol for E.

The values indicate that the adsorption pattern for Sn2+ on nano-HAp followed thrid the Langmuir isotherm (R2 > 0.997), the D–K–R isotherm (R2 > 0.935), and the Freundlich isotherm (R2 > 0.929) at all experiments. It is clear that the Langmuir isotherm is the best fit for the sorption of Sn2+ on nano-HAp. When the system is in a state of equilibrium, the distribution of Sn2+ between nano HAp and the Sn2+ solution is of fundamental importance in determining the maximum sorption capacity of nano HAp for the Sn2+ ion from the isotherm. The E values are 15.81 kJ/mol for Sn2+, on the nano HAp. It is the orders of an ion-exchange mechanism, in which the sorption energy lies within 8–16 kJ/mol.

Generally, HAp selectivity toward divalent metal cations is a result of the ion-exchange process with Ca2+ ions (Monteil Rivera and Fedoroff, 2002). Ionic radius of Sn2+ (0.71 Å) slightly differ from that of Ca2+ (0.99 Å), and it can substitute Ca2+ in the HAp crystal lattice. Fig. 7 presents the XRD patterns of the Sn2+-loaded sample. No structural changes of nano HAp were detected by the powder X-ray diffraction analysis of the solid residue with the maximum amount of uptake capacity of Sn2+, obtained after interaction of 0.01 g/L of nano HAp with 140 mg/L Sn2+ solution, at pH 6.5, at 300 rpm with 120 min of contact time. The sample was indexed in the hexagonal system with space group P63/m. The diffractogram’s evidence clarifies that all XRD peaks were shifted toward upper diffraction angles for Sn-HAp particles (maximum peak: from 2 θ = 31.94 ° to 2 θ = 32.11 ° ). These shifts are indicative of the decrease in unit cell dimensions which is due to the replacement of Sn2+ (ionic radius 0.71 Å), which is smaller than Ca2+ (ionic radius 0.99 Å), into the crystal lattice of apatite molecules.

XRD pattern of the solid residue with maximum amount of uptake capacity of Sn2+ (pH 6.5, initial metal concentration = 140 mg/L, adsorbent dosage = 0.01 g/L, 300 rpm agitating rate).
Figure 7
XRD pattern of the solid residue with maximum amount of uptake capacity of Sn2+ (pH 6.5, initial metal concentration = 140 mg/L, adsorbent dosage = 0.01 g/L, 300 rpm agitating rate).

The reaction mechanism corresponds to the equimolar exchange of tin and calcium yielding Ca10−xSnx(PO4)6(OH)2, where x can vary from 0 to 10 depending on the reaction time and experimental conditions. Our results of synthesized nano-HAp agreed with those described elsewhere that the proposed mechanism for Sn2+ removal by HAp comprises two steps: firstly, rapid surface complexation of Sn2+ on the ≡POH sites of HAp which causes the decrease of the pH (from pH = 6.5 to pH = 6.0 for initial metal concentration = 140 mg/L, dosage = 0.01 g/L, 300 rpm agitating rate) and secondly, partial dissolution of calcium followed by the precipitation of an apatite with formula: SnxCa10−x(PO4)6(OH)2.

In which Sn2+ ions are first adsorbed on the nano HAp surface and substitution with Ca2+ ion occurs as described by the following equation:

(11)
Ca 10 ( PO 4 ) 6 ( OH ) 2 + x Sn 2 + Ca 10 - x Sn x ( PO 4 ) 6 ( OH ) 2 + x Ca 2 +

3.6

3.6 Sorption kinetics

Sorption kinetic studies were carried out in order to understand the behavior of nano-HAp toward Sn2+ metal. The sorption kinetics includes two phases: a rapid metal sorption stage followed by a much slower stage before the equilibrium was established. It was found that the mass transfer was the key factor in the metal sorption (Chen and Wang, 2004). The sorption kinetics describes the metal sorption rate, which in turn governs the residence time of the sorption reaction and also the efficiency of the sorption process. Out of the several kinetic models available to examine the controlling mechanism of the sorption kinetic process and to test the experimental data, the Lagrangian equation or pseudo-first-order equation and pseudo-second-order equation have been used for the metal sorption kinetics of nano-HAp.

The linearized form of the pseudo-first-order equation:

(12)
log ( q e - q ) = log q e cal - k 1 2.303 t where qe is the metal sorbed at equilibrium (mg g−1), q is the amount of the metal adsorbed (mg g−1) at any time t, k1 is the first-order rate constant. The first-order rate constant k1 and q were determined from the slopes and intercept of plots of log (qe − q) vs t at different metal concentration.

The linearized form of the pseudo-second-order equation for the kinetics of absorption described by Ho and Chiang (2001) is as follows:

(13)
t qt = 1 k 2 q e cal 2 + 1 q e cal

The second-order rate constant (k2) and qe cal were determined from the slope and intercept of the plot obtained by plotting t/qt vs t.

Linear plots of log (qe − qt) versus t and t/qt versus t are depicted in Fig. 8a and Fig. 8b at 25, 45 and 65  C, respectively. A comparison of the results with the correlation coefficients is shown in Table 3. The pseudo-second-order kinetic model obtained for Sn2+ sorption at various temperatures showed a better correlation of the result than the pseudo-first-order equation model. The correlation coefficients for the second order kinetic model obtained at 140 ppm concentrations at different temperatures were high. The values of k2 at 25, 45 and 65  C were varied from 0.00023 to 0.000191 min−1. The higher rate of metal sorption in the beginning (Fig. 3b) could be due to the presence of the active site in the nano-HAp surface, available for the sorption of metals. Once the sorptive sites are exhausted, the uptake rate may be controlled by the rate of intra particle diffusion. The activation energy Ea was determined using the Arrhenius equation (Aksu, 2002):

(14)
Ln K ad = Ln A - E a RT where kad(k2) is the rate constant value for the metal adsorption, Ea the activation energy in kJ mol−1, T the temperature in Kelvin, and R is the gas constant (=8.314 kJ mol−1 K−1). When Ln kad is plotted versus 1/T, a straight line with slope −Ea/R is obtained. The activation energy for the adsorption system of Sn2+ onto nano-HAp was found as 4.125 kJ mol−1 from the slope of this plot, indicating the physical adsorption. As known when the rate is controlled by intra-particle diffusion mechanism, the activation energy is low and hence it can be concluded that the process is controlled by intra-particle diffusion, which is a physical step in the adsorption process (Dogan and Alkan, 2003).
Linear fit of experimental data obtained using pseudo-first order kinetic model (a) and pseudo-second order kinetic model (b) (pH 6.5, initial metal concentration = 140 mg/L, adsorbent dosage = 0.01gr/L, 300 rpm agitating rate).
Figure 8
Linear fit of experimental data obtained using pseudo-first order kinetic model (a) and pseudo-second order kinetic model (b) (pH 6.5, initial metal concentration = 140 mg/L, adsorbent dosage = 0.01gr/L, 300 rpm agitating rate).
Table 3 A comparison of the first and second order kinetic rate constants and calculated qe cal values obtained at different temperatures (pH 6.5, initial metal concentration = 140 mg/L, adsorbent dosage = 0.01gr/L, 300 rpm agitating rate).
T [°C] K1 [min−1] qe cal [mg/g] R2
Pseudo-first-order kinetic model
25 0.09672 1312.20 0.67761
45 0.12436 1377.21 0.73865
65 0.17042 2157.74 0.80748
Pseudo-second-order kinetic model
25 0.00023 2440.45 0.99975
45 0.00024 2577.78 0.99985
65 0.00019 2790.49 0.99993

4

4 Conclusions

The present investigation shows that nano HAp is an effective adsorbent for the removal of Sn2+ from aqueous Sn2+ solutions. The adsorption process is a function of the adsorbent dosage, the initial Sn2+ concentration and the temperature. The efficiency of Sn2+ adsorption increased with an increase in the adsorbent dosage. Isotherm studies indicate that the Langmuir model fits the experimental data better than Freundlich and D–K–R models. The adsorption equilibrium was described well by the Langmuir isotherm model with a maximum adsorption capacity of 2500 mg/g of Sn2+ ions on nano HAp. The results of XRD analysis strongly support the ion-exchange as a main mechanism for Sn2+ removal by nano HAp. The results show that the Sn2+ uptake by nano hydroxyapatite proceeds with a rapid surface complexation of the Sn2+ on the ≡POH site before the formation of a compound of formula Ca10−xSnx(PO4)6(OH)2. Thermodynamic calculations showed that the Sn2+ sorption process of nano HAp has an endothermic and spontaneous nature. The kinetic behavior of the metal toward nano-HAp demonstrated pseudo-second-order kinetics rather than pseudo-first-order kinetics. The second order kinetic model was successfully applied to the experimental data, confirming that adsorption was controlled by intra-particle diffusion.

Acknowledgment

This research was supported by the Research Project Unit at the Maybod Branch, Islamic Azad University, Maybod, Iran under the Project No. P/870358.

References

  1. , . Determination of the equilibrium, kinetic and thermodynamic parameters of the batch biosorption of nickel(II) ions onto Chlorella vulgaris. Process Biochem.. 2002;38:89-99.
    [Google Scholar]
  2. , , . Biosorption of reactive dyes on the green alga Chlorella vulgaris. Process Biochem.. 2005;40:1347-1361.
    [Google Scholar]
  3. , , . Application of biosorption for penicilin G removal: comparison with activated carbon. Process Biochem.. 2005;40:831-847.
    [Google Scholar]
  4. , , . Characterization of metal adsorption kinetic properties in batch and fixed bed reactors. Chemosphere. 2004;54:397-404.
    [Google Scholar]
  5. , , , , . Effects of pH on heavy metal sorption on mineral apatite. Environ. Sci. Technol.. 1997;31(3):624-631.
    [Google Scholar]
  6. Czerniczyniec, M., Farias, S., Magallanes, J., Cicerone, D. 2003. Arsenic adsorption on biogenic HAP: Solution composition effects. In: 11th International Conference on Surface and Colloid Science, Fozdo Iguazu, Brazil, 269.
  7. , , . Adsorption kinetics of methyl violet onto perlite. Chemosphere. 2003;50:517-528.
    [Google Scholar]
  8. , . Structure and Chemistry of the Apatites and Other Calcium Orthophosphates. Amsterdam: Elsevier; .
  9. , , , , . Mechanisms of uranium interactions with hydroxyapatite: implications for groundwater remediation. Environ. Sci. Technol.. 2002;36(2):158-165.
    [Google Scholar]
  10. , , , , . Sorption and thermodynamic behavior of zinc(II)-thiocyanate complexes onto polyurethane foam from acidic solutions. Nucl. Chem.. 2002;252(3):477-484.
    [Google Scholar]
  11. , . Removal of copper ions from aqueous solution by tree fern. Water Res.. 2003;37:2323-2330.
    [Google Scholar]
  12. , , . Sorption studies of acid dye by mixed sorbents. Adsorption. 2001;7:139-147.
    [Google Scholar]
  13. , , , . Adsorption of nickel(II) from aqueous solution onto activated carbon prepared from coirpith. Sep. Purif. Technol.. 2001;24:497-505.
    [Google Scholar]
  14. , , , . Adsorption of chromium(III), chromium(VI) and silver(I) on bentonite. Waste Manage.. 1995;15(4):271-282.
    [Google Scholar]
  15. , , , . Mechanism of aqueous uranium(VI) uptake by a natural zeolitic tuff. Miner. Eng.. 2004;16(12):1363-1370.
    [Google Scholar]
  16. , , , . Thermodynamics of chromium(VI) anionic species sorption onto surfactant-modified montmorillonite clay. J. Colloid Interface Sci.. 2000;229(1):230-236.
    [Google Scholar]
  17. , . The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc.. 1918;40(9):1361-1403.
    [Google Scholar]
  18. , , , , , . Chemical and mineralogical characterizations of Pb in a contaminated soil: reactions with synthetic apatite. Environ. Sci. Technol.. 1996;30:3321-3326.
    [Google Scholar]
  19. , , , , . Sorption of antimony onto hydroxyapatite. Environ. Sci. Technol.. 2001;35(18):3669-3675.
    [Google Scholar]
  20. , , . Heavy metal removal from water by sorption using surfactant-modified montmorillonite. J. Hazard. Mater.. 2002;B 92(3):315-326.
    [Google Scholar]
  21. , , , , . In situ lead immobilization by apatite. Environ. Sci. Technol.. 1993;27:1803-1810.
    [Google Scholar]
  22. , , , , . Effects of Aqueous Al, Cd, Cu, Fe(II), Ni, and Zn on Pb immobilization by hydroxyapatite. Environ. Sci. Technol.. 1994;28:1219-1228.
    [Google Scholar]
  23. , , . The removal of chromium(VI) from synthetic wastewater by Ulothrix zonata. Fres. Environ. Bull.. 2003;12(4):376-381.
    [Google Scholar]
  24. , . Studies on the mechanisms of lead immobilization by hydroxyapatite. Environ. Sci. Technol.. 2002;36:1625-1629.
    [Google Scholar]
  25. , , , , , . Influence of the sorption protocol on the uptake of Cd ions in calcium hydroxyapatite. Sep. Purif. Technol.. 2001;24:129-138.
    [Google Scholar]
  26. , , , , . Synthesis of nanocrystalline hydroxyapatite by using precipitation method. Jalcom. 2007;430:330-333.
    [Google Scholar]
  27. , , , . Removal of divalent cadmium cations by means of synthetic nano crystallite hydroxyapatite. Desalination. 2011;266:142-148.
    [Google Scholar]
  28. , , . Sorption of inorganic species on apatites from aqueous solutions. In: Encyclopedia of Surface and Colloid Science. New York: Marcel Dekker Inc.; .
    [Google Scholar]
  29. , , . Calcium phosphate stabilization of fly ash with chloride extraction. Waste Manag.. 2002;2002:235-239.
    [Google Scholar]
  30. , , . An evaluation of hydroxyapatite-based filters for removal of heavy metal ions from aqueous solutions. J. Mater. Sci.. 1996;31(5):1231-1241.
    [Google Scholar]
  31. , , . Removal of coexisting Pb2+, Cu2+, Cd2+ ions from water by addition of hydroxyapatite powder. J. Chem. Eng. Jpn.. 1990;23:75-80.
    [Google Scholar]
  32. , , , , , . Structure of synthetic calcium hydroxyapatite particles modified with pyrophosphoric acid. J. Colloid Interface Sci.. 2005;283(2):609-612.
    [Google Scholar]
  33. , , , . Interaction of oxovanadium(IV) with crystalline calcium hydroxyapatite: surface mechanism with no structural modification. J. Phys. Chem. Solids. 1999;60(6):759-766.
    [Google Scholar]
  34. , . Effects of exchanged surfactant cations on the pore structure and adsorption characteristics of montmorillonite. J. Colloid Interface Sci.. 2004;280(1):27-35.
    [Google Scholar]
  35. , , , . Adsorption of tin using granular activated carbon. J. Environ. Pro. Sci.. 2009;3:41-52.
    [Google Scholar]
Show Sections