Вы находитесь на странице: 1из 7

Journal of Colloid and Interface Science 317 (2008) 402408

www.elsevier.com/locate/jcis

Cadmium removal from single- and multi-metal (Cd + Pb + Zn + Cu)


solutions by sorption on hydroxyapatite
Alessia Corami, Silvano Mignardi , Vincenzo Ferrini
Department of Earth Sciences, University of Rome La Sapienza, Piazzale Aldo Moro 5, 00185 Rome, Italy
Received 10 July 2007; accepted 28 September 2007
Available online 2 October 2007

Abstract
Heavy metal contamination of waters and soils is particularly dangerous to the living organisms. Different studies have demonstrated that
hydroxyapatite has a high removal capacity for divalent heavy metal ions in contaminated waters and soils. The removal of Cd from aqueous
solutions by hydroxyapatite was investigated in batch conditions at 25 2 C. Cadmium was applied both as single- or multi-metal (Cd + Pb +
Zn + Cu) systems with initial concentrations from 0 to 8 mmol L1 . The adsorption capacity of hydroxyapatite in single-metal system ranged
from 0.058 to 1.681 mmol of Cd/g of hydroxyapatite. In the multi-metal system competitive metal sorption reduced the removal capacity by 63
83% compared to the single-metal system. The sorption of Cd by hydroxyapatite follows the Langmuir model. Cadmium immobilization occurs
through a two-step mechanism: rapid surface complexation followed by partial dissolution of hydroxyapatite and ion exchange with Ca resulting
in the formation of a cadmium-containing hydroxyapatite.
2007 Elsevier Inc. All rights reserved.
Keywords: Cadmium; Hydroxyapatite; Competitive sorption; Surface complexation; Ion exchange

1. Introduction
The removal of toxic heavy metals from industrial wastewaters is one of the most important issues of environmental
remediation. Heavy metals such as Pb, Cd, Cu, Zn, Hg, Cr,
and Ni are the main contaminants of surface water, groundwater, and soils. The main sources of these elements are metal
plating industries, abandoned disposal sites, and mining industries [1]. The presence of toxic heavy metals in water has caused
several health problems with animals, plants, and human being [2]. Among toxic heavy metals, cadmium is one of the
most dangerous for human health [3]. Cadmium is an irritant
to the respiratory tract and exposure to this pollutant can lead
to anaemia, renal damage, osseous disease with effects similar
to osteoporosis, and Itai-Itai disease [4,5]. However, cadmium
has also practical applications: e.g., it is highly corrosion resistant and is used as a protective coating for iron, steel, and
copper. The industrial uses of cadmium are increasing in plas* Corresponding author. Fax: +39 064454729.

E-mail address: silvano.mignardi@uniroma1.it (S. Mignardi).


0021-9797/$ see front matter 2007 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2007.09.075

tics, paint pigments, electroplating, batteries, mining, and alloy


industries [6].
Many methods have been proposed for the removal of toxic
heavy metals from wastewaters. Chemical precipitation, filtration, complexing, solvent extraction, electrochemical technique, ion exchange, and adsorption are some of the commonly
used processes [79]. However, all these procedures vary in effectiveness and cost. In particular, chemical immobilization is
one of the most used for reducing the bioavailability of heavy
metals. The aim is to remove toxic metals through the formation
of new minerals with a lower solubility and which are stable in
the environment in a wide range of conditions [10].
Many studies have recognized the ability of hydroxyapatite
(HA) [Ca10 (PO4 )6 (OH)2 ] to bind divalent heavy metal ions and
previous studies have shown that synthetic HA has a high removal capacity for Pb, Zn, Cu, Cd, Co, and Sb in aqueous
solutions [5,1120].
Mechanisms such as ion exchange [13,15], surface complexation [13,14], dissolution of HA and precipitation of metal
phosphates [12,13], substitution of Ca in HA by metals during
recrystallization (coprecipitation) [13,21] have been proposed
in order to describe the uptake of heavy metals from aque-

A. Corami et al. / Journal of Colloid and Interface Science 317 (2008) 402408

403

Table 1
Chemical composition, specific surface area (SSA) and point of zero charge (pHPZC ) of HA
Ca
(wt%)

P
(wt%)

Mg
(wt%)

Si
(wt%)

Cd
(ppm)

Cu
(ppm)

Zn
(ppm)

Pb
(ppm)

Ca/P

SSA
(m2 g1 )

pHPZC

38.3

16.8

0.3

0.3

1.76

50

7.4

ous solutions by synthetic HA. However, controversy connected


with the relative contribution of each process in removing heavy
metals still exists in the literature [13,22]. Moreover, it is difficult to detect the formation of Cd complexes or modifications
in the HA structure by X-ray diffraction due to the similar ionic
radii of Ca (0.99 ) and Cd (0.97 ). Therefore, in the present
work the general term sorption defines the attachment of heavy
metals from a solution to its coexisting solid surfaces.
Contaminated waters commonly contain many metals including, besides Cd, Pb, Cu, Zn, Cr, and Ni [23]. Competition
among these metals affecting synthetic HA overall removal capacity has not been extensively evaluated [12,13].
The purposes of this study were (a) to evaluate the uptake
of Cd2+ ions by synthetic HA both in single- and multi-metal
(Cd + Pb + Zn + Cu) systems; (b) to examine the effect of the
other three metals on the sorption of Cd2+ ions by HA.
2. Materials and methods
2.1. Materials
Table 1 shows the chemical composition, the specific surface
area and the point of zero charge (pHPZC ) of the commercial
synthetic HA (Alfa Aesar) used in this study. The material
consists only of crystalline HA as determined by X-ray diffraction (XRD) analyses (ICDD 9-432) [24]. Scanning electron
microscopy (SEM) analyses showed that HA was formed of
spherical-like particles with average diameter 10 m and energy dispersive spectroscopy (EDS) analyses determined only
Ca and P.
The specific surface area of the HA was measured by the
BET-method (adsorption gas N2 , carrier gas He, heating temperature 150 C).
Analytical reagent grade compounds and deionized water
were used to prepare all solutions and suspensions. Single
(Cd(II)) and multi-metal (Cd(II) + Pb(II) + Zn(II) + Cu(II))
aqueous solutions were prepared from their nitrate salts. All the
experiments were carried out in duplicate.
To measure pHPZC the batch equilibration technique [25]
was used.
Solution pH adjustment was not used during the sorption
experiments to simulate the conditions existing during the remediation of contaminated water, where the pH control is either
not necessary or difficult to be realized. As well as, a background electrolyte was not applied.
2.2. Sorption experiments
Sorption experiments were performed in batch at 25 2 C.
In high-density polyethylene beakers (Nalgene) 200 mL of the

Fig. 1. Time variation plot for sorption of Cd on HA. (a) ([Cd2+ ] =


0.89 mmol L1 , [Pb2+ ] = 0.48 mmol L1 , [Zn2+ ] = 1.53 mmol L1 ,
[Cu2+ ] = 1.57 mmol L1 , HA = 0.2 g, T = 25 2 C) and intraparticle
diffusion plot (b). Closed symbols = single-metal system; open symbols =
multi-metal system (Cd + Pb + Zn + Cu).

heavy metal solutions containing initial concentrations ranging


from 0 to 8 mmol L1 were equilibrated with HA (0.2 g). The
HA suspensions and appropriate blanks were stirred at 500 rpm
for 24 h to reach apparent equilibrium based on our preliminary
kinetic experiments (Fig. 1a) and as reported by Gmez del Ro
et al. [15] and Peld et al. [16].
Assuming wastewaters from industries contain multiple
heavy metals, in multi-metal sorption experiments solutions of
Cd, Pb, Zn, and Cu were mixed in the following way: concentration of Cd ranging from 0.089 to 4.45 mmol L1 , Pb
ranging from 0.048 to 2.48 mmol L1 , Zn ranging from 0.153 to
7.65 mmol L1 , and Cu ranging from 0.157 to 7.85 mmol L1 .
In the multi-metal suspensions initial and final pH was measured but not adjusted [8,13,22].

404

A. Corami et al. / Journal of Colloid and Interface Science 317 (2008) 402408

To infer the contribution of the intraparticle diffusion process


in the sorption of heavy metals, the model by Weber and Morris [26] was used. A linear plot of the amounts of sorbed metals
versus the square root of time suggests the involvement of intraparticle diffusion in the overall sorption mechanism and if
the line passes through the origin the intraparticle diffusion is
the rate-controlling step of the process.
2.3. Analytical methods
The HA suspensions and blanks were filtered through a
0.20 m Nucleopore polycarbonate membrane filter and analyzed for total Cd, Pb, Zn, Cu, and Ca by inductively coupled
plasma atomic emission spectrometry (ICP-AES) using a Varian Vista RL CCD Simultaneous spectrometer. For each of the
analyzed element the detection limits were 0.1, 1.4, 0.6, 0.3, and
0.3 mol L1 , respectively. Analytical errors were estimated to
be on the order of 3%.
In order to calculate the amounts of metal ions sorbed by HA
per mass unit, the following expression was used:
V
(1)
,
M
where q is the amount of contaminant removed from solution
(mmol g1 ), C0 and Ce are the concentrations (mmol L1 ) of
the metal ions in the initial solution and at the equilibrium after the experiment, respectively, V (mL) is the volume of the
solution and M (g) the amount of HA used.
Selected solid residues were analyzed by X-ray diffractometry using a Seifert MZ IV diffractometer operating with CuK
radiation at 35 kV and 20 mA. Scans were made from 5 to
60 with a rate of 2 s/0.05 2 . All XRD analyses were performed using back-filled, randomly oriented mounts. The solid
samples were also examined with a Zeiss MD 940 scanning
electron microscope operating at 25 kV equipped with X-ray
energy dispersive spectroscopy (EDS).
FTIR spectra of the solid residues were also collected at
room temperature using the DRIFT (Diffuse Reflectance Infrared Fourier Transform) technique with an Interferometer
Equinox 55 Bruker after the samples being dispersed in KBr
excess (sample/KBr ratio about 1/100) and having the sample compartment carefully purged. The spectra resolution was
1 cm1 or better, cumulating at least 200 scans.
pH was measured using a pH 510 Eutec pH-meter to obtain
information about the change of hydrogen ion concentrations
during the metal uptake.
q = (C0 Ce )

3. Results and discussion


3.1. Characterization of solid materials
The diffraction patterns did not detect new crystalline
phases. However, the precipitation of new crystalline phases
cannot be determined as XRD can normally detect phases at
concentration higher than 1% of the sample matrix.
SEM results did not reveal morphological differences in the
solid residues with respect to the starting HA (Figs. 2a and 2b).

Fig. 2. SEM photographs of HA before (a) and after (b) Cd sorption; (c) EDS
spectrum showing the Cd uptake.

No new solid phases on the surface of HA grains were shown


by EDS analyses, whereas a uniform distribution of heavy metals was detected (Fig. 2c). White zones on the HA grains have
been observed (Fig. 2b), corresponding to the sorbed heavy
metals [27].
The infrared spectra of HA before and after interaction with
heavy metals in the 4000400 cm1 region are shown in Fig. 3.
The similar FTIR spectra suggest no other phases formed during the heavy metal sorption. The sharp band at 3566 cm1 is
assigned to the stretching vibrations of the bulk OH ions [14,
28]. The bands 3 near 1027 cm1 and 4 near 575 cm1 rep-

A. Corami et al. / Journal of Colloid and Interface Science 317 (2008) 402408

405

Fig. 3. FTIR spectra of HA before and after uptake of Cd.

resent the vibration mode of the tetrahedral PO3


4 [29,30]. The
splitting of the 4 vibrational band indicates the low site symmetry of molecules, as the observed bands confirm the presence
of more than one distinction site for the phosphate group [30].
Also the group of bands around 2000 cm1 are assignable to
1 and the broad band at
PO3
4 [28]. The band at 1635 cm
34003200 cm1 could be assigned to water [28,31]. The 2
1 and near
and 3 peaks of CO2
3 are located near 875 cm
1
1400 cm [14].
3.2. Sorption of Cd in single- and multi-metal systems
The kinetics of Cd removal by HA in both systems shows a
similar pattern (Fig. 1a), although the process in the multi-metal
system was more rapid. The Cd sorption can be decomposed in
two steps: in the first step a high rate of sorption occurred, the
second step being slower before reaching the equilibrium. The
plot in Fig. 1b, showing two almost straight lines with different
slopes, suggests the contribution of the intraparticle diffusion in
the uptake of Cd for both single- and multi-metal systems.
In single-metal system the amount of sorbed Cd ranged
from 0.058 (C0 = 0.089 mmol L1 ) to 1.681 mmol (C0 =
4.45 mmol L1 ) per gram of HA. This is a result of the increase
in the driving force, i.e., the concentration gradient, as an increase in the initial metal ion concentrations [32].
In the Cd + Pb + Zn + Cu system, competitive sorption
among the heavy metals affected retention of Cd2+ by HA.
Sorption was lower compared to single-metal system, with values of sorbed metal ranging from 0.010 to 0.618 mmol of Cd
per gram of HA. Competition among the four heavy metals determined reduction of the amount of sorbed Cd by 63% (C0 =
4.45 mmol L1 )83% (C0 = 0.089 mmol L1 ) compared to the
single-metal system. The results of our experiments showed
that the sorption behavior of Cd was altered by the competitive
effects among the metals in agreement with Chen et al. [33].

Fig. 4. Langmuir isotherm plots for sorption of Cd. Symbols as in Fig. 1.

Indeed, in single-metal sorption only internal competition (between ions of the same metal) and competition with H+ for
adsorption sites affected heavy metal sorption by HA, whereas
in the multi-metal system competition occurs also among each
heavy metal for precipitation and for adsorption sites.
Metal uptake resulted in a pH reduction from 7.04 to 4.93.
According to the results of previous studies [13,16,34] the maximum Cd sorption capacity of HA occurs at pH values included
in the observed range.
Several mathematical models have been developed to quantitatively express the relationship between the extent of sorption
and the residual solute concentration. One of the most widely
used is the Langmuir adsorption isotherm model, representing the most suitable model of monolayer adsorption [35]. The
model is based on the assumption that there is a finite number
of adsorption sites. All sites are equivalent and there is no interaction between adsorbed ions.
The general form of the Langmuir isotherm is
Ce
Ce
1
=
+
,
qe
qm bqm

(2)

406

A. Corami et al. / Journal of Colloid and Interface Science 317 (2008) 402408

Table 2
Langmuir parameters for Cd sorption in single- and multi-metal systems together with R2 and qmix /qm
System

qm (mmol g1 )

b (L mmol1 )

RL

R2

qmix /qm

Single-metal
Multi-metal

2.58
2.02

3.91
1.78

0.28
0.71

0.983
0.982

0.78

Table 3
The isotherm shape represented by the RL value
RL value

Type of isotherm

RL > 1
RL = 1
0 < RL < 1
RL = 0

Unfavorable
Linear
Favorable
Irreversible
Fig. 5. Metal disappearance versus final solution pH. Symbols as in Fig. 1.

(mmol L1 )

where Ce
is the equilibrium metal concentration,
qe (mmol g1 ) denotes the amount of Cd sorbed per unit weight
of HA, qm (mmol g1 ) means the Langmuir maximum sorption capacity of the surface and b (L mmol1 ) indicates the
adsorption constant related to bonding-energy of the adsorbate
to the adsorbent. The Langmuir linear isotherms (Fig. 4) indicate that the experimental data fitted well the model. Table 2
reports qm and b values obtained by the application of the Langmuir equation to single- and multi-metal sorption. The values of
qm and b were higher in single-metal than in multi-metal system according to Mohan and Singh [36] and Cao et al. [22].
The dimensionless constant separation factor RL represents the
main characteristics of Langmuir isotherm and is defined by
RL =

1
.
(1 + bC0 )

The shape of the isotherms is represented by the RL value as


follows in Table 3.
Our RL values (Table 2) were 0 < RL < 1 suggesting favorable sorption of Cd onto HA in both systems [17,37].
The ratio between qmix (maximum sorption capacity for the
metal in a multi-metal system) and qm (maximum sorption
capacity of the same metal when it is the only metal ion in
solution) shows the result of the competition between Cd and
Pb + Zn + Cu. Indeed, when the ratio is >1 the sorption is easier for the presence of the other metals; when qmix /qm = 1 there
are not observable competitive effects; when the ratio is <1 the
sorption of the metal is hindered by the presence of the other
metal ions.
The qmix /qm value (Table 2) is found to be <1, indicating
that the sorption capacity of HA decreases in multi-metal system compared to single-metal system.
The maximum sorption capacity per SSA of HA has to
be considered a constant value for any HA according to the
conditions of sample preparation [34]. In spite of our experimental qm value is different from those of previous studies [13,15,34], normalizing this datum to SSA the value is
5.2 102 mmol m2 , it is almost close to those of the above
studies.

3.3. Cd uptake mechanism


The sorption kinetic results (Fig. 1a) are in agreement with
those of previous studies [5,8,34,38]. The overall removal of
Cd by HA appears to be due to a two-step mechanism. The first
step involves the rapid surface complexation of Cd ions on the
surface of HA particles. In the second step the diffusion of metal
ions within the HA particles through the ion exchange with Ca
occurs, leading to the formation of a Cd-containing HA.
Support to the surface complexation mechanism is provided
by our XRD results as they did not detect any new crystalline
phase in the solid residues, FTIR analysis since the spectra of
the blanks and the solid residues are identical, SEM results
which did not show any kind of morphological differences in
the solid residues compared with the original HA, whereas EDS
analysis confirm the presence of Cd on the HA surface.
The pH decrease (Fig. 5) was already observed in previous studies [13,22,39]. Wu et al. [40] showed that when pH <
pHPZC POH are the significant surface species, whereas at pH
close and higher than pHPZC the dominant surface species are
PO and Ca(OH)+
2 , respectively. Our pH values <pHPZC
suggest that POH should be the dominant surface species.
Therefore, the observed pH decrease can be explained by proton
leaching from POH sites of HA because of surface complexation of Cd [13,22,39].
The complexation of Cd on the HA surface displaced partially the H+ ions [13,22,40], determining pH decrease and
calcium release (Fig. 6). This mechanism can be represented
by the following general reaction:
HAOH + Cd2+ = HAOCd+ + H+ .

(3)

The linear part of the plot in Fig. 1b for contact times 24


48 h suggests the diffusion of Cd2+ within the HA particles
took place, after the surface complexation was completed, in
the second step of the sorption process. However, the intraparticle diffusion process was not the rate-determining step because
of the deviation of the plot from the origin.
Further contribution to retrieve the sorption mechanisms has
been provided by the values of the molar ratios (Qs ) of Cd2+
fixed by HA to Ca2+ released. Equal amounts of bound and

A. Corami et al. / Journal of Colloid and Interface Science 317 (2008) 402408

407

formula Cdx Ca10x (PO4 )6 (OH)2 in agreement with previous


studies [5,8].
The end product of the two-step mechanism would be a
cadmium-containing hydroxyapatite.
4. Conclusions
This study investigated the sorption of aqueous Cd onto hydroxyapatite surfaces.The results have shown that:

Fig. 6. Relationship between amount of Ca2+ released in solution and the


amount of sorbed Cd2+ . Symbols as in Fig. 1.

released cations (Qs = 1) indicate the ion exchange of cations


between the HA and the solution. A quantity of fixed metal ions
larger than that of the released ones (Qs > 1) suggests surface
complexation was the main sorption mechanism. Qs < 1 indicates dissolution of HA and deposition of new phosphate with
lower cation to phosphate molar ratio. Our Qs values in the
single-metal system were generally greater than one, although
some values are close to one. In the multi-metal system the
sorption of Pb, Cu, and Zn besides Cd determined a larger release of Ca by HA compared to single-metal system, resulting
in Qs lower than one. These results are compatible with both
surface complexation and ion exchange.
Fig. 6 shows the relation between Cd2+ uptake and Ca2+
release. A linear correlation exists for both single- and multimetal systems with correlation coefficients higher than 0.98.
The observed relationship is consistent with the ion exchange
mechanism between Cd2+ and Ca2+ , which could be depicted
by the following general reaction:
HACa2+ + Cd2+ = HACd2+ + Ca2+ .

(4)

However, the slopes of the curves in Fig. 6 were 0.96 and 1.13,
respectively, suggesting an almost equimolar exchange of Cd2+
and Ca2+ in the single-metal system and a non-equimolar exchange in the multi-metal system. The apparent non-equimolar
exchange for the multi-metal system depends on the higher
amounts of Ca2+ in solution compared to the single-metal systems, which could be explained by the simultaneous sorption
of Pb, Cu, and Zn besides Cd. However, the determining of the
contribution of the ion exchange process to the overall metal
immobilization is complicated by the partial HA dissolution,
produced by the pH decrease, which partly contributed to the
total amount of Ca2+ in solution. The substitution of Ca by
other cations in HA crystal structure is the product of the ion
exchange process [41]. The proposed mechanism for the second step of the Cd overall uptake is through either the ion
exchange with Ca2+ and the partial dissolution of HA with subsequent precipitation of a Cd-containing hydroxyapatite with

Synthetic hydroxyapatite is able to immobilize Cd from


aqueous solutions both in single- and multi-metal (Cd +
Pb + Zn + Cu) systems. In single-metal system the amount
of sorbed Cd ranged from 0.058 to 1.681 mmol per gram of
HA. In the Cd + Pb + Zn + Cu system, competitive sorption among the heavy metals affected uptake of Cd2+ by
HA, resulting in the reduction of the amount of sorbed Cd
by 6383% compared to the single-metal system.
Our results support a two-step mechanism involved in the
removal of Cd by HA. Firstly, a rapid surface complexation
of Cd on the POH site of HA and secondly either an ion
exchange with Ca2+ and the precipitation of Cd-containing
hydroxyapatite.
The low effectiveness of HA in the multi-metal system
must be considered for a possible role of hydroxyapatite
for heavy metal polluted water processing.
Acknowledgments
This study formed part of the Ph.D. research of A.C. Funding for this work was partly provided by CNR-IGG-Roman
Branch. FTIR analyses were carried out in the laboratories
of CNR-IGG-Roman Branch at the Department of Earth Sciences, Sapienza University of Rome. The authors are thankful
to T. Coppola, S. Nunziante-Cesaro, and S. Stellino for their
laboratory assistance. Thanks also go to L. Conforto for her
kind help. Thanks are also extended to Jerry Bingham, School
of Natural Resources, Ohio State University, Cincinnati (USA)
for his help in BET measurements.
References
[1] C.N. Mulligan, R.N. Yong, B.F. Gibbs, Eng. Geol. 60 (2001) 193.
[2] A. zer, H.B. Pirini, J. Hazard. Mater. B 137 (2006) 849.
[3] S. Mandjiny, K.A. Matis, A.I. Zouboulis, M. Fedoroff, J. Jeanjean, J.C.
Rouchaud, N. Toulhoat, V. Potocek, C. Loos-Neskovic, P. Maireles-Torres,
D. Jones, J. Mater. Sci. 33 (1998) 5433.
[4] T. Miyahara, M. Miyakoshi, Y. Saito, H. Kozuka, Toxicol. Appl. Pharm. 55
(1980) 477.
[5] D. Marchat, D. Bernache-Assollant, E. Champion, J. Hazard. Mater. 139
(2007) 453.
[6] S. Al-Asheh, Z. Duvnjak, J. Hazard. Mater. 56 (1997) 35.
[7] S.F. Montanher, E.A. Oliveira, M.C. Rollemberg, J. Hazard. Mater. B 117
(2005) 207.
[8] S. Raicevic, T. Kaludjerovic-Radoicic, A.I. Zouboulis, J. Hazard. Mater.
B 117 (2005) 41.
[9] A.B. Prez-Marn, V. Meseguer Zapata, J.F. Ortuo, M. Aguilar, J. Sez,
M. Llorns, J. Hazard. Mater. B 139 (2007) 122.
[10] B.S. Crannell, T.T. Eighmy, J.E. Krzanowski, J.D. Eudsen Jr., E.L. Shaw,
C.A. Francis, Waste Manage. 20 (2000) 135.

408

A. Corami et al. / Journal of Colloid and Interface Science 317 (2008) 402408

[11] T. Suzuki, T. Hatsushika, M. Michihiro, J. Chem. Soc. Faraday


Trans. 1 (78) (1982) 3605.
[12] Q.Y. Ma, T.J. Logan, S.J. Traina, J.A. Ryan, Environ. Sci. Technol. 28
(1994) 1219.
[13] Y. Xu, F.W. Schwartz, S.J. Traina, Environ. Sci. Technol. 28 (1994) 1472.
[14] A.G. Leyva, J. Marrero, P. Smichowski, D. Cicerone, Eviron. Sci. Technol. 35 (2001) 3669.
[15] J.A. Gmez del Ro, P.J. Morando, D.S. Cicerone, J. Environ. Manage. 71
(2004) 169.
[16] M. Peld, K. Tnsuaadu, V. Bender, Environ. Sci. Technol. 38 (2004)
5626.
[17] A. Corami, S. Mignardi, V. Ferrini, J. Hazard. Mater. 146 (2007) 164.
[18] I. Smiciklas, S. Dimovic, I. Pleca, M. Mitric, Water Res. 40 (2006)
2267.
[19] A. Yasukawa, T. Yokoyama, K. Kandori, T. Ishikawa, Colloids Surf. A 299
(2007) 203.
[20] R.R. Sheha, J. Colloid Interface Sci. 310 (2007) 18.
[21] J. Jeanjean, U. Vincent, M. Fedoroff, J. Solid State Chem. 108 (1994) 68.
[22] X. Cao, L.Q. Ma, D.R. Rhue, C.S. Appel, Environ. Pollut. 131 (2004) 435.
[23] Q.Y. Ma, T.J. Logan, S.J. Traina, Environ. Sci. Technol. 29 (1995) 1118.
[24] Powder Diffraction File, 1998.
[25] S.K. Milonjic, A.R. Ruvarac, M.V. uic, Thermochim. Acta 11 (1975)
261.
[26] W.J. Weber, J.C. Morris, J. Sanit. Eng. Div. Am. Soc. Civ. Eng. 89 (1963)
31.

[27] S. McGrellis, J.-N. Serafini, J. Jeanjean, J.-L. Pastol, M. Fedoroff, Sep.


Purif. Technol. 24 (2001) 129.
[28] Z.H. Cheng, A. Yasukawa, K. Kandori, T. Ishikawa, Langmuir 14 (1998)
6681.
[29] D.S. Soejoko, M.O. Tjia, J. Mater. Sci. 38 (2002) 2087.
[30] I. Rehman, W. Bonfield, J. Mater. Sci. Mater. Med. 8 (1997) 1.
[31] S. Rayanaud, E. Champion, D. Bernache-Assollant, P. Thomas, Biomaterials 23 (2002) 1065.
[32] K. Chojnacka, Chemosphere 59 (2005) 315.
[33] X. Chen, J.V. Wright, J.L. Conca, L.M. Peurrung, Environ. Sci. Technol. 31 (1997) 624.
[34] N.C.C. Da Rocha, R.C. De Campos, A.M. Rossi, E.L. Moreira, A.F. Barbosa, G.T. Moure, Environ. Sci. Technol. 36 (2002) 1630.
[35] R. Han, H. Li, Y. Li, J. Zhang, H. Xiao, J. Shi, J. Hazard. Mater. 137 (2006)
1569.
[36] D. Mohan, K.P. Singh, Water Res. 36 (2002) 2304.
[37] G. McKay, H.S. Blair, J.R. Garden, J. Appl. Polym. Sci. 27 (1982) 3043.
[38] J. Gmez del Ro, P. Sanchez, P.J. Morando, D.S. Cicerone,
Chemosphere 64 (2006) 1015.
[39] E. Mavropoulos, A.M. Rossi, A.M. Costa, C.A.C. Perez, J.C. Moreira,
M. Saldanha, Environ. Sci. Technol. 36 (2002) 1625.
[40] L. Wu, W. Forsling, P.W. Schindler, J. Colloid Interface Sci. 147 (1991)
178.
[41] E. Monteil-Rivera, M. Fedoroff, in: Encyclopedia of Surface and Colloid
Science, Dekker, New York, 2002, p. 1.

Вам также может понравиться