Sei sulla pagina 1di 34

Journal of Hazardous Materials 252–253 (2013) 428–461

Contents lists available at SciVerse ScienceDirect

Journal of Hazardous Materials


journal homepage: www.elsevier.com/locate/jhazmat

Review

A review on zinc and nickel adsorption on natural and modified


zeolite, bentonite and vermiculite: Examination of process
parameters, kinetics and isotherms
S. Malamis a,b,∗ , E. Katsou a,b
a
School of Chemical Engineering, National Technical University of Athens, 9 Iroon Polytechniou St., Zographou Campus, 15773, Athens, Greece
b
Department of Biotechnology, University of Verona, Strada Le Grazie 15, 37134, Verona, Italy

h i g h l i g h t s

• Important parameters: sorbate and sorbent type and concentration, pH, ionic strength.
• Presence of competing ions adversely impacts on process.
• Variability in nickel and zinc removal due to parameters variation.
• Langmuir is the isotherm that best suits data in most cases.

a r t i c l e i n f o a b s t r a c t

Article history: Adsorption and ion exchange can be effectively employed for the treatment of metal-contaminated
Received 7 January 2012 wastewater streams. The use of low-cost materials as sorbents increases the competitive advantage of
Received in revised form 22 February 2013 the process. Natural and modified minerals have been extensively employed for the removal of nickel
Accepted 11 March 2013
and zinc from water and wastewater. This work critically reviews existing knowledge and research on
Available online xxx
the uptake of nickel and zinc by natural and modified zeolite, bentonite and vermiculite. It focuses on the
examination of different parameters affecting the process, system kinetics and equilibrium conditions.
Keywords:
The process parameters under investigation are the initial metal concentration, ionic strength, solution
Zinc
Nickel pH, adsorbent type, grain size and concentration, temperature, agitation speed, presence of competing
Adsorption ions in the solution and type of adsorbate. The system’s performance is evaluated with respect to the
Ion exchange overall metal removal and the adsorption capacity. Furthermore, research works comparing the process
Minerals kinetics with existing reaction kinetic and diffusion models are reviewed as well as works examining the
Kinetics performance of isotherm models against the experimental equilibrium data.
Equilibrium studies © 2013 Published by Elsevier B.V.
Process parameters

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
2. Metal removal mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
3. Mineral characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
4. Parameters impacting on adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
4.1. Initial metal concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
4.2. Liquid medium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
4.3. Solution pH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
4.4. Temperature and thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 434
4.5. Adsorbent grain size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 436
4.6. Adsorbent concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437
4.7. Adsorbent type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437

∗ Corresponding author at: School of Chemical Engineering, National Technical University of Athens, 9 Iroon Polytechniou St., Zographou Campus, 15773, Athens, Greece.
Tel.: +30 210 7723108; fax: +30 210 7723285.
E-mail addresses: malamis.simos@gmail.com (S. Malamis), ekatsou@yahoo.gr (E. Katsou).

0304-3894/$ – see front matter © 2013 Published by Elsevier B.V.


http://dx.doi.org/10.1016/j.jhazmat.2013.03.024
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 429

4.8. Competitive adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438


4.9. Ionic strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
4.10. Agitation speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
4.11. Evaluation of parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
5. Removal efficiencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
6. Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
7. Isotherms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452
8. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 457
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 457

1. Introduction applications, adsorption and ion exchange can be grouped together


as sorption for a unified consideration.
Intense industrial activity results in the contamination of waste-
water with several heavy metals, including zinc and nickel. The 2. Metal removal mechanisms
extensive use of nickel and zinc in industry means that consider-
able amounts of these metals can find their way into the aquatic To investigate the suitability of an adsorbent, it is necessary
environment. Nickel is employed in several industrial activities, to characterize its porous structure and its surface composition
including mineral processing, electroplating, production of paints which consists of several functional groups. Adsorption of metals
and batteries, manufacturing of sulphate and porcelain enamelling on the active sites results in the formation of outer and inner-
[1–3]. Zinc is released into the aquatic environment through several sphere complexes. Outer-sphere complexes occur if at least one
industrial activities, such as mining, metal coating, battery produc- water molecule is interposed between the surface functional group
tion and its use in paints, ceramics, wood, fabrics, drugs, sun blocks and the bound ion. On the contrary, inner-sphere complexes occur
and deodorants [4–6]. Human exposure to these heavy metals at when water molecules are not interposed between the bound ion
significant levels is associated with serious health effects. Nickel and the functional group of the adsorbent and the sorbate forms a
is associated with dermatitis, nausea, coughing, chronic bronchi- direct coordinate-covalent bond with surface functional groups on
tis, gastrointestinal distress, reduced lung function and lung cancer the variable charge surface [19]. Specific adsorption involves the
[3,7–9]. Zinc is associated with short-term “metal-fume fever”, nau- formation of inner-sphere complexes, while non-specific adsorp-
sea, diarrhoea, depression, lethargy, and neurological signs, such as tion involves the formation of weaker outer-sphere complexes.
seizures and ataxia [7–9]. Consequently, there is a need to effec- Outer-sphere complexes are formed through electrostatic interac-
tively remove these metals from wastewater in order to ensure tions between metal cations and the negatively charged centres.
adequately treated effluent quality for various uses. Several pro- Therefore, metal binding takes place through ion exchange at the
cesses are available for the removal of heavy metals from water planar sites of the adsorbent [26,27]. At the earlier adsorption
and wastewater including chemical precipitation, adsorption, ion stages, outer-sphere complexes are formed at the external sur-
exchange, flotation, membrane filtration, electrochemical treat- face sites. As metal concentration increases, metal ions are forced
ment and coagulation–flocculation. The advantages and disadvan- into internal surface sites forming also inner-sphere complexes.
tages of each process have been summarized by O’Connell et al. [10]. These complexes are more stable compared to the outer-sphere
Adsorption and ion exchange processes can be used for the complexes due to the formation of covalent bonds or combinations
removal of heavy metals from wastewater. The competitive advan- of covalent and ionic bonding. During their formation, protons are
tage of such processes can increase when low-cost sorbents/ion released and the process causes a total decrease in solution pH.
exchangers are used. A wide variety of heavy metals can be removed The type of surface complexes affects the rate and the reversibil-
from the liquid phase at significant quantities and kinetics are ity of the adsorption. Outer-sphere complexation is usually rapid
usually fast. The drawbacks are that the performance depends and reversible, while inner-sphere complexation is slower and may
on sorbent’s selectivity towards the metal, while in many cases be irreversible [28]. Inner-sphere complexation is influenced by the
pre-treatment is required to improve its adsorption capacity [10]. ionic strength of the solution. Outer-sphere complexation is partic-
Various natural minerals have been applied for the removal of ularly influenced by pH, since at pH < 4 most silanol and aluminol
nickel and zinc. In many cases these minerals have been chem- groups on the edges are protonated and thus metal adsorption
ically and/or thermally modified to increase the metal uptake. decreases [26,29]. Also, the metals can be transferred from the bulk
Also, in some cases a mixture of minerals with other sorbents is solution to the solid phase due to precipitation on the mineral’s
employed to treat wastewater [11]. Some reviews summarize the surface. Mineral addition in electrolyte solution can reduce metal
work conducted on the removal of heavy metals by different sor- solubility and enhance metal precipitation on the mineral surface
bents [12–19]. Certain reviews have focused on the examination even in a non-saturated environment. At low surface coverage com-
of kinetics and others on equilibrium [20–25]. Nevertheless, the plexation tends to dominate. As the surface coverage increases
uptake of zinc and nickel by natural and modified minerals has nucleation occurs and results in the formation of distinct entities or
not been adequately summarized. The present work aims to crit- aggregates on the surface; as loading increases further, surface pre-
ically review information concerning the factors that impact the cipitation becomes the dominant mechanism [6,28]. Precipitation
adsorption and ion exchange, the system kinetics and equilibrium mostly depends on solution pH and initial metal concentration [14].
with respect to various models that have been applied in the liter- Moreover, co-precipitation of metals may play an important role in
ature. Much of the information summarized in this work has been the surface precipitation mechanism, where the metal is drifted
tabulated; this way it is easy to compare the experimental condi- from another metal that is contained in high concentration in the
tions and results of various works. Since both ion exchange and solution [30]. Surface precipitation is favoured by the pH increase.
adsorption are usually involved in the uptake of heavy metals by Hence, the addition of minerals that significantly increases the pH
minerals, in the context of this work hereafter the term adsorp- enhances metal removal from the bulk solution [14,19]. Precipita-
tion also includes the ion exchange that takes places. In practical tion is a much slower process than complex formation [31].
430 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

3. Mineral characteristics physicochemical factors related to the sorbate, the sorbent, and
the liquid phase characteristics including the solution pH, initial
Zeolites are natural, hydrated aluminosilicate minerals with metal concentration in the solution, presence of other cations, ionic
crystalline microporous structure and belong to the mineral type of strength, temperature, type of the liquid phase, mineral type, grain
‘tectosilicates’. The zeolite structure consists of three-dimensional size and concentration, type of sorbent pre-treatment and agitation
frameworks of SiO4 and AlO4 that are linked to each other by speed.
sharing oxygen atoms. The replacement of tetravalent silicon by
trivalent aluminium in the mineral’s structure creates a net neg- 4.1. Initial metal concentration
ative charge that is counterbalanced by the presence of cations
(usually Ca2+ , Na+ and K+ ) which are situated in cavities [32,33]. The initial metal concentration in the liquid phase significantly
These cations are exchangeable with other cations including heavy affects the adsorption process. The impact depends on several
metals. The zeolite’s framework structure encloses cavities occu- parameters, such as the type of metal and the liquid medium, the
pied by large ions and water molecules which are able to move, presence of competing cations, the availability of the functional
allowing ion exchange. Zeolites can be classified by considering the groups on the adsorbent’s surface and the ability of these groups to
topology of the framework. Zeolites having the same topology con- bind metal ions. In most cases, the increase of the initial metal con-
stitute a zeolite framework type. Currently there are 206 known centration results in an increase in the amount of metal adsorbed
framework types recognized by the Structure Commission of the per unit mass of adsorbent until a plateau is reached and in a
International Zeolite Association [33]. An important feature is the decrease in the overall removal efficiency [40–43].
effective width of the channels which determines the accessibility At low initial metal concentrations the ratio of metal cations
of the channels to incoming cations. Zeolite channels are formed by to adsorbent mass is low and thus adsorption does not depend on
different combinations of linked rings of tetrahedra. The channels initial concentration [44,45]. Increasing the initial concentration
that are wider at their narrowest parts can receive larger cations means that more metals are available and thus, more metal ions
into their structure [32]. are adsorbed for constant adsorbent mass. At higher initial metal
Bentonite is an aluminium phyllosilicate (clay) consisting concentrations the driving force to overcome the mass transfer
mostly of montmorillonite. The main characteristic of montmoril- resistance for the migration of the metals from the bulk solution
lonite is its ability to absorb water molecules between its sheets, to the mineral surface increases [9]. However, each unit mass of
resulting in a significant expansion [34]. The crystal structure of adsorbent is subjected to a larger number of metal cations, which
montmorillonite is composed of two layers of silica tetrahedra with gradually fill up the sites until saturation is reached. In such a
a central layer of alumina octahedron between them [30]. The tetra- case, further increase of metal concentrations is not accompanied
hedral and octahedral sheets are combined in such a way that the with an increase in the amount of metal adsorbed per unit mass
tips of the tetrahedra of each silica sheet and one of the hydroxyl of adsorbent. The determination of the maximum metal concen-
layers of the octahedral sheet form a common layer [35]. The dis- tration where total saturation of the adsorbent occurs is important
tance between the layers is not fixed since the layers can expand. for practical applications. Furthermore, the level of surface precip-
Due to its layered structure the mineral is characterized by internal itation strongly depends on the initial metal concentration. At low
(i.e. interlayer) and external surfaces (i.e. edge). Different exchange- metal concentrations, the adsorbent surface coverage is low and
able cations can replace the ions in the montmorillonite structure. the formation of surface complexes is the main mechanism. The
In particular, in the tetrahedral sheets the tetravalent silicon can increase of metal concentration favours the concentration of com-
be replaced by trivalent aluminium, while in the octahedral sheets pounds and aggregates on mineral surface. A further increase in
the aluminium can be substituted by other ions such as divalent metal concentration results in saturation of adsorption sites and
magnesium. As a result, the layer charge becomes negative and surface precipitation is the main uptake mechanism [46]. The sat-
is mostly neutralized by hydrated cations in the interlayer space. uration of active sites is usually faster in the cases where minerals
These cations are bonded by electrostatic forces to the internal exhibit low selectivity for a metal. However, this is not always the
surfaces and can be exchanged with other cations [30]. fact, since the metal complexes formed or entrapped in the min-
Vermiculite is a type of clay that is usually found in nature in its eral’s surface may accelerate the surface saturation even at low
magnesium form. It is composed of two layers of silicon tetrahedra, concentrations of metals that present high affinity for the specific
in which silicon is partially replaced by aluminium, and a layer of mineral.
OH− groups and magnesium ions, forming a strongly bound mica Several studies have documented the effect of initial metal
stack. A hydrated layer is located between the stacks consisting concentration on metal uptake by minerals as shown in Table 1
of divalent metals. Thus, the two layers are bonded together by [40,42,44,47–65]. In most cases the increase of initial concentration
one octahedral layer and the structure is commonly referred to as resulted in an increase of the adsorption capacity and a decrease
2:1 phyllosilicate [36,37]. The mineral’s structural analysis shows of the overall removal efficiency. Ijagbemi et al. [42] examined
that the water molecule sites form a distorted hexagonal arrange- the adsorption of nickel on montmorillonite and found that the
ment with each site being linked to single oxygen on the silicate adsorption capacity increased with increasing metal concentration,
layer surface. The magnesium ions lie between the water sheets while after a certain concentration a plateau was reached. Çoruh
and are octahedrally coordinated with the water molecules [38]. and Ergun [40] found that by increasing the initial nickel concen-
When tetravalent silicon is replaced by trivalent aluminium in the tration the equilibrium solid phase concentration (qe ) increased
tetrahedral layer of the vermiculite sheet, a negative charge is gen- for both natural and chemically modified clinoptilolite, while the
erated on the layer which is counterbalanced by interlayer cations; removal efficiency decreased. The latter has also been reported for
in the case of vermiculite these are mainly magnesium cations that the adsorption of nickel on montmorillonite [44,47], on natural
are easily exchanged with other cationic species [32,39]. zeolite [48] and on calcined Bofe bentonite [49].
A similar picture was obtained for zinc adsorption on natural
and modified minerals. Erdem et al. [62] examined the adsorption
4. Parameters impacting on adsorption of Zn2+ on natural zeolite for varying initial metal concentration
and found that the adsorption efficiency decreased with increasing
The mineral’s adsorption capacity and the system’s kinet- metal concentration. Kaya and Ören [58] found that the increase of
ics for metal uptake from the liquid phase depends on several zinc initial concentration in the solution resulted in an increase of
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 431

Table 1
Studies on the effect of initial zinc and nickel concentration (C0 ) on the adsorption process.

Metal Mineral C0 range C0 for which highest C0 for which highest Reference
(mg L−1 ) adsorption is obtained (mg L−1 ) (%) removal occurs
(mg L−1 )

Ni Clinoptilolite 12.5–200 200 12.5 [40]


NaCl clinoptilolite
HCl and NaCl clinoptilolite
Ni Montmorillonite 50–300 200–300 [42]
Ni Montmorillonite 10–60 60 10 [44,47]
ZrO-montmorillonite
TBA-montmorillinite
Ni Natural zeolite 1–100 100 [48]
Ni Calcined Bofe bentonite 100–400 100 [49]
Nia Clinoptilolite 1–30 10 [50,51]
Chabazite
Ni Clinoptilolite 10–800 10 [52]
Nia Jordanian zeolite 50–400 400 [53]
(phillipsite) tuff
Nia Clinoptilolite 50–3000 10 [54]
Ni Zeolite 5.87, 31.69, 5.87 [55]
61.63
Ni Na-montmorillonite 50–100 100 50 [56]
Acid modified montmorillonite 75
a
Zn Clinoptilolite 1–30 10 [50,51]
Chabazite
Zna Clinoptilolite 50–3000 10 [54]
Zn Zeolite 6.54, 30.73, 6.54 [55]
63.42
Zn Vermiculite 25–500 500 25 [57]
Zn Natural bentonite 1–200 ∼60–200 [58]
Na-bentonite ∼80–200
Zn Natural zeolite 37.27–659.03 32.27 [59]
Zn Gordes zeolite 1–20 14 (14–20) [60]
Bigadic zeolite 14 (14–20)
Zn Cankiri bentonite 20–160 ∼80–160 [61]
Zn Zeolite 50–400 50 [62]
Zn Bentonite 30–50 50 [63]
Zn Zeolite 5–20 20 [64]
Zn Turkish zeolite 78.59–1012.78 5.63–78.59 [65]
a
Mixed metal solution.

qe on natural bentonite and Na-bentonite up to a certain point and • presence of competing ions in wastewater for which minerals
thereafter remained almost constant. A similar trend was observed may exhibit higher selectivity [68,69],
by Ören and Kaya [60] for zinc uptake by Gordes and Bigadic zeo- • adsorption of metals on suspended solids contained in wastewa-
lite. Veli and Alyüz [61] studied the adsorption of zinc on Cankiri ter and/or metals precipitation [70,71]; this reduces the soluble
bentonite and reported that adsorption increased up to a certain metal concentration available for mineral adsorption
level, and thereafter remained stable with further concentration • formation of soluble, stable complexes that are not adsorbed on
increase. Perić et al. [59] reported that the increase of initial zinc the mineral’s surface [27]. In this case the formation of solu-
concentration resulted in a decrease of metal adsorption on natural ble complexes becomes competitive to surface complexation and
zeolite. Wu et al. [57] reported that the increase of initial zinc con- metal adsorption decreases [28,72] due to the hindrance of metals
centration led to an increase of qe of vermiculite and a reduction to reach the active sites and/or to the capturing of adsorption sites
of the removal efficiency. Increased removal efficiencies at low ini- by ligands [73].
tial metal concentrations have also been reported by Erdem et al.
[62] and Cabrera et al. [55]. Ouki and Kavannagh [50,51] studied In some other cases, the presence of ligands may enhance
the adsorption of nickel and zinc on chabazite and found that the adsorption through the [26,27]:
maximum removal efficiency occurred at the medium range of the
• formation of negatively charged complexes between the ligands
initial metal concentrations.
and the metals, which interact with positively charged, mineral
4.2. Liquid medium surface sites,
• complexation of metals with ligands that bond to the mineral
The prediction of metals adsorption on minerals in aqueous surface,
solutions may not reflect the true mineral behaviour when complex • outer-sphere complexation between metal ions and negatively
liquids are treated. Thus, adsorption must be evaluated in the liquid charged ligand–surface complexes,
where the adsorption process actually takes place [66]. The pres- • formation of positively charged metal–ligand complexes, which
ence of ligands in wastewater may enhance or suppress adsorption, are ion exchanged with the mineral’s cations,
depending on the ligand, the adsorbent and the metal [67]. The • accumulation of negatively charged ligands in the double layer
mineral’s performance is usually lower in wastewater compared to of positively charged particles, which reduces the electrostatic
aqueous solutions, due to the: repulsion between metals and the surface.

• competition among ligands and metals for the available mineral Malandrino et al. [27] studied the impact of ligands on heavy
surface sites, metal adsorption for vermiculite and concluded that ligands having
432 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

high complexation constants resulted in significant reduction of motorway stormwater. Katsou et al. [85] investigated the removal
the mineral’s adsorption capacity. Abollino et al. [74] found that of nickel, zinc, lead and copper from multi-metal wastewater using
the presence of ligands in solutions influenced the adsorption aluminosilicate minerals combined with ultrafiltration membranes
of heavy metals on Na-montmorillonite due to the formation of and found that although the overall metal removal obtained was
metal–ligand complexes that hinder adsorption. Abollino et al. [26] higher in wastewater compared to aqueous solutions, the adsorp-
concluded that ligands and their concentration were more criti- tion capacity of minerals was lower. This was mainly attributed
cal to the adsorption process than the type of clay. The presence to the formation of particulate metal forms that limited the liquid
of anions may affect the metal adsorption due to: the formation phase metal concentration available for the adsorption process, but
of metal complexes that exhibit higher affinity for the adsorbent were successfully retained by the membranes. Researchers have
than the free metals forms; binding of metals with anions form- also studied nickel and zinc removal from wastewater using natu-
ing soluble substances that remain in the liquid phase and cannot ral and Na-modified bentonite [82], vermiculite [83], acid-alkaline,
be adsorbed by the minerals; interaction of anions with the min- physically and chemically modified clay containing montmoril-
eral, changing the state of the active sites [75,76]. Yang et al. [68] lonite and hydromica [86], natural and modified zeolite [87].
examined the effect of Cl− , NO3 − and ClO4 − on nickel adsorption on
bentonite from aqueous solutions and concluded that their impact 4.3. Solution pH
was negligible. The low removal of Ni, Hg, Cu, Fe3+ and Cr3+ was
attributed to the presence of SO4 2− , HPO4 2− and NO3 − that form It is well recognized that the pH is an important parameter
stable complexes with metal ions and decrease adsorption [77–79]. impacting on the adsorption process. Fig. 1, shows the different
The decrease in metal adsorption on bentonite due to the increase in species of zinc and nickel in aqueous solutions. At pH < 6 the Ni2+
the concentration of NaCl and CaCl2 has also been documented. The and Zn2+ forms of nickel and zinc dominate [88,89]. The solution pH
formation of stable metal complexes decreases adsorption. Also, significantly affects the ionisation degree, the metal chemistry and
the competition of Na+ and Ca2+ with heavy metals for active sites the mineral’s surface properties [6,90–92]; it impacts on the extent
may decrease heavy metal uptake [80]. of metal precipitation and the type of complexes formed between
Álvarez-Ayuso et al. [81] used natural and synthetic zeo- the ligands and metals. It is one of the most important parameters
lite (NaP1) to treat electroplating wastewater. Synthetic zeolite influencing not only site dissociation, but also the solution chem-
effectively treated acid zinc and nickel electroplating process istry of the heavy metals since hydrolysis, complexation by organic
wastewater. However, synthetic zeolite performed poorly when it and/or inorganic ligands, redox reactions, precipitation and avail-
treated zinc cyanide electroplating wastewater; this is attributed ability of heavy metals are all influenced by it [93,94]. The different
to the formation of complexes between Zn2+ and CN− . Álvarez- metal species dominating in the solution at various pH differ in
Ayuso and García-Sánchez [82] found that the performance of their charge and ability to adsorb on the mineral.
Na-bentonite for Zn removal from acid electroplating effluents The activity of the adsorbent’s functional groups is strongly
decreased by 10% compared to that obtained for the treatment affected by solution pH. The point of zero charge pH (pHpzc ) is the
of synthetic solutions, while the capacity of Na-bentonite for Zn pH of the solution at which the overall observed charge on the sur-
uptake from cyanide-containing wastewater was very low, due face of the mineral is zero. At pH < pHpzc , the functional groups are
to the formation of stable Zn(CN)4 2− complexes and the low protonated and the positively charged surfaces dominate. In this
selectivity of the mineral for these complexes. On the contrary, case, the attraction of negatively charged ions is possible to occur. In
Na-bentonite exhibited similar efficiency for Ni uptake from syn- highly acidic environment, several functional groups become pro-
thetic solutions and Ni-electroplating effluents. In another study, tonated and act as positively charged species, resulting in reduced
Álvarez-Ayuso and García-Sánchez [83] employed natural vermic- attraction between the metals and the minerals. Deprotonation of
ulite in columns to treat wastewater originating from a nickel functional groups occurs at increasing pH and these behave as neg-
electroplating process and achieved similar purification levels to atively charged moieties, attracting heavy metals [3,95]. However,
those obtained for the treatment of synthetic solutions. Pitcher at alkaline environment, the solubility of metals decreases allowing
et al. [84] found that the performance of synthetic zeolite for the precipitation, which may complicate the sorption process [45]. Pre-
removal of Zn from synthetic solutions and motorway stormwa- cipitation occurring at alkaline environment masks the true extent
ter containing Zn, Cd, Pb and Cu was similar, while zinc uptake of metal sorption on minerals; this is why it is important when
by natural zeolite (modernite) from motorway stormwater was the mineral behaviour is examined at high pH values to deter-
lower compared to that obtained for synthetic solutions. The lat- mine the soluble metal concentration prior to the addition of the
ter was attributed to the presence of dissolved contaminants in mineral.

Fig. 1. (a) Zn and (b) Ni speciation as a function of pH in dilute aqueous solutions, T = 25 ◦ C [88,89].
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 433

Table 2 [11,26,40,42,44,47–57,60,61,63,68,81–83,92,96–108] maximum removal due to adsorption is in some cases reported


summarizes the effect of pH on adsorption of nickel and zinc on zeo- to be higher than 8, this also includes precipitation that occurs
lite, bentonite and vermiculite. In all studies, the lowest adsorption at such basic pH values. Argun [98] found that the optimal solu-
occurred at acidic environment (pH ≤ 3), while the highest adsorp- tion pH for adsorption of Ni2+ on natural zeolite was 6, while at
tion was usually observed in the pH range of 5.0–8.0. Although the an acidic environment the performance of clinoptilolite reduced

Table 2
Studies showing the effect of pH on the adsorption of nickel and zinc on zeolite, bentonite and vermiculite.

Metal Mineral Examined pH range pH of highest pH of lowest Reference


adsorption adsorption

Ni Vermiculite 2.5–8.0 8.0 2.5 [26]


Montmorillonite 3.5–8.0 3.0
Ni Clinoptilolite 3–9 7–8 3 [40]
NaCl clinoptilolite
HCl and NaCl clinoptilolite
Ni Montmorillonite 2.5–9.0 6.0–8.0 2.5 [42]
Ni Montmorillonite 1–10 10 1 [44]
Ni Montmorillonite 1 [47]
ZrO-montmorillonite 1–10 10
TBA-montmorillinite
Ni Zeolite 1–10 6 1–2 [48]
Ni Zeolitea 3–6 4 3 [50,51]
Ni Chabazitea 3–6 5 3, 6 [50,51]
Ni Clinoptilolite 3.4–7.5 7.5 3.4 [52]
Ni Jordanian zeolite tuff 2.5–4.5 4.0 2.5 [53]
(phillipsite)
Ni Clinoptilolitea 1–5 4 1 [54]
Ni Zeolite ∼2.3–7.8 ∼2.3 ∼7.8 [55]
∼2.5–7.9 ∼2.5 ∼7.9
Ni Na-montmorillonite 3–9.3 9.3 3 [56]
Acid modified montmorillonite
Ni GMZ Na-bentonite 2–12 ≈8.5 2 [68]
Ni Synthetic zeolite 3–6 4–6 3 [81]
Ni Na-bentonite 3–6 6 3 [82]
Ni Natural vermiculite 3–6 6 3 [83]
Ni Zeolite 4A 3, 4 4 3 [92]
Ni Zeolite 3–9 4–7 3 [96]
Bentonite
Vermiculite
Ni Clinoptilolite 2–9 6–8 2 [98]
Ni Chitosan-clinoptilolite 2.0–6.0 5.0 2.0 [99]
composite
Ni Natural bentonite 1–5 3 1 [100]
Ni Na-bentonite 2–9 6–8 2 [101]
Ni Na-montmorillonite 3–9 9 3 [102]
Ni Vermiculite 1.2–9.0 5.2–9.0 1.2 [103]
Zn Zeolite–Portland cement 3–8 6.0 3.0 [11]
mixture
Zn Vermiculite 2.5–8.0 7.0–8.0 2.5 [26]
Montmorillonite 3.5–8.0 3.0
Zn Bentonite 2–10 6 2 [49]
Zn Clinoptilolitea 3–6 4–5 3 [50,51]
Zn Chabazitea 3–6 4–5 6 [50,51]
Zn Clinoptilolitea 1–5 1 5 [54]
Zn Zeolite ∼2.3–7.2 ∼7.2 ∼2.3 [55]
Zn Vermiculite 1.3–6.3 4.4–6.3 1.3 [57]
Zn Gordes zeolite 3–8 4–6 3 [60]
Bigadic zeolite 6
Zn Cankiri bentonite 2–10 8 2 [61]
Zn Bentonite 3.81–7.69 6.76 3.81 [63]
Zn Synthetic zeolite 3–6 6 3 [81]
Zn Na-bentonite 3–6 6 3 [82]
Zn Zeolite 4A 3, 4 4 3 [92]
Zn Zeolite 4–9 6–7 4 [97]
Bentonite
Vermiculite
Zn Zeolite 2–7 7 2 [104]
Zn Zeolite 4A 1–8 6.0 1.0 [105]
Zeolite 13X
Bentonite
Zn Natural bentonite 2–9 3 [106]
Composite E20 bentonite 5–6 2
Composite A85 bentonite 2
Zn Na-dodecylsulphate 1.5–3.5 ≈3.3 1.5 [107]
montmorillonite
Zn Bentonite <3–10 6 <3 [108]
a
Multi-metal solution.
434 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

due to the presence of protons that competed with nickel ions for removal. Purna Chandra Rao et al. [105] studied the adsorption of
the available adsorption sites. Precipitation was observed at pH zinc on zeolite 4A, zeolite 13X and bentonite and concluded that
higher than 8. Ijagbemi et al. [42] reported similar findings for nickel the highest adsorption occurred at pH 6.0. Álvarez-Ayuso et al. [81]
adsorption on montmorillonite. In another study, Ijagbemi et al. reported that NaP1 zeolite exhibited stable adsorption capacity for
[56] stated that the increasing adsorption on Na-montmorillonite Ni2+ uptake at pH ranging from 4 to 6, while the highest adsorp-
and acid modified montmorillonite that was observed when the tion capacity for Zn2+ was obtained at pH 6. Sen and Gomez [63]
solution pH increased, was attributed to the lower concentration of reported that the increase of pH up to 6.67 increased the adsorp-
H3 O+ since more sites became available to nickel ions. The authors tion capacity of bentonite towards zinc, while the further increase
also mention that surface complexation significantly contributed resulted in a subsequent reduction. Ok et al. [11] reported that the
to the adsorption process; the latter could explain the dependence highest adsorption of zinc on a zeolite–Portland cement mixture
of adsorption on solution pH. Çoruh and Ergun [40] found that occurred at pH 6. Ouki and Kavannagh [50,51] found that the low-
the highest adoption capacity of natural and modified clinoptilo- est zinc adsorption on chabazite occurred at the highest pH that
lite occurred at the pH range of 7–8, while the adsorption was was examined (i.e. pH 6), while for clinoptilolite the lowest adsorp-
also high and stable at the pH range of 4–6. Sprynskyy et al. [52] tion occurred at highly acidic environment. The selectivity order
reported that the optimal pH for adsorption of nickel on clinoptilo- of minerals for specific metals is also influenced by solution pH
lite was 7.5, while the lowest performance was obtained at pH 3.4 [26,74].
due to competition with hydrogen ions. Song et al. [102] reported
a gradual increase of sorption of Ni2+ on Na-montmorillonite as 4.4. Temperature and thermodynamics
the pH increased from 3 to 9. However, at pH 9, precipitation sig-
nificantly contributed to nickel removal, masking the true metal Temperature is directly related to the kinetic energy of the
adsorption. Yang et al. [68] concluded that at pH < 8, the sorption metal ions in the solution. An elevation in temperature subse-
of nickel on GMZ Na-bentonite was dominated by ion exchange quently results in an increase in the diffusion rate of the sorbate.
while at pH > 8 inner-sphere surface complexation was the dom- In addition, temperature alternations will affect the equilibrium
inant uptake mechanism. Katsou et al. [96] found that at the pH adsorption capacity of the sorbent for a particular sorbate [110].
range of 4–7 the removal of nickel due to adsorption on ben- Usually, at elevated temperatures the metal uptake is higher due
tonite, zeolite and vermiculite was stable, while the highest nickel to higher affinity of the mineral for the metal and/or an increase
removal obtained at pH ≥ 8 was mainly attributed to precipita- in the active sites of the solid. At higher temperature the energy
tion. Kocaoba et al. [48] obtained the maximum adsorption for of the system facilitates the attachment of metals on the mineral’s
nickel on zeolite at pH ≥ 6, but the authors mention that at the surface [111,112].
alkaline region the true adsorption capacity was lower since pre- Temperature increase results in changes related to both kinet-
cipitation also occurred. Cabrera et al. [55] also reported that the ics and equilibrium attributed to: (i) increase in the kinetic energy
highest adsorption of nickel on zeolite was obtained at pH close to 8, which facilitates the access of the metal ions to the active adsorp-
with highly acidic environment resulting in the lowest adsorption. tion sites, (ii) increase in the surface activity of the mineral resulting
Ouki and Kavannagh [50,51] found that the lowest nickel adsorp- in higher affinity or increase in the active adsorption sites, (iii)
tion on chabazite occurred at the lowest and highest pH (i.e. 3 decrease in the mass transfer resistance. The increase in tem-
and 6), while for clinoptilolite the lowest adsorption occurred at perature is accompanied with a decrease in the thickness of the
acidic environment. Da Fonseca et al. [103] found that the max- boundary layer surrounding the mineral, so that the mass trans-
imum amount of Ni2+ adsorbed on vermiculite was observed at fer resistance of the adsorbate in the boundary layer decreases,
pH 5.2–9.0. However, at pH 9 significant amount of nickel was facilitating the diffusion of metals in the adsorbent. Thus, the effec-
removed through the formation of precipitates. Gupta and Bhat- tive diffusion coefficient of ions in solid phase usually increases
tacharyya [44,47] observed a gradual increase of Ni2+ on natural and an increase in external mass transport is observed [113,114].
montmorillonite, poly(hydroxyl-zirconium) modified montmoril- As temperature is increased, the retarding specific or electro-
lonite and TBA-montmorillonite up to pH 8.0, while at higher pH static interactions become weaker and the ions become smaller
there was a much rapid rise indicating precipitation [109]. The because solvation is reduced [115]. However, at very high tem-
authors explained the variations in nickel ions adsorbed with pH peratures physical damage to the mineral can occur, reducing its
on the basis of competition between the metal and H3 O+ ions for adsorption capacity [45,116]. In most cases it is desirable to eval-
capturing the available adsorption sites on the clay surface. uate the adsorption capacity of minerals at room temperature,
Mockovčiaková et al. [106] concluded that at pH 5–6 adsorp- since at higher temperature the operational cost of the process
tion was stable for zinc uptake on natural and modified bentonite. increases.
Mishra and Patel [108] reported that at pH < 3 there was practically Variations in temperature cause changes in the thermodynamic
no removal of zinc by bentonite due to the high proton concen- parameters (i.e. G◦ , H◦ and S◦ ). These parameters can be
tration, while the highest adsorption capacity was obtained at pH calculated using the equations of Table 3 [9,98,116–118]. Small
6. Kaya and Ören [58] studied the impact of pH on the adsorption fluctuations in temperature are usually not considered critical. A
of zinc on natural and Na-modified bentonite and found that at pH positive value of the standard enthalpy change (H◦ > 0) shows
4–7 a plateau was reached with negligible difference in the adsorp- that adsorption is an endothermic process, while negative value
tion of zinc, while at higher pH precipitation enhanced the metal shows that the process is exothermic (H◦ < 0). A negative value of

Table 3
Thermodynamic parameters used to describe adsorption systems.

Equation Comment

G = H − TS Gibbs free energy (G) as a function of enthalpy (H) and entropy (S) change at constant temperature

G = − RTlnKeq
o
Correlation of standard Gibbs free energy G◦ with the equilibrium adsorption constant Keq

lnKeq = − Ho /(RT) + So /R Calculation of H◦ and S◦ from the slope and intercept of the Van’t Hoff plot

lnka = lnAa − Ea /(RT) Apparent activation energy (Arrhenius)


S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 435

the standard Gibbs free energy of adsorption (i.e. G◦ < 0) means adsorption is of the same order of magnitude observed in physical
that the process is spontaneous. A temperature rise results in a processes [119].
higher negative value of the Gibbs free energy; this indicates that The effect of temperature on Ni2+ and Zn2+ uptake by
temperature elevation also increases the spontaneity [110]. H◦ minerals and the thermodynamic parameters of the adsorp-
is a measure of the interaction level between the metal and the tion system are summarized in Table 4 [40,42,44,47,49,53,57,63,
adsorbent implying the strength of bonding. Physical adsorption is 68,96,98,100–102,104,107,120–127]. Most studies of nickel and
always an exothermic process; as temperature increases the bonds zinc adsorption show that an increase in temperature enhances
between the metals and the adsorbent become weaker. However, the adsorption process. Rajic et al. [120] studied the removal of
in some cases in literature it has been observed that although the nickel using natural zeolitic tuff and found that the increase in
process is characterized as endothermic the heat evolved during temperature from 25 to 65 ◦ C resulted in a threefold increase

Table 4
Studies on the effect of temperature on nickel and zinc uptake by zeolite, bentonite and vermiculite, thermodynamic parameters.

Metal Mineral Examined temperature Temperature of highest Thermodynamic parameters Reference


range (◦ C) adsorption (◦ C)

Ni Clinoptilolite 10–90 90 G◦ : −8.68 to 17.27 kJ mol−1 [40]


NaCl clinoptilolite
HCl and NaCl clinoptilolite
Ni Montmorillonite 15–40 40 G◦ : 4.30 kJ mol−1 (at 25 ◦ C) [42]
H◦ : 28.9 kJ mol−1
S◦ : 82.5 J mol−1 K−1
Ni Montmorillonite 30–40 30 G◦ : −57.1 kJ mol−1 (at 30 ◦ C) [44,47]
H◦ : −45.1 kJ mol−1
S◦ : −146.4 J mol−1 K−1
Ni ZrO-montmorillonite 30–40 30 G◦ : −33.3 kJ mol−1 (at 30 ◦ C)
H◦ : −37.7 kJ mol−1
S◦ : −133.8 J mol−1 K−1
Ni TBA-montmorillinite 30–40 30 G◦ : −37.7 kJ mol−1 (at 30 ◦ C)
H◦ : −43.1 kJ mol−1
S◦ : −160.5 J mol−1 K−1
Ni Calcined Bofe bentonite 20–75 20 G◦ : −1.39 kJ mol−1 (at 20 ◦ C) [49]
H◦ : −7.33 kJ mol−1
S◦ : −20.33 J mol−1 K−1
Ni Zeolite 20–35 35 – [53]
Ni GMZ Na-bentonite 30–60 60 G◦ : −17.24 kJ mol−1 (at 30 ◦ C) [68]
H◦ : 10.92 kJ mol−1
S◦ : 92.8 J mol−1 K−1
Ni Zeolite 15–35 35 – [96]
Bentonite
Vermiculite
Ni Clinoptilolite 20–60 20 G◦ : −2.89 kJ mol−1 (at 20 ◦ C) [98]
H◦ : −4.32 kJ mol−1
S◦ : 49 J mol−1 K−1
Ni Bentonite 20–60 60 G◦ : −23.85 kJ mol−1 (at 20 ◦ C) [100]
H◦ : 14.21 kJ mol−1
S◦ : 81.46 J mol−1 K−1
Ni Na-bentonite 25–55 55 – [101]
Ni Na-montmorillonite 20–67 67 G◦ : −17.3 kJ mol−1 (at 20 ◦ C) [102]
H◦ : 28.1 kJ mol−1
S◦ : 148.1 J mol−1 K−1
Ni Zeolite tuff 25–65 65 G◦ : −9.7 kJ mol−1 (at 25 ◦ C) [120]
H◦ : 37.9 kJ mol−1
S◦ : 160 J mol−1 K−1
Ni Zeolite 25–100 100 – [121]
Ni Bentonite 25–50 50 G◦ : −7.34 kJ mol−1 (at 25 ◦ C) [122]
H◦ : 2.58 kJ mol−1
S◦ : 32.7 J mol−1 K−1
Ni Bentonite 20–60 60 G◦ : −17.3 kJ mol−1 (at 20 ◦ C) [123]
H◦ : 13.9 kJ mol−1
S◦ : 110 J mol−1 K−1
Zn Vermiculite 15–45 45 – [57]
Zn Bentonite 30–65 30 G◦ : −2.36 kJ mol−1 (at 30 ◦ C) [63]
H◦ : 12.31 kJ mol−1 (at 30 ◦ C)
S◦ : 32.85 J mol−1 K−1 (at 30 ◦ C)
Zn Zeolite 25–45 45 – [104]
Zn Na-dodecylsulphate 25–55 55 G◦ : −9.17 kJ mol−1 (at 25 ◦ C) [107]
montmorillonite H◦ : 7.39 kJ mol−1
S◦ : 6.39 J mol−1 K−1
Zn Zeolite A 25–60 60 – [124]
Zn Na-clinoptilolite 25–65 65 G◦ : −14.1 kJ mol−1 (at 25 ◦ C) [125]
H◦ : 22.4 kJ mol−1
S◦ : 123 J mol−1 K−1
Zn Bentonite 20–50 20 – [126]
Zn Bentonite 20–80 20 – [127]
436 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

of adsorption, showing that the process was endothermic and suggested that this occurred due to the increase of active sites
spontaneous. Enhanced removal of nickel at elevated tempera- and due to the change in pore size and the enhanced rate of intra-
tures has also been documented in several studies involving nickel particle diffusion. Similar findings were reported by Minceva et al.
uptake by natural minerals [42,53,68,96,100–102,122]. On the con- [104] for adsorption on zeolite, by Lin and Juang [107] for sorption
trary, Gupta and Bhattacharyya [44] found that the temperature on Na-dodecylsulphate montmorillonite and by Wu et al. [57] for
increase from 30 to 40 ◦ C resulted in a decrease of nickel adsorp- vermiculite. On the contrary, Bereket et al. [126], Sen and Gomez
tion on montmorillonite indicating an exothermic reaction. The [63] and Mellah and Chegrouche [127] found that the temperature
authors attributed this behaviour to the increased solubility of decrease enhanced the adsorption of zinc on minerals.
the metal ions in the solution which favoured the migration of
metal ions from the solid to the liquid phase. Similar results 4.5. Adsorbent grain size
have been obtained by Argun [98] for the removal of nickel from
clinoptilolite. The effect of mineral particle size on nickel and zinc removal
The effect of temperature has also been examined for the uptake has been investigated by several researchers. The most promi-
of zinc from natural minerals. El-Kamash et al. [124] reported that nent results are summarized in Table 5 [27,52,53,60,64,65,
the increase of temperature resulted in an increase of the adsorp- 83,97,100,128,129]. The specific area of the mineral depends on
tion capacity on synthetic zeolite A towards zinc. The authors particle size, pore size distribution and surface roughness. Particles

Table 5
Studies on the effect of grain size on nickel and zinc uptake by zeolite, bentonite and vermiculite.

Metal Mineral Examined grain size range Grain size of highest adsorption (␮m) Reference
(␮m)

Ni Vermiculite <90 <90 (small effect) [27]


90–300
<300
Ni Clinoptilolite 1400–2000 120–350 [52]
500–710
120–350
Ni Zeolite 355–710 45–90 [53]
180–355
90–180
45–90
Ni Jordanian zeolite tuff (phillipsite) 355–710 45–90 [53]
180–355
90–180
45–90
Ni Natural vermiculite 1000–2000 <100 [83]
600–1000
200–600
<100
Ni Natural bentonite 53–300 53 [100]
Ni Scolecite >250 No effect [128]
177–250
149–177
62–149
<62
Zn Vermiculite <90 <90 [27]
30–90
<300
Zn Bigadic zeolite Gordes zeolite 212–300 53–106 [60]
106–212
53–106
Zn Zeolite 250–300 90–125 [64]
150–250
125–200
90–125
Zn Turkish zeolite 1000–2000 300–420 [65]
840–1000
590–840
500–590
420–500
300–420
Zn Vermiculite 1400–2000 <180 [97]
500–1400
180–500
<180
Zn Scolecite >250 <149 (small impact) [128]
177–250
149–177
62–149
<62
Zn Zeolite 2500–5000 <300 (small impact) [129]
1000–2500
300–1000
<300
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 437

with smaller diameter are characterized by larger specific surface adsorbent mass, the increase in the adsorbent amount in a given
area, and thus adsorption is enhanced and shorter time is required liquid volume reduces the number of available sites per unit mass
to reach equilibrium. Each adsorption site is characterized by its of adsorbent, as the effective surface area is likely to decrease [47].
‘availability’ and ‘accessibility’ with respect to the incoming ion. Moreover, the addition of mineral affects the solution pH, which
Availability is related to equilibrium characteristics and accessi- usually increases with increasing mineral concentration [137]. This
bility to kinetics. The availability of active sites remains almost may be attributed to the increase of the amount of negatively
constant between larger and smaller particles, since the concentra- charged sites which can adsorb more H+ resulting in an increase
tion of the available sites is constant. However, the accessibility of in the pH of the final solution.
metal ions to available sites is facilitated when the adsorbent con- Table 6 [40,42,44,47–49,53–55,58,60,61,63,65,68,96,98,100–
sists of smaller particles, since the diffusion paths of ions become 102,105,108,123,127,138] shows several studies reporting the
shorter [130]. Thus, the reduction in particle size is expected to impact of mineral concentration on nickel and zinc removal.
favour kinetics. Most studies report that higher mineral concentrations resulted in
Although the grain size can affect the adsorption capacity, usu- increased metal removal [40,44,47,49,55,61,68,105,108,138] and
ally the differences are not critical. Most studies showed that the reduced amount of metal adsorbed per unit mass of adsorbent
smaller grain size favoured the adsorption capacity due to the [44,47,49,53,58,60,63,138]. Two reasons can explain this decrease
mineral’s larger specific surface area [52,53,60,97,100,129]. Some in the adsorption capacity: the first is associated with the unsatu-
studies reported that the grain size did not impact, or had a small rated adsorption sites and the second one with particle aggregation
impact on the adsorption capacity [27,53,128]. In porous materials, resulting in a decrease of the total surface area and an increase of
the internal surface area is much more critical than the external one. the diffusion path length [139]. These interactions between min-
Therefore, changes in the external surface imposed by the grain size eral particles become important when the adsorbent mass in the
will have a limited effect on the mineral’s adsorption capacity [60]. solution is high and may physically block some adsorption sites
In terms of kinetics, it has been reported that the impact of grain size from the incoming metal ions [63]. Kaya and Ören [58] found that
is more important during the early stages of adsorption [129,131]; the increase in mineral concentration resulted in a decrease of
this can be related to the fact that external diffusion is usually criti- zinc uptake. The authors mentioned that lower mineral concen-
cal at the initial stages of adsorption. In the adsorption processes in trations resulted in enhanced adsorption due to the increase in
which film diffusion is dominant, the grain size seems to influence the solid/liquid interface area, resulting from the dilution of the
more the adsorption process, while in the cases where intraparticle suspension [140]. Wu et al. [57] stated that for given initial metal
diffusion is dominant the impact of the mineral grain size attenu- concentration the equilibrium adsorption capacity decreases as the
ates. Inglezakis and Poulopoulos [130] reported that the conduction mineral concentration in the solution increases until it approaches
of experiments with adsorbents of different diameter reveals the a critical value at which all ions have been adsorbed on the mineral.
control stage of adsorption. The rate is proportional to the particle
radius when film diffusion is the control stage and inversely propor- 4.7. Adsorbent type
tional to the square root of the particle radius, when intraparticle
diffusion controls the process. The adsorbent type has a profound effect on adsorption.
Table 7 [26,40,47,56,58,60,81–83,92,96,97,105,106,123,141–155]
4.6. Adsorbent concentration summarises the studies comparing the performance of different
types of natural and modified minerals for nickel and zinc uptake.
The mineral mass in the solution also affects the adsorption pro- In most cases the chemical and/or thermal treatment of the mineral
cess, since it determines the availability of active sites. The increase resulted in an increase of its adsorption capacity, while in certain
of the adsorbent concentration results in more sites being avail- cases the opposite was observed. Çoruh and Ergun [40] reported
able in the same solution volume. At low mineral concentrations that chemical conditioning of zeolite with HCl and NaCl increased
high amount of sorbate is readily available to be captured by few its adsorption capacity. Na-rich forms of clinoptilolite are known to
available sites. An increase in the solid concentration increases have an enhanced exchange capacity, since part of the more tightly
the surface area of the adsorbent, which in turn increases the bound K+ and Ca2+ cations are exchanged with strong Na+ solutions.
number of binding sites for the same liquid volume and thus the Conditioning with HCl and NaCl solutions can remove fine dust par-
total amount of metal that is removed increases [132]. However, ticles from the surface of clinoptilolite crystals making the channel
the amount of metal adsorbed per unit mineral mass decreases. apertures more easily accessible to metal ions [40,128,156,157].
At high mineral concentrations the available metal concentration Álvarez-Ayuso et al. [81] found that the adsorption of synthetic
is insufficient to cover completely the exchangeable sites on the zeolite NaP1 was 10 times higher than that of natural zeolite for
adsorbent, usually resulting in low metal uptake [133]. Moreover, nickel and zinc. Moreover, Álvarez-Ayuso and García-Sánchez [82]
the interference between binding sites due to increased adsorbent concluded that Ca-bentonite exhibited lower adsorption for Ni and
concentration can result in a low specific uptake [134]. The inter- Zn uptake than Na-bentonite. Esposito et al. [93] concluded that
actions between mineral particles become more important when coal fly ash zeolite 4A and commercial zeolite 4A exhibited sim-
the mass of the adsorbent in the liquid phase is higher and may ilar performance for zinc and nickel removal. Vieira dos Santos
cause physical blockage of some adsorption sites, decreasing the and Masini [66] showed that in multi-metal solutions (contain-
adsorption efficiency. These interactions can create electrostatic ing zinc and nickel) the performance of chabazite was superior
interferences such that the electrical surface charges on the closely to that of clinoptilolite, while the exposure of zeolite to concen-
packed particles diminish attractions between the adsorbed solutes trated NaCl solutions converted them to a homoionic state in the
and surfaces of individual grains [63,135]. In parallel, the interac- Na-form which improved their exchange capacity. Blais et al. [146]
tions between particles may cause desorption of metal ions, when found that the alkaline treatment of vermiculite with 0.75 M NaOH
their adsorption on mineral is reversible. At low mineral concen- increased its performance, while acid treatment with 0.75 M H2 SO4
trations, metal ions in the solution would not only be adsorbed resulted in inferior performance. The mineral’s adsorption capac-
on the surface of the sorbent, but would also enter into its inner ity was significantly affected by the pH variation resulting from
parts, thus facilitating the concentration gradient of metal ions alkaline and acid treatment. Bhattacharyya and Gupta [142,143]
[136]. Hence, although the number of adsorption sites per unit reported that acid activation of montmorillonite resulted in a slight
mass of an adsorbent should remain constant, independent of the increase of its adsorption capacity towards nickel that could be
438 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

Table 6
Studies on the effect of adsorbent concentration on nickel and zinc uptake by zeolite, bentonite and vermiculite.

Metal Mineral Mineral concentration (g L−1 ) Concentration of highest Concentration of highest Reference
adsorption (g L−1 ) removal efficiency (g L−1 )

Ni Clinoptilolite 2.4–40 30–40 [40]


NaCl clinoptilolite
HCl and NaCl clinoptilolite
Ni Montmorillonite 2–30 30 [42]
Ni Montmorillonite 2–6 2 6 [44,47]
ZrO-montmorillonite
TBA-montmorillinite
Ni Natural zeolite 2–22 22 [48]
Ni Calcined Bofe bentonite 5–40 5 40 [49]
Ni Jordanian zeolite tuff 1–8 1 [53]
(phillipsite)
Ni Clinoptilolite 25–100 2 [54]
Ni Zeolite 5–100 100 [55]
Ni GMZ Na-bentonite −1.2 1–1.2 [68]
Ni Zeolite 5–15 15 [96]
Bentonite
Vermiculite
Ni Clinoptilolite 0.5–30 15–30 [98]
Ni Natural bentonite 10–100 100 [100]
Ni Na-bentonite 1–6 6 [101]
Ni Na-montmorillonite −0.9 0.8–0.9 [102]
Ni Bentonite 0.1–5 0.1 5 [123]
Ni Bentonite 1–8 8 [127]
Zn Clinoptilolite 25–100 2 [54]
Zn Zeolite 5–100 100 [55]
Zn Natural bentonite 1–10 1 1 [58]
Na-bentonite 1–2 1
Zn Gordes zeolite 1–10 1 1 [60]
Bigadic zeolite
Zn Cankiri bentonite 1–9 4–9 [61]
Zn Bentonite 0.25–0.75 0.25 [63]
Zn Turkish zeolite 5–50 25–50 [65]
Zn Zeolite 4A 5–30 25–30 [105]
Zeolite 13X
Bentonite
Zn Bentonite 5, 10, 20 20 [108]
Zn Vermiculite 50–150 50 150 [138]

attributed to the increase in the mineral surface area and pore [96,97] found that the mineral order for nickel removal was ver-
volume. Cayllahua [89] found that zeolite modification with NaCl miculite > bentonite > zeolite, while for zinc removal the order was
and CH3 COONa increased nickel adsorption by 25–30% compared bentonite > zeolite > vermiculite, showing the high affinity of ver-
to natural zeolite. Stefanović et al. [149] found that chemical condi- miculite towards nickel and of bentonite towards zinc. Zamzow
tioning of clinoptilolite tuff with NaCl and CaCl2 resulted in higher and Murphy [159] found that the adsorption capacity of different
zinc uptake. Olu-Owolabi and Unuabonah [154] reported that ben- types of zeolite for zinc uptake in wastewater exhibited signifi-
tonite modification with sulphate and phosphate ions increased cant variability as it ranged from 5.6 to 44.1 mg g−1 . Pitcher et al.
the adsorption capacity for zinc uptake. Athanasiadis and Helm- [84] concluded that synthetic zeolite was more effective (>96%)
reich [148] reported that the NaCl clinoptilolite exhibited up to for the removal of Zn2+ from both multi-metal synthetic solutions
100% higher sorption capacity for zinc compared to that of natural and stormwater compared to natural zeolite (10–42%). The higher
clinoptilolite. Cincotti et al. [147,152] and Cerjan-Stefanović et al. performance of synthetic zeolite was attributed to its higher alu-
[158] found that Na-modified clinoptilolite performed better than minium content, its purity, its exchangeable and homoionic form,
natural clinoptilolite with respect to zinc removal. Kaya and Ören its lower grain size and its larger internal area [84,131].
[58] also concluded that Na-modified bentonite performed better
than natural bentonite for zinc removal. Purna Chandra Rao et al. 4.8. Competitive adsorption
[105] found that synthetic zeolite 4A and 13X had a higher affinity
towards zinc uptake compared to bentonite. The presence of other cations apart from the one that must be
Some studies report that mineral modification adversely removed from wastewater has a negative influence on the adsorp-
impacted on its performance. Gupta and Bhattacharyya [47] tion process [136]. The adsorbent’s performance depends on the
found that natural montmorillonite performed better than ZrO- type and concentration of other cations that co-exist in the liquid
montmorillonite and TBA-montmorillonite as the introduction of phase and more importantly with the selectivity of the adsorbent
ZrO and TBA groups in clay blocked some of the available adsorption for cations compared to its selectivity towards the heavy metal
sites. Similarly, Mockovčiaková et al. [106] found that natural ben- under investigation. The presence of other cations in the solution
tonite performed better than synthetic bentonite with respect to can reduce the adsorption capacity compared to that obtained for
zinc removal. Abollino et al. [26] compared the adsorption capacity the uptake of each metal individually in single metal systems, par-
of Na-montmorillonite and vermiculite for nickel and zinc removal. ticularly when the mineral is more selective for the competing
The authors concluded that montmorillonite performed better than cations rather than the heavy metal. Malamis et al. [160] stud-
vermiculite at low pH in the presence of effluents having high ionic ied the competitive adsorption of Ni and Zn on zeolite, bentonite
strength, while vermiculite was more suitable for the removal of and vermiculite in multi-metal solutions containing Ni, Zn, Cu and
heavy metals from the examined effluents at high pH. Katsou et al. Pb. Katsou et al. [96,97] studied the adsorption of Ni and Zn on
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 439

Table 7
Impact of adsorbent type on nickel and zinc adsorption.

Metal Sorbent Adsorption Reference


capacitya
(mg g−1 )

Ni Vermiculite 25.33 [26]


Na-montmorillonite 3.64
Ni Natural zeolite 8.69 [40]
Modified zeolite (HCl + NaCl) 10.75
Modified zeolite (NaCl) 10.46
Ni Montmorillonite 21.14 [47]
ZrO-montmorillonite 10.45
TBA-montmorillonite 8.50
Ni Na-montmorillonite 10.04 [56]
Acid modified montmorillonite 7.31
Ni Zeolite 1.98 [81]
Ni Synthetic (NaP1) zeolite 20.1
Ni Na-bentonite 24.2 [82]
Ni Ca-bentonite 6.3
Ni Natural vermiculite 19.3 [83]
Exfoliated vermiculite 5.91
Ni Commercial zeolite 4A 7.90 [92]
Ni Coal fly ash prepared zeolite 4A 8.96
Ni Vermiculite 25.51 [96]
Bentonite 19.92
Zeolite 16.86
Ni Natural bentonite 50 [123]
Washed bentonite 20
Calcined bentonite 22
Ni Na-zeolite 1.30 [141]
Gördes (natural) zeolite 1.44
Bigadic (natural) zeolite 2.74
Bigadic (natural) zeolite 1.26
Ni Acid-activated montmorillonite 29.5 [142]
Montmorillonite 28.4
Ni Acid-activated montmorillonite 21.3 [143]
Montmorillonite 21.1
Ni TBA-montmorillonite 19.7 [144,145]
Montmorillonite 28.4
Ni Vermiculite 0.229 [146]
Ni Acid (H2 SO4 ) treated vermiculite 0.139
Ni Alkaline (NaOH) treated vermiculite 0.643
Zn Vermiculite 23.40 [26]
Na-montmorillonite 3.61
Zn Na-modified bentonite 54 [58]
Natural bentonite 24
Zn Gordes zeolite 6 [60]
Bigadic zeolite 3
Zn Zeolite 3.45 [81]
Zn Synthetic (NaP1) zeolite 32.6
Zn Na-bentonite 23.1 [82]
Zn Ca-bentonite 5.8
Zn Commercial zeolite 4A 31.58 [92]
Zn Coal fly ash prepared zeolite 4A 30.80
Zn Vermiculite 31.65 [97]
Bentonite 39.5
Zeolite 18.66
Zn Zeolite 4A 42.82 [105]
Zeolite 13X 38.31
Bentonite 35.17
Zn Natural bentonite 36.63 [106]
Composite E20 bentonite 29.67
Composite A85 bentonite 14.10
Zn Vermiculite 0.266 [146]
Acid (H2 SO4 ) treated vermiculite 0.137
Alkaline (NaOH) treated vermiculite 0.737
Zn Sardinian clinoptilolite 3.27 [147]
Na-modified zeolite 14.06
Zn Clinoptilolite 4.40 [148]
NaCl-modified clinoptilolite 8.00
Zn Zeolite 17.00 [149]
Na-modified zeolite 49.69
Zn Natural clinoptilolite 17.50 [150]
Na-clinoptilolite 14.20
NH4 -clinoptilolite 19.70
Zeolon P4A clinoptilolite 154.97
Na-zeolon 4A clinoptilolite 134.20
NH4 -zeolon P4A clinoptilolite 111.90
440 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

Table 7 (Continued)

Metal Sorbent Adsorption Reference


capacitya
(mg g−1 )

Zn Clinoptilolite 3.16 [151]


Philipsite 1.70
Chabazite 7.98
Analcime 3.67
Bentonite 3.08
Zn Clinoptilolite 4.58 [152]
Na-clinoptilolite 8.17
Zn Heated (110 ◦ C) bentonite 3.32 [153]
Heated (200 ◦ C) bentonite 4.54
Heated (200 ◦ C) bentonite 2.88
0.5 mol L−1 H2 SO4 bentonite 2.29
2.5 mol L−1 H2 SO4 bentonite
Zn Bentonite 16.36–24.48 [154]
Sulphate-modified bentonite 17.17–26.19
Phosphate-modified bentonite 17.50–26.63
Zn Montmorillonite 215.63 ± 11.12 [155]
HyA-montmorillonite 395.88 ± 25.52
HAS-montmorillonite 396.02 ± 10.81
a
Adsorption capacity in most cases refers to qm and in some cases to qe .

the same natural minerals in single metal solutions. In the latter mineral’s selectivity for the uptake of different heavy metals in a
case the adsorption was significantly higher due to the absence of multi-metal environment follows the order Pb > Zn > Cu > Ni, which
other ions in the solutions. However, in several cases the mineral is the same as the one observed in single metal solutions. Sprynskyy
adsorption capacity for the uptake of all metals remains relatively et al. [52] found a significant decrease of Ni adsorption on clinop-
constant. The reduction in the adsorbent’s performance depends on tilolite in multi-component solutions (Pb, Cu, Cd and Ni) compared
its selectivity and affinity for the respective ion in comparison with to single solutions which was probably caused by the propinquity
the competing ions, as well as the ionic properties and the concen- of the sorbents to the other metals in the solution. The other metals
tration of each ion. The parameters that impact on the performance in the solution were only slightly affected by the presence of com-
of an adsorbent are related to the adsorbent characteristics, the peting ions. Specifically, the selectivity order of clinoptilolite in the
properties of the metal ions and the liquid medium characteristics presence of other metals has the following order Pb > Cu > Cd > Ni.
[93,161]. The preference of Na-bentonite for Cu can explain the significant
The metal ions having smaller hydrated ionic radius have easier decrease of Ni adsorption which is observed when both metals (i.e.
access to the mineral’s surface and can diffuse more easily inside Cu and Ni) co-exist in the solution [101]. Futalan et al. [164] drew
its pores. On the other hand, the presence of ions having a larger similar conclusions since they mention that the adsorption of Ni on
hydrated ionic radius results in a more rapid saturation of the bentonite was impeded due to the presence of Pb and Cu in the solu-
adsorption sites. Furthermore, electronegativity and ionic poten- tion. Mishra and Patel [108] stated that the capacity of bentonite
tial provides an indication of the power of adherence of the metal for Zn uptake was lower when Pb was present in the solution com-
ions to the solid. The metals exhibiting the highest electronega- pared to single metal solutions, while the adsorption capacity of
tivety exhibit the greatest uptake by the adsorbent. Furthermore, bentonite for Pb was the same in single and binary (Pb–Zn) metal
the metals having the highest ionic potential are strongly retained solutions.
by the adsorption sites exhibiting high negatively charged density. The temperature significantly influences the performance of
Nevertheless, their interaction with certain functional groups of the competitive adsorption. Donat et al. [100] found that in the pres-
adsorbent can be weaker than that of other metals having lower ence of other metals (i.e. Pb, Cd and Cu) nickel adsorption on
ionic potential [162]. The adsorption of metals exhibiting higher bentonite was lower. The temperature increase contributed to a
affinity to the minerals is not impeded much in a competitive envi- reduction of adsorption inhibition caused by the presence of both
ronment of metal ions. The presence of competing metals impacts Ni and Cu in the solution. Lead adsorption was found to be more
to a much greater level in the cases where the metals are retained sensitive in the presence of Ni and Cu compared to Cd, while the
by the minerals at the active adsorption sites with weak bonds. For adsorption order of metals on bentonite from multi-metal solutions
example, the selectivity of zeolite 4A for the uptake of metals from was Pb > Cu > Ni > Cd.
a multi-metal solution followed the order: Cu > Cr3+ > Zn > Co > Ni. The initial concentration of competing cations in the solution
The increase in the concentration of the metal ions in the solu- influences the adsorption of heavy metals on minerals. The pres-
tion resulted in a significant reduction of the adsorption of Zn, Co ence of a metal having higher initial concentration in the solution
and Ni. The lower zeolite selectivity for these metals compared to is usually accompanied by higher adsorption due to its competitive
Cu and Cr3+ resulted in reduced adsorption capacity. The afore- advantage compared to the other metals that co-exist in the solu-
mentioned selectivity order cannot be attributed to the hydration tion. Ibrahim et al. [165] attribute the selectivity order of zeolite
energy of ions, neither to the free hydration energy. The metals (Ni > Cr3+ > Cu > Fe > Zn) for the uptake of different metals to a large
having the highest free energy of hydration should prefer the liq- extent to the initial metal concentration contained in galvaniz-
uid rather than the solid phase. Researchers attribute the higher ing wastewater (Ni > Cu > Cr3+ > Fe > Zn). However, the initial metal
removal of Cu and Cr3+ to the higher precipitation of metal hydrox- concentration is only one of the several parameters that can have an
ides on the mineral surface and pores [93]. A similar case has been influence on competitive adsorption. Other properties such as the
recorded by Panayotova and Velikov [163]. The higher selectivity ionic size, valence, hydrated energy, ionic potential are parameters
of clinoptilolite for lead contributes to a significant reduction of that can determine the selectivity order of minerals for metals. Zeo-
adsorption of Ni, Cu and Zn when these metals co-exist, with the lites having high Si/Al ratio have a preference towards monovalent
lower decrease being observed for the adsorption of Zn [54]. The metal cations having low charge density. Among divalent metal
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 441

cations, the mineral’s selectivity depends on hydration energy, NH4 + and Cs+ when monovalent cations are exchanged and (ii) ions
with the mineral being more selective towards the cations hav- with low hydration energy such as Pb2+ when divalent cations are
ing lower hydration energy. The mineral has higher selectivity for exchanged with monovalent ones.
the metals that have larger valence (electrostatic interaction), fact The selectivity order of metal cations in clinoptilolite with
that explains the greater adsorption of Cr3+ compared to Cu2+ . Si/Al ratio of 4.2 is given by Caputo and Pepe [169] as:
Among two metals having the same valence, selectivity usually Cs+ > NH4 + > Pb2+ > Na+ > Sr2+ > Cd2+ > Zn2+ ≈ Cu2+ . This order con-
increases with a reduction in the hydrated ionic radius. Wu et al. firms the high selectivity of clinoptilolite for monovalent cations
[138] concluded that the addition of Cd and H+ in the solution con- with low charge density. Furthermore, the high mineral selec-
tributes to the reduction of Zn adsorption on vermiculite, since tivity for Pb2+ is attributed to the low hydration energy of the
the increase of the concentration of competing cations affects cation. Langella et al. [69] have found similar results as the clinop-
more the adsorption process. Similar results were observed for tilolite selectivity order is: NH4 + > Pb2+ > Na+ > Cd2+ > Cu2+ ≈ Zn2+ .
the study of Zn adsorption on vermiculite in Zn–Cd–K solution Consequently, clinoptilolite selectivity is mainly determined by
[43]. Covelo et al. [166] investigated the competitive adsorption of the free energy of hydration of the cations. Tsitsishvili et al.
heavy metals on vermiculite and found that the selectivity order [170] found the following selectivity order for natural zeo-
was: Cu > Zn > Pb > Cd > Cr > Ni. The adsorption of Cu is stronger lites: Ba2+ > Pb2+ > Cd2+ > Zn2+ > Cu2+ . According to Blanchard et al.
compared to the other metals and as a result it is preferred by [171], the selectivity of Na-clinoptilolite follows the order
vermiculite in the presence of competing cations. Abollino et al. Pb2+ > NH4 + > Cu2+ , Cd2+ > Zn2+ , Co2+ > Ni2+ > Hg2+ . Similar results
[74] mention that the higher charge density results in greater were obtained by Ouki and Kavannagh [50,51], since the perfor-
electrostatic attraction of the metals ions on Na-vermiculite. The mance of clinoptilolite for the removal of heavy metals from multi-
mineral selectivity order was Cr3+ > Ni > Mn > Zn > Cu > Cd = Pb. In metal solutions reduces in the order: Pb > Cu > Cd > Zn > Cr > Co > Ni.
this case the metals having the larger ionic radius and the smaller Wang and Peng [18] summarize the selectivity order of clinoptilo-
charge density have unfavourable adsorption characteristics. At lites having different origin for the competitive adsorption of heavy
the same time these metals are influenced more by the proto- metals and confirm the high mineral selectivity for lead. From the
nation of the surface groups of the adsorbent at low pH and above discussion, it can be concluded that the hydration energy, the
thus their uptake by minerals becomes more difficult. Ouki and charge density (ratio of charge/ionic radius) and the dimensions of
Kavannagh [50,51] investigated the performance of zeolite and the hydrated ions provide an indication of the mineral’s preference
chabazite for the uptake of Ni and Zn from multi-metal aqueous for the different metals during competitive adsorption. However,
solutions and found that the clinoptilolite selectivity sequence was the mineral’s selectivity also depends on other parameters such as
Pb > Cu > Cd > Zn > Cr > Co > Ni, while the corresponding selectivity the geometry and/or the orientation of the ion. For example, clinop-
for chabazite was Pb > Cd > Zn > Co > Cu > Ni > Cr. tilolite is very selective for the ammonium ion, despite its high
The presence of alkali metals and alkaline earth metals in the hydration energy (−1329 kcal mol−1 ). This happens because NH4 +
solution can influence the adsorption process; the increase of the can easily penetrate through the clinoptilolite’s channels by suit-
alkali metal concentration usually limits the adsorption process. ably transforming its orientation [172]. Wingenfelder et al. [129]
Yang et al. [68] found that the simultaneous presence of Na+ , attribute the higher hindrance of Zn and Cd adsorption on zeo-
Ca2+ and K+ in the solution reduces the adsorption of Ni on GMZ- lite and the low hindrance of Pb adsorption that occurs with the
bentonite. Calcium was found to have greater influence in the increase of Ca2+ concentration in the solution to the low Si/Al ratio
adsorption of Ni on bentonite compared to the other cations. El- and the low charge density of zeolite. Lead is characterized by the
Bayaa et al. [167] found that the presence of Na+ in the solution lowest hydration energy and as a result its adsorption on zeolite
has a smaller impact on the adsorption of Cr3+ and Cu2+ than is favoured. At the same time its presence in the solution reduces
that of NH4 + and K+ . The latter is attributed to the lower mineral the adsorption of Cd and Zn, while the presence of other cations
selectivity towards Na+ . Sodium is characterized by a large hydra- (i.e. Ca2+ ) has a small impact on lead adsorption. Panayotova and
tion radius and forms outer-sphere complexes with the mineral. Velikov [173] compared the removal of Ni and Zn obtained by a
The adsorption of Ni and Cd on zeolite and bentonite decreases natural Bulgarian zeolitic rock in single and in multi (Ni, Zn, Pb,
significantly due to the presence of Mg2+ , Ca2+ and Fe3+ [105]. Cu, Cd) metal solutions and concluded that there was a significant
The influence of competing cations is greater for zeolite (4A and decrease in metal uptake in multi-metal solutions. Moreover, the
13X) compared to bentonite, while the presence of Mg results in authors reported that the presence of Ca2+ in single and multi-
greater reduction of Zn and Cd uptake by minerals compared to Ca metal solutions decreased metal uptake by zeolite in the order
and Fe. Zn ≈ Cd > Ni > Cu; the presence of Mg2+ in multi-metal solutions did
According to the Eisenmann–Sherry model, zeolite selectivity not significantly impact on metal removal, while the availability
for the various cations is a function of their free hydration energy of Mg2+ in single heavy metal solutions resulted in a significant
and of the electrostatic interactions among the free ions and those decrease in Ni and Zn uptake by zeolite. In the study of Panay-
located in the zeolite structure [168]. The preference of a mineral otova and Velikov [163] the presence of Mg2+ in wastewater slightly
for a metal ion as opposed to another one depends upon whether decreased the uptake of Ni and Zn by NaCl-zeolite from multi-metal
the electrostatic (Coulombic) energy of the ionic interaction with solutions (Pb, Cd, Cu, Zn, Ni), while a significant decrease in the
the anionic grid prevails over its difference in the free energy of adsorption of both metals by the modified zeolite was observed
hydration. In zeolites having low Si/Al ratio which are character- due to the presence of Ca2+ that is characterized by lower hydration
ized by a high charge grid and thus high capacity, the selectivity energy compared to the two heavy metals.
reduces with the increase of the ionic radius and cations having a Most documented studies show that nickel and zinc adsorption
higher charge density are preferred when monovalent cations are on minerals decreases in the presence of other cations. On the
exchanged. When monovalent cations are exchanged with diva- other hand, certain researchers mention that the presence of alkali
lent ones, the zeolite having low Si/Al ratio usually prefers divalent metals and alkaline earths does not significantly impact on the
cations. In zeolites having high Si/Al ratio which are characterized adsorption of heavy metals [165]. In some works, researchers
by low charge grid and thus lower capacity, their selectivity reduces have found that although the individual adsorption capacity of
with the increase (in absolute value) of the free energy of hydra- the mineral for each metal decreases in multi-metal solutions
tion. In the clinoptilolite grid a low anionic field is created which compared to single metal solutions, the total metal uptake by the
enhances its selectivity for (i) ions with low charge density such as mineral is actually higher. For example, Jha et al. [174] found that
442 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

the adsorption of Ni and Cu on composite activated carbon-zeolite mineral uptake of heavy metals. The impact of ionic strength can be
decreases when the two metals co-exist in the same solution, attributed to the competition among cations, the metal’s activity,
while this decrease becomes more significant when Cd is added the interface potential, the solution pH and to the properties of the
in the solution. Nevertheless, the total adsorption capacity for electric double layer. When the adsorbent’s surface is in contact
all metals is larger in multi-component solutions, showing that with metals, the cations that are adsorbed are surrounded by an
the adsorption of a metal from single metal solutions leaves electric double layer due to electrostatic interactions. These inter-
certain adsorption sites empty. This happens because certain actions decrease when the ionic strength increases. This decrease
adsorption sites have the ability to attract specific metals. Table 8 is attributed to the weaker repulsive forces of the electric double
[27,50,51,54,57,92,96,97,101,108,128,138,146,161,166,174–176] layer since the latter is compressed at high ionic strength, while
presents the adsorption capacities of minerals for nickel and zinc attractive van der Waal forces remain independent of the solution’s
ions in multi-component solutions; in most cases a comparison is ionic strength. The surface potential decreases with the compres-
made with the respective ones obtained in single metal solutions. sion of the double layer, while the surface charge density remains
constant. The compression of the double layer may favour the
4.9. Ionic strength uptake of metal ions by the adsorbent. Thus, in some cases high
adsorption capacity is observed with high ionic strength [177,178].
Wastewater contains various cations (e.g. Na+ , K+ , Ca2+ , Mg2+ , However, the competition between the electrolyte cations (e.g. Na+ ,
NH4 + ) and anions (e.g. Cl− , SO4 2− , PO4 3− , HCO3 − ) that may limit the K+ ) and the heavy metals for the available adsorption sites can

Table 8
Competitive adsorption of nickel and zinc on zeolite, bentonite and vermiculite.

Solution Adsorbent Ni2+ adsorption Zn2+ adsorption Reference


(mg g−1 ) (mg g−1 )

Multi (Ni, Zn, Cd, Cu, Mn, Pb) Vermiculite 25.33 23.40 [27]
Multi (Pb, Cd, Cu, Zn, Cr, Ni, Co) Clinoptilolite 0.9 2.7 [50,51]
Multi (Pb, Cd, Cu, Zn, Cr, Ni, Co) Chabazite 4.5 5.5 [50,51]
Single Zn Clinoptilolite – 8.21 [54]
Single Ni 5.08 –
Multi (Pb, Zn, Cu, Ni) Clinoptilolite 0.5 3.53 [54]
Single Zn Vermiculite 1.06 [57]
Ternary (Zn, Cd, K) 0.70
Multi (Co, Cr, Cu, Zn, Ni) Zeolite 4A 7.9 [92]
Multi (Co, Cr, Cu, Zn, Ni) 31.6
Multi (Co, Cr, Cu, Zn, Ni)
Multi (Co, Cr, Cu, Zn, Ni) Coal fly ash prepared 8.96 30.80 [92]
Multi (Co, Cr, Cu, Zn, Ni) zeolite 4A
Multi (Co, Cr, Cu, Zn, Ni)
Single Ni and Zn Vermiculite 15.76 22.76 [96,97]
Zeolite 5.52 12.9
Bentonite 8.93 24.20
Single Ni Na-bentonite 13.97 [101]
Binary (Ni, Cu) 10.91
Single Zn Bentonite – 9.12 [108]
Binary (Pb–Zn) 7.56
Binary (Pb–Zn) 5.80
Multi (Cu, Zn, Ni, Co, Pb, Cd) Scolecite 0.9 2.1 [128]
Single Zn Vermiculite 0.85 [138]
Binary (Zn, Cd) 0.10
Binary (Zn, H) 0.10
Multi (Al, Ca, Cd, Cr, Cu, Fe, Mg, Mn, Ni, Vermiculite 0.23 [146]
Pb, Zn) 0.27
Multi (Al, Ca, Cd, Cr, Cu, Fe, Mg, Mn, Ni, Acid (H2 SO4 ) treated 0.139 [146]
Pb, Zn) vermiculite 0.137
Multi (Al, Ca, Cd, Cr, Cu, Fe, Mg, Mn, Ni, Alkaline (NaOH) 0.643 0.737 [146]
Pb, Zn) treated vermiculite
Multi (Ni, Cu, Pb, Zn) Vermiculite 9.54 7.14 [161]
Zeolite 3.71 4.11
Bentonite 7.44 8.30
Multi (Cd, Cr, Cu, Ni, Pb, Zn) Vermiculite 0.32 2. 64 [166]
Single Ni Activated 70.43 [174]
Binary (Ni, Cu) carbon–zeolite 20.54
Ternary (Ni, Cu, Cd) composite (type 1) 21.72
Single Ni Activated 62.22 [174]
Binary (Ni, Cu) carbon–zeolite 22.30
Ternary (Ni, Cu, Cd) composite (type 2) 22.30
Single Zn Zeolite 3.926 [175]
Binary (Zn, Cd) 2.02
Binary (Zn, Cd) 0.881
Binary (Ni, Pb) Vermiculite 13.04 – [176]
Binary (Ni, Ag) 14.27
Binary (Ni, Cd) 9.01
Tertiary (Ni, Ag, Pb) 12.77
Tertiary (Pb, Ni, Cd) 8.73
Tertiary (Ag, Ni, Cd) 9.16
Multi (Ag, Ni, Cd, Pb) 8.09
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 443

decrease the amount of metals that is adsorbed on the minerals column studies had little influence on adsorption, while in batch
[45,179,180]. experiments a significant reduction in adsorption performance
The increase in ionic strength can influence adsorption in two was observed when NaClO4 concentration increased from 0.03 to
ways: (i) it decreases the solution-phase activity of the metal and 0.04 M. Wu et al. [138] were able to obtain a linear positive cor-
(ii) it increases the concentration of competing ions [19,64]. The relation between the ionic strength and the amount of zinc that
decrease in the metal’s activity resulting from the elevated ionic is adsorbed on zeolite. Álvarez-Ayuso et al. [81] mention that the
strength is due to the increase in the concentration of the accompa- increase of Ca(NO3 )2 ·4H2 O concentration in the solution reduces
nying anions of the electrolyte solution. These anions create stable, the adsorption of Ni, Zn and Cd, indicating that ion exchange is the
soluble complexes with the heavy metals which are not easily dominant process. The presence of Ca(NO3 )2 ·4H2 O impacts on the
adsorbed by the mineral. Several studies have shown that at low adsorption of Ni and Zn, contributing to the decrease of their uptake
ionic strength, metal adsorption on minerals occurs both at the pla- by 55–60%, since the hydration energy of Ca2+ is significantly lower
nar sites of the adsorbent (i.e. outer-sphere complexes) as well as than that of Ni and Zn.
at the mineral edges (i.e. inner-sphere complexes). The increase
in the ionic strength results in an increase of the concentration of 4.10. Agitation speed
the cations (e.g. Na+ , K+ , Ca2+ , Mg2+ ) that compete with the heavy
metals particularly for the planar mineral sites, which in several The rate of the adsorption process may be controlled by external
cases exhibit lower affinity for heavy metals [26,167,181]. There- mass transfer or/and intraparticle diffusion. External mass trans-
fore, at high ionic strength metal adsorption takes place mainly fer is usually important at the initial stages of the process. The
through the formation of inner-sphere complexes with the func- application of appropriate agitation increases ion mobility in the
tional groups at the mineral edges which have high affinity for the solution and thus, the mass transfer resistance reduces [45,116].
heavy metals. Several salts have been employed to examine the At high agitation rates, the boundary layer becomes thinner, which
effect of ionic strength on adsorption including NaCl, NaNO3 , KCl, usually enhances the rate of solute diffusion through the boundary
MgCl2 , CaCl2 , NaClO4 [9]. Most studies show that an increase in layer [186]. Under high agitation speed, the external diffusion coef-
ionic strength decreases the mineral’s adsorption capacity [64,81], ficient increases [187]. In most cases the increase in stirring speed
while there are also some studies showing that the ionic strength is associated with an increase in the rate of adsorption, particu-
has a small impact on adsorption [64,182] or even that it has a larly during the early stages of the process. Shawabkeh et al. [188]
positive contribution [138]. used zeolite produced by Jordanian oil shell ash to remove zinc
Álvarez-Ayuso et al. [81] studied the effect of ionic strength from wastewater and found that the rate of adsorption increased
by employing 0–0.1 M Ca(NO3 )2 ·4H2 O as electrolyte at pH 6 for with increasing agitation speed. Kocaoba et al. [48] investigated the
nickel and zinc adsorption on synthetic zeolite. The authors found effect of agitation speed (5–250 rpm) on adsorption of nickel on
that the presence of calcium reduced the adsorption performance natural zeolite and found that the uptake increased up to 150 rpm
by 55–60% for both metals. Song et al. [102] investigated the and thereafter remained constant. Other researchers found that
adsorption of nickel on Na-montmorillonite in the presence of high agitation rates do not influence or have a negative impact on
0.01–0.1 M NaClO4 . The authors found that at low pH (<7) the metal adsorption. In these cases it seems that diffusion of metal
ionic strength affected the adsorption process, while at higher pH ions from the bulk solution to the mineral surface through the
it did not influence it. Yang et al. [68] investigated the adsorp- boundary layer is not the rate limiting step, or a breaking up of
tion of nickel on GMZ Na-bentonite for NaClO4 concentrations of mineral particles may have occurred. Mellah and Chegrouche [127]
0.001, 0.01 and 0.1 M and found that ionic strength influenced the examined the impact of various agitation speeds (i.e. 200, 350
adsorption process for pH < 8, while at pH > 8 the ionic strength and 500 rpm) on zinc removal by bentonite and found that dur-
was not an important factor. Respectively, the adsorption of Ni on ing the first 15 min the adsorption was higher for 200 rpm. The
Na-montmorillonite was found to be influenced by the increase authors mention that the lack of homogeneity and/or the break-
in the concentration of NaClO4 at pH < 7, while at higher pH the ing up of particles at higher agitation speeds were the most likely
ionic strength did not impact on the metal removal process. This reasons for this behaviour. On the contrary, Trgo and Perić [182]
behaviour can be associated to outer and inner-sphere complexa- stated that the potential breaking up of mineral particles enhanced
tion. At low pH the main mechanism of adsorption is that of ion the adsorption performance since freshly broken adsorption sites
exchange (formation of outer-sphere complexes), while at higher were revealed; thus greater stirring speeds increased the number
pH the formation of inner-sphere complexes prevails. The metals of available adsorption sites. Taparcevska et al. [189] investigated
that are connected with strong bonds to the available adsorption the impact of agitation speed (150–190 rpm) on Zn2+ removal from
sites are not significantly affected by the ionic strength. On the other aqueous solutions by employing zeolite and concluded that the rate
hand, the outer-sphere complexes are affected by the solution’s of zinc adsorption on zeolite was not significantly influenced by the
ionic strength [26,28]. It is postulated that when ionic strength degree of stirring. Thus, mass transfer across the boundary layer did
influences the adsorption process outer-sphere complexation is not impact the adsorption rate of the process. According to McKay
important, while in the opposite case inner-sphere complexation [190] and Meshko et al. [191] the variation in the external mass
is important [183,184]. Consequently, the impact of ionic strength transfer coefficient with agitation speed was more significant only
on adsorption can be employed to distinguish between the forma- when low (<50 rpm) stirring speeds were employed.
tion of inner and outer-sphere complexes. The electrolyte’s cations
compete with heavy metals for the formation of bonds with the 4.11. Evaluation of parameters
functional groups of the adsorbent that are mainly electrostatic in
nature, while they do not influence significantly the adsorption of The prominent parameters affecting the adsorption process are
metals that are linked with strong bonds (e.g. covalent bonds) with the solution pH, the adsorbate type and concentration and ionic
the adsorption sites [185]. strength. Furthermore, the presence of competing ions in the liq-
Trgo and Perić [182] concluded that the dissolution of zeolitic uid medium and/or certain wastewater compounds can adversely
tuff is affected more by ionic strength than by the pH. Baker et al. impact on the adsorption process. The chemical and/or thermal
[64] investigated the effect of ionic strength in batch and column pre-treatment of natural mineral also modifies its cation exchange
experiments for zinc adsorption on zeolite. The authors found that capacity. However, mineral modification is not always associated
that the change in ionic strength (i.e. 0.01–0.05 M NaClO4 ) in the with increased metal uptake. The adsorbent properness for the
444 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

Table 9
Evaluation of the effect of several parameters on mineral adsorption capacity and on heavy metal removal.

Parameter Evaluation result

Adsorption capacity Metal removal (%)

Presence of competing ions ↓ It depends on metal type ↓


Composition of liquid medium where adsorption takes place ↓ In wastewater compared to aqueous solutions ↑ In wastewater compared to aqueous solutions
Presence of ligands ↓ It depends on metal and ligand type ↑ Insoluble metals ↓ Soluble metal complexes
Increase in initial metal concentration ↑ Until a specific value ↓
Decrease in mineral grain size ↑ Or relatively constant ↑
Increase in solution pH ↑ Until a specific value and then↓ ↑
Increase in mineral concentration ↓ Or remains almost stable ↑
Increase in total suspended solids of wastewater ↑
Increase in temperature ↑ ↑

removal of one or more metals must be examined in the real envi- place [214]. The reaction kinetic equation limits the uptake rate
ronment (i.e. true effluents) that it will be applied. The evaluation of in only one mechanism [215]. Nevertheless, metal adsorption on
the adsorbent efficiency must be based on the comparative assess- minerals is considered a diffusion process in which more than one
ment of the impact of several parameters (e.g. pH, composition of stage may be observed to take place.
liquid medium, presence of other ions and compounds). Table 9 Table 13 [40,44,47,49,56,63,86,87,98,99,120,124,125,142,189,
summarizes the common tendencies concerning the impact of sev- 204] lists the diffusion model constants including the effective
eral parameters on the adsorption of metals on bentonite, zeolite diffusion coefficients (Deff ) concerning Ni and Zn adsorption on
and vermiculite in wastewater and aqueous solutions. minerals. The adsorption of metal ions on a solid surface occurs
due to the concentration difference between the liquid and the
5. Removal efficiencies solid phase. The transfer of metal ions from the liquid to the solid
phase is a complex process, which involves the diffusion into the
It is evident that the adsorption performance depends on boundary layer that surrounds the solid particles, as well as, on
several parameters which have been reviewed in this work. the surface and the interior of the solid particles. The adsorption
Depending on the operating conditions, the removal efficiencies of metal ions on the solid surface includes the following stages
of zinc and nickel by zeolite, bentonite and vermiculite are vari- [63,216]: (i) the transfer of the metal ions from the bulk solution
able and are summarized in Table 10 [6,11,26,40,47–52,54–57, to the liquid–solid interface which takes place through convection,
59–62,64,65,68,74,80–82,84,96–98,100–102,105,108,123,125,126, (ii) the diffusion of the metal ions through the boundary layer that
128,129,141–146,153,155,163,173]. surrounds the solid particles, reaching the solid surface (i.e. film
diffusion), (iii) the diffusion of metal ions on the surface (surface
6. Kinetics diffusion) and the interior (pore diffusion) of the solid particle (iv)
the adsorption of the metal ions to active adsorption sites on the
The prediction of the rate at which adsorption takes place is an mineral’s surface. This stage is considered to be an equilibrium
essential factor for the successful design of an adsorption system. reaction.
The adsorbate residence time and the reactor dimensions depend The stages that define the process may take place either in par-
on the adsorption system kinetics [21]. The examination of system allel or in series. If they take place in parallel and are independent
kinetics can reveal the adsorption mechanism. In adsorption of each other, then the total rate is given by the sum of the rate
studies it is important to determine the underlining mecha- of each stage. On the contrary, if they take place in series then at
nisms which may include external diffusion, internal diffusion steady state conditions these have the same rate which is equal
and/or chemical reactions. To identify these mechanisms several to the total rate. In solid–liquid systems and in non-catalytic reac-
reaction kinetics and diffusion models are available. Table 11 tions the stages are considered to take place in series [172]. The
[63,98,187,192–209] shows the reaction kinetic and diffusion slowest stage is the rate controlling step. According to Helfferich
models employed to study the adsorption process. The kinetic [217] stage (iv) takes place very quickly and does not influence the
models that are usually examined consist of the pseudo-first-order, rate of diffusion. In addition, stage (i) occurs very quickly since the
the pseudo-second-order and Elovich’s model. These models are metal ions are readily available at the surface of the boundary layer.
widely employed for examining the kinetics of adsorption. Table 12 Therefore, the total rate of adsorption is defined either by the rate
[40,42,44,47,49,56,61,63,68,92,96–99,101,105,120,123–125,142, of the boundary layer diffusion (rate ii) or by the rate of the intra-
154,174,176,210] summarizes the results obtained by employ- particle diffusion (rate iii) or by the rates of both stages. If the third
ing reaction kinetic models to predict the kinetics of Ni and stage is the only one that controls the process, then a concentra-
Zn uptake by minerals. Several researchers have reported that tion gradient exists only within the solid, while when the process
the adsorption of these metals on minerals follows the pseudo- is controlled only by the second stage, then there is a concentra-
second-order model [44,56,63,68,105,108,120,123–125], while tion gradient only in the boundary layer [217–220]. In Fig. 2 [220],
some researchers have indicated the applicability of the pseudo- stages (ii) and (iii) are represented diagrammatically. The top part
first-order model [107,173]. The fit of the linear forms of these of Fig. 2 shows the external mass transfer resistance that occurs at
equations to experimental data is conducted using linear regres- the liquid–solid interface. At the lower part of Fig. 2 intraparticle
sion analysis. The non-linear equations are solved by minimizing diffusion is shown to take place.
a specific error function among the experimental and the model Sprynskyy et al. [52] noted that the adsorption of nickel on
values. The second-order equation describes the chemical sorption clinoptilolite was a three-stage process associated with fast adsorp-
and thus some type of chemical reaction is considered to take tion at the initial stage, an inversion stage with the prevalence
place that includes valence forces with the exchange of ions or of desorption at the second stage and moderate adsorption of
the formation of covalent bonds [211–213]. However, the good fit nickel inside the mineral’s micro-crystal interior at the third stage.
of the second-order model with the experimental data does not Al-Degs et al. [219] investigated the uptake of zinc by a natural
necessarily reveal the true nature of the mechanism that takes aluminosilicate and carbonate sorbent and found that the sorption
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 445

Table 10
Summary of removal efficiencies achieved by adsorption of nickel and zinc on zeolite, bentonite and vermiculite.

Metal Solution/ Adsorbent C0 (mg L−1 ) pH/T (◦ C) Adsorbent Removal Reference


Wastewater concentration (g L−1 ) efficiency (%)

Ni Aqueous Montmorillonite 5.87 5.5/Room temp. 1 g/column >99 [26]


Ni Aqueous Vermiculite 5.87 5.5/Room temp. 1 g/column 33 [26]
Ni Aqueous Vermiculite 5.87 8.0/Room temp. 1 g/column 78 [26]
Ni Aqueous Clinoptilolite 50 8/23 10 68.2 [40]
NaCl clinoptilolite 77.1
HCl and NaCl clinoptilolite 92.5
Ni Aqueous ZrO-montmorillonite 50 1.0–8.0/30 2.0 17.7–38.6 [47]
TBA-montmorillinite 17.1–33.1
Ni Aqueous Zeolite 100 10/25 8 91.9 [48]
6/25 8 86.5
Ni Aqueous Calcined Bofe bentonite 50 2.18–9.94/20 10 15.0–87.5 [49]
Ni Aqueous Calcined Bofe bentonite 50 5.3/20 5–40 28.31–81.93 [49]
Ni Aqueous Calcined Bofe bentonite 2.66–177 5.3/20 10 93.23–13.56 [49]
Nia Aqueous Clinoptilolite 10 5/– 5 30 [50,51]
Zna 88
Nia Aqueous Chabazite 10 5/– 5 99 [50,51]
Zna 99
Ni Aqueous Clinoptilolite 50 5/25 50 57 [54]
Zn 98
Ni Aqueous Clinoptilolite 10–40 6.2/25 2.5 [52]
Ni Nickel Synthetic zeolite 50 4.0/Room temp. 2.5 86 [52]
electroplating
wastewater
Ni Aqueous Zeolite 31.69 7.8/20 20 80 [55]
Ni Aqueous Zeolite 6.22 7.5/20 100 75 [55]
Ni Aqueous Na-montmorillonite 50 6.5/25 6 90 [56]
Acid modified 75 58
montmorillonite
Ni GMZ-Bentonite 6.0 8.5/30 0.5 >90 [68]
Aqueous 6.4/30 1.0–1.2 52
Ni EDTA Na-montmorillonite 5.87 5.5/25 Fixed bed column 8.5 [74]
NTA 5.9
Tartaric acid 65.3
Oxalic acid 93.9
Manolic acid 100
Succinic acid 96.3
Glutaric acid 98.6
Citric acid 95.3
Ni Aqueous Na-bentonite 100 6/22 10 100 [82]
Ca-bentonite 100
Na-bentonite 100
Ca-bentonite 100
Ni Aqueous Vermiculite 320 6/25 10 51.3 [96]
Bentonite 29.8
Zeolite 19.0
Ni Aqueous Clinoptilolite 25 7/20 15 93.6 [98]
Ni Aqueous Bentonite 100 3.0/60 2.5 72.72 [100]
Ni Na-bentonite 50 9.0/25 1 80 [101]
Aqueous –/25 6 >95
Ni Aqueous Na-montmorillonite 5.87 9/20 0.6 ≈95 [102]
5.8/20 0.8 42
Ni Aqueous Bentonite 500 –/25 100 98.4 [123]
8/25 10 77.8
Ni Aqueous Scolecite 30 6/ 5 10 [128]
Room temp.
Ni Synthetic and Na-zeolite 50 7/25 10 36 [141]
wastewater Gördes zeolite 10 40
Bigadic zeolite 5 38
Bigadic zeolite 10 35
Ni Aqueous Montmorillonite 10 5.7/30 2.0 84.3 [142]
50 62.7
Ni Aqueous Acid-activated 10 5.7/30 2.0 91.5 [142]
montmorillonite 50 69.2
Ni Aqueous Montmorillonite 50 1.0–10.0/30 2.0 50.1–96.4 [143]
Acid-activated 56.3–99.2
montmorillonite
Ni Aqueous Montmorillonite 50 5.7/30 2.0 62.7 [144,145]
TBA-montmorillonite 26.5
Ni Multi metal Vermiculite 14.67 2.0/25 20 33.3 [146]
aqueous Acid treated vermiculite 21.1
solution
Ni Aqueous Zeolite 50 6/Room temp. 10 45 [163]
NaCl-zeolite 58
Ni Multi (Pb, Cd, Zeolite 50 6/Room temp. 10 15 [163]
Cu, Zn, Ni) NaCl-zeolite 22
446 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

Table 10 (Continued)

Metal Solution/ Adsorbent C0 (mg L−1 ) pH/T (◦ C) Adsorbent Removal Reference


Wastewater concentration (g L−1 ) efficiency (%)

Ni Aqueous Zeolite-clinoptilolite 50 –/Room temp. 10 44 [173]


Zn 57
Zn Aqueous Clinoptilolite 5 6.74–6.70/25 10.5 98 [6]
10–100 94–66
Zn Aqueous Fe-oxide clinoptilolite 10–100 6.90–6.75/25 10.5 >99 [6]
Zn Aqueous Zeolite–Portland cement 600 8.0/23 40 90 [11]
mixture
Aqueous Montmorillonite 6.54 5.5/ 1 g/column >99 [26]
Zn Room temp.
Zn Aqueous Vermiculite 6.54 5.5/ 1 g/column 32 [26]
Room temp.
Zn Aqueous Vermiculite 6.54 8.0/ 1 g/column >99 [26]
Room temp.
Zn Aqueous Zeolite 30.73 7.4/20 20 71 [55]
Zn Aqueous Zeolite 6.93 7.0/20 100 90 [55]
Zn Aqueous Vermiculite 25–500 Unadjusted/25 10 74.04–9.92 [57]
150 99.42–62.39
Zn Aqueous Zeolite 37.27–659 5.53–5.26/23 10 92 [59]
Zn Aqueous Gordes zeolite 20 4/ 1 68 [60]
Bigadic zeolite Room temp. 68
Zn Cankiri bentonite 100 8.0/23 5 >98 [61]
Aqueous
Zn Aqueous Zeolite 50 6–7/25 20 45.96 [62]
Zn Aqueous Jordanian zeolite (column) 5 6/27 5 97.6 [64]
20 99.0
Zn Aqueous Jordanian zeolite 5 6/27 5 77 [64]
(batch system) 20 85
Zn Aqueous Turkish zeolite 21 5.6/25 20 100 [65]
Zn EDTA Na-montmorillonite 6.54 5.5/25 Fixed bed column 0.3 [74]
NTA 0.1
Tartaric acid 49.4
Oxalic acid 63.9
Manolic acid 81.2
Succinic acid 97.4
Glutaric acid 99.3
Citric acid 94.4
Zn Aqueous Humic acid-immobilized- 25 10/30 2 99 [80]
amine-modified
polyacrylamide/bentonite
composite
Zn Acid zinc Synthetic zeolite 50 6.0/ 2.5 66 [81]
wastewater Room temp.
Zn Aqueous Na-bentonite 100 6/22 10 100 [82]
Ca-bentonite 100
Na-bentonite 100
Ca-bentonite 100
Zn Multi Synthetic Zeolite (synthetic MAP) 0.5 3.2/– 10 96.8 [84]
(Pb, Cu, Zn, solution Zeolite (natural mordenite) 41.8
Cd)
Zn Multi Motorway Zeolite (natural mordenite) 0.5 7.1/– 10 96.5 [84]
(Pb, Cu, Zn, stormwater 10.1
Cd)
Zn Aqueous Vermiculite 320 6/25 10 75.3 [97]
Bentonite 82.4
Zeolite 49.3
Zn Aqueous Zeolite 4A 25–100 6.5/30 25 97–87 [105]
Zeolite 13X 94.9–78.8
Bentonite 92.4–72.0
Zn Aqueous Bentonite 100 6/25 10 ≈66 [108]
40 ≈80
Zn Aqueous Serbian clinoptilolite 100–600 6/25 1 13–26 [125]
Zn Aqueous Bentonite – 5/20 10 33.9 [126]
Zn Acidic mine Natural zeolite 49 2.2 20 23.4 [129]
water 41 5.5 93.6
Weakly acidic
mine water
Zn Aqueous Vermiculite 16.35 2.0/25 20 36.42 [146]
Acid treated vermiculite 18.42
Zn Aqueous Heated (110 ◦ C) bentonite 58.22 3.32/30 0.25 13.9 [153]
Heated (200 ◦ C) bentonite 4.54/30 0.50 19.9
0.5 mol L−1 H2 SO4 2.88/30 0.25 11.2
bentonite 2.29/30 0.25 8.2
2.5 mol L−1 H2 SO4
bentonite
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 447

Table 10 (Continued)

Metal Solution/ Adsorbent C0 (mg L−1 ) pH/T (◦ C) Adsorbent Removal Reference


Wastewater concentration (g L−1 ) efficiency (%)

Zn Aqueous Montmorillonite 0.059 6.5/25 2.5 49 [155]


HyA-montmorillonite ∼100
HAS-montmorillonite ∼100
Zn Aqueous Zeolite 50 6/Room temp. 10 58 [163]
NaCl-zeolite 93
Zn Multi (Pb, Cd, Zeolite 50 6/Room temp. 10 18 [163]
Cu, Zn, Ni) NaCl-zeolite 55
a
Mixed metal solution.

adsorption capacity (qe ). On the contrary, the pseudo-first-order


model could not adequately predict qe , despite its high correla-
tion with the experimental data. The application of the intraparticle
diffusion model showed that this mechanism was important; the
authors also concluded that liquid film diffusion was influential in
the process of nickel adsorption on natural and modified montmo-
rillonite [44,47]. Trgo et al. [204] used the homogeneous diffusion
model and Vermeulen’s approximation to determine Deff for zinc
adsorption on Na-modified clinoptilolite and found that it was
of the order of 10−12 m2 s−1 and was independent of the initial
metal concentration, while it decreased with increasing adsorp-
tion time. The authors identified two linear areas for both models.
The first area was associated with diffusion through the mineral
macropores, while the second area was associated with diffusion
through the crystal lattice. Taparcevska et al. [189] applied the
parabolic diffusion model and Vermeulen’s approximation on their
experimental data of zinc adsorption on clinoptilolite and found
an effective diffusion coefficient that was two orders of magni-
tude higher than that obtained by Trgo et al. [204] and concluded
that Deff depended on initial metal concentration. Rajic et al. [120]
found that the adsorption process on clinoptilolite involved film
diffusion, intraparticle diffusion and cation exchange. Two distinct
adsorption regions were identified; the first one corresponding
to film diffusion characterized by high diffusion rates and the
second one to intraparticle diffusion having lower diffusion rate.
The authors also mentioned that the pseudo-second-order model
closely matched the experimental results. Sen and Gomez [63] con-
cluded that the kinetics of zinc adsorption on bentonite was a
two-stage process involving a rapid adsorption of the metal ions
Fig. 2. Diagrammatic representation of stages (ii) and (iii) of the diffusion of metal on the mineral’s external surface followed by intraparticle diffu-
ions in the adsorbent [220].
sion. The two-stage adsorption process has also been observed
by Stojakovic et al. [125]. Katsou et al. [96,97] found that the
rate was slow with only 10–20% of the adsorption occurring dur- adsorption of zinc and nickel on zeolite, bentonite and vermicu-
ing the first 30 min. The experimental kinetics data fitted well lite followed a three-stage diffusion process, where external mass
to the pseudo-first-order and to the external diffusion models transfer was dominant at the early stages and intraparticle diffu-
and thus external diffusion was more important than intraparti- sion at the later stages of the process. The authors found that Deff
cle diffusion. Ijagbemi et al. [42] found that the rate constant of was of the order of 10−13 –10−14 m2 s−1 and its value followed the
the pseudo-second-order model decreased with increasing initial order vermiculite > bentonite > zeolite. A three-stage diffusion pro-
metal concentration and thus more time was required to reach cess has also been identified by Ijagbemi et al. [56] for nickel uptake
equilibrium. The authors also examined the intraparticle diffusion by Na-montmorillonite and acid modified montmorillonite, with
model and identified two distinct plots indicating that intraparticle limited contribution of mass transfer and boundary layer diffusion.
diffusion was not the only rate limiting mechanism. They applied Taparcevska et al. [189] found that the increase of metal concen-
Boyd’s model and concluded that film diffusion is also important tration until a certain point is not accompanied by a decrease in the
at least to some extent. Gupta and Bhattacharyya [44] compared effective diffusion coefficient. Other researchers did not find a clear
the experimental data with various kinetic models and found that trend between Deff and metal concentration [63,204]. The effect of
the pseudo-second-order model closely matched the experimental competing ions and of the adsorbate type on the effective diffusion
data and could provide a reasonable prediction for the equilibrium coefficient has not been documented.

Table 11a
Reaction kinetic models.

Model Non-linear equation Linear equation Model parameters Reference


−k1 t
Pseudo-first-order qt = qe (1 − e ) ln(qe − qt ) = ln qe − k1 t qe , k1 [192]
Pseudo-second-order qt = q2e k2 t/(1 
+ qe k2 t)   2 qe ) + t/qe
t/qt = 1/(k 2
qe , k2 [193]
Elovich qt = (1/ˇe )ln ˛ˇe t + 1 qt = ln ˛ˇe /ˇe + ln t/ˇe ˛, ˇe [194,195]
448
Table 11b

S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461


Diffusion models.

Model Equation Model parameters Reference

Intraparticle qt  0.5
= kid t  kid [196]
External mass transfer ln (C0 − Cs ) / (Ct − Cs ) = kf am t; am = 6ma / (dp b ) kf , am , Cs [187,197,198]
 0.5
F ≤ 0.85 ⇒ Bt = 2 − 2 F/3 − 2 1 − F/3 ;
Reichenberg F > 0.85 ⇒ Bt = −0.4977 − ln (1 − F); Deff [199,200]
B = 2 Deff /r 2 ; F = qt /qe
 

   
Boyd F = 1 − 6/2 1/n2 exp −n2 Bt [201]

n=1   0.5
Non-linear: F = 1 − exp
 −
2 2
Deff t /r
Vermeulen   Deff [202]
Linear: ln 1/ 1 − F 2 = 2 Deff t/r 2
   
Homogeneous diffusion model (t→∞) ln 1 − qt /qe = ln 6/2 − Deff 2 t/r 2 Deff [203]
 0.5  √ −1  1.5
Parabolic diffusion Original formula F = 4−0.5 Deff t/r 2 − Deff t/r 2 − 3  Deff t/r 2 Deff [204,205]
−0.5
   0.5
Parabolic diffusion Simplified formula F/t = 4 2
Deff / r t − Deff /r
2
Deff [204,205]
Liquid film diffusion ln (1 − F) = −kfd t kfd [201]


 
F =1− ai exp −bi t/ ;  = r /Deff ;
2

Homogeneous diffusion Deff [204–206]


i=0
Spherical particles and t → t0 ⇒ F = 6−0.5 Deff
0.5 0.5
t /r
 
Spherical particles and t → t∞ ⇒ ln(1 − F) = ln 6/2 − 2 Deff t/r 2
Film diffusion equation Df = 0.23rqe ı/t1/2 Df [98,207,208]
Pore diffusion equation Dp = 0.03r 2 /t1/2 Dp [98,207,208][63,209]
Table 12
Applications of reaction kinetic models on nickel and zinc adsorption on zeolite, bentonite and vermiculite.

Metal Mineral C0 (mg L−1 )/pH/T Experimental qe Pseudo-first-order Pseudo-second-order Elovich Equilibrium Reference
(◦ C) (mg g−1 ) time (h)

Predicted qe R2 Predicted qe R2 R2
(mg g−1 ) (mg g−1 )
Ni Clinoptilolite 50/–/23 3.56 0.9998 0.9585 3–3.33 [40]
NaCl clinoptilolite 4.19 1.000 0.8876
HCl and NaCl clinoptilolite 4.22 0.9775 0.9642
Ni Montmorillonite 50/6/25 1.52 2.56 0.887 1.53 0.998 [42]
100/6/25 3.11 4.36 0.721 3.15 0.999
200/6/25 5.68 5.66 0.583 5.70 0.998
300/6/25 5.61 6.05 0.688 5.62 0.999

S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461


Ni Montmorillonite 50/5.7/30 15.70 12.69 0.97 16.52 0.99 0.99 3 [44,47]
ZrO-montmorillonite 7.43 13.32 0.96 9.00 0.99 0.99
TBA-montmorillinite 6.38 11.71 0.97 8.00 0.99 0.99
Ni Calcined Bofe bentonite 50/5.3/20 ≈1.8 0.35 0.9009 1.80 0.9971 2.5 [49]
Ni 50/6.5/25 7.54 5.04 0.988 7.81 0.999 0.975 3.83 [56]
Na-montmorillonite 75/6.5/25 8.29 7.84 0.860 8.37 0.999 0.985
100/6.5/25 10.04 9.75 0.998 10.18 0.999 0.987
Ni Acid modified montmorillonite 50/6.5/25 4.57 3.77 0.972 4.95 0.999 0.981 [56]
75/6.5/25 7.09 5.44 0.978 7.47 0.999 0.984
100/6.5/25 7.31 8.32 0.979 7.61 0.999 0.988
Ni GMZ Na-bentonite 6.0/6.4/30 4.56 0.9997 24 [68]
Ni (Co, Cr, Cu, Zn, Ni) Zeolite 4A 50/3/25 7.59 7.63 0.963 [92]
100/3/25 3.98 3.87 0.942
50/4/25 11.51 10.49 0.956
100/4/25 6.14 4.95 0.989
Ni Zeolite 320/6/25 6.08 4.12 0.9046 6.51 0.9993 0.9792 2 [96]
Bentonite 9.52 5.72 0.9831 10.02 0.9995 0.9732 2
Vermiculite 16.42 5.99 0.9499 16.81 0.9998 0.8726 2
Ni Clinoptilolite 25/7/20 1.61 0.174 1.60 0.9998 [98]
25/7/40 1.58 0.15 1.59
25/7/60 1.56 0.12 1.58
Ni Chitosan–clinoptilolite composite 4109/5/25 375.9 536.9 0.940 416.8 0.999 24 [99]
Ni Na-bentonite 50/–/25 13.02 0.9844 13.66 0.9967 3.33 [101]
Ni Zeolitic tuff 100/6/25 1.9 0.964 2.23 0.988 >24 [120]
100/6/35 3.1 0.988 3.35 0.990
100/6/45 3.8 0.978 4.11 0.991
100/6/55 4.9 0.990 5.20 0.989
Ni Zeolitic tuff 200/6/25 2.5 0.974 2.78 0.992 >24 [120]
200/6/35 4.0 0.996 4.35 0.990
200/6/45 5.4 0.984 5.61 0.993
200/6/55 6.9 0.992 7.18 0.988
Ni Zeolitic tuff 300/6/25 3.0 0.998 3.31 0.999 >24 [120]
300/6/35 4.9 0.984 5.12 0.991
300/6/45 6.0 0.987 6.27 0.990
300/6/55 8.3 0.992 8.47 0.995
Ni Zeolitic tuff 400/6/25 3.3 0.989 3.57 0.992 >24 [120]
400/6/35 5.2 0.990 5.36 0.997
400/6/45 6.2 0.962 6.42 0.990
400/6/55 8.9 0.978 9.15 0.989
Ni Bentonite 200/–/25 ≈9 0.56 0.06 9.19 0.99 0.5 [123]
Ni Montmorillonite acid-activated 50/5.7/30 15.7 12.7 0.97 16.5 0.99 0.99 3 [142]
montmorillonite 17.3 12.9 0.96 18.1 0.99 0.99
Ni Activated carbon – zeolite composite 500/–/25 115.63 65.74 [174]
(type 1)

449
450
Table 12 (Continued)

Metal Mineral C0 (mg L−1 )/pH/T Experimental qe Pseudo-first-order Pseudo-second-order Elovich Equilibrium Reference
(◦ C) (mg g−1 ) time (h)

Predicted qe R2 Predicted qe R2 R2
(mg g−1 ) (mg g−1 )
Ni (Ni, Pb) Vermiculite Variable/5.0/20 ≈12.6 0.9379 13.04 0.9940 0.8729 1 [176]
Ni (Ni, Ag) ≈14.2 0.9522 14.23 0.9931 0.8549
Ni (Ni, Cd) ≈8.9 0.9269 9.01 0.9967 0.8405
Ni (Ni, Ag, Pb) ≈12.6 0.9917 12.77 0.9932 0.8878
Ni (Pb, Ni, Cd) ≈8.5 0.5206 8.73 0.9998 0.5762
Ni (Ag, Ni, Cd) ≈9.0 0.9350 9.16 0.9986 0.4768
Ni (Ag, Ni, Cd, Pb) ≈7.8 0.9634 8.09 0.9987 0.7921
Zn Cankiri bentonite 20/8/23 3.6 0.377 0.9211 3.60 0.9998 0.17 [61]
100/8/23 19.6 1.116 0.8162 19.60 0.9999
160/8/23 31.6 2.298 0.8411 31.64 1.000
Zn Bentonite 30/6.76/30 7.5 7.97 0.9852 1.33 [63]

S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461


40/6.76/30 10.0 9.89 0.9765
50/6.76/30 47.2 47.17 0.9958
Zn Bentonite 30/3.81/30 11.8 11.93 0.9847 1.33 [63]
30/6.76/30 47.4 46.95 0.9954
30/7.69/30 28.8 28.49 0.9978
Zn (Co, Cr, Cu, Zn, Ni) Zeolite 4A 50/3/25 26.58 30.14 0.997 [92]
100/3/25 9.22 10.54 0.975
50/4/25 40.38 41.17 1.000
100/4/25 19.59 20.10 0.997
Zn Zeolite 320/6/25 15.76 13.23 0.9803 18.35 0.9951 0.8565 2 [97]
Bentonite 26.36 15.76 0.9901 27.10 0.9988 0.9479 1
Vermiculite 24.08 10.05 0.8850 24.21 0.9993 0.9825 1
Zn Zeolite 4A Low value 0.9975 1.5 [105]
Zeolite 13X 25/6.5/30 0.9963
Bentonite 0.9964
Zn Zeolite A 500/–/25 158.87 0.947 263.29 0.993 0.75 [124]
500/–/40 185.03 0.927 355.67 0.991
500/–/60 207.91 0.941 402.54 0.995
Zn Na-clinoptilolite 100/6/25 ≈5.6 0.995 6.67 0.998 >24 [125]
100/6/35 ≈6.5 0.985 8.30 0.996
100/6/45 ≈7.6 0.980 8.81 0.994
100/6/55 ≈8.3 0.998 8.71 0.999
Zn Na-clinoptilolite 200/6/25 ≈8.0 0.987 9.98 0.995 >24 [125]
200/6/35 ≈10.3 0.983 12.50 0.996
200/6/45 ≈11.5 0.997 13.87 0.998
200/6/55 ≈11.8 0.998 12.7 0.999
Zn Na-clinoptilolite 300/6/25 ≈9.7 0.909 12.0 0.985 >24 [125]
300/6/35 ≈11.8 0.996 14.3 0.998
300/6/45 ≈13.0 0.994 15.9 0.998
300/6/55 ≈13.6 0.969 13.9 0.999
Zn Na-clinoptilolite 400/6/25 ≈10.0 0.941 12.1 0.990 >24 [125]
400/6/35 ≈12.8 0.993 14.1 0.999
400/6/45 ≈14.0 0.996 15.5 0.999
400/6/55 ≈14.8 0.999 15.2 0.999
Zn Bentonite 200/5.5/28 16.36 16.18 0.9996 1 [154]
Sulphate-modified bentonite 17.17 17.21 0.9997
Phosphate-modified bentonite 17.50 17.76 0.9996
Zn Bentonite 300/5.5/28 24.48 26.11 0.9941 1 [154]
Sulphate-modified bentonite 26.19 26.95 0.9948
Phosphate-modified bentonite 26.63 28.01 0.9953
Zn Natural zeolite 50/const./25 ≈0.8 0.169 0.244 [210]
100/const./25 ≈1.7 0.385 0.267
120/const./25 ≈2.0 0.616 0.623
150/const./25 ≈2.7 0.762 0.766
180/const./25 ≈2.7 0.897 0.899
200/const./25 ≈2.0 0.945 0.950
250/const./25 ≈2.3 0.980 0.980
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 451

Table 13
Intraparticle diffusion model and effective diffusion coefficients for nickel and zinc adsorption on zeolite, bentonite and vermiculite.

Metal Adsorbent C0 (mg L−1 )/pH/T Intraparticle diffusion Deff (m2 s−1 ) Model Reference
(◦ C)

R2 k (mg g−1 min−1/2 )

Ni Clinoptilolite 0.9650 44.45 [40]


NaCl clinoptilolite 0.8997 64.12
HCl and NaCl 0.9762 68.42
clinoptilolite
Ni Montmorillonite 50/5.7/30 0.95 0.417 [44,47]
ZrO-montmorillonite 0.95 0.436
TBA-montmorillinite 0.94 0.400
Ni Calcined Bofe 50/5.3/20 0.9557 0.0245 [49]
bentonite
Ni 50/6.5/25 0.978 0.122/0.118/0.020 [56]
Na-montmorillonite 75/6.5/25 0.987 0.429/0.204/0.031
100/6.5/25 0.987 0.738/0.384/0.036
Ni Acid modified 50/6.5/25 0.981 0.285/0.259/0.046 [56]
montmorillonite 75/6.5/25 0.984 0.334/0.311/0.052
100/6.5/25 0.989 0.584/0.343/0.058
Zn Bentonite 30/6.76/30 7.33 × 10−13 Film diffusion [63]
40/6.76/30 1.10 × 10−13 equation
50/6.76/30 1.41 × 10−12
Ni Zeolite 320/6/25 0.998/0.961/0.983 0.7/0.3/0.1 5.57 × 10−14 Vermeulen [86]
Bentonite 0.999/0.889/0.909 1.3/0.4/0.2 6.77 × 10−14
Vermiculite 0.990/0.957/0.947 3.0/1.1/0.1 7.18 × 10−14
Zn Zeolite 320/6/25 0.961/0.983/0.977 1.3/1.9/0.6 [87]
Bentonite 0.987/0.974/0.946 4.2/1.6/0.4
Vermiculite 0.980/0.986/0.966 5.5/1.3/0.3
Ni Clinoptilolite 25/7/20 2.5 × 10−13 [98]
25/7/40 2.3 × 10−13 Pore diffusion
25/7/60 2.0 × 10−13 equation
Ni Clinoptilolite 25/7/20 3.1 × 10−13 [98]
25/7/40 2.4 × 10−13 Film diffusion
25/7/60 1.9 × 10−13 equation
Ni Chitosan-clinoptilolite 4109/5/25 0.942 6.04 [99]
composite
Ni Zeolitic tuff 100/6/25 0.999 0.0272 [120]
100/6/35 0.958 0.0372
100/6/45 0.999 0.0370
100/6/55 0.929 0.0336
Ni Zeolitic tuff 200/6/25 0.974 0.0367 [120]
200/6/35 0.938 0.0390
200/6/45 0.999 0.0351
200/6/55 0.922 0.0361
Ni Zeolitic tuff 300/6/25 0.957 0.0373 [120]
300/6/35 0.926 0.0493
300/6/45 0.972 0.0383
300/6/55 0.960 0.0358
Ni Zeolitic tuff 400/6/25 0.997 0.0337 [120]
400/6/35 0.967 0.0330
400/6/45 0.989 0.0297
400/6/55 0.944 0.0328
Zn Zeolite A 500/-/25 0.999 26.54 1.83 × 10−12 [124]
500/-/40 0.993 30.79 2.34 × 10−12
500/-/60 0.998 33.34 3.05 × 10−12
Zn Na-clinoptilolite 100/6/25 0.999 0.0932 [125]
100/6/35 0.999 0.0590
100/6/45 0.999 0.0590
100/6/55 0.991 0.0827
Zn Na-clinoptilolite 200/6/25 0.997 0.136 [125]
200/6/35 0.999 0.161
200/6/45 0.986 0.153
200/6/55 0.972 0.123
Zn Na-clinoptilolite 300/6/25 0.957 0.107 [125]
300/6/35 0.991 0.165
300/6/45 0.995 0.228
300/6/55 0.999 0.0745
Zn Na-clinoptilolite 400/6/25 0.986 0.116 [125]
400/6/35 0.991 0.140
400/6/45 0.980 0.174
400/6/55 0.984 0.0961
Ni Montmorillonite 50/5.7/30 0.95 0.42 [142]
Acid-activated 0.95 0.42
montmorillonite
Zn Clinoptilolite 50/5–7/25 0.57 × 10−12 Vermeulen [189]
100/5–7/25 0.35 × 10−12
150/5–7/25 0.25 × 10−12
250/5–7/25 0.13 × 10−12
452 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

Table 13 (Continued)

Metal Adsorbent C0 (mg L−1 )/pH/T Intraparticle diffusion Deff (m2 /s) Model Reference
(◦ C)

R2 k (mg g−1 min−1/2 )

Zn Clinoptilolite 50/5–7/25 3.33 × 10−12 Parabolic [189]


100/5–7/25 2.40 × 10−12 diffusion
150/5–7/25 2.25 × 10−12
250/5–7/25 0.63 × 10−12
Zn Na-modified 194.8/5.56/23 3.83 × 10−14 Vermeulen [204]
clinoptilolite 396.1/5.36/23 6.22 × 10−14
490.5/5.42/23 5.70 × 10−14
600.5/5.33/23 3.08 × 10−14
718.7/5.26/23 3.58 × 10−14
Zn Na-modified 194.8/5.56/23 4.30 × 10−14 Homogeneous [204]
clinoptilolite 396.1/5.36/23 7.17 × 10−14 diffusion
490.5/5.42/23 6.23 × 10−14
600.5/5.33/23 4.68 × 10−14
718.7/5.26/23 4.45 × 10−14

Allen et al. [220] also attributed the initial stage of adsorption to of the system’s performance and the optimisation of adsorbent use.
external mass transfer through the boundary layer surrounding the Sorption equilibrium provides fundamental physicochemical data
particles, while the following stages represented intraparticle dif- for evaluating the applicability of sorption processes as a unit oper-
fusion into the branched porous network of the adsorbent particle ation. Sorption equilibrium is usually described by an isotherm
[42,63,120,221–224]. The diffusion mechanism depends on several equation whose parameters express the surface properties and
parameters including the structure and the type of adsorbent, the affinity of the sorbent, at a fixed temperature and pH. Hence, an
physicochemical properties of the adsorbent and adsorbate, as well accurate mathematical description of the equilibrium isotherm,
as, other conditions that may affect the process performance. which is based on a correct adsorption mechanism, is essential
for the effective design of adsorption systems [20]. Equilibrium
7. Isotherms models can be classified into empirical and mechanistic models.
Empirical models cannot represent the mechanisms of the sor-
Sorption isotherm is the equation or curve that connects the bate uptake but can be used to predict the experimental results,
metal concentration that has been adsorbed on the solid phase with while the mechanistic models can represent the system mecha-
metal concentration in the solution at equilibrium for specific tem- nisms [116,225,226]. Therefore, mechanistic models describe the
perature. The application of experimental data on equations that underlying interactions that take place between the metal ions in
describe the isotherms is valuable, since it results in the prediction the solution and the charged surface. Empirical models are usually

Table 14
Equilibrium models.

Isotherm Equation Linear Form Model Parameters Reference

Langmuir qe = qm KL Ce / (1 + KL Ce ) Ce /qe = 1/ (qm KL ) + Ce /qm KL , qm [227]


  [228]
Freundlich qe = KF Ce1/nF ln qe = ln KF + 1/nF ln Ce KF ,nF [229]
Langmuir-Freundlich qe = qm KLF C
1/nLF
e  + KLF Ce LF )
/ (1 1/n
KLF , qm , nLF [226,230]
qe = qm exp −ˇε2 ;
Dubinin–Radushkevich (DR)    −0.5 lnqe = lnqm − ˇε2 ˇ, qm [231]
ε = RT ln 1 + 1/Ce ; E = 2ˇ
     
Temkin qe = RT/b ln (ACe ) qe = RT/b ln A + RT/b ln Ce A, b [232]
ns ns
Sips qe = qm (as Ce ) / (1 + (as Ce ) ) ns , qm , aS [20]
  [230]
Redlich–Peterson qe = KR Ce / (1 + ˛R CeˇR ) ln KR Ce /qe − 1 = ln ˛R + ˇR ln Ce KR , ˛R , ˇR [233]
1    
Toth qe = KT Ce /(AT + CeTa ) ⁄T a ln qe /KT = ln Ce − 1/T ln (AT + CeTa ) AT , KT , Ta [234]
nK
Khan qe = qm bK Ce / (1+ bK Ce )    bK , qm , nK [235]
p p−1
Radke–Prausnitz qe = ˛RP rRP Ce / ˛RP + rRP Ce ln 1/qe − 1/ (˛RP Ce ) = − ln rRP − p ln Ce ˛RP , rRP , p [236]
 
Generalized qe = qm Ceng / (K + Ceng ) ln qm /qe − 1 = ln K − ng ln Ce K, ng [237]
BBET Ce qm Ce
 1   B −1   Ce 
BET qe = = + B BET qm Csb , BBET , qm [238]
   
(Csb −Ce )[(1+(BBET −1)Ce )/Csb ]
nFH   BBET qm
(Csb −Ce )qe BET Csb

Flory–Huggins /C0 = 1/ KFH 1 −  log /C0 = log KFH + nFH log(1 − ) , nFH , KFH [211,239]

     1 − Ce /C0
= 
Frumkin KFF Ce =  exp −f / 1 −  ln qe / (Ce (qF − qe )) = ln KFF + 2fqe /qF f, KFF , [240,241]

N

Competitive Redlich–Peterson qe,i = KR,i Ce,i 1+ ˇR,i


˛R,i Ce,i KR,i , ˛R,i , ˇR,i [233,242]

 i=1


N

Competitive Langmuir qe,i = qm,i KL,i Ce,i 1+ KL,i Ce,i KL,i , qm,i [229]

 i=1


N

Competitive combined qe,i = 1/ni


qm,i KL,i Ce,i 1+ 1/ni
KL,i Ce,i ni , KL,i , qm,i [232]
Langmuir–Freundlich
i=1
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 453

Table 15
Summary of isotherms investigation for nickel and zinc adsorption on zeolite, bentonite and vermiculite.

Metal Metal Sorbent C0 (mg L−1 ) Dosage pH/T (◦ C) qm (mg g−1 ) Isotherm Reference
(g L−1 )

Ni Single Natural zeolite 12.5–200 10 7–8/23 8.7 (L) L [40]


Conditioned zeolite 10.8 (L)
Ni Single Montmorillonite 50–300 20 6/25 12.89 (L) L [42]
Ni Single Montmorillonite 10–50 2 5.7/30 21.14 (L) L, F [44,47]
ZrO-montmorillonite 10.45 (L)
TBA-montmorillonite 8.50 (L)
Ni Single Montmorillonite 10–250 2 31.05 (L) L, F [44,47]
ZrO-montmorillonite 5.7/30 31.44 (L)
TBA-montmorillonite 30.67 (L)
Ni Single Turkish zeolite 1–100 8 5–6/20 119.7 (L) L, F [48]
Ni Single Calcined Bofe bentonite 3–200 10 5.3/20 2.875 (L) L [49]
5.3/40 3.072 (L)
5.3/50 3.047 (L)
5.3/75 3.893 (L)
Ni Single Ukraine clinoptilolite 10–800 2.5 6.2/25 13.03, 15.55 (E, L) L, F [52]
Ni Single Jordanian zeolitic tuff 50–400 1–8 4/20 25.59, 22.30 (L, E) L [53]
(phillipsite)
Ni Single Turkish clinoptilolite 50–3000 50 5/25 8.21, 8.24 (E, L) F, L [54]
Ni Single Na-montmorillonite 50–100 6 6.5/25 10.65, 9.345 (L, DR) RP [56]
Acid modified 11.02, 8.566 (L, DR) DR
montmorillonite
Ni Single GMZ Na-bentonite 0.5 6.4/30 14.38; 104.47 (L, L, DR, F [68]
6.4/45 DR)
6.4/60 14.44; 108.00 (L,
DR)
14.61; 112.69 (L,
DR)
Ni Single Natural zeolite 10–200 2.5 6/22 2.00 (L) L [81]
Synthetic zeolite 20.07 (L) L
Ni Single Na-bentonite 10–150 2.5 6/22 24.2 (L) L [82]
Ca-bentonite 10 6.32 (L)
Ni Single Natural vermiculite 10–100 2.5 19.3 (L) L, F [83]
Exfoliated vermiculite 10 6/22 5.91 (L)
Ni Single Zeolite 10–500 10 16.86, 9.83 (L, DR) L [96,97]
Bentonite 6/25 19.92, 11.96 (L, DR)
Vermiculite 25.51, 20.91 (L, DR)
Ni Single Heated (350 ◦ C) clay 10–200 0.5 5/20 80.9 (L) L [86]
Ternary containing 23.3 (L)
(Ni, Cu, Zn) montmorillonite and
hydromica
Ni Multi (Co, Commercial zeolite 4A 50–300 each 5 3/25 7.9 (L) L [92]
Cr3+ , Cu, Coal fly ash prepared metal 8.96 (L) L
Zn, Ni) zeolite 4A
Ni Single Zeolite 0.1–100 15 7/20 3.28 (L) [98]
7/40 2.97 (L) L
7/60 2.65 (L)
Ni Single Chitosan-clinoptilolite 147–5869 10 5/25 466.03 (L) L [99]
Ni Single Bentonite 50–300 25 3/30 4.31 (L) L [100]
Ni Single Na-montmorillonite 0.6 5.8/20 3.82 (L) L [102]
5.8/37 7.04 (L)
5.8/67 10.56 (L)
Ni Single Natural zeolitic tuff 100–600 10 6/25 4.48, 16.64 (L, S) S [120]
6/35 6.15, 6.43 (L, S)
6/45 7.04, 6.52 (L, S)
6/55 10.42, 10.71 (L, S)
6/65 11.44, 11.46 (L, S)
Ni Single Bentonite 108–427 15 5/25 819.7, 155.1 (L, DR) L [122]
Ni Single Natural bentonite 40–2000 10 –/20 25 (L) L [123]
–/40 27 (L)
–/60 33 (L)
Ni Single Montmorillonite 10–250 2 5.7/30 28.4 (L) L, F [142,145]
Acid-activated 29.5 (L)
montmorillonite
Ni Single Montmorillonite 10–250 2 5.7/30 21.1 (L) L, F [143]
Acid-activated 21.3 (L)
montmorillonite
Ni Single Montmorillonite 10–250 2 5.7/30 28.4 (L) L, F [144]
TBA-montmorillonite 19.7 (L)
Ni Multi (Cd, Vermiculite 5–500 60 4.5/25 0.32 (L) L [166]
Cr, Cu, Ni,
Pb, Zn)
Ni Single Activated carbon-zeolite 9-293 2 25 70.43 (L) L [174]
composite (type 1) 62.22 (L)
Activated carbon-zeolite
composite (type 2)
454 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

Table 15 (Continued)

Metal Metal Sorbent C0 (mg L−1 ) Dosage pH/T (◦ C) qm (mg g−1 ) Isotherm Reference
(g L−1 )

Ni Single Zeolitized vermiculite 29–2935 6–8/22 16.5 (L) L [251]


Clinoptilolite 2.3 (L) L
Zn Single Zeolite–Portland cement 50–1600 40 8/25 12.85 (L) L [11]
mixture
Zn Single Turkish clinoptilolite 50–3000 50 5/25 8.21, 8.24 (E, L) F, L [54]
Zn Single Na-enriched bentonite 12.5–200 1.0 4/25 57.4, 54 (L, E) L [58]
Natural bentonite 30.7, 24 (L, E)
Zn Single Croatian zeolitic tuff 37–659 10 5.3–5.5/25 13.5, 13.1 (DR, DR, MDR [59]
MDR)
Zn Single Gordes zeolite 2.5–20 1 4/25 6.60 (L) F [60]
Bigadic zeolite 4/25 3.59 (L)
Zn Single Cankiri bentonite 20–160 5 8/23 80.64 (L) L, F [61]
Zn Single Zeolite 100–400 20 6–7/30 8.75; 0.02 (L, DR) L, DR [62]
Zn Single Bentonite 10–90 0.25 6.76/30 68.49 (L) L [63]
Zn Single Zeolite 5–35 5 6/27 41.84, 62.90 (L, DR) F, DR [64]
Zn Single Turkish zeolite 78.59–1012.78 15 5.6/ 19.49 (L) L [65]
Zn Single Humic acid-immobilized- 10–300 2 9/30 52.93 L [80]
amine-modified
polyacrylamide/bentonite
composite
Zn Single Natural zeolite 10–200 2.5 6/22 3.47 (L) L [81]
Synthetic zeolite 32.42 (L) L
Zn Single Na-bentonite 10–150 2.5 6/22 23.1 (L) L [82]
Ca-bentonite 10 5.75 (L)
Zn Single Heated (350 ◦ C) clay 10–200 0.5 5/20 63.2 (L) L [86]
Ternary containing 26.6 (L)
(Ni, Cu, Zn) montmorillonite and
hydromica
Zn Multi (Co, Commercial zeolite 4A 50–300 each 5 3/25 31.6 (L) L [92]
Cr3+ , Cu, Coal fly ash prepared metal 30.80 (L) L
Zn, Ni) zeolite 4A
Zn Single Vermiculite 327 2.5 5.2/22 86.30, 60.80 (L, E) L [103]
Zn Single Zeolite 10–500 4–16 Not 3.926 (L) L [104]
adjusted/25 4.097 (L)
Not 4.172 (L)
adjusted/35
Not
adjusted/45
Zn Single Zeolite 4A 25–100 25 2.3/30 42.82 (L) L [105]
Zeolite 13X 38.31 (L)
Bentonite 35.17 (L)
Zn Single Natural bentonite 10–750 2 5/20 36.63 (L) L [106]
Composite E20 bentonite 29.67 (L)
Composite A85 bentonite 14.10 (L)
Zn Single Surfactant-modified [107]
montmorillonite 294–2942 2.5 3/25 13.27 (L) L, DR
Surfactant-modified
montmorillonite
Zn Single Bentonite 50–100 5 6/25 9.12 (L) L [108]
Binary 5.80 (L)
(Zn–Pb)
Zn Single Synthetic zeolite A 10–2000 0.5 –/25 165.41, 281.53 (L, F, DR [124]
–/40 DR)
–/60 166.72, 290.61 (L,
DR)
185.68, 293.16 (L,
DR)
Zn Single Serbian natural 100–600 10 6/25 12.0, 170 (L, S) L [125]
clinoptilolite 6/35 15.5, 22.6 (L, S)
6/45 15.8, 19.5 (L, S)
6/55 16.6, 21.7 (L, S)
6/65 16.8, 17.5 (L, S)
Zn Single Bentonite 300 0.1–10 4.5/20 52.91 (L) L [127]
4.5/40 33.11 (L)
4.5/80 25.77 (L)
Zn Single Clinoptilolite 10 4/25 46.22 (L) L [149]
Na-clinoptiloloite 51.19 (L)
Ca-clinoptilolite 43.09 (L)
Zn Single Clinoptilolite 0–500 20 –/20 3.15 (L) L [151]
Philipsite 1.70 (L)
Chabazite 7.98 (L)
Analcime 3.67 (L)
Bentonite 3.08 (L)
Zn Single Clinoptilolite 1–5 4.54, 4.58 (L, E) L [152]
Na-clinoptilolite 8.14, 8.17 (L, E)
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 455

Table 15 (Continued)

Metal Metal Sorbent C0 (mg L−1 ) Dosage pH/T (◦ C) qm (mg g−1 ) Isotherm Reference
(g L−1 )

Zn Single Heated (110 ◦ C) treated 10–140 1 8/30 3.32 (L) [153]


bentonite 0.5 7.9/30 4.54 (L) L
Heated (200 ◦ C) treated 3.7/30 2.88 (L)
bentonite 3.5/30 2.29 (L)
H2 SO4 (0.5 mol L−1 ) treated 1
bentonite 1
H2 SO4 (2.5 mol L−1 ) treated
bentonite
Zn Multi Vermiculite 5–500 60 4.5/25 2.64 (L) L [166]
(Cd–Cr–Cu–Ni–Pb–Zn)
Zn Single Zeolite 10–500 4–16 Not 3.93 (L) [175]
Binary (Cd, adjusted/25 2.02 (L) L
Zn) 0.881 (L)
Binary (Pb,
Zn)
Zn Single Natural zeolite 50–250 15 Constant/25 3.644, 4.077 (L, LF) LF [210]
Zn Single Natural aluminosilicate 0–1000 187.0 (L) L [219]
and carbonate 1 <7.1/25
Zn Single Clinoptilolite 33–3269 6–8/22 18.0 (L) L [251]
2.5 (L)
Zn Single Zeolite NaY 3–98 4.5/30 92.51 (L) F [252]
Zn Single Zeolite 261–6538 40 5.5/room 96.11 (L) L [253]
temp.

L = Langmuir.
E = Experimental.
F = Freundlich.
DR = Dubinin–Radushkevich.
MDR = Modified Dubinin–Radushkevich.
S = Sips.
RP = Redlich–Peterson.

based on simple mathematical relationships between the liquid than 8 kJ mol−1 [40,56,61,96,122]. Purna Chandra Rao et al. [105]
phase equilibrium concentration and the solid phase equilibrium found that the isotherm shapes for zinc adsorption on zeolite
concentration [14]. Table 14 [20,211,226–242] summarizes the 4A, zeolite 13X and bentonite were identical indicating a similar
main two and three parameter isotherm models that have been sorption process. Several studies showed that the Langmuir model
applied for the investigation of nickel and zinc adsorption on min- provided the best fit to the experimental data [63,96,104,151,175]
erals. The assumptions on which these equations are based have other studies showed that the Freundlich was the most suitable
been mentioned in several works [20,45,116,211,239,243–250]. model [60,252] and some studies showed that both the Langmuir
Isotherm models have also been developed to describe competitive and Freundlich models could adequately predict the experimental
adsorption when more than one type of heavy metals is present data [44,52,54,83,142,144]. The high correlation of both the Lang-
in the solution. These models can be simple equations that are muir and the Freundlich models with experimental data can be
associated with the individual parameters only (i.e. non-modified interpreted as a possible monolayer and heterolayer formation of
models) and more complex models exploiting the individual model metallic compounds on the mineral surface [254]. The majority
parameters together with suitable correction factors (modified of research studies examine two-parameter equations, since the
adsorption models) [161,213]. complexity of three parameter equations limits their application.
Table 15 [11,40,42,44,47–49,52–54,56,58–65,68,80–83,86,92, Perić et al. [59] investigated the performance of various isotherms
98–100,102–108,120,122–125,127,142–145,149,151–153,158,166, (Langmuir–Freundlich, Toth, DR, modified DR, Redlich–Peterson
174,175,210,219,251–253] summarizes the equilibrium studies and Lineweawer–Burk) for the adsorption of zinc on natural zeolite
concerning adsorption of nickel and zinc ions on natural and and found that the best fit to the experimental results was provided
modified minerals in single and multi-metal systems. The Table by the models that assume that the metal ions bind first at the
lists the experimental conditions, the maximum adsorption energetically most favourable sites, with multi-layer adsorption
capacity and the isotherm model that provides the best fit to taking place subsequently. Rajic et al. [120] found that the Sips
the experimental data. Usually, the Langmuir and the Freundlich isotherm best described the equilibrium data for nickel adsorp-
models are the more frequently applied models. Oter and Akcay tion on natural zeolite for a wide range of temperatures since
[54] found that the Langmuir and the Freundlich isotherm models it accounts for heterogeneity. Stojakovic et al. [125] also found
could adequately describe the adsorption of nickel and zinc ions a good correlation between the experimental data and the Sips
on Turkish clinoptilolite. El-Kamash et al. [124] mentioned that isotherm for zinc adsorption on natural clinoptilolite. However,
both Freundlich and Dubinin–Radushkevich (DR) models provided the authors concluded that the Langmuir model could describe
adequate fit for the adsorption of zinc on synthetic zeolite A, while better the experimental data since it gave a better prediction of
the values of the mean energy of adsorption E were in the range the maximum adsorption capacity compared to the Sips model.
of 8–16 kJ mol−1 , indicating that ion exchange was the prevailing The latter model also could not adequately describe the changes
mechanism. Other researchers have also reported values of E in the in qm due to temperature increase. Katsou et al. [96] examined
range of 8–16 kJ mol−1 [62,64,68,107]. Donat et al. [100] found that several equilibrium models (Langmuir, Freundlich, Temkin, DR)
the Langmuir model provided the best fit to the experimental data, and concluded that the Langmuir isotherm provided the best fit
while the application of the DR model showed that E was in the to the experimental data concerning the adsorption of nickel on
range of ion exchange (8–16 kJ mol−1 ). Other equilibrium studies zeolite, bentonite and vermiculite. Meshko et al. [210] studied the
have shown that adsorption is physical with E taking values lower adsorption isotherms of Langmuir, Freundlich, Redlich–Peterson
456 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

and Langmuir–Freundlich of zinc ions on clinoptilolite and con- ai parameter that depends on the shape of the adsorbent
cluded that the Langmuir–Freundlich equation provided the am specific area of the adsorbent (m2 m−3 )
best fit to the experimental data. Finally, Ijagbemi et al. [56] aS Sips isotherm constant (L g−1 )
applied the isotherm models of Langmuir, Freundlich, Temkin, AT Toth isotherm constant (L mg−1 )
Redlich–Peterson, Florry–Huggins and DR to describe the equilib- B time constant
rium sorption of nickel on Na-montmorillonite and acid modified b constant related to the heat of adsorption in the Temkin
montmorillonite and concluded that Redlich–Peterson provided model (J mol−1 )
the best fit to the experimental data for Na-montmorillonite and BBET BET adsorption constant related to the energy of interac-
DR for acid treated montmorillonite. As it is seen from Table 13, the tion with the surface
maximum adsorption capacity concluded by the different studies bi parameter that depends on the shape of the adsorbent
for the same adsorbent and metal has significant differences and bk Khan isotherm constant
depends on various parameters. These parameters include solu- Ce liquid phase equilibrium metal concentration (mg L−1 )
tion pH, temperature, adsorbent’s concentration, grain size, initial Ce,i equilibrium concentration of solute i
metal concentrations, the adsorbent’s origin and its composition, Cs metal concentration at the surface of the adsorbent
the application of modified or natural adsorbents, the presence (mg L−1 )
of other cations, the liquid phase characteristics, the isotherm Csb saturation concentration of the metal corresponding to
employed to determine qm . Usually, the presence of other cations monolayer saturation (mg L−1 )
in the solution results in the reduction of qm compared to the one Ct liquid phase metal concentration at time t (mg L−1 )
obtained for each metal in single metal solutions [86,108,174]. C0 initial metal concentration in the solution (mg L−1 )
Furthermore, in most studies the system’s equilibrium is examined Deff effective diffusion coefficient (m2 s−1 )
using the linear forms of the isotherm models. The application Df film diffusion coefficient (m2 s−1 )
of the non-linear forms can provide different estimated for the Dp pore diffusion coefficient (m2 s−1 )
maximum adsorption capacity. Currently, there is no comparative dp average particle size of adsorbent (m)
study applying both linear and non-linear two-parameter isotherm E mean energy of sorption per mole of adsorbate for the DR
equations for the adsorption of Ni and Zn on minerals. model (kJ mol−1 )
Ea Arrhenius activation energy (kJ mol−1 )
8. Conclusions F fractional attainment of equilibrium at time t
f interaction coefficient of the Frumkin equation
Adsorption and ion exchange using natural and modified min- K saturation constant of Generalized isotherm (mg L−1 )
erals is effective for the removal of nickel and zinc ions from water ka adsorption rate constant
and wastewater. Extensive studies have been conducted in order to Keq equilibrium adsorption constant
evaluate the impact of several parameters on adsorption. The above KF Freundlich constant (mg1−1/n L1/n g−1 )
analysis has shown that the adsorbate type, solution pH, metal con- kf mass transfer coefficient between bulk solution and solid
centration and ionic strength as well as the adsorbent type and surface (m min−1 )
concentration are usually the most influential parameters. Solution kfd rate of liquid film diffusion (min−1 )
temperature, mineral grain size and agitation speed also seem to KFH Flory-Huggins equilibrium constant
affect the process. The presence of other ions in the solution and/or KFF equilibrium constant of Frumkin equation
certain wastewater compounds adversely impact on the process. kid constant of intraparticle diffusion (mg g−1 min−1/2 )
The chemical and/or thermal modification of natural minerals KL Langmuir constant (L mg−1 or L mol−1 )
can increase its adsorption capacity. However, mineral modifica- KLF Langmuir-Freundlich isotherm constant (L mg−1 )
tion is not always associated with increased metal uptake since KL,i Langmuir constant for single solute i (L mg−1 )
some cases have been reported where modified minerals exhibited KR Redlich–Peterson isotherm constant (L g−1 )
an inferior performance compared to natural ones. The influence KR,i Redlich–Peterson isotherm constant derived for a single
of a wide pool of parameters on adsorption is also reflected by solute i (L g−1 )
the variability in the mineral adsorption capacities for zinc and KT Toth isotherm constant (L mg−1 )
nickel uptake in the summarized scientific works. Several stud- k1 rate constant of the pseudo-first-order model (min−1 )
ies use the pseudo-first-order and pseudo-second-order kinetics k2 rate constant of the pseudo-second-order model
in order to model the adsorption kinetics, with most studies con- (g mg−1 min−1 )
cluding that the pseudo-second-order model is in close agreement ma adsorbent concentration (g L−1 )
with experimental data. Diffusion models such as the intraparticle N number of data points
diffusion model, external mass transfer and Vermeulen’s approxi- n ascending integer number
mation have also been applied and the effective diffusion coefficient nF Freundlich affinity constant
has been determined. The latter usually lies within the range of nFH Flory–Huggins isotherm exponent
10−12 –10−14 m2 s−1 . In terms of equilibrium examination the Lang- ng cooperative binding constant of Generalized isotherm
muir isotherm model is in several cases in good agreement with the ni Freundlich affinity constant for a single solute i
experimental data, while the Freundlich model has also been found nK Khan isotherm exponent
to fit the experimental data in several cases. Some researchers nLF Langmuir–Freundlich isotherm exponent
have also successfully applied other isotherm models, such as the ns Sips isotherm exponent
Temkin and DR models to predict the adsorption equilibrium. p Radke-Prausnitz isotherm exponent
qe solid phase equilibrium metal concentration (mg g−1 )
qe,i solid phase equilibrium concentration of solute i (mg g−1 )
Nomenclature qF theoretical monolayer saturation capacity (mg g−1 )
qm maximum adsorption capacity (mg g−1 )
qm,i maximum adsorption capacity for single solute i (mg g−1 )
A equilibrium binding constant of Temkin model (L mg−1 ) qt solid phase concentration of metal at time t (mg g−1 )
Aa pre-exponential factor, Arrhenius constant R universal gas constant (Jmol−1 K−1 )
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 457

r radius of the adsorbent particle considered to be spherical [10] D.W. O’Connell, C. Birkinshaw, T.F. O’Dwyer, Heavy metal adsorbents pre-
(m) pared from the modification of cellulose: a review, Bioresour. Technol. 99
(2008) 6709–6724.
rRP Radke–Prausnitz isotherm constant (L mg−1 ) [11] Y.S. Ok, J.E. Yang, Y.-S. Zhang, S.-J. Kim, D.-Y. Chung, Heavy metal adsorption by
T absolute temperature (K) a formulated zeolite-Portland cement mixture, J. Hazard. Mater. 147 (2007)
t time (min) 91–96.
[12] S.E. Bailey, T.J. Olin, R.M. Bricka, D.D. Adrian, A review of potentially low-cost
Ta Toth isotherm exponent sorbents for heavy metals, Water Res. 33 (1999) 2469–2479.
t1/2 half time for the adsorption process (s or min) [13] R.D. Harter, R. Naidu, An assessment of environmental and solution parameter
˛ initial adsorption rate of Elovich equation (mg g−1 min−1 ) impact on trace-metal sorption by soils, Soil Sci. Soc. Am. J. 65 (2001) 597–612.
[14] H.B. Bradl, Adsorption of heavy metal ions on soils and soils constituents, J.
˛R Redlich–Peterson isotherm constant (L mg−1 )
Colloid. Interface Sci. 277 (2004) 1–18.
˛R,i Redlich–Peterson isotherm isotherm constant for a single [15] G.S. Miguel, S.D. Lambert, N.J.D. Graham, A practical review of the perfor-
solute i (L mg−1 ) mance of organic and inorganic adsorbents for the treatment of contaminated
˛RP Radke-Prausnitz isotherm constant (L g−1 ) waters, J. Chem. Technol. Biotechnol. 81 (2006) 1685–1696.
[16] D. Sud, G. Mahajan, M.P. Kaur, Agricultural waste material as potential adsor-
ˇ constant of the DR model related to the mean energy of bent for sequestering heavy metal ions from aqueous solutions – a review,
sorption (mol2 kJ−2 ) Bioresour. Technol. 99 (2008) 6017–6027.
ˇe desorption constant of Elovich equation (g mg−1 ) [17] S.K.R. Yadanaparthi, D. Graybill, R.V. Wandruszka, Adsorbents for the removal
of arsenic, cadmium, and lead from contaminated waters, J. Hazard. Mater.
ˇR Redlich–Peterson isotherm exponent 171 (2009) 1–15.
ˇR,i Redlich–Peterson isotherm exponent for a single solute i [18] S. Wang, Y. Peng, Natural zeolites as effective adsorbents in water and waste-
ı bounbary layer thickness (m) water treatment, Chem. Eng. J. 156 (2010) 11–24.
[19] Y.-F. Zhou, R.J. Haynes, Sorption of heavy metals by inorganic and organic
G Gibbs free energy of adsorption (kJ mol−1 ) components of solid wastes: significance to use of wastes as low-cost adsor-
G◦ standard Gibbs free energy (kJ mol−1 ) bents and immobilizing agents, Crit. Rev. Environ. Sci. Technol. 40 (2010)
H enthalpy change (kJ mol−1 ) 909–977.
[20] Y.S. Ho, J.F. Porter, G. McKay, Equilibrium isotherm studies for the sorption
H◦ standard enthalpy change (kJ mol−1 ) of divalent metal ions onto peat: copper, nickel and lead single component
S entropy change (J mol−1 K−1 ) systems, Water Air Soil Pollut. 141 (2002) 1–33.
S◦ Standard entropy change (J mol−1 K-1 ) [21] Y.-S. Ho, Review of second-order models for adsorption systems, J. Hazard.
Mater. 136 (2006) 681–689.
ε Polanyi sorption potential of the DR model
[22] C. Gerente, V.K.C. Lee, P. Le-Cloirec, G. McKay, Application of chitosan for the
 degree of surface coverage of adsorbent surface removal of metals from wastewaters by adsorption – mechanisms and models
b adsorbent bulk density (g L−1 ) review, Crit. Rev. Environ. Sci. Technol. 37 (2007) 41–127.
[23] W. Plazinski, W. Rudzinski, A. Plazinska, Theoretical models of sorption kinet-
 the time needed for diffusion of the metal from the adsor-
ics including a surface reaction mechanism: a review, Adv. Colloid Interface
bent’s surface to the exchangeable site (s) Sci. 152 (2009) 2–13.
[24] H. Qiu, L. Lv, B.-c. Pan, Q.-j. Zhang, W.-m. Zhang, Q.-x. Zhang, Critical review
in adsorption kinetic models, J. Zhejiang Univ. Sci. A 10 (2009) 716–724.
Acknowledgements [25] B. Saha, C. Orvig, Biosorbents for hexavalent chromium elimination
from industrial and municipal effluents, Coord. Chem. Rev. 254 (2010)
2959–2972.
[26] O. Abollino, A. Giacomino, M. Malandrino, E. Mentasti, Interaction of metal
ions with montmorillonite and vermiculite, Appl. Clay Sci. 38 (2008) 227–236.
[27] M. Malandrino, O. Abollino, A. Giacomino, M. Aceto, E. Mentasti, Adsorption
of heavy metals on vermiculite: influence of pH and organic ligands, J. Colloid
Interface Sci. 299 (2006) 537–546.
[28] M.K. Doula, A. Ioannou, The effect of electrolyte anion on Cu adsorption-
This research has been co-financed by the European Union (Euro- desorption by clinoptilolite, Micropor. Mesopor. Mater. 58 (2003) 115–130.
pean Social Fund – ESF) and Greek national funds through the [29] P.W. Schindler, B. Fürst, R. Dick, P.U. Wolf, Ligand properties of surface silanol
Operational Program “Education and Lifelong Learning” of the groups. I. Surface complex formation with Fe3+ , Cu2+ , Cd2+ and Pb2+ , J. Colloid
Interface Sci. 55 (1976) 469–475.
National Strategic Reference Framework (NSRF) – Research Fund- [30] N.M. Nagy, J. Kónya, Interfacial Chemistry of Rocks and Soils, CRC Press, Taylor
ing Program: Heraclitus II. Investing in knowledge society through & Francis, New York, 2009.
the European Social Fund. [31] M. Hamidpour, M. Kalbasi, M. Afyuni, H. Shariatmadari, P.E. Holm, H.C.B.
Hansen, Sorption hysteresis of Cd(II) and Pb(II) on natural zeolite and ben-
tonite, J. Hazard. Mater. 181 (2010) 686–691.
[32] W.A. Deer, R.A. Howie, J. Zussman, An Introduction to the Rock-Forming Min-
References erals, Wesley Longman, Hong Kong, 1992, pp. 520–529.
[33] S. Kulprathipanja, Zeolites in Industrial Separation and Catalysis, Wiley-VCH
[1] R.P. Beliles, The lesser metals, in: F.W. Oehme (Ed.), Toxicity of Heavy Metals Verlag GmbH & Co KGaA, Weinheim, 2010.
in the Environment, Part 2, Marcel Dekker, New York, 1978, pp. 547–616. [34] S. Cornelius, J.R. Hurlbut, Mineral, John Wiley & Sons Inc, London, 1963.
[2] N. Akhtar, J. Iqbal, M. Iqbal, Removal and recovery of nickel(II) from aque- [35] K.G. Bhattacharyya, S.S. Gupta, Adsorption of a few heavy metals on natural
ous solution by loofa sponge-immobilized biomass of Chlorella sorokiniana: and modified kaolinite and montmorillonite: a review, Adv. Colloid Interface
characterization studies, J. Hazard. Mater. B 108 (2004) 85–94. Sci. 140 (2008) 114–131.
[3] U. Farooq, J.A. Kozinski, M.A. Khan, M. Athar, Biosorption of heavy metal ions [36] M. Fleischer, J.A. Mandarino, Glossary of Mineral Species, The Mineralogical
using wheat based biosorbents – a review of the recent literature, Bioresour. Record Inc, Tuscon, 1991, p. 27.
Technol. 101 (2010) 5043–5053. [37] M.J. Potter, Vermiculite, US Geological Survey, Minerals Yearbook, 2002, pp.
[4] P.N. Cheremisinoff, Handbook of Water and Wastewater Treatment Technol- 82.1–82.3.
ogy, Marcel Dekker Inc, New York, 1995. [38] C.E. Weaver, Developments in Sedimentology 44, Clays, Muds and Shales,
[5] T. Zoumis, W. Calmano, U. Förstner, Demobilization of heavy metals from Elsevier, Atlanta, 1989, pp. 82–84.
mine wastewaters, Acta Hydrochim. Hydrobiol. 28 (2000) 212–218. [39] G. Xueyi, K. Inoue, Elution of copper from vermiculite with environmentally
[6] A. Dimirkou, Uptake of Zn2+ ions by a fully iron-exchanged clinoptilolite. Case benign reagents, Hydrometallurgy 70 (2003) 9–21.
study of heavily contaminated drinking water samples, Water Res. 41 (2007) [40] S. Çoruh, O.N. Ergun, Ni2+ removal from aqueous solutions using conditioned
2763–2773. clinoptilolites: kinetic and isotherm studies, Environ. Prog. Sustain. Energy
[7] T.A. Kurniawan, G.Y.S. Chan, W.H. Lo, S. Babel, Comparisons of low-cost adsor- 28 (2009) 162–172.
bents for treating wastewaters laden with heavy metals, Sci. Total Environ. [41] B. Erdem, A. Özcan, O. Gök, A.S. Özcan, Immobilization of 2,2 -dipyridyl onto
366 (2006) 409–426. bentonite and its adsorption behavior of copper(II) ions, J. Hazard. Mater. 163
[8] T.A. Kurniawan, G.Y.S. Chan, W.H. Lo, S. Babel, Physico-chemical treatment (2009) 418–426.
techniques for wastewater laden with heavy metals, Chem. Eng. J. 118 (2006) [42] C.O. Ijagbemi, M.-H. Baek, D.-S. Kim, Montmorillonite surface properties and
83–98. sorption characteristics for heavy metal removal from aqueous solutions, J.
[9] V.O. Arief, K. Trilestari, J. Sunarso, N. Indraswati, S. Ismadji, Recent progress Hazard. Mater. 166 (2009) 538–546.
on biosorption of heavy metals from liquids using low cost biosorbents: char- [43] D. Wu, Y. Sui, S. He, X. Wang, C. Li, H. Kong, Removal of trivalent chromium
acterization, biosorption parameters and mechanism studies, Clean Soil Air from aqueous solution by zeolite synthesized from coal fly ash, J. Hazard.
Water 36 (2008) 937–962. Mater. 155 (2008) 415–423.
458 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

[44] S.S. Gupta, K.G. Bhattacharyya, Immobilization of Pb(II), Cd(II) and Ni(II) ions [77] V.J. Inglezakis, A.A. Zorpas, M.D. Loizidou, H.P. Grigoropoulou, The effect of
on kaolinite and montmorillonite surfaces from aqueous medium, J. Environ. competitive cations and anions on ion exchange of heavy metals, Sep. Purif.
Manag. 87 (2008) 46–58. Technol. 46 (2005) 202–207.
[45] K. Vijayaraghavan, Y.-S. Yun, Bacterial biosorbents and biosorption, Bio- [78] V.J. Inglezakis, A.A. Zorpas, M.D. Loizidou, H.P. Grigoropoulou, Simultaneous
technol. Adv. 26 (2008) 266–291. removal of metals Cu2+ , Fe3+ and Cr3+ with anions SO4 2− and HPO4 2− using
[46] M.K. Doula, A. Dimirkou, Use of an iron-overexchanged clinoptilolite for the clinoptilolite, Micropor. Mesopor. Mater. 61 (2003) 167–171.
removal of Cu2+ ions from heavily contaminated drinking water samples, J. [79] P. Papachristou, K.J. Haralambous, M. Loizidou, N. Spyrellis, Studies on the
Hazard. Mater. 151 (2008) 738–745. nickel removal from aqueous solutions, J. Environ. Sci. Health A Environ. Sci.
[47] S.S. Gupta, K.G. Bhattacharyya, Adsorption of Ni(II) on clays, J. Colloid Interface Eng. 28 (1993) 135–142.
Sci. 295 (2006) 21–32. [80] T.S. Anirudhan, P.S. Suchithra, Heavy metals uptake from aqueous solutions
[48] S. Kocaoba, Y. Orhan, T. Akyüz, Kinetics and equilibrium studies of heavy metal and industrial wastewaters by humic acid-immobilized polymer/bentonite
ions removal by use of natural zeolite, Desalination 214 (2007) 1–10. composite: kinetics and equilibrium modelling, Chem. Eng. J. 156 (2011)
[49] M.G.A. Vieira, A.F.A. Neto, M.L. Gimenes, M.G.C. da Silva, Sorption kinetics and 146–156.
equilibrium for the removal of nickel ions from aqueous phase on calcined [81] E. Álvarez-Ayuso, A. García-Sánchez, X. Querol, Purification of metal electro-
Bofe bentonite clay, J. Hazard. Mater. 177 (2010) 362–371. plating waste waters using zeolites, Water Res. 37 (2003) 4855–4862.
[50] S.K. Ouki, M. Kavannagh, Performance of natural zeolites for the treatment of [82] E. Álvarez-Ayuso, A. García-Sánchez, Removal of heavy metals from waste
mixed metal-contaminated effluents, Waste Manag. Res. 15 (1997) 383–394. waters by natural and Na-exchanged bentonites, Clays Clay Miner. 51 (2003)
[51] S.K. Ouki, M. Kavannagh, Treatment of metals-contaminated wastewaters by 475–480.
use of natural zeolites, Water Sci. Technol. 69 (1999) 115–122. [83] E. Álvarez-Ayuso, A. García-Sánchez, Removal of heavy metals from waste
[52] M. Sprynskyy, B. Buszewski, A.P. Terzyk, J. Namieśnik, Study of the selec- waters by vermiculites, Environ. Technol. 24 (2003) 5615–5625.
tion mechanism of heavy metal (Pb2+ , Cu2+ , Ni2+ , and Cd2+ ) adsorption on [84] S.K. Pitcher, R.C.T. Slade, N.I. Ward, Heavy metal removal from motorway
clinoptilolite, J. Colloid Interface Sci. 304 (2006) 21–28. stormwater using zeolites, Sci. Total Environ. 334 (2004) 161–166.
[53] A. Al-Haj Ali, R. El-Bishtawi, Removal of lead and nickel ions using zeolite tuff, [85] E. Katsou, S. Malamis, K.J. Haralambous, Industrial wastewater pre-treatment
J. Chem. Technol. Biotechnol. 69 (1997) 27–34. for heavy metal reduction by employing a sorbent-assisted ultrafiltration
[54] O. Oter, H. Akcay, Use of natural clinoptilolite to improve, water quality: sorp- system, Chemosphere 82 (2011) 557–564.
tion and selectivity studies of lead(II), copper(II), zinc(II), and nickel(II), Water [86] T. Vengris, R. Binkiene, A. Sveikauskaite, Nickel, copper and zinc removal from
Environ. Res. 79 (2007) 329–335. waste water by a modified clay sorbent, Appl. Clay Sci. 18 (2001) 183–190.
[55] C. Cabrera, C. Gabaldón, P. Marzal, Sorption characteristics of heavy metal ions [87] M. Panayotova, Kinetics and thermodynamics of removal of nickel ions from
by a natural zeolite, J. Chem. Technol. Biotechnol. 80 (2005) 477–481. wastewater by use of natural and modified zeolite, Fresenius Environ. Bull.
[56] C.O. Ijagbemi, M.-H. Baek, D.-S. Kim, Adsorptive performance of un-calcined 10 (2001) 267–272.
sodium exchanged and acid modified montmorillonite for Ni2+ removal: equi- [88] T.G.P. Vásquez, A.E.C. Botero, L.M.S. de Mesquita, M.L. Torem, Biosorptive
librium, kinetics, thermodynamics and regeneration studies, J. Hazard. Mater. removal of Cd and Zn from liquid streams with a Rhodococcus opacus strain,
174 (2010) 746–755. Minerals Eng. 20 (2007) 939–944.
[57] X. Wu, F. Zhao, M. Chen, Y. Zhang, C. Zhao, H. Zhou, Factors affecting the [89] J.E.B. Cayllahua, R.J. Carvalho, M.L. Torem, Evaluation of equilibrium, kinetic
adsorption of Zn2+ and Cd2+ ions from aqueous solution onto vermiculite, and thermodynamic parameters for biosorption of nickel(II) ions onto bacte-
Adsorpt. Sci. Technol. 26 (2008) 145–155. ria strain Rhodococcus opacus, Miner. Eng. 22 (2009) 1318–1325.
[58] A. Kaya, A.H. Ören, Adsorption of zinc from aqueous solutions to bentonite, J. [90] C.Y. Lin, D.H. Yang, Removal of pollutants from wastewater by coal bottom
Hazard. Mater. 125 (2005) 183–189. ash, J. Environ. Sci. Health A Tox. Hazard. Subst. Environ. Eng. 37 (2002)
[59] J. Perić, M. Trgo, N.V. Medvidović, Removal of zinc, copper and lead by nat- 1509–1522.
ural zeolite – a comparison of adsorption isotherms, Water Res. 38 (2004) [91] M.I. Ansari, F. Masood, A. Malik, Bacterial biosorption: a technique for reme-
1893–1899. diation of heavy metals in microbes and microbial technology, in: I. Ahmad,
[60] A.H. Ören, A. Kaya, Factors affecting adsorption characteristics of Zn2+ on two F. Ahmad, J. Pichtel (Eds.), Agricultural Environmental Applications, Springer,
natural zeolites, J. Hazard. Mater. 131 (2006) 59–65. New York, 2011.
[61] S. Veli, B. Alyüz, Adsorption of copper and zinc from aqueous solutions by [92] K.S. Hui, C.Y.H. Chao, S.C. Kot, Removal of mixed heavy metal ions in waste-
using natural clay, J. Hazard. Mater. 149 (2007) 226–233. water by zeolite 4A and residual products from recycled coal fly ash, J. Hazard.
[62] E. Erdem, N. Karapinar, R. Donat, The removal of heavy metal cations by Mater. 127 (2005) 89–101.
natural zeolites, J. Colloid Interface Sci. 280 (2004) 309–314. [93] A. Esposito, F. Pagnanelli, F. Vegliò, pH-related equilibria models for biosorp-
[63] T.K. Sen, D. Gomez, Adsorption of zinc (Zn2+ ) from aqueous solution on natural tion in single metal systems, Chem. Eng. Sci. 57 (2002) 307–313.
bentonite, Desalination 267 (2011) 286–294. [94] J. Yang, B. Volesky, Modeling the uranium-proton ion exchange in biosorption,
[64] H.M. Baker, A.M. Massadeh, H.A. Younes, Natural Jordanian zeolite: removal Environ. Sci. Technol. 33 (1999) 4079–4085.
of heavy metal ions from water samples using column and batch methods, [95] M.A. Acheampong, R.J.W. Meulepas, P.N.L. Lens, Removal of heavy metals and
Environ. Monit. Assess. 157 (2009) 319–330. cyanide from gold mine wastewater, J. Chem. Technol. Biotechnol. 85 (2010)
[65] N. Beyazit, I. Peker, O.N. Ergun, Removal of lead and zinc ions from aqueous 590–613.
solution using Amasya zeolites from Turkey, Int. J. Environ. Pollut. 19 (2003) [96] E. Katsou, S. Malamis, K.J. Haralambous, M. Loizidou, Use of ultrafiltration
160–170. membranes and aluminosilicate minerals for nickel removal from industrial
[66] A.C. Vieira dos Santos, J.C. Masini, Evaluating the removal of Cd(II), Pb(II) and wastewater, J. Membr. Sci. 360 (2010) 234–249.
Cu(II) from a wastewater sample of a coating industry by adsorption onto [97] E. Katsou, S. Malamis, K.J. Haralambous, Examination of zinc uptake in a com-
vermiculite, Appl. Clay Sci. 37 (2007) 167–174. bined system using sludge, minerals and ultrafiltration membranes, J. Hazard.
[67] A. Katsoyiannis, C. Samara, The fate of dissolved organic carbon (DOC) in the Mater. 182 (2010) 27–38.
wastewater treatment process and its importance in the removal of waste- [98] M.E. Argun, Use of clinoptilolite for the removal of nickel ions from water:
water contaminants, Environ. Sci. Pollut. Res. Int. 14 (2007) 284–292. kinetics and thermodynamics, J. Hazard. Mater. 150 (2008) 587–595.
[68] S. Yang, J. Li, Y. Lu, Y. Chen, X. Wang, Sorption of Ni(II) on GMZ bentonite: [99] M.V. Dinu, E.S. Dragan, Evaluation of Cu2+ , Co2+ and Ni2+ ions removal from
effects of pH, ionic strength, foreign ions, humic acid and temperature, Appl. aqueous solution using a novel chitosan/clinoptilolite composite: kinetics and
Radiat. Isot. 67 (2009) 1600–1608. isotherms, Chem. Eng. J. 160 (2010) 157–163.
[69] A. Langella, M. Pansini, P. Cappelletti, B. De Gennaro, M. De Gennaro, C. Colella, [100] R. Donat, A. Akdogan, E. Erdem, H. Cetisli, Thermodynamics of Pb2+ and Ni2+
NH4 + , Cu2+ , Zn2+ , Cd2+ and Pb2+ exchange for Na+ in a sedimentary clinoptilo- adsorption onto natural bentonite from aqueous solutions, J. Colloid Interface
lite, North Sardinia, Italy, Micropor. Mesopor. Mater. 37 (2000) 337–343. Sci. 286 (2005) 43–52.
[70] D. Ziolko, D. Hala, J.N. Lester, M.D. Scrimshaw, The effectiveness of conven- [101] Z.-R. Liu, S.-Q. Zhou, Adsorption of copper and nickel on Na-bentonite, Process
tional trickling filter treatment plants at reducing concentrations of copper Saf. Environ. Prot. 88 (2010) 62–66.
in wastewaters, Sci. Total Environ. 407 (2009) 6235–6241. [102] X. Song, S. Wang, L. Chen, M. Zhang, Y. Dong, Effect of pH, ionic strength
[71] S. Kempton, R.M. Sterritt, J.N. Lester, Heavy metal removal in primary sedi- and temperature on the sorption of radionickel on Na-montmorillonite, Appl.
mentation. II. The influence of metal speciation and particle size distribution, Radiat. Isot. 67 (2009) 1007–1012.
Sci. Total Environ. 63 (1987) 247–258. [103] M.G. Da Fonseca, M.M. De Oliveira, L.N.H. Arakaki, J.G.P. Espinola, C. Airoldi,
[72] M. Karvelas, A. Katsoyiannis, C. Samara, Occurrence and fate of heavy metals Natural vermiculite as an exchanger support for heavy cations in aqueous
in the wastewater treatment process, Chemosphere 53 (2003) 1201–1210. solution, J. Colloid Interface Sci. 285 (2005) 50–55.
[73] K. Lackovic, M.J. Angove, J.D. Wells, B.B. Johnson, Modelling the adsorption of [104] M. Minceva, R. Fajgar, L. Markovska, V. Meshko, Comparative study of Zn2+ ,
Cd(II) onto goethite in the presence of citric acid, J. Colloid Inteface Sci. 269 Cd2+ , and Pb2+ removal from water solution using natural clinoptilolitic zeo-
(2004) 37–45. lite and commercial granulated activated carbon. Equilibrium of adsorption,
[74] O. Abollino, M. Aceto, M. Malandrino, C. Sarzanini, E. Mentasti, Adsorption of Sep. Sci. Technol. 43 (2008) 2117–2143.
heavy metals on Na-montmorillonite. Effect of pH and organic substances, [105] G. Purna Chandra Rao, S. Satyaveni, A. Ramesh, K. Seshaiah, K.S.N. Murthy, N.V.
Water Res. 37 (2003) 1619–1627. Choudary, Sorption of cadmium and zinc from aqueous solutions by zeolite
[75] G.M. Naja, V. Murphy, B. Volesky, Biosorption, metals, in: Encyclopedia of 4A, zeolite 13X and bentonite, J. Environ. Manag. 81 (2006) 265–272.
Industrial Biotechnology: Bioprocess, Bioseparation, and Cell Technology, [106] A. Mockovčiaková, Z. Orolínová, J. Škvarla, Enhancement of the bentonite
Wiley, 2010. sorption properties, J. Hazard. Mater. 180 (2010) 274–281.
[76] B. Volesky, S. Schiewer, Biosorption of metals, in: M. Flickinger, S.W. Drew [107] S.-H. Lin, R.-S. Juang, Heavy metal removal from water by sorption using
(Eds.), Encycl. of Bioprocess Technol., Wiley, New York, 1999, pp. 433–453. surfactant-modified montmorillonite, J. Hazard. Mater. 92 (2002) 315–326.
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 459

[108] P.C. Mishra, R.K. Patel, Removal of lead and zinc ions from water by low cost [139] A. Shukla, Y.-H. Zhang, P. Dubey, J.L. Margrave, S.S. Shukla, The role of sawdust
adsorbents, J. Hazard. Mater. 168 (2009) 319–325. in the removal of unwanted materials from water, J. Hazard. Mater. 95 (2002)
[109] B. Bayat, Comparative study of adsorption properties of Turkish fly ashes: 137–152.
I. The case of nickel(II), copper(II) and zinc(II), J. Hazard. Mater. 95 (2002) [140] F. Bordas, A. Bourg, Effect of solid/liquid ratio on the remobilization of Cu,
251–273. Pb, Cd and Zn from polluted river sediment, Water Air Soil Pollut. 128 (2001)
[110] L. Nouri, I. Ghodbane, O. Hamdaoui, M. Chiha, Batch sorption dynamics and 391–400.
equilibrium for the removal of cadmium ions from aqueous phase using wheat [141] A.A. Türkman, S. Aslan, I. Ege, Treatment of metal containing wastewaters by
bran, J. Hazard. Mater. 149 (2007) 115–125. natural zeolites, Fresenius Environ. Bull. 13 (2004) 574–580.
[111] N. Goyal, S.C. Jain, U.C. Banerjee, Comparative studies on the microbial adsorp- [142] K.G. Bhattacharyya, S.S. Gupta, Uptake of Ni(II) ions from aqueous solution by
tion of heavy metals, Adv. Environ. Res. 7 (2003) 311–319. kaolinite and montmorillonite: influence of acid activation of the clays, Sep.
[112] P.R. Puranik, K.M. Paknikar, Biosorption of lead, cadmium and zinc by Cit- Sci. Technol. 43 (2008) 3221–3250.
robacter strain MCM B-181: characterization studies, Biotechnol. Prog. 15 [143] K.G. Bhattacharyya, S.S. Gupta, Influence of acid activation on adsorption of
(1999) 228–237. Ni(II) and Cu(II) on kaolinite and montmorillonite: kinetic and thermody-
[113] G. Crini, P.-M. Badot, Application of chitosan, a natural aminopolysaccha- namic study, Chem. Eng. J. 136 (2008) 1–13.
ride, for dye removal from aqueous solutions by adsorption processes using [144] K.G. Bhattacharyya, S.S. Gupta, Calcined tetrabutylammonium kaolinite and
batch studies: a review of recent literature, Prog. Polym. Sci. 33 (2008) montmorillonite and adsorption of Fe(II), Co(II) and Ni(II) from solution, Appl.
399–447. Clay Sci. 46 (2009) 216–221.
[114] E. Malkoc, Ni(II) removal from aqueous solutions using cone biomass of Thuja [145] K.G. Bhattacharyya, S.S. Gupta, Kaolinite and montmorillonite as adsorbents
orientalis, J. Hazard. Mater. 137 (2006) 899–908. for Fe(III), Co(II) and Ni(II) in aqueous medium, Appl. Clay Sci. 41 (2008) 1–9.
[115] V.J. Inglezakis, M.M. Loizidou, H.P. Grigoropoulou, Ion exchange studies on [146] J.-F. Blais, S. Shen, N. Meunier, R.D. Tyagi, Comparison of natural adsor-
natural and modified zeolites and the concept of exchange site accessibility, bents for metal removal from acidic effluent, Environ. Technol. 24 (2003)
J. Colloid Inteface Sci. 275 (2004) 570–576. 205–215.
[116] D. Park, Y.-S. Yun, J.M. Park, The past, present, and future trends of biosorption, [147] A. Cincotti, A. Mameli, A.M. Locci, R. Orrù, G. Cao, Heavy metals uptake by
Biotechnol. Bioprocess Eng. 15 (2010) 86–102. Sardinian natural zeolites: experiment and modelling, Ind. Eng. Chem. Res.
[117] A. Sari, M. Tuzen, Kinetic and equilibrium studies of biosorption of Pb(II) and 45 (2006) 1074–1084.
Cd(II) from aqueous solution by macrofungus (Amanita rubescens) biomass, J. [148] K. Athanasiadis, B. Helmreich, Influence of chemical conditioning on the ion
Hazard. Mater. 164 (2009) 1004–1011. exchange capacity and on kinetic of zinc uptake by clinoptilolite, Water Res.
[118] G. Uslu, M. Tanyol, Equilibrium and thermodynamic parameters of single and 39 (2005) 1527–1532.
binary mixture biosorption of lead(II) and copper(II) ions onto Pseudomonas [149] S.C. Stefanović, N.Z. Logar, K. Margeta, N.N. Tušar, I. Arčon, K. Maver, J. Kovač,
putida: effect of temperature, J. Hazard. Mater. 135 (2006) 87–93. V. Kaučič, Structural investigation of Zn2+ sorption on clinoptilolite tuff from
[119] Y. Sağ, T. Kutsal, Determination of the biosorption heats of heavy metal the Vranjska Banja deposit in Serbia, Micropor. Mesopor. Mater. 105 (2007)
ions on Zoogloea ramigera and Rhizopus arrhizus, Biochem. Eng. J. 6 (2000) 251–259.
145–151. [150] A. Bujnova, J. Lesny, Sorption characteristics of zinc and cadmium by some
[120] N. Rajic, D. Stojakovic, M. Jovanovic, N.Z. Logar, M. Mazaj, V. Kaucic, Removal natural, modified, and synthetic zeolites, Hung. Electron. J. Sci., Environ. Eng.
of nickel(II) ions from aqueous solutions using the natural clinoptilolite and (2004) 1–10.
preparation of nano-NiO on the exhausted clinoptilolite, Appl. Surf. Sci. 257 [151] A.S. Sheta, A.M. Falatah, M.S. Al-Sewailem, E.M. Khaled, A.S.H. Sallam, Sorption
(2010) 1524–1532. characteristics of zinc and iron by natural zeolite and bentonite, Micropor.
[121] M.A. Keane, The removal of copper and nickel from aqueous solution using Y Mesopor. Mater. 61 (2003) 127–136.
zeolite ion exchangers, Colloids Surf. A: Physicochem. Eng. Aspects 138 (1998) [152] A. Cincotti, N. Lai, R. Orrú, G. Cao, Sardinian natural clinoptilolites for heavy
11–20. metals and ammonium removal: experimental and modelling, Chem. Eng. J.
[122] S.S. Tahir, N. Rauf, Thermodynamic studies of Ni(II) adsorption onto bentonite 84 (2001) 275–282.
from aqueous solution, J. Chem. Thermodyn. 35 (2003) 2003–2009. [153] E.G. Pradas, M.V. Sanchez, F.C. Cruz, M.S. Viciana, M.F. Perez, Adsorption of
[123] N.M. Alandis, O.A. Aldayel, W.K. Mekhemer, J.A. Hefne, H.A. Jokhab, Thermo- cadmium and zinc from aqueous solution on natural and activated bentonite,
dynamic and kinetic studies for the adsorption of Fe(III) and Ni(II) ions from J. Chem. Technol. Biotechnol. 59 (1994) 289–295.
aqueous solution using natural bentonite, J. Dispers. Sci. Technol. 31 (2010) [154] B.I. Olu-Owolabi, E.I. Unuabonah, Kinetic and thermodynamics of the removal
1526–1534. of Zn2+ and Cu2+ from aqueous solution by sulphate and phosphate-modified
[124] A.M. El-Kamash, A.A. Zaki, M.A.E. Geleel, Modelling batch kinetics and ther- bentonite clay, J. Hazard. Mater. 184 (2010) 731–738.
modynamics of zinc and cadmium ions removal from waste solutions using [155] U.K. Saha, K. Iwasaki, K. Sakurai, Desorption behavior of Cd, Zn and Pb
synthetic zeolite A, J. Hazard. Mater. 127 (2005) 211–220. sorbed on hydroxyaluminum and hydroxyaluminosilicate-montmorillonite
[125] D. Stojakovic, J. Hrenovic, M. Mazaj, N. Rajic, On the zinc sorption by the Ser- complexes, Clays Clay Miner. 51 (2003) 481–492.
bian natural clinoptilolite and the disinfecting ability and phosphate affinity [156] L. Ćurković, S. Cerjan-Stefanović, T. Filipan, Metal ion exchange by natural and
of the exhausted sorbent, J. Hazard. Mater. 185 (2011) 408–415. modified zeolites, Water Res. 31 (1997) 1379–1382.
[126] G. Bereket, A.Z. Aroğuz, M.Z. Özel, Removal of Pb(II), Cd(II), Cu(II), and Zn(II) [157] M.V. Mier, R.L. Callejas, R. Gehr, B.E.J. Cisneros, P.J.J. Alvarez, Heavy metal
from aqueous solutions by adsorption on bentonite, J. Colloid Interface Sci. removal with Mexician clinoptilolite: multi-component ionic exchange,
187 (1997) 338–343. Water Res. 35 (2001) 373–378.
[127] A. Mellah, S. Chegrouche, The removal of zinc from aqueous solutions by [158] S. Cerjan-Stefanović, L. Ćurković, T. Filipan, Metal ion exchange by natural
natural bentonite, Water Res. 31 (1997) 621–629. zeolites, Croat. Chem. Acta 69 (1996) 281–290.
[128] S.T. Bosso, J. Enzweiler, Evaluation of heavy metal removal from aqueous [159] M.J. Zamzow, J.E. Murphy, Removal of metal cations from water using zeolites,
solution onto scolecite, Water Res. 36 (2002) 4795–4800. Sep. Sci. Technol. 27 (1992) 1969–1984.
[129] U. Wingenfelder, C. Hansen, G. Furrer, R. Schulin, Removal of heavy metals [160] S. Malamis, E. Katsou, K.J. Haralambous, Study of Ni(II), Cu(II), Pb(II), and Zn(II)
from mine waters by natural zeolites, Environ. Sci. Technol. 39 (2005) removal using sludge and minerals followed by MF/UF, Water Air Soil Pollut.
4606–4613. 218 (2011) 81–92.
[130] V.J. Inglezakis, S.G. Poulopoulos, Adsorption, Ion Exchange and Catalysis: [161] Z. Aksu, U. Acikel, E. Kabasakal, S. Tezer, Equilibrium modelling of individ-
Design of Operation and Environmental Applications, Elsevier, UK, 2006. ual and simultaneous biosorption of chromium(VI) and nickel(II) onto dried
[131] E. Malliou, M. Loizidou, N. Spyrellis, Uptake of lead and cadmium by clinop- activated sludge, Water Res. 36 (2002) 3063–3073.
tilolite, Sci. Total Environ. 149 (1994) 139–144. [162] D. Mohan, C.U. Pittman, P.H. Steele, Single binary and multi-component
[132] A. Esposito, F. Pagnanelli, A. Lodi, C. Solisio, F. Vegliò, Biosorption of heavy adsorption of copper and cadmium from aqueous solutions on Kraft lignin
metals by Sphaerotilus natans: an equilibrium study at different pH and – a biosorbent, J. Colloid Interface Sci. 297 (2006) 489–504.
biomass concentrations, Hydrometallurgy 60 (2001) 129–141. [163] M. Panayotova, B. Velikov, Influence of zeolite transformation in a homoionic
[133] J. Tangaromsuk, P. Pokethitiyook, M. Kruatrachue, E.S. Upatham, Cadmium form on the removal of some heavy metal ions from wastewater, J. Environ.
biosorption by Sphingomonas paucimobilis biomass, Bioresour. Technol. 85 Sci. Health A Tox. Hazard. Subst. Environ. Eng. 38 (2003) 545–554.
(2002) 103–105. [164] C.M. Futalan, C.-C. Kan, M.L. Dalida, K.-J. Hsien, C. Pascua, M.-W. Wan, Compar-
[134] G.M. Gadd, C. White, L. De Rome, Heavy metal and radionuclide uptake by ative and competitive adsorption of copper, lead, and nickel using chitosan
fungi and yeasts, in: P.R. Norri, D.P. Kelly (Eds.), Biohydrometallurgy, A. Rowe, immobilized on bentonite, Carbohydr. Polym. 80 (2010) 891–899.
Chippenham, Wilts, UK, 1988. [165] K.M. Ibrahim, T.N. Ed-Deen, H. Khoury, Use of natural chabazite-phillipsite
[135] A. Hammaini, F. González, A. Ballester, M.L. Blázquez, J.A. Muñoz, Biosorption tuff in wastewater treatment from electroplating factories in Jordan, Environ.
of heavy metals by activated sludge and their desorption characteristics, J. Geol. 41 (2002) 547–551.
Environ. Manag. 84 (2007) 419–426. [166] E.F. Covelo, F.A. Vega, M.L. Andrade, Competitive sorption and desorption of
[136] J. Wang, C. Chen, Biosorption of heavy metals by Saccharomyces cerevisiae: heavy metals by individual soil components, J. Hazard. Mater. 140 (2007)
a review, Biotechnol. Adv. 24 (2006) 427–451. 308–315.
[137] W. Wang, H. Chen, A. Wang, Adsorption characteristics of Cd(II) from aqueous [167] A.A. El-Bayaa, N.A. Badawy, E.A. AlKhalik, Effect of ionic strength on the
solution onto activated palygorskite, Sep. Purif. Technol. 55 (2007) 157–164. adsorption of copper and chromium ions by vermiculite pure clay mineral, J.
[138] X. Wu, H. Zhou, F. Zhao, C. Zhao, Adsorption of Zn2+ and Cd2+ ions on vermic- Hazard. Mater. 170 (2009) 1204–1209.
ulite in buffered and unbuffered aqueous solutions, Adsorpt. Sci. Technol. 27 [168] H.S. Sherry, A series of advances, in: J.A. Marinsky (Ed.), Ion Exchange, Marcel
(2010) 907–919. Dekker, New York, 1969, p. 89.
460 S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461

[169] D. Caputo, F. Pepe, Experiments and data processing of ion exchange equilibria [202] V.C. Srivastava, I.D. Mall, I.M. Mishra, Characterization of mesoporous rice
involving Italian natural zeolites: a review, Micropor. Mesopor. Mater. 105 husk ash (RHA) and adsorption kinetics of metal ions from aqueous solution
(2007) 222–231. onto RHA, J. Hazard. Mater. B 134 (2006) 257–267.
[170] G.V. Tsitsishvili, T.G. Andronikashvilli, G.V. Kirov, L.D. Filizova, Natural Zeo- [203] M.H. Kalavathy, T. Karthikeyan, S. Rajgopal, L.R. Miranda, Kinetic and isotherm
lites, Ellis Horwood, Chichester, 1991. studies of Cu(II) adsorption onto H3 PO4 – activated rubber wood sawdust, J.
[171] G. Blanchard, M. Maunaye, G. Martin, Removal of heavy metals from waters Colloid Interface Sci. 292 (2005) 354–362.
by means of natural zeolites, Water Res. 18 (1984) 1501–1507. [204] M. Trgo, J. Perić, N.V. Medvidović, A comparative study of ion exchange kinet-
[172] V.J. Inglezakis, Design of Ion Exchange Columns Using Natural Minerals as ics in zinc/lead-modified zeolite-clinoptilolite systems, J. Hazard. Mater. 136
Packing Materials, Dep. Chem. Eng., PhD Thesis, National Technical University (2006) 938–945.
of Athens, Athens, 2002. [205] C. Aharoni, D.L. Sparks, S. Levinson, I. Ravina, Kinetics of soil chemical reac-
[173] M. Panayotova, B. Velikov, Kinetics of heavy metal ions removal by use of tions: relationship between empirical equations and diffusion models, Soil
natural zeolite, J. Environ. Sci. Health A Tox. Hazard. Subst. Environ. Eng. 37 Sci. Soc. Am. J. 55 (1991) 1307–1312.
(2002) 139–147. [206] D.L. Sparks, Kinetics of sorption/release reactions on natural particles, in:
[174] V.K. Jha, M. Matsuda, M. Miyake, Sorption properties of the activated carbon- P.M. Huang, N. Senesi, J. Buffle (Eds.), Structure and Surface Reactions of Soil
zeolite composite prepared from coal fly ash for Ni2+ , Cu2+ , Cd2+ and Pb2+ , J. Particles, John Wiley and Sons, London, 1998, pp. 413–448.
Hazard. Mater. 160 (2008) 148–153. [207] C. Namasivayam, R.T. Yamuna, Adsorption of chromium (VI) by a low cost
[175] M. Minceva, L. Markovska, V. Meshko, Removal of Zn2+ , Cd2+ and Pb2+ from adsorbent: biogas residual slurry, Chemosphere 30 (1995) 561–578.
binary aqueous solution by natural zeolite and granulated activated carbon, [208] T.S. Anirudhan, M. Ramachandran, Adsorptive removal of tannin from aque-
Maced. J. Chem. Chem. Eng. 26 (2007) 125–134. ous solutions by cationic surfactant-modified bentonite clay, J. Colloid
[176] Y. Liu, H. Li, X.-H. Zhu, Competitive adsorption of Ag+ , Pb2+ , Ni2+ , and Cd2+ ions Interface Sci. 299 (2006) 116–124.
on vermiculite, Sep. Sci. Technol. 45 (2010) 277–287. [209] M. Doğan, Y. Özdemir, M. Alkan, Adsorption kinetics and mechanism of
[177] L.H. Gai, S.G. Wang, W.X. Gong, X.W. Liu, B.Y. Gao, H.Y. Zhang, Influence of cationic methyl violet and methylene blue dyes onto sepiolite, Dyes and
pH and ionic strength on Cu(II) biosorption by aerobic granular sludge and Pigments 75 (2007) 701–713.
biosorption mechanism, J. Chem. Technol. Biotechnol. 83 (2008) 806–813. [210] V. Meshko, L. Markovska, M. Marinkovski, Experimental study and modelling
[178] M.W.P. Esparza-Soto, Biosorption of humic and fulvic acids to live activated of zinc adsorption by granular activated carbon and natural zeolite, Int. J.
sludge biomass, Water Res. 37 (2003) 2301–2310. Environ. Pollut. 27 (2006) 285–299.
[179] E. Eren, A. Tabak, B. Eren, Performance of magnesium oxide-coated bentonite [211] J. Febrianto, A.N. Kosasih, J. Sunarso, Y.-H. Ju, N. Indraswati, S. Ismadji, Equi-
in removal process of copper ions from aqueous solution, Desalination 257 librium and kinetic studies in adsorption of heavy metals using biosorbent: a
(2010) 163–169. summary of recent studies, J. Hazard. Mater. 162 (2009) 616–645.
[180] G. Dönmez, Z. Aksu, Removal of chromium(VI) from saline wastewaters by [212] X.-S. Wang, C. Sun, Removal of copper(II) ions from aqueous solutions using
Dunaliella species, Process Biochem. 38 (2002) 751–762. Na-mordenite, Sep. Sci. Technol. 42 (2007) 1215–1230.
[181] T. Undabeytia, S. Nir, G. Rytwo, C. Serban, E. Morillo, C. Maqueda, Modeling [213] M.Y. Pamukoglu, F. Kargi, Effects of operating parameters on kinetics of
adsorption–desorption processes of Cu on edge and planar sites of montmo- copper(II) ion biosorption onto pre-treated powdered waste sludge (PWS),
rillonite, Environ. Sci. Technol. 36 (2002) 2677–2683. Enzym. Microb. Technol. 42 (2007) 76–82.
[182] M. Trgo, J. Perić, Interaction of the zeolitic tuff with Zn-containing simulated [214] S. Schiewer, S.B. Patil, Pectin-rich fruit wastes as biosorbents for heavy metal
pollutant solutions, J. Colloid Interface Sci. 260 (2003) 166–175. removal: equilibrium and kinetics, Bioresour. Technol. 99 (2008) 1896–1903.
[183] J. Lützenkirchen, Ionic strength effects on cation sorption to oxides: macro- [215] A.B. Pérez Marín, M.I. Aguilar, V.F. Meseguer, J.F. Ortuño, J. Sáez, M. Lloréns,
scopic observations and their significance in microscopic interpretation, J. Biosorption of chromium (III) by orange (Citrus cinensis) waste: batch and
Colloid Interface Sci. 195 (1997) 149–155. continuous studies, Chem. Eng. J. 155 (2009) 199–206.
[184] J. Echeverría, J. Indurain, E. Churio, J. Garrido, Simultaneous effect of pH, tem- [216] Z. Aksu, A.I. Tatli, O. Tunç, A comparative adsorption/biosorption study of Acid
perature, ionic strength, and initial concentration on the retention of Ni on Blue 161: effect of temperature on equilibrium and kinetic parameters, Chem.
illite, Colloids Surf. A: Physicochem. Eng. Aspects 218 (2003) 175–187. Eng. J. 142 (2008) 23–39.
[185] C. Jeon, Y. Je Yoo, W.H. Hoell, Environmental effects and desorption charac- [217] F. Helfferich, Ion Exchange, Dover Publications, New York, 1995, pp. 166–168.
teristics on heavy metal removal using carboxylated alginic acid, Bioresour. [218] K.T. Valsaraj, Elements of Environmental Engineering: Thermodynamics and
Technol. 96 (2005) 15–19. Kinetics, Lewis Publisher, Boca Raton, Florida, 2000.
[186] J.R. Evans, W.G. Davids, J.D. MacRae, A. Amirbahman, Kinetics of cadmium [219] Y.S. Al-Degs, M.I. El-Barghouthi, A.A. Issa, M.A. Khraisheh, G.M. Walker, Sorp-
uptake by chitosan-based crab shells, Water Res. 36 (2002) 3219–3226. tion of Zn(II), Pb(II), and Co(II) using natural sorbents: equilibrium and kinetic
[187] J. Shen, Z. Duvnjak, Adsorption kinetics of cupric and cadmium ions on corn- studies, Water Res. 40 (2006) 2645–2658.
cob particles, Process Biochem. 40 (2005) 3446–3454. [220] S.J. Allen, G. McKay, K.Y.H. Khader, Intraparticle diffusion of a basic dye during
[188] R. Shawabkeh, A. Al-Harahsheh, A. Al-Otoom, Production of zeolite from Jor- adsorption onto sphagnum peat, Environ. Pollut. 56 (1989) 39–50.
danian oil shale ash and application for zinc removal from wastewater, Oil [221] M.F.F. Sze, G. McKay, An adsorption diffusion model for removal of
Shale 21 (2004) 125–136. parachlorophenol by activated carbon derived from bituminous coal, Environ.
[189] J. Taparcevska, L. Markovska, B. Koumanova, V. Meshko, Diffusion models for Pollut. 158 (2010) 1669–1674.
adsorption kinetics of Zn2+ , Cd2+ and Pb2+ onto natural zeolite, Water Sci. [222] B. Koumanova, P. Peeva, S.J. Allen, Variation of intaparticle diffusion param-
Technol. 62 (2010) 1136–1142. eter during adsorption of p-cholorophenol onto activated carbon made from
[190] G. McKay, Adsorption of dyestuffs from aqueous solutions with activated apricot stones, J. Chem. Technol. Biotechnol. 78 (2003) 582–587.
carbon I: equilibrium and batch contact-time studies, J. Chem. Technol. Bio- [223] V. Ponnusami, K.S. Rajan, S.N. Srivastava, Application of film-pore diffusion
technol. 32 (1982) 759–772. model for methylene blue adsorption onto plant leaf powders, Chem. Eng. J.
[191] V. Meshko, L. Markovska, M. Mincheva, A.E. Rodrigues, Adsorption of basic 163 (2010) 236–242.
dyes on granular activated carbon and natural zeolite, Water Res. 35 (2001) [224] R. Apiratikul, P. Pavasant, Sorption of Cu2+ , Cd2+ , and Pb2+ using modified
3357–3366. zeolite from coal fly ash, Chem. Eng. J. 144 (2008) 245–258.
[192] Y.-S. Ho, Citation review of Lagergren kinetic rate equation on adsorption [225] B. Volesky, Biosorption process simulation tools, Hydrometallurgy 71 (2003)
reaction, Scientometrics 59 (2004) 171–177. 179–190.
[193] Y.S. Ho, G. McKay, Kinetic models for the sorption of dye from aqueous solu- [226] J. Wang, C. Chen, Biosorbents for heavy metals removal and their future,
tion by wood, J. Environ. Sci. Health B: Process Saf. Environ. Protect. 76 (1998) Biotechnol. Adv. 27 (2009) 195–226.
183–191. [227] I. Langmuir, The constitution and fundamental properties of solids and liquids,
[194] S.H. Chien, W.R. Clayton, Application of Elovich equation to the kinetics of J. Am. Chem. Soc. 38 (1916) 2221–2295.
phosphate release and sorption in soils, Soil Sci. Soc. Am. J. 44 (1980) 265–268. [228] K.R. Hall, L.C. Eagleton, A. Acrivos, T. Vermeulen, Pore- and solid-diffusion
[195] D.L. Sparks, Kinetics and mechanisms of chemical reactions at the soil min- kinetics in fixed-bed adsorption under constant-pattern conditions, Ind. Eng.
eral/water interface, in: D.L. Sparks (Ed.), Soil Physical Chemistry, CRC Press, Chem. Fundam. 5 (1966) 212–223.
Boca Raton, FL, 1999, pp. 135–192. [229] H.M.F. Freundlich, Uber die adsorption in losungen, J. Phys. Chem. 57 (1906)
[196] E. Guibal, C. Milot, J.M. Tobin, Metal-anion sorption by chitosan beads: equi- 385–470.
librium and kinetic studies, Ind. Eng. Chem. Res. 37 (1998) 1454–1463. [230] R. Sips, On the structure of a catalyst surface, J. Chem. Phys. 16 (1948) 490–495.
[197] Z. Cheng, X. Liu, M. Han, W. Ma, Adsorption kinetic character of copper ions [231] M.M. Dubinin, The potential theory of adsorption of gases and vapors for
onto a modified chitosan transparent thin membrane from aqueous solution, adsorbents with energetically non-uniform surface, Chem. Rev. 60 (1960)
J. Hazard. Mater. 182 (2010) 408–415. 235–266.
[198] G.M. Walker, L. Hansen, J.A. Hanna, S.J. Allen, Kinetics of a reactive dye adsorp- [232] M.J. Temkin, V. Pyzhev, Kinetics of ammonia synthesis on promoted iron
tion onto dolomitic sorbents, Water Res. 37 (2003) 2081–2089. catalyst, Acta Physicochem. URSS 12 (1940) 217–256.
[199] D. Reichenberg, Properties of ion-exchange resin in relation to their structure, [233] O. Redlich, D.L. Peterson, A useful adsorption isotherm, J. Phys. Chem. 63
III. Kinetics of exchange, J. Am. Chem. Soc. 75 (1953) 589–597. (1959) 1024.
[200] R. Lagoa, J.R. Rodrigues, Kinetic analysis of metal uptake by dry and gel alginate [234] J. Toth, State equations of the solid gas interface layer, Acta Chem. Acad. Hung.
particles, Biochem. Eng. J. 46 (2009) 320–326. 69 (1971) 311–317.
[201] G.E. Boyd, A.W. Adamson, L.S. Mayers, The exchange adsorption of ions from [235] A.R. Khan, R. Ataullah, A. Al-Haddad, Equilibrium adsorption studies of some
aqueous solutions on organic fly ash, kinetic II, J. Am. Chem. Soc. 69 (1947) aromatic pollutants from dilute aqueous solutions on activated carbon at
2836–2848. different temperatures, J. Colloid Interface Sci. 194 (1997) 154–165.
S. Malamis, E. Katsou / Journal of Hazardous Materials 252–253 (2013) 428–461 461

[236] C.J. Radke, J.M. Prausnitz, Adsorption of organic solutes from dilute aque- [246] U.R. Malik, S.M. Hasany, M.S. Subhani, Sorptive potential of sunflower stem
ous solutions on activated carbon, Ind. Eng. Chem. Fundam. 11 (1972) for Cr(III) ions from aqueous solutions and its kinetic and thermodynamic
445–450. profile, Talanta 66 (2005) 166–173.
[237] F. Kargi, S. Ozmihci, Biosorption performance of powdered activated sludge [247] P. Miretzky, C. Muñoz, A. Carillo-Chavez, Cd(II) removal from aqueous solution
for removal of different dyestuffs, Enzym. Microb. Technol. 35 (2004) by Eleocharis acicularis biomass, equilibrium and kinetic studies, Bioresour.
267–271. Technol. 101 (2010) 2637–2642.
[238] S. Bruanuer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular [248] D.H.K. Reddy, K. Seshaiah, A.V.R. Reddy, M.M. Rao, M.C. Wang, Biosorption of
layers, J. Am. Chem. Soc. 60 (1938) 309–316. Pb+2 from aqueous solutions by Moringa oleifera bark: equilibrium and kinetic
[239] K. Vijayaraghavan, T.V.N. Padmesh, K. Palanivelu, M. Velan, Biosorption studies, J. Hazard. Mater. 174 (2010) 831–838.
of nickel(II) ions onto Sargassum wightii: application of two-parameter [249] S.J. Allen, Q. Gan, R. Matthews, P.A. Johnson, Comparison of optimised
and three-parameter isotherm models, J. Hazard. Mater. 133 (2006) isotherm models for basic dye adsorption by kudzu, Bioresour. Technol. 88
304–308. (2003) 143–152.
[240] T. Grchev, M. Cvetkovska, T. Stafilov, J.W. Schultze, Adsorption of polyacryl- [250] K. Vijayaraghavan, H.Y.N. Winnie, R. Balasubramanian, Biosorption charac-
amide on gold and iron from acidic aqueous solutions, Electrochim. Acta 36 teristics of crab shell particles for the removal of manganese(II) and zinc(II)
(1991) 1315–1323. from aqueous solutions, Desalination 266 (2011) 195–200.
[241] Y. Liu, Y.-J. Liu, Biosorption isotherms, kinetics and thermodynamics, Sep. [251] C.D. Johnson, F. Worrall, Novel granular materials with microcrystalline active
Purif. Technol. 61 (2008) 229–242. surfaces-waste water treatment applications of zeolite/vermiculite compos-
[242] J.C. Bellot, J.S. Condoret, Modelling of liquid chromatography equilibrium, ites, Water Res. 41 (2007) 2229–2235.
Process Biochem. 28 (1993) 365–376. [252] I.C. Ostroski, M.A.S.D. Barros, E.A. Silva, J.H. Dantas, P.A. Arroyo, O.C.M.A. Lima,
[243] S.J. Allen, G. McKay, J.F. Porter, Adsorption isotherm models for basic dye A comparative study for the ion exchange of Fe(III) and Zn(II) on zeolite NaY,
adsorption by peat in single and binary component systems, J. Colloid Inter- J. Hazard. Mater. 161 (2009) 1404–1412.
face Sci. 280 (2004) 322–333. [253] P. Castaldi, L. Santona, S. Enzo, P. Melis, Sorption processes and XRD analysis
[244] T.A. Davis, B. Volesky, A. Mucci, A review of the biochemistry of heavy metal of a natural zeolite exchanged with Pb2+ , Cd2+ and Zn2+ cations, J. Hazard.
biosorption by brown algae, Water Res. 37 (2003) 4311–4330. Mater. 156 (2008) 428–434.
[245] B.A. Shah, A.V. Shah, R.R. Singh, Sorption isotherms and kinetics of chromium [254] M. Achak, A. Hafidi, N. Ouazzani, S. Sayadi, L. Mandi, Low cost biosorbent
uptake from wastewater using natural sorbent material, Int. J. Environ. Sci. “banana peel” for the removal of phenolic compounds from olive mill waste-
Technol. 6 (2009) 77–90. water: kinetic and equilibrium studies, J. Hazard. Mater. 166 (2009) 117–125.

Potrebbero piacerti anche