Sie sind auf Seite 1von 19

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

Available online at www.sciencedirect.com

ScienceDirect
journal homepage: www.elsevier.com/locate/watres

Review

Aqueous mercury adsorption by activated carbons


Pejman Hadi a, Ming-Ho To a, Chi-Wai Hui a, Carol Sze Ki Lin b,
Gordon McKay a,c,*
a

Chemical and Biomolecular Engineering Department, Hong Kong University of Science and Technology,
Clear Water Bay Road, Hong Kong
b
School of Energy and Environment, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong
c
Division of Sustainable Development, College of Science, Engineering and Technology, Hamad Bin Khalifa
University, Qatar Foundation, Doha, Qatar

article info

abstract

Article history:

Due to serious public health threats resulting from mercury pollution and its rapid dis-

Received 14 October 2014

tribution in our food chain through the contamination of water bodies, stringent regula-

Received in revised form

tions have been enacted on mercury-laden wastewater discharge. Activated carbons have

19 December 2014

been widely used in the removal of mercuric ions from aqueous effluents. The surface and

Accepted 9 January 2015

textural characteristics of activated carbons are the two decisive factors in their efficiency

Available online 21 January 2015

in mercury removal from wastewater. Herein, the structural properties and binding affinity
of mercuric ions from effluents have been presented. Also, specific attention has been

Keywords:

directed to the effect of sulfur-containing functional moieties on enhancing the mercury

Activated carbon

adsorption. It has been demonstrated that surface area, pore size, pore size distribution

Adsorption

and surface functional groups should collectively be taken into consideration in designing

Mercury

the optimal mercury removal process. Moreover, the mercury adsorption mechanism has

Porous structure

been addressed using equilibrium adsorption isotherm, thermodynamic and kinetic

Sulfur functional groups

studies. Further recommendations have been proposed with the aim of increasing the
mercury removal efficiency using carbon activation processes with lower energy input,
while achieving similar or even higher efficiencies.
2015 Elsevier Ltd. All rights reserved.

Contents
1.
2.
3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Preparation of activated carbon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Effect of treatment techniques on mercury removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.1. Physical and chemical activation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2. Sulfurization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

* Corresponding author. Chemical and Biomolecular Engineering Department, Hong Kong University of Science and Technology, Clear
Water Bay Road, Hong Kong. Tel.: 852 23588412; fax: 852 23580054.
E-mail address: kemckayg@ust.hk (G. McKay).
http://dx.doi.org/10.1016/j.watres.2015.01.018
0043-1354/ 2015 Elsevier Ltd. All rights reserved.

38

4.

5.
6.
7.

1.

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

Effect of adsorption parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44


4.1. Equilibrium contact time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.2. Initial concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.3. pH value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.4. Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.5. Adsorbent dosage and particle size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Mercury affinity to various functional groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Equilibrium adsorption isotherms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Conclusion and future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

Introduction

Mercury is categorized as an extremely toxic substance whose


health hazards have primarily been associated with inhalation of mercury vapor or ingestion of organic mercury via
aquatic organisms causing mercury poisoning, widely known
as Minamata disease (Harada, 1982). The absorption of this
hazardous substance into the bloodstream, its distribution to
the entire tissues and its bioaccumulation in the receptive
sites result in many well-recognized adverse effects, such as
potent neurotoxicity, blood vessel congestion and kidney
damages (Kidd and Batchelar, 2012).
The presence of mercury in the environment as a result of
naturogenic sources, such as geothermal eruptions and
seismic activities, represents only a small fraction of the total
annual mercury emission. The anthropogenic activities are
the major contributors for mercury pollution in the
ecosystem, among which the fossil fuel combustion plays the
most dominant role. An overarching estimate of anthropogenic input of 16 elements into the environment by Nriagu
and Pacyna has revealed that mercury releases into the atmosphere and aquatic bodies are in the same range
(Ebinghaus et al., 1999; Nriagu and Pacyna, 1988). The
anthropogenic sources of mercury can be divided into primary
and secondary sources. The former involves the mobilization
and release of mercury of geological origin to the environment, such as mining, industrial processing of ores or fossil
fuel combustion and more specifically coal. The latter is
associated with the direct use of mercury in industrial processes, including vinyl chloride monomer production as
catalyst, batteries as cathodes and chlor-alkali production as
well (Pacyna et al., 2010). Improper discharge of effluents and
exhaust emissions from both primary and secondary sources
accounts for the environmental concerns over this toxic
compound. Despite the constraints in the usage of certain
toxic compounds by the Restriction of Hazardous Substances
Directive (RoHS), some substances, including mercury, have
been granted exemption to narrowly-defined applications due
to the technical/scientific impracticality of the substance
prohibition (United Nations Environment Programme, 2010).
Hence, it can be evidently remarked that complete elimination of mercury from the ecosystem is unrealistic due to the
uncontrollable discharge of mercury by primary anthropogenic sources and the inevitable, though limited, utilization of

mercury in industrial processes (secondary anthropogenic


sources).
Accordingly, as a result of the catastrophic impacts of
mercury presence in the ecosystem and its potential fatal
health consequences, many stringent regulations and directives have been enacted regarding the control of the mercury discharge into the environment (Sunderlan and Chmura,
2000). Therefore, many researchers have been engaged with
the removal of mercury from aqueous media. Numerous
techniques employed for this purpose include precipitation
(Blue et al., 2010; Hutchison et al., 2008; Matlock et al., 2001),
coagulation (Henneberry et al., 2011; Nanseu-Njiki et al., 2009),
cementation (Ku et al., 2002; Lo and Yu, 1988), ultrafiltration
(Barron-Zambrano et al., 2002; Han et al., 2014; Uludag et al.,
 et al.,
1997), solvent extraction (Huebra et al., 2003; Sevdic
re et al.,
1980), photocatalysis (Botta et al., 2002; de la Fournie
 pez-Mun
~ oz et al., 2011), adsorption (Aguado et al.,
2007; Lo
2005; Antochshuk and Jaroniec, 2002; Bandaru et al., 2013;
Cui et al., 2013; Di Natale et al., 2011; Li et al., 2011, 2013;
Mondal et al., 2013), ion exchange (Anirudhan et al., 2008;
Chiarle et al., 2000; Gash et al., 1998; Lloyd-Jones et al., 2004)
or a combination of these methods (Barron-Zambrano et al.,
2004; Byrne and Mazyck, 2009). However, the challenges concerning some of these methods include high energy demand
for process operation, large amounts of chemicals used, high
operation and/or capital costs, removal inefficiency and
unselectivity (Ismaiel et al., 2013; Li et al., 2011; Taurozzi et al.,
2013; Wajima and Sugawara, 2011). Among all these removal
techniques, adsorption has been found to be a very promising
method for the removal of heavy metals from wastewater
streams owing to the ease of operation, heavy metal removal
efficiency, high adsorption rate, selective removal and the
availability of a wide range of adsorbent materials (Hadi et al.,
2014a, 2013a, 2013b, 2013c, 2013d; Ismaiel et al., 2013; Nabais
et al., 2006; Ramadan et al., 2010; Xu et al., 2014). Among all
types of adsorbent materials including extracellular biopolymers (Inbaraj et al., 2009), cellulosic materials (Takagai
et al., 2011), zeolites and aluminosilicates (Liu et al., 2013;
Somerset et al., 2008), nanomaterials (Bandaru et al., 2013)
and activated carbons (Anoop Krishnan and Anirudhan, 2002;
Di Natale et al., 2011; Ranganathan, 2003), the latter has gained
considerable attention both in research-based studies and
practical industrial applications. Also, some studies have
argued the high cost of activated carbon materials is still an
important problem and have resolved the issue by using

39

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

waste-derived activated carbons to diminish the production


cost of this adsorbent (Kadirvelu et al., 2004; Mohan et al.,
2001; Rao et al., 2009; Zabihi et al., 2010; Zhang et al., 2005).
The current work aims at reviewing the adsorptive removal
of aqueous mercury from effluents by activated carbons. The
focus of this paper is to present a comprehensive overview of
the activated carbon preparation by chemical and physical
methods, their modification techniques, their mercury
removal efficiencies and the effect of various parameters,
such as pH, initial concentration, activated carbon amount,
adsorbent particle size and temperature, on mercury uptake.
These factors are of the utmost significance, as any change in
these parameters may considerably change the mercury
removal efficiency of an adsorbent. Therefore, a general
knowledge of the effect of these parameters is critical in
designing the appropriate mercury-laden wastewater treatment facilities.

2.

Preparation of activated carbon

Activated carbons have long been used as absorbent for the


removal of various pollutants. Textural properties of activated
carbons and functional groups on their surface are two of the
principal characteristics which should be enhanced by certain
modification processes in order to make them exhibit high
pollutant removal efficiencies.
Physical and chemical activation techniques are the most
commonly-used approaches to develop high internal porosity
and desired pore size and also to introduce certain functional
groups onto the adsorbent surface (Hadi et al., 2015).
The physical activation mainly involves two steps,
carbonization and activation. Carbonization includes a heat
treatment of a carbonaceous precursor at moderate temperatures, mainly below 700  C, in an inert atmosphere to pyrolize the precursor material into a low surface area char.
Carbonization leads to the partial evolution of the volatile
matter from the carbonaceous precursor, enriching the produced char in carbon content and developing the preliminary
branched porous structure. For a carbonization temperature
up to 700  C, the volume of the micropores gradually increases, while further increase in the temperature reduces the
pore volume of the material. The produced char is subsequently activated at elevated temperatures, usually above
700  C, under a partial oxidizing atmosphere (primarily steam
or carbon dioxide as gasifying agents) (Hadi et al., 2015). The
aim of the activation stage is to produce a highly porous
structure from the weakly-developed porous char. The activation process entails the reaction of the oxidizing agent with
tar decomposition products blocking the pores to volatilize
carbon oxide, which, in turn, opens the closed pores, widens
the existing small pores and forms new pores. The resulting
activated product at this stage has high pore volume and
surface area, therefore enhancing its ability for the removal of
pollutants by capturing absorbate molecules in its porous
network (Budinova et al., 2006; Hadi et al., 2014b). The
simplified gasification reactions of the oxidizing agents with
carbon have been illustrated according to the following stoichiometric equations:

1

C CO2 /2CO DH 159 KJ mol


C H2 O/CO H2

(R1)
1

DH 117 KJ mol

(R2)

The chemical activation technique is a one-step process in


which the carbonization and activation processes occur
simultaneously. During chemical activation, the source material is submerged in a dehydrating compound (such as
phosphoric acid, zinc chloride, alkaline hydroxides or alkaline
carbonates) resulting in the diffusion of the chemical reagent
into the particles and its incorporation into the carbon structure. The resultant slurry is then heated up to temperatures in
the range of 400e600  C under an inert atmosphere. This leads
to the depolymerization of cellulose, hemicellulose or lignin
catalyzed by the chemical reagent, followed by dehydration
and condensation leading to the formation of more aromatic
and reactive products. Also, in some cases, the alkaline metals
are intercalated between the graphene layers while creating
some porosity by the oxidation of carbon into carbon oxides
(Marsh and Reinoso, 2006). This intercalation inhibits the
contraction of the carbonaceous precursor during the heat
treatment.
Energy required for chemical activation process is considerably lower than that of physical activation method. Significantly lower activation temperatures, shorter reaction time
and employment of a single-step treatment in the case of
chemical activation process account for this low energy consumption. Moreover, it has been reported that, in most cases,
both carbon yield and specific surface area for the materials
produced by chemical activation are higher than those of the
physical activation method. These two reasons account for
the more extensive application of the chemical activation
 -Agullo
 et al.,
method compared with the physical one (Macia
2004).
Although, commercial activated carbons have been widely
used in various industries to remove mercury from wastewater streams, their use has been recently challenged by their
high cost (Di Natale et al., 2006). Hence, much research has
recently been devoted to the exploration of alternative wastebased carbon sources as precursor for activated carbon production. Toles et al. found that it is highly profitable to produce almond shell-based granular activated carbon (GAC) at
US$20/ton compared with two comparable commercial GACs
at about US$3.3/kg using it as a basis for economic feasibility
study (Toles et al., 2000). Considering such a great financial
incentive, numerous researches have been conducted on
certain inexpensive waste precursors to produce costeffective activated carbons with high pollutant removal efficiencies. Among certain precursor materials to be used in
activated carbon production, high mercury adsorption capacities have been reported using coirpith (Namasivayam and
Kadirvelu, 1999), furfural (Yardim et al., 2003), walnut shell
(Zabihi et al., 2010), agricultural solid waste (Kadirvelu et al.,
2003), silk cotton hull (Roberts and Rowland, 1973), sago
waste (Kadirvelu et al., 2004), coconut tree sawdust, maize cob
and banana pith (Kadirvelu et al., 2003) while moderate to low
adsorption capacities have been observed for activated carbons derived from pazzolana and yellow tuff (Di Natale et al.,
2006), Fullers earth (Oubagaranadin et al., 2007), Ceibapentandra hulls, Phaseolus aureus hulls and Cicerarietinum

40

(Nabais et al., 2006)


(Choi and Jang, 2008)
(Lu et al., 2014)
22.5
1.967

(Inbaraj and Sulochana, 2006)


 -Agullo
 et al., 2004)
(Macia

(Ekinci et al., 2002)

0.004

2.32

N2 & CO2
N2 & CO2
N2 & CO2

N2
H2O
H2O

N2 & H2O

1528
2487
997
702
870

174
134
92
56
105
37
153
174

140
4.93

0.681
0.89
0.64

0.64
0.86
0.37
2.7
0.428

0.297
0.425
0.21

0.328

0.321
0.11
0.085

(cm g )
(cm g )
(mg g )

25.88
23.66
22.88

521
325
280
761
1100
480
600
460
720
520
1000
1100

(m g )
Atm.

H2O

3.1.

5
5
15

3
10

20

1

HR ( C min )

(Budinova et al., 2006)


(Budinova et al., 2003)

(Rao et al., 2009)

(nm)
(cm g )

1

1
850

N2
N2

H 2O

9
22.5
5

1
2
600
750
400

HR ( C min )

890
890
900

T (h)

750

T ( C)
Atm.

1
5
800
900

2
750

500e950

T (h)

200

T ( C)

HR denotes heating rate.

10

10

1

N2

Vmeso

1
3

Vmicro

1
3

Vt
qm

1
1
2

SBET

3.
Effect of treatment techniques on mercury
removal

Activation conditions

(Rao et al., 2009). Hence, it can be inferred that the type of


precursor is an influential factor in determining the adsorption efficiency of the activated carbons.

a 

Carbonization conditions

Table 1 e Effect of the precursor type and physical activation conditions on the mercury removal efficiencies of the activated carbons.

Dp

Ref.

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

Physical and chemical activation

The nature of precursor materials and their activation


methods directly influences the surface area, pore size and
functional groups on the surface of the prepared activated
carbons which can, in turn, have an influence on the mercury
removal efficiency. Tables 1 and 2 show the effect of the precursor type and activation conditions on mercury adsorption
capacities of the prepared activated carbons.
A suitable activation procedure leads to the production of a
high surface area adsorbent material and thus, reasonably
high mercury removal efficiency. Budinova et al. have
demonstrated the effect of steam activation and air oxidation
on the specific surface area of a carbonaceous biomass waste.
Sample activated with steam showed significantly higher
surface area than the one oxidized in air atmosphere. This
resulted in a higher mercury adsorption capacity for the
steam-activated adsorbent as compared with the air-oxidized
adsorbent (Budinova et al., 2003). A more detailed study of the
activated carbon preparation conditions under various atmospheres has been reported elsewhere (Budinova et al.,
2006). Water vapor has been shown to penetrate into the
carbon structure, react with the carbon at the internal surface
of the carbon and extract the carbon from the pore walls,
resulting in the enlargement of the existing pores and creation
of new pores (as shown in the reaction R2) and thereby an
increase in the adsorption capacity of the produced adsorbent.
It has been demonstrated that the pore size and pore size
distribution of activated carbons can be manipulated by using
different activating atmospheres. Molina-Sabio et al. have
shown that steam-activated carbons exhibit larger mesopores
compared to carbons activated under a CO2 atmosphere
(Molina-Sabio et al., 1996). This will, undoubtedly, influence
the adsorption behavior of these adsorbents towards various
adsorbates. However, no comprehensive study has been
conducted on the effect of activation atmosphere, which, in
turn, results in a change in the surface functionalities of the
adsorbents, to maximize mercury removal from effluents.
In addition to the activation conditions, both surface area
and functional groups are predominantly affected by the nature of the precursor material used. As shown in Table 1,
Ekinci et al. have found that, under similar activation conditions, the surface areas range from 460 to 720 m2 :g1 for coalbased activated carbons and 1000 and 1110 m2 :g1 for apricot
stone and furfural-based activated carbons, respectively
(Ekinci et al., 2002). The difference in the textural properties of
these produced activated carbons is reflected in their mercury
adsorption capacities. Furthermore, Rao et al. compared the
mercury removal efficiencies of activated carbons prepared
from three different carbonaceous precursors under similar
activation conditions and observed huge differences in their
surface areas and adsorption capacities (Rao et al., 2009).

Table 2 e Effect of the precursor type and chemical activation conditions on the mercury removal efficiencies of the activated carbons.
Carbonization/Activation
conditions
T ( C)

T (h)

600/600
700
600

1/1
2

200/450

600/800

HRa
( C min1)
3
10

Atm.
H2O/N2
H2O
N2

Air

N2/H2O

0.5/1
N2

Chem.

H2SO4
H2SO4

NaOH
ZnCl2

K2CO3
KOH

NaOH

1/5

900

900
900
200
200/400

800
850

2/1

0.5
0.5

5/15

N2 & CO2
N2

10

Conc.
Conc.

H2O2
H2SO4 & (NH4)2S2O8 50%
H2SO4
Conc.
H2SO4 & (NH4)2S2O8 Conc.

H2SO4

800/900

Conc.

S
H2S
Pyrrole

N2
H2S
SO2, H2S & N2
H2O, SO2 & H2S
H2O & SO2
H2O,& H2S
H2O
K2S
N2

Imp. Rb

1:200
3: 5 & 1:33

1:1
1:2
1:4
1:6
1:8
1:2
1:4
1:6
1:8
1:3
1:3
43:100
11:20

1:3.6

qm

Vt

Vmicro

Vmeso

Imp.tc (h) (m2 g1) (mg g1) (cm3 g1) (cm3 g1) (cm3 g1)
1290
1360
1050

1:20

20% (W/V)
98%
1:0.5
1:1
Conc.
1:1.8
33-75%

SBET

0.5
12
2

24
2
1
1
1
1
1
1
1
1
2
1

629
625
1100
592
379.4
780
803
208.1
1260
1090
1635
2225
2420
1130
2000
2541
3033
848
905
413
346
158
922
785
764
500.5
506.5
530.2
536.5
30.0
30.0

160
154
10
10
10
10
10
18.1
55.6
174
154

0.939
1.026
0.827

0.471
0.485
0.38

0.425

0.486
0.541
0.19

0.11

52.7
151.5
100.9
109.9
129

410
450
541
682
441
301
351
351
227.3
222.2
217.4
208.3
235.7
243.9

Ref.

(nm)
1.4
1.28

(Budinova et al., 2006)


(Budinova et al., 2003)
(Kadirvelu et al., 2003)

>7.5

(Mohan et al., 2001)


(Kadirvelu et al., 2004)
(Yardim et al., 2003)
(Namasivayam and
Kadirvelu, 1999)
(Wahi et al., 2009)
(Zabihi et al., 2009)
(Namasivayam
et al., 1993)
(Budinova et al., 2008)
 -Agullo

(Macia

0.492
0.49
0.78
0.9
0.94
0.51
0.81
0.96
1.02
0.33
0.33
1.52
1.04
0.54
0.87
0.78
0.72
0.43
0.48
0.48
0.52

Dp

et al., 2004)

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

400/700

Chemical treatment

(Nabais et al., 2006)


(Choi and Jang, 2008)

0.37
0.31
0.3
0.04
0.06
0.13

0.16
0.13
0.16
0.33
0.29
0.28
0.2

(Gomez-Serrano
et al., 1998)
(Anoop Krishnan
and Anirudhan, 2002)

(Wajima and
Sugawara, 2011)

41

(continued on next page)

42

Table 2 e (continued )
Carbonization/Activation
conditions
T ( C)

T (h)

900
1000

0.5
0.5

HRa
( C min1)

50

Atm.

N2

Chem.

H2O2
H2 & CS2
H2O2 & He
H2O2, He & CS2
He
S

HCl
H2SO4
H2SO4
HNO3, SOCl2 &
C2H4(NH2)2
H2O2

600
600
600
600
600
600
600
500
700
800
900

10

N2

b
c

HR denotes heating rate.


Imp. R denotes impregnation ratio.
Imp. t denotes impregnation time.

Imp. Rb

3:10

1N
13 M
13 M
12 M
12 M
5 M, 5%
& 0.05 M
30%

0.01 M

1:5
1:5
1:9
1:9
3: 20 & 3:8
1:2
2:5
1:25
1:25
1:25
1:25
1:50
2:25
1:25
1:25
1:25
1:25
1:25
1:0.3

qm

Vt

Vmicro

Vmeso

Imp.tc (h) (m2 g1) (mg g1) (cm3 g1) (cm3 g1) (cm3 g1)

168

2:1
2:1
2:1
2:1
2:1

SO2

TOMATS
HNO3
HSCH2COOH,
(CH3CO)2O & H2SO4
a

Conc.

SBET

4
0.33
0.33
0.5
0.5
7

0.5
1
2
3
1
1
1
1
1
1
1
48
72
13

28.0
30.0
836.0
778.0
950.0
776.0
880.0
645.0
594.0

254.4
224.1
100
170
110
175
110
467
507

1070.0
1045.0
311.0
66.0

827
427
9.32
303.03
384
385
526
120

19.0

829.0
825.0
720.9
772.5
776.4
790.7
773.7
518.5
757.2
764.1
751.3
1087.0
1057.0
107.7
136.7
193.7

5
4.2
122.8
129.8
130.5
131.6
128.2
114.8
125.7
135.9
184.2
196.8
207.8
83.3
315.8
694.9

0.23
0.18

Dp

Ref.

(nm)

(Lopez-Gonzalez
et al., 1982)

0.22
(Wang et al., 2009)

(Feng et al., 2004)


(El-Shafey, 2010)
(Cox et al., 2000)
(Zhu et al., 2009)
0.413
0.408

0.318
0.315

0.01
0.013

1.994 (Lu et al., 2014)


1.98
(Asasian et al., 2014)

(Ismaiel et al., 2013)


(Li et al., 2013)

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

700

Chemical treatment

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

Nonetheless, although specific surface area is one of the factors affecting the mercury adsorption, it is not the only
influential parameter. Zabihi et al. prepared two activated
carbons with surface areas of 780 and 803 m2 :g1 , while their
mercury adsorption capacities were 151 and 101 m2 :g1 ,
respectively (Zabihi et al., 2009). Also, the surface areas of the
activated carbons prepared by Roa et al. ranged from 280 to
521 m2 :g1 , whereas their mercury adsorption capacities
under similar adsorption conditions did not exhibit a significant difference (Rao et al., 2009). Wang et al. demonstrated
that an activated carbon with a surface area of 1896 m2 :g1
had a much smaller adsorption capacity than an activated
carbon with a surface area of 1070 m2 :g1 (160 vs 827 m2 :g1 ,
respectively) (Wang et al., 2009). Since the size of the solvated
mercury is much larger than the nitrogen molecules, the pore
size and pore size distribution of the produced activated carbons besides their specific surface area is of significance.
Hence, it can be concluded that other factors, such as surface
functional groups, pore size and pore size distribution also
have considerable effects besides surface area in mercury
removal. Nevertheless, few studies have simultaneously
investigated the effects of pore structure and functional
groups on mercury removal.
Considering the surface areas listed in Tables 1 and 2, it is
noticeable that the chemical activation technique has greater
performance in pore formation at appropriate chemical reagent to adsorbent ratio. Comparing the surface areas obtained by chemical and physical activation of coal tar pitch
carbon fibers, it is obvious that although high surface area
activated carbon (2487 m2 :g1 ) can be obtained by physical
activation, it is not an economical option in terms of energy
consumption due to prolonged activation time at a high
temperature (22 h at 890  C) causing an excessive carbon burnoff (94%). On the contrary, chemical activation of this material
using an alkaline solution as activating reagent (KOH or NaOH)
at an impregnation ratio of 6:1 (w/w) yields activated carbon
with similar pore width and higher surface area. Other advantages of the chemically activated carbon are higher product yield (60% and 27%, respectively), lower activation
temperature (750  C) and shorter activation time (1 h). The
highest surface area (3033 m2 :g1 ) can be obtained using NaOH
as activating agent at an impregnation ratio of 8:1 (w/w)
 -Agullo
 et al., 2004).
(Macia
In order to enhance mercury adsorption, several authors
have studied the combination of chemical and physical activation techniques. Budinova et al. confirmed that when the
H3PO4-impregnated carbonaceous sample was treated under
steam atmosphere, both the surface area and iodine number
were considerably higher than the samples pyrolyzed under
nitrogen atmosphere. They also revealed that the concentration of the chemical reagent used for impregnation has a
significant effect on pore development. When the concentration of phosphoric acid was increased from 20% to 50%, the
mercury adsorption capacity of the activated carbon was
enhanced considerably (Budinova et al., 2006; Yardim et al.,
2003). Although no in-depth reason was provided for this
phenomenon, we believe that increasing the acid concentration increases the rate of the pyrolytic decomposition of the
precursor and enhances the density of the cross-linked
structure due to the catalytic effect of the phosphoric acid

43

and thus results in the modification of the textural properties


of the activated carbon.

3.2.

Sulfurization

The binding ability of the carbonaceous compound surfaces


with sulfur-containing functional groups is well-recognized
(Cai and Jia, 2010; Hsi et al., 2001; Korpiel and Vidic, 1997;
Vitolo and Pini, 1999; Wang et al., 2009). It has been widely
verified that sulfurization of activated carbons results in
enhanced adsorption capacity and selectivity towards mercury. Therefore, the application of sulfur-functionalized activated carbons in the removal of mercury has become a
common practice. A variety of techniques have been
employed to immobilize sulfur on the surface of adsorbents,
including treatment with carbon disulfide (CS2), sodium sulfide (Na2S), hydrogen sulfide (H2S), sulfur dioxide (SO2) or
sulfur powder, with the aim of increasing their mercury uptakes (Feng et al., 2006a; Fouladi Tajar et al., 2009; Mohan et al.,
2001; Vitolo and Pini, 1999; Wajima et al., 2009; Zhang et al.,
2003). Nabais et al. have applied two modification techniques, namely impregnation with elemental sulfur and using
hydrogen sulfide gas as modifying agent. Both of these adsorbents exhibited higher adsorption capacities than the unsulfurized activated carbons (Nabais et al., 2006). Asasian
et al. found a 50% increase in mercury adsorption capacity by
sulfurizing the activated carbon with 4% sulfur dioxide gas
stream (Asasian et al., 2014). Wang et al. have studied the effect of the impregnation of activated carbon with elemental
sulfur. Elemental sulfur not only can directly deposit on the
adsorbent surface and interact with mercury, but also can
react with the adsorbent surface and lead to the formation of
new functional groups to enhance mercury adsorption (Wang
et al., 2009). They found that the mercury adsorption capacity
of the unmodified and modified activated carbons were
190 mg:g1 and 820 mg:g1 , respectively. The chemical reaction between elemental sulfur and the surface of the adsorbent leads to the formation of disulfide, thiophene, sulfoxide
and sulfone groups that have more affinity to mercuric ions
and can enhance the overall mercury adsorption capacity and
selectivity (Cai and Jia, 2010; Wang et al., 2009). Mohan et al.
have observed a doubled mercury uptake after soaking the
adsorbent in carbon disulfide (Mohan et al., 2001). Despite the
well-acknowledged sulfur effect on mercury sequestration,
the process of sulfur binding onto the activated carbon surface
and consequently, the mechanism of mercury adsorption
onto sulfur-containing moieties have not been satisfactorily
exploited and established. An in-depth understanding of the
activation and mercury adsorption mechanisms will assist in
designing a proper activation/functionalization procedure in
order to achieve high mercury abatement. Pillay et al. have
investigated the activation and adsorption mechanisms using
Raman spectroscopy as an analytical tool to monitor the
changes in the functional groups before and after the
adsorption process (Pillay et al., 2013). They verified the
presence of S]CeS bonds (at 475 cm1 , 495 cm1 and
503 cm1 ) associated with thiol and thioester groups after
treating the virgin carbon nanotube with phosphorus pentasulfide. Subsequent mercury adsorption revealed significantly
diminished intensities of the bands corresponding to the thiol

44

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

groups and appearance of new band assigned to Hg(SH)2 and


Hg2(SH)2 bonds on the used adsorbent material. In addition to
interaction with thiol moieties, weak chemisorption between
the mercury and hydroxyl groups were also noticed by the
reduction of hydroxyl peak intensity at 620 cm1 and the
formation of new peak at 550 cm1 corresponding to HgeO
bond, indicating strong binding of mercury ions to the thiol
groups rather than oxygen functional groups. Furthermore,
Nabais et al. have identified the presence of SeS, C]S, CeS
and SeO bonds by FT-IR analysis after sulfurization, but
changes in the intensities of these peaks after the mercury
adsorption have not been provided (Nabais et al., 2006).
Due to the affinity of sulfur functional groups with mercury, higher amounts of sulfur moieties will theoretically be
advantageous in mercury removal. Several researchers have
reported the direct linear relationship between mercury uptake and sulfur content (Cai and Jia, 2010; Pillay et al., 2013).
However, Wang et al. have ruled out this hypothesis and have
demonstrated that the activated carbon with lower sulfur
amount (22%) on its surface had a higher mercury uptake than
the adsorbent with higher sulfur content (34%) (Wang et al.,
2009), but no justification has been provided in the paper.
However, the aggregation of sulfur within the large pores,
rather than the uniform distribution of sulfur on the activated
carbon, may account for this phenomenon. Also, Nabais et al.
have compared several sulfur introduction methods and
identified that mixing of activated carbon fibre (ACF) with
solid sulfur at a ratio of 1:3 (w/w), followed by treatment at
600e800  C resulted in the production of an adsorbent with
higher sulfur content compared with the introduction of sulfur to ACF via gas stream H2S. This may be reasonable due to
the melting, recrystallization and deposition of the solid sulfur on the activated carbon surface at such high temperatures.
However, subsequent mercury adsorption tests showed that
higher mercury uptake was obtained by the latter method
which elucidated the importance of the type of sulfur functional groups on the carbon surface besides its quantity
(Nabais et al., 2006). This may also occur due to the aggregation of the sulfur on the activated carbon when solid sulfur is
used as the surface modifying agent which diminishes the
effect of sulfur functional groups in mercury removal.
Furthermore, despite similar sulfur contents of K2S-impregnated coal samples were prepared at three distinct temperatures (800e1000  C), whereas the activated carbon sample
prepared at 900  C exhibited the highest and fastest adsorption (Wajima and Sugawara, 2011). This indicates that the
sulfur content on the adsorbent surface, type of sulfur functional groups and porous structure of the activated carbons
collectively influence their mercury adsorption efficiency.
In addition to higher capacity, exceptional affinity between
mercury and sulfur has been demonstrated in a multicomponent system of mercury, cadmium and lead where
highly-selective adsorption towards mercury was achieved
(Gomez-Serrano et al., 1998). The superior adsorption of
mercury on sulfur-grafted adsorbent is believed to originate
from the Pearson acid-base concept in which the hard acids
prefer to coordinate with hard bases and soft acids react in a
higher rate with soft bases. Accordingly, the soft acid mercury
species in the solution, such as HgCl2, (HgCl2)2, Hg(OH)2 and
HgOHCl tend to predominantly react with sulfur groups (soft

bases) on the adsorbent surface (Cai and Jia, 2010). This phenomenon has also been confirmed by comparing the interaction of HgX2 (where X is a halide) with C4H8O and C4H8S. It
has been reported that HgX2 interacts weakly with C4H8O, but
much stronger with C4H8S (Farhangi and Graddon, 1973;
 zquezet al. have related the
Fisher and Drago, 1975). Va
higher affinity towards mercury in a multi-component system
of cadmium, zinc and mercury to the higher electronegativity
 zquez et al., 2002).
of mercury (Va

4.

Effect of adsorption parameters

4.1.

Equilibrium contact time

Equilibrium contact time is the period of time required for the


adsorption and desorption processes to reach equilibrium.
When the equilibrium is reached, the amount of adsorption
from the solution to the adsorbent surface equals the amount
of desorption from the adsorbent surface to the solution and
no further increase in the uptake occurs. The adsorption
process involves several steps including mass transfer from
bulk fluid phase to the particle surface across the boundary
layer, adsorption on the surface of the adsorbent and diffusion
within the pores (Wang et al., 2011). Depending on which of
these steps is the rate determining step and also depending on
the boundary layer thickness and diffusion rate, the contact
time required to reach equilibrium will be different.
Adsorption of mercury has been shown to comply with a
general trend in which the mercury uptake rate is very fast at
the beginning because of the large number of vacant functional group sites on the surface of activated carbon available
for the mercury ions. As the sites are occupied in the course of
time, the uptake rate is gradually slowed down until a plateau
is reached upon equilibrium (Zabihi et al., 2009). Shorter
contact time required by the adsorbent to reach equilibrium is
economically more favorable in industry.
Kadirvelu et al. have suggested that the rate of adsorption
depends on several factors such as the type of precursor used
for adsorbent production, pore size and pore size distribution
and concentration of functional groups (Kadirvelu et al., 2004).
Namasivayam and Periasamy have reported that the activated
carbon from bicarbonate-treated peanut hull (BPHC) exhibited
7 times higher adsorption rate compared with commercial
activated carbon. They ascribed this high adsorption rate to
higher porosity and ion exchange ability of BPHC resulting in
less adsorption time required to acquire a certain mercury
removal percentage and thus more cost effectiveness
(Namasivayam and Periasamy, 1993). Also, rapid mercury
adsorption of less than 20 min to reach equilibrium was reported for activated carbons prepared from antibiotic waste
and rice husk ash as precursors (Budinova et al., 2008; Feng
et al., 2004). It can be hypothesized that as the pore size increases up to a certain extent, the diffusion path is reduced
and the adsorption rate increases. Also, higher concentration
of adsorption sites will increase the probability of contact
between mercury molecules and functional moieties and
therefore increase the uptake rate. Hence, it is believed that
more abundant adsorption sites with an optimum pore size
for mercury will increase the rate of mercury adsorption. More

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

studies are necessary to be conducted to prove these hypotheses regarding the relationship between mercury
adsorption rate and textural and surface properties of the
adsorbents.
It has been further demonstrated that as the initial concentration of mercury increases, the Lagergren rate constant
decreases and thus, longer time is required to achieve equilibrium (Namasivayam and Kadirvelu, 1999; Namasivayam
and Periasamy, 1993). This could be due to the saturation of
sites present on the exterior of adsorbent surface by adsorbate
at an initial stage of adsorption. Further adsorption can only
occur by the diffusion of the mercury ions into the pores and
adsorption in the interior surface of the pores which requires
relatively longer contact time (Hameed, 2007). Hence, in
modeling the mercury adsorption kinetics, a combination of
pseudo-type models with diffusion models is worthy of
consideration to elucidate the adsorption mechanism. However, for certain applications, these models have been shown
to be inconclusive (Plazinski et al., 2009).

4.2.

Initial concentration

In general, the mercury adsorption experiments display a


direct relationship between the metal uptake and initial concentration of the metal ions present in the solution up to a
certain limiting initial concentration and inverse relationship
between the removal percentage and initial metal concentration. An apparent distinction has to be drawn between
removal percent and adsorption capacity. The former term
does not reflect the efficiency of the material in mercury
removal at various initial concentrations and adsorbent dosages, whereas the latter takes into account the adsorbent
dosage and reveals the genuine mercury adsorption efficiency
of the material at different initial concentrations. Several authors have reported complete mercury removal (Mohan et al.,
2001; Rao et al., 2009; Wahi et al., 2009). But when the initial
concentration and adsorbent amount are taken into consideration, the adsorption capacity is found very small in some
cases. Therefore, reporting mercury removal percentage is
highly discouraged due to misleading results (Hadi et al.,
2015). As the initial concentration of mercury in the solution
increases, the percent removal of the adsorbate decreases
because of the presence of more mercury ions and limited
adsorption sites on the adsorbent materials. On the other
hand, at low mercury concentrations, the adsorption capacity
of the adsorbent material is low, while it increases by
increasing the initial concentration. This has been related to
the fact that at low mercury concentrations, the adsorption
sites are not completely occupied (Budinova et al., 2008, 2003;
Zabihi et al., 2009), whereas increasing the initial concentration of mercury results in higher collision probability between
the adsorbate molecules and adsorbent active sites, higher
occupation of active sites and thus higher adsorption capacity
(Zabihi et al., 2010). When the initial mercury concentration is
sufficiently increased, the adsorption capacity reaches a
plateau and does not increase anymore by increasing the
initial mercury concentration. This has been attributed to the
full occupation of the active sites on the surface of the
adsorbent at a certain initial concentration above which no
more adsorption enhancement can be achieved. Inbaraj and

45

Sulochana have observed a similar trend and have suggested


that this effect is caused by an increase in the driving force
offered by concentration gradient at high mercury concentrations (Inbaraj and Sulochana, 2006).

4.3.

pH value

Adsorption of mercury is a highly pH dependent process. As


the pH value of the solution increases, more mercury uptake
occurs. The increased adsorption of mercury ion has been
shown to be related to the species of mercury present in the
solution at various pH values and their solubility. Higher pH
values of the solution results in the presence of more soluble
mercuric species which, in turn, promotes the effective contact between the adsorbate molecules and the adsorbent
materials thus enhancing the possibility of the mercury uptake by the porous adsorbent particles (Adams, 1991; Lopes
et al., 2010; Namasivayam and Periasamy, 1993). Moreover,
lower pH values increase the solubility of the mercuric ions
and thus their subsequent desorption from the activated
carbon surface into the solution. Therefore, the relative
attraction between the adsorbent and adsorbate is lower than
between the adsorbate and the solvent phase at lower pH
values, leading to the lower adsorption of the mercuric ions.
Solution acidity also plays an important role in the ionization of the functional groups on the adsorbent surface. In
acidic environment, high concentration of hydronium ion
(H3O) in the solution drives the equilibrium ionization reaction (reaction R4) to the left and prevents the formation of
ionized functional groups, thereby hampering the ion exchange reaction between metal ions and adsorbent surface
functional groups (reaction R5). When the pH level of the solution increases to above 4, the hydronium ion concentration
in the solution decreases. This shifts the equilibrium reaction
R4 to the right resulting in the availability of more ionized
functional groups for ion exchange and therefore an increase
in the metal uptake (Eligwe et al., 1999).
Adsorbent  COOH4Adsorbent  COO H
aq

(R4)

Adsorbent  COO Mn
aq 4Adsorbent  COO M

(R5)

It is noteworthy that, the surface properties of adsorbent


can significantly affect the adsorption of mercury. The
adsorbent can be positively or negatively charged depending
on its point of zero charge (PZC). When the pH of the medium
is lower than the PZC, the adsorbent surface becomes positively charged leading to electrostatic repulsion of the mercury ions and the adsorbent surface and reduction in mercury
adsorption (Budinova et al., 2008; Rao et al., 2009).

4.4.

Temperature

Many researchers have shown that increasing the temperature results in higher mercury uptake due to the endothermic
nature of this process. Inbaraj and Sulochana have used the
thermodynamic parameters to study the effect of temperature
on the mercury adsorption behavior of fruit shell-based activated carbon and found a decrease in Gibbs free energy, DG, as
well as a positive enthalpy value, DH, by raising the

46

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

temperature which revealed that the adsorption process is


endothermic (Inbaraj and Sulochana, 2006). Also, Giles et al.
believe that high temperatures increase the mobility of the
mercuric ion and widen the pore on the sorbent surface
leading to enhanced intra-particle diffusion rate (Giles et al.,
1974). However, no justification has been provided regarding
the pore widening phenomenon. The effect could not be due
to physical widening, but apparent widening because of an
increase in compressibility of the particles with increasing
temperature. Since porous structure of carbonaceous materials usually forms at very high temperatures, it is particularly
implausible to alter the pore sizes at adsorption temperatures
as low as 20e80  C.
On the contrary, Mohan et al. have reported the exothermic
nature of the mercury adsorption using activated carbon
derived from fertilizer waste. This is confirmed with the
negative enthalpy value, DH, and an increase in Gibbs free
energy DG by an increase in temperature. They have related
this behavior to the physical adsorption mechanism of mercury by the adsorbent material. Physical adsorption caused by
the van der Waals forces between the adsorbent surface and
the adsorbate molecules is typically favored at low temperatures. This explains the higher adsorption capacity of the
fertilizer-based adsorbent for mercury at low temperatures
(Mohan et al., 2001).

4.5.

Adsorbent dosage and particle size

Well-documented researches have proven that an increase in


the dosage of adsorbent at a constant pH and adsorbate concentration has positive effect on the removal of pollutants
from wastewater (Gupta et al., 2003; Namasivayam et al.,
2001). Although many researchers have reported that the
mercury removal percent increases as the adsorbent dosage is
increased, as discussed in preceding sections, removal percentage is an entirely relative term changing by initial concentration of mercury and adsorbent dosage and thus it is not
appropriate to evaluate the efficiency of an adsorbent using
this parameter. Percentage mercury removal increased from
40% to nearly 100% when the C. pentandra hull adsorbent dose
increased from 25 to 200 mg (equal to 0.5 g:L1 and 4 g:L1 ,
respectively) (Rao et al., 2009). Increasing the dose of Indian
almond fruit shell from 0.05 to 5 g:L1 also led to a maximum
mercury removal of 99.5% (Inbaraj and Sulochana, 2006).
Similar mercury adsorption trends have been reported using
activated carbon from sago waste and commercial activated
carbon (Kadirvelu et al., 2004). Typically, an increase in the
adsorbent dosage results in the availability of higher surface
area and larger number of functional groups for ion exchange
in the system and leads to more chemisorption and/or physisorption as well as higher rate of adsorbate removal (Wahi
et al., 2009). It is suggested that more tangible adsorption capacities should be reported in this context instead of simply
quoting the percent removal.
On the other hand, although the removal percentage increases by increasing adsorbent load, the mercury adsorption
capacity of the adsorbent has been shown to steadily
decrease. This has been related to the decrease in the availability of mercury ions in aqueous phase per adsorbent site
and unsaturation of the adsorbent surface active sites. Rao

et al. used three types of adsorbents for mercury removal and


all of the adsorbents exhibited a decrease in the adsorption
capacity and an increase in the mercury removal percentage
by increasing the adsorbent dosage (Rao et al., 2009).
In addition, particle size also plays an important role in
altering the rate and capacity of mercury adsorption. It has
been demonstrated that when the size of the adsorbent particles decreased from 1.25e2.5 mm to 0.21e1 mm, the mercury
adsorption capacity showed a two-fold increase (430 mg:g1
versus 815 mg:g1 , respectively) (Mckay et al., 1989; PenicheCovas et al., 1992). Similarly, Kadirvelu et al. have also
demonstrated that a stepwise decrease of activated carbon
particle size produced from sago waste (750e500 mm,
500e250 mm and 250e125 mm) resulted in an increase in
mercury adsorption (85%, 90% and 93% removal, respectively)
(Kadirvelu et al., 2004). Similar results have also been reported
by Mohan et al. (Mohan et al., 2001) and Feng et al. (Feng et al.,
2004). It has been shown that reducing the adsorbent particle
size increases the effective surface area and enhances the
availability of adsorption sites (Kara et al., 2007). Also, the
diffusion path becomes shorter and the adsorbate molecules
can more easily penetrate into the internal pores of the
adsorbent (Gupta et al., 2011).

5.
Mercury affinity to various functional
groups
The adsorption of adsorbate by activated carbon can be categorized into chemical and physical adsorption. Briefly, physical adsorption is mediated by the weak van der Waal
interaction between the adsorbate and adsorbent, while
chemical adsorption is governed by the bonding between the
functional groups on the adsorbent surface and adsorbate.
Weak van der Waal interaction have been proven inefficient in
promoting mercury adsorption, however surface functional
groups, specially the oxygen containing groups of the adsorbent, exhibit a key role on the adsorption of mercury (Sun
et al., 2011). The behavior of enhanced mercury adsorption
by oxygen-containing functional groups has been explained
by the Lewis characteristic of Hg (II) which can be bonded to
the basic functional groups of the adsorbent surface (Nabais
et al., 2006). In the aqueous medium, the oxygen-containing
functional groups on the surface of the adsorbent tend to
lose their protons and become ionized, thus leading to unbalanced charge on the adsorbent surface where ion exchange
with the mercuric ion can occur (Sun et al., 2011). This accounts for the critical effect of pH level of the mercury-laden
solution on the adsorption of mercuric ions. As discussed in
the preceding sections, changing the pH level significantly
changes the adsorbent surface charge, thus resulting in a
considerable difference in the adsorption capacity of the
adsorbent. Moreover, electron lone pairs on nitrogencontaining functional groups can interact with mercury ions
and assist in their removal (Zhu et al., 2009).
Although functional groups of the activated carbons have
long been considered to be crucial in chemisorption, the effect
of oxygen-containing functional group quantity on mercury
adsorption capacity and rate has not been comprehensively
examined. It is also noteworthy that despite an inverse

47

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

relationship between the percentage of oxygen functional


groups and the total surface area of the adsorbent material,
both of which are regarded positive factors for mercury
adsorption capacity, no trade-off graph between these two
crucial parameters has been provided to optimize the efficiency of adsorption.
As functional groups largely determine the surface properties and thus the intensity of ion exchange, manipulation of
the surface functional group have been of great interest. Sulfur group has been widely reported to promote mercury
adsorption. Rao et al. have observed an increase in the mercury adsorption capacity of activated carbon with the introduction of sulfur groups on the activated carbon surface and
ascribed the higher removal efficiency of the sulfurcontaining activated carbon to the interaction of various
Hg(II) species, such as HgCl2, (HgCl2)2, Hg(OH)2 and HgOHCl,
with surface sulfur groups. The following redox reaction has
been proposed as the adsorption mechanism of the activated
carbon for mercury (Rao et al., 2009).

2
2
2Hg2 SO2
3 2OH 4Hg2 SO4 H2 O

(R6)

This reaction is in good agreement with the effect of pH


value on the adsorption capacity, where increasing the pH
level of the solution increases the hydroxide ion content,
driving the reaction to the right side, and thus leads to a higher
mercury adsorption capacity.
Enhancement in the removal of mercury has also been
carried out by grafting thiol group onto the surface of the
activated coke, where the adsorption capacity has been
increased from 315.8 mg:g1 for the unmodified material to
694.9 mg:g1 for the modified material (Li et al., 2013). Anoop
Krishnan and Anirudhan have verified the effect of sulfur

modification by H2S and SO2 on the adsorption capacity of


mercury (Anoop Krishnan and Anirudhan, 2002). They
observed that irrespective of the type of modifying agent
(either H2S or SO2), the mercury adsorption capacity of the
activated carbon increases. This can be due to the similar
types of sulfur functional groups doped on the adsorbent
surface by gas surface modification. These results are
different from the gas-phase adsorption of mercury which can
be related to the different mechanism in gas-phase and
aqueous-phase mercury adsorption (Feng et al., 2006b).
Studies concerning the effect of activation parameters on
the type and quantity of the functional groups are listed in
Table 3. Toles et al. have identified that the type of precursor
has minor effect on the functional groups of the produced
activated carbons and implicated the importance of activation
temperature in the formation of the functional groups (Toles
et al., 1999). The oxygen-containing functional groups can be
formed by exposing the carbonaceous precursor to oxygen at
temperatures between 200 and 700  C (Bansal et al., 1988).
Also, more carbonyl groups can be formed by oxidizing the
activated carbon at 400  C, but subsequent oxidization of the
activated carbon destroys the carbonyl group and produces
more phenol, lactones and carboxylic acid group (Toles et al.,
1999). Furthermore, it has been observed that the carboxylic
groups begin to decompose at 200e500  C and all the acid sites
are destroyed at 700  C (Budinova et al., 2008). Although such
manipulation can be carried out on the surface functional
groups of activated carbons, it can be criticized that there is no
comprehensive study to compare the effects of various functional groups on mercury removal.
The type of the activating agent has also been considered
effective in the manipulation of the surface functional groups

Table 3 e Acid-base neutralization capacity (meq/g) of the activated carbon adsorbents.


Precursor material

Mengen
Seyitomer
Some
Bulluca
Apricot Stones
Furfural
Furfural
Mixture of steam pyrolysis
tar and furfural (30:70)
Air oxidized Furfural
Woody biomass birch

Walnut shell

Coconut activated carbon

Sago
Furfural
Antibiotic waste
a

Below detection limit.

Name

Carbon A
Carbon B
Carbon C
N600-1
NS600-1
S700-2
Walnut shell
Carbon A
Carbon B
AC
AC1
AC1-1
AC1-2

Base uptake

Acid uptake

Reference

NaHCO3

Na2CO3

NaOH

EtONa

HCL

0.092
e
0.100
e
0.130
0.120
0.120
0.030

0.120
0.120
0.110
0.110
0.210
0.160
0.160
0.080

0.184
0.250
0.183
0.320
0.360
0.230
0.230
0.250

1.900
2.040
1.570
1.900
1.350
1.500
1.500
1.300

1.120
2.670
1.120
2.010
0.842
0.600
0.600
0.560

1.900
0.744
0.124
e
0.450
0.540
0.720
1.097
1.174
1.295
1.328
1.200
0.120
BDLa

1.930
0.126
0.034
0.123
0.490
0.480
0.420
0.627
0.594
0.341
0.099
1.800
0.160
BDL

3.660
0.480
0.572
0.422
0.390
0.350
0.300
0.561
0.693
0.726
0.869
0.900
0.230
0.230

6.340
2.234
2.530
2.355
0.520
0.420
0.290
0.495
0.341
0.572
0.594
1.600
1.500
2.300

e
0.083
1.100
0.902
0.520
0.420
0.290
e

(Lu et al., 2014)

1.100
0.600
1.300

(Kadirvelu et al., 2004)


(Yardim et al., 2003)
(Budinova et al., 2008)

(Ekinci et al., 2002)

(Budinova et al., 2003)

(Budinova et al., 2006)

(Zabihi et al., 2009)

48

Table 4 e Equilibrium model constants for the Hg adsorption by activated carbons.


Activated carbon

Isotherm models tested

Langmuir

Freundlich

Temp ( C)

qm (mg g1)

aL (L mg1)

R2

KF

1=n

R2

e
1.0
0.2
0.2
142.7
66.3
10.7
1.7
28.8

e
1.110
6.300
1.060
0.710
0.750
1.380
0.221
1.818
e

0.927
0.925
0.962
0.978
0.923
0.942
0.955
0.984
0.987

Surface area
(m2 g1)

Langmuir and Freundlich


Langmuir and Freundlich

30
e

52.7

0.0260
e

0.9010

Langmuir and Freundlich

Activated sago waste carbon


Activated furfural
Carbon A (furfural)
Carbon B (Apricot stone and furfural
(30:70) (w/w))
Carbon C (Air oxidized Furfural)
Impregnation of Walnut shell
with ZnCl2 at 1:0.5 (w/w) (Carbon A)
Impregnation of Walnut shell
with ZnCl2 at 1:1 (w/w) (Carbon B)
Activated coirpith

Langmuir and Freundlich


Langmuir and Freundlich
Langmuir

27
45
65
30
RT
RT

3.6
1.9
0.4
55.6
174
174
154

0.0010
0.0006
0.0004
0.3750
e
1.4000
0.5400

0.9927
0.9903
0.9896
0.9999
0.9992
0.9976
0.9999

Langmuir and Freundlich

RT

134
151.5

0.1100
0.0091

0.9994
0.9981

1.9

0.823

0.997

480
780

100.9

0.0087

0.9998

1.2

0.820

0.995

803

154

0.6300

0.9991

50

0.549

0.998

592

109.9

0.3640

0.9956

42.2

0.286

0.994

208.1

30

12.4

0.5050

0.9949

4.7

0.316

0.990

354.6

Langmuir and Freundlich

32

94.4

0.4900

0.9956

38.3

4.000

0.913

4.6

Langmuir
Langmuir

RT
e

129
92
56
105
37
153
174
160

0.5000
0.3400
1.4600
1.1900
0.1800
0.2300
1.4000
e

0.9949

0.9992
0.9957

25.9
23.7
22.9
60.1
121.6
54

0.4500
0.5100
0.3600
0.0503
0.0993
0.0260

0.8167
0.9016
0.9273
0.9494
0.9115
0.9700

11.2
9.5
8.4
7.3
21.5
12.3

0.243
0.268
0.275
2.205
2.427
0.240

0.969
0.966
0.966
0.965
0.912
0.910

521
325
280
650
e
732

112

0.0096

0.9300

5.7

0.480

0.960

1222

455

0.0050

0.9000

8.8

0.600

0.900

828

Bicarbonate-treated peanut
hulls (BPHC)
Commercial granular activated
carbon (CAC)
Activated Terminalia catappa
Mixture of antibiotic production waste
Mengen-coal
Seyitomer-coal
Some-coal
Bolluca-coal
Apricot Stone
Furfural
Woody biomass impregnated with
H3PO4 and pyrolized at 700

C in steam
C. pentandra
P. aureus
C. arietinum
Activated carbon (AC)
AC with ethylenediamine (MAC)
Sulfur-impregnated activated carbon
treated with KOH
Sulfur-impregnated activated
carbon treated with KOH & SO2
Activated carbon (AC)

Langmuir and Freundlich

RT

Langmuir and Freundlich

Langmuir

20

Langmuir and Freundlich

30

Langmuir and Freundlich

25

Langmuir and Freundlich

25

Langmuir and Freundlich

30

379.4
1.4
0.99
1.07
629

625
1100
1100
1050

1260
600
460
720
520
1000
1100
1360

e
e

0.9997

(Wahi et al., 2009)


~ a-Rodrguez
(Pen
et al., 2013)
(Mohan et al., 2001)

(Kadirvelu et al., 2004)


(Yardim et al., 2003)
(Budinova et al., 2003)

(Zabihi et al., 2010)

(Namasivayam
and Kadirvelu, 1999)
(Namasivayam
et al., 1993)

(Inbaraj and
Sulochana, 2006)
(Budinova et al., 2008)
(Ekinci et al., 2002)

(Budinova et al., 2006)

(Rao et al., 2009)

(Zhu et al., 2009)


(Cai and Jia, 2010)

(Asasian et al., 2014)

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

Palm oil empty fruit bunch


Calcineed mussel shell
Finely groung mussel shell
Coarsely ground mussel shell
Waste slurry

Reference

Sulfur impregnation of (AC) with


4%(v/v) SO2 at 700  C for 60 min
Steam activation of bagasse
pith (SA-C)

Langmuir

Steam activation of bagasse


pith in presence of SO2
(SAeSO2eC)
Steam activation of bagasse
pith in presence of H2S
(SAeH2SeC)

Rice husk ash

Langmuir and Freundlich

Sulfuric acid treated with rice


husk (dry sorbent)

Langmuir and Freundlich

Sulfuric acid treated with rice


husk (wet sorbent)
Sulfuric acid treated with
flax shave (dry sorbent)
Sulfuric acid treated with
flax shave (wet sorbent)
Commercial activated carbon (AC)

Langmuir

Langmuir and Freundlich

Steam activated AC (AC-1)

AC1 was oxidized with H2O2


at 1:2 (m:v)
(AC1-1)
AC1 was oxidized with H2O2
at 2:5 (m:v)
(AC1-2)
PSAC grafted with TOMATS

Langmuir and Freundlich

0.0090
0.0080
0.0220
0.0740
0.0860
0.0072
0.008
0.0086
0.0106
0.0195
0.0202
0.0229
0.0262
0.0113
0.0123
0.0164
0.0229
0.0281
0.0273
0.0367
0.0480
0.0115
0.0158
0.0052
0.0107
0.0219
0.0052
0.0088
0.0129
0.0805

0.9800
0.8900
0.9600
0.9200
0.9500
e

0.9868
0.9900
0.9990
0.9992
0.9988
0.9991
0.9998
0.9993
0.9980

344

0.0468

0.9990

4.1
3.5
3
4.9
4.5
4.1
5.0
5
4.6
5.2
5.1
4.6
76.9
76.9
83.3
83.3
315.8
694.9

0.2083
0.1053
0.1330
0.6870
0.1540
0.3862
0.2329
0.2000
0.2618
0.7795
0.4491
0.9287
0.0588
0.0778
0.1100
0.1250
0.0500
0.0600

0.9522
0.9827
0.9828
0.9889
0.9914
0.9866
0.9828
0.9914
0.9949
0.9835
0.9841
0.9521
0.9920
0.9940
0.9930
0.9910
0.9770
0.9840

15.9
15
71.2
129.2
148.2

0.42
0.54
7.1
28.7
48.9
8.7
15.0
18.1

0.560
0.600
0.330
0.250
0.230
e

0.940
0.890
0.990
0.980
0.960

0.493
0.469
0.5579
0.377
0.3407
0.4607
0.428
0.448

0.973
0.965
0.9812
0.987
0.986
0.9523
0.928
0.959

751.3

(Anoop Krishnan
and Anirudhan, 2002)

311

(Feng et al., 2004)

66

(El-Shafey, 2010)

19

(Cox et al., 2000)

e
0.8
0.3
0.4
2.8
0.6
1.4
1.1
1
1.6
3.4
2.3
2.7
5.5
6.9
9.4
10.2
12.2
71.1

0.974
1.262
1.150
0.790
1.248
0.750
0.914
0.858
0.857
0.823
1.031
0.720
0.630
0.590
0.560
0.570
0.714
0.526

0.956
0.971
0.970
0.975
0.982
0.952
0.986
0.984
0.985
0.976
0.979
0.980
0.916
0.906
0.892
0.898
0.980
0.954

797

(Lu et al., 2014)

870

829

825

107.7

(Ismaiel et al., 2013)

136.7
193.7

(Li et al., 2013)

49

Activated coke (AC)


Thiol-functionalized activated coke (SH-AC)

Langmuir and Freundlich

10
25
50
10
25
50
10
25
50
10
25
50
20
25
30
35
25

472
578
523
496
510
172.4
181.8
200
208.3
185.2
204.1
208.3
222.2
181.8
200
204.1
217.4
188.7
208.3
212.8
227.3
9.3
6.7
303
336.7
384.6
227.3
270.3
303
416

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

Steam activation of bagasse


pith in presence of SO2 & H2S
(SAeSO2eH2SeC)

45
60
30
45
60
30
40
50
60
30
40
50
60
30
40
50
60
30
40
50
60
15
30
25
35
45
25
35
45
45

50

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

for better mercury uptake. Nabais et al. have demonstrated


that the type of sulfur doped onto the adsorbent surface is a
more critical factor than its quantity. They observed that the
increase in the sulfur content of the activated carbon using
solid sulfur as the modifying agent did not improve mercury
uptake, whereas the modification of the adsorbent by H2S
resulted in a considerable increase in mercury adsorption.
They related this phenomenon to better accessibility of the
sulfur to mercury by gas modification compared with the solid
modification (Nabais et al., 2006).
The activation atmosphere can also affect the surface
functionality of activated carbon. Budinovaet al. have found
that activated carbons pyrolyzed under nitrogen atmosphere
have high carboxylic group on the material surface, but
consecutive pyrolysis and steam activation results in a significant drop in the content of carboxylic and lactone groups
and formation of more hydroxyl and carbonyl groups
(Budinova et al., 2006). It has also been highlighted that activation in the presence of air and water vapor results in an
increase and decrease of oxygen content of the final modified
product, respectively (Budinova et al., 2003).
Besides physical activation, chemical activation can also
alter the functional groups on the activated carbon surface.
Most activated carbons contain varying amounts of functional
groups such aseOH, eCH]O and COOH without any treatment. When activated carbons are treated with oxidizing
agent such as HNO3, H2O2, or (NH4)2S2O8, chemical reaction
occurs between the activating agent and the adsorbent surface which alters the surface functionality and pKa of the
activated carbons as well as their porous structure and
adsorption capacity (Bandosz et al., 1993; Montagnaro and
Santoro, 2009).
X-ray photoelectron spectroscopy has demonstrated that
that oxygen- and nitrogen-containing functional groups act
as electron donors during mercury adsorption and it has
been hypothesized that chemical coordination of mercury
with these functional groups are accountable for mercury
adsorption (Zhu et al., 2009). Therefore, higher oxygen- and
nitrogen-containing functional groups favor the mercury
adsorption. Zhu et al. studied the effect of activating agent
and chemical activation time on the surface functionality of
the activated carbons and detected the formation of hydroxyl, carboxylic and carboxylic anhydride group by nitric
acid treatment (Zhu et al., 2009). When the contact time between activated carbon and the nitric acid increases, significant amount of phenolic group forms while the content of
lactone group is reduced. Increase in the concentration of
nitric acid also leads to the formation of higher carboxylic
acid and carbonyl group content (Xianglan et al., 2011).
Xianglan et al. have investigated the use of hydrogen
peroxide as modifying agent and found that the lactone
moieties are decomposed into carbonyl and phenol groups by
increasing the chemical concentration and reaction time
(Xianglan et al., 2011). Danish et al. have explored the effect
of two chemical agents and have found that phosphoric acidtreated activated carbon contains higher amount of acidic
functional group as compared with using zinc chloride
(Danish et al., 2013). However, when acid treatment becomes
more intense, no further oxygen functional groups, crucial
for the adsorption capacity of the activated carbons, are

detected. Thus altering the type and concentration of the


activating agent is an important factor to be studied for
adsorption purposes. Ahmad et al. used 2 M hydrochloric acid
to treat cocoa shell for 2 h at a relatively high temperature
and found that, despite the high surface area obtained, the
oxygen functional group was not detected. It has been hypothesized that the oxygen attached to the minerals are
removed by intense acid treatment (Ahmad et al., 2013).

6.

Equilibrium adsorption isotherms

Langmuir and Freundlich isotherm models have been mostly


applied to describe the equilibrium adsorption of mercury (II)
on the adsorbent.
The Langmuir isotherm model was originally developed to
describe gasesolid adsorption onto adsorbent. The model
assumes irreversible homogeneous monolayer adsorption
and each adsorbate being adsorbed only to one adsorption
site. It also assumes that all the adsorption sites are identical.
Therefore, the affinity of each adsorbate to adsorbent is
equivalent resulting in constant enthalpies and sorption
activation energy, without lateral interaction and steric hindrance between the adsorbed molecules on the adjacent sites
(Langmuir, 1918). The mathematical expression for this model
can be represented as:
qe

qm aL Ce
1 aL Ce

(1)

where qe is the adsorption capacity of the adsorbent (mg:g1 );


qm is the maximum adsorption capacity of the adsorbent
(mg:g1 ); Ce is the equilibrium concentration of the adsorbate
in the solution (mg:L1 ); and aL is the Langmuir constant. A
mathematical expression developed by Webber and Chakkravorti has been used to test the favourability of the adsorption process (Weber and Chakravorti, 1974):
RL

1
1 aL C0

(2)

where RL, called separation factor, is a dimensionless term


and Co is the adsorbate initial concentration (mg:L1 ). The
separation factor (RL) indicates the adsorption nature to be
unfavorable (RL > 1), linear (RL 1), favorable (0 < RL < 1), or
irreversible (RL 0).
The Freundlich model, on the other hand, has been used to
fit a heterogeneous surface with non-identical adsorption
sites (Freundlich, 1906). This empirical model can be applied
to multilayer adsorption with uneven distribution of heat and
affinities when adsorption occurs over the heterogeneous
surface (Adamson and Gast, 1967). Although Freundlich model
is a widely-used empirical isotherm model to fit the experimental adsorption results, it has been criticized for its
inability to comply with the Henry's law at low adsorbate
concentrations (Ho et al., 2002). The expression of the
Freundlich isotherm can be described as:
1

qe KF Cen

(3)

where KF is the Freundlich constant and 1=n is the heterogeneity factor. When the value of n is between 1 and 10, the

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

adsorption process is favorable and the adsorbent surface is


considered heterogeneous (Jodeh et al., 2014).
Although both the Langmuir and Freundlich isotherm
models are popular tools in the prediction of equilibrium
adsorption isotherm, these models are sometimes criticized
due to their limitations of oversimplified assumptions for the
former and lack of fundamental basis for the latter isotherm
model. In addition, other hybrid forms of the Langmuir and
Freundlich adsorption models are also established. RedlichPeterson and Sips isotherm models are the modified
isotherm models that incorporate the features of both the
Langmuir and Freundlich equations.
The Redlich-Peterson model can be applied either in homogeneous and heterogeneous adsorption (Redlich and
Peterson, 1959):
qe

qm aR Ce
1 aR Cbe

(4)

where aR is the Redlich-Peterson isotherm constant and b is


the Redlich-Peterson isotherm exponent, which lies between
0 and 1. The two extremes of the exponent b transform this
equation to the Henry's law and Langmuir equations when its
value is the lowest and highest, respectively. For any other
exponent values, this equation can be considered as an
incorporation of the features of the Langmuir and Freundlich
models.
Sips isotherm model is applied for the prediction of heterogeneous adsorption system and assumes the occurrence of
dissociative adsorption (Sips, 1948). At low adsorbate concentration, the isotherm can be reduced to Freundlich
isotherm, while at high concentration, it complies with homogenous adsorption characteristic of the Langmuir isotherm
(Diaz et al., 2007). The Sips isotherm is still criticized not to
follow the Henry's law at low adsorbent concentration. The
Sips isotherm model is expressed through the following
equation:
qe

qm aS Cbe s
1 aS Cbe s

(5)

where as is the Sips isotherm model constant and bs is the Sips


isotherm model exponent.
Table 4 summarizes the Langmuir and Freundlich constants obtained for the removal of aqueous mercury by
various adsorbent materials. Pena-Rodriguez et al. have
tested the mercury adsorption behavior of three adsorbents
obtained from calcined mussel shell, finely ground mussel
shell and coarsely ground mussel shell and identified that
the KF value obtained from Freundlich isotherm model is
not closely correlated with the surface area, given that the
highest KF value corresponds to the calcined shell with
~ asurface area lower than that of finely ground mussel (Pen
Rodrguez et al., 2013). Cai and Jia showed the S-shaped
relationship between SBET and mercury adsorption and
suggested that the high porosity contributed by micropore
might not be accessible to mercury and its species, such as
HgCl2 and HgOHCl, instead a better linear correlation was
found between the adsorption capacity and mesopore surface area (Cai and Jia, 2010). This phenomenon suggested
that the surface area alone is not sufficient to reflect the

51

adsorption capacity of the adsorbent, but several other


factors such as pore size, surface structure and adsorbate
species will also interfere with the overall adsorption
capacity.
Cai and Jia have found a positive correlation between the
mercury adsorption capacity of activated carbons and its
sulfur content (Cai and Jia, 2010). Wang et al. have also
determined that although sulfur impregnation results in significant reduction in SBET of the activated carbons, a staggering enhancement in the mercury adsorption capacity was
achieved (Wang et al., 2009). The strong affinity between
mercury and sulfur has also been reported by Asasian et al.
(Asasian et al., 2014).
Further information could be deduced regarding the
mechanism of mercury adsorption by comparing the
isotherm shapes as a means of fitting mercury adsorption
isotherms, and correlating these with Giles isotherm classification (Giles et al., 1960). To date, no authors have attempted
to do this for aqueous mercury adsorption.

7.

Conclusion and future perspectives

Removal of mercury from wastewater using activated carbons


has been shown to be very promising if proper combination of
properties is possessed by the adsorbent materials. The effects of surface area and functional groups of the activated
carbons on mercury uptake have been examined in numerous
studies. In this study, the generic misconception that higher
surface area of an adsorbent leads to higher mercury
adsorption has been criticized. Herein, it has been demonstrated that a combination of medium-to-high surface area
with well-functionalized surface properties are collectively
crucial in enhancing the mercury removal. Despite these
findings, very limited research has been carried out on
simultaneous optimization of surface and textural characteristics of activated carbons. Hence, further study is necessary for such optimization to save the high amount of energy
required to obtain unnecessary very high surface area activated carbons.
Furthermore, although a lot of studies are concerned with
sulfurization of activated carbons with the aim of higher
mercury uptake, the mechanism of sulfurization process,
including the type and quantity of sulfur-containing moieties
doped onto the activated carbon surfaces and their functionalization process, and consequently the mercury adsorption
mechanism are not thoroughly examined. A profound insight
into the activation and adsorption mechanisms will assist in
designing a proper adsorbent-adsorbate system for optimal
mercury abatement from effluents.
In addition, the mercury adsorption is believed to take
place in two stages; initially the surface active sites are
involved in the adsorption process and when the surface sites
are less available, the mercuric ions have to diffuse into the
pores. Therefore, a combination of pseudo and diffusion
models has to be considered for modeling the mercury
adsorption kinetic results. Nevertheless, this modeling strategy has been disregarded.
Further recommendations can be presented as follows:

52

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

 Mercury forms complexes in real wastewater systems and


is rarely found in ionic form. Therefore, study of the efficiency of the activated carbon adsorbents using industrial
wastewater seems to be of high priority.
 The presence of other metallic compounds in the effluent
will unequivocally affect the mercury removal efficiency of
the activated carbon samples. Therefore, more detailed
study of the multi-component adsorption systems have to
be carried out.
 The regeneration of the activated carbon samples has to be
conducted for economic feasibility enhancement. Unfortunately, in adsorption processes, the regeneration is
overlooked.
 If not regenerated, due to its hazardous nature, mercuryloaded activated carbon needs to be stabilized or vitrified
and then disposed of in hazardous landfill. Nonetheless,
this landfilled carbon is not reusable, therefore emphasizing the importance of the production cost of the activated carbon.
 Typically, the focus of the research in mercury adsorption
systems is lab-scale batch adsorption studies. However,
the studies should not be confined only to these lab-scale
experiments and column studies have to be performed
for better understanding the industrial-scale operation
challenges.

references

Adams, M.D., 1991. The Mechanisms of Adsorption of Hg ( CN ) 2


and HgC12 on to Activated Carbon, vol. 26, pp. 201e210.
Adamson, A.W., Gast, A.P., 1967. Physical Chemistry of Surfaces.
Aguado, J., Arsuaga, J.M., Arencibia, A., 2005. Adsorption of
aqueous Mercury(II) on propylthiol-functionalized
mesoporous silica obtained by cocondensation. Ind. Eng.
Chem. Res. 44, 3665e3671.
Ahmad, F., Daud, W.M.A.W., Ahmad, M.A., Radzi, R., 2013. The
effects of acid leaching on porosity and surface functional
groups of cocoa (Theobroma cacao)-shell based activated
carbon. Chem. Eng. Res. Des. 91, 1028e1038.
Anirudhan, T.S., Divya, L., Ramachandran, M., 2008. Mercury(II)
removal from aqueous solutions and wastewaters using a
novel cation exchanger derived from coconut coir pith and its
recovery. J. Hazard. Mater. 157, 620e627.
Anoop Krishnan, K., Anirudhan, T.S., 2002. Removal of
mercury(II) from aqueous solutions and chlor-alkali industry
effluent by steam activated and sulphurised activated carbons
prepared from bagasse pith: kinetics and equilibrium studies.
J. Hazard. Mater. 92, 161e183.
Antochshuk, V., Jaroniec, M., 2002. 1-Allyl-3-propylthiourea
modified mesoporous silica for mercury removal. Chem.
Commun. 258e259.
Asasian, N., Kaghazchi, T., Faramarzi, A., Hakimi-Siboni, A.,
Asadi-Kesheh, R., Kavand, M., Mohtashami, S.-A., 2014.
Enhanced mercury adsorption capacity by sulfurization of
activated carbon with SO2 in a bubbling fluidized bed reactor.
J. Taiwan Inst. Chem. Eng. 45, 1588e1596.
Bandaru, N.M., Reta, N., Dalal, H., Ellis, A.V., Shapter, J.,
Voelcker, N.H., 2013. Enhanced adsorption of mercury ions on
thiol derivatized single wall carbon nanotubes. J. Hazard.
Mater. 261, 534e541.
Bandosz, T.J., Jagieuo, J., Schwarz, J.A., 1993. Effect of Surface
Chemical Groups on Energetic Heterogeneity of Activated
Carbons, pp. 2518e2522.

Bansal, R.C.C., Donnet, J.-B.J.-B., Stoeckli, F., 1988. Active Carbon.


Macel Dekker, New York.
Barron-Zambrano, J., Laborie, S., Viers, P., Rakib, M., Durand, G.,
2002. Mercury removal from aqueous solutions by
complexationdultrafiltration. Desalination 144, 201e206.
Barron-Zambrano, J., Laborie, S., Viers, P., Rakib, M., Durand, G.,
2004. Mercury removal and recovery from aqueous solutions
by coupled complexationeultrafiltration and electrolysis. J.
Memb. Sci. 229, 179e186.
Blue, L.Y., Jana, P., Atwood, D.A., 2010. Aqueous mercury
precipitation with the synthetic dithiolate, BDTH2. Fuel 89,
1326e1330.
Botta, S.G., Rodriguez, D.J., Leyva, A.G., Litter, M.I., 2002. Features
of the transformation of Hg II by heterogeneous
photocatalysis over TiO2. Catal. Today 76, 247e258.
rnbom, E.,
Budinova, T., Ekinci, E., Yardim, F., Grimm, a, Bjo
Minkova, V., Goranova, M., 2006. Characterization and
application of activated carbon produced by H3PO4 and water
vapor activation. Fuel Process. Technol. 87, 899e905.
Budinova, T., Petrov, N., Parra, J., Baloutzov, V., 2008. Use of an
activated carbon from antibiotic waste for the removal of
Hg(II) from aqueous solution. J. Environ. Manage 88, 165e172.
Budinova, T., Savova, D., Petrov, N., Razvigorova, M., Minkova, V.,
Ciliz, N., Apak, E., Ekinci, E., 2003. Mercury adsorption by
different modifications of furfural adsorbent. Ind. Eng. Chem.
Res. 42, 2223e2229.
Byrne, H.E., Mazyck, D.W., 2009. Removal of trace level aqueous
mercury by adsorption and photocatalysis on silica-titania
composites. J. Hazard. Mater. 170, 915e919.
Cai, J.H., Jia, C.Q., 2010. Mercury removal from aqueous solution
using coke-derived sulfur-impregnated activated carbons. Ind.
Eng. Chem. Res. 49, 2716e2721.
Chiarle, S., Ratto, M., Rovatti, M., 2000. Mercury removal from
water by ion exchange resins adsorption. Water Res. 34,
2971e2978.
Choi, M., Jang, J., 2008. Heavy metal ion adsorption onto
polypyrrole-impregnated porous carbon. J. Colloid Interface
Sci. 325, 287e289.
Cox, M., El-shafey, E.I., Pichugin, A.A., Appleton, Q., 2000. Removal
of Mercury ( II ) from Aqueous Solution on a Carbonaceous
Sorbent Prepared from Flax Shive, vol. 435, pp. 427e435.
Cui, H., Qian, Y., Li, Q., Wei, Z., Zhai, J., 2013. Fast removal of Hg(II)
ions from aqueous solution by amine-modified attapulgite.
Appl. Clay Sci. 72, 84e90.
Danish, M., Hashim, R., Ibrahim, M.N.M., Sulaiman, O., 2013.
Effect of acidic activating agents on surface area and surface
functional groups of activated carbons produced from Acacia
mangium wood. J. Anal. Appl. Pyrolysis 104, 418e425.
re, E.M., Leyva, A.G., Gautier, E.A., Litter, M.I., 2007.
De la Fournie
Treatment of phenylmercury salts by heterogeneous
photocatalysis over TiO(2). Chemosphere 69, 682e688.
Di Natale, F., Erto, A., Lancia, A., Musmarra, D., 2011. Mercury
adsorption on granular activated carbon in aqueous solutions
containing nitrates and chlorides. J. Hazard. Mater. 192,
1842e1850.
Di Natale, F., Lancia, a, Molino, a, Di Natale, M., Karatza, D.,
Musmarra, D., 2006. Capture of mercury ions by natural and
industrial materials. J. Hazard. Mater. 132, 220e225.
Diaz, E., Ordonez, S., Vega, a, 2007. Adsorption of volatile organic
compounds onto carbon nanotubes, carbon nanofibers, and
high-surface-area graphites. J. Colloid Interf. Sci. 305, 7e16.
Ebinghaus, R., Turner, R., de Lacerda, L., Vasiliev, O.,
Salomons, W., 1999. Mercury Contaminated Sites:
Characterization, Risk Assessment and Remediation.
Ekinci, E., Budinova, T., Yardim, F., Petrov, N., Razvigorova, M.,
Minkova, V., 2002. Removal of mercury ion from aqueous
solution by activated carbons obtained from biomass and
coals. Fuel Process. Technol. 77e78, 437e443.

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

Eligwe, C.A., Okolue, N.B., Nwambu, C.O., Nwoko, C.I.A., 1999.


Adsorption thermodynamics and kinetics of mercury (II),
cadmium (II) and lead (II) on lignite. Chem. Eng. Technol. 22,
45e49.
El-Shafey, E.I., 2010. Removal of Zn(II) and Hg(II) from aqueous
solution on a carbonaceous sorbent chemically prepared from
rice husk. J. Hazard. Mater. 175, 319e327.
Farhangi, Y., Graddon, D.P., 1973. Thermodynamics of metalligand bond formation. V. Adducts of mercury (II) halides with
Lewis bases. Aust. J. Chem. 26, 983e989.
Feng, Q., Lin, Q., Gong, F., Sugita, S., Shoya, M., 2004. Adsorption
of lead and mercury by rice husk ash. J. Colloid Interface Sci.
278, 1e8.
Feng, W., Borguet, E., Vidic, R.D., 2006a. Sulfurization of a carbon
surface for vapor phase mercury removal e II: sulfur forms
and mercury uptake. Carbon N. Y. 44, 2998e3004.
Feng, W., Kwon, S., Feng, X., Borguet, E., Vidic, R.D., Asce, M.,
2006b. Sulfur impregnation on activated carbon fibers through
H2S oxidation for vapor phase mercury removal. J. Environ.
Eng. 132, 292e300.
Fisher, K.J., Drago, R.S., 1975. Trends in the acidities of the zinc
family elements. Inorg. Chem. 14, 2804e2808.
Fouladi Tajar, A., Kaghazchi, T., Soleimani, M., 2009. Adsorption
of cadmium from aqueous solutions on sulfurized activated
carbon prepared from nut shells. J. Hazard. Mater. 165,
1159e1164.
Freundlich, H.M.F., 1906. Over the adsorption in solution. J. Phys.
Chem. A 57, 385e470.
Gash, A.E., Spain, A., Dysleski, L.M., Flaschenriem, C.J.,
Kalaveshi, A., Dorhout, P.K., Strauss, S.H., 1998. Efficient
recovery of elemental mercury from Hg (II)-Contaminated
aqueous Media using a material. Environ. Sci. Technol. 32,
1007e1012.
Giles, C.H., MacEwan, T.H., Nakhwa, S.N., Smith, D., 1960. Studies
in adsorption. Part XI. A system of classification of solution
adsorption isotherms, and its use in diagnosis of adsorption
mechanisms and in measurement of specific surface areas of
solids. J. Chem. Soc. 3973e3993.
Giles, C.H., Smith, D., Huitson, A., 1974. A general treatment and
classification of the solute adsorption isotherm. I. Theoretical.
J. Colloid Interface Sci. 47, 755e765.
Gomez-Serrano, V., Macias-Garcia, A., Espinosa-Mansilla, A.,
Valenzuela-Calahorro, C., 1998. Adsorption of mercury,
cadmium and lead from aqueous solution on heat-treated and
sulphurized activated carbon. Water Res. 32, 1e4.
Gupta, V., Ali, I., Mohan, D., 2003. Equilibrium uptake and
sorption dynamics for the removal of a basic dye (basic red)
using low-cost adsorbents. J. Colloid Interface Sci. 265,
257e264.
Gupta, V.K., Gupta, B., Rastogi, A., Agarwal, S., Nayak, A., 2011. A
comparative investigation on adsorption performances of
mesoporous activated carbon prepared from waste rubber tire
and activated carbon for a hazardous azo dye-Acid Blue 113. J.
Hazard. Mater. 186, 891e901.
Hadi, P., Barford, J., McKay, G., 2013a. Toxic heavy metal capture
using a novel electronic waste-based material-mechanism,
modeling and comparison. Environ. Sci. Technol. 47,
8248e8255.
Hadi, P., Barford, J., McKay, G., 2013b. Synergistic effect in the
simultaneous removal of binary cobaltenickel heavy metals
from effluents by a novel e-waste-derived material. Chem.
Eng. J. 228, 140e146.
Hadi, P., Barford, J., McKay, G., 2013c. Electronic waste as a new
precursor for adsorbent production. SIJ Trans. Ind. Financ.
Bus. Manag. 1, 128e135.
Hadi, P., Barford, J., McKay, G., 2014a. Selective toxic metal uptake
using an e-waste-based novel sorbenteSingle, binary and
ternary systems. J. Environ. Chem. Eng. 2, 332e339.

53

Hadi, P., Gao, P., Barford, J.P., McKay, G., 2013d. Novel application
of the nonmetallic fraction of the recycled printed circuit
boards as a toxic heavy metal adsorbent. J. Hazard. Mater.
252e253, 166e170.
Hadi, P., Ning, C., Ouyang, W., Lin, C.S.K., Hui, C.-W., McKay, G.,
2014b. Conversion of an aluminosilicate-based waste
material to high-value efficient adsorbent. Chem. Eng. J. 256,
415e420.
Hadi, P., Xu, M., Ning, C., Lin, C., McKay, G., 2015. A critical review
on preparation, characterization and utilization of sludgederived activated carbons for wastewater treatment. Chem.
Eng. J. 260, 895e906.
Hameed, B.H., 2007. Equilibrium and kinetics studies of 2,4,6trichlorophenol adsorption onto activated clay. Colloids
Surfaces A Physicochem. Eng. Asp. 307, 45e52.
Han, D.S., Orillano, M., Khodary, A., Duan, Y., Batchelor, B., AbdelWahab, A., 2014. Reactive iron sulfide (FeS)-supported
ultrafiltration for removal of mercury (Hg(II)) from water.
Water Res. 53, 310e321.
Harada, M., 1982. Minamata Disease in: Adverse Effects of Foods.
Springer, US.
Henneberry, Y.K., Kraus, T.E.C., Fleck, J. a, Krabbenhoft, D.P.,
Bachand, P.M., Horwath, W.R., 2011. Removal of inorganic
mercury and methylmercury from surface waters following
coagulation of dissolved organic matter with metal-based
salts. Sci. Total Environ. 409, 631e637.
Ho, Y.S., Porter, J.F., Mckay, G., 2002. Equilibrium isotherm studies
for the sorption of divalent metal ions onto peat: copper,
nickel and Lead single component systems. Water. Air. Soil
Pollut. 141, 1e33.
Hsi, H.C., Rood, M.J., Rostam-Abadi, M., Chen, S., Chang, R., 2001.
Effects of sulfur impregnation temperature on the properties
and mercury adsorption capacities of activated carbon fibers
(ACFs). Environ. Sci. Technol. 35, 2785e2791.
Huebra, M., Elizalde, M., Almela, a, 2003. Hg(II) extraction by LIX
34. Mercury removal from sludge. Hydrometallurgy 68, 33e42.
Hutchison, A., Atwood, D., Santilliann-Jiminez, Q.E., 2008. The
removal of mercury from water by open chain ligands
containing multiple sulfurs. J. Hazard. Mater. 156, 458e465.
Inbaraj, B.S., Sulochana, N., 2006. Mercury adsorption on a carbon
sorbent derived from fruit shell of Terminalia catappa. J.
Hazard. Mater. 133, 283e290.
Inbaraj, B.S., Wang, J.S., Lu, J.F., Siao, F.Y., Chen, B.H., 2009.
Adsorption of toxic mercury(II) by an extracellular biopolymer
poly(gamma-glutamic acid). Bioresour. Technol. 100, 200e207.
Ismaiel, A.A., Aroua, M.K., Yusoff, R., 2013. Palm shell activated
carbon impregnated with task-specific ionic-liquids as a novel
adsorbent for the removal of mercury from contaminated
water. Chem. Eng. J. 225, 306e314.
Jodeh, S., Basalat, N., Obaid, A.A., Bouknana, D., Hammouti, B.,
Hadda, T.B., Jodeh, W., Warad, I., 2014. Adsorption of some
organic phenolic compounds using activated carbon from
cypress products. J. Chem. Pharm. Res. 6, 713e723.
Kadirvelu, K., Kavipriya, M., Karthika, C., Radhika, M.,
Vennilamani, N., Pattabhi, S., 2003. Utilization of various
agricultural wastes for activated carbon preparation and
application for the removal of dyes and metal ions from
aqueous solutions. Bioresour. Technol. 87, 129e132.
Kadirvelu, K., Kavipriya, M., Karthika, C., Vennilamani, N.,
Pattabhi, S., 2004. Mercury (II) adsorption by activated carbon
made from sago waste. Carbon N. Y. 42, 745e752.
Kara, S., Aydiner, C., Demirbas, E., Kobya, M., Dizge, N., 2007.
Modeling the effects of adsorbent dose and particle size on the
adsorption of reactive textile dyes by fly ash. Desalination 212,
282e293.
Kidd, K., Batchelar, K., 2012. Mercury. In: Wood, C., Farrell, A.P.,
Brauner, C.J. (Eds.), Homeostasis and Toxicology of Nonessential Metals. Elsevier Ltd., US, pp. 237e295.

54

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

Korpiel, J.A., Vidic, R.D., 1997. Effect of sulfur impregnation


method on activated carbon uptake of Gas-Phase mercury.
Environ. Sci. Technol. 31, 2319e2325.
Ku, Y., Wu, M.-H., Shen, Y.-S., 2002. Mercury removal from
aqueous solutions by zinc cementation. Waste Manag. 22,
721e726.
Langmuir, I., 1918. The adsorption of gases on the plane surfaces
of glass, mica and platinum. J. Am. Chem. Soc. 40, 1361e1403.
Li, S.-X., Zheng, F.-Y., Yang, H., Ni, J.-C., 2011. Thorough removal
of inorganic and organic mercury from aqueous solutions by
adsorption on Lemna minor powder. J. Hazard. Mater. 186,
423e429.
Li, Z., Wu, L., Liu, H., Lan, H., Qu, J., 2013. Improvement of aqueous
mercury adsorption on activated coke by thiolfunctionalization. Chem. Eng. J. 228, 925e934.
Liu, M., Hou, L.-A., Xi, B., Zhao, Y., Xia, X., 2013. Synthesis,
characterization, and mercury adsorption properties of hybrid
mesoporous aluminosilicate sieve prepared with fly ash. Appl.
Surf. Sci. 273, 706e716.
Lloyd-Jones, P.J., Rangel-Mendez, J.R., Streat, M., 2004. Mercury
sorption from aqueous solution by chelating ion exchange
resins, activated carbon and a biosorbent. Process Saf.
Environ. Prot. 82, 301e311.
Lo, K.S.L., Yu, Y.H., 1988. The removal of soluble mercury by
cementation processes. In: Proceeding of Second IAWPRC
Asian Conference on Water Pollution Control, pp. 441e447.
Lopes, C.B., Otero, M., Lin, Z., Silva, C.M., Pereira, E., Rocha, J.,
Duarte, A.C., 2010. Effect of pH and temperature on Hg2
water decontamination using ETS-4 titanosilicate. J. Hazard.
Mater. 175, 439e444.
Lopez-Gonzalez, J.D.D., Moreno-Castilla, C., Guerrero-Ruiz, A.,
Rodriguez-Reinoso, F., 1982. Effect of carbon-oxygen and
carbon-sulphur surface complexes on the adsorption of
mercuric chloride in aqueous solutions by activated carbons. J.
Chem. Technol. Biotechnol. 32, 575e579.
 pez-Mun
~ oz, M.J., Aguado, J., Arencibia, a, Pascual, R., 2011.
Lo
Mercury removal from aqueous solutions of HgCl2 by
heterogeneous photocatalysis with TiO2. Appl. Catal. B
Environ. 104, 220e228.
Lu, X., Jiang, J., Sun, K., Wang, J., Zhang, Y., 2014. Influence of the
pore structure and surface chemical properties of activated
carbon on the adsorption of mercury from aqueous solutions.
Mar. Pollut. Bull. 78, 69e76.
 -Agullo
 , J. a, Moore, B.C., Cazorla-Amoro
 s, D., LinaresMacia
Solano, a, 2004. Activation of coal tar pitch carbon fibres:
physical activation vs. chemical activation. Carbon N. Y. 42,
1367e1370.
Marsh, H., Reinoso, F.R., 2006. Activated Carbon. Elsevier Ltd, UK.
Matlock, M.M., Howerton, B.S., Atwood, D. a, 2001. Irreversible
precipitation of mercury and lead. J. Hazard. Mater. 84,
73e82.
Mckay, G., Blair, H.S., Findon, A., 1989. Equilibrium studies for the
sorption of metal-ions onto chitosan. Indian J. Chem. Sect. aInorg. Bio-Inorg. Phys. Theor. Anal. Chem. 28, 356e360.
Mohan, D., Gupta, V.K., Srivastava, S.K., Chander, S., 2001.
Kinetics of mercury adsorption from wastewater using
activated carbon derived from fertilizer waste. Colloids
Surfaces A Physicochem. Eng. Asp. 177, 169e181.
Molina-Sabio, M., Gonzalez, M.T., Rodriguez-Reinoso, F.,
Sepulveda-Escribano, A., 1996. Effect of steam and carbon
dioxide activation in the micropore size distribution of
activated carbon. Carbon N. Y. 34, 505e509.
Mondal, D.K., Nandi, B.K., Purkait, M.K., 2013. Removal of mercury
(II) from aqueous solution using bamboo leaf powder:
equilibrium, thermodynamic and kinetic studies. J. Environ.
Chem. Eng. 1, 891e898.
Montagnaro, F., Santoro, L., 2009. Reuse of coal combustion ashes
as dyes and heavy metal adsorbents: effect of sieving and

demineralization on waste properties and adsorption


capacity. Chem. Eng. J. 150, 174e180.
Nabais, J.V., Carrott, P.J.M., Carrott, M.M.L.R., Belchior, M.,
Boavida, D., Diall, T., Gulyurtlu, I., 2006. Mercury removal from
aqueous solution and flue gas by adsorption on activated
carbon fibres. Appl. Surf. Sci. 252, 6046e6052.
Namasivayam, C., Kadirvelu, K., 1999. Uptake of mercury (II) from
wastewater by activated carbon from an unwanted
agricultural solid by-product: coirpith. Carbon N. Y. 37, 79e84.
Namasivayam, C., Periasamy, K., Division, E.C., Nadu, T., 1993.
Bicarbonate-treated peanut hull carbon for mercury (II)
removal from aqueous solution. Water Res. 27, 1663e1668.
Namasivayam, C., Radhika, R., Suba, S., 2001. Uptake of dyes by a
promising locally available agricultural solid waste: coir pith.
Waste Manag. 21, 381e387.
Nanseu-Njiki, C.P., Tchamango, S.R., Ngom, P.C., Darchen, A.,
Ngameni, E., 2009. Mercury(II) removal from water by
electrocoagulation using aluminium and iron electrodes. J.
Hazard. Mater. 168, 1430e1436.
Nriagu, J.O.J.O., Pacyna, J.M.J.M., 1988. Quantitative assessment of
worldwide contamination of air, water and soils by trace
metals. Nature 333, 134e139.
Oubagaranadin, J.U.K., Sathyamurthy, N., Murthy, Z.V.P., 2007.
Evaluation of Fuller's earth for the adsorption of mercury from
aqueous solutions: a comparative study with activated
carbon. J. Hazard. Mater. 142, 165e174.
Pacyna, E.G., Pacyna, J.M., Sundseth, K., Munthe, J., Kindbom, K.,
Wilson, S., Steenhuisen, F., Maxson, P., 2010. Global emission
of mercury to the atmosphere from anthropogenic sources in
2005 and projections to 2020. Atmos. Environ. 44, 2487e2499.
~ a-Rodrguez, S., Bermudez-Couso, A., No
 voa-Mun
~ oz, J.C.,
Pen

vez, M., Ferna
 ndez-Sanjurjo, M.J., AlvarezArias-Este
~ ez-Delgado, A., 2013. Mercury removal
Rodrguez, E., Nun
using ground and calcined mussel shell. J. Environ. Sci. 25,
2476e2486.
Peniche-Covas, C., Alvarez, L.W., Arguelles-Monal, W., 1992. The
adsorption of mercuric ions by chitosan. J. Appl. Polym. Sci. 46,
1147e1150.
Pillay, K., Cukrowska, E.M., Coville, N.J., 2013. Improved uptake of
mercury by sulphur-containing carbon nanotubes.
Microchem. J. 108, 124e130.
Plazinski, W., Rudzinski, W., Plazinska, A., 2009. Theoretical
models of sorption kinetics including a surface reaction
mechanism: a review. Adv. Colloid Interface Sci. 152, 2e13.
Ramadan, H., Ghanem, A., El-Rassy, H., 2010. Mercury removal
from aqueous solutions using silica, polyacrylamide and
hybrid silicaepolyacrylamide aerogels. Chem. Eng. J. 159,
107e115.
Ranganathan, K., 2003. Adsorption of Hg (II) ions from aqueous
chloride solutions using powdered activated carbons. Carbon
N. Y. 41, 1087e1092.
Rao, M.M., Reddy, D.H.K.K., Venkateswarlu, P., Seshaiah, K., 2009.
Removal of mercury from aqueous solutions using activated
carbon prepared from agricultural by-product/waste. J.
Environ. Manage 90, 634e643.
Redlich, O., Peterson, D.L., 1959. A useful adsorption isotherm. J.
Phys. Chem. 63, 1024e1024.
Roberts, E.J., Rowland, S.P., 1973. Removal of mercury from
aqueous solutions by nitrogen-containing chemically
modified cotton. Environ. Sci. Technol. 7, 552e555.
, D., Fekete, L., Meider, H., 1980. Macrocyclic polythiaethers
Sevdic
as solvent extraction reagentsdIII: extraction and complex
formation of silver(I) and mercury(II) picrates. J. Inorg. Nucl.
Chem. 42, 885e889.
Sips, R., 1948. Combined form of Langmuir and Freundlich
equations. J. Chem. Phys. 16, 490e495.
Somerset, V., Petrik, L., Iwuoha, E., 2008. Alkaline hydrothermal
conversion of fly ash precipitates into zeolites 3: the removal

w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5

of mercury and lead ions from wastewater. J. Environ. Manage


87, 125e131.
Sun, X., Hwang, J.-Y., Xie, S., 2011. Density functional study of
elemental mercury adsorption on surfactants. Fuel 90,
1061e1068.
Sunderlan, E.M., Chmura, G.L., 2000. An inventory of historical
mercury emissions in maritime canada: implications for
present and future contamination. Sci. Total Environ. 256,
39e57.
Takagai, Y., Shibata, A., Kiyokawa, S., Takase, T., 2011. Synthesis
and evaluation of different thio-modified cellulose resins for
the removal of mercury (II) ion from highly acidic aqueous
solutions. J. Colloid Interface Sci. 353, 593e597.
Taurozzi, J.S., Redko, M.Y., Manes, K.M., Jackson, J.E.,
Tarabara, V.V., 2013. Microsized particles of Aza222 polymer
as a regenerable ultrahigh affinity sorbent for the removal of
mercury from aqueous solutions. Sep. Purif. Technol. 116,
415e425.
Toles, C.A., Marshall, W.E., Johns, M.M., 1999. Surface functional
groups on acid-activated nutshell carbons. Carbon N. Y. 37,
1207e1214.
Toles, C.A., Marshall, W.E., Johns, M.M., Wartelle, L.H.,
McAloon, A., 2000. Acid-activated carbons from almond shells:
physical, chemical and adsorptive properties and estimated
cost of production. Bioresour. Technol. 71, 87e92.
Uludag, Y., Ozbelge, H., Yilmaz, L., 1997. Removal of mercury from
aqueous solutions via polymer-enhanced ultrafiltration. J.
Memb. Sci. 129, 93e99.
United Nations Environment Programme, 2010. Options for
Regulating Mercury in Products (Choice Reviews Online).
 zquez, G., Gonza
 lez-Alvarez, J., Freire, S., Lo
 pez-Lorenzo, M.,
Va
Antorrena, G., 2002. Removal of cadmium and mercury ions
from aqueous solution by sorption on treated Pinus pinaster
bark: kinetics and isotherms. Bioresour. Technol. 82, 247e251.
Vitolo, S., Pini, R., 1999. Deposition of Sulfur From H 2 S on Porous
Adsorbents And e V Ect on Their Mercury Adsorption
Capacity, vol. 28, pp. 341e354.
Wahi, R., Ngaini, Z., Jok, V.U., 2009. Removal of Mercury, Lead and
Copper from Aqueous Solution by Activated Carbon of Palm
Oil Empty Fruit Bunch, vol. 5, pp. 84e91.

55

Wajima, T., Murakami, K., Kato, T., Sugawara, K., 2009. Heavy
metal removal from aqueous solution using carbonaceous K2
S-impregnated adsorbent. J. Environ. Sci. 21, 1730e1734.
Wajima, T., Sugawara, K., 2011. Adsorption behaviors of mercury
from aqueous solution using sulfur-impregnated adsorbent
developed from coal. Fuel Process. Technol. 92, 1322e1327.
Wang, J., Deng, B., Wang, X., Zheng, J., 2009. Adsorption of
aqueous Hg(II) by sulfur-impregnated activated carbon.
Environ. Eng. Sci. 26, 1693e1699.
Wang, X.J., Xu, X.M., Liang, X., Wang, Y., Liu, M., Wang, X.,
Xia, S.Q., Zhao, J.F., Yin, D.Q., Zhang, Y.L., 2011. Adsorption of
copper(II) onto sewage sludge-derived materials via
microwave irradiation. J. Hazard. Mater. 192, 1226e1233.
Weber, T.W., Chakravorti, R.K., 1974. Pore and solid diffusion
models for fixed-bed adsorbers. AICHE J. 20, 228e238.
Xianglan, Z., Shengfu, D., Qiong, L., Yan, Z., Lei, C., 2011. Surface
functional groups and redox property of modified activated
carbons. Min. Sci. Technol. 21, 181e184.
Xu, M., Hadi, P., Chen, G., McKay, G., 2014. Removal of cadmium
ions from wastewater using innovative electronic wastederived material. J. Hazard. Mater. 273, 118e123.
Yardim, M.F., Budinova, T., Ekinci, E., Petrov, N., Razvigorova, M.,
Minkova, V., 2003. Removal of mercury (II) from aqueous
solution by activated carbon obtained from furfural.
Chemosphere 52, 835e841.
Zabihi, M., Ahmadpour, a, Asl, a H., 2009. Removal of mercury
from water by carbonaceous sorbents derived from walnut
shell. J. Hazard. Mater. 167, 230e236.
Zabihi, M., Haghighi Asl, A., Ahmadpour, A., 2010. Studies on
adsorption of mercury from aqueous solution on activated
carbons prepared from walnut shell. J. Hazard. Mater. 174,
251e256.
Zhang, F.-S., Nriagu, J.O., Itoh, H., 2005. Mercury removal from
water using activated carbons derived from organic sewage
sludge. Water Res. 39, 389e395.
~ Solidification
Zhang, J., Bishop, P.L., Asce, F., 2003. Stabilization O
of High Mercury Wastes with Reactivated Carbon, pp. 31e36.
Zhu, J., Deng, B., Yang, J., Gang, D., 2009. Modifying activated
carbon with hybrid ligands for enhancing aqueous mercury
removal. Carbon N. Y. 47, 2014e2025.

Das könnte Ihnen auch gefallen