aqueous mercury adsorption by activated carbon

Upload: grace-pooley

Post on 22-Feb-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    1/19

    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 Kongb School of Energy and Environment, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kongc

    Division of Sustainable Development, College of Science, Engineering and Technology, Hamad Bin KhalifaUniversity, Qatar Foundation, Doha, Qatar

    a r t i c l e i n f o

    Article history:

    Received 14 October 2014

    Received in revised form

    19 December 2014

    Accepted 9 January 2015

    Available online 21 January 2015

    Keywords:Activated carbon

    Adsorption

    Mercury

    Porous structure

    Sulfur functional groups

    a b s t r a c t

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

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

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

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

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

    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

    directed to the effect of sulfur-containing functional moieties on enhancing the mercuryadsorption. It has been demonstrated that surface area, pore size, pore size distribution

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

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

    been addressed using equilibrium adsorption isotherm, thermodynamic and kinetic

    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. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

    2. Preparation of activated carbon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

    3. 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, ClearWater Bay Road, Hong Kong. Tel.:852 23588412; fax: 852 23580054.

    E-mail address:[email protected](G. McKay).

    Available online atwww.sciencedirect.com

    ScienceDirect

    j o u r n a l h o m e p a g e : w w w . e l s e v i er . c o m / l o c a t e / wa t r e s

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

    http://dx.doi.org/10.1016/j.watres.2015.01.018

    0043-1354/2015 Elsevier Ltd. All rights reserved.

    mailto:[email protected]://www.sciencedirect.com/science/journal/00431354http://www.elsevier.com/locate/watreshttp://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://www.elsevier.com/locate/watreshttp://www.sciencedirect.com/science/journal/00431354http://crossmark.crossref.org/dialog/?doi=10.1016/j.watres.2015.01.018&domain=pdfmailto:[email protected]
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    2/19

    4. 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

    5. Mercury affinity to various functional groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

    6. Equilibrium adsorption isotherms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 507. Conclusion and future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

    1. Introduction

    Mercury is categorized as an extremely toxic substance whose

    health hazards have primarily been associated with inhala-

    tion of mercury vapor or ingestion of organic mercury viaaquatic 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 themost dominant role. An overarching estimate of anthropo-

    genic input of 16 elements into the environment by Nriagu

    and Pacyna has revealed that mercury releases into the at-

    mosphere and aquatic bodies are in the same range

    (Ebinghaus et al., 1999; Nriagu and Pacyna, 1988). The

    anthropogenic sources of mercurycan be divided intoprimary

    and secondary sources. The former involves the mobilization

    and release of mercury of geological origin to the environ-

    ment, 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 pro-

    cesses, including vinyl chloride monomer production as

    catalyst, batteries as cathodes and chlor-alkali production aswell (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 elimina-

    tion of mercury from the ecosystem is unrealistic due to the

    uncontrollable discharge of mercury by primary anthropo-

    genic 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 di-

    rectives have been enacted regarding the control of the mer-cury 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.,

    1997), solvent extraction (Huebra et al., 2003; Sevdic et al.,

    1980), photocatalysis (Botta et al., 2002; de la Fourniere et al.,

    2007; Lopez-Mu~noz et al., 2011), adsorption (Aguado et al.,

    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 con-

    cerning 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 theavailability 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 bio-

    polymers (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

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 538

    http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    3/19

    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 mercuryremoval 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 treat-

    ment 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 heattreatment of a carbonaceous precursor at moderate temper-

    atures, mainly below 700 C, in an inert atmosphere to pyro-

    lize 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 pro-

    duced 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 in-

    creases, while furtherincrease in the temperature reduces the

    pore volume of the material. The produced char is subse-

    quently 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). Theaim of the activation stage is to produce a highly porous

    structure from the weakly-developed porous char. The acti-

    vation 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 stoi-

    chiometric equations:

    C CO2/2CO DH 159 KJ mol1 (R1)

    C H2O/CO H2 DH 117 KJ mol1 (R2)

    The chemical activation technique is a one-step process in

    which the carbonization and activation processes occur

    simultaneously. During chemical activation, the source ma-

    terial is submerged in a dehydrating compound (such asphosphoric 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 struc-

    ture. The resultant slurry is then heatedup to temperatures in

    the rangeof 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 heattreatment.

    Energy required for chemical activation process is consid-

    erably lower than that of physical activation method. Signifi-

    cantly 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 con-

    sumption. 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

    method compared with the physical one (Macia-Agulloet al.,

    2004).Although, commercial activated carbons have been widely

    used in various industries to remove mercury from waste-

    water 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 waste-

    based carbon sources as precursor for activated carbon pro-

    duction. Toles et al. found that it is highly profitable to pro-

    duce 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 cost-effective activated carbons with high pollutant removal effi-

    ciencies. Among certain precursor materials to be used in

    activated carbon production, high mercury adsorption ca-

    pacities 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 car-

    bons derived from pazzolana and yellow tuff (Di Natale et al.,

    2006), Fullers earth (Oubagaranadin et al., 2007), Ceiba-

    pentandra hulls, Phaseolus aureus hulls and Cicerarietinum

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5 39

    http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    4/19

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

    precursor is an influential factor in determining the adsorp-

    tion efficiency of the activated carbons.

    3. Effect of treatment techniques on mercuryremoval

    3.1. 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 2show the effect of the pre-

    cursor 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 oxidationon 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 at-

    mospheres 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 anincrease 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 areaand functional groups are predominantly affected by the na-

    ture of the precursor material used. As shown in Table 1,

    Ekinci et al. have found that, under similar activation condi-

    tions, the surface areas range from 460 to 720 m2:g1 for coal-

    based 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).Table1e

    Effectoftheprecursortype

    andphysicalactivationconditionson

    themercuryremovalefficienciesof

    theactivatedcarbons.

    Carbonizationconditions

    Activationconditions

    SBET

    qm

    Vt

    Vmicro

    Vmeso

    Dp

    Ref.

    T(C)

    T(h)

    HRa

    (Cmin

    1)

    Atm.

    T(C)

    T(h)

    HR(Cmin

    1)

    Atm

    .

    (m2

    g

    1)

    (mgg

    1)

    (cm

    3

    g

    1)

    (cm

    3

    g

    1)

    (cm

    3

    g

    1)

    (nm)

    200

    2

    750

    2

    20

    H2O

    521

    25.8

    8

    (Raoeta

    l.,

    2009)

    325

    23.6

    6

    280

    22.8

    8

    600

    1

    3

    N2

    761

    0.6

    81

    0.2

    97

    0.3

    21

    2.3

    2

    (Budinovaetal.,

    2006)

    750

    2

    10

    H2O

    1100

    174

    0.8

    9

    0.4

    25

    0.1

    1

    (Budinovaetal.,

    2003)

    400

    H2O

    480

    134

    0.6

    4

    0.2

    1

    0.0

    85

    750

    2

    10

    H2O

    600

    92

    (Ekincie

    tal.,

    2002)

    460

    56

    720

    105

    520

    37

    1000

    153

    1100

    174

    500e950

    10

    N2

    (Inbaraj

    andSulochana,

    2006)

    890

    9

    5

    N2

    &

    CO2

    1528

    0.6

    4

    (Macia-A

    gulloetal.,

    2004)

    890

    22.5

    5

    N2

    &

    CO2

    2487

    0.8

    6

    800

    1

    5

    N2

    900

    5

    15

    N2

    &

    CO2

    997

    0.3

    7

    (Nabaisetal.,

    2006)

    900

    5

    N2

    702

    140

    2.7

    22.5

    (Choian

    dJang,

    2008)

    850

    1

    N2

    &

    H2O

    870

    4.9

    3

    0.4

    28

    0.3

    28

    0.0

    04

    1.9

    67

    (Luetal.,

    2014)

    a

    HRdenotesheatingrate.

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 540

    http://-/?-http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://-/?-
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    5/19

    Table 2 e Effect of the precursor type and chemical activation conditions on the mercury removal efficiencies of the activated carbo

    Carbonization/Activationconditions

    Chemical treatment SBET qm Vt Vmicro

    T (C) T (h) HRa

    (C min1)Atm. Chem. Conc. Imp. Rb Imp.tc (h) (m2 g1) (mg g1) (cm3 g1) (cm3 g1)

    600/600 1/1 3 H2O/N2 1290 0.939 0.471

    700 2 H2O 1360 160 1.026 0.485

    600 10 N2 H2SO4 Conc. 1:20 1050 154 0.827 0.38

    H2SO4 Conc. 10

    10

    10

    10

    10

    200/450 1 Air H2O2 629 18.1

    H2SO4 & (NH4)2S2O8 50% 1:200 0.5 625 55.6

    600/800 1 N2/H2O H2SO4 Conc. 1100 174 0.425

    H2SO4 & (NH4)2S2O8 Conc. 3: 5 & 1:33 12 592 154

    400/700 0.5/1 NaOH 20% (W/V) 2 379.4 52.7

    N2 ZnCl2 98% 1:0.5 780 151.5

    1:1 803 100.9

    H2SO4 Conc. 1:1.8 24 208.1 109.9

    K2CO3 33-75% 1:1 2 1260 129 0.492

    KOH 1:2 1 1090 0.49

    1:4 1 1635 0.78

    1:6 1 2225 0.9

    1:8 1 2420 0.94

    NaOH 1:2 1 1130 0.51

    1:4 1 2000 0.81

    1:6 1 2541 0.96

    1:8 1 3033 1.02

    800/900 1/5 5/15 N2 & CO2 S 1:3 2 848 410 0.33

    H2S 1 905 450 0.33

    900 5 N2 Pyrrole 1:3 413 541 1.52

    43:100 346 682 1.04

    11:20 158 441 0.54

    900 N2 922 301 0.87 0.37 900 H2S 785 351 0.78 0.31

    200 SO2, H2S & N2 764 351 0.72 0.3

    200/400 2/1 10 H2O, SO2 & H2S 500.5 227.3 0.43

    H2O & SO2 506.5 222.2 0.48 0.04

    H2O,&H2S 530.2 217.4 0.48 0.06

    H2O 536.5 208.3 0.52 0.13

    800 0.5 N2 K2S 1:3.6 30.0 235.7

    850 0.5 30.0 243.9

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    6/19

    Table 2 e (continued)

    Carbonization/Activationconditions

    Chemical treatment SBET qm Vt Vmicro

    T (C) T (h) HRa

    (C min1)Atm. Chem. Conc. Imp. Rb Imp.tc (h) (m2 g1) (mg g1) (cm3 g1) (cm3 g1)

    900 0.5 28.0 254.4

    1000 0.5 30.0 224.1

    H2O2 3:10 168 836.0 100 0.23

    H2 & CS2 778.0 170 0.18H2O2 & He 950.0 110

    H2O2, He & CS2 776.0 175 0.22

    He 880.0 110

    2 50 N2 S 2:1 645.0 467

    2:1 594.0 507

    2:1

    2:1 1070.0 827

    2:1 1045.0 427

    700 4 HCl 1 N 4 311.0 9.32

    H2SO4 13 M 1:5 0.33 66.0 303.03

    13 M 1:5 0.33 384

    H2SO4 12 M 1:9 0.5 19.0 385

    12 M 1:9 0.5 526

    HNO3, SOCl2 &

    C2H4(NH2)2

    5 M, 5%

    &0.05 M

    3: 20 & 3:8 7 120

    H2O2 30% 1:2 829.0 5 0.413 0.318

    2:5 825.0 4.2 0.408 0.315

    600 10 N2 SO2 1:25 0.5 720.9 122.8

    600 1:25 1 772.5 129.8

    600 1:25 2 776.4 130.5

    600 1:25 3 790.7 131.6

    600 1:50 1 773.7 128.2

    600 2:25 1 518.5 114.8

    600 1:25 1 757.2 125.7

    500 1:25 1 764.1 135.9

    700 1:25 1 751.3 184.2

    800 1:25 1 1087.0 196.8

    900 1:25 1 1057.0 207.8

    TOMATS 1:0.3 48 107.7 83.3

    HNO3 0.01 M 72 136.7 315.8 HSCH2COOH,

    (CH3CO)2O & H2SO4

    13 193.7 694.9

    a HR denotes heating rate.b Imp. R denotes impregnation ratio.c Imp. t denotes impregnation time.

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    7/19

    Nonetheless, although specific surface area is one of the fac-

    tors 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 capacitiesunder similar adsorption conditions did not exhibit a signifi-

    cant 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 car-

    bons 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 simultaneouslyinvestigated the effects of pore structure and functional

    groups on mercury removal.

    Considering the surface areas listed inTables 1 and 2, it is

    noticeable that the chemical activation technique has greater

    performance in pore formation at appropriate chemical re-

    agent to adsorbent ratio. Comparing the surface areas ob-

    tained 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 burn-

    off (94%). On the contrary, chemical activation of this materialusing 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 ad-

    vantages of the chemically activated carbon are higher prod-

    uct yield (60% and 27%, respectively), lower activation

    temperature (750 C) and shorter activation time (1 h). The

    highestsurface area (3033 m2:g1) can be obtainedusing NaOH

    as activating agent at an impregnation ratio of 8:1 (w/w)

    (Macia-Agulloet al., 2004).

    In order to enhance mercury adsorption, several authors

    have studied the combination of chemical and physical acti-

    vation techniques. Budinova et al. confirmed that when the

    H3PO4-impregnated carbonaceous sample was treated understeam atmosphere, both the surface area and iodine number

    were considerably higher than the samples pyrolyzed under

    nitrogen atmosphere. They also revealed that the concentra-

    tion of the chemical reagent used for impregnation has a

    significant effect on pore development. When the concentra-

    tion 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 concentra-

    tion 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

    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 mer-

    cury. Therefore, the application of sulfur-functionalized acti-

    vated 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 sul-

    fide (Na2S), hydrogen sulfide (H2S), sulfur dioxide (SO2) or

    sulfur powder, with the aim of increasing their mercury up-

    takes (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 tech-niques, namely impregnation with elemental sulfur and using

    hydrogen sulfide gas as modifying agent. Both of these ad-

    sorbents exhibited higher adsorption capacities than the un-

    sulfurized 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 ef-

    fect 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 capacityof the unmodified and modified activated carbons were

    190 mg:g1 and 820 mg:g1, respectively. The chemical reac-

    tion between elemental sulfur and the surface of the adsor-

    bent 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 processof sulfurbinding onto the activated carbon surface

    and consequently, the mechanism of mercury adsorption

    onto sulfur-containing moieties have not been satisfactorilyexploited 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 penta-

    sulfide. Subsequent mercury adsorption revealed significantly

    diminished intensities of the bands corresponding to the thiol

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5 43

    http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    8/19

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

    Hg2(SH)2bonds 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 mer-

    cury, higher amounts of sulfur moieties will theoretically be

    advantageous in mercury removal. Several researchers have

    reported the direct linear relationship between mercury up-

    take 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 sul-

    fur to ACF via gas stream H2S. This may be reasonable due to

    the melting, recrystallization and deposition of the solid sul-

    furon the activated carbon surface at such high temperatures.

    However, subsequent mercury adsorption tests showed thathigher mercury uptake was obtained by the latter method

    which elucidated the importance of the type of sulfur func-

    tional groups on the carbon surface besides its quantity

    (Nabais et al., 2006). This may also occur due to the aggrega-

    tion 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-impreg-

    nated coal samples were prepared at three distinct tempera-

    tures (800e1000 C), whereas the activated carbon sample

    prepared at 900 C exhibited the highest and fastest adsorp-

    tion (Wajima and Sugawara, 2011). This indicates that the

    sulfur content on the adsorbent surface, type of sulfur func-tional 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 multi-

    component 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 phe-

    nomenon has also been confirmed by comparing the inter-

    action of HgX2(where X is a halide) with C4H8O and C4H8S. It

    has been reported that HgX2interacts weakly with C4H8O, but

    much stronger with C4H8S (Farhangi and Graddon, 1973;

    Fisher and Drago, 1975). Vazquezet al. have related the

    higher affinity towards mercury in a multi-component system

    of cadmium, zinc and mercury to the higher electronegativityof mercury (Vazquez et al., 2002).

    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 adsorptionprocess 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 stepand 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 func-

    tional group sites on the surface of activated carbon available

    for the mercury ions.As the sitesare occupied in the course of

    time, the uptake rate is gradually slowed down until a plateauis 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 Periasamyhave 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 mercuryremoval percentage and thus more cost effectiveness

    (Namasivayam and Periasamy, 1993). Also, rapid mercury

    adsorption of less than 20 min to reach equilibrium was re-

    ported 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 in-

    creases 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 mercuryadsorption. More

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 544

    http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    9/19

    studies are necessary to be conducted to prove these hy-

    potheses regarding the relationship between mercury

    adsorption rate and textural and surface properties of the

    adsorbents.

    It has been further demonstrated that as the initial con-

    centration of mercury increases, the Lagergren rate constant

    decreases and thus, longer time is required to achieve equi-

    librium (Namasivayam and Kadirvelu, 1999; Namasivayamand Periasamy, 1993). This could be due to the saturation of

    sites presenton 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. How-

    ever, 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 con-

    centration 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 concen-

    tration. 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 dos-

    ages, whereas the latter takes into account the adsorbent

    dosage and reveals the genuine mercury adsorption efficiency

    of the material at different initial concentrations. Several au-thors 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 consid-

    eration, 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 tothe 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 concentra-

    tion 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

    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 concen-

    trations (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 con-

    tact between the adsorbate molecules and the adsorbent

    materials thus enhancing the possibility of the mercury up-

    take 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 ioni-

    zation of the functional groups on the adsorbent surface. In

    acidic environment, high concentration of hydronium ion

    (H3O) in the solution drives the equilibrium ionization reac-

    tion (reaction R4) to the left and prevents the formation of

    ionized functional groups, thereby hampering the ion ex-

    change reaction between metal ions and adsorbent surface

    functional groups (reaction R5). When the pH level of the so-

    lution 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 Haq (R4)

    Adsorbent COO Mnaq4Adsorbent COOM (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 posi-tively charged leading to electrostatic repulsion of the mer-

    cury 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 tempera-

    ture 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 acti-

    vatedcarbon and found a decrease in Gibbs free energy,DG, as

    well as a positive enthalpy value, DH, by raising the

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5 45

    http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    10/19

    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 anincrease in compressibility of the particles with increasing

    temperature. Since porous structure of carbonaceous mate-

    rials 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 mer-

    cury by the adsorbent material. Physical adsorption caused by

    the van der Waals forces between the adsorbent surface andthe adsorbate molecules is typically favored at low tempera-

    tures. 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 con-

    centration 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 isincreased, as discussed in preceding sections, removal per-

    centage is an entirely relative term changing by initial con-

    centration 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. pentandrahull 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 theadsorbent 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 phys-

    isorption as well as higher rate of adsorbate removal (Wahi

    et al., 2009). It is suggested that more tangible adsorption ca-

    pacities should be reported in this context instead of simply

    quoting the percent removal.

    On the other hand, although the removal percentage in-

    creases 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 avail-

    ability 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 par-

    ticles decreased from 1.25e2.5mmto0.21e1 mm,the mercuryadsorption capacity showed a two-fold increase (430 mg:g1

    versus 815 mg:g1, respectively) (Mckay et al., 1989; Peniche-

    Covas 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 moleculescan more easily penetrate into the internal pores of the

    adsorbent (Gupta et al., 2011).

    5. Mercury affinity to various functionalgroups

    The adsorption of adsorbate by activated carbon can be cate-

    gorized into chemical and physical adsorption. Briefly, phys-

    ical 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 beenproven inefficient in

    promoting mercury adsorption, however surface functional

    groups, specially the oxygen containing groups of the adsor-

    bent, 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 un-

    balanced charge on the adsorbent surface where ion exchange

    with the mercuric ion can occur (Sun et al., 2011). This ac-counts 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 nitrogen-

    containing 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 consideredto 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

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 546

    http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    11/19

    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 effi-

    ciency of adsorption.

    As functional groups largely determine the surface prop-

    erties and thus the intensity of ion exchange, manipulation ofthe surface functional group have been of great interest. Sul-

    fur group has been widely reported to promote mercury

    adsorption. Rao et al. have observed an increase in the mer-

    cury adsorption capacity of activated carbon with the intro-

    duction of sulfur groups on the activated carbon surface and

    ascribed the higher removal efficiency of the sulfur-

    containing 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).

    2Hg2

    SO23 2OH

    4Hg

    22 SO

    24 H2O (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 theright 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 canbe 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 theactivated 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 acidsites

    are destroyed at 700 C(Budinova et al., 2008). Although such

    manipulation can be carried out on the surface functional

    groups of activatedcarbons, it can be criticized that there is no

    comprehensive study to compare the effects of various func-

    tional 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 Name Base uptake Acid uptake Reference

    NaHCO3 Na2CO3 NaOH EtONa HCL

    Mengen 0.092 0.120 0.184 1.900 1.120 (Ekinci et al., 2002)

    Seyitomer e 0.120 0.250 2.040 2.670

    Some 0.100 0.110 0.183 1.570 1.120

    Bulluca e 0.110 0.320 1.900 2.010

    Apricot Stones 0.130 0.210 0.360 1.350 0.842

    Furfural 0.120 0.160 0.230 1.500 0.600

    Furfural Carbon A 0.120 0.160 0.230 1.500 0.600 (Budinova et al., 2003)

    Mixture of steam pyrolysis

    tar and furfural (30:70)

    Carbon B 0.030 0.080 0.250 1.300 0.560

    Air oxidized Furfural Carbon C 1.900 1.930 3.660 6.340 e

    Woody biomass birch N600-1 0.744 0.126 0.480 2.234 0.083 (Budinova et al., 2006)

    NS600-1 0.124 0.034 0.572 2.530 1.100

    S700-2 e 0.123 0.422 2.355 0.902

    Walnut shell Walnut shell 0.450 0.490 0.390 0.520 0.520 (Zabihi et al., 2009)

    Carbon A 0.540 0.480 0.350 0.420 0.420

    Carbon B 0.720 0.420 0.300 0.290 0.290

    Coconut activated carbon AC 1.097 0.627 0.561 0.495 e (Lu et al., 2014)

    AC1 1.174 0.594 0.693 0.341

    AC1-1 1.295 0.341 0.726 0.572

    AC1-2 1.328 0.099 0.869 0.594

    Sago 1.200 1.800 0.900 1.600 1.100 (Kadirvelu et al., 2004)

    Furfural 0.120 0.160 0.230 1.500 0.600 (Yardim et al., 2003)

    Antibiotic waste BDLa BDL 0.230 2.300 1.300 (Budinova et al., 2008)

    a

    Below detection limit.

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5 47

    http://-/?-http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://-/?-
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    12/19

  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    13/19

    45 472 0.0090 0.9800 15.9 0.560 0.940

    60 578 0.0080 0.8900 15 0.600 0.890

    Sulfur impregnation of (AC) with

    4%(v/v) SO2at 700 C for 60 min

    30 523 0.0220 0.9600 71.2 0.330 0.990

    45 496 0.0740 0.9200 129.2 0.250 0.980

    60 510 0.0860 0.9500 148.2 0.230 0.960

    Steam activation of bagasse

    pith (SA-C)

    Langmuir 30 172.4 0.0072 e e

    40 181.8 0.008

    50 200 0.0086

    60 208.3 0.0106

    Steam activation of bagassepith in presence of SO2

    (SAeSO2eC)

    30 185.2 0.019540 204.1 0.0202

    50 208.3 0.0229

    60 222.2 0.0262

    Steam activation of bagasse

    pith in presence of H2S

    (SAeH2SeC)

    30 181.8 0.0113

    40 200 0.0123

    50 204.1 0.0164

    60 217.4 0.0229

    Steam activation of bagasse

    pith in presence of SO2 & H2S

    (SAeSO2eH2SeC)

    30 188.7 0.0281

    40 208.3 0.0273

    50 212.8 0.0367

    60 227.3 0.0480

    Rice husk ash Langmuir and Freundlich 15 9.3 0.0115 0.9868 0.42 0.493 0.973

    30 6.7 0.0158 0.9900 0.54 0.469 0.965

    Sulfuric acid treated with rice

    husk (dry sorbent)

    Langmuir and Freundlich 25 303 0.0052 0.9990 7.1 0.5579 0.981

    35 336.7 0.0107 0.9992 28.7 0.377 0.98745 384.6 0.0219 0.9988 48.9 0.3407 0.986

    Sulfuric acid treated with rice

    husk (wet sorbent)

    25 227.3 0.0052 0.9991 8.7 0.4607 0.952

    35 270.3 0.0088 0.9998 15.0 0.428 0.928

    45 303 0.0129 0.9993 18.1 0.448 0.959

    Sulfuric acid treated with

    flax shave (dry sorbent)

    Langmuir 45 416 0.0805 0.9980

    Sulfuric acid treated with

    flax shave (wet sorbent)

    344 0.0468 0.9990

    Commercial activated carbon (AC) Langmuir and Freundlich 10 4.1 0.2083 0.9522 0.8 0.974 0.956

    25 3.5 0.1053 0.9827 0.3 1.262 0.971

    50 3 0.1330 0.9828 0.4 1.150 0.970

    Steam activated AC (AC-1) 10 4.9 0.6870 0.9889 2.8 0.790 0.975

    25 4.5 0.1540 0.9914 0.6 1.248 0.982

    50 4.1 0.3862 0.9866 1.4 0.750 0.952

    AC1 was oxidized with H2O2at 1:2 (m:v)

    (AC1-1)

    10 5.0 0.2329 0.9828 1.1 0.914 0.986

    25 5 0.2000 0.9914 1 0.858 0.984

    50 4.6 0.2618 0.9949 1.6 0.857 0.985

    AC1 was oxidized with H2O2at 2:5 (m:v)

    (AC1-2)

    10 5.2 0.7795 0.9835 3.4 0.823 0.976

    25 5.1 0.4491 0.9841 2.3 1.031 0.979

    50 4.6 0.9287 0.9521 2.7 0.720 0.980

    PSAC grafted with TOMATS Langmuir and Freundlich 20 76.9 0.0588 0.9920 5.5 0.630 0.916

    25 76.9 0.0778 0.9940 6.9 0.590 0.906

    30 83.3 0.1100 0.9930 9.4 0.560 0.892

    35 83.3 0.1250 0.9910 10.2 0.570 0.898

    Activated coke (AC) Langmuir and Freundlich 25 315.8 0.0500 0.9770 12.2 0.714 0.980

    Thiol-functionalized activated coke (SH-AC) 694.9 0.0600 0.9840 71.1 0.526 0.954

  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    14/19

    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 thesulfurto 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 sig-

    nificant 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 acti-

    vation 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 treat-

    ment. 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 sur-

    face 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 actas 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 hy-

    droxyl, carboxylic and carboxylic anhydride group by nitric

    acid treatment (Zhu et al., 2009). When the contact time be-

    tween activated carbon and the nitric acid increases, signif-

    icant amount of phenolic group forms while the content of

    lactone group is reduced. Increase in the concentration ofnitric 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 acid-

    treated 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 Mhydrochloric 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 hy-

    pothesized 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 hin-

    drance between the adsorbed molecules on the adjacent sites

    (Langmuir, 1918). The mathematical expression for this model

    can be represented as:

    qe qmaLCe1 aLCe

    (1)

    whereqeis 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. Amathematical expression developed by Webber and Chakk-

    ravorti has been used to test the favourability of the adsorp-

    tion process (Weber and Chakravorti, 1974):

    RL 1

    1 aLC0(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

  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    15/19

    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 andFreundlich adsorption models are also established. Redlich-

    Peterson 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 ho-

    mogeneous and heterogeneous adsorption (Redlich and

    Peterson, 1959):

    qe qmaRCe

    1 aRCbe

    (4)

    whereaR 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 thisequation 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 het-

    erogeneous adsorption system and assumes the occurrence of

    dissociative adsorption (Sips, 1948). At low adsorbate con-

    centration, the isotherm can be reduced to Freundlich

    isotherm, while at high concentration, it complies with ho-

    mogenous 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. TheSips isotherm model is expressed through the following

    equation:

    qe qmaSCbse1 aSC

    bse

    (5)

    where as is the Sips isotherm model constant andbsisthe Sips

    isotherm model exponent.

    Table 4 summarizes the Langmuir and Freundlich con-

    stants 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 musselshell 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

    surface area lower than that of finely ground mussel ( Pe~na-

    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 sur-

    face area (Cai and Jia, 2010). This phenomenon suggested

    that the surface area alone is not sufficient to reflect the

    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 sig-nificant reduction in SBET of the activated carbons, a stag-

    gering 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 classifi-

    cation (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 ef-

    fects 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 demon-

    strated that a combination of medium-to-high surface area

    with well-functionalized surface properties are collectivelycrucial in enhancing the mercury removal. Despite these

    findings, very limited research has been carried out on

    simultaneous optimization of surface and textural charac-

    teristics of activated carbons. Hence, further study is neces-

    sary for such optimization to save the high amount of energy

    required to obtain unnecessary very high surface area acti-

    vated 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 function-

    alization process, and consequently the mercury adsorptionmechanism 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 strat-

    egy has been disregarded.

    Further recommendations can be presented as follows:

    w a t e r r e s e a r c h 7 3 ( 2 0 1 5 ) 3 7 e5 5 51

    http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018http://dx.doi.org/10.1016/j.watres.2015.01.018
  • 7/24/2019 Aqueous mercury adsorption by activated carbon

    16/19

    Mercury forms complexes in real wastewater systems and

    is rarely found in ionic form. Therefore, study of the effi-

    ciency 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 tobe carried out.

    The regeneration of the activated carbon samples has to be

    conducted for economic feasibility enhancement. Unfor-

    tunately, in adsorption processes, the regeneration is

    overlooked.

    If not regenerated, due to its hazardous nature, mercury-

    loaded activated carbon needs to be stabilized or vitrified

    and then disposed of in hazardous landfill. Nonetheless,

    this landfilled carbon is not reusable, therefore empha-

    sizing the importance of the production cost of the acti-

    vated 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.

    r e f e r e n c e s

    Adams, M.D., 1991. The Mechanisms of Adsorption of Hg ( CN ) 2and HgC12on 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-functionalizedmesoporous silica obtained by cocondensation. Ind. Eng.Chem. Res. 44, 3665e3671.

    Ahmad, F., Daud, W.M.A.W., Ahmad, M.A., Radzi, R., 2013. Theeffects of acid leaching on porosity and surface functionalgroups of cocoa (Theobroma cacao)-shell based activatedcarbon. Chem. Eng. Res. Des. 91, 1028e1038.

    Anirudhan, T.S., Divya, L., Ramachandran, M., 2008. Mercury(II)removal from aqueous solutions and wastewaters using anovel cation exchanger derived from coconut coir pith and itsrecovery. J. Hazard. Mater. 157, 620e627.

    Anoop Krishnan, K., Anirudhan, T.S., 2002. Removal ofmercury(II) from aqueous solutions and chlor-alkali industryeffluent by steam activated and sulphurised activated carbonsprepared 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 ofactivated carbon with SO2in 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 onthiol derivatized single wall carbon nanotubes. J. Hazard.Mater. 261, 534e541.

    Bandosz, T.J., Jagieuo, J., Schwarz, J.A., 1993. Effect of SurfaceChemical 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 bycomplexationdultrafiltration. Desalination 144, 201e206.

    Barron-Zambrano, J., Laborie, S., Viers, P., Rakib, M., Durand, G.,2004. Mercury removal and recovery from aqueous solutionsby 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. Featuresof the transformation of Hg II by heterogeneousphotocatalysis over TiO2. Catal. Today 76, 247e258.

    Budinova, T., Ekinci, E., Yardim, F., Grimm, a, Bjornbom, E.,Minkova, V., Goranova, M., 2006. Characterization andapplication of activated carbon produced by H3PO4and watervapor activation. Fuel Process. Technol. 87, 899e905.

    Budinova, T., Petrov, N., Parra, J., Baloutzov, V., 2008. Use of anactivated carbon from antibiotic waste for the removal ofHg(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 bydifferent modifications of furfural adsorbent. Ind. Eng. Chem.Res. 42, 2223e2229.

    Byrne, H.E., Mazyck, D.W., 2009. Removal of trace level aqueousmercury by adsorption and photocatalysis on silica-titaniacomposites. J. Hazard. Mater. 170, 915e919.

    Cai, J.H., Jia, C.Q., 2010. Mercury removal from aqueous solutionusing coke-derived sulfur-impregnated activated carbons. Ind.Eng. Chem. Res. 49, 2716e2721.

    Chiarle, S., Ratto, M., Rovatti, M., 2000. Mercury removal fromwater by ion exchange resins adsorption. Water Res. 34,2971e2978.

    Choi, M., Jang, J., 2008. Heavy metal ion adsorption ontopolypyrrole-impregnated porous carbon. J. Colloid InterfaceSci. 325, 287e289.

    Cox, M., El-shafey, E.I., Pichugin, A.A., Appleton, Q., 2000. Removalof Mercury ( II ) from Aqueous Solution on a CarbonaceousSorbent 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 surfacefunctional groups of activated carbons produced from Acaciamangium wood. J. Anal. Appl. Pyrolysis 104, 418e425.

    De la Fourniere, E.M., Leyva, A.G., Gautier, E.A., Litter, M.I., 2007.Treatment of phenylmercury salts by heterogeneousphotocatalysis over TiO(2). Chemosphere 69, 682e688.

    Di Natale, F., Erto, A., Lancia, A., Musmarra, D., 2011. Mercuryadsorption 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 andindustrial materials. J. Hazard. Mater. 132, 220e225.

    Diaz, E., Ordonez, S., Vega, a, 2007. Adsorption of volatile organiccompounds onto carbon nanotubes, carbon nanofibers, andhigh-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 aqueoussolution by activated carbons obtained from biomass andcoals. 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 552

    http://refhub.elsevier.com/S0043-1354(15)00038-X/sref1http://refhub.elsevier.com/S0043-1354(15)00038-X/sref1http://refhub.elsevier.com