You are on page 1of 29

Critical Reviews in Environmental Science and

Technology

ISSN: 1064-3389 (Print) 1547-6537 (Online) Journal homepage: http://www.tandfonline.com/loi/best20

A review of biochar as a low-cost adsorbent for


aqueous heavy metal removal

Mandu I. Inyang, Bin Gao, Ying Yao, Yingwen Xue, Andrew Zimmerman,
Ahmed Mosa, Pratap Pullammanappallil, Yong Sik Ok & Xinde Cao

To cite this article: Mandu I. Inyang, Bin Gao, Ying Yao, Yingwen Xue, Andrew Zimmerman,
Ahmed Mosa, Pratap Pullammanappallil, Yong Sik Ok & Xinde Cao (2016) A review of biochar as
a low-cost adsorbent for aqueous heavy metal removal, Critical Reviews in Environmental Science
and Technology, 46:4, 406-433, DOI: 10.1080/10643389.2015.1096880

To link to this article: http://dx.doi.org/10.1080/10643389.2015.1096880

Accepted author version posted online: 25


Sep 2015.
Published online: 25 Sep 2015.

Submit your article to this journal

Article views: 1688

View related articles

Citing articles: 46 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=best20

Download by: [GGS Indraprastha University] Date: 21 August 2017, At: 00:03
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY
2016, VOL. 46, NO. 4, 406433
http://dx.doi.org/10.1080/10643389.2015.1096880

A review of biochar as a low-cost adsorbent for aqueous


heavy metal removal
Mandu I. Inyanga, Bin Gaoa,, Ying Yaoa, Yingwen Xueb, Andrew Zimmermanc,
Ahmed Mosad, Pratap Pullammanappallila, Yong Sik Oke, and Xinde Caof
a
Department of Agricultural and Biological Engineering, University of Florida, Gainesville, Florida, USA;
b
School of Civil Engineering, Wuhan University, Wuhan, China; cDepartment of Geological Sciences,
University of Florida, Gainesville, Florida, USA; dSoils Department, Faculty of Agriculture, Mansoura
University, Mansoura, Egypt; eKorea Biochar Research Center and Department of Biological Environment,
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Kangwon National University, Chuncheon, Korea; fSchool of Environmental Science


and Engineering, Shanghai Jiaotong University, Shanghai, China

ABSTRACT KEYWORDS
As a low-cost adsorbent, biochar can be used as a low-cost Black carbon; charcoal; trace
adsorbent for wastewater treatment, particularly with respect to element; wastewater;
treating heavy metals in wastewater. A number of studies have pyrogenic carbon
demonstrated effective removal of heavy metals from aqueous
solutions by biochar and, in some cases, proven the superiority
of biochars to activated carbons. Among several factors
affecting the sorption ability of biochars, feedstock materials
play a signicant role. This review incorporates existing literature
to understand the overall sorption behavior of heavy metals on
biochar adsorbents. Depending on the biochar type, heavy
metal can be removed by different mechanisms such as
complexation, physical sorption, precipitation and electrostatic
interactions. Mathematical sorption models can be used to
understand the efciency of biochar at removing heavy metals,
and promote the application of biochar technology in water
treatment.

1. Introduction
Biochar is pyrogenic black carbon derived from thermal degradation (e.g., pyrolysis) of
carbon-rich biomass in an oxygen-limited environment. In recent years, biochar has
received increasing attention due to its multifunctionality including carbon sequestra-
tion and soil fertility enhancement (Laird et al., 2010), bioenergy production (Field et
al., 2013), and environmental remediation (Mohan et al., 2014a). Several recent publi-
cations have provided evidences of biochars excellent ability to immobilize organic
(Inyang et al., 2014; Lattao et al., 2014; Rajapaksha et al., 2014) and inorganic pollutants
(Meng et al., 2014; Mohan et al., 2014b; Yao et al., 2011a) in soil and water systems.

CONTACT Bin Gao bg55@u.edu Department of Agricultural and Biological Engineering, University of Flor-
ida, Gainesville, FL 32611, USA.
Color versions of one or more of the gures in the article can be found online at www.tandfonline.com/best.
2016 Taylor & Francis Group, LLC
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 407

While most organic pollutants are biodegradable, inorganic pollutants, mainly heavy
metals, are nonbiodegradable and may be passed along the food chain through bioaccu-
mulation. Biochar is increasingly being considered as an alternative agent in water treat-
ment technologies for metal removal.
The presence of elevated concentrations of heavy metals in waste streams poses
challenges in environmental remediation. Table 1 summarizes several heavy metals
ranked in order of their toxicity by the Comprehensive Environmental Response
Compensation and Liability Act (Srivastava and Majumder, 2008). It also lists the
current maximum allowable contaminant levels (MCLs) provided by the Environ-
mental Protection Agency (EPA) for these metals. To meet the EPA standards, vari-
ous removal techniques (e.g., ion exchange, precipitation, electrocoagulation,
activated carbon adsorption, packed bed ltration) have been employed in wastewater
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

treatment facilities. Most of these techniques, however, are expensive and there is a
need to develop more cost-effective technologies with low-cost adsorbents (Kailash et
al., 2010).
Development of biochar technology provides opportunities to satisfy the need of
low-cost adsorbents for aqueous heavy metals. Mohan et al. (2007) noted that
removal efciencies of lead (Pb) and cadmium (Cd) by oak bark char are compara-
ble to that of Calgon F-400, a commercial activated carbon. Chen et al. (2011)
reported that biochar produced from wood or corn straw can effectively adsorb
copper (Cu) and zinc (Zn) in aqueous solutions. Similarly, Kong et al. (2011)
reported 7587% removal of mercury from aqueous solution by soybean stalk bio-
char. Further, the application of different engineering methods in biochar produc-
tion, such as pretreatment of feedstocks or modication of char surfaces, has
resulted in many high efciency and cost-effective engineered biochars with
adsorption capacities comparable to or even surpassing some commercial activated
carbons. For instance, previous studies (Inyang et al., 2011, 2012) showed that bio-
chars produced from anaerobic digested biomass have a much higher Pb sorption
capacity than commercial activated carbon. Physical treatment of the feedstock, by
pulverization prior to pyrolysis, can also greatly enhance the sorption ability of the
biochars to Cu in aqueous solution (Tong et al., 2011).

Table 1. Top heavy metals and their maximum allowable contaminant level.
Heavy metals Rank (CERCLA) Maximum contaminant level by EPA (mg/l)

Lead 2 0.015
Mercury 3 0.002
Cadmium 7 0.005
Chromium 17 0.1
Cobalt 52
Nickel 57
Zinc 75 5.0
Uranium 97
Copper 125 1.3
Manganese 140 0.05

Note. Maximum contaminant level was derived from goals from CERCLAs 2011 list of priority hazardous substances
and EPAs drinking water standards (Srivastava and Majumder, 2008).
408 M. I. INYANG ET AL.

Many aspects of biochar including its production, physical and biological proper-
ties, stability, and its applications have been assembled in a book, Biochar for Environ-
mental Management: Science and Technology (Lehmann, 2012). A recent study by
Laird et al. (2009) also reviewed the process of pyrolysis as a practical and effective
platform for producing renewable bioenergy products, bio-oil and biochar, as well as
simultaneously controlling greenhouse emissions. As for biochars applications, there
are several good reviews about the potential environmental applications of biochar to
immobilizing contaminants, including metals and organics, in soils (Beesley et al.,
2011; Kookana, 2010; Mohan and Pittman, 2007). Recently, two comprehensive
reviews (Ahmad et al., 2014; Mohan et al., 2014a) were published on several aspects
of biochar production and heavy metal sorption, but certain aspects such as heavy
metal sorption kinetic and thermodynamic behaviors on different biochar types are
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

yet to be reviewed in depth.


The overarching objective of this review is to provide a comprehensive review of
recent research ndings and theory developments on the role of biochar in removing
heavy metals from aqueous solutions. The specic objectives of this work are as fol-
lows: (a) assemble data on the removal of heavy metals by different types of biochar
sorbents, (b) discuss the effect of biochar feedstock materials on the sorption behavior
of heavy metals on biochars, and (c) review the mechanisms and mathematical mod-
els that can be used to describe heavy metal removal by biochar.

2. Biochar production and heavy metal removal ability


Biochar can be produced from a wide range of feedstock materials, including agricul-
tural and forest residues, industrial by-products and wastes, municipal solid waste
materials, and nonconventional materials, such as waste tires, papers, and even bones.
Table 2 shows a summary of the heavy metal removal ability of different biochars pro-
duced from various types of feedstocks from published studies.

2.1 Feedstock materials


2.1.1 Agricultural and forest residues
The growing use of agricultural and forest residues as biochar feedstock is due to
their abundance and low cost. In addition, conversion of waste biomass to value-
added biochar products can reduce operation cost associated with disposal of these
abundant waste agricultural and forest materials. Global annual production of
agricultural residues has been estimated to be 500 million tons (Duku
et al., 2011). The majority of agricultural and forest residues generated are waste or
by-products from harvesting and processing of crops such as sugarcane, corn, pea-
nut, sorghum, and oil palm for food and bioenergy production. Often, most of the
residues generated in many countries, including the United States, are scarcely uti-
lized, and thus can serve as feedstocks for biochar production. Furthermore, agri-
cultural and forest residues, such as wood, yard, and forest waste, are attractive
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 409

Table 2. Heavy metal adsorption capacities for biochar adsorbents.


Adsorption capacity for metals (mg/g)

Biochar Temp. ( C) Cr Hg Cd Ni Pb Cu Zn U Co Reference

Parent Biomass
Pinewood 300 4.3 (Liu and Zhang, 2009)
Rice husk 300 2.4 (Liu and Zhang, 2009)
Corn straw 600 12.5 11.0 (Chen et al., 2011)
Hardwood 450 6.8 4.5 (Chen et al., 2011)
Switch grass 300 2.1 (Kumar et al., 2011)
Pine wood 400/450 4.1 (Mohan et al., 2007)
Oak wood 400/450 0.4 2.6 (Mohan et al., 2007)
Pine bark 400/450 0.3 3.0 (Mohan et al., 2007)
Oak bark 400/450 5.4 13.1 (Mohan et al., 2007)
Sugar beet tailings 300 123.0 (Dong et al., 2011)
Soybean stalk 700 0.7 (Kong et al., 2011)
Flax shive 160 14.3 21.1 19.2 87.0 31.6 18.2 17.3 (El-Shafey et al., 2002)
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

200 25.5 39.1 33.4 147.1 52.1 32.5 31.2 (El-Shafey et al.,
2002)
Oak wood 400450 3.0 (Mohan et al., 2011)
Oak bark 400450 4.6 (Mohan et al., 2011)
Rice husk 300 4.6 (Pellera et al., 2012)
600 0.3 (Pellera et al., 2012)
Olive pomace 300 5.1 (Pellera et al., 2012)
600 0.7 (Pellera et al., 2012)
Orange waste 300 4.9 (Pellera et al., 2012)
600 0.4 (Pellera et al., 2012)
Compost 300 7.9 (Pellera et al., 2012)
600 3.4 (Pellera et al., 2012)
Sugarcane bagasse 600 6.5 (Inyang et al., 2011)
Dairy manure 200 140.9 (Cao et al., 2009)
350 101.7 (Cao et al., 2009)
Industrial by-products
Sewage sludge 550 30.88 (Lu et al., 2012)
Anaerobically digested 600 135.5 (Inyang et al., 2011)
sugarcane bagasse
Anaerobically digested 600 40.8
whole sugar beet
Anaerobically digested 600 51.4 (Inyang et al., 2012)
animal waste
Dried sewage sludge 650 3.00 40.30 6.70 (Otero et al., 2009)
Chemically activated
materials
H2O2 Peanut hull 300 22.8 (Xue et al., 2012)
H2SO4 Corn stalk 220 36.4 (Youssef et al., 2004)
ZnCl2 Corn stalk 600 32.4 (Youssef et al., 2004)
H3PO4/HNO3 Peanut shell 550 35.5 (Xu and Liu, 2008)
Nonconventional
materials
Waste tire rubber ash 500 22.4 34.4 (Mousavi et al., 2010)
Bone char 500 47.6 34.7 (Cheung et al., 2000a)

feedstock for the production of biochar due to their ease of collection from residen-
tial collection services and wood mills.
The strong sorption ability of biochars produced from agricultural and forest
residues may be attributed to their surface properties originating from the feed-
stock materials (Sun et al., 2014; Wang et al., 2015a). For example, the abundance
of oxygen containing groups (CDO, C O, OH) in oak bark biochars originat-
ing from polyphenolic tannins, avonoids, and suberin contents in the partially
410 M. I. INYANG ET AL.

aromatized bark materials provided negatively charged surface sites (COO and
OH) for the attraction of Pb2C(Mohan et al., 2007). Kumar et al. (2011) reported
that volatilization of large amounts of cellulose from switch grass during pyrolysis
creates porous carbon materials with surface functional groups strongly afliated
to aqueous U (VI). Pyrolysis of other agricultural residues, such as sugar beet tail-
ings, may also yield electron donor functional groups (C-OH, C-O, C-O-R)
because the feedstocks have complex heteropolysaccharides containing galactur-
onic acid, arabinose, galactose, and several pectin substances (Aksu and Isoglu,
2005). These oxygen functional groups can promote the sorption of aqueous chro-
mium, Cr (VI) on biochar by reducing Cr (VI) to Cr (III) to facilitate surface
adsorption (Dong et al., 2011). It has also been reported that elevated levels of cat-
ionic nutrients, such as Na, K, and Mg, in oak and pine wood feedstock materials
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

can also increase the cation exchange capacities of the corresponding biochars and
enhance their sorption of Pb through ion exchange process under acidic pH condi-
tions (Mohan et al., 2007).
Domesticated animal wastes are another abundant source of potential biochar
feedstock. Biochars produced from animal wastes such as poultry litter (including
bedding materials, spilled feed, and feathers) and dairy manure generally have
high amounts of ash and inorganic components that can bind heavy metals (Cao
et al., 2011; Duku et al., 2011; Uchimiya et al., 2010). Multiple mechanisms, includ-
ing precipitation and surface complexation, have been reported to be responsible
for the strong afliations between animal waste-derived biochar and aqueous
heavy metals (Cao et al., 2011; Uchimiya et al., 2010).

2.1.2 Industrial by-products


Production of biochar from industrial wastes and by-products has attracted a
growing interest recently. The feasibility of using waste materials collected from
bioenergy facilities (e.g., digested residues) and wastewater treatment plants (sew-
age sludge) for biochar production has been investigated (Phuengprasop et al.,
2011; Yao et al., 2011a; Yao et al., 2011b). Anaerobic digestion or sewage treatment
involves biodegradation of waste residues by a variety of microorganisms. The deg-
radation of carbonaceous substrates results in concentrating cationic or metallic
elements within the residues which when converted into biochars may have high
ion exchange capacity to remove heavy metals from aqueous solutions (Gu and
Wong, 2004; Hanay et al., 2008).

2.1.3 Nonconventional materials


Nonconventional materials such as waste tires (Karakoyun et al., 2011), periwinkle
shells (Bello and Ahmad, 2011a, 2011b), invasive plants (Liao et al., 2013), diatoms
and algae (Bird et al., 2012; Grierson et al., 2011), municipal solid waste (Ryu et al.,
2007; Hwang et al., 2008), newspaper scraps (Li and Zhang, 2004), plastics (Gil
et al., 2010), bones (Dimovic et al., 2011; Cheung et al., 2000a; Ko et al., 2004), bio-
energy residues (Inyang et al., 2010; Yao et al., 2015) and food wastes (Ahmed and
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 411

Gupta, 2010; Rhee and Park, 2010) have also been explored for producing carbon
sorbents. In particular, the use of waste tires as feedstock for biochar production is
attractive because waste tires are abundant, and pyrolysis of waste tires reduces the
risk and operation cost of waste tire disposal. Several studies have also shown that
pyrolyzed waste tire is an effective sorbent for heavy metals, such as Pb in aqueous
solutions (Mousavi et al., 2010), partly due to electrostatic interactions between the
positively charged cations and negatively charged biochar surface. Bone biochars
are heterogeneous sorbent materials derived from the combustion of crushed and
distilled animal bones (Pan et al., 2009). Cheung et al. (2000a) showed strong
uptake of Cu and Zn by bone biochars and suggested that high amounts of calcium
hydroxyapatite and calcium carbonate in the bone biochars promote ion-exchange
reactions between heavy metals (Cu, Zn) in solution and lattice Ca2C ions of the
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

apatite within the biochar (Cheung et al., 2000a; Pan et al., 2009).

3. Heavy metal removal mechanisms


A number of mechanisms may play a role in controlling the removal of heavy met-
als from aqueous solutions by biochar, including precipitation, complexation, ion
exchange, electrostatic interaction (chemisorption), and physical sorption
(Figure 1). Similar to activated carbon, biochars may have high sorption capacity
for metallic contaminants because of their surface heterogeneity (Kasozi et al.,
2010). In addition, many biochars have been reported to have high surface area
with well-distributed pore network including, micropores (<2 nm), mesopores (2
50 nm), and macropores (>50 nm; Mukherjee et al., 2011). Biochars with high sur-
face area and pore volumes have a high afnity for metals because metallic ions can
be physically sorbed onto char surface and retained within the pores (Kumar et al.,
2011). Many biochars have negatively charged surfaces and can sorb positively
charged metals through electrostatic attractions. Specic ligands and functional
groups on biochars can also interact with various metals to form complexes (Dong
et al., 2011; Wang et al., 2015b) or precipitates of their solid mineral phases (Cao

Figure 1. Removal mechanisms for biochar.


412 M. I. INYANG ET AL.

Table 3. Summary of removal mechanisms for biochar-based adsorbents.


Highest treatment
Biochar temperature ( C) Heavy metal Removal mechanism Reference

Cotton seed hull 350 Cd, Cu, Ni, and Pb Complexation of metals by (Uchimiya et al.,
char oxygen functional groups 2011)
Rice husk carbon 175180 Hg, Zn Ion exchange between HC and (El-Shafey, 2010)
Zn2C and Hg2C
Flax shive 200 Cd Ion exchange between HC and (El-Shafey et al.,
carbon Cd2C 2002)
Beech wood 550 Cu Complexation of Cu by carboxyl (Borchard et al.,
char functional groups 2012)
Wheat and rice open eld burning Pb Electrostatic interaction (Qiu et al., 2008)
char between Pb2C and
negatively charged
functional groups on biochar
Sugar beet 300 Cr Electrostatic attraction between (Dong et al.,
tailings char Cr(VI) and biochar and 2011)
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

complexation with carboxyl


and phenolic groups
Anaerobically 600 Pb Precipitation, complexation of (Inyang et al.,
digested Pb2C with carboxyl 2011)
bagasse functional groups
biochar
Digested animal 600 Pb, Cu, Ni, Cd Precipitation (Inyang et al.,
waste char 2012)
Digested whole 600 Pb, Cu, Ni, Cd Precipitation (Inyang et al.,
sugar beet 2012)
char
Waste tire char 500 Pb Precipitation and formation of (Mousavi et al.,
lead hydroxide 2010)
Dairy manure 25 and 200 Pb Precipitation and complexation (Cao et al., 2009)
biochar with phosphate ligand
Hydrothermal 300 U Surface adsorption and binding (Kumar et al.,
switch grass of U(VI) pores of biochar 2011)
char
Wood and bark 400 and 450 As, Cd, and Pb Precipitation, complexation of (Mohan et al.,
chars Pb(II) and Cd(II) with 2007)
hydroxyl functional groups
Hydrothermal 300 Pb Complexation of Pb2C with (Liu and Zhang,
rice husk char oxygen functional groups 2009)
Bone char Not reported Cu, Zn Ionic exchange between Cu (II) (Cheung et al.,
and Zn(II) with Ca(II) from 2000a)
calcium hydroxyapatite and
calcium carbonate in bone
char
Waste tire char 550 Cu Precipitation, surface adsorption (Quek et al.,
and diffusion of Cu into 2010)
pores of biochar
Bone char Not reported Cd, Cu, Zn Surface adsorption and diffusion (Choy and
into pores McKay, 2005)
Hydrothermal 300 Cu Ion exchange (Liu et al., 2010)
pinewood
char
Pyrolytic 700 Cu Surface adsorption (Liu et al., 2010)
pinewood
char
Swine bone char 800 Co Ion exchange (Pan et al., 2009)
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 413

et al., 2009; Inyang et al., 2011; 2012). Table 3 summarizes the governing mecha-
nisms for the immobilization of aqueous metallic species by various biochars.

3.1 Physical sorption

Physical or surface sorption describes the removal of heavy metals by diffu-


sional movement of metal ions into sorbent pores without formation of chem-
ical bonds. For both animal and plant biochars, increasing temperatures of
carbonization ( 300 C) will favor high surface areas and pore volumes in
biochars. For example, biochars converted from switch grass (300 C) and pine
wood (700 C) can both effectively remove aqueous uranium (U) and Cu
through diffusive surface adsorption process (Kumar et al., 2011; Liu et al.,
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

2010). Animal bone chars were also reported (Choy and McKay, 2005) to
retain Cd (53.6 mg/g), Cu (45.04 mg/g), and Zn (33.03 mg/g) in their pore
networks, and their sorption data was well described by lm pore diffusion
controlled model.

3.2 Ion exchange


Sorption of heavy metals through exchange of ionizable protons/cations on biochar
surfaces with dissolved metal species is another possible mechanism. The efciency of
the ion exchange process in retaining heavy metal contaminants on biochar is closely
related to the size of the metal contaminant and surface functional group chemistry of
the biochar. Ionic exchange of metal species on biochar occurs by a selective replace-
ment of positively charged ions on biochar surfaces with target metal species. Following
the Goldschmidts geochemical classication, most metal species within groups 13
(e.g., Na, K, Ca, Li, Mg, Be, and Sc) on the periodic table are more likely to be substi-
tuted or exchanged with other metallic elements within this group based on the similar-
ities of their ionic radii, charge differences, and bond characteristics (Krauskopf, 1967).
Several transition elements may also have a strong binding ability to these exchange
sites. For most plant materials, cation exchange capacity (CEC) is mainly controlled by
surface functional groups. High CECs of plant biochars are indicative of good heavy
metal removal ability. In general, nonwoody and grassy biochars with high O contents
and acidic surface sites have higher CECs than woody plant chars having low O func-
tionalities (Harvey et al., 2011). The highest CECs of chars have been observed in plant
biochars produced at 300 or 350 C due to higher contents of carboxylic acids. In con-
trast, low CEC values are observed at relatively high temperatures (>350 C). Accord-
ingly, rice husk biochar produced at temperatures of 175180 C showed strong
sorption of Zn and Hg mainly through deprotonation of HC and subsequent exchange
of Zn2C and HgC (El-Shafey, 2010). Bone biochars rich in cation species can also sorb
heavy metals such as cobalt (Co) from aqueous solutions through the ion exchange
mechanism (Pan et al., 2009).
414 M. I. INYANG ET AL.

3.3 Electrostatic interactions


Electrostatic interaction between surface charged biochars and metal ions is another
mechanism for the immobilization of heavy metals. Prevalence of this mechanism in
biochar-metal sorption process is dependent on the solution pH and point of zero
charge (PZC) of biochar (Dong et al., 2011; Mukherjee et al., 2011). High tempera-
tures (>400 C) of carbonization also promote the formation of graphene structures
in the chars to favor electrostatic attractions sorption mechanisms (Keiluweit and
Kleber, 2009). High retention of Pb in wheat and rice char was reported (Qiu et al.,
2008), due to the attraction between the positively charged Pb and negatively charged
biochars. Electrostatic attractions have also been reported for negatively charged Cr
(VI) to positively charged biochar surfaces at pH of 2 (Dong et al., 2011).
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

3.4 Complexation
Complexation (outer- and innersphere) involves the formation of multiatom structures
(i.e., complexes) with specic metal-ligand interactions. This binding mechanism is
likely important for metals, such as transition metals with partially lled d-orbitals, hav-
ing a high afnity for ligands (Crabtree, 2009). In particular, oxygen functional groups
(carboxyl, phenolic, and lactonic) in low-temperature biochars have been demonstrated
to effectively bind with heavy metals (Liu and Zhang, 2009; Mohan et al., 2007). Oxygen
content of biochars has also been shown to increase over time, likely due to oxidation of
biochar surface and the formation of carboxyl group (Harvey et al., 2011), thus, metal
complexation may increase over time.
From Table 3, it can be inferred that complexation of metallic species is more popular
with plant-derived than animal-derived biochars. As a result, plant-derived chars readily
bind with several heavy metals, such as Cd, Cu, Ni, and Pb to precipitate phenolic and
carboxylic-metal complexes on the chars. Similarly, some animal derived chars in partic-
ular, dairy manure and chicken litter carbons can also bind metals such as Pb with phos-
phate ligands which originate from the nutrient rich, feeds to form metallic complexes
such as pyromorphite (Cao et al., 2009; Guo et al., 2010).

3.5 Precipitation
Precipitation is the formation of solid(s), either in solution or on a surface, during
the sorption processes. Precipitation has been commonly cited as one of the main
mechanisms responsible for the immobilization of heavy metals by biochar sorb-
ents. Metals and rare earth elements with intermediate ionization potentials
between 2.5 and 9.5 (e.g., Cu, Zn, Ni, Pb) are more likely to precipitate (Krauskopf,
1967) on biochar surfaces than other elements. The thermal degradation of cellu-
lose and hemicellulose in plant biomass at relatively high temperatures (>300 C)
often produces alkaline biochars (Cao and Harris, 2010). When in solution, these
biochars may trigger the precipitation of metallic species. For instance, previous
study had showed that digested bagasse biochar (pH of 10.93) was able to
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 415

precipitate Pb by the formation hydrocerrusite [(Pb3(CO3)2(OH)2] on the surface


of the char (Inyang et al., 2011).
In addition to high pH of the biochars, precipitation of heavy metals may arise
from the reaction of metallic ions with several mineral phases entrained in both plant
and animal feedstock. Animal and plant biochars produced at high temperatures
have high mineral matter content with signicant levels of calcium (Ca), magnesium
(Mg), iron (Fe), Cu, and silicon (Si) elements. This is particularly true for biochars
from animal wastes, such as chicken litter biochars (45% mineral matter; Koutcheiko
et al., 2007) and bone biochars (84% mineral matter) (Purevsuren et al., 2004). These
mineral phases including, slyvite (KCl), quartz (SiO2), amorphous silica, calcite
(CaCO3), hydroxyapatite (Ca10PO4)6(OH)2), and calcium anhydrite (CaSO4) exist in
either free forms or are intercalated within the carbon matrix of the biochars. Cao
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

and Harris (2010) suggested that soluble forms of these phases also exist in low tem-
perature biochars (<200 C), and can directly react with heavy metal species to form
insoluble metallic precipitates such as pyromorphite [Pb5(PO4)3Cl] (Cao et al., 2009).
Less soluble forms of these mineral phases exist at higher temperatures (350600 C),
and are more likely to be slowly released during sorption reactions with heavy metals
to form precipitates on biochar surfaces (Inyang et al., 2011).

4. Mathematical modeling
4.1 Adsorption thermodynamics
Thermodynamic models are powerful tools to describe metal sorption processes and to
explore governing mechanisms. Thermodynamic parameters of metal sorption on bio-
chars reported in the literatures have been summarized in Table 4. The thermodynamic
behavior of heavy metal sorption on biochars can be described as either exothermic
(sorption decreases with increasing temperature) or endothermic sorption processes
(sorption increases with increasing temperature). Three thermodynamic parameters are
often used to characterize the thermodynamics: (a) enthalpy, DH ; (b) entropy, DS ; and
(c) Gibbs free energy, DG . According to Chen et al. (2011), these thermodynamic
parameters can be computed from the following equations:

qe
Ke D (1)
Ce
DG0 D RTlnKe (2)
B
DG D DH TDS
0
(3)
DS 0
DH 0
lnKe D (4)
R R
where qe (mg/g) is the amount of heavy metals adsorbed on biochar at equilibrium,
Ce (mg/l) is the equilibrium concentration of heavy metals in the solution, R (J/
mol_K) is the gas constant (8.314), T (K) is the absolute temperature, and Ke (l/g) is
the adsorption equilibrium constant. By plotting ln Ke against 1/T, the values of
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

416
Table 4. Summary of thermodynamic parameters by biochar-based sorbents.
Thermodynamics parameters

Biochar (temperature [ C]) Metal Temperature (K) DG0 (KJ/mol) DH0 (KJ/mol) DS0 (J/mol K) Sorption type Reference

Pine wood (300) Pb 298 14.68 41.1 187.15 Endothermic, spontaneous (Liu and Zhang, 2009)
308 16.55
318 18.42
Rice husk (300) Pb 298 13.20 37.13 168.88 Endothermic, spontaneous (Liu and Zhang, 2009)
308 14.89
318 16.58
Corn straw (600) Cu 295 -0.80 25.27 88.21 Endothermic, spontaneous (Chen et al., 2011)
303 1.39
310 2.13
Zn 295 1.45 30.94 99.85 Endothermic (Chen et al., 2011)
303 0.68
310 0.05
Hard wood (450) Cu 295 5.38 9.22 13.09 Endothermic, non spontaneous (Chen et al., 2011)
303 5.2
310 5.19
Zn 295 6.39 7.55 3.99 Endothermic, nonspontaneous (Chen et al., 2011)
303 6.31
310 6.34
Barrabus barkcarbon Fea 303 2.59 0.92 5.42 Endothermic, spontaneous (Arivoli et al., 2008)
313 2.81
323 3
333 3.18
Coa 303 1.8 2.16 8.82 Endothermic, spontaneous (Arivoli et al., 2008)
313 2.21
323 2.52
333 2.74
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Table 4. (Continued ).
Thermodynamics parameters

Biochar (temperature [ C]) Metal Temperature (K) DG0 (KJ/mol) DH0 (KJ/mol) DS0 (J/mol K) Sorption type Reference
a
Ni 303 0.615 1.15 5.21 Endothermic, spontaneous (Arivoli et al., 2008)
313 0.655
323 0.708
333 0.77
Oxidized peanut shell carbon Cr(VI) 293 28.38 0.04 0.1 Endothermic and spontaneous (Al-Othman et al., 2012)
303 29.39
313 30.32
Unoxidized peanut shell carbon Cr(VI) 293 11.73 0.02 0.04 Endothermic, spontaneous (Al-Othman et al., 2012)
303 11.13
313 12.53
a
Thermodynamic parameter values were obtained at a concentration of 30 mg/l.

417
418 M. I. INYANG ET AL.

DH and DS can be determined from the slopes and intercepts, respectively. The
values of DG can then be calculated from the corresponding values of DH and
DS . Positive DH values indicate that the sorption reaction is endothermic, sup-
ported by an increase in qe (mg/g) with increasing temperature. On the other hand,
negative DG values suggest a spontaneous sorption process with increasing metal
sorption at higher temperatures. Positive values of DS may reect an afnity of the
carbon sorbent for the metal ions (Lu et al., 2009). Free energy, DG (KJ/mol) could
also provide information to distinguish physical sorption (DG , 20 to 0 KJ/mol)
from chemisorption (DG , 400 to 80 KJ/mol) processes (Liu and Zhang, 2009).
Traditionally, thermodynamics of biochar sorption to heavy metals have been
investigated in conjunction with exploring the effect of temperature on the sorption
capacity of biochar. For instance, adsorption studies by Liu and Zhang (2009) with
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

hydrothermally carbonized biochars demonstrated that sorption capacities for pine


and rice husk biochars to Pb increased with increasing temperatures. This suggests
that Pb sorption process was endothermic and is in agreement with the positive
enthalpy values of pine (41.10 KJ/mol) and rice husk (37.13 KJ/mol) biochars
obtained. Higher temperatures would provide sufcient energies for the transfer of
Pb to the interiors of the biochar pores. Similarly, sorption of Cu and Zn by corn
straw and hard wood biochars was also endothermic with positive free energies
(Table 5), but the sorption was not consistently spontaneous with increasing tem-
peratures (22, 30, and 37 C).

4.2 Adsorption kinetics


Several mathematical models have been developed to describe the sorption kinetics
of metals on biochar sorbents. These kinetic models have been classied into two
main types: reaction-based and diffusion based models (Ho et al., 2000). While dif-
fusion models are focused on the diffusive transport of metal ions from aqueous
solutions into pore networks and active sites of the biochar sorbent, reaction mod-
els describe interaction rates between biochar and specic metal ions (Qiu et al.,
2009). The most commonly used adsorption kinetic models are the rst or pseudo
rst order, second or pseudo second order, Elovich, and intraparticle diffusion
models. Mathematical expressions describing these models are the following:
Pseudo first order qt D qe .1 e k1 t / (5)
k2 qe 2 t
Pseudo second order qt D (6)
1 C k 2 qe t
1
Elovich qt D ln.abt C 1/ (7)
b
Intraparticle diffusion qt D K1 t 1=2 C W (8)

where qt (mmol kg1) and qe (mmol kg1) are the amounts of metal sorbed at time
t and at equilibrium respectively; k1 (h1) and k2 (kg mmol1 h1) are the rst-
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Table 5. Best t model parameters for biochar sorption of heavy metals.


Biochar type
(temperature
[ C]) Metal Model Parameter 1 Parameter 2 Parameter 3 Parameter 4 Reference
1 1
Digested animal Pb(II) First order k1 (h ) D 0.280 qe (mmol kg ) D 266 (Inyang et al., 2012)
waste (600) Second order k2 (kgmmol1 h1) D 0.00100 qe (mmol kg 1) D 311
Elovich a (mmol kg1 h1) D 174 b (kgmmol1) D 0.0140
Langmuir K (l/mmol) D 928 Smax (mmol kg1) D 248
Freundlich Kf (mmol(1n)Lnkg1) D 248 n D 0.0619
Double Langmuir K 1 (l/mmol) D 1120 Smax1 (mmol kg1) D 234 K2(l/mmol) D 0.00233 Smax2(mmol kg-1) D 1.90
Digested whole Pb(II) First order k1 (h1) D 0.181 qe (mmol kg1) D 203
sugarbeet Second order k2 (kgmmol1h1) D 0.000760 qe (mmol kg1) D 250
1 1
(600) Elovich a (mmol kg1 h1) D 67.7 b (mmol kg h ) D 0.0150
Langmuir K (l/mmol) D 266 Smax (mmol kg1) D 197
Freundlich Kf (mmol(1n)l kg1) D 189 n D 0.140
Double Langmuir K 1 (l/mmol) D 351 Smax1 (mmol kg-1) D 172 K2 /l/mmol) D 2.87 Smax2(mmol kg-1) D 43.9
Soybean stalk Hg(II) Langmuir K (l/mg) D 0.034 Smax (mg/g) D 674.909 (Kong et al., 2011)
(700) Freundlich Kf (l/g) D 72.353 n D 0.382
Pinewood Pb(II) Intraparticle diffusion Kd (mg g-1h-1/2) D 0.78 C (mg/g) D 2.01 (Liu and Zhang, 2009)
hydrochar First order k1 (h1) D 0.08 qe (mg/g) D 0.87
(300) Second order k2 (g mg1 h1) D 0.86 qe (mg/g) D 2.50
Langmuir K (l/mg) D 0.36 Smax (mg/g) D 3.89
Freundlich Kf ((mg/g)(l/mg)1/n) D 1.75 n D 4.77
Ricehusk Pb (II) Intraparticle diffusion Kd (mg g-1h-1/2) D 0.47 C (mg/g) D 1.80
hydrochar First order k1 (h1) D 0.12 qe (mg/g) D 0.89
(300) Second order k2 (g mg1 h1) D 1.30 qe (mg/g) D 1.76
Langmuir K (l/mg) D 0.21 Smax (mg/g) D 1.84
Freundlich Kf ((mg/g)(l/mg)1/n) D 0.35 n D 2.07
Corn straw (300) Cu(II) First order k1 (h1) D 0.083 qe (mg/g) D 8.93 (Chen et al., 2011)
Second order k2 (g mg1 h1) D 0.003 qe (mg/g) D 10.76
Langmuir K (l/mg) D 0.682 Smax (mg/g) D 12.52
Freundlich Kf ((mg/g)(l/mg)1/n) D 3.71 n D 3.605
Zn(II) First order k1 (h1) D 0.13 qe (mg/g) D 7.11
Second order k2 (g mg1 h1) D 0.006 qe (mg/g) D 8.20
Langmuir K (l/mg) D 0.232 Smax (mg/g) D 11.00
Freundlich Kf ((mg/g)(l/mg)1/n) D 2.84 n D 3.336
(Continued)

419
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

420
Table 5. (Continued ).
Biochar type
(temperature
[ C]) Metal Model Parameter 1 Parameter 2 Parameter 3 Parameter 4 Reference

Hardwood (450) Cu(II) First order k1 (h1) D 0.100 qe (mg/g) D 3.70


Second order k2 (g mg1 h1) D 0.011 qe (mg/g) D 4.31
Langmuir K (l/mg) D 0.048 Smax (mg/g) D 6.79
Freundlich Kf ((mg/g)(l/mg)1/n) D 0.71 n D 2.294
Zn(II) First order k1 (h1) D 0.141 qe (mg/g) D 2.63
Second order k2 (g mg1 h1) D 0.009 qe (mg/g) D 3.14
Langmuir K (l/mg) D 0.061 Smax (mg/g) D 4.54
Freundlich Kf ((mg/g)(l/mg)1/n) D 0.72 n D 2.827
Pine bark (400/ Pb (II) Langmuir K (l/mg) D 22.67102 Smax (mg/g) D 3.00 (Mohan et al., 2007)
450) Cd (II) K (l/mg) D 0.02102 Smax (mg/g) D 0.34
As(III) K (l/mg) D 0.0007102 Smax (mg/g) D 12.15
Pb (II) Freundlich Kf (mg/g) D 1.28 n D 0.15
Cd (II) Kf (mg/g) D 0.40 n D 0.35
As(III) Kf (mg/g) D 0.041 n D 0.69
Oak wood (400/ Pb (II) Langmuir K (l/mg) D 16.36102 Smax (mg/g) D 2.62
450) Cd (II) K (l/mg) D 3.77102 Smax (mg/g) D 0.37
As(III) K (l/mg) D 0.0009102 Smax (mg/g) D 5.85
Pb (II) Freundlich Kf (mg/g) D 0.77 n D 0.22
Cd (II) Kf (mg/g) D 0.23 n D 0.12
As(III) Kf (mg/g) D 0.021 n D 0.09
Olive pomace Cu(II) First order k1 (min1) D 0.006 qe (mg/g) D 0.101 (Pellera et al., 2012)
(300) Second order k2 (g mg1 min1) D 0.037 qe (mg/g) D 1.305
Langmuir K (l/mg) D 0.089 Smax (mg/g) D 5.118
Freundlich Kf (l/g) D 0.857 n D 2.788
Intraparticle diffusion Kd (mg g1min1/2) D 0.78 C (mg/g) D 1.0575
Orange waste Cu(II) First order k1 (min1) D 0.0002 qe (mg/g) D 0.357
(300) Second order k2 (g mg1 min1) D 0.029 qe (mg/g) D 1.404
Langmuir K (l/mg) D 0.077 Smax (mg/g) D 4.921
Freundlich Kf (l/g) D 0.783 n D 2.827
Intraparticle diffusion Kd (mg g1min1/2) D 0.011 C (mg/g) D 0.9271
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Table 5. (Continued ).
Biochar type
(temperature
[ C]) Metal Model Parameter 1 Parameter 2 Parameter 3 Parameter 4 Reference
1
Compost (300) Cu(II) First order k1 (min ) D 0.003 qe (mg/g) D 0.543
Second order k2 (g mg1 min1) D 0.107 qe (mg/g) D 3.630
Langmuir K (l/mg) D 0.359 Smax (mg/g) D 7.937
Freundlich Kf (l/g) D 1.701 n D 1.744
Intraparticle diffusion
Kd (mg g1min1/2) D 0.018 C (mg/g) D 3.0078
Sludge (650) Hg(II) Langmuir K (l/mg) D 0.044 Smax (mg/g) D 64.9 (Otero et al., 2009)
Pb(II) K (l/mg) D 0.019 Smax (mg/g) D 40.3
Cu(II) K (l/mg) D 0.022 Smax (mg/g) D 6.7
Cr(III) K (l/mg) D 0.015 Smax (mg/g) D 3.0
Oak bark (400/ Cr(VI) First order k1 (h1) D 1.49 qe (mg/g) D 1.49 (Mohan et al., 2011)
450) Second order k2 (g mg1 h1) D 1.63 qe (mg/g) D 1.63
Langmuir K (l/mg) D 0.073 Smax (mg/g) D 4.619
Freundlich Kf (l/g) D 0.523 n D 2.016
Langmuir-Freundlich KL-F (l/g) D 0.182 aL-F (l/mg) D 0.048 nL-F D 1.461
Dairy manure Pb (II) Langmuir K (l/mmol) D 61.3 Smax (mmol/kg) D 452 (Cao et al., 2009)
(350) Double Langmuir K1 (l/mmol) D 0.91 Smax1 (mmol/kg) D 78.3 K2 (l/mmol) D 69.3 Smax2 (mmol/kg) D 413

421
422 M. I. INYANG ET AL.

and second-order apparent sorption rate constants, respectively; a (mmol kg1


h1) and b (kg mmol1) are the initial Elovich sorption and desorption rate con-
stant at time t, respectively; K1 is the intraparticle diffusion rate constant (mmol
kg1h0.5); and W (mmol/kg) is a constant.

4.2.1 Pseudo rst-order model


Among all adsorption kinetic models, the Lagergren pseudo rst-order model is
perhaps the simplest and earliest form, describing the rate of adsorption in liquid
phase systems (Gupta and Bhattacharyya, 2011). The model is applied in its linear-
ized form as follows

ln.qe qt / D ln.qe / k1 t (9)


Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

The pseudo rst-order adsorption rate coefcient, k1 (h1), can be evaluated as the
slope from the plot of ln (qeqt) versus ln qe. The value of qe is unknown and the
amount sorbed is signicantly smaller than the actual equilibrium amount (Gerente
et al., 2007). As such, the pseudo rst-order may only apply for the initial 2030 min
of most sorption processes, after which, the model may not t the whole sorption
data within the entire range of contact times (Gupta and Bhattacharyya, 2011). When
this occurs, the plot of ln (qe qt) versus ln qe ceases to be linear and an accurate
representation of the sorption data would require a more complex model, such as the
second order kinetic model. Despite limitations of the pseudo rst order model, this
model has been successfully applied to describe the sorption kinetics of heavy metals
on various biochar sorbents (Table 5). For instance, Zuo et al. (2012) reported that
the pseudo rst-order model closely reproduced data from the sorption of Cu, Cd,
and Zn by carbonized mulch. Plotted kinetic curves for Cd, Zn, and Cu on the car-
bonized mulch, revealed a rapid sorption of these metals for the initial 30 min (typical
for rst order reactions). However, as more adsorption sites progressively lled up,
the sorption rate gradually reduced until equilibrium.

4.2.2 Pseudo second-order model


Pseudo second-order kinetic models are used for chemisorption or sorption pro-
cesses involving chemical bonding between metal ions and functional groups of
the adsorbent (Ho, 2006). In contrast to the rst order model, the linearized form
of the pseudo second-order model does not require assigning a value for qe
(Gerente et al., 2007).

t=qt D 1=k2 qe 2 C 1=qe t (10)

A plot of t/qt against t gives a linear relationship, from which values of qe and k2
can be determined from the slope and intercept of the plot, respectively. Typically,
the rate of the second order reaction depends on the amount of adsorbate (metal
ion) on the surface of the adsorbent (biochar) and the amount of adsorbate in
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 423

solution at equilibrium (Ho, 2006). The pseudo second-order kinetic model has
been used to effectively correlate sorption data for the removal of Cu and Zn by
corn straw and hardwood biochars with coefcients of determination above 0.98
(R2 > 0.98; Table 5; Chen et al., 2011). The close t of the pseudo second-order
model to the adsorption data suggests that the model could be applied to the whole
time range of the sorption process with chemisorption being the rate-limiting
mechanism for the retention of Cu (II) and Zn (II) on the biochars. Other reported
pseudo second-order model parameters of heavy metal sorption on biochars are
summarized in Table 5.

4.2.3 Elovich model


The Elovich model was originally developed to describe chemisorption kinetics
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

(Wu et al., 2009a), and has been proposed for modeling sorption processes of
adsorbents with heterogeneous surfaces (Cheung et al., 2000b). The model can be
used as a combination of two or more simultaneous rst order kinetic equations to
describe experimental data that is not accurately described by a single rst-order
equation (Chien and Clayton, 1980). It is also a powerful tool in evaluating homo-
geneity of adsorbent surfaces during the whole course of sorption reaction time.
Thus, any deviations in linearity of the linearized Elovich equation would reect
changes in surface reactivity (Chien and Clayton, 1980). There have been some
attempts at interpreting experimental data from the sorption of heavy metals by
biochars with the Elovich model (Table 5). For instance, Pb sorption kinetics study
showed that the Elovich model reproduced experimental data of Pb sorption on
digested whole sugar beet biochar much better than either the rst- or second-
order models (Inyang et al., 2012).

4.2.4 Diffusion model


Metal sorption process on biochars could proceed in the following order (a) exter-
nal mass transport of the metal solute across the boundary layer surrounding the
biochar particles; (b) diffusional mass transfer of solute within internal and exter-
nal pores of the biochar, or a combination of both; and (c) adsorption on surface
site (Ho and McKay, 1998). Diffusion models are developed to describe the sorp-
tion kinetics of heavy metals on biochar during the initial sorption process when
the interaction is limited by diffusional mass transfer rate (Gerente et al., 2007).
Previously, diffusional mass transport models have been extensively applied to the
removal of organics and dyes from wastewaters (Ahmad et al., 2013; Dizge et al.,
2008; Ofomaja, 2008; Wu et al., 2009b). But recently, diffusional models, particu-
larly the intraparticle diffusion model, have been increasingly applied to describe
the kinetics of heavy metal sorption on adsorbents (Ofomaja, 2010; Ranjan et al.,
2009; Sag and Aktay, 2000).
The intraparticle diffusion equation can be modeled by plotting the amount of
solute adsorbed at any time (qt) against the square root of time (t1/2) to obtain a
straight line that may not necessarily pass through the origin (Choy et al., 2004;
424 M. I. INYANG ET AL.

Wu et al., 2009b). Previous study (Inyang et al., 2011) noted that intraparticle dif-
fusion may more likely occur in microporous than in macroporous adsorbents. In
addition, this study reported that pre-equilibrium sorption of lead on bagasse bio-
char had a strong linear dependency on the square of time.

4.3 Adsorption isotherms

Equilibrium models and sorption isotherms are used to explore how an adsorbate
interacts with an adsorbent (Ai et al., 2011). Common equilibrium models used to
describe isotherms of metal sorption on biochars include Langmuir, Freundlich,
Langmuir-Freundlich, and double Langmuir model, which are expressed as the fol-
lowing:
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Smax KC
Langmuir S D (11)
1 C KC
Freundlich S D Kf C n (12)
Smax .KC/n
Langmuir Freundlich S D (13)
1 C .KC/n
Smax1 K1 C Smax2 K2 C
Double Langmuir S D C (14)
1 C K1 C 1 C K2 C

where S (mmol/kg) is the amount of metal adsorbed and Smax (mmol/kg) is the
maximum amount of metal adsorbed in mmol/kg; Smax1 (mmol/kg) and Smax2
(mmol/kg) are the maximum amounts of metal adsorbed related to the sorption
and precipitation processes, respectively; K (l/mmol) is the Langmuir adsorption
constant related to the interaction bonding energy; Kf (mmol(1n)Ln/kg) is the
Freundlich equilibrium constant in l/mmol; K1 and K2 are the Langmuir bonding
terms related to sorption and precipitation energies in l/mmol, respectively; C is
the equilibrium solution concentration in mmol/l of the sorbate; and n is the
Freundlich linearity constant.

4.3.1 Langmuir model


The Langmuir model is probably the best known and most widely used sorption
isotherm model (Gerente et al., 2007). This mechanistic model is based on several
assumptions, including monolayer adsorption (adsorbed layer is one molecule of
thickness), in which sorption only occurs at a nite number of identical active sites.
Also, there are no lateral interactions or steric hindrances between adsorbed mole-
cules and adjacent sites (Foo and Hameed, 2010; Vijayaraghavan et al., 2006). In
addition, the Langmuir model implicitly assumes that binding to an adsorbents
surface is driven by physical forces and all the sites are energetically equivalent
with equal afnity for the sorbate (Perez-Marin et al., 2007; Vijayaraghavan et al.,
2006). The Langmuir equation has been used to describe the sorption of metals by
biochar (Table 5). For example, the Langmuir model was successfully applied to
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 425

describe the sorption isotherms for Pb sorption on bagasse biochars (Inyang et al.,
2011).

4.3.2 Freundlich model


The Freundlich model is often applied to nonideal sorption on heterogeneous sur-
faces as well as multilayer sorption (Gerente et al., 2007). Unlike the Langmuir
model, the Freundlich equation does not assume that the sorption sites are ener-
getically equivalent and homogenous. On the contrary, the model assumes that the
sorption sites of the adsorbent are of varied afnities, and the stronger binding sites
are occupied rst such that the binding strength decreases with increasing degree
of site occupation (Vijayaraghavan et al., 2006). Thus, the amount adsorbed is a
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

summation of adsorption on all sites (each site having its own bond energy) and
the energy distribution of the sorption sites exponentially decays, up until the com-
pletion of the adsorption process (Foo and Hameed, 2010; Gerente et al., 2007).
The Freundlich equation has been successfully applied to describe the sorption iso-
therms of heavy metals such as Cu, Zn, and Pb on various biochars (Table 5).

4.3.3 Langmuir-Freundlich and Langmuir-Langmuir model


The Langmuir-Freundlich (L-F) model, also known as the Sips equation, is a ver-
satile isotherm expression that can simulate both Langmuir and Freundlich type
adsorption behavior (Jeppu and Clement, 2012). The L-F model derives its name
from the limiting behavior of the model, and predicts surface saturation at high
concentrations (Mohan et al., 2011). According to Gerente et al. (2007), the L-F
isotherm is a exible equation that reduces to the Freundlich model at low equilib-
rium concentrations (Ce) or low pressures, while, it reduces to the Langmuir
model when the Freundlich linearity constant, n D 1. One drawback to the L-F
model, however, is that it suffers thermodynamic inconsistencies at zero surface
coverage because it does not exhibit a nite Henrys constant, similar to the Lang-
muir model (Al-Muhtaseb et al., 2008). Nonetheless, the model was used in tting
equilibrium data for the sorption of Cr(VI) by oak bark chars at different tempera-
tures (2545 C; Mohan et al., 2011). Findings from Mohan et al. (2011) suggested
that the Sips model was a better t (R2 D 0.98 at 25 C) for the sorption of Cr(VI)
because at low concentrations, adsorption was controlled by diffusion, while at
higher concentrations, monomolecular saturation occurred on the surface of the
char conrming the Sips assumption of surface saturation at high concentrations.
In contrast, the Langmuir-Langmuir (L-L) model describes sorption on hetero-
geneous surfaces and assumes that sorption occurs on two types of surfaces with
contrasting bonding energies (Holford et al., 1974). The L-L model was used to
describe sorption of Pb on dairy manure biochar and model parameters on the
sorption sites were differentiated by assigning the rst site to physical adsorption
(K1, Smax1) and second site to precipitation (K2, Smax2; Cao et al., 2009).
426 M. I. INYANG ET AL.

4.4 Filtration model


Large-scale industrial operations often rely on xed-bed ltration systems to
remove contaminants from wastewater streams. As a result, ltration models have
been developed to optimize the xed-bed systems and to describe the fate and
transport of contaminants (Gerente et al., 2007). Most xed-bed systems are
packed with adsorbents as porous media to retain pollutants from the ow passing
through. Small size porous medium columns are often used in laboratory to mimic
the ltration systems and to generate data to validate the ltration models, such as
the advection-dispersion equation (ADE)based models (Genc-Fuhrman et al.,
2005; Jaradat et al., 2009; Scheytt et al., 2004). In cases, where biochar is used as l-
ter media in xed-bed settings to remove heavy metals, the ltration model of con-
taminants in porous media is still applicable. Thus, the governing equation of
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Figure 2. (a) Comparison of breakthrough curve data for the sorption of lead on modied peanut
hull hydrochar and unmodied peanut hull hydrochar (PHHC) and (b) breakthrough curve data for
sorption of multicomponent sorption of Cd, Ni, Cu, and Pb on modied peanut hull hydrochar
(mPHHC (Xue et al., 2012)).
Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the
rightsholder.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 427

heavy metals in biochar-based lters can be written as (Xue et al., 2012)

@Cw @ 2 Cw @Cw
DD 2 v S (15)
@t @z @z

where Cw is the heavy metal concentration in pore water (mg l1); D is the disper-
sion coefcient (cm2 min1); v is the velocity of pore water (cm min1); and S is
the sink term, which reects the removal rate in the biochar column. Presently,
there is still a paucity of studies being conducted to study the ltration of heavy
metals by biochars in xed-bed settings. Xue et al. (2012) reported effective
removal of Pb, Cd, Ni, and Cu by laboratory ltration columns packed with hydro-
chars. They found that the ADE model can accurately predict the ltration and
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

transport of four heavy metals both in single- and mixture-metal mode (Figure 2).

5. Conclusions and future direction


This review has presented the effect of feedstock on the removal of aqueous heavy
metals by biochar. Thermodynamic, kinetic, equilibrium parameters that often
describe and interpret sorption characteristics of biochar to heavy metals have also
been discussed in depth, along with governing sorption mechanisms. Predominant
heavy metal removal mechanisms vary for different biochars and metal contami-
nants. For instance, transition metals are commonly adsorbed by precipitation and
complexation mechanisms on alkaline biochars (plant or animal) produced above
300 C. However, nontransition metals can be adsorbed on lower treatment tem-
peratures biochars via ion exchange mechanisms. Most thermodynamic sorption
studies show that heavy metal removal by biochar is increasingly favorable in
endothermic systems. Thermodynamic model parameters, particularly free energy,
DG values can be used to differentiate between physical (DG , 20 to 0 KJ/mol)
or chemical sorption (DG , 400 to 80 KJ/mol) processes. Due to the paucity of
thermodynamic sorption studies on biochar, it is unclear whether these values can
be applied to all heavy metals. Mathematical models can accurately describe the
interaction of biochar with heavy metals. However, ltration studies for heavy
metal removal by biochars are lacking. Experimental and modeling studies on l-
tration of heavy metals in packed columns are highly recommended for future
studies.

Funding
This research was partially supported by the NSF through grants: CBET-1054405.

References
Ahmad, M., Lee, S. S., Oh, S. E., Mohan, D., Moon, D. H., Lee, Y. H., and Ok, Y. S. (2013). Modeling
adsorption kinetics of trichloroethylene onto biochars derived from soybean stover and peanut
shell wastes. Environmental Science and Pollution Research 20, 83648373.
428 M. I. INYANG ET AL.

Ahmad, M., Rajapaksha, A. U., Lim, J., Zhang, M., Bolan, N., Mohan, D., Vithanage, M., Lee, S.,
and Ok, Y. (2014). Biochar as a sorbent for contaminant management in soil and water: A
review. Chemosphere 99, 1933.
Ahmed, I. I., and Gupta, A. K. (2010). Pyrolysis and gasication of food waste: Syngas character-
istics and char gasication kinetics. Applied Energy 87, 101108.
Ai, L., Zhang, C., Liao, F., Wang, Y., Li, M., Meng, L., and Jiang, J. (2011). Removal of methylene
blue from aqueous solution with magnetite loaded multi-wall carbon nanotube: Kinetic, iso-
therm and mechanism analysis. Journal of Hazardous Materials 198, 282290.
Aksu, Z., and Isoglu, I. A. (2005). Removal of copper(II) ions from aqueous solution by biosorp-
tion onto agricultural waste sugar beet pulp. Process Biochemistry 40, 30313044.
Al-Muhtaseb, S. A., El-Naas, M. H., and Abdallah, S. (2008). Removal of aluminum from aque-
ous solutions by adsorption on date-pit and BDH activated carbons. Journal of Hazardous
Materials 158, 300307.
Al-Othman, Z. A., Ali, R., and Naushad, M. (2012). Hexavalent chromium removal from aque-
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

ous medium by activated carbon prepared from peanut shell: Adsorption kinetics, equilib-
rium and thermodynamic studies. Chemical Engineering Journal 184, 238247.
Arivoli, S., Hema, M., and Barathiraja, C. (2008). Comparative study on metal ions adsorption
on a low cost carbonaceous adsorbent kinetic equilibrium and mechanistic studies. Iranian
Journal of Environmental Health Science & Engineering 5, 110.
Beesley, L., Moreno-Jimenez, E., Gomez-Eyles, J. L., Harris, E., Robinson, B., and Sizmur, T.
(2011). A review of biochars potential role in the remediation, revegetation and restoration
of contaminated soils. Environmental Pollution 159, 32693282.
Bello, O. S., and Ahmad, M. A. (2011a). Removal of remazol brilliant violet-5R dye using peri-
winkle shells. Chemistry and Ecology 27.
Bello, O. S., and Ahmad, M. A. (2011b). Response surface modeling and optimization of rema-
zol brilliant blue reactive dye removal using periwinkle shell-based activated carbon. Separa-
tion Science and Technology 46, 23672379.
Bird, M. I., Wurster, C. M., de Paula Silva, P. H., Paul, N. A., and de Nys, R. (2012). Algal bio-
char: effects and applications. Global Change Biology Bioenergy 4, 6169.
Borchard, N., Wolf, A., Laabs, V., Aeckersberg, R., Scherer, H. W., Moeller, A., and Amelung,
W. (2012). Physical activation of biochar and its meaning for soil fertility and nutrient leach-
ing - a greenhouse experiment. Soil Use and Management 28, 177184.
Chen, X., Chen, G., Chen, L., Chen, Y., Lehmann, J., McBride, M. B., and Hay, A. G. (2011).
Adsorption of copper and zinc by biochars produced from pyrolysis of hardwood and corn
straw in aqueous solution. Bioresource Technology 102, 88778884.
Cao, X., and Harris, W. (2010). Properties of dairy-manure-derived biochar pertinent to its
potential use in remediation. Bioresource Technology 101, 52225228.
Cao, X., Ma, L., Gao, B., and Harris, W. (2009). Dairy-manure derived biochar effectively sorbs
lead and atrazine. Environmental Science & Technology 43, 32853291.
Cao, X., Ma, L., Liang, Y., Gao, B., and Harris, W. (2011). Simultaneous immobilization of lead
and atrazine in contaminated soils using dairy-manure biochar. Environmental Science &
Technology 45, 48844889.
Cheung, C. W., Porter, J. F., and McKay, G. (2000a). Sorption kinetics for the removal of copper
and zinc from efuents using bone char. Separation and Purication Technology 19, 5564.
Cheung, C. W., Porter, J. F., and McKay, G. (2000b). Elovich equation and modied second-
order equation for sorption of cadmium ions onto bone char. Journal of Chemical Technol-
ogy and Biotechnology 75, 963970.
Chien, S. H., and Clayton, W. R. (1980). Application of Elovich equation to the kinetics of phos-
phate release and sorption in soils. Soil Science Society of America Journal 44, 265268.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 429

Choy, K. K. H., and McKay, G. (2005). Sorption of cadmium, copper, and zinc ions onto bone
char using Crank diffusion model. Chemosphere 60, 11411150.
Choy, K. K. H., Ko, D. C. K., Cheung, C. W., Porter, J. F., and McKay, G. (2004). Film and intra-
particle mass transfer during the adsorption of metal ions onto bone char. Journal of Colloid
and Interface Science 271, 284295.
Crabtree, R. H. (2009). The organometallic chemistry of the transition metals. New York, NY:
Wiley.
Dimovic, S. D., Smiciklas, I. D., Sljivic-Ivanovic, M. Z., Plecas, I. B., and Slavkovic-Beskoski, L.
(2011). The effect of process parameters on kinetics and mechanisms of Co2C removal by
bone char. Journal of Environmental Science and Health Part a-Toxic/Hazardous Substances
& Environmental Engineering 46, 15581569.
Dizge, N., Aydiner, C., Demirbas, E., Kobya, M., and Kara, S. (2008). Adsorption of reactive
dyes from aqueous solutions by y ash: Kinetic and equilibrium studies. Journal of Hazard-
ous Materials 150, 737746.
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Dong, X. L., Ma, L. N. Q., and Li, Y. C. (2011). Characteristics and mechanisms of hexavalent chromium
removal by biochar from sugar beet tailing. Journal of Hazardous Materials 190, 909915.
Duku, M. H., Gu, S., and Hagan, E. B. (2011). Biochar production potential in GhanaA
review. Renewable and Sustainable Energy Reviews 15, 35393551.
El-Shafey, E. I. (2010). Removal of Zn(II) and Hg(II) from aqueous solution on a carbonaceous
sorbent chemically prepared from rice husk. Journal of Hazardous Materials 175, 319327.
El-Shafey, E. I., Cox, M., Pichugin, A. A., and Appleton, Q. (2002). Application of a carbon sor-
bent for the removal of cadmium and other heavy metal ions from aqueous solution. Journal
of Chemical Technology and Biotechnology 77, 429436.
Field, J. L., Keske, C. M. H., Birch, G. L., Defoort, M. W., and Cotrufo, M. F. (2013). Distributed
biochar and bioenergy coproduction: a regionally specic case study of environmental bene-
ts and economic impacts. Global Change Biology Bioenergy 5, 177191.
Foo, K. Y., and Hameed, B. H. (2010). Insights into the modeling of adsorption isotherm sys-
tems. Chemical Engineering Journal 156, 210.
Genc-Fuhrman, H., Bregnhoj, H., and McConchie, D. (2005). Arsenate removal from water
using sand-red mud columns. Water Research 39, 29442954.
Gerente, C., Lee, V. K. C., Le Cloirec, P., and McKay, G. (2007). Application of chitosan for the
removal of metals from wastewaters by adsorption - Mechanisms and models review. Critical
Reviews in Environmental Science and Technology 37, 41127.
Gil, M. V., Fermoso, J., Pevida, C., Pis, J. J., and Rubiera, F. (2010). Intrinsic char reactivity of
plastic waste (PET) during CO2 gasication. Fuel Processing Technology 91, 17761781.
Grierson, S., Strezov, V., and Shah, P. (2011). Properties of oil and char derived from slow pyrol-
ysis of Tetraselmis chui. Bioresource Technology 102, 82328240.
Gu, X. Y., and Wong, J. W. C. (2004). Identication of inhibitory substances affecting bioleach-
ing of heavy metals from anaerobically digested sewage sludge. Environmental Science &
Technology 38, 29342939.
Guo, M. X., Qiu, G. N., and Song, W. P. (2010). Poultry litter-based activated carbon for remov-
ing heavy metal ions in water. Waste Management 30, 308315.
Gupta, S. S., and Bhattacharyya, K. G. (2011). Kinetics of adsorption of metal ions on inorganic
materials: A review. Advances in Colloid and Interface Science 162, 3958.
Hanay, O., Hasar, H., Kocer, N. N., and Aslan, S. (2008). Evaluation for agricultural usage with
speciation of heavy metals in a municipal sewage sludge. Bulletin of Environmental Contami-
nation and Toxicology 81, 4246.
Harvey, O. R., Herbert, B. E., Rhue, R. D., and Kuo, L.-J. (2011). Metal interactions at the bio-
char-water interface: energetics and structure-sorption relationships elucidated by ow
adsorption microcalorimetry. Environmental Science & Technology 45, 55505556.
430 M. I. INYANG ET AL.

Ho, Y.-S. (2006). Review of second-order models for adsorption systems. Journal of Hazardous
Materials 136, 681689.
Ho, Y. S., and McKay, G. (1998). A comparison of chemisorption kinetic models applied to pollutant
removal on various sorbents. Process Safety and Environmental Protection 76, 332340.
Ho, Y. S., Ng, J. C. Y., and McKay, G. (2000). Kinetics of pollutant sorption by biosorbents:
Review. Separation and Purication Methods 29, 189232.
Holford, I. C. R., Wedderburn, R. W. M., and Mattingly, G. E. G. (1974). A Langmuir two-surface
equation as a model for phosphate adsorption by soils. Journal of Soil Science 25, 242255.
Hwang, I. H., Nakajima, D., Matsuto, T., and Sugimoto, T. (2008). Improving the quality of
waste-derived char by removing ash. Waste Management 28, 424434.
Inyang, M., Gao, B., Pullammanappallil, P., Ding, W., and Zimmerman, A. R. (2010). Biochar
from anaerobically digested sugarcane bagasse. Bioresource Technology 101, 88688872.
Inyang, M., Gao, B., Ding, W., Pullammanappallil, P., Zimmerman, A. R., and Cao, X. (2011).
Enhanced lead sorption by biochar derived from anaerobically digested sugarcane bagasse.
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Separation Science and Technology 46, 19501956.


Inyang, M., Gao, B., Yao, Y., Xue, Y., Zimmerman, A. R., Pullammanappallil, P., and Cao, X.
(2012). Removal of heavy metals from aqueous solution by biochars derived from anaerobi-
cally digested biomass. Bioresource Technology 110, 5056.
Inyang, M. D., Gao, B., Zimmerman, A., Zhou, Y., and Cao, X. (2014). Sorption and cosorption
of lead and sulfapyridine on carbon nanotube-modied biochars. Environmental Science and
Pollution Research 22, 18681876.
Jaradat, A. Q., Fowler, K., Grimberg, S. J., and Holsen, T. M. (2009). Transport of colloids and
associated hydrophobic organic chemicals through a natural media lter. Journal of Environ-
mental Engineering 135, 3645.
Jeppu, G. P., and Clement, T. P. (2012). A modied Langmuir-Freundlich isotherm model for simu-
lating pH-dependent adsorption effects. Journal of Contaminant Hydrology 129, 4653.
Kailash, D., Dharmendra, P., and Anil, V. (2010). Low cost adsorbents for heavy metal removal
from wastewater: a review. Research Journal of Chemistry and Environment 14, 100103.
Karakoyun, N., Kubilay, S., Aktas, N., Turhan, O., Kasimoglu, M., Yilmaz, S., and Sahiner, N.
(2011). Hydrogelbiochar composites for effective organic contaminant removal from aque-
ous media. Desalination 280, 319325.
Kasozi, G. N., Zimmerman, A. R., Nkedi-Kizza, P., and Gao, B. (2010). Catechol and humic acid
sorption onto a range of laboratory-produced black carbons (biochars). Environmental Sci-
ence & Technology 44, 61896195.
Keiluweit, M., and Kleber, M. (2009). Molecular-level interactions in soils and sediments: the
role of aromatic pi-systems. Environmental Science & Technology 43, 34213429.
Ko, D. C. K., Cheung, C. W., Choy, K. K. H., Porter, J. F., and McKay, G. (2004). Sorption equi-
libria of metal ions on bone char. Chemosphere 54, 273281.
Kong, H., He, J., Gao, Y., Wu, H., and Zhu, X. (2011). Cosorption of phenanthrene and mercury
(II) from aqueous solution by soybean stalk-based biochar. Journal of Agricultural and Food
Chemistry 59, 1211612123.
Kookana, R. S. (2010). The role of biochar in modifying the environmental fate, bioavailability, and
efcacy of pesticides in soils: a review. Australian Journal of Soil Research 48, 627637.
Koutcheiko, S., Monreal, C. M., Kodama, H., McCracken, T., and Kotlyar, L. (2007). Prepara-
tion and characterization of activated carbon derived from the thermo-chemical conversion
of chicken manure. Bioresource Technology 98, 24592464.
Krauskopf, B. K. (1967). Introduction to geochemistry. New York, NY: McGraw Hill.
Kumar, S., Loganathan, V. A., Gupta, R. B., and Barnett, M. O. (2011). An assessment of U(VI)
removal from groundwater using biochar produced from hydrothermal carbonization. Jour-
nal of Environmental Management 92, 25042512.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 431

Laird, D. A., Brown, R. C., Amonette, J. E., and Lehmann, J. (2009). Review of the pyrolysis platform
for coproducing bio-oil and biochar. Biofuels Bioproducts & Biorening 3, 547562.
Laird, D. A., Fleming, P., Davis, D. D., Horton, R., Wang, B., and Karlen, D. L. (2010). Impact of
biochar amendments on the quality of a typical Midwestern agricultural soil. Geoderma 158,
443449.
Lattao, C., Cao, X., Mao, J., Schmidt-Rohr, K., and Pignatello, J. J. (2014). Inuence of molecular
structure and adsorbent properties on sorption of organic compounds to a temperature
series of wood chars. Environmental Science & Technology 48, 47904798.
Lehmann, J., and Joseph, S. (Eds.). (2012). Biochar for environmental management: Science and
technology. New York, NY: Routledge.
Liao, R., Gao, B., and Fang, J. (2013). Invasive plants as feedstock for biochar and bioenergy pro-
duction. Bioresource Technology 140, 439442.
Li, L., and Zhang, H. X. (2004). Preparing levoglucosan derived from waste material by pyroly-
sis. Energy Sources 26, 10531059.
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Liu, Z., and Zhang, F. (2009). Removal of lead from water using biochars prepared from hydro-
thermal liquefaction of biomass. Journal of Hazardous Materials 167, 933939.
Liu, Z., Zhang, F.-S., and Wu, J. (2010). Characterization and application of chars produced
from pinewood pyrolysis and hydrothermal treatment. Fuel 89, 510514.
Lu, C. Y., Liu, C. T., and Su, F. S. (2009). Sorption kinetics, thermodynamics and competition of
Ni2C from aqueous solutions onto surface oxidized carbon nanotubes. Desalination 249,
1823.
Lu, H., Zhang, W., Yang, Y., Huang, X., Wang, S., and Qiu, R. (2012). Relative distribution of
Pb2C sorption mechanisms by sludge-derived biochar. Water Research 46, 854862.
Meng, J., Feng, X., Dai, Z., Liu, X., Wu, J., and Xu, J. (2014). Adsorption characteristics of Cu(II)
from aqueous solution onto biochar derived from swine manure. Environmental Science and
Pollution Research 21, 70357046.
Mohan, D., Kumar, H., Sarswat, A., Alexandre-Franco, M., and Pittman, C. U. Jr. (2014b). Cad-
mium and lead remediation using magnetic oak wood and oak bark fast pyrolysis bio-chars.
Chemical Engineering Journal 236, 513528.
Mohan, D., and Pittman, C. U. (2007). Arsenic removal from water/wastewater using adsorb-
ents: A critical review. Journal of Hazardous Materials 142, 153.
Mohan, D., Pittman, C. U. Jr., Bricka, M., Smith, F., Yancey, B., Mohammad, J., Steele, P. H.,
Alexandre-Franco, M. F., Gomez-Serrano, V., and Gong, H. (2007). Sorption of arsenic, cad-
mium, and lead by chars produced from fast pyrolysis of wood and bark during bio-oil pro-
duction. Journal of Colloid and Interface Science 310, 5773.
Mohan, D., Rajput, S., Singh, V. K., Steele, P. H., and Pittman, C. U. Jr. (2011). Modeling and
evaluation of chromium remediation from water using low cost bio-char, a green adsorbent.
Journal of Hazardous Materials 188, 319333.
Mohan, D., Sarswat, A., Ok, Y., and Pittman, C. U. Jr. (2014a). Organic and inorganic contami-
nants removal from water with biochar, a renewable, low cost and sustainable adsorbent: A
critical review. Bioresource Technology 160, 191202.
Mousavi, H. Z., Hosseynifar, A., Jahed, V., and Dehghani, S. A. M. (2010). Removal of lead from
aqueous solution using waste tire rubber ash as an adsorbent. Brazilian Journal of Chemical
Engineering 27, 7987.
Mukherjee, A., Zimmerman, A. R., and Harris, W. (2011). Surface chemistry variations among a
series of laboratory-produced biochars. Geoderma 163, 247255.
Ofomaja, A. E. (2008). Sorptive removal of Methylene blue from aqueous solution using palm
kernel bre: Effect of bre dose. Biochemical Engineering Journal 40, 818.
Ofomaja, A. E. (2010). Intraparticle diffusion process for lead(II) biosorption onto mansonia
wood sawdust. Bioresource Technology 101, 58685876.
432 M. I. INYANG ET AL.

Otero, M., Rozada, F., Moran, A., Calvo, L. F., and Garca, A. I. (2009). Removal of heavy metals
from aqueous solution by sewage sludge based sorbents: competitive effects. Desalination
239, 4657.
Pan, X., Wang, J., and Zhang, D. (2009). Sorption of cobalt to bone char: Kinetics, competitive
sorption and mechanism. Desalination 249, 609614.
Pellera, F. M., Giannis, A., Kalderis, D., Anastasiadou, K., Stegmann, R., Wang, J. Y., and Gidar-
akos, E. (2012). Adsorption of Cu(II) ions from aqueous solutions on biochars prepared
from agricultural by-products. Journal of Environmental Management 96, 3542.
Perez-Marin, A. B., Zapata, V. M., Ortuno, J. F., Aguilar, M., Saez, J., and Llorens, M. (2007).
Removal of cadmium from aqueous solutions by adsorption onto orange waste. Journal of
Hazardous Materials 139, 122131.
Phuengprasop, T., Sittiwong, J., and Unob, F. (2011). Removal of heavy metal ions by iron oxide
coated sewage sludge. Journal of Hazardous Materials 186, 502507.
Purevsuren, B., Avid, B., Gerelmaa, T., Davaajav, Y., Morgan, T. J., Herod, A. A., and Kandiyoti,
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

R. (2004). The characterisation of tar from the pyrolysis of animal bones. Fuel 83, 799805.
Qiu, H., Lv, L., Pan, B.-C., Zhang, Q.-J., Zhang, W.-M., and Zhang, Q.-X. (2009). Critical review
in adsorption kinetic models. Journal of Zhejiang University-Science A 10, 716724.
Qiu, Y. P., Cheng, H. Y., Xu, C., and Sheng, D. (2008). Surface characteristics of crop-residue-
derived black carbon and lead(II) adsorption. Water Research 42, 567574.
Quek, A., Zhao, X. S., and Balasubramanian, R. (2010). Mechanistic insights into copper
removal by pyrolytic tire char through equilibrium studies. Industrial & Engineering Chemis-
try Research 49, 45284534.
Rajapaksha, A. U., Vithanage, M., Zhang, M., Ahmad, M., Mohan, D., Chang, S. X., and Ok, Y.
S. (2014). Pyrolysis condition affected sulfamethazine sorption by tea waste biochars. Biore-
source Technology 166, 303308.
Ranjan, D., Talat, M., and Hasan, S. H. (2009). Biosorption of arsenic from aqueous solution
using agricultural residue rice polish. Journal of Hazardous Materials 166, 10501059.
Rhee, S.-W., and Park, H.-S. (2010). Effect of mixing ratio of woody waste and food waste on the charac-
teristics of carbonization residue. Journal of Material Cycles and Waste Management 12, 220226.
Ryu, C., Shari, V. N., and Swithenbank, J. (2007). Waste pyrolysis and generation of storable
char. International Journal of Energy Research 31, 177191.
Sag, Y., and Aktay, Y. (2000). Mass transfer and equilibrium studies for the sorption of chro-
mium ions onto chitin. Process Biochemistry 36, 157173.
Scheytt, T., Mersmann, P., Leidig, M., Pekdeger, A., and Heberer, T. (2004). Transport of phar-
maceutically active compounds in saturated laboratory columns. Ground Water 42, 767773.
Srivastava, N. K., and Majumder, C. B. (2008). Novel bioltration methods for the treatment of
heavy metals from industrial wastewater. Journal of Hazardous Materials 151, 18.
Sun, Y., Gao, B., Yao, Y., Fang, J., Zhang, M., Zhou, Y., Chen, H., and Yang, L. (2014). Effects of
feedstock type, production method, and pyrolysis temperature on biochar and hydrochar
properties. Chemical Engineering Journal 240, 574578.
Tong, X. J., Li, J. Y., Yuan, J. H., and Xu, R. K. (2011). Adsorption of Cu(II) by biochars gener-
ated from three crop straws. Chemical Engineering Journal 172, 828834.
Uchimiya, M., Chang, S., and Klasson, K. T. (2011). Screening biochars for heavy metal reten-
tion in soil: Role of oxygen functional groups. Journal of Hazardous Materials 190, 432441.
Uchimiya, M., Lima, I. M., Klasson, K. T., Chang, S., Wartelle, L. H., and Rodgers, J. E. (2010). Immo-
bilization of heavy metal ions (Cu(II), Cd(II), Ni(II), and Pb(II)) by broiler litter-derived biochars
in water and soil. Journal of Agricultural and Food Chemistry 58, 55385544.
Vijayaraghavan, K., Padmesh, T. V. N., Palanivelu, K., and Velan, M. (2006). Biosorption of
nickel(II) ions onto Sargassum wightii: Application of two-parameter and three-parameter
isotherm models. Journal of Hazardous Materials 133, 304308.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 433

Wang, S. S., Gao, B., Zimmerman, A. R., Li, Y. C., Ma, L. N., Harris, W. G., and Migliaccio, K.
W. (2015a). Physicochemical and sorptive properties of biochars derived from woody and
herbaceous biomass. Chemosphere 134, 257262.
Wang, S. S., Gao, B., Zimmerman, A. R., Li, Y. C., Ma, L., Harris, W. G., and Migliaccio, K. W.
(2015b). Removal of arsenic by magnetic biochar prepared from pinewood and natural
hematite. Bioresource Technology 175, 391395.
Wu, F. C., Tseng, R. L., and Juang, R. S. (2009a). Characteristics of Elovich equation used for the
analysis of adsorption kinetics in dye-chitosan systems. Chemical Engineering Journal 150,
366373.
Wu, F. C., Tseng, R. L., and Juang, R. S. (2009b). Initial behavior of intraparticle diffusion model used
in the description of adsorption kinetics. Chemical Engineering Journal 153, 18.
Xu, T., and Liu, X. Q. (2008). Peanut shell activated carbon: Characterization, surface modica-
tion and adsorption of Pb2C from aqueous solution. Chinese Journal of Chemical Engineer-
ing 16, 401406.
Downloaded by [GGS Indraprastha University] at 00:03 21 August 2017

Xue, Y. W., Gao, B., Yao, Y., Inyang, M., Zhang, M., Zimmerman, A. R., and Ro, K. S. (2012).
Hydrogen peroxide modication enhances the ability of biochar (hydrochar) produced from
hydrothermal carbonization of peanut hull to remove aqueous heavy metals: Batch and col-
umn tests. Chemical Engineering Journal 200, 673680.
Yao, Y., Gao, B., Inyang, M., Zimmerman, A. R., Cao, X. D., Pullammanappallil, P., and Yang, L. Y.
(2011a). Removal of phosphate from aqueous solution by biochar derived from anaerobically
digested sugar beet tailings. Journal of Hazardous Materials 190, 501507.
Yao, Y., Gao, B., Inyang, M., Zimmerman, A. R., Cao, X. D., Pullammanappallil, P., and Yang, L.
(2011b). Biochar derived from anaerobically digested sugar beet tailings: Characterization
and phosphate removal potential. Bioresource Technology 102, 62736278.
Yao, Y., Gao, B., Wu, F., Zhang, C.Z. and Yang, L.Y. (2015). Engineered Biochar from Biofuel
Residue: Characterization and Its Silver Removal Potential. ACS Applied Materials & Interfa-
ces 7, 1063410640.
Youssef, A. M., El-Nabarawy, T., and Samra, S. E. (2004). Sorption properties of chemically-
activated carbons 1. Sorption of cadmium(II) ions. Colloids and Surfaces A-Physicochemical
and Engineering Aspects 235.
Zuo, X.-J., Fu, D.-F., and Li, H. (2012). Adsorption removal of copper, zinc and cadmium in
aqueous solutions and road runoff by carbonized mulch: heavy metal removal by carbonized
mulch. Proceedings of the 2012 International Conference on Biomedical Engineering and Bio-
technology (iCBEB), 11801185.

You might also like