Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Toxicological & Environmental Chemistry

ISSN: 0277-2248 (Print) 1029-0486 (Online) Journal homepage: https://www.tandfonline.com/loi/gtec20

Metal organic frameworks as adsorbents for


dye adsorption: overview, prospects and future
challenges

Aderonke Ajibola Adeyemo , Idowu Olatunbosun Adeoye & Olugbenga


Solomon Bello

To cite this article: Aderonke Ajibola Adeyemo , Idowu Olatunbosun Adeoye & Olugbenga
Solomon Bello (2012) Metal organic frameworks as adsorbents for dye adsorption: overview,
prospects and future challenges, Toxicological & Environmental Chemistry, 94:10,
1846-1863, DOI: 10.1080/02772248.2012.744023

To link to this article: https://doi.org/10.1080/02772248.2012.744023

Published online: 22 Nov 2012.

Submit your article to this journal

Article views: 2301

View related articles

Citing articles: 68 View citing articles

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


https://www.tandfonline.com/action/journalInformation?journalCode=gtec20
Toxicological & Environmental Chemistry
Vol. 94, No. 10, December 2012, 1846–1863

Metal organic frameworks as adsorbents for dye adsorption: overview,


prospects and future challenges
Aderonke Ajibola Adeyemo, Idowu Olatunbosun Adeoye and Olugbenga
Solomon Bello*

Department of Pure and Applied Chemistry, Ladoke Akintola University of Technology,


P. M. B. 4000, Ogbomoso, Oyo State, Nigeria
(Received 23 May 2012; final version received 19 October 2012)

Considering the amount of colored waste water generated from many industries
(textile, leather, paper, printing, dyestuff, and plastic) that are sent to various
water bodies and the ecosystem, the search for efficient and better methods of
purification still continues. With the recent research into metal organic frame-
works (MOFs), there is a steady growing interest worldwide for their various
applications. This article presents a review of MOFs, their application in dye
adsorption and their various challenges and future prospects. It was concluded
that with the current interest, research and development for various applications,
there are possibilities which will bring to limelight more laboratory, industrial and
environmental usage of MOFs as dye adsorbents.
Keywords: metal organic frameworks; dyes; adsorption; wastewater; adsorbents

Introduction
Synthetic organic dyes have been widely used for dying textile fibers such as cotton and
polyesters (Tsai et al. 2007). Recently, considerable amount of waste water has been
generated from industries such as leather, paper, printing, dyestuff, plastic, and so on
(Haque, Khan, Park, et al. 2010). Colored dye waste water arises as a direct result of the
production of dye and also as a consequence of its use in the textile and other industries
(Robinson et al. 2001; Allen and Koumanova 2005). Dyes are chemicals which on binding
with a material gives color to the material. Dyes are ionic, aromatic organic compounds
with structures including aryl rings which have delocalized electron systems. The color of
the dye is provided by a chromophore group. A chromophore is a radical configuration
consisting of conjugated double bonds containing delocalized electron. Other common
chromophoric configurations include azo (–N¼N–); carbonyl (¼C¼O); carbon (¼C¼C¼);
carbon – nitrogen (4C¼NH or –CH¼N–); nitroso (–NO or N–OH); nitro (–NO2 or NO–
OH); and sulphur (C¼S). The chromogen, in which the aromatic structure normally
contains benzene, naphthalene or anthracene rings, is a part of chromogen – chromophore
structure along with an auxochrome. The presence of ionizing groups known as
auxochromes results in a much stronger alteration of the maximum absorption band of
the compound and provides bonding affinity. Some of the common auxochrome groups

*Corresponding author. Email: osbello@yahoo.com

ISSN 0277–2248 print/ISSN 1029–0486 online


ß 2012 Taylor & Francis
http://dx.doi.org/10.1080/02772248.2012.744023
http://www.tandfonline.com
Toxicological & Environmental Chemistry 1847

include –NH3, –COOH, –HSO3, –OH. There are more than 100,000 commercially
available dyes with over 7  105 tonnes produced annually (Allen and Koumanova 2005).
The rapid growth rate in the use of reactive dyes is due to increasing use of cellulosic
fibers and the technical and economic limitations of other dyes used for these fibers
(Bonneau 1995; O’Neill et al. 1999; Allen and Koumanova 2005). Textile factories make
use of various types of dyes depending on production orders; thus, the characteristics of
the textile wastewater are different for each factory. These wastewater contains a variety of
organic compounds and toxic substances, which are harmful to fish and other aquatic
organisms. Some are even considered to be carcinogenic; dyes may have acute and/or
chronic effects on exposed organisms depending on exposure time and concentration
(Pierce 1994; Inthorn et al. 2004). Dyed wastewater discharged to natural receiving waters
makes them unacceptable for public consumption (Inthorn et al. 2004). These chemical
materials pose certain health hazards and environmental pollution (Tsai et al. 2007).
Neglecting the aesthetic problem, the greatest environmental concern with dyes is their
absorption and reflection of sunlight entering the water which interferes with the growth of
bacteria to levels insufficient to biologically degrade impurities in water (Strickland and
Perkins 1995; Slokar and Le Marechal 1998; Allen and Koumanova 2005).

Methods of dye removal


The methods of dye removal from industrial effluents may be traditionally divided into
three main categories: biological, physical and chemical processes. Biological process is
used to treat domestic and industrial wastewater. This can be done in two ways: aerobic
and anaerobic processes. Chemical processes can also be used for dye removal in
wastewater treatment e.g. by precipitation with an alkaline material such as sodium or
calcium hydroxide or lime precipitation, destroying organic chemicals by oxidizing them
using ozone or hydrogen peroxide or adsorption process. Physical process usually treats
suspended rather than dissolved pollutants. It may be a passive process simply allowing
suspended pollutants to settle out to the top or bottom naturally, depending on whether
they are more or less dense than water. Chemical flocculants may also be added to produce
large particles. This process is known as flocculation. Filtration through a medium such as
sand at a final treatment stage can result in very clean water. Ultrafiltration,
nanofiltration, and reverse osmosis are processes which force water through membranes
thereby removing colloidal materials (very fine, electrically charged particles, which will
not settle) and even some dissolved matter (Tsai et al. 2007).
Among these methods, physical adsorption is generally considered to be the most
efficient method for quickly lowering the concentration of dissolved dyes in an effluent. In
this regard, activated carbon is the most widely used adsorbent for the removal of dyes
from aqueous solutions (Allen and Koumanova 2005; Tsai et al. 2007). The most common
porous organic material is activated carbon. These are usually prepared by pyrolysis of
carbon – rich materials which have high surface areas and high adsorption capacities, yet
do not possess ordered structures. Despite this lack of order, porous carbon materials have
many uses, such as separation and storage of gases, the purification of water and solvent
removal and recovery (Manocha 2003; Kuppler et al. 2009). Different physical forms of
activated carbon are produced depending on their applications: Granular activated carbon
is used in adsorption columns and powder activated carbon is used in batch adsorption
followed by filtration (Allen and Koumanova 2005; Tsai et al. 2007). Despite the prolific
use of this adsorbent throughout water/wastewater treatment and other industrial
1848 A.A. Adeyemo et al.

applications, the removal of organic pollutants by activated carbon adsorption remains an


expensive process because the adsorbent is still expensive and has high regeneration cost
when exhausted. For these reasons, there is growing interest in the use of low – cost
alternatives to carbon adsorbents (Tsai et al. 2007).

Adsorption as a dye removal method


Adsorption is the attachment of molecules to a surface; the surface is also known as
adsorbent. The adsorbent may be any material that provides surface area with
adsorption sites for attachment of molecules or particles. There are two types of
adsorbents: natural and synthetic adsorbents. Natural adsorbents are obtained from
plant and animal sources e.g. banana stalk (Bello, Ahmad, and Ahmad 2012), groundnut
hull (Din, Hameed, and Ahmad 2009; Bello, Fatona, et al. 2012), oil palm ash (Foo and
Hameed 2009a), cocoa pod husks (Bello and Ahmad 2011a), mango peels (Bello and
Ahmad 2011b), rice husk (Foo and Hameed 2009b; Yahaya et al. 2010a, 2010b, 2011a,
2011b), periwinkle shell (Bello, Adeogun, and Ajaelu 2008; Bello and Ahmad 2011c,
2011d), palm kernel shell (Zawani, Luqman, and Choong 2009), coconut shell (Bello and
Ahmad 2012b), Imperata cylindrica leaf (Bello and Semire 2012), rubber seed coat (Bello
and Ahmad 2012a) and so many others while the synthetic adsorbents are synthesized or
obtained through chemical reactions, e.g., zeolite (Baek, Kim, and Ihm 2004), activated
carbon (Choy, McKay, and Porter 1999; Mohammad-Khan and Ansari 2009), metal
organic frameworks (MOFs) (Haque, Khan, Park, et al. 2010; Haque, Jong, and Sung
2011; Chen et al. 2012; Huo and Yan 2012). The molecules or particles that itself attach
to the adsorbent are called adsorbate. The process of adsorption can remove dyes,
pigments, viruses, bacteria, colloidal particles and can also be used to control
biochemical oxygen demand.
Adsorption techniques for waste water treatment have become more popular in recent
years owing to their efficiency in the removal of pollutants too stable for biological
methods. Adsorption can produce high quality water while also being a process that is
economically feasible (Choy, McKay, and Porter 1999). There are two types of adsorption
mechanism: physical adsorption and chemical adsorption. Physical adsorption occurs
when weak interparticle bonds exist between the adsorbate and adsorbent. Examples of
such bonds are van der Waals, hydrogen and dipole – dipole. In majority of cases, physical
adsorption is easily reversible. Chemical adsorption occurs when strong interparticle
bonds are present between the adsorbate and the adsorbent due to an exchange of
electrons. Examples of such bonds are covalent and ionic bonds. Chemisorption is deemed
to be reversible in major cases (Ruthven 1984). Most adsorbents are highly porous
materials. The pores are generally very small; the internal surface area is orders of
magnitude greater than the external area. Separation occurs because of the difference in
molecular mass, shape, or polarity causing some molecules to be held more strongly on the
surface than others or because the pores are too small to admit the large molecules (Suzuki
1990; Allen and Koumanova 2005).

Metal organic frameworks


Metal organic frameworks are a new class of crystalline porous materials, the structure of
which is composed of metal – oxide units joined by organic linkers through strong covalent
bonds (Britt, Tranchemontagne, and Yaghi 2008). The interest in porous coordination
Toxicological & Environmental Chemistry 1849

polymers and MOFs started only around 1990. Hoskins and Robson set the basis for the
future of MOFs. In their paper, they already envisioned what has been subsequently
shown by many scientists around the world: the formation of a large range of crystalline,
microporous, stable solids, possibly using structure-directing agents, with ion-exchange,
gas sorption, or catalytic properties that further allow introducing functional groups by
post - synthetic modification (Robson 2008). The term MOF was popularized by Yaghi
and co-workers in 1995 for a layered Co-trimesate that showed reversible sorption
properties (Britt, Tranchemontagne, and Yaghi 2008; Stock and Biswas 2011). Design and
synthesis of functional materials of MOFs type has become a fascinating area of interest in
last years. The polymeric complexes consisting of transition metal ions and organic ligands
have attracted considerable attention because of their promising applications in various
fields like gas storage or catalysis (Rowsell et al. 2005). A MOF is composed of two major
components: a metal ion or cluster of ions and organic molecule called a linker. The
organic units are typically mono-, di-, tri-, or tetravalent ligands. MOF is a porous
material based on molecular open structure which have potential in a myriad of
applications in adsorption and chemical separation when compared with conventional
zeolites and carbon based adsorbent (Czaja, Natalia, and Ulrich 2009). It is applied in gas
storage, heterogeneous catalysis, optical, electronic, and magnetic materials (Qu et al.
2011). The flexibility with which these components can be varied has led to an extensive
class of MOF structures with ultra high surface areas, turnable pore size, and adjustable
internal surface properties, far exceeding those achieved for porous carbons (Britt,
Tranchemontagne, and Yaghi 2008; Chemical Review Editorial 2012).

MOFs as adsorbents for synthetic dye removal


The difficulty encountered in degrading dye materials because they are very stable to light
and oxidation reactions has formed the basis for the search of adsorbents suitable to
remove dyes from contaminated water (Chen et al. 2010). Removal of dye materials from
contaminated water is very important because water quality is adversely influenced by
colorants (Crini 2006). However, so far, there has been scanty report on the use of MOFs
in the removal of dye materials. In recent times, Haque and co-workers suggested the
potential applications of MOFs for the adsorptive removal of methyl orange (MO)
especially MOFs having a positive charge such as protonated ethylenediamine-grafted Cr-
terephthalate (MIL-101) which gave high adsorption capacity, rapid uptake and ready
regeneration for the MO (Haque, Khan, Park, et al. 2010). Iron terephthalate, an MOF
has also been used for the removal of harmful dyes (anionic dye MO and cationic dye
methylene blue (MB)) from contaminated water via adsorption. The adsorption capacities
of this MOF were much higher than the commercially available activated carbon (Figure
1). The performance of iron terephthalate having high adsorption capacity was remarkable
because the adsorption of MO and MB was as a result of specific electrostatic interactions
between the dyes and the adsorbent (MOF). Their study showed that adsorption of MO
and MB at various temperatures was a spontaneous and endothermic process and that the
entropy increased with adsorption of MO and MB. Thus, MOF-type materials were
suggested as potential adsorbents in removing harmful materials (dyes inclusive) in the
liquid phase (Haque, Jong, and Sung 2011).
Recently, adsorption of dyes from wastewater became a more prominent application in
research on transition MOFs. Haque and co-workers reported the results of
the adsorption of not only an anionic dye (MO) but also a cationic dye (MB) using
1850 A.A. Adeyemo et al.

Figure 1. Effect of contact time and initial MO concentration on the adsorption of MO over the five
adsorbents: (a) Ci ¼ 30 ppm; (b) Ci ¼ 40 ppm; (c) Ci ¼ 50 ppm (Haque, Lee, Jang, et al. 2010).

MOF-235 (Haque, Jong, and Sung 2011) because removal of both cationic and anionic
dyes are not readily achieved simultaneously (Rocher et al. 2008; Haque, Jong, and Sung
2011) and the adsorption can be understood with a comparison of the two adsorbates.
MOF-235, [Fe3O(terephthalate)3(DMF)3][FeCl4] (DMF stands for N,N-dimethylfor-
mide), was composed of non-toxic Fe3O clusters (Sudik, Cote, and Yaghi 2005). The
MOF-235 frameworks have positive charge (þ1 per formula unit), which is balanced with
Toxicological & Environmental Chemistry 1851

Table 1. The pseudo-second-order kinetic constants (k2) with correlation coefficients (R2) at
various initial MO and MB concentrations and the maximum adsorption capacities (Qo) of MOF-
235 and activated carbon (Haque, Jong, and Sung 2011).

Pseudo-second-order kinetics constants k2


(g mg1 min1)

20 ppm 30 ppm 40 ppm

Adsorption
capacity,
Adsorbents Dyes k2 R2 k2 R2 k2 R2 Qo (mg g1)

MOF-235 MO 7.67  104 0.999 8.79  104 0.998 9.10  104 0.998 477
MB 9.58  105 0.995 1.67  104 0.993 2.18  104 0.998 187
Activated MO 1.95  104 0.994 2.17  104 0.987 2.34  104 0.981 11.2
carbon
MB 9.71  105 0.997 1.14  104 0.979 1.48  104 0.972 26.0

[FeCl4] ion (Sudik, Cote, and Yaghi 2005; Haque, Jong, and Sung 2011). The adsorptive
removal of MO and MB was interesting because both anionic and cationic dyes are
adsorbed in liquid phase even though MOF-235 was regarded as a non-porous material as
nitrogen is hardly adsorbed over the MOF-235 at low temperature (Haque, Jong, and
Sung 2011). The scanning electron microscope image and nitrogen adsorption isotherm
showed that the MOF-235 has homogeneous morphology and adsorbs little nitrogen
compared with activated carbon (Haque et al. 2011). It was observed that the adsorbed
quantity of MO and MB over MOF-235 was much higher than those over an activated
carbon. The amount of adsorbed dyes slightly increased with the initial concentration,
showing the favorable adsorption at high dye concentration (Table 1). The adsorption
kinetic constants (k2) for MO adsorption over MOF-235 were larger than the constants
over activated carbon. However, the kinetic constants for MB over MOF-235 were nearly
similar to those over activated carbon. Interestingly, the kinetic constants for MO over
MOF-235 were smaller than those over MIL-101s (Haque, Lee, Jang, et al. 2010) even
though the adsorption capacities over MOF-235 were larger than those over MIL-101s.
The kinetic constants over MOF-235 and activated carbon increased slightly with
increasing initial dye concentrations, which showed rapid adsorption in the presence of
dyes at high concentration, similar to previous reports (Srivastava et al. 2006; Hameed and
Rahman 2008; Din, Hameed, and Ahmad 2009; Haque, Khan, Talapaneni, 2010; Haque,
Jong, and Sung 2011). The adsorption data were also analyzed using the pseudo-first-order
kinetic model and the kinetics results were in accordance with the results of pseudo-
second-order kinetic model. The adsorption isotherms were obtained after adsorption for
12 h. The amount of adsorbed dyes over MOF-235 was higher than that over activated
carbon for the experimental conditions, suggesting the efficiency of the MOF-235. The
adsorption capacity values for MO and MB were generally larger than those over other
adsorbents like activated carbon (Singh et al. 2003), cellulose-based wastes (Annadurai,
Juang, and Lee 2002) and magnetic cellulose bead (Luo and Jhang 2009).
To shed more light on dye adsorption over MOF-235; the adsorption free energy,
enthalpy, and entropy changes were calculated from the adsorption of MO and MB over
MOF-235 at various temperatures. The results obtained (Table 2) showed that the
adsorption capacity increased with increasing adsorption temperature, suggesting an
endothermic adsorption similar to previous results (Wang et al. 2005; Haque, Jong, and
1852 A.A. Adeyemo et al.

Table 2. The maximum adsorption capacities and thermodynamic parameters of MO and MB


adsorption over MOF-235 at different temperatures (Haque, Jong, and Sung 2011).

Temperature Qo DG DH DS
Dyes ( C) (mg g1) (kJ mol1) (kJ mol1) (J mol1 K1)

MO 25 477 31.1 99.6 441


35 448 35.9
45 501 41.0
MB 25 187 26.2 63.1 300
35 230 29.5
45 252 32.2

Sung 2011). It was quite remarkable that MOF-235 readily adsorbed MO and MB in
liquid phase because the MOF-235 has little porosity as confirmed with nitrogen
adsorption. Haque and co-workers concluded that representative dyes (anionic MO and
cationic MB) in contaminated water can be efficiently removed with a MOF-type material.
MOF-235, an iron terephthalate adsorbed very large amount of both MO and MB via a
specific interaction like electrostatic interaction between the dyes and the adsorbent. The
adsorption of MO or MB over MOF-235 at various temperatures showed that the
adsorption was spontaneous and endothermic and the randomness increased probably
with desorption of pre-adsorbed water. The driving force of MO or MB adsorption over
MOF-235 was mainly due to an entropy effect rather than an enthalpy change. It was
remarkable that the MOF-235 readily adsorbed both cationic and anionic dyes in liquid
phase even though the MOF-235 hardly adsorbed nitrogen (Haque, Jong, and Sung 2011).
Another MOF based on porous chromium – benzenedicarboxylates (Cr-BDCs) (MIL-
101) (MIL stands for Material of Institute Lavoisier) used for the adsorption of MO has
also been reported (Haque, Lee, Jang, et al. 2010). The adsorption kinetic constants over
activated carbon and Cr-BDCs were larger than those over cross-linked polymer (Huang
et al. 2008) and activated carbon (obtained from agricultural product, Phragmites
australis) (Chen et al. 2010). On the contrary the constants were smaller than the constants
over NHþ 3 -MCM-41, activated carbon (entrapping magnetic cellulose bead) (Luo and
Jhang 2009), lemon peel (Bhatnagar et al. 2009), and alginate bead (Rocher et al. 2008).
The kinetic constants over activated carbon and Cr-BDCs increased slightly with
increasing initial MO concentration, showing rapid adsorption in the presence of MO at
high concentration. The increase of the kinetic constant with increasing initial adsorbate
concentration has also been reported (Srivastava et al. 2006; Hameed and Rahman 2008;
Din, Hameed, and Ahmad 2009). The kinetic constants of adsorption over Cr-BDCs
showed that the adsorption over (protonated ethylenediamine-grafted MIL-101)
PED-MIL-101 is the fastest and that adsorption kinetics increased with modification of
virgin MIL-101. The kinetic constant over PED-MIL-101 was 10 times greater than that of
the activated carbon under the experimental conditions (Haque, Lee, Jang, et al. 2010).
The adsorption kinetic constants for MO adsorption were in the order of activated
carbon 5MIL-535MIL-1015ED-MIL-101 (ethylenediamine-grafted MIL-101)5PED-
MIL-101, similar to the adsorbed quantity (Table 3). Therefore, PED-MIL-101 was the
most effective adsorbent for MO removal in the viewpoint of adsorption amount and rate.
The fast adsorption of MO over MIL-101, compared with the adsorption over MIL-53,
was probably due to the large pore size of MIL-101 as the kinetic constant of adsorption
generally increases with the increasing pore size of a porous material not only in liquid-
phase adsorption (Jhung et al. 2007; Xu et al. 2008). This was also observed in gas phase
Table 3. Textural properties, the pseudo-second-order kinetic constants (k2) with correlation constants (R2) at various MO concentrations and the
maximum adsorption capacities (Qo) of the five adsorbents (Haque, Lee, Jang, et al. 2010).

Pseudo-second order kinetics constants k2 (g mg1 min1)

BET surface Total pore 30 ppm MO 40 ppm MO 50 ppm MO Maximum adsorption


Adsorbent area volume capacity, Qo
(pore size, nm) (m2 g1) (cm3 g1) k2 R2 k2 R2 k2 R2 (mg g1)

Activated carbon (51.0) 1068 0.50 2.17  104 0.987 2.34  104 0.981 2.59  104 0.978 11.2
MIL-53 (51.0) 1438 0.55 7.23  104 0.999 7.70  104 0.999 7.84  104 0.999 57.9
MIL-101 (1.6, 2.1) 3873 1.70 9.01  104 0.999 9.19  104 0.996 9.29  104 0.998 114
ED-MIL-101 (1.4, 1.8) 3491 1.37 1.06  103 0.998 1.19  103 0.999 1.27  103 0.999 160
PED-MIL-101 (1.4, 1.8) 3296 1.18 2.75  103 1.00 2.95  103 1.00 3.04  103 0.999 194
Toxicological & Environmental Chemistry
1853
1854 A.A. Adeyemo et al.

Table 4. The maximum adsorption capacities and thermodynamic parameters of MO adsorption


over PED-MIL-101 and MIL-101 at different temperatures (Haque, Lee, Jang, et al. 2010).

Temperature Qo DG DH DS
Adsorbent ( C) (mg g1) (kJ mol1) (kJ mol1) (J mol1 K1)

PED-MIL-101 25 194 32.5 29.5 209


35 219 34.5
45 241 36.6
MIL-101 25 114 31.9 4.00 120
35 121 33.1
45 140 34.3

adsorption (Jhung et al. 2006). However, the kinetic constants of the MIL-101s showed
that the pore size of the MIL-101s does not have any effect on the kinetics of adsorption,
suggesting the presence of a specific interaction between PED-MIL-101 or ED-MIL-101
and MO. So far many adsorbents have been evaluated as candidates for the removal of
MO from water and their adsorption capacities varied widely from 2.1 to about 366 mg g1
depending on the adsorbent (Annadurai, Juang, and Lee 2002; Singh et al. 2003; Crini
2006; Ayar, Gezici, and Kucukosmanoglu 2007; Ni et al. 2007; Huang et al. 2008; Rocher
et al. 2008; Bhatnagar et al. 2009; Qin, Ma, and Liu 2009; Chen et al. 2010).
Generally, the isotherms and Langmuir plots were similar to those of PED-MIL-101.
However, it should be mentioned that the adsorbed amount of MO over PED-MIL-101 was
high at very low concentration (especially at 45 C), which was very helpful in the removal of
MO even at a low concentration. Similar to PED-MIL-101, the adsorption of MO over
MIL-101 was a spontaneous process (negative DG). However, the absolute value of DH was
very small for the case of MIL-101 as seen in Table 4. This may be due to a physical
adsorption of MO and little desorption of pre-adsorbed water. The small free energy change
(DS) supports this assumption (small number of desorbed water molecules). The adsorption
of dye depends highly on the pH of the dye solution. MO adsorption at various pH values
were measured after equilibration with PED-MIL-101 and activated carbon. The amounts
adsorbed decreased with increasing pH of the MO solution, which was quite similar to
previously reported results (Singh et al. 2003). The decrease in the amount of adsorbed MO
with increasing pH was due to the fact that the concentration of positively charged
adsorbents decreased with increasing pH (Haque, Lee, Jang, et al. 2010).
It was concluded that the adsorption capacity and adsorption kinetic constant of MIL-
101 were greater than those of MIL-53, showing the importance of porosity and pore size
for adsorption. MIL-101, after being modified to have a positive charge, can be efficiently
used to remove MO via liquid phase adsorption. Based on the rate constants (pseudo-
second- or pseudo-first-order kinetics for adsorption) and adsorption capacity, it was
suggested that there was a specific interaction like electrostatic interaction between MO
and the adsorbent for rapid and high uptake of the dye. The adsorption of MO over PED-
MIL-101 at various temperatures showed that the adsorption was spontaneous and
endothermic and the randomness increased with the adsorption of MO. The driving force
of MO adsorption over PED-MIL-101 was mainly due to an entropy effect rather than an
enthalpy change (Haque, Lee, Jang, et al. 2010).
The adsorption of xylenol orange (XO) onto MIL-101(Cr) has been studied. The pH
value effect on the adsorption was carried out at 298 K. At a concentration of 400 ppm, the
pH value 2–12 of the solution was adjusted with NaOH or HCl solution. The amount
adsorbed decreased with increasing pH value of the XO solution, which was quite similar to
Toxicological & Environmental Chemistry 1855

previously reported results of MO adsorbed onto various adsorbents (Haque, Khan,


Talapaneni, et al. 2010). When the pH values were 4–10, the adsorbed amounts were much
closer. As the acidity became enhanced to pH value 2, the percentage removal of XO reached
almost 90%; however, in comparison, when the pH value increased to 12, the amount
adsorbed was zero. The Zeta potential increased with the increase of the pH value in the
acidic region, however, in the alkaline region the Zeta potential decreased with increase of
the pH value. The mechanism might be described as one in which a microstructure is
stripped from the structure of MIL-101, and the R  SO 3 group of XO were responsible for
the pH value effect on the adsorption of XO onto MIL-101. This explains the behavior of all
dye stuffs with a SO 3 group that are adsorbed onto MIL-101 (Chen et al. 2012).
The used adsorbent (MIL-101) was regenerated with 0.01 M NaOH solution.
Considering the patterns of powdered XRD, it was discovered that NaOH solution
destroyed the nature of the structure and decreased the specific surface, but the reused MIL-
101 still had the capacity to adsorb XO. Moreover, the performance of the adsorbent was
above 90% even when reused for the second time, showing an excellent capacity for
regeneration (Haque, Lee, Jang, et al. 2010; Chen et al. 2012). From these results, Chen et al.
(2012) concluded that pseudo-second-order kinetic model fitted much better with the
adsorption of XO onto MIL-101 compared with pseudo-first order kinetic model. The
Langmuir model fits the data better compared with the Freundlich model in terms of
regression coefficients. The thermodynamic parameters, including free energy, enthalpy,
and entropy of adsorption, were calculated from the result of isotherms, suggesting that the
adsorption of XO onto MIL-101 was a process with negative free energy change, negative
enthalpy change, and positive entropy change. Comparing the adsorption capacity of MIL-
101 to MCM-41 and activated carbon, the amount of adsorption of MCM-41 was lesser
than the other two adsorbents; activated carbon was only suitable for dye adsorption at low
concentration, whereas MIL-101 showed good capacity for dye adsorption over a wide
concentration range. The material can also be regenerated for reuse (Chen et al. 2012).
Huo and Yan reported the use of MOF MIL-100(Fe) for the adsorption of malachite
green (MG) from aqueous solution. The prepared MIL-100(Fe) gave a BET surface area
of 1626 m2 g1 with a pore volume of 0.79 cm3 g1. The pore size distribution of MIL-
100(Fe) estimated by using the Barrett–Joyner–Halenda method gave a pore diameter of
maxima at 31.6 Å, which was slightly larger than the large pore (29 Å) of MIL-100(Fe)
obtained from the crystal structure. The (thermogravimetric analysis) TGA data revealed
that the MIL-100(Fe) was stable up to 330 C (Horcajada et al. 2007; Huo and Yan 2012).
The solution pH may affect the surface charge of MIL-100(Fe), the stability and ionization
degree of MG in solution. The adsorption capacity of MIL-100(Fe) for MG increased as
the pH increased from 1 to 4, then leveled off in the pH range of 4–7, and decreased with
further increase of pH from 7 to 10. MG (pKa 10.3) was protonated in acidic medium and
deprotonated at higher pH (Mall et al. 2005; Crini et al. 2007). As the solution pH
increased from 1 to 5, the negative charges on the surface of MIL-100(Fe) increased, i.e.,
more negative zeta potential, which was favorable for the adsorption of positively charged
MG due to electrostatic attraction. A significant decrease of the adsorption capacity of
MIL-100(Fe) for MG in the pH range of 8–10 likely resulted from the instability of MG
and the decrease of negative charges on the surface of MIL-100(Fe) (less negative zeta
potential) (Huo and Yan 2012).
The effect of the concentration of NaCl and CaCl2 on the adsorption of MG on MIL-
100(Fe) showed that the adsorbed capacity of MIL-100(Fe) for MG suddenly dropped
once 0.01 mol L1 NaCl or CaCl2 was added. Furthermore, the divalent electrolyte (CaCl2)
had more negative effect on the adsorption of MG than the univalent electrolyte (NaCl).
1856 A.A. Adeyemo et al.

However, such a negative effect became negligible as the concentration of NaCl or CaCl2
further increased to 0.1 mol L1. The above results indicated that electrostatic interaction
was one of the mechanisms for the adsorption of MG on MIL-100(Fe) (Huo and Yan
2012). The values of KF (Freundlich adsorption constant) and adsorption capacity
increased with adsorption temperature, indicating that the adsorption of MG on MIL-
100(Fe) was endothermic. The values of 1/n (a measure of adsorption intensity) were less
than 1 (0.579–0.809) in the adsorption temperature range of 25–50 C, suggesting that
MIL-100(Fe) possessed heterogeneous surface caused by the presence of different
functional groups on the surface (Haghseresht and Lu 1998; Hameed and El-Khaiary
2008). The lack of plateau in desorption isotherms indicated multi-layer adsorption and
higher adsorption capacity (Huo and Yan 2012).
The values of ln Ko, DG , DH , and DS are shown in Table 5. Negative DG means
spontaneous adsorption of MG on MIL-100(Fe) in the temperature range studied. The
positive DH , which was supported by increased adsorption capacity of MG at high
temperatures, suggested that the adsorption of MG on MIL-100(Fe) was endothermic.
However, positive DS was favorable for spontaneous adsorption of MG on MIL-100(Fe),
which resulted from the release of more water molecules desolvated from one MG
molecule after adsorption on MIL-100(Fe). Therefore, the driving force for the adsorption
of MG on MIL-100(Fe) was controlled by an entropy effect rather than an enthalpy
change. The adsorption capacity of MG increased significantly in the first two hours and
reached equilibrium gradually. Moreover, the adsorption capacity significantly increased
as the initial concentration of MG increased, indicating the favorable adsorption at high
concentrations of MG. The time dependence for the adsorption of MG on MIL-100(Fe)
was well described by a versatile pseudo-second-order kinetic model which was based on
the adsorption capacity on the solid phase and was in agreement with a chemisorption
mechanism being the rate controlling step (Crini 2007; Huo and Yan 2012).
Moreover, the rate constant k2 decreased as the initial MG concentration increased,
indicating that chemisorption was significant in the rate-limiting step, involving valence
forces through sharing or exchange of electrons between MG and MIL-100(Fe) (Ho and
McKay 1999). This mechanism was supported by the XPS spectra (Figure 2) which
showed the shift of the N 1s peak of the MG and the Fe 2p peak in MIL-100(Fe) after the
adsorption of MG on MIL-100(Fe). The XPS spectra showed that the adsorption of MG
on MIL-100(Fe) led to a shift to higher energy for the N 1s peak of the MG (from 399.3 to
400.2 eV), a shift to lower energy for the Fe 2p peak in MIL-100(Fe) (from 725.3 and 711.7
to 724.9 and 711.2 eV, respectively). Although the open metal sites in MIL-100(Fe) have
been occupied by water molecules through the coordination in aqueous solution, the water
molecules can still be replaced by other stronger Lewis bases. Thus, the interaction
between the Lewis base –N(CH3)2 in MG and the Lewis acid Fe sites of MIL-100(Fe) can
occur due to the replacement of the water molecules by the Lewis base –N(CH3)2 of MG.
MG molecule cannot completely enter the pore of MIL-100(Fe) as the pore window is
smaller than the size of MG molecule. However, the MG molecule can partially enter the
pore of MIL-100(Fe) as the pore window is large enough for its –N(CH3)2 and –C6H5
groups. Besides, the possibility to establish – interaction between the benzene rings in
MG and MIL-100(Fe) may also play important roles (Huo and Yan 2012).
The adsorption isotherms of the MG on MIL-101(Cr), MIL-53(Al) and activated
carbon were also studied for comparison. MIL-100(Fe), MIL-101(Cr) and MIL-53(Al)
gave significantly larger specific surface area and total pore volume than activated carbon,
but only MIL-100(Fe) offered higher adsorption capacity for MG than activated carbon
as MG can be strongly adsorbed on the surface of MIL-100(Fe) via surface complexation
Table 5. Characteristic parameters for the adsorption of MG on MIL-100 (Fe) (Huo and Yan 2012).

Freundlich
constants
Temperature qexp DG DH DS
( C) (mg g1) 1/n KF R2 ln Ko (kJ mol1) (kJ mol1) (J mol1 K1)

25 205 0.768  0.052 6.49  0.47 0.994 0.93 2.31 14.5 57.0
30 266 0.580  0.057 12.6  4.65 0.993 1.21 3.05
35 325 0.579  0.033 20.6  3.38 0.977 1.24 3.18
40 361 0.609  0.038 24.7  3.88 0.963 1.33 3.48
45 475 0.805  0.031 26.4  2.28 0.982 1.37 3.64
50 485 0.809  0.031 33.5  2.88 0.976 1.44 3.87

Note: qexp: experimental adsorption capacity.


Toxicological & Environmental Chemistry
1857
1858 A.A. Adeyemo et al.

Figure 2. XPS spectra of: (a) N 1s in pure MG and adsorbed on MIL-100(Fe) and (b) Fe 2p in MIL-
100 and after adsorption of MG on surface of MIL-100(Fe) (Huo and Yan 2012).

and electrostatic interactions between cationic MG and negatively charged MIL-100(Fe).


The larger surface area of MIL-100(Fe) produced more active adsorption site than MIL-
53(Al) and activated carbon. Furthermore, MIL-100(Fe) provided water molecule
coordinated metal centers (Fe3þ) to interact with Lewis acid – base sites of MG, whereas
MIL-53(Al) had no such metal sites. The surface positive charge of MIL-101(Cr) below
pH 10 led to lower adsorption capacity for MG due to the electrostatic repulsion between
the surface of MIL-101(Cr) and MG though MIL-101(Cr) also has water molecule
coordinated metal sites and even larger pore and surface area than MIL-100(Fe).
Meanwhile, the adsorption capacity of MIL-100(Fe) was higher than MIL-53(Al),
indicating that the water molecule coordinated metal sites played an important role in
adsorption. Activated carbon has lower adsorption capacity than MIL-100(Fe) and it
reaches saturation capacity easily. However, saturation has not been reached in the
adsorption isotherm of MG on MIL-100(Fe), indicating that MIL-100 has the potential
for even much higher adsorption capacity. The adsorption capacity of MIL-100(Fe) for
MG was much higher than those of other adsorbents such as MIL-53(Al), activated
carbon, natural zeolite, and chitosan bead (Huo and Yan 2012).
Desorption and regeneration of MG-loaded MIL-100(Fe) were achieved using ethanol
solution (including 0.5% HCl) with a higher desorption efficiency of 97.3%. The
framework of MIL-100(Fe) remained stable after regeneration with the solutions of salt,
acid and ethanol under ultrasound assisted desorption. The good reusability of spent MIL-
100(Fe) also demonstrated the potential of MIL-100(Fe) for the adsorption and removal
of MG. Conclusively, they reported that the high adsorption capacity, good solvent
Toxicological & Environmental Chemistry 1859

stability and excellent reusability made MIL-100(Fe) promising as a novel adsorbent for
the adsorption and removal of MG from aqueous solution (Huo and Yan 2012).

Challenges and future prospects


Unfortunately, the lack of a high-throughput synthesis, i.e. a process whereby gram or
kilogram quantities can be synthesized in a matter of hours at ambient pressure, has thus
far been an impediment in the study of novel applications for MOFs. Formerly, many
MOFs were synthesized in water or through slow-diffusion methods. While these are
effective, these approaches have drawbacks. Many of the non-aqueous synthesis methods
involve slow-diffusion processes that take days or weeks to complete, which eliminates the
possibility of an industrially relevant process. Other alternatives involve the use of an
autoclave, which is not desirable from an industrial standpoint (Carson et al. 2009).
Another main drawback of using MOFs is the small pore size with diameters that
usually fall in the micropore range. This limits the amount of dyes that can be adsorbed in
the framework. Thus MOFs containing pores in the mesoporous range need to be
synthesized. However, the process of regeneration and reuse of the MOFs are of high
interest. Extensive research is needed to find better physical methods for regeneration for
the used adsorbents for them to be more economical.

Conclusion
From the above discussions on MOFs, it can be concluded that a lot of research has been
done on MOFs in the last two decades although the early studies were focused on the
structural diversities of MOFs and its application in gas storage. Extensive researches are
still ongoing about the various applications of MOFs but more priorities needs to be given
to MOF used as dye adsorbents. Availability of the starting materials and the desirability
of obtaining products with high yield and high purity should be considered when designing
new MOFs. Research into the workable synthetic methods of MOFs should be explored so
as to obtain low – cost MOFs which will be very useful as synthetic dye adsorbents. This
will serve as alternatives to the expensive commercially activated carbon. Furthermore, the
choice of ligands and metal salts must be considered in designing novel MOFs to be used
as dye adsorbents. Attention has to be given to the reaction conditions (temperature,
pressure, choice and amount of solvent, etc.) so as to obtain adjustable pore size, shape,
and better surface morphology in general. It is to be noted that novel MOFs used as
adsorbents are still being synthesized. With continual interest, research and development
now and in the near future for other applications, other than the earlier mentioned, there
are possibilities which will bring to limelight more laboratory, industrial, medicinal, and
environmental usage of MOFs.

References

Allen, S. J., and B. Koumanova. 2005. ‘‘Decolourisation of Water/Wastewater Using Adsorption


(Review).’’ Journal of the University of Chemical Technology and Metallurgy 40 (3): 175–92.
Annadurai, G., R. S. Juang, and D. J. Lee. 2002. ‘‘Use of Cellulose-based Wastes for Adsorption of
Dyes from Aqueous Solutions.’’ Journal of Hazardous Materials 92: 263–74.
Ayar, A., O. Gezici, and M. Kucukosmanoglu. 2007. ‘‘Adsorptive Removal of Methylene Blue and
Methyl Orange from Aqueous Media by Carboxylated Diaminoethane Sporopollenin: On the
1860 A.A. Adeyemo et al.

Usability of an Aminocarboxylic Acid Functionality-bearing Solid-stationary Phase in Column


Techniques.’’ Journal Hazardous Materials 146: 186–93.
Baek, S.-W., J.-R. Kim, and S.-K. Ihm. 2004. ‘‘Design of Dual Functional Adsorbent/Catalyst
System for the Control of VOC’s by Using Metal-loaded Hydrophobic Y-zeolites.’’ Catalysis
Today–Elsevier 93–95: 574–81.
Bello, O. S., and M. A. Ahmad. 2011a. ‘‘Adsorptive Removal of a Synthetic Textile Dye Using
Cocoa Pod Husks.’’ Toxicological and Environmental Chemistry 93 (7): 1298–308.
Bello, O. S., and M. A. Ahmad. 2011b. ‘‘Adsorption of Dyes from Aqueous Solution Using
Chemical Activated Mango Peels.’’ 2nd International Conference on Environmental Science and
Technology (ICEST) 2: 103–6.
Bello, O. S., and M. A. Ahmad. 2011c. ‘‘Response Surface Modelling and Optimization of Remazol
Brilliant Blue Reactive Dye Removal Using Periwinkle Shell Based Activated Carbon.’’
Separation Science and Technology 46: 2367–79.
Bello, O. S., and M. A. Ahmad. 2011d. ‘‘Removal of Remazol Brilliant Violet–5R Dye Using
Periwinkle Shell.’’ Chemistry and Ecology 27 (5): 481–92.
Bello, O. S., and M. A. Ahmad. 2012a. ‘‘Preparation and Characterisation of Activated
Carbon Derived from Rubber Seed Coat.’’ Bulgarian Journal of Science Education 21 (3):
389–95.
Bello, O. S., and M. A. Ahmad. 2012b. ‘‘Coconut (Cocos nucifera) Shell Based Activated Carbon for
the Removal of Malachite Green Dye from Aqueous Solutions.’’ Separation Science and
Technology 47 (6): 903–12.
Bello, O. S., and B. Semire. 2012. ‘‘Equilibrium, Kinetic and Quantum Chemical Studies on the
Adsorption of Congo Red Using Imperata cylindrica Leaf Powder Activated Carbon.’’
Toxicological and Environmental Chemistry 94 (6): 1114–24.
Bello, O. S., I. A. Adeogun, and J. C. Ajaelu. 2008. ‘‘Adsorption of Methylene Blue onto Activated
Carbon Derived from Periwinkle Shells: Kinetics and Equilibrium Studies.’’ Chemistry and
Ecology 24 (4): 285–95.
Bello, O. S., M. A. Ahmad, and N. Ahmad. 2012. ‘‘Adsorptive Features of Banana (Musa
paradisiaca) Stalk-based Activated Carbon for Malachite Green Dye Removal.’’ Chemistry and
Ecology 28 (2): 153–67.
Bello, O. S., T. A. Fatona, F. S. Falaye, O. M. Osuolale, and V. O. Njoku. 2012. ‘‘Adsorption of
Eosin Dye from Aqueous Solution Using Groundnut Hull-based Activated Carbon: Kinetic,
Equilibrium, and Thermodynamic Studies.’’ Environmental Engineering Science 29 (3): 186–94.
Bhatnagar, A., E. Kumar, A. K. Minocha, B. H. Jeon, H. Song, and Y. C. Seo. 2009. ‘‘Removal of
Anionic Dyes from Water Using Citrus limonum (lemon) Peel: Equilibrium Studies and Kinetic
Modelling.’’ Separation Science and Technology 44: 316–34.
Bonneau, M. C. 1995. ‘‘The Chemistry of Fabric Reactive Dyes.’’ Journal of Chemical Education 72:
724–5.
Britt, D., D. Tranchemontagne, and O. M. Yaghi. 2008. ‘‘Metal-organic Frameworks with High
Capacity and Selectivity for Harmful Gases.’’ Proceedings of the National Academy of Science
105 (33): 11623–7.
Carson, C. G., K. Hardcastle, J. Schwartz, X. Liu, C. Hoffmann, R. A. Gerhardt, and
R. Tannenbaum. 2009. ‘‘Synthesis and Structure Characterization of Copper Terephthalate
Metal–Organic Frameworks.’’ European Journal of Inorganic Chemistry 16: 2338–43.
Chemical Review Editorial. 2012. ‘‘Introduction to Metal-Organic Frameworks.’’ American
Chemical Society 112: 673–4.
Chen, C., M. Zhang, Q. Guan, and W. Li. 2012. ‘‘Kinetic and Thermodynamic Studies on the
Adsorption of Xylenol Orange onto MIL-101 (Cr).’’ Chemical Engineering Journal 183: 60–7.
Chen, S., J. Zhang, C. Zhang, Q. Yue, Y. Li, and C. Li. 2010. ‘‘Equilibrium and Kinetic Studies of
Methyl Orange and Methyl Violet Adsorption on Activated Carbon Derived from Phragmites
austries.’’ Desalination 252: 149–56.
Choy, K. K. H., G. McKay, and J. F. Porter. 1999. ‘‘Sorption of Acid Dyes from Effluents Using
Activated Carbon.’’ Resources Conservation and Recycling 27: 57–71.
Toxicological & Environmental Chemistry 1861

Crini, G. 2006. ‘‘Nonconventional Low-cost Adsorbents for Dye Removal: A Review.’’ Bioresource
Technology 97: 1061–85.
Crini, G., H. N. Peindy, F. Gimbert, and C. Robert. 2007. ‘‘Removal of C. I. Basic Green 4
(Malachite Green) from Aqueous Solutions by Adsorption Using Cyclodextrin-based Adsorbent:
Kinetic and Equilibrium Studies.’’ Separation and Purification Technology 53: 97–110.
Czaja, U. A., T. Natalia, and M. Ulrich. 2009. ‘‘Industrial Applications of Metal-organic
Frameworks.’’ Chemical Society Reviews 38 (5): 1284–93.
Din, A. T. M., B. H. Hameed, and A. L. Ahmad. 2009. ‘‘Batch Adsorption of Phenol onto
Physiochemical-activated Coconut Shell.’’ Journal of Hazardous Material 161: 1522–9.
Foo, K. Y., and B. H. Hameed. 2009a. ‘‘Value-added Utilization of Oil Palm Ash: A Superior
Recycling of the Industrial Agricultural Waste.’’ Journal of Hazardous Materials 172: 523–31.
Foo, K. Y., and B. H. Hameed. 2009b. ‘‘Utilization of Rice Husk Ash as Novel Adsorbent:
A Judicious Recycling of the Colloidal Agricultural Waste.’’ Advances in Colloid and Interface
Science 152: 39–47.
Haghseresht, F., and G. Q. Lu. 1998. ‘‘Adsorption Characteristics of Phenolic Compounds onto
Coal-reject-derived Adsorbents.’’ Energy Fuels 12 (6): 1100–7.
Hameed, B. H., and M. I. El-Khaiary. 2008. ‘‘Equilibrium, Kinetics and Mechanism of Malachite
Green Adsorption on Activated Carbon Prepared from Bamboo by K2CO3 Activation and
Subsequent Gasification with CO2.’’ Journal of Hazardous Materials 157: 344–51.
Hameed, B. H., and A. A. Rahman. 2008. ‘‘Removal of Phenol from Aqueous Solutions by
Adsorption onto Activated Carbon Prepared from Biomass Material.’’ Journal of Hazardous
Materials 160: 576–81.
Haque, E., W. J. Jong, and H. J. Sung. 2011. ‘‘Adsorptive Removal of Methyl Orange and
Methylene Blue from Aqueous Solution with a Metal Organic Framework Material, Iron
Terephthalate (MOF–235).’’ Journal of Hazardous materials 185: 507–11.
Haque, E., N. A. Khan, J. H. Park, and S. H. Jhung. 2010. ‘‘Synthesis of a Metal-organic
Framework Material, Iron Terephthalate, by Ultrasound, Microwave, and Conventional Electric
Heating: A Kinetic Study.’’ Chemistry – European Journal 16: 1046–52.
Haque, E., N. A. Khan, S. N. Talapaneni, A. Vinu, J. Jegal, and S. H. Jhung. 2010. ‘‘Adsorption of
Phenol on Mesoporous Carbon CMK-3: Effect of Textural Properties.’’ Bulletin of Korean
Chemical Society 31: 1638–42.
Haque, E., J. E. Lee, I. N. Jang, Y. K. Hwang, J. S. Chang, J. Jegal, and S. H. Jhung. 2010.
‘‘Adsorptive Removal of Methyl Orange from Aqueous Solution with Metal-organic Frameworks,
Porous Chromium-benzenedicarboxylates.’’ Journal of Hazardous Materials 181: 535–42.
Ho, Y. S., and G. McKay. 1999. ‘‘Pseudo-second Order Model for Sorption Processes.’’ Process
Biochemistry 34: 451–65.
Horcajada, P., S. Surble, C. Serre, D. Y. Hong, Y. K. Seo, J. S. Chang, J. M. Greneche,
I. Margiolaki, and G. Ferey. 2007. ‘‘Synthesis and Catalytic Properties of MIL-100(Fe), an
Iron(III) Carboxylate with Large Pores.’’ Chemical Communications 27: 2820–2.
Huang, J. H., K. L. Huang, S. Q. Liu, A. T. Wang, and C. Yan. 2008. ‘‘Adsorption of Rhodamine B
and Methyl Orange on a Hypercrosslinked Polymeric Adsorbent in Aqueous Solution, Colloids
and Surfaces A: Physicochemical and Engineering.’’ Aspects 330: 55–61.
Huo, S.-H., and X.-P. Yan. 2012. ‘‘Metal–Organic Framework MIL-100(Fe) for the Adsorption of
Malachite Green from Aqueous Solution.’’ Journal of Material Chemistry 22: 7449–54.
Inthorn, D., S. Singhtho, P. Thiravetyan, and E. Khan. 2004. ‘‘Decolourisation of Basic, Direct and
Reactive Dyes by Pre-treated Narrow-leaved Cattail (Typha angustifolia Linn.).’’ Bioresource
Technology 94: 299–306.
Jhung, S. H., H. K. Kim, J. W. Yoon, and J. S. Chang. 2006. ‘‘Low-temperature Adsorption of
Hydrogen on Nanoporous Aluminophosphates: Effect of Pore Size.’’ Journal of Physical
Chemistry B 110: 9371–4.
Jhung, S. H., J. H. Lee, J. W. Yoon, C. Serre, G. Ferey, and J. S. Chang. 2007. ‘‘Microwave
Synthesis of Chromium Terephthalate MIL-101 and its Benzene Sorption Ability.’’ Advanced
Materials 19: 121–4.
1862 A.A. Adeyemo et al.

Kuppler, R. J., D. J. Timmons, Q.-R. Fang, J.-R. Li, T. A. Makal, M. D. Young, D. Yuan, D. Zhao,
W. Zhuang, and H.-C. Zhou. 2009. ‘‘Potential Applications of Metal-organic Frameworks.’’
Coordination Chemistry Reviews 253: 3042–66.
Luo, X., and L. Jhang. 2009. ‘‘High Effective Adsorption of Organic Dyes on Magnetic Cellulose
Beads Entrapping Activated Carbon.’’ Journal of Hazardous Materials 171: 340–7.
Mall, I. D., V. C. Srivastava, N. K. Agarwal, and I. M. Mishra. 2005. ‘‘Adsorptive Removal of
Malachite Green Dye from Aqueous Solution by Bagasse Fly Ash and Activated Carbon-Kinetic
Study and Equilibrium Isotherm Analyses.’’ Colloids and Surfaces, A Physicochemical and
Engineering aspect 264: 17–28.
Manocha, S. M. 2003. ‘‘Porous Carbons.’’ Sadhana 28: 335–48.
Mohammad-Khan, A., and R. Ansari. 2009. ‘‘Activated Charcoal: Preparation, Characterization
and Applications: A Review Article.’’ International Journal of ChemTech Research 1 (4): 859–64.
Ni, Z. M., S. J. Xia, L. G. Wang, F. F. Xing, and G. X. Pan. 2007. ‘‘Treatment of Methyl Orange by
Calcined Layered Double Hydroxides in Aqueous Solution: Adsorption Property and Kinetic
Studies.’’ Journal of Colloid and Interface Science 316: 284–91.
O’Neill, C., F. R. Hawkes, D. L. Hawkes, N. D. Lourenco, H. M. Pinheiro, and W. J Delee. 1999.
‘‘Colour in Textile Effluents–Sources, Measurement, Discharge Consents and Simulation: A
Review.’’ Journal of Chemical Technology and Biotechnology 74 (11): 1009–18.
Pierce, J. 1994. ‘‘Colour in Textile Effluents: The Origins of Problem.’’ Journal of the Society of
Dyers and Colourists 110: 131–3.
Qin, Q., J. Ma, and K. Liu. 2009. ‘‘Adsorption of Anionic Dyes on Ammonium-functionalized
MCM-41.’’ Journal of Hazardous Materials 162: 133–9.
Qu, H., L. Qiu, X.-K. Leng, M.-M. Wang, S.-M. Lan, L.-L. Wen, and D.-F. Li. 2011. ‘‘Structures
and Photocatalytic Activities of Metal-organic Frameworks Derived from Rigid Aromatic
Dicarboxylate Acids and Flexible Imidazole-based Linkers.’’ Inorganic Chemistry Communications
14: 1347–51.
Robinson, T., G. McMullan, R. Marchant, and P. Nigam. 2001. ‘‘Remediation of Dyes in Textile
Effluent: A Critical Review on Current Treatment Technologies with a Proposed Alternative.’’
Bioresource Technology 77 (3): 247–55.
Robson, R. 2008. ‘‘Design and its Limitations in the Construction of Bi- and Poly-nuclear
Coordination Complexes and Coordination Polymers (aka MOFs): A Personal View.’’ Dalton
Transactions 38: 5113–31.
Rocher, V., J. M. Siaugue, V. Cabuil, and A. Bee. 2008. ‘‘Removal of Organic Dyes by Magnetic
Alginate Beads.’’ Water Research 42: 1290–8.
Rowsell, J. L. C., E. C. Spencer, J. Eckert, J. A. K. Howard, and O. M. Yaghi. 2005. ‘‘Gas
Adsorption Sites in a Large-Pore Metal-organic Framework.’’ Science 309: 1350–4.
Ruthven, D. M. 1984. Principles of Adsorption and Adsorption Processes. New York: Wiley Inter
Science.
Singh, K. P., D. Mohan, S. Sinha, G. S. Tondon, and D. Gosh. 2003. ‘‘Colour Removal from
Wastewater Using Low-cost Activated Carbon Derived from Agricultural Waste Material.’’
Industrial and Engineering Chemistry Research 42: 1965–76.
Slokar, Y. M., and M. Le Marechal. 1998. ‘‘Methods of Decolouration of Textile Wastewaters.’’
Dyes Pigments 37: 335–56.
Srivastava, V. C., M. M. Swamy, I. D. Mall, B. Prasad, and I. M. Mishra. 2006. ‘‘Adsorptive
Removal of Phenol by Bagasse Fly Ash and Activated Carbon: Equilibrium, Kinetics and
Thermodynamics.’’ Colloids and Surfaces A: physicochemical and Engineering Aspects 272:
89–104.
Stock, N., and S. Biswas. 2011. ‘‘Synthesis of Metal-organic Frameworks (MOFs): Routes to
Various MOF Topologies, Morphologies, and Composites.’’ American Chemical Society -
Chemical Reviews 112: 933–69.
Strickland, A. F., and W. C. Perkins. 1995. ‘‘Decolourisation of Continuous Dyeing of Wastewater
by Ozonation.’’ Textile Chemist and Colourist 27 (5): 11–5.
Toxicological & Environmental Chemistry 1863

Sudik, A. C., A. P. Cote, and O. M. Yaghi. 2005. ‘‘Metal-organic Frameworks Based on Trigonal
Prismatic Building Blocks and the New ‘‘ACS’’ Topology. 2005.’’ Inorganic Chemistry 44:
2998–3000.
Suzuki, M. 1990. Adsorption Engineering. Amsterdam: Elsevier.
Tsai, W.-T., T.-Y. Su, K.-Y. Lin, C.-M. Lin, and T.-H. Dai. 2007. ‘‘The Adsorption of Cationic Dye
from Aqueous Solution onto Acid-activated Andesite.’’ Journal of Hazardous Materials 147 (3):
1056–62.
Wang, S., L. Li, H. Wu, and Z. H. Zhu. 2005. ‘‘Unburned Carbon as a Low-cost Adsorbent for
Treatment of Methylene Blue-containing Wastewater.’’ Journal of Colloid and Interface Science
292: 336–43.
Xu, D. P., S. H. Yoon, I. Mochida, W. M. Qiao, Y. G. Wang, and L. C. Ling. 2008. ‘‘Synthesis of
Mesoporous Carbon and its Adsorption Property to Biomolecules.’’ Microporous and Mesoporous
Materials 115: 461–8.
Yahaya, N. E. M., M. F. Pakir, M. Latiff, I. Abustan, O. S. Bello, and M. A. Ahmad. 2010a.
‘‘Process Optimisation for Zn(II) Removal by Activated Carbon Prepared from Rice Husk Using
Chemical Activation.’’ International Journal of Engineering and Technology 10 (6): 132–6.
Yahaya, N. E. M., M. F. Pakir, M. Latiff, I. Abustan, O. S. Bello, and M. A. Ahmad. 2010b. ‘‘Effect
of Preparation Conditions of Activated Carbon Prepared from Rice Husk by CO2 Activation for
Removal of Cu(II) from Aqueous Solution.’’ International Journal of Engineering and Technology
10 (6): 47–51.
Yahaya, N. E. M., M. F. Pakir, M. Latiff, I. Abustan, O. S. Bello, and M. A. Ahmad. 2011a.
‘‘Adsorptive Removal of Cu(II) Using Activated Carbon Prepared from Rice Husk by ZnCl2
Activation and Subsequent Gasification with CO2.’’ International Journal of Engineering and
Technology 11 (1): 207–11.
Yahaya, N. E. M., M. F. Pakir, M. Latiff, I. Abustan, O. S. Bello, and M. A. Ahmad. 2011b. ‘‘Fixed
Bed Column Study for Cu(II) Removal from Aqueous Solutions Using Rice Husk Based
Activated Carbon.’’ International Journal of Engineering and Technology 11 (1): 248–52.
Zawani, Z., C. A. Luqman, and T. S. Y. Choong. 2009. ‘‘Equilibrium, Kinetics and Thermodynamic
Studies: Adsorption of Remazol Black 5 on the Palm Kernel Shell Activated Carbon (PKS-AC).’’
European Journal of Scientific Research 37 (1): 63–71.

You might also like