You are on page 1of 36

9

Metal–Organic Frameworks
for Heavy Metal Removal
Anam Asghar1*, Mustapha Mohammed Bello2 and Abdul Aziz Abdul Raman3
1
Instrumental Analytical Chemistry, Faculty of Chemistry, University Duisburg-Essen,
Universitaetsstraße 5, Essen, Germany
2
Centre for Dryland Agriculture, Bayero University, Kano, Nigeria
3
Department of Chemical Engineering, Faculty of Engineering,
University of Malaya, Kuala Lumpur, Malaysia

Abstract
Heavy metals are among the major sources of environmental pollution, where they
pose threat to human health and the environment. Thus, there is increasing effort
to develop effective technology for heavy metals removal. Adsorption is considered
one of the effective process for heavy metal removal. However, the development of
effective adsorbents is still a major challenge. Although metal-organic frameworks
have potentials for heavy metals adsorption due to their excellent properties, their
application is still not fully established. Herein, we present the use of metal-organic
frameworks as adsorbents for heavy metals removal. We discussed the problem of
heavy metals in the environment and their common removal technologies, high-
lighting adsorption as an effective technology. We then presented the development
of metal-organic frameworks and their applications in removing specific heavy
metals. The effectiveness of metal-organic frameworks is largely attributed to their
ultrahigh porosity (around 90% free volume), specific surface areas (often exceed-
ing 7000 m2/g) and tunability. Their tunability permits their easy modification
to introduce specific functional groups for the target metal ions. Although pris-
tine metal-organic frameworks have been studied, we noted a growing interest in
using metal-organic framework-based composites and metal-organic framework
derived advanced materials for heavy metals removal. Materials such as graphene
oxide, carbon nanotubes, metal nanoparticles and conducting polymers have been
combined with metal-organic frameworks, forming composites with synergistic
adsorption capacity for heavy metals removal. To effectively utilize the potentials of

*Corresponding author: chem.uet@hotmail.com; anam.asghar@gast.uni-due.de

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Tauseef Ahmad Rangreez (eds.) Applied
Water Science Volume 2: Remediation Technologies, (321–356) © 2021 Scrivener Publishing LLC

321
322  Applied Water Science Volume 2

metal-organic frameworks for heavy metals removal, their synthesis methods and
stability in aqueous solutions need to be enhanced.

Keywords:  Adsorption, advanced materials, composites, environment, heavy


metals, metal–organic frameworks, nanoparticles, public health

9.1 Introduction
The growing trend of industrialization and economic globalization has
placed an increasing pressure on the environment. Wastewaters discharged
from several industries such as coal-fired power plants [1, 2] and mining
[3], along with agricultural and solid waste disposal activities [4, 5] are
major contributors to environmental pollution. These wastewaters contain
toxic organic and inorganic compounds that are often recalcitrant [6–12].
The situation is even worse in developing regions of the world, where
wastewater treatment technologies are inefficient [13]. Estimates indicate
that 70–80% of the problems in developing regions are linked with water
pollution [14]. According to the World Health Organisation (WHO), ~844
million of the population have no access to clean drinking water [15, 16].
This number is expected to increase to around 3.9 billion by 2030 based on
an estimate by the World Water Council [17].
According to the United Nations, ~80% of all industrial and munici-
pal wastewater in developing countries need end-of-pipe treatment [18].
The proliferation of heavy metals in fresh water is of particular concern
due to their health effects on humans [19]. Heavy metals are stable and
non-degradable, thus accumulating in the environment [20]. In biologi-
cal systems, the toxicity level of heavy metals is due to their interactions
with proteins to form complexes with proteins [21]. Heavy metals could
also affect cellular organelles and other components such as cell mem-
brane [22]. Furthermore, heavy metals may interact deoxyribonucleic acid
(DNA), causing structural changes and cell cycle modulations [22, 23]. In
this regard, many heavy metals have been ranked among priority metals
and “probable” human carcinogens [24].
Figure 9.1 summarizes the impacts of heavy metals on human health.
For instance, intake of water having arsenic concentration greater than
0.01 ppm can cause skin, circulatory and nervous system problems [21].
Lead can impair liver, kidney, reproductive and nervous system [25,
26]. Cadmium accumulates in human organs like livers, kidneys, pan-
creas and bones, causing anaemia, hypertension and secretion disor-
der [27]. Chromium is classified as carcinogenic and its exposure can
Metal–Organic Frameworks for Heavy Metal Removal  323

Skin Bones
Lungs Liver
Brain Kidneys
Kidneys Lungs
Liver Testes
Metabolic Brain
Cardiovascular Immunological
Immunological Cardiovascular
Endocrine

As Cd
Bones
Skin Liver
Lungs Kidneys
Kidneys Brain
Liver Heavy Lungs
Brain Cr Pb Spleen
Pancreas metals Immunological
Testes Haematological
Gastrointestinal Cardiovascular
Reproductive Reproductive
Hg Cu

Brain Liver
Lungs Brain
Kidneys Kidneys
Liver Cornea
Immunological Gastrointestinal
Cardiovascular Lungs
Endocrine Immunological
Reproductive Haematological

Figure 9.1  Impact of some heavy metals on human health. Heavy metals have been
reported to affect the major body organs and systems of humans (Reproduced from
García-Niño & Pedraza-Chaverrí (2014) with permission from Elsevier [32]).

cause asthma attacks, nasal ulcer, anaemia, eye and respiratory irritation
[28]. High concentrations of copper have detrimental effects on human
health and its exposure has been linked to neurodegenerative diseases.
Consequently, the WHO has recommended 2 mg/L as the maximum
permissible limit of copper ions in drinking water [15]. Mercury, known
for its high toxicity, induces severe alteration in body tissues [29] and
may impair the central nervous system [30].
Considering the detrimental effects of heavy metals, the WHO has set
maximum contamination levels (MCLs) to ensure threshold concentration
or zero discharge of heavy metals [31]. For instance, expect for specific
applications, lead and mercury are banned. At governmental levels, MCLs
of heavy metals have also been set to minimize the effects of these toxic ele-
ments on the ecosystem. The regulation of metal concentration in drinking
water has become a priority for many countries, with some even extending
the regulation to irrigation systems [33, 34].

9.2 Heavy Metals in Environment


Heavy metals are produced from natural and anthropogenic activities and
their distribution in the environment depends not only on the proximity
324  Applied Water Science Volume 2

to emission sources but also on the distribution media [34]. The natural
sources of heavy metals include rock weathering, erosion and soil for-
mation. Heavy metals such as copper, nickel, cobalt and manganese are
found in rock forming minerals [35]. The concentration and nature of
heavy metal in the soil depend on the soil origin and its interaction with
the fluid medium [36]. Among the natural sources, surface waters are
the main carriers of heavy metals. Therefore, their compositions depend
on several factors such as pH, redox potential, temperature, biological,
physical and chemical factors [34].
In addition to natural processes, anthropogenic activities such as indus-
trial, mining, metallurgical, landfills, sewage systems and agricultural pro-
cesses introduce heavy metals into the environment [34, 35]. Heavy metals
from both diffuse and point sources are introduced into the environment
through aqueous, gaseous and solid phases [37]. These elements are ulti-
mately introduced into groundwater reservoirs through natural water
cycle and pose risk to both human and aquatic life [20]. Figure 9.2 depicts
a schematic representation of heavy metals transportation routes in the
environment [38].
Among the anthropogenic sources, industrial processes repre-
sent the main sources of heavy metals pollution. The waste streams

Heavy metal sources and transportation routes within environment.


Surface

Weathering Bioaccumulation
and erosion
Anthropogenic
sources

Large immobile Plant


crystals uptake

Invertebrate
Mineral phase Input and fish Output

Discharge and leaching


Sorption

Phytoplankton
Soil

Metal ion aggregation

Benthic
organisms

Desorption Leaching

Sediment

Groundwater

Figure 9.2  An overview showing redistribution of heavy metals in groundwater


(Reproduced from Kobielska et al. [38] with permission from Elsevier).
Metal–Organic Frameworks for Heavy Metal Removal  325

produced from industrial operations such as mining, coal combustion


and oil refining contain high concentrations of heavy metals, which
may end up in the environment [39]. Other anthropogenic activities
that significantly contribute to heavy metal pollution include agricul-
tural use of insecticides, fertilizers and pesticides. This is of particular
concern especially in developing regions that are yet to introduce and
implement current legislation [33, 34].

9.3 Heavy Metals Removal Technologies


Heavy metals are non-biodegradable and therefore persist in the environ-
ment, posing threat to public health and the environment. It is therefore
necessary to have effective technologies for heavy metals removal. Over the
years, different types of technologies have been developed for heavy metal
removal. These technologies can be broadly grouped into physicochemical
and biological processes as shown in Figure 9.3.
In physicochemical technologies, physical and chemical processes are
used to remove heavy metals from aqueous solution. Typical physicochem-
ical processes of heavy metal removal, adsorption, floatation, ion exchange,
chemical precipitation, chemical deposition and membrane filtration [40,
41]. Biological processes utilized microorganism either to sequester the
heavy metals through bio-sorption or to modify their chemical structure,
making them less toxic [42]. The use of microorganisms (microremedia-
tion) and plants (phytoremediation) are the common biological processes

Heavy metal removal


technologies

Biological processes
Physico-chemical technologies
(bioremediation)

• Chemical precipitation • Microremediation


• Adsorption • Phytoremediation
• Floatation
• Ion exchange
• Membrane filtration
• Chemical deposition
• Coagulation-flocculation

Figure 9.3  Classification of heavy metals removal technologies.


326  Applied Water Science Volume 2

for heavy metals removal. Although bioremediation is receiving a signifi-


cant attention due to its cost-effectiveness and sustainability, the physico-
chemical processes are still the more widely utilized technologies.

9.3.1 Adsorption of Heavy Metals


Adsorption has received the widest attention among the physicochemical
processes due to its cost-effectiveness and the availability of vast materials
that can be used as adsorbents [43–45]. Adsorption involves the use of suit-
able materials to adsorb heavy metals from aqueous media. Depending on
the adsorbent type, the adsorption process can occur either through phys-
ical or chemical processes. Physical adsorption (physiosorption) involves
the interaction of the adsorbent and heavy metal through physical forces
(such as the van der Waals) while chemical adsorption (chemisorption) is
controlled by valence forces. In most cases, however, both chemical and
physical interactions occurred simultaneously during adsorption and it is
difficult to demarcate them.
The chemical composition and surface properties of adsorbent are
important considerations in adsorption. In this regard, an effective adsor-
bent is required to have a large specific surface area, high porosity and
abundant functional groups. Activated carbons are widely used as adsor-
bents due to their high surface areas and porosity [46–48]. Due to the high
cost of activated carbon, however, there is growing interest to develop
adsorbents from cheaper materials such as agricultural wastes and other
low-cost materials. Activated carbons derived low-cost materials such as
fly ash [49], pecan nutshell [50], sludge [51], shrimp shell [52], coffee husk
[53], date [54], palm kernel shell [55] and chestnut oak shell [56] have
been used as adsorbents for metal ions. The adsorption capacities of these
low-cost adsorbents, however, are mostly lower than commercial activated
carbon.
Recently, the use of advanced engineered materials as adsorbents for
heavy metals removal is receiving attention. Advanced materials such as
carbon nanotubes [17, 57], graphene oxide [20, 58], polymeric materials
[59, 60] and metal oxides [61, 62] have been investigated. These materi-
als are designed to have large surface areas, porosity and surface proper-
ties suitable for heavy metals uptake. The adsorption capacity of advanced
materials can be enhanced by additional surface modification and combin-
ing different materials into composites with synergistic performance. An
example of this is combining metal ions with organic molecules to form
metal–organic frameworks (MOFs), which are a class of advanced mate-
rials with excellent porosities, large surface areas and tunability [63–65].
Metal–Organic Frameworks for Heavy Metal Removal  327

These unique properties make MOFs an excellent class of adsorbents for


environmental applications.

9.3.2 Metal–Organic Frameworks as Adsorbent for Heavy


Metals Removal
Metal–organic frameworks are formed by combining metal-containing
units with organic linkers, forming highly porous crystalline frame-
works with excellent stability, tunability and organic functionality [63,
66]. Figure 9.4 depicts a typical structure of MOFs. Their structure can
be view as a multidimensional lattice of interconnected inorganic nodes
(metal ions) and organic linkers that are connected through coordina-
tion bonds [67]. The unique properties of MOFs are due to the synergy
of the combined properties of metal ions and organic components [65].
The structures of MOFs can be easily modified through different strate-
gies to obtained desirable chemical functionality, framework topology,
surface area and porosity for specific applications [63]. The ease with
which MOFs can be tuned to form different compounds with unique
properties represents the major advantage of this class of materials. Due
to these excellent properties, MOFs have potentials for a wide range of
applications.
MOFs are commonly prepared through hydrothermal and solvothermal
processes. These processes involve high pressure and temperature reac-
tions in the presence of organic solvents or water. The major limitations
of these conventional synthesis methods are the long period of reaction
and the use of solvents that may have environmental implications [69].

Organic Metal
Linker Node

Figure 9.4  A typical structure of metal–organic framework containing organic linkers


and metal nodes. The organic and metals components are so connected, forming a well-
defined framework with unique properties (Reproduced from Burnett et al. [68] with
permission from Elsevier).
328  Applied Water Science Volume 2

Consequently, there are growing efforts to develop alternative methods


that could eliminate the need for solvents and reduce process time. In this
regard, alternative preparation methods based on mechanochemical [70],
sonochemical [71], electrochemical [72] and microwave-assisted [73] have
been reported. Details on these preparation methods can be found in pre-
vious works [74–76].
Since there are endless choice of starting materials (both organic and
inorganic), different types of MOFs can be developed. The nomenclature
of MOFs is such that they are categorized based on the precursor materials,
place of discovery/development, geometry and topology. For example, ZIF
(such ZIF 8, ZIF 90) represent zeolite imidazolate frameworks, a class of
MOFs based on the topology of zeolite while HKUST-1 (Cu3(BTC)2) rep-
resents a type of metal–organic framework developed in the Hong Kong
University of Science and Technology.
MOFs typically consist of single metal ion/oxide. However, multi-metallic
metal–organic frameworks are nowadays more commonly prepared. Due
to the synergy of the combined metal oxides, multi-metallic MOFs offer
higher performances than single-metallic MOFs [77, 78]. The current
trend in MOFs focuses on the development of multifunctional MOF-based
composites/hybrids, where other functional materials are combined with
MOFs. Zhu and Xu [79] have presented a comprehensive review on MOF–
composites, which include MOF–metal nanoparticles, MOF–metal oxide,
MOF–silica, MOF–organic polymer, MOF–quantum dot, MOF–polyoxo-
metalate and MOF–advanced carbon. Combining MOFs with other mate-
rials resulted in composites with excellent synergistic properties that are
superior to either of the materials.
The excellent properties of MOFs and the ease with which they can be
easily tuned to obtain targeted characteristics have led to their wide poten-
tial applications. For applications in adsorption, two excellent properties of
MOFs are particularly attractive. The ultrahigh porosity (around 90% free
volume) and specific surface areas (often exceeding 7,000 m2/g) [79, 80]
of MOFs make them excellent adsorbents for environmental applications.
In addition, their surface can be easily modified to incorporate additional
functional groups that can assist in adsorbing targeted pollutants. Many
studies have been reported on the efficacy of MOFs as adsorbents (Figure
9.5). However, our discussion focuses on heavy metals adsorption using
MOFs.
MOFs and MOF-based composites have been widely investigated as
potential adsorbents owing to their unique features. In these applications,
the MOFs are specifically tuned to be receptive to the cationic metal ions.
The mechanism of metal ions uptake depends on the composition and
Metal–Organic Frameworks for Heavy Metal Removal  329

Cu2+ Zn2+ Ni2+


Heavy metal ions
Cr3+ Hg2+ Pb2+
s

Ha gas
CP

rm es
PP

fu
l
MOFs COFs

es ic
dy an
rg
Ra

O
di
onu
cl
id
es

Others

Figure 9.5  Typical applications of Metal–organic frameworks (MOFs) to remove


environmental pollutants. MOFs can be used to remove pollutants found in both water
and air (COFs, covalent organic frameworks; PPCPs, pharmaceutical and personal care
products) (Reproduced from Lv et al. [81] with permission from Elsevier).

Cd2+

Figure 9.6  Typical mechanism of heavy metal removal by Cu3(BTC)2-SO3H metal–


organic framework. The metal ion (Cd2+) is diffused into the porous structure of the
metal–organic framework (Reproduced from Wang et al. [82] with permission from The
Royal Society of Chemistry).

topology of the MOFs as well as the metal speciation. The uptake of the
metal ions could occur through diffusion into the interior of the MOF
or through interactions at the surface of the framework [38]. Figure 9.6
depicts a typical mechanism of heavy metal removal where the metallic
ions diffused into the porous structure of the MOF.
330  Applied Water Science Volume 2

9.4 Applications of Metal–Organic Framework


in Heavy Metals Removal
MOFs have been widely applied for heavy metals adsorption. In the reported
studies, both pristine MOFs and modified-MOFs have been utilized for
heavy metals adsorption. In this section, we discussed these specific appli-
cations based on heavy metal types. The heavy metals considered are mer-
cury, copper, chromium, lead, arsenic and cadmium as shown in Table 9.1.

9.4.1 Mercury
Mercury (Hg(II)) has a detrimental effect on human nervous system, brain
and kidneys. Due to its affinity to the thiol group of protein, Hg(II) can
easily accumulate in human body and an exposure to even a small con-
centration of Hg(II) can cause harm [83, 84]. Adsorption is among the
widely used technology and many studies have reported the application of
MOF-based adsorbents to remove Hg (II) from water (Table 9.1). While
pristine MOF can be used for Hg(II) adsorption, in most of these appli-
cations, the MOFs are functionalized to enhance their selectivity towards
Hg(II). The most commonly used materials to functionalize MOFs for
Hg(II) adsorption are sulfur-based functional groups due to the potential
Hg–S interactions.
Hg(II) has a strong affinity towards sulphur-containing moieties such
thiols. Thus, pristine MOF can be functionalized with such materials in
order to enhance their adsorption capacity towards Hg(II). Many stud-
ies have shown that the introducing thiol groups into MOFs strongly
enhanced their adsorption capacity towards Hg(II) due to Hg−S and Hg−N
interactions [116–119]. Li et al. [70] prepared a thiol-functionalized-­MOF
by a post-synthesis modification of pristine MOF. The synthesis involved
the preparing NH2-MiL-68(In) MOF via solvothermal method and the
post-synthetic addition of thiol group (–SH) to obtain SH-MiL-68(In).
This procedure is depicted in Figure 9.7. The presence of the thiol group
increased the selectivity of the MOF towards Hg(II) due to the strong acid-
base interactions, resulting in an optimum adsorption capacity of 450
mg/g. The possible mechanism of Hg(II) removal is the coordination of
Hg(II) ions with –SH group, where the H atoms is replaced by the metal
ions.
Due to the requirement of large amounts of reagents and extreme opera-
tional conditions in the solvothermal process of introducing sulphur func-
tional groups to MOFs, few studies have attempted to developed alternative
Table 9.1  Applications of Metal–organic framework for heavy metals removal.
Adsorbent Contact Adsorption
Heavy dosage Metal conc time capacity
metal MOF (g/L) (mg/L) pH (min) (mg/g) Ref.
Hg SH-MIL-68(In) 1 10 4 120 450 [70]
UiO-66-50Benz 0.5 10 6.8 120 824 [85]
In2S3@MIL-101 10 7 60 518.2 [86]
ZIF-90-THP - 100 7 10 596 [87]
Fe-BTC/PDA 0.5 1 - 24 h 1634 [88]
γ-Fe2O3@CTF-1 4 10 7 24 h 165.8 [89]
Cu MOF-2 (Cd) 1 50 5 3h 769.23 [90]
(Zn3L3(H2O)6][(Na)(NO3) 0.13 100 6 - 379.13 [91]
Cd-MOF-74 0.2 10 6.7 10 189.5 [92]
IRMOF-3/GO 0.8 406 5 8h 254.14 [93]
(Continued)
Metal–Organic Frameworks for Heavy Metal Removal  331
Table 9.1  Applications of Metal–organic framework for heavy metals removal. (Continued)
Adsorbent Contact Adsorption
Heavy dosage Metal conc time capacity
metal MOF (g/L) (mg/L) pH (min) (mg/g) Ref.
Cr IMF-Cr-MOF 0.5 50 - 50 321 [94]
Nitrosomas 100 5 4 180 921 [95]
modified-UiO-66
PAN/chitosan/ 0.4 50 3 60 99.5 [96]
UiO-66-NH2
[Co3(tib)2(H2O)12](SO4)3 1.25 35 4 8h 121 [97]
(BUC-17)
332  Applied Water Science Volume 2

HKUST-1 2 8 6 60 24.20 [98]


UiO-66 0.4 50 3 350 85.6 [99]
ZIF-8 (150 0.4 50 3 350 150 [99]
Cu-BTC 0.5 40 7 - 48 [100]
Cd γ-CD MOF-NPC 1 50 7 60 140.85 [101]
PAN/chitosan/ 0.4 50 6 60 107.6 [96]
UiO-66-NH2
(Continued)
Table 9.1  Applications of Metal–organic framework for heavy metals removal. (Continued)
Adsorbent Contact Adsorption
Heavy dosage Metal conc time capacity
metal MOF (g/L) (mg/L) pH (min) (mg/g) Ref.
[Co3(tib)2(H2O)12](SO4)3 1.25 35 4 8h 121 [97]
Cu3(BTC)2-SO3H 1 200 6 10 88.7 [102]
TMU-16-NH2 0.05 200 6 30 min 126.6 [103]
Pb Cu(TCPBDA) 1 5 - 60 300 [104]
Fe3O4-NHSO3H@ 0.2 400 7 90 384.6 [105]
HKUST-1
MIL-101(Fe)/GO 1 400 6 15 128.6 [106]
UiO-66-NH2@CA 0.4 100 - 64 h 89.40 [107]
Amide-based COF 0.3 300 6.5 24 h 185.7 [108]
PAN/chitosan/ 0.4 50 6 60 115 [96]
UiO-66-NH2
UiO-66-RSA 1 100 4 3H 189.8 [109]
Fe-BTC/PDA 0.5 1 - 24 h 394 [88]
(Continued)
Metal–Organic Frameworks for Heavy Metal Removal  333
Table 9.1  Applications of Metal–organic framework for heavy metals removal. (Continued)
Adsorbent Contact Adsorption
Heavy dosage Metal conc time capacity
metal MOF (g/L) (mg/L) pH (min) (mg/g) Ref.
MOF-2 (Cd) 1 50 5 3h 434.78 [90]
Cu-MOFs/Fe3O4 10 1000 - 60 219 [110]
As
Fe-BTC 5 - 4 10 12.3 [111]
NH2-MIL-88(Fe) 0.2 245 6 60 125 [112]
UiO-66 0.5 50 2 48 h 303 [82]
334  Applied Water Science Volume 2

Fe3O4@UiO-66 0.005 3-150 7 100 73.2 [113]


UiO-66-(SH)2 0.01 50 5 24 h 40 [114]
Zn-MOF-74 1 1000 As(V): 7 4 As(V): 325 [115]
As(III): As(III): 211
12
γ-Fe2O3@CTF-1 4 10 7 24 h 198 [89]
(4-carboxyphenyl) biphenyl-4, 4`-diamine; CA: cellulose aerogels; COF: covalent organic framework; THP: tetrahydropyran; BTC/PDA:
1,3,5-benzenetricaboxylate/polydopamine; CD: cyclodextrin; NPC: nanoporous carbon; CTF: covalent triazine framework; RSA: resorcyl
aldehyde.
Metal–Organic Frameworks for Heavy Metal Removal  335

O
HO
NH2
In(NO3)3 +
OH

al
O

m
er
th SH
H
lvo
2
NH N
So

Functionalization
Mercaptoacetic acid

NH2-MiL-68(In) SH-MiL-68(In)

Figure 9.7  Preparation steps for thiol-functionalized metal–organic framework (MOF).


It involves the synthesis of the MOF through solvothermal method and a post-synthesis
modification to introduce a thiol group (–SH) into the MOF. The –SH replaced the NH2
present in the pristine MOF (Reprinted with permission from Li et al. [70] Copyright
(2019) American Chemical Society).

approaches. Liang et al. [86] introduced indium(III) sulphide nanoparti-


cles (In2S3) into the cavity of MOF (MIL-101), producing In2S3@MIL-101
composite with ultrafast adsorption of Hg(II). The procedure involved
converting pristine MIL-101 into In3+@MIL-101 using In(NO3)3 solution,
which was subsequently converted into In2S3@MIL-101 through solid–
gas reaction with H2S. The composite achieved an adsorption capacity of
518 mg/g for Hg(II) adsorption. This approach produced a more effective
adsorbent compared to some previous studies using the conventional sol-
vothermal approach to prepare sulphur-functionalized MOF [70, 117].
A photo-assisted post-synthetic modification approach to introduce
pyrimidine–thione fragment onto ZIF-90 was reported by Yin et al. [87].
The synthesis method involved a multicomponent reaction between pris-
tine ZIF-90, thioureas, and tetrahydropyran (THP) or tetrahydrofuran
(THF) under UV irradiation. The functionalized materials exhibited high
adsorption of Hg(II), up to 596 and 403 mg/g for ZIF-90-THP and ZIF-90-
THF respectively. These adsorption capacities are superior to the pristine
ZIF-90, whose adsorption capacity was only 47 mg/g. The advantage of
this approach is that multi-functional groups can be incorporated into the
MOF in a single stage, eliminating the multi-stage procedure existing in
other reported procedures.
MOF-derived advanced materials have also been used as adsor-
bents for Hg(II). Li et al. [120] derived ZrOx, ZrOxyPhos and ZrSulf
336  Applied Water Science Volume 2

from MOFs through a ligand extraction method. UiO-66-50Benz


(Zr6(OH)4O4(Benzene-1,4-dicarboxylate)6,) was used as the precursor
where specific functional groups were added to it, while maintaining its
original morphology and porosity. The ligand extraction process involved
replacing the organic ligand of the pristine MOF with inorganic moieties.
In this case, NaOH, Na3PO4 and Na2S.9H2O were used as the inorganic
moieties to produced ZrOx, ZrOxyPhos and ZrSulf respectively. ZrSulf
exhibited the highest adsorption capacity of 824 mg/g, while ZrOxyPhos
and ZrOx achieved 663 and 485 mg/g respectively. These adsorption
capacities are higher than that of the pristine UiO-66-50Benz (363 mg/g).
MOF-based materials have potentials for Hg(II) removal based on the
attained adsorption capacities reported above. It is evident from the previ-
ous studies that the major driving factor for the excellent removal of Hg(II)
is the presence of functional groups in the MOFs, particularly sulphur-based
moieties due to the strong Hg-S interactions. Consequently, recent studies
on Hg(II) adsorption using MOFs are mostly on post-synthesis modifica-
tion using sulfur-based materials. Thus, selection of appropriate function-
alities and development/refinement of functionalization methods are the
major research focus.

9.4.2 Copper
Although copper is an essential trace element required by plants and
animals, it has been linked to many health problems, including neuro/­
hepatodegenerative disorders such as Alzheimer’s and Wilson’s diseases
[121, 122]. Due to the extensive industrial usage of copper, it is commonly
found at high concentrations in the environment. To protect human
health, stringent guidelines have been recommended for the concentration
of Cu(II) in drinking water. For example, the United Sates Environmental
Protection Agency has set the 1.3 ppm as the permissible level of Cu(II)
in drinking water [123]. The removal of Cu(II) from the environment is
therefore of great importance.
MOFs have been investigated for Cu(II) removal (Table 9.1). While the
adsorption of Cu(II) by some MOFs occurs through ion-exchange with
the MOF’s metallic ions [124], the presence of amino functional groups
(–NH2) is the driving force for Cu(II) adsorption in most cases. Cu(II) is
a Lewis acid with a lone pair electrons that can accept an electron from
Lewis base such as amino group, resulting in a stable chemical coordina-
tion [124, 125]. The adsorption of Cu(II) can also occur through inter-
actions with uncoordinated carboxylate oxygen atoms presence in some
MOFs (Yu et al., 2018).
Metal–Organic Frameworks for Heavy Metal Removal  337

Zheng et al. [92] prepared Cd-based MOF (Cd-MOF-74) through


a solvothermal synthesis method and applied it for Cu(II) removal. The
prepared MOF achieved an adsorption capacity of 190 mg/g. Besides the
conventional solvothermal approach, ultrasonic synthesis methods could
also be used to prepare Cd-based MOF for Cu(II) removal. Ghaedi et al.
[90] prepared a Cd-based MOF through ultrasonic synthesis method for
Cu(II) adsorption. The procedure involved a room temperature ultrasonic
driven reaction of Cd(CH3COO)2·2H2O, terephthalate and dimethylfor-
mamide to produce the Cd-TPA template. The prepared MOF achieved
a metal uptake of 435 mg/g for Cu(II) and could be used for three cycles
while maintaining its effectiveness. A recent study has also shown that
microwave-assisted methods could provide an alternative to the conven-
tional solvothermal/hydrothermal methods [124].
The use of MOF-based hybrids/composites have also been investi-
gated as a possible way to enhance the performance of MOFs for Cu(II)
removal. Rao et al. [93] prepared an MOF/graphene oxide composite

DMF
Zn2+ 100ºC, 24h

graphene oxide (GO)


2-aminoterephthalic acid
IRMOF-3/GO

on
rpti
adso O O O O
O Zn Zn O O Zn Zn O
O O
O Zn Zn O O Zn Zn O
O O H2N O O
Cu2+
NH2
Cu2+
Cu2+
H2N Cu2+
O O NH2 O O
O Zn Zn O O Zn Zn O
O O
Cu2+ O Zn Zn O O Zn Zn O
O O O O

Figure 9.8  Schematic presentation of the preparation 2-aminoterephthalic acid-ZnO4


MOF (IRMOF-3)/graphene oxide composite and its application for Cu(II) removal. The
IRMOF-3 is synthesized through solvothermal and then incorporated with graphene
oxide using dimethylformamid (DMF). The adsorption of Cu(II) is coordinated by the
amino group (–NH2) of the composite (Reprinted with permission from Rao et al. [93]
Copyright (2017) American Chemical Society).
338  Applied Water Science Volume 2

for Cu(II) removal, where graphene oxide was incorporated into a pris-
tine 2-aminoterephthalic acid-ZnO4 MOF (IRMOF-3). Figure 9.8 shows
the schematic of the preparation of the IRMOF-3/graphene oxide and the
plausible mechanism of Cu(II) uptake by the composite. The adsorption
capacity and selectivity of the pristine IRMOF-3 was enhanced by the
addition of the graphene oxide, with the adsorption capacity reaching 254
mg/g. Recently, Wang et al. [124] combined amino group-functionalized
Zr-MOF and ceramic membrane for Cu(II) removal. While the pristine
Zr-MOF removed 60 mg/g of Cu(II), the combined Zr-MOF/ceramic
membrane resulted Cu(II) uptake of 988 mg/g. Although the adsorption
performance is significantly enhanced, the stability of the MOF remains a
challenge.
The removal of Cu(II) can be through ion-exchange with metallic ions
such as Ca(II) or through chemical coordination with amino groups pres-
ent in the MOFs. The removal of Cu(II) through exchange with the MOF’s
metallic ions can affect the stability of the MOFs. Thus, the interactions
with the MOF’s anions is a better option to remove Cu(II) while maintain-
ing the stability of the adsorbent. Since Cu(II) is a Lewi’s acid with lone
electrons, it can easily coordinate with Lewis’s base such as –NH2, resulting
in effective removal of Cu(II).

9.4.3 Chromium
Chromium ions occurs mainly as hexavalent (Cr(VI)) or trivalent (Cr(III))
ions in the environment. Thus, the characteristics and toxicity of chro-
mium depend on its oxidation state. While Cr(III) is considered relatively
harmless, Cr(VI) is highly toxic and has been classified as carcinogenic,
with its recommended maximum permissible level in drinking water set
at 0.1 ppm [95, 126]. Thus, the removal of Cr(VI) has received more atten-
tion, with adsorption being widely reported.
Noraee et al. [99] reported the use of pristine Uio-66 and ZIF-8 for
Cr(VI) adsorption. The adsorption capacities for ZIF-8 and Uio-66 were
150 and 85.6 mg/g respectively. Though the authors have attributed the
superior performance of ZIF-8 to its higher specific surface area, the possi-
ble influence of the MOF’s composition was not highlighted. Maleki et al.
[100] prepared copper-benzenetricarboxylates (Cu-BTC) through a solvo-
thermal method and applied it for Cr(VI) adsorption. The prepared MOF
exhibited an effective adsorption of Cr(VI) at neutral pH. The adsorption
of Cr(VI) by a silver-triazolate MOF was also reported, where a maximum
uptake of 38 mg/g was achieved [127]. The adsorption of Cr(VI) was pre-
dominantly through anion-exchange with the cationic MOF.
Metal–Organic Frameworks for Heavy Metal Removal  339

Modified-MOFs have also been applied for Cr(VI) adsorption. Guo et al.
[97] tuned a BUC-17 MOF [Co3(tib)2(H2O)12](SO4)3] into an advanced
powder for Cr(VI) uptake from simulated wastewater. An adsorption
capacity of 121 mg/g was achieved, which was attributed to electrostatic
interactions and ion-exchange between the MOF and Cr(VI). Sathvika
et al. [95] prepared a Nitrosomonas-modified Uio-66 through a micro-
wave assisted method for the adsorption of Cr(VI). The Cr(VI) uptake
was predominantly through electrostatics interactions between the func-
tional groups of the modified-MOF and the chromate ions as shown in
Figure 9.9. While the Nitrosomonas sp. and pristine Uio-66 MOF exhibited
Langmuir adsorption capacities of 8.98  and 13.33 mg/g respectively, the
Nitrosomonas-modified Uio-66 achieved 23.69 mg/g. The enhancement in
adsorption performance is due to the synergistic increase in the specific
surface area and functional groups of the combined materials.
Recently, Jamshidifard et al. [96] reported a hybrid of UiO-66-NH2,
chitosan and polyacrylonitrile for Cr(VI) adsorption. The hybrid was
prepared first by synthesizing Uio-66-NH2 through microwave-assisted
method and then incorporating it into the chitosan/polyacrylonitrile solu-
tion using ultrasonication and electrospinning processes. The maximum
adsorption capacity achieved by the adsorbent was 373 mg/g, which could
be effectively maintained over five cycles of adsorption–desorption. The
excellent performance of the adsorbent is due to the enhanced surface area
and surface functional groups of the composite.

HCrO4¯

HCrO4¯
HCrO4¯
HCrO4¯ HCrO

HCrO4¯

pH4 HCrO4¯
+
HCrO4¯

HCrO4¯

HCrO

Cr(VI) solution HCrO4¯


HCrO4¯
HCrO4¯

HCrO4¯

HCrO4¯
HCrO4¯
Nitrisomonas immobilized MOF
After Cr(VI) adsorption

Figure 9.9  Possible mechanism of Cr(VI) uptake by Nitrosomonas-modified Uio-66.


The adsorption occurs through electrostatic interactions between the surface functional
groups of the modified-MOF and the Cr(VI) (pink rod = microbe, blue = Zr, green =
O, orange = C, violet = H, yellow = Cr(VI)) (Reprinted from Sathvika et al. [95] with
permission from Elsevier).
340  Applied Water Science Volume 2

9.4.4 Lead
Lead (Pb(II)), one of the most commonly used heavy-metal in industrial
activities such as smelting, metal plating, oil refining, battery manufactur-
ing and painting, causes toxic and carcinogenic effects on human health
[128]. Lead ingestion at higher levels mainly results in kidney damage, slow
bone growth, metal decline and neurodegenerative diseases. The WHO, the
European Union and US Environmental Protection Agency have restricted
Pb(II) concentration to 0.01, 0.01 and 0.015 mg/L respectively [129, 130].
Huang et al. [131] synthesized zeolite-imidazolate frameworks, ZIF-8
and ZIF-67, for Pb(II) removal. The MOFs exhibited adsorption capacities
of 1,119 and 1,348 mg/g respectively, which were higher than some avail-
able porous materials. Qin et al. [132] developed another Pb (II) adsorbing
MnO2-MOF. The MOF was synthesized via simple oxidation process and
the metal uptake efficiency was determined by a fast equilibrium obtained
within 1 h. The material exhibited an adsorption capacity of 917 mg/g,
which was attained as a result of inner-sphere complexation of hydroxyl
groups with the metal ions.
Hasankola et al. [133] synthesized Cu-BTC and Zn-BTC MOFs by
solvothermal reaction using benzene-1,3,5-tricarboxylic acid as a linker.
The produced composites attained maximum adsorption capacities of
333 and 312 mg/L respectively. The frameworks exhibited similar func-
tions in adsorbing Pb(II) ions, which could be maintained over three
adsorption-desorption cycles. Li et al. [70] evaluated the performance of
two amide-based COFs (COF-TP and COF-TE) for Pb(II) and obtained
adsorption capacities of 140 and 185.7 mg/g respectively. The Pb(II)
uptake was enhanced by the amide group, which behaves as active adsorp-
tion site for the metal ions via multi-coordination (Figure 9.10). Sun et al.
[88] reported polymer-based MOF composite, Fe-BTC/PDA, t for Pb(II)
adsorption, which achieved an adsorption capacity of 394 mg/g. Abbasi
et al. [134] developed 3D Co-MOF composites for Pb(II) adsorption. The
performance of the adsorbent was affected by the solution pH, metal ion
concentration and treatment time.
Zhang et al. [130] synthesized HS-mSi@MOF-5 framework, a silica
coated thiolated MOF-5 derivative, which achieved Pb(II) removal of
312 mg/g. The performance of the MOF was strongly affected by the solu-
tion ph, with the highest performance at low pH values. A schematic rep-
resentation of the effect of pH on the adsorption is shown as Figure 9.11.
Chakraborty et al. [135] introduced Zinc(II) and tetracarboxylate based-
MOF, which showed a maximum metal uptake of 71 mg/g. Rivera et al.
[136] reported adsorption characteristics of MOF-5 for Pb(II) removal.
Metal–Organic Frameworks for Heavy Metal Removal  341

Pb2+

Pb2+
C
O
N

Figure 9.10  Graphical representation of Pb (II) adsorption mechanism by COF based


MOFs (Reprinted from Li et al. [70] with permission from Elsevier).

Thiol-coordination interaction Thiol-coordination interaction Thiol-coordination interaction


+ + +
Electrostatic repulsion Weak electrostatic interaction Strong electrostatic interaction
M2+ M2+ M2+,M(OH)+

+ + + – – – – – –
HS-mSi@MOF-5 HS-mSi@MOF-5 HS-mSi@MOF-5
+ + + – – – – – – –
pH
pH=2 pH=3 pH=6

Figure 9.11  Adsorption mechanism of Pb(II) at different pH values (Reproduced from


Zhang et al. [130] with permission from Elsevier).

The adsorption capacity showed decreasing trend with increase in pH, with
the adsorption capacities of 660 and 750 mg/g at pH 6 and 4 respectively.

9.4.5 Arsenic
Arsenic occurs mainly in an inorganic form in the environment [137]. It
commonly exists in oxidation states of +5, +3, +1 and −3 valences [21]
and is component of roughly 245 minerals including elemental arsenic,
arsenides, arcentates, etc. [138]. Daily intake of water with more than 0.01
ppm of arsenic can cause skin, circulatory and nervous system problems
[21]. The maximum acceptable concentration of arsenic in drinking water
342  Applied Water Science Volume 2

recommended by the WHO is 10 μg/l [20]. According to estimates, over 60


kilo tons of arsenic is released into the environment annually [139]. MOFs
are promising materials for the treatment of arsenic-containing water and
wastewater. Figure 9.12 depicts the development stages of MOFs used for
arsenic removal application.
The first reported application of MOF for the removal of aqueous arse-
nic was Fe-BTC [111]. The authors compared the schematic illustration
of the adsorption strategies for bulk material, nano-particles and MOFs
(Figure 9.13). The MOF contains iron nodes and 1,3,5-benzenetrocarbox-
ylic linkers. Metal ions linkage with specific polyatomic organic bridging
ligands and high thermal and mechanical stability were the two listed
advantages of the MOF. The material could adsorb As(V) with adsorption
capacity of 12.3 mg/g.
Classic MOFs were developed by several researchers for adsorption of
arsenic. For instance, Li et al. [140] synthesized ZIF-8 adsorbent, which
showed an adsorption capacity of 76.5 mg/g, which was attributed to the
presence of abundant Zn-OH functional group. To further improve the
adsorption capacity of ZIF-8, Massoudinejad et al. [141] functionalized it
with ethylenediamine and obtained adsorption capacity of 83.5 mg/g. Vu
et al. [142] investigated another form of classic MOF i.e. MIL-53 (Fe), syn-
thesized via solvothermal method. The adsorbent achieved an adsorption
capacity of 21 mg/g. Other examples include MIL-53 (Al), MIL-88A and
MIL-88B with adsorption capacities of 106, 145 and 156 mg/g respectively
[112, 143, 144].

+ H2O
First MOF reported MOF applied for ions adsorption MOF with superior As adsorption
*J. Am. Chem. Soc. 1989, 111, 5962 *Angew. Chem. Int. Ed. 2010, 49, 1057 *Sci. Rep. 2015, 5, 16613

1989 2009 2010 2012 2015 2017 Future

Water stable MOF MOF applied for As adsorption MOF membrane for As rejection
*J. Am. Chem. Soc. 2009, 131, 8784 *J. Phys. Chem. 2012, 116, 8601 *J. Memb. Sci. 2017, 541, 262

Figure 9.12  Development of MOFs for Arsenic removal (Reprinted from Wang et al.
[138] with permission from Elsevier).
Metal–Organic Frameworks for Heavy Metal Removal  343

Absorb

Bulk material absorbents

Absorb

Nanoparticle absorbents

Absorb

Metal-organic coordination
polymer absorbents

Figure 9.13  Graphical demonstration of adsorption using different materials (Reprinted


from Zhu et al. [111] with permission from ACS).

Indium-based AUBM-1 [145], Co-MOF [146], Ni-MOF [126], MOF-


808 [147], Zr-MOF UiO-66 [102] have also shown high adsorption
capacities for arsenic. Among them, the UiO-66 exhibited the highest
adsorption capacity of 303 mg/g. Wang et al. [102] suggested surface OH
and organic ligands as the major active sites for adsorption of heavy metal.
The adsorbent worked efficiently over a wide pH range (1–10). Huo et al.
[154] synthesized Fe3O4@UiO-66 for the arsenic removal. The composite
demonstrated a metal ion uptake of 73 mg/g and efficient separation in the
presence of magnetic field.
The adsorption of arsenic by MOFs has been investigated and a series of
thermodynamic and kinetic investigation have been conducted to develop
the mechanism of the adsorption process. Through spectroscopic stud-
ies, the important role of terminal OH for arsenic complexation, resulting
in the formation of M–O–As complex has been highlighted. In addition,
amino (–NH2) and thiol (–SH2) groups are the potential candidates for the
adsorption of arsenic. Besides, electrostatic interactions, hydrogen bond-
ing, van der Waals forces are few attractive interactions which assist the
arsenic uptake process [120].
344  Applied Water Science Volume 2

9.4.6 Cadmium
Cadmium is a heavy metal with a high toxicity and is introduced into the
environment through various industrial activities. Mainly produced as a
by-product from mining, smelting and industrial activities [148, 149], it
generally exist as divalent cations and complexed with other elements (e.g.
CdCl2, CdCO3, and Cd(OH)2) [137, 150]. Its compounds can exist in several
phases and form chemical compounds due to its low solubility. In nature,
it forms complexes with inorganic (e.g. Cl¯, SO4−2 , HCO3−1 , F¯) and organic
ligands (amino acids, citrate, oxalate and humic acid). Although Cd exists
in environment in traces, it is bio-accumulative and is demonstrated to have
detrimental effects on several organisms [151]. Cadmium has been classi-
fied as human carcinogen, with its recommended maximum contaminant
level (MCL) in drinking water set at 3 μg/L [130].
Roushani et al. [103] reported the adsorption of Cd(II) by TMU-
16-NH2 MOF. The material achieved a maximum adsorption capacity of
126.6 mg/g. Wang et al. [102] reported the adsorptive uptake of Cd(II)
using Cu3(BTC)2-SO3H framework, a material synthesized through oxi-
dation of Cu3(BTC)2. The adsorption process was marginally controlled
by pH fluctuation, with the optimum results obtained at the pH value of
6. The adsorption process was chelation reaction between sulfonic groups
of MOF and Cd (II) ions. The sulfonic groups resulted in an improved
selectivity for Cd(II) due to the multiple bonding sites and coordina-
tion modes of SO3H [102]. Another Cd(II) adsorbing MOF, AMOF-1,
was introduced by Chakraborty et al. [135]. The MOF was based on Zn
and tetracarboxylate linkers, and achieved an adsorption capacity of 41
mg/g. Feng et al. [153] presented PCN-100 MOF synthesized from TATAB
([[Zn4O(C24H15N6O6)2(H2O)2]·6H2O·DMF]n  (1) based on 4,4′,4″-s-tri-
azine-1,3,5-triyltri-p aminobenzoate) linkers and Zn4O(CO2)6 second-
ary building units. The MOF was found to interact with metal ions in
chelating coordination mode through linkers. In addition, a magnetic
Cu-terephthalate MOF was evaluated for Cd(II) adsorption [152]. The
adsorption process occurs via chemical adsorption process, with the max-
imum adsorption capacity of 100 mg/g. The adsorption performance was
greatly enhanced by the COOH groups of the MOF.

9.5 Conclusion
Metal–organic frameworks are excellent adsorbents due to their high
porosity, large surface areas and tunability. For applications in heavy metals
Metal–Organic Frameworks for Heavy Metal Removal  345

removal, two excellent properties of MOFs are particularly attractive. The


ultrahigh porosity (around 90% free volume) and specific surface areas
(often exceeding 7,000 m2/g) make MOFs suitable adsorbents for heavy
metals. Numerous studies have shown the efficacy of MOFs in the adsorp-
tion of heavy metals such as mercury, copper, chromium, lead, arsenic and
cadmium as discussed herein. While there have been studies on the use
of pristine MOFs, we observed a growing trend in the use of MOF-based
composites and MOF-based derived advanced materials for heavy metals
adsorption. It is evident that MOFs represent one of the attractive routes
for removing heavy metals remediation. To fully utilize their potentials,
future research must focus on the refinement of synthesis methods and
stability of MOF-based adsorbents.

References
1. Demirak, A., Yilmaz, F., Levent Tuna, A., Ozdemir, N., Heavy metals in
water, sediment and tissues of Leuciscus cephalus from a stream in south-
western Turkey. Chemosphere, 63, 9, 1451–1458, 2006.
2. Zhao, S., Duan, Y., Chen, L., Li, Y., Yao, T., Liu, S., Liu, M., Lu, J., Study on
emission of hazardous trace elements in a 350 MW coal-fred power plant.
Part 1. Mercury. Environ. Pollut., 229, 863–870, 2017.
3. Wang, C., Luan, J., Wu, C., Metal–organic frameworks for aquatic arsenic
removal. Water Res., 158, 370–382, 2019.
4. Bola, O., Adefoye, T., Osibote, A., Heavy Metals Contamination of Water,
Soil, and Plants around an Electronic Waste Dumpsite. Pol. J. Environ. Stud.,
22, 1431–1439, 2013.
5. Herat, S. and Agamuthu, P., E-waste: A problem or an opportunity? Review
of issues, challenges and solutions in Asian countries. Waste Manage. Res.,
30, 11, 1113–1129, 2012.
6. Akan, J.C., Abdulrahman, F.I., Ayodele, J.T., Ogugbuaja, V.O., Impact of tan-
nery and textile effluent on the chemical characteristics of Challawa River,
Kano State, Nigeria. Aust. J. Basic Appl. Sci., 3, 3, 1933–1947, 2009.
7. Faryal, R. and Hameed, A., Isolation and characterization of various fungal
strains from textile effluent for their use in bioremediation. Pak. J. Bot., 37, 4,
1003–1008, 2005.
8. Gao, B.Y., Yue, Q.Y., Wang, Y., Zhou, W.Z., Color removal from dye-contain-
ing wastewater by magnesium chloride. J. Environ. Manage., 82, 2, 167–172,
2007.
9. Kim, T.-H., Park, C., Yang, J., Kim, S., Comparison of disperse and reac-
tive dye removals by chemical coagulation and Fenton oxidation. J. Hazard.
Mater., 112, 1–2, 95–103, 2004.
346  Applied Water Science Volume 2

10. Pala, A. and Tokat, E., Color removal from cotton textile industry waste­
water in an activated sludge system with various additives. Water Res., 36, 11,
2920–2925, 2002.
11. Robinson, T., McMullan, G., Marchant, R., Nigam, P., Remediation of dyes
in textile effluent: a critical review on current treatment technologies with a
proposed alternative. Bioresour. Technol., 77, 3, 247–255, 2001.
12. Savin, I.I. and Butnaru, R., Wastewater characteristics in textile finishing
mills. Environ. Eng. Manage. J., 7, 6, 859–864, 2008.
13. Joseph, L., Jun, B.-M., Flora, J.R.V., Park, C.M., Yoon, Y., Removal of heavy
metals from water sources in the developing world using low-cost materials:
A review. Chemosphere, 229, 142–159, 2019.
14. Vardhan, K.H., Kumar, P.S., Panda, R.C., A review on heavy metal pollu-
tion, toxicity and remedial measures: Current trends and future perspectives.
J. Mol. Liq., 290, 111197, 2019.
15. WHO, Guidelines for Drinking-Water Quality, 2017a.
16. WHO, Progress on Drinking Water, Sanitation and Hygiene: 2017 Update and
SDG Baseline, UNICEF, 2017b.
17. Xu, J., Cao, Z., Zhang, Y., Yuan, Z., Lou, Z., Xu, X., Wang, X., A review of
functionalized carbon nanotubes and graphene for heavy metal adsorption
from water: Preparation, application, and mechanism. Chemosphere, 195,
351–364, 2018.
18. UN-Water, The United Nations World Water Development Report 2018:
Nature-Based Solutions for Water, 2018.
19. Chowdhury, S., Mazumder, M.A.J., Al-Attas, O., Husain, T., Heavy metals in
drinking water: Occurrences, implications, and future needs in developing
countries. Sci. Total Environ., 569–570, 476–488, 2016.
20. Sherlala, A.I.A., Raman, A.A.A., Bello, M.M., Asghar, A., A review of the
applications of organo-functionalized magnetic graphene oxide nanocom-
posites for heavy metal adsorption. Chemosphere, 193, 1004–1017, 2018.
21. Uwamariya, V., Adsorptive removal of heavy metals from groundwater by iron
oxide based Adsorbents, (Doctoral Degree), Delf University of Technology,
CRC Press/Balkema, 2013.
22. Beyersmann, D. and Hartwig, A., Carcinogenic metal compounds: recent
insight into molecular and cellular mechanisms. Arch. Toxicol., 82, 8, 493,
2008.
23. Wang, S. and Shi, X., Molecular mechanisms of metal toxicity and carcino-
genesis. Mol. Cell. Biochem., 222, 1, 3–9, 2001.
24. Tchounwou, P.B., Yedjou, C.G., Patlolla, A.K., Sutton, D.J., Heavy metal tox-
icity and the environment. Experientia Suppl., 101, 133–164, 2012.
25. Fu, F. and Wang, Q., Removal of heavy metal ions from wastewaters: A
review. J. Environ. Manage., 92, 3, 407–418, 2011.
26. Rebelo, F.M. and Caldas, E.D., Arsenic, lead, mercury and cadmium: Toxicity,
levels in breast milk and the risks for breastfed infants. Environ. Res., 151,
671–688, 2016.
Metal–Organic Frameworks for Heavy Metal Removal  347

27. Purkayastha, D., Mishra, U., Biswas, S., A comprehensive review on Cd(II)
removal from aqueous solution. J. Water Process Eng., 2, 105–128, 2014.
28. Jin, W., Du, H., Zheng, S., Zhang, Y., Electrochemical processes for the envi-
ronmental remediation of toxic Cr(VI): A review. Electrochim. Acta, 191,
1044–1055, 2016.
29. Bhan, A. and Sarkar, N.N., Mercury in the Environment: Effect on Health
and Reproduction. Rev. Environ. Health, 20, 1, 39–56, 2005.
30. Duan, Y., Han, D.S., Batchelor, B., Abdel-Wahab, A., Application of a reactive
adsorbent-coated support system for removal of mercury(II). Colloids Surf.
A: Physicochem. Eng. Asp., 509, 623–630, 2016
31. Abdullah, N., Yusof, N., Lau, W.J., Jaafar, J., Ismail, A.F., Recent trends of
heavy metal removal from water/wastewater by membrane technologies.
J. Ind. Eng. Chem., 76, 17–38, 2019.
32. García-Niño, W.R. and Pedraza-Chaverrí, J., Protective effect of curcumin
against heavy metals-induced liver damage. Food Chem. Toxicol., 69, 182–
201, 2014.
33. Pourrut, B., Shahid, M., Dumat, C., Winterton, P., Pinelli, E., Lead Uptake,
Toxicity, and Detoxifcation in Plants, in: Reviews of Environmental
Contamination and Toxicology, vol. 213, D.M. Whitacre (Ed.), pp. 113–136,
Springer New York, New York, NY, 2011.
34. Vareda, J.P., Valente, A.J.M., Durães, L., Assessment of heavy metal pollu-
tion from anthropogenic activities and remediation strategies: A review.
J. Environ. Manage., 246, 101–118, 2019.
35. Bradl, H.B., Chapter 1 Sources and origins of heavy metals, in: Interface
Science and Technology, vol. 6, pp. 1–27, Elsevier, 2005.
36. Heavy Metals in Soils, in: Sources of Heavy Metals and Metalloids in Soils, B.J.
Alloway (Ed.), Springer Netherlands, 2012.
37. Liu, Y., Xing, J., Wang, S., Fu, X., Zheng, H., Source-specifc speciation pro-
files of PM2.5 for heavy metals and their anthropogenic emissions in China.
Environ. Pollut., 239, 544–553, 2018.
38. Kobielska, P.A., Howarth, A.J., Farha, O.K., Nayak, S., Metal–organic frame-
works for heavy metal removal from water. Coord. Chem. Rev., 358, 92–107,
2018.
39. Sun, L., Guo, D., Liu, K., Meng, H., Zheng, Y., Yuan, F., Zhu, G., Levels,
sources, and spatial distribution of heavy metals in soils from a typical coal
industrial city of Tangshan, China. CATENA, 175, 101–109, 2019.
40. Barakat, M.A., New trends in removing heavy metals from industrial waste-
water. Arabian J. Chem., 4, 4, 361–377, 2011.
41. Bolisetty, S., Peydayesh, M., Mezzenga, R., Sustainable technologies for water
purification from heavy metals: Review and analysis. Chem. Soc. Rev., 48, 2,
463–487, 2019.
42. Gardea-Torresdey, J.L., de la Rosa, G., Peralta-Videa, J.R., Use of phytofil-
tration technologies in the removal of heavy metals: A review. Pure Appl.
Chem., 76, 4, 801–813, 2004.
348  Applied Water Science Volume 2

43. Bello, M.M. and Raman, A.A.A., Adsorption and Oxidation Techniques to
Remove Organic Pollutants from Water, in: Green Adsorbents for Pollutant
Removal: Fundamentals and Design, G. Crini and E. Lichtfouse (Eds.),
pp. 249–300, Springer International Publishing, Cham, 2018.
44. Uddin, M.K., A review on the adsorption of heavy metals by clay minerals,
with special focus on the past decade. Chem. Eng. J., 308, 438–462, 2017.
45. Yagub, M.T., Sen, T.K., Afroze, S., Ang, H.M., Dye and its removal from
aqueous solution by adsorption: A review. Adv. Colloid Interface Sci., 209,
172–184, 2014.
46. Bello, M.M. and Raman, A.A.A., Synergy of adsorption and advanced oxida-
tion processes in recalcitrant wastewater treatment. Environ. Chem. Lett., 17,
2, 1125–1142, 2019.
47. Bhatnagar, A., Hogland, W., Marques, M., Sillanpää, M., An overview of the
modification methods of activated carbon for its water treatment applica-
tions. Chem. Eng. J., 219, 499–511, 2013.
48. Foo, K.Y. and Hameed, B.H., An overview of landfill leachate treatment via
activated carbon adsorption process. J. Hazard. Mater., 171, 1, 54–60, 2009.
49. Xiyili, H., Çetintaş, S., Bingöl, D., Removal of some heavy metals onto
mechanically activated fy ash: Modeling approach for optimization, iso-
therms, kinetics and thermodynamics. Process Saf. Environ. Prot., 109, 288–
300, 2017.
50. Aguayo-Villarreal, I.A., Bonilla-Petriciolet, A., Muñiz-Valencia, R.,
Preparation of activated carbons from pecan nutshell and their application
in the antagonistic adsorption of heavy metal ions. J. Mol. Liq., 230, 686–695,
2017.
51. Ni, B.-J., Huang, Q.-S., Wang, C., Ni, T.-Y., Sun, J., Wei, W., Competitive
adsorption of heavy metals in aqueous solution onto biochar derived from
anaerobically digested sludge. Chemosphere, 219, 351–357, 2019.
52. Núñez-Gómez, D., Rodrigues, C., Lapolli, F.R., Lobo-Recio, M.Á.,
Adsorption of heavy metals from coal acid mine drainage by shrimp shell
waste: Isotherm and continuous-flow studies. J. Environ. Chem. Eng., 7, 1,
102787, 2019.
53. Hernández Rodiguez, M., Yperman, J., Carleer, R., Maggen, J., Dadi, D.,
Gryglewicz, G., Bruggen, B.V.d., Hernandez, J.F., Calvis, A.O., Adsorption
of Ni(II) on spent coffee and coffee husk based activated carbon. J. Environ.
Chem. Eng., 6, 1, 1161–1170, 2018.
54. Norouzi, S., Heidari, M., Alipour, V., Rahmanian, O., Fazlzadeh, M.,
Mohammadi-moghadam, F., Nourmoradi, H., Goudarzi, B., Dindarloo, K.,
Preparation, characterization and Cr(VI) adsorption evaluation of NaOH-
activated carbon produced from Date Press Cake; an agro-industrial waste.
Bioresource Technol., 258, 48–56, 2018.
55. Karri, R.R. and Sahu, J.N., Modeling and optimization by particle swarm
embedded neural network for adsorption of zinc(II) by palm kernel shell
Metal–Organic Frameworks for Heavy Metal Removal  349

based activated carbon from aqueous environment. J. Environ. Manage., 206,


178–191, 2018.
56. Niazi, L., Lashanizadegan, A., Shariffard, H., Chestnut oak shells activated
carbon: Preparation, characterization and application for Cr (VI) removal
from dilute aqueous solutions. J. Cleaner Prod., 185, 554–561, 2018.
57. Ihsanullah, Abbas, A., Al-Amer, A.M., Laoui, T., Al-Marri, M.J., Nasser, M.S.,
Khraisheh, M., Atieh, M.A., Heavy metal removal from aqueous solution by
advanced carbon nanotubes: Critical review of adsorption applications. Sep.
Purif. Technol., 157, 141–161, 2016.
58. Peng, W., Li, H., Liu, Y., Song, S., A review on heavy metal ions adsorption
from water by graphene oxide and its composites. J. Mol. Liq., 230, 496–504,
2017.
59. Sajid, M., Nazal, M.K., Ihsanullah, Baig, N., Osman, A.M., Removal of heavy
metals and organic pollutants from water using dendritic polymers based
adsorbents: A critical review. Sep. Purif. Technol., 191, 400–423, 2018.
60. Vunain, E., Mishra, A.K., Mamba, B.B., Dendrimers, mesoporous silicas and
chitosan-based nanosorbents for the removal of heavy-metal ions: A review.
Int. J. Biol. Macromol., 86, 570–586, 2016.
61. Hua, M., Zhang, S., Pan, B., Zhang, W., Lv, L., Zhang, Q., Heavy metal
removal  from water/wastewater by nanosized metal oxides: A review.
J. Hazard. Mater., 211–212, 317–331, 2012.
62. Sharma, M., Singh, J., Hazra, S., Basu, S., Adsorption of heavy metal ions by
mesoporous ZnO and TiO2@ZnO monoliths: Adsorption and kinetic stud-
ies. Microchem. J., 145, 105–112, 2019.
63. Furukawa, H., Cordova, K.E., O’Keefe, M., Yaghi, O.M., Te Chemistry and
Applications of Metal–Organic Frameworks. Science, 341, 6149, 1230444,
2013.
64. Gao, Q., Xu, J., Bu, X.-H., Recent advances about metal–organic frameworks
in the removal of pollutants from wastewater. Coord. Chem. Rev., 378, 17–31,
2019.
65. Zhou, H.-C., Long, J.R., Yaghi, O.M., Introduction to Metal–Organic
Frameworks. Chem. Rev., 112, 2, 673–674, 2012.
66. Yaghi, O.M., O’Keefe, M., Ockwig, N.W., Chae, H.K., Eddaoudi, M., Kim,
J., Reticular synthesis and the design of new materials. Nature, 423, 6941,
705–714, 2003.
67. Drout, R.J., Robison, L., Chen, Z., Islamoglu, T., Farha, O.K., Zirconium
Metal– Organic Frameworks for Organic Pollutant Adsorption. Trends
Chem., 1, 3, 304–317, 2019.
68. Burnett, B.J., Barron, P.M., Choe, W., Recent advances in porphyrinic metal–
organic frameworks: Materials design, synthetic strategies, and emerging
applications. CrystEngComm, 14, 11, 3839–3846, 2012.
69. Chen, D., Zhao, J., Zhang, P., Dai, S., Mechanochemical synthesis of metal–
organic frameworks. Polyhedron, 162, 59–64, 2019.
350  Applied Water Science Volume 2

70. Li, G., Ye, J., Fang, Q., Liu, F., Amide-based covalent organic frameworks
materials for efficient and recyclable removal of heavy metal lead (II). Chem.
Eng. J., 370, 822–830, 2019.
71. Masoomi, M.Y. and Morsali, A., Sonochemical synthesis of nanoplates of two
Cd(II) based metal–organic frameworks and their applications as p ­ recursors
for preparation of nano-materials. Ultrason. Sonochem., 28, 240–249, 2016.
72. Zhang, F., Zhang, T., Zou, X., Liang, X., Zhu, G., Qu, F., Electrochemical
synthesis of metal organic framework films with proton conductive property.
Solid State Ionics, 301, 125–132, 2017.
73. Efome, J.E., Rana, D., Matsuura, T., Lan, C.Q., Metal–organic frameworks
supported on nanofibers to remove heavy metals. J. Mater. Chem. A, 6, 10,
4550–4555, 2018.
74. Cheong, V.F. and Moh, P.Y., Recent advancement in metal–organic frame-
work: Synthesis, activation, functionalisation, and bulk production. Mater.
Sci. Technol., 34, 9, 1025–1045, 2018.
75. Gangu, K.K., Maddila, S., Mukkamala, S.B., Jonnalagadda, S.B., A review on
contemporary Metal–Organic Framework materials. Inorg. Chim. Acta, 446,
61–74, 2016.
76. Safaei, M., Foroughi, M.M., Ebrahimpoor, N., Jahani, S., Omidi, A., Khatami,
M., A review on metal–organic frameworks: Synthesis and applications.
TrAC Trends Anal. Chem., 118, 401–425, 2019.
77. Lu, M., Li, Y., He, P., Cong, J., Chen, D., Wang, J., Wu, Y., Xu, H., Gao, J., Yao,
J., Bimetallic metal–organic framework nanosheets as efficient electrocata-
lysts for oxygen evolution reaction. J. Solid State Chem., 272, 32–37, 2019.
78. Zheng, F., Xiang, D., Li, P., Zhang, Z., Du, C., Zhuang, Z., Li., X., Chen, W.,
Highly Conductive Bimetallic Ni–Fe Metal Organic Framework as a Novel
Electrocatalyst for Water Oxidation. ACS Sustain. Chem. Eng., 7, 11, 9743–
9749, 2019.
79. Zhu, Q.-L. and Xu, Q., Metal–organic framework composites. Chem. Soc.
Rev., 43, 16, 5468–5512, 2014.
80. Farha, O.K., Eryazici, I., Jeong, N.C., Hauser, B.G., Wilmer, C.E., Sarjeant,
A.A., Snurr, R.Q., Nguyen, S.T., Yazaydin, A.O., Hupp, J.T., Metal–Organic
Framework Materials with Ultrahigh Surface Areas: Is the Sky the Limit?
J. Am. Chem. Soc., 134, 36, 15016–15021, 2012.
81. Lv, S.-W., Liu, J.-M., Li, C.-Y., Zhao, N., Wang, Z.-H., Wang, S., A novel and
universal metal–organic frameworks sensing platform for selective detection
and efficient removal of heavy metal ions. Chem. Eng. J., 375, 122111, 2019.
82. Wang, C., Liu, X., Chen, J.P., Li, K., Superior removal of arsenic from water
with zirconium metal–organic framework UiO-66. Sci. Rep., 5, 16613, 2015.
83. Halder, S., Mondal, J., Ortega-Castro, J., Frontera, A., Roy, P., A Ni-based
MOF for selective detection and removal of Hg2+ in aqueous medium: A
facile strategy. Dalton Trans., 46, 6, 1943–1950, 2017.
84. Omichinski, J.G., Toward Methylmercury Bioremediation. Science, 317,
5835, 205, 2007.
Metal–Organic Frameworks for Heavy Metal Removal  351

85. Li, J., Li, X., Alsaedi, A., Hayat, T., Chen, C., Synthesis of highly porous inor-
ganic adsorbents derived from metal–organic frameworks and their appli-
cation in efficient elimination of mercury(II). J. Colloid Interface Sci., 517,
61–71, 2018.
86. Liang, L., Liu, L., Jiang, F., Liu, C., Yuan, D., Chen, Q., Wu, D., Jiang, H.-L.,
Hong, M., Incorporation of In2S3 Nanoparticles into a Metal–Organic
Framework for Ultrafast Removal of Hg from Water. Inorg. Chem., 57, 9,
4891–4897, 2018.
87. Yin, W.H., Xiong, Y.Y., Wu, H.Q., Tao, Y., Yang, L.X., Li, J.Q., Tong, X.L.,
Luo, F., Functionalizing a Metal–Organic Framework by a Photoassisted
Multicomponent Postsynthetic Modification Approach Showing Highly
Effective Hg(II) Removal. Inorg. Chem., 57, 15, 8722–8725, 2018.
88. Sun, D.T., Peng, L., Reeder, W.S., Moosavi, S.M., Tiana, D., Britt, D.K., Oveisi,
E., Queen, W.L., Rapid, Selective Heavy Metal Removal from Water by a
Metal–Organic Framework/Polydopamine Composite. ACS Cent. Sci., 4, 3,
349–356, 2018.
89. Leus, K., Folens, K., Nicomel, N.R., Perez, J.P.H., Filippousi, M., Meledina,
M., Dirtu, M.M., Turner, S., Tendeloo, G.V., Garcia, Y., Liang, G.D., Van Der
Voort, P., Removal of arsenic and mercury species from water by covalent
triazine framework encapsulated γ-Fe2O3 nanoparticles. J. Hazard. Mater.,
353, 312–319, 2018.
90 Ghaedi, A.M., Panahimehr, M., Nejad, A.R.S., Hosseini, S.J., Vafaei, A.,
Baneshi, M.M., Factorial experimental design for the optimization of highly
selective adsorption removal of lead and copper ions using metal organic
framework MOF-2 (Cd). J. Mol. Liq., 272, 15–26, 2018.
91. Yu, C., Shao, Z., Liu, L., Hou, H., Efficient and Selective Removal of
Copper(II) from Aqueous Solution by a Highly Stable Hydrogen-Bonded
Metal–Organic Framework. Cryst. Growth Des., 18, 5, 3082–3088, 2018.
92. Zheng, T.-T., Zhao, J., Fang, Z.-W., Li, M.-T., Sun, C.-Y., Li, X., Wang, X.-L.,
Su, Z.-M., A luminescent metal organic framework with high sensitivity for
detecting and removing copper ions from simulated biological fluids. Dalton
Trans., 46, 8, 2456–2461, 2017.
93. Rao, Z., Feng, K., Tang, B., Wu, P., Surface Decoration of Amino-
Functionalized Metal–Organic Framework/Graphene Oxide Composite
onto Polydopamine-Coated Membrane Substrate for Highly Efficient Heavy
Metal Removal. ACS Appl. Mater. Interfaces, 9, 3, 2594–2605, 2017.
94. Zou, Y.-H., Liang, J., He, C., Huang, Y.-B., Cao, R., A mesoporous cat-
ionic metal–organic framework with a high density of positive charge for
enhanced removal of dichromate from water. Dalton Trans., 48, 20, 6680–
6684, 2019.
95. Sathvika, T., Balaji, S., Chandra, M., Soni, A., Rajesh, V., Rajesh, N., A co-­
operative endeavor by nitrifying bacteria Nitrosomonas and Zirconium
based metal organic framework to remove hexavalent chromium. Chem.
Eng. J., 360, 879–889, 2019.
352  Applied Water Science Volume 2

96. Jamshidifard, S., Koushkbaghi, S., Hosseini, S., Rezaei, S., Karamipour, A.,
Jafari rad, A., Irani, M., Incorporation of UiO-66-NH2 MOF into the PAN/
chitosan nanofibers for adsorption and membrane filtration of Pb(II), Cd(II)
and Cr(VI) ions from aqueous solutions. J. Hazard. Mater., 368, 10–20, 2019.
97. Guo, J., Li, J.-J., Wang, C.-C., Adsorptive removal of Cr(VI) from simulated
wastewater in MOF BUC-17 ultrafine powder. J. Environ. Chem. Eng., 7, 1,
102909, 2019.
98. Zheng, Y., Li, Y., Huang, L., Nan, H., Wu, Y., A facile fabrication of MOF for
selective removal of chromium (III) from aqueous solution. J. Dispersion Sci.
Technol., 40, 6, 918–924, 2019.
99. Noraee, Z., Jafari, A., Ghaderpoori, M., Kamarehie, B., Ghaderpoury, A., Use
of metal–organic framework to remove chromium (VI) from aqueous solu-
tions. J. Environ. Health Sci. Eng., 2019.
100. Maleki, A., Hayati, B., Naghizadeh, M., Joo, S.W., Adsorption of hexavalent
chromium by metal organic frameworks from aqueous solution. J. Ind. Eng.
Chem., 28, 211–216.
101. Liu, C., Wang, P., Liu, X., Yi, X., Liu, D., Zhou, Z., Ultrafast Removal of
Cadmium(II) by Green Cyclodextrin Metal–Organic-Framework-Based
Nanoporous Carbon: Adsorption Mechanism and Application. Chem.—
Asian J., 14, 2, 261–268, 2019.
102. Wang, Y., Ye, G., Chen, H., Hu, X., Niu, Z., Ma, S., Functionalized metal–
organic framework as a new platform for efficient and selective removal of
Cadmium(ii) from aqueous solution. J. Mater. Chem. A, 3, 29, 15292–15298,
2015.
103. Roushani, M., Saedi, Z., Baghelani, Y.M., Removal of cadmium ions from
aqueous solutions using TMU-16-NH2 metal organic framework. Environ.
Nanotechnol. Monit. Manage., 7, 89–96, 2017.
104. Wu, Y., Ma, Y., Xu, G., Xia, T., Liu, W., Dong, Z., Yuan, Q., Zhang, C., Hu, Q.,
Synthesis of two novel H4TCPBDA-based metal–organic frameworks and
their application in lead ion adsorption. J. Mater. Sci., 54, 3, 2093–2101, 2019.
105. Karimi, M.A., Masrouri, H., Karami, H., Andishgar, S., Mirbagheri, M.A.,
Pourshamsi, T., Highly efficient removal of toxic lead ions from aqueous
solutions using a new magnetic metal–organic framework nanocomposite.
J. Chin. Chem. Soc., 0, 0, 2019.
106. Lu, M., Li, L., Shen, S., Chen, D., Han, W., Highly efficient removal of Pb2+
by a sandwich structure of metal–organic framework/GO composite with
enhanced stability. New J. Chem., 43, 2, 1032–1037, 2019.
107. Lei, C., Gao, J., Ren, W., Xie, Y., Abdalkarim, S.Y.H., Wang, S., Ni, Q., Yao,
J., Fabrication of metal–organic frameworks@cellulose aerogels composite
materials for removal of heavy metal ions in water. Carbohydr. Polym., 205,
35–41, 2019.
108. Li, G.-P., Zhang, K., Zhang, P.-F., Liu, W.-N., Tong, W.-Q., Hou, L., Wang,
Y.-Y., Tiol-Functionalized Pores via Post-Synthesis Modification in a
Metal–Organic Frameworks for Heavy Metal Removal  353

Metal–Organic Framework with Selective Removal of Hg(II) in Water. Inorg.


Chem., 58, 5, 3409–3415, 2019.
109. Fu, L., Wang, S., Lin, G., Zhang, L., Liu, Q., Zhou, H., Kang, C., Wan, S., Li, H.,
Wen, S., Post-modification of UiO-66-NH2 by resorcyl aldehyde for selective
removal of Pb(II) in aqueous media. J. Cleaner Prod., 229, 470–479, 2019.
110. Shi, Z., Xu, C., Guan, H., Li, L., Fan, L., Wang, Y., Liu, L., Meng, Q., Zhang, R.,
Magnetic metal organic frameworks (MOFs) composite for removal of lead
and malachite green in wastewater. Colloids Surf. A: Physicochem. Eng. Asp.,
539, 382–390, 2018.
111. Zhu, B.-J., Yu, X.-Y., Jia, Y., Peng, F.-M., Sun, B., Zhang, M.-Y., Luo, T., Liu,
J.-H., Huang, X.-J., Iron and 1,3,5-Benzenetricarboxylic Metal–Organic
Coordination Polymers Prepared by Solvothermal Method and Their
Application in Efficient As(V) Removal from Aqueous Solutions. J. Phys.
Chem. C, 116, 15, 8601–8607, 2012.
112. Xie, D., Ma, Y., Gu, Y., Zhou, H., Zhang, H., Wang, G., Zhnag, Y., Zhao, H.,
Bifunctional NH2-MIL-88(Fe) metal–organic framework nanooctahedra
for highly sensitive detection and efficient removal of arsenate in aqueous
media. J. Mater. Chem. A, 5, 45, 23794–23804, 2017.
113. Hou, S., Wu, Y.-n., Feng, L., Chen, W., Wang, Y., Morlay, C., Li, F., Green syn-
thesis and evaluation of an iron-based metal–organic framework MIL-88B
for efficient decontamination of arsenate from water. Dalton Trans., 47, 7,
2222–2231, 2018.
114. Audu, C., Nguyen, H.G., Chang, C.-Y., Katz, M., Mao, L., Farha, O., Hupp, J.,
Nguyen, S., The dual capture of As V and As III by UiO-66 and analogues.
Chem. Sci., 7, 6492–6498, 2016.
115. Yu, W., Luo, M., Yang, Y., Wu, H., Huang, W., Zeng, K., Luo, F., Metal–organic
framework (MOF) showing both ultrahigh As(V) and As(III) removal from
aqueous solution. J. Solid State Chem., 269, 264–270, 2019.
116. Mon, M., Lloret, F., Ferrando-Soria, J., Martí-Gastaldo, C., Armentano, D.,
Pardo, E., Selective and Efficient Removal of Mercury from Aqueous Media
with the Highly Flexible Arms of a BioMOF. Angew. Chem. Int. Ed., 55, 37,
11167–11172, 2016.
117. Rudd, N.D., Wang, H., Fuentes-Fernandez, E.M.A., Teat, S.J., Chen, F.,
Hall, G., Chabal, Y.J., Li, J., Highly Efficient Luminescent Metal–Organic
Framework for the Simultaneous Detection and Removal of Heavy Metals
from Water. ACS Appl. Mater. Interfaces, 8, 44, 30294–30303, 2016.
118. Yee, K.-K., Reimer, N., Liu, J., Cheng, S.-Y., Yiu, S.-M., Weber, J., Stock, N., Xu,
Z., Effective Mercury Sorption by Tiol-Laced Metal–Organic Frameworks:
in Strong Acid and the Vapor Phase. J. Am. Chem. Soc., 135, 21, 7795–7798,
2013.
119. Zhou, X.-P., Xu, Z., Zeller, M., Hunter, A.D., Reversible uptake of HgCl2 in a
porous coordination polymer based on the dual functions of carboxylate and
thioether. Chem. Commun., 36, 5439–5441, 2009.
354  Applied Water Science Volume 2

120. Li, J., Wang, X., Zhao, G., Chen, C., Chai, Z., Alsaedi, A., Hayat, T., Wang, X.,
Metal–organic framework-based materials: superior adsorbents for the cap-
ture of toxic and radioactive metal ions. Chem. Soc. Rev., 47, 7, 2322–2356,
2018.
121. Bagheri, S., Squitti, R., Haertlé, T., Siotto, M., Saboury, A., Role of Copper in
the Onset of Alzheimer’s Disease Compared to Other Metals. Front. Aging
Neurosci., 9, 446, 2018.
122. Purchase, R., The link between copper and Wilson’s disease. Sci. Prog., 96, 3,
213–223, 2013.
123. USA, E., Lead and Copper Rule, 2017, Link: https://www.epa.gov/dwreginfo/
lead-and-copper-rule#rule-summary.
124. Wang, K., Tian, Z., Yin, N., Significantly Enhancing Cu(II) Adsorption onto
Zr-MOFs through Novel Cross-Flow Disturbance of Ceramic Membrane.
Ind. Eng. Chem. Res., 57, 10, 3773–3780, 2018.
125. Luo, J., Gao, F., Kamasamudram, K., Currier, N., Peden, C.H.F., Yezerets, A.,
New insights into Cu/SSZ-13 SCR catalyst acidity. Part I: Nature of acidic
sites probed by NH3 titration. J. Catal., 348, 291–299, 2017.
126. Lv, Z., Fan, Q., Xie, Y., Chen, Z., Alsaedi, A., Hayat, T., Wang, X., Chen, C.,
MOFs-derived magnetic chestnut shell-like hollow sphere NiO/Ni@C com-
posites and their removal performance for arsenic(V). Chem. Eng. J., 362,
413–421, 2019.
127. Li, L.-L., Feng, X.-Q., Han, R.-P., Zang, S.-Q., Yang, G., Cr(VI) removal via
anion exchange on a silver-triazolate MOF. J. Hazard. Mater., 321, 622–628,
2017.
128. Ho, S.-H., Chen, Y.-d., Yang, Z.-k., Nagarajan, D., Chang, J.-S., Ren, N.-q.,
High-efficiency removal of lead from wastewater by biochar derived from
anaerobic digestion sludge. Bioresour. Technol., 246, 142–149, 2017.
129. Awual, M.R., An efficient composite material for selective lead(II) moni-
toring and removal from wastewater. J. Environ. Chem. Eng., 7, 3, 103087,
2019.
130. Zhang, J., Xiong, Z., Li, C., Wu, C., Exploring a thiol-functionalized MOF for
elimination of lead and cadmium from aqueous solution. J. Mol. Liq., 221,
43–50, 2016.
131. Huang, Y., Zeng, X., Guo, L., Lan, J., Zhang, L., Cao, D., Heavy metal ion
removal of wastewater by zeolite-imidazolate frameworks. Sep. Purif.
Technol., 194, 462–469, 2018.
132. Qin, Q., Wang, Q., Fu, D., Ma, J., An efficient approach for Pb(II) and Cd(II)
removal using manganese dioxide formed in situ. Chem. Eng. J., 172, 1,
68–74, 2011.
133. Hasankola, Z.S., Rahimi, R., Safarifard, V., Rapid and efcient ultrasonic-­
assisted removal of lead(II) in water using two copper- and zinc-based
metal–organic frameworks. Inorg. Chem. Commun., 107, 107474, 2019.
Metal–Organic Frameworks for Heavy Metal Removal  355

134. Abbasi, A., Moradpour, T., Van Hecke, K., A new 3D cobalt (II) metal–
organic framework nanostructure for heavy metal adsorption. Inorg. Chim.
Acta, 430, 261–267, 2015.
135. Chakraborty, A., Bhattacharyya, S., Hazra, A., Ghosh, A.C., Maji, T.K., Post-
synthetic metalation in an anionic MOF for efficient catalytic activity and
removal of heavy metal ions from aqueous solution. Chem. Commun., 52, 13,
2831–2834, 2016.
136. Rivera, J.M., Rincon, S., Youssef, C.B., Zepeda, A., Highly Efficient Adsorption
of Aqueous Pb(II) with Mesoporous Metal–Organic Framework-5: An
Equilibrium and Kinetic Study. J. Nanomater., 9, 2016.
137. Faust, S.D. and Aly, O.M., Chemistry of Water Treatment, Lewis Publisher,
Boca Raton, 1998.
138. Wang, Y., Dong, R., Zhou, Y., Luo, X., Characteristics of groundwater dis-
charge to river and related heavy metal transportation in a mountain min-
ing area of Dabaoshan, Southern China. Sc. Total Environ., 679, 346–358,
2019.
139. Sarkar, A. and Paul, B., The global menace of arsenic and its conventional
remediation—A critical review. Chemosphere, 158, 37–49, 2016.
140. Li, J., Wu, Y.-n., Li, Z., Zhang, B., Zhu, M., Hu, X., Zhang, Y., Li, F., Zeolitic
Imidazolate Framework-8 with High Efficiency in Trace Arsenate Adsorption
and Removal from Water. J. Phys. Chem. C, 118, 47, 27382–27387, 2014.
141. Massoudinejad, M., Ghaderpoori, M., Shahsavani, A., Jafari, A., Kamarehie,
B., Ghaderpoury, A., Amini, M.M., Ethylenediamine-functionalized cubic
ZIF-8 for arsenic adsorption from aqueous solution: Modeling, isotherms,
kinetics and thermodynamics. J. Mol. Liq., 255, 263–268, 2018.
142. Vu, T.A., Le, G.H., Dao, C.D., Dang, L.Q., Nguyen, K.T., Nguyen, Q.K., Dang,
P.T., Tran, H.T.K., Duong, Q.T., Nguyen, T.V., Lee, G.D., Arsenic removal
from aqueous solutions by adsorption using novel MIL-53(Fe) as a highly
efficient adsorbent. RSC Adv., 5, 7, 5261–5268, 2015.
143. Li, J., Wu, Y.-n., Li, Z., Zhu, M., Li, F., Characteristics of arsenate removal
from water by metal–organic frameworks (MOFs). Water Sci. Technol., 70, 8,
1391–1397, 2014.
144. Wu, H., Ma, M.-D., Gai, W.-Z., Yang, H., Zhou, J.-G., Cheng, Z., Xu, P., Deng,
Z.-Y., Arsenic removal from water by metal–organic framework MIL-88A
microrods. Environ. Sci. Pollut. Res., 25, 27, 27196–27202, 2018.
145. Atallah, H., Elcheikh Mahmoud, M., Jelle, A., Lough, A., Hmadeh, M., A
highly stable indium based metal organic framework for efficient arsenic
removal from water. Dalton Trans., 47, 3, 799–806, 2018.
146. Zhang, C., Xiao, Y., Qin, Y., Sun, Q., Zhang, S., A novel highly efficient
adsorbent {[Co4(L)2(μ3-OH)2(H2O)3(4,4′-bipy)2]·(H2O)2}n: Synthesis, crystal
structure, magnetic and arsenic (V) absorption capacity. J. Solid State Chem.,
261, 22–30, 2018.
356  Applied Water Science Volume 2

147. Gallegos-Garcia, M., Ramírez-Muñiz, K., Song, S., Arsenic Removal from
Water by Adsorption Using Iron Oxide Minerals as Adsorbents: A Review.
Miner. Process. Extr. Metall. Rev., 33, 5, 301–315, 2012.
148. Arbneshi, T., Rugova, M., Berisha, L., The level concentration of lead, cad-
mium, copper, zinc and phenols in the water river of Sitnica. J. Int. Environ.
Appl. Sci., 3, 2, 66–73, 2008.
149. Farooq, O., Ashizawa, A., Wright, S., Tucker, P., Jenkins, K., Ingerman, L.,
Rudisill, C., Toxicological Profile for Cadmium. Agency for Toxic Substances
and Disease Registry (US), Atlanta (GA), 2012.
150. Bernhof, R.A., Cadmium Toxicity and Treatment. Sci. World J., 1–7, 2013.
151. Zhang, H. and Reynolds, M., Cadmium exposure in living organisms: A
short review. Sci. Total Environ., 678, 761–767, 2019.
152. Rahimi, E. and Mohaghegh, N., Removal of Toxic Metal Ions from Sungun
Acid Rock Drainage Using Mordenite Zeolite, Graphene Nanosheets, and a
Novel Metal–Organic Framework. Mine Water Environ., 35, 1, 18–28, 2016.
153. Fang, Q.-R., Yuan, D.-Q., Sculley, J., Li, J.-R., Han, Z.-B., Zhou, H.-C.,
Functional Mesoporous Metal-Organic Frameworks for the Capture of
Heavy Metal Ions and Size-Selective Catalysis. Inorg. Chem., 49, 24, 11637–
11642, 2010.
154. Huo, J., Xu, L., Chen, X., Zhang, Y., Yang, J.E., Yuan, B., Fu, M., Direct epitax-
ial synthesis of magnetic Fe3O4@UiO-66 composite for efficient removal of
arsenate from water. Microporous and Mesoporous Mater, 276, 68–75, 2019.

You might also like