You are on page 1of 15

Chemosphere 285 (2021) 131432

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Recent advances in the application of water-stable metal-organic


frameworks: Adsorption and photocatalytic reduction of heavy metal
in water
Zhongwu Li a, Lei Wang a, Lei Qin b, *, Cui Lai b, **, Zhihong Wang b, Mi Zhou a, Linhui Xiao a,
Shiyu Liu b, Mingming Zhang b
a
College of Geographic Science, Hunan Normal University, Changsha, Hunan, 410081, PR China
b
College of Environmental Science and Engineering, Hunan University and Key Laboratory of Environmental Biology and Pollution Control (Hunan University), Ministry
of Education, Changsha, 410082, PR China

A R T I C L E I N F O A B S T R A C T

Handling Editor: X. Cao Heavy metals pollution in water is a global environmental issue, which has threatened the human health and
environment. Thus, it is important to remove them under practical water environment. In recent years, metal-
Keywords: organic frameworks (MOFs) with water-stable properties have attracted wide interest with regard to the cap­
Heavy metals ture of hazardous heavy metal ions in water. In this review, the synthesis strategy and postsynthesis modification
Water-stable MOFs
preparation methods are first summarized for water-stable MOFs (WMOFs), and then the recent advances on the
Adsorption
adsorption and photocatalytic reduction of heavy metal ions in water by WMOFs are reviewed. In contrast to the
Photocatalytic reduction
conventional adsorption materials, WMOFs not only have excellent adsorption properties, but also lead to
photocatalytic reduction of heavy metal ions. WMOFs have coupling and synergistic effects on the adsorption
and photocatalysis of heavy metal ions in water, which make it more effective in treating single pollutants or
different pollutants. In addition, by introducing appropriate functional groups into MOFs or synthesizing MOF-
based composites, the stability and ability to remove heavy metal ions of MOFs can be effectively enhanced.
Although WMOFs and WMOF-based composites have made great progress in removing heavy metal ions from
water, they still face many problems and challenges, and their application potential needs to be further improved
in future research. Finally, this review aims at promoting the development and practical application of heavy
metal ions removal in water by WMOFs.

1. Introduction endangering the human health (Fu and Xi, 2020; Zheng et al., 2020).
Due to the complex and changeable actual water body conditions, the
Heavy metals pollution in water is a global problem, which has purification of toxic metal ions remains a tough challenge. Thus, it is
attracted widespread attention (Fu and Wang, 2011; Hader et al., 2020). important to explore suitable methods for the removal of heavy metal
The most common heavy metal pollution in water includes cations (e.g., ions in actual water bodies.
Hg2+, Pb2+, Cd2+, and Cu2+) and oxygen anions (e.g., CrO2− 2−
4 /Cr2O7 , In the past decades of study on the treatment of heavy metal ion
SeO2−
3 /SeO 2−
4 and HAsO 2−
4 /AsO3−
4 ) (Dinda et al., 2013; Bora and Dutta, pollution in water, many treatment methods have been developed, such
2019; Humelnicu et al., 2020; Upadhyay et al., 2021). These heavy as adsorption (Wen et al., 2013; Cheng et al., 2019; Chai et al., 2021),
metal ions present in water are difficult to be degraded into environ­ photocatalysis (Hoch et al., 2008), biological treatments (Narayani and
mentally friendly substances by biodegradation, compared to organic Shetty, 2013), and membrane filtration (Gao et al., 2014), etc. It is worth
pollutants (Joseph et al., 2019). Gradually, heavy metals in water are noting that adsorption technology has many advantages in capturing
easily absorbed by organisms, then enter the food chain and accumulate heavy metal ions, such as high efficiency, simple design, and convenient
in organisms, thus causing toxic hazards to organisms and eventually operation, which endow it one of the most widely used practical

* Corresponding author. College of Environmental Science and Engineering, Hunan University, Changsha, Hunan, 410082, China.
** Corresponding author.
E-mail addresses: leiqin@hnu.edu.cn (L. Qin), laicui@hnu.edu.cn (C. Lai).

https://doi.org/10.1016/j.chemosphere.2021.131432
Received 5 May 2021; Received in revised form 23 June 2021; Accepted 1 July 2021
Available online 6 July 2021
0045-6535/© 2021 Elsevier Ltd. All rights reserved.
Z. Li et al. Chemosphere 285 (2021) 131432

technologies for heavy metal pollution remediation. Traditional properties and the synergistic effect of adsorption and photocatalysis.
adsorption materials include metal oxides (Tan et al., 2009), zeolite Therefore, this paper summarizes the post-synthesis modification
molecular sieves (Santasnachok et al., 2015; He et al., 2020), activated method to synthesize WMOFs, systematically reviews the applications of
carbon (Ren et al., 2013), carbon nanotubes (Li et al., 2014a), nano WMOFs and their composites in the adsorption removal and photo­
zero-valent iron (Zou et al., 2016), organic resin (Bhatti et al., 2014), etc. catalytic reduction of heavy metals in aqueous solution. Finally, the
Most of these traditional adsorbents are organic or inorganic porous synergistic effect of WMOFs in the removal of heavy metal ions from
materials, which have certain disadvantages in the process of absorbing water is discussed.
heavy metals in water (Kumar et al., 2017). For example, metal oxides
have slow adsorption kinetics and limited selectivity, activated carbon 2. Water stable metal¡organic frameworks (WMOFs)
has small pore sizes, and organic resin shows poor reusability (Li et al.,
2018). Therefore, it is necessary to explore more efficient porous ma­ 2.1. Synthesis strategy of WMOFs
terials to overcome the limitations of traditional adsorption materials.
Besides, a single adsorbent material is insufficient to target heavy According to the definition, WMOFs are categorized as MOFs that do
metals in water and results in efficient adsorption of heavy metals, not undergo structural damage and maintain functional stability when
because some heavy metals are highly toxic in the high-valence state and exposed to water (Wang et al., 2016a). To determine the water stability
low-toxic or non-toxic in the low-valence state (Feng et al., 2018; of MOFs, the degree of change in typical chemical characteristics of
Veerakumar and Lin, 2020). Hence, a matching and effective method is MOFs after exposure to aqueous environment was used as a criterion.
needed to remove heavy metal ions whose toxicity varies greatly with The factors affecting the water stability of MOFs can be broadly classi­
valence. Fortunately, photocatalytic technology can transform light fied into internal and external operational factors. Internal factors
energy into chemical energy to degrade and reduce pollutants in water, include metal ions, organic ligands, metal-organic coordination bond
which is green, efficient, and sustainable (Qin et al., 2021; Wang et al., strength, metal-ligand coordination geometry, pore surface hydropho­
2021). Photocatalytic reduction of such heavy metal ions is an efficient bicity, crystallinity and porosity; external factors include temperature,
way to reduce high-valence state and highly toxic heavy metal ions into pH, humidity, etc. (Canivet et al., 2014). Among them, the strength of
low-valence state and low-toxic or non-toxic heavy metal ions. There­ metal-organic coordination bonds (thermodynamic stability), hydro­
fore, the combination of adsorption and photocatalysis technology is a phobicity of the pore surface and spatial potential resistance (kinetic
good idea to remove these heavy metal ions and reduce the toxicity of stability) are important factors affecting the water stability of MOFs
them. In this respect, metal-organic frameworks (MOFs) are ideal ma­ (Burtch et al., 2014). Therefore, in order to prevent the structure of
terials for heavy metal adsorption and photocatalytic reduction, and MOFs from being destroyed by water molecules, it is necessary to fully
have the potential to become dual-functional materials. understand the interaction between water molecules and MOFs (Liu
MOFs is one kind of organic-inorganic hybrid material formed by et al., 2020a).
self-assembly of organic ligands and metal ions or clusters through co­ In an aqueous solution, the weaker metal-organic coordination
ordination bonds (Li et al., 1999; Liu et al., 2020b). It has the unique bonds are easily attacked by water molecules, and the organic ligands
characteristics of porosity, large specific surface area, and open metal are replaced by water molecules or hydroxide, destroying the structure
sites. MOFs enables structural and functional diversity through changing of MOFs (Wang et al., 2016a). Thus, enhancing the strength of
metal center and organic ligand (Furukawa et al., 2010; Wang et al., metal-organic coordination bonds can enhance the water stability of
2020d). These features, coupled with the ordered porous structure and MOFs (Yuan et al., 2018). Under the same coordination environment,
modifiable properties of MOFs, make it a great advantage in removing metal-organic coordination bonds formed by high-valent metal ions
heavy metal ions from water (Wang et al., 2016b; Lai et al., 2021). In with high charge density are stronger (Yuan et al., 2018). According to
addition, MOFs is photoresponsive and can absorb light through organic Pearson’s hard/soft acid/base (HSAB) theory, the strong interaction
linkers or metal centers. As a photocatalyst, MOFs has the advantages of between hard acid and hard base or soft acid and soft base can form
good topology and high specific surface area, which can promote the strong coordination bonds, thereby building a stronger framework
rapid translocation and good adjustment of target molecules (Zhang (Miranda-Quintana, 2017). Usually, the high-valent metal units of MOFs
et al., 2020e; Fu et al., 2020). It has shown great potential in the form a larger rigid structure, allowing the metal center to resist the
application of photocatalytic reduction of heavy metal ions. However, attack of water molecules to a certain extent (Qadir et al., 2015; Ding
the water stability of MOFs is an important factor for their practical et al., 2019). Hence, WMOFs can be constructed using carboxylate-based
application in the treatment of heavy metal pollution in water (Wang ligands (hard Lewis bases) and high-valent metal ions (hard Lewis acids)
et al., 2016a; Zhang et al., 2020c). The presence of water in the reaction (Yuan et al., 2018). For example, Ferey and his colleagues have devel­
environment of water sensitive materials will have harmful effects on oped Material Institut Lavoisier (MIL) series: MIL-53(AI), MIL-100(Fe)
the materials, and even lead to the structural degradation of the mate­ (Ferey et al., 2004), MIL-101(Cr) (Hupp, 2005), etc., which are chemi­
rials. Water-stable MOFs (WMOFs) is able to maintain structural and cally stable in various solvents for months (Liu et al., 2020a). In addi­
functional stability in the aqueous environment to ensure that its reac­ tion, there is the University of Oslo (UiO) series, based on the Zr4+
tion of removing heavy metal ions from water can proceed smoothly synthesis of UiO-66, which remains stable under acidic conditions
(Wang et al., 2016a). Due to the unique characteristics and advantages (Cavka et al., 2008). Soft azolate ligands are regarded as soft Lewis
of WMOFs, there have been many studies on WOMFs adsorption/pho­ bases, and soft divalent metal ions are regarded as soft Lewis acids.
tocatalytic reduction of toxic heavy metals in wastewater. WMOFs can also be constructed using carboxylate ligands and
In previous studies, several reviews have focused on the synthesis high-valent metal ions (Park et al., 2006). The imidazoline zeolite
strategies and applications of WMOFs (Wang et al., 2016a; Feng et al., frameworks (ZIFs) in this type of structure are the most representative,
2018; Yuan et al., 2018; Liu et al., 2020a; Zhang et al., 2020c). Among which use Zn2+ and imidazolate linkers to construct stable crystals
them, Liu et al. summarized the construction strategy of WMOF and similar to the zeolite topology (Demessence et al., 2009).
introduced the synthesis and application of representative WMOFs (Liu According to the HSAB theory, the strengthening of metal-organic
et al., 2020a). Wang et al. discussed the application of WMOFs in the coordination bonds determines the thermodynamic stability, making
fields of adsorption, catalysis, sensing, membrane separation, and pro­ the metal center and organic ligands be less susceptible to attacking by
ton conduction. A comprehensive discussion of the studies related to water molecules. Besides, kinetic factors are also considered. The link
moisture content in these fields was presented by Wang et al. (2016a). length of the metal-organic coordination bond is negatively related to
So far, there is no article systematically discussing the WMOFs adsorp­ the stability of the framework (Yuan et al., 2018). The increase of the
tion and photocatalytic reduction of heavy metals with different bond length will lead to the decrease of stability, which is caused by an

2
Z. Li et al. Chemosphere 285 (2021) 131432

increase in the length of the coordination bond affecting its rigidity and (Nguyen and Cohen, 2012). It was first proposed by Hoskins and Robson
thus its dynamic stability (Cavka et al., 2008). In addition, it prevents when they reported on metallo-cyanides that functionalized rod-like
water molecules from approaching and attacking by functionalizing substituents could be introduced into the framework (Hoskins and
enhanced steric hindrance and designs hydrophobic surfaces or adding Robson, 1990). Until 2007, when Cohen’s group synthesized the
an additional layer of protection from metal ions (Fig. 1) (Yang and amino-functionalized IRMOF-3 and realized the post-synthetic modifi­
Park, 2012) cation of NH2, the post-synthetic modification approach was increas­
ingly used in MOFs (Batten et al., 2013). The synthesis of functional
MOFs by PSM has the following advantages: (1) When the target func­
2.2. Enhancement of water stability of MOFs by postsynthesis tional group is incompatible with the synthesis conditions, the ligand
modification (PSM) can be pre-functionalized to construct MOFs and then the target func­
tional group will be generated to synthesize the MOF material with
Due to the characteristics of MOF materials and PSM methods, the desired function; (2) MOFs are highly porous materials with uniform
PSM methods are particularly suitable for use with MOFs, which can pores, which can be modified not only by the outer surface of the ma­
improve the water stability and expand application range of MOFs, and terial, but also by the inner surface of the pores; (3) The topology of
also can save time and cost (Mandal et al., 2020). The concept of PSM is MOFs is the same before and after modification, because PSM only reacts
the introduction of target functional group into an already synthesized on the skeleton of MOFs and does not change the original skeleton
material through coordination or functional group conversion reactions

Fig. 1. Mechanisms and experimental methods for improving the water stability in existing MOFs. Reproduced with permission (Burtch et al., 2014). Copyright
2014, American Chemical Society.

3
Z. Li et al. Chemosphere 285 (2021) 131432

structure of them. Therefore, the water stability of MOFs can be surface with a hydrophobic amorphous-carbon could prevent the hy­
enhanced by hydrophobization, ligand functionalization, and metal drolysis of MOFs to a certain extent. Liu et al. achieved hydro­
clustering modifications using PSM (Joyaramulu et al., 2019). phobization on the external surface of ZIF-8 through a shell-ligand
exchange reaction. ZIF-8 hydrolysis was still observed in excessive
2.2.1. Hydrophobization water, although it was considered as a stable MOF (Liu et al., 2013).
Enhancement of hydrophobicity is the most direct way to enhance However, the 2-methylimidazole ligand in the outermost layer of ZIF-8
the water stability of MOFs, and the use of PSM to enhance the hydro­ is replaced with hydrophobic 5,6-dimethylbenzimidazole (DMBIM)
phobicity of MOFs has received much attention. The internal and through a ligand exchange reaction. The water contact angle of ZIF-8
external surfaces of MOFs can be hydrophobized using the PSM method. increased from 60◦ to 121◦ , which indicates the enhancement of hy­
The hydrophobicity of the internal pores of MOFs is internal surface drophobicity and water stability of ZIF-8-DMBIM. This is because
hydrophobicity, which can be judged through detecting the crystal DMBIM ligands are hydrophobic and produce steric hindrance. Some
structure of MOFs. The hydrophobic surfaces of MOFs crystals are studies have shown that MOFs can be hydrophobized by coating some
external surface hydrophobic, which can be judged by measuring the hydrophobic polymers. For instance, Zhang et al. reported a method for
water contact angle. preparing MOFs with polydimethylsiloxane (PDMS) -coating. It was
achieved by thermal degradation of PDMS to deposit volatile silicon
(1) Internal surface hydrophobization molecules and then cross-linking them to the surface of MOF (Fig. 2)
(Zhang et al., 2014). The water contact angle of MOF-5, HKUST-1, and
The internal surface of MOFs can be hydrophobized through the PSM ZnBT are all close to 0◦ . After PDMS modification, their water tentacles
of the ligands. In 2008, Cohen et al. performed postsynthetic hydro­ increase to 128◦ , 130◦ and 130◦ , respectively. The BET surface area of
phobization of IRMOF-3 ([Zn4O(1,4-bdc-NH2)3]) and MIL-53(Al)–NH2 HKUST-1 coated with PDMS remains essentially unchanged after being
([Al(OH)(1,4-bdc-NH2)]) (Cohen and Cohen, 2010). The PSM reaction soaked in water for 3 days. The PDMS coating can effectively prevent
of amine groups and alkylic anhydride generated amide substituents, water molecules from attacking the fragile coordination bonds, thereby
which introduced a bulky or long hydrophobic alkyl group on its pore ensuring the stability of the MOF.
surface. The integration of long or bulky alkyl groups within IRMOF-3
transformed it into a hydrophobic material, and water contact angle 2.2.2. Ligand functionalization and metallic cluster modification
increased from 0◦ to 105◦ –125◦ . Similarly, the water contact angles of The hydrophobic and spatial properties of metallic ligands have been
MIL-53(Al)–NH2 modified by alkyl anhydrides became very large. Its greatly improved by the functionalization of organic ligands. The water
crystallite size and the combination of hydrophobic functional groups absorption isotherm can also be used to determine whether the
are the main reasons for the superhydrophobicity. In addition, Rubin enhanced water stability is due to the strengthened framework stability
et al. modulated the hydrophobicity of [Cu3(NH2BTC)2] by the reaction or increased hydrophobicity (Burtch et al., 2014). In addition, the water
of alkyl anhydrides with amine groups to form amides (Rubin and stability of pillared MOFs can be enhanced by properly utilizing and
Reynolds, 2017). Hence, the water contact angle of Cu3(NH2BTC)2 controlling the catenation (Jiang et al., 2013). For example, Lee et al.
samples with long alkyl chains was greatly improved. reported a novel microporous metal organic framework JCM-1 ([Co
The coordination of new ligands with metal centers via PSM is also a (PzIm)(NO3)]) that possessed the pyrazolate metal binding site and
way to improve the hydrophobicity of MOFs. Bae and his colleagues imidazolium groups (Lee et al., 2018). The structure of JCM-1 is basi­
modified Ni-MOF-74 with pyridine molecules by the dative PMS cally unchanged after being soaked in water for 2 months. The strong
method. The pyridine molecules occupy open metal sites, resulting in water stability of JCM-1 was due to the methyl groups in the ligand and
hydrophobization of the hydrophilic Ni-MOF-74 internal surface (Bae the strong metal-pyrazolate bond.
et al., 2014). Drache et al. exchanged fluorinated monocarboxylates to Thermodynamically stable MOFs do not collapse after chronically
monocarboxylate ligand of DUT-67([Zr6O6(OH)2(tdc)4(CH3COO)2]) by exposing in an aqueous environment. Inert metal cluster is the key
the PSM method (Drache et al., 2016). The experimental results showed structural property of thermodynamically stable MOFs, which prevents
that the exchange of ligands achieved internal hydrophobicity of irreversible hydrolysis reactions under high water loads, and the metal
DUT-67. It was also discovered that water stability was improved by clusters of MOFs can be replaced by inert metals through cation ex­
PSM-exchange of DUT-67. change (Burtch et al., 2014). For example, Li et al. synthesized MOF-5s
The hydrophobicity of MOFs is regulated by covalent bonding be­ doped with Ni ions to improve their water stability (Li et al., 2012). The
tween the active substance and the inorganic SBUs. Decoste et al. re­ new NixZn4-xO6+ secondary units were more stable than the Zn4O6+
ported that the PSM of Cu-BTC (1,3,5 benzenetricarboxylic acid, BTC) units, therefore, the water resistance of the framework was significantly
was obtained by a plasma-enhanced chemical vapor deposition of per­ improved by the Ni-doped MOF-5s. In addition, ZIF-8 (Li et al., 2014b),
fluorohexane (Decoste et al., 2012). Monte Carlo simulation revealed MIL-100(Cr) (Hall and Bollini, 2020), and Ni3(BTP)2 (Shearer et al.,
that the aggregation of water clusters was hindered by fluorohexane, 2013) all have strongly bound ligands, or in the absence of strongly
thereby preventing the hydrolysis of Cu-BTC. Ding et al. polymerized 1, bound ligands, they all have highly inert metallic species (Cr3+), in order
2-divinylbenzene in the internal channel of MOF-5 by the PSM method, to maintain good water stability.
which divided the channels of MOFs to form closed hydrophobic com­ The synergistic effect of ligand functionalization and metal cluster
partments. The surface of PN@MOF-5 exhibited hydrophobicity, and modification also enables the synthesis of more stable MOFs. Zhang
the water contact angle increased from 0◦ to 135◦ (Ding et al., 2016). et al. designed new Zr(IV)-MOFs using hexacarboxylate ligands as
functional group (Zhang et al., 2018b). The modification of metal
(2) External surface hydrophobization clusters using different functional groups such as HCOO− , PhCOO− ,
H2O/OH, and CH3COO− could adjust their water absorption properties.
The hydrophobicity and stability of MOFs can also be improved by All four MOFs exhibited good regeneration and water adsorption cycling
modifying their external surface. Modification of the outer surface of properties under mild conditions. Thus, the synergistic effect can
MOFs can be done using inorganic materials, hydrophobic small mole­ enhance the stability of MOFs while optimizing other properties.
cules/ligands and organic polymers. Yang and Park prepared
amorphous-carbon coated MOF-5 in a N2 atmosphere at 480–530 ◦ C.
MOF-5 coated with amorphous carbon could still maintain a stable
structure after being soaked in aqueous solution for 2 h (Yang and Park,
2012). All the findings indicated that the coating of MOF-5 crystal

4
Z. Li et al. Chemosphere 285 (2021) 131432

Fig. 2. Coating PDMS on the surface of MOFs and improving the moisture/water resistance of MOFs. Reproduced with permission (Zhang et al., 2014). Copyright
2014, American Chemical Society.

3. Heavy metals adsorption by WMOFs functionalized MOF([Zn(hip)(L)]⋅(DMF)(H2O)) to remove Hg2+ from


aqueous solutions (Luo et al., 2015). [Zn(hip)(L)]⋅(DMF)(H2O) had 1-D
3.1. Cation adsorption hexagonal channels with a pore size of 5.4 Å. The pore walls were
decorated with a large number of acylamines and hydroxyl groups
3.1.1. Hg2+ (Fig. S1). Data from EDS and FT-IR indicated that Hg2+ was inserted in
Mercury is widespread in nature and human society and affects the form of HgCl2; Hg2+ species had strong interactions with imides and
human health due to its high toxicity. Organic mercury is more toxic hydroxyl functional groups. In 2019, Wu et al. synthesized
than inorganic mercury. Microorganisms in soil and water induce the UiO-66-EDTA multifunctional adsorbent with octahedral structure and
methylate inorganic mercury to form methylmercury and eventually amine and carboxyl groups, which remained stable under acidic con­
accumulate in animals and humans through the food chain (Rahman and ditions at pH 1 (Wu et al., 2019). Notably, UiO-66-EDTA achieved a
Singh, 2019; Wang et al., 2020c). High concentrations of mercury can removal rate of more than 99% for 11 metal ions in the coexisting sys­
damage the cells, inhibit the enzyme activity and even threaten the tem. The results of XPS analysis showed that Hg2+ coordination with
nervous and immune systems (Ali et al., 2019). In recent years, the carboxyl groups and tertiary amines was the reason for its high removal
adsorption and removal of mercury ions (Hg2+) from water by various efficiency.
WMOFs have been studied, and WMOFs have been found to have great Recently, Tian et al. prepared an adaptive MOF FJI-H30 using a
potential as effective adsorbents. flexible TPMA[tris(pyridine- 4-ylmethyl)amine] ligand and Co(SCN)2
Sulfur functionalization or sulfide modification of MOFs can effec­ (Tian et al., 2020). FJI-H30 had anti-interference ability and could resist
tively and selectively facilitate the adsorption and removal of Hg2+ various interference ions. Most interestingly, the Hg2+ ions adsorbed by
because Hg2+ can bond with S and form strong affinity with it (Lai et al., FJI-H30 would bind to the S atom of the SCN group, leading to the in-situ
2018; Qin et al., 2018). In 2014, Liu et al. synthesized alkenyl com­ deformation of FJIH-30 to form an extended framework (FJI-H30-Hg).
pounds functionalized MOFs by PSM method. Cr-MIL-101-AS with This adaptive deformation decreased the potential repulsion between
alkenyl side chains and thiol side chains had excellent adsorption per­ adsorbed Hg2+, which further promoted the adsorption of Hg2+ by
formance of Hg2+ from water and had good water stability (Liu et al., FJIH-30. Therefore, FJI-H30 with in situ adaptive deformation perfor­
2014). In 2016, Mon et al. reported a newly synthesized Bio-MOF with a mance was a promising and efficient heavy metal ion adsorption ma­
honeycomb structure for mercury ions adsorption. The alkyl thiol chains terial. Compared with previous studies, the development of MOFs is not
functionalized MOF not only have a special adsorption capacity for only about the enhancement of the adsorption capacity of Hg2+, but also
mercury ions, but more importantly, it can be reused by using dimethyl the potential of MOFs to be applied in the actual environment. Finally, a
sulfide to reverse the adsorption process (Mon et al., 2016). In 2019, Li summary of WMOFs that have been used in recent years for Hg2+
et al. prepared NH2-MIL-68(In) to more clearly reflect the function of the removal from water is presented in Table 1.
thioglycolic acid group. The exposed –NH2 group in the NH2-MIL-68(In)
channel provided accessible sites for surface functionalization by bind­ 3.1.2. Pb2+
ing to the aldehyde/carboxylic acid. Generation of aramid groups from Lead is generally available in both inorganic and organic forms, and
thioglycolic acid was fixed on the channel wall of NH2-MIL-68(In) by divalent lead compounds include water-soluble and insoluble salts, such
PSM method. Isothermal adsorption experiments revealed that the as lead acetate and lead oxide. In humans and animals, lead exposure
introduction of mercaptoacetic acid groups made a key contribution to and ingestion may have significant neurological, developmental, im­
the improved removal of Hg2+ from SH-MIL-68(In) (Li et al., 2019). mune, and renal health effects. Through biochemical processes, lead
Hg2+ not only binds strongly to sulfur, but also to carboxyl, acyla­ inhibits or mimics the behavior of calcium and protein interactions to
mide, and hydroxyl groups. Therefore, the modification of MOFs with exert lead toxicity, and Pb2+ emission is a serious problem and its
these groups can greatly improve the adsorption capacity of MOFs to removal from water is critical to the human health (Shen et al., 2019;
Hg2+. In 2015, Luo et al. reported an acylamine and hydroxyl- Kumar et al., 2020).

5
Z. Li et al. Chemosphere 285 (2021) 131432

Table 1 MOFs with sulfur functionalization can also improve the adsorption
Adsorption data of Hg2+ by typical WMOFs in aqueous systems. of Pb2+ from aqueous solutions. In 2016, Zhang et al. prepared thiol-
Type of MOF Adsorption Equilibrium Optimal Ref. functionalized HS-mSi@MOF-5 and thiol-modified MOF-5 using silica
capacity time (min) pH gel layer as the Pb2+ adsorption medium (Zhang et al., 2016a). The
(mg g− 1) or n-value in the adsorption kinetics of HS-mSi@MOF-5 was larger than 1,
adsorption
which was much larger than the n value of the unmodified material,
rate
indicating that the thiol modification could significantly improve the
Cr-MIL-101-AS 99%(20 mg, 300 – Liu et al. adsorption capacity. The water stability test exhibited that the stability
10 ppm) (2014)
MIL-101-Thymine 52 200 6 Luo et al.
of the modified material was greatly improved compared to the original
(2016) sample.
In2S3/MIL-101 518 260 7 Liang et al. In addition, there are some MOFs composites with high adsorption
(2018) properties to Pb2+ while maintaining water stability. Zhao et al. syn­
SH-MIL-68(In) 450 360 4 Li et al.
thesized a highly loaded UiO-66-NH2-PAM-PET (PAM, polyacrylamide;
(2019)
UiO-66-NHC(S) 769 240 – Saleem et al. PET, polyethylene terephthalate) composite by growing MOFs on a
NHMe (2016) modified PET carrier, which was able to efficiently remove Pb2+ from
UiO-66-EDTA 371.6 30 7 Wu et al. water (Zhao et al., 2020a). The composites contained many –COOH,
(2019) –NH2, and –CO–NH2 groups for rapid capture of Pb2+ by ion exchange
UiO-66-DMTD 670.5 600 3 Fu et al.
(2019)
and electrostatic surface complexation. In general, the composites had
AC-UiO-66 205 1440 7 Solis et al. the advantages of simple synthesis, strong adsorption capacity for Pb2+,
(2020) which had a wide application prospect in practical water treatment.
Zr-DMBD 171.5 10 6 Ding et al. Finally, a summary of WMOFs that have been used in recent years for
(2018)
Pb2+ removal from water is presented in Table 2.
Zr-MAS 734 50 0–7 Yang et al.
(2019)
DUT-67 (Zr) 90%(10 mg, 60 6 Chen et al. 3.1.3. Cd2+
20 μg L− 1) (2018) Cadmium is used in a variety of industrial activities, including the
MOF-MA 1080 10 4 Wang et al. production of alloys, plastics and batteries, resulting in steadily
(2020b)
increasing concentrations of cadmium in the environment (Zhao et al.,
AMOF-1′ 78 1440 – Chakraborty
et al. (2016) 2020b). As a heavy metal, cadmium is a huge hazard to human health.
3D Co(II) MOF 70%(200 100 6 Abbasi et al. Most of the cadmium that enters the human body will accumulate in the
mg, 10 (2015) kidneys, liver and bones, causing various diseases such as anemia, high
ppm)
blood pressure, and nephritis (Dukic-Cosic et al., 2020).
2D-NCS 1698 270 4/9 Wang et al.
(2020a) Sulfonic acid functionalized MOF has a good adsorption and removal
LMOF-263 380 30 – Rudd et al. effect on Cd2+ in water, especially the selective adsorption capacity of
(2016) Cd2+. In 2015, Wang et al. functionalized HKUST-1 (Cu3(BTC)2) with
MFC-S 282 30 3 Huang et al. sulfonic acid and immobilized –SO3H on the surface of HKUST-1 by two
(2016)
post-synthetic modifications for the adsorption of Cd2+ (Wang et al.,
Fe-BTC/PDA 1634 1 7 Sun et al.
(2018) 2015b). Firstly, dithioglycol units were introduced into HKUST-1 and
MTV-MOF 99%(20 3 7 Mon et al. then converted to sulfonic acid by oxidation and acid treatment. The
mg, 1 ppm) (2019) chelation between Cd2+ and –SO3H group made Cu3(BTC)2-SO3H had a
FJI-H30 705 60 7 Tian et al.
fast-kinetic adsorption capacity of Cd2+. In 2016, Moradi et al. first
(2020)
Fe3O4@SiO2HKUST- 264 10 3 Huang et al.
functionalized water-stable MOF-235(Fe) by sulfonic acid and com­
1 (2015) plexed it with iron oxide nanoparticles to form a magnetic nanosorbent
Zn(hip)(L)⋅(DMF) 278 30 5 Luo et al. (Fe3O4@MOF-235(Fe)–OSO3H), which not only had high stability, low
(H2O) (2015) toxicity, and good reusability, but also could extract Cd2+ from water by
0.5HgCl2@CuII4[(S, 99.95%(50 15 Mon et al.
magnetic solid phase extraction (Moradi et al., 2016).

S)-methox]2}⋅ mg, 10 (2017)
5H2O ppm) In a stable zinc-based MOF, Ahmadijokani et al. synthesized
{[Ni1.5(L)(NH2- 93.69 300 3 Shao et al. ethylenediamine-functionalized UiO-66-EDA and investigated its ability
bpy)-(H2O)]⋅ (2018) to adsorb Cd2+ in water (Ahmadijokani et al., 2020). The adsorption of
7.5H2O}n
Cd2+ on UiO-66-EDA was consistent with the pseudo-second-order ki­
netic model and the Langmuir isotherm model, indicating that the
The amino functionalization of MOFs can effectively improve the adsorption was chemisorption and monolayer adsorption. The strong
adsorption performance of Pb2+ adsorption. In 2015, Ricco et al. first adsorption of Cd2+ by UiO-66-EDA was mainly due to the interaction
reported an amino-functionalized MIL-53(Al) water-resistant magnetic between metal ions, and the abundant functional groups on its surface.
composite for Pb2+ adsorption in aqueous samples by adjusting the NH2 The removal of Cd2+ by UiO-66-EDA could be improved by increasing
group concentration (Ricco et al., 2015). The magnetic frame composite the reaction temperature.
(MFC) reacts rapidly to external magnetic force, so it is easier to recycle. In addition, Esrafili investigated the adsorption of heavy metal ions
In 2018, Yin et al. synthesized novel melamine-modified MOFs to adsorb by MOFs with diverse functional groups in the pillar structure (Esrafili
Pb2+ in low-salt solutions (Yin et al., 2018). The calculations showed et al., 2018). Among them, the adsorption efficiency of MOFs with
that there was charge transfer and orbital transition between melamine amide functional group was higher than that of MOFs with imine
and MOFs. Furthermore, MOFs containing N atoms in the ligand showed functional group for metal ions. This was due to the stronger alkalinity of
better adsorption capacity for Pb2+. In 2019, Zhang et al. reported a the amide-biased amine group, which leads to a stronger interaction of
WMOFs (CAU-7-TATB) with ligands containing N to provide binding the Cd2+ with the MOFs coordination sites. Moreover, the electronic
sites for metal ions (Zhang et al., 2019a). CAU-7-TATB remained highly structure of the ligands of MOF is the main reason for its different
selective adsorption towards Pb2+ in the presence of multiple interfering adsorption modes. TMU-23 had suitable functional groups, high
ions. porosity and water stability. DFT calculations showed that the main
reason for the adsorption of Cd2+ by MOFs was the electrostatic

6
Z. Li et al. Chemosphere 285 (2021) 131432

Table 2
Adsorption data of Pb2+ by typical WMOFs in aqueous systems.
Type of MOF Adsorption capacity (mg g− 1) Equilibrium time (min) Optimal pH Ref.

HKUST-1-MW@H3PW12O40 98.18 120 7 Zou et al. (2013)


TMU-5 251 5 10 Tahmasebi et al. (2015)
MIL-53(Al) MFC 492.4 360 – Ricco et al. (2015)
MIL-101(Fe)/GO 128.6 15 6 Lu et al. (2019)
ZIF-8-0 1321.21 120 5 Song et al. (2019)
UiO-66-NH2 1795.3 40–45 4.5 Yin et al. (2016)
UiO-66-NHC(S)NHMe 232 240 – Saleem et al. (2016)
Melamine-modified Zr-MOFs 122 120 5 Yin et al. (2018)
UiO-66-NH2@CA 89.4 3660 – Lei et al. (2019)
UiO-66-EDA 243.9 360 7 Song et al. (2019)
UiO-66-NH2-PAM-PET 711.99 240 5 Zhao et al. (2020a)
Zn(Bim)(OAc)-NS 253.8 60 5.5 Xu et al. (2020)
MOF-MA 510 960 4 Wang et al. (2020b)
Dy-MOF 3.05 10 6.5 Jamali et al. (2016)
HS-mSi@MOF-5 312 30 6 Zhang et al. (2016a)
AMOF-1′ 71 1440 – Chakraborty et al. (2016)
Fe-BTC/PDA 394 1 7 Sun et al. (2018)
CAU-7-TATB 63 45 5 Zhang et al. (2019a)

interaction between the amine group of TMU-23 and metal ions. Finally, et al., 2016b). In the pH range of 3–6, ZIF-8 can effectively adsorb both
a summary of WMOFs that have been used in recent years for Cd2+ high and low concentrations of Cu2+. In aqueous solutions with low
removal from water is presented in Table 3. concentrations of Cu2+, ZIF-8 adsorbs Cu2+ by ion exchange. At high
Cu2+ concentrations, the main reason for adsorption is the coordination
3.1.4. Cu2+ of Cu2+ with the nitrogen group on 2-methylimidazole. In 2019, Wei
Copper ion is a heavy metal element required in humans and ani­ et al. loaded ZIF-8 onto graphene oxide (GO) and then modified the
mals, mainly in the two oxidation states of Cu+ and Cu2+. However, the composite material using 3-amino amine. ZIF-8 complexed with GO
cycling of copper between oxidized and reduced copper may generate could improve the adsorption capacity of MOFs and eliminate the
hydroxyl radicals and superoxide, resulting in its potential toxicity. In agglomeration of MOFs to a certain extent (Wei et al., 2019). Finally, a
addition, excessive exposure and ingestion of copper have damaging summary of WMOFs that have been used in recent years for Cu2+
effects on the stomach, liver and kidneys, with a narrow concentration removal from water is presented in Table 4.
range of beneficial and harmful effects (Kumar et al., 2021). Thus,
developing efficient method to remove excessive copper is very impor­ 3.2. Anion adsorption
tant and MOFs-based adsorption technology is efficient on this.
In 2015, Tahmasebi et al. reported three MOFs modified with azine 3.2.1. CrO2−4 and Cr2O7
2-

and imine groups, which were used to remove Cu2+ in water (Tahmasebi Chromium is a naturally occurring element whose valence states
et al., 2015). The azine and imine groups had easily accessible Lewis range from trivalent to hexavalent chromium. Among them, trivalent
base sites, which facilitated the adsorption of heavy metal ions. In chromium is more stable and is always found in natural ores. The
TMU-6, two nitrogen atoms are in resonance with the benzene ring, toxicity of chromium and its compounds depends mainly on the oxida­
which reduce the charge density of TMU-6. In addition, because of the tion state and solubility. The toxicity of hexavalent chromium is far
presence of a space between the two nitrogen atoms, they are less stronger than that of trivalent chromium, and it is a highly toxic
accessible for metal ions than the nitrogen atoms in the azine groups of carcinogen. Hexavalent chromium and its compounds are released into
TMU-5, and so, they have less interaction with metal ions. Therefore, the the environment mainly as oxygen anions (CrO2− 2−
4 , Cr2O7 ) from in­
azine group was more alkaline than the imine group, and the adsorption dustrial activities such as welding, dyes, and leather tanning (Pavesi and
efficiency of TMU-5 and TMU-4 for metal ions was almost the same, Moreira, 2020).
while that of TMU-6 was lower. Among them, TMU-5 has better The exchange of anions into cation hosts is a common anion capture
adsorption capacity. method, and its capture capacity is limited by the process balance. A new
In addition, zinc-based MOFs have a good effect on the adsorption method for selective capture of anions was first proposed by Fei et al.
removal of Cu2+ from water. Without any pretreatment or surface using the single-crystal-to-single-crystal (SC-SC) mode, which replacing
functionalization, the water-stabilized zeolite imidazolite framework the trap with the metastable state of the cationic material and recrys­
has a high adsorption capacity for Cu2+ in water (800 mg g− 1) (Zhang tallizing it to a new structure (Fei et al., 2011). In 2014, Fu et al.

Table 3
Adsorption data of Cd2+ by typical WMOFs in aqueous systems.
Type of MOF Adsorption capacity(mg g− 1) or adsorption rate Equilibrium time (min) Optimal pH Ref.

AMOF-1′ 41 1440 – Chakraborty et al. (2016)


TUM-5 43 15 10 Tahmasebi et al. (2015)
TUM-23 55 10 8 Esrafili et al. (2018)
MOF-88/PAN 225.05 10 – Efome et al. (2018)
3D Co(II) MOF 70%(200 mg,10 ppm) 10 6 Abbasi et al. (2015)
HS-mSi@MOF-5 98 30 6 Zhang et al. (2016a)
HKUST-1-MW@H3PW12O40 33 80 7 Zou et al. (2013)
UiO-66-NHC(S)NHMe 49 240 – Saleem et al. (2016)
UiO-66-EDA 217.39 360 7 Ahmadijokani et al. (2020)
ED-MIL-101(Cd) 63.15 180 8 Elaiwi and Sirkecioglu (2020)
Cu3(BTC)2-SO3H 88.7 10 6 Wang et al. (2015b)
Fe3O4@MOF-235(Fe)–OSO3H 163.9 5 3 Moradi et al. (2016)

7
Z. Li et al. Chemosphere 285 (2021) 131432

Table 4
Adsorption data of Cu2+ by typical WMOFs in aqueous systems.
Type of MOF Adsorption capacity (mg g− 1) or adsorption rate Equilibrium time (min) Optimal pH Ref.

ZIF-8 800 30 4 Zhang et al. (2016b)


f-ZIF-8@GO 1872.24 300 6 Wei et al. (2019)
TUM-5 57 15 10 Tahmasebi et al. (2015)
UiO-66(Zr)–2COOH 11 60 6 Zhang et al. (2015)
UiO-66-EDA 208.33 360 7 Ahmadijokani et al. (2020)
ED-MIL-101(Cd) 69.9 180 6 Elaiwi and Sirkecioglu (2020)
Dy-MOF 5 10 6.5 Jamali et al. (2016)
MOF-1 379.13 – 6 (Yu et al., 2018)
MOF-2(Cd) 439.1 20 5–6 Ghaedi et al. (2018)
HKUST-1-MW@H3PW12O40 225 120 7 Zou et al. (2013)

synthesized large nanotube channels of FIR-53 ([Zn2(Tipa)2(OH)]⋅ the contact of FJI-C11 with Cr2O2− 7 in aqueous solution, thus enhancing
3NO3⋅12H2O) and FIR-54([Zn(Tipa)]⋅2NO3⋅DMF⋅4H2O)) via Zn(II) ions the adsorption of Cr2O2−7 . In addition, the high positive charge density of
and Tipa(tris(4-(1H-imidazole-1-yl)phenyl)amine) ligands (Fu et al., FJI-C11 framework had a stronger coulombic attraction with Cr2O2− 7 ,
2014). Single-crystal x-ray diffraction of the samples after ion exchange thus it had a high selective adsorption to Cr2O2− 7 . In 2020, Jiang et al.
showed that Cr2O2− 7 filling the nanotube channels was able to equili­ reported a kind of zirconium-based cation MOFs (Zr–C-MOFs) that
brate the skeletal charge, suggesting a SC-SC transition of ion exchange introduced methylated 2,5-pyridinedicarboxylic acid as a linker, which
(Fig. 3). Later, Fu et al. synthesized a water-stable MOF [Zn2(TI­ had strong coordination bonds and high metal groups (Jiang et al.,
PA)2(OH)(NO3)3]⋅5H2O, consisting of four small channels. This MOFs 2020). The cluster structure could remain stable in acid, and the pyri­
not only exhibited high selectivity to CrO2− 4 , but more importantly, it dine salt could be used as an anion exchanger. Zr–C-MOF enhanced the
had a space-limiting effect on CrO2− 2−
4 . It could capture CrO4 in water acid resistance, selectivity, and adsorption capacity of CrO2− 4 .
through SC-SC transformation, which proved that MOFs with The construction of the cationic MOF by zwitterionic ligands is also
space-limited properties were powerful single-crystal tools for CrO2− 4 an effective way to enhance the adsorption capacity of CrO2− 4 and
trapping (Fu et al., 2018). In 2019, Othong et al. synthesized flexible Cr2O2−
7 in aqueous solutions. In 2018, Zhang et al. synthesized a cationic
MOFs, which not only maintained good adsorption capacity and selec­ MOF ([Cu2L(H2O)2]⋅(NO3)2⋅5.5H2O) based on Cu2+ and amphoteric
tivity for CrO2− 2−
4 and Cr2O7 , but more importantly, expanded the pH ligand 1,1′ -bis(3,5-dicarboxyphenyl)-4,4′ -bipyridinium chlorine ([H4L]
range for effective adsorption of CrO2− 2−
4 and Cr2O7 (Fig. S2) (Othong Cl2) (Zhang et al., 2018a). Such 3D cation framework had the expected
et al., 2019). nbo-type network of a typical [Cu2(O2C)4] paddlewheel configuration.
By functionalizing the ligands, it is possible to make some MOFs The adsorption properties were due to ion exchange by the interaction of
materials exhibit excellent adsorption performance to CrO2− 4 and Cr2O7
2−
the 3D cation framework with anions. Later, Zhang et al. synthesized
in water. In 2018, He et al. designed a functionalized T-shaped ligand MOFs([In3(ipbp)2(μ2-OH)(μ2-O)3]) with cuboctahedra and triangular
4,4′ ,4′′ -(1Hbenzo[d]imidazole-2,4,7-triyl)tribenzoic acid (H3BTBA) and antiprismatic structures based on the zwitterionic ligand 1-(3,5-dicar­
synthesized the luminescent Zr(IV)- MOF (BUT-39, BUT = Beijing Uni­ boxyphenyl)-4,4′ -bipyridinium bromide (H2(ipbp)Br) (Zhang et al.,
versity of Technology) (He et al., 2018). BUT-39 had a unique porous 2020a). Because of the surface electrostatic interaction of the MOFs
framework structure with BTBA3− as a T-shaped 3-connected linker, and crystal surface with the CrO2− 2−
4 and Cr2O7 oxygen anions. A summary of
Zr6 cluster as an unusual low-symmetric 9-connected node. It could WMOFs that have been used in recent years for CrO2− 4 and Cr2O7
2−

exhibit excellent selective sensing and high adsorption performance for removal from water is presented in Table S1.
Cr2O2−7 at low concentration. In 2019, Zou et al. synthesized
imidazoline-functionalized MOF FJI-C11, which was a positively 3.2.2. SeO2−
3 and SeO4
2-

charged cationic mesoporous MOF (Zou et al., 2019). The high-density Selenium is an essential element for the human body. An appropriate
positive charge of the imidazole functionalized framework facilitated amount of selenium can help to improve the body’s immunity and other

Fig. 3. The structures of FIR-53 before and after ion exchange. a) 3D structure of FIR-53 containing NO3− along the c-axis; (b) 3D structureof FIR-53 containing NO3−
along the b-axis; (c) 3D structure of FIR-53 containing Cr2O2− 2−
7 along the c-axis; (d) 3D structure of FIR-53 containing Cr2O7 along the b-axis. Reproduced with
permission (Fu et al., 2014). Copyright 2014, American Chemical Society.

8
Z. Li et al. Chemosphere 285 (2021) 131432

functions. An intake of less than 40 μg per day will lead to selenium MOFs monomers, Xie et al. synthesized the amino-functionalized
deficiency in the human body, and an intake of more than 400 μg per day MIL-88(Fe), which can accurately monitor trace arsenic in aqueous so­
is toxic (Li et al., 2018; Rayman, 2020). Selenium enters the water body lution and has excellent arsenic removal performance (Xie et al., 2017).
through many ways, such as the erosion of natural sediments, mining Its excellent adsorption performance is due to the open pores, large
and oil refinery discharge, etc., making the selenium pollution of the specific surface area and abundant active sites of NH2-MIL-88(Fe).
water body become increasingly serious. Selenium is mainly present in Bimetallic MOFs can further enhance the adsorption performance of
water in the form of SeO2− 2−
3 and SeO4 . In recent years, the use of MOFs arsenic-oxygen anions. In 2019, Gu et al. prepared a bimetallic
as adsorbents to remove SeO2− 2−
3 and SeO4 from water has emerged and Fe/Mg-MIL-88B by introducing Mg ions into MOFs (Gu et al., 2019). Due
has shown great potential (Kumar et al., 2017). to the replacement of Fe3+ by some Mg2+, the crystal lattice of
The large pore size and large number of node adsorption sites are Fe/Mg-MIL-88B changes and its specific surface area increases.
very important for the MOF to adsorb SeO2− 2−
3 and SeO4 from aqueous Fe/Mg-MIL-88B(0.5) formed Fe–O–As and Mg–O–As bonds due to the
solutions quickly and efficiently. In 2015, Howarth et al. explored the interaction of H2AsO4 with metal ions, which rapidly reached adsorp­
removal of SeO2− 2−
3 and SeO4 from water by a series of Zr-based WMOFs tion equilibrium within 30 min. In addition, there are composites with
and found that NU-1000([Zr6(l3-OH)8(OH)8(TBAPy)2], TBAPy = iron-based MOFs, which also improve their adsorption properties. In
1,3,6,8-tetrakis(p-benzoic acid)pyrene]) had the highest adsorption. 2020, Pandi et al. prepared Fe-MIL-88B@PGCx% composite material by
The NU-1000 was a 3D structure consisting of eight Zr6 nodes bridged by growing MOFs on alginate-derived porous graphitic carbon (PGC)
TBAPy4+. These nodes were rich in hydroxyl groups, resulting in strong (Pandi et al., 2020). After As(III) adsorbed on Fe-MIL-88B@PGC20%
adsorption of SeO2− 2−
3 and SeO4 by NU-1000 (Fig. S3). The NU-1000 had composite, the Fe 2p peak marginally shifted to a higher EB, indicating
triangular pores with a diameter of 12 Å and hexagonal pores with a that iron actively participated in the adsorption of As(III)/As(V) through
diameter of 30 Å, which resulted in fast adsorption speed because of the chemical contacts as a result of the formation of Fe–O–As bonds. As
large pores. In 2017, Li et al. designed Zr6 MOFs with different defects (III)/As(V) exchanged with hydroxyl groups of iron octahedral to suc­
and investigated the effect of defects on the capture of SeO2− 4 from cessfully form an internal spherical bidentate/monodentate complex. As
aqueous solutions (Li et al., 2017). The displaceable ligands and (III) was adsorbed and cum oxidized into As(V) during the adsorption
weak-binding capping ligands on Zr6 nodes could be replaced by SeO2− 4 experiment. In short, As ions are adsorbed by oxygen fraction and
via an anion-exchange mechanism. The efficiency and rate of adsorption combined with iron clusters and exchanged for ligands through hy­
of SeO2−4 by Zr6 MOFs increased with increasing density of missing droxyl groups on the adsorbent surface.
cluster defects and missing linkage defects. According to the experi­ Zirconium-based MOFs were also studied for As(V) sorption. In 2015,
mental results and DFT calculations, smaller Cl-, H2O- or HO- compen­ Wang et al. synthesized the zirconium-based WMOFs (UiO-66), which
sation ligands were easier to be substituted. This modulator synthesis was first used as an adsorbent material to remove arsenate from water
strategy can effectively expand the pore size and introduce defects as (Wang et al., 2015a). The excellent adsorption capacity of UiO-66 is due
binding sites, thus improving the adsorption performance of MOFs. to the porous crystal structure of zirconia clusters, which affords a
In addition, some MOF-based composite materials have excellent number of active sites per unit space. UiO-66 has a two arsenic binding
adsorption properties for SeO2− 2−
3 and SeO4 in water. In previous studies, sites, namely hydroxyl and phthalate ligands. In equilibrium, a Zr6
it was found that due to the low Fe density of some Fe-MOF channel cluster captures 7 equivalent arsenic by forming Zr–O–As coordination
adsorption sites, the oxygen anion could not be adsorbed using these bonds. In 2019, Muthu Prabhu et al. synthesized a benzoic acid-driven
channels (Wang et al., 2018). Wang et al. reported the use of Al3+ in situ zirconium fumarate (Zr-fum) MOF (Muthu Prabhu et al., 2019). In the
modified Fe-MOF to encapsulate aluminum in the pores (Wang et al., absence of benzoic acid, zirconium and fumarate quickly precipitate to
2019b). Because of the increase in adsorption site density, Al@Fe-MOF form amorphous MOF. Using different amounts of benzoic acid to adjust
improved its adsorption performance on SeO2− 3 . Compared with the surface charge of the MOF is beneficial to obtain defect-free Zr-fum
Fe-MOF, the adsorption capacity of Al@Fe-MOF for selenite is increased MOF and enhance the adsorption of AsO3− 3−
4 and AsO3 . In 2019, Huo
by 77%. In 2020, Solis synthesized AC-UiO-66 complex with large sur­ et al. prepared a magnetic Fe3O4@UiO-66 composite, which has a
face area (Solis et al., 2020). The scanning electron microscope-energy unique core-shell structure with more active sites and abundant pores
spectrum analysis results showed that the Se zone and the Zr-rich zone (Huo et al., 2019). In addition, its magnetic properties facilitate the
completely overlap, indicating that the exposed UIO-66 surface was the separation of composites from aqueous solutions. Finally, a summary of
main reason for the adsorption of SeO2− 3 . The preparation of composite WMOFs that have been used in recent years for As(III) and As(V)
materials is an effective way to improve the adsorption performance of removal from water is presented in Table S3.
SeO2− 2−
3 and SeO4 , because it can be easier to achieve large pores and
more adsorption sites. Finally, a summary of WMOFs that have been 4. Heavy metals catalytic reduction by WMOFs
used in recent years for SeO2− 2−
3 and SeO4 removal from water is pre­
sented in Table S2. 4.1. Photocatalytic mechanism

3.2.3. As(III) and As(V) The common semiconductor photocatalyst is an n-type semi­
Arsenic is a very toxic heavy metal element that is widely distributed conductor material, which has an energy band structure of valence
in the natural environment. Arsenic exists in the environment in band, conduction band, and forbidden band. The photocatalyst absorbs
different oxidation states, forming different species. The toxicity of light, and when the photon energy is above the absorption threshold of
inorganic arsenic is far greater than that of organic arsenic. In the water the catalyst, the photocatalyst’s valence band electrons undergo an
environment, arsenic mainly appears in the form of oxygen anions. inter-band jump from the valence band to the conduction band, result­
Arsenic pollution has been recognized as a catastrophic problem, and ing in photogenerated electrons (e− ) and holes (h+). The e− or h+ are in
arsenic in low concentrations in drinking water can cause serious health a bound state due to Coulomb gravity. In the presence of electric and
effects (Sodhi et al., 2019). The mercapto reaction of arsenic with pro­ "chemical" fields (When the photocatalyst is immersed in a solution, a
teins and enzymes inhibits the function of sensitive enzymes, and the "self-built electric field" is created due to the formation of a double
liver, kidney, gastrointestinal tract, nervous system and circulatory electric layer; the oxidant or reductant at the particle/solution interface
system of the human body are seriously affected (Rahman and Singh, captures e− or h+, which can be regarded as a "chemical field"), e− or h+
2019). are separated and migrated to the surface of the photocatalyst by electric
The removal of arsenic from water by MOFs was first reported by Zhu field forces or diffusion (Li et al., 2020). The cavities that migrate to the
et al. (2012). Because of the low adsorption properties of iron-based catalyst surface are usually captured by OH- and H2O to produce

9
Z. Li et al. Chemosphere 285 (2021) 131432

hydroxyl radical (OH•), which is a free radical with strong oxidizing regarded as quantum dots, electron-hole pairs are excited when the
ability and can oxidatively degrade many organic substances. The e− energy of the incident photon exceeds the energy band. In 2019, Zhang
that migrate to the surface of the catalyst are usually captured by the et al. prepared ammonia-functionalized UiO-66 and then loaded silver
electron acceptor O2 adsorbed on the surface, generating a series of free nanoparticles to form Ag/UiO-66-NH2 composites (Zhang et al., 2019b).
radicals such as O2− and HOO-, which can participate in several redox The local surface plasmon resonance effect of Ag facilitates the photo­
processes and can be used to reduce heavy metal ions. Meanwhile, e− or catalytic reduction of Cr(VI) by the composite. However, excessive Ag
h+ can also be compounded inside or on the surface of the catalyst and loading reduces the separation efficiency of UiO-66-NH2, but Ag oc­
release the excitation energy as heat, reducing the quantum efficiency of cupies the NH2 site, resulting in a relatively stable reduction rate. In
the photocatalytic reaction, which is very detrimental to the photo­ 2020, Hussain et al. synthesized UiO-66-NH2@ZnIn2S4 composite with
catalytic reaction. Therefore, the photocatalytic activity of the catalyst high visible light-driven activity (Hussain et al., 2020). UiO-66-NH2
can be enhanced by increasing the light absorption and decreasing the octahedron is bound by ZnIn2S4 nanoflakes. From the analysis of pho­
complexation rate of e− or h+ (Fig. 4) (Henderson, 2011; Banerjee et al., toluminescence spectra and ultraviolet and visible spectra, it can be seen
2014; Schultz and Yoon, 2014; Litter, 2015; Loeb et al., 2018; Buzzetti that the formed UiO-66-NH2@ZnIn2S4 heterostructure has a synergistic
et al., 2019). effect.
Among the many MOFs series, iron-based MOFs exhibit better
property in photocatalytic reduction of Cr(VI) in water, and many iron-
4.2. Heavy metals photocatalytic reduction based MOFs and various composite materials have emerged for this
application. In 2015, Liang et al. immobilized precious metal nano­
According to theoretical calculations, MOFs are semiconductors or particles on MIL-100(Fe) by photo-deposition technique (Liang et al.,
insulators with a band gap of 1.0–5.5 eV (Pham et al., 2014). When 2015). Under visible light irradiation, the reduction activity of the
MOFs are used as photocatalysts, they can afford a number of dispersed synthesized M@MIL-100(Fe) for Cr(VI) was enhanced compared to
active sites, which can reach the substrate/product through open blank-MIL-100(Fe). The photoelectrochemical analysis showed that the
channels. In addition, each active site in MOFs can be seen as a single charge separation efficiency was effectively improved by the
quantum dot, acting as a light absorber, charge generator, and catalytic noble-metal deposition. In 2018, Lin et al. synthesized MIL-53(Fe)/CQD
site, just like a small semiconductor. The organic connectors in MOFs act composite photocatalytic reduction of Cr(VI) by generating carbon
like antennas to absorb light (UV light, visible light and infrared light) quantum dots (CQDs) in situ in MIL-53(Fe) (Lin et al., 2018). The CQDs
and transfer energy to the active site through ligand-metal charge generated in the MOF pores have more opportunities to contact the
transfer. Due to the open framework structure of MOFs, their photo­ photocatalytic unit than those on the surface and are able to form many
catalytic activity can be enhanced by modifying and functionalizing small heterojunction structures. The internally generated CQDs can act
MOFs and by synthesizing MOF-based composites that increase light as a local electron acceptor and have a better ability to inhibit the
absorption and weaken the complexation rate of e− and h+ (Liu et al., complexation of photogenerated electrons with metal-oxygen clusters
2019; Wang et al., 2019a; Wu et al., 2020). Therefore, MOFs as photo­ (Fig. S4). In addition, CQDs can also act as photosensitizers to sensitize
catalysts can effectively reduce heavy metal ions in aqueous solutions. In part of the visible light, improve the light absorption of the composites
recent years, MOFs and MOFs-based composites have been used as in the long-wave visible region, and improve its catalytic activity (Zhang
photocatalysts to reduce heavy metals, and the photocatalysis of chro­ et al., 2020b). In 2019, Yuan et al. synthesized TiO2@NH2-MIL-88B(Fe)
mium in aqueous solutions is the main focus. We take chromium ions as composites for the photocatalytic reduction of Cr(VI) in neutral aqueous
the main object to discuss the research of WMOFs photocatalytic solutions (Yuan et al., 2019). Because H+ is required to reduce Cr(VI),
reduction of heavy metal ions in water. the pH of the reaction solution has a strong influence on the photo­
Ammonia functionalized UiO-66 was first used by Shen and his catalytic activity. The composites reduce Cr(VI) by trapping hydrogen in
colleagues for the photocatalytic reduction of Cr(VI) to Cr(III) in water water or pore scavengers. In addition, it was found that the
(Shen et al., 2013). The introduction of amino groups on the organic small-molecule organic substances could help to inhibit the recombi­
ligand of UiO-66 extends its light absorption range to the visible region. nation of photogenic carriers. The hole scavenger added during the
UiO-66 is almost in an inactive state, while the photocatalytic reduction reduction process, especially ammonium oxalate, significantly improves
reaction of UiO-66 (NH2) proceeds very smoothly under visible light the reduction efficiency of Cr(VI) (Fig. S5).
irradiation. Because the zirconium oxide clusters in UiO-66 (NH2) can be

Fig. 4. Schematic of semiconductor photocatalysis. Reproduced with permission (Banerjee et al., 2014). Copyright 2014, American Chemical Society.

10
Z. Li et al. Chemosphere 285 (2021) 131432

Except for Cr(VI), there is no example of using MOFs as photo­ rhodamine B. The synergistic effect of the photocatalytic reduction and
catalysts to catalyze the reduction of other heavy metal ions. In previous oxidation processes significantly enhanced the photocatalytic activity of
reports, some studies on the reduction of heavy metals by TiO2 and its MIL-53(Fe) in the mixed solution system. The organic dyes are oxidized
composites as photocatalysts have been found, such as photocatalytic to hole scavengers and the Cr(VI) ions are reduced to photogenerated
reduction of Ni(II) in water by TiO2 (Saien et al., 2019), TiO2 hollow electron acceptors. In 2019, Wang et al. synthesized BUC-21 composites
sphere photocatalytic reduction of As(III) and Cr(VI) in water (Cai and with Titanate nanotube composites (TNTs) by a ball milling process,
Li, 2020), Pb(II) in Cu–TiO2 photocatalytic reduction of water (Sree­ which can achieve photocatalytic reduction of Cr(VI) to Cr(III) and
kantan et al., 2014), and other materials photocatalytic reduction of Zn adsorption of the reduced Cr(III) products in one system (Wang et al.,
(II) in water, Cu (II), Hg(II), etc. (Mohamed and Aazam, 2012; He et al., 2019d). First, the composite photocatalytically reduces Cr(VI) ions to Cr
2019; Shahzad et al., 2019). Therefore, photocatalytic materials have (III) ions, and then the formed Cr(III) ions accumulates around BUC-21
potential for photocatalytic reduction of heavy metals in water. MOFs through electrostatic interactions. Due to the high concentration on the
materials have unique advantages, and its photocatalytic reduction of BUC-21 surface, even Cr(OH)3 precipitates were formed, which further
heavy metals in water needs further development. Finally, a summary of inhibited the photocatalytic reduction of Cr(VI). However, TNTs can
WMOFs as photocatalysts for the reduction of Cr(VI) is presented in trap the Cr(III) formed on the BUC-21 surface, eliminating the possible
Table S4. Cr(OH)3 precipitate. The transfer of Cr(III) from BUC-21 to TNTs
resulted in release of the photocatalytic site of BUC-21, which promotes
the photocatalytic reduction of Cr(VI) (Fig. 5). The synergistic effect of
4.3. Synergistic effect of adsorption and photocatalysis adsorption and photocatalysis allows the removal of Cr(III) adsorption
generated during photocatalysis, while enhancing the performance of Cr
Adsorption and photocatalytic reduction of heavy metal ions in (VI) photocatalytic reduction. Recently, Valizadeh et al. developed an
water are not independent, there is always a coupling and synergistic integrated adsorption photoreduction system for capturing Cr(VI) in
effect between them. WMOFs with unique properties have a better effect actual water and reducing it to Cr(III) (Valizadeh et al., 2020). For this
by using both of adsorption and photocatalytic reduction properties purpose, the researchers functionalized the UiO-66 with a double amino
when treating a single pollutant or different pollutants. group, makes the newly synthesized Zr-BDC-(NH2)2 have the dual
In 2015, Liang et al. synthesized dual-functional MIL-53(Fe) photo­ functions of adsorption and photocatalysis. It is assumed that the ideal
catalytic degradation of organic dyes and reduction of Cr(VI) in water application is based on separation from water. The polymer needs to be
(Liang et al., 2015c). In a mixed solution system, MIL-53(Fe) can oxidize carefully selected to ensure that the polymer is sufficiently hydrophilic.
organic dyes and reduce Cr(VI) to Cr(III) at the same time. The presence Since polyethersulfone (PES) is highly hydrophilic, PES is selected and
of Cr(VI) promotes the photocatalytic oxidation of malachite green and

Fig. 5. (a) The reusability of BT-1 for the removal of Cr ions. (b) PXRD patterns of BT-1 before and after three recycles for Cr removal. (c) XPS scanning spectra before
and after the photocatalysis-adsorption process of BT-1. (d) Possible photocatalytic and adsorption reaction mechanism of BT-1. Reproduced with permission (Wang
et al., 2019d). Copyright 2019, Elsevier Ltd.

11
Z. Li et al. Chemosphere 285 (2021) 131432

modified with carboxyl groups to enhance its hydrophilicity. A com­ performance of WMOFs and WMOF-based composites, which synergizes
parison between un-modified and modified PES indicates a significant the adsorption performance and photocatalytic reduction performance.
decrease in the water contact angle from 89◦ to 25◦ , respectively. Next, The photocatalytic reduction can be obtained during the adsorption of
Zr-BDC-(NH2)2 was combined with polyethersulfone into MOF@pol­ heavy metal ions, forming a green and sustainable removal process.
ymer beads (M@PB). The researchers designed a simple separation
process, including two glass columns for placing the material and an Declaration of competing interest
excitation light source. Zr-BDC-(NH2)2@PB provides the highest Cr(VI)
absorption capacity reported so far, fast extraction rate, high Cr(VI) The authors declare that they have no known competing financial
selectivity in actual water, and complete recovery capability. The ca­ interests or personal relationships that could have appeared to influence
pacity of Zr-BDC-(NH2)2@PB is attributed to the Zr-BDC-(NH2)2 crystals the work reported in this paper.
and the synergistic effect between the MOF and polymer. Due to the
synergistic effect between MOF and polymer, which might promote the Acknowledgments
formation of molecular pockets in the polymer matrix at the MOF
polymer interface; the latter could introduce new adsorption sites within This study was financially supported by the Program for the National
the polymer matrix. Natural Science Foundation of China (51879103, 51779090), Hunan
Natural Science Foundation (2020JJ3009), Hunan Researcher Award
5. Conclusions and prospects Program (2020RC3025), and the Fundamental Research Funds for the
Central Universities (531118040083, 531118010473).
By reviewing the WMOFs and its application in the removal of heavy
metal ions in water in the past few years, we have a clearer under­ Appendix A. Supplementary data
standing of its development. First, water stability is an important
consideration factor for any material used in an aquatic environment. In Supplementary data to this article can be found online at https://doi.
this paper, we focus on PSM method to improve the water stability of org/10.1016/j.chemosphere.2021.131432.
MOFs, and discuss the strategies to improve the water stability of MOFs
in terms of hydrophobization, ligand functionalization, and modifica­
References
tion of metal clusters. A great number of WMOFs have been successfully
constructed, and their applications in the water environment have been Abbasi, A., Moradpour, T., Van Hecke, K., 2015. A new 3D cobalt (II) metal–organic
further developed. Then, WMOFs and MOF-based composites are green framework nanostructure for heavy metal adsorption. Inorg. Chim. Acta. 430,
261–267.
materials with excellent performance in removing cationic and anionic
Ahmadijokani, F., Tajahmadi, S., Bahi, A., Molavi, H., Rezakazemi, M., Ko, F.,
heavy metal ions from an aqueous environment. It not only is an Aminabhavi, T.M., Arjmand, M., 2020. Ethylenediamine-functionalized Zr-based
excellent adsorption material, but also can exert the performance of MOF for efficient removal of heavy metal ions from water. Chemosphere 264,
photocatalytic reduction of heavy metal ions. By introducing appro­ 128466.
Ali, H., Khan, E., Ilahi, I., 2019. Environmental chemistry and ecotoxicology of
priate functional groups into MOFs or synthesizing MOF-based com­ hazardous heavy metals: environmental persistence, toxicity, and bioaccumulation.
posites, the ability to remove heavy metal ions can be effectively J. Chem. 1–14.
enhanced, and the stability of MOFs can be improved. MOFs as adsor­ Bae, Y.S., Liu, J., Wilmer, C.E., Sun, H., Dickey, A.N., Kim, M.B., Benin, A.I., Willis, R.R.,
Barpaga, D., LeVan, M.D., Snurr, R.Q., 2014. The effect of pyridine modification of
bents can quickly and effectively adsorb heavy metal ions in an aqueous Ni-DOBDC on CO2 capture under humid conditions. Chem. Commun. 50,
solution through a large number of nodes and pores. MOFs as photo­ 3296–3298.
catalysts can produce electron and hole pairs under visible light Banerjee, S., Pillai, S.C., Falaras, P., O’Shea, K.E., Byrne, J.A., Dionysiou, D.D., 2014.
New insights into the mechanism of visible light photocatalysis. J. Phys. Chem. Lett.
photoexcitation, and electrons and superoxide radicals as strong oxi­ 5, 2543–2554.
dants can effectively reduce heavy metal ions. It is worth paying Batten, S.R., Champness, N.R., Chen, X.-M., Garcia-Martinez, J., Kitagawa, S.,
attention to the application of the photocatalytic reduction performance Öhrström, L., O’Keeffe, M., Paik Suh, M., Reedijk, J., 2013. Terminology of
metal–organic frameworks and coordination polymers (IUPAC Recommendations
of WMOFs and WMOF-based composite materials. They synergize the 2013). Pure Appl. Chem. 85, 1715–1724.
adsorption performance and the photocatalytic reduction performance Bhatti, A.A., Memon, S., Memon, N., 2014. Dichromate extraction by calix [4] arene
to better remove heavy metal ions in the water. appended amberlite XAD-4 resin. Separ. Sci. Technol. 49, 664–672.
Bora, A.J., Dutta, R.K., 2019. Removal of metals (Pb, Cd, Cu, Cr, Ni, and Co) from
Although the construction of WMOFs and their application in the
drinking water by oxidation-coagulation-absorption at optimized pH. J. Water
water environment have been well developed, we still found problems in Process Eng. 31, 100839.
their development. Firstly, it is limited by the synthesis cost of MOFs. At Burtch, N.C., Jasuja, H., Walton, K.S., 2014. Water stability and adsorption in metal-
present, most MOFs are still in the experimental stage, and the high price organic frameworks. Chem. Rev. 114, 10575–10612.
Buzzetti, L., Crisenza, G.E.M., Melchiorre, P., 2019. Mechanistic studies in
of organic linkers such as H3BTC and H2BDC used in the synthesis photocatalysis. Angew. Chem. Int. Ed. 58, 3730–3747.
process leads to the high cost of their synthesis. Secondly, many MOFs Cai, J., Li, H., 2020. Electrospun polymer nanofibers coated with TiO2 hollow spheres
require organic solvents such as methanol and N, N-dimethylformamide catalyze for high synergistic photo-conversion of Cr(VI) and As(III) using visible
light. Chem. Eng. J. 398, 125644.
in the synthesis process, and the toxicity of using organic solvents to the Canivet, J., Fateeva, A., Guo, Y., Coasne, B., Farrusseng, D., 2014. Water adsorption in
environment should also be fully considered. Last but not least, MOFs MOFs: fundamentals and applications. Chem. Soc. Rev. 43, 5594–5617.
are influenced by pH, ionic strength, and other ions in a complex real Cavka, J.H., Jakobsen, S., Olsbye, U., Guillou, N., Lamberti, C., Bordiga, S., Lillerud, K.P.,
2008. A new zirconium inorganic building brick forming metal organic frameworks
water environment. In addition to practical application issues, many of with exceptional stability. J. Am. Chem. Soc. 130, 13850–13851.
the reported porous MOFs are microporous, which is not conducive to Chai, W.S., Cheun, J.Y., Kumar, P.S., Mubashir, M., Majeed, Z., Banat, F., Ho, S.-H.,
the migration of large metal ions within the framework. The slow Show, P.L., 2021. A review on conventional and novel materials towards heavy
metal adsorption in wastewater treatment application. J. Clean. Prod. 296, 126589.
adsorption kinetics of many MOFs is also an important issue. Chakraborty, A., Bhattacharyya, S., Hazra, A., Ghosh, A.C., Maji, T.K., 2016. Post-
Faced with these problems and challenges, the application potential synthetic metalation in an anionic MOF for efficient catalytic activity and removal of
of WMOFs needs to be further improved in the future research. The heavy metal ions from aqueous solution. Chem. Commun. 52, 2831–2834.
Chen, S., Feng, F., Li, S., Li, X.-X., Shu, L., 2018. Metal-organic framework DUT-67 (Zr)
development of green synthesis and inexpensive organic linkers for
for adsorptive removal of trace Hg2+ and CH3Hg+ in water. Chem. Speciat.
aqueous phase is the direction in which WMOFs can be practically Bioavailab. 30, 99–106.
applied in the future. The adsorption and catalytic properties of WMOFs Cheng, S.Y., Show, P.L., Lau, B.F., Chang, J.S., Ling, T.C., 2019. New prospects for
are improved and structural stability is enhanced by modifying various modified algae in heavy metal adsorption. Trends Biotechnol. 37, 1255–1268.
Cohen, J.G., Cohen, S.M, 2010. Moisture-resistant and superhydrophobic metal-organic
functional groups and combining them with other materials. It is worth frameworks obtained via postsynthetic modification. J. Am. Chem. Soc. 132,
paying attention to the application of the photocatalytic reduction 4560–4561.

12
Z. Li et al. Chemosphere 285 (2021) 131432

Decoste, J.B., Peterson, G.W., Smith, M.W., Stone, C.A., Willis, C.R., 2012. Enhanced nanoparticles loading efficiency on defect-free high silica ZSM-5 zeolite for the
stability of Cu-BTC MOF via perfluorohexane plasma-enhanced chemical vapor reduction of nitrophenols. Chemosphere 256, 127083.
deposition. J. Am. Chem. Soc. 134, 1486–1489. Henderson, M.A., 2011. A surface science perspective on TiO2 photocatalysis. Surf. Sci.
Demessence, A., D’Alessandro, D.M., Foo, M.L., Long, J.R., 2009. Strong CO2 binding in a Rep. 66, 185–297.
water-stable, triazolate-bridged metal-organic framework functionalized with Hoch, L.B., Mack, E.J., Hydutsky, B.W., Hershman, J.M., Skluzacek, J.M., Mallouk, T.E.,
ethylenediamine. J. Am. Chem. Soc. 131, 8784–8786. 2008. Carbothermal synthesis of carbon-supported nanoscale zero-valent iron
Dinda, D., Gupta, A., Saha, S.K., 2013. Removal of toxic Cr(VI) by UV-active particles for the remediation of hexavalent chromium. Environ. Sci. Technol. 42,
functionalized graphene oxide for water purification. J. Mater. Chem. 1, 2600–2605.
11221–11228. Hoskins, B.F., Robson, R., 1990. Design and construction of a new class of scaffolding-
Ding, N., Li, H., Feng, X., Wang, Q., Wang, S., Ma, L., Zhou, J., Wang, B., 2016. like materials comprising infinite polymeric frameworks of 3D-linked molecular
Partitioning MOF-5 into confined and hydrophobic compartments for carbon capture rods. J. Am. Chem. Soc. 112, 1546–1554.
under humid conditions. J. Am. Chem. Soc. 138, 10100–10103. Huang, L., He, M., Chen, B., Hu, B., 2015. A designable magnetic MOF composite and
Ding, L., Luo, X., Shao, P., Yang, J., Sun, D., 2018. Thiol-functionalized Zr-based facile coordination-based post-synthetic strategy for the enhanced removal of Hg2+
metal–organic framework for capture of Hg(II) through a proton exchange reaction. from water. J. Mater. Chem. 3, 11587–11595.
ACS Sustain. Chem. Eng. 6, 8494–8502. Huang, L., He, M., Chen, B., Hu, B., 2016. A mercapto functionalized magnetic Zr-MOF
Ding, M.L., Cai, X.C., Jiang, H.L., 2019. Improving MOF stability: approaches and by solvent-assisted ligand exchange for Hg2+ removal from water. J. Mater. Chem. 4,
applications. Chem. Sci. 10, 10209–10230. 5159–5166.
Drache, F., Bon, V., Senkovska, I., Marschelke, C., Synytska, A., Kaskel, S., 2016. Humelnicu, D., Lazar, M.M., Ignat, M., Dinu, I.A., Dragan, E.S., Dinu, M.V., 2020.
Postsynthetic inner-surface functionalization of the highly stable zirconium-based Removal of heavy metal ions from multi-component aqueous solutions by eco-
metal-organic framework DUT-67. Inorg. Chem. 55, 7206–7213. friendly and low-cost composite sorbents with anisotropic pores. J. Hazard Mater.
Dukic-Cosic, D., Baralic, K., Javorac, D., Djordjevic, A.B., Bulat, Z., 2020. An overview of 381, 120980.
molecular mechanisms in cadmium toxicity. Curr. Opin Toxicol. 19, 56–62. Huo, J.-B., Xu, L., Chen, X., Zhang, Y., Yang, J.-C.E., Yuan, B., Fu, M.-L., 2019. Direct
Efome, J.E., Rana, D., Matsuura, T., Lan, C.Q., 2018. Insight studies on metal-organic epitaxial synthesis of magnetic Fe3O4@UiO-66 composite for efficient removal of
framework nanofibrous membrane adsorption and activation for heavy metal ions arsenate from water. Microporous Mesoporous Mater. 276, 68–75.
removal from aqueous solution. ACS Appl. Mater. Interfaces 10, 18619–18629. Hupp, J.T., 2005. Chemistry: enhanced: Better living through nanopore chemistry.
Elaiwi, F.A., Sirkecioglu, A., 2020. Amine-functionalized metal organic frameworks MIL- Science 309, 2008–2009.
101(Cr) adsorbent for copper and cadmium ions in single and binary solution. Separ. Hussain, M.B., Azhar, U., Loussala, H.M., Razaq, R., 2020. Synergetic effect of ZnIn2S4
Sci. Technol. 55, 3362–3374. nanosheets with metal-organic framework molding heterostructure for efficient
Esrafili, L., Safarifard, V., Tahmasebi, E., Esrafili, M.D., Morsali, A., 2018. Functional visible- light driven photocatalytic reduction of Cr(VI). Arabian J. Chem. 13,
group effect of isoreticular metal–organic frameworks on heavy metal ion 5939–5948.
adsorption. New J. Chem. 42, 8864–8873. Jamali, A., Tehrani, A.A., Shemirani, F., Morsali, A., 2016. Lanthanide metal-organic
Fei, H., Bresler, M.R., Oliver, S.R., 2011. A new paradigm for anion trapping in high frameworks as selective microporous materials for adsorption of heavy metal ions.
capacity and selectivity: crystal-to-crystal transformation of cationic materials. Dalton Trans. 45, 9193–9200.
J. Am. Chem. Soc. 133, 11110–11113. Jiang, H.-L., Makal, T.A., Zhou, H.-C., 2013. Interpenetration control in metal–organic
Feng, M., Zhang, P., Zhou, H.C., Sharma, V.K., 2018. Water-stable metal-organic frameworks for functional applications. Coord. Chem. Rev. 257, 2232–2249.
frameworks for aqueous removal of heavy metals and radionuclides: a review. Jiang, C., Sun, R., Du, Z., Singh, V., Chen, S., 2020. A cationic Zr-based metal organic
Chemosphere 209, 783–800. framework with enhanced acidic resistance for selective and efficient removal of
Ferey, G., Serre, C., Mellot-Draznieks, C., Millange, F., Surble, S., Dutour, J., CrO2−4 . New J. Chem. 44, 12646–12653.
Margiolaki, I., 2004. A hybrid solid with giant pores prepared by a combination of Joseph, L., Jun, B.M., Flora, J.R.V., Park, C.M., Yoon, Y., 2019. Removal of heavy metals
targeted chemistry, simulation, and powder diffraction. Angew. Chem. 43, from water sources in the developing world using low-cost materials: a review.
6296–6301. Chemosphere 229, 142–159.
Fu, F.L., Wang, Q., 2011. Removal of heavy metal ions from wastewaters: a review. Joyaramulu, K., Geyer, F., Schneemann, A., Kment, S., Otyepka, M., Zboril, R.,
J. Environ. Manag. 92, 407–418. Vollmer, D., Fischer, R.A., 2019. Hydrophobic metal-organic frameworks. Adv.
Fu, Z.S., Xi, S.H., 2020. The effects of heavy metals on human metabolism. Toxicol. Mater. 31, 1900820.
Mech. Methods 30, 167–176. Kumar, P., Pournara, A., Kim, K.-H., Bansal, V., Rapti, S., Manos, M.J., 2017. Metal-
Fu, H.-R., Xu, Z.-X., Zhang, J., 2014. Water-stable metal–organic frameworks for fast and organic frameworks: challenges and opportunities for ion-exchange/sorption
high dichromate trapping via single-crystal-to-single-crystal ion exchange. Chem. applications. Prog. Mater. Sci. 86, 25–74.
Mater. 27, 205–210. Kumar, A., Kumar, A., Cabral-Pinto, M.M.S., Chaturvedi, A.K., Shabnam, A.A.,
Fu, H.R., Wang, N., Qin, J.H., Han, M.L., Ma, L.F., Wang, F., 2018. Spatial confinement of Subrahmanyam, G., Mondal, R., Gupta, D.K., Malyan, S.K., Kumar, S.S., Khan, S.A.,
a cationic MOF: a SC-SC approach for high capacity Cr(VI)-oxyanion capture in Yadav, K.K., 2020. Lead toxicity: health hazards, influence on food chain, and
aqueous solution. Chem. Commun. 54, 11645–11648. sustainable remediation approaches. Int. J. Environ. Res. Publ. Health 17, 2179.
Fu, L.K., Wang, S.X., Lin, G., Zhang, L.B., Liu, Q.M., Fang, J., Wei, C.H.N., Liu, G., 2019. Kumar, V., Pandita, S., Sidhu, G.P.S., Sharma, A., Khanna, K., Kaur, P., Bali, A.S.,
Post-functionalization of UiO-66-NH2 by 2,5-Dimercapto-1,3,4-thiadiazole for the Setia, R., 2021. Copper bioavailability, uptake, toxicity and tolerance in plants: a
high efficient removal of Hg(II) in water. J. Hazard Mater. 368, 42–51. comprehensive review. Chemosphere 262, 127810.
Fu, Y.K., Zeng, G.M., Lai, C., Huang, D.L., Qin, L., Yi, H., Liu, X.G., Zhang, M.M., Li, B.S., Lai, C., Liu, S., Zhang, C., Zeng, G., Huang, D., Qin, L., Liu, X., Yi, H., Wang, R., Huang, F.,
Liu, S.Y., Li, L., Li, M.F., Wang, W.J., Zhang, Y.J., Pi, Z.J., 2020. Hybrid architectures Li, B., Hu, T., 2018. Electrochemical aptasensor based on sulfur-nitrogen Co-doped
based on noble metals and carbon-based dots nanomaterials: a review of recent ordered mesoporous carbon and thymine Hg2+-thymine mismatch structure for Hg2+
progress in synthesis and applications. Chem. Eng. J. 399, 125743. detection. ACS Sens. 12, 2–35.
Furukawa, H.F., Ko, N., Go, Y.B., Aratani, N., Choi, S.B., 2010. Ultrahigh porosity in Lai, C., Wang, Z., Qin, L., Fu, Y., Li, B., Zhang, M., Liu, S., Li, L., Yi, H., Liu, X., Zhou, X.,
metal-organic frameworks. Science 329, 424–428. An, N., An, Z., Shi, X., Feng, C., 2021. Metal-organic frameworks as burgeoning
Gao, J., Sun, S.P., Zhu, W.P., Chung, T.S., 2014. Chelating polymer modified P84 materials for the capture and sensing of indoor VOCs and radon gases. Coord. Chem.
nanofiltration (NF) hollow fiber membranes for high efficient heavy metal removal. Rev. 427, 213565.
Water Res. 63, 252–261. Lee, J., Chuah, C.Y., Kim, J., Kim, Y., Ko, N., Seo, Y., Kim, K., Bae, T.H., Lee, E., 2018.
Ghaedi, A.M., Panahimehr, M., Nejad, A.R.S., Hosseini, S.J., Vafaei, A., Baneshi, M.M., Separation of acetylene from carbon dioxide and ethylene by a water-stable
2018. Factorial experimental design for the optimization of highly selective microporous metal-organic framework with aligned imidazolium groups inside the
adsorption removal of lead and copper ions using metal organic framework MOF-2 channels. Angew. Chem. 130, 7995–7999.
(Cd). J. Mol. Liq. 272, 15–26. Lei, C., Gao, J., Ren, W., Xie, Y., Abdalkarim, S.Y.H., Wang, S., Ni, Q., Yao, J., 2019.
Gu, Y., Xie, D., Wang, Y., Qin, W., Zhang, H., Wang, G., Zhang, Y., Zhao, H., 2019. Facile Fabrication of metal-organic frameworks@cellulose aerogels composite materials for
fabrication of composition-tunable Fe/Mg bimetal-organic frameworks for removal of heavy metal ions in water. Carbohydr. Polym. 205, 35–41.
exceptional arsenate removal. Chem. Eng. J. 357, 579–588. Li, H., Eddaoudi, M., O’Keeffe, M., Yaghi, O.M., 1999. Design and synthesis of an
Hader, D.P., Banaszak, A.T., Villafane, V.E., Narvarte, M.A., Gonzalez, R.A., Helbling, E. exceptionally stable and highly porous metal-organic framework. Nature 402,
W., 2020. Anthropogenic pollution of aquatic ecosystems: emerging problems with 276–279.
global implications. Sci. Total Environ. 713, 9742–9746. Li, H., Shi, W., Zhao, K., Li, H., Bing, Y., Cheng, P., 2012. Enhanced hydrostability in Ni-
Hall, J.N., Bollini, P., 2020. Enabling access to reduced open-metal sites in metal-organic doped MOF-5. Inorg. Chem. 51, 9200–9207.
framework materials through choice of anion identity: the case of MIL-100(Cr). ACS Li, J., Chen, C., Zhang, S., Wang, X., 2014a. Surface functional groups and defects on
Mater. Lett. 2, 838–844. carbon nanotubes affect adsorption–desorption hysteresis of metal cations and
He, T., Zhang, Y.Z., Kong, X.J., Yu, J., Lv, X.L., Wu, Y., Guo, Z.J., Li, J.R., 2018. Zr(IV)- oxoanions in water. Environ. Sci. Nano 1, 488–495.
based metal-organic framework with T-shaped ligand: unique structure, high Li, R., Ren, X.Q., Ma, H.W., Feng, X., Lin, Z.G., Li, X.G., Hu, C.W., Wang, B., 2014b.
stability, selective detection, and rapid adsorption of Cr2O2- 7 in Water. ACS Appl. Nickel-substituted zeolitic imidazolate frameworks for time-resolved alcohol sensing
Mater. Interfaces 10, 16650–16659. and photocatalysis under visible light. J. Mater. Chem. 2, 5724–5729.
He, F., Lu, Z., Song, M., Liu, X., Tang, H., Huo, P., Fan, W., Dong, H., Wu, X., Han, S., Li, J., Liu, Y., Wang, X., Zhao, G., Ai, Y., Han, B., Wen, T., Hayat, T., Alsaedi, A.,
2019. Selective reduction of Cu2+ with simultaneous degradation of tetracycline by Wang, X., 2017. Experimental and theoretical study on selenate uptake to zirconium
the dual channels ion imprinted POPD-CoFe2O4 heterojunction photocatalyst. Chem. metal–organic frameworks: effect of defects and ligands. Chem. Eng. J. 330,
Eng. J. 360, 750–761. 1012–1021.
He, J.F., Lai, C., Qin, L., Li, B.S., Liu, S.Y., Jiao, L.J., Fu, Y.K., Huang, D.L., Li, L.,
Zhang, M.M., Liu, X.G., Yi, H., Chen, L., Li, Z.W., 2020. Strategy to improve gold

13
Z. Li et al. Chemosphere 285 (2021) 131432

Li, J., Wang, X., Zhao, G., Chen, C., Chai, Z., Alsaedi, A., Hayat, T., Wang, X., 2018. Nguyen, J.G., Cohen, S.M., 2012. Postsynthetic methods for the functionalization of
Metal-organic framework-based materials: superior adsorbents for the capture of metal-organic frameworks. Chem. Rev. 112, 970–1000.
toxic and radioactive metal ions. Chem. Soc. Rev. 47, 2322–2356. Othong, J., Boonmak, J., Youngme, S., 2019. Highly selective Cr2O2- 7 removal in aqueous
Li, G.P., Zhang, K., Zhang, P.F., Liu, W.N., Tong, W.Q., Hou, L., Wang, Y.Y., 2019. Thiol- medium by using a flexible 2D metal-organic framework through single-crystal-to-
functionalized pores via post-synthesis modification in a metal-organic framework single-crystal transformation. J. Environ. Chem. Eng. 7, 102998.
with selective removal of Hg(II) in water. Inorg. Chem. 58, 3409–3415. Pandi, K., Prabhu, S.M., Ahn, Y., Park, C.M., Choi, J., 2020. Design and synthesis of
Li, B.S., Liu, S.Y., Lai, C., Zeng, G.M., Zhang, M.M., Zhou, M.Z., Huang, D.L., Qin, L., biopolymer-derived porous graphitic carbon covered iron-organic frameworks for
Liu, X.G., Li, Z.W., An, N., Xu, F.H., Yi, H., Zhang, Y.J., Chen, L., 2020. Unravelling depollution of arsenic from waters. Chemosphere 254, 126769.
the interfacial charge migration pathway at atomic level in 2D/2D interfacial Park, K.S., Ni, Z., Cote, A.P., Choi, J.Y., Huang, R., Uribe-Romo, F.J., Yaghi, O.M., 2006.
Schottky heterojunction for visible-light-driven molecular oxygen activation. Appl. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc.
Catal., B 266, 118650. Natl. Acad. Sci. U.S.A. 103, 10186–10191.
Liang, R., Jing, F., Shen, L., Qin, N., Wu, L., 2015. M@MIL-100(Fe) (M = Au, Pd, Pt) Pavesi, T., Moreira, J.C., 2020. Mechanisms and individuality in chromium toxicity in
nanocomposites fabricated by a facile photodeposition process: efficient visible-light humans. J. Appl. Toxicol. 40, 1183–1197.
photocatalysts for redox reactions in water. Nano Res 8, 3237–3249. Pham, H.Q., Mai, T., Pham-Tran, N.N., Kawazoe, Y., Mizuseki, H., Nguyen-Manh, D.,
Liang, R., Jing, F., Shen, L., Qin, N., Wu, L., 2015c. MIL-53(Fe) as a highly efficient 2014. Engineering of band gap in metal organic frameworks by functionalizing
bifunctional photocatalyst for the simultaneous reduction of Cr(VI) and oxidation of organic linker: a systematic density functional theory investigation. J. Phys. Chem. C
dyes. J. Hazard Mater. 287, 364–372. 118, 4567–4577.
Liang, L., Liu, L., Jiang, F., Liu, C., Yuan, D., Chen, Q., Wu, D., Jiang, H.L., Hong, M., Qadir, N.u., Said, S.A.M., Bahaidarah, H.M., 2015. Structural stability of metal organic
2018. Incorporation of In2S3 nanoparticles into a metal-organic framework for frameworks in aqueous media – controlling factors and methods to improve
ultrafast removal of Hg from water. Inorg. Chem. 57, 4891–4897. hydrostability and hydrothermal cyclic stability. Microporous Mesoporous Mater.
Lin, R., Li, S., Wang, J., Xu, J., Xu, C., Wang, J., Li, C., Li, Z., 2018. Facile generation of 201, 61–90.
carbon quantum dots in MIL-53(Fe) particles as localized electron acceptors for Qin, L., Zeng, G., Lai, C., Huang, D., Xu, P., Zhang, C., Cheng, M., Liu, X., Liu, S., Li, B.,
enhancing their photocatalytic Cr(vi) reduction. Inorg. Chem. Front. 5, 3170–3177. 2018. “Gold rush” in modern science: fabrication strategies and typical advanced
Litter, M.I., 2015. Mechanisms of removal of heavy metals and arsenic from water by applications of gold nanoparticles in sensing. Coord. Chem. Rev. 359, 1–31.
TiO2-heterogeneous photocatalysis. Pure Appl. Chem. 87, 557–567. Qin, L., Wang, Z., Fu, Y., Lai, C., Liu, X., Li, B., Liu, S., Yi, H., Li, L., Zhang, M., Li, Z.,
Liu, X., Li, Y., Ban, Y., Peng, Y., Jin, H., Bux, H., Xu, L., Caro, J., Yang, W., 2013. Cao, W., Niu, Q., 2021. Gold nanoparticles-modified MnFe2O4 with synergistic
Improvement of hydrothermal stability of zeolitic imidazolate frameworks. Chem. catalysis for photo-Fenton degradation of tetracycline under neutral pH. J. Hazard
Commun. 49, 9140–9142. Mater. 414, 125448.
Liu, T., Che, J.X., Hu, Y.Z., Dong, X.W., Liu, X.Y., Che, C.M., 2014. Alkenyl/thiol-derived Rahman, Z., Singh, V.P., 2019. The relative impact of toxic heavy metals (THMs) (arsenic
metal-organic frameworks (MOFs) by means of postsynthetic modification for (As), cadmium (Cd), chromium (Cr)(VI), mercury (Hg), and lead (Pb)) on the total
effective mercury adsorption. Chemistry 20, 14090–14095. environment: an overview. Environ. Monit. Assess. 191, 419.
Liu, Y., Liu, Z.F., Huang, D.L., Cheng, M., Zeng, G.M., Lai, C., Zhang, C., Zhou, C.Y., Rayman, M.P., 2020. Selenium intake, status, and health: a complex relationship. Horm.
Wang, W.J., Jiang, D.N., Wang, H., Shao, B.B., 2019. Metal or metal-containing Int. J. Endocrinol. Metab. 19, 9–14.
nanoparticle@MOF nanocomposites as a promising type of photocatalyst. Coord. Ren, X., Li, J., Tan, X., Wang, X., 2013. Comparative study of graphene oxide, activated
Chem. Rev. 388, 63–78. carbon and carbon nanotubes as adsorbents for copper decontamination. Dalton
Liu, B., Vikrant, K., Kim, K.-H., Kumar, V., Kailasa, S.K., 2020a. Critical role of water Trans. 42, 5266–5274.
stability in metal–organic frameworks and advanced modification strategies for the Ricco, R., Konstas, K., Styles, M.J., Richardson, J.J., Babarao, R., Suzuki, K., Scopece, P.,
extension of their applicability. Environ. Sci. Nano 7, 1319–1347. Falcaro, P., 2015. Lead(II) uptake by aluminium based magnetic framework
Liu, S., Lai, C., Liu, X., Li, B., Zhang, C., Qin, L., Huang, D., Yi, H., Zhang, M., Li, L., composites (MFCs) in water. J. Mater. Chem. 3, 19822–19831.
Wang, W., Zhou, X., Chen, L., 2020b. Metal-organic frameworks and their Rubin, H.N., Reynolds, M.M., 2017. Functionalization of metal-organic frameworks to
derivatives as signal amplification elements for electrochemical sensing. Coord. achieve controllable wettability. Inorg. Chem. 56, 5266–5274.
Chem. Rev. 424, 213520. Rudd, N.D., Wang, H., Fuentes-Fernandez, E.M., Teat, S.J., Chen, F., Hall, G., Chabal, Y.
Loeb, S.K., Alvarez, P.J.J., Brame, J.A., Cates, E.L., Choi, W., Crittenden, J., Dionysiou, D. J., Li, J., 2016. Highly efficient luminescent metal-organic framework for the
D., Li, Q., Li-Puma, G., Quan, X., Sedlak, D.L., David Waite, T., Westerhoff, P., simultaneous detection and removal of heavy metals from water. ACS Appl. Mater.
Kim, J.-H., 2018. The technology horizon for photocatalytic water treatment: sunrise Interfaces 8, 30294–30303.
or sunset? Environ. Sci. Technol. 53, 2937–2947. Saien, J., Azizi, A., Ghamari, F., 2019. Simultaneous photocatalytic reduction/
Lu, M., Li, L., Shen, S., Chen, D., Han, W., 2019. Highly efficient removal of Pb2+ by a degradation of divalent nickel/naphthalene pollutants in aqueous solutions. Water
sandwich structure of metal–organic framework/GO composite with enhanced Sci. Technol. 79, 240–250.
stability. New J. Chem. 43, 1032–1037. Saleem, H., Rafique, U., Davies, R.P., 2016. Investigations on post-synthetically modified
Luo, F., Chen, J.L., Dang, L.L., Zhou, W.N., Lin, H.L., Li, J.Q., Liu, S.J., Luo, M.B., 2015. UiO-66-NH2 for the adsorptive removal of heavy metal ions from aqueous solution.
High-performance Hg2+ removal from ultra-low-concentration aqueous solution Microporous Mesoporous Mater. 221, 238–244.
using both acylamide- and hydroxyl-functionalized metal–organic framework. Santasnachok, C., Kurniawan, W., Hinode, H., 2015. The use of synthesized zeolites from
J. Mater. Chem. 3, 9616–9620. power plant rice husk ash obtained from Thailand as adsorbent for cadmium
Luo, X., Shen, T., Ding, L., Zhong, W., Luo, J., Luo, S., 2016. Novel thymine- contamination removal from zinc mining. J. Environ. Chem. Eng. 3, 2115–2126.
functionalized MIL-101 prepared by post-synthesis and enhanced removal of Hg2+ Schultz, D.M., Yoon, T.P., 2014. Solar synthesis: prospects in visible light photocatalysis.
from water. J. Hazard Mater. 306, 313–322. Science 343, 1239176.
Mandal, S., Natarajan, S., Mani, P., Pankajakshan, A., 2020. Post-synthetic modification Shahzad, K., Tahir, M.B., Sagir, M., Kabli, M.R., 2019. Role of CuCo2S4 in Z-scheme
of metal-organic frameworks toward applications. Adv. Funct. Mater. 4, 2006291. MoSe2/BiVO4 composite for efficient photocatalytic reduction of heavy metals.
Miranda-Quintana, R.A., 2017. Note: the minimum electrophilicity and the hard/soft Ceram. Int. 45, 23225–23232.
acid/base principles. J. Chem. Phys. 146, 064101. Shao, Z., Huang, C., Dang, J., Wu, Q., Liu, Y., Ding, J., Hou, H., 2018. Modulation of
Mohamed, R.M., Aazam, E.S., 2012. Enhancement of photocatalytic activity of ZnO–SiO2 magnetic behavior and Hg2+ removal by solvent-assisted linker exchange based on a
by nano-sized Ag for visible photocatalytic reduction of Hg(II). Desalin. Water Treat. water-stable 3D MOF. Chem. Mater. 30, 7979–7987.
50, 140–146. Shearer, G.C., Colombo, V., Chavan, S., Albanese, E., Civalleri, B., Maspero, A.,
Mon, M., Lloret, F., Ferrando-Soria, J., Martí-Gastaldo, C., Armentano, D., Pardo, E., Bordiga, S., 2013. Stability vs. reactivity: understanding the adsorption properties of
2016. Selective and efficient removal of mercury from aqueous media with the Ni3(BTP)2 by experimental and computational methods. Dalton Trans. 42,
highly flexible arms of a BioMOF. Angew. Chem. 128, 11333–11338. 6450–6458.
Mon, M., Qu, X., Ferrando-Soria, J., Pellicer-Carreño, I., Sepúlveda-Escribano, A., Ramos- Shen, L., Liang, S., Wu, W., Liang, R., Wu, L., 2013. Multifunctional NH2-mediated
Fernandez, E.V., Jansen, J.C., Armentano, D., Pardo, E., 2017. Fine-tuning of the zirconium metal-organic framework as an efficient visible-light-driven photocatalyst
confined space in microporous metal–organic frameworks for efficient mercury for selective oxidation of alcohols and reduction of aqueous Cr(VI). Dalton Trans. 42,
removal. J. Mater. Chem. 5, 20120–20125. 13649–13657.
Mon, M., Bruno, R., Tiburcio, E., Viciano-Chumillas, M., Kalinke, L.H.G., Ferrando- Shen, X.Y., Chi, Y.K., Xiong, K.N., 2019. The effect of heavy metal contamination on
Soria, J., Armentano, D., Pardo, E., 2019. Multivariate metal-organic frameworks for humans and animals in the vicinity of a zinc smelting facility. PloS One 14, 1–15.
the simultaneous capture of organic and inorganic contaminants from water. J. Am. Sodhi, K.K., Kumar, M., Agrawal, P.K., Singh, D.K., 2019. Perspectives on arsenic
Chem. Soc. 141, 13601–13609. toxicity, carcinogenicity and its systemic remediation strategies. Environ. Technol.
Moradi, S.E., Haji Shabani, A.M., Dadfarnia, S., Emami, S., 2016. Sulfonated metal Innovation 16, 100462.
organic framework loaded on iron oxide nanoparticles as a new sorbent for the Solis, K.L.B., Kwon, Y.H., Kim, M.H., An, H.R., Jeon, C., Hong, Y., 2020. Metal organic
magnetic solid phase extraction of cadmium from environmental water samples. framework UiO-66 and activated carbon composite sorbent for the concurrent
Anal. Methods 8, 6337–6346. adsorption of cationic and anionic metals. Chemosphere 238, 124656.
Muthu Prabhu, S., Kancharla, S., Park, C.M., Sasaki, K., 2019. Synthesis of modulator- Song, Y.C., Wang, N., Yang, L.Y., Wang, Y.G., Yu, D., Ouyang, X.K., 2019. Facile
driven highly stable zirconium-fumarate frameworks and mechanistic investigations fabrication of ZIF-8/calcium alginate microparticles for highly efficient adsorption of
of their arsenite and arsenate adsorption from aqueous solutions. CrystEngComm 21, Pb(II) from aqueous solutions. Ind. Eng. Chem. Res. 58, 6394–6401.
2320–2332. Sreekantan, S., Lai, C.W., Mohd Zaki, S., 2014. The influence of lead concentration on
Narayani, M., Shetty, K.V., 2013. Chromium-resistant bacteria and their environmental photocatalytic reduction of Pb(II) ions assisted by Cu-TiO2 nanotubes. Int. J.
condition for hexavalent chromium removal: a review. Crit. Rev. Environ. Sci. Photoenergy 1–7.
Technol. 43, 955–1009.

14
Z. Li et al. Chemosphere 285 (2021) 131432

Sun, D.T., Peng, L., Reeder, W.S., Moosavi, S.M., Tiana, D., Britt, D.K., Oveisi, E., Yang, S.J., Park, C.R., 2012. Preparation of highly moisture-resistant black-colored metal
Queen, W.L., 2018. Rapid, selective heavy metal removal from water by a metal- organic frameworks. Adv. Mater. 24, 4010–4013.
organic framework/polydopamine composite. ACS Cent. Sci. 4, 349–356. Yang, P., Shu, Y., Zhuang, Q., Li, Y., Gu, J., 2019. A robust MOF-based trap with high-
Tahmasebi, E., Masoomi, M.Y., Yamini, Y., Morsali, A., 2015. Application of density active alkyl thiol for the super-efficient capture of mercury. Chem. Commun.
mechanosynthesized azine-decorated zinc(II) metal-organic frameworks for highly 55, 12972–12975.
efficient removal and extraction of some heavy-metal ions from aqueous samples: a Yin, N., Wang, K., Wang, L., Li, Z., 2016. Amino-functionalized MOFs combining ceramic
comparative study. Inorg. Chem. 54, 425–433. membrane ultrafiltration for Pb (II) removal. Chem. Eng. J. 306, 619–628.
Tan, X., Fan, Q., Wang, X., Grambow, B., 2009. Eu(III) sorption to TiO2 (anatase and Yin, N., Wang, K., Xia, Y.a., Li, Z., 2018. Novel melamine modified metal-organic
rutile): batch, XPS, and EXAFS studies. Environ. Sci. Technol. 43, 3115–3121. frameworks for remarkably high removal of heavy metal Pb (II). Desalination 430,
Tian, J., Shi, C., Xiao, C., Jiang, F., Yuan, D., Chen, Q., Hong, M., 2020. Introduction of 120–127.
flexibility into a metal-organic framework to promote Hg(II) capture through Yu, C., Shao, Z., Liu, L., Hou, H., 2018. Efficient and selective removal of copper(II) from
adaptive deformation. Inorg. Chem. 59, 18264–18275. aqueous solution by a highly stable hydrogen-bonded metal–organic framework.
Upadhyay, U., Sreedhar, I., Singh, S.A., Patel, C.M., Anitha, K.L., 2021. Recent advances Cryst. Growth Des. 18, 3082–3088.
in heavy metal removal by chitosan based adsorbents. Carbohydr. Polym. 251, Yuan, S., Feng, L., Wang, K., Pang, J., Bosch, M., Lollar, C., Sun, Y., Qin, J., Yang, X.,
117000. Zhang, P., Wang, Q., Zou, L., Zhang, Y., Zhang, L., Fang, Y., Li, J., Zhou, H.C., 2018.
Valizadeh, B., Nguyen, T.N., Kampouri, S., Sun, D.T., Mensi, M.D., Stylianou, K., Smit, B., Stable metal-organic frameworks: design, synthesis, and applications. Adv. Mater.
Queen, W.L., 2020. A novel integrated Cr(VI) adsorption-photoreduction system 30, 1704303.
using MOF@polymer composite beads. J. Mater. Chem. 8, 9629–9637. Yuan, R., Yue, C., Qiu, J., Liu, F., Li, A., 2019. Highly efficient sunlight-driven reduction
Veerakumar, P., Lin, K.C., 2020. An overview of palladium supported on carbon-based of Cr(VI) by TiO2@NH2-MIL-88B(Fe) heterostructures under neutral conditions.
materials: synthesis, characterization, and its catalytic activity for reduction of Appl. Catal., B 251, 229–239.
hexavalent chromium. Chemosphere 253, 126750. Zhang, W., Hu, Y., Ge, J., Jiang, H.-L., Yu, S.-H., 2014. A facile and general coating
Wang, C., Liu, X., Chen, J.P., Li, K., 2015a. Superior removal of arsenic from water with approach to moisture/water-resistant metal–organic frameworks with intact
zirconium metal-organic framework UiO-66. Sci. Rep. 5, 16613. porosity. J. Am. Chem. Soc. 136, 16978–16981.
Wang, Y., Ye, G., Chen, H., Hu, X., Niu, Z., Ma, S., 2015b. Functionalized metal–organic Zhang, Y., Zhao, X., Huang, H., Li, Z., Liu, D., Zhong, C., 2015. Selective removal of
framework as a new platform for efficient and selective removal of cadmium(II) from transition metal ions from aqueous solution by metal–organic frameworks. RSC Adv.
aqueous solution. J. Mater. Chem. 3, 15292–15298. 5, 72107–72112.
Wang, C., Liu, X., Keser Demir, N., Chen, J.P., Li, K., 2016a. Applications of water stable Zhang, J., Xiong, Z., Li, C., Wu, C., 2016a. Exploring a thiol-functionalized MOF for
metal-organic frameworks. Chem. Soc. Rev. 45, 5107–5134. elimination of lead and cadmium from aqueous solution. J. Mol. Liq. 221, 43–50.
Wang, C.C., Du, X.D., Li, J., Guo, X.X., Wang, P., Zhang, J., 2016b. Photocatalytic Cr(VI) Zhang, Y., Xie, Z., Wang, Z., Feng, X., Wang, Y., Wu, A., 2016b. Unveiling the adsorption
reduction in metal-organic frameworks: a mini-review. Appl. Catal., B 193, 198–216. mechanism of zeolitic imidazolate framework-8 with high efficiency for removal of
Wang, R., Gong, Y., Xu, H.J., Wei, S.Y., Wu, D.Y., 2018. Characteristics and adsorption copper ions from aqueous solutions. Dalton Trans. 45, 12653–12660.
properties of Se(IV) for MOF-Fe prepared with different conditioning agents. Chin. J. Zhang, C., Liu, Y., Sun, L., Shi, H., Shi, C., Liang, Z., Li, J., 2018a. A zwitterionic ligand-
Inorg. Chem. 34, 906–916. based cationic metal-organic framework for rapidly selective dye capture and highly
Wang, C.C., Yi, X.H., Wang, P., 2019a. Powerful combination of MOFs and C3N4 for efficient Cr2O2-7 removal. Chem. Eur J. 24, 2718–2724.
enhanced photocatalytic performance. Appl. Catal., B 247, 24–48. Zhang, Y.Z., He, T., Kong, X.J., Lv, X.L., Wu, X.Q., Li, J.R., 2018b. Tuning water sorption
Wang, R., Xu, H., Zhang, K., Wei, S., Deyong, W., 2019b. High-quality Al@Fe-MOF in highly stable Zr(IV)-metal-organic frameworks through local functionalization of
prepared using Fe-MOF as a micro-reactor to improve adsorption performance for metal clusters. ACS Appl. Mater. Interfaces 10, 27868–27874.
selenite. J. Hazard Mater. 364, 272–280. Zhang, R., Liu, Y., An, Y., Wang, Z., Wang, P., Zheng, Z., Qin, X., Zhang, X., Dai, Y.,
Wang, X., Liu, W., Fu, H., Yi, X.-H., Wang, P., Zhao, C., Wang, C.-C., Zheng, W., 2019d. Huang, B., 2019a. A water-stable triazine-based metal-organic framework as an
Simultaneous Cr(VI) reduction and Cr(III) removal of bifunctional MOF/Titanate efficient adsorbent of Pb(II) ions. Colloids Surf., A 560, 315–322.
nanotube composites. Environ. Pollut. 249, 502–511. Zhang, W., Wang, L., Zhang, J., 2019b. Preparation of Ag/UiO-66-NH2 and its
Wang, C., He, C., Luo, Y.H., Su, S., Wang, J.Y., Hong, D.L., He, X.T., Chen, C., Sun, B.W., application in photocatalytic reduction of Cr(VI) under visible light. Res. Chem.
2020a. Efficient mercury chloride capture by ultrathin 2D metal-organic framework Intermed. 45, 4801–4811.
nanosheets. Chem. Eng. J. 379, 122337. Zhang, C., Shi, H., Yan, Y., Sun, L., Ye, Y., Lu, Y., Liang, Z., Li, J., 2020a. A zwitterionic
Wang, C., Lin, G., Xi, Y., Li, X., Huang, Z., Wang, S., Zhao, J., Zhang, L., 2020b. ligand-based water-stable metal-organic framework showing photochromic and Cr
Development of mercaptosuccinic anchored MOF through one-step preparation to (VI) removal properties. Dalton Trans. 49, 10613–10620.
enhance adsorption capacity and selectivity for Hg(II) and Pb(II). J. Mol. Liq. 317, Zhang, M.M., Lai, C., Li, B.S., Xu, F.H., Huang, D.L., Liu, S.Y., Qin, L., Fu, Y.K., Liu, X.G.,
113896. Yi, H., Zhang, Y.J., He, J.F., Chen, L., 2020b. Unravelling the role of dual quantum
Wang, L.W., Hou, D.Y., Cao, Y.N., Ok, Y.S., Tack, F.M.G., Rinklebe, J., O’Connor, D., dots cocatalyst in 0D/2D heterojunction photocatalyst for promoting photocatalytic
2020c. Remediation of mercury contaminated soil, water, and air: a review of organic pollutant degradation. Chem. Eng. J. 396, 13.
emerging materials and innovative technologies. Environ. Int. 134, 105281. Zhang, X., Wang, B., Alsalme, A., Xiang, S., Zhang, Z., Chen, B., 2020c. Design and
Wang, Z., Lai, C., Qin, L., Fu, Y., He, J., Huang, D., Li, B., Zhang, M., Liu, S., Li, L., applications of water-stable metal-organic frameworks: status and challenges. Chem.
Zhang, W., Yi, H., Liu, X., Zhou, X., 2020d. ZIF-8-modified MnFe2O4 with high Soc. Rev. 423, 213507.
crystallinity and superior photo-Fenton catalytic activity by Zn-O-Fe structure for TC Zhang, Y.J., Mao, F.X., Wang, L.J., Yuan, H.Y., Liu, P.F., Yang, H.G., 2020e. Recent
degradation. Chem. Eng. J. 392, 124851. advances in photocatalysis over metal-organic frameworks-based materials. Sol. RRL
Wang, L.L., Xie, L.B., Zhao, W.W., Liu, S.J., Zhao, Q., 2021. Oxygen-facilitated dynamic 4, 29.
active-site generation on strained MoS2 during photo-catalytic hydrogen evolution. Zhao, F., Su, C., Yang, W., Han, Y., Luo, X., Li, C., Tang, W., Yue, T., Li, Z., 2020a. In-situ
Chem. Eng. J. 405, 127028. growth of UiO-66-NH2 onto polyacrylamide-grafted nonwoven fabric for highly
Wei, N., Zheng, X., Ou, H., Yu, P., Li, Q., Feng, S., 2019. Fabrication of an amine- efficient Pb(II) removal. Appl. Surf. Sci. 527, 146862.
modified ZIF-8@GO membrane for high-efficiency adsorption of copper ions. New J. Zhao, Y.J., Deng, Q.Y., Lin, Q., Zeng, C.Y., Zhong, C., 2020b. Cadmium source
Chem. 43, 5603–5610. identification in soils and high-risk regions predicted by geographical detector
Wen, T., Wu, X., Tan, X., Wang, X., Xu, A., 2013. One-pot synthesis of water-swellable method. Environ. Pollut. 263, 114338.
Mg-Al layered double hydroxides and graphene oxide nanocomposites for efficient Zheng, S.N., Wang, Q., Yuan, Y.Z., Sun, W.M., 2020. Human health risk assessment of
removal of As(V) from aqueous solutions. ACS Appl. Mater. Interfaces 5, 3304–3311. heavy metals in soil and food crops in the Pearl River Delta urban agglomeration of
Wu, J., Zhou, J., Zhang, S., Alsaedi, A., Hayat, T., Li, J., Song, Y., 2019. Efficient removal China. Food Chem. 316, 126213.
of metal contaminants by EDTA modified MOF from aqueous solutions. J. Colloid Zhu, B.-J., Yu, X.-Y., Jia, Y., Peng, F.-M., Sun, B., Zhang, M.-Y., Luo, T., Liu, J.-H.,
Interface Sci. 555, 403–412. Huang, X.-J., 2012. Iron and 1,3,5-benzenetricarboxylic metal–organic coordination
Wu, T., Liu, X.J., Liu, Y., Cheng, M., Liu, Z.F., Zeng, G.M., Shao, B.B., Liang, Q.H., polymers prepared by solvothermal method and their application in efficient As(V)
Zhang, W., He, Q.Y., Zhang, W., 2020. Application of QD-MOF composites for removal from aqueous solutions. J. Phys. Chem. C 116, 8601–8607.
photocatalysis: energy production and environmental remediation. Coord. Chem. Zou, F., Yu, R., Li, R., Li, W., 2013. Microwave-assisted synthesis of HKUST-1 and
Rev. 403, 213097. functionalized HKUST-1-@H3PW12O40: selective adsorption of heavy metal ions in
Xie, D., Ma, Y., Gu, Y., Zhou, H., Zhang, H., Wang, G., Zhang, Y., Zhao, H., 2017. water analyzed with synchrotron radiation. ChemPhysChem 14, 2825–2832.
Bifunctional NH2-MIL-88(Fe) metal–organic framework nanooctahedra for highly Zou, Y., Wang, X., Khan, A., Wang, P., Liu, Y., Alsaedi, A., Wang, X., 2016.
sensitive detection and efficient removal of arsenate in aqueous media. J. Mater. Environmental remediation and application of nanoscale zero-valent iron and its
Chem. 5, 23794–23804. composites for the removal of heavy metal ions: a review. Environ. Sci. Technol. 50,
Xu, R., Jian, M., Ji, Q., Hu, C., Tang, C., Liu, R., Zhang, X., Qu, J., 2020. 2D water-stable 7290–7304.
zinc-benzimidazole framework nanosheets for ultrafast and selective removal of Zou, Y.H., Liang, J., He, C., Huang, Y.B., Cao, R., 2019. A mesoporous cationic metal-
heavy metals. Chem. Eng. J. 382, 122658. organic framework with a high density of positive charge for enhanced removal of
dichromate from water. Dalton Trans. 48, 6680–6684.

15

You might also like