You are on page 1of 19

Review Article

ISSN 1998-0124 CN 11-5974/O4


2019, 12(1): 000–000 https://doi.org/10.1007/s12274-019-2309-8

Nanomaterials for the abatement of cadmium (II) ions from water/


wastewater
Kumar Vikrant1, Vanish Kumar2, Kowsalya Vellingiri3, and Ki-Hyun Kim1 ()
1
Department of Civil and Environmental Engineering, Hanyang University, 222 Wangsimni-ro, Seoul 04763, Republic of Korea
2
National Agri-Food Biotechnology Institute (NABI), S.A.S. Nagar, Punjab 140306, India
3
Environmental and Water Resources Engineering Division, Department of Civil Engineering, IIT Madras, Chennai 600036, India

© Tsinghua University Press and Springer-Verlag GmbH Germany, part of Springer Nature 2019
Received: 20 October 2018 / Revised: 18 January 2019 / Accepted: 20 January 2019

ABSTRACT
The rapid rise of modern industry is the source of unchecked effluents containing many pernicious heavy metals (e.g., cadmium). To rehabilitate
the ecology, food resources, and health of humans and animals, various conventional methodologies are being used in wastewater treatment
facilities for the abatement of cadmium. Nonetheless, the development of advanced, economical, and efficient adsorbents is needed because
of the many shortcomings of conventional methods (e.g., high cost, intensive operation, and inefficiency). Recent advancements in materials
science and chemistry have introduced the use of nanomaterials, which possess very high specific surface areas and multiple functionalities,
for the removal of specific targets such as cadmium. This review explores the recent developments and trends in nanomaterial adsorption
technology for the mitigation of cadmium. The paper further surveys the present obstacles and future opportunities for the advancement of
nanomaterial-based technologies in the area of water treatment.

KEYWORDS
nanomaterials, cadmium, adsorption, pollution control, water purification, heavy metal

1 Introduction accumulate in the food chain [11].


The prime sources of anthropogenic cadmium are ceramics,
The regulation and removal of cadmium from water is a high- electroplating, metallurgical plants, mining, paint and pigment
priority environmental control enterprise because of the deleterious plants, plastic production facilities, and dry battery manufacturing
potential of cadmium species toward animal and human health and plants [12, 13]. Given the harm caused by cadmium species, a
overall ecosystems in both the short and long terms [1, 2]. Although permissible limitation of 0.003 mg·L−1 for Cd(II) ions was set for
only a quarter of the Earth’s surface is covered by landmass [3], drinking water by the World Health Organization [14]. To tackle the
about 97% of all water resources are stored in the oceans, making pressing issue of cadmium pollution, various conventional methods
them ill-suited for most purposes [4]. Moreover, of the remaining have been used in wastewater treatment facilities, as detailed in
3% of available water, about 2.4% is confined as ice, glaciers, and section 2; however, most conventional methods cannot yet reduce
atmospheric/soil moisture, remaining out of reach for normal, highly contaminated water to the permissible level [15, 16]. Also,
everyday applications. Consequently, only 0.6% of the Earth’s total the conventional options are generally uneconomical because they
water resources can be classified as drinkable freshwater (groundwater, require secondary treatments before the discharge of effluents [17].
lakes, and rivers) [4] . The adsorptive remediation of cadmium rich water has been
Recently, the research community has become markedly interested recommended as the most fitting and simple option [18]. Finding
in the abatement of heavy metals (e.g., cadmium, mercury, lead, an adequate sorbent for the effective removal of Cd(II) ions has thus
chromium, and arsenic) from water/wastewater because it has become a paramount issue [19]. To that end, many nanomaterials
recognized the environmental significance of their toxic effects. For (NMs) have been developed as highly efficient, economical, and
instance, taking in cadmium-laden water can damage the cardiovascular environmentally benign adsorbents for cadmium [20]. In comparison
and urinary systems and lead to diverse diseases, including skeletal to conventional sorbents (e.g., activated carbon and zeolite), NMs
deformation, high blood pressure, infertility, and emphysema [5, 6]. have distinct physicochemical characteristics (e.g., high specific
Cadmium has also been recognized as a human carcinogen by the surface area, short passage for diffusion, and tunable active surface
United States Environmental Protection Agency (US EPA) [7]. sites) that provide excellent sorption capacities [18, 21]. Although
Essentially, the presence of cadmium within living cells can lead many toxic heavy metals (e.g., cadmium, mercury, lead, chromium,
to the generation of H2O2 [8]. A rise in peroxide concentration and arsenic) may occur together in wastewater, the sources of each
within a cell can induce lipid peroxidation (oxidative stress) and metal may vary greatly depending on conditions [20, 22–24]. Likewise,
glutathione and ascorbate depletion [9]. Moreover, H2O2 can cause as the concentration of particular metallic species might be
nuclear DNA to defunctionalize proteins through the conversion of distinctively different from others in wastewater systems, it is often
thiol groups into sulfones [10]. Cadmium has been recognized as a desirable to adopt selective separation approaches to efficiently
high-priority target by the US EPA because of its tendency to treat them [20, 23]. Moreover, the extent of interactions between

Address correspondence to kkim61@hanyang.ac.kr


2 Nano Res.

NMs and individual heavy metallic species (e.g., surface complexation agents (e.g., trimercaptotriazine, sodium dimethyldithiocarbamate,
and cation exchange) may vary significantly by their intrinsic chemical and potassium/sodium thiocarbonate) have been specifically prepared
properties [24–26]. Hence, it is imperative to properly evaluate the for the treatment of heavy metal-rich water systems [31]. However,
performances of NMs for sorptive interactions with individual these commercially available chelating agents are environmentally
heavy metallic species (e.g., Cd(II) ions). risky and have limited binding sites for heavy metals. A few other
For this review article, we have collected insights into various heavy metal chelating materials (e.g., 1,3-benzenediamidoethanethiol
topics and factors that affect the use of NMs for the removal of dianion, 1,3,5-hexahydrotriazinedithiocarbamate, N,Nʹ-bis-
cadmium (Cd(II)) ions from water. To do that, we initially analyzed (dithiocarboxy)piperazine, potassium ethyl xanthate, and dipropyl
the feasibility of various NMs for their use in Cd(II) ion-rich water/ dithiophosphate) have also been tested [32, 34–36]. Although those
wastewater. A section has been devoted to assessing the performance chelating agents exhibited good removal efficiency for heavy metals,
of the various conventional methods for comparison with NM they still require further refinement in terms of environmental safety
technology. Each section of the review describes a variety of NMs and efficiency.
to promote understanding of this subject. Moreover, we discuss The ion exchange method for the treatment of water systems rich
the underlying hindrances and barriers to current nanotechnology in heavy metals, especially Cd(II) ions, has the advantages of rapid
to propel further advancement in this challenging area of research. kinetics, a large capacity for treatment, and great removal efficiency
On the whole, we furnish a thorough account of the current [27, 37, 38]. Ion-exchange materials that have been explored for
state-of-the-art in using nanotechnology for the abatement of Cd(II) the removal of heavy metal (e.g., Cd(II) ions) from water include
ions from wastewater/water, and we provide a detailed roadmap for natural/synthetic resins, zeolites, and silicate minerals [37, 39–42].
the further growth of this pollutant treatment technology. Likewise, the membrane filtration approach to the removal of heavy
metals has also been explored. On the basis of the filtration
2 Conventional approaches to the treatment of membrane used, membrane filtration is divided into four methods:
reverse osmosis, ultrafiltration, electrodialysis, and nanofiltration
Cd(II) ions [43–46]. The membrane used for ultrafiltration is suitable for the
The presence of heavy metals in the water system exerts harmful removal of colloidal materials and metal complexes. However, the
effects on both aquatic and terrestrial systems. To resolve those pore size in ultrafiltration membranes is larger than metal ions, so
problems, many approaches (e.g., chemical precipitation, ion exchange several modifications (e.g., polymer enhanced ultrafiltration and
filtration, chelation, electrochemical techniques, osmosis, coagulation, micellar enhanced ultrafiltration) have been made to increase the
adsorption, and electrodialysis) (Fig. 1) have been developed and efficiency of ultrafiltration. Materials such as ceramics, cellulose,
used. polycarbonates, polysulfones, polyethersulfone, and their modifications
The chemical precipitation method is often used to treat heavy are widely used in ultrafiltration-based abatement of heavy metallic
metal-contaminated water because of its simple operation and low ions from water systems [44, 47–52]. A semi-permeable membrane
cost [27]. Insoluble precipitates of heavy metals are formed through has been used for the reverse osmosis approach. Reverse osmosis is
chemical reactions and then removed from the treated water through an efficient technique for the abatement of heavy metallic ions from
filtration or sedimentation. Conventionally, two major approaches wastewater/water that has been widely used in water purification
have been used for heavy metal chemical precipitation, sulfide and systems [53, 54]. However, its limitations (e.g., large power
hydroxide precipitation [27–30]. Those approaches require iron requirements to maintain the pumping pressure, cost, and need for
sulfide/sulfate reducing bacteria and Ca(OH)2/NaOH, respectively, frequent membrane replacements due to its poisoning) have led
and both of them have important limitations. For instance, the researchers to explore alternatives [27, 53]. The use of nanofiltration
hydroxide precipitation method is limited by the generation of low decreases the energy required to use the membrane filtration
density sludge, precipitation at different pH for different metals, approach [27]. Another membrane filtration approach, electrodialysis,
and the inhibition of metal-oxide precipitation in the presence of uses ion-exchange membranes [55]. The coagulation and flocculation
complexing agents. On the other hand, the precipitates formed approach has demonstrated good potential in the abatement of
heavy metallic ions from aqueous systems [56, 57]. However, all
using the sulfide precipitation approach are less soluble than those
those techniques require additional treatment, such as precipitation,
formed using the hydroxide precipitation method, and sulfide
to remove heavy metallic ions from wastewater/water [27]. The
precipitation can generate H2S gas and colloidal precipitates that are
floatation and electrochemical approaches are also commonly used
difficult to separate. Chelating precipitation is another method used
for this purpose [58–63].
for the chemical precipitation of metal ions [31–33]. A few chelating
Adsorption is one of the most promising options for the abatement
of heavy metals from aqueous systems, with benefits such as low
cost and high efficiency, flexibility, and simplicity [64–66]. Many
conventional materials (e.g., activated carbon, amorphous silica,
clay minerals, diatomite, biomass, zeolites, and polymers) have been
tested for the adsorptive removal of heavy metals [38, 64, 67–72].
All the conventional techniques remove Cd(II) ions from water, but
they are subject to some specific limitations [27]. For instance,
chemical precipitation cannot remove low concentrations of heavy
metals, and it forms large amounts of sludge. Chemical regeneration
is required for ion exchange resins, and the high cost of activated
carbon (a commonly used adsorbent) limits the use of adsorption.
Likewise, the membrane filtration approach faces problems such as
membrane fouling, high cost, and process complexity. Economic
factors (e.g., the initial capital requirements and high operational
and maintenance costs) limit the applicability of floatation- and
  electrochemical-based heavy metal removal systems. Advancements
Figure 1 List of methods explored for the removal of heavy metals. in materials have produced dramatic progress in the development

| www.editorialmanager.com/nare/default.asp
Nano Res. 3

of systems for removing heavy metals, especially for adsorption. preferably due to their faster uptake kinetics and resulting superior
The rest of this review provides detailed information about using performance [20] (Tables 1 and 2). As summarized in Table 1, the
nanomaterials to remove Cd(II) ions from contaminated water. inorganic NMs can be generally classified into three categories: (i)
transition metal oxides, (ii) transition metals, and (iii) carbon-
3 Nanotechnological approaches based materials. The recovery of spent nanoparticles (NPs) after
wastewater treatment poses a practical constraint to the application
For the abatement of Cd(II) ions from wastewater, the use of NMs of nanotechnology [73]. 
composed of inorganic constituents has been recommended In this regard, NMs and their composites with magnetic

Table 1 Removal of Cd(II) ions from water through inorganic nanomaterials and their composites
Optimum Initial Cd Maximum Maximum Final Cd Distribution
Order Nanomaterial adsorption concentration removal adsorption capacity concentration coefficient Reference
−1
condition (°C, pH) (μM) efficiency (%) (mg·g ) (μM) (mg·g−1 ·μM−1)
[A] Transition
metal oxide
nanomaterials
1 Fe3O4 25, 6 88.96 90 9.68 8.896 1.09 [91]
2 Fe3O4 25, 4 117.92 88.7 10.9 13.325 0.82 [90]
3 Fe3O4 25, 5.5 1334.4 — 11 — — [215]
4 Fe3O4-SO3H 25, 7 88.96 94 80.9 5.338 15.16 [82]
5 CNSR-Fe3O4 30, 10 444.8 99.9 54.6 0.445 122.75 [92]
6 Ni-αFe2O3 55, 7 177.92 98 90.91 3.558 25.55 [93]
7 CoFe2O4/SiO2 25, 8 8.89 95 5 0.445 11.25 [94]
8 γ-Fe2O3/MgSNTs 25 444.8 — 200 — — [98]
9 NiO — 117.92 — 97 — — [102]
10 CuO nanoblades 25, 7 889.6 99.2 192.3 7.117 27.02 [105]
11 MNP-OPP 45, 7 142.33 96 76.92 5.693 13.51 [103]
12 ZrO2 20, 7 8.89 99.9 3.18 0.009 357.71 [104]
13 NZP 30, 6 500 — 214.7 — — [121]
14 SM-MNP 25, 5 8.89 92 5.18 0.711 7.28 [216]
15 ZnO nanorods 30, 8 889.6 91 147.25 80.064 1.84 [79]
16 PANI/CoHCF 25, 8 889.6 64.15 27.17 318.922 0.09 [108]
17 GAC-NSIO 25, 6 444.8 — 7.84 — — [217]
18 Ti-C 25, 7 444.8 — 31.7 — — [106]
19 HA-TiO2 25, 9 0.35 — 0.36 — — [116]
20 TiO2 dandelions 25, 5 1779.2 80 396 355.840 1.11 [80]
[B] Transition
metal nanoparticles
1 nZVI 25, 9 444.8 78 188 97.856 1.92 [129]
2 FeSSi 25, 6 600 — 105 — — [135]
3 MoS2/TMPyP — 2.49 91 — — — [139]
4 CdS Np 30, 6 444.8 — 141.5 — — [144]
5 Ag NP 25, 6.5 44.48 — 4.67 — — [218]
[C] Carbon based
nanomaterials
1 CNT 25, 7 8.89 27 0.65 6.490 0.10 [199]
2 MWCNT 25, 8 889.6 95.1 181.8 43.590 4.17 [148]
3 e-MWCNT 45 44.48 — 25.7 — — [219]
4 N-doped CNT 25, 7 600 — 7.86 — — [149]
5 Oxidized N-doped 25, 5 700 — 9.33 — — [152]
MWCNT
6 MWCNT-COOH 25, 7 444.8 96 16.34 17.792 0.92 [153]
7 CNT-PAC 26, 5.5 20.19 98.35 11.16 0.333 33.50 [154]
8 γ-Fe2O3/MWCNT 25, 8 133.44 97.2 78.81 3.736 21.09 [157]
9 PEG-MWCNT 25, 4 88.96 97 77.6 2.669 29.08 [155]
10 MGO 25, 6 1,779.2 85 91.29 266.880 0.34 [160]
11 Cys-MGO nanosheet 55, 9 133.44 — 30.3 — — [163]
12 aGO 20, 5 177.92 — 156 — — [166]
13 MDFGO 25, 6.2 444.8 99.4 — — — [164]

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


4 Nano Res.

(Continued)
Optimum Initial Cd Maximum Maximum Final Cd Distribution
Order Nanomaterial adsorption concentration removal adsorption capacity concentration coefficient Reference
condition (°C, pH) (μM) efficiency (%) (mg·g−1) (μM) (mg·g−1 ·μM−1)
14 M-Fe3O4-GO 25, 5 222.4 — 125 — — [165]
15 SG-PySA-GO 22 1,067.52 — 10 — — [169]
16 TCAS-rGO 46, 7 88.96 90 128 8.896 14.39 [172]
17 nZVI/rGO 25.31, 6.48 179.34 76.11 25.37 42.844 0.59 [174]
18 3D-SRGO 55, 6 17.79 93 234.8 1.245 188.55 [177]
19 EDA-GCNP 35, 6 88.96 99 18.7 0.890 21.02 [220]
20 MT-CD/SiO2 25, 6 88.96 90 0.3 8.896 0.03 [203]

Table 2 Removal of Cd(II) ions from water through organic/miscellaneous nanomaterials and their composites
Optimum Initial Cd Maximum Maximum Final Cd Distribution
Order Nanomaterial adsorption concentration removal adsorption capacity concentration coefficient Reference
condition (°C, pH) (μM) efficiency (%) (mg·g−1) (μM) (mg·g−1·μM−1)
[A] Organic
nanocomposites
1 MECT 25, 5 889.6 82 209.2 160.13 1.31 [182]
2 P(MB-IA)-g-MNCC 50, 6 889.6 99 262.27 8.90 29.48 [178]
3 CTS-VMT 30, 4 444.8 97 58.48 13.34 4.38 [190]
4 Nanocellulose fibers 25, 6 222.4 90.7 9.7 20.68 0.47 [179]
5 nHApC 25, 5.6 889.6 92 122 71.17 1.71 [189]
6 carboxycellulose 25, 7 2,224 84 2,550 355.84 7.17 [180]
nanofibers
[B] Miscellaneous
nanomaterials
1 Polylysine-resorcinol 25, 8 88.96 98 220 1.78 123.65 [191]
alumina nanotube
2 TPDP/SiO2 25, 5.5 88.96 93 148.32 6.23 23.82 [192]
3 MgO NP 25, 7 889.6 — 1,500 — — [193]
4 MNCLAP 25, 5 2,000 — 123.89 — — [194]
5 nano γ-Al2O3 25, 5 444.8 87 17.22 57.82 0.30 [195]
6 nHAp 25, 5.5 224.82 — 64.07 — — [196]

characteristics (e.g., Fe3O4 and Fe3O4-SO3H; Table 1) have received that can be used for Cd(II) ion removal. In fact, they combine high
widespread attention. The magnetic NMs typically consist of a Cd(II) ion sorption capacity with great ion exchange capability and
core–shell structure. The inner layer is generally made up of the ability to be magnetically separated [20]. In addition to iron
metals such as iron, nickel, and cobalt and their ferromagnetic/ oxide–based NMs, other transition metal oxides such as TiO2 and
superparamagnetic alloys/oxides, and the outer layer contains ZnO have been used for the removal of Cd(II) ions from water [79,
polymers, organic compounds, silica, or alumina [20, 74] (Tables 1 80] (Table 1).
and 2). The novelty of hybrid inorganic magnetic NMs lies in the
fact that they can selectively treat heavy metals (e.g., Cd(II) ions) in 3.1.1 Iron oxide-based sorbents
wastewater, and their recovery upon treatment is made easy by the The most commonly used transition metal oxide NM is magnetite
application of an external magnetic field [75]. Selecting a suitable (Table 1). Magnetite allows easy functionalization for specific
support material during the synthesis of NMs is often crucial to the interactions with Cd(II) ions due to its superb magnetic characteristics
stability and facile regeneration of the system [20]. Interestingly, a and high specific surface area [81] (Table 1). The main advantage of
wide variety of supports (inorganic, organic, and hybrids of transition the magnetic NMs is their easy recovery by applying an external
elements and carbon-based systems (e.g., carbon nanotubes (CNTs) magnetic field upon completion of the sorptive treatment [73]. The
and graphene)) have been used in the fabrication of NMs [21]. recovered NMs can then be regenerated using a strong base or acid
(e.g., sodium hydroxide or hydrochloric acid) [82]. Nevertheless,
3.1 Transition metal oxide-based nanomaterials as Fe3O4 NMs are prone to corrosion and oxidation under very
Metal oxides of transition elements have been actively used in low pH conditions, they may require suitable functionalization or
wastewater treatment because of their high surface area, great modification of the parent material (e.g., modification with amino-
adsorption capacity for heavy metals, and easy reuse upon recovery methylene-phosphonic groups) to overcome the attack of strong
from the treated aqueous systems [76, 77]. Moreover, transition acids [83].
metal-oxides are particularly attractive for removing toxic metal The conventional synthesis methods for Fe3O4 NMs (e.g., sol-gel
species such as Cd(II) ions due to their eco-friendliness, easy method, co-precipitation, and micro emulsion) require huge
availability, and low price [78]. According to the analysis of the data amounts of expensive and toxic organic solvents (e.g., benzene and
summarized in Table 1, magnetite (Fe3O4) and other iron oxides hydrazine), making them highly unfavorable [84, 85]. In addition,
compose a large portion of the transition metal oxide-based NMs the commonly employed synthesis protocols are tedious and

| www.editorialmanager.com/nare/default.asp
Nano Res. 5

arduous to downgrade the practicality [86]. Also, iron oxide NMs sorption capacity) [94] (Table 1). Interestingly, the CoFe2O4/SiO2
succumb to agglomeration under van der Waals forces during NMs were exceptionally reusable (up to 45 cycles) because the silica
separation from the bulk phase, resulting in large particles [87]. To coating protected the iron core from corrosive eluents [95]. Similar
overcome the latter limitation, organic or inorganic surfactants and to silica (SiO2, a non-negatively charged compound that requires no
hydrophilic polymers are often used as capping or dispersion agents coupling ions), silicates (SiO44−, compounds bearing an anion and at
[87]. However, most of those agents are also highly toxic. As a least one cation with silicon as the central atom and electronegative
consequence, green synthesis approaches for magnetic NMs have groups as the ligands) have attracted widespread attention in
been explored extensively [88, 89]. Tangerine peel extract effectively wastewater treatment technologies because of their environmental
induced co-precipitation for the synthesis of Fe3O4 NMs [90]. The benignity, low cost, and high stability [96].
tangerine peel extract also effectively reduced NM agglomeration, Mesoporous one-dimensional magnesium silica nanotubes
reducing the NM size from 200 to 50 nm as the concentration (MgSNTs) have shown good capacity for the adsorption of heavy
of tangerine peel extract increased from 2% to 6%. Overall, the metallic ions due to their high mechanical sturdiness and short
performance of those green NMs was comparable with that of diffusion pathway for the target analyte [97]. For instance,
other NMs (e.g., 88.7% removal efficiency and 10.9 mg·g−1 sorption ferromagnetic γ-Fe2O3/MgSNTs with a high specific surface area
capacity) [90] (Table 1). and porous nanotube structure displayed a maximum sorption
The functionalization of an NM plays a pivotal role in maximizing capacity of 200 mg·g−1 for Cd(II) ions [98] (Fig. 4). The surface of
its sorption capacity toward Cd(II) ions [20]. For instance, the the synthesized γ-Fe2O3/MgSNTs was seen to bear a negative charge
performance of pristine Fe3O4 NMs was moderately good (e.g., 90% density that electrostatically attracted the cationic Cd(II) ions [98].
removal efficiency and 9.68 mg·g−1 sorption capacity) but poorer Cation exchange between Mg2+ ions and Cd(II) ions present on the
than that of many functionalized transition metal oxide NMs [91] surface of the γ-Fe2O3/MgSNTs was primarily responsible for the
(Table 1). Therefore, post-synthetic alterations (e.g., fixing specific abatement of Cd(II) ions [95].
functional groups (such as thiols) on the residual functionalities The Cd(II) ions in aqueous systems can exist in diverse forms
(such as hydroxyls) present on the iron oxide NM surface) have (Cd(OH)+, Cd(OH)3–, Cd(OH)2, and Cd2+) depending on the
been widely explored to enhance the affinity of iron-oxide NMs solution pH [91]. For the sorptive abatement of Cd(II) ions via iron
toward Cd(II) ions [92, 93]. In principle, the presence of functional oxide NMs, the use of acidic conditions (pH below 2) should be
groups such as –SO3H, –SH, and –OH on the adsorbent surface avoided. At such acidic pH, H+ cations might compete with the
may facilitate cation exchange and complexation reactions with positively charged Cd(II) ions to occupy functional sites on the NM
Cd(II) ions to promote their removal processes (Eqs. (1)–(3)). In surface [21]. Likewise, basic conditions (pH > 8) are suspected to
this regard, a free radical polymerization approach was reported to hamper the sorption process through the formation of insoluble
anchor SO3H groups on magnetite NMs (Fe3O4-SO3H) [82] (Fig. 2(a)). cadmium hydroxides [82]. Fe3O4-SO3H NMs possessing –OH,
The Fe3O4-SO3H NMs demonstrated good performance toward –SO3H, and –SH groups exhibit the best performance in neutral or
Cd(II) ion removal (94% removal efficiency, 80.9 mg·g−1 sorption alkaline pH (Eqs. (1)–(3)) (Fig. 2(b)).
capacity with recyclability for up to 10 cycles), which was attributed
to complexation between the sulfo functionalities and cadmium 2 [Fe3O4-SO3H](s) + Cd2+(aq) → [(Fe3O4-SO3H)2Cd](s) + 2 H+ (aq)
species [82] (Table 1). (1)
Magnetite NMs functionalized with cashew nut shell resin (CNSR)
provided a cost effective option that simultaneously used waste 2 [Fe3O4-OH](s) + Cd2+(aq) → [(Fe3O4-O)2Cd](s) + 2 H+(aq)
biomass (Fig. 3) [92]. The CNSR-Fe3O4 NMs had good performance (2)
toward Cd(II) ions (99.9% removal efficiency and 54.6 mg·g−1
sorption capacity), which was attributed to enhancement of the 2 [Fe3O4-SH](s) + Cd2+(aq) → [(Fe3O4-S)2Cd](s) + 2 H+(aq)
surface active sites for Cd(II) ion sorption [92] (Table 1). Hematite (3)
(αFe2O3) was doped with 15 wt.% Ni to produce Ni-αFe2O3 NMs
that displayed good performance for the abatement of Cd(II) ions 3.1.2 Miscellaneous
(98% removal efficiency and 90.91 mg·g−1 sorption capacity) [93] Along with iron-oxide-based NMs, other transition metal oxide-
(Table 1). based NMs (e.g., TiO2, ZnO, CuO, and NiO) have also been explored
Magnetic CoFe2O4 NMs were coated with SiO2 because silica is for the abatement of Cd(II) ions (Table 1). Diverse synthesis routes,
widely known to be biocompatible and chemically inert, making it such as the sol-gel approach, homogeneous solution precipitation,
a great support material [94]. The CoFe2O4/SiO2 NMs displayed a and reverse micellar routes, have been applied to the production
moderate tendency for Cd(II) ion sorption compared with other of transition metal oxide-based NMs [99, 100]. However, new and
transition metal oxide NMs (95% removal efficiency and 5 mg·g−1 advanced synthesis methods need to be developed to prevent the

 
Figure 2 Adsorption of Cd(II) ions by Fe3O4-SO3H NMs. (a) Schematic preparation of Fe3O4-SO3H NMs. (b) Effect of pH on Cd(II) ion sorption. (Reproduced with
permission from [82], © Elsevier 2017).

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


6 Nano Res.

efficient removal of heavy metal ions through ion exchange [95]. For
that reason, ZnO nanorods were synthesized via a hydrothermal
process, an ecofriendly, economical, catalyst-less growth process
that provided easy control over the NP morphology [79]. The high
removal capacity of the ZnO nanorods against Cd(II) ions (147.25 mg·g−1
and 91% removal efficiency) was attributed to complexation and
ion exchange [79] (Table 1). CuO nanoblades demonstrated good
performance for the removal of Cd(II) (99.2% removal efficiency
and 192.3 mg·g−1 sorption capacity), with lower viability than the
control (Fig. 5(a)) [105]. Titanium hybrid nanocomposites (Ti-C)
formed by functionalizing TiO2 NPs with 1,5 diphenyl-carbazone
exhibited a maximum sorption capacity of 31.7 mg·g−1 for Cd(II) ions
[106] (Table 1). Likewise, TiO2 dandelion NPs were synthesized by
critically managing reaction conditions such as pH, pressure, the
control agent for preventing aggregation, and the alcoholysis rate
[80]. The excellent sorption capacity (396 mg·g−1 and 80% removal
 
Figure 3 Preparation of CNSR-functionalized Fe3O4 nanoparticles. (Reproduced
efficiency) of the TiO2 dandelion NPs was attributed to the presence of
with permission from [92], © Elsevier 2017). rich hydroxyl functionalities on the surface, which provided a large
number of chelating sites to coordinate with Cd(II) ions [80, 95].
The past decade has produced prolific growth in the use of
adsorbents analogous to Prussian blue (CkM3[M(CN)6]2·nH2O, in
which M is a transition metal and C is an alkali ion) because of
their good sorption capabilities for wastewater treatment [66]. Also,
various chelating polymers such as polyaniline (PANI, rich in nitrogen
atoms bearing positive charges) are being increasingly applied
for sorptive removal due to their high selectivity toward heavy
metals, great reusability, and ease of handling [16, 107]. PANI was
synergistically combined with cobalt hexacyanoferrate (a Prussian
blue analogue) to form PANI-CoHCF nanocomposites that displayed
average sorption capacity (27.17 mg·g−1) for the sorptive removal of
Cd(II) ions (64.2% removal efficiency) (Fig. 6) [108].
For the proper application of NM technology, the cytotoxic
characteristics of a given NM need to be specified [109]. Because
the size of most NMs is comparable to that of biological cells, NMs
can penetrate biological systems and result in severe toxicity [110]. For
  example, transition metal oxide-based NMs induce great oxidative
Figure 4 The programmed synthesis of magnetic magnesium silicate nanotubes stress in biological cells (e.g., mitochondria) through the formation
(MgSNTs) from carbon nanotubes (CNTs). TREG = triethylene glycol, TEOS = of reactive oxygen species (e.g., OH and OOH free radicals) [111].
tetraethoxysilane. (Reproduced with permission from [98], © Wiley-VCH 2013). Such oxygen species can disrupt DNA, alter essential protein
molecules, and target the central nervous system [112]. As a con-
agglomeration of the produced NMs and control their morphology sequence, the optimum amounts of NMs recommended for wastewater
through simple reaction conditions such as temperature and reactant treatment applications boost their agglomeration, which reduces
concentrations [89].
Interestingly, metal oxalates have often been used for the synthesis
of complex and simple metal oxides because they have the tendency
to decompose at fairly low temperatures [101]. In this way, NiO
NPs were synthesized from nickel oxalates prepared via a single
stage solvothermal approach without toxic surfactants [102]. The
NiO NPs demonstrated good performance with a peak sorption
capacity of 97 mg·g−1 for Cd(II) ions [102] (Table 1). On a similar
note, orange peels (rich in pectin, hemi-cellulose, cellulose, carboxyls,
and hydroxyls) were powdered and synthesized into magnetic NPs
(MNP-OPP) using a co-precipitation approach with magnetite [103].
The MNP-OPP displayed good performance for Cd(II) ions (96%
removal efficiency and 76.92 mg·g−1 sorption capacity), which was
attributed to the involvement of ion-exchange and complexation
(Table 1) (Eq. (4)).
–Fe–O–(C=O) –(OH)n + Cd2+ → –Fe–O–(C=O)–(O–Cd)n+ + n H+
(4)
 
Biocompatible and chemically inert ZrO2 NPs rich in hydroxyl Figure 5 Adsorption of Cd(II) ions by CuO nanoblades and multi-walled carbon
functionalities exhibited average performance in the abatement of nanotubes. (a) Effect of adsorbent concentration on Cd removal by CuO nanoblades.
Cd(II) ions from water (99.9% removal efficiency and 3.18 mg·g−1 (b) Cytotoxicity analysis of synthesized CuO nanoblades. (c) Effect of solution pH
sorption capacity), with regeneration for up to 3 cycles [104] (Table 1). on percentage removal of cadmium by multi-walled carbon nanotubes. (Reproduced
ZnO-based NMs possessed hydroxyl-rich functionalities for the with permission from [105, 148], © Elsevier 2017).

| www.editorialmanager.com/nare/default.asp
Nano Res. 7

NMs [120]. For that reason, a nanocomposite designated as NZP was


developed by impregnating ZrO2·nH2O into a commercial cation
exchanger (D-001) made up of a highly porous polystyrene framework
[121]. The maximum sorption capacity of NZP toward Cd(II) ions
(214.7 mg·g−1) was attributed to the efficient dispersion of the ZrO2 NPs
in the crosslinked and nanoporous framework of the polymeric
cation exchanger [121] (Table 1).

3.2 Transition metal-based nanomaterials


A major portion of inorganic NMs are transition metals used
for the treatment of wastewaters rich in heavy metals [20, 95].
Interestingly, transition metal–based NM composites are relatively
inert on the bulk scale (because of relativistic contractions, molecular
orbital theory, and the electrochemical potential on the nanoscale),
thus elevating their practical usability [21, 122]. The great specific
surface area of transition metal-based NPs can be effectively
exploited for the abatement of water pollutants by dispersing those
NPs in the treatment system (homogeneous chemical approach)
[123]. However, that approach is impractical because recovering the
spent NPs is difficult [124]. Therefore, a heterogeneous chemical
approach is often applied in which transition metal-based NPs are
synergistically coupled with suitable support materials such as oxides
and polymers [78]. Interestingly, such synergistic combinations
elevate the number of active surface functionalities on the NM
surface and strengthen the affinity of the NM toward the target
pollutants [125]. Also, the support material adds stability and prevents
NPs from leaching into the water system (nanoscale particles are
  highly unstable because of their very high surface energy) [126].
Figure 6 Schematic of the formation of PANI-CoHCF nanocomposites. Zerovalent iron (nZVI) NPs have often been used to remove
(Reproduced with permission from [108], © Taylor & Francis, 2016). heavy metallic ions from aqueous solutions [127, 128]. nZVI NPs
typically consist of an Fe(0)-based core and an iron oxide-based outer
their cytotoxicity but also deteriorates their performance [105]. For layer [74]. The performance of nZVI NPs was moderate for the
example, high concentrations of CuO nanoblades (0.75–1.0 mg·mL−1) abatement of Cd(II) ions (78% removal efficiency and 188 mg·g−1
were recommended for the abatement of Cd(II) ions from water- sorption capacity) [129] (Table 1). Interestingly, because cadmium
based systems to conform with cytotoxicity standards (Fig. 5(b)). has more positive standard electrode potential (E0, E10 = −0.403 V)
The agglomerated NMs do not disperse easily in aqueous solutions, than iron (E20 = −0.447 V), Cd(II) is readily reduced to Cd(0) upon
but they also do no not penetrate biological cells [105]. interaction with iron-based NPs [130]. A sharp elevation in the
From a practical perspective, it is essential to understand the abatement of Cd(II) ions by nZVI NPs was found at pH ≥ 8.5 (Fig. 7)
interactions between NMs and natural organic matter, such as [129]. Essentially, at pH values higher than 8.0, the Cd(II) ions
humic acid (HA), which is abundant in natural aquatic environments precipitate out as Cd(OH)2 and are also hydrolyzed to form Cd2(OH)3+
[109]. Interactions between NMs and HA can involve cation bridging, and Cd(OH)+ ions [22]. The formation of those cadmium species
anion exchange, hydrophobic interactions, ligand exchange, hydrogen might imply a heightened deposition and electrostatic attraction
bonding, and entropic effects [113]. Essentially, HA can reduce the with nZVI NPs [131]. The isoelectric point of nZVI (around pH 8.1)
number of active sites present on the NM surface or increase the plays a vital role in the abatement of Cd(II) ions. As the nZVI
affinity of NMs toward the heavy metals, increasing their dispersibility surface becomes negatively charged (above the isoelectric point),
by reducing the size of aggregate NM clusters [114, 115]. Nano-TiO2 electrostatic interactions with the Cd(II) ions increase (Fig. 7) [129].
particles were coated with HA via a solution precipitation-based The practicality of using pristine iron NPs for the abatement
approach [116]. Interestingly, the HA-TiO2 NPs demonstrated a of Cd(II) ions is largely limited by their chemical instability in the
maximum sorption capacity of only 0.36 mg·g−1, the smallest capacity presence of other commonly encountered ions in wastewater [126].
value reported for any NM examined for the abatement of Cd(II) For example, Cl− ions severely inhibit the sorption of Cd(II) ions
ions [116] (Tables 1 and 2). onto nZVI [132]. Therefore, sulfidation of iron-based NPs by
The practical application of most transition metal oxide NMs is synthesizing Fe(0)/FeS (S-nZVI) nanocomposites has been touted
limited in wastewater treatment because their ultrafine size causes as a good option to improve their stability and performance toward
elevated pressure drops and poor mechanical sturdiness in fixed heavy metal removal [133, 134]. In this regard, nano-SiO2-seeded
beds and similar setups [20, 117]. However, this drawback can be S-nZVI nanocomposites (FeSSi) showcased a peak adsorption capacity
curbed by impregnating the NMs into materials with larger pores, of 105 mg·g−1 for Cd(II) ions, which was achieved through precipitation
such as zeolites and activated carbon [118]. Such hybrid systems and chemisorption [135] (Table 1).
provide better permeability and an easy separation of phases in In recent years, the great potential of transition metal dichalcogenides
flow systems [20]. Polymer-based ion exchangers are also a great (TMDs, e.g., MoSe2, WSe2, and WS2) has been recognized in various
possibility for hybrid systems because they possess both sufficient environmental applications, such as the sensing and removal of
rigidity (due to a polymeric backbone) and the Donnan membrane hazardous pollutants [118, 136]. With an MX2 stoichiometry and
effect (due to the presence of covalently bonded charged functionalities two dimensional structure, TMDs are made up of hexagonal sheets
on the polymer surface) [21, 119]. Such polymeric ion exchange of transition metals (M) packed between two layers of chalcogen
systems can provide effective preconcentration and elevated species (X) [136]. Interestingly, nano MoS2 tends to exhibit very
permeation of Cd(II) ions prior to their removal by metal oxide high chemical versatility coupled with great dispersibility (no

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


8 Nano Res.

the sorptive abatement of Cd(II) ions because of their high surface


area, chemical inertness, high porosity framework with many voids,
π-π electrostatic interactions, light mass density, and strong binding
tendency toward Cd and other heavy metal ions due to their rich
surface functionalities with great hydrophobicity [20, 145]. Pristine
MWCNTs exhibited good performance toward Cd(II) ions (181.8 mg·g−1
and 95.1% removal efficiency) [148] (Table 1). The removal efficiency
for MWCNTs increased as the solution became more alkaline
(Fig. 5(c)). As explained above, under acidic conditions, H+ ions
compete with the Cd(II) ions for the active surface sites. Also, at
low pH, CNTs exist in carboxylated or hydroxylated forms that are
unfavorable for cadmium abatement [147, 148].
To improve the removal characteristics of CNTs, the pristine
forms are often doped with specific functionalities that favorably
  attract the target pollutant(s) [147]. For example, nitrogen-doped
Figure 7 Cd(II) removal rate of nZVI at different pH values with and without the CNTs (CNx) obtain structural defects (heptagons and pentagons) in
addition of nitrate. (Reproduced with permission from [129], © Elsevier 2014).
the structural framework or vacancies pertaining to the hexagonal
network of carbon when they undergo oxidative treatments with
requirement for an assisting surfactant) [137]. Unique conjugated nitrogen oxide, ozone, or oxygen at temperatures ≤ 400 °C during their
systems with cyclic structures have allowed porphyrin derivatives synthesis [149, 150]. Those defects then favorably accommodate
to be used in the synthesis of nanocomposites, improving their surface functionalities and provide reactive anchoring sites for
removal characteristics for heavy metals [16, 138]. For example, Cd(II) ions [131, 151]. Because the presence of localized electrons
MoS2 was synergistically combined with 5, 10, 15, and 20-tetrakis from nitrogen atoms in the CNT framework helps distorts its π
(1-methyl-4-pyridinio) porphyrin tetra (p-toluenesulfonate) (TMPyP) cloud, its chemical reactivity can be greatly improved [134].
to form MoS2/TMPyP nanocomposites that displayed a maximum To increase the dispersibility of CNTs in aqueous solutions,
removal efficiency of 91% for Cd(II) ions [139] (Table 1). As various oxidative treatments can reduce their hydrophobicity, which
electrostatic interactions between the TMPyP on the MoS2 surface originates from intermolecular van der Waals attractions, while
and Cd(II) ions increased, cadmium removal proceeded efficiently purifying and chemically altering their structural framework by
through complexation [138]. introducing lactone, hydroxyls, carboxyls, carbonyls, pyrones, and
Templates are widely used in the synthesis of NPs [140]. Interestingly, chromenes [26, 128]. CNx demonstrated a peak adsorption capacity
NPs synthesized via templates showed excellent dispersibility in of 7.86 mg·g−1, which was attributed to the conjugation of cadmium
aqueous solutions, which systematically increased the accessibility species with oxygen functionalities on the CNx surface [149] (Table 1).
of the NPs to Cd(II) ions [141]. Among the many templates available In a similar study, CNx oxidized with nitric acid demonstrated a
for NP synthesis, biological ones have attracted great attention because peak sorption capacity of 9.33 mg·g−1 for Cd(II) ions because of
they can allow the resulting NPs to specifically recognize molecules abundant carboxylic functional groups on the CNT surface [152]
by imparting certain favorable spatial organizations during growth (Table 1). In a similar way, MWCNTs were treated with nitric acid
(distinct binding interactions, nucleation, and assembly) [142, 143]. to yield MWCNT-COOH, which showed an average performance
Also, the biological templates provide good control over the size and in the removal of Cd(II) ions (16.34 mg·g−1 and 96% removal
shape of NPs and increase the number of favorable functionalities efficiency) [153] (Table 1). The sorptive removal of cadmium by
[142]. For example, Aspergillus versicolor mycelia, an inexpensive MWCNT-COOH was explained as the chelation between cadmium
fungus with easy handling and a rapid growth rate, was functionalized species and the hydroxyl/carboxyl functionalities in MWCNT-
with xanthate, which has low solubility and high stability with heavy COOH [153].
metal complexes, in the synthesis of CdS NPs that demonstrated CNT structures were also grown on powdered activated carbon
a maximum Cd (II) ion sorption capacity of 141.5 mg·g−1 [144] via a fixed catalyst chemical vapor deposition approach to yield
(Table 1). a synergistic hybrid material (CNT-PAC) [154]. The CNT-PAC
displayed an average performance in the removal of Cd(II) ions
3.3 Carbon-based nanomaterials
(11.16 mg·g−1 with 98.35% removal efficiency) [154] (Table 1). Amino
NMs based on carbon have garnered widespread attention in recent groups are often incorporated into the CNT structure to boost its
years as prospective platforms for the abatement of heavy metals affinity toward specific target pollutants [128] and have demonstrated
from water-based systems [145, 146]. Graphene and CNTs are two great reactivity with heavy metals such as Cd(II) ions for effective
major components of carbon-based NMs for the abatement of Cd(II) removal. The amino molecules were synergistically combined with
ions because of their great surface area, tunable functional sites, the oxygen functionalities of CNTs to form active electrostatic and
many inherent oxygen functionalities, and resistance toward corrosion chelation sites on the CNT surface [15, 147]. For example, amino
[26, 128]. However, advanced carbon-based nanocomposites work polyethylene glycol (PEG) was anchored on MWCNTs via oxidation
better for the abatement of Cd(II) ions than their pristine forms followed by SOCl2 treatment and amination to yield hybrid
primarily due to synergistic effects and the presence of highly nanocomposite PEG-MWCNTs (Fig. 8) [155]. The PEG-MWCNTs
favorable surface functionalities [20, 24, 147]. efficiently removed Cd(II) ions (97 mg·g−1 with 77.6% removal
efficiency) from an aqueous solution [155] (Table 1).
3.3.1 Carbon nanotube-based sorbents
Along with oxidative strategies, the introduction of specific
CNTs are essentially graphene sheets rolled up in a tube-like functionalities has been touted as a good strategy for increasing the
structure [147] and can be formed from a single layer of graphene dispersibility of CNTs [156]. Also, the recovery of spent CNTs is a
(single-walled, SWCNTs) or multiple layers of graphene (multi- major issue that can be effectively addressed by designing magnetic
walled, MWCNTs) [26]. CNTs display great mechanical sturdiness, CNT composites. The recovery of such composites can be carried out
primarily through the presence of sp2 hybridization carbons, as efficiently by applying an external magnetic field [26]. For example,
compared to the sp3 bonds prevalent in other carbonaceous substances maghemite NPs were fixed on the surface of MWCNTs treated with
[128]. MWCNTs have received more attention than SWCNTs for nitric acid via coprecipitation [157]. The γ-Fe2O3/MWCNT composites

| www.editorialmanager.com/nare/default.asp
Nano Res. 9

 
Figure 8 Reaction pathways applied to obtain PEG-MWCNTs. (Reproduced with permission from [155] , © Forum of Chalcogeniders, 2013).

demonstrated good performance toward Cd(II) ions (78.81 mg·g−1 functionalized with cysteine, a water soluble, non-essential amino
and 97.2% removal efficiency), with easy recovery and regeneration acid bearing –COOH, –NH2, and –SH functionalities, to boost the
capability for up to 5 cycles [157] (Table 1). Essentially, as soon as affinity of MGO for Cd(II) ions [163]. The Cys-MGO nanosheets
the Fe2O3 NPs came in contact with water molecules, they used demonstrated a peak sorption capacity of 30.3 mg·g−1 for Cd(II)
hydroxyl groups to complete their shells (strong dependency on ions with easy regeneration for up to 10 cycles [163] (Table 1).
the solution pH). The surface of the maghemite thus became rich Interestingly, the cys-MGO nanosheets displayed great selectivity
in O−, OH2, and OH functionalities that acted synergistically with for Cd(II) ions in the presence of potentially interfering ions
the inherent carboxyl groups of the MWCNTs to efficiently adsorb commonly encountered in wastewater (e.g., Pb2+, Zn2+, Ni2+, Cu2+,
Cd(II) ions [20, 145]. Cr3+, Fe3+, Ca2+, Mg2+, Na+, and K+), supporting their enhanced
practical applicability [163]. In a similar approach, MGO NPs
3.3.2 Graphene-based sorbents
were functionalized with ethylene diamine to yield hybrid nano-
Graphite oxide (GrO) organized in a monolayered fashion is known composites (MDFGO) that displayed a maximum removal efficiency
as graphene oxide (GO) [158]. GO is typically synthesized via of 99.4% for Cd(II) ions [164] (Table 1). Similarly, MGO NPs were
sonication or mechanical stirring of exfoliated GrO [140]. The GO functionalized with 3-mercaptopropyl trimethoxysilane to yield hybrid
surface is rich in oxygen functionalities such as epoxies, hydroxyls, nanocomposites (M-Fe3O4-GO) with a peak sorption capacity of
carbonyls, and carboxyls [158]. Interestingly, the GO surface 125 mg·g−1 for Cd(II) ions [165] (Table 1). The excellent removal
acquires a negative surface-charge density in aqueous solutions due performance of M-Fe3O4-GO was attributed to complexation between
to deprotonation of the –COOH and –OH functionalities on its edge the GO surface and the Cd(II) ions due to strong Lewis acid-base
[95]. Essentially, the negative surface-charge density of GO NPs interactions along with ion-exchange due to the presence of thiol
enables them to be dispersed homogenously in aqueous solutions groups in the nanocomposite structure [128, 165].
due to electrostatic repulsions between them [21]. Moreover, An alkaline hydrothermal approach was used to synthesize alkaline
the negative charge on the GO surface makes GO NPs excellent deoxygenated GO (aGO) that had a peak adsorption capacity
candidates for the sorptive removal of positively charged pollutants of 156 mg·g−1 for Cd(II) ions [166] (Table 1). The electrostatic
(e.g., Cd(II) ions) through electrostatic attraction [26]. attraction between Cd(II) ions and the negatively charged aGO
The excellent potential of GO in the removal of various water surface (containing –CO− due to deprotonation in the aqueous
pollutants, including Cd(II) ions, has been recognized, along with solution) was described as the primary abatement mechanism, with
its numerous other merits (e.g., eco-friendliness, high specific cationic π interactions between the Cd(II) ions and benzene rings/
surface area (≥ 1,000 m2·g−1), and easily tailorable functionalities) phenolic hydroxyls also playing a major role (Fig. 9) [166]. Note that
[128]. However, like several NMs discussed above, GO suffers from various porous carbonaceous materials are used for the removal of
a serious shortcoming in terms of recovering spent NPs [159]. heavy metals. Among them, biochar (produced by the pyrolysis
Therefore, magnetic GO (MGO) was proposed and synthesized via of waste biomass) has attracted widespread attention as a cheap
coprecipitation with iron oxides [160]. MGO efficiently removed adsorbent [161, 167]. Because NPs are often combined with support
Cd(II) ions from water systems (91.29 mg·g−1 and 85% removal materials to prevent their elution from the treatment system [95],
efficiency) while maintaining regenerative capabilities for up to 3 cycles biochar has been used to stabilize GO NPs [168]. The GO-biochar
[160] (Table 1). nanocomposite (SG-PySA-GO) was synthesized by the pyrolysis
Interestingly, various amino acids have been extensively incorporated of sweetgum (Liquidambar styraciflua) pretreated with GO [169].
as ligands in the preparation of advanced materials for the abatement However, the SG-PySA-GO nanocomposite demonstrated only a
of pollutants because of their strong affinity toward heavy metal moderate peak sorption capacity of 10 mg·g−1 for Cd(II) ions. The
species, inexpensive availability, biocompatibility, and highly durable/ observed affinity was ascribed to complexation, cation exchange,
flexible structure [161, 162]. Therefore, MGO nanosheets were electrostatic attraction, and π interactions in the GO basal plane

 
Figure 9 Proposed adsorption mechanism of Cd(II) adsorption on alkaline deoxygenated GO (aGO). (Reproduced with permission from [166], © IWA Pub. 2015).

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


10 Nano Res.

[169] (Table 1). nanocellulose (NC), exhibits enhanced functional properties (especially
Calixarenes are third-generation supramolecular compounds hydrophilicity, biodegradability, regenerability, mechanical charac-
that have been widely used as building blocks and host molecular teristics, and specific surface area) compared to normal cellulose.
platforms for carriers of various metal ions as well as receptor NC can be prepared from a variety of raw materials (rice straw,
compounds for heavy metals [170, 171]. A wet chemical approach spinifex grass, poplar alkaline peroxide mechanical pulp, lignocellulosic
was used to functionalize thiacalix[4]arenetetrasulfonate (TCAS) fibers, and so on) [86, 179–181]. Rice straw nanofibers were prepared
on reduced graphene oxide (rGO) to yield the nanocomposite using a physical-chemical treatment [179]. The NC fibers prepared
TCAS-rGO, which demonstrated good performance for the from rice straw demonstrated better removal efficiency for Cd(II)
abatement of Cd(II) ions (128 mg·g−1 and 90% removal efficiency) than bulk cellulose and unprocessed rice straw. Their increased
[172] (Table 1) (Fig. 10). The Pearson acid-base concept was used adsorption of Cd(II) was ascribed to two major factors: (1) the removal
to explain the sorptive removal of Cd(II) ions by TCAS-rGO [172]. of non-cellulosic amorphous constituents during preparation of the
Essentially, the TCAS structure consisted of bridging sulfur atoms, NC fibers, which increased the availability of hydroxyl groups on
which are regarded as soft bases with great affinity for Cd(II) the cellulose chains, and (2) increasing the negative charge on the
ions [173]. surface of the NC fibers through acid hydrolysis, which enhanced
As elucidated in section 3.2, nZVI has been extensively used for the electrostatic interactions with cationic metals ions. The maximum
the sorptive abatement of heavy metals due to its great reactivity removal efficiency for the NC fibers was 90.7% (adsorption capacity
and large surface area [20]. However, nZVI NPs suffer from many of 9.7 mg·g−1) [179]. On the other hand, NC synthesized form
shortcomings, such as frequent agglomeration, poor stability, poor Australian spinifex grass exhibited a maximum removal efficiency
durability, and poor mechanical sturdiness [145]. To correct those of 84%, with a significantly high peak sorption capacity of 2,550 mg·g−1
shortcomings, nZVI was functionalized with rGO to form a for Cd(II) ions [180]. On the basis of Cd(II) ion concentrations,
nanocomposite (nZVI/rGO) with average performance for the Sharma, Chattopadhyay, Sharma, Geng, Amiralian, Martin and Hsiao
removal of Cd(II) ions (25.37 mg·g−1 and 76.11% removal efficiency) [180] proposed two types of removal mechanisms for Cd(II) by NC.
[174] (Table 1). At lower levels of Cd(II) ions (below 500 ppm), the interaction
Three-dimensional (3D) graphene-based aerogels have attracted between the metal ions and the carboxylic acid present on NC
great attention for the abatement of heavy metals from water systems dominated, whereas at higher levels of Cd(II) ions (above 1000 ppm),
mineralization (i.e., forming Cd(OH)2 nanocrystals) predominated.
[175]. These 3D aerogels possess a high specific surface area, high
The change in the surface morphology of the NC is shown via
strength-to-weight ratio, great mechanicalsturdiness, highly porous
scanning electron microscope (SEM) images in Fig. 11.
structure, and easily accessible active surface sites for the target
The effectiveness of NC in removing Cd(II) ions could be upgraded
pollutant [175, 176]. Therefore, a low-temperature gel, freeze-dried
further by the addition of support materials that can be easily
approach was used to covalently anchor sulfonic acid bearing aryl
functionalized with suitable ligands, e.g., sulfhydryl and carboxyl
radicals onto rGO [177]. The synthesized hybrid nanocomposite
groups [178]. The functionalities could be incorporated into the NC
(3D-SRGO) demonstrated good performance for the removal of Cd(II)
structure by modification with 2-mercaptobenzamide. Moreover,
ions (234.8 mg·g−1 and 93% removal efficiency) [177] (Table 1).
the structural properties, reusability, and functionality of NC
3.4 Organic polymeric nanomaterials/supported nano- were improved by adding iron nanoparticles to form magnetite
nanocellulose (MNCC). Further improvements in the structure
composites of MNCC were achieved by grafting itaconic acid (IA) with a
Many nanostructures have been tested for the abatement of Cd(II) crosslinking agent, e.g., ethylene glycol dimethacrylate. The final
ions from aquatic systems. The use of a biocompatible material such as structure is called 2-mercaptobenzamide modified IA-grafted-magnetite
cellulose or chitosan is considered beneficial in the development of nanocellulose composite (P(MB-IA)-g-MNCC). The interaction
heavy-metal removal systems because the raw materials are highly between the carboxyl and hydroxyl groups of the nanocomposite
abundant and inexpensive. Organic materials such as cellulose fit and Cd(II) was assumed to cause the removal of Cd(II) from the
those criteria [178–180]. In general, nanoscale cellulose, termed water system via ion exchange and complexation.

 
Figure 10 Schematic of synthesis procedures for the TCAS-rGO nanocomposites. (Reproduced with permission from [172]. © Royal Society of Chemistry 2016).

| www.editorialmanager.com/nare/default.asp
Nano Res. 11

 
Figure 11 Scanning electron microscope images of nanocellulose after exposure to (a) 500 ppm and (b) 1,000 ppm cadmium nitrate solution. At the higher concentration
of Cd ions, the nanocrystals of Cd were observed in the form of aggregates on the fibers due to mineralization. (Reproduced with permission from [180], © American
Chemical Society 2018).

The MNCC-based organic adsorption system demonstrated a material is special because it exhibited four routes of heavy
peak sorption capacity of 262.3 mg·g−1 for Cd(II) ions. The P(MB- metal immobilization: 1) surface complexation, 2) ion exchange,
IA)-g-MNCC followed pseudo-second-order kinetics for the sorption 3) dissolution, and 4) precipitation and co-precipitation [196]. The
of Cd(II) and fitted the Sips model well. The P(MB-IA)-g-MNCC adsorption properties of HAp could be enhanced by using a nanoform
sorption system displayed good performance during multiple uses, of hydroxyapatite (nHAp). The higher surface area of nHAp caused
decreasing from 97.2% to 86.3% after five cycles. The used adsorbents greater Cd(II) ion immobilization compared with normal HAp.
were regenerated using 0.1 M HCl. The maximum sorption amount of Cd by nHAp was 1.41 mmol·g−1
Chitosan is another biopolymer found abundantly in nature. (~ 158.5 mg·g−1), which is far higher than that from HAp (0.63 mmol·g−1
This organic structure exhibited several advantages, such as (~ 70.8 mg·g−1)) [196, 197]. Apatite materials, especially chlorapatite
biocompatibility, biodegradability, and chelation properties with (Ca10(PO4)6Cl2 (CLAP)), are very difficult to remove from water
metal ions via functional groups, e.g., amino and hydroxyl groups after treatment [194]. Therefore, the use of magnetic chlorapatite
[182]. However, a few modifications are suggested to improve (MNCLAP) NPs could be beneficial. The MNCLAP system was
the adsorption properties, stability, and reusability of chitosan as an successfully tested on the abatement of Cd(II) ions from wastewater,
effective adsorbent for metal ions. In general, adding magnetic and the spent structures were easily recovered from the water system
materials to chitosan would be a good option to enhance the with an external magnetic field [194].
reusability of the adsorption system [182, 183]. On the other hand, Alumina NPs are another promising nanostructure explored for
the adsorption characteristics of chitosan could be enhanced by the removal of Cd(II) ions from water [191, 195]. These structures
modifications with thiourea, isatin, and xanthate groups as well were explored as both NPs and nanotubes. The alumina NPs
as chelating agents such as diethylenetriaminepentaacetic acid, synthesized using the sol-gel approach demonstrated a peak sorption
ethylenediaminetetraacetic acid (EDTA), and ethylene glycol capacity of 17.22 mg·g−1 [195]. The surface modification of those
tetraacetic acid [182, 184–188]. structures is critical for maximizing Cd(II) ion adsorption [191].
A removal system for Cd(II) ions was developed using cross- Alumina nanotubes were modified with polylysine and resorcinol
linked magnetic EDTA/chitosan/TiO2 (MECT) [182]. The MECT to achieve a peak sorption capacity of 220 mg·g−1 for Cd(II) ions.
nanocomposite followed the Langmuir model for the sorption of Likewise, the modification of mesoporous silica with a ligand
Cd(II) ions, exhibiting a peak adsorption capacity of 209.2 mg·g−1. (4-tert-octyl-4-((phenyl)diazenyl) phenol (TPDP)) enhanced the
The sorption capacity of this nanocomposite-based system dropped Cd(II) ion sorption capacity from 3.62 to 148.3 mg·g−1 [192]. The
to 90% after five cycles of regeneration. Likewise, hydroxyapatite/ excellent removal capacity of magnesium oxide (MgO) nanostructures
chitosan nanorods (nHApC) were also proposed and used for the was also tested for Cd(II) [193]. Those structures adsorbed Cd(II)
sorption of Cd(II) ions [189]. The nHApC composite exhibited a ions via electrostatic and ion exchange mechanisms. The MgO
peak sorption capacity of 122 mg·g−1 for Cd(II) ions. The formation nanostructures demonstrated a peak sorption capacity of 1,500 mg·g−1
of a biocomposite between chitosan and clay materials not only for Cd(II) ions.
increased the adsorption characteristics but also lowered the overall A wide range of nanomaterials has been tested for the abatement
cost [190]. Therefore, the excellent ion exchange property of of Cd(II) ions from aqueous systems (Tables 1 and 2). Those materials
vermiculite (VMT), a phyllosilicate, was explored by forming a have demonstrated promising sorption capacity for the development
composite with chitosan. The chitosan/vermiculite biocomposite of future heavy metal removal systems. Nevertheless, the final
(CTS-VMT) was synthesized using epichlorohydrin as the cross- performance cannot be assessed by simply considering adsorption
linking agent. The composite demonstrated a peak sorption capacity capacity. Therefore, in addition to adsorption capacity, we also analyzed
of 58.5 mg·g−1 for Cu(II) ions at pH 4. This chitosan-based biocomposite the distribution (or partition) coefficient of the Cd removal systems
followed the chelation mechanism for the sorption of heavy metallic summarized in this review to reduce bias stemming
ions. The biocomposite displayed excellent performance in terms of
reusability, maintaining > 90% of its adsorption efficiency after four 4 Comparison between nanomaterials and com-
uses, with HCl used for regeneration. The Cd(II) ion adsorption
parameters for the organic nanocomposites are given in Table 2. mercial adsorbents
According to kinetics studies, the adsorption capacity of a given
3.5 Other nanomaterials sorbent is directly and sensitively affected by the concentration (or
Nanomaterials are an efficient way to remove Cd(II) from polluted amount) of pollutant loaded onto the system. Loading higher
water. In addition to the nanomaterials described above, a few other concentrations of pollutants on a sorption system will cause the
nanomaterials, such as alumina NPs, magnesium oxide, meso- sorbates to occupy more sorption sites and yield a higher sorption
porous silica, hydroxyapatite (HAp)-based nanostructures, and capacity per given mass than loading lower concentrations. In this
their composites have also been used to remove Cd(II) ions from respect, the selection of an equilibrium sorption capacity is not a
water-based systems [191–196]. The exploration of the HAp-based useful tool for comparing sorption capacities determined at varying

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


12 Nano Res.

levels of pollutant concentration. To fairly evaluate the sorption also reported low Kp values. Overall, our literature survey indicated
strengths of different sorbents, we use the distribution coefficient (Kp). that NMs demonstrate better performance in capturing Cd2+ ions
The formula for Kp is the sorption capacity divided by the final than traditional materials because of their fascinating structures
concentration of sorbate remaining at the equilibrium (or maximal) and associated advantageous properties.
removal condition [198]. Because this procedure can reduce the ZrO2 performed almost 35.7 times better than zeolite for cadmium
bias of the concentration effect just described, an evaluation based removal. The cost-effectiveness of sorbents used to remove Cd(II)
on Kp values is meaningful irrespective of the diverse experimental ions is also important and does not necessarily trend with Kp values.
conditions used across many different studies. For example, according For instance, although the sorption capabilities of CuO nanoblades
to the data summarized in Table 1, the maximum sorption capacity and PEG-MWCNT are comparable, the former has much smaller
for Cd(II) ions (396 mg·g−1) was found when TiO2 dandelions were synthesis costs than the latter [105, 155], making it a better option
used as the sorbent [80]. However, when the relative patterns for each for Cd(II) removal.
sorbent were compared based on Kp, ZrO2 exhibited the maximum
capability of all the materials listed in Table 1. The calculated Kp values 5 Challenges in nanotechnology for Cd(II) ion
thus contrast sharply with those based on adsorption capacity. The
Kp value of ZrO2 was 358 mg·g−1·μM−1, whereas that for the TiO2 removal
dandelions was only 1.11 mg·g−1·μM−1. As shown by this comparison, A brief self-explanatory description is provided for the practical
ZrO2 is the best sorbent material for Cd(II) ions listed in Table 1. and/or proven advantages of NM-based technologies towards diverse
Interestingly, the transmission electron microscopy (TEM) image of fields of environmental applications with the emphasis on water/
ZrO2 confirmed that they are agglomerated with an irregular shape wastewater treatment in Fig. 13. As such, nanotechnology has
(~ 13 nm diameter) (Fig. 12(a)). Likewise, their SEM image also received great attention from the research and industrial community
supported such observations (Fig. 12(b)) [104]. Apart from ZrO2, the because of the infinite merits of NMs based on significantly
best adsorbents for the capture of Cd(II) ions in the aqueous phase enhanced porosity and surface flexibility that offer multifunctional
were 3D-SRGO (189 mg·g−1·μM−1) [177], CNSR-Fe3O4 (123 mg·g−1·μM−1) properties unavailable with conventional adsorbents (e.g., activated
[92], and polylysine-resorcinol alumina nanotubes (124 mg·g−1·μM−1) carbons, polymeric organic resins, silica gel, zeolites, molecular sieves,
[191]. Several sorbents have moderately strong adsorption capacities and alumina). However, because of low cost and facile production,
for the removal of Cd(II) ions. Although their synthesis procedures traditional sorbent materials continue to be used (with or without
and structural properties differ, the performance of many NMs (e.g., modification) for the removal of diverse pollutants. Despite their
Ni-αFe2O3, CNT, and-PAC) appears to be highly comparable to one limited efficacy, conventional adsorbents (e.g., activated carbons,
another. For example, we calculated Kp values in the range of < polymeric organic resins, silica gel, zeolites, molecular sieves, and
34–21 mg·g−1·μM−1 for Ni-αFe2O3 (25.6 mg·g−1·μM−1) [93], CuO alumina) are still commonly used for several reasons: (a) cost-
nanoblades (27 mg·g−1·μM−1) [105], CNT-PAC (33.5 mg·g−1·μM−1) [154], effectiveness, (b) easy availability, (c) no need for particular synthesis
PEG-MWCNT (29.1 mg·g−1·μM−1) [155], and γ-Fe2O3/MWCNT or activation, (d) recyclability, (e) easy disposability, (f) relatively low
(21.1 mg·g−1·μM−1) [98] (Table 1). All those sorbents had Kp values toxicity, and (g) no harmful end-product caused by sorbent–sorbate
that exceeded those of the traditional adsorbents: activated carbon interactions. Among those factors, cost-related issues (e.g., the
(0.15 mg·g−1·μM−1) [199], activated clay (0.04 mg·g−1·μM−1) [200], expense of synthesis and regeneration) remain the main limitation
perlite (0.16 mg·g−1·μM−1) [201], and zeolite (10.4 mg·g−1·μM−1) [202]. for NMs. Additionally, concern remains that some NMs could
Thus, the Kp values confirm the limited efficacy of the traditional generate toxicity either through sorbent–sorbate interactions or at
sorbents. Only a few NMs recorded such low Kp values: MT-CD/ the disposal stage.
SiO2 (0.03 mg·g−1·μM−1) [203] was even lower than activated clay To date, a few types of NMs have been synthesized in large
(0.04 mg·g−1·μM−1), and CNTs (0.10 mg·g−1·μM−1) [199], PANI/CoHCF enough quantities to lower the market price. For instance, the price
(0.09 mg·g−1·μM−1) [108], nZVI/rGO (0.59 mg·g−1·μM−1) [174], Fe3O4 of 1 kg of MWCNTs (> 90% carbon, dimension × length: 110–
(0.82 mg·g−1·μM−1) [90], and MWCNT-COOH (0.92 mg·g−1·μM−1) 170 nm × 5–9 μm) is 40,317 USD [204], whereas the price of 1 kg of

Table 3 Removal of Cd(II) ions from water through commercial adsorbents


Initial Cd Maximum Maximum Final Cd Distribution
Optimum adsorption
Order Adsorbent concentration removal adsorption capacity concentration coefficient Reference
condition (°C, pH)
(μM) efficiency (%) (mg·g−1) (μM) (mg·g−1·μM−1)
1 Activated carbon 25, 7 8.89 38 0.8 5.51 0.15 [199]
2 Zeolite 25, 6 62.27 96 25.9 2.49 10.40 [202]
3 Activated clay 30, 6 889.6 75.2 8.71 220.62 0.04 [200]
4 Perlite 22, 6 8.89 55 0.64 4.00 0.16 [201]

Figure 12 Microscopic images of ZrO2. Panel (a) and (b) correspond to transmission electron microscopy (TEM) image and scanning electron microscopy (SEM)
image, respectively. (Reproduced with permission from [104]. © Elsevier 2016).

| www.editorialmanager.com/nare/default.asp
Nano Res. 13

zeolite is 478 USD [205]. Although MWCNTs cost almost 100 times Further, as the organic NMs tend to be subject to relatively slow
more than zeolite, the performance of some functionalized MWCNTs uptake kinetics (e.g., requirement of a longer operation to achieve
(e.g., MWCNT-COOH) is found to be even less effective than that the desired level of removal relative to other NMs), it makes them
of zeolite (Table 1). The abnormally high cost of CNTs arises from less favorable for practical applications [20, 209]. In addition, as some
the usage of standard catalytic chemical vapor deposition-based organic polymers (e.g., chitosan) may dissolve under highly acidic
techniques which involve the breakdown of hydrocarbons (chemically conditions, the structural breakdown of the composite material can
induced) onto a substrate [24]. In addition to the high cost and be accompanied with the reduction of their performance [24, 210].
energy intensive production procedure, the catalysts used during For the optimal use of NMs, an accurate assessment of their
their synthesis often deposit on the CNT surface to leave potential life-cycle, from synthesis to regeneration and disposal, should be
risk to the environment (e.g., due to their toxicity) [23, 206]. Also, done. Unlike catalytic degradation, adsorption transfers a pollutant
because of the high van der Waals attraction, pristine CNTs have from one phase (pollutant contaminating a substance) to another
low solubility in water to undergo eventual agglomeration which may phase (pollutant loaded onto a sorbent). Thus, adsorption cannot
lower their performance with uneasy handling [23, 207]. In this provide complete mineralization of the pollutant, as catalytic
respect, the use of suitable modification (e.g., functionalization) degradation does. Only a few studies have considered the safe
may offer a solution to the problems associated with CNTs or other disposal or regeneration of adsorbents loaded with Cd(II) [211].
NMs [207, 208]. Therefore, despite the excellent performance of Among them, phytoremediation was one of the options considered;
certain NMs, they are not yet ready to meet the practical needs in the Cd(II)-loaded adsorbent was used as a fertilizer for plant growth
the field. Energy intensive production procedure, the catalysts used [211]. The Cd(II) enriched plants/trees were then used as fuel/bio-ore
during their synthesis often deposit on the CNT surface to leave in an industrial mining process. But phytoremediation and similar
potential risk to the environment (e.g., due to their toxicity) [23, 206]. options apply mainly to traditional adsorbents (zeolite, activated
Also, because of the high van der Waals attraction, pristine CNTs carbon, and bio-char) [212]. The toxicity of the functional moieties
have low solubility in water to undergo eventual agglomeration on modern adsorbents has not yet been fully investigated. Incineration
which may lower their performance with uneasy handling [23, 207]. In or pyrolysis of the spent sorbent might be required. For instance,
this respect, the use of suitable modification (e.g., functionalization)
using pyrolysis researchers recovered > 95% of the Cd(II) from a
may offer a solution to the problems associated with CNTs or other
spent bio-sorbent [213]. This high recovery rate arose from the low
NMs [207, 208]. Therefore, despite the excellent performance of
melting point (321 °C) of Cd(II), which propelled the conversion
certain NMs, they are not yet ready to meet the practical needs in
rate of Cd(II) into its gas and liquid forms [214]. However, only a
the field. In recent years, the utilization of organic polymeric NMs
small amount of information is available about the use of pyrolysis
has received numerous attention in the domain of water purification
to treat spent bio-sorbents [211]. Likewise, recycling strategies for
owing to their distinct advantages of biodegradability, biocompatibility,
NMs used to treat Cd(II) are still in their infancy. Consequently, the
and eco-friendliness [209]. Nonetheless, the performance of most
two main challenges to using NMs for the capture of Cd(II) are (a)
organic polymeric NMs is generally observed to be inferior to that
of inorganic NMs (Tables 1 and 2). the high cost of nanomaterials and (b) a lack of recycling or disposal
technologies for spent sorbents.

6 Conclusions and future prospects


We organized this review article to accentuate the practical
application of NMs to remove Cd(II) ions from water. Recent years
have witnessed a prolific rise in attempts to apply nanotechnology
to water management because of the unique advantages of NMs
(e.g., tailorable functionalities, great porosity, and high specific surface
area). To this end, we described the application of both organic and
inorganic NMs to the sorptive removal of Cd(II) ions. Essentially,
the efficiency of such abatement was sensitively affected by the
combined effects of the types of functional groups present on the
NM surface and the structural characteristics of nanocomposites
(e.g., short paths for the diffusion of the pollutant). Extensive
investigations have been made to functionalize NM surfaces with
target-specific functional groups to improve the affinity of NMs for
the target pollutant and boost the sorption capacity. The specific
functional groups immobilized on the NM surface should facilitate
the abatement of Cd(II) ions, primarily through ion-exchange and
complexation reactions involving soft–soft interactions.
The synthesis of nanocomposites has been regarded as an effective
and practical strategy for boosting the regenerative and sorptive
capabilities of NMs. The use of support materials such as silica is
a highly efficient option for stabilizing NPs in the aqueous phase
by distributing the excess surface energy. Also, many support
materials (e.g., GO, which bears oxygen-rich surface functional
sites) synergistically combine with the primary NM, introducing
new functionalities and upgrading the sorption performance. Most
reported NMs outperform traditional adsorbents (e.g., zeolites
and activated clay) for the abatement of Cd(II) ions. The excellent
Figure 13 A list of proven nanomaterial-based technologies towards environ- performance of NMs in removing Cd(II) ions from water justifies
mental applications [221–223]. the further research needed to enable their real-world application.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


14 Nano Res.

Despite their excellent performance, NMs suffer from some M. I. Assessment of biotransfer and bioaccumulation of cadmium, lead
significant limitations: extremely high synthesis/operating costs, and zinc from fly ash amended soil in mustard–aphid–beetle food chain.
challenges with recyclability/disposability, and the need for toxic Science of The Total Environment 2017, 584-585, 1221–1229.
substances during their preparation. Most of those challenges can [12] Bravo, D.; Pardo-Díaz, S.; Benavides-Erazo, J.; Rengifo-Estrada, G.; Braissant,
O.; Leon-Moreno, C. Cadmium and cadmium-tolerant soil bacteria in cacao
be resolved through the development of advanced and economic
crops from northeastern Colombia. J. Appl. Microbiol. 2018, 124, 1175–1194.
synthesis routes using green synthesis approaches and the [13] White, A. J.; O'Brien, K. M.; Jackson, B. P.; Karagas, M. R. Urine and
functionalization/modification of pure NMs. The technical hindrances toenail cadmium levels in pregnant women: A reliability study. Environ.
to the practical application of NMs for water treatment can be Int. 2018, 118, 86–91.
overcome through active collaborations among environmental [14] Idrees, N.; Tabassum, B.; Abd_Allah, E. F.; Hashem, A.; Sarah, R.;
scientists, material engineers, and chemists. Interdisciplinary research Hashim, M. Groundwater contamination with cadmium concentrations in
efforts are expected to propel nanotechnology toward real-world some West U.P. Regions, India. Saudi J. Biol. Sci. 2018, 25, 1365–1368.
applications. [15] Kobielska, P. A.; Howarth, A. J.; Farha, O. K.; Nayak, S. Metal–organic
frameworks for heavy metal removal from water. Coord. Chem. Rev. 2018,
358, 92–107.
Acknowledgements [16] Zare, E. N.; Motahari, A.; Sillanpää, M. Nanoadsorbents based on conducting
The authors acknowledge the support made by the R&D Center polymer nanocomposites with main focus on polyaniline and its derivatives
for removal of heavy metal ions/dyes: A review. Environ. Res. 2018, 162,
for Green Patrol Technologies through the R&D for Global Top
173–195.
Environmental Technologies funded by the Ministry of Environment
[17] Shakya, M.; Rene, E. R.; Nancharaiah, Y. V.; Lens, P. N. L. Fungal-based
(MOE 2018001850001) as well as a grant from the National Research nanotechnology for heavy metal removal. In Nanotechnology, Food
Foundation of Korea (NRF) funded by the Ministry of Science, ICT Security and Water Treatment. Gothandam, K. M.; Ranjan, S.; Dasgupta, N.;
& Future Planning (No. 2016R1E1A1A01940995). This work was Ramalingam, C.; Lichtfouse, E., Eds.; Springer International Publishing:
also supported by “Cooperative Research Program for Agriculture Cham, 2018; pp 229–253.
Science & Technology Development (Project No. PJ014297)” Rural [18] Burakov, A. E.; Galunin, E. V.; Burakova, I. V.; Kucherova, A. E.; Agarwal,
Development Administration, Republic of Korea. V. K. acknowledges S.; Tkachev, A. G.; Gupta, V. K. Adsorption of heavy metals on conventional
the support from the Department of Science and Technology, New and nanostructured materials for wastewater treatment purposes: A review.
Delhi, India, in the form of an INSPIRE Faculty Award. Ecotoxicol. Environ. Saf. 2018, 148, 702–712.
[19] Vikrant, K.; Kumar, V.; Kim, K.-H.; Kukkar, D. Metal-organic frameworks
(MOFs): Potential and challenges for capture and abatement of ammonia.
Notes J. Mater. Chem. A 2017, 5, 22877–22896.
[20] Lu, F.; Astruc, D. Nanomaterials for removal of toxic elements from water.
There are no conflicts to declare.
Coord. Chem. Rev. 2018, 356, 147–164.
[21] Azzaza, S.; Kumar, R. T.; Vijaya, J. J.; Bououdina, M. Nanomaterials for
References heavy metal removal. In Advanced Environmental Analysis: Applications
of Nanomaterials, Volume 1. The Royal Society of Chemistry, 2017; pp
[1] Borah, R.; Kumari, D.; Gogoi, A.; Biswas, S.; Goswami, R.; Shim, J.;
139–166.
Begum, N. A.; Kumar, M. Efficacy and field applicability of Burmese
[22] Ahmad, Z.; Gao, B.; Mosa, A.; Yu, H. W.; Yin, X. Q.; Bashir, A.; Ghoveisi,
grape leaf extract (BGLE) for cadmium removal: An implication of metal
H.; Wang, S. S. Removal of Cu(II), Cd(II) and Pb(II) ions from aqueous
removal from natural water. Ecotoxicol. Environ. Saf. 2018, 147, 585–593.
solutions by biochars derived from potassium-rich biomass. J. Clean.
[2] Davis, A. D.; Webb, C. J.; Sorensen, J. L.; Dixon, D. J.; Hudson, R.
Prod. 2018, 180, 437–449.
Geochemical thermodynamics of cadmium removal from water with
[23] Ihsanullah; Abbas, A.; Al-Amer, A. M.; Laoui, T.; Al-Marri, M. J.; Nasser,
limestone. Environ. Earth Sci. 2018, 77, 37. M. S.; Khraisheh, M.; Atieh, M. A. Heavy metal removal from aqueous
[3] Ilic, M.; Jovic, S.; Spalevic, P.; Vujicic, I. Water cycle estimation by neuro- solution by advanced carbon nanotubes: Critical review of adsorption
fuzzy approach. Comp. Electron. Agric. 2017, 135, 1–3. applications. Sep. Purif. Technol. 2016, 157, 141–161.
[4] Vikrant, K.; Kim, K.-H.; Ok, Y. S.; Tsang, D. C. W.; Tsang, Y. F.; Giri, B. [24] Vikrant, K.; Kim, K.-H. Nanomaterials for the adsorptive treatment of
S.; Singh, R. S. Engineered/designer biochar for the removal of phosphate Hg(II) ions from water. Chem. Eng. J. 2019, 358, 264–282.
in water and wastewater. Sci. Total Environ. 2018, 616–617, 1242–1260. [25] Ray, P. Z.; Shipley, H. J. Inorganic nano-adsorbents for the removal of
[5] Gustin, K.; Tofail, F.; Vahter, M.; Kippler, M. Cadmium exposure and heavy metals and arsenic: A review. RSC Adv. 2015, 5, 29885–29907.
cognitive abilities and behavior at 10 years of age: A prospective cohort [26] Xu, J.; Cao, Z.; Zhang, Y. L.; Yuan, Z. L.; Lou, Z. M.; Xu, X. H.; Wang, X.
study. Environ. Int. 2018, 113, 259–268. K. A review of functionalized carbon nanotubes and graphene for heavy
[6] Luca, F.-A.; Ciobanu, C.-I.; Andrei, A. G.; Horodnic, A. V. Raising awareness metal adsorption from water: Preparation, application, and mechanism.
on health impact of the chemicals used in consumer products: Empirical Chemosphere 2018, 195, 351–364.
evidence from east-central europe. Sustainability 2018, 10, 209. [27] Fu, F. L.; Wang, Q. Removal of heavy metal ions from wastewaters: A
[7] Zheng, W.; Xu, Y.-M.; Wu, D.-D.; Yao, Y.; Liang, Z.-L.; Tan, H. W.; review. J. Environ. Manag. 2011, 92, 407–418.
Lau, A. T. Y. Acute and chronic cadmium telluride quantum dots-exposed [28] Mirbagheri, S. A.; Hosseini, S. N. Pilot plant investigation on petrochemical
human bronchial epithelial cells: The effects of particle sizes on their wastewater treatmentfor the removal of copper and chromium with the
cytotoxicity and carcinogenicity. Biochem. Biophys. Res. Commun. 2018, objective of reuse. Desalination 2005, 171, 85–93.
495, 899–903. [29] Özverdi, A.; Erdem, M. Cu2+, Cd2+ and Pb2+ adsorption from aqueous solutions
[8] Jacquet, A.; Arnaud, J.; Hininger-Favier, I.; Hazane-Puch, F.; Couturier, K.; by pyrite and synthetic iron sulphide. J. Hazard. Mater. 2006, 137, 626–632.
Lénon, M.; Lamarche, F.; Ounnas, F.; Fontaine, E.; Moulis, J.-M. et al. [30] Tünay, O.; Kabdaşli, N. Hydroxide precipitation of complexed metals. Water
Impact of chronic and low cadmium exposure of rats: Sex specific disruption Res. 1994, 28, 2117–2124.
of glucose metabolism. Chemosphere 2018, 207, 764–773. [31] Matlock, M. M.; Henke, K. R.; Atwood, D. A. Effectiveness of commercial
[9] Belhaj, D.; Athmouni, K.; Ahmed, M. B.; Aoiadni, N.; El Feki, A.; Zhou, reagents for heavy metal removal from water with new insights for future
J. L.; Ayadi, H. Polysaccharides from Phormidium versicolor (NCC466) chelate designs. J. Hazard. Mater. 2002, 92, 129–142.
protecting HepG2 human hepatocellular carcinoma cells and rat liver [32] Matlock, M. M.; Howerton, B. S.; Van Aelstyn, M. A.; Nordstrom, F. L.;
tissues from cadmium toxicity: Evidence from in vitro and in vivo tests. Atwood, D. A. Advanced mercury removal from gold leachate solutions
Int. J. Biol. Macromol. 2018, 113, 813–820. prior to gold and silver extraction: A field study from an active gold mine
[10] Skipper, A.; Sims, J. N.; Yedjou, C. G.; Tchounwou, P. B. Cadmium chloride in Peru. Environ. Sci. Technol. 2002, 36, 1636–1639.
induces DNA damage and apoptosis of human liver carcinoma cells via [33] Abumaizar, R. J.; Smith, E. H. Heavy metal contaminants removal by soil
oxidative stress. Int. J. Environ. Res. and Public Health 2016, 13, 88. washing. J. Hazard. Mater. 1999, 70, 71–86.
[11] Dar, M. I.; Green, I. D.; Naikoo, M. I.; Khan, F. A.; Ansari, A. A.; Lone, [34] Matlock, M. M.; Howerton, B. S.; Atwood, D. A. Chemical precipitation

| www.editorialmanager.com/nare/default.asp
Nano Res. 15

of heavy metals from acid mine drainage. Water Res. 2002, 36, 4757–4764. combined sewer overflow: Heavy metal removal and treatment optimization.
[35] Fu, F. L.; Chen, R. M.; Xiong, Y. Application of a novel strategy— Water Res. 2008, 42, 951–960.
Coordination polymerization precipitation to the treatment of Cu2+-containing [58] Rubio, J.; Tessele, F. Removal of heavy metal ions by adsorptive particulate
wastewaters. Sep. Purif. Technol. 2006, 52, 388–393. flotation. Miner. Eng. 1997, 10, 671–679.
[36] Fu, F. L.; Zeng, H. Y.; Cai, Q. H.; Qiu, R. L.; Yu, J.; Xiong, Y. Effective removal [59] Blöcher, C.; Dorda, J.; Mavrov, V.; Chmiel, H.; Lazaridis, N. K.; Matis,
of coordinated copper from wastewater using a new dithiocarbamate-type K. A. Hybrid flotation—membrane filtration process for the removal of
supramolecular heavy metal precipitant. Chemosphere 2007, 69, 1783–1789. heavy metal ions from wastewater. Water Res. 2003, 37, 4018–4026.
[37] Da̧browski, A.; Hubicki, Z.; Podkościelny, P.; Robens, E. Selective [60] Aldrich, C.; Feng, D. Removal of heavy metals from wastewater effluents
removal of the heavy metal ions from waters and industrial wastewaters by biosorptive flotation. Miner. Eng. 2000, 13, 1129–1138.
by ion-exchange method. Chemosphere 2004, 56, 91–106. [61] Polat, H.; Erdogan, D. Heavy metal removal from waste waters by ion
[38] Erdem, E.; Karapinar, N.; Donat, R. The removal of heavy metal cations flotation. J. Hazard. Mater. 2007, 148, 267–273.
by natural zeolites. J. Colloid Interface Sci. 2004, 280, 309–314. [62] Merzouk, B.; Gourich, B.; Sekki, A.; Madani, K.; Chibane, M. Removal
[39] Wong, C.-W.; Barford, J. P.; Chen, G. H.; McKay, G. Kinetics and equilibrium turbidity and separation of heavy metals using electrocoagulation–
studies for the removal of cadmium ions by ion exchange resin. J. Environ. electroflotation technique: A case study. J. Hazard. Mater. 2009, 164,
Chem. Eng. 2014, 2, 698–707. 215–222.
[40] Ahmed, S.; Chughtai, S.; Keane, M. A. The removal of cadmium and lead [63] Belkacem, M.; Khodir, M.; Abdelkrim, S. Treatment characteristics of
from aqueous solution by ion exchange with Na-Y zeolite. Sep. Purif. textile wastewater and removal of heavy metals using the electroflotation
Technol. 1998, 13, 57–64. technique. Desalination 2008, 228, 245–254.
[41] da Fonseca, M. G.; de Oliveira, M. M.; Arakaki, L. N. H. Removal of [64] Bailey, S. E.; Olin, T. J.; Bricka, R. M.; Adrian, D. D. A review of potentially
cadmium, zinc, manganese and chromium cations from aqueous solution low-cost sorbents for heavy metals. Water Res. 1999, 33, 2469–2479.
by a clay mineral. J. Hazard. Mater. 2006, 137, 288–292. [65] Babel, S.; Kurniawan, T. A. Low-cost adsorbents for heavy metals uptake
[42] Sanchez, A. G.; Ayuso, E. A.; De Blas, O. J. Sorption of heavy metals from from contaminated water: A review. J. Hazard. Mater. 2003, 97, 219–243.
industrial waste water by low-cost mineral silicates. Clay Miner. 1999, 34, [66] Zhou, G. Y.; Luo, J. M.; Liu, C. B.; Chu, L.; Crittenden, J. Efficient heavy
469–477. metal removal from industrial melting effluent using fixed-bed process
[43] Barakat, M. A. New trends in removing heavy metals from industrial based on porous hydrogel adsorbents. Water Res. 2018, 131, 246–254.
wastewater. Arab. J. Chem. 2011, 4, 361–377. [67] Kobya, M.; Demirbas, E.; Senturk, E.; Ince, M. Adsorption of heavy metal
[44] Trivunac, K.; Stevanovic, S. Removal of heavy metal ions from water by ions from aqueous solutions by activated carbon prepared from apricot
complexation-assisted ultrafiltration. Chemosphere 2006, 64, 486–491. stone. Bioresour. Technol. 2005, 96, 1518–1521.
[45] Jakobsen, M. R.; Fritt-Rasmussen, J.; Nielsen, S.; Ottosen, L. M. Electrodialytic [68] Mureseanu, M.; Reiss, A.; Stefanescu, I.; David, E.; Parvulescu, V.; Renard,
removal of cadmium from wastewater sludge. J. Hazard. Mater. 2004, 106, G.; Hulea, V. Modified SBA-15 mesoporous silica for heavy metal ions
127–132.
remediation. Chemosphere 2008, 73, 1499–1504.
[46] Kheriji, J.; Tabassi, D.; Hamrouni, B. Removal of Cd(II) ions from aqueous
[69] Khraisheh, M. A. M.; Al-degs, Y. S.; Mcminn, W. A. M. Remediation of
solution and industrial effluent using reverse osmosis and nanofiltration
wastewater containing heavy metals using raw and modified diatomite.
membranes. Water Sci. Technol. 2015, 72, 1206–1216.
Chem. Eng. J. 2004, 99, 177–184.
[47] Cañizares, P.; Pérez, Á.; Camarillo, R. Recovery of heavy metals by
[70] Yan, G. Y.; Viraraghavan, T. Heavy-metal removal from aqueous solution
means of ultrafiltration with water-soluble polymers: Calculation of design
by fungus Mucor rouxii. Water Res. 2003, 37, 4486–4496.
parameters. Desalination 2002, 144, 279–285.
[71] Brown, M. J.; Lester, J. N. Metal removal in activated sludge: The role of
[48] Vijayalakshmi, A.; Arockiasamy, D. L.; Nagendran, A.; Mohan, D. Separation
bacterial extracellular polymers. Water Res. 1979, 13, 817–837.
of proteins and toxic heavy metal ions from aqueous solution by CA/PC
[72] Inyang, M. I.; Gao, B.; Yao, Y.; Xue, Y. W.; Zimmerman, A.; Mosa, A.;
blend ultrafiltration membranes. Sep. Purif. Technol. 2008, 62, 32–38.
Pullammanappallil, P.; Ok, Y. S.; Cao, X. D. A review of biochar as a
[49] Mbareck, C.; Nguyen, Q. T.; Alaoui, O. T.; Barillier, D. Elaboration,
low-cost adsorbent for aqueous heavy metal removal. Crit. Rev. Environ.
characterization and application of polysulfone and polyacrylic acid blends
Sci. Technol. 2016, 46, 406–433.
as ultrafiltration membranes for removal of some heavy metals from water.
[73] Akın Sahbaz, D.; Yakar, A.; Gündüz, U. Magnetic Fe3O4-chitosan micro-
J. Hazard. Mater. 2009, 171, 93–101.
and nanoparticles for wastewater treatment. Particul. Sci. Technol., in press,
[50] Adams, F. V.; Nxumalo, E. N.; Krause, R. W. M.; Hoek, E. M. V.; Mamba,
B. B. Preparation and characterization of polysulfone/β-cyclodextrin DOI: 10.1080/02726351.2018.1438544.
polyurethane composite nanofiltration membranes. J. Membrane Sci. [74] Ling, L.; Huang, X.-Y.; Zhang, W.-X. Enrichment of precious metals
2012, 405–406, 291–299. from wastewater with core–shell nanoparticles of iron. Adv. Mater. 2018,
[51] Landaburu-Aguirre, J.; Pongrácz, E.; Perämäki, P.; Keiski, R. L. Micellar- 30, 1705703.
enhanced ultrafiltration for the removal of cadmium and zinc: Use of response [75] Korina, E.; Stoilova, O.; Manolova, N.; Rashkov, I. Polymer fibers with
surface methodology to improve understanding of process performance magnetic core decorated with titanium dioxide prospective for photocatalytic
and optimisation. J. Hazard. Mater. 2010, 180, 524–534. water treatment. J. Environ. Chem. Eng. 2018, 6, 2075–2084.
[52] Camarillo, R.; Llanos, J.; García-Fernández, L.; Pérez, Á.; Cañizares, P. [76] Castro, L.; Blázquez, M. L.; González, F.; Muñoz, J. A.; Ballester, A. Heavy
Treatment of copper (II)-loaded aqueous nitrate solutions by polymer metal adsorption using biogenic iron compounds. Hydrometallurgy 2018,
enhanced ultrafiltration and electrodeposition. Sep. Purif. Technol. 2010, 179, 44–51.
70, 320–328. [77] Lee, S. C.; Jeong, Y.; Kim, Y. J.; Kim, H.; Lee, H. U.; Lee, Y.-C.; Lee, S. M.;
[53] Mukherjee, R.; Bhunia, P.; De, S. Impact of graphene oxide on removal Kim, H. J.; An, H.-R.; Ha, M. G. et al. Hierarchically three-dimensional
of heavy metals using mixed matrix membrane. Chem. Eng. J. 2016, 292, (3D) nanotubular sea urchin-shaped iron oxide and its application in heavy
284–297. metal removal and solar-induced photocatalytic degradation. J. Hazard.
[54] Ibrahim, G. P. S.; Isloor, A. M.; Inamuddin; Asiri, A. M.; Ismail, A. F.; Mater. 2018, 354, 283–292.
Kumar, R.; Ahamed, M. I. Performance intensification of the polysulfone [78] Bagheri, S.; Aghaei, H.; Ghaedi, M.; Asfaram, A.; Monajemi, M.; Bazrafshan,
ultrafiltration membrane by blending with copolymer encompassing novel A. A. Synthesis of nanocomposites of iron oxide/gold (Fe3O4/Au) loaded
derivative of poly(styrene-co-maleic anhydride) for heavy metal removal on activated carbon and their application in water treatment by using
from wastewater. Chem. Eng. J. 2018, 353, 425–435. sonochemistry: Optimization study. Ultrason. Sonochem. 2018, 41, 279–287.
[55] Nataraj, S. K.; Hosamani, K. M.; Aminabhavi, T. M. Potential application [79] Kumar, K. Y.; Muralidhara, H. B.; Nayaka, Y. A.; Balasubramanyam, J.;
of an electrodialysis pilot plant containing ion-exchange membranes in Hanumanthappa, H. Hierarchically assembled mesoporous ZnO nanorods
chromium removal. Desalination 2007, 217, 181–190. for the removal of lead and cadmium by using differential pulse anodic
[56] Duan, J. C.; Lu, Q.; Chen, R. W.; Duan, Y. Q.; Wang, L. F.; Gao, L.; Pan, stripping voltammetric method. Powder Technol. 2013, 239, 208–216.
S. Y. Synthesis of a novel flocculant on the basis of crosslinked Konjac [80] Zha, R. H.; Nadimicherla, R.; Guo, X. Cadmium removal in waste water by
glucomannan-graft-polyacrylamide-co-sodium xanthate and its application nanostructured TiO2 particles. J. Mater. Chem. A 2014, 2, 13932–13941.
in removal of Cu2+ ion. Carbohydr. Polym. 2010, 80, 436–441. [81] Lung, I.; Stan, M.; Opris, O.; Soran, M.-L.; Senila, M.; Stefan, M. Removal
[57] El Samrani, A. G.; Lartiges, B. S.; Villiéras, F. Chemical coagulation of of lead(II), cadmium(II), and arsenic(III) from aqueous solution using

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


16 Nano Res.

magnetite nanoparticles prepared by green synthesis with box–behnken [102] Behnoudnia, F.; Dehghani, H. Anion effect on the control of morphology
design. Anal. Lett. 2018, 51, 2519–2531. for NiC2O4·2H2O nanostructures as precursors for synthesis of Ni(OH)2
[82] Chen, K.; He, J. Y.; Li, Y. L.; Cai, X. G.; Zhang, K. S.; Liu, T.; Hu, Y.; Lin, and NiO nanostructures and their application for removing heavy metal
D. Y.; Kong, L. T.; Liu, J. H. Removal of cadmium and lead ions from water ions of cadmium(II) and lead(II). Dalton Trans. 2014, 43, 3471–3478.
by sulfonated magnetic nanoparticle adsorbents. J. Colloid Interface Sci. [103] Gupta, V. K.; Nayak, A. Cadmium removal and recovery from aqueous
2017, 494, 307–316. solutions by novel adsorbents prepared from orange peel and Fe2O3
[83] Wang, D.; Guan, K. W.; Bai, Z. P.; Liu, F. Q. Facile preparation of acid- nanoparticles. Chem. Eng. J. 2012, 180, 81–90.
resistant magnetite particles for removal of Sb(Ⅲ) from strong acidic [104] Gusain, D.; Singh, P. K.; Sharma, Y. C. Kinetic and equilibrium modelling
solution. Sci. Technol. Adv. Mater. 2016, 17, 80–88. of adsorption of cadmium on nano crystalline zirconia using response
[84] Beheshtkhoo, N.; Kouhbanani, M. A. J.; Savardashtaki, A.; Amani, A. M.; surface methodology. Environ. Nanotechnol. Monit. Manag. 2016, 6,
Taghizadeh, S. Green synthesis of iron oxide nanoparticles by aqueous 99–107.
leaf extract of Daphne mezereum as a novel dye removing material. Appl. [105] Bhanjana, G.; Dilbaghi, N.; Singhal, N. K.; Kim, K.-H.; Kumar, S. Copper
Phys. A 2018, 124, 363. oxide nanoblades as novel adsorbent material for cadmium removal.
[85] Parveen, S.; Wani, A. H.; Shah, M. A.; Devi, H. S.; Bhat, M. Y.; Koka, Ceram. Int. 2017, 43, 6075–6081.
J. A. Preparation, characterization and antifungal activity of iron oxide [106] Mahdavi, S. Nano-TiO2 modified with natural and chemical compounds
nanoparticles. Microb. Pathog. 2018, 115, 287–292. as efficient adsorbents for the removal of Cd+2, Cu+2, and Ni+2 from water.
[86] Heiligtag, F. J.; Niederberger, M. The fascinating world of nanoparticle Clean Technol. Environ. Policy 2016, 18, 81–94.
research. Mater. Today 2013, 16, 262–271. [107] Li, X.; Zhao, K.; You, C. Y.; Linghu, W.; Ye, F.; Yu, M.; Alsaedi, A.; Hayat,
[87] Seo, K.; Sinha, K.; Novitskaya, E.; Graeve, O. A. Polyvinylpyrrolidone (PVP) T.; Pan, H.; Luo, J. et al. Nanocomposites of polyaniline functionalized
effects on iron oxide nanoparticle formation. Mater. Lett. 2018, 215, 203–206. graphene oxide: Synthesis and application as a novel platform for removal
[88] Gholami, L.; Kazemi Oskuee, R.; Tafaghodi, M.; Ramezani Farkhani, A.; of Cd(II), Eu(III), Th(IV) and U(VI) in water. J. Radioanal. Nucl. Chem.
Darroudi, M. Green facile synthesis of low-toxic superparamagnetic iron 2018, 315, 509–522.
oxide nanoparticles (SPIONs) and their cytotoxicity effects toward Neuro2A [108] Moosavian, M. A.; Moazezi, N. Removal of cadmium and zinc ions from
and HUVEC cell lines. Ceram. Int. 2018, 44, 9263–9268. industrial wastewater using nanocomposites of PANI/ZnO and PANI/
[89] Plachtová, P.; Medříková, Z.; Zbořil, R.; Tuček, J.; Varma, R. S.; Maršálek, CoHCF: A comparative study. Desalin. Water Treat. 2016, 57, 20817–20836.
B. Iron and iron oxide nanoparticles synthesized with green tea extract: [109] Rai, P. K.; Kumar, V.; Lee, S.; Raza, N.; Kim, K.-H.; Ok, Y. S.; Tsang, D.
differences in ecotoxicological profile and ability to degrade malachite C. W. Nanoparticle-plant interaction: Implications in energy, environment,
green. ACS Sustainable Chem. Eng. 2018, 6, 8679–8687. and agriculture. Environ. Int. 2018, 119, 1–19.
[90] Ehrampoush, M. H.; Miria, M.; Salmani, M. H.; Mahvi, A. H. Cadmium [110] Cao, Y. The toxicity of nanoparticles to human endothelial cells. In
removal from aqueous solution by green synthesis iron oxide nanoparticles Cellular and Molecular Toxicology of Nanoparticles. Saquib, Q.; Faisal,
with tangerine peel extract. J. Environ. Health Sci. Eng. 2015, 13, 84. M.; Al-Khedhairy, A. A.; Alatar, A. A., Eds.; Springer International
[91] Ebrahim, S. E.; Sulaymon, A. H.; Saad Alhares, H. Competitive removal Publishing: Cham, 2018; pp 59–69.
of Cu2+, Cd2+, Zn2+, and Ni2+ ions onto iron oxide nanoparticles from [111] Girigoswami, K. Toxicity of metal oxide nanoparticles. In Cellular
wastewater. Desalin. Water Treat. 2016, 57, 20915–20929. and Molecular Toxicology of Nanoparticles. Saquib, Q.; Faisal, M.;
[92] Devi, V.; Selvaraj, M.; Selvam, P.; Kumar, A. A.; Sankar, S.; Dinakaran, K. Al-Khedhairy, A. A.; Alatar, A. A., Eds.; Springer International Publishing:
Preparation and characterization of CNSR functionalized Fe3O4 magnetic Cham, 2018; pp 99–122.
nanoparticles: An efficient adsorbent for the removal of cadmium ion from [112] Kovalishyn, V.; Abramenko, N.; Kopernyk, I.; Charochkina, L.; Metelytsia,
water. J. Environ. Chem. Eng. 2017, 5, 4539–4546. L.; Tetko, I. V.; Peijnenburg, W.; Kustov, L. Modelling the toxicity of a
[93] OuldM'hamed, M.; Khezami, L.; Alshammari, A. G.; Ould-Mame, S. M.; large set of metal and metal oxide nanoparticles using the OCHEM
Ghiloufi, I.; Lemine, O. M. Removal of cadmium(II) ions from aqueous platform. Food Chem. Toxicol. 2018, 112, 507–517.
solution using Ni (15 wt.%)-doped αFe2O3 nanocrystals: Equilibrium, [113] Zhang, Y.; Meng, T. T.; Guo, X.; Yang, R. X.; Si, X. H.; Zhou, J. T. Humic
thermodynamic, and kinetic studies. Water Sci. Technol. 2015, 72, 608–615. acid alleviates the ecotoxicity of graphene-family materials on the
[94] Wang, Y.; Tian, T.; Wang, L.; Hu, X. Solid-phase preconcentration of freshwater microalgae Scenedesmus obliquus. Chemosphere 2018, 197,
cadmium(II) using amino-functionalized magnetic-core silica-shell nano- 749–758.
particles, and its determination by hydride generation atomic fluorescence [114] Lau, W.-J.; Emadzadeh, D.; Shahrin, S.; Goh, P. S.; Ismail, A. F.
spectrometry. Microchim. Acta 2013, 180, 235–242. Ultrafiltration membranes incorporated with carbon-based nanomaterials
[95] Simeonidis, K.; Martinez-Boubeta, C.; Zamora-Perez, P.; Rivera-Gil, P.; for antifouling improvement and heavy metal removal. In Carbon-Based
Kaprara, E.; Kokkinos, E.; Mitrakas, M. Nanoparticles for heavy metal Polymer Nanocomposites for Environmental and Energy Applications.
removal from drinking water. In Environmental Nanotechnology. Dasgupta, Ismail, A. F.; Goh, P. S., Eds.; Elsevier: Netherlands, 2018; pp 217–232.
N.; Ranjan, S.; Lichtfouse, E., Eds.; Springer International Publishing: [115] Yang, T.; Hodson, M. E. Investigating the potential of synthetic humic-
Cham, 2018; pp 75–124. like acid to remove metal ions from contaminated water. Sci. Total Environ.
[96] Huang, R. Y.; He, L.; Zhang, T.; Li, D. Q.; Tang, P. G.; Feng, Y. J. Novel carbon 2018, 635, 1036–1046.
paper@magnesium silicate composite porous films: Design, fabrication, [116] Chen, Q. Q.; Yin, D. Q.; Zhu, S. J.; Hu, X. L. Adsorption of cadmium(II)
and adsorption behavior for heavy metal ions in aqueous solution. ACS on humic acid coated titanium dioxide. J. Colloid Interface Sci. 2012,
Appl. Mater. Interfaces 2018, 10, 22776–22785. 367, 241–248.
[97] Huang, R. Y.; Wu, M. J.; Zhang, T.; Li, D. Q.; Tang, P. G.; Feng, Y. J. [117] Crini, G.; Lichtfouse, E.; Wilson, L. D.; Morin-Crini, N. Adsorption-
Template-free synthesis of large-pore-size porous magnesium silicate oriented processes using conventional and non-conventional adsorbents
hierarchical nanostructures for high-efficiency removal of heavy metal for wastewater treatment. In Green Adsorbents for Pollutant Removal:
ions. ACS Sustainable Chem. Eng. 2017, 5, 2774–2780. Fundamentals and Design. Crini, G.; Lichtfouse, E., Eds.; Springer
[98] Cao, C. Y.; Wei, F.; Qu, J.; Song, W. G. Programmed synthesis of magnetic International Publishing: Cham, 2018; pp 23–71.
magnesium silicate nanotubes with high adsorption capacities for lead and [118] Ji, Y. J.; Yang, M. Y.; Lin, H. P.; Hou, T. J.; Wang, L.; Li, Y. Y.; Lee, S.-T.
cadmium ions. Chem.—Eur. J. 2013, 19, 1558–1562. Janus structures of transition metal dichalcogenides as the heterojunction
[99] Garadkar, K. M.; Kadam, A. N.; Park, J. Microwave-assisted sol-gel synthesis photocatalysts for water splitting. J. Phys. Chem. C 2018, 122, 3123–3129.
of metal oxide nanomaterials. In Handbook of Sol-Gel Science and Technology: [119] Alabi, A.; AlHajaj, A.; Cseri, L.; Szekely, G.; Budd, P.; Zou, L. D. Review
Processing, Characterization and Applications. Klein, L.; Aparicio, M.; Jitianu, of nanomaterials-assisted ion exchange membranes for electromembrane
A., Eds.; Springer International Publishing: Cham, 2018; pp 483–504. desalination. npj Clean Water 2018, 1, 10.
[100] Yang, Z.-F.; Li, L.-Y.; Hsieh, C.-T.; Juang, R.-S. Co-precipitation of [120] Deshmukh, M. A.; Shirsat, M. D.; Ramanaviciene, A.; Ramanavicius, A.
magnetic Fe3O4 nanoparticles onto carbon nanotubes for removal of copper Composites based on conducting polymers and carbon nanomaterials
ions from aqueous solution. J. Taiwan Inst. Chem. Eng. 2018, 82, 56–63. for heavy metal ion sensing (review). Crit. Rev. Anal. Chem. 2018, 48,
[101] Aseem, A.; Jeba, G. G.; Conato, M. T.; Rimer, J. D.; Harold, M. P. Oxidative 293–304.
coupling of methane over mixed metal oxide catalysts: Steady state [121] Hua, M.; Jiang, Y. N.; Wu, B.; Pan, B. C.; Zhao, X.; Zhang, Q. X.
multiplicity and catalyst durability. Chem. Eng. J. 2018, 331, 132–143. Fabrication of a new hydrous Zr(IV) oxide-based nanocomposite for

| www.editorialmanager.com/nare/default.asp
Nano Res. 17

enhanced Pb(II) and Cd(II) removal from waters. ACS Appl. Mater. Interfaces [141] Sadegh, H.; Ali, G. A. M.; Gupta, V. K.; Makhlouf, A. S. H.; Shahryari-
2013, 5, 12135–12142. ghoshekandi, R.; Nadagouda, M. N.; Sillanpää, M.; Megiel, E. The
[122] Yang, S. M.; Morozovska, A. N.; Kumar, R.; Eliseev, E. A.; Cao, Y.; role of nanomaterials as effective adsorbents and their applications in
Mazet, L.; Balke, N.; Jesse, S.; Vasudevan, R. K.; Dubourdieu, C. et al. V. wastewater treatment. J. Nanostruct. Chem. 2017, 7, 1–14.
Mixed electrochemical–ferroelectric states in nanoscale ferroelectrics. [142] Krajina, B. A.; Proctor, A. C.; Schoen, A. P.; Spakowitz, A. J.; Heilshorn,
Nat. Phys. 2017, 13, 812–818. S. C. Biotemplated synthesis of inorganic materials: An emerging paradigm
[123] Naeem, H.; Ajmal, M.; Muntha, S.; Ambreen, J.; Siddiq, M. Synthesis and for nanomaterial synthesis inspired by nature. Prog. Mater. Sci. 2018,
characterization of graphene oxide sheets integrated with gold nanoparticles 91, 1–23.
and their applications to adsorptive removal and catalytic reduction of [143] Qiu, L.; McCaffrey, R.; Zhang, W. Synthesis of metallic nanoparticles
water contaminants. RSC Adv. 2018, 8, 3599–3610. using closed-shell structures as templates. Chem.— Asian J. 2018, 13,
[124] Mauter, M. S.; Zucker, I.; Perreault, F.; Werber, J. R.; Kim, J.-H.; Elimelech, 362–372.
M. The role of nanotechnology in tackling global water challenges. Nat. [144] Das, S. K.; Shome, I.; Guha, A. K. Surface functionalization of Aspergillus
Sustain. 2018, 1, 166–175. versicolor mycelia: In situ fabrication of cadmium sulphide nanoparticles
[125] Kunduru, K. R.; Nazarkovsky, M.; Farah, S.; Pawar, R. P.; Basu, A.; and removal of cadmium ions from aqueous solution. RSC Adv. 2012, 2,
Domb, A. J. Nanotechnology for water purification: Applications of 3000–3007.
nanotechnology methods in wastewater treatment. In Water Purification. [145] Gopalakrishnan, I.; Sugaraj Samuel, R.; Sridharan, K. Nanomaterials-
Grumezescu, A. M., Ed.; Academic Press: London, UK, 2017; pp 33–74. based adsorbents for water and wastewater treatments. In Emerging Trends
[126] Amil Usmani, M.; Khan, I.; H. Bhat, A. H.; Pillai, R. S.; Ahmad, N.; of Nanotechnology in Environment and Sustainability: A Review-Based
Mohamad Haafiz, M. K.; Oves, M. Current trend in the application of Approach. Sridharan, K., Ed.; Springer International Publishing: Cham,
nanoparticles for waste water treatment and purification: A review. Curr. 2018; pp 89–98.
Org. Synth. 2017, 14, 206–226. [146] Lim, J. Y.; Mubarak, N. M.; Abdullah, E. C.; Nizamuddin, S.; Khalid, M.;
[127] Xue, W. J.; Huang, D. L.; Zeng, G. M.; Wan, J.; Zhang, C.; Xu, R.; Cheng, Inamuddin. Recent trends in the synthesis of graphene and graphene
M.; Deng, R. Nanoscale zero-valent iron coated with rhamnolipid as an oxide based nanomaterials for removal of heavy metals—A review. J.
effective stabilizer for immobilization of Cd and Pb in river sediments. J. Ind. Eng. Chem. 2018, 66, 29–44.
Hazard. Mater. 2018, 341, 381–389. [147] Sarkar, B.; Mandal, S.; Tsang, Y. F.; Kumar, P.; Kim, K.-H.; Ok, Y. S.
[128] Yang, F.; Zhang, S. S.; Sun, Y. Q.; Cheng, K.; Li, J. S.; Tsang, D. C. W. Designer carbon nanotubes for contaminant removal in water and
Fabrication and characterization of hydrophilic corn stalk biochar-supported wastewater: A critical review. Sci. Total Environ. 2018, 612, 561–581.
nanoscale zero-valent iron composites for efficient metal removal. [148] Bhanjana, G.; Dilbaghi, N.; Kim, K.-H.; Kumar, S. Carbon nanotubes
Bioresource Technol. 2018, 265, 490–497. as sorbent material for removal of cadmium. J. Mol. Liq. 2017, 242,
[129] Su, Y. M.; Adeleye, A. S.; Huang, Y. X.; Sun, X. Y.; Dai, C. M.; Zhou, X. F.; 966–970.
Zhang, Y. L.; Keller, A. A. Simultaneous removal of cadmium and nitrate [149] Diaz-Flores, P. E.; López-Urías, F.; Terrones, M.; Rangel-Mendez, J. R.
in aqueous media by nanoscale zerovalent iron (nZVI) and Au doped Simultaneous adsorption of Cd2+ and phenol on modified N-doped carbon
nZVI particles. Water Res. 2014, 63, 102–111. nanotubes: Experimental and DFT studies. J. Colloid Interface Sci. 2009,
[130] Li, Z. T.; Wang, L.; Meng, J.; Liu, X. M.; Xu, J. M.; Wang, F.; Brookes, P. 334, 124–131.
Zeolite-supported nanoscale zero-valent iron: New findings on simultaneous [150] Rocha, R. P.; Soares, O. S. G. P.; Gonçalves, A. G.; Órfão, J. J. M.;
adsorption of Cd(II), Pb(II), and As(III) in aqueous solution and soil. J. Pereira, M. F. R.; Figueiredo, J. L. Different methodologies for synthesis
Hazard. Mater. 2018, 344, 1–11. of nitrogen doped carbon nanotubes and their use in catalytic wet air
[131] Lin, J. J.; Su, B. L.; Sun, M. Q.; Chen, B.; Chen, Z. L. Biosynthesized oxidation. Appl. Catal. A: Gen. 2017, 548, 62–70.
iron oxide nanoparticles used for optimized removal of cadmium with [151] Pei, H. N.; Wang, J.; Yang, Q. F.; Yang, W. X.; Hu, N.; Suo, Y. R.; Zhang,
response surface methodology. Sci. Total Environ. 2018, 627, 314–321. D. H.; Li, Z. H.; Wang, J. L. Interfacial growth of nitrogen-doped carbon
[132] Boparai, H. K.; Joseph, M.; O’Carroll, D. M. Cadmium (Cd2+) removal with multi-functional groups on the MoS2 skeleton for efficient Pb(II)
by nano zerovalent iron: Surface analysis, effects of solution chemistry removal. Sci. Total Environ. 2018, 631–632, 912–920.
and surface complexation modeling. Environ. Sci. Pollut. Res. 2013, 20, [152] Perez-Aguilar, N. V.; Muñoz-Sandoval, E.; Diaz-Flores, P. E.; Rangel-
6210–6221. Mendez, J. R. Adsorption of cadmium and lead onto oxidized nitrogen-
[133] Fan, D. M.; Lan, Y.; Tratnyek, P. G.; Johnson, R. L.; Filip, J.; O’Carroll, D. doped multiwall carbon nanotubes in aqueous solution: Equilibrium and
M.; Nunez Garcia, A.; Agrawal, A. Sulfidation of iron-based materials: kinetics. J. Nanopart. Res. 2010, 12, 467–480.
A review of processes and implications for water treatment and remediation. [153] Tahermansouri, H.; Ahi Roghayeh, M.; Kiani, F. Kinetic, equilibrium
Environ. Sci. Technol. 2017, 51, 13070–13085. and isotherm studies of cadmium removal from aqueous solutions by
[134] Podyacheva, O. Y.; Cherepanova, S. V.; Romanenko, A. I.; Kibis, L. S.; oxidized multi-walled carbon nanotubes and the functionalized ones with
Svintsitskiy, D. A.; Boronin, A. I.; Stonkus, O. A.; Suboch, A. N.; Puzynin, thiosemicarbazide and their toxicity investigations: A comparison. J. Chin.
A. V.; Ismagilov, Z. R. Nitrogen doped carbon nanotubes and nanofibers: Chem. Soc. 2014, 61, 1188–1198.
Composition, structure, electrical conductivity and capacity properties. [154] AlSaadi, M. A.; Al Mamun, A.; Alam, M. Z.; Amosa, M. K.; Atieh, M. A.
Carbon 2017, 122, 475–483. Removal of cadmium from water by CNT–PAC composite: Effect of
[135] Su, Y. M.; Adeleye, A. S.; Huang, Y. X.; Zhou, X. F.; Keller, A. A.; Zhang, functionalization. Nano 2015, 11, 1650011.
Y. L. Direct synthesis of novel and reactive sulfide-modified nano iron [155] Veličković, Z. S.; Bajić, Z. J.; Ristić, M. D.; Djokić, V. R.; Marinković, A.
through nanoparticle seeding for improved cadmium-contaminated water D.; Uskoković P. S.; Vuruna, M. M. Modification of multi-wall carbon
treatment. Sci. Rep. 2016, 6, 24358. nanotubes for the removal of cadmium, lead and arsenic from wastewater.
[136] Hu, Z. H.; Wu, Z. T.; Han, C.; He, J.; Ni, Z. H.; Chen, W. Two-dimensional Dig. J. Nanomater. Bios. 2013, 8, 501–511.
transition metal dichalcogenides: Interface and defect engineering. Chem. [156] Song, X. Y.; Guo, H.; Tao, J. B.; Zhao, S. L.; Han, X.; Liu, H. L.
Soc. Rev. 2018, 47, 3100–3128. Encapsulation of single-walled carbon nanotubes with asymmetric pyrenyl-
[137] Roobakhsh, S.; Rostami, Z.; Azizian, S. Can MoS2 nanosheets be used as gemini surfactants. Chem. Eng. Sci. 2018, 187, 406–414.
adsorbent for water treatment? Sep. Purif. Technol. 2018, 200, 23–28. [157] Pashai Gatabi, M.; Milani Moghaddam, H.; Ghorbani, M. Efficient removal
[138] Khalili, S. S.; Dehghani, H.; Afrooz, M. New porphyrin-doped silica of cadmium using magnetic multiwalled carbon nanotube nanoadsorbents:
monolith: An effective adsorbent for heavy metal ions in aqueous solution. Equilibrium, kinetic, and thermodynamic study. J. Nanopart. Res. 2016,
J. Sol-Gel Sci. Technol. 2018, 85, 290–301. 18, 189.
[139] Yin, W. Y.; Dong, X. H.; Yu, J.; Pan, J.; Yao, Z. Y.; Gu, Z. J.; Zhao, Y. L. [158] Samaddar, P.; Son, Y.-S.; Tsang, D. C. W.; Kim, K.-H.; Kumar, S. Progress
MoS2-nanosheet-assisted coordination of metal ions with porphyrin for in graphene-based materials as superior media for sensing, sorption, and
rapid detection and removal of cadmium ions in aqueous media. ACS separation of gaseous pollutants. Coord. Chem. Rev. 2018, 368, 93–114.
Appl. Mater. Interfaces 2017, 9, 21362–21370. [159] Sherlala, A. I. A.; Raman, A. A. A.; Bello, M. M.; Asghar, A. A review
[140] Evans, J. S.; Guo, T. B.; Sun, Y. R.; Liu, W.; Peng, L.; Xu, Z.; Gao, C.; of the applications of organo-functionalized magnetic graphene oxide
He, S. L. Shape-controlled of ten-nanometer-thick graphite and worm-like nanocomposites for heavy metal adsorption. Chemosphere 2018, 193,
graphite by lithographic exfoliation. Carbon 2018, 135, 248–252. 1004–1017.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


18 Nano Res.

[160] Deng, J.-H.; Zhang, X.-R.; Zeng, G.-M.; Gong, J.-L.; Niu, Q.-Y.; Liang, J. fibers for biosorption of cadmium, nickel, and lead ions from aqueous
Simultaneous removal of Cd(II) and ionic dyes from aqueous solution solution. Clean Technol. Environ. Policy 2014, 16, 385–393.
using magnetic graphene oxide nanocomposite as an adsorbent. Chem. [180] Sharma, P. R.; Chattopadhyay, A.; Sharma, S. K.; Geng, L. H.; Amiralian,
Eng. J. 2013, 226, 189–200. N.; Martin, D.; Hsiao, B. S. Nanocellulose from spinifex as an effective
[161] Shaheen, S. M.; Niazi, N. K.; Hassan, N. E. E.; Bibi, I.; Wang, H. L.; adsorbent to remove cadmium(II) from water. ACS Sustainable Chem.
Tsang, D. C. W.; Ok, Y. S.; Bolan, N.; Rinklebe, J. Wood-based biochar Eng. 2018, 6, 3279–3290.
for the removal of potentially toxic elements in water and wastewater: A [181] Abraham, E.; Deepa, B.; Pothan, L. A.; Jacob, M.; Thomas, S.; Cvelbar, U.;
critical review. Int. Mater. Rev., in press, DOI: 10.1080/09506608.2018. Anandjiwala, R. Extraction of nanocellulose fibrils from lignocellulosic
1473096. fibres: A novel approach. Carbohydr. Polym. 2011, 86, 1468–1475.
[162] Xu, G.; Wang, L.; Xie, Y. J.; Tao, M. L.; Zhang, W. Q. Highly selective [182] Alizadeh, B.; Delnavaz, M.; Shakeri, A. Removal of Cd(II) and phenol
and efficient adsorption of Hg2+ by a recyclable aminophosphonic acid using novel cross-linked magnetic EDTA/chitosan/TiO2 nanocomposite.
functionalized polyacrylonitrile fiber. J. Hazard. Mater. 2018, 344, Carbohydr. Polym. 2018, 181, 675–683.
679–688. [183] Zhang, S.; Lü, T.; Qi, D. M.; Cao, Z. H.; Zhang, D.; Zhao, H. T. Synthesis of
[163] Banazadeh, A.; Mozaffari, S.; Osoli, B. Facile synthesis of cysteine quaternized chitosan-coated magnetic nanoparticles for oil-water separation.
functionalized magnetic graphene oxide nanosheets: Application in solid Mater. Lett. 2017, 191, 128–131.
phase extraction of cadmium from environmental sample. J. Environ. [184] Zhou, L. M.; Wang, Y. P.; Liu, Z. R.; Huang, Q. W. Characteristics of
Chem. Eng. 2015, 3, 2801–2808. equilibrium, kinetics studies for adsorption of Hg(II), Cu(II), and Ni(II)
[164] Ghorbani, M.; Shams, A.; Seyedin, O.; Afshar Lahoori, N. Magnetic ions by thiourea-modified magnetic chitosan microspheres. J. Hazard.
ethylene diamine-functionalized graphene oxide as novel sorbent for Mater. 2009, 161, 995–1002.
removal of lead and cadmium ions from wastewater samples. Environ. [185] Zhu, Y. H.; Hu, J.; Wang, J. L. Competitive adsorption of Pb(II), Cu(II)
Sci. Pollut. Res. 2018, 25, 5655–5667. and Zn(II) onto xanthate-modified magnetic chitosan. J. Hazard. Mater.
[165] Liu, J.; Du, H. Y.; Yuan, S. W.; He, W. X.; Liu, Z. H. Synthesis of 2012, 221–222, 155–161.
thiol-functionalized magnetic graphene as adsorbent for Cd(II) removal [186] Monier, M.; Ayad, D. M.; Wei, Y.; Sarhan, A. A. Adsorption of Cu(II),
from aqueous systems. J. Environ. Chem. Eng. 2015, 3, 617–621. Co(II), and Ni(II) ions by modified magnetic chitosan chelating resin. J.
[166] Liu, J.; Du, H. Y.; Yuan, S. W.; He, W. X.; Yan, P. J.; Liu, Z. H. Alkaline Hazard. Mater. 2010, 177, 962–970.
deoxygenated graphene oxide as adsorbent for cadmium ions removal [187] Repo, E.; Warchol, J. K.; Kurniawan, T. A.; Sillanpää, M. E. T. Adsorption
from aqueous solutions. Water Sci. Technol. 2015, 71, 1611–1619. of Co(II) and Ni(II) by EDTA-and/or DTPA-modified chitosan: Kinetic
[167] Poo, K.-M.; Son, E.-B.; Chang, J.-S.; Ren, X. H.; Choi, Y.-J.; Chae, K.-J. and equilibrium modeling. Chem. Eng. J. 2010, 161, 73–82.
Biochars derived from wasted marine macro-algae (Saccharina japonica [188] Zhao, F. P.; Repo, E.; Yin, D. L.; Sillanpää, M. E. T. Adsorption of Cd(II)
and Sargassum fusiforme) and their potential for heavy metal removal in and Pb(II) by a novel EGTA-modified chitosan material: Kinetics and
aqueous solution. J. Environ. Manag. 2018, 206, 364–372. isotherms. J. Colloid Interface Sci. 2013, 409, 174–182.
[168] Zhang, Y.; Cao, B.; Zhao, L. L.; Sun, L. L.; Gao, Y.; Li, J. J.; Yang, F. [189] Salah, T. A.; Mohammad, A. M.; Hassan, M. A.; El-Anadouli, B. E.
Biochar-supported reduced graphene oxide composite for adsorption
Development of nano-hydroxyapatite/chitosan composite for cadmium
and coadsorption of atrazine and lead ions. Appl. Surf. Sci. 2018, 427,
ions removal in wastewater treatment. J. Taiwan Inst. Chem. Eng. 2014,
147–155.
45, 1571–1577.
[169] Liu, T. Z.; Gao, B.; Fang, J.; Wang, B.; Cao, X. D. Biochar-supported carbon
[190] Chen, L. Y.; Wu, P. X.; Chen, M. Q.; Lai, X. L.; Ahmed, Z.; Zhu, N. W.;
nanotube and graphene oxide nanocomposites for Pb(II) and Cd(II)
Dang, Z.; Bi, Y. Z.; Liu, T. Y. Preparation and characterization of the
removal. RSC Adv. 2016, 6, 24314–24319.
eco-friendly chitosan/vermiculite biocomposite with excellent removal
[170] Jlassi, K.; Abidi, R.; Benna, M.; Chehimi, M. M.; Kasak, P.; Krupa, I.
capacity for cadmium and lead. Appl. Clay Sci. 2018, 159, 74–82.
Bentonite-decorated calix [4] arene: A new, promising hybrid material
[191] Hossein Beyki, M.; Ghasemi, M. H.; Jamali, A.; Shemirani, F. A novel
for heavy-metal removal. Appl. Clay Sci. 2018, 161, 15–22.
polylysine–resorcinol base γ-alumina nanotube hybrid material for effective
[171] Liu, Y. N.; Zhong, Z. M. Extraction of heavy metals, dichromate anions and
adsorption/preconcentration of cadmium from various matrices. J. Ind.
rare metals by new calixarene-chitosan polymers. J. Inorg. Organomet.
Eng. Chem. 2017, 46, 165–174.
Polym. Mater. 2018, 28, 962–967.
[192] Awual, M. R.; Khraisheh, M.; Alharthi, N. H.; Luqman, M.; Islam, A.;
[172] Liu, C.; Zhang, D. X.; Zhao, L. T.; Lu, X.; Zhang, P.; He, S. N.; Hu, G. W.;
Rezaul Karim, M.; Rahman, M. M.; Khaleque, M. A. Efficient detection
Tang, X. Q. Synthesis of a thiacalix[4]arenetetrasulfonate-functionalized
reduced graphene oxide adsorbent for the removal of lead(II) and and adsorption of cadmium(II) ions using innovative nano-composite
cadmium(II) from aqueous solutions. RSC Adv. 2016, 6, 113352–113365. materials. Chem. Eng. J. 2018, 343, 118–127.
[173] Vikrant, K.; Kumar, V.; Ok, Y. S.; Kim, K.-H.; Deep, A. Metal-organic [193] Cao, C.-Y.; Qu, J.; Wei, F.; Liu, H.; Song, W.-G. Superb adsorption capacity
framework (MOF)-based advanced sensing platforms for the detection of and mechanism of flowerlike magnesium oxide nanostructures for lead
hydrogen sulfide. TrAC Trend. Anal. Chem. 2018, 105, 263–281. and cadmium ions. ACS Appl. Mater. Interfaces 2012, 4, 4283–4287.
[174] Fan, M.; Li, T.; Hu, J.; Cao, R.; Wei, X.; Shi, X.; Ruan, W. Artificial neural [194] Keochaiyom, B.; Wan, J.; Zeng, G. M.; Huang, D. L.; Xue, W. J.; Hu, L.;
network modeling and genetic algorithm optimization for cadmium removal Huang, C.; Zhang, C.; Cheng, M. Synthesis and application of magnetic
from aqueous solutions by reduced graphene oxide-supported nanoscale chlorapatite nanoparticles for zinc(II), cadmium(II) and lead(II) removal
zero-valent iron (nZVI/rGO) composites. Materials (Basel) 2017, 10, E544. from water solutions. J. Colloid Interface Sci. 2017, 505, 824–835.
[175] Kong, Q. P.; Wei, C. H.; Preis, S.; Hu, Y.; Wang, F. Facile preparation of [195] Tabesh, S.; Davar, F.; Loghman-Estarki, M. R. Preparation of γ-Al2O3
nitrogen and sulfur co-doped graphene-based aerogel for simultaneous nanoparticles using modified sol-gel method and its use for the adsorption
removal of Cd2+ and organic dyes. Environ. Sci. Pollut. Res. 2018, 25, of lead and cadmium ions. J. Alloy. Compd. 2018, 730, 441–449.
21164–21175. [196] Zhang, Z. Z.; Li, M. Y.; Chen, W.; Zhu, S. Z.; Liu, N. N.; Zhu, L. Y.
[176] Dong, C. C.; Lu, J.; Qiu, B. C.; Shen, B.; Xing, M. Y.; Zhang, J. L. Immobilization of lead and cadmium from aqueous solution and
Developing stretchable and graphene-oxide-based hydrogel for the removal contaminated sediment using nano-hydroxyapatite. Environ. Pollut. 2010,
of organic pollutants and metal ions. Appl. Catal. B: Environ. 2018, 222, 158, 514–519.
146–156. [197] da Rocha, N. C.; de Campos, R. C.; Rossi, A. M.; Moreira, E. L.; do F.
[177] Wu, S. B.; Zhang, K. S.; Wang, X. L.; Jia, Y.; Sun, B.; Luo, T.; Meng, Barbosa, A.; Moure, G. T. Cadmium uptake by hydroxyapatite synthesized
F. L.; Jin, Z.; Lin, D. Y.; Shen, W. et al. Enhanced adsorption of cadmium in different conditions and submitted to thermal treatment. Environ. Sci.
ions by 3D sulfonated reduced graphene oxide. Chem. Eng. J. 2015, 262, Technol. 2002, 36, 1630–1635.
1292–1302. [198] Tran, H. N.; You, S.-J.; Chao, H.-P. Thermodynamic parameters of
[178] Anirudhan, T. S.; Shainy, F. Adsorption behaviour of 2-mercaptobenzamide cadmium adsorption onto orange peel calculated from various methods:
modified itaconic acid-grafted-magnetite nanocellulose composite for A comparison study. J. Environ. Chem. Eng. 2016, 4, 2671–2682.
cadmium(II) from aqueous solutions. J. Ind. Eng. Chem. 2015, 32, [199] Al-Khaldi, F. A.; Abu-Sharkh, B.; Abulkibash, A. M.; Atieh, M. A.
157–166. Cadmium removal by activated carbon, carbon nanotubes, carbon
[179] Kardam, A.; Raj, K. R.; Srivastava, S.; Srivastava, M. M. Nanocellulose nanofibers, and carbon fly ash: A comparative study. Desalin. Water

| www.editorialmanager.com/nare/default.asp
Nano Res. 19

Treat. 2015, 53, 1417–1429. [213] Lievens, C.; Yperman, J.; Vangronsveld, J.; Carleer, R. Study of the potential
[200] Wasewar, K. L.; Kumar, P.; Chand, S.; Padmini, B. N.; Teng, T. T. Adsorption valorisation of heavy metal contaminated biomass via phytoremediation
of cadmium ions from aqueous solution using granular activated carbon by fast pyrolysis: Part I. Influence of temperature, biomass species
and activated clay. CLEAN 2010, 38, 649–656. and solid heat carrier on the behaviour of heavy metals. Fuel 2008, 87,
[201] Mathialagan, T.; Viraraghavan, T. Adsorption of cadmium from aqueous 1894–1905.
solutions by perlite. J. Hazard. Mater. 2002, 94, 291–303. [214] Stals, M.; Thijssen, E.; Vangronsveld, J.; Carleer, R.; Schreurs, S.;
[202] Taamneh, Y.; Sharadqah, S. The removal of heavy metals from aqueous Yperman, J. Flash pyrolysis of heavy metal contaminated biomass
solution using natural Jordanian zeolite. Appl. Water Sci. 2017, 7, 2021–2028. from phytoremediation: Influence of temperature, entrained flow and
[203] Yang, T.; Li, Y.-K.; Chen, M.-L.; Wang, J.-H. Supported carbon dots wood/leaves blended pyrolysis on the behaviour of heavy metals. J. Anal.
decorated with metallothionein for selective cadmium adsorption and Appl. Pyrol. 2010, 87, 1–7.
removal. Chin. Chem. Lett. 2015, 26, 1496–1501. [215] Nassar, N. N. Kinetics, equilibrium and thermodynamic studies on the
[204] Sigma-Aldrich. Carbon nanotube, multi-walled [Online]. Merck KGaA: adsorptive removal of nickel, cadmium and cobalt from wastewater by
South Korea, 2018; https://www.sigmaaldrich.com/catalog/product/aldrich/ superparamagnetic iron oxide nanoadsorbents. Can. J. Chem. Eng. 2012,
659258?lang=ko&region=KR (accessed Oct 10, 2018). 90, 1231–1238.
[205] Sigma-Aldrich. Zeolite [Online]. Merck KGaA: South Korea, 2018; [216] Jafarinejad, S.; Faraji, M.; Norouz, Z.; Mokhtari-Aliabad, J. Application
https://www.sigmaaldrich.com/catalog/product/sigma/96096? of sulfur-modified magnetic nanoparticles for cadmium removal from
lang=ko&region=KR (accessed Oct 10, 2018). aqueous solutions. J. Water Environ. Nanotechnol. 2018, 3, 58–69.
[206] Kabir, E.; Kumar, V.; Kim, K.-H.; Yip, A. C. K.; Sohn, J. R. Environmental [217] Xu, Z. H.; Zhang, D. F.; Chen, W. F.; Li, Y. R.; Yuan, S. J. Nanoscale iron
impacts of nanomaterials. J. Environ. Manag. 2018, 225, 261–271. oxides loaded granular activated carbon (GAC-NSIO) for cadmium
[207] Mubarak, N. M.; Sahu, J. N.; Abdullah, E. C.; Jayakumar, N. S. Removal removal. Desal. Water Treat. 2016, 57, 3559–3571.
of heavy metals from wastewater using carbon nanotubes. Sep. Purif. [218] Zuo, Y.; Chen, G. Q.; Zeng, G. M.; Li, Z. W.; Yan, M.; Chen, A. W.; Guo,
Rev. 2014, 43, 311–338. Z.; Huang, Z. Z.; Tan, Q. Transport, fate, and stimulating impact of silver
[208] Das, R.; Ali, M. E.; Hamid, S. B. A.; Ramakrishna, S.; Chowdhury, Z. Z. nanoparticles on the removal of Cd(II) by Phanerochaete chrysosporium
Carbon nanotube membranes for water purification: A bright future in in aqueous solutions. J. Hazard. Mater. 2015, 285, 236–244.
water desalination. Desalination 2014, 336, 97–109. [219] Vuković, G. D.; Marinković, A. D.; Čolić, M.; Ristić, M. Đ.; Aleksić, R.;
[209] Awual, M. R. Novel nanocomposite materials for efficient and selective Perić-Grujić, A. A.; Uskoković, P. S. Removal of cadmium from aqueous
mercury ions capturing from wastewater. Chem. Eng. J. 2017, 307, solutions by oxidized and ethylenediamine-functionalized multi-walled
456–465. carbon nanotubes. Chem. Eng. J. 2010, 157, 238–248.
[210] Naseeruteen, F.; Hamid, N. S. A.; Suah, F. B. M.; Ngah, W. S. W.; [220] Tshwenya, L.; Arotiba, O. A. Ethylenediamine functionalized carbon
Mehamod, F. S. Adsorption of malachite green from aqueous solution by nanoparticles: Synthesis, characterization, and evaluation for cadmium
using novel chitosan ionic liquid beads. Int. J. Biol. Macromol. 2018, removal from water. RSC Adv. 2017, 7, 34226–34235.
107, 1270–1277. [221] National Nanotechnology Initiative. Benefits and Applications of nano-
[211] Han, Z. Y.; Guo, Z. H.; Zhang, Y.; Xiao, X. Y.; Xu, Z.; Sun, Y. Adsorption- materials [Online]. United States: National Nanotechnology Initiative,
pyrolysis technology for recovering heavy metals in solution using 2018; https://www.nano.gov/you/nanotechnology-benefits (accessed Oct 8,
contaminated biomass phytoremediation. Resour. Conservat. Recycl. 2018, 2019).
129, 20–26. [222] Valli, F.; Tijoriwala, K.; Mahapatra, A. Nanotechnology for water purification.
[212] Mahar, A.; Wang, P.; Ali, A.; Awasthi, M. K.; Lahori, A. H.; Wang, Q.; Int. J. Nucl. Desalin. 2010, 4, 49–57.
Li, R. H.; Zhang, Z. Q. Challenges and opportunities in the phytoremediation [223] Baruah, S.; Dutta, J. Nanotechnology applications in pollution sensing
of heavy metals contaminated soils: A review. Ecotoxicol. Environ. Saf. and degradation in agriculture: A review. Environ. Chem. Lett. 2009, 7,
2016, 126, 111–121. 191–204.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

You might also like