You are on page 1of 19

Applied Clay Science 198 (2020) 105838

Contents lists available at ScienceDirect

Applied Clay Science


journal homepage: www.elsevier.com/locate/clay

Review article

Biopolymer-clay nanocomposites as novel and ecofriendly adsorbents for T


environmental remediation
María del Mar Ortaa, Julia Martínb, , Juan Luis Santosb, Irene Apariciob,

Santiago Medina-Carrascoc, Esteban Alonsob


a
Departamento de Química Analítica, Facultad de Farmacia, Universidad de Sevilla, C/ Profesor García González 2, 41012 Sevilla, Spain
b
Departamento de Química Analítica, Escuela Politécnica Superior, Universidad de Sevilla, C/ Virgen de África 7, E-41011 Sevilla, Spain
c
Laboratorio de Rayos-X (CITIUS), Universidad de Sevilla, Avenida Reina Mercedes 4B, 41012 Sevilla, Spain

ARTICLE INFO ABSTRACT

Keywords: Water pollution is usually a result of human activities and it has become an issue of concern because of the toxic
Biopolymers and carcinogenic effects associated with some pollutants which may affect all living organisms. Among all the
Clays water treatment technologies adsorption is considered the most effective technique due to its affordability,
Bionanocomposites universal availability and ease of operation. Recent advances in nanotechnology have allowed the development
Adsorption
of bionanocomposites based on clay minerals and biopolymers which have emerged as promising materials for
Hazardous pollutants
Remediation
the removal of pollutants from contaminated water. Their improved properties regarding the biocompatibility
and biodegradability of biopolymers, the high surface area, increased number of active sites, and the excellent
sorption capacity of clays make the use of bionanocomposites very attractive. This review focuses on the use of
biopolymer-clays nanocomposites for environmental remediation. First, the physicochemical and structural
properties of clays and biopolymers as well as the main characterization and preparation techniques are dis-
cussed. Subsequently, the main applications of bionanocomposites as adsorbent materials for the removal of
emerging organic and inorganic pollutants from water samples are described.

1. Introduction potential adsorption materials due to their excellent properties such as


high surface area and cation exchange capacity (CEC), swelling, mi-
The introduction of toxic substances of synthetic or natural origin croporosity, layered structure, nano-size, as well as their lower cost
(pesticides, hydrocarbons, heavy metals, dyes, detergents, and others) compared to conventional adsorbents like activated carbon. Ad-
into the environment is an alarming sign of the degradation of the ditionally, clays are suitable for large-scale applications due to their
quality of the environment. The removal of these toxic chemicals from excellent stability and safety (Awad et al., 2019).
wastewater before the discharge into aquatic ecosystems has become However, natural inorganic clays are not effective for the removal of
imperative and, in this respect, the development of efficient technolo- hydrophobic organic micropollutants from water due to the hydrophilic
gies for the remediation of water contamination has become a main nature of clay surfaces. However, the nature of the interlayer space can
field of study in environmental research. Major technologies, with be modified from hydrophilic to hydrophobic by the intercalation of
varying degrees of efficiency, include advanced oxidation, ion ex- alkyl long chains or functional groups to the interlayer space, which
change, membrane filtration, ozonation or adsorption (Tarpani and provides clays with a wide array of new and interesting adsorbent
Azapagic, 2018). Adsorption is considered the most effective technique properties (Buruga et al., 2019; Kotal and Bhowmick, 2015). The bio-
for the treatment of contaminated water because of the ease of opera- compatibility of clays and biodegradability properties of biopolymers
tion, low investment costs and the improved accessibility to different makes their combination a successful approach for the development of
adsorbents (Sophia and Lima, 2018; Tran et al., 2015). Much research is bionanocomposites that can be used for environmental remediation
being done on the exploration of new low-cost, more stable, safer, (Ruiz-Hitzky et al., 2008a, 2008b; Darder et al., 2007; Ruiz-Hitzky and
ecofriendly and efficient adsorbents that can be used for pollutant re- Van Meerbeek, 2006; Ruiz-Hitzky, 2004). Among the biopolymers,
moval from water in batch mode processes (Sophia and Lima, 2018; polysaccharides and polipeptides are considered the best candidates
Bhat et al., 2018). Among these adsorbents, clays are considered because of their inexpensive and non-toxic properties. Since the first


Corresponding author.
E-mail address: jbueno@us.es (J. Martín).

https://doi.org/10.1016/j.clay.2020.105838
Received 4 March 2020; Received in revised form 10 August 2020; Accepted 6 September 2020
Available online 14 September 2020
0169-1317/ © 2020 Elsevier B.V. All rights reserved.
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

Fig. 1. Structure of 2:1 phyllosilicates (modified from Awad et al., 2019).

investigations on gelatin-montmorillonite (Talibudeen, 1950), collagen- 2. Clay minerals


sepiolite (Herrera et al., 1995), or starch-montmorillonite (Besun et al.,
1997) hybrid nanocomposites, the research in this field of study has The structure of most clay minerals is composed of well-defined
advanced rapidly. A wide variety of bionanocomposites based on the layers of linked silicate tetrahedra and octahedra shaping the clay as-
intercalation of high molecular mass biopolymers such as chitosan semblages. Tetrahedral layers consist of oriented, linked silica–oxygen
(CTS), cellulose (CELL), starch (ST), alginate (ALG), sacran, zein, ge- tetrahedra sharing three oxygens with neighboring tetrahedra and the
latin (Gel) or polylactic acid (PLA) into clay minerals such as smectites, fourth apical oxygen is shared with an adjacent octahedral sheet via a
sepiolite, palygorskite, micas, kaolinites, and synthetic layered double covalent bond. The octahedral layers are composed of aluminum, ferric
hydroxides (LDH) have been developed over the last 20 years or magnesium ions in central position surrounded by a closely packed
(Alcântara et al., 2016a; Alcântara et al., 2014a; Avérous, 2013; arrangement of six oxygen atoms or hydroxyl groups. Aluminum ions
Avérous and Pollet, 2012; Alcântara et al., 2012; Ray, 2012; Chivrac only hold two-thirds of central positions of the octahedral sheets and
et al., 2009; Chivrac et al., 2008; Darder et al., 2007; Darder et al., the remaining one-third centers remain empty (Murray, 1991; Charles
2006; Mayer, 2006; Ruiz-Hitzky et al., 2005; Siqueira et al., 2010; and Lin, 1937). This property makes the octahedral layers prone to
Dujardin and Mann, 2002). isomorphic substitutions by metals like magnesium, which can fill all
The control of biopolymer-clay interactions will help in the devel- the metal center positions in the sheet. These layers arrange themselves
opment of smart nanomaterials for environmental applications. To date into stacks with a regular van der Waals gap between them called the
this new type of biohybrid materials has been used for the removal of interlayer or the gallery spacing. The isomorphic substitutions in the
heavy metals, reactive dyes, pesticides and even emerging pollutants octahedral sheets produce negative charges that are counterbalanced by
such as bisphenol A (BPA) in aqueous media (Mukhopadhyay et al., alkali or alkaline earth cations in the interlayer (Lee and Tiwari, 2012;
2020; Buruga et al., 2019; Alcântara and Darder, 2018; Shabtai and Ruiz-Hitzky et al., 2004).
Mishael, 2018). Additionally, some of the individual limitations of clays It is the arrangement of the sheets and/or layers that differentiates
and polymers (low specificity, sensitivity to pH and particle size as well the clays, providing them with their defining characteristics and
as low wettability) are improved when both materials are assembled properties. Based on their mineralogical composition, there are ap-
together to form bionanocomposites. This review describes the physi- proximately 30 types of nanoclays with different properties that allow
cochemical and structural properties of clay-biopolymer bionano- different applications. The main subdivisions of the layered silicates are
composites and it highlights the advances for their preparation, the done according to the type of combinations of the octahedral and tet-
methods of modification and characterization techniques as well as the rahedral sheets and the way they stack upon each other, whether the
main applications in the field of environmental remediation. octahedral sheet contains two or three cations per half unit cell, and
lastly to the number of isomorphous replacements and the cations in-
volved (Charles and Lin, 1937). There are seven groups of natural clay

2
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

minerals: (i) kaolinites, where one octahedral sheet is linked with one be improved by ion exchange with cationic surfactant molecules
tetrahedra (1:1 ratio) (e.g., kaolinite, halloysite, serpentine); (ii) non- (usually [(CH3)3NR]+ or [(CH3)2NR2]+ where R is a hydrocarbon
expanding clays, where one octahedral sheet intercalates between two substituent) intercalated between the interlayer space which results in
tetrahedral sheets (2:1 ratio) (e.g., mica and illite); (iii) limited ex- the surface shifting from highly hydrophilic to increasingly hydro-
panding clays (2:1 ratio) (e.g., vermiculite); (iv) strongly expanding phobic (Ogawa and Kuroda, 1997). In addition, in swelling clays the
clays (2:1 ratio) (e.g., montmorillonite [Mt)]); (v) uncharged group (2:1 intercalation of a cationic surfactant leads to increased basal spacing
ratio) (e.g., pyrophyllite and talc); (vi) 2:1:1 group with a brucite oc- and exposure of new sorption sites (Park et al., 2011). In recent years,
tahedral layer between a 2:1 mineral (e.g., chlorites); and (vii) fibrous the adsorption capacity of swelling organomodified clays with primary
silicates (e.g., palygorskite and sepiolite) (Mukhopadhyay et al., 2020). alkylamine and quaternary ammonium cations for organic pollutants
The binding forces that hold the sheets forming an ‘interlayer’ are weak removal has been extensively studied (Park et al., 2011; Jlassi et al.,
van der Waals bonds. The pH and isomorphous substitutions in the 2017; Awad et al., 2019, Martín et al., 2019a). Recently, Jlassi et al.
tetrahedral and octahedral sheets generate negative and positive sur- (2017) presented a review about preparation, modification and ad-
face charges. The clay minerals included in the classification provided sorption properties of organoclays. Awad et al. (2019) reviewed the
above are defined as nanoscale materials having at least one dimension information available about the technical viability of raw and modified
at the nanometric scale (Ruiz-Hitzky et al., 2004; Murray, 1991). clay sorption processes. They explored the adsorption of phenolic
Strongly expanding (2:1 ratio) clays such as Mt., hectorite or ben- compounds, aromatic compounds, pesticides and herbicides, and other
tonite, are the most commonly used layered silicates. As the forces that organic contaminants and concluded that adsorption of clays is largely
hold the stacks together are relatively weak, the intercalation of small affected by the clay properties, the kind of cationic surfactant mole-
molecules between the layers is easy. The structure of 2:1 phyllosili- cules, size and shape of the target pollutants, as well as the adsorption
cates is shown in Fig. 1. The adsorption capacity of these natural clays conditions (pH, adsorbent dosage, surfactant loading, organic con-
has been extensively assessed for the removal of heavy metals and polar centration, and others). The role-played for all these variables need to
organic pollutants from aqueous solutions (Park et al., 2011). One of be considered when assessing the sorption capacity of clays for the
the advantages that make clays superior to other conventional ad- removal of specific contaminants.
sorbents like activated carbon is their low cost and environmental
disposal of their byproducts as compared to other biological and che-
mical processes. 3. Biopolymer-clay nanocomposites
However, as stated in the Introduction, the hydrophilic nature of the
surface of naturally occurring clays makes them ineffective for the New materials with enhanced structural and functional properties
adsorption of anionic, hydrophobic or non-polar organic pollutants have appeared as the result of the development of innovative synthesis
(Park et al., 2011). For instance, the removal efficiency of phenolic strategies. A bionanocomposite is a type of hybrid material composed
compounds with clays is lower than the efficiency of state-of-the-art by a mixture of a natural polymer or biopolymer and an inorganic solid
technologies; however, it has been reported that adsorption of 2,4 di- with at least one dimension at the nanometric scale (Ruiz-Hitzky et al.,
chlorophenol and bisphenol with clays is 10% and 40% higher than 2008a, 2008b). These materials have been successfully incorporated
with activated carbon and graphene/zeolites, respectively (Awad et al., into different fields such as advanced biomedical materials (tissue en-
2019). gineering, artificial bones or gene therapy), in controlled drug and
In natural clays adsorption, is surface-related and the surface pesticides delivery, food processing, water purification, oxygen barrier
characteristics can be modified by organic cation exchange reactions films or food package (Alcântara and Darder, 2018; Ucankus et al.,
which give rise to organoclays (Lee and Tiwari, 2012; Ruiz-Hitzky et al., 2018; Ahmed and Varshney, 2010; Rhim, 2007; Ruiz-Hitzky et al.,
2004; Vaia et al., 1994; Theng, 1974; Murray, 1924a, 1924b). The 2005; Siqueira et al., 2010).
negatively charged aluminosilicate layer of clays is compensated by Among inorganic solids, special attention is given to layered mate-
inorganic exchange cations (e.g. Na+ and Ca+2), which are strongly rials with capacity to intercalate biopolymers and form hybrids with
hydrated in the presence of water. The properties of clay minerals can functional properties (Alcântara and Darder, 2018; Darder et al., 2007;
Ruiz-Hitzky et al., 2005). Nanoclays are the main inorganic components

Fig. 2. Classification of clay materials.

3
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

Fig. 3. Chemical structures of the biopolymers most commonly used in biocomposite preparation.

used owing to their affordability, availability, high CEC, adsorption the hydroxyl and amino groups present in CTS can improve its physical
capacity and high surface area. The intercalation of polymers in the and chemical properties. Some CTS properties include hydrophilicity,
interlayer space greatly enhances the adsorption characteristics of biocompatibility, biodegradability, chelating ability, antimicrobial ac-
polymer nanocomposites. Note that some of the clays shown in Fig. 2 tivity and film forming capacity, making this polysaccharide especially
have not been studied, since not all of the clays are useful for polymeric suited for application in the food and cosmetics industry, water treat-
composite preparation. The 2:1-type layer silicates (Mt, sepiolite, ver- ment, and in the biomedical field for tissue engineering and drug de-
miculite or mica) and MgeAl LDH have been the types of clays more livery (Sethy and Sahoo, 2019).
extensively studied (Mukhopadhyay et al., 2020). As for biopolymers, CTS is characterized by its acetylation degree and by its molecular
the most commonly used for bionanocomposite preparation are related weight, which afecct its viscosity and solubility. In acid conditions,
to polysaccharides such as CTS, CELL, ST and ALG (Abbasian et al., when the amino groups are protonated, CTS becomes a water soluble
2017; Zafar et al., 2016; Avérous and Pollet, 2014; Chivrac et al., 2009; polycation. This property has been extensively used to elaborate CTS-
Siqueira et al., 2010) proteins such as Zein or Gel (Alcântara et al., Mt hybrid materials (Chivrac et al., 2009). The binding forces between
2016a; Alcântara et al., 2012; Darder et al., 2006; Ruiz et al., 2006) and the protonated amino groups of the CTS solution and the negative
PLA (Avérous, 2013; Ray, 2012).. The structure of the main biopoly- charge on the silicate layers are mainly electrostatic interactions, but
mers used is shown in Fig. 3. From an environmental standpoint, hydrogen bonds may also play a role (Alcântara and Darder, 2018).
polysaccharides are the preferred biopolymers because of their inherent A great number of CTS-clay bionanocomposites have been elabo-
biodegradability, production from renewable sources, availability, ab- rated, studied and reported in the literature (Alba et al., 2019; Chen
sence of toxicity, and relatively low cost (Azhar et al., 2014; Arora et al., 2018; Padilla-Ortega et al., 2016; Liu et al., 2015; Deng et al.,
et al., 2018). Table 1 shows a summary of the characterization of the 2012; Darder et al., 2005a; Wang et al., 2005; Darder et al., 2003).
main biopolymers and clays used as adsorbent materials. Experiments carried out by the Ruiz-Hitzky group (Darder et al., 2005a;
Darder et al., 2003) revealed the possibility to control the intercalation
3.1. Polysaccharides-based bionanocomposites of the biopolymer as a mono- or a bilayer. They intercalated both,
monolayer and bilayer of CTS (with varying amounts) between the
3.1.1. Chitosan-based bionanocomposites individual silicate layers of Mt. and found that intercalation resulted in
CTS [poly-b(1,4)-2-amino-2-deoxy-D-glucose], the deacetylated an increases in the basal spacing (d001) from 1.20 nm to 1.45 nm or
product of chitin, is a cationic biopolymer and the second most abun- 2.04 nm, which corresponds the thickness of one or two polysaccharide
dant polysaccharide in nature after cellulose. Chemical modifications of chains. The first CTS layer is adsorbed through a cationic exchange

4
Table 1
Main biopolymers and clays used in the synthesis of bionanocomposites.
Clay/Organoclay Polymer % w:w (Clay/ Preparation method Characterization Reference
biopolymer)
M.d.M. Orta, et al.

Montmorillonite Chitosan 96:4 Solution blending method XRD; FTIR; TGA; Zeta potential Zhu et al., 2017
Chitosan/Ag 1:1 Solution blending method FTIR; XRD; SEM Azzam et al., 2016
Chitosan 1:1 Solution blending method XRD; FTIR; SEM Zhu et al., 2015
Chitosan 1:1 Solution blending method XRD; FTIR; TEM; TGA Azhar et al., 2014
Chitosan 1.23:1 Solution blending method XRD; FTIR; TEM Celis et al., 2012
Chitosan – Solution blending method XRD; FTIR; TEM; SEM Pandey and Mishra, 2011
Chitosan – Solution blending method XRD; FTIR; Zeta potencial Bleiman and Mishael, 2010
Chitosan 1:1,33 Solution blending method XRD; FTIR; TEM; SEM; TGA; AFM Yao et al., 2010
Chitosan 1:1.6 Solution blending method XRD; FTIR Monvisade and Siriphannon,
2009
Chitosan/Poly 1:7.2 In situ intercalated polymerization XRD; IR Wang et al., 2008
acrylic acid
Alginate 4:1 Solution blending method XRD; FTIR;SEM; TGA; BET Etcheverry et al., 2017
Alginate 1:1 Solution blending method XRD; FTIR; TEM; TGA Azhar et al., 2014
Alginate 4:1 Ionic gelation method FTIR Barreca et al., 2014
Alginate 3:7 Solution blending method XRD; FTIR; SEM; BET Shawky, 2011
Alginate-Aniline 1:1 Chemical oxidative polymerization XRD; FTIR; TEM; TGA Azhar et al., 2014
Starch 1:1 Solution blending method XRD; FTIR; TEM; TGA Azhar et al., 2014
Starch-Aniline 1:1 Chemical oxidative polymerization XRD; FTIR; TEM; TGA Azhar et al., 2014
Guar Gum 1:1 Solution blending method XRD; FTIR; TEM; SEM/EDX; TGA; DTG Ahmad and Mirza, 2018a
β-cyclodextrin – Solution blending method XRD; SEM; TGA;BET Shabtai and Mishael, 2018
Cellulose 2:1 Solution blending method FTIR; EDX; XRD; BET Kumar et al., 2012
Gelatin 1:1 Solution blending method XRD; FTIR; TEM; TGA Azhar et al., 2014
Gelatin-Aniline 1:1 Chemical oxidative polymerization XRD; FTIR; TEM; TGA Azhar et al., 2014

5
Kappa-carrageenan/ poly(vinyl alcohol) – – XRD; FTIR; TEM; SEM Hosseinzadeh et al., 2015
Polylactic acid – Melt-blending method WAXD; SAXS; TEM; DSC Neppalli et al., 2014
Polylactic-co-glycolic acid 3:1 Solution blending method XRD; TGA; DSC Seema, 2013
Xylan 1,25:1 Solution blending method XRD; FTIR; RC Ünlü et al., 2012
Phosphatidydylcholine – Solution blending method XRD; FTIR; TGA; NMR Wicklein et al., 2010
Bentonite Chitosan 1:1 Solution blending method FTIR; SEM; XRD; TGA Liu et al., 2015
Chitosan 20:1 Solution blending method TGA; BET Futalan et al., 2011
Chitosan 1:1 Solution blending method FTIR; SEM; BET; BJH Ngah et al., 2010a
Chitosan 1.5:1 Solution blending method XRD; FTIR; SEM Ngah et al., 2010b
Starch 1:1 Solution blending method XRD; Zeta potencial Koriche et al., 2014
Guar gum 1:1 Solution blending method XRD; FT-IR; TEM; SEM; EDX; TGA; BET Ahmad and Mirza, 2018b
Na-mica-n Chitosan 1:5 Solution blending method XRD; FTIR; TGA; MAS-NMR; Zeta- Alba et al., 2019
potential
Mica Alginate/Au 2:1 Solution blending method XRD; FTIR; TEM; SEM/EDX; TGA Ahmad and Mirza, 2017
Cellulose; Fructose 1:0,15 Solution blending method XRD; FTIR; TGA; DTA Sharma and Komarneni, 2009
Vermiculite Chitosan – Solution blending method FTIR; SEM; BET; TG-DTG; Zeta Chen et al., 2018
potential; XPS
Chitosan 35:65 Solution blending method involving XRD; ATR-IR; TEM; FE-SEM Padilla-Ortega et al., 2016
ultrasound irradiation
Chitosan 1:1 Solution blending method ATR-IR; FTIR; SEM Alcântara et al., 2014a
Hexadecyltrimethylammonium- Vermiculite Chitosan 35:65 Solution blending method involving XRD;ATR-IR; TEM; FE-SEM Padilla-Ortega et al., 2016
ultrasound irradiation
Sepiolite Sacram 1:2.7 Solution blending method ATR-IR; FTIR; SEM Alcântara et al., 2014a
Kappa-carrageenan/poly(acrylamide) – In situ-polymerization FTIR; SEM; TEM; TGA Mahdavinia and Asgari, 2013
Phosphatidydylcholine – Solution blending method XRD; FTIR; TGA; NMR Wicklein et al., 2010
Palygorskite Starch 1:1 Solution blending method ATR-IR; FTIR; SEM Alcântara et al., 2014a
Alginate 1:1 Solution blending method ATR-IR; FTIR; SEM Alcântara et al., 2014a
(continued on next page)
Applied Clay Science 198 (2020) 105838
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

izations; RSS: Raman spectroscopy studies; SEM: Scanning electron microscopy; TEM: Transmission electron microscopy; TGA: Thermogravimetric analysis; XRD: X-ray diffraction; XPS: X-ray Photoelectron Spectroscopy.
AFM: Atomic force microscope; ATR-IR: Attenuated total reflection-Infrared; BET: Brunauer–Emmett–Teller method; BJH: Barrett, Joyner, and Halenda method; DSC: Differential scanning calorimetry; EDX: Energy-
dispersive X-ray spectroscopy; FE-SEM: Field Emission Scanning Electron Microscopes; FTIR: Fourier transform infrared; MAS-NMR: Solid State Nuclear Magnetic Resonance spectroscopy; RC: Rheological character-
procedure, while the second layer is adsorbed in the acetate salt form
(Darder et al., 2003). Additional intercalation of more CTS layers re-
Dinari and Dadkhah, 2020
sulted in exfoliated structures (Wang et al., 2005; Darder et al., 2003).

Auta and Hamed, 2014


Moreover, the modification in the number of CTS layers allowed the
Swain et al., 2018 shift from CEC to an anionic exchange capacity (AEC). When there is an
Deng et al., 2012

Zhu et al., 2010


exchange higher than CEC, the adsorption of CTS on Mt. improves the
Li et al., 2018

absorpion of anions because the –NH3+ groups of CTS not involved in


Reference

the interaction with the negatively clay surfaces can act as anionic
exchangers (Ruiz-Hitzky et al., 2005; Wang et al., 2005). Fourier-
transform infrared spectroscopy (FTIR) analyses have confirmed the
electrostatic interactions established between both components of the
nanocomposite. The bands related to the deformation vibration (dNH3)
XRD; FTIR; TGA; DSC; Z-potencial

of the protonated amino group in the CTS structure are shifted towards
XRD; FTIR; EDX; TEM; SEM

low frequency values. Solid-state 13C nuclear magnetic resonance


(NMR) studies confirmed the exchange of the starting acetate anions
XRD; FTIR; TEM; TGA
XRD; FTIR; SEM; TGA

after treatment of this nanocomposite in a sodium nitrate solution


Characterization

XRD; TEM; SEM

(Darder et al., 2005a).


Wang et al. (2006) introduced carboxymethyl groups into CTS to
FTIR; SEM

enhance the hydrogen-bonding reaction between the matrix and –OH


group located at the edges of Mt. The Mt. was exfoliated into a large
excess of water and then mixed with a solution of carboxymethylated
CTS, but this resulted in the formation of highly flocculated structures.
According to the authors, these flocculated structures are formed by the
hydroxylated edge–edge interaction of the silicate layers.
Padilla-Ortega et al. (2016) developed functional bionanocompo-
sites based on the intercalation of CTS in natural or organically mod-
ified vermiculite (with hexadecyltrimethylammonium (HDTMA). Ul-
method
method
method
method
method
method

trasound was used for homogenization and the resulting materials were
processed as low density macroporous foams. FTIR characterization
Preparation method

blending
blending
blending
blending
blending
blending

showed that the intercalation of CTS was performed by electrostatic


attraction between the protonated amino groups and the negative
charge in the natural vermiculite layers. In the organically modified
Solution
Solution
Solution
Solution
Solution
Solution

vermiculite, stabilization was achieved by other types of interactions


with the intercalated HDTMA. Both bionanocomposites presented a
similar hydrophilic behavior, textural properties, pore distribution and
superficial morphology, however the CTS-vermiculite showed higher
thermal stability than the organovermiculite. More recently, Alba et al.
(2019) have explored the combination of brittle swelling micas and CTS
% w:w (Clay/
biopolymer)

for the preparation of eco-friendly bionanocomposites. The results from


the XRD, FTIR, and NMR spectroscopy analyses demonstrated that
1:10

when CTS is intercalated between mica layers, part of the CTS remains
1:1

1:1
1:5

bound to the mica surface by hydrogen bonds with SieO and AleO of
the basal plane. The surface coverage is quite high and the zeta-po-
tentials changed from a negative value in Na-mica to a positive value
for CTS-mica. Some authors have also combined CTS–clay nano-
composites with a hydrophilic polymer such as polyacrylic acid to
improve superabsorbent materials (Qiu et al., 2005).

3.1.2. Cellulose-based bionanocomposites


Chitosan/Fe2O3

CELL is a polysaccharide consisting in a linear polymer chain of b(1,


Starch/Ag

4)-linked D-glucopyranose monomers, and it is one of the most in-


Chitosan

Chitosan
Chitosan
Polymer

Starch

novative biopolymers for the development of adsorbents. Some of its


most attractive properties are biodegradation, biocompatibility, ab-
sence of toxicity, affordability, and production from renewable sources.
Several studies have shown that CELL nanoparticles can be used as
fillers to improve the mechanical and barrier properties of biocompo-
sites (Siqueira et al., 2010). However, CELL processing is usually
Mg-Al layered double hydroxide

complex. CELL has a very high melting point (260–270 °C) and is in-
soluble in water or in common solvents due to its partially crystalline
structure. Additionally, CELL chains are very densely packed through
Table 1 (continued)

hydrogen bonds. To overcome such drawbacks, chemical modification


Modified Ball clay
Clay/Organoclay

of the CELL structure is necessary, but because CELL decomposition


occurs at temperatures below its melting point any melt-compounding
Attapulgite

Kaolinite

technique should be avoided. Different methods can be used to obtain


non-flocculated dispersions of CELL nanocrystals in an appropriated
organic medium (Siqueira et al., 2010). One of the easiest methods for

6
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

grafting CELL is free-radical graft copolymerization which can be used intercalation between the silica sheets of clays via cationic exchange.
to change the behaviors and properties of CELL, allowing us for instance Due to an excess of positively charge from the quaternary ammonium
to alter its hydrophobic nature and to enhance clay exfoliation process groups in the cationic ST structure, these bionanocomposites showed
(Abbasian et al., 2017). Park et al. (2006) used CELL acetate butyrate- high adsorption capacity towards anionic pollutants. Overall, nanoclay-
grafted-maleic anhydride as a compatibilizer to synthesize CELL-based reinforced ST composites exhibit improved physical and chemical
nanocomposites. This compatibilizer was synthesized by radical graft properties without decreasing their biodegradation. In this respect, De
polymerization of maleic anhydride monomers onto cellulose acetate Carvalho et al. (2001) and Park et al. (2002) showed that dispersed clay
butyrate in the presence of 2,5-dimethyl-2,5-di(tert-butylperoxy) in thermoplastic ST matrix improves the mechanical, thermal, and
hexane. Yoshioka et al. (2006) obtained grafted poly(ɛ-caprolactone) barrier properties of composites.
chains to facilitate the clay delamination process using a combination of Pinto et al. (2016) synthetized a carbonaceous bionanocomposite
in situ polymerization and solvent intercalation. Abbasian et al. (2017) based on synthetic hectorite and cassava ST by pyrolysis at 850 °C. After
obtained a terpolymer of CELL-graft-Polychloromethylstyrene-graft- pyrolysis and acid digestion of clay, the resulting material had a partial
Polyacrylonitrile/Organomontmorillonite (OMt) by grafting chlor- graphitic structure with almost twice the specific surface area, and up
omethylstyrene onto CELL using free-radical polymerization. XRD, to 20 times more mesopore volume than the carbonaceous material
Transmission Electron Microscopy (TEM), and Scanning Electron Mi- obtained without the clay. Zubair et al. (2018) prepared ST-NiFe-LDH
croscopy (SEM) imaging of resulting bionanocomposite showed the (1:1) and ST-NiFe-LDH (2:1) by the coprecipitation method and their
absence of the peaks at 2θ = 5° characteristic of the OMt. SEM imaging adsorption capacities were compared to NiFe-LDH. The FTIR and
clearly showed the dispersion of the silicate layers into the polymer thermogravimetric analysis (TGA) showed better intercalation of ST
matrix owing to intercalated or partially exfoliated structures. into the NiFe-LDH interlayer and increased hydroxyl and carboxyl
Yadollahi et al. (2014) intercalated carboxymethyl cellulose (CMCELL) functional groups on the surface of ST-NiFe-LDH composites. SEM
into MgeAl LDH and NieAl LDH by co-precipitation methods. The d- images showed a rough and porous microstructure in ST-NiFe-LDH
spacing increased from 0.862 nm and 0.816 nm to 1.73 nm and 2.23 nm composites and confirmed the presence of small cracks on the surface of
for Mg-Al-LDH and Ni-Al-LDH CMCELL bionancomposites, respectively. the composites, which might be associated with the reaction between
Lastly, they found that changes in the pH from 2 to 10 resulted in in- the hydroxyl groups of ST and the functional groups of NiFe-LDH
creased swelling of the bionanocomposites, with a dramatic increase in during precipitation. More recently, Dinari and Dadkhah (2020) pre-
swelling when pH ≥10. pared an easily recoverable Fe3O4-Ag-LDH-ST bionanocomposite via an
in situ growth method. XRD and elemental mapping analysis revealed
3.1.3. Starch-based bionanocomposites the presence of Fe3O4 with a cubic inverse spinel structure and Ag
ST is a polysaccharide that is the primary energy reserve in plants nanoparticles. SEM and TEM images showed that the as-synthesized
where the excess glucose is stored as starch in roots, tubers, seeds, and nanoparticles were spherical in shape and the TGA showed that ST
stems. ST is composed of two glucose polymers: amylose (around of the played an effective role in stabilizing the hybrid nanoparticles.
granule) and amylopectin, accounting for 10%–30% and 70%–90% of
the mass of a ST granule, respectively. In amylose a series of α (1–4)- 3.1.4. Alginate-based bionanocomposites
linked D-glucose units form a long linear chain with polymerization ALG is a natural polysaccharide biosynthesized by bacteria and
ranging 300–10,000. Amylopectin is composed of amylose chains cross- brown seaweeds from Phaeophyceae species. ALG consists of linear co-
linked with α (1–6) bonds (Arora et al., 2018; Avérous and Pollet, 2014; polymers of units of 1,4-β-D-mannuronic acid and 1,4-β-L-guluronic
Xie et al., 2013). acid. Its physical properties are dependent on the sequences of these
A higher number of ST-based bionanocomposites have been pro- two acids and the polymer molecular weight. ALG has numerous ap-
duced, studied and reported in the literature in comparison to CELL- plications including environmental remediation due to its properties
clay biocomposites (Dinari and Dadkhah, 2020; Alcântara et al., 2014a; including its biocompatibility, low toxicity, chelating activity, and hy-
Avérous and Pollet, 2014; Azhar et al., 2014; Koriche et al., 2014; Xie drophilicity (Ucankus et al., 2018).
et al., 2013; Kampeerapappun et al., 2007; Xu et al., 2005). Xu et al. Following a co-precipitation method, ALG has been successfully
(2005) synthesized four different ST-organoclays nanocomposites by trapped between the ZnAl LDH layers. This method consists of the
melt-intercalation method and found that the glass transition tem- “coorganized assembly” synthesis of the LDH in the presence of the
perature increased by 6–14 °C depending on the type of clay. The in- polymer (Leroux et al., 2004). An interesting feature of the resulting
tercalation of ST acetate improved the thermal stability and mechanical ALG-LDH nanocomposites is that the characteristic AEC of the pristine
properties of the nanocomposite. In order to enhance the nano-disper- LDH is reversed to a CEC (Darder et al., 2005b; Leroux et al., 2004;
sion state, bionanocomposites based on plasticized ST with polyol have Leroux and Besse, 2001) due to the strong electrostatic interactions
been prepared with Mt. There is a strong influence of the polyol plas- between the negatively charged carboxylate groups of ALG and the
ticizer on the exfoliation process due to the hydrogen bonds between positively charged sites of the LDH sheets (Darder et al., 2005b; Leroux
glycerol and Mt. platelets (Chivrac et al., 2009). Kampeerapappun et al. et al., 2004). Alcântara et al. (2014a) evaluated the effect of the in-
(2007) prepared cassava ST-CTS-Mt bionanocomposites to promote Mt. corporation of sepiolite and palygorskite (fibrous clays) into neutral
platelet exfoliation, which led to few clay aggregates and improved (ST), cationic (CTS), and negatively anionic (ALG) polysaccharide
mechanical properties. Chivrac et al. (2008) also used this method matrices and found that interactions occur between the hydroxyl
known as “cationic polysaccharide surfactant” to promote clay ex- groups in the biopolymer backbone and the silanol groups on the clay
foliation, using cationic ST and OMt. After organo modification, the surface. In the case of ALG and CTS the presence of other functional
nanofiller was incorporated into wheat ST plasticized with glycerol by groups, carboxylate and amino groups, respectively, may be also in-
mechanical kneading. XRD characterization showed no diffraction volved in the interactions, as shown by FTIR spectra. The incorporation
peak, suggesting an exfoliated morphology. Moreover, TEM confirmed of clays also resulted in increased viscosity, indicative of the formation
the nanoscaled dispersion and showed that the use of this surfactant led of hydrogen bonds. These interactions led to improved mechanical
to a non-aggregated structure. The authors assumed that this dispersion properties and higher degree of water resistance of these biomaterials
state is due to the preferential interactions between the OH groups of compared to pure biopolymer films.
the different starch-based biomacromolecules. Koriche et al. (2013, Ahmad and Mirza (2017) developed a novel and ecofriendly ALG-
2014) also prepared bionanocomposites by previous modification of ST Au-mica bionanocomposite with improved mechanical strength. Au
with cationic groups and layered clays (Mt and bentonite). XRD analysis nanoparticles were incorporated due to the excellent biocompatibility,
showed that the presence of these cationic groups promotes the easy preparation and adsorbent capacity owing to their large specific

7
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

surface area. Au nanoparticles were synthesized via green procedures than the height of a βCD cavity (0.8 nm), which indicated that the
using glutathione and oxalic acid. SEM images revealed a hetero- polymer is tightly intercalated between the clay layers. XRD showed no
geneous cross-linked structure with dispersed Au nanoparticles. TGA clear diffraction and increasing intensity at low angles upon adsorption
analysis also showed that the synthesized bionanocomposite decom- of the pCD0, indicating exfoliation of the clay structure resulting from
poses at higher temperature with good thermal properties as compared polymer adsorption in a very extended conformation. pCD+-Mt was
to pure ALG. characterized by a dense, compact structure of closely packed clay
In another attempt, Darder et al. (2017) prepared a bionano- tactoids while pCD0-Mt was characterized by an open, fluffy structure
composite based on ALG, potato ST and sepiolite as reinforcing filler by of loosely aggregated clay tactoids. Prior to the investigations of
lyophilization and that exhibited low density and mechanical and fire Shabtai et al., hybrids of sepiolite with modified CD (randomly me-
resistance. FTIR analysis confirmed the strong interaction between the thylated) had been explored by Mura et al. (2016) for improving ox-
clay and the polysaccharide chains through a highly perturbed aprozin dissolution properties and, consequently, its therapeutic effi-
stretching vibration band of −OH in the silanol groups of sepiolite. The ciency and safety.
authors also observed that the porosity of the biomaterial can be
modulated by varying the sepiolite content, with decreased porosity as 3.1.7. Carrageenan-based bionanocomposites
the sepiolite content increased thereby allowing the modulation of Carrageenans are a family of linear sulfated polysaccharides ob-
certain properties like mechanical resistance. tained from a red seaweed which are also widely used in the food in-
dustry. Electrostatic interactions occur between the negatively charged
3.1.5. Xanthan gum-based bionanocomposites carrageenan and positively charged molecules. There are many studies
Xanthan gum (XG) is a natural polysaccharide whose backbone is a available on the interactions of carrageenans with different cationic
β-(1–4)-D-glucose (CELL). Every alternate glucose residue has a side compounds, but there is little literature about their use as adsorbents
chain of two mannoses and one glucuronic acid residues. XG is con- for the removal of organic pollutants from water.
sidered an anionic polyelectrolyte. Its excellent properties (availability, As for ALG, bionanocomposites based on i-carrageenan-Zn–Al LDH
affordability, biodegradability and the presence of hydrophilic func- have also been prepared by co-precipitation methods (Darder et al.,
tional groups–OH and –COOH) make this natural gum suited for dif- 2005b). The incorporation of carrageenans resulted in non-intercalated
ferent applications (Ahmad and Mirza, 2018a, 2018b, 2018c). and poorly crystallized materials (Darder et al., 2005b; Ruiz-Hitzky
Ahmad and Mirza (2018a) developed a novel XG-Mt bionano- et al., 2005). XRD analysis showed that i-carrageenan seems to be in-
composite that appeared as a densely bounded network structure with corporated as a bilayer or, most likely, as a double helix (ΔdL of 1.2 nm)
porous and irregular surface on SEM images and highly dispersed within the LDH structure.
particles on TEM images. The same authors also reported that grafting Since kappa-carrageenan contains high contents of sulfate and hy-
with functional glutathione (with amine, thiol, and carboxylate groups) droxyl functional groups, it may potentially be miscible with poly(vinyl
could increase the adsorption capacity of XG (Ahmad and Mirza, alcohol) (PVA) due to the formation of hydrogen bonds. PVA has been
2018a). They prepared and characterized a XG-Glutathione-Zeolite extensively used in many biomaterial applications due to its unique
bionanocomposite which was effectively applied for the adsorption of properties such as easy preparation, good biodegradability, and ex-
heavy metals. They also synthetized a XG-n-acetyl cysteine-modified cellent mechanical and chemical resistance (Park et al., 2001).
mica which increased metal chelation and the number of binding sites Hosseinzadeh et al. (2015) developed a hybrid kappa-carrageenan-
(Ahmad and Mirza, 2018b). The SEM image of the bionanocomposite PVA-Mt nanocomposite by crosslinking the polymers with the free-
revealed a highly irregular surface that became smoother after Pb(II) zing–thawing technique and subsequently with K+ ions. The FTIR
loading, suggesting a successful adsorption of Pb(II). analysis revealed the interactions of Mt. with the functional groups of
Liu et al. (2011) prepared a macroporous foam from CTS-XG-Mt via kappa-carrageenan and PVA and XRD and TEM images showed a
freeze-drying. Their results suggested that the main factor governing layered morphology due to the penetration of polymer chains into Mt.,
the structural hardness was the freezing rate during preparation. Slow which showed an intercalated and exfoliated structure.
cooling rate during freezing led to larger pore sizes and a more tilted
pore arrangement. XRD and TEM characterization was consistent with 3.2. Protein-based bionanocomposites
an exfoliated structure. Mt. also improved the hardness of the prepared
foams. 3.2.1. Zein-based bionanocomposites
Zein is the primary storage protein of maize. This globular protein,
3.1.6. Cyclodextrin-based bionanocomposites with its greater concentration of nonpolar amino acids, is 50-fold more
β-Cyclodextrin (βCD) is an oligosaccharide formed by seven glucose hydrophobic than other proteins like albumin or fibrinogen. Zein is
units linked by α-(1,4) glycosidic bonds in a cyclic structure. βCD has insoluble in water but soluble in aqueous ethanol (60–95% (v/v)),
an internal hydrophobic cavity (0.78 nm width) that allows the for- aqueous solution at pH > 11, some organic polar solvents (e.g., pro-
mation of a size-specific inclusion complex between βCD and hydro- pylene glycol and acetic acid) and certain anionic detergents (Shukla
phobic molecules. The possibility of cross-linking or anchoring βCD to and Cheryan, 2001). Its molecular weight ranges between
solid supports allows the formation of insoluble sorbents with higher 23,000–27,000 Da, and this protein has already been used in many
adsorption capacity for the removal of micropollutants than activated applications such as adhesives and cosmetics, and to generate drug-
carbon (Shabtai and Mishael, 2018). delivery matrices alone or in combination with other compounds. Zein
Shabtai and Mishael (2018) designed a novel composite sorbent is also used in the food industry because of its ability to form hydro-
based on Mt. and neutral (pCD0) or cationic (pCD+) CD by crosslinking phobic grease-proof coatings and biodegradable films that provide a
βCD with epichlorohydrin and incorporated cationic glycidyl tri- barrier to moisture and oxygen. Because of these characteristics, the use
methylammonium chloride (GTMAC) groups. The incorporation of of zein might be a promising method to reduce the hydrophilicity of
GTMAC significantly decreased the polymer molar mass and hydro- pristine clays, as in conventional organoclays with long-chain alky-
dynamic radius, confirmed by 1H NMR and FTIR analysis. pCD0 showed lammonium cations (Alcântara, 2013; Alcântara et al., 2012).
better adsorptive properties than pCD+ due to the electrostatic repul- In 2012, Alcântara et al. combined zein with two fibrous clays
sion between the highly charged chains of pCD+ and because of the (sepiolite and palygorskite) and found that the amount of adsorbed zein
preferential adsorption of higher molar mass pCD0 (Shabtai and can be adjusted to ensure the complete coverage of the clay fibers, but
Mishael, 2018). The XRD analysis showed that the increase in basal zein aggregates were found at high equilibrium concentration of zein.
spacing upon adsorption of pCD+ (1.92 nm) was only slightly larger Due to the higher external surface area of the zein-sepiolite

8
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

nanocomposites, they exhibit larger amount of assembled protein than modulus, tensile strength, break elongation, crystallization rate, and
zein-palygorskite. The assembling mechanism is related to the inter- other mechanical properties. More importantly, the presence of clay
action of the amide groups of zein with silanol groups on the surface of decreases the gas and water vapor permeation, increases the heat dis-
the fibrous clays. In the same work the authors also studied the prop- tortion temperature and scratch resistance, and controls the biode-
erties of zein–fibrous clay biohybrids as additives in the preparation of gradation of the PLA matrix.
nanocomposites when using ALG as a polymer matrix (Alcântara et al., Neppalli et al. (2014) prepared different (PLA)-based nanocompo-
2016a; Alcântara et al., 2012). The ALG-Zein-fibrous clay bionano- sites filled with different kinds of cationic (Mt) and anionic (perkalite)
composite showed reduced water vapor transmission and improved clays. A similar degree of dispersion was achieved in both types of clays
barrier properties towards oxygen even at high humidity conditions. but different structural and morphological changes. Perkalite induced
The intercalation of zein on Mt. is a complex process which is in- higher crystallinity, a faster crystallization rate and also a modification
fluenced by the structural stability of the protein, the ionic strength, the of the crystallization mechanism with preservation of the lamellar
pH value as well as the surface properties. The same authors (Alcântara framework of PLA while Mt-based nanocomposites resulted in dis-
et al., 2016b) prepared a zein-Mt biohybrid following two synthesis ordered lamellar stacks. Perkalite-PLA also underwent faster degrada-
methods (Alcântara et al., 2016b): 1) using zein dissolved in 80% (v/v) tion than either the PLA matrix or the Mt-PLA. These findings show that
ethanol/water solution, and 2) using a sequential process that involves the choice of the chemical nature of the nanofiller allows us to mod-
the previous separation of zein components in absolute etanol. The ulate both the degradation rate of PLA and its structure.
latter resulted the most effective method. Their results also suggested Bionanocomposites based on PLA-LDH with improved mechanical
that these bionanocomposites could be associated with other polymers and thermal properties have also been investigated but to a lesser extent
of different nature, representing a promising eco-friendlyalternative to than natural clays. The studies availabe are very limited and because of
organoclays based on alkylammonium cations. the strong electrostatic interactions between the highly charged hy-
droxyl layers and intercalated anions the delamination of LDH layers at
3.2.2. Gelatin based bionanocomposites the nanolevel is very difficult to achieve (Ray, 2012; Chiang and Wu,
Gelatin (Gel), a natural polymer, is the denatured form of the pro- 2010; Zhou and Xanthos, 2010).
tein collagen and its constituents are amino acids such as glycine,
proline, hydroxyproline, glutamic acid and arginine. The growing in- 4. Methods for preparation of biopolymer-clay nanocomposites
terest in Gel is due to its low antigenic capacity. However, like carra-
geenan, there is little literature about its application in the environ- Nanofiller dispersion is largely dependent on the physical and
mental remediation field. chemical affinitty between the clay material and the polymer matrix
The investigations on the interactions between Gel and layered si- and the preparation method. Depending on the interactions between
licates in 1950 (Talibudeen, 1950) laid the basis for the development of nanofiller and polymer matrix, the following types of composite can be
novel Gel-based bionanocomposites derived from clay minerals and described (see Fig. 4):
layered perovskite (Darder et al., 2006; Ruiz et al., 2006). The assembly Phase-separated microcomposites are formed when the polymer is
Gel-perovskite results in a dielectric permittivity greater than the in- unable to intercalate between the silicate sheets. As the individual
crease in the dielectric loss values in biohybrid films, which makes it layers are not separated, the clay agglomerates act as microparticles
useful in the microwave industry or in high-frequency devices (Ruiz dispersed in the polymeric matrix. For this reason, these composites are
et al., 2006). frequently named “microcomposites”. This lack of uniform dispersion
The experiments conducted by the Ruiz-Hitzky group showed that makes their properties particularly poor (Valapa et al., 2017).
the uptake of Gel by clay minerals is influenced by the pH, whose values Intercalated nanocomposites are obtained when polymer chains are
must be adjusted below the isoelectric point of the biopolymer. For inserted between the sheets of the layered clay resulting in an enhanced
instance, the intercalation of Gel in homoionic smectites or in Mt. oc- interlayer distance (Valapa et al., 2017).
curs at low pH values required to protonate the amino groups of the Exfoliated nanocomposites are produced when individual clay layers
protein. Thus, the positively charged biopolymer is able to replace to- are separated from one another and are homogeneously dispersed
tally or partially the sodium ions mainly located on the interlayer space within the polymer matrix. This homogeneous dispersion provides na-
of the clays (Ruiz-Hitzky et al., 2005; Weiss, 1969; Talibudeen, 1950). It nocomposites with enhanced mechanical, thermal, and barrier prop-
has been suggested that intercalation of proteins into Mt. may result in erties. However, complete exfoliation is usually challenging (Valapa
distortions and other structural disturbances in the proteins due to their et al., 2017).
confinement in the interlayer space of the clay (Theng, 1979). Ruiz- To enhance the intercalation/exfoliation process into a polymer
Hitzky et al. (2005) suggests that these parameters and those related matrix, a chemical modification of the clay surface to match the
with the expected loss of entropy associated to the intercalation process polymer polarity, is often conducted (Alexandre and Dubois, 2000).
should be investigated using advanced spectroscopic techniques such as Cationic exchange is the most common technique, but other methods
solid state high resolution NMR combined with microcalorimetric such as organosilane grafting (Dai and Huang, 1999), the use of iono-
measurements. mers (Shen et al., 2002) or block copolymers adsorption (Fisher et al.,
1999) have also been used (Valapa et al., 2017; Chivrac et al., 2009). In
3.3. Polylactide-based bionanocomposites this sense, Chivrac et al. (2009) provided us with an in-depth review of
the main strategies used to develop polysaccharide-based biocompo-
PLA is formed by polymerization of monomers of lactic acid (LA) sites with Mt. with a discussion of their dispersion state and properties.
obtained from bacterial fermentation of carbohydrates. PLA is a ther- There are three main methods for the clay incorporation into the
moplastic polyester completely biodegradable and bioabsorbable that polymer matrix: (i) solution blending; (ii) polymer melt blending; and
derives from renewable sources (Avérous, 2013, Ray, 2012; Ahmed and (iii) in situ polymerization method (Bhawani et al., 2018; Guo et al.,
Varshney, 2010). High molecular weight PLA is synthesized by ring- 2018; Ojijo and Ray, 2013; Chivrac et al., 2009; Alexandre and Dubois,
opening polymerization of LA, normally using aluminum and tin alk- 2000). As shown in Table 1, solution blending is the most appropriate
oxides as catalysts. PLA was one of the first biodegradable man-made for biopolymers.
polymers because of its renewable origin, controlled synthesis, me-
chanical properties, and inherent biocompatibility (Ray, 2012). The 4.1. Solution-blending
intercalation of clays with PLA allows the development of bionano-
composites with enhanced properties such as improved storage In this method, the polymer is dissolved in a suitable solvent,

9
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

Fig. 4. Schematic showing the polymer-clay nanocomposite morphologies: flocculated, intercalated and exfoliated (modified from Valapa et al., 2017).

typically water for biopolymers, and the clay is simultaneously dis- morphology. The latter helps to induce platelet delamination from the
persed in the same solvent or a different one to obtain a miscible so- clay tactoids. This method is eco-friendly because no solvent is in-
lution. The polymer solution is then intercalated between the clay volved, but the thermal or thermo-mechanical inputs result in partial
dispersion. The resulting solution is homogenized and the solvent degradation of chains and the high residence times required for clay
evaporated (Valapa et al., 2017). For non water-soluble polymers, this exfoliation induce matrix degradation. These drawbacks make this ap-
process involves the use of large amount of organic solvents, which is proach unsuitable for biopolymers, except for PLA or the thermoplastic
environmentally unfriendly and costly (Bhawani et al., 2018). More- ST modified by additives (Dang et al., 2020; Ren et al., 2018; Neppalli
over, a small amount of solvent may remain in the final product at the et al., 2014). Fig. 5B shows a schematic of the melt-blending method.
polymer-clay interface thus reducing the interactions between the
polymer and the clay surfaces (Jin et al., 2002).
As mentioned above, solution blending is the most common method 4.3. In-situ polymerization
(see Table 1) because many of the biopolymers used to produce hybrid
bionanocomposites, such as CTS, CELL or pectin, cannot be melt pro- In a first stage clays are swollen into a monomer solution where
cessed due to high thermal or thermo-mechanical degradations. The use polimerization starts and propagates (Chivrac et al., 2009). The poly-
of in-situ polymerization is also limited as most biopolymers are natu- merization initiating agent can be heat or a suitable chemical
rally availabe in the polymeric form rather than as monomers (Chivrac (Soetaredjo et al., 2018). The molecular weight of macromolecules in-
et al., 2009). Fig. 5A shows a schematic of the solution-blending creases, leading to a d001 increase and sometimes to an almost fully
method. exfoliated morphology (Ray and Okamoto, 2003). This method has
been applied as an alternative to melt blending for in situ poly-
merization of intercalated LA monomers for the formation of exfoliated
4.2. Melt-blending systems (Paul et al., 2005). As mentioned above, the polysaccharide
chains are naturally formed during plant growth and are extracted from
Melt blending is a high temperature process where the clay is in- plants already in their polymeric form making this technique unsuitable
corporated simultaneously to the annealed polymer matrix into a melt for the preparation of polysaccharide-based bionanocomposites
mixing device and the resulting composite is kneaded for homogeneous (Chivrac et al., 2009). Fig. 5C shows a schematic of the in-situ poly-
distribution (Valapa et al., 2017). Two parameters can be used for merization method.
optimization of clay dispersion (Soetaredjo et al., 2018): the residence Other non-traditional preparation techniques such as electrospin-
time and the shearing. The first one is needed to allow diffusion of the ning and processing under supercritical conditions (e.g., supercritical
polymer chains into the inter-layer gallery to obtain an exfoliated carbon dioxide), microwave-induced, one-template-directed, self-

10
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

Fig. 5. Techniques for bionanocomposite preparation: (A) solution blending; (B) melt blending; (C) in-situ polymerization (reproduced from Valapa et al., 2017 with
permission).

assembly or intermatrix synthesis have been developed in order to cannot access the interlayer space (Bee et al., 2017).
shape the material in the desired configuration (Bhawani et al., 2018). However, in some cases the XRD analysis alone is not reliable en-
ough to determine nanocomposite structure because it does not provide
information about the spatial orientation of the silicate layers, which
5. Characterization techniques of biopolymer-clay complicates the interpretation of data. For example, when mixed-clay
nanocomposites morphologies are obtained the measurement of the interlayer spacing
might be inaccurate as the basal reflections originate from different
A key issue in polymer-clay nanocomposite research is the de- components. Additionally, the absence of diffraction peaks in randomly
termination of the properties related to the adsorption behavior such as distributed intercalated clays might lead to wrongly identify structures
morphology, expansion capacity, layer charge and charge arrangement, as exfoliated. Lastly, clay dilution can also result in peak broadening or
pore size and pore arrangement. The most commonly methods applied peak absence even without delamination (Bee et al., 2017).
for characterization of clays are XRD, TEM, SEM, FTIR, NMR, and TGA XRD should be complemented with TEM because this technique can
(Alexandre and Dubois, 2000).. provide qualitative information about the morphological features and
Due to its ease of operation and availability, XRD has been widely spacial arrangement of the clay as well as visual information of any
employed to evaluate the resulting hybrid structure after dispersing the internal flaws. The main advantages of TEM are its magnifica-
nanofiller in the polymeric matrix (Bee et al., 2017). Through the tion—typically between 50,000 to 10,000,000×—and the ability to
evaluation of the presence and shift of the basal reflections we can visualize specimens and to provide diffraction data. TEM images allow
determine the type of nanocomposite structure (phase-separated, in- the visualization of the spatial distribution of clay layers. Clay layers
tercalated, or exfoliated). XRD can also be used to examine the kinetic show strong contrast because they consist of elements with high atomic
properties of the polymer melt blending. The intensity and shape of the number (Si, Al and O), whereas the polymer matrix basically contains
basal reflections are indicative of the degree of dispersion of the na- C, H and N. Consequently, the intersections of clay layers appear as
nofiller within the polymer matrix. For example, for exfoliated struc- dark lines. TEM images of phase-separated composites are darker owing
tures, the disappearance of the diffraction peak suggests the complete to the presence of clay aggregates resulting in electron scattering. In
distribution of clay aggregates in the polymer matrix with the disrup- contrast, TEM of exfoliated composites shows brighter images owing to
tion of the layer structure of the clay. For intercalated structures, the the homogeneous dispersion of clay layers within the polymer matrix.
expansion of the interlayer spacing after polymer intercalation is de- As with TEM, SEM also relies on the use of high-energy electrons to
monstrated by the shift of the basal reflections towards lower diffrac- generate images of the surface of solid specimens as well as providing
tion angles. In contrast, for conventional microcomposite structures, the information about their dispersion, crystalline structure, composition
basal reflections remain unchanged. This is due to the clay aggregates and orientation (Bee et al., 2017). However, due to the limited
retaining their stacking arrangement because the polymer chains

11
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

resolution of SEM instruments (around 1 nm), detailed information of clay used in the formulation. In general, the breakdown of the biona-
the nanoparticle distribution within the polymer matrix is difficult to nocomposite begins at approximately 250 °C with the decomposition of
obtain. In contrast, it is useful for imaging fractured surfaces. Recently, organic components and finishes at approximately 600–630 °C with the
a combination of SEM and energy dispersive X-ray spectroscopy is destruction of the crosslinked network structure and the complete
being used to determine the degree of dispersion of nanoparticles on the breakage of clay mineral chains (Li et al., 2018; Swain et al., 2018;
fractured surface of bionanocomposites (Ojijo and Ray, 2013). Darder et al., 2017; Mahdavinia et al., 2016; Azhar et al., 2014; Ray and
FTIR spectroscopy is also used to understand the structure of the Okamoto, 2003). Lastly, several authors have examined the thermal
polymer nanocomposite allowing us to differentiate between the stability of ST-based nano-biocomposites to establish the relationships
bonding in a polymer from the bonding in a nanocomposite. This between clay dispersion and thermal stability. Park et al. (2002) used
technique is fast and involves minimal sample preparation and as a TGA to show the higher degradation temperature of potato ST-Mt and
result has become a well-accepted method for sample characterization. OMt in comparison to the neat matrix. Moreover, thermal stability of
However, the bonding differentiation may be difficult to understand potato ST-Mt was higher than OMt which suggests a relation between
even though intercalation has already occurred. As mentioned in pre- Mt. dispersion and thermal stability.
vious sections, multiple studies have used FTIR for characterizing the
bands attributed to the intercalated biopolymer in the clays structure 6. Applications of biopolymer-clay nanocomposite as adsorbents
(Alba et al., 2019; Wang et al., 2005; Darder et al., 2003). This tech-
nique is very helpful to determine the electrostatic interactions between Technological advances in analytical methods have made possible
byopolimers and clays. Darder et al. (2017) used FTIR to assess the to detect and quantify a wide range of chemicals in the various en-
interactions between sepiolite and ST and ALG and found a consider- vironmental compartments. These chemicals include pesticides, heavy
able perturbation of the stretching vibration band of the silanol groups metals, polycyclic hydrocarbons, dioxins, and others as well as those
located at the sepiolite surface, which suggest that the interactions known as “emerging contaminants” (plasticizers, flame retardants,
occur between the external OH groups of the silicate and the hydroxyl perfluorinated compounds). An extensive number of publications have
groups of the polysaccharide chains. The band was shifted towards described levels of contaminations worldwide and in all the environ-
lower frequency values, leading to a decrease in absorbance with re- mental compartments. Adsorption is considered one of the most effec-
spect to that assigned to the stretching vibration of Mg-OH that remains tive techniques for the treatment of contaminated water, with the de-
unaltered. In another attempt, Darder et al. (2005b) and Leroux et al. velopment of novel adsorbents in recent years (Guo et al., 2018). In this
(2004) observed using FTIR measurements strong electrostatic inter- respect, biopolymer-clay nanocomposites have shown high adsorption
actions between the negatively charged carboxylate groups of ALG and capacity and long life cycle owing to their easy regeneration, effective
pectin biopolymers and the positively charged sites of the LDH sheets, cation exchange, large surface area, and relative affordability and lack
while a weak interaction was observed between the i-carrageenan and of toxicity.
the LDH. The adsorption capacity, selectivity and recovery of bionano-
Bulk characterization techniques such as NMR are also be used to composites is largely affected by (1) the modification technique, the
evaluate the morphology, surface chemistry, and the quality of clay presence of exchangeable cations, nanoclay morphology, and the pre-
dispersion by the measurement of the 1H, 13C, 23Na, 27Al or 29Si spin- sence of water molecules in the interlayer; (2) pH, temperature, ionic
lattice relaxation time (T1) (Ray and Okamoto, 2003). VanderHart et al. strength, pollutant concentration, and the presence of other ions or
(2001) first used NMR (1H and 13C) to quantify the degree of clay ex- compounds. There are multiple mechanisms involved in the adsorption
foliation of nylon 6-Mt nanocomposites and found a potential re- process, including electrostatic attraction, Lewis acid-base interaction,
lationship between measured T1H values and the quality of the clay hydrogen bonds or ion-exchange. Table 2 shows a summary of scientific
dispersion. When the clay particles are stacked and poorly dispersed in papers on bionanocomposites as adsorbent materials for the removal of
the polymer matrix, the average distances between polymer-clay in- emerging organic and inorganic pollutants such as pesticides, BPA,
terfaces are greater, and the average paramagnetic contribution to T1H hydrophobic organic pollutants, dyes or heavy metals, from water
is weaker (Ray and Okamoto, 2003). Alba et al. (2019) used 29Si and samples. The adsorption of the different contaminants is discussed
23
Na magic angle spinning (MAS)-NMR spectroscopy to show CTS in- below.
corporation into mica but part of the biopolymer remains at the outer
surface where it binds to the SieO and AleO of the basal plane. On the 6.1. Pesticides
other hand, as mentioned above, both FTIR and 13C NMR determina-
tions also help to confirm the ability of CTS–smectite nanocomposite to Pesticides have been extensively used to evaluate the efficacy of
act as an anionic exchanger when the biopolymer is arranged as a bi- bionanoacomposites as adsorbent materials. Organophosphate pesti-
layer in the interlayer space (Darder et al., 2005a). cides have largely replaced organochlorine pesticides because of the
Finally, TGA provides information about the thermal stability of persistence and accumulation of the latter in the environment. One of
bionanocomposites. There are many studies that have shown the in- the first studies on bionanocomposite adsorption capacity was the one
creased stability of biopolymers in bionanocomposites. Intercalation of by Alcântara (2013) who developed a bionanocomposite with zein-se-
biopolymers into clay mineral layers increases the decomposition piolite modified with magnetite nanoparticles. The main properties of
temperature in 40–50 °C as the clay acts as a heat barrier in the na- this magnetosorbent are its water stability and the mechanical prop-
nocomposite structure, which enhances the overall thermal stability of erties owing to the reinforcement provided by sepiolite fibers. The re-
the system, and helps in the formation of char after thermal decom- sulted material has been successfully applied to adsorb 4-chloro-2-me-
position. Initially, this heat barrier effect increases the decomposition thylphenoxy-acetic acid, indicating their application in environmental
temperature and subsequently it serves as a heat source to accelerate remediation (Alcântara, 2013). One of the main advantages of this type
the decomposition process, in conjunction with the heat flow supplied of adsorbent is that it can be easily recovered from the aqueous media
by the outside heat source (Ray and Okamoto, 2003). Typical TGA by applying an external magnetic field.
curves show degradation, measured as mass loss, as a function of Acidic pesticides are weakly sorbed to soil, which increases the risk
temperature (dynamic heating) or time (isothermal heating). The mass of contamination of ground and surface water. Sorption depends on pH
loss that occurs between ambient temperature to around 170 °C is as- and acidic pesticides are usually found in their anionic form at the pH of
sociated with the loss of water from the surface and interlayers, water soil and water environments (Celis et al., 2012). Clay-based adsorbents
retained in the biopolymers and intersticial water. The process of usually show a high affinity for cationic and very polar pesticides, but
thermal decomposition of the system depends on the biopolymer and low affinity for anionic pesticides owing to the negative charges on the

12
Table 2
Adsorption experiments on biopolymer-clay composites as adsorbent materials for the removal of organic and inorganic pollutants from water.
M.d.M. Orta, et al.

Group Analyte Bionanocomposite Qmax (mg/g) Isotherme Kinetic Reference

Pesticides Paraquat Mt/ALG 74.3 Langmuir – Etcheverry et al., 2017


Clopyralid Mt/CTS 96.0 Freundlich – Celis et al., 2012
Dicamba Laponite/ST 251.9 Langmuir Pseudo- second order Pinto et al., 2016
Phenols and hydrophobic organic pollutants Bisphenol A Mt/β-CD+ 0.1 Langmuir Pseudo-second order Shabtai and Mishael, 2018
4-nitrophenol Ag/LDH/ST – – – Dinari and Dadkhah, 2020
Phenol Mt/CTS 11.0 – – Zhu et al., 2017
PCB 3-Cl Mt-ALG 332.6 Langmuir Pseudo-second order Barreca et al., 2014
PCB 4 Cl Mt-ALG 70.0 Langmuir Pseudo-second order Barreca et al., 2014
PCB 5 Cl Mt-ALG 55.0 Langmuir Pseudo-second order Barreca et al., 2014
PCB 6 Cl Mt-ALG 44.9 Langmuir Pseudo-second order Barreca et al., 2014
Dyes Crystal violet Bentonite/Guar gum 167.9 Langmuir Pseudo-second order Ahmad and Mirza, 2018a
Conger red Mt/CTS 350.0 – – Zhu et al., 2017
Crystal violet Mt/CTS 450.0 – – Zhu et al., 2017
Amido Black 10B Bentonite/CTS 323.6 Langmuir Pseudo-second order Liu et al., 2015
Reactive black 5 Mt/CTS – – – Zhu et al., 2015
Crystal violet Sepiolite-Kappa-carrageenan-g-poly(acrylamide) 47.0 Langmuir Pseudo-second order Mahdavinia and Asgari, 2013
Methylene blue Modified Ball clay/CTS 142.0 Langmuir Pseudo-second order Auta and Hamed, 2014
Basic blue 9 Mt/CTS 700.0 – – Monvisade and Siriphannon, 2009
Basic blue 66 977.0
Basic Yelow 300.0
Methylene blue Mt/CTS-g-poly 1859.0 Langmuir Pseudo-second order Wang et al., 2008
(acrylic acid)
Tartrazine Bentonite/CTS 294.1 Langmuir Pseudo-second order Wan Ngah et al., 2010a

13
Malachite green Bentonite/CTS 435.0 Langmuir Pseudo-second order Wan Ngah et al., 2010b
Methyl Orange Chitosan Kaolin/γ-Fe2O3 14.2 – – Zhu et al., 2010
Inorganic methals Cr(VI) Red palygorskite/CTS-g-PMMA 92.5 – – Sethy and Sahoo, 2019
Pb(II) Bentonite/Guar gum 18.9 Freundlich Pseudo-second order Ahmad and Mirza, 2018a
Cu(II) Mica/ALG 169.8 Langmuir Pseudo-second order Ahmad and Mirza, 2018b
Pb(II) Vermiculite/CTS 166.7 Langmuir Pseudo-second order Chen et al., 2018
Cd(II) Vermiculite/CTS 58.5 Langmuir Pseudo-second order Chen et al., 2018
Se(IV) LDH/CTS 17.0 Langmuir Pseudo-second order Liu et al., 2015
Se(VI) 12.0 Langmuir Pseudo-second order
Pb(II) Mica/ALG 25.7 Freundlich Pseudo-second order Ahmad and Mirza, 2017
Cd(II) Mt/CTS 8.0 – – Zhu et al., 2017
Cu(II) Mt/Fe3O4-ST-polyamidoxime 158.7 Langmuir Pseudo-second order Mahdavinia et al., 2016
Cd(II) Vermiculite/CTS 2047.0 Langmuir Pseudo-second order Padilla-Ortega et al., 2016
Cd(II) Vermiculite-HDTMA/CTS 1211.0 Langmuir Pseudo-second order Padilla-Ortega et al., 2016
Nd(III) Sepiolite/Sacram 1929.0 Langmuir Pseudo-second order Alcântara et al., 2014b
Pb(II) Sepiolite/CTS 100.0 – – Alcântara et al., 2014a
Cu(II) Sepiolite/CTS 75.0 – – Alcântara et al., 2014a
Pb(II) Palygorskite/ALG 290.0 – – Alcântara et al., 2014a
Cu(II) Palygorskite/ALG 65.0 – – Alcântara et al., 2014a
Cr(VI) Bentonite/ST 4.0 Langmuir Pseudo-second order Koriche et al., 2014
Cr(VI) Mt/CELL 22.2 Langmuir Pseudo-second order Kumar et al., 2012
Pb(II) Bentonite/CTS 26.4 Freundlich Pseudo-second order Futalan et al., 2011
Cr(VI) Mt/CTS 357.1 Langmuir Pseudo-second order Pandey and Mishra, 2011
Pb(II) Mt/ALG 244.6 Langmuir Pseudo-second order Shawky, 2011
Se Mt/CTS 18.4 Langmuir Pseudo-second order Bleiman and Mishael, 2010

ALG: alginate; β-CD: β-Cyclodextrin; CELL: cellulose; CTS: Chitosan; LDH: Layered double hydroxide; Mt.: Montmorichonite; ST: Starch.
Applied Clay Science 198 (2020) 105838
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

clay surfaces. The interaction of some organic cations, like alky- equipment, medical devices, dental fillings, thermal paper, toys, and
lammonium cations, with clay minerals can modify the affinity for others. There is a large body of research on the adverse effects of BPA
anionic pesticides by changing the nature of the clay mineral surfaces on reproduction and development, neural networks, metabolism, car-
from hydrophilic to hydrophobic or, if adsorbed in excess of the CEC of diovascular and immune system in in vitro assays and animal models.
the clay, even producing charge reversal. Adsorption of CTS on Mt. with BPA can enter the environment through different pathways such as
excess of the CEC of the clay mineral results in structures with good wastewater treatment plant effluents, industrial discharge and landfill
adsorption properties for anions because the –NH3+ groups not directly leachate (Martín et al., 2019b).
involved in the interaction with the clay surfaces can act as anionic Shabtai and Mishael, 2018synthetized a regenerable sorbent con-
exchange sites. Along these lines, Celis et al. (2012) synthetized a CTS- sisting of a polymer of βCD units modified with a cationic group
Mt bionanocomposite to be used as an adsorbent of the herbicide clo- (pCD+) and Mt. The design of this sorbent allows simultaneous ad-
pyralid from aqueous solutions and soil/water suspensions. The CTS-Mt sorption of BPA (which fills the βCD cavities) and anionic organic
composite proved to be a useful adsorbent under mild acidic conditions, matter (through electrostatic interactions). BPA adsorption was not
that is, when the herbicide is mainly in its anionic form (pKa = 2.3) affected by the high organic matter adsorption. Most of the works
and CTS in its cationic form (pKa = 6.3). High NaCl concentrations available on pollutant removal are not conducted using real wastewater
promoted the desorption of the adsorbed clopyralid from CTSeMt, samples, and they only test sorbents in suspension, no by column fil-
supporting ion exchange as the primary adsorption mechanism. tration, and they do not deal with sorbent regeneration. In contrast,
Amendment with CTS-Mt (5–10%) to acidic soil resulted in an in- Shabtai and Mishael (2018) used the selective elution of BPA and or-
creased clopyralid adsorption, whereas this effect was negligible for ganіc matter on the sorbent using alkaline solutions and brine, re-
basic soils. Moreover, the presence of easily reversible ad- spectively, to achieve successful in-column regeneration. They also
sorption–desorption interactions suggests that CTS-Mt-based formula- demonstrated that the removal performance of pCD+–Mt is superior to
tions of pesticides can yield slow release and consequently reduce an- that of activated carbon (Shabtai and Mishael, 2018) by assessing the
ionic pesticide mobility into soil and ground and surface water. A ST- simultaneous removal of 38 micropollutants, commonly found in
hectorite nanocomposite has also been assessed as adsorbent for di- treated wastewater, such as pesticides and pharmaceuticals. The au-
camba, an herbicide, from aqueous solutions and the findings show thors found that pCD + -Mt exhibited higher adsorption capacity to-
adsorption to be highly pH-dependent (Pinto et al., 2016) with higher wards 34 pollutants than activated carbon.
adsorption observed at pH values of 2.0–4.0 (pH < pHpzc) when the
pesticide (negatively charged) is attracted to the positive surface of the 6.3. Hydrophobic organic pollutants
bionanocomposite via electrostatic interactions. While at pH values of
6.0–11.5, the bionanocomposite surface is negatively charged (pH > Zhu et al. (2017) evaluated the capacity of Mt. modified with dif-
pHpzc), which explains the low herbicide adsorption capacity. More- ferent volumes of HDTMA-Mt and CTS (H/CTS-Mt) to adsorb hydro-
over, at pH values above 7.0, competition for the surface of the ST- phobic organic contaminants, heavy metals and dyes from water. Mt. is
hectorite bionanocomposite was observed between the negatively usually ineffective for the adsorption of hydrophobic organic pollutants
charged herbicide and OH groups of the medium. The isotherm studies owing to the hydrophilic characteristics of its surface and charges (Zhu
followed the typical Langmuir behavior. The capacity of this bionano- et al., 2017), but HDTMA-Mt can adsorb phenol thanks to the formation
composite to adsorb dicamba was 251.9 mg/g, related to a pseudo- of more effective organic phases (Zhu et al., 2017). After modification
second-order kinetic. with 2% and 4% CTS, the capacity to adsorb phenol was increased, but
Bionanocomposites have also been applied for the removal of ca- when the amount of CTS was higher than 6%, the adsorption decreased.
tionic pesticides from aqueous solutions. Etcheverry et al. (2017) pre- The adsorption isotherms followed a partition-dominated mechanism.
pared an ALG-Mt nanocomposite for the removal of paraquat, a qua- These authors suggest that cationic polymers might modify the ar-
ternary ammonium compound widely used as a contact herbicide to rangement of HDTMA (increase in packing density) within the clay,
control broad-leaved weeds. The adsorbed paraquat was linearly cor- which in turn would increase the amount of adsorbed hydrophobic
related with Mt. content, the maximum adsorption capacities (qmax) organic contaminants on the OMt. In contrast, an excess of cationic
increased from 0.093 to 0.278 mmol/g for Mt. contents from 0.5 to polymers results in higher packing density of HDTMA, which might
70%, respectively. The adsorption capacity of ALG is much lower than reduce the adsorption of hydrophobic organic contaminants due to
Mt., but ALG is also necessary as it acts as a support for clay particles. steric hindrance within the interlayer spaces of Mt.
As mentioned previously, a sustained release of pesticides into soils Polychlorinated biphenyls (PCBs) are another example of hydro-
is required for efficient bioavailability and environmental protection. phobic organic pollutants which consist of one to ten chlorine atoms on
Recently, Wang et al. (2019) have used an hydrophobic derivative of two benzene rings. Some PCB congeners do not easily degrade in the
sodium ALG, polyacrylamide and Mt. to synthetize stretchy nano- environment and although they are no longer manufactured, they
composite hydrogels with double networks to achieve a slower release persist for many years. Barreca et al. (2014) investigated the use of
of l-cyhalothrin. They also found that the hydrophobic interactions ALG-Mt beads for the removal of PCBs from aqueous solutions and their
between l-cyhalothrin molecules and the hydrophobic groups grafted findings revealed that higher initial PBC concentrations and longer
on the side chains of sodium ALG have an important role in the loading adsorption times improve adsorption. Adsorption of trichlorobiphenyls
efficiency and in the controlled pesticide release. The hydrogel con- on the bionanocomposite followed a Freundlich isotherm model, while
taining 5% Mt. showed the lowest cumulative release percentage of adsorption of tetra-, penta-, and hexachlorobiphenyls were best fitted to
6.68% over 87 h and the pesticide release profile followed the Weibull Chapman model. Kinetic followed a pseudo-second order model. By
model. These novel materials combine the advantages of ALG deriva- comparing the adsorption % using the same number of beads (10 beads,
tives (gelation) and nanoclays (adsorption capacity) to become carriers same volumen), best removals were obtained by using hybrid ALGeMt.
for enhancing the distribution efficiency of water-insoluble pesticides.
6.4. Dyes
6.2. Bisphenol A
Natural clays such as Mt. exhibit weak adsorption capacity of an-
BPA is a chemical produced in large volumes worldwide. Its main ionic dyes due to the repulsive electrostatic interactions occurring be-
uses are as a monomer for the synthesis of polycarbonate (~80%) and tween the negatively charged surface of the clays and the dye (Wang
epoxy (~18%) resins. These polymers are used as coatings in almost all and Wang, 2007a, 2007b). The work by Zhu et al. (2017) investigated
food contact materials, in the production of CDs and DVDs, electronic the adsorption capacity of Mt. modified with HDTMA and CTS for

14
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

Congo red (CR) and crystal violet (CV). Both HDTMA-Mt and HDTMA 2:1) for the removal of methyl orange (MO) from aqueous phase and
(60%)/CTS (2%)-Mt showed good adsorption capacity towards CR. their results showed that dye removal increased when the adsorbent
Their results showed an effective adsorption capacity of Mt. and dosage was increased (from 2 to 10 mg) for 1:1 and 2:1 bionano-
HDTMA (60%)/CTS (2%)-Mt to CR despite both adsorbents having composites probably due to the increase in the number of adsorption
negative zeta potential in their surfaces. This suggests that there are sites. In addition, the removal % decreased when initial MO con-
stronger interactions capable of overcoming the repulsive forces be- centration was increased. They reported 99% and 90% removal with
tween the dye and the adsorbents. Moreover, it was observed that the adsorption capacities of 387.59 mg/g and 358.42 mg/g at pH = 3 for
increase in CTS% resulted in higher adsorption capacity of CR due to an ST-NiFeLDH (1:1) and ST-NiFeLDH (2:1), respectively. The adsorption
increase in the zeta potential, which is indicative of the role played in kinetics followed a pseudo-second order model. The irregular and
the adsorption process by electrostatic interactions. Similar studies rough surface of ST-NiFe-LDH composites, mentioned in Section 3, re-
have reported that high CTS% significantly increases the adsorption of sults in an increased number of adsorption sites which in turn increases
anionic dyes onto positively charged CTS-Mt (Liu et al., 2013; Wang the trapping of dye molecules. Recovery of ST-NiFeLDH with little loss
and Wang, 2007a, 2007b). of its adsorption capacity was achieved with NaOH solution. Lastly, the
As for CV, their results found good adsorption capacity of Mt. which authors compare the adsorption capacity of the prepared bionano-
can be explained by cation exchange between the dye and the re- composite with other absorbents such as active carbon, graphene, Ni-
maining Ca2+ on Mt. surface and by the electrostatic interactions be- FeLDH, MgAlLDH, cork powder and bottom ash, and conclude that ST-
tween the negatively charged surface of Mt. and the dye. The hydro- NiFeLDH exhibits optimal capacity for the adsorption of MO.
phobic interactions between HDTMA aggregates and CV might also be
responsible for dye adsorption (Zhu et al., 2017). The adsorption ca- 6.5. Metal ions
pacity of HDTMA/CTS-Mt to CV was lower than that of Mt. and
HDTMA-Mt, with higher CTS% resulting in lower adsorption capacity. Heavy metals are among the most toxic compounds that can enter
This can be explained by the reduced negative charge of Mt. surface the environment. They are not biodegradable and accumulate in the
(that can even be reversed to possitive) associated with higher CTS% environment producing toxic effects even at low concentrations
and lower Ca2+ remaining on HDTMA/CTS-Mt. Authors also observed (Ahmad and Mirza (2018a). The United States Environmental Protec-
that the co-existing dyes (CR and CV) enhanced the adsorption sites for tion Agency (US EPA) has established the health advisory levels of Pb
phenol probably due to the π-π interaction between them (Zhu et al., (II), Cu(II) and Ni(II) at 0.05 mg/L, 2 mg/L and 0.015 mg/L, respec-
2014; Wei et al., 2009). tively (Ahmad and Mirza, 2018a, 2018b, 2018c). The development of
The removal of CR from aqueous solution with cationic ST-bento- novel bionanocomposites used as adsorbents will play an important role
nite was close to 100% compared to 10% removal of pristine bentonite in the protection against environmental pollution by heavy metals.
(Koriche et al., 2013). Ahmad and Mirza. (2018a) evaluated guar gum- Numerous studies have reported on the use of bionanocomposites for
bentonite bionanocomposite for the removal of Pb (II) and CV dye from the adsorption of heavy metals such as Pb(II), Cd(II), Ni(II), As(V), Cr
wastewater samples and found that the adsorption of both pollutants is (III) or Cr(VI) as well as their chemically modified derivatives.
affected by pH, amount of contaminants, contact time and temperature. Recently, Mukhopadhyay et al. (2020) provided a complete review
The optimal adsorption capacity was observed at pH 5.1 and 7.6, of different polymer-clay nanocomposites and their applications in
contact time 240 min and 300 min for Pb(II) and CV, respectively. The water remediation. The authors found that some of the individual
adsorption kinetics followed the pseudo-second order model for both limitations of clays and polymers (low specificity, sensitivity to pH and
contaminants which indicates that chemisorption is a rate determining particle size as well as low wettability) are improved by nanocompo-
step and the adsorption isotherm was best fitted to a Freundlich model sites. In this respect, several authors have investigated the use of bio-
for Pb(II) and Langmuir model for CV. They also reported that the polymers in nanocomposites. For instance, Kumar et al. (2012) assessed
adsorption process was endothermic and spontaneous with increased Cr(VI) removal from industrial effluents with clay modified with CELL
randomness at solid-solution interface. In addition, guar gum-bentonite and optimized various batch parameters like pH, adsorbent loading,
adsorbent was easily regenerated (reused in five more cycles) with and adsorption time prior to conducting column experiments in was-
0.1 M HCl. This bionanocomposite was also used for the removal of tewater spiked with the metal ions. They reported a maximum ad-
heavy metals from wastewater from various industrial and medical sorption capacity of 22.2 mg/g in accordance with the Langmuir iso-
processes before and after spiking with CV solution. Vinod and therm model. The absorption kinetics followed the pseudo-second order
Anirudhan (2003) used humic acid immobilized on modified Mt. by model. The adsorbent could be reused with quantitative recovery for 10
exchange with polymeric hydroxy cations as adsorbent of the dyes CV, adsorption-desorption cycles with sodium hydroxide. The column stu-
methylene blue (MB) and rhodamine B from aqueous solution and dies were further extended to treat a tannery leather effluent rich in
found that the adsorption capacity was two-fold higher for the modified basic chromium sulfate. Their results showed the high capacity of the
Mt. than for the pristine clay. The adsorption % was concentration nanocomposite for the removal of Cr(III) (> 95%) from the diluted
dependent, decreasing with an increase in dye concentration. The ad- effluent sample, with negligible anionic and cationic interferences.
sorption process was pH dependent: maximum adsorption capacities When compared with other adsorbents such as HDMA-Mt, CTS-Mt or
were found at pH 5.0–7.0. Dyes adsorption was best described by the acid activate kaolinite, the prepared bionanocomposite showed a rea-
Freundlich model. Finally, the authors conducted competitive adsorp- sonably good adsorption capacity for Cr(III). Pandey and Mishra (2011)
tion studies to determine the adsorption behavior of each dye from used solution blending for the development of a CTS-Cloisite 10A bio-
binary mixtures. nanocomposite for the removal of Cr(VI) from aqueous solution. CTS-
Auta and Hamed, 2014prepared and assessed a CTS-modified ball Cloisite 10A takes advantage of the hydrophilicity of the polycationic
clay (MBC) composite that was highly effective for batch and fixed-bed polysaccharide and adsorption capacity of inorganic polyanion. Al-
adsorption of MB. Ball clay has over 70% of kaolin and small amounts though the adsorption occurs in a wide pH range, pH 3 was the optimal
of quartz and other impurities. The adsorption of MB on CTS-MBC in- pH value. Based on the Langmuir model, the maximum adsorption
creased with dye concentration and pH 4.0–12.0. The adsorption pro- capacity was 357.14 mg/g. The absorption kinetics followed a pseudo-
cess was spontaneous and endothermic. Sulphate anions had greater second order model. CTS-Mt also showed efficient uptake of As(V) and
inhibition effect than those of sodium chloride and bicarbonate on CTS- Cr(VI) ions, with adsorption capacities of 8.9 mg/g (pH 4–5) and
MBC. Lastly, the authors reported that after five regeneration cycles the 9.3 mg/g (pH 3.5), respectively (An and Dultz, 2008). Adsorption iso-
adsorption performance was above 50%. therms were fitted to the Langmuir and Dubinin-Radushkevich equa-
Zubair et al. (2018) used ST-NiFeLDH bionanocomposite (1:1 and tions. The adsorbed ionic species were pH dependent, as did the degree

15
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

of protonation of the amine groups, and this played a decisive role in ~12 mg/g for Se(VI) which are higher than the capacities obtained with
the amount of anions adsorbed. By increasing the concentration of the CTS embedded with other nanocrystalline metal oxide fillers. Alcântara
competitive anion, Cl-, in solution, the amount of Cr(VI) and As(V) et al. (2014b) investigated the properties of a recently discovered
adsorbed remained almost constant, whereas sulphate anion had a polysaccharide called sacran, extracted from fresh water algae, which
more pronounced competitive effect. CTS-vermiculite bionanocompo- can modify sepiolite fibers to form a biosorbent used for the uptake of
site foams were developed by Padilla-Ortega et al. (2016) for the re- lanthanide ions from aqueous solutions. The authors reported that the
moval of Cd(II). The synthesis of vermiculite-based nanocomposites is orientation of the sacran chains as liquid crystals can be controlled by
normally complicated because the high charge density per unit cell in modifying sacran % in the bionanocomposite and by the preparation
its layers prevents high molecular weight polymers to intercalate be- method. This liquid crystal arrangement seems to be responsible for the
tween the vermiculite layers. In order to improve CTS intercalation, the higher adsorption capacity and selectivity for Nd(III) ions compared to
authors used ultrasound which caused exfoliation of the nanocompo- Ce(III), Eu(III) and Gd(III) lanthanides. Bionanocomposites prepared
sites owing to the high degree of intercalation achieved. The adsorption using sacran solutions (> 0.5%, w/v) with sepiolite (43%, w/w) showed
capacity of CTS-vermiculite and HDTMA-vermiculite-CTS was three a maximum adsorption of 13.4 mmol/g for Nd(III). Lastly, the authors
times higher than the absorption capacity of the pristine vermiculite suggest that the preservation of the self-orientation of sacran chains in
and it was increased with incrased pH, which promotes chelation me- the liquid crystalline state after assembling with the sepiolite fibers is
chanisms. Electrostatic interactions, cation exchange and chelation responsible for the enhanced adsorption.
were the primary mechanism. Azzam et al. (2016) used solution mixing
for the development of a nanocomposite based on CTS and gold and 7. Conclusion and future potential
silver nanoparticles (NPs) for the removal of Cu(II) ions from aqueous
solution. Their results revealed higher Cu(II) adsorption using CTS- Nanoscale assembly of biopolymers on clay minerals for bionano-
AgNPs-clay and CTS-AuNPs-clay than using CTS-clay composite, which composite formation has allowed the development of ecofriendly
is indicative of the role played by the metal nanoparticles in the im- functionalized adsorbents that are considered a promising approach in
provement of adsorption. The maximum adsorption capacity of CTS- water remediation. In the last decade, different biopolymer-clay na-
AgNPs-clay composite was 181.5 mg/g. Lastly, the authors reported nocomposites having similar to or higher adsorption capacity than
that contact time, Cu(II) initial concentration, adsorbent dose or solu- commercial adsorbents have been developed. The clays can be modified
tion pH did not have an impact on Cu(II) removal. to enhance their properties and dispersion within the polymer matrix.
In three micro-scale experiments Anirudhan and Suchithra (2010a, The most commonly used biopolymers are the polysaccaharides CTS,
2010b) used a humic acid-immobilized amine-modified poly- CELL, ST, ALG and CD. The resulting properties will be determined by
acrylamide–bentonite composite for the removal of heavy metals from the modification method and the agent used for clay functіonalіzation.
simulated wastewater. The experiments were conducted for de- In this respect, solution blending seems to adequately enhance the
termining the removal capacity for specific heavy metals from three dispersіon of the clay within the matrіx.
types of synthetic wastewater: (i) Cu(II) removal from wastewater The control of the interactіons between biopolymer and clay will
coming from surface electroplating operations; (ii) Zn(II) from in- allow the creation of smart nanomaterials with exceptional properties
dustrial wastewater; and (iii) Co(II) from water used as coolant in nu- like thermal stability and hydrophobic surface, which will widen the
clear reactors (Anirudhan and Suchithra, 2010b). Their findings number of environmental applications. Many authors have expressed
showed the high capacity of low doses of nanocomposite (10 g/L) for the necessity to further investigate the role played by contact tіme,
the removal (close to 100%) of the different metals, which suggests the іnіtіal polymer concentration, polymer/clay ratio and the nature of the
possibility of applying this modified bentonite for the removal of heavy clay in the optimization of the polymer loadіng. The main factor in-
metals from actual industrial wastewater. However, paremeters such as fluencing the optimization is the nature of the clay as the presence of
pH and initial concentration of metals should be further investigated. electrostatic attraction enhances the adsorption capacity. In addition,
Ahmad and Mirza (2018b) developed a nanocomposite using mica the presence of magnetіc nanoparticles in the polymer matrіx can also
modified with XG and n-acetyl cysteine for the removal of Pb(II), Cu(II) optimize the functionality of bionanocomposites. Thіs review opens up
and Ni(II) from aqueous solutions. Their findings showed that higher the possibility of exploring novel biopolymers that will allow the im-
temperatures increase the monolayer adsorption capacity and that the provement of nanocomposites to act as adsorbents of environmental
adsorption process is an endothermic and spontaneous reaction occur- pollutants.
ring more randomly at the solid-solution interface. The adsorption The adsorption capacity of these bionanocomposites for the removal
process followed pseudo-second order model and Langmuir adsorption of different classes of organic contaminants, emerging pollutants, and
isotherm with 530.54, 177.2 and 51.48 mg/g for Pb(II), Cu(II) and Ni inorganic heavy metals from water has been tested at laboratory scale.
(II), respectively, at pH values 4–5 and contact time between 1 and 2 h. However, the application of these nanocomposites as adsorbents in
The adsorbent was easily regenerated with HCl for all metal ions. The industrial and environmental settings is still in the research stage. Most
same authors also developed a recyclable XG-Mt for the removal of Pb of the experіments have been conducted using synthetic contaminated
(II) from synthetic and industrial wastewater. The kinetic study re- water (obtained by dissolution of pollutants in deionized water) and
vealed that the adsorption process of Pb(II) followed the pseudo-second previously established experimental conditions. However, settings with
order model, confirming chemisorption. The adsorption isotherm was actual contaminated water are complex systems in which a diverse
consistent with the Freundlich model for Pb(II). Thermodynamic stu- collection of biological and interference species and organic compo-
dies showed that absorption was driven by a physisorption process that nents can be present in large amounts. After our literature review, we
was spontaneous and endothermic with increased randomness. The have found no full-scale experiment on the use of bionanocomposites
bionanocomposite was easily regenerated for five more cycles with for environmental remediation. In order to be able to conduct full-scale
0.05 M HCl and incinerated when exhaust, which suggests that this XG- experiments, a deep understandіng of the structural stability of biona-
based nanocomposite might be an inexpensive potential adsorbent for nocomposites and their environmental toxicity is paramount to de-
actual wastewater treatments. monstrate their practical applications.
Li et al. (2018) used a CTS-LDH bionanocomposite for the removal In addition, few studies have addressed the regeneration capacity of
of selenate and selenite oxoanion and reported higher removal of Se(VI) biopolymer-clay-based adsorbents, which allows the reduction of the
with CTS-LDH (> 80%) than with CTS (57%) due to the increased amount of waste generated and the removal of bulk amounts of con-
adsorption provided by LDH incorporation. The maximum adsorption taminants. In this regard, these affordable clay mіnerals can be suc-
capacities from Langmuir isotherms were 17 mg/g for Se(IV) and cessfully used in regeneratіon studies. Financial analyses would also

16
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

help to understand the feasibility and durability of these adsorbents at S., 2019. Adsorption of organic pollutants by natural modified clays: a comprehen-
ground level as there is little financial information on the sustainable sive review. Sep. Purif. Technol. 228, 115719.
Azhar, F.F., Olad, A., Mirmohseni, A., 2014. Development of novel hybrid nanocompo-
use of bionanocomposites for water remediation. In this respect, the sites based on natural biodegradable polymer–montmorillonite/ polyaniline: pre-
utilization of naturally occuring clay mіnerals will help to reduce the paration and characterization. Polym. Bull. 71, 1591–1610.
production costs. Azzam, E.M.S., Eshaq, G., Rabie, A.M., Bakr, A.A., Abd-Elaal, A.A., El Metwally, A.E.,
Tawfik, S.M., 2016. Preparation and characterization of chitosan-clay nanocompo-
sites for the removal of Cu(II) from aqueous solution. Int. J. Biol. Macromol. 89,
Declaration of Competing Interest 507–517.
Barreca, S., Orecchio, S., Pace, A., 2014. The effect of montmorillonite clay in alginate gel
beads for polychlorinated biphenyl adsorption: Isothermal and kinetic studies. Appl.
The authors declare that they have no known competing financial Clay Sci. 99, 220–228.
interests or personal relationships that could have appeared to influ- Bee, S.-L., Abdullah, M., Mamat, M., Bee, S.-T., Sin, L.T., Hui, D., Rahmat, A., 2017.
Characterization of silylated modified clay nanoparticles and its functionality in
ence the work reported in this paper.
PMMA. Compos. Part B Eng. 110, 83–95.
Besun, N., Peker, S., Kokturk, U., Yilmaz, H., 1997. Structure of starch-bentonite gels.
Acknowledgement Colloid Polym. Sci. 275, 378–389.
Bhat, A.H., Rehman, W.U., Khan, I.U., Khan, I., Ahmad, S., Ayoub, M., Usmani, M.A.,
2018. Nanocomposite membrane for environmental remediation. In: Jawaid, M.,
This work was supported by the Spanish Ministry of Economy, Khan, M.M. (Eds.), Polymer-based Nanocomposites for Energy and Environmental
Industry and Competitiveness (Project No. CTM2017-82778-R). Applications. Elsevier, U.K, pp. 407–440 chapter 15.
Bhawani, S.A., Bhat, A.H., Ahmad, F.B., Ibrahim, M.N.M., 2018. Green polymer nano-
composites and their environmetal applications. In: Jawaid, M., Khan, M.M. (Eds.),
References Polymer-based Nanocomposites for Energy and Environmental Applications. Elsevier,
U.K, pp. 617–633 Chapter 23.
Abbasian, M., Pakzad, M., Nazari, K., 2017. Synthesis of cellulose-graft-poly- Bleiman, N., Mishael, Y.G., 2010. Selenium removal from drinking water by adsorption to
chloromethylstyrene-graft-polyacrylonitrile terpolymer/organoclay bionanocompo- chitosan–clay composites and oxides: batch and columns tests. J. Hazard. Mater. 183,
site by metal catalyzed living radical polymerization and solvent blending method. 590–595.
Polym.-Plast. Technol. Eng. 56, 857–865. Buruga, K., Song, H., Shang, J., Bolan, N., Kalathi, J.T., Kim, K., 2019. A review on
Ahmad, R., Mirza, A., 2017. Adsorption of Pb(II) and Cu(II) by Alginate-Au Mica bio- functional polymer-clay based nanocomposite membranes for treatment of water. J.
nanocomposite: Kinetic, isotherm and thermodynamic studies. Process. Saf. Environ. Hazard. Mater. 379, 120584.
109, 1–10. Celis, R., Adelino, M.A., Hermosín, M.C., Cornejo, J., 2012. Montmorillonite-chitosan
Ahmad, R., Mirza, A., 2018a. Synthesis of Guar gum/bentonite a novel bionanocompo- bionanocomposites as adsorbents of the herbicide clopyralid in aqueous solution and
site: Isotherms, kinetics and thermodynamic studies for the removal of Pb (II) and soil/water suspensions. J. Hazard. Mater. 209-210, 67–76.
Cristal violet dye. J. Mol. Liq. 249, 805–814. Charles, E., Lin, D., 1937. Introduction. In: Charles, E., Lin, D. (Eds.), The Chemistry of
Ahmad, R., Mirza, A., 2018b. Application of Xanthan gum/ n-acetyl cysteine modified Clay Minerals. vol.15. Elsevier, U.K, pp. 1–4.
mica bionanocomposite as an adsorbent for the removal of toxic heavy metals. Chen, L., Wu, P., Chen, M., Lai, X., Ahmed, Z., Zhu, N., Dang, Z., Bi, Y., Liu, T., 2018.
Groundwater Sustain. Develop. 7, 101–108. Preparation and characterization of the eco-friendly chitosan/vermiculite bio-
Ahmad, R., Mirza, A., 2018c. Adsorptive removal of heavy metals and anionic dye from composite with excellent removal capacity for cadmium and lead. Appl. Clay Sci.
aqueous solution using novel Xanthan gum-Glutathione/Zeolite bionanocomposite. 159, 74–82.
Groundw. Sustain. Dev. 7, 305–312. Chiang, M.-F., Wu, T.-M., 2010. Synthesis and characterization of biodegradable poly(L-
Ahmed, J., Varshney, S.K., 2010. Polylactides-chemistry, properties and green packaging lactide)/layered double hydroxide nanocomposites. Compos. Sci. Technol. 70,
technology: a review. Int. J. Food Prop. 14, 37–58. 110–115.
Alba, M.D., Cota, A., Osuna, F.J., Pavón, E., Perdigón, A.C., Raffin, F., 2019. Chivrac, F., Pollet, E., Schmutz, M., Averous, L., 2008. New approach to elaborate ex-
Bionanocomposites based on chitosan intercalation in designed swelling high- foliated starch-based nanobiocomposites. Biomacromolecules 9, 896–900.
charged micas. Sci. Rep. 9, 10265. Chivrac, F., Pollet, E., Avérous, L., 2009. Progress in nano-biocomposites based on
Alcântara, A.C.S., 2013. PhD dissertation Autonomous University of Madrid, Madrid. polysaccharides and nanoclays. Mater. Sci. Eng. 67, 1–17.
Alcântara, A.C.S., Darder, M., 2018. Building up functional bionanocomposites from the Dai, J.C., Huang, J.T., 1999. Surface modification of clays and clay–rubber composite.
Assembly of Clays and Biopolymers. Chem. Rec. 18, 696–712. Appl. Clay Sci. 15, 51–65.
Alcântara, A.C.S., Darder, M., Aranda, P., Ruiz-Hitzky, E., 2012. Zein–fibrous clays bio- Dang, K.M., Yoksan, R., Pollet, E., Avérous, L., 2020. Morphology and properties of
hybrid materials. Eur. J. Inorg. Chem. 1, 5216–5224. thermoplastic starch blended with biodegradable polyester and filled with halloysite
Alcântara, A.C.S., Darder, M., Aranda, P., Ruiz-Hitzky, E., 2014a. Polysaccharide-fibrous nanoclay. Carbohydr. Polym. 242, 116392.
clay bionanocomposites. Appl. Clay Sci. 96, 2–8. Darder, M., Colilla, M., Ruiz-Hitzky, E., 2003. Biopolymer-clay nanocomposites inter-
Alcântara, A.C.S., Darder, M., Aranda, P., Tateyama, S., Okajima, M.K., Kaneko, T., calated in montmorillonite based on chitosan. Chem. Mater. 15, 3774–3780.
Ogawa, M., Ruiz-Hitzky, E., 2014b. Clay-bionanocomposites with sacran mega- Darder, M., Colilla, M., Ruiz-Hitzky, E., 2005a. Chitosan-clay nanocomposites: applica-
molecules for the selective uptake of neodymium. J. Mater. Chem. A 2, 1391–1399. tion as electrochemical sensors. Appl. Clay Sci. 28, 199–208.
Alcântara, A.C.S., Darder, M., Aranda, P., Ayral, A., Ruiz-Hitzky, E., 2016a. Darder, M., López-Blanco, M., Aranda, P., Leroux, F., Ruiz-Hitzky, E., 2005b.
Bionanocomposites based on polysaccharides and fibrous clays for packaging appli- Bionanocomposites based on layered double hydroxides. Chem. Mater. 17,
cations. J. Appl. Polym. Sci. 133, 42362. 1969–1977.
Alcântara, A.C.S., Darder, M., Aranda, P., Ruiz-Hitzky, E., 2016b. Effective intercalation Darder, M., Ruiz, A.I., Aranda, P., Van Damme, H., Ruiz-Hitzky, E., 2006. Bio-
of zein into Na-montmorillonite: role of the protein components and use of the de- Nanohybrids based on layered inorganic solids: gelatin nanocomposites. Curr.
veloped biointerfaces. Beilstein J. Nanotechnol. 7, 1772–1782. Nanosci. 2, 231–241.
Alexandre, M., Dubois, P., 2000. Polymer-layered silicate nanocomposites: preparation, Darder, M., Aranda, P., Ruiz-Hitzky, E., 2007. Bionanocomposites: a new concept of
properties and uses of a new class of materials. Mater. Sci. Eng. 28, 1–63. ecological, bioinspired, and functional hybrid materials. Adv. Mater. 19, 1309–1319.
An, J.H., Dultz, S., 2008. Adsorption of Cr(VI) and As(V) on chitosan-montmorillonite: Darder, M., Santos Matosa, C.R., Aranda, P., Figueredo Gouveia, R., Ruiz-Hitzky, E., 2017.
Selectivity and pH dependence. Clay Clay Miner. 56, 549–557. Bionanocomposite foams based on the assembly of starch and alginate with sepiolite
Anirudhan, T.S., Suchithra, P.S., 2010a. Humic acid-immobilized polymer/bentonite fibrous clay. Carbohydr. Pol. 157, 1933–1939.
composite as an adsorbent for the removal of copper(II) ions from aqueous solutions De Carvalho, A.J.F., Curvelo, A.A.S., Agnellib, J.A.M., 2001. A first insight on composites
and electroplating industry wastewater. J. Ind. Eng. Chem. 16, 130–139. of thermo-plastic starch and kaolin. Carbohydr. Polym. 45 (2), 189–194.
Anirudhan, T.S., Suchithra, P.S., 2010b. Heavy metals uptake from aqueous solutions and Deng, Y., Wang, L., Hu, X., Liu, B., Wei, Z., Yang, S., Sun, C., 2012. Highly efficient
industrial wastewaters by humic acid-immobilized polymer/bentonite composite: removal of tannic acid from aqueous solution by chitosan-coated attapulgite. Chem.
kinetics and equilibrium modeling. Chem. Eng. J. 156, 146–156. Eng. J. 181, 300–306.
Arora, B., Bhatia, R., Attri, P., 2018. Bionanocomposites: Green materials for a sustainable Dinari, M., Dadkhah, F., 2020. Swift reduction of 4-nitrophenol by easy recoverable
future. In: Hussain, C.M., Mishra, A.K. (Eds.), New Polymer Nanocomposites for magnetite-Ag/layered double hydroxide/starch bionanocomposite. Carbohydr.
Environmental Remediation. Elsevier, UK, pp. 699–712 chapter 28. Polym. 228, 115392.
Auta, M., Hamed, B.H., 2014. Chitosan-clay composite as highly effective and low-cost Dujardin, E., Mann, S., 2002. Bioinspired materials chemistry. Adv. Mater. 14, 775–788.
adsorbent for batch and fixed-bed adsorption of methylene blue. Chem. Eng. J. 237, Etcheverry, M., Cappa, V., Trelles, J., Zanini, G., 2017. Montmorillonite-alginate beads:
352–361. natural mineral and biopolymers based sorbent of paraquat herbicides. J. Environ.
Avérous, L., 2013. Synthesis, properties, environmental and biomedical applications of Chem. Eng. 5, 5868–5875.
polylactic acid. In: Ebnesajjad, S. (Ed.), Handbook of Biopolymers and Biodegradable Futalan, C.M., Kan, C., Dalida, M.L., Hsien, K., Pascua, C., Wan, M., 2011. Comparative
Plastics. Elsevier, U.K., pp. 171–188. and competitive adsorption of copper, lead, and nickel using chitosan immobilized on
Avérous, L., Pollet, E., 2012. Environmental Silicate Nano-Biocomposites, 1st ed. Springer- bentonite. Carbohydr. Polym. 83, 528–536.
Verlag, London. Fischer, H.R., Gielgens, L.H., Koster, T.P.M., 1999. Nanocomposites from polymers and
Avérous, L., Pollet, E., 2014. Nanobiocomposites based on plasticized starch. In: Halley, layered minerals. Acta Polym 50, 122–126.
P.J., Avérous, L. (Eds.), Starch Polymers. Elsevier, U.K, pp. 211–239 Chapter 8. Guo, F., Aryana, S., Han, Y., Jiao, Y., 2018. A review of the synthesis and applications of
Awad, A.M., Shaikh, S.M.R., Jalab, R., Gulied, M.H., Nasser, M.S., Benamor, A., Adham, polymer–nanoclay composites. Appl. Sci. 8, 1696.

17
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

Herrera, J.I., Olmo, N., Turnay, J., Sicilia, A., Lizarbe, M.A., 1995. Implantation of se- 38, 1543–1589.
piolite-collagen complexes in surgically created rat calvaria defects. Biomaterials 16, Padilla-Ortega, E., Darder, M., Aranda, P., Gouveia, R.F., Leyva-Ramos, R., Ruiz-Hitzky,
625–631. E., 2016. Ultrasound assisted preparation of chitosan-vermiculite bionanocomposite
Hosseinzadeh, H., Zoroufi, S., Mahdavinia, G.R., 2015. Study on adsorption of cationic foams for cadmium uptake. Appl. Clay Sci. 130, 40–49.
dye on novel kappa-carrageenan/poly (vinyl alcohol)/montmorillonite nanocompo- Pandey, S., Mishra, S.B., 2011. Organic–inorganic hybrid of chitosan/organoclay biona-
site hydrogels. Polym. Bull. 72, 1339–1363. nocomposites for hexavalent chromium uptake. J. Colloid Interface Sci. 361,
Jin, Y.H., Park, H.J., Im, S.S., Kwak, S.Y., Kwak, S., 2002. Polyethylene/Clay nano- 509–520.
composite by in-situ exfoliation of montmorillonite during ziegler-natta poly- Park, J.S., Park, J.W., Ruckenstein, E., 2001. Thermal and dynamic mechanical analysis of
merization of ethylene. Macromol. Rapid Commun. 23, 135–140. PVA/MC blend hydrogels. Polymer 42, 4271–4280.
Jlassi, K., Krupa, I., Chehimi, M., 2017. Overview: Clay preparation, properties, mod- Park, H.-M., Li, X., Jin, C.-Z., Park, C.-Y., Cho, W.-J., Ha, C.-S., 2002. Preparation and
ification. In: Jlassi, K., Chehimi, M.M., Thomas, S. (Eds.), Clay-Polymer properties of bio-degradable thermoplastic starch/clay hybrids. Macromol. Mater.
Nanocomposites. Elsevier, U.K, pp. 1–28 Chaper 1. Eng. 287 (8), 553–558.
Kampeerapappun, P., Ahtong, D., Pentrakoon, D., Srikulkit, K., 2007. Preparation of Park, H.M., Mohanty, A.K., Drzal, L.T., Lee, E., Mielewski, D.F., Misra, M., 2006. Effect of
cassava starch/montmorillonite composite film. Carbohydr. Polym. 67, 155–163. sequential mixing and compounding conditions on cellulose acetate/layered silicate
Koriche, Y., Darder, M., Aranda, P., Semsari, S., Ruiz-Hitzky, E., 2013. Efficient and nanocomposites. J. Polym. Environ. 14, 27–35.
ecological removal of anionic pollutants by cationic starch-clay bionanocomposites. Park, Y., Ayoko, G.A., Frost, R.L., 2011. Application of organoclays for the adsorption of
Sci. Adv. Mater. 5, 994–1005. recalcitrant organic molecules from aqueous media. J. Colloid Interface Sci. 354,
Koriche, Y., Darder, M., Aranda, P., Semsari, S., Ruiz-Hitzky, E., 2014. Bionanocomposites 292–305 Academic Press.
based on layered silicates and cationic starch as eco-friendly adsorbents for hex- Paul, M.-A., Delcourt, C., Alexandre, M., Degée, P., Monteverde, F., Rulmont, A., Dubois,
avalent chromium removal. Dalton Trans. 43, 10512–10520. P., 2005. (Plasticized) polylactide/(organo-) clay nanocomposites by in situ inter-
Kotal, M., Bhowmick, A.K., 2015. Polymer nanocomposites from modified clays: recent calative polymerization. Macromol. Chem. Phys. 206, 484–498.
advances and challenges. Prog. Polym. Sci. 51, 127–187. Pinto, M.E., Gonçalves, R.G.L., Menezes dos Santos, R.M., Araújo, E.A., Perotti, G.F.,
Kumar, A.S.K., Kalidhasan, S., Rajesh, V., Rajesh, N., 2012. Application of cellulose-clay Macedo, R., Bizeto, M.A., Constantino, V.R.L., Pinto, F.G., Tronto, J., 2016.
composite biosorbent toward the effective adsorption and removal of chromium from Mesoporous carbon derived from a biopolymer and a clay: Preparation, character-
industrial wastewater. Ind. Eng. Chem. Res. 51, 58–69. ization and application for an organochlorine pesticide adsorption. Microporous
Lee, S.M., Tiwari, D., 2012. Organo and inorgano-organo-modified clays in the re- Mesoporous Mater. 225, 342–354.
mediation of aqueous solutions: an overview. Appl. Clay Sci. 59–60, 84–102. Qiu, H., Yu, J., Zhu, J., 2005. Polyacrylate/(chitosan modified montmorillonite) nano-
Leroux, F., Besse, J.-P., 2001. Polymer interleaved layered double hydroxide: a new composite: water absorption and photostability. Polym. Polym. Compos. 13,
emerging class of nanocomposites. Chem. Mater. 13, 3507–3515. 167–172.
Leroux, F., Gachon, J., Besse, J.-P., 2004. Biopolymer immobilization during the crys- Ray, S.S., 2012. Polylactide-based bionanocomposites: a promising class of hybrid ma-
talline growth of layered double hydroxide. J. Solid State Chem. 177, 245–250. terials. Acc. Chem. Res. 45, 1710–1720.
Li, M., Dopilka, A., Kraetz, A.N., Jing, H., Chan, C.K., 2018. Layered double hydroxide/ Ray, S.S., Okamoto, M., 2003. Polymer/layered silicate nanocomposites: a review from
chitosan nanocomposite beads as sorbents for selenium oxoanions. Ind. Eng. Chem. preparation to processing. Prog. Polym. Sci. 28, 1539–1641.
Res. 57, 4978–4987. Ren, J., Dang, K.M., Pollet, E., Avérous, L., 2018. Preparation and characterization of
Liu, H., Nakagawa, K., Chaudhary, D., Asakuma, Y., Tadé, M., 2011. Freeze-dried mac- thermoplastic potato starch/halloysite nano-biocomposites: effect of plasticizer
roporous foam prepared from chitosan/xanthan gum/montmorillonite nanocompo- nature and nanoclay content. Polymers 10, 808.
sites. Chem. Eng. Res. Des. 89, 2356–2364. Rhim, J.W., 2007. Potential use of biopolymer-based nanocomposite films in food
Liu, B., Wang, D., Yu, G., Meng, X., 2013. Adsorption of heavy metal ions, dyes and packaging applications. Food Sci. Biotechnol. 16, 691–709.
proteins by chitosan composites and derivatives – a review. J. Ocean Univ. China 12, Ruiz, A.I., Darder, M., Aranda, P., Jiménez, R., Van Damme, H., Ruiz-Hitzky, E., 2006.
500–508. Bio-nanocomposites by assembling of gelatin and layered perovskite mixed oxides. J.
Liu, Q., Yang, B., Zhang, L., Huang, R., 2015. Adsorption of an anionic azo dye by cross- Nanosci. Nanotechnol. 6, 1602–1610.
linked chitosan/bentonite composite. Int. J. Biol. Macromol. 72, 1129–1135. Ruiz-Hitzky, E., 2004. In: Sanchez, C., Gómez-Romero, P. (Eds.), Functional Hybrid
Mahdavinia, G.R., Asgari, A., 2013. Synthesis of kappa-carrageenan-g-poly(acrylamide)/ Materials. Wiley VCH, Weinheim.
sepiolite nanocomposite hydrogels and adsorption of cationic dye. Polym. Bull. 70, Ruiz-Hitzky, E., Van Meerbeek, A., 2006. Clay mineral- and organoclay–polymer nano-
2451–2470. composite. In: Bergaya, F., Theng, B.K.G., Lagaly, G. (Eds.), Handbook of Clay Science
Mahdavinia, G.R., Hasanpour, S., Behrouzi, L., Sheykhloie, H., 2016. Study on adsorption Development in Clay Science. Elsevier, Amsterdam, pp. 583–621.
of Cu(II) on magnetic starch-g-polyamidoxime/montmorillonite/Fe3O4 nano- Ruiz-Hitzky, E., Aranda, P., Serratosa, J.M., 2004. In: Aucherbach, S., Carrado, K.A.,
composites as novel chelating ligands. Starch/Stärke 68, 188–199. Dutta, P. (Eds.), Handbook of Layered Materials. Marcel Dekker, New York.
Martín, J., Orta, M.-M., Medina-Carrasco, S., Santos, J.L., Aparicio, I., Alonso, E., 2019a. Ruiz-Hitzky, E., Darder, M., Aranda, P., 2005. Functional biopolymer nanocomposites
Evaluation of a modified mica and montmorillonite for the adsorption of ibuprofen based on layered solids. J. Mater. Chem. 15, 3650–3662.
from aqueous media. Appl. Clay Sci. 171, 29–37. Ruiz-Hitzky, E., Aranda, P., Darder, M., 2008a. Bionanocomposites. In: Hoboken, N.J.
Martín, J., Santos, J.L., Malvar, J.L., Aparicio, I., Alonso, A., 2019b. Determination of (Ed.), Kirk- Othmer Encyclopedia of Chemical Technology. John Wiley & Sons, pp.
bisphenol A, its chlorinated derivatives and structural analogues in vegetables by 1–28.
focussed ultrasound solid-liquid extraction and GC–MS/MS. Environ. Chem (IN Ruiz-Hitzky, E., Darder, M., Aranda, P., 2008b. In: Ruiz-Hitzky, E., Ariga, K., Lvov, Y.
PRESS). (Eds.), Bio-Inorganic hybrid materials: Strategies, syntheses, characterization and
Mayer, G., 2006. New classes of tough composite materials-Lessons from natural rigid applications. Wiley-VCH, Weinheim, Germany.
biological systems. Mater. Sci. Eng. 26, 1261–1268. Seema, M.D., 2013. Clay–polymer nanocomposites as a novel drug carrier: synthesis,
Monvisade, P., Siriphannon, P., 2009. Chitosan intercalated montmorillonite: prepara- characterization and controlled release study of Propranolol Hydrochloride. Appl.
tion, characterization and cationic dye adsorption. Appl. Clay Sci. 42, 427–431. Clay Sci. 80, 85–92.
Mukhopadhyay, R., Bhaduri, D., Sarkar, B., Rusmin, R., Hou, D., Khanam, R., Sarkar, S., Sethy, T.R., Sahoo, P.K., 2019. Highly toxic Cr (VI) adsorption by (chitosan-g PMMA)/
Biswas, J.K., Vithanage, M., Bratnager, A., Ok, Y.S., 2020. Clay–polymer nano- sílica bionanocomposite prepared via emulsifier-free emulsion polymerisation. Int. J.
composites: progress and challenges for use in sustainable water treatment. J. Biol. Macromol. 122, 1184–1190.
Hazard. Mater. 383, 1–17. Shabtai, I.A., Mishael, Y.G., 2018. Polycyclodextrin-clay composites: regenerable dual-
Mura, P., Maestrellia, F., Aguzzi, C., Viseras, C., 2016. Hybrid systems based on “drug – in site sorbents for bisphenol A removal from treated wastewater. ACS Appl. Mater.
cyclodextrin – in nanoclays” for improving oxaprozin dissolution properties. Int. J. Interfaces 10, 27088–27097.
Pharm. 509, 8–15. Sharma, S., Komarneni, S., 2009. Synthesis and characterization of synthetic mica-bio-
Murray, H.H., 1924a. Structure and composition of the clay minerals and their physical nanocomposites. Appl. Clay Sci. 42, 553–558.
and chemical propierties. In: Murray, H.H. (Ed.), Applied Clay Mineralogy. vol. 2. Shawky, H.A., 2011. Improvement of water quality using alginate/montmorillonite
Elsevier, U.K, pp. 24–25. composite beads. J. Appl. Polym. Sci. 119, 2371–2378.
Murray, H.H., 1924b. Bentonite applications. In: Murray, H.H. (Ed.), Applied Clay Shen, Z., Simon, G.P., Cheng, Y.-B., 2002. Comparison of solution intercalation and melt
Mineralogy. Elsevier, U.K., pp. 124–125. intercalation of polymer-clay nanocomposites. Polymer 43, 4251–4260.
Murray, H.H., 1991. Overview-clay mineral applications. Appl. Clay Sci. 5, 379–395. Shukla, R., Cheryan, M., 2001. Zein: the industrial protein from corn. Ind. Crop. Prod. 13,
Neppalli, R., Causin, V., Marega, C., Modesti, M., Adhikari, R., Scholtyssek, S., Ray, S.S., 171–192.
Marigo, A., 2014. The effect of different clays on the structure, morphology and Siqueira, G., Bras, J., Dufresne, A., 2010. Cellulosic bionanocomposites: a review of
degradation behavior of poly (lactic acid). Appl. Clay Sci. 87, 278–284. preparation, properties and applications. Polymers 2, 728–765.
Ngah, W.S.W., Md Ariff, N.M., Hashim, A., Hanafiah, M.A.K.M., 2010a. Hanafiah, pre- Soetaredjo, F.E., Ismadji, S., Foe, K., Yi-Hsu, J., 2018. Recent advances in the application
paration, characterization, and environmental application of crosslinked chitosan- of polymer-based nanocomposites for removal of hazardous substances from water
coated bentonite for tartrazine adsorption from aqueous solutions. Water Air Soil and wastewater. New Polim. Nonocompos. Environ. Rem. 21, 499–540.
Pollut. 206, 225–236. Sophia, A.C., Lima, E.C., 2018. Removal of emerging contaminants from the environment
Ngah, W.S.W., Md Ariff, N.M., Hashim, A., Hanafiah, M.A.K.M., 2010b. Malachite green by adsorption. Ecotoxicol. Environ. Saf. 150, 1–17.
adsorption onto chitosan coated bentonite beads: isotherms, kinetics and mechanism. Swain, S.K., Barik, S., Pradhan, G.C., Behera, L., 2018. Delamination of Mg-Al layered
Clean Soil Air Water 38, 394–400. double hydroxide on starch: change in structural and thermal properties. Polym.-
Ogawa, M., Kuroda, K., 1997. Preparation of inorganic-organic nanocomposites through Plast. Technol. Eng. 57, 1585–1591.
intercalation of organoammonium ions into layered silicates. Bull. Chem. Soc. Jpn. Talibudeen, O., 1950. Interlamellar adsorption of protein monolayers on pure mon-
70, 2593–2618. tmorillonoid clays. Nature 166, 236.
Ojijo, V., Ray, S.S., 2013. Processing strategies in bionanocomposites. Prog. Polym. Sci. Tarpani, R.R.Z., Azapagic, A., 2018. Life cycle environmental impacts of advanced

18
M.d.M. Orta, et al. Applied Clay Science 198 (2020) 105838

wastewater treatment techniques for removal of pharmaceuticals and personal care Wei, J., Zhu, R., Zhu, J., Ge, F., Yuan, P., He, H., Ming, C., 2009. Simultaneous sorption of
products (PPCPs). J. Environ. Manag. 215, 258–272. crystal violet and 2-naphthol to bentonite with different CECs. J. Hazard. Mater. 166,
Theng, B.K.G., 1974. The Chemistry of Clay-Organic Reactions, 1st ed. Wiley, New York. 195–199.
Theng, B.K.G., 1979. Formation and Properties of Clay–Polymer Complexes, 1st ed. Weiss, A., 1969. In: Eglinton, G., Murphy, M.T.J. (Eds.), Organic geochemistry. Springer
Elsevier, New York. Verlag, Berlin, pp. 737–781.
Tran, V.S., Ngo, H.H., Guo, W., Zhang, J., Liang, S., Ton-That, C., Zhang, X., 2015. Typical Wicklein, B., Darder, M., Aranda, P., Ruiz-Hitzky, E., 2010. Bio-organoclays based on
low cost biosorbents for adsorptive removal of specific organic pollutants from water. phospholipids as immobilization hosts for biological species. Langmuir 26,
Bioresour. Technol. 182, 353–363. 5217–5225.
Ucankus, G., Ercan, M., Uzunoglu, D., Culha, M., 2018. Methods for preparation of na- Xie, F., Pollet, E., Halley, P.J., Avérous, L., 2013. Starch-based nano-biocomposites. Progr.
nocomposites in environmental remediation. New Polim. Nanocompos. Environ. Pol. Sci. 38, 1590–1628.
Rem. 1, 1–28. Xu, Y., Zhou, J., Hanna, M.A., 2005. Melt-intercalated starch acetate nanocomposite
Ünlü, C.H., Günister, E., Atici, O., 2012. Effect of acidity on xylan-montmorillonite bio- foams as affected by type of Organoclay. Cereal Chem. 82 (1), 105–110.
nanocomposites. Mater. Chem. Phys. 136, 653–660. Yadollahi, M., Namazi, H., Barkhordari, S., 2014. Preparation and properties of carbox-
Vaia, R.A., Teukolsky, R.K., Giannelis, E.P., 1994. Interlayer structure and molecular ymethyl cellulose/layered double hydroxide bionanocomposite films. Carbohydr.
environment of alkylammonium layered silicates. Chem. Mater. 6, 1017–1022. Polym. 108, 83–90.
Valapa, R.B., Loganathan, S., Pugazhenthi, G., Tomas, S., Varghese, T.O., 2017. An Yao, H., Tan, Z.H., Fang, Y.H., Yu, S.H., 2010. Artificial nacre-like bionanocomposite
overview of polimer-clay nanocomposites. In: Jlassi, K., Chehim, M.M., Thomas, S. films from the self-assembly of chitosan-montmorillonite hybrid building blocks.
(Eds.), Clay-Polymer Nanocomposites. Elsevier, U.K, pp. 29–81 Chapter 2. Angew. Chem. Int. Ed. 49, 10127–10131.
VanderHart, D.L., Asano, A., Gilman, J.W., 2001. NMR measurements related to clay Yoshioka, M., Takabe, K., Sugiyama, J., Nishio, Y., 2006. Newly developed nano-
dispersion quality and organic-modifier stability in nylon 6/clay nanocomposites. composites from cellulose acetate/layered silicate/poly εoly y developed Synthesis
Macromolecules 34, 3819–3822. and morphological characterization. J. Wood Sci. 52, 121–127.
Vinod, V.P., Anirudhan, T.S., 2003. Adsorption behaviour of basic dyes on humic acid Zafar, R., Zia, K.M., Tabasum, S., Jabeen, F., Noreen, A., Zuber, M., 2016. Polysaccharide
immobilized pillared clay. Water Air Soil Pollut. 150, 193–217. based bionanocomposites, properties and applications: a review. Int. J. Biol.
Wang, L., Wang, A., 2007a. Removal of Congo red from aqueous solution using a chit- Macromol. 92, 1012–1024.
osan/organomontmorillonite nanocomposite. J. Chem. Technol. Biotechnol. 82, Zhou, Q., Xanthos, M., 2010. Effects of cationic and anionic clays on the hydrolytic de-
711–720. gradation of polylactides. Polym. Eng. Sci. 51, 320–330.
Wang, L., Wang, A., 2007b. Adsorption characteristics of Congo red onto the chitosan/ Zhu, R., Chen, Q., Liu, H., Ge, F., Zhu, L., Zhu, J., He, H., 2014. Montmorillonite as a
montmorillonite nanocomposite. J. Hazard. Mater. 147, 979–985. multifunctional adsorbent can simultaneously remove crystal violet, cetyl-
Wang, S.F., Shen, L., Tong, Y.J., Chen, L., Phang, I.Y., Lim, P.Q., Liu, T.X., 2005. trimethylammonium, and 2-naphthol from water. Appl. Clay Sci. 88–89, 33–38.
Biopolymer chitosan/montmorillonite nanocomposites: Preparation and character- Zhu, J., Tian, M., Zhang, Y., Zhang, H., Liu, J., 2015. Fabrication of a novel “loose”
ization. Polym. Degrad. Stab. 90, 123–131. nanofiltration membrane by facile blending with Chitosan–Montmorillonite na-
Wang, S.F., Shen, L., Tong, Y.J., 2006. Structure-property relationship in chitosan-based nosheets for dyes purification. Chem. Eng. J. 265, 184–193.
biopolymer/montmorillonite nanocomposites. J. Pol. Sci. Part A: Pol. Chem. 44, Zhu, L., Wang, L., Xu, Y., 2017. Chitosan and surfactant co-modified montmorillonite: a
686–696. multifunctional adsorbent for contaminant removal. Appl. Clay Sci. 146, 35–42.
Wang, L., Zhanga, J., Wang, A., 2008. Removal of methylene blue from aqueous solution Zhu, H.-Y., Jiang, R., Xiao, L., 2010. Adsorption of an anionic azo dye by chitosan/kaolin/
using chitosan-g-poly (acrylic acid)/montmorillonite superadsorbent nanocomposite. γ-Fe2O3 composites. Appl. Clay Sci. 48, 522–526.
Colloids Surf. A: Physicochem. Eng. Aspects 322, 47–53. Zubair, M., Jarraha, N., Ihsanullah, Khalid, A., Manzar, M.S., Kazeeme, T.S., Al-Harth,
Wang, L., Yu, G., Li, J., Feng, Y., Peng, Y., Zhao, X., Tang, Y., Zhang, Q., 2019. Stretchable M.A., 2018. Starch-NiFe-layered double hydroxide composites: efficient removal of
hydrophobic modified alginate double-network nanocomposite hydrogels for sus- methyl orange from aqueous phase. J. Mol. Liq. 249, 254–264.
tained release of water-insoluble pesticides. J. Clean. Prod. 226, 122–132.

19

You might also like