You are on page 1of 43

Journal Pre-proof

Surfactants: Recent advances and their applications

Samy M. Shaban, Joohoon Kang, Dong-Hwan Kim

PII: S2452-2139(20)30265-5
DOI: https://doi.org/10.1016/j.coco.2020.100537
Reference: COCO 100537

To appear in: Composites Communications

Received Date: 24 August 2020


Revised Date: 15 October 2020
Accepted Date: 16 October 2020

Please cite this article as: S.M. Shaban, J. Kang, D.-H. Kim, Surfactants: Recent advances and their
applications, Composites Communications (2020), doi: https://doi.org/10.1016/j.coco.2020.100537.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


Mini-Review

Surfactants: Recent Advances and Their Applications


Samy M. Shaban1,2,3, Joohoon Kang4,*, and Dong-Hwan Kim1,2,*
1. School of Chemical Engineering, Sungkyunkwan University (SKKU), Suwon 16419,
Republic of Korea
2. Biomedical Institute for Convergence at SKKU (BICS), Sungkyunkwan University
(SKKU), Suwon 16419, Republic of Korea
3. Petrochemicals Department, Egyptian Petroleum Research Institute, Egypt
4. School of Advanced Materials Science and Engineering, Sungkyunkwan University

of
(SKKU), Suwon 16419, Republic of Korea

ro
E-mail: dhkim1@skku.edu (D.-H. K.) and joohoon@skku.edu (J.K.)
-p
Abstract
re
Surfactants have been widely used in many industrial products such as detergents,
lP

medicines, and anti-corrosive treatments due to their unique structures consisted


na

of two different molecular parts and the broad range of selection. Surfactants
have also contributed significantly to many research fields, especially in
ur

nanotechnology. For example, the amphiphilic nature of surfactants has been


Jo

exploited to stabilize hydrophobic nanomaterials in water, affording access to a


plethora of solution-processed nanomaterial-based scalable applications.
Surfactants are also a key additive in the production of well-controlled
nanoparticles. Surfactant-assisted metallic nanoparticle production has enabled
the realization of colorimetric sensors, which are of massive interest in numerous
interdisciplinary applications owing to their simplicity, practical applicability,
cost-effective production, high stability, and high selectivity. Surfactants added
during nanoparticle synthesis play a critical role in optimizing the sensor
sensitivity and selectivity because they directly affect the nanoparticle properties.

1
Mini-Review

Further, a new category of surfactants with magnetic properties has been


introduced for drug delivery applications. In this short review, we provide an
overview of the fundamentals of surfactants and their applications for the
development of nanotechnology.

Keywords: Surfactant; amphiphilic; stabilizer; nanoparticle synthesis; nano-


sensors; magneto surfactants

of
ro
-p
re
lP
na
ur
Jo

2
Mini-Review

1. Introduction

Surfactants are organic compounds composed of two chemical parts with


different polarities, a head group with affinity for polar phases, and a tail group
that is attracted to nonpolar phases [1-7]. Due to their structural uniqueness,
surfactants can be widely used to reduce the surface and interfacial tension
between two or more phases [8-11]. Their tendency to generate self-assembled
structures in solution can also lead to the formation of micelles with diameters

of
ranging from nanometers to microns [12, 13]. The amphiphilic nature of

ro
surfactants makes them suitable for use in numerous industrial products,
-p
including medicines [14-16], corrosion inhibitors for protecting steel and other
re
corrosive metals [17-20], detergents [21-23], de-emulsifiers [24, 25], wetting
lP

agents [26, 27], oil recovery enhancers [28-31], pour-point depressants [32-34],
pharmaceutical formulations [35-37], and drug delivery [38-40].
na

In addition to industrial applications, surfactants have great potential to


overcome the current limitations in nanotechnology [41-46]. For example,
ur

amphiphilic surfactants have been reported as stabilizers for the preparation of


Jo

stable dispersions of hydrophobic inorganic nanomaterials such as carbon


nanotubes [47-49], graphene [50], transition metal dichalcogenides [51], and
black phosphorus [52]. Since the first graphene exfoliation method,
micromechanical exfoliation, was discovered in 2004 [53], researchers have
attempted scale-up of the process by stabilizing exfoliated graphene in solvents.
Considering the accessibility and environmental toxicity of organic solvents [48],
water may be an ideal solvent system for preparing nanomaterial dispersions, but
most inorganic nanomaterials have hydrophobic surfaces and thus exhibit
repellent behavior in water. To overcome this limitation, various organic solvents

3
Mini-Review

have been exploited for stabilizing nanomaterials, such as N-methyl-pyrrolidone


and dimethylformamide [54-58]. Although these organic solvents can
successfully disperse nanomaterials, their high boiling points leads to the
requirement for additional high-temperature annealing to remove the solvent,
which may induce chemical degradation of the nanomaterials and obscure their
toxicity limits for use in biological applications [48]. This hurdle was overcome
by adding amphiphilic surfactants to water, together with inorganic nanomaterials.

of
The hydrophobic parts of the surfactants interact with the surfaces of the

ro
nanomaterials, and the hydrophilic parts stabilize the surfactant-nanomaterial
-p
composite in water. Furthermore, the buoyant densities of nanomaterials can be
controlled based on the type of surfactants, which enables the separation of
re
nanomaterials in the structure via isopycnic density gradient ultracentrifugation
lP

[59]. The resulting structurally sorted nanomaterials have contributed


significantly to many research fields in nanotechnology, especially electronics
na

[60-62], optoelectronics [57, 63, 64], energy [58, 65, 66], biology [48], and
ur

catalysis [54].
Jo

Due to the potential advantage of surfactants for tuning the surface/interface


tensions between solid/liquid interfaces and further improving the dispersion
stability, surfactants may be a key agent in the synthesis of nanoparticles with
well-controlled geometries. Recently, researchers have focused on nanoparticle-
based sensors that detect organic and inorganic pollutants and biomolecular
compounds in serum or urine samples as biomarkers for the early diagnosis of
diseases. One approach is to use localized surface plasmon resonance phenomena
on a nanoparticle array, where the nanoparticle size and shape, degree of
aggregation, and refractive index affect the sensitivity of the sensors [67, 68].
Furthermore, advanced sensing systems have been developed for enhancing both

4
Mini-Review

the identification and transduction processes by using many novel nanoparticles


rather than conventional organic dye-based sensing assays, where the former
offer advantages in sensitivity, selectivity, stability, reusability, and practical
applicability [69-75]. Among the available methods, colorimetric methods are
simple, highly efficient, and exhibit great potential for point-of-care diagnosis
because the detection process is facile as responses can be discerned by the naked
eyes or via simple instrumental techniques [76-81]. For colorimetric applications,

of
the unique physical and chemical properties (size, shape, large active surface

ro
area, and higher activity) of nanoparticles afford tremendous catalytic activities
-p
that differ from those of their bulk counterparts due to size-induced quantum
confinement effects in the former [80, 82-85]. For example, colloidal plasmonic
re
gold and silver nanoparticles have been most intensively studied as they produce
lP

rapid and vivid responses to incident light, and can be efficiently used for the
sensitive detection of target analytes through visible color changes in the solution
na

phase [86-91].
ur

In this short review, we provide an overview of surfactants for nanoparticle


Jo

synthesis and discuss the detailed relationships between the molecular structure of
the surfactant and the resulting particle structure distributions for optimizing
ultrasensitive detection in sensing application devices. We also discuss a new
category of surfactants that have magnetic properties for drug delivery, protein
separation, DNA extraction and catalyst applications.

2. Surfactant structure dependent synthesis and properties

2.1 Surfactant modification for nanoparticle synthesis

5
Mini-Review

The amphiphilic characteristics of surfactants play an important role in


nanoparticle synthesis by preventing aggregation of the synthesized nanoparticles
and providing more stability in colloidal systems [43, 92-97].
Surfactant-assisted nanoparticle synthesis has attracted considerable
attention in the development of nanoscale sensors for detecting toxic materials
[98-100], biomaterials, and biomarkers [101-103]. Because nanoscale sensors
should distinguish a trace amount of material, it is necessary to achieve high

of
sensitivity and selectivity for a specific analyte. The sensitivity and selectivity are

ro
strongly related to the structural parameters of the nanoparticles, including the
-p
precise size distribution. Consequently, controlling the size by modifying the
molecular structure of the surfactant can provide a new strategy for maximizing
re
the sensitivity enhancement. One such approach in which the tail length of the
lP

surfactant was altered to control the nanoparticle size, growth, and colloidal
stability was recently reported [95, 96, 104].
na

The colloidal stability of surfactant-nanoparticle hybrid systems can be


ur

described by the thermodynamic Gibbs free energy changes (ΔG°), which


Jo

strongly depend on the free energy for surfactant micellization (ΔG°mic) and
adsorption (ΔG°ads), surface tension, and critical micelle concentration. Such
factors are highly correlated to the molecular structure of surfactants, from head
to tail [105]. To understand the impact of the molecular structure of the surfactant
on the stability of the synthesized nanoparticles in the colloidal state, we must
first understand the behavior of surfactants in solution systems.
Surfactants consist of two parts with opposing polarities, which makes the
surfactant solution thermodynamically unstable due to the repulsive force
between the hydrophobic tail of the surfactant and the polar medium in which the
surfactant is dissolved. This repulsion forces the surfactant to be adsorbed on the

6
Mini-Review

nanoparticle and undergo micellization to stabilize the system by lowering the


free energy. In the surfactant-nanoparticle hybrid system, the surfactant
monomers direct themselves toward the surface of the nanoparticles through
interaction of the hydrophobic surfactant tail with the nanoparticles via
electrostatic interaction or van der Waals forces during the synthesis process,
while the hydrophilic surfactant head is in contact with the polar solvent,
consequently reducing the interfacial tension between the nanoparticles and

of
solvent. This arrangement decreases the free energy of the surfactant-nanoparticle

ro
hybrid system and enhances stabilization of the nanoparticles in the surfactant
-p
solution [104, 106-110]. Thus, increasing the surfactant hydrophobicity by
chemical structure modification will enhance both the adsorption and self-
re
assembly of the surfactant on the nanoparticle surface, and greater stability in
lP

solution can be attained.


Ordóñez et al. reported the effects of nonionic (polyvinyl pyrrolidone,
na

PVP), cationic (cetyltrimethylammonium bromide, CTAB), and anionic


ur

surfactants (sodium dodecylbenzene sulfonate, SDBS) on the stability of


Jo

zirconium oxide nanoparticles, ZrO2. The results outlined an improvement in the


stability, and reduced agglomeration of the synthesized ZrO2 with the surfactant
[95]. Both SDBS and CTAB surfactants conferred higher stability to the ZrO2
suspension, with zeta potentials of −50.38 ± 2.13 and 42.22 ± 3.27 mV compared
to that induced by PVP with a zeta potential of −19.99 ± 2.9 mV. A similar trend
was observed by Grządka et al. who reported that some cationic surfactants such
as CTAB, silicone surfactants, and fluorocarbon (S-106-A) surfactants improved
the stability of alginic acid/ZrO2 nanofluid [111].
The colloidal stability of the as-prepared AgNPs was found to be dependent
on the length of the hydrocarbon tail of a series cationic surfactants, as reported

7
Mini-Review

by Shaban et al. [112]. The surfactant with a longer tail (C16Dim) produced more
stable AgNPs with a zeta potential of 52 ± 10.7 mV, whereas the shorter-tail
surfactant C10Dim produced less stable AgNPs with a relatively lower zeta
potential of 34.1 ± 10.4 mV. Another study on the effect of modifying the
surfactant chain length on the synthesized AgNPs was also conducted by Abd-
Elaal et al., where they found that the aggregation and particle size mainly
depended on the length of the hydrophobic tail of three different nonionic

of
surfactants, namely, HTOPD, HTOPT, and HTOPH, with 12, 14, and 16 carbon

ro
atoms, respectively [113]. In addition, the surfactant HTOPH with a longer chain
-p
length yielded more stable AgNPs, whereas the zeta potentials of the AgNP
colloids obtained with HTOPD, HTOPT, and HTOPH were ‒23.7, ‒32.7, and ‒
re
35.4 mV, respectively. Similar behaviors have also been reported elsewhere [114-
lP

116]. Aiad et al. reported the dependence of the stability of AgNPs on the
amphipathic alkyl chain of two series of antipyrine cationic surfactants [117]. By
na

utilizing a series of surfactants (APC8, APC12, and APC16, where 8, 12, and 16
ur

refer to the number of carbon atoms in the alkyl chain), the zeta potential
Jo

increased to +31, +34, and +40 mV for the series APB8, APB12, and APB16,
respectively, indicating an increase in the nanoparticle stability with the
surfactants [117].

The effects of the spacer length of a synthesized Gemini cationic surfactant


with structures 18-s-18 (s equal to 3, 4, 6, 8, 10, and 12) on the stability of AuNPs
were reported by Qian Liu et al. [118]. As the spacer length increased, the
stability of the synthesized AuNPs increased, as confirmed by the increased zeta
potential of the AuNPs. The maximum zeta potential was obtained when s = 8
carbon atoms. Similar results have been reported by Pisárčik Martin et al. [119],

8
Mini-Review

who investigated the correlation between the surfactant structure and stability of
as-synthesized AgNPs. Increasing the length of the hydrophobic carbon
substituents associated with the ammonium head led to the formation of AgNPs
with a higher zeta potential, as an indication of their higher stability [119]. Morsi
et al. reported that using CTAB and SDS as respective cationic and anionic
surfactants improved the stability of yttrium oxide nanoparticles, where the
measured zeta potentials were 37.3 mV and ‒61.7 mV for the Y2O3/CTAB and

of
Y2O3/SDS systems, respectively, indicating stable nanofluid systems. In

ro
accordance with previous reports, Peng et al. reported that the stability of C60
-p
nanoparticles was improved by SDBS, where the zeta potential of the C60-SDBS
complex was ‒49.5 mV [96].
re
Based on previous reports, it is concluded that the surfactant strongly affects the
lP

stability of the synthesized nanoparticles. The stability is also influenced by the


hydrophobic tail of the surfactant and the spacer, as in the case of the Gemini
na

surfactant, where these factors improved the stability of the produced nanofluid
ur

system.
Jo

In general, modification of the chain length of surfactants and the hydrophobic


spacers between the hydrophilic heads are found to control the size distribution of
the as-synthesized nanoparticles, which can provide a route for enhancing the
sensitivity of nanoscale sensors. Adsorption of surfactant molecules onto the
surface of the nanomaterials not only controls their stability, but also significantly
controls the particle size. The surfactant is characterized by excellent adsorption
on the surface of the nanoparticles and can readily form micelles. The presence of
the surfactant during nucleation of the nanoparticles can control the growth of the
nanoparticles. As the hydrophobicity of the surfactant increases, the rate of
migration onto the surfaces of nanoparticles in nanomaterials increases, forming a

9
Mini-Review

more protective layer, thereby maintaining the dispersion of the nanoparticles in


solution with less agglomeration. Consequently, a larger particle size is attained.
In another study on the effect of modifying the surfactant chain length on
the particle size of the synthesized AgNPs, Abd-Elaal et al. found that the particle
size mainly depended on the length of the hydrophobic tail of three different
nonionic surfactants, HTOPD, HTOPT, and HTOPH, having 12, 14, and 16
carbon atoms, respectively. It has been reported that the size of AgNPs is reduced

of
as the tail length increases from 12 to 16 (Figure 1b) [113].

ro
-p
re
lP
na
ur
Jo

Figure 1. (a) Characterization of Au nanoparticles formed with Gemini surfactant


(18-s-18) having different spacer lengths. Reproduced from Ref. [118]. (b)
Surfactant tail length-dependence of aggregation tendency and size of AgNPs:
(A) HTOPD, (B) HTOPT, and (C) HTOPH nonionic surfactants. Reproduced
from Ref. [113].

10
Mini-Review

Shaban et al. revealed the effect of the length of the hydrophobic tail of the
surfactant by exploring the effect of cationic, Gemini cationic, and polymeric
anionic surfactants on the geometric modifications and size of silver nanoparticles
(AgNPs) prepared in situ via the photochemical reduction method [104, 110,
120](Figure 2a-c). Increasing the hydrocarbon chain length of the surfactant tail
led to the formation of smaller AgNPs.

of
ro
-p
re
lP
na
ur
Jo

Figure 2. TEM images and DLS size distribution data for AgNPs synthesized
using different surfactants with various hydrophobic chain lengths; the
hydrophobic chain length increases from left to right. (a) Gemini cationic

11
Mini-Review

surfactant-AgNP colloidal system, (b) anionic polymeric surfactant-AgNP


colloidal system, and (c) cationic surfactant-AgNP colloidal system. Reproduced
from Refs. [104, 110, 120].

Similar results have been reported by Pisárčik Martin et al. [119], who
investigated the correlation between the surfactant structure and particle size of
as-synthesized AgNPs. Increasing the length of the hydrophobic carbon

of
substituents associated with the ammonium head led to smaller AgNPs. The size

ro
of the AgNPs was controlled by changing the length of the surfactant tail, as
-p
reported by Feng Xu, et al. A new series of Gemini surfactants Cn-C4-Cn.2Br,
where n = 12, 14, 16, and 18 carbon atoms, was prepared, and it was shown that
re
the size of the AgNPs decreased as the number of carbon atoms attached to the
lP

hydrophilic head increased; the size was 5.6 ± 1.9, 5 ± 1.8, 3.9 ± 1.4, and 4 ± 1.4
nm, respectively [121]. Similar behaviors have also been reported elsewhere
na

[114-116]. Aiad et al. reported size control of AgNPs based on the amphipathic
ur

alkyl chain of two series of antipyrine cationic surfactants [117]. By utilizing the
Jo

APC8, APC12, and APC16 series, where 8, 12, and 16 refer to the number of
carbon atoms in the alkyl chain, the size of the AgNPs decreased from 25 to 15
nm as the length of the hydrophobic tail increased [117]. Surfactant-dependent
size modification has also been demonstrated for ZnS nanoparticles, where
particle diameters of 55.5 ± 0.5, 13.4 ± 0.5, and 11.6 ± 0.5 nm were obtained
using different surfactants, namely DTAB, TTAB, and CTAB, respectively,
where the formation of smaller nanoparticles was triggered by the hydrophobicity
of the surfactant tail [122]. The effect of the nonionic chain length (polyethylene
glycol chains) on the preparation of a silver nanohybrid was reported by Negm et
al. [123]. The particle diameter was influenced by the polyethylene glycol chains

12
Mini-Review

joined to the polyurethane nonionic surfactant. The size of the AgNPs decreased
from 152 to 14 nm as the molecular weight of polyethylene oxide in the nonionic
chains increased from 400 to 100000 g/mol. El-Dib et al. reported that the use of
a surfactant during the synthesis of magnetite nanoparticles (Fe2O4) produced
smaller particles than those achieved without surfactants [124].
The head groups of surfactants can be replaced that affects to size,
hydrophobicity/hydrophilicity, and degree of aggregation of nanoparticles. Mehta

of
et al. have reported that the size of ZnS nanoparticles and their dispersion stability

ro
can be controlled by variation of the hydrophilic head group [125]. Two cationic
-p
surfactants with different hydrophilic cetyltrimethylammonium chloride (CTAC)
and cetyltrimethylpyridinium chloride (CPyC) were explored, where the large
re
head surfactant CPyC shows higher affinity and yields smaller ZnS nanoparticles
lP

with higher colloidal stability compared to CTAC [125]. Based on the approach,
it was proved that the smaller size of ZnS can be synthesized using CPyC due to
na

more hydrophobic pyridine ring in CPyC compared to the trimethyl groups in


ur

CTAC, and thus the hydrophobicity of the CPyC is much higher than CTAB,
Jo

which can be the controlling factor of ZnS nanoparticles size.


In addition to the size control, shapes of nanoparticles can be controlled
based on a molecular structure of incorporated surfactant via a controlling the
facet growth method. In this case, surfactants play a role as a shape directing
agent [126-129]. A hydrophobic tail length of a surfactant significantly affects a
shape of gold nanorods as reported by Gao et al., as the length of the alkyl chain
increases from 10 to 16 carbons atoms, an aspect ratio of synthesized gold
nanorods increases [127]. A longer hydrocarbon chain produces a compacted
bilayer because of the stronger hydrophobic tail interactions that allows less
possibility of freshly generated gold atoms penetration to nucleate laterally,

13
Mini-Review

therefore both ends of the nanorod are preferred for further growth resulting in
the formation of higher aspect ratio gold nanorods [129, 130]. Furthermore,
surfactants control the nucleation and growth of nanoparticles during a seed-
mediated growth, and yield various shapes of gold and silver nanoparticles (e.g.,
rods, spheres, cubes, and octahedra) [131-134]. The Kou et al. used the seed
mediated growth method for gold nanorods and bipyramids synthesis. They
revealed that cetyltriethylammonium bromide (CTEAB) can control the shape of

of
the gold nanorods while absence of the CTAEB and using citrate as stabilizing

ro
agent can produce both gold nanorods and gold bipyramids [135].
-p
2.2 Surfactant structure dependent nanoparticle properties
re
lP

Modification of the geometry of as-synthesized nanoparticles based on the


molecular structures of the surfactants, where the size distribution and the
na

colloidal stability are strongly correlated to the alkyl chain length, can open a new
avenue for optimizing the desired properties of nanoparticles. In this section, we
ur

discuss the effects of varying the surfactant structure (i.e., alkyl chain length) on
Jo

the nanoparticle properties.


Dadwal and Joy et al. reported the colloidal stability of Fe3O4 nanoparticles in
toluene solvent using myristic, palmitic, and stearic acid as surfactants with long
alkyl chains. It was demonstrated that different chain lengths influenced the
thermal conductivity of the magnetite nanofluid under an applied magnetic field.
The thermal conductivity decreased as the length of the surfactant chain increased
[136]. Modification of the alkyl chain length also affects the antimicrobial
activity of nanoparticles, as reported by Shaban et al. It was demonstrated that
increasing the tail length of the surfactant changed the antibacterial activity of the

14
Mini-Review

AgNP nanohybrid system [112, 114-116]. In another study by Aiad et al., a


similar result was demonstrated wherein the surfactant tail length modified not
only the size of the AgNPs, but also their antimicrobial activity [117]. In a
different material obtained by Negm et al. [137], it was shown that cationic
polymeric surfactants with different alkyl chain lengths (8 and 12 carbon atoms)
did not affect the size of copper nanoparticles, but influenced the antimicrobial
activity. In addition, increasing the number of ethylene oxide units in a

of
synthesized cationic quaternary polymer produced smaller silver nanoparticles

ro
with high microbial activity against sulfate-reducing bacteria [138]. Zhou et al.
-p
investigated the synergetic effect between SDS and TiO2 nanoparticles on the
viscosity of water-based nanofluids at different temperatures. Zhou reported that
re
the resulting viscosity of the suspension when both SDS and TiO2 were added
lP

together in water was lower than when they were added separately [139]. Kumari
et al. investigated the effect of some surfactants, CTAB, SDS, and TX-100, on the
na

toxicity of AgNPs toward the crop Fagopyrum esculentum L [140]. The


ur

phytotoxic effect of the AgNPs formed without any surfactant induced 77.4% cell
Jo

death compared to the control without AgNPs. Treatment of the plant crop
Fagopyrum esculentum with AgNPs with incorporated surfactants (TX-100,
CTAB, and SDS) led to a significant reduction in the cell death (by 37%). As
illustrated in Figure 3, that study provides promising data on the significant role
of surfactants in enhancing plant tolerance to AgNP stress and ensuring food
safety [140].

15
Mini-Review

of
ro
Figure 3. Schematics showing the change in the phytotoxic effect of AgNPs
-p
on the crop Fagopyrum esculentum with the use of different surfactants.
re
Reproduced from Ref. [140].
lP

Ghahfarokhi et al. investigated the effects of different molecular weights of


na

ethylene oxide surfactants on the magnetic properties of spinel-strontium ferrite


(SrFe2O4) nanoparticles [141]. The addition of ethylene glycol surfactants
ur

decreased the particle size of SrFe2O4, thereby increasing the magnetism of


Jo

SrFe2O4, as evaluated using a vibrating sample magnetometer, owing to the


reversible relationship between the magnetic coercivity and particle size of
SrFe2O4, which greatly depends on the ethylene glycol surfactant.

3. Ionic surfactants with magnetic properties

In the past decade, an enthralling new class of ionic liquid surfactants with
magnetic responsivity, termed magnetic ionic liquids or magneto surfactants, has
been reported [142-144]. These new categories of surfactants exhibit a unique
magnetic response in addition to the common properties of the surfactants (i.e.,

16
Mini-Review

adsorption and micellization). Magneto surfactants were first discovered by


Brown et al. in 2012 [145, 146], and have been rapidly developed in the fields of
biomedicine and magneto nanomaterial immobilization [147, 148]. In this
section, we discuss the synthesis, properties, and applications of magneto
surfactants.

3.1. Synthesis of magneto surfactants

of
Magneto surfactants were synthesized by stirring a metal trihalide with a

ro
surfactant in an alcoholic solution for 12 h, as shown in Scheme 1a [149]. Kevin
-p
et al. reported the successful synthesis of a new series of paramagnetic ionic
liquid surfactants based on N,N'-bis(alkyl)imidazolium bromotrichloroferrate.
re
Scheme 1b illustrates the protocol for the synthesis of [(Cn)2Im][FeCl3Br] and
lP

[CnC1Im][FeCl3Br], where C1 refers to a methyl group; Cn refers to ethyl, butyl,


hexyl, octyl, decyl, and dodecyl groups in succession; and Im refers to imidazole
na

[150].
ur
Jo

17
Mini-Review

of
ro
-p
re
lP
na
ur
Jo

Scheme 1. (a) Synthesis of magneto surfactant alkyltrimethylammonium


bromotrichloroferrate, reproduced from Ref. [149]. (b) Preparation of
[(Cn)2Im][FeCl3Br] and [CnC1Im][FeCl3Br] magnetic surfactants. Reproduced
from Ref. [150].

Subsequently, Brown et al. reported a new set of magnetic surfactants based on


a lanthanide metal as a counter ion (Table 1), where one mole of metal trichloride
was added to the methanolic solution of nonmagnetic surfactant DTAB or
C10mimCl, the mixture was stirred overnight (12 h) at room temperature, and then

18
Mini-Review

dehydrated under vacuum at 80 °C overnight to obtain the desired magnetic


nanoparticles [145].

Table 1. Chemical structure of some prepared lanthanide magnetic surfactants.


Reproduced from Ref. [145]

of
ro
-p
re
lP
na
ur

Under an applied magnetic field, the surface tension changes reversibly, and this
property of magneto surfactants allows a switching “on” and “off” behavior by
Jo

controlling the external magnetic field [151]. A change in the surface parameters
was also proved by Brown et al., where the surface tension could be reduced by
12.1% compared to that achieved with the corresponding nonmagnetic surfactant
under identical conditions, as illustrated in Figure 4. Here, H refers to holmium
metal [145].

19
Mini-Review

of
ro
Figure 4. Pendant drop profiles of magnetic surfactants (0.1 mol) with and
without an external magnet. Reproduced from Ref. [145].
-p
re
Another new series of cationic and anionic magneto surfactants based on a
lP

gadolinium counter ion [152] with CTAG (cationic


hexadecyltrimethlyammonium bromotrichlorogadolinate) and DTAG
na

(decyltrimethylammonium bromotrichlorogadolinate) was demonstrated. These


ur

magnetic surfactants were synthesized by stirring overnight using an equivalent


quantity of GdCl3.6H2O and either DTAB or CTAB in methanol, following which
Jo

the solvent was evaporated under vacuum at 80 °C. The anionic magneto
surfactant based on gadolinium Gd(AOT)3 (tris-(1,4-bis(2-ethylhexoxy)-1,4-
dioxobutane-2-sulfonate) was synthesized via a liquid−liquid ion exchange
method. The GdCl3.6H2O was dissolved in an ethanol and water mixture (75:25
v/v, ethanol:water); 3 M aqueous NaAOT was added to the aforementioned
solution and stirred for 5 h. The solvent was then removed under low-pressure
and the sample was dried for 2 d at 80 °C. The anionic magneto surfactant was
then resolubilized in dry dichloromethane to remove the insoluble salts, and
finally, the pure product was obtained by centrifugation [152]. The synthesized

20
Mini-Review

magneto surfactants were functionalized with proteins; consequently, protein


separation and extraction was enabled by simply using an external magnetic field.
It is noted that the toxicity of magneto surfactants, in some cases, is lower than
that of conventional halide-based surfactant analogs [152] which may make
magneto surfactants suitable for use in contaminated or infected sites to enable
selective microbial reduction without influencing normal tissues and cells [153].

of
3.2. Applications of magneto surfactants

ro
For cancer therapy, the emerging technology of smart nano-vehicles has
-p
received significant attention for improving chemotherapy efficiency and
preventing side effects. This has been increased by directing drugs to the targeted
re
area and controlling drug release. Nucleic acids are crucial biological building
lP

blocks as they are characterized by molecular self-recognition [154-156].


Although DNA acts as a better intercalating platform to carry anticancer drugs, it
na

is not considered a good nano-vehicle for delivering cancer drugs to infected


ur

tissues for specific targeting [157] because DNA with high molecular weight
Jo

does not readily pass across the cell membranes independently, resulting in poor
intracellular uptake due to repulsive interaction between the negatively charged
DNA and cytomembrane.
When the surfactant concentration is increased beyond the critical micelle
concentration limit, the surfactant tends to aggregate as cationic micelles in the
vicinity of DNA [158-160]. These multivalent cationic micelles interact with
DNA through hydrophobic and electrostatic interactions to decrease the repulsion
between the DNA segments. This induces interaction between the multivalent
cationic micelles and at least three positive centers and DNA, and consequently
further induces compaction from the extended linear structure to the compacted

21
Mini-Review

form [161]. Although this approach can provide a solution for DNA delivery, it
still presents obstacles with respect to sensitivity and controllability, which are
required for more specific tissue targeting. By replacing the conventional
surfactants with magneto surfactants, a solution can be provided, where magnetic
switching allows the surfactants to reach an effective concentration under an
external magnetic field, resulting in a more condensed DNA structure at lower
concentration [148, 162]. The magneto surfactant with a magnetic counter ion has

of
more functionalities that interact with the DNA [163]. Hao et al. reported that 10

ro
mM (high concentration) of the conventional surfactants CTAB or CTAC could
-p
promote the de-compaction of 0.15 mM of DNA, whereas the same concentration
of a magneto surfactant in which the counter ions (Cl¯ or Br¯ ) were replaced by
re
[FeCl3Br]¯ afforded completely compacted DNA [163].
lP

Ling Wang et al. synthesized multiple responsive nano-vehicles that


effectively combine with DNA by utilizing a magneto surfactant and glutathione
na

GSH-reductive triggers to control the delivery of the anticancer drug doxorubicin


ur

(DOX) with precise targeting, as illustrated in Figure 5a. The magneto surfactant
Jo

CTAF, where the counter ion is [FeCl3Br]¯ , was synthesized with fullerenes
(C60) to increase the solubility and dispersity and increase the overall positive
charge around the cationic micelles to produce magnetic fullerenes
(C60@CTAF). DNA loaded with the anticancer drug DOX and treated with the
modified polysaccharide hyaluronic acid was then fabricated with as-synthesized
C60@CTAF to form C60@CTAF/DNA/HA-SS-COOH. The composite was then
used as an electrostatic platform to load the anticancer drug DOX, and the DNA-
DOX backbones were assembled with C60@CTAF to construct
C60@CTAF/DNA complexes. Under an applied magnetic field,

22
Mini-Review

C60@CTAF/DNA/HA-SS-COOH afforded controlled transport of the anticancer


drug DOX to the target tissue [147].

of
ro
-p
re
lP
na
ur
Jo

Figure 5. (a) Schematic representation of magnetic fullerene-DNA vehicles.


Reproduced from Ref. [147]. (b) Schematic of ordered magneto surfactant/DNA
nanosphere hybrid: synthesis and structure. Reproduced from Ref. [164].

23
Mini-Review

Xu et al. synthesized well-ordered surfactant-DNA hybrid nanospheres using


a magnetic cationic surfactant (C16H33(CH3)3N+[FeCl3Br]−, CTAFe) and a
cationic ionic liquid C2H5O-azobenzene-OC2H4(CH3)3N+[FeCl3Br]−, azoTAFe.
The synthesized hybrid nanospheres could be targeted using a magnetic field, as
shown in Figure 5b. The procedure for synthesizing the final nanosphere
products was via DNA compaction, aggregation, and coagulation, and the
synthesized nanospheres could be used as promising drug vehicles in the field of

of
nanomedicine [164]. Subsequently, Xu et al. reported the synthesis of AuNPs

ro
using the magnetic surfactant CH3(CH2)14CH2N(CH3)3+ [FeCl3Br]‒ (Figure 6a)
-p
[148]. These magnetic AuNPs enabled DNA delivery under the influence of a
low-strength magnetic field.
re
lP
na
ur
Jo

24
Mini-Review

of
ro
-p
re
lP
na
ur
Jo

Figure 6. (a) Schematic representation of synthesis of AuNPs using magneto


surfactant (A) and DNA and protein migration under low-strength external
magnetic field (B). Reproduced from Ref. [148]. (b) Synthesis of magnetic network
clusters and subsequent in situ AgNP formation on network templates (A);
synthesis of magnetic network for capturing specific sequence of DNA (B),
Reproduced from Ref. [165]. (c) Schematic illustrating control of biomolecule
migration by magneto-surfactant-coated carbon nanotubes. Reproduced from Ref.
[166].

25
Mini-Review

Wang et al. utilized fluorescent carbon quantum dots (CQDs) conjugated with
cysteine to produce a ligand (cys-CQDs), which then self-assembled with the
synthesized magnetic surfactants (CTAHo and CTAGd) via a facile non-covalent
method [165]. This magnetic network can separate DNA molecules with the
assistance of an external magnet. The cysteine protein on the CQD-magneto
surfactant network was used for in situ synthesis of AgNPs, as shown in Figure
6b (A), through three interaction centers in the cysteine molecules (carboxylate,

of
ammonium, and sulfide groups). The AgNPs produced on the CQD-magneto

ro
surfactant network were functionalized with DNA modified-thiol through the
-p
bond between Ag and sulfide. The thus-constructed magnetic vehicle was used to
select specific sequences of DNA by DNA base pairing, as illustrated in Figure
re
6b (B), thereby furnishing magnetic assemblies with high capture efficiency and
lP

recovery [165].
Xu et al. prepared cetyltrimethyl ammonium trichloromonobromocerate
na

(CTACe). The cationic surfactant CTAB was stirred with CeCl3 in a molar ratio
ur

of 1:1 in methanol for 12 h under environmental conditions, and the solvent was
Jo

evaporated to obtain the magneto surfactant (CTACe), which was then mixed
with carbon nanotubes under ultrasonication to obtain carbon nanotubes
functionalized with the magneto surfactant (CNTs@CTACe) [166].
CNTs@CTACe was used as an efficient nano-vehicle for drug delivery, as
illustrated in Figure 6c. The synthesized CNTs@CTACe can electrostatically
interact with proteins and DNA and also make compact structures of proteins and
DNA. Under an external magnetic force, the migration of DNA and proteins
could be controlled [166].
Zhao et al. synthesized magnetic-aggregated surfactants comprised of a
layered structure by co-assembling a double-tailed magneto surfactant with a

26
Mini-Review

chemical structure comprising (C18)2C2N+[FeCl4]‒ and polyoxometalate (POM)


[167]. The POM-magneto surfactant (Figure 7a) exhibits high transportation
efficiency for delivering myoglobin (protein) under an applied magnetic field.

of
ro
-p
re
lP
na

Figure 7. (a) Fabrication of magnetic POM-magneto surfactant (A) and


ur

schematic of magnet-facilitated protein delivery B), reproduced from Ref. [167].


Jo

(b) Magnetic response of magnetic surfactant−graphene oxide systems; all


samples contain graphene oxide in water, but only the right-hand vials contain
additional magneto surfactant (DTA and min), reproduced from Ref. [168].

Graphene oxide (GO) has been intensively studied in a wide range of


applications, including water treatment [169], optics [170], and interface
stabilization [171], owing to its high dispersability, stability, and massive surface
area. Furthermore, GO has been modified and functionalized with magnetic
nanocomposites in many studies for targeted purposes. For example, modified

27
Mini-Review

GO can remover water pollutants such as arsenic [172], selenium [173],


chromium, and mercury [174, 175]. All of the aforementioned methods require
multi-step preparation, which undoubtedly affects the GO structure. To overcome
the issue, Thomas et al. developed a new method for implanting GO with an as-
synthesized magneto surfactant, where the magnetic ionic liquid surfactants, 1-
methyl-3-butylimidazolium tetrachloroferrate [mim] and
dodecyltrimethylammonium trichloromonobromoferrate [DTA], were

of
synthesized with the GO nanosheets, providing a simple and economic method

ro
for GO recycling by using an external magnetic field [168], as shown in Figure
7b. -p
In drug delivery, cytotoxicity of surfactants should also be considered and
re
discussed in elsewhere [176]. In addition to the properties of magnetic surfactants
lP

in the field of drug delivery as discussed above, magnetic surfactants have


advantages for their applications in many research fields of nanotechnology
na

compared to the non-magnetic conventional surfactants. For example, their


ur

unique magnetic response can be easily controlled under the external magnetic
Jo

field which can be particularly applicable for protein and DNA extraction and
self-aggregation-free molecular magnets. Beyond nanotechnology, the magnetic
surfactants have a great potential to use in environmental clean-up, water-
treatment. etc [177].

Conclusion and Future outlook


In this short review, we discussed the recent advances in surfactants and
their tunable, structure-dependent properties for optimized applications in
nanotechnology. During the synthesis of nanoparticles in solution, surfactants
play an important role in improving the dispersion stability and controlling the

28
Mini-Review

size distributions of the as-synthesized nanoparticles. In more detail, as the tail


length of the surfactant used for nanoparticle synthesis increases, smaller, less
aggregated, and more stable nanoparticles are produced. This development can
significantly promote colorimetric sensor applications based on size-tunable
nanoparticles. However, the correlation between the surfactant structure and
nanoparticle geometry should still be defined through intensive studies to provide
direct evidence of this correlation and molecular fingerprints. Furthermore, a new

of
category of surfactants, namely, magneto surfactants, has been introduced. By

ro
forming a mixture of conventional surfactants and magnetic elements, a new
-p
series of surfactants having unique magnetic properties can be successfully
synthesized. Magneto surfactants have been further utilized in the field of drug
re
delivery because of their high selectivity and sensitivity to target a certain
lP

position. Although the emerging category of surfactants has been intensively


studied, it is still necessary to elucidate the fundamental mechanisms, provide
na

technical solutions, achieve deeper understanding of their toxicity, and enable the
ur

possibility of large-scale production for commercial uses, considering the short


Jo

history of surfactants since their initial discovery.

29
Mini-Review

Acknowledgement

This study was supported by the National Research Foundation of Korea (NRF) grant, funded

by the Korean Government (MSIT) (2020R1A5A1018052, 2020R1C1C1009381, and

2020R1A4A3079710).

of
ro
-p
re
lP
na
ur
Jo

30
Mini-Review

References

1. Yekeen, N., et al., Influence of surfactant and electrolyte concentrations on surfactant


Adsorption and foaming characteristics. Journal of Petroleum Science and Engineering, 2017.
149: p. 612-622.
2. Zhang, D., et al., Synthesis and properties study of novel fluorinated surfactants with
perfluorinated branched ether chain. Journal of Fluorine Chemistry, 2019. 219: p. 62-69.
3. Alam, M.S., et al., Physicochemical properties and bioactivity studies of synthesized counterion
coupled (COCO) gemini surfactant, 1,6-bis(N,N-hexadecyldimethylammonium) adipate. Journal
of Molecular Liquids, 2019. 273: p. 16-26.
4. Hussain, S.M.S., M.S. Kamal, and L.T. Fogang, Synthesis and physicochemical investigation of

of
betaine type polyoxyethylene zwitterionic surfactants containing different ionic headgroups.
Journal of Molecular Structure, 2019. 1178: p. 83-88.
5. Pal, N., K. Samanta, and A. Mandal, A novel family of non-ionic gemini surfactants derived from

ro
sunflower oil: Synthesis, characterization and physicochemical evaluation. Journal of
Molecular Liquids, 2019. 275: p. 638-653.
6. -p
Tehrani-Bagha, A.R., Cationic gemini surfactant with cleavable spacer: Emulsion stability.
Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2016. 508: p. 79-84.
re
7. Shaban, S.M., et al., Some alginates polymeric cationic surfactants; surface study and their
evaluation as biocide and corrosion inhibitors. Journal of Molecular Liquids, 2019. 273: p. 164-
lP

176.
8. Zhang, F., et al., Adsorption of different types of surfactants on graphene oxide. Journal of
Molecular Liquids, 2019. 276: p. 338-346.
na

9. Chernysheva, M.G., et al., Cationic surfactant coating nanodiamonds: Adsorption and


peculiarities. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2019. 565: p.
25-29.
ur

10. Chang, Z., X. Chen, and Y. Peng, The adsorption behavior of surfactants on mineral surfaces in
the presence of electrolytes – A critical review. Minerals Engineering, 2018. 121: p. 66-76.
Jo

11. Casandra, A., et al., Adsorption kinetics of the partially dissociated ionic surfactants: The effect
of degree of dissociation. Journal of the Taiwan Institute of Chemical Engineers, 2018. 92: p. 2-
7.
12. Möbius, D., R. Miller, and V.B. Fainerman, Surfactants: chemistry, interfacial properties,
applications. Vol. 13. 2001: Elsevier.
13. Rosen, M.J., Surfactants and Interfacial Phenomena. 2004: wiley.
14. Karsa, D.R., Industrial applications of surfactants IV. 1999: Elsevier.
15. Clarke, J.G., S.R. Wicks, and S.J. Farr, Surfactant mediated effects in pressurized metered dose
inhalers formulated as suspensions. I. Drug/surfactant interactions in a model propellant
system. International Journal of Pharmaceutics, 1993. 93(1): p. 221-231.
16. Torchilin, V.P., Structure and design of polymeric surfactant-based drug delivery systems.
Journal of Controlled Release, 2001. 73(2): p. 137-172.
17. El Achouri, M., et al., Corrosion inhibition of iron in 1 M HCl by some gemini surfactants in the
series of alkanediyl-α,ω-bis-(dimethyl tetradecyl ammonium bromide). Progress in Organic
Coatings, 2001. 43(4): p. 267-273.
18. Hegazy, M.A., M. Abdallah, and H. Ahmed, Novel cationic gemini surfactants as corrosion
inhibitors for carbon steel pipelines. Corrosion Science, 2010. 52(9): p. 2897-2904.

31
Mini-Review

19. Elachouri, M., et al., Some surfactants in the series of 2-(alkyldimethylammonio) alkanol
bromides as inhibitors of the corrosion of iron in acid chloride solution. Corrosion Science,
1995. 37(3): p. 381-389.
20. Qiu, L.-G., et al., Synergistic effect between cationic gemini surfactant and chloride ion for the
corrosion inhibition of steel in sulphuric acid. Corrosion Science, 2008. 50(2): p. 576-582.
21. Wu, H.-Y., et al., Development and validation of an analytical procedure for quantitation of
surfactants in dishwashing detergents using ultra-performance liquid chromatography-mass
spectrometry. Talanta, 2019. 194: p. 778-785.
22. Yada, S., et al., Emulsification, solubilization, and detergency behaviors of homogeneous
polyoxypropylene-polyoxyethylene alkyl ether type nonionic surfactants. Colloids and Surfaces
A: Physicochemical and Engineering Aspects, 2019. 564: p. 51-58.
23. Lee, S., et al., Synthesis of environment friendly nonionic surfactants from sugar base and

of
characterization of interfacial properties for detergent application. Journal of Industrial and
Engineering Chemistry, 2016. 38: p. 157-166.

ro
24. Atta, A.M., et al., Demulsification of heavy crude oil using new nonionic cardanol surfactants.
Journal of Molecular Liquids, 2018. 252: p. 311-320.
25. Abdulraheim, A.M., Green polymeric surface active agents for crude oil demulsification.

26.
-p
Journal of Molecular Liquids, 2018. 271: p. 329-341.
Xu, C., et al., Experimental investigation of coal dust wetting ability of anionic surfactants with
re
different structures. Process Safety and Environmental Protection, 2019. 121: p. 69-76.
27. Alexandrova, L., et al., Effects of pH on wetting behavior of ‘star-like’ block copolymer
lP

surfactant solutions. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2017.
519: p. 78-86.
28. Wang, Y., et al., Halide-free PN ionic liquids surfactants as additives for enhancing tribological
na

performance of water-based liquid. Tribology International, 2018. 128: p. 190-196.


29. Zhao, M., et al., Study on the synergy between silica nanoparticles and surfactants for
enhanced oil recovery during spontaneous imbibition. Journal of Molecular Liquids, 2018. 261:
ur

p. 373-378.
30. Howe, A.M., et al., Visualising surfactant enhanced oil recovery. Colloids and Surfaces A:
Jo

Physicochemical and Engineering Aspects, 2015. 480: p. 449-461.


31. Nourafkan, E., Z. Hu, and D. Wen, Controlled delivery and release of surfactant for enhanced
oil recovery by nanodroplets. Fuel, 2018. 218: p. 396-405.
32. Hafiz, A.A. and T.T. Khidr, Hexa-triethanolamine oleate esters as pour point depressant for
waxy crude oils. Journal of Petroleum Science and Engineering, 2007. 56(4): p. 296-302.
33. El Mehbad, N., Efficiency of N-Decyl-N-benzyl-N-methylglycine and N-Dodecyl-N-benzyl-N-
methylglycine surfactants for flow improvers and pour point depressants. Journal of Molecular
Liquids, 2017. 229: p. 609-613.
34. Gu, X., et al., Investigation of cationic surfactants as clean flow improvers for crude oil and a
mechanism study. Journal of Petroleum Science and Engineering, 2018. 164: p. 87-90.
35. Weiszhár, Z., et al., Complement activation by polyethoxylated pharmaceutical surfactants:
Cremophor-EL, Tween-80 and Tween-20. European Journal of Pharmaceutical Sciences, 2012.
45(4): p. 492-498.
36. Steinhilber, D., et al., Surfactant free preparation of biodegradable dendritic polyglycerol
nanogels by inverse nanoprecipitation for encapsulation and release of pharmaceutical
biomacromolecules. Journal of Controlled Release, 2013. 169(3): p. 289-295.

32
Mini-Review

37. Bouchemal, K., et al., Nano-emulsion formulation using spontaneous emulsification: solvent,
oil and surfactant optimisation. International Journal of Pharmaceutics, 2004. 280(1): p. 241-
251.
38. Drummond, C.J. and C. Fong, Surfactant self-assembly objects as novel drug delivery vehicles.
Current Opinion in Colloid & Interface Science, 1999. 4(6): p. 449-456.
39. Uchegbu, I.F. and S.P. Vyas, Non-ionic surfactant based vesicles (niosomes) in drug delivery.
International Journal of Pharmaceutics, 1998. 172(1): p. 33-70.
40. Rivera, A. and T. Farías, Clinoptilolite–surfactant composites as drug support: A new potential
application. Microporous and Mesoporous Materials, 2005. 80(1): p. 337-346.
41. Zhao, Q., et al., Interaction of nano carbon particles and anthracene with pulmonary
surfactant: The potential hazards of inhaled nanoparticles. Chemosphere, 2019. 215: p. 746-
752.

of
42. Hasheminejad, M. and A. Nezamzadeh-Ejhieh, A novel citrate selective electrode based on
surfactant modified nano-clinoptilolite. Food Chemistry, 2015. 172: p. 794-801.

ro
43. Huang, B., et al., Size-controlled synthesis and morphology evolution of Nd2O3 nano-powders
using ionic liquid surfactant templates. Journal of Alloys and Compounds, 2017. 712: p. 164-
171.
44.
-p
Amer, W., et al., Synthesis of mesoporous nano-hydroxyapatite by using zwitterions
surfactant. Materials Letters, 2013. 107: p. 189-193.
re
45. Li, X., et al., Effect of surfactants on the aggregation and stability of TiO2 nanomaterial in
environmental aqueous matrices. Science of The Total Environment, 2017. 574: p. 176-182.
lP

46. Asikin-Mijan, N., Y.H. Taufiq-Yap, and H.V. Lee, Synthesis of clamshell derived Ca(OH)2 nano-
particles via simple surfactant-hydration treatment. Chemical Engineering Journal, 2015. 262:
p. 1043-1051.
na

47. Hersam, M.C., Progress towards monodisperse single-walled carbon nanotubes. Nature
Nanotechnology, 2008. 3(7): p. 387-394.
48. Wang, X., et al., Toxicological Profiling of Highly Purified Single-Walled Carbon Nanotubes with
ur

Different Lengths in the Rodent Lung and Escherichia Coli. Small, 2018. 14(23): p. 1703915.
49. Yoon, K., et al., Metal-Free Carbon-Based Nanomaterial Coatings Protect Silicon Photoanodes
Jo

in Solar Water-Splitting. Nano Letters, 2016. 16(12): p. 7370-7375.


50. Green, A.A. and M.C. Hersam, Solution Phase Production of Graphene with Controlled
Thickness via Density Differentiation. Nano Letters, 2009. 9(12): p. 4031-4036.
51. Kang, J., et al., Thickness sorting of two-dimensional transition metal dichalcogenides via
copolymer-assisted density gradient ultracentrifugation. Nature Communications, 2014. 5(1):
p. 5478.
52. Kang, J., et al., Solvent Exfoliation of Electronic-Grade, Two-Dimensional Black Phosphorus.
ACS Nano, 2015. 9(4): p. 3596-3604.
53. Novoselov, K.S., et al., Electric Field Effect in Atomically Thin Carbon Films. Science, 2004.
306(5696): p. 666.
54. Kang, J., et al., Solution-Based Processing of Monodisperse Two-Dimensional Nanomaterials.
Accounts of Chemical Research, 2017. 50(4): p. 943-951.
55. Hu, G., et al., Functional inks and printing of two-dimensional materials. Chemical Society
Reviews, 2018. 47(9): p. 3265-3300.
56. Kang, J., et al., Solution-Processed Layered Gallium Telluride Thin-Film Photodetectors. ACS
Photonics, 2018. 5(10): p. 3996-4002.

33
Mini-Review

57. Kang, J., et al., Solution-Based Processing of Optoelectronically Active Indium Selenide.
Advanced Materials, 2018. 30(38): p. 1802990.
58. Lam, D., et al., Anhydrous Liquid-Phase Exfoliation of Pristine Electrochemically Active GeS
Nanosheets. Chemistry of Materials, 2018. 30(7): p. 2245-2250.
59. Kang, J., et al., Stable aqueous dispersions of optically and electronically active phosphorene.
Proceedings of the National Academy of Sciences, 2016. 113(42): p. 11688-11693.
60. Kim, S., et al., Carbon nanotube ferroelectric random access memory cell based on omega-
shaped ferroelectric gate. Carbon, 2020. 162: p. 195-200.
61. Zhu, J., et al., Solution-Processed Dielectrics Based on Thickness-Sorted Two-Dimensional
Hexagonal Boron Nitride Nanosheets. Nano Letters, 2015. 15(10): p. 7029-7036.
62. Choi, Y., et al., Capacitively Coupled Hybrid Ion Gel and Carbon Nanotube Thin-Film Transistors
for Low Voltage Flexible Logic Circuits. Advanced Functional Materials, 2018. 28(34): p.

of
1802610.
63. Zhong, C., et al., Hot Carrier and Surface Recombination Dynamics in Layered InSe Crystals. The

ro
Journal of Physical Chemistry Letters, 2019. 10(3): p. 493-499.
64. Husko, C., et al., Silicon-Phosphorene Nanocavity-Enhanced Optical Emission at
Telecommunications Wavelengths. Nano Letters, 2018. 18(10): p. 6515-6520.
65.
-p
Engel, M., et al., Graphene-enabled and directed nanomaterial placement from solution for
large-scale device integration. Nature Communications, 2018. 9(1): p. 4095.
re
66. Li, Q., et al., Revealing the Effects of Electrode Crystallographic Orientation on Battery
Electrochemistry via the Anisotropic Lithiation and Sodiation of ReS2. ACS Nano, 2018. 12(8):
lP

p. 7875-7882.
67. Guo, L.H., et al., Strategies for enhancing the sensitivity of plasmonic nanosensors. Nano
Today, 2015. 10(2): p. 213-239.
na

68. Jain, P.K., et al., Noble metals on the nanoscale: optical and photothermal properties and
some applications in imaging, sensing, biology, and medicine. Acc Chem Res, 2008. 41(12): p.
1578-86.
ur

69. Stewart, M.E., et al., Nanostructured Plasmonic Sensors. Chemical Reviews, 2008. 108(2): p.
494-521.
Jo

70. Song, S., et al., Functional nanoprobes for ultrasensitive detection of biomolecules. Chemical
Society Reviews, 2010. 39(11): p. 4234-4243.
71. Howes, P.D., R. Chandrawati, and M.M. Stevens, Colloidal nanoparticles as advanced biological
sensors. Science, 2014. 346(6205): p. 1247390.
72. Choi, S., et al., Recent Advances in Flexible and Stretchable Bio-Electronic Devices Integrated
with Nanomaterials. Advanced Materials, 2016. 28(22): p. 4203-4218.
73. Wu, F., P. Yu, and L. Mao, Self-powered electrochemical systems as neurochemical sensors:
toward self-triggered in vivo analysis of brain chemistry. Chemical Society Reviews, 2017.
46(10): p. 2692-2704.
74. Kumar, P., et al., Hybrid porous thin films: Opportunities and challenges for sensing
applications. Biosensors and Bioelectronics, 2018. 104: p. 120-137.
75. Paladiya, C. and A. Kiani, Nano structured sensing surface: Significance in sensor fabrication.
Sensors and Actuators B: Chemical, 2018. 268: p. 494-511.
76. Gunnlaugsson, T., et al., Colorimetric “Naked Eye” Sensing of Anions in Aqueous Solution. The
Journal of Organic Chemistry, 2005. 70(26): p. 10875-10878.
77. Song, Y., W. Wei, and X. Qu, Colorimetric Biosensing Using Smart Materials. Advanced
Materials, 2011. 23(37): p. 4215-4236.

34
Mini-Review

78. Verma, M.S., et al., Colorimetric biosensing of pathogens using gold nanoparticles.
Biotechnology Advances, 2015. 33(6, Part 1): p. 666-680.
79. Kangas, M.J., et al., Colorimetric Sensor Arrays for the Detection and Identification of Chemical
Weapons and Explosives. Critical Reviews in Analytical Chemistry, 2017. 47(2): p. 138-153.
80. Tsogas, G.Z., et al., Recent Advances in Nanomaterial Probes for Optical Biothiol Sensing: A
Review. Analytical Letters, 2018. 51(4): p. 443-468.
81. Piriya V.S, A., et al., Colorimetric sensors for rapid detection of various analytes. Materials
Science and Engineering: C, 2017. 78: p. 1231-1245.
82. J. Duan and J. Zhan, Recent developments on nanomaterials-based optical sensors for Hg2+
detection. Science China Materials, 2015. 58(3): p. 223-240.
83. Yang, T., et al., Recent Progresses in Nanobiosensing for Food Safety Analysis. Sensors, 2016.
16(7): p. 1118.

of
84. Chinta, J.P., Coinage metal nanoparticles based colorimetric assays for natural amino acids: A
review of recent developments. Sensors and Actuators B: Chemical, 2017. 248: p. 733-752.

ro
85. Ma, X.-M., et al., Progress of Visual Biosensor Based on Gold Nanoparticles. Chinese Journal of
Analytical Chemistry, 2018. 46(1): p. 1-10.
86. Vilela, D., M.C. González, and A. Escarpa, Sensing colorimetric approaches based on gold and
-p
silver nanoparticles aggregation: Chemical creativity behind the assay. A review. Analytica
Chimica Acta, 2012. 751: p. 24-43.
re
87. Zarlaida, F. and M. Adlim, Gold and silver nanoparticles and indicator dyes as active agents in
colorimetric spot and strip tests for mercury(II) ions: a review. Microchimica Acta, 2017.
lP

184(1): p. 45-58.
88. Jain, P.K., et al., Review of Some Interesting Surface Plasmon Resonance-enhanced Properties
of Noble Metal Nanoparticles and Their Applications to Biosystems. Plasmonics, 2007. 2(3): p.
na

107-118.
89. Guo, L., et al., Strategies for enhancing the sensitivity of plasmonic nanosensors. Nano Today,
2015. 10(2): p. 213-239.
ur

90. Tang, L. and J. Li, Plasmon-Based Colorimetric Nanosensors for Ultrasensitive Molecular
Diagnostics. ACS Sensors, 2017. 2(7): p. 857-875.
Jo

91. Sabela, M., et al., A Review of Gold and Silver Nanoparticle-Based Colorimetric Sensing Assays.
Advanced Engineering Materials, 2017. 19(12): p. 1700270.
92. Khan, Z., et al., Effects of shape-controlling cationic and anionic surfactants on the morphology
and surface resonance plasmon intensity of silver@copper bimetallic nanoparticles. Journal of
Molecular Liquids, 2019. 275: p. 354-363.
93. Chekuri, R.D. and S.R. Tirukkovalluri, Synthesis of cobalt doped titania nano material assisted
by gemini surfactant: Characterization and application in degradation of Acid Red under visible
light irradiation. South African Journal of Chemical Engineering, 2017. 24: p. 183-195.
94. Yuenyongsuwan, J., et al., Surfactant effect on phase-controlled synthesis and photocatalyst
property of TiO2 nanoparticles. Materials Chemistry and Physics, 2018. 214: p. 330-336.
95. Ordóñez, F., et al., Synthesis of ZrO2 nanoparticles and effect of surfactant on dispersion and
stability. Ceramics International, 2020. 46(8, Part B): p. 11970-11977.
96. Peng, X., et al., Aqueous stability and mobility of C60 complexed by sodium dodecyl benzene
sulfonate surfactant. Journal of Environmental Sciences, 2016. 42: p. 89-96.
97. Li, D., et al., Stability properties of water-based gold and silver nanofluids stabilized by cationic
gemini surfactants. Journal of the Taiwan Institute of Chemical Engineers, 2019. 97: p. 458-
465.

35
Mini-Review

98. Ullah, N., et al., Nanomaterial-based optical chemical sensors for the detection of heavy
metals in water: Recent advances and challenges. TrAC Trends in Analytical Chemistry, 2018.
100: p. 155-166.
99. Rana, M. and P. Chowdhury, L-glutathione capped CdSeS/ZnS quantum dot sensor for the
detection of environmentally hazardous metal ions. Journal of Luminescence, 2019. 206: p.
105-112.
100. Xiao, N., et al., Multifunctional fluorescent sensors for independent detection of multiple
metal ions based on Ag nanoclusters. Sensors and Actuators B: Chemical, 2018. 264: p. 184-
192.
101. Jiang, W., et al., A novel fluorescence “turn off-on” nano-sensor for detecting Cu2+ and
Cysteine in living cells. Journal of Photochemistry and Photobiology A: Chemistry, 2018. 362: p.
14-20.

of
102. Salahandish, R., et al., Nano-biosensor for highly sensitive detection of HER2 positive breast
cancer. Biosensors and Bioelectronics, 2018. 117: p. 104-111.

ro
103. Yang, H., et al., A cancer cell turn-on protein-CuSMn nanoparticle as the sensor of breast
cancer cell and CH3O-PEG-phosphatide. Chinese Chemical Letters, 2018. 29(10): p. 1528-1532.
104. Badr, E.a., et al., Synthesis of anionic chitosan surfactant and application in silver nanoparticles
-p
preparation and corrosion inhibition of steel. International Journal of Biological
Macromolecules, 2020. 157: p. 187-201.
re
105. Bricha, M. and K. El Mabrouk, Effect of surfactants on the degree of dispersion of MWNTs in
ethanol solvent. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2019. 561:
lP

p. 57-69.
106. Naderi, O., et al., Synthesis and characterization of silver nanoparticles in aqueous solutions of
surface active imidazolium-based ionic liquids and traditional surfactants SDS and DTAB.
na

Journal of Molecular Liquids, 2019. 273: p. 645-652.


107. Yu, J., et al., Controlling the dispersion of multi-wall carbon nanotubes in aqueous surfactant
solution. Carbon, 2007. 45(3): p. 618-623.
ur

108. Wang, H., et al., Dispersing Single-Walled Carbon Nanotubes with Surfactants: A Small Angle
Neutron Scattering Study. Nano Letters, 2004. 4(9): p. 1789-1793.
Jo

109. Shaban, S.M., et al., Studying surface and thermodynamic behavior of a new multi-hydroxyl
Gemini cationic surfactant and investigating their performance as corrosion inhibitor and
biocide. Journal of Molecular Liquids, 2020. 316: p. 113881.
110. Shaban, S.M. and D.-H. Kim, The influence of the Gemini surfactants hydrocarbon tail on in-situ
synthesis of silver nanoparticles: Characterization, surface studies and biological performance.
Korean Journal of Chemical Engineering, 2020. 37(6): p. 1008-1019.
111. Grządka, E., J. Matusiak, and E. Godek, Alginic acid as a stabilizer of zirconia suspensions in the
presence of cationic surfactants. Carbohydrate Polymers, 2020. 246: p. 116634.
112. Shaban, S.M., et al., One step green synthesis of hexagonal silver nanoparticles and their
biological activity. Journal of Industrial and Engineering Chemistry, 2014. 20(6): p. 4473-4481.
113. Abd-Elaal, A.A., et al., Studying the corrosion inhibition of some prepared nonionic surfactants
based on 3-(4-hydroxyphenyl) propanoic acid and estimating the influence of silver
nanoparticles on the surface parameters. Journal of Molecular Liquids, 2018. 249: p. 304-317.
114. Shaban, S.M. and A.A. Abd-Elaal, Studying the silver nanoparticles influence on
thermodynamic behavior and antimicrobial activities of novel amide Gemini cationic
surfactants. Materials Science and Engineering: C, 2017. 76: p. 871-885.

36
Mini-Review

115. Shaban, S.M., Studying the effect of newly synthesized cationic surfactant on silver
nanoparticles formation and their biological activity. Journal of Molecular Liquids, 2016. 216:
p. 137-145.
116. Shaban, S.M., et al., Synthesis of newly cationic surfactant based on dimethylaminopropyl
amine and their silver nanoparticles: Characterization; surface activity and biological activity.
Chinese Chemical Letters, 2017. 28(2): p. 264-273.
117. Aiad, I., et al., Antipyrine cationic surfactants capping silver nanoparticles as potent
antimicrobial agents against pathogenic bacteria and fungi. Journal of Molecular Liquids, 2017.
243: p. 572-583.
118. Liu, Q., et al., Spacer-Mediated Synthesis of Size-Controlled Gold Nanoparticles Using Geminis
as Ligands. Langmuir, 2008. 24(5): p. 1595-1599.
119. Pisárčik, M., et al., Silver nanoparticles stabilised with cationic single-chain surfactants.

of
Structure-physical properties-biological activity relationship study. Journal of Molecular
Liquids, 2018. 272: p. 60-72.

ro
120. Shaban, S.M., et al., The Tail Effect of Some Prepared Cationic Surfactants on Silver
Nanoparticle Preparation and Their Surface, Thermodynamic Parameters, and Antimicrobial
Activity. Journal of Surfactants and Detergents, 2019. 22(6): p. 1445-1460.
121.
-p
Xu, F., Q. Zhang, and Z. Gao, Simple one-step synthesis of gold nanoparticles with controlled
size using cationic Gemini surfactants as ligands: Effect of the variations in concentrations and
re
tail lengths. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2013. 417: p.
201-210.
lP

122. Mehta, S.K., S. Kumar, and M. Gradzielski, Growth, stability, optical and photoluminescent
properties of aqueous colloidal ZnS nanoparticles in relation to surfactant molecular structure.
Journal of Colloid and Interface Science, 2011. 360(2): p. 497-507.
na

123. Negm, N.A., et al., Novel Nonionic Polyurethane Surfactants and Ag Nanohybrids: Influence of
Nonionic Polymeric Chains. Journal of Surfactants and Detergents, 2017. 20(1): p. 173-182.
124. El-Dib, F.I., et al., Study the adsorption properties of magnetite nanoparticles in the presence
ur

of different synthesized surfactants for heavy metal ions removal. Egyptian Journal of
Petroleum, 2020. 29(1): p. 1-7.
Jo

125. Mehta, S.K., et al., Effect of Cationic Surfactant Head Groups on Synthesis, Growth and
Agglomeration Behavior of ZnS Nanoparticles. Nanoscale Research Letters, 2009. 4(10): p.
1197.
126. Murphy, C.J., et al., The Many Faces of Gold Nanorods. The Journal of Physical Chemistry
Letters, 2010. 1(19): p. 2867-2875.
127. Gao, J., C.M. Bender, and C.J. Murphy, Dependence of the Gold Nanorod Aspect Ratio on the
Nature of the Directing Surfactant in Aqueous Solution. Langmuir, 2003. 19(21): p. 9065-9070.
128. Murphy, C.J. and N.R. Jana, Controlling the Aspect Ratio of Inorganic Nanorods and Nanowires.
Advanced Materials, 2002. 14(1): p. 80-82.
129. Jana, N.R., L. Gearheart, and C.J. Murphy, Wet chemical synthesis of silver nanorods and
nanowires of controllable aspect ratio. Chemical Communications, 2001(7): p. 617-618.
130. Bakshi, M.S., How Surfactants Control Crystal Growth of Nanomaterials. Crystal Growth &
Design, 2016. 16(2): p. 1104-1133.
131. Sanchez-Gaytan, B.L., et al., Controlling the Topography and Surface Plasmon Resonance of
Gold Nanoshells by a Templated Surfactant-Assisted Seed Growth Method. The Journal of
Physical Chemistry C, 2013. 117(17): p. 8916-8923.

37
Mini-Review

132. Nikoobakht, B. and M.A. El-Sayed, Preparation and Growth Mechanism of Gold Nanorods
(NRs) Using Seed-Mediated Growth Method. Chemistry of Materials, 2003. 15(10): p. 1957-
1962.
133. Zhang, Q. and Y. Yin, Beyond spheres: Murphy's silver nanorods and nanowires. Chemical
Communications, 2013. 49(3): p. 215-217.
134. Guerrero-Martínez, A., et al., Nanostars shine bright for you: Colloidal synthesis, properties
and applications of branched metallic nanoparticles. Current Opinion in Colloid & Interface
Science, 2011. 16(2): p. 118-127.
135. Kou, X., et al., Growth of Gold Nanorods and Bipyramids Using CTEAB Surfactant. The Journal
of Physical Chemistry B, 2006. 110(33): p. 16377-16383.
136. Dadwal, A. and P.A. Joy, Influence of chain length of long-chain fatty acid surfactant on the
thermal conductivity of magnetite nanofluids in a magnetic field. Colloids and Surfaces A:

of
Physicochemical and Engineering Aspects, 2018. 555: p. 525-531.
137. Sabbah, I.A., et al., Synthesis, characterization and antimicrobial activity of colloidal copper

ro
nanoparticles stabilized by cationic thiol polyurethane surfactants. Journal of Polymer
Research, 2018. 25(12): p. 252.
138. Khowdiary, M.M., et al., Synthesis, characterization and biocidal efficiency of quaternary
-p
ammonium polymers silver nanohybrids against sulfate reducing bacteria. Journal of
Molecular Liquids, 2017. 230: p. 163-168.
re
139. Zhou, J., M. Hu, and D. Jing, The synergistic effect between surfactant and nanoparticle on the
viscosity of water-based fluids. Chemical Physics Letters, 2019. 727: p. 1-5.
lP

140. Kumari, R. and D.P. Singh, Ameliorating effect of surfactants against silver nanoparticle toxicity
in crop Fagopyrum esculentum L. Environmental Nanotechnology, Monitoring & Management,
2019. 12: p. 100254.
na

141. Mousavi Ghahfarokhi, S.E. and E. Mohammadzadeh Shobegar, An investigation of the ethylene
glycol surfactant on the structural, microstructure, magnetic and optical properties of SrFe2O4
nanoparticles. Journal of Magnetism and Magnetic Materials, 2020. 495: p. 165866.
ur

142. Hayashi, S. and H.O. Hamaguchi, Discovery of a magnetic ionic liquid [bmim]FeCl4. Chemistry
Letters, 2004. 33(12): p. 1590-1591.
Jo

143. Clark, K.D., et al., Magnetic ionic liquids in analytical chemistry: A review. Anal Chim Acta,
2016. 934: p. 9-21.
144. Mallick, B., et al., Dysprosium room-temperature ionic liquids with strong luminescence and
response to magnetic fields. Angew Chem Int Ed Engl, 2008. 47(40): p. 7635-8.
145. Brown, P., et al., Properties of new magnetic surfactants. Langmuir, 2013. 29(10): p. 3246-51.
146. Brown, P., et al., Magnetic control over liquid surface properties with responsive surfactants.
Angew Chem Int Ed Engl, 2012. 51(10): p. 2414-6.
147. Wang, L., et al., Magnetic Fullerene-DNA/Hyaluronic Acid Nanovehicles with
Magnetism/Reduction Dual-Responsive Triggered Release. Biomacromolecules, 2017. 18(3): p.
1029-1038.
148. Xu, L., et al., Magnetic controlling of migration of DNA and proteins using one-step modified
gold nanoparticles. Chemical Communications, 2015. 51(45): p. 9257-9260.
149. Dai, X., et al., Design and functionalization of magnetic ionic liquids surfactants (MILSs)
containing alkyltrimethylammonium fragment. 2018.

38
Mini-Review

150. Greeson, K.T., et al., Synthesis and properties of symmetrical N, N′-bis (alkyl) imidazolium
bromotrichloroferrate (III) paramagnetic, room temperature ionic liquids with high short-term
thermal stability. 2018.
151. Brown, P., et al., Magnetic surfactants. 2015. 20(3): p. 140-150.
152. Brown, P., et al., Magnetic Surfactants and Polymers with Gadolinium Counterions for Protein
Separations. Langmuir, 2016. 32(3): p. 699-705.
153. de la Fuente-Nunez, C., et al., Magnetic Surfactant Ionic Liquids and Polymers With
Tetrahaloferrate (III) Anions as Antimicrobial Agents With Low Cytotoxicity. Colloid and
Interface Science Communications, 2018. 22: p. 11-13.
154. Ruiz-Hernandez, E., A. Baeza, and M. Vallet-Regi, Smart Drug Delivery through DNA/Magnetic
Nanoparticle Gates. Acs Nano, 2011. 5(2): p. 1259-1266.
155. Li, J., et al., Smart drug delivery nanocarriers with self-assembled DNA nanostructures. Adv

of
Mater, 2013. 25(32): p. 4386-96.
156. Langer, R., Drug delivery and targeting. Nature, 1998. 392(6679 Suppl): p. 5-10.

ro
157. Le Ny, A.-L.M. and C.T.J.J.o.t.A.C.S. Lee, Photoreversible DNA condensation using light-
responsive surfactants. 2006. 128(19): p. 6400-6408.
158. -p
Bhattacharya, S. and S.S. Mandal, Interaction of surfactants with DNA. Role of hydrophobicity
and surface charge on intercalation and DNA melting. Biochim Biophys Acta, 1997. 1323(1): p.
re
29-44.
159. Cassell, A.M., W.A. Scrivens, and J.M. Tour, Assembly of DNA/Fullerene Hybrid Materials.
Angew Chem Int Ed Engl, 1998. 37(11): p. 1528-1531.
lP

160. Dias, R., et al., DNA phase behavior in the presence of oppositely charged surfactants.
Langmuir, 2000. 16(24): p. 9577-9583.
161. Bloomfield, V.A., DNA condensation by multivalent cations. Biopolymers, 1997. 44(3): p. 269-
na

82.
162. Brown, P., et al., Magnetizing DNA and Proteins Using Responsive Surfactants. Advanced
ur

Materials, 2012. 24(46): p. 6244-6247.


163. Xu, L., et al., Compaction and decompaction of DNA dominated by the competition between
counterions and DNA associating with cationic aggregates. Colloids and Surfaces B-
Jo

Biointerfaces, 2015. 134: p. 105-112.


164. Xu, L., et al., Ordered DNA-Surfactant Hybrid Nanospheres Triggered by Magnetic Cationic
Surfactants for Photon- and Magneto-Manipulated Drug Delivery and Release.
Biomacromolecules, 2015. 16(12): p. 4004-4012.
165. Wang, L., et al., Magnetic networks of carbon quantum dots and Ag particles. Journal of
Colloid and Interface Science, 2019. 539: p. 203-213.
166. Xu, L., et al., Carbon nanotubes modified by a paramagnetic cationic surfactant for migration
of DNA and proteins. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2018.
559: p. 201-208.
167. Zhao, W., et al., Co-assemblies of polyoxometalate {Mo72Fe30}/double-tailed magnetic-
surfactant for magnetic-driven anchorage and enrichment of protein. Journal of Colloid and
Interface Science, 2019. 536: p. 88-97.
168. McCoy, T.M., et al., Noncovalent Magnetic Control and Reversible Recovery of Graphene Oxide
Using Iron Oxide and Magnetic Surfactants. Acs Applied Materials & Interfaces, 2015. 7(3): p.
2124-2133.

39
Mini-Review

169. Zhao, G.X., et al., Few-Layered Graphene Oxide Nanosheets As Superior Sorbents for Heavy
Metal Ion Pollution Management. Environmental Science & Technology, 2011. 45(24): p.
10454-10462.
170. Chang, H. and H.J.A.F.M. Wu, Graphene-based nanomaterials: synthesis, properties, and
optical and optoelectronic applications. 2013. 23(16): p. 1984-1997.
171. Kim, J., et al., Graphene Oxide Sheets at Interfaces. Journal of the American Chemical Society,
2010. 132(23): p. 8180-8186.
172. Chandra, V., et al., Water-dispersible magnetite-reduced graphene oxide composites for
arsenic removal. ACS Nano, 2010. 4(7): p. 3979-86.
173. Fu, Y., et al., Water-dispersible magnetic nanoparticle–graphene oxide composites for
selenium removal. 2014. 77: p. 710-721.
174. WooáLee, J. and S.J.N. BináKim, Enhanced Cr (VI) removal using iron nanoparticle decorated

of
graphene. 2011. 3(9): p. 3583-3585.
175. Sreeprasad, T.S., et al., Reduced graphene oxide-metal/metal oxide composites: facile

ro
synthesis and application in water purification. J Hazard Mater, 2011. 186(1): p. 921-31.
176. Sagar, G.H., M.A. Arunagirinathan, and J.R. Bellare, Self-assembled surfactant nano-structures
important in drug delivery: A review. Indian Journal of Experimental Biology, 2007. 45(2): p.

177.
133-159.
-p
Brown, P., T.A. Hatton, and J. Eastoe, Magnetic surfactants. Current Opinion in Colloid &
re
Interface Science, 2015. 20(3): p. 140-150.
lP
na
ur
Jo

40
Mini-Review

Recent advances in surfactants-assisted nanotechnology


Samy M. Shaban1,2,3, Joohoon Kang4,*, and Dong-Hwan Kim1,2,*
1. School of Chemical Engineering, Sungkyunkwan University (SKKU), Suwon 16419,
Republic of Korea
2. Biomedical Institute for Convergence at SKKU (BICS), Sungkyunkwan University
(SKKU), Suwon 16419, Republic of Korea
3. Petrochemicals Department, Egyptian Petroleum Research Institute, Egypt
4. School of Advanced Materials Science and Engineering, Sungkyunkwan University

of
(SKKU), Suwon 16419, Republic of Korea

ro
E-mail: dhkim1@skku.edu (D.-H. K.) and joohoon@skku.edu (J.K.)
-p
Highlights
re
- Surfactant-assisted technological advances in nanotechnology
lP

- Surfactant structure-oriented nanoparticle synthesis and dispersion stability


na

- Emerging category of surfactant with magnetic properties; magneto-


surfactants
ur

- Magneto-surfactants based applications; drug delivery and biomolecules


Jo

separation

1
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like