You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/275167712

Surfactant adsorption and aggregate structure of silica nanoparticles: A


versatile stratagem for the regulation of particle size and surface modification

Article  in  Materials Research Express · March 2014


DOI: 10.1088/2053-1591/1/1/015011

CITATIONS READS

3 421

3 authors:

Savita Chaudhary Deepak Rohilla


Panjab University Panjab University
91 PUBLICATIONS   835 CITATIONS    5 PUBLICATIONS   7 CITATIONS   

SEE PROFILE SEE PROFILE

Surinder K Mehta
Panjab University
266 PUBLICATIONS   4,558 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Publication View project

Luminescent Carbon dots: Fabrication, characterization and application as potential photocatalyst and fluorescent sensor View project

All content following this page was uploaded by Savita Chaudhary on 31 October 2015.

The user has requested enhancement of the downloaded file.


Home Search Collections Journals About Contact us My IOPscience

Surfactant adsorption and aggregate structure of silica nanoparticles: a versatile stratagem for

the regulation of particle size and surface modification

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2014 Mater. Res. Express 1 015011

(http://iopscience.iop.org/2053-1591/1/1/015011)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 14.139.246.28
This content was downloaded on 31/01/2014 at 04:12

Please note that terms and conditions apply.


Surfactant adsorption and aggregate structure of
silica nanoparticles: a versatile stratagem for the
regulation of particle size and surface modification
Savita Chaudhary1, Deepak Rohilla2 and S K Mehta2
1
Centre for Nanoscience & Nanotechnology, University Institute of Emerging area in Science &
Technology, Panjab university, Chandigarh-160014, India
2
Department of Chemistry and Centre of Advanced Studies in Chemistry, Panjab University,
Chandigarh 160014, India
E-mail: schaudhary@pu.ac.in

Received 29 November 2013, revised 29 November 2013


Accepted for publication 17 December 2013
Published 30 January 2014
Materials Research Express 1 (2014) 015011
doi:10.1088/2053-1591/1/1/015011

Abstract
The area of silica nanoparticles is incredibly polygonal. Silica particles have
aroused exceptional deliberation in bio-analysis due to great progress in parti-
cular arenas, for instance, biocompatibility, unique properties of modifiable pore
size and organization, huge facade areas and pore volumes, manageable mor-
phology and amendable surfaces, elevated chemical and thermal stability.
Currently, silica nanoparticles participate in crucial utilities in daily trade
rationales such as power storage, chemical and genetic sensors, groceries dis-
pensation and catalysis. Herein, the size-dependent interfacial relation of anionic
silica nanoparticles with twelve altered categories of cationic surfactants has
been carried out in terms of the physical chemical facets of colloid and interface
science. The current analysis endeavours to investigate the virtual consequences
of different surfactants through the development of the objective composite
materials. The nanoparticle size controls, the surface-to-volume ratio and surface
bend relating to its interaction with surfactant will also be addressed in this
work. More importantly, the simulated stratagem developed in this work can be
lengthened to formulate core–shell nanostructures with functional nanoparticles
encapsulated in silica particles, making this approach valuable and extensively
pertinent for employing sophisticated materials for catalysis and drug delivery.

S Online supplementary data available from stacks.iop.org/MRX/1/015011/


mmedia

Keywords: cationic surfactant, surface-to-volume, size control, catalysis, drug


delivery

Materials Research Express 1 (2014) 015011


2053-1591/14/015011+17$33.00 © 2014 IOP Publishing Ltd
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

1. Introduction

The arena of silica nanoparticles is phenomenally multilateral. Silica particles have stimulated
special consideration in bio-analysis due to their immense improvements in various
characteristics such as biocompatibility, distinctive properties of amendable pore size and
structure, large surface areas (1000 m2 g−1) and pore volumes (near 1.0 cm3 g−1), controllable
morphology and adjustable surfaces, high chemical and thermal steadiness [1–5]. These
characteristics permit the silica nanoparticles to be extensively employed as the solid-supporting
or entrapping matrix. Silica particles play an essential function in everyday commercial
purposes, for instance, energy storage, chemical and biological sensors, food processing and
catalysis [6–10] etc. Their distinctive topology offers three distinctive realms that can be
autonomously functionalized: the silica skeleton, the hexagonal nanochannels/pores, and the
nanoparticle’s exterior corona. By themselves, silica nanoparticles are principally appropriate to
the assignment of integrating the vital competence of a theranostic podium in a solo particle,
with divided provinces for (i) the contrast agent that facilitates perceptible imaging of
theranostic targeting (ii) the drug consignment for therapeutic intrusion and (iii) the
biomolecular ligand for extremely specific targeted delivery [11, 12]. In addition to these
aspects, silica nanoparticles have confirmed high in vivo and in vitro biocompatibility,
unproblematic endocytosis by a wide variety of cell types and biodegradability [13].
Morphologically different silica materials have been created using a broad spectrum of
synthesis protocols, including the well-known Stöber method and other approaches such as soft/
hard-templating approach [14–17]. The hard-template method characteristically comprises the
fabrication of sacrificial hard templates and their surface functionalization pursued by the
deposition of mesoporous silica shell and selective etching of the core templates [16].
Conversely, the soft-template method using micelles and vesicles will lessen the working route
of the after-treatment. In the formation of uniform, morphologically controllable, spherical
silica materials various surfactants have been extensively utilized as capping agents to apply
exquisite control over the nucleation and growth of silica nanoparticles. Until now, a
variety of surfactants with different headgroups, hydrophobic chains, counterions, and
molecular architectures, have been used for the shape-controlled synthesis of silica nanocrystals
[18–22]. However, a few questions such as how surfactants interact with the nanoparticle
surface, what type of structures are formed and the effect of surfactant adsorption on toxicity,
have not been tackled appropriately in literature. From the viewpoint of both elementary
research and realistic relevancies of surfactants in silica nanoparticle synthesis, there is a
significant need for a fundamental understanding of the surfactant adsorption on silica
nanoparticles. Moreover, knowledge of the surface charge of the surfactant customized silica
nanoparticles and the structure of the adsorbed layer is vital to determine the stability of silica
nanoparticles.
Herein, the size-dependent interaction of anionic silica nanoparticles with twelve different
types of cationic surfactants has been studied in terms of the physical chemical aspects of
colloid and interface science. The surfactants used consist of the same charges i.e. cationic on
the head group and different hydrophobic tails as well as the different counterions (Scheme A,
supplementary data stacks.iop.org/MRX/1/015011/mmedia).
The present investigation aims to explore the comparative effect of different surfactants
during the formation of the target composite materials. The nanoparticle size controls, the
surface-to-volume ratio and surface curvature with respect to its interaction with surfactant will

2
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

also be discussed in this work. By investigating the roles of different surfactants, we proposed
the stabilization mechanism for the formation of silica nanoparticles and revealed the
importance of surfactants in stabilizing mesostructures. More importantly, the synthetic strategy
developed in this work will be extended to fabricate core–shell nanostructures with functional
nanoparticles encapsulated in silica, and making the strategy useful and widely applicable for
preparing advanced mesoporous materials for catalysis and drug delivery.

2. Experimental

2.1. Preparation

2.1.1. Chemicals. The cationic surfactants, dodecyl trimethyl ammonium bromide (DTMAB),
dodecyl trimethyl ammonium chloride (DTMAC), dodecyl ethyl dimethyl ammonium bromide
(DEDAB), hexadecyltrimethylammonium bromide (CTAB), hexadecyltrimethylammonium
chloride (CTAC), cetyl pyridinium chloride (CPC), cetyl pyridinium bromide (CPB), didodecyl
dimethyl ammonium bromide (DDDMAB), dimethyl ditetradecyl ammonium bromide
(DMDTDAB), dihexa decyl dimethyl amonium bromide (DHDDMAB), dimethyl
dioctadecyl ammonium bromide (DMDOAB), dimethyl dioctadecyl ammonium chloride
(DMDOAC) were purchased from Sigma Aldrich with purity >99%. Tetraethoxysilane (TEOS)
and aqueous ammonia were procured from Merck. Absolute ethanol was obtained from
Changshu Yangyuan chemical China with purity >99%. The syntheses of particles were carried
out in triply distilled water having conductivity lower than 3 μS.

2.1.2. Synthesis of silica nanoparticles. Surfactant modified silica nanoparticles of different


particle sizes were prepared by means of a tailored Stöber process [23]. In a distinctive
synthesis, 0.022 mol of TEOS were quickly added into a mixture of surfactant aqueous solution
(7 mM), ethanol (1.3 mol), deionized water (50 mL), and ammonium aqueous solution
(0.25 mol). The blend of the reactants was then subsequently stirred at room temperature for
1 h, resulting in the creation of a white silica colloidal suspension. The solution was set aside
under medium stirring for 2 h; and endorsed to place for 24 h to form homogeneous gel. The
silica particles were then centrifugally detached from the dispersion and sluiced with deionized
water and ethanol and left for parching overnight at 90 °C. The dried material thus obtained was
termed as uncalcined SiO2. The calcination of SiO2 has been performed with muffle furnace of
AICIL at the rate of 1 °C min−1 from room temperature to 550 °C and held for 3 h at 550 °C.

2.2. Methods

2.2.1. X-ray diffraction (XRD). The structural characterization of silica nanoparticles was
performed with a Panalytical, D/Max-2500 x-ray Diffractometer equipped with Cu–Kα
radiation (λ = 1.5418 Å). Data were collected in a step scan mode between 2θ values 10° to 70°,
with a step size of 0.02° s−1. The size of the particles was calculated using the Scherrer equation
[24]
αλ
D= . (1)
βCos θ

3
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

where D is crystallite diameter, λ is the x-ray wavelength (in nanometre), β is full width at half
maxima), and θ the Bragg angle (in degrees 2θ), value of α is 0.94.

2.2.2. Thermo gravimetric analysis. The synchronized TG–DTG plots were attained through
thermal analysis method; model SDT Q-600, from TA Instruments. The purge gas was nitrogen
flow of 100.0 ml min−1. A heating pace of 10 °C min−1 was accepted, with samples weighing
about 10.0 mg. This procedure consists of heating the sample to a specified temperature at a
predetermined heating rate (β). Alumina crucibles were employed for verifying the curves.

2.2.3. Differential scanning calorimetry (DSC). DSC measurements were performed with a
DSC TA Q20 instrument equipped with a refrigerated cooling system (TA Instruments, New
Castle, DE). Nitrogen with a flow rate of 50 ml min−1 was used as purge gas. Approximately
6–11 mg of sample was weighed precisely into hermetically sealable aluminum pans. An empty
hermetically sealed pan was used as a reference.

2.2.4. FTIR spectroscopy.


The capping ability of the surfactants was qualitatively evaluated
with the help of a Perkin-Elmer (RX1) FTIR spectrometer in the frequency range
4400–350 cm−1.

2.2.5. Scanning electron microscopy (SEM) and EDX analysis. A JEOL (JSM-6610) scanning
electron microscope (SEM) equipped with EDX analysizer operating at 20 kV was used to
establish the surface morphology as well as the elemental analysis of the powdered silica
nanoparticles fabricated from different surfactants.

2.2.6. Transmission electron microscopy (TEM). Samples were prepared by drying a drop of a
dilute dispersion of the synthesized silica nanoparticles using different cationic surfactants onto
a carbon-coated grid and analyzed using a Hitachi H 7500 electron microscope operating at
80 kV.

2.2.7. Particle size analysis.


Particle size estimations (90° optics) were carried out with a
Nano S 90 (Red badge) Malvern Instruments Corporation model no. ZEN 1690.

2.2.8. CHN analysis. A Perkin Elmer 2400 CHN Elemental Analyser instrument has been
used to determine the CHN Analyzer percentages of C, H and N with an accuracy of 0.3%.

2.2.9. Zeta analysis. The zeta potential of the nanoparticles was determined from their
electrophoretic mobilities according to Smoluchowski’s approximation using a Malvern
Zetasizer NanoZS (Zen 3600) at 25 °C using a folded capillary cell (DTS 1060).

2.2.10. Refractive index measurements. The refractive index for silica nanoparticles modified
with different surfactants has been estimated using an Anton Paar Abbemat 500 refractometer
with accuracy of ±0.000 02 nD.

4
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Figure 1. SEM images of the silica nanoparticles (0.022 M) stabilized in aqueous


solutions of different surfactants (0.007 M).

3. Results and Discussion

3.1. Structure and morphology of silica nanoparticles

The sizes and morphologies of the silica nanoparticles (0.022 M) stabilized in different
surfactants (0.007 M) have been measured using SEM (figure 1, S1). Almost mono-diffused
spherical and well-separated silica nanoparticles are apparent in all the cationic surfactants,
although in the case of CPC and DDDMAB stabilized silica nanoparticles, some distorted or
agglomerated pattern has been created. The aqueous distributions of the SiO2 nanoparticles in
CPC and DDDMAB surfactants were uniformly homogeneous, as compared to other
surfactants, and therefore these prototypes of the nanoparticles are believed to be bred during
the sample aeration procedure due to the diverse natures of the surfactants. A slight deformation
from the spherical profile in the case of SiO2 nanoparticles amalgamated in DMDOAB may be
attributed to the weak stabilizing propensity of DMDOAB. The results of sizes and shapes
variation of the silica nanoparticles have been found to be strongly surfactant dependent. The
morphology of the particles in different surfactant media has also been verified using TEM
analysis (supplementary figure S2). The uniform capping of the surfactants has also been clearly
visualized. In addition, the elemental analysis of the samples obtained has been depicted in the
supplementary data (figure S3).
The size distributions of the aqueous silica nanoparticles have also been judged from DLS
analysis (figure S4). The bimodal allocations with one main and one minor peak have been
scrutinized in all the cases. The strength and disparity of each peak appreciably depends upon

5
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Figure 2. XRD patterns of the uncalcined powdered silica nanoparticles recovered from
aqueous solutions of C16 chained surfactants.

the surfactant type. Figure S4 clearly specifies the existence of a few larger cluster formations,
although the majority of the nanoparticles are in a nanometer range in all the surfactants. The
average particle sizes, analogous to the upper limit of the most concentrated peak in the DLS
size distribution contours, have been given in figure S4. This procedure quantifies the
hydrodynamic radius of the nanodispersions which comprises the size of the adsorbed
surfactant layer/micelles and water of hydration as well.

3.2. X-ray powder diffraction (XRD)


Wide angle XRD (WAXRD) patterns for the uncalcined and calcined silica nanoparticles
obtained after solvent evaporation from C16 chained surfactants have been depicted in figures 2
and 3. The broad XRD peaks at the 2θ angles of around 22° indicated that the silica
nanoparticles were amorphous with 〈100〉 planes. XRD spectra for other surfactant modified
particles are depicted in figures S5 and S6 for uncalcined and calcined samples. Significantly, it
is evident from figures S5 and S6 that all the XRD patterns of uncalcined and calcined silica
particles appear to be approximately analogous, irrespective of the surfactants employed in the
synthesis. This behaviour apparently signifies that all the surfactants give nanoparticles with
identical average size in powder form and that the structure dependent propensities of the
surfactants towards stabilization of the nanoparticles are considerable only in aqueous solutions.
The degree of crystallinity (K) of silica nanoparticles has also been calculated using the
following equation for different surfactant modified particles [25]
AI
D= x100. (2)
AT
where AI is the area of the broad diffraction peak and AT is the total area of diffractogram. The
obtained average crystallite sizes and K have been tabulated in table S1 in the supplementary
data. It has been anticipated that the enhanced value of K corresponds to the better chain

6
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Figure 3. XRD patterns of the calcined powdered silica nanoparticles recovered from
aqueous solutions of C16 chained surfactants.

ordering of surfactant over silica nanoparticle surface. The sizes of the nanoparticles, as
acquired from the powder XRD patterns, show some variation from the size obtained by DLS
analysis. This discrepancy of sizes computed via altered procedures is vindicated as the XRD
line augmentation does not take into account the various other involvements e.g. lattice defects,
dislocations, faulting and lattice strain.

3.3. FTIR analysis


FTIR analyses have been carried out for surfactant stabilized silica nanoparticles to understand
the interatomic bonding of surfactant moieties with silica nanoparticles. Figure 4 shows the
FTIR spectra of the uncalcined silica nanoparticles obtained after functionalization with
different cationic surfactants. The spectrum for the uncalcined particles displays a group of
strong intense peaks at around 3500, 2800, 1700, and 1500 cm−1 as well as a collection of
bands in the section below 1400 cm−1. Table 1 catalogues the key facets recognized in the IR
spectra for silica nanoparticles modified with different cationic surfactants. The bands at around
3500 and 1600 cm−1 are allocated for the stretching mode and bending mode of adsorbed water
molecules. The strong sp3 hybridised C-H stretching peaks appeared between 2700 and
3000 cm−1, which come from all the bound template species. The bending mode for the
bounded surfactant moieties appeared between 1400–1500 cm−1. The band at 1093 cm−1 is
allotted to ѵas(-Si-O-Si), the weak band at 809 cm−1 is assigned to ѵs (Si-O-Si), and the band at
463 cm−1 is assigned to δ(Si-O-Si) [26–28].
The stretching mode of alkyl groups from different surfactants can further be quantitatively
analysed to access the amount of surfactant bound to the surface of silica nanoparticle. Since the
sp3 C-H peaks come from all these surfactants, the integrated area between 2800–3000 cm−1
therefore allows us to estimate the relative amount of surfactants bound to the nanoparticle
surface [29, 30]. Comparing the change in the C-H area for different surfactants over

7
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Figure 4. FTIR spectra of surfactant modified silica nanoparticles.

Table 1. Mode assignment of different IR peaks of various surfactants in the presence of


silica nanoparticles.
δ
νs(C–H) ѵas(- (Si-
νsym (C–H) νasym(C–H) methylene Si-O- ѵs(Si- O-
Surfactant stretching stretching bending Si) O-Si) Si)
DDTMAB 2926 2855 1490 1076 796 463
DDTMAC 2927 2856 1479 1069 793 466
DEDAB 2955 2851 1488 1063 794 462
CTAB 2919 2850 1487 1070 796 462
CTAC 2915 2851 1472 1066 777 457
CPC 2924 2852 1488 1071 778 456
CPB 2918 2850 1470 1072 779 457
DDAB 2956 2853 1468 1068 793 468
DMDTDAB 2956 2852 1470 1063 796 467
DDDMAB 2955 2851 1488 1069 793 456
DMDOAB 2955 2850 1471 1055 794 453
DMDOAC 2919 2851 1470 1069 798 463

8
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Figure 5. Integrated peak areas of the sp3 C-H peak for surfactant modified silica
nanoparticles.

nanoparticles allows us to estimate how the surface atoms affect native surfactant binding
(figure 5).
On comparing the C12 chain carbon surfactants, the variation of the integrated peak area
shows the order DEDAB > DDTMAC > DDTMAB. The C-H peak increases by ~66% in the
case of DEDAB. Concomitantly, the DDTMAB peak decreases by ~18% in comparison to
DDTMAC. This implies that there is more DEDAB surfactant bound over the silica sample as
compared to DDTMAC and DDTMAB. The case of DDTMAC and DDTMAB is also
comparable where the former reveals considerably larger value of integrated peak area in
contrast with the latter. This variation is mainly explained on the basis of the counterion
polarizibility. The Cl− ion possesses lower polarizibility and has a higher tendency to stabilize
the nanoparticles as compared to Br− ion. Moreover, the effect of counterion size may also play
a role in stabilization of nanoparticles. However, in case of C-16 chain length surfactants, i.e.
CTAB, CTAC, CPC and CPB, the variation of the integrated peak area follows:
CPC > CPB > CTAC > CTAB. The variation can be explained by the presence of pyridinium
ring CPC and CPB in comparison with CTAB and CTAC. This pyridinium ring of CPC and
CPB can act as a better Lewis acid (as compared to the trimethylammonium head group of
CTAB and CTAC) due to resonance on the ring, and can stabilize the electrons to be transferred
from the nanoparticle surface more efficiently. Comparing the integrated peak area variation of
C-12 and C-16 surfactants, it is observed that the former possess higher peak area, which can be
explained on the basis of bending rigidity of the surfactant layer over the nanoparticles. The
longer hydrophobic chains may lead to more rigid interface, and result in slower growth rate.
Therefore, lower area per molecule adopted by the lower chain surfactant leads to higher
packing efficiency over nanoparticle surface. Moreover, the shorter chain length surfactants
ensure stable dispersion of particles and reduce the polydisperse character of the particles. The
effect of dimeric surfactants, i.e. DDAB, DMDTDAB, has also been assessed for the integrated
peak area values. These cationic dimeric surfactants act as a good structure directing agents for
the synthesis of mesoporous nanoparticles. The C-H peak area increases by ~72% for DDAB in
comparison to DDTMAB. This mainly arises due to the high propensity of dimeric surfactants
to assemble over nanoparticle surface. In addition, the shorter spacer makes the distance

9
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Figure 6. DSC curves for the calcined and uncalcined powdered silica nanoparticles
recovered from aqueous solutions of C16 chained surfactants.

between the head groups (caused by the repulsion among the cationic heads) decrease.
Secondly, this also increases the van der Waals interactions between their alkyl chains. This
implies that there are more DDAB monomers bound over the nanoparticle surface. The
behaviour of higher chain length dimeric surfactants has also been assessed by studying their
integrated peak area variation. The C-H peak area decreases as the chain length for dimeric
surfactant increases from DDAB to DMDOAB. This variation is explained on the basis of their
relative microviscosity behaviour. The microviscosity increases with increase in chain length,
therefore, the bending rigidity of the surfactant is also affected relative to increase in chain
length. The behaviour of counter ion remains the same in dimeric surfactants as in the case of
single chained surfactant.

3.4. DSC data for surfactant adsorption on SiO2 nanoparticles

Silica exterior surface in aqueous milieu turns out to be negatively charged [31] once pH value
is larger than 2, which crafts the ionic acquaintance with cationic surfactants achievable. It is
clear that in aqueous media, the close association of surfactants over silica nanoparticles results
in the formation of monolayers or multilayers of surfactant over the particles [32]. However, the
comparative role of chain length of different surfactants, the effect of counterions as well as the
effect of different head groups of the surfactants on the thermal properties, especially with
reference to melting and crystallization transitions processes of surfactants with nanoparticles,
has not been performed. Here in the present manuscript, we consider the adsorption mechanism
of different surfactants over the silica surface using DSC analysis. For comparison sake, the
weight dependent heating profile of calcined, uncalined silica nanoparticles synthesized with
different surfactants was taken under identical conditions by keeping concentrations of
surfactant (7 × 10−3 M) as well as that of TEOS (2.2 × 10−4 M) the same in each system.
Figure 6 illustrates the thermograms for melting and crystallization transitions for calcined and
uncalcined silica nanoparticles with different surfactants.
Additionally, the thermograms for other calcined and uncalcined silica particles with
various other surfactants are demonstrated in figures S7, S8 and S9 and that of pure surfactants

10
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

are depicted in figure S8. The vertical scales of the thermograms were adjusted to clearly show
the temperature of alterations. On comparing the heating scans (figure S7) C12 chain carbon
surfactants exhibited major endothermic peaks at around 100.6 °C and 280.6 °C for DDTMAB,
136.8 °C and 298.1 °C for DDTMAC and 151.9 °C and 246.3 °C for DEDAB, respectively.
Moreover, the neat DEDAB surfactant also exhibits pre-transition peaks around 22 and 49 °C,
in addition to the main transition. These pre- and main transitions are allied to the transitions
from the gel phase to the rippled gel phase and from this to the liquid-crystalline (LC) phase,
correspondingly on escalating the temperature [33]. For surfactant modified silica nanoparticles,
the melting transitions for the DDTMAB tail showed a peak at 87.3 °C and 232.9 °C, 106 °C
and 249 °C for DDTMAC and 89.3 °C and 235.3 °C for DEDAB. Similar behaviour has been
illustrated by C16 and dimeric surfactants in presence of silica nanoparticles (table S2). These
absorption-induced reductions in the transition temperature and changes in the transition
enthalpy of surfactants have been explained via the restricted degree of freedom of the
surfactant tail pinned to a solid surface. As a result, entropy of the surfactant decreases, rather
that the enthalpy upon adsorption over the nanoparticles. The width of the transition peak has
also been enlarged when surfactants were adsorbed over the silica nanoparticles. The increase in
the width of the transition peaks can be due to the conformational heterogeneity in the
surfactant-bound silica nanoparticle.
The thermal effect of the counterions, i.e. Cl− or Br− ions, is also studied where the former
exhibits larger value of Tm in contrast with the latter. The scaling description of thermal
behavioural changes can be related to the (i) radius of the hydrated counterion and (ii) their
polarizability value. Since Br− and Cl− have the same hydrated radius (2.06 and 1.82 Å) but
differing polarizability (4.27 and 3.01 Å3) [34]. Therefore, the polarizability is probably
explaining the higher Tm for surfactants with Cl− as a counterion as compared to Br−.
The enthalpy and entropy variations during melting of surfactant tail were also quantified
for pure surfactant as well as in the existence of nanoparticles from the area under the transition
curve (table S2). From the statistical data, it has been scrutinized that both the transition
enthalpy and entropy are approximately linearly dependent on the chain length. Moreover, the
divergence in alkyl chain length appreciably persuades surfactant adsorption at the solid
interface, with small chain surfactants attaching vertically to the surface and long chain
surfactants anchoring at an angle, thus partially exposing the hydrophobic chains of the
surfactant to the surrounding environment. This diversity in arrangement of surfactants at the
nanoparticle surface may be one of the reasons for creating differences in the thermal behaviour
of surfactants.
The effect of different surfactants on the glass transition temperature (Tg) and heat capacity
(ΔCp) values have also been scrutinized (table S3). It has been observed from the data that ΔCp
values for surfactants adsorbed on silica nanoparticles are significantly less than the ΔCp for the
neat surfactants. These results have been qualified from the viewpoint of restricted degree of
freedom of the surfactant monomers over the nanoparticle surface. Moreover, the variation in
the ΔCp and Tg over the nanoparticle surface is also associated with the sizes of the
nanoparticles. Therefore, under comparable concentration of surfactant ΔCp value is higher and
Tg is smaller for small-sized nanoparticle. These variations further indicate that there is a
smaller number of surfactant head group contacts for the smaller size nanoparticle. The
difference in the Tg arises due to the smaller number of contacts of surfactant and more
flexibility on the smaller sized nanoparticles. Further, the variation in ΔCp is associated with the
increased degree of freedom for surfactant over small-sized silica nanoparticles.

11
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Figure 7. TGA curves for the uncalcined powdered silica nanoparticles recovered from
aqueous solutions of C16 chained surfactants.

3.5. Thermogravimetric analysis of SiO2 nanoparticles

In order to investigate the underlying adsorption and stabilization mechanisms, thermogravi-


metric analysis (TGA) analysis has also been carried out. Figure S10 shows the mass loss
curves for pure surfactants. The start and onset decomposition temperature of these surfactants
has also been determined by TGA analysis. The variation of Tstart and Tonset is further related to
the stability of the surfactant. It has been found that DDTMAC is thermally more stable with
Tstart of 203 °C and Tonset of 239.9 °C, whereas DMDOAC shows the Tstart value of 170 °C and
Tonset of 194.9 °C. The Tstart and Tonset temperature of all the other surfactants under
consideration lie between these two extreme values (table S4). The decrease in these
temperature values is further associated with the increase in alkyl chain length of the
surfactants. The heating effect of the counterions i.e. Cl− or Br− ions is also studied using TGA
analysis. It has been found that Cl− ions reveal larger value of Tstart and Tonset in association with
the Br− ions. The extent of this thermal behavioural transformation can be associated to the
higher polarizability of Cl− ions in contrast to the Br− ions. The TGA analysis also provides
comprehensive information for viewing the quantity of energy necessitated to eradicate
surfactant molecules from the nanoparticle exterior facade. This energy depends on whether the
surfactant is present as a mono- or bi-layer on nanoparticle surface. In consequence, the
temperature at which meticulous mass losses take place can be interrelated to surfactant
organization on nanoparticle surface. Figure 7 represents the thermographs for uncalcined silica
nanoparticles synthesized with C16 chained surfactants.
Additionally, the thermograms for other calcined and uncalcined silica particles with
various surfactants are demonstrated in figures S11 and S12. The spectra for all the different
surfactant modified silica particles show mass loss in a step-wise manner. It has also been
observed that each step limiting the mass loss is very sensitive for the type of surfactant
employed for the synthesis of silica nanoparticles (table S4). Based on the detailed thermal

12
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Table 2. Thermal characterization surfactants in presence of silica nanoparticles.


Decomposition Weight no. of
Temperature range loss Graft density groups χ
Surfactant (°C) (wt%) δ (mmol g−1) (per nm2)
DDTMAB 31–154 6.4 0.133 2.63
155–293 24.9
294–528 3.4
529–1000 6.2
DDTMAC 22–182 9.4 0.144 2.99
183.306 18.5
307–383 2.6
384–1000 7.5
DEDAB 30–175 9.2 0.124 2.74
176–309 21.9
310–532 4.0
533–1000 4.9
CTAB 35–186 8.7 0.098 1.81
187–317 17.2
318–525 3.7
526–1000 6.3
CTAC 35–206 4.6 0.108 2.01
207–304 18.9
305–573 8.7
574–1000 4.56
CPC 28–198 6.4 0.142 2.94
199–349 24.9
350–505 3.4
506–1000 6.2
CPB 28–193 9.4 0.141 2.90
194–304 18.5
305–496 2.6
497–1000 7.5
DDAB 30–137 9.2 0.129 2.71
138–281 21.9
282–352 4.0
353–1000 4.9
DMDTDAB 30–130 8.7 0.117 2.63
131–306 17.2
307–378 3.7
379–1000 6.3
DDDMAB 28–186 4.6 0.146 3.34
187–234 18.9
235–337 8.7
338–1000 4.56
DMDOAB 28–87 6.4 0.114 2.84
88–288 24.9
289–526 3.4
527–1000 6.2

13
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Table 2. (Continued ).
Decomposition Weight no. of
Temperature range loss Graft density groups χ
Surfactant (°C) (wt%) δ (mmol g−1) (per nm2)
DMDOAC 28–139 9.4 0.113 2.94
139–593 18.5
594–1000 2.6
7.5

analysis studies, first weight loss step, up to 200 °C is attributed to the loss of adsorbed and
interlayer water [35], whereas the subsequent two major weight losses occur at 250 to 600 °C
due to the removal of template surfactant. For a single layer of surfactant-coated particles [36],
the mass profile exhibits a well-defined decrease over a temperature range of 250–600 °C,
whereas for all bi-layer surfactant-coated particles, the heating curves are characterized by mass
losses with two distinctive strides, disclosing an altered pattern from that observed for
monolayer surfactant-coated particles. These weight-loss steps are further attributed to different
states, i.e. bi-layer formation of surfactants over silica nanoparticles. Table 2 shows the
corresponding weight losses for silica nanoparticles modified with different surfactants.
To assess the surfactant packing concentration, the mass losses were allocated to certain
fractions of the surfactant coatings. For the monolayer surfactant-coated particles, the weight
losses were due to desorption and successive disappearance of the primary surfactant from the
particle surface. For bi-layer surfactant-coated particles, the weight losses are endorsed to
quantitative removal of the outer (secondary surfactant) and inner (primary surfactant) layer,
respectively. This point has further been verified by comparing the TGA curves for C12 chain
carbon surfactants. DDTMAB, DDTMAC and DEDAB show sharp steps of 26%, 19% and
24% weight loss within 145–311 °C, respectively, which corresponds to the decomposition of
outer layer of surfactant. While the subsequent loss of 6%, 4% and 4% for DDTMAB,
DDTMAC and DEDAB in 310–582 °C is attributed to inner monolayer decomposition of
surfactant. Similar behaviour has been demonstrated by C16 and dimeric surfactants in the
presence of silica nanoparticles. Moreover, all the surfactant bound silica nanoparticles display
higher mass loss peak in comparison to pure surfactant. This thermal behavioural change
signifies the strapping bonding of the surfactant to the silica nanoparticle surface and is mainly
possible due to the electrostatic binding of cationic head group to the electronegative silica
particle surface (first layer of surfactant). The first layer of the surfactant over nanoparticle
surface acts as a template for the hydrophobic–hydrophobic inter-digitations of the second
layer. Additionally, the proportion of the molecules in inner and outer layer is thus
approximately calculated for every surfactant by evaluating the number of weight losses from
these two steps. The ratio of molecules in inner and outer layer is roughly estimated as 1:7 for
C12 chain length surfactant, whereas as the chain length increases, the ratio shows significance
changes. For C16 the ratio comes out to be 1:5. On the other hand CPC and CPB surfactants
also show variation in the ratio as compared to CTAB and CTAC. The variation arises due the
presence of pyridinium ring in the head group of CPC and CPB. On this basis, the capping
model of surfactant shell on silica nanoparticle has also been established (scheme 1).
The drafted density of surfactants have also been obtained using equation (3) from the
residual weight fraction [37, 38] after heating at 800 °C by assuming that the remainder consist

14
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

Scheme 1. Schematic structure of adsorbed surfactant on silica nanoparticles.

of silica particles only (table 2)


ΔW
Grafted density(mmol g−1) = , (3)
Mgroup
where ΔW is weight loss of sample, Mgroup is the molecular weight of the capping agent
δ AN
no of groups χ (per nm2) = . (4)
S
Surfactant groups nm−2 (χ) (equation (4)) [39, 40] has been calculated from Avogadro’s
number (AN), the specific surface area (S), and the average particle radius (r).
These outcomes have been proficient from the viewpoint of restricted degree of freedom of
the surfactant monomers over the nanoparticle surface which restricts the growth of the
nanoparticles. To the best of our knowledge, this is the first time that detailed thermal behaviour
of silica nanoparticles has been studied in surfactant media of different chain length.

4. Conclusions

In the present study, the association structures of twelve different cationic surfactants have been
carried out in the presence of silica nanoparticles. The comparative role of chain length of
different surfactants, the effect of counterions as well as the effect of different head groups of
the surfactants on the thermal properties, especially with reference to melting and crystallization
transitions processes of surfactants with nanoparticles has also been performed. The one-step
synthetic process is straightforward, reliable and holds assurance for the synchronized
stabilization of the silica nanoparticles and amendment of their surface properties. Detailed
FTIR, DSC and TGA analysis has been carried out to develop an understanding of the
interatomic bonding of surfactant moieties with silica nanoparticles. The thermal behavioural
change signifies the strapping bonding of the surfactant to the silica nanoparticle surface with
little disorganised initial monolayer and somewhat organized second layer of surfactant. The
effect of different surfactants on the glass transition temperature (Tg) and heat capacity (ΔCp)
values have also been scrutinized. These results have been qualified from the viewpoint of
restricted degree of freedom of the surfactant monomers over the nanoparticle surface. Through
this manuscript, we present a simple model to account for surfactant association in the presence
of nanoparticles. Hence, this study can contribute to an improved perception of the factors
controlling the self-assembly of surfactants at nanoparticles. This will be valuable in the

15
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

formulation of nanoparticle dispersions and their application in particle nanotechnology.


Furthermore, the referred nanoparticles can be prepared in small or large amounts through this
process and used as is or applied to other supports. This unlocks huge prospects for its
advantages for various surface science applications.

Acknowledgements

Savita Chaudhary is grateful to DST for the Inspire faculty award and to UGC for the Start up
grant, and S K Mehta is grateful to DST for financial assistance.

References

[1] Kresge C T, Leonowicz M E, Roth W J, Vartuli J C and Beck J S 1992 Ordered mesoporous molecular sieves
synthesized by a liquid-crystal template mechanism Nature 359 710–2
[2] Beck J S et al 1992 A new family of mesoporous molecular sieves prepared with liquid crystal templates
J. Am. Chem. Soc. 114 10834–43
[3] Sonwane C G and Bhatia S K 1999 Structural characterization of MCM-41 over a wide range of length scales
Langmuir 15 2809–16
[4] Cassiers K et al 2002 A detailed study of thermal, hydrothermal, and mechanical stabilities of a wide range of
surfactant assembled mesoporous silicas Chem. Mater. 14 2317–24
[5] Taguchi A and Schüth F 2005 Ordered mesoporous materials in catalysis Micropor. Mesopor. Mater. 77
1–45
[6] Naik B, Prasad V S and Ghosh N N 2010 A simple aqueous solution based chemical methodology for
synthesis of Ag nanoparticles dispersed on mesoporous silicate matrix Powder Technol. 199 197–201
[7] Fukuoka A, Kimura J, Oshio T, Sakamoto Y and Ichikawa M 2007 Preferential oxidation of carbon
monoxide catalyzed by platinum nanoparticles in mesoporous silica J. Am. Chem. Soc. 129 10120–5
[8] Kumai Y, Tsukada H, Akimoto Y, Sugimoto N, Seno Y, Fukuoka A, Ichikawa M and Inagaki S 2006 Highly
ordered platinum nanodot arrays with cubic symmetry in mesoporous thin films Adv. Mater. 18 760–2
[9] Sarkar B, Venugopal V, Tsianou M and Alexandridis P 2013 Adsorption of pluronic block copolymers on
silica nanoparticles Colloid Surf. A 422 155–64
[10] He Q J and Shi J L 2011 Mesoporous silica nanoparticle based nano drug delivery systems: synthesis,
controlled drug release and delivery, pharmacokinetics and biocompatibility J. Mater. Chem. 21 5845–55
[11] Mortera R, Vivero-Escoto J, Slowing L, Garrone E, Onida B and Lin V S Y 2009 Cell-induced intracellular
controlled release of membrane impermeable cysteine from a mesoporous silica nanoparticle-based drug
delivery system Chem. Commun. 22 3219–21
[12] Yang P P, Quan Z W, Hou Z Y, Li C X, Kang X J, Cheng Z Y and Lin J 2009 A magnetic, luminescent and
mesoporous core–shell structured composite material as drug carrier Biomaterials 30 4786–95
[13] Lu J, Choi E, Tamanoi F and Zink J I 2008 Light-activated nanoimpeller-controlled drug release in cancer
cells Small 4 421–6
[14] Grün M, Unger K K, Matsumoto A and Tsutsumi K 1999 Novel pathways for the preparation of mesoporous
MCM-41 materials: control of porosity and morphology Micropor. Mesopor. Mater. 27 207–16
[15] Matsumoto A, Chen H, Tutsumi K, Grün M and Unger K 1999 Novel route in the synthesis of MCM-41
containing framework aluminium and its characterization Micropor. Mesopor. Mater. 32 55–62
[16] Ribeiro Carrott M M L, Conceicao F L, Lopes J M, Carrott P J M, Bernardes C, Rocha J and
Ramoa Ribeiro F 2006 Comparative study of Al-MCM materials prepared at room temperature with
different aluminium sources and by some hydrothermal methods Micropor. Mesopor. Mater. 92 270–85
[17] Zheng Y, Li Z, Zheng Y, Shen X and Lin L 2006 Synthesis and characterization of Fe–Ce–MCM-41 Mater.
Lett. 60 3221–3

16
Mater. Res. Express 1 (2014) 015011 S Chaudhary et al

[18] Fowler C E, Khushalani D and Mann S 2001 Interfacial synthesis of hollow microspheres of mesostructured
silica Chem. Commun. 19 2028–9
[19] Wang J G et al 2006 Budded, mesoporous silica hollow spheres: hierarchical structure controlled by kinetic
self-assembly Adv. Mater. 18 3284–8
[20] Yu M H, Wang H N, Zhou X F, Yuan P and Yu C Z 2007 One template synthesis of raspberry-like
hierarchical siliceous hollow spheres J. Am. Chem. Soc. 129 14576–7
[21] Liu J, Hartono S B, Jin Y G, Li Z, Lu G Q and Qiao S Z 2010 A facile vesicle template route to multi-shelled
mesoporous silica hollow nanospheres J. Mater. Chem. 20 4595–601
[22] Wu X J and Xu D S 2010 Soft template synthesis of yolk/silica shell particles Adv. Mater. 22 1516–20
[23] Grün M, Lauer I and Unger K K 1997 The synthesis of micrometer- and submicrometer-size spheres of
ordered mesoporous oxide MCM-41 Adv. Mat. 9 254–7
[24] Warren B E 1969 X-ray Diffraction (New York: Addison-Wesley) chapter 13
[25] Hussain A M P, Kumar A, Saikia D, Singh F and Avasthi D K 2005 Study of 160 MeV Ni12+ ion irradiation
effects on electrodeposited polypyrrole films Nuclear Inst. Methods: B 240 871–80
[26] Kadgaonkar M D, Laha S C, Pandey R K, Kumar P, Mirajkar S P and Kumar R 2004 Cerium-containing
MCM-41 materials as selective acylation and alkylation catalysts Catal. Today 97 225–31
[27] Laha S C, Mukherjee P, Sainkar S R and Kumar R 2002 Cerium containing MCM-41-type mesoporous
materials and their acidic and redox catalytic properties J. Catal. 207 213–23
[28] Yao W, Chen Y, Min L, Fang H, Yan Z, Wang H and Wang J 2006 Liquid oxidation of cyclohexane to
cyclohexanol over cerium-doped MCM-41 J. Mol. Catal. A: Chem. 246 162–6
[29] Weer J G and Scheuing D R 1991 Structure/performance relationships in monoalkyl/dialkyl cationic
surfactant mixtures J. Colloid Interface Sci. 145 563–80
[30] Kung K S and Hayes K F 1993 Fourier transform infrared spectroscopic study of the adsorption of
cetyltrimethylammonium bromide and cetylpyridinium chloride on silica Langmuir 9 263–7
[31] Horr T J, Arora P S, Ralston J and Smart R S 1995 XPS film thickness and adsorption studies of
alkyltrimethylammonium bromides and organosilanes on silica surfaces Colloid Surf. A 102 181–90
[32] Tyrode E, Rutland M W and Bain C D 2008 Adsorption of CTAB on hydrophilic silica studied by linear and
nonlinear optical spectroscopy J. Am. Chem. Soc. 130 17434–45
[33] Evans D F and Wennerstrom H 1994 The Colloidal Domain (New York: Wiley)
[34] Okada H, Kajiwara Y, Tanaka K and Chujo Y 2013 Rapid heat generation under microwave irradiation by
imidazolium-presenting silica nanoparticles Colloid Surf. A 428 65–9
[35] Khalil K M S 2007 Cerium modified MCM-41 nanocomposite materials via a nonhydrothermal direct
method at room temperature J. Colloid Interface Sci. 315 562–8
[36] Kleitz F, Schmidt W and Schüth F 2003 Calcination behavior of different surfactant-templated
mesostructured silica materials Micropor. Mesopor. Mater. 65 1–29
[37] Pamela M, Visintin F, Carbonell R G, Schauer C K and DeSimone J M 2005 Chemical functionalization of
silica and alumina particles for dispersion in carbon dioxide Langmuir 21 4816–23
[38] Zhao B, Hu H, Yu A, Perea D and Haddon R C 2005 Synthesis and characterization of water soluble single-
walled carbon nanotube graft copolymers J. Am. Chem. Soc. 127 8197–203
[39] Slostowski C, Marre S, Babot O, Toupance T and Aymonier C 2012 Near- and supercritical alcohols as
solvents and surface modifiers for the continuous synthesis of cerium oxide nanoparticles Langmuir 28
16656–63
[40] Abdollahi S N, Naderi M and Amoabediny G 2013 Synthesis and characterization of hollow gold
nanoparticles using silica spheres as templates 2013 Colloid Surf. A 436 1069–75

17

View publication stats

You might also like