You are on page 1of 10

Article

Cite This: Langmuir 2018, 34, 366−375 pubs.acs.org/Langmuir

New Aspects of the Gold Nanorod Formation Mechanism via Seed-


Mediated Methods Revealed by Molecular Dynamics Simulations
Jose ́ Adriano da Silva and Mario R. Meneghetti*
Grupo de Catálise e Reatividade QuímicaGCaR, Instituto de Química e Biotecnologia da Universidade Federal de Alagoas, Av.
Lourival de Melo Mota, s/n, Cidade Universitária, CEP, 57072-970 Maceió-AL, Brazil
*
S Supporting Information

ABSTRACT: New aspects of the formation and growth


mechanism of gold nanorods (AuNRs) during seed-mediated
colloidal synthesis are revealed from the results of molecular
dynamics simulation. The model systems consist of
cetyltrimethylammonium bromide (CTAB) units adsorbed
on low-index [Au(110), Au(100), and Au(111)] and high-
index [Au(250)] gold surfaces. The CTAB units are adsorbed
as adjacent cylindrical micelles when the relative number of
adsorbed bromide ions is small. At later AuNR growth stages,
the number of bromide ions increases as the [AuBr2]− species
pass through the channels between the adsorbed micelles on
the gold surface. Thus, the mature AuNRs have a high concentration of bromide ions at their surface, which appears to change
the organization of the CTAB units on the particle surface from adsorbed micelles to a compact CTAB bilayer.

■ INTRODUCTION
Anisotropic metal nanoparticles are characterized by several
standing the synthesis mechanism. Despite advances in
elucidating the formation mechanism of AuNRs, especially
unique properties that originate from nanoparticle size effects the development of mechanistic models that are consistent
because of their high surface areas and electron confinement. with the experimental data, some gaps in understanding still
In particular, gold nanorods (AuNRs) exhibit remarkable remain.4
properties and are promising for numerous applications.1 In 2005, Murphy and co-workers15 proposed the first model
The most common and successful strategy for synthesizing explaining the origin of the anisotropic growth mechanism of
AuNRs is thought to be a seed-mediated colloidal method, colloidal AuNRs, that is, the zipping-like growth mechanism. It
which was originally used by Murphy et al. in 2001 to produce states that the CTAB surfactants preferentially adsorb on the
pentatwinned AuNRs.2 This method was later modified by El- lateral facets of AuNRs. An AuNR begins to develop
Sayed and Nikoobakht in 20033 to produce single-crystal anisotropy when a CTAB bilayer forms on the lateral sides
AuNRs in higher yields (see Table 1).4 The crystallographic of the growing particle. The bilayer hinders the access of the
facet indices of the AuNRs obtained by both methodologies Au(I) species to the gold surface, where they are reduced,
are depicted in Figure 1A.5,6 whereas the Au(I) species can still diffuse into the tips of the
Briefly, in the seed-mediated synthesis of AuNRs, the growth AuNR (Figure 1B).
of these anisotropic nanoparticles is mediated by the addition Jana16 proposed another model for the anisotropic growth of
of metallic gold atoms at the surface of previously prepared AuNRs, namely, the soft template model. In this model,
small gold nanoparticles (seeds), by a controlled reduction of elongated CTAB micelles in the growth solution generate a
Au(I) species at the surface of the seeds that grow in an template into which seeds with diameters of 1−3 nm can
anisotropic way due to the presence of cetyltrimethylammo- penetrate. Then, the shape of the micelle, that is, the soft
nium bromide (CTAB) as a growth-driving agent.2,3,7 Indeed, template, induces the anisotropy (Figure 1C). According to
many studies of these standard synthesis methods and this model, the temperature and CTAB concentration during
adaptations thereof have appeared in the literature. In the synthesis must be within a certain range to generate
particular, the effects of the seed size,8 cosurfactants,9 elongated micelles that induce the formation of AuNRs.
additives,10 temperature,11,12 surfactant chain length,13 and Additionally, additives that can promote micellar transitions to
headgroup structure14 have been investigated. more elongated CTAB micelles, for example, salicylate
Because the properties of AuNRs depend on their shape, the aromatic additives,10 can be used to fabricate AuNRs with a
particle symmetry breaking and anisotropic growth mechanism
must be understood to optimize the synthesis conditions and Received: October 26, 2017
thus improve the shape and size uniformity of the particles. Revised: December 13, 2017
Accordingly, several research groups are focused on under- Published: December 15, 2017

© 2017 American Chemical Society 366 DOI: 10.1021/acs.langmuir.7b03703


Langmuir 2018, 34, 366−375
Langmuir Article

Table 1. General Details for Preparing AuNRs via Seed-Mediated Synthesis


aspect ratio selectivitya
synthesis method seed solution seed shape growth solution crystallinity of the AuNRs (AR) (%)
Murphy and co- NaBH4, HAuCl4, and decahedral ascorbic acid, HAuCl4, and pentatwinned with pentagonal 6−20 ca. 15
workers2 citrate CTAB cross section
El-Sayed and NaBH4, HAuCl4, and cuboctahedral ascorbic acid, HAuCl4, CTAB, single crystal with octagonal 1.5−5.0 ca. 90
Nikoobakht3 CTAB and AgNO3 cross section
a
Relative to the number of nanoparticles formed.

higher AR. Thus, on the basis of this model, seeds with AuNR surface. In contrast, an organized bilayer pattern was
diameters larger than 3.5 nm cannot generate AuNRs, proposed to exist at higher CTAB concentrations. It should be
explaining the observation that AuNRs cannot be fabricated noted that separating and redispersing the AuNRs generated in
from large seeds because they are larger than the micelle this system in a dilute aqueous solution of CTAB resulted in
template. the formation of a “collapsed bilayer” of CTAB on the
It is essential to determine how CTAB is adsorbed and particles.
structurally organized on the surfaces of growing and mature Gómez-Graña and co-workers used transmission electron
AuNRs to understand the nanoparticle growth mechanism. microscopy, small-angle neutron scattering (SANS), and small-
Analogous CTAB surfactants with a shorter alkyl chain angle X-ray scattering to measure the thickness of the CTAB
(CnTAB, 10 ≤ n < 16) were used to drive growth in the structure adsorbed on the single-crystal AuNR. The measured
pentatwinned AuNR synthesis and led to AuNRs with smaller value of 3.2 ± 0.2 nm is significantly smaller than twice the
ARs. Using the zipping-like mechanism, these results were extended surfactant chain length (4.34 nm), suggesting
explained by the different capacities of these surfactants to significant tilting or interdigitation of the hydrocarbon chains
generate a stable bilayer structure on the lateral surfaces of the of the surfactants.18
AuNRs; the longer chains provided greater stability and thus In both seed-mediated synthesis protocols,2,3 the concen-
generated AuNRs with higher ARs.13 tration of CTAB in the growth solution is generally 0.1 M;
Hafner and co-workers used surface-enhanced Raman however, the CTAB second critical micelle concentration in an
spectroscopy to confirm the existence of Au−Br− interactions aqueous solution is 0.27 M,19 meaning cylindrical micelles are
and show that the arrangement of the adsorbed CTAB on the not present in the growth solution. Walsh and co-workers20
gold surface depends on its concentration.17 They verified that showed that when a standard synthesis protocol for single-
the adsorbed CTAB forms a “collapsed bilayer” at relatively crystal AuNRs is employed, that is, when silver is present, the
low CTAB concentrations by detecting the interactions symmetry breaking only occurs in seeds that are 4−6 nm in
between the alkane chains of the CTA+ moiety and the diameter. In a 0.1 M CTAB solution, the micelles are

Figure 1. (A) Crystallographic facets of different types of AuNRs: (1) intermediate growth stage of a single-crystal AuNR, (2) mature single-crystal
AuNR, and (3) mature pentatwinned AuNR. (B) Illustration of the AuNR zipping-like growth mechanism. (C) Illustration of the role of the soft
template during the growth process.

367 DOI: 10.1021/acs.langmuir.7b03703


Langmuir 2018, 34, 366−375
Langmuir Article

ellipsoidal, with minor and major axes of 2.6 and 4.4 nm, the simulations, the migration of bromides to the gold surfaces
respectively.21 Then, the micelles in the solution under is monitored. In this study, systems with a number of bromides
standard synthesis conditions are smaller than or at least inserted at the gold/CTA+ interface were constructed. In
approximately the same size as the seed particles and cannot particular, the Br/Au ratio found by Meena and Sulpizi (0.1
serve as templates during growth. Thus, evidence supporting Br/Au) and a slightly larger ratio than that measured
the soft template mechanism is lacking, and it is not currently experimentally (0.2 Br/Au) were employed.
accepted as an explanation of the anisotropic growth. The systems were constructed with CTAB adsorbed on
Vaia and co-workers mapped the crystallographic facets different surfaces, namely, the low-index Au(110), Au(100),
formed during single-crystal AuNR growth and divided the and Au(111) surfaces and the high-index Au(250) surface,
growth process of the particles into five stages based on the with bromides at the gold/CTA+ interface in ratios of 0.1 Br/
results.22 Thus, using the zipping-like growth mechanism as the Au (low concentration of bromides) and 0.2 Br/Au (high
main model, it is assumed that a bilayer exists on the lateral concentration of bromides).
facets of the growing AuNR in all the stages. It should be In this work, we demonstrate that the adsorption patterns
noted, however, that this assumption is supported by observed in the simulations are consistent with the
experimental studies of mature AuNRs. In fact, the existence experimental measurements and the results are used to
of a bilayer coating during all the growth stages has not been propose the growth mechanisms of the pentatwinned and
verified experimentally. The existence of a compact bilayer single-crystal AuNRs.
during the entire growth process conflicts with the It should be noted that the role of silver, which is present
experimental results because growth occurs at the side facets, during the production of single-crystal AuNRs, was not
although to a lesser degree than at the tips, in the early investigated in this study (silver was not included in the
stages.22 However, in the later stages of the growth process, the simulation models), which focuses on the gold surface/
formation of a compact bilayer on the lateral facets appears to bromide/surfactant/bromide/water interactions. This analysis
occur, making it difficult for the most abundant source of gold, is reasonable because Funston and co-workers demonstrated
that is, [AuBr2]−, to pass through the hydrophobic core of the that the presence of silver nitrate in the medium during the
CTAB bilayer.23 synthesis of single-crystal AuNRs only affects the symmetry-
Sulpizi and co-workers24−26 used molecular dynamics (MD) breaking event.31
simulations to identify the CTAB adsorption modes on low-
index and high-index gold surfaces. They observed adjacent
cylindrical micelles on all the surfaces, as well as channels
■ COMPUTATIONAL METHODOLOGY
System Setup. The topology and atomic parameters for
(called intermicellar channels), that allow the molecular gold the CTA+ and Br− units32 used in the simulations were
source to flow to the surface of the AuNRs. According to these validated in our previous work;30 they were shown to lead to
researchers, the width of these channels depends on the facet structures exhibiting CTAB/gold surface interactions and
on which the micelles are adsorbed and decreases in the CTAB micelles in aqueous solution and to predict the
following order: Au(111) > Au(110) > Au(100) > Au(250). counterion dissociation degree. The parameters for the gold
Thus, the Meena−Sulpizi growth mechanism states that a surface33 were used in previous simulations (Table 2).24−26,30
series of cylindrical CTAB micelles, not a CTAB bilayer,
adsorb on all facets of the gold surface, resulting in the Table 2. Lennard-Jones Parameters Used for the Bromide
formation of channels filled with water molecules. Indeed, the Ions and Gold Atoms
presence of these channels is consistent with the growth
process during the early stages of AuNR formation because species C6 (kJ·mol−1·nm6) × 10−2 C12 (kJ·mol−1·nm6) × 10−5

both the lateral facets and the tips of the AuNRs grow, Br 5.8243749 5.10031
although at different rates. However, the adsorption of adjacent Au 2.92057 0.964326
cylindrical CTAB micelles observed in the simulations is
contradictory to the experimental results,17,27,28 which confirm The initial configurations (input) for all the simulations
the existence of a bilayer structure. Nevertheless, it should be included: (i) a gold surface, which could be any of the different
noted that all the experimental data confirming the presence of facets, (ii) a specific number of CTAB units, with the bromide
a compact bilayer of CTAB units were obtained using mature ions distributed below or above of the CTA+ bilayer (Br-1 and
AuNRs. Br-2, respectively), and (iii) a suitable number of water
Furthermore, the number of bromide ions adsorbed on the molecules (see Figure 2). More specifically, for the simulations
surfaces in the Meena and Sulpizi simulations (a Br/Au ratio of of the Au(250) surface, two different amounts of Br-1 (Br-1/
ca. 0.10) is smaller than that measured experimentally for Au molar ratios of 0.2 and 0.1) were placed on the gold
AuNRs.24−26 Using X-ray photoelectron spectroscopy, Zhang surface. For the low-index Au(110), Au(100), and Au(111)
and co-workers29 found that the Br/Au ratio on AuNR surfaces surfaces, a 0.1 Br-1/Au ratio was employed, and for the
is ca. 0.17. This difference in the simulated and experimental Au(100) surface, a 0.2 Br/Au ratio was also used. The CTA+
Br/Au ratios motivated this study to determine the influence of packing density in the simulations was slightly higher than that
this ratio on the formation of a bilayer of CTAB in MD calculated from the cross-sectional area of the CTA +
simulations. headgroup (0.32 nm2).34 Each system was filled with SPC
Therefore, the adsorption of CTAB on gold surfaces is water molecules35 and the number of the bromide ions (Br-2)
reexamined using MD simulations.30 Previous theoretical required to neutralize the charged polar groups of the bilayers.
studies typically employed an initial configuration of the The characteristics of the systems are summarized in Table 3.
CTAB species in which the CTA+ cations are perpendicular to Simulation Procedures. The energies of the initial
the gold surface and the bromide anions are on top of the configurations were minimized by running 5000 steps of a
CTA+ bilayer, dispersed in the aqueous solution.24−26 During steepest descent procedure. The systems were simulated under
368 DOI: 10.1021/acs.langmuir.7b03703
Langmuir 2018, 34, 366−375
Langmuir Article

■ RESULTS AND DISCUSSION


Simulation of CTAB Adsorption on the Au(250)
Surface with 0.1 Br/Au. In this simulation, the initial
configuration has a relatively low number of bromide ions (Br-
1) near the gold surface, compared to the total number of
bromide ions (Br-1 + Br-2). After the simulation begins,
adjacent cylindrical micelles of CTAB are formed on the gold
surface and these structures are observed until the end of the
simulation (see Figure 3A). Figure 3B shows the water
molecule arrangement near the gold surface. Clearly, the space
between the adsorbed CTAB micelles, that is, the intermicellar
channels, is filled with water molecules.
The densities of nitrogen, bromide-1, bromide-2, the last
carbon of the hydrophobic tail, and the water molecules were
measured along the axis perpendicular to the surface
Figure 2. Initial configurations of the surfactant bilayer on the gold (estimated for the last 50 ns of the simulation). These results
surfaces. (A) 0.1 Br/Au and (B) 0.2 Br/Au. Gold atoms (yellow), are presented in Figure 3C. A density peak is observed for the
nitrogen atoms (blue), hydrophobic tail (gray), Br-1 (magenta), and last carbon of the CTA+ chain (indicated by the black circle),
Br-2 (green). The water molecules were omitted for clarity. suggesting a constant carbon-tail adsorption on the gold
surface. This interaction was experimentally detected by
isothermal−isobaric NPT conditions, and periodic boundary Hafner and co-workers,17 and the related adsorption structure
conditions were applied to the rectangular boxes in all corresponds to the formation of adjacent cylindrical CTAB
directions. During the equilibration and production phases, micelles at the surface, which the authors referred to as a
the leap-frog algorithm was used to integrate Newton’s “collapsed bilayer”. The distribution of nitrogen atoms
equations of motion with a time step of 0.002 ps. Each permeates the intermicellar channel walls. The Br-1 ions
simulation trajectory was 200 ns long. The center of mass remain close to the surface, whereas the channel is filled with
motion was removed every five steps. The Berendsen the Br-2 ions (Figure 3A,C).
thermostat was used to maintain the system temperature at Simulation of CTAB Adsorption on the Au(250)
300 K by independently coupling the temperatures of the Surface with 0.2 Br/Au. In the simulation of CTAB
solute and solvent using a time constant of 0.4 ps.36 The adsorption on the Au(250) surface with a relatively high
pressure was maintained using the Parrinello−Rahman
number of bromide ions on the gold surface, a bilayer structure
barostat by weakly coupling the particle coordinates and box
of CTAB units is observed on the gold surface (see Figure 4A).
dimensions to a pressure bath of 1.0 bar.37 Anisotropic
Figure 4B shows that only a few water molecules reach the
coordinate scaling was used with a relaxation time of 0.4 ps and
a compressibility of 4.5 × 10−5 bar−1, which is appropriate for gold surface. This bilayer adsorption pattern of the CTAB
water. No bond constraints were applied during the units is described in the literature.17,27,28
simulations. A generalized reaction field correction and a In this simulation, the densities of nitrogen, bromide-1,
cutoff of 1.4 nm were used for both the van der Waals and bromide-2, the last carbon of the hydrophobic tail, and the
long-range electrostatic interactions, and the permittivity water molecules were measured along the axis perpendicular to
(dielectric constant) was 66.38 In all the simulations, the pair the surface (estimated for the last 50 ns of the simulation). The
lists for the short-range nonbonded and long-range electro- distributions of these elements are shown in Figure 4C. The
static interactions were updated every five time steps. The general distributions of the two types of bromide ions (Br-1
configurations of the trajectory were recorded every 1 ps. and Br-2) are nearly the same as those in the initial
The time-dependent distributions of nitrogen, bromide-1, configuration (see Figure 2 and Table 3), that is, no migration
bromide-2, the last carbon of the hydrophobic tail, and the occurs between the two layers of bromide ions.
water molecules were derived from the MD simulations and On the basis of the difference between the two highest
analyzed. GROMACS 4.5.5 was used for all the MD densities of nitrogen atoms indicated in Figure 4C, the
simulations and trajectory analyses.39 The VMD software thickness of the adsorbed surfactant bilayer is estimated to be
version 1.9.1 was used to visualize the trajectories and prepare 3.20 nm. This value is in agreement with the thickness on the
the figures.40 surface of AuNRs measured by SANS (3.2 ± 0.2 nm).18

Table 3. Parameters of the Simulated Systems


type of number of gold numbers of bromide-1 and number of water CTA+ box dimensions x/y/z thickness of the gold surface
surface atoms bromide-2a molecules units (nm) (nm)
Au(250) 1310 40 and 156 4141 196 4.00/4.07/16.39 1.59
Au(250) 1310 80 and 116 4181 196 4.00/4.07/16.39 1.59
Au(100) 1600 40 and 156 3785 196 4.10/4.10/15.61 1.43
Au(100) 1600 80 and 116 3785 196 4.10/4.10/16.61 1.43
Au(111) 1332 40 and 156 3955 196 4.10/4.10/15.60 1.26
Au(110) 1400 40 and 156 3782 196 4.08/4.08/15.90 1.45

a
The total number of bromide ions in all the calculations was 196.

369 DOI: 10.1021/acs.langmuir.7b03703


Langmuir 2018, 34, 366−375
Langmuir Article

Figure 3. (A) Snapshot of the MD simulation of CTAB adsorption on the Au(250) surface with 0.1 Br/Au. The yellow and dark blue spheres
represent gold and nitrogen atoms, respectively. The gray lines are the apolar tails. The bromide-1 and bromide-2 ions are represented by magenta
and green spheres, respectively. The water molecules were omitted for clarity. (B) Snapshot of the MD simulation of water molecules on the
Au(250) surface. (C) Measurements of the densities of nitrogen (blue), the last carbon of the hydrophobic tail (gray), bromide-1 (magenta),
bromide-2 (green), and the water molecules (red and white spheres) relative to the gold surface.

Simulation of CTAB Adsorption on the Au(110), ns of the simulation) are presented in Figure 6C. The general
Au(100), and Au(111) Surfaces with 0.1 Br/Au. In these distribution of the Br-1 ions indicates that they remain at the
simulations of a relatively low concentration of bromide ions gold/CTA+ interface, and likewise, the Br-2 ions remain near
on the gold surfaces, the adsorbed CTAB units are organized the CTA+/aqueous solution interface (see Figure 6A,C).
as adjacent cylindrical micelles and intermicellar channels are Furthermore, as shown in this figure, the difference between
observed between them. These channels are filled with water the two nitrogen peaks indicates that the thickness of the
molecules, as shown in Figure 5. adsorbed surfactant bilayer is 3.43 nm in this case. This
Indeed, a similar pattern of CTAB adsorption was also distance is slightly larger than that observed in the simulation
described by Meena and Sulpizi on low-index gold surfaces.24 of the Au(250) surface under similar conditions. Figure 6C
They also observed a similar Br/Au ratio on the gold surfaces also shows that the nitrogen distribution relative to the gold
after simulation. Moreover, they observed that the channel surface exhibits two main peaks in the number density at
thickness depends on the facet; it decreases in the following approximately 0.5 and 3.9 nm and a third, smaller density peak
order: (110) > (100) > (111). In these simulations, alkyl chain at approximately 1.0 nm (indicated by the black circle). These
interactions on the gold surface, which were not detected in
features might be attributed to the weaker binding interactions
previous studies, are observed (see the distribution of the last
between the Br-1 ions and the Au(100) surface, if compared
carbon of the hydrophobic tail along the axis perpendicular to
with these interactions with the Au(250) surface (see Table S1
the surface for the last 50 ns of the simulation in Figure S1).
Simulation of CTAB Adsorption on the Au(100) and Figures S2 and S3), which enable greater mobility of the
Surface with 0.2 Br/Au. A relatively high number of CTA+ polar groups. Thus, it is suggested that the density of the
bromide ions were placed on the gold surface in the initial bilayer depends on the strength of the bromide interaction
configuration for this simulation. Consequently, the CTAB with the gold surface.
units are arranged in a bilayer structure on this gold surface Proposed Growth Mechanism for the Pentatwinned
(see Figure 6A). As shown in Figure 6B, the water molecules AuNR. Figure 7 shows a model of the evolution of the
do not reach the gold surface because of the formation of a crystallographic facets during pentatwinned AuNR growth.6 In
compact bilayer of CTAB units. This pattern is also observed stage I, the seed symmetry breaks and anisotropy develops.
for the Au(250) system with a relatively high content of Here, the pentatwinned AuNR must grow by the mechanism
bromide ions (see Figure 4). described by Meena and Sulpizi,24 in which adjacent cylindrical
The distributions of nitrogen, bromide-1, bromide-2, the last micelles are adsorbed on the lateral and tip facets. The
carbon of the hydrophobic tail, and the water molecules along [AuBr2]− species reach the surface through the intermicellar
the axis perpendicular to the surface (estimated for the last 50 channels.
370 DOI: 10.1021/acs.langmuir.7b03703
Langmuir 2018, 34, 366−375
Langmuir Article

Figure 4. (A) Snapshot of the MD simulation of CTAB adsorption on the Au(250) surface with 0.2 Br/Au. The yellow and dark blue spheres
represent gold and nitrogen atoms, respectively. The gray lines are the apolar tails. The bromide-1 and bromide-2 ions are represented by magenta
and green spheres, respectively. The water molecules were omitted for clarity. (B) Snapshot of the MD simulation of water molecules on the
Au(250) surface. (C) Measured densities of nitrogen (blue), the last carbon of the hydrophobic tail (gray), bromide-1 (magenta), bromide-2
(green), and the water molecules (red and white spheres) relative to the gold surface.

Figure 5. Intermicellar channels observed in the MD simulations of low-index surfaces with 0.1 Br/Au. The yellow spheres represent gold atoms.
(A) Au(100), (B) Au(111), and (C) Au(110). The water molecules are represented by red and white spheres.

Previous studies of CTAB adsorption patterns on a bilayer on the lateral (100) facet of mature pentatwinned
pentatwinned AuNRs revealed the presence of a bilayer.27 AuNRs (Figure 6). Thus, it is suggested that bromide must
Our simulation model shows that a higher concentration of accumulate on the gold surface during the growth of the
bromides on the gold surface is necessary for the formation of particle. An increase in the number of bromides on the gold
371 DOI: 10.1021/acs.langmuir.7b03703
Langmuir 2018, 34, 366−375
Langmuir Article

Figure 6. (A) Snapshot of the MD simulation of CTAB adsorption on the Au(100) surface with 0.2 Br/Au. The yellow and dark blue spheres
represent gold and nitrogen atoms, respectively. The gray lines are the apolar tails. The bromide-1 and bromide-2 ions are represented by magenta
and green spheres, respectively. The water molecules were omitted for clarity. (B) Snapshot of the MD simulation of water molecules on the
Au(100) surface. (C) Measurements of the densities of nitrogen (blue), the last carbon of the hydrophobic tail (gray), bromide-1 (magenta),
bromide-2 (green), and the water molecules (red and white) relative to the gold surface.

surface induces the rearrangement of the CTAB units on the intermicellar channels to close, which sustains the anisotropic
gold particle, causing the adsorbed cylindrical micelles (stages growth (stages II−IV). This event is caused by the
II−III) to transform into a compact bilayer (stage IV). This reorganization of the CTA+ ions as the concentration of
hypothesis is reasonable because the continuous reduction of bromide ions, on the gold surface, increases. Finally, the
the Au(I) ions of the [AuBr2]− species on the gold surface formation of the high-index Au(250) facet is induced in stage
releases bromide ions in loco.23 V. This transition from the Au(100) to the Au(250)
Proposed Growth Mechanism for the Single-Crystal crystallographic facet is probably due to the reorganization of
AuNR. Figure 8 shows the single-crystal AuNR growth the gold atoms at the surface to generate a crystalline facet with
stages.22 The growth begins with seed symmetry breaking a greater ability to stabilize bromide ions.
(stage I) and the adsorption of the CTAB units via the Studies of the CTAB arrangement on mature single-crystal
bromide ions on the Au(100) and Au(111) surfaces. AuNRs reported the presence of a CTAB bilayer.17,28 The
Vaia and co-workers22 reported growth on the lateral facets
simulation results indicate that this structure appears when the
of AuNRs during the early growth stages, which is consistent
concentration of bromides on the gold surface is high (Figure
with the presence of intermicellar channels on the lateral facets
4). Therefore, the viability of the proposed growth mechanism,
(stages II−IV). Thus, at these stages, because of the low
concentration of bromide ions on the gold surfaces, the which assumes that a transition from cylindrical micelles to a
adsorbed CTAB molecules form adjacent cylindrical micelles bilayer occurs, relies on the assumption that bromide
on all the facets, based on the results of Meena and Sulpizi24 accumulates at the gold/CTA+ interface. This assumption
and this study (see Figure 5). must be verified by experimental evidence to confirm the
Again, because the flux of [AuBr2]− to the tips of the AuNR proposed mechanism.
and their subsequent rate of reduction are higher, many It should be noted that in the synthesis of single-crystal
bromide ions are released at the tips. This excess of bromide AuNRs, the majority of the Au(I) ions present during the
ions must move from the surface of the tips [Au(111) facets] growth is not completely reduced. This result must be due to
to the sides of the rods [Au(100) facets] because of the higher the formation of a compact bilayer of CTAB units on all the
affinity of the latter facet for bromide ions.41 The increasing facets of mature AuNRs, thus preventing the Au(I) species
concentration of bromide ions on the lateral facets causes the from accessing the surface of the gold particle.42
372 DOI: 10.1021/acs.langmuir.7b03703
Langmuir 2018, 34, 366−375
Langmuir Article

Figure 7. Illustration of the growth stages of a pentatwinned AuNR during seed-mediated colloidal synthesis.

Figure 8. Illustration of the growth stages of a single-crystal AuNR during seed-mediated colloidal synthesis.

■ CONCLUSIONS
An MD simulation model that captures the different
The formation of the adsorbed cylindrical micelles with
intermicellar channels allows the [AuBr2]− ions to reach the
gold surface and is therefore proposed to be the driving force
arrangements of CTAB units adsorbed on various crystallo- of the anisotropic growth, in accordance with the results of
graphic facets of AuNRs in an aqueous solution was previous theoretical studies.24,25 This mechanism is consistent
constructed. On the basis of the model, a detailed mechanism with the growth of the particles (length and width) in the early
for the colloidal AuNR formation in an aqueous phase that stages and with the zipping-like growth mechanism in the later
accounts for the effect of the number of bromides at the gold stages of AuNR fabrication. The driving force that separates
the early and later stages of the AuNR growth process appears
surface on the CTAB adsorption structure, which can be to be related to the concentration of bromide ions adsorbed at
anchored cylindrical micelles or a compact bilayer, was the gold surface. A relatively small number of bromide ions on
proposed. the gold surface leads to the formation of adsorbed adjacent
373 DOI: 10.1021/acs.langmuir.7b03703
Langmuir 2018, 34, 366−375
Langmuir Article

micelles of CTAB on it. As the concentration of bromide ions (4) Lohse, S. E.; Murphy, C. J. The Quest for Shape Control: A
at the gold surface increases during the later stages of particle History of Gold Nanorod Synthesis. Chem. Mater. 2013, 25, 1250−
growth, the arrangement of CTAB units on the gold particle 1261.
changes to a bilayer structure. This mechanism can apply to (5) Carbó-Argibay, E.; Rodríguez-González, B.; Gómez-Graña, S.;
both methods of AuNR synthesis, that is, for the single-crystal Guerrero-Martínez, A.; Pastoriza-Santos, I.; Pérez-Juste, J.; Liz-
Marzán, L. M. The Crystalline Structure of Gold Nanorods Revisited:
and pentatwinned methods.
Evidence for Higher-Index Lateral Facets. Angew. Chem., Int. Ed.
It is concluded that the concentration of bromide ions at the 2010, 49, 9397−9400.
gold surface during the reduction of the [AuBr2]− species must (6) Johnson, C. J.; Dujardin, E.; Davis, S. A.; Murphy, C. J.; Mann, S.
be known to understand the AuNR growth process in aqueous Growth and form of gold nanorods prepared by seed-mediated,
solutions. Furthermore, the change in the CTAB adsorption surfactant-directed synthesis. J. Mater. Chem. 2002, 12, 1765−1770.
pattern during the growth process provides insight into the role (7) da Silva, M. G. A.; Nunes, Á . M.; Meneghetti, S. M. P.;
that the surfactant plays in causing anisotropic growth. Meneghetti, M. R. New aspects of gold nanorod formation via seed-
This work hopefully contributes to the understanding of the mediated method. C. R. Chim. 2013, 16, 640−650.
growth mechanism of AuNRs and sheds more light on the (8) Gole, A.; Murphy, C. J. Seed-Mediated Synthesis of Gold
fascinating studies of nanoparticle growth. Nanorods: Role of the Size and Nature of the Seed. Chem. Mater.


2004, 16, 3633−3640.
ASSOCIATED CONTENT (9) Ye, X.; Zheng, C.; Chen, J.; Gao, Y.; Murray, C. B. Using Binary
Surfactant Mixtures To Simultaneously Improve the Dimensional
*
S Supporting Information
Tunability and Monodispersity in the Seeded Growth of Gold
The Supporting Information is available free of charge on the Nanorods. Nano Lett. 2013, 13, 765−771.
ACS Publications website at DOI: 10.1021/acs.lang- (10) Ye, X.; Jin, L.; Caglayan, H.; Chen, J.; Xing, G.; Zheng, C.;
muir.7b03703. Doan-Nguyen, V.; Kang, Y.; Engheta, N.; Kagan, C. R.; Murray, C. B.
Graph of the overlap of the last carbon of the tail on low- Improved Size-Tunable Synthesis of Monodisperse Gold Nanorods
through the Use of Aromatic Additives. ACS Nano 2012, 6, 2804−
index surfaces with 0.1 Br/Au; MD simulations of 2817.
bromide adsorption on Au(250) and Au(100) surfaces; (11) Zijlstra, P.; Bullen, C.; Chon, J. W. M.; Gu, M. High-
and model details of the gold surface and sodium Temperature Seedless Synthesis of Gold Nanorods. J. Phys. Chem. B
bromide solution (PDF) 2006, 110, 19315−19318.


(12) Burrows, N. D.; Harvey, S.; Idesis, F. A.; Murphy, C. J.
Understanding the Seed-Mediated Growth of Gold Nanorods
AUTHOR INFORMATION through a Fractional Factorial Design of Experiments. Langmuir
Corresponding Author 2017, 33, 1891−1907.
*E-mail: mrm@qui.ufal.br. (13) Gao, J.; Bender, C. M.; Murphy, C. J. Dependence of the Gold
Nanorod Aspect Ratio on the Nature of the Directing Surfactant in
ORCID Aqueous Solution. Langmuir 2003, 19, 9065−9070.
Mario R. Meneghetti: 0000-0002-0722-8599 (14) da Silva, M. G. A.; Meneghetti, M. R.; Denicourt-Nowicki, A.;
Notes Roucoux, A. Tunable hydroxylated surfactants: an efficient toolbox
The authors declare no competing financial interest. towards anisotropic gold nanoparticles. RSC Adv. 2014, 4, 25875−


25879.
(15) Murphy, C. J.; Sau, T. K.; Gole, A. M.; Orendorff, C. J.; Gao, J.;
ACKNOWLEDGMENTS Gou, L.; Hunyadi, S. E.; Li, T. Anisotropic Metal Nanoparticles:
The authors acknowledge Dr. Santosh Kumar Meena at Synthesis, Assembly, and Optical Applications. J. Phys. Chem. B 2005,
Johannes Gutenberg University, Germany, for kindly providing 109, 13857−13870.
the atomic coordinates of the gold surfaces and Prof. Dr. Paulo (16) Jana, N. R. Gram-Scale Synthesis of Soluble, Near-
A. Netz at Federal University of Rio Grande do Sul for Monodisperse Gold Nanorods and Other Anisotropic Nanoparticles.
important contributions in solving technical problems with the Small 2005, 1, 875−882.
simulations, as well as, Prof. Dr. Thereza A. Soares at Federal (17) Lee, S.; Anderson, L. J. E.; Payne, C. M.; Hafner, J. H.
University of Pernambuco for encouraging us to carry out Structural Transition in the Surfactant Layer that Surrounds Gold
theoretical studies in the field of nanoparticle syntheses. The Nanorods as Observed by Analytical Surface-Enhanced Raman
Spectroscopy. Langmuir 2011, 27, 14748−14756.
authors are grateful for the financial support of the Brazilian (18) Gómez-Graña, S.; Hubert, F.; Testard, F.; Guerrero-Martínez,
funding agencies CNPq, Capes, Fapeal, and INCT-Catalise A.; Grillo, I.; Liz-Marzán, L. M.; Spalla, O. Surfactant (Bi)Layers on
and for the computational resources provided by the Cenapad- Gold Nanorods. Langmuir 2012, 28, 1453−1459.
Unicamp. M.R.M. thanks CNPq for the research fellowships.


(19) Goyal, P. S.; Dasannacharya, B. A.; Kelkar, V. K.; Manohar, C.;
Rao, K. S.; Valaulikar, B. S. Shapes and sizes of micelles in CTAB
REFERENCES solutions. Physica B: Condensed Matter 1991, 174, 196−199.
(1) Falagan-Lotsch, P.; Grzincic, E. M.; Murphy, C. J. New Advances (20) Walsh, M. J.; Barrow, S. J.; Tong, W.; Funston, A. M.;
in Nanotechnology-Based Diagnosis and Therapeutics for Breast Etheridge, J. Symmetry Breaking and Silver in Gold Nanorod Growth.
Cancer: An Assessment of Active-Targeting Inorganic Nanoplatforms. ACS Nano 2015, 9, 715−724.
Bioconjugate Chem. 2017, 28, 135−152. (21) Berr, S. S. Solvent isotope effects on alkytrimethylammonium
(2) Jana, N. R.; Gearheart, L.; Murphy, C. J. Seed-Mediated Growth bromide micelles as a function of alkyl chain length. J. Phys. Chem.
Approach for Shape-Controlled Synthesis of Spheroidal and Rod-like 1987, 91, 4760−4765.
Gold Nanoparticles Using a Surfactant Template. J. Adv. Mater. 2001, (22) Park, K.; Drummy, L. F.; Wadams, R. C.; Koerner, H.; Nepal,
13, 1389−1393. D.; Fabris, L.; Vaia, R. A. Growth Mechanism of Gold Nanorods.
(3) Nikoobakht, B.; El-Sayed, M. A. Preparation and Growth Chem. Mater. 2013, 25, 555−563.
Mechanism of Gold Nanorods (NRs) Using Seed-Mediated Growth (23) Moiraghi, R.; Douglas-Gallardo, O. A.; Coronado, E. A.;
Method. Chem. Mater. 2003, 15, 1957−1962. Macagno, V. A.; Pérez, M. A. Gold nucleation inhibition by halide

374 DOI: 10.1021/acs.langmuir.7b03703


Langmuir 2018, 34, 366−375
Langmuir Article

ions: a basis for a seed-mediated approach. RSC Adv. 2015, 5, 19329−


19336.
(24) Meena, S. K.; Sulpizi, M. Understanding the Microscopic
Origin of Gold Nanoparticle Anisotropic Growth from Molecular
Dynamics Simulations. Langmuir 2013, 29, 14954−14961.
(25) Meena, S. K.; Sulpizi, M. From Gold Nanoseeds to Nanorods:
The Microscopic Origin of the Anisotropic Growth. Angew. Chem.,
Int. Ed. 2016, 55, 11960−11964.
(26) Meena, S. K.; Celiksoy, S.; Schäfer, P.; Henkel, A.; Sönnichsen,
C.; Sulpizi, M. The role of halide ions in the anisotropic growth of
gold nanoparticles: a microscopic, atomistic perspective. Phys. Chem.
Chem. Phys. 2016, 18, 13246−13254.
(27) Nikoobakht, B.; El-Sayed, M. A. Evidence for bilayer assembly
of cationic surfactants on the surface of gold nanorods. Langmuir
2001, 17, 6368−6374.
(28) Matthews, J. R.; Payne, C. M.; Hafner, J. H. Analysis of
Phospholipid Bilayers on Gold Nanorods by Plasmon Resonance
Sensing and Surface-Enhanced Raman Scattering. Langmuir 2015, 31,
9893−9900.
(29) Zhang, Q.; Zhou, Y.; Villarreal, E.; Lin, Y.; Zou, S.; Wang, H.
Faceted Gold Nanorods: Nanocuboids, Convex Nanocuboids, and
Concave Nanocuboids. Nano Lett. 2015, 15, 4161−4169.
(30) da Silva, J. A.; Dias, R. P.; da Hora, G. C. A.; Soares, T. A.;
Meneghetti, M. R. Molecular dynamics simulations of cetyltrimethy-
lammonium bromide (CTAB) micelles and their interactions with a
gold surface in aqueous solution. J. Braz. Chem. Soc. 2018, 29, 191−
199.
(31) Tong, W.; Walsh, M. J.; Mulvaney, P.; Etheridge, J.; Funston, A.
M. Control of Symmetry Breaking Size and Aspect Ratio in Gold
Nanorods: Underlying Role of Silver Nitrate. J. Phys. Chem. C 2017,
121, 3549−3559.
(32) Reiser, S.; Deublein, S.; Vrabec, J.; Hasse, H. Molecular
dispersion energy parameters for alkali and halide ions in aqueous
solution. J. Chem. Phys. 2014, 140, 044504.
(33) Heinz, H.; Vaia, R. A.; Farmer, B. L.; Naik, R. R. Accurate
Simulation of Surfaces and Interfaces of Face-Centered Cubic Metals
Using 12-6 and 9-6 Lennard-Jones Potentials. J. Phys. Chem. C 2008,
112, 17281−17290.
(34) Nakahara, H.; Shibata, O.; Moroi, Y. Examination of surface
adsorption of cetyltrimethylammonium bromide and sodium dodecyl
sulfate. J. Phys. Chem. B 2011, 115, 9077−9086.
(35) Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P. The missing
term in effective pair potentials. J. Phys. Chem. 1987, 91, 6269−6271.
(36) Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.;
DiNola, A.; Haak, J. R. Molecular dynamics with coupling to an
external bath. J. Chem. Phys. 1984, 81, 3684.
(37) Parrinello, M.; Rahman, A. Polymorphic transitions in single
crystals: A new molecular dynamics method. J. Appl. Phys. 1981, 52,
7182.
(38) Tironi, I. G.; Sperb, R.; Smith, P. E.; van Gunsteren, W. F. A
generalized reaction field method for molecular dynamics simulations.
J. Chem. Phys. 1995, 102, 5451.
(39) Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark,
A. E.; Berendsen, H. J. C. GROMACS: fast, flexible, and free. J.
Comput. Chem. 2005, 26, 1701−1718.
(40) Humphrey, W.; Dalke, A.; Schulten, K. VMD: visual molecular
dynamics. J. Mol. Graphics 1996, 14, 33−38.
(41) Magnussen, O. M. Ordered Anion Adlayers on Metal Electrode
Surfaces. Chem. Rev. 2002, 102, 679−726.
(42) Orendorff, C. J.; Murphy, C. J. Quantitation of Metal Content
in the Silver-Assisted Growth of Gold Nanorods. J. Phys. Chem. B
2006, 110, 3990−3994.

375 DOI: 10.1021/acs.langmuir.7b03703


Langmuir 2018, 34, 366−375

You might also like