You are on page 1of 9

J Nanopart Res (2016)18:242

DOI 10.1007/s11051-016-3557-0

RESEARCH PAPER

A high-yield, one-step synthesis of surfactant-free gold


nanostars and numerical study for single-molecule
SERS application
S. Chatterjee . A. B. Ringane . A. Arya .
G. M. Das . V. R. Dantham . R. Laha .
S. Hussian

Received: 3 March 2016 / Accepted: 6 August 2016


Springer Science+Business Media Dordrecht 2016

Abstract We report a high-yield synthesis of starshaped gold nanostructures in one step, using a new
surfactant-free wet chemistry method. Compared to
the existing reports, these nanostars were found to
have longer and sharper spikes anchored uniformly on
the surface of the spherical core, allowing at least a
few hot spots irrespective of the incident light
polarization. The average experimental values of core
radius and spike length were found to be 88.5 and
72 nm, respectively. Using these values in numerical
simulations, the local electric field enhancement (g)
and localized surface plasmon resonance (LSPR)
spectrum were obtained. Moreover, the single-molecule surface-enhanced Raman scattering (SERS)
enhancement factor was found to vary from 1010 to
1013 depending on the excitation wavelengths. Our
theoretical calculations suggest that these nanostructures can be used to fabricate efficient SERS-based
biosensors for the detection of single molecules in real
time and for predicting structural information of single
molecules.

S. Chatterjee  A. B. Ringane  A. Arya 


G. M. Das  V. R. Dantham (&)  R. Laha
Department of Physics, Indian Institute of Technology
Patna, Bihta 801103, India
e-mail: dantham@iitp.ac.in
S. Hussian
Department of Chemistry, Indian Institute of Technology
Patna, Bihta 801103, India

Keywords Nanostars  Finite element method  Wet


chemistry method  Localized surface plasmons 
Lightning rod effect  Plasmonic hot spot

Introduction
Recently, single-molecule surface-enhanced Raman
spectroscopy has become a fascinating tool for
studying the structural properties, static and dynamic
events of single biomolecules (as opposed to ensemble
average), with the help of efficient plasmonic nanostructures (Lee et al. 2013; Matsushita and Kiguchi
2015). This tool is extremely useful in the field of
proteomics as the structures of protein molecules are
heterogeneous in nature (Sadana 1998). However,
high-quality surface-enhanced Raman scattering
(SERS) substrates are essential for studying singlemolecule Raman scattering. According to the electromagnetic theory of SERS, the power/intensity of
expected Raman scattering signal from the single
molecule strongly depends on the local electric field
intensity (Iloc) of the plasmonic particles (Kneipp et al.
2002; Kneipp and Kneipp 2006). According to the
quasi-static theory of nanostructures, the value of Iloc
strongly depends on the size, shape, relative dielectric
permittivity of the nanostructures and polarization of
the excitation light (Sarid and Challener 2010). For
these reasons, researchers have been trying to improve
the signal-to-noise ratio (S/N) of the experimental

123

242

Page 2 of 9

signal using various kinds of nanostructures such as


nanodimers (Le and Etchegoin 2009; Guo et al. 2013),
nanotriangles (Haes and Van Duyne 2002; Scarabelli
et al. 2014), nanoprisms (Ciou et al. 2009) and
nanorods (Yang et al. 2012; Sivapalan et al. 2013;
Greeneltch et al. 2013). However, the value of Iloc for
these nanostructures is extremely sensitive to their
orientation with respect to the polarization of excitation light. Due to this, if SERS substrates are prepared
using these nanostructures by self-assembly method,
then one cannot have the control on the orientation of
nanostructures. Therefore, it will not be possible to
excite localized surface plasmons in all nanostructures
existing at focal volume of the objective lens of the
Raman microscope. In contrast to the above-mentioned nanostructures, the star-shaped nanostructures
contain spikes in all directions. Therefore, the giant
electric field intensity could be expected at the end of
at least a few spikes irrespective of the polarization of
the excitation light. This enables one to make efficient
SERS substrates using nanostars for single-molecule
SERS application. In addition, these nanostructures
are found to be good candidates for nanosensing,
thermal therapy and drug delivery applications
(Chirico et al. 2015). All these applications, however,
need the nanostars to be prepared with relative ease
and with proper control. Following is an account of the
status of different methods used for preparing gold
nanostars.
In, Collen et al. 2006 first reported the synthesis of
star-shaped gold nanostructures using two-step seedmediated growth process. However, those nanostars
were polydispersed, and their yield was not so good
(914 %). In 2008, Kumar et al. succeeded in
improving the yield significantly. Later, many
research groups have synthesized nanostars of various
sizes using their own approach to improve its quality,
monodispersity and yield (Wu et al. 2009; Barbosa
et al. 2010; Trigari et al. 2011). Few researchers have
also engaged for the green synthesis of gold nanostars
after looking at their importance in various fields
(Kaneko et al. 2011). However, in all these synthesis
procedures, a surfactant was used to restrict the growth
of the nanoparticles during the synthesis. But, the
surfactant is not favourable for few applications like
single-molecule SERS and fluorescence as the surfactant keeps the molecule under test away from the
nanostar spikes. This obviously reduces the strength of
the interaction between local field of the nanostar and

123

J Nanopart Res (2016)18:242

the molecule. Hence, the power/intensity of singlemolecule SERS/fluorescence signal decreases drastically. The surfactant-free synthesis of the nanostars in
two-step seed-mediated method was first reported by
Yuan et al. in 2012. In Kedia et al. 2013, demonstrated
the process of reshaping the gold nanostars and its
corresponding plasmon tuning mechanism. Later,
Minati et al. (2014) have synthesized gold nanostars
by one-step reduction process. During the same
period, some of the researchers reported the synthesis
and characterization of branched gold nanocrystals
using biomolecules (Xie et al. 2007; Yi et al. 2013;
Sajitha et al. 2015). The gold nanostars synthesized in
all these works, however, lack longer and sharp spikes
in all directions, making them less effective for SERS
applications. Therefore, herein, we report a high-yield,
one-step synthesis of surfactant-free gold nanostars
using a new wet chemistry method. In addition, we
performed numerical study on single gold nanostars
using finite element method (FEM) to realize the
importance of the synthesized nanostars for singlemolecule SERS application.

Experimental section
Chemicals
Chloroauric acid (HAuCl4) (99.1 %) solution, silver
nitrate (AgNO3) (99.2 %), ascorbic acid (C6H8O6)
(99.7 %), hydrochloric acid (HCl) (3537 %) and
0.22 lm nitrocellulose membrane were purchased
from Sigma Aldrich and used as received without
any further purification. The water used throughout
this synthesis process was of reagent grade, produced
using a Milli-Q SP ultrapure water purification system
purchased from Nihon Millipore Ltd., Tokyo.
Synthesis
In this study, gold nanostars were synthesized in a
simple, seed-less, one-step, surfactant-free wet chemistry method. First of all, 10 ml of 0.25 mM HAuCl4
solution (in the presence of 10 ll of 0.1 M HCl
solution) was taken in a 20-ml glass vial. The solution
was stirred at room temperature under moderate
stirring (780 rpm). Then, 100 ll of AgNO3 solution
of 1 mM concentration and 50 ll of C6H8O6 solution
of 100 mM concentration were added simultaneously

J Nanopart Res (2016)18:242

Page 3 of 9 242

with the above HAuCl4 solution at room temperature


under moderate stirring (780 rpm). After 30 s, the
colour of the solution changed rapidly from light red to
greenish black. After that, one centrifugal wash at
5000 rcf for 15 min was performed in a 15-ml tube to
halt the nucleation. The solution was redispersed in DI
water and then filtered through a 0.22-lm nitrocellulose membrane, and the produced nanostars solution
were maintained at very low temperature for longterm storage. In this synthesis process, C6H8O6 acts as
a reducing agent, and the presence of Ag? is necessary
for the nanostar formation. The main function of Ag?
is to expedite the anisotropic growth of Au branches
on certain crystallographic facets (Yuan et al. 2012).
Without the presence of Ag?, this synthesis procedure
gives polydispersed nanorods and nanospheres. A
small amount of HCl was used to decrease the pH of
the solution in order to enable the plasmon band red
shift (Yuan et al. 2012). The synthesized nanostructures were characterized morphologically and structurally by field emission scanning electron microscope
(FE-SEM, Hitachi S-4800) and X-ray diffractometer
(XRD, Rigaku Trax-III), respectively.

core radius (rc) and spike length (ls) were estimated to


be 88.5 and 72 nm, respectively. It is noteworthy that
it is very difficult to find out the total number of spikes
of nanostars (3-dimensional in nature) exactly from
the two-dimensional SEM images.
Using energy-dispersive X-ray analysis (EDXA)
unit attached to FE-SEM, the elemental analysis has
been done for a few drops of synthesized solution by
dispersing them on carbon tape. Inset table of Fig. 2
shows the weight percentage of the elements present in
the solution. It is to be noted that the intense peak in
this spectrum is corresponding to the substrate. After
carbon, gold shows the second highest weight percentage (22.51 %) in the solution confirming that the
synthesized nanostars are of gold.
Figure 3 shows a typical XRD pattern of the
synthesized nanostructures dispersed on a silicon
wafer. The characteristic peaks corresponding to
(111), (200), (220) and (311) planes of gold are
marked accordingly after comparing the pattern with
Joint Committee on Power Diffraction Standard
(JCPDS) data. The XRD data confirm that Au present
in gold nanostar is in single elemental phase with fcc
structure as per JCPDS File No. 04-0784.

Results and discussion

FEM simulations for estimating local electric field


enhancement and LSPR wavelength

Sample characterization
The synthesized nanostructures were dispersed on a
glass slide, and SEM images were recorded. Panels
(a) and (b) of Fig. 1 show the typical SEM images at
lower and higher magnifications, respectively. It is
apparent that all the nanostars have long and sharp
spikes in all directions, and no other shapes of
nanoparticles are found other than nanostars. It is
difficult to find the exact value of the yield of
nanostars. However, based on the present observations
and the comparison with the earlier reports, it can be
safely concluded that the yield of nanostars in the
present case is well above 90 %. In order to estimate
the average values of core size and spike length of the
synthesized nanostars, the values of core size and
spike length of different nanostars in 6 different SEM
images have been estimated with the help of the
imaging processing software Gwyddion (Necas and
Klapetek 2012), and the corresponding histograms
have been shown in panels (c) and (d) of Fig. 1,
respectively. From these histograms, the average of

Although the LSPR wavelength of ensemble of


nanostructures can be easily predicted by recording
the absorption spectrum using a UVvisible spectrophotometer, such a spectrum is of less use in singlemolecule spectroscopy mainly due to the following
reason: In the experiments related to single-molecule
spectroscopy, a single molecule would adsorb on the
surface of single plasmonic nanostructure. Therefore,
the information about the local electric field enhancement (g) and localized surface plasmon resonance
(LSPR) wavelength (kLSPR) of single nanostructure is
important rather than ensemble of nanostructures. This
prompted us to estimate the g and kLSPR of single-gold
nanostars using FEM (in COMSOL Multiphysics 5.0),
which is discussed in the following.
The 3D modelling of a nanostar having 18 sharp
spikes in COMSOL is shown in panel (a) of Fig. 4.
Here, the solid core was made by a spherical shape
object, and the spikes were made by placing sharp
cones on the surface of the core. The core size and
spike length were fixed at 88.5 and 72 nm,

123

242

Page 4 of 9

J Nanopart Res (2016)18:242

177 nm (rc = 88.5 nm)

72 nm
30
25

146 nm
15

Number of spikes

Number of nanostars

20

10

88 nm

20
15
10

5
5

50

100

150

200

20

250

40

60

80

100 120

Spike length (nm)

Core size (nm)

Fig. 1 a, b Typical field emission scanning electron microscopic (FE-SEM) images of the synthesized nanostructures with lower and
higher magnifications. c, d Histograms of the core size and spike length of synthesized nanostars, respectively

Carbon substrate peak

Au

Intensity (arb. units)

Fig. 2 Energy-dispersive
X-ray analysis (EDXA)
spectrum of the sample
dispersed on carbon tape.
Inset table shows weight
percentage (%) of elements
present in the sample

Weight %
66.25

6.08

F
Na

1.13
0.23

Al

0.36

Cl
K

1.50
0.66

Ca

0.33

Ag
Au
Total

0.94
22.51
100.00

Ag
Cl

O
Na

Al Au
Au

Ag
Cl

Au

Au

Energy (keV)

123

Element
C

Au

10

J Nanopart Res (2016)18:242

Page 5 of 9 242

Intensity (arb. units)

(11 1)

(220)
(311)

(200)

20

30

40

50

60

70

80

2 (Degrees)
Fig. 3 XRD pattern of the synthesized gold nanostars

respectively, in order to correlate the theoretical


findings to the average values obtained from histograms shown in panels (c) and (d) of Fig. 1. Panel
(b) of Fig. 4 shows the enhancement in the local
electric field of a nanostar illuminated with a plane
wave of wavelength 650 nm. In this figure, it is
apparent that the electric field is enhanced drastically
at the tip of the spikes which are aligned in the plane of
vibration of incident electric field, and from these
simulations, it has been found that definitely a few hot
spots would develop irrespective of the incident light
polarization. In the present case, the expression for g
can be written as follows:
g  gr gnr ;

where gr is the resonant enhancement factor due to the


resonant excitation of localized surface plasmons.
Hence, this factor strongly depends on the size, shape,
relative dielectric permittivity of the nanostructure and
its orientation relative to the polarization of the
excitation light (Sarid and Challener 2010). gnr is the
geometric or non-resonant enhancement factor due to
an enhanced charge density localization near sharp
tips, which is popularly known as the lightning rod
effect, and this allows a nanostructure to act as an
optical antenna (Jackel and Feldmann 2012; Futamata
and Maruyama 2007). This is the primary reason for
developing larger g at the end of the spikes.
Blue curve in Panel (c) shows the variation in
maximum electric field at the tip of the spike of a
nanostar (rc = 88.5 nm and ls = 72 nm) with

excitation wavelength. Based on this curve, we can


conclude that the localized surface plasmons inside
synthesized nanostructures could be excited easily at
several wavelengths, indicating additional advantage
of synthesized gold nanostars. By comparing this
curve with the red curve, which represents the
maximum electric field values at the tip of the spike
of a nanostar of rc = 73 nm and ls = 72 nm, it is clear
that the electric field values are increased significantly,
and red shift is observed in the wavelength location of
peaks upon decreasing the rc value.
From histograms shown in panels (c) and (d) of
Fig. 1, the synthesized nanostars also have rc = 88.5
nm, ls = 88 nm and rc = 73 nm, ls = 88 nm. In order
to have the knowledge regarding the local electric field
enhancement and LSPR of the nanostars having these
dimensions, FEM simulations have been carried out.
Red and blue curves in panel (d) show the effect of
excitation wavelength on maximum electric field at
the tip of the spikes of two different nanostars. From
these curves, it is clear that these nanostars also could
be excited easily using different excitation wavelengths. In addition, the significant red shift in
wavelength locations of the peaks is observed when
the value of rc is decreased from 88.5 to 73 nm.
FEM simulations for estimating SERS
enhancement
The SERS enhancement originates due to electromagnetic enhancement effect, chemical enhancement
effect and enhancement of photonic density of states
(Kneipp et al. 2002). However, it has been found that
the electromagnetic enhancement is much larger than
those obtained from the other two effects. The
electromagnetic enhancement arises due to the
enhanced electric fields originating from the LSPR
effect on both the incident laser and Raman scattered
radiation from the molecules adsorbed on the surface
of metal nanoparticles. The theoretical SERS enhancement (n) is given by (Choi et al. 2010; Wei et al. 2008)



Eloc xL 2 Eloc xR 2

;
nxL ; xR 
2
E0 xL E0 xR
where Eloc xL is the amplitude of the local electric
field at the laser excitation frequency xL , E0 xL is the
amplitude of the incident electric field of the laser
light, Eloc xR is the amplitude of the local electric

123

242

Page 6 of 9

J Nanopart Res (2016)18:242

499

= 650 nm

450

72nm

400

350
300
177nm
250
200
150
100

50

5.96 x 10

d
1.4x10

rc = 88.5 nm, ls = 88 nm

rc = 88.5 nm, ls = 72 nm

rc = 73 nm, ls = 72 nm

2.0x10

1
8

1.5x10

1.0x10

5.0x10

1.2x10
1.0x10
8.0x10
6.0x10
4.0x10
2.0x10

rc = 73 nm, ls = 88 nm

0.0

0.0
400

Electric field enhancement

Electric field enhancement

3.0x10
2.5x10

-7

600

800

1000

1200

1400

1600

600

400

Wavelength (nm)

800

1000

1200

1400

1600

Wavelength (nm)

Fig. 4 a3D modelling of a nanostar (rc = 88.5 nm and


ls = 72 nm) having 18 spikes, in COMSOL software. bElectric
field enhancement on the surface of a gold nanostar illuminated
with the plane wave of wavelength 650 nm. c, d Variation of

maximum electric field enhancement with the optical excitation


wavelength at the tip of the spikes of 4 different nanostars. For
all the simulations, the dielectric constant values of gold have
been taken from Johnson and Christy (1972)

field at the Raman scattered frequency xR and E0 xR


is the amplitude of the electric field at xR . Equation 2
is valid as long as the frequency of the scattered
radiation lies within the LSPR band. Otherwise, the
value of second factor becomes 1, and in this
circumstance, it is almost impossible to obtain the
Raman spectrum of a single molecule experimentally.
Based on these, one can conclude that a nanoparticle
having wider LSPR band would be more suitable for
obtaining all the characteristic frequencies (vibrational modes) of a molecule. From Fig. 4, it is clear
that our synthesized nanostructures fulfil this criterion.

When the value of xL is very close to xR, then the


above expression can be simplified as follows:


Eloc xL 4
:

n
3
E x

123

The power of the Raman scattered radiation


originating from a molecule adsorbed on the surface
of a plasmonic nanostructure (or in the vicinity of
plasmonic nanostructure) due to the electromagnetic
enhancement is given by (Kneipp et al. 2002; Kneipp
and Kneipp 2006)
P(xR ) = rR I(xL ) n;

J Nanopart Res (2016)18:242

Page 7 of 9 242
8.81 x 1013

= 635.3 nm

8 x 10

1.23 x 1011
13

7 x 10

13

6 x 10

13

= 750 nm

1.0 x 10

5 x 10

13

4 x 1013
3 x 10

13

2 x 10

13

2.64 x 10-18
11

6x1011

0.8 x 10

11

0.6 x 10

11

0.4 x 10

11

1.01 x 10-25
4.74x1010

6.43x10

11

0.2 x 1011

1 x 1013

=1080 nm

11

1.2 x 10

=1210nm

10

4.5x10

4.0x1010

11

5x10

3.5x1010
10

11

3.0x10

4x10

10

3x10

11

2x10

11

2.5x10

2.0x1010
1.5x1010
10

1.0x10

1x1011

0.5x1010

-25

6.22x10-25

5.2x10

Fig. 5 The SERS enhancement factor at the end of a spike of the nanostar [shown in a of Fig. 1) calculated at 4 different wavelengths.
The values of rc and ls of this nanostar are 88.5 and 72 nm, respectively

where rR is the Raman scattering cross section of an


adsorbed molecule and I (xL ) is the intensity of
incident laser light. From Eq. 4, it is clear that the
value of P (xR ) is proportional to n. Using FEM, the
values of n have been calculated for a nanostar
(rc = 88.5 nm and ls = 72 nm) for four different
wavelengths and are shown in Fig. 5. From this figure,
it is apparent that the maximum value of n varies from
1010 to 1013 depending on the excitation wavelength.
This means that the power of the Raman scattered
radiation from a molecule in the vicinity of the
synthesized nanostar would be enhanced almost 1010
to 1013 times which makes it possible for studying the
single-molecule Raman scattering in real time.
In, Indrasekara et al. 2014 have reported the
fabrication of SERS substrates using nanostars and

demonstrated chemical sensing using these substrates


in the femto Molar regime. However, the S/N of some
of the experimental signals was found to be very poor.
This could be improved by several times easily using
our synthesized nanostars because the yield of our
nanostars is much higher as compared to the nanostars
shown in their report (4045 % of the nanostructures
were not in nanostar shape). Therefore, our future goal
is to fabricate efficient SERS substrates using these
nanostars and demonstrate the detection of single
molecules in real time in their natural state.
Conclusions
We have successfully synthesized the surfactant-free
gold nanostars in one step, using a new wet chemistry

123

242

Page 8 of 9

method. In contrast to the existing reports, these


nanostructures have longer and sharper spikes in all
directions. This could be considered as a unique
advantage of the nanostar, because at least a few hot
spots will be developed at the end of the spikes
irrespective of the polarization of incident light. To the
best of our knowledge, the yield of the synthesized
nanostars is visibly much larger than those reported in
the literature so far. Using FEM simulations, it has
been found that the localized surface plasmons inside
nanostructures could be excited using a wide range of
excitation wavelengths. The local electric field values
are found to be increased, and the red shift in
wavelength locations of the peaks is observed, on
decreasing the core size. The numerical SERS
enhancement was found to be varying from 1010 to
1013 depending on excitation wavelengths. Thus, these
nanostars are useful to fabricate efficient SERS
substrates for examining single molecules in their
natural form, which is holy-grail in bioscience and
biomedicine fields.
Acknowledgments We thank the Department of Science and
Technology, India for providing financial support.

References
Barbosa S, Agrawal A, Rodrguez-Lorenzo L, Pastoriza-Santos
I, Alvarez-Puebla RA, Kornowski A, Weller H, Liz Marzan
LM (2010) Tuning size and sensing properties in colloidal
gold nanostars. Langmuir 26:1494314950
Chirico G, Borzenkov M, Pallavicini P (2015) Gold nanostars
synthesis. Properties and biomedical application. Springer
International Publishing, New York City
Choi CJ, Xu Z, Wu H-Y, Liu GL, Cunningham BT (2010)
Surface-enhanced raman nanodomes. Nanotechnology
21:415301415308
Ciou SH, Cao YW, Huang HC, Su DY, Huang CL (2009) SERS
enhancement factors studies of silver nanoprism and
spherical nanoparticle colloids in the presence of bromide
ions. J Phys Chem C 113:95209525
Colleen LN, Liao H, Hafner JH (2006) Optical properties of
star-shaped gold nanoparticles. Nano Lett 6:683688
Futamata M, Maruyama Y (2007) Tip Enhancement. In: Kawata
S, Shalaev VM (eds). Elsevier, Amsterdam. p.50
Greeneltch NG, Blaber MG, Henry AI, Schatz GC, Van Duyne
RP (2013) Immobilized nanorod assemblies: fabrication
and understanding of large area surface-enhanced Raman
spectroscopy substrates. Anal Chem 85:22972303
Guo H, Jiang D, Li H, Xu S, Xu W (2013) Highly efficient
construction of silver nanosphere dimers on
poly(dimethylsiloxane) sheets for surface-enhanced
Raman scattering. J Phys Chem C 117:564570
Haes AJ, Van Duyne RP (2002) A nanoscale optical biosensor:
sensitivity and selectivity of an approach based on the

123

J Nanopart Res (2016)18:242


localized surface plasmon resonance spectroscopy of triangular silver nanoparticles. J Am Chem Soc
124:1059610604
Indrasekara ASDS, Meyers S, Shubeita S, Feldman LC,
Gustafssonc T, Fabris L (2014) Gold nanostar substrates
for sers-based chemical sensing in the femtomolar regime.
Nanoscale 6:88918899
Johnson PB, Christy RW (1972) Optical constants of the noble
metals. Phys Rev B 6:43704379
Jackel F, Feldmann J (2012) Complex shaped metal nanoparticles. In: Sau TK, Rogach AL (eds). Wiley, Weinheim.
p. 439
Kaneko T, Chen Q, Hatakeyama R (2011) Structural and reactive kinetics in gas-liquid interfacial plasmas. Condens
Matter 20:14011409
Kedia A, Kumar PS (2013) Controlled reshaping and plasmon
tuning mechanism of gold nanostars. J Mater Chem C
1:45404549
Kneipp K, Kneipp H (2006) Single molecule Raman scattering.
Appl Spec 60:322A334A
Kneipp K, Kneipp H, Itzkan I, Dasari RR, Feld MS (2002)
Surface-enhanced Raman scattering and biophysics. J Phys
Condens Matter 14:R597R624
Kumar PS, Pastoriza-Santos I, Rodriguez-Gonzalez B, Abajo
GD, Javier F, Liz-Marzan LM (2008) High-yield synthesis
and optical response of gold nanostars. Nanotechnology
19:0156061
Le RCL, Etchegoin PG (2009) Principles of Surface Enhance
Raman Spectroscopy. Elsevier, Amsterdam 185
Lee HM, Jin SM, Kim HM, Suh YD (2013) Single-molecule
surface-enhanced Raman spectroscopy: a perspective on
the current status. Phys Chem Chem Phys 15:52765287
Matsushita R, Kiguchi M (2015) Surface enhanced Raman
scattering of a single molecular junction. Phys Chem Chem
Phys 17:2125421260
Minati L, Benetti F, Chiappini A, Speranza G (2014) One-Step
synthesis of star-shaped gold nanoparticles. Colloids and
surfaces A: physicochem. Eng Aspects 441:623628
Necas D, Klapetek P (2012) Gwyddion: an open-source software for SPM data analysis. Cent Eur J Phys 10:181188
Sadana A (1998) Bioseperation of proteins: unfolding and
folding and validations. Academic press, London. p. 236
Sajitha M, Vindhyasarumi A, Gopi A, Yoosaf K (2015) Shape
controlled synthesis of multi-branched gold nanocrystals
through a facile one-pot bifunctional biomolecular
approach. RSC Adv 5:9831898324
Sarid D, Challener W (2010) Modern introduction to surface
plasmons: theory of mathematica modelling and applications. Cambridge University Press, Cambridge
Scarabelli L, Coronado-Puchau M, Giner-Casares JJ, Langer J,
Liz-Marzan LM (2014) Monodisperse gold nanotriangles:
size control, large-scale self-assembly, and performance in
surface-enhanced Raman scattering. ACS Nano
8:58335842
Sivapalan ST, DeVetter BM, Yang TK, Dijk TV, Schulmerich
MV, Carney PS, Bhargava R, Murphy CJ (2013) Off-resonance surface-enhanced Raman spectroscopy from gold
nanorod suspensions as a function of aspect ratio: not what
we thought. ACS Nano 7:20992105
Trigari S, Rindi A, Margheri G, Sottini S, Dellepiane G, Giorgetti E (2011) Synthesis and modelling of gold nanostars

J Nanopart Res (2016)18:242


with tunable morphology and extinction spectrum. J Mater
Chem 21:65316540
Wei H, Hao F, Huang Y, Wang W, Nordlander P, Xu H (2008)
Polarization dependence of surface-enhanced Raman
scattering in gold nanoparticle-nanowire systems. Nano
Lett 8:24972502
Wu HL, Chen CH, Huang MH (2009) Seed-mediated synthesis
of branched gold nanocrystals derived from the side growth
of pentagonal bipyramids and the formation of gold
nanostars. Chem Mater 21:110114
Xie J, Lee JY, Wang DIC (2007) Seedless, surfactantless, highyield synthesis of branched gold nanocrystals in HEPES
buffer solution. Chem Mater 19:28232830

Page 9 of 9 242
Yang L, Liu H, Ma Y, Liu J (2012) Solvent-induced hot spot
switch on silver nanorod enhanced Raman spectroscopy.
Analyst 137:15471549
Yi S, Sun L, Lenaghan SC, Wang Y, Chong X, Zhang Z, Zhang
M (2013) One-step synthesis of dendritic gold nanoflowers
with high surface-enhanced Raman scattering (SERS)
properties. RSC Adv 3:1013910144
Yuan H, Khoury CG, Hwang H, Wilson CM, Grant GA, VoDinh T (2012) Gold nanostars: surfactant-free synthesis,
3D modelling, and two-photon photoluminescence imaging. Nanotechnology 23:075102075111

123

You might also like