Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Continuous In-flight Synthesis for On-Demand Delivery of Ligand-Free Colloidal Gold

Nanoparticles

Paul Maguire*†, David Rutherford†, Manuel Macias-Montero†, Charles Mahony†, Colin Kelsey†, Mark
Tweedie†, Fátima Pérez-Martin†, Harold McQuaid†, Declan Diver‡, Davide Mariotti†.

NIBEC, University of Ulster, Belfast, BT37 0QB, N. Ireland

SUPA, School of Physics and Astronomy, University of Glasgow, Glasgow G12 8QQ, United Kingdom

*corresponding author: pd.maguire@ulster.ac.uk

SI-1: Setup & Operational Parameters


Details of the experimental setup have been reported previously.1 A schematic and images of the
setup are shown in FIG. S 1. A Burgener Mira Mist X-175 nebuliser, driven by neon gas flows up to
1.0 sLm via a mass flow controller, was used to deliver a continuous stream of near mono-sized
droplets to the plasma region. The liquid (H2O + 1 mM HAuCl4) was supplied by a syringe pump at a
flow rate of 10 μl/min. The divergent (20°) aerosol output of the nebuliser was interfaced to a 2 mm
inner diameter quartz tube via a custom-designed acrylic manifold, which focussed the aerosol/Ne
gas flow using a concentric helium gas curtain which allows controlled gas flows up to 4 slm. The
aerosol droplet transport efficiency was optimized by varying the water, neon, and helium gas flows.
All experiments were conducted with gas flows of 3.5 slm (He) and 1.0 slm (Ne).

The He/Ne gas stream transported the droplets along the tube to the plasma region formed by
concentric ring electrodes at the outer wall and driven by rf power. The electrode position is variable
and the set power (at supply) was 80W at 13.56 MHz. For all experiments, the length of the plasma
region between electrodes was ∼2 mm with a further ∼10 mm to tube exit. Low resolution images
show a parallel droplet spray at the tube outlet, indicating persistent laminar flow beyond the
capillary, with a slight reduction in spray diameter when the plasma is turned on.

FIG. S 1 Reactor and plasma assembly


Images and schematic of plasma setup consisting of ring electrodes around 2 mm diameter quartz tube and nebuliser
interface. (a) Segment of plasma-aerosol assembly showing the quartz tube and the shielded plasma enclosure below
which the location of the TEM grid (not shown) is indicated; (b) inside the shielded plasma enclosure where the plasma
region between the two ring electrodes is visible; (c) assembly schematic.
Plasma treated droplets were collected on TEM grids, exposed to the droplet stream for 30
seconds. The grids were located ~10mm downstream of the quartz tube exit i.e. ~20 mm
downstream of the plasma region. For UV-vis analysis, 30 minutes of plasma treated droplets (0.3
mL) were collected in 2mL of deionised water. For these experiments, the quartz tube was
extended, using PTFE tubing, into the collection liquid within a closed cuvette, to limit losses.

Size, velocity measurement


A Questar QM-100 microscope based on the Maksutov Cassegrain Telescope design and capable
of x34 magnification at the image plane, and an image intensified Andor iStar DH734-18U-03
camera were used, FIG. S 2, to obtain high resolution images of in-flight droplets, at various
distances from the quartz tube outlet, FIG. S 1(c). The magnification range was x6 - x50, with a
working distance of 160 mm and an exposure time of 10 µs – 20 µs. The depth of field was
measured at 120 m, centred on the tube axis. All out of focus droplets were ignored. The camera
gating speed was < 2 ns and capable of capturing droplet speeds up to 60 ms -1. A fixed
magnification of 1.73 m per pixel was used throughout. The measured droplet velocity and size
distributions are shown in FIG. S 3. 1

(a) Optical configuration for the in-flight measurement of (b) Image showing droplet streak, from which velocity, size
plasma-exposed droplet velocity, size and location at the and position can be obtained.
exit of the quartz tube.
FIG. S 2 Optical setup

(a) Ratio of measured individual droplet velocity to the (b) Log probability graph of ln(diameter in m) vs probit
average gas speed (17 ms-1) in 3.5 slm (He) and 1.0 slm values obtained from cumulative frequency plots. The
(Ne) gas flows. ln(count median diameter) is determined from the intercept
value on the x = 0 axis. Inset: droplet size distribution.
FIG. S 3 Droplet size and velocity
SI-2: Precursor reduction
Au NP TEM images
Plasma reduction of HAuCl4 precursor
Theoretical and experimental analysis of gold colloid UV-vis spectra of has been used to determine
nanoparticle size and concentration. Direct determination of nanoparticle size from UV-vis spectra is
possible, under appropriate conditions, using the surface plasmon resonance (SPR) peak 2,3 while
Liu et al.4 provide experimental values of extinction coefficient versus diameter at 506 nm. Haiss et
al.5 developed a model using Mie theory from which extinction coefficient dependence on diameter
at 450 nm is derived and experimentally verified. To avoid interference from the SPR peak, Hendel
et al. have proposed the use of absorption at 400 nm to determine Au colloid characteristics. 6 At
this wavelength, absorption is due to the allowed direct interband transition energies from 5d to 6sp
in bulk gold. The extent of the precursor reduction over the exposure time represents the conversion
rate to metal and hence the rate of nanoparticle synthesis. The UV−vis spectrum of HAuCl4 shows
two bands at around 220 nm and 300 nm due to the [AuCl4]− ligand – metal charge transfer (LMCT)
band.7-9 The exact peak wavelengths can depend on ligand exchange involving surfactants and
solution pH. These have been used in plasma synthesis studies to estimate the amount of precursor
reduction.10

3
(a) Plasma treated, dilute
HAuCl4 0.1mM
HAuCl4 0.5mM
2
HAuCl4 1mM
Absorbance

0
200 300 400 500
Wavelength (nm)

4
(b) 217nm
y = 18.561x y = 2.7477x
R² = 0.9859 300nm
R² = 0.9877
3 400nm
450nm
Absorbance

1 y = 0.479x
R² = 0.9938
y = 0.0299x
0 R² = 0.8746
0.0 0.5 1.0
Conc HAuCl4 (mM)

FIG. S 4 UV-vis spectra for HAuCl4


(a) UV-vis spectra for plasma treated droplets with 1 mM HAuCl4, and diluted (6.67:1) in collecting liquid. Also shown are
precursor calibration spectra without plasma treatment. (b) Absorbance vs concentration calibration curves and fitting
parameters for HAuCl4 at wavelengths 217 nm, 300 nm, 400 nm and 450 nm.

The plasma treated droplets are collected over a 30 min period and the total sample volume per run
is 0.3 mL which is collected in 2 mL H2O. The UV-vis spectra were obtained immediately afterwards
and then analysed at a number of wavelengths to estimate nanoparticle concentration and the
amount of precursor reduction. Figure FIG. S 4a shows the UV-vis spectra for plasma treated
droplets and precursor solutions at various concentrations without plasma treatment and in Figure
FIG. S 4b the HAuCl4 absorbance vs concentration calibration curves at wavelengths 217 nm, 300
nm, 400 nm and 450 nm are given. With an average nanoparticle size of 4.4 nm, from TEM, we
obtain values for the nanoparticle colloid extinction coefficient of 6.8 x106 M-1 cm-1 (506 nm)4 and
4.6 x106 M-1 cm-1 (450 nm)5 while the Au0 molar extinction coefficient is 2.29 M-1 cm-1 (400 nm)6. The
contribution of HAuCl4 to the spectra at wavelengths above 400 nm is < 10%, assuming 50%
reduction, and is subtracted. Estimates of the amount of HAuCl4 reduction are 50% (506 nm), 90%
(450 nm) and 75% (400 nm) from the colloidal characteristics while from the precursor absorbance
bands, the estimates are 66% (300 nm) and 85% (217 nm). Given the small size of the
nanoparticles, i.e. below the plasmon size threshold, the limitations of sample size from the droplet
stream and the high dilution required during collection, the sample absorption is therefore low and it
is not surprising that estimates of precursor reduction vary between 50% and 90%. We therefore
assume the lower bound value of 50% reduction in subsequent analysis.

H2O2 reduction of HAuCl4 precursor


We have previously observed DC plasma synthesis of gold nanoparticles in bulk liquid where the
evidence clearly suggests that reduction by H 2O2 is an important factor.11,12 The appearance of the
Au SPR peak, indicative of Au NP growth above a threshold size, happens over a timescale of a few
minutes and is complete by 10 min. H2O2 levels of ~0.3 mM were measured after 10 min plasma
processing. By scavenging the plasma produced H 2O2 in the liquid we observe no nanoparticle
growth. This demonstrates the importance, of H2O2 in the reduction of the gold precursor, for this
particular plasma configuration.

Chen et al.13 observed plasma synthesis of Au NPs in liquid, via a mechanism they attribute to
HAuCl4 reduction by H2O2. They used a bipolar pulsed plasma at reduced gas pressure and the
evolution of the Au SPR peak occurred over a number of hours. The kinetics of Au NP growth by
direct H2O2 reduction of HAuCl4 have been reported by Pacławski et al. 14 at high pH (>13) and via
the use of pre-loaded Au NP seeds, at lower pH, by Liu et al. 15 In both cases, high concentrations of
H2O2 were used and the reduction timescales were of the order of at least 60 s. At low pH, without
seeds, Pacławski et al.14 observed no growth while Liu et al. 15 reported growth timescales of
~2000 s. Although the average pH in the plasma synthesis case is 3 – 5, simulations of plasma-
induced chemical reactions suggest local enhancement of the pH may occur in the liquid interface
region adjacent to the plasma.16

In plasma – droplet synthesis, the measured average plasma generated H2O2 is ~ 10 mM. It is now
accepted that H2O2 is primarily generated in the gas / plasma region and subsequently dissolves
into the liquid.17,18 A local change in pH at the droplet surface is also possible, whereby electron
reactions yield an excess of OH- while H2O2 production increases acidity.19 The effective H2O2
surface concentration and penetration depth during droplet flight depends on the surface flux and
the advection rate within the droplet. For the case of simple diffusion, without convection, H2O2 will
diffuse ~800 nm during the plasma flight time and the average surface layer concentration increases
to ~30 mM. From solubility considerations, the H2O2 surface concentration will be limited by the gas
phase density of OH via20

OHgas + OHgas +M  (H2O2)gas + M (k350K = 5.12 x10-12 cm-3 s-1)

where M is a third body, typically the inert He but possibly also a water cluster or molecule. The
competitive gas phase reactions are

OHgas + OHgas  H2Ogas + Ogas (k350K = 1.4 x10-12 cm-3 s-1)


OHgas + Ogas  O2 gas + Hgas (k350K = 32.9 x10-12 cm-3 s-1)

Direct OH measurements in RF plasmas indicate peak OH densities of ~2 x1020 m-3.20,21 At this OH


density, assuming all available OHgas recombines to form (H2O2)gas, and with a Henry constant of
1.07 M Pa-1, the maximum surface (H2O2)liq concentration is ~0.45 M. For an average droplet
concentration of 10 mM, this requires a surface layer depth > 50 nm. The maximum solubility may
be lower than this since Henry’s Law applies only at low concentrations. Also high concentration
H2O2 tends to dissociate into H2O and O2. Note that the above analysis also ignores the possibility
of droplet mixing caused by its interaction with the gas flow. Droplet mixing is likely in order to
supply the surface with precursor complexes as their diffusion rate is insufficient over the flight time.
Nevertheless, in considering the possibility of precursor reduction by H2O2, we ignore mixing and
assume an enhanced surface concentration.

In order to explore the possibility of reduction by plasma generated H2O2, we have investigated the
direct reduction without plasma of HAuCl4 by H2O2 for a range of concentrations. The HAuCl4
reduction time and the Au NP synthesis time were determined from UV-vis measurements at
300 nm and 565 nm, FIG. S 5. These wavelengths correspond to the decrease in the HAuCl4 peak
and the appearance of the SPR peak, respectively. For a high concentration of H2O2 (6.7 M), i.e. ten
times the maximum possible concentration at the liquid surface, the SPR peak appeared at 565 nm
after ~100 s and reached its maximum intensity at 340 s while the absorbance of the precursor
reached its minimum around a similar time. The decrease in the SPR peak beyond the maximum is
likely due to NP precipitation. At this concentration (6.7 M H2O2), the reduction time for 1 mM
HAuCl4, as obtained from UV-vis analysis at 300 nm, was at least 200 s. With lower H2O2
concentration (0.1 M), no SPR peak was detected up to 1200 s.

1.5
6.7 M, 300nm
6.7 M, 565 nm 0.02
Absorbance 310 nm

Absorbance 565 nm
1.0

0.01
0.5

0.0 0.00
0 250 500 750 1000
Time (s)

FIG. S 5 Absorbance vs time


Absorbance versus time for HAuCl4 precursor (300 nm) and Au NP SPR peak (565 nm) for a H2O2 concentration of 6.7 M.

Assuming a first order reaction with HAuCl4, i.e. that the concentration of H2O2 is not significantly
diminished over this reaction period, we estimate the initial rate constant for HAuCl 4 reduction to be
0.005 s-1, FIG. S 6.

The timescales required for significant precursor reduction by H2O2 appear to be much too long
compared to the flight time between the plasma region and the TEM collection grid, which is < 1 ms.
The appearance of an Au SPR peak was not observed for plasma-treated droplets collected in liquid
and stored for many hours. Since Pacławski et al.3 and Liu et al.4 both demonstrate the importance
of (auto-) catalytic processes in the reduction of HAuCl4 by H2O2, we believe that at the initial stages
of plasma synthesis, small nanoparticles and clusters are formed which, if given sufficient time, can
facilitate subsequent growth via H2O2 reduction. Therefore in the plasma droplet synthesis we see
only the initial small sized nanoparticles due to the very short exposure time. However with bulk
liquid synthesis, there is sufficient time for seed catalysed growth into the much larger nanoparticles
we observe. We can conclude therefore that the role of H 2O2 in precursor reduction is not significant
in this case.
1.5

0.02

Absorbance 565 nm
-ln([HAuCl4] ) mM

1.0
ln[HAuCl4]/[HAuCl4]0
6.7 M 565 nm
0.01
0.5

y = 0.0052x
R² = 0.898
0.0 0.00
0 500 1000
Time (s)
FIG. S 6 Rate coefficient
Plot of log reduction of [HAuCl4]/ [HAuCl4]0 with time obtained from absorbance at 300nm. The precursor reduction rate of
-1
0.005 s is estimated from a linear fit to the curve over the initial 200 s. The concentration of H2O2 was 6.7M while that of
the HAuCl4 precursor was 1 mM.

SI-3: Nanoparticle synthesis rates


Colloid chemistry
The nanoparticle synthesis rate can be obtained from the reduction of the precursor and the plasma
exposure time.

FIG. S 7 NP synthesis rate vs reactor volume


(a) Nanoparticle synthesis rate of plasma-exposed droplets, determined from the rate of reduction of
HAuCl4, compared with equivalent rates extracted from published literature for different reactor
architectures and volumes, namely (b) and (c) microfluidic reactors, (d) turbulent capillary flow
device, (e) – (i) stopped or continuous flow devices, (j) gas-segmented flow device, (k) laminar
capillary flow device and (l) – (n) batch reactor processes.

For a 1 mM HAuCl4 initial precursor concentration and 50% reduction after ~120 s exposure time
for each droplet, the Au3+ conversion rate to Au0 metal atoms that contribute to NP growth, i.e. the
NP synthesis rate, is ~1024 atoms s-1 L-1 . This is significantly higher, by many orders of magnitude,
than observed with traditional colloidal chemistry processes. A comparison is made with
representative literature reports for precursor reduction / nanoparticle synthesis rates via colloidal
chemistry processes involving a number of reactor architectures and volumes, FIG. S 7. This figure
is an annotated version of Figure 3 in the main manuscript. Note that division by Avogardo’s No.
converts the unit atoms s-1 L-1 to M s-1.

Comparison data is obtained from either the reported fraction of precursor reduced within the given
reaction time or in (e) the NP concentration along with average diameter and reaction time. Where
the reaction is non-linear, the fastest rate has been chosen. Brief details and references are
provided in Table S1.
FIG. S Reference Details Vol Reaction [precursor] Rate
7 (mL) time (s) (M) (atom s-1 L-1)
b Wagner22,23 Ascorbic acid: IPHT Jena mixer chip; 8 x10-3 0.06 1 x10-3 9 x1021
c Tsunoyama 22,23 NaBH4/PVP SIMM-V2 mixer chip 8 x10-3 0.14 1 x10-3 4 x1021
24 -4 20
d Han. Millifluidic tubing, turbulent mixing 0.1 1 5 x10 3 x10
f Ftouni 25 Turkevich, stopped flow 0.05 35 5 x10-3 9 x1019
26 o -5 18
g Paclawski Glucose; rate equation (5), 35 C 0.25 4 4 x10 6 x10
27
h Luty-Blocho Ascorbic acid; rate eqn (8) 0.25 440 1 x10-4 1 x1017
28 -5 19
i Wojnicki DMAB (BH4) Stopped flow, fig 5 0.25 0.2 2.5 x10 7 x10
29 -3 19
j Sebastian NaBH4: Segmented flow, air-water 0.1 10 1 x10 6 x10
30 -2 17
k Lohse Lohse millifluidic capillary, laminar 150 180 1 x10 3 x10
31,32 -4 16
l Polte Turkevich, Fig 2(h) Phase III 14000 3600 2.5 x10 4 x10
33 -4 18
m Wuithschick Turkevich, high conc precursor 200 180 5 x10 2 x10
34 -5 17
n Streszewski Hydrazine sulphate: rate eqn 13 1000 90 1 x10 1 x10
35 -4 20
o Polte NaBH4: continuous flow: complete 0.15 0.2 3 x10 9 x10
0
reduction to Au .
Vol Reaction [NP] dNP Rate
-1 -1
(mL) time (s) (M) (nm) (atom s L )
e Abécassis36,37 TBAB: Stopped flow mixer, fig 3 0.2 0.5 2x10-7 1.2 2 x1020
Table S1 Chemical synthesis rate comparison details

Radiolysis, electron beam

FIG. S 8 Nanoparticle synthesis rate versus dose rate


(a) Nanoparticle synthesis rate against dose rates of plasma-exposed droplets, determined from the rate of reduction of
HAuCl4, compared with equivalent rates extracted from published literature for (b) – (j) steady-state gamma radiolysis, (k)
– (m) pulsed electron beam and (n) – (q) high dose rate TEM electron beam into liquid cell The dose rate is calculated
from the measured power deposited into the plasma x ratio of droplet to plasma volume.

The plasma delivers a high flux of low energy ions and electrons to the droplet surface during flight
through the plasma. The latter are highly reducing and comparison with other solvated electron
reduction techniques are given in FIG. S 8, where the rate of reduction to Au 0 (i.e. NP synthesis
rate) is plotted against dose rate (kGy s-1), a standard metric within radiation chemistry. The dose
rate can also be expressed as deposited power per unit mass, with equivalent units of kJ kg -1 s-1.
For steady-state gamma radiation FIG. S 8 (b) – (j) and pulsed electron beam studies FIG. S 8 (k) –
(m), where a reaction time is not given, the NP synthesis rate is obtained from the value of dose at
the absorbance plateau, indicating end of reaction, and the given dose rate and precursor
concentration. Alternatively dose, dose rate, NP size and concentration are used. With regard to
TEM beam studies, FIG. S 8 (n) – (q); in Alloyeau et al.38 the NP growth rate is determined from a
linear fit to the observed radius versus time graphs, in figure S3, for each dose rate, while in Park et
al.39 the dose rate is estimated from the beam current and stopping power and the NP synthesis
rate is obtained from the given Au 0 precipitation rate and an estimate of the interaction volume
between beam and liquid.

label Reference Details Dose Rate Rate


kGy s-1 (atom L-1 s-1)
(kJ kg-1 s-1)
b Dey40 60
Co CW γ-rays: Methanol, fig2 (inset) 5 x10-4 2 x1016
41 60 -4 16
c Vo Co CW γ-rays: Au, chitosan 3 x10 2 x10

d Dey42 60
Co CW γ-rays: 2-propanol, fig2b(inset) 5 x10-4 3 x1016

Dey43
60
e Co CW γ-rays: 2-propanol 5 x10-4 4 x1016

Hanzic44
60
f Co CW γ-rays: Au/citrate, N2, long exposure, Table 1 2 x10-3 8 x1016

Gascard45
60
g Co CW γ-rays: 2-propanol, fig 2 3 x10-3 1 x1017
h Hanzic44 60
Co CW γ-rays: Au/citrate, N2, Table 1 3 x10-3 5 x1017

Meyre46
137
i Cs CW γ-rays: Nanoreactor, 5 x10-4 6 x1017
Hanzic44
60
j Co CW γ-rays: Au/citrate, air. Table 1 2x10-3 7 x1017

k Tregeur47 Pulsed electron beam (10 MeV–20 kW, 350Hz) 2 x100 9 x1019
45 0 20
l Gascard Pulsed electron beam (10 MeV–20 kW, 350Hz) 2 x10 2 x10
m Vo41 Pulsed electron beam (10 MeV–20 kW, 500Hz), fig 3 1.5 x101 2 x1021
n Alloyeau38 200kV TEM electron beam, liquid cell, low dose, fig S3 5 x102 2 x1019

o Alloyeau38 200kV TEM electron beam, liquid cell, mid- dose, fig S3 1 x103 1 x1020
38 3 20
p Alloyeau 200kV TEM electron beam, liquid cell, high dose, fig S3 4 x10 2 x10
39,48,49 5 21
q Park 300kV TEM electron beam, liquid cell 1 x10 7 x10

Table S2 Radiolysis and electron beam synthesis details for FIG. S 8

References

(1) Maguire, P. D.; Mahony, C. M. O.; Kelsey, C. P.; Bingham, A. J.; Montgomery, E. P.; Bennet, E.
D.; Potts, H. E.; Rutherford, D. C. E.; McDowell, D. A.; Diver, D. A.; Mariotti, D. Controlled
microdroplet transport in an atmospheric pressure microplasma. Appl. Phys. Lett. 2015,
106 "DOI:"10.1063/1.4922034.

(2) Jain, P. K.; Lee, K. S.; El-Sayed, I. H.; El-Sayed, M. A. Calculated absorption and scattering
properties of gold nanoparticles of different size, shape, and composition: Applications in
biological imaging and biomedicine. J Phys Chem B 2006, 110, 7238-
7248 "DOI:"10.1021/jp057170o.

(3) Khlebtsov, N. G. Determination of size and concentration of gold nanoparticles from extinction
spectra. Anal. Chem. 2008, 80, 6620-6625 "DOI:"10.1021/ac800834n.

(4) Liu, X.; Atwater, M.; Wang, J.; Huo, Q. Extinction coefficient of gold nanoparticles with different
sizes and different capping ligands. Colloids and Surfaces B: Biointerfaces 2007, 58, 3-
7 "DOI:"10.1016/j.colsurfb.2006.08.005.

(5) Haiss, W.; Thanh, N. T. K.; Aveyard, J.; Fernig, D. G. Determination of size and concentration of
gold nanoparticles from UV-Vis spectra. Anal. Chem. 2007, 79, 4215-
4221 "DOI:"10.1021/ac0702084.

(6) Hendel, T.; Wuithschick, M.; Kettemann, F.; Birnbaum, A.; Rademann, K.; Polte, J. In situ
determination of colloidal gold concentrations with uv-vis spectroscopy: Limitations and
perspectives. Anal. Chem. 2014, 86, 11115-11124 "DOI:"10.1021/ac502053s.

(7) Bai, T.; Tan, Y.; Zou, J.; Nie, M.; Guo, Z.; Lu, X.; Gu, N. AuBr2–-Engaged Galvanic Replacement
for Citrate-Capped Au–Ag Alloy Nanostructures and Their Solution-Based Surface-Enhanced
Raman Scattering Activity. The Journal of Physical Chemistry C 2015, 119, 28597-
28604 "DOI:"10.1021/acs.jpcc.5b10095.

(8) Sharma, V.; Park, K.; Srinivasarao, M. Colloidal dispersion of gold nanorods: Historical
background, optical properties, seed-mediated synthesis, shape separation and self-assembly.
Mater. Sci. Eng. R-Rep. 2009, 65, 1-38 "DOI:"10.1016/j.mser.2009.02.002.

(9) Esumi, K.; Suzuki, A.; Yamahira, A.; Torigoe, K. Role of poly (amidoamine) dendrimers for
preparing nanoparticles of gold, platinum, and silver. Langmuir 2000, 16, 2604-2608.

(10) Liang, X.; Wang, Z. -.; Liu, C. -. Size-Controlled Synthesis of Colloidal Gold Nanoparticles at
Room Temperature Under the Influence of Glow Discharge. Nanoscale Res. Lett. 2010, 5, 124-
129 "DOI:"10.1007/s11671-009-9453-0.

(11) Mariotti, D.; Patel, J.; Švrček, V.; Maguire, P. Plasma-liquid interactions at atmospheric
pressure for nanomaterials synthesis and surface engineering. Plasma Processes Polym.
2012, 9, 1074-1085 "DOI:"10.1002/ppap.201200007.

(12) Patel, J.; Němcová, L.; Maguire, P.; Graham, W. G.; Mariotti, D. Synthesis of surfactant-free
electrostatically stabilized gold nanoparticles by plasma-induced liquid chemistry.
Nanotechnology 2013, 24 "DOI:"10.1088/0957-4484/24/24/245604.

(13) Chen, Q.; Kaneko, T.; Hatakeyama, R. Reductants in Gold Nanoparticle Synthesis Using Gas -
Liquid Interfacial Discharge Plasmas. Applied Physics Express 2012, 5, 086201.

(14) Paclawski, K.; Fitzner, K. Kinetics of reduction of gold(III) complexes using H2O2. Metall.
Mater. Trans B 2006, 37, 703-714 "DOI:"10.1007/s11663-006-0054-3.

(15) Liu, X.; Xu, H.; Xia, H.; Wang, D. Rapid seeded growth of monodisperse, quasi-spherical,
citrate-stabilized gold nanoparticles via H2O2 reduction. Langmuir 2012, 28, 13720-
13726 "DOI:"10.1021/la3027804 [doi].

(16) Gopalakrishnan, R.; Kawamura, E.; Lichtenberg, A. J.; Lieberman, M. A.; Graves, D. B.
Solvated electrons at the atmospheric pressure plasma - water anodic interface. J. Phys. D
2016, 49, 295205.

(17) Winter, J.; Tresp, H.; Hammer, M. U.; Iseni, S.; Kupsch, S.; Schmidt-Bleker, A.; Wende, K.;
Dünnbier, M.; Masur, K.; Weltmann, K. -.; Reuter, S. Tracking plasma generated H2O2 from
gas into liquid phase and revealing its dominant impact on human skin cells. J. Phys. D 2014,
47 "DOI:"10.1088/0022-3727/47/28/285401.
(18) Gorbanev, Y.; O'Connell, D.; Chechik, V. Non-Thermal Plasma in Contact with Water: The
Origin of Species. Chem. Eur. J. 2016, 22, 3496-3505 "DOI:"10.1002/chem.201503771.

(19) Rumbach, P.; Witzke, M.; Sankaran, R. M.; Go, D. B. Decoupling interfacial reactions between
plasmas and liquids: Charge transfer vs. plasma neutral reactions. J. Am. Chem. Soc. 2013,
135, 16264-16267 "DOI:"10.1021/ja407149y.

(20) Benedikt, J.; Schröder, D.; Schneider, S.; Willems, G.; Pajdarová, A.; Vlcek, J.; Schulz-von der
Gathen, V. Absolute OH and O radical densities in effluent of a He/H2O micro-scaled
atmospheric pressure plasma jet. Plasma Sources Sci. Technol. 2016, 25,
045013 "DOI:"10.1088/0963-0252/25/4/045013.

(21) Bruggeman, P.; Cunge, G.; Sadeghi, N. Absolute OH density measurements by broadband UV
absorption in diffuse atmospheric-pressure He - H2O RF glow discharges. Plasma Sources Sci.
Technol. 2012, 21, 035019.

(22) Wagner, J.; Köhler, J. M. Continuous synthesis of gold nanoparticles in a microreactor. Nano
Lett. 2005, 5, 685-691 "DOI:"10.1021/nl050097t.

(23) Tsunoyama, H.; Ichikuni, N.; Tsukuda, T. Microfluidic synthesis and catalytic application of pvp-
stabilized, ∼1 nm gold clusters. Langmuir 2008, 24, 11327-11330 "DOI:"10.1021/la801372j.

(24) Jun, H.; Fabienne, T.; Florent, M.; Coulon, P. -.; Nicolas, M.; Olivier, S. Understanding of the
size control of biocompatible gold nanoparticles in millifluidic channels. Langmuir 2012, 28,
15966-15974 "DOI:"10.1021/la303439f.

(25) Ftouni, J.; Penhoat, M.; Addad, A.; Payen, E.; Rolando, C.; Girardon, J. -. Highly controlled
synthesis of nanometric gold particles by citrate reduction using the short mixing, heating and
quenching times achievable in a microfluidic device. Nanoscale 2012, 4, 4450-
4454 "DOI:"10.1039/c2nr11666a.

(26) Paclawski, K.; Streszewski, B.; Jaworski, W.; Luty-Blocho, M.; Fitzner, K. Gold nanoparticles
formation via gold(III) chloride complex ions reduction with glucose in the batch and in the flow
microreactor systems. Colloids Surf. A Physicochem. Eng. Asp. 2012, 413, 208-
215 "DOI:"10.1016/j.colsurfa.2012.02.050.

(27) Luty-Blocho, M.; Paclawski, K.; Jaworski, W.; Streszewski, B.; Fitzner, K. Kinetic studies of gold
nanoparticles formation in the batch and in the flow microreactor system. Prog Colloid Polym
Sci 2010,138, 39-44. "ddx "10.1007/978-3-642-19038-4_7.

(28) Wojnicki, M.; Rudnik, E.; Luty-Blocho, M.; Paclawski, K.; Fitzner, K. Kinetic studies of gold(III)
chloride complex reduction and solid phase precipitation in acidic aqueous system using
dimethylamine borane as reducing agent. Hydrometallurgy 2012, 127-128, 43-
53 "DOI:"10.1016/j.hydromet.2012.06.015.

(29) Sebastian Cabeza, V.; Kuhn, S.; Kulkarni, A. A.; Jensen, K. F. Size-controlled flow synthesis of
gold nanoparticles using a segmented flow microfluidic platform. Langmuir 2012, 28, 7007-
7013 "DOI:"10.1021/la205131e.

(30) Lohse, S. E.; Eller, J. R.; Sivapalan, S. T.; Plews, M. R.; Murphy, C. J. A simple millifluidic
benchtop reactor system for the high-throughput synthesis and functionalization of gold
nanoparticles with different sizes and shapes. ACS Nano 2013, 7, 4135-
4150 "DOI:"10.1021/nn4005022.

(31) Polte, J.; Ahner, T. T.; Delissen, F.; Sokolov, S.; Emmerling, F.; Thünemann, A. F.; Kraehnert,
R. Mechanism of gold nanoparticle formation in the classical citrate synthesis method derived
from coupled in situ XANES and SAXS evaluation. J. Am. Chem. Soc. 2010, 132, 1296-
1301 "DOI:"10.1021/ja906506j.

(32) Polte, J.; Erler, R.; Thünemann, A. F.; Emmerling, F.; Kraehnert, R. SAXS in combination with a
free liquid jet for improved time-resolved in situ studies of the nucleation and growth of
nanoparticles. Chem. Commun. 2010, 46, 9209-9211 "DOI:"10.1039/c0cc03238g.

(33) Wuithschick, M.; Birnbaum, A.; Witte, S.; Sztucki, M.; Vainio, U.; Pinna, N.; Rademann, K.;
Emmerling, F.; Kraehnert, R.; Polte, J. Turkevich in New Robes: Key Questions Answered for
the Most Common Gold Nanoparticle Synthesis. ACS Nano 2015, 9, 7052-
7071 "DOI:"10.1021/acsnano.5b01579.

(34) Streszewski, B.; Jaworski, W.; Paclawski, K.; Csapó, E.; Dékány, I.; Fitzner, K. Gold
nanoparticles formation in the aqueous system of gold(III) chloride complex ions and hydrazine
sulfate-Kinetic studies. Colloids Surf. A Physicochem. Eng. Asp. 2012, 397, 63-
72 "DOI:"10.1016/j.colsurfa.2012.01.031.

(35) Polte, J.; Erler, R.; Thünemann, A. F.; Sokolov, S.; Ahner, T. T.; Rademann, K.; Emmerling, F.;
Kraehnert, R. Nucleation and growth of gold nanoparticles studied via in situ small angle X-ray
scattering at millisecond time resolution. ACS Nano 2010, 4, 1076-
1082 "DOI:"10.1021/nn901499c.

(36) Abécassis, B.; Testard, F.; Spalla, O.; Barboux, P. Probing in situ the nucleation and growth of
gold nanoparticles by small-angle X-ray scattering. Nano Lett. 2007, 7, 1723-
1727 "DOI:"10.1021/nl0707149.

(37) Abécassis, B.; Testard, F.; Kong, Q.; Francois, B.; Spalla, O. Influence of monomer feeding on
a fast gold nanoparticles synthesis: Time-resolved XANES and SAXS experiments. Langmuir
2010, 26, 13847-13854 "DOI:"10.1021/la1020274.

(38) Alloyeau, D.; Dachraoui, W.; Javed, Y.; Belkahla, H.; Wang, G.; Lecoq, H.; Ammar, S.; Ersen,
O.; Wisnet, A.; Gazeau, F.; Ricolleau, C. Unravelling Kinetic and Thermodynamic Effects on the
Growth of Gold Nanoplates by Liquid Transmission Electron Microscopy. Nano Lett. 2015, 15,
2574-2581 "DOI:"10.1021/acs.nanolett.5b00140.

(39) Park, J. H.; Schneider, N. M.; Grogan, J. M.; Reuter, M. C.; Bau, H. H.; Kodambaka, S.; Ross,
F. M. Control of Electron Beam-Induced Au Nanocrystal Growth Kinetics through Solution
Chemistry. Nano Lett. 2015, 15, 5314-5320 "DOI:"10.1021/acs.nanolett.5b01677.

(40) Dey, G. R. Reduced Au3+ species in methanol in presence of iodide ions: A radiation chemical
study. Radiat. Phys. Chem. 2014, 102, 44-48 "DOI:"10.1016/j.radphyschem.2014.04.018.

(41) Vo, K. D. N.; Kowandy, C.; Dupont, L.; Coqueret, X.; Hien, N. Q. Radiation synthesis of
chitosan stabilized gold nanoparticles comparison between e- beam and γ irradiation. Radiat.
Phys. Chem. 2014,94, 84-87. "ddx "10.1016/j.radphyschem.2013.04.015.

(42) Dey, G. R.; El Omar, A. K.; Jacob, J. A.; Mostafavi, M.; Belloni, J. Mechanism of trivalent gold
reduction and reactivity of transient divalent and monovalent gold ions studied by gamma and
pulse radiolysis. J Phys Chem A 2011, 115, 383-391 "DOI:"10.1021/jp1096597.

(43) Dey, G. R. Gold and gold-copper nanoparticles in 2-propanol: A radiation chemical study.
Radiat. Phys. Chem. 2011,80, 1216-1221. "ddx "10.1016/j.radphyschem.2011.06.006.

(44) Hanžić, N.; Jurkin, T.; Maksimović, A.; Gotić, M. The synthesis of gold nanoparticles by a
citrate-radiolytical method. Radiat. Phys. Chem. 2015, 106, 77-
82 "DOI:"10.1016/j.radphyschem.2014.07.006.
(45) Gachard, E.; Remita, H.; Khatouri, J.; Keita, B.; Nadjo, L.; Belloni, J. Radiation-induced and
chemical formation of gold clusters. New J. Chem. 1998, 22, 1257-1265.

(46) Meyre, M. -.; Tréguer-Delapierre, M.; Faure, C. Radiation-induced synthesis of gold


nanoparticles within lamellar phases. Formation of aligned colloidal gold by radiolysis.
Langmuir 2008,24, 4421-4425. "ddx "10.1021/la703650d.

(47) Treguer, M.; de Cointet, C.; Remita, H.; Khatouri, J.; Mostafavi, M.; Amblard, J.; Belloni, J.; de
Keyzer, R. Dose Rate Effects on Radiolytic Synthesis of Gold-Silver Bimetallic Clusters in
Solution. The Journal of Physical Chemistry B 1998, 102, 4310
4321 "DOI:"10.1021/jp981467n.

(48) Grogan, J. M.; Schneider, N. M.; Ross, F. M.; Bau, H. H. Bubble and pattern formation in liquid
induced by an electron beam. Nano Lett. 2014, 14, 359-364 "DOI:"10.1021/nl404169a.

(49) Schneider, N. M.; Norton, M. M.; Mendel, B. J.; Grogan, J. M.; Ross, F. M.; Bau, H. H. Electron-
Water interactions and implications for liquid cell electron microscopy. J. Phys. Chem. C 2014,
118, 22373-22382 "DOI:"10.1021/jp507400n.

You might also like