You are on page 1of 10

Article

pubs.acs.org/Langmuir

Gaining Control over Radiolytic Synthesis of Uniform Sub-3-


nanometer Palladium Nanoparticles: Use of Aromatic Liquids in the
Electron Microscope
Patricia Abellan,*,†,‡ Lucas R. Parent,‡,§ Naila Al Hasan,∥ Chiwoo Park,⊥ Ilke Arslan,‡ Ayman M. Karim,#
James E. Evans,▽ and Nigel D. Browning‡

SuperSTEM Laboratory, SciTech Daresbury Campus, Keckwick Lane, Daresbury WA4 4AD, United Kingdom

Fundamental and Computational Sciences Directorate, ∥Institute for Integrated Catalysis, and ▽Environmental Molecular Science
LaboratoryPacific Northwest National Laboratory, Post Office Box 999, Richland, Washington 99352, United States
§
Department of Chemistry & Biochemistry, University of California, San Diego, La Jolla, California 92093, United States

Department of Industrial and Manufacturing Engineering, Florida State University, Tallahassee, Florida 32306, United States
#
Department of Chemical Engineering, Virginia Polytechnic Institute and State University, Blacksburg, Virginia 24061, United States
*
S Supporting Information

ABSTRACT: Synthesizing nanomaterials of uniform shape


and size is of critical importance to access and manipulate the
novel structure−property relationships arising at the nanoscale,
such as catalytic activity. In this work, we synthesize Pd
nanoparticles with well-controlled size in the sub-3 nm range
using scanning transmission electron microscopy (STEM) in
combination with an in situ liquid stage. We use an aromatic
hydrocarbon (toluene) as a solvent that is very resistant to
high-energy electron irradiation, which creates a net reducing
environment without the need for additives to scavenge
oxidizing radicals. The primary reducing species is molecular
hydrogen, which is a widely used reductant in the synthesis of
supported metal catalysts. We propose a mechanism of particle
formation based on the effect of tri-n-octylphosphine (TOP)
on size stabilization, relatively low production of radicals, and autocatalytic reduction of Pd(II) compounds. We combine in situ
STEM results with insights from in situ small-angle X-ray scattering (SAXS) from alcohol-based synthesis, having similar
reduction potential, in a customized microfluidic device as well as ex situ bulk experiments. This has allowed us to develop a
fundamental growth model for the synthesis of size-stabilized Pd nanoparticles and demonstrate the utility of correlating different
in situ and ex situ characterization techniques to understand, and ultimately control, metal nanostructure synthesis.

■ INTRODUCTION
The chemical and physical properties of metallic nanomaterials
and sensing.13,14 Typically, the redox synthesis of particles with
well-controlled size involves the introduction of a protective
can differ significantly from those of the bulk. These differences stabilizer during particle formation: surfactants, polymers,
arise from the relatively large number of surface atoms per dendrimers, and ligands are common additives used.3,15,16
particle volume and from quantum confinement effects on the Additionally, factors such as nature and concentration of
nanoscale,1−3 and they can be tuned during synthesis.4 For reducing agent and precursors, nucleation sites, and processing
example, the catalytic efficiency of metallic nanoparticles and temperature also determine the nucleation and growth kinetics
clusters is known to be strongly particle-size-dependent.3,5,6 of metallic nanocrystals of uniform size and shape.15,17−19
True control of particle size during synthesis is, therefore, Methods that exploit radiation chemistry have been
crucial to achieving optimized performance and to precisely previously applied to the synthesis of metallic nanomateri-
investigate structure−property relationships. For the specific als,6,20−24 including Pd nanostructures.25 Radiolytic synthesis
case of Pd, nanoparticles are widely used as catalysts and has also been successful in preparing metal-supported clusters
electrocatalysts,7,8 where high efficiency in various fuel cell for catalytic applications.26−28 These synthesis routes use the
types has been demonstrated. Pd is an especially effective
catalyst, for instance, of ethanol oxidation in alkaline media9,10 Received: November 15, 2015
since it is not easily poisoned by carbon monoxide and, hence, Revised: January 7, 2016
is a key material for applications such as hydrogen storage11,12 Published: January 7, 2016

© 2016 American Chemical Society 1468 DOI: 10.1021/acs.langmuir.5b04200


Langmuir 2016, 32, 1468−1477
Langmuir Article

chemical effects of the absorption of high-energy radiation, relatively constant. Platinum spacers of 70 nm thickness were built
typically 60Co γ-rays or 2−20 MeV electrons,29,30 by precursor- on one of the chips by using focused ion beam (FIB) lithography in a
containing liquids to form nanostructures by reproducing a dual-beam FIB/SEM (FEI Quanta). Prior to sample loading, the
selective reducing/oxidizing environment. Conditions to silicon nitride chips were submitted to plasma cleaning for 1 min in
argon atmosphere at 150 mTorr pressure. Loading of the sample in the
generate free radicals of strong reducing potential can be fluid stage was carried out inside an MBraun glovebox filled with
found and exploited for different applications.31,32 Some purified argon where the solution was allowed to flow into the cell at 5
strengths of radiation chemical synthesis are that the nuclei μL/min for 15 min. Prior to solution flow, argon was allowed to flow
formed are homogeneously distributed in the whole volume, directly from the environment for 10 min to ensure no residual oxygen
growth rate can be easily controlled, and synthesis can be was present in the lines. The electron beam current values reported
performed at room temperature.6,33 have been measured by use of the screen dose meter of the
New characterization methods have been developed to microscope, previously calibrated to give the value at the sample plane
investigate colloidal synthesis and gather real-time information by use of the calibration curve shown in Figure S1. Calibration of the
at the solid−liquid interface.34−38 Among them, in situ liquid screen dose meter current reading of the microscope was done by use
of an analytical holder equipped with a Faraday cup near the specimen
cells used in combination with (scanning) transmission electron
location at the tip of the holder (Gatan, Inc.).47 An electron current of
microscopy [(S)TEM] uniquely allow for direct imaging of the 5.7 pA and pixel dwell time of 3 μs were used for all STEM data sets,
formation of nanoparticles with millisecond temporal reso- giving rise to values of the electron dose rate of 31 and 116 e−/Å2 per
lution and subnanometer spatial resolution.39−44 Control of the frame for 450K× and 910K× magnification, respectively. Formation
experimental parameters and reproducibility of the results kinetics were not accurately detected for magnifications below 450K×.
requires an understanding of the effects of high-energy A direct conversion of electron dose rates into radiation chemistry
electrons on the solution.45−49 Thus far, most experiments in units yields 3.1 × 107 Gy/s (450K×) and 11.5 × 107 Gy/s (910K×),
the fluid stage involve the use of water as a solvent. In water, where the stopping power for 300 kV electrons in toluene is 2.352
the radiation chemical yields are relatively large where highly MeV·cm2/g51 and total frame time was 3.78 s. In situ data sets 1−4 are
reactive oxidizing and reducing radicals and species are created included as Movies S1, S2. S3, and S4, respectively. The liquid
thickness was not measured in situ. Thin liquid layer thicknesses of
in about equal amounts.50 Thus, reproducing net reducing 100−150 nm were considered for our volume calculation in Figure 4
conditions for particle growth involves the addition of (see Supporting Information for further details and the effect of
substances that convert primary radicals into free reducing thickness uncertainty in the calculations).
radicals (use of OH• scavengers, for instance).20,50 Since the Precursor Solution Preparation and Nanoparticle Chemical
electron beam in the STEM is used as both irradiation source Synthesis. Toluene (99.8%) and recrystallized palladium(II) acetate
and imaging probe to resolve dynamic processes, larger incident (99.98%) were purchased from Sigma−Aldrich (St. Louis, MO).
electron doses may be required to achieve high-magnification Palladium acetate was dissolved in toluene in septum-sealed
imaging in the case of small (<3 nm) nanoparticles. An ideal scintillation vials to maintain an airtight environment and degassed
environment that can be understood and thus tuned in situ in with ultra-high-purity helium (99.999% purity from Oxarc) prior to
addition of the capping agent, TOP. The concentration of Pd in
the STEM would, therefore, consists of net reducing (or net toluene was 5 mM for in situ STEM experiments as well as ex situ bulk
oxidizing) conditions, where minimum (ideally none) additives synthesis and 50 mM for in situ small-angle X-ray scattering (SAXS)
are present, and where the yields of radical production are the experiments. The precursor to capping agent ratio was 1:1.5 Pd/TOP.
smallest possible for a given incident electron dose. Storage and addition of TOP (Aldrich, 90%) was carried out in an
Aromatic hydrocarbons are known to be very resistant to MBraun glovebox supplied with nitrogen. Hexanol (anhydrous,
high-energy electron irradiation,30 with toluene being able to ≥99%) was also purchased from Sigma−Aldrich (St. Louis, MO)
create a net reducing environment, as discussed below. and sparged with ultra-high-purity helium prior to solution
Furthermore, the use of organic solvents instead of water preparation.
allows for tuning the polarity of the medium, giving access to a Small-Angle X-ray Scattering. SAXS studies of TOP-capped Pd
nanoparticle bulk synthesis were carried out at the Argonne National
broader range of synthesis conditions. Here, we report a study Laboratory (ANL) Advanced Photon Source (APS) with the pinhole
aimed at finding more suitable Pd nanoparticle synthesis setup at beamlines 12-ID-C and 12-ID-B with incident X-ray energies
environments that accomplish all the conditions above for of 18 and 12 keV, respectively. In situ SAXS data was also collected in
nanostructure control. We demonstrate the ability to control a microfluidic reactor with similar results. The reacted solution from
the nucleation kinetics as well as the final size of Pd the microfluidic reactor (at 50 mM) was diluted with toluene to 5 mM
nanoparticles, ∼2.2 nm diameter, during in situ STEM and drop-cast on a lacey carbon film supported on a copper
irradiation, which is matched in bulk solution synthesis using transmission electron microscopy (TEM) grid (Ted Pella, Inc.) to
hexanol reduction of the same precursors. Synthesis is investigate the reaction products by STEM. From the time evolution
performed starting from a precursor solution of palladium of particles from SAXS, a linear fit was used to obtain the nucleation
acetate [Pd(OAc)2] dissolved in toluene, to which tri-n- rate (arbitrary units/s) as the slope of the number of particles versus
time at the early reaction times. The slope was scaled by a factor
octylphosphine (TOP) is added as a capping agent to promote estimated by taking into account the initial concentration of solution,
growth stabilization. final average nanoparticle size, and final number of nanoparticles

■ EXPERIMENTAL SECTION
Scanning Transmission Electron Microscopy and Fluid
(where full conversion of the precursor after an extended time period
is assumed). A step-by-step estimation of the initial nucleation rate
from SAXS for similar reactions and details on microreactor fabrication
Stage. In situ STEM data sets were acquired on an 80−300 kV and operation, SAXS data acquisition and analysis, and materials
FEI probe Cs-corrected Titan electron microscope equipped with an preparation can be found elsewhere.38
electron gun monochromator and a Gatan Quantum ERS Multitarget Particle Tracking Analysis. The multitarget particle
spectrometer. All the experiments were performed with 300 kV tracking approach has been previously applied to analysis of video
electrons. Liquid experiments in the STEM were performed on a frames of silver nanoparticles grown via in situ liquid STEM44,52 and
Hummingbird Scientific fluid stage. Silicon nitride (SiNx) windows, 10 has provided insights into the attack mechanisms that occur during
nm thick and supported on silicon frames (Norcada, Inc.), were used degradation of electrolyte solutions in Li ion batteries.53 First, an
to maximize image contrast while keeping the liquid thickness image segmentation algorithm was applied to successive images in

1469 DOI: 10.1021/acs.langmuir.5b04200


Langmuir 2016, 32, 1468−1477
Langmuir Article

order to extract a set of image pixel locations for nanoparticles,54 and


an object tracking algorithm was applied55 to associate the pixel
locations over multiple time frames.56 Further details can be found in
refs 52,55, and 56.

■ RESULTS
First we consider the radiation chemistry of the solvent,
toluene. It is indeed solvent radiolysis that predominantly
dictates the species and yields involved in the chemical
processes leading to metal nanostructure synthesis in the
STEM fluid stage. Aromatic hydrocarbons such as toluene are
very resistant to radiation damage because the excitation energy
is not localized in a particular bond but shared by all π-electrons
in the molecule.30 Additionally, the radiation process is
dominated by electron−ion recombination,30,57,58 limiting the
overall production of free radicals in these liquid media. An
introduction to the concept of geminate recombination and
some relevant numbers for toluene and other solvents to
illustrate this fact are given in Supporting Information. Overall, Figure 1. (a) Cropped sequence from a multiframe movie of the
the yield of ions escaping geminate recombination in toluene formation and size stabilization of several Pd particles from solution.
The first frame corresponds to the first scan after initial exposure to
(yield of free ions) is very low (Gfi = 0.01 μmol·J−1).59 This the electron beam. The time lapse between frames is 3.78 s.
means that the amount of major ionic species produced by Magnification used is M = 910000×, pixel dwell time is 3 μs, beam
radiolysis of liquid toluene will be much smaller than that of current is 5.7 pA, and image size is 1024 × 1024 pixels, which
liquid water or other polar liquids such as alcohols. correspond to an electron dose rate of 116 e−/Å2 per frame. The
Previous work has shown that the major products in multitarget particle tracking approach was applied to determine: (b)
radiolysis of toluene are hydrogen and dimeric materials mainly number of particles detected in each frame (red circles) and total
consisting of bibenzyl, bitolyls, and benzyltoluenes.60−63 The number of particles formed over time (black squares), which accounts
rupture of C−H bonds on the methyl group is the dominant for detection fluctuations via a multiway data association method, and
process, leading to the formation of bibenzyl.63 The overall (c) particle size evolution over time for individual particles.
reaction for toluene giving rise to molecular hydrogen and
bibenzyl (see Supporting Information for a discussion on the volume (STEM imaging area), with formation of Pd particles
production of major products in radiolysis of toluene and their being characterized by slow nucleation rate and fast growth
interaction with Pd2+ ions) would be rate.
2C6H5CH3 ⇝ H 2 + (C6H5CH 2)2 For quantitative evaluation of particle nucleation kinetics, we
have applied multitarget particle tracking statistical analysis55 to
Such neutral dimerized products [(C6H5CH2)2] are inert our raw STEM data sets to determine the number of particles
toward the reduction of palladium. Thus, we consider in each movie frame (see Experimental Section for details of the
dihydrogen (H2) as the main primary species involved in the analysis). The number of particles detected in each separate
reduction process leading to Pd0 particle formation in our frame per unit time is plotted as solid red circles in Figure 1b
experiments. H2 is a widely used reductant in the synthesis of (same data set as for Figure 1a). These are particles formed and
supported metal catalysts.64 It is characterized by reduction continuously observed after formation. To more accurately
potential of E0[H2/H+] = 0 V, while for the case of Pd2+, account for all formation events, a second measurement was
E0[Pd/Pd2+] = 0.915 V versus standard hydrogen electrode performed (solid black squares in Figure 1b), which also
(SHE).65 These values will vary when measured in a different accounts for particles that appear and quickly disappear, some
solvent rather than in water,66,67 however, a relatively large of which might be image artifacts. Both measurements make
potential difference should remain in toluene. use of a multiway data association method56 for the analysis,
Figure 1a is a series of cropped dark-field (DF) STEM originally developed for the case of multitarget tracking of
images showing successive and concomitant in situ nucleation, interacting objects. This methodology has been previously
growth, and size stabilization of individual Pd nanoparticles applied to systems displaying coalescence of individual particles
induced by the electron beam. The time interval between into larger ones or splitting of particles into different
frames is 3.78 s, and the electron dose rate is 116 e−/Å2 per subparts.52,53,56 The multitracking algorithm used is capable
frame (see Movie S1 for complete data set showing growth of of detecting particles that appear and shortly disappear and can
the entire population). After nucleation, the individual discard them to prevent the analysis results from being
nanoparticles rapidly stabilize at around 2.2 nm diameter with corrupted by image artifacts. For our case, where particle
a spherical geometry. Particles indicated by arrows in frames 3 motion out of the viewing area and deatchment/reatachment to
and 7 are first detected with diameters of 1.6 and 1.5 nm, the SiNx window has been observed (see Figure S2), both
respectively, and both stabilize to around 2.2 nm by the analyses are informative. Quantification of nucleation kinetics in
subsequent frame. Figure 1a shows that each successive movie this work has been done only for the period of time where both
frame contains newly nucleated particles that grow and stabilize measurements overlap (≤40 s). In addition to tracking the
in size to about 2.2 nm even upon further electron beam nucleation of particles over time, individual growth kinetics can
irradiation (see also the complete STEM data sets in Movies also be obtained from the raw STEM data sets. Figure 1c shows
S1, S2, S3, and S4). Multiple individual nucleation events take the projected diameter as a function of time for some individual
place simultaneously and continuously within the irradiated particles from the data set shown in Figure 1a. All the particles
1470 DOI: 10.1021/acs.langmuir.5b04200
Langmuir 2016, 32, 1468−1477
Langmuir Article

grow and stabilize within 1−2 frames to diameters that oscillate chemical reductive agent rather than by the effect of the
only slightly around the system average of 2.2 nm. electron beam. Hexanol was used as a reducing agent, being of
The overall nucleation and growth of the particle ensemble relatively similar reducing potential [for alcohols E(V) ∼ 0 V
was also evaluated by statistical treatment of all the individual versus SHE in water]68,69 to molecular H2. Figure 3 shows a
events. The average growth curves (average diameter over
time) for particle synthesis at an incident electron dose rate of
31 e−/Å2 per frame are shown in Figure 2a. The ensemble

Figure 3. DF-STEM image of Pd particles grown ex situ by chemical


reduction at 100 °C of 1:1.5 Pd/TOP at 5 mM in toluene (same
concentration and Pd/TOP ratio as for the STEM experiments) in the
presence of an excess of hexanol and corresponding particle size
distribution.

postmortem STEM image and corresponding particle size


Figure 2. (a) Average growth of the Pd nanoparticles ensemble distribution of a population of Pd particles grown at 100 °C
measured for a dose rate of 31 e−/Å2 per frame. (Inset) Cropped DF- reaction temperature (see Experimental Section for details on
STEM image of Pd particles formed after 120 s exposure and taken in chemical synthesis of the particles). The average ex situ particle
the fluid stage. Scale bar is 20 nm. (b) Total number of particles size of 2.1 nm diameter matches well with the average particle
formed per unit area as a function of time for two different dose rates, size formed by the electron beam during in situ STEM
showing a difference in slope scaling linearly with dose rate. synthesis experiments. The size of the particles is comparable
for both chemical and radiation-induced reduction processes,
particle size for this case stabilizes to ∼2.3 nm. Note we thus suggesting that the Pd/TOP ratio mainly determines
measure only a slight difference in ensemble particle size of 0.1 particle size for both cases; that is, TOP is responsible for the
nm between data sets produced with electron dose rates arrested growth mechanism during radiation-induced synthesis.
differing by a factor of 4 (see Figure S3 for particle size These ex situ results also suggest that additional factors, such as
distributions). The inset in Figure 2a shows a DF-STEM image formation of polymeric molecules from the breakdown of
of a population of Pd particles formed after 120 s of electron organic dimeric species or of TOP itself, are most likely not
beam exposure (data set 3 in the growth curve) as seen in acting as cluster stabilizers.
toluene in the fluid stage. First signs of damage of the organic In Figure 4 we have plotted together, using a logarithmic
solvent are found after 50 s of exposure and can be seen in the scale, the number of particles formed per unit volume versus
center and left-hand side of the image. Polymerization of the time from the in situ STEM experiments and an estimation of
organic solvent has been previously observed in the fluid the nucleation rate based on time evolution of the particles
stage.53 Figure 2b shows the total number of individual particles from in situ SAXS experiments with hexanol in a customized
per unit area as a function of time for two different electron microfluidic device38 (see Figure S4 for linear fits to number of
dose rates. Interestingly, the nucleation rate for all data sets particles per unit volume versus time and Experimental Section
remains fairly constant during the experiment, even for the for details on measurements and calculations). The nucleation
highest dose rates. A change in the rate of particle formation rate of hexanol-based synthesis increases with reaction
nearly proportional to the electron dose rate was measured. temperature, from 80 to 100 °C. As already mentioned, the
As a control experiment, we performed a bulk ex situ nucleation rate by STEM irradiation increases with dose rate.
synthesis using the same precursor solution (identical molarity Notably, the nucleation kinetics in the case of hexanol-based
and Pd/TOP ratio) to grow Pd nanoparticles with a different synthesis is 1−2 orders of magnitude slower than in the case of
1471 DOI: 10.1021/acs.langmuir.5b04200
Langmuir 2016, 32, 1468−1477
Langmuir Article

association kinetics were accelerated, for instance by the


catalytic effect of specific locations, such as surfaces. Reduction
kinetics is thus key to understand the amount of initial nuclei
formed.
Reduction kinetics of palladium(II) acetate [Pd(OAc)2] by
H2 have been previously investigated.72 The overall chemical
equation can be written as follows:
Pd(OAc)2 + H 2 = Pd0 + 2HOAc
This is, in fact, a multistep reaction [see Supporting
Information for a discussion on the kinetics of reduction of
Pd(OAc) 2 and a description of possible autocatalytic
mechanisms]. In previous work, the kinetics of reduction of
Pd(OAc)2 by H2 in solution and on the surface of a carbon
Figure 4. Plot curves of number of particles per unit volume versus
time, measured for radiolytic synthesis in STEM for two different support were compared.72 Faster reduction of Pd(II) species
electron dose values (squares and triangles) and estimated for was observed on the support surface. Additionally, reduction of
chemical synthesis with hexanol at two different reaction temperatures the Pd(II) to Pd0 could occur without intermediate formation
from in situ SAXS (dashed lines). A fluid thickness of 125 nm has been of Pd(I) compounds, as opposed to reduction in bulk solution.
used for volume estimation in STEM (see Supporting Information). In both cases, an autocatalytic mechanism, mainly driven by
Results of a linear fit to the SAXS data (25 mM Pd concentration) for Pd(0), was proposed. Differences in reaction kinetics due to the
T = 100 and 80 °C are shown. Experimental data can be found in presence of the stabilizer, TOP, as well as due to the different
Figure S4. A logarithmic scale has been used to highlight the order of support material, SiNx, are expected. However, it is very likely
production kinetics of the different syntheses. that a similar autocatalytic reduction mechanism by the already
formed Pd0 at the nuclei surface takes place. Also, the SiNx
the radiation-induced synthesis in the STEM, even at 100 °C. support may be partly responsible for the faster overall
We present next a discussion on the formation mechanisms of reduction kinetics found by in situ STEM, as compared to
particles, including nucleation and growth kinetics and size bulk hexanol-based reduction (see Figure 4). It can be
stabilization mechanisms. discussed that once the Pd metal starts to dimerize and form
Typically, at high irradiation dose rates, radiation-induced small aggregates, the autocatalytic reduction pathway may
synthesis of metallic clusters occurs by direct reduction activate, promoting very fast growth of a few aggregates (see
followed by coalescence.70 All reducing species are produced Figures 1c and 2a) and rapidly consuming the precursors at the
and scavenged within a short time, and a sudden generation of surface of the particle.
a high concentration of isolated atoms is expected, resulting in a The yield of H2 upon radiolysis of toluene has been
large number of smaller nuclei and a decreasing nucleation rate determined by use of γ-radiolysis [linear energy transfer (LET)
over time.70 Instead, our observations show a continuous about 0.2 eV/nm] and 1.8 MeV electron irradiation, giving in
increase in the density of particles with no sudden decay in the both cases a value of G(H2) = 0.14 molecule/100 eV,73,74
nucleation rate after the first nucleation event. Continuous which is much lower than production yields of more reactive
nucleation is also found for particles following slower formation species (of stronger reduction potential) from radiolysis of
kinetics by hexanol (see Figure S4 for experimental data water. We estimate that about 8 × 104 primary electrons have
resulting in nucleation rates plotted in Figure 4). This synthesis been delivered to the irradiated area for each Pd0 incorporated
also yielded a homogeneous particle size distribution, as shown in a particle (see details in Supporting Information). Note that
in a previous work.71 Nucleation typically needs to occur within an experimental value of about 6 × 103 primary electrons was
a very short period of time for control of the final particle size found for production of one Au0 in water.49
to be possible (LaMer model).71 On the contrary, the linear Finally, there is one more possibility that could lead to low
trend of nucleation of Pd particles over time reported here does production of Pd atoms: namely, rapid depletion of precursors
not follow the LaMer model for formation of uniform particles. inside the irradiated volume, being continuously compensated
Continuous nucleation in the STEM could be explained by by fast diffusion from the bulk solution. An increase of
the simultaneous occurrence of two conditions: (1) formation nucleation kinetics proportional to the increase in dose rate has
of a small number of initial stable nuclei and (2) a robust size been observed (Figures 2b and 4), indicative of a formation
stabilization mechanism. Indeed, if a small number of particles process limited by the rate of production of neutral species
is initially formed in the absence of a size stabilizer, growth within the irradiated area, rather than by the arrival of species
would continue over time, giving rise to fewer larger from the bulk solution. Formation of particles limited by the
particles.20,44,70 Instead, growth stops at about 2.2 nm size, reaction of radiolytic species and metal ions has been previously
and newly produced Pd atoms are thus unable to incorporate observed in the liquid cell for growth of metallic particles from
into the already formed particles. Conditions for nucleation of water with no stabilizer.44,49,75 Consistent with previous
stable nuclei and growth will then be met in multiple observations where fewer faceted bidimensional platelike
subsequent occasions. crystals were grown,49 no decay of formation rate of particles
Formation of a small number of initial nuclei upon electron is observed after the total amount of reduced Pd incorporated
beam irradiation implies a slower production rate of reducing into particles becomes larger than what was initially available in
radicals as compared to the rate of Pd ions−Pd0 and Pd0−Pd0 the precursor solution. We have estimated the time when
association mechanisms. This could simply be a consequence of depletion of precursors occurs. Consumption of all palladium in
a relatively low production rate of reducing agent by the solution within the irradiated area for the lowest dose rate of 31
incident electron beam. A similar effect is expected if palladium e−/Å2 per frame occurs upon formation of approximately 36
1472 DOI: 10.1021/acs.langmuir.5b04200
Langmuir 2016, 32, 1468−1477
Langmuir Article

particles (see Supporting Information for calculation details). surface,72 Pd metal ions with unstable intermediate oxidation
When it is taken into account that growth occurs on both state, Pd+, which are a powerful reductant themselves.30
membranes, solution depletion is estimated to occur after about From our growth curves of individual particles (Figure 1c),
30 s of irradiation. Slow nucleation, rather than a burst of two distinct stages may be defined, where growth proceeds
nucleation, is observed from the first formation event after just quickly at first, followed by a second stage where growth is
<4 s exposure, where 83−95% of precursors are still available in interrupted and particle size is almost unchanged. Polar head
the irradiated area. Additionally, the same gradual nucleation groups of the TOP capping agent may bind to the nucleus
rate continues for at least a few tens of seconds beyond 30 s. surface, with Pd nuclei initially showing only partial surface
This supports our conclusion that slow nucleation is limited by coverage by the capping ligand, and full coverage at the time
the reduction rate of metallic cations rather than by diffusion of growth is arrested (Figure 5b). This mechanism is consistent
species. Furthermore, continuous nucleation has also been with previous time-resolved X-ray absorption fine structure
observed in the case of synthesis by hexanol,38 presenting much (EXAFS) analysis showing that the TOP/Pd ratio in solution
slower kinetics of formation. increases as more atoms are incorporated into the nano-
Altogether, we propose the following scenario (Figure 5) for particles.38 It is worth noting that no oxygen is expected to
formation of supported Pd particles in toluene by radiolysis. participate in the reactions above. We should expect different
species and yields for radiolysis of toluene in the presence of air
or moisture.77,78 For this reason, and because TOP is air-
sensitive, all experiments have been performed in an inert
environment (see details in Experimental Section). Additional
results and further discussion on the stability of the solution
when exposed to air and the reaction mechanisms can be found
in Supporting Information (see Figures S5 and S6 for
agglomeration of nanoparticles due to the presence of moisture
in the solution).
As previously noted for growth of gold in water,49 the
constant formation rate of particles, even for the highest dose
rate (Figure 1c) and after consumption of more Pd than
initially present, requires continuous mobility of Pd2+ into the
reaction zone. Typically, in the fluid stage, the formation of
particles in the bulk solution and subsequent precipitation on a
SiNx surface is not observed. Instead, formation of particles is
systematically favored directly at the SiNx support surface. In
other words, the reaction between cations and radiolytic species
mainly occur at/near the window surfaces. A hemispherical
diffusion profile, in analogy to peculiar electrochemistry with
Figure 5. (a) Pd salt precursor reduction pathway by H2 (multistep ultramicroelectrodes (UME) and small currents applied, may
process). Reduction of Pd(II) species directly to Pd0 could occur, most be expected on both SiNx windows. A steady state for cation
likely promoted at the SiNx membranes. Reduction can follow a two- diffusion could be achieved over a very small reaction area
step process from Pd(II) to Pd(I) and from Pd(I) to Pd0. An (near the membranes as compared to the bulk liquid),
autocatalytic reduction mechanism by Pd0 is also expected on the
nucleus surface. (b) Schematics of different stages of nanoparticle
preventing cation precursor depletion near the windows.
formation shown along a representational in situ STEM growth curve. The presence of membranes is known to have a strong
Continuous nucleation at the SiNx membranes can be due to low influence on formation of particles in the STEM. However,
production of H2 and an autocatalytic reduction mechanism leading to their effects are not well understood. Among the effects of
few initial nuclei formed, followed by the effect of TOP on particle size membranes in radiolytic synthesis outside the microscope,
stabilization. adsorption has been observed to facilitate the redox process and
to slow down aggregation by strongly inhibiting diffusion.28
The incident electrons effectively ionize the solvent,6,30,76 Most generally, differences in radiolytic yields, reduction and
which is the most abundant molecule in the irradiated volume association kinetics, as well as possible electrostatic effects, can
(for diluted precursor solutions up to 5 mM over a broad range occur in the STEM. For the case of preformed supported
of pH values).76 Pd(II) complexes [Pd(OAc)2], solvated particles, electrostatic effects produced by the electron beam
divalent Pd ions, Pd2+, and TOP are minimally ionized through can even produce changes in particles’ mobility proportional to
knock-on damage and, therefore, almost exclusively react with the applied electron dose rates.79
species generated from toluene, predominantly H2. In the case The fact that the amount of Pd(II) reduced to Pd0 to be
of TOP, being an electron donor ligand, no reaction with H2 is incorporated into particles (i.e., nucleation rate) increases with
expected. Reduction of Pd(II) species in the presence of TOP increasing dose rate is interesting and is an advantage of the e−
could proceed via multiple pathways. Reaction of Pd(II) species radiolysis synthesis method, since it makes the synthesis
with H2 produces neutral Pd atoms (see Figure 5a). The effect process more efficient and allows a means to control reaction
of the SiNx support in the STEM could promote the reduction kinetics in situ. Also, kinetic rates are directly measured instead
of Pd2+ metal ions to the neutral Pd0 state,72 speeding up of using indirect fitting procedures. However, the time
formation kinetics compared to synthesis outside the liquid cell resolution of STEM scans did not allow us to detect the first
(Figure 4). TOP could act as a reducing agent as well (this seed nuclei formed or fully resolve the individual growth rates,
possibility is not included in Figure 5a). Finally, Pd0 can also which could range from t1/2 or faster to t1/3, depending on
produce, through an autocatalytic process at the particle whether the first growth stage is fully contained within one
1473 DOI: 10.1021/acs.langmuir.5b04200
Langmuir 2016, 32, 1468−1477
Langmuir Article

movie frame or two, respectively. Slower growth kinetics would particles in the presence of a polar capping agent. Finally, we
allow us to access important information on the initial stages. correlate fluid-stage observations in the STEM with in situ
The kinetics of reduction is known to influence strongly the SAXS and confirmed that similar general trends in the
intrinsic structure of the nanoparticles formed, for example, nanoparticles formation kineticsfast growth and slow
through the creation of defects or twinning events,20,75 which nucleation kinetics of Pd particles from tolueneare observed
leads to very different properties of the final particles.6 The in not only for different dosages in the electron microscope but
situ STEM nucleation rate is several orders of magnitude larger also for different experimental conditions in a microfluidic
than that by wet chemistry with an alcohol (see Figure 4). reactor. We show that the in situ STEM nucleation rate can be
Slowing down the reduction kinetics of in situ STEM controlled by tuning experimental liquid STEM parameters.
experiments to match those measured outside the microscope Under the conditions explored in this work, the in situ STEM
may allow us to study formation of particles with the relevant nucleation rate is several orders of magnitude larger than that
structure and final properties. measured for synthesis by wet chemistry with alcohol. We
Further methods to reduce dose and thus slow down highlight the importance of controlling the kinetics of reaction
reaction kinetics, and to increase acquisition times, are needed. with the electron beam, since it determines the intrinsic
Suitable solvents could be selected on the basis of a tradeoff structure of the nanoparticles and thus their properties.
between image contrast and the amount, and type, of radicals
produced per incident electron. Technologies that enable more
efficient use of the transmitted electrons, such as direct electron

*
ASSOCIATED CONTENT
S Supporting Information
detectors or new image processing and computational methods The Supporting Information is available free of charge on the
to retain image information at much lower doses, would also ACS Publications website at DOI: 10.1021/acs.lang-
enable faster acquisition times. Methods for reducing radiolytic muir.5b04200.
species formation would include use of solvents that minimize Additional text and equations and eight figures,
the production of radiolytic species (this work), choice of initial describing geminate recombination in radiolysis of
pH,49 low-dose imaging techniques, thinner SiNx membrane toluene, reaction mechanisms during radiolysis of
thickness, or finding ways to control the liquid thickness on toluene, effect of moisture on Pd nanoparticles, and
truly two-dimensional materials like graphene.


calculations on number of incident electrons per Pd atom
incorporated into stable particles and total amount of
CONCLUSION Pd2+ions initially available in the irradiation volumes
We have demonstrated control over Pd nanoparticle size (<3 (PDF)
nm size) during real-time imaging in the fluid stage and have Movie S1 (data set 1) (AVI)
used the results from our observations to develop a model for Movie S2 (data set 2) (AVI)
size-controlled Pd nucleation and growth in the presence of a Movie S3 (data set 3) (AVI)
capping agent (TOP). Quantification of particle formation has Movie S4 (data set 4) (AVI)


been achieved with and without taking into account single/few
frames events by incorporating a multiway data association AUTHOR INFORMATION
method into the multitarget tracking analysis. Comparison of
these two measurements is important for shedding light on Corresponding Author
kinetics of small nanometric-size particles undergoing lateral *E-mail pabellan@superstem.org.
motion out of the viewing area, intermittent attachment/ Notes
The authors declare no competing financial interest.


detachment to the SiNx membranes, and other sources of local
contrast change. As opposed to radiolytic synthesis outside the
microscope, where fine control of particle size requires ACKNOWLEDGMENTS
increasing the production of radicals, the challenge in the The work involving the development of in situ stages and
STEM, where incident doses are inherently higher by orders of solutions preparation was supported by the Chemical Imaging
magnitude, is to drastically reduce such production. An Initiative, under the Laboratory-Directed Research and
aromatic hydrocarbon (toluene) has been used as an efficient Development Program at Pacific Northwest National Labo-
solvent to reduce the production of radiolytic species by the ratory (PNNL). PNNL is a multiprogram national laboratory
electron beam drastically as compared to more polar liquids. operated by Battelle for the U.S. Department of Energy (DOE)
For the specific case of toluene, Pd2+ ions are subjected to a net under Contract DE-AC05-76RL01830. Use of beamlines (12-
reducing environment defined mainly by a single reducing ID-B and 12-ID-C) at the Advanced Photon Source is
species, dihydrogen. Direct observation of the formation of supported by the U.S. DOE, Office of Science, and Office of
nanometric supported Pd particles by molecular hydrogen is of Basic Energy Sciences, under Contract DE-AC02-06CH11357.
special interest, since this is a common reductant for supported A portion of the research was performed by use of the
metal catalysts by different synthesis methods. We have Environmental Molecular Sciences Laboratory (EMSL), a
discussed formation mechanisms of particles based on the national scientific user facility sponsored by the DOE’s Office
relatively low radiolytic production of H2, an autocatalytic of Biological and Environmental Research and located at
reduction mechanism at the particle surface, and the effect of PNNL. C.P. was supported by the National Science
the surfactant TOP on size stabilization of particles. We Foundation, Division of Civil, Mechanical, and Manufacturing
reproduced size-stabilized Pd nanoparticles via ex situ synthesis Innovation, under Contract NSFCMMI-1334012, by Florida
with hexanol, which has a similar reducing potential as H2, and State University Committee on Faculty Research Support
achieved similar final size of the particles. Using insights from in Award 032968, and by the Ralph E. Powe Junior Faculty
situ experiments, supported by ex situ results, we propose a Enhancement Award. SuperSTEM is the U.K. National Facility
fundamental growth process for the synthesis of sub-3 nm Pd for Aberration-Corrected STEM, supported by the Engineering
1474 DOI: 10.1021/acs.langmuir.5b04200
Langmuir 2016, 32, 1468−1477
Langmuir Article

and Physical Research Council (EPSRC). P.A. thanks dicobalt octacarbonyl: II. Nucleation and growth of particles. J. Colloid
Emmanuel Maisonhaute (Sorbonne Universities, UPMC) and Interface Sci. 1983, 94 (1), 220−228.
Ivan T. Lucas (Sorbonne Universities, UPMC) for helpful (19) Xia, Y.; Yang, P.; Sun, Y.; Wu, Y.; Mayers, B.; Gates, B.; Yin, Y.;
discussions. Kim, F.; Yan, H. One-Dimensional Nanostructures: Synthesis,


Characterization, and Applications. Adv. Mater. 2003, 15 (5), 353−
389.
ABBREVIATIONS (20) Belloni, J.; Mostafavi, M.; Remita, H.; Marignier, J. L.; Delcourt,
STEM, scanning transmission electron microscopy; TEM, M. O. Radiation-induced synthesis of mono- and multi-metallic
transmission electron microscopy; SAXS, small-angle X-ray clusters and nanocolloids. New J. Chem. 1998, 22 (11), 1239−1255.
scattering; FIB, focused ion beam; SEM, scanning electron (21) Surendran, G.; Ramos, L.; Pansu, B.; Prouzet, E.; Beaunier, P.;
microscope; APS, advanced photon source; BF, bright field; Audonnet, F.; Remita, H. Synthesis of Porous Platinum Nanoballs in
Soft Templates. Chem. Mater. 2007, 19 (21), 5045−5048.
Pd(OAc)2, palladium acetate; TOP, tri-n-octylphosphine; SHE,
(22) Krishnaswamy, R.; Remita, H.; Impéror-Clerc, M.; Even, C.;
standard hydrogen electrode


Davidson, P.; Pansu, B. Synthesis of Single-Crystalline Platinum
Nanorods within a Soft Crystalline Surfactant−PtII Complex.
REFERENCES ChemPhysChem 2006, 7 (7), 1510−1513.
(1) Bourdon, J., Ed. Growth and Properties of Metal Clusters: (23) Yamamoto, T. A.; Nakagawa, T.; Seino, S.; Nitani, H. Bimetallic
Applications to Catalysis and the Photographic Process: Proceedings of nanoparticles of PtM (M = Au, Cu, Ni) supported on iron oxide:
the 32nd International Meeting of the Societe de Chimie Physique, Radiolytic synthesis and CO oxidation catalysis. Appl. Catal., A 2010,
Villeurbanne, 24−28 September 1979; Elsevier: Amsterdam, 1980. 387 (1−2), 195−202.
(2) Roduner, E. Size matters: why nanomaterials are different. Chem. (24) Yamamoto, T. A.; Kageyama, S.; Seino, S.; Nitani, H.;
Soc. Rev. 2006, 35 (7), 583−592. Nakagawa, T.; Horioka, R.; Honda, Y.; Ueno, K.; Daimon, H.
(3) Klabunde, K. J.; Richards, R. M. Nanoscale Materials in Chemistry; Methanol oxidation catalysis and substructure of PtRu/C bimetallic
2nd ed.; Wiley−Interscience: New York, 2009. nanoparticles synthesized by a radiolytic process. Appl. Catal., A 2011,
(4) Murray, C. B.; Norris, D. J.; Bawendi, M. G. Synthesis and 396 (1−2), 68−75.
characterization of nearly monodisperse CdE (E = sulfur, selenium, (25) Redjala, T.; Apostolecu, G.; Beaunier, P.; Mostafavi, M.;
tellurium) semiconductor nanocrystallites. J. Am. Chem. Soc. 1993, 115 Etcheberry, A.; Uzio, D.; Thomazeau, C.; Remita, H. Palladium
(19), 8706−8715. nanostructures synthesized by radiolysis or by photoreduction. New J.
(5) Somorjai, G. A.; Park, J. Y. Molecular factors of catalytic Chem. 2008, 32 (8), 1403−1408.
selectivity. Angew. Chem., Int. Ed. 2008, 47 (48), 9212−9228. (26) Belloni, J.; Delcourt, M.; Leclere, C. Radiation induced
(6) Belloni, J. Nucleation, growth and properties of nanoclusters preparation of metal catalysts: iridium aggregates. Nouv. J. Chim.
studied by radiation chemistry - Application to catalysis. Catal. Today 1982, 6 (1), 507−509.
2006, 113 (3−4), 141−156. (27) Bruneaux, J.; Cachet, H.; Froment, M.; Amblard, J.; Belloni, J.;
(7) Pérez-Lorenzo, M. Palladium Nanoparticles as Efficient Catalysts Mostafavi, M. Improvement of charge transfer kinetics at transparent
for Suzuki Cross-Coupling Reactions. J. Phys. Chem. Lett. 2012, 3 (2), SnO 2 counterelectrodes by means of a radiolytic grafting of metallic
167−174. nanoaggregates. Electrochim. Acta 1987, 32 (10), 1533−1536.
(8) Cookson, J. The preparation of palladium nanoparticles. Platinum (28) Platzer, O.; Amblard, J.; Marignier, J.; Belloni, J. Nanosecond
Met. Rev. 2012, 56 (2), 83−98. pulse radiolysis study of metal aggregation in polymeric membranes. J.
(9) Shen, P. K.; Xu, C. Alcohol oxidation on nanocrystalline oxide Phys. Chem. 1992, 96 (5), 2334−2340.
Pd/C promoted electrocatalysts. Electrochem. Commun. 2006, 8 (1), (29) Buxton, G. V.; Mulazzani, Q. G.; Ross, A. B. Critical Review of
184−188. Rate Constants for Reactions of Transients from Metal Ions and Metal
(10) Liu, J.; Ye, J.; Xu, C.; Jiang, S. P.; Tong, Y. Kinetics of ethanol Complexes in Aqueous Solution. J. Phys. Chem. Ref. Data 1995, 24 (3),
electrooxidation at Pd electrodeposited on Ti. Electrochem. Commun. 1055−1349.
2007, 9 (9), 2334−2339. (30) Buxton, G. V. An overview of the radiation chemistry of liquids.
(11) Kishore, S.; Nelson, J. A.; Adair, J. H.; Eklund, P. C. Hydrogen In Radiation Chemistry: From Basics to Applications in Material and Life
storage in spherical and platelet palladium nanoparticles. J. Alloys
Sciences; Spotheim-Maurizot, M.; Mostafavi, M.; Douki, T.; Belloni, J.,
Compd. 2005, 389 (1−2), 234−242.
Eds.; EDP Sciences: Les Ulis, France, 2008; pp 3−16.
(12) Yamauchi, M.; Ikeda, R.; Kitagawa, H.; Takata, M. Nanosize
(31) Radiation Chemistry: From Basics to Applications in Material and
Effects on Hydrogen Storage in Palladium. J. Phys. Chem. C 2008, 112
Life Sciences; Spotheim-Maurizot, M.; Mostafavi, M.; Douki, T.;
(9), 3294−3299.
(13) Mubeen, S.; Zhang, T.; Yoo, B.; Deshusses, M. A.; Myung, N. V. Belloni, J., Eds.; EDP Sciences: Les Ulis, France, 2008; p 307.
Palladium Nanoparticles Decorated Single-Walled Carbon Nanotube (32) Mozumder, A. Fundamentals of Radiation Chemistry; Academic
Hydrogen Sensor. J. Phys. Chem. C 2007, 111 (17), 6321−6327. Press: San Diego, CA, 1999.
(14) Wu, G.-h.; Wu, Y.-f.; Liu, X.-w.; Rong, M.-c.; Chen, X.-m.; Chen, (33) Choi, S.-H.; Zhang, Y.-P.; Gopalan, A.; Lee, K.-P.; Kang, H.-D.
X. An electrochemical ascorbic acid sensor based on palladium Preparation of catalytically efficient precious metallic colloids by γ-
nanoparticles supported on graphene oxide. Anal. Chim. Acta 2012, irradiation and characterization. Colloids Surf., A 2005, 256 (2), 165−
745 (0), 33−37. 170.
(15) Richards, R.; Bönnemann, H. Synthetic Approaches to Metallic (34) Somorjai, G. A.; Beaumont, S. K.; Alayoglu, S. Determination of
Nanomaterials. In Nanofabrication Towards Biomedical Applications; molecular surface structure, composition, and dynamics under reaction
Wiley−VCH: Weinheim, Germany, 2005; pp 3−32. conditions at high pressures and at the solid-liquid interface. Angew.
(16) Kim, S.-W.; Park, J.; Jang, Y.; Chung, Y.; Hwang, S.; Hyeon, T.; Chem., Int. Ed. 2011, 50 (43), 10116−29.
Kim, Y. W. Synthesis of Monodisperse Palladium Nanoparticles. Nano (35) Blasco, T. Insights into reaction mechanisms in heterogeneous
Lett. 2003, 3 (9), 1289−1291. catalysis revealed by in situ NMR spectroscopy. Chem. Soc. Rev. 2010,
(17) Duff, D. G.; Baiker, A.; Edwards, P. P. A New Hydrosol of Gold 39 (12), 4685−702.
Clusters 0.1. Formation and Particle-Size Variation. Langmuir 1993, 9 (36) Ellis, P. J.; Fairlamb, I. J.; Hackett, S. F.; Wilson, K.; Lee, A. F.
(9), 2301−2309. Evidence for the surface-catalyzed Suzuki-Miyaura reaction over
(18) Papirer, E.; Horny, P.; Balard, H.; Anthore, R.; Petipas, C.; palladium nanoparticles: an operando XAS study. Angew. Chem., Int.
Martinet, A. The preparation of a ferrofluid by decomposition of Ed. 2010, 49 (10), 1820−4.

1475 DOI: 10.1021/acs.langmuir.5b04200


Langmuir 2016, 32, 1468−1477
Langmuir Article

(37) Lee, A. F.; Ellis, P. J.; Fairlamb, I. J.; Wilson, K. Surface catalysed (56) Park, C.; Woehl, T. J.; Evans, J. E.; Browning, N. D. Minimum
Suzuki-Miyaura cross-coupling by Pd nanoparticles: an operando XAS Cost Multi-way Data Association for Optimizing Multitarget Tracking
study. Dalton Trans 2010, 39 (43), 10473−82. of Interacting Objects. IEEE Trans. Pattern Anal. Mach. Intell. 2015, 37
(38) Karim, A. M.; Al Hasan, N.; Ivanov, S.; Siefert, S.; Kelly, R. T.; (3), 611−624.
Hallfors, N. G.; Benavidez, A. D.; Kovarik, L.; Jenkins, A.; Winans, R. (57) Katsumura, Y.; Tabata, Y.; Tagawa, S. The formation process of
E.; Datye, A. K. Synthesis of 1nm Pd Nanoparticles in a Microfluidic the excited state of cycloalkane liquid using picosecond pulse
Reactor: Insights from In Situ X-ray Absorption Fine Structure radiolysis. Radiat. Phys. Chem. 1982, 19 (4), 267−276.
Spectroscopy and Small-Angle X-ray Scattering. J. Phys. Chem. C 2015, (58) Jonah, C. D. Decay of geminate ions in hexanes. Radiat. Phys.
119, 13257. Chem. 1983, 21 (1), 53−56.
(39) Williamson, M. J.; Tromp, R. M.; Vereecken, P. M.; Hull, R.; (59) Allen, A. O. Yields of free ions formed in liquids by radiation;
Ross, F. M. Dynamic microscopy of nanoscale cluster growth at the National Standard Reference Data System, 1976; http://www.nist.
solid-liquid interface. Nat. Mater. 2003, 2 (8), 532−536. gov/data/nsrds/NSRDS-NBS57.pdf.
(40) Evans, J. E.; Jungjohann, K. L.; Browning, N. D.; Arslan, I. (60) Hoigné, J.; Gäumann, T. Strahlungschemie der Kohlenwasser-
Controlled Growth of Nanoparticles from Solution with In Situ Liquid stoffe. 4. Mitteilung. Toluol. Helv. Chim. Acta 1961, 44 (7), 2141−
Transmission Electron Microscopy. Nano Lett. 2011, 11 (7), 2809− 2150.
2813. (61) Gale, L. H.; Gordon, B. E.; Steinberg, G.; Wagner, C. D.
(41) de Jonge, N.; Ross, F. M. Electron microscopy of specimens in Radiolysis of Toluene: Mechanism of Formation of Benzyl Radicals. J.
liquid. Nat. Nanotechnol. 2011, 6 (11), 695−704. Phys. Chem. 1962, 66 (8), 1538−1539.
(42) Liao, H. G.; Cui, L. K.; Whitelam, S.; Zheng, H. M. Real-Time (62) Hoigné, J.; Burns, W. G.; Marsh, W. R.; Gäumann, T.
Imaging of Pt3Fe Nanorod Growth in Solution. Science 2012, 336 Strahlungschemie von Kolhenwasserstoffen 10. Mitteilung Let-Effekte
(6084), 1011−1014. in Toluol. Helv. Chim. Acta 1964, 47 (1), 247−259.
(43) Parent, L. R.; Robinson, D. B.; Woehl, T. J.; Ristenpart, W. D.; (63) Sagert, N. H.; Quinn, M. J.; Sanipelli, G. G.; Kremers, W.;
Evans, J. E.; Browning, N. D.; Arslan, I. Direct in Situ Observation of Buchannon, W. D.; Westmore, J. B. Gamma-radiolysis of toluene and
Nanoparticle Synthesis in a Liquid Crystal Surfactant Template. ACS deuterated toluenesII. Liquid products. Int. J. Radiat. Appl. Instrum.
Nano 2012, 6 (4), 3589−3596. Part C 1991, 38 (5), 449−455.
(44) Woehl, T. J.; Evans, J. E.; Arslan, L.; Ristenpart, W. D.; (64) Toebes, M. L.; van Dillen, J. A.; de Jong, K. P. Synthesis of
Browning, N. D. Direct in Situ Determination of the Mechanisms supported palladium catalysts. J. Mol. Catal. A: Chem. 2001, 173 (1),
Controlling Nanoparticle Nucleation and Growth. ACS Nano 2012, 6 75−98.
(10), 8599−8610. (65) Bard, A. J.; Faulkner, L. R. Electrochemical Methods:
(45) Woehl, T. J.; Jungjohann, K. L.; Evans, J. E.; Arslan, I.; Fundamentals and Applications; Wiley: New York, 1980; Vol. 2.
Ristenpart, W. D.; Browning, N. D. Experimental procedures to (66) Elving, P. J.; Spritzer, M. S. Polarographic reduction of hydrogen
mitigate electron beam induced artifacts during in situ fluid imaging of ion in non-aqueous solvents. Talanta 1965, 12 (12), 1243−1258.
nanomaterials. Ultramicroscopy 2013, 127, 53−63. (67) Izutsu, K. Electrochemistry in Nonaqueous Solutions, 2nd ed.; John
(46) Grogan, J. M.; Schneider, N. M.; Ross, F. M.; Bau, H. H. Bubble Wiley & Sons, 2009; DOI: 10.1002/9783527629152.
and Pattern Formation in Liquid Induced by an Electron Beam. Nano (68) Hapiot, P.; Pinson, J.; Francesch, C.; Mhamdi, F.; Rolando, C.;
Lett. 2014, 14 (1), 359−364. Schneider, S. One-electron redox potentials for the oxidation of
(47) Abellan, P.; Woehl, T. J.; Parent, L. R.; Browning, N. D.; Evans, coniferyl alcohol and analogues. J. Electroanal. Chem. 1992, 328 (1),
J. E.; Arslan, I. Factors influencing quantitative liquid (scanning) 327−331.
transmission electron microscopy. Chem. Commun. 2014, 50 (38), (69) Iwasita, T. Electrocatalysis of methanol oxidation. Electrochim.
4873−4880. Acta 2002, 47 (22), 3663−3674.
(48) Schneider, N. M.; Norton, M. M.; Mendel, B. J.; Grogan, J. M.; (70) Belloni, J.; Mostafavi, M. Radiation chemistry of nanocolloids
Ross, F. M.; Bau, H. H. Electron−Water Interactions and Implications and clusters. Stud. Phys. Theor. Chem. 2001, 87, 411−452.
for Liquid Cell Electron Microscopy. J. Phys. Chem. C 2014, 118 (38), (71) LaMer, V. K.; Dinegar, R. H. Theory, Production and
22373−22382. Mechanism of Formation of Monodispersed Hydrosols. J. Am.
(49) Park, J. H.; Schneider, N. M.; Grogan, J. M.; Reuter, M. C.; Bau, Chem. Soc. 1950, 72 (11), 4847−4854.
H. H.; Kodambaka, S.; Ross, F. M. Control of Electron Beam-Induced (72) Berenblyum, A.; Al-Wadhaf, H. A.; Katsman, E.; Flid, V. Kinetics
Au Nanocrystal Growth Kinetics through Solution Chemistry. Nano and mechanism of palladium (II) acetate reduction by hydrogen on
Lett. 2015, 15 (8), 5314−5320. the surface of a carbon support. Kinet. Catal. 2011, 52 (2), 296−304.
(50) Farhataziz; Rodgers, M. A. J. Radiation Chemistry: Principles and (73) LaVerne, J. A.; Baidak, A. Track effects in the radiolysis of
Applications; VCH Publishers: New York, 1987; p 641. aromatic liquids. Radiat. Phys. Chem. 2012, 81 (9), 1287−1290.
(51) Berger, M. J.; Coursey, J. S.; Zucker, M. A.; Chang, J. Stopping- (74) Hentz, R. R.; Burton, M. Studies in Photochemistry and
power and range tables for electrons, protons, and helium ions; NISTIR Radiation Chemistry of Toluene, Mesitylene and Ethylbenzene1, 2, 3.
4999; NIST Physical Measurement Laboratory, 1998; http://www. J. Am. Chem. Soc. 1951, 73 (2), 532−536.
nist.gov/pml/data/star/. (75) Alloyeau, D.; Dachraoui, W.; Javed, Y.; Belkahla, H.; Wang, G.;
(52) Woehl, T. J.; Park, C.; Evans, J. E.; Arslan, I.; Ristenpart, W. D.; Lecoq, H.; Ammar, S.; Ersen, O.; Wisnet, A.; Gazeau, F.; Ricolleau, C.
Browning, N. D. Direct observation of aggregative nanoparticle Unravelling Kinetic and Thermodynamic Effects on the Growth of
growth: kinetic modeling of the size distribution and growth rate. Gold Nanoplates by Liquid Transmission Electron Microscopy. Nano
Nano Lett. 2014, 14 (1), 373−8. Lett. 2015, 15 (4), 2574−2581.
(53) Abellan, P.; Mehdi, B. L.; Parent, L. R.; Gu, M.; Park, C.; Xu, (76) Fricke, H. The Reduction of Oxygen to Hydrogen Peroxide by
W.; Zhang, Y.; Arslan, I.; Zhang, J.-G.; Wang, C.-M.; Evans, J. E.; the Irradiation of Its Aqueous Solution with X-Rays. J. Chem. Phys.
Browning, N. D. Probing the Degradation Mechanisms in Electrolyte 1934, 2 (9), 556−557.
Solutions for Li-Ion Batteries by in Situ Transmission Electron (77) Christensen, H.; Sehested, K.; Hart, E. Formation of benzyl
Microscopy. Nano Lett. 2014, 14 (3), 1293−1299. radicals by pulse radiolysis of toluene in aqueous solutions. J. Phys.
(54) Park, C.; Huang, J. H. Z.; Ji, J. X.; Ding, Y. Segmentation, Chem. 1973, 77 (8), 983−987.
Inference, and Classification of Partially Overlapping Nanoparticles. (78) Yamamoto, Y.; Takamuku, S.; Sakurai, H. Gas-phase radiolysis
IEEE Trans. Pattern Anal. Mach. Intell. 2013, 35 (3), 669−681. of toluene. J. Phys. Chem. 1970, 74 (18), 3325−3332.
(55) Park, C. Estimating Multiple Pathways of Object Growth Using (79) Woehl, T. J.; Prozorov, T. The Mechanisms for Nanoparticle
Non-longitudinal Image Data. Technometrics 2014, 56 (2), 186−199. Surface Diffusion and Chain Self-Assembly Determined from Real-

1476 DOI: 10.1021/acs.langmuir.5b04200


Langmuir 2016, 32, 1468−1477
Langmuir Article

Time Nanoscale Kinetics in Liquid. J. Phys. Chem. C 2015, 119 (36),


21261−21269.

1477 DOI: 10.1021/acs.langmuir.5b04200


Langmuir 2016, 32, 1468−1477

You might also like