ArticlePDF Available

Determination of the yield, mass and structure of silver patches on colloidal silica using multiwavelength analytical ultracentrifugation

Authors:
  • Indian Institute of Technology Goa, Goa, India

Abstract

Anisotropic nanoparticles offer considerable promise for applications but also present significant challenges in terms of their characterization. Recent developments in the electroless deposition of silver patches directly onto colloidal silica particles have opened up a simple and scalable synthesis method for patchy particles with tunable optical properties. Due to the reliance on patch nucleation and growth, however, the resulting coatings are distributed in coverage and thickness and some core particles remain uncoated. To support process optimization, new methods are required to rapidly determine patch yield, thickness and coverage. Here we present a novel approach based on multiwavelength analytical ultracentrifugation (MWL-AUC) which permits simultaneous hydrodynamic and spectroscopic characterization. The patchy particle colloids are produced in a continuous flow mixing process that makes use of a KM-type micromixer. By varying the process flow rate or metal precursor concentration we show how the silver to silica mass ratio distribution derived from the AUC-measured sedimentation coefficient distribution can be influenced. Moreover, through reasoned assumptions we arrive at an estimation of the patch yield that is close to that determined by arduous analysis of scanning electron microscopy (SEM) images. Finally, combining MWL-AUC, electrodynamic simulations and SEM image analysis we establish a procedure to estimate the patch thickness and coverage.
1
Determination of the yield, mass and structure of silver patches on colloidal silica using
multiwavelength analytical ultracentrifugation
Thomas Meincke, Johannes Walter, Lukas Pflug, Thaseem Thajudeen, Andreas Völkl, Paola
Cardenas Lopez, Maximilian Uttinger, Michael Stingl, Satoshi Watanabe, Wolfgang Peukert
and Robin Klupp Taylor*
T. Meincke, J. Walter, A. Völkl, P.I. Cardenas Lopez, M.J. Uttinger, Prof, W. Peukert, Prof.
Robin N. Klupp Taylor
Institute of Particle Technology, Friedrich-Alexander University of Erlangen-Nürnberg,
Cauerstrasse 4, 91058 Erlangen, Germany
E-mail: robin.klupp.taylor@fau.de
Dr. L. Pflug, Prof. M. Stingl
Applied Mathematics 2, Friedrich-Alexander University of Erlangen-Nürnberg, Cauerstrasse
11, 91058 Erlangen, Germany
Prof. T. Thajudeen
Mechanical Engineering Department, IIT Goa, Ponda, Goa, 403401, India
Prof. S. Watanabe
Chemical Engineering Department, Kyoto University, Katsura, Nishikyo, Kyoto, 615-8510
Japan
Keywords: ultracentrifugation, localized surface plasmon resonance, patchy particles,
continuous flow synthesis, multidimensional characterization
Abstract
Anisotropic nanoparticles offer considerable promise for applications but also present
significant challenges in terms of their characterization. Recent developments in the
electroless deposition of silver patches directly onto colloidal silica particles have opened up a
simple and scalable synthesis method for patchy particles with tunable optical properties. Due
to the reliance on patch nucleation and growth, however, the resulting coatings are distributed
in coverage and thickness and some core particles remain uncoated. To support process
optimization, new methods are required to rapidly determine patch yield, thickness and
coverage. Here we present a novel approach based on multiwavelength analytical
ultracentrifugation (MWL-AUC) which permits simultaneous hydrodynamic and
spectroscopic characterization. The patchy particle colloids are produced in a continuous flow
mixing process that makes use of a KM-type micromixer. By varying the process flow rate or
2
metal precursor concentration we show how the silver to silica mass ratio distribution derived
from the AUC-measured sedimentation coefficient distribution can be influenced. Moreover,
through reasoned assumptions we arrive at an estimation of the patch yield that is close to that
determined by arduous analysis of SEM images. Finally, combining MWL-AUC,
electrodynamic simulations and SEM image analysis we establish a procedure to estimate the
patch thickness and coverage.
1. Introduction
Patchy particles represent a special class of anisotropic colloids.[1] In the present context, this
anisotropy is manifested through the presence of regions of different material at the surface of
a colloidal spherical core particle. Due to this arrangement, patchy particles can exhibit the
bulk and interfacial physicochemical properties of two different materials. The type of patchy
particle of concern in the present work comprises a colloidal silica core particle, produced by
the Stöber method,[2] having a certain fraction of its surface coated by a spherical cap of silver
up to a few tens of nanometers in thickness (Figure 1, upper row), produced by electroless
plating. Such types of patchy or Janus particles, the latter defined as having a single patch
with 50% coverage are typically synthesized by template based processes.[39] These permit
the fabrication of well-defined, monodisperse patchy or Janus particles but lack considerably
in their scalability. Since anisotropic and especially patchy particles are highly promising for
several fields of applications like sensors,[10] functional films,[9,11] environment protection,[12]
in medicine,[13,14] e.g. for drug delivery[15] and cancer treatment[16,17] or the design of self-
assembled hierarchical particle structures,[1,1824] scalable fabrication techniques are urgently
sought.
3
Figure 1. Schematic showing 14 nm thick silver patches with various coverages (cpatch) on
255 nm diameter silica spheres (upper row) and the equivalent patch volume core-shell
particles with shell thickness tshell and having the same silver to silica mass ratio (ϕ) as the
patch particles (lower row).
The recently introduced colloidal method for patchy particle synthesis used in the present
work follows the principle of heterogeneous nucleation and surface conformal growth of
electroless reduced silver on sub-micrometer silica spheres.[25] Due to the effect of mixing and
the stochastic nature of nucleation, some core particles fail to nucleate a patch, and the
patches that do form are distributed in terms of their coverage and thickness. Recently our
original batch process was successfully transferred into a continuous T-mixer based flow
process.[26] The main reason for doing this was to reduce the influence of mixing on the patch
formation reaction and thus obtain more narrowly dispersed patches. By variation of reaction
parameters it was shown that both patchy and almost complete silver coatings, i.e. where the
patch grows so far that it leads to a closed concentric shell, could be reproducibly synthesized.
In addition, it was shown how the combination of optical UV-Vis-NIR spectroscopy, SEM
(scanning electron microscopy) image analysis and FEM (finite element method) based
simulations of the silver patchy particle extinction spectra allow the characterization of a
given sample according to its patch yield (defined as the fraction of all core particles
possessing a patch).[26] Unfortunately, due to very strong dependency of the patch extinction
spectrum on morphology, such optical patch yields can only be derived by UV-Vis-NIR
4
spectroscopy within a set of samples showing very similar optical properties. Moreover, an
additional reference sample of known statistical patch yield, as derived by SEM image
analysis independently, is required. In turn, the result of such image analysis is highly
dependent on sample quality, orientation of individual patchy particles and to some extent on
the operator of the microscope. Moreover, it is highly challenging to obtain the patch
coverage distribution, a critical parameter to be tuned for applications. Hence, additional
characterization techniques are needed in order to generate a rapid and full understanding of
the patch coverage and thickness distributions and patch yield.
In the present study, the combination of particle sedimentation analysis, through analytical
ultracentrifugation (AUC) with simultaneous spectroscopic characterization, is shown to be a
very promising non-microscopical technique for determining the disperse morphology of
patchy particles.
An AUC equipped with a multiwavelength detector (MWL-AUC) allows the separation of
different particle fractions according to their sedimentation properties and the extinction
spectra corresponding to those fractions extracted. Using this technique, we have recently
demonstrated the derivation of species-specific extinction spectra for gold nanorods enabling
the retrieval of the 2D (length and width) size distributions with excellent agreement to TEM
data.[27] For the case of a population of silver patches grown on monodisperse core particles as
used in this work, the relative effective density of each particle is determined solely by the
mass of patch material attached to the core. This means that, in principle, the patch yield can
be readily identified since the slowest sedimenting fraction will be that of uncoated core
particles. In practice, however, the strong difference in optical extinction between the bare
and coated core particles prohibits this. However, provided that a median patch to core mass
ratio for the coated particles can be measured, and by making reasonable assumptions about
the reaction stoichiometry, it is possible to arrive at an estimate of the patch yield which
avoids the need to detect the uncoated particles. Moreover, if patches are assumed to have a
5
constant thickness and density of bulk silver their mass will depend only on the thickness and
coverage of the patch. Due to the highly morphology-dependent plasmon resonance of silver
nanostructures,[28] dispersity in thickness and coverage at constant effective density would be
associated with broadened plasmonic features. The benefit of MWL-AUC is that by
separating the particles according to density and measuring optical spectra of resulting
fractions the latter effect can be rationalized, leading to an understanding of the width of the
patch distribution. Hence, MWL-AUC is expected to be a valuable technique able to
determine patch yield, thickness and coverage, essential information for the optimization of
the synthetic process in order to target certain morphologies and properties in the future.
Moreover, the measurements also allow the determination of process relevant aspects like the
weight ratio of patches per particle and the reaction yield of the silver reduction reaction.
In this article it is demonstrated for the first time how MWL-AUC can be used to obtain
realistic values of yield and disperse geometrical parameters of a distribution of
asymmetrically coated nanoparticles (NPs). The theoretical background on the
characterization methodology is provided and it is shown for two exemplary process
parameter variation studies how MWL-AUC can assist with bringing an in-depth
understanding on this promising class of new material.
2. Theoretical background for analytical ultra-centrifugation and data evaluation
2.1. Theory of sedimentation and gravitational sweep analysis
When exposed to a centrifugal field, particles will sediment or float according to their size,
density and shape. Moreover, particles will diffuse depending on temperature, solvent
viscosity and their hydrodynamic diameter. AUC allows to measure the sedimentation and
diffusion properties of NPs by means of sedimentation velocity (SV) experiments with high
resolution and excellent statistical confidence. For large, rapidly sedimenting colloids, such as
the patchy particles of the present work, gravitational sweep (GS) experiments are an
effective alternative.[27,2931] In such case, the measurement position is kept constant while the
6
rotor speed is continuously increased. Even though diffusion information is not accessible by
GS experiments, much larger NPs and much broader particle size distributions (PSDs) can be
studied. Besides, for the large particle sizes (> 20 nm) typically studied in GS runs, diffusion
plays a negligible role.
The most important quantity derived by an AUC experiment is the sedimentation coefficient.
It is defined as the sedimentation velocity of a particle divided by the centrifugal acceleration.
The sedimentation coefficient s is a function of the time t, rotor angular velocity , radial
measurement position r and the meniscus position rm:
=

(1)
During a MWL-AUC GS experiment, the optical transmission is monitored at multiple
wavelengths as a function of the sedimentation coefficient. GS data can be evaluated using the
software HDR-MULTIFIT, which provides the extinction weighted density or cumulative
distributions of the sedimentation coefficient.[31] As the density distributions are extinction
weighted, integration immediately provides the associated extinction spectra. This allows the
derivation of the spectra of individual species in a mixture as has been shown before for
plasmonic NPs.[27,31] Notably, this data is equivalent to band sedimentation, for which details
are discussed in Section VI of the Supplementary Material. However, GS experiments require
no special centrepieces, are less prone to uncertainties, at least for the given boundary
conditions, and are thus the method of choice for patchy particle characterization.
The analysis in HDR-MULTIFIT is based on a model independent approach and basically
applies Equation 1 while correcting for radial dilution effects.[30,31] As data acquired by the
MWL-AUC is noisy, smoothing has to be applied. Direct boundary modelling (DBM)
constitutes an interesting alternative to the model independent approach as excessive
smoothing is not necessary. Here, the ls-g*(s) approach is based on the method developed by
Schuck and Rossmanith,[32] and was previously also transferred to analytical centrifugation.[33]
7
In this work we also applied this technique to GS data. In short, the amplitudes of the
simplified solution of the Lamm equation, neglecting all diffusional effects, are fitted to the
experimental data by means of a least-squares approach. Regularization is used and results in
a continuous sedimentation coefficient distribution. One drawback of the DBM is that it is
computationally more expensive. For further details the reader is referred to the original
publication and our previous work on this topic.
The sedimentation equivalent particle diameter d can be calculated using Stokes’ equation:
=
 (2)
where is the particle density, is the solvent density and η is the solvent viscosity. In the
present work the cumulative distributions of the sedimentation coefficient only represent the
fraction of silver-coated particles. This is due to the similar refractive indices of bare silica
and the 33 wt% aqueous sucrose solution used which lead to the uncoated silica particles not
being detected at the low concentration used.
2.2. Data evaluation strategy for patchy particles
2.2.1. General strategy
As stated in the previous section, AUC is sensitive to the size and density of the analyte, since
it measures its sedimentation rate. In the present work the overall size and mass of
monodisperse silica NPs are increased by the formation of a silver patch on their surface.
Since silver is roughly five times denser than colloidal silica, the presence of a patch results in
more pronounced sedimentation rates, which can be detected by AUC. However, for a patchy
particle, the change in size is typically small compared to the bare particle as the patch
thickness is about one order of magnitude smaller than the core particle diameter (Figure 1).
This is the main reason why light scattering or particle tracking techniques being sensitive to
the size or diffusion properties are of limited applicability when studying patchy particles of
the type discussed here. Nevertheless, the patch does slightly increase the hydrodynamic
diameter and, in addition, due to its asymmetrical shape will alter the friction the particle
8
experiences as it sediments. However, regarding the latter effect, we demonstrated through
simulations that the asymmetric distribution of silver on the silica cores has a negligible effect
on the sedimentation behavior (Figure S1 in the Supplementary Material). This opens up the
possibility to analyze the AUC data according to a simpler core-shell model.
The strategy pursued for the patchy particle characterization using MWL-AUC in this work is
depicted in Figure 2. First, MWL-AUC provides the sedimentation coefficient distributions
of the patchy particle samples, which can be converted to distributions of the patch-to-core
mass ratio or the patch mass per particle. Relating the median patch mass per particle with the
mass balance allows calculation of the patch yield. Second, MWL-AUC provides spectral
information about the different species being present in the sample. In principle, any species
having a different sedimentation coefficient could be examined. This would result in hundreds
of species per sample. However, the derived spectra would be very noisy in such case and the
data would not be of any use. Thus, only five averaged fractions will be considered for each
sample. Still, future improvements of the detection system will extend the number of
accessible species. As will be shown later in this work, the UV-peak position obtained from
the UV-Vis spectra by MWL-AUC can be related to the patch thickness. This is achieved by
empirical calibration via SEM image analysis. Combination of the patch mass and patch
thickness provides the patch coverage. Hence, in this work a methodology was established
that enables the determination of patch thicknesses and coverages from MWL-AUC.
Figure 2. Strategy established in this work for the multidimensional characterization of
patchy particles using MWL-AUC.
GravitationalSweep
Multiwavelength Analytical
Ultracentrifugation
Sedimentation
coeffici ent
distribution
Silver-silica
mass ratio ()
distribution
Extinction spectra
deconvolutedfro m
selected range of
distribution
Median patch thick ness t
patch,50
Median patch coverage c
patch,50
Correlat ion of UV-peak posit ion and
thicknes s:
Calibration via SEM image analy sis
Comparison with FEM based
simulation of extinction s pectra
UV-peak
positionλ
UV-peak
Median mass of
patches, m
patch,50
Patch yield Y
AUC
Core-sh ell equivalency model Mass balance for Ag and Si O
2
9
2.2.2. Determination of the Ag/SiO2 mass ratio distribution from the patchy particle
sedimentation coefficient distribution
The most important equations which allow calculation of the silver to silica mass ratio from
the AUC-determined sedimentation coefficient distribution will be provided in the following
section. The methodology is based on a core-shell equivalency model. In this the volume of a
silver patch is redistributed over the core particle surface as a uniformly thick shell (compare
upper and lower rows in Figure 1). Such an approximation is possible since, as noted above,
the core diameter is very large compared to the patch thickness. As such the presence of the
patch will not introduce significant shape anisotropic effects and differing hydration of the
silver and silica can also be neglected.
To establish the core-shell equivalency model we first define a spherical core particle with
diameter dcore and a patch on it having constant thickness  and volume . From
this we can derive the patch coverage  as the ratio of the patch volume to that of a
hypothetical complete shell with thickness  (Figure 1, upper row):
=

(3)
We now define the equivalent volume shell particle diameter as the diameter of a spherical
core-shell particle having a shell with the same volume,  as the patch:
 =
[
]= (4)
From a simple mass balance the effective density of the particle comprising core and patch is
given as:
= 


(5)
where  and  are the core and patch material densities, respectively. This expression
allows Equation 2 to be rewritten:



18 = 0 (6)
10
For each sedimentation coefficient in the AUC-measured distribution, Equation 6 is solved to
determine the diameter of the equivalent patch volume core-shell particle defined above. The
density of the uncoated core particle, which is microporous and hydrated, has to be
determined experimentally prior to the patchy particle studies if not known, e.g. by combining
AUC and DLS or SEM. An alternative is Kratky’s method which relies on densitometer
measurements.[34] However, this technique is of limited applicability to microporous particles,
such as Stöber silica as it leads to the partial specific volume which cannot be readily
associated with a macroscopic density.[3537] The partial specific volume is defined as the
change in volume when adding a certain mass of particles to the solution. Here, we are
interested in the effective density of the non-coated core particle, including all contributions
due to solvent filled and empty pores. Moreover, the size of the non-coated particle must be
determined prior to the patchy particles studies. In the Supplementary Material we show how
this information can be obtained by AUC measurements of silica dispersions in sucrose
solution (Figure S2) and also demonstrate that the finite width of the core PSD negligibly
influences the determined silver to silica mass ratio distribution (Figure S4). In addition, it is
shown that there is no measurable influence of heterogeneity in density for the uncoated core
particles (Figure S3).
Once the equivalent core-shell particle size and core particle size are known, the equivalent
volume shell thickness can be determined (Figure 1, lower row):
 =()/2 (7)
To determine the Ag/SiO2 mass ratio we determine the masses of the silver patch and silica
core:
=
(
) (8)
 =

 (9)
11
The Ag/SiO2 mass ratio is then calculated via the following expression for each value of d
obtained by solution of Equation 6 for each measured sedimentation coefficient:
=
 =


 (10)
The cumulative sedimentation coefficient distribution can thus be converted to a cumulative
Ag/SiO2 mass ratio distribution, ϕ. In this study the median of this distribution, ϕ50,AUC in
addition to its span will be of primary interest. The latter parameter is defined as:
 =,,
, (11)
where ϕ10,AUC and ϕ90,AUC correspond to the values of ϕ at the 10th, and 90th percentiles of the
cumulative distribution of ϕ, respectively.
2.2.3. Calculation of theoretical patch yield based on Ag/SiO2 mass ratio derived by AUC
data and mass balance
Assuming mass conservation for the absolute mass of silver and silica particles injected into
the reactor within one experiment and 100 % reaction yield for the reduction reaction of silver
nitrate to metallic silver patches, the median isolated absolute mass of all silica particles that
were coated with silver can be calculated:
,  =
, (12)
Knowing the mass of silica particles having a patch from the AUC distribution data and
knowing the total mass of silica particles used for silver coating reaction, allows the definition
of the AUC-determined patch yield:
 =, 
,  (13)
In this work we also determined the statistical patch yield  derived by manual analysis of
SEM images.  is determined by dividing the number of patchy particles  in images
of a sample by the total number of particles  observed. In this analysis we counted at
least 300 particles per sample.
12
 =
 (14)
2.2.4. Calculation of median patch coverage ,
Provided the patch thickness and median Ag/SiO2 mass ratio are known, an estimate of the
median patch coverage cpatch,50 can be made. From Equations 3 and 10 we obtain:

,
=


=
, 


((



)


)
(15)
As will be explained in Section 4.5,  can be obtained from a combination of
spectroscopy, electrodynamic simulations and SEM image analysis.
3. Results and Discussion
3.1. The influence of mixing conditions on the median and width of cumulative Ag/SiO2
mass ratio distribution
To demonstrate the benefit of AUC for the analysis of patchy particles with distributed patch
volumes we consider firstly the effect of mixing, the characteristic time for which can be
selected in our setup through the educt flow rates. Figure 3a and Figure 3b show SEM images
of samples with silver patchy particles synthesized at the highest (20 mL min-1) and the lowest
(2 mL min-1) flow rates used within this study. To illustrate the typical patch morphology,
these images were acquired at positions on the SEM sample holder with very high local patch
yield. More representative images taken at lower magnification for all samples at five
different flow rates are shown in Figure S9 in the Supplementary Material. In both Figure 3a
and Figure 3b silver patches of varying coverages can be seen alongside apparently uncoated
silica particles. While some silica particles appear fully coated, it cannot be ruled out that
these are high coverage patches oriented such that the exposed silica is directed towards the
substrate. Apart from the appearance of homogeneously nucleated silver particles for the
2 mL min-1 flow rate sample, rather little difference in patch thickness and coverage can be
inferred by the naked eye. The similar appearance of the patches can be explained by the
concentrations chosen for synthesis and especially by the very short characteristic mixing
times of ~16 ms and ~1 ms for highest and lowest flow rates, respectively, compared to our
13
earlier work (see Section VII of the Supplementary Material for details of the characteristic
mixing time determination).[38]
Figure 3. a) and b) SEM images of patchy particles synthesised at the lowest and highest flow
rates used in the study. c) Cumulative extinction weighted distributions of ϕ
at 600 nm
wavelength for samples synthesised at five different flow rates. d) Flow rate dependencies of
the median ϕ50,AUC and span of the ϕ distributions shown in c). e) Flow rate dependencies of
the SEM-determined statistical patch yield Yst and the AUC-determined yield YAUC derived
from ϕ50,AUC,. The scalebars in a) and b) correspond to 200 nm.
14
Despite the similar outward appearance of the samples prepared at low and high flow rate,
AUC measurements of the sedimentation coefficient at a single wavelength (600 nm) were
able to identify a clear difference in the silver loading of the patchy particles. Here, the
cumulative sedimentation coefficient distributions (Figure S9f) for the samples synthesized at
different flow rates were converted into distributions of the Ag/SiO2 mass ratio according to
the method described in the theoretical section. In addition, the distribution of the bare silica
particles is depicted for reference. The distributions, and their medians and spans are shown in
Figure 3c and Figure 3d, respectively. With increasing flow rate there is an initial slight
increase in ϕ50,AUC. This can be attributed to suppression of the formation of free silver
particles by increasing the flow rate i.e. the better-mixed reactions never reach the critical
supersaturation for homogeneous nucleation. Since more silver becomes available for patch
growth there is a commensurate increase in the median Ag/SiO2 mass ratio for the patchy
particle population. Interestingly, ϕ50,AUC decreases again for further increases in the flow rate.
Since the ratio of silver to silica in the educts is constant, according to Equations 12 and 13,
this decrease in ϕ50,AUC must correspond to an increase in patch yield  as indicated in
Figure 3e (red circles). This trend, which results from better mixing conditions leading to a
higher heterogeneous nucleation rate, is corroborated by the statistical patch yield 
determined manually from SEM images (black squares). Given that errors exist in both
methods of determining the yield, the differences in the absolute values of the yield are less
important. What is significant is that both show a similar trend, when the process variable
(mixing time) is adjusted. This means that the considerably more rapid determination of patch
yield by AUC can be used to optimize the process for maximum yield and arduous analysis of
micrographs can be avoided.
The effect of mixing on the width of the distribution of ϕ is also shown in Figure 3d (blue
squares). While at the lowest flow rate the span is ~1.6, for higher flow rates it is lower and
rather constant in the range 1.33 - 1.39. As verified in Figure S3, any influence of the size
15
distribution of the silica core particles on this span trend can be neglected. Rather, our
observations can be explained by two effects. For the sample synthesized at a low flow rate,
free silver particles with diameters larger than 100 nm were present (Figure 3b). Such large
silver particles will contribute to the cumulative extinction at 600 nm and thus influence the
sedimentation coefficient distribution at its lower end, which will in turn broaden the
distribution of ϕ. Moreover, as shown in earlier work, lower flow rates lead to a blurring of
the patch nucleation and patch growth periods, resulting in a broader range of patch
volumes.[26] For samples synthesized with higher flow rates corresponding to mixing times of
1 ms to 8 ms, the ϕ distribution width is narrower and the span is rather constant. This results
from the low amount of homogeneously nucleated silver particles, especially those of a size
having a strong extinction at 600 nm, and from better separation of patch nucleation and patch
growth due to the better mixed reaction conditions.
The results shown in Figure 3 therefore confirm that the median and span of the AUC-
determined ϕ distribution enable a simple evaluation of the effect of mixing on patch yield.
Since samples produced at 20 mL min-1 showed negligible free silver particles and had small
values for ϕ50,AUC and span, all further experiments reported here were carried out at this flow
rate.
3.2. Influence of silver nitrate concentration on the ϕ distribution
Silver patches were grown on silica cores particles using silver nitrate concentrations ranging
from 200 µM to 600 µM. Figure 4a and Figure 4b show the SEM images of the samples
synthesized at the lowest and highest concentrations, respectively. By qualitatively comparing
these two images, a difference in patch thickness can be observed. For the higher
concentration the metal coating is generally thicker. Nevertheless, as in the case of the flow
rate variation study described above, an objective assessment of the patch coverage cannot be
obtained from these images. This is also the case when comparing SEM images of samples
from all silver nitrate concentrations used (see Figure S10).
16
Figure 4. a) and b) SEM images of patchy particles synthesised with the lowest and highest
silver nitrate concentrations used in the study. c) Cumulative extinction weighted distribution
of ϕ at 600 nm wavelength for samples synthesised with five different silver nitrate
concentrations. d) Silver nitrate concentration dependencies of the median ϕ50,AUC and span of
the ϕ distributions shown in c). e) Silver nitrate concentration dependencies of the SEM-
determined patch yield Yst and the AUC-determined yield YAUC derived from ϕ50,AUC,. The
scalebars in a) and b) correspond to 200 nm.
Once again, however, AUC measurements can identify clear trends in the distributions of ϕ
(Figure 4c). Notably, there is a marked shift of the distribution towards larger values of ϕ with
increasing silver nitrate concentration. The extracted ϕ5 0,AUC values, as shown in Figure 4d,
17
clearly underline this trend which confirms that the patches produced at higher silver nitrate
concentrations contain more silver. Although this might appear an obvious statement, it
should be reminded that the heterogeneous nucleation rate and thus the yield will also be
affected by the silver concentration. Indeed, the derived values of yield  (red circles in
Figure 4e) show an increase up to 500 µM silver nitrate concentration followed by a slight
decrease. This trend is matched by the statistical yield  determined from SEM images
(black squares in Figure 4e).
The decrease in yield at the highest silver concentration indicates that the heterogeneous
nucleation rate is beginning to decrease again. Since homogeneously nucleated silver was not
observed in this sample we suggest that this decrease in yield is due to the mixing time
starting to again play a more significant role at the high silver ion concentration. In other
words, while mixing is rapid enough that in the bulk solution a supersaturation sufficient for
homogeneous nucleation is not achieved, the conditions for heterogeneous nucleation are not
reached uniformly across all silica particles. This in turn leads to the situation of some patches
nucleating and growing (and thus depleting silver) at the same time that others are still yet to
nucleate. The resulting broad patch volume distribution is corroborated by the increase in the
span of the ϕ distribution for the highest silver nitrate concentration compared to the lower
concentrations (blue squares in Figure 4d).
To conclude, in this section we have demonstrated the effective use of AUC at a single
wavelength for the determination of the Ag/SiO2 mass ratio distribution and estimation of the
patch yield for a range of patchy particles samples produced at different silver concentrations.
The yield trends are fully consistent with those determined by arduous image analysis.
3.3. Acquisition of Ag/SiO2 mass ratio dependent optical extinction spectra via MWL-
AUC
In this section we focus on deriving further information on the patchy particles in a sample
from their extinction spectrum. MWL-AUC offers the unique capability to address the optical
18
properties of different species in a mixture without any further purification and by means of
only a single experiment.[31,3941] Due to the highly morphology-dependent nature of the
plasmon resonance of silver nanostructures, we expect MWL-AUC to be especially suitable
for studying systems like the present patchy particles. Hence, a comparison between the set of
samples synthesized with increasing silver nitrate concentration and the optical properties of
the particle fractions having different ϕ within the same sample was carried out.
Figure 5a shows raw spectra obtained by deconvoluting data corresponding to all values of ϕ
for the five samples studied. Due to the high level of noise on these MWL-AUC acquired
spectra, we verified their validity with measurements of the same samples on a standard dual-
beam spectrophotometer (Figure S7 in the Supplementary Material). While minor differences
are observed due to the lower spectral resolution of the MWL-AUC, the agreement of the
spectra is very good. The higher deviations seen at wavelengths > 600 nm are due to the low
signal-to-noise-ratio of the MWL-AUC detector in that range (see marked increase in noise in
Figure 5a). Most importantly, it is clear that the amount of noise in the spectral region of
interest (UV-peak) does not preclude a quantitative analysis.
19
Figure 5. a) Raw spectra for patchy particle samples produced at different silver nitrate
concentrations and obtained by deconvoluting MWL-AUC data over the full range of
measured sedimentation coefficients. The UV-peak in the spectral region from 320 nm to 380
nm is fit with an asymmetric double sigmoidal function. b) Dependency of the UV-peak
position obtained from the spectra in a) on the median Ag/SiO2
mass ratio of the five samples.
c) Deconvoluted MWL-AUC spectra of five fractions of the patchy particle sample
synthesised at 600 µM silver nitrate concentration. Each spectrum corresponds to a fraction of
patchy particles with increasing median Ag/SiO2
mass ratios which contribute equally with
20% to the cumulative extinction at 600 nm. The legend corresponds to the percentage ranges
of the total extinction. For the sake of clarity, the spectra are plotted with y-offset. d) UV-peak
positions of MWL-AUC extinction spectra for samples synthesised with different silver
nitrate concentrations for fractions of the Ag/SiO2 mass ratio with equal 20% contributions to
the total extinction at 600 nm. The spectra used in the analysis are shown in c) and Figure
S12. The UV-peak positions were obtained by fitting an asymmetric double sigmoidal
function in the wavelength range 320 to 380 nm. The abscissa labels correspond to the
percentage ranges of the total extinction.
Figure 5a indicates that with increasing silver nitrate concentration the extinction increases
and shifts towards shorter wavelengths and some peaks appear in the spectral region around
500 to 600 nm. These observations are consistent with both the already-observed patch yield
increase (Figure 4e) and FEM simulated spectra of patches with increasing coverage or
thickness (Figure S11 in the Supplementary Material). A further, and, as we shall see, highly
20
useful feature of these spectra is the peak between 320 nm and 360 nm (hereafter referred to
as the “UV-peak”). Figure 5b shows that the position of this peak, determined from the
MWL-AUC spectra in Figure 5a by fitting with an asymmetric double sigmoidal function,
shifts monotonously from about 333 nm to 345 nm with increasing ϕ50,AUC. The position of
this spectral feature therefore appears to be strongly correlated with the amount of silver
attached to the silica sphere. This dependency is also observed when the spectra of individual
fractions of a sample are investigated. Figure 5c shows, exemplarily for the sample
synthesized at 600 µM silver nitrate, the deconvoluted and smoothed extinction spectra of
fractions within five intervals of cumulatively increasing ϕ which contribute almost equally
with ~20% relative extinction each at 600 nm i.e. the extinction of the fractional spectra are
equal at this wavelength. Figure S12 shows the same analysis for the other silver
concentration samples with similar trends apparent throughout. The lowest ϕ fraction of each
sample generally has a spectrum with just the UV-peak and a broad peak at around 650-700
nm. With an increase in concentration the spectra show the appearance of features between
400 nm and 600 nm with increasing ϕ fraction. As with the complete sample spectra of Figure
5a we analyzed the position of the UV-peak for all five fractions of each sample. Figure 5d
shows the UV-peak position plotted against the cumulative fraction for each sample. It is clear
that the individual fractions generally show an increase (red-shift) in UV-peak position for
both increasing silver nitrate concentration and fraction. For the heaviest fraction (80-100%)
there is a levelling off or even a blue-shift of this peak which seems to contradict the trend in
Figure 5b. There may be various reasons for this observation. First, the deconvoluted
spectrum is in general noisier at the upper end of the distribution of ϕ which makes it difficult
to determine the exact peak position. This is because of the higher absolute extinction
measured and thus lower signal-to-noise ratio at the beginning of the measurement, where all
species are still present in the dispersion. Second, the interval with highest sedimentation
coefficients might also be influenced by the presence of agglomerates of patchy particles
21
having themselves a lower ϕ and thus a smaller UV-peak. Similarly, particles having multiple
thin patches might contribute to the extinction of this fraction.
In conclusion, in this section we have shown that useful spectral information can be obtained
for ranges of the Ag/SiO2 mass ratio distribution by deconvoluting GS data acquired by
MWL-AUC. From these spectra we see that there is a particularly strong correlation between
the MWL-AUC determined UV-peak position and Ag/SiO2 mass ratio for the silver patchy
particles considered in this work.
3.4. Determination of patch thickness and coverage from MWL-AUC data
To generate a deeper understanding of the meaning of the UV-peak, FEM based simulations
of the extinction spectra of idealized cup-like patchy particles having patch coverages
between 10% and 90% and patch thicknesses between 5 nm to 35 nm were performed (see
Section IX of the Supplementary Material for more details). The simulated silver patches
were assumed to have a perfect smooth surface, constant thickness and regular sharp edges.
To obtain each wavelength point, multiple simulations covering different particle orientations
and light polarizations were averaged. The simulated extinction spectra of the UV region for
patchy particles with different patch coverage and thickness indicated that the UV-peak
position is almost completely independent of the patch coverage (Figure S11a). This is
consistent with the fact that this feature has been attributed to the out-of-plane quadrupole
resonance.[42] Hence we can infer that the correlation between the UV-peak position and the
Ag/SiO2 mass ratio (Figure 5b) implies that the patch thickness must increase with increasing
silver nitrate concentration. To obtain an expression for the relation between the UV-peak
position and patch thickness we first plotted the average position of the peak in the
simulations against the patch thickness used as input (see Figure S14). The resulting curve
was fitted with a Boltzmann sigmoid function leading to an expression relating UV-peak
position to “optical” patch thickness ,. We used this expression to estimate the patch
thickness of the samples in the silver nitrate concentration series. Rather than taking the UV-
22
peak position from the spectra in Figure 5a we analyzed spectra obtained by summating the
three central fractions i.e. 20%-80% in Figure 5c and S12. This was done so as to avoid
potential errors introduced by patchy particles at the extremities of the cumulative ϕ
distribution, particularly the heaviest fraction, as already noted above. Figure S12 shows the
fitting of the UV-peaks in these spectra with an asymmetric double sigmoidal function. The
resulting UV-peak positions were used to estimate , for each silver nitrate
concentration. The results are plotted in Figure 6a (blue squares) and indicate a monotonous
increase in thickness with concentration from 13.2 nm up to 24.1 nm.
Figure 6. a) Patch thickness versus silver nitrate concentration obtained by SEM image
analysis (black crosses) and two different methods of processing UV-peak data from MWL-
AUC spectra. The blue squares correspond to patch thickness values obtained using an FEM-
simulation determined relation (see Figure S14) while the red circles are for patch thickness
values obtained using Equation 16, an empirical fit of the UV-peak position versus SEM-
determined patch thickness relation (see Figure S17). For the UV-peak position derived data
the original spectra were deconvoluted from the MWL-AUC data for the central 20%-80%
fraction of the cumulative extinction at 600 nm. Fitting is shown in Figure S13. b) Patch
coverage versus silver nitrate concentration determined using patch thicknesses from UV-
peak positions through Equation 16 (red circles in a). c) Patch thickness and d) patch coverage
determined for individual fractions of the concentration series samples and plotted against the
23
Ag/SiO2 mass ratio. Horizontal error bars correspond to the mass ratio ranges of the fractions.
For clarity the lightest fraction was taken from a starting mass ratio corresponding to 2% of
the cumulative extinction at 600 nm while the heaviest fraction was truncated at a mass ratio
corresponding to 98% of the cumulative extinction at 600 nm.
To verify the accuracy of the determined patch thicknesses we carried out arduous, systematic
image analysis of SEM micrographs. The procedure followed is described in Section VIII of
the Supplementary Material and illustrated with Figure S15 showing the various stages of
processing for one image. Briefly, we made use of the fact that for all samples across the
silver nitrate concentration range used, a large proportion of the patchy particles had high
coverage patches, and, for many of those, the exposed silica was hidden from view. For our
image analysis we selected only those particles and assumed their diameter corresponded to
the core particle diameter plus twice the patch thickness. Analysis of at least 120 particles in
each sample, along with a sample of only core particles, enabled us to determine mass
weighted PSDs (Figure S16a). These were fitted with Boltzmann sigmoidal functions and
thus the cumulative mass weighted distributions of , could be determined (Figure
S16b). The median values of these distributions are plotted in Figure 6a (black crosses). It is
evident that the patch thicknesses estimated using the UV-peak position and the simulation-
derived relation are somewhat lower than the “real” SEM-determined patch thicknesses.
There are several reasons why this might be the case. On the one hand, it is very likely due to
the fact that the simple geometry used in the FEM simulation does not reflect reality. As seen
in the SEM images of Figure 4a and 5a, the patches are neither smooth nor single crystalline.
On the other hand, we cannot rule out at this stage that they are not exactly concentric i.e.
their thickness tapers towards their edges. Hence it was recognized that rather than using a
simulation-based calibration of the UV-peak to patch thickness relation, an empirical
expression should be developed. We obtained this by plotting the UV-peak positions
(determined in Figure S13) against the SEM-determined median patch thicknesses
(determined in Figure S16b). Figure S17 in the Supplementary Material shows these data
24
mostly lie on a line slightly offset to the simulated data. We performed a linear fit to the
empirical data, ignoring the outlying data point corresponding to the 200 µM sample. The
latter exclusion is reasonable since the largest error in the SEM image analysis is likely in this
sample due to it having the thinnest patches. Moreover, the UV-peak is the least prominent of
the five samples, making its analysis most inaccurate. The linear fit of Figure S17 provided
the following empirical relation between the UV-peak position  and the patch
thickness ,:
, = 1.067  340.9  (16)
From this relation a new estimate of the patch thickness from the MWL-AUC spectra could
be made. As shown in Figure 6a (red circles) this data closely follows the SEM-measured
patch thickness. This good agreement is of course due to the fact that the latter data set was
used to develop the empirical relation.
Once the median silver to silica mass ratio ϕ and median patch thickness is known, the
median patch coverage can be estimated from Equation 15. This was done for the five
samples of the silver nitrate concentration series using the patch thickness as determined from
Equation 16. The results are shown in Figure 6b. While the median patch coverages are rather
high (between 86% and 93% for concentrations between 300 µM and 600 µM) it is not
possible to identify a clear trend. This is rather unsurprising since SEM images for all samples
(Figure S10) indicate predominantly high coverage patches. In other words, the patch
coverage is not strongly dependent on the silver nitrate concentration.
To demonstrate that information on the dispersity of patchy particle morphological features
can be acquired by MWL-AUC we applied the relations for patch thickness (Equation 16) and
coverage (Equation 15) to fractionized spectroscopic data from the silver nitrate concentration
series of Figure 5c and S10. The full derived data is listed in Table S4 in the Supplementary
Material and is plotted in Figure 6c (patch thickness) and 6d (patch coverage). These plots
25
show the Ag/SiO2 mass ratio on the abscissa and horizontal error bars show the width of the
fractions. Thus it is possible to identify that for increasing silver nitrate concentration the
dispersity of patch thicknesses over all fractions increases from being rather narrow at 13 to
17 nm for the 200 µM sample to being rather broad at 20 to 34 nm for the 600 µM sample.
Due to the inverse relation between thickness and coverage in Equation 15, the variation of
patch coverage is rather complementary to the thickness. Thus the lowest concentration
sample has a large spread of coverages over a given mass ratio range, while the highest
concentration sample has a comparatively narrower range of coverages. It can also be seen
that the heaviest fraction in each sample has a coverage of above 100%. As already noted
above, this fraction is likely to contain aggregates of patchy particles or multi-patch particles
and the unexpectedly low thickness leads to this non-physical coverage value.
The above analysis demonstrates that realistic trends in patch thickness and coverage and
information on the dispersity of these parameters within individual samples can be obtained
by MWL-AUC measurements alone. However, more work is required in the future to refine
our approach. As shown in the electrodynamic simulations, the NIR region of the spectrum
contains information about the patch coverage (Figure S11a). Hence, upgrading of the
spectrometer in our setup to obtain spectra deeper into the NIR will open up an additional
means to further improve the determination of patch coverage. Overall, due to the fact that it
avoids arduous microscopy studies, we believe our method will greatly assist with the
systematic exploration and optimization of process variables in order to target other patchy
particle morphologies not discussed here e.g. narrowly dispersed, small coverage patches or
even multiple patch particles.
26
4. Conclusion
In this work, a method based on multiwavelength analytical ultracentrifugation (MWL-AUC)
for the characterization of silver patches on silica spheres, synthesized via continuous a flow
process, has been shown. This technique helps to rationally investigate the patch yield in
addition to the amount and geometrical distribution of silver on the silica particles.
First, the use of AUC measurements to experimentally investigate the effect of the process
flow rate was demonstrated. The combination of a low median Ag/SiO2 mass ratio and a
narrow width of the mass ratio distribution indicated a high sample quality regarding patch
yield and patch uniformity for the sample synthesized at highest flow rate, i.e. shortest
characteristic mixing time.
In the second part, patchy particles were synthesized at constant high flow rate and with
increasing silver nitrate concentration. The results showed that the patch yield in addition to
the median Ag/SiO2 mass ratio can be estimated from AUC measurements. The trends in
these parameters compared favourably with those obtained by arduous analysis of electron
micrographs.
Since the silver patchy particles possess plasmonic properties which depend sensitively on the
patch thickness and patch coverage, MWL-AUC-determined extinction spectra were
investigated in detail. Making use of FEM-based electrodynamic simulations it was found that
the observed UV-peak, being present in the range of 330 nm to 360 nm, is directly related to
the patch thickness. While simulations could provide a relation between the peak position and
thickness, when applied to real data from MWL-AUC spectra the patch thicknesses were
underestimated compared to SEM-determined thicknesses. Thus, an empirical relationship
was obtained by correlating UV-peak position with patch thicknesses determined by SEM
image analysis. Besides patch thickness, this enabled, with use of the already-known Ag/SiO2
mass ratio, the patch coverage to be estimated. Finally, we demonstrated the true value of
MWL-AUC for analysing the dispersity of this new type of material by determining the
27
median patch thicknesses and coverages of five individual fractions of samples produced at
different silver nitrate concentrations.
In conclusion, our work presents a powerful, fast and versatile method to analyze complex
plasmonic anisotropic particles by means of just a single measurement. Future work is
addressed to improve the optical equipment installed inside the AUC in order to further
improve the quality of the data obtained, especially at longer wavelengths. This is critical in
order to validate the derived information on the patch coverage, since the corresponding
plasmonic features are mainly located at longer wavelengths.
5. Materials and Methods
Synthesis of colloidal silica particles:
Heat-treated colloidal silica particles, produced by the Stöber method,[43] having a mean
diameter of 255 nm, the same as those used in our earlier work,[26] were applied as core
particles for patchy particles synthesis. Their synthesis proceeded following the method of
Stöber et al.[43] First, a mixture of 497 g absolute ethanol (Carl Roth GmbH, Germany), 26.4 g
aqueous ammonium hydroxide (32 %, Carl Roth GmbH, Germany) and 62.71 g ultrapure
water was prepared. By the subsequent addition of 50 mL tetraethylorthosilicate (98 %,
Merck KGaA, Germany) the formation of colloidal silica was initiated. The reaction was
allowed to proceed for 5 h at room temperature before it was finally stopped by four times
washing of the synthesized silica particles with ultrapure water. After drying of the derived
particles, the powder was calcined at 1000 °C for 6 h.
Continuous flow synthesis and stabilisation with sucrose of patchy- and complete nanoshell
particles: In extension to our earlier work,[26] patchy particles were produced here in a
continuous process using a KM-micromixer (Figure 7), similar to that recently published by
S. Watanabe et al. for gold nanoshell and gold patch synthesis.[44,45] This type of mixer
combines reasonable flow rates for lab-scale synthesis with higher mixing performance
28
(characteristic mixing times of 1-20 ms) than that of a simple T-mixer used earlier
(characteristic mixing times greater than ~90 ms), as verified by the Villermaux-Dushmann
reaction.[38] Two educt streams were injected into the microreactor via a syringe pump. The
first stream contained aqueous ammonia solution (32 %, Carl Roth GmbH, Germany) at a
concentration of 27 mM and for the second stream a dispersion of colloidal silica (25 mg/L),
silver nitrate (99 %, Carl Roth GmbH, Germany) with concentrations in the range of 200 µM
to 600 µM and 152 mM formaldehyde (37 % aqueous solution, methanol stabilized, Carl Roth
GmbH, Germany) was prepared. As the preferred temperature for high yield silver patch
formation has already been shown to be 50 °C,[25] the microreactor was immersed into a
stirred water bath and two stainless steel heating coils (inner diameter = 1.6 mm, length of
tubing = 2 m) were installed before its inlets to ensure the educt streams were at the correct
temperature prior to mixing. After the reactor, a coiled polyamide tube (inner diameter =
2 mm, length of tubing = 6 m) was used to ensure sufficient residence time for the reduction
reaction of ammoniacal silver nitrate to metallic silver by formaldehyde at all flow rates used.
For each syringe, the flow rates were varied within the range of 2 mL min-1 to 20 mL min-1.
An overview of synthetic conditions used in all experiments can be found in Table S1 in the
Supplementary Material.
Figure 7. Schematic of the KM-microreactor used for the growth of silver patches on silica
NPs.
After synthesis, an aliquot of 2 mL product dispersion was added to 1 g D(+) sucrose (>99,5%
p.a.) in powder form (Carl Roth GmbH, Germany). The dissolution of sucrose was achieved
by short sonication and shaking by hand.
29
Scanning electron microscopy: SEM samples were prepared directly after synthesis before the
addition of sucrose and without any further purification. 3 µL of product particles were
dropped onto a silicon wafer and were allowed to dry at ambient conditions. The patchy silver
particles were analysed using a ULTRA 55 field emission SEM (Carl Zeiss NTS GmbH,
Germany). Statistical patch yields were obtained by manual counting of particles with and
without a patch and using Equation 14.
UV-Vis-NIR spectroscopy: Since UV-Vis-NIR measurements were done of samples after the
addition of sucrose, a reference sample of 1 g sucrose and 2 mL ultrapure water was prepared.
The extinction spectra were recorded in the wavelength range of 300 nm to 1350 nm using a
PerkinElmer Lambda 950 UV-Vis-NIR spectrophotometer.
Analytical ultracentrifugation: GS experiments of patchy particles were performed using a
modified preparative ultracentrifuge, type Optima L-90K (Beckman Coulter, USA), equipped
with an integrated UV-Vis MWL detector.[39,46] For all runs, titanium centrepieces
(Nanolytics, Germany) with path lengths of 12 mm were used. GS data were acquired at
20 °C and rotor speeds ranging from 2000 rpm to 20000 rpm depending on the sample
investigated. MWL data between 250 nm and 750 nm was recorded and saved in binary
format.[31] Figure S5 exemplarily shows GS raw data for a wavelength of 600 nm for all
samples discussed in this work. The bare silica particles in 33 wt% aqueous sucrose solution
were measured in a SV run in the same experimental setup. The rotor speed was 3000 rpm
and the temperature was 20 °C. It has to be emphasised that such low rotor speeds will not
allow a sucrose density gradient to build up during the GS or SV experiments. Thus, it can be
safely assumed that the sucrose is equally distributed in the measurement cell. Moreover, it
should be pointed out that the necessary low scanning speed of the MWL setup at a typical
rotor speed of 2000 rpm prohibit SV or band sedimentation experiments to be performed as an
alternative for spectral characterization for sedimentation coefficient-resolved fractions of the
patchy particle samples considered in this work.
30
Data were evaluated with HDR-MULTIFIT (version 1.06).[31] Sedimentation coefficient
distributions were obtained using DBM based on the works by Schuck et al.[32] and Walter et
al.[33] Distributions are either represented as cumulative distributions or as their first
derivative, herein referred to as density distributions according to ISO 9276-1. As noted above
the DBM analysis of GS experiments allows for less noisy distributions compared to the
traditional analysis. Details on the evaluation procedure can be found in the original
publications.[32,33] Most importantly, both strategies result in very similar results and could be
used interchangeably for the given investigations (see Figure S6 and Section IV in the
Supplementary Material for further details). For the extraction of extinction spectra, the model
independent procedure implemented in HDR-MULTIFIT was chosen due to computational
limitations of the DBM analysis when applied to all wavelengths.[31] The SV data obtained for
the bare silica particles was evaluated using the ls-g*(s) model in SEDFIT (version 14.81)
(Figure S2). The sedimentation coefficient distribution of the bare silica particles was
corrected for Mie scattering in HDR-MULTIFIT. Correlation of the AUC and SEM size data
resulted in an effective density of 2.1327 g cm-3 for the bare silica, which was used
throughout this work.
Acknowledgements
The authors gratefully acknowledge the Deutsche Forschungsgemeinschaft (DFG, German
Research Foundation) for funding of the Collaborative Research Centre 1411 “Design of
Particulate Products” (Project-ID 416229255) and the Cluster of Excellence “Engineering of
Advanced Materials” (Project-ID 53244630). The authors thank the Deutscher Akademischer
Austauschdienst (DAAD) and Japan Society for the Promotion of Science (JSPS) for financial
and travel support. Additionally, the authors acknowledge Shuji Ohsaki for his support in
experimental work at the Department of Chemical Engineering, Kyoto University. Thomas
Meincke and Johannes Walter contributed equally to this work.
31
Received: ((will be filled in by the editorial staff))
Revised: ((will be filled in by the editorial staff))
Published online: ((will be filled in by the editorial staff))
References
[1] Zhang, S. C. Glotzer, Nano Lett. 2004, 4, 1407.
[2] P. P. Ghimire, M. Jaroniec, J. Colloid Interface Sci. 2021, 584, 838.
[3] D. Rodríguez-Fernández, L. M. Liz-Marzán, Part. Part. Syst. Charact. 2013, 30, 46.
[4] M. Frederiksen, V. E. Bochenkov, M. B. Cortie, D. S. Sutherland, J. Phys. Chem. C
2013, 117, 15782.
[5] Z. He, I. Kretzschmar, Langmuir 2012, 28, 9915.
[6] Z. He, I. Kretzschmar, Langmuir 2013, 29, 15755.
[7] L. Bradley, Y. Zhao, Langmuir 2016, 32, 4969.
[8] D. Rodriguez-Fernandez, J. Perez-Juste, I. Pastoriza-Santos, L. M. Liz-Marzan,
ChemistryOpen 2012, 1, 90.
[9] G. K. Larsen, N. Stom, Y. Zhao, RSC Adv 2015, 5, 101257.
[10] M. E. Stewart, C. R. Anderton, L. B. Thompson, J. Maria, S. K. Gray, J. A. Rogers, R. G.
Nuzzo, Chem. Rev. 2008, 108, 494.
[11] B. Ai, Y. Yu, H. Mohwald, G. Zhang, B. Yang, Adv. Colloid Interface Sci. 2014, 206, 5.
[12] Y. Hou, Y. Li, L. Wang, D. Chen, M. Bao, Z. Wang, J. Colloid Interface Sci. 2019, 556,
54.
[13] Y. Yi, L. Sanchez, Y. Gao, K. Lee, Y. Yu, Chem. Mater. 2017.
[14] C. Kaewsaneha, P. Tangboriboonrat, D. Polpanich, M. Eissa, A. Elaissari, ACS Appl.
Mater. Interfaces 2013, 5, 1857.
[15] L.-T.-C. Tran, S. Lesieur, V. Faivre, Expert opinion on drug delivery 2014, 11, 1061.
[16] W. He, J. Frueh, Z. Wu, Q. He, ACS Appl. Mater. Interfaces 2016, 8, 4407.
[17] W. He, J. Frueh, Z. Wu, Q. He, Langmuir 2016, 32, 3637.
32
[18] F. Chang, S. Ouhajji, A. Townsend, K. Sanogo Lacina, B. G. P. van Ravensteijn, W. K.
Kegel, J. Colloid Interface Sci. 2021, 582, 333.
[19] M. A. Boles, M. Engel, D. V. Talapin, Chem. Rev. 2016, 116, 11220.
[20] B. Bharti, D. Rutkowski, K. Han, A. U. Kumar, C. K. Hall, O. D. Velev, J. Am. Chem.
Soc. 2016, 138, 14948.
[21] G. Loget, A. Kuhn, J. Mater. Chem. 2012, 22, 15457.
[22] A. Walther, A. H. E. Muller, Chem. Rev. 2013, 113, 5194.
[23] A. B. Pawar, I. Kretzschmar, Macromol. Rapid Commun. 2010, 31, 150.
[24] Q. Chen, S. C. Bae, S. Granick, Nature 2011, 469, 381.
[25] H. Bao, W. Peukert, R. N. Klupp Taylor, Adv. Mater. 2011, 23, 2644.
[26] T. Meincke, H. Bao, L. Pflug, M. Stingl, R. N. Klupp Taylor, Chem. Eng. J. 2017, 308,
89.
[27] S. E. Wawra, L. Pflug, T. Thajudeen, C. Kryschi, M. Stingl, W. Peukert, Nat. Commun.
2018, 9, 4898.
[28] X. Lu, M. Rycenga, S. E. Skrabalak, B. Wiley, Y. Xia, Annual review of physical
chemistry 2009, 60, 167.
[29] H. G. Müller, Colloid Polym. Sci. 1989, 267, 1113.
[30] W. Mächtle, Biophysical journal 1999, 76, 1080.
[31] J. Walter, W. Peukert, Nanoscale 2016, 8, 7484.
[32] P. Schuck, P. Rossmanith, Biopolymers 2000, 54, 328.
[33] J. Walter, T. Thajudeen, S. Suss, D. Segets, W. Peukert, Nanoscale 2015, 7, 6574.
[34] O. Kratky, H. Leopold, H. Stabinger, in Part D: Enzyme Structure, Vol. 27, Elsevier
1973, p. 98.
[35] S. Li, Q. Wan, Z. Qin, Y. Fu, Y. Gu, Langmuir 2015, 31, 824.
[36] S. R. Parnell, A. L. Washington, A. J. Parnell, A. Walsh, R. M. Dalgliesh, F. Li, W. A.
Hamilton, S. Prevost, J. P. A. Fairclough, R. Pynn, Soft matter 2016, 12, 4709.
33
[37] S. Li, Q. Wan, Z. Qin, Y. Fu, Y. Gu, Langmuir 2016, 32, 9180.
[38] J.-M. Commenge, L. Falk, Chemical Engineering and Processing: Process
Intensification 2011, 50, 979.
[39] J. Walter, K. Lohr, E. Karabudak, W. Reis, J. Mikhael, W. Peukert, W. Wohlleben, H.
Colfen, ACS Nano 2014, 8, 8871.
[40] J. Walter, P. J. Sherwood, W. Lin, D. Segets, W. F. Stafford, W. Peukert, Analytical
chemistry 2015, 87, 3396.
[41] E. Karabudak, E. Brookes, V. Lesnyak, N. Gaponik, A. Eychmuller, J. Walter, D. Segets,
W. Peukert, W. Wohlleben, B. Demeler, H. Colfen, Angew. Chem. Int. Ed. 2016, 55,
11770.
[42] K. L. Kelly, E. A. Coronado, L. L. Zhao, G. C. Schatz, J. Phys. Chem. B 2003, 107, 668.
[43] W. Stöber, A. Fink, E. Bohn, J. Colloid Interface Sci. 1968, 26, 62.
[44] S. Watanabe, T. Hiratsuka, Y. Asahi, A. Tanaka, K. Mae, M. T. Miyahara, Part. Part.
Syst. Charact. 2015, 32, 234.
[45] S. Watanabe, Y. Asahi, H. Omura, K. Mae, M. T. Miyahara, Adv. Powder Technol. 2016,
27, 2335.
[46] J. Pearson, J. Walter, W. Peukert, H. Cölfen, Anal. Chem. 2018, 90, 1280.
34
List of Figure Captions
Figure 1. Schematic showing 14 nm thick silver patches with various coverages (cpatch) on
255 nm diameter silica spheres (upper row) and the equivalent patch volume core-shell
particles with shell thickness tshell and having the same silver to silica mass ratio (ϕ) as the
patch particles (lower row).
Figure 2. Strategy established in this work for the multidimensional characterization of
patchy particles using MWL-AUC.
Figure 3. a) and b) SEM images of patchy particles synthesised at the lowest and highest flow
rates used in the study. c) Cumulative extinction weighted distributions of ϕ
at 600 nm
wavelength for samples synthesised at five different flow rates. d) Flow rate dependencies of
the median ϕ50,AUC and span of the ϕ distributions shown in c). e) Flow rate dependencies of
the SEM-determined statistical patch yield Yst and the AUC-determined yield YAUC derived
from ϕ50,AUC,. The scalebars in a) and b) correspond to 200 nm.
Figure 4. a) and b) SEM images of patchy particles synthesised with the lowest and highest
silver nitrate concentrations used in the study. c) Cumulative extinction weighted distribution
of ϕ at 600 nm wavelength for samples synthesised with five different silver nitrate
concentrations. d) Silver nitrate concentration dependencies of the median ϕ50,AUC and span of
the ϕ distributions shown in c). e) Silver nitrate concentration dependencies of the SEM-
determined patch yield Yst and the AUC-determined yield YAUC derived from ϕ50,AUC ,. The
scalebars in a) and b) correspond to 200 nm.
35
Figure 5. a) Raw spectra for patchy particle samples produced at different silver nitrate
concentrations and obtained by deconvoluting MWL-AUC data over the full range of
measured sedimentation coefficients. The UV-peak in the spectral region from 320 nm to 380
nm is fit with an asymmetric double sigmoidal function. b) Dependency of the UV-peak
position obtained from the spectra in a) on the median Ag/SiO2
mass ratio of the five samples.
c) Deconvoluted MWL-AUC spectra of five fractions of the patchy particle sample
synthesised at 600 µM silver nitrate concentration. Each spectrum corresponds to a fraction of
patchy particles with increasing median Ag/SiO2
mass ratios which contribute equally with
20% to the cumulative extinction at 600 nm. The legend corresponds to the percentage ranges
of the total extinction. For the sake of clarity, the spectra are plotted with y-offset. d) UV-peak
positions of MWL-AUC extinction spectra for samples synthesised with different silver
nitrate concentrations for fractions of the Ag/SiO2 mass ratio with equal 20% contributions to
the total extinction at 600 nm. The spectra used in the analysis are shown in c) and Figure
S12. The UV-peak positions were obtained by fitting an asymmetric double sigmoidal
function in the wavelength range 320 to 380 nm. The abscissa labels correspond to the
percentage ranges of the total extinction.
Figure 6. a) Patch thickness versus silver nitrate concentration obtained by SEM image
analysis (black crosses) and two different methods of processing UV-peak data from MWL-
AUC spectra. The blue squares correspond to patch thickness values obtained using an FEM-
simulation determined relation (see Figure S14) while the red circles are for patch thickness
values obtained using Equation 16, an empirical fit of the UV-peak position versus SEM-
determined patch thickness relation (see Figure S17). For the UV-peak position derived data
the original spectra were deconvoluted from the MWL-AUC data for the central 20%-80%
fraction of the cumulative extinction at 600 nm. Fitting is shown in Figure S13. b) Patch
coverage versus silver nitrate concentration determined using patch thicknesses from UV-
36
peak positions through Equation 16 (red circles in a). c) Patch thickness and d) patch coverage
determined for individual fractions of the concentration series samples and plotted against the
Ag/SiO2 mass ratio. Horizontal error bars correspond to the mass ratio ranges of the fractions.
For clarity the lightest fraction was taken from a starting mass ratio corresponding to 2% of
the cumulative extinction at 600 nm while the heaviest fraction was truncated at a mass ratio
corresponding to 98% of the cumulative extinction at 600 nm.
Figure 7. Schematic of the KM-microreactor used for the growth of silver patches on silica
NPs.
37
Supplementary Material
Determination of the yield, mass and structure of silver patches on colloidal silica using
multiwavelength analytical ultracentrifugation
Thomas Meincke, Johannes Walter, Lukas Pflug, Thaseem Thajudeen, Andreas Völkl, Paola
Cardenas Lopez, Maximilian Uttinger, Michael Stingl, Satoshi Watanabe, Wolfgang Peukert
and Robin Klupp Taylor*
Table of contents
I
Influence of patchiness on sedimentation behaviour
II
Influence of core size distribution on core-shell correlation
III
Raw data of gravitational sweep (GS) experiments
IV
Comparison of data analysis methods for gravitational sweep (GS) data
V
Comparison of offline UV-VIS spectra and AUC derived spectra
VI
Comparison of GS data and band sedimentation
VII
Characterisation of mixing via iodate-iodide reaction
VIII
Image analysis to determine patch thickness distribution from SEM images
IX
FEM based simulations of optical extinction spectra of silver patchy particles
X
Supplementary tables
XI
Further supplementary figures
38
I. Influence of patchiness on sedimentation behaviour
In this work sedimentation coefficient distributions derived by analytical ultracentrifugation
(AUC) are converted to distributions of the patch-to-core mass ratio through implementation
of a core-shell model. In this it is assumed that the patch, which partially coats the silica
particle, is equivalent, in a hydrodynamic sense, to a complete shell of uniform thickness and
equal volume to the patch. As the patch thickness is small compared to the overall size of the
patchy particle, this is a valid assumption as the hydrodynamic diameter is only slightly
affected when altering the shell morphology. Nevertheless, we evaluated the possible error
encountered in making this assumption.
The error can be exemplarily approximated when computing the sedimentation coefficients
for hemispherical or quadrant patches with the help of hydrodynamic modelling (e.g. bead
modelling or path integration technique) to account for all frictional effects. In addition, the
sedimentation coefficients for a complete shell of constant thickness but identical mass can be
derived. As the latter is used for the core-shell correlation, the deviation in both sedimentation
coefficients is a direct measure for the error encountered in the correlation. The sedimentation
coefficient for a non-uniform patch is expected to be smaller as its hydrodynamic diameter
will be larger compared to an ideal spherical particle, owing to the increased frictional effects
of the anisotropic patch morphology.
Figure S1 depicts the results of the error calculations for the cases of hemispherical and
quadrant patches. The maximum patch thickness one would expect is about 30 nm. It can be
seen from the data that the maximum relevant Ag/SiO2 mass ratio is ~ 225 % for the
hemispherical patch and ~ 125 % for the quadrant patch. This relates to a maximum error of ~
1.5 % for the sedimentation coefficient in the case of a hemispherical patch. For a quadrant
patch, the maximum error is ~ 2.5 % and thus somewhat larger. This is because the relative
hydrodynamic diameter is larger compared to the hemispherical patch. However, it has to be
taken into account that one single quadrant patch is not very representative for the patchy
39
particles studied herein. Spherical caps would be more representative but are difficult to
simulate as there are too many degrees of freedom. Still, the error would be smaller as the
resulting patchy particles would be smoother. Hence, the maximum error is expected to be
smaller than for the quadrant patch and more in the order of the hemispherical patch. Overall,
we believe that it is reasonable to conclude that the errors encountered in the core-shell-
correlation when assuming a shell of constant thickness are very small and in the order of the
other possible errors relevant for this correlation (i.e., solvent parameters, uncertainty in
sedimentation coefficient, etc.).
Figure S1. a) Ratio of “true” sedimentation coefficient for the hemispherical patch and
theoretical sedimentation coefficient when assuming a complete shell of constant thickness
with identical mass as a function of Ag/SiO2 mass ratio. b) Ratio of “true” sedimentation
coefficient for the quadrant patch and theoretical sedimentation coefficient when assuming a
complete shell of constant thickness with identical mass as a function of Ag/SiO2 mass ratio.
This correlation is independent of the core size.
40
I. Influence of core size distribution on core-shell correlation
Figure S2. a) Sedimentation velocity data (points) of bare SiO2 nanoparticles together with
the fit profiles (lines) and residuals as derived by the least squares model ls-g*(s) in SEDFIT.
b) Corresponding sedimentation coefficient distribution derived by the ls-g*(s) model in
SEDFIT. c) Sedimentation coefficient distribution (black, dashed line) and log-normal fit
(red, straight line) for the main fraction of the bare SiO2 nanoparticles. The derived
distribution parameters are µ = 9117 S and σ = 0.09712. A low amount of larger fractions
indicating agglomerates can be observed. d) Particle size distribution of the SiO2 core
particles. The distribution was calculated from the log-normal fit of the sedimentation
coefficient distribution. The density used for calculating the particle diameters is 2.1327
g/cm3 and was derived by relating the mean of the AUC derived distribution with the mean
value of the SEM derived distribution.
41
Figure S3. Comparison of number weighted particle size distributions of bare SiO2
nanoparticles as derived by AUC analysis and SEM analysis. The extinction weighted
sedimentation coefficient distribution was corrected for Mie scattering in HDR-MULTIFIT to
determine the number weighted distribution. Both distributions are found to be very similar.
Slight differences are due to the limiting statistics of the SEM analysis and a small amount of
doublets and triplets found via AUC analysis. As varying density of the SiO2 nanoparticles
due to different porosities would have resulted in a broadening of the sedimentation
coefficient distributions and different widths of the distributions, it can be excluded that
heterogeneity in density had a measureable influence.
Figure S4. a) Cumulative distributions of the Ag/SiO2 mass ratio for different shell thickness
or Ag/SiO2 mass ratios when taking into account the particle size distribution of the SiO2 core
(See Figure S2). The dashed lines represent the ideal cases, when no polydispersity in the core
size would be present. b) Span as calculated for the different Ag/SiO2 mass ratio distributions.
The span is representative for the error in the core-shell correlation and is observed due to the
polydispersity of the SiO2 core particles. It decreases with larger shell thicknesses and
Ag/SiO2 mass ratios. Most importantly, the relevant spans are lower than the ones derived for
the patchy particles (approximately 1.2 – 1.7). Thus, the error due to the polydispersity of the
core size can be neglected in the majority of cases. It should be noted that the purpose of the
analysis shown in this figure was not to simulate the actual Ag/SiO2 seen in our measurements
but rather to illustrate and assess the influence of the core size distribution, independent from
the morphology of the patch and thus the layer thickness distribution.
100 200 300 400
0.000
0.005
0.010
0.015
0.020
0.025
0.030
number density / nm
-1
particle size / nm
SEM
AUC
42
II. Raw data of gravitational sweep (GS) experiments
Figure S5. Raw data derived by the gravitational sweep experiments conducted in the AUC.
Even though multiwavelength data was recorded, only data at a single wavelength (600 nm) is
plotted herein for the sake of clarity. Plots A) – E) show data relating to the flow rate study,
whereas plots F) – J) show data relating to the silver nitrate concentration study.
43
III. Comparison of data analysis methods for gravitational sweep (GS) data
Gravitational sweep data can be analysed in two ways. First, the method introduced by W.
Mächtle can be used,1 which has recently been implement in HDR-MULTIFIT for analysis of
multiwavelength (MWL) data.2 It directly processes the measurement data while correcting
for radial dilution effects. Thus, no fitting is performed and the derived distributions are a
direct reflection of the raw data. As the data acquired by the MWL detector are quite noisy by
default (see Figure S5), smoothing is typically applied which improves the subsequent
processing and analysis of the distributions but decreases the resolution. MWL analysis can
be easily conducted which provides the isolated extinction spectra of the different fractions in
the distribution.
An alternative approach is direct boundary modelling, where a least-squares approach is used
to fit the data to the simplified solution of the Lamm equation (mass transport due to diffusion
is neglected). This approach has been first presented by P. Schuck3 and has been inter alia
applied by J. Walter and co-workers to other cases (such as analytical centrifugation)4. Here
we use the same procedure to analyse the gravitational sweep experiments. Regularization
replaces the smoothing in the direct boundary model and leads to more robust results. In this
study, direct boundary modelling was used for all single wavelength analyses but not applied
for the MWL analyses as it is much more time consuming. It is important to note that both
approaches lead to very similar results for the given data, as can be seen in Figure S6. Thus,
they can be used interchangeably.
1 W. Mächtle, Coupling Particle Size Distribution Technique. A New Ultracentrifuge Technique for Determination
of the Particle Size Distribution of Extremely Broad Distributed Dispersions. Angew. Makromol. Chem. 1988, 162,
35.
2 J. Walter, W. Peukert, Dynamic range multiwavelength particle characterization using analytical
ultracentrifugation. Nanoscale 2016, 8, 7484.
3 P. Schuck, P. Rossmanith, Determination of the sedimentation coefficient distribution by least-squares boundary
modeling. Biopolymers 2000, 54, 328.
4 J. Walter, T. Thajudeen, S. Süß, D. Segets, W. Peukert, New possibilities of accurate particle characterisation by
applying direct boundary models to analytical centrifugation. Nanoscale 2015, 7, 6574.
44
For details about the different procedures it is referred to the original publications where
further information is provided.
Figure S6. Cumulative distributions of the Ag/SiO2 mass ratio for the (a) flow rate and (b)
silver nitrate concentration series. Two different analysis approaches were used. The results
obtained by the standard procedure first proposed by W. Mächtle and later adapted by J.
Walter and co-workers is shown as dashed lines, whereas the results of the direct boundary
analysis is given as solid lines. The results reveal that both approaches lead to very similar
results and can thus be used interchangeably.
45
IV. Comparison of offline UV-VIS spectra and AUC derived spectra
We verified how well the MWL-AUC spectrometer performs compared to “offline”
measurement on a standard benchtop UV-VIS spectrometer. MWL-AUC allows to assess the
spectra of individual fractions in a single sample. For each sample in the silver nitrate
concentration series, five fractions of equal extinction at 600 nm were obtained. To compare
both instruments, the cumulative signal of the MWL-AUC data (sum of all five fractions) was
calculated and plotted together with the spectra measured by the benchtop spectrophotometer.
The results are shown in Figure S7. In general a very good agreement between the two
spectra is found in all cases. Slight differences are mainly observed due to the lower spectral
resolution of the MWL-AUC spectrometer, which is about 2.0 nm for the given setup. Higher
deviations are observed for wavelengths > 600 nm which is due to the low signal-to-noise-
ratio of the MWL-AUC in that range (low intensity output of the xenon flash lamp, see for
example 5). Most importantly, the peak located at 330 to 350 nm, which is required for the
determination of the patch thickness, is well reproduced by the MWL-AUC spectrometer.
5 J. Walter, D. Segets, W. Peukert, Extension of the Deep UV-Capabilities in Multiwavelength Spectrometry in
Analytical Ultracentrifugation: The Role of Oil Deposits. Part. Part. Syst. Charact. 2016, 33, 184.
46
Figure S7. Comparison of UV-VIS spectra derived by MWL-AUC and an offline benchtop
spectrophotometer. Note that the MWL-AUC spectra were smoothed to assist with the
comparison.
47
V. Comparison of GS data and band sedimentation
Data analysis of GS experiments relies on the first derivative of the dilution corrected
cumulative extinction distributions to determine the spectra of the species possessing different
sedimentation coefficients. Previous theoretical and experimental studies have proven that this
works well for nanoparticles.6,7 An alternative approach to the herein conducted GS
experiments constitute band sedimentation experiments, where a particle dispersion is layered
on top of a solution with slightly higher density. During rotation, the particle dispersion is
pushed out of a small sample reservoir, which leads to the formation of a sedimentation band.
In consequence, the particles are not only separated but also isolated based on their different
sedimentation coefficients. From a theoretical point of view, band sedimentation experiments
provide the first derivative of standard sedimentation velocity (SV) or GS experiments.8
For further validation of our GS studies, we measured patchy particles in band sedimentation
mode to correlate the results with the GS experiments. Due to the fast sedimentation of the
patchy particles, we had to conduct the band sedimentation measurements at a fixed detector
position of 7.0 cm. As a consequence of this experimental design, there are several
considerations for band sedimentation experiments which must be made. First, the solution
where the sample is laid upon needs to have a higher density to prevent convection effects and
ensure a stable lamella. For the given case, this can be achieved by using D2O instead of H2O.
However, as we use sucrose to stabilise the particles and to reduce the sedimentation
coefficient, this will also affect the sucrose content as this is adjusted based on a fixed mass
concentration. In consequence, the extinction spectra of the patchy particles will change and
need thus to be compared with GS and off-line measurements conducted in D2O with the
same sucrose content. A direct comparison with existing optical simulations is hence not
possible anymore as these would have to be repeated with the new solvent parameters.
6 J. Walter and W. Peukert, Nanoscale, 2016, 8(14), 7484.
7 S. E. Wawra, L. Pflug, T. Thajudeen, C. Kryschi, M. Stingl and W. Peukert, Nat. Commun., 2018, 9(1), 4898.
8 C.M. Schneider, H. Cölfen, Eur. Biophys. J., 2018, 47, 799.
48
Second, due to high sedimentation coefficients, the experiments had to be conducted at 2500
rpm, which is close to the rotor speed where the overlay of the samples occurs. Moreover, the
band passes the detector within less than 5 minutes. While the exact time of overlay is not
problematic for small particles spun at high rotor speeds,9 it is a considerable source of
uncertainty for these patchy particle studies. Third, due to the dispersion of the sample, the
local concentration is low as is the signal of the individual species. As a consequence, the
particle signal could not be sufficiently distinguished from the background noise. However,
an immediate advantage of band sedimentation experiments are that higher particle
concentrations can be studied because the signal is not limited by the loading concentration as
it would be for SV or GS experiments. In principle, this will grant much better signal-to-noise
ratios and improved spectral information. In our studies, we used 10 µL of sample as an
overlay volume as smaller volumes would have decreased the signal-to-noise ratio further.
The finite width of the lamella reduces resolution further as this results in a slightly undefined
starting point for the sedimentation.
For the given patchy particles, the upper concentration is determined by the synthesis process
and higher concentrations could only be achieved when enriching the particles via
centrifugation. When comparing the retrieved sedimentation coefficient distributions by GS
and band sedimentation at 600 nm shown in Figure S8a, we observed in general a good
agreement between both experiments. However, a small portion of larger particles was
retrieved by the band sedimentation experiment. We believe that these constitute patchy
particle agglomerates which formed during the enrichment process and which could not be
destroyed via ultrasonication. This assumption is supported when extracting the spectra from
the GS and band sedimentation experiments and comparing them to the off-line measured
spectrum in Figure S8b. In principle, both AUC measurements must yield the original
9 C.M. Schneider, D. Haffke, H. Cölfen, Anal. Chem. 2018 90 (18), 10659.
49
spectrum, when integrating over the specific signal of all species. However, we observed a
slight increase in extinction in the VIS-NIR region for the band sedimentation experiment,
which can be explained by the presence of agglomerates, which scatter light - due to their
larger size - more predominantly. Most importantly, the peak located at 330 to 350 nm is not
affected by those agglomerates and further matches for all three measurements.
Figure S8. a) Cumulative extinction-weighted sedimentation coefficient distributions
retrieved at 600 nm. b) Integral extinction spectra of all species in the samples as derived by
MWL-AUC for patchy particles using GS and band sedimentation as well as an off-line
measurement using a spectrophotometer.
In summary, we are confident to conclude that GS and band sedimentation provide equivalent
data, with the limitations for band sedimentation described above. For the purpose of this
study, GS is well-suited to retrieve the sedimentation coefficient distributions as well as the
position of the characteristic UV peak. For analysis of the full extinction spectra, band
sedimentation experiments constitute a promising alternative as higher concentrations can be
used. However, formation of agglomerates would need to be prevented, preferably by
synthesis of higher concentrations right away.
50
VI. Characterisation of mixing via iodate-iodide reaction
The flow rate dependent mixing conditions and corresponding values of characteristic mixing
time tm within the experimental setup were determined by the iodate-iodide reaction.
Experiments were carried out in the flow rate range of 1 mL/min to 20 mL/min for each
educts stream following the Villermaux-Dushmann protocol, as established by J. M.
Commenge and L. Falk.10 The concentration set used is shown in Table S2.
VII. Image analysis to determine patch thickness distribution from SEM images
The following workflow was used to obtain patch thickness distributions with imageJ
(obtained from https://imagej.nih.gov/ij/):
1. Raw image opened in imageJ (Figure S15a)
2. Image scale acquired from metadata
3. Image processing to optimize image for thresholding (Figure S15b)
a. Subtract background with sliding paraboloid and 20 pixel rolling ball radius
b. Enhance contrast function with 0.3% saturated pixels
c. Median filter with radius 0.5 pixel radius
4. Threshold image and apply watershed algorithm (Figure S15c)
5. Select patchy particles with negligible or no core particle exposed (Figure S15d)
6. Measure selected particle’s areas and determine the circle equivalent area diameter, xc
7. Concatenate datasets from multiple images of the same sample11 and determine mass
weighted cumulative size distribution Q3(xc) (Figure S16)
8. To determine cumulative mass weighted patch thickness distributions, Q3(tpatch,SEM), the
core particle and patchy particle size distributions in Figure S16 were fitted with
10 J.-M. Commenge and L. Falk, VillermauxDushman protocol for experimental characterization of micromixers,
Chemical Engineering and Processing: Process Intensification, 2011, 50, 979.
11 Per sample, multiple images were processed so that at least 120 suitable patch-coated particles were used in the
analysis
51
Boltzmann functions. For each value of Q3(xc) the corresponding size from the core
particle distribution was subtracted from the value for the patchy particle. To obtain
patch thickness the result was halved.
52
VIII. FEM based simulations of optical extinction spectra of silver patchy particles
The time-harmonic formulation of Maxwell’s equation describes the interaction of harmonic
incident light and matter. To simulate the extinction cross-section spectra of non-spherical
multi-material particles we developed our own FEM time-harmonic-Maxwell solver12 based
on the work of P. Monk13 and J. Schöberl et al.14. The computational domain is spatially
discretized into tetrahedral elements using Tetgen15. Based on this mesh the partial-
differential equation is discretized using Nedelec basis functions where the mandatory
absorbing boundary condition on the artificial outer boundary of the computational domain is
modelled by a perfectly matched layer (PML)16. The resulting complex valued linear system
of equations is solved by the highly parallel direct solver HSL_MA86 HSL17. Since
Extinction spectra are in most measurements an average over all particle orientations, this is
taken into account by using a quadrature rule to numerically integrate over all possible
orientations and polarization states. By using the rotational-symmetry of the particle as well
as the point-symmetry of the extinction cross-section with respect to the incident direction,
the surface integral can be reduced to a line integral over the quarter-circle. This is done for
64 incident directions with 2 polarizations each. The refractive index of bulk silver is taken
from P. B. Johnson and R. W. Christy18 while the correction of the refractive index of due to
damping in reduced dimension nanostructures is used according to S. Kawata et al.19
12 A. Pratelli, G. Leugering. New trends in shape optimization, Birkhäuser, Swizerland 2015.
13 P. Monk. Finite Element Methods for Maxwell’s Equations, Clarendon Press, Oxford, England 2003.
14 J. Schöberl, S. Zaglmayr, High order Nédélec elements with local complete sequence properties, COMPEL 24
(2005) 374-384.
15 H. Si, TetGen, a Delaunay-Based Quality Tetrahedral Mesh Generator, ACM Trans. Math. Software 41 (2015)
11.
16 J. Berenger, A perfectly matched layer for the absorption of electromagnetic waves, J. Comput. Phys. 114 (1994)
184-200
17 Science & Technology Facilities Council, HSL, A collection of Fortran codes for large scale scientific
computation, http://www.hsl.rl.ac.uk/, accessed: June, 2016
18 P. B. Johnson, R. W. Christy, Optical Constants of the Noble Metals, Phys. Rev. B 6 (1972) 4370-4379
19 S. Kawata, M. Ohtsu, M. Irie, Near-Field Optics and Surface Plasmon Polaritons, Springer Science & Business
Media, Berlin Heidelberg, Germany 2003
53
IX. Supplementary tables
Table S1. Conditions for silver patchy particles flow synthesis.
Syringe 1
Syringe 2
Study
AgNO
3
M]
SiO
2
[mg/L]
CH
2
O
[mM]
NH
4
OH
[mM]
Flow rate per syringe
[mL/min]
Flow rate
400
25
152
27
2,5,10,15,20
AgNO3 concentration
200, 300, 400, 500, 600
25
152
27
20
Table S2. Concentrations used for iodate-iodide reaction to determine characteristic mixing
time.
Syringe
Substance
Concentration [
M
]
1
H+ / H2SO4
0.06 / 0.03
2
NaI
0.032
KIO3
0.006
NaOH
0.09
H3BO3
0.09
54
Table S3. Summary of results of the AUC analysis of Ag/SiO2 mass ratio and MWL-AUC
analysis of patch thickness and coverage for samples prepared at different silver nitrate
concentrations. ,, was obtained from λUV-peak according to the calibration function
determined in Figure S17. λUV-peak corresponds to the peak at between 320 nm and 380 nm
determined by fitting the MWL-AUC spectral data corresponding to the fraction of the
cumulative mass ratio distribution defined as having between 20% and 80% of the cumulative
extinction at 600 nm (see Figure S13). ,,, (median of the mass weighted patch
thickness distribution) was obtained by SEM image analysis (see Section VIII and Figures
S15 and 16) and is shown for comparison. cpatch,50 was obtained by Equation 15.
[AgNO
3
] / µ
M
50,AUC
/
%
λ
UV-peak
/ nm
,,,
/ nm
,,
/ nm
,,
/ %
200
149
334.4
13.4
15.9
71.8
300
214
335.8
17.6
17.4
93.0
400
251
339.4
21.2
21.2
86.7
500
295
342.4
24
24.5
86.6
600
405
348.6
31.2
31.1
89.1
Table S4. Summary of results of MWL-AUC analysis regarding individual fractions of the
cumulative Ag/SiO2 mass ratio. The parameters of the fractions were determined similarly to
those in Table S3.
[AgNO3]
/ µM
Extinction (600 nm)
cumulative fraction of
AUC-determined
distribution
50,
FRACTION
,AUC
/
%
λUV-peak,fraction /
nm ,,,
/ nm ,.,
/ %
200
0 % - 20 %
67
331.9
13.2
39.3
20 % - 40 %
113
332.2
13.6
64.7
40 % - 60 %
149
335.3
16.9
67.2
60 % - 80 %
191
335.3
16.9
86.0
80 % - 100 %
262
334.5
16.0
124.8
300
0 % - 20 %
88
333.0
14.4
47.3
20 % - 40 %
166
335.5
17.1
73.5
40 % - 60 %
214
334.7
16.2
100.6
60 % - 80 %
266
338.1
19.9
99.3
80 % - 100 %
347
338.5
20.3
126.5
400
0 % - 20 %
107
334.0
15.5
53.0
20 % - 40 %
190
337.2
18.9
75.1
40 % - 60 %
251
339.0
20.8
88.7
60 % - 80 %
317
344.0
26.2
85.9
80 % - 100 %
427
339.5
21.4
146.7
500
0 % - 20 %
135
335.6
17.2
59.3
20 % - 40 %
229
338.1
19.9
85.6
40 % - 60 %
295
345.6
27.9
74.2
60 % - 80 %
370
345.5
27.8
93.4
80 % - 100 %
498
345.5
27.8
125.8
600
0 % - 20 %
145
338.0
19.8
54.6
20 % - 40 %
295
345.6
27.9
74.0
40 % - 60 %
405
348.2
30.6
90.7
60 % - 80 %
522
351.5
34.2
102.1
80 % - 100 %
742
351.3
33.9
146.3
55
X. Further supplementary figures
Figure S9. a)-e) Low magnification SEM images of samples synthesised with flow rates in
the range 2 mL/min to 20 mL/min. f) AUC-determined cumulative distributions of
sedimentation coefficient of samples corresponding to a)-e) as well as to the bare SiO2
nanoparticles.
10000 100000
0.0
0.2
0.4
0.6
0.8
1.0
ext. cumulative (600 nm) / -
sedimentation coefficient / S
20 mL·min-1
15 mL·min-1
10 mL·min-1
5 mL·min-1
2 mL·min-1
SiO2
f)
56
Figure S10. a)-e) SEM images of samples synthesised at silver nitrate concentrations from
200 µM to 600 µM. f) Cumulative distributions of sedimentation coefficient s of samples
corresponding to a)-e) as well as to the bare SiO2 nanoparticles.
10000 100000
0.0
0.2
0.4
0.6
0.8
1.0
f)
200 µM
300 µM
400 µM
500 µM
600 µM
SiO2
ext. cumulative (600 nm) / -
sedimentation coefficient / S
57
Figure S11. FEM simulations of silver patchy particles extinction spectra for a) constant
patch thickness of 20 nm and varying patch coverage and b) constant patch coverage of 70%
and varying patch thickness.
58
Figure S12. AUC in situ extinction spectra of sample synthesised with 200 µM - 600 µM
silver nitrate concentration. Each spectrum represents particles which contribute 20 % of the
total extinction at 600 nm in the cumulative distribution of the Ag/SiO2 mass ratio. The
spectra shown are smoothed versions of the raw data.
59
Figure S13. Example of the fitting of raw spectral data with an asymmetric double sigmoidal
function in the wavelength range 320 to 380 nm. The spectra correspond to the central 20%-
80% fraction of the cumulative extinction at 600 nm. The positions of the fitted peaks are
listed in Table S2 and are correlated to SEM-determined patch thickness in Figure S17.
60
Figure S14. a) FEM simulation of the UV peak of idealised patchy particles for different
patch thicknesses and coverages. b) Dependency of the UV-peak position on patch thickness
derived from the data in a). Error bars indicate the effect of patch coverage on UV-peak
position at given patch thicknesses. The data is fitted by a Boltzmann growth function and led
to the calibration function to convert the UV-peak position into patch thickness:
, =ln
.

.
16.5974+19.6625
.
61
Figure S15. Workflow for image analysis (see Section VIII).
Figure S16. (a) Cumulative mass weighted particle size distributions for core particles and
visually fully coated patchy particles determined by analysis of SEM images. (b) Cumulative
mass weighted patch thickness distributions derived from the data in (a).
62
Figure S17. Plot showing the calibration of the patch thickness determined from the UV-peak
position. The squares correspond to the SEM image analysis determined patch thickness
plotted against the corresponding UV-peak positions for the 5 samples produced by varying
silver nitrate concentration. The UV-peak positions were determined by fitting the MWL-
AUC spectral data corresponding to the fraction of the cumulative mass ratio distribution
defined as having between 20% and 80% of the cumulative extinction at 600 nm (see Figure
S13). The data is fit with the linear function: , = 1.067  434.9.  (the
likely spurious lowest UV-peak position datapoint was ignored in the fitting). The crosses
show, for comparison, the relationship between UV-peak position and patch thickness
determined by FEM simulations (see Section IX and Figures S11 and S14)
... Thus, attempts to control the amount of growth in order to obtain silver patches with targeted size invariably lead to an unwanted influence on the nucleation conditions, resulting in undesirable numbers and sizes of patches on the silica cores, or even a large fraction of bare cores. 47 Through the use of a cascaded two-T mixer process and making use of the kinetic control afforded to the reaction by the ammonia concentration, we were able to partially separate nucleation from growth, but not sufficiently for control of patch size. 46 Such control, however, is essential in order to design arbitrary patchy particle morphologies for application as anisotropic building blocks, 48 or, as already noted above, to achieve a wide tunability of the LSPR. ...
... This peak was observed in our earlier studies with silver patches and corresponds to the out-of-plane quadrupole resonance, similar to that seen for flat silver nanoprisms. 41,55 Since this feature depends almost exclusively on the patch thickness, 47 we can conclude that the value of ρ 0, Seed has rather little influence on that aspect of the patch morphology. It follows that for a lower seeding density, the coverage of patches should increase. ...
... MWL detection also proved useful for the characterization of the smallest silver clusters between 1 and 5 atoms [65], carbon nanotube-intercalant-surfactant complexes [66], gold nanorods, different species of single-wall carbon nanotubes and CdTe nanoparticles [67], yield, mass, and structure of silver patches on colloidal silica [68], as well as West Nile Virus RNA-protein interactions where the protein and nucleic acid signals could be separated by spectral decomposition such that sedimentation information from each individual species, including any complexes, could be dearly identified based on their spectral signatures [69]. The UltraScan as well as the SEDANAL and HDR-MULTIFIT software packages are the currently available evaluation programs, which can directly work with multiwavelength datasets [38,70,71]. ...
Article
Full-text available
Analytical ultracentrifugation (AUC) is a classical polymer and colloid analysis technique invented by Theodor Svedberg 100 years ago. Modern hard- and software and powerful computers make it now possible to develop the methodology beyond what was possible with this technique before. This perspective aims to describe new possibilities, which extend the possibilities of AUC beyond the classical repertoire of the determination of distributions of sedimentation coefficient, particle size, and molar mass as well as stoichiometries and interaction constants of interacting systems. High-resolution simultaneous characterization of particle size and optical property distributions, investigation of nucleation by reaction in the AUC cell, characterization of particle interactions at a very high concentration, and characterization of complex fluids or osmotic pressures over large concentration ranges even crossing phase boundaries are among the discussed topics. They show that even after 100 years of successful application, AUC still has much yet unexplored potential in colloid and polymer science. Graphical Abstract This perspective paper spans from the days of invention of analytical ultracentrifugation to now including nonmainstream methodology and instrumentation, which has a huge potential for the future. This includes multiwavelength detectors, high-resolution particle size distributions, chemical reactions in the ultracentrifuge, high-concentration work, osmotic pressure distributions, and characterization of complex fluids.
... The modeling and efficient numerical approximations of multi-dimensional (MD) nanoparticle (NP) synthesis are increasingly required as the properties of anisotropic particles play a major role in various applications, including bio-nanosensors or catalysts [2,3,4]. In recent years, automated analysis of MD particle shape distributions -essential to validate and calibrate MD process models -has been demonstrated for a range of composite and anisotropic NPs [5,6,7], unlocking the possibility of using predictive modeling for MD-NP synthesis. ...
Preprint
Full-text available
The unique properties of anisotropic and composite particles are increasingly being leveraged in modern particulate products. However, tailored synthesis of particles characterized by multi-dimensional dispersed properties remains in its infancy and few mathematical models for their synthesis exist. Here, we present a novel, accurate and highly efficient numerical approach to solve a multi-dimensional population balance equation, based on the idea of the exact method of moments for nucleation and growth \cite{pflug2020emom}. The transformation of the multi-dimensional population balance equation into a set of one-dimensional integro-differential equations allows us to exploit accurate and extremely efficient numerical schemes that markedly outperform classical methods (such as finite volume type methods) which is outlined by convergence tests. Our approach not only provides information about complete particle size distribution over time, but also offers insights into particle structure. The presented scheme and its performance is exmplified based on coprecipitation of nanoparticles. For this process, a generic growth law is derived and parameter studies as well as convergence series are performed.
... This could explain the blue-shifts in the UV plasmon resonance which were observed for the cases of ammonia ageing (Fig. 3e) and nitric acid addition (Fig. S4e) and not for NH 4 NO 3 addition (Fig. 4e). This aspect requires further study in the future, for instance making use of multidimensional characterization techniques like multi-wavelength analytical ultracentrifugation which can reveal trends in the yield, thickness and coverage (Meincke et al., 2022). The objective of this work was to identify and mitigate for the cause of the sporadic appearance of spiky nanostructures in our patchy particle product. ...
Article
Patchy particles, which comprise a core particle of one material partially covered by a thin layer of another, offer great promise for tailored particle-based materials and devices in the future. This study concerns the investigation and improvement of a potentially gram-scalable continuous flow process based on a cascade of two T-mixers for the synthesis of thin, conformal silver patches on colloidal silica particles. The starting point of the work is the observation that on seemingly random occasions the process results in patches having undesired spiky protrusions. Through a systematic study of reagent ageing supported by scanning electron microscopy, optical spectroscopy and chemical equilibrium calculations it can be ascertained that the desired product morphology is only obtained through carbon dioxide absorption by the dilute ammonia solution used in the reaction. It is shown that the ageing process can be emulated for fresh reagents by addition of ammonium nitrate or nitric acid.
... nitric acid addition (Fig. S3e) and not for NH4NO3 addition (Fig. 4e). This aspect requires further study in the future, for instance making use of multidimensional characterization techniques like multi-wavelength analytical ultracentrifugation which can reveal trends in the yield, thickness and coverage (Meincke et al., 2021). ...
Preprint
Full-text available
Patchy particles, which comprise a core particle of one material partially covered by a thin layer of another, offer great promise for tailored particle-based materials and devices in the future. This study concerns the investigation and improvement of a potentially gram-scalable continuous flow process based on a cascade of two T-mixers for the synthesis of thin, conformal silver patches on colloidal silica particles. The starting point of the work is the observation that on seemingly random occasions the process results in patches having undesired spiky protrusions. Through a systematic study of reagent ageing supported by scanning electron microscopy, optical spectroscopy and chemical equilibrium calculations it can be ascertained that the desired product morphology is only obtained through carbon dioxide absorption by the dilute ammonia solution used in the reaction. It is shown that the ageing process can be emulated for fresh reagents by addition of ammonium nitrate or nitric acid.
Article
Full-text available
We describe the development of a T-mixer based continuous flow process for the coating of 86–500 nm diameter spherical polystyrene particles with thin gold patches by heterogeneous nucleation and growth. After establishing a suitable flow rate for good mixing and sufficiently uniform product morphology we systematically investigate the main reaction parameters. This reveals a considerable tunability of the patch morphology and, by virtue of the localized surface plasmon resonance of gold, the optical properties of the product dispersions. In order to further widen the range of nanostructures accessible by our process, a second T-mixer was added. This introduced new gold precursor, leading to further growth of the patches that were formed after the first mixer. By this approach, nearly-complete gold nanoshells could be produced in high yield on both small and large core particles, without the unwanted production of free-standing gold nanoparticles. Due to the pronounced optical properties of nearly-complete gold nanoshells on small core particles, we could estimate from electrodynamic simulations the equivalent shell thickness to be as low as 8.6 nm. This is significantly thinner than can be routinely achieved using the standard seeded growth approach to synthesise gold nanoshells. Our results are therefore highly promising for the gram-scale synthesis of plasmon resonant nanostructures with designed optical properties.
Article
Full-text available
Hypothesis Self-assembly using anisotropic colloidal building blocks may lead to superstructures similar to those found in molecular systems yet can have unique optical, electronic, and structural properties. To widen the spectrum of achievable superstructures and related properties, significant effort was devoted to the synthesis of new types of colloidal particles. Despite these efforts, the preparation of anisotropic colloids carrying chemically orthogonal anchor groups on distinct surface patches remains an elusive challenge. Experiments We report a simple yet effective method for synthesizing patchy particles via seed-mediated heterogeneous nucleation. Key to this procedure is the use of 3-(trimethoxysilyl)propyl methacrylate (TPM) or 3-(trimethoxysilyl)propyl acrylate (TMSPA), which can form patches on a variety of functional polymer seeds via a nucleation and growth mechanism. Findings A family of anisotropic colloids with tunable numbers of patches and patch arrangements were prepared. By continuously feeding TPM or TMSPA the geometry of the colloids could be adjusted accurately. Furthermore, the patches could be reshaped by selectively polymerizing and/or solvating the individual colloidal compartments. Relying on the chemically distinct properties of the TPM/TMSPA and seed-derived domains, both types of patches could be functionalized independently. Combining detailed control over the patch chemistry and geometry opens new avenues for colloidal self-assembly.
Article
Full-text available
Properties of nanoparticles are influenced by various parameters like size, shape or composition. Comprehensive high throughput characterization techniques are urgently needed to improve synthesis, scale up to production and make way for new applications of multidimensional particulate systems. In this study, we present a method for measuring two-dimensional size distributions of plasmonic nanorods in a single experiment. Analytical ultracentrifuge equipped with a multiwavelength extinction detector is used to record the optical and sedimentation properties of gold nanorods simultaneously. A combination of sedimentation and extinction properties, both depending on diameter and length of the dispersed nanorods, is used to measure two-dimensional distributions of gold nanorod samples. The length, diameter, aspect ratio, volume, surface and cross-sectional distributions can be readily obtained from these results. As the technique can be extended to other non-spherical plasmonic particles and can be used for determining relative amounts of particles of different shapes it provides complete and quantitative insights into particulate systems.
Article
Full-text available
Dielectric-metal core-shell particles with morphologically tunable optical properties are highly promising candidates for applications ranging from theranostics, energy harvesting and storage to pigments and sensors. Most structures of interest have, until now, been produced in small volume wet chemical batch approaches which are difficult to scale. In extension to the growing interest in continuous flow process for the scalable synthesis of single phase nanomaterials, here we describe the formation of two-phase core-shell particles. Specifically, the coating of silver onto colloidal silica particles using a static T-mixer approach is presented. The coating is achieved in a single reaction step in which careful control of educt concentration and process conditions leads to nucleation and surface conformal growth of silver. Through two generations of reactor, good coating yield and process reproducibility are demonstrated. In particular, the ability to control the characteristic time of mixing, as verified by the Villermaux-Dushmann reaction, was shown to be crucial in narrowing the range of morphologies obtained in the product particles. Moreover we could tune, through process parameters, the degree of coating of the core particles, leading to patchy particles, Janus particles and even complete nanoshells. Since silver nanostructures have a strong structure-dependent plasmon resonance, we could verify the coating morphologies via optical spectroscopy and electrodynamic simulations. Optical measurements were also used in order to optimize, via trial and error, the continuous flow process parameters to produce a high coating yield. This opens up future potential for feedback-driven optimization of particle coating. Finally, we demonstrate that the silver coatings can be morphologically stabilized, opening up the possibility of effective formulation for applications.
Article
Colloidal silica particles have received a widespread interest because of their potential applications in adsorption, ceramics, catalysis, drug delivery and more. Among many approaches towards fabrication of these colloidal particles, Stöber, Fink and Bohn (SFB) method, known as Stöber synthesis is an effective sol–gel strategy for production of uniform, monodispersed silica particles with highly tailorable size and surface properties. This review, after a brief introduction showing the importance of colloidal chemistry, is focused on the Stöber synthesis of silica spheres including discussion of the key factors affecting their particle size, porosity and surface properties. Next, further developments of this method are presented toward fabrication of polymer, carbon, and composite spheres.
Article
Application of dispersants is one of the widely used oil spill response strategies. With the growing concern for particulate emulsifiers, applying amphiphilic Janus particles to emulsify oil in seawater has been examined in this paper. Amphiphilic Janus particles with a combination of colloidal particles and amphiphilic characteristic are found to show a highly effective emulsification for various oils in seawater and overcome several drawbacks of conventional chemical dispersants. We report the synthesis of amphiphilic Janus silica particles (JSPs) as a novel oil spill dispersant and confirm their amphiphilicity by various techniques. The oil-in-seawater emulsion stabilized by JSPs presents small and stable oil droplets in one month, even at a higher volume ratio of oil to seawater and a wide temperature range. It is believed that the small size and high stability of droplets are very important to oil dispersion and eventual biodegradation. Through confocal laser scanning microscopy and scanning electron microscope observations, JSPs were found to anchor tightly at droplet surface and form a dense interfacial particle film to prevent droplets coalescence. The appropriate wettability of JSPs and the reduced interfacial tension contribute to the high interfacial adsorption energy of JSPs. Additionally, JSPs show weak inhibitory effect on oil-degrading bacteria, which is confirmed by bacterial growth curve, resazurin test and oil biodegradation experiment. This study reveals that amphiphilic Janus silica particles hold an impressive capability for environmentally friendly treatment of oil spills.
Article
This work highlights significant advancements in detector hardware and software for multiwavelength analytical ultracentrifugation (MWL-AUC) experiments, demonstrating improvement in both, the spectral performance and UV capabilities of the instrument. The hardware is an extension of the Open AUC MWL detector developed in academia and first introduced in 2006 by Bhattacharya et al. Additional modifications as well as new analytical methods available for MWL data have since been reported. The present work describes new and continuing improvements to the MWL detector, including mirror source and imaging optics, UV sensitive acquisition modes and revised data acquisition software. The marked improvement of experimental data promises to provide access to increasingly complex systems, especially semiconductor nanoparticles, synthetic polymers, biopolymers and other chromophores absorbing in the UV. Details of the detection system and components are examined to reveal the influences on data quality, and to guide further developments. The benchmark comparisons of data quality across platforms will also serve as a reference guide for evaluation of forthcoming commercial absorbance optics.
Article
Janus particles have two distinct surfaces or compartments. This enables novel applications that are impossible with homogeneous particles, ranging from the engineering of active colloidal metastructures to creating multimodal therapeutic materials. Recent years have witnessed a rapid development of novel Janus structures and exploration of their applications, particularly in the biomedical arena. It, therefore, becomes crucial to understand how Janus particles with surface or structural anisotropy might interact with biological systems and how such interactions may be exploited to manipulate biological responses. This perspective highlights recent studies that have employed Janus particles as novel toolsets to manipulate, measure, and understand cellular functions. Janus particles have been shown to have biological interactions different from uniform particles. Their surface anisotropy has been used to control the cell entry of synthetic particles, to spatially organize stimuli for the activation of immune cells, and to enable direct visualization and measurement of rotational dynamics of particles in living systems. The work included in this perspective showcases the significance of understanding the biological interactions of Janus particles and the tremendous potential of harnessing such interactions to advance the development of Janus structure-based biomaterials.
Article
Janus and patchy particles are emerging as models for studying complex directed assembly patterns and as precursors of new structured materials and composites. Here we show how lipid induced capillary bridging could serve as a new and non conventional method of assembling patchy particles into ordered structures. Iron oxide surface patches on latex microspheres were selectively wetted with liquid lipids, driving the particle assembly into 2D and 3D clusters via interparticle capillary bridge formation. The liquid phase of the bridges allows local reorganization of the particles within the clusters and assists in forming true equilibrium configurations. The temperature driven fluid-to-gel and gel-to-fluid phase transitions of the fatty acids within the bridge act as a thermal switch for cluster assembly and disassembly. By complementing the experiments with Monte-Carlo simulations we show that the equilibrium cluster morphology is determined by the patch characteristics, namely; their size, number, and shape. The study demonstrates the ability of capillary bridging as a versatile tool to assemble thermoresponsive clusters and aggregates. This method of binding particles is simple, robust and generic, and can be extended further to assemble particles with non-spherical shapes and complex surface chemistries enabling the making of sophisticated colloidal molecules.
Article
Silica@Au core-shell type composite particles are promising for biomedical applications due to their unique optical and chemical characteristics, although the problem lies in the synthetic processes. They are typically synthesized via batch type reactions including multiple steps, and the complexity in the synthetic process is an obstacle to practical applications. A robust and facile process is thus required. In the present study, we developed a flow synthetic process by applying a microreactor with a high mixing performance to silica@Au nanoshells and patchy particles. Following a concept of selective nucleation and subsequent growth of the gold on the core surface, we mixed a suspension of surface-modified silica particles containing gold ions with a reducing agent solution in the microreactor, and investigated the effects of the core particle size and the type of reducing agents on the composite particle synthesis. Through the investigation, we found that the ratio of the gold ion amount to the surface area of core particles is a key factor for obtaining uniform core-shell particles. In addition, the morphology of gold formed on the core surface was demonstrated to change from spherical nanoparticles to dendritic patches depending on the reducing ability of reducing agents.
Article
Chemical methods developed over the past two decades enable preparation of colloidal nanocrystals with uniform size and shape. These Brownian objects readily order into superlattices. Recently, the range of accessible inorganic cores and tunable surface chemistries dramatically increased, expanding the set of nanocrystal arrangements experimentally attainable. In this review, we discuss efforts to create next-generation materials via bottom-up organization of nanocrystals with preprogrammed functionality and self-assembly instructions. This process is often driven by both interparticle interactions and the influence of the assembly environment. The introduction provides the reader with a practical overview of nanocrystal synthesis, self-assembly, and superlattice characterization. We then summarize the theory of nanocrystal interactions and examine fundamental principles governing nanocrystal self-assembly from hard and soft particle perspectives borrowed from the comparatively established fields of micrometer colloids and block copolymer assembly. We outline the extensive catalog of superlattices prepared to date using hydrocarbon-capped nanocrystals with spherical, polyhedral, rod, plate, and branched inorganic core shapes, as well as those obtained by mixing combinations thereof. We also provide an overview of structural defects in nanocrystal superlattices. We then explore the unique possibilities offered by leveraging nontraditional surface chemistries and assembly environments to control superlattice structure and produce nonbulk assemblies. We end with a discussion of the unique optical, magnetic, electronic, and catalytic properties of ordered nanocrystal superlattices, and the coming advances required to make use of this new class of solids.