Measurement: G. Derkachov, D. Jakubczyk, K. Kolwas, K. Piekarski, Y. Shopa, M. Woz Niak

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Measurement 158 (2020) 107681

Contents lists available at ScienceDirect

Measurement
journal homepage: www.elsevier.com/locate/measurement

Dynamic light scattering investigation of single levitated


micrometre-sized droplets containing spherical nanoparticles
G. Derkachov a, D. Jakubczyk a, K. Kolwas a, K. Piekarski b, Y. Shopa c,⇑, M. Woźniak a
a
Institute of Physics, Polish Academy of Sciences, al. Lotników 32/46, 02-668 Warsaw, Poland
b
Białystok University of Technology, Faculty of Computer Science, Wiejska 45A, 15-351 Białystok, Poland
c
Cardinal Stefan Wyszyński University in Warsaw. Faculty of Mathematics and Natural Sciences, Wóycickiego 1/3, 01-938 Warsaw, Poland

a r t i c l e i n f o a b s t r a c t

Article history: We applied the dynamic light scattering (DLS) technique for studying single microdroplets of dispersions
Received 30 March 2019 of SiO2 and TiO2 nanoparticles. The evaporating microdroplets were levitating in an electrodynamic
Received in revised form 10 February 2020 quadrupole trap. Exponentially decaying autocorrelation functions (ACFs) were obtained for the light
Accepted 24 February 2020
scattered by evaporating microdroplets as well as for microaggregates of nanoparticles formed at the
Available online 4 March 2020
end of the evaporation process. It was found that the temporal variation of the ACF generally reflects
the evolution of the microdroplet of dispersion. At the initial stage of the evaporation, apart from the
Keywords:
optical morphology resonances of the droplet (dielectric sphere), it was possible to identify the charac-
Dynamic light scattering
Autocorrelation function
teristic times corresponding to the Brownian motion of the dispersed nanoparticles. At the end of evolu-
Nanoparticle tion, when the drying microdroplet transforms into a (non-rigid) microaggregate, the Brownian motion of
Microdroplet the dispersed nanoparticles was masked by the rotational Brownian motion of the microaggregate as a
Electrodynamic trap whole.
Ó 2020 Elsevier Ltd. All rights reserved.

1. Introduction properties have utmost importance (see e.g. [6] and references
therein) and require in-depth studies.
Nowadays, a wide variety of micro- and nanoparticles is pro- In order to effectively investigate evaporating composite dro-
duced for applications in technology and biosciences (see e.g.: plets it is necessary to use a technique that provides non-
[1–4] for reviews). The diversity concerns shape, size, material invasive and non-contact means of droplet characterisation. There
and its distribution. Dispersions of such particles gain additional are rather few methods that might enable characterisation of the
functionality at the expense of their complexity. Furthermore, dis- internal or at least the surface structure of a composite droplet
persions are commonly encountered as droplets and microdroplets (compare e.g.: [7]). For instance, static light scattering-based meth-
in particular. Such shape of the gas-liquid interface additionally ods were investigated: white light scattering by Matic et al. [8],
imposes spherical symmetry onto the system modifying its proper- rainbow refractometry by Saengkaew et al. [9] and Li et al. [10]
ties. Therefore, efficient characterisation methods are pursued, and digital in-line holography by Coëtmellec et al. [11]. The appli-
both for bulk materials and droplets. On the other hand, droplets cability of dynamic light scattering (DLS) was investigated e.g. by
of complex suspensions are formed in the atmosphere via the pro- Bronk et al. [12] and Krieger et al. [13]. However, the conclusion
cess of so-called ‘‘scavenging” [5] – water droplets of clouds (also of [13] was rather discouraging. Since our preliminary studies
often condensing on aerosols), fog and rain absorb gases and vari- [14,15] suggested otherwise, in this work we’ve revisited the issue.
ous atmospheric aerosols, both soluble and insoluble. The complex For bulk fluid dispersions, DLS method is predominantly used
droplets of suspension/mixture undergo further evolution, in par- for measuring particle size and size distribution, studying relax-
ticular, evaporation, which leads to self-organisation of inclusions. ation and aggregation processes, indirect viscosity measurement
Inclusions such as micrograins of sand or spores can aggregate into etc. (see e.g. [16,17] for review). The DLS technique is based on
fairly regular structures. Since aerosols have a significant impact the analysis of the intensity fluctuations of the scattered light,
upon (radiative) energy transfer in planetary atmospheres (see caused by Brownian motions of the dispersed particles. It requires
e.g. radiative forcing but also visibility impairment), their optical measuring the autocorrelation function (ACF) of the fluctuations
[18,19]. A standard procedure of the ACF analysis for polydisperse
⇑ Corresponding author. samples is the application of the method of cumulants [18,20].
E-mail address: i.shopa@uksw.edu.pl (Y. Shopa). However, it must be kept in mind that the results obtained with

https://doi.org/10.1016/j.measurement.2020.107681
0263-2241/Ó 2020 Elsevier Ltd. All rights reserved.
2 G. Derkachov et al. / Measurement 158 (2020) 107681

the DLS technique depend to a large extent not only on the data nanoparticles – fine grains of sand, while the other represents
quality but also on the data processing method. The reliability of nanoparticles in technological application. Water, diethylene glycol
the results depends on the completeness of the applied model of (DEG) and tetraethylene glycol (TetraEG) were used as the disper-
the observed processes. sion media. The droplets, with initial diameters in the 20  30 lm
In this paper we adapted the DLS technique, usually used for range, were introduced into the electrodynamic trap by means of
macroscopic samples in cuvettes, to study single levitated a droplet-on-demand injector [24,25]. The trap was placed in a
micrometre-sized droplets of liquids containing spherical nanopar- small chamber with dry nitrogen atmosphere at room temperature
ticles. Bronk et al. [12] reported the application of ACF technique to ( 22  C) and atmospheric pressure. The dry nitrogen is obtained
characterisation of the levitating droplets of liquid/dispersion from LN boiling freely in a vessel. On the way to the chamber, it
under similar experimental conditions. However, the results were passes through the heat exchanger in the thermostatic bath in order
obtained for non-evaporating droplets, only for one order of mag- to equalise the gas temperature with the temperature of the cham-
nitude in the ACF time scale. Krieger et al. extended the idea pro- ber. The chamber is flushed with dry nitrogen before each droplet
posed by Bronk et al., though they concluded that ‘‘diffusional injection and the dry nitrogen flow is stopped only for the duration
motion of an inclusion within the aqueous solution of a host dro- of the droplet measurement. The droplet carried an electric charge
plet is masked by rotational Brownian motion”. In our previous (gained by charge separation in the electric field on passing through
research [14,15], we found that we are probably capable of finer the injector nozzle) which enabled stabilisation of its position using
measurements. We found some features of the DLS signals that a combination of the static (DC) and oscillating (AC) electric fields.
could be connected with the properties of a microdroplet of disper- There are several mature droplet levitation techniques (see e.g. [26]
sion. In this work we further follow this track and carefully study for a review). We’ve chosen electrodynamic levitation rather than
the relationship of the measured ACF to the changes in the size optical or acoustic (compare [7]), since it enables trapping of
and composition of the evaporating droplet. (sub)microdroplets and causes smallest droplet deformation
[27,28]. The AC field was in a quadrupolar arrangement, so in the
2. Basic relations and experimental setup very centre of the trap, the AC field strength (ideally) went to zero.
In consequence, the micromotion (with the frequency of the AC
2.1. Autocorrelation function in DLS method field) of a particle kept in the centre of the trap is minimal. Further-
more, a particle in the ground state of the trap performs no secular
The random motion of particles in dispersion with the temper- motion (macromotion). The DC field is used to balance the particle
ature T and the viscosity g is characterised by the diffusion coeffi- weight and actively stabilise the particle in the centre in a vertical
cient D, which by means of the well-known Stokes-Einstein direction. The levitated droplet was illuminated with a collimated
relation D ¼ kB T=ð6pgRH Þ is related to the so-called hydrodynamic vertically polarised blue Arþ laser beam (497 nm, 50 mW) and/or
radius RH . In order to obtain the hydrodynamic radius with the DLS horizontally polarised red semiconductor laser beam (654 nm, 20
method it is necessary to measure the time dependence of the mW). The in-focus image of the droplet, registered with a CCD cam-
autocorrelation function GðsÞ of the scattered light intensity IðtÞ. era, was used to determine the droplet position and measure its dis-
For a delay time s, the normalised ACF is given by [17,21] tance from the centre of the trap (reference position) in arbitrary
hIðtÞIðt þ sÞi units (pixels). This measurement provided the error signal for
GðsÞ ¼ : ð1Þ feed-back loop (operating at 50 cycles/s), which controlled the DC
hIðtÞi2
voltage source and stabilised the particle at the reference position.
For a monodisperse system of particles with Brownian motion the It should be noted that the applied DC voltage is proportional to the
ACF exhibits the following exponential decay [19]: droplet mass-to-charge ratio and with proper calibration can be
used for finding the (effective) droplet mass/radius (electrostatic
GðsÞ ¼ A0 þ A1 expð2Dq2 sÞ; ð2Þ
weighting). See e.g. [25] for the details of our trap arrangement.
where A0 is the baseline of the ACF for s ! 1, and A1 is the ampli- The optical scheme for the scattered light detection consisted of
tude of the ACF at s ¼ 0. The scattering vector a two-lens objective (using aspherised lenses) and an optical fibre
4pn (Fig. 1). The droplet was in the focus of an objective and a dia-
q¼ sinðh=2Þ ð3Þ phragm of 1-mm-diameter limited the numerical aperture to
k
ensure the scattering angle very close to 90 . For the detection of
is connected with the refraction index of the dispersion medium n, the fluctuations of the light scattered by a single micron-sized dro-
vacuum wavelength of the incident light k and with the scattering plet, we used a fast photon-counting PMT (R649, Hamamatsu) with
angle h. In many cases the intensity correlation function (2) can a thermoelectric cooler and an amplifier-discriminator (F-100T,
be written in the form of the so-called Siegert relation [18,22], in Advanced Research Instruments Corporation) that sends the TTL-
which A0 ¼ 1 and A1 ¼ b is the coherence factor, which is deter- level 10-ns pulses to the real-time digital correlator (Photocor-
mined by the experimental conditions. The exponential time con- FC, Photocor Instruments) [29]. The autocorrelation functions were
 
stant s0 ¼ 1= 2Dq2 can be easily found from the GðsÞ curve for analysed with Alango DynaLS software [30] to obtain the corre-
the monodisperse sample. This is the characteristic time of the con- sponding characteristic times by discrete component analysis.
sidered diffusion process. However, in most cases the samples are Although only a small part of the 2 mm laser beam was scattered
polydisperse and the ACF becomes a sum of single exponential by a microdroplet, the sensitivity of the DLS setup is sufficient to
decay functions with different decay time constants (characteristic get the signal level of 105 —106 cps. As a result, we could obtain a
times) si [23]. high-quality autocorrelation function after a time of only 10 s.
However, in order to expand the temporal range of the ACF, accu-
2.2. Experimental details mulation time sometimes reached t a  103 s. The light scattered by
the droplet was also collected with a CCD camera as out-of-focus
In this work, single droplets of diverse dispersions of images. For (nearly) homogeneous and spherical microdroplets,
nanoparticles, levitating in an electrodynamic quadrupole trap, such as usually encountered at the beginning of their evolution
are investigated with light scattering techniques. We used two in the trap, the Mie-scattering patterns [31] could be used for very
types of nanoparticles: SiO2 and TiO2. The first represents natural accurate droplet radius measurement [25]. Such measurement was
G. Derkachov et al. / Measurement 158 (2020) 107681 3

Fig. 1. The optical scheme of the dynamic light scattering experiment in the electrodynamic trap.

used for precise calibration of the droplet electrostatic weighting. for a TetraEG dispersion of the 250 nm titanium dioxide (TiO2 )
Then the electrostatic (DC field) weighting enables measuring the nanospheres. Here and in the following figures, the ACF is pre-
droplet/aggregate (effective) radius during the whole evolution sented in normalised form as ðGðsÞ  1Þ=b according to the relation
(evaporation), in particular for the rigid aggregate. (1), and in semi-log scale, unless specified otherwise. The
A conventional DLS setup with a cuvette was used for reference coherence factor b is determined in our setup by the real-time
particle size measurements for the given dispersion. It utilised a correlator.
diode laser working at 405 nm wavelength [14,15]. These experi- The autocorrelation function for the accumulation time
ments were also performed at room temperature. ta1 ¼ 10; 000 s (Fig. 3, curve (1)) exhibits a periodic character (with
The coherence factor b of the measured ACF is strongly depen- period T ¼ 296 s). The corresponding light intensity oscillations are
dent on the geometry of the experiment, in particular on the num- presented in the right panel of Fig. 2 in a ten times shorter time-
ber N of coherence areas Acoh that is defined by the following scale, to show the structure of the morphology resonances in
relation [17,32]: greater detail. In order to identify the Brownian motion of TiO2
2 nanospheres we used a short-time part of this ACF obtained after
k2 l
NAcoh ¼ : ð4Þ subtraction the long-time A0 level (according to relation (2)) and
pa2 normalisation. Such procedure was possible since the periodic part
Here l is the distance of the detector from the scattering volume, of the ACF changes very little over the first few orders of magnitude
and a is the scattering radius (i.e. the microdroplet radius). It is on a time scale.
important for a good signal quality that the radius of detector b is The short-time part is represented in Fig. 3 by the curve (2).
chosen to have N  1 and Acoh ¼ pb . Increasing the number of
2
There are several characteristic times that can be identified in this
coherence areas above unity leads to a decrease of the ACF ampli- part of the full ACF. The longest characteristic time of about 1 s is
tude [32,33] that is associated with the decrease of the coherence the remnant of the long-time part and should be associated with
factor. The geometry of our experimental setup (Fig. 1) meets the the optical resonances, which can be seen in Fig. 2. The shortest
condition of N  1. The optical fibre of 2b ¼ 400 lm core diameter characteristic time (s1 ¼ 0:034 s) corresponds to the Brownian
defines a much smaller viewed-volume and at the same time motion of titania nanospheres in the microdroplet.
increases the measurement sensitivity [34,35]. The small local maxima that are present on the curve (2) corre-
At the initial stage of a microdroplet’s evolution, the optical res- spond to the frequency of the AC field of the electrodynamic trap.
onances of a dielectric sphere (also known as Morphology Depen- In our experiment, the AC frequency varied from 200 Hz at the
dent Resonances) [36,37] dominate in the temporal changes of the beginning of the droplet evolution to about 700 Hz at the final
scattered light intensity (see Fig. 2). The analysis of these reso- stage, when most of the liquid has evaporated and the mass of
nances enables an accurate relative or even absolute measurement the resulting aggregate was significantly smaller. The presence of
of the temporal evolution of the droplet radius [38,39]. Alterna- such maxima in the ACF confirms the high sensitivity of the tech-
tively, as mentioned above, it is possible to analyse the angular dis- nique to small changes in the intensity of scattered light, such as
tribution of the scattered light using a high-resolution CCD camera those connected with the small droplet’s oscillations at the AC fre-
with a wide dynamic range [24,25,40]. quency, relative to the equilibrium position – so-called micromo-
tion. Since the droplet/aggregate was in the ground-state of the
3. ACFs for levitating microdroplets of diverse dispersions of trap, there was no droplet macro-motion at the eigenfrequency
nanospheres of the trap.
In order to eliminate the effects of long-time signal changes
3.1. ACFs for the evaporation stage of microdroplet upon the ACF, smaller accumulation times could be used (see
curves (3) and (4) in Fig. 3). On the other hand, the Brownian
Fig. 3 shows representative autocorrelation functions that were motion of the titania nanospheres in TetraEG is very slow and a
obtained over a broad (nine orders of magnitude) temporal range rather long measurement is required. Therefore, the accumulation
4 G. Derkachov et al. / Measurement 158 (2020) 107681

Fig. 2. The optical morphology resonances manifesting in the temporal dependence of the intensity of light scattered by a very slowly evaporating droplet of tetraethylene
glycol with titanium dioxide nanospheres. The structure of several consecutive resonances in 3000–4100 s range is shown in detail in the right panel.

Fig. 3. Normalised autocorrelation functions for 250 nm TiO2 nanospheres in TetraEG. (1–4) pertain to the microdroplet and wavelength k ¼ 497 nm, while (5) pertains to the
bulk sample in a cuvette and wavelength k ¼ 405 nm. (1, 2): ACF with the accumulation time ta1 ¼ 10; 000 s, (3, 4): subsequent ACF with ta3 ¼ 20:5 s and t a4 ¼ 7:1 s
respectively. The inset shows a typical snapshot of the Mie-scattering pattern for wavelength k ¼ 497 nm at the beginning of the droplet evaporation.

time t a4  7 s (see curve (4)) seems not sufficiently long to get a nanospheres there are respectively  300 of TiO2 and  500 of
good quality ACF. SiO2 nanospheres inside a 20 lm microdroplet.
The ACF labelled as (5) was obtained with the conventional DLS Assuming a uniform distribution of nanoparticles in the volume,
setup (in a cuvette) for the same dispersion. Taking into account the average distance between them is  3 lm then. Thus, the num-
the difference in the wavelength of the scattered light, we found ber of scattering centres in the microdroplet is small in the optical
s5 ¼ 0:023 s and got a good agreement with the results obtained path (image of such droplet is similar to that of a pure liquid dro-
for a microdroplet. This indicates that, in principle, in our DLS plet, see Fig. 4a), which explains the small amplitude of the ACF at
experiment with the levitating droplet it is possible to detect the the initial stage of temporal droplet evolution. In the case of water
Brownian motion of nanoparticles in a droplet of micrometre size. as the dispersion medium, a sensible ACF accumulation time sig-
However, it must be kept in mind, that this may not be possible for nificantly exceeds the interval between the optical morphology
quickly evaporating microdroplets (dispersion media of higher resonances. As mentioned above, since the autocorrelation of a
volatility) since relatively long ACF accumulation times are periodic signal from the photon counter system is also periodic
necessary. with the same period, the ACF analysis is significantly hindered
at this stage as longer accumulation times are not accessible. Thus,
3.2. ACFs for microaggregates for a quickly evaporating droplet (e.g. of aqueous dispersion) the
ACF with exponential decay can be obtained only at the end of
In order to prevent multiple light scattering, the concentration the droplet evolution.
of nanoparticles in the microdroplet should be small. We used var- The evaporation of the dispersion medium leads to an increase
ious initial concentrations from 0.25 down to 0.005 wt% of silica in the particles’ concentration and ultimately to the formation of
and titania particles dispersed in water. Knowing the density of sil- (non-rigid) aggregates. The images in panels (b), (c) and (d) in
ica and titania in nanoparticles, as well as of water, it can be found Fig. 4 correspond to an aggregate consisting of  300 titania parti-
that, for example, for 0.25 wt% concentration of the 250 nm cles, after the water evaporation (the initial concentration of
G. Derkachov et al. / Measurement 158 (2020) 107681 5

ACFs obtained for microaggregates exhibit properties, which do


not correspond to the standard assumption of the Brownian
motion of the nanoparticles.
Fig. 6 shows the ACFs corresponding to final stages of evolution
of four different droplets that contained silica nanospheres dis-
persed in DEG and TetraEG respectively. First of all, the opposite
shift in the distribution of the characteristic times versus the diam-
eter of the nanospheres (100 and 450 nm) can be noticed. Sec-
ondly, despite the significant difference of viscosity between DEG
and TetraEG the obtained characteristic times hardly differ for
the same nanoparticle diameter (see the characteristic time distri-
butions in the insets in Fig. 6).
Naturally, the rotational Brownian motion of a non-spherical
aggregate should be considered as a source of the observed ACF
properties. There is a lot of theoretical research on this topic (see
e.g. [43–45]). Usually however it is applied to the liquid medium.
For example, the rotational Brownian motion of isolated
micrometre-sized ellipsoidal particles in water was experimentally
studied by Han et al. [46] in the framework of 2D random walk tra-
jectories. Application to gaseous medium – droplets trapped with
electric fields in air – was implemented experimentally by Krieger
et al. [13]. The authors claim that in their experiment the charac-
Fig. 4. (a) Experimental in-focus image of a levitated, pure TetraEG,  40 lm teristic time of rotational Brownian motion was on a time scale
droplet and (b, c, d) consecutive in-focus images of an aggregate – at the final stage of seconds. We estimate that in our experiment it was on a time
of evolution – obtained from a water droplet containing  300 TiO2 spheres of scale of tens of milliseconds, as visualised in the real-time, 50 fps
250 nm diameter. Image (a) was obtained with additional back-illumination with movie of a dry aggregate [47]. The distributions of characteristic
unpolarised white light, in order to visualise the droplet boundaries.
times corresponding to different dispersions (10 ; 20 ; 30 ; 40 in Fig. 6)
show peeks at large values (10–100 ms), while displaying a signif-
nanoparticles of 250 nm diameter was 0.025 wt%). The roughness icant variation in this region. The characteristic time of rotational
of the surface of the aggregate can be inferred from the images. It Brownian motion calculated with Einstein–Smoluchowski relation
was shown in the SEM study of the aggregates deposited on a sub- with the rotational frictional drag coefficient for a sphere at vis-
strate [41] that their overall shape is regular - close to spherical - cosities comparable to that of air seems to be in agreement with
and the irregularities of the surface are minor. these results. It must be kept in mind that microaggregates are
The small size of the aggregate precludes the application of 2D- not (perfectly) spherical and their surface is highly structured.
autocorrelation analysis, as it was done in [42]. We obtained the Based on the in-focus images (see the movie [47] and snapshots
ACFs of diverse aggregates using different initial concentrations in panels (b) – (d) in Fig. 4), it can also be inferred that the random
of silica and titania spherical nanoparticles. Systematic observa- rotations of the microaggregate in the electrodynamic trap
tions have shown that the ACF almost always had an exponential induce/facilitate changes of its shape. We haven’t noticed any
decay like in the case of a monodisperse scattering bulk medium. dependence of the rotational Brownian motion of aggregates upon
The regular rotational motion of a rigid aggregate, if present, is also the AC frequency. Thus we do not expect that they are significantly
expected to introduce a periodicity into the autocorrelation func- assisted by the AC field. We expect that ACFs corresponding to dif-
tion for the large accumulation times. However, we have never ferent levitating aggregates (Fig. 5 and 6) reflect primarily the
found such periodicity. Small maxima visible on curves (1), (2) in changes in the distribution of mass and charge in the aggregate
Fig. 3 and 5 correspond to the frequency of the AC field of the trap and its random rotation rather than the parameters of the original
and were by default eliminated by the proper selection of the DC dispersion.
voltage and AC frequency as described in Section 2.2. It is however
expected that the frequency of free aggregate rotation would be
independent and different from the AC frequency. 4. Conclusions
As the reference, the ACFs was also found in the standard DLS
setup in a cuvette for the laser wavelength of 405 nm (curve (5) We used the dynamic light scattering technique adapted to
in Fig. 5). It confirmed the monodisperse character of the studied study single levitating microdroplets of dispersion. Three laser
wavelengths, different sizes of silica and titania nanoparticles
dispersion, with the characteristic time s5 ¼ 6:1  104 s. In accor-
and their concentrations in different dispersion media (liquids)
dance with relation (3), the characteristic times obtained from
were used. We registered ACFs at the evaporation stage of micro-
standard DLS experiments for the same conditions should be
droplets and at the final stage of evolution corresponding to
scaled with the wavelengths of the illumination source. For
microaggregates.
instance, for 497 nm and 654 nm, the exponential ACF decays with
At the initial stage of the microdroplet temporal evolution it
the characteristic times of 9:2  104 and 16:3  104 respectively
seems possible to find the characteristic times corresponding to
can be expected. Then, sð405Þ : sð497Þ : sð654Þ ¼ 1 : 1:5 : 2:7.
the Brownian motion of the dispersed nanoparticles, though the
However, in our DLS experiment with microdroplets (curves (3)
optical morphology resonances of a dielectric sphere manifest
and (4) in Fig. 5) we obtained s3 ð654Þ ¼ 1:23  103 s and strongly as periodic changes of the scattered light intensity.
s4 ð497Þ ¼ 0:61  10 3
s, which corresponds to the ratios of At the final stage of the droplet evaporation, the formation of
1 : 2:2 : 4:4, significantly differing from the theoretical prediction non-rigid aggregates of nanoparticles is confirmed by the in-
for the Brownian motion. It can also be noticed in the inset in focus images. We obtained the ACFs of the aggregates using differ-
Fig. 5 (in ln-lin scale) that curves (3) and (4) exhibit a somewhat ent sizes of silica and titania spherical nanoparticles in water, DEG
different character (curvature) than curve (5). As expected, the and TetraEG. The autocorrelation functions measured with the DLS
6 G. Derkachov et al. / Measurement 158 (2020) 107681

Fig. 5. A series of the ACFs that were measured for 0.004 wt% aqueous dispersion of 250 nm TiO2 nanoparticles in droplets and in cuvette. Curves (1), (2) correspond to ACFs
for a larger droplet, while curves (3), (4) correspond to a smaller droplet. Curves (1), (3) correspond to the wavelength of 654 nm, while (2), (4) correspond to the wavelength
of 497 nm. Curve (5) corresponds to the measured in a cuvette at the wavelength of 405 nm. The inset presents the curves (3), (4) and (5) in ln-lin scale in a short range of
delay time.

Fig. 6. The ACFs obtained at the wavelength of 497 nm for four droplets of SiO2 nanoparticles dispersions at the final evolution stage – (nearly) dry microaggregates. The left
panel corresponds to DEG as a dispersion medium, while the right panel corresponds to TetraEG. Curves (1, 3) correspond to nanoparticle diameter of 100 nm, while (2, 4)
correspond 450 nm diameter. The insets (10 ; 30 ) and (20 ; 40 ) show the characteristic time distributions for each of the four functions respectively.

technique for microdroplets exhibit exponential decay. However, Declaration of Competing Interest
the temporal dependencies of these ACFs are significantly different
from those found for the nanoparticles dispersion in bulk. It fol- The authors declare that they have no known competing finan-
lows from the rotational Brownian motion and shape changes of cial interests or personal relationships that could have appeared to
the final non-rigid aggregates in the AC electric field of the trap. influence the work reported in this paper.
The analysis showed that the dynamics of a levitated droplet of
dispersion is rather complex and, apart from ACF, it requires simul- References
taneous tracking of additional parameters such as size, weight,
number of nanoparticles, electric charge (distribution), etc. The dif- [1] V.J. Mohanraj, Y. Chen, Nanoparticles-a review, Tropical J. Pharm. Res. 5 (2006)
ficulties encountered in experiments on single levitating droplets 561–573, https://doi.org/10.4314/tjpr.v5i1.14634.
[2] M. Haase, H. Schäfer, Upconverting nanoparticles, Angew. Chem. Int. Ed. 50
and analysing the obtained results call for further detailed
(2011) 5808–5829, https://doi.org/10.1002/anie.201005159.
research. Also, the theoretical basis for dynamic light scattering [3] D.L. Fedlheim, C.A. Foss, Metal Nanoparticles: Synthesis, Characterization, and
on a single microdroplet awaits an additional consideration. Applications, CRC Press, 2001.
[4] R.H. Kodama, Magnetic nanoparticles, J. Magn. Magn. Mater. 200 (1999) 359–
372, https://doi.org/10.1016/S0304-8853(99)00347-9.
[5] P. Herckes, K.T. Valsaraj, J.L. Collett, A review of observations of organic matter
in fogs and clouds: origin, processing and fate, Atmos. Res. 132–133 (2013)
CRediT authorship contribution statement
434–449, https://doi.org/10.1016/j.atmosres.2013.06.005.
[6] H. Moosmüller, R.K. Chakrabarty, W.P. Arnott, Aerosol light absorption and its
G. Derkachov: Software. D. Jakubczyk: Supervision, Writing - measurement: a review, J. Quant. Spectrosc. Radiat. Transf. 110 (2009) 844–
review & editing. K. Kolwas: Conceptualization. K. Piekarski: For- 887, https://doi.org/10.1016/j.jqsrt.2009.02.035.
[7] U.K. Krieger, C. Marcolli, J.P. Reid, Exploring the complexity of aerosol particle
mal analysis. Y. Shopa: Investigation, Writing - original draft. M. properties and processes using single particle techniques, Chem. Soc. Rev. 41
Woźniak: Investigation. (2012) 6631–6662, https://doi.org/10.1039/C2CS35082C.
G. Derkachov et al. / Measurement 158 (2020) 107681 7

[8] S. Mitic, M.Y. Pustylnik, G.E. Morfill, E. Kovačević, In situ characterization of [27] P.C.F. Møller, L.B. Oddershede, Quantification of droplet deformation by
nanoparticles during growth by means of white light scattering, Opt. Lett. 36 electromagnetic trapping, EPL, 88 (2009) 48005(6p). https://doi.org/10.1209/
(2011) 3699–3701, https://doi.org/10.1364/OL.36.003699. 0295-5075/88/48005
[9] S. Saengkaew, T. Charinpanitkul, H. Vanisri, W. Tanthapanichakoon, Y. Biscos, [28] E. Giglio, B. Gervais, J. Rangama, B. Manil, B.A. Huber, D. Duft, R. Müller, T.
N. Garcia, G. Lavergne, L. Mees, G. Gouesbet, G. Gréhan, Rainbow Leisner, C. Guet, Shape deformations of surface-charged microdroplets, Phys.
refractrometry on particles with radial refractive index gradients, Exp. Fluids Rev. E 77 (2008) 36319(7p), https://doi.org/10.1103/PhysRevE.77.036319.
43 (2007) 595–601, https://doi.org/10.1364/AO.37.003227. [29] Multi-angle dynamic and static light scattering instrument Photocor complex.
[10] C. Li, W. Xue-cheng, C. Jian-zheng, C. Ling-hong, G. Gréhan, C. Ke-fa, http://www.photocor.com/dls-instrument.
Application of rainbow refractometry for measurement of droplets with [30] Software for Particle Size Distribution Analysis in Photon Correlation
solid inclusions, Opt. Laser Technol. 98 (2018) 354–362, https://doi.org/ Spectroscopy. http://www.softscientific.com.
10.1016/j.optlastec.2017.07.026. [31] C.F. Bohren, D.R. Huffman, Absorption and Scattering of Light by Small
[11] S. Coëtmellec, W. Wichitwong, G. Gréhan, D. Lebrun, M. Brunel, A.J.E.M. Particles, Wiley-VCH, 1998.
Janssen, Digital in-line holography assessment for general phase and opaque [32] J.W. Goodman, Statistical Properties of Laser Speckle Patterns (Laser Speckle
particle, J. Eur. Opt. Soc. 9 (2014) 14021(15pp). https://doi.org/10.2971/jeos. and Related Phenomena. Topics in Applied Physics 9) ed. J.C. Dainty, Berlin,
2014.14021 Springer, Heidelberg, 1975.
[12] B. Bronk, M. Smith, S. Arnold, Photon-correlation spectroscopy for small [33] A. Zardecki, C. Delisle, J. Bures, Generalized motion of the coherence area, Opt.
spherical inclusions in a micrometer-sized electrodynamically levitated Commun. 5 (1972) 298–300.
droplet, Opt. Lett. 18 (1993) 93–95, https://doi.org/10.1364/OL.18.000093. [34] A.J. MacFadyen, B.R. Jennings, Fibre-optic systems for dynamic light scattering
[13] U.K. Krieger, A.A. Zardini, Using dynamic light scattering to characterize mixed – a review, Opt. Laser Technol. 22 (1990) 175–187, https://doi.org/10.1016/
phase single particles levitated in a quasi-electrostatic balance, Faraday 0030-3992(90)90105-D.
Discuss. 137 (2007) 377–388, https://doi.org/10.1039/B702148H. [35] V. Petta, N. Pharmakakis, G.N. Papatheodorou, S.N. Yannopoulos, Dynamic
[14] G. Derkachov, D. Jakubczyk, K. Kolwas, Y. Shopa, M. Wozniak, Dynamic light light scattering study on phase separation of a protein-water mixture:
scattering with nanoparticles: setup and preliminary results, in: O. Fesenko, L. application on cold cataract development in the ocular lens, Phys. Rev. E 77
Yatsenko (eds.), Nanophysics, Nanomaterials, Interface Studies, and (2008) 061904, https://doi.org/10.1103/PhysRevE.77.061904.
Applications. NANO 2016. Springer Proceedings in Physics 195 (2017) 275– [36] H. van de Hulst, Light Scattering by Small Particles, Wiley, New York, 1957.
282. https://doi.org/10.1007/978-3-319-56422-7_19 [37] A. Ashkin, J. Dziedzic, Observation of optical resonances of dielectric spheres
[15] G. Derkachov, D. Jakubczyk, K. Kolwas, Y. Shopa, M. Wozniak, T. by light scattering, Appl. Opt. 20 (1981) 1803–1814, https://doi.org/10.1364/
Wojciechowski, Application of dynamic light scattering for studying the AO.20.001803.
evolution of micro- and nano-droplets, in: Proc. SPIE 10612, Thirteenth [38] M.B. Hart, V. Sivaprakasam, J.D. Eversole, L.J. Johnson, J. Czege, Optical
International Conference on Correlation Optics, 2018, 106120U. https://doi. measurements from single levitated particles using a linear electrodynamic
org/10.1117/12.2305397 quadrupole trap, Appl. Opt. 54 (2015) F174–F181, https://doi.org/10.1364/
[16] H.G. Merkus, Particle Size Measurements: Fundamentals, Practice, Quality, AO.54.00F174.
Springer Particle Technology Series, 2009. [39] J. Archer, PhD Thesis: Investigation of Surfaces of Evaporating microdroplets of
[17] R. Pecora (Ed.), Dynamic Light Scattering: Applications of Photon Correlation Colloidal Suspensions, Institute of Physics, Polish Academy of Sciences, 2017.
Spectroscopy, Springer Science & Business Media, 2013. [40] N. Riefler, R. Schuh, T. Wriedt, Investigation of a measurement technique to
[18] W. Brown, Dynamic light scattering: the method and some applications. estimate concentration and size of inclusions in droplets, Meas. Sci. Technol.
Clarendon Press; New York: Oxford University Press, Oxford [England], 1993. 18 (2007) 2209–2218, https://doi.org/10.1088/0957-0233/18/7/054.
[19] W.I. Goldburg, Dynamic light scattering, Am. J. Phys. 67 (1999) 1152–1160, [41] M. Woźniak, G. Derkachov, K. Kolwas, J. Archer, T. Wojciechowski, D.
https://doi.org/10.1119/1.19101. Jakubczyk, M. Kolwas, Formation of highly ordered spherical aggregates
[20] D.E. Koppel, Analysis of macromolecular polydispersity in intensity correlation from drying microdroplets of colloidal suspension, Langmuir 31 (2015) 7860–
spectroscopy: the method of cumulants, J. Chem. Phys. 57 (1972) 4814–4820, 7868, https://doi.org/10.1021/acs.langmuir.5b01621.
https://doi.org/10.1063/1.1678153. [42] J. Kielar, P. Lemaitre, C. Gobin, Y. Wua, E. Porcheron, S. Coetmellec, G. Grehan,
[21] B.J. Berne, R. Pecora, Dynamic Light Scattering: With Applications to M. Brunel, Simultaneous interferometric in-focus and out-of-focus imaging of
Chemistry, Biology, and Physics, Krieger, Malabar, FL, 1976. ice crystals, Opt. Commun. 372 (2016) 185–195, https://doi.org/10.1016/j.
[22] A. Mailer, P. Clegg, P. Pusey, Particle sizing by dynamic light scattering: non- optcom.2016.04.004.
linear cumulant analysis, J. Phys.: Condens. Matter. 27 (2015) 145102, https:// [43] B.E. Dahneke, D.K. Hutchins, Characterization of particles by modulated
doi.org/10.1088/0953-8984/27/14/145102, 8p. dynamic light scattering. I. Theory, J. Chem. Phys. 100 (1994) 7890–7902,
[23] B.J. Frisken, Revisiting the method of cumulants for the analysis of dynamic https://doi.org/10.1063/1.466835.
light-scattering data, Appl. Opt. 40 (2001) 4087–4091, https://doi.org/ [44] G. Weber, Rotational Brownian Motion and polarization of the fluorescence of
10.1364/AO.40.004087. solutions, Adv. Protein Chem. 8 (1953) 415–459, https://doi.org/10.1016/
[24] D. Jakubczyk, M. Kolwas, G. Derkachov, K. Kolwas, M. Zientara, Evaporation of S0065-3233(08)60096-0.
microdroplets: the ”Radius-Square-Law” Revisited, Acta Phys. Pol. A 122 [45] P.S. Hubbard, Rotational Brownian Motion, Phys. Rev. A 6 (1972) 2421–2433,
(2012) 709–716, https://doi.org/10.12693/APhysPolA.122.709. https://doi.org/10.1103/PhysRevA.6.2421.
[25] D. Jakubczyk, G. Derkachov, M. Kolwas, K. Kolwas, Combining weighting and [46] Y. Han, A.M. Alsayed, M. Nobili, J. Zhang, T.C. Lubensky, A. Yodh, Brownian
scatterometry: application to a levitated droplet of suspension, J.Q.S.R.T. 126 motion of an ellipsoid, Science 314 (2006) 626–630, https://doi.org/
(2013) 99–104, https://doi.org/10.1016/j.jqsrt.2012.11.010. 10.1126/science.1130146.
[26] E.J. Davis, A history of Single Aerosol Particle Levitation, Aerosol Sci. Tech. 26 [47] https://youtu.be/K0P-XGKghGY.
(1997) 212–254, https://doi.org/10.1080/02786829708965426.

You might also like