You are on page 1of 16

Particle transport and deposition near the contact line of an

evaporating sessile droplet containing mono- or bi-dispersed


colloidal particles
Nagesh D. Patil1,*, Boris Stoeber2
1
Department of Mechanical Engineering, Indian Institute of Technology Bhilai,
Raipur 492015, India.
2
Department of Mechanical Engineering, Department of Electrical and Computer Engineering,
The University of British Columbia, Vancouver, V6T 1Z4, BC, Canada.
*
Corresponding author (nageshpatil@iitbhilai.ac.in)

Abstract

1. Introduction
The evaporation of a water droplet containing colloidal particles typically , such as a spilled
coffee droplet, frequently results in a ring-like deposition pattern on the substrate for in the case
the contact angle is smaller than 90° and the contact line is stationary; a common example is a
spilled coffee droplet. The evaporation mass flux across the liquid-gas interface is non-uniform
along the droplet radius, and it is highest near the contact line, where diffusive transport of water
molecules away from the surface occurs more rapidly, . As a resulting in a, particles move
radially outward along with the capillary flow. Likewise, the geometric change of an evaporating
droplet with a stationary contact line and a spherical cap surface is associated with flow towards
the contact line. Both effects lead to a transport of suspended particles toward until they deposit
near the contact line where they can deposit and form a ring-shaped pattern. This so-called
“coffee-ring” effect was first explained by Deegan et al. 1 in 1997. Since then the transport and
deposition of particles associated with the evaporation of droplets has been an active area of
research. This topic is relevant to diverse applications in the areas of biosensing 2, ink-jet
printing3_ENREF_3, surface coating4, and biochemistry5.

1
The evaporation of a sessile droplet on a substrate can involve two modes:, during the constant
contact radius (CCR) mode, where the contact radius (r) stays constant, meaning the contact line
stays stationary (often referred to as “pinned”) while the contact angle (θ) decreases as the
droplet evaporates;, and in contrast, during the constant contact angle (CCA) mode, where the
contact angle remains constant while the contact radius decreases as the contact line moves
towards the droplet center6. During an evaporation process, the droplet can experience one of
these two modes, or the CCR mode can be followed by the CCA mode. In the last stage of
evaporation, a third mode is often observed; during this so-called mixed mode the contact radius
as well as the contact angle decrease. In general, the CCR mode usually occurs on hydrophilic
substrates (i.e. contact angle below 90°) and leads to a ring-deposit, while the CCA mode occurs
on hydrophobic substrates (i.e. contact angle above 90°) and results in an “inner-deposit”.
Several studies have investigated techniques to control the flow dynamics of evaporating
droplets to manipulate the resulting particle deposition7. In this contextAs an example, the
temperature-induced Marangoni flow offers a way to control the particle deposit, when a droplet
evaporates on a heated substrate. The thermal Marangoni flow is generated due to a surface
tension gradient at the liquid-air interface that results from the temperature difference between
the droplet contact line and the droplet apex 8. The surface tension of water drops decreases with
increasing temperature. This results in a flow along the droplet surface away from the warmer
high-temperature contact line that contributes to an inner deposit named refered to as “coffee
eye” pattern as particles are advected back from the contact line to the droplet center, where they
descend towards the substrate and then deposit8. Moreover, as explained by Li et al.8, during the
evaporation of an aqueous droplet on a heated substrate, a stagnation region develops on the air-
water interface near the contact line due to the outward capillary flow and inward circulatory-
Marangoni flow. Parsa et al.9 and Patil et al.10 observed that during the evaporation of a droplet
with bi-dispersed particles, the particles can separate in the stagnation region and form a deposit
pattern with smaller particles depositing closer to the contact line in comparison to the larger
particles. However, the physical understanding with respect to of the combined effect of
substrate temperature induced inward-Marangoni flow and outward-Capillary flow, associated
with, and the particle sizes of suspended particles on during the particle transport, separation, and
deposition near the contact line of an evaporating sessile droplet is complex and has not been
fully explained yet. To date, researchers have reported the effect of either the substrate

2
temperature or of the particles size on the deposition pattern for mono- and bi-dispersed colloidal
particles.
In the context of mono-dispersed colloidal particles, the effect of substrate temperature on the
deposit pattern has been studied by several researchers. For example, Li et al.8 used a glass
substrate with an equilibrium contact angle, eq = 24-30°, and droplets containing polystyrene
particles of diameter, dP = 0.1 µm at a volume concentration c = 0.25% (v/v). They studied only
the effect of substrate temperature (Ts = 30-80°C) on the particle deposition patterns, and showed
that at 30°C a ring-like pattern forms and with increase in temperature the pattern changes to a
thin ring with an inner deposit pattern. Parsa et al. 11 performed experiments on a silicon substrate
(eq = 30-40°) and varied the substrate temperature (Ts = 25−99°C). They used a copper oxide
powder with dP = 50 nm (at a weight concentration c = 0.05%(w/w)) and observed a uniform
deposit pattern at 25°C substrate temperature, which changes to a dual-ring and multi-ring
patterns with an increase in temperature. Zhong and Duan12 observed a uniform deposit to dual-
ring pattern with increase in substrate temperature on silicon (Ts = 10−50°C), they used graphite
nanopowders of dP = 2−3 nm (c = 0.05% (v/v)). Patil et al.13 studied the deposit patterns for
droplets with polystyrene particles of dP = 0.46 μm (c = 0.05, 0.1, and 1.0% (v/v) on glass (θeq =
34−38°) and silicon (θeq = 94−97°) for various substrate temperatures (Ts = 27−90°C). They
observed that on glass, a thick-ring deposit changes to a thin ring with inner deposit; while on
silicon, an inner deposit changes to a thin ring with inner deposit, with increase in substrate
temperature. Recently, Lama et al. 14 also studied the effect of substrate temperature on deposit
patterns for TiO2 particles (dP = 22 nm, c = 2.5 % (w/w) on glass (Ts = 25−70°C). They obtained
a ring-like deposit at ambient substrate temperature, a ring with inner deposit at Ts = 45−60°C,
and a ring with uniform deposit at Ts = 70°C. Their study focused on altering the inner deposit to
uniform deposit via an increase in substrate temperature. In all above studies, the alteration in the
deposit patterns from a thick-ring to the thin ring with inner deposit/dual ring/ring with uniform
deposit is formed due to the thermal-induced Marangoni convection that transports the particles
from the contact line region to the inner region of the droplet. Very recently, the effect of
substrate temperature (Ts = 40−70°C) on deposit patterns was studied by Katre et al. 15 for
droplets of ethanol-water binary mixtures with and without Al 2O3 particles (dP = 20-30 nm) and
observed that due to the presence of particles contact line pinning happens on heated substrate in
comparison to droplets without particles.

3
In the context of mono-dispersed colloidal particles, the effect of particle sizes was studied on
various nonheated substrates. For example, Biswas et al.16 studied polystyrene particles of dP =
0.02, 0.2, and 1.1 μm (c = 0.05% w/w) on glass (having different receding contact angles, θrec =
6, 17, 89°). They concluded that the final deposit pattern varies based on particle size and
substrate wettability. On more hydrophilic substrates a ring-like pattern forms for smaller
particles, while on more hydrophobic substrates an inner deposit forms for all tested particles.
They also found that the location of particle deposition in the wedge-shaped region at the contact
line depends upon the particle size. Weon and Je17 compared the deposit patterns on glass (θeq =
15°) for two mono-dispersed colloidal particles with size dP = 0.1 and 1.0 μm (c = 1% w/w).
They found that 0.1 μm particles deposit are more closely and densely packed at the contact line
as compared to that of 1.0 μm particles deposit. Yu et al.18 studied the effect of particle
concentration c = 0.02−1.28% (w/w) of polystyrene particles of dP = 20 μm in water and ethanol-
water droplets on PDMS substrate (θeq = 70−110°). They found that for low particle
concentration an inner deposit forms, which changes to a ring-like pattern with increasing
particle concentration. They also developed a self-pinning model considering the force balance
on the particles at the contact line. Patil et al.10 tested the effect for four particle sizes dP = 0.1,
0.46, 1.1 and 3 μm (c = 0.05% v/v) on silicon (θeq = 85°). They observed that an inner deposit
pattern forms for 0.1-1.1 μm and a ring-like pattern forms for 3 μm. They also reported that the
contact line pinning and depinning depends on the particle size and the substrate temperature.
In the context of bi-dispersed colloidal particles, there are very few studies which investigated
the effect of substrate temperature on particle deposits. For example, Hendarto and
Gianchandani19 studied the size separation of hollow microspheres of glass of dP = 5−200 μm via
isopropyl alcohol droplets evaporation on heated substrate (Ts = 55 to 85°C), and better
separation was seen at the higher substrate temperature. Parsa et al. 9 obtained five different
deposit patterns for evaporation of droplet containing 1 and 3.2 μm carboxylate modified
polystyrene particles on a substrate at different temperatures (Ts = 22−99°C). They observed that
at the contact line smaller particles (1 μm) deposited at the outer edge and a mixture of 1 and 3.2
μm particles form an inner ring. Patil et al. 10 studied size separation of particles on heated silicon
substrate (Ts = 27−90°C), and reported that the separation of particles was dependent on the
contact line pinning due to the substrate temperature and particle diameter ratio, and particles get
separated at the contact line region according to the their sizes. Carreón et al.20 considered the

4
evaporation of droplet containing proteins (bovine and lysozyme) on heated glass (Ts =
25−63°C), and obtained a ring deposit at ambient which alters to a ring with inner deposit pattern
(coffee-eye pattern) with increase in substrate temperature.
In the context of bidispersed colloidal particles for the effect of particle-sizes, several studies
observed the separation of the particles near the contact line region on nonheated substrates. Han
et al.21 studied the deposit patterns for droplets containing particles of dP = 50 and 500 nm on
silicon. Chhasatia and Sun22 studied the deposit patterns for droplets containing particles of dP =
100 nm and 1.1 μm on various glass types. Devlin et al.23 studied deposit patterns for droplets
24 10
containing 1 and 3 μm colloidal particles on silicon. Monteux and Lequeux and Patil et al.
investigated the effect of particle diameter ratio (R) on a nonheated and heated substrate,
24
respectively. Monteux and Lequeux explained that at the droplet edge a depletion region is
generated based upon the particle size and the contact angle of the liquid-gas interface with the
solid. The separation of particles take place in this depletion region. They achieved particle
separation for different combinations of bidispersed particles (dP = 100 nm and 1 μm (R = 0.1),
10
100 nm and 5 μm (R = 0.02), and 1 and 5 μm (R = 0.2)). Patil et al. observed that particle
separation occurs in the stagnation region based upon the particle diameter ratio for different
combinations of particle sizes (dP = 0.1 μm and 3 μm (R = 0.03), 0.1 μm and 1.1 μm (R = 0.09),
0.46 μm and 3 μm (R = 0.15), 0.1 μm and 0.46 μm (R = 0.21), and 1.1 μm and 3 μm (R = 0.36)).
25
Iqbal et al. observed the particle separation behavior on hydrophilic and hydrophobic
substrates for two different combinations of particle sizes (dP = 0.2 μm and 3 μm (R = 0.06), and
1 μm and 6 μm (R = 0.16). In all these studies they reported that smaller diameter particles
deposit closer to the contact line in comparison to larger diameter particles; they also showed
that the particle separation in the deposit pattern decreases with a decrease in the substrate
wettability.
Owing to the above literature, although various studies have observed particle deposition
patterns from evaporating sessile droplets with mono- and bi-dispersed colloidal particles of
different sizes, the physical explanation specifically during particle transport and their separation
as well as deposition near the contact line in view of substrate temperature and particle sizes is
still not clear. Thus, the objectives of this work are: (a) to study the particle transport dynamics
inside an evaporating sessile water droplet containing mono- and bi-dispersed particles on
nonheated and heated glass substrates, and (b) to understand the size-based particle separation

5
and deposition pattern near the contact line considering the effect of thermal-Marangoni flow
and cCapillary flow and the particle size. The effect of substrate temperature is studied by
varying the glass substrate temperature from 21 to 90°C. The effect of particle size is studied by
keeping the size of smaller particles constant (1 μm) and varying the size of larger particles (3.2
or 5 μm). The flow field in droplets is measured dDuring the evaporation to identify , as
explained in the literature a the stagnation region that develops near the contact line due to the
outward capillary flow and inward circulatory-Marangoni flow. The experiments are performed
and data analyzed to see this region inside the droplet where particle separation occurs. The
experiments permit exploring the effect of substrate temperature on the stagnation region and on
particle separation is also explored.

2. Experimental Details
[2.1.] Preparation of colloidal solutionsuspensions
An aqueous colloidal monodispersed suspensions of red fluorescent polystyrene particles were
purchased from different vendors: red particles of diameterof dP = 3.2 μm, with concentration c
= 2.5% (w/v) of monodispersed colloidal particles were purchased from Magsphere Inc, and .
Fluoro-Max dyed blue and green fluorescent polystyrene particles of with dP = 1 and 5 μm (with
1% (w/v) concentration) of monodispersed colloidal particles, respectively were purchased from
Thermofisher Scientific. In the present work, each of tThese solutions suspensions were diluted
to c = 0.05% by adding DI water. The Bbi-dispersed colloidal solutions susppensions were made
prepared by mixing equal volumes of the dP = 1 solutions suspension and either the dP = 3.2 μm
or the dP = 5 μm of two different particle sizessuspension. Prepared The mono- and bi-dispersed
colloidal particles solutions suspensions were stabilized with by 1 hr sonication (Branson, 2800
Ultrasonic Bath) to ascertain that no sedimentation as well as agglomeration of particles
occurred. For bi-dispersed solutions, two particle size combinations were considered: one was 1
and 3.2 µm, and another was 1 and 5 µm.

2.1.[2.2.] Substrate preparation and droplet generation


Microscopeic glass slides of 75 mm x 25 mm x 1 mm and acrylic slides of 75 mm x 25 mm x 1.2
mm were used as a substrate. The slides were cleaned first using isopropanol and then allowed to
dry out completely in under the ambient conditions. Using above prepared solutions, droplets of

6
2±0.2 μL volume were created using a micropipette (Make- Cole-Parmer Inc.) and were gently
deposited on a slide. The experiments were performed at different substrate temperatures ranging
from Ts = 30 to 90°C. To heat the substrate a transparentn ITO (iIndium-tTin-oOxide) coated
glass wasere used as heaters. The glass substrate slide was kept on the ITO based heaters, which
that waswere heated through a DC power supply source meter (Make- Agilent Technologies).
The sSubstrate temperatures wasere measured using an infra-red spot thermal camera (Make-
FLIR Systems Inc.). Once the desired constant substrate temperature was attained and was
showing constant temperature value, thereafter only the droplets with colloidal particles were
deposited on the substrate.droplets of 2±0.2 μL volume of the prepared suspensions were gently
deposited using a micropipette (Cole-Parmer Inc.).

2.2.[2.3.] Visualization of droplet interface and particle deposition near the contact line
Figure 1 shows the schematic of an the experimental set up used for side-visualization of a
droplet interface and visualization of particle deposition near the contact line. During
evaporation, the time-varying droplet shapes iswere recorded from side using a high-speed
camera (Phantom, Miro 4.0) connected to a stereo microscope with an optical resolution of
256x150 pixels at 10 and 100 frames per second for non-heated and heated substrates cases,
respectively. The particles deposition near the contact line was recorded from the bottom using a
confocal microscope (Eclipse Ti, Nikon Instruments). The particles were distinguished based on
their size difference and fluorescence color, as shown schematically in Figure 1. The equilibrium
contact angles (θeq) observed on the surface in the side view for sessile droplets of mono-
dispersed as well as bi-dispersed susppensionssolutions for non-heated/heated glass waswere
around 26±3°, and that for acrylic it waswere around 72±2°. The experiments were performed at
the ambient temperature and relative humidity of 21.0±2°C and 46±6%, respectively.

7
Smaller diameter
Larger diameter particles
particles

Particle deposition
near contact line

Camera with
stereo mircoscope
Side-view of droplet
Substrate
ITO
heater Computer
Confocal
Microscope

Figure 1: Schematic of the experimental set up for side-visualization of droplet interface and
visualization of particle deposition near the contact line.

[2.4.] Visualization of flow field of particles inside the droplets


To qualitatively visualize the flow field inside the droplet, Thean optical coherence tomography
(OCT) imaging system (SD-OCT, TEL1300V2-BU, Thorlabs) was usedserved to visualize the
position of particles inside a droplet over time. OCT is an optical technique based on light
interferometry that, which allows capturing cross-sectional images through the a mostly
transparent target object while recording scatter sources such as particles. The OCT system used
here, with has a beam width of 13 μm and . It employs broadband illumination, centered at 1300
nm wavelength, having a maximum depth of the scan of 3.5 mm and an axial resolution of 4.2
μm. a is Tthe schematic diagram of the experimental setup used to record the cross-sectional
images of the evaporating droplets at along their central vertical plane is shown in a. Images
were captured with 1200x482 resolutions at a rate of 50 frames per second. The inset iIn a, an
inset is showsn schematically for the cross-sectional image of flow field of particles and the
flow field inside a droplet on a heated substrate. Note that in these experiments, the evaporation
of only droplets with only monodispersed colloidal particles (dP = 1 μm) were considered to
observe the flow field trajectories of particles inside the droplets. Due to the small size of the
particles compared to the resolution of the OCT system it is not possible to distinguish the
particle images from particles with different sizes. Bi-dispersed colloidal solution was considered
since the scattering of light with bi-dispersed particles would have created difficulty in tracking

8
the flow field of each of the particles sizes. In b, Aa typical sample cross-sectional image
obtained at the central cross-sectional plane for of an evaporating droplet containing 1 μm
particles on a heated glass substrate is shown in In b.

Mono-dispersed Droplet-air
particles interface
Pipette
0.5 mm

ITO Droplet Vertical cross-sectional view at red line location


heater Substrate

DC Power
Sourcemeter
Computer
Optical Coherence 1 mm
Tomography (OCT)

(a) (b) Top-view

Figure 2: (a) Schematic of the experimental set up for visualization of flow field of particles
inside the evaporating droplet. (b) Typical cross-sectional image at the centre of evaporating
droplet containing colloidal particles captured using OCT imaging.
2.3.[2.5.] PIV analysis
The OCT image sequences showing the particles in cross sections of evaporating droplets To
quantitatively visualize the flow field of particles, above obtained time-varying cross-sectional
images of evaporating droplets are utilized to perform analyzed using particle image velocimetry
(PIV) analysis with using the commercial software package DAVIS 8.1.5 by LaVision to yield
the particle velocity fields, a commercially available software package. PIV analysis gives the
flow field from the displacement of seeded colloidal particles inside the droplet. This determines
the particle displacement field was measured by cross-correlating corresponding interrogatiojn
regions in the time-varying OCT cross-sectional image sequences of corresponding regions in
consecutive images. Further, noting the time delay between two consecutive images and the a
displacement field can eb converted into a particle vector, the velocity vector at each region is
calculatedfield. For this analysis, anThe interrogation window size of 32x32 pixels with multi-
pass processing (4 passes) and with an overlap percentage of 75% was used for velocity vector
calculation. In the analysis, the interrogation window was an elliptical 25 Gaussian weighting
function since the evaporation of droplet on nonheated as well as heated substrate is
axisymmetric and considering that the velocity components are dominant in radial direction.

9
Before performing the PIV analysis on time-varying cross-sectional images of the OCT image
sequencesing, those the images were pre-processed first in ImageJ 1.53i software and then in
MATLAB 2018 to reduce the noise and enhance the image quality. For the analysis, the time
interval considered between the two successive images was 0.2 s.

3. Results and Discussion


3.1. Flow field analysis during droplet evaporation
3.1.1. Flow field inside the droplet and particle deposition pattern
We recorded flow inside the droplet during its evaporation using OCT and performed the PIV
analysis to observe the flow field inside the droplet. We observed the resulting particle
deposition pattern consistent with the literature studies. We performed experiment on nonheated
and heated glass and obtained ring-like and a ring with inner deposit pattern respectively. Inner
deposit is attributed to inward Marangoni circulation inside droplet. Since, our objective was to
analyze the particle deposition near contact line, we needed to analyze the flow field closer to
contact line inside the droplet (presented in next subsection).

Figure 3: Flow field during evaporation of a 2 µL sessile droplet on nonheated and heated
substrate, and the corresponding particle deposition patterns.
3.1.2. Flow field inside the droplet closer to the contact line
We analyzed the flow field closer to the contact line of the droplet to observe the stagnation
point inside the droplet. The stagnation point separates the outward radial flow and inward
Marangoni flow. We observed that as the substrate temperature increases the Marangoni velocity
increases and the stagnation distance decreases. We also found that with increase in substrate
10
temperature the outward radial flow increases as well. We hypothesis that the particle deposition
may get affected by this and performed further analysis for particle deposition near the contact
line for different substrate temperature.

Figure 4: Evaporation dynamics inside the droplet near the contact point region during droplet
evaporation on the heated substrate.

Figure 5: Variation of (a) stagnation distance and Marangoni velocity with substrate temperature,
(b) radial velocity with substrate temperature.

11
3.2. Particle deposition near contact line at fixed time (T = 0.4te) of the droplet evaporation
3.2.1. Effect of substrate temperature and particle size
We performed droplet evaporation on a glass kept at different substrate temperature. It is found
that the particle separation decreases with increase in substrate temperature. The particle
separation decreases corresponding to decreased stagnation distance and increased outward
radial flow. Further, if particle size is larger the particles deposit much away from the contact
line in comparison to smaller size of particles. In general, particle deposits according to their
geometric sizes, smaller size particles deposit closer to contact line and larger size particles next
to that.

Figure 6: Qualitative comparison of the particle distance from the contact line due to the effect of
substrate temperature for different particle sizes from a bidispersed colloidal solution.

12
Figure 7: Droplet contact angle and contact diameter as a function of time during evaporation of
a 2 µL sessile droplet containing bidispersed particles, on a substrate, maintained at different
temperatures.
3.2.2. Comparison of particle distance from contact line

Figure 8: (a) Schematic for the definition of measurement of the particle distance from the
contact line, Si (b) quantitative comparison of particle distance from the contact line due to the
effect of substrate temperature for different particle sizes from a bidispersed colloidal solution.
3.2.3. Comparison of particle separation distance

13
Figure 9: Quantitative comparison of the measured particle separation distance for the
evaporation of a droplet containing bidispersed colloidal particles on a substrate, maintained at
different temperatures.

4. Conclusions

5. Acknowledgments

6. References

1. R. D. Deegan; O. Bakajin; T. F. Dupont; G. Huber; S. R. Nagel; Witten, T. A., Capillary flow


as the cause of ring stains from dried liquid drops. Nature 1997, 389, 827-829.
2. Wen, J. T.; Ho, C.-M.; Lillehoj, P. B., Coffee ring aptasensor for rapid
protein detection. Langmuir 2013, 29 (26), 8440-8446.
3. Al‐Milaji, K. N.; Secondo, R. R.; Ng, T. N.; Kinsey, N.; Zhao, H.,
Interfacial Self‐Assembly of Colloidal Nanoparticles in Dual‐Droplet Inkjet
Printing. Advanced Materials Interfaces 2018, 5 (10), 1701561.
4. Layani, M.; Gruchko, M.; Milo, O.; Balberg, I.; Azulay, D.; Magdassi,
S., Transparent conductive coatings by printing coffee ring arrays obtained
at room temperature. ACS nano 2009, 3 (11), 3537-3542.
5. Brutin, D.; Sobac, B.; Loquet, B.; Sampol, J., Pattern formation in
drying drops of blood. Journal of fluid mechanics 2011, 667, 85-95.
6. Picknett, R. G.; Bexon, R., The evaporation of sessile or pendant drops in still air. J. Colloid
Interface Sci. 1977, 61 (2), 336-350.
7. Zang, D.; Tarafdar, S.; Tarasevich, Y. Y.; Choudhury, M. D.; Dutta,
T., Evaporation of a droplet: From physics to applications. Physics Reports
2019, 804, 1-56.
8. Li, Y.; Lv, C.; Li, Z.; Quéré, D.; Zheng, Q., From coffee rings to coffee
eyes. Soft matter 2015.
9. Parsa, M.; Harmand, S.; Sefiane, K.; Bigerelle, M.; Deltombe, R.,
Effect of substrate temperature on pattern formation of bidispersed
particles from volatile drops. The Journal of Physical Chemistry B 2017,
121 (48), 11002-11017.
10. Patil, N. D.; Bhardwaj, R.; Sharma, A., Self-Sorting of Bidispersed
Colloidal Particles Near Contact Line of an Evaporating Sessile Droplet.
Langmuir 2018.

14
11. Parsa, M.; Harmand, S.; Sefiane, K.; Bigerelle, M.; Deltombe, R.,
Effect of Substrate Temperature on Pattern Formation of Nanoparticles
from Volatile Drops. Langmuir 2015, 31 (11), 3354-3367.
12. Zhong, X.; Duan, F., Disk to dual ring deposition transformation in
evaporating nanofluid droplets from substrate cooling to heating. Physical
Chemistry Chemical Physics 2016, 18 (30), 20664-20671.
13. Patil, N. D.; Bange, P. G.; Bhardwaj, R.; Sharma, A., Effects of
Substrate Heating and Wettability on Evaporation Dynamics and
Deposition Patterns for a Sessile Water Droplet Containing Colloidal
Particles Langmuir 2016, 32, 11958–11972.
14. Lama, H.; Satapathy, D. K.; Basavaraj, M. G., Modulation of central
depletion zone in evaporated sessile drops via substrate heating. Langmuir
2020, 36 (17), 4737-4744.
15. Katre, P.; Balusamy, S.; Banerjee, S.; Chandrala, L. D.; Sahu, K. C.,
Evaporation dynamics of a sessile droplet of binary mixture laden with
nanoparticles. Langmuir 2021.
16. Biswas, S.; Gawande, S.; Bromberg, V.; Sun, Y., Effects of particle
size and substrate surface properties on deposition dynamics of inkjet-
printed colloidal drops for printable photovoltaics fabrication. Journal of
Solar Energy Engineering 2010, 132 (2), 021010.
17. Weon, B. M.; Je, J. H., Self-pinning by colloids confined at a contact
line. Physical review letters 2013, 110 (2), 028303.
18. Yu, Y.-S.; Wang, M.-C.; Huang, X., Evaporative deposition of
polystyrene microparticles on PDMS surface. Scientific Reports 2017, 7 (1),
14118.
19. Hendarto, E.; Gianchandani, Y. B., Size sorting of floating spheres
based on Marangoni forces in evaporating droplets. Journal of
Micromechanics and Microengineering 2013, 23 (7), 075016.
20. Carreón, Y. J.; Ríos-Ramírez, M.; Vázquez-Vergara, P.; Salinas-
Almaguer, S.; Cipriano-Urbano, I.; Briones-Aranda, A.; Díaz-Hernández,
O.; Santos, G. J. E.; González-Gutiérrez, J., Effects of substrate
temperature on patterns produced by dried droplets of proteins. Colloids
and Surfaces B: Biointerfaces 2021, 203, 111763.
21. Han, W.; Byun, M.; Lin, Z., Assembling and positioning latex
nanoparticles via controlled evaporative self-assembly. Journal of Materials
Chemistry 2011, 21 (42), 16968-16972.
22. Chhasatia, V. H.; Sun, Y., Interaction of bi-dispersed particles with
contact line in an evaporating colloidal drop. Soft Matter 2011, 7 (21),
10135-10143.

15
23. Devlin, N. R.; Loehr, K.; Harris, M. T., The separation of two different
sized particles in an evaporating droplet. AIChE Journal 2015, 61 (10),
3547-3556.
24. Monteux, C. c.; Lequeux, F. o., Packing and sorting colloids at the
contact line of a drying drop. Langmuir 2011, 27 (6), 2917-2922.
25. Iqbal, R.; Majhy, B.; Shen, A. Q.; Sen, A., Evaporation and
morphological patterns of bi-dispersed colloidal droplets on hydrophilic and
hydrophobic surfaces. Soft matter 2018, 14 (48), 9901-9909.

16

You might also like