0% found this document useful (0 votes)
28 views170 pages

Heat Transfer and Fluid Dynamics in Boiling Systems

Transferencia de calor y dinámica de fluidos en sistemas de ebullición.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
28 views170 pages

Heat Transfer and Fluid Dynamics in Boiling Systems

Transferencia de calor y dinámica de fluidos en sistemas de ebullición.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 170

Special Issue Reprint

Heat Transfer and Fluid


Dynamics in Boiling Systems

Edited by
Vladimir Serdyukov and Fedor Ronshin

mdpi.com/journal/energies
Heat Transfer and Fluid Dynamics in
Boiling Systems
Heat Transfer and Fluid Dynamics in
Boiling Systems

Editors
Vladimir Serdyukov
Fedor Ronshin

Basel • Beijing • Wuhan • Barcelona • Belgrade • Novi Sad • Cluj • Manchester


Editors
Vladimir Serdyukov Fedor Ronshin
Kutateladze Institute of Kutateladze Institute of
Thermophysics Thermophysics
Novosibirsk Novosibirsk
Russia Russia

Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland

This is a reprint of articles from the Special Issue published online in the open access journal Energies
(ISSN 1996-1073) (available at: https://www.mdpi.com/journal/energies/special issues/HT FD
BS).

For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:

Lastname, A.A.; Lastname, B.B. Article Title. Journal Name Year, Volume Number, Page Range.

ISBN 978-3-7258-0717-8 (Hbk)


ISBN 978-3-7258-0718-5 (PDF)
doi.org/10.3390/books978-3-7258-0718-5

© 2024 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license. The book as a whole is distributed by MDPI under the terms
and conditions of the Creative Commons Attribution-NonCommercial-NoDerivs (CC BY-NC-ND)
license.
Contents

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Evgeny A. Chinnov, Sergey Ya. Khmel, Victor Yu. Vladimirov, Aleksey I. Safonov, Vitaliy V.
Semionov, Kirill A. Emelyanenko, et al.
Boiling Heat Transfer Enhancement on Biphilic Surfaces
Reprinted from: Energies 2022, 15, 7296, doi:10.3390/en15197296 . . . . . . . . . . . . . . . . . . . 1

Alexander V. Fedoseev, Mikhail V. Salnikov, Anastasiya E. Ostapchenko and


Anton S. Surtaev
Lattice Boltzmann Simulation of Optimal Biphilic Surface Configuration to Enhance Boiling
Heat Transfer
Reprinted from: Energies 2022, 15, 8204, doi:10.3390/en15218204 . . . . . . . . . . . . . . . . . . . 20

Vladimir Serdyukov, Ivan Malakhov and Anton Surtaev


The Influence of Pressure on Local Heat Transfer Rate under the Vapor Bubbles during Pool
Boiling
Reprinted from: Energies 2023, 16, 3918, doi:10.3390/en16093918 . . . . . . . . . . . . . . . . . . . 34

Jéssica Martha Nunes, Jeferson Diehl de Oliveira, Jacqueline Biancon Copetti,


Sameer Sheshrao Gajghate, Utsab Banerjee, Sushanta K. Mitra and Elaine Maria Cardoso
Thermal Performance Analysis of Micro Pin Fin Heat Sinks under Different Flow Conditions
Reprinted from: Energies 2023, 16, 3175, doi:10.3390/en16073175 . . . . . . . . . . . . . . . . . . . 48

Semyon Vostretsov, Anna Yagodnitsyna, Alexander Kovalev and Artur Bilsky


Experimental Study of Mass Transfer in a Plug Regime of Immiscible Liquid–Liquid Flow in a
T-Shaped Microchannel
Reprinted from: Energies 2023, 16, 4059, doi:10.3390/en16104059 . . . . . . . . . . . . . . . . . . . 61

Oleg Volodin, Nikolay Pecherkin and Aleksandr Pavlenko


Combining Microstructured Surface and Mesh Covering for Heat Transfer Enhancement in
Falling Films of Refrigerant Mixture
Reprinted from: Energies 2023, 16, 782, doi:10.3390/en16020782 . . . . . . . . . . . . . . . . . . . . 74

Sergey Misyura, Andrey Semenov, Yulia Peschenyuk, Ivan Vozhakov and Vladimir Morozov
Nonisothermal Evaporation of Sessile Drops of Aqueous Solutions with Surfactant
Reprinted from: Energies 2023, 16, 843, doi:10.3390/en16020843 . . . . . . . . . . . . . . . . . . . . 91

Alexander Ashikhmin, Nikita Khomutov, Roman Volkov, Maxim Piskunov and


Pavel Strizhak
Effect of Monodisperse Coal Particles on the Maximum Drop Spreading after Impact on a Solid
Wall
Reprinted from: Energies 2023, 16, 5291, doi:10.3390/en16145291 . . . . . . . . . . . . . . . . . . . 112

Artem N. Kotov, Aleksandr L. Gurashkin, Aleksandr A. Starostin, Kirill V. Lukianov and


Pavel V. Skripov
Nucleation of a Vapor Phase and Vapor Front Dynamics Due to Boiling-Up on a Solid Surface
Reprinted from: Energies 2023, 16, 6966, doi:10.3390/en16196966 . . . . . . . . . . . . . . . . . . . 130

Maksim A. Pakhomov and Viktor I. Terekhov


Modeling of Turbulent Heat-Transfer Augmentation in Gas-Droplet Non-Boiling Flow in
Diverging and Converging Axisymmetric Ducts with Sudden Expansion
Reprinted from: Energies 2022, 15, 5861, doi:10.3390/en15165861 . . . . . . . . . . . . . . . . . . . 144

v
Preface
Boiling is a highly efficient heat transfer mechanism, which has wide-ranging applications in
diverse industrial and technological sectors such as thermal desalination, heat and nuclear energies,
heat pipes, and heat exchangers. It is particularly acknowledged as the most effective method for
the thermal management of high-heat-flux devices like high-power electronics. Ongoing research
endeavors include the investigation of pool and flow boiling mechanisms and the development of
techniques to enhance boiling performance. The process of boiling is intricate and influenced by
various factors, with surface topology and wettability playing a particularly critical role. Accordingly,
the modification of surfaces for boiling applications constitutes a prominent area of research.
This Special Issue presents a collection of carefully selected papers that delve into heat
transfer and fluid dynamics during boiling phenomena. The Special Issue comprises state-of-the-art
experimental and numerical research contributions, with a specific emphasis on recent advancements
in enhancing heat transfer during boiling under different conditions, especially through novel
approaches to surface modification for heating purposes. The Guest Editors extend their gratitude
to the authors for submitting high-quality research articles and to the referees for providing valuable
reviews that facilitated the selection process. It is anticipated that this Special Issue will prove
invaluable to the community and inspire further progress in the rapidly evolving field of boiling. Last
but not least, we want to thank Mr. Chester Zheng—the Assistant Editor, MDPI—for his dedication
to this Special Issue. He has been a major supporter of this Special Issue, and we are indebted to him.
One of the ways to enhance nucleate boiling performance and increase critical heat fluxes
is using wettability patterned surfaces or coatings, so-called biphilic surfaces. In the research of
Chinnov et al., flat surfaces with different patterns of hydrophobic spots were employed for the
experimental investigation of water boiling heat transfer. The results show that boiling heat transfer
on the test biphilic surfaces was significantly higher (up to 600%) than on base surfaces. The highest
heat transfer efficiency was detected for the surface with the largest number of hydrophobic spots.
After long-term experiments (up to 3 years), the heat transfer coefficient on the obtained surfaces
remained higher than on the smooth copper surface. Biphilic surfaces with arrays of cavities formed
due to laser ablation turned out to be the most stable during prolonged contact with boiling water.
To study the processes of boiling on a smooth surface with contrast wettability, Fedoseev et al.
developed a hybrid model based on the Lattice Boltzmann method and heat transfer equation. To
find the optimal configuration of the biphilic surface, at the first stage, a numerical simulation was
carried out for a single lyophobic zone on a lyophilic surface. The dependences of the bubble
departure frequency and the departure diameter of the bubble on the width of the lyophobic zone
were obtained, and its optimal size was determined. At the next stage, the boiling process on an
extended surface was studied in the presence of several lyophobic zones of a given size with different
distances between them. It was shown that, in the region of moderate surface superheat, the intensity
of heat transfer on biphilic surfaces can be several times (more than four times) higher compared to
surfaces with homogeneous wettability. Based on numerical calculations, an optimal configuration of
the biphilic surface with the ratios of the lyophobic zones’ width of the order of 0.16 and the distance
between the lyophobic zones in the range of 0.9–1.3 to the bubble departure diameter was found.
In the research of Serdyukov et al., the results of an experimental study on the evolution of a
nonstationary temperature field during ethanol pool boiling in a pressure range of 12–101.2 kPa are
presented. Experimental data were obtained using infrared thermography with high temporal and
spatial resolutions, which made it possible to reconstruct the distribution of the heat flux density
and to study the influence of pressure reduction on the local heat transfer rate in the vicinity of the

vii
triple contact line under vapor bubbles for the first time. It was shown that, for all studied pressures,
a significant heat flux density is removed from the heating surface due to microlayer evaporation,
which exceeds the input heat power by a factor of 3.3–27.7, depending on the pressure. Meanwhile,
the heat transfer rate in the area of the microlayer evaporation significantly decreases with the
pressure reduction. Estimates of the microlayer profile based on the heat conduction equation were
made, which showed the significant increase in the microlayer thickness with the pressure reduction.
Nunes et al. analyzed the influence of different heights of square micro pin fins with an aligned
array and investigates their influence on pressure drop and heat transfer behavior. The HFE-7100 was
used as the working fluid, and the pressure drop and surface temperature behavior are analyzed for
different mass fluxes and inlet subcooling. There is good agreement between the experimental results
and the numerical analysis, with a mean absolute error of 6% for all the considered parameters. For
the two-phase flow condition, experimental tests were performed, and for the highest subcooling, an
increase in mass flux causes an enhancement in the heat transfer for low heat flux; by increasing heat
flux, there is a gradual predominance of boiling heat transfer over convection as the heat transfer
mechanism. The pressure drop drastically increases with the vapor amount flowing into the system,
regardless of the pin fin height; the boiling curves for the higher fin height show a much smaller
slope and a smaller wall superheat than the fin with the smallest height, and consequently, a high
heat transfer performance.
In the research of Vostretsov et al., the influence of parameters such as the total flow rate of
phases, the ratio of flow rates, and residence time on mass transfer during the two-phase flow of
immiscible liquids in a T-shaped microchannel were investigated using the micro-LIF technique.
The study focused on the plug flow regime, where a 70% water–glycerol solution was used as the
dispersed phase, and tri-n-butyl phosphate (TBP) was used as the carrier phase. Using the obtained
data, the extraction efficiency and overall volumetric mass transfer coefficient and established
dependencies were determined demonstrating the effect of the flow-rate ratio, total flow rate, and
the residence time on the mass transfer rate and extraction efficiency. Finally, a model for the overall
volumetric mass transfer coefficient corresponding to the set of liquids used with an R-squared value
of 0.966 was proposed.
Volodin et al. presented the experimental results of combining a basic microstructure with partly
closed pores and a mesh covering for heat transfer enhancement at the film flow of a refrigerant
mixture. All experimental series were carried out using a binary mixture of R114 and R21 refrigerants.
It was shown that a microstructured surface with a fin pitch of 200 μm, fin height of 220 μm,
and longitudinal knurling pitch of 160 μm, created using deformational cutting, demonstrates
significant heat transfer enhancement: up to four times compared to a smooth surface. However,
adding a mesh covering with an aperture of 220 μm and a wire diameter of 100 μm reduces the
intensification. The mesh covering overlaid on a smooth surface also does not provide heat transfer
enhancement as compared to the smooth surface itself. The heat transfer coefficient values obtained
for basic microstructured surfaces were compared with the dependencies available in the literature
for predicting pool boiling heat transfer on microfinned surfaces.
In the research of Misyura et al., the results of the experimental study of the effect of the droplet
evaporation rate on wall temperature in the range 20–90 °C and of the concentration of surfactant in
an aqueous solution of sodium lauryl sulfate (SLS) from 0 to 10,000 ppm are presented. It is shown
for the first time that an inversion of the evaporation rate related to the droplet diameter occurs with
increasing wall temperature. The influence of key factors on the evaporation of a water droplet with
SLS changes with temperature. Thus, at a slightly heated wall, the growth of the droplet diameter
becomes predominant. At high heat flux, the role of nonisothermality is predominant. To determine
the individual influence of the surfactant on the partial pressure of water vapor, experiments on the

viii
evaporation of a liquid layer were carried out.
The effect of coal hydrophilic particles in water-glycerol drops on the maximum diameter
of spreading along a hydrophobic solid surface is experimentally studied by Ashikhmin et al.
via analyzing the velocity of internal flows using Particle Image Velocimetry (PIV). The impact
of particle-laden drops on a solid surface occurred at Weber numbers (We) from 30 to 120. It
revealed the interrelated influence of We and the concentration of coal particles on changes in the
maximum absolute velocity of internal flows in a drop within the kinetic and spreading phases of the
drop–wall impact. The kinetic energy of the translational motion of coal particles in a spreading drop
compensates for the energy expended by the drop in sliding friction along the wall. An increase
in We contributes to more noticeable differences in the convection velocities in spreading drops.
When the drop spreading diameter rises at the maximum velocity of internal flows, a growth of
the maximum spreading diameter occurs. The presence of coal particles causes a general tendency to
reduce drop spreading.
Kotov et al. studied the effect of temperature and pressure on the nucleation of the vapor
phase and the velocity of the vapor front in the initial stage of activated boiling-up of n-pentane
on the surface of a quartz fiber. Using a developed approach combining the “pump-probe” and
laser Doppler velocimetry methods, this velocity was tracked in the course of sequential change in
the degree of superheating with respect to the liquid–vapor equilibrium line. The studied interval
according to the degree of superheating was 40–100 °C (at atmospheric pressure). An increase in
temperature at a given pressure was found to lead to an increase in the speed of the transition process
with a coefficient of about 0.2 m/s per degree, while an increase in pressure at a given temperature
leads to a decrease in the transition process speed with a coefficient of 25.8 m/s per megapascal.
The advancement of the vapor front velocity measurements to sub-microsecond intervals from
the first signs of boiling-up did not confirm the existence of a Rayleigh expansion stage with a
constant velocity.
The effect of positive (adverse) and negative (favorable) longitudinal pressure gradients on the
structure and heat transfer of gas-droplet (air and water) flow in an axisymmetric duct with sudden
expansion were examined by Pakhomov and Terekhov. The superimposed pressure gradient has a
large influence on the flow structure and heat transfer in a two-phase mist flow in both a confuser
and a diffuser. A narrowing of the confuser angle leads to significant suppression of flow turbulence
(more than four times that of the gas-drop flow after sudden pipe expansion without a pressure
gradient at ϕ = 0°). Recirculation zone length decreases significantly compared to the gas-droplet flow
without a longitudinal pressure gradient (by up to 30%), and the locus of the heat-transfer maximum
shifts slightly downstream, and roughly aligns with the reattachment point of the two-phase flow.
Growth of the diffuser opening angle leads to additional production of kinetic energy of gas flow
turbulence (almost twice as much as gas-droplet flow after a sudden pipe expansion at ϕ = 0°). The
length of the flow recirculating region in the diffuser increases significantly compared to the separated
gas-droplet flow without a pressure gradient (ϕ = 0°), and the location of maximum heat transfer
shifts downstream in the diffuser.

Vladimir Serdyukov and Fedor Ronshin


Editors

ix
energies
Article
Boiling Heat Transfer Enhancement on Biphilic Surfaces
Evgeny A. Chinnov 1 , Sergey Ya. Khmel 1 , Victor Yu. Vladimirov 1, *, Aleksey I. Safonov 1 , Vitaliy V. Semionov 1 ,
Kirill A. Emelyanenko 2 , Alexandre M. Emelyanenko 2 and Ludmila B. Boinovich 2

1 Kutateladze Institute of Thermophysics, Siberian Branch of the Russian Academy of Sciences, 1 Lavrentyev
Ave., 630090 Novosibirsk, Russia
2 A. N. Frumkin Institute of Physical Chemistry and Electrochemistry, Russian Academy of Sciences, Leninsky
Prospect 31 Bldg. 4, 119071 Moscow, Russia
* Correspondence: [email protected]; Tel.: +7-9137723236

Abstract: Flat surfaces with different patterns of hydrophobic spots were employed for experimental
investigation of boiling heat transfer. In one case, hydrophobic spots were created on a smooth
copper surface and on a surface coated with arrays of micrococoons from silicon oxide nanowires by
vapor deposition of a fluoropolymer. In the second case, a hydrophobic coating was deposited on
heater surfaces with cavity microstructures formed by laser ablation and chemisorption of fluorinated
methoxysilane. Water under saturation conditions at atmospheric pressure was used as the working
liquid. The temperature of the heating surface was varied from 100 to 125 ◦ C, and the maximum value
of the heat flux was 160 W/cm2 . Boiling heat transfer on the test biphilic surfaces was significantly
(up to 600%) higher than on non-biphilic surfaces. Surface texture, the shape of hydrophobic regions,
and the method of their creation tested in this study did not show a significant effect on heat transfer.
The boiling heat transfer rate was found to depend on the size of hydrophobic spots, the distance
between them, and hence the number of spots. The highest heat transfer efficiency was detected
for the surface with the largest number of hydrophobic spots. After long-term experiments (up to
3 years), the heat transfer coefficient on the obtained surfaces remained higher than on the smooth
copper surface. Biphilic surfaces with arrays of cavities formed by laser ablation turned out to be the
most stable during prolonged contact with boiling water.

Keywords: wettability; biphilic; surface modification; boiling; heat transfer enhancement


Citation: Chinnov, E.A.; Khmel, S.Y.;
Vladimirov, V.Y.; Safonov, A.I.;
Semionov, V.V.; Emelyanenko, K.A.;
Emelyanenko, A.M.; Boinovich, L.B.
1. Introduction
Boiling Heat Transfer Enhancement
on Biphilic Surfaces. Energies 2022, 15,
Boiling heat transfer makes it possible to remove high heat fluxes at low overheating
7296. https://doi.org/10.3390/ of the heated surface relative to the saturated vapor temperature because of the high latent
en15197296 heat of the phase transition. Boiling is widely used to cool electronic devices, nuclear
reactors, electric power generators, etc. Biphilic surfaces can be suitable for cooling smooth
Academic Editor: Marco Marengo
surfaces of chips and optical devices of high-power lasers.
Received: 3 September 2022 Boiling is a complex process that depends on many factors, in particular, the roughness
Accepted: 29 September 2022 (structure) and wettability of the surface on which this process occurs. Recently, there has
Published: 4 October 2022 been an increase in boiling studies associated with the development of new micro/nano
surface modification techniques that provide heat transfer enhancement [1–7]. Micro/nano
modification includes both micro/nano surface structuring and application of thin coatings
(nanocoatings). Both approaches change the physicochemical properties of boiling surfaces,
Copyright: © 2022 by the authors.
in particular, wettability.
Licensee MDPI, Basel, Switzerland.
It is known that nucleate boiling begins earlier on hydrophobic surfaces than on
This article is an open access article
hydrophilic surfaces because hydrophobic surfaces have more centers of vapor bubble
distributed under the terms and
nucleation. This is due to the higher gas concentration in submicron cavities on hydrophobic
conditions of the Creative Commons
Attribution (CC BY) license (https://
surfaces; in addition, after bubble detachment, a larger amount of vapor remains on
creativecommons.org/licenses/by/
hydrophobic surfaces compared to hydrophilic surfaces [8]. As a result, at low heat fluxes,
4.0/). the boiling heat transfer coefficient on a hydrophobic surface is higher than on a hydrophilic

Energies 2022, 15, 7296. https://doi.org/10.3390/en15197296 1 https://www.mdpi.com/journal/energies


Energies 2022, 15, 7296

surface. With an increase in the heat flux, the number of bubbles increases and they begin
to coalesce, which prevents liquid access to the surface and deteriorates heat transfer,
leading to the appearance of dry spots and, ultimately, to a boiling crisis. Boiling on
hydrophilic surfaces starts later than on hydrophobic surfaces; however, at increased heat
fluxes, the boiling heat transfer coefficient of hydrophilic surfaces is higher due to good
surface wettability, leading to an enhanced liquid supply to the surface. For the same
reason, the critical heat flux is higher. Thus, the use of hydrophobic surfaces is preferred at
low heat fluxes, and the use of hydrophilic surfaces at high heat fluxes.
Several approaches have been proposed to simultaneously utilize the advantages of
both types of surfaces in boiling processes. One approach is to use surfaces with mixed
(heterogeneous, patterned) wettability or biphilic surfaces. Although this approach has
been proposed previously [9,10], increased interest of researchers in the effect of biphilic
surfaces on pool boiling heat transfer enhancement arose after the publication of a study [11]
in 2010. Apparently, this is due to the above-mentioned development of new micro/nano
surface modification techniques. These techniques emerged primarily in microelectronics
and MEMS and were developed for silicon. Therefore, a significant part of the studies
of biphilic surfaces was carried out on silicon substrates [8,11–14]. Hydrophilic surfaces
were obtained, as a rule, by oxidation of silicon substrates [8,11–14]. Hydrophobic surfaces
were obtained by applying a fluoropolymer coating using the spin-coating method [11–14].
Particle deposition can be analyzed using the method presented in [15]. Surface patterns
were made by photolithography.
Pool boiling experiments have shown [11] an increase in both the heat transfer coeffi-
cient and the critical heat flux on surfaces with hydrophobic spots (regions) located on a
hydrophilic background (surrounding) surface. Apparently, a large number of vaporization
centers are formed on the hydrophobic surface of the spots, which provides rapid boiling
at low overheating and a high heat transfer coefficient. In turn, the background hydrophilic
surface between hydrophobic spots prevents bubble coalescence and increases the critical
heat flux. In addition, the liquid and vapor flows are separated, which facilitates liquid
access to the surface and contributes to an increase in the critical heat flux. Regular hexago-
nal spots were formed at the nodes of a uniform square lattice. The distance between the
centers of the spots (pitch) was varied from 50 to 200 μm, and the size of hydrophobic spots
was 40–60% of this distance.
Jo et al. [12] also investigated pool boiling heat transfer on biphilic surfaces consist-
ing of hydrophobic spots patterned on a hydrophilic background surface. The effect of
the hydrophobic spot size (50–500 μm), spot pitch (0.5 and 1 mm), and the number of
hydrophobic spots (150 and 600) was studied. It has been shown that at low heat fluxes
(<30 W/cm2 ), the heat transfer coefficient is determined by the size of hydrophobic spots
and the distance between them. At high heat fluxes, it is determined by the number of spots
and their size. It is argued that the area ratio (the ratio of the total hydrophobic surface area
to the total boiling area) has little effect on the heat transfer coefficient.
It is known that surface roughness has a significant effect on wettability; in particular,
an increase in roughness leads to an increase in both the hydrophilicity of hydrophilic
surfaces and the hydrophobicity of hydrophobic surfaces [10,16]. Betz et al. [13] used this
circumstance to create biphilic surfaces with a maximum wettability gradient. Superbiphilic
surfaces exhibited exceptional pool boiling performance with high critical heat fluxes
(>100 W/cm2 ) and very high heat transfer coefficients (10 W/cm2 K). In particular, this
coefficient was more than twice the heat transfer coefficient for biphilic surfaces with
contact angles of 20◦ and 120◦ .
Jo et al. [14] systematically studied the dependences of the critical heat flux and boiling
heat transfer on biphilic surfaces on the size and number of hydrophobic dots, pitch, and
area ratio. It was found that the critical heat flux is inversely proportional to the area ratio
and reaches maximum values for small values of this ratio. Boiling heat transfer depends
on the size and number of hydrophobic dots and the pitch between them, but weakly
depends on the area ratio. Therefore, for a significant enhancement of pool boiling heat

2
Energies 2022, 15, 7296

transfer, a biphilic surface should contain a large number of hydrophobic dots 50–100 μm
in size and at the same time should have a relatively small total hydrophobic area.
Motezakker et al. [8] studied boiling parameters on biphilic surfaces for different area
ratios. Maximum values of the critical heat flux and the heat transfer coefficient were
achieved at an area ratio of 38.46%, which corresponded to a hydrophilic spot diameter of
700 μm.
Other common materials used to fabricate substrates (heaters) when creating biphilic
surfaces are metals and, above all, copper [17–21]. In this case, the hydrophilic surface
was a metal surface or a thin film of another material deposited on metal. Hydrophobic
surfaces were obtained by applying a fluoropolymer coating or silicone. Surface patterns
were made by photolithography or other methods using a mask.
Takata et al. [17] created two types of biphilic surfaces: in one case, superhydrophobic
spots of a nickel–fluoropolymer composite were deposited on a clean copper surface; in the
other case, fluoropolymer spots were deposited on a superhydrophilic surface of titanium
oxide deposited on a copper surface. In both cases, an enhancement of water pool boiling
heat transfer was achieved. The determining parameters were the hydrophobic spot size
and the area ratio. It was shown that a decrease in the spot size at a constant pitch leads
to an increase in the heat transfer coefficient. For the same area ratio, the heat transfer
coefficient was the same regardless of the type of coating, the shape of spots, etc. The best
results were obtained for surfaces with minimum spot size and area ratio.
In a study [18], hydrophobic spots were formed from a nickel–fluoropolymer compos-
ite on a copper surface. The heat transfer coefficient was shown to be higher for smaller
spot diameters and smaller pitches between spots.
The authors of [19,20] applied an original approach to the fabrication of biphilic
surfaces aimed at minimizing their production cost. Hydrophobic spots and stripes were
formed by spreading a high-temperature silicone sealant on a copper surface through
a template. The spots were 2 and 4 mm in size with a pitch of 6 mm. Boiling heat
transfer enhancement was obtained, with the enhancement being higher on spots of smaller
diameter. However, the aging problem for such surfaces was not discussed in these papers.
Može et al. [21] created superbiphilic surfaces on an aluminum surface. Superhy-
drophobic spots were produced by CVD deposition of fluorinated silane on a nanos-
tructured metal surface, and a superhydrophilic background (surrounding) surface was
obtained by laser texturing. The heat transfer was found to significantly depend on the
pitch and area ratio but weakly on the spot size. The maximum heat transfer coefficient
was obtained for a superbiphilic surface with a spot size of 0.5 mm, a pitch of 1 mm, and an
area ratio of 23%.
Promising methods for producing biphilic surfaces are laser techniques. These meth-
ods use beams with a minimum diameter of tens of microns. Therefore, a biphilic pattern
for heat transfer enhancement can be easily achieved by scanning a substrate without
using photolithography or other mask techniques. This approach was employed in [21,22],
where certain surface areas were laser textured to obtain a biphilic pattern, and in [23],
the wettability of the applied coating was changed by laser beam heating, resulting in a
biphilic pattern.
Thus, experiments with silicon, aluminum, and copper substrates (heaters) have
shown that the use of biphilic surfaces consisting of hydrophobic spots on a hydrophilic
background surface leads to an enhancement of pool boiling heat transfer. Most stud-
ies have been performed on silicon substrates; studies on metal substrates, e.g., copper
substrates, are markedly fewer in number. In particular, metal substrates with small hy-
drophobic spot sizes (less than 250 μm) have not been investigated, although it follows
from the literature that the heat transfer enhancement is greater for smaller spot sizes.
At the same time, the findings of the previous studies are contradictory: for example,
in [12,14], it is stated that the heat transfer coefficient (HTC) depends weakly on the area
ratio, whereas in [8,21], the optimal values of the area ratio are given, but they are different:
38.46% and 23%, respectively. We also note that the aging problem is not considered in

3
Energies 2022, 15, 7296

papers on biphilic surfaces, although it is fundamental, as for any micro/nanomodified


surface [24,25]. Further improvement in boiling heat transfer on biphilic surfaces can be
achieved by preliminary micro/nanostructuring [8,13,26]; in particular, laser methods
are promising for micro/nanostructuring [21]. New interesting approaches to enhancing
boiling heat transfer have also been proposed—for example, the simultaneous use of a
biphilic surface and nanofluids. [27].
In this paper, we present the results of a study of pool boiling heat transfer on biphilic
copper surfaces with small hydrophobic spots. Spots of different shapes with an equivalent
diameter of 50 to 100 μm were used. Boiling was investigated on three types of such surfaces
with different micro/nanostructuring (micro/nanotexturing): a smooth copper surface, a
copper surface coated with micro/nanostructures in the form of arrays of micrococoons
from silicon oxide nanowires, and a copper surface with cavity microstructures formed by
laser ablation. In addition, long-term experiments were carried out to study the problem of
aging of biphilic surfaces.

2. Materials and Methods


2.1. Biphilic Surfaces and Their Fabrication Process
In this work, we used copper heaters with surface texture of three types: a smooth
surface, a surface with micro/nanostructures in the form of arrays of micrococoons from
silicon oxide nanowires, and a surface with arrays of microcavities formed by laser abla-
tion. Heaters with biphilic surfaces were fabricated by two methods. In one method, a
fluoropolymer coating was applied to the surface of heaters of the first and second types
through a mask using hot wire chemical vapor deposition (HWCVD) with the formation of
hydrophobic spots. In the other method, a hydrophobic coating was applied to the surface
of heaters of the third type by chemisorption of fluorinated methoxysilane from the vapor
phase at a temperature of 100–110 ◦ C [28,29]. Detailed information on the characteristics of
the surfaces used is given in Table 1.

Table 1. Characteristics of the surfaces used.

Surface Hydrophobic Shape and Size of Spot Pitch and Area


Type of Surface Texture
Number Coating Hydrophobic Spots Number of Spots Ratio
Micrococoons from SiOx
1 No No No 0
nanowires
Micrococoons from SiOx Fluoropolymer 0.5 mm,
2 0.1 mm diameter circle 2.8%
nanowires HWCVD 69
Micrococoons from 0.5 mm,
3 Fluoropolymer HWCVD 0.1 mm diameter circle 2.8%
nanowires SiOx 69
Fluoropolymer 0.5 mm,
4 Smooth copper surface 0.1 mm diameter circle 2.8%
HWCVD 69
Micrococoons from SiOx Fluoropolymer 0.2 mm,
5 0.05 mm diameter circle 5.1%
nanowires HWCVD 506
Fluoropolymer 0.2 mm,
6 Smooth copper surface 0.05 mm diameter circle 5.1%
HWCVD 506
Hydrophobization by Triangular cross-section
Copper surface with an
chemisorption of cavity with an 0.78 mm,
7 array of cavities formed 2.1%
fluorinated effective diameter of 32
by laser ablation
methoxysilane 0.07 mm
Hydrophobization by
Copper surface with an Rectangular cross-section
chemisorption of 0.78 mm,
8 array of cavities formed cavity with an effective 3.4%
fluorinated 32
by laser ablation size of 0.1 mm
methoxysilane

4
Energies 2022, 15, 7296

A detailed description of the HWCVD for a fluoropolymer coating is given in [30,31].


We used two masks made of stainless steel with holes 50 μm and 100 μm in diameter
located at the nodes of a square lattice with a pitch (distance between the centers of the
holes) of 200 μm and 500 μm, respectively. These masks were tightly pressed against the
heater surface during the deposition of the fluoropolymer coating. As a result, round
hydrophobic spots 50 μm in diameter with a pitch of 200 μm and 100 μm in diameter with
a pitch of 500 μm were formed on a smooth copper surface.
The second type of surface texture was obtained using arrays of micrococoons from
silicon oxide nanowires. A micrococoon is a submicron-sized tin catalyst particle covered on
all sides with silicon oxide nanowires. Tin catalyst particles were formed from a tin island
film deposited by vacuum thermal evaporation. Before the deposition of tin, the copper
heater surface was coated with a barrier layer of tungsten about 100 nm thick. Silicon
oxide nanowires on tin catalyst particles were synthesized by gas-jet electron beam plasma
chemical vapor deposition. The procedure for synthesizing arrays of micrococoons from
SiOx nanowires on copper heater surfaces is described in detail in [25]. SiOx was chosen
for the synthesis due to its high chemical and thermal stability. Such micro/nanostructures
show the greatest resistance to destruction by boiling compared to other structures [25,31].
Figure 1a shows the surface morphology with arrays of micrococoons from silicon oxide
nanowires before the deposition of the hydrophobic layer (fluoropolymer film). In Table 1,
this surface is numbered 3.

Figure 1. Surface morphology No. 3 before (a) and after the deposition of fluoropolymer (b).

Fluoropolymer film was deposited on these heaters by HWCVD, as in the previous


case, using the same masks and under the same conditions. As a result, round hydrophobic
spots 50 μm in diameter with a pitch of 200 μm and 100 μm in diameter with a pitch of
500 μm were formed on the copper heater surface with arrays of micrococoons. Figure 1b
shows the surface morphology with arrays of micrococoons after the deposition of the fluo-
ropolymer film. Features of the formation of hydrophobic spots on arrays of micrococoons
from silicon oxide nanowires are given in [32]. Figure 2 shows SEM images at different
magnifications of the surface after the deposition of fluoropolymer spots. It can be seen
that the deposition through the mask results in round fluoropolymer spots.

5
Energies 2022, 15, 7296

Figure 2. SEM images of surface No. 5 after the deposition of the fluoropolymer coating through
a 50–200 μm mask (hole diameter 50 μm, pitch 200 μm): (a) magnification 150, scale bar 300 μm,
(b) magnification 1000, scale bar 50 μm.

The elemental composition of these coatings was analyzed by EDS. Figure 3 shows
surface mapping images for different elements. Fluorine and carbon (fluoropolymer
components) are clearly detected in the region of spots (Figure 3a,b), and silicon dominates
in the region between the spots (Figure 3c). This confirms that the composition of the spots
corresponds to the fluoropolymer and that there is no fluoropolymer in the space between
the spots.


(a) (b) (c)

Figure 3. Elemental composition map of the biphilic heater surface obtained using a 50–200 μm mask:
(a) fluorine, (b) carbon, and (c) silicon. Scale bar: 200 μm.

The third type of surface texture was obtained by laser ablation with the formation of
arrays of cavities with an equivalent diameter of about 70–100 μm on the boiling surface.
For texturing, we used an Argent-M laser marking system (Laser Technology Center, Russia)
based on an infrared ytterbium fiber laser with a wavelength of 1.064 μm, a focused laser
spot diameter (at the 1/e2 level) of 40 μm, and a peak power of up to 0.95 mJ in TEM00
mode. Local surface regions were subjected to a single or multiple laser treatment with a
pulse duration of 200 ns and a pulse frequency of 20 kHz. Surfaces with arrays of cavities
were fabricated. In the single pass mode, 1568 pulses were used to fabricate one cavity and
the resulting cavities had a triangular cross section. In the multiple pass mode, 7840 pulses
were used and rectangular cross-section cavities were obtained. Figure 4 shows images of
individual triangular and rectangular cavities.

6
Energies 2022, 15, 7296

Figure 4. SEM image of an individual cavity on surface No. 7 (a) and surface No. 8 (b). Scale bar:
10 μm.

The cross-section of the triangular cavity has an equivalent diameter of about 70 μm,
and the cross-section of the rectangular cavity has an approximate size of 100 μm. For both
surfaces, the pitch was 780 μm. To create biphilic surfaces, copper surfaces were hydropho-
bized by UV/ozone activation, followed by chemisorption of fluorinated methoxysilane
from the vapor phase at a temperature of 100–110 ◦ C. This resulted in the formation
of a layer of two-dimensional chemically cross-linked fluoroxysilane which is bonded
through hydroxyl groups to the surface at ablation sites [28,29]. It was expected that the
hydrophobized layer on the smooth copper surface would be quickly destroyed, whereas in
laser-treated areas it will remain for a long time. The main factor in this process should be
different types of adsorption of fluoroxysilane molecules: stronger chemical adsorption in
laser-treated areas and less strong physical adsorption in untreated areas. Thus, a biphilic
surface is formed.
Figure 5 shows SEM images at different magnifications of the side wall of the rectan-
gular cavity. It can be seen that the laser-treated surface becomes rough and porous.

Figure 5. SEM image of the wall of an individual cavity on surface No. 8 at different magnifications.
Scale bar: (a) 1 μm, 5000 magnification; (b) 100 nm, 50,000 magnification.

7
Energies 2022, 15, 7296

The contact angle for the smooth copper surface was 54 ± 2◦ , and that for the copper
surface with the deposited fluoropolymer was 140 ± 2◦ . The contact angle for the array
of micrococoons from silicon oxide nanowires was 36 ± 2◦ , and after fluoropolymer
deposition, it was 164 ± 2◦ . For hydrophobized with fluorinated methoxysilane samples,
the contact angle was about 170 ± 2◦ in ablated regions [32]. The roughness Ra of the
smooth copper surface was no more than 0.17 μm.
The characterization of the samples was carried out using a JEOL JSM-6700F and a
Hitachi S-3400N scanning electron microscopes equipped with energy dispersive X-ray
spectrometers (EDS) for element analysis. The water contact angle was measured using a
KRUSS DSA-100 device.

2.2. Experimental Setup


The experimental setup is shown schematically in Figure 6 with an indication of the
main components. The main component of the setup is a thermostated chamber (see
Figures 7 and 8) with five optical windows for monitoring the processes occurring during
the experiments. The walls of the chamber are parallel steel plates, between which liquid
circulates. The chamber consists of three independent circuits: a box with side walls, an
upper wall (lid), and a lower wall (bottom). Each circuit includes a system of connecting
pipes and valves and is connected to a thermostat to maintain a uniform and constant
temperature distribution on the walls of the chamber. The spatial temperature distribution
is monitored by two thermocouples brought into the chamber through a common sealed
port and mounted on each wall. The chamber is filled to the required level with the
working liquid through the liquid supply channel. At the bottom of the chamber there is
a cylindrical working section with a copper core fixed with a special hold-down device.
Thermocouples used to measure and calculate the heat flux from the heater are distributed
over the working section.

Figure 6. Sketch of the experimental setup.

8
Energies 2022, 15, 7296

Figure 7. View of the thermostated chamber: 1—lateral thermal stabilization circuit,


2—upper thermal stabilization circuit, 3—lower thermal stabilization circuit, 4—working section,
5—heating element, 6—hold-down device, 7—thermocouple entry port, 8—sensor entry port, and
9—optical window.

Figure 8. View of the experimental setup.

On the upper wall of the chamber, there is a finned tube steam condenser connected
to a cooling circuit with a thermostat. A steam pressure sensor and a pressure relief valve
are also located in the upper part of the chamber. A high-speed video camera for high-
precision recording of the processes under study is placed opposite the side optical window.
An illumination system is mounted on the other side of the thermostated chamber to
provide sufficient lighting conditions of the working section during high-speed shooting.
An additional general view camera for visualization and monitoring of the experiment is
placed above the upper optical window.

9
Energies 2022, 15, 7296

The instrument and control system consisted of an NI-9214 data acquisition system and
software. The temperature measurement sensitivity of this unit is 0.1 ◦ C. The temperature
of the heater, liquid, and environment was measured using thermocouples. Thermocouples
were connected with a cold junction (no automatic cold junction compensation) to improve
accuracy. The temperature of the heater, liquid, and environment is measured using
thermocouples. All thermocouples of the setup were individually calibrated from 0 ◦ C to
150 ◦ C with an interval of 10 ◦ C. A KS 100-1 dry-well calibrator was used below 100 ◦ C,
and a Termex VT-8 glycerin thermostat in the range from 100 ◦ C to 150 ◦ C. In both cases,
temperature was monitored using a V7-99 meter with ETS-100 resistance thermometers.
The calibration error was 0.1 ◦ C. Distilled, deionized, and degassed water (Milli-Q) was
used as the working liquid.
The synthesis and deposition of nanocoatings by different methods requires the use of
sufficiently small heaters that can be placed in synthesis and deposition chambers, quickly
mounted into the setup without damage to the nanostructures on the end heating surface,
removed, and regularly monitored with an electron microscope. To perform experiments
under the same carefully controlled conditions, it is necessary to fabricate a large series of
heaters with identical characteristics.
Comparison of our method with other methods, e.g., gluing silicon wafers with
nanostructures to a heater, showed the advantages of the approach used. The heating
element is a cylindrical copper core with a head of 5 mm diameter (Figure 9). The boiling
surface is the flat top end of the cylinder. The tip of the core is tightly inserted into the hole
of the fluoropolymer base and is aligned in one plane with the top of the base. The heat
source is a Nichrome wire tightly wound around the core stem with a resistance of 4 Ohm.
The heating element is carefully insulated to minimize heat loss. Glass fabric wrapped in
several layers around the core of the heating element is used as a heat insulator.

Figure 9. Sketch (a) and photo (b) of the typical heating element (photo was created
before modifications).

Two thermocouples T1 and T2 were mounted in grooves at different depths in the


copper core (Figure 9) to measure the surface temperature and determine the heat flux
according to the Fourier law. To monitor the heat flux distribution, additional thermo-
couples T3 and T4 were mounted in the region of the core and used for winding the
Nichrome heater.
The heat flux supplied to the test section was calculated by Fourier’s law [33] as:

T2 − T1
q = −k Cu (1)
X1

10
Energies 2022, 15, 7296

where q is the heat flux, X1 is the distance between thermocouples T2 and T1, and kCu is
the thermal conductivity of copper.
The surface temperature was determined from the equation [33]:
 
X2
Tw = T1 − q (2)
k Cu

where Tw is the temperature of the test surface and X2 is the distance between the test
surface and thermocouple T1.
The main source of the measurement error was inaccuracy in determining the real heat
flux. This inaccuracy is due to the calibration error of thermocouples T1 and T2, the error in
determining the depth of grooves for thermocouple placement, and the inaccuracy of their
positioning in the grooves. The calibration error was 0.1 ◦ C, as already mentioned above.
The total positioning error of the thermocouples was about 0.01 mm. The measurement
error was estimated using the well-known formulas mentioned in [33]. The final error of
temperature determination was no more than 0.2 ◦ C. The heat flux error did not exceed
3 W/cm2 .

3. Results
3.1. Validation of the Experimental Setup and Measurement Technique
Before the main experiments, a series of test measurements was performed on smooth
copper surfaces. The results of the series for the three experiments of pool boiling on
smooth copper are shown in Figure 10.

Figure 10. Dependence of the heat flux on the temperature difference between the heating surface
and the liquid under saturation conditions for the smooth surface.

The well-known Rohsenow correlation [34] was used for comparison. For the correla-
tion, the coefficients for surfaces with the closest characteristics were used. The data are in
good agreement with the theoretical curve. The average deviation does not exceed 15%.
The second experiment has the least standard deviation (8 W/cm2 ) from the theoretical
curve. These results exceed the previously calculated experimental errors but this can be
explained by the incompletely matched surface parameters (roughness, degree of oxida-
tion). Thus, the experimental setup and the measurement technique can be considered
valid and to provide reliable results. Further, all experimental results are compared with
the Rohsenow correlation.

11
Energies 2022, 15, 7296

3.2. Experimental Results


The effect of the size of round hydrophobic spots and the pitch between them on
the boiling heat transfer rate was studied. In the first series of experiments, surfaces
with hydrophobic fluoropolymer regions 100 μm in diameter and a pitch of 500 μm were
used. A boiling curve was first obtained for a smooth copper surface with fluoropolymer
spots (surface No. 4). The result is shown in Figure 11. It can be seen from the figure
that the heat transfer on the biphilic surface is enhanced compared to the homogeneous
smooth copper surface (Rohsenow correlation). A similar result was obtained for surfaces
covered with arrays of micrococoons: boiling heat transfer on surfaces No. 2 and No. 3
with fluoropolymer spots is more intense than on surface No. 1 (a homogeneous array of
micrococoons without hydrophobic spots). Note that the results for surfaces No. 2 and
No. 3 are approximately the same, which indicates that the technology is reproducible.
Thus, biphilicity enhances heat transfer on both smooth surfaces and surfaces covered
with micrococoons from silicon oxide nanowires. It is additionally notable that the results
on smooth surface No. 4 are slightly superior to the results on surfaces No. 2 and No. 3,
despite the initial assumption that pre-coating with micrococoons would enhance boiling
heat transfer on the biphilic surface.

Figure 11. Boiling curves for surface No. 1 and surfaces Nos. 2–4 from Table 1 with deposited
fluoropolymer spots 100 μm in diameter with a pitch of 500 μm.

Surfaces with hydrophobic fluoropolymer regions 50 μm in diameter with a pitch of


200 μm exhibit a similar behavior in experiments (Figure 12). The heat transfer performance
of smooth copper surface No. 6 with hydrophobic spots is higher than that of surface No. 5
covered with arrays of micrococoons with hydrophobic spots. However, in this case, the
effect is much less pronounced and is observed at heat fluxes of about 60 to 130 W/cm2 .

12
Energies 2022, 15, 7296

Figure 12. Boiling curves for surface Nos. 5 and 6 from Table 1 with deposited fluoropolymer spots
50 μm in diameter with a pitch of 200 μm.

A comparison of the efficiency of using surfaces with fluoropolymer spots 100 μm in


diameter with a pitch of 500 μm and surfaces with spots 50 μm in diameter with a pitch of
200 μm shows a fairly significant increase of up to 20% in heat transfer rate for surfaces
with smaller spots and a smaller pitch. Moreover, this effect is observed both for surfaces
with arrays of micrococoons (Nos. 2, 3, and 5) (Figure 13) and for smooth copper surfaces
(Nos. 4 and 6) (Figure 14).

Figure 13. Comparison of boiling curves for surfaces Nos. 2 and 3 (spot diameter 100 μm, pitch
500 μm) and surface No. 5 (spot diameter 50 μm, pitch 200 μm) with arrays of micrococoons.

13
Energies 2022, 15, 7296

Figure 14. Boiling curves for surfaces Nos. 4, 6, 7, and 8.

Figure 14 shows data for surfaces No. 4 (spot diameter 100 μm, pitch 500 μm),
No. 6 (spot diameter 50 μm, pitch 200 μm), and for surfaces Nos. 7 and 8 (triangular and
rectangular cavities, respectively) obtained by laser methods.
It can be seen that the heat transfer rate on surface No. 6 is 20% higher than on the
other surfaces, including surface No. 4. Thus, in view of the data in Figure 13, it can be
concluded that for all biphilic samples with fluoropolymer spots of 50 μm in diameter
and a pitch of 200 μm, the heat transfer efficiency is higher than for samples with spots of
100 μm in diameter and a pitch of 500 μm.

4. Discussion
4.1. Effect of Surface Properties on Boiling
All biphilic surfaces show an intensity of heat transfer that is more than twice as high
(up to 600%) as a smooth copper surface. Heat transfer intensification on surfaces obtained
using a 50/200 mask (Nos. 5 and 6) is 20% higher than on the other surfaces.
Biphilic surfaces Nos. 7 and 8 are most comparable in the spot area (in the plane of
the surface) and spot pitch to surface No. 4. The heat transfer rates on these three surfaces
are almost identical, although their shape, texture, and depth of cavities are significantly
different. Apparently, the initial assumption that the hydrophobic layer is rapidly destroyed
on the main surface and has good stability in cavities obtained by ablation turns out to
be valid.
It was already mentioned above that the boiling heat transfer rate on biphilic surfaces
is significantly affected by the following factors: the size of hydrophobic spots, the distance
between them (pitch), and the area ratio (the ratio of the total hydrophobic surface area to
the total boiling area). A significant difference of surface No. 6 is that it has a much larger
(by an order of magnitude) number of hydrophobic spots: 506 versus 69 for surface No. 4
and 32 for surfaces No. 7 and No. 8. The larger area ratio for surface No. 6 (5.1%) compared
to the area ratio for surfaces No. 4 (2.8%), No. 7 (2.1%), and No. 8 (3.4%) may also affect the
heat transfer enhancement.
It can be assumed that the use of a small number of large spots is indeed less effective
since closely spaced vaporization centers in the same spot interfere with each other, and it is
worth isolating them from each other by reducing the spot size in parallel with decreasing

14
Energies 2022, 15, 7296

the spot pitch. However, in general, the question of the optimal characteristics of biphilic
surfaces remains open and requires further detailed study.

4.2. Comparison with Literature Data


A review of the literature showed that the available data on pool boiling of water on
flat copper heaters at atmospheric pressure are few in number. Systemic data on the optimal
geometry of hydrophobic regions and the optimal area ratio are absent or inconsistent. A
comparison with the results of the studies cited above is shown in Figure 15. It can only
be noted that our data on the degree of hydrophobicity are closest to those reported in the
paper by Yamada et al. [18], where hydrophobic regions occupy approximately 2.2%. It can
be seen that the data obtained for some of the most characteristic surfaces (Nos. 4, 6, and 8)
are superior to the results of Yamada.

Figure 15. Boiling curves for surfaces Nos. 4, 6, and 8 and some literature data.

4.3. Stability of the Biphilic Surfaces


An important problem in the operation of various micro/nanomodified surfaces is
their aging. During operation, many surfaces undergo morphological changes that can be
associated with mechanical destruction by boiling due to the effect of growing bubbles on
micro/nanostructures and/or due to different thermal expansion coefficients of surface
materials during layer-by-layer deposition of films of several materials. In addition, intense
hydrothermal loads can cause chemical changes on surfaces.
In the literature, insufficient attention is paid to these issues and there is no mention
of the reproducibility of experiments.
In this work, we performed a series of long-term (two-year) experiments for surfaces
Nos. 7 and 8 (with cavity microstructures formed by laser ablation), in which samples of
these surfaces demonstrated high stability. As an example, Figure 16 shows boiling curves
for surface No. 7 measured on different dates. It can be seen from these data that the pool
boiling heat transfer rate of the surface remains high throughout the period of operation
and the boiling curves do not show significant changes.

15
Energies 2022, 15, 7296

Figure 16. Boiling curves for surface No. 7 showing its stability. DD/MM/YYYY format was used.

Surfaces obtained by local deposition of the fluoropolymer through a mask onto


smooth copper did not exhibit high stability compared to surfaces No. 7 and No. 8. As an
example, Figure 17 shows some data for surface No. 4. After the best result obtained on
7 June 2019, the heat transfer rate decreased in the subsequent experiments.

Figure 17. Boiling curves for surface No. 4.

Surfaces obtained by local deposition of the fluoropolymer on an array of micrococoons


also did not exhibit high stability, despite the initial assumptions that the adhesion of the
fluoropolymer deposited on microstructures would be higher than in the case of deposition
on smooth copper. Some data for surface No. 3 are given in Figure 18. Just as for surface
No. 4, the heat transfer rate decreased after the first experiments.

16
Energies 2022, 15, 7296

Figure 18. Boiling curves for surface No. 3.

The standard deviations between the experimental curves in both cases exceed
11 W/cm2 , which is much larger than the experimental error.

5. Conclusions
1. Boiling experiments were carried out using two types of biphilic surfaces obtained
by different techniques. The degrees of heat transfer enhancement on nano-modified
surfaces were determined. The effect of the size (0.05, 0.07, 0.1 mm), shape (circle,
triangular, rectangular), and number of hydrophobic spots (32 to 506) on the boiling
heat transfer rate on eight copper surfaces was studied.
2. Both techniques demonstrated their effectiveness and provided a significant en-
hancement of boiling heat transfer. The heat transfer coefficient on biphilic surfaces
was significantly (up to 600 %) higher than on smooth surfaces or surfaces coated
with cocoons.
3. The boiling heat transfer rate depended on the size of hydrophobic spots, the dis-
tance between them, and hence on their number. The results complement previous
data [7,18] and show that heat transfer enhancement can be obtained on surfaces
with a small diameter of hydrophobic regions (0.05 mm) and a relatively small area
ratio (5%). At the same time, no significant effect of surface texture and shape of
hydrophobic regions tested in this study on heat transfer was found.
4. After long-term experiments (over 1 to 3 years), the heat transfer coefficient on the
obtained surfaces remained higher than on a smooth copper surface. During 3 years
of operation, the best stability of enhanced heat transfer was observed for biphilic
surfaces with cavities formed by laser ablation.

Author Contributions: Conceptualization, E.A.C. and S.Y.K.; funding acquisition, E.A.C.; investiga-
tion, V.Y.V. and V.V.S.; methodology, S.Y.K., A.I.S., K.A.E., A.M.E. and L.B.B.; project administration,
E.A.C.; resources, S.Y.K., A.I.S., K.A.E., A.M.E. and L.B.B.; supervision, E.A.C.; validation, V.Y.V. and
V.V.S.; visualization, S.Y.K. and V.Y.V.; writing—original draft, E.A.C., S.Y.K. and V.Y.V.; writing—
review and editing, E.A.C., S.Y.K., V.Y.V., K.A.E., A.M.E. and L.B.B. All authors have read and agreed
to the published version of the manuscript.
Funding: This work was funded by the Russian Science Foundation (grant number 22-19-20090) and
the Government of the Novosibirsk Region (agreement number p-13).

17
Energies 2022, 15, 7296

Institutional Review Board Statement: Not applicable.


Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design
of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or
in the decision to publish the results.

Nomenclature
q—heat flux, W/cm2 ; kCu —thermal conductivity of copper, W/(m*K); T—temperature, K, ◦ C;
X—distance, mm; Contact angle, ◦ .

References
1. Liang, G.; Mudawar, I. Review of pool boiling enhancement by surface modification. Int. J. Heat Mass Transf. 2019, 128, 892–933.
[CrossRef]
2. Li, W.; Dai, R.; Zeng, M.; Wang, Q. Review of two types of surface modification on pool boiling enhancement: Passive and active.
Renew. Sustain. Energy Rev. 2020, 130, 109926. [CrossRef]
3. Chen, J.; Ahmad, S.; Cai, J.; Liu, H.; Lau, K.T.; Zhao, J. Latest progress on nanotechnology aided boiling heat transfer enhancement:
A review. Energy 2021, 215, 119114. [CrossRef]
4. Kamel, M.S.; Lezsovits, F.; Hussein, A.K. Experimental studies of flow boiling heat transfer by using nanofluids A critical recent
review. J. Therm. Anal. Calorim. 2019, 138, 4019–4043. [CrossRef]
5. Surtaev, A.S.; Serdyukov, V.S.; Pavlenko, A.N. Nanotechnologies for thermophysics: Heat transfer and crisis phenomena at
boiling. Nanotech. Russia 2016, 11, 696–715. [CrossRef]
6. Dedov, A.V. A Review of Modern Methods for Enhancing Nucleate Boiling Heat Transfer. Therm. Eng. 2019, 66, 881–915.
[CrossRef]
7. Serdyukov, V.; Patrin, G.; Malakhov, I.; Surtaev, A. Biphilic surface to improve and stabilize pool boiling in vacuum. Appl. Therm.
Eng. 2022, 209, 118298. [CrossRef]
8. Motezakker, A.R.; Sadaghiani, A.K.; Çelik, S.; Larsen, T.; Villanueva, L.G.; Koşar, A. Optimum ratio of hydrophobic to hydrophilic
areas of biphilic surfaces in thermal fluid systems involving boiling. Int. J. Heat Mass Transf. 2019, 135, 164–174. [CrossRef]
9. Hummel, R.L. Means for Increasing the Heat Transfer Coefficient between a Wall and Boiling Liquid. U.S. Patent No. 3207209,
21 September 1965. Available online: https://patents.google.com/patent/US3207209A/en (accessed on 5 July 2022).
10. Attinger, D.; Frankiewicz, C.; Betz, A.R.; Schutzius, T.M.; Ganguly, R.; Das, A.; Kim, C.-J.; Megaridis, C.M. Surface engineering for
phase change heat transfer: A review. MRS Energy Sustain. 2014, 1, 4. [CrossRef]
11. Betz, A.R.; Xu, J.; Qiu, H.; Attinger, D. Do surfaces with mixed hydrophilic and hydrophobic areas enhance pool boiling? Appl.
Phys. Lett. 2010, 97, 141909. [CrossRef]
12. Jo, H.; Ahn, H.S.; Kang, S.; Kim, M.H. A study of nucleate boiling heat transfer on hydrophilic, hydrophobic and heterogeneous
wetting surfaces. Int. J. Heat Mass Transf. 2011, 54, 5643–5652. [CrossRef]
13. Betz, A.R.; Jenkins, C.-J.; Kim, C.-J.; Attinger, D. Boiling heat transfer on superhydrophilic, superhydrophobic, and superbiphilic
surfaces. Int. J. Heat Mass Transf. 2013, 57, 733–741. [CrossRef]
14. Jo, H.J.; Kim, S.H.; Park, H.S.; Kim, M.H. Critical heat flux and nucleate boiling on several heterogeneous wetting surfaces:
Controlled hydrophobic patterns on a hydrophilic substrate. Int. J. Multiph. Flow 2014, 62, 101–109. [CrossRef]
15. Colombo, L.; Alberti, L.; Mazzon, P.; Antelmi, M. Null-space Monte Carlo particle backtracking to identify groundwater
tetrachloroethylene sources. Front. Environ. Sci. 2020, 8, 142. [CrossRef]
16. Wenzel, R.N. Resistance of solid surfaces to wetting by water. Ind. Eng. Chem. 1936, 28, 988–994. [CrossRef]
17. Takata, Y.; Hidaka, S.; Kohno, M. Effect of surface wettability on pool boiling: Enhancement by hydrophobic coating. Int. J.
Air-Cond. Refrig. 2012, 20, 1150003. [CrossRef]
18. Yamada, M.; Shen, B.; Imamura, T.; Hidaka, S.; Kohno, M.; Takahashi, K.; Takata, Y. Enhancement of boiling heat transfer under
sub-atmospheric pressures using biphilic surfaces. Int. J. Heat Mass Transf. 2017, 115, 753–762. [CrossRef]
19. Sarode, A.; Raj, R.; Bhargav, A. Scalable macroscale wettability patterns for pool boiling heat transfer enhancement. Heat Mass
Transf. 2020, 56, 989–1000. [CrossRef]
20. Liang, G.; Chen, Y.; Wang, J.; Wang, Z.; Shen, S. Experiments and modeling of boiling heat transfer on hybrid-wettability surfaces.
Int. J. Multiph. Flow 2021, 144, 103810. [CrossRef]
21. Može, M.; Zupančič, M.; Golobič, I. Pattern geometry optimization on superbiphilic aluminum surfaces for enhanced pool boiling
heat transfer. Int. J. Heat Mass Transf. 2020, 161, 120265. [CrossRef]
22. Xia, Y.; Gao, X.; Li, R. Surface effects on sub-cooled pool boiling for smooth and laser-ablated silicon surfaces. Int. J. Heat Mass
Transf. 2022, 194, 123113. [CrossRef]

18
Energies 2022, 15, 7296

23. Zupančič, M.; Steinbücher, M.; Gregorčič, P.; Golobič, I. Enhanced pool-boiling heat transfer on laser-made hydropho-
bic/superhydrophilic polydime thylsiloxane-silica patterned surfaces. Appl. Therm. Eng. 2015, 91, 288–297. [CrossRef]
24. Može, M. Effect of boiling-induced aging on pool boiling heat transfer performance of untreated and laser-textured copper
surfaces. Appl. Therm. Eng. 2020, 181, 116025. [CrossRef]
25. Khmel, S.; Baranov, E.; Vladimirov, V.; Safonov, A.; Chinnov, E. Experimental study of pool boiling on heaters with nanomodified
surfaces under saturation. Heat Transf. Eng. 2021. Available online: https://www.tandfonline.com/doi/full/10.1080/01457632.2
021.2009211(accessed on 5 July 2022).
26. Cho, H.R.; Park, S.C.; Kim, D.; Joo, H.-m.; Yu, D.I. Experimental study on pool boiling on hydrophilic micro/nanotextured
surfaces with hydrophobic patterns. Energies 2021, 14, 7543. [CrossRef]
27. Freitas, E.; Pontes, P.; Cautela, R.; Bahadur, V.; Miranda, J.; Ribeiro, A.P.C.; Souza, R.R.; Oliveira, J.D.; Copetti, J.B.; Lima, R.; et al.
Pool boiling of nanofluids on biphilic burfaces: An experimental and numerical study. Nanomaterials 2021, 11, 125. [CrossRef]
28. Boinovich, L.B.; Emelyanenko, A.M. The behaviour of fluoro-and hydrocarbon surfactants used for fabrication of superhydropho-
bic coatings at solid/water interface. Coll. Surf. A Phys. Eng. Asp. 2015, 481, 167–175. [CrossRef]
29. Boinovich, L.B.; Emelyanenko, K.A.; Domantovsky, A.G.; Chulkova, E.V.; Shiryaev, A.A.; Emelyanenko, A.M. Pulsed laser
induced triple layer copper oxide structure for durable polyfunctionality of superhydrophobic coatings. Adv. Mater. Interfaces
2018, 5, 1801099. [CrossRef]
30. Safonov, A.I.; Sulyaeva, V.S.; Gatapova, E.Y.; Starinskiy, S.V.; Timoshenko, N.I.; Kabov, O.A. Deposition features and wettability
behavior of fluoropolymer coatings from hexafluoropropylene oxide activated by NiCr wire. Thin Solid Films 2018, 653, 165–172.
[CrossRef]
31. Chinnov, E.A.; Shatskiy, E.N.; Khmel, S.Y.; Baranov, E.A.; Zamchiy, A.O.; Semionov, V.V.; Kabov, O.A. Enhancement of heat
transfer at pool boiling on surfaces with silicon oxide nanowires. J. Phys. Conf. Series 2017, 925, 012033. [CrossRef]
32. Vladimirov, V.Y.; Khmel, S.Y.; Safonov, A.I.; Semionov, V.V.; Chinnov, E.A. Effect of fluoropolymer spots on pool boiling heat
transfer. J. Phys. Conf. Series 2021, 2119, 012087. [CrossRef]
33. Može, M.; Zupančič, M.; Golobič, I. Investigation of the scatter in reported pool boiling CHF measurements including analysis of
heat flux and measurement uncertainty evaluation methodology. App. Therm. Eng. 2020, 169, 114938. [CrossRef]
34. Rohsenow, W.M. A Method of Correlating Heat Transfer Data for Surface Boiling of Liquids; Technical Report No 5 (Heat Transfer
Laboratory); MIT Division of Industrial Cooperation: Cambridge, MA, USA, 1951. Available online: https://dspace.mit.edu/
handle/1721.1/61431 (accessed on 5 July 2022).

19
energies
Article
Lattice Boltzmann Simulation of Optimal Biphilic Surface
Configuration to Enhance Boiling Heat Transfer
Alexander V. Fedoseev 1, *, Mikhail V. Salnikov 1 , Anastasiya E. Ostapchenko 1 and Anton S. Surtaev 1,2

1 Institute of Thermophysics SB RAS, Lavrentyev Ave. 1, Novosibirsk 630090, Russia


2 Physics Department, Novosibirsk State University, Pirogova Str. 2, Novosibirsk 630090, Russia
* Correspondence: [email protected]

Abstract: To study the processes of boiling on a smooth surface with contrast wettability, a hybrid
model was developed based on Lattice Boltzmann method and heat transfer equation. The model
makes it possible to describe the phenomena of natural convection, nucleate boiling, and transition to
film boiling, and, thus, to study heat transfer and the development of crisis phenomena in a wide
range of surface superheats and surface wetting characteristics. To find the optimal configuration of
the biphilic surface, at the first stage a numerical simulation was carried out for a single lyophobic
zone on a lyophilic surface. The dependences of the bubble departure frequency and the departure
diameter of the bubble on the width of the lyophobic zone were obtained, and its optimal size was
determined. At the next stage, the boiling process on an extended surface was studied in the presence
of several lyophobic zones of a given size with different distances between them. It is shown that
in the region of moderate surface superheat, the intensity of heat transfer on biphilic surfaces can
be several times (more than 4) higher compared to surfaces with homogeneous wettability. Based
on numerical calculations, an optimal configuration of the biphilic surface with the ratios of the
lyophobic zones’ width of the order of 0.16 and the distance between the lyophobic zones in the range
of 0.9–1.3 to the bubble departure diameter was found.

Keywords: pool boiling; heat transfer enhancement; biphilic surface; lattice Boltzmann method

Citation: Fedoseev, A.V.; Salnikov, 1. Introduction


M.V.; Ostapchenko, A.E.; Surtaev, A.S.
To date, issues related to the enhancement of heat transfer during liquid boiling have
Lattice Boltzmann Simulation of
received great attention from scientists and engineers, which is confirmed by numerous
Optimal Biphilic Surface
papers and reviews published in the past few years [1–5]. This is due to a wide range of
Configuration to Enhance Boiling
practical applications in which the boiling process plays a key role. One of the most popular
Heat Transfer. Energies 2022, 15, 8204.
and effective methods for the enhancement of the heat transfer and increasing the critical
https://doi.org/10.3390/en15218204
heat flux (CHF) during boiling is the modification of the heat exchange surface. Surface
Academic Editor: Marco Marengo modification can be implemented both by structuring the original surface using, for exam-
Received: 12 October 2022
ple, mechanical machining or laser texturing, and by fabrication of micro/nanostructured
Accepted: 1 November 2022
porous coatings [5]. As a result of heat exchange surface modifications, not only the mor-
Published: 3 November 2022
phology, but also the wetting characteristics can change. It is known that wetting properties
have a significant effect on the heat transfer intensity and the development of crisis phenom-
ena during liquid boiling. For example, the use of superhydrophilic micro- or nanoporous
surfaces leads to a significant increase in the CHF value at parameters corresponding to
Copyright: © 2022 by the authors. water boiling (up to 2–3 times) [6,7]. In turn, the use of weakly hydrophobic coatings leads
Licensee MDPI, Basel, Switzerland. to a decrease in the temperature of the onset of nucleate boiling (ONB), to an increase in
This article is an open access article the nucleation site density and the heat transfer enhancement in the regions of low heat
distributed under the terms and
fluxes [8–11]. However, the CHF significantly decreases when using hydrophobic and
conditions of the Creative Commons
superhydrophobic coatings [11–13]. In particular, in [10,11] during water boiling on flat
Attribution (CC BY) license (https://
and tubular hydrophobic heat exchange surfaces (contact angle θ ≈ 125◦ ), heat transfer
creativecommons.org/licenses/by/
enhancement was found in the region of low heat fluxes (q < 100–150 kW/m2 ) compared to
4.0/).

Energies 2022, 15, 8204. https://doi.org/10.3390/en15218204 20 https://www.mdpi.com/journal/energies


Energies 2022, 15, 8204

heat transfer on the base hydrophilic surface. Moreover, the use of the developed hydropho-
bic coatings can significantly reduce the temperature of the onset of water boiling (up to
2–3◦ K), especially during boiling at subatmospheric pressures [11]. However, the CHF de-
creases significantly when using hydrophobic and superhydrophobic coatings [11–13]. For
example, in an experimental study [12] it was shown that the usage of a superhydrophobic
heated surface (contact angle θ ≈ 165◦ ), vapor film is almost immediately formed at 1–2◦ K
of superheating, covering the entire surface, thus the heat transfer is reduced just after the
onset of boiling.
One of the modern trends in the field of material science in relation to the problems
of improvement of pool boiling performance is the fabrication of biphilic surfaces or, in
other words, surfaces with contrast wettability. The use of biphilic surfaces, as shown by
the analysis of recent papers [14–17], permits simultaneously increasing the heat transfer
intensity by reducing the bubbles activation threshold and increasing CHF due to more
efficient wetting of dry spots in precrisis modes and modulation of the two-phase flows
near the heated wall. In particular, it was shown in experiments [14,15] that the use of
biphilic coatings can lead to an increase in the heat transfer coefficient and CHF values up
to 5 and 2 times, respectively. At the same time, the task of determination of the biphilic
surface configuration for the optimal boiling conditions remains open. These configurations
vary for different liquids and boiling regimes including pressure change.
The search for the optimal configuration of a surface with contrast wettability based
on experimental studies is a rather laborious task, the complexity of which is primarily
determined by the creation of stable coatings with the ability to control the geometric
parameters of the structure at various scales, including micro and nanoscales. An additional
complication is the requirement for high-precision high-speed measurements to study
the evolution of vapor bubbles and a two-phase layer near the surface using a set of
modern experimental methods. With the rapid development of computer technology,
computational fluid dynamics (CFD) methods have proven to be a good alternative to
experiment for studying two-phase flows, including those in which phase transitions
occur [18–26]. The classical approach to modeling two-phase flows, including boiling, is
reduced to the numerical solution of the Navier–Stokes equations in combination with an
additional method for tracking the two-phase interface. These methods include the level
set method [19–22], the volume of liquid (VOF) method [23,24], or the combined VOSET
method [25,26]. A common problem with these methods is that they do not allow modeling
of the nucleation process, which should be considered at the micro or nano scale. This
process is crucial for modeling the life cycle of vapor bubbles, in particular, the stage of
waiting for the appearance of nuclei, the temperature of the onset of boiling, and the density
of nucleation sites. Without the ability to model these characteristics, it is impossible to
calculate the boiling curves for an ensemble of bubbles that form on extended surfaces.
To simulate micro-nanoscale phenomena, such as the process of nucleation, the molecular
dynamics method is well applicable [27–29]. However, its application on a macro scale is
difficult due to usually limited computational power.
Lattice Boltzmann Method (LBM) implies a different approach to multiphase flows
simulation. In this method, the medium is represented in the form of ensembles of pseu-
doparticles, for which the kinetic equation is solved in the discrete formulation. Only
a small set of pseudoparticle velocities is possible, such that the velocities are directed
towards neighboring nodes of the spatial lattice. This method belongs to the mesoscopic cat-
egory, and occupies an intermediate position between the microscopic method of molecular
dynamics and the macroscopic CFD approaches. Despite the complexity of the Boltzmann
kinetic equation in the classical formulation, its discrete counterpart is simple and ideally
suited for parallel computing. Over 30 years of LBM development, it has proven itself
well for modeling both single-phase and multi-phase flows [30,31]. In [32], the LBM with a
pseudopotential was presented, in which the interaction function between pseudoparticles
is specified. With this approach, phase separation is obtained in a natural way as a conse-
quence of the forces of interaction of pseudoparticles at the lattice sites. The action of body

21
Energies 2022, 15, 8204

forces can be specified in different ways, of which the Guo method [33], He method [34]
and the exact difference method (EDM) [35] are the most effective. Another important
advantage of LBM is the way of specifying the condition at the fluid-solid boundary, the
so-called bounce-back condition [36]. This makes it possible to simply set a solid boundary
of an arbitrary shape [36], as well as to set an arbitrary local value of the wetting angle [37].
The advantages stated above made it possible to use LBM to model and study the pool
boiling processes. Using this method, it is possible to simulate the process of vapor phase
nucleation without the need to specify additional initial conditions obtained empirically.
The authors of [38,39] used LBM to simulate the processes of growth and detachment of
single bubbles during boiling. The dependence of the departure diameter and the frequency
of bubble departure on the wetting contact angle and the Jacob number was studied.
In [40–44], boiling curves were calculated for smooth surfaces with different uniform
wettability. It was shown that the boiling crisis for a lyophobic heater occurs at lower
superheats than for a lyophilic one, and an increase in the wetting contact angle leads to a
decrease in the CHF. Recently, works have been presented on the LBM simulation of heat
transfer processes during boiling on surfaces with contrast wettability [45,46]. The papers
simulated boiling on a lyophilic surface with the inclusion of lyophobic zones. The authors
demonstrated the possibility of increasing the heat flux compared to a homogeneous
lyophilic surface.
The purpose of the paper is to develop a hybrid model based on LBM and heat transfer
equations capable of simulating the process of boiling on surfaces with contrast wettability
in a wide range of surface superheats up to the boiling crisis development. The model will
permit obtaining the boiling curves for heat exchange surfaces with homogeneous and
heterogeneous wettability, to study the effect of the size of lyophobic zones on a lyophilic
base surface and of the distance between them, and to determine the optimal biphilic
surface configuration for efficient heat transfer.

2. Model
For the study of pool boiling processes, a hybrid model based on the standard pseu-
dopotential multiphase Lattice Boltzmann method coupled with the heat transfer equation
is used. Below a brief overview of the model is presented while a more detailed description
can be found in [39,41]. The basic equation for the time evolution of the density distribution
function fi in a discrete form can be presented in the form:
→ → → → →
f i ( x + c i Δt, t + Δt) = f i ( x , t) + Ωi ( x , t) + Si ( x , t), (1)
→ →
where x is the position and t is the dimensionless time, c i are the unit lattice vectors, Ωi is
the collision operator, which describes the particles’ interaction, and Si is the change of the
distribution function due to the action of the bulk forces. In the paper, a two dimensional
→ √
D2Q9 lattice with nine directions is used: | c i | = (0, 1, 2), i = {1, 2, . . . , 5, 6, . . . , 9}.

The density ρ and the velocity u of the medium in each point can be obtained from the
moments of the distribution function:
→ → → → → →
ρ( x , t) = ∑ f i ( x , t ), ρ u ( x , t) = ∑ c i f i ( x , t ). (2)
i i

The Bhatnagar–Gross–Krook (BGK) approximation for a collision operator is used:

→ f i − f i eq
Ωi ( x , t ) = − . (3)
τ
where f i eq is the equilibrium distribution function, τ is the relaxation time connected with
the kinematic viscosity ν as τ = 0.5 − ν/c2s , and where cs 2 = (1/3)Δx2 /Δt2 is the lattice
speed of sound; Δx and Δt are the spatial and time steps.

22
Energies 2022, 15, 8204

For the body force term, the exact difference method [35] is used:
→ → →
Si = f i eq (ρ, u + Δ u ) − f i eq (ρ, u ) (4)

→ → →
where the change of velocity u is determined by the force acting on the node: Δ u = F /ρ.

The total force F acting on the particles consists of the following components:
→ → → → → → → →
F ( x ) = F int ( x ) + F s ( x ) + F g ( x ), (5)
→ → → →
where F int ( x ) is the particles’ interaction force responsible for phase separation, F s ( x )
→ →
describes the interaction of the medium with the solid surface, and F g ( x ) = g(ρ − ρ avg ) is
the gravity force, where ρ avg is the average density of the liquid/vapor medium; g is free
fall acceleration.
→ →
The equation of state (EOS) of the simulated fluid is defined by F int ( x ) written as a
gradient of a pseudopotential Ψ as proposed in [32]. The gradient of the pseudopotential
is approximated with the finite-difference scheme [35]. In the paper, the Peng–Robinson
equation of state [47] is used to determine the pressure dependence P(ρ,T) on the density ρ
and the temperature T of the liquid/vapor medium.
In the absence of the fluid-solid interaction (adhesive force), a liquid-solid contact
angle θ is equal to 90◦ . To adjust the value of the contact angle, an approach proposed
→ →
in [37] is used, and an extra force F s ( x ) is added to the nodes near the solid boundary.
→ →
The contact angle is modeled by varying the strength and direction of the force F s ( x ) as
described in detail in [37,41]. In the present paper, the contact angles θ = 38◦ , 90◦ and 116◦
are used for the simulation of the lyophilic, neutral and lyophobic surfaces.
Evolution of the heat transfer in the computational region is described by the heat
conduction equation taking into account diffusion, convection, work of pressure forces, as
well as phase transition [41,42]. In the model, the dimensionless parameters in the lattice
units are used. The values of pressure P, temperature T and density ρ are expressed in
units Pc , Tc and ρc , which correspond to the parameters of the fluid at the critical point. In
dimensionless form this equation reads:
   
∂T → → 1 → → Δt T Pc ∂p EOS → →
+ u f ∇T = ∇ λ∇ T − ∇ · u f, (6)
∂t ρc ρcv Δx2 ρc ρcv Tc ∂T ρ

where cv and λ are the heat capacity and thermal conductivity coefficient. Here,
→ → →
u f = u + Δ u /2 is the physical velocity of the continuum. The last term in the equa-
tion implicitly accounts for the latent heat of vaporization, which is the main source of
enhancement of the heat flux under the onset of nucleate boiling.
The computational domain consists of a liquid/vapor medium and a metal heater (see
Figure 1). In the paper, 800 spatial cells in a horizontal direction (axis x) and 500 spatial cells
in a vertical direction (axis y) are used. The metal heater with a thickness of nh = 30 spatial
cells is placed on the bottom boundary. Spatial and time steps are equal to Δx = 20 × 10−6 m
and Δt = 2.5 × 10−6 s, respectively. Periodic boundary conditions are applied to the left
and the right boundaries. On the top boundary, a constant temperature T0 = 0.9 Tc and
corresponding pressure P0 are specified. Inside the metal heater, only heat diffusion
Equation (6) is solved, with constant temperature Th condition imposed at the bottom. The
calculations were performed for different values of superheat ΔT, i.e., the temperature
difference between the top and the bottom boundaries of the solution region, ΔT = Th − T0 .
Thermal conductivity and heat capacity of the metal heater are set to λh = 20 W/m/K
and ch = 3 × 106 J/m3 , respectively. The thermodynamic properties of the fluid, e.g., heat
capacity, thermal conductivity, and viscosity are evaluated based on the current density of
the liquid/vapor medium and are parameterized based on the properties of the liquid and
its vapor.

23
Energies 2022, 15, 8204

Figure 1. Computational domain and boundary conditions. Layout of the lyophobic zones on
the lyophilic surface: Dphob is the width of the lyophobic zones, L is the distance between the
lyophobic zones.

The phase diagram for some instant at boiling on a biphilic surface is also presented in
Figure 1. Areas with the gas phase are shown in blue while the liquid phase is shown in red.
The distance between the neighbor lyophobic zones is L and the width of each lyophobic
zone is Dphob .

3. Results
3.1. Boiling on the Surfaces with Homogenous Wettability
At the initial stage, the boiling curves <q>(ΔT  ) were calculated for the smooth sur-
faces with homogeneous wettability including lyophobic and lyophilic states, where
ΔT  = <Tw > − Tsat is the surface superheat and Tsat is the saturation temperature (Figure 2).
Average heater surface temperature <Tw > is obtained by averaging the temperature Tw (x,t)
over time and the heater surface. In calculations, to obtain each point of the dependence
q(ΔT  ), temperature value is set on the lower wall of the heater Th = const. The heater
surface temperature Tw (x,t) depends on the heat transfer processes and evolves according to
the Equation (6), see [41]. In turn, the time and the surface averaged heat flux <q> through
the metal heater of the height Hh = nh h is determined as <q> = −λh (<Tw > − Th )/Hh .
As can be seen from the data presented in Figure 2, at low superheats, the boiling
curves q(ΔT  ) do not depend on the value of the wetting angle, since the heat transfer in
this region occurs in the mode of single-phase natural convection. After the onset of boiling,
the slope of the q(ΔT  ) curves sharply increases, which corresponds to an increase in the
intensity of heat transfer with the transition from natural convection to nucleate boiling.
It can be seen that the temperature corresponding to ONB (onset of nucleate boiling)
noticeably decreases with increasing contact angle, which is in qualitative agreement
with the results of experimental studies [8–13]. The analysis of the curves in Figure 2
shows that the heat transfer depends in a complex way on the surface superheats and
surface wettability. For example, in the region of low surface superheat, the removal
heat flux increases with increasing contact angle, which qualitatively agrees with the
results of experiments [9–11], in which it was shown that for weakly hydrophobic surfaces
(θ = 116–130◦ ) at low input heat fluxes, an enhancement of heat transfer is observed in
comparison with hydrophilic samples. However, as calculations show, the situation changes
dramatically in the area of high wall superheats. It can be seen that the value of the CHF
decreases significantly with an increase in the contact angle, which also qualitatively agrees
with the experimental results [11–13]. Moreover, the boiling crisis on the lyophobic surface
can be observed even before the onset of nucleate boiling on the lyophilic surface at a given
wall superheat.

24
Energies 2022, 15, 8204

Figure 2. Boiling curves q(ΔT  ) for the smooth surfaces with homogeneous wettability: lyophilic
surface (θ = 38◦ ), neutral surface (θ = 90◦ ), lyophobic surface (θ = 116◦ ).

Figure 3 shows density contour plots of boiling at low wall superheat (ΔT = 0.045 Tc )
on surfaces with different contact angles. As can be seen from the pictures, at low surface
superheats the boiling on homogeneous lyophobic surface (θ = 116◦ ) resembles a transi-
tional boiling regime, when part of the surface is occupied by vapor phase with local sites
of film boiling. In general, this picture is similar to the results of experimental studies,
in which it was shown that boiling on weakly hydrophobic surfaces in the region of low
heat fluxes is characterized by the presence of large-scale sessile bubbles with periodic
separation of the vapor phase from liquid-vapor interface [10–12]. It has also been shown
in the vast majority of experiments [10–12,48,49] that for lyophobic surfaces, the angle of
inclination of the liquid meniscus with respect to the solid body exceeds 90◦ , and the vapor
bubble detaches at some distance from the heater surface. This leads to the fact that after
the departure of the vapor bubble, part of the vapor always remains on the surface, which
leads to the degeneration of the waiting stage of the appearance of a bubble in the life cycle
of the nucleation site. Indeed, the results of LBM simulation on lyophobic surfaces (θ ≥ 90◦ )
demonstrate a qualitatively similar behavior of the triple contact line during evolution of
vapor bubbles on the heating surface. At the same wall superheat, as can be seen form
Figure 3c, the heat transfer on the lyophilic surface (θ = 38◦ ) occurs in the free convection
mode without activation of nucleation sites.

Figure 3. The density contour plots illustrating the boiling process at low surface superheat,
ΔT = 0.045 Tc . (a) lyophobic surface (θ = 116◦ ), (b) neutral surface (θ = 90◦ ) (c) lyophilic surface
(θ = 38◦ ).

25
Energies 2022, 15, 8204

With an increase in wall superheat ΔT = 0.06 Tc , the boiling behavior on surfaces with
different wettability changes (Figure 4). As can be seen from Figure 4a, in this case for
lyophobic surface, the stable film boiling mode is observed, and the entire surface of the
heater is covered with a vapor phase with periodic detachment of vapor bubbles from the
film surface. For a neutral surface (Figure 4b), with an increase in the input heat flux, several
nucleation sites can simultaneously be activated on the surface, which also leads to an
increase of vaporization rate compared to the case of low superheat. However, large bubbles
with the area bounded by the triple contact line larger than the bubble departure diameter
may also appear. For a lyophilic surface at given superheat (Figure 4c), vapor bubbles are
already activated in comparison with the case shown in Figure 3c. However, the boiling
process seems to have a periodic character, when several nucleation sites are activated
at once, and after the detachment of vapor bubbles, a long waiting stage is observed.
This leads to noticeable pulsations in the space averaged wall temperature over time. This
character of boiling can be related to the fact that an absolutely smooth surface is specified in
the simulation model, when there are no cavities which are potential sites for the activation
of vapor bubbles. At the same time, the evolution of individual bubbles qualitatively
corresponds to experimental observations of the dynamics of growth and detachment of
bubbles during boiling of a wide class of liquids on lyophilic surfaces [19–22,38,39].

Figure 4. The density contour plots illustrating the boiling process at moderate surface superheat,
ΔT = 0.06 Tc . (a) lyophobic surface (θ = 116◦ ), (b) neutral surface (θ = 90◦ ) (c) lyophilic surface
(θ = 38◦ ).

3.2. Bubble Dynamic at Boiling above a Single Lyophobic Zone


In our opinion, the choice of the configuration of the biphilic pattern surface is de-
termined primarily by the selection of the optimal sizes of the lyophobic zones and the
distances between them on the lyophilic base. Therefore, at the first stage, the process
of vapor bubble growth and detachment for a single lyophobic zone with different sizes
(contact angle θ = 116◦ ) located in the center of the smooth lyophilic surface of the heater
(θ = 38◦ ) was simulated. In LBM simulation, the following range of widths of the lyophobic
zone Dphob = 0.3–3 mm was considered, which is noticeably larger than the critical radius of
vapor nuclei and smaller or comparable with the bubble departure diameter. Moreover,
this range of sizes of the lyophobic zone is, as a rule, used to create biphilic surfaces in
experimental works [14–17].
Figure 5 shows the characteristic temporal evolution of a single bubble for a given
width of the lyophobic zone Dphob = 1.5 mm and a heater temperature Th = 0.95 Tc . The size
of the computational domain was 400 cells horizontally (12 mm) and 600 cells vertically
(18 mm). In general, the observed picture of the process is similar to boiling on a lyophobic
surface; however, there is some limitation on the size of the contact line region associated
with the finite size of the lyophobic spot.
Then, various key characteristics of bubble evolution were studied and dependences
of the departure diameter Dd and the departure frequency νd of bubbles on the width of
the lyophobic zone Dphob were plotted (Figure 6). It can be seen that the bubble departure
diameter Dd increases almost linearly with the width of the lyophobic zone Dphob and
changes by 25% in the studied range (Figure 6a). At the same time, on the dependence of
the bubble departure frequency νd on Dphob , shown in Figure 6b, a maximum is observed

26
Energies 2022, 15, 8204

at Dphob = 450 μm. The higher the bubble departure frequency, the more energy is spent
on vaporization averaged over time. The thermal analysis of calculations results also
demonstrate that the maximum heat flux removed from the surface is observed also for a
lyophobic zone with a width of Dphob = 450 μm. Therefore, further we will consider this
width of the lyophobic zone as the optimal.

Figure 5. Temporal evolution of the growth and departure of a single bubble: (a) t = 0.15 s,
(b) t = 0.175 s, (c) t = 0.375 s, (d) t = 0.475 s, (e) t = 0.5 s, (f) t = 0.525 s. Lyophobic zone size
Dphob = 1.5 mm, horizontal region size is 12 mm, heater temperature Th = 0.95 Tc .

(a) (b)

Figure 6. (a) Bubble departure diameter Dd and (b) departure frequency νd depending on the
lyophobic zone width Dphob . Heater temperature Th = 0.95 Tc .

27
Energies 2022, 15, 8204

3.3. Boiling on the Surfaces with Contrast Wettability


At the next stage, the boiling process was simulated on the surfaces with contrast
wettability by changing the distance L between the lyophobic zones (i.e., pitch size) with
an optimal spot width Dphob = 450 μm. The size of the computational domain was 800 cells
horizontally (24 mm) and 500 cells vertically (15 mm). Simulation was carried out for
different numbers N = 3, 5, 7, 10, 15 and 20 of lyophobic zones equidistantly located
on the lyophilic surface, which corresponded to the range of the distances between the
spots L = 1.2–12 mm, and the ratio L/Dd = 0.46–4.61. Thus, such characteristic cases were
considered when the distance between the centers of bubble nucleation is less L/Dd < 1,
corresponds to L/Dd ~ 1, and is greater L/Dd > 1 than the departure diameter of the bubble.
An additional analysis of the density contour plots and temperature fields was carried
out in a wide range of input heat flux, which permitted determination of the cause-and-
effect relationships between the dependence of the heat transfer efficiency and the boiling
behavior for the various configurations of the patterned biphilic surface.
The analysis of boiling curves presented in Figure 7 shows that in the regime of single
phase convection (ΔT ≤ 0.044 Tc ) the dependencies for surfaces with contrast wettability
coincide with the surface with homogeneous wettability. The onset of nucleate boiling on
the surfaces with contrast wettability is observed at the same temperature of the heater as
on the pure lyophobic surface (TONB,phob ≈ 0.944 Tc ), but much lower than on the lyophilic
surface (TONB,phil ≈ 0.955 Tc ). The calculated curves show that with the development of
boiling at ΔT > 0.044 Tc , the removed heat flux <q> from the biphilic surfaces increases
significantly compared to surfaces with homogeneous wettability at a given wall superheat.
Calculations also show that at certain surface superheats (ΔT ≈ 0.07 Tc ), the intensity of
heat transfer on biphilic surfaces begins to decrease in relation to boiling heat transfer on
lyophilic surface. It is also seen that at high surface superheat ΔT > 0.07 Tc , the removed
heat flux <q> from the surfaces with contrast wettability decreases with a decrease of the
distance L between the lyophobic zones that corresponds to larger number of the zones. It
should be noted that the CHF on the surfaces with contrast wettability is lower than the
CHF on the lyophilic surface, and the value of the CHF decreases with a decrease of the
distance between the lyophobic zones.

Figure 7. Boiling curves q(ΔT  ) for the smooth surfaces with contrast wettability. Lyophobic zone
width Dphob = 450 μm, space separation between the zones L/Dd = 0.46, 0.62, 0.92, 1.3, 1.8 and 4.61.

28
Energies 2022, 15, 8204

Figure 8 shows the density contour plots of boiling on surfaces with contrast wettability
for various numbers of lyophobic zones on the lyophilic substrate, N = 2, 7 and 20, at
moderate surface superheat ΔT = 0.065 Tc . On a surface with N = 2 lyophobic zones, the
formation of bubbles occurs mainly above these zones, and after the bubble departure, the
vapor phase always remains above them. The bubble base then may extend beyond the
boundaries of the lyophobic zone due to the small width of the zone itself. The distance
between the zones for N = 2 exceeds the separation diameter of the bubble by several times,
L/Dd = 4.61, and therefore the bubbles do not interact with each other. For a surface with
N = 7 lyophobic zones, the distance between lyophobic zones becomes comparable to the
bubble departure diameter, L/Dd = 1.3. Neighbor bubbles begin to interact with each other.
As a result of the merging of two adjacent vapor bubbles, some part of the heating surface
can be covered by a vapor phase. The usage of the surface with large number of lyophobic
zones, N = 20, and, accordingly, a small distance between the spots, L/Dd = 0.62, leads
to coalescence of bubbles from neighboring lyophobic zones in the early stages of their
evolution, which ultimately leads to an increase in the area occupied by dry spots and a
decrease in the intensity of heat transfer during boiling compared to other configurations
of biphilic surfaces described above. Figure 7 shows that at moderate surface superheat
ΔT = 0.065 Tc , the heat flux removed from the surface with large distance between the
lyophobic zones, L/Dd = 4.61, is much lower than from the surfaces with L/Dd = 1.3 and
L/Dd = 0.62. However, the heat flux removed from the surface with low distance between
the zones L/Dd = 0.62 is less than from a surface with L/Dd = 1.3. At high surface superheat
ΔT > 0.07 Tc , a significant part of the heating surface is covered by a vapor film due to
merging of several bubbles. These local areas of film boiling are quite stable over time,
which leads to a noticeable decrease in the intensity of heat transfer for biphilic surfaces
with L/Dd < 1.8.

Figure 8. The density contour plots of boiling at ΔT = 0.065 Tc for different numbers of the lyophobic
zones: (a) N = 2 (L/Dd ≈ 4.61), (b) N = 7 (L/Dd ≈ 1.3), (c) N = 20 (L/Dd ≈ 0.62).

It should be noted that the degradation of heat transfer on modified surfaces with
contrast properties compared to the unmodified lyophilic surface was also observed ex-
perimentally. In [50], degradation of the heat transfer has been observed with increasing
the size of hydrophobic spots. In [51], the formation of a “gigantic single bubble” on the
pattered surface was observed as a result of bubbles merging at high surface superheats. A
similar effect of heat transfer degradation was observed on a structured surface at a small
distance between cavities [52].

3.4. Discussion
Finally, we tried to generalize the obtained results on the enhancement of heat transfer
at pool boiling on surfaces with contrast and homogeneous wettability. Let us consider the
ratio of the heat transfer coefficient on a biphilic surface to the heat transfer coefficient on a
lyophilic surface, HTC/HTCphil . Figure 9 shows the dependences of enhancement factor
HTC/HTCphil on the distance between the lyophobic zones for various values of the heater
temperature Th . It can be seen that for moderate heater temperatures, Th < 0.96 Tc , the
enhancement factor on surfaces with contrast wettability is greater than unity. However, at
high heater temperatures, Th > 0.97 Tc , the enhancement factor on the modified surfaces de-

29
Energies 2022, 15, 8204

creases due to the presence of wide unwetted areas. It was also found that the dependences
of the enhancement factor on the distance between zones L/Dd have a maximum around
L/Dd ~ 1.1. The maximum values of the enhancement factor HTC/HTCphil ≈ 4 at heater
temperature Th = 0.955 Tc were obtained for surfaces with parameters L/Dd ≈ 0.92 and 1.3.
Thus, as a result of the performed numerical simulations, it can be concluded that for mod-
erate surface superheat the optimal distance between the lyophobic zones should be of the
order of the bubble departure diameter, 0.9 ≤ L/Dd ≤ 1.3 and the ratio of the width of the
lyophobic zone to the bubble departure diameter should be of the order of Dphob /Dd ~ 0.16.
Such distance between the lyophobic zones relative to the bubble departure diameter is
consistent with the results of experiments [53], in which heat transfer and bubble dynamics
were simultaneously analyzed on biphilic surfaces with different configurations.

Figure 9. Enhancement factor for heat transfer on the surfaces with contrast wettability depending
on the distance between the lyophobic zones L/Dd for different heater temperatures Th .

It should be emphasized that the simulation results obtained in the paper with the
help of hybrid LBM model for the boiling on ideally smooth surfaces do not take into
account a number of important aspects related to real experimental situations. First of all,
the roughness of heat exchange surfaces leads to some stabilization of the bubbles and
influences the nucleation process. Although the liquid parameters given in the model are
taken as close as possible to water physical properties, the obtained results correspond to
some kind of a model fluid. In the future, it is necessary to carry out simulations varying
the properties of the fluid in order to search for universal regularities. It should be also
noted that in the model a low liquid to vapor densities ratio (ρl /ρv ~ 25) is considered,
that corresponds to the case close to the critical point. Furthermore, 3D modeling could
lead to a better agreement with available experimental data than obtained in this work
with 2D simulations. Nevertheless, the presented results qualitatively describe the main
characteristics of the boiling process, and the abovementioned shortcomings set a task for
improvement of the model in the future.

4. Conclusions
In this study, to simulate the pool boiling on smooth surfaces with contrast wettability,
a hybrid model based on the lattice Boltzmann method and the heat transfer equation in
a two-phase medium was used. As a result of numerical simulation, boiling curves were
calculated for surfaces with uniform wettability (lyophilic, neutral and lyophobic surfaces).
The process of boiling over a single lyophobic zone on a lyophilic surface was studied, and

30
Energies 2022, 15, 8204

dependences of the bubble departure frequency and bubble departure diameter on the
width of the spot were obtained that permit calculation of its optimal size. The pool boiling
on the surfaces with different configuration of the patterns was studied and the boiling
curves were calculated.
The following conclusions were drawn:
- The calculations showed that the use of the surfaces with contrast wettability permits
a substantial decrease in the onset of nucleate boiling compared to a bare lyophilic
surface. The ONB of biphilic surfaces occurs approximately at the same superheats as
on a homogeneous lyophobic surface.
- Based on the simulations of a bubble dynamic at boiling above a single lyophobic
zone and on analysis of key characteristics of bubble evolution, i.e., the bubble depar-
ture diameter Dd and the bubble departure frequency νd , the optimal width of the
lyophobic spot Dphob /Dd ~ 0.16 was obtained.
- It was shown that at moderate surface superheat, the heat transfer at boiling on
surfaces with contrast wettability is significantly higher than the heat transfer on the
surfaces with uniform wettability.
- In terms of heat transfer performance, the optimal configuration of the biphilic
surface at moderate surface superheats was determined: the ratio of the width of
the lyophobic zone Dphob /Dd ~ 0.16 and the distance between the lyophobic zones
0.9 ≤ L/Dd ≤ 1.3 to the bubble departure diameter permits enhancement of the heat
transfer by more than 4 times. The presented comprehensive studies made it possible
to better understand the physics of the boiling process on surfaces with contrast wet-
tability, and to identify the mechanisms of heat transfer enhancement/degradation.
Based on the obtained simulation results, an optimal configuration of lyophobic zones
on a lyophilic heating surface is proposed for a significant enhancement of boiling
heat transfer. This information can already be used for future experiments to simplify
the search for optimal configurations of biphilic surfaces for enhancement of boiling
heat transfer. At the same time, it is necessary to further improve the numerical model,
which would allow not only to control the surface wetting properties, but also to set
morphological parameters, as well as to vary the properties of the liquid to simulate
real coolants, heat exchange devices and thermal stabilization systems operating in
various boiling conditions.

Author Contributions: Conceptualization, A.V.F. and A.S.S.; methodology, A.V.F. and A.E.O.; formal
analysis, A.V.F. and A.S.S.; writing—original draft preparation, A.V.F., M.V.S., A.E.O. and A.S.S.;
writing—review and editing, A.V.F., M.V.S., A.E.O. and A.S.S. All authors have read and agreed to
the published version of the manuscript.
Funding: This research was funded by Russian Science Foundation grant number 22-29-01251.
Data Availability Statement: Data are available from the authors.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Liang, G.; Mudawar, I. Review of pool boiling enhancement by surface modification. Int. J. Heat Mass Transf. 2019, 128, 892–933.
[CrossRef]
2. Sajjad, U.; Sadeghianjahromi, A.; Muhammad, H.A.; Wang, C.C. Enhanced pool boiling of dielectric and highly wetting
liquids—A review on enhancement mechanisms. Int. J. Heat Mass Transf. 2020, 119, 104950. [CrossRef]
3. Mehralizadeh, A.; Shabanian, S.; Bakeri, G. Effect of modified surfaces on bubble dynamics and pool boiling heat transfer
enhancement: A review. Therm. Sci. Eng. Prog. 2019, 15, 100451. [CrossRef]
4. Khan, S.A.; Atieh, M.A.; Koç, M. Micro-nano scale surface coating for nucleate boiling heat transfer: A critical review. Energies
2018, 11, 3189. [CrossRef]
5. Surtaev, A.; Serdyukov, V.; Pavlenko, A. Nanotechnologies for thermophysics: Heat transfer and crisis phenomena at boiling.
Nanotech. Russ. 2016, 11, 696–715. [CrossRef]
6. Rahman, M.; Olceroglu, E.; McCarthy, M. Role of wickability on the critical heat flux of structured superhydrophilic surfaces.
Langmuir 2014, 30, 11225–11234. [CrossRef]

31
Energies 2022, 15, 8204

7. Xie, S.; Beni, M.S.; Cal, J.; Zhao, J. Review of critical-heat-flux enhancement methods. Int. J. Heat Mass Transf. 2018, 122, 275–289.
[CrossRef]
8. Bourdon, B.; Rioboo, R.; Marengo, M.; Gosselin, E.; De Coninck, J. Influence of the wettability on the boiling onset. Langmuir 2012,
28, 1618–1624. [CrossRef]
9. Betz, A.; Jenkins, J.; Kim, C.; Attinger, D. Boiling heat transfer on superhydrophilic, superhydrophobic, and superbiphilic surfaces.
Int. J. Heat Mass Transf. 2013, 57, 733–741. [CrossRef]
10. Safonov, A.I.; Kuznetsov, D.V.; Surtaev, A.S. Fabrication of Hydrophobic Coated Tubes for Boiling Heat Transfer Enhancement.
Heat Transf. Eng. 2020, 42, 1390–1403. [CrossRef]
11. Surtaev, A.; Koşar, A.; Serdyukov, V.; Malakhov, I. Boiling at subatmospheric pressures on hydrophobic surface: Bubble dynamics
and heat transfer. Int. J. Therm. Sci. 2022, 173, 107423. [CrossRef]
12. Teodori, E.; Valente, T.; Malavasi, I.; Moita, A.S.; Marengo, M.; Moreira, A.L.N. Effect of extreme wetting scenarios on pool boiling
conditions. Appl. Therm. Eng. 2017, 115, 1424–1437.
13. Phan, H.; Caney, N.; Marty, P.; Colasson, S.; Gavillet, J. Surface wettability control by nanocoating: The effects on pool boiling
heat transfer and nucleation mechanism. Int. J. Heat Mass Transf. 2009, 52, 5459–5471. [CrossRef]
14. Motezakker, A.; Sadaghiani, A.K.; Çelik, S.; Larsen, T.; Villanueva, L.G.; Kosar, A. Optimum ratio of hydrophobic to hydrophilic
areas of biphilic surfaces in thermal fluid systems involving boiling. Int. J. Heat Mass Transf. 2019, 135, 164–174.
15. Može, M.; Zupančič, M.; Golobič, I. Pattern geometry optimization on superbiphilic aluminum surfaces for enhanced pool boiling
heat transfer. Int. J. Heat Mass Transf. 2020, 161, 120265.
16. Cheng, H.C.; Jiang, Z.X.; Chang, P.H.; Chen, P.H. Effects of difference in wettability level of biphilic patterns on copper tubes in
pool boiling heat transfer. Exp. Therm. Fluid Sc. 2021, 120, 110241.
17. Serdyukov, V.; Patrin, G.; Malakhov, I.; Surtaev, A. Biphilic surface to improve and stabilize pool boiling in vacuum. Appl. Therm.
Eng. 2022, 209, 118298.
18. Kharangate, C.R.; Mudawar, I. Review of computational studies on boiling and condensation. Int. J. Heat Mass Transf. 2017, 108,
1164–1196. [CrossRef]
19. Son, G.; Dhir, V.K.; Ramanujapu, N. Dynamics and Heat Transfer Associated With a Single Bubble during Nucleate Boiling on a
Horizontal Surface. ASME J. Heat Transf. 1999, 121, 623–632.
20. Mukherjee, A.; Dhir, V.K. Study of lateral merger of vapor bubbles during nucleate pool boiling. J. Heat Transf. 2004, 126,
1023–1039.
21. Mukherjee, A.; Kandlikar, S.G. Numerical study of single bubbles with dynamic contact angle during nucleate pool boiling. Int. J.
Heat Mass Transf. 2007, 50, 127.
22. Duan, X.; Phillips, B.; McKrell, T.; Buongiorno, J. Synchronized High-Speed Video, Infrared Thermometry, and Particle Image
Velocimetry Data for Validation of Interface-Tracking Simulations of Nucleate Boiling Phenomena. Exp. Heat Transf. 2013, 26,
169–197.
23. Kunkelmann, C.; Stephan, P. CFD simulation of boiling flows using the volume of fluid method within OpenFOAM. Numer. Heat
Transf. A Appl. 2009, 56, 631–646. [CrossRef]
24. Jiaqiang, E.; Zhang, Z.; Tu, Z.; Zuo, W.; Hu, W.; Han, D.; Jin, Y. Effect analysis on flow and boiling heat transfer performance of
cooling water-jacket of bearing in the gasoline engine turbocharger. Appl. Therm. Eng. 2018, 130, 754–766.
25. Sun, D.L.; Tao, W.Q. A coupled volume-of-fluid and level set (VOSET) method for computing incompressible two-phase flows.
Int. J. Heat Mass Transf. 2010, 53, 645–655. [CrossRef]
26. Ling, K.; Li, Z.Y.; Tao, W.Q. A direct numerical simulation for nucleate boiling by the VOSET method. Numer. Heat Transf. A Appl.
2014, 65, 949–971. [CrossRef]
27. She, X.; Shedd, T.A.; Lindeman, B.; Yin, Y.; Zhang, X. Bubble formation on solid surface with a cavity based on molecular
dynamics simulation. Int. J. Heat Mass Transf. 2016, 95, 278–287.
28. Chen, Y.; Zou, Y.; Sun, D.; Wang, Y.; Yu, B. Molecular dynamics simulation of bubble nucleation on nanostructure surface. Int. J.
Heat Mass Transf. 2018, 118, 1143–1151.
29. Zhang, L.; Xu, J.; Liu, G.; Lei, J. Nucleate boiling on nanostructured surfaces using molecular dynamics simulations. Int. J. Therm.
Sci. 2020, 152, 106325.
30. Succi, S. The Lattice Boltzmann Equation for Fluid Dynamics and Beyond; Oxford University Press: Oxford, MI, USA, 2001; p. 304.
31. Chen, S.; Doolen, G.D. Lattice Boltzmann method for fluid flows. Ann. Rev. Fluid Mech. 1998, 30, 329–364.
32. Shan, X.; Chen, H. Lattice Boltzmann model for simulating flows with multiple phases and components. Phys. Rev. E 1993, 47,
1815–1820. [CrossRef] [PubMed]
33. Guo, Z.; Zheng, C.; Shi, B. Discrete lattice effects on the forcing term in the lattice Boltzmann method. Phys. Rev. E 2002, 65, 46308.
[CrossRef] [PubMed]
34. He, X.; Shan, X.; Doolen, G. Discrete Boltzmann equation model for nonideal gases. Phys. Rev. E Rapid Comm. 1998, 57, 13.
[CrossRef]
35. Kupershtokh, A.L. Incorporating a body force term into the lattice Boltzmann equation. Vestnik NGU Quart. J. Novosibirsk State
Univ. Ser. Math. Mech. Inform. 2004, 4, 75–96.
36. Chen, S.; Martínez, D. On boundary conditions in lattice Boltzmann methods. Phys. Fluid. 1996, 8, 2527–2536. [CrossRef]

32
Energies 2022, 15, 8204

37. Li, Q.; Luo, K.H.; Kang, Q.J.; Chen, Q. Contact angles in the pseudopotential lattice Boltzmann modeling of wetting. Phys. Rev. E
2014, 90, 053301.
38. Gong, S.; Cheng, P. Lattice Boltzmann simulation of periodic bubble nucleation, growth and departure from a heated surface in
pool boiling. IJHMT 2013, 64, 122–132. [CrossRef]
39. Fedoseev, A.V.; Surtaev, A.S.; Moiseev, M.I.; Ostapchenko, A.E. Lattice Boltzmann simulation of bubble evolution at boiling on
surfaces with different wettability. J. Phys. Conf. Ser. 2020, 1677, 012085.
40. Gong, S.; Cheng, P. Lattice Boltzmann simulations for surface wettability effects in saturated pool boiling heat transfer. Int. J. Heat
Mass Transf. 2015, 85, 635–646. [CrossRef]
41. Moiseev, M.I.; Fedoseev, A.; Shugaev, M.V.; Surtaev, A.S. Hybrid thermal lattice Boltzmann model for boiling heat transfer on
surfaces with different wettability. Int. Phenom. Heat Transf. 2020, 8, 81. [CrossRef]
42. Li, Q.; Kang, Q.J.; Francois, M.M.; He, Y.L.; Luo, K.H. Lattice Boltzmann modeling of boiling heat transfer: The boiling curve and
the effects of wettability. Int. J. Heat Mass Transf. 2015, 85, 787–796. [CrossRef]
43. Fang, W.Z.; Chen, L.; Kang, Q.J.; Tao, W.Q. Lattice Boltzmann modeling of pool boiling with large liquid-gas density ratio. Int. J.
Therm. Sci. 2017, 114, 172–183. [CrossRef]
44. Zhang, C.; Cheng, P. Mesoscale simulations of boiling curves and boiling hysteresis under constant wall temperature and constant
heat flux conditions. Int. J. Heat Mass Transf. 2017, 110, 319. [CrossRef]
45. Gong, S.; Cheng, P. Numerical simulation of pool boiling heat transfer on smooth surfaces with mixed wettability by lattice
Boltzmann method. Int. J. Heat Mass Transf. 2015, 80, 206–216. [CrossRef]
46. Li, Q.; Yu, Y.; Zhou, P.; Yan, H.J. Enhancement of Boiling Heat Transfer Using Hydrophilic-Hydrophobic Mixed Surfaces: A Lattice
Boltzmann Study. Appl. Therm. Eng. 2018, 132, 490–499. [CrossRef]
47. Peng, Y.; Laura, S. Equations of state in a lattice Boltzmann model. Phys. Fluids 2006, 18, 042101.
48. Nam, Y.; Wu, J.; Warrier, G.; Ju, Y.S. Experimental and numerical study of single bubble dynamics on a hydrophobic surface.
J. Heat Transf. 2009, 131, 121004. [CrossRef]
49. Kim, S.H.; Lee, G.C.; Kang, J.Y.; Moriyama, K.; Park, H.S.; Kim, M.H. The role of surface energy in heterogeneous bubble growth
on ideal surface. Int. J. Heat Mass Transf. 2017, 108, 1901–1909. [CrossRef]
50. Jo, H.; Ahn, H.S.; Kang, S.; Kim, M.H. A study of nucleate boiling heat transfer on hydrophilic, hydrophobic and heterogeneous
wetting surfaces. Int. J. Heat Mass Transf. 2011, 54, 5643. [CrossRef]
51. Choi, C.H.; David, M.; Gao, Z.; Chang, A.; Allen, M.; Wang, H.; Chang, C.H. Large-scale Generation of Patterned Bubble Arrays
on Printed Bi-functional Boiling Surfaces. Sci. Rep. 2016, 6, 23760. [CrossRef]
52. Nimkar, N.D.; Bhavnani, S.H.; Jaeger, R.C. Effect of nucleation site spacing on the pool boiling characteristics of a structured
surface. Int. J. Heat Mass Transf. 2006, 49, 2829. [CrossRef]
53. Lim, D.Y.; Bang, I.C. Controlled bubble departure diameter on biphilic surfaces for enhanced pool boiling heat transfer perfor-
mance. Int. J. Heat Mass Transf. 2020, 150, 119360. [CrossRef]

33
energies
Article
The Influence of Pressure on Local Heat Transfer Rate under the
Vapor Bubbles during Pool Boiling
Vladimir Serdyukov *, Ivan Malakhov and Anton Surtaev

Kutateladze Institute of Thermophysics, 630090 Novosibirsk, Russia


* Correspondence: [email protected]

Abstract: This paper presents the results of an experimental study on the evolution of a nonstationary
temperature field during ethanol pool boiling in a pressure range of 12–101.2 kPa. Experimental
data were obtained using infrared thermography with high temporal and spatial resolutions, which
made it possible to reconstruct the distribution of the heat flux density and to study the influence of
pressure reduction on the local heat transfer rate in the vicinity of the triple contact line under vapor
bubbles for the first time. It is shown that, for all studied pressures, a significant heat flux density
is removed from the heating surface due to microlayer evaporation, which exceeds the input heat
power by a factor of 3.3–27.7, depending on the pressure. Meanwhile, the heat transfer rate in the
area of the microlayer evaporation significantly decreases with the pressure reduction. In particular,
the local heat flux density averaged over the microlayer area decreases by four times as the pressure
decreases from 101.3 kPa to 12 kPa. Estimates of the microlayer profile based on the heat conduction
equation were made, which showed the significant increase in the microlayer thickness with the
pressure reduction.

Keywords: boiling; heat transfer; microlayer; infrared thermography; subatmospheric pressures

1. Introduction
Boiling is one of the most efficient heat transfer regimes, widely used in various
technologies. To date, quite a large number of experimental and theoretical studies devoted
Citation: Serdyukov, V.; Malakhov, I.;
to various aspects of boiling have been presented. Despite this, today there is no complete
Surtaev, A. The Influence of Pressure
theory of heat transfer during nucleate boiling [1–3]. In particular, to calculate the heat
on Local Heat Transfer Rate under
transfer coefficients (HTC) during the boiling of various liquids under various conditions,
the Vapor Bubbles during Pool
dozens of semi-empirical correlations are given in the literature, which usually describe
Boiling. Energies 2023, 16, 3918.
only a narrow change range of regime parameters. This is due to the fact that boiling is a
https://doi.org/10.3390/en16093918
multiscale nonstationary process, and for its correct description it is necessary to take into
Academic Editor: Moghtada Mobedi account the effects that take place on various scales, down to micro- and even nanoscales
Received: 14 April 2023
(Figure 1).
Revised: 30 April 2023
One of the phenomena that occur on the microscale during boiling is the formation
Accepted: 3 May 2023 and further evaporation of a liquid layer under a vapor bubble, which is called a microlayer
Published: 5 May 2023 due to its small thickness [4–6]. As can be seen from Figure 1, this region is limited on
one side by the wall–liquid–vapor contact line and the surrounding liquid (the so-called
macrolayer) on the other. To date, it has generally been accepted that the process of
formation and evaporation of a microlayer is an essential and important part of a vapor
Copyright: © 2023 by the authors. bubble dynamics [7–9]. Moreover, it was shown in a number of papers [10–13] that the
Licensee MDPI, Basel, Switzerland. evaporation of a microlayer is one of the key mechanisms explaining the high intensity
This article is an open access article of heat transfer during nucleate boiling. Finally, understanding the nature and influence
distributed under the terms and of various parameters on the microlayer evaporation rate is important to develop and
conditions of the Creative Commons fabricate micro- and nanostructured heating surfaces to achieve maximum heat transfer
Attribution (CC BY) license (https://
rates and critical heat fluxes during boiling [14–16].
creativecommons.org/licenses/by/
4.0/).

Energies 2023, 16, 3918. https://doi.org/10.3390/en16093918 34 https://www.mdpi.com/journal/energies


Energies 2023, 16, 3918

Figure 1. Multiscale nature of the nucleate boiling.

However, despite the rather large number of papers devoted to both the experimen-
tal study on microlayer evolution and structure [5–12] and the numerical simulation of
its evaporation [17–20], a number of questions remain open. One of such issues is the
influence of system parameters (pressure, subcooling degree, gravity level, etc.), as well
as the properties of a heating surface (wettability, capillary wicking, structure, etc.) on the
evolution and structure of a microlayer.
Pressure is one of the most important parameters in a boiling system, the change in
which has a complex effect on the nucleation, dynamics of vapor bubbles, heat transfer rate
and the development of boiling crisis phenomena. At the same time, it is known that the
pressure reduction below atmospheric level leads not only to a quantitative, but also to a
qualitative change in the boiling character [21–25]. In particular, a decrease in pressure is
accompanied by a decrease in the intensity of heat transfer and a decrease in the critical
heat flux, the nucleation of bubbles becomes more unstable and takes place at larger surface
superheating. As a result, a noticeable change in the local characteristics of boiling (bubble
growth rates, departure diameters, nucleation frequency, etc.) is observed. However, as
the literature analysis shows, in contrast to boiling at atmospheric pressure, today there
is practically no information about the features of local heat transfer and the dynamics of
the contact line under vapor bubbles during boiling at subatmospheric pressures. Among
other things, this is due to the fact that in most experimental studies the analysis of the
temperature of a heating surface during boiling was carried out using thermocouples or
temperature sensors. This method of temperature measurement can only be used to analyze
the surface temperature averaged over the heater area.
Today, to measure the temperature field distribution of various objects, including
boiling systems, the method of high-speed infrared thermography is widely used, which
is devoid of the above-mentioned disadvantages. As an analysis of the papers of various
authors (e.g., [10,23,26–34]) shows, this technique makes it possible to obtain fundamentally
new information about the local and integral multiscale characteristics of boiling, as well
as about the development of crisis phenomena under various conditions. In particular,
infrared thermography allows the features of heat transfer in the vicinity of the triple contact
line to be studied, including the evaporation of the microlayer during boiling [10,11].
In the present paper, an experimental study on the evolution of a nonstationary tem-
perature field during the pool boiling of ethanol at pressures 12–101.2 kPa was performed.
Experimental data were obtained using infrared thermography with high temporal and
spatial resolutions, which made it possible to reconstruct the distribution of the heat flux

35
Energies 2023, 16, 3918

density and to study the influence of pressure reduction on the local heat transfer rate in
the vicinity of the triple contact line under the vapor bubbles.

2. Materials and Methods


2.1. Experimental Setup
The experiments were performed using the setup, a schematic view of which is shown
in Figure 2. The setup consists of two sealed parts. To maintain a constant temperature
of the working fluid, the inner volume was placed in thermostatic chamber filled with
deionized water. The water temperature in the thermostatic volume was set at the required
level using the Danfoss EKC-102 electronic temperature controller and two pre-heaters,
each with a power of 1.2 kW. The temperature of the liquid in the inner volume (T1 ) and the
outer chamber (T2 ) was monitored using Honeywell HEL 700 and NTC thermistors, respec-
tively. In order to avoid the system pressure increase in the working area during boiling
experiments, the internal volume was equipped with a water-cooled vapor condenser.

Figure 2. The scheme of the experimental setup.

To perform boiling experiments under subatmospheric pressure conditions, the setup


was evacuated according to DIN 28400-1-1990 norm. The required pressure level in the
working volume was set using a rotary vane vacuum pump EVP 2XZ-1C, equipped with a
liquid nitrogen trap, and controlled using a digital piezoresistive vacuum gauge Thyracont
VD81. The value of the reduced pressure ps in the experiments was calculated based on
the readings of the vacuum gauge (p), taking into account the influence of the hydrostatic
pressure of the working fluid column with height h:

ps = p + ρl gh (1)

As the working fluid, 95% ethanol was used under the saturation conditions at a given
pressure ps . During experiments on boiling at low pressures, it is extremely important
to correctly calculate the system conditions. In the present study, for such conditions (ps ,
Tsat ), the ones near a heating surface were considered. To do this, in each experiment
the value of the reduced pressure was calculated according to (1) and the corresponding
saturation temperature of the working fluid Tsat was taken from the tabular data. The
following reduced pressures were studied in the paper: 12 kPa (Tsat = 32.4 ◦ C), 21.8 kPa
(Tsat = 44.3 ◦ C), 35 kPa (Tsat = 53.6 ◦ C), 57.6 kPa (Tsat = 64.7 ◦ C), 80 kPa (Tsat = 72.4 ◦ C) and
101.3 kPa (Tsat = 78.6 ◦ C).

36
Energies 2023, 16, 3918

2.2. Heating Surface


The heating surface used in the study (Figure 3) was a conductive film of indium-tin
oxide (ITO), 1 μm thick, deposited by ion sputtering onto a sapphire substrate 60 mm in
diameter and 3 mm thick. The area of the heat release was 30 × 28 mm2 . Samples were
resistively heated by a DC power supply Elektro Automatik PS 8080-60 DT via thin (2 μm)
chromium-nickel electrodes vacuum deposited onto the ITO film and brought to the reverse
side of the sapphire substrate.

Figure 3. The scheme of the heating surface used in the study.

An important advantage of using indium–tin oxide as a heater material in experi-


ments to study boiling characteristics is its transparency in the visible wavelength range
(380–750 nm) and opacity in the mid-IR range (3–5 μm). In turn, the integral transmission
capacity of a sapphire in the 0.3–5 μm wavelength range exceeds 80%. The combination
of these properties makes it possible to measure the non-stationary temperature field on
the ITO film using an infrared camera with corresponding spectral range and to visually
record the vapor bubble dynamics on the heating surface using a video camera [23,26–31].
In the present study, the transient temperature field of the ITO heater was recorded
using a FLIR X6530sc high-speed infrared camera with a spectral range of 1.5–5.1 μm,
160 × 128 resolution and a frame rate of 1500 fps. The spatial resolution of the IR recording
in the experiments was 250 μm per pixel. Prior to the boiling experiments, the IR camera
was calibrated using a resistance temperature detector located near the film heater [23].
Prior to the experiments, an analysis of the physicochemical properties of the heat-
releasing surface was performed. According to the profilometry performed using the
Bruker Contour GT-K1 optical profilometer, the ITO film roughness was Ra = 3 nm, which
classifies it as an “ultra” smooth heating surface. Using the KRUSS DSA 100 setup, an
analysis of the wetting properties of the ITO film was performed for the working fluid to
determine the value of the static contact angle (θ). According to these measurements, the
value of θ was about 10◦ . To additionally check the working fluid for the absence of harmful
insoluble impurities, the wetting properties of the heating surface were also investigated
after boiling experiments. An analysis of these results showed that the measured values of
the contact wetting angles before and after the experiments are consistent with each other
within the error (±2◦ ), which indicates sufficient purity of the used liquid and the absence
of harmful impurities in it.

2.3. Local Heat Flux Calculation


To analyze the local density of the heat flux coming from the heater to the liquid, a
numerical calculation was performed using the experimental data of infrared thermography
on the evolution of the ITO film temperature field. Below, the problem of heat conduction
to recover local heat fluxes will be described. In the study, a numerical solution of the
non-stationary heat conduction equation was performed, which has the following form:
 2 
∂T ∂ T ∂2 T ∂2 T
= λsap + + . (2)
∂t ∂x2 ∂y2 ∂z2

37
Energies 2023, 16, 3918

The three-dimensional internal volume of the sapphire substrate (Figure 4a) was used
as the computational domain, bounded from above and below by the outer boundaries of
the sapphire, and from the sides by the boundaries of the heat release region. The boundary
conditions were given as follows:

T ( x, y)z=zmax = TIR ( x, y), (3)

∂T ∂T ∂T
= 0, = 0, = 0. (4)
∂z z=0 ∂x x=0,x= xmax ∂y y=0.y=ymax
These conditions correspond to the absence of heat losses to the sides, which, as the
analysis of the experimental results showed, is quite correct. The temperature field of the
wall–liquid contact surface changed with time, as happened in the experiment.

(a)

(b)
Figure 4. (a) The calculation domain including boundary conditions to reconstruct heat flux density
distribution maps during boiling on the ITO heater based on the infrared thermography data; (b) an
example of the reconstruction of the heat flux distribution map (ps = 21.8 kPa, qinput = 150 kW/m2 ).

The calculations were performed using an explicit finite difference scheme. The
splitting of the calculated area along the X and Y axes corresponded to the splitting of the
IR frame to the pixels. The splitting along the Z axis and in time were chosen based on the
analysis of the decrease in the calculation error with an increase in the amount of splitting.
For calculations at intermediate moments between recorded IR frames, the temperature
field was recovered using linear interpolation. The temperature field measured with IR
camera was set as the initial temperature distribution in the calculated volume on all layers.
As a result of the numerical calculations, the non-stationary temperature distribution
in the volume of the sapphire substrate was recovered. To calculate the density of the heat
flux incoming to the liquid, the energy conservation condition was used, which takes into
account the heat release in the ITO film qinput :

ql = qinput − qsap , (5)

38
Energies 2023, 16, 3918

where qsap is the heat flux density going into the sapphire substrate, determined on the
basis of the calculated temperature distribution in the volume of sapphire:

∂T
qsap = λ (6)
∂z
As an example, Figure 4b shows an IR recording frame and the corresponding cal-
culated distribution of the heat flux density during the boiling of ethanol at ps = 21.8 kPa
(qinput = 150 kW/m2 ). As can be seen, the obtained experimental data and the heat fluxes
map calculated on their basis make it possible to identify and analyze various heat transfer
areas in the vicinity of the contact line (Figure 1). In particular, the area in the center of the
vapor bubble with a higher temperature and with a lower heat flux density consequently is
a dry spot bounded by a triple contact line. In turn, the region along the periphery of the
bubble with the minimum temperature and the maximum heat flux density corresponds to
the region of a microlayer evaporation.

2.4. Measurements Uncertainties


The uncertainty of the pressure measurement consists of the sum of the pressure
sensor accuracy and the accuracy of the liquid column height h measurement. The total
pressure measurement error was about 1% in the experiments. In turn, the uncertainty of
the input heat flux measurement includes inaccuracies associated with the current, voltage
gauging, and heat losses. According to the apparatus data sheets, the total error in current
and voltage measurements is no more than 1%. In turn, the 2D numerical calculation in
Comsol Multiphysics was performed to analyze the lateral heat losses for the steady-state
conditions. The results showed that the heat losses contribute about 5%, since the sapphire
substrate has a relatively low thermal conductivity and the heat transfer coefficients on the
ITO-liquid surface are rather high. Therefore, the total uncertainty of the heat flux density
measurement in the experiments was about 6%. The major contributions to the uncertainty
of the surface temperature measurements using IR camera are an uncertainty associated
with the calibration procedure and camera sensitivity. According to the analysis, the total
uncertainty of the surface temperature measurement was no more than 1.5 K.
To verify the correctness of the performed data gathering and curation, a comparison
of the obtained results on the heat transfer rate during boiling at atmospheric pressure
with the models of Rohsenow, 1952 [35] and Yagov, 1988 [36] was made (Figure 5). As the
literature analysis shows, these models are some of the most used to describe boiling data
for various liquids and heating surfaces.

Figure 5. Comparison of the obtained experimental data on the heat transfer rate during ethanol pool
boiling under atmospheric pressure with the models of Rohsenow (1952) [35] and Yagov (1988) [36].

39
Energies 2023, 16, 3918

As can be seen from Figure 5, the experimental data show the same trend as the
used dependencies, which confirms that the data were collected correctly. However, it is
also seen that the experiments show lower values of HTC compared to the calculations.
Apparently, this is due to a rather low roughness of the heating surface used in the study. In
particular, it is well known [37] that smooth and ultrasmooth surfaces are characterized by
much lower heat transfer rates during boiling compared to the so-called technical surfaces,
for which the semi-empirical models shown in Figure 5 were developed.

3. Results
3.1. Heat Transfer Rate
First of all, the effect of the pressure reduction on the heat transfer rate during ethanol
boiling was studied. To do this, the values of the integral temperature of the heating
surface (averaged over the area and recording time (10 s)) were obtained based on the
IR recording data. As the result, the corresponding boiling curves q(ΔT) were plotted for
various pressures (Figure 6a). As can be seen from the figure, the pressure reduction in
a given range is accompanied by a significant increase in the onset of nucleate boiling
(ONB). In particular, if for atmospheric pressure conditions the onset of nucleate boiling
corresponds to surface superheating ΔT = 14 K, then for the lowest studied pressure (12 kPa)
this value is 37 K.

(a) (b)

Figure 6. (a) Ethanol boiling curves at various pressures ps ; (b) the pressure influence on the heat
transfer coefficient during ethanol boiling at various heat flux densities.

Using the obtained data, the heat transfer coefficients HTC were also determined and
dependencies of HTC(ps ) were constructed for different heat flux densities qinput (Figure 6b).
An analysis of the obtained curves shows that for all presented heat flux densities, the
dependence of HTC(ps ) has a linear form. The pressure reduction from 101.3 kPa to 12 kPa
leads to the heat transfer deteriorating by approximately 40%. The literature analysis shows
that the results on the influence of pressure reduction on heat transfer rate during boiling
obtained in the present study are consistent with the data of other authors obtained using
different liquids, including ethanol [38].
As noted in the introduction, such a decrease in the heat transfer rate during boiling
with the pressure reduction is directly related to a significant change in the nucleation
and vapor bubbles dynamics. This is clearly demonstrated by the analysis of heat flux
distribution maps obtained according to the algorithm described in Section 2.3 (Figure 7).

40
Energies 2023, 16, 3918

Figure 7. Heat flux density distribution maps during boiling of ethanol at various pressures.

In particular, Figure 7 demonstrates that a noticeable decrease in the number of


nucleation sites happens with the pressure reduction, while the departure diameters of the
vapor bubbles and their lifetime on the heating surface increase. In addition, it is clearly
seen that in the area of microlayer evaporation, the local heat flux density decreases with
pressure reduction. At the same time, it is also seen that a change in pressure affects the
heat transfer rate in the regions free of vapor bubbles, i.e., the convective component of the
heat transfer according to the mechanistic approach [39,40]. Thus, for atmospheric pressure,
the average heat flux density removed by the convection does not exceed 50 kW/m2 , while
for pressures below 35 kPa this value is about 150 kW/m2 , which is comparable to the
input heat power. In the future, the authors plan to perform a more detailed analysis of
the effect of pressure reduction on the contribution of various heat transfer mechanisms
(e.g., convection, microlayer evaporation, and quenching) during boiling, based on the
machine learning algorithms [41] of infrared thermography dataset analysis. Further, a
more detailed analysis of the features of local heat transfer under single nucleation sites
during boiling at various pressures will be presented.

3.2. Local Heat Transfer under the Vapor Bubbles


Figure 8 demonstrates the heat flux density distribution under a vapor bubble at different
moments for various pressures and the same input heat flux density (qinput = 150 kW/m2 ). It
can be seen that for all pressures, the maximum local heat flux is observed in the region
of microlayer evaporation at the initial stage of vapor bubble growth (t = 0.67–1.3 ms
depending on pressure). Additionally, the appearance of a region with reduced heat transfer
under the vapor bubble’s center is clearly visible for all bubbles, which corresponds to the
formation and growth of a dry spot on a heating wall. As a result, the local heat flux spent
to evaporate the microlayer decreases as the bubble grows and becomes comparable to
the input heat power at the final stage of the bubble’s lifecycle. In addition, the presented
figures reveal a significant effect of a pressure reduction on the complete depletion time of
the microlayer, which is associated with a significant increase in the lifetime of a bubble
with a decrease in pressure. In particular, for atmospheric pressure, the microlayer fully
evaporated in 6 ms. In turn, as the pressure decreased to ps = 12 kPa, this value increased
to more than 26 ms.

41
Energies 2023, 16, 3918

Figure 8. Evolution of the heat flux density distribution under vapor bubbles during ethanol boiling
at various pressures ps (qinput = 150 kW/m2 ).

In the further analysis of the microlayer region, the following criteria were used to
determine its boundaries: the inner microlayer boundary corresponds to the maximum
intensity of local heat transfer. In turn, the outer boundary of the microlayer was deter-
mined from the condition that the local heat flux does not change with distance from the
bubble center.

42
Energies 2023, 16, 3918

In addition, the presented figures show that the region of microlayer evaporation
noticeably increases in size with pressure reduction. In particular, if its maximum transverse
size during boiling at atmospheric pressure is about 2.5 mm, then at 12 kPa this value
is more than 10 mm. On the one hand (also taking into account the longer lifetime of
the microlayer), this suggests that during boiling at low pressures, the evaporation of
the microlayer should make a greater contribution to the overall heat transfer compared
to boiling at atmospheric conditions. On the other hand, the results obtained indicate a
noticeable decrease in the density of the heat flux removed from the heating surface due to
the microlayer evaporation with pressure reduction. As noted earlier, the authors plan to
use the obtained data for a more accurate analysis of the effect of pressure reduction on the
contribution of various mechanisms, including microlayer evaporation, to the heat transfer
during boiling.
An analysis of the results shows that with pressure reduction, the maximum heat
flux density spent to microlayer evaporation decreases significantly (Figure 9a). Thus,
the ratio of this value (qml max ) to the input heat flux density decreases from 27.7 times
(qml max ≈ 4.36 × 103 kW/m2 ) at boiling under atmospheric pressure to 3.3 times
(qml max ≈ 484 kW/m2 ) at ps = 12 kPa.

(a) (b)

Figure 9. (a) Dependence of the maximum heat flux density in the microlayer evaporation area on
pressure; (b) evolution of the local heat flux density averaged over the microlayer area during ethanol
boiling at various pressures (qinput = 150 kW/m2 ).

Additionally, two trends are observed in Figure 9a. If the pressure reduction from
atmospheric to 35 kPa demonstrates close to linear dependence qml max (ps ), then a further
decrease in pressure to 12 kPa no longer leads to such a significant drop in the heat transfer
intensity in the microlayer evaporation region. It is interesting to note that in the previous
studies of the authors [23], a change in the dynamics of the contact line under vapor bubbles
was also experimentally found during water boiling at a similar pressure. In particular,
the authors showed that if pressure reduction to 42 kPa leads to a decrease in the growth
rate of the dry spot (which is directly related to the evaporation of the microlayer), then a
further pressure reduction, on the contrary, is characterized by an increasing growth rate of
the dry spot. This indicates that, apparently, at these pressure levels, there is a change in
the determining mechanisms of the intensity of microlayer evaporation and the dynamics
of the triple contact line. Nevertheless, an accurate answer to this question requires a
comprehensive analysis of both the dynamics of the contact line and the structure of the
microlayer during the boiling of various liquids at various pressures.
Using the obtained experimental data, the evolution of the local heat flux averaged
over the area of microlayer evaporation was also studied for various pressures (Figure 9b).

43
Energies 2023, 16, 3918

It can be seen that for all studied pressures, the main contribution to heat transfer is made
by the evaporation of the microlayer at the initial stage of vapor bubble growth. This figure
again indicates that the pressure reduction leads to a noticeable decrease in the heat flux
density spent to evaporate the microlayer. In particular, the heat flux density averaged over
the microlayer area decreases by four times with a pressure reduction from atmospheric
level to 12 kPa.

3.3. Microlayer Thickness Estimation


Using the obtained data on the heat flux densities spent to evaporate the microlayer,
it becomes possible to reconstruct its profile δml (Rb ), where δml is the thickness of the
microlayer at a distance Rb from the center of a bubble. A fairly simple estimate can be
made using the following heat conduction equation:

kl
qml = ( Tw − Ti ) (7)
δml

where Tw is the temperature of a heating surface and Ti is the temperature at the vapor–
liquid interface. In this case, the temperature Ti was taken to be the saturation temperature
for a given pressure. Figure 10 shows the microlayer profiles at the initial moment of its
formation for pressures ps = 12, 57.6 and 101.3 kPa. Additionally, the values of the so-called
initial thickness of the microlayer δml 0 (Figure 1) are presented. The obtained profiles show
that with a pressure reduction from atmospheric pressure to the lowest one, the microlayer
becomes more than eight times thicker. Moreover, it can be seen that at ps = 12 kPa, the
minimum thickness of the microlayer in the vicinity of the triple contact line is noticeably
larger and is about 10 μm. This may be due to a significant change in the shape of vapor
bubbles during boiling at subatmospheric pressures and their strong deviation from the
spherical shape characteristic of boiling at atmospheric pressure. In particular, a number of
authors showed via high-speed video recordings [22,23,25] that the vapor bubbles formed
during boiling in vacuum have a “mushroom” shape with a pronounced vapor stem, which
connects the bubble body with a heating surface.

Figure 10. Microlayer profiles during ethanol boiling at various pressures.

It is important to note here that the performed analysis is only an estimate of the
microlayer thickness during boiling at various pressures, which does not claim to be
quantitatively accurate. In particular, a number of authors [42,43] showed that the heat
flux spent to evaporate the microlayer depends significantly on the so-called evaporation
resistance associated with the molecular dynamics of the evaporation process at the liquid–
vapor interface. Ignoring it and calculating the microlayer thickness only using the heat

44
Energies 2023, 16, 3918

conduction Equation (7) shows overestimated values of δml (Rb ). For an accurate description
of the microlayer profile, it is necessary to perform direct experimental observations using
interferometric methods [6,7,10,12], which make it possible to study the process of its
formation and growth with high temporal and spatial resolutions and to study the δml (Rb )
dependence in detail. Here, the method of LED interferometry proposed recently in [44]
should be especially noted, which is a simpler alternative to laser interferometry for
studying the structure and evolution of a microlayer during boiling. Using the example
of water flow boiling, the authors of [44] showed that this relatively easy-to-implement
technique makes it possible to obtain distinct interference patterns during the formation
and growth of bubbles, which makes it possible to analyze the microlayer thickness. The
usage of LED interferometry simultaneously with the high-speed infrared thermography is
promising for studying the effect of pressure on the microlayer structure during boiling.
The authors of the present paper plan to perform such a study in the near future.

4. Conclusions
In the present paper, the effect of pressure reduction on the local heat transfer in the
vicinity of the contact line during pool boiling under conditions was studied for the first
time. Experimental data on the evolution of the non-stationary temperature field of the
heating surface during the ethanol boiling in the pressure range of 12–101.3 kPa were
obtained using high-speed infrared thermography. The developed numerical algorithm
made it possible to reconstruct the distribution of the heat flux density over the heater for
various pressures and to study the heat transfer in the area of microlayer evaporation. The
analysis of the data showed the following:
• The dependence of heat transfer coefficients on pressure during ethanol pool boiling
has a linear form. A decrease in pressure from 101.3 kPa to 12 kPa leads to a decrease
in the intensity of heat transfer of about 40%. At the same time, the onset of nucleate
boiling noticeably increases—from 14 to 37 K.
• The data obtained indicate the effect of pressure reduction on the contribution of
various mechanisms to the integral heat transfer rate during boiling. The results can
be further used to perform a more accurate analysis of this influence, including using
machine learning algorithms.
• For all studied pressures, a significant heat flux density is removed from the heating
surface due to microlayer evaporation (qml max = 484 kW/m2 –4.36 × 103 kW/m2 ),
which exceeds the input heat power by a factor of 3.3–27.7, depending on pressure.
• The heat transfer rate in the area of the microlayer evaporation significantly decreases
with reduction in pressure. Thus, local heat flux density averaged over the microlayer
area decreases by four times as the pressure decreases from 101.3 kPa to 12 kPa.
• Estimates of the microlayer thickness based on the heat conduction equation were
made. The results showed that the pressure reduction from atmospheric level to
12 kPa leads to the microlayer thickness increasing by more than eight times—from
7 μm to 57 μm.

Author Contributions: Conceptualization, V.S. and A.S.; methodology, V.S. and A.S.; software, I.M.;
validation, V.S. and I.M.; formal analysis, V.S. and I.M.; investigation, I.M. and V.S.; resources, V.S.
and A.S.; data curation, V.S. and I.M.; writing—original draft preparation, V.S.; writing—review
and editing, V.S.; visualization, V.S. and I.M.; supervision, V.S.; project administration, V.S.; funding
acquisition, V.S. All authors have read and agreed to the published version of the manuscript.
Funding: The work was supported by the Russian Science Foundation (Grant No. 22-79-00174). The
experimental setup was designed and made within the framework of the state assignment of the IT
SB RAS (№ 121031800216-1).
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

45
Energies 2023, 16, 3918

References
1. Yagov, V.V. Nucleate boiling heat transfer: Possibilities and limitations of theoretical analysis. Heat Mass Transf. 2009, 45, 881–892.
[CrossRef]
2. Koizumi, Y.; Shoji, M.; Monde, M.; Takata, Y.; Nagai, N. Boiling: Research and Advances; Elsevier: Amsterdam, The Netherlands, 2017.
3. Liang, G.; Mudawar, I. Pool boiling critical heat flux (CHF)—Part 2: Assessment of models and correlations. Int. J. Heat Mass
Transf. 2018, 117, 1368–1383. [CrossRef]
4. Moore, F.D.; Mesler, R.B. The measurement of rapid surface temperature fluctuations during nucleate boiling of water. AIChE J.
1961, 7, 620–624. [CrossRef]
5. Cooper, M.G.; Lloyd, A.J.P. The microlayer in nucleate pool boiling. Int. J. Heat Mass Transf. 1969, 12, 895–913. [CrossRef]
6. Voutsinos, C.M.; Judd, R.L. Laser Interferometric Investigation of the Microlayer Evaporation Phenomenon. ASME J. Heat Transf.
1975, 97, 88–92. [CrossRef]
7. Jawurek, H.H. Simultaneous determination of microlayer geometry and bubble growth in nucleate boiling. Int. J. Heat Mass
Transf. 1969, 12, 843–848. [CrossRef]
8. Chen, Z.; Utaka, Y. On heat transfer and evaporation characteristics in the growth process of a bubble with microlayer structure
during nucleate boiling. Int. J. Heat Mass Transf. 2015, 81, 750–759. [CrossRef]
9. Srivastava, A. On the identification and mapping of three distinct stages of single vapor bubble growth with the corresponding
microlayer dynamics. Int. J. Multiph. Flow 2021, 142, 103722.
10. Jung, S.; Kim, H. An experimental study on heat transfer mechanisms in the microlayer using integrated total reflection, laser
interferometry and infrared thermometry technique. Heat Transf. Eng. 2015, 36, 1002–1012. [CrossRef]
11. Serdyukov, V.S.; Surtaev, A.S.; Pavlenko, A.N.; Chernyavskiy, A.N. Study on local heat transfer in the vicinity of the contact line
under vapor bubbles at pool boiling. High Temp. 2018, 56, 546–552. [CrossRef]
12. Utaka, Y.; Hu, K.; Chen, Z.; Morokuma, T. Measurement of contribution of microlayer evaporation applying the microlayer
volume change during nucleate pool boiling for water and ethanol. Int. J. Heat Mass Transf. 2018, 125, 243–247. [CrossRef]
13. Narayan, L.S.; Srivastava, A. Non-contact experiments to quantify the microlayer evaporation heat transfer coefficient during
isolated nucleate boiling regime. Int. Commun. Heat Mass Transf. 2021, 122, 105191. [CrossRef]
14. Zou, A.; Singh, D.P.; Maroo, S.C. Early evaporation of microlayer for boiling heat transfer enhancement. Langmuir 2016, 32,
10808–10814. [CrossRef]
15. Ding, W.; Zhang, J.; Sarker, D.; Hampel, U. The role of microlayer for bubble sliding in nucleate boiling: A new viewpoint for heat
transfer enhancement via surface engineering. Int. J. Heat Mass Transf. 2020, 149, 119239. [CrossRef]
16. Bongarala, M.; Hu, H.; Weibel, J.A.; Garimella, S.V. Microlayer evaporation governs heat transfer enhancement during pool
boiling from microstructured surfaces. Appl. Phys. Lett. 2022, 120, 221602. [CrossRef]
17. Urbano, A.; Tanguy, S.; Huber, G.; Colin, C. Direct numerical simulation of nucleate boiling in micro-layer regime. Int. J. Heat
Mass Transf. 2018, 123, 1128–1137. [CrossRef]
18. Guion, A.; Afkhami, S.; Zaleski, S.; Buongiorno, J. Simulations of microlayer formation in nucleate boiling. Int. J. Heat Mass Transf.
2018, 127, 1271–1284. [CrossRef]
19. Bureš, L.; Sato, Y. Comprehensive simulations of boiling with a resolved microlayer: Validation and sensitivity study. J. Fluid
Mech. 2022, 933, A54. [CrossRef]
20. Lakew, E.; Sarchami, A.; Giustini, G.; Kim, H.; Bellur, K. Thin film evaporation modeling of the liquid microlayer region in a
dewetting water bubble. Fluids 2023, 8, 126. [CrossRef]
21. Kutateladze, S.S.; Mamontova, N.N. Critical heat fluxes in the pool boiling of liquids at reduced pressure. J. Eng. Phys. 1967, 12,
86–90. [CrossRef]
22. Van Stralen, S.J.D.; Zijl, W.; De Vries, D.A. The behaviour of vapour bubbles during growth at subatmospheric pressures. Chem.
Eng. Sci. 1977, 32, 1189–1195. [CrossRef]
23. Surtaev, A.; Serdyukov, V.; Malakhov, I. Effect of subatmospheric pressures on heat transfer, vapor bubbles and dry spots
evolution during water boiling. Exp. Therm. Fluid Sci. 2020, 112, 109974. [CrossRef]
24. Emir, T.; Ourabi, H.; Budakli, M.; Arik, M. Parametric effects on pool boiling heat transfer and critical heat flux: A critical review.
J. Electron. Packag. 2022, 144, 040801. [CrossRef]
25. Mahmoud, M.M.; Karayiannis, T.G. Bubble growth on a smooth metallic surface at atmospheric and sub-atmospheric pressure.
Int. J. Heat Mass Transf. 2023, 209, 124103. [CrossRef]
26. Gerardi, C.; Buongiorno, J.; Hu, L.W.; McKrell, T. Study of bubble growth in water pool boiling through synchronized, infrared
thermometry and high-speed video. Int. J. Heat Mass Transf. 2010, 53, 4185–4192. [CrossRef]
27. Su, G.Y.; Wang, C.; Zhang, L.; Seong, J.H.; Kommajosyula, R.; Phillips, B.; Bucci, M. Investigation of flow boiling heat transfer and
boiling crisis on a rough surface using infrared thermometry. Int. J. Heat Mass Transf. 2020, 160, 120134. [CrossRef]
28. Surtaev, A.; Serdyukov, V.; Malakhov, I.; Safarov, A. Nucleation and bubble evolution in subcooled liquid under pulse heating.
Int. J. Heat Mass Transf. 2021, 169, 120911. [CrossRef]
29. Ronshin, F.V.; Dementiev, Y.A.; Chinnov, E.A. Investigation of dielectric liquid FC-72 boiling in a slit microchannel. Thermophys.
Aeromech. 2022, 29, 975–980. [CrossRef]
30. Surtaev, A.; Koşar, A.; Serdyukov, V.; Malakhov, I. Boiling at subatmospheric pressures on hydrophobic surface: Bubble dynamics
and heat transfer. Int. J. Therm. Sci. 2022, 173, 107423. [CrossRef]

46
Energies 2023, 16, 3918

31. Kangude, P.; Srivastava, A. Experiments to understand bubble base evaporation mechanisms and heat transfer on nano-coated
surfaces of varying wettability under nucleate pool boiling regime. Int. J. Multiph. Flow 2022, 152, 104098. [CrossRef]
32. Može, M.; Hadžić, A.; Zupančič, M.; Golobič, I. Boiling heat transfer enhancement on titanium through nucleation-promoting
morphology and tailored wettability. Int. J. Heat Mass Transf. 2022, 195, 123161. [CrossRef]
33. Surtaev, A.; Malakhov, I.; Serdyukov, V. Explosive vaporization of ethanol on microheater during pulse heating. Heat Transf. Eng.
2022, 44, 502–511. [CrossRef]
34. Sielaff, A.; Mangini, D.; Kabov, O.; Raza, M.Q.; Garivalis, A.I.; Zupančič, M.; Dehaeck, S.; Evgenidis, S.; Jacobs, C.; Van Hoof,
D.; et al. The multiscale boiling investigation on-board the International Space Station: An overview. Appl. Therm. Eng. 2022,
205, 117932. [CrossRef]
35. Rohsenow, W.M. A method of correlating heat-transfer data for surface boiling of liquids. Trans. Am. Soc. Mech. Eng. 1952, 74,
969–975. [CrossRef]
36. Yagov, V.V. Heat transfer with developed nucleate boiling of liquids. Therm. Eng. 1988, 35, 65.
37. Jones, B.J.; McHale, J.P.; Garimella, S.V. The influence of surface roughness on nucleate pool boiling heat transfer. J. Heat Transf.
2009, 131, 121009. [CrossRef]
38. Kalani, A.; Kandlikar, S.G. Enhanced pool boiling with ethanol at subatmospheric pressures for electronics cooling. J. Heat Transf.
2013, 135, 111002. [CrossRef]
39. Kurul, N.; Podowski, M.Z. Multidimensional effects in forced convection subcooled boiling. In Proceedings of the International
Heat Transfer Conference Digital Library, Jerusalem, Israel, 19–24 August 1990; Begel House Inc.: Danbury, CT, USA, 1990.
40. Benjamin, R.J.; Balakrishnan, A.R. Nucleate pool boiling heat transfer of pure liquids at low to moderate heat fluxes. Int. J. Heat
Mass Transf. 1996, 39, 2495–2504. [CrossRef]
41. Malakhov, I.; Seredkin, A.; Chernyavskiy, A.; Serdyukov, V.; Mullyadzanov, R.; Surtaev, A. Deep learning segmentation to analyze
bubble dynamics and heat transfer during boiling at various pressures. Int. J. Multiph. Flow 2023, 162, 104402. [CrossRef]
42. Giustini, G.; Jung, S.; Kim, H.; Walker, S.P. Evaporative thermal resistance and its influence on microscopic bubble growth. Int. J.
Heat Mass Transf. 2016, 101, 733–741. [CrossRef]
43. Chen, Y.; Jin, S.; Yu, B.; Ling, K.; Sun, D.; Zhang, W.; Jiao, K.; Tao, W. Modeling and study of microlayer effects on flow boiling in a
mini-channel. Int. J. Heat Mass Transf. 2023, 208, 124039. [CrossRef]
44. Kossolapov, A.; Phillips, B.; Bucci, M. Can LED lights replace lasers for detailed investigations of boiling phenomena? Int. J.
Multiph. Flow 2021, 135, 103522. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

47
energies
Article
Thermal Performance Analysis of Micro Pin Fin Heat Sinks
under Different Flow Conditions
Jéssica Martha Nunes 1, Jeferson Diehl de Oliveira 2, Jacqueline Biancon Copetti 3, Sameer Sheshrao Gajghate 4 ,
Utsab Banerjee 5 , Sushanta K. Mitra 5 and Elaine Maria Cardoso 1, *

1 Post-Graduation Program in Mechanical Engineering, School of Engineering, UNESP—São Paulo State University,
Av. Brasil, 56, Ilha Solteira 15385-000, SP, Brazil
2 Department of Mechanical Engineering, FSG—University Center, Os Dezoito do Forte, 2366,
Caxias do Sul 95020-472, RS, Brazil
3 LETEF—Laboratory of Thermal and Fluid Dynamic Studies, UNISINOS—University of Vale do Rio dos Sinos,
São Leopoldo 93022-750, RS, Brazil
4 Mechanical Engineering Department, G H Raisoni College of Engineering & Management, Pune 412207,
Maharashtra, India
5 Micro & Nano-Scale Transport Laboratory, Department of Mechanical and Mechatronics Engineering,
Waterloo Institute for Nanotechnology, University of Waterloo, Waterloo, ON N2L 3G1, Canada
* Correspondence: [email protected]

Abstract: Due to microscale effects, the segmented microchannels or micro pin fin heat sinks emerged
as a high thermal management solution. In this context, the present work analyzes the influence
of different heights of square micro pin fins with an aligned array and investigates their influence
on pressure drop and heat transfer behavior. The HFE-7100 is used as the working fluid, and the
pressure drop and surface temperature behavior are analyzed for different mass fluxes and inlet
subcooling. The single-phase flow was analyzed numerically using the computational fluid dynamics
(CFD) software ANSYS FLUENT® for comparing the simulation results with the experimental data,
showing that the highest micro pin fins configuration provides a more uniform and lowest wall
temperature distribution compared to the lowest configuration. There is a good agreement between
Citation: Nunes, J.M.; de Oliveira,
the experimental results and the numerical analysis, with a mean absolute error of 6% for all the
J.D.; Copetti, J.B.; Gajghate, S.S.;
considered parameters. For the two-phase flow condition, experimental tests were performed, and
Banerjee, U.; Mitra, S.K.; Cardoso,
for the highest subcooling, an increase in mass flux causes an enhancement in the heat transfer for
E.M. Thermal Performance Analysis
of Micro Pin Fin Heat Sinks under
low heat flux; by increasing heat flux, there is a gradual predominance of boiling heat transfer over
Different Flow Conditions. Energies convection as the heat transfer mechanism. The pressure drop drastically increases with the vapor
2023, 16, 3175. https://doi.org/ amount flowing into the system, regardless of the pin fin height; the boiling curves for the higher fin
10.3390/en16073175 height show a much smaller slope and a smaller wall superheat than the fin with the smallest height,
and consequently, a high heat transfer performance. A larger region of the heat sink is filled with
Academic Editors: Kyung Chun Kim,
vapor for lower inlet subcooling temperatures, degrading the heat transfer performance compared to
Silvia Ravelli and Vladimir
Serdyukov
higher inlet subcooling temperatures.

Received: 22 February 2023 Keywords: convective boiling; two-phase flow; pin fin geometry; heat transfer coefficient; pressure drop
Revised: 21 March 2023
Accepted: 27 March 2023
Published: 31 March 2023

1. Introduction
An alternative to modifying the configuration of microchannels to minimize the insta-
Copyright: © 2023 by the authors. bilities presented in this system is using segmented or microfinned microchannels (micro
Licensee MDPI, Basel, Switzerland. pin fins). Such compact heat sinks can be used in high-power laser cooling systems, high-
This article is an open access article concentration photovoltaic cell cooling, microreactor cooling, fuel cells, and microchips.
distributed under the terms and In addition to the advantage of segmented microchannels in terms of increasing the heat
conditions of the Creative Commons exchange area to volume ratio, which provides an ability to dissipate higher heat rates, they
Attribution (CC BY) license (https:// can also be manufactured on the chip scales of electronic devices. The recent development
creativecommons.org/licenses/by/
of microfabrication techniques has allowed complex geometries on a reduced scale; thus,
4.0/).

Energies 2023, 16, 3175. https://doi.org/10.3390/en16073175 48 https://www.mdpi.com/journal/energies


Energies 2023, 16, 3175

in recent years, several studies involving heat transfer in micro pin-fin heat sinks have
been carried out in order to characterize the heat transfer mechanisms and to predict the
thermal and fluid dynamic behavior of these systems. The micro fins can have different
shapes and sizes and be arranged in different patterns to improve heat transfer [1–3]. It
is also noteworthy that the ideal spacing of the fins depends on the working fluid and its
subcooling in the system. The mini and microfinned channel arrangements are considered
promising structures for compact heat sinks [4].
Deng et al. [5] proposed a new type of heat sink with pin fin-interconnected reentrant
microchannels (PFIRM) and tested it in convective boiling, using deionized (DI) water
and ethanol as working fluids. Several tests were performed under different subcooling
conditions (40 and 10 ◦ C) and mass fluxes (125 to 300 kg/m2 s). An increase of 39–284%
was observed in the heat transfer coefficient (HTC) for water and 29–220% for ethanol
compared to parallel microchannels. Authors attributed this enhancement to the intercon-
nected microchannels, which provide different paths for the vapor bubbles reducing the
confinement effect. In addition, the interconnected spaces provided ideal conditions for the
nucleation of vapor bubbles, contributing to the heat transfer improvement for the PFIRM.
For pressure drop, Deng et al. [5] reported an increase with increasing heat flux and vapor
quality; moreover, the mass flux strongly influenced pressure drop at moderated and high
heat fluxes.
Recently, Asrar et al. [6] conducted an experimental investigation of convective boiling
using R245fa in micro gaps improved with the arrays of micro pin fins made of silicon
(cylindrical pin fins 150 μm in diameter and 200 μm of interfin space in a staggered
arrangement). Several tests were performed at different mass flux conditions (between 781
and 5210 kg/m2 s) and inlet temperatures (between 13 and 18 ◦ C). The authors reported
that HTC increased with increasing mass flux for the single-phase flow regime. For the
two-phase flow regime, they compared the results with their previous works [7]; the new
device showed better thermal performance than the previous one. Regarding pressure drop
and vapor quality, Asrar et al. [6] found the same behavior as Woodcock et al. [8] and Chien
et al. [9], in which these parameters were independent of the heat flux in the single-phase
regime but increased remarkably with the intensification of convective boiling.
The scientific community has extensively studied heat transfer in segmented mi-
crochannels. However, there are still several challenges that require further research to
understand and optimize the process, such as (i) the good balance between heat transfer
and pressure drop to have an efficient heat sink, (ii) the appropriate material for the heat
sink and the micro pin fins in order to have high thermal conductivity and good mechanical
strength, and (iii) the optimal configuration for a micro pin fin heat sink to improve the
heat transfer capacity.
Many of the developed works take into account different dimensions and configura-
tions of micro pin fins to understand the physical mechanisms responsible for heat transfer
enhancement in an attempt to develop models to be applied on an industrial scale capable
of predicting the heat transfer coefficient, the behavior of the critical heat flux, and the
pressure drop. As mentioned by Jung et al. [10], since water is more widely available and
has superior thermophysical properties, most studies are conducted with it; however, its
high electrical conductivity limits its use for embedded cooling.
In this context, the current work investigates the thermal performance and pressure
drop of HFE-7100 in micro pin fin heat sinks with different heights and their influence
on pressure drop and heat transfer behavior. The effects of geometrical parameters and
operating conditions on the thermal performance and pressure drop were analyzed experi-
mentally and, for single-phase flow, numerically. The computational model used in the
current study validates the thermal performance and pressure drop determined from the
experiments for single-phase flow. The current study contributes to better comprehending
heat removal capability, factors impacting heat transfer performance, and mechanisms
responsible for enhancing heat transfer in such compact heat exchangers.

49
Energies 2023, 16, 3175

2. Materials and Methods


2.1. Experimental Apparatus
Figure 1 shows the experimental apparatus used in the present study. The working
fluid is pumped from a reservoir to the flow loop; the HFE-7100 flow rate is set by a Coriolis
mass flow meter (Yokogawa ROTAMASS Total Insight with 0.2% mass flow accuracy)
installed just upstream of the preheater (consisting of a horizontal copper tube heated by an
electrical tape resistance). There is a bypass line used for the test facility maintenance. The
pressure drop between inlet and outlet plenums is measured by two pressure transducers
(OMEGA PX309 model, with 0.25% accuracy). The flow temperature is measured using
previously calibrated K-type thermocouples (uncertainty of 0.3 ◦ C) in the inlet and outlet
plenums (both in contact with the fluid). The working fluid is cooled by a condenser and
then returned to the reservoir, as shown in Figure 1.

Figure 1. Schematic diagram of the experimental apparatus.

The design of the heat sink test section in different views is shown in Figure 2, including
a cutaway to view internal details and elemental descriptions (Figure 2b). Five holes
(1 mm diameter) were drilled in the copper block to accommodate K-type thermocouples
(Figure 2a, A) to determine the wall temperature and verify the one-dimensional heat
conduction along the copper block. The heat flux is provided by electrical resistance
(cartridge type, 250 W/220 V) embedded in the copper block (Figure 2a, B) and controlled
by a DC power supply.
The one-dimensional conduction law of Fourier was determined with the temperature
readings from the thermocouples fixed along the vertical direction of the copper block;
therefore, the wall temperature was obtained by extrapolating the linear temperature profile,
which exhibits an R-square error of approximately 1.0 regardless of heat flux (Figure 2c).
Moreover, the heat losses were lower than 15%, corroborating the one-dimensional steady-
state heat flux assumption.
The heat sink (Figure 2) consists of a copper block with a 20 × 15 mm footprint with
972 micro pin fins. The square micro pin fins (300 μm in width and 250 μm of interfin space)
were manufactured using a CNC precision milling machine (Hermile, model C800U). The
micro fins were manufactured in an aligned array with the fluid flow direction (Figure 3),
and different heights (H) of micro pin fins—160 μm named S1 and 350 μm named S2—were
analyzed in the current work.

50
Energies 2023, 16, 3175

(a) (b)

(c)

Figure 2. Design of the microfinned heat sink. (a) Isometric view; (b) front view with internal details
(measurements in mm); (c) linear temperature profiles used to estimate the wall temperatures.

Figure 3. The constructive array of the micro pin fins.

As shown in Figure 2b, the thermal insulation is made with ceramic and polytetraflu-
oroethylene (PTFE); the working fluid is not heated before contacting the micro pin fins
since the inlet and outlet plenums are machined on the PTFE with 10 × 15 × 10 mm. Two
K-type thermocouples, one in the inlet and another in the outlet, measure the working
fluid temperature (Ti and To , respectively). Flow homogenization channels with a depth
of 0.75 mm were manufactured between the plenums and the heat sink to minimize flow
entrance turbulence. Flow visualization (using a high-speed camera Photron SA3 model
with 1000 fps and 1024 × 1024 resolution) is allowed by a polycarbonate plate (8 mm thick)
covering the heat sink.

51
Energies 2023, 16, 3175

The geometric characterization of the micro pin fin heat sinks was performed by Zeiss®
SteREO DiscoveryV8 and SEM EVO LS15 Zeiss® (Table 1).

Table 1. Structural characterization of the micro pin fin heat sinks.

STEREO
Surface SEM (100×)
Top View Side View

S1
H = 160 μm

S2
H = 350 μm

The experimental uncertainty was calculated using the free package developed in
Python, called Uncertainties (© 2010–2016, EOL), based on the Taylor series method and
standardized by the Bureau International des Poids et Mesures (BIPM). Consequently, for
all tests carried out in the current study, the uncertainty of the heat flux, the heat transfer
coefficient, and pressure drop varied from 4 to 16%, 7 to 21%, and 3 to 9%, respectively. It is
worth mentioning that all analyses take into account the effective heat flux, determined by
subtracting the heat loss to the surroundings from the power supplied; in the current study,
the heat losses are less than 22% for all tests performed.

2.2. Experimental Procedure


The consistency analysis aims to verify the coherence of the results obtained experimen-
tally; thus, the results for the single-phase flow regime were compared to those obtained
from a numerical analysis considering the same conditions. The simulation was based on
the mass, momentum, and energy conservation equations with the second-order upwind
scheme for energy and pressure and the first-order for momentum [11]. The fluid flow was
assumed to be steady-state, incompressible, and laminar, and it was solved by adopting
the Finite-Volume Method implemented in ANSYS® Fluent 2020 R2. The computational
domain with appropriate boundary conditions is shown in Figure 4.
As a reference pressure, atmospheric pressure was defined; considering the character-
istics of a low-pressure system, the outlet pressure was set up as zero. The no-slip condition
was considered on all the surfaces. The input parameters for the simulation, such as inlet
mass flux and the dimensions of the micro pin fins, were taken from the experimental
approach. The heat flux was distributed through the micro pin fins, except for the top side,
which was thermally insulated by a polycarbonate piece; for the inlet and outlet plenums,
adiabatic wall conditions were considered.
The hex-dominant meshing grid scheme with a free-face mesh type combining trian-
gles and quadrilaterals was used to mesh the systems (for both cases), as shown in Figure 5.
The mesh was accomplished in the meshing module with minimum mesh orthogonality of
0.311 (S1, H = 160 μm) and 0.346 (S2, H = 350 μm) and a maximum skewness of 0.56 (S1,
H = 160 μm) and 0.61 (S2, H = 350 μm).

52
Energies 2023, 16, 3175

Figure 4. Computational domain of the heat sink with main boundary conditions.

Figure 5. Grid view with fluid domain on the right side.

The convergence occurred for meshes with 511.09 k elements; the finer mesh was
achieved when residuals were less than 10−5 for the continuity equation and 10−6 for
momentum and energy equations. The simulations used the segregated algorithm with the
SIMPLE algorithm for pressure-velocity coupling.
A study on grid independence was conducted based on wall temperature as a criterion
to ensure the results were independent of the mesh. This analysis consisted of three
different meshes (around 120 k; 370 k; 550 k elements), aiming to obtain a heated wall
temperature range of a maximum of 1.2 K (less than 10% of the maximum wall temperature,
according to [12]); it was noticed that the difference in wall temperature was around 1.2 K
between the last two meshes. Hence, the mesh chosen aimed to save computational time.
|φexp −φ num |
The mean absolute errors (MAE = N1 ∑1N φexp × 100% ) of the surface tempera-
ture and total pressure drop were 5.6% and 7.1%, respectively. The computational results
were consistent with the experimental data for the heat transfer coefficient; for both S1
and S2, the MAE was 5%. Therefore, the mean absolute errors of the experimental and
simulation data for pressure drop and heat transfer coefficient are within the experimental
uncertainty range.
After the validation analysis for the single-phase flow, two-phase flow tests were
performed for two different subcooling values, 10 ◦ C and 20 ◦ C, for mass fluxes of 1000 and
1200 kg/m2 s, and for different footprint heat fluxes from 10 kW/m2 to the system limit,
characterized by intense instability in the flow (reverse flow). The gear pump’s rotation was
set to achieve the desired mass flux; the preheater was adjusted until its outlet temperature
was equal to the desired subcooling. A data acquisition system (Agilent 34970A) recorded
the data every 2 s after the system achieved the steady-state regime, characterized by tem-
perature variations lower than the thermocouples uncertainties (±0.3 ◦ C). At least 250 data
points were recorded, corresponding to 500 s steady-state. The pressure, temperatures,
mass flux, and electrical voltage are constantly monitored. Flow visualization was carried

53
Energies 2023, 16, 3175

out using a high-speed camera. The same procedure is adopted during all the experimental
tests to ensure repeatability.

2.3. Data Reduction


The heat transfer coefficient is calculated based on Equations (1)–(5), similar to the
approach adopted by Prajapati et al. [13]:
. . .
Qloss = Qin − m·c p ( To − Ti ) (1)
.
where m corresponds to the mass flow rate [kg/s]; cp to the specific heat capacity [J/kg·K];
and Ti and To are the coolant temperature at the inlet and outlet, respectively. In the current
.
study, heat loss (Qloss ) varied from 15 to 30% over the range of varying parameters. The
heat flux, q , dissipated by the test section is given by:
. .
 Qin − Qloss
q f ootprint = (2)
Ap

where Ap is the footprint area of the heating surface. The effective heat flux, qe f f [W/m2 ],
based on the total surface area in contact with the working fluid (At ), is calculated by:
. .
 Qin − Qloss
qe f f = (3)
At

In order to calculate the total surface area, At , the fin parameters and efficiency (η)
concepts have been calculated considering the adiabatic fin tip, since a polycarbonate plate
is used to cover the heat sink. Thus, At is given by Equation (4), where N is the total number
of micropillars.  
At = A p − N · Ac + η · N · Pma · H (4)
where Ac is the cross-sectional area, Pma is the pin fin perimeter, and H is the height of the
micro pin fins.
Therefore, it is possible to calculate the heat transfer coefficient or HTC (h) through
Equation (5), where Tw is the average temperature of the heat sink provided by three K-type
thermocouples fixed within the heat sink wall. The Tf is the average temperature of the
fluid given by the same procedure as Leão et al. [14]

qe f f
h= (5)
Tw − T f

Pressure transducers (at inlet and outlet plenums, Pi and Po , respectively) measure the
pressure drop in the region between the inlet and outlet plenums; thus, the pressure drop
through the microchannels is given by ΔP = ( Pi − Po ) − ΔPcontraction − ΔPexpansion where
the pressure drop due to contraction and expansion is obtained by the method described in
Chalfi and Ghiaasiaan [15].

3. Results and Discussion


3.1. Effect of the Inlet Subcooling Temperature
Figure 6 shows the effect of different inlet subcooling temperatures (10 and 20 ◦ C)
on the flow boiling heat transfer for both surfaces (S1 and S2). The increase in subcooling
shifted the boiling curve to the left regardless of micro pin fin height. The HTC continuously
increased with heat flux for all mass fluxes and higher inlet subcooling temperature, while
the HTC slightly decreased with high heat fluxes for lower inlet subcooling temperature.
According to Yin et al. [16], such HTC behavior is due to the flow pattern transition into a
confined annular flow, where partial dryout occurs on the surface as heat flux increases,
leading to the rise in the wall temperature (being more pronounced for S1).

54
Energies 2023, 16, 3175

 
(a)

 
(b)
Figure 6. Effect of inlet subcooling temperature on flow boiling heat transfer of HFE-7100 for S1 and
S2. (a) G = 1000 kg/m2 s; (b) G = 1200 kg/m2 s.

Analyzing the boiling curves of samples S1 and S2 is possible to observe the beginning
of the nucleate boiling regime, indicated in Figure 6 as ONB (Onset of Nucleate Boiling),
and characterized by the sudden change in the slope of the boiling curve, reducing the
surface temperature.
Figure 7 shows the effect of different inlet subcooling temperatures (10 and 20 ◦ C) on
the pressure drops for both surfaces (S1 and S2). One can observe that the effect of inlet
subcooling temperatures on both surfaces’ pressure drops in the single-phase flow region
(for q < 30 kW/m2 ) is negligible. However, in the case of the two-phase flow region, the
pressure drop becomes more significant as the inlet subcooling temperature decreases,
regardless of the mass flux and micro pin fin height; a lower inlet subcooling temperature
leads to a higher vapor quality through the heat sink, which increases the pressure drop.

55
Energies 2023, 16, 3175


(a)

 
(b)
Figure 7. Effect of inlet subcooling on the pressure drops for S1 and S2. (a) G = 1000 kg/m2 s;
(b) G = 1200 kg/m2 s.

3.2. Effect of the Mass Flux


Figure 8 shows the effect of different mass fluxes on the boiling curves for S1 and S2,
with different subcooling temperatures at the inlet of the heat sink. The influence of mass
flux, G, on the convective flow boiling heat transfer was negligible for the inlet subcooling
of 10 ◦ C and the lowest micro pin fin height (S1). On the contrary, for the highest fin height,
the fluid has more space to flow between the fins, and the convective effects (mass flux
influences) are more pronounced in the single-phase flow region.
For the inlet subcooling of 20 ◦ C, an increase in the mass flux shifted the curves to
the left, characterized by an HTC enhancement. Cheng and Wu [17] indicated a gradual
predominance of boiling heat transfer over convection as heat flux increases; furthermore,
the micro pin fins induced flow turbulence and strengthened convection heat transfer, the
primary heat dissipation component in subcooled convective boiling [18].
For both S1 and S2, increasing the mass flux increased the pressure drops for low
heat flux values (single-phase flow region); however, no significant influence of mass flux
on pressure drop was observed in the single-phase flow region for both inlet subcooling
temperatures (10 and 20 ◦ C). As the heat flux increased (two-phase flow region), the
pressure drop became more pronounced due to the vapor mass flowing through the heat
sink; thus, the pressure drop was more influenced by the void fraction than by mass flux.

56
Energies 2023, 16, 3175

 
(a)

 
(b)

Figure 8. Effects of mass flux on the flow boiling heat transfer of HFE-7100. (a) ΔTsub = 10 ◦ C;
(b) ΔTsub = 20 ◦ C.

3.3. Effect of the Fin Height


Figure 9 shows the effect of pin fin height on the boiling curves for different mass
fluxes and a subcooling inlet temperature of 10 ◦ C. Considering the effective heat exchange
area, we can infer that the increase in the effective area leads to an increase in the HTC,
characterized by the shift of the boiling curve to the left. The same was reported by
Kiyomura et al. [19], who evaluated different configurations of micro fin surfaces during
pool boiling of the HFE-7100. One can observe in Figure 9 that S2 presents a better HTC,
since its effective heat exchange area is approximately 55% greater than S1.
In order to discuss the flow boiling behavior, Figure 10 presents the boiling curve for
S1 with G = 1200 kg/m2 s and subcooling of 20 ◦ C, with the respective visualization points.
It is worth mentioning that similar behavior was observed for all test conditions. Flow
boiling videos under these conditions can be found in the Supplementary Material.

57
Energies 2023, 16, 3175

(a) (b)
Figure 9. Effect of pin fin height on flow boiling heat transfer of HFE-7100 for ΔTsub = 10 ◦ C.
(a) G = 1000 kg/m2 s; (b) G = 1200 kg/m2 s.

Figure 10. Boiling curve and high-speed camera images for S1. G = 1200 kg/m2 s and ΔTsub = 20 ◦ C.

Initially, the single-phase flow regime is predominant at lower heat flux, with no vapor
bubbles (point (a), Figure 10). By increasing heat flux, isolated vapor bubbles nucleate
preferentially between the adjacent fins, even though the working fluid temperature is lower
than the saturation temperature, i.e., subcooled boiling condition (point (b), Figure 10). In
the nucleate boiling region, after the ONB, nucleation sites are activated over the entire
heating surface (point (c), Figure 10), increasing the departure frequency and the coalescence
of vapor bubbles near the heat sink outlet. For high heat fluxes, the vapor core fills the
entire length of the heat sink (point (d), Figure 10), and the annular flow regime becomes
pronounced. A high void fraction is observed at the outlet of the heat sink, promoting
thermal instabilities, a high pressure drop and the occurrence of reverse flow, which is
mainly observed for lower inlet subcooling temperature (reverse flow visualization).

4. Conclusions
The current work experimentally studied the thermal and fluid dynamic behaviors, in
terms of heat transfer coefficient and pressure drop, of convective boiling using HFE-7100

58
Energies 2023, 16, 3175

as the working fluid in a heat sink based on square micro pin fins. Different micro pin fins
were tested (heights of 160 and 350 μm in an aligned array) at different mass fluxes (1000
and 1200 kg/m2 s) and two levels of inlet subcooling temperatures (10 and 20 ◦ C). The
boiling heat transfer and pressure drop behaviors were evaluated for each test condition.
The visualization of the experimental tests was performed using a high-speed camera to
observe the transition from single-phase to two-phase flow and to identify possible flow
patterns and the occurrence of reverse flow. The main conclusions are summarized below:
 As the mass flux increases, HTC increases in the region where the effects of forced
convection are dominant for each sample. However, when the effects of nucleate
boiling overlap, the increase in mass flux does not guarantee a gain in HTC, especially
for aligned arrays.
 The lower the inlet subcooling temperature, the lower the heat flux for the ONB
occurrence, and a larger region of the heat sink is filled with vapor, which can promote
the dryout incipience (decreasing the maximum heat flux).
 With a lower mass flux and inlet subcooling, the system becomes more sensitive to
the effects of nucleate boiling, with significant gains in HTC due to the phase-change
heat transfer (for S1 with G = 1000 kg/m2 s and ΔTsub = 10 ◦ C, the HTC was increased
about 39% compared to ΔTsub = 20 ◦ C for a heat flux of 30 kW/m2 ). However, this
can lead to the early dryout process.
 Pressure drop increases substantially with an increase of vapor amount flowing into
the heat sink, which becomes more pronounced for lower subcooling, leading to the
fluid dynamic limit of the system at lower heat fluxes compared to higher subcooling.
 An increase in the effective area leads to an increase in the HTC; thus, the taller the
micro pin fins, the higher the heat exchange area, leading to an HTC enhancement.
 The reverse flow occurrence was observed more intensely for the lowest inlet subcool-
ing temperature; the high vapor core acts as a barrier to the flow, degrading the HTC,
increasing the pressure drop, and causing thermal and fluid dynamic instabilities.
This study indicates that further attention must be given concerning physical param-
eters (related to surface and working fluid) to the development of new technologies for
thermal management systems. An isolated analysis of the effects of surface characteristics
or flow parameters is not sufficient to explain the HTC and pressure drop behavior. The
optimal configuration for a micro pin fin heat sink will depend on several factors, including
the heat transfer requirements, the fabrication process, and the fluid flow properties. More
analyses can be conducted in future works, such as developing new correlations to evaluate
the HTC; for that, an extensive experimental database is needed considering different
design configurations and operating conditions.

Supplementary Materials: The following supporting information can be downloaded at https://www.


mdpi.com/article/10.3390/en16073175/s1: Video S1: point (a), Figure 10; Video S2: point (b), Figure 10;
Video S3: point (c), Figure 10; Video S4: point (d), Figure 10; Video S5: reverse flow visualization.
Author Contributions: All authors contributed equally to developing the manuscript. All authors
have read and agreed to the published version of the manuscript.
Funding: This research was funded by Conselho Nacional de Desenvolvimento Científico e Tec-
nológico (CNPq), grants numbers 458702/2014-5 and 309848/2020-2, and Fundação de Amparo
à Pesquisa do Estado de São Paulo (FAPESP), grants numbers 2013/15431-7, 2019/02566-8, and
2022/03946-1.
Data Availability Statement: The data supporting this study’s findings are available upon request.
Acknowledgments: The authors are grateful for the financial support from the UNESP, CAPES,
CNPq, and FAPESP. The authors also thank Alessandro Roger Rodrigues from Escola de Engenharia
de São Carlos/USP and Ricardo Arai from IFSP/São Carlos for their important contributions to
this work.

59
Energies 2023, 16, 3175

Conflicts of Interest: The authors declare no conflict of interest, and the funders had no role in the
study’s design; in the collection, analyses, or interpretation of data; in the writing of the manuscript;
or in the decision to publish the results.

References
1. Tullius, J.; Tullius, T.; Bayazitoglu, Y. Optimization of short micro pin fins in minichannels. Int. J. Heat Mass Transf. 2012, 55,
3921–3932. [CrossRef]
2. Liang, G.; Mudawar, I. Review of pool boiling enhancement by surface modification. Int. J. Heat Mass Transf. 2019, 128, 892–933.
[CrossRef]
3. Li, W.; Dai, R.; Zeng, M.; Wang, Q. Review of two types of surface modification on pool boiling enhancement: Passive and active.
Renew. Sustain. Energy Rev. 2020, 130, 109926. [CrossRef]
4. McNeil, D.A.; Raeisi, A.H.; Kew, P.A.; Hamed, R.S. An investigation into flow boiling heat transfer and pressure drop in a
pin–finned heat sink. Int. J. Multiph. Flow 2014, 67, 65–84. [CrossRef]
5. Deng, D.; Chen, L.; Wan, W.; Fu, T.; Huang, X. Flow boiling performance in pin fin-interconnected reentrant microchannels heat
sink in different operational conditions. Appl. Therm. Eng. 2019, 150, 1260–1272. [CrossRef]
6. Asrar, P.; Ghiaasiaan, S.M.; Joshi, Y.K. Two-Phase Heat Transfer and Flow Regimes in Pin Fin-Enhanced Microgaps—Effect of Pin
Spacing. ASME J. Heat Transf. 2021, 143, 023001. [CrossRef]
7. Asrar, P.; Zhang, X.; Green, C.E.; Bakir, M.; Joshi, Y.K. Flow boiling of R245fa in a microgap with staggered circular cylindrical pin
fins. Int. J. Heat Mass Transf. 2018, 121, 329–342. [CrossRef]
8. Woodcock, C.; Yu, X.; Plawsky, J.; Peles, Y. Piranha Pin Fin (PPF)—Advanced flow boiling microstructures with low surface
tension dielectric fluids. Int. J. Heat Mass Transf. 2015, 90, 591–604. [CrossRef]
9. Chien, L.H.; Cheng, Y.T.; Lai, Y.L.; Yan, W.M.; Ghalambaz, M. Experimental and numerical study on convective boiling in a
staggered array of micro pin-fin microgap. Int. J. Heat Mass Transf. 2020, 149, 119203. [CrossRef]
10. Jung, D.; Lee, H.; Kong, D.; Cho, E.; Jung, K.W.; Kharangate, C.R.; Iyengar, M.; Malone, C.; Asheghi, M.; Lee, H.; et al. Thermal
design and management of micro-pin fin heat sinks for energy-efficient three-dimensional stacked integrated circuits. Int. J. Heat
Mass Transf. 2021, 175, 121192. [CrossRef]
11. Ortegon, J.A.A.; Souza, R.R.; Silva, J.B.C.; Cardoso, E.M. Analytical, experimental, and numerical analysis of a microchannel
cooling system for high-concentration photovoltaic cells. J. Braz. Soc. Mech. Sci. Eng. 2019, 41, 255. [CrossRef]
12. Computational Fluid Dynamics Committee. Guide for the Verification and Validation of Computational Fluid Dynamics Simulations
(AIAA G-077-1998(2002)); American Institute of Aeronautics and Astronautics, Inc.: Washington, DC, USA, 1998.
13. Prajapati, Y.K.; Pathak, M.; Khan, M.K. Bubble dynamics and flow boiling characteristics in three different microchannel
configurations. Int. J. Therm. Sci. 2017, 112, 371–382. [CrossRef]
14. Leão, H.L.S.L.; Nascimento, F.J.; Ribatski, G. Flow boiling heat transfer of r407c in a microchannels based heat spreader. Exp.
Therm. Fluid Sci. 2014, 59, 140–151. [CrossRef]
15. Chalfi, T.Y.; Ghiaasiaan, S. Pressure drop caused by flow area changes in capillaries under low flow conditions. Int. J. Multiph.
Flow 2008, 34, 2–12. [CrossRef]
16. Yin, L.; Chauhan, A.; Recinella, A.; Jia, L.; Kandlikar, S.G. Subcooled flow boiling in an expanding microgap with a hybrid
microstructured surface. Int. J. Heat Mass Transf. 2020, 151, 119379. [CrossRef]
17. Cheng, X.; Wu, H. Improved flow boiling performance in high-aspect-ratio interconnected microchannels. Int. J. Heat Mass Transf.
2021, 165, 120627. [CrossRef]
18. Yin, L.; Jiang, P.; Xu, R.; Hu, H. Water flow boiling in a partially modified microgap with shortened micro pin fins. Int. J. Heat
Mass Transf. 2020, 155, 119819. [CrossRef]
19. Kiyomura, I.S.; Nunes, J.M.; de Souza, R.R.; Gajghate, S.S.; Bhaumik, S.; Cardoso, E.M. Effect of microfin surfaces on boiling heat
transfer using HFE-7100 as working fluid. J. Braz. Soc. Mech. Sci. Eng. 2020, 42, 366. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

60
energies
Article
Experimental Study of Mass Transfer in a Plug Regime of
Immiscible Liquid–Liquid Flow in a T-Shaped Microchannel
Semyon Vostretsov 1,2 , Anna Yagodnitsyna 1,2, *, Alexander Kovalev 1,2 and Artur Bilsky 1, *

1 Kutateladze Institute of Thermophysics SB RAS, 630090 Novosibirsk, Russia


2 Physics Department, Novosibirsk State University, 630090 Novosibirsk, Russia
* Correspondence: [email protected] (A.Y.); [email protected] (A.B.)

Abstract: In the presented work, the influence of parameters such as the total flow rate of phases,
the ratio of flow rates, and residence time on mass transfer during the two-phase flow of immiscible
liquids in a T-shaped microchannel was investigated using the micro-LIF technique. The study
focused on the plug flow regime, where a 70% water–glycerol solution was used as the dispersed
phase, and tri-n-butyl phosphate (TBP) was used as the carrier phase. We determined the transition
boundary between the dispersed and parallel flow patterns and calculated the plug length and
velocities to develop a mass transfer model. Furthermore, we measured the partition coefficient for
the set of liquids used in the experiments and analyzed the concentration fields inside the slugs of
the continuous phase at various distances downstream of the T-junction. Using the obtained data, we
determined the extraction efficiency and overall volumetric mass transfer coefficient and established
dependencies demonstrating the effect of the flow-rate ratio, total flow rate, and the residence time on
mass transfer rate and extraction efficiency. Finally, we developed a model for the overall volumetric
mass transfer coefficient corresponding to the set of liquids used with an R-squared value of 0.966.

Keywords: immiscible liquids; extraction; mass transfer; micro-LIF; microchannel

Citation: Vostretsov, S.; 1. Introduction


Yagodnitsyna, A.; Kovalev, A.; Bilsky, Microchannels are ubiquitous in various fields under their unique properties. Owing
A. Experimental Study of Mass
to their small size, microchannels exhibit a high surface-to-volume ratio, which allows for
Transfer in a Plug Regime of
achieving high efficiency in heat and mass transfer processes. Therefore, microchannels
Immiscible Liquid–Liquid Flow in a
have found their application as micro-heat exchangers, micromixers, microextractors [1–5],
T-Shaped Microchannel. Energies
and integrated systems for screening and sorting biological objects [6]. Standard mass
2023, 16, 4059. https://doi.org/
transfer equipment, such as columns and settlers, used in these processes have several
10.3390/en16104059
significant disadvantages, including high reagent consumption, low process intensity, and
Academic Editor: Dmitry Eskin potential danger to personnel in the case of leaks. In contrast, microchannels are devoid of
Received: 10 April 2023
these disadvantages and can intensify mass transfer. Therefore, studying the influence of
Revised: 10 May 2023
various parameters on mass transfer in microchannels is necessary.
Accepted: 11 May 2023
The increasing demand for low-carbon, cost-effective energy sources has led to the in-
Published: 12 May 2023 tensive development of new power plants worldwide [7,8]. As a result, the amount of spent
nuclear fuel is constantly growing, and its regeneration is most acute for radiochemical
technology. Typically, uranium(VI) and plutonium(IV) are recovered from nitric acid solu-
tions of spent nuclear fuel through the PUREX process, with mixtures of organic solvents
Copyright: © 2023 by the authors. with tributylphosphate (TBP) as an extractant [9]. By using microchannels as reservoirs
Licensee MDPI, Basel, Switzerland. for extraction, the technology can be improved in terms of the safety and efficiency of the
This article is an open access article extraction process.
distributed under the terms and The slug regime is of great interest for extraction in a microchannel with a two-phase
conditions of the Creative Commons flow of immiscible liquids. This is because plugs and slugs provide a greater interfacial
Attribution (CC BY) license (https://
area-to-volume ratio. In addition, due to vortices in slugs and plugs, mass transfer intensi-
creativecommons.org/licenses/by/
fication occurs both inside the phases and between them [10]. Thus, various models have
4.0/).

Energies 2023, 16, 4059. https://doi.org/10.3390/en16104059 61 https://www.mdpi.com/journal/energies


Energies 2023, 16, 4059

been developed to assess mass transfer for gas–liquid [11,12] and liquid–liquid [13–16]
systems in this flow regime. Based on the results obtained in [17–20], conclusions can
be drawn regarding the dependence of mass transfer efficiency in microchannels on the
residence time, temperature, and viscosity of the liquids used, as well as the shape of the
microchannel. A review conducted by Ganguli and Pandit [21] summarized the effect of
microchannel size on mass transfer. It was concluded that reducing the size of the cross
section of the microchannel increases the intensity of mass transfer. Kashid et al. [17]
investigated the influence of microchannel geometry on mass transfer efficiency and found
that the microchannels with obstacles inside could intensify mass transfer due to additional
flow circulation.
The influence of temperature on mass transfer in a liquid–liquid system was described
by Zhang et al. [18]. The highest mass transfer coefficients corresponded to the highest
temperatures studied, however, circulation within the phases decreased. The authors
suggest that this effect was caused by the lower viscosity of the continuous phase, which
led to a lower shear stress acting on the dispersed phase. The influence of the viscosity
of working liquids on mass transfer was studied in [18,19]. Based on experiments using
various sets of liquids, it was found that the viscosity of the dispersed phase had a negligible
effect on mass transfer, in contrast to the viscosity of the continuous phase.
The effect of residence time on mass transfer in the slug flow regime was investigated
by Angeli et al. [20]. The authors concluded that mass transfer mainly occurs at short resi-
dence times. At longer times, it becomes noticeable that the extraction efficiency increases
more with a higher total flow rate, indicating the influence of the flow structure inside the
slugs and plugs on mass transfer. Thus, it is possible to distinguish two main stages of mass
transfer in the two-phase flow of immiscible liquids. At short residence times, there is a
high solute concentration gradient between phases, so the diffusion coefficient determines
the mass transfer. At long residence times, the layer near the interfacial boundary becomes
poorer, and the hydrodynamics of the flow becomes decisive to ensure the inflow of solute
to the interfacial boundary.
There are different approaches to assessing mass transfer in microchannels which
can be classified as local and integral techniques. In local measurements, the values
characterizing the mass transfer are determined based on concentration fields obtained
by the microresolution laser-induced fluorescence technique (micro-LIF) or colorimetric
techniques. In micro-LIF measurements, the transferred solute itself is used as a fluores-
cent dye [20,22]. In colorimetric techniques, the dye added to the flow changes color due
to the reaction with the transferred solute or pH change occurring when the compound
in the flow reacts with the solute [14,19,23]. However, integral measurements of the
solute concentration in phases are often performed. For example, gas chromatography
was used by Kashid et al. [17]. After phase separation, chromatography of the exam-
ined liquid is carried out, and the amount of solute is determined. Another method of
integral measurement is UV spectroscopy, performed by Priest et al. [5] but, in this case,
it is necessary to separate the phases before analysis. Moreover, integral techniques
are not applicable if the residence time in microchannels is much lower than the phase
separation time.
Currently, existing studies on mass transfer in the flow of immiscible liquids in mi-
crochannels aim to determine the influence of parameters such as microchannel geometry,
flow pattern, total flow rate, the flow rate ratio of the continuous and dispersed phases, and
residence time. However, mass transfer studies using local measurement techniques are
limited to a narrow range of flow rates and phase flow ratios. The present work uses a micro-
LIF technique to study the effect of flow parameters and residence time on the efficiency of
mass transfer in the slug flow regime of immiscible liquids in a T-shaped microchannel.

62
Energies 2023, 16, 4059

2. Materials and Methods


2.1. Mass Transfer Characteristics
Extraction is a process of transferring a solute between two non-miscible media. In
our experiment, the dye is extracted from the dispersed phase into the continuous phase,
so the equations presented in this article refer to the slugs of the continuous phase. The
main parameters studied during mass transfer are the overall volumetric mass transfer
coefficient (1) and extraction efficiency (2):
 
Ccin −Cc∗
ln Ccout −Cc∗
kL a = , (1)
τ

Ccout
%E = , (2)
Cc∗
where Ccin and Ccout —the solute concentration in the continuous phase at the microchannel
inlet and outlet, respectively, Cc∗ —equilibrium concentration, and τ—residence time.
The ratio of equilibrium concentrations of the solute in the phases is called partition
coefficient K: ∗
Corg C∗
K = ∗ = c∗ (3)
Caq Cd
where subscripts org and aq denote organic and aqueous phases.
The equilibrium concentration for various flow rates is determined from the following
considerations. Initially, the solute is presented only in the dispersed phase. Its amount
Nin entering the channel is determined by the flow rate of the dispersed phase Qd and the
initial concentration of the solute in it C0 :

Nin = Qd ·C0 (4)

At the outlet of the channel, the solute is presented in both phases in concentrations
Cc and Cd . Due to the conservation of the amount of solute at the inlet and outlet of the
microchannel, it is possible to write:

Qd ·C0 = Qd ·Cd + Qc ·Cc (5)

Assuming that equilibrium is established between the phases in the microchannel


with respect to mass transfer, we substitute the equilibrium concentrations and express the
equilibrium concentration in the continuous phase:

C0
Cc∗ =   (6)
Qc
1
K + Qd

Thus, to obtain the extraction efficiency and the overall volumetric mass transfer
coefficient, it is necessary to have values of the partition coefficient K, the concentration of
the solute in the inlet and outlet of the measurement area of the microchannel Cin and C out ,
and the residence time τ.

2.2. Experimental Setup and Techniques


The mass transfer was studied in a plug regime of immiscible liquids in a T-shaped
microchannel with a square cross-section of 370 μm × 200 μm. The schemes of the ex-
perimental setup and the microchannel are presented in Figure 1. The experimental
setup consisted of a Carl Zeiss Axio Observer.Z1 microscope with a lens magnification of
M = 10× and a numerical aperture of NA = 0.25. A T-shaped microchannel was mounted
on the microscope stage. Liquid flow rates were set using a KDS Gemini 88 double syringe
pump with an accuracy of 0.3%. The experiments were performed at room temperature
T = 23 ◦ C.

63
Energies 2023, 16, 4059

Figure 1. The schemes of the experimental setup and a T-shaped microchannel.

The microresolution laser-induced fluorescence technique was applied to measure the


solute concentration fields. The method is based on the relation between dye concentration
in the fluid and its fluorescence intensity. In the standard LIF method at the macroscale,
the flow is illuminated with a laser sheet, hence the concentration field is determined
by the emission intensity of the fluorescent dye in the measurement plane. However,
creating a thin laser sheet is impossible on the microscale; therefore, the entire flow volume
was illuminated by a mercury lamp, obtaining the concentration field averaged over the
channel depth. Rhodamine 6G fluorescent dye was used as a solute added initially to the
dispersed phase. The solute concentration measurements were performed in the slugs
of the continuous phase since the measurements in the plugs of the dispersed phase are
unreliable due to the presence of the film of the continuous phase on the bottom and top
walls of the microchannel. The light from the mercury lamp passed through a band-pass
filter of 546 ± 12 nm, reflected from a dichroic mirror with 560 nm edge wavelength, and
illuminated the measurement area of the microchannel. The re-emitted light from the solute
passed through the dichroic mirror and a band-pass filter of 575–640 nm. Fluorescence
intensity was recorded by an IMPERX CCD camera with a resolution of 4 MPix in a 16-bit
format. The spatial resolution was 0.43 μm/pixel. The fluorescence images were processed
in ActualFlow software, and the calculated concentration fields were analyzed with the
help of a script written in Python.
The concentration field calculation during the experiment requires a calibration curve
to determine the dependence of the registered light intensity on the dye concentration
in the flow. Calibration was performed pixel by pixel, so the measurement area retained
the same position during the curve construction and the experiment. For calibration,
the microchannel was filled with the liquid with known fluorescent dye concentrations
including zero concentration. One hundred fluorescence intensity images were registered
for each concentration and averaged before proceeding to the calibration step. Additionally,
the dark noise of the camera was captured and subtracted from the calibration images and
the fluorescence images of the slugs.
To reveal the flow patterns and plug properties, high-speed flow visualization was
performed. The flow was illuminated by a halogen lamp. A microscope lens with a
magnification of M = 5× and a numerical aperture of NA = 0.12 was used. Flow images
were recorded by a high-speed PCO camera with 1 MPix resolution and a frame rate of up
to 1 kHz. The spatial resolution was 2.17 μm/pixel.
An aqueous solution was used as the dispersed phase, and tri-n-butyl phosphate
(TBP) was used as the continuous phase. The properties of liquids are presented in Table 1.
Water–glycerol solution properties were calculated by the parametrization in [24] with
adjustments described in [25]. The physical properties of TBP are provided by the supplier.
The interfacial tension of the liquids was evaluated using Antonov’s rule [26], the calculated
value coincides well with the measured value in [27]. The percentage for the water–glycerol

64
Energies 2023, 16, 4059

solution was chosen so that its refractive index nd = 1.428 was close to the refractive index
of the carrier phase nc = 1.425. Therethrough, the distortion of the rays of the re-emitted
light is avoided, significantly reducing the measurement error.

Table 1. Physical properties of the liquids.

Water–Glycerol Solution TBP


density, kg/m3 1181 973
dynamic viscosity, mPa·s 23 3.4
interfacial tension, mN/m 12.6

3. Results
3.1. Slug Flow Properties
To distinguish the boundary between parallel and dispersed flow patterns high-speed
visualization was performed in the range of superficial velocities of the continuous phase
(100 μm/s ≤ Uc ≤ 6700 μm/s) and dispersed phase (100 μm/s ≤ Ud ≤ 2200 μm/s).
Subsequently, a flow pattern map was created, as shown in Figure 2a. Initially, the flow
rate of the continuous phase was kept constant, after which the flow rate of the dispersed
phase was increased until a parallel flow pattern was established in the channel.
The dimensionless criterion Weα ·Ohb is suitable for unifying flow pattern maps. Here,
We is the Weber number, which expresses the ratio of inertia forces to interfacial tension
forces, and Oh is the Ohnesorge number, which characterizes the properties of liquids
in a two-phase flow [28]. In works by Kovalev et al. [29,30], the parameters a = 0.4 and
b = 0.6 were proposed to draw a universal flow pattern map. Additionally, the equation
for the boundary between the segmented and continuous flow patterns was suggested as
 0.4 0.6 0.4
d ·Ohd = 0.052· Wec Ohc
We0.4 0.6 , which describes the transition between flow patterns
for liquid–liquid sets with different properties. We compared the boundary between slug
and parallel flow in our experiment with the equation by Kovalev et al. in Figure 2b.
The boundary is located at the higher values of the We0.4 d ·Ohd which coincides with the
0.6

findings by Kovalev et al. that the proposed equation works well for pairs of liquids with a
viscosity ratio of less than unity. In our case, the viscosity ratio is 6.8.

(a) (b)

Figure 2. (a) Flow pattern map in terms of superficial velocities of the phases; (b) the boundaries
between parallel and dispersed flow patterns [29,30].

Based on the flow images obtained during high-speed visualization, we measured


the lengths and velocities of plugs, which are necessary for further developing the mass

65
Energies 2023, 16, 4059

transfer model. For each flow rate of the continuous and dispersed phase, the averaging
was performed among at least ten slugs. The slug lengths were approximated by models
from the works of Garstecki et al. [31] (7) and Xu et al. [32] (8). The first approximation
corresponds to the ‘squeezing’ mechanism of plug breakup, which depends crucially on
the blockage of the channel by a liquid plug. Xu’s model is designed for the so-called
‘dripping’ regime of plug formation dominated by the balance between shear force and
interfacial force.
Q
L plug = 1 + 1.33· d (7)
Qc
   
Qd 0.42 1 0.02
L plug = 2.34 ∗ (8)
Qc Ca
where Ca = (μc Ubulk )/σ—the capillary number based on the dynamic viscosity of the
continuous phase and bulk velocity. The parity plots of the measured plug lengths and
approximated data according to Equations (7) and (8) are presented in Figure 3a,b. The
R-squared was 0.76 and 0.86 for each of the models, respectively. As we obtained good
results for plug approximation by the Xu model, we can conclude that the capillary number
of the continuous phase influences the plug length, and the formation of plugs is in the
‘dripping’ regime.

(a) (b)

Figure 3. Comparison of experimental plug lengths with models (a) Equation (7); (b) Equation (8).

The plug velocity was approximated using linear and power functions of bulk ve-
locity. We obtained nearly equal R-squared for each approximation: 0.996 for the linear
approximation and 0.997 for a power function. The function plots are presented in Figure 4.
Further, we used a linear approximation of the plug length.

66
Energies 2023, 16, 4059

Figure 4. The dependence of plug velocity on bulk velocity.

3.2. Mass-Transfer Assessment


3.2.1. Partition Coefficient Measurement
Three samples were prepared to determine the partition coefficient K. These samples
were mixtures of equal volumes of TBP and water–glycerol solution, 30 mL each. The initial
concentrations of the fluorescent dye in the dispersed phase were 10 mg/L, 20 mg/L, and
30 mg/L, corresponding to the linear part of the dependence of the emitted light intensity
on the dye concentration according to the work of Zehentbauer et al. [33]. Each sample was
mixed on a magnetic stirrer for 72 h and allowed to settle until the interfacial layer became
completely uniform, indicating complete coalescence of the droplets. The liquids were then
manually separated.
The residual concentration of the solute in the aqueous phase was determined
using the micro-LIF technique in a microchannel. To construct the calibration curve, the
microchannel was filled with prepared samples of the aqueous phase with the known
fluorescent-dye concentrations: 0 mg/L, 10 mg/L, 20 mg/L, and 30 mg/L. Based on the
obtained intensity values corresponding to specific concentrations, a calibration curve
for the aqueous phase was constructed (Figure 5a). The residual concentrations in the
separated samples of the aqueous phase were determined using the calibration curve.
Since equal volumes of water–glycerol solution and TBP were used, it was possible to
calculate the volume concentration of the solute in the TBP samples:
∗ ∗
Corg = C0 aq − Caq (9)

The partition coefficient was calculated according to Equation (3). Its average value
was K = 3.9 ± 0.7.

67
Energies 2023, 16, 4059

(a) (b)

Figure 5. Calibration curves of the registered fluorescent intensity on fluorescent dye concentration:
(a) water–glycerol solution; (b) TBP at first measurement area.

3.2.2. Solute Concentration Measurements in Slugs of the Continuous Phase


To measure solute concentration in slugs of the continuous phase in two-phase flow,
the calibration curve of the dependence of Rhodamine 6G fluorescence intensity on its
concentrations was derived using TBP samples obtained after phase separation in the
measurements of the partition coefficient. The calibration curve is presented in Figure 5b.
The maximum deviation of the points from the approximating linear function was 4.6%,
and the random measurement error did not exceed 1.2%.
The measurement scheme of solute concentration in TBP in two-phase flow is pre-
sented in Figure 6. The continuous phase without a fluorescent dye and the dispersed phase
with a Rhodamine 6G concentration of 30 mg/L were fed to the microchannel inlets, which
formed slugs and plugs in the mixing zone. The flow rates of the continuous and dispersed
phases were chosen according to the plug regime in the flow pattern map obtained during
high-speed visualization and the condition that slugs of the continuous phase fit entirely
into the CCD camera frame. Thus, the flow rate ranges of the dispersed and continuous
phases were 0.5 μL/min ≤ Qd ≤ 6 μL/min and 0.5 μL/min ≤ Qc ≤ 3 μL/min, and the
flow rate ratio Qd /Qc varied in the range of 1–6.

Figure 6. The measurement scheme of the solute concentration in slugs of the continuous phase.

68
Energies 2023, 16, 4059

A series of fluorescence images were recorded at points located at distances of 5.6 mm,
8.5 mm, 13.6 mm, and 17 mm, which corresponded to 20, 32, 52, and 65 hydraulic diameters
downstream of the T-junction. At least 15 slugs of the continuous phase were recorded for
each studied regime and position. Then, concentration fields in the slugs were calculated ac-
cording to the calibration curve, and averaging was performed over 15 concentration fields
for the studied flow rates in each position. Afterward, with the help of a script written in
Python, the average values of the concentrations in the slugs and their standard deviations
were calculated. Since the measured concentration field in the slug is depth-averaged, we
numerically averaged concentration fields point by point, obtaining the average volumetric
concentration. An example of the evolution of concentration fields of solute in slugs of the
continuous phase for the fixed dispersed phase flow rate Qd = 3 μL/min and different flow
rates of the continuous phase downstream from the flow is shown in Figure 7.

Figure 7. The evolution of concentration fields with increasing Qc at a fixed Qd = 3 μL/min.

4. Discussion
As a result of the measurement of the concentration fields in the slugs of the continuous
phase, the dependences of the extraction efficiency and the overall volumetric mass transfer
coefficient on the residence time and the total flow rate of the phases were obtained. The
calculation was made using Equations (1) and (2) and the equilibrium solute concentration
in the continuous phase for the studied flow rates was determined by Equation (6). The
residence time τ was assessed as the distance from the T-junction to the measurement area
L divided by plug velocity U plug .
Figure 8 shows the dependence of the extraction efficiency and the mass transfer
coefficient at a fixed flow rate ratio Qd /Qc = 2. The error bars correspond to the standard
deviation of the values calculated based on 15 slugs. The extraction efficiency and the
overall volumetric mass transfer coefficient were found to increase with an increase in
the total flow rate. We assume that is due to the increasing intensity of vortices inside
the slugs and plugs, providing a higher concentration gradient at the interfacial area
between phases.

69
Energies 2023, 16, 4059

(a) (b)

Figure 8. (a) Extraction efficiency and (b) overall volumetric mass transfer coefficient as a function of
residence time at a fixed flow rate ratio.

Figure 9 shows the dependence of the extraction efficiency at 17 mm downstream


of the T-junction and the overall volumetric mass transfer coefficient at the distance of
5.6 mm from the T-junction on the total flow rate of the phases for various flow rate ratios.
Increasing the flow rate ratio Qd /Qc at a fixed total flow rate Qtot decreases the extraction
efficiency and mass transfer coefficient, possibly due to the fact that increasing Qd /Qc leads
to an increase in the equilibrium concentration but the average concentration in the slug
remains approximately the same. For the range of studied flow rates, the highest mass
transfer efficiency was achieved at a minimal studied flow rate ratio Qd /Qc close to unity.

(a) (b)

Figure 9. (a) Extraction efficiency at 17 mm from the T-junction as a function of the total phase flow
rate at different phase ratios; (b) overall volumetric mass transfer coefficient at 5.6 mm from the
T-junction as a function of the total phase flow at different phase ratios.

70
Energies 2023, 16, 4059

We correlated the obtained values of the total volume mass transfer coefficient with
the model from Kashid et al. [13], which was developed to assess the mass transfer intensity
between microchannel input and output. The authors used a fixed flow rate ratio and
considered the part containing the length of the plug to be a constant. In our case, the flow
rate ratio is changing, so we cannot neglect this part of the model:
 d  
Dh L plug e
k L a·τ = a· (Ca) b · ( Re)c · · (10)
L Dh
μ U ρ U D
where Ca = M σ plug —capillary number and Re = M μplug M
h
—Reynolds number, which
includes dynamic viscosity (μ M ) and density (ρ M ) of the mixture, Dh —hydraulic diameter
of the microchannel, L—length between the beginning of the channel and a point down-
stream of the T-junction, and L plug —plug length. Mixture properties were calculated by
the following equations:
 
εc ε −1
μM = + d (11)
μc μd
 
εc ε −1
ρM = + d , (12)
ρc ρd
where ε c and ε d —phase fraction of the continuous and dispersed phases.
Constants a − e are adjustable parameters determined by fitting experimental data to
the correlation (10) using the Levenberg–Marquardt algorithm. As a result, the following
parameter values were obtained:
 −0.59  
Dh L plug −0.04
k L a·τ = 0.042· (Ca)−0.92 · ( Re)0.64 · · (13)
L Dh

The experimental data versus the model prediction are plotted in Figure 10. The
R-squared value was 0.966. The median deviation between the experimental and predicted
data was 8%. The analysis of the obtained parameters shows the dependence on the
residence time and the total flow rate to be aligned with the experimentally obtained
results and previously conducted studies [20,34]. The obtained coefficients correspond to
the dependencies on the flow-rate ratio and total flow rate in Bai et al. [11]. At the same
time, compared to the values of the parameters in Kashid et al. [7], the discrepancy in the
dependence from the Reynolds number is observed.

Figure 10. Experimental data of overall volumetric mass-transfer coefficient versus the data predicted
by the model.

71
Energies 2023, 16, 4059

5. Conclusions
Experiments were conducted to investigate mass transfer during the flow of immiscible
liquids in a T-type microchannel using the micro-LIF method. The concentration fields
of the fluorescent dye in the slugs of the continuous phase were measured. To calculate
the extraction efficiency and the overall volumetric mass transfer coefficient, the partition
coefficient for the studied set of liquids was experimentally determined.
The results of the experiments led to conclusions regarding the effect of the total flow
rate and the phase flow ratio on the efficiency of mass transfer and overall volumetric mass
transfer coefficient. It was found that an increase in the flow rate ratio leads to a decrease in
mass transfer efficiency. Meanwhile, an increase in the total flow rate results in an increase
in both the extraction efficiency and mass transfer coefficient. A model for the overall
volumetric mass transfer coefficient was also developed with an R2 value of 0.966. This
model supports the conclusions about the impact of the total flow rate and flow-rate ratio
on the intensity of mass transfer between phases.

Author Contributions: Conceptualization, A.Y.; methodology, A.Y. and A.K.; investigation, S.V.;
resources, A.Y.; data curation, S.V.; writing—original draft preparation, S.V. and A.Y.; writing—review
and editing, A.B. and A.K.; supervision, A.B.; project administration, A.Y.; funding acquisition, A.Y.
and A.B. All authors have read and agreed to the published version of the manuscript.
Funding: The slug flow properties study (Section 3.1) was supported by a grant from the Russian
Science Foundation (project No 21-79-10307). Mass transfer measurements (Section 3.2) have been
financially supported by the Ministry of Science and Higher Education of the Russian Federation,
Project No 075-15-2022-1043.
Data Availability Statement: The data that support the findings of this study are available from the
corresponding author upon reasonable request.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Abdollahi, A.; Sharma, R.N.; Vatani, A. Fluid flow and heat transfer of liquid-liquid two phase flow in microchannels: A review.
Int. Commun. Heat Mass Transf. 2017, 84, 66–74. [CrossRef]
2. Kandlikar, S.G.; Garimella, S.; Li, D.; Colin, S.; King, M.R. Heat Transfer and Fluid Flow in Minichannels and Microchannels; Elsevier:
Amsterdam, The Netherlands, 2014; ISBN 9780080983462.
3. Chen, G.-B.; Chao, Y.-C.; Chen, C.-P. Enhancement of hydrogen reaction in a micro-channel by catalyst segmentation. Int. J.
Hydrogen Energy 2008, 33, 2586–2595. [CrossRef]
4. Wang, K.; Li, L.; Xie, P.; Luo, G. Liquid–liquid microflow reaction engineering. React. Chem. Eng. 2017, 2, 611–627. [CrossRef]
5. Priest, C.; Zhou, J.; Klink, S.; Sedev, R.; Ralston, J. Microfluidic Solvent Extraction of Metal Ions and Complexes from Leach
Solutions Containing Nanoparticles. Chem. Eng. Technol. 2012, 35, 1312–1319. [CrossRef]
6. Tran, T.M.; Lan, F.; Thompson, C.S.; Abate, A.R. From tubes to drops: Droplet-based microfluidics for ultrahigh-throughput
biology. J. Phys. D Appl. Phys. 2013, 46, 114004. [CrossRef]
7. Chen, R.; Su, G.H.; Zhang, K. Analysis on the high-quality development of nuclear energy under the goal of peaking carbon
emissions and achieving carbon neutrality. Carbon Neutrality 2022, 1, 1–12. [CrossRef]
8. Voumik, L.C.; Science, N. Impact of Renewable and Non-Renewable Energy on EKC in SAARC Countries: Augmented Mean
Group Approach. Energies 2023, 16, 2789. [CrossRef]
9. Swanson, J.L. PUREX Process Flowsheets. In Science and Technology of Tributyl Phosphate; Schulz, W.W., Burger, L.L., Navratil, J.D.,
Bender, K.P., Eds.; CRC Press: Boca Raton, FL, USA, 1984; p. 55.
10. Ma, S.; Sherwood, J.M.; Huck, W.T.S.; Balabani, S. On the flow topology inside droplets moving in rectangular microchannels. Lab
Chip 2014, 14, 3611–3620. [CrossRef]
11. Abiev, R.S.; Butler, C.; Cid, E.; Lalanne, B.; Billet, A.-M. Mass transfer characteristics and concentration field evolution for
gas-liquid Taylor flow in milli channels. Chem. Eng. Sci. 2019, 207, 1331–1340. [CrossRef]
12. Butler, C.; Cid, E.; Billet, A.-M. Modelling of mass transfer in Taylor flow: Investigation with the PLIF-I technique. Chem. Eng. Res.
Des. 2016, 115, 292–302. [CrossRef]
13. Kashid, M.N.; Gupta, A.; Renken, A.; Kiwi-Minsker, L. Numbering-up and mass transfer studies of liquid–liquid two-phase
microstructured reactors. Chem. Eng. J. 2010, 158, 233–240. [CrossRef]
14. Dietrich, N.; Loubière, K.; Jimenez, M.; Hébrard, G.; Gourdon, C. A new direct technique for visualizing and measuring gas–liquid
mass transfer around bubbles moving in a straight millimetric square channel. Chem. Eng. Sci. 2013, 100, 172–182. [CrossRef]

72
Energies 2023, 16, 4059

15. van Baten, J.M.; Krishna, R. Corrigendum to “CFD simulations of mass transfer from Taylor bubbles rising in circular capillaries”.
Chem. Eng. Sci. 2004, 59, 2535–2545. [CrossRef]
16. Yue, J.; Luo, L.; Gonthier, Y.; Chen, G.; Yuan, Q. An experimental study of air–water Taylor flow and mass transfer inside square
microchannels. Chem. Eng. Sci. 2009, 64, 3697–3708. [CrossRef]
17. Kashid, M.; Renken, A.; Kiwi-Minsker, L. Influence of Flow Regime on Mass Transfer in Different Types of Microchannels. Ind.
Eng. Chem. Res. 2011, 50, 6906–6914. [CrossRef]
18. Zhang, Q.; Liu, H.; Zhao, S.; Yao, C.; Chen, G. Hydrodynamics and mass transfer characteristics of liquid–liquid slug flow in
microchannels: The effects of temperature, fluid properties and channel size. Chem. Eng. J. 2019, 358, 794–805. [CrossRef]
19. Yao, C.; Ma, H.; Zhao, Q.; Liu, Y.; Zhao, Y.; Chen, G. Mass transfer in liquid-liquid Taylor flow in a microchannel: Local
concentration distribution, mass transfer regime and the effect of fluid viscosity. Chem. Eng. Sci. 2020, 223, 115734. [CrossRef]
20. Angeli, P.; Tsaoulidis, D.; Weheliye, W.H. Studies on mass transfer of europium(III) in micro-channels using a micro Laser Induced
Fluorescence technique. Chem. Eng. J. 2019, 372, 1154–1163. [CrossRef]
21. Ganguli, A.A.; Pandit, A.B. Hydrodynamics of Liquid-Liquid Flows in Micro Channels and Its Influence on Transport Properties:
A Review. Energies 2021, 14, 6066. [CrossRef]
22. Bai, L.; Zhao, S.; Fu, Y.; Cheng, Y. Experimental study of mass transfer in water/ionic liquid microdroplet systems using micro-LIF
technique. Chem. Eng. J. 2016, 298, 281–290. [CrossRef]
23. Kuhn, S.; Jensen, K.F. A pH-Sensitive Laser-Induced Fluorescence Technique to Monitor Mass Transfer in Multiphase Flows in
Microfluidic Devices. Ind. Eng. Chem. Res. 2012, 51, 8999–9006. [CrossRef]
24. Cheng, N.-S. Formula for the Viscosity of a Glycerol−Water Mixture. Ind. Eng. Chem. Res. 2008, 47, 3285–3288. [CrossRef]
25. Volk, A.; Kähler, C.J. Density model for aqueous glycerol solutions. Exp. Fluids 2018, 59, 75. [CrossRef]
26. Demond, A.H.; Lindner, A.S. Estimation of interfacial tension between organic liquids and water. Environ. Sci. Technol. 1993, 27,
2318–2331. [CrossRef]
27. Rader, C.A.; Schwartz, A.M. The Migration of Liquids in Textile Assemblies: Two-Component and Two-Phase Liquid Systems.
Text. Res. J. 1962, 32, 140–153. [CrossRef]
28. Yagodnitsyna, A.A.; Kovalev, A.V.; Bilsky, A.V. Flow patterns of immiscible liquid-liquid flow in a rectangular microchannel with
T-junction. Chem. Eng. J. 2016, 303, 547–554. [CrossRef]
29. Kovalev, A.V.; Yagodnitsyna, A.A.; Bilsky, A.V. Viscosity Ratio Influence on Liquid-Liquid Flow in a T-shaped Microchannel.
Chem. Eng. Technol. 2020, 44, 365–370. [CrossRef]
30. Kovalev, A. The Influence of Viscosity on the Hydrodynamics of Immiscible Liquid-Liquid Flows in Rectangular Microchannels.
Ph.D. Thesis, Kutateladze Institute of Thermophysics SB RAS, Novosibirsk, Russia, 2022.
31. Garstecki, P.; Fuerstman, M.J.; Stone, H.A.; Whitesides, G.M. Formation of droplets and bubbles in a microfluidic T-junction—
Scaling and mechanism of break-up. Lab Chip 2006, 6, 437–446. [CrossRef]
32. Xu, J.H.; Li, S.W.; Tan, J.; Luo, G.S. Correlations of droplet formation in T-junction microfluidic devices: From squeezing to
dripping. Microfluid. Nanofluidics 2008, 5, 711–717. [CrossRef]
33. Zehentbauer, F.M.; Moretto, C.; Stephen, R.; Thevar, T.; Gilchrist, J.R.; Pokrajac, D.; Richard, K.L.; Kiefer, J. Fluorescence
spectroscopy of Rhodamine 6G: Concentration and solvent effects. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 121,
147–151. [CrossRef]
34. Qian, J.-Y.; Li, X.-J.; Wu, Z.; Jin, Z.-J.; Sunden, B. A comprehensive review on liquid–liquid two-phase flow in microchannel: Flow
pattern and mass transfer. Microfluid. Nanofluidics 2019, 23, 1–30. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

73
energies
Article
Combining Microstructured Surface and Mesh Covering for
Heat Transfer Enhancement in Falling Films of
Refrigerant Mixture
Oleg Volodin *, Nikolay Pecherkin and Aleksandr Pavlenko *

Kutateladze Institute of Thermophysics, Siberian Branch of the Russian Academy of Sciences, 1 Lavrentyev Ave.,
630090 Novosibirsk, Russia
* Correspondence: [email protected] (O.V.); [email protected] (A.P.)

Abstract: The article presents the experimental results of combining a basic microstructure with
partly closed pores and a mesh covering for heat transfer enhancement at the film flow of a refrigerant
mixture. To reveal the effect of the combined structure, heat transfer on a microstructured surface
without a covering as well as on a smooth surface with a mesh covering only has been studied. All
experimental series were carried out using a binary mixture of R114 and R21 refrigerants. The mixture
film flowed down the outer surface of a vertical cylinder in the undeveloped turbulence regime,
when the film Reynolds number varied from 400 to 1300. It is shown that a microstructured surface
with a fin pitch of 200 μm, fin height of 220 μm, and longitudinal knurling pitch of 160 μm, created
by deformational cutting, demonstrates significant heat transfer enhancement: up to four times as
compared to a smooth surface. However, adding a mesh covering with an aperture of 220 μm and
a wire diameter of 100 μm reduces the intensification. The mesh covering overlaid on a smooth
surface also does not provide heat transfer enhancement as compared to the smooth surface itself.
The absence or even deterioration of heat transfer enhancement on surfaces with mesh covering can
be primarily associated with the low thermal conductivity of the mesh material and shortcomings
of the applied method of mesh mounting. The possibility of deteriorating vapor removal due to
the incorrect selection of mesh covering parameters was also analyzed. The heat transfer coefficient
Citation: Volodin, O.; Pecherkin, N.;
values obtained for basic microstructured surfaces were compared with the dependencies available
Pavlenko, A. Combining
in the literature for predicting pool boiling heat transfer on microfinned surfaces.
Microstructured Surface and Mesh
Covering for Heat Transfer
Keywords: nucleate boiling; combined coating; deformational cutting; mesh covering; falling films;
Enhancement in Falling Films of
Refrigerant Mixture. Energies 2023,
refrigerant mixture
16, 782. https://doi.org/
10.3390/en16020782

Academic Editors: Fedor Ronshin


1. Introduction
and Vladimir Serdyukov
The study of heat transfer in falling films of liquids and their mixtures is important for
Received: 29 November 2022 improving the efficiency of numerous systems using film flows—from natural gas liquefac-
Revised: 5 January 2023 tion (LNG) plants, distillation plants, and absorption apparatuses, as well as evaporative
Accepted: 6 January 2023 equipment of the chemical and food industries—to desalination plants and electronic equip-
Published: 10 January 2023 ment cooling systems (e.g., spray and falling-film cooling system [1]). The intensification of
heat transfer at boiling and evaporation of liquids by structuring the heat-generating sur-
face is a key method for increasing the efficiency of both traditional [2–5] and renewable [6]
energy systems. Despite the fact that studies on heat transfer enhancement have been
Copyright: © 2023 by the authors.
actively carried out since the middle of the last century, their relevance at the current pace
Licensee MDPI, Basel, Switzerland.
This article is an open access article
of technology development is only increasing. Among various types of structured coatings
distributed under the terms and
(including capillary-porous), wire mesh coverings—which can be attributed to the simplest
conditions of the Creative Commons porous coatings—are among the most accessible and easily modified. The advantages
Attribution (CC BY) license (https:// of these coverings also include: ease of installation, high scalability, low production cost,
creativecommons.org/licenses/by/ and reproducibility of the geometric parameters of wire meshes. All this determines the
4.0/). renewed interest in recent years in the use of mesh coverings in various configurations [5].

Energies 2023, 16, 782. https://doi.org/10.3390/en16020782 74 https://www.mdpi.com/journal/energies


Energies 2023, 16, 782

For a rather long period of research on the use of mesh coverings of heating surfaces
for heat transfer enhancement, i.e., from about 1975 and until the end of the first decade
of the 21st century, not so many promising results were obtained on increasing the heat
transfer coefficient (HTC) or delaying of reaching the critical heat flux (CHF) [5]. It was
generally believed that mesh coverings enhance heat transfer at low heat fluxes and impair
it at high heat fluxes. The enhancement was associated with the increased area of the mesh-
covered surface and consequently the higher number of active nucleation sites, whereas at
higher heat fluxes heat transfer becomes dependent on the vapor removal, and in this case,
the mesh coatings (especially multilayer ones) impair heat transfer [5,7].
However, the results of studies conducted over the past decade [5] show that mesh
coverings can be an effective means of heat transfer enhancement in a wide range of heat
flux changes. Some authors have also expanded previous ideas [7] about the causes of
the HTC improvement by introducing fresh concepts like “micro-chimney effects” and
“gradient mesh coatings” [6]. The desired enhancement can be realized provided that the
mesh geometrical parameters, the number of mesh layers, the wire material, and the mesh
mounting method are optimally chosen. Thus, significant heat transfer enhancement was
obtained using the mentioned above gradient mesh coatings [6] (HTC enhancement up
to 6.6 times), as well as uniform layers mesh coatings at atmospheric (HTC enhancement
from about 3 [8] to 10 times [9]) and subatmospheric pressures (HTC enhancement up to
22 times [9]). Promising results have also been achieved by combining mesh covering with
other types of surface treatment: microfinning, nanostructuring, mesh with cells filled with
powder from micro-nanoparticles, as well as when processing wire mesh itself.
Below we briefly review the works devoted to the study of heat transfer enhancement
and the increase of the critical heat flux using multilayer or combined mesh coatings, which
contain the most interesting or promising results in the opinion of the authors.
In the works [10,11], it was demonstrated that covering a rough surface with a mesh
can reduce the intensity of heat transfer. Trying to use simple and practical methods of
enhancement, Tsay et al. [10] were among the first to combine the roughening of heat
transfer surface with a mesh covering (from stainless steel AISI 304) to enhance boiling in
thin layers of water on a horizontal plate (also from AISI 304) of 10 cm length, 2.5 cm width,
and 0.1 cm thickness. The authors also investigated the effect of applying various mesh
coverings (mesh 16, 24, and 50) to a smooth plate. As a result, it received up to a sevenfold
HTC increase (at liquid level H = 5 mm and mesh 16) at low heat fluxes. The thinner the
liquid layer, the greater the observed enhancement. However, no additional enhancement
for the mesh-covered rough surfaces as compared with the mesh-covered smooth ones
was received.
The use of combined coatings by Brautsch and Kew [11] was also unsuccessful in terms
of heat transfer enhancement. The authors used low carbon stainless steel AISI 304L meshes
(mesh 50, 100, 150, and 200), as well as two heat-releasing surfaces with Ra = 0.42 μm and
Ra = 1.04 μm, to study heat transfer enhancement in the saturated pool boiling of water on
the vertical test section. They showed that both techniques—roughening of surface and
covering the heater with a mesh, are effective. However, combining them was shown to be
ineffective, resulting in the HTC deterioration even as compared to the smooth surface.
Despite these and other unsuccessful experiments on heat transfer enhancement [5],
changes in the material and method of mounting mesh coverings allowed Sloan et al. [9]
to achieve significant intensification of HTC over the entire range of heat fluxes. The
authors [9] studied the pool boiling of water at subatmospheric pressure on a 4 cm2
vertically oriented copper circular disk, covered with 1–8 layers of copper mesh. It was
shown that eight-layer mesh 145 covering demonstrated about a 10-fold HTC increase (at
ΔT = 10 K) at atmospheric pressure and a 22-fold increase (at ΔT = 8 K) at subatmospheric
pressure (0.2 atm). It is worth noting that Sloan et al. used copper mesh coverings previously
cold-rolled to increase the available surface area for the diffusion bonding process by
producing flat spots on the high points of each wire.

75
Energies 2023, 16, 782

Chien and Tsai [12] studied heat transfer at film flow and pool boiling of R-245fa at
different saturation temperatures on horizontal finned copper tubes (of 0.4 mm fin height,
60 FPI) and the same tubes, covered with copper mesh. They achieved a notable HTC
enhancement—up to four and seven times at pool boiling of R-245fa for 5 and 20 ◦ C,
respectively, and up to five times at film flow of R-245fa, compared to a smooth tube. For
comparison, the uncovered finned tube in the latter case provided intensification only
3.5 times.
The authors [13] used a combined surface with brass or copper meshes (mesh 80,
100, and 120) covered finned horizontal tubes with fins of 0.2–0.4 mm high. Pool boiling
of R-134a at saturation temperatures of 5, 10, and 26.7 ◦ C was enhanced 2–3 times. The
best performance was achieved by wrapping a brass mesh 100 on a tube with 0.4 mm fin
height—up to 8 times (peak value) compared to an uncovered tube.
Kim et al. [14] achieved 84% CHF enhancement by using single-layered stainless steel
mesh (with wire diameter 0.29 ÷ 0.7 mm and mesh aperture 1.29 ÷ 2.67 mm range) with
micro/nano-sized pore structure of meshes applied to SiO2 heating surface at water pool
boiling. The authors underlined that their method of increasing CHF does not require
any modifications of the heating surface and can be easily adopted in different technical
applications (for example, to create IC chip coatings).
Dabek
˛ et al. [15] studied the pool boiling of water and ethyl alcohol on a horizontal
copper heater of 30 mm diameter with copper or bronze one- or two-layer mesh coatings at
atmospheric pressure. The possibility of sevenfold and fourfold enhancement as compared
to a smooth surface (at superheat of about 8 K) for copper meshes with apertures of 0.75 mm
and 0.2 mm, respectively, has been demonstrated.
The work of [16] is notable for the fact that the authors used 3D-printed mesh structures
for heat transfer enhancement at the pool boiling of water at the saturation line. A total
of 12 samples were divided into two groups of printed meshes: “thin” (0.75 mm wall
height) and “thick” (1.5 mm wall height), with pitch varied in a range of 0.4 ÷ 1.3 mm.
Stainless steel 316 L powder was used in the process of selective laser melting (SLM).
Zhang et al. [16] showed the possibility of a threefold CHF enhancement as compared to a
smooth surface by applying a mesh with 1.1 mm pitch (“thick”), also HTC enhancement of
two to three times by applying meshes with 0.5 mm pitch (“thin”) and 0.7–1.1 mm pitch
(“thick”) as compared to the smooth surface was achieved. This work [16] demonstrates
the perspectives of additive manufacturing (AM) for creating prototypes of samples with
precisely controlled structure parameters.
Pastuszko et al. [17] used micro-finned surfaces with copper mesh covering as well as
micro-finned surfaces covered by copper perforated foil at pool boiling of water, ethanol,
Novec-649, and FC-72 at atmospheric pressure. Microfins covered with wire mesh produced
the highest HTC among studied surfaces at medium and high heat fluxes for water, low
and medium heat fluxes for ethanol, and medium heat fluxes for FC-72.
The authors of [6] have demonstrated that multilayer mesh coatings can be highly
efficient intensifiers by studying the pool boiling of distilled water at the saturation line on
a heated surface covered with multi-layer mesh. Four configurations of six-layer copper
mesh coverings with gradient (direct or inverse) or uniform porosity were studied. It was
shown that a six-layer mesh (3 + 3) with coarser three upper layers gives a maximum HTC
enhancement—up to 6.6 times (261 kW/m2 K), along with a three-fold enhancement in
CHF (outstanding 2719 kW/m2 ). The authors associate the obtained enhancement results
with so-called “micro-chimney effects” [6], taking place in gradient porous micro meshes.
Huang et al. [18] applied four hybrid surfaces for heat transfer enhancement at sub-
cooled water flow boiling in channels (at a pressure of 0.5 MPa, a flow velocity of 1–5 m/s,
and an inlet temperature of 298 K). The authors used wire mesh coatings combined with a
powder mixture of micro/nanoparticles of Ag, Cu, and Ti. Heat flux removed by combined
surfaces was two to three times higher than heat flux removed by the smooth surface, CHF
for the investigated surfaces increased by 80–200%.

76
Energies 2023, 16, 782

The authors [19], as well as [6] demonstrate the effectiveness of multilayer gradient
meshes, but in terms of enhancement of wicking capability. It was shown that the wicking
capability of a multilayer gradient mesh consisting of three lower layers of a mesh 100 and
three upper layers of a mesh 300 is significantly enhanced compared to wicks consisting of
a multilayer mesh with identical layer characteristics.
In the previously mentioned work [8] the authors proposed a surface, sintered with
multilayer copper meshes having identical geometrical characteristics (mesh 200 with
30 μm wire diameter), studying the water pool boiling. It is shown that an increase in the
number of layers (up to 5) can reduce the size of the micropores, increasing the density
of nucleation sites and improving capillary wicking performance, thus improving the
HTC and delaying the boiling crisis development. The multilayer mesh with 5 layers
demonstrates optimal boiling performance, providing the highest CHF of 208 W/cm2 and
the highest HTC of 16 W/(cm2 K). The authors made the conclusion that the remarkable
boiling performance along with the low cost, simplicity, and high durability of mesh
coatings show the industrial prospects for commercial compact microelectronics cooling.
Hu et al. [20] along with the authors of [8,9] demonstrate the perspectives of using
uniform multilayer copper micromeshes for heat transfer enhancement, in particular, to
increase heat transfer proportion of liquid film boiling in spray cooling of 10 × 10 mm2
target surface. It is shown that a four-layer mesh 100 covering (mesh covering with
50 μm wire diameter and 204 μm aperture) fabricated by diffusion bonding, exhibits
the best heat transfer performance with CHF of 605 W/cm2 and maximum HTC of
71 kW/(m2 K), which corresponds to enhancing by 127% and 176%, compared with the
uncovered surface, respectively.
Thus, the above shows that wire mesh coatings in various combinations and modifica-
tions can be an effective means of boiling heat transfer for a wide spectrum of technical appli-
cations, including apparatuses using pool boiling [6–9,11–17], working at low pressures [9],
using spray cooling [20], thin layers of liquid [10], film flow [7,12], or microchannels [18].
The aim of this work is to initiate the investigation of the efficiency of combined coatings
for heat transfer enhancement in the binary refrigerant mixture films falling down the outer
surface of a vertical cylinder. Despite the fact that the cooling of heating surfaces by falling
films does not allow the removal of large heat fluxes as in the case of the recognized leader
among cooling methods—spray irrigation, which makes it possible to remove heat fluxes
up to 1000 W/cm2 [21], tubular heat exchangers operating at low and moderate heat fluxes
are widely used and in demand in the industry (for example, in LNG systems), and the
possibility of enhancing heat transfer and improving the ergonomics of film evaporators by
using tubes with modified surfaces requires systematic research in this field.

2. Experimental Setup and Procedure


2.1. Test Sections
To create the basic microstructure with halfway-closed micropores, providing more
effective nucleate boiling, the previously well-established method of deformational cutting
(MDC) was used [22–24]. The microstructure, created on the outer surface of a copper cylinder
with a diameter of 50 mm and wall thickness of 1.5 mm, has the next geometrical parameters:
fin pitch—200 μm, fin height—220 μm, and longitudinal knurling pitch—160 μm (Figure 1a,b).
The knurling over microfinning, crushing the tops of the fins with a decrease in its height by
35%, is carried out in order to create the halfway-closed pores surface. In the work of [22], it
was shown the effectiveness of such structure type for nucleate boiling enhancement. In this
study, we slightly modified the parameters of the most effective MDC-surface (microstructure
No. 1 from [22]), increasing its fin pitch (from 100 μm to 200 μm)—with the aim of improving
vapor removal, and reducing its knurling pitch (from 318 μm to about 160 μm)—in attempt to
increase the amount of nucleation centers.

77
Energies 2023, 16, 782

(a) (b)
Figure 1. (a) photograph of the surface, structured by deformational cutting, (b) sketch of the element
of the MDC-surface.

At the next stage of test section preparation, the stainless steel mesh (AISI 304) with
an aperture of 220 μm and wire diameter of 100 μm was tightly wrapped around the
microstructured tube, fixed, and then soldered with a slight overlap of layer upon layer.
Resulting in a combined structure (Figure 2a), the single element of which is shown (accord-
ing to scale) in Figure 2b. The characteristics of mesh were chosen so as to approximately
correspond to the parameters of the basic structured surface since a finer upper mesh
would impair vapor removal, while a too-coarse upper mesh would not have an effect on
the nucleate boiling process.

(a) (b)
Figure 2. (a) photograph of the combined surface, (b) sketch of the element of the combined surface.

In addition to applied steel mesh coverings (which did well in [10] and some other
works, despite the low thermal conductivity of steel), in the next stage of the study, it is
planned to use copper and brass meshes; the geometric parameters of meshes and applied
installation methods will also vary.
To reveal the effect of the combined coating structure on heat transfer intensification,
the heat transfer on the microstructured surface without mesh covering as well as on a
smooth surface with mesh covering alone and on an uncovered smooth surface were also
studied. All types of test sections used in the study are listed in Table 1.

78
Energies 2023, 16, 782

Table 1. Types of test sections used in the study.

Test Section Type Material Geometrical Parameters


smooth tube with roughness
Smooth tube Copper M1
Ra = 2.5 μm
microstructure with fin pitch
Microstructured tube Copper M1 200 μm, fin height 220 μm and
knurling pitch 160 μm
smooth tube with Ra = 2.5 μm
Smooth tube + mesh Copper M1 + AISI 304 + mesh with aperture 220 μm
and wire diameter 100 μm
microstructure with fin pitch
200 μm, fin height 220 μm and
Microstructured tube + mesh Copper M1 + AISI 304 knurling pitch 160 μm + mesh
with aperture 220 μm and
wire diameter 100 μm

2.2. Experimental Setup and Parameters


The experiments were carried out in the test column (1) 1.5 m high and with an inner
diameter of 0.27 m under saturation conditions at an absolute pressure of about 2 bar
(Figure 3a,b). A more complete arrangement of the experimental setup is given in [25].
The R114/R21 refrigerant mixture is fed into a constant-level tank (2) from where it comes
through the slot distributor to the test section (3). The local temperatures of the heating
wall are controlled by five copper-constantan thermocouples (4) with a wire diameter of
0.15 mm, embedded flush with the wall surface and located vertically along the heated
area having a length of 70 mm with a step of 13 mm. Thermocouples are handcrafted and
calibrated over a range of operating temperatures (20–55 ◦ C) using the proven reference
thermometer (LT-300). The diameter of the thermocouple bead does not exceed 0.7 mm.
The sensitivity of copper-constantan (CuKn) thermocouples or type T thermocouples is
about 40 microvolts per degree. The use of a highly sensitive voltmeter APPA 207 made
it possible to measure thermo-EMF values with an accuracy of ±8 μV. The cold junctions
of thermocouples (5) are located at the bottom of the column, immersed in the liquid
layer. The temperatures in the different sections of the column are controlled by platinum
thermoresistors HEL-711-U-0-12-00 (6) by Honeywell. The HEL-711 temperature sensors
allow the measurement of temperature in the range of −200 to +260 ◦ C with an accuracy
of ±0.1%. One of the thermistors located at the bottom of the column together with cold
junctions of thermocouples in order to measure the temperature of the liquid to correctly
build the dependence of the wall temperature on the thermo-EMF, taking into account the
temperature fluctuations of the liquid at the column bottom.
The observations and high-speed shooting (by Phantom VEO 410 L) of processes,
developing in the heated area, are carried out through three quartz optical windows (7).
The pressure in the test volume is measured using a manometer Metran-100 (8).
Stabilized supply source Mastech HY10010E was used to heat the test section, capable
of providing heat fluxes up to critical. The load was raised stepwise, the readings were taken
while moving up the heat flux branch after the heater temperature reached a stationary
value (which took about 10–15 min, depending on the heat transfer regime). The binary
mixture film flowed down the vertical cylinder in the undeveloped turbulence regime with
the film Reynolds number at the inlet of the test section varied from 400 to 1300. The film
Reynolds number was determined as:

4G
Re = , (1)
πdμ

here G is mass rate, d—tube diameter, and μ—dynamic viscosity. The mass flow rate was
measured by the Coriolis flow meter CORI-FLOW M55.

79
Energies 2023, 16, 782

(a) (b)
Figure 3. The scheme of the test column (a): 1—sealed column; 2—constant level tank; 3—test section;
4—heated zone of the test section with thermocouples; 5—thermal resistors; 6—thermocouples cold
junctions; 7—optical windows; 8—a manometer. The photograph of the test column without heat
shield (b).

The distance from the slot distributor to the beginning of the heated area was equal to
100 mm, which noticeably exceeded the distance of hydrodynamic stabilization of the flow
in all studied ranges of the flow rates; a corresponding estimation was carried out in [22].
The selected R114/R21 refrigerant mixture has low surface tension, high wettability,
and low viscosity, which makes it convenient for modeling heat transfer processes for
a wide class of low-viscosity technical fluids. In addition, the presence of the volatile
component of the mixture (R114) can notably enhance the heat transfer process. In the
work of [26], it was shown that for a 10% mixture of R114/R21, the heat flux density
corresponding to the onset of boiling (ONB) was reduced by half as compared to pure
R21, and an increase in HTC up to 2.5 times was observed in the range of heat fluxes q:
0.1 ÷ 10 W/cm2 . It can also be noted that the use of a non-azeotropic mixture can increase
the critical heat flux value, due to the later evaporation of the low-boiling component [27,28].
With this in mind, a mixture of R114 and R21 refrigerants with an initial concentration
of volatile component R114 of about 12% (the mixture behaves as non-azeotropic at this
concentration) was chosen as the working fluid.

2.3. Experimental Procedure and Data Reduction


Before the start of each experimental series (and after the tightness test), vacuum
evacuation of the test volume was performed to degas the volume and the micropores of
the test section. To evacuate the test setup (remove air) before filling it with the working
fluid, an RV-40 forevacuum pump was used. Air was pumped out to a residual pressure of
6–8 Pa. After air was evacuated from the test setup, the setup volume was filled with vapors
of the working fluid from the storage vessel. At room temperature of about (20–22) ◦ C,
the excess pressure in the setup is maintained at the level of (0.5–0.6) bar, which excludes
air suction into the setup. Measurement of vapor pressure and vapor and liquid phase

80
Energies 2023, 16, 782

temperatures in the test volume made it possible to control the state of saturation of the
working fluid and the absence of impurities (like air) during the experiments.
To measure the local wall temperatures, the thermo-EMF values from five thermocou-
ples were recalculated into temperature values using a two-dimensional regression—the
function of thermo-EMF and the temperature of cold junctions.
The heat flux density was calculated by the formula:

I2 R
q=C , (2)
A
here, C is a correction factor taking into account heat losses due to the heat conductivity
of the test section. According to the numerical calculations performed for the studied test
sections, the heat loss due to longitudinal heat conductivity of cylinder walls did not exceed
5%. The areas of the modified heat-releasing surfaces A during calculations were taken
equal to the area of the covered smooth surface.
The local heat transfer coefficients were determined by the standard formula:
q
h= , (3)
( Tw − Ts )

here TW —temperature of the wall, TS —saturation temperature.


The uncertainty in determining the local heat transfer coefficient h consisted of uncer-
tainties in determining the heat flux q and the wall superheat temperature ΔT = ( Tw − Ts ).
The final uncertainty of the heat transfer coefficient h = f (q, ΔT) was calculated by
the formula:
N  
∂f 2 2
U 2 (h) = ∑ U ( xi ) (4)
i =1
∂xi
or  2  2  2
U (h) U (q) U (ΔT )
= + , (5)
h q ΔT
where U 2 (ΔT ) = U 2 ( Tw ) + U 2 ( Ts ).
According to the assessment, the maximum error when determining ΔT was intro-
duced by the relative uncertainty of Tw , measured by CuKn thermocouples, that did not
exceed 10% at ΔT > 2 ◦ C. So the uncertainty of ΔT = ( Tw − Ts ) will also not exceed 10% up
to the integer. The relative uncertainty of the heat flux q taking into account heat losses did
not exceed 6%. Thus the relative uncertainty of the local heat transfer coefficient h did not
exceed 12% at ΔT > 2 ◦ C.

3. Results and Analysis


Figure 4a–d illustrates the boiling process of the R114/R21 mixture falling film on stud-
ied test sections at the initial stages of nucleate boiling. The test sections structured by defor-
mational cutting, both uncovered (Figure 4b) and having mesh covering (Figure 4d), demon-
strate a more effective nucleation process than sections without basic MDC-structuring
(Figure 4a,c), with nucleation sites activated from the beginning of heating area—so it
becomes visible—to its end.
Figure 5a–d demonstrate the nucleate boiling of binary mixture falling film on studied
sections at developed boiling regime—with the presence of very large bubbles (see for
example Figure 5c) characteristic of the mixtures and according to our observations does
not occur in pure refrigerants. Again it can be noted that both test sections with substrate
structured by deformational cutting (Figure 5b,d) demonstrate a more effective nucleation
process with nucleation sites evenly distributed over the entire heat-releasing surface. Next
will be shown that the heat transfer data confirm the visual observation.

81
Energies 2023, 16, 782

(a) (b)

(c) (d)

Figure 4. Inception of nucleate boiling of R114/R21 mixture on: (a) smooth (q = 1.7 W/cm2 , Re = 640),
(b) microstructured by MDC (q = 1.7 W/cm2 , Re = 640), (c) smooth covered by mesh (q = 1.7 W/cm2 ,
Re = 845) and (d) combined (q = 1.7 W/cm2 , Re = 640) surfaces.

Boiling curves obtained for the investigated surfaces are shown in Figure 6a–d.
Figure 6a shows the boiling curve for the smooth surface. In the film evaporation
regime (q ≤ 1 W/cm2 ), the flow rate does not affect heat transfer. For all Reynolds numbers,
ONB takes place at q ≥ 1 W/cm2 and incipience superheat ΔT ≈ 8 K, respectively. Above
ΔT ≈ 8 K and up to the developed boiling regime (ΔT ≈ 10.5 K), an insignificant effect of
the flow rate on heat transfer is observed—the thinner the film, the higher heat transfer, as a
rule. This could be due to the different contributions of evaporation to heat transfer, which
takes place for different thicknesses of the falling film in the developing boiling regime.
However, this effect is weakly expressed, and the presence of any kind of microtexture
(Figure 6b–d) immediately shuffles the cards.
Microstructured surface by the method of DC (Figure 6b) demonstrates the greatest
heat transfer efficiency, while the presence of the mesh covering does not affect (Figure 6c),
or even worsens (Figure 6d), the heat transfer process. In the first case, we suppose,
this is due to the fact that stainless steel (AISI 304) mesh used in the experimental series
has a low thermal conductivity (15 W/mK), and the method of mesh attachment used is
apparently not effective enough to create an ideal mesh-heated surface contact. In the case
of combined surface, this also can be connected with the fact that we did not choose the

82
Energies 2023, 16, 782

mesh covering parameters well enough for our experimental conditions, which may cause
the deterioration of vapor removal.
Thus, despite the fact that the presence of an applied steel mesh covering to some
extent orders nucleate boiling (one can compare boiling patterns on mesh-covered surfaces
in Figure 4c,d and on the smooth one in Figure 4a), taking into account the deterioration
of heat transfer, at the next stage of the experiment, measures should take into account
the existing shortcomings of the mesh coating used. It is planned to use highly thermally
conductive copper mesh coatings and the process of sintering to ensure contact of the mesh
covering with the heated surface; a more careful selection of mesh covering parameters
also should be provided.
A comparison of experimental data on HTC and its enhancement on the studied
surfaces for the range of Reynolds number 400 ÷ 1300 is presented in Figure 7a,b. As
noted above, the presence of mesh with the given characteristics does not enhance the
boiling process when covering the smooth surface: data on HTC coincide with data for a
smooth surface, Figure 7a. The microstructured surface, created by deformational cutting
(Figure 7b), demonstrates the greatest heat transfer enhancement (up to four times as
compared with the smooth one). In the case of mesh overlay of the microstructured surface
(i.e. in the case of combined coating), even impairment (up to two times) in heat transfer as
compared to high-performance MDC-surface is observed, Figure 7b.

(a) (b)

(c) (d)

Figure 5. Developed nucleate boiling of R114/R21 mixture on: (a) smooth (q = 5.6 W/cm2 , Re = 845),
(b) microstructured by MDC (q = 6.2 W/cm2 , Re = 1260), (c) covered by mesh (q = 5 W/cm2 , Re = 845)
and (d) combined (q = 5 W/cm2 , Re = 1260) surfaces.

83
Energies 2023, 16, 782

(a) (b)

(c) (d)

Figure 6. Boiling curves for: smooth (a), microstructured by MDC (b), smooth covered by mesh (c),
and combined (d) surfaces.

(a) (b)

Figure 7. HTC vs. q dependencies for: smooth tube and smooth tube, covered by mesh (a); mi-
crostructured tube and microstructured tube, covered by mesh (b).

A brief discussion on the problem of the correct selection of the upper mesh covering
parameters is given at the end of the section.

84
Energies 2023, 16, 782

There are not many correlations for predicting pool boiling heat transfer coefficients for
microstructured surfaces. We were able to find the calculation dependencies for predicting
the pool boiling HTC for microfinned surfaces in [29,30]. Next, we tried to compare the
obtained data for the MDC surface with these correlations, assuming that the process of
developed nucleate boiling should be very similar for falling films and pool boiling. The
main difference is brought by the evaporation of the film, however, in a developed boiling
regime, its role decreases to a negligible value. We also neglected the presence of additional
longitudinal knurling, present on the MDC-surface over the microfinning, since there are
no calculation dependencies for such geometry in the literature (the correlation proposed
in [29] for the prediction of boiling on micropin surfaces also gives greatly overestimated
values in case of our MDC-surface).
The calculation dependence of Aksyanov et al. [29] has the form:
 0.554  0.19  0.201
θ h Δ
α/α0 = 6Kq−0.2 (δ/l0 )−0.394 , (6)
90 l0 l0

where α, α0 are the heat transfer coefficients of microstructured and smooth surfaces;
Kq is the dimensionless criterion, namely, the scale of the averaged velocity of liquid
resulting from a vapor generation process: Kq = ql0 /(rρ ν ); l0 is the Laplace constant

l0 = σ/( g(ρ − ρ )); ρ and ρ” are the liquid and vapor density; ν is the fluid kinematic
viscosity; r is the latent heat of vaporization; σ is the surface tension coefficient. Geometrical
parameters are the following: the angle of fin inclination—θ; fin height—h; gap between
the fins—Δ; average fin thickness—δ.
The formula for α0 expresses a standardized heat transfer coefficient:
 5    
1 1 P 0.1 P 1.16 2/3
α0 = 872Pcr3 / Tcr6 M 6 1 + 4.64 q , (7)
Pcr Pcr

where Pcr and Tcr are the critical pressure and critical temperature of the coolant, and M
is the molecular weight of the coolant. The calculation dependence proposed in [29] by
Formulas (6) and (7) makes it possible to predict heat transfer without requiring empirical
parameters. However, it is worth noting that Formula (7) in our case gave overestimated
values of α0 , so we used our own experimental data for smooth surface (Figure 6a) as
reference values.
The dependence proposed by Huang [30] has the form of:
  √ 0.36  0.3  −0.2  −0.5
kl ql0 2 ρ Pf in Δ
α = 180 √ √ √ , (8)
l0 2 k l TS ρ l0 2 l0 2

where kl is the liquid thermal conductivity; TS —saturation temperature and Pfin is fin pitch.
Comparison of received experimental data on HTC vs. heat flux for microstructured
by MDC surface with the correlations for microfinned surfaces (6,8) shown in Figure 8.
For reference, Figure 8 also shows the experimental data obtained for a smooth surface.
As noted, the heat transfer enhancement on the microstructured tube is about 3–4 times
more as compared to the smooth surface even at small heat fluxes, and ONB begins at heat
flux values about 2 times less than those for the smooth surface. In fact, the maximal heat
transfer enhancement up to 4 times takes place, namely, in the region of small heat fluxes
(q: 1 ÷ 2.6 W/cm2 ) because in this case, we compare the quite developed boiling on the
MDC-surface with a highly underdeveloped boiling regime on the smooth surface.
The dependence (6) for a microfinned structure quite well describes the obtained
experimental data for the microstructured MDC-surface, even better than within a deviation
of 30%, as stated by the authors of [29]. Some excess of the experimental data over the
predicted data (here the logarithmic fit of calculated points is implied) in the region of heat
fluxes q ≤ 4 W/cm2 can be associated with the presence of additional knurling, which

85
Energies 2023, 16, 782

creates partly closed pores and intensifies heat transfer as compared to microfinned surfaces
(according to our assumptions), so with the fact that we used the dependence (6) developed
for a pool boiling to describe the falling films data: it is known that the contribution of
falling film evaporation can give an additional increase in heat transfer as compared to
pool boiling, especially at low and moderate heat fluxes, reaching 20–30%. The appearance
of a step in the calculated HTC values by (6) in Figure 8 in the region of low heat fluxes
(1 ÷ 1.8 W/cm2 ) is due to the presence of an underdeveloped boiling regime on the smooth
surface with low HTCs, which are substituted into (6) giving low values of α.

Figure 8. Comparison of HTC vs. q dependence for microstructured by deformational cutting tube
with correlations for predicting pool boiling HTC for microfinned surfaces [29,30].

The dependence (8), on the contrary, does not work in our case, about two times
exceeding the HTC data obtained, although it can be noted that both correlations have very
similar character of the dependence of HTC on the heat flux.
In conclusion, it can be added that next exceeding the heat flux values provided in
Figure 8 leads to the flattening of HTC vs. q dependence, associated with the appearance of
washable dry spots, beginning from the lower part of the heated tube. Dry spots appear
after reaching of CHF in places where the film is thinned, for example, such regions are
visible in Figure 5a like bare zones (but in fact, these regions are still covered with a
microlayer of liquid film). The microstructured surfaces, similar to those used in this work,
in addition to increasing the heat transfer coefficient and reducing the ONB, provide the
capillary replenishment of nucleation centers with liquid, making it possible to delay the
development of the crisis phenomena. According to our estimates, based on a study of heat
transfer on similar MDC-surfaces earlier [22], the minimal increase in CHF on the MDC-
surface used in this work may be about 2 times as compared to the smooth surface, namely,
to reach about 12 W/cm2 for the upper values of the passed range of liquid flow rates.
Below, we briefly discuss the possible causes of reducing heat transfer when using the
combined coverage.
As already mentioned we attribute the absence of heat transfer enhancement for the
stainless steel (AISI 304) mesh-covered smooth copper surface primarily to the low thermal
conductivity of the used mesh material (λ ≈ 15 W/mK) and the imperfect contact of the
mesh with the heated surface (allowing the existence of gaps ~10 microns at the points of
contact (despite our attempts to tighten the mesh covering on the tube as closely as possible).
Maybe the second reason is even more important because there are works in which it is
shown that steel coverings can be effective means of heat transfer. For example, in [31] it is
shown that HTC provided by 3D-printed capillary-porous stainless steel coating (LPW 155,
λ ≈ 20 W/mK) can be higher than HTC of the brass one (AISI C836000, λ ≈ 89 W/mK)
with the same parameters of porous structure, at boiling of n-dodecane in horizontal liquid
layers at reduced pressures. Thus, these issues require additional research.

86
Energies 2023, 16, 782

Also, especially in the case of combined surfaces, we should not discount the possi-
bility that guided by the choice of mesh parameters based on the empirical data of other
authors [5,6,8] as well as our own previous data for microstructured surfaces [22], we may
not choose the mesh coverage parameters well enough for our experimental conditions,
which may cause the deterioration of vapor removal.
Studying the one-layer mesh coverings with an aperture size larger than the bubble
departure diameter as well as the order of the departure diameter and smaller than the
departure diameter, Tolubinskiy et al. [32] showed that the maximum HTC is realized at an
aperture of the order of the bubble departure diameter. Based on this, we can assess that it
is also true for the upper layer of two-layer coverings (or combined coatings with mesh
covering the basic microstructure, like in our case). Then, the basic microstructure—or
fine mesh in the case of two-layer mesh covering—is responsible for the intensification
of the nucleation process and the upper coverage at least should not prevent effective
vapor removal.
To estimate the value of departure diameter for the R114/R21 refrigerant mixture use
the Labuntsov formula [33]:
 1/3
σd0
D = 1.8 (9)
g(ρ − ρ )

where d0 is the characteristic size of the microroughness of the heating surface.


Based on Mahmoud and Karayiannis [34] calculation, who used the Hsu model [35]
to predict the range of active cavity size for different liquids at atmospheric pressure and
5 K subcooling (5.5–126 μm, 0.7–108 μm, 0.4–79 μm—for water, HFE-7100 and FC-72,
respectively), we can assume, without greatly deviating from the truth, that the range of
active nucleation sites will be 1–100 μm for the R114/R21 refrigerant mixture (note, that
this range includes the transverse dimensions of the gap between microfins, Figure 1b).
Substituting the minimum and maximum values of the range into Formula (9) as d0 , we
have: D ≈ 190 μm for the smallest cavities and D ≈ 890 μm for the biggest ones.
Thus, according to the calculation, by using the mesh covering with an aperture of
220 μm perhaps we are preventing vapor removal. When choosing the mesh parame-
ters, besides basing the choice on our own results on enhancing surfaces microstructure
parameters [22], we were guided by the parameters of enhancing covering from [6] (consist-
ing of three upper layers of mesh with aperture 254 μm and diameter 50 μm) as well as by
the uniform covering parameters from [8] (with aperture 100 μm and diameter 30 μm)—in
both cases, the aperture of upper coverings was significantly less than the bubble departure
diameter for water (D ≈ 2 mm under standard conditions). As a result, our preference
was given to mesh with an aperture of 220 μm (available with a wire diameter of 100 μm),
slightly exceeding the pitch of the lower microfinning by MDC (200 μm). However, we
did not take into account the fact that in previous work [6,8] boiling was carried out at
high heat fluxes ~100 W/cm2 , an order of magnitude higher than maximum operating heat
fluxes ~10 W/cm2 in case of boiling refrigerant mixture R114/R21 films. At such high heat
fluxes, cavities with the smallest dimensions can be activated. So, the different operating
ranges of heat fluxes used in different works may be a possible cause of the discrepancy
with respect to the choice of optimal aperture values of upper mesh coverings.
However, these are just some preliminary considerations, the issue of choosing cover-
ing geometrical parameters requires more precise analysis and calculations.

4. Conclusions
The experimental data on heat transfer coefficients at evaporation and boiling of a
refrigerant mixture R114/R21 film falling over the smooth or microstructured surfaces and
the same surfaces, covered by micromesh, are obtained. It is shown that:
• A microstructured surface created by deformational cutting demonstrates significant
heat transfer enhancement—up to four times as compared to the smooth surface.

87
Energies 2023, 16, 782

• Adding mesh covering with an aperture of 220 μm reduces the enhancement reached
on the microstructured surface alone by up to two times. Thus, the heat transfer
enhancement, provided by a combined surface, is about two times as compared to the
smooth surface.
• The chosen mesh coverage overlaying the smooth surface does not give heat transfer
enhancement compared to the smooth surface.
• The absence or deterioration of heat transfer enhancement on studied mesh-covered
surfaces can be associated with the low thermal conductivity of used mesh material
and especially the imperfectness of the mesh mounting method. The applied mesh
covering possibly prevents the effective vapor removal during the boiling of the
R114/R21 mixture falling film.
• The correlation, proposed by Aksyanov et al. for predicting the pool boiling heat
transfer on microfinned surfaces, describes the obtained film flow data on heat transfer
on the MDC surface quite well, while the correlation of Huang gives twice the values
of the heat transfer coefficient.

Author Contributions: Conceptualization, O.V. and N.P.; methodology, O.V. and N.P.; validation,
O.V. and N.P.; investigation, O.V. and N.P.; resources, A.P.; writing—original draft preparation, O.V.;
writing—review and editing, O.V. and N.P.; visualization, O.V.; project administration, A.P.; funding
acquisition, A.P. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by a mega-grant from the RF Ministry of Science and Higher
Education (No. 075-15-2021-575); studies of heat transfer on the MDC-surface were carried out within
the framework of the state assignment of the IT SB RAS.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design
of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or
in the decision to publish the results.

Nomenclature

A heat releasing surface area


C correction factor
D bubble departure diameter
d tube diameter; characteristic size of microroughness
G mass flow rate
g gravity acceleration
h heat transfer coefficient; fin height
I current through the heating element
kl liquid thermal conductivity
l0 Laplace constant
P Pressure
Pfin fin pitch
q heat flux density
R heating element resistance
Ra mean roughness
Re film Reynolds number, Re = πdμ 4G

T temperature
U uncertainty
Greek symbols
α heat transfer coefficient
Δ gap between the fins
δ liquid film thickness, average fin thickness
δgap size of gap between fin tips

88
Energies 2023, 16, 782

θ the angle of fin inclination


λ thermal conductivity
μ dynamic viscosity
ρ Density
σ surface tension
Subscripts
cr Critical
s Saturation
w Wall
Superscripts
 liquid phase
 vapor phase

References
1. Huang, Y.; Wang, M.; Xu, L.; Deng, J. Experimental study on a spray and falling-film cooling system. Case Stud. Therm. Eng. 2021,
26, 101057. [CrossRef]
2. Liang, G.; Mudawar, I. Review of pool boiling enhancement by surface modification. Int. J. Heat Mass Transf. 2019, 128, 892–933.
3. Dedov, A.V. A Review of Modern Methods for Enhancing Nucleate Boiling Heat Transfer. Therm. Eng. 2019, 66, 881–915.
[CrossRef]
4. Volodin, O.A.; Pecherkin, N.I.; Pavlenko, A.N. Heat Transfer Enhancement at Boiling and Evaporation of Liquids on Modified
Surfaces—A Review. High Temp. 2021, 59, 405–432. [CrossRef]
5. Volodin, O.A.; Pavlenko, A.N.; Pecherkin, N.I. Heat Transfer Enhancement on Multilayer Wire Mesh Coatings and Wire Mesh
Coatings Combined with Other Surface Modifications—A Review. J. Eng. Thermophys. 2021, 30, 563–596. [CrossRef]
6. Zhang, S.; Jiang, X.; Li, Y.; Chen, G.; Sun, Y.; Tang, Y.; Pan, C. Extraordinary Boiling Enhancement Through Micro-Chimney Effects
in Gradient Porous Micromeshes for High-Power Applications. Energy Convers. Manag. 2020, 209, 112665. [CrossRef]
7. Asakavičius, J.P.; Zukauskas, A.A.; Gaigalis, V.A.; Eva, V.K. Heat Transfer from Freon-113, Ethyl Alcohol and Water with Screen
Wicks. Heat Transf. Sov. Res. 1979, 11, 92–100.
8. Tang, K.; Bai, J.; Chen, S.; Zhang, S.; Li, J.; Sun, Y.; Chen, G. Pool Boiling Performance of Multilayer Micromeshes for Commercial
High-Power Cooling. Micromachines 2021, 12, 980. [CrossRef]
9. Sloan, A.; Penley, S.; Wirtz, R.A. Sub-Atmospheric Pressure Pool Boiling of Water on a Screen-Laminate Enhanced Surface. In
Proceedings of the 25th Annual IEEE Semiconductor Thermal Measurement and Management Symposium, San Jose, CA, USA,
15–19 March 2009.
10. Tsay, J.Y.; Yan, Y.Y.; Lin, T.F. Enhancement of Pool Boiling Heat Transfer in a Horizontal Water Layer Through Surface Roughness
and Screen Coverage. Heat Mass Transf. 1996, 32, 17–26. [CrossRef]
11. Brautsch, A.; Kew, P.A. The Effect of Surface Conditions on Boiling Heat Transfer from Mesh Wicks. In International Heat Transfer
Conference Digital Library; Begel House Inc.: Danbury, CT, USA, 2002.
12. Chien, L.H.; Tsai, Y.L. An Experimental Study of Pool Boiling and Falling Film Vaporization on Horizontal Tubes in R-245fa. Appl.
Therm. Eng. 2011, 31, 4044–4054. [CrossRef]
13. Chien, L.H.; Hwang, H.L. An Experimental Study of Boiling Heat Transfer Enhancement of Mesh-on-Fin Tubes. J. Enhanc. Heat
Transf. 2012, 19, 75–86. [CrossRef]
14. Kim, H.; Park, Y.; Kim, H.; Lee, C.; Jerng, D.W.; Kim, D.E. Critical Heat Flux Enhancement by Single-Layered Metal Wire Mesh
with Micro and Nano-Sized Pore Structures. Int. J. Heat Mass Transf. 2017, 115, 439–449. [CrossRef]
15. Dabek,
˛ L.; Kapjor, A.; Orman, Ł.J. Distilled Water and Ethyl Alcohol Boiling Heat Transfer on Selected Meshed Surfaces. Mech.
Ind. 2019, 20, 701. [CrossRef]
16. Zhang, C.; Zhang, L.; Xu, H.; Li, P.; Qian, B. Performance of Pool Boiling with 3D Grid Structure Manufactured by Selective Laser
Melting Technique. Int. J. Heat Mass Transf. 2019, 128, 570–580. [CrossRef]
17. Pastuszko, R.; Kaniowski, R.; Wójcik, T.M. Comparison of Pool Boiling Performance for Plain Micro-Fins and Micro-Fins With a
Porous Layer. Appl. Therm. Eng. 2020, 166, 114658. [CrossRef]
18. Huang, S.; Wang, L.; Pan, Z.; Zhou, Z. Experimental Investigation of a New Hybrid Structured Surface for Subcooled Flow
Boiling Heat Transfer Enhancement. Appl. Therm. Eng. 2021, 192, 116929. [CrossRef]
19. Chen, G.; Fan, D.; Zhang, S.; Sun, Y.; Zhong, G.; Wang, Z.; Wan, Z.; Tang, Y. Wicking capability evaluation of multilayer composite
micromesh wicks for ultrathin two-phase heat transfer devices. Renew. Energy 2021, 163, 921–929. [CrossRef]
20. Hu, Y.; Lei, Y.; Liu, X.; Yang, R. Heat transfer enhancement of spray cooling by copper micromesh surface. Mater. Today Phys.
2022, 28, 100857. [CrossRef]
21. Smakulski, P.; Sławomir, P. A review of the capabilities of high heat flux removal by porous materials, microchannels and spray
cooling techniques. Appl. Therm. Eng. 2016, 104, 636–646. [CrossRef]
22. Volodin, O.; Pecherkin, N.; Pavlenko, A.; Zubkov, N. Surface Microstructures for Boiling and Evaporation Enhancement in Falling
Films of Low-Viscosity Fluids. Int. J. Heat Mass Transf. 2020, 55, 119722. [CrossRef]

89
Energies 2023, 16, 782

23. Zubkov, N.N.; Ovchinnikov, A.I.; Vasil’ev, S.G. Tool–workpiece interaction in deformational cutting. Russ. Eng. Res. 2016, 36,
209–212. [CrossRef]
24. Zubkov, N.; Poptsov, V.; Vasiliev, S. Surface Hardening by Turning without Chip Formation. Jordan J. Mech. Ind. Eng. 2017, 11,
13–19.
25. Pecherkin, N.I.; Pavlenko, A.N.; Volodin, O.A. Heat transfer and critical heat flux at evaporation and boiling in refrigerant mixture
films falling down the tube with structured surfaces. Int. J. Heat Mass Transf. 2015, 90, 149–158. [CrossRef]
26. Zhukov, V.E.; Mezentseva, N.N.; Pavlenko, A.N. Teplootdacha Pri Kipenii Na Modifitsirovannoy Poverhnosti Vo Freone R21
I Smesi Freonov R114/R21 (Boiling Heat Transfer on Modified Surface in R21 Freon and in R114/R21 Freon Mixture). In
Proceedings of the XXXVIII Siberian Thermophysics Seminar dedicated to the 65th Anniversary of the Kutateladze Institute of
Thermophysics SB RAS, Novosibirsk, Russia, 29–31 August 2022; pp. 106–111. (In Russian).
27. Dang, C.; Jia, L.; Peng, Q.; Huang, Q.; Zhang, X. Experimental and analytical study on nucleate pool boiling heat transfer of
R134a/R245fa zeotropic mixtures. Int. J. Heat Mass Transf. 2018, 119, 508–522. [CrossRef]
28. Shamirzaev, A. On the Pressure Drop Calculation During the Flow of Two-Phase Non-Azeotropic Mixtures. Int. J. Multiph. Flow
2022, 104314. [CrossRef]
29. Aksyanov, R.A.; Kokhanova, Y.S.; Kuimov, E.S.; Gortyshov, Y.F.; Popov, I.A. Recommendations for Improving the Efficiency of
Radio-Electronic Equipment Cooling Systems. Russ. Aeronaut. 2021, 64, 291–296. [CrossRef]
30. Huang, L.D. Pool Boiling Correlations for Structured Fin Tubes. In Proceedings of the 10th International Conference on Boiling
and Condensation Heat Transfer, Nagasaki, Japan, 12–15 March 2018.
31. Shvetsov, D.A.; Pavlenko, A.N.; Brester, A.E.; Zhukov, V.I. Inversiya Crivoy Kipeniya Na Mikrostrukturirovannyh Poristyh
Pokrytiyah (Boiling Curve Inversion on Microstructured Porous Coatings). In Proceedings of the 8th Russian national conference
on heat transfer “PHKT-8”, Moscow, Russia, 17–22 October 2022; pp. 83–84. (In Russian).
32. Tolubinskiy, V.I.; Antonenko, V.A.; Ivanenko, G.V. Crisis Phenomena in Boiling on Submerged Wire Mesh-Wrapped Wall. Heat
Transf. Sov. Res. 1989, 21, 531–535.
33. Labuntsov, D.A. Heat Transfer and Vapor Dynamics. In Current Concepts of the Mechanism of Nucleate Boiling of Liquids; Institute of
High Temperatures: Moscow, Russia, 1974; Available online: https://inis.iaea.org/collection/NCLCollectionStore/_Public/06/1
71/6171615.pdf (accessed on 28 November 2022).
34. Mahmoud, M.M.; Karayiannis, T.G. Pool boiling review: Part I–Fundamentals of boiling and relation to surface design. Therm.
Sci. Eng. Prog. 2021, 25, 101024. [CrossRef]
35. Hsu, Y.Y. On the size range of active nucleation cavities on a heating surface. J. Heat Transf. 1962, 84, 207–213. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

90
energies
Article
Nonisothermal Evaporation of Sessile Drops of Aqueous
Solutions with Surfactant
Sergey Misyura 1, *, Andrey Semenov 1,2 , Yulia Peschenyuk 1,2 , Ivan Vozhakov 1,2 and Vladimir Morozov 1

1 Kutateladze Institute of Thermophysics SB RAS, 630090 Novosibirsk, Russia


2 Faculty of Physics, Novosibirsk State University, 630090 Novosibirsk, Russia
* Correspondence: [email protected]

Abstract: In recent decades, electronic devices have tended towards miniaturization, which neces-
sitates the development of new cooling systems. Droplet cooling on a heated wall is effectively
used in power devices with high heat flux densities. The use of a surfactant leads to an increase in
the diameter of the wetted spot and the rate of droplet evaporation. Despite the wide interest and
numerous works in this area, there are still unexplored questions regarding the influence of surfactant
and wall temperature on convection, of nonisothermality, and of the decrease in the partial pressure
of vapor with increasing surfactant concentration. This work experimentally studies the effect on the
rate of droplet evaporation of wall temperature in the range 20–90 ◦ C and of the concentration of
surfactant in an aqueous solution of sodium lauryl sulfate (SLS) from 0 to 10,000 ppm. It is shown for
the first time that an inversion of the evaporation rate related to the droplet diameter occurs with
increasing wall temperature. The influence of key factors on the evaporation of a water droplet with
SLS changes with temperature. Thus, at a slightly heated wall, the growth of the droplet diameter
becomes predominant. At high heat flux, the role of nonisothermality is predominant. To determine
the individual influence of the surfactant on the partial pressure of water vapor, experiments on the
evaporation of a liquid layer were carried out. The obtained results and simplified estimates may be
used to develop existing calculation models, as well as to optimize technologies for cooling highly
heated surfaces.

Keywords: droplet evaporation; heated wall; surfactant; heat transfer; free convection
Citation: Misyura, S.; Semenov, A.;
Peschenyuk, Y.; Vozhakov, I.;
Morozov, V. Nonisothermal
Evaporation of Sessile Drops of 1. Introduction
Aqueous Solutions with Surfactant.
In the modern world, the main tasks of technical design, such as weight reduction
Energies 2023, 16, 843. https://
and object miniaturization, face implementation problems due to the disadvantages of
doi.org/10.3390/en16020843
conventional heat removal systems. Efficient dissipation of high heat fluxes requires
Academic Editor: Chi-Ming Lai new approaches to cooling [1]. At present, various two-phase heat exchangers such as
Received: 7 December 2022
spray systems [2], film cooling [3], and microchannel heat exchangers [4] are being widely
Revised: 26 December 2022
investigated. In addition, some researchers are considering the possibility of changing the
Accepted: 29 December 2022
properties of the working fluid with the help of various additives [5,6].
Published: 11 January 2023 One of the new directions in the study of the effectiveness of cooling systems is the
use of aqueous solutions with surfactants as a working fluid [7]. However, the addition of
a surfactant can reduce the liquid evaporation rate, leading to a decrease in the equipment
efficiency. The easiest way to find out the influence of this factor is to study the evaporation
Copyright: © 2023 by the authors. of an immobile sessile drop. The problem of an evaporating drop has attracted the attention
Licensee MDPI, Basel, Switzerland. of researchers since the end of the last century. A well-known problem is that of the “coffee
This article is an open access article drop” [8]: the drop is deposited on a rough substrate and its contact line remains stationary
distributed under the terms and during evaporation, when the capillary flow carries the material from the center to the
conditions of the Creative Commons edge and, on drying, results in a circular trail. Another case is when the drops are applied
Attribution (CC BY) license (https://
to a smooth, well-wettable surface. In this case, the drop spreads, i.e., the contact line is
creativecommons.org/licenses/by/
movable [9,10].
4.0/).

Energies 2023, 16, 843. https://doi.org/10.3390/en16020843 91 https://www.mdpi.com/journal/energies


Energies 2023, 16, 843

The theoretical description of evaporating droplets is complicated by a large num-


ber of physical effects that must be taken into account. The problem of droplet spread-
ing and evaporation is the subject of many works, which are summarized in a sound
review by Bonn et al. [11]. In the classical mathematical formulation of the mass trans-
fer model in single-component spherical droplets evaporating in still air, Maxwell’s ap-
proach was used [12]. It was assumed that the ratio of diffusion time to evaporation time
R2d /Dv τe ≈ ρvs (1 − RH )/ρ a is small, and then evaporation may be considered as a quasi-
stationary process. In this case, the mass flow is limited by diffusion and for a spherical
drop, it is equal to:
J M = 4πRd Dv (ρvs − ρva )
Assuming that the vapor obeys the ideal gas law, the vapor density can be expressed
in terms of the partial pressure:

p v Mv RHPsat ( T ) Mv
ρv = =
RT RT
In the case of evaporation of a liquid drop in air, the resulting vapor on the droplet
surface displaces the surrounding vapor-air mixture and forces it to move away from the
drop at a certain velocity. This phenomenon is called the Stefan flow. The evaporation rate
of a spherical droplet is limited by the Stefan flow [13]:

JS = 4πRd Dv ρ a ln(1 + B M )
B M = (ρvs − ρva )/(ρ a − ρvs )

Tonini and Cossali [14] extended this model to the case of nonisothermal evaporation
of a spherical drop in the surrounding atmosphere.
When considering
 small sessile drops with a contact radius less than the capillary
length λc = σ/ρl g, the gravitational forces are insignificant, so the liquid takes the form
of a spherical cap. The height hd , surface area Sd and volume Vd of a spherical cap can be
calculated using the wet spot radius Rd and the contact angle θ:

hd = Rd tan(θ/2)
Sd = π (h2d + R2d )
Vd = πh6 d (3R2d + h2d )

Hu and Larsen [15] developed an isothermal evaporation model for a sessile drop. An
analytical dependence of the mass flow on the contact angle was obtained. The developed
theory is in good agreement with the experimental data and the results of numerical
calculations for contact angles from 0 to π/2. According to this model, the mass flow is:

JH = πRd Dv (ρvs − ρva )(0.27θ 2 + 1.3)

Experimental studies of the evaporation of a sessile drop have been carried out for
a long time and have resulted in extensive and versatile data. The papers [16–18] deal
with water drop evaporation in the open air, showing the influence of such parameters as
substrate material, angle of inclination, and environmental conditions on the evaporation
process. By now, a relatively large amount of information on the evaporation of various
liquids has been collected.
Tarasevich [19] performed an analytical analysis of hydrodynamics inside a sessile
drop at the initial stage of evaporation. The qualitative picture of the velocity field was
shown to be independent of the ratio between the drop height and the radius of the contact
line. Saada et al. [20] numerically simulated the evaporation of a drop lying on a heated
substrate. The simulation took into account the convection of the surrounding air as a result
of heating from the substrate, but heat was considered to be transferred inside the drop only
due to thermal conduction. The results revealed an underestimation of the evaporation

92
Energies 2023, 16, 843

rate compared to the experimental data, and the error increased with increasing substrate
temperature. Timm et al. [21] presented data on the evaporation rate in comparison with
the diffusion dependence. This assumption was shown to be valid when the time scale of
vapor concentration diffusion was much smaller than the total time of droplet evaporation.
However, these conditions are valid for experiments in open air.
Misyura [22] demonstrated that at high corrosion rates and intense gas evolution,
the evaporation rate decreases due to a decrease in the rate of heat supply to the droplet
surface. Modeling the corrosion kinetics also requires taking into account the statistical
nature of the corrosion process, which depends on the wettability and droplet diameter.
It was shown for the first time that the corrosion kinetics is determined by the wetting
regimes. Misyura et al. [23] also studied the effect of the thermocapillary convection and
surfactant in a sessile droplet during nonisothermal evaporation.
Misyura [24] investigated the influence of the free convection. The Marangoni
thermal convection was shown to play a predominant role in the heat transfer in the
liquid evaporation.
Truskett and Stebe [25] investigated the effect of an insoluble surfactant monolayer on
the evaporation rate. In this work, no difference from pure water was found. Kim et al. [26]
studied the flash evaporation of water droplets with the addition of a surfactant. The latter
was shown to strongly affect the shape of the evaporating droplet due to Marangoni forces.
Semenov et al. [27] conducted an experimental and theoretical study of the contact angle
during the evaporation of drops with surfactant solutions. Guti’errez et al. [28] proved that
during vacuum evaporation, a surfactant reduces the rate of water evaporation.
For normal liquids, in particular pure water and most aqueous surfactant solutions [29],
the surface tension decreases with increasing temperature. This means that the derivative
of the surface tension coefficient with respect to temperature is negative. Hence, thermocap-
illary Marangoni forces arise with a nonuniform temperature distribution on the surface.
The result is an outflow of fluid from warm area to cold area.
The term “self-wetting” was introduced by Abe et al. [30], who studied the thermo-
physical properties of dilute aqueous solutions of high-carbon alcohols. Due to thermo-
capillary stresses and the shape of the surface tension–temperature curve, the investigated
liquids spread in a “self-wetting” manner, spontaneously propagating towards hot areas,
and thereby preventing hot surfaces from drying out and increasing the rate of heat transfer.
Because of these properties, “self-wetting” liquids were associated with significantly higher
critical heat fluxes compared to water [31]. Savino et al. studied the behavior of vapor
locks inside wickless heat pipes [32]. They found that the plugs were significantly smaller
than those of liquids such as water. The study of self-wetting liquids was carried out in
microgravity for space applications on the International Space Station [33]. Hu et al. [34]
demonstrated that the use of these fluids in micro-oscillatory heat pipes resulted in an
increase in heat pipe efficiency.
The use of alcoholic solutions with water may be undesirable, since they can signifi-
cantly differ in thermophysical properties from pure water. At the same time, a very small
amount of this substance is used for creating a surfactant solution, which does not affect its
thermophysical properties, but noticeably changes its surface properties.
The performed analysis of existing works has shown that there are very few exper-
imental and theoretical data on the influence of various factors on the evaporation of a
drop of an aqueous solution with surfactant in a wide range of wall temperatures. The
complexity of these studies is related to the fact that at high-temperature evaporation, the
isothermal model of diffusion is unacceptable and evaporation is influenced simultaneously
by many key factors. In addition, the properties of the surfactant change strongly with
increasing liquid temperature.
Previous experimental studies have demonstrated that the addition of a surfactant
leads to a decrease in the evaporation rate of a liquid droplet. However, the specific factors
influenced by the surfactant have not been established. In addition, there has been no

93
Energies 2023, 16, 843

comprehensive study of the effect of the surfactant in a wide range of wall temperatures
and droplet base diameters.
Research objectives of this article are as follows. (1) Experimental determination of
geometrical parameters of a droplet and measurement of temperature fields of a liquid
droplet. The obtained experimental data are used to predict the evaporative behavior
of droplets in a wide range of wall temperatures. (2) On the basis of experimental data,
approximate calculations are made to estimate the influence of separate key factors on the
droplet evaporation rate. The findings may be useful for the development of the existing
models for determining the evaporation rate of droplets of solutions with surfactants,
which is important for the creation of droplet irrigation technologies, widely used for heat
exchange intensification in devices with high energy efficiency.

2. Experimental Setup
Preparation for the experimental study of droplet evaporation involved obtaining
aqueous solutions with surfactants. In the presented work, three different surfactants
were selected: polyoxyethylene sorbitan monooleate (TWEEN 80), cetrimonium bromide
(CTAB), and sodium lauryl sulfate (SLS). They differed in properties depending on the
charge component of their hydrophilic components: nonionic (TWIN80), cationic (CTAB),
or anionic (SLS).
Using a high-precision micro analytical balance AND BM-252 with a readability of 0.01
mg, the weight of the surfactant was measured. It was then mixed with deionized nanofil-
tered Milli-Q pure water, and the solution was thoroughly stirred at room temperature in a
closed cuvette, using a magnetic stirrer with medium speed.
Attempts to obtain an aqueous solution of TWIN80 have demonstrated that producing
the solutions of nonionic surfactants requires a special procedure, described in chemical
protocols, including special conditions and additional reagents. Concluding that TWIN80
is poorly soluble in water, it was decided not to use it for our experiments. The study of the
remaining solutions is carried out using two experimental setups described below.

2.1. Tensiometer KRÜSS K100


The analysis of the properties of aqueous solutions with surfactants was carried out
on a precision tensiometer KRÜSS K100 with KRÜSS laboratory software. Using a Huber
thermostat, a cuvette with a working liquid was heated or cooled in the temperature range
from 294 K to 343 K. Measurements were performed using the Wilhelmy plate method. A
standard 10 mm × 20 mm × 0.2 mm platinum plate was used to measure surface tension
since the contact angle of such a plate with most liquids is approximately zero. The second
plate with dimensions of 10 mm × 20 mm × 2.8 mm was made of copper of industrial
grade A1 and was used to measure contact angles. The copper surface was technically
treated. In most cases, copper with high thermal conductivity and low cost is used as a
main material for heat pipes. In the case of using a nonstandard plate, a special holder was
used to ensure a strictly vertical position for the copper plate.
The tensiometer placed the platinum strip in the solution, then slowly removed it
at a constant rate until it was completely out of the solution. During this procedure, the
tensiometer measured the force f with which the plate was removed. Furthermore, the
value of the surface tension coefficient σ was calculated by the formula: σ = f /(l · cosθ ).
In this case, the contact angle θ between the platinum plate and the investigated
fluid was taken to be zero. To measure the value of the contact angle with the copper
plate, a similar procedure was followed. To verify the obtained data, measurements were
carried out in series of 5 runs. As a result, statistically averaged data for the surface tension
coefficient σ(T) and the contact angle θ were obtained.

2.2. Drop Shape Analyzer KRÜSS DSA100


Experiments for the sessile liquid drop evaporation were performed with the use of
the DSA100 drop shape analysis system, produced by KRÜSS and presented in Figure 1.

94
Energies 2023, 16, 843

This device consists of three main parts: a high-precision dosing system with the dosing
step of 0.1 μL, a motorized object table with program-driven movement in 2 horizontal
axes, and the optical shadow system which includes a 50 W light source and a CCD
camera with a resolution of 780 × 580 pixels (field of view varied from 3.7 mm × 2.7 mm
to 23.2 mm × 17.2 mm). A Peltier chamber open to the atmosphere was used for heating
substrates. It allowed the specified temperature of the bottom wall to be maintained
with an accuracy of about 0.1 ◦ C. The thermostat was used to cool the chamber and
maintain the temperature of the substrate. The needle with the selected solution was
moved into the chamber and a drop was deposited onto the substrate surface.

Figure 1. KRÜSS DSA 100 experimental setup: 1—dosing system; 2—camera; 3—three-axis position-
ing system; 4—processed image of a droplet; 5—automatic inclination system.

Measurement of the surface temperature of the copper substrate was carried out using
a platinum resistance thermometer, fixed at the distance of 1–2 mm from the liquid drop. A
special platinum resistance thermometer pt100 with a diameter of 2.5 mm, produced by
KRÜSS, was applied. The accuracy of pt100 was 1/3 DIN class B (±0.1 ◦ C at 0 ◦ C to 0.8 ◦ C
at 400 ◦ C). To increase the reliability of temperature measurements, the wall temperature
was also measured by a thermocouple, located at a distance of 0.3–0.5 mm from the surface
of the copper substrate on which a droplet was placed. The measurement error for the
wall temperature (Tw ) made by this thermocouple did not exceed 0.5–1 ◦ C (taking into
account the estimated distance of the thermocouple from the wall). The substrate surface
temperature in each experiment was kept constant. The temperature and relative humidity
of the ambient air during the experiments was 20–21 ◦ C and 30–32%, respectively.
The measurement error of the droplet diameter did not exceed 5%. In the range of
contact angle of 90–20◦ , the inaccuracy of the contact angle of the drop ranged within 3–5%.
The maximum error of the evaporation rate did not exceed 5–7%.
The initial liquid drop volume was about 2–3 μL. The KRÜSSDSA Advance software
allowed the processing of the shadow images of evaporating liquid droplets in fully
automatic mode 2 (real-time determination of the droplet volume, base diameter and
contact angles). Figure 2 shows a photograph of the liquid droplet shape analysis.

95
Energies 2023, 16, 843

Figure 2. Photograph of liquid droplet shape analysis.

3. Experimental Results
The presented results of the experimental studies are divided into two main parts. To
begin with, the analysis of aqueous solutions of SLS and CTAB on the KRÜSS tensiometer
is considered. The second part is devoted to the results of the study of droplet evaporation.

3.1. Surface Tension


The wide temperature range specific for the operation of a pulsating heat pipe imposes
restrictions on the applicability of working fluids. One of the determining roles in heat
transfer and efficient operation of a pulsating heat pipe is played by the properties of the
working fluid used.
At the outset, researchers obtained data on the dependence of surface tension on
the concentration of surfactants in an aqueous solution at a room temperature of about
20 ◦ C. For this purpose, a wide range of concentrations of surfactant solutions was used.
The maximum concentration exceeded the CMC (critical micelle concentration), while the
minimum concentration was similar in terms of surface tension to that of pure water. The
components of the experimental setup were carefully prepared before each measurement.
To reduce contamination caused by the surfactants used, the cuvette for the investigated
working fluid was cleaned with a 70% isopropanol solution and washed several times with
clean water.
Aqueous solutions with four different concentrations of SLS were investigated and
pure water was examined for comparison. Figure 3 shows temperature dependence of
surface tension of aqueous solutions with surfactants.

Figure 3. Surface tension vs. temperature for different concentrations of SLS in aqueous solutions.

96
Energies 2023, 16, 843

The general dependence σ(T) for CTAB is similar to that for water. That is, the first
derivative of surface tension with respect to temperature is less than zero. In a heat
exchange facility, for example, such as a pulsating heat pipe, liquid flows from the more
heated area to the less heated one occurring on the surface due to Marangoni forces. This
leads to the formation of dry spots and a decrease in the heat transfer coefficient, which
may lead to overheating of the equipment. However, the SLS solution does not exhibit
regular dependence, as CTAB, water and other normal liquids do [35]. Figure 3 shows that
dσ/dT is positive over a wide range of SLS concentrations. At the same time, it should
be noted that at low SLS concentrations, a rapid change in the sign of the derivative is
observed with increasing temperature. The temperature of the break point depends on
the concentration. At higher temperature, the surface tension of the solution completely
coincides with that of water.
Thus, aqueous solutions of SLS have the properties of a self-wetting liquid in a wide
range of concentrations and temperatures. However, self-wetting liquids are quite often
known to be unstable, which is also confirmed by the results of the current study.
From the obtained data, it may be inferred that SLS solutions are useful in solving
the problem of dry spots and expanding the working range into the region of higher heat
fluxes—with the proviso, of course, that for each task it is necessary to pay attention to the
permissible range of applicability for each specific concentration of the solution.

3.2. Evaporation
In droplet evaporation, it is important to take into account the contact angle, since this
determines the shape of the droplet surface. Therefore, the contact angles of evaporating
droplets were measured for different concentrations of surfactant solutions and different
substrate temperatures.
Figure 4 shows the evolution of a pure water drop during evaporation under isother-
mal conditions at a room temperature of about 21 ◦ C. Since the drop has a small volume of
the order of 1 μL, its shape is a spherical cap. In this case, evaporation apparently occurs
at a fixed contact line, with the so-called “pinning” of the contact line. The evaporation
is as a result of the diffusion of vapor from the drop surface, where it is saturated, into
the atmosphere with a constant vapor concentration (in the experimental conditions, the
humidity was about 30–40%). The evaporation is almost uniform over the entire surface
area and is well described by Stefan’s theoretical model [13]. Droplet evaporation on the
heated surface qualitatively coincides with isothermal evaporation, but there are important
differences. In this case, the mass flow from the droplet surface is substantially nonuniform.
A significant mass flow is provided by intense evaporation in the area of the contact line,
called the microregion [36].

Figure 4. Evolution of an evaporating water droplet at a temperature of 21 ◦ C (the initial droplet


radius Rd = 0.7mm, the initial contact angle θ 0 = 97◦ ). Time step equals 300 s.

97
Energies 2023, 16, 843

In a further step, the evaporation of drops of surfactant solutions was considered in


order to find out the effect of the surfactant concentration on the evaporation rate (heat
exchange with the substrate) and the behavior of the contact line during evaporation.
The data for a small set of drops are presented below, whereby all the given graphs of
diameter and contact angle correspond to the same drops.
Figures 5 and 6 show values of the droplet contact angle and the droplet height versus
time during evaporation. At the initial moment of time, the contact angle of droplets of
surfactant solutions is in the range from 35◦ to 100◦ . The initial droplet height is 0.4–1 mm.
Upon evaporation, the contact angle decreases, which is associated with a change in the
droplet volume, but the contact line remains almost immovable. It is clearly seen that pure
water corresponds to the maximum contact angle for both low (30 ◦ C) and higher (80 ◦ C)
temperatures.

Figure 5. Evolution of contact angle over time of evaporation.

Figure 6. Evolution of drop height over time of evaporation.

Meanwhile, at a low surfactant concentration, the contact angle does not differ from
that of pure water, and with an increase in the surfactant concentration, it monotonically
decreases. There are two reasons for this. First, the surface tension reduction leads to a
decrease in capillary forces and spreading of the droplet. Secondly, at the initial stage, the
temperature of the liquid is close to room temperature and, upon contact with the heated
substrate, it begins to heat up rapidly. It is supposed that the drop heating in the vicinity
of the contact line occurs faster than in the main volume of the drop, which gives rise to
the Marangoni force. The force vector is directed along the surface from lower to higher
temperature. Estimating the time required to heat the drop to the substrate temperature,

98
Energies 2023, 16, 843

previous researchers obtained the formula τ = R2 ρcp /λ ≈ 6 s. Thus, the droplet diameter
should grow noticeably in the first seconds of heating.
Figure 7 shows the droplet diameters as a function of time. At low surfactant concen-
trations in pure water, the diameter is seen to be constant during the entire evaporation
time, except for the last seconds of the droplet life.

Figure 7. Evolution of dropletsdiameter over time of evaporation.

With increasing concentration, the area where the change in diameter occurs is clearly
observed to stretch over time. Moreover, for a high concentration (above CMC), the droplet
diameter changes throughout the entire evaporation process: in the first seconds, the drop
diameter increases as the drop itself spreads, and then decreases as it dries.

4. Predicting the Evaporative Behavior of a Water Droplet with a Surfactant


Direct numerical simulation of the unsteady and nonisothermal evaporation of a
water droplet with a surfactant is very complicated due to the large number of uncertain
factors that affect each other. In the presence of a surfactant, it is hard to estimate the
equilibrium diameter of the droplet, which can change markedly over time when the
droplet evaporates. Any change in the diameter, contact angle or height of the droplet alters
free convection in the gas and liquid phase. The surfactant affects the partial pressure of
water vapor and, accordingly, the evaporation rate. To date, there are no reliable methods
for computing all these factors. Calculating the temperature of the free surface of the
droplet is extremely difficult, since the Marangoni flow and convection in a droplet depend
on the surfactant concentration. Therefore, assessment of the influence of various key
factors on the evaporative behavior of the droplet becomes crucial at the first stage. For
this purpose, it is advisable to use experimental data on the geometry of the droplet, as
well as on the thermal field measured by a thermal imager (Ts ) and a thermocouple (Tw ).

4.1. Key Factors Effecting the Droplet Evaporation


In general, the rate of evaporation depends on the geometry of the droplet, the
conditions of heat transfer and on free convection. It is possible to distinguish the following
key parameters that determine the intensity of evaporation of a water droplet with the
addition of a surfactant:
(1) Drop diameter;
(2) Natural convection of gas;
(3) Natural convection of liquid;
(4) Nonisothermality;
(5) Buoyancy of water vapor in air and the Stefan flow;
(6) Evaporation rate of surfactant solution.

99
Energies 2023, 16, 843

With a small droplet diameter (1–3 mm) and in the absence of wall or air heating,
convection in the liquid and gaseous phases can be neglected. When a droplet evaporates
on a hot wall, free convection has to be taken into account. The influence of these key
factors on the evaporation of liquid are considered separately below.
(1) Let us take a detailed look at the influence of the wetted droplet diameter on the
evaporation rate. In most experimental curves, there is an attached contact line for most
of the evaporation time (evaporation mode of constant contact radius, or CCR). With the
growth of the surfactant concentration, the droplet radius increases. Despite the decrease
in the contact angle with time, for most of the curves, a quasi-linear decrease in the droplet
volume is realized (i.e., at a fixed radius, the evaporation rate (dV/dt) is constant, despite a
significant decrease in θ). Thus, the effect of the surfactant on evaporation due to the change
in radius will be proportional to the ratio of the radii of the two drops (J1 /J2 ∼ Rd1 /Rd2 ).
To analyze the influence of other key factors, it is necessary to exclude the influence of the
radius being a part of the ratio J/Rd .
(2) Let us estimate the effect of gas convection on the evaporation of a drop. When
a liquid layer evaporates, the convective mass flow of water vapor through air can be
estimated using Equation (1) [37],

Sh = 2 + 0.55( Re)0.5 (Sc)0.33 (1)

where the Sherwood number Sh = βRd /Dv , the Schmidt number Sc = νg /Dv , the Reynolds
number Re = Uca Rd /νg , in which β is the mass transfer coefficient, Rd is the radius of the
droplet, Dv is the diffusion coefficient of water vapor in air, Uca is the rate of thermograv-
itational convection of air at the temperature difference (ΔTs ) between the temperature
of the free surface of the droplet Ts and the temperature of the ambient air Ta , and νg is
the kinematic viscosity of the gas. The evaporation rate J is related to the mass transfer
coefficient through the concentration difference ρvs − ρva given in Equation (2),

J = βF (ρvs − ρva ) (2)

where F is the interface area of the layer.


For intense convection, when the Sherwood number is much greater than 1,
Sh = 0.55( Re)0.5 (Sc)0.33 , then in accordance with Equations (1) and (2),

J ∼ β ∼ (Uc a )0.5 (3)

The air velocity due to thermogravitational convection can be approximated from


Equation (4),
Uca ∼ ( gβΔTs R)0.5 (4)
and the evaporation rate is related to the droplet radius according to Equation (5):

J ∼ ( Rd )0.5 (5)

As indicated in Equation (5), the presence of the surfactant leads to an increase in the
droplet radius; and an increase in the radius contributes to an increase in evaporation.
(3, 4) Let us evaluate the joint influence of factors 3 and 4, since they are interrelated.
For convenience of analysis, the effect of these factors will be denoted as the effect of
nonisothermality. The heat flux in the liquid ql is supplied to the interface of the drop
and spent on evaporation (qe = rJ, J = Δm/Δt), radiation (qr = εσ(T4 s − T4 a )) and free air
convection (qc = αa (Ts − Ta )). The heat balance for the droplet surface corresponds to
Equation (6),
 
ql = qe + qr + qc = rJ/F + εσ T 4 s − T 4 a + α a ( Ts − Ta ) (6)

100
Energies 2023, 16, 843

where r is the latent heat of vaporization, and F is the area of the liquid interface. The air heat
transfer coefficient αa is determined by the Nusselt number Nu = αa Rd /λa = 0.54 Ra0.2 , where
Ra is the Rayleigh number [37]. The heat supplied to the liquid from the wall is determined
by the conductive and convective heat transfer in the liquid according to Equation (7) [38–40],

ql = αl ( Tw − Ts )
 0.3
 2 0.3
λl
αl = h
Ucl h
+1
al ∼ λhl khal + 1 ∼ 0.4
1
, if Uacl h 1 (7)
 0.3  2 0.3 h l

αl = λhl Uacl h + 1 ∼ λhl kha + 1 ∼ 1h , if Uacl h 1


l l l

where αl is the liquid heat transfer coefficient, Ucl ∼ c(Ma + Ra), Ul is the average convective
velocity in the liquid droplet, the Marangoni number Ma = (ΔTw h/μl al )(dσ/dTs ), the
Rayleigh number Ra = gβT ΔTw h3 /νl al , βT is the thermal liquid expansion, g is the gravity
acceleration, ΔTw = Tw − Ts , νl is the water kinematic viscosity, al is the liquid thermal
diffusivity, and μL is the water dynamic viscosity. Empirical expressions for evaporation
intensification with an increase in the Ra number are given in [41]. For very small droplets
(droplet height below 1 mm), low values of Ra and Ma and low values of Ul are realized,
and the condition αl ∼ 1/h (conductive heat transfer) is satisfied. Therefore, the ratio of
heat fluxes for a pure water drop (ql1 ) to a water drop with surfactant (ql2 ) will be inversely
proportional to the ratio of heights ql1 /ql2 ∼ h2 /h1 . To determine the relationship between
ΔTw and the heat flux, it is necessary to solve a system of differential equations and it
is difficult to make a simplified estimate for the change in Δρs due to nonisothermality.
Therefore, the effect of nonisothermality can only be determined after the evaluation of
other parameters (key factors 1, 2, 5, 6) and taking into account experimental data. At
different wall temperatures, the droplet free surface temperature and ρs were determined
using thermal imaging measurements and the equilibrium curve for the equilibrium partial
pressure of water vapor.
(5) Let us consider the effect of water vapor buoyancy. Since water vapor is lighter
than air, the effect of the additional flow must be considered. The vapor convection velocity
Uρ corresponds to Equations (8) and (9) [42],

Uρ = ((ρ a − ρm ) gL/ρm )0.5 (8)

ρm = ρ a − (ρ a − ρv )/ρv )Ccat ( Ta ) (9)


where ρa is the density of the air, ρv is the density of the vapor, ρm is the density of
the air + vapor mixture, and L is the characteristic length of the diffusion vapor layer.
Experimental and calculated data in [42] show that the increase in the evaporation rate due
to the buoyancy of water vapor (for a wall with high thermal conductivity) is about 40%. It
does not depend on the droplet radius (Rd > 0.8 mm). Thus, despite the change in droplet
diameter due to the surfactant, the effect of buoyancy on J will be the same for a water
droplet with and without surfactant.
(6) Let us consider the influence of the surfactant on the droplet evaporation rate. Since
the surfactant alters the radius of the drop, the simultaneous influence of several factors
appears. To eliminate the effect of the droplet radius, one may consider the evaporation of
a water layer (with and without surfactant). According to Equation (10),

J = πRd Dv (ρvs − ρva )(0.27θ 2 + 1.3) (10)

at zero contact angle, the evaporation rate does not depend on the angle: in this case, the
expression for evaporation of the liquid layer thus coincides with the expression for a
droplet with a very small height, Equation (11):

J = 4.1Rd Dv (ρvs − ρva ) (11)

101
Energies 2023, 16, 843

4.2. Evaporation of a Layer of Water with Surfactant


Figure 8 presents experimental data on the evaporation of a layer of water with
(10,000 ppm) and without surfactant. Figure 8a,b gives experimental curves for layer
evaporation of Tw = 21 ◦ C, and Tw = 90 ◦ C, respectively. All curves demonstrate a linear
dependence of the layer mass on time.

(a)




m J



 


    
tV
(b)
Figure 8. Evaporation of water layer (curve 1) and water layer with 10,000 ppm surfactant (curve 2)
for layer diameter 40 mm and air humidity 30%: (a) Tw = Ta = 21 ◦ C, (b) Tw = 90 ◦ C, Ta = 21 ◦ C.

The measurement of the water layer mass is carried out using the gravimetric method.
The working section made of copper with a layer of liquid is placed on the scales. The
wall temperature is measured with a thermocouple. Although the cell radius is constant
(40 mm), the characteristic radius Rd in Equation (11) is different according to the scenario
considered. In the case of water and surfactant, the length of the meniscus on the side wall
of the cuvette is greater than for pure water, and the total radius Rd1 is thus larger than
for the water layer by about 11–12% at Tw = 21 ◦ C and by 9–10% at Tw = 90 ◦ C. Thus, to
simplify the variables to a single radius when calculating the evaporation rate of the water
with surfactant layer, it is necessary to divide the value of J/Rd by 1.11–1.12 and 1.09–1.1
(taking into account the length of the meniscus).
Figure 9 shows thermal images of the liquid layer and the profiles of Ts along the line
0L. The average surface temperature Ts of the layer increases during evaporation due to
the decrease in layer height. The central region temperature is lower than at the edges. For
a water layer with surfactant, the average temperature over the layer surface is higher than
the temperature for a pure water layer.

102
Energies 2023, 16, 843

Figure 9. Thermal images of the layer interface of (a) pure water and (b) the surfactant solution.
Temperature profiles of Ts along the 0L line for (c) the water layer and (d) the surfactant solution
layer, (a–d) Tw = 90 ◦ C, Ta = 21 ◦ C.

4.3. Thermal Imaging Measurements of the Free Surface of a Water Droplet with and
without Surfactant
A droplet of water with surfactant shows a 15–20% decrease in Jρ (compared to pure
water) at Tw = 21 ◦ C and a 9–11% decrease at Tw = 90 ◦ C. The obtained data correspond to
the previously obtained measurements in other works. In [43], the maximum decrease in
the rate of droplet evaporation with a surfactant is 25–30%. The presence of surfactant also
leads to a decrease in Jρ for the liquid layer [44]. The decrease in Jρ is due to the competing
effect of two factors—a decrease in the instantaneous area of the free surface of a drop
occupied by water molecules, and a weakening of the molecular bonds between water
molecules and surfactant molecules. The integral value of the two parameters for the entire
free surface of the liquid decreases the equilibrium partial pressure of water vapor ρvs and,
in accordance with Equation (11), the evaporation rate.
To assess the influence of nonisothermality, the temperature of the free surface of the
drop Ts was measured using a thermal imager and an optical lens providing a tenfold
increase in the image (Figure 10a). Evaporation leads to a decrease in Ts . The minimum drop
temperature corresponds to the maximum height, i.e., the center of the drop (Figure 10b).
Since the diameter of the water with surfactant drop is 1.8–2 times larger than the pure
water drop diameter, the height is significantly less. As a result, the temperature difference
ΔTw = Tw − Ts is lower by 0.4–0.7 ◦ C for a water with surfactant droplet than for water
alone. In other words, a drop with a lower height has higher values of Ts and JT . When
adding a surfactant, the increase in JT (the effect of nonisothermality) is 2–3%.

103
Energies 2023, 16, 843

Figure 10. (a) Thermal images of the droplet interface for t = 20 s (without heating): water is on
the left; water with surfactant (10,000 ppm) is on the right; and temperature profiles of the droplet
interface at a cross section of 0L: (b) water; (c)-water with surfactant (10,000 ppm).

With high temperature heating (Tw = 90 ◦ C), the droplet diameter increases by 50–60%
due to the addition of surfactant. Due to the fast evaporation rate on the heated wall,
the temperature Ts depends much more on the drop height. For a water with surfactant
droplet, the average temperature Ts (Figures 11 and 12 for t = 7 s) for the droplet surface is
4–5 ◦ C higher than that for water during almost the entire evaporation time. This effect
of nonisothermality (JT ) leads to a 14–17% increase in the evaporation rate of a droplet
of water with surfactant, compared to a pure water drop (according to the water vapor
equilibrium curve). Figures 11 and 12 show thermal images of the surface of a drop located
on a heated wall, as well as temperature distributions of Ts along the 0L axis. As the wall
temperature and Ts increase, the difference ΔTs increases (comparison with Figure 10a). The
small height of the drop and the high heat flux from the wall contribute to the temperature
field uniformity over the drop surface.

104
Energies 2023, 16, 843

Figure 11. Thermal images of the water droplet interface and temperature profiles of the droplet
interface at the cross section of 0L (Tw = 90 ◦ C).

Figure 12. Thermal images of the droplet interface and temperature profiles at the cross section of 0L
(Tw = 91 ◦ C, water with surfactant at 10,000 ppm concentration).

105
Energies 2023, 16, 843

The temperature profile of the droplet surface at a high wall temperature (Tw = 90 ◦ C)
is qualitatively similar to that at an unheated wall. (Figure 10b). The minimum temper-
ature is at the center of the drop and the maximum temperature is at the edges. This
distribution is associated with a negligible effect of free convection in the liquid, i.e., the
temperature field is determined by the conductive heat transfer due to the small values
of the Rayleigh numbers. In this case, the temperature Ts is determined by the height
of the droplet (Ts ~ 1/h). The surfactant results in a higher temperature Ts and a more
uniform temperature distribution over the droplet surface.

4.4. Assessing the Impact of Key Factors on Evaporation of the Water Droplet with Surfactant
To evaluate the effect of several factors on droplet evaporation, it is convenient to
exclude the influence of the droplet base diameter, which strongly depends on the surfactant.
For this purpose, it is convenient to attribute the value of the evaporation rate to the
diameter of the droplet.
Figure 13 provides experimental data on the dependence of J/2Rd on the wall tem-
perature Tw . Points are given for pure water (points 1), water with surfactant 1000 ppm
(points 2) and water with surfactant 10,000 ppm (points 3).

Figure 13. Drop evaporation rate depending on the wall temperature Tw . Experimental results:
1—water, 2—water with surfactant (1000 ppm), 3—water with surfactant (10,000 ppm); and by
calculation from Equation (10) for the droplet contact angle: 4—of 60◦ , and 5— of 27◦ .

For droplet evaporation at room temperature, as well as at small wall heating


(Tw = 313 ◦ K), a water drop evaporates faster than a water with surfactant one (by
14–17% at Tw = 293 ◦ K). At a temperature of Tw = 333 ◦ K, the evaporation rates for points
1–3 are close to each other. At a wall temperature of Tw = 363 ◦ K, a droplet of water
with surfactant with a concentration of 10,000 ppm evaporates 25–27% faster. Thus,
the simultaneous influence of several factors considered above leads to an inversion
of J/2Rd c with increasing temperature. The calculated curve 4 is constructed for the
maximum value of the contact angle. This angle corresponds to the experiment with the
maximum angle, which is taken as an average value over time (θ = (97◦ + 20◦ )/2 = 60◦ ).
The curve of angle change over time has a quasi-linear character. Curve 5 is plotted for
the lowest contact angle over time (θ = (35◦ + 20◦ )/2 = 27◦ ). The drop radius is also
taken as the time average in the calculation and processing of the experimental data.

106
Energies 2023, 16, 843

The diffusion coefficient of vapor in air is determined at a temperature T = (Ts + Ta )/2,


where Ts is the temperature of the free surface of the drop, Ts is the temperature of the
external air, and the temperature Ts is measured with a thermal imager. The equilibrium
partial pressure of water vapor and the value of the vapor density are determined from
the temperature of the free surface of the drop Ts .
Based on the above analysis, diagrams illustrating the influence of various factors on
the evaporation of a drop are presented in Figure 14.

(a) (b)
Figure 14. Influence of key factors on droplet evaporation, Δ = 100% (Jw,s − Jw )/Jw . (a) Tw = 20–21 ◦ C,
(b) Tw = 90 ◦ C. Key: Jw (kg/s) —pure water mass flux, Jw,s —solution mass flux for surfactant
concentration of 10,000 ppm, Jd —drop diameter effect, Jρ —partial pressure of water vapor effect,
Jc —thermogravitational convection effect, and JT —nonisothermality effect.

In Figure 14, the influence of various factors is denoted as follows: Jd is the influence
of the droplet diameter, Jρ is the influence of the equilibrium partial pressure of water
vapor, Jc is the influence of thermogravitational convection, and JT is the influence of non-
isothermality. Both at Tw = 20 ◦ C and at Tw = 90 ◦ C, the predominant effect on evaporation
is the growth of the droplet diameter due to the surfactant (Jd ). At high temperature, Jc and
JT are comparable. The decrease in the evaporation rate (Jρ ) as a result of a decrease in the
partial pressure of vapor at room temperature is double that at high-temperature heating.
As the temperature Ts increases, the effect of Jρ on evaporation decreases. The influence of
the natural convection of gas (Jc ∼ (Rd )0.5 ) in Equation (5) noticeably manifests itself at high
temperatures only (in Figure 10b, 18–21%) and at room temperature, thermogravitational
convection can be neglected. As the wall temperature increases from 20 ◦ C to 90 ◦ C, the
effect of JT on the enhancement of evaporation increases approximately six times (from
2–3% to 14–17%). Thus, the J/2Rd inversion with increasing wall temperature (when
comparing the evaporation rate of a water drop and a drop of water with surfactant) is
associated with a strong dependence of all these key factors on temperature.
At Tw = 90 ◦ C, the experimental points J/2Rd are located 67–71% above the calculated
curves. According to Figure 14b, the influence of gas convection leads to a 67–70% excess of
the experimental values of J/2Rd , compared to the calculated diffusion model (Equation 12),
if the measured free surface temperature Ts is taken for the diffusion model. As mentioned
above, 40% corresponds to an increase in the evaporation rate due to the buoyancy of
water vapor. In Figure 14, this factor is not specified, since it has the same effect on a
water drop with or without a surfactant. The increase in the evaporation rate due to
thermogravitational convection is approximately 30% (40% + 30% = 70%). The effect
of nonisothermality is expressed in a decrease in temperature Ts compared to the wall

107
Energies 2023, 16, 843

temperature Tw . Thus, for Tw = 90, 60, 40, and 21 ◦ C, the average temperature for the droplet
surface Ts is approximately 78–80 ◦ C, 54–55 ◦ C, 38–39 ◦ C, and 19–19.5 ◦ C, respectively. At
the maximum wall temperature (Tw ), the decrease in ρvs due to the decrease in Ts is 40–50%
(the maximum effect of nonisothermality on droplet evaporation).
At Tw = 60 ◦ C, the influence of convection and nonisothermality decreases. At
Tw = 40 ◦ C, the calculation is slightly lower than the experimental points. During
evaporation without heating (Tw = 20 ◦ C), the Rayleigh number for the gas and liquid
phases is low (there is no natural convection, since there is a small droplet height and
small temperature gradients due to droplet evaporation), and the experimental data are
quite accurately modeled by the diffusion model.

5. Conclusions
An experimental study of the evaporation of sessile drops of aqueous solutions with
surfactant under nonisothermal conditions has been carried out. The choice of an immobile
liquid drop as the object of study enabled the behavior of the contact line of three phases
to be investigated, and the flow of an evaporating liquid to be measured over a wide
temperature range from 20 ◦ C to 90 ◦ C. Using the optical method, droplet contact angles,
droplet heights and contact diameters of evaporating droplets were measured in the
temperature range.
An analysis of the contact diameter of the evaporating droplets has shown that pure
water and weakly concentrated solutions on a copper substrate are characterized by fixed
contact lines (so-called pinning). However, for a highly concentrated solution (above critical
micelle concentration), pinning of the droplets under study was absent. At the initial stage,
when the droplet was heated, an intense increase in the droplet diameter was observed.
Surfactant presence has a complex effect on the rate of a droplet’s evaporation due
to several factors, some of which decrease the evaporation rate while others, vice versa,
increase it:
• The largest effect on the increase in the evaporation rate of a droplet is associated with
an increase in its diameter due to a decrease in surface tension.
• It is shown that the evaporation of the surfactant solution layer is slower than for a
pure water layer due to the drop in the partial vapor pressure. The magnitude of the
effect depends on the substrate temperature. At room temperature, the decrease in the
evaporation rate is 15–20%, and for a high-temperature substrate, the change is 9–11%.
• The contribution of nonisothermality increases by a factor of six as the substrate
temperature rises from 20 ◦ C to 90 ◦ C.
• The effect of natural gas convection is significant only for nonisothermal evaporation
with a contribution of about 20%.
• The effect of free gas convection leads to an excess of the experimental values of J/2Rd
by 67–70% compared to the calculated diffusion model.
• The effect of surfactants on the specific evaporation rate is inverted with an increase in
the substrate temperature. In the isothermal case, the surfactant solution evaporates
more slowly than pure water. In the nonisothermal case, the solution evaporates faster
than pure water. This is due to the different dependence of key factors on temperature.

Author Contributions: Conceptualization, S.M. and I.V.; methodology, A.S. and V.M.; software, A.S.
and Y.P.; validation, S.M., V.M. and I.V.; formal analysis, S.M.; investigation, A.S., Y.P. and V.M.; resources,
Y.P.; data curation, S.M. and I.V.; writing—original draft preparation, S.M. and I.V.; writing—review
and editing, S.M.; visualization, V.M. and A.S.; supervision, S.M.; project administration, I.V.; funding
acquisition, I.V. All authors have read and agreed to the published version of the manuscript.
Funding: This work was supported by the grants of the Russian Science Foundation, RSF 20-79-10096.
The tensiometer KRÜSS K100 and KRÜSS DSA 100 were provided in accordance with the state
contract of Kutateladze Institute of Thermophysics SB RAS.
Conflicts of Interest: The authors declare no conflict of interest.

108
Energies 2023, 16, 843

Nomenclature

α Heat transfer coefficient


β Convective mass transfer coefficient
βT The thermal liquid expansion
λ Thermal conductivity
λc Capillary length
σ Surface tension
μ Dynamic viscosity
ν Kinematic viscosity
ρg Gas density
ρl Liquid density
ρvs Vapor density on the droplet interface
ρa Atmospheric density
ρva Vapor density in the ambient atmosphere
θ Contact angle
τe Characteristic time of evaporation
Δm The change in mass
Δt Period of time
The temperature difference between the temperature of the droplet interface
ΔTs
and the temperature of the ambient air
The temperature difference between the wall temperature and the
ΔTw
droplet interface temperature
g Gravitational acceleration
pv Partial vapor pressure
ql Total heat flux in the liquid
qe Evaporative heat flux
qr Radiative heat flux
qc Natural convection heat flux
r Latent heat of vaporization
h Droplet height
a Thermal diffusivity
j Specific mass transfer rate
J Evaporation rate
Sd Drop surface
Vd Drop volume
Rd Drop radius
F The interface area of the liquid layer
L The characteristic length of the diffusion vapor layer
BM Spalding mass transfer number
T Temperature
Ta Ambient atmosphere temperature
Ts Droplet surface temperature
Tw Substrate temperature
Uρ The vapor convection velocity
Mv Vapor molar mass
Ucl The average convective velocity in the droplet
Uca The air velocity due to thermogravitational convection
RH Relative humidity
Ma Marangoni number
Nu Nusselt number
Ra Rayleigh number
Re Reynolds number
Sh Sherwood number
Sc Schmidt number

109
Energies 2023, 16, 843

References
1. Smakulski, P.; Pietrowicz, S. A review of the capabilities of high heat flux removal by porous materials, microchannels and spray
cooling techniques. Appl. Therm. Eng. 2016, 104, 636–646. [CrossRef]
2. Wang, J.X.; Guo, W.; Xiong, K.; Wang, S.N. Review of aerospace-oriented spray cooling technology. Prog. Aerosp. Sci. 2020, 116,
100635. [CrossRef]
3. Zhang, J.; Zhang, S.; Chunhua, W.; Xiaoming, T. Recent advances in film cooling enhancement: A review. Chin. J. Aeronaut. 2020,
33, 1119–1136. [CrossRef]
4. Khan, M.G.; Fartaj, A. A review on microchannel heat exchangers and potential applications. Int. J. Energy Res. 2011, 35, 553–582.
[CrossRef]
5. Smakulski, P.; Ishimoto, J.; Pietrowicz, S. The cooling performance of the micro-solid nitrogen spray technique on the cryopreser-
vation vitrification process: A qualitative study. Int. J. Heat Mass Transf. 2022, 184, 122253. [CrossRef]
6. Pandya, N.S.; Shah, H.; Molana, M.; Tiwari, A.K. Heat transfer enhancement with nanofluids in plate heat exchangers: A
comprehensive review. Eur. J. Mech.-B/Fluids 2020, 81, 173–190. [CrossRef]
7. Gandomkar, A.; Kalan, K.; Vandadi, M.; Shafii, M.; Saidi, M. Investigation and visualization of surfactant effect on flow pattern
and performance of pulsating heat pipe. J. Therm. Anal. Calorim. 2020, 139, 2099–2107. [CrossRef]
8. Deegan, R.D.; Bakajin, O.; Dupont, T.F.; Huber, G.; Nagel, S.R.; Witten, T.A. Capillary flow as the cause of ring stains from dried
liquid drops. Nature 1997, 389, 827–829. [CrossRef]
9. Cahile, M.; Benichou, O.; Cazabat, A. Evaporating droplets of completely wetting liquids. Langmuir 2002, 18, 7985–7990.
[CrossRef]
10. Shahidzadeh-Bonn, N.; Rafai, S.; Azouni, A.; Bonn, D. Evaporating droplets. J. Fluid Mech. 2006, 549, 307–313. [CrossRef]
11. Bonn, D.; Eggers, J.; Indekeu, J.; Meunier, J.; Rolley, E. Wetting and spreading. Rev. Mod. Phys. 2009, 81, 739. [CrossRef]
12. Maxwell, J.C. Diffusion, Collected Scientific Papers; Encyclopedia Britannica: Cambridge, UK, 1877.
13. Fuchs, N.A.; Pratt, J.N.; Sabersky, R.H. Evaporation and droplet growth in gaseous media. J. Appl. Mech. 1960, 27, 759–760.
[CrossRef]
14. Tonini, S.; Cossali, G. An analytical model of liquid drop evaporation in gaseous environment. Int. J. Therm. Sci. 2012, 57, 45–53.
[CrossRef]
15. Hu, H.; Larson, R.G. Evaporation of a sessile droplet on a substrate. J. Phys. Chem. B 2002, 106, 1334–1344. [CrossRef]
16. Gatapova, E.Y.; Semenov, A.A.; Zaitsev, D.V.; Kabov, O.A. Evaporation of a sessile water drop on a heated surface with controlled
wettability. Colloids Surf. A Physicochem. Eng. Asp. 2014, 441, 776–785. [CrossRef]
17. Shekriladze, I.G. Boiling heat transfer: Mechanisms, models, correlations and the lines of further research. Open Mech. Eng. J.
2008, 2, 104–127. [CrossRef]
18. Erbil, H.Y. Control of stain geometry by drop evaporation of surfactant containing dispersions. Adv. Colloid Interface Sci. 2015, 222,
275–290. [CrossRef] [PubMed]
19. Tarasevich, Y.Y. Simple analytical model of capillary flow in an evaporating sessile drop. Phys. Rev. E 2005, 71, 027301. [CrossRef]
[PubMed]
20. Ait Saada, M.; Chikh, S.; Tadrist, L. Numerical investigation of heat and mass transfer of an evaporating sessile drop on a
horizontal surface. Phys. Fluids 2010, 22, 112115. [CrossRef]
21. Timm, M.L.; Dehdashti, E.; Darban, A.J.; Masoud, H. Evaporation of a sessile droplet on a slope. Sci. Rep. 2019, 9, 19803.
[CrossRef]
22. Misyura, S. The dependence of drop evaporation rate and wettability on corrosion kinetics. Colloids Surf. A Physicochem. Eng. Asp.
2021, 610, 125735. [CrossRef]
23. Misyura, S.Y.; Volkov, R.S.; Filatova, A.S. Interaction of two drops at different temperatures: The role of thermocapillary convection
and surfactant. Colloids Surf. A 2018, 559, 275–283. [CrossRef]
24. Misyura, S.Y. The influence of convection on heat transfer in a water layer on a heated structured wall. Int. Commun. Heat Transf.
2019, 102, 14–21. [CrossRef]
25. Truskett, V.N.; Stebe, K.J. Influence of surfactants on an evaporating drop: Fluorescence images and particle deposition patterns.
Langmuir 2003, 19, 8271–8279. [CrossRef]
26. Kim, D.O.; Rokoni, A.; Kaneelil, P.; Cui, C.; Han, L.H.; Sun, Y. Role of surfactant in evaporation and deposition of bisolvent
biopolymer droplets. Langmuir 2019, 35, 12773–12781. [CrossRef] [PubMed]
27. Semenov, S.; Trybala, A.; Agogo, H.; Kovalchuk, N.; Ortega, F.; Rubio, R.G.; Starov, V.M.; Velarde, M.G. Evaporation of droplets of
surfactant solutions. Langmuir 2013, 29, 10028–10036. [CrossRef]
28. Guti’errez, G.; Benito, J.M.; Coca, J.; Pazos, C. Vacuum evaporation of surfactant solutions and oil-in-water emulsions. Chem.
Eng. J. 2010, 162, 201–207. [CrossRef]
29. Guo, D.S.; Li, X.B.; Zhang, H.N.; Li, F.C.; Ming, P.J.; Oishi, M.; Oshima, M. Experimental study on the characteristics of temperature
dependent surface/interfacial properties of a non-ionic surfactant aqueous solution at quasi-thermal equilibrium condition. Int. J.
Heat Mass Transf. 2022, 182, 122003. [CrossRef]
30. Abe, Y.; Iwasaki, A.; Tanaka, K. Microgravity experiments on phase change of self-rewetting fluids. Ann. N. Y. Acad. Sci. 2004,
1027, 269–285. [CrossRef]

110
Energies 2023, 16, 843

31. Suzuki, K.; Nakano, M.; Itoh, M. Subcooled Boiling of Aqueous Solution of Alcohol. In Proceedings of the 6th KSME-JSME Joint
Conference on Thermal and Fluid Engineering Conference, Jeju City, Republic of Korea, 20–23 March 2005; pp. 21–23.
32. Savino, R.; Cecere, A.; Di Paola, R. Surface tension-driven flow in wickless heat pipes with self-rewetting fluids. Int. J. Heat Fluid
Flow 2009, 30, 380–388. [CrossRef]
33. Savino, R.; Cecere, A.; Van Vaerenbergh, S.; Abe, Y.; Pizzirusso, G.; Tzevelecos, W.; Mojahed, M.; Galand, Q. Some experimental
progresses in the study of self-rewetting fluids for the SELENE experiment to be carried in the Thermal Platform 1 hardware.
Acta Astronaut. 2013, 89, 179–188. [CrossRef]
34. Hu, Y.; Liu, T.; Li, X.; Wang, S. Heat transfer enhancement of micro oscillating heat pipes with self-rewetting fluid. Int. J. Heat
Mass Transf. 2014, 70, 496–503. [CrossRef]
35. Semenov, A.; Peschenyuk, Y.A.; Vozhakov, I. Application of Aqueous Solutions of Surfactants in Pulsating Heat Pipe. J. Eng.
Thermophys. 2021, 30, 58–63. [CrossRef]
36. Sobac, B.; Brutin, D. Thermal effects of the substrate on water droplet evaporation. Phys. Rev. E 2012, 86, 021602. [CrossRef]
[PubMed]
37. Kutateladze, S. Fundamentals of Heat Transfer Theory; Atomizdat: Moscow, Russia, 1979; p. 416.
38. Misyura, S.Y. Dependence of wettability of microtextured wall on the heat and mass transfer: Simple estimates for convection
and heat transfer. Int. J. Mech. Sci. 2020, 170, 105353. [CrossRef]
39. Misyura, S.; Morozov, V.; Egorov, R. Water evaporation on structured surfaces with different wettability. Int. J. Heat Mass Transf.
2022, 192, 122843. [CrossRef]
40. Volkov, R.; Strizhak, P.; Misyura, S.; Lezhnin, S.; Morozov, V. The influence of key factors on the heat and mass transfer of a sessile
droplet. Exp. Therm. Fluid Sci. 2018, 99, 59–70. [CrossRef]
41. Carrier, O.; Shahidzadeh-Bonn, N.; Zargar, R.; Aytouna, M.; Habibi, M.; Eggers, J.; Bonn, D. Evaporation of water: Evaporation
rate and collective effects. J. Fluid Mech. 2016, 798, 774–786. [CrossRef]
42. Dunn, G.; Wilson, S.; Duffy, B.; David, S.; Sefiane, K. The strong influence of substrate conductivity on droplet evaporation.
J. Fluid Mech. 2009, 623, 329–351. [CrossRef]
43. Doganci, M.D.; Sesli, B.U.; Erbil, H.Y. Diffusion-controlled evaporation of sodium dodecyl sulfate solution drops placed on a
hydrophobic substrate. J. Colloid Interface Sci. 2011, 362, 524–531. [CrossRef]
44. Zhang, J.; Wang, B. Study on the interfacial evaporation of aqueous solution of SDS surfactant self-assembly monolayer. Int. J.
Heat Mass Transf. 2003, 46, 5059–5064. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

111
energies
Article
Effect of Monodisperse Coal Particles on the Maximum Drop
Spreading after Impact on a Solid Wall
Alexander Ashikhmin, Nikita Khomutov, Roman Volkov, Maxim Piskunov * and Pavel Strizhak

Heat Mass Transfer Laboratory, School of Energy & Power Engineering,


National Research Tomsk Polytechnic University, 30 Lenin Ave., 634050 Tomsk, Russia;
[email protected] (A.A.); [email protected] (N.K.); [email protected] (R.V.); [email protected] (P.S.)
* Correspondence: [email protected]

Abstract: The effect of coal hydrophilic particles in water-glycerol drops on the maximum diameter
of spreading along a hydrophobic solid surface is experimentally studied by analyzing the velocity
of internal flows by Particle Image Velocimetry (PIV). The grinding fineness of coal particles was
45–80 μm and 120–140 μm. Their concentration was 0.06 wt.% and 1 wt.%. The impact of particle-
laden drops on a solid surface occurred at Weber numbers (We) from 30 to 120. It revealed the
interrelated influence of We and the concentration of coal particles on changes in the maximum
absolute velocity of internal flows in a drop within the kinetic and spreading phases of the drop-wall
impact. It is explored the behavior of internal convective flows in the longitudinal section of a drop
parallel to the plane of the solid wall. The kinetic energy of the translational motion of coal particles
in a spreading drop compensates for the energy expended by the drop on sliding friction along the
wall. At We = 120, the inertia-driven spreading of the particle-laden drop is mainly determined
by the dynamics of the deformable Taylor rim. An increase in We contributes to more noticeable
differences in the convection velocities in spreading drops. When the drop spreading diameter rises
at the maximum velocity of internal flows, a growth of the maximum spreading diameter occurs. The
presence of coal particles causes a general tendency to reduce drop spreading.

Keywords: coal particle; drop impact; maximum spreading; PIV; slurry; velocity field
Citation: Ashikhmin, A.; Khomutov,
N.; Volkov, R.; Piskunov, M.; Strizhak,
P. Effect of Monodisperse Coal
Particles on the Maximum Drop 1. Introduction
Spreading after Impact on a Solid In many technologies, e.g., spraying of composite liquid fuels (CLF) into combustion
Wall. Energies 2023, 16, 5291. https:// chambers [1–3], 3D printing, including bioprinting [4,5] and printed electronics [6], spray-
doi.org/10.3390/en16145291
ing of liquid friction modifiers [7], the particle-laden drops interact with various media.
Academic Editors: Vladimir The development of these technological processes requires knowledge of the particle distri-
Serdyukov and Fedor Ronshin butions in drops spreading after impact on the wall and of the effect of these particles on
the flow. For example, for CLF drops, results on the effect of carbon-containing particles
Received: 24 June 2023
on drop spreading and splashing are important since the analysis of such results will help
Revised: 5 July 2023
optimize the secondary atomization when hitting the wall [8], when 3D printing composite
Accepted: 6 July 2023
materials, tasks related to the ordering of particles on the target (functional surface) are
Published: 10 July 2023
critical [9].
In the case of particle-laden drop-wall collisions, the particle size, volume fraction,
and wettability of the particles and the surface are considered as influencing factors. In this
Copyright: © 2023 by the authors. regard, the particle-laden drops can be in the form of liquid marbles [10] and slurries [11],
Licensee MDPI, Basel, Switzerland. and surfaces, respectively, can be hydrophobic [12] and hydrophilic [13]. This study focuses
This article is an open access article on the particle-laden drops interacting with a hydrophobic surface.
distributed under the terms and There is quite a wide range of works devoted to suppressing the regimes known
conditions of the Creative Commons from pure drop impact, such as jetting [14], rebound [15,16], and drop break-up during
Attribution (CC BY) license (https:// rebound [17]. It is also known that such regimes as prompt and corona splashing, which
creativecommons.org/licenses/by/ are typical for drops without particles [18–21], are significantly modified [14] since the
4.0/).

Energies 2023, 16, 5291. https://doi.org/10.3390/en16145291 112 https://www.mdpi.com/journal/energies


Energies 2023, 16, 5291

destruction of a thin liquid layer can occur far from the drop contact line, being more
similar to splashing of viscous liquids [22].
Particles in drops also affect the initial phase of interaction—spreading [11], decreasing
the maximum spreading diameter Dmax [14,15]. However, this result is rather ambiguous
since there is a contradiction in which an increase in Dmax occurs with a growth of the
volume fraction of particles, but at the same time, an increase in the effective viscosity
should lead to less spreading [11,23]. Obviously, in the case of particle-laden drops, spread-
ing cannot be controlled by viscosity alone; it is a complex of influencing factors and
related effects listed above. The examination of the spreading drop morphology and the
distribution of internal particles results in the exploration of the spreading and receding
phases [11]. In particular, Grishaev et al. demonstrated that the particle patterns strongly
depend on the surface wettability, particle size, and initial impact velocity, as well as the
Reynolds (Re) and We numbers [11]. These parameters also determine the spreading and
receding phases.
Nicolas explored the patterns and distribution of the particles in the case of dilute
slurries [23]. He observed a ring and disk-like distribution of particles as a function of Re
and particle diameter (dp ). The manifestation of ring distribution was associated with the
movement of the liquid toward the center when the drop receded. In general, the effect of
particles on spreading was mainly determined by the ratio between inertial and viscous
forces (Re). With a significant predominance of inertia (Re > 5000–6000), particle-laden
drops spread more significantly compared to the drops without particles. However, the
shape of the drop contact line was significantly distorted relative to the shape of the circle.
Since there is a rather limited amount of research on the morphology of the particle-
laden drop impacting on a surface and the effect of particles on the drop spreading control,
a number of issues indicated by Grishaev et al. [11] remain relevant and motivate further
research. These include issues related to the influence of surface wettability, particle size,
and initial drop velocities on particle distribution and drop spreading. We believe that
all these issues for dilute slurries should be considered comprehensively (with empirical
expressions derived), with due regard for the effective viscosity. The latter depends to
a greater extent on the viscosity of the carrier medium, the buoyancy of monodisperse
particles, their volume fraction, and the material. Thus, the purpose of the study is to
experimentally determine the conditions for reducing or increasing the maximum spreading
diameter of water-glycerol drops laden with hydrophilic coal particles colliding with the
hydrophobic surface by analyzing the velocity fields using Particle Image Velocimetry; in
addition, the research will consider the morphology of the particle-laden drop impacting
on a surface—contact line deformation.

2. Materials
The solid component of the suspension was coking coal from the Berezovskaya mine,
Kemerovo region, Russia. Table 1 shows the results of its elemental and technical analysis.
The data was measured by the Vario microcube Elementar device. The carrier medium
of a slurry was a mixture of distilled water and glycerol (Merck, Germany, CAS number:
56–81-5). The use of glycerol is caused by the need to artificially increase the viscosity
of the slurry, bringing the conditions of particle movement in the flow closer to highly
concentrated slurries typical of the energy sector [24,25].

Table 1. Elemental composition and main characteristics of the coking coal.

Main Characteristics Elemental Composition

Ash Carbon Hydrogen Nitrogen Oxygen


Specific Sulfur
content of content content content content
Fuel Content of heat of content in
the fuel on calculated calculated calculated calculated
Substance moisture volatiles combus- dry matter
on a dry on a dry on a dry on a dry
Wa , % a dry basis V daf , % tion Qa s,V ,
ash-free (daf) ash-free (daf) ash-free (daf) St d , % ash-free (daf)
Ad , % MJ/kg
mass Cdaf , % mass Hdaf , % mass Ndaf , % mass Odaf , %
Coking coal 2.05 14.65 27.03 29.76 79.79 4.486 1.84 0.868 13.016

113
Energies 2023, 16, 5291

The solids were kept in an air bath at 378.15 K for two hours before mixing. The dried
solids were crushed in a Pulverisette 14 high-speed rotary mill with a grinding degree of
0.08–6 mm and a rotor speed of 6000–20,000 rpm. To obtain the required grinding degree,
the crushed solids were sieved in an ANALYSETTE 3 SPARTAN vibrating screen with a
potential sieving time of 3–20 min and using the sieves with a mesh width of 20–90 μm. The
solids were weighed using Vibra AF 225DRCE analytical scales with a resolution of 10−6 g.
The slurries were prepared in several stages. 12 g of glycerol was added to the
carrier medium—water weighing 8 g. Mixing of the aqueous solution with additives was
carried out using an AIBOTE ZNCLBS-2500 magnetic stirrer with a stirring temperature
of 298.15 ± 2 K, a rotation speed of 1500 rpm, and a mixing time of 7 min. Solid particles
were added gradually to the mixed volume to prevent the formation of agglomerates and
obtain a more uniform structure. Despite the presence of solid particles and varying their
concentration, the properties of liquids within different compositions were assumed to
be the same and corresponded to the water-glycerol solution under examination. This is
due to the relatively low proportion of solids in the slurry. Table 2 lists the properties of
the water-glycerol solution. Grinding fineness and mass concentration of coal particles
varied. Two groups of coal particles by grinding fineness with a noticeable difference in
size (Table 3) were chosen to reveal the additional effect of particle size on the maximum
spreading diameter and the velocities in a drop. In addition, the experiments were carried
out at different drop discharge heights relative to the impact surface (h) and, consequently,
at different drop velocities before impact (U0 ) and Weber numbers (We = ρD0 U02 σ−1 , where
D0 is the diameter of the drop before impact on a wall, m; ρ is the density of the liquid,
kg/m3 ; σ is the surface tension of the liquid, N/m). Different Weber numbers (Table 3)
separated conditionally by the four groups (We = 30; 60; 90; 120) were examined to ensure
proper confidence in the potential relationships while analyzing results. The minimal one
(We = 30) is defined by the experimental set-up capability, while the highest threshold
(We = 120) is due to complexing the adjustment of the used optical and laser system (see
Sections 3 and 4) in the case of increasing the impact droplet velocity U0 . Table 3 introduces
the initial conditions for conducting experiments.

Table 2. Properties of the water-glycerol solution.

Temperature Density, ρ Dynamic Viscosity, μ Surface Tension, σ


K kg/m3 Pa·s N/m
293.15 1154 10.8 × 10−3 0.06058

Table 3. Initial conditions for conducting experiments.

Coal grinding fineness Sample name Coal grinding fineness Sample name Number of experiments
Discharge height—2 cm, We ≈ 30
45–80 μm - 120–140 μm - pcs.
0.06 Slurry 1 0.06 Slurry 3 6
Particle concentration, wt.%
1 Slurry 2 1 Slurry 4 6
Coal grinding fineness Sample name Coal grinding fineness Sample name Number of experiments
Discharge height—7 cm, We ≈ 60
45–80 μm - 120–140 μm - pcs.
0.06 Slurry 1 0.06 Slurry 3 6
Particle concentration, wt.%
1 Slurry 2 1 Slurry 4 6
Discharge height—15 cm, Coal grinding fineness Sample name Coal grinding fineness Sample name Number of experiments
We ≈ 90
45–80 μm - 120–140 μm - pcs.
0.06 Slurry 1 0.06 Slurry 3 6
Particle concentration, wt.%
1 Slurry 2 1 Slurry 4 6
Discharge height—22 cm, Coal grinding fineness Sample name Coal grinding fineness Sample name Number of experiments
We ≈ 120
45–80 μm - 120–140 μm - pcs.

114
Energies 2023, 16, 5291

Table 3. Cont.

0.06 Slurry 1 0.06 Slurry 3 6


Particle concentration, wt.%
1 Slurry 2 1 Slurry 4 6
Total 48

3. Experimental Set-Up
During the research, an experimental set-up was used, the scheme of which is demon-
strated in Figure 1a. The PIV technique determined the flow velocity inside the drop during
isothermal contact with the surface. The method estimates the velocity of convective flows
in the longitudinal section of a spreading (after impact on a wall) drop parallel to the
plane of the solid wall at a distance of 0.2–0.3 mm from it. The PIV method is based on
recording the movement of special particles (tracers) over a very short period of time.
For this purpose, “tracer” microparticles are introduced into the measured medium. The
particles are illuminated by a powerful laser source, and their images are recorded by a
video camera. In this case (Figure 1b), the flat laser knife generated by the radiation source
cuts the drop in a plane parallel to the solid wall, and the optical axis of the video camera
is perpendicular to the plane of the laser knife (Figure 1a). This method is applied due to
the peculiarities of the absorption and scattering of laser radiation in the internal volume
of the drop, related to the presence of solid carbon-containing particles and their mass
concentration in a slurry.
The experimental set-up (Figure 1a) included the following aggregated positions:
measurement (registration) part, linear displacement module, drop generation system,
lighting system, high-speed video camera, and a computer for data collection and subse-
quent processing (not shown in Figure 1a). A platform of 0.1 m in diameter and 0.01 m
in thickness with a through hole of 0.02 m in diameter was manufactured for placing an
optically transparent substrate above the hole. The substrate has a diameter of 0.025 m
and a thickness of 0.005 m and is made of sapphire glass (Thorlabs). The platform and the
substrate with a particle-laden drop conventionally represent the measuring part.

(a)
Figure 1. Cont.

115
Energies 2023, 16, 5291

(b)
Figure 1. (a) Scheme of the experimental set-up for the implementation of the PIV method: 1—linear
displacement module; 2—test tube with a liquid; 3—peristaltic pump; 4—hollow needle-nozzle;
5—silicone feed tube; 6—liquid drop; 7—optically transparent substrate; 8—a platform with a through
hole for placing the substrate. 9—high-speed CMOS video camera; 10—continuous DPSS laser; 11—
collimator; 12—laser power supply; 13—lens; 14—laser beam (knife); 15—micro-positioning system
for the platform; (b) frame with the image of tracer particles in the drop under study, the planes of
the laser radiation and video recording, as well as the masked and analyzed areas.

The drop discharge height varies due to the linear displacement module with an
integrated servo drive and a power supply. This module is used for linear vertical (relative
to the substrate surface) movement of the drop generation system, which consists of a feed
tube and a hollow needle. The temperature of the liquid in the test tube corresponds to the
temperature in the laboratory and is 295.15–296.15 K. The liquid in the system is pumped
by a LongerPump BT100-1F dosing peristaltic pump with a set dosage volume of 0.01 mL.
The drop velocity U0 to the moment of impact on the surface could change as the drop
generation system moved vertically and was 0.63 m/s, 1.17 m/s, 1.72 m/s, and 2.08 m/s.
The systematic error in velocity measurement is 0.1 m/s. The diameter of the generated
drop D0 to the moment of impact on the surface remained constant and was 2.9 ± 0.05 mm.
The latter was defined as the arithmetic mean of the drop diameters measured vertically
and horizontally (in a frame) before the drop came into contact with the wall.
The set-up (Figure 1a) was equipped with a high-speed Phantom Miro M310 video
camera recording the drop spreading and the speed of tracer particles with a sample rate
of 10,000 fps and a resolution of 512 × 512 pixels. When using PIV (Figure 1a), a Nikon
200 mm f /4 AF-D Macro lens with a focal length of 200 mm and a relative aperture of
f /4 was engaged. A light filter was installed on the lens, which is a laboratory orange opti-
cal glass with a bandwidth of more than 590 nm. A Thorlabs PT1B/M linear manipulator
with a maximum shift of 25 mm and an accuracy of ±5 μm representing a micro-positioning
system allowed the adjustment of the optics relative to the platform in the vertical plane.

116
Energies 2023, 16, 5291

The minimum and maximum dimensions of the registration area were 4 × 4 mm and 25
× 25 mm, respectively. A continuous diode-pumped solid-state (DPSS) laser KLM-532A
(radiation wavelength—532 nm, maximum power—5 W, power stability—3%) was em-
ployed to illuminate particle-laden drops. When implementing PIV, the generated laser
radiation was transformed into a flat laser knife with an opening angle of 12◦ by means of
the collimator based on a set of spherical and cylindrical lenses. The width and thickness
of the laser knife in the measuring area were 60 mm and 0.2 mm, respectively. The plane
of the laser knife was set parallel to the plane of the substrate surface. The laser knife cut
the drop in the longitudinal section at a minimum distance (about 200–300 μm) from the
substrate surface (Figure 1a).
Polyamide fluorescent microparticles with a diameter of 5 μm acted as tracers nec-
essary for recording the velocity of internal convective flows in a drop. The particles
absorbed laser radiation at a wavelength of 532 nm (close to the maximum of the absorp-
tion spectrum) and emitted (re-emitted) light at a wavelength of more than 550 nm. The
microparticles were added to the drop at the stage of slurry preparation (see Section 2).
The concentration of fluorescent microparticles was 1 g/L in accordance with the results
and recommendations in Ref. [26]. The tracers were introduced into the slurry sample
together with coal particles. The utilization of fluorescent microparticles, together with
the optical light filter, contributed to reaching the following positive aspects. First, it made
it possible to filter out reflected and refracted (by drop and substrate) laser radiation at a
wavelength of 532 nm. It was possible to exclude the glare on the drop image. Second, it
made it possible to filter out the light reflected from the coal particles inside the slurry drop
and, as a result, to visualize only the tracers inside the drop, i.e., to determine exactly a
liquid velocity.
For each liquid sample, at least three experiments were performed under identical
initial conditions: drop size and velocity before the drop-wall collision. A calibration prism
(Edmund Optics) with a minimum division value of 10 μm helped to determine the depth
of field of the lens and the scale factor. The depth of field in the case of the PIV method
was about 2 mm, and the minimum value of the scale factor for the considered cases was
0.05 mm/pixel.

4. Method of Drop Spreading Research


PIV allowed the recording of instantaneous velocity fields in a longitudinal section
of the drop by fluorescent microparticle motion for an inter-frame delay. Thus, we were
able to study the effect of coal particle concentration on the velocity of internal convective
flows in a spreading drop. High-power laser radiation used as illumination and high-speed
video recording with an exposure time of 4 μs made it possible to observe the movement of
microparticles in the drop with great detail, as well as to record the microparticle velocity
in spreading and receding phases. The Actual Flow v1.18 software enabled to process
of experimental data by a cross-correlation algorithm when constructing instantaneous
velocity fields of microparticles in a drop. The data processing included several consecutive
stages (Figure 2):
- the frames recorded by the high-speed video camera were imported into Actual Flow;
- the average background intensity of the image was determined for the area without a
drop; the obtained intensity was subtracted from each frame, i.e., the intensity of each
pixel of the image was reduced by this value of the intensity;
- when the laser light beam is strongly distorted by a drop or when laser light is strongly
absorbed in a slurry drop, a geometric mask was applied to the original frame to avoid
absorption and refraction of the laser light so that only the half of the drop that the
light beam enters is analyzed (Figure 1);
- each frame was divided into elementary regions of 32 × 32 pixels in size.
- for each elementary region, the correlation function was calculated, after which the
coordinates of its maximum were estimated;

117
Energies 2023, 16, 5291

- the shift of the coordinates of the maximum of the correlation function in each ele-
mentary region for the time between each of two consecutive frames was determined
with an accuracy of 0.2 pixels; the displacement of the coordinates of the maximum
corresponds to the most probable movement of particles within the elementary region;
- using the scale factor and the time delay between two consecutive frames, the velocity
of fluorescent microparticles was calculated, and plotting the corresponding velocity
vector for each elementary region took place;
- using a set of velocity vectors, an instantaneous two-dimensional two-component
field of the velocity of fluorescent microparticles was reconstructed at each frame;
- an interpolation procedure was performed for the obtained velocity fields, during
which the modulus and direction of each velocity vector were compared with the
corresponding modulus and direction of neighboring vectors, as well as vectors
located in the same elementary region in the previous and subsequent velocity fields;
If the differences in the modulus and direction of the velocity vector were more than
20%, the correction of these parameters for this vector occurred.

Figure 2. Illustration of the image processing and construction of convective flow velocity fields:
(I)—primary image; (II)—image after subtracting the background intensity; (III)—instantaneous
velocity field; (IV)—instantaneous velocity field after the vector interpolation procedure.

The following limitations and disadvantages of PIV for measuring the velocity of
internal convective flows in a spreading drop were typical of preliminary experiments:
• The curvature of the drop surface causes the angle between the plane of the laser knife
incident on it and the drop surface to be different from 90◦ [27]. At the same time,
this angle also changes up during the drop spreading. This leads to the refraction
of the laser knife inside the drop and, as a result, its deviation from the direction
parallel to the substrate surface. In other words, the knife begins to hit the surface
of the substrate and reflect off it. This fact may introduce an additional error in the
measurement results;
• Due to the sphericity of the drop (curvature of its surface), the latter works as a
collecting lens, focusing the laser knife falling on it [27]. This leads to the appearance
of two non-laser-illuminated (shaded) “dead zones”, where the convection velocity
cannot be detected (Figure 1).
• When a drop spreads radially after it collides with a solid wall, waves (“horns”) are
often formed along its surface, the crests of which also focus the laser beam as local
collecting lenses. This results in the appearance of alternating laser-illuminated and
non-illuminated regions and makes the registration of convection velocities inside the
drop also impossible;
• If the content of coal particles is higher than 1–2 wt.%, the laser knife does not penetrate
into the drop to a depth of more than 0.2–0.5 mm due to the absorption and reflection
of light by these particles. This makes it impossible to record the convection velocities
inside the drop.
Thus, PIV allows measuring the convection velocity fields in a drop at any stage of
its spreading over the substrate under conditions of insignificant curvature of the drop
surface, even when it is heated, but only at small mass concentrations of coal particles, less

118
Energies 2023, 16, 5291

than 1 wt.%. If at least one of the above disadvantages occurred, the video frame and its
corresponding velocity field were not considered.
The PIV-derived results enabled analyzing the trends of maximum absolute convection
velocities of internal flows in the spreading drop (Umax ), as well as the average velocities
of these flows (Umean ) from the contact of the surface by a drop to its maximum spread-
ing, i.e., until the drop reaches the maximum spreading diameter Dmax . Accordingly, the
primary data are presented in the form of instantaneous maximum and average velocities
of fluorescent microparticles in a drop impacting the wall (U). Instantaneous maximum
velocities (Umax ) were determined based on averaging 20 maximum velocity vectors in the
frame at a given impact time (t). Instantaneous average velocities (Umean ) were calculated
as the arithmetic mean of all velocity vectors contained in the velocity field at a specific
time moment. The values of Umax and Umean were analyzed by processing the interpolated
instantaneous velocity fields of fluorescent microparticles. The data of velocity distribu-
tions from Actual Flow were additionally processed through the Wolfram Mathematica
customized algorithm of a moving average filter. The results were analyzed by introducing
a parameter Umax /U0 , representing Umax scaled by the initial drop velocity before the
impact U0 , to describe the behavior of the maximum absolute velocities of internal flows in
the spreading drop at various We.
Since the drop-transparent wall impact was recorded from below by means of high-
speed photography, it became possible to simultaneously measure the hydrodynamic
characteristics of the drop spreading, in particular, Dmax and the time to reach Dmax —
tmax . The value of Dmax was the arithmetic mean of the horizontal Dmax hor and vertical
Dmax vert diameters since the drop spreading occurs radially and conventionally evenly
from the center. To determine Dmax hor and Dmax vert , the shadow photography method
captures the collision process (Figure 3). In order to derive the empirical expressions on
the spreading process, the factor of maximum spreading βmax = Dmax /D0 widely used in
typical studies [18,28–30] was under examination.

Figure 3. Illustration of the approach to analyzing the maximum drop spreading diameter Dmax in
shadow photography of the collision process.

By combining the PIV-derived data on Umax and the results of the drop spreading
diameters recorded by shadow photography, it became possible to determine the drop
diameter at which the maximum absolute velocity of internal flows DUmax is reached. In
addition, a parameter DUmax /D0 was introduced for quantitative and qualitative interpre-
tation of the effect of the maximum absolute velocity of internal flows on βmax .

5. Results and Discussion


5.1. Morphological Observations
The results of measuring the velocity of internal fluid flow during a drop-wall collision
are analyzed within four generally accepted phases (Figure 4a), such as kinetic, spreading,
receding, and relaxation phases [31]. The kinetic one is characterized by a sharp increase
in the maximum and average velocities of internal flows in the longitudinal section of the
drop and the achievement of peak velocities Umax and Umean . Thus, in the drop spreading

119
Energies 2023, 16, 5291

phase, at the initial stage, the values of flow velocities correspond to peak values, after
which they steadily tend to values close to zero. At the same time, both phases show
an almost linear character of the growth and fall of Umax and Umean values for all the
considered initial experimental conditions (size and velocity of droplets, liquid viscosity).
In most cases, the duration of the kinetic phase is about two times shorter relative to the
spreading phase. The end of the spreading phase occurs when the drops reach Dmax .
At this point, mostly due to the viscous dissipation, the kinetic energy in the system is
exhausted. The evolution of velocity fields in the longitudinal section of a spreading drop
before kinetic energy depletion is presented in inserts above each velocity distribution over
time (Figure 4). The inserts allow us to characterize the velocity field of internal flows
as significantly inhomogeneous with a shift of increased velocities in the radial direction
towards the drop spreading at the end kinetic phase and the beginning of the spreading one.
During the spreading, the flow velocity decreases (see all inserts # 4 in Figure 4), starting
from the periphery (rim), in the opposite direction towards the point of initial contact of
the drop with the wall.
The subsequent receding phase is characterized by inertial processes, mainly in the
drop rim, which leads to a local, less significant increase in the velocity of internal flows.
This is clearly seen in Figure 4 from the values of Umax at t = 0.0075–0.0085 s. The presence
of coal particles in a drop noticeably weakens the flow velocities during receding due
to the effect of inhibiting the outflow of liquid by solids. This is clearly observed in the
time distributions of flow velocities when comparing the ends of the receding phase in
Figure 4a–e. The values of Umax for the compared cases are significantly different. The
relaxation phase (rather, its initial stage) begins immediately after the local extremum (Umax )
at the end of receding and is accompanied by a monotonous attenuation of internal flows.
The addition of coal particles generally affects the velocity of internal convective flows
in a drop in a variety of ways. In addition, the experimental results made it possible to
identify a number of influencing factors, namely, the grinding fineness of coal particles,
their concentration, and We.

Figure 4. Cont.

120
Energies 2023, 16, 5291

Figure 4. Instantaneous maximum and average velocities of fluorescent microparticles in a spreading


drop at We ≈ 30 as a function of the spreading time, as well as the insertion of velocity fields in
the longitudinal section of the spreading drop at the following time points: 1—0.001 s; 2—0.0015 s;
3—0.0024 s; 4—0.004 s for Slurry 1 (a), Slurry 2 (b), Slurry 3 (c), Slurry 4 (d), water-glycerol solution
(e). The confidence intervals for the experimental data were no more than 3.3%.

5.2. Effect of We and the Concentration of Coal Particles on the Velocity of Internal Flows in a
Spreading Drop
The analysis of the results on the distributions of maximum velocities in the drop over
the spreading time (Figure 5) revealed the interrelated effect of We and the concentration of
coal particles on changes in the peak values of Umax in the kinetic and spreading phases. At
We = 120, it is difficult to distinguish the effect of coal particles on the development of inter-
nal flows (Figure 5b,d), regardless of their grinding fineness and concentration. The radial

121
Energies 2023, 16, 5291

motion of the drops occurs with the spreading velocity Uspr (Figure 5b,d) approximately
twice as high as at We = 30 (Figure 5a,c). At such high values of Uspr (about 3.5 m/s), the
inertia-driven spreading of the particle-laden drop is mainly determined by the dynamics
of the Taylor rim, whose diameter becomes relatively smaller. Then, the rim begins to
deform (Figure 6a,b) due to the Rayleigh-Taylor instability [32–34]. The contribution of
solids is insignificant, so the values of Umax are quite close. While with We = 30 and a
grinding fineness of 45–80 μm (Figure 5a), the addition of coal particles contributes to a
rather significant increase in Umax . However, at a coal particle concentration of 0.06 wt.%,
the values of Umax were significantly higher than at 1 wt.%. The relative decrease in Umax
in the case of a higher concentration can be physically associated with the formation of
the internal structure (Figure 6c) and, accordingly, an increase in shear stresses between
the liquid layers during the drop spreading. The latter leads to the expenditure of more
energy to move the liquid. At a lower concentration, solids, having more physical space
between them, cannot form the structure, and therefore the particles can act as single
“accelerators” in the velocity field, which get the inertia-driven acceleration from the inter-
nal translational flow of the liquid (Figure 6d). When We = 30, and the particle grinding
fineness is 120–140 μm (Figure 5c), then even in the case of a lower concentration, a certain
effect of inhibition of internal flows is observed both in absolute values of Umax and in the
spreading time t relative to the case without coal particles. This phenomenon is presumably
caused by the immediate sedimentation of coal particles upon contact with the surface
and their restraining disturbance of the laminar flow (Figure 6e). At 1 wt.%, the number
of the particles of 120–140 μm in size becomes larger; they can not only restrain laminar
flow but also mechanically deform the liquid-gas interface both on the free surface of the
liquid and near the contact line after particle collisions (Figure 6f). The latter can lead to
additional local liquid flows that affect the overall distribution of U over the spreading time
(Figure 5c). A remarkable thing happened for the drops of Slurry 4, which characterizes
the difficult-to-predict nature of particle motion. In particular, coal particles were often
grouped in a rather limited area of the radially moving flow. In this case, most of the
spreading particle-laden drop contained almost no coal particles.
The effect of coal particles and U0 on Umax is demonstrated in the simplest format
in Figure 7a. First of all, it can be seen that the values of U0 and Umax are quite close,
i.e., as the values of U0 increase, Umax grows in direct proportion, and the momentum
conservation law in a drop is satisfied. However, with more detailed observation, it is
noticeable that in the absence of coal particles, the drop consumes energy by the sliding
friction force when spreading along a solid surface for all We considered. In Figure 7a, this
moment is expressed in the values of Umax for water-glycerol drops, which are lowered
relative to this characteristic for particle-laden drops and the values of U0 . All other
things being equal, coal particles (due to their mass) allow the development of a relatively
high maximum absolute velocity of internal flows, which almost does not differ from the
values of U0 . Essentially, this means that the kinetic energy (mainly) of the translational
motion of coal particles in the drop compensates for the energy spent by the drop on
sliding friction along the wall. Linear functions that describe the behavior of Umax with a
change in U0 have the following form: Umax = 1.02U0 —0.11 for the particle-laden drops
and Umax = 0.91U0 —0.11 for the water-glycerol drops. The coefficient of determination
R2 is 0.94 and 0.99, respectively. Figure 7b, when introducing the parameter Umax /U0 ,
clearly illustrates in which cases the values of Umax are closest to the values of U0 . In
addition, Figure 7b also assumes that, based on the location of the experimental points, a
further increase in We will not result in a directly proportional increase in the maximum
absolute velocity of internal flows. Most likely, Umax will remain constant, about 2 m/s, for
the considered sizes and concentrations of coal particles in the particle-laden drops. The
behavior of the parameter Umax /U0 depending on We for water-glycerol and particle-laden
drops is well described by third-order polynomial functions, Equations (1) and Equation (2),
respectively. The value of R2 for water-glycerol drops is 0.97 for the particle-laden drops—
R2 = 0.77. The value of R2 for the particle-laden drops can be considered satisfactory since

122
Energies 2023, 16, 5291

all the slurries under study are taken into account, i.e., with different concentrations and
grinding fineness of coal particles.

Umax
= 2.54 × We3 − 7.2 × We2 + 0.06 × We − 0.77 (1)
U0

Umax
= 9.3 × We3 − 2.78 × We2 + 0.03 × We − 0.03 (2)
U0

0.8 3.5 3.5


0.06 wt.% 1 wt.% 1.8
1 wt.% 0.06 wt.% 1 wt.%
1.6 3.0 3.0
No particles No particles 0.06 wt.%
0.6 1.4 No particles
2.5 0.06 wt.% 2.5
1.2 1 wt.%
No particles
Umax, m/s

Uspr, m/s

Umax, m/s
2.0 2.0

Uspr, m/s
0.4 1.0
0.8 1.5 1.5

0.6
0.2 1.0 1.0
0.4
0.5 0.5
0.2
0.0
0.0 0.0 0.0
0.000 0.002 0.004 0.006 0.008 0.010 0.000 0.002 0.004 0.006 0.008 0.010

t, s t, s

(a) (b)
0.6 2.0 3.5 3.5
0.06 wt.% 1 wt.% 1 wt.%
1 wt.% 0.06 wt.% 1.8 0.06 wt.%
No particles No particles 3.0 3.0
1.6 No particles
0.06 %
0.4 1.4 2.5 1% 2.5

1.2 No particles
Umax, m/s

Uspr, m/s
Umax, m/s

2.0 2.0
Uspr, m/s

1.0

0.8 1.5 1.5


0.2
0.6 1.0 1.0
0.4
0.5 0.5
0.2
0.0
0.0 0.0 0.0
0.000 0.002 0.004 0.006 0.008 0.010 0.000 0.002 0.004 0.006 0.008 0.010
t, s t, s

(c) (d)

Figure 5. Effect of the Weber number and the concentration of coal particles on the distributions of
the instantaneous maximum velocity of fluorescent microparticles in a drop over the spreading time:
(a)—grinding fineness of coal particles 45–80 μm, We = 30; (b)—45–80 μm, We = 120; (c)—120–140 μm,
We = 30; (d)—120–140 μm, We = 120. The confidence intervals for the experimental data were no
more than 3.3%.

123
Energies 2023, 16, 5291

(a) (b) (c)

(d) (e)

(f)

Figure 6. Frames of the forming rim during the Slurry 4 drop spreading with a particle concentration
of 1 wt.% and a grinding fineness of 120–140 μm at We = 120 and t = 1.3 ms (a), We = 30 and
t = 3 ms (b); frame illustrating the appearance of a Slurry 2 drop with a particle concentration of
1 wt.% and a grinding fineness of 45–80 μm at We = 120 and t = 0.6 ms (c); frames illustrating the
relative acceleration of a single coal particle in a drop of Slurry 1 (d); frames illustrating coal particles
deposited at the liquid–substrate interface during most of the Slurry 3 drop spreading (e); frame of
the mechanical deformation of the Slurry 4 drop rim due to perturbation by coal particles (f).

5.3. Effect of the Internal Flow Velocity on the Maximum Drop Spreading Diameter
5.3.1. Weber Number Factor
In the previous subsection, it was shown that coal particles in a spreading drop can
affect the internal flow velocities. If the velocity of internal flows in the drop increases, this
should affect the spreading characteristics, in particular, the spreading diameter. Therefore,
one of the key tasks of the study was to establish an implicit relationship between the
velocity of internal flows in the longitudinal section of the drop and the factor of its
maximum spreading βmax . To test this relationship, it was necessary to make sure that

124
Energies 2023, 16, 5291

for all the liquids under study, the behavior of βmax is mainly determined by the initial
drop velocity with the constancy of other terms within We. This confidence was achieved
after summarizing the results in the framework of the relationship βmax = f (We) depicted
in Figure 8a. The behavior of βmax for particle-laden and water-glycerol drops is governed
by the power function of βmax = 0.45We0.4 with R2 = 0.95.
2.5 1.4
45-80 μm 0.06 wt.% Acceleration relative to the initial droplet
45-80 μm 1 wt.%
1.2 velocity
2.0 120-140 μm 0.06 wt.%
120-140 μm 1 wt.%
1.0
Water-glycerol solution
Umax (m/s)

1.5

Umax / U0
0.8
Umax=ʝ0.12ÂU0
0.6 45-80 μm 0.06 wt.%
1.0
45-80 μm 1 wt.%
0.4 120-140 μm 0.06 wt.%
0.5 Linear fit for slurries
120-140 μm 1 wt.%
Umax=ʝ0.17ÂU0 Linear fit for solution 0.2 Water-glycerol solution
of glycerol Fit for solution of glycerol
0.0 Fit for slurries
0.0
0.5 1.0 1.5 2.0 40 60 80 100 120
U0 (m/s) We
(a) (b)

Figure 7. Maximum absolute velocities of internal flows in a drop Umax as a function of the initial
velocity of the drop before impact U0 (a), parameter Umax /U0 as a function of We (b).

3.0 3.0

βmax=0.45 We 0.4

2.5 2.5
ȕmax
βmax

2.0 Slurry 1 2.0


Slurry 2 We=30
Slurry 3 We=60
Slurry 4 We=90
Water-glycerol solution We=120
1.5 1.5
30 60 90 120 0.0000 0.0003 0.0006 0.0009
We tmax
(a) (b)
4
45-80 μm 0.06 wt.%
45-80 μm 1 wt.%
120-140 μm 0.06 wt.%
120-140 μm 1 wt.%
3 Water-glycerol solution
Overall fit
βmax

βmax = 40Â(DUmax/D0)0.35
1
10-5 10-4 10-3
DUmax / D0
(c)

Figure 8. Maximum spreading factor βmax and time to reach the maximum spreading diameter tmax
for the considered liquids at We = 30, 60, 90, and 120 (a); values of βmax as a function of We (b); values
of βmax as a function of the dimensionless diameter of the spreading drop at the maximum velocity
of internal flows DUmax /D0 (c).

125
Energies 2023, 16, 5291

Data analysis in Figure 8b suggests that the time to reach the maximum spreading
diameter tmax for the liquids under consideration begins to vary more strongly with in-
creasing We. At the same time, there are cases when tmax stays almost the same for different
We. Thus, the higher We, the more noticeable the differences in the internal flow velocities
in the spreading drops become. This is clearly demonstrated in Figure 7a. Another feature
suggests that an increase in DUmax , which depends on We, causes a growth of the maximum
spreading diameter (Figure 8c) that is qualitatively described by a power function according
to Equation (3). This result is not obvious since the later achievement of the maximum
absolute velocity of internal flows in a drop is naturally in no way connected with the
achievement of a higher maximum spreading diameter. Nevertheless, the established
feature allows quantitative (Equation (3), R2 = 0.78) and qualitative characterization of the
effect of the internal flow velocity in the drop on its maximum spreading diameter.
 0.35
DUmax
β max = 40 × (3)
D0

Despite the same initial conditions (i.e., drop velocity before impacting the wall) for
drops of all liquids within the same We, the experimental results in Sections 5.3.1 and 5.3.2
introduce a very ambiguous effect of coal particles on the velocity of internal convective
flows in the drop. Thus, the potential energy of particles in the particle-laden drops,
expressed for different slurries in the form of a change in the flow velocity when colliding
with a wall, should establish a certain pattern with respect to the movement of the contact
line at least until inertia-driven motion ceases (i.e., until the maximum spreading diameter
is reached). This will be discussed in more detail in the next subsection.

5.3.2. Factor of Coal Particles in a Drop


The predicted pattern on the effect of the internal flow velocities Umax and Umean on
Dmax (Dmax = βmax D0 ) was initially obtained from the point of view of the presence of coal
particles in the drop, considering their grinding fineness and all the studied concentrations.
In Figure 9a,b, the trend lines (Dmax = aUmax b and Dmax = aUmean b , respectively) make it
clear that Dmax for the particle-laden drops decrease relative to the water-glycerol drops at
the same velocities of internal flows and drops before impacting the wall.
9 9
45-80-μm particles 45-80-μm particles
120-140-μm particles 120-140-μm particles
Water-glycerol solution Water-glycerol solution
8 8

7 7
Dmax, mm

Dmax, mm

6 6

5 Best fit as Dmax=axb for slurries 5


Best fit as Dmax=axb for slurries
Best fit as Dmax=axb for water-glycerol Best fit as Dmax=axb for water-glycerol
solution solution
4 4
0.0 0.5 1.0 1.5 2.0 0.1 0.2 0.3 0.4 0.5 0.6
Umax, m/s Umean, m/s

(a) (b)

Figure 9. Effect of Umax (a) and Umean (b) on the maximum spreading diameter at different grinding
finenesses of coal particles.

The differences in Dmax tend to grow with increasing internal flow velocities Umax and
Umean . Thus, the presence of coal particles, in general, regardless of the considered particle

126
Energies 2023, 16, 5291

fineness and their concentration, causes a general tendency to decrease the intensity of
liquid drop spreading.

6. Conclusions
• The PIV-derived results of measuring the velocities of internal convective flows in
the longitudinal section of a water-glycerol solution drop laden with hydrophilic coal
particles and spreading over a hydrophobic surface allowed the exploration of the
conditions for reducing the maximum spreading diameter.
• At We = 30, the particle grinding fineness and their concentration strongly affect the
internal flow velocities, contributing both to their increase and decrease, depending
on the combination of the initial parameters of a slurry. At We = 120, the spreading
velocity of the particle-laden drops is approximately twice as high as at We = 30.
Given this fact, the inertia-driven spreading of the particle-laden drop is mainly
determined by the dynamics of the deformable Taylor rim, and the contribution of
solids is insignificant, causing the closeness of the values of the maximum absolute
velocity of internal flows for various combinations of the initial parameters of a slurry.
Relying on the experimental data obtained by the shadow photography and PIV, the
behavior of internal convective flows in the longitudinal section of a particle-laden
drop is characterized. It is revealed that the kinetic energy of the translational motion
of coal particles in a drop compensates for the energy expended by the drop on sliding
friction along the wall.
• The behavior of the maximum spreading factor βmax for particle-laden and water-
glycerol drops is mainly defined by the initial drop velocity with the constancy of
other terms within the Weber number and is governed by the power function of
βmax = 0.45We0.4 with a coefficient of determination R2 = 0.95. Further, it is revealed the
peculiarity of a noticeable increase in the differences in the velocities of internal flows
in spreading drops with an increase in Weber number. Finally, as the Weber number
grows, an increase in the spreading drop diameter at the maximum absolute velocity
of internal flows causes the elevated values of the maximum spreading diameter and
 0.35
is described by an expression of β max = 40 × DUmax D0 . In addition, the presence
of coal particles causes a general tendency to reduce the liquid drop spreading.

Author Contributions: Conceptualization, R.V., M.P. and P.S.; methodology, A.A., R.V. and M.P.;
investigation, A.A., N.K., R.V. and M.P.; data curation, A.A. and M.P.; formal analysis, A.A., R.V., M.P.
and P.S.; writing—original draft preparation, A.A., R.V. and M.P.; writing—review and editing, N.K.,
R.V. and P.S.; resources, R.V. and P.S.; project administration, M.P. and P.S. All authors have read and
agreed to the published version of the manuscript.
Funding: The study was supported by the grant of the Ministry of Science and Higher Education of
the Russian Federation, Agreement No 075-15-2020-806 (Contract No 13.1902.21.0014).
Data Availability Statement: The raw/processed data can be provided by the corresponding author
if required.
Conflicts of Interest: The authors declare no conflict of interest.

Nomenclature

Ad ash content of the fuel on a dry basis (%);


Cdaf carbon content calculated on a dry ash-free (daf) mass, (%);
D0 initial droplet diameter before impact (m);
Dmax maximum spreading diameter (m);
maximum diameter of the drop spreading measured horizontally
Dmax hor
in a frame (m);
Dmax vert maximum diameter of the drop spreading measured vertically in a frame (m);

127
Energies 2023, 16, 5291

drop diameter corresponding to the maximum absolute velocity of internal


DUmax
flows (m);
Hdaf hydrogen content calculated on a dry ash-free (daf) mass, (%);
Ndaf nitrogen content calculated on a dry ash-free (daf) mass, (%);
Odaf oxygen content calculated on a dry ash-free (daf) mass, (%);
Qa s,V specific heat of combustion (MJ/kg);
Re Reynolds number (–);
St d sulfur content in dry matter, (%);
t impact time (s);
tmax time at the maximum spreading diameter (s);
instantaneous maximum and average velocities of fluorescent particles in a
U
drop impacting the wall (m/s);
U0 initial drop velocity (m/s);
maximum absolute velocity of internal flows in the drop during
Umax
spreading (m/s);
average speed of internal flows for the period from the contact of the surface
Umean
with a drop and up to the maximum spreading (m/s);
Uspr drop spreading velocity (m/s);
V daf amount of volatiles (%);
Wa fuel moisture (%);
We Weber number (–).
Greek symbols
βmax maximum spreading factor (–);
ρ density (kg/m3 );
σ coefficient of surface tension (N/m);
μ dynamic viscosity (Pa·s).

References
1. Wu, X.; Guo, Q.; Gong, Y.; Cheng, C.; Ding, L.; Wang, F.; Yu, G. Visualization Study on Particle Flow Behaviors during Atomization
in an Impinging Entrained-Flow Gasifier. Chem. Eng. Sci. 2020, 225, 115834. [CrossRef]
2. Wu, X.; Gong, Y.; Guo, Q.; Xue, Z.; Yu, G. Experimental Study on the Atomization and Particle Evolution Characteristics in an
Impinging Entrained-Flow Gasifier. Chem. Eng. Sci. 2019, 207, 542–555. [CrossRef]
3. Kuznetsov, G.V.; Strizhak, P.A.; Valiullin, T.R.; Volkov, R.S. Atomization Behavior of Composite Liquid Fuels Based on Typical
Coal Processing Wastes. Fuel Process. Technol. 2022, 225, 107037. [CrossRef]
4. Zhang, Y.; Kumar, P.; Lv, S.; Xiong, D.; Zhao, H.; Cai, Z.; Zhao, X. Recent Advances in 3D Bioprinting of Vascularized Tissues.
Mater. Des. 2021, 199, 109398. [CrossRef]
5. Shen, E.M.; McCloskey, K.E. Affordable, High-Resolution Bioprinting with Embedded Concentration Gradients. Bioprinting 2021,
21, e00113. [CrossRef]
6. Brian, D.; Ahmadian-Yazdi, M.-R.; Barratt, C.; Eslamian, M. Impact Dynamics and Deposition of Perovskite Droplets on
PEDOT:PSS and TiO2 Coated Glass Substrates. Exp. Therm. Fluid Sci. 2019, 105, 181–190. [CrossRef]
7. Suda, Y.; Iwasa, T.; Komine, H.; Tomeoka, M.; Nakazawa, H.; Matsumoto, K.; Nakai, T.; Tanimoto, M.; Kishimoto, Y. Development
of Onboard Friction Control. Wear 2005, 258, 1109–1114. [CrossRef]
8. Thoraval, M.J.; Schubert, J.; Karpitschka, S.; Chanana, M.; Boyer, F.; Sandoval-Naval, E.; Dijksman, J.F.; Snoeijer, J.H.; Lohse, D.
Nanoscopic Interactions of Colloidal Particles Can Suppress Millimetre Drop Splashing. Soft Matter 2021, 17, 5116–5121. [CrossRef]
9. Visser, C.W.; Kamperman, T.; Karbaat, L.P.; Lohse, D.; Karperien, M. In-Air Microfluidics Enables Rapid Fabrication of Emulsions,
Suspensions, and 3D Modular (Bio)Materials. Sci. Adv. 2018, 4, eaao1175. [CrossRef] [PubMed]
10. Laborie, B.; Lachaussée, F.; Lorenceau, E.; Rouyer, F. How Coatings with Hydrophobic Particles May Change the Drying of Water
Droplets: Incompressible Surface versus Porous Media Effects. Soft Matter 2013, 9, 4822–4830. [CrossRef]
11. Grishaev, V.; Iorio, C.S.; Dubois, F.; Amirfazli, A. Impact of Particle-Laden Drops: Particle Distribution on the Substrate. J. Colloid
Interface Sci. 2017, 490, 108–118. [CrossRef] [PubMed]
12. Nguyen, T.V.; Ichiki, M. Bubble Entrapment during the Recoil of an Impacting Droplet. Microsyst. Nanoeng. 2020, 6, 36. [CrossRef]
[PubMed]
13. Almohammadi, H.; Amirfazli, A. Droplet Impact: Viscosity and Wettability Effects on Splashing. J. Colloid Interface Sci. 2019,
553, 22–30. [CrossRef] [PubMed]
14. Grishaev, V.; Iorio, C.S.; Dubois, F.; Amirfazli, A. Complex Drop Impact Morphology. Langmuir 2015, 31, 9833–9844. [CrossRef]
[PubMed]
15. Ok, H.; Park, H.; Carr, W.W.; Morris, J.F.; Zhu, J. Particle-Laden Drop Impacting on Solid Surfaces. J. Dispers. Sci. Technol. 2005,
25, 449–456. [CrossRef]

128
Energies 2023, 16, 5291

16. Ashikhmin, A.E.; Khomutov, N.A.; Piskunov, M.V.; Yanovsky, V.A. Secondary Atomization of a Biodiesel Micro-Emulsion Fuel
Droplet Colliding with a Heated Wall. Appl. Sci. 2020, 10, 685. [CrossRef]
17. Ueda, Y.; Yokoyama, S.; Nomura, M.; Tsujino, R.; Iguchi, M. Bouncing Behaviors of Suspension Liquid Drops on a Superhy-
drophobic Surface. J. Vis. 2010, 13, 281–283. [CrossRef]
18. Piskunov, M.; Semyonova, A.; Khomutov, N.; Ashikhmin, A.; Yanovsky, V. Effect of Rheology and Interfacial Tension on Spreading
of Emulsion Drops Impacting a Solid Surface. Phys. Fluids 2021, 33, 83309. [CrossRef]
19. Bertola, V. An Impact Regime Map for Water Drops Impacting on Heated Surfaces. Int. J. Heat Mass Transf. 2015, 85, 430–437.
[CrossRef]
20. Piskunov, M.; Khomutov, N.; Semyonova, A.; Ashikhmin, A.; Misyura, S. Unsteady Convective Flow of a Preheated Water-in-Oil
Emulsion Droplet Impinging on a Heated Wall. Phys. Fluids 2022, 34, 93311. [CrossRef]
21. Semyonova, A.; Khomutov, N.; Misyura, S.; Piskunov, M. Dynamic and Kinematic Characteristics of Unsteady Motion of a
Water-in-Oil Emulsion Droplet in Collision with a Solid Heated Wall under Conditions of Convective Heat Transfer. Int. Commun.
Heat Mass Transf. 2022, 137, 106277. [CrossRef]
22. Bolleddula, D.A.; Berchielli, A.; Aliseda, A. Impact of a Heterogeneous Liquid Droplet on a Dry Surface: Application to the
Pharmaceutical Industry. Adv. Colloid Interface Sci. 2010, 159, 144–159. [CrossRef] [PubMed]
23. Nicolas, M. Spreading of a Drop of Neutrally Buoyant Suspension. J. Fluid Mech. 2005, 545, 271–280. [CrossRef]
24. Zhao, Z.; Wang, R.; Ge, L.; Wu, J.; Yin, Q.; Wang, C. Energy Utilization of Coal-Coking Wastes via Coal Slurry Preparation: The
Characteristics of Slurrying, Combustion, and Pollutant Emission. Energy 2019, 168, 609–618. [CrossRef]
25. Kuznetsov, G.V.; Romanov, D.S.; Vershinina, K.Y.; Strizhak, P.A. Rheological Characteristics and Stability of Fuel Slurries Based
on Coal Processing Waste, Biomass and Used Oil. Fuel 2021, 302, 121203. [CrossRef]
26. Volkov, R.S.; Strizhak, P.A. Using Planar Laser Induced Fluorescence and Micro Particle Image Velocimetry to Study the Heating
of a Droplet with Different Tracers and Schemes of Attaching It on a Holder. Int. J. Therm. Sci. 2021, 159, 106603. [CrossRef]
27. Volkov, R.S.; Strizhak, P.A.; Misyura, S.Y.; Lezhnin, S.I.; Morozov, V.S. The Influence of Key Factors on the Heat and Mass Transfer
of a Sessile Droplet. Exp. Therm. Fluid Sci. 2018, 99, 59–70. [CrossRef]
28. Breitenbach, J.; Roisman, I.V.; Tropea, C. From Drop Impact Physics to Spray Cooling Models: A Critical Review. Exp. Fluids 2018,
59, 55. [CrossRef]
29. Liang, G.; Mudawar, I. Review of Drop Impact on Heated Walls. Int. J. Heat Mass Transf. 2017, 106, 103–126. [CrossRef]
30. Piskunov, M.; Ashikhmin, A.; Khomutov, N.; Semyonova, A. Effects of Wall Temperature and Temperature-Dependent Viscosity
on Maximum Spreading of Water-in-Oil Emulsion Droplet. Int. J. Heat Mass Transf. 2022, 185, 122442. [CrossRef]
31. Rioboo, R.; Marengo, M.; Tropea, C. Time Evolution of Liquid Drop Impact onto Solid, Dry Surfaces. Exp. Fluids 2002, 33, 112–124.
[CrossRef]
32. Piskunov, M.; Breitenbach, J.; Schmidt, J.B.; Strizhak, P.; Tropea, C.; Roisman, I.V. Secondary Atomization of Water-in-Oil Emulsion
Drops Impinging on a Heated Surface in the Film Boiling Regime. Int. J. Heat Mass Transf. 2021, 165, 120672. [CrossRef]
33. Burzynski, D.A.; Roisman, I.V.; Bansmer, S.E. On the Splashing of High-Speed Drops Impacting a Dry Surface. J. Fluid Mech. 2020,
892, A2. [CrossRef]
34. Sharp, D.H. An Overview of Rayleigh-Taylor Instability. Phys. D Nonlinear Phenom. 1984, 12, 3–18. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

129
energies
Article
Nucleation of a Vapor Phase and Vapor Front Dynamics Due to
Boiling-Up on a Solid Surface
Artem N. Kotov, Aleksandr L. Gurashkin, Aleksandr A. Starostin, Kirill V. Lukianov and Pavel V. Skripov *

Institute of Thermal Physics, Ural Branch, Russian Academy of Sciences, Yekaterinburg 620016, Russia;
[email protected] (A.N.K.); [email protected] (A.L.G.); [email protected] (A.A.S.);
[email protected] (K.V.L.)
* Correspondence: [email protected]

Abstract: The effect of temperature and pressure on the nucleation of the vapor phase and the velocity
of the vapor front in the initial stage of activated boiling-up of n-pentane on the surface of a quartz
fiber was studied. Using a developed approach combining the “pump-probe” and laser Doppler
velocimetry methods, this velocity was tracked in the course of sequential change in the degree of
superheating with respect to the liquid–vapor equilibrium line. The studied interval according to
the degree of superheating was 40–100 ◦ C (at atmospheric pressure). In order to spatiotemporally
localize the process, the activation of boiling-up at the end of the light guide was applied using a short
nanosecond laser pulse. A spatial locality of measurements was achieved in units of micrometers,
along with a time localization at the level of nanoseconds. An increase in temperature at a given
pressure was found to lead to an increase in the speed of the transition process with a coefficient of
about 0.2 m/s per degree, while an increase in pressure at a given temperature leads to a decrease in
the transition process speed with a coefficient of 25.8 m/s per megapascal. The advancement of the
vapor front velocity measurements to sub-microsecond intervals from the first signs of boiling-up
did not confirm the existence of a Rayleigh expansion stage with a constant velocity.

Keywords: pump-probe; superheated liquid; n-pentane; activation boiling-up; laser pulse

Citation: Kotov, A.N.; Gurashkin,


A.L.; Starostin, A.A.; Lukianov, K.V.;
Skripov, P.V. Nucleation of a Vapor
1. Introduction
Phase and Vapor Front Dynamics
Due to Boiling-Up on a Solid Surface.
The superheating of a liquid precedes and accompanies the liquid–vapor phase transi-
Energies 2023, 16, 6966. https:// tion used in various technological applications [1–6]. Interest in the issue of boiling liquids
doi.org/10.3390/en16196966 has received an additional impetus due to the development of mini- and micro-sized
devices [6,7]. An important component of this problem is the stage of the vapor phase
Academic Editors: Vladimir
nucleation onset [8,9]. It is registered by some macroscopic response of the system, which
Serdyukov, Fedor Ronshin and
is monitored until the superheating is removed.
Moran Wang
The practically significant case of nucleate boiling-up on a heated surface begins with
Received: 18 August 2023 bubble nucleation in the near-wall layer of the superheated liquid. Boiling-up modes
Revised: 11 September 2023 are determined by the superheat value of the liquid with respect to the liquid–vapor
Accepted: 3 October 2023 equilibrium line. Essentially, the initial stage of boiling represents an isothermal process
Published: 6 October 2023 of a vapor-bubble expansion, which is limited by the rate at which a mechanical impulse
is transmitted to the surrounding liquid (inertial bubble growth). Since turbulence of the
liquid occurs near the heating surface, the growth rate of the vapor bubble determines
the hydrodynamic regime in the wall layer and the intensity of heat exchange [10,11]. In
Copyright: © 2023 by the authors.
connection with the development of microelectronics, it is timely to study heat transfer
Licensee MDPI, Basel, Switzerland.
This article is an open access article
from locally heated “hot spots” at the boundary with the coolant [12].
distributed under the terms and
From the first experiments on the superheating of liquids, the attention of researchers
conditions of the Creative Commons was attracted to the initial stage of the boiling-up relative to the liquid–vapor equilibrium
Attribution (CC BY) license (https:// line [13]. By reducing the product V liq ·texp (where V liq —volume of the superheated liquid;
creativecommons.org/licenses/by/ texp —observation time of the superheated state), it was possible to break through the
4.0/). background of the ready vaporization centers to achieve significant superheating. Due

Energies 2023, 16, 6966. https://doi.org/10.3390/en16196966 130 https://www.mdpi.com/journal/energies


Energies 2023, 16, 6966

to the characteristic random nature of spontaneous boiling-up [14], it is difficult to study


the processes in detail. The use of recording equipment is complicated by a combination
of uncertainty in the waiting time for boiling-up and the relatively rapid transition to
the saturation line. This is due to the difficulty of identifying the site and moment of the
spontaneous boiling-up onset in a liquid at the necessary microsecond resolution for the
recording of rapid processes. A significant advance in this regard was achieved with the
use of low-inertia metal heaters comprised of thin wires [14,15] and films [16,17]. Due to
the relatively rapid heating of wires and films, it became possible to reduce the uncertainty
at the time of boiling-up. The superheating of liquid was achieved in a thin layer adjacent
to the surface of the heater during non-stationary heating. Consequently, the boiling site
turned out to be localized on the heating surface. Numerous experiments have been carried
out with oscillography and high-speed recording of the surface boiling processes [14–24];
extensive data have been obtained on the attainable superheating of liquids during their
pulsed heating [4–6]. However, there was uncertainty associated with the randomness
of the boiling site on the surface and the influence of the thickness of the heated layer
(the temperature gradient normal to the heating surface) on the dynamics of the observed
processes. In addition, the method was mainly applied at high superheating rates to reduce
the waiting time for boiling-up onset in light of the influence of surface treatment quality
over longer time intervals [17,18,25].
In pursuit of further localization in terms of the location and time of observation of
the initial stage of boiling-up, a technique was developed for the pulsed laser activation of
the liquid–vapor transition in a miniature bubble chamber [26]. The bubble chamber [4,14]
is characterized by the uniform superheating of the liquid in the capillary. Since, as a result,
there is no restriction on the thickness of the heated layer, the value of the superheating
temperature can be set more precisely. The development of the boiling-up process can be
traced by the reflection of a probing beam following the activation pulse. This approach,
when the first powerful pulse transfers the system under study to a nonequilibrium state,
and the subsequent probing serves to track the relaxation process of the system, is known
as the “pump-probe” method [27]. The advantage of this method lies in the strict synchro-
nization of the processes of normalized exposure and observation, which allows for the rate
of occurrence and repeatability of the observed phenomena to be assessed even at short
time intervals. The purpose of the present work is to demonstrate the capabilities of the
“pump-probe” laser method by identifying the features of the initial stage of activated liq-
uid boiling-up (on the example of n-pentane) under the conditions of uniform superheating
in a miniature bubble chamber.

2. Background
The action of the bubble chamber consists of transferring the liquid to the area of
superheated states by resetting the initial pressure in the liquid p0 below the equilibrium
pressure liquid/vapor ps (see Figure 1, A–C transition).
Thus, the degree of superheating was given by the final pressure value pexp < ps ; the
parameter monitored in the experiment was the lifetime of the superheated state prior to its
decay by spontaneous boiling-up at given thermodynamic parameters. A practical basis for
the detailed verification of the theory of homogeneous nucleation [8], and the measurement
of different properties of substances in superheated states [4,28–30] including fuel-in-water
emulsions and fuel blends [31], consisted of the determination of the temperature, pressure,
and volume of the superheated liquid. To track the rapid boiling processes, a fiber optic
sensor is placed in the capillary of the bubble chamber [9,29,32,33]. As an example, Figure 2
shows records of the spontaneous boiling-up signals of hexane carried out using the fiber
optic sensor in previous works [9,32,33]. The change in signal voltage at the output of the
photodetector is associated with a change in the density of the medium surrounding the
fiber optic sensor. The letters indicating the stages of the process correspond to the states
and transitions noted in Figure 1.

131
Energies 2023, 16, 6966

Figure 1. Liquid–vapor phase diagram designating the states of matter in the course of experiments
carried out in the bubble chamber. A—initial state (p0 > ps ); C—superheated state (pexp < ps ); E—two-
phase state on the saturation line; CP—critical point. Arrows B and D schematically represent the
pressure relief and boiling-up development processes. 1—saturation line; 2—attainable superheat
line; 3—spinodal.

Figure 2. Voltage U of the fiber optic sensor photodetector against time in the course of experiment
to study the spontaneous boiling-up of n-hexane. A—initial state (p0 > ps ); B—depressurization;
C—superheated (metastable) state (pexp < ps ); D—boiling-up and transition to the saturation line;
E—two-phase state on the saturation line.

The initial equilibrium state is indicated by A (at initial pressure p0 > ps and tempera-
ture T0 = Texp ). Transition B denotes the pressure release to pexp < ps . In state C, the liquid
is superheated (metastable) until boiling-up occurs, followed by a two-phase transition, D.
The process terminates in equilibrium state E on the saturation line. The described cycle
can be repeated many times for a dataset to permit its statistical processing. In state C,

132
Energies 2023, 16, 6966

spontaneous boiling can be expected or activated by some external action. The region of
vapor phase nucleation of interest to us is located at the transition from state C to state D
(marked with an arrow in Figure 2). Our approach to the boiling-up activation of n-pentane
in a bubble chamber is described in [26]. The activation by a short laser pulse at the end of
the light guide in the transparent liquid was found to have a threshold character in terms
of the intensity of the excitation pulse. Increasing the energy of the pulse by changing its
duration within wide limits (up to two orders of magnitude) with insufficient intensity did
not lead to boiling-up. The results of a study into the effects accompanying the nanosecond
excitation pulse suggest that the activation mechanism may be of a non-thermal nature. It is
likely that an electrostriction mechanism of the liquid is activated at a certain intensity of the
electromagnetic field of the laser pulse. Accordingly, the boiling-up activation is generated
by a mechanical impulse from electrostriction. On the basis of the detected phenomenon,
a method was developed for activating and studying transient the boiling-up processes
across a wide area of superheated states with a precisely defined superheating temperature
in the bubble chamber. The method of observing transient processes is described in [29,32].
A light guide in the bubble chamber was used to observe the liquid–vapor transition by
changing the reflection signal from the end of the light guide with changes in the density
of the medium [9,29,32]. Since the speed of the available photodetectors is measured in
hundreds of megahertz, it is possible to record processes at nanosecond resolution.

3. Method
The application of the pump-probe method is illustrated in Figure 3. A pump pulse
beam with a duration of 1–10 ns was generated by pulsed laser 1530 nm. A probe beam
was continuously generated by CW laser 1550 nm. Combining the emissions for the
transmission over a single fiber was achieved using the add-drop multiplexer. The pump
pulse beam acted on the end of the “optofiber” on the liquid superheated in the capillary
and, at a sufficient intensity, activated its boiling. The probe beam reflected from the
end was returned back through the add-drop multiplexer and optical circulator to the
photodetector. The intensity of the reflected signal depends on the density of the medium
at the end of the optofiber and increased as the density decreased. Therefore, the boiling
of the liquid was accompanied by a sharp increase in the reflected signal. In the course of
boiling, a two-phase system was formed near the end of the optofiber, as shown in Figure 3.
Then, the probe beam had a double reflection from the end and from the vapor–liquid
interface, followed by interference. Since this interface moved as the bubble grows, the
photodetector signal exhibited oscillations. The oscillation frequency is related to the speed
of the interface movement according to the Doppler effect.

Figure 3. Scheme for the formation of optical signals by the pump-probe method for activating the
boiling-up of a superheated liquid at the end of a quartz light guide.

133
Energies 2023, 16, 6966

To activate the boiling of n-pentane through the light guide at different temperatures
and pressures, pulses of laser radiation (pump beam pulses) were applied. The required
intensity and energy of the activating pulses increased with a decrease in the superheating
temperature and an increase in the residual pressure pexp [26]. An increase in the work of
the bubble formation with a decrease in the superheating temperature and with an increase
in external pressure [4,5,14] predetermined an increase in the intensity of the activating
pulses. The development of the phase transition was monitored by changing the intensity
of the probe beam reflected from the end of the light guide. Here, the measured intensity
of the reflected radiation increased with a decrease in the refractive index of the medium
according to the Fresnel formula. In turn, the refractive index decreased with a reduction in
the density of the medium. The dynamics of the changes in the density of the medium were
assessed by tracking the changes in the average intensity of the reflected radiation according
to the photodetector signal following the activating pulse [26,34]. During the fiber-optic
densitometry experiments with superheated liquid following activation of boiling-up,
high-frequency oscillations were observed at the output of the high-speed photodetector
of reflected radiation. However, the observed oscillations are not associated with changes
in the density of the medium, appearing rather as a result of interference of the reflected
rays. The traveling interference pattern shown in Figure 4, which is characteristic of the
interferometers with a moving reflector, is due to the formation of an additional reflected
signal as a result of the superposition of the origin and propagation of the phase interface
with the reflected light waves from the stationary end of the fiber [34]. Each oscillation
period in the recorded signal (Figure 4) corresponds to a shift of the vapor front by half the
wavelength of the probing radiation.

Figure 4. Structure of the reflected signal with interference of reflected rays from the end of the fiber
and the moving phase interface following the activation of boiling-up of n-pentane at 120 ◦ C using a
laser pulse.

In this case, the frequency of the observed oscillations depended on the speed of the
movement of the reflector according to the Doppler effect. Thus, the speed of movement of

134
Energies 2023, 16, 6966

the reflecting phase interface can be determined by measuring the frequency of the recorded
oscillations. This laser measurement method is called laser Doppler velocimetry [35]. In our
experiments, the Doppler frequency shift increased in accordance with an intensification
of in the bubble growth rate due to the higher superheating temperature of n-pentane at
atmospheric pressure.
The values of the velocity of movement of the phase interface were determined by the
Doppler velocimetry ratio:
V(t) = FD (t)·λ/2n(t), (1)
where FD —current Doppler frequency shift (MHz); λ—wavelength of the laser diode
radiation (1.55 microns); and n(t)—current refractive index of the medium:

n(t) = nliq − (nliq — nvap )·[U(t) − Umin ]/(Umax − Umin ), (2)

where nliq , nvap —refractive indices of the liquid and vapor phases on the saturation line at
the experimental temperature, respectively; Umin ; Umax —minimum and maximum average
values of the photodetector output signal corresponding to the values of nliq , nvap ; and
U(t)—current average value of the output signal of the photodetector.
Since the oscillations developed during decompression to the vapor phase, in most
cases, n(t) => 1.

4. Installation
According to the phase diagram (Figure 1), two stages are necessary to obtain a
superheated state in the bubble chamber. At the first stage, the sample is heated and held
in a stable liquid state (point A) at elevated pressure. At the second stage, there was a
rapid decrease in the pressure (to atmospheric pressure) and the transfer of the liquid to a
superheated state (point C). An experimental setup implementing such a mode is shown
in Figure 5. The test liquid was used to fill a glass capillary with an internal diameter of
1 mm and a heated section of 30 mm. Inside the capillary, the liquid pressure could be set
by the external pressure of nitrogen vapor on the separation membrane. To achieve this,
gas was supplied from the N2 gas tank to the gas–liquid separation block via the pressure
control block. The movable membrane, which separated the nitrogen and the liquid under
study while maintaining their hydraulic connection, was installed inside the gas–liquid
separation block. The mobility of the membrane and the low compressibility of the liquid
were necessary to ensure the equality of the pressures in the capillary and in the gas cavity
of the gas–liquid separation block. Such a system allowed for two possible pressure values
to be selected for the liquid in the capillary: atmospheric and that provided from the N2
gas tank via the pressure control block. The pressure control block is controlled from a PC.
In order to heat the sample, the glass capillary was immersed in a transparent ther-
mostat (heat carrier) with a coolant fluid (thermostat liquid). The coolant fluid was heated
using an electric heater.
A precision temperature sensor was installed to monitor the temperature near the
capillary. The sensor was a type K thermocouple with measurement error of 0.1 ◦ C.
To increase the accuracy of measurements, the calibration curve of thermocouples was
obtained on a special bench by comparing it with the indications of a platinum thermometer.
The thermostat control block maintained the set temperature of the coolant by reading the
sensor indications and adjusting the electric power level of the heater. This temperature
was regulated directly on the unit or via software (PC).

135
Energies 2023, 16, 6966

Figure 5. Block diagram of an experimental setup for the study of activated boiling-up.

A fiber optic probe was installed inside the capillary to supply the activation effect
to the superheated liquid. The probe was a standard single-mode quartz fiber light guide,
which was previously cleaned from the outer shell, and whose free end was located in
the capillary. The outer diameter of the light guide was 125 microns; the diameter of the
fiber core was 10 microns. The probe was connected to the electronic radiation supply
and reception units via a fiber optic device. This scheme implemented the “pump-probe”
principle. For the pump, a pulsed laser of 1530 nm, having an adjustable power of up to
30 W and a variable pulse duration on a scale from 1 ns to 100 ns, was used. The probing
beam was generated by a permanent laser (CW laser 1550 nm) having a wavelength of
1550 nm and a power of 2.5 mW. Both kinds of radiation—pumping and probing—were
combined and fed into the fiber probe using fiber elements comprising an adder and an
optical circulator, which make up the fiber optic device. The probing radiation reflected
from the free end of the light guide was separated, filtered, and then fed to the photodetector.
The output of the photodetector generated an electrical signal, which was recorded using
a high-speed oscilloscope. The recorded waveform files were transferred to the PC for
processing. The photodetector received signals in the frequency range 0–100 MHz with a
sensitivity to optical power of 0.1 V/μW. The Doppler frequency measurements using a
photodetector and a Rigol 5354 oscilloscope have an uncertainty of 5% over the frequency
range 10–40 MHz.
Due to the synchronous activation of the boiling-up with the pumping pulse and the
subsequent growth of the vapor bubble, it was possible to capture the stages of bubble
growth on a microsecond time scale using the stroboscopic video method. For this purpose,
a video camera with a frame rate of 60 Hz in stroboscopic mode was used. The mode was
provided by the operation of a pulsed light source, with the generation of a short (from
0.1 μs to 0.5 μs) light pulse (strobe) delayed relative to the pump pulse for a specified time.
By changing the pause time between the strobe and the pump pulse, various stages of
bubble growth could be recorded in a series of experiments (Figure 6).

136
Energies 2023, 16, 6966

Figure 6. A vapor bubble at the end of the light guide in n-pentane at a delay of 50 μs from the
moment of activation and an experimental temperature of 123 ◦ C. The duration of illumination
(strobe) was 0.5 μs.

5. Results
The experiments in the bubble chamber were carried out in a wide area of superheated
states created by the pressure drop in a heated liquid. The mean “lifetime” of n-pentane
before spontaneous boiling-up under the given conditions (pexp < ps , Texp > Ts ) ranged from
units of seconds to tens of minutes. This time was sufficient to establish thermodynamic
equilibrium following the pressure drop [14]. The activation of boiling was carried out by a
single nanosecond pulse pump pulse (1–10 ns) with an intensity exceeding the activation
threshold by ~10%. The values of the intensity and the duration of the pulse were selected
experimentally from the condition of the absence of their influence on the recorded data.
Figure 7 shows frames from the video footage obtained using the stroboscopic method.
The duration of synchronous illumination with the pump pulse was 0.5 μs. The delay
from the pump pulse increased by 1 microsecond for each subsequent frame. The video
sequence shows the regular growth of the vapor film at the end of the light guide over time
and the formation of a bubble after 3 μs at a temperature of 130 ◦ C and at atmospheric
pressure. Despite the image distortion due to light refraction, it is possible to estimate
the size of the vapor cavity relative to the size of the light guide (125 μm) after 3 μs at a
value of about 60 μm (average growth rate is 20 m/s). It is noteworthy that, for a short
period of time, the surface tension forces hold the bubble at the end of the fiber. The
subsequent development of the process was less predictable with the increasing influence
of hydrodynamic disturbances.
An example of the recorded dependences of the voltage in the photodetector on time
is shown in Figure 8. The waveforms of signals for n-pentane temperature values of 80 ◦ C,
95 ◦ C, and 130 ◦ C are shown (superheating Texp − Ts is 43, 58, and 93 degrees, respectively).

137
Energies 2023, 16, 6966

Figure 7. Bubble growth in n-pentane at the end of the light guide after the pump pulse. The
images were obtained by stroboscopic shooting with an increase in the pulse illumination delay by
1 microsecond for each subsequent image. The experiment was carried out at a temperature of 130 ◦ C
and atmospheric pressure. The duration of illumination (strobe) was 0.5 μs.

Figure 8. Measured reflection signals from the end of the light guide following activation of n-pentane
boiling at different temperatures and atmospheric pressure on a general (a) and enlarged (b) scale with
different frequency of Doppler shift at an activated phase transition with different conversion rates.

A change in the dynamics of signal growth following the activation of the boiling-up
is shown in Figure 8a, while Figure 8b depicts the change in the frequency of the Doppler
shift depending on temperature. The transition process to the steady-state average value
occurred over a time from 1 μs for 130 ◦ C to 6 μs for 80 ◦ C. The frequency of Doppler
oscillations varied accordingly from 12 to 25 MHz. The change in the oscillation frequency
of up to 10% observed at intervals of estimating the rate of processes was associated with a

138
Energies 2023, 16, 6966

known decrease in the bubble growth rate over time [11]. For the subsequent calculations,
averaged values were used.
Figure 9 shows the dependences calculated from the frequency of oscillations of the
vapor front velocity on time at different temperatures Texp . Since a determination of the
frequency of oscillations and velocity was possible only when a sufficient amplitude of the
signal was reached, there is a delay in the beginning of measurements in Figure 9.

Figure 9. Change in the velocity of the vapor front over time at different initial temperatures of
superheated n-pentane. The sampling delay increases for lower temperatures due to a decrease in the
amplitude of the recorded signal at the beginning of the process.

The developed Doppler velocimetry technique was validated for measuring the veloc-
ity of the evaporation front on samples of pure n-pentane at various degrees of superheating
relative to the liquid–vapor equilibrium temperature when the pressure drops to atmo-
spheric values was carried out (Figure 10a). In particular, for a temperature of 130 ◦ C,
the average growth rate of 19.2 m/s agrees quite well with the photographic estimate of
20 m/s (see Figure 7). Figure 10b shows the effect of the final pressure pexp in the course of
its release at a given temperature Texp .

Figure 10. Temperature and baric dependencies of the initial velocity of the phase boundary displace-
ment calculated by the Doppler velocimetry method at atmospheric pressure (a) and at a temperature
of 135 ◦ C (b).

139
Energies 2023, 16, 6966

6. Discussion
Boiling-up is commonly defined as the process of separation of the liquid and vapor
phases inside a superheated liquid. For our case of relatively high superheating (40–100 ◦ C
for pentane), the formation of a vapor film on the surface of the fiber end face was observed
(see Figure 7). The dynamics of the appearance and growth of a vapor film at the boundary
of a solid surface and a highly superheated liquid significantly determine the intensity of
heat transfer at the liquid boiling-up onset [36]. The approach combining pump-probe and
laser Doppler velocimetry made it possible to track the vapor front velocity at the initial
stage of n-pentane boiling in a miniature bubble chamber with a successive change in the
degree of superheat.
The first obtained result consists of the possibility of such a local impact occurring in
which the subsequent boiling process proceeds independently of the pump pulse parame-
ters. In this case, a fairly accurate replication of the video frames and waveforms is capable
of being repeatedly reproduced. The values of the initial growth rate of the vapor phase at
the end of the fiber calculated from the measured Doppler frequencies differ significantly
(by about 2 times) from those calculated using the Rayleigh formula for the inertial stage.
Rayleigh’s formula for inertial bubble growth [11] assumes the constancy of the growth
.
rate R for bubbles with a radius R of more than 1 μm in a superheated under the conditions
of our experiments liquid:
. 2Δp
R= (3)

where Δp is the difference between the pressure in the bubble and the pressure in the liquid,
and ρ is the density of the liquid.
Despite the advance of measurements of the vapor front velocity to sub-microsecond
intervals from the beginning of boiling-up, we were not able to identify a region of ex-
pansion with a constant velocity as predicted by Rayleigh’s formula. Figure 8a shows a
gradual increase in the signal with time at the output of the fiber-optic densitometer, which
corresponds to the gradual density decrease in the medium near the end of the fiber. The
dependences of the vapor cavity growth rate on time and temperature turned out to be
more consistent with thermally controlled bubble growth (see Figures 9 and 10). Despite the
process occurring near the hot surface of the end of the fiber, the obtained results confirm
the assumptions made earlier [37] about a lower vapor pressure and the probable cooling
of the vapor film already at the initial stage of growth. The obtained dependencies of the
initial velocity of the phase boundary displacement on temperature and pressure are close
to linear (see Figure 10). The opposite effects of the experimental temperature and pressure
are consistent with the change in the degree of superheating of the liquid (see Figure 1). An
increase in temperature increased the speed of the boundary displacement by a factor of
about 0.2 m/s per degree, while an increase in pressure reduced this speed by a factor of
~25.8 m/s per megapascal.

7. Conclusions
A method for measuring the velocity of the vapor front in the course of the activation
of liquid boiling-up at the end of a light guide due to a short laser pulse has been developed
and tested. The experiments were carried out in a miniature bubble chamber of the type
traditionally used to determine the dependencies of the mean “lifetime” of a liquid and
rate of nucleation on the temperature of superheating under strictly controlled conditions.
The main novelty of our approach consists of the achievement of spatial localization in
the units of micrometers with localization in time at the level of units of nanoseconds.
The combination of a fiber-optic sensor with a bubble chamber according to the pump-
probe method has opened up new opportunities for studying the fast-flowing processes
during initial stages of liquid–vapor phase transitions using densitometric and velocimetric
techniques. The obtained results, which do not fit into the Rayleigh scheme, can be
interpreted as confirming the assumption of a lower vapor pressure and cooling of the

140
Energies 2023, 16, 6966

vapor film already at the early stage of vapor phase growth. By applying this approach to
the study of the initial velocity of the vapor front, it becomes possible to test theoretical
ideas about the development of a phase transition at the interface between a liquid and a
solid surface depending on the degree of superheating of the liquid.
Activation by nanosecond pulses was used to involve a wide area of moderately
superheated states in the study, which were previously considered difficult to access due
to the long waiting times for spontaneous boiling-up. Further research will be carried out
in the field of relatively low superheating, which has significance for various practical
applications, such as alternative fuel technologies based on rapid secondary atomization of
composite droplets and modern minimally invasive medical tools, see references [31,34]
and bibliography therein.

Author Contributions: Conceptualization, A.A.S. and P.V.S.; methodology, A.N.K. and A.A.S.; soft-
ware K.V.L.; hardware A.N.K. and A.L.G.; validation, A.N.K., A.A.S. and P.V.S.; writing—original
draft preparation, A.N.K., A.A.S., A.L.G. and P.V.S.; writing—review and editing, A.N.K., A.A.S. and
P.V.S. All authors have read and agreed to the published version of the manuscript.
Funding: The investigation has been conducted at the expense of a grant of the Russian Science
Foundation (project № 19-19-00115-P).
Data Availability Statement: Data are available on request due to organization rules.
Conflicts of Interest: The authors declare no conflict of interest.

Abbreviations
Definition
Subscripts
D Doppler
exp Experimental
liq Liquid
max Maximum
min Minimum
s Saturation
vap Vapour
Variables and functions
Δp Difference between the pressure in the bubble and in the liquid
λ Wavelength of the probing radiation
ρ Density of the liquid
FD (t) Doppler frequency shift
n(t) Current refractive index of the medium
nliq Refractive index of the liquid phases
nvap Refractive index of the vapor phases
p0 Initial pressure in the liquid
ps Equilibrium pressure liquid/vapor
.
R Bubbles growth rate
T0 Initial liquid temperature
Texp Experimental temperature
Ts Equilibrium temperature liquid/vapor
texp Observation time of the superheated state
U Voltage of the fiber optic sensor photodetector
U(t) Current value of the output signal of the photodetector
Umax Maximum average values of the photodetector output signal
Umin Minimum average values of the photodetector output signal
V liq Volume of the superheated liquid
V(t) Velocity of movement of the phase interface

141
Energies 2023, 16, 6966

Abbreviations
CW Continuous wave
CP Critical point
States
A Initial state (p0 > ps )
B Depressurization
C Superheated state (pexp < ps )
D Boiling-up and transition to the saturation line
E Two-phase state on the saturation line

References
1. Berthoud, G. Vapor Explosions. Annu. Rev. Fluid Mech. 2000, 32, 573–611. [CrossRef]
2. Sazhin, S.S. Droplets and Sprays: Simple Models of Complex Processes; Springer: Cham, Switzerland, 2022.
3. Antonov, D.V.; Fedorenko, R.M.; Yanovskiy, L.S.; Strizhak, P.A. Physical and Mathematical Models of Micro-Explosions: Achieve-
ments and Directions of Improvement. Energies 2023, 16, 6034. [CrossRef]
4. Skripov, V.P.; Sinitsyn, E.N.; Pavlov, P.A.; Ermakov, G.V.; Muratov, G.N.; Bulanov, N.V.; Baidakov, V.G. Thermophysical Properties of
Liquids in the Metastable (Superheated) State; Gordon and Breach Science Publishers: London, UK, 1988.
5. Debenedetti, P.G. Metastable Liquids: Concepts and Principles; Princeton University Press: Princeton, NJ, USA, 1996. [CrossRef]
6. Ching, E.J.; Avedisian, C.T.; Cavicchi, R.C.; Chung, D.H.; Rah, K.J.; Carrier, M.J. Rapid Evaporation at the Superheat Limit of
Methanol, Ethanol, Butanol and n-Heptane on Platinum Films Supported by Low-Stress SiN Membranes. Int. J. Heat Mass Transf.
2016, 101, 707–718. [CrossRef] [PubMed]
7. Khan, S.; Atieh, M.; Koç, M. Micro-Nano Scale Surface Coating for Nucleate Boiling Heat Transfer: A Critical Review. Energies
2018, 11, 3189. [CrossRef]
8. Ermakov, G.V.; Lipnyagov, E.V.; Perminov, S.A. Classical Theory of Homogeneous Nucleation in Superheated Liquids and Its
Experimental Verification. Thermophys. Aeromechanics 2012, 19, 667–678. [CrossRef]
9. Gurashkin, A.L.; Starostin, A.A.; Ermakov, G.V.; Skripov, P.V. Communication: High Speed Optical Investigations of a Character
of Boiling-up Onset. J. Chem. Phys. 2012, 136, 021102. [CrossRef] [PubMed]
10. Pavlenko, A.N. Life Devoted to Science. Thermophys. Aeromechanics 2014, 21, 265–278. [CrossRef]
11. Avdeev, A.A. Thermally Controlled Bubble Growth. In Bubble Systems.; Avdeev, A.A., Ed.; Springer International: Cham,
Switzerland, 2016; Volume 4, pp. 99–132.
12. Khandekar, S.; Sahu, G.; Muralidhar, K.; Gatapova, E.Y.; Kabov, O.A.; Hu, R.; Luo, X.; Zhao, L. Cooling of High-Power LEDs by
Liquid Sprays: Challenges and Prospects. Appl. Therm. Eng. 2021, 184, 115640. [CrossRef]
13. Wismer, K.L. The Pressure-Volume Relation of Super-Heated Liquids. J. Phys. Chem. 1922, 26, 301–315. [CrossRef]
14. Skripov, V.P. Metastable Liquids; Halsted Press: Sydney, Australia; John Wiley & Sons: New York, NY, USA, 1974.
15. Glod, S.; Poulikakos, D.; Zhao, Z.; Yadigaroglu, G. An Investigation of Microscale Explosive Vaporization of Water on an Ultrathin
Pt Wire. Int. J. Heat Mass Transf. 2002, 45, 367–379. [CrossRef]
16. Asai, A. Bubble Dynamics in Boiling Under High Heat Flux Pulse Heating. J. Heat Transf. 1991, 113, 973–979. [CrossRef]
17. Iida, Y.; Okuyama, K.; Sakurai, K. Boiling Nucleation on a Very Small Film Heater Subjected to Extremely Rapid Heating. Intern.
J. Heat Mass Transf. 1994, 37, 2771–2780. [CrossRef]
18. Nikitin, E.D. Mechanisms of Vaporization in Superheated Water; Ural Polytechnical Institute: Sverdlovsk, Russia, 1981.
19. Strenge, P.H.; Orell, A.; Westwater, J.W. Microscopic Study of Bubble Growth during Nucleate Boiling. AIChE J. 1961, 7, 578–583.
[CrossRef]
20. Hong, Y.; Ashgriz, N.; Andrews, J. Experimental Study of Bubble Dynamics on a Micro Heater Induced by Pulse Heating. J. Heat
Transf. 2004, 126, 259–271. [CrossRef]
21. Kuznetsov, V.V.; Oreshkin, V.I.; Zhigalin, A.S.; Kozulin, I.A.; Chaikovsky, S.A.; Rousskikh, A.G. Metastable States and Their
Disintegration at Pulse Liquid Heating and Electrical Explosion of Conductors. J. Eng. Thermophys. 2011, 20, 240–248. [CrossRef]
22. Skripov, P.V.; Puchinskis, S.E.; Begishev, V.P.; Lipchak, A.I.; Pavlov, P.A. Heat Pulse Monitoring of Curing and Polymer–Gas
Systems. J. Appl. Polym. Sci. 1994, 51, 1607–1619. [CrossRef]
23. Avedisian, C.T.; Cavicchi, R.E.; Tarlov, M.J. New Technique for Visualizing Microboiling Phenomena and Its Application to Water
Pulse Heated by a Thin Metal Film. Rev. Sci. Instr. 2006, 77, 063706. [CrossRef]
24. Serdyukov, V.; Malakhov, I.; Surtaev, A. The Influence of Pressure on Local Heat Transfer Rate under the Vapor Bubbles during
Pool Boiling. Energies 2023, 16, 3918. [CrossRef]
25. Skripov, P.V.; Bar-Kohany, T.; Antonov, D.V.; Strizhak, P.A.; Sazhin, S.S. Approximations for the Nucleation Temperature of Water.
Int. J. Heat Mass Transf. 2023, 207, 123970. [CrossRef]
26. Kotov, A.N.; Gurashkin, A.L.; Starostin, A.A.; Skripov, P.V. Low-energy activation of superheated n-pentane boiling-up by laser
pulse at the fiber-liquid interface. Interf. Phen. Heat Transf. 2022, 10, 15–23. [CrossRef]
27. Kotov, A.N.; Starostin, A.A.; Gorbatov, V.I.; Skripov, P.V. Thermo-Optical Measurements and Simulation in a Fibre-Optic Circuit
Using an Extrinsic Fabry–Pérot Interferometer under Pulsed Laser Heating. Axioms 2023, 12, 568. [CrossRef]

142
Energies 2023, 16, 6966

28. Skripov, P.V.; Skripov, A.P. The Phenomenon of Superheat of Liquids: In Memory of Vladimir P. Skripov. Int. J. Thermophys. 2010,
31, 816–830. [CrossRef]
29. Gurashkin, A.L.; Starostin, A.A.; Uimin, A.A.; Yampol’skiy, A.D.; Ermakov, G.V.; Skripov, P.V. Experimental Determination of
Superheated Liquid Density by the Optical Fiber Method. J. Eng. Thermophys. 2013, 22, 194–202. [CrossRef]
30. Lipnyagov, E.V.; Parshakova, M.A. Investigation of the Kinetics of Spontaneous Boiling-up of Superheated n-Pentane in a Glass
Tube with Defects of the Inner Surface. I. Monitoring the Liquid-Vapor Surface Tension. Int. J. Heat Mass Transf. 2022, 196, 123254.
[CrossRef]
31. Antonov, D.V.; Fedorenko, R.M.; Strizhak, P.A. Micro-Explosion Phenomenon: Conditions and Benefits. Energies 2022, 15, 7670.
[CrossRef]
32. Davitt, K.; Arvengas, A.; Caupin, F. Water at the Cavitation Limit: Density of the Metastable Liquid and Size of the Critical Bubble.
Europhys. Lett. 2010, 90, 16002. [CrossRef]
33. Gurashkin, A.L.; Yampol’skii, A.D.; Starostin, A.A.; Skripov, P.V. Optical Studies of the Initial Stage of Spontaneous Boiling-Up.
Tech. Phys. Lett. 2013, 39, 751–754. [CrossRef]
34. Zubalic, E.; Vella, D.; Babnik, A.; Jezeršek, M. Interferometric Fiber Optic Probe for Measurements of Cavitation Bubble Expansion
Velocity and Bubble Oscillation Time. Sensors 2023, 23, 771. [CrossRef]
35. McMillan, C.F.; Goosman, D.R.; Parker, N.L.; Steinmetz, L.L.; Chau, H.H.; Huen, T.; Whipkey, R.K.; Perry, S.J. Velocimetry of Fast
Surfaces Using Fabry–Perot Interferometry. Rev. Sci. Instr. 1988, 59, 1–21. [CrossRef]
36. Surtaev, A.S.; Serdyukov, V.S.; Zhou, J.; Pavlenko, A.N.; Tumanov, V.V. An experimental study of vapor bubbles dynamics at
water and ethanol pool boiling at low and high heat fluxes. Int. J. Heat Mass Transf. 2018, 126, 297–311. [CrossRef]
37. Vinogradov, V.E.; Pavlov, P.A. Rate of Bubble Growth at Limiting Superheats of a Stretched Liquid. Heat Transf. Res. 2007, 38,
389–398. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

143
energies
Article
Modeling of Turbulent Heat-Transfer Augmentation in
Gas-Droplet Non-Boiling Flow in Diverging and Converging
Axisymmetric Ducts with Sudden Expansion
Maksim A. Pakhomov * and Viktor I. Terekhov

Laboratory of Thermal and Gas Dynamics, Kutateladze Institute of Thermophysics, Siberian Branch of Russian
Academy of Sciences, Acad. Lavrent’ev Avenue 1, 630090 Novosibirsk, Russia
* Correspondence: [email protected]

Abstract: The effect of positive (adverse) and negative (favorable) longitudinal pressure gradients on
the structure and heat transfer of gas-droplet (air and water) flow in axisymmetric duct with sudden
expansion are examined. The superimposed pressure gradient has a large influence on the flow
structure and heat transfer in a two-phase mist flow in both a confuser and a diffuser. A narrowing of
the confuser angle leads to significant suppression of flow turbulence (more than four times that of
the gas-drop flow after sudden pipe expansion without a pressure gradient at ϕ = 0◦ ). Recirculation
zone length decreases significantly compared to the gas-droplet flow without a longitudinal pressure
gradient (by up to 30%), and the locus of the heat-transfer maximum shifts slightly downstream, and
roughly aligns with the reattachment point of the two-phase flow. Growth of the diffuser opening
angle leads to additional production of kinetic energy of gas flow turbulence (almost twice as much as
gas-droplet flow after a sudden pipe expansion at ϕ = 0◦ ). The length of the flow recirculating region
in the diffuser increases significantly compared to the separated gas-droplet flow without a pressure
gradient (ϕ = 0◦ ), and the location of maximum heat transfer shifts downstream in the diffuser.

Keywords: heat transfer; droplets evaporation; turbulence; droplet-laden flow; confuser; diffuser;
pipe; sudden expansion; RANS

Citation: Pakhomov, M.A.; Terekhov,


V.I. Modeling of Turbulent
Heat-Transfer Augmentation in
1. Introduction
Gas-Droplet Non-Boiling Flow in
Diverging and Converging Two-phase flows in pipes or channels with a backward-facing step (BFS) are often
Axisymmetric Ducts with Sudden used in energy and chemical equipment. They have a rather simple flow geometry and are
Expansion. Energies 2022, 15, 5861. one of the classical types of shear flows, but their flow structure is quite complex. A flow
https://doi.org/10.3390/en15165861 detaches from the sharp edge at the flow SE station, thus forming a region of shear mixing
layer. A large recirculation flow region (a few step heights) develops (see comprehensive
Academic Editor: Gianpiero
reviews [1,2]).
Colangelo
The complexity of modeling flow and heat transfer is exacerbated after BFS in the
Received: 11 July 2022 presence of a longitudinal pressure gradient (LPG) in an expanding (diffuser) or narrowing
Accepted: 9 August 2022 (confuser) subsonic turbulent two-phase flow (see Figure 1). An overview of the state of
Published: 12 August 2022 research on flows in a diffuser or confuser without sudden expansion of a pipe [3] or a
channel [4,5] has been presented. The study of the effect of LPG behind a pipe or channel
with SE on mean and fluctuational flow and heat transfer is an important for mechanical
engineering. There are several studies on the development of separated flows with the
Copyright: © 2022 by the authors.
influence of longitudinal pressure gradient for a single-phase flow, yet only a few of these
Licensee MDPI, Basel, Switzerland.
experimental works concerned the flow in diffusers and confusers with a BFS [6–9].
This article is an open access article
distributed under the terms and
An effect of flow separation in the field of LPG was experimentally evaluated in said
conditions of the Creative Commons
studies. The position of an “upper” duct wall was changed, which caused narrowing or
Attribution (CC BY) license (https://
expansion of a cross-section, whereas the “lower” wall with the SE remained unchanged.
creativecommons.org/licenses/by/ The most detailed structures of the turbulent flow were assessed in [6] using the LDA
4.0/). method along the length of the diffuser channel. The authors measured the profiles of

Energies 2022, 15, 5861. https://doi.org/10.3390/en15165861 144 https://www.mdpi.com/journal/energies


Energies 2022, 15, 5861

averaged longitudinal and transverse velocities and their fluctuations, Reynolds stresses,
length of the recirculation region, triple correlations, and turbulent viscosity. The authors
then compared their experimental and numerical data.

Figure 1. Schematic view of the flow in diffuser (APG, +ϕ), confuser (FPG, − ϕ), and in the separated
flow in pipe sudden expansion (ZPG, ϕ = 0). 1 is the droplet-laden flow.

The experimental results on the effect of favorable pressure gradient (FPG) and adverse
pressure gradient (APG) in a channel behind a BFS on heat transfer and wall pressure
distributions at Reynolds numbers ReH = Um1 H/ν = (0.4–1.2) × 104 were presented in [10].
The diffuser opening angle varied in the range of ϕ = 0–4◦ , and confuser narrowing angle
was varied within ϕ = 0–−7.5◦ . The magnitude of the Nusselt number increases as the
LPG increases for a narrowing channel, and it decreases for the diffuser. The locus of the
heat-transfer maximum moves downstream with diffuser expansion, and shifts upstream
towards the step as the confuser narrows. In [11], a quantitative study assessed the effect of
an APG on mean flow, turbulence, and heat transfer in an axisymmetric diffuser in a pipe
with SE. The literature also presents experimental [12] and numerical [13–18] studies of
fluid flow and heat transfer in single-phase turbulent flows without SE of a pipe or channel
in the presence of APG and FPG for a single-phase flow.
Solid particles addition to a turbulent flow in a BFS have large effect on reduction of
turbulent kinetic energy (TKE) in backward-facing step flow [19]. Droplets evaporation in
turbulent flow behind a BFS [20] or after a pipe with SE [21] causes significant intensification
of heat transfer (by several times in comparison with a single-phase flow). Authors of this
work have published numerical investigations of heat-transfer augmentation in gas-droplet
flows behind a pipe with SE [21]. There are few papers concerning numerical simulation
of gas-liquid flow [22,23] and droplet-laden [24] flows in a converge or divergent channel
without sudden expansion; we know of only one work on the numerical study of heat
transfer in two-phase flows after pipe sudden expansion with LPG [25], where the effect of
evaporation of water droplets on heat transfer in an axisymmetric diffuser was studied.
Heat transfer in turbulent droplet-laden flow with SE with APG and FPG has not been
previously performed. The influence of LPG on flow and heat transfer in the confuser and
diffuser after pipe SE is evaluated in the present study.

2. Mathematical Methods and Numerical Solution


The motion and heat transfer of a two-phase turbulent gas-droplet flow in a pipe with
SE is numerically considered. A sketch of the flow is given in Figure 1. To simulate the
dispersed phase dynamics, the Eulerian approach [26–28] is used. The Eulerian approach
is widely used for the simulations of two-phase confined flows [21,22,29,30]. The system of
axisymmetric stationary Reynolds-averaged Navier–Stokes (RANS) equations accounts
for the effect of vaporizing drops on mean and fluctuational transport processes [21,25].
The set of governing equations both for gas and dispersed phases have been provided in
detail [21,25]. The volume fraction of the droplets is low (Φ1 = ML1 ρ/ρL < 1.2 × 10−4 for
the highest mass fraction studied ML1 = 10%). Drops are rather small (d1 < 100 μm), so
effects of their collisions can be neglected. Gas phase turbulence is predicted using the
elliptical Reynolds stress model [31] by taking the dispersed phase influence on TKE [32].

145
Energies 2022, 15, 5861

Break-up and coalescence of droplets in flow is not taken into account due to their rarity
(Φ1 < 1.2 × 10−4 ) [33]. The Weber number We = ρ(US − U L )2 d/σ << 1 and the bag break-
up are ignored [33,34]. Here, US = U + uS is the gas velocity seen by the droplet, and
uS is the drift velocity between fluid flow and drops [35]. This assumption is applicable
when the pipe cross-section expands for a diffuser. The use of this approach seems less
obvious for a confuser, even when taking into account the preliminary pipe with SE. Effect
of break-up and coalescence in the flow can be neglected due to a low droplet volume
fraction at the inlet according to preliminary author’s estimations.
The technique for numerical implementation of the Eulerian approach for two phases
is described in detail in [21,25]. The numerical solution was obtained using the finite
volume method on staggered grids. The QUICK scheme of third-order ode accuracy was
utilized for solution of convective terms. Central differences of second-order accuracy
were evaluated for diffusion fluxes. Pressure–velocity fields were corrected according to
SIMPLEC procedure.
All simulations were carried out on a “basic” mesh containing 550 × 200 control
volumes (CV) for the diffuser with the largest opening angle, and for the confuser with the
largest convergence angle of 550 × 100. The information about meshes for the confuser and
diffuser is summarized in Table 1. The difference in calculations of the Nusselt number for
the two-phase gas-droplet flow did not exceed 0.1%. A further increase in their number
does not significantly affect the results of numerical calculations. The grid verification for
the case of droplet-laden flow in pipe with SE was presented in [21]. The grid independence
tests for two-phase flows in APG and FPG are given in Figure 2 for the smallest constriction
angle in the confuser and for the largest opening angle in diffuser. The Nusselt number at a
constant wall temperature is determined by dependence:

Nu = −(∂T/∂y)W H/( TW − Tm ),

where Tm is mass-averaged temperature of gas in the considered cross-section.

Table 1. Meshes for two-phase flow in the confuser (FPG) and diffuser (APG).

Flow Type “Basic” “Coarse” “Fine”


Confuser 550 × 100 300 × 50 850 × 150
Diffuser 550 × 200 300 × 100 850 × 300

The convergence criteria for all residual levels in this study were up to 10− 5 . The
differences in Nusselt number and gas-phase kinetic energy of turbulence for gas-droplet
APG and FPG flows were up to 10− 6 .
The model was validated against experimental results on the flow and heat transfer
for the single-phase axisymmetric diffuser downstream of a pipe with SE. The difference
between our predictions and measured results of previous experiments did not exceed 15%.
These results were given in our previous paper [25], but this comparison is not presented
here. We did not find measured or numerical results concerning the study of an APG
or FPG of gas-droplet flow in a pipe or duct with sudden expansion. We performed the
comparisons with experimental two-phase droplet-laden mist and solid particle-laden
turbulent flow behind the BFS and a pipe with SE. These results were published in a
previous paper [21] but are not included here. We believe that the validation analysis of
two-phase solid particle-laden and droplet-laden flows behind backward-facing step or
pipe sudden expansion without LPG have been fully completed.

146
Energies 2022, 15, 5861

(a)

(b)

Figure 2. Grid independence tests for confuser at ϕ = −1◦ (a) and diffuser at ϕ = 5◦ (b).

3. Results and Discussion


The primary concern of this study was shown the effect of diffuser opening and con-
fuser narrowing angles on the characteristics and heat transfer in the two-phase mist with
vaporized water droplets after the pipe with SE. Drop diameter and mass fraction decreased
due to evaporation both in the axial and radial directions after the flow detachment section.
The diffuser expanding angle was ϕ = 0–5◦ and the confuser convergence angle
ϕ = 0–−3◦ . The pipe diameter before SE was 2R1 = 20 mm, after SE it was 2R2 = 60 mm,
and the step height was H = 20 mm. The computational domain after pipe expan-
sion was 25H = 0.5 m. Mass-average air velocity before separation was Um1 = 15 m/s,
and the Reynolds number was ReH = HUm1 /ν ≈ 2 × 104 . The wall temperature was
TW = const = 373 K, and the temperatures of air and droplets at the inlet were T1 =TL1 = 293 K.
Water droplets were added to a single-phase air turbulent flow at the inlet, and their initial
velocity was set constant over pipe cross-section: UL1 = 0.8Um1 . Inlet droplet size was
constant d1 = 1–100 μm, and mass fraction ML1 = 0.01–0.1. The Stokes number in mean
motion was Stk = τ/τ f = 0.03–3, where τ f = 5H/Um1 is the turbulent time macroscale [19,20].
Here,τ = ρ L d2 /(18ρνW ) is the particle relaxation time, W = 1 + Re2/3 L /6 and
Re L = |US − U L |d/ν is the dispersed-phase Reynolds number. The Stokes number
StkK = τ/fK = 0.2–20, where τ K is the Kolmogorov timescale. While the value of inter-
facial velocity in our previous works [21,25] was based only on the average velocity of the
carrier phase, it is based on the actual value in the present study.

3.1. The Wall Friction and Pressure Coefficients


 2 
The distributions of wall friction coefficient C f /2 = τW / ρUm1 and pressure coeffi-
 2 
cient CP = 2( PW − P1 )/ ρUm1 along the length of diffuser and confuser in gas-droplet

147
Energies 2022, 15, 5861

flow with variation of expansion and contraction angles are shown in Figure 3. Here, τ W is
the wall friction; PW , P1 are the mean static pressures on the wall in considered and inlet
cross-sections. The data for the flow after a pipe with SE, without an effect of LPG (ϕ = 0◦ ),
are also shown in this figure for comparison.

(a)

(b)

Figure 3. The evolution of pressure CP (a) and wall friction Cf (b) coefficients along the axial
coordinates in ZPG (ϕ = 0◦ ), diffuser (ϕ > 0◦ , APG), and confuser (ϕ < 0◦ , FPG). ML1 = 0.05, d1 = 30 μm.

The distributions of non-dimensional pressure coefficients along axial coordinates with


the development of a separated flow in the diffuser and confuser are shown in Figure 3a. In
the diffuser, directly behind the flow separation point, a negative pressure region is formed,
with a length of x/H ≈ 7. The pressure coefficient also increases with an increase in the
diffuser opening angle, which can be attributed mainly to flow deceleration. The zone
with pressure attenuation is formed in the confuser directly behind the flow detachment
cross-section, and its length axial direction is x/H = 5–7. With growth of the confuser
convergence angle, the presence of a significant region of pressure attenuation is observed,
and the absolute value of pressure attenuation increases noticeably as the convergence angle
increases (more than 5.5 times at ϕ = −2◦ ). Obviously, the main reason for a significant
pressure decrease in confuser is flow acceleration. The wall friction coefficient Cf decreases
significantly (several times over) with growth of APG, and a sharp increase in the flow
recirculation zone is observed (see Figure 3b). With the increase in the magnitude of
FPG, the wall friction coefficient increases noticeably (almost doubling) after the zone of
flow relaxation.

3.2. The Flow Structure in Confuser (FPG) and Diffuser (APG)


Profiles of the mean axial velocity, temperature, and turbulent kinetic energy of the
gas phase in a cross-section at x/H = 15 are shown in Figure 4. The predictions are carried
out for different values for the diffuser (APG, ϕ > 0◦ ), the confuser (FPG, ϕ < 0◦ ), and in the

148
Energies 2022, 15, 5861

separated flow behind sudden expansion of the pipe (ϕ = 0◦ ). A large effect of two-phase
flow detachment with a zero pressure gradient (ZPG) and with FPG and APG on the mean
axial velocity distributions is revealed in two-phase flow. Obviously, the increase in the
diffuser opening angle leads to a reduction of gas velocity in the core zone (see Figure 4a).
It should be noted that in a cylindrical duct, as well as at small diffuser opening angles
(ϕ ≤ 2◦ ), the separated flow is reattached in cross-section (x/H = 15) and the flow is relaxed.
Air velocity and the velocity gradient in the radial direction in the core region increase in
the confuser.

(a)

(b)

(c)
Figure 4. Mean streamwise velocity component (a), temperature (b), TKE (c) of the gas phase in
ZPG (ϕ = 0◦ ), confuser (FPG, ϕ < 0◦ ), and diffuser (APG, ϕ > 0◦ ). The results for the confuser are the
dashed lines, for the diffuser are the solid curves, and the separated flow with ϕ = 0◦ are the bolded
lines. ML1 = 0.05, d1 = 30 μm.

Gas temperature distributions Θ = ( T − TW )/( T0 − TW ) over the pipe radius depend,


to a lesser extent, on the longitudinal pressure gradient rather than on distributions of the

149
Energies 2022, 15, 5861

axial gas velocity (see Figure 4b). Here, T0 and TW are gas phase temperatures on a pipe
axis and on a wall. The slightly changing diverging angle of the diffuser (ϕ ≤ 2◦ ) and the
converging angle of the confuser (ϕ ≥ −20 ) have little effect on the gas phase temperature in
droplet-laden flow. The temperature increases for the diffuser and decreases for the confuser.
This leads to heat-transfer enhancement in the confuser and heat-transfer suppression in
the diffuser. These conclusions qualitatively concur with the results of simulations [11] for
a single-phase flow in a diffuser behind a pipe with SE. Gas temperature becomes lesser for
gas-droplet flow compared to the case at ϕ = 0◦ .
Turbulent kinetic energy (TKE) of the gaseous phase is significantly enhanced (by
up to two times over) by an increase in the diffuser opening angle (see Figure 4c). The
TKE of the gas phase is calculated for an axisymmetric flow using a known formula:
2k = u2 + v2 + w2 ≈ u2 + 2 v2 . This is not an effect of the dispersed phase;
it is known that the presence of a finely dispersed phase suppresses the carrier-phase
turbulence in the separated flow, both behind the BFS [19,20] and with sudden expansion
of the pipe [21,22]. Particles or droplets are involved in the mean gas movement and a part
of the turbulent energy of a carrier flow is spent on this process [19,32]. The maximum
kinetic energy of turbulence is observed in the mixing layer, and the same phenomena
were found for the gas-droplet flow in the pipe with sudden expansion at ZPG (ϕ = 0◦ ) [21].
This effect was shown previously, in our recent study of an axisymmetric diffuser with
a sudden pipe expansion [25]. An increase in the LPG in an axisymmetric diffuser with
pipe SE causes additional flow turbulization. The maximum value of the turbulent kinetic
energy of the carrier phase for a confuser decreases almost twice as much compared to the
turbulence level of a separated two-phase flow at ZPG. Such a significant TKE suppression
of the carrier phase cannot be explained only by the effect of the dispersed phase.
The transverse distributions of mean axial water droplet velocity UL /UL1 (a), drops in
temperature Θ L = ( TL − TL,max )/( TL,0 − TL,max ) (b), and the mass fraction ML /ML1 (c) of
dispersed phase in the confuser and diffuser in a pipe with SE are presented in Figure 5.
Here TL , TL , and TL ,max are the droplet temperature, the droplet temperature on pipe axis,
and maximum droplet temperature in corresponding cross-section, respectively.
With growth of the confuser convergence angle, a significant increase in the longitudi-
nal averaged velocity of droplets occurs (by more than double at ϕ = −2◦ as compared to
the separated flow at ZPG) (see Figure 5a). The droplet temperature profile has a qualita-
tively similar form for all three types of ducts (ZPG, APG, and FPG) studied previously
(see Figure 5b). On the whole, droplet temperature distributions are similar to those for
the gas phase. The maximum value of droplet mass fraction is obtained in the axial region
of the pipe, and the minimum value is obtained in its near-wall region (see Figure 5c).
The simulations for droplets’ mass fractions ML1 > 10% were not successful due to the
possible effect of droplets deposition in reality. Most likely, the distribution of the mass
fraction of droplets is qualitatively similar to those for ML1 = 10%, but there are quantitative
differences. It is also necessary to take into account the effect of droplet deposition on the
wall from a two-phase flow, and the possible formation of liquid spots and films on the wall
surface. The influence of droplet deposition on transport processes and heat transfer are
not taken into account for our numerical results obtained for ML1 = 5%. Obviously, for high
values of the droplets’ mass fraction at the inlet, it is necessary to account for the influence
of the deposition process and the entrainment of liquid droplets into the droplet-laden flow
from the liquid film or spots.

3.3. The Effect of LPG on the Mean Parametrs of the Two-Phase Mist Flow
A significant increase in the length of recirculating area xR is observed in two-phase
flow in diffuser (see Figure 6). The locus of the heat-transfer peak xmax moves in the
downstream direction. A slight increase in the flow recirculation region is shown for
small expanding angles (ϕ ≤ 1◦ ), and the position of the heat-transfer maximum is close
to the locus of the reattachment point of two-phase flow. The coordinate of xmax moves
downstream by almost double (ϕ = 5◦ ) in comparison with the case of ϕ = 0◦ . The presence

150
Energies 2022, 15, 5861

of FPG leads to a reduction in flow recirculation area and the coordinates of xmax move
upstream by about 30–35% compared to the case of ϕ = 0◦ . The significant displacement of
flow reattachment points in the diffuser and confuser is caused by deformation of the gas
phase velocity profile due to the effect of LPG.

(a)

(b)

(c)
Figure 5. Mean axial velocity componet (a), temperature (b), and mass fraction (c) of the dispersed
phase. The results for confuser (FPG, ϕ < 0◦ ) are the dashed lines, for the diffuser (APG, ϕ > 00 ) are
the solid curves, and the separated flow with ϕ = 0◦ are the bold lines. ML1 = 0.05, d1 = 30 μm.

The TKE of the carrier phase increases almost two times over in the diffuser at ϕ = 5◦
as compared with the case without longitudinal pressure gradient ϕ = 0◦ . Changing the
confuser convergence angle causes suppression of the level of turbulence more than three
times over. The heat transfer decreases significantly with expanding of diffuser opening
angle (almost by a factor of 1.5 as compared to the separated flow in the pipe at ϕ = 0◦ ).

151
Energies 2022, 15, 5861

For the confuser, an increase in the relative value of the maximum heat transfer is observed
at ϕ = −3◦ (by approximately 20%). The heat transfer for the confuser (FPG) case has the
greatest value, and the diffuser (APG) has the smallest.

Figure 6. Effect of LPG on recirculating length xR , location of heat transfer maximum xmax , value
of maximal Nusselt number Numax , and maximum of TKE kmax in gas-droplet flow in pipe SE.
ML1 = 0.05, d1 = 30 μm.

The effect convergence (confuser) and divergence (diffuser) on Nusselt number dis-
tributions along the axial coordinate for the separated flow (ZPG), confuser (FPG), and
diffuser (APG) are shown in Figure 7. Initially, for two-phase mist flows with APG and
FPG, the attenuation of heat transfer rate is observed. This is typical both for both types of
flows and for the case of gas-droplet flow in pie with SE at ϕ = 0◦ . Then there is a sharp
increase in heat transfer with the achievement of a maximum heat transfer. In the zone of
flow relaxation, the observed decrease in Nusselt number is similar to a single-phase flow.

Figure 7. Nusslet numbers distributions along streamwise coordinate in ZPG (ϕ = 0◦ ), FPG (ϕ < 0◦ ),
and APG (ϕ > 0◦ ). ML1 = 0.05, d1 = 30 μm.

The influence of water droplets’ mass concentration on the maximal magnitude of heat
transfer Numax for a diffuser and confuser after pipe SE is presented in Figure 8. For all
types of flow behind the pipe with SE at ϕ = 0◦ , in the diffuser (ϕ = 2◦ ), and in the confuser
(ϕ = −2◦ ), an increase in the maximum heat-transfer value (up to 75% in a single-phase
airflow) was obtained with increasing mass fraction of droplets to ML1 = 10%. The heat
transfer in confuser enhances compare to the diffuser and for two-phase separated flow
with ZPG at ϕ = 0◦ .

152
Energies 2022, 15, 5861

Figure 8. The magnitude of maximal heat transfer in the diffuser (ϕ = 2◦ ), confuser (ϕ = −2◦ ) and
separated flow (ϕ = 0◦ ) vs water droplets mass concentration. d1 = 30 μm.

4. Conclusions
The numerical results of the effects of favorable and adverse longitudinal pressure
gradients on the flow and heat transfer augmentation, in a droplet-laden flow in a pipe
with SE, are presented. Elliptical second-moment closure was used to predict the gas phase
turbulence with taking into account the effect of droplets presence. While this study does
not have a direct application, it shows potential ways to control the turbulence level and
to enhance heat transfer performance in APG and FPG flow behind a backward-facing
step. Thus, these data may be of interest for various practical applications. The scope of the
model’s use is limited the inlet droplet diameter d1 = 100 μm and their initial mass fraction
ML1 ≤ 10%. This can be explained by noting that the model does not take into account the
formation and evolution of a liquid film on the pipe wall, as drops break up and coalesce.
The presence of flow expansion (diffuser) and constriction (confuser) of pipe with SE
shows significant effect on the mean and fluctuational flow characteristics, and heat transfer
in an axisymmetric. The increase of the confuser constriction angle causes considerable
reduction of the pressure coefficient. The length of the flow recirculating area noticeably
shortens compared to the gas-droplet flow behind the pipe with SE at angle ϕ = 0◦ , and
the point of maximum of heat transfer slightly shifts downstream. The heat transfer
augmentation and the suppression of turbulence in a two-phase flow in a confuser are
mainly due to the FPG. The large growth of flow recirculating area (up to 3.5 times at
ϕ = 5◦ ) compared to the gas-droplet flow downstream of pipe SE at ϕ = 0◦ is obtained. The
expansion of the diffuser leads to reduction of the wall friction coefficient. Two-phase flow
does not reattach to the wall at angle ϕ = 5◦ . Points of flow reattachment and maximum
heat transfer are significantly shifted downstream by an increase in the opening angle of
the diffuser. The significant heat-transfer suppression (by up to 1.5 times) and turbulence
production (by up to two times) are observed for the two-phase mist flow in a diffuser.

Author Contributions: Conceptualization, M.A.P. and V.I.T.; methodology, M.A.P. and V.I.T.; Investi-
gation, M.A.P.; data curation, M.A.P. and V.I.T.; formal analysis, M.A.P. and V.I.T.; writing—original
draft preparation, M.A.P. and V.I.T.; writing—review and editing, M.A.P. and V.I.T.; resources, M.A.P.
and V.I.T.; project administration, V.I.T. All authors have read and agreed to the published version of
the manuscript.
Funding: This work was financially supported by the grant of the Russian Science Foundation
(project code 21-19-00162).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

153
Energies 2022, 15, 5861

Nomenclature
d droplet diameter
H step height
ML mass fraction
Nu = −(∂T/∂y)W H/( TW − Tm ) Nusselt number
ReH = Um1 H/ν Reynolds number
Stk = τ/τ f mean Stokes number
T temperature
U average velocity vector
Ui , Uj mean gas velocities components
US = U + uS gas velocity vector seen by the droplet
uS drift velocity between fluid flow and drops
We = ρ(US − U L )2 /σ Weber number
x streamwise coordinate
xmax location of heat-transfer maximum
xR reattachment length
Subscripts
0 single-phase fluid (air) flow
1 initial condition
L liquid
m mean
max maximal value
W wall
Greek
Φ volume fraction
λ thermal conductivity
ρ density
ν kinematic viscosity
τ particle relaxation time
τW wall shear stress
ϕ diffuser opening angle (ϕ > 0) or confuser (ϕ < 0) narrowing angle
Acronym
APG adverse pressure gradient
BFS backward-facing step
FPG favorable pressure gradient
LPG longitudinal pressure gradient
CV control volume
SE sudden expansion
ZPG zero pressure gradient

References
1. Eaton, J.K.; Johnston, J.P. A review of research on subsonic turbulent flow reattachment. AIAA J. 1981, 19, 1093–1100. [CrossRef]
2. Chen, L.; Asai, K.; Nonomura, T.; Xi, G.N.; Liu, T.S. A review of backward-facing step (BFS) flow mechanisms, heat transfer and
control. Thermal Sci. Eng. Progr. 2018, 6, 194–216. [CrossRef]
3. Klein, A. Review: Effects of inlet conditions on conical-diffuser performance. ASME J. Fluids Eng. 1981, 103, 250–257. [CrossRef]
4. Azad, R.S. Turbulent flow in a conical diffuser: A review. Exp. Therm. Fluid Sci. 1996, 13, 318–337. [CrossRef]
5. Apsley, D.D.; Leschziner, M.A. Advanced turbulence modelling of separated flow in a diffuser. Flow Turbul. Combust. 1999, 63,
81–112. [CrossRef]
6. Driver, D.M.; Seegmiller, H.L. Features of a reattaching turbulent shear layer in divergent channel flow. AIAA J. 1985, 23, 163–171.
[CrossRef]
7. Ra, S.H.; Chang, P.K. Effects of pressure gradient on reattaching flow downstream of a rearward-facing step. J. Aircr. 1990, 27,
93–95. [CrossRef]
8. Iftekhar, H.; Agelin-Chaab, M. Structure of turbulent flows over forward facing steps with adverse pressure gradient. ASME J.
Fluids Eng. 2016, 138, 111202. [CrossRef]
9. Wang, L.B.; Tao, W.Q.; Wang, Q.W.; Wong, T.T. Experimental study of developing turbulent flow and heat transfer in ribbed
convergent/divergent square ducts. Int. J. Heat Fluid Flow 2001, 22, 603–613. [CrossRef]

154
Energies 2022, 15, 5861

10. Terekhov, V.; Dyachenko, A.; Smulsky, Y. The effect of longitudinal pressure gradient on heat transfer in a separated flow behind
a sudden expansion of the channel. Heat Transf. Eng. 2020, 41, 1404–1416.
11. Terekhov, V.I.; Bogatko, T.V. Aerodynamics and heat transfer in a separated flow in an axisymmetric diffuser with sudden
expansion. J. Appl. Mech. Techn. Phys. 2015, 56, 471–478. [CrossRef]
12. Güemes, A.; Sanmiguel Vila, C.; Örlü, R.; Vinuesa, R.; Schlatter, P.; Ianiro, A.; Discetti, S. Flow organization in the wake of a rib in
a turbulent boundary layer with pressure gradient. Exp. Therm. Fluid Sci. 2019, 108, 115–124. [CrossRef]
13. Leont’ev, A.I.; Lushchik, V.G.; Reshmin, A.I. Heat transfer in conical expanding channels. High Temp. 2016, 54, 270–276. [CrossRef]
14. Lushchik, V.G.; Reshmin, A.I. Heat transfer enhancement in a plane separation-free diffuser. High Temp. 2018, 56, 569–575.
[CrossRef]
15. Lushchik, V.G.; Makarova, M.S.; Reshmin, A.I. Laminarization of flow with heat transfer in a plane channel with a confuser. Fluid
Dyn. 2019, 54, 67–76. [CrossRef]
16. Sakhnov, A.Y. Local laminarization within the mild pressure gradient flow over the heated wall. Int. J. Heat Mass Transf. 2021,
165, 120698. [CrossRef]
17. Sakhnov, A.Y.; Naumkin, V.S. Velocity overshoot within the accelerated subsonic boundary layer over the heated wall. Int. J. Heat
Mass Transf. 2020, 161, 120249. [CrossRef]
18. Hajaali, A.; Stoesser, T. Flow separation dynamics in three-dimensional asymmetric diffusers. Flow Turbul. Combust. 2022, 108,
973–999. [CrossRef]
19. Fessler, J.R.; Eaton, J.K. Turbulence modification by particles in a backward-facing step flow. J. Fluid Mech. 1999, 314, 97–117.
[CrossRef]
20. Hishida, K.; Nagayasu, T.; Maeda, M. Augmentation of convective heat transfer by an effective utilization of droplet inertia. Int. J.
Heat Mass Transf. 1995, 38, 1773–1785. [CrossRef]
21. Pakhomov, M.A.; Terekhov, V.I. Second moment closure modelling of flow, turbulence and heat transfer in droplet-laden mist
flow in a ver pipe with sudden expansion. Int. J. Heat Mass Transf. 2013, 66, 210–222. [CrossRef]
22. Ahmadpour, A.; Noori Rahim Abadi, S.M.A.; Kouhikamali, R. Numerical simulation of two-phase gas–liquid flow through
gradual expansions/contractions. Int. J. Multiph. Flow 2016, 79, 31–49. [CrossRef]
23. Kopparthy, S.; Mansour, M.; Janiga, G.; Thévenin, D. Numerical investigations of turbulent single-phase and two-phase flows in a
diffuser. Int. J. Multiph. Flow 2020, 130, 103333. [CrossRef]
24. Golubkina, I.V.; Osiptsov, A.N. Compressible gas-droplet flow and heat transfer behind a condensation shock in an expanding
channel. Int. J. Therm. Sci. 2022, 179, 107576. [CrossRef]
25. Pakhomov, M.A.; Terekhov, V.I. Gas-droplet flow structure and heat transfer in an axisymmetric diffuser with a sudden expansion.
J. Appl. Mech. Techn. Phys. 2020, 61, 787–797. [CrossRef]
26. Drew, D.A. Mathematical modeling of two-phase flow. Ann. Rev. Fluid Mech. 1983, 15, 261–291. [CrossRef]
27. Reeks, M.W. On a kinetic equation for the transport of particles in turbulent flows. Phys. Fluids A 1991, 3, 446–456. [CrossRef]
28. Derevich, I.V.; Zaichik, L.I. Particle deposition from a turbulent flow. Fluid Dyn. 1988, 23, 722–729. [CrossRef]
29. Wu, H.; Yang, C.; Zhang, Z.; Zhang, Q. Simulation of two-phase flow and syngas generation in biomass gasifier based on
two-fluid model. Energies 2022, 15, 4800. [CrossRef]
30. Zeng, Y.; Xu, W. Investigation on bubble diameter distribution in upward flow by the two-fluid and multi-fluid models. Energies
2021, 14, 5776. [CrossRef]
31. Fadai-Ghotbi, A.; Manceau, R.; Boree, J. Revisiting URANS computations of the backward-facing step flow using second moment
closures. Influence of the numerics. Flow Turbul. Combust. 2008, 81, 395–410. [CrossRef]
32. Elgobashi, S. On predicting particle-laden turbulent flows. Appl. Scient. Res. 1994, 52, 309–329. [CrossRef]
33. Lin, S.P.; Reitz, R.D. Drop and spray formation from a liquid jet. Ann. Rev. Fluid Mech. 1998, 30, 85–105. [CrossRef]
34. Elgobashi, S. An updated classification map of particle-laden turbulent flows. In IUTAM Symposium on Computational Approaches
to Multiphase Flow, Fluid Mechanics and Its Applications; Balachandar, S., Prosperetti, A., Eds., Springer: Dordrecht, The Netherlands,
2006; Volume 81, pp. 3–10.
35. Mukin, R.V.; Zaichik, L.I. Nonlinear algebraic Reynolds stress model for two-phase turbulent flows laden with small heavy
particles. Int. J. Heat Fluid Flow 2012, 33, 81–91. [CrossRef]

155
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
www.mdpi.com

Energies Editorial Office


E-mail: [email protected]
www.mdpi.com/journal/energies

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are
solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s).
MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from
any ideas, methods, instructions or products referred to in the content.
Academic Open
Access Publishing

mdpi.com ISBN 978-3-7258-0718-5

You might also like