Dissertation Tenzer
Dissertation Tenzer
Dissertation Tenzer
Dissertation
vorgelegt von
aus Heppenheim
Darmstadt 2020
D 17
Tenzer, Fabian Michael:
Heat transfer during transient spray cooling: An experimental and analytical
study
Darmstadt, Technische Universität Darmstadt
Jahr der Veröffentlichung der Dissertation auf TUprints: 2020
Tag der mündlichen Prüfung: 26.03.2020
i
Theoretical models to predict the heat flux during spray cooling are devel-
oped and validated with experimental results. A model for the film boiling
regime accounts for spray and wall properties and predicts the temporal evo-
lution of surface temperature and heat flux. It agrees well with experimental
data.
A theory for the nucleate boiling regime indicates that the heat flux is
limited by the thermal inertia of the substrate material and is not a function
of the spray properties.
Furthermore, the Leidenfrost point is found to be nearly independent of
the spray properties. Instead, it is strongly influenced by the material of
the substrate and fluid: A high thermal effusivity leads to a low Leidenfrost
temperature and vice versa. This influence is captured in a newly developed
theoretical prediction.
ii
Kurzfassung
Die Sprühkühlung zeichnet sich durch eine sehr hohe und gleichmäßig verteilte
Kühlleistung aus. Deshalb wird sie in einer großen Bandbreite von industriellen
Prozessen eingesetzt, wie beispielsweise bei der Kühlung von Hochleistungs-
elektronik, zum Abschrecken während der Metallherstellung oder zur Kühlung
von Schmiedewerkzeugen in der Massivumformung. Die Effizienz der Kühlung
wird von einer Vielzahl von Größen beeinflusst. Dazu zählen unter anderem:
Tropfendurchmesser und -geschwindigkeit, Beaufschlagungsdichte, Oberflä-
chentemperatur, das Spraymedium sowie dessen Temperatur, Material und
Beschaffenheit der Oberfläche und viele mehr. Der gesamte Prozess ist extrem
komplex und bis jetzt existieren nur sehr wenige physikalische Modelle, die
diesen beschreiben. Stattdessen müssen zur Vorhersage des Wärmestroms
empirische Korrelationen verwendet werden, die auf Grund von fehlender
Universalität nicht immer zur entsprechenden Anwendung passen.
iii
Die Kühlung einer schrägen Oberfläche sowie das Kühlen mit heißem Wasser
reduziert die Kühlleistung, während Oberflächenrauigkeit und Verwendung
eines Substrats mit hoher Wärmeleitfähigkeit die Wärmestromdichte erhöhen.
Weiterhin wurden während dieser Arbeit theoretische Modelle zur Vorhersa-
ge der Wärmestromdichte entwickelt und mit den experimentellen Ergebnissen
validiert. Ein Modell für das „film boiling regime“ berücksichtigt den Einfluss
verschiedener Sprays und Substrateigenschaften. Es beschreibt die zeitliche
Entwicklung der Wärmestromdichte und Oberflächentemperatur und stimmt
sehr gut mit den experimentellen Ergebnissen überein.
Weiterhin zeigt eine Theorie für das „nucleate boiling regime“, dass die
Wärmestromdichte durch die thermische Trägheit des Substrats limitiert wird
und nahezu unabhängig von den Sprayeigenschaften ist.
Der Leidenfrostpunkt ist ebenfalls unabhängig von den Sprayeigenschaften.
Stattdessen wird er stark vom Material des Substrats und Fluids beeinflusst:
Ein großer Wärmeeindringkoeffizient erzeugt eine geringe Leidenfrosttempera-
tur und umgekehrt. Dieser Einfluss wird durch die entwickelte Theorie gut
vorhergesagt.
iv
Danksagung
Zuallererst möchte ich mich bei Prof. Dr.-Ing. Cameron Tropea für die Anstel-
lung am Fachgebiet „Strömungslehre und Aerodynamik“ und die Möglichkeit
diese Dissertation anzufertigen bedanken. Ich konnte mich jederzeit auf seine
Zusagen und einen strukturellen Rahmen verlassen, der wenige bis keine
Einschränkungen bedeutete. Weiterhin geht mein ganz besonderer Dank an
Apl. Prof. Dr. Ilia V. Roisman, der mir beim Verständnis der physikalischen
Vorgänge während der Sprühkühlung und deren Modellierung essenziell gehol-
fen hat. Die freudvollen Diskussionen endeten häufig mit neuen Erkenntnissen
und manchmal nur mit noch mehr Fragen. Bei Prof. Dr.-Ing. Eckehard Specht
möchte ich mich vielmals für die Übernahme des Korreferats dieser Arbeit
bedanken.
Weiterhin möchte ich mich bei der Deutschen Forschungsgemeinschaft
(DFG) bedanken, die diese Forschung im Rahmen des Sonderforschungsbe-
reichs SFB-TRR 75 erst finanziell möglich gemacht hat. Ohne die tat- und
sachkräftige Förderung durch den Industrieverband Massivumformung e.V.
wäre dies ebenfalls nicht möglich gewesen. Vielen Dank hierfür und die vielen
Projekttreffen der Patengruppe, die immer wieder aufs Neue den Bezug zur
industriellen Anwendung sichergestellt haben.
Ich danke weiterhin den Studenten Julian Hofmann, Till Kaupe, Minh
Khang Pham und Sebastian Wolter, die durch ihre Mühe und ihren Ehrgeiz
diese Arbeit wesentlich unterstützt haben. Ein ganz besonderer Dank geht
an die Griesheimer Jungs, die durch zahlreiche Diskussionen, gemeinsame
Kochaktionen zur Mittagspause, allerlei Unfug während und nach der Arbeit
und ein ganz besonderes Arbeitsklima das Arbeiten in der Exklave lebens-
wert, äußerst produktiv und sozial verträglich gemacht haben. Dazu zählen:
Andreas Bauer, Alexander Beck, Jan Breitenbach, Sebastian Brulin, Johannes
Feldmann, Julian Hofmann, Till Kaupe, Johannes Kissing, Maximilian Kuhn-
henn, Max Luh, Klaus Schiffmann, Benedikt Schmidt, Bernhard Simon und
Martin Stenger. Danke für die immer wiederkehrende gegenseitige Motivation!
v
Weiterhin danke ich dem Werkstatt-Team in Griesheim um Timm Geelhaar,
Joachim Heyl, Ilona Kaufhold und Martin Weiß für die Unterstützung beim
Bau der Versuchsanlage.
Obwohl zuletzt genannt, geht mein allergrößter Dank an meine Familie
und Freunde. Ich möchte meinen Eltern Georg und Marina Tenzer und
meiner Schwester Anne Tenzer für die nicht endende Unterstützung und
den immer währenden Rückhalt danken. Ihr habt durch eure Erziehung und
Verbundenheit den Grundstein für alles gelegt. Mein größter Dank geht an
Meike, die mich unzählige Male bestärkt hat weiterzumachen sowie sich ebenso
oft bahnbrechenden Erkenntnisse über Sprühkühlung anhören durfte, was sie
sicherlich immer brennend interessiert hat. Ohne dich wäre mir vieles deutlich
schwerer gefallen.
Danke,
Fabian
vi
Contents
Abstract i
Kurzfassung iii
Danksagung v
1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Fundamentals of spray cooling . . . . . . . . . . . . . . . . . 4
1.3 Objective and outline of this thesis . . . . . . . . . . . . . . . 10
vii
Contents
6 Leidenfrost point 87
6.1 Influences on the Leidenfrost point . . . . . . . . . . . . . . . 87
6.2 Comparison with literature . . . . . . . . . . . . . . . . . . . 91
6.3 Theoretical prediction of the Leidenfrost point . . . . . . . . 94
Nomenclature 107
Bibliography 111
viii
1 Introduction
The present chapter provides a short introduction into the topic of spray
cooling. The first part gives a motivation, why research and progress in the
field of spray cooling is of interest for a broad community. In the subsequent
section an overview of the literature indicates questions which remain open.
The present chapter ends with the objectives and outline of this thesis. Parts
of this chapter have already been published in Tenzer et al. (2019b) and
appeared in Tenzer et al. (2020).
1.1 Motivation
Spray cooling is a process capable of achieving a very high, nearly uniform
heat flux, and therefore high cooling performance. A comparison of the heat
transfer coefficients of different cooling technologies is shown in Fig. 1.1. Of all
shown technologies spray cooling clearly has the highest cooling performance.
Therefore it is used in various industrial applications, like cooling of micro-
chips and other high powered electronics or electrical parts (Mudawar, 2001;
Bar-Cohen et al., 2006; Ebadian & Lin, 2011), cooling of metal products in
metallurgy during quenching processes, in metalworking (Chen & Tseng, 1992),
cooling of tools for hot forging (Pola et al., 2013), of solar panels (Nižetić
et al., 2016; Sargunanathan et al., 2016) and in many other technological
processes.
These completely different applications indicate the complex nature and
versatility of the term spray cooling. A rough classification regarding the
temperature range, range of spray properties and temporal evolution helps
to understand the different features of each application. For example, spray
cooling of high power electronics is performed at rather low surface temper-
atures. The mass flux as well as the sprayed surface is small. The process
is stationary, therefore the surface temperature remains nearly constant. In
contrast, spray cooling for quenching during steel-making exhibits a com-
1
1 Introduction
pletely different behavior. The surface temperatures are very high. The mass
flux and the sprayed surface are large. The process is transient, meaning the
surface temperature continuously decreases with increasing time and therefore
different hydrodynamic and thermodynamic phenomena occur over time at
the surface.
The motivation for this project is focused on cooling during hot forging.
Here metal blanks are heated up to 1260 ◦ C and pressed between forging dies
to reach the desired geometry. This process is used for the production of
various parts that require high strength. Exemplary parts are crankshafts,
gears and other parts used in industrial machinery and consumer products.
Since the heated parts stay in contact with the forging dies, the temperature
of the dies also rise; therefore, to prevent overheating and damage they need
to be cooled. This is most often achieved by spray cooling. By shortening the
necessary cooling time, the entire process time for the production of each part
can be reduced. This directly leads to an increased productivity and economic
efficiency. One goal of the present work is to help to overcome the bottleneck
of spray cooling by achieving a better understanding of the physical processes
which determine the cooling rate and by developing reliable models describing
these processes, eventually leading to shortened cycle times in the cooling
phase.
2
1.1 Motivation
3
1 Introduction
4
1.2 Fundamentals of spray cooling
Figure 1.3: Exemplary evolution of the heat flux as a function of the surface
temperature during spray cooling.
example for empirical correlations can be found in Mudawar & Deiters (1994).
Among the governing parameters are spray properties, like droplet velocity,
droplet diameter, mass flux, or liquid properties (Puschmann & Specht, 2004;
Wendelstorf et al., 2008; Yang et al., 1996; Chen et al., 2002; Estes & Mudawar,
1995). Moreover, experiments of Cebo-Rudnicka et al. (2016) with different
target materials demonstrated that the heat transfer is influenced also by the
thermal conductivity of the surface.
Most of the models for the heat flux and for the critical heat flux are
completely empirical. The main goal of the present study is to develop a
predictive theoretical model for fast transient cooling of a very hot thick
substrate by spray impact. The model should be based on the identification
of the main influencing physical parameters. These influencing parameters
are different for the film boiling regime and for the nucleate boiling regime.
The following subsections address some selected issues and influencing
parameters during spray cooling which are covered to a lesser extent in the
review literature cited above.
5
1 Introduction
The physics of the transition from the nucleate boiling regime to the film
boiling regime at the Leidenfrost point is not yet completely known in that
the Leidenfrost temperature cannot be reliably predicted. Several theoretical
models have been developed based on the hydrodynamic stability analysis of
the vapor/liquid interface Jerome (1960); Zuber (1958); Kakac & Bon (2008)
or thermocapillary stability Aursand et al. (2018). Some authors assume that
the Leidenfrost temperature is determined by the foam limit Spiegler et al.
(1963); Wang et al. (2019b), which is the maximum temperature to which
a liquid can be superheated, or by the limiting minimum vapor thickness
Cai et al. (2020) when it becomes comparable with the surface roughness.
However, the influence of the surface roughness is not yet clearly delineated
and requires further investigations.
During single drop impacts: Numerous studies deal with the Leidenfrost
point during single drop impact (Quéré, 2013; Biance et al., 2003; Castanet
et al., 2015; Tran et al., 2012), showing a difference between the static
Leidenfrost temperature TLs of a sessile droplet and the dynamic Leidenfrost
temperature TLd of an impacting droplet. In these studies the dynamic
Leidenfrost point is often observed as the transition between a wet and dry
rebound. A review on the static Leidenfrost temperature can be found in
Bernardin & Mudawar (1999).
Although, there is an ongoing discussion about the influencing parameters
on this point (Liang & Mudawar, 2017a), there exists some agreement that
the dynamic Leidenfrost point depends on the Weber number. Therefore,
common correlations for the dynamic Leidenfrost temperature in the form of
can be found (Bertola, 2015; Yao & Cai, 1988). Here the dynamic Leidenfrost
temperature is linked to the static one by an influence of the Weber number.
Although Eq. (1.1) is used in many studies, the values found for the empirical
constants C1 and C2 span a large range, depending on which study is quoted.
The studies have in common that C2 < 0.5. This leads to a somewhat diffuse
picture of the physics involved in the Leidenfrost point. Nevertheless, since
C2 < 0.5, Eq. (1.1) indicates that for a sufficiently large Weber number the
6
1.2 Fundamentals of spray cooling
second term in Eq. (1.1) will converge to a constant. Therefore the influence
of the Weber number becomes small and the Leidenfrost temperature is nearly
constant for sufficiently large Weber numbers.
Wang et al. (2019b) performed experiments with various substrate materials
for single drop impacts having a Weber number in the range We = 30 − 120.
The results indicate no influence of the Weber number. Instead different
substrate materials lead to different Leidenfrost temperatures. The authors
conclude that a higher thermal effusivity leads to a higher Leidenfrost tem-
perature. They further compared their experimental data to a Leidenfrost
point model proposed by Bjornard & Griffith (1977) which is based on the
homogeneous nucleation theory. They find a good agreement between their
experimental data and the theory.
In sprays: The studies of Sozbir et al. (2010) and Sozbir et al. (2003) indi-
cate an higher Leidenfrost temperature for higher mass fluxes. Furthermore,
their study shows an increasing Leidenfrost temperature for increased air
velocity of a pneumatic atomizer. The same influence of the mass flux is
reported in Hoogendoorn & den Hond (1974). Here the range of Leidenfrost
temperatures is 350 to 900 ◦ C. The authors also conclude that outcomes
concerning the Leidenfrost point of single droplets cannot be transferred to
the Leidenfrost point in sprays. The same influence of rising Leidenfrost
temperature with rising mass flux is confirmed by Gottfried et al. (1966);
Al-Ahmadi & Yao (2008). Al-Ahmadi & Yao (2008) report a insignificant
influence of the droplet diameter and velocity. They correlated their data for
ṁ = 1.5 − 30 kg/m2 s with the expression
7
1 Introduction
Figure 1.4: Dependence of the Leidenfrost temperature on the spray Weber number
and application of correlation Eq. (1.5). (Adapted from Yao & Cox (2002), with
permission of Taylor & Francis. © 2002 Taylor & Francis.)
ṁD
ReS = and (1.3)
η
ṁ2 D
WeS = , (1.4)
ρσ
TL = 1400We0.13
S (1.5)
8
1.2 Fundamentals of spray cooling
9
1 Introduction
We can summarize: First, the fluid temperature strongly influences the heat
flux during spray and jet cooling. Second, it has an impact on the Leidenfrost
point for spray and jet cooling. Third, there is no influence on the Leidenfrost
point of a single drop.
10
1.3 Objective and outline of this thesis
11
12
2 Experimental setup and methods
This chapter describes the different experimental setups and measurement
techniques used for the acquisition of the experimental data in this study,
with some exemplary results shown for illustration.
Parts of this chapter have already been published in Tenzer et al. (2018,
2019a,b). In addition Wolter (2018) supported the construction of the experi-
mental setup and performed first measurements.
13
2 Experimental setup and methods
Figure 2.1: Schematic of experimental setup. Left: Heat flux measurements with
thermocouples and visualization with HS-camera. Right: Spray characterization
with phase Doppler measurement system and patternator. (Adapted from Tenzer
et al. (2019b), with permission of Cambridge University Press. © 2019 Cambridge
University Press.)
the type of atomizer, the supply pressure of the atomizer and the distance
between the atomizer and the heated target, different kinds of sprays can
be produced. Geometric constraints limit the distance between the atomizer
and target to 50 − 350 mm. The entire setup is completely automated and
controlled by a PC running Labview, which is also used for acquiring and
recording the data.
A more detailed description of this equipment follows in the subsequent
sections of this chapter.
14
2.2 Spraying test rig
Figure 2.2: Detailed sketch of the atomizer and water supply system used for spray
generation.
atomizer (9) is separated from the supply line by a check valve (8). The supply
line is connected to the recirculation line through a directional valve (7). A
movable shutter (10) is located directly below the orifice of the atomizer. This
device is able to collect all spray that exits the atomizer. It is quickly pushed
into or out of the spray stream by a pneumatic cylinder. The shutter is opened
only when the spray has reached its steady state. This avoids larger fluid
ligaments, which are typical for the unsteady starting phase of an atomizer,
from reaching the heated surface or patternator.
The setup allows the water to circulate through the entire system when the
directional valve is in open position. The check valve remains closed since the
pressure upstream is lower than its opening pressure and therefore no water
leaves the atomizer. During this circulation phase the circulating water heats
all components of the system to the desired water temperature to ensure a
constant spray water temperature during spraying. When the directional
valve is closed the pressure upstream of the check valve rises until finally the
check valve opens and the water leaves the atomizer as a spray. The check
valve creates a pressure loss of 0.2 bar which is registered during operation.
However, since the same spraying system is used for all experiments the
pressure drop does not have to be considered further. The desired atomizer
operation pressure is kept constant using a controller, which controls the
speed of the gear pump based on the pressure sensor data.
15
2 Experimental setup and methods
Table 2.1: Overview of atomizers used for the experiments and their operational
parameters.
Atomizer Bore diame- Spray Operation Mass flow
ter, mm angle pressure, bar rate, kg/h
Lechler 490.403 1.25 45◦ 1.5 − 10 51 − 135
Lechler 490.603 2 45◦ 2−5 186 − 266
Spraying Systems 0.79 30◦ 1.5 − 10 46 − 102
3002.5
Spraying Systems 0.79 30◦ 2 − 10 28 − 64
3001.4
The atomizers are mainly one component, pressure swirl, full-cone type,
which are available from various companies. For this work four different atom-
izers are chosen, with the aim of obtaining a broad span of spray parameters.
The designation and typical operational parameters of the four atomizers are
summarized in Table 2.1. The bore diameter and spray angle are taken from
the manufacturer’s data sheet. The operation pressure and the mass flow
rates were acquired during the phase Doppler measurements. For this purpose
a Coriolis mass flow meter (Optimass 7400 C from Krohne) was installed
upstream of the atomizer. Due to the limited capacity of the gear pump, the
pressure range for the largest atomizer is limited.
Preliminary experiments were performed using a pneumatic atomizer, which
uses a secondary air stream to enhance the atomization process. The desig-
nation of that atomizer is Lechler 136.115. Generally a pneumatic atomizer
produces a spray having smaller and faster droplets and a lower mass flux.
Since only a few experiments were performed with this atomizer, a detailed
description of its spray properties have been omitted below.
16
2.2 Spraying test rig
et al. (1996). The relevant system parameters are summarized in Table 2.2.
Laser power and other operational parameters were adjusted according to
Araneo & Tropea (2000) to ensure the best measurement results and validation
rates. The interchangeable receiving aperture mask was selected depending on
the droplet size range of each atomizer and operating pressure. More details
about the setup and experimental procedure can be found in Wolter (2018).
2.2.2 Patternator
A measurement device for the acquisition of the local mass flux is called a
patternator (Lefebvre, 1988). During this work two different custom made
mechanical patternators were used. Both devices use tubes to collect the
amount of spray water m that hits a part of an surface area A in a given time
t. Since the collection area A is small compared to the overall sprayed surface,
the mass flux ṁ used in this work is a local quantity. It can be calculated by
m
ṁ = . (2.1)
At
The difference of the two patternators employed in this study lies in the
method with which the amount of collected water was measured. The first
device measures the mass of the collected liquid by weighing. The water is
gathered by 17 small tubes of inner diameter of 4 mm, placed in a row with a
17
2 Experimental setup and methods
distance of 6 mm from one another. These tubes feed the water to containers
which are then weighed. A more detailed description can be found in Wolter
(2018).
Due to the necessity of weighing each container, these measurements take
a long time. To overcome this drawback an automated patternator was
developed. The collection surface was increased to 6 × 6 mm2 which is still
small enough to capture any spray inhomogeneities. The weighing is replaced
by a capacitive measurement of the fluid level in each of the cells. Details
of the working principle and the construction can be found in Kaupe (2017)
and Hofmann (2019). With the addition of an automated drain system the
measurement time is greatly reduced.
The measurement axis of the patternator was aligned with the position at
which the temperature measurements were performed, ensuring that both
quantities are acquired at the same position.
18
2.2 Spraying test rig
Figure 2.3: Mean droplet diameter and velocity as a function of radial position.
Position r = 0 mm corresponds to the geometrical center below the orifice.
(Adapted from Tenzer et al. (2019b), with permission of Cambridge University
Press. © 2019 Cambridge University Press.)
In Fig. 2.3b the mean droplet velocity in the main flow direction U is shown
at the same positions and distances as considered before. In the center the
velocity attains a maximum and decreases at the outer positions. An increased
pressure results in an increased droplet velocity.
19
2 Experimental setup and methods
is the mass flux and the previously mentioned differences are rather small,
we conclude that it is reasonable to use the droplet diameter and velocity
acquired under free stream conditions as input data for the present work.
Details of the experiments and results regarding the influence of the target
can be found in Wolter (2018). Perhaps even more significant is however the
fact that the measurement of the mass flux using the patternator does mimic
exactly the conditions prevailing with the heated target.
Exemplary results of mass flux In Fig. 2.4 the mass flux at different
radial positions is shown for different operating pressures of the atomizer
at 100 mm distance to the nozzle. The measurements are performed with
the patternator. The plot for 2 bar demonstrates the expected trend of a
full-cone atomizer: the highest mass flux appears in the center with a strong
decrease towards the outside. This behaviour changes with increasing pressure.
Normally one would expect the shape to remain nearly constant with a shift
towards higher mass fluxes. From the measurements we see instead a decrease
of the mass flux in the center and an increase at the outer region. The
entire spray pattern changes with rising pressure and the atomizer appears to
behave more like a hollow cone atomizer. This means that the main spray is
concentrated in a ring of the radius 40 to 50 mm. This behaviour is typical of
a pressure swirl atomizer.
20
2.2 Spraying test rig
Figure 2.5: Droplet velocity versus diameter for various atomizer in the center
region below the atomizer. Operation pressure and distance also varied but not
indicated in the figure.
It is interesting that the mass flux at the center depends only weakly on
the injection pressure. For the nozzles used in the experiments the pressure
influences mainly the mass flux distribution in the outer ring of the spray
cross-section.
The local mass flux spans the range ṁ = 0.5 − 29.5 kg/m2 s. An isolated
trend for the dependence of the local mass flux on operation conditions cannot
be identified, as shown in Fig. 2.4. Nevertheless, decreasing the distance
between the atomizer and target obviously increases the mass flux.
21
2 Experimental setup and methods
Since the outcomes of isothermal drop wall interaction are governed by the
droplet velocity and diameter, it is reasonable to introduce the dimensionless
parameters, Reynolds and Weber number, as follows:
U D10
Re = (2.2)
νf
ρf D10 U 2
We = (2.3)
σf
Here νf is the kinematic viscosity, ρf the density and σf the surface tension of
the liquid. The range of parameters used for this experimental study can also
be described as Re = 403 − 1732 and We = 26 − 549.
For the sake of completeness, it is mentioned that the spray parameters of the
pneumatic atomizer span the ranges: ṁ = 0.2 − 3.3 kg/m2 s, D10 = 12 − 24 µm
and U = 12.7 − 23.4 m/s or Re = 200 − 434 and We = 32 − 125.
These ranges of spray parameters are also used for the cooling experiments.
The heated target The heated surface of the spray impact target is the top
end of a circular cylinder (diameter dT = 100 mm and height hT = 53.2 mm).
A detailed drawing is found in Fig. 2.8. During the course of this study
different targets having the same geometry were used. They were built from
various materials and have different surface conditions. Table 2.3 summarizes
the different targets and their material properties. Here ρ, cp and λ are the
density, specific heat capacity and thermal conductivity respectively. Each
22
2.3 Cooling test rig
Figure 2.6: Heated target equipped with watertight housing and ventilation slit.
Table 2.3: Overview of targets and material parameters used of cooling experiments.
Name Material Surface ρ, kg/m3 cp , J/kgK λ, W/mK
condi-
tion
Stainless steel 1.4841 smooth 7900 542 16
Stainless steel 1.4841 rough 7900 542 16
Hot-work 1.2365 smooth 7800 510 33
tool steel
Nickel 2.4068 smooth 8900 500 63
value corresponds to the calculated mean between 100 and 500 ◦ C. The smooth
surface condition corresponds to a mirror polished surface made by lapping
and polishing. The average roughness of the these surfaces is < 0.03 µm. The
rough surface was produced by sandblasting, resulting in an average roughness
of ≈ 10 µm.
Both stainless steel and nickel exhibit good resistance against corrosion and
oxidation. The majority of experiments were performed with stainless steel.
This target endured more than 300 cooling experiments. Since the surface
suffered from thermo-shock, changes of the surface, like less brightness, are
visible. Despite this, the repeatability between the first and last experiments
is good.
23
2 Experimental setup and methods
In contrast, the target built of hot-work tool steel exhibits much more
corrosion. Therefore only a few exemplary experiments were performed to fill
the gap of thermal properties between stainless steel and nickel.
• Typically insulation materials are porous and therefore not suitable for
contact with water.
24
2.3 Cooling test rig
Figure 2.7: Heated target and additional components inside the watertight housing.
25
2 Experimental setup and methods
The target (1) rests with its collar on the top cover. A metallic seal (10) is
installed between the target collar and stainless steel cover to watertight this
interface. The seal is a metallic high temperature duct seal type “E-Ring”. It
has the benefit of not requiring high sealing forces. These seals are usually
custom built parts, except some stock models used in the aviation industry.
For simplicity this setup is built using such a stock seal (AS1895/7-450 ),
which can be purchased from aviation aftermarket suppliers.
High temperature disk springs (11) built by Vinsco Spring Limited are
installed at the bottom of the heated target to ensure a nearly constant sealing
force, independent of the thermal expansion. Theses disk springs act between
the threaded rods (4) and a compression structure. The structure consists of
a plate (12) built of Kelutherm 800 M and two hollow cylinders (13) built of
Kelutherm 700. A spacer (14) limits the compression of the seal to its range of
operation. Due to this construction, the insulation material is only subjected
to pressure, and the heated parts can freely expand and contract.
The free space inside the cylinder (8) is filled with Insulfrax S insulating
blanket for insulation propose. Additionally, the inside is filled with Argon
gas to protect all materials and especially the copper disk from oxidation.
Due to the seals, the housing is nearly gas tight and the pressure inside can
be raised to ≈ 0.2 bar above atmosphere. Therefore, failure of any seal or part
can be detected by an increased volume flow of the Argon gas. Thus, any
damage of the watertight housing is detected before water enters the inside
and leads to further destruction.
A temperature controller based on Labview is implemented in the measuring
PC and controls the desired temperature inside the copper disk by taking
temperature readings from a thermocouple inside the copper disk as an input
value and by adjusting the heating power.
26
2.3 Cooling test rig
Figure 2.8: Sketch of the sectional view of the heated target showing the thermo-
couple positions. Dimensions are in mm.
27
2 Experimental setup and methods
Figure 2.9: Sketch of the thermocouples having an open measuring tip which is
aligned with the shield.
for any radial distribution of the heat flux. The problem is assumed to be
two-dimensional, axisymmetric and having adiabatic boundary conditions at
the curved surface area.
The IHCP is solved using a procedure published by Monde et al. (2003).
Furthermore, Monde and coworkers at Saga University developed and made
available an inverse heat conduction analysis tool “Invers2D”. In the present
work, this tool is used for calculating the heat flux and surface temperature
from temperature readings of the thermocouples.
The procedure is described in detail in Woodfield et al. (2006). It solves
the problem in Laplace space. Therefore the two-dimensional heat conduction
equation is transformed into Laplace space. The measured temperature inside
the target is approximated in time as a series of half-power polynomials and in
space direction as Fourier-Bessel series. After solving the problem in Laplace
space, an inverse Laplace transform leads to the solution.
The thermocouples are type K, class 1, with 0.5 mm shield diameter. The
measuring tip is open and aligned with the shield, as shown in Fig. 2.9. This
results in a fast rise time.
The holes, inside which the thermocouples are placed, are produced using
the spark erosion technique. The resulting hole diameter is 0.6 mm. The
sensors are bonded inside the holes using a thermally high conductive adhesive
(Aremco Ceramabond 569 VFG) to ensure good thermal contact.
The temperatures are sampled at a sample rate of 95 Hz using National
Instruments NI 9212 thermocouple input modules attached to a National
Instruments cRio 9074.
28
2.3 Cooling test rig
Figure 2.10: Field of view for configuration with Tamron 180 mm macro optical
lens.
Figure 2.11: Field of view for configuration with Questar QM-100 long distance
microscope objective lens.
Since there are only qualitative and no quantitative outcomes based on the
visualization, a long description of the visualization system is omitted and
only a brief summary of the hardware is given.
The observation system consists of a high-speed camera equipped with two
different lenses and a back light illumination source. The camera is either a
Phantom v12.1 or Phantom v2012. Depending on the desired field of view
a Tamron 180 mm macro optical lens (far field) or a Questar QM-100 long
distance microscope (near field) is used. An exemplary image of the surface
and scale for both configurations is shown in Figs. 2.10 and 2.11. In addition
to the different magnification the depth of view is much smaller in the case of
the long distance microscope. This helps to mask the droplets that are outside
of the focal plane and therefore prevent these droplets from obstructing the
optical accessibility.
29
2 Experimental setup and methods
The back light illumination consists of a light source and a diffusor plate.
In the case of the objective lens, a high power LED is used as light source.
For the configuration with the long distance microscope an infrared Cavilux
HF laser is necessary to achieve short illumination times to prevent motion
blur while maintaining high light intensity.
The visualization and heat flux measurements are temporally matched and
therefore visual observations can be directly associated with the instantaneous
local heat flux and target surface temperature.
30
3 Typical results of transient spray
cooling
This chapter describes the experimental results obtained in this study. First
some general outcomes of the heat transfer during spray cooling are shown
and afterwards these are connected to the visualization of the hydrodynamics
involved in spray impact. The subsequent parts show the influence of various
parameters on the heat transfer.
Unless otherwise stated, the underlying experiments of the following results
were performed using the stainless steel target having a smooth surface.
Parts of this chapter have already been published in Tenzer et al. (2018,
2019a,b). Furthermore Pham (2018) and Hofmann (2019) supported the
experimental investigation and performed some of the measurements.
31
3 Typical results of transient spray cooling
The typical evolution of the temperature field inside the substrate, the
surface temperature and the heat flux are illustrated in Figs. 3.1 to 3.4. The
initial wall temperature is Tw0 = 450 ◦ C. The spray parameters for this case
are: ṁ = 2.9 kg/m2 s, D10 = 55 µm and U = 10.3 m/s.
The evolution of the temperature profiles T (z) is calculated by solving the
inverse heat conduction problem and is shown in Fig. 3.1 for different times
t. The z coordinate coincides with the spray axis while the position z = 0
corresponds to the wall surface, where the spray impact takes place. The
bottom of the heated plate corresponds to z = 53.2 mm.
The vertical dashed lines in Fig. 3.1 indicate the position of the thermocou-
ples at z1 = 0.5 mm and z2 = 20 mm. The temperature measurements of the
thermocouples are used as the input data for the solution of the inverse heat
conduction problem.
The corresponding time series of the surface temperature and heat flux are
plotted in Fig. 3.2. In Fig. 3.3 the heat flux is plotted as a function of the
surface temperature, which shows good agreement to the schematic diagram
in Fig. 1.3. The vertical dashed lines in Figs. 3.2 and 3.3 correspond to the
boundaries of the boiling regimes: film boiling, transition boiling and nucleate
boiling regime, which are described in more detail in Figs. 3.4 and 3.5.
32
3.1 Measurements of the heat flux
Figure 3.2: Typical temporal evolution of the surface temperature Ti (t) and in-
stantaneous local heat flux q̇(t). The vertical dashed lines indicate the boundaries
between film boiling, transition and nucleate boiling. (Adapted from Tenzer
et al. (2019b), with permission of Cambridge University Press. © 2019 Cambridge
University Press.)
Figure 3.3: Typical evolution of the instantaneous local heat flux q̇(t) as a function
of the surface temperature Ti (t). The vertical dashed lines indicate the boundaries
between film boiling, transition and nucleate boiling.
33
3 Typical results of transient spray cooling
Figure 3.4: Spray cooling regimes at different surface temperatures Ti (t) around the
Leidenfrost point. a) Measured heat flux q̇ as a function of surface temperature Ti ;
b) visualized spray impact in the film boiling regime; c) at the Leidenfrost point,
characterized by the first appearance of liquid patches; d) the fast expansion
of the liquid spots. (Adapted from Tenzer et al. (2019b), with permission of
Cambridge University Press. © 2019 Cambridge University Press.)
34
3.2 Visualization of the spray impact
In Fig. 3.4b, 3.4c and 3.4d images of spray impact and hydrodynamic
phenomena at the surface captured at different instants after the spray cooling
begins are shown. The time instants chosen for these images are marked on
the graph in Fig. 3.4a.
The hydrodynamic phenomena visualized in Fig. 3.4b, 3.4c and 3.4d are
each different, since they correspond to different drop and spray impact
thermodynamic regimes. In Fig. 3.4b each drop impact onto the wall leads
to its break up, formation of multiple secondary droplets (Roisman et al.,
2018) and rebound. The contact time is short and there is no remaining
wetting of the surface. As a result the heat flux is low, which is typical for
the film boiling regime. At the Leidenfrost point, illustrated in Fig. 3.4c, a
few impacting drops start to wet and spread on the surface - a part of the
surface is covered by initial liquid patches. At the next instant the area of the
wet patches increases, Fig. 3.4d, and the heat flux starts to rapidly increase.
This phenomena correspond to the transition boiling regime.
Similar phenomena in the film and transition regimes are observed in
Fig. 3.5b and 3.5c, respectively. The images are of higher contrast, since
the mass flux of the spray is smaller in the experiment shown in Fig. 3.5.
In Fig. 3.5a the measured heat flux is plotted as a function of the surface
temperature measured during continuous spraying. In the illustrated case the
spray properties are: ṁ = 0.9 kg/m2 s, D10 = 43 µm and U = 9.8 m/s. In this
example the target is initially heated to a surface temperature of 306 ◦ C.
Shortly before the point where the maximum heat flux is achieved a large
portion of the surface is wetted by a liquid water film, as shown in Fig. 3.5c.
Small nucleation bubbles form, grow and collapse. The heat flux increases
significantly, since the wetted area of the substrate increases rapidly. Heat
goes into the overheating of the liquid (sensible heat) and into the formation
of bubbles.
At the instant corresponding to the critical heat flux, Fig. 3.5d Ti = 170 ◦ C,
the surface area is almost completely wetted by liquid. The appearance of a
corona of an impacting drop is clear evidence that the drop impacts onto a
liquid film (Yarin et al., 2017).
35
3 Typical results of transient spray cooling
36
3.2 Visualization of the spray impact
in Figs. 2.3a, 2.3b and 2.4. All data and images are taken from the same
continuous experiment but at different instants in time. The time t, starting at
the moment when the shutter opens, is indicated in the caption of each figure.
The images at the top show pictures of the surface where the spray impact
takes place. The corresponding radial distribution of surface temperature
is indicated in the plot below. The circular markers at the radial positions
r = 3.5, 7, 10.5, ..., mm correspond to the positions of the thermocouples.
Since the radial positions of picture and plot are matched, a direct comparison
between hydrodynamics at the surface and corresponding surface temperature
is possible. The vertical dashed line in the picture at the top indicates the
area at the surface where the hydrodynamic behavior changes. It is detected
from visual observation of the conditions at the surface. The vertical dashed
line in the plot at the bottom is simply an elongation and thus corresponds to
the same radial position r. In the plot at the bottom, the horizontal dashed
line indicates the Leidenfrost temperature, which is determined from the
minimum heat flux in the area around r = 14 mm.
In Fig. 3.6a on the left-hand side of the dashed vertical line a part of
the liquid of the impacting droplets remains at the surface after the drop
has impacted. A persisting wetting of the surface is present and results in a
continuous increase of the amount of water at the surface. This is equivalent to
the previously mentioned nucleate boiling regime. The right-hand side of the
dashed vertical line corresponds to the film boiling regime: Impacting droplets
spread at the surface, atomize and form secondary droplets. However, there
is no remaining wetting at the surface. The surface remains dry. Therefore,
the vertical dashed line corresponds to the area where the Leidenfrost point is
located, which marks the transition between dry and wet droplet impact. The
previously discussed transition boiling regime cannot be explicitly identified
in this representation. Although not clearly identifiable, this regime also takes
place in the area of the dashed line together with the Leidenfrost point.
These observations at the surface correspond to the temperature distribution
in the plot at the bottom. On the right side of the vertical dashed line, in
the film boiling regime, the temperature is nearly constant. It approximately
corresponds to the calculated Leidenfrost temperature (horizontal dashed
line). On the left side of the vertical dashed line the temperature strongly
decreases. Therefore, the Leidenfrost point corresponds to the point at which
the temperature starts to decrease.
37
3 Typical results of transient spray cooling
(a) t = 75.2 s
(b) t = 75.5 s
(c) t = 76 s
38
3.2 Visualization of the spray impact
(d) t = 76.2 s
(e) t = 77.1 s
Figure 3.6: Position of the wetting front that follows the Leidenfrost point at
different instants of time t. Top image shows the hydrodynamics at the surface.
Bottom plot indicates the surface temperature distribution along the radial coor-
dinate r. The vertical dashed line corresponds to the approximate position of the
wetting front. The horizontal dashed line indicates the Leidenfrost temperature.
Figs. 3.6b to 3.6d show the same phenomena. Comparing these figures
with each other, the movement of the wetting front is visible. Starting at
r = 18 mm in Fig. 3.6a, the wetting front is at r = 17 mm in Fig. 3.6b, at
r = 14 mm in Fig. 3.6c and finally at r = 12 mm in Fig. 3.6d. The movement
of the wetting front is a result of the inhomogeneous distribution of the spray
properties. As shown in Section 2.2.3, the sprays in this investigation, which
were produced using high water supply pressure, tend to have a higher mass
39
3 Typical results of transient spray cooling
flux at the outer edges of the spray. Therefore at the outer region (left-hand
part of the surface in the image) the cooling is more intensive due to the
higher mass flux. At the inside (right-hand part of the surface in the image)
the mass flux is lower resulting in a less intensive cooling. This results in a
wetting front that moves from the outside to the inside with increasing time.
A slightly different behavior can be observed in Fig. 3.6e, which is still the
same experiment, but at some instants later. Here a second wetted spot is
visible at the inside of the target (right-hand side of the dashed line), which
is again a result of the inhomogeneous spray distribution. At this instant the
wetting pattern at the surface directly corresponds to the radial distribution
of the mass flux, as shown in Fig. 2.4. In summary, we identify a moving
wetting front that starts at the outer region. It moves towards the inside and
meets at approximately r = 7 mm with a second wetting front that forms in
the center and moves towards the outside.
The previously shown temporal evolution of the wetting front at the surface
is of exemplary nature. Depending on the atomizer, its supply pressure or the
distance between atomizer and surface, different wetting patterns are possible.
For example in the case of a low atomizer supply pressure, the wetting starts
in the center and moves towards the outside, due to the highest mass flux
located in the center.
Figure 3.7 summarizes the different hydrodynamic phenomena that occur
during spray impact at a hot surface: Film boiling features single drop impacts
onto a dry surface. At the Leidenfrost point a small amount of liquid begins
to remain on the surface immediately after the drop impact. A first wetting
occurs. During the transition boiling regime the amount of liquid wetting the
surface increases. Finally, in the nucleate boiling regime a continuous liquid
film is formed, which covers the entire surface.
40
3.3 Influence of various parameters
Figure 3.7: Sketch of the hydrodynamics occurring during spray impact. For
clarity, the height of the heated target is much smaller than the one used in the
experiments. (Adapted from Hofmann (2019).)
41
3 Typical results of transient spray cooling
42
3.3 Influence of various parameters
(a) Heat flux against time. (b) Surface temperature against time.
Figure 3.8: Influence of mass flux ṁ on the heat transfer. (Adapted from Tenzer
et al. (2019b), with permission of Cambridge University Press. © 2019 Cambridge
University Press.)
43
3 Typical results of transient spray cooling
(a) Heat flux against surface temperature. (b) Heat flux against time.
Figure 3.9: Influence of the mean droplet velocity on the heat transfer.
(red curve) results in a higher heat flux, compared to the lower velocity (blue
curve). The Leidenfrost point is slightly different, but the overall trend of the
curves is comparable. The corresponding plot showing the dependence of the
heat flux on time is found in Fig. 3.9b.
Due to the previously mentioned interdependence of the spray parameters,
it was not possible to generate sprays with different droplet diameters, but
the same velocity and mass flux. Therefore an explicit presentation of the
individual influence of the droplet diameter is not possible.
Although for illustration propose only one set of experimental data is
shown here, generally the same outcomes are found in the other experiments.
Furthermore, the influence of both single drop quantities, droplet diameter
and velocity, is rather small compared to that of the mass flux.
44
3.3 Influence of various parameters
(a) Heat flux versus surface temperature. (b) Heat flux versus time.
45
3 Typical results of transient spray cooling
Uβ = cos(β)U, (3.1)
46
3.3 Influence of various parameters
(a) Heat flux against surface temperature. (b) Heat flux against time.
where U is the mean velocity measured by the phase Doppler system. There
is no reason for a changing droplet diameter and the mass flux is assumed to
be constant.
In Fig. 3.13 the scaled heat flux in the film boiling regime q̇β /q̇ is plotted
as a function of the scaled impact velocity Uβ /U for different experiments.
Here the indices β correspond to the different impact angles (0◦ , 15◦ , 30◦
and 45◦ ) and q̇ is the heat flux obtained during wall-normal impact (0◦ case).
The spray impact parameters of the different experiments are indicated in the
figure by different colors. All experimental data collapse onto one straight line
having a slope of unity. This indicates that the heat flux in the film boiling
regime of an oblique spray impact scales with the wall normal spray impact
velocity. Knowing the heat flux of the normal impact q̇, the heat flux for the
oblique impact can be calculated according to
Uβ
q̇β = q̇ = q̇ cos β. (3.2)
U
47
3 Typical results of transient spray cooling
Figure 3.13: Scaled heat flux in the film boiling regime q̇β /q̇ as a function of the
scaled impact velocity Uβ /U for different spray impact angles β. The different
colored lines represent different kinds of sprays.
Figure 3.14: Heat flux dependence on the surface temperature for different spray
water temperatures. The spray parameters are: ṁ = 0.9 kg/m2 s, D10 = 43 µm
and U = 9.9 m/s. (Adapted from Tenzer et al. (2019a).)
48
3.3 Influence of various parameters
In Fig. 3.14 the heat flux q̇ is plotted as a function of the surface tempera-
ture Ti . Since the surface is continuously cooled and the heating is turned
off at the beginning of the spraying process, the curves in this figure can
be read as increasing time (t) from higher to lower surface temperatures.
The spray parameters are ṁ = 0.9 kg/m2 s, D10 = 43 µm and U = 9.9 m/s.
For all different spray water temperatures Tf0 the commonly known boiling
regimes can be identified: film boiling regime, where the heat flux linearly
slightly decreases, at high surface temperatures from the beginning of the
experiment (Ti = 450 ◦ C) until Leidenfrost point is reached (Ti ≈ 340 ◦ C).
After passing the Leidenfrost point the transition boiling regime follows, which
is characterized by a strong increase of the heat flux. It ends at the critical
heat flux point (maximum heat flux) at Ti = 215 ◦ C. After that point the
nucleate boiling regime follows, after which the heat flux again decreases.
We can conclude that an increase of the spray water temperature yields
a decrease of the heat flux at all surface temperatures and therefore for all
boiling regimes. Especially in the film boiling regime and at the critical heat
flux, the heat flux is obviously decreased by the higher water temperature. We
can further identify the non-linearity of the effect: the drop of the heat flux
between Tf0 = 30 and 50 ◦ C is higher than that between Tf0 = 50 and 79 ◦ C.
For further investigation, in Fig. 3.15 we expand the variation of the different
water temperatures and only show the film boiling regime and the beginning of
the transition boiling regime. The previously mentioned outcomes can easily
be confirmed. In addition, the non-linearity results in nearly no difference
at high water temperatures (between Tf0 = 69 and 79 ◦ C). The strongest
influence is at low water temperatures and continuously decreases towards
higher water temperatures.
Figure 3.15 also leads to a deeper insight into the Leidenfrost point: For
the lowest water temperature the Leidenfrost temperature is TL ≈ 345 ◦ C and
for the highest water temperature TL ≈ 350 ◦ C. The results of the other spray
water temperatures are in-between these boundaries. Keeping in mind that
the Leidenfrost point is not a distinct point, but more a region where the slope
of the curve changes, it is obviously difficult to determine the Leidenfrost
point. Therefore, and because of the deviation of only 5 ◦ C being very small
compared to the change of the water temperature (50 ◦ C), the influence of the
water temperature on the Leidenfrost point is also weak or even nonexistent.
49
3 Typical results of transient spray cooling
Figure 3.15: Heat flux dependence on the surface temperature for different spray
water temperatures. Only film boiling regime, Leidenfrost point and beginning of
transition boiling regime are shown. Spray parameters are equivalent to Fig. 3.14.
(Adapted from Tenzer et al. (2019a).)
In Fig. 3.16 similar results for a different spray are shown. The inserts show
representative pictures of the surface at the corresponding surface temperature
and are taken from videos capturing the spray impact. This time the spray
properties are ṁ = 2.8 kg/m2 s, D10 = 54 µm and U = 10.6 m/s. Especially
the strong influence of the higher mass flux results in an overall higher
heat flux compared to Fig. 3.14. The general trends are comparable to the
previously mentioned observations: Higher spray water temperature results
in a lower heat flux. The Leidenfrost temperature is only weakly influenced,
TL ≈ 320 ◦ C.
From the insert on the right we can identify single drop impact onto a dry
wall as the leading spray impact mechanism during film boiling at temperatures
above Leidenfrost. The insert on the left indicates that at temperatures below
Leidenfrost, first small patches of liquid remain and wet the surface (marked
in red). From the videos we can identify the Leidenfrost point as the instant,
when for the first time a very small wetted spot remains shortly after the
impacting drop left the surface, which was already discussed in the previous
sections. Also, shown only for one spray water temperature, these observations
are independent of the water temperature. We can conclude that different
spray water temperatures do not change the hydrodynamics at the surface.
50
3.3 Influence of various parameters
Figure 3.16: Heat flux dependence on the surface temperature for different spray
water temperatures. The inserts show the conditions at the surface at the
corresponding surface temperature. The shape of the remaining liquid patch
at the surface below Leidenfrost is marked in red. The spray parameters are:
ṁ = 2.8 kg/m2 s, D10 = 54 µm and U = 10.6 m/s. (Adapted from Tenzer et al.
(2019a).)
51
3 Typical results of transient spray cooling
Compared to Fig. 3.14 the influence of the water temperature on the heat
flux, as well as the non-linearity of the effect, are stronger. In addition, the
Leidenfrost temperature is lower, TL ≈ 320 ◦ C in contrast to TL ≈ 345 ◦ C in
Fig. 3.16.
These results are in strong contrast to the findings in literature. Both
experimental studies of Nimi et al. (2012) and Xu & Gadala (2006) conclude
a strong decrease of the Leidenfrost temperature for hot water. In our
experiments this decrease is not observed, which is confirmed by data from both
measurement techniques (inverse heat conduction based on thermocouples
and visual observation). From the visual observation we identify single drop
impact in the film boiling regime. This leads to the comparison of our spray
experimental results to those experimental results of single drop impact:
The Leidenfrost temperature of a single drop is not influenced by the water
subcooling (Bernardin & Mudawar, 1999), which is in accordance to our
results. The Leidenfrost point in sprays and that of single drops share this
common feature.
The previously described different Leidenfrost temperature for both series of
experiments must be caused by the different sprays of the experiments, since
nothing else was changed. The visual observations indicate single drop impact
and no film building above Leidenfrost. Therefore we deduce no influence
of the mass flux on the Leidenfrost temperature. The remaining possible
influencing factors are the drop diameter and velocity. At this stage, from
our experiments shown in Figs. 3.14 to 3.16 we identify, that drops having a
larger diameter and being faster result in a lower Leidenfrost temperature.
This is in contrast to the findings of the influence of the drop velocity on the
Leidenfrost temperature of a single drop in the experimental studies of Celata
et al. (2006); Bernardin & Mudawar (2004); Testa & Nicotra (2009). Here a
higher drop velocity results in a higher drop Leidenfrost temperature.
Let us further try to understand the mechanisms leading to the influence
of the water temperature on the heat flux in the film boiling regime. The
impacting drop hits the surface and stays in contact with the wall for a very
short time. Immediately, heat is transferred from the hot wall to the liquid
of the drop by heat conduction. The sensible heat of the drop increases. A
thermal boundary layer develops. When the saturation temperature in the
liquid is reached at the interface between the drop and the wall, a vapor layer
forms and separates the drop from the wall. There is no contact between the
52
3.3 Influence of various parameters
Figure 3.17: Heat flux as a function of the initial spray water temperature. The
data and spray parameters are equal to Fig. 3.15. (Adapted from Tenzer et al.
(2019a).)
drop and the wall anymore. The drop breaks up due to thermal atomization
(Roisman et al., 2018) or inertia. The sensible heat which is necessary to
heat the liquid to saturation temperature, so vapor is generated, is higher for
a cold water drop. This means more heat from the wall goes into sensible
heat of the drop. Therefore, less vapor is generated in the case of cold water
compared to the case of a higher initial water temperature. Less vapor results
in a smaller vapor layer thickness (Breitenbach et al., 2017b) and a smaller
thermal resistance. Therefore the heat flux is higher. Since spray impact in
the film boiling regime is a superposition of single drop impacts, the previously
discussed behavior of single drop impact can be directly transferred to spray
impact.
In Fig. 3.17 we replot the data from Fig. 3.15. This time the heat flux at
different surface temperatures Ti is plotted as a function of the spray water
temperature. For low water temperatures Tf0 < 70 ◦ C the influence is nearly
linear. For higher water temperatures the slope starts to flatten which leads
to non-linearity. The slope is comparable for all surface temperatures Ti .
From the experimental results it is obvious, that the sensible heat plays an
important role during spray impact at temperatures above Leidenfrost.
53
3 Typical results of transient spray cooling
(a) Heat flux against time. (b) Surface temperature against time.
Figure 3.18: Influence of the initial wall temperature Tw0 for a small mass flux
of ṁ = 1.6 kg/m2 s. (Adapted from Tenzer et al. (2019b), with permission of
Cambridge University Press. © 2019 Cambridge University Press.)
In Fig. 3.18 the heat flux and surface temperature are shown as a function
of time for various initial substrate temperatures: 350, 400 and 450 ◦ C. The
spray properties are ṁ = 1.6 kg/m2 s, D10 = 64 µm and U = 8 m/s. To
highlight the transient behavior the temporal axis is limited to t < 50 s. Some
minor influence of the initial temperature on the heat flux in the film boiling
regime can be identified. However, the overall trend of the curves remain
the same. The time shift is a direct result of the different initial substrate
temperatures.
Similar curves, but this time for a larger mass flux, are shown in Fig. 3.19.
The spray properties are ṁ = 9.3 kg/m2 s, D10 = 48 µm and U = 15.6 m/s.
Again the slopes are comparable but the heat flux at the critical heat flux is
higher for the higher initial substrate temperatures. Furthermore, no typical
film boiling regime is visible. This can again be explained by the small time
scales in the film boiling regime which are of the same order as the rise time
of the thermocouples.
54
3.3 Influence of various parameters
(a) Heat flux against time. (b) Surface temperature against time.
Figure 3.19: Influence of the initial wall temperature Tw0 for a large mass flux
of ṁ = 9.3 kg/m2 s. (Adapted from Tenzer et al. (2019b), with permission of
Cambridge University Press. © 2019 Cambridge University Press.)
55
3 Typical results of transient spray cooling
(a) Heat flux against surface temperature. (b) Heat flux against time.
Figure 3.20: Influence of the substrate material on the heat transfer for a constant
rather sparse spray.
in the film boiling regime follow the same trend. Comparing the Leidenfrost
temperature for the three different materials, a negative correlation between
this temperature and the thermal conductivity can be identified. This means
a higher thermal conductivity results in a lower Leidenfrost temperature. In
Fig. 3.20b the corresponding dependence of the heat flux on time is shown.
Here we can again identify the same general slopes for all curves. Since all
curves end when a surface temperature of 100 ◦ C is reached, the curves show
an influence of the substrate material on the cooling time. The experiment
performed with stainless steel takes a shorter time, followed by hot-work tool
steel and nickel. This temporal difference arises mainly during the film boiling
regime. The time between Leidenfrost point, critical heat flux and end of the
experiment is comparable for the different materials.
Fig. 3.21 shows a second exemplary set of experiments having the same
spray parameters but different substrate materials. This time the spray is
much more dense. The properties are ṁ = 11.9 kg/m2 s, D10 = 79 µm and
U = 10.2 m/s. The curve shows obviously different general trends. From
the curve measured with the stainless steel target (blue) no film boiling can
be distinguished. This curve does not exhibit the typical shape and boiling
regimes. As previously stated in Section 3.3.1, this is caused by the limited
temporal response of the measurement system in combination with the short
time until the critical heat flux is reached, as seen in Fig. 3.21b. The curve
56
3.3 Influence of various parameters
(a) Heat flux against surface temperature. (b) Heat flux against time.
Figure 3.21: Influence of the substrate material on the heat transfer for a constant
rather dense spray.
of hot-work tool steel (yellow) slightly resembles the typical trend we know
from previous experiments. However, the timescales are rather short, so that
the data inside the film boiling regime is not considered so reliable. The
experiment performed with nickel (red) clearly shows the typical behavior: We
can identify the film boiling, transition and nucleate boiling regime. The slope
in the film boiling regime is rather steep compared to previously discussed
results. This appears to be connected to the heat flux inside the film boiling
regime: A higher heat flux results in a steeper slope. Furthermore this slope
is highest at the beginning of the experiment, except for the part indicated by
dashed line, where the rise time prevents valid data. It flattens with increasing
time. Since the time in the film boiling regime is long (Fig. 3.21b), this data
is regarded as valid and can be used for further analysis.
57
3 Typical results of transient spray cooling
(a) Heat flux against time. (b) Surface temperature against time.
Figure 3.22: Influence of the surface roughness on the heat transfer, for a small
mass flux of 1 kg/m2 s.
58
3.3 Influence of various parameters
59
3 Typical results of transient spray cooling
(a) Heat flux against time. (b) Surface temperature against time.
Figure 3.23: Influence of the surface roughness on the heat transfer, for a small
mass flux of 3.1 kg/m2 s.
60
4 Heat transfer in film boiling regime
In this chapter a predictive model for the heat flux and surface temperature
in the film boiling regime is developed. Originally the model was derived for
the heated target built of stainless steel, which is discussed in the first part of
this chapter. In the second part the model is validated against experimental
data. Here we also take into account experimental data acquired with other
target materials, to show the universality of the theoretical model.
Parts of this chapter have already been published in Tenzer et al. (2019b).
61
4 Heat transfer in film boiling regime
is
" #
Z t
z
T (t) = Tw0 + A(τ )erfc dτ, (4.3)
2 α(t − τ )
p
0
T (τ )
Z t Z t 0
w
Ti (t) = Tw0 + A(τ )dτ, q̇(t) = − √ √i dτ, (4.4)
0 π 0 t−τ
where
w = (4.5)
p
λρcp
62
4.2 Modelling of temperature and heat flux evolution
in the liquid region and in an expanding thin vapor layer emerging between
the impacting drop and the very hot substrate. The mass flux of the vapor
generated at the lower liquid interface is determined from the energy balance
at this interface. This model has already been validated by comparison with
experimental results from literature (Wendelstorf et al., 2008). Moreover, the
predicted vapor layer thickness agrees well with the direct measurements of
Tran et al. (2012) and Chaze et al. (2019). Predictions for the evolution of
the heat flux q̇(t) also agree with the accurate measurements based on the
infrared technique (Chaze et al., 2019).
The total heat transferred during the impact of a single drop Qsingle is
determined by integration of the heat flux q̇(t) over the “apparent contact
area” during the contact time. It should be noted that the contact area cannot
be based on the drop spreading diameter, since the free lamella in the remote
regions can levitate (Roisman et al., 2018). Therefore the values of D2 and
D/U are used as the scales for the contact area and for the contact duration,
where U and D are the impact velocity and drop diameter, respectively.
This analysis (Breitenbach et al., 2017b) allows the heat flux during spray
impact in the film boiling regime to be predicted
63
4 Heat transfer in film boiling regime
In this study the model (Eqs. (4.6) to (4.8)) is used for prediction of the
evolution of the wall temperature in time. The predictions are then compared
with our experimental data.
For typical experimental conditions of this study, shown in Table 4.1, the
estimated values for w and b defined in Eq. (4.8) are w ≈ 700 and b ≈ 25.
Since b 1 and b2 and w are of the same order of magnitude, the effect of
the dependence of S on the changing temperature Ti in the expression for w
can be neglected. The value of S can be estimated by using Tw0 instead of
Ti (t) in the Eq. (4.8).
Equations (4.4) and (4.6) lead to the following integral equation for the
surface temperature, presented in the dimensionless form
ξ
Θ0 (ζ)
Z
Θ(ξ) + √ dζ = 0, (4.9)
0 ξ−ζ
Ti (t) − Tsat
Θ= , ξ = tπS 2 , ζ = τ πS 2 . (4.10)
Tw0 − Tsat
64
4.2 Modelling of temperature and heat flux evolution
(a) Dimensionless surface temperature Θ(ξ) (b) Dimensionless heat flux Φ(ξ).
and its derivative Θ0 (ξ).
This equation can be solved subjected to the initial condition Θ(0) = 1. The
analytical solution for Θ(ξ) can be represented as a series
∞
X
Θ(ξ) = 1 + ai ξ i/2 , (4.11)
i=1
2 1 4
a1 = − , a2 =, a3 = − 2 , ... ,
π π 3π
2−i Γ(i + 1)
ai+1 = −ai i+1 i+3 (4.12)
Γ 2 Γ 2
where Γ is the gamma function. The series Eq. (4.12) converges on the interval
0 < i < 22. Therefore it is sufficient to calculate Eq. (4.11) in this interval.
The corresponding heat flux q̇ can be estimated using Eq. (4.4)
The solution for the dimensionless surface temperature Θ(ξ), its derivative
Θ0 (ξ) and the dimensionless heat flux Φ(ξ) are shown in Fig. 4.1.
65
4 Heat transfer in film boiling regime
(a) Dimensionless surface temperature Θ(ξ) (b) Dimensionless heat flux Φ(ξ)
(Eq. (4.10)). (Eq. (4.13)).
Figure 4.2: Evolution of the dimensionless quantities for the stainless steel target
as a function of the dimensionless time ξ for experimental data in the film boiling
regime and for the theoretical solution (Eqs. (4.9) and (4.14)). (Adapted from
Tenzer et al. (2019b), with permission of Cambridge University Press. © 2019
Cambridge University Press.)
Stainless steel target In Fig. 4.2a the measured evolution of the dimen-
sionless surface temperature Θ(ξ) as a function of dimensionless time ξ is
plotted for the theoretical prediction (Eq. (4.11)) and for experimental results.
The computation of the theoretical solution is based on the first 50 terms of
the series. It converges on the interval 0 < ξ < 22. The line shown for the
experiments is the mean and the error bars indicate the standard deviation
computed over 49 experiments. For the reduction of the experimental data Ti
and t, χ in (Eq. (4.7)) is fitted to the experimental data using a least square
fit, resulting in χ = 2.2. Only the experiments exhibiting clear film boiling
behaviour are chosen and the experimental data comprising the film boiling
regime are plotted. We skip the experiments showing no film boiling behaviour
because of the previously mentioned limited temporal response of the measure-
66
4.3 Model validation
67
4 Heat transfer in film boiling regime
(a) Dimensionless surface temperature Θ(ξ) (b) Dimensionless heat flux Φ(ξ)
(Eq. (4.10)). (Eq. (4.13)).
Figure 4.3: Evolution of the dimensionless quantities for the hot-work tool steel
target as a function of the dimensionless time ξ for experimental data in the film
boiling regime and for the theoretical solution (Eqs. (4.9) and (4.14)).
Nickel target Fig. 4.4 shows again the validation of the model for the film
boiling regime against experimental data. This time the data is obtained with
the target built of nickel. The parameter χ is again fitted to experiments
yielding χ = 2.2, which is the same as for the stainless steel target. As
previously stated in Section 3.3.5, the nickel target leads to a much larger span
of different sprays which can be used for the cooling experiments. This becomes
particularly obvious from the much larger mass flux. The spray properties
underlying the 61 experiments are: ṁ = 0.6−29.5 kg/m2 s, D10 = 41−117 µm,
U = 5.2 − 23.7 m/s, Tw0 = 300 − 462 ◦ C and Tf0 = 18 − 23 ◦ C.
Again there is a very good agreement between experimental data and
theoretical prediction. Especially the strong decrease of the theoretical pre-
diction for small dimensionless times ξ in Fig. 4.4b is well represented by the
experiments. In Fig. 4.4a some small discrepancy between experiment and
prediction of the dimensionless surface temperature exist. These are caused
68
4.3 Model validation
(a) Dimensionless surface temperature Θ(ξ) (b) Dimensionless heat flux Φ(ξ)
(Eq. (4.10)). (Eq. (4.13)).
Figure 4.4: Evolution of the dimensionless quantities for the nickel target as a
function of the dimensionless time ξ for experimental data in the film boiling
regime and for the theoretical solution (Eqs. (4.9) and (4.14)).
by the long duration of the experiments: The longer the experiments last,
the larger the thermal boundary layer grows. The higher thermal effusivity
supports this increase. Since the developed model is based on the assumption
of a semi-infinite substrate; hence, the long duration experiments are no longer
covered by the model. Nevertheless, the agreement between experiment and
model is still good. In contrast, the dimensionless heat flux is not affected
that much by the violation of the semi-infinity of the target. This is because
the dependence of the heat flux on the surface temperature (the slope in the
film boiling regime in e.g. Fig. 3.3) is rather small for low mass fluxes. Since
only the experiments performed with small mass fluxes are the ones that last
the longest, the influence of the violated assumption of semi-infinity on the
predicted heat flux is rather small.
For pneumatic atomizer For the sake of completeness, Fig. 4.5 shows
the comparison between experimental data and theoretical prediction for
experiments performed with the nickel target and the pneumatic atom-
izer. The spray properties of these 5 experiments cover the following nar-
row range: ṁ = 0.1 − 2.1 kg/m2 s, D10 = 12 − 24 µm, U = 12.7 − 20.8 m/s,
Tw0 = 448 − 450 ◦ C and Tf0 = 20 − 23 ◦ C.
69
4 Heat transfer in film boiling regime
(a) Dimensionless surface temperature Θ(ξ) (b) Dimensionless heat flux Φ(ξ)
(Eq. (4.10)). (Eq. (4.13)).
Figure 4.5: Evolution of the dimensionless quantities for the nickel target spayed
with a pneumatic atomizer as a function of the dimensionless time ξ for experi-
mental data in the film boiling regime and for the theoretical solution (Eqs. (4.9)
and (4.14)).
70
4.4 Application of the model
Examining the initial instants of spray impact (small t in Fig. 4.6a and
high Ti in Fig. 4.6c), the theory leads to a better understanding of the process.
At the very first instants of the spray impact, the heat flux is very high.
Subsequently, it decreases and converges to the well-known linear slope. The
very high heat flux in the first instants can be explained by the fact that
the first droplets impact onto a substrate having no thermal boundary layer,
resulting in an extremely high temperature gradient at the solid-liquid contact
interface. With time the thermal boundary layer in the substrate grows and
the temperature gradient decrease, leading to a near steady state heat flux.
This knowledge is of special interest for pulsed spray cooling, where the
spray is continuously switched on and off. During the period when no spray
impact occurs, heat flows from the deeper part of the substrate towards the
surface and decreases the thermal boundary layer. At the next instant, when
the spray is switched on again, spray impact takes place again at a more
homogeneously heated surface (compared to the case of continuous spraying).
This leads to a heat flux that again starts at a high level. Continuous pulsing
of the spray can therefore increase the amount of heat extracted from the
substrate while using an overall smaller amount of spray liquid.
71
4 Heat transfer in film boiling regime
(a) Heat flux evolution with time. (b) Surface temperature evolution with
time.
Figure 4.6: Exemplary comparison between one experiment and the theoretical
prediction in the film boiling regime. In a) and b) times are only shown up to
shortly after the critical heat flux point is reached.
72
4.4 Application of the model
Figure 4.7: Heat flux as a function of the initial spray water temperature. Com-
parison between experimental data and theoretical prediction. Experimental data
is equivalent to those in Fig. 3.17.
Initial spray water temperature In Fig. 4.7 the application of the theo-
retical prediction for experiments performed with various spray water temper-
atures Tf0 is shown. Here the heat flux, evaluated at the surface temperature
Ti = 370 ◦ C, is plotted as a function of Tf0 . The experimental data is equivalent
to those in Fig. 3.17.
73
4 Heat transfer in film boiling regime
Figure 4.8: Evolution of the heat flux as a function of time for an exemplary
experiment. The parameters of the spray are: ṁ = 1.4 kg/m2 s, D10 = 43 µm
and U = 11 m/s. The dashed vertical lines bound the transition boiling regime.
(Adapted from Tenzer et al. (2019b), with permission of Cambridge University
Press. © 2019 Cambridge University Press.)
74
4.5 Transitional spray cooling regime below the Leidenfrost point
75
76
5 Heat transfer in nucleate boiling
regime
This chapter describes the development and application of a model for the heat
flux in the nucleate boiling regime. First, the underlying assumptions of the
model are presented, followed by the formulation, validation and exemplary
application of the model. Initially the model was developed for the stainless
steel target; however, other target materials are also taken into account.
Parts of this chapter have already been published in Tenzer et al. (2019b).
On the right-hand side of Eq. (5.1) the first term is associated with the
thermal history during the film boiling regime, the second term is associated
with the temperature jump ∆TL during the transition regime at τ = tL , and
the last term is based on the temperature evolution during the nucleate boiling
at times tL < τ < t.
In order to model the heat flux in the nucleate boiling regime the values of
∆TL and the evolution of the surface temperature Ti (τ ) (which is required for
the computation of the time derivative Ti0 (τ )) are necessary.
77
5 Heat transfer in nucleate boiling regime
In the estimation of an upper bound for heat flux during the nucleate boiling
regime of single drop impact the temperature at the wetted part of the wall
interface is approximated by the saturation temperature Tsat (Breitenbach
et al., 2017a).
The nucleate boiling regime is characterized by intensive nucleation and
expansion of vapor bubbles. The temperature in the vicinity of the contact line
of the each expanding bubble is close to the saturation temperature. However,
some overheating of the surrounding liquid is required for the bubble growth.
The heat transfer in the liquid phase during the nucleate boiling regime is
governed by the convection in the liquid flow between the bubbles. The
heat mainly goes into vaporization at the bubble interfaces where T = Tsat .
Therefore, the upper bound for the heat flux during nucleate boiling regime
can be estimated invoking the assumption that the temperature at the wetted
wall interface is Tsat . Using this rough estimation of q̇ the predicted duration
of a single drop evaporation in the nucleate boiling regime agrees very well
with numerous experimental data (Abu-Zaid, 2003; Itaru & Kunihide, 1978;
Tartarini et al., 1999).
In this study the upper bound for the heat flux q̇ during spray cooling is
also estimated, as in the case of a single drop impact, using the assumption
that the substrate temperature is equal to the saturation temperature at
t > tL . The third term on the right-hand side of Eq. (5.1), associated with
the time gradient of the surface temperature at t > tL , can be neglected in
comparison to the effect of the temperature jump at the Leidenfrost point.
The temperature jump during the transition boiling regime can be estimated
as ∆TL = TiL − Tsat .
The heat flux can be obtained from Eq. (5.1), neglecting the value of the
last term, in the form
∞
ξL i 1 i−1
w TiL − Tsat Tw0 − Tsat X
q̇ = √ √ − Sw ai iB ; , ξ 2 , (5.2)
π t − tL 2 i=1
ξ 2 2
where B[·; ·, ·] is the incomplete beta function. Equation (5.2) is valid only for
very fast substrate cooling, when the time interval between the Leidenfrost
point and the point corresponding to the critical heat flux is very short. It
78
5.2 Model validation
2w ∆T 2
T ≡k 2 ≈ t − tL , (5.5)
π q̇(t)
79
5 Heat transfer in nucleate boiling regime
Figure 5.1: Scaling of the heat flux, expressed through T defined in Eq. (5.5),
plotted as a function of time for the stainless steel target. The straight line
has a slope of unity and indicates good agreement between the theory and the
experimental data. (Adapted from Tenzer et al. (2019b), with permission of
Cambridge University Press. © 2019 Cambridge University Press.)
Hot-work tool steel target Similarly, Fig. 5.2 shows the comparison
between theoretical prediction and experiential data for the target built of hot-
work tool steel. The spray properties of these 6 experiments cover the following
narrow range: ṁ = 1.1 − 4.6 kg/m2 s, D10 = 43 − 79 µm, U = 7.8 − 14.3 m/s,
Tw0 = 410 − 449 ◦ C and Tf0 = 19 − 21 ◦ C.
The constant k = 1.05 is the same as in the previous case of the stainless
steel target. Again we can identify very good agreement between theoretical
prediction and experimental data for larger times, when the critical heat flux
has been passed.
Nickel target In Fig. 5.3a the same comparison is shown for the target built
of nickel. The spray properties of these 59 experiments span the following
ranges: ṁ = 0.6 − 30.3 kg/m2 s, D10 = 43 − 114 µm, U = 5.1 − 22.5 m/s,
Tw0 = 300 − 462 ◦ C and Tf0 = 18 − 23 ◦ C. In contrast to the other materials,
we identify good agreement between theory and experimental data only for
small times t < 20 s. At larger times the experiments differ from the theoretical
prediction. This can be explained by the thermal boundary layer inside the
substrate. Since the thermal diffusivity of nickel is significantly larger than
those of stainless steel and hot-work tool steel, the thermal boundary layer
growths much faster for experiments performed with nickel. At some instant
80
5.2 Model validation
Figure 5.2: Scaling of the heat flux, expressed through T , defined in Eq. (5.5),
plotted as a function of time for the target built of hot-work tool steel.
the boundary layer reaches the bottom of the target and the assumption of
semi-infinity is violated. Therefore the theory (Eq. (5.5)) is no longer valid
and experimental data and theoretical predictions differ.
In Fig. 5.3b we show only those experiments where the assumption of semi-
infinity is justifiable. If the temperature difference at the bottom between the
beginning and the end of the experiment is larger than 30 ◦ C, the experiments
are treated as significantly not semi-infinite and are not included in Fig. 5.3b.
It turns out that this constraint leads to basically the same amount of
experimental data that spans across the same temporal range (t < 20 s) as
observed in Fig. 5.3a.
Finally, we observe good agreement between experimental data and the-
oretical predictions as long as the assumption of semi-infinity is valid. Our
experiments show that the heat flux in the fully developed nucleate boiling
regime depends significantly on time. It is estimated as
w Tw0 − Tsat
q̇ ≈ √ √ . (5.6)
π t − tL
Note that the measured values for T deviate significantly from the theoret-
ical predictions at small times, associated with the film boiling and transition
regimes, for which the scaling (Eq. (5.5)) is not applicable.
81
5 Heat transfer in nucleate boiling regime
(a) All experiments, including those where (b) Only experiments, where the target is
the target is not semi-infinite. semi-infinite.
Figure 5.3: Scaling of the heat flux, expressed through T defined in Eq. (5.5),
plotted as a function of time for the nickel target.
The remote asymptotic solution (Eq. (5.6)) is valid only for a semi-infinite
hot substrate and uniform spray. In practice this means that the thickness
and the width of the substrate are larger than the thickness of the thermal
√
boundary layer in the substrate αt.
82
5.4 Application of the model
Figure 5.4: Exemplary comparison between one experiment and the theoretical
prediction in the nucleate boiling regime (Eq. (5.6)). Heat flux evolution with
time.
√
event is hbubble ∼ αw tbubble . This value is approximately 70 µm in our
case. In any case, our measurement system is not able to detect such rapid
temperature fluctuations.
At any given location on the wall surface the temperature jumps to the
saturation temperature each time the contact line propagates through this
position. Our measurements allow estimation of only the averaged value of
the temperature fluctuations above the saturation temperature. Nevertheless,
the contribution of the temperature fluctuations (consisting of positive and
negative jumps above Tsat ) to the time averaged heat flux q̇ is negligibly small.
83
5 Heat transfer in nucleate boiling regime
Although the curve shows the entire experiment, the theoretical prediction
is only valid after the Leidenfrost point is passed. This range corresponds to
t > tL = 40 s. At these first instants the predicted heat flux starts at infinity
and strongly decreases. The agreement between experiment and model is
obviously not good in this interval. Keeping in mind the limited temporal
resolution of the measurement system, the disagreement is not unexpected.
A few instants later, roughly at the critical heat flux (maximum heat flux),
the prediction matches almost perfectly the experimental data. The excellent
agreement persists until the end of the experiment.
In summary, we find excellent agreement between experiment and theoretical
prediction (Eq. (5.6)) after the first instants. Furthermore, the very good
agreement leads to a new understanding of the heat transfer in the nucleate
boiling regime during transient spray cooling: The heat flux is limited by the
thermal effusivity w of the substrate and the initial wall temperature Tw0 .
It depends on time. There is no or only a very weak dependence of the heat
flux on the spray properties. In other words, the heat flux is limited by the
thermal inertia of the material. This means that there is no possibility to
influence the heat transfer in the nucleate boiling regime with different spray
properties. Even the smallest mass flux of all our experiments is sufficiently
large enough to lead to this state of saturation where the material limits the
heat flux even if the spray could lead to a higher one.
The complete heat flux dependence on time for the entire experiment is
calculated by combining the heat fluxes predicted by the models for both
regimes (Eqs. (4.9) to (4.14) and (5.6)) as
(
Sw (Tw0 − Tsat )Φ(ξ) for t < tL
q̇S (t) = (5.7)
√
π
√ −Tsat
w Tw0
t−tL
for t > tL .
As described and justified in Section 4.5 the transition boiling regime is skipped
and replaced by a jump at the Leidenfrost point. Again, the only necessary
addition information is tL = 40 s, which is taken from the experiments. For
the sake of completeness, the temporal evolution of the resulting overall heat
flux expressed by Eq. (5.7) is shown in Fig. 5.5. Obviously, the heat flux in
the transition boiling regime is strongly over-estimated due to replacing this
regime with a jump. From the experimental point of view we are not able
to confirm or disprove whether this jump is reasonable or in which way the
84
5.4 Application of the model
Figure 5.5: Exemplary comparison between one experiment and the theoretical
prediction of the entire cooling process (Eq. (5.7)). Heat flux evolution with time.
85
86
6 Leidenfrost point
The Leidenfrost point corresponds to the instant when the minimum heat flux
is reached and afterwards strongly increases. For this reason it is of major
interest to reliably predict this point in any spray cooling process. Moreover,
it is the last remaining part of the boiling curve which has yet to be described
to finally predict the heat flux during the entire spray cooling process.
The following chapter is dedicated to a better understanding of the Lei-
denfrost point. In the first part the experiential results of this study are
analyzed to identify any possible dependence on the various spray param-
eters. Subsequently, our results are compared to published predictions in
literature. Finally we identify the dependence on the substrate and derive a
new understanding and universal prediction for the Leidenfrost point.
Parts of this chapter have already appeared in Tenzer et al. (2020).
87
6 Leidenfrost point
by a typical linear increasing heat flux in the film boiling regime. Other
experiments not showing this behavior are discarded, since in these cases
the cooling is too intensive and the timescales are too short to be properly
captured by the thermocouples. Although the linear slope is also present
in these discarded experiments, it cannot be detected by the measurement
technique, as previously discussed in Section 3.3.1.
Even though the literature survey in Section 1.2 reveals no clear picture of
the dependencies of the Leidenfrost point, there is general agreement that the
mass flux plays an important role. To check whether this influence is present
in our experimental data Fig. 6.2 shows the Leidenfrost temperature TL as a
function of the local mass flux ṁ. Here the different materials of the target,
stainless steel and nickel, are indicated by different markers and colors. Since
only a few experiments were performed with hot-work tool steel, these results
are not shown here and in the following discussion. Although the data scatters
a lot, there is no dependence of the Leidenfrost temperature on the mass
flux for both materials. This clearly becomes apparent from the experiments
performed with nickel. For the very large range of mass fluxes the Leidenfrost
temperature remains constant. Due to high cooling rate in the experiments
performed with stainless steel, the data only spans a narrow range of mass flux.
But even in this case no influence is visible. Compared to literature we detect
no influence of the mass flux on the Leidenfrost temperature. Additionally,
88
6.1 Influences on the Leidenfrost point
Figure 6.2: Leidenfrost temperature TL as a function of the mass flux ṁ. (as can
be read in Tenzer et al. (2020).)
the Leidenfrost temperature is higher for stainless steel than for nickel. This
leads to the assumption that the Leidenfrost point is dependent on the target
material.
The observations in Section 3.2 and the modeling in the film boiling regime
in Chapter 4 have shown that spray impact in the film boiling regime consists of
a superposition of single drop impacts. Since the Leidenfrost point temporally
follows the film boiling regime and starts to take place during this regime, we
can assume that the Leidenfrost point is connected to the impact of a single
drop.
To check this assumption Fig. 6.3 shows the dependence of the Leidenfrost
temperature on the spray properties: mean drop diameter D10 in Fig. 6.3a
and mean drop velocity U in Fig. 6.3b. In both cases no dependence can be
observed. Despite the scatter, the Leidenfrost temperature is constant and
independent of the mean drop diameter or mean drop velocity of the spray.
The shown data also includes the experiments performed with the pneumatic
atomizer. Due to the smallest drop diameter of ≈ 20 µm, these points are
clearly visible in Fig. 6.3a. In this case the Leidenfrost temperature is also
constant and comparable to the other experiments. It is remarkable that even
for sprays produced by different principles of atomization, the Leidenfrost
temperature is constant.
89
6 Leidenfrost point
(a) Weber number We. (as can be read in (b) Reynolds number Re.
Tenzer et al. (2020).)
90
6.2 Comparison with literature
91
6 Leidenfrost point
Figure 6.5: Comparison between own data and the correlation proposed by Al-
Ahmadi & Yao (2008) (Eq. (1.2)).
Figure 6.5 shows the experimental data of this study in comparison to the
correlation proposed by Al-Ahmadi & Yao (2008) which presumes an influence
of the mass flux (Eq. (1.2)). As discussed before, the data of this study shows
no influence of the mass flux on the Leidenfrost temperature. Therefore
there is obviously no agreement between our experiments and the proposed
correlation. Instead the correlation strongly over-predicts the Leidenfrost
temperature of up to 800 ◦ C.
In Fig. 6.6 the experimental data of this thesis is compared to the model
proposed by Bernardin & Mudawar (2004) (Eq. (1.6)). This model predicts an
influence of the droplet velocity. As can be seen, the Leidenfrost temperature
predicted by the model and the experimental data span approximately the
same range. But again, the experimental data show no influence of the droplet
velocity. In contrast the model forecasts that the Leidenfrost temperature
increases with an increasing droplet velocity. We can identify no agreement
between prediction and our experimental data.
As already shown in Section 1.2, Yao & Cox (2002) plotted the experimental
data of various researchers (Yao & Choi, 1987; Choi & Yao, 1987; Hoogendoorn
& den Hond, 1974; Mizikar, 1970; Ito et al., 1991; Shoji et al., 1984; Cox &
Yao, 1999) as a function of the spray Weber number Wes . They assumed that
the Leidenfrost temperature depends on Wes and fitted a correlation resulting
in Eq. (1.5). Figure 6.7 shows an adaption of their original illustration. To
provide a good comparison to the original illustration, the logarithmic axes
92
6.2 Comparison with literature
Figure 6.6: Comparison between own data and the correlation proposed by
Bernardin & Mudawar (2004) (Eq. (1.6)).
Figure 6.7: Comparison between own data, those of other researchers and the
correlation proposed by Yao & Cox (2002) (Eq. (1.5)). (Adapted from Yao &
Cox (2002), with permission of Taylor & Francis. © 2002 Taylor & Francis.)
93
6 Leidenfrost point
and their scale are kept the same. In addition the experimental data of the
present study is added. The correlation seems to fit quite well to the ensemble
of all data. Taking a closer look at the individual data sets of each group, the
agreement is however much worse. Especially the data of this study (stainless
steel and nickel), is not well described by their correlation. Instead, the
data presented from each study could be considered as being approximately
constant and independent of the spray Weber number.
In summary, there is no agreement between models and empirical correla-
tions published in the literature and the data of this study. All publications
maintain that the Leidenfrost temperature is significantly influenced by spray
properties. This assertion cannot be corroborated by the data from the present
study, in fact the present results indicates that the Leidenfrost temperature
is independent of any spray property.
94
6.3 Theoretical prediction of the Leidenfrost point
95
6 Leidenfrost point
Figure 6.9: A sessile drop on a heated copper surface at different surface tempera-
tures Ti .
a spray (Ito et al., 1991), which are both lower than the limit for superheat of
distilled water. Moreover, neither of these hypotheses can explain the observed
significant influence of nano-structures, surface morphology or wettability of
the substrate on the Leidenfrost temperature, (Auliano et al., 2016; Kruse
et al., 2013; Romashevskiy et al., 2016; Kim et al., 2012; Takata et al., 2005).
Consider however the first stage of heat transfer, just after the contact of
the liquid with the solid substrate. In the nucleate boiling regime the contact
is followed by the emergence of numerous bubbles on the solid substrate as a
result of heterogeneous nucleation. The heat transfer can be described by the
heat conduction in two expanding thermal boundary layers in the liquid and
the substrate respectively. The thickness of the thermal boundary layer is
√
hw ∼ αw t where αw is the thermal diffusivity of the corresponding material,
liquid or solid. This phenomenon is depicted schematically in Fig. 6.8c.
The typical heat flux at the solid-liquid interface in the nucleate boiling
regime is determined by the thickness of the boundary layer in the solid
substrate and by the fact that the temperature of the vapor bubble interface
is close to the saturation temperature Tsat
w (Tw0 − Tsat )
q̇ ∼ √ , (6.1)
t
96
6.3 Theoretical prediction of the Leidenfrost point
where Tw0 is the initial substrate temperature prior to contact with the liquid
and w is the thermal effusivity of the wall material. Equation (6.1) has been
recently validated by comparison with the experimental data for single drops
(Breitenbach et al., 2017a) and sprays (Tenzer et al., 2019b).
At some elevated wall temperature, characterized by intensive heating,
vaporization fronts appear instead of local bubbles and expand along the
substrate with a certain velocity. For a high enough velocity, the front leads
to the formation of a thin, near-wall vapor layer (as shown in the sketch in
Fig. 6.8d). The propagation velocity of the vaporization front, depending on
the heating rate and liquid properties, can be constant (Okuyama et al., 2006;
Avksentyuk & Ovchinnikov, 2000; Stutz & Simões-Moreira, 2013; Aktershev
& Ovchinnikov, 2011) or can grow exponentially (Staszel & Yarin, 2018) in
time.
We can roughly estimate the characteristic velocity Uv ∼ q̇/ρf L of the
vaporization front using Eq. (6.1)
w (Tw0 − Tsat )
Uv ∼ √ , (6.2)
ρf L t
where L is the latent heat of evaporation and ρf is the density of the liquid.
A similar approach has been successfully used to predict the velocities of
secondary drops in the thermal atomization regime (Breitenbach et al., 2018a;
Roisman et al., 2018).
The only available characteristic thickness in the liquid region associated
with the vaporization front is the thickness of the thermal boundary layer in
√
the fluid hf ∼ αf t, where αf is the thermal diffusivity of the fluid. Therefore,
the Weber and the capillary numbers based on the typical vaporization front
velocity (Eq. (6.2)) are singular at t → 0, which means the influence of surface
tension on the formation of the vaporization front is initially negligibly small.
The Reynolds number Rev = ρf hf Uv /νf , in contrast, is finite
√
w αf (Tw0 − Tsat )
Rev = . (6.3)
νf L
The viscous stresses therefore play an important role during the entire process
of the vaporization front formation. The characteristic superheat ∆T ∗ at which
the inertia and viscous terms are comparable can be determined assuming
97
6 Leidenfrost point
νf L
∆T ∗ = √ (6.4)
w αf
In Fig. 6.10 the experimental data for TL − Tsat of sessile drops of different
liquids on various substrates is shown as a function of ∆T ∗ . The liquid
properties for calculation of ∆T ∗ are taken at the saturation temperature.
For data taken from the literature, the substrate materials and their thermal
properties, type of experiment and the resulting Leidenfrost temperature are
summarized in Table 6.1. The thermal properties of the designated substrate
materials are taken from general literature and handbooks. Since these
properties are temperature dependent, they have been calculated at the
corresponding Leidenfrost temperature. In all cited studies the surface is
polished. Since the spray or drop impact occurs at the surface, the thermal
properties of the plating material is used in the case of a plated substrate,
(Yao & Choi, 1987; Choi & Yao, 1987; Yao & Cox, 2002; Bernardin et al.,
1997). Data from Wang et al. (2019a) were measured for different Weber
numbers: those with a large We were treated as spray data, whereas those
with a small We is treated as sessile drops.
98
6.3 Theoretical prediction of the Leidenfrost point
It seems that for a wide range of targets the value of TL − Tsat can be
predicted rather well as TL − Tsat = ∆T ∗ . For wall materials with very
low thermal effusivity the values for the Leidenfrost temperature deviate
significantly from the prediction. In both cases, for water and for ethanol, the
measured wall temperatures are above the critical temperature. In this case
additional physics has to be accounted for.
The corresponding data for spray impact or high Weber number drop
impact are shown in Fig. 6.11. The superheat associated with the Leidenfrost
temperature also follows a linear dependence on ∆T ∗ . The best fit of the
data is T L − Tsat = 1.51∆T ∗ . The Leidenfrost temperature in the impacting
drops increases, since the thickness of the thermal boundary in the liquid and
the propagation of the vaporization front are influenced by the flow in the
spreading drop (Roisman, 2010).
The fact that the value of TL − Tsat is comparable with ∆T ∗ for a range
of different liquids and substrates indicates that the initiation of the film
boiling can indeed be explained by the propagation of the vaporization front
along the substrate, assuming that the wall temperature does not exceed the
critical value for the studied liquid. Note that propagation of the vaporization
front is also influenced by the conditions at the moving contact line. This
99
6 Leidenfrost point
Table 6.1: Overview the thermal properties of the materials and the value of the
Leidenfrost temperature. (as can be read in Tenzer et al. (2020).)
Study Type of Fluid Substrate w , TL ,
experiment material Ws1/2 /m2 K ◦
C
This study Spray Water Stainless steel 8.8501e+03 342
1.4841
This study Spray Water Nickel Alloy 1.7892e+04 286
201
Hoogendoorn Spray Water Stainless steel 9.0193e+03 371
& den Hond 1.4301
(1974)
Shoji et al. Spray Water Nickel 1.8569e+04 321
(1984)
Ito et al. Spray Water Copper 3.6476e+04 134
(1991)
Yao & Choi Spray Water Copper with 1.7831e+04 222
(1987) chrome plating
Choi & Yao Spray Water Copper with 1.7831e+04 253
(1987) chrome plating
Yao & Cox Spray Water Copper with 1.7831e+04 266
(2002) chrome plating
Bernardin Drop Water Copper with 2.7701e+04 225
et al. (1997) chain gold plating
Tran et al. Drop Water Silicon wafer 9.7544e+03 480
(2012)
Wang et al. Drop Water FeCrAl 6.5664e+03 445
(2019a)
Wang et al. Drop Water Sintered SiC 1.5733e+04 350
(2019a)
Wang et al. Drop Water Zr-4 5.1511e+03 531
(2019a)
100
6.3 Theoretical prediction of the Leidenfrost point
101
6 Leidenfrost point
102
7 Conclusion and outlook
The present study deals with the investigation of transient spray cooling of a
heated thick target. Spray cooling is used in a great variety of applications like
cooling of high power electronics or quenching during steel making. Although
widely used, the overall physical picture is not complete and predictions for the
heat transfer are mainly of empirical nature. The drawback of these predictions
is the lack of universality and therefore limited range in which they produce
reliable results. To overcome this lack of knowledge several experiments of
transient spray cooling were conducted. The initially homogeneously heated
target was continuously cooled by various kinds of sprays. The local heat flux
and surface temperature were acquired and compared to visual observations of
the hydrodynamics during spray impact captured by a high speed observation
system.
In the film boiling regime the spray impact can be treated as single drop
impacts onto a dry surface and thus, can be regarded as a superposition of
single drop impacts. With decreasing temperature, and after the Leidenfrost
point has been passed, first small patches of water remain immediately after
the drop impact. Subsequently, in the transition boiling regime, the amount
of liquid remaining on the surface increases. Since the time spent in transition
boiling is very short, it is reasonable to simplify and replace this condition by
a jump of the heat flux. After passing the maximum heat flux (critical heat
flux) the entire surface is covered by a liquid film.
The mass flux of the impinging spray turns out to be the key factor for
achieving high cooling performance, whereas the mean droplet velocity and
diameter have only a smaller influence. Oblique spray cooling of a surface,
when the main spray direction is inclined towards the surface normal, results
in a lower heat transfer compared to the perpendicular impact. The decrease
turns out to be linearly dependent on the normal component of the spray
mean velocity. Furthermore, the heat transfer is influenced by the initial spray
water temperature. Cooling with hot water results in an decreased heat flux
103
7 Conclusion and outlook
The analysis of the heat transfer during spray cooling leads to a predictive
model for the heat flux and surface temperature in the film boiling regime. The
model is based on the superposition of single drop impacts and the assumption
of a semi-infinite target. It accounts for different surface materials, spray
properties, spray temperature and initial wall temperatures. There is excellent
accordance between experimental results and theoretical predictions for all
experimental data, covering a large range of parameters.
In the nucleate boiling regime the heat flux dependence on time indicates
some kind of a universal character, which is nearly independent of the spray
parameters. Assuming an instantaneous jump of the surface temperature
to saturation temperature and a semi-infinite wall, results in a heat flux
dependence on time, which is only influenced by the substrate material and
the initial temperature. This upper bound of the heat flux in the nucleate
boiling regime agrees well with experimental data as long as the semi-infinite
assumption is valid. The upper bound is reached, even when spraying with
low mass fluxes. This leads to the conclusion that in many cases the material
is the limiting factor for heat flux and that further increasing the spray mass
flux is of little benefit in the nucleate boiling regime.
104
Measurements of the Leidenfrost point, identified as the minimum heat
flux, indicate an independence on the spray characteristics, which is in strong
contrast to findings in literature. Instead, the Leidenfrost point is solely
influenced by the material of the target surface. A theoretical prediction
based on the formation of a vaporization front, assuming that the inertial
and viscous effects are comparable, agrees well with the experimental data
for sessile drops and sprays.
The general outcomes and derived models from this study provide a com-
prehensive set of predictions and new knowledge for developing and improving
reliable cooling processes which can be adapted to the special needs of each
individual cooling application.
The fact that the heat flux at the first instants of spray impact above the
Leidenfrost point is extremely high suggests another novel cooling strategy
based on the results of this study. Since the thermal boundary layer inside
the substrate has a significant influence on the cooling performance, it is
promising to use pulsating sprays for increasing the cooling efficiency, i.e. the
spray is continuously switched on and off. During the off-phase heat from the
lower portions of the substrate can flow to the top and decrease the thermal
boundary layer thickness. When the spray is again switched on, the spray
impact takes place on a substrate which has a thin boundary layer, leading
to an initial higher heat flux. Depending on the exact material properties
and temperature range, this may lead to an increased overall heat flux while
decreasing the required amount of spray fluid. The model developed in this
study for the film boiling regime accounts for the very high heat flux at the
initial phase of spray impact and can therefore be used to design and assess
such spray strategies.
Nevertheless, a number of topics should be addressed in future investiga-
tions: Due to the short timescales in the transition boiling regime it seems
reasonable to neglect this regime in some cases. But in others this simplifi-
cation leads to large errors. Since the physics of the formation and growth
of the liquid patches is not yet completely understood, no predictive models
exist. Furthermore, there must be some lower bound under which no complete
wetting of the surface can accrue. This boundary has to be identified and
described. Therefore future studies should deal with the transition boiling
regime.
105
7 Conclusion and outlook
Although the upper bound for the heat flux in the nucleate boiling regime
agrees well with the experimental data in this study, it is in fact an upper bound
which can only be valid if certain conditions are met. These circumstances
must be investigated so that models can be developed which describe when
to used this upper boundary. Additionally, a prediction for the heat flux, if
the upper boundary is not reached, has to be developed.
The theory for the Leidenfrost temperature is only valid for temperatures
below the critical temperature of the fluid, which is sufficient for many
industrial applications. Future investigations should find explanations for the
physical processes to achieve a prediction covering also the cases when the
temperature exceeds the critical value. Furthermore, the experimental data
found in literature for different fluids is often quite old and rather sparse.
Acquisition of an accurate and complete set of Leidenfrost temperatures with
dependence on various fluids and substrates could further enhance research
in this field.
106
Nomenclature
107
7 Conclusion and outlook
108
Φ - dimensionless heat flux
Lowercase greek letters
α m2 /s thermal diffusivity
αi m2 /s thermal diffusivity of the phase i, i ∈ [f, w]
β ◦
spray impact angle
J/s1/2 m2 K thermal effusivity
i J/s1/2 m2 K thermal effusivity of the phase i, i ∈ [f, w]
ζ - dimensionless paramter
η kg/ms dynamic viscosity
λ W/mK thermal conductivity
λv W/mK thermal conductivity of the phase v
ν m2 /s kinematic viscosity
νf m2 /s kinematic viscosity of phase f
ξ - dimensionless time
ξL - dimensionless time at Leidenfrost point
ρ kg/m3 density
ρf kg/m3 density of phase f
σ kg/s2 surface tension
σf kg/s2 surface tension of phase f
τ s parameter
χ - fitting parameter
Abbreviations
CHF critical heat flux
DPSS diode pumped solid state
f fluid
film film boiling regime
HS high speed
IHCP inverse heat conduction problem
L Leidenfrost
min minimum
nucleate nucleate boiling regime
PC personal computer
PD phase Doppler
PDA phase Doppler analyzer
109
7 Conclusion and outlook
v vapor
w wall
110
Bibliography
Abu-Zaid, M. (2003). An experimental study of the evaporation characteristics
of emulsified liquid droplets. Heat and Mass Transfer, 40(9):737–741.
Albrecht, H.-E., Borys, M., Damaschke, N., & Tropea, C. (2013). Laser
Doppler and Phase Doppler Measurement Techniques. Springer, Berlin
Heidelberg.
Auliano, M., Fernandino, M., Zhang, P., & Dorao, C. A. (2016). The Leiden-
frost phenomenon on silicon nanowires. In ASME 2016 14th International
Conference on Nanochannels, Microchannels, and Minichannels, ICNMM
2016. American Society of Mechanical Engineers.
111
Bibliography
Bar-Cohen, A., Arik, M., & Ohadi, M. (2006). Direct liquid cooling of
high flux micro and nano electronic components. Proceedings of the IEEE,
94(8):1549–1570.
Baumeister, K. J. & Simon, F. F. (1973). Leidenfrost temperature—Its
correlation for lipid metals, cryogens, hydrocarbons, and water. Journal of
Heat Transfer, 95(2):166–173.
Bernardin, J. D. & Mudawar, I. (1996). An experimental investigation into
the relationship between temperature-time history and surface roughness in
the spray quenching of aluminum parts. Journal of Engineering Materials
and Technology, Transactions of the ASME, 118(1):127–134.
Bernardin, J. D. & Mudawar, I. (1999). The Leidenfrost point: experimen-
tal study and assessment of existing models. Journal of Heat Transfer,
121(4):894.
Bernardin, J. D. & Mudawar, I. (2004). A Leidenfrost point model for
impinging droplets and sprays. Journal of Heat Transfer, 126(2):272.
Bernardin, J. D., Stebbins, C. J., & Mudawar, I. (1997). Mapping of impact
and heat transfer regimes of water drops impinging on a polished surface.
International Journal of Heat and Mass Transfer, 40(2):247–267.
Bertola, V. (2015). An impact regime map for water drops impacting on heated
surfaces. International Journal of Heat and Mass Transfer, 85:430–437.
Biance, A.-L., Clanet, C., & Quéré, D. (2003). Leidenfrost drops. Physics of
Fluids, 15(6):1632.
Bjornard, T. A. & Griffith, P. (1977). PWR blowdown heat transfer. In
Winter annual meeting of the American Society of Mechanical Engineers,
pages 17–41, Atlanta, GA, USA.
Breitenbach, J., Kissing, J., Roisman, I. V., & Tropea, C. (2018a). Character-
ization of secondary droplets during thermal atomization regime. Experi-
mental Thermal and Fluid Science, 98:516–522.
Breitenbach, J., Roisman, I. V., & Tropea, C. (2017a). Drop collision with a
hot, dry solid substrate: Heat transfer during nucleate boiling. Physical
Review Fluids, 2(7):074301.
112
Bibliography
Breitenbach, J., Roisman, I. V., & Tropea, C. (2017b). Heat transfer in the
film boiling regime: Single drop impact and spray cooling. International
Journal of Heat and Mass Transfer, 110:34–42.
Breitenbach, J., Roisman, I. V., & Tropea, C. (2018b). From drop impact
physics to spray cooling models: a critical review. Experiments in Fluids,
59(3):55.
Cai, C., Mudawar, I., Liu, H., & Si, C. (2020). Theoretical Leidenfrost point
(LFP) model for sessile droplet. International Journal of Heat and Mass
Transfer, 146:118802.
Castanet, G., Caballina, O., & Lemoine, F. (2015). Drop spreading at the
impact in the Leidenfrost boiling. Physics of Fluids, 27(6).
Celata, G. P., Cumo, M., Mariani, A., & Zummo, G. (2006). Visualization
of the impact of water drops on a hot surface: effect of drop velocity and
surface inclination. Heat and Mass Transfer, 42(10):885–890.
Chaze, W., Caballina, O., Castanet, G., Pierson, J. F., Lemoine, F., & Maillet,
D. (2019). Heat flux reconstruction by inversion of experimental infrared
temperature measurements – Application to the impact of a droplet in
the film boiling regime. International Journal of Heat and Mass Transfer,
128:469–478.
Chen, R.-H. H., Chow, L. C., & Navedo, J. E. (2002). Effects of spray charac-
teristics on critical heat flux in subcooled water spray cooling. International
Journal of Heat and Mass Transfer, 45(19):4033–4043.
113
Bibliography
Chen, S. J. & Tseng, A. A. (1992). Spray and jet cooling in steel rolling.
International Journal of Heat and Fluid Flow, 13(4):358–369.
Cheng, W. L., Zhang, W. W., Chen, H., & Hu, L. (2016). Spray cooling
and flash evaporation cooling: The current development and application.
Renewable and Sustainable Energy Reviews, 55:614–628.
Choi, K. J. & Yao, S. C. (1987). Mechanisms of film boiling heat transfer of
normally impacting spray. International Journal of Heat and Mass Transfer,
30(2):311–318.
Cox, T. L. & Yao, S. C. (1999). Heat transfer of sprays of large water
drops impacting on high temperature surfaces. Journal of Heat Transfer,
121(2):446–450.
Ebadian, M. A. & Lin, C. X. (2011). A review of high-heat-flux heat removal
technologies. Journal of Heat Transfer, 133(11):110801.
Emmerson, G. S. (1975). The effect of pressure and surface material on the
Leidenfrost point of discrete drops of water. International Journal of Heat
and Mass Transfer, 18(3):381–386.
Estes, K. A. & Mudawar, I. (1995). Correlation of sauter mean diameter and
critical heat flux for spray cooling of small surfaces. International Journal
of Heat and Mass Transfer, 38(16):2985–2996.
Gottfried, B. S., Lee, C. J., & Bell, K. J. (1966). The Leidenfrost phenomenon:
film boiling of liquid droplets on a flat plate. International Journal of Heat
and Mass Transfer, 9(11):1167–1188.
Hiroyasu, H., Kadota, T., & Senda, T. (1974). Droplet evaporation on a
hot surface in pressurized and heated ambient gas. Bulletin of JSME,
17(110):1081–1087.
Hofmann, J. (2019). Experimental investigation of the Leidenfrost point during
spray- and single drop impact. Master’s thesis, Technische Universität
Darmstadt, Darmstadt, Germany.
Hoogendoorn, C. J. & den Hond, R. (1974). Leidenfrost temperature and
heat transfer coefficients for water sprays impinging on a hot surface. In
IFTH International Heat Transfer Conference, volume 4, pages 135–138.
114
Bibliography
Ito, T., Takata, Y., Mousa, M. M. M., & Yoshikai, H. (1991). Studies on the
water cooling of hot surfaces (experiment of spray cooling). Memoirs of the
Faculty of Engineering, Kyushu University, 51(2):119–144.
Kim, H., Truong, B., Buongiorno, J., & Hu, L. W. (2012). Effects of
micro/nano-scale surface characteristics on the Leidenfrost point tempera-
ture of water. Journal of Thermal Science and Technology, 7(3):453–462.
Kim, J. (2007). Spray cooling heat transfer: The state of the art. International
Journal of Heat and Fluid Flow, 28(4):753–767.
Klinzing, W. P., Rozzi, J. C., & Mudawar, I. (1992). Film and transition
boiling correlations for quenching of hot surfaces with water sprays. Journal
of Heat Treating, 9(2):91–103.
Kruse, C., Anderson, T., Wilson, C., Zuhlke, C., Alexander, D., Gogos, G., &
Ndao, S. (2013). Extraordinary shifts of the Leidenfrost temperature from
multiscale micro/nanostructured surfaces. Langmuir, 29(31):9798–9806.
115
Bibliography
Monde, M., Arima, H., Liu, W., Mitutake, Y., & Hammad, J. A. (2003). An
analytical solution for two-dimensional inverse heat conduction problems
using Laplace transform. International Journal of Heat and Mass Transfer,
46(12):2135–2148.
Nimi, M., Pri, O., Ulivanju, K., Raudensky, M., Hnizdil, M., Hwang, J. Y.,
Lee, S. H., & Kim, S. Y. (2012). Influence of the water temperature on
the cooling intensity of mist nozzles in continuous casting. Materiali in
tehnologije, 46(3):311–315.
Nižetić, S., Čoko, D., Yadav, A., & Grubišić-Čabo, F. (2016). Water spray cool-
ing technique applied on a photovoltaic panel: The performance response.
Energy Conversion and Management, 108:287–296.
116
Bibliography
117
Bibliography
Shoji, M., Wakunaga, T., & Kodama, K. (1984). Heat transfer from a heated
surface to impinging subcooled droplets : Heat transfer characteristics
in non-wetting region. Transactions of the Japan Society of Mechanical
Engineers Series B, 50(451):716–723.
Sienski, K., Eden, R., & Schaefer, D. (1996). 3-D electronic interconnect pack-
aging. In IEEE Aerospace Applications Conference Proceedings, volume 1,
pages 363–373. IEEE.
Sozbir, N., Chang, Y. W., & Yao, S. C. (2003). Heat transfer of impacting
water mist on high temperature metal surfaces. Journal of Heat Transfer,
125(1):70.
Sozbir, N., Yigit, C., Issa, R. J., Yao, S.-C., Guven, H. R., & Ozcelebi,
S. (2010). Multiphase spray cooling of steel plates near the Leidenfrost
temperature-experimental studies and numerical modeling. Atomization
and Sprays, 20(5):387–405.
Spiegler, P., Hopenfeld, J., Silberberg, M., Bumpus, C. F., & Norman, A.
(1963). Onset of stable film boiling and the foam limit. International
Journal of Heat and Mass Transfer, 6(11):987–989.
Staat, H. J. J., Tran, T., Geerdink, B., Riboux, G., Sun, C., Gordillo, J. M.,
& Lohse, D. (2015). Phase diagram for droplet impact on superheated
surfaces. Journal of Fluid Mechanics, 779:R3.
Takata, Y., Hidaka, S., Cao, J. M., Nakamura, T., Yamamoto, H., Masuda,
M., & Ito, T. (2005). Effect of surface wettability on boiling and evaporation.
Energy, 30:209–220.
118
Bibliography
Tenzer, F. M., Hofmann, J., Roisman, I. V., & Tropea, C. (2020). Leiden-
frost temperature in sprays: role of the substrate and liquid properties.
Manuscript submitted for publication, (arXiv:2001.05426).
Tenzer, F. M., Pham, M. K., Roisman, I. V., & Tropea, C. (2019a). Fast spray
cooling: Effect of the liquid temperature. In HEFAT 2019, 14th Interna-
tional Conference on Heat Transfer, Fluid Mechanics and Thermodynamics,
pages 706–710, Wicklow, Ireland.
Tenzer, F. M., Roisman, I. V., & Tropea, C. (2018). Spray cooling in the
nucleate boiling regime: Hydrodynamics and heat transfer. In ICLASS
2018, 14th Triennial International Conference on Liquid Atomization and
Spray Systems, page 123, Chicago, IL.
Tenzer, F. M., Roisman, I. V., & Tropea, C. (2019b). Fast transient spray
cooling of a hot thick target. Journal of Fluid Mechanics, 881:84–103.
Tran, T., Staat, H. J. J., Prosperetti, A., Sun, C., & Lohse, D. (2012). Drop
impact on superheated surfaces. Physical Review Letters, 108(3):1–5.
Tropea, C., Xu, T.-H., Onofri, F., Géhan, G., Haugen, P., & Stieglmeier,
M. (1996). Dual-mode phase-Doppler anemometer. Particle & Particle
Systems Characterization, 13(2):165–170.
Wachters, L. H., Bonne, H., & van Nouhuis, H. J. (1966). The heat transfer
from a hot horizontal plate to sessile water drops in the spherodial state.
Chemical Engineering Science, 21(10):923–936.
119
Bibliography
Wang, Z., Qu, W., Xiong, J., Zhong, M., & Yang, Y. (2019a). Investigation on
effect of surface properties on droplet impact cooling of cladding surfaces.
Nuclear Engineering and Technology.
Wang, Z. F., Xiong, J., Yao, W., Qu, W., & Yang, Y. (2019b). Experimental
investigation on the Leidenfrost phenomenon of droplet impact on heated
silicon carbide surfaces. International Journal of Heat and Mass Transfer,
128:1206–1217.
Wendelstorf, J., Spitzer, K. H., & Wendelstorf, R. (2008). Spray water cooling
heat transfer at high temperatures and liquid mass fluxes. International
Journal of Heat and Mass Transfer, 51(19-20):4902–4910.
Yang, J., Chow, L. C., & Pais, M. R. (1996). Nucleate boiling heat transfer
in spray cooling. Journal of Heat Transfer, 118(3):668.
120
Bibliography
121
122
List of Figures
123
List of Figures
124
List of Figures
5.1 Scaling of the heat flux plotted as a function of time for the
stainless steel target. . . . . . . . . . . . . . . . . . . . . . . . 80
5.2 Scaling of the heat flux plotted as a function of time for the
target built of hot-work tool steel. . . . . . . . . . . . . . . . 81
5.3 Scaling of the heat flux plotted as a function of time for the
nickel target. . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.4 Exemplary comparison between one experiment and the theo-
retical prediction in the nucleate boiling regime. . . . . . . . . 83
125
List of Figures
126
List of Tables
2.1 Overview of atomizers used for the experiments and their
operational parameters. . . . . . . . . . . . . . . . . . . . . . 16
2.2 Parameters of the phase Doppler measurement system. . . . . 17
2.3 Overview of targets and material parameters used of cooling
experiments. . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
6.1 Overview the thermal properties of the materials and the value
of the Leidenfrost temperature. . . . . . . . . . . . . . . . . . 100
127