You are on page 1of 376

Nucleate Boiling on

Micro-Structured
Surfaces

Mieczyslaw E. Poniewski
and
John R. Thome

Warsaw, Poland and Lausanne, Switzerland


2008
Nucleate Boiling on Micro-Structured Surfaces

Published by Heat Transfer Research, Inc. (HTRI),


150 Venture Drive, College Station, TX 77845, USA,
www.HTRI.net.

We dedicate this book to our wives, Dorota and Carla.

Figures prepared by Dmitry Lomako, Minsk, Belorussia.

Copyright © 2008 M. E. Poniewski and J. R. Thome. All rights reserved.

You may download and print one (1) copy of this e-book for your personal use.
Prior to any additional printing or redistributing of this text,
you must obtain the authors’ written approval.
Nucleate Boiling on Micro-Structured Surfaces

Contents
About the Authors.........................................................................................v

Foreword....................................................................................................vii

1. Introduction: Characteristic Properties of Boiling on


Micro-Structured Surfaces..........................................................................1

2. Formation of Active Boiling Sites..................................................................7


2.1. Conditions for vapor bubble formation on a heated surface .................................................................................... 7
2.2. Surface geometry characteristics at a boiling site .................................................................................................. 8
2.3. Boiling site density on a heated surface ................................................................................................................ 19
2.4. Wetting angle hysteresis ........................................................................................................................................ 24
2.5. Summary................................................................................................................................................................. 33

3. Enhancement of Boiling Heat Transfer: Techniques and


Experimental Results................................................................................35
3.1. Types of micro-structured surface enhancements for boiling................................................................................. 35
3.2. Experimental results and empirical prediction methods ....................................................................................... 37
3.2.1. Industrially manufactured micro-structured surfaces ................................................................................ 37
Correlations for high flux porous coated surface....................................................................................... 50
Correlation for Gewa-T surfaces ............................................................................................................... 51
Correlation for Thermoexcel-E surfaces ................................................................................................... 54
3.2.2. Sintered powder structures......................................................................................................................... 55
3.2.3. Thermally sprayed structures .................................................................................................................... 70
Plasma spraying......................................................................................................................................... 71
Electric arc spraying................................................................................................................................... 80
Flame spraying .......................................................................................................................................... 85
3.2.4. Mesh structures.......................................................................................................................................... 90
3.2.5. Metal fibrous structures.............................................................................................................................110
Other researchers’ investigations . ...........................................................................................................110
Investigations by Poniewski and coworkers............................................................................................. 124
3.2.6. Surfaces developed with other technologies . ......................................................................................... 147
3.2.7. Combined surfaces................................................................................................................................... 149
3.3. Conclusions ......................................................................................................................................................... 153

4. Modeling Nucleate Boiling on Microporous Surfaces ................................... 157


4.1. Visualization of boiling on microporous surfaces – capillary-porous structures ................................................. 157
4.2. Heat transfer models............................................................................................................................................. 162
4.2.1. Attempt to classify the heat transfer models . .......................................................................................... 162
4.2.2. Heat transfer models – theoretical basis, experimental verification ....................................................... 165
Ferrell and Alleavitch model..................................................................................................................... 165
Moss and Kelly model............................................................................................................................... 167
O’Neill et al. models.................................................................................................................................. 169
Cornwell et al. model.................................................................................................................................174
Man’kovski et al. model . .......................................................................................................................... 175
Smirnov et al. model ................................................................................................................................ 181
Nishikawa et al. models............................................................................................................................ 190
Tehver et al. model.................................................................................................................................... 191
Orlov and Savelev model.......................................................................................................................... 197
Nakayama et al. model............................................................................................................................. 201
Rannenberg and Beer model.................................................................................................................... 204
Xin and Chao models................................................................................................................................ 206
Kovalev et al. model...................................................................................................................................211

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved iii


Contents

Ayub and Bergles model . ........................................................................................................................ 223


Kravchenko and Ostrovski model............................................................................................................. 225
Wang et al. model..................................................................................................................................... 230
Webb and Haider model........................................................................................................................... 235
Chien and Webb model............................................................................................................................. 243
4.3. Heat transfer hysteresis ....................................................................................................................................... 252
4.3.1. Kinds of boiling hysteresis . ...................................................................................................................... 252
4.3.2. Nucleation hysteresis................................................................................................................................ 253
Technically smooth surfaces – experimental investigations and theoretical analysis ............................ 253
Industrially developed microporous surfaces – experimental data . ....................................................... 257
Models for microporous surfaces............................................................................................................. 258
4.3.3. I kind of hysteresis – analysis of experimental data................................................................................. 270
Sintered powder and plasma sprayed porous coverings ........................................................................ 270
Metal fibrous capillary-porous structures................................................................................................. 277
4.3.4. II kind hysteresis....................................................................................................................................... 278
Experimental data analysis....................................................................................................................... 278
II kind hysteresis application to heat transfer control .............................................................................. 295
4.4. Boiling crisis and the critical heat flux................................................................................................................... 296
4.4.1. Definitions, experimental data and correlations . ..................................................................................... 296
Industrially manufactured developed microsurfaces................................................................................ 297
Sintered powder structures....................................................................................................................... 299
Thermally sprayed structures................................................................................................................... 302
4.4.2. Attempts at phenomenological modeling...................................................................................................311
Hydrodynamic models.............................................................................................................................. 312
4.5. Summary............................................................................................................................................................... 334

Bibliography.............................................................................................335

Notation...................................................................................................359

iv © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

About the Authors


Mieczyslaw E. Poniewski joined the faculty of Warsaw University of
Technology, Plock Campus, Poland, in 2001 and now he is an Ordinary
Professor and Director of the Institute of Mechanical Engineering (staff
above 50). His professional career focuses on different aspects of heat
transfer, heat management, and nonequilibrium thermodynamics. He
received his Ph.D. in 1975 and D.Sc. in 1989 at Warsaw University of
Technology, Faculty of Power and Aeronautical Engineering. He spent
the academic year 1977/78 at UC Berkeley, USA as a Fulbright scholar
and the year 1988/89 at the University of Houston, USA as a research
associate. Currently he is conducting research projects related to
pool boiling heat transfer enhancement on porous layers and forced
convective boiling in minichannels, concentrating on experimental and
theoretical approaches to boiling incipience and various hysteresis
phenomena. His main industrial achievement is heading the project for
the central heating system master plan for the city of Kielce, Poland
(about 230,000 inhabitants), 1997. In addition to the present book, he
is a coauthor of three other books in Polish, including the academic
handbook on nonequilibrium thermodynamics, which appeared this
year by the WUT Publishing House. He has also published more than
140 papers and numerous reports in the fields of his scientific interest.
He is the founder and organizer of the first three editions (1996, 1999,
2002) of a successful international conference in Poland, called HEAT
that is devoted to transport phenomena in multiphase systems. In 2007
he was elected to the Committee of Thermodynamics and Combustion
of Polish Academy of Sciences for the fourth term.

e-mail: meponiewski@pw.plock.pl

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved v


About the Authors

John R. Thome has been Professor of Heat and Mass Transfer at


the Ecole Polytechnique Fédérale de Lausanne (EPFL), Switzerland
since 1998, where his primary interests of research are two-phase
flow and heat transfer related to refrigerants, covering both macro-
scale and micro-scale heat transfer and enhanced heat transfer, and
directs a research staff of about 20. He received his Ph.D. at Oxford
University, England in 1978. In addition to the present book, he is the
author of three other books: Enhanced Boiling Heat Transfer (1990),
Convective Boiling and Condensation, 3rd Edition (1994) with J. G.
Collier, and Wolverine Engineering Databook III (2004). His Databook
III e-book is also available for free on the web at www.wlv.com/
products. He received the ASME Heat Transfer Division’s Best Paper
Award in 1998 for a three-part paper on two-phase flow and boiling
of refrigerants in horizontal tubes, published in the Journal of Heat
Transfer and received the J&E Hall Gold Medal from the UK Institute of
Refrigeration in 2008 for his contributions to refrigeration heat transfer.
He has published over 100 journal papers since joining the EPFL in
1998. He is the Swiss delegate to the Assembly of the International
Heat Transfer Conference, which organizes the International Heat
Transfer Conference every four years. He is also the chairman of
ALPEMA (Brazed Aluminium Plate-Fin Heat Exchanger Manufacturers’
Association). He is the author of the software Enhanced Heat Transfer
that is integrated into HTRI software for the design of enhanced heat
exchangers.

e-mail: john.thome@epfl.ch

vi © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Foreword

Boiling heat transfer is an intensive method of heat dissipation, yet the advancement of modern
technologies requires removing such large heat fluxes that boiling on a smooth surface does not
guarantee sufficient cooling. The efforts made by many researchers, therefore, have focused on
developing manufacturing technology and investigating the characteristics of micro-surfaces,
which significantly intensify boiling heat transfer. The latter possesses a number of properties
different from those typical of boiling on a technically smooth surface.

The issue of vapor bubble generation in active nucleation centers on the heating surface is of
fundamental importance for boiling enhancement. Chapter 2 thus provides the analysis of the
conditions, under which vapor bubbles are generated. Particular emphasis was put on the surface
geometrical characteristics at the site where nucleation occurs, the density of boiling sites, as
well as on the impact of wetting angle hysteresis on the nucleation phenomenon. The discussion
of the present state of knowledge leads to the statement that the density of active boiling sites
of nucleate boiling is proportional to the heat flux squared. The hypotheses concerning the
generation of vapor nuclei inside the heating surface irregularities are well, but only indirectly
confirmed by experiments. The lack of accurate knowledge of surface characteristics or the
possibility of a mathematical description of bubble interaction precludes the construction of a
unified theory of vapor bubble generation.

Chapter 3 is devoted to the discussion of experimental investigations, their results and


empirical correlations for the calculation of the Nusselt number, heat flux and the coefficient
of heat transfer or the heating surface superheating. Subsequently, experimental results for
industrially manufactured micro-surfaces of modified structure were presented, such as High
Flux, Thermoexcel-E, Turbo-B, Gewa-T, Gewa-TX, Gewa-TXY and others. The manufacturing
technologies of those micro-surfaces are different, yet all of them are representative of an on-
surface capillary-porous structure, which enhances the boiling process. The above-mentioned
industrial micro-surfaces are the most frequently used in heat exchangers at present.

The density of active boiling sites on those surfaces was found to be far higher than on smooth
surfaces, which leads to much higher heat transfer coefficients. The increase is a function of the
industrial micro-surface structural parameters, resulting from the manufacturing technology.

In order to obtain a developed micro-surface, various capillary-porous structures are applied to


form a layer on the substrate surface. They are made of sintered powders by means of thermal
(plasma, electric arc or flame) spraying or coating with mesh or metal fibrous sintered layers.
Metal fibrous structures demonstrate a number of advantages such as: a high value of heat

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved vii


Foreword

transfer coefficient, the possibility of manufacturing coatings of pre-set structural parameters,


having only open pores and the possibility of overlaying the base of complex shapes.

M. E. Poniewski’s investigations into those structures focused on their properties, particularly


on hysteresis phenomena. Extensive experimental work was conducted and, as a result, a new
correlation for the Nusselt number was found and a new II-kind hysteresis model was proposed.

All the above-mentioned porous surfaces, except for the mesh ones, enhance boiling heat
transfer. For the majority of porous coatings, an optimum layer thickness can be found, which
ensures the highest value of the heat transfer coefficient.

A specific feature of both industrially modified surfaces and porous coatings is the occurrence of
three kinds of hysteresis, i.e. nucleation hysteresis, I-kind hysteresis following the intralayer boiling
crisis and II-kind hysteresis characteristic of heat fluxes lower than the critical one.

There exists no general theory explaining the reasons for heat transfer coefficient increase,
the lowering of the surface initial superheating or the intensity and quantity of boiling hysteresis
phenomena on developed micro-surfaces. Because of only fragmentary knowledge of the boiling
process mechanisms inside the capillary-porous structure, the heat transfer coefficient prediction
methods are of empirical character and also sometimes based on the theory of similarity.

Chapter 4 deals with the modeling of nucleate boiling heat transfer on surfaces with a porous
coating. This is a complex thermodynamic process including phenomena of the change of
phase, heat as well as liquid and vapor flows. All known theoretical models, which aimed at the
explanation of the phenomena in both pool boiling and liquid capillary feeding of the porous layer.
The models described three kinds of vaporization characteristic of porous structures:

• vaporization of a thin liquid micro-layer covering the surfaces of capillaries;

• vaporization out of liquid-vapor menisci inside the structures;

• vaporization on the external surface of the porous coating.

The models put forward so far could be grouped according to the geometry of porous structures
they were developed for. Another possible division would be to base their categorization on the
common properties of the mechanisms of heat and mass transfer, which would yield:

• heat conduction through a liquid micro-layer covering the surfaces of micro-fins,


capillaries and internal tunnels, conveying the vapor to the liquid pool, vaporization on
the liquid micro-layer surface;

• vaporization out of liquid-vapor menisci in tunnels (capillaries) inside the layer;

• vaporization inside tunnels (capillaries) of the porous structure and convection on the
structure external surface;

viii © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

• porous coating substitution with equivalent micro-fins, conduction through micro-fins


and the liquid micro-layer covering them;

• applying the assumption that the sum of the pressure drop in the liquid and vapor layer
cannot exceed the capillary pressure increment;

• other singular models.

Models presented by the same authors can belong to two or even more groups differentiated
above, which makes those divisions quite arbitrary. They were, therefore, discussed in
chronological order, thus reflecting the broadening of the knowledge about boiling in capillary-
porous structures.

Special emphasis was laid on the detailed presentation of all assumptions concerning the
physics of boiling, substitute geometry simulating the capillary-porous structure and calculation
procedures. The similarities and differences of models were demonstrated and their usability for
engineering calculations was accentuated. Calculation results were compared with experimental
data selected by the authors of models.

The diversity of the boiling heat transfer hysteresis phenomena on surfaces covered with capillary-
porous structures became another subject of detailed considerations in Chapter 4. Experimental
data on nucleation hysteresis and then I and II kind hysteresis were discussed and compared.
Methods were also presented in an attempt to model nucleation hysteresis and II kind hysteresis.
Three various hypotheses explaining nucleation hysteresis were put forward, namely: the total
filling of the inactive pores with the liquid and the resistance of medium flow through the porous
layer, the dynamic change in the wetting angle as a function of the direction of the liquid motion
and the various geometries of pores occupied by vapor for boiling initiation and deactivation.

All forms of boiling heat transfer hysteresis are normally regarded as disadvantageous
phenomena, which excludes the thermal stabilization of systems dissipating large heat fluxes.

In this work, one of the authors (Poniewski) advocates a novel approach to the problem. It
consists of an attempt at the practical application of II kind hysteresis to heat transfer process
control. The boiling heat transfer model in metal fibrous porous coatings was presented, which
resulted in obtaining a boiling curve with II kind hysteresis. It was assumed that the increase in the
heating surface temperature and the heat flux is accompanied by a change in the participation of
individual heat transfer processes, which coexist or replace one another. The sequence of events
is different for increasing or decreasing the heat flux, which constitutes the physical cause of the
II kind hysteresis occurrence. The capillary-porous structure was substituted with an equivalent
system of vertical micro-fins where free space was proportional to porosity of the structure. It was
assumed that vaporization took place on the micro-layer surface, the liquid covering the micro-
fins, and on that basis the heat flux was calculated.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved ix


Foreword

Heterogeneity of pore dimensions causes the activation of pores of smaller and smaller
dimensions, while the surface temperature increases. With the temperature drop, the pores
already activated remain active, which leads to the II kind hysteresis. Theoretical and experimental
boiling curves demonstrated good agreement with this approach.

Hence, application of this idea means that II kind hysteresis could be applied to maintain the
constant temperature of the heating surface, in spite of an uncontrolled increase in the heat flux
(the patent of Poniewski).

The crisis of nucleate boiling is one of the most important heat transfer issues. In porous layers,
the surface temperature increases under the crisis conditions much less rapidly than in the case
of smooth surfaces and the heat transfer coefficient diminishes gently. The multitude of factors
affecting boiling crisis, such as the porous structure geometry, its structural parameters, the type
of liquid and the way the liquid is fed to the layer results in a significant divergence between the
trends observed experimentally. Apart from the nucleate boiling crisis, understood as the creation
of a continuous vapor film above the porous layer, there occurs also the so-called intralayer crisis,
which is thought to be the progressive filling of the porous layer inside with vapor.

The two boiling crisis models presented in the literature are based on the hydrodynamic
hypothesis of vertical vapor stream stability, with the critical heat flux for smooth surfaces acting
as the reference quantity. The models were applied to mesh structures and those of sintered
powders.

A totally new approach to intralayer boiling crisis was presented, that is Poniewski’s probabilistic
model. It is based on the concept of the maximum point process. Both the distribution of pores
on the heating surface and the process of their filling are thought to be Poisson point processes.
Using the concept of the sequence of thinning point processes, the dynamics of the drying of
the capillary-porous structure was described and the condition sufficient for the intralayer crisis
occurrence was established. The results were congruent with experimental data for water boiling
on a surface covered with a metal fibrous porous structure.

Although many studies dealing with the subject of boiling on micro-structured surfaces are
available (including those by the authors), they do not offer a uniform view or explanation of
the topic. None of the models discussed in the monograph completely explains the results of
experiments obtained by competing laboratories and/or capillary-porous structures of different
geometry manufactured by a different technology.

Unquestionable economic benefits can be realized from the tenfold or more enhancement of
boiling heat transfer accompanying the application of a micro-structured surface. This in itself
provides sufficient stimuli for the continuation of research in the field.

x © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Chapter 1
1. Introduction: Characteristic Properties of
Boiling on Micro-Structured Surfaces

Nucleate boiling is a basic two-phase heat transfer process in which heat is transferred from
a heated surface to a saturated fluid in contact with it. Nucleate boiling is capable of attaining
relatively large heat transfer coefficients from the intensity of the phase change process that
takes place. The intense heat transfer process is caused by the dynamics of vapor bubble
formation, growth and departure from the heating surface. The numerous rapidly growing bubbles
trap thin evaporating liquid microlayers on the heat wall and this significantly enhances the
transient conduction process from the wall to the evaporating liquid-vapor interface. Furthermore,
convection from the wall to the liquid is intensified by the high frequency motion of the liquid
adjacent to the heated wall imparted by the growing and departing bubbles. The motion also
cyclically removes the thermal boundary layer from the wall. All this decreases the thermal
resistance relative to single-phase natural convection and thus increases heat transfer. When
nucleate boiling occurs in an otherwise quiescent liquid, this is referred to as nucleate pool boiling.
This two-phase natural convection process is much more effective for heavy duty cooling of a
heated wall than single-phase natural convection to a pool of liquid.

At the upper limit of this nucleate boiling process, the so-called boiling crisis is reached when the
imposed heat flux is raised high enough. Here, a drastic change of the boiling process from the
nucleate boiling regime occurs (when the bubbles originate from nucleation sites formed by micro-
cavities in the surface change to unstable vapor jets) and the process jumps to the film boiling
process (where the surface is covered by an insulating film of vapor). This condition is normally
referred to as the critical heat flux and it marks the highest heat flux that can be dissipated within
the nucleate boiling regime. At the opposite extreme at very low heat fluxes, the temperature
difference between the heated surface and the saturated liquid is just sufficient to activate boiling
sites to begin the nucleate boiling process while the single-phase natural convection regime
prevails at lower wall superheats. This threshold temperature difference marks the onset of boiling
on the heated surface and is often called the nucleation or boiling incipience superheat. These
two extremes mark the upper and lower boundaries of the so called nucleate boiling curve, which
is the portion of the complete pool boiling curve which is of our present interest in this book on
enhancement of nucleate boiling.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 1


Chapter 1

While heat transfer in nucleate boiling on simple, plain surfaces is very intensive, new
technologies and even higher performance electronic devices require even better heat dissipation
solutions to carry away ever larger heat flux densities at relatively small wall to fluid superheats for
which boiling on plain, smooth surfaces does not usually prove to be sufficient. Consequently, a
genuine and challenging need for enhancing boiling heat transfer arises. This can be achieved by
various means, namely:

a. choosing an appropriate working fluid whose physical properties provide for higher
heat transfer coefficients;

b. pumping the cooling agent, thus superimposing forced convection on top of the
nucleate boiling process to increase heat transfer;

c. forming a micro-structure on the heat transfer surface to intensify the nucleate boiling
process and thus enhance heat transfer.

Option (a) is often not a viable choice since the selection of the working fluid depends on
the thermodynamic parameters of the system to be cooled (matching of the fluid’s saturation
temperature and resulting saturation pressure to the temperature of the device to be cooled, for
instance) and various practical considerations: material compatibility, safety, freezing point of
the fluid, etc. Option (b) introduces forced convection by inclusion of an additional device in the
system (a pump) and is often a viable choice but involves an additional cost for the pump/motor
package and the electrical power necessary to drive it.

Option (c) is the most attractive option and also the most effective one but also the least
understood. For instance, enhancement of the boiling process can be achieved by a change in
the roughness of the heating surface (although this approach does not yield large advantages),
covering it with a porous layer, changing its structure by making small openings in it, formation
of small fins on it, or modifying its wetting characteristics. The idea behind all these possible
changes is to greatly increase the number of active nucleation sites and to form bubbles at higher
frequencies, both of which enhance the underlying heat transfer mechanisms. However, the
methods for selecting the type of enhancement geometry, optimizing its dimensions and predicting
its thermal performance, especially for the more complex ones that yield the best results, are only
partially developed.

Within the above context, the principal focus of the present book is on the fundamentals
of enhancing heat transfer by use of micro-structured surfaces which can be fabricated by
deformation of the original surface to obtain new micro-geometries favorable to the boiling
process or by applying thin metallic layers made from meshes, particulate sprays, etc. Both
approaches lead to the formation of capillary-porous layers somewhat similar to those that exist
in heat pipes (where their objective is to pump the liquid from the condenser to the evaporator)
whose thicknesses are dependent on the depth of penetration of the boiling liquid into the micro-
structured surface and which typically range from a few microns to a few millimeters thick.

2 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Regarding applications to cooling technologies, it is essential to be able to determine the following


quantities that characterize nucleate boiling heat transfer when trying to dissipate large heat fluxes
at relatively low wall-to-fluid temperature differences:

a. the temperature difference between the heated wall and the fluid required to initiate
nucleate boiling incipience, i.e. the border between natural convection and nucleate
boiling, and thus that needed to attain the enhanced boiling performance regime;

b. the relationship between the wall-to-fluid temperature difference and the resulting heat
transfer coefficient;

c. the critical heat flux that identifies the onset of the boiling crisis, which is a condition to
be avoided during startup, normal operation and any transient changes in the imposed
heat flux.

There are in fact a myriad of possibilities of how to change a heating surface’s structure to
enhance its boiling performance, so research on the topic of boiling enhancement achieved by
such means is really vast and numerous patents have been filed. Although much work has been
devoted to the subject, it is difficult to draw clear-cut conclusions about the best approach or how
to predict or optimize their performance because of the multitude of parameters involved, which
affect the individual boiling phenomena and thus the intensity of the heat transfer process. This
situation poses a real challenge for researchers and heat transfer engineers alike and fully justifies
further extensive investigation into the field.

The present book thus attempts to bring together the wide variety of experiences reported in
numerous publications in the literature into a more coherent picture of the state of the art than
that currently available. The first book dedicated solely to this subject, Enhanced Boiling Heat
Transfer, written by one of the present authors (J. R. Thome), was published in 1990 and covered
both enhanced pool boiling and enhanced convective boiling. The first author (M. E. Poniewski)
has worked in the field of enhanced pool boiling heat transfer for several decades. He brought
that experience and also his access over those years to the extensive Eastern European literature
into his own book on this topic written in Polish, “Wrzenie pęcherzykowe na rozwiniętych
mikropowierzchniach”, that was published in 2001. The current book is our joint effort to put
together a new state-of-the-art on the specific topic of nucleate pool boiling on micro-structured
surfaces.

Due to the importance of bubble formation on boiling enhancement, in the present book the
process of boiling nucleation is discussed in detail for both natural cavities in plain surfaces and
for micro-structured surfaces in particular. An extensive and critical review of the numerous
experimental investigations aimed at increasing nucleate boiling heat transfer enhancement and
measuring their performances is also given, along with the corresponding empirical correlations
developed to describe these results.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 3


Chapter 1

When boiling on a surface covered with a capillary–porous structure, hysteresis of the nucleate
boiling curve is observed when cycling of the heat flux between low and high values. It has
important ramifications on thermal performance and manifests itself much more prominently on
micro-structured surfaces than when boiling on a smooth surface. Another feature of capillary-
porous structures is the diversity of the forms that hysteresis takes. Generally, a notable hysteresis
effect is thought to be a disadvantageous, since it may prevent thermal stabilization of a cooling
system (and that of the device to be cooled) when the heat flux to be dissipated is constantly
changing. The authors’ proposal to apply one of the hysteresis types described here to control the
heat transfer process will be an exception here.

The construction of a general model to predict the boiling performance of micro-structures


surfaces is still an unrealized objective of many researchers, including the authors. In the present
book, particular attention is paid to the modeling of boiling heat transfer on these micro-structured
surfaces. Such surfaces can be manufactured in many ways and, as a result, their geometries
differ considerably. Consequently, boiling heat transfer models have to be closely related to the
geometrical characteristics and dimensions of the particular form of micro-structure. Usually, it is
difficult to transfer or extend such models for one geometric class of micro-structures to another
fabricated by a different technique. Thus, many competing models are presented here such that
an overall picture of the different approaches is obtained while at the same time their similarities
and differences can be made evident.

One type of micro-structured geometry, which has been insufficiently investigated so far, is the
combination of an extended surface with a porous coating. A high density of fins significantly
enlarges the area available for heat transfer while the porous coating on the fins very favorably
affects the nucleate boiling process itself. Therefore, it seems reasonable to assume that this
combination of fins with a porous coating should substantially increase both the maximum level of
heat flux that can be dissipated from the original base area and also the heat transfer coefficient.
The extensive experimental results of one of the authors (M. E. Poniewski) on this type of
enhanced surface are presented in detail in this book, and confirm the assumption.

The book is structured as follows.

Chapter 2 addresses the numerous issues and theoretical aspects of boiling


nucleation and the formation of stable boiling sites on plain surfaces with natural
cavities and on micro-structured surfaces with their re-entrant types of cavities.
Furthermore, the parameters affecting the active boiling site density are reviewed and
the effect of wetting angle hysteresis is discussed.

Chapter 3 begins with a description of the numerous types of micro-structured


surfaces that have been proposed and tested, together with an extensive presentation
of the associated experimental boiling results. Then, a survey of the empirical

4 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

correlations proposed to describe nucleate pool boiling on these surfaces is


presented.

Chapter 4 is dedicated to describing the analytical and phenomenological models


proposed for predicting nucleate boiling heat transfer coefficients on micro-structured
surfaces. This chapter begins by presenting some visualization studies to help better
understand how these surfaces actually function. Then, the important issue of boiling
hysteresis is discussed in detail. Finally, the critical heat flux of micro-structured
surfaces is described and discussed.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 5


Chapter 1

6 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Chapter 2
2. Formation of Active Boiling Sites

Any discussion of boiling on a heated surface must begin with how the bubbles come to originate
on the surface, a process which is typically referred to as boiling nucleation. Numerous detailed
presentations and experimental investigations on boiling nucleation are available in the literature
and thus the present chapter addresses only the issues of importance here.

2.1. Conditions for vapor bubble formation on a heated surface

A vapor bubble constitutes an element of a new phase (gaseous phase) in the liquid (liquid phase). If
the liquid is superheated, vapor bubbles form spontaneously. Furthermore, a liquid superheating
ΔTsl corresponds to a certain critical value of the bubble radius Rcr, which equals:

2 σlg Tsat (vg – vl)


Rcr = (2.1)
hlg ∆Tsl

where the liquid and the vapor remain in the state of thermal equilibrium [150]. Formula (2.1) is
derived on the basis of thermodynamic considerations. It says that a bubble of the radius
R < Rcr is not stable, so that the vapor condenses and the bubble disappears. For a vapor bubble
of radius Rcr to be maintained at the state of equilibrium, it is necessary for the liquid superheating
to be equal to ΔTsl. If, on the other hand, R ≥ Rcr, the bubble will grow from the heat transferred by
thermal diffusion from the superheated liquid to the bubble’s surface.

In the event of boiling proceeding on the surface to which heat is transferred from the surface
to the liquid, bubbles are formed at certain cavities or nucleation sites in the surface and the
number of such active boiling nuclei is a function of the heat flux passing through the surface.
Furthermore, active boiling nuclei usually operate in a somewhat regular fashion, that is, bubbles
depart from active sites with a frequency that is approximately constant at a particular operating
condition whereas their frequency typically tends to rise with increasing heat flux.

The analysis of the conditions under which active boiling nuclei can exist and operate has been
the subject of many research works [21, 173, 245, 305 and 306]. Investigations into this field are
difficult to carry out because of the small dimensions of elementary bubbles and the surface
cavities they emerge from (on the order of 1 µm). It is generally assumed that a necessary
condition for a boiling site to remain active is its capacity to retain within its volume a small vapor
(or vapor-gas) nucleus. Later this vapor nucleus expands and grows until it attains a size larger
than the original cavity in the surface and then at some point departs from the surface, creating a
stable boiling site if this process continues to reproduce itself.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 7


Chapter 2

The following information about boiling sites in the heated surface is essential to the
understanding of the nucleate boiling heat transfer process:

a. the geometric characteristics of the cavity operating as the boiling site;

b. the density of boiling sites on the surface;

c. the effect of wetting angle hysteresis on the nucleation process.

2.2. Surface geometry characteristics at a boiling site

Numerous investigations, in particular those where electron microscopes were used, demonstrate
that all sorts of cavities, micro-cracks, scratches, etc., become active boiling sites. There have
also been experiments in which the heating surface has been artificially roughened with a
regularly repeated pattern. Regarding micro-surface geometries, active micro-cavities have been
found to take various shapes, such as conical, cylindrical, non-symmetrical, re-entrant, etc. Due
to the microscopic dimensions of these cavities, it is not possible to observe minute vapor bubbles
inside them or the early stage of the growth of such bubbles as they emerge from the mouth of
these cavities.

On the other hand, it has been observed that at a constant heat flux, the majority of active boiling
sites become a source of bubble formation, growth and departure at fairly regular time intervals.
Sometimes, a periodic operation of an active site is observed, when a series or burst of bubbles
is followed by a break in their formation, after which a new series starts. Moreover, active and
inactive periods of site operation can change in an irregular manner.

The operation of a perspective boiling site depends on how it becomes and remains active. If the
cavity is totally flooded with liquid, a considerable temperature difference is needed to initiate a
new bubble that is capable of further growth. If the bubble grows large enough, it departs from the
surface. That is followed by one of three possible scenarios:

a. the surrounding liquid totally floods the boiling cavity and therefore a certain period of
time is necessary so that this site can become active again;

b. the liquid only partially floods the boiling cavity and after a relatively short break,
another bubble starts growing;

c. a microscopic bubble always remains within the cavity and it is capable of


instantaneous growth.

Thus the activity of a given boiling nucleus depends primarily on its capability of maintaining a
residual vapor bubble, which becomes the seed for the next one, etc.

There are two means of evaluating the mechanism involved in retaining a residual vapor bubble.
The first is a phenomenological approach that depends on the boiling cavity geometry, i.e., on its

8 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

opening, depth, size and shape. The second involves a thermodynamic approach to modeling of
the process.

Bankoff [21] provided one of the first analyses based on a geometrical approach. He determined
the following condition for gas or vapor to be maintained in a conical cavity:

θ > 2 β (2.2)

where θ is the wetting or contact angle between the surface of the inflowing liquid and the wall
of the cone and 2β is the aperture angle of the conical section, as shown in Fig. 2.1 of [173]. This
condition was later used by Lorenz [149] for determining the critical conical radius of a boiling site.

The manner in which the boiling cavity operates depends on the liquid’s capability to wet the
surface, and thus on the surface tension and the wetting angle. Moreover, there can be a pre-
existing non-condensable gas nucleus in the boiling cavity, trapped there by the entering liquid
or from degassing of the liquid itself. The gas is either trapped by the heating surface or liberated
from the boiling liquid when it is heated.

Nishikawa and Fujita [173] discussed the operational stability of active boiling sites of various
geometries. Two cases were illustrated with the example of a conical opening in the heat transfer
surface, shown here in Fig. 2.1 for (a) a wetting fluid where θ < �/2 + β and (b) a non-wetting fluid
where θ > �/2 + β.

Fig. 2.1. Convex and concave interfacial shapes for vapor (gas) trapped inside a
conical boiling site under good (a) and weak (b) surface wettability with
boiling liquid, according to Nishikawa and Fujita [173]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 9


Chapter 2

The curvature of radius of the liquid - vapor interface can be expressed by the following formula:

(2.3)

where the plus sign corresponds to the geometrical condition

π
θ< + β (2.4)
2

while the minus sign corresponds to the condition

π
θ> +β (2.5)
2

The difference in vapor and liquid pressures on opposite sides of the interface is described by the
equation known as the Laplace equation:

pg – pl = ± 2 σlg /R (2.6)

Moreover, the plus or minus signs correspond to the same conditions which appear in the
equation (2.3).

If, in the event of condition (a) in Figure 2.1, the vapor volume is reduced, the length x gets
smaller as the radius R is reduced. As a result, the vapor pressure increases in accordance with
(2.6) which induces a corresponding increase in the value of the saturation temperature of the
trapped vapor. If this temperature is higher than that of the liquid in contact with the vapor, then
condensation will take place, and the vapor volume in the nucleus will diminish even further. There
is a possibility that the nucleus will get completely de-activated or the vapor will, at some point,
start to collect heat from the surrounding liquid, whose temperature may increase due to the
enthalpy being released by the condensation process or by some external influence. Summing
up, a conical boiling cavity with a wetting liquid may be stable or unstable. It can be demonstrated
that, contrary to the phenomenon discussed above, a conical boiling cavity with the non-wetting
liquid is always stable.

Another boiling nucleus, also discussed by Nishikawa and coworkers [173, 174], had a more
generalized shape, as shown in Fig. 2.2. The change in pressure difference pg – pl depends on
the location of the meniscus as shown in this figure. It demonstrates a peculiarity at the point
where the meniscus progresses from the cylindrical zone to the conical one.

10 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Considering the boiling cavity pictured in Fig. 2.2, the radius of curvature of the meniscus in the
conical part, for the wetting liquid, is expressed by the equation:

(2.7)

where the pressure difference pg – pl is determined from (2.6). The signs “+” and “–” in (2.6) and
(2.7) correspond to the following conditions, as illustrated in Fig. 2.2: (a) the sign + is for when
θ < �/2 – β and (b) the sign – corresponds to the situation when θ > �/2 – β.

Similarly to the previous case, condensation can take place and possibly a new equilibrium
condition can be established at a new location in the boiling cavity. Here, we have to take
into account the following circumstances: firstly, owing to the different geometry, the pressure
difference in the conical part will be much smaller than that in Fig. 2.2, and secondly, when 2θ ≥ �,
the boiling nucleus is always stable.

Griffith and Wallis [92, 245] investigated other conditions for boiling incipience, where boiling
cavities of pre-set diameter were stable and contained residual vapor bubbles. The relationship
they obtained between the radius of curvature of the vapor-liquid interface and the superheat of
the liquid required to attain thermal equilibrium, obtained by combining the Laplace and Clausius-
Clapeyron equations, was confirmed experimentally and reads as follows:

2 σlg Tsat (vg – vl)


Rint = (2.1.a.)
hlg ∆Tsl

where ΔTsl represents the difference between the wall temperature and saturation temperature at
the instant when the nucleus becomes active. The dependence is identical with equation (2.1), and
it describes the critical radius of the interface at a liquid superheating of ΔTsl.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 11


Chapter 2

Fig. 2.2. Pressure variation inside a diminishing size of vapor bubble for cases (a)
and (b) from Fig. 2.1, according to Kraus [131] and Nishikawa and Fujita [173]

a) Inflowing liquid b) Liquid


θ
θ
h

Entrapped vapor

2R0

Fig. 2.3 Vapor trapping process according to Singh, Mikic and Rohsenow [245].
a) Mechanism suggested for flooding of a cylindrical boiling cavity and
trapping of vapor; b) Meniscus of vapor trapped inside a cylindrical boiling
cavity

12 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Drawing on Bankoff’s considerations [21], Singh, Mikic and Rohsenow presented a model for
trapping of vapor in a cavity [245]. It is shown in Fig. 2.3 for a cylindrical cavity in the wall and the
liquid wetting the surface, i.e., for the wetting angle θ < �/2. The case where the liquid wets the
surface is more important for practical applications than the one in which the liquid does not as
most commonly used fluids and surface combinations result in a wettable condition.

Applying the above model to a growing and departing bubble, heat is transported by convection to
the liquid in contact with the heating surface. This liquid layer thus becomes superheated above
the saturation temperature. The bubble grows due to the heat supplied by conduction (thermal
diffusion) from the superheated liquid layer surrounding the growing bubble and thermodynamic
equilibrium is assumed to be maintained at the interface. After the bubble departs, the outside
liquid flows into the cavity through the cylindrical opening, pushed by inertia forces produced by
the bubble breaking away from the surface. Due to the system seeking to attain thermodynamic
equilibrium between the vapor and liquid near the opening, the small fraction of the vapor may
partially condense during the flooding process and eventually the vapor will be compressed by the
inertia forces.

At the instant of the liquid contact with the whole perimeter of the cavity opening, a meniscus is
formed and a certain amount of vapor is entrapped, which is shown in Fig. 2.3 (on the right), and
the radius of the liquid surface curvature R relative to the radius of the cylindrical cavity R0 is
expressed by the equation:

R0
R= (2.8)
cos θ

Singh, Mikic and Rohsenow [245] considered the movement of the liquid-vapor interface, taking
into account the friction forces on the opening surface, pressure forces on the phase division
boundary and also inertia forces resulting from liquid motion inside the opening. Thus, they
provided the analysis of the dynamic motion inside the cavity that occurs during the entrapment of
a boiling nucleus.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 13


Chapter 2

5
-6
R0 = 2.45⋅10 m

4 Water Methanol
θ = π/4 3 θ = 2/45 π

1 - w0 = 0
3 2 - w0 = 0.508 m⋅s-1
3 - w0 = 1.016 m⋅s-1
Tstat , K
act

2 2
3
1
1 1

0
0 5 ⋅10-6 10 ⋅10-6 15 ⋅10-6 20 ⋅10-6
Depth of boiling nuclei , m

Fig. 2.4. Impact of boiling cavity depth and flooding velocity on the liquid
superheating necessary for site activation, from Singh, Mikic and
Rohsenow [245]

The authors [245] termed radius R0 of equation (2.8) a static radius, whereas the value of the
radius resulting from the dynamic analysis was denoted Ref. Then they introduced the quantity

act ∆Tact R0
∆Tstat = =
∆Tstat Ref

which is a ratio of the actual superheating ΔTact to the static superheating ΔTstat, corresponding to
the radius R0 of (2.8) and calculated on the basis of (2.1).

14 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Fig. 2.4 shows the results of calculations for water and methanol, where w0 stands for the initial
velocity of the flooding through the opening and R0 is the radius of the cylindrical opening.
Considerations presented in [245] allow us to draw the following conclusions:

a. a small value of the wetting angle results in a deeper liquid penetration and hinders
the boiling nucleation process, causing an increase in the liquid superheating
necessary to activate the small resulting vapor nucleus;

b. an increase in the initial velocity of the liquid penetration diminishes the stability of the
nucleation process as it causes deeper liquid penetration into the cavity;

c. an increase in the cavity depth at the given value of R0 enhances the stability of the
boiling nucleation process.

These authors [245] also conducted experimental investigations for water and methanol for
cylindrical boiling cavities manufactured with a laser. The measured values of Ref they obtained
were congruent with their theoretical considerations.

Wang and Dhir [306] proposed a model of operation for various shapes of cavities. Their model
was based on a thermodynamic analysis, where they took into account changes in the Helmholtz
free energy for the liquid-gas interface as it moved inside the cavity and in its surroundings. The
criterion for trapping a residual gas or vapor nucleus in the boiling cavity has the following form:

θ > βmin (2.9)

where βmin denotes the minimum angle between the horizontal and the tangent to the surface
of the boiling cavity at its edge for conical, spherical and sinusoidal shapes, respectively, as
illustrated in Fig. 2.5.

The nucleation criterion takes on the form of a modified Griffith-Wallis dependence:

2 σlg Tsat
∆Tsl = K (2.10)
ρg hlg R0

~
~ / V
K = R0 / Rlg – m (2.11)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 15


Chapter 2

where K is a dimensionless bubble curvature and R0 is the radius of the boiling cavity’s opening,
~
Rlg is the radius of curvature of the interface and m is a dimensionless vapor bubble mass.
Moreover, the criterion of stability of an elementary bubble was established to be:

dK
~ > 0 (2.12)
dV

~
where V is a dimensionless bubble volume, being the ratio of the vapor bubble volume divided by
the volume of the cavity, that is Vg /Vp.

Fig. 2.5. Examples of various shapes of liquid-gas interfaces for spherical, conical
and sinusoidal boiling cavities as a function of the interface’s position and
the ratio of the wetting angle to the minimum angle between the horizontal
and the tangent to the edge of the cavity according to Wang and Dhir [306]

The parameter K reaches the maximum value Kmax = 1 for the wetting angle θ ≤ �/2 and the value
Kmax = sin θ for θ > �/2. The boiling incipience in a given nucleus is expressed by the dependence:

2 σlg Tsat
∆Tsl = Kmax (2.13)
ρg hlg R0

16 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

This means that nucleation starts when the dimensionless bubble curvature reaches its
maximum value.

pg pg pg

Surface simulating
the heating surface
pl
Liquid A

2R
pl
pl C
pl
2Rcap 2Rcap
B
Vapor bubble

pg pg

Fig. 2.6. Model for the theoretical analysis of the nucleation process inside a channel
of a capillary porous structure from Konev and Mitrovich [110]

Konev and Mitrovich [110] analyzed the nucleation process in capillary-porous structures. The
model of the system is depicted in Fig. 2.6 and shows a bubble inside a capillary tube of the radius
Rcap. The liquid filling the capillary tube was divided at the cross section C, indicated in Fig. 2.6b,
which was assumed to operate as the heating surface. Thus, the capillary tube in Fig. 2.6c can be
assumed to represent the vapor in a channel inside a porous structure. The radius of the bubble
in the state of equilibrium with the superheated liquid in the capillary tube of Fig. 2.6c can be
expressed as:

(2.14)

where

(2.15)

The above expressions are correct for the situation where the liquid surrounding the bubble has a
flat interface.

In the case shown in Fig. 2.6, the interface is concave and the respective equation adopts a
modified form:

(2.16)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 17


Chapter 2

where Rcap denotes the radius of the capillary tube. If Rcap → ∞, equation (2.16) reduces to (2.14).
The temperature difference corresponding to a given value of R can be determined by rearranging
equation (2.16) to obtain:

(2.17)

This expression can be further simplified because ω is very small, whereas ρ g can be neglected
with respect to ρl for low and moderate pressures. Thus, the above expression can be reduced to
a relatively simple form as follows:

(2.18)

The above expression shows that the temperature difference ΔTsl is not only affected by physical
properties but also by the capillary tube radius Rcap.

8
Water, T water = 333 K
7
Temperature difference DTsl , K

6
2R cap
5

4
Rcap = ∞
3
2R
2

1 Rcap = 5⋅10-5 m Rcap = 10-4 m

0
0 1 2 3 4 5 6 7 8 9 10 11 12
Bubble radius in equilibrium R ⋅10 5 , m

Fig. 2.7. Boiling liquid superheating ΔTsl calculated as the function of vapor bubble
radius for various capillary radii from Konev and Mitrovich [110]

Fig. 2.7 presents the results of calculations of ΔTsl as a function of the bubble radius for different
values of the capillary radius Rcap. ΔTsl is seen to decrease when the bubble radius grows, yet the
fall is sharper for small values of Rcap. It should be noted that ΔTsl can reach zero for certain Rcap,
then R equals Rcap. Then even a small temperature difference is sufficient to initiate bubble growth
and activate the boiling surface and thus transfer a large quantity of heat, which is typical of what
is observed in capillary structures.

18 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

2.3. Boiling site density on a heated surface

Boiling site density on heated surfaces, N, has been measured by many researchers using
different methods over the years. Based on their results, Gaertner and Westwater [88] put forward
the following relationship:

2.1
N~q (2.19)

On the basis of numerous earlier works, Hsu and Graham [99] concluded that the following
functional dependency could be formulated as:

1 - 2.1
N~q (2.20)

where the exponent on q depends on the heat flux. For very small heat fluxes, the exponent
equals 1, whereas for larger values it goes to 2.1.

Nishikawa and Fujita [173] proposed a similar relationship to that of Hsu and Graham, making the
exponent on q equal to 2 for large heat fluxes. For heat flux values that are relevant to engineering
practice, it seems that the exponent has a value of about 2.

To complete the expressions for N above, a proportionality coefficient needs to be added to the
expression. Finding such values by fitting the expressions to experimental data, the values of
the proportionality coefficient to introduce into equations (2.19) and (2.20) differ from each other
significantly, demonstrating a strong dependence on the type of surface, its roughness or the kind
of boiling liquid wetting the heating surface. Up to now, no reliable model has been proposed to
calculate this empirical coefficient.

According to experimental observations, the relationship between the heat transfer coefficient and
heat flux approximately takes on the form:

2
α ~ q3 (2.21)

2
1 2/3 1/3 3
For small q from (2.20): N ~ q . From (2.21): α ~ q3 → q/ΔT ~ q → ΔT ~ q → ΔT ~ q →
3
ΔT ~ N. Thus, the following relationship can also be discerned:

3
N ~ ∆Tsl (2.22)

The corresponding proportionality coefficient to close the expression is still indeterminate.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 19


Chapter 2

Wang and Dhir [305] conducted very thorough experimental investigations which aimed, among
other things, to obtain boiling site density data and they confirmed the relationship given by
(2.22). It should be emphasized that the majority of experiments described in the literature, where
the measurement of boiling site number is taken directly (Wang and Dhir’s work belongs to this
category), yields the relationship given by (2.22).

Apart from considering the impact of parameters characterizing the functional dependence of
the boiling site density on the surface wettability, a number of works dealt with the analysis of the
dependence of active site number on the total number of the potential boiling sites and also with
the statistical analysis of the boiling sites.

Mikic and Rohsenow [162] put forward the relationship in the following form:

(2.23)

where Rmax is the radius of the largest surface deformation constituting a potential boiling cavity
–2
for hosting an active nucleation site, which corresponds to N = 1 m , whereas R0 is the radius
of the critical cavity size resulting from the difference between the surface temperature and the
saturation temperature. This quantity can be calculated from the modified dependence (2.11). The
exponent m is determined empirically.

Bier et al. [30] gave the following relationship for describing N:

(2.24)

where Nmax denotes the maximum value of N, corresponding to R0 = 0, where Rmax has the same
meaning as in equation (2.23), whereas the value of exponent m depends on the type of the
surface treatment. This model presented by Bier et al. [30] is based on the premise that the heat
flux transferred by the boiling process consists of two components. One component corresponds
to pure free convection and can be calculated from appropriate correlations provided in the
literature. The other component covers heat transportation by bubbles departing from the surface.
After making assumptions concerning the heat conduction in the liquid layer near the bubble,
defining the area affected by the boiling nucleus and the statistical distribution of active boiling site
diameters, they arrived at the above equation (2.24). The value of m for R-113 and R-11 was found
to be 0.42 for their chemically etched surface and 0.26 for their mechanically treated surface.

20 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Cornwell and Brown [64] carried out a synthetic analysis of active boiling site distribution for water
on a copper surface at atmospheric pressure. The analysis was restricted to heat fluxes that were
not very large but accounted for the influence of surface roughness. Their results were summed
up by the following relationship:

4.5
N ~ ∆Tsl (2.25)

where the proportionality coefficient to close the relationship depends on the surface roughness.

The analysis conducted by Yang and Kim [322] relied on known dimensions and the conical angle
of cavities in the heating surface that provided active boiling sites. For cavities with openings
R0 ranging from 0.43 to 2.8 μm the distribution of the surface roughness dimensions could be
expressed by the Poisson equation:

–λR0
f(R0) = λe (2.26)

whereas the distribution of conical half angles of roughness could be described by a normal
distribution

(2.27)

where λ and s are statistical parameters and β represents the mean value of the conical angles β.

The number of surface deformations which might continuously maintain gas or vapor and operate
as active nucleation sites was calculated with the use of the Bankoff criterion (2.2). They are
the cavities whose radii at the heating surface range from Rmin to Rmax. Rmin is calculated from
expression (2.8) and Rmax is determined on the basis of a statistical analysis of the surface
properties [322]. The resulting expression has the following form:

(2.28)

where N̅ denotes the mean density of boiling sites determined experimentally. The values of the
statistical parameters depend on the material and the method of surface preparation.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 21


Chapter 2

Investigations carried out by Wang and Dhir [305] aimed at defining the relationship between
active boiling sites relative to all potential boiling sites (cavities in the heating surface).
Measurements were implemented as follows:

a. first of all, the following were defined for each surface: the measurement method of
boiling site density, the distribution of their dimensions and the conical angle at the
mouth of the cavity and the heating surface;

b. surfaces of different wettabilities were prepared and the appropriate wetting angles
measured;

c. the relationship between the dimensions of the active boiling sites and the heating
surface superheat temperature was verified by experiment;

d. the functional dependency between the density of active boiling sites and the heat flux
was established.

The results of their investigations can be summed up as follows:

1. Three surfaces were examined for water at atmospheric pressure and their wetting
π 7 π
angles were , π, and . When the wetting angle was equal to �/2, only 1 – 10%
2 36 10
of the cavities observed operated as active boiling sites.

2. In the area under consideration, three different distributions of cavity density in the
heating surface were found, namely:
7 –2 –2
a. N = 9 ∙ 10 D , m
D ≥ 5.8 ∙ 10 μm
6 –5.2 4 –2
b. N = (10.3 + 2.5 ∙ 10 D ) ∙ 10 , m (2.29)
3.5 ∙ 10 μm ≤ D ≤ 5.8 ∙ 10 μm
6 –5.4 4 –2
c. N = (2213.5 + 1.0 ∙ 10 D ) ∙ 10 , m
D ≤ 3.5 ∙ 10 μm

where D is the diameter of the mouth of the cavity at the heating surface measured in
µm. The density of active boiling sites, that is the density of cavities for all diameters
(2.29) satisfying the condition that the angle β, (Fig. 2.5), between the horizontal line
and the tangent to the boiling cavity surface at its edge is smaller than �/2, was found
to be:

10 –54 –2
N = 5.8 ∙ 10 D ,m β < π/2 (2.30)

In Fig. 2.8, the results of experimental investigations for water on a copper surface are
compared with the curves resulting from equations (2.29) and (2.30).

22 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

3. The number of active boiling sites rises as the heating surface’s wettability increases,
i.e., when the wetting angle diminishes.
2
4. The density of active boiling sites varies approximately proportional to q , independent
of the heating surface’s wetting angle, which is illustrated in Fig. 2.9.

5. The distribution of active boiling sites is described by a Poisson distribution.

6. The distribution of the distances between the nearest neighboring boiling sites is also
described by a Poisson distribution.
͞͞
7. The distance between the nearest neighboring sites approximately equals 0.84 √N.

9
10
Experimental density of hollows on the surface N
data density of active boiling nuclei N a
108

107

106
-2
N, N a , m

1
105

104 Copper/water
2
p = 1.013 ⋅10 Pa
5

103
1 - N from eq. (2.29)
2 - N a from eq. (2.30 )
2
10
2 ⋅10 -6 5 ⋅10
-6
10
-7
5 ⋅10
-7

Diameter of the hollow opening at the heating surface


D, m

Fig. 2.8. The density of active boiling sites as a function of the cavity diameters in the
heating surface – D, taken from Wang and Dhir [305]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 23


Chapter 2

107

q = p/2
q = (7/36) p
106
q = (1/10) p

5
Na , m-2 10

104

103
Na ~ q 2

102 3 4 5
10 10 10 10 6
q, W⋅m-2

Fig. 2.9. Density of active boiling sites as the function of heat flux from Wang and
Dhir [305]

2.4. Wetting angle hysteresis

The boiling heat transfer process is subject to hysteresis, which is manifested by the fact that
the curves of the dependence q = f (ΔT) differ from one another when the heat flux is decreased
and increased as shown later in section 4.2. This phenomenon has not been fully investigated and
lacks a quantitative description. The latter fact is undoubtedly connected with the wetting effects, and
it is generally believed that the decisive factor is attributable to variable characteristics of the surface
tension and the wetting angle [83]. In the theory of capillary phenomena, static and dynamic wetting
angles are defined. They differ from each other and occur, respectively, for motionless and mobile
phase boundaries where there is a three-way contact of solid, liquid and gas.

24 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Fig. 2.10. Surface tension on the surfaces of a phase boundary: solid, liquid and gas in
the state of static equilibrium

The static wetting angle can be calculated from the Young equation (Fig. 2.10):

σlg cos θ = σsg – σsl (2.31)

Despite this simple mathematical formula, the equation cannot be verified experimentally due to
difficulties in measuring σsg and σsl. It is usually σlg and θ that are determined experimentally and
on that basis, the value σsg – σsl is given. The equation (2.31) is not entirely correct as mechanical
equilibrium requires both the vertical and horizontal constituents of the surface tension. There
should therefore exist a vertical reaction force in the solid body equal to σlg sin θ. On the other
hand, if the system is in the state of equilibrium, it should have the minimum value of free energy.

The presence of surface tension calls for the introduction of the notion of surface energy for the
energy conservation equation. The conclusion that can be drawn is that the wetting angle should
become a parameter of a given cavity, which is fully confirmed by observations.

There are many methods for measuring wetting angles, in which most often the static angle is
not measured. The measurements are taken when the motion of the phase boundary is only
slight, which accounts for much of the uncertainty in the measurements. If measurements are
made at different velocities of movement, the wetting angle generally depends on this velocity.
Measurements can be made for positive (in accordance with Fig. 2.11) or negative values of the
velocity of the phase boundary. This diagram presents the typical results of such measurements
where w is the velocity of the phase boundary.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 25


Chapter 2

q+

q-

0 w

Fig. 2.11. The influence of the dynamic wetting angle θ on the sense and the velocity of
the phase boundary from Dusan [83]

∂θ
It was found out that usually ∂w ≥ 0.

The value θ+ constitutes the extrapolation of the results for the motion of the phase boundary to
the right (forward movement of the liquid); similarly θ – is the extrapolated value for the motion of
phase boundary to the left (receding movement of the liquid).

For many liquid-vapor-solid systems, it has been noted that there is in fact a range of boundary
angles [θ –, θ+], characterized by the fact that if the angle θ lies in this range, the phase boundary
does not change its position. This phenomenon is called wetting angle hysteresis.

Fig. 2.12. Surface tension and the wetting angle for the liquid moving over a smooth
surface from Cornwell [63]

Surface roughness also affects the value of the phase boundary angle. Fig. 2.12 shows a flat,
smooth surface with a liquid moving over it, forming the wetting angle whose real value is θreal and
–2
its increment of Δθreal when the liquid is moving. Free surface energies (J m ) or surface tensions
are denoted by symbols σ with appropriate subscripts. If the liquid moves in such a way that it
additionally covers the surface ΔFs, the change in Gibbs free energy amounts to:

∆Gs = σsl ∆Fs – σsg ∆Fs + σlg cos(θreal + ∆θreal) ∆Fs (2.32)

26 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

In the state of equilibrium

(2.33)

from which the Young equation is obtained

σsg – σsl = σlg cos θreal (2.34)

In this equation, it was assumed that the vapor pressure is its saturation pressure. Thus the
quantity σsg illustrates the case where the whole wall surface is covered with a layer of adsorbed
vapor and can be expressed as follows:

σsg = σs – πg (2.35)

where σs stands for the surface energy of a stationary partition in a vacuum, whereas �g is
proportional to the pressure of the adsorbed vapor layer.

Fig. 2.13 shows a fragment of a rough surface. Here the change in the surface energy equals:

∆Gs = σsl ∆Fs – σsg ∆Fs + σlg cos(θ + ∆θ) ∆Fpa (2.36)

Fig. 2.13. Rough surface wetted by the moving liquid front from
Cornwell [63]

where θ is the mean value of the wetting angle for a given surface. If the measure of surface
roughness is defined by the ratio of actual rough area with respect to the projected area then:

actual rough area ∆Fs


s= = (2.37)
projected area ∆Fpa

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 27


Chapter 2

In the state of equilibrium

cos θ
σsg – σsl = σlg (2.38)
s

Moreover, from equations (2.34) and (2.38), it can be concluded that

cos θ = s cos θreal (2.39)

The quantity θentp is the wetting angle determined most frequently experimentally, whose value
lies in the interval from θ – to θ+. The angle θ is a function of s and θreal and it can be assumed
to be the macroscopic value given in the literature. Equation (2.39) is often called the Wenzel
equation [16]. From equation (2.39) it can be deduced that for θreal > �/2, the roughness increases
the value of the angle θ but for θreal < �/2, it reduces the value of θ.

The wetting angle hysteresis phenomenon described above affects the operation of boiling sites,
which can be explained by the analysis presented in [63], presented below.

Fig. 2.14 presents four possible diagrams of conical boiling site operation. Case (a) corresponds
to a wetting liquid that has a changeable wetting angle, (b) reflects the case where a wetting
liquid has a constant wetting angle, (c) refers to the situation where a non-wetting liquid has a
changeable wetting angle and (d) shows a non-wetting liquid with a constant wetting angle.

a) b) c) d)
Stable Unstable
2
2
1 1 Interfaces 2 2
1 1
Gas and vapor
entrapped inside
the boiling nucleus

θ changes θ constant θ changes θ constant


p gas + pvapor > pl (R > 0) pgas + pvapor < p l (R < 0)

Fig. 2.14. Diagram of conical boiling site operation; θ – wetting angle, R – radius of
phase boundary curvature in the state of equilibrium from Cornwell [63]

The condition for pressure equilibrium in the nucleus is as follows:

pg = pgas + pvapor (2.40)

28 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

and the Laplace equation also gives:

2 σlg
pg – pl = (2.41)
R

Thus, the phase boundary is stable for a small perturbation dp, if the following criterion is satisfied:

(2.42)

which means that the signs of the perturbation dp and dR are the same. The stability condition
can be rewritten as follows:

(2.43)

because changes in the volume of gas and vapor mixture contained in a conical hollow are
proportional to changes in pressure. On the basis of the above relationships, it can be concluded
that the nuclei (a), (c) and (d) in Fig. 2.14 are stable whereas the nucleus (b) is conditionally
unstable.

Liquid
θ Recycling interface

Solid
θreal
θ

θreal
Moving forward
interface
x
Gas (vapor)
y ϕ
Conical hollow
β

Fig. 2.15. Surface roughness inside a conical cavity, taken from Cornwell [63]

Furthermore, changes in the wetting angle caused by the micro-roughness inside the conical
opening can be analyzed. Fig. 2.15 shows, in the form of a diagram, a simplified structure of a
conical cavity whose walls have a roughness with a simple geometrical shape that can be easily

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 29


Chapter 2

described mathematically. Combining equation (2.39) with the surface geometry (Fig. 2.15) yields
the equations:

(2.44a)

cos θ = s cos θreal (2.44b)

θ+ = θreal + φ θ – = θreal – φ (2.45)

The solution of this system of equations is shown in Fig. 2.16 for the cone half angle β = π/18
for three different values of the angle θ. The values of angles θ – and θ+ constitute the hysteresis
boundary. The probable condition that restricts stable entrapment of gas in such a conical cavity is
the stability of the gaseous phase at the opening, that is:

π
θ+ ≥ +β (2.46)
2

Under such conditions, the mean value of θentp amounts to, from equations (2.39), (2.44) and
(2.45):

(2.47)

or making use of the equation (2.44)

(2.48)

30 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

2.0
1
β= π
θ- θ real θ+ 18

1.8

2π 9 θ = 2π 9 2π 9
1.6 π/3 π/3 π/3
4π 9 4π 9 4π 9
s

θentp
1.4 Surface
Entrapment Entrapment shapes
impossible possible
ϕ
1.2

1
s=
cos ϕ
1.0
1 1 1 2 5
0 π π π π π π
6 3 2 3 6
θ − , θ real , θ + , θentp

Fig. 2.16. Impact of the heating surface roughness on wetting angle hysteresis from
Cornwell [63]

The critical angle θentp is also presented in Fig. 2.16 as well as the boundary of vapor entrapment
in the boiling cavity. In the event of a cylindrical cavity whose opening has β = 0, then

(2.49)

The variation of angles θ – and θ+ as a function of surface roughness was experimentally


confirmed in the investigations conducted by Shepherd and Bartell [244] and also Dettre and
Johnson [75]. Although the coefficient s, in the way of reasoning adopted in those works, is treated
as a roughness measure, it cannot, however, be directly connected with roughness as pointed out
by Wenzel [316], the researcher who originally proposed to introduce that coefficient.

Experimental determination of the coefficient s is usually carried out by investigating gas


adsorption on the surface. Metal surfaces with machine finishes show values of s from 1.1 to
1.3, yet the data should be approached with caution due to the complex way of taking such
measurements and the results are sometimes ambiguous.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 31


Chapter 2

The considerations presented above show that for a wetting liquid in a boiling cavity, the shape
of the meniscus can change from convex to concave. Such a local change may result from the
wetting angle hysteresis and surface roughness. Fig. 2.17 shows how the liquid meniscus changes
inside a conical boiling cavity when changes take place in the meniscus curvature at position 2.
Position 2 corresponds to diagram (a) of Fig. 2.14, case (b) starts at position 3, whereas cases
(c) and (d) correspond roughly to the surroundings of the position 1, in which the change in shape
(meniscus curvature) occurred. Position 4 shows an intermediate situation as the interface moves
towards the mouth of the cavity while 5, 6 and 7 show the progressive change in shape and
meniscus radius at the mouth of the cavity until it reaches its stable position at 7.

Location of the center for meniscus radius


7
7
6
θ- 5
6
θ-
5
4

4
θ-
Variation of the 3
2
meniscus shape
3
(meniscus curvature )
1 2

1→ ∞

Meniscus radius

Fig. 2.17. Movement of a meniscus inside a conical boiling cavity from Cornwell [63]
where the numbers inside the cavity refer to the location of the interface
while the graph at the right depicts the corresponding radius of the
meniscus

In accordance with the stability criterion (2.46), the area inside the nucleus where the changes in
the meniscus position are indicated with a numbered line at the right, constitutes the zone in which
the growth of the nucleus is unstable. For the case shown in Fig. 2.17, there are two maxima in
the meniscus radius at points 3 and 6. The meniscus radius at point 3 is smaller than at point 6.
Therefore the radius at point 3 can be regarded as the critical one for the bubble growth process
and forms a barrier to the occurrence of boiling incipience. The result is that the boiling nucleus
radius at the heating surface boundary is usually bigger than the radius obtained from (2.11).

32 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Apart from the hysteresis caused by the surface microstructure itself, a hysteresis from
macroscopic effects is also found, resulting from impurities and heterogeneities in the surface.
Even if the surface is carefully prepared, it can get oxidized or covered with a coating formed by
contact of the boiling liquid or vapor with the surface.

If such effects are taken into account, the wetting angle is as follows:

σsg – σsl – πs
cos θreal = (2.50)
σlg

where �s depends on the character the surface impurities. The value �s is difficult to define and
calculate, consequently there are no means to make a qualitative evaluation of its influence when
a heating surface gets covered with an absorbed compound. Similarly, it is not possible to specify
changes in the surface characteristics with time, i.e., aging effects.

2.5. Summary

Despite a significant number of publications devoted to the subject, the review of the most
important factors characterizing active nucleation sites reveals that the phenomena occurring on
the heating surface are not fully understood, neither is a reliable qualitative description available.
The reason for this is the huge complexity of the phenomena, difficulties with conducting
microscopic observations of liquid-vapor interfaces inside minute cavities or providing a
mathematical description.

Summing up, very generally, the present state of knowledge on boiling site operation, the following
comments can be made:

1. The topology of active boiling sites on the heat transfer surface is fairly well
understood.

2. Hypotheses concerning boiling site activation inside complex shaped cavities in


a heating surface have been validated indirectly. The fact that it is not possible to
completely define all the surface characteristics makes it impossible to formulate a
general theory describing the process in an unambiguous way.

3. Numerous works have dealt with boiling site density. The only conclusion, confirmed
many times, is the relationship of the form:

2
N~q

This relationship, however, lacks a satisfactory description of the proportionality


coefficient required to convert this relationship into an equation. This arises from our
insufficient knowledge of the heat transfer surface characteristics.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 33


Chapter 2

Recently, attempts have been made at providing a quantitative description of the


mutual effect the neighboring boiling sites have on each other. Although the approach
seems successful to some extent, it does not yet result in a proven mathematical
description of the phenomenon.

4. The most serious drawback of all theories, including those presented here, is the
ambiguity and high dependence on the exact experiment conditions. In particular,
the influence of the wetting angle hysteresis has to be accounted for. The hysteresis
effect, although understood, has not so far been described quantitatively in a very
satisfactory manner.

34 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Chapter 3
3. Enhancement of Boiling Heat Transfer:
Techniques and Experimental Results

3.1. Types of micro-structured surface enhancements for boiling

Boiling heat transfer enhancement is one of the major issues in modern thermal technologies.
Large-scale integrated circuits, aerospace avionics and nuclear technology devices need cooling
systems that are highly effective and reliable. Heat transfer enhancement is also important in
refrigeration, chemical engineering, air conditioning, heat pipes, etc. Heat transfer enhancement
also has important economic aspects as it leads to much more compact and lighter weight heat
exchangers. They may also be cheaper to operate if they consume less energy by reducing
pressure drops. The research effort is therefore aimed at developing new enhancement
geometries, at working out technologies for their manufacture and obtaining test data with the
refrigerants and process fluids of interest.

The methods of boiling heat transfer enhancement can be divided into two categories:

1. Active methods: these refer to vibrating or rotating the heating surface, regular
scraping of the surface, ultrasonic vibration of the evaporating liquid, application of
an electrostatic field and other such techniques, usually rather complicated and not
particularly productive.

2. Passive methods: these refer to micro-structured surfaces of heightened roughness,


e.g. pitting obtained chemically, non-wetting spots applied to the surface (e.g. of
PTFE), finned or knurled surfaces, finely cut grooves or cuts, depositing a porous
layer, etc. There are also techniques, where the modification of the heating surface
is not necessary, e.g. wire, metal tapes or wire meshes wrapped onto the outside of
boiling tubes.

The so-called “combined” surfaces rely on two or more of the above-mentioned heat transfer
enhancement techniques, used at the same time. The literature on the subject is quite extensive
[131, 173, 271, 288, 311, 312]. The discussion of the results produced by the monograph’s authors
and those obtained by other researchers as well as the publications arranged in tables can also
be found in Polish publications [58, 202, 252, 261, 317].

The enhancement is achieved through the change in the heating surface structure, which consists
first in increasing the number of active boiling nuclei and the stabilization of their operation,
secondly in increasing the wetted surface area relative to the original base area, and thirdly in

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 35


Chapter 3

modifying the heat transfer mechanisms themselves to increase their effectiveness. Another factor
that has to be taken into account in the investigations is that the choice of micro-structured surface
is limited to those that can be manufactured on a mass scale at reasonable prices. The diversity
of enhancements already used is enormous. Some of them are patented and have been given
names by their manufacturers. Some examples of such micro-structured surfaces are presented
in Fig. 3.1, according to [166, 196, 288].

a) Porous layers made of sintered metal particles from 10 to 100 µm in diameter – High
Flux (formerly of Union Carbide and now belonging to UOP); b) Deformed finned tubes with
cut and bent over edges, with pore diameters of approximately 100 µm – Thermoexcel-E
(Hitachi); c) Knurled and compressed fins – Turbo-B (Wolverine Tube); d), e) and f) Split and
compressed fins: d) Gewa-T; e) Gewa-TX; f) Gewa-TXY (Wieland-Werke)

Fig. 3.1. Examples of industrial enhancements for boiling from Nakayama [166], Pate,
Ayub, and Kohler [196] and Thome [288]

Numerous works have dealt with boiling surfaces, on which cavities, fins, cylindrical openings
with different opening inclination angles, etc. were made and arranged in an ordered manner.
Researchers have often attempted to look for the optimum geometry and dimensions of the micro-
structured surfaces (some of which are capillary-porous structures) that would maximize heat
transfer. Some of the investigated surfaces had their roughness increased through mechanical
methods, e.g. by sanding [58, 131, 288, 311].

It should be stressed that the diversity of geometries and forms of roughness that can be given
to a surface is endless and that is why no one can expect to get a clear answer to the following
question: What kind of micro-structured surface is the best for enhancing boiling heat transfer?

The existing types of micro-structured surfaces have both advantages and drawbacks. Metal,
fibrous and sintered porous coverings have a high skeletal thermal conductivity and porosity.
6 –2
Therefore, it is possible to attain heat fluxes over 10 W m . Their manufacturing technology,
however, is complex, although it is possible to obtain and control the desired dimensions of the
structure (porosity and layer thickness, for example).

36 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Meshes have been widely applied due to the wide range available and the fact that the
production of such structures has been mastered; however, they are typically still too expensive
or complicated to use in industrial applications, except in heat pipes. Structures made of sintered
powders and particles have numerous advantages but are characterized by increased brittleness,
and therefore they are not appropriate for dynamic loads. Grooved structures have high thermal
conductivities and they can operate at high heat fluxes, but they are labor-intensive to manufacture
and the capillary pressure formed in such structures is small. Structures obtained with plasma
or flame spraying typically provide very good thermal performance but are not widely applicable,
because it is not easy to manufacture them to precise dimensions and it is difficult to obtain high
porosities.

3.2. Experimental results and empirical prediction methods

3.2.1. Industrially manufactured micro-structured surfaces

Addressing our attention to practical applications, the most frequent experimental investigations
are those that focus on industrially manufactured enhancements, such as those shown earlier in
Fig. 3.1.

Bergles and Chyu [26, 27] for example investigated heat transfer on metallic High Flux (Trademark
of UOP) surfaces. They were manufactured by covering a copper surface with copper particles
in such a manner so that the porosity ranged from 50 to 65%. The samples had a constant layer
thickness of 0.38 mm and were made with particles that ranged from 44 to 74 µm in diameter.
Investigations were carried out for water and R-113 and Fig. 3.2 depicts their boiling curves. In
Fig. 3.2, the natural convection curves were computed in accordance with the McAdams
expression and the investigated samples were denoted as follows:

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 37


Chapter 3

The conclusions drawn from their study were as follows:

1. On the porous coated surfaces, the boiling process was very stable.

2. In comparison with a smooth surface, boiling incipience on the porous coated


surfaces required much smaller superheating.

3. Nucleation hysteresis was observed on the boiling curve, and in addition, stronger
hysteresis effects manifested themselves for R-113 than for water.

4. Heat transfer coefficients on the porous coated surfaces were up to 8 times those for
smooth surfaces.

The authors attributed the boiling hysteresis to the effect of the porous surface being flooded with
a liquid in such a way that only small diameter boiling cavities could get activated. In accordance
with the accepted theory of boiling nucleation, the cavities getting activated should have a shape
that allows the return of vapor to them. Furthermore, the nucleation process is affected by the
wetting ability, characterized by the wetting angle. For R-113, the wetting angle is smaller than for
water, and therefore penetration of the boiling cavities by the liquid is easier, which results in a
more prominent boiling hysteresis.

Czikk, O’Neill and Gottzmann [69] conducted extensive investigations on different High Flux
surfaces, changing fluids, pressures, heat fluxes, coatings, etc. Measurements were also taken for
mixtures, for subcooled boiling and for forced convection conditions. The summary of the results
for pool boiling with pure fluids is presented in Table 3.1.

Yilmaz and Westwater [323] investigated boiling on a smooth tube and various industrial
enhanced tubes. The surfaces were denoted: ECR-40, Gewa-K, Gewa-T, High Flux, CSBS and
Thermoexcel-E. The investigated liquid was isopropyl alcohol. The authors found both an increase
in the heat transfer coefficient and the critical heat flux on these enhanced surfaces compared to
the smooth tube.

38 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

a) b)
5
10 105
Experiment with Empty symbol - heat flux
temperature leap increase
Full symbol - heat flux
decrease
10 4 104
-2

q, W⋅m -2
q, W⋅m

1-F-D-22-C

on
c ti
1-F-D-4-C

ve
n
co
10 3 103
n
2A-W-B-56-C

ral
o
cti

tu
2A-W-B-39-C
ve

Na
n

2A-W-B-69-C
co
al

2A-W-B-1.5-C
r
tu
Na

10 2 102
10-1 10 0 10
1
102 10-1 10 0 10
1
10 2
DTsl , K DTsl , K

Fig. 3.2. a) and b): Results of experimental investigations for water and R-113 boiling
on surfaces with High Flux coating from Bergles and Chyu [26, 27]

c) d)

10 5 105
2B-F-B-15-C
2A-F-B-0.5-S 2B-F-B-2-C
2A-F-B-12-C 2B-F-B-2.6-S
2B-F-D-12-C
2A-F-D-4-C
10 4 2A-F-D-0.5-S 104 2B-F-D-9-S
-2

q, W⋅m-2
q, W⋅m

on
cti

103
ve
n

10 3
tio

n
co
ec
nv

al
co

tur
l

Na
ra
atu
N

10 2 102
10 -1 10 0 10
1
102 10-1 100 10
1
102
DTsl, K DTsl , K

Fig. 3.2. c) and d): Results of experimental investigations for water and R-113 boiling
on surfaces with High Flux coating from Bergles and Chyu [26, 27]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 39


Chapter 3

Table 3.1. Comparison of properties of the High Flux coating to smooth surfaces,
according to Czikk, O’Neill and Gottzmann [69]

Properties High Flux Surface Smooth Surface


Effect of pressure α increases proportional to pressure to Usually changes with pressure to the
on heat transfer the power 0.5 or higher power of 0.3 to 0.4
Slope of boiling Typical slope is 3 or more but varies Typical slope is about 3
curve (α vs. q depending on the kind of porous surface
or q vs. ΔT in
logarithmic scale)
Surface roughness Boiling cavity dimensions could be Increase of roughness leads to an
and distribution of changed when manufacturing the increase in α. The distribution of boiling
boiling cavity sizes porous surface. Both slopes of α and cavity dimensions affects the slope of the
q vs. ΔT change as a function of cavity boiling curve.
dimensions
Boiling hysteresis Temperature leap observed in certain Temperature leap observed for certain
fluids for heat fluxes usually below combinations of surface and boiling liquid
–2
3kWm
Critical heat flux At least equal to the value for the smooth Independent of the surface character
surface but up to 70% for some surfaces
Heating surface Seldom observed with non-corrosive Generally observed
aging liquids
Heat transfer Small influence only Small influence on α for pool boiling
surface orientation

Nishikawa and Fujita [173], apart from their own results, showed a composite graph of selected
results of investigations on High Flux surfaces conducted by the manufacturer [67, 90, 181], shown
here in Fig. 3.3. For all the investigated liquids (R-11, propylene, ethanol and water), considerable
heat transfer enhancement was seen when compared to boiling on a smooth surface. In fact,
enhancement ratios larger than 10 were achieved for these fluids using a porous coating
optimized for each fluid.

40 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

2·10 5
2
3 4
5
1 4 3 2 1
10

-2
q, W⋅m

104 High Flux Smooth surface


3
6·10 1
2·10 -1 100 101 4·10
DTsl , K

Fig. 3.3. Boiling heat transfer enhancement with High Flux surfaces: 1 – R-11,
2 – propylene, 3 – ethanol, 4 – water with diagram taken from Nishikawa
and Fujita [173]

Marto and Lepere [156] carried out experiments with three types of enhanced tubes, all
15.8 mm in diameter: High Flux, Thermoexcel-E and Gewa-T. Measurements were taken for R-113
2 5 –2
and FC-72 for boiling at atmospheric pressure over heat fluxes from 10 to 2 ∙ 10 W m . Prior
to the measurements, the micro-surfaces had been aged by boiling for long periods of time. The
investigations demonstrated that with R-113 boiling improved heat transfer 2 to 10 times when
compared with a smooth tube. For FC-72, the increase was from 2 to 5 times. The degree of
superheating necessary to reach boiling incipience depended on the prior history of the boiling
enhancement and the properties of the boiling fluid. The authors also provided an assessment of
the efficiency of various kinds of micro-structured enhanced boiling surfaces. The authors [156]
declared they obtained the best results while using the High Flux surface, which in their tests was
a tube coated with small copper particles (46% of the diameters smaller than 44 µm and 54% of
the diameters ranging from 44 µm to 74 µm). The layer thickness was 0.08 mm.

Fig. 3.4a shows their measurements for a High Flux surface with R-113 as compared with the
results of Bergles and Chyu [26, 27] at similar test conditions. The comparison shows good
agreement of the data sets. Fig. 3.4b presents their data for the Thermoexcel-E type of surface
(this surface had a layer thickness of 0.19 mm and an average pore diameter of 0.1 mm).

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 41


Chapter 3

a) b)
10 6 106 Aged surface- A
Empty symbol- heat flux
increase Aged surface- B
Full symbol - heat flux Aged surface- C
decrease Smooth tube
10 5 105 data

Boiling
incipience
q, W⋅m -2

q, W⋅m -2
Natural
4 convection [37] 4
10 10
Bergles´ data
[29, 30]

10 3 103

R 113 R 113

10 2 -1 102 -1
10 100 101 10 2 10 10 0 10 1 102
DTsl, K DTsl , K

Fig. 3.4. a) Comparison of Marto and Lepere [156] data for High Flux surface to those
obtained by Bergles and Chyu [26, 27], including the natural convection
curve; b) Comparison of Thermoexcel-E data to a smooth pipe where letters
A, B, C denote various means of sample aging from Marto and Lepere [156]

Later investigations of Marto and co-researchers [158] showed that the performance of the
Gewa-T surface depended on the distance between T-fins and their density per the tube length
unit. Their optimum values for a copper pipe and R-113 were determined.

30
q = 10 4 W
⋅m –2
Heating surface superheating DTsl, K

1.5⋅10 4 q = 2⋅10 4 W 0. 5⋅ 10 4
10 ⋅m –2
1⋅ 10 4
0.1⋅10 4

q = 2⋅10 4
W ⋅m –2
1
Water, rough surface 0.5⋅ 10 4

R-11, rough copper surface

R-11, Thermoexcel-E
0.1 2
10 10 3 104 105 3⋅105
Active boiling nuclei density N a, m-2

Fig. 3.5. Boiling superheat as a function of the boiling site density, according to
Nakayama [166]

42 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

In his survey paper [166], Nakayama provided information on boiling heat transfer with reference
to the cooling of electronic devices. Fig. 3.5 shows the dependence of boiling site density on the
temperature difference ΔTsc for water and R-11 boiling. The dashed lines at the right represent
boiling on the Thermoexcel-E type surface. This kind of micro-structured surface is seen to have
the best characteristics, initiating boiling at superheats even below 1 K. Nakayama emphasizes
another important aspect of electronics cooling, namely the onset of boiling. If the onset of boiling
is manifested as a large temperature jump occurring at relatively high heat fluxes, such operation
would be dangerous to the device being cooled. The detailed data on the phenomenon bear a
large amount of uncertainty and with a note of caution, we can conclude that micro-structured
surfaces usually have smaller incipient superheats than smooth surfaces. Moreover, the type of
liquid affects the onset of boiling. For instance, the transition from natural convection to boiling in
the liquid FC-72 takes place at a smaller temperature difference than for R-11.

2⋅106
6
10 WSP

WEP
22EC

105
12EC
113SC
-2

11EC
q, W⋅m

22SC
N 2 EP
4 12SC
10
HeEP 113EC

11SC

103 HeSP N2SP

3⋅102
10-1 100 10 1 3⋅10 1
DTsl , K

WSP — Water, smooth, plane; WEP – Water, Thermoexcel, plane; 11SC – R-11, smooth, cylinder;
11EC – R-11, Thermoexcel, cylinder; 22SC – R-22, smooth, cylinder; 22EC – R-22, Thermoexcel,
cylinder; 113SC – R-113, smooth, cylinder; 113EC – R-113, Thermoexcel, cylinder; N2SP –
Nitrogen, smooth, plane; N2EP – Nitrogen, Thermoexcel, plane; HeSP – Helium, smooth, plane;
HeEP – Helium, Thermoexcel, plane

Fig. 3.6. Heat transfer enhancement for selected liquid/enhancement combinations


for boiling on Thermoexcel-E geometries, according to Nishikawa and
Fujita [173]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 43


Chapter 3

Heat transfer enhancement was also investigated with respect to cryogenic liquids with a view
to improving the operation of cooling devices operating at very low temperatures. Experiments
were conducted primarily for helium and nitrogen. In their work [173], Nishikawa and Fujita quoted
earlier results obtained by Aria et al. [14], which dealt with the boiling of different liquids, including
helium and nitrogen, on flat and cylindrical surfaces, both smooth and with Thermoexcel-E
surfaces, which are summarized in Fig. 3.6. The Thermoexcel-E surface diminishes by almost
ten times the superheat necessary for boiling incipience and significantly increases the heat
dissipated, especially for small superheats.

Besides the above tests with commercially available enhancements, numerous investigations
have been performed for surfaces covered with capillary–porous structures of pre-set dimensions
created in laboratories. The pre-set quantities are the density and the dimensions of artificially
created boiling pores in the form of channels, openings and micro-fins obtained by the mechanical
working of the heating surface, which are somewhat similar to industrially manufactured surfaces.

44 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

2.0
2

-2
q, W⋅m
4 5
1.5
1
6
3
1.0
5 10 20 25
5 R 11 DTsl, K
psat = 0.23 MPa
Tsat = 323 K
1 - smooth surface
105 2 - U5
q, W⋅m -2

5 3 - C10-5-1
4 - C10-5-2
5 - U10
6 - U15
104
5 6 5

1
103 4 3 2

5
5 10-1 5 100 5 101 5
DTsl , K

(a) Pore diamters (µm) and


# Designation (b) Relative number densities – RND (%)
(a) Large (b) RND (a) Small (b) RND
2 U5 — — 50 100
3 C10-5-1 100 8.3 50 91.7
4 C10-5-2 100 33.3 50 66.7
5 U10 100 100 — —
C15-10 150 8.3 100 91.7
6 U15 150 100 — —

Fig. 3.7. Boiling curves for different surfaces from Nakayama, Daikoku and
Nakajima [169]

Nakayama, Daikoku and Nakajima [169] conducted measurements for R-11 boiling on the surfaces
with the enhancement structures made as follows. First, a number of parallel tunnels were made in
a flat plate. These tunnels, spaced at equal distances from one another (tunnel pitch 0.5 mm), had
cross-sections of 0.25 × 0.4 mm. Then this plate was covered with another one. The cover plate
was thin and attached with the previous one. Pores of 50, 100 and 150 µm in diameter (pore pitch
was 0.7 mm) connected the resulting re-entrant channels with the outside. The openings were
distributed in a regular grid. The relative number densities of large and small pores are given in the
table in Fig. 3.7. In this way, a system of artificial boiling sites, connected with one another through
a set of tunnels, was constructed.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 45


Chapter 3

Fig. 3.7 shows the results of their boiling heat transfer measurements. The symbols at individual
curves denote different distributions of the diameters of the openings on the surface. The following
conclusions were drawn from their work:

1. The density and diameters of pores determined the character of the boiling process at
4 –2
heat fluxes larger than 3 to 4 · 10 W m . On surfaces with different pore diameters,
the pores with the highest density controlled the level of heat transfer enhancement.

2. At low heat fluxes, the largest pore determined the heat transfer performance.
Saturation pressure also affected the boiling process to a considerable extent. The
combination of elevated pressure and appropriate pore dimensions made it possible
for boiling incipience to take place at very small superheats.

3. The occurrence of boiling at very small superheats can be explained due to the
contact of the saturated liquid with the heat transfer surface inside the pores, which
causes the evaporation of the liquid sucked in to the re-entrant channels.

4. Capillary superheating, resulting from menisci formation in the pores, can partially
explain why the boiling curves show different trends at different distributions of pore
dimensions. An appropriate proportion of the number of active and inactive pores can
constitute an important factor affecting the shape of the boiling curves.

Fig. 3.8. Spot drilling of holes in a vertical plate: a) horizontal orientation of the
holes, b) vertical orientation, c) obliquely drilled holes. Dimensions in mm,
according to Anderson and Mudawar [10]

46 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Anderson and Mudawar [10] conducted an investigation into cooling with dielectric fluids with an
enhanced boiling surface. Their investigation focused on a system using a simulated electronic
element that constituted a vertical plate with FC-72 as a cooling liquid. The measurements were
taken for three kinds of surfaces: a smooth one, one with artificial openings (drilled holes) and
one with small fins. Fig. 3.8 shows the surface arrangement with the drilled holes. Moreover, the
smooth surface was also roughened so that the effect of micro-roughing could be obtained.

Fig. 3.9 shows the results of investigations for a smooth and a micro-roughened surface and also
a surface with a row of artificial openings (cf. Fig. 3.8). We can see that the presence of artificial
boiling sites causes a greater temperature drop at the transition from natural convection to boiling.
It should be mentioned that only a slight change in the value of the critical heat flux was observed.

a) b)
3·105 3·105
Critical heat
flux
105 Boiling 10 5
7 7
incipience
5 5

3 3
-2

q, W⋅m-2
q, W⋅m

104 10 4
7 7
5 5

3 3
FC-72 FC-72

3 3
10 10
100 101 102 100 101 102
DTsl, K DTsl, K

a) Δ – polished surface, ○ – sand blasted surface, □ – surface smoothed by liquid


honing

b) Δ – horizontal hole orientation (Fig. 3.8), □ – vertical hole orientation, ○ – oblique


spot drillings

Fig. 3.9. The impact of roughness and micro-drilled holes in a flat vertical surface on
nucleate boiling, according to Anderson and Mudawar [10]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 47


Chapter 3

106
Critical heat flux

105
q, W⋅ m -2

Boiling
incipience

3 2 1
104

FC-72

103
100 10 1 102
DTsl , K

1 – micro-fins; 2 – micro-pin fins; 3 – micro-fins, data by Nakayama et al.,


according to Nakayama, Nakajima and Hirasawa [170]

Fig. 3.10. The impact of micro-fins and micro-pin fins on the heating surface on
nucleate boiling, according to [10]

Fig. 3.10 shows the results of investigations into surfaces with micro-fins and micro-pin fins,
according to Nakayama, Nakajima and Hirasawa [170]. In this case, the temperature drop at the
transition from natural convection to boiling is bigger than before, whereas the value of the critical
heat flux increased slightly. Moreover, in the boiling zone, the temperature difference is much
smaller than for the smooth surface at the same heat flux. The conclusion drawn by these authors
is that the generally accepted boiling incipience theories based on boiling site diameter do not
apply to liquids that wet the heating surface well, such as FC-72.

Peterson and Ortega [198], in their survey paper on the cooling of electronic devices, also
discussed the issues of boiling heat transfer enhancement, commenting on the aforementioned
results presented by Bergles and Chyu [26, 27] and also those of Anderson and Mudawar [10].
The authors provided some comments concerning the consequences for electronic systems of the
temperature overshoot and subsequent drop accompanying the transition from natural convection
to boiling. Though the investigations quoted in [198] demonstrated that, on developed micro-
structured surfaces, an increase in the number of active boiling sites takes place and these sites

48 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

resist flooding, no effective means was found to reduce the temperature overshoot. It is possible,
however, to estimate the magnitude of the overshoot for dielectric liquids and check whether the
device operating temperature limit is surpassed or not according to Bar-Cohen and Simon [23].

Investigations into the boiling properties of micro- structured surfaces manufactured by industries
have continued, focusing on increasingly detailed issues. Some exemplary topics could be
mentioned: the impact of oxidation on micro-surface effectiveness by Imadojemu, Hong and Webb
[101] for water and R-11 boiling on Gewa-SE copper tubes, the impact of surface-active agents on
the boiling of mixtures by Tang and Wang [270] for the boiling of water solutions of organic liquids
on a Gewa-T tube, the impact of oil on refrigerant boiling by Webb, Chien and McQuade [313]
for boiling on a Gewa-SE, Turbo-B and smooth tubes and the very important problem of solid
particle deposition on enhanced boiling surfaces by Bergles and Sommerscales [28] for boiling on
industrial High Flux, Thermoexcel-E, Gewa-T and Turbo-B tubes.

Webb, Chien and McQuade [313] demonstrated that R-123 contamination with a small amount of
oil, up to 2% in a mixture, does not affect the heat transfer coefficient values in a perceivable way
for a smooth surface or for the Gewa-SE and Turbo-B tubes.

Investigations into the effects of boiling mixtures of refrigerants R-11 and R-113 show that heat
transfer coefficients obtained on Turbo-B and High Flux tubes are always higher than those
obtained on smooth surfaces. With the mixtures of R-11 and R-113, the results were worse than
for pure refrigerants and High Flux (Trevin, Jensen and Bergles [297]).

Strong interest in industrially manufactured Gewa type of surface of Wieland-Werke stimulated


scientific co-operation of researchers from Germany, Belarus and Ukraine, as described by Mertz
et al. [160]. The subject of this joint investigation was capillary-porous structure of the Gewa-K,
Gewa-TWX and Gewa-PB tubes made of copper, steel and copper-nickel alloy. Simultaneously,
porous surfaces, manufactured from powders with the flame spraying technique and deposited
on steel pipes, were investigated. The third group of tubes tested were those with pores that
were spaced both symmetrically and non-symmetrically and of predetermined diameters and
depths. Their results confirmed those of earlier studies that with the enhancement structures
manufactured by industry, it is possible to obtain heat transfer augmentation ratios as high as four.
Sprayed coatings also produced similar results. It was found out that for the whole range of the
heat fluxes, the best results were achieved for the Gewa-K and Gewa-TWX geometries [160].

Experimenting with Gewa-K and Gewa-T tubes, Ayub and Bergles [19] tried to experimentally
determine the optimum fin gap for water and R-113. For each liquid, different values of the fin
arrangement were obtained. One optimum gap was determined for R-113, equal 0.25 mm,
whereas for water, there were two optima: 0.15 mm and 0.35 mm. The most extensive set of
experimental data, presented in an orderly manner, is provided by the work of Webb and Pais
[315], which deals with cylindrical heating surfaces.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 49


Chapter 3

Correlations for high flux porous coated surface

Gottzman and O’Neill with co-researchers [68, 91, 182, 288, 310, 311, 312] derived two empirical
correlations necessary to compute the relationship between the boiling heat transfer coefficient
and the heating surface superheat. The correlations were developed on the basis of two heat
transfer models, put forward by them, of boiling in capillary-porous-structures, simulated by
layers of spheres. The models are constructed on the concept of two thermal resistances, the
first associated with the superheat necessary for the nucleation incipience and the second for the
superheat necessary for the evaporation of a liquid film which covers the channels of the capillary-
porous structure.

In accordance with the first model [91], the heat transfer coefficient correlation has the form:

(3.1)

where, on the basis of (2.6), the derivative (dp/dT) sat is described by the Gibbs equation:

(3.2)

The derivative (dp/dT) sat is the slope of the saturation curve p - T for a given liquid at the
experiment conditions.

The values of αmax, computed from (3.1), were compared with the experimental data for the High
Flux surface and the following liquids: water, R-11, R-22, R-113, nitrogen, oxygen, propylene,
ammonia, ethanol and toluene, see Fig. 4.7 (section 4.2.2). The investigations were conducted
5 5 5
at three pressures: 0.14 ∙ 10 Pa; 0.39 ∙ 10 Pa and 1.03 ∙ 10 Pa. The experimental constant m
in (3.1) was determined with respect to the experimental data with the method of least squares.
The majority of experimental results showed relatively good congruence with computed results
assuming a regular sphere packing. Still, there was a lack of congruence with some of data,
namely for ammonia and toluene with errors on the order of 100%. These large errors can be
partially attributed to the inaccuracy of the heat transfer coefficient measurement at the very small
superheats observed, below 1 K.

On the basis of the other model [182], which was an improvement on the previous one, Webb [310,
311] introduced a more detailed relationship for describing the heating surface superheat:

(3.3)

50 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where dgrn is the particle diameter of the sintered powder and δ is the porous layer thickness.
Equation (3.3) yields a considerable discrepancy with respect to the data given by Shakir et al.
[242] and also Webb [310]. For instance, the error in the calculation of the surface superheat for
water amounted to 56% and 74% for methanol [242]. Significant discrepancies were also found for
R-11 and R-113 for various other micro-structured surfaces [310].

Correlation for Gewa-T surfaces

Xin and Chao [318] built a physical model of boiling in the channels of a flat Gewa-T surface,
discussed also in 4.2.2. According to their model, they assumed the behavior of a liquid film
covering the channel walls inside the structure depends on the dynamics of the counter-current
two-phase flow between the re-entrant channel and the external boiling liquid volume. On this
basis, the general dependence for the Nusselt number was derived as follows:

(3.4)

where h is the channel height underneath the T-shaped fin, l is the width of the re-entrant channel
and D is the gap opening between the T-fins through which the vapor bubble departs from the re-
entrant channel. The similarity numbers in (3.4) were defined as follows where s is the pitch of the
T-fins:

Nusselt number
(3.5)

Archimedes number
(3.6)

Reynolds number
(3.7)

Weber number
(3.8)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 51


Chapter 3

The relationship between the total heat flux and that used to evaporate the film is as follows:

(3.9)

and the Prandtl number is

(3.10)

Applying a regression method, Xin and Chao carried out the analysis of their experimental data
obtained at atmospheric pressure for ten different copper Gewa-T surfaces for water and ethanol
as well as six Gewa-T surfaces for R-113 [318]. In this study, the gap D changed from 0.09 to 24
mm, two channel widths were tested of 0.6 and 0.8 mm, two pitches of 0.8 and 1.2 mm, and two
heights h of 0.5 and 1.0 mm. The following empirical parameters were obtained eventually:

C = 3.76 m = – 0.15 n = 0.29 p = 0.76 (3.11)

The ranges of similarity numbers for the examined liquids were:

0.016 < Rel < 7.03 0.0000014 < We < 0.17 1.76 < Prl < 7.86 (3.12)

Expression (3.4) with the empirical constants given by (3.11) correlated with the experimental data
with the accuracy of ± 30%.

Thome [288] compared the above correlation to data for Gewa-T tubes with respect to the
experimental data for refrigerants (R-11, R-12, R-22, R-113) as well as for isopropanol and
p-xylene. The mean errors ranged from 74% to 2%. It was stated in the conclusion that Xin and
Chao’s correlation did not in general yield good results for refrigerants.

Wang, Cheng and Zhang [304] conducted investigations for boiling on Gewa-T and smooth tubes.
5 5
The liquids tested were R-113 and ethanol at pressures ranging from 1 ∙ 10 Pa to 6 ∙ 10 Pa.
3 5 -2
Heat fluxes ranged from 3 ∙ 10 to 1.5 ∙ 10 Wm . The Gewa-T tube increased the heat transfer
coefficient by 1.5 - 6 times in comparison with a smooth tube as well as exhibiting a pressure
influence in the investigated range.

52 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The model put forward by Wang, Cheng and Zhang [304] further developed the concept of Xin
and Chao [318], also discussed in 4.2.2. Applying dimensional analysis, the following similarity
numbers were isolated and defined:

(3.13)

where the notation for We was presented in formulas (3.4) to (3.10), Ja is Jakob number and dpipe
is the diameter of the tube over the enhancement. Using regression analysis on their experimental
data, the following correlation was obtained:

(3.14)

The experimental ranges of the similarity numbers were as follows:

(3.15)

According to the authors [304], 90% of their experimental data fell within ± 30% of their correlation
(3.14).

Ayub and Bergles [20] made an attempt, on the basis of experimental investigations on water and
R-113 boiling at atmospheric pressure, to develop a correlation for the Gewa-TW surface. The
authors, however, did not provide a method to determine their two empirical constants, which
are functions of fluid properties. In order to calculate the heat flux for a given wall superheat, it is
also necessary to know the density of nucleation sites, which Ayub and Bergles [20] determined
experimentally for the surfaces they investigated as a function of ΔTsl and fit in the form of a
polynomial. The above-mentioned factors include the interesting effect of boiling site density on
the heat transfer process but on the other hand limits the applicability of this approach. The model
presented by Ayub and Bergles is also discussed later in this book in section 4.2.2.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 53


Chapter 3

Correlation for Thermoexcel-E surfaces

The approach of Xin and Chao [318] discussed above was also applied to the Thermoexcel-E
enhanced boiling surface in [318]. The following parameters were used to characterize the
geometry: the channel width l, its height h, the fin pitch s, the pore diameter Dp and the pore
pitch along the channel sp. Proceeding as before, the empirical constants in the Nusselt number
correlation were determined using the experimental data by Nakayama et al. [167, 168] for water
5
and R-11 at the pressure of 1 ∙ 10 Pa. The following expression was obtained:

(3.16)

where the Nusselt, Archimedes and Reynolds numbers are defined by equations (3.5), (3.6) and
3 2
(3.7). The gap width D in the Weber number (3.8) was replaced with the ratio Dp / sp, which yields

(3.17)

The accuracy of this correlation was about ± 20%. According to Thome [288], it is not possible to
verify (3.16) with respect to other Thermoexcel-E surfaces as the information concerning the pitch
sp are not available.

The presented survey shows that several correlations to calculate the Nusselt number or the
heating surface superheat Tsl – Tsat have been proposed for a few industrially manufactured
micro-structured surfaces. They are valid for very restricted ranges of similarity numbers and one
type of geometry for each correlation, which means that experimental investigations will have to
continue and are indispensable as these geometries are changed by their manufacturers to further
improve their thermal performance.

Summing up, it can be stated that the experimental investigations on industrially manufactured
enhanced boiling surfaces result in the following general conclusions:

a. the density of boiling sites on such surfaces is much higher than on smooth ones;

b. hysteresis at boiling incipience is clearly evident, being more evident the better is the
heating surface wetting;

c. in some cases, the hysteresis causes a significant shift of the boiling incipience
towards greater differences between the surface temperature and the boiling liquid
saturation temperature;

d. if hysteresis does not occur or is negligible, boiling starts at considerably lower


superheats on micro-structured surfaces than on smooth surfaces;

54 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

e. boiling on industrial micro-structured surfaces yields much higher heat transfer


coefficients than smooth surfaces, especially at low heat fluxes;

f. The critical heat flux is found to be equal to or higher than the values obtained for
smooth surfaces.

3.2.2. Sintered powder structures

Although powder structures have been investigated experimentally and theoretically for many
years and are quite easy to make in laboratories, they have not been manufactured on industrial
scale. Modeling of the boiling process in such structures often assumes that they are made of
connected spheres, which simplifies the considerations considerably.

2
R 113 - atmospheric pressure

1 4
105
-2
q, W⋅m

5
3
4
10

2
3 10 0 3 3
10 1
DTsl, K

1 – copper, dgrn = 250 µm, δ = 1 mm; 2 – bronze, dgrn = 100 µm, δ = 1 mm; 3 – bronze,
dgrn = 250 µm, δ = 1 mm; 4 – bronze, dgrn = 500 µm, δ = 3 mm; 5 – horizontal copper
cylinder, rough surface (abrasive paper), according to Nishikawa and Fujita [173]

Fig. 3.11. Heat transfer data for surfaces covered with sintered powder structures
characterized by the following two dimensions: dgrn – powder particle
diameter, δ – porous layer thickness

In their survey, Nishikawa and Fujita [173] presented their results depicted here in Fig. 3.11 for
boiling curves of q vs. ΔT for R-113 boiling at atmospheric pressure [176]. Their capillary-porous
structures were produced by sintering bronze or copper particles to the surface of a tube. The
heating surface superheat was greatly diminished whereas the heat transfer coefficient increased
correspondingly, especially at low heat fluxes (and small superheats). The results for surfaces with
porous coverings were compared with those for a rough surface.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 55


Chapter 3

Fig. 3.12 shows the impact of the thickness of a layer made of sintered copper particles on the
heat transfer coefficient. Here we can see there exists the optimum layer thickness at which the
highest heat transfer performance is obtained.

δ/d
0 4 8 12 16
4
7⋅10

6⋅104
1⋅10 4 W⋅m-2
4 -2
5⋅104 3⋅10 W⋅m
7⋅10 W⋅m-2
4
α , W⋅m-2⋅K -1

4⋅104 15⋅104 W⋅m-2

3⋅104

2⋅104

1⋅104

0
0 1 2 3 4
δ, mm

Fig. 3.12. Impact of the thickness of a porous layer of sintered copper powder on the
nucleate boiling heat transfer coefficient for R-113 at atmospheric pressure
with a powder particle diameter dgrn = 250 µm, according to Nishikawa and
co-researchers [173, 177]

Nishikawa and co-researchers [173, 177] proposed two empirical correlations, presented below,
resulting from their investigations on R-11, R-113 and benzene boiling on surfaces with sintered
layers of bronze and copper particles of the diameter dgrn = 100 to 1000 µm. Sintered layer
thicknesses were from δ = 1 to 5 mm and porosities from ε = 0.38 to 0.71:

(3.18)

I C = 0.639 a = 0.0658 b = 0.626 c = 0.665


d = – 0.692 e = 0.904

–3
II C = 1.00 · 10 a = 0.0284 b = 0.560 c = 0.593
d = – 0.708 e = 1.67

where

λef = ε λl + (1 – ε) λs (3.18)

56 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The index s refers to the porous material while the values for I are from [177] and those for II are
from [173]. Correlation II is the one most frequently quoted and is that shown in Fig. 3.13. The
comparison of the same results with correlation I yields a similar congruence [175].

102

7
qδ / (l ef DTsl )

3
0%
+3

3
10
0%

7
-3

5
2 4 8 2 4 6 8 2 4
102 10 3
0.56 0.593 −0.708 1.67
 δ   q d grn   ρl 
0.0284
 σlg h lg   λ ef   
 2 2       
 d grn   ε h lg μ g   ρg 
 q δ       λl   

Fig. 3.13. Comparison between correlation (3.18, II) and experimental results,
according to Nishikawa and Fujita [173]

Figure 3.14 (a and b) from Nishikawa and Fujita [173], on the other hand, compares different kinds
of enhanced boiling surfaces (those manufactured industrially and those prepared in laboratories),
based on R-113 and water boiling results at atmospheric pressure. The boiling curves show
that surfaces with porous sintered layers give the best results. Generally, both laboratory-made
porous sintered layers and industrial ones (and mechanically formed porosity) like High Flux (and
Thermoexcel-E), significantly increase heat transfer in comparison with rough surfaces.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 57


Chapter 3

a) b)
5⋅105 2 5⋅10 5
R 113
Atmospheric pressure

105 10 5 3
1
q, W⋅m-2

-2
q, W⋅m
3 4 5 6 7
1 2 4
4
104 10
Water
Atmospheric pressure 5 6
3
2⋅103 2⋅10
2⋅10-1 100 101 5⋅101 2⋅10 -1 100 101 5⋅101
DTsl, K DTsl, K

a) 1 – porous layer of copper, dgrn = 350 µm, δ = 0.4 mm; 2 – Thermoexcel-E;


3 – Thermoexcel-E; 4 – porous layer of bronze, dgrn = 350 µm, δ = 0.94 mm;
5 – High Flux; 6 – rough surface (machine made); 7 – rough surface (abrasive paper)

b) 1 – porous layer of bronze, dgrn = 350 µm, δ = 0.94 mm; 2 – porous layer of copper,
dgrn = 3 µm, δ = 0.4 mm; 3 – High Flux; 4 – Thermoexcel-E; 5 – Thermoexcel-E;
6 – rough surface (abrasive paper)

Fig. 3.14. Comparison of boiling heat transfer enhancement properties for water (left)
and R-113 (right) between sintered powder layers and the industrial surfaces
High Flux and Thermoexcel-E, according to Nishikawa and Fujita [173]

Ito et al. [103] conducted a comprehensive investigation on boiling of nitrogen on a horizontal


cylinder covered with a porous layer obtained by sintering of bronze particles of 127 to 390 µm
in size. The investigations showed that thinner layers gave higher heat transfer coefficients.
Moreover, it was found that for each layer thickness, there exists an optimum size of sintered
particle diameter. Surprisingly, the authors also observed a new kind of boiling hysteresis in which
the boiling curves for increasing superheating are to the left of those for decreasing superheating
as shown in Fig. 3.15 for curves 1, 2 and 3. Usually the heat flux for decreasing superheating is
greater than that for increasing superheating. The authors suggested that there might exist two
kinds of hysteresis, the “normal” one at lower values of the heat flux (Fig. 3.15, curve 4) and the
inverse one – at greater heat flux values. The second type of hysteresis type was also observed
by Afgan and Jovic [7]. Afgan attributed this hysteresis to the capillary crisis, which can occur
at a thick porous layer made of small particles. The crisis could be caused by the liquid being
substituted with the vapor in the porous structure at heat fluxes lower that a critical value.

nd
Similar hysteresis phenomena, termed 2 type of hysteresis, were observed by Poniewski and
Wojcik for boiling on metal fibrous structures [203, 205, 211, 214, 216, 217, 218]. These authors
[203, 205, 211] presented a phenomenological interpretation of the process and proposed a model
that is coherent with their experimental data, described later in chapter 4.

58 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

5
10
8
6 Nitrogen
6
4
1
2
2

104
8
6
-2

3
q, W⋅m

4
5
2 4

103
8
6
4

2 Empty symbol - heat flux increase


Full symbol - heat flux decrease
102
10
-1 2 4 6 8 10 0 2 4 6 8 101 2
DT sl, K

1 – δ = 0.74 mm; 2 – δ = 0.96 mm; 3 – δ = 1.91 mm; 4 – copper (technically smooth); 5 and
6 – curves determined from the theoretical model, according to Ito, Tanaka and Tamari [104,
105] for δ = 0.74 mm and δ = 1.91 mm

Fig. 3.15. Boiling curves for sintered powder coverings of changeable thickness;
porous layers of bronze with dgrn = 127 µm, according to Ito et al. [103]

Fig. 3.15 also shows lines 5 and 6 predicted by the boiling model put forward by Ito, Tanaka and
Tamari [104, 105] for a porous structure constructed of spheres of known diameter distribution.
The model’s boiling curves give a good qualitative prediction of the results, but quantitative
agreement is not so satisfactory.

In another porous layer study, the liquids investigated by Zhang and Chen [327] were water
and ethanol and the experiments were conducted at atmospheric pressure. The dimensions of
the particles sintered to their flat copper surface ranged 100 to 400 µm and the porous layer
thicknesses were from 0.5 to 4 mm. The heat transfer coefficients were found to be 3 to 10
times those of a smooth surface. The existence of an optimum thickness of the porous layer was
observed. A qualitative (descriptive) analysis of the boiling mechanism was given and relationships
for calculating the heat transfer coefficient and the optimum layer thickness as a function of the
heat flux were proposed. Fig. 3.16 shows the results of their measurements, which are similar to
those by Nishikawa and Fujita, shown earlier in Fig. 3.12.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 59


Chapter 3

7 7
Water 3 4 Ethanol
6 4 -2 6
1 - q = 3⋅10 W⋅ m 1 - q = 3⋅104 W⋅ m-2
2 - q = 7⋅104 W⋅ m -2 2 - q = 7⋅104 W⋅ m-2
5 5
3 - q = 15⋅104 W⋅ m -2 3 - q = 15⋅104 W⋅ m-2
-1

α, W⋅m-2⋅K-1
4 - q = 30⋅104 W⋅ m -2 4 - q = 30⋅104 W⋅ m-2
α, W⋅m ⋅K

4 4 2
-2

4
1
3 3 3
2
2 2
1
1 1

0
0 1 2 3 4 0 1 2 3 4
d , mm d , mm

Fig. 3.16. Heat transfer coefficients plotted versus the porous layer thickness and heat
flux for water and ethanol from Zhang and Chen [327]

In order to express the heat transfer coefficient, Zhang and Chen [327] proposed the following
polynomial expression:

2 3 4 5
α=A+Bδ+Cδ +Dδ +Eδ +Fδ (3.19)

where δ is the porous thickness layer whereas the coefficients A to F depend on the fluid and
the heat flux, for a particular set of diameters of the sintered material. For the four different heat
fluxes and two boiling liquids tested, whose results are shown in Fig. 3.16, Zhang and Chen
[327] proposed eight different sets of empirical coefficients A to F, which means that (3.19) is not
particularly useful in engineering practice. The values of the optimum thickness of the porous layer
are described by the equations:

• for water:

2 –3 3 –4 4
δopt = 1.314 − 0.215 q + 0.280 q – 1.920 ∙ 10 q + 6.308 ∙ 10 q –
–7 5 –9 6 –11 7
– 5.624 ∙ 10 q – 3.464 ∙ 10 q + 6.234 ∙ 10 q , mm (3.20)

• for ethanol:

–3 2 –4 3 –5 4
δopt = 0.733 – 0.050 q + 6.209 ∙ 10 q – 4.207 ∙ 10 q + 1.374 ∙ 10 q –
–7 5 –9 6 –11 7
– 1.192 ∙ 10 q – 3.464 ∙ 10 q + 6.134 ∙ 10 q , mm (3.21)

The expressions (3.20) and (3.21) are valid exclusively for the range of the measured heat fluxes:
4 –2 5 –2
3 ∙ 10 W m ≤ q ≤ 3 ∙ 10 W m [327].

60 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Extensive experimental and theoretical investigations on boiling heat transfer on surfaces covered
with capillary porous structures made of powders with sintering techniques were conducted in
the former Soviet Union. The work by Sirotin et al. [246] was one of the first publications on the
subject. It discussed the experimental results of n-butane boiling on the surface of a steel pipe
(steel 1H18N10T) covered with a powder of the same material. They obtained a boiling curve
α = α(q), presented in Fig. 3.17, for porous layer thicknesses ranging from δ = 0.22 to 0.25 mm,
porosities from 0.5 to 0.6, and a mean powder particle diameter dgrn = 0.63 mm. A considerable
increase in the heat transfer coefficient in comparison with a smooth surface was seen, especially
3 –2 4 –2
for small heat fluxes (approx. 7.5 times for q = 5 ∙ 10 W m and 3 times for q = 5 ∙ 10 W m ).

4
3⋅10

2
104
-1
α, W⋅m ⋅K
-2

1
3
10

2⋅10 2
103 10 4 105
q, W⋅m-2

Fig. 3.17. Heat transfer coefficient plotted versus heat flux for a smooth pipe and the
same pipe covered with a porous layer (dpipe = 0.12 m): 1 – smooth pipe;
2 – porous layer of sintered powder, according to Sirotin et al. [246]

Interesting comparative experimental investigations were conducted by Rojzen and co-


researchers [223] for nitrogen and R-113 boiling on surfaces covered with capillary-porous
structures made by two different techniques: electric arc spraying and powder sintering. Copper
and aluminum powders were deposited on a 22 mm diameter pipe where the powder particle
diameters ranged 50 to 250 µm, the porous layer thickness was 0.72 to 2.20 mm, and the porosity
was 0.66 to 0.74. Apart from these authors’ own investigations, the results of other researchers
were also included in their analysis. In the conclusions it was stated that porous coatings
made with the electric arc do not always produce favorable results with respect to heat transfer
enhancement. All types of coatings of sintered powders, however, yielded a visible increase in the
heat transfer coefficient when compared with a technically smooth surface. The best results were
given by a coating of approximately 2 mm thickness and particle diameters of 40 to 50 µm [223].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 61


Chapter 3

4
5⋅10
1 6
2 7
3 8
4 9
q, W⋅m -2 5

10 4

2⋅10 3
2 4 6 8 10
-1 2 4 6 8 10
-2 2

DTsl , K

1 – R-12; 2 – R-22; 3 – ammonia; 4 – smooth pipe; 5 – low finned tube (pitch – 0.76 mm,
height – 1.5 mm); 6 – electro-chemical coating; 7 – layer of glass cloth (δ = 0.3 mm);
8 – electric arc sprayed coating; 9 – sintered powder coating, according to Djundin, Danilova
and Borishanskaja [76, 77]

Fig. 3.18. Experimental results of refrigerants boiling on tubular surfaces with porous
coatings, Tsat = 283 K

Borishanskaja, Danilova, Djundin and coworkers [35, 37, 71, 76, 77] conducted extensive
experimental investigations on R-12, R-22 and ammonia boiling on surfaces with porous
coatings obtained with various fabrication techniques, such as powder sintering, electric arc
spraying, electrochemical plating and covering the surface with glass cloth. Fig. 3.18 shows their
collected experimental results, presented in a detailed manner in the works [76, 77]. The general
conclusion drawn by these authors is that the highest heat transfer coefficients are found for the
porous structures made with the sintering and the electric arc spraying techniques. The choice
of the optimum porous structure should account for the following: boiling liquid properties, heat
exchanger operational conditions where boiling takes place and the structure’s manufacturing
technique.

Borishanskaja and co-researchers [35, 71] examined R-12 and R-22 boiling on tubes with coatings
made of sintered powders. Table 3.2 presents the data concerning the investigated porous
coverings. The corresponding boiling measurement results are presented in Fig. 3.19. They were
obtained for steel tubes (1H18N10T) covered with a powder of the same material (Table 3.2). The
saturation temperature was varied from –253 K to +293 K, reduced pressure from p/pcr = 0.037
3 4 –2
to 0.184 and heat fluxes from q = 2 · 10 to 3 · 10 W m . The largest heat transfer enhancement

62 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

ratio occurred for Sample 1 in Table 3.2; it was 3 to 7 times higher when compared with a smooth
tube and 2 to 4 times higher when compared with a low finned tube.

Table 3.2. Properties of porous coverings, according to Borishanskaja [35]

Property Sample
Covering thickness, δ, mm 0.3 0.5 1.0 1.0
Porosity, ε 0.46 0.51 0.50 0.52

Mean powder particle diameter, dgrn, mm 0.081 0.205 0.150 0.300

The following correlation was derived for two characteristic intervals of reduced pressure [35]
based on their results:

(3.22)

I. p/pcr = 0.184 to 0.103 m = –0.10 C = 49


II. p/pcr = 0.071 to 0.037 m = 0.35 C = 13.3

where

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 63


Chapter 3

4
Sample
Liquid p/pcr
1 2 3 4
0.138 0.037
R 12
2 0.103 0.053
0.049
R 22 0.184
0.137 0.071
Curve 1 2
103
8
6

4
B

2
1

2
10
8

6
2
4
1 − B = 49 Re −0.86
2 − B = 13.3 Re − 0.86

2
2 4 6 8 2 4
10 -2 10-1
Re

Fig. 3.19. Experimental data compared to the correlation (3.22) by Borishanskaja [35]

In the formula (3.22), the diameter De is the equivalent diameter for the mean dimension of the
capillary channels in the capillary-porous structure and the capillary constant k describes the
impact of the surface tension on the boiling process. The three quantities inside parenthesis in
formula (3.22) are in fact well-known similarity numbers:

(3.23)

As we can see from equation (3.22), α increases when the pressure rises at higher pressures
while α decreases when the pressure rises at lower pressures. This means that, with respect to
pressure, the character of the heat transfer process changes and is explained as follows. The
same dependence, similar to (3.22), occurs when boiling in narrow channels, where two kinds
of boiling are observed: the first one in which small isolated bubbles are generated at a high
frequency and the second where coalescing bubbles are formed that depart with low frequency
[35].

64 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The results of Borishanskaja’s investigations [35] can also be presented in a slightly different
manner, where the heat transfer coefficient is explicitly dependent on the heat flux [71] as follows:

(3.24)

where

The values of the constants appearing in (3.24) are given in Table 3.3. The authors [71] estimate
that the correlation (3.24) represents the experimental data with an accuracy of about ± 30%.

Table 3.3. Values of constants in (3.24), according to [71]

Values of constants
Liquid Tsat , °C
C1 C2 m
R-12 293 to 283 49.0 1.94 0.10
263 to 253 13.3 1.94 -0.35
R-22 293 to 283 49.0 2.31 0.10
263 to 253 13.3 2.31 -0.35

Some data concerning water boiling on a surface coated with sintered powder layers, as
the function of pressure, were provided by Kuzma-Kichta and co-researchers [138]. These
investigations were carried out with coatings made of stainless steel powders with particle
diameters of approximately 0.06 mm and a porosity of about 0.5, with the tests conducted at two
5 5 –2
fixed heat fluxes of 3 ⋅ 10 and 9 ⋅ 10 W m . The advantageous influence of the porous coating
on the heat transfer coefficient was found to diminish with rising pressure and heat flux. For
5 –2
q = 9 ⋅ 10 W m and pressures above 6 MPa, the surfaces with the powder coating had heat
transfer coefficients lower than those of the technically smooth surfaces. This was explained as
follows: when the pressure is increased, the critical radius of departing bubbles diminished, which
according to the authors, changes its ratio to the characteristic pore diameter and adversely
influences the boiling process intensity [138].

Andrianov, Malyshenko and Styrikovich [11, 12, 13, 154, 264] carried out extensive experimental
investigations on the boiling hysteresis phenomenon on surfaces with porous coatings made of
two materials – corundum (Al2O3) and nichrome (NiCr). The porous coating was deposited on a
flat nichrome sample with two techniques: plasma spraying (corundum and nichrome) and powder
sintering (nichrome). The coating thicknesses ranged from 0.2 to 1.8 mm and the powder particle
diameters from 130 to 400 µm but the porosity did not change much, namely it remained between
0.4 to 0.5 for the powder structures and 0.2 to 0.6 for the sprayed structures. The investigations

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 65


Chapter 3

were conducted for water boiling at pressures ranging form 0.05 to 0.31 MPa. The results obtained
by Malyshenko and Styrikovich [11, 12, 13, 151, 152, 153, 154, 262, 263, 264] are very close to the
st
1 kind of boiling hysteresis recorded by Poniewski and Wojcik [203, 205, 211, 214, 216, 217, 218]
for metal fibrous structures described earlier.

The analysis of the boiling curves in [11, 12, 13, 154, 264] led to the distinction of two kinds of
nucleate boiling on the surfaces covered with porous structures made with sintered powders or
st 5 –2
plasma spraying. Boiling of the 1 kind occurs when q ≤ 10 W m and where the vapor bubbles
nd
are generated inside the pores at the surface of the capillary-porous base. Boiling of the 2 kind
is characteristic of higher heat fluxes, for which a stable vapor film, separating the original base
surface from the vapor bubble formation zone, is formed inside the porous structure.

Both kinds of nucleate boiling are characterized by different indexes m in the heat flux vs. the
temperature difference relationship between the heating surface and the boiling superheat, i.e.,
m
q ~ DTsl , where ΔTsl = Ts – Tsat. Differences between the above-mentioned kinds of boiling are
discussed later in 4.3.3 [11, 12, 13, 151, 152, 153, 154, 262, 263, 264].

st
Andrianov and Malyshenko [11] suggest that for nucleate boiling of the 1 kind, the functional
dependency of the heat flux on the heating surface superheat should be written in the
dimensionless form:

(3.25)

where

(3.26)

66 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

1.0
0.9
0.8 1 10
0.7 2 11
3 12
0.6 4 13
0.5 5 14
6 15
0.4 7 16
8 17
q~

0.3 9

m=1.65
m=1.05
0.2

0
0 0.2 0.4 0.6 0.8 1.0
~
DT

a.b. = as before
1 – water, sintered powders NiCr, psat = 0.1 MPa, dgrn = 0.14 mm, δ = 0.8 mm; 2 – a.b.,
dzrn = 0.4 mm, δ = 1.8 mm; 3 – a.b., psat = 0.05 MPa, dgrn = 0.14 mm, δ = 0.8 mm;
4 – a.b., dzrn = 0.27 mm, δ = 1.8 mm; 5 – water, plasma spraying, psat = 0.1 MPa, AL 2O3,
ε = 0.2 to 0.3, δ = 0.8 mm; 6 – a.b., ε = 0.5 to 0.6, δ = 1.4 mm; 7 – a.b., NiCr, ε = 0.32 to 0.36,
δ = 0.2 mm; 8 – a.b., δ = 0.5 mm; 9 – a.b., δ = 1.0 mm; 10 – a.b., ε = 0.5 to 0.6,
δ = 1.4 mm; 11 – ethanol water solution, psat = 0.1 MPa, plasma spraying, NiCr, ε = 0.32
to 0.36, δ = 0.8 mm; 12 – a.b., sintered powders, dgrn = 0.14 mm, δ = 0.8 mm; 13 – R-113,
according to Tehver and Sui [279]; 14 – a.b., according to [245]; 15 – water, according
to [27]; 16 – a.b., according to Abhat and Seban [1]; 17 – FC-72, according to Marto and
Lepere [156]
st
Fig. 3.20. Experimental data by Andrianov et al. [11] for nucleate boiling of the 1 kind

The heat flux qcon(ΔTsl) was obtained by extrapolating the convection part of the boiling curve
into the developed nucleate boiling zone, i.e. into the zone q > qincip, where qincip corresponds to
*
the nucleate boiling incipience. The heat flux q and temperature difference DTsl * characterize the
change in the boiling kind.

All analyzed experimental data describe curves of the type (3.25), where the exponent ranges
from 1.05 ≤ m ≤ 1.65 and for the constant C = 1, as shown in Fig. 3.20. At small heat fluxes,
significant differences are observed for various capillary-porous structures because of their
differing thermal resistances, which is reflected in the resulting differences in temperature drop
from the heating surface base to vapor bubble generation zone inside the structure. This causes
significant differences in the value of m. For ~
q ≥ 0.3, equation (3.25) with the mean value of the

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 67


Chapter 3

m = 1.35 and the constant C = 1 approximates the majority of the experimental data with an error
which diminishes as the heat flux grows.

1.6
Water
1.4

1.2
-2
q⋅10 , W⋅m

1.0
-4

0.8

0.6

0.4 d = 0.85 mm
d = 1.00 mm
0.2 d = 2.00 mm
d = 4.00 mm
0
0 10 20 30 40 50 60 70 80
DTsl , K

Fig. 3.21. Heat flux as the function of the heating surface superheat, according to
Abramenko and Kanonchik [2]

The results obtained by Abramenko and Kanonchik [2] were slightly different from those by
Nishikawa and co-researchers [173, 177] or Zhang and Chen [327]. Abramenko and Kanonchik
investigated the impact of the capillary-porous structure thickness on the shape of the boiling
curve q = (ΔTsl). Their experiments focused on water boiling on a flat copper plate (0.1 × 0.05 m)
coated with a structure formed by sintering bronze particles. The diameter of powder particles
ranged from 0.2 to 0.315 mm, the porosity was approximately 0.3 and the four coating thicknesses
tested were 0.85 mm, 1 mm, 2 mm and 4 mm. The results are presented in Fig. 3.21. The highest
critical heat flux was recorded for the porous layer of the smallest thickness. This means that the
vapor escape mechanism from the porous layer is decisive for the course of the boiling process
and it deteriorates as the covering thickness increases [2].

Lisovskij [147] investigated nitrogen boiling on a capillary-porous structure made of bronze


powders, whose thickness was 2 mm whereas two particle diameters were tested: 0.2 and 0.6
mm. The pressure ranged from 14.6 kPa to 58.6 kPa in the tests while the heat fluxes were
4 4 –2
from 0.1 ∙ 10 to 0.4 ∙ 10 W m . The porosity of the layers was not cited. In spite of the fact that
Lisovskij did not collect a large amount of experimental data, he put forward a correlation to
determine Nusselt number. The correlation has the following form:

0.3 0.36
Nu = 1.5 (Re Np) Pr (3.27)

68 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The similarity numbers in (3.27) are defined as follows:

(3.28)

where Dh is the pore hydraulic diameter and φ denotes the porous structure dryness degree.
Lisovskij however did not describe the method to determine this parameter. The dimensionless
number Np represents the impact of pressure on the boiling process.

Zhang and colleagues [331, 332] conducted experimental investigations for water, ethanol and
R-113. They proposed a correlation representing their experimental data with an error of ± 23.5%.
The porous coating was made of bronze powder and deposited on a copper cylinder of 40 mm
diameter. The thickness δ of the porous layer ranged from 0.94 to 2.63 mm, the porosity ε from
0.3444 to 0.5517 and the mean diameter of powder dgrn from 0.106 to 0.529 mm. The heat flux
3 6 –2
ranged from 0.5 ∙ 10 to 1 ∙ 10 W m . In [331], however, some mistakes were made in defining
their similarity numbers. These mistakes were carried through in their subsequent formula
transformations.

The correlation has the form:

(3.29)

where individual similarity numbers should be defined as follows:

(thermal conductivity was not defined)

(in [331] written erroneously as

(in [331] written erroneously as (3.30)

(in [331] written erroneously as

(in [331] written erroneously as

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 69


Chapter 3

Zhang and Zhang [331] explained that Fr/We represents the ratio of inertia forces to surface
tension forces and thus specifies the impact of surface tension on vapor flow. The number We1
is the ratio of the surface tension and hydrostatic lift forces and describes the influence of both
of these quantities on liquid film flow in nucleation centers. We2 describes liquid penetration into
pores.

On the assumption that the similarity numbers mentioned in (3.30) were derived correctly, the
range of validity of their correlations is as follows:

(3.31)

Summarizing the conclusions of the above investigations, they confirm that the incipience of
nucleate boiling on sintered layers of metallic powder takes place at superheats that are much
lower than on technically smooth surfaces. Furthermore, sintered micro-structured surfaces
significantly enhance boiling heat transfer. The majority of experiments confirm that there exists
an optimum porous layer thickness, for which the highest value of the heat transfer coefficient is
found at a fixed value of the heat flux. Generally, better results are obtained for thin layers, which
points to the dominant influence of vapor escape routes on the boiling process. Hydraulic resistance
to vapor outflow grows as the layer gets thicker.

A characteristic feature of capillary-porous structures of sintered powders is the occurrence of


st
hysteresis phenomena, i.e. nucleation hysteresis at increasing heat flux density as well as 1 and
nd
2 kind hysteresis, also observed on fibrous metal coatings. Hysteresis phenomena hamper the
application of sintered structures due to the ambiguous shape of their boiling curves q(ΔTsl).

The advantage of porous coatings of sintered powders is the possibility of producing a


predetermined porosity of open pores through the selection of particle size and the compression
of the powder. Powder layers, however, are brittle and difficult to deposit on the surfaces of
complex shapes. The technique of depositing porous layers made of sintered powders has not
been applied to industries on a large scale, except for the High Flux tube.

3.2.3. Thermally sprayed structures

A number of techniques can be used to make a porous coating by means of thermal spraying.
These include plasma spraying, electric arc spraying and flame spraying. These processes are
described briefly below.

Creation of a plasma can be achieved with an electric arc, and the power necessary is usually
approximately 100 kW. The material to be sprayed is fed to the burner between the arc electrodes
as a powder or a wire and blown towards the surface to be coated by the working gas. The latter

70 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

is, most frequently, pure argon or an argon mixture containing helium, nitrogen and hydrogen.
Plasma spraying can be performed at atmospheric pressure, at reduced pressure (vacuum
spraying) or in the atmosphere of an inert gas. The plasma most often reaches a temperature of
–1
about 14,000 K and its velocity at the burner nozzle is about 800 ms or more.

Instead, in electric arc spraying, the material being sprayed comes from two arcs forming the
electrodes that get used up. Therefore, only materials that conduct electric current can be sprayed
with this technique. Droplets of the material melted in the electric arc are blown towards the target
surface by compressed gas. This technique of spraying can be used in air, in the atmosphere of
inert or a reacting gas and also in a vacuum. The electric arc power usually amounts to 5 to 10 kW
and the voltage ranges from 20 to 40 V.

In flame spraying technology, the chemical energy released by gases burnt in oxygen is used to
heat the sprayed material. Inflammable gases and oxygen are usually fed to the burner at the ratio
of 1:1 and the material to be sprayed is usually a powder or a wire. The flame temperature usually
–1
reaches 3200 K and its velocity is from 80 to 100 ms . The flow rates of the inflammable gas and
oxygen depend on the burner design. Detailed technical data on burner designs, the properties of
powders and wires used in this technique as well as thermal and electrical parameters of spraying
are available in Kostrzewa and Nowak [114] and also in Pawlowski [197].

All techniques of thermal spraying produce coatings whose porosity usually does not exceed
ε = 0.2. In order to achieve higher porosities, some components can be added to the sprayed
materials. For example, Cieslinski [58] worked out a modified technique of flame spraying, which
yielded porosities up to ε = 0.6 on flat surfaces.

Plasma spraying

Porous coverings deposited on the base heating surface are most frequently made with the
same material as the base or with a material of higher thermal conductivity, e.g. a copper porous
coating on a steel pipe. The thermal plasma spraying technique also uses materials of low thermal
conductivity, such as corundum (Al2O3) or polymerized tetrafluoroethylene (TFE). Those materials
are, for example, resistant to oxidation in elevated temperatures typical of plasma spraying.
Porous coatings made of such materials are characterized by a considerable resistance to
chemical action.

One of the first works dealing with the application of porous coatings made with plasma spraying
was for enhancing the boiling of liquid nitrogen by Warner, Park and Mayhank [308]. A layer of
TFE polymer was deposited on a gilded copper disc. The porous layer thickness ranged from
2 to 8 µm. The best results were recorded for the thickest TFE polymer layer, which are shown
in Fig. 3.22. Apart from an increase in the heat flux for small superheats of the heating surface,
it was also found that nucleate boiling passes into film boiling less dramatically on such a porous
coating than on a technically smooth surface.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 71


Chapter 3

}
30
- porous coverings
25
-uncoated surfaces
reference curves
20

q⋅10-4, W⋅m-2 15

10

0
0 1 2 3 4 5 6 7 8 9
DTsl , K

Fig. 3.22. Pool boiling results for nitrogen boiling for a surface with TFE porous
coating, 8 µm in thickness (four measurement series) and reference curves
for all uncoated surfaces, according to Warner, Park and Mayhank [308]

In the former Soviet Union, the investigations into the properties of boiling on capillary-porous
structures made with plasma spraying were conducted mainly at two centers: the Institute of High
Temperatures of the USSR Academy of Sciences in Moscow [11, 12, 13, 151, 152, 152, 154, 262,
263, 264] and the Institute of Thermophysics and Electrophysics of the Estonian Academy of
Sciences in Tallinn [148, 200, 274, 276 – 281, 284 – 287].

Andrianov, Malyshenko and Styrikovich [11, 12, 13, 151, 152, 153, 154, 262, 263, 264] and
co-researchers focused on examining and analyzing the peculiarities of boiling on capillary-
porous structures made with two techniques, namely nichrome (NiCr) powder sintering and
plasma spraying of nichrome and corundum (Al2O3). With respect to heat transfer enhancement,
experimental investigations into pool boiling of water at pressures from 0.05 to 3.10 MPa showed
that plasma sprayed capillary-porous structures gave higher heat transfer coefficients than those
for technically smooth surfaces, but lower than the values obtained with sintered powder layers
and High Flux surfaces [11], as can be seen in Fig. 3.23.

72 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

α⋅10-4, W/m -2⋅K-1


2

1 1 6
2 7
0.7 8
3
0.5 4 9
5 10
0.3
104 2 4 6 8 10 5 2 4
q, W/m-2

a.b. = as before
1 – sintered powders, dgrn = 0.14 mm, δ = 0.8 mm; 2 – a.b., two-layer, anisotropic covering,
δ = 1.5 mm; 3 – a.b., dgrn ≈ 0.063 mm, δ = 1.5 mm; 4 – plasma spraying, NiCr, ε = 0.5 to 0.6,
δ = 0.4 mm; 5 – a.b., ε = 0.32 to 0.36, δ = 0.5 mm; 6 – a.b., Al2O3, ε = 0.5 to 0.6, δ = 1.4 mm;
7 – a.b., anistropic structure with channels, δ = 1.4 mm; 8 – a.b., ε = 0.2 to 0.3, δ = 0.4 mm;
9 – High Flux, according to [90]; 10 – smooth surface

Fig. 3.23. Heat transfer coefficient as a function of heat flux for porous coatings
obtained with various techniques for water at psat = 0.1 MPa from Andrianov
and Malyshenko [11]

Both for sintered structures and for those made with plasma spraying, an optimum of the coating
thickness is observed, which is similar in thickness to porous coatings manufactured with other
techniques as illustrated in Fig. 3.24 [11]. The generalization of experimental results collected
for plasma spraying and sintered powder structures was presented before in Fig. 3.20, which
illustrates the previously discussed correlation (3.22).

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 73


Chapter 3

3.0

1
2.5
2
3

α⋅ 10-4, W/ m- 2 ⋅K- 1
2.0 4
5
6
1.5 7

1.0

0.5

0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
δ , mm

a.b. = as before
1, 2, 3 – sintered coverings, dgrn = 0.14 - 0.4 mm; 4 – plasma spraying NiCr, ε = 0.32 to 0.36;
5 – a.b., ε = 0.5 to 0.6; 6 – plasma spraying Al2O3, ε = 0.2 to 0.3; 7 – a.b., ε = 0.5 to 0.5,
according to Andrianov and Malyshenko [11]

Fig. 3.24. Heat transfer coefficients plotted as a function of the porous layer
4 –2
thickness at psat = 0.1 MPa where q = 5 ∙ 10 W m is for empty symbols
5 –2
and q = 10 W m is for solid symbols

The research interests of Tehver and co-workers [148, 200, 274, 276 – 281, 284 – 287] were
similar to those of Andrianov, Malyshenko, Styrikovich et al. [11, 12, 13, 151, 152, 153, 154, 262,
263, 264]. They also investigated the peculiarities of boiling on porous coatings, paying attention
to the problems of hysteresis, boiling crisis and the modeling of the boiling process on porous
coatings made by plasma spraying of metal and ceramic powders. Tehver’s experimental heat
transfer investigations on boiling of R-113 on a copper tube of 22 mm coated with a copper porous
layer confirmed that the optimum layer thickness decreases with increasing heat flux, which is also
congruent with the theoretical considerations in [285, 286].

The work of Tehvera, Sui and Temkina [281] presents a thorough survey of experimental results
for R-113 boiling. The authors observed the boiling process on flat heated surfaces coated with
porous layers of aluminum, bronze, copper and corundum, made with plasma spraying. They
analyzed the impact of the layer thickness, porosity and mean pore diameter on the shape of
the boiling curve, in other words, on such phenomena as the superheat necessary for boiling
incipience, hysteresis and boiling crisis (critical) heat flux. Fig. 3.25 shows the heating surface
superheat plotted as a function of the porosity, the layer thickness and the pore mean diameter
for copper capillary-porous structures on an aluminum base plate at a fixed heat flux [281]. The
experimental data were approximated with parabolic curve fits, with the exponents determined
with the least squares method. For all three cases, the optimum values of ε, δ and Dp were
determined, for which the superheat is at its minimum value, at the heat flux tested.

74 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

15 15
D p = 4 ÷ 6.4 µm a) D p = 2 ÷ 6 µm b)
δ = 0.05÷0.17 mm ε = 0.02÷0.35
1 − DT = 57.7ε2 − 35.5ε + 7.8 1 − DT = 57.2δ 2 − 25.5δ + 5.2
10 10
2 − DT = 80.1ε2 − 45.6ε + 9.4 2 − DT = 53.3δ 2 − 20δ + 5.4

DTsl , K
DTsl, K
2
2
1
5 1 5

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 0.1 0.2 0.3 0.4 0.5 0.60.7
ε δ
15
ε = 0.3÷0.38 c)
δ = 0.1÷0.2 mm
1 − DT = 0.062 Dp − 0.85D p + 5.2
2
10
2 − ΔT = 0.123Dp − 1.88D p + 9.3
2
DTsl, K

5 2

0
0 5 10 15
Dp , µm

a) porosity; b) layer thickness; c) mean pore diameter

Fig. 3.25. Impact of the coating on the heating surface superheat for a copper
5 –2
capillary-porous structure on an aluminum base where q1 = 10 W m and
5 –2
q2 = 2 ∙ 10 W m according to Tehver, Sui and Temkina [281]

The experiments in [281] were conducted for only one liquid, i.e. R-113 and various compositions
of the heating surface (base) and porous coating (aluminum – bronze, copper – bronze, copper
– copper, aluminum – copper, aluminum – corundum, aluminum – aluminum) with 113 samples
tested altogether. In these tests, the porosity ranged from 0.05 ≤ ε ≤ 0.5, the covering thickness
from 0.03 mm ≤ δ ≤ 0.60 mm, and the pore diameter from 1 µm ≤ Dp ≤ 20 µm.

On the basis of the above, Tehver, Sui and Temkina [281] drew the following conclusions:

1. The porous coating virtually does not affect the heating surface superheat necessary
to initiate boiling. At the same time, a smaller superheat is required to maintain stable
nucleate boiling, which results in the increase in the initial superheat and hence
causes boiling hysteresis.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 75


Chapter 3

2. The structural parameters (porosity ε, thickness δ and the pore mean diameter Dp)
should be chosen (optimized) depending on the heat flux, as can be clearly seen in
Fig. 3.25.

3. Boiling heat transfer hysteresis phenomena strongly manifest themselves on porous


coatings. They are caused by the influence of the coating structural parameters and
initial conditions on boiling incipience.

4. The heterogeneity of the boiling process on a surface with a porous covering can
cause a drop in the mean value of the heat transfer coefficient and lower the critical
heat flux.

5. The initial superheat of the heating surface, which brings about boiling hysteresis, can
be decreased through artificial activation of nucleation centers and a proper selection
of the coating thickness and the mean pore diameter.

Dorochov and Bochagov [81] examined the impact of various plasma-sprayed porous coatings
on boiling of R-21. The coatings, made of two alloys of nickel and titanium, pure nickel and pure
corundum, were deposited on a steel tube 22 mm in diameter. The porosity was determined only
for the corundum coating and it was ε = 0.12. It might be assumed that it was of the same order
for the remaining samples. It was stated that when the heat flux increased, the heat transfer
coefficients were always lower than for the smooth surface. This demonstrated that a low porosity
coating, ε ~ 0.1, does not enhance heat transfer at elevated heat fluxes due to the difficulty of the
vapor to flow out of the layer.

Scurlock [231] presented results of investigations on boiling of liquid nitrogen, argon, oxygen and
R-12 on flat aluminum heated surfaces with plasma sprayed porous layers made of aluminum
powders or an aluminum-silicone mixture (10%). Polyester, which was vaporized in the course
of the spraying, was added to the powders, thus increasing the porosity of coatings. The coating
thicknesses ranged from 0.13 to 1.32 mm. The values of the porosity were not cited. Heat transfer
coefficients increased by as much as ten times when compared with a smooth surface, which
for boiling of nitrogen are illustrated in Fig. 3.26. Also note that boiling on the porous coatings
occurred at superheats as little as 0.1 to 0.2 K. Similar to what the investigations of Tehver et al.
[148, 200, 274, 276 – 281, 284 – 287] demonstrated, it was stated that at a fixed heat flux, there
exists an optimum thickness of the porous coating at which the highest value of the heat transfer
coefficient is found.

76 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

6
10

105 δ, mm

q, W⋅m -2
smooth
0.13
0.25
0.38
104 0.50
0.81
1.04
1.32
103
0.1 1 10 100
DT sl , K

Fig. 3.26. Boiling curves for liquid nitrogen as a function of the porous coating
thickness from Scurlock [231]

Hsieh and Weng [98] investigated R-134a and R-407C boiling on an electrically heated copper
tube whose diameter was 9 mm. The tube was covered with porous layers made with plasma
and flame spraying techniques to deposit copper, molybdenum, aluminum and zinc powders.
Layer thicknesses ranged from 30 to 300 µm and the porosities ranged from 0.077 to 0.113. The
authors also investigated boiling on rough surfaces produced by rubbing the surface with abrasive
papers of different roughness. The occurrence of hysteresis phenomenon in nucleate boiling was
confirmed. There was a 2.5-fold increase in the heat transfer coefficient when compared with a
technically smooth surface, for which the results are presented in Fig. 3.27.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 77


Chapter 3

105
1 2 Tube
Smooth
Cu plasma
Mo plasma
4 Al(L) flame
10
Al(H) flame
Zn flame
-2
q, W/m

103

2
10

R-134a
1– heat flux increase
2 – heat flux decrease
101
0.5 1.0 10.0 20.0
DTsl , K

Fig. 3.27. R-134a boiling curves for tubes with porous coatings made with plasma and
flame spraying techniques from Hsieh and Weng [98]

Emulating the work of Rudemiller and Lindsay [225], Hsieh and Weng [98] put forward an
empirical correlation based on the following similarity numbers: Jakob – Ja, Reynolds – Re,
constant heat flux similarity number – Ncf and a geometrical scaling coefficient – λ. These
quantities are defined as follows:

(3.32)

The correlation takes on the following form:

0.292 –0.19 0.065


Ja = 0.137 Re λ Ncf (3.33)

78 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

It is valid over the following ranges:

–5 –1
6.91 ∙ 10 ≤ Re ≤ 1.90 ∙ 10
–3 –1
6.67 ∙ 10 ≤ λ ≤ 1.67 ∙ 10 (3.34)
–4 –3
7.77 ∙ 10 ≤ Ncf ≤ 4.79 ∙ 10

It is applied to porous coatings obtained with plasma and flame spraying, where R-134a and
R-407C are the boiling liquids. Fig. 3.28 shows a comparison of their correlation (3.33) with
respect to the experimental data.

1.000

± 25%

0.100
Ja / (λ-0.190 Ncf 0.065)

0.010

1 0.292

Experimental data
Hsieh and Weng´s correlation , 1997
0.001
0.00001 0.0001 0.001 0.01 0.1 1
Re

Fig. 3.28. Hsieh and Weng correlation for plasma and flame sprayed coatings
compared to experimental data [98]

Experimental investigations into water boiling on flat and cylindrical surfaces with porous coatings
produced with plasma and flame spraying were also conducted at the Gdansk University of
Technology by Cieslinski [51 – 60]. The porous coatings were made of aluminum powder and acid
resistant steel 316L. The coating thicknesses ranged from 0.15 mm to over 0.90 mm. Experimental
results for plasma sprayed coatings differed considerably from those obtained for flame spraying.
Observation of the coatings revealed that the plasma sprayed layers were not porous and thus in
further investigations flame spraying was the only technique used. This emphasizes an important
issue concerning the difficulty in achieving considerable porosity with plasma spraying and is a
major drawback of this technique.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 79


Chapter 3

Electric arc spraying

Spraying with an electric arc is a technique supplied to metal heating surfaces coated with porous
layers that are also made of metal [197]. Numerous experimental and theoretical works devoted
to investigations of boiling on porous coatings made with the electric arc spraying technique were
conducted in the former Soviet Union [37, 71, 76 – 79, 89, 134, 183, 184, 220, 221, 223].

In the first important experimental investigation, conducted by Djundin et al. [78], three refrigerants
(R-11, R-12 and R-22) were boiled on the surface of a steel pipe with a porous copper coating.
They demonstrated that the coating significantly diminished the heating surface superheat ΔTsc
required to achieve boiling incipience and resulted in a considerable increase in the heat transfer
coefficient when compared with a technically smooth pipe. The investigations involved four
coatings whose thicknesses were 0.075, 0.125, 0.275 and 0.3 mm and the tests were conducted at
two pressures, 2.7 and 3.75 bars. They found that the heat transfer coefficient increased with the
thickness of the coating as shown in Fig. 3.29.

The porosity of coatings made with electric arc spraying is relatively low and it ranged from
0.144 ≤ ε ≤ 0.368 in the investigations by Djundin, Danilova and Borishanskaja [76, 77]. Fig. 3.18
presents boiling curves for their porous coatings made with various techniques. It should be
noted that the lowest surface superheats and hence highest heat transfer coefficients occurred
for sintered coatings and, subsequently, for those sprayed by electric arc. When commenting on
their experimental investigations, Rojzen et al. [223] arrived at a conclusion that sintered structure
porosity is of “volumetric” character, whereas that of arc sprayed is “superficial”. This means that
the porosity of a sintered layer is nearly uniform through its thickness while that of an arc sprayed
layer is only porous at its surface. Sintered structures, therefore, yield far better results for boiling
enhancement. Those conclusions were confirmed by Bukin et al. [37] for experiments with R-12
and capillary-porous structures sprayed by the electric arc technique.

80 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

2·10 4

10
4 2
8 3 4 5 6

6
-2
q, W⋅m
4

10 3
0.3 0.4 0.6 0.8 1 2 3 4 6 7
DTsl, K

Copper coating thicknesses tested: 1 – 0.3 mm; 2 – 0.275 mm; 3 – 0.125 mm; 4 – 0.075 mm;
5 – technically smooth pipe; 6 – low finned pipe with fin pitch s = 0.76 mm

Fig. 3.29. Boiling curves for R-12 boiling on a steel pipe of drura =19 mm at p = 3.75 bars
from Djundin et al. [78]

Further experimental investigations carried out by the researchers noted above [89] dealt with
the way a porous coating is made by the electric arc technique and the impact its characteristics
have on heat transfer enhancement with boiling of R-12 and R-22. A sprayed structure was found
to possess clearly marked layers, large visible pores and internal channels. Its effectiveness
was much higher than that of a packed structure with no differentiable layers or channels. These
investigations also confirmed that good vapor escape routes in the capillary-porous structure play
a decisive role in the advantageous effect the structure produces on the boiling enhancement
process.

On the basis of the experimental investigations with boiling of R-12 and R-22, Danilova et al. [71]
proposed a relatively simple correlation to calculate the heat transfer coefficient on electric arc
sprayed coatings:

(3.35)

where pressure is input in bars and the thickness in millimeters. The correlation covers the
following range of parameters: coating thicknesses from 0.14 mm ≤ δ ≤ 0.58 mm, porosities from
3 4 –2
0.144 ≤ ε ≤ 0.37, heat fluxes from 2 ∙ 10 ≤ q ≤ 2 ∙ 10 W m and saturation temperatures from

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 81


Chapter 3

253 K ≤ Tsat ≤ 293 K. In accordance with the authors’ statement, the experimental scatter with
respect to their correlation (3.35) was ± 30%.

4
1
6 Number of tubes in a cluster
25

3 Ammonia
Pool boiling
ps / sm

1
1 2 3 4 5 6 7 8 9 10
q, kW·m -2

Fig. 3.30. Ratio of mean boiling heat transfer coefficients for tube bundles with 1,
6 and 25 tubes with porous coatings versus technically smooth tubes for
ammonia from Solovev, Djundin and Danilova [257]

The research by Solovev, Djundin and Danilova [257] studied the impact of the bundle effects
with tubes with porous coatings. Experiments covered ammonia boiling on steel tubes covered
with a capillary porous structure of aluminum alloy AMC (Russian notation). The basic data on
the coatings, made with the electric arc spraying technique, are as follows: porosity ε = 0.279,
thickness δ = 0.8 mm, and equivalent pore diameter De = 24.4 µm. The conclusion the authors
drew is that the advantageous influence of the porous coating decreases when the number of
tubes increases, which is shown in Fig. 3.30. We can see that the augmentation ratio αps ⁄ αsm
is 3.5 for a single tube with a porous coating, whereas for a cluster of 25 tubes, it is less than two.
This can be explained by the fact that convective boiling is important in the heat transfer process
on plain tube bundles as explained in detail in Thome [288] and, hence, this decreases the
advantage of the porous coated tubes within a tube bundle.

Extensive experimental investigations by Poznjak, Orlov and Savelev [183, 184, 220, 221]
provided the basis for constructing a model of boiling in a capillary-porous structure and deriving
a complex correlation to calculate the Nusselt number for surfaces coated with these structures.
The investigations covered nitrogen, oxygen and hydrogen boiling on flat and cylindrical surfaces
made of aluminum (and its alloys), copper and acid resistant steel 1H18N10T. Sample porosities ε

82 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

ranged from 0.19 to 0.413; however, only open pores were taken into account in the measurements
while totally closed pores were not. The coating thickness δ changed from 0.05 to 0.7 mm.
The equivalent pore diameters were 40 to 60 µm and the pressures tested ranged from 0.012
to 0.095 MPa. In order to develop the correlation for cylindrical surfaces covered with capillary-
porous structures resulting from electric arc spraying, the following similarity numbers and
dimensionless numbers were used:

Nusselt number
(3.36)

Similarity number for boiling inside a capillary channel:


(3.37)

Biot number
(3.38)

Bond number
(3.39)

Galileo number
(3.40)

Grashof number
(3.41)

Prandtl number
(3.42)

Reynolds number
(3.43)

Other dimensionless numbers:


(3.44)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 83


Chapter 3

where B is the number of capillary channels feeding a single channel filled with vapor, Dg
is the capillary vapor channel diameter, De,fin is the fin equivalent diameter for the capillary-
porous structure’s skeleton and αg,chnl is the heat transfer coefficient inside a vapor channel.
Characteristics of the investigated samples are cited in Table 3.4.

Table 3.4 Characteristics of porous coatings, according to Orlov and Savelev [184]

Porous coating material


Parameter
Aluminum alloy AMC (Russian notation) Copper 12H18N10T

Sample no 1 no 2 no 3 no 4 no 5 no 6 no 7 no 8 no 9 no 10 no 11

Porosity (open
0.21 0.19 0.25 0.29 0.32 0.288 0.235 0.28 0.285 0.413 0.248
pores)
Layer thickness,
595 490 450 490 400 515 360 468 430 120 140
µm
Equivalent pore
41 42 43 44.5 47 49.5 43.5 45.5 53 36 42
diameter, µm

The Nusselt number (3.36) is given by the following expression from [184] where “ch” means
hyperbolic cosine:

(3.45)

In Fig. 3.31, one can see the correlated data for nitrogen at the pressure p = 0.098 MPa and the
samples described in Table 3.4; additionally, data for hydrogen and oxygen and also for nitrogen
at two other pressures 0.0196 MPa and 0.049 MPa are shown. Satisfactory agreement with
experimental data was achieved with the following empirical constants: B = 25 and m = 4. A more
detailed description of the Orlov and Savelev model [183, 184] is given later in 4.2.2.

84 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

20.0
-1 -8
-2 -9
-3 - 10
10.0 -4 - 11
8.0 -5 - 12
+30% -6 - 13
6.0 -7 - 14
- 15
-20%
4.0
Nuexp

2.0

1.0
0.8
0.6
0.06 0.08 0.1 0.2 0.4 0.6 0.8 1.0 2.0
Z = Nu-1

Numbers 1 - 11 correspond to the samples in Table 3.4 for nitrogen at p = 0.098 MPa;
12 – hydrogen, 13 – oxygen, and 14 – nitrogen, all at p = 0.0196 MPa; 15 – nitrogen at
p = 0.049 MPa from Orlov and Savelev [184]

Fig. 3.31. Experimental data comparison with correlation (3.45)

Flame spraying

Up to now in experimental investigations, flame spraying has been used less frequently than other
techniques to make porous coatings to enhance boiling heat transfer. This is probably due to the
difficulty of producing porous coatings of sufficiently high porosity, that also should contain mainly
open pores.

The first experiments by Djundin et al. [80] on flame sprayed coatings gave negative boiling
results. Further investigations, made with an improved spraying technique, led to porous coatings
that significantly enhanced boiling heat transfer in Danilova et al. [72]. The selected flame spraying
parameters were as follows: the velocity of the spraying apparatus relative to the base surface, the
rate of the wire feed (rate of material sprayed), the pressure and temperature of the compressed
air and the number of layers sprayed on the surface to form the coating. The heat transfer surface
in this case was a steel pipe (d = 20 mm). Two samples, described in Table 3.5, yielded significant
heat transfer enhancement ratios, which are shown in Fig. 3.32. As with other porous coatings, the
beneficial influence of the porous coating diminished with increasing heat flux density.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 85


Chapter 3

Table 3.5 Characteristics of two porous coatings made with flame spraying technique
using aluminum wire of d = 2 mm from Danilova et al. [72]
Sample δ, mm ε De, µm Number of layers
1 0.45 0.425 29.1 1
2 0.80 0.279 24.4 4

Hsieh and Weng [98] provided a correlation for the Nusselt number that was given earlier as
equation (3.33) that referred to coatings produced by both plasma spraying and flame spraying.
Fig. 3.27 show earlier demonstrates that both techniques yield porous coatings that enhance
boiling heat transfer equally well.

Mertz et al. [160], in co-operation with the Minsk and Kiev research centers, also investigated
flame sprayed porous coatings. They conducted investigations for boiling of propane on cylindrical
surfaces (d = 20 mm) of stainless steel covered with a capillary porous structure of the same
material. The coating thicknesses δ ranged from 0.1 mm to 0.3 mm and the porosities ε were from
0.04 to 0.17. There were 10 samples investigated in all. A clear boiling hysteresis was observed
and so was significant increase in heat transfer coefficient when compared with a smooth pipe
–2
(2.5 to 3 times). Correlating this experimental data for heat fluxes from 3 to 70 kW m and
reduced pressures p ~ = p/p from 0.081 to 0.323, the following dimensional correlation was put
crit
forward:

(3.46)

–2 –1
It predicts their propane data to within ± 25% accuracy where α has the dimensions of kW m K
–2
and q is in kW m ).

9
8
1
7
α /α sm

6
2
5
R 22
4
3 3
10 2 4 6 8 104 2 4⋅104
q, W⋅m -2

Fig. 3.32. Boiling heat transfer enhancement ratios for two flame sprayed porous
surfaces, 1, 2 – sample numbers as in Table 3.5 (from Danilova et al. [72])

Cieslinski [51 – 60] conducted numerous investigations for water boiling on flat and cylindrical
surfaces covered with capillary-porous structures which were made with three techniques:

86 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

electrolytic deposition, plasma spraying and flame spraying. After performing initial tests, the first
two techniques were rejected and further investigations concerned only the porous coatings made
with flame spraying.

An important characteristic of thermally sprayed coatings is that they provide good adhesion to the
base surface (resisting forces from 20 to 80 MPa) and also other good mechanical properties. A
major drawback of all thermal spraying techniques is the difficulty in achieving desirable porosity
with open pores. In order to obtain porosities ε > 0.4, Cieslinski and Mikos [59, 60] modified the
technique of flame spraying. Their modification consisted in joining the metal and an additive in a
controlled way. The first stream of melted metal, made in the spraying apparatus, mixes with the
additive while being deposited on the base. After mixing with the metal and reaching the saturation
temperature, the additive evaporates from the melted metallic layer and producing pores in it. This
technique was used to produce a uniform porous coating on the cylindrical surface of a tube
(d = 8.15 mm). Porous coatings were made of aluminum, copper, molybdenum and stainless steel
1H18N9T on flat surfaces while coatings on cylindrical surfaces were made only of aluminum.

The majority of Cieslinski’s experiments with boiling surfaces made with the modified flame
spraying technique [52, 53, 57, 58] were performed for flat surfaces, therefore the conclusions
below refer to these results. The thickness of porous coverings ranged from 0.06 mm ≤ δ ≤ 2.00 mm,
the porosity from 0.051 ≤ ε ≤ 0.639 and the pore effective diameter from 1.79 µm ≤ De ≤ 10.5 µm.
The test fluid was water, boiling at atmospheric pressure.

In his tests, it was noted that the boiling on flat surfaces is distorted by the so-called heel
effect, which results in boiling incipience occurring at the edge rather than on a smooth surface
itself, which yields smaller superheats ΔTsc. For boiling on a flat surface facing upwards, the
bubbles depart freely, whereas on a cylindrical surface, bubbles from the lower perimeter float
upwards along the perimeter, affecting those generated above. In spite of those differences, the
conclusions drawn regarding flat and cylindrical porous surfaces manufactured with his new flame
spraying technique are similar to those that refer to surfaces obtained with other techniques,
namely [52, 53, 57, 58]. They are as follows:

a. nucleate boiling incipience on all porous surfaces took place at superheats that were
much lower than those for smooth surfaces;

b. a considerable increase in heat transfer coefficient was achieved in comparison with


the smooth surface, but this beneficial effect decreases as the heat flux density grows;

c. aluminum is the best material to manufacture porous coatings with the modified
flame spraying technique as it produces a greater and more secure nucleate boiling
enhancement than copper or the steel 1H18N9T; molybdenum proved to be useless
with this spraying technique according to [52, 58]; all the results are shown in
Fig. 3.33.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 87


Chapter 3

The impact of the coating thickness on heat transfer was thought to be ambiguous as regards
coatings of high porosity (0.41 ≤ ε ≤ 0.64) as can be seen in Fig. 3.34 from [58]. In this diagram,
4 –2
the heat flux density is the parameter. For small or medium heat fluxes (1 ∙ 10 W m and
4 –2
4 ∙ 10 W m ), the heat transfer coefficient has weak maxima, corresponding to the optimum
porous thickness. This effect is observed for porous coatings produced with various techniques. For
5 –2
higher heat fluxes (q ≥ 2 ∙ 10 W m ), however, the thinner is the coating, the higher is the heat
transfer coefficient. This means that the capability of the capillary-porous structure to “exhaust”
the vapor is decisive for the structure to attain a good boiling enhancement and vapor flow is
impeded as the coating gets thicker.

4
5⋅10

5
α, W⋅m -2⋅K-1

4 3 2
104

α = 1.25DTsl3.147

103
1 10 20
DTsl , K

1 – smooth, copper; 2 – 1H18N9T, ε = 0.5, De = 8.31 µm, δ = 0.48 mm;


3 – molybdenum, ε = 0.29, De = 3.85 µm, δ = 0.49 mm ; 4 – aluminum, ε = 0.2,
De = 3.58 µm, δ = 0.5 mm; 5 – copper, ε = 0.29, De = 3.16 µm, δ = 0.51 mm

Fig. 3.33. Impact of the porous coating material on the heat transfer coefficient for
water boiling at atmospheric pressure from Cieslinski [58]

Cieslinski [53, 58] compared his experimental results with the correlations proposed by other
authors for porous coatings obtained with techniques other than flame spraying. The correlation
by Ayub and Bergles [20], for which it is necessary to know two experimental constants and the
density of nucleation centers N, see Fig. 3.35, yielded only partial agreement. The boiling site
density was determined experimentally and the adopted experimental constants were like those in
Zuber and Nakayama [167, 168].

The experimental results obtained by Zhang and Zhang [328] for bronze porous coatings and
tests with water, ethanol and R-113 at atmospheric pressure for three layers of different thickness

88 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

(δ = 0.3 mm, 0.9 mm and 1.14 mm) are in agreement, in the qualitative sense, with the results
discussed previously [51 – 58, 72, 98, 160].

31·103
q = 2 ⋅ 105 W⋅ m −2
q = 4 ⋅ 104 W⋅ m −2
q = 1 ⋅ 10 4 W⋅ m −2
26·103

α, W⋅m -2⋅K -1 21·103

16·103

3
11·10

6·103

1·103
0 0.3 0.6 0.9 1.2 1.5 1.8 2.1
δ, mm

Fig. 3.34. Heat transfer coefficient α plotted versus the thickness δ of the aluminum
coating of high porosity (0.41 ≤ ε ≤ 0.64) from Cieslinski [58]

The basic difficulty arising in all techniques of thermal spraying is the production of capillary-
porous structures of considerable and controllable (repeatable) porosity and with open pores. In
order to achieve heightened porosity, some additives must be added to the materials sprayed,
which evaporate as the sprayed layer cools down. Using this spraying technique, it is possible to
get porous coatings of uniform thickness only for simple shapes of the base surface, i.e., for flat
and cylindrical surfaces.

Boiling incipience in sprayed structures occurs for the superheats that are lower or the same as
those that are characteristic of technically smooth surfaces. Similar to sintered powder structures,
in order to get the highest values of the heat transfer coefficient, it is necessary to optimize the
coating’s structural parameters such as porosity, thickness and pore diameter. Experimental
results do not provide the basis for unambiguous evaluation of the influence of each of the above-
mentioned parameters. The ratio of heat transfer augmentation with relative to a smooth surface
decreases as the heat flux rises.

st nd
The occurrence of a marked hysteresis phenomena (that of 1 and 2 kind) is characteristic of
the majority of thermally sprayed coatings. As a result, their industrial applications are restricted
because of this operating complication, at the same time the occurrence of the phenomena
reveals the complexity of heat transfer processes taking place inside porous coatings.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 89


Chapter 3

a) b)
5
5⋅10 5⋅105

105
105

q, W⋅m-2
q, W⋅m-2

104
4
10

5⋅103 2⋅10 3
1 10 20 0.5 1 10
DT sl, K DTsl, K

a) 1H18N9T covering, δ = 0.48 mm, ε = 0.5


b) aluminum covering, δ = 0.23 mm, ε = 0.59

Fig. 3.35. Comparison of experimental data of Cieslinski [53, 58] with the Ayub and
Bergles correlation [19, 20]

3.2.4. Mesh structures

Layers of metal mesh are frequently used as heat pipe wicks. Their basic task in that case is not to
enhance heat transfer but to increase the capillary transport of the liquid from the heat pipe’s cold
zone to the evaporator. The majority of the research on boiling heat transfer for mesh structures
has therefore been conducted under the conditions of capillary liquid feed or, less frequently, for
pool boiling.

Capillary–porous mesh structures are constructed in a simple way by means of deposing


subsequent layers of mesh. Inside a cylindrical heat pipe, those layers are pressed against
one another as a result of their being bent and hence they have a residual flex. When the base
surface is flat, mesh layers are sintered or pressed on the surface. Pressing does not guarantee
a good thermal contact between layers, hence the value of contact resistance is often high.
Mesh structures offer the possibility of producing capillary-porous structures of pre-selected pore
diameters and a homogeneous distribution, which facilitates the modeling of boiling heat transfer.
Another advantage of mesh structures is that all of their pores are open.

The works of Sasin et al. [226 – 229] as well as Hasegawa et al. [96] were undoubtedly among
the first publications dealing with experimental investigations on boiling heat transfer for mesh
structures.

90 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Sasin et al. [227] was interested in water and ethanol boiling inside a heat pipe wick. Two or three
layers of brass or acid resistant steel 1H18N9T mesh were used to make the wick of the heat pipe.
To model the process, one-dimensional differential equations of mass, momentum and energy
conservation were written in a dimensionless form for a unit element of the evaporator.

On that basis, the following similarity numbers were obtained:

(3.47)

The number Nφ describes the filtration properties of the capillary porous structure and Np
represents the surface tension to pressure ratio where c1 is a constant that depends on the liquid-
vapor meniscus form and c1 = 2 when the meniscus has a spherical shape.

Their investigations involved heat pipes, for which external diameters were from dpipe = 10 to
25 mm, lengths were from Lhp = 200 to 1200 mm, the mesh diameters (measured as a clearance)
–10 2
varied from Reyt = 0.04 to 0.20 mm, permeability changed from K = (2 to 6) 10 m , and
4
pressures were for pg ≥ 10 Pa. Having analyzed numerous experimental data, they arrived at the
following equation:

(3.48)

where Fevaporator is the surface of the heat pipe evaporator, Fwick is the mesh filing cross- section
area, and Reyt,opt is the optimum dimension of the mesh opening for water and ethanol
(Reyt,opt = 0.14 mm). The exponent n = 4.25 for Reyt < Reyt,opt, whereas for Reyt > Reyt,opt n = 0.43.
Equation (3.48) correlates the experimental data for water and ethanol in the zone of the heat pipe
evaporator with an accuracy of ± 20%, as shown in Fig. 3.36 [227].

In equation (3.48), the liquid thermal properties are determined for the mean temperature of the
evaporator surface temperature and the vapor temperature. In the case considered, the heat
5 –2
flux was found to be as high as 2 to 3 ∙ 10 W m , i.e. an order higher than in the case when the
evaporation in the evaporator proceeds only from the wick’s external surface.

In their earlier work, Sasin, Fedorov and Sorokin [227] analyzed the impact of mesh diameter
on the boiling process. It was stated that at a fixed number of mesh layers, there occurs the
optimum mesh dimension Deyt,opt, for which the highest value of the heat flux is obtained. The
drop in the heat flux dissipated for Deyt < Deyt,opt is explained by the increase in hydraulic resistance
for the vapor to depart from the capillary porous structure, while for Deyt > Deyt,opt it is attributed to
insufficient capillary feed.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 91


Chapter 3

Fig. 3.36. Prediction of experimental data for evaporators of heat pipes with mesh
wicks from Sasin, Fedorov and Sorokin [227]

Discussing boiling heat transfer in mesh structures, it is impossible to pass over the work of Abhat
and Seban [1]. They conducted experimental investigations for cylindrical surfaces, both smooth
and wrapped up with meshes or metal felt. The tubes were either immersed in a large liquid
volume or the porous structures were fed in a capillary flow. The experimental data in the form
q = q(ΔTsl) were collected for water, acetone and ethanol at atmospheric pressure for the heat
3 –2 5 –2
fluxes from 1.6 ∙ 10 W m to 1.6 ∙ 10 W m . Wrapping pipe in metal mesh did not cause a
considerable change in the heat transfer coefficient for water. When acetone and ethanol were the
boiling liquids, a drop in the heat transfer coefficient was observed as the thickness of the metal
mesh increased. For both pool boiling and the porous layer capillary feeding, similar values of
the heat transfer coefficient were obtained. Nucleate boiling incipience on the tubes wrapped in
porous coatings made from the meshes took place at smaller superheats of the heating surface
than for a smooth pipe.

Meshes made of non-metallic materials are hardly ever used. Djundin et al. [76, 77] investigated
how glass mesh affected nucleate boiling incipience (a single layer, δ = 0.3 mm in thickness). It
was found that, similarly to metal coverings, nucleate boiling incipience takes place at superheats

92 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

that are smaller than for a smooth surface. The other conclusion drawn by the authors, that the
pressing of the mesh against the base is decisive for the heat transfer intensity, seems obvious but
is nonetheless important.

A physical analysis of the boiling process inside capillary-porous structures of different mesh
diameters was made in the experimental work by Fedorov [151]. The author examined both
homogeneous structures and those with vertical channels drilled in order to facilitate vapor escape.
Meshes were welded as a stack to the base. Horizontal heating surfaces had a cylindrical shape,
70 mm in diameter. Degassed and distilled water supplied to the porous structure by capillary feed
was the boiling fluid. The stainless steel square mesh dimensions were as follows:
Rsk = 40 – 80 – 140 – 160 – 200 µm, the uniform porous layer thickness was 1 ± 0.1 mm.

Exemplary results are given in Fig. 3.37. It can be seen that, especially for the finest mesh
(Rsk = 40 µm), perforation intended to facilitate the vapor departure from the porous structure
increases the heat transfer coefficient by up to 20 times for surface 3 compared to the smooth
surface. The highest values of the heat transfer coefficient in the region of nucleate boiling are
obtained for the finer meshes, Reyt = 40 – 80 µm, with additional perforation. As the heat flux in
the fine meshes grows, there is a gentle transition to the film boiling zone in Fig. 3.37. The heat
transfer coefficient increase with decreasing mesh size, when compared with the results for the
larger mesh, could be explained by an increase in the thermal conductivity of the porous layer due
to diminished mesh dimensions. That enhances evaporation from the layer upper surface. At the
same time, vapor is more easily transported away from the layer due to the perforations. Semena
and Zaripov drew similar conclusions when they investigated metal fibrous structures [238]. Both
the experimental results in Fig. 3.37 and observations of the boiling process in [85] suggest that
the hydraulic resistance to vapor flow and the porous layer’s thermal conductivity are of primary
importance to the heat transfer intensity obtained.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 93


Chapter 3

4
3⋅10

2
3

4
4 5
10
8
α, W⋅m-2⋅K-1

6
6

2 1

103
4⋅104 6 8 105 2 4 6 8⋅105
-2
q, W⋅m

1 – Reyt = 40 µm, without additional perforation; 2 – Reyt = 40 µm, perforation


D = 1 mm, distance between openings l = 2.5 mm; 3 – Reyt = 40 µm, perforation,
l > 2.5 mm; 4 – Reyt = 140 µm, without perforation; 5 – Reyt = 140 µm, perforation,
l = 5 mm; 6 – technically smooth

Fig. 3.37. Heat transfer for water boiling on surfaces with mesh porous coverings from
Fedorov [85]

Smirnov, Afanasev and colleagues [3, 5, 247 – 251, 254, 255], including Poniewski and others [4,
6, 252, 256], conducted extensive experimental research and theoretical considerations on heat
transfer from meshes. These studies were concerned with water, ethanol and R-113 boiling on
mesh capillary-porous structures. The porous layer was fed with liquid by capillary flow.

A voluminous set of experimental data was presented in Afanasev’s doctoral thesis [3]. The
pressures tested ranged from 0.005 to 0.1 MPa; the meshes were made of stainless steel, brass
and copper; the mesh dimensions were Deyt = 40 – 45 – 60 – 80 – 125 – 200 – 450 – 1600 µm;
the number of mesh layers varied from 1 to 24. The liquid height above the porous structure
equaled h = 50 mm and it was maintained constant; the liquid suction height ranged from 0 to 140
3 6 –2
mm; the heat flux was changed from 3⋅10 to 0.97⋅10 W m ; the dimensions of the grid pressing
the meshes onto the heating surface were altered as well [3]. The mesh structures were both
homogeneous (the same mesh dimensions) and non-homogeneous. Homogeneous meshes
were found to produce results that were almost identical for pool boiling (liquid above the porous
structure) as for capillary feed, as shown in Fig. 3.38.

94 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

5
2·10
Water
p = 0.1 MPa

105
8 -1
-2
6
q, W⋅m-2

Technically smooth surface


Pool boiling

104
2 4 6 8
1 10 20
DTsl, K

Fig. 3.38. Examples of experimental results for three-layered, homogeneous mesh


structures: material – stainless steel; Reyt = 40 µm; 1 – pool boiling;
2 – capillary feed from Afanasev [3] and Smirnov, Afanasev and
Poniewski [252]

Boiling curves obtained with non-homogeneous meshes are shown in Fig. 3.39 and demonstrated
a characteristic change in slope at a certain value of the superheat ΔTsl, below which a significant
boiling enhancement takes place, where region I is that to the left of point A and region II is that to
the right. It is characteristic of both homogeneous and non-homogeneous structures that for small
superheats the heat flux dissipated through the mesh structure is higher than for a technically
smooth surface. Therefore, the value of the heat transfer coefficient is higher. Also nucleate
boiling incipience occurs for smaller ΔTsl in both Fig. 3.38 and Fig. 3.39. Similar trends were
observed by Poniewski and Wojcik in metal fibrous structures [217, 218].

Non-homogeneous cell distribution formed by imperfect alignment in the overlay of successive


mesh layers is also characteristic of multi-layer mesh structures of the same mesh dimensions.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 95


Chapter 3

4·105
Pool boiling

105 Technically smooth surface II


8
6
q, W⋅m-2

4
A

I
2
Ethanol
-1
104 -2
8 -3

3
4·10
2 4 6 8 10 20 40
DTsl, K

Fig. 3.39. Boiling heat transfer on non-homogeneous mesh structure: material –


stainless steel, Reyt,1 = 60 µm + Reyt2 = 200 µm + Reyt,3 = 125 µm;
1 – p = 0.01 MPa, 2 – 0.03 MPa, 3 – 0.1 MPa from Afanasev [3] and Smirnov,
Afanasev and Poniewski [252]

In this case, the “cell” simply means an opening for vapor flow. If the deposition of subsequent
mesh layers is not perfect, which is the most common case, the openings for vapor flow will vary
between layers. Our Russian colleagues call this phenomenon non-homogeneous cell distribution,
which has a stochastic character.

Depositing a few mesh layers onto one another in such a way that all cells are identical is, in
practice, rather impossible to do. Photographic observations confirmed the above conclusion,
Fig.3.40. With pool boiling and a detailed analysis of the data in Fig. 3.38 (points denoted 1), a
change in the slope of the boiling curve is seen for higher values of the heat flux.

96 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

1 – mesh wire; 2 – heating surface; 3 – liquid layer; 4 – vapor bubble; 5 – additional


nucleation center from Afanasev [3] and Poniewski, Wojcik and Afanasev [218]

Fig. 3.40. Photograph of water boiling at atmospheric pressure on a metal mesh,


4 –2
Deyt = 0.5 mm, 3 layers: a) region I, q ~ (5 to 10) 10 W m (450 frames/s);
5 –2
b) region II, q ~ 3.5 ∙ 10 W m (3500 frames/s)

The shape of the boiling curve (on a logarithmic scale) in the left region I in Fig. 3.39 could
indicate that the heat transfer coefficient is nearly constant, which is the consequence of an
almost constant number of nucleation centers.

The following reasoning is behind this conclusion:

1. Let’s assume first that the boiling heat flux is constant when the heating surface
superheating, ΔT, increases so that in this case the heat transfer coefficient, α, has to
decrease. The only “possible” physical reason would be a decrease in the number of
active sites.

2. Now, let’s take into consideration the second case. The boiling heat flux slightly
increases when ΔT increases. In this case, it is quite rational to assume that the heat
transfer coefficient is constant, because of a constant number of active sites.

3. For a certain value of ΔT, the slope of the boiling curve increases substantially. Such
a change can be caused only by a sharp change in the intensity of the nucleation
process, that is, a sharp increase in the number of the nucleation centers.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 97


Chapter 3

Further increasing the heat flux above the point of the inclination of the boiling curve in Fig. 3.39,
denoted as point A, leads to a change in the heat transfer mechanism in region II, which becomes
similar to that of boiling on a smooth surface (note that these data above point A fall on a line
parallel to that of pool boiling on a smooth surface indicated by the dashed line). Fig. 3.40 shows
photographs of the boiling process in regions I and II in Figure 3.39, obtained with a high-speed
photographic camera [3, 218]. Photograph 3.40a shows the process in region I, where vapor is
generated in all cells of the capillary–porous structure adjacent to the heating surface.

Some cells are closed and heat transfer proceeds only at their bottom, whereas vapor
condensation takes place in the upper part of the cells. The cells behave like micro-heat pipes.

Photograph 3.40b shows the process in region II. Additional nucleation centers are visible at
the bottom of the capillary–porous structure, a fact that confirms the similarity of the boiling
heat transfer mechanisms in the mesh structure and on a smooth surface. The boiling feature
characteristic of both processes is the number of nucleation centers with respect to the heat flux,
where N = f (q). Owing to differences in the dimensions of porous cells, when the heating surface
superheat ΔTsl grows, cells of smaller and smaller dimensions get activated.

Smirnov based his boiling heat transfer model for mesh structures on the concept of the
substitution of the structural skeleton material with a collection of microfins [247, 249, 252, 256].
An additional assumption was made that a liquid microlayer covering the surfaces of equivalent
microfins was decisive for heat transfer. From the solution of the heat conduction equation for an
equivalent microfin, as well as from the Bernoulli equation for a liquid micro-film moving over the
microfin surface, taking also into account the pressure drop in the vapor stream moving between
the fins, the following relationship was obtained:

(3.49)

4σlg Tsat
where the nucleation criterion is ∆T* = (see (2.1) and (2.11));
hlg ρg Deyt

ΔTsl = Ts – Tsat is the heating surface superheat with respect to the saturation temperature;
4 0.2
Le = [Deyt (f/O) ] is the characteristic dimension of elementary repeatable cells in the capillary
2
porous structure, in which Dskl is the open pore diameter (mesh diameter), f = (Deyt + dwire) – Deyt
2

is the sectional area of the skeleton structure element (elementary, equivalent microfin); O = π Deyt
is the wetted circumference; ε* is the superficial porosity (porosity of the wetted part of the
capillary-porous structure taking part in vapor generation) [251].

98 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The correlation (3.49) describes the author’s experimental data with the accuracy ± 50%, which
is illustrated in Fig. 3.41, according to [3, 5, 250 – 252]. The accuracy of (3.49) was initially verified
with respect to data from other authors, yielding similar results, namely an accuracy of ± 50%
[1, 3, 62, 86, 157, 227, 247]. That means that (3.49) correctly correlates a significant amount of
experimental data. The model, its simplifying assumptions and results are discussed in more
detail in Chapter 4.

Among notable achievements of Smirnov and co-workers [4, 5, 6], we count the model of nucleate
boiling as well as a semi-empirical expression for the calculation of the critical heat flux. The
model and experimental methods are presented in 4.4.2.

Eva, Asakavichjus and Gajgalis [16, 84] analyzed the properties of multilayer mesh structures
with a view to making use of them as heat pipe wicks. Copper and stainless steel structures were
investigated with mesh dimensions of 0.07 and 0.1 mm and made with 2, 8 and 12 number of
layers. Water, ethanol and R-113 were the boiling test fluids. The liquid was fed to the porous layer
by means of gravitational flow or by immersing the layer in the liquid. Heat transfer enhancement,
when compared with the surface without a porous covering, for R-113 in the examined heat flux
3 4 –2
density range from 6 ∙ 10 to 2 ∙ 10 W m and also for water and ethanol in the range from
5 5 –2
1.5 ∙ 10 to 3 ∙ 10 W m , was clearly observed. The enhancement effect is weakened as the heat
flux grows, which is also characteristic of other porous structures.

As in the investigations conducted by Smirnov et al. [3, 5, 247, 250, 251], as illustrated in Fig. 3.38,
it was found out that the heat transfer coefficient did not depend on the manner the capillary-
porous structure was fed with liquid.

The analysis also covered the influence of the boiling liquid thermal conductivity and the capillary-
porous structure skeleton material on the heat transfer coefficient. Heat transfer coefficients
obtained for copper structures were on average 1.3 times higher than those for stainless steel.
When all other parameters were constant, heat transfer coefficients for water were 1.8 times
higher than those for ethanol and R-113. Furthermore, heat transfer coefficients were inversely
proportional to mesh dimensions and their highest values were recorded for single-layer coverings
with increasing mesh dimensions as the distance from the base grew. The results confirmed
earlier suggestions that the micro-structure’s hydraulic resistance to vapor outflow was decisive
for heat transfer enhancement in the capillary-porous layers.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 99


Chapter 3

a.b. = as before, s.s. = stainless steel, ls = layers, m.o.p.i. = mesh openings per inch
1 – s.s, 3 ls, Deyt = 60 µm, p = 0.1 MPa; 2 – a.b., 2 ls, p = 0.01 MPa; 3 – a.b., 3 ls;
4 – a.b., 2 ls, p = 0.03 MPa; 5 – a.b., 3 ls; 6 – copper, 2 ls, Deyt = 45 µm, p = 0.1 MPa;
7 – s.s., 3 ls, Deyt = 40 µm, p = 0.1 MPa; 8 – a.b., Deyt = 125 µm; 9 – a.b., 325 m.o.p.i.,
p = 0.1 MPa; 10 – nickel, 125 m.o.p.i., p = 0.1 MPa [62]; 11 – s.s., 3 ls, Deyt = 60 µm,
p = 0.1 MPa; 12 – a.b., 2 ls, Deyt = 125 µm, p = 0.01 MPa; 13 – a.b., p = 0.03 MPa;
14 – copper, 2 ls, Deyt = 40 µm, p = 0.1 MPa; 15 – a.b., p = 0.03 MPa; 16 – a.b.,
p = 0.01 MPa; 17 – s.s., 2 ls, Deyt = 70 µm, p = 0.1 MPa [16]; 18 – s.s., 2 ls,
Deyt = 40 µm, p = 0.1 MPa; 19 – a.b., Deyt = 60 µm; 20 – s.s., 3 ls, Dskl = 40 µm +
Deyt = 125 µm + Deyt = 125 µm, p = 0.1 MPa; 21 – s.s., 8 ls, Deyt = 100 µm, p = 0.1 MPa

Fig. 3.41. Comparison of (3.49) to experimental data of Smirnov and Afanasev [3, 5,
249 – 252]

100 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

For the sake of experimental data generalization, the following similarity numbers were put
forward:

(3.50)

cp,l σlg ρl Tsat


where l* = 2
, and λef is the effective thermal conductivity computed on the basis of
(ρg hlg)

the empirical dependence λef = c λnl [84]. Making use of these similarity numbers, the following
correlation was proposed:

0.24 4.2
Nu = 0.24 Re Pr (3.51)

It predicted nearly all the experimental data within an error of ± 22%, which is illustrated in
Fig. 3.42.

Yamaguchi and James [321] used the definitions of the similarity numbers given by (3.50) to
generalize their investigations with a correlation whose form was similar to (3.51):

0.41 0.1
Nu = 35 Re Pr (3.52)

It was developed from heat transfer experimental results for water and R-113 boiling in the
5 5
pressure range from 0.25 ∙ 10 Pa to 1 ∙ 10 Pa. However, the correlation (3.52) accounted for
geometrical parameters and thermal properties of the mesh capillary-porous structure in a much
more precise way. The porosity of the mesh structure was computed from the expression:

(3.53)

where ε1 is the porosity of a single mesh and s is the array coefficient defined by

(3.54)

where δ is the real thickness of the porous structure, n is the number of meshes and δ1 is the
thickness of a single layer. Finally, the effective thermal conductivity of the mesh porous structure
was computed in accordance with the formula:

(3.55)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 101


Chapter 3

a.b. = as before, s.s. = stainless steel, ls = layers


1 – R-113, copper, 8 ls, Deyt = 100 µm, wire rings, structure not immersed; 2 – a.b., held
down with a special grip; 3 – a.b., immersed structure; 4 – see 1, immersed structure;
5 – ethanol, s.s., 12 ls, Deyt = 70 µm, held down with a special grip, not immersed; 6 – a.b.,
copper; 7 – a.b., 8 ls, Deyt = 100 µm, wire rings; 8 – a.b., held down with a special grip,
immersed; 9 – a.b., not immersed; 10 – see 7, immersed; 11 – see 6, 2 ls, rings; 12 – a.b.,
not immersed; 13 – see 11, Deyt = 40 µm; 14 – a.b., immersed; 15 – see 10, water; 16 – a.b.,
held down with a special grip; 17 – see 5, not immersed; 18 – water, s.s., 11 ls, Deyt = 40 µm,
immersed [1]; 19 – a.b., capillary transport; 20 – water, Monel spheres, capillary transport
[86]

Fig. 3.42. Comparison of (3.42) to experimental data on mesh structures by Eva,


Asakavichjus and Gajgalis [84]

The formulas (3.53) to (3.55) were introduced to the definitions of similarity numbers (3.50) and
then to the correlation (3.52). The new method compared to Yamaguchi and James’s experimental
results [321] however gave a major error band (up to 100%), which also resulted from a
considerable dispersion of the experimental data themselves.

102 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Summing up the results of their investigation, Yamaguchi and James [321] concluded that both
the roughness of the surface and pressure significantly affected heat transfer, especially for
R-113. The visual observations they conducted of the boiling process convinced them that vapor
generation began with nucleation on the surface of the mesh capillary-porous structure base.

Tolubinskij and co-researchers [289 – 295] analyzed heat transfer on heating surfaces covered
with one layer of metal mesh (the majority of their investigations), metal felt or perforated foil.
Meshes were made of bronze, nickel, stainless steel and brass, while mesh dimensions Deyt
ranged from 0.04 to 3.0 µm. The examined liquids were: distilled water (most investigations), butyl
alcohol, n-heptane, acetone and benzene. Test pressures ranged from 0.01 to 0.5 MPa and the
feed was by capillary flow. The mesh contact with the base surface was checked throughout the
experiment.

These researchers gave special attention to the mechanism of heat transport from the heating
surface to the liquid. They chose to focus exclusively on the analysis of the properties of one-
layer coverings based on their intuitive observation that it was this part of the porous layer directly
adjacent to the heating surface that controlled the heat flux transported by this layer. Furthermore,
an increase in the porous layer thickness merely hampers vapor transport out of the layer.

An important task undertaken by Tolubinski’s team [289 – 295] was to experimentally determine
the relationship between the critical heat flux and the pressure and with respect to the structural
parameters of the layer. This issue is discussed in detail later in 4.4.

Owing to photographs and observations of the boiling heat transfer process, it was found that:

5 –2
1. For elevated heat fluxes (e.g. 10 W m for water), evaporation in the pores of the
capillary-porous structure took place solely from the surfaces of the liquid menisci,
that is, the boiling process from micro-pores in the base surface did not occur.

2. Evaporation from menisci surfaces was accompanied by the departure and


entrainment of liquid droplets, caused by non-uniformities of the structure’s
hydrodynamic properties.

3. The entrainment of liquid droplets could lead to dry out of the structure, which
ultimately leads to a larger heating surface superheat.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 103


Chapter 3

The absence of nucleate boiling inside the porous layer in spite of the considerable heat flux was
attributed by Tolubinskij et al. [294] to the following reasons:

a. the formation of a very thin and stable liquid film in the central part of the meniscus
due to the action of surface tension, which heightens the intensity of heat transport
through the meniscus. According to their observations, surface tension stabilizes the
central part of the meniscus. It results in heat being transferred by conduction through
the liquid layer and causes evaporation from the meniscus interface;

b. a large temperature gradient in the capillary-porous structure skeleton adjacent to the


evaporating liquid meniscus, which made it difficult to initiate stable nucleate boiling;

c. high effective thermal conductivity of the structure, which resulted in the heat being
conducted through its skeleton to the adjacent liquid layers.

A sharp drop in the heat transfer coefficient was observed at various heat fluxes depending on
the fluid and surface, as shown by the data in Fig. 3.43. This phenomenon was explained by
the drying up of portions of the capillary-porous structure due to insufficient liquid feed (too low
capillary pressure increment) and the entrainment of drops from menisci surfaces.

3
8⋅10

4
α, W⋅m -2⋅K-1

-1 -4
-2 -5
-3 -6
10 3

8⋅10 2 4
10 2 4 6 8 105 2⋅10 5
-2
q, W⋅m

Fig. 3.43. Heat transfer coefficient plotted versus heat flux for mesh structures and
butyl alcohol (1, 2), n-heptane (3, 4), benzene (5) and acetone (6) for the
following conditions: pressure = 0.1 MPa: 1, 3 – brass mesh, Deyt = 0. 4 µm;
2, 4 – brass mesh, Dskl = 0.08 µm; 5, 6 – stainless steel mesh, Deyt = 1.8 µm
from Tolubinskij et al. [293, 294]

104 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Summing up the results of these investigations, on the basis of dimensional analysis, the following
dependence for heat transfer coefficient computation was put forward, where a comparison of the
experimental results and the correlation are shown in Fig. 3.44:

(3.56)

As can be seen in Fig. 3.44, the capillary-porous structure material or geometry did not
significantly influence the magnitude of the heat transfer coefficient and the wide range of physical
properties of the fluids tested was well captured. The empirical constants in (3.56) were: C = 11.5;
n = 0.8; m = 0.2 and u = 0.4. The correlation (3.56) is applicable when the structures are fully
saturated with liquid, i.e. without dry spots. For water, the dependence was applied to heat fluxes
4 6 –2 5
ranging from 2 ∙ 10 to 1.5 ∙ 10 W m and for organic liquids, to those ranging from 1.5 ∙ 10 to
6 –2
1.5 ∙ 10 W m .

Water, 0.05 - 0.4 MPa: 1 – stainless steel, Deyt = 1.8 mm; 2 – stainless steel, mesh eyelet
2
0.42 × 0.49 mm ; 3 – brass, Deyt = 0.58 mm; 4 – copper felt; 5 – brass, 2 layers, D1, eyt = 0.58
mm; D2, eyt = 1 mm; 6 – nickel plate, perforated, D = 0.7 mm, the heating surface clearance
is 0.2 mm; 7 – data from [234]; 8 – butyl alcohol, brass, Deyt = 0.4 mm; 9 – benzene,
stainless steel, Deyt = 0.8 mm from Tolubinskij et al. [294]

Fig. 3.44. Comparison of correlation (3.56) to experimental data

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 105


Chapter 3

Later works by Tolubinskij and co-researchers [292] confirmed other scientists’ conclusions that
depositing a mesh porous coating generally produced an advantageous effect on the boiling
process. Furthermore, a mesh structure also produced a significant increase in the critical heat
flux when compared with a smooth surface. The increase diminished as the pressure grew. The
results obtained earlier by Sasin et al. [227] were also confirmed, namely that at a pre-set mesh
thickness, an optimum of the mesh dimensions was found for which the highest value of the
critical heat flux was noted. The mesh structure also led to a drop in the heating surface superheat
indispensable for boiling incipience relative to smooth surface, as illustrated by the data in
Fig. 3.45.

1, 2, 3 – water, 1 layer, Deyt = 0.5, 1 and 2.5 mm; 4 – ethanol, 1 layer, Deyt = 1 mm;
5 – acetone, 1 layer, Deyt = 1 mm; 6, 7 – water, 2 and 3 layers, Deyt = 1 mm;
8, 9 – nitrogen* and R-113*; 10 – nitrogen*; 11 – water*; 12 – R-22; 13 – R-12*;
* denotes data by other researchers, according to Tolubinskij et al. [292]

Fig. 3.45. Impact of pressure on boiling incipience for mesh structures

A problem faced with the application of mesh porous structures is how to appropriately hold them
down on the base surface. Tolubinskij et al. [290] investigated the impact of faulty contact on
the heat flux. They found that for small heat fluxes poor contact actually enhanced boiling heat
transfer if the clearances or gaps between the mesh and the base were not too large.

The investigations conducted by Tsay and co-researchers [298] were similar to those carried out
by Tolubinskij et al. [289 – 295], as the former also focused on the analysis of boiling on a heating
surface covered with only one layer of a steel mesh. Water, at atmospheric pressure, was the
investigated liquid. At the same time, they observed the boiling process on the porous surface
with and without a steel covering. It was noted that the heat flux decreased when the height of
the liquid level above the porous layer diminished from 60 to 10 mm, reaching the minimum in
the range from 10 to 5 mm and then grew for 2 mm. Covering the heating surface with a mesh
layer significantly enhanced boiling heat transfer, if the mesh dimension was comparable with a

106 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

departing bubble diameter, especially for small liquid heights. It was observed that an increase in
the heating surface roughness enhanced boiling heat transfer, but it did not play any role when the
surface was covered with a mesh layer [298].

Rannenberg and Beer [222] made an attempt at the classification of the dominant heat transfer
mechanisms occurring during boiling in a mesh structure as the function of threads/inch and the
heating surface temperature. The structures were made of steel and bronze meshes ranging from
38 to 198 threads per inch, which consisted of two to nine layers. Heat transfer processes taking
place in R-11 and R-113 at atmospheric pressure were investigated. Boiling curves, as those in
Fig.3.46, underwent qualitative analysis. The qualifying factor was a value of the exponent n in the
heat flux equation: q = DTsln.

10 6
Smooth
surface
106 mesh
157 mesh
Pool boiling

10 5 q = CDTsln
q, W⋅m-2

n = 2.57 n = 1.53

10 4
n = 4.68

n = 1.00

n = 1.33
n = 1.22
10 3
1 2 5 10 20 50
DTsl, K

Fig. 3.46. Characteristic boiling curves fit from experimental data for mesh porous
structures from Rannenberg and Beer [222]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 107


Chapter 3

On the basis of the above results and also from observations of the boiling process, four zones
were identified for various heat transfer mechanisms, the function of mesh threads per a length
unit and the dimensionless temperature, the latter being the quotient of the real temperature of
the heating surface, Ts, and the surface temperature Tphase, at which the change of phase of the
boiling liquid takes place, Fig. 3.47. They are the zones:

a. pure heat conduction without visible change of phase, n = 1;

b. convection-conduction heat transfer, change of phase can be seen, n > 1;

c. heat conduction with visible change of phase, n ≈ 1;

d. convection heat transfer with visible change of phase, n >> 1.

Fig. 3.47 illustrates that for the 38 to 106 mesh densities, conduction is dominant at low surface
temperatures only, followed by the convective-conductive process up to Ts/Tphase = 1. For the
135 to 198 mesh densities, pure conduction prevails up to Ts/Tphase = 1 because the small size of
the structure prevents convection. Except for the 198 mesh density, convection with visible phase
change is dominating at Ts/Tphase > 1. For the 198 mesh density, heat transfer takes place by pure
conduction with visual formation of vapor.

Fig. 3.47. Heat transfer mechanism inside the mesh porous structure as the function
of the number of mesh threads per a length unit and dimensionless
temperature Ts /Tphase, according to Rannenberg and Beer [222]

On the basis of similarity theory and the analysis of the equations of mass, momentum and energy
conservation, it was found out that the controlling dimensionless numbers are the products of two
similarity numbers: Re Pr and Re Eu*. The results of the experimental investigations were generalized
in the form of the following correlation:

(3.57)

108 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where the dimensionless groups are defined as:

The effective conductivity of the capillary-porous structure was calculated in accordance with the
formula (3.55) and the exponents in (3.57) were determined with the least squares regression
method. The definition of the effective pore radius Ref was not provided, and neither was the
relation between the thickness of a single mesh layer δeyt and the wire diameter dwire. It was stated
that the transition from pure conduction to convection and conduction took place at

(3.58)

which determined the control area (3.57). The correlation (3.57) approximated experimental data
for both convection areas, Fig. 3.47, i.e. for convection and conduction without the change of
phase (b) as well as for convection with the change of phase (d), where the error was ± 15% [222].

Owing to the use of mesh structures as heat pipe wicks, researchers have taken a considerable
amount of experimental data and made many correlations for heat transfer coefficient calculations.
Measurements were, however, taken for meshes obtained with various techniques. Moreover, the
ways meshes are joined with each other and held against the heating surface are nearly always
different. As a result, most of the vast experimental database cannot be grouped together to
develop a generalized correlation to compute the heat transfer coefficient.

One-layer mesh structures enhance heat transfer, whereas a greater number of layers often
worsen it relative to one-layer. The heat flux that can be dissipated increases when the effective
thermal conductivity of mesh structure grows, yet the manner of feed does not significantly affect
the heat transfer process inside the porous structure if the whole layer is saturated with liquid. An
increase in the inclination of the curve q(ΔTsl) for higher heat fluxes is particularly characteristic
of inhomogeneous mesh structures, which represents the activation of additional cells inside the
structure.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 109


Chapter 3

3.2.5. Metal fibrous structures

Other researchers’ investigations

Metallic fibrous structures have a completely different basic structure than the porous layers
discussed previously. They are made of thin metal (copper, brass, steels, nickel, aluminum and
other) or metal alloy fibers rather than particles or powders or meshes discussed in the previous
section. The fibers have diameters in the range of dfbr = 10 to 100 µm and their lengths are on the
order of Lfbr = 1 to 10 mm [232].

They are used as heat pipe wicks in the countries of the former Soviet Union, particularly in
the Ukraine. They have not been produced on an industrial scale because their manufacturing
process is time consuming. Moreover, heat pipes with metal fibrous capillary-porous wicks have
never entered into broad spread production. First, fibers of a set length and diameter are prepared
to form a flat layer. Then the fibers are sintered in a reducing atmosphere (e.g. hydrogen) and
pressed to obtain the intended thickness and porosity. The layer is then welded or soldered to the
base [113, 232].

The main research centers which manufacture and investigate such enhancements are the
Technical University of Kiev and the Kiev Metal Science Institute of the Ukrainian Academy of
Sciences. Such metallic fibrous structures show a number of advantages as follows:

• high value of boiling heat transfer coefficient confirmed by numerous experimental


investigations;

• the possibility of manufacturing coatings which have set structural parameters, i.e.,
thickness, porosity and skeletal conductivity;

• it is possible to obtain coatings of high porosity with all pores open;

• porous coatings can be deposited on base surfaces of complex shapes.

Four important notions are used for the sake of the assessment of the properties of metallic
fibrous coverings: skeletal conductivity – λskl, effective conductivity – λef, effective diameter of
the porous cell – Dp,ef and the maximum porosity – εmax. The above-mentioned quantities are
determined experimentally [113, 232] and each is discussed in more detail below.

110 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Skeletal conductivity

λskl is the thermal conductivity dependent on the pre-set geometry of the porous structure, the
thermal conductivity of the fiber and the porosity. It does not account for the thermal properties
of the liquid in the voids (it assumes the voids are occupied by vacuum or that the liquid has
zero thermal conductivity). Skeletal conductivity decreases as the porosity increases, which in
accordance with the assumptions that the thermal conductivity of the saturating liquid λ1 = 0 and
that the structure is ordered, can be written approximately as follows [82]:

2
λskl = λm(1 – ε) (3.59)

Experimental results obtained for metal fibrous capillary-porous structures confirm the validity of
the following expression expanded version of (3.59) [239] as:

(3.60)

where the exponential term takes into account fiber contact resistance.

Effective conductivity

λef is not a property of the capillary-porous structure as it depends instead on the thermal
conductivity of the liquid occupying the voids. In Zaripov [322] and also in monographs written by
Semena et al. [232], Kostornov [113] as well as Dulnev and Zarichnjak [82] one can find detailed
analysis on the skeletal and effective thermal conductivity and its modeling. These authors also
described measurement procedures used to determine the above-mentioned conductivities, λskl
and λef, which also depend on the fiber dimensions and contact resistances between fibers. In
the monographs [82, 113, 232], attempts were made at providing an analytical description of both
thermal conductivities λskl and λef on the basis of experimental results [238, 239].

Effective diameter

Dp,ef is simply an average hydraulic diameter of a porous cell, which characterizes a structure with
pore diameter of random distribution. It is determined by measuring the structure permeability to
gases [113, 232, 237].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 111


Chapter 3

Maximum porosity

εmax is understood to be the maximum porosity attainable for a structure made of a given sample
of fibers of pre-set diameter and length. Owing to numerous experimental data, Semena et al.
[232, 326] proposed the following relationship to calculate the maximum porosity:

–6 –4
εmax = exp (–dfbr / Lfbr) 10 ≤ dfbr / Lfbr ≤ 10 (3.61)

Semena and co-researchers from the Kiev Technical University, the Technical Thermophysics
Institute and the Materials Science Institute of the Ukrainian Academy of Sciences in Kiev [141,
178, 185, 232 – 241, 266 – 269, 325, 326] started and developed investigations into metal fibrous
structures. The investigations were later continued by Poniewski (one of the authors of this book)
and his former co-workers [4, 6, 187 – 189, 191 – 195, 202 – 208, 210 – 219, 252 – 261, 317].

The coatings examined by Semena et al. [141, 178, 185, 232 – 241, 266 – 269, 325, 326] were
made of metals (alloys) such as copper, nickel and stainless steels, which considerably differed in
thermal conductivity. Different boiling liquids such as nitrogen, water, ethanol, ammonia, acetone,
R-22 and R-113 were used in the experiments. The latter were conducted over a wide range of
coatings of changeable thickness δ (from 0.5 to 2.8 mm) and porosity ε (from 0.6 to 0.91), for pool
boiling and for capillary feeding of the liquid to the porous structure.

The detailed description of various research test facilities and measurement procedures can be
found in [141, 185, 232, 235, 241, 266 – 268]. A special research stand that simulated boiling and
condensation as well as liquid capillary transport specific to a heat pipe was also designed and
built. With it, it was possible to carry visual observations of the boiling process [232].

The basic conclusions drawn from these investigations on boiling under capillary feed conditions
were presented by Semena et al. in [232, 240]. The investigations were primarily geared towards
the design of heat pipes with wicks of capillary-porous metal fibrous structures. Measurements
were taken for cylindrical heat pipes of changeable diameters (from 6 to 38 mm) and flat pipes of
changeable dimensions (from 3 × 13 mm and 3 × 45 mm to 10 × 23 mm), and pipe lengths ranged
2 5 –2
from 100 to 2200 mm. The heat fluxes tested changed from 10 to 7 ∙ 10 W m and the pressure
3
varied from 2 ∙ 10 to 106 Pa.

112 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Measurements and visual observations led to a better understanding of the general characteristics
of the boiling heat transfer processes taking place in a heat pipe, where the heat transfer results
are depicted in Fig. 3.48 and can be described as follows. At small heat fluxes q, heat transport
through a capillary-porous structure filled with a liquid proceeds by means of conduction with
evaporation on the external surface of the structure. In the case considered, the heat transfer
coefficient α is about an order of magnitude higher than the corresponding natural convection
coefficient on a technically smooth surface, immersed in a large liquid volume. It could be
determined from the relationship:

(3.62)

where Ω is the thermal contact resistance of the structure’s fibers. Both λef and Ω are determined
experimentally [238] or in accordance with the expression developed by Dulnev and Zarichnjak
[82]:

(3.63)

where η is the ratio of thermal conductivities of the liquid and the fiber material.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 113


Chapter 3

64

log α
48

α×103, W⋅m -2⋅K -1


32

16
-4
-5
0
0 0.6 1.2 1.8 2.4 3.0 3.6
q⋅104, W⋅m-2
3 Film
Zone II 2 boiling
Zone I 1
Conduction

Natural Transition
convection boiling

log q

Point 1 – technically smooth surface, pool boiling; Points 2 and 3 – heating surfaces covered
with a metal fibrous capillary-porous structure (2 – low porosity, ε ≤ 0.8 ≤ εmax; 3 – high
porosity, ε close to εmax. Points 4 and 5 – experimental data from heat pipe investigations:
4 4
4 – ε = 0.91, δ = 1.7 mm, psat = 5 ∙ 10 Pa; 5 – ε = 0.68, δ = 1.7 mm, psat = 2 · 10 Pa by
Semena et al. [232, 240]

Fig. 3.48. Heat transfer coefficient plotted versus heat flux (psat = const)

(3.64)

The validity of (3.62) is confirmed by the experimental results provided by Semena and Zaripov
[238]. Contact resistance depends on porosity and for ε ≥ 0.8 it is assumed that Ω ≈ 0, whereas for
–5 2 –1
ε < 0.8 it has a constant value of Ω = 5 ∙ 10 Km W .

The density of the heat flux q and the heating surface superheat ΔTsl, for which the transition
from the evaporation zone to the nucleate boiling zone occurs, depend on the capillary-porous
structure characteristics and the boiling liquid thermal properties. The superheat necessary for
nucleate boiling to take place is derived from a known criterion for bubble nucleation [21, 143, 173,
305, 306], which has the form:

(3.65)

114 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where the pore diameter is determined experimentally or from [232]:

(3.66)

In the boiling zone, heat transfer coefficients for the surface with metal fibrous porous layer are
significantly higher than for a technically smooth surface, as illustrated by the comparison in Fig.
3.48. The boiling zone on the metal fibrous porous layer can be divided into two regions. In region
I, the slope of α = α (q) is similar to that for a smooth surface. When q and psat grow, there is an
increase in the number of potential boiling sites. If the critical diameter of the bubble, resulting
from (3.65) where Dp,cr = 4 σlg Tsat /hlg ρlg ΔTsl, becomes smaller than Dp,min, that is Dp,cr ≤ Dp,min,
the transition from region I to region II takes place. With a further increase in the heating surface
superheat and heat flux, when Dp,cr > Dp,min, the cavities in the fiber surface, resulting from the
surface roughness, become active boiling sites.

At a certain value of q, pores close to the base heating surface begin to get filled with a
continuous vapor layer and then heat transfer proceeds primarily by means of conduction through
the capillary-porous structure skeleton. This leads, in turn, to a decrease in the heat flux that
can be dissipated, despite further increasing of the superheat of the heating surface, shown by
the sharp fall in the heat transfer coefficient in Fig. 3.48. Critical heat fluxes, dependent on the
structure porosity, can be both higher as well as lower than the those for a smooth surface, as
again illustrated in Fig. 3.48.

Observations of the boiling process in a metal fibrous porous structure led Semena and co-
researchers [232, 240] to the conclusion that vapor bubbles develop in it, because a superheated
liquid microlayer is constantly present and occupies a substantial part of the porous covering,
which quickens the bubble growth.

Following Labuncov [140], it was assumed that heat transfer through the microlayer takes place by
conduction, and hence the heat transfer coefficient could be represented by:

(3.67)

where δ1,ef is an average thickness of the superheated liquid microlayer. On the basis of [140],
it was also assumed that the microlayer thickness was supposed to diminish when there was
an increase in the number of nucleation centers inside the porous layer, the frequency of vapor
bubble generation grew and the liquid viscosity got lower. An indication of the increase in the

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 115


Chapter 3

number of potential boiling sites in region I in Fig. 3.48 is the ratio of the maximum diameter of the
pores in the structure to the critical diameter, for a given superheat of the heating surface:

N ~ Dp,max /Dp,cr 1 < Dp,max /Dp,cr < Dp,max /Dp,min (3.68)

In accordance with (3.65), the ratio Dp,max /Dp,cr = 1 corresponds to the beginning of region I for
nucleate boiling in Fig. 3.48, whereas Dp,max /Dp,cr = Dp,max /Dp,min represents the transition from
region I to region II.

The ratio of vapor generation rate to the porous layer thickness can be adopted as a measure of
the bubble departure frequency:

(3.69)

According to Labuncov [140], an average thickness of a superheated liquid microlayer amounts to:

(3.70)

Making use of (3.65), (3.68) and (3.69), we obtain from the last relationship:

(3.71)

After inserting (3.71) into (3.67), we find an expression for the heat transfer coefficient for boiling
region I in a metal fibrous porous layer:

(3.72)

which holds in the range

(3.73)

The thermal properties in (3.72) and (3.73) are determined at the saturation temperature, which
is in K. The constant C in (3.72) is computed on the basis of experimental data where “th” means
hyperbolic tangent:

4
C = C1 th (1.5 ∙ 10 ρg /ρl) (3.74)

116 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where the constant C1 is dependent on the heating surface material and fiber material (for copper
C1 = 0.8, for stainless steel C1 ≈ 0.4).

For nucleate boiling region II in Fig. 3.48, the heat transfer coefficient is given by:

|| 0.25 0.1
α = αlmt (q/qlmt) (A/A lmt) (3.75)

2
A = (λl) / (νl σlg Tsat) (3.76)

where αlmt, qlmt and A lmt correspond to the region I to region II transition boundary. They are
specified with the method of successive approximations, under the condition
ΔTsl = 4 σlg Tsat /hlg ρlg Dp,min.

Full line – dependence (3.72) and (3.75)


1 – ε = 0.91, δ = 1.3 mm, water; 2 – ε = 0.91, δ = 1.7 mm, water; 3 – ε = 0.91, δ = 2.8 mm,
water; 4 – ε = 0.86, δ = 2 mm, R-113; 5 – ε = 0.85, δ = 0.9 mm, methanol; 6 - ε = 0.83,
δ = 0.5 mm, water; 7 – ε = 0.82, δ = 0.5 mm, nitrogen; 8 – ε = 0.81, δ = 0.5, acetone;
9 – ε = 0.8, δ = 1 mm, water; 10 – ε = 0.78, δ = 1.8 mm, water; 11 – ε = 0.76, δ = 1.1 mm,
water; 12 – ε = 0.75, δ = 1 mm, nitrogen; 13 – ε = 0.74, δ = 1 mm, ethanol; 14 – ε = 0.72,
δ = 0.6 mm, R-22; 15 – ε = 0.7, δ = 0.9 mm, water; 16 – ε = 0.68, δ = 1.7 mm, water;
17 – ε = 0.63, δ = 0.6 mm, ammonia; 18 – ε = 0.6, δ = 0.8 mm, water. Heating surface and
fiber material: 1 – 11, 13, 15, 16 – copper; 12, 14, 17 – stainless steel, 18 – nickel from
Semena et al. [232, 240]

Fig. 3.49. Experimental data for boiling under the conditions of liquid capillary feed

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 117


Chapter 3

Equation (3.75) holds for the condition:

(3.77)

where ΔTcr is the temperature drop corresponding to the critical heat flux. Fig. 3.49 compares the
values of the heat transfer coefficient α calculated with (3.72) and (3.75) to an extensive set of
experimental data. Data scatter with respect to the theoretical analysis is mostly within ± 30%.

7⋅10 3
6 Nitrogen -1 -4
-2 -5
4 -3
α, W⋅m -2⋅K-1

10 3
8

4⋅102 2 4 6 8 2 4 6 8
1⋅10 -1 1⋅100 1⋅101
δ, mm

Fig. 3.50. Heat transfer coefficients for boiling of nitrogen on porous layers of the
following thicknesses: 1 – ε = 0.8; 2 – ε = 0.7; 3 – ε = 0.6; 4 – ε = 0.5; 5 – ε = 0.4.
4 –2
Data refer to q = 2.8 ∙ 10 W m from Levterov, Semena and Zaripov [141]

As was the case with other capillary-porous structures, these authors [141, 235, 266, 268, 269,
326] also investigated the impact of the coating thickness and porosity on heat transfer and the
critical heat flux. Their experimental results are presented in Fig. 3.50 and Fig. 3.51. It was found
that in order to maximize heat transfer on surfaces with a porous coating for nitrogen, obtaining
heat transfer coefficients on the order of ten times those for a smooth surface, it was necessary
to use layers whose thickness and porosity were in the range δ = 0.6 to 1 mm and ε = 0.4 to 0.6,
respectively, made of metals of the highest skeletal conductivity possible λskl, e.g. copper [141,
268]. Coatings having a porosity ε > 0.6 and thickness δ = 0.2 to 0.8 mm gave instead the highest
values of the critical heat flux. Also, Shapoval [266, 269] obtained similar results for water.

118 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Broadening the scope of investigations to cover fluids other than nitrogen, i.e. water and acetone,
led Zaripov et al. [326] to draw the conclusion that:

(3.78)

while holding the other structural parameters fixed. At the heat transfer is at its maximum value so
that , where for water and for nitrogen [141, 326].

5
7⋅10
6 -1 -4
-2 -5
4 -3
qcr , W⋅m-2

5
10
8
Nitrogen
6
5⋅10 4 2 4 6 8 2 4 6 8
10-1 10
0
101
δ, mm

Fig. 3.51. The critical heat flux characterizing nucleate boiling crisis as a function of
porosity of the porous coating from Levterov, Semena and Zaripov [141]
(notation as in Fig. 3.50)

Another important issue is the impact of pressure on the boiling heat transfer coefficient for
water on metallic fibrous layers, This has been investigated over the pressure range of 0.1 to
5 –2
1 ∙ 10 N m in [235, 269]. A complex and somewhat ambiguous influence of this parameter on α
was accounted for in the following experimental correlation:

n m l (3.79)
α=Cq δ p

where

–0.12 –0.25
n = 2.8 δ p
4 –0.17
C = 1.17 ∙ 10 m = 6.8 p l = 0.28 0.1 < δ < 1 mm
4 –0.26
C = 1.58 ∙ 10 m = 7.6 p l = 0.7 1 < δ < 10 mm

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 119


Chapter 3

Further experimental investigations conducted by Semena, Shapoval and coworkers [233, 235,
241, 266, 267, 269, 326] resulted in experimental correlations that took into account the liquid
thermal properties, porous coating structural parameters and test conditions of the heat transfer
surface.

Semena and Kiselev [233] proposed the following correlation applicable to gravitational
(thermosyphon) heat pipes with wicks made of metal fibrous porous coverings:

0.35 0.17 0.7


Nu = 1.875 K (dfbr/k) (dp/δ) (3.80)

where

which holds for R-11, R-113, water and ethanol; the range of validity of the dimensionless groups in
(3.80) was not given [233].

Shapoval [266] and also Semena et al. [241] analyzed heat transfer issues in water, ethanol and
acetone for pool boiling, when the porous layer was capillary fed with liquid. They presented two
correlations in successive papers for capillary feed conditions:

(3.81)

where

C = 200 m = 0.65 0.4 mm ≤ δ ≤ 1.2 mm


C = 0.5 m = –0.2 1.2 mm ≤ δ ≤ 9.0 mm

and

(3.82)

where

–2
C = 9.43 ∙ 10 m = 0.65 0.4 mm ≤ δ ≤ 1.2 mm
–2
C = 3.10 ∙ 10 m = –0.2 1.2 mm ≤ δ ≤ 9.0 mm

120 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Correlations (3.81) and (3.82) apply to boiling region I in Fig. 3.48, which is expressed by
the condition (3.73). Equation (3.81) applies the same range of the porous coating structural
parameters as for correlation (3.83).

Note to the reader: Most of the methods presented in this chapter are not dimensionless and
hence the units of the various properties are needed to implement the methods. This information
is not available in almost any of the quoted Soviet papers. The problem can be resolved by
“scientific guess” and/or “trial and error” approaches only. According to our experience, they
followed some unspoken rules: generally they applied SI units with some rare exceptions, like
porous layer thickness and pore diameter in mm, pressure in bars, etc. Usually, if they applied a
non-SI unit, it was somehow stated or suggested in the paper. Thus, we are not able to say with
certainty what the units actually were in these methods.

A slightly different approach was proposed for pool boiling conditions in [266, 269, 326].
The following expression was arrived at for water and acetone in the pressure range 0.1 to
5 –2
1 ∙ 10 N m :

(3.83)

where

4 –0.14
C = 2 ∙ 10 n = 0.15 δ 0.1 mm ≤ δ < 0.8 mm
–0.28
n = 0.0535 δ 0.8 mm ≤ δ < 10 mm

This correlation is applicable for the following ranges:

–1 –1
0.40 < ε ≤ 0.93 19 μm < Dp,ef ≤ 230 μm 0.5 < λskl < 69 W m K

The correlation (3.83) describes the experimental data well in the zone of fully developed nucleate
boiling, i.e. for heat transfer coefficients satisfying the condition α ≤ αmax. The mean relative error
of correlation (3.83) with respect to 95% of experimental data is ± 28 %.

Still another correlation by these authors in [326] covering pool boiling of four fluids (water,
nitrogen, acetone and ethanol) has the form:

n m k 0.4 0.5
α = C q δ λef B Am N (3.84)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 121


Chapter 3

where “cth” in the expression below means hyperbolic cotangent and

0.7 5 2.7
n=5δ cth (0.0035 + 104 δ + 7 ∙ 10 δ ) 0.1 mm ≤ δ ≤ 3 mm
n = 0.21432 3 mm ≤ δ ≤ 10 mm

The dependence of the exponent n in (3.84) on the coating thickness for four fluids is illustrated in
Fig. 3.52 [326]. The remaining quantities in equation (3.84) are as follows

(3.85)

(3.86)

the following heating surface materials:

(3.87)

In correlation (3.84) and in complementary expressions (3.85) to (3.87), defining the validity
range of (3.84), Semena, Shapoval and coworkers took into account the results of their previous
experimental investigations [141, 232, 241, 268, 269].

A total of 79 samples, all 30 mm in diameter, were investigated. The thickness of the disc on
which the capillary-porous structures were overlaid was 3.5 mm. Discs were made of copper
and stainless steel and porous structures of copper-nickel and stainless steel fibers. Fiber
diameters dfbr ranged from 20 to 70 µm, whereas their lengths were from 3 and 9 mm. The porous
coating parameters were as follows: porosity ε = 0.4 to 0.93; layer thickness δ = 0.1 to 10 mm;
–12 –9 2
effective pore diameters Dp,ef = 20 to 300 µm; and liquid permeability K1 = 10 to 10 m . The
investigations were conducted at atmospheric and subatmospheric pressures and heat fluxes
4 6 –2
ranged from q = 1 ∙ 10 to 2 ∙ 10 W m . According to the authors’ assessment, the correlation
(3.84) predicts values of the heat transfer coefficient α within an error of ±35% for 95% of their
experimental data.

122 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

8
-1
6
-2
-3
-4
4

1
10-1 2 4 6 8100 2 4 6 8 101
δ, mm

Fig. 3.52. Effect of metal fibrous porous coating thickness on exponent n: 1 – water;
2 – nitrogen; 3 – acetone; 4 – ethanol. Full lines were calculated with (3.84),
according to Zaripov et al. [326]

Following the correlation proposed by Rannenberg and Bier [222] for mesh structures, Danilova
and Tichonov [74] made an extensive investigation for R-113. They investigated metal fibrous
steel structures of fibers that were Lfbr = 3 mm in length, whose diameter was dfbr = 30 µm, whose
coating thicknesses δ were 0.2 and 0.7 mm whereas their porosities were ε = 0.5, 0.6 and 0.9.
Their correlation for the Nusselt number is:

(3.88)

where

Equation (3.88) predicts their experimental data with an error of ± 20 % [74]. At the same time, the
authors declared that the highest values of the heat transfer coefficient α were obtained for the
coating thickness δ = 0.2 mm and with the porosity in the range ε = 0.5 to 0.6.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 123


Chapter 3

Investigations by Poniewski and coworkers

General description of test parameters

Investigations of the application of porous coatings to boiling heat transfer enhancement


were started in Poland in the early 1990s by Cieslinski and others at the Gdansk University of
Technology [51 – 60] and by Poniewski and others at the Kielce University of Technology [4, 6,
187 – 189, 191 – 195, 202 – 208, 210 – 219, 252, 261, 317]. Because of the advantages
demonstrated by metal fibrous capillary-porous structures described earlier in this chapter,
Poniewski’s team at Kielce focused on investigations into the structural properties of such
structures, especially those leading to the occurrence of hysteresis under the conditions of
increasing and then decreasing heat flux. The hysteresis effect, due to the ambiguity of porous
structures’ characteristics on the boiling curve q = q(ΔTsl), considerably complicates the
application of these coatings.

The majority of investigations conducted by Semena et al. [141, 178, 185, 232 – 241, 266 – 269,
325, 326] dealt with the capillary feeding of liquid to the porous layer, which corresponds to the
operating conditions of a wick in a heat pipe. In order to better understand the enhancing role of a
porous coating on boiling heat transfer, it is justifiable to also conduct experimental investigations
under pool boiling conditions.

In their experiments, Poniewski’s team used three liquids: distilled water, ethanol and R-113. The
parameters of the coatings under investigation were as follows:

• material: copper fiber, dfbr = 50 µm, Lfbr = 3 mm

• diameter of the coated heating surface = 30 mm

• layer thicknesses δ, mm: 0.20, 0.35, 0.60, 0.80, 1.20, 1.50, 2.0

• porosities ε: 0.4, 0.7, 0.85

• effective diameters Dp,ef, mm: 0.20, 0.35, 0.60


–1 –1
• skeletal conductivities λskl, W m K : 2.0, 8.7, 68.0

Fig. 3.53 presents an enlarged photograph of porous metal fibrous surface made with this copper
fiber. Measurements were taken at two pressures, 0.03 and 0.1 MPa for water, while heat fluxes
4 6 –2
ranged from 2 ∙ 10 to 1 ∙ 10 W m .

The concept of a superheated liquid layer [140] advocated by Semena et al. [232, 240] to predict
the heat transfer coefficient, does not however offer a model that is specific to boiling heat transfer
in a capillary-porous structure. It was assumed that Smirnov’s model of substitute fins would
better serve the purpose [247, 256] since it could be adopted to account for fins of finite height
(the coating thickness in the present case), with heat transfer proceeding through the fin tip (the

124 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

external surface of the porous layer in this case) as well as the stochastic distribution of fins on the
heating surface (distribution of micro-channels within the layer where a channel is a void while a
fin is part of the structure).

Fig. 3.53. Photograph of metal fibrous capillary-porous structure: δ = 0.2 mm; ε = 0.7;
enlargement – 10 x of Poniewski and coworkers

The experimental test stand

The design of the experimental heat transfer test stand of Poniewski and others resulted in part
from a thorough analysis of previous test facilities described in the literature [187, 202, 261, 317].
A detailed description of the resulting test facility that they constructed was given in [187, 202, 261,
317]. Figures 3.54 and 3.55 present a diagram of the individual systems and a schematic of the
entire set-up. The two photographs in Fig. 3.56 show the apparatus arranged in the laboratory.
The design worked out by the team importantly includes two methods of determining the test
section heat flux – the principal one based on the temperature gradient method and the second
one relying on monitoring the supplied electric power (adjusting the stand to measurements with
both methods offers the possibility of checking the correctness of calculations).

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 125


Chapter 3

2
Heating system
with electric power
measurement

5 3
1
Fluid supply Temperature
Basic
and condensate measurement
module
recovery

7 4 6
Signal and power cutoff Subatmospheric Data digital
when exceeding the pressure acquisition
maximum permissible generation and and processing
temperature measurement system

Fig. 3.54. Measurement stand modules of Poniewski and coworkers

1 – Basic module; 2 – Heat losses compensation; 3 – Heating system with electric power
measurement; 4 – Measurement of pressure in a glass container; 5 – Subatmospheric
pressure generation and measurement; 6 – Vapor cooling and condensate recovery;
7 – Signal and power cutoff when exceeding the maximum permissible temperature;
8 – Temperature measurement; 9 – Data digital acquisition and processing system

Fig. 3.55. Diagram of the measurement stand with its associated systems of
Poniewski and coworkers

126 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

1 – Basic module; 2 – Condenser; 3 – Autotransformer; 4 – Wattmeter; 5 – Millivoltmeter;


6 – Thermocouple switch; 7 – System equalizing tank; 8 – Zero ice; 9 – Temperature
acquisition system

Fig. 3.56. Test stand for boiling heat transfer on surfaces covered with capillary-
porous structures of Poniewski and coworkers

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 127


Chapter 3

1 – Heated copper rod; 2 – Main heating coil; 3 – Additional heating coil compensating for
thermal losses; 4 – Ceramic cover; 5 – Teflon housing; 6 – Thermocouple for the liquid
saturation temperature measurement; 7 – Quartz glass vessel; 8 – Upper cover;
9 – Fixture elements (screws, nuts, spring washers); 10 and 11 – Vacuum apparatus;
12 – Pressure tap; 13 – Rubber gaskets; 14 – Test sample with capillary-porous structure
overlaid on it; 15 – Thermocouples; 16 – Insulation material; 17 – Ceramic holder;
18 – Metal base

Fig. 3.57. Boiling heat transfer test section design of Poniewski and coworkers

The schematic diagram of the test section is depicted in Fig. 3.57. The main heating coil (2) was
wound on the copper rod (1); the former is supplied with alternating current of up to 220 V voltage
through an autotransformer. The power measurement is taken with a wattmeter. The heater is
insulated from the environment with a ceramic cover (4) and a packing of insulation material (16).
An additional protection against heat losses into the environment is provided by a compensation
heater (3) on the external surface of the insulation layer. The compensating heater is controlled
by the signal of the temperature difference between the external and internal pipe walls.

128 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Thermocouples (15) are placed under the sample (14) in the copper heating rod to measure the
temperature profile (1) and also in the boiling liquid. The liquid is contained by the thermo-resistant
glass vessel (7), where boiling process observations can be made. Using an external condenser,
the condensed vapor is returned back to the vessel (7). The boiling vessel (7) is capable of
holding vacuum and operating at or below ambient pressure.

Heat fluxes transferred in the boiling process from the test surface to the liquid are substantial
6 –2
and can exceed 10 W m . Generating such large heat fluxes created serious technological
difficulties and the solution was as follows. The copper heating module is shown in more detail in
Fig. 3.58. It had a cylindrical groove, 6.5 mm wide and 8.7 mm in pitch, cut in the rod in a helix so
that contact between the resistance wire and the rod would be enhanced. A Kanthal resistance
wire of 0.7 mm in diameter was used with a total resistance of about 37 ohm, which made it
possible to attain a heating power of about 1300 W. Prior to winding the wire onto the rod, the rod
was sprayed with a thin layer of a ceramic insulating material capable of resisting temperature
of 1600 K. A thermo-resistant cement adhesive was then applied to seal the wire to the rod and
cured in vacuum oven.

Heat transfer investigations of porous and smooth surfaces were conducted in the range of
nucleate pool boiling for distilled water, ethyl alcohol and R-113. The liquid level above the sample
surface always exceeded 50 mm.

Before an experiment, the porous layer to be tested was rinsed with acetone, washed with ethanol
and dried and then soldered together with the thermocouples to the heating rod. The heating
module was then placed in a vacuum drier at 550 K and a pressure of 0.02 MPa. The samples
were then left in the test liquid for at least 24 hours before measurements. Prior to tests, intensive
5 –2
nucleate boiling (q ~10 W m ) for a half an hour removed any dissolved gases from the liquid.
Then the power of the heating coil was reduced until nucleate boiling faded away. The heat flux
–2 –1
was then increased from q = 0 to q = qmax at the rate of dq / dτ up to approximately 230 W m s ,
maintaining constant power at points set in the experiment plan.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 129


Chapter 3

1 – copper rod; 2 – sprayed ceramic layer; 3 – heating coil; 4 –ceramic filling

Fig. 3.58. Details of the copper rod and heating coil of Poniewski and coworkers

The heat flux was determined on the basis of measurements of the electric power supplied
to the main heater while taking into account heat losses to the environment and also with the
temperature gradient method. In order to calculate the heat flux supplied to the sample, the
following expression was used:

(3.89)

where C is the correction coefficient of heat losses to the environment (C = 0.918); Qel is the
electric power read from the wattmeter; and F is the heat transfer area covered with the porous
2
structure (0.000707 m ). The loss coefficient C was determined experimentally with a specially
prepared calorimeter shown in Fig. 3.59.

130 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

1 – calorimeter; 2 – thermocouples; 3 – ice; 4 – channel switch; 5 – millivoltmeter;


6 – three-way valve; 7 – water-pipe network; 8 – thermostat; 9 – measurement vessel;
10 – stop watch

Fig. 3.59. Test stand for heat loss measurements of Poniewski and coworkers

Fig 3.60 shows that heat losses to the environment were about 8% and remained unchanged in
the whole measurement range, hence the constant C could be introduced in (3.90). In order to take
heat flux measurements with the temperature gradient method, two thermocouples were installed in
the heating rod at its central axis at the distances of 6.5 mm and 11.5 mm from the heating surface,
as shown in Fig. 3.61. Fig. 3.62 shows temperature values measured in the rod under the heat
transfer surface in a few measurements. The slope of the line yields the heat flux whereas the
point of intersection with the axis of ordinates determines the temperature of the heating surface
under the porous layer.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 131


Chapter 3

1.2

1.0

Heat losses 8%
qcal ⋅10 -6, W⋅m-2
0.8

0.6

0.4
q
0.2 qcal

0
0 0.2 0.4 0.6 0.8 1.0 1.2
q⋅10-6, W⋅m-2

Fig. 3.60. Energy balance results on the test section of Poniewski and coworkers

Fig. 3.61. Temperature measurements in the copper heating rod of Poniewski and
coworkers

132 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Tpg T1 T2 T3
405

400 1
δ2
395
δ1
390
T, K

2
385

380 3 4 5

375

370
0 2 4 6 8 11 12
Distance between thermocouples

5 –2 5 –2 5 –2
1 – q1 = 2.409 ∙ 10 W m ; 2 – q2 = 1.704 ∙ 10 W m ; 3 – q3 = 1.327 ∙ 10 W m ;
5 –2 4 –2
4 – q4 = 1.1096 ∙ 10 W m ; 5 – q5 = 5.85 ∙ 10 W m

Fig. 3.62. Temperature distribution in the copper heating rod for boiling water at
atmospheric pressure of Poniewski and coworkers, ε = 0.4 and δ = 0.2 mm

The tests involved the determination of the boiling curve for a reference (technically smooth)
surface and comparison with literature data. In the heat transfer investigations, it is assumed that
the roughness of the technically smooth surface is ≤ 5 microns. Pool boiling test results for water
are shown in Fig. 3.63 together with the typical range of similar data taken from the literature. The
reference pool boiling curve was determined with the well-known Rohsenow correlation [24]:

(3.90)

where n = 1 for water and n = 1.7 for other liquids, Csl = a constant dependent on the surface
material–liquid combination (e.g. copper – water, Csl = 0.013).

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 133


Chapter 3

6
10
Literature based
experimental data
scatter

-2
q, W⋅m

105

according to
Rohsenow
initial results
final result
104
1 10 100
DT sl, K

Fig. 3.63. Heat flux plotted versus temperature difference for horizontal reference
(technically smooth) copper surface for pool boiling of water of Poniewski
and coworkers

Measurement of pore diameter distribution

Enlarged photographs (20-times magnification) of the upper surface of capillary-porous structure


(example – Fig. 3.53) were taken and covered with a square grid whose quadrants were bigger
than the largest pore dimension. The grid nodes were numbered and checked whether they
coincided with a pore location. If a pore was located at a node, its largest and smallest dimensions
were measured. The pore’s equivalent diameter was calculated as a geometric mean of measured
dimensions:

(3.91)

where sk is the factor of magnification. Samples of 50 measurements were collected from the
enlarged image of each structure analyzed, and they were grouped into 7 size categories. Results
of measurements and statistical calculations [33, 159] are presented in Chapter 4.3.4 [205 – 208,
211].

134 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Assessment of measurement accuracy

The assessment of the propagation of errors was made in accordance with the principles of
measurement accuracy analysis for experimental investigations presented in [275]. Details of
measurement error calculations are given in [187] and [317] and are summarized below:

• error in the C constant in the equation (3.89) was from 2.8% to 5%;

• error in the heat flux measurement by the electric power was about 5%;

• error in the heat flux measurement with the temperature gradient method was about
10 to 11%;

• the extrapolation error for the heating surface temperature at the contact point with the
capillary-porous layer was from 0.06% to 0.2%.

These error calculations were made for the small electric power Qel = 50 W, because relative
measurement errors become larger as the heat flux is lowered.

Boiling test procedures and results

The procedure of determining boiling curves, which accounts for hysteresis phenomena, is
presented in Fig. 3.64. At the beginning of the measurement session, the electric power to the
heating is systematically increased and when the pre-set maximum heat flux is reached – qz1
(see Fig. 3.64a), the power is decreased until reaching a value at which nucleate boiling vanished.
The process was then repeated increasing the heat flux to a higher value than the previous one
– qz2, which was again followed by a decrease (see Fig. 3.64b). The procedure continued until
the critical heat flux qmax was reached (see Figs. 3.64c and 3.64d), and then the heat flux was
systematically lower until boiling faded away.

For some samples, the critical heat flux was not reached because its value was higher than that
possible with the electric heater. Heat fluxes dissipated under pre-crisis conditions in metal fibrous
6 –2
capillary-porous structures were very high and they exceeded 10 W m . At the highest heat
fluxes tested, the heating rod’s hot end temperature was higher than 900 K, which sometime lead
to failure of the heater.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 135


Chapter 3

q q
a) b)
qz2
qz1 qz1

DTsl DT sl

q c) c)
qmax
qz2 qmax
qz2
qz1

DTsl DT sl

Fig. 3.64. Test procedure to determine boiling curves with hysteresis (a, b and c
curves refer to hysteresis occurring for q < qmax, d curve – to hysteresis that
occurs after qmax has been reached) of Poniewski and coworkers

A characteristic feature of the boiling process on surfaces with porous coatings is an intrinsic
increase in the surface temperature when the heat flux reaches qmax. Heating surface temperature
time traces were recorded by the data acquisition system. If after about 15 minutes after increasing
the imposed heat flux to a new value the temperature continued to rise, it meant that the critical heat
flux was reached. In this case, then the frequency of temperature sampling was increased. When the
surface temperature exceeded about 450 K, the heating coil power was reduced, as per Fig. 3.64d.

Applying the procedure shown in Fig. 3.64, three kinds of boiling curves were found:

• no hysteresis as in Fig. 3.65


st
• with hysteresis of the 1 kind, occurring for q ≤ qmax as in Fig. 3.66;
nd
• with hysteresis of the 2 kind, occurring for q > qmax as in Fig. 3.67.

Fig. 3.68 shows a composite diagram illustrating the impact of the boiling fluid and the porous
coating structural parameters on the occurrence of both kinds of hysteresis.

136 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

106 106
Water Water
0.1 MPa 0.1 MPa

q, W⋅m-2 1

q, W⋅m-2
5
10 105
2 2
1

ε = 0.7 ε = 0.85
δ = 0.2 mm δ = 0.35 mm
104 104
1 10 100 1 10 100
DTsl, K DTsl, K

106 106
Water Ethanol
0.1 MPa 0.1 MPa 1

1
q, W⋅m-2

q, W⋅m -2

105 105
2
2

ε = 0.4 ε = 0.85
δ = 0.6 mm δ = 1.5 mm
104 104
1 10 100 1 10 100
DTsl, K DTsl, K

Fig. 3.65. Examples of boiling curves without hysteresis (broken line – smooth
surface) of Poniewski and coworkers

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 137


Chapter 3

106 106
Water Water
0.03 MPa 0.1 MPa

q, W⋅m-2

-2
q, W⋅m
1
105 105
2
1 2

ε = 0.4 ε = 0.7
δ = 0.8 mm δ = 2 mm
10 4 10 4
1 10 100 1 10 100
DT sl, K DTsl, K

106 106
Water Water
0.1 MPa -2 0.1 MPa
q, W⋅m-2

1
q, W⋅m

2
105 105
2
1

ε = 0.85 ε = 0.85 Structure of carbon


δ = 2 mm δ = 2 mm fiber and copper
powder
104 104
1 10 100 1 10 100
DTsl, K DT sl, K

st
Fig. 3.66. Examples of boiling curves with 1 kind of hysteresis (broken line – smooth
surface) of Poniewski and coworkers

138 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

106 106
R 113
R 113
0.1 MPa
0.1 MPa

q, W⋅m-2
q, W⋅m-2 1
2
105 2
105

ε = 0.85 ε = 0.85
δ = 0.6 mm δ = 0.35 mm
10 4 2⋅104
1 10 100 5 10 50
DTsl, K DTsl, K
106 106
Ethanol Ethanol
0.1 MPa 0.1 MPa
q, W⋅m -2

q, W⋅m-2

1 2 1
105 105

ε = 0.85 ε = 0.4
δ = 0.35 mm δ = 0.6 mm
104 104
1 10 100 1 10 100
DTsl, K DTsl, K
nd
Fig. 3.67. Examples of boiling curves with 2 kind of hysteresis (broken line – smooth
surface) of Poniewski and coworkers

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 139


Chapter 3

Fig. 3.68. Impact of the porous covering structural parameters and boiling fluid
on hysteresis occurrence: I – I kind hysteresis; II – II kind hysteresis of
Poniewski and coworkers

Impact of the porous coating structural parameters


and pressure on the heat transfer coefficient

To investigation the impact of the porous coating structural parameters and pressure on the heat
transfer coefficient, the boiling experiments need to be done at test conditions that produce a
unique boiling curve so that the results can be compared in an unambiguous manner. Hence, in
this study the tests were conducted for the combinations of boiling liquid – porous coating whose
boiling curves had an unambiguous shape in the nucleate boiling zone. This means that the heat
transfer coefficient had the same value, independent of the direction of the heat flux during the
tests. Thus, the following cases were tested:

• hysteresis did not occur at all (curve 1 in Fig. 3.69);


st
• hysteresis of the 1 kind occurred, so for q < qmax the boiling curve had a unique
variation between the conduction and convection zone (A) and the boiling crisis zone
(B) (curve 2 in Fig. 3.69).

nd
This procedure eliminated the added complexity of the 2 kind of hysteresis effects on heat
transfer and porous layer performance.

140 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

st
1 - boiling curve without 1 kind hysteresis (curve 1 corresponds to boiling curves in Fig.
st
3.65); 2 - boiling curve with 1 kind hysteresis (curve 2 refers to boiling curves in Fig.3.66);
A and B - regions of ambiguous boiling curves; — — — — limiting lines of boiling curves
ambiguity regions

Fig. 3.69. Boiling curves ranges, included in the analysis of the influence of pressure
and porous coatings structural parameters on the heat transfer coefficient of
Poniewski and coworkers:

Poniewski and coworkers [203 – 208, 210 – 219, 317] via their experimental investigations
confirmed a significant influence of the porous coating structural parameters on boiling heat
transfer. Fig. 3.70a shows exemplary heat transfer data for water. A strong dependence of the
heat transfer coefficient on the layer thickness at constant porosity can be seen. The heat transfer
coefficient first increases and later decreases when the covering becomes thicker, which confirms
earlier results in [141]. The best results were obtained for the porosity ε = 0.85 in Fig. 3.70a for the
layer thickness δ = 0.8 mm. For such structural parameters, the heat transfer coefficient is over
5.5 times higher when compared with a technically smooth surface (dashed line in the diagram)
at the same heat flux. Here and in subsequent figures, the boiling curves for technically smooth
surfaces are calculated with the Rohsenow correlation (3.91) [24]. Similar heat transfer coefficient
results were also observed for porosities of 0.4 and 0.7 where the maximum values of the heat
transfer coefficient were recorded for thicknesses from 0.6 – 0.8 mm.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 141


Chapter 3

a) b)
105 105
Water ε = 0.85 Water δ = 0.8 mm
0.1 MPa 0.1 MPa
-1

-1
α, W⋅m ⋅K

α, W⋅m ⋅K
-2

-2
104 104
0.2 mm ε = 0.40
0.6 mm ε = 0.70
0.8 mm ε = 0.85
1.5 mm
2.0 mm
Smooth surface
Smooth surface
103 103 4
104 105 106 10 105 106
q, W⋅m -2 q, W⋅m-2

a) Tests with layers of various thicknesses at constant porosity


b) Tests at various porosities at constant porous layer thickness

Fig. 3.70. Pool boiling data for water of Poniewski and coworkers

a) b)
10 5 10 5
Ethanol R 113 0.20 mm
0.1 MPa 0.1 MPa 0.35 mm
ε = 0.4 ε = 0.85 0.60 mm
0.80 mm
-1

α, W⋅m-2⋅K -1
α, W⋅m ⋅K
-2

10 4 10 4
Boiling crisis
0.2 mm
0.6 mm
2.0 mm
Smooth surface
10 3 4 10 3 4
10 10 5 10 6 10 10 5 10 6
-2 -2
q, W⋅m q, W⋅m

Fig. 3.71. Heat transfer coefficient data for layers of constant porosity and changeable
thickness of Poniewski and coworkers

Additional tests were also run holding the layer thickness fixed and varying the porosity of
the layer. Some of these results are shown in Fig. 3.70b. Here it can be noted that for a fixed
thickness, the heat transfer coefficient increased as porosity got lower. Furthermore, the heat
transfer coefficient tends to a fixed value for all three porosities as the heat flux is increased
5 –2
towards a value of q ≈ 5 ∙ 10 W m , which is also observable in Fig. 3.70b.

142 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Similar trends in the heat transfer coefficient with layer thickness were also observed for ethanol
and R-113 as shown in Figs. 3.71a and 3.71b. It was revealed that the boiling crisis for R-113 takes
place at a much lower heat flux when the layer thickness is equal to 0.8 mm as shown in
Fig. 3.71b.

a) b)
105 Water, p = 0.1MPa 105
Water, p = 0.1MPa
8 q = 7·104 W·m-2 8 q = 6·105 W·m-2
6 6
ε = 0.40

α, W⋅ m -2 ⋅ K -1
ε = 0.70
α, W⋅m-2⋅K-1

ε = 0.85
4 ε = 0.85 4
ε = 0.40

2 2

104 104
0 1 2 0 1 2
δ, mm δ, mm

Fig. 3.72. Impact of the layer thickness and porosity on heat transfer coefficients of
Poniewski and coworkers

Fig. 3.72 shows combined influence of porosity and layer thickness on the heat transfer coefficient
at selected heat fluxes. A fall off in the heat transfer coefficient value is seen as the layer thickness
gets bigger and the heat flux increases. It is connected with the higher resistance to vapor flow out
through the porous layer. There is also an influence of porosity on the location of the maximum
heat transfer performance.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 143


Chapter 3

6
10
Water Boiling
ε = 0.7 crisis
δ = 2 mm

q, W⋅m-2
105

1
2
3

104
1 10 100
DT sl, K

1 – p = 0.1 MPa; 2 – p = 0.03 MPa; 3 – smooth surface, p = 0.1 MPa

Fig. 3.73. Influence of subatmospheric pressures on the boiling curve of Poniewski


and coworkers

Boiling curves for water were determined for two pressures that are typical of low-temperature
heat pipes, 0.03 and 0.1 MPa, and the results are shown in Fig. 3.73, where the sample is for
a layer δ = 2 mm in thickness and porosity ε = 0.7. Overall, the analysis of experimental results
obtained for thirteen samples indicated that lowering the pressure into vacuum had no significant
influence on heat transfer. While reducing the pressure did not affect heat transfer coefficient
values for the whole of the boiling curve, nevertheless, the critical heat flux was lowered. Similarly,
for ethanol there was no effect of vacuum pressures on the heat transfer coefficient values when
the pressure was lowered to 0.03 MPa.

For the development of a heat transfer correlation to describe the results, Poniewski and
coworkers [203 – 208, 210 – 219, 317] relied on the Nishikawa work [175] as their starting point.
Similarity numbers were defined as follows:

(3.92)

The following modification was introduced into Nishikawa equation with respect to the definition of
similarity numbers: the effective pore diameter Dp,ef was substituted for the diameter of a particle
or a wire, which was thought to better characterize the vapor flow velocity in the Reynolds number
than a constant wire dimension. The conductivity λskl and the diameter Dp,ef were measured
experimentally, whereas λef was calculated from [175]:

144 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

λef = ε λ1 + (1 – ε) λskl (3.93)

The new correlation took on the form:

(3.94)

where A = an empirical constant, Ks = the similarity number describing the impact of the surface
tension, Kp = the similarity number describing the geometrical properties of the porous coating,
Re = Reynolds number, K λ describes the ratio of the thermal conductivity of the porous layer
soaked in liquid to the thermal conductivity of the liquid, whereas Kc describes the thermal
properties of the boiling fluid.

In order to estimate the empirical parameters in (3.94) 212 experimental results were collected for
three liquids (water, ethanol, R-113), two pressure values (0.1 and 0.03 MPa) and twelve samples
of different porosity and layer thickness. The least square method was used and the constants in
(3.94) are:

A = 457.145 k = –0.389 l = 0.322 (3.95)


m = –0.49 (3.95) n = –1.273 0 = –0.511

The correlation is compared with experimental data in Fig. 3.74. The correlation (3.94) is
applicable to porous coatings 0.2 to 2 mm in thickness for water and 0.6 to 2 mm for ethanol and
R-113, and for porosities of 0.4, 0.7 and 0.85. It covers the following range of similarity numbers:
0.2 ≤ Re ≤ 13; 0.02 ≤ Ks ≤ 55; 0.01 ≤ Kp ≤ 90; 2 ≤ K λ ≤ 620; 200 ≤ Kc ≤ 1600.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 145


Chapter 3

100

10

+50%
Nueks

1 -50%

0.1

0.05
0.05 0.1 1 10 100
-0.389 0.322 -0.490
457.145 Ks Kp Re Kλ-1.273 Kc -0.511

Fig. 3.74. Comparison of correlation (3.95) for nucleate boiling on surfaces with metal
fibrous porous layers by Poniewski and coworkers

It is characteristic of metal fibrous capillary-porous structures to have high heat transfer


coefficients in nucleate boiling over the whole range of heat fluxes. The heat transfer coefficient
strongly depends on the structural parameters of the porous layer such as its thickness δ and
porosity ε. For a set porosity and heat flux, it is possible to experimentally determine the optimum
layer thickness, for which the highest value of the heat transfer coefficient α can be obtained. The
above correlation can be used for this purpose too.

st nd
Certain irregularities taking the form of 1 and 2 kinds of hysteresis occur in the nucleate boiling
on the heating surface with metal fibrous coating. The phenomena manifest themselves far more
strongly than in boiling on a technically smooth surface or on porous coatings manufactured
st
with other technologies. Hysteresis of the 1 kind seems to occur for liquids of weak wettability
nd
(water) and thick layers (0.8 to 2 mm). Hysteresis of the 2 kind seems to occur for liquids of good
wettability (that is, those with a small contact angle such as ethanol and R-113) for thin layers
(0.2 to 0.6 mm).

The manufacture of fibrous capillary-porous structures under laboratory conditions is neither


complicated nor costly. Attempts at their manufacture at the Kielce University laboratories have
proved successful. Those factors together with the already mentioned thermal performance
advantages of such coatings explains the continued interest in this type of porous structures for
enhanced boiling.

146 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

3.2.6. Surfaces developed with other technologies

The results of the experimental investigations and heat transfer correlations presented in previous
sections of this chapter refer mostly to micro-structured surfaces (capillary-porous structures).
Apart from the types of porous layers discussed so far, there are other means to make porous
layers, such as surface painting with mixtures of metal particles and epoxy resins [40, 42 – 44,
106, 179, 180], electronic (galvanic) deposition of a porous layer precipitated out of metal salts [51,
58, 76, 77, 154, 172], covering with microfins, grooves and pitting [10, 80, 160, 169], creasing the
heating surface covering mesh [106] and others [324].

An interesting new technology of the porous structure manufacture was worked out by Wlosinski
and co-researchers at the Warsaw University of Technology [202]. The porous layer is made of a
carbon fiber-copper powder composite. In the sample investigated, the carbon fibers were
dfbr = 7 µm in diameter and Lfbr = 1 to 2 cm in length whereas the copper powder had particle
diameters dgrn = 40 to 120 µm. They were fabricated by the Synthetic Fiber Institute in Lodz. The
composite is produced by diffusion volumetric welding at a temperature of 1125 K. Hence, this
composite enhancement is made from a combination of powder and fibers of different materials
and sizes. The results of investigations into the composite structure compared with the boiling
curves obtained for metal fibrous coverings are presented in the lower right graph in Fig. 3.66.
Heat transfer enhancement is similar to the effects observable for metal fibrous structures but with
higher values, for example in comparison to the results in the lower left graph in Fig. 3.66. If two
structures of the same porosity and thickness are compared experimentally, a higher heat transfer
coefficient in the nucleate boiling range and a higher critical heat flux can be obtained for the
composite structure [202, 211].

Electrolytic porous coatings are manufactured with a dispersion method, in which a dispersed
material is added to the standard galvanic bath. The dispersed material’s particles are built into a
porous layer by means of mechanical action in the electrolysis process [51, 58]. Capillary-porous
coatings made with this method significantly enhance boiling heat transfer as shown earlier in
Fig. 3.18, according to [51, 58, 76, 77]. The mechanical life of electrolytic coatings is however very
short (they can be removed by hand); therefore, they are not applicable to use in industry [58,
196].

Microsurfaces with cavities obtained mechanically or chemically also enhance boiling heat
transfer, as was proven earlier in Figs. 3.8 to 3.10, according to [10, 98, 160, 169]. However, as
their manufacture is costly and time-consuming, they have not been applied on an industrial scale.
On the other hand, they are useful for modeling the influence of micro-surface geometry on the
nucleate boiling process and intensity.

In the mid-1990s, a new technique for depositing of micro-porous coatings on a heating surface,
namely spraying, was developed. Mixtures of diamond or metal powder, epoxy resin and isopropyl
alcohol (or with epoxy-ceramic resins and ethyl-methyl ketone) are sprayed on the heating surface

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 147


Chapter 3

and then the surface is cured in an oven [40, 42 – 44, 179, 180]. The resin functions as a binding
material. Powder particle diameters range from 1 to 70 µm and micro-porous layers of 30 to 250
µm thickness are made.

A new parameter, called a dimensionless effective volume, was introduced for describing porous
layers of this kind. It was defined as follows:

(3.96)

(3.97)

where mpr and ρpr are the powdered material’s mass and density, ε is porosity of the powder in the
layer, whereas Vsp is the binder volume. The dimensionless effective volume gives the ratio of the
powder volume to the binder volume in the mixture. In the above-mentioned works it ranges from
about 1.1 to 6.2.

Chang and You [40, 42 – 44] determined boiling curves for a few combinations of powders and
binders. The investigations were conducted for the dielectric fluid FC-72 at atmospheric pressure.
The boiling curves obtained for such a microporous surface, a technically smooth surface and an
industrially available High Flux surface are shown in Fig. 3.75. As can be seen, the High Flux surface
outperforms the other enhancement over nearly the entire heat flux range but the two types of
surfaces have nearly the same critical heat flux.

Summarizing the work in [40, 42 – 44, 179, 180], it was found that with micro-porous coverings,
in comparison with a smooth surface, that it is possible to obtain lower boiling incipience
superheats as well as higher heat transfer coefficients and the critical heat fluxes. Micro-porous
coverings, deposited on the heating surface by brush painting or spraying have been proven to be
mechanically stable.

148 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

6
10
1 - High Flux
7
2 - Aluminum powder+ brushable ceramic+
5
+ methylethylketone
3 Nucleate
~ boiling crisis
Vef = 4.6
5
10
7
q, W⋅m-2 5
3 1 2

104
7
5 Smooth
3 surface

103 0
10 2 4 6 8 101 2 4 6 8 102
DTsl, K

Fig. 3.75. Comparison of boiling curves for a sprayed micro-porous surface, a High
Flux surface and a technically smooth surface from Chang and You [43]

3.2.7. Combined surfaces

Micro-structured surfaces manufactured industrially typically fall into two categories, those
made by finning and/or knurling the surface of a tube (such as the various types of the Turbo-B,
Gewa-B, Thermoexcel-E, etc.) or obtained by covering the surface with capillary-porous
structures deposited and sintered on the surface (such as the High Flux tube). In order to enhance
boiling heat transfer to even a greater extent, it is a natural extension to industrial practice to
consider depositing a porous layer on finned surfaces. Such a combination would increase the
base surface area (low finned tubes have up to 4 times the surface area of the original plain tube)
while the porous coating would still augment the heat transfer mechanisms, and thus perhaps
create an exceptionally good boiling heat transfer coefficient if the right combination can be found.
Such options are classified here as combined surfaces. Such enhanced boiling surfaces have
been tried over the years and this section describes some of their experimental results.

Xiuling et al. [319] investigated liquid nitrogen boiling on the external surfaces of horizontal and
vertical copper tubes with notched fins 1.5 mm in height and with triangular cross-section (thread)
that were covered with a porous layer of 0.5 mm thickness and 0.50 porosity of sintered powders.
The pool boiling data for this combined surface tube are shown in Fig. 3.76. A significant increase,
6.4 to 8.2 times, in the heat transfer coefficient was observed when compared with a smooth
tube (estimated by a plain tube correlation). In comparison with a tube coated with a porous layer
but without fins, heat transfer was augmented by a factor of 1.4 to 1.8, illustrating the potential of

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 149


Chapter 3

combined surfaces. A heat transfer correlation was put forward for both horizontal and vertical
tube orientations [319]. The correlations covered only one geometrical pattern of the fins, one
boiling liquid and a single coating of one thickness and porosity and hence are not able to be
applied to other conditions.

5⋅10 4
Nitrogen

104
-2
q, W⋅m

10 3

-1
-2

10 2
0.1 2 4 6 8 1.0 2.0
DTsl, K

1) Finned tube covered with porous layer of sintered powder;


2) Plain tube covered with same porous layer of sintered powder

Fig. 3.76. Boiling curves for liquid nitrogen Xiuling et al. [319]

Rosenfeld et al. [224] investigated boiling heat transfer on pin fins of copper and aluminum,
covered with a layer of sintered powder, making the internal surface for the evaporator zone of
a heat pipe used for electronic circuit cooling. Water, methanol, acetone and ammonia were the
6 –2 –1
test liquids. Heat transfer coefficients up to as high as 10 W m K were obtained! The authors
presented two models for predicting heat transfer on the combined structures they investigated.
In the first one, for small heat fluxes, heat transfer was assumed to proceed through the structure
skeleton by conduction with evaporation occurring on the boundary of the porous coating and the
heat pipe vapor space. In the other model for higher heat fluxes, boiling inside the porous layer
was thought to be decisive for heat transfer. The boiling curves obtained for the two models were
compared with scarce experimental data but no satisfactory congruence was noted.

Extensive experimental investigations on boiling heat transfer on finned surfaces covered with
metal fibrous capillary porous structures were conducted by Pastuszko and Poniewski [187–195].
The investigations focused on large rectangular copper fins on a flat base, whose fin heights hfin
were 5 mm and 10 mm and fin thicknesses δż were 1.5, 3.0 and 5 mm. The following porous layer
thicknesses δ were tested: 0.2 mm; 0.3 mm; 0.6 mm and 0.7 mm (with porosities from
0.47 ≤ ε ≤ 0.85). Water, ethanol and R-113 were the boiling liquids and the investigations were

150 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

conducted at atmospheric pressure. Significant heat transfer augmentation ratios were obtained,
which are listed in Table 3.6.

Table 3.6 Heat transfer enhancement on fins with metal fibrous porous coatings from
Pastuszko [187]

Augmentation Boiling liquids


5 –2
For q = 10 W m Ratio Water Ethanol R-113

Relative to flat smooth surface αfin+p / αsm 2...3.4 3...5.5 2.1...3.4

Relative to plain fins αfin+p / αfin,sm 1.3...2.2 2.5...4 1.5...2.5

These authors summed up their various results of experimental investigations [187, 188, 194, 195]
by proposing the following correlation for the Nusselt number:

0.47 –0.62 0.30 –1.10 1.14


Nu = 0.01 Re Pr Kp Kφ–0.35 We* K λ (3.99)

The similarity numbers and other characteristic quantities were defined as follows:




(3.100)

The name or meaning of most of these numbers is well-known or easily discerned. The lower right
number l is the bubble characteristic dimension. Other expressions are as follows: λef = ε λ1 + λskl
is the effective conductivity of the capillary-porous structure, λCu is thermal conductivity of the
– is the mean fin width and s is the interfin spacing. The following are the
fin material (copper), w
empirically determined parameters:

0.3 ≤ < Nu ≤ 83; 39 ≤ Re ≤ 5462; 1.74 ≤ Pr ≤ 6.65;


1.8 ≤ Kp ≤ 20; 6.3 ≤ Kφ ≤ 20.9; 0.01 ≤ We* ≤ 2095; (3.101)
9.4 ≤ K λ ≤ 184.4

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 151


Chapter 3

330 experimental data


two - dimensional model, n = 1.3
329 one - dimensional model, n = 1.0
one - dimensional model, insulated face
one - dimensional model, uninsulated face
328

327

326
R 113
325

324

323

322
αfin + ps = C (Tfin − Tsat ) n −1
321

320
0 2 4 6 8 10
Distance from the fin base, mm

Fig. 3.77. Comparison of computed temperature distributions along a selected vertical


line inside a fin with experimental data of Pastuszko [187] and Pastuszko and
Poniewski [190]

Good agreement with the experimental data was obtained, as shown in Fig. 3.77, for this two-
dimensional model of heat transfer in a two-layer fin, accounting for the variation of the heat
transfer coefficient along the fin height, according to [187, 190]. The correlation predicted
about 86% of the data within ±35%. The experimental data presented in Fig. 3.77 were collected
from thermographs taken with an infrared camera installed to make observations of temperature
distribution in the fin section [187, 190].

With only a limited amount of experimental data available for combined surfaces, it is not
possible to predict the viability of these surfaces for industrial application to boiling heat transfer
enhancement.

152 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

3.3. Conclusions

Experimental investigations into enhanced boiling heat transfer led to the creation of various
capillary-porous structures that differ in manufacture technologies, selected materials and
geometrical parameters. Manufactured in a mechanical way or deposited on heating surfaces,
they generally yield a significant increase in the heat transfer coefficient while also reducing the
superheat of the heating surface necessary to achieve boiling incipience.

Although capillary-porous structures are widely applied at present, there is no generally accepted
physical explanation of the mechanism controlling the vapor formation process inside these
layers. A vast amount of literature discussing the results of experimental investigations on boiling
heat transfer on micro-structured surfaces (capillary-porous structures) is available. The studies
discussed in the present chapter have dealt with both nucleate pool boiling conditions and boiling
under the conditions of capillary feed. In their investigations, Semena and Shapoval [178, 232,
240, 241, 266, 269, 326] clearly differentiated between these two liquid feed cases, putting
forward two separate correlations for the heat transfer coefficient depending on the type of feed.

Furthermore, studies of other authors [5, 62, 86, 226] indicate that the liquid level does not affect
the heat transfer coefficient, therefore it is possible to apply the results obtained for pool boiling to
the working conditions in heat pipes.

Covering the base of a heated surface with a capillary-porous structure generally enhances
boiling heat transfer process in comparison with a smooth surface, for example as was shown
in Fig. 3.78, curves 1 to 8 and 11. A certain set of structural parameters, however, can in fact
produce a decrease in the heat transfer coefficient , refer to curves 9 and 10 in Fig. 3.78. In
addition, overlaying a capillary-porous structure on the base of a heated surface also changes the
functional dependence between the heat transfer coefficient and the heat flux. The increase in α
as q grows is not as considerable as for a smooth surface (indicated by the smaller slopes of the
enhanced boiling curves in α vs. q). The heat transfer coefficient can also have a constant value
over a wide range of q or even decrease when the heat flux grows (see curve 3 in Fig. 3.78).

There is also substantial evidence that there is a strong link between heat transfer performance
and the structure’s dimensions and parameters. For instance, an increase in the stainless steel
powder layer thickness causes changes in both the heat transfer coefficient and the shape of the
relationship α = f(q), (see Fig. 3.78, curves 1, 2 and 3). When the layer thickness grows (curves 2
and 3) the heat transfer coefficient falls. Further growth of the layer thickness, for a small porosity
increase (curve 1), leads to an increase in the heat transfer coefficient and another change in the
functional relationship α = f(q), namely from a decreasing to an increasing effect. On the other
hand, the literature on this subject lacks an explanation of the phenomena that create this complex
relationship between α and q. The considerable differences in the shapes of boiling curves can be
attributed to the specific features of heat transfer in various kinds of capillary-porous structures.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 153


Chapter 3

Boiling curves shown in Fig. 3.78 are an illustration of a considerable impact of three structural
dimensions and parameters (porosity, thickness, porous layer thermal conductivity) on boiling heat
transfer.

4
3⋅10
1 2
2
11
3
5 4
4
10
8 6
α, W⋅m-2⋅K -1

7
8
6
9
4

10

2
Smooth surface

103 4 5
10 2 4 6 8 10
q, W⋅m-2

Material: mi – copper, mo – brass, n – nickel, sn – stainless steel; kind of structure:


s – mesh, mw – metal, fibrous, p – powder; parameters: w – number of layers,
δ – thickness, ε – porosity

1 – p, sn, ε = 0.67, δ = 1.1 mm; 2 – p, sn, ε = 0.5, δ = 0.15 mm; 3 – p, sn, ε = 0.5,
δ = 0.5 mm; 4 – mw, n, ε = 0.8, δ = 2.1 mm; 5 – s, sn, 4w, 100 mesh threads/inch;
6 – (a) s, mo + 5w (80 µm) + sn + 1w (60 µm) + 2 w (130 µm); (b) s, sn + 1w (40 µm) +
2w (130 µm) + m + 1w (50 µm) + sn + 2 w (130 µm); 7 – s, mi, 8w (100 µm); 8 – s, sn, 4w,
325 mesh threads /inch); 9 – s, sn, δ = 1 mm, 10 – s, sn, 2w (0.2 to 0.28 µm); 11 – mw,
mi, ε = 0.85, δ = 0.8 mm

Fig. 3.78. Compilation of boiling heat transfer results on surfaces coated with
capillary-porous structures from [210, 232]

154 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The present analysis of the numerous experimental studies detailed in this chapter reveals
common features shared by micro-structured surfaces, obtained by industrial technologies and
by overlaying the base heat transfer surface with capillary-porous structures. These can be
summarized as follows:

a. the number of possible and active nucleation sites is larger on such surfaces when
compared with a technically smooth surface, which generally leads to a considerable
increase in the heat transfer coefficient, but the heat transfer coefficient on these
surfaces tends to decrease with increasing heat flux:

b. the critical heat flux is most often equal to or higher than the corresponding values of
smooth surfaces;

c. on the majority of micro-structured surfaces, nucleate boiling incipience occurs at


significantly lower superheats than on smooth surfaces unless nucleation hysteresis
takes place;

d. an optimum porous layer thickness exists for the majority of types of capillary-porous
structures, for which the highest value of the heat transfer coefficient (that is the
optimum) depends on the selected heat flux and elevated values of the heat transfer
coefficient are more frequently obtained for smaller layer thicknesses, which is an
indirect proof that vapor escape plays a decisive role in the intensity of boiling heat
transfer through the hydraulic resistance to vapor outflow, which increases as the layer
gets thicker;

e. a characteristic feature of all micro-structured surfaces is the occurrence of three


st
kinds of hysteresis, namely: nucleation hysteresis, 1 kind of hysteresis following the
nd
nucleate boiling crisis and 2 kind of hysteresis typical of heat flux densities lower
than the critical one;

f. on the basis of published experimental results, it is difficult to unambiguously specify


the impact of porosity on the heat transfer coefficient and on the initial superheat
of the heating surface, yet it can be stated that for each kind of the capillary-porous
structure, an optimum porosity and thickness can be found, for which the highest
values of the heat transfer coefficient are obtained;

g. if micro-structures are manufactured with the same technique but differ in thickness
and porosity, the heat transfer coefficient increases as the skeletal thermal
conductivity increases;

h. changing the boiling saturation pressure affects the shape of the boiling curve and the
heat transfer coefficient only to a small extent, especially in the region of high heat
fluxes.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 155


Chapter 3

The studies discussed in this chapter also dealt with detailed matters like, for instance, the impact
of the structure cell dimensions and the distribution of pores of various dimensions along the
structure thickness on the resulting heat transfer coefficient. Experimental results presented in
[85] for mesh structures with cells 0.2 mm in diameter indicated that for such dimensions the heat
transfer coefficients corresponded to those obtained in pool boiling on a smooth surface and thus
small mesh structures can in fact deteriorate heat transfer.

The work by Tehver [148], Poniewski and co-researchers [216, 218] and Smirnov [256] confirmed
the impact of the density function of pore dimension distribution on the heat transfer hysteresis
nd
phenomenon. The 2 kind of hysteresis, which results in the heat transfer coefficient increasing
as the heat flux decreases, occurred for thin layered structures with a considerable non-
homogeneous distribution of pore sizes.

The value of the heat transfer coefficient as well as the shape of its variation with respect
to heat flux are connected with the specific type of the capillary-porous structure. Structural
parameters and the thermal properties of the structural material are of crucial significance. The
coating thickness, its porosity, the pore distribution and the skeletal thermal conductivity are
the most important parameters. These parameters, whose possible variations cover a wide
range, considerably influence the boiling heat transfer coefficient. The present comprehensive
analysis of these studies, however, does not yet an unambiguous explanation of the impact of
these parameters. The reason why such discrepancies are found lies in differences in the high
complexity of the underlying boiling heat transfer mechanisms in different kinds of capillary-porous
structures.

156 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Chapter 4
4. Modeling Nucleate Boiling on Microporous Surfaces

4.1. Visualization of boiling on microporous surfaces – capillary-porous


structures

Heat transfer enhancement in boiling on microporous surfaces with industrially modified (treated)
structures or those obtained by depositing a capillary porous structure cannot be attributed totally
to the development of the heating surface. The following benefits are given by the capillary-porous
structure:

a. A significant decrease in the heating surface superheating that is necessary for


boiling incipience;

b. A considerable increase in the heat transfer coefficient, especially in the range of


lower heat flux densities.

Up to now, no uniform concept has been put forward to explain what causes the increase in the
heat transfer coefficient on such microporous surfaces. The reason lies in the complexity of
physical processes and significant differences in experimental results, see Chapter 3.

Many authors like, e.g. Djundin et al. [76, 77], attribute heat transfer enhancement to the increase
in the number of nucleation centers and to the fact that the capillary porous structure hinders
vapor nuclei removal by the suction of the liquid departing in the wake of a bubble [311] that results
from vapor generation inside the porous layer [76, 77, 135], as it happens on a smooth surface.
The presence of the capillary porous structure considerably increases the ratio of the surface
giving up sensible heat to the liquid volume, increasing the liquid superheating in the surroundings
of the heating surface, which facilitates vapor bubble formation [155].

With the porous surface capillary feed mechanism, in the transition from evaporation to boiling,
the liquid meniscus was observed to move deeper inside the layer and thus multiply the
evaporation surface by many times. The liquid microlayer covering the walls of the pores, which
contacts the meniscus, enhances the evaporation process [300].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 157


Chapter 4

The conditions for vapor phase formation on a rough surface or that covered with porous layer
result from the change in thermodynamic potential [139, 301]:

(4.1)

where ψ are the respective chemical potentials of vapor and liquid, V is the vapor bubble volume,
F is the smooth surface area, Fps is the surface area of pores, and θ is the wetting angle. The
larger is the ratio Fps/F, the lower is the required thermodynamic potential threshold Δψ, so the
faster vapor bubbles grow. In accordance with [139, 301], the process stabilization and heat
transfer enhancement can be achieved by diminishing the surface area F and increasing the ratio
Fps/F, i.e. through a transition change from a smooth to a porous surface.

Investigations into boiling mechanisms in capillary-porous structures are difficult to conduct as


vapor formation takes place inside opaque thin layers. The data obtained through observation
of the external (usually upper) surface of the porous coating provides us with only some indirect
knowledge to make hypotheses about what is proceeding inside. The few experimental studies
on the phenomena occurring inside the capillary-porous structure do not provide unequivocal
answers regarding the means of vapor escape and replenishing with liquid, the sites of the
structure where evaporation occurs or where and how interfacial surfaces are located [9, 155].

158 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

1 – glass cylinder (internal diameter 75 mm), 2 – glass plate (δ = 3 mm, Pyrex);


3 – condenser; 4 and 5 – bar and plate; 6 – polyurethane sponge; 7 – silicon carbide
heaters; 8 – heater; 9 – mirror; 10 – camera

Fig. 4.1. Diagram of the system for the visualization of the boiling process on a
surface with a porous coating of Cornwell et al. [65]

Moss and Kelly [165] and Cornwell et al. [65] have tried to answer the questions posed above.
Moss and Kelly [165] applied neutronoscopy to a mesh capillary–porous structure that operated
as a heat pipe wick. They found a stable vapor film inside the porous layer and measured the film
thickness as a function of the heat flux density. Moreover, it was revealed that the frequency of
bubble formation inside the capillary-porous structure was much lower than in nucleate boiling on
a smooth surface.

Observing the interface of contact of the polyurethane foam and the transparent heating surface,
Cornwell et al. [65] confirmed the co-existence of the liquid and vapor at this surface for water,
within a wide range of the heat flux densities. Their diagram of their system for boiling process
visualization is shown in Fig. 4.1, in which a linear dependence of the heat flux density on the
ratio of the surface covered with vapor to the total surface area was found, as shown in Fig. 4.2.
In conclusion, they stated that the linear dependence of heat flux density on the ratio Fg /F proved
that the pressure drop across the porous layer thickness is constant and the vapor flux growth
results from the increase in the surface on which it is generated. Hence, the critical (maximum)
density of the heat flux corresponds to Fg /F = 1, Fig. 4.2, in accordance with [65].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 159


Chapter 4

11
Range of the critical
10 (maximum) heat flux. 10
9 Water, atmospheric pressure 9
8 8
7 7

-2
q⋅10 , W⋅m
6

-5
5
4
3
2
1 Experimental data
0
0 0.2 0.4 0.6 0.8 1.0
Fg / F

Fig. 4.2. Heat flux density dependence on the fraction of the surface covered with a
vapor film from Cornwell et al. [65]

Nakayama and co-researchers [167] visualized R-11 boiling in a rectangular capillary of a


Thermoexcel-E microsurface, at atmospheric pressure, using the setup shown in Fig. 4.3. In the
first series of experiments, the heat was supplied from the bottom; in the other, the cover with
holes acted as an electric heater and, at the same time, as a resistance thermometer. The boiling
process was recorded with a fast speed film camera or a photographic camera. Their observations
provided the following conclusions [167]:

1. At ambient temperature, a large R-11 vapor bubble was maintained in the capillary
without it being additionally heated. The bubble volume decreased with the lowering of
the environment temperature.

2. The vapor volume grew after heating started. The liquid film persisted on capillary
surfaces. The liquid was also found in capillary corners.

3. When the superheating of the capillary walls was higher than 0.6 K, vapor bubbles
generation started at the frequency of approx. 0.125 Hz.

4. Vapor bubble departure frequency and the number of active pores grew as the wall
superheating increased.

5. The above-mentioned phenomena were observed for all capillary dimensions, except
for the largest pores (0.5 mm). In that case, the liquid was never pushed out of the
capillary volume, covering the capillary internal surfaces and taking up approximately
10% to 50% of its volume.

160 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Fig. 4.3. Diagram of the stand for observations of boiling in capillaries with pores by
Nakayama and co-researchers [167]

Arshad and Thome [15] built a test stand to observe the boiling process in the capillary channels
similar to that of Nakayama et al. shown in Fig. 4.3, but with an axial view as opposed to a side
view, i.e. the observations were made at the face of the capillary that was closed with a glass
plate. Capillary channels of triangular, rectangular and cylindrical cross-sectional shapes were
milled in a copper plate and covered with a very thin copper plate (0.005 mm) with holes from 0.19
to 0.25 mm in diameter. Fig. 4.4 shows, in the form of a diagram, the process of bubble formation
and development inside the capillary. The experiments yielded the following findings [15]:

1. Evaporation of a thin film is a basic mechanism of vapor generation inside a capillary.

2. The capillary cross-sectional shape affects the thin liquid film shape and the way
it forms.

3. Evaporation of all the liquid inside capillary film makes the surface behave like a
smooth surface.

Fig. 4.4. Process of a bubble formation and growth in a capillary of triangular shape
according to Arshad and Thome [15]

Typically, simple models are built in order to explain complex processes taking place in boiling.
Correctly taking into account significant properties of the capillary-porous structure and those of
the liquid wetting it determines then the practicality of the model. So far a large number of models,

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 161


Chapter 4

both for pool boiling and that with liquid capillary feed, have been proposed. Relatively speaking, it
is easier to model industrially manufactured microporous surfaces as they have precisely defined
geometries and are repeatable. A few works presented a review of some models, which has
already been reported in the present monograph [58, 202, 261, 288, 311, 317].

4.2. Heat transfer models

4.2.1. Attempt to classify the heat transfer models

Heat transfer on industrially manufactured microporous surfaces and those covered with capillary-
porous structures is a complex thermodynamic process in which the phenomena of phase change,
heat transport and liquid and vapor motion are all interconnect. Due to incomplete knowledge of
the boiling process inside capillary-porous structures, the explanations of experimental trends in
their heat flux or heat transfer coefficients are either of the empirical character or are based on
similarity theory, as noted earlier in Chapter 3.

Presently, as already mentioned before, no uniform concept is available to explain the reasons
for heat transfer coefficient increase or the lowering of the heating surface superheating that
initiates the boiling process on industrial microporous surfaces or those with porous coatings. A
large number of theoretical models have been presented, which tried to account for the boiling
phenomena both in pool boiling or when the porous layer is capillary fed with liquid. Models
are classified according to various criteria, mainly on the basis of the capillary-porous structure
geometry or the mechanism of vapor generation inside the structure.

Three kinds of evaporation, characteristic of capillary-porous structures, can be differentiated:

1. evaporation of a thin liquid micro-layer covering surfaces of the capillaries;

2. evaporation from liquid – vapor menisci inside the structures;

3. evaporation on the external surface of the porous coating.

While grouping models in accordance with the structure’s geometry, to which theoretical
considerations referred or for which computation results were compared with experimental data, it
is possible to segregate the models as follows:

1. Gewa-T industrial microsurfaces of deformed fins:


–– Xin and Chao model, 1987 [318]
–– Ayub and Bergles model, 1987 [20]
–– Wang et al. model, 1991 [304]

162 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

2. Thermoexcel-E industrial microsurfaces of deformed fins:


–– Nakayama et al. model, 1980 [167, 168]
–– Xin and Chao model, 1987 [318]
–– Webb and Haider model, 1992 [314]

3. High Flux industrial microporous layered surfaces:


–– O’Neill et al. models, 1971 – 9 [68, 91, 182]

4. microfins covered with foil with holes (tunnel structures equivalent to the Gewa-T and
Thermoexcel-E geometries):
–– Chien and Webb model, 1998 [46]

5. mesh porous layers:


–– Moss and Kelly model, 1970 [165]
–– Smirnov et al. model, 1977 – 9 [5, 247, 250, 254]
–– Rannenberg and Beer model, 1980 [222]

6. porous layers of sintered powders:


–– Nishikawa et al. models, 1979 [177]
–– Tehver et al. model, 1978 – 83 [200, 285, 287]
–– Kovalev and Solovev model, 1985 – 90 [121, 124, 126 – 128]

7. electric arc sprayed porous layers:


–– Orlov and Savelev model, 1980 [183, 184]

8. layers of particles and other structures:


–– Ferrell and Alleavitch model, 1970 [86]
–– O’Neill et al. models, 1971 – 9 [68, 91, 182]
–– Man’kovski et al. model, 1976 [155]
–– Cornwell et al. model, 1976 [65]

Models put forward by various authors could also be arranged in accordance with the
assumptions made about the physics of boiling process inside capillary-porous structures. Each
division proposed in this respect must be arbitrary in character because individual models comprise
characteristics pertaining also to other models.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 163


Chapter 4

The shared features of the models, which could be differentiated in the literature on the subject,
provide the basis for the following classification based on heat transfer mechanisms:

1. heat conduction through a liquid microlayer covering the surfaces of microfins,


capillaries and internal tunnels and evaporation on the liquid microlayer surface with
the resulting vapor carried away to a pool of liquid:
–– Ferrell and Alleavitch model, 1970 [86]
–– O’Neill et al. models, 1971 – 9 [68, 91, 182]
–– Man’kovski et al. model, 1976 [155]
–– Smirnov et al. model, 1977 – 9 [5, 247, 250, 254]
–– Tehver et al. model, 1978 – 83 [200, 285, 287]
–– Orlov and Savelev model, 1980 [183, 184]
–– Xin and Chao model, 1987 [318]
–– Ayub and Bergles model, 1987 [20]
–– Kravchenko and Ostrovski model, 1990 [133]
–– Wang et al. model, 1991 [304]

2. evaporation from liquid-vapor menisci (capillaries) inside the capillary-porous


structure:
–– Kovalev and Solovev model, 1985 – 90 [121, 124, 126 – 128]
–– Webb and Haider model, 1992 [314]
–– Chien and Webb model, 1998 [46]

3. evaporation inside tunnels (capillaries) of the capillary-porous structure and


convection on the external surface of the structure:
–– Nakayama et al. model, 1980 [167, 168]
–– Ayub and Bergles model, 1987 [20]
–– Webb and Haider model, 1992 [314]
–– Chien and Webb model, 1998 [46]

4. porous layer substitution with equivalent microfins; conduction through minifins and
the liquid micro-layer covering them:
–– Man’kovski et al. model, 1976 [155]
–– Smirnov et al. model, 1977 – 9 [5, 247, 250, 254]
–– Orlov and Savelev model, 1980 [183, 184]

164 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

5. assumption that the sum of pressure drops in the liquid layer and vapor layer cannot
exceed the increment of capillary pressure
–– Moss and Kelly model, 1970 [165]
–– Smirnov et al. model, 1977 – 9 [5, 247, 250, 254]

6. other models
–– Cornwell et al. model, 1976 [65]
–– Nishikawa et al. models, 1979 [177]
–– Rannenberg and Beer model, 1980 [222]

It can be seen that models of the same authors can belong to two or more different categories.
That means that individual authors building their own models adopted similar simplifying
assumptions arranging them in different ways, according to their interpretation of the boiling heat
transfer process inside capillary-porous structures. Thus, it is possible to distinguish yet other
model groupings than those presented above. Therefore, it seems most reasonable to discuss the
published theoretical models chronologically, showing how the knowledge and insights on the boiling
process in capillary-porous structures has evolved.

4.2.2. Heat transfer models – theoretical basis, experimental verification

Ferrell and Alleavitch model

Ferrell and Alleavitch [86] built a porous deposit of Monel particles placed on a flat steel heating
surface. The deposit they obtained had a uniform porosity of ε = 0.4 but changeable number of
pores per unit area, i.e. from 20 pores per square inch (20 mesh) to 200 pores (200 mesh) per
–12 2 –12 2
square inch. Permeability ranged from 10.96 ∙ 10 m to 411.56 ∙ 10 m . The deposit height
also varied from 3.2 mm to 38.1 mm. The deposit was flooded by a water column of constant
height 76.2 mm.

Their theoretical analysis of boiling on this porous deposit was based on two quantities, measured
independently, characterizing a porous layer: (i) permeability, which is proportional to the
hydraulic pressure of the deposit and (ii) the height of the capillary draw, which is a function of
surface tension forces. Good congruence of experimental data with computation results for both
above-mentioned quantities suggested it was possible to make an assumption that the particles
formed a regular cubic configuration in the porous deposit. Based on observations of the boiling
process in the deposit, Ferrell and Alleavitch [86] assumed that the conduction through a thin
layer of saturated liquid, being in contact with the heating surface, was a decisive heat transfer
mechanism. They solved a three-dimensional heat conduction equation numerically, taking into
account the following assumptions and phenomena:

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 165


Chapter 4

1. Heat is transferred through a thin layer of liquid covering the immersed particles that
are in contact with the heating surface. The thickness of this layer is defined by the
location of a pore of the smallest diameter, resulting from the configuration formed by
a row of particles adjacent to the heating surface.

2. All the conducted heat is used to evaporate the liquid, a process which takes place
from the meniscus of the pore of the smallest diameter.

3. The minimum pore diameter can be computed from the equilibrium height of the
capillary draw, based on the measurements.

4. The interfacial surface (meniscus), on which evaporation proceeds can be described


by a single curvature radius.

5. The interfacial surface temperature equals the saturation temperature for the pressure
prevailing in the deposit.

6. In contact with the heating surface, particles form regular, cubic cells.

7. The heating surface temperature is constant on the whole surface.

For the sake of computation, it was assumed that that the heating surface – particle contact was
a point and the contact area amounted to 5% of the heating surface. Such an assumption did not
affect the results of calculations in a significant way according to the authors.

The heat conduction equation together with all the necessary boundary conditions were not
written explicitly in their paper and neither was the numerical procedure given. On the basis of
numerical calculations, a constant value of the heat transfer coefficient was obtained, inversely
proportional to the particle diameter:

–1
α = C Dpellet (4.2a)

which makes the heat flux density increase linearly with the heating surface temperature
difference:

–1 –1
q = C Dpellet (Ts − Tsat) = C Dpellet ∆Tsl (4.2b)

In spite of numerous simplifying assumptions, the model yielded exceptionally good agreement
with their experimental results, as shown in Fig. 4.5.

166 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

2⋅10
6

Monel pellets deposit height


5
10 3.17 mm
7 12.70 mm
25.40 mm

q, W⋅m-2
5
38.10 mm
3

4 Numerically calculated curve


10
3
7⋅10 -1
5⋅10 10
0 3 5 7
10
1
3⋅101
DTsl, K

Fig. 4.5. Water boiling in a layer of Monel particles of changeable deposit height and
constant number of pores per unit area 20 – 30 pores per square inch (20 –
30 mesh) from Ferrell and Alleavitch [86]

Ito et al. [104, 105] tried later to adopt a similar theoretical approach, i.e. solve the heat conduction
equation for a layer of heaped particles. It was necessary, however, to assume a number of
arbitrary experimental constants in their model and it showed poor agreement with experimental
data, see Fig. 3.15, and therefore it was not developed any further [104, 105].

Moss and Kelly model

Moss and Kelly [165] built two models in order to compute the thickness of a vapor layer in a
mesh wick in the heated part of a heat pipe (evaporator). It was assumed that the vapor layer was
formed at the contact surface of the porous layer and the heat pipe housing. In the first model,
the pressure drop in the liquid layer in the cold part of the heat pipe was calculated from Darcy’s
law based on the assumption that all the heat provided to the liquid was used for its evaporation
that occurred exclusively on the layer’s upper surface. Heat was supposed to be transferred
through the vapor layer by means of conduction. Moreover, it was assumed that the superheated
liquid temperature on the interfacial surface is higher than the saturation temperature due to the
capillary increase in pressure. On the basis of the above-mentioned assumptions, the vapor film
thickness was determined to be:

(4.3)

where L1 is the pipe segment length over which the liquid is transported to the evaporation section,
L 2 is the length of evaporation section, φ is the wetting angle, and θ is the pipe inclination angle
with respect to the horizontal.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 167


Chapter 4

In their other model, the pressure drop in the liquid layer was calculated from Darcy’s law as
before. A constant thickness of vapor film was also assumed. Vapor was assumed to leave the
porous layer only at its edges. While integrating the two-dimensional continuity equation, the
condition was used that the sum of pressure drops in the liquid and vapor layer could not exceed
the capillary increment in pressure. The following expression was obtained as a result:

(4.4)

The vapor layer thicknesses were measured with a neutron radiography technique [165]. Results
of the measurements are presented in Fig. 4.6. As can be seen in Fig. 4.6, only the second model
can be regarded as the proper one, although the error was significant. The discrepancy could
result from their very simplified model of heat transport in the porous layer, where it was assumed
that the all the heat supplied to the liquid was used for its evaporation and heat transport in the
vapor film occurred only through conduction while neglecting natural convection.

5
Model I Model II

4 ϕ=0
ϕ = 60°
ϕ=0
ϕ = 60°
3
δg , mm

2 Wick thickness (porous


layer) δ = 6.35 mm

1 Measurement error

Experimental data
0
0 0.25 0.50 0.75 1.00 1.25 1.50 1.75
q×10-5, W⋅m-2

Fig. 4.6. Water vapor film thickness as a function of heat flux for a 100 grade
mesh (mesh is one inch long) and a pipe inclination of θ = 20° from
Moss and Kelly [165]

168 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

O’Neill et al. models

O’Neill and co-researchers [68, 91, 182] proposed three models for boiling on surfaces with
porous structures. The capillary-porous structures they examined were those they gave the
tradename High Flux (of UOP). They thought that their microporous surfaces reference geometry
could be represented as a layer of spherical particles and they conducted detailed considerations
for such geometries.

Model I

The following assumptions were made for the sake of theoretical considerations:

• The vapor bubbles are generated inside the porous layer, in free spaces between solid
elements, and fresh liquid flows in through those connected pores, in which vapor is
not being generated at a given instant;

• The vapor is generated owing to the evaporation of a liquid microlayer covering the
layer solid elements;

• The total liquid superheating that is necessary for it to evaporate equals the sum of
temperature drops caused by (i) conduction through the liquid film and (ii) equilibrium
superheating that is characteristic of curved liquid-vapor interfacial surfaces:

∆Tsl = ∆Tsat + ∆Tl (4.5)

• The effects produced by the increase in the density of nucleation centers inside the
porous layer when compared with smooth surface, temperature drop through the layer
in the direction normal to the smooth surface and convection outside the layer were all
disregarded.

For the first two models, O’Neill et al. [91, 182] did not provide derivation of their full formula, but
the details can be found in Webb [310, 311] and Thome [288]. The superheating necessary for
bubble generation is described by the dependence (3.2) given earlier:

The temperature drop through the liquid film is expressed by the equation:

(4.6)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 169


Chapter 4

where the geometric coefficient B is a structural parameter of the porous layer, R p is the mean
pore radius, and the heat flux density q refers to that of the nominal heating surface (the surface of
the porous layer base).

It is possible to determine the minimum heating surface superheating necessary to initiate the
process of vapor generation for the pre-set values of the heat flux density and pore diameter. The
superheating is computed while differentiating (4.6) with respect to Rp, equating the derivative to
zero and eliminating Rp by means of algebraic transformations:

(4.7)

6
10
7 R 113 (0.14·10 5 Pa)
5 Water (0.14·105 Pa )
Water (0.39·105 Pa)
3 R 11 (0.39·10 5 Pa)
-1
α max, W⋅m ⋅K
-2

Nitrogen
105 Oxygen
7 Propylene
1.03 ·105 Pa

5 R 22
Ammonia
3
Ethanol
R 11
Water
R 113
Toluen
4
10 4
10 3 5 7 10 5 3 5 7 10 6
α exp, W⋅m-2⋅K-1

Fig. 4.7. Comparison of experimental and computed values of αmax for High Flux
microsurfaces, various liquids and pressures from O’Neill et al. [91, 288]

The maximum value of the heat transfer coefficient for the pre-set value of the heat flux density
amounts to:

(4.8)

170 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The constant in the equation (4.9), see equation (3.1), was determined on the basis of
experimental data for three different pressures and nine investigated liquids. The scatter of
experimental results with respect to equation (4.8) is shown in Fig. 4.7, which shows good
agreement with the data.

Model II

While discussing the improved version of the model proposed by O’Neill et al. [181], we should
refer to a detailed study of the model prepared by Webb [310] rather than to the original work.
The new model was based, as the earlier one, on the concept of two thermal resistances, yet the
impact of the porous layer properties on the liquid film evaporation was taken into account more
accurately. The heating surface superheating indispensable for boiling incipience is expressed as
follows:

∆Tsl = Ts − Tsat = (Ts − Tlg) + (Tlg − Tsat) (4.9)

where Tlg is the temperature of the liquid film interfacial surface. The second term in equation
(4.9) is described by the dependence of equation (3.2), whereas the first term results from the
solution to the one-dimensional Fourier transform for a liquid film. The heat flux density is written
as follows:

(4.10)

where Fl is the liquid microlayer surface, F is the heating surface, and δl is the microlayer
thickness. Convection in the liquid microlayer is disregarded. It is also assumed that the
temperature does not drop over the porous coating thickness and that the coating temperature
equals the heating surface temperature. In order to calculate the Fl/F, the following assumptions
are made:

1. The porous coating is made of particles of the same diameter.

2. Capillary channels inside the porous layer are of identical dimensions and are
connected with one another.

3. Pore dimension is defined as the diameter of the biggest sphere that can be inscribed
into a free space between the particles. The geometry of the particle packing is known
and two alternatives are possible, as illustrated in Fig. 4.8.

4. Each pore is an active nucleation center and offers a stable retention of vapor.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 171


Chapter 4

a) dgrn b) d grn

Dp

Dp

Fig. 4.8. Definition of the porous layer particle diameter dgrn and departure bubble
diameter Dp (pore diameter) for (a) staggered packing and (b) inline packing
(b) from [182]

The porous covering volume is:

V = F δ (4.11)

which, after inserting into (4.10), gives

(4.12)

Accounting for (4.12) and (3.2), and rearranging (4.9), this gives:

(4.13)

The term δl V/δ Fl can be computed if we know the type of particle packing [310]. O’Neill et al.
[182] proposed the following definition of the geometric coefficient B, which accounted for all
quantities dependent on the porous layer geometry:

(4.14)

Inserting (4.14) to (4.13) gives:

(4.15)

172 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Webb [310] demonstrated that the geometric coefficient B could be determined from the
dependence:

(4.16)

if the particle diameter dgrn and the type of packing are known.

Porosity ε measurements on selected High Flux microporous surfaces showed that it ranged from
0.50 < ε < 0.65, which corresponds approximately to the porosity for inline packing for which
ε = 0.48. Inserting numerical values (4.9) and (4.10) leads to the equation (3.3):

Comments on the model were made earlier while equation (3.3).

Model III

Czikk and O’Neill [68] proposed a model which provided a detailed account of the porous coating
geometry. Structural examinations and computations were made for tubes coated with the
patented High Flux microporous layer. In order to obtain detailed information on pore distribution
inside the capillary-porous structure, three samples underwent the following examinations:

• samples were sliced many times and metallographic sections were prepared;

• using a specialized apparatus and software (“Quantimet Image Analyzing Computer”


[68]), the pore distribution density was determined.

• making assumptions about the physical phenomena, similar to those made before, the
authors used detailed information on porous structure geometry to compute boiling
curves q = q(ΔT)

The agreement between the theoretical and experimental curves was only satisfactory, which
means that improvement on calculation compatibility with the experiment could be sought in
further developing the model with respect to the physics of the boiling process. That is what
Thome [288] suggested in his comments on the previous model by O’Neill et al. [182]. Thome’s
proposals [288] focused on the necessity of accounting for convection inside and outside the
capillary-porous structure and making the geometric coefficient B dependent on liquid thermal
properties and on the heat flux, comments supported by the Arshad and Thome investigations
[15], which indicated that the liquid film thickness in concave pores decreased with increasing
heat flux.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 173


Chapter 4

In all models proposed by O’Neill et al. [68, 91, 182, 310], a constant inclination of the boiling
curve q = q(ΔT) is obtained, which contradicts many experimental results, especially for low
superheating of the heating surface where the slope changes significantly.

Cornwell et al. model

Cornwell et al. [65] decided that the heat flux could be connected with the vapor mass rate of flow
generated inside the porous layer, i.e. with pressure drop in the vapor across the layer thickness.
The friction coefficient for vapor flow through the porous layer was written in the form:

m
F = C Re (4.17)

where the definitions of the coefficient and the Reynolds number were as follows:

(4.18)

(4.19)

In equations (4.18) and (4.19), ∆pg is the vapor pressure drop across the porous coating thickness
δ, Dp is the mean pore diameter and wg is the vapor flow velocity in the capillary.

On the basis of experimental investigations it was found out that in boiling in a wetted porous
structure, the coefficient f decreases with increasing flow velocity, hence m = – 0.71. Moreover,
for a large vapor flux and a thin porous layer, the area Fg, occupied by departing vapor does not
change across the layer thickness. The mass flow rate of vapor amounts to:

Gg = ρg wg (4.20)

174 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

After inserting (4.17), (4.18) and (4.19), this becomes:

(4.21)

where C1 is a constant containing the constant C, density ρ g, viscosity μg, pore diameter D P and
the exponent m. The heat flux, referred to the total heating surface, amounts to:

(4.22)

Inserting (4.21) in (4.22) yields:

(4.23)

Experimental heat flux values as a function of Fg /F ratio were presented earlier in Fig. 4.2,
together with comments on the experimental results. Good agreement of the theoretical
calculations with the authors’ experimental data was obtained.

Inserting the experimental value of the boiling curve inclination q/(Fg /F) in Fig. 4.2 into (4.23), the
authors obtained the value of Δpg, whose value was much lower than that of counterbalancing the
water column height above the heating surface.

Conclusions drawn in [65] from their experimental and theoretical investigation were as follows:

• areas taken up by liquid and vapor inside the porous layer coexist for a wide range of
heat fluxes;

• heat flux can be correlated with the magnitude of the areas occupied by vapor
and liquid;

• the critical (maximum) value of the heat flux can be obtained by extrapolating the
curve q = q (Fg /F) in Fig. 4.2.

Man’kovski et al. model

The concept of treating a capillary-porous structure as numerous vertical channels (“chimneys”),


connected horizontally and carrying away vapor from inside the structure was first put forward by
Cohen [61]. He proposed simplified models of heat and mass transfer, with which it was possible
to get analytical solutions, for the analysis of boiling process occurring on the surface of steam
generator pipe covered with boiler scale. In accordance with Cohen’s calculations [61], the heat
conducted through the skeleton of the capillary-porous structure was almost totally used (99,7%)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 175


Chapter 4

for liquid evaporation that proceeds at the sites where horizontal minichannels (feeding the liquid)
are connected with the vertical ones.

1 – porous covering with capillary channels; 2 – heating surface from Man’kovskij [155]

Fig. 4.9. Capillary-porous structure: a) real one; b) structure model

Although Man’kovskij et al. [155] did not quote Cohen’s work [61], their model can be regarded as
a development of his original concept. The authors [155] simplified the porous layer by presenting
it as a system of vertical microchannels sucking in liquid and carrying away vapor, which is shown
in Fig. 4.9. The model assumes that the basic mechanism of boiling in the porous layer is vapor
generation inside the layer. The vapor is carried away outside through the vapor channels and
other liquid channels suck in the liquid, as illustrated in Fig. 4.9.

The much larger values of the boiling heat flux obtained on surfaces with porous coatings in
comparison with smooth surfaces are attributed to the following:

a. the interfacial surface inside the porous coating, which considerably lowers the
heating surface superheating necessary for boiling incipience;

b. the high values of the heat transfer coefficient of the laminar liquid flow inside the
pores (liquid channels);

c. the developed surface of the porous covering.

176 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

In accordance with the model presented in Fig. 4.9, the authors [155] treated porous coating
as a system of cylindrical capillaries connected with one another. The vertical capillaries must
resemble a real porous layer with respect to flow and thermal characteristics and also with respect
the capillary suction, and therefore it was necessary to introduce three diameters:

a. equivalent capillary diameter

Dcap = 0.41 dgrn (4.24)

b. hydraulic diameter

(4.25)

c. equivalent thermal diameter, which results from the volume and area of pores and
substitute capillaries being equal

(4.26)

In order to calculate the boiling heat flux, the authors [155] treated the walls of the layer between
capillaries as rectangular fins of the width lfin calculated from assuming that the volume of porous
structure particles and the volume of capillaries were equal:

(4.27)

It is assumed that the vapor-liquid interfacial surface is located in the proximity of the heating
surface, as the conditions are the most favorable there. It is also assumed that there are m + 1
capillaries to each nucleation center, out of which m are filled with liquid and one capillary carries
the vapor away outside. Thus, the number of capillary channels per unit area amounts to:

(4.28)

and the number of nucleation centers N is:

(4.29)

If qcap is the heat flux in the channel filled with liquid and generating vapor, heat transported per
one nucleation center is:

Q = � DT δ qcap m (4.30)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 177


Chapter 4

The heat flux referred to the flat base surface is then:

(4.31)

and after inserting (4.26) and (4.27), this becomes

(4.32)

The pressure drop due to liquid and vapor flow in the capillaries is counterbalanced by the
pressure increment resulting from the action of surface tension forces:

Δpcap = Δpg + Δpl = (m + 1) Δpl (4.33)

Assuming that the flow inside the capillaries is laminar, it is possible to determine Δpc from
Poiseuille’s equation:

(4.34)

From an energy balance, the liquid flow rate in a capillary is:

(4.35)

The number of capillaries is determined by inserting (2.6), (2.9), (4.24), (4.34) and (4.35) into
(4.33):

(4.36)

178 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Fig. 4.10. Dimensions and temperature distribution in equivalent microfins, in


accordance with [155]

In order to determine the heat flux qcap in a capillary, the porous coating is replaced with a system
of equivalent microfins of the height δ, thickness lfin, length �DT and thermal conductivity equal to
the covering skeletal conductivity λskl, as shown in Fig. 4.10. It is assumed that, taking into account
the liquid meniscus in the capillary, that the liquid temperature along the microfin is equal to the
saturation temperature. Then temperature difference at the microfin base is:

ΔTsl = Ts – Tlg = (Ts – Tsat) – (Tlg – Tsat) = ΔTs,sat – ΔT* (4.37)

where, on the basis of (2.1), the term ΔT* can be written

(4.38)

In accordance with all the assumptions made above, the heat flux on the capillary surface
(equivalent microfin with insulated tip) amounts to (where “th” = hyperbolic tangent):

(4.39)

If flow in the cylindrical capillaries is assumed to occur for small Re, α is described by the
dependence of laminar flow in a uniform temperature channel:

(4.40)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 179


Chapter 4

For the problem under consideration, it can be proven that the argument in the hyperbolic tangent
is large, so it is possible to assume the tangent equals one. Adopting this conclusion in (4.39), and
also inserting expressions (4.39) and (4.40) into (4.32), the following equation is obtained:

(4.41)

where m can be computed on the basis of equations (4.36) and (4.40).

Heat transfer in the capillaries carrying vapor away was disregarded as negligibly small in this
model. The friction coefficient for vapor flow inside the capillary-porous structure was also
neglected. The skeletal thermal conductivity λskl is difficult to determine and can vary by an order
of magnitude, depending on the way the structure was manufactured and on the character of how
the particles contact with one another [301].

Notably, the expression (4.41) yields almost a linear dependence of q on ΔTs,sat, especially for
m >> 1. However, many experimental investigations have shown that the boiling curve q(ΔTs,sat)
inclination changes with increasing q. The range in which (4.41) is applicable as regards the heat
flux, layer thickness, type of porous material, etc., was not given in their paper. The boiling curve
(4.37) shows satisfactory agreement with the Ferrell and Alleavitch experimental data [86] in
Fig. 4.11.

Comparison of the authors’ experimental data for tubes with porous coatings of porosity ε = 0.46
to 0.6 and thickness δ = 0.22 to 1 mm with boiling curves obtained in accordance with (4.41) can
be found in [73]. Only qualitative agreement was obtained. Quantitative agreement was observed
–1 –1
only in a small range of low values of skeletal conductivities λskl from 0.01 to 0.05 W m K .

180 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

3⋅105
Water
Atmospheric pressure
2

105

q, W⋅m -2
8

2
-1
-2

104
1 2 4 6 8 10 20 30
DTsl, K

dgrn = 0.507 mm; 1 – δ = 3.15 mm; 2 – δ = 25.4 mm

Fig. 4.11. Comparison of the Ferrell and Alleavitch experimental data [86] with the
boiling curve from the Man’kovski et al. model [155]

Smirnov et al. model

This model seems to further develop that of Man’kovski et al. [155], in which the capillary porous
structure was substituted with a microfin system. The model evolved in successive publications of
Smirnov and co-researchers and was verified against many experimental data [3, 5, 247 – 252,
254 – 256]. Smirnov’s investigations [247, 248, 254] deal with porous coatings that are both fully
and partially flooded with a boiling liquid. According to [247, 254], four characteristic means of heat
transfer through a capillary-porous structure occur, which are shown in Fig. 4.12.

Following Ferrell and Johnson [87], Smirnov and co-researchers adopted the following simplifying
assumptions about the heat transfer process:

1. Heat transfer occurs between the capillary-porous structure skeleton and superheated
liquid. The conduction from the skeleton to the liquid is the source of the main thermal
resistance.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 181


Chapter 4

a) evaporation; heat conduction through the structure of conductivity λ = λef;


b) micro heat pipes inside the structure; c) boiling inside the structure totally
flooded with liquid; d) boiling inside the structure with capillary feed

Fig. 4.12. Heat transfer mechanisms in a capillary-porous structure from Smirnov and
coworkers [247, 254]

2. Vapor generation and its flow out of the structure affect only the way liquid motion is
organized. Heat transfer in capillaries filled with vapor is disregarded because the
number of capillaries filled with liquid is much higher.

3. The liquid temperature inside the capillaries is constant and corresponds to that of
capillary pressure resulting from the liquid-vapor interfacial surface curvature.

The following heat transfer mechanism are assumed to operate:

1. For heating surface superheating ΔT < ΔT*, where ΔT* is computed from the formula
(4.38) and vapor is not present inside the layer, the only possible heat transfer
mechanism is conduction from the heating surface to the phase separation boundary
as in Fig. 4.12a. The presence of gas inside the structure can cause the formation
of vapor-gas bubbles, which function as thermal insulators diminishing the effective
thermal conductivity of the structure λef.

2. For ΔT > ΔT*, two heat transfer mechanisms can be identified. The first one, shown
in Fig. 4.12b, is termed a micro heat pipe because the vapor formed at the heating
surface completely condenses in the upper part of the capillary-porous structure.
The higher are ΔT* and λef, the less flooded is the structure and the more likely the
occurrence of this heat pipe mechanism. In the other case, as shown in Fig. 4.12c and
4.12d, vapor exits the structure no matter how the layer is fed with liquid. In [247], the

182 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

authors state that heat fluxes in both cases are approximately equal and vapor leaving
the structure results from excess capillary pressure at the heating surface (2.6), where
Dp is the pore diameter:

4 σlg
∆p =
Dp

Influenced by this excess pressure (2.6), curvilinear interfacial surfaces are formed at the sites of
the capillary-porous structure in contact with the heating surface and also the structure elements
in contact with one another. Vapor flow out of the porous structure induces a liquid flow into the
structure. In this model of vapor and liquid flow, the main thermal resistance is that of conduction
through the liquid film wetting the capillary-porous structure skeleton.

The structure elementary cell is presented as a fragment of a finned surface, where D is a side
of a square pore, as shown in Fig. 4.13. The thickness of a conventional microfin is linearly
proportional to 1-ε. Microfins are covered with a liquid microlayer of the mean thickness δ c.
In order to simplify calculations, it is assumed that δc = const over the whole microfin height,
i.e. the porous coating thickness. For simple heat conduction through the film, the heat transfer
coefficient is:

(4.42)

The equations describing one-dimensional heat conduction in the fin in Fig. 4.13 and its boundary
conditions are as follows:

(4.43)

(4.44)

(4.45)

x=L ϑ=0 (4.46)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 183


Chapter 4

The notation of equations (4.43) to (4.45) results from assuming the following relationships:

Tfin = Ts (4.47)

cell
ϑ = Tfin − Tlg = (Ts − Tsat) − (Tlg − Tsat) = ∆Tsl − ∆T* = Ts − Tsat (4.48)

cell
Tsat = Tsat + ∆T* (4.49)

cell
Tsat is the saturation temperature inside the porous layer, decreased by the temperature
increment resulting from the capillary force, where in accordance with (4.38), is:

Heat transfer through the fin tip is disregarded assuming that Tfin = Tsat, which leads to the
boundary condition (4.45b).

Fig. 4.13. Simplified model of an elementary cell of a capillary-porous structure from


Smirnov and coworkers [247, 254]

cell
The microfin surface temperature is equal to the saturation temperature Tfin = Tsat , which results
from the boundary condition (4.46), and this constitutes the boundary condition for vapor bubble
generation inside an elementary cell between the fins. From the condition (4.46), the length of the
segment L, measured at the base of the microfin on which boiling occurs, is determined. Above
x = L, the liquid superheating is lower than the minimum ΔT* value necessary to initiate nucleation

184 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

as indicated in Fig. 4.13. From equations (4.43) to (4.45), the heat flux transferred through a single
microfin in boiling is computed to be:

(4.50)

Many works quoted earlier, especially those on mesh structures, indicate that the boiling heat flux
weakly depends on the porous layer thickness. The length of the active segment L satisfies this
condition when mL >> 1, owing to which it is possible to write (4.50) in a simplified form:

cell
q = λef m (∆Tsl − ∆T*) (4.51)

The mean temperature drop in an elementary porous cell can then be expressed as follows:

∆T = ∆Tsl (1 − ε*) + ∆T2 ε* (4.52)

where ε* is superficial porosity of the structure and ΔT2 is the temperature drop in the part of the
cell that is covered with a liquid microlayer. When computing the heat flux referred to the whole
heating surface, temperature drop across the thickness of the microlayer covering the porous
cell is disregarded, which yields: ΔT2 = ΔT*. This heat flux is defined in relation to the mean
temperature drop in the cell ∆Tsc and it takes into account the fraction of the surface occupied by
the fins. Then, equation (4.51) is rearranged to the form:

q = λef m (1 − ε*) (∆Tsl − ∆T*) (4.53)

If the heat transferred by the part of the heating surface not covered with microfins is taken into
account, expression (4.53) takes on the form:

(4.54)

The liquid flows inside the capillary–porous structure to the sites where menisci are formed by
the action of capillary forces. Hydraulic resistance to the liquid flow in the evaporation zone Δpl
and hydraulic resistance of departing vapor Δpg are related through the capillary pressure by the
dependence:

(4.55)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 185


Chapter 4

For laminar flow:

(4.56)

where ffin is fin intersection, L1 is liquid path towards the meniscus, L 2 is vapor path towards the
meniscus and O is cell circumference. From expression (4.55), the following can be obtained:

(4.57)

Assuming that the vapor channel length L 2 is of the same order as the length as that of the liquid
inflow, we obtain:

(4.58)

which results from vapor and liquid pressure drops

(4.59)

Taking into account (4.58), the inequality (4.57) takes on the form:

(4.60)

If uniform distribution of menisci in the structure cells is assumed, then the larger is the coefficient
m, the shorter is the length L1 over which the evaporation process occurs:

(4.61)

From the expressions (4.60) and (4.61), the equation for liquid microlayer thickness is derived:

(4.62)

186 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Next from (4.51), (4.52), (4.53), (4.58), (4.61) and (4.61), the expression (3.49) presented in an
earlier chapter, is obtained:

Quantities found in the formula above are explained in the description of equation (3.49).
Relying on the Afanasev’s experimental results [3] for water, ethanol and R-113 for meshes with
homogeneous cell distribution, the authors determined the constant in the equation (3.49), see
Fig. 3.41.

In their next papers, Smirnov et al [248, 252, 256] made an attempt to generalize the model
presented above so that it would cover capillary-porous structures with a stochastic pore
distribution. The general equation (3.49) can be written in the form:

(4.63)

where Le = kDp, whereas k is a proportionality coefficient, dependent on the capillary-porous


structure cell geometry, Φ (psat) is a function of boiling liquid thermal properties, calculated from
the saturation pressure

(4.64)

In accordance with their model [248, 256], for boiling in metal fibrous capillary-porous structures
with random pore distribution, only those pores, whose superheating is greater than the minimum
described by the equation (4.38) will be activated. This means that when the heat flux q and
superheating ΔTsl grow, pores of smaller and smaller diameter Dp,i get activated. According to
Smirnov and Vinogradova [256], those pores will be successively become active, for which the
minimum flux increment is sufficient to provide for their continuous and stable activity. Hence, the
necessary and sufficient condition to determine the sequence of pore activation has the following
form:

(4.65)

It is assumed that a real, discrete pore diameter distribution is described by the distribution density
function γ (Dp,i), whereas the step changes in pore diameter amounts to ΔDp,i. It should also be

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 187


Chapter 4

noted that quantities λef, ffin, and Oi are functions of Dp,i, so they will depend on their distribution.
The number of pores per unit area is:

(4.66)

where

(4.67)

Relationship between surface porosity and that of the whole capillary porous-structure volume is
as follows:

(4.68)

It is assumed that the section of the skeleton elementary cell structure, connected with a single
pore, is square and a vapor channel inside such a cell is cylindrical in shape with the diameter Dp,i.
Hence, the section area of a conventional fin is:

(4.69)

Consequently, the characteristic cell dimension is:

(4.70)

According to the assumptions made above, pores satisfying the condition (4.65), will successively
get activated, owing to which, after accounting for the relationships given by (4.63), (4.66), (4.67),
(4.68), (4.69) and (4.70), it is possible to determine the equation describing the sequence of pore
activation as a function of diameter:

(4.71)

188 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where A1 and A 2 are certain functions λef, λl and ε* (Dp,i). The analysis of experimental data
presented in [239] makes it possible to assume that pore diameter distribution in metal fibrous
capillary-porous structures is a normal or logarithmic normal distribution. If given properties of
pore distribution are known, in order to write the function γ (Di), it is enough to determine limiting
values of Dmin and Dmax. If equations (4.63) and (4.71) are considered jointly:

(4.72)

where

(4.73)

and Dp is the most probable pore diameter.

For a normal or logarithmic normal distribution of pore diameters, distribution functions are as
follows:

(4.74)

(4.75)

Inserting equations (4.74) and (4.75) to (4.71) yields:

a. for normal distribution

3 2
χ − χ (0.5 + b) + χ (0.5b − 0.02) − 0.014b = 0 (4.76)

b. for logarithmic normal distribution

5 + (b χ − 1) (3 − C1 ln χ) = 0 (4.77)

36
C1 = (4.78)
Dmax / Dmin

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 189


Chapter 4

Solving simultaneously equations (4.72), (4.74) and (4.76) or (4.72), (4.75) and (4.77) leads to the
determination of the boiling curve q = q (ΔTsl).

Further considerations in Smirnov and Vinogradova [256] lack sufficient clarity to understand
precisely what was assumed, and moreover the authors did not provide the data necessary to
compare unequivocally the computation results with quoted experimental data. For instance,
it was not stated which of the proposed pore diameter distributions was finally adopted for
calculations or what were the pore diameter distribution characteristics of the porous coatings
examined by other researchers.

Nishikawa et al. models

Nishikawa and co-researchers [177] proposed two completely different simple models of boiling in
capillary-porous structures of sintered powders. Their first model is based on the analysis of vapor
flow through a porous layer. It is assumed that the vapor staying in the lower part of the layer,
which is adherent to the heating surface, is superheated correspondingly to its pressure drop in
the layer. Combining the Burke-Plummer equation [202] describing pressure drop in the layer with
Clapeyron equation results in the following relationship:

(4.79)

The heat flux q calculated on the basis of (4.79) is about a hundred times greater than
experimental values, as illustrated in Fig. 4.14.

In their second model, they assumed that, due to a very small Rayleigh’s number, free convection
in boiling inside the porous layer can be neglected. Heat transfer from the heating surface to the
liquid thus takes place through one-dimensional heat conduction, in which the substitute heat
transfer coefficient is calculated using the following relationship based on the porosity of the layer:

λm = ε λ1 + (1 – ε) λs (4.80)

and the heat flux is then calculated as:

-1
q = λm δ (Ts – Tsat) (4.81)

Although the expression (4.81) is obviously simple, it provides a fairly good top-down estimation of
the heat flux’s dependence on the heating surface superheating as shown in Fig. 4.14.

190 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

7
3⋅ 10
4 ε 3 d grn ρ g2 h 3lg Ts
q= ln
10 7 7 (1 − ε )δ Tsat

5
3 a Dgrn = 500 µm, ε = 0.427

Dgrn = 1000 µm, ε = 0.525


10 6
R 113, atmospheric pressure
5 Sintered bronze powder layer ,
3 thickness δ = 2 mm
q, W⋅m -2

q = λ m δ −1 (Ts − Tsat )
10 5
ε =0.427 Dgrn = 500 µm
5
ε =0.525 Dgrn = 1000 µm
3

10 4 b
ε = 0.427 Dgrn = 500 µm
5
ε = 0.525 Dgrn = 1000 µm
3
Experimental data

10 3 -1
3 5 100 3 5 101
10 3·10 1
DTsl, K

Fig. 4.14. Comparison of experimental values of heat flux for powder capillary-porous
structures with values computed from the model taking into account
(a) vapor flow resistance in the structure and (b) one-dimensional heat
conduction model (b) from Nishikawa et al. [177]

Tehver et al. model

Tehver and co-researchers have dealt with modeling boiling heat transfer in porous layers,
modeling of the boiling crisis, the hysteresis phenomena and associated experimental
investigations in numerous publications [148, 200, 273, 274, 276 – 281, 283 – 287, 299]. In the
Tehver et al. model [200, 285, 287], it was assumed that liquid evaporation inside the porous
coating occurs in the layer parallel to the heating surface, directly adjacent to the heating surface.
Three kinds of boiling process were differentiated. In the first one, the evaporation surface comes
in contact with the heating surface located below the porous coating. In the second kind, the
evaporation surface inside the coating is separated from the heating surface by a layer of dry,
non-wetted porous material. This kind of boiling is termed intralayer boiling crisis [287]. In the third
kind of boiling, the space between the heating surface and the evaporation surface within the

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 191


Chapter 4

porous coating is filled with vapor and this kind of boiling is called boiling crisis [287]. Each kind
of boiling is accompanied by the respective relationship between the heat flux and the heating
surface superheating. The real capillary-porous structure was substituted with a model containing
simple cylindrical capillaries of a length equal to the layer thickness, the same pore dimension
distribution and the same porosity, as illustrated in Fig. 4.15.

If it is assumed that all the heat flux is used to generate vapor, Darcy’s equation describing
pressure drop in the porous covering will take the form of:

(4.82)

where p1 is pool pressure and p is mean vapor pressure inside the porous layer. Capillaries of the
radius R < 2σlg /Δp are assumed to be filled with the liquid and the remaining ones are assumed
to be filled with vapor with a liquid film covering their surfaces. Unlike in other models, it is not
the boiling heat flux that is determined but the total temperature drop inside the porous coating.
The latter depends on the structural parameters of the porous coating, such as its thermal heat
conductivity, its porosity and its thickness.

The total temperature drop in the porous layer has three components as indicated in Fig. 4.15:

1. The temperature difference between the heating surface and vapor in the porous
layer, ΔT1;

2. The temperature difference between vapor in the porous layer and vapor in the
pool, ΔT2;

3. The temperature drop in the dry (non-wetted) part of the porous layer at the heating
surface, ΔT3.

In order to determine temperature difference ΔT1, the one dimensional heat conduction equation
for a rod of length δ, covered with liquid microlayer of the thickness δ micro was solved. The
equation has the form and boundary conditions as follows:

(4.83)

192 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

1 – heating surface; 2 – dry (non-wetted) sublayer; 3 – liquid pores; 4 – vapor pores

Fig. 4.15. Temperature distribution in the porous layer from Tehver et al. [285]

(4.84)

where T0 is the cooling liquid temperature at the heating surface and in the vapor pores, T1 is the
temperature at the base of the rod, α is the heat transfer coefficient on the lateral surface, and
αfs is the heat transfer coefficient on the base surface of the rod (external surface of the porous
covering). The solution to (4.83) yields [161] the following relationship:

(4.85)

(4.86)

where O is the circumference of the rod per unit external surface, m and n are coefficients
characterizing heat transfer on its lateral surface and root surface, and δ micro is the thickness of
the liquid microlayer covering the surface of the vapor capillaries.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 193


Chapter 4

The conventional rod circumference per unit sublayer surface, that is proper internal open surface,
is computed using Kozena’s equation:

(4.87)

The porosity of the open pores εopn and the permeability Kg are determined experimentally or
computed on the basis of the known function of pore dimensions and density:

(4.88)

The pressure at the bottom of the porous layer (at the heating surface) depends, in accordance
with Darcy’s equation (4.82), on the heat flux q and the vapor layer’s permeability Kg. The
temperature of the interfacial surface inside the porous layer is higher than Tsat by the value
resulting from the excess pressure. Inserting (4.82) to the Clausius-Clapeyron equation gives:

(4.89)

When the liquid stops flowing to the heating surface, a thin dry (non-wetted) sublayer is formed in
the porous coating. The sublayer thickness Δδ is calculated from the following expression:

(4.90)

In this expression, qintcs is the heat flux characteristic of the intralayer crisis, which occurs when
the sum of the pressure drops due to friction of the liquid and vapor flow through the porous
layer is higher than the maximum negative pressure of capillary suction [115, 118, 287]. The
temperature drop over the height of the dried sublayer amounts to:

(4.91)

Thus, the total temperature drop between the heating surface and the liquid boiling in the pool is
the sum of three temperature drops:

ΔTsl = ΔT1 + ΔT2 + ΔT3 (4.92)

194 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

After using (4.85), (4.89) and (4.90) the following expression is obtained:

(4.93)

Temperature difference given by (4.93) is primarily a function of the porous coating thickness.
The curve ΔTsl = ΔTsl (δ) has a minimum, which at a set value of q determines the optimum
layer thickness δ, for which the optimum heat transfer conditions are obtained, that is
αmax = q / ΔTsl (δ) min.

Tehver and Tunik [285] provided some example calculations in accordance with equations (4.85),
(4.89), (4.90) and (4.93) for R-113 for the following boiling conditions:

5 –2 –12 2
q = 3.92 ∙ 10 W m ; ε = 0.3; εopn = 0.073; Kg = 0.276 ∙ 10 m ;

–6 –1 –1 3 –2 –1
δmicro = 5 ∙ 10 m; λs = 45 W m K ; αfs = 5 ∙ 10 W m K .

Their resulting simulations are shown in Fig. 4.16.

50
R 113
λs=45 W⋅m-1⋅K-1
DTsl, K

25

4
3
1
2
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
δopt δ, mm

1 – ΔT1; 2 – ΔT2; 3 – ΔT3; 4 – ΔTsl, from Tehver and Tunik [285]

Fig. 4.16. Dependence of temperature difference ΔTsl between the heating surface and
the liquid on the porous coating thickness

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 195


Chapter 4

4
8⋅10

3
7

6
2

-1
α, W⋅m ⋅K
-2
4
1

4
1⋅10
Smooth surface

7⋅10 3
0 0.2 0.4 0.6 0.8 1.0
δ, mm

5 –2 –1 5 –2 –1 5 –2 –1
1 – q = 1 ∙ 10 W m K ; 2 – q = 2 ∙ 10 W m K ; 3 – q =3 ∙ 10 W m K

Fig. 4.17. Heat transfer coefficient dependence on the porous coating (sintered
powder) thickness and heat flux from Tehver and Tunik [285]

The calculations showed that the ΔT1 component decreases with an increase in the porous
coating thickness. The ΔT2 component depends linearly on the coating thickness up to a certain
value, above which the menisci curvature radius becomes minimal and ΔT2 has a constant value.
The pre-set value of the heat flux q becomes that which corresponds to the intralayer boiling crisis
qintcs for the thickness δintcs = 1.8 mm. When δ > δintcs, the ΔT3 component is greater than zero.

Experiments demonstrated, as shown in Fig. 4.17, that the optimum coating thickness becomes
smaller with increasing heat flux, which confirms the correctness of the expression (4.93). Their
investigations were conducted for flat copper porous structures made of sintered powders [285,
286]. As Cieslinski [58] rightly pointed out, the model contains many parameters that are difficult
to determine. The major ones are the thickness δ micro of the microlayer covering vapor capillaries
and the heat flux qintcs at which the intralayer boiling crisis occurs.

196 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Pichlak together with Tehver [200] continued development of this model, additionally taking into
account:

1. The pore diameter distribution curve that originated in Tunika et al.’s investigations
[296] for capillary-porous structures of sintered powders (it was used for computing
the porosity of open pores εopn and the liquid and vapor layer permeabilities, K l
and Kg);

2. The influence of the thermal conductivity of the porous coating on the porosity and the
pores filled with liquid [29].

Computation results obtained by Pichlak and Tehver [200] for ΔT = ΔT (δ, q), assuming a
selection of values for the remaining unknown parameters, confirmed previously drawn qualitative
conclusions with respect to the drop in the value of δ opt with the increase in q. The influence of the
other structural parameters, such as the pore diameter or the thermal conductivity of the coating
material, on the value of ΔT were also analyzed. These computed results, however, were not
verified experimentally. They concluded that a porous coating can be optimized with respect to
three structural parameters: thickness, porosity and pore diameter distribution, whereas
an increase in the structure material thermal conductivity always favorably affects the heat
transfer process.

Orlov and Savelev model

Based on an analysis of the experimental and theoretical investigations conducted by previous


researchers, Orlov and Savelev [183, 184] advanced a thesis that in boiling in slits and capillary
channels, it is the resistance of the liquid microlayer separating a vapor bubble from the channel
wall that is decisive for the thermal resistance for heat transfer. Like in Man’kovski et al. [155]
and Smirnov et al. [247, 254], the capillary-porous structure performs the function of microfins
operating over the length from the base to the boiling liquid layer found above the porous coating.
It was also assumed that liquid motion in capillary channels occurs due to the action of surface
forces which are far greater than mass forces. Owing to investigations conducted by Grigorev et
al. [93], the microlayer thickness was determined from the relationship:

(4.94)

where wg is saturated vapor velocity in the vapor channel

(4.95)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 197


Chapter 4

In the formula (4.95), Korneev’s dependence [111, 112] describing the dryness level is used:

(4.96)

The velocity wg and the dryness level φ are determined using a balance of forces acting on a
vapor bubble inside the capillary channel and K is an experimental constant.

In accordance with Fig. 4.18, it is assumed that the porous coating has two kinds of channels with
diameters Dl and Dg, where Dl < Dg. The majority of the open capillary channels have a diameter
of Dc; moreover, on the basis of experimental data coming from various sources, it has been found
that the following relation holds:

2
Dg = m D2l (4.97)

where 0 < m < 10 [184]. Furthermore, it is assumed that that heat transfer takes place primarily
in the vapor channels, Dg. Channels with the diameter Dl are those feeding the liquid into the
porous layer. This assumption means that the local heat transfer coefficients inside the porous
layer satisfies the relationship: αg >> αl. They are constant along the length of capillary channels,
i.e. across the coating’s thickness. If the dimensions of feeding (liquid) channels and porosity are
2
known, it is possible to specify the number of open capillaries per m :

(4.98)

Fig. 4.18. Diagram of boiling heat transfer process in the capillary-porous structure,
according to Orlov and Savelev [184]

198 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The number of vapor channels is determined using:

(4.99)

where B is the number of capillary liquid channels feeding one vapor channel and m is an
experimental constant.

The heat flux supplied to a single channel, in which vapor bubbles are generated, is then:

(4.100)

where q is the heat flux on the base surface of the porous structure. Using the assumption that
αg >> αl, the mean value of the heat transfer coefficient inside the porous channel is determined:

(4.101)

The mean temperature difference between the vapor channel wall and the vapor bubble at the
height δ/2 is then:

(4.102)

Temperature difference between the base (fin foot) and the liquid above the porous structure (fin
tip) is determined using the known function for a single fin:

(4.103)

where

(4.104)

and O = � dfin is the fin circumference, λm = λs (1 – ε) is the substitute thermal conductivity of the
capillary-porous structure solid filling and Ffin is the area of the microfin cross section.

The number of microfins is assumed to equal the number of vapor generation centers, hence:

(4.105)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 199


Chapter 4

Thus the equivalent microfin diameter is:

(4.106)

Furthermore, it is assumed that the temperature of the liquid wetting the conventional microfin is
constant along its height. After taking into account the temperature increment due to the action of
capillary forces, the computed temperature difference between the base and the saturated liquid
is

∆Tsl = ∆Tfin,sat − ∆T* (4.107)

where, in accordance with (2.1):

The boiling heat transfer coefficient for a surface with a capillary-porous coating can then be
determined on the basis of Newton’s law of cooling:

(4.108)

Introducing expressions describing the vapor velocity wg (4.95), the liquid microlayer thickness
δmicro (4.94), the degree of dryness φ (4.96) and temperature difference ΔT into ΔTsl (4.107), the
Nusselt number correlation for boiling on a heating surface with a porous coating is obtained
(3.45):

Definitions of similarity numbers and other dimensionless numbers are provided by equations
(3.37) to (3.44). Fig. 3.31 presents the comparison of computed values of the Z number with
experimental data.

200 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Nakayama et al. model

Nakayama together with co-workers [167, 168] presented a complex model of the boiling process
for the industrial enhanced surface Thermoexcel-E and alike. This kind of porous structure is
characterized by small openings connecting the internal channels with the surrounding liquid
and also a large number of channels inside the structure. Three kinds of boiling processes in
the channels were differentiated as illustrated in Fig. 4.19: (i) for the lowest range of heat flux,
Nakayama et al. [167, 168] proposes a kind of boiling called “flooding”, in which the majority of
internal channels is occupied by liquid. The active pores function as individual nucleation centers;
(ii) for higher heat fluxes, they suggested boiling was characterized by a “suction - evaporation”
process. Here, the liquid is sucked inside the channels through the inactive pores due to vapor
being pushed outside the capillary-porous structure by bubbles growing in active pores. The
entering liquid spreads out inside the channels and evaporates from menisci in the corners and
(iii) for a further increase in heat flux causes the “drying” of channel walls, i.e. makes them totally
vapor-filled. The vapor in this case forms from bubbles growing on the external surface of the
porous covering as shown at the bottom of Fig. 4.19.

Fig. 4.19. a) Thermoexcel-E surface (see Fig. 3.1.b); b) kinds of evaporation in the
internal channels from Nakayama et al. [167, 168]

Observations of the boiling process for this type of capillary-porous structure were described
earlier and resulted in Nakayama et al. [167, 168] building the model for the bubble growth
dynamics, presented in graphic form in Fig. 4.20, which refers to the “suction - evaporation” boiling
process shown in Fig. 4.19. Three basic phases are addressed:

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 201


Chapter 4

1. Phase I (pressure increase period): evaporation inside the tunnel causes the vapor
pressure to increase until the instant when the vapor nucleus located at the pore
reaches the shape of a hemisphere, directed outside into the liquid pool.

Fig. 4.20. Model of bubble growth dynamics by Nakayama et al. [167, 168]

2. Phase II (pressure drop period): a certain number of pores activate and their vapor
nuclei grow to become bubbles. Initially, bubble growth results from the pressure
difference on the interfacial surface, which makes vapor flow into bubbles from the
channel and induces a drop in the pressure in the tunnel. Bubbles then continue to
grow due to the action of mass forces on the surrounding liquid.

3. Phase III (liquid suction period): while a bubble grows, the pressure inside the channel
decreases to become lower than the pressure of the surrounding fluid, which is then
sucks liquid inside the channels at the inactive pores. Finally, the bubble departs,
all pores are closed with liquid menisci, and the liquid advances inside the channels
because of the action of capillary forces and the whole cycle is repeated.

Their heat transfer model corresponds to the cyclic process of bubble growth and departure
presented above. It is assumed that heat transfer proceeds by two mechanisms: external
convection and internal evaporation of the liquid layer. The external convection process is
enhanced by the departing vapor bubbles, which causes liquid mixing. The liquid microlayer
covering the internal surface of channels is evaporated and the heat of vaporization is carried away

202 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

by the vapor bubbles leaving the structure through the pores. Evaporation on the external surface
is neglected as well as any convection inside the channels. Thus the total heat flux density is the
sum of:

q = qcon + qtun (4.109)

where qcon is the heat density for convection on the external surface, qtun is the heat flux for
evaporation inside the channels referred to the same surface.

The convective heat transfer process is assumed to have a heat flux qcon as follows:

(4.110)

where C, a, and b are empirical constants determined experimentally and N is the density of
nucleation centers, which can also be determined only by experimental means. The formula
expressing the evaporation heat flux takes on the form:

(4.111)

where mI and mII are liquid masses evaporating during the times τI and τII, which specify the time
span of phase I and phase II.

In order to determine the heat flux q in (4.109), Nakayama et al. [167, 168] established an
eight-step procedure, according to which the times τI and τII, the masses mI and mII , and also
the density of nucleation centers N should be calculated in proper order. The fundamental
shortcoming of the Nakayama et al. model is the necessity to determine seven empirically
determined constants for each of the investigated liquids (water, nitrogen, R-11). The model does
not account for the impact of the wetting angle of the fluid, and therefore its practical applicability
seems doubtful [314].

The model deals with only one kind of boiling termed “suction – evaporation”, which occurs for the
heat fluxes that are relatively high and for a range that is not precisely defined. The assumption
that the liquid microlayer occurs only in the corners of the internal tunnel, with the remaining
volume being filled with vapor, can be considered another major weakness of the model, which
was pointed out in critical comments expressed by Webb and Haider [314]. The process of bubble
formation inside the layer and the active pore distribution on its external surface are not accounted
for either. Nevertheless, detailed considerations given to the matter by Nakayama et al. [167, 168]
are extremely valuable as regarding the physics of the boiling process inside the porous covering.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 203


Chapter 4

Rannenberg and Beer model

Rannenberg and Beer [222] made an attempt to generalize the results of their experimental work
by modifying classic relationships used to compute heat transfer coefficients for boiling on smooth
surfaces. Their attempt proved to be in vain as it was impossible to predict such quantities as
the density of active nucleation centers, the frequency of bubble generation or the diameters of
departing vapor bubbles. As the vapor generation process inside a capillary-porous structure
cannot be observed, they claimed that the only available means of analysis is the application of
similarity theory. Vapor bubble generation and entrainment as well as the structure filling with
liquid were treated as a convective process in the liquid phase, with forced flow created by the
motion of the bubbles. They assumed that:

• the flow is stationary and incompressible;

• inertia and mass forces can be disregarded;

• no sources or sinks are present;

• the liquid thermal properties are constant.

For the above mentioned assumptions, a system of equations takes on the following form:

(4.112)

(4.113)

(4.114)

On the basis of classic similarity theory considerations, the following similarity numbers were
obtained after introducing scaling coefficients:

(4.115)

(4.116)

Hence, the Nusselt number can be written in the form:

Nu = Nu (Re Pr, Re Eu, L) (4.117)

204 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where the dimensionless number L describes geometric similarity. The Nusselt number Nu and
the hydraulic diameter Dh are defined by equations (3.57). The Reynolds number has the form:

(4.118)

where the characteristic velocity is computed on the basis of vapor flow rate (q/hlg) and the
structure porosity

(4.119)

The similarity number in (4.115) is further analyzed in a detailed manner, with the use of Darcy’s
law to calculate the pressure drop of the flow through a porous body (Re ≤ 1):

(4.120)

where K is permeability defined by the formula (3.57). Using Dh from (3.57) in (4.120) as the
characteristic dimension for computing pressure drop in the process of vapor bubble formation
and then inserting (4.120) to (4.115), the following relationship is obtained:

(4.121)

Subsequent insertion of the expression defining permeability K (3.57) into (4.121) leads to the
conclusion that the above product of similarity numbers is a constant, independent of the porous
coating structural parameters. It cannot be therefore a defining number, owing to which the
expression (4.117) yields the form:

Nu = Nu (Re Pr, L) (4.122)

In order to define the Prandtl and Reynolds numbers, the liquid thermal conductivity is used in the
above considerations. The ratios λef / λl and Deyt / δeyt, where λef defines the dependence (3.55),
are introduced to the resulting Nusselt number correlation to account for the capillary-porous
structure properties.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 205


Chapter 4

Though the product Re Eu was eliminated as a defining number, while correlating the dependence
of Nusselt number with respect to experimental data, it turned out that taking the product into
account leads to a better statistical fit, under the condition that the effective radius of the pore Ref
is used in place of the hydraulic diameter Dh, so that

(4.123)

Finally, the following expression is obtained using (3.57):

Its properties are discussed in section 3.2.4.

Xin and Chao models

The Xin and Chao [318] theoretical analysis focuses on boiling heat transfer in capillary-porous
structures that are characterized by tunnels with longitudinal internal connections. Two kinds of
industrially enhanced boiling surface geometries were analyzed: the Gewa-T and Thermoexcel-E
tubes. Basic theoretical investigations and experimental work concerned a flat plate version of
the Gewa-T surface. Analysis was restricted to a single repeatable cell of the enhancement, as is
shown in Fig. 4.21.

Fig. 4.21. Simplified model of reentrant channels and counter-current flow of vapor
and liquid for a flat Gewa-T surface of Xin and Chao [318]

Stationary evaporation of the liquid microlayer covering the internal surface of the T-shaped
channel is assumed. The vapor generated leaves the channel inside through the small opening as
shown in Fig. 4.21. T-shaped channels are repetitive grooves in the surface, filled with saturated

206 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

liquid and trapping vapor inside. The process of the reentrant channel filling with liquid and
generating vapor is assumed to have a continuous, steady-state counter-current flow of both liquid
and vapor. The model does not account for the cyclic character of bubble formation and departure,
i.e. their dimensions and frequency. It is also assumed that vapor generation proceeds over the
whole surface of the tunnel and, consequently, the density of nucleation centers does not affect
the boiling process. Though convenient for making a theoretical analysis, the above-mentioned
assumptions contradict the numerous observations of vapor nuclei formation, vapor bubble cyclic
growth and bubble departure from such enhanced boiling surfaces.

Like in many other models, it was assumed that the total heat flux conveyed by the surface to the
fluid has two components: a convective heat flux and an evaporative heat flux connected with thin
film evaporation inside the channel. The thickness of the liquid microlayer covering the channel in
Fig. 4.21 results from the dynamics of the counter-current two-phase flow and is given from [303]
to be:

m
δ C1 = C Rel

–2 1/3
C1 = [g (ρl − ρg) ρl μl ] (4.124)


Rel =
μl

where δ is the liquid microlayer thickness, Γ is the mass flux per channel length and C, C1 and
m are constants. It is further assumed that the slit height b is equal to half the channel width
(Fig. 4.21) so that b = 0.5 l, and that the tunnel bottom is dry and therefore inactive for vapor
generation. On such a basis, the microlayer penetration depth along the channel wall is expressed
by the dependence:

ξ H = ξ (h + 0.5 l) (4.125)

where h and l are tunnel height and width, respectively, and ξ is an empirical coefficient that
accounts for the impact of the vapor shear and liquid-vapor surface tension on the liquid film
thickness. The above-mentioned assumption is however contradictory to the results of the
investigation conducted by Arshad and Thome [15], who observed such channels being entirely
wetted around their perimeter.

On the basis of the relationships mentioned above, for an arbitrary position x from the bottom of
the channel, the mass flux is:

1/m 1/m
Γx = (C1/C) δx μl/4 (4.126)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 207


Chapter 4

An energy balance for a liquid microlayer of the length dx, which accounts for the conduction
through the liquid and the heat of evaporation, after taking into account the expression (4.126),
yields:

(4.127)

Integrating (4.127) for the initial boundary condition

x=0 δ = 0 (4.128)

the following is obtained:

(4.129)

Hence, the evaporative heat flux per channel length is the integral along the penetration depth of
the liquid microlayer:

(4.130)

In accordance with (4.126) and (4.130), the following expression is obtained:

(4.131)

where ΓξH describes the mass flux per channel length for x = ξ H.

As the liquid flows into the tunnel in only one direction and vapor can only depart from the
channel through the slit, the mass flux of liquid evaporating on both tunnel walls, as illustrated in
Fig. 4.21, is:

(4.132)

208 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where s is a pitch for internal channels of the Gewa-T surface and qev is the heat flux referred to
the porous surface projection onto a flat surface, i.e. that of the nominal flat surface. Inserting
(4.132) into (4.131) yields

(4.133)

where the dimensionless liquid microlayer thickness is

(4.134)

and Ar D is defined by the formula (3.6).

The impact of surface tension σlg and shearing stresses τ on the interfacial surface in this model
are accounted for by means of the empirical coefficient ξ. The vapor shear affects both the
thickness and penetration depth of the liquid microlayer down the wall, and thus all are directly
dependent on the vapor velocity and, consequently, the heat flux. That is, the higher is the vapor
velocity, the smaller is the penetration depth and the higher is the heat flux. The conclusions are
confirmed experimentally.

The vapor shear and surface tension induce the liquid to flow into the corners of the channel,
where they cause an increase in the liquid microlayer thickness. Their combined effect is
described by the relation

(4.135)

whereas the vapor shear stress can be computed using the well known expression [303]:

(4.136)

where f is the friction coefficient and ug is vapor velocity in the slit, which can be determined on
the basis of an energy balance to be:

(4.137)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 209


Chapter 4

Hence:

(4.138)

Inserting (4.135) and (4.138) into (4.133) leads to the relation:

(4.139)

where the Weber number is defined by the formula (3.8).

Experimental results quoted in [318] indicate that the total heat flux is approximately linearly
proportional to the evaporative heat flux through the Prandtl number for the liquid (3.9):

p
q ~ qev Prl

which when introduced into (4.139) yields:

(4.140)

or

(4.141)

'
Constants C 4, C4, m, n, p, m1, n1 and p1 were then determined experimentally, and the similarity
numbers Nu, Ar, Rel, We and Pr l are defined by equations (3.5), (3.6), (3.7), (3.8) and (3.10),
respectively. Equations (3.4), (3.11) and (3.12) show the results of a regression analysis with
respect to experimental data for the industrial surface Gewa-T.

As already mentioned before, Xin and Chao [318] extended their model to cover the
Thermoexcel-E enhancement geometry. Characteristic dimensions of this enhance boiling surface
were described earlier when deriving its equation (3.16). Like in the above Gewa-T model, the
liquid inflow was assumed to be continuous, steady-state and counter-current whereas the vapor
was thought to flow out of a single pore opening when breaking the reentrant channel into

210 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

individual cells, each with its own pore opening to the outside. The vapor velocity was obtained
from an energy balance like before:

(4.142)

Now the characteristic dimension for the corrective coefficient ξ is the pore diameter, hence

(4.143)

or after applying (4.136) and (4.142):

(4.144)

Proceeding as before, the empirical constants were determined experimentally with respect to the
data of Nakayama et al. [167, 168], obtaining the correlation (3.16).

Kovalev et al. model

The model of Kovalev et al. [116 – 130, 186, 201, 258 – 260] further developed the hydrodynamic
theory of evaporation in a capillary-porous structure put forward by Krochin and Kulikov [136]. It
differs to a great extent from the static models proposed by Man’kovski et al. [155], Xin and Chao
[315] and O’Neill et al. [68, 91, 182]. The above-mentioned models assumed static liquid and
vapor inside the enhancement structure, which is hardly consistent with the physics of the boiling
process. The dynamics of vapor bubble growth, i.e. phenomena such as bubble departure, its
frequency and nucleation center density, were disregarded in those studies.

Two basic features differentiate the Kovalev et al. model [116-130, 186, 201, 258 – 260] from the
former ones, namely:

• evaporation from interfacial surfaces of liquid-vapor menisci inside the porous


capillary structure;

• the dynamics of counter-current liquid and vapor flow in the capillaries of the structure
are accounted for.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 211


Chapter 4

The capillary-porous structure is assumed to be a system of vertical and horizontal capillaries,


as illustrated in Fig. 4.22a. The following assumptions were made in order to derive equations
describing the liquid and vapor motion and heat transport inside the layer:

1. Evaporation inside the structure is stationary and stable.

2. The liquid penetrates the structure inside through small pores (capillaries), whereas
vapor escapes through big pores, Fig. 4.22a.

3. The liquid ideally wets the material of the porous capillary structure, i.e. the wetting
angle equals zero.

4. Heat is transferred by means of conduction exclusively by the structure’s skeleton


(λskl >> λl) to interfacial menisci formed at the sites where vertical and horizontal
capillaries are connected, as per Fig. 4.22a. Menisci are heat sinks whose capacities
are described by the volumetric heat transfer coefficient αvol.

5. The skeleton temperature changes only in the direction normal to the heating surface,
across the porous layer thickness.

6. Vapor moves from the heating surface towards the liquid volume and its pressure
decreases along the porous layer thickness, as shown in Fig. 4.22b. The pressure
of the liquid entering the layer is equal to the saturation pressure. The pressure of
the vapor leaving the layer is equal to the pressure in the vapor bubbles and is thus
slightly higher than the saturation pressure.

7. The difference between the liquid and vapor pressures results from the action of
surface tension forces 2 σlg / Rlmt, where Rlmt is the meniscus radius that varies across
the porous layer thickness. Like in Tehver et al. [199, 282, 284], all pores (capillaries)
of the radius R < Rg are filled with the liquid and those where R > Rlmt are filled with
vapor, as illustrated in Fig. 4.22c.

The maximum possible difference in vapor–liquid pressures pg − pl | max corresponds to the


minimum pore diameter and, at the same time, specifies the maximum depth of liquid suction into
the porous layer. Rg reaches its maximum at the external surface and its value decreases with
depth in the porous layer. For high heat fluxes (close to the critical heat flux), the skeleton layer
adjacent to the heating surface is dry (filled with vapor) and Rlmt = Rmin.

212 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Equation of motion

In accordance with assumption 7, for each section x-x, the boundary meniscus radius results from
the equation:

pg(x) − pl(x) = 2 σlg / Rlmt(x) (4.145)

and sections of capillaries are functions of Rlmt(x). Filtration equations for both phases should,
therefore, account for acceleration of the liquid and vapor streams [136]:

(4.146)

where β is the porous layer inertia coefficient, and K is the given phase permeability. Writing
equations (4.145) and (4.146) jointly, while the inertia term for the liquid is disregarded, yields:

(4.147)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 213


Chapter 4

a) diagram of liquid and vapor circulation in the structure; b) pressure


distributions in both phases; c) distribution of liquid and vapor in capillaries in
the selected structure section; d) temperature distribution

Fig. 4.22. Model of nucleate boiling inside a porous capillary structure by Kovalev
et al. [116 – 130, 186, 201, 258, 259]

Velocities of both phases are determined from the equation of continuity for mass flow rate
G(x) > 0 to be:

(4.148)

where ∏(Rlmt) is the integral of the pore diameter distribution γ(R). In contrast to the
considerations by Krochin and Kulikov [136], the change in flow rate along the capillary length is
taken into account in equations (4.148).

214 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

According to filtration theory [45, 270], the porous body’s permeability is expressed by the
dependence:

(4.149)

where γ(R) = d∏ / dR is the function of pore diameter distribution, and ∏(R) is its integral

(4.150)

0.7
Π (R)
0.6

0.5 0.02

γ (R), mm-1
Π (R)

0.4 γ (R)

0.3 0.01
0.2
100 80 60 40 20 0
R, mm

Fig. 4.23. Function of pore diameter distribution γ (R) and its integral ∏ (R), sintered
stainless steel powder, δ = 1.1 mm, Dgrn = 100 µm, from Kovalev and Lenkov
[117, 118]

Fig. 4.23 shows the function of pore diameter distribution γ(R) and its integral ∏(R) for a selected
capillary-porous structure made by sintering stainless steel powder. The permeability of each
phase after eliminating the constant C with the use of (4.149) is written as follows:

(4.151)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 215


Chapter 4

Inserting (4.148) and (4.151) into (4.147) leads to the equation describing the dynamics of motion of
both phases in boiling in the capillary-porous structure:

(4.152)

where

(4.153)

In order to solve (4.152), it is necessary to know the boundary condition for the external surface
of the porous covering x = δ, Rlmt = Rlmt(δ) and the function G(x). The value of the mass rate flow
of vapor G(x) is determined on the basis of the solution of heat transfer problem inside the porous
covering.

The boundary condition for expression (4.152) is determined as follows. The liquid pressure for
x = δ, see Fig. 4.22b, is equal to the saturation pressure, which after introducing into (4.145) yields:

–1
pg (δ) − psat = 2 σlg (Rlmt (δ)) (4.154)

The pressure drop (4.154) is the sum of losses resulting from the dynamics of bubble motion and
the pressure increment necessary for liquid volume penetration [45]:

2
pg (δ) − psat = ρg wg (δ) + ∆pbd (4.155)

Combining (4.154) and (4.155), it is possible to determine Rgr(δ):

(4.156)

216 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Computation of volumetric heat transfer coefficients

The one-dimensional equation of heat conduction in the porous capillary structure’s skeleton takes
on the form:

(4.157)

where qvol (x) is heat generation per unit volume of the internal heat sources and αvol is the
volumetric heat transfer coefficient. The necessity of accounting for qvol results from the fact that in
experimental investigations the structure is heated by directly heating it with electric current. While
writing (4.157), it was assumed that pg – psat << psat, and therefore it is possible to assume that Tsat
has a constant value. For the sake of computation of the volumetric heat transfer coefficients, heat
transfer in the structure’s capillaries is thought to proceed as shown in Fig. 4.24.

In accordance with earlier considerations, pores of R > Rgr are filled with vapor, whereas those of
R < Rgr are filled with liquid. The amount of heat stored due to evaporation between sections x and
x + dx amounts to, Fig. 4.24:

dQ x,x+dx (x) = F dx αg,vol (Tskl − Tsat) (4.158)

In their work [259], Solovev and Kovalev made detailed theoretical investigations on the amount of
heat transferred to the liquid in evaporation from the meniscus of the radius R* formed in the pore
of the radius R:

(4.159)

where αcph is heat transfer coefficient for the change of phase and Ref,σ is the effective radius of
the action of surface forces. In the case under consideration, R* = Rlmt.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 217


Chapter 4

a) section of a capillary-porous structure; b) developed view of vapor channel


(menisci are blackened)

Fig. 4.24. Model for computation of volumetric heat transfer coefficient of the vapor
αg, vol from Kovalev, Solovev and Ovodkov [127]

The amount of heat supplied to the liquid due to evaporation from the surfaces of the outlets of the
capillaries of radii from Rmin to Rlmt, see Fig. 4.24, distributed on the lateral surface of the radius
R > Rlmt, is equal to:

(4.160)

The summation of the heat (4.160) transferred by evaporation from menisci on lateral surfaces of
the capillaries filled with vapor in all these capillaries leads to the relationship:

(4.161)

After integrating (4.161) and inserting the result into (4.158), the following expression is obtained:

(4.162)

For Rlmt → Rmax and Rlmt → Rmin, αg,vol → 0, which corresponds to the porous structure being
totally filled with vapor (high heat flux q and temperature difference ΔTsl) or completely filled with
liquid (low heat flux and temperature difference).

218 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Continuing with the analysis, the coefficient αl, vol describes the convective heat transfer in the
capillaries of radius R < Rlmt. Superheating of the liquid moving inside the porous covering takes
place in accordance with the equation:

(4.163)

where

x=δ Tl = Tsat (4.164)

For stabilized laminar flow in a round pipe, which begins at the pipe inlet for Peclet numbers
approaching zero,

(4.165)

(4.166)

and boundary conditions of the Ist kind, the heat transfer coefficient for the liquid can be
determined using the dependence [199] to be:

–1
αl = 3.66 λl dpipe (4.167)

It should be mentioned that the forms of Peclet number definitions given in [122, 124, 129] are
erroneous.

The volumetric heat transfer coefficient αg,vol is determined on the basis of the balance of the heat
transferred from the capillary of the radius R to the liquid

dQl = αl 2 Π R dx (Tskl – Tl) (4.168)

which after summation of all capillaries filled with the liquid and taking into account (4.167) gives:

(4.169)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 219


Chapter 4

Hence,

(4.170)

The relationship between the equation of motion (4.152) and the heat conduction equation (4.157)
is given by the expression describing the vapor mass flux:

(4.171)

Equation (4.171) is based on the obvious physical fact that the vapor flow rate through the
section x-x as is that generated across the segment from the capillary-porous structure base to
the height x.

Boundary conditions for (4.157) are as follows:

(4.172)

Mathematical model of boiling heat transfer on the surface with a porous coating includes the
following equations: the equation of motion (4.152) with the boundary condition (4.156), the
heat conduction equation (4.157) with boundary conditions (4.172) and the expressions for the
volumetric heat transfer coefficients (4.162) and (4.170), and also the coupling equation for vapor
mass flux (4.171).

In accordance with the model presented in Fig. 4.22, counter-current flows of liquid and vapor
occur inside the porous layer. The liquid penetrates the layer inside from the pool volume, flowing
to the base surface whereas vapor generated in pores is removed through big pores. Ovodkov
and Kovalev [119, 186] built a special test stand to investigate counter-current flows of both phases
through porous capillary structures. They used water and nitrogen and also sintered powder
structures of the same parameters as those in boiling investigations. The diagrams Δp – G,
obtained experimentally, were applied to determine the distribution functions of γ(R) and ∏(R),
that are indispensable for solving the equation of motion (4.152) and the heat conduction equation
(4.157).

According to Kovalev and Solovev [121 – 124, 126 – 128], the heat flux transferred by convection
through the liquid can be neglected in the equation (4.157). Thus it is possible to assume that
αl, vol = 0, owing to which the heat conduction equation is simplified and the equation (4.163)
disappears.

220 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Figs 4.25 and 4.26 show the results of experimental investigations and numerical calculations.
Fig. 4.25 indicates good agreement between the experimental and numerically computed boiling
curves q = q(ΔTsl) for selected experimental data. It is assumed that ΔTsl = Tskl,0 – Tl, where
Tskl,0 = Ts, Tl = Tsat.

6
4⋅10

106
8

6
q, W⋅m-2

105 1
8 2
6
3
4

2⋅104 0
10 2 4 6 8 101 2 4 6 8 102 2 2
4⋅10
DTsl, K

Parameter 1 2 3
δ, mm 1 0.45
ε 6.4 0.4 Numerical
computation
Dgrn, µm 400 - 615 100 - 200

Fig. 4.25. Comparison of computed results to experimental data for water at


atmospheric pressure for sintered stainless steel powder layer from Kovalev
and coworkers [122, 127, 128]

Fig. 4.26 presents the results of the numerical simulations. The comparison of curves 2 and
4 implies that the decrease in the porous coating permeability enhances heat transfer for
small values of q and decreases it at large heat fluxes. That results from the relationship
αg, vol = ƒ(q), which was discussed earlier. An increase in the skeleton conductivity (curves 1 – 3)
causes an increase in the heat flux q in the whole range of nucleate boiling. An increase in the
layer thickness (curves 2 and 5) leads to a higher heat flux dissipation in the range of its smaller
values, whereas in the range of high values, it has an unfavorable effect.

Kovalev and Solovev [122, 129] claimed that all coefficients and constants that occur repeatedly
in the set of equations: (4.152), (4.156), (4.157), (4.163), (4.170) and (4.172) can be determined
experimentally or found in publications presenting experimental data. It is, however, reasonable

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 221


Chapter 4

to voice the opinion expressed by Webb [311] that the authors do not provide sufficiently accurate
and detailed definitions of many variables used in the model’s equations nor clear procedures to
calculate such quantities as the departing vapor bubble diameter Ddep, the heat transfer coefficient
on the surface of the evaporating meniscus αf or the effective radius of surface forces acting Ref,σ.
No clue, even a general one, is given as regards to the application of the presented numerical
procedure.

6
4⋅10
2 1
6 2
10
6
q, W⋅m-2

2 5

10 5 4
6
3
4

2⋅10 4 -1 2 4 6 0 2 4 6 10 1 2 4 6 2 2 2
10 10 10 4⋅10
DT sl, K

Parameter 1 2 3 4 5
–1 –1
λ, Wm K 20 7.29 2.0 7.29 7.29
12 2
K ∙ 10 , m 1.35 1.35 1.35 0.1 1.35
δ, mm 0.45 0.45 0.45 0.45 1.0

Fig. 4.26. Results of numerical simulations designed to investigate the impact of


porous layer structural characteristics on boiling heat transfer from Kovalev
et al. [122, 127, 128]

The presented comments indicate that it is quite difficult for researchers to apply the Kovalev
and Solovev method of modeling boiling heat transfer in the porous capillary structures to their
calculations aimed at comparing numerically obtained boiling curves with a larger number of
various experimental data.

Nevertheless, the model built by Kovalev, Solovev and coworkers [116 – 129, 186, 201, 258,
259] still makes an important contribution to the understanding of the complex process of heat
transfer and also that of liquid and vapor flows in boiling in porous capillary structures. A decided
advantage of the approach presented by the authors is the possibility of predicting not only the
shape of the boiling curve q = q(ΔTsl) as a function of the liquid thermal properties and the porous
layer structural parameters but also the critical heat flux for nucleate boiling.

222 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Ayub and Bergles model

The geometric similarity of the Thermoexcel-E and Gewa-T enhanced boiling surfaces (both
structures have regular horizontal channels inside, one with pores connecting the channels to the
outside pool of liquid and the second with a continuous slit) encouraged Ayub and Bergles [20] to
adopt the concept of Nakayama et al. [167, 168] while building their model of boiling heat transfer
on the industrially manufactured surface Gewa-T. Similar to Nakayama et al. [167, 168], they
assumed that two components qev and qcon, described by the equation (4.109), can be identified
as contributing to the total heat flux. The basic simplifying assumptions of their model are as
follows:

• the formation and development of a vapor bubble is a continuous process and it is not
necessary to distinguish three different development phases, like in Nakayama et al.
[167, 168];

• to describe the heat transfer process, it is sufficient to rely only on an energy balance
taking into account two components of the heat flux: qev and qcon.

a) boiling model; b) the section of channel and nucleation center on such a surface

Fig. 4.27. Gewa-T microsurface structure and model of Ayub and Bergles [20]

A simple model of vapor bubble generation is assumed, namely that in the horizontal channel,
near some of the pores, and since the vapor pressure exceeds the pool liquid pressure, this
causes a deformation of the internal liquid film phase, as shown in Fig. 4.27a. Such pores become
nucleation centers, while the remaining pores in the surface are inactive. Vapor flowing out of the
active pores in departing vapor bubbles causes a slight lowering of the liquid meniscus at inactive
pores towards the inside of the channel and causes external liquid to flow into the channel. The
liquid spreads over the internal surface of the channel, for example as in Fig. 4.27b, and starts
evaporating. After bubble departure, the interfacial surface becomes flat again and the whole
cycle is repeated.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 223


Chapter 4

The heat balance for the channel in which evaporation occurs is as follows:

(4.173)

which can also be written

Ql = λl Cev ∆T (4.174)

In the expressions above, Fcap is the channel’s total surface area, while Fmicro and δmicro are the
liquid microlayer surface area and its thickness, respectively, in Fig. 4.27b. The constant Cev is
the ratio of the surface area to the thickness of the liquid microlayer covering the channel interior.
In equations (4.173) and (4.174), both the liquid and vapor are assumed to be at the saturation
temperature while the wall temperature is constant and uniform. In accordance with (4.173) and
(4.174), the evaporation heat flux for the liquid microlayer inside the channels amounts to:

qev = λl Cev ∆T/F (4.175)

where F is the heating surface.

Like in Zuber [334] and Nakayama et al. [167, 168], the convective heat flux is assumed to be the
same as in boiling on the flat smooth surface:

1/y –x/y
qcon = (∆T/Ccon) (N/F) (4.176)

where x = –1/5 and y = 3/5. The density of nucleation centers for both investigated liquids, namely,
water and R-113, are similar in their experimental investigations and the constant is assumed to be
–0.8 –3/5 4/5
Ccon = 3 ∙ 10 KW m .

The total density heat flux is thus:

1/y –x/y
q = λl Cev ∆T/F + (∆T/Ccon) (N/F) (4.177)

224 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where free convection is assumed to proceed over the entire heating surface. Ayub and Bergles
experimentally determined the density of nucleation centers to be a function of the heating
surface’s superheating, separately for each of the investigated liquids:

for water
2 4
N = (–42.94 + 40.96 ∆T – 2.53 ∆T ) 10
(4.178)
for R-113
2 4
N = (–32.13 + 20.25 ∆T + 0.85 ∆T ) 10

The constant Cev is estimated for an arbitrarily selected Gewa-T 19D/W microsurface and
4 –2
experimental conditions: ΔT = 4K, q = 4.5 ∙ 10 Wm . Its value for water is Cev = 10.12779 m. The
constant Cev, expressing the ratio of the surface area to the liquid microlayer thickness, strongly
depends on the value of the surface tension σlg. Hence, Cev for R-113 is obtained by multiplying
its value for water by the quotient σR-113/σwater, which yields Cev = 32.14 m. After accounting for
both constants Ccon and Cev and also expression (4.178), the values of the heat flux obtained from
equation (4.177) are in agreement with the experimental data for water, with the error of ±28% and
for R-113 with an error of ±23%.

Ayub and Bergles, however, did not describe their method to determine the empirical expressions
in (4.178) or the functional dependency of Ccon and Cev on the liquid thermal properties, the
enhancement geometry and its material properties. Hence, their equation (4.177) is only
applicable to the liquids and the surface used in the authors’ experiment. Nevertheless, their
model illustrates a simple relationship between the boiling process and the two heat transfer
mechanisms.

Kravchenko and Ostrovski model

The feature shared by the models by Smirnov et al. [247, 254], Xin and Chao [318], Wang et al.
[304], Ayub and Bergles [20] and also Kravchenko and Ostrovski [133] is the common assumption
that for porous layers that conduct heat well (metallic ones), the liquid microlayer covering the
surfaces of microfins [247, 254], internal horizontal channels [20, 318] or capillaries [133] provides
the main thermal resistance, Fig. 4.28.

As in Xin and Chao [318], it is assumed by Kravchenko and Ostrovski [133] that counter-current,
steady-state liquid inflow and vapor outflow are determined by the counterbalance of the surface
tension versus the vapor shear stress (hydraulic resistance), largely exceeding gravitational
forces. In the steady-state, the capillary in which vapor is generated, as shown in Fig. 4.28,
functions as a heat pipe. Vapor flows along its center whereas liquid flows down over the capillary
wall. Furthermore, the liquid microlayer thickness δl depends on the capillary diameter, the liquid
thermal properties and the general conditions of the boiling process that are not specified.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 225


Chapter 4

Fig. 4.28. Model of counter-current vapor and liquid flow inside the capillary in the
porous layer from Kravchenko and Ostrovski [133]

Heat is transferred by conduction to the evaporating surface of the flowing, cylindrical liquid
microlayer. The heat transferred through any dry surface at the capillary base and the heat
transferred from the external surface of the capillary are both disregarded in the considerations.
The liquid and the vapor flows are assumed to be laminar, forced by capillary pull-up. Two cases
of temperature difference between the capillary wall and flowing vapor are distinguished:
a) constant, ΔTsl = const; b) linearly decreasing along the capillary height, ΔTsl = ΔT0 (1 – z/h).
Having thus accounted for the assumptions discussed above, it is possible to write three equations
describing the heat dissipation as well as liquid and vapor transportation in the capillary.

First of all, the heat flux passing through the conductive cylindrical liquid microlayer of thickness δl,
as indicated in Fig. 4.28, can be written as follows:

(4.179)

The introduction of the definition of dryness level (essentially the cross-sectional vapor void
fraction) for a capillary to (4.179)

(4.180)

226 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

and then integrating the equation with respect to the height h for the two previously mentioned
characteristic temperature differences ΔT yields:

(4.181)

(4.182)

where φ is the mean dryness level over the segment [0, z].

The heat flux conducted through the liquid microlayer evaporates liquid on the phase interface in
accordance with the equation:

(4.183)

Combining (4.183) with (4.181) and (4.182), two vapor velocities in the capillary are thus derived for
these two conditions:

(4.184)

Applying a mass balance to the liquid and vapor flows in the capillary results in:

(4.185)

and hence the relationship between the liquid and vapor velocities is established to be:

(4.186)

The capillary pressure increment, induced by the action of surface tension, causes the motion of
liquid and vapor in the capillary, which yields:

∆pcap = ∆pg + ∆pl (4.187)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 227


Chapter 4

The capillary pressure increment can be approximately estimated in the following way:

(4.188)

The pressure gradients in the liquid and the vapor are described by the equations:

(4.189)

(4.190)

Integrals of the equations (4.189) and (4.190), for the two characteristic temperature differences
ΔT, lead to the following expressions:

(4.191)

∆Tsl = const, Cl,2 = 16, Cl,1 = 24

∆Tsl = ∆T0 (1 – z/h), Cg,2 = 63/3, Cl,2 = 32

where wl and wg are mean velocities of the liquid and the vapor at the mouth of the capillary.

By inserting (4.191) into (4.187) and accounting for the two solutions given in (4.184), the final form
of the equation of motion is obtained:

(4.192)

ΔTsl = const, C1 = 1/64 ΔTsl = ΔT0 (1 – z/h), C2 = 3/128

The introduction of the Archimedes (3.6), Prandtl (3.10) and Jakob (3.13) similarity numbers and
the capillary constant k (3.20) to (4.192) makes it possible to write the equation in the form:

(4.193)

228 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where

~ ~
v = vg /vl k = k/Dcap
~ ~
ρ = ρ /ρ
g l h = h/Dcap
~2 ~2 ~ –1
K = Pr Ar k (Ja h p)

The equation (4.193) connects the liquid microlayer thickness and thermal properties with the
capillary dimensions and temperature difference along the capillary height. This temperature
difference depends both on the porous layer’s geometric parameters, its thermal properties and
also the heat flux.

Both sides of the equation (4.193) can be rewritten as:

(4.194)

owing to which it possible to show that (4.193) can have: (a) no solution; (b) one solution and
(c) two solutions, which are illustrated in Fig. 4.29. Case (a) means that the capillary pressure
increment is too small to fill the capillary with the liquid. In case (b), the capillary pressure
increment is exactly equal to the hydraulic resistance of the liquid and vapor flow. As regards case
(c), a more accurate analysis can show that the liquid microlayer is stable and it corresponds to
the minimum value of φ in Fig. 4.29.

η
γ η

c
b
a

ϕc1 ϕb ϕc 2 1 ϕ

Fig. 4.29. Graphs of auxiliary functions η(φ) and γ(φ) from Kravchenko and
Ostrovski [133]

According to Kravchenko and Ostrovski [133], the heat flux in the capillary reaches a maximum
for a constantly wetted capillary when the dryness level φ also reaches its maximum, which
corresponds to case (b) in Fig. 4.29. This means that if other structural and boiling process
parameters are set, it is possible to determine the optimum capillary diameter, for which the heat
flux transferred in the capillary is the greatest. However, their theoretical results were unfortunately

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 229


Chapter 4

not checked against any experimental data. A further weak point of Kravchenko and Ostrovski’s
investigation is the assumption that the liquid and vapor flow in the capillary is continuous and
steady, which is inconsistent with the cyclic process of vapor bubbles growth and departure.

Wang et al. model

Wang et al. [304] extended Xin and Chao’s concept [318] to cover cylindrical surfaces with the
industrially made Gewa-T type of surface. A sectional view of the Gewa-T surface analyzed is
shown in Fig. 4.30.

Fig. 4.30. Section view along the axis of a tube with a Gewa-T type of surface
analyzed by Wang et al. [304]

Like in Xin and Chao [318], it was assumed that the heat conducted into the liquid microlayer
covering the internal channels, as shown earlier in Fig. 4.21, is carried away to the external liquid
pool by two vapor streams as in illustrated in Fig. 4.31. In their steady-state developed boiling
process, they divided the vapor flow as follows in Fig. 4.31:

a. vapor stream I inside the channel around the circumference of the tube towards
the top;

b. vapor stream II leaving the channel locally through the slit and moving upwards
around the pipe circumference.

The vapor streams I and II then join together on the upper surface of the horizontal tube as
indicated in Fig. 4.31.

230 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Fig. 4.31. Model of the liquid microlayer evaporation and two vapor streams flowing
along the horizontal cylindrical Gewa-T heating surface of Wang et al. [304]

Following Xin and Chao [318], the authors assumed that vapor stream I and the liquid microlayer
move with respect to each other in a single porous cell like in Fig. 4.32. The following assumptions
are thought to be satisfied:

1. the fluid flow and heat transfer proceed in a steady-state;

2. the liquid and vapor have constant thermal properties;

3. the liquid level in the vessel in which the Gewa-T tube is immersed is constant;

4. the liquid microlayer acts as a Poiseuille flow with linear distributions of velocity and
temperature.

Fig. 4.32. Vapor stream and liquid microlayer flow of Wang et al. [304]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 231


Chapter 4

Fig. 4.33. Unit control volume for conservation equations analysis of Wang et al. [304]

In order to analyze this process, an elementary cell is adopted as depicted in Fig. 4.33. First of all,
a momentum balance per unit channel length can be written as follows:

(4.195)

with the assumption that the thickness δg and mass flow rate M are constant with respect to the
angle β. Next, the vapor flow is assumed to be laminar, on which basis, following Burmeister [38],
it is assumed that

(4.196)

The static pressure outside the vapor layer is given by:

(4.197)

The mean vapor velocity over the segment from the section 1-1 to the section 2-2 in Fig. 4.33 can
be computed from the equation:

(4.198)

232 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where Δx is the tube segment length in the direction normal to the surface as in Fig. 4.33. The
relationship between the thicknesses of the liquid and vapor layers is as follows:

n
δl ~ f (β) δg (4.199)

where n is a constant, whereas the function f (β) is of complex character and makes δ l decrease
with increases in δg and β. Then, from an energy balance, the following is obtained:

(4.200)

where ζ is the partition coefficient specifying the fraction of the heat flux which is carried away by
the vapor stream I in Fig.4.31.

Combining equations (4.195) to (4.200), together with integration with respect to x, it is possible to
determine the vapor stream flow rate M to be:

(4.201)

where C1 is a constant.

The mean value of the heat transfer coefficient is determined from the heat transfer balance:

M hlg = αl π dtube L ∆Tsl (4.202)

where L is the tube length. Similar to condensation on a horizontal tube, the Nusselt number can
be written as:

(4.203)

Dimensional analysis indicates that equation (4.203) is dimensionless for n = 1, and hence

(4.204)

–1
The three expressions on the right side of (4.204) are the similarity numbers Ja , Pr l and Ar (see
(3.6) and (3.13)).

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 233


Chapter 4

Observations show that vapor stream II flow over the external surface of the Gewa-T tube is
similar to flow around a smooth tube. Vapor bubbles flowing around the external surface stimulate
turbulence and enhance heat transfer. Thus, the Nusselt number for stream II is proportional to
the Prandtl and Reynolds numbers as:

NuII ~ Pr l, Re (4.205)

where the Reynolds number is defined by the formula (3.13).

The partition coefficient ζ (4.200) depends on the liquid microlayer shape and evaporation rate
inside the channel. The liquid flows into the channel through the slit opening due to the action
of capillary forces – σlg /D. At the same time, vapor leaves the channel and its shear τ in the slit
prevents the channel from being flooded with the liquid.

Equation (4.138), which results from the analysis of Xin and Chao [318], can be written in the form
of the heat flux as:

(4.206)

The heat flux here refers to the tube base surface, from which heat is conducted through the liquid
microlayer, and therefore the area ratio 2 (h + l)/s must be introduced into the expression above,
which finally gives the Weber number:

(4.207)

The conclusion that can be drawn from the present discussion is that the complete expression for
the Nusselt number involves the following similarity numbers:

Nu = f (Ja, Pr l, Ar, Re, We) (4.208)

Detailed definitions of similarity numbers are given by the formulae (3.13). Then, correlating
(4.208) with respect to experimental data, the detailed correlation (3.14) is obtained, for which the
applicability range was also determined.

In summary, the Wang et al. model [304] overcomes some of the shortcomings from their
predecessor’s work [318] with respect to neglecting the fact that vapor bubbles are generated and
depart in cycles.

234 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Webb and Haider model

Webb and Haider [309, 311, 314] aimed at working out an entirely analytical model for boiling
heat transfer for all types of integral enhancements and surfaces with porous coatings in which
the principal geometric structure is subsurface tunnels, e.g. Thermoexcel-E. In Webb and
Haider [314], their model explains the boiling process inside the capillary-porous structures to
be controlled by “flooding”, as shown in Fig. 4.19. The model does not contain experimental
constants, except for constants in their formula for departing bubble diameter [167, 168]. It is
assumed in the model that the tunnels inside the enhancement layer are alternately filled with
liquid or vapor. Following Nakayama et al. [167, 168], they assumed that the total heat flux is the
sum of the evaporation heat flux inside the tunnels and that of heat transfer by convection and
conduction to the single-phase area at the external (upper) surface of the enhancement:

q = qtun + qext (4.209)

Fig. 4.34. Geometry of the enhancement structure with round tunnels


analyzed by Webb and Haider [314]

Round tunnels, like those in Fig. 4.34, are regarded as the reference geometry, though the heat
transfer model put forward below can also refer to rectangular tunnels. Characteristic dimensions,
such as vertical slit width Dp (pore dimension), vertical slit height b (fin thickness at the external
surface of the capillary-porous structure), transverse slit pitch stun (tunnel pitch) and horizontal
tunnel radius Rtun are given in Fig. 4.34. Fig. 4.35 shows the hypothetical boiling process they
assumed in the tunnel and liquid pool. The notation refers to: Dg – departing vapor bubble
diameter, sg – longitudinal pitch of vapor lock in the tunnel, Rg,tun – radius of curvature of the
meniscus of the vapor lock in the tunnel, Ts – the heating surface (base) temperature and
Ts1 – temperature of the external surface of the enhancement.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 235


Chapter 4

Fig. 4.35. Model of boiling process in reentrant tunnels of Webb and Haider [314]

The evaporation of the tunnel-filling liquid takes place on surfaces of the menisci restraining vapor
locks as in Fig. 4.35. This in turn causes a pressure increase in the tunnel and the menisci shift.
The heating surface temperature Ts and the saturation temperature of the boiling liquid Tsat are
assumed to be given. The preset temperature difference ΔTsc is written in the form of the sum of
three temperature differences:

ΔTsl = Ts – Tsat = (Ts – Tl,tun) + (Tl,tun – Tsat,tun) + (Tsat,tun – Tsat) (4.210)

In this expression, Tsat,tun – Tsat is the superheating indispensable for vapor bubble formation in the
slit of the diameter Dp:

(4.211)

236 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where the derivative dp/dT can be computed using Clausius-Clapeyron equation [102].
Tl,tun – Tsat,tun is the superheating indispensable for maintaining the liquid vapor meniscus inside
the tunnel:

(4.212)

where

(4.213)

According to Webb and Haider [314] both temperature differences (4.212) and (4.213) are small
when compared with the difference Ts – Tsat.

If Qm is the heat flux that causes evaporation on the meniscus surface, the heat flux transferred by
the liquid in the tunnel amounts to:

(4.214)

where stun is the tunnel pitch and sg is the pitch between active pores along the slit, the latter
which should be determined on the basis of considerations on boiling in the tunnel.

While computing the heat flow rate Qms, it is assumed that:

a. the position of liquid-vapor meniscus is fixed and that heat is transferred by means of
steady-state conduction;

b. conduction through the liquid layer is to be disregarded when compared with


conduction in the radial direction near the menisci;

c. heat transfer to the liquid in the tunnel takes place only near the menisci as in
Fig. 4.36.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 237


Chapter 4

Each meniscus is “sliced into indefinitely small pieces” in Fig. 4.36 where Ts and Tl,tun are
boundary temperatures of the ring, for which the heat conduction equation is integrated. After
summation over the whole meniscus area, the following is obtained:

Qms = 2 � Rtun λl (Ts – Tl,tun) S(θ) (4.215)

(4.216)

where θ is the wetting angle.

Fig. 4.36. Heat transfer between the enhancement structure and the liquid in
the tunnel from Webb and Haider [314]

Vapor formation on surfaces of vapor lock menisci in a single cycle of bubble generation is:

(4.217)

where τ is the time of vapor bubble formation. Hence

τ = 1/ω (4.218)

238 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

and ω is the frequency of bubble departure. It is assumed to equal the frequency of shifts of the
location of vapor lock menisci.

The mass of vapor contained in a single departing bubble of diameter Dg amounts to:

(4.219)

Inserting (4.217) and (4.218) into (4.219), the frequency ω expression is obtained:

(4.220)

As (4.214) indicates, the density of active nucleation centers is:

(4.221)

The longitudinal pitch of vapor locks along a tunnel is the sum of three quantities:

sg = Ll + Lg,min + δam (4.222)

where Ll is the liquid lock length in the tunnel, Lg,min is the minimum length of the vapor lock
contact with the heating surface and δam is the meniscus oscillation amplitude. The amplitude δam
is computed from the equation of liquid lock motion along the current line, prior to which numerous
simplifying assumptions are made. Consequently, the following is obtained [314]:

(4.223)

where Re0 and Re0,m are dimensionless real-valued Bessel functions and Im 0 and Im0,m are
dimensionless imaginary Bessel functions while Δpmax is the maximum pressure drop along the
liquid lock that induces vapor bubble motion.

Following the same considerations of Nakayama et al. [167, 168], the maximum pressure
difference is estimated as follows:

(4.224)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 239


Chapter 4

In accordance with (4.219) and (4.220), the maximum increment in vapor mass in the process of
vapor lock volume increase is:

(4.225)

The corresponding to mg,min volume of the vapor lock is computed by summing the volume of
vapor lock along the contact with the heating surface xg,min with the volume of two spheroids
corresponding to the menisci:

2 2 3 –3 2
Vg,min = π Rtun Lg,min + π Rtun (cos θ) (2 + sin θ) (1 – sin θ) (4.226)
3

where Lg,min remains an unknown quantity. The maximum vapor volume equals the previously
calculated volume increased by the volume corresponding to the menisci oscillation amplitude and
the volume of a hemispherical vapor bubble of the diameter Dp:

2 1 3
Vg,max = Vg,min + 2π Rtun δam + π Dp (4.227)
12

Differences in vapor masses and volumes can be written in the following way:

mg,max – mg,min = ρg,tun (Vg,max – Vg,min) (4.228)

which, after the application of (4.225), (4.226) and (4.227), leads to the computation of the menisci
pulsation amplitude:

1 3 3
δam = 2 (Dg – Dp) (4.229)
24 Rtun

The bubble diameter Dg in the equations above is computed on the basis of a modified formula
worked out by Nakayama et al. [167, 168]:

(4.230)

where

(4.231)

240 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The length of the liquid lock Ll is computed on the basis of a comparison between the expressions
(4.223) and (4.229). Webb and Haider [314] assume the vapor is pushed from the tunnel into the
growing bubble until two opposing menisci, for which Lg,min equals zero, come into contact.

To perform the calculations, Ts, Tsat, stun, Rtun, Dp, the liquid thermal properties ρg, ρl, λl, σlg, θ,
(dp/dT) sat and hlg are given. The following are computed in succession: Qmk on the basis of
(4.215), sg using (4.229), (4.223) and (4.222) and consequently qtun with (4.214).

Webb and Haider [314] assume that natural convection on the single-phase area on the external
surface of the capillary-porous structure is negligible. Instead, it is transient conduction in the
liquid adjacent to the wall, induced by a step-change of the liquid temperature following vapor
bubble departure, that is decisive for the external heat flux transferred from this surface to the
liquid. Those issues were discussed by Mikic and Rohsenow [162].

Thus, transient conduction in the direction normal to the external surface is described by the
equation:

(4.232)

with the initial condition

τ=0 T(n, 0) = Tsat (4.233)

and boundary conditions

n=0 T(0, τ) = Ts n→∞ T(∞, τ) = Tsat (4.234)

The analytical solution of (4.241) has the form:

(4.235)

The mean heat flux referred to the area of vapor bubble effect and its formation cycle amounts
to [162]:

(4.236)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 241


Chapter 4

If the flux Q λ is known, it is possible to determine the heat flux from the external surface of the
porous structure by referring this heat flux to the density of nucleation centers N given by (4.221).
After the frequency of the process is accounted for using (4.218), qext takes on the form:

(4.237)

5⋅105
R 11

105

104
2
-2
q, W⋅m

1 3
103 4
Tsat = 26.7 oC
5
D p = 0.10 mm
D tun = 0.41 mm
stun = 0.55 mm
102

Experiment

101 -1
10 100 101
DTsl , K

a) broken lines – Nakayama et al. model [167, 168]: 1 – q; 2 – qext;


b) full lines – Webb and Haider model [314]: 3 – q; 4 – qtun; 5 – qext

Fig. 4.37. Comparison of theoretical computations with experimental data from


Webb and Haider [314]

In Fig. 4.37, theoretical computations of qtun, qext and q for the model under consideration and
also those for Nakayama et al. model [167, 168] are compared with the experimental data obtained
by Nakayama et al. [167, 168] for R-11. In the Nakayama model, the qext formula for R-11 has
the form:

(4.238)

242 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The solid lines in Fig. 4.37 represent curves obtained for the model under discussion. It can be
seen that the computed quantities differ greatly from the experimental data in value at the lower
and higher superheats and also with respect to the slope of the curves obtained but on the other
hand intersect the experimental data at intermediate values. This means that their attempt at
building a model without any experimental constants was only partially successful. Furthermore,
the model was designed primarily for low heat fluxes, where discrepancies are the greatest in
Fig. 4.37 (over an order of magnitude difference).

Chien and Webb model

Theoretical analyses provided by Chien and Webb [46] were preceded by experimental
investigations into the impact of geometric parameters of their enhancement structures, specially
manufactured for that occasion, on boiling heat transfer [47, 48]. Structures with subsurface
tunnels were produced by overlaying an externally micro-finned tube with a perforated foil. The
dynamics of bubbles departing from the porous cover was observed [49] and the boiling process
was photographed with an ultra fast camera [50]. Based on their observations, a semi-analytical
nucleate boiling heat transfer model was proposed for this type of enhancement structure with
subsurface tunnels.

Chien and Webb model [46], despite bearing many similarities to the Webb and Haider model
[314], makes a different basic assumption about the structure of the two-phase mixture filling the
subsurface tunnels. In the new model, the tunnels are assumed to be almost completely filled
with vapor, apart from the liquid menisci in the corners. Furthermore, the elementary quantity
determined in this model is the instantaneous evaporation intensity inside the tunnel, which is
obtained by analyzing the meniscus dimensions, the bubble growth process and the bubble
departure diameter and also transient convection on the external surface.

Like in earlier models [167, 168, 314], it is assumed that the total heat flux is the sum of qtun and
qext, as in (4.209). Evaporation from the surface of menisci decides the magnitude of the heat flux
qtun, whereas qext is strongly affected by transient conduction and convection, the latter caused
by departing vapor bubbles. The entire bubble growth cycle is divided into three periods: a waiting
period Δτwait, a bubble growth period Δτgth and a liquid suction period Δτsct as illustrated in
Fig. 4.38 and explained below.

In the waiting period Δτwait, the liquid evaporates inside the tunnel. The beginning and end of the
period are shown in Figs. 4.38a and 4.38b. Over that time, the meniscus radius decreases from
the initial value of Rms,wait to the value Rms,gth. In the waiting period, the tunnel becomes totally
filled with vapor apart from the liquid menisci areas in the corners.

In the bubble growth period, the vapor from the tunnel permeates through the surface pores into
the bubbles, whose radii increase from the initial value of R = Dp/2 to the departure bubble radius
of R = Dg /2 as shown in Figs. 4.38b and 4.38c. The meniscus radius, at the same time, changes

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 243


Chapter 4

from Rms,gth to Rms,sct. Evaporation from the meniscus does not occur if the saturation temperature
increases, the latter caused by the capillary pressure σlg /R on the heating surface, to equal the
surface superheating. This limit indicates that the radius Rms,sct should be at least equal to the
radius of the bubble, on the surface of which evaporation does not occur:

(4.239)

Fig. 4.38. Evaporation in the subsurface tunnel of an enhanced surface and the
bubble growth process of Chien and Webb [46]

Vapor bubble departure is followed by liquid being sucked into the tunnel and held up in its
corners. Experimental investigations conducted by Nakayama et al. [167, 168] and also by Chien
and Webb [49] indicate that the period is much shorter than the two previous ones, and it is
therefore neglected in their model when computing the bubble departure frequency. The analysis
of the parameters of the model proceeds then as follows.

244 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Heat flux in the tunnel – qtun

The following assumptions are made regarding heat transfer in the tunnels of the enhancement
structure:

1. Tunnels are filled with vapor except for liquid menisci in the corners. Menisci are
uniformly distributed along the tunnel length.

2. The latent heat of vaporization is used to form vapor inside the tunnels and
generate bubbles.

3. Evaporation rate can be determined by solving the one-dimensional equation of heat


conduction in the area surrounding the menisci.

4. In the rectangular tunnel section, four menisci are found in the corners. In the
circular section, see Fig. 4.38, only two menisci occur, and they are located at the
upper corners.

5. The temperature of the microfins is assumed to be constant for structures made of


materials characterized by good thermal conductivity, e.g. copper.

Fig. 4.39. Liquid-vapor meniscus in an enhancement tunnel corner of


Chien and Webb [46]

In accordance with the assumptions made above, the heat flux in the tunnel is the product of the
heat of evaporation from the meniscus surface for one cycle of bubble formation and departure,
so that:

qtun Ftun = Qms ω (4.240)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 245


Chapter 4

The heat flux of evaporation from the meniscus can be determined using the heat conduction
equation:

(4.241)

where δms (τ, φ) is the local thickness of the meniscus, like in Fig. 4.39. Following the geometry
proposed:

–1
δms (τ, φ) = [Rms (τ) + δn–ev] (cos φ) Rms (τ) (4.242)

where the thickness of non-evaporating liquid microlayer is determined using the relationship

(4.243)

–12
The Hamaker constant CH in (4.243) is assumed to equal 2 ∙ 10 J for all investigated liquids.
Elementary change in the surface area is:

dF = L ∙ Nms ∙ Rms ∙ dφ (4.244)

Therefore, on the basis of (4.250) and (4.251), it can be written:

(4.245)

where L is the total tunnel length and Nms is the number of menisci in a given tunnel.

The total heat amount needed for evaporation while a single bubble is formed amounts to:

(4.246)

Numerical integration of (4.254) and (4.255) is performed, changing, at each time step, the
meniscus radius value in accordance with the equation:

(4.247)

246 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where ΔVl represents a change in the liquid volume at a given step. The balance of mass and
energy gives a simple dependence:

∆Vl = ∆Qms h–1


lg ρl
–1
(4.248)

On the basis of (4.245), the change in the liquid volume at a given time step is

(4.249)

Integration of (4.245) to (4.249) starts with the initial value of the radius Rms = Rms,int, see
Fig. 4.38a, proper to the incipience of vapor bubble formation cycle. The meniscus radius
decreases at successive time steps until the value Rms = Rms,sct is reached, as in Fig. 4.38c.
Rms,int depends on the amount of the liquid flowing into the tunnel during the cycle of vapor bubble
formation. Geometric considerations lead to the following:

(4.250)

where Δfl,cycl is the change in the meniscus section area for one cycle of bubble formation.
It is proportional to the amount of the liquid flowing into the tunnel during this cycle. In order
to determine Rms,int and Δfl,cycl, Chien and Webb [46] developed a procedure based on their
experimental results [47, 48], which is discussed later in this chapter.

Vapor bubble departure diameter

On the basis of the results of their experimental investigations with the use of a fast video camera
[49, 50], the authors proposed an equation to compute the diameter of a departing vapor bubble:

(4.251)

where the Bond number is defined in a way similar to that in (3.39):

2 –1
Bo = Dp (ρl – ρg) g σlg (4.252)

The frequency of vapor bubble departure is assumed to be the inverse of the sum of the waiting
time Δτwait and the growth time Δτtgh.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 247


Chapter 4

Vapor bubble waiting and growth time

While analyzing the waiting time with respect to vapor bubble, Nakayama et al. [167, 168], using
balances of mass and energy, obtained the following:

(4.253)

where

(4.254)

Following the approach presented in [167, 168], Chien and Webb [46] introduced the equation of
state and the Clausius-Clapeyron equation to (4.253), which yields:

(4.255)

where Vg is the mean vapor volume at the waiting time, Tg,1 and Tg,2 represent the initial and
final temperature of the vapor during that period, Rg is the gas constant, ΔTtun stands for the
temperature change in the tunnel and Vg,2 is the vapor volume at the end of the waiting period.
According to the authors [46], the right side of (4.255) can be computed if a value for ΔTsl is
specified. The left side of (4.255) is computed on the basis of (4.245) – (4.249) until the values of
both sides are equalized (4.255). The value of Δτwait is determined in this way.

The growth time is computed with the use of a general expression presented by Mikic and
Rohsenow [163], in which vapor bubble growth is assumed to be dependent on increasing vapor
pressure inside the bubble, inertia forces and surface tension:

(4.256)

where the constant C = �/7 for bubbles of spherical shape observed by Chien and Webb [49].
Applying the Clausius-Clapeyron equation and also the authors’ experimental results [49], it was
possible to transform (4.256) to the form:

(4.257)

248 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Then, by integrating the above equation (4.257) over the time interval: τ = 0, R = Dp/2 to τ = Δτgth,
R = Dg /2, we receive:

(4.258)

Nucleation site density

Vapor generated inside the tunnel departs through the pores in the cover plate. The bubble
generation frequency ω is the inverse of the sum of the waiting time τwait (4.255) and the growth
time τgth (4.258). Owing to a simple balance of mass and energy, the expression for the nucleation
site density is obtained:

(4.259)

where the bubble diameter Dg is given by (4.251).

External heat flux – qext

It is assumed that the external heat flux is contained between two asymptotes. One of them results
from the assumption that it is transient conduction that controls heat transfer. In accordance with
the solution provided by Mikic and Rohsenow [162], qext is given by (4.237). The other asymptotic
limit corresponds to steady-state convection induced by departing vapor bubbles. Considering a
smooth surface, Haider and Webb [95] obtained:

(4.260)

where the constant value C = 6.42 is obtained on the basis of experimental data presented by
Chien and Webb [49] and also Nakayama et al. [167, 168].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 249


Chapter 4

Computational procedure

The total heat flux q is computed while assuming that ΔTs,sat is given as well as the characteristic
dimensions of the enhancement: Dp – pore diameter, sp – pore pitch, stun – microfin (tunnel) pitch,
htun – tunnel height. The procedure to implement their model is as follows:

1. Compute the vapor bubble departure diameter Dg (4.251) for the chosen pore diameter
and known thermal properties of the boiling liquid.

2. Compute the bubble growth time Δτgth using (4.258) for the chosen ΔTsl and diameter
Dg calculated in step 1.

3. Compute the initial value of the meniscus radius Rms,int (4.250). For the Δfl,cycl
computation, the correlation (4.261) is used.

4. Integrating (4.245) and (4.248), calculate the amount of latent heat of vaporization
necessary for the formation of a single vapor bubble.

5. On the basis of (4.245) – (4.249) and (4.255), compute the waiting time Δτwait and the
value of the meniscus radius Rms,gth at the beginning of the growth period.

6. For the times Δτwait < τ < Δτwait + Δτgth and decreasing meniscus radius Rmk, continue
calculating the heat flux Qtun.
–1
7. Compute the frequency of vapor bubble departure ω = (Δτwait + Δτgth) and the value
of qtun using (4.240)..

8. Compute the density of nucleation sites N using (4.259).

9. Compute the density of the flux qext in accordance with (4.260).

10. The total heat flux then obtained from the sum of q = qext + qtun.

Meniscus cross-section change – Δfl,cycl

Chien and Webb [46] computed the change in the liquid layer section in the liquid-vapor meniscus
surroundings on the basis of the following experimental correlation:

(4.261)

which was determined on the basis of experimental values Rms,int. The values were obtained
while repeating the computational procedure 1 to 10 above until compatibility of the values q = qexp
was obtained.

Experimental investigations were conducted for pool boiling on cylindrical surfaces with external
microfins making tunnels of a semi-circular cross-section after being covered with a perforated
foil, as shown in Fig. 4.38. R-11, R-123, R-134a and R-22 were the liquids boiled. The heat flux

250 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

4 4 2
ranged from 1 ∙ 10 to 6.5 ∙ 10 W m , which constituted a range of 20 to 100% of the critical heat
flux qcr [47, 48]. The heat flux value corresponding to the maximum heat transfer coefficient, which
is accompanied by the tunnel completely filled with vapor, was assumed to be the critical one. The
comparison of the computed value of q with experimental data qexp is presented in Fig. 4.40.

1.0
R 11
0.8
R 123
0.6 R 22
R 134a
0.4
(q - q exp )/qexp

0.2

-0.2

-0.4
Tunnels partially flooded
-0.6 R 11, 0.30 qcr
R 123, 0.36 qcr
-0.8
R 22, 0.35 qcr
-1.0
0 1⋅104 2⋅10 4 3·10 4 4⋅104 5·104 6⋅10 4 7⋅104
qexp , W⋅m-2

Fig. 4.40. Comparison of computed and experimental heat flux density values from
Chien and Webb [46]

Fig. 4.40 shows that the model discussed above, together with the computational procedure
designed to accompany it, can predict the experimental heat flux with an accuracy of about ±33%.
Fig. 4.40, however, does not compare shapes of computed curves with experimental ones. It can
be concluded from looking at the results for R-11 and R-123 that the theoretical and experimental
curves intersect. This means that theoretical curves do not properly follow the slope of the
boiling curves, perhaps by incorrectly modeling the increase in the density of nucleation sites or
vapor bubble departure frequency with the increase in the heat flux. Both the model [46] and the
previous model presented by Webb and Haider [314] have very complex computational codes, so
it is difficult to determine which assumption of the model strongly affects the final result, i.e.
q = q(ΔT).

Direct comparison of the two above-mentioned models shows that the difference in the basic
physical assumptions regarding how the porous structure tunnels fill with liquid and vapor does
not produce a significant change in the computed results and their level of agreement with the
experimental data.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 251


Chapter 4

4.3. Heat transfer hysteresis

4.3.1. Kinds of boiling hysteresis

The application of microporous enhanced boiling surfaces leads to significant boiling heat
transfer enhancement. The fact is confirmed by numerous publications presenting the results
of experimental investigations, discussed in Chapter 3. The majority of experimental data found
in the literature were obtained by determining the boiling curve for increasing values of the heat
flux, from the minimum to near the critical (maximum) heat flux. Far fewer works focused on the
hysteresis phenomenon, or rather phenomena, revealed when the heat flux is initially increased
and then decreased.

Hysteresis phenomena manifest themselves more clearly in boiling on enhanced boiling


surfaces than on smooth surfaces. Experimental boiling curves with hysteresis have diversified
shapes, depending on the enhancement structure’s geometry and thermal parameters and also
on the liquid physical properties. To apply boiling enhancements in heat transfer equipment,
it is necessary for these surfaces to have stable and unambiguous shapes of boiling curves.
Hysteresis is generally regarded as a disadvantageous phenomenon, which prevents thermal
stabilization of systems emitting high heat fluxes.

Research work is therefore designed mainly to determine the conditions under which hysteresis
occurs and find the means to eliminate it [25, 27, 107, 142, 151]. The majority of investigations into
hysteresis, however, refer to only one kind, which results from the fact that most investigations
were conducted for only one class of porous coating and a few boiling liquids. A different
approach, which makes an attempt at practically applying one kind of hysteresis to control the
boiling heat transfer process, is presented in the author’s works [195, 205 – 208, 218, 219].

Fig. 4.41 shows the juxtaposition of known shapes of boiling heat transfer hysteresis. Analyzing
the curves in Figs. 4.41 and 4.42, it is possible to differentiate characteristic types as follows:

a. Figs. 4.41a and 4.42a: the phenomenon is termed zero boiling crisis, TOS –
temperature overshoot or nucleation hysteresis [10, 11, 17, 22, 25-27, 36, 40 – 43, 94,
107, 109, 122, 127, 128, 132, 143 – 145, 151, 153, 154, 156, 265, 273, 276, 296, 297,
329, 331, 332];

b. Figs. 4.41b and 4.42b: I kind of hysteresis, or intralayer (internal) boiling crisis, or
inverse hysteresis loop, characterized by a decrease in the heat transfer coefficient
after the heat flux is reduced, occurring after the heat flux reaches the critical heat flux
[7, 11 – 13, 54, 118, 122, 127, 128, 132, 151, 153, 154, 203, 207, 211, 212, 214 – 218,
265, 273, 276];

c. Figs. 4.41c and 4.42c: II kind of hysteresis, or simple hysteresis loop, or hysteresis with
temperature deviation, distinguishable by an increase in the heat transfer coefficient

252 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

when the heat flux prior to its reaching the critical heat flux [18, 132, 203, 205 – 208,
211, 214, 216, 218, 219, 273, 274, 278, 281, 302, 307];

d. Figs. 4.41d and 4.42d: hysteresis characterized by a decrease in the heat transfer
coefficient when the heat flux decreases prior to its reaching the critical heat flux (rarely
observed) [320];

e. Figs. 4.41e and 4.42e: hysteresis characterized by first a decrease and later an increase
in the heat transfer coefficient for the heat flux decreasing after reaching the critical
heat flux (rarely found) [7];

f. Figs. 4.41f and 4.42f: boiling crisis hysteresis characterized by an increase in the critical
heat flux with each repeated increase in the heat flux supplied to the heating surface,
illustrated later in Figs. 4.60 and 4.61 [273, 274, 276].

q
f e

c b

a
Pool boiling on a smooth
surface

DTsl

Notation a, b, c, d, e, f explained in text

Fig. 4.41. Juxtaposition of different kinds of hysteresis in boiling heat


transfer on microporous surfaces

4.3.2. Nucleation hysteresis

Technically smooth surfaces – experimental investigations and


theoretical analysis

Nucleation hysteresis, as shown in Figs. 4.41a and 4.42a, for boiling on a smooth surface was
discovered by Corty [66] and van Camp [39]. The phenomenon occurs in the initial segment of the
boiling curve, at the transition from natural convection to nucleate boiling. Gradual increase in the
heat flux does not initiate nucleate boiling and natural convection continues to elevated superheats
of the heating surface ∆Tsc, normally characteristic of nucleate boiling. Vehement transition to

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 253


Chapter 4

nucleate boiling is often accompanied by fluctuations of the heating surface temperature. This is
the most commonly type of hysteresis found experimentally, observed both on smooth surfaces
and microporous surfaces.

a) b)
q q

DT 1 DT2 DT sl DT sl
c) d)
q q

DT sl DT sl
e) f)
q q Dqcr

2 1

DTsl DT sl

Notation a, b, c, d, e, f explained in text

Fig. 4.42. Types of boiling heat transfer hysteresis (see Fig. 4.41)

Numerous results of experimental investigations, and attempts at providing theoretical explanation


for the boiling incipience phenomena and nucleation hysteresis accompanying it for technically
smooth surfaces of various roughness and geometry, are presented in review works by Bräuer
and Mayinger [36] and also in Bar-Cohen [22]. Investigations concerned both pool boiling and flow
boiling.

Bräuer and Mayinger [36] differentiated two kinds of models for explaining the necessary heating
surface superheating above the saturation temperature for boiling to initiate, namely thermal
(thermodynamic) models and mechanic ones. The authors count as thermal models those, in

254 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

which it is assumed that the loss of thermodynamic equilibrium by a vapor bubble nucleus located
in a microcavity in the heating surface leads to boiling incipience.

For boiling incipience to occur, it is necessary to superheat both the heating surface and the liquid
layer in its surroundings. A frequently applied expression relating the heat flux and the necessary
heating surface superheating is an inequality proposed by Sato and Matsumura [230]:

(4.262)

In deriving (4.262), the Clapeyron equation, a balance of forces acting on a spherical vapor bubble
and the equation of heat conduction in the liquid (to compute the thickness of the superheated
liquid layer) are used. Boiling incipience is assumed to occur at the minimum thickness of the
superheated liquid layer at which a vapor bubble of the critical radius can be formed. Equation
(4.262) yields results compatible with the experiment for liquids poorly wetting the heating surface
[22, 36].

Boiling incipience in mechanical models is assumed to be dependent on boiling damping


phenomena (e.g. initial setting of a high pressure), the heating surface temperature, the cavity
geometry and its dimensions, the wetting angle, any non-condensing gas in the boiling liquid and the
heat flux. The model of boiling incipience worked out by Cornwell [63], in reference to the hysteresis
of the dynamic wetting angle, is described in Chapter 2, see Figs. 2.12 – 2.17.

For liquids wetting the heating surface well, boiling incipience takes a different course than is the
case for poorly wetting liquids. Two important factors should be mentioned in this respect:

1. A liquid with a small wetting angle penetrates the heating surface cavities due to
its increased flooding ability. The result is a considerably smaller number of active
nucleation sites in the same surface area when compared with a poorly wetting liquid.

2. In well wetting liquids, residual vapor bubbles remaining in surface cavities are small,
and they therefore require a large superheating to initiate their growth.

In accordance with the remarks above, Hahne et al. [94] assumed that in well wetting liquids,
nucleation sites of the largest diameters decide boiling incipience. They proposed the following
expressions to compute the heating surface superheating and the corresponding flux density at
boiling incipience:

(4.263)

(4.264)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 255


Chapter 4

(4.265)

where T∞ is the temperature of the subcooled liquid flowing above the heating surface. The
second term in both pairs of square brackets describes the vapor temperature superheat with
respect to the saturation temperature inside the bubble of characteristic radius R*. In accordance
with the definition put forward by Mizukami et al. [164], the radius R* describes the curvature of
the interfacial liquid-vapor surface, under the conditions of the greatest damping. Its magnitude
decreases as the surface tension σlg decreases, thus also as the pressure increases [36]. Good
agreement between the measured and computed heat flux values using (4.264) was obtained for
R-12 [36].

If the heating surface is immersed in a well wetting liquid, it is possible there are no vapor nuclei,
which inhibits heterogeneous nucleation from occurring. In this case, the liquid superheat grows
until it reaches a point at which a spontaneous molecule agglomeration of increased internal
energy forms a stable vapor bubble. Lienhard and Karimi [143-145] demonstrate that the liquid
superheat, indispensable for boiling incipience to proceed according to the described mechanism,
reaches, for homogeneous nucleation, values close but lower than those characteristic of the
spinodal curve:

(4.266)

Tong [296] showed that (4.266) gave results that are close to experimental data for many well
wetting liquids, such n-heptane, n-pentane and perfluoro-derivatives of hydrocarbons (e.g.
FC-72). However, while such relationships for determining the surface superheat and the heat flux
accompanying it yield results compatible with experiments for increasing heat flux, they do not
explain the nucleation hysteresis phenomenon itself.

Bar-Cohen [22] analyzed the impact of the following parameters on nucleation hysteresis: liquid
thermal properties, pressure, flow velocity, liquid subcooling, the heating surface roughness,
the presence of the porous coating and non-condensing gases in the boiling liquid and also the
mixture composition. The analysis of numerous experimental data provided by many sources
[22] for well wetting liquids leads to the statement that the temperature overshoot is smaller for
perfluoro-derivatives of hydrocarbons (fluorinerts, e.g. FC-72) than for chlorofluorohydrocarbons
(e.g. R-113). The temperature overshoot can be decreased or even eliminated by increasing the
saturation pressure, the liquid flow velocity and non-condensing gas concentration. Another way
to reduce nucleation hysteresis, that is diminish the heating surface temperature overshoot, is to
increase the liquid subcooling, cover the surface with a microporous layer, in other words increase

256 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

the surface roughness, and also, in some cases, by adding a liquid of higher boiling point (5% to
10% of the volumetric content).

Bilicki [31] and also Bohdal and Czapp [34] conducted experimental investigations and made
theoretical analysis of zero boiling crisis (nucleation hysteresis) in refrigerants flowing through
cylindrical sections. On the basis of thermodynamic analysis of flow boiling, Bilicki demonstrated
that the phenomenon is caused by the transport of latent heat by vapor bubbles formed in the
liquid adjacent to the wall that functions as a heat sink.

Industrially developed microporous surfaces – experimental data

Regarding industrially manufactured microporous surfaces, the largest collection of data on


nucleation hysteresis is available for the High Flux surface, examined by Bergles and Chyu [26,
27] for water and R-113, see Fig. 3.2, Marto and Lepere [156] for R-113, see Fig. 3.4, Trewin et al.
[297] also for R-113 as well as Chang and You [40 – 43] for FC-72, FC-87 and R-123, in Fig. 3.75
and Shakir and Thome [243] for water, ethanol, acetone and methanol.

Nucleation hysteresis was also observed on Turbo-B surfaces with fluorinert FC-87 as the boiling
liquid [41, 42] and for R-113 [297]. Marto and Lepere [156] conducted extensive experimental
investigations into R-113 and FC-72 boiling heat transfer for a number of different industrially
enhanced surfaces. They showed that nucleation hysteresis occurred on all the surfaces tested,
such as High Flux and Thermoexcel-E, in Fig. 3.4, and Gewa-T.

For a few different enhanced surfaces of Gewa-T type and R-113, Ayub and Bergles [17] recorded
“stepwise” nucleation hysteresis that occurred in many small steps, as depicted in Fig. 4.43. Its
occurrence was attributed to gradual extension of the heating surface areas on which boiling
takes place. The activation of successive zones proceeds “step by step”, due to consecutive
temperature increments. They claimed that concave spiral channels inside Gewa-T structures
enhanced free convection and delayed boiling incipience. Consequently, one active channel does
not activate the neighboring ones, which may be responsible for the small “stepwise” nucleation
hysteresis they observed.

Nucleation hysteresis for water and ethanol [151, 153, 329, 331, 332] and also R-113 [25, 108,
109] was observed on porous coatings made of sintered powders. Kravchenko and Ostrovskij
[132] noticed this kind of hysteresis on surfaces made with a plasma spray technique for ethanol,
acetone and the mixture of the two liquids. Nucleation hysteresis also occurs on microporous
surfaces and those made from microfins and microposts, see Fig. 3.10, in accordance with [10] in
Fig. 3.75 and by others [40 – 43]. In those investigations, FC-72, FC-87 and R-113 were the
test liquids.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 257


Chapter 4

5
10
R 113
Gewa - T19D

D C

q, W⋅m-2 4 B
10
A
q-

Smooth
surface

10 3
0.1 1.0 10.0 30.0
DTsl , K

Fig. 4.43. “Stepwise” nucleation hysteresis on a cylindrical Gewa-T19D surface


from Ayub and Bergles [17]

Models for microporous surfaces

Zhang and Zhang model

An attempt was made at providing a theoretical explanation for nucleation hysteresis on


microporous surfaces (capillary-porous structures) made of sintered powders [109, 151 – 154,
329]. Zhang and Zhang [329] assumed that the pore openings have shapes of inverted cones, see
Fig. 4.44, in which residual amounts of vapor or inert gas are found. The condition of the liquid,
vapor and inert gas pressure balance can be written summing equations (2.40) and (2.41):

2 σlg
pgas + pvapor – pl = (4.267)
R

The partial pressure of the gas (or vapor) entrapped in the conical pore is determined using the
ideal gas equation:

3 3
pgas = (mgas Rgas Tvapor/gas / π R ) (4.268)
4

where mgas is the mass of residual amounts of the gas (vapor).

258 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

In accordance with Clausius-Clapeyron equation, the slope of the saturation curve is:

(4.269)

Fig. 4.44. Equilibrium vapor nucleus from Zhang and Zhang [329]

Making use of the fact that vg >> vl and substituting differentials with finite differences in (4.269),
the latter is transformed to the form

pvapor – pl = (Tvapor – Tsat) ρg / Tsat (4.270)

Inserting (4.267) and (4.268) into (4.270) and transforming the expression makes it possible to
determine the heating surface superheat indispensable for a vapor nucleus to be at equilibrium:

ΔTsl = Tsl – Tsat = Tvapor – Tsat (4.271)

which takes on the form

3
∆T1 = ∆Tsl = Tsat (2 σlg /R – 3 mgas Rgas Tvapor/gas /4π R ) /hlg ρg (4.272)

If residual amounts of the vapor or gas do not exist in the pores, mgas = 0, then the superheat is
expressed by:

∆T2 = ∆Tsl = 2 σlg Tsat /hlg ρg R (4.273)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 259


Chapter 4

Obviously ΔT2 > ΔT1, which means that the heating surface superheat necessary to initiate boiling
is, in this case, much higher than when residual amounts of the vapor or gas are contained in a
pore. Thus the temperature overshoot of nucleation hysteresis amounts to:

3
∆Thst = ∆T2 – ∆T1 = 3 mgas Rgas Tsat Tvapor/gas / 4 π R ρg hlg (4.274)

According to Zhang and Zhang [329], the occurrence of nucleation hysteresis is conditioned by
the process of pore wetting, which, depending on the wetting properties of the liquid, can involve
residual amounts of vapor or gas being either entrapped in or pushed out of the pore, as illustrated
in Fig. 4.45. If the wetting angle satisfies one of the two conditions in Fig. 4.45a:

θ < 2β θ < � – 2β (4.275)

Fig. 4.45. Liquid and vapor flow in the surroundings of a pore opening from
Zhang and Zhang [329]

then the liquid floods the cone, pushing the vapor (or gas) out of it. The cavity (pore) is inactive
then and falls in zone II of their diagram in Fig. 4.46. If the wetting angle satisfies two other
conditions in Fig. 4.45b:

θ > 2β θ > � – 2β (4.276)

then the liquid is not able to push the vapor (or gas) out of the cone and the cavities (pores)
become stable nucleation centers, falling in zone I in Fig. 4.46. In the remaining zones III and IV in

260 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Fig. 4.46, the conical cavity (pore) activation depends on the shape of the liquid-solid state system
inside each of the cavities.

Fig. 4.46. Nucleation zone diagram from Zhang and Zhang [329]

Prior to boiling incipience, residual amounts of inert gas or vapor in porous layers of sintered
powders are not able to leave the layer despite considerable superheating of the heating surface.
This is due to the following factors: static pressure of the liquid, high resistance to flow through the
porous structure and capillary forces. Increasing the superheat causes more vapor accumulation
inside the porous coating. The superheating ΔT2 given by (4.273) is therefore greater than ΔT1
given by (4.272) that is sufficient to sustain boiling.

When pressure builds up, a part of the vapor leaves the porous coating, which causes the
pressure to then drop and flashing of some of the liquid to occur. This flashing phenomenon
enhances heat transfer and the surface superheating drops to its stable value ΔT1.

The above interpretation of the temperature overshoot process led Zhang and Zhang [329] to the
conclusion that the main reasons for nucleation hysteresis are: the total filling of inactive pores
with the liquid and the medium resistance to flow through the porous coating. The temperature
overshoot accompanying nucleation hysteresis ΔThst is properly expressed by the relationship
(4.274). The overshoot depends on the pore’s ability to accumulate gas (or vapor) (mgas), the liquid
thermal properties (hlg ρg) and the thermodynamic state of residual amounts of gas (or vapor)
(Tvapor/gas, Tsat).

Zhang and Zhang [329] also carried out extensive experimental investigations on nucleation
hysteresis. Boiling heat transfer investigations were conducted on heating surfaces made of
sintered copper powders made with the following parameters: powder grain = 40 – 300 grains
per inch (mesh); thickness δ = 0.5275 – 2.099 mm; porosity ε = 0.351 – 0.5246; relative coating
thickness δ/Dp = 4.068 – 19.99. Water and ethanol were the test liquids and experiments were
conducted at atmospheric pressure and changeable initial liquid and porous sample subcooling
with respect to the saturation temperature.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 261


Chapter 4

The impact of the following parameters was analyzed: the liquid wetting properties, liquid
subcooling and the porous coating structural parameters on temperature overshoot at nucleation
hysteresis.

Fig. 4.47 compares boiling curves for water and ethanol. For the ethanol–copper arrangement, the
wetting angle θethanol ranges from 8 to 15°, for the water–copper arrangement, θ water > 50°. That
means, for ethanol, the majority of pores are located in zone II in Fig. 4.46, i.e. they are flooded
with the liquid. Consequently, nucleation hysteresis is much stronger than that for water. The
likelihood of flashing is low for water, after a part of the vapor contained inside the porous coating
is liberated, hence the stepwise character of the hysteresis phenomenon in Fig. 4.47b. Similar
occurrences were also observed by Ayub and Bergles [17] for R-113 in Fig. 4.43.

a) b)
1⋅10 5 4⋅105
3
5

2 1⋅10 5
q, W⋅m-2

q, W⋅m-2

1⋅104 5

5 3

2
1⋅104
1⋅103

5⋅102 4⋅103
0.3 0.5 1 3 5 0.5 1 3 5 10
DTsl, K DTsl, K

a) ethanol: ○ – ∆Tsub = 75.0 K ● – ∆Tsub = 10.0 K


▲ – ∆Tsub = 30.0 K ▼ – decreasing heat flux,
sintered copper powder, δ = 1.07 mm, ε = 0.5091, δ/Dp = 4.068

b) water: ○ – ∆Tsub = 51.8 K ● – ∆Tsub = 5.1 K;


▲ – ∆Tsub = 20.0 K ▼ – decreasing heat flux;
sintered copper powder, δ = 0.6485 mm, ε = 0.8978, δ/Dp = 4.266

Fig. 4.47. Boiling curves with nucleation hysteresis of Zhang and Zhang [329] for
a) ethanol and b) water

The temperature overshoot ΔThst grows with decreasing porosity ε in coatings of sintered
powders as shown in Fig. 4.48b. When the porosity ε is increased, the residual amount of gas (or
vapor) entrapped inside the coating also increases, which facilitates nucleation and diminishes
the temperature overshoot ∆Thst. The relative coating thickness δ/Dp, being a measure of flow
resistance, to all practical purposes does not affect nucleation hysteresis according to Fig. 4.48b.

262 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

On the one hand a larger thickness hinders pores from flooding, while on the other it delays
vapor outflow.

a) b)
4⋅104 1⋅105

2 5

1⋅104
2
q, W⋅m-2

q, W⋅m-2
5
1⋅104

2 5
3
1⋅10
2

2
4⋅10 1⋅103
0.3 1.0 2 4 6.0 0.3 1.0 3 5 10.0
DTsl , K DTsl , K

a) 1. δ = 0.6083 mm ε = 0.4785 δ/Dp = 5.7993


▲ – ΔTsub = 20 K ○ – ΔTsub = 5 K
2. δ = 1.074 mm ε = 0.5091 δ/Dp = 4.068
▼ – ΔTsub = 20 K □ – ΔTsub = 5 K

– δ = 1.074 mm ε = 0.5091 δ/Dp = 4.068


b)
Δ
○ – δ = 0.866 mm ε = 0.4771 δ/Dp = 5.697
□ – δ = 0.6083 mm ε = 0.4785 δ/Dp = 5.7933
Δ – δ = 0.957 mm ε = 0.5246 δ/Dp = 6.2961
◊ – δ = 0.874 mm ε = 0.3753 δ/Dp = 8.3238

Fig. 4.48. (a) Impact of the boiling liquid subcooling and (b) structural
parameters of the porous coating on ethanol nucleation hysteresis
from Zhang and Zhang [329]

According to the Bergles and Chyu [26, 27] investigations, shown in Fig. 3.2, subcooling hardly
affects nucleation hysteresis for R-113 and has a weak effect on water. Fig. 4.48a indicates that
subcooling changes the boiling curve shape for ethanol. Zhang and Zhang [331] attribute this to
the condensation of residual amounts of vapor contained inside the porous coating, which causes
pores to deactivate due to their flooding. Thus the greater is the subcooling ΔTsub, the larger is
the temperature overshoot ΔThst in Fig. 4.48a. If this explanation is correct, it also follows that
subcooling’s impact on nucleation hysteresis increases with increasing wetting angle θ. Since the
wetting angle for R-113 is small, θ R-113 < 5°, hence subcooling hardly affects its hysteresis.

The results of experimental investigations conducted by Zhang and Zhang [329] for sintered
powder layers bring the same conclusions as those previously discussed, drawn from results for

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 263


Chapter 4

technically smooth surfaces and structures made with other techniques. The wetting angle θ
and penetrating properties of the boiling liquid decide the magnitude of the temperature
overshoot ΔThst.

Ko et al. model

The results of experimental investigations of Ko et al. [109] are congruent with those obtained by
Zhang and Zhang [329], with their results shown in Fig. 4.49. The experiments were conducted
with water (Fig. 4.49a) and R-113 (Fig. 4.49b) and confirm earlier conclusions above that the
wetting angle decides the magnitude of the temperature overshoot ΔThst, i.e. the smaller is the
angle θ, the greater is the overshoot ΔThst. Ko et al. [109] presented their interpretation of the
temperature overshoot ΔThst partially based on the same assumptions as in the model by Zhang
and Zhang [329]. For the sake of simplification, it was assumed that the capillary-porous structure
of sintered powder was a composition of vertical capillaries. Other assumptions were as follows:

1. all grains of metal powder have the same diameter;

2. all free spaces (pores) have the same shape;

3. each free space (pore) is an active nucleation site;

4. the thermal conductivity of the porous layer is very large, i.e. the temperature gradient
across the layer can be neglected.

Because of nucleation hysteresis, the surface superheat ΔT2, see Fig. 4.42a, is much higher than
the value necessary to maintain boiling. At the instant of vapor out flow from the capillary-porous
structure, vapor pressure amounts to, as per Figs. 2.1 and 2.3:

pg = pl + 2 σlg /R (4.277)

where, in accordance with (2.9)

R = Ropn /cos θ (4.278)

The Clapeyron equation (4.269) is written in the form:

(4.279)

Δρ = ρl – ρg

264 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Thus:

(4.280)

~ ∆p/∆T and p = h g ρ holds, it is possible to transform (4.279) to


If the assumption that dp/dT = l l
the following:

(4.281)

a)
105
Water
- q↑
- q↓
q, W⋅m-2

4
10
[3 7]
l n
ra
atu ec tio
N nv
co

10 3 -1 0 1
10 10 10 102
DT sl , K
b)
105
R 113
- q↑
- q↓
q, W⋅m-2

104
]
[37
al n
atu r e ctio
N nv
co
103 -1
10 10 0 101 102
DTsl , K

Copper powder – 60 – 80 grains/inch (mesh), δ = 1.5 mm,


ε = 0.60 – 0.65 from Ko et al. [109]

Fig. 4.49. Nucleation hysteresis on porous coverings of sintered powders

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 265


Chapter 4

After boiling incipience, the heating surface superheat is reduced to the value ΔT1, see Fig. 4.42a.
It is assumed that for ΔT = ΔTl large vapor bubbles whose radii equal Rp are generated in free
spaces (pores). The liquid covering the porous surface is presumed to be superheated hence, in
accordance with Gibbs equation:

(4.282)

where ϑ = (dp/dT) sat is the saturation curve p-T inclination.

The heating surface superheating ΔT1 can be written as:

∆T1 = Ts – Tsat = (Ts – Tl) + (Tl – Tsat) (4.283)

The temperature difference Ts – Tl is determined using the Fourier equation for the liquid layer of
thickness δl, covering the inside of the pore:

(4.284)

where F is the surface area of the porous coating base (the heating surface) and Fp is the pore
surface area.

Following Gottzmann et al. [91], the authors adopted a coefficient characterizing the porous
coating geometry

(4.285)

owing to which, after transformation, it is possible to write (4.284) in the form

2 –1
Ts – Tl = Λ q Rp λl (4.286)

Inserting (4.282) and (4.286) into (4.283) yields:

2 –1
∆T1 = Λ q Rp λl + 2 σlg / ϑ Rp (4.287)

On the basis of (4.281) and (4.287), the temperature overshoot for nucleation hysteresis is:

(4.288)

266 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

From equation (4.288), it follows that the greater is the temperature overshoot for nucleation
hysteresis, the higher is the liquid column above the heating surface (h g ρ l), the lower is the value
of the wetting angle (θ), the higher is surface tension (σlg) and the lower is the latent heat (hlg).
Structural parameters of the porous covering also affect the value of ΔThst.

The conclusions drawn by Ko et al. [109] are that the wetting angle and its dynamic changes as
the function of the liquid motion direction have a decisive effect on hysteresis, as seen in Fig. 2.11
and Fig. 4.45. Their experimental investigations and theoretical considerations are congruent with
those presented by other authors with regard to the basic mechanics of nucleation hysteresis.
Neither of the two models presented, however, offers the possibility of predicting the temperature
overshoot ΔThst for predetermined parameters of the porous coating or the boiling liquid
thermal properties.

Malyshenko et al. model

Malyshenko et al. [151 – 154, 262, 263] have suggested that differences in the values of the heat
flux and the heating surface superheat for boiling incipience and stable boiling are decreased as
a result of the different geometry of pores occupied by vapor in each of those processes. The
external radius of the microcavity opening in the heating surface – Ropn (see Figs. 2.1 and 2.3) is
decisive for boiling incipience on technically smooth surfaces. Instead for porous surfaces, two
characteristic radii are found: the radius Ropn and the minimum radius of the pore filled with vapor
when its vapor starts exiting the porous layer – Rp. The smaller of the two radii decides the heating
surface superheat necessary to initiate boiling. In accordance with equations (2.1), (2.8) and (2.9),
the superheat is expressed by:

(4.289)

where R = min {Ropn; Rp}. As for capillary-porous structures, Ropn and Rp are much greater than
Ropn for smooth surfaces, so the boiling incipience for such structures occurs at smaller ΔTincip.
Numerous experimental data confirm this conclusion.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 267


Chapter 4

5
10
Water
7
0.1 MPa
5

q, W⋅m-2
2
-1
4 -2
10
-3
7 -4
5 -5
-6
3⋅103
0.5 1 2 3 6
DTsl , K

Nichrome, δ = 0.8 mm, Dgrn = 130 to 140 µm, MPa; 1 – 3 – deactivation through
preliminary suction process; 4 – 6 – deactivation through preliminary suction
2 –2
process and prolonged boiling with cooling; 1, 4 – increase from q = 10 W m ;
2, 5 – q lowered after developed nucleate boiling was obtained; 3, 6 – successive
q increase, from Malyshenko et al. [151, 154]

Fig. 4.50. Initial segment of the boiling curve on a porous coating of


sintered powders

At the same time, deactivation of active pores proceeds much easier because the curvature
radius at the pore entrance is a function of the diameter of particles comprising the structure, i.e.
it is large enough. Under such conditions, thermocapillary convection, induced by superficially
inactive and non-evaporating liquid impurities, can also develop in the liquid microlayer covering
the structure skeleton walls [262]. This mechanism can cause nucleation hysteresis that manifests
itself much stronger than on a smooth surface [153].

The magnitude of the nucleation hysteresis loop depends on the mutual interaction between
vapor outlets Ng and the number of active pores Np per unit surface area of the porous coating. A
change in their mutual relationship, due to various nucleation site deactivation procedures applied,
alters the shape of the hysteresis loop, as can be seen in Fig. 4.50.

The minimum radius of a pore filled with vapor can be computed if the function of pore distribution
density and the geometry of pore spatial lattice (owing to percolation theory [45]) are known. The
percolation threshold for connections between pores of known spatial lattice is determined using
the relationship:

(4.290)

268 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where w is the number of dimensions for pore spatial lattice and z is the coordination number.

5.0
-1
4.0 -2
-3
-4

DTincip, K
3.0
-5
-6
2.0
-7
-8
1.0

0
0.02 0.05 0.1 3.1
psat, MPa

1 to 5 – sintered powders; 6 to 8 – plasma spraying: 1, 2 – Dgrn = 130 to 140 µm,


δ = 1.5 mm; 3 – Dgrn = 130 to 140 µm, δ = 0.8 mm; 4 – Dgrn = 380 to 400 µm,
δ = 1.8 mm; 5 – Dgrn = 380 to 400 µm, δ = 0.8 mm; 6, 7, 8 – ε = 0.32 to 0.36;
6 – h = 0.2 mm; 7 – h = 0.5 mm; 8 – h = 1 mm; full lines – computation results,
from Malyshenko et al. [151, 152]

Fig. 4.51. ΔTincip dependence on pressure for water and nichrome porous coverings

The assumption that vapor propagates inside the porous layer along the largest pores available
leads to the dependence:

(4.291)

on the basis of which, if f (R) is known, it is possible to compute the pore radius Rp for the
vapor that starts departing from the layer. Owing to (4.289), (4.290) and (4.291), it is possible to
compute the heating surface superheat ΔTincip for nucleate boiling incipience. Fig. 4.51 presents
the comparison of experimental data for water and nichrome porous coatings with computational
results of ΔTincip as a function of pressure [151 – 154]. Good agreement in the pressure range from
0.05 to 3.1 MPa is shown.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 269


Chapter 4

4.3.3. I kind of hysteresis – analysis of experimental data

Sintered powder and plasma sprayed porous coverings

The I kind of hysteresis is observed primarily for water, see Fig. 3.68 [206, 211, 317], [7, 12,
118] and also for liquids that wet the heating surface well [132]. I kind hysteresis as depicted in
Figs. 4.41b and 4.42b, nucleation hysteresis as depicted in Figs. 4.41a and 4.42a and transient
phenomena in boiling on surfaces with a porous coating are the subject of a collection of works
by Styrikovic, Andrianov, Malyshenko and co-researchers [11, 12, 13, 151 – 154, 263 – 265]. The
authors present both the results of experimental investigations, shown in Figs. 3.23 and 3.24, and
also make an attempt to provide a theoretical interpretation of nucleation and I kind hysteresis
phenomena.

Experiments were conducted for pool boiling of two liquids: water and ethanol, mainly at
atmospheric pressure (0.1 MPa) and several other pressures: 0.02, 0.05 and 3.1 MPa. Heating
surfaces were made of nichrome sheets of 0.15 to 0.30 mm thickness with an area of 10 mm ×
40 mm. Two technologies, sintered powders and plasma spraying, were used to make the
nichrome (NiCr) and corundum (Al2O3) porous coatings of thickness δ = 0.15 – 3 mm and porosity
ε = 0.2 – 0.5. Particle diameters Dgrn ranged from 0.63 – 450 µm, the coordination number was
11 11 2
z ≥ 6 [45], the coating permeability K was 0.0014 ∙ 10 – 4.1 ∙ 10 m , the conductivity ranged
–1 –1
from λskl = 0.7 – 3.1 W m K . The wetting angle θ was approximately 40° for water and
close to zero for the ethanol aqueous solution. A zinc (Zn) porous coating, obtained with an
2
electrochemical precipitation method (ε = 0.2, K = 108 m ), was used in investigations for boiling of
liquid helium [154]. The nichrome sheet surface was heated by direct current, so depending on the
material and the porous coating thickness, 70 to 100% of the total heat flux was released [12].

Boiling curves presented in Fig. 4.52 [11, 154] sum up and generalize the results of experimental
investigations discussed in [12, 13, 148, 151, 153, 263, 265]. Covering the heating surface with
a porous layer of low thermal conductivity altered the boiling curve shape considerably when
compared with the shape characteristic of a technically smooth surface also shown in Fig. 4.52.
The changes are as follows:

1. Boiling incipience on the porous layers occurred at lower superheats of the heating
surface, ΔTincip < ΔTincip,sm and lower heat fluxes qincip < qincip,sm than the smooth
surface. Over the initial segment of the boiling curve, the heat transfer coefficient
α = q/ΔTsl is 5 to 10 times higher than for the smooth surface.

2. Two types of nucleate boiling were observed, denoted by Roman numerals I and II in
Fig.4.52. Fig. 4.53 shows a graphic representation of types of boiling found inside the
porous covering.

270 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

log q
III
qcr,nb
q cr,fb

qcr,sm,nb II

q*

q cr,sm,fb
I

q**

DTincip,sm,1 DT* DT incip,sm,fb DT cr,fb

DT incip,1 DTincip,2 DTcr,sm,nb DTcr,nb log DTsl

Fig. 4.52. Shape of boiling curves on heating surfaces with porous coatings of low
thermal conductivity – full lines; boiling curves for smooth surfaces –
broken lines, from Malyshenko and co-researchers [11, 154]

For I type of nucleate boiling, when q < q* < qcr,nb in Fig. 4.52, areas of vapor
generation inside the coating do not coalesce and the liquid remains in contact with
the base surface, i.e. as shown in Fig. 4.52(I). The heat flux in the area I demonstrates
2
a functionality of q ~ ∆Tsl, which is the same as for technically smooth surfaces. The
heat transfer coefficient is 4 to 10 times greater than for smooth surfaces.

I – nucleate I; II – nucleate II; III – film, from Malyshenko [151]

Fig. 4.53. Kinds of boiling inside the porous covering

For the heat flux satisfying the condition q* < q < qcr,nb, areas of vapor generation
inside the coating coalesce, as depicted in Fig. 4.53(II), which causes the liquid layer
to be pushed from the heating surface upwards but remaining within the porous
coating. In zone II shown in Fig. 4.52, the heat flux is linearly proportional to the

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 271


Chapter 4

surface superheat q ~ ΔTsl. The nucleate boiling mechanism of vapor release from the
porous coating, the same as in zone I, continues to function. Heat transfer coefficients
α are 10 to 15 times greater in zone II when compared with a technically smooth
surface.

3. For a particular characteristic heat flux q = q*, a dissipation phase change from type
I to type II nucleate boiling takes place [12, 13, 264, 265]. It is accompanied by a
dramatic increase in relaxation time when q → q*, transient phenomena and I kind
hysteresis occur, see Fig. 4.42b, when q is lowered.

4. Film boiling occurs when a continuous vapor layer fills the inside of the porous coating.
That means a crisis of the II kind of nucleate boiling inside the capillary-porous
structure, as seen in Fig. 4.52. Above the coating, a continuous vapor film is formed,
from which bubbles depart as illustrated in Fig. 4.53(III). Critical heat fluxes, both for
nucleate boiling crisis qcr,nb and film boiling crisis qcr,fb are a few times higher than their
respective heat fluxes for a technically smooth surface. Heat transfer coefficients for
film boiling as in Fig. 4.52(III) are close to the values obtained for technically smooth
surfaces.

5. When the heat flux diminishes, q → q**, a dissipation phase change occurs again, now
from type II to type I nucleate boiling. It is accompanied by a dynamic decrease in the
heating surface temperature to the value characteristic of type I nucleate boiling. After
steady-state is reached, in the interval q** < q < q*, q increases again in accordance
2
with the proportion q ~ ∆Tsl.

Coordinates of the characteristic points of the boiling curve q = q(ΔTsl) in Fig. 4.52 and also the
magnitude and position of the segments corresponding to individual types of boiling depend
mainly on the structural parameters of the porous coatings and the boiling liquid’s thermal
properties.

On the basis of their experimental investigations into porous coatings of sintered powders, Zhang
and Zhao [330] obtained the same boiling curve shape as the one in Fig. 4.52, where the total
capillary-porous structure filling with vapor was assumed to be the nucleate boiling crisis.

272 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

5⋅105
Water
0.1 MPa

105

q, W⋅m-2
5
-1
-2
-3
4
10

3⋅103
3 5 10 50 100 500
DTsl , K

Plasma-sprayed nichrome porous layer, δ = 1.4 mm, ε = 0.5 to 0.6. Blank symbols
– type I nucleate boiling, increasing q; blackened symbols – type II nucleate
5 –2
boiling, increasing and decreasing q: 1 – hysteresis at q* = 2.3 ∙ 10 W m ;
steady-state for type II nucleate boiling obtained after 2 to 3 hrs; 2 – as above,
after waiting 1 hr, the heat flux began to be reduced; 3 – hysteresis for q ≥ 2 q*,
steady-state for type II nucleate boiling obtained after approx. 0.5 hr

Fig. 4.54. Type I hysteresis with memory effects – transition from type I to II
nucleate boiling illustrated in Figs. 4.52 and 4.53 from Malyshenko and
Styrikovich [154]

Regarding type I hysteresis, Andrianov, Malyshenko, Styrikovich et al. [151, 152, 154] also term it
hysteresis with memory for the trends illustrated in Fig. 4.54. This is due to the fact that the shape
of the type II boiling curve in Fig. 4.54 depends on the remembered value of the heating surface
superheat ΔTsc, at which the heat flux starts being decreased. The experimental results presented
by Andrianov, Malyshenko, Styrikovich et al. [151, 152, 154] are in agreement with those obtained
earlier by Poniewski and Wojcik for metal fibrous structures, see Fig. 3.66 [206, 211, 317].

According to the authors [151, 152, 154, 263], type I hysteresis occurs when the capillary-porous
structure is characterized by considerable inhomogeneities with respect to pores filling with vapor.
That occurs for structures of low permeability and thermal conductivity. If openings to facilitate
vapor release are made in the porous coating, type I hysteresis can vanish as seen in Fig. 4.55.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 273


Chapter 4

6
10
5
Water
5⋅10 0.1 MPa

105

q, W⋅m -2
5⋅104
}1
4
10 }2
5⋅103 3

1 5 10 50 100 500
DT sl, K

1 – porous coating without additional openings, type I hysteresis; 2 – porous


coating as above with additional openings to facilitate vapor release, lack of type I
hysteresis; 3 – technically smooth surface

Fig. 4.55. Change in the boiling curve shape due to the presence of openings
facilitating vapor release from the porous coating, plasma spraying,
corundum, δ = 1.4 mm, ε = 0.5 to 0.6. Blank symbols – increasing q;
blackened symbols – increasing and decreasing q from Malyshenko and
co-workers [151, 154, 263]

These experimental results are also compatible with the data obtained by Poniewski and Wojcik
[211, 317] for metal fibrous structures, where type I hysteresis was observed for the coatings of the
greatest thickness, shown in Fig. 3.69. A flat segment of the boiling curve, shown in Fig. 4.55, was
also perceived by Cieśliński [54] for pipes with a flame-sprayed aluminum porous coating.

In their work, Andrianov, Malyshenko, Styrikovich [151, 152, 154, 263] assumed that phase
transition from type I to type II nucleate boiling, shown in Fig. 4.52, corresponds to the percolation
threshold, when the vapor penetrates the capillary-porous structure, i.e. permeates the pores
that became occupied by the liquid at nucleate boiling incipience. This means that pores initially
occupied by vapor and corresponding to the percolation threshold ∏ (4.290) should be excluded
from the spatial lattice of stochastically distributed pores. Under such conditions, the percolation
threshold and the coordination number become:

(4.292)

z* = z (1 – Π) (4.293)

274 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Like in (4.291), the minimum diameter of pores occupied by vapor under type I hysteresis is
computed using the equation:

(4.294)

For a known R*, ∆T*sl, 1 is computed on the basis of equation (4.289). The full heating surface
superheat for the phase transition from type I → II nucleate boiling ∆T*sl is the sum of ∆Tsl,
* 1 and
∆Tsl,
* 2, where the second temperature difference represents the mean temperature drop in the
vapor layer from the heating surface to the evaporation zone. It is assumed that the distance from
the heating surface, δev, is primarily a function of the position of the pore openings in the most
upper layer of the porous coating. According to [11, 151, 152, 154], for a known mean particle
dimension Dgrn, known geometry of particle packing and known skeleton conductivity λszk, it is
* 2 = q* δev /λskl,
possible to assume that for close-packed capillary-porous structures that ∆Tsl,
δev ≈ Dgrn/4. As shown in work [11], values of ∆Tsl
* computed in accordance with the above
procedure are close to experimental data.

4⋅105
Water
2 0.1 MPa

105
-2

8
q, W⋅m

6 1 4
4

3 5
2
2
104
8 100 2 4 6 8 101 2 4 6 8⋅101
DTsl , K

1, 2, 3 – surface with porous coating; 4, 5 – technically smooth


surface (other explanation in the text)

Fig. 4.56. Type I hysteresis on a sintered powder, porous surface made of 1H18N9T
steel with δ = 1.1 mm, ε = 0.67, Dgrn ≈ 100 µm and Dp ≈ 55 µm from Kovalev
and Lenkov [118]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 275


Chapter 4

Kovalev and Lenkov [118] also observed hysteresis of the I kind for boiling water on the external
surface of 1H18N10T alloy steel pipes, with the results shown in Fig. 4.56. The curves 1, 2, 3 in
Fig. 4.56 differ depending on the time duration at which the critical heat flux qmax is maintained.
For curve 2, the duration is 15 minutes, which causes an increase in the heating surface
temperature by 30 K, while for curve 3 it is 30 minutes, which produces an increase in ΔTsc of 75K.
The effect of time duration can be explained as follows: intralayer boiling crisis or capillary crisis
takes place when the capillary forces holding the liquid in place are smaller than the resistance
to liquid and vapor flow through the porous layer. This means that condition (4.55) is not satisfied.
They interpret the I kind hysteresis phenomenon in a way similar to Andrianov, Malyshenko,
Styrikovich et al. [11, 151, 152, 154, 263], see Fig. 4.53. Namely, they treat the formation of a vapor
film inside the porous coating at the heating surface as an internal boiling crisis [118].

In their successive works, Kovalev and Lenkov [122, 127, 128] present generalized boiling curves
characteristic of porous surfaces as shown in Fig. 4.57. They consider nucleation hysteresis,
segment A”–A of curve 2, to be characteristic of capillary-porous structures of small pore
diameters, which is also the reason why ΔTsl,A” is greater for porous surfaces than technically
smooth ones. Point E denotes the beginning of the nucleate boiling crisis.

E
log q

2
D
C

H F
1

A'
A G
A"
1 - smooth surface

log DTsl

Fig. 4.57. Generalized boiling curves for capillary-porous structures of Kovalev and
Lenkov [122, 127, 128]

If it is difficult for vapor to escape from the capillary-porous structure, so reducing the heat
flux takes place along the curves EFG or DHB. The curve BFE occurs for structures of low
permeability. According to Kovalev and Lenkov [122, 127, 128], the hysteresis loops EFG and DHB
result from the capillary hysteresis phenomenon, which consists in restricting the liquid flow from
small diameter pores to the larger ones by surface tension forces.

276 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Capillary hysteresis, according to the description given by Chejfec and Nejmark [45], results from
pressure differences for the adsorption (condensation) and desorption (evaporation) processes
occurring on heating surfaces of variable geometry and under conditions of non-ideal wetting.
Hence, according to Kovalev et al. [115, 118], intralayer boiling crisis or capillary crisis occurs
when the capillary forces holding the liquid in place are smaller than the resistance to liquid and
vapor flow through the porous layer, i.e. when condition (4.55) is not satisfied.

Metal fibrous capillary-porous structures

I kind hysteresis was also analyzed by Poniewski and Wojcik, whose results are depicted in Fig. 3.66
and Fig. 3.68, taken from [206, 211, 317]. It was observed only for water and the thickest porous
coatings investigated, which were for δ ≤ 2 mm.

Their characteristic boiling curve, for which the hysteresis phenomenon is discussed, is
presented in Fig. 4.58. Like for other porous coatings discussed earlier, nucleate boiling begins at a
superheating that is smaller than a technically smooth surface (ΔTsl below 2 K). A steep increase
in the heat flux with increasing superheat (with respect to a smooth surface) is a characteristic
process feature for low temperature differences (range I in Fig. 4.58). The heat transfer coefficient
4 –2
at low heat flux (q ~ 5 ∙ 10 W m ) is over 5 times greater than that for the smooth surface,
resulting from its high density of active nucleation sites and high frequency of bubble formation.

5 –2
After the heat flux reaches the critical heat flux, q ~ 1.7 ∙ 10 W m in Fig. 4.58, a continuous
increase in the heating surface temperature sets in, denoted as range II. It is a slow process
in which the temperature increases by about 1.5 to 30 K/h. The flat trend in the boiling curve
indicates a fall in the heat transfer coefficient value. A similar phenomenon was observed earlier
by Andrianov, Malyshenko, Styrikovich et al. [11, 151, 154] in Figs. 4.52, 4.54 and 4.55 and also by
Kovalev and Lenkov [118] in Fig. 4.56.

If even a small step increase in the heat flux is applied in range II and this heat flux is maintained
as shown in range III, the heating surface temperature will increase in an uncontrolled way, like in
range II; moreover, the rate of temperature rise is higher than in range II. A similar phenomenon
was observed by Malyshenko and Styrikovich [154]. A decrease in the heat transfer coefficient in
ranges II and III indicates that a vapor layer is formed inside the capillary-porous structure, see
Fig. 4.53.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 277


Chapter 4

2⋅106
Water
106
0.1 MPa

III
II

q, W⋅m-2
105
I IV
V
1
2
4
10
1 10 100
DTsl , K

1 – copper, fibrous capillary-porous covering, ε = 0.4, δ = 2 mm;


2 – technically smooth surface, copper from Poniewski and Wojcik [211]

Fig. 4.58. I kind hysteresis

After the heat flux is reduced in range III, the transition to lower surface temperature proceeds
along a different path that in the case of increasing heat flux, as shown by the range IV that is
shifted to much higher ΔTsl values. Range IV’s heat transfer coefficient is nearly constant and
thus represents a nearly constant thermal resistance in this segment of the boiling curve, i.e.
this can be interpreted that a vapor layer of constant thickness is being maintained. The same
phenomenon is observed after the heat flux is reduced in range II, creating range V in Fig. 4.58.
Thus, a “memory” effect occurs, noticed earlier by Andrianov, Malyshenko, Styrikovich et al. [151,
152, 154], and shown in Fig. 4.54. If the heat flux is reduced in range I, however, no inclination
of the boiling curve is found and the heating surface temperature follows the same curve as for
increasing heat flux. The description of the boiling curve in Fig. 4.58 can be considered typical of
curves with I kind hysteresis.

4.3.4. II kind hysteresis

Experimental data analysis

Nucleation hysteresis, shown in Fig. 4.42a and I kind hysteresis, shown in Fig. 4.42b, are the
most frequently encountered types of hysteresis observed for various capillary-porous structures
and different boiling liquids. Tehver et al. [273, 276] also analyzed a second kind of hysteresis,
illustrated by their results summed up in Fig. 4.59. Tehver et al. [273, 276], like Kovalev and Lenkov
[118], treat I kind hysteresis as an intralayer nucleate boiling crisis. Apart from hysteresis loops
previously recorded by other researchers, Tehver [273] also obtained a loop of “simple” hysteresis

278 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

given by segments AA’B’ in Fig. 4.59. This indicates the possibility of different kinds of nucleate
boiling when the heat flux is much higher than the maximum qmax (or the critical qcr) value.

q D

C'
B'

B
A'
A
DTsl

AA’B – nucleation hysteresis; AA’B’ – simple hysteresis loop;


CDEC’ – inverse hysteresis loop

Fig. 4.59. Hysteresis phenomena on surfaces with porous coverings


observed by Tehver et al. [273]

A simple hysteresis - II kind hysteresis loop was plotted on the basis of the results of experimental
investigations on R-113 boiling on a flat aluminum surface with a plasma-sprayed bronze porous
coating shown in Fig. 4.60 [273, 274, 278, 281]. The boiling curve “a” in Fig. 4.60 originates in the
situation when the heat flux solely increases. Lowering the heat flux at an arbitrary point q, ΔTsc of
the “a” curve leads to a new boiling curve “c”, along which it is possible to proceed in both directions
of heat flux changes. The nucleate boiling process in the “c” curves continues for small heat fluxes,
4 –2
qnb ≈ 1 to 2 ∙ 10 W m . The position of these curves depends solely on the initial values of q*,
∆Tsl
* in curve “a”. In Figs. 4.60 and 4.61, the boiling crisis hysteresis is also visible. The highest
curves in Fig. 4.60 (curve “b”) and Fig. 4.61 are obtained after previously attaining conditions
characteristic of film boiling. The position of curve “a” and its shape depend on the time after
which successive heat flux values are changed, which is illustrated in Fig. 4.61.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 279


Chapter 4

5 R 113
3⋅10 0.1 MPa

q↑
q↓
2⋅105 b a
q, W⋅ m-2
c

1⋅105 q*

qnb

0 DT*
nb 5 10 15 DT*sl 20
DTsl, K

δ = 0.15 mm, ε = 0.59, Dp = 5.1 µm, a, b and c curves are described in the text,
from Tehver and co-workers [274, 278, 281]

Fig. 4.60. Boiling curves with II kind hysteresis loops on a flat aluminum heating
surface with a plasma-sprayed bronze porous coating

Tehver et al. [273, 274, 278, 281] attribute changes in the shape of boiling curves, Figs. 4.60 and
4.61, to changes in the numbers of active pores. This results from a non-uniform distribution of
pore dimensions in the porous coating. The concept was developed by Poniewski and Wojcik in
their II kind hysteresis model [203, 205 – 208, 211, 214, 216, 218, 219]. A distinctive feature of II
kind hysteresis is an increase in the value of the heat transfer coefficient for boiling curves of “c”
type in Fig. 4.60, i.e. when the heat flux is reduced.

Qualitative results compatible with experimental data by Tehver et al. [273, 274, 278, 281] were
obtained by Kravchenko and Ostrovskij [132] for acetone, ethanol and also an acetone and ethanol
mixture. Investigations were conducted for tubular surfaces of stainless steel, 6 mm in diameter,
covered with plasma-sprayed porous coating of stainless steel or nickel. The coating thicknesses
were from 45 to 200 µm with a porosity ε ≈ 0.15. Both nucleation hysteresis and II kind hysteresis
were observed as shown in Fig. 4.62.

Kravchenko and Ostrovskij [132], like Tehver et al. [273, 274, 278, 281] point to a non-uniformity
of pore diameter distribution as the reason for hysteresis phenomena. Another factor strongly
affecting nucleation hysteresis is the wetting angle of the boiling liquid on a given heating surface
and the liquid viscosity. Therefore, strong hysteresis occurs for acetone as shown in Fig. 4.62,
whereas it is much weaker for ethanol and does not take place at all for water [132].

280 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

R 113
3⋅105
0.1 MPa
t = 1 minute
t = 5 minutes
t = 30 minutes
t = 3 hours
2⋅105
-2
q, W⋅m

1⋅105

0 5 10 15 20
DTsl , K

Fig. 4.61. Boiling curves for the heating surface in Fig. 4.60 at different times of the
heat flux changes, from Tehver and co-workers [274, 278, 281]

4·10 1
Acetone
2
1
0.1 MPa
10
8
6
α, W⋅m-2⋅K-1

-1
2 -2
-3
100 -4
8 -5
6 -6
4·10-
1
6⋅103 10 4 2 4 6 105 2 4⋅10 5
q, W⋅m-2

Blackened symbols indicate q↑, blank symbols indicate q↓; 1, 2 – δ = 60 µm;


3, 4 – δ = 90 µm; 5, 6 – 45 µm, from Kravchenko and Ostrovskij [132]

Fig. 4.62. Nucleation hysteresis and II kind hysteresis

Ayub and Bergles [18] found II kind hysteresis for water at atmospheric pressure on a Gewa-T
type tube, where re-entrant channels between its T-fins were filled with polystyrene foam. They
put forward a hypothesis [18] that hysteresis can be caused by the fact that once activated, pores

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 281


Chapter 4

remain active when the heat flux is reduced. The same interpretation of II kind hysteresis is
assumed by Poniewski and Wojcik [205 – 208, 211].

Wang et al. [132] term II kind hysteresis as temperature deviation hysteresis, TD hysteresis. These
authors encountered it for water on flat surfaces whose roughness was increased by using an
abrasive paper. The physical interpretation of hysteresis put forward by Wang et al. [132] is the
same as that of Ayub and Bergles [18] and Poniewski and Wojcik [205 – 208, 211]. Vasilev et al.
[302] also observed II kind hysteresis for propane boiling on a cylindrical surface with a porous
coating of small thickness: δ = 0.3 mm, porosity ε = 0.17 and reduced pressure p/pcr = 0.111.

Poniewski et al. model – experiment, theory, computation results

Poniewski, Wojcik and coworkers [203, 205 – 208, 211, 214, 216, 218, 219] conducted extensive
experimental and theoretical investigations into II kind hysteresis for copper fibrous porous
coatings. As it was already shown in Figs. 3.67 and 3.68, II kind hysteresis occurred for coatings
of small thickness (δ ≤ 0.6 mm) and when the wetting angle was smaller than that of water, i.e. for
R-113 and ethanol. II kind hysteresis was not found to occur for water in their studies. The features
of II kind hysteresis are discussed using the results for ethanol as the example, shown in Fig. 4.63.

Like in the works discussed earlier [18, 132, 273, 274, 278, 281, 307], hysteresis occurs while
reducing the heat flux that has not exceeded the maximum values. The transition to lower values
of the surface temperature proceeds along a different path than that for increasing the heat flux.
The boiling curve is shifted towards lower ΔTsc values. The character of the boiling curve shape,
1
q ~ ∆Tsl, in segment “a”, Fig. 4.63, points to practically invariable thermal resistance. While further
6
increasing the heat flux, at point “A”, the boiling curve inclination changes to q ~ ∆Tsl. That means
the number of nucleation centers grows with increasing the heat flux, range “b”.

Under the conditions of pore diameter distribution, a successive activation of cells of smaller
dimensions takes place inside the capillary-porous structure when the wall superheating grows.
Approximate computation indicates (equation (4.295)) that at point “A”, Fig. 4.63, a delayed
activation of pores inside the structure takes place.

Hysteresis is revealed after the heat flux is reduced in range “b”, Fig. 4.63. It means that when
additional nucleation centers have been activated, they remain active after the heat flux is
reduced. The phenomenon does not occur for a smooth surface in pool boiling. Results obtained
by Wang et al. [307] make an exception.

The boiling curve segments in range “c” are stable in temperature ranges given in Fig. 4.63, as
the heat flux increases and decreases. The transition to the boiling curve located higher in range
“c” is possible via the boiling curve “b”. Obtaining smaller heat fluxes for the same temperature
difference (the lower curve in ranges “c” or curve “a”) is possible only when boiling decay is
reached. Increasing the heat flux higher in range “b” leads to the occurrence of the so-called

282 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

intralayer boiling crisis – range “d”. It is characterized by an intrinsic increase in the heating
surface temperature at constant heat flux. In all recorded cases of II kind hysteresis, ranges ”a”, “b”
n
and “c” of the boiling curve were observed. In ranges “b”, the exponent n in the dependence q ~ ΔT
was varied from 3 to 9 [203, 205 – 208, 211, 214, 216, 218, 219].

With the analysis of experimental data concerning 9 thin-layered samples of three porosities
(0.4, 0.7 and 0.85) and two liquids (ethanol and R-113), it was possible to determine characteristic
ranges of the boiling curves represented by a full line in Fig. 4.64. The dotted line represents
ranges at very low superheating (ΔT ~ 1K), not shown in experimental curves because of the
distortion of the results by the measurement error at such small superheats. Nevertheless, their
occurrence and shape result from the heat transfer process in thin-layered capillary-porous
structures.

2⋅106
Ethanol
10 6
0.1 MPa
d
b
c
-2
q, W⋅m

a
10 5
a

1
2
10 4
1 10 100
DTsl, K

1 – copper fibrous porous covering, ε = 0.40, δ = 0.2 mm; 2 – copper, technically


smooth surface, from Poniewski, Wojcik and coworkers [211]

Fig. 4.63. II kind hysteresis

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 283


Chapter 4

c d
b

o A
k
a

DT1 DT2 DT3 DT4 DTsl

D(DTdel )

Fig. 4.64. Characteristic boiling curve ranges for ethanol and R-113 boiling
and thin-layered capillary-porous structures from Poniewski,
Wojcik and coworkers [208]

First of all, range “k” in Fig. 4.64 covers the liquid convection and conduction in the porous layer.
In range “o”, activation of pores on the external surface of the porous layer begins. For a non-
homogeneous pore dimension distribution, a gradual activation of nucleation sites of increasingly
smaller diameter Da takes place when the heating surface temperature increases. In accordance
with (4.273), substituting R = Da /2 we get:

(4.295)

which is illustrated in Fig. 4.65 for three investigated liquids.

Due to the mechanism of II kind hysteresis, ranges “a”, “b” and “c” are of special importance, as
discussed below:

Range a. Nucleate boiling clearly manifests itself in this range. Segment “a” of the boiling curve
1
(q ~ ∆Tsl) with its linear relationship points to a constant thermal resistance and constant number
of active cells on the external surface of the porous coating in Fig. 4.66.

Range b. An abrupt change in the slope of the boiling curve at point “A” indicates that the number
of nucleation sites is growing with increasing heat flux, so N = f (q). Successive cells of smaller
and smaller dimensions get activated inside the capillary-porous structure, as illustrated in
Fig. 4.66b.

284 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Fig. 4.65. Magnitude of nucleation site diameters as a function of the heating surface
superheating from Poniewski, Wojcik and coworkers [208]

Fig. 4.66. Heat transfer mechanisms in ranges “a” and “b” from Poniewski,
Wojcik and coworkers [208]

Range c. Hysteresis phenomenon, represented by curves “c” in Figs. 4.63 and 4.64, is revealed
after the heat flux is reduced in range “b”. The shape of the boiling curve indicates that the pores
activated in range “b” remain active, despite the fact that the heat flux has been reduced. This is
the physical cause of II kind hysteresis.

Building their model of boiling with II kind hysteresis in ranges “b” and “c”, the authors relied on the
works of Smirnov [5, 247], who presented a model of vapor generation in regular capillary-porous
structures (mesh structures). In his model, Smirnov assumed that the thermal resistance of the
liquid microlayer covering vapor channels determines the rate of heat transfer. The model was

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 285


Chapter 4

extended to cover the case of thin-layered structures by making an assumption on the finite height
of the equivalent microfin, see Fig. 4.13, and non-homogeneity of the pore dimensions.

F, F1 and f1 denote respectively: the total area of the heating surface, the area occupied by
convection and conduction and the fraction of the heating surface occupied by the quoted heat
transfer mechanisms.

For very low superheats, ΔTsc < ΔTsc1, (see range “k” in Fig. 4.64), on the whole capillary-porous
structure surface F = F1, since f1 = F1/F = 1, so the heat flux is given by:

q1 = αc+c ΔTsl (4.296)

where αc+c denotes the overall heat transfer coefficient for conduction and convection.

Beginning at ΔTsl = ΔTsl1, another heat transfer mechanism comes into play – boiling on the
external surface of the structure, as noted earlier in the description of Fig. 4.64 for range “o”.

As the superheat grows, the active sites cover a greater and greater fraction f 2 of the total surface
F. Assuming the notation γ = γ (Dp) for the function of pore diameter distribution density, the
fraction of the porous surface affected by boiling is

(4.297)

where the minimum diameter Da of active pores depends on the superheating ∆T*sl, in accordance
with (4.295). If the geometric parameters of the capillary-porous structure are known, it is possible
to determine characteristic superheats in Fig. 4.64, inserting the maximum and the minimum
pore diameters Dmax and Dmin, respectively in equation (4.295).

Due to a very small thickness and relatively high thermal conductivity of the coating, the thermal
resistance of conduction is negligibly low. Thus, the heat transfer process in the area F2 is
described by the dependence:

q2 = αsb ΔTsl (4.298)

286 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where αsb is the experimentally determined heat transfer coefficient for boiling on the external
surface of the structure. Taking this into account in the superheat range ΔTsl1 < ΔTsl < ΔTsl2, i.e.
segment “o” in Fig. 4.64, the sum of heat fluxes on the surface amounts to:

q = q1 (1 – f2) + q2 f2 (4.299)

In range “a”, i.e. when ΔTsl2 < ΔTsl < ΔTsl3 and F = F2 (f2 = F2 /F = 1), the total heat flux takes on
the form:

q = q 2 (4.300)

In accordance with Smirnov’s work [5, 247], in zone “b” the vapor generation takes place on the
external surface of the superheated liquid microlayer covering a vapor channel surface. A vapor
channel together with the adjacent skeleton constitutes an elementary cell regarded as a fragment
of the finned surface, see Fig. 4.13, while and mutual interaction of the microfins, however, is not
accounted for.

Equivalent dimensions of horizontal cross-sections, microfin – d and cell – Dp, are determined
as geometric means of their extreme dimensions (the minimum and maximum width for sections
whose shape is similar to a rectangle or an ellipse). Areas of those sections will be proportional
2
to d and D2p, respectively, with the same proportionality coefficients. If the microfin height and
pore depth (equal to the coating thickness δ) are thought to be constant, the porosity ε is assumed
to be the ratio of the area of the pore horizontal sections to the total area. Thus on the basis of
measurement of pore diameters, it is possible to determine the microfin dimension d:

(4.301)

It is assumed that the microfins are covered with a liquid microlayer of the thickness δc that is
constant along the whole of their height. Using the condition (4.55), that the sum of pressure drops
in the vapor and liquid streams moving in opposite directions should be smaller or equal to the
capillary pressure increment 4 σlg /Dp, the boundary thickness of the liquid microlayer δl is stated
by equation (4.62) according to [5, 247].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 287


Chapter 4

In order to get a better fit of the Smirnov et al. model [5, 247] to the boiling mechanism in thin-
layered metal fibrous coatings, a finite microfin height is assumed, which yields the following
boundary conditions for the heat conduction equation (4.43):

for (4.302)

After solving the boundary value problem (4.302) and finding the function ϑ = ϑ(x), the
cell
dependence q = –λ (dϑ/dx)x=0 is used to determine the heat flux q transferred in boiling in a
single cell:

(4.303)

λl
For simple conduction across a liquid film yields α = δ (4.42) and δ1, in accordance with formula
l
cell cell
(4.62), depends on q , so (4.303) has an implicit character (i.e. q cannot be computed explicitly)
because of the heat flux transferred in the cell. Applying a MATHCAD software package, which
solves implicit equations, approximate solutions to equation (4.303) can be found for each set of
geometric and physical parameters. The equation describes quantitatively the heat transfer process
in a single cell, in which boiling occurs.

Like was the case for the external side of the porous coating for range “o” discussed above, a
gradual activation of cells of smaller and smaller diameters D*a takes place inside the coating as
well. Additionally, due to the “flooding” of potential nucleation sites when the boiling liquid has
a small wetting angle, the activation of pores inside is delayed in accordance with the following
relationship:

(4.304)

where Δ(ΔTdel) is the delay of pore activation inside the capillary porous structure determined
experimentally, see Fig. 4.64. Hence, the fraction of the surface affected by internal boiling is:

(4.305)

288 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

whereas the heat flux received from active cells equals:

(4.306)

where the member 1 – ε is proportional to the area occupied by the microfins.

On the basis of the model’s assumptions, a conclusion can be drawn that the sum of the heat
fluxes in range “b” of the boiling curve can be computed using the relationship:

q = q2 (1 – f3) + q3 f3 (4.307)

Summing up the reasoning so far, the heat transfer process on the thin-layered porous surface in
ranges “k”, “o”, “a” and “b” of the boiling curve can be described with one equation:

q = [q1 (1 – f2) + q2 f2] (1 – f3) + q3 f3 (4.308)

where quantities q1, q2 and q3 as well as F2 and F3 are described by equations (4.296), (4.297),
(4.298), (4.305) and (4.306), respectively.

Fig. 4.67. Occurrence and scope of individual heat transfer mechanisms


for successive ranges of the boiling curve, from Poniewski and
coworkers [208]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 289


Chapter 4

Fig. 4.67 shows, in a symbolic way, the sequence and co-occurrence of individual boiling
mechanisms in the heat transfer process together with the fraction of the total area they occupy
for consecutive ranges of the boiling curve.

If the heat flux is reduced before the superheat ΔT4 is reached in Fig. 4.64, in range “c” the cells
participating in the boiling process remain active, although their diameters can become smaller
than Da*. In other words, if the heat flux reduction takes place at the superheat ΔThst, the area
occupied by boiling inside the structure does not change and equals Fhst. A constant fraction of
the active area fhst = Fhst /F can now be computed by inserting Da* = Dhst
* into (4.305).

Thus, the dependence of the heat flux on the superheat in range “c” of the boiling curve can be
computed on the basis of the following summation:

q = q2 (1 – fhst) + q3 fhst (4.309)

The above equation together with equation (4.308) gives the mathematical description of the
boiling curve with II kind hysteresis.

In order to determine the function of pore diameter distribution γ(Dp) necessary to compute
the area fraction occupied by active pores, (4.297) and (4.309), statistical investigations were
conducted in accordance with the procedure described in section 3.2.5 [33, 159, 171]. Parameters
of investigated porous coatings together with the results of tests of chi-square empirical
distributions of consecutive sets of pore diameters compatibility with theoretical probability
distributions are presented in Table 4.1.

Table 4.1. Significance level of chi-square test for some theoretical distributions

Significance level lower than 0.05 is denoted as “rejected”. Basic parameters of


investigated fibrous copper structures: 1 – δ = 0.6 mm, ε = 0.85; 2 – δ = 0.35 mm,
ε = 0.85; 3 – δ = 0.6 mm, ε = 0.40; 4 – δ = 0.35 mm, ε = 0.4; 5 – δ = 0.2 mm, ε = 0.40.

Coating Logarithmic- Chi-


Uniform Exponential Normal Gamma
no. normal square
1 rejected rejected 0.1522 0.2023 0.554 rejected
2 rejected rejected 0.1882 0.414 0.4041 rejected
3 rejected rejected 0.1226 rejected 0.0826 rejected
4 rejected rejected 0.6384 rejected 0.6247 rejected
5 rejected rejected 0.7403 0.7278 0.7423 0.649

The following families of probability distributions: uniform, exponential, normal, logarithmic-


normal, gamma and chi-square were accounted for in the testing process [33, 159, 171]. Due to
the small thickness δ ≤ 0.6 mm of the porous coatings that underwent statistical investigations,

290 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

the pore diameter distribution on the surface could be considered representative of the whole
coating volume. For the sake of modeling diameter distribution in thin porous fibrous structures, a
gamma distribution is chosen as shown in Fig. 4.68. Such a choice has been made because of the
on average higher significance value in chi-square goodness of fit test (Table 4.1) and the non-
negative set of gamma variable values. The probability density of gamma distribution has the
form [33]:

(4.310)

where w and z are positive constants that characterize the distribution. They are bound by the
value Dp and the mean standard deviation s in the following way:

(4.311)

Fig. 4.69 shows an example of the empirical distribution of the equivalent diameter of the
nucleation center (pore) and the density of fitted gamma distribution.

1.00
Empirical distribution function

0.75

Covering
0.50 number:
-1
-2
-3
0.25 -4
-5

0
0 0.25 0.50 0.75 1.00
Theoretical gamma distribution function

Fig. 4.68. Probability for gamma distributions and empirical distributions for samples
1, 2, 3, 4 and 5 in Table 4.1, from Poniewski and coworkers [208]

An analysis of the effects of the parameters Dp and s of the pore diameter distribution density on
the heat transfer process of the copper fibrous porous coating with ε = 0.4 and δ = 0.2 mm was
made and is shown in Figs. 4.70a and 4.70b. An increase in the mean value of pore diameters Dp,
when other quantities are unchanged, accelerates pore activation and results in the occurrence

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 291


Chapter 4

of full boiling development (f3 ≈ 1) at smaller heating surface superheats – range “b” in Fig. 4.64
– 4.70a. It should be added that the impact of the mean value of pore diameter Dp on boiling heat
transfer in range “o”, in Fig. 4.64, is analogous to that in range “b” due to assuming a similar model
of surface boiling propagation.

Fig. 4.69. Empirical distribution of pore diameters with respect to fitted gamma
distribution for coating 2 in Table 4.1 from Poniewski and coworkers [208]

A fall in diversification norm of horizontal pore sections or, in other words, a fall in the mean
standard deviation of pore diameters s, while other parameters are unchanged, also accelerates
pore activation, i.e. reduces the time interval ΔTsl for which f3 ≈ 1 is obtained in Fig.4.70b. In
accordance with the assumed model, this means that the limited decrease in pore dimension
scatter should lead to a simultaneous and step activation of all pores.

Fig. 4.71 shows the results of computation and experimental data for three selected cases of II
kind hysteresis. The obtained results are qualitatively and, to a considerable extent, quantitatively
in agreement with experimental results.

292 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

a)
1.0

↑ Dp
f 3 (ΔTsl , Dp1 )
f 3 (ΔTsl , Dp 2 )
f 3 (ΔTsl , Dp 3 )
0.5
f3
f 3 (ΔTsl , Dp 4 )

0
10.5 11.0 11.5 12.0
DTsl, K
b)
1.0

↑s

f 3 (ΔTsl , s1 )
f3

0.5
f 3 (ΔTsl , s 2 )
f 3 (ΔTsl , s3 )
f 3 (ΔTsl , s 4 )

0
10.5 11.0 11.5 12.0
DTsl, K

a) changeable value of the mean Dp and the constant mean standard deviation
1 2 3 4
s = 6 µm; Dp = 13 µm; Dp = 15 µm; Dp = 17 µm; Dp = 19 µm; b) changeable mean
standard deviation s and the constant value of Dp = 18 µm; s1 = 3 µm; s2 = 5 µm;
s3 = 7 µm; s4 = 9 µm. Calculations are for sample 2, Table 4.1, ε = 0.4;
δ = 0.2 mm with gamma distribution, from Poniewski and coworkers [208]

Fig. 4.70. Dependence of area fraction f3 occupied by boiling inside the porous coating
on the heating surface superheat ΔTsl as the function of Dp and s

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 293


Chapter 4

106 106
R 113 R113
0.1 MPa 0.1 MPa

q, W·m-2

q, W·m-2
105 105

104 104
1 10 100 1 10 100
DTsl , K DTsl, K

106 106
Ethanol Ethanol
0.1 MPa 0.1 MPa

q, W·m-2
q, W·m-2

105 105

104 104
1 10 100 1 10 100
DT sl, K DTsl, K

106 106
R 113 R 113
0.1 MPa 0.1 MPa
q, W·m-2

q, W·m-2

105 105

104 104
1 10 100 1 10 100
DTsl, K DTsl, K

Fig. 4.71. Comparison of experimental and theoretical boiling curves with II kind
hysteresis, from Poniewski and coworkers [208]

294 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

II kind hysteresis application to heat transfer control

Every kind of hysteresis is considered a disadvantageous phenomenon that impedes thermal


stabilization of systems emitting high heat fluxes. An attempt at practical application of II kind
hysteresis, investigated and patented by Poniewski and co-researchers [219], provides an
exception. Owing to stability and repeatability of the boiling curves obtained with II kind hysteresis,
see Fig. 3.67, it is possible to develop a new approach to the control of boiling heat transfer on
heating surfaces covered with porous layers. Both above-mentioned features of II kind hysteresis are
responsible for terming it “controlled hysteresis”.

The phenomenon can be applied to control the heating surface temperature. The heat source that
is to be controlled is cooled due to the boiling process on the surface covered with a capillary-
porous structure. The structure’s characteristics are selected in such a way so that II kind
hysteresis is obtained. The means of control is presented in Fig. 4.72.

q
B

qd
b
q1
c
qn
A

DTn DT2 DTsl

Fig. 4.72. Means of heat transfer control of Poniewski and co-researchers [219]

When the heat flux emitted by the source changes (for instance from qn to q1), it is always possible
to find a point “B” on curve “b” such that, after the heat flux is lowered another time, the original
constant temperature difference between the heating surface and the liquid saturation ΔTn is
re-established. This process requires turning on an additional heat source emitting the heat flux qd
for the time necessary to reach point “B” [219].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 295


Chapter 4

4.4. Boiling crisis and the critical heat flux

4.4.1. Definitions, experimental data and correlations

Industrially manufactured high performance boiling surfaces, such as High Flux, Thermoexcel-E
and others as well as capillary-porous structures imposed with various techniques do not only
enhance boiling heat transfer but can also alter the critical (maximum) heat flux when compared
with technically smooth surfaces, as noted in Chapter 3 [25, 69, 141, 330, 332]. An increase
in qcr on microporous surfaces is thought to be caused by (i) the smaller diameter of departing
vapor bubbles and higher departure frequency, (ii) the greater number of nucleation sites than on
smooth surfaces and (iii) a damping of bubble coalescence and hence an added resistance to
the formation of a vapor film, which might hamper liquid inflow into the porous coating [8, 55, 58,
86, 134].

The boiling crisis on porous coatings cannot be explained in the same way as it is for smooth
surfaces. Results of investigations presented earlier, see Figs. 3.37, 3.43, 3.66, 3.67, 4.54, 4.55,
4.56, 4.58, 4.62, 4.63 from [85, 117, 118, 132, 151, 154, 211, 293, 294], indicate that the heating
surface temperature increase is not so abrupt and the heat transfer coefficient decreases less
drastically for microporous surfaces. The critical heat flux density qcr for microporous surfaces
is therefore often termed the maximum density qmax [3, 55, 58, 172, 251, 323]. Other proposed
names are: qburnout [275] and qboundary for boiling under the conditions of capillary feed [292, 295].

The relaxation time for the heat flux q ≤ qcr, which is understood to be the time interval between
the instant of q increase and that of abrupt growth in the heating surface superheat ΔTsl, does not
usually exceed 1 to 2 seconds for technically smooth surfaces, whereas it can range from 100 to
over 1000 seconds for microporous surfaces [55, 134, 152, 263]. According to Kravchenko et al.
[134], the activation of new nucleation sites in capillary-porous structures takes place gradually as
vapor nuclei propagate inside the structure, which can significantly prolong the relaxation time. For
example, the shapes of boiling curves obtained by Fedorov [85] in Fig. 3.37, Tolubinski et al. [293,
294] in Fig. 3.43, Poniewski and Wojcik [211] in Figs. 3.66 and 4.58, Malyshenko et al. [11, 154] in
Fig. 4.54, and also Kovalev and Lenkov [118] in Fig. 4.56 point to extended relaxation times.

Malyshenko, Styrikovich et al. [151, 152, 154, 263] also report that when q ≈ qcr, the rate of the
increase in the temperature of the surface coated with capillary-porous structures was at least ten
times lower than for a smooth surface. The rate of vapor film propagation on the external surface
–5 –1
of the porous coating under those conditions was 10 ms [151].

296 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Furthermore, the definitions of the notion of nucleate boiling crisis show some divergence as a
great variety of industrially manufactured enhanced microsurfaces and imposed capillary-porous
structures results in a number of differences in the character of the phenomenon [58, 117, 118,
141, 151, 152, 263, 317, 330]. Describing the phenomenon, for instance, researchers mention a
fall in the heat transfer coefficient value [141], the structure filling with vapor [65], intralayer boiling
crisis or capillary crisis preceding the nucleate boiling crisis [115, 118, 138, 154, 287] or the heat
fluxes corresponding to the maximum heat transfer coefficient [269].

Such a state results from the multiplicity of factors affecting the boiling crisis, such as

1. the kind of industrial surface or capillary-porous structure,

2. the structure’s thickness, size and distribution of pore diameters,

3. the thermal properties of the boiling liquid,

4. and the manner in which the porous coating is fed with the liquid.

By analogy with technically smooth surfaces, the formation of a stable, continuous, thin vapor
layer above the external surface of a capillary-porous structure is most frequently regarded as the
incipience of the nucleate boiling crisis [151].

Industrially manufactured developed microsurfaces

The review of the maximum heat flux densities qmax, obtained for industrial enhancements
Gewa-K, Gewa-T, Thermoexcel-E, High Flux and others can be found in the works by Yilmaz
and Westwater [323] and also Czikk, O’Neill and Gottzmann [69]. Yilmaz and Westwater [323]
determined qmax for isopropanol and p-xylene on surfaces of tubes heated inside with steam.
Fig. 4.73 shows experimental boiling curves for these industrial enhanced tubes. The maximum,
in the majority of curves, is rather flat but, at the same time, clearly defined, in accordance
with [323].

5
Table 4.2 presents their experimental data for isopropanol at a pressure of 1 ∙ 10 Pa. The
maximum heat flux density qmax,ind and the heating surface superheating ΔTsl,ind for industrial
surfaces are compared with qmax,sm and ΔTsl,sm for technically smooth tubes. The ratios for
Gewa-T, Thermoexcel-E, ECR-40 and High Flux tubes show that qmax,ind is from 1.4 to almost
twice as high as qmax,sm, whereas for the two others, CSBS and Gewa-K, it is lower by approx.
10% and 35%, respectively. In the majority of cases, the maximum heat flux density qmax,ind
occurs at superheats lower than those for technically smooth surfaces.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 297


Chapter 4

Table 4.2. Experimental data for tubes with industrial enhancements compared
with technically smooth pipes extracted from the work of Yilmaz and
Westwater [323]

qmax,ind ΔTsl,ind
Microsurface kind
qmax,sm ΔTsl,sm
Gewa-K 0.643 0.873
Gewa-T 1.468 0.915
Thermoexcel-E 1.470 1.541
CSBS 0.898 0.432
High Flux 1.968 0.687
ECR-40 1.404 0.730

6⋅10 6
4 Smooth CSBS
GEWA-T ECR-40
2 GEWA-K HIGH FLUX
6 THERMOEXCEL-E
10
6
-2

4
q, W⋅m

10 5
6
4
Isopropanol
2
1⋅10 5 Pa
104 0
10 2 4 6 10 1 2 4 6 102 2⋅102
DTsl, K

Fig. 4.73. Boiling curves for industrial enhanced boiling tubes from Yilmaz and
Westwater [323]

Results of investigations into the boiling properties of the High Flux microporous surface for
R-113, R-114, FC-88 and trichloroethylene are presented in Table 3.1 [69]. The heat flux qmax
for this surface with these fluids is at least equal to or greater than qmax for smooth surfaces.
The increase amounts to a minimum of 15% – 20%, a maximum up to 70% and is independent of
5 –2
the structural microporous surface parameters. A considerable scatter, from 1.84 ∙ 10 W m to
5 –2
5.74 ∙ 10 W m , of the maximum values of the heat flux is also reported in [69].

298 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Sintered powder structures

According to Zhao and Zhang [332], the boiling crisis consists in the total filling of the porous
coating with vapor. Their experimental investigations covered pool boiling of ethanol and R-113
at atmospheric pressure, on the external surface of a copper tube of d = 0.04 m coated with a
porous layer of bronze powder, whose parameters are given in Table 4.3. The heat fluxes qcr they
measured were higher than those for smooth surfaces qcr,sm, as shown in Fig. 4.74.

Table 4.3. Structural parameters of porous layers tested by Zhao and Zhang [332]

Number of pores
Sample no. δ, mm dgrn, µm ε
per inch (mesh)
1 60–80 1.0 370.4 0.4155
2 60–80 1.8 370.4 0.3943
3 60–80 2.56 370.4 0.3419
4 100–120 1.0 232.8 0.3748
5 140–180 1.0 161.3 0.4028

6
10
R 113
5 Atmospheric pressure

5
10

5
q, W⋅m-2

Smooth
2

10 4

5 -1
-4
-5
2

10 35·10- 5
2 5 2
1 10 0 10 1 102
DTsl, K

Fig. 4.74. Boiling curves obtained by Zhao and Zhang [332] for the porous sintered
powder coatings listed in Table 4.3

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 299


Chapter 4

According to Zhao and Zhang [332], the reason qcr increases on surfaces with porous coatings
is that the feed of liquid to the heating surface occurs much more intensively than on smooth
surfaces. This is due to the action of capillary and negative pressure forces caused by vapor
streams leaving the layer and not undergoing coalescence. Both above-mentioned factors
intensify the liquid suction into the porous coating.

The authors proposed the following causes of boiling crisis: (i) considerable superheating of the
heating surface is accompanied by vehement evaporation that leads to the porous coating filling
with vapor due to insufficient liquid feed, (ii) vapor bubble streams departing from the coating have
a considerable velocity, which tend to impede liquid inflow in the opposite direction and tends to
make the bubbles coalesce and (iii) this results in the formation of a vapor film above the upper
surface of the layer of low thermal conductivity, which leads to a significant increase in the surface
superheat at the imposed heat flux and hence initiation of the boiling crisis.

Their results for qcr for the porous coatings described in Table 4.3 are summarized in Table 4.4.
The analysis of numerical data in Table 4.3 and Table 4.4 leads to a conclusion that the size of
powder and, consequently, pore diameters, has a decisive effect on the qcr value. The larger are
particles and pores, the easier is the inflow of liquid into the porous coating interior, which results
in the increase in qcr. Under the conditions of boiling crisis, the porous layer is filled with vapor
and hence the porous layer thickness and its thermal conductivity decide the heating surface
superheat. This explains why the superheat grows with increasing the coating thickness, while the
increasing thickness does not affect the value of qcr [332].

Table 4.4. Experimental values of qcr and ΔTsl,cr from Zhao and Zhang [332]

Sample no. Ethanol R-113


Table 4.3 5 –2 5 –2
qcr, 10 W m ΔTcr, K qcr, 10 W m ΔTcr, K
1 6.336 – 7.5702 26.546 3.605 – 3.7207 24.0
2 7.137 – 7.3006 53.423 3.525 – 3.8813 35.411
3 7.373 – 7.5482 63.896 3.633 – 3.6945 38.637
4 6.260 – 6.3445 21.427 3.177 – 3.33928 20.633
5 5.493 – 6.0197 18.649 2.826 – 3.0562 21.24

A different result was obtained by Malyshenko and Andrianov [151, 263], who investigated the
boiling crisis for water at two different pressures: 0.1 and 0.2 MPa. Experiments were conducted
for porous coatings of sintered nichrome powders with particle diameters dgrn from 80 µm to
120 µm and porosities of ε = 0.4-0.5. The porous layer thickness ranged from 0.3 mm to 1.0
mm. The optimum coating thickness, for which the qcr value was the highest, was found for both
investigated pressures.

300 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Extending the scope of investigations by Zhang and Zhao [330] to include capillary-porous
structures with 40 – 300 pores per square inch (mesh), with thicknesses δ from 0.4 mm to 4 mm
and porosities ε from 0.3 to 0.6, confirmed previous conclusions [332] and resulted in a boiling
crisis model, presented later in this Chapter.

On the basis of theoretical investigations and experimental data analysis, the following correlation
was put forward by Zhang and Zhao [330]:

(4.312)

where the Weber number, defined as the ratio of surface tension to gravity forces, describes the
conditions of the liquid inflow to the porous coating

(4.313)

Experimental data scatter with respect to correlation (4.312) is ± 20% and the comparison is
shown in Fig. 4.75.

10-4
Ethanol
7 R 113

5
qcrν l (σlg)-1 (hlg) -1

+20%
3 -20%

Atmospheric
pressure

10-5 -5
10 3 5 7
10 -4
We −0.089
(ν g νl ) 1.427
(ρ g ρl ) 2.217

Fig. 4.75. Comparison of (4.312) with the experimental data by Zhang and Zhao [330]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 301


Chapter 4

Thermally sprayed structures

Plasma spray

Andrianov, Malyshenko and Styrikovich [11, 151, 152, 154, 263] differentiate between two kinds
of nucleate boiling crisis for nichrome and corundum porous coatings of low thermal conductivity,
Fig. 4.52 and Fig. 4.53. The authors assume the complete dryout of the porous coating and the
formation of a continuous vapor film above it to be the total nucleate boiling crisis phenomenon.
For a known particle size of the sprayed powder, the optimum coating thickness can be found as
depicted in Fig. 4.76, for which the ratio of the critical heat flux qcr is plotted normalized by those
for smooth surfaces qcr,sm. A similar effect of the coating thickness is also revealed for water at
pressures below and above atmospheric pressure, both for plasma-sprayed coatings and those of
sintered powders [152].

Tehver and co-researchers [148, 273, 275, 277, 279] investigated boiling heat transfer on thin
porous coatings of aluminum, bronze and copper. Most often R-113 at atmospheric pressure and
liquid nitrogen were the test fluids.

Fig. 4.76. Impact of the porous coating thickness on the critical heat flux qcr at two
pressures for helium for dgrn = 100 µm and a nichrome layer, from Andrianov,
Malyshenko and Styrikovich [151, 154]

By analogy with smooth surfaces, Tehver et al. [273, 275, 277, 279] considered the boiling crisis
to be equivalent to the separation of the heated surface from the boiling liquid by a vapor film.
For porous coatings, this means the formation of a vapor film on the external surface of the
coating. They propose two terms for that reason, the critical heat flux density qcr or the burnout
heat flux density.

302 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

On the basis of the analysis of the effect of structural parameters of the coating, the following
experimental expression was proposed to compute qcr:

qcr = A ε Dp + B (4.314)
10 –3 5 –2
A = 5.1 ∙ 10 Wm B = 2.5 ∙ 10 W m

–2
where Dp should be input in meters and qcr is given in W m . Equation (4.314) was checked
for 46 porous coatings for a range of porosities ε, the mean pore diameters Dp and the coating
thicknesses δ as shown in Fig. 4.77. The mean square deviation in Fig. 4.77 from the straight line
(4.314) is 0.05 and the correlation coefficient is 0.89. If boiling crisis hysteresis occurs for a given
porous structure, see Fig. 4.60 and Fig. 4.61, the highest value of the heat flux qcr is used in the
analysis (equation (4.314) and Fig. 4.77).

5
5⋅10

R 113

4⋅10 5
qcr , W⋅m-2

3⋅10 5

2⋅10 5 -6 -6 -6 -6
0 1⋅10 2⋅10 3⋅10 4⋅10
εDp, m

Fig. 4.77. Relationship of the critical heat flux density qcr on structural parameters of
the porous coating from Tehver et al. [279]

In their later works, Tehver et al. [273, 275, 277] presented a modification of the A and B
coefficients in equation (4.314) to fluid specific values, notably:

a. R-113 [273, 275]:

10 5
qcr = 2.66 ∙ 10 ε Dp + 3.08 ∙ 10 (4.315)

b. R-113 [277]:

10 5
qcr = 3.02 ∙ 10 ε Dp + 3.08 ∙ 10 (4.316)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 303


Chapter 4

c. Nitrogen on the aluminum surface [273, 275]:

10 5
qcr = 1.97 ∙ 10 ε Dp + 1.45 ∙ 10 (4.317)

d. Nitrogen on the aluminum surface [277]:

10 5
qcr = 1.38 ∙ 10 ε Dp + 1.63 ∙ 10 (4.318)

4 –2
The mean square deviation of the experimental points with respect to (4.315) was 2.5 ∙ 10 W m
4 –2
and with respect to (4.317) it was 1.31 ∙ 10 W m , as shown in Fig. 4.78.

The qualitative conclusions that can be drawn together for the nucleate boiling crisis on porous
surfaces are as follows. The absence of thickness δ in the correlations for qcr, equations (4.315)
to (4.318), means that only the external layers of the porous coating affect the boiling crisis.
The interpretation is based on the reasoning that, under the conditions prior to boiling crisis, the
duration of contact between the liquid and the external surface is relatively short, and therefore
the possibility of the liquid penetration into the porous coating is rather restricted. Hence, one
can surmise that only low values of the layer thickness should produce some effect on the
phenomenon. Furthermore, this means that for small coating thicknesses and also coatings with
small pore sizes, qcr can be considerably diminished due to non-uniform boiling on the external
surface [148]. The possibilities of the porous layer feed with the liquid grow as the porosity ε and
pore diameters Dp increase, hence qcr increases with increasing their product ε Dp.

304 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

5
4.8⋅10

4.4⋅105

4.0⋅105

3.6⋅105
3
R 11
5
3.2⋅10
qcr, W⋅m-2

q cr = 2.66 ⋅ 1010 ε D p + 3.08 ⋅ 105


2.8⋅105

q cr = 1.97 ⋅ 1010 ε D p + 1.45 ⋅ 105


2.4⋅105

2.0⋅105
gen
1.6⋅105 Nitro

1.2⋅105
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5
εD p ⋅106 , m

Fig. 4.78. Impact of structural parameters of the porous coating on the critical heat
flux qcr from Tehver et al. [273, 275]

Flame spray

Only a few investigations have dealt with determining qcr for flame-sprayed porous coatings [55,
58, 328]. Both Cieśliński [55, 58] and also Zhang and Zhao [328] report that the critical heat flux
was almost equal to or only slightly higher than that on smooth surfaces. In [328], an increase in
qcr was attributed to the same reasons that cause qcr to increase in capillary-porous structures of
sintered powders [332]. The smaller enhancement in qcr for flame sprayed surfaces with respect
to qcr,sm, when compared with powder structures, is thought to be caused by the lower porosity
obtainable with flame spraying.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 305


Chapter 4

Mesh structures

Tolubinskij, Antonenko and coworkers [289, 290, 292] conducted extensive investigations and
analyzed the effect of meshes overlaying the heating surface on boiling heat transfer. As it was
already stated in Chapter 3.2.4, imposing layers of meshes leads to a fall in the heating surface
mesh
superheat that initiates pool boiling, when compared with a technically smooth, Tsl ≥ ΔTsl,sm.
The difference grows with increasing pressure as was shown in Fig. 3.45. The critical heat flux
mesh
satisfies the condition qcr > qcr,sm but the difference diminishes as the pressure increases as
illustrated by their data in Fig. 4.79.

1 – water with heating wire d = 1.1 mm; 2 – water with d = 2.5 mm; 3 – water data from [96];
4 – acetone with d = 2.5 mm; 5 – ethanol with d = 1.1 mm; 6 – ethanol with d = 2.5 mm

Fig. 4.79. Impact of pressure on the critical heat flux (one mesh layer 1 mm × 1 mm)
from Tolubinskij, Antonenko and coworkers [292]

Similar to the investigations by Afanasev and Smirnov [251], the authors found that qcr diminishes
as the number of layers imposed on the heating surface is increased as shown here in Fig. 4.80a.
They also reported the optimum mesh cell dimension, for which it was possible to obtain the
highest qcr value for a chosen liquid and pressure, see Fig. 4.80b, which was also confirmed by
investigations by Sasin et al. [227] and Hasegawa et al. [96].

The relationship between the mesh cell dimension in a single mesh layer D and the diameter of
a vapor bubble departing from a smooth surface Ddep has been analyzed at constant pressure
[289], resulting in the following conclusions:

1. When D << Ddep, in the initial phase the vapor bubble grows tangentially to the
external mesh surface. The bubble growth is intensive and unbalanced, while the
capillary forces are much smaller than the inertia forces. The presence of the mesh
slightly affects the bubble growth in the final asymptote phase of growth.

2. When D ≈ Ddep, the bubble is generated inside the mesh and moves above it only
in the asymptotic phase of growth. The action of capillary forces is caused by

306 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

the formation of interfacial surfaces inside the capillary-porous structure. They


significantly affect the diameter Ddep and the bubble nucleation frequency.

3. When D > Ddep, the majority of bubbles generated on the heating surface depart
without contact with the mesh (i.e. because the spacing of the mesh is much larger
than the size of the bubble). If this is the case, boiling has an almost identical
character to that on smooth surfaces.

The classification shown above applies exclusively to single meshes as each new layer of mesh
blanks out a part of free space left by the previous meshes. Experimental results in Fig. 4.81
confirm the validity of the above conclusions. The critical heat flux qcr reaches its maximum for
mesh cell diameters D approaching the diameters of the departing vapor bubbles according
to [289].

1 – mesh 1 mm × 1 mm with heating wire of d = 1.1 mm; 2 – mesh 1 mm × 1 mm


with d = 2.5 mm; 3 – data from [96]; 4 – mesh 0.5 mm × 0.5 mm with d = 1.1 mm;
5 – one mesh layer; 6 – two layers

Fig. 4.80. The impact of mesh layer number (a) and mesh cell dimensions (b) on the
critical heat flux qcr from Tolubinskij, Antonenko and coworkers [292]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 307


Chapter 4

1, 2, 3 – water; 4 – acetone; pressures: 1 – 0.02 MPa; 2 – 0.1 MPa;


3 – 0.5 MPa; 4 – 0.1 MPa
mesh
Fig. 4.81. Impact of mesh cell diameters D on the critical heat flux qcr from
Tolubinskij, Antonenko and coworkers [289]

A single layer deposited on the heated surface causes qcr to increase because the thermal
boundary layer adjacent to the wall is less disturbed by free convection in comparison with a
smooth surface according to [289, 292]. That leads to an increase in the layer’s superheat, which
results in a higher rate of vapor bubble growth and departure frequency. Adding a mesh to the
surface also adds an additional force, where the newly added capillary pressure inside the mesh
cell tends to quicken bubble departure. The effect produced is to reduce the departure diameter
Ddep and increment the departure frequency. Furthermore, the presence of the mesh impedes
condensation of vapor bubbles, which also increases the value of qcr.

One of the major difficulties in applying meshes to make porous boiling structures is their weak
adherence to the heating surface, as proven experimentally in [289]. An increase in the gap
mesh
between the mesh and the base surface causes qcr to drop as shown in Fig. 4.82. Above a
certain size of gap, vapor bubbles are generated mainly inside the gap, hindering liquid inflow to
the heating surface [290]. This results in coating the base heating surface with a vapor film and
mesh
thus lowering qcr values below those when the mesh adheres well. If the gap between the mesh
and the base is smaller than a certain threshold value, additional capillary forces acting inside
mesh
the gap push bubbles outside the mesh. This effect additionally increases qcr when the heating
surface is covered by a single mesh layer [289].

308 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

For mesh cell dimensions 1 – 1.05 mm × 1.05 mm; 2 – 0.4 mm × 0.4 mm;
3 – 0.2 mm × 0.2 mm

Fig. 4.82. Impact of the thickness of the gap between a single mesh layer and the
mesh
heating surface on the critical heat flux qcr from Tolubinskij, Antonenko
and coworkers [289]

Metal fibrous structures

Capillary-porous structures of this kind demonstrate a dependence of the critical heat flux on
both porosity ε and the coating thickness δ, as shown in Fig. 3.51 [141]. The highest qcr values
were obtained for the porosity ε = 0.8 and the lowest for ε = 0.4. This proves that the resistances
of vapor outflow from and liquid inflow into the coating interior, both of which decrease with
increasing porosity, controls the magnitude of qcr. The effect produced by porosity is however
rather ambiguous. At a selected porosity ε, an increase in δ causes first an increase and then a
fibrous
decrease in qcr. In general, qcr > qcr,sm from 1.4 to 2.2 times according to [141].

The analysis of the heat flux corresponding to the maximum heat transfer coefficient, as
influenced by the coating thickness δ and skeleton conductivity λskl (3.60), led Shapoval et al.
[269] to put forward the following empirical expression to compute the critical heat flux:

(4.319)

where the maximum porosity εlmt values are determined using (3.61), whereas the critical heat
flux for smooth surfaces qcr,sm is obtained with the Kutateladze-Zuber hydrodynamic model [137,
–1 –1
333, 334]. In eq. (4.319) δ should be input in mm, λskl in W m K and the mean pore diameter
D in µm.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 309


Chapter 4

The correlation (4.319) was developed on the basis of experimental investigations with water and
acetone at atmospheric pressure. The porous coatings were made of copper, nickel and stainless
steel, for layer thicknesses of δ = 0.2 – 1.0 mm and porosities of ε = 0.4 – 0.6. The statistical
scatter of the experimental data with respect to equation (4.319) was not given. Curiously, taking
the empirical expression (4.319) to its limit, qcr → ∞ when δ → 0, which is not physically possible.

Electrolytically deposited capillary-porous structures

This kind of porous coatings is not applied in practice due to poor mechanical properties [51, 58].
The characteristic feature of this technique of depositing a porous coating is the possibility of
obtaining very thin layers.

On the basis of the shapes of boiling curves for liquid helium shown in Fig. 4.83, Malyshenko and
Styrikovich [154] make a hypothesis that the thermal resistance in a capillary-porous structure
totally filled with vapor results from the resistance of the skeleton thermal conductivity and natural
convection in vapor. Natural convection is controlled by the thermal conductivity, especially for
liquids of low boiling point and small viscosity, such as helium. In the case under consideration
in Fig. 4.83, although the coating thickness δ grew, the critical heat flux qcr also increased, which
indicates that the thermal resistance due to convection in the vapor has decreased [154].

4
6·10
4 Helium

10 4
-2

8
q, W⋅m

6
4
-1
2 -2
-3
103 -4

4
0.1 0.4 0.6 1.0 4.0 6.0 10.0 40.0
DTsl , K

1 – smooth surface; 2 – δ = 10 µm; 3 – δ = 25 µm; 4 – δ = 50 µm (porosities not given)

Fig. 4.83. Boiling curves for electrolytically deposited porous coatings of various
thickness of Malyshenko and Styrikovich [154]

310 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

In porous coatings in which the coating is deposited by means of electrolysis, similar to porous
structures made with other techniques, the critical heat flux qcr increases up to 80% relative to
qcr,sm [172]. The correlation proposed by Nikolaev and Tokalov [172] to compute qcr as the function
of pressure is based on a small number of experimental data, for fluids that are rarely used for
heating, i.e. sulfur hexafluoride (SF6) and carbon dioxide (CO2). Investigations were conducted
exclusively at elevated pressures, close to the critical pressure for both fluids. The correlation
uses the critical heat flux at the pressure p = 0.9 pcr as its reference value, which means that the
applicability range of the correlation is very limited and hence of limited application in practice.

Summary

Summing up all the experimental results, it should be noted that coating a heating surface with a
porous layer manufactured with all the techniques discussed above results in an increase in the
critical heat flux qcr. The increment of the ratio of qcr to that of smooth surfaces qcr,sm depends
on the structural parameters of the coating, namely porosity ε, thickness δ and the mean pore
dimension D. For meshes, sintered powders, plasma-sprayed and metal fibrous coatings, it is
possible to select structural parameters of the porous layer in such a way so as to maximize the
critical heat flux for a particular fluid at a particular saturation condition.

4.4.2. Attempts at phenomenological modeling

The diversity in the nucleate boiling crisis for industrial enhanced boiling surfaces and those made
by depositing capillary-porous structures together with the multiplicity of factors affecting the
phenomenon account for difficulties one faces in trying to propose a model for its prediction. The
majority of boiling heat transfer models, where it is possible to determine the curves q = q (ΔTsl),
discussed in Chapter 4.2, do not apply to heat fluxes satisfying the condition q ≈ qcr.

Experimental investigations aimed at explaining the evaporation mechanism inside the porous
structure, especially those conducted with fast cameras, point to a hydrodynamic character of
boiling crisis [3, 251, 273, 275, 283]. The hydrodynamic hypothesis, put forward by Kutateladze [137]
and Zuber [333, 334], is referred to in theoretical considerations in the works by Smirnov, Afanasev
et al. [3, 4, 6, 202, 251, 253], Kovalev, Solovev et al. [115, 122, 124, 129, 201] and also Tehver et al.
[273, 275, 277, 286]. In their models of nucleate boiling crisis, the critical heat flux for technically
smooth surfaces provides the reference quantity, which is modified with respect to the porous
coating properties.

Cornwell et al. [65] are convinced that the boiling crisis occurs when the ratio of the heating
surface coated with vapor to the total surface approaches the value of one, i.e. Fg /F ≈ 1. This
leads to the conclusion that the critical heat flux can be obtained by extrapolating the curve
q = q (Fg /F) in Fig. 4.2. This concept was not verified experimentally, however [65].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 311


Chapter 4

Poniewski et al. [204, 209], on the other hand, proposed that the boiling crisis should be treated as
a statistical maximum point process, which develops inside the porous layer, a proposition that is
described later in this section.

Hydrodynamic models

Smirnov, Afanasev et al. model [3, 4, 6, 201, 251, 253]

Mesh structures, being the major wicks of heat pipes manufactured in former Soviet Union
countries, count among those most thoroughly investigated and theoretically analyzed, and in
particular the nucleate boiling crisis model built by Smirnov, Afanasev and coworkers [3, 4, 6, 202,
251, 253].

In accordance with hydrodynamic hypothesis, the crisis occurs when the stability of the two-phase
layer adjacent to the wall is disturbed. This happens when the kinetic energy of vapor generated
inside the capillary-porous coating is greater than or equal to that of the two-phase layer at the
heating surface [137]. The energies involved can be written as follows:

(4.320)

(4.321)

where wg is the evaporation rate.

In order to emphasize the difference in the character of boiling crisis on heating surfaces coated
with capillary-porous structures, the characteristic heat flux is denoted as qmax.

Note to the reader: Afanasev and Smirnov say that the boiling crisis inside a porous layer has a
different character than that on a smooth surface and it gives a different boiling curve in this zone.
In fact, the curve is rather flat, slowly growing with temperature. Because there is no temperature
jump like that which occurs for a smooth surface, they therefore strongly recommend the notion
“maximum heat flux” be used for porous coatings. Adhering to that recommendation, qmax will be
used for the rest of this section for porous coatings.

After accounting for the friction force for a vapor flow through the porous structure, equation
(4.320) can be written as:

2
E ≥ k1 Π2f – k 2 Πt (4.322)

312 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

where E is the kinetic energy of vapor stream, ∏2f is the potential energy of the stable two-phase
layer and ∏f is the work of friction forces of the vapor flow through the porous layer. Continuing
with this line of reasoning adopted in writing (4.320) and (4.322), the maximum heat flux qmax,
characteristic of the boiling crisis in the porous layer, can be determined using the equation

(4.323)

The potential energy of the stable two-phase layer at the heating surface can be derived from a
similar expression:

(4.324)

2
Expression (4.320) suggests the occurrence of only one “source” of potential energy – k1 ∏2f.
Therefore, in accordance with (4.322), the presence of capillary-porous structures should always
lead to a lowering of qmax, which is inconsistent with experimental data [121, 154, 281, 289, 290,
292]. Boiling in cells inside capillary-porous structures is accompanied by the formation of a
curvilinear phase boundary, which is due to the action of capillary forces. The boundary can be
treated as a “source” of potential energy and estimated in accordance with Laplace’s equation
to be:

(4.325)

where D is the diameter of a cell in the zone adjacent to the wall of the capillary-porous structure
(cell diameter in mesh adherent to the heating).

Boiling inside porous layers, especially for capillary feed, is accompanied by intensive droplet
ejection outside the porous layer [3, 293]. This causes an additional energy loss ΔEcon. The
modified stability condition for the two-phase layer at the heating surface can be presented in the
following way:

2
E ≥ k1 Π2f – k 2 Πf + k3 Πσ – ∆Econ (4.326)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 313


Chapter 4

Another extreme case of boiling heat transfer in porous layers, which occurs mainly in heat
pipes, is termed “hydrodynamic boundary”. A certain maximum heat flux qmax corresponds to the
boundary. The hydrodynamic balance for qmax is determined on the basis of the condition of a
difference between the capillary pressure and liquid flow resistance in the porous layer, increased
or decreased by the value of the hydrostatic pressure:

(4.327)

where L is the path of the heat-carrying liquid and h is the difference in the liquid levels. The liquid
mass flux is equal to the vapor mass flux generated inside the layer, hence the liquid velocity wl is
computed from an energy balance as:

(4.328)

where F and O are the heating surface area and its circumference, respectively, δ is a single mesh
thickness and n is the number of meshes in the layer.

All the factors occurring in inequalities (4.326) and (4.327) jointly affect the boiling crisis in the
capillary-porous structure. Satisfying both conditions may lead to the occurrence of the nucleate
boiling crisis, the complex character of which is illustrated in Fig. 4.84 from Afanasev et al. [4 – 6].

Fig. 4.84. Illustration of nucleate boiling crisis model in mesh structures in pool
boiling and with the liquid capillary feed, from Afanasev et al. [4, 6]

314 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

In the case under consideration, energy losses from friction depend on the structure of the two-
phase flow in the porous layer, i.e. the dryness level φ (volumetric content of the vapor phase). As
a first approximation it can be assumed that when q → qmax, also φ → φmax, so the term k 2 ∏f can
be estimated as follows:

k 2 ∏f ≈ C2 Δp1f (4.329)

where Δp1f is the hydraulic resistance of one-phase vapor flow.

A porous coating built of a few mesh layers functions as a filtering layer for the vapor flowing
through it. Thus, Δp1f can be computed as a result of the local filtration resistance to a vapor
stream [100]:

(4.330)

where

(4.331)

ξ is a local friction coefficient and ψ(ε) is a porosity function.

Regarding the capillary feed of the porous coating, liquid droplets are carried away outside the
layer when vapor bubbles burst. The stream of droplets entrained outside the porous layer zone
is estimated using the equilibrium condition for momentums at vapor release from the bursting
bubble:

(4.332)

where Gcon is mass flow rate of droplets carried away outside the porous coating. Hence:

(4.333)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 315


Chapter 4

It should also be noted that the smaller the stream of entrained droplets, the greater the thickness
of the porous layer and the smaller the permeability K. If those remarks are taken into account, the
kinetic energy of droplets can be estimated as follows:

(4.334)

where ff is the function of flow resistances.

When equations (4.326) and (4.327) are considered jointly and (4.323), (4.324), (4.325), (4.328)
and (4.334) are accounted for, we obtain the following:

(4.335)

where nδ is the porosity layer thickness (mesh set).

For the sake of simplification and on the basis of the authors’ experimental data [3, 201], it is
assumed that:

ff (n, K) K n δ ≈ const1 F ∙ L ≈ const 2 (4.336)

Owing to the simplifications above, it is possible to describe the share of droplet entrainment in
equation (4.335) as additional heat transfer that compensates droplet loss outside the porous
layer. This additional heat transfer causes an increase in the hydraulic resistance, which making
use of (4.336), can be written in the form of a correction coefficient:

(4.337)

8
where the value const3 is estimated to be 4 × 10 m. After accounting for (4.337), the second term
of the left side of equation (4.335) has the form:

(4.338)

316 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

In experimental investigations intended to verify the model under consideration, the condition
ρl g h ≥ 0 is satisfied for the majority of cases. Then, for the first term of the right side of (4.335),
it can be written:

(4.339)

2
Further analysis leads to the conclusion that the sum of the terms of (4.335) containing (qmax) is
close to zero. Having accounted for the above mentioned simplifications, the solution to equation
(4.335), in reference to qmax, is presented in the following form:

(4.340)

where

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 317


Chapter 4

10-1
Water
-2 -1
10 -50%
-2
-3
p = 0.01; 0.03; 0.1 MPa -50%
10-3 -4
-5
-6
10-4 Ethanol
Z

-7
10-5 -8
-9
p = 0.005; 0.01; 0.03; 0.1 MPa
- 10
10-6 - 11
- 12
-7
10
10 -8 10-6 10-4 10 -2 10 0 10 1
Y

1 – stainless steel, D = 0.04 mm, n = 2, 3; 2 – stainless steel, D = 0.06 mm,


n = 2, 3, 4, 5, 10, 24; 3 – stainless steel, D = 0.125 mm, n = 2, 3, 4, 5, 9;
4 – stainless steel, D = 0.2 mm, n = 2, 3, 4, 5, 9; 5 – brass, D = 0.08 mm, n = 2,
3, 5; 6 – copper, D = 0.045 mm, n = 2, 3, 5; 7 – stainless steel, D = 0.06 mm,
n = 2, 9; 8 – stainless steel, D = 0.125 mm, n = 2, 3, 8; 9 – stainless steel,
D = 0.45 mm, n = 2, 3; 10 – brass, D = 0.08 mm, n = 3, 9; 11 – brass,
D = 0.16 mm, n = 2, 3; 12 – copper, D = 0.045 mm, n = 2

Fig. 4.85. Generalization of experimental data for the maximum heat flux qmax and
mesh capillary-porous structures from Afanasev et al. [3, 4, 6, 251]

–5 5
The following values of constants B1 = 1.9 ∙ 10 and B2 = 2.21 ∙ 10 were obtained for water and
ethanol for boiling in mesh capillary-porous structures [3, 250]. The most difficult problem was
to determine the permeability K. In equation (4.337), it is computed using the formula for mesh
capillary-porous structures in heat pipes [3, 251]:

2.09
K = 0.035 D (4.341)

where D should be expressed in meters. Using this expression to obtain K, the experimental
results are compared with (4.340) in Figure 4.85, where the maximum data scatter is ± 50%.

It can be seen that the semi-empirical correlation (4.339) agrees fairly well with the experimental
data in Fig. 4.85, with an error acceptable for measurements of boiling heat flux. This indicates
that the hydrodynamic hypothesis about nucleate boiling crisis is able to predict the correct trend
and magnitude of qmax for mesh capillary-porous structures.

318 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The expression (4.339) can be applied to estimate the maximum operating heat flux in heat
exchangers with mesh coatings and in heat pipes. It can also be used to optimally select structural
parameters of the coating so that the highest qmax value is obtained for a heat exchanger of a
given design.

1 – vacuum pump; 2 – clamp device; 3 – liquid container; 4 – windows;


5 – thermocouples; 6 – main heater; 7 – condenser; 8 – capillary-porous
structure; 9 – copper cylinder; 10 – liquid; 11 – compensation heater

Fig. 4.86. Diagram of experimental stand to investigate boiling heat transfer in mesh
coatings from Afanasev et al. [3, 4, 6]

Experimental investigations were conducted using the measurement stand depicted in Fig. 4.86
of Afanasev et al. [3, 4, 6]. The copper heating surface, 30 mm in diameter, was coated with
capillary-porous structures made of meshes, the number of which ranged from n = 1 to n = 24.
The mean mesh cell dimension D changed from 0.04 mm to 0.5 mm and pressure from 0.01
MPa to 0.1 MPa. Experiments were made for water and ethanol for pool boiling (+ρ l g h) and for
capillary feed of the porous layer (–ρl g h). The experimental stand was used to analyze the impact
of selected parameters on the maximum value of the heat flux qmax.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 319


Chapter 4

a) b)
6 6
4 -1 4
-2

qmax, W⋅m -2
2 -3 2
-4
-5
10 5 105 -1 -4
8 8
6 6 -2 -5
-3
4 Water 4 Water
p = 0.1 MPa p = 0.1 MPa
Ethanol
2 2
1 2 4 6 8 10 2 4 4 6 8 10 2 2 4 6
c) n D, mm
4⋅10 5
Ethanol

3⋅105
qmax, W⋅m-2

2⋅105

-1 -3 -5
105 -2 -4 -6
0 0.04 0.08 1.00 1.02
p, MPa

a) effect of mesh number (coating thickness); capillary feed; stainless steel:


1 – D = 0.06 mm; 2 – copper, D = 0.045 mm; 3 – D = 0.06 mm (pool boiling);
4 – D = 0.2 mm; 5 – D = 0.125 mm;

b) effect of mesh cell dimensions D: water, 1 – n = 2; 2 – n = 3; 3 – n = 5; ethanol,


4 – n = 2; 5 – n = 5;

c) effect of pressure, stainless steel: 1 – n = 2, D = 0.45 mm; 2 – n = 2, D = 0.125 mm;


3 – n = 1, D = 0.06 mm; 4 – n = 2, D = 0.06 mm; 5 – brass, n = 2, D = 0.16 mm;
6 – n1 = 1, D = 0.06 mm + n2 = 2, D = 0.2 mm

Fig. 4.87. Impact of selected parameters on the maximum heat flux qmax for mesh
coatings from Afanasev et al. [4, 6]

Like in the investigations conducted by Tolubinski, Antonenko and coworkers [289, 290, 292], qmax
was found to decrease with a growing number of mesh layers n, see Fig. 4.87a and for decreasing
mesh cell dimensions D, see Fig. 4.87b. The effect of thickness was attributed to a change in
hydraulic resistance to vapor flow, term k 2 ∏t in (4.326). The increase in mesh cell dimensions D
has a dual effect on the qmax value. Firstly, it causes a drop in the hydraulic resistance through
term k 2 ∏f, which results in an increase in qmax. Secondly, the increase in D leads to a reduction in
capillary pressure increment in term k3 ∏σ, which tends to decrease the value of qmax.

320 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

A change in saturation pressure alters the boiling liquid’s thermal properties, creating a complex
influence on qmax as shown in Fig. 4.87c for ethanol. Nevertheless, an optimum pressure can
be found experimentally, at which it is possible to get the highest qmax values for pre-selected
structural parameters of the porous coating.

Kovalev, Solovev and coworkers model


[115, 117, 118, 122, 124, 129, 201]

The hydrodynamic model of the nucleate boiling crisis built by Kovalev, Solovev and coworkers
[115, 117, 118, 122, 124, 129, 201] is an extension of the boiling heat transfer model inside
capillary-porous structures presented in Chapter 4.2.2.13. It is assumed, as before, that
evaporation proceeds inside the structure, which makes a system of vertical and horizontal
capillaries, as shown in Fig. 4.22a and Fig. 4.88.

Theoretical considerations are referred to experimental results for sintered powder layers. For
heat fluxes q ≤ qmax in Fig. 4.88, condition (4.55) must be satisfied, which ensures stable vapor
and liquid circulation inside the coating.

Tehver and Tunik [286] make an assumption that in order to maintain stable nucleate boiling it
is sufficient to feed the porous coating with sufficient liquid, which forms a stream of individual
droplets. The estimation presented in [286] indicates that the critical heat flux, obtained on the
basis of the stability condition of opposite directed streams of droplets and vapor, is approximately
ten times greater than the critical heat flux for a smooth surface qcr,sm.

~ 2 to 4 times q
The increment of qmax for surfaces with a porous coating is much lower (qmax = cr,sm)
[120, 130, 260] than ten, so therefore Kovalev [115] makes an assumption that boiling crisis for
such surfaces is also of a hydrodynamic character.

For heat fluxes close to the critical value j, q ≤ qmax, a vapor layer, which adheres to the heating
surface, is generated inside the porous coating as in Fig. 4.88. The heat flux is transferred to the
boiling front inside the coating through the vapor layer. Hence the heating surface superheat is
high, much greater than for a smooth surface. That refers, in particular, to materials of low thermal
conductivity, though the coating skeleton temperature in the evaporation zone is almost equal to
the boiling liquid saturation temperature. Vapor streams penetrate the coating through large pores,
see Fig. 4.22 and Fig. 4.88, while the liquid wets the coating through the small pores. The position
of vapor streams is fixed and the Taylor instability does not control the formation of the vertical
vapor streams. This fact constitutes a basic difference in comparison with boiling on the smooth
surface where the theory of Zuber applies the Taylor theory to the find the wavelength between
neighboring vapor streams leaving a plain surface.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 321


Chapter 4

On a smooth surface, vapor streams of the radius λT/4 are located in the corners of the square
having sides of length λT. The radius of vapor streams in boiling on surfaces with porous coatings
is considerable less than λT/4. Therefore, in accordance with Helmholtz’s hypothesis [13, 333],
vapor streams are stable for much higher vapor velocities than in the Taylor stability model, which
leads to an increase in qmax with respect to qcr,sm.

a) “evaporation front” location inside the capillary-porous coating; b) temperature distribution

Fig. 4.88. Heat transfer model for q ≤ qmax from Kovalev, Solovev and coworkers [122,
124, 129]

In order to compute qmax, it is necessary to determine the threshold vapor velocity and the content
of pores acting as vapor channels. A real capillary-porous structure is substituted with a system of
vertical cylindrical capillaries, ideally connected by horizontal channels [45]. The maximum vapor
velocity is computed using the onset of instability in accordance with Helmholtz’s model [333]:

(4.342)

where the stream radius is assumed to equal the pore radius, R = Rp.

322 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Prior to determining the content of pores acting as vapor channels, it is necessary to compute
the liquid meniscus on the external surface of the porous coating. Equation (4.156), written in the
following form:

(4.343)

describes the equilibrium of the liquid and vapor phase at departure from the porous coating.
Pores with radii R > Rlmt (δ) are filled with vapor and those with R < Rlmt (δ) are filled with the
liquid.

In each porous layer section parallel to the heating surface, the pressure is constant, which results
from the assumed model of cylindrical, vertical ideally connected capillaries. Consequently, the
vapor velocity in each pore when leaving the porous coating is constant as well (4.148), so that:

= const (4.344)

The maximum vapor velocity at the departure from the porous coating is related to the radius of
the liquid volume being “perforated” by the departing vapor stream, R = Rbd, which is determined
with the use of equations (4.342) and (4.343):

(4.345)

~ 0,76 R .
where for ρg << ρl, Rlmt (δ) = bd

The value of the maximum heat flux is obtained by summing the vapor streams with respect to the
unit area of the porous coating:

(4.346)

~ Π(R ) is determined owing to percolation theory and on


The approximate value of Π(Rlmt (δ)) = bd
the basis of experimental data originating from numerical investigations into sintered structures of
spherical particles [45]:

(4.347)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 323


Chapter 4

According to Kovalev and Solovev [122, 124], the most frequently used structures of sintered
powders satisfy condition (4.347b), so that

(4.348)

Table 4.5. Comparison of experimental and computed critical heat fluxes, qmax, for
sintered powder capillary-porous structures from [120, 122]

Rp δ Dgrn qmax,exp qmax Boiling Structure


No ε –2 Source
µm mm µm MW m qmax,exp liquid material

1 0.15 5.7 1.0 – 2.90 0.90


2 0.35 40 1.0 – 4.40 0.73

3 0.35 40 2.0 – 4.30 0.74 Water [130]

4 0.40 65 2.0 – 4.10 0.83


5 0.47 100 2.0 – 3.70 1.05
Tin bronze
6 0.15 5.7 1.0 – 1.45 0.69
7 0.35 40 1.0 – 2.50 0.52

8 0.35 40 2.0 – 2.40 0.54 Ethanol [260]

9 0.40 65 2.0 – 2.30 0.57


10 0.47 100 2.0 – 2.30 0.54

11 0.70 – 1.34 200-250 0.45 1.16 R-113


[223]
12 0.70 6.2 1.34 200-250 0.28 1.44 Nitrogen

13 0.64 6.2 1.0 – 0.55 0.65


Copper
14 0.64 6.2 2.2 – 0.38 0.95
R-113 [299]
15 0.64 6.2 4.1 – 0.30 1.20

16 0.64 6.2 6.2 – 0.38 0.95

Values of qmax, computed on the basis of (4.346) and (4.348) are compared with experimental data
for water, ethanol, R-113 and nitrogen for boiling on coatings made of sintered powders of copper
and tin bronze listed in Table 4.5 [130, 223, 260, 299]. The porous coatings compared in Table 4.5
–0.5 –0.5
satisfy the condition (1 – Π (Rbd)) ε Rbd < 15 m with the mean square error equal to 0.31. If the
above-given condition is not satisfied, the values of the expression 1 – ∏ (Rbd) computed on the
basis of (4.347) considerably exceed those resulting from real pore dimension distributions.

324 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

In order to determine the impact of pressure on the maximum heat flux, the values for technically
smooth surfaces and those with a porous coating were compared [120, 122]. For water and
0.3
reduced pressures p/pcr < 0.07, both values were practically the same and qmax ~ p . As the
pressure grows, differences between qmax and qcr,sm grow and qcr,sm > qmax. This indicates a
weaker pressure effect on the maximum heat flux for surfaces with a porous coating than for
smooth surfaces.

Kovalev and Solovev [120, 122] discussed the range of applicability of the proposed model to
compute the maximum heat flux qmax. Firstly, it is necessary to presume that the assumptions
concerning the physics of the boiling process, put forward in Chapter 4.2.2.13, are satisfied. The
basic limitation is the applicability range of Helmholtz’s theory, which was formulated for a single
gas stream moving vertically adjacent to an infinite liquid volume. Therefore, the assumption that
there is no interaction of neighboring vapor streams is not fulfilled for structures of high porosity ε
and small pores Dp.

Furthermore, the model assumes the liquid permeates the porous coating interior up until
q ≤ qmax. The assumption cannot always be satisfied. It is known that the nucleate boiling crisis on
the smooth surface occurs when it reaches the required superheat, Ts – Tsat ≥ ΔTmax. With respect
to boiling inside the porous coating, if the condition Tskl – Tsat ≥ ΔTmax is satisfied, the boiling liquid
and the skeleton material do not come in direct contact with each other due to the formation of
a vapor layer that separates them. Then the skeleton area, inside which the liquid evaporation
occurs, is restricted to the part that fulfills the condition Tskl – Tsat ≥ ΔTmax. The depth of liquid
penetration inside the coating depends on the skeleton thermal conductivity λskl. The higher is λskl,
the larger is the evaporation zone (i.e. that of liquid penetration) and the more vapor is generated
in it. Regarding coatings of low thermal conductivity, the area of the evaporation zone is small and
too little vapor is generated in it for condition (4.342) to be satisfied, on the basis of which equation
(4.346) was derived. This means that in this case (4.346) gives the upper value of the critical heat
flux qmax.

According to Kovalev and Solovev [120, 122], the above considerations explain why the
maximum values of critical heat fluxes for boiling on surfaces with porous coatings of both low
and high skeleton conductivity do not differ much from values specific to smooth surfaces [125,
130, 260, 299].

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 325


Chapter 4

Model of intralayer boiling crisis based on


the concept of the maximum point process

Poniewski, Wiśniewski and Wojcik [204, 209] proposed a probabilistic model of the intralayer
crisis. The latter is a process of vapor layer formation inside the coating, which is characteristic of
some capillary-porous structures. The layer isolates the heating surface from the remaining part of
the coating flooded with the liquid. It manifests itself as considerably smaller slopes of inclination
of the boiling curve or the occurrence of horizontal segments of the boiling curve, shown earlier in
Fig. 4.52 [11, 154] and Fig. 4.57 [115, 122, 127, 271].

When the heat flux, denoted earlier as qintcs or q*, initiating the intralayer boiling crisis is exceeded,
it does not cause as dramatic surface temperature rise as when exceeding qcr, yet attaining a
steady-state is impossible and a given surface is excluded from further operation.

The distribution of the pores of the capillary-porous structure at the heating surface can be always
described with a two-dimensional point process. On the basis of analogy with a smooth surface
[204, 209], it is assumed to be a Poisson point process, denoted further as N. A Poisson point
process is fully probabilistically characterized by its intensity. For the sake of simplification, it
is possible to assume that the intensity λ of process N is the same on the whole of the heating
surface and then λ is proportional (with a certain proportionality constant dependent on unit area)
to the number of pores per unit area.

The process of filling the capillary-porous structure cells with vapor depends on the thermal
properties of the liquid and the skeleton material and also on the value of the superheat and the
pore dimensions. For a given structure and liquid, at a set superheat, it is assumed that the rule
that determines if a cell of a certain size is filled with vapor can be expressed stochastically. The
rule has the form of the so- called thinning function p = p (x), which ascribes the probability p(x)
of being filled with vapor under given conditions to each cell x. In other words, the distribution
of the dry sites of the structure under given conditions can be modeled mathematically with the
use of the so-called thinned point process Np, originating in process N, in which its localization is
randomly “cut out”, in accordance with probability p. It should be noted that the process formed in
this way is also a Poisson process of intensity λ, which can be easily computed if λ and p
are known.

Let us assume that the process of drying the capillary-porous structure is observed for the values
of the heating surface superheats ΔT1,..., ΔTn. Point processes of the distribution of dried cells,
which correspond to those values, are denoted as N1,..., Nn. As already stated, they are “thinnings”
of process N with the thinning probability equal to p1,..., pn, respectively. On the basis of a series
of point processes: N1, N2,..., Nn, the process M, termed the maximum point process, is defined.
Process M has localizations (points) distributed everywhere, where at least one of processes
N1, N2,..., Nn has a localization. For such a defined process, the probability that the number of its

326 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

points in a given zone will exceed a certain value is equal to the probability that the number of the
points of any of processes N1, N2,..., Nn in this zone will exceed the value under consideration.

A certain critical level of drying intensity ν exists, the exceeding of which in an arbitrary zone
A of surface F causes the drying of the capillary-porous structure in zone A. The level ν (A) is
a measure proportional to the area of zone A, moreover, the proportionality constant b gives
information as to how many cells per unit area, at the least, must be dried so that a vapor film
would be formed inside the coating.

Furthermore, it is assumed that a certain critical value of area c exists, such that the occurrence,
wherever inside the capillary-porous structure, of the zone covered with a vapor film of the area
equal to c, inevitably initiates the extension of the area, which leads to the occurrence of the
boiling crisis.

That means the occurrence of the intralayer heat transfer crisis is equivalent to the maximum point
process exceeding the critical intensity level in an arbitrary structure zone of an area larger than
the critical one.

In accordance with the assumptions made above, when a, b, c, p1,..., pn, dependent on the
structure’s geometrical properties and thermal properties of both the structure and the liquid, are
presumed to be known, the probability p* of intralayer boiling crisis non-occurrence for the series
of superheats ΔT1,..., ΔTn is estimated with the following formula [70]:

(4.349)

where

p* = 1 – p

(4.350)

whereas the symbol [x] denotes the integral part of number x. The present model indicates that
the probability of a crisis occurrence is, in particular, a function of the range of superheat of the
heating surface ΔT, owing to which, for a given structure and liquid, it is possible to:

1. find the value of permissible superheating, for which the probability of boiling crisis
occurrence will be lower than the pre-set security level;

2. compute the mean superheating value, for which intralayer boiling crisis occurs.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 327


Chapter 4

Using this methodology it is possible to simulate numerically the process of formation of nucleate
boiling crisis in porous structures.

Initial verification of the model was provided by a sample of porosity ε = 0.40, thickness δ = 2 mm
(metal fibrous structure) and water as the test fluid. The following critical heat fluxes qmax were
–2 5 5 5 5
obtained experimentally for the sample (in W m ): 1.7 · 10 ; 1.6 ∙ 10 ; 2.0 ∙ 10 ; 1.5 ∙ 10 ;
5 5
1.7 ∙ 10 and 1.8 ∙ 10 . In the experiment, the heat flux was gradually increased until the boiling
crisis occurred, while other process parameters remained unchanged. Assuming that probabilities
–1
of thinnings p = p(q,...) are generated by quantiles q = Φ (p) of the normal distribution Φ,
estimated on the basis of empirical qcr values and assuming, arbitrarily, what could be typical
values of other model parameters on the basis of (4.349):

(4.351)

The probabilities of intralayer boiling crisis occurrence at the particular heat flux values were
estimated and are plotted in Fig. 4.89a. A monotonic increasing trend was obtained. As the mean
5 –2
value of the critical heat flux for the sample under consideration was 1.7 ∙ 10 Wm , to which
close to a probability of 1.0 corresponded, the results obtained for the model should be regarded
as describing the intralayer boiling crisis phenomenon with a good approximation.

328 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

a)
1.0

P robability p
0.8

0.6

0.4

0
1.4⋅105 1.5⋅105 1.6⋅105 1.7⋅105
-2
b) q, W⋅m
1.0

0.8
Probability p

0.6

0.4

0
3.0⋅10-6 8.0⋅10-6 1.3⋅10-5 1.8⋅10-5 2.3⋅10 -5
Critical area c, m 2

Fig. 4.89. Probability of a) intralayer boiling crisis occurrence versus heat flux,
obtained on the basis of (4.349) and b) intralayer boiling crisis occurrence
p = 1 – p* plotted versus the critical area value c, obtained on the basis of
(4.349), according to Poniewski, Wiśniewski and Wojcik [204, 209]

The effect of the magnitude of the critical area c on the probability of intralayer crisis occurrence
was examined for the above data. A monotonically decreasing trend was obtained as shown in
Fig. 4.89b, in accordance with the intuitional behavior of the examined phenomenon.

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 329


Chapter 4

a)
1.0

0.8

Probability p
0.6

0.4

0.2

0
0.10 0.15 0.20 0.25
Ratio b/a
b)
1.0
0.8
Probability p

0.6

0.4
0.2

0
-3
0 5.0⋅10 1.0⋅10-2 1.5⋅10 -2 2.0⋅10-2
Total area F, m2

Fig. 4.90. Probability of intralayer boiling crisis occurrence p = 1 – p* versus a)


the ratio b/a, obtained on the basis of (4.349) and b) the total area of the
heating surface F, obtained on the basis of (4.349), according to Poniewski,
Wiśniewski and Wojcik [204, 209]

The effect of the ratio b/a, i.e. the fraction of the dry cells in the structure, necessary for the
formation of a vapor layer insulating the heating surface from the boiling liquid in the zone under
consideration, on the probability of intralayer crisis occurrence was shown. While the remaining
parameters were set and not determined, the monotonic decreasing trend obtained in Fig. 4.90a is
consistent with the intuitional behavior of the examined phenomenon. For the qualitative change in
the probability of crisis occurrence, the interval (15% to 20%) turned out to be the key range of the
investigated fraction.

Finally, the impact of the total area of the heating surface on the probability of crisis occurrence
was examined while other parameters were fixed in such a way so that the probability of drying an
arbitrary single cell at an arbitrary instant would be 0.01. For this, a monotonic increasing trend was
obtained as shown in Fig. 4.90b, which suggests functionally an increasing relationship between
the total area of the heating surface and the critical area value c.

330 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

The concept of a series of thinned Poisson point processes was used to describe the dynamics
of the dryout of a capillary-porous structure, whereas with the notion of the maximum point
process of the above-mentioned series, it was possible to formulate the condition sufficient for
the occurrence of the intralayer boiling crisis. Thus, the presented model can be used to estimate
the probability of boiling crisis occurrence, depending on the geometrical and thermal properties
of the porous layer, the liquid thermal properties and the heat flux. In particular, it is possible
to compute the levels of the heat flux which result in boiling crisis occurrence with a pre-set
probability. Such an approach offers the possibility of investigating the effect individual structural
parameters and liquid properties have on the phenomenon. The model is only introductory in
character, so it is necessary further investigate its appropriateness and verify it empirically.

The maximum (critical) heat flux qmax (that is qcr for a porous coated surface, see comment below
eq. 4.321) for porous coatings depends, in a complex manner, on the porous layer structural
parameters, thermal properties of the boiling liquid and their interaction. Experimentally obtained
qcr (or qmax) values and correlations and theoretical models designed to compute them have a
restricted applicability range and cannot readily be compared with one another. Both experiments
and theoretical models predict an increase qmax when facilitating vapor escape, which can be
achieved by increasing porosity and/or reducing the coating thickness.

An interesting and sophisticated approach, both theoretical and experimental, to the maximum
(critical) heat flux enhancement was proposed by Liter and Kaviany [146]. They applied modulated
porous coatings with periodically non-uniform, designed variations in the layer’s thickness,
Fig.4.91. The coatings were made from sintered spherical copper particles by dry-phase diffusion
in a reducing atmosphere. The imposed modulation was assumed to create alternating regions of
low resistance to vapor escape and highly capillary-assisted liquid suction, which should facilitate
liquid-vapor counter flow within the capillary-assisted porous layer and enhance boiling heat
transfer from the substrate to the liquid pool in a manner similar to heat pipes.

For the investigated porous coating design, various possible physical limitations causing nucleate
boiling crisis were considered. The limitations ranged from a counter flow limit for a deep
porous layer (the lowest qmax) to a kinetic evaporation limit (the highest qmax). Some simplifying
assumptions were made regarding the flow paths, without rigorous experimental validation in
order to develop models that would allow one to calculate the boiling curve q = q(ΔTsl) and qmax.
The most important assumptions were as follows:

a. separated liquid and vapor mass fluxes within the layer’s space, Fig. 4.91;

b. local thermal equilibrium in all phases;

c. Darcy-Ergun momentum equation with Leveret J-function relating the liquid saturation,
porosity, permeability and wettability to capillary pressure;

d. energy equation for volume-averaged values of the liquid local temperature and the
stagnant thermal conductivity of the porous coating;

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 331


Chapter 4

e. viscous-drag numerical model, which defines the evaporation rate at the liquid-vapor
interface and the heat flux at the bottom of the porous coating;

f. hydrodynamic liquid-choking limit, qmax = qmax, h (subscript h refers to hydrodynamic),


which determines the ability of the liquid to flow towards the heating surface through
the escaping vapor; the limit determines the highest heat flux qmax at which the liquid
is able to reach the heating surface.

Numerical solution of the momentum and energy equations for the viscous drag numerical model
and the wetted surface regime, Fig. 4.91, allowed for the prediction of the slope of qvs (subscript
vs refers to viscous drag model) versus ΔTsl curve. The model underestimates the value of qvs as
a function of ΔTsl, but it can be considered to be in fair agreement with the experimental dashed
line of Fig. 4.92, despite many assumptions that were made. Liter and Kaviany [146] concluded
that the assumptions regarding liquid flow paths and volume averaging of the porous media should
be reconsidered to make the result of numerical calculations show better agreement with the
experiment. It was also assumed that modulation of the porous coating imposes a geometrically
determined, i.e. modulated, critical wave length λR-T, Fig. 4.91, which supersedes the dependence
on the Rayleigh-Taylor wavelength and extends the hydrodynamic choking limit. This way of
thinking led to the following very simple final expression for qmax:

1/2
qmax, h = (π/8) hlg (σlg ρg /λR – T) (4.352)

For the porous coatings tested by Liter and Kaviany [146], qmax was shown to be hydrodynamically
limited, qmax = qmax, h as shown in Fig.4.92. The agreement between the theory and the
qmax, h experimental data is presented in a slightly obscure manner. For the dual-height modulation
coating, the experimental qmax corresponds to qmax, h for λR-T = 1.46 mm predicted from eq. (4.352)
and the real λm = 1.6 mm. For single-height modulation, the appropriate values of λm are 1.68 mm
and 1 mm, respectively, where the difference is quite substantial.

One of the obvious advantages of this hydrodynamic model is good agreement between the
measured and calculated values of qc,pl (which refers to the critical heat flux for plane surfaces)
for a flat porous coating, Fig. 4.92. The investigated modulated capillary-porous coatings provide
a noticeable increase in qmax and a decrease in ΔTsl when compared with a flat coatings tested
in this study and reported in the literature [4, 121, 154, 204]. Liter and Kaviany [146] predict
that further enhancement of qmax in the modulated capillary-porous coating can be done by
reduction of the modulated wavelength for λR-T and particle size D and by removing the limiting
hydrodynamic mechanism.

332 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Fig. 4.91. Scheme of the physical model of liquid-vapor interface hydrodynamic


instability limit for liquid reaching the surface at boiling on modulated
porous coatings from Liter and Kaviany [146]

1 – single modulated, δ = 6 D, λm = 5 D; 2 – double modulated, δ = 5 D and 9 D,


λm = 8 D; 3 – uniform layer, δ = 3 D; 4 – plain surface

Fig. 4.92. Experimental data for various porous coatings from Liter and Kaviany [146]

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 333


Chapter 4

4.5. Summary

The complex mechanism of boiling heat transfer on industrial enhanced boiling surfaces
and those produced by depositing capillary-porous structures is responsible for a number of
peculiarities that are not observable on technically smooth surfaces. The intralayer boiling crisis and
a variety of hysteresis phenomena should be mentioned here first of all.

In spite of numerous theoretical and experimental studies available on the subject, no unified
description of the boiling heat transfer mechanism controlling heat transfer from microporous
surfaces is yet possible. None of the models discussed can explain the inconsistency in
experimental results from different research centers or arising from different geometrical and
technological capillary-porous structures. Considerable simplifications in models, frequently
disregarded effects produced by some properties of the boiling liquid or the structural parameters
of the microporous surface, neglect of the heat transfer hysteresis phenomena, etc. result in
limited applicability of models to selected heat transfer conditions or provide a description of only a
segment of the boiling curve.

None of the discussed models was supported by microscale experimental investigations into heat
transfer in a single cell or a set of neighboring cells of a capillary-porous structure.

The conclusion drawn from the present theoretical and experimental findings is that it is not
possible to formulate a general theory of boiling in capillary-porous structures because:

• no agreement can be reached regarding the basic mechanism of heat transport (liquid
layer evaporation or evaporation from liquid-vapor menisci);

• it is not possible to account for complex and greatly diversified geometrical structures
within a single universal model.

Only thorough microscale experimental investigations, e.g. visualization of heat and mass
transport in a single pore or a small set of neighboring pores, can contribute to the advancement
of understanding the mechanisms of heat and mass transfer inside capillary-porous structures.
Without this, the wide diversity of models proposed so far will continue to grow without attaining a
unified, general model based on the actual micro-mechanics of the process.

Indisputable benefits resulting from considerable boiling heat transfer enhancement due to the
application of industrial enhanced boiling surfaces or those produced by depositing capillary-
porous structures provide a strong stimulus to continue research efforts in this field.

334 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Bibliography

1. Abhat A., Seban R.A., Boiling and Evaporation from Heat Pipe Wicks with Water and
Acetone, Trans. ASME, ser. C., J. Heat Transfer, vol.96 (1974), p.74, (Russian translation,
Teploperedacha (Mir),1974, No.3, pp.74-82)

2. Abramenko A.N., Kanonchik L.E., Liquid Boiling Heat Transfer in Porous Materials, in
collective work: “Heat and mass transfer in systems with porous elements”, Minsk,1981,
pp.13-19 (in Russian)

3. Afanasev B.A., Investigation of Vapor Generation Process in Capillary – Porous Mesh


Structures (heat transfer, limiting heat fluxes, vapor generation mechanism), Ph.D. thesis,
Odessa Technological Institute of Refrigeration Industry, Odessa, 1981 (in Russian)

4. Afanasev B.A., Poniewski M.E., Smirnov G.F., Boiling Crisis Heat Flux in Capillary-Porous
Layer, Archives of Thermodynamics, vol.16 (1995), No.3-4, pp.165-175

5. Afanasev B.A., Smirnov G.F., Investigation of Heat Transfer and Limiting Heat Fluxes at
Boiling in Capillary – Porous Structures, Teploenergetika, 1979, No.5, pp.65-7 (in Russian)

6. Afanasev B.A., Smirnov G.F., Poniewski M.E., Wojcik T.M., The Boiling Crisis Phenomenon
on Capillary-Porous Coverings, ASME, Proc. XXX Nat. Heat Transfer Conf., Portland, USA,
1995, ASME-HTD, vol.309, pp.17-21

7. Afgan N.W, Jovich L.A., Kovalev S.A., Lenkov V.A., Boiling Heat Transfer from Surfaces with
Porous Layers, Int. J. Heat Mass Transfer, vol.28(1985), pp.415-21

8. Ali S.M., Thome J.R., Boiling of Ethanol – Water and Ethanol – Benzene Mixture on
Enhanced Boiling Surface, Heat Transfer Engineering, vol.5 (1984), pp.70-81

9. Allingham W., McEntire J., Determination of Boiling Film Coefficient for Heated Horizontal
Tube in Water-Saturated Wick Material, J. Heat Transfer, vol.83 (1969), pp.71-76

10. Anderson T.M., Mudawar I., Microelectronic Cooling by Enhanced Pool Boiling of a
Dielectric Fluorocarbon Liquid, J. Heat Transfer, vol.111 (1989), pp.752-759

11. Andrianov A.B., Malyshenko S.P., Effects of Porous Coating Characteristics on Boiling Heat
Transfer, Izvestia AN SSSR, ser. Power Engineering and Transport, 1989, No.1, pp.139-149
(in Russian)

12. Andrianov A.B., Malyshenko S.P., Sirenko E.I., Styrikovich M.A., Hysteresis and Transient
Phenomena at Boiling on Surfaces with Porous Coatings, Doklady AN SSSR, vol.256
(1981), No.3, pp.591-595 (in Russian)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 335


Bibliography

13. Andrianov A.B., Malyshenko S.P., Styrikovich M.A., Talaev I.V., Transient Processes
Peculiarities and Boiling Curve Shape on Surfaces with Porous Coatings, Doklady AN
SSSR, vol.273 (1983), No.4, pp.867-870 (in Russian)

14. Arai N., Fukushima T., Arai A., Nakajima T., Fujie, Nakayama Y., Heat Transfer Tubes
Enhancing Boiling and Condensation in Heat Exchangers of a Refrigerating Machine,
ASHRAE Trans., vol.83(1977), p.58

15. Arshad J., Thome J.R., Enhanced Boiling Surfaces: Heat Transfer Mechanism and Mixture
Boiling, Proc. ASME-JSME Thermal Eng. Joint Conf., 1983, vol.1, pp.191-197

16. Asakavichjus I.P., Zhukauskas A.A., Gajgalis V.A., Eva V.K., Heat transfer of Freon-113,
Ethanol and Water in Mesh Wicks, Trudy AN Litovskoj SSR, ser. B, 1978, vol.1 (104), pp.87-
93 (in Russian)

17. Ayub Z.H., Bergles A.E., Nucleate Boiling Curve Hysteresis for Gewa-T Surfaces in
Saturated R-113, ASME-HTD, vol.26 (1988), pp.515-21

18. Ayub Z.H., Bergles A.E., Pool Boiling Enhancement of Modified GEWA-T Surface in Water,
J. Heat Transfer, vol.110 (1998), pp.266-267

19. Ayub Z.H., Bergles A.E., Pool Boiling from GEWA Surfaces in Water and R-113, Bishop
P.J.(ed.), Augmentation of Heat Transfer in Energy Systems, ASME, HTD -vol.52 (1985),
pp.57-66

20. Ayub Z.H., Bergles A.E., Pool Boiling from GEWA Surfaces in Water and R-113, Wärme-
und Stoffübertragung, vol.21 (1987), pp. 209-219

21. Bankoff S.B., Entrapment of Gas in the Spreading a Liquid Over a Rough Surface, AIChE J.,
vol.4 (1958), pp.24-28

22. Bar-Cohen A., Hysteresis Phenomena at the Onset of Nucleate Boiling, eds. Dhir V.J.,
Bergles A.E., ASME-HTD, Pool and External Flow Boiling, 1992, pp.1-14

23. Bar-Cohen A., Simon T.W., Wall Superheat Excursions in the Boiling Incipience of Dielectric
Fluids, Bar-Cohen A. (ed.), Heat Transfer in Electronic Equipment, ASME, HTD-vol.57
(1986), pp.83-94

24. Bayazitoglu Y., Özisik M.N., Elements of Heat Transfer, McGraw-Hill, 1988

25. Bergles A.E., Some Perspectives on Enhanced Heat Transfer – Second Generation of Heat
Transfer Technology, J. Heat Transfer, vol.110 (1998), pp.1082-1096

26. Bergles A.E., Chyu M.C., Characteristics of Nucleate Pool Boiling from Porous Metallic
Coatings, Advances in Enhanced Heat Transfer, ASME, HTD-vol.18 (1981), pp.61-72

27. Bergles A.E., Chyu M.C., Characteristics of Nucleate Pool Boiling from Porous Metallic
Coatings, J. Heat Transfer, vol.104 (1982), pp.279-285

28. Bergles A.E., Sommerscales E.F.C., The Effect of Fouling on Enhanced Heat Transfer
Equipment, J. Enhanced Heat Transfer, vol.2 (1995), pp.157-166

336 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

29. Biceroglu O., Mujumdar A.S., van Heinigen A.R.P., Douglas W.J.M., Thermal Conductivity
of Sintered Metal Powders at Room Temperature, Letters in Heat and Mass Transfer, vol.3
(1976), pp.183-192

30. Bier K., Gorenflo D., Salem M., Tanes Y., Pool Boiling Heat Transfer and Size of Active
Nucleation Centers for Horizontal Plates with Different Surface Roughness, Proc. VI Int.
Heat Transfer Conf., Toronto, 1978, vol.1, pp.151-156

31. Bilicki Z., Latent Heat Transfer in Forced Boiling Flow, Int. J. Heat Mass Transfer, vol.26
(1983), pp.559-565

32. Bird B.R., Steward W.E., Lightfoot E.N., Transport Phenomena, John Wiley & Sons, New
York, 1960

33. Bobrowski D., Probabilistcs in Technological Applications, WNT Publishers, Warsaw, 1986
(in Polish)

34. Bohdal T., Czapp M., Investigation of Zero Flow Boiling Crisis, Proc. Nat. Symp. on Heat and
Mass Transfer, Warsaw-Jablonna, 1986, pp.34-40 (in Polish)

35. Borishanskaja A.V., Freons Boiling Heat Transfer on Surfaces with Metallic Porous Coatings,
Holodil’naja Tehnika, 1979, No.12, pp.17-20 (in Russian)

36. Bräuer H., Mayinger F., Onset of Nucleate Boiling and Hysteresis Effects Under Forced
Convection and Pool Boiling, eds. Dhir V.J., Bergles A.E., ASME-HTD, Pool and External
Flow Boiling, 1992, pp.15-36

37. Bukin V.G., Danilova G.N., Djundin V.A., Freon-12 Boiling Heat Transfer on Banks of
Sprinkled Horizontal Tubes with Heat Transferring Surface Porous Coating, in collective
work: “Boiling and Condensation”, Riga, 1981, pp.79-82 (in Russian)

38. Burmeister L. C., Convective Heat Transfer, John Wiley & Sons, Inc., 1983

39. Camp W.M., van, Ph.D. Thesis, Prude Univ., La Fayette, (Indiana), 1952

40. Chang J.Y., You S.M., Boiling Heat Transfer Phenomena from Micro-Porous and Porous
Surfaces in Saturated FC-72, Int. J. Heat Mass Transfer, vol.40 (1997), pp.4437-4447

41. Chang J.Y.,You S.M., Enhanced Boiling Heat Transfer from Micro-Porous Cylindrical
Surfaces in Saturated FC-87 and R-123, ASME-HTD, vol.330 (1996), pp. 19-27

42. Chang J.Y., You S.M., Enhanced Boiling Heat Transfer from Micro-Porous Cylindrical
Surfaces in Saturated FC-87 and R–123, J. Heat Transfer, vol.119 (1997), pp.319-325

43. Chang J.Y., You S.M., Enhanced Boiling Heat Transfer from Micro-Porous Surfaces: Effects
of a Coating Composition and Method, Int. J. Heat Mass Transfer vol.40 (1997), pp.4449-60

44. Chang J.Y., You S.M., Heater Orientation Effects on Pool Boiling of Micro-Porous
–Enhanced Surfaces in Saturated FC-72, J. Heat Transfer, vol.118 (1996), pp.937-43

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 337


Bibliography

45. Chejfec L.I., Nejmark A.V., Multiphase Processes in Porous Media, Himija, Moscow, 1982 (in
Russian)

46. Chien L.-H., Webb R.L., A Nucleate Boiling Model for Structured Enhanced Surfaces, Int. J.
Heat Mass Transfer, vol.41 (1998), pp.2183-95

47. Chien L.-H., Webb R.L., A Parametric Study of Nucleate Boiling on Structured Surfaces.
Part I: Effect of Tunnel Dimensions, ASME-HTD, vol.326 (1996), Nat. Heat Transfer Conf.,
vol.4, pp.129-36

48. Chien L.-H., Webb R.L., A Parametric Study of Nucleate Boiling on Structured Surfaces.
Part II: Effect of Pore Diameter and Pore Pitch, ASME-HTD, vol.326 (1996), Nat. Heat
Transfer Conf., vol.4, pp.137-43

49. Chien L.-H., Webb R.L., Measurement of Bubble Dynamics on Enhanced Boiling Surface,
Experimental Thermal and Fluid Science, vol.16 (1998), pp.177-86

50. Chien L.-H., Webb R.L., Visualization of Pool Boiling on Enhanced Surfaces, Experimental
Thermal and Fluid Science, vol.16 (1998), pp.332-41

51. Cieslinski J.T., An Experimental Study of Nucleate Pool Boiling Heat Transfer from a Flat
Horizontal Plate Covered with Porous Coatings, Archives of Thermodynamics, vol.12 (1991),
pp.69-76

52. Cieslinski J.T., Augmentation of Heat Transfer in Boiling of Water on Surfaces with Capillary
Porous Layers, eds. Sunden B. et al, Proc. II Baltic Heat Transfer Conf., 1995, Jurmala
(Latvia), pp.395-402

53. Cieslinski J.T., Experimental Validation of Predictive Models for Boiling on Porous Coatings,
III Int. Conf. on Multiphase Flow – ICMF’98, Lyon, 1998, pp.1-6

54. Cieslinski J.T., Improved Heat Transfer to Water Boiling Tubes Covered with Porous Metallic
Coatings, Archives of Thermodynamics, vol.16 (1995), pp.97-105

55. Cieslinski J.T., Pool Boiling Crisis on Porous Surfaces, Proc. 9th Symp. on Heat and Mass
Transfer, KTS – PAN, Augustow, 1995, vol.1, pp.223-32 (in Polish)

56. Cieslinski J.T., Nucleate Pool Boiling Heat Transfer of Water from Gas – Thermally Coated
Surfaces, Archives of Thermodynamics, vol.13 (1992), pp.49-57

57. Cieslinski J.T., Pool Boiling of Water on Rough and Porous Coated Surfaces, Proc. Int.
Conf. Heat Transfer with Change of Phase, Kielce, 1996, Sci. Papers Kielce Univ. Techn., s.
Mechanics, vol.61 (1996), pp.153-160

58. Cieslinski J.T., Study of Nucleate Boiling on Metallic Porous Surfaces, Gdansk Technical
Univ. Sci. Papers, no. LXXVI (547), 1996 (in Polish)

59. Cieslinski J.T., Mikos M., A New Method of Manufacturing Boiling Heat Transfer Enhancing
Surfaces, Proc. of the 5th Polish - German Symp. “Science for practice”, Gdansk, 1994,
pp.39-44 (in Polish)

338 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

60. Cieslinski J.T., Mikos M., Surfaces with Gas – Thermal Spray Coatings, Proc. of Sci. and
Techn. Conf. – “Mechanics’95 – Science and practice”, Gdansk, 1995, vol.1, Science A-6,
pp.27-28 (in Polish)

61. Cohen P., Heat and Mass Transfer for Boiling in Porous Deposits with Chimneys, AIChE
Symp. Series., vol.70 (1974), No.138, pp.71-80

62. Corman L.C., Walmet G.E., Vaporization from Capillary Wick Structures, ASME Paper,
1971, 71-HT-35, pp.1-8

63. Cornwell K., On Boiling Incipience Due to Contact Angle Hysteresis, Int. J. Heat Mass
Transfer, vol.25 (1982), pp.205-211

64. Cornwell K., Brown R.D., Boiling Surface Topography, Proc. VI Int. Heat Transfer Conf.,
Toronto, 1978, vol.1, pp.157-161

65. Cornwell K., Nair B.G., Patten T.D., Observation of Boiling in Porous Media, Int. J. Heat
Mass Transfer, vol.19 (1976), pp.236-238

66. Corty C., Ph. D. Thesis, Univ. Michigan, Ann Arbor, 1951

67. Czikk A.M., Gottzmann C.F., Ragi E.G., Withers J.G., Hadbbas E.P., Performance of
Advanced Heat Transfer Tubes in Refrigerant – Flooded Liquid Coolers, ASHRAE Trans.,
vol.76 (1970), part I, pp.96-109

68. Czikk A.M., O’Neill P.S., Correlation of Nucleate Boiling from Porous Metal Films,
Chenoweth J.M., Kaelis J., Michel J., Shenkman S. (eds.), Advances in Enhanced Heat
Transfer, ASME, 1979, pp.53-60

69. Czikk A.M., O’Neill P.S., Gottzmann C.F., Nucleate Boiling from Porous Metal Films: Effect
of Primary Variables, Advances in Enhanced Heat Transfer, ASME, HTD-vol.18 (1981),
pp.109-122

70. Daley D.J., Vere-Jones D., An Introduction to the Theory of Point Processes, Springer-
Verlag, New York, 1988

71. Danilova G.N., Djundin V.A., Bogdanov S.N., Kuprijanova A.V., Borishanskaja A.V., Kozyrev
A.A., Heat Transfer Enhancement in Shell-and-Tube Evaporators, Holodil’naja Tehnika,
1981, No.5, pp.36-40 (in Russian)

72. Danilova G.N., Djundin V.A., Borishanskaja A.V., Solov’ev A.G., Vol’nyh Ju.A., Kozyrev A.A.,
Effect of Surface Conditions on Boiling Heat Transfer of Refrigerants in Shell – and – Tube
Evaporators, Heat Transfer – Soviet Research, vol.22 (1990), No.1, pp.56-65

73. Danilova G.N., Ioffe O.B., Djundin V.A., Borishanskaja A.V., Kozyrev A.A., Nikiforova V.V.,
Povolockij V.M., Fridgant L.G., Liquid Boiling Heat Transfer on Heating Surfaces with
Capillary – Porous Coatings, in Martynova O.I., Borishanskij V.M. (eds), “Vapor generators
thermal regime and hydraulics”, Leningrad, Nauka, 1978, pp.73-78 (in Russian)

74. Danilova G.N., Tihonov A.V., R-113 Boiling Heat Transfer Enhancement on Various
Surfaces, Holodil’naja Tehnika, 1984, No.5, pp.33-37 (in Russian)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 339


Bibliography

75. Dettre R.M., Johnson, Jr. R.E., Contact Angle Hysteresis. II. Contact Angle Measurements
on Rough Surfaces, Adv. Chem. Series, vol.43 (1964), pp.136-144

76. Djundin V.A., Danilova G.N., Borishanskaja A.V., Heat Transfer to Refrigerants Boiling on
Tubes Covered with Porous Coatings, Heat Transfer – Soviet Research, vol. 13 (1981), No.1,
pp.61-74

77. Djundin V.A., Danilova G.N., Borishanskaja A.V., Cooling Agents Boiling Heat Transfer on
Surfaces Covered with Porous Coatings, in „Heat Transfer and Hydrodynamics” – Proc.
5th Federal Conf. on Two-Phase Flow Heat Transfer and Hydraulic Resistance in Power
Machinery and Apparatuses, pp.15-30, Leningrad, 1977 (in Russian)

78. Djundin V.A., Danilova G.N., Borishanskaja A.V., Krotkov V.N., Gogolin V.A., Bahalin V.A.,
Protasov G.A., Cooling Agents Boiling Heat Transfer Enhancement on Surfaces with Gas –
Thermal Coatings, Himicheskoe i Neftjanoe Mashinostroenie, 1975, No.9, pp.22-23 (in Russian)

79. Djundin V.A., Solovev A.G., Protasov G.A., Lopuhin V.I., Vol’nyh Ju.A., Kljueva K.D., Gas
– Thermal Aluminum Coatings Application to Refrigerators Evaporators, Himicheskoe i
Neftjanoe Mashinostroenie, 1986, No.12, pp.13-15 (in Russian)

80. Djundin V.A., Solov’ev A.G., Vol’nyh Ju.A., Surface Type Impact on Cooling Agents
Efficiency, Holodil’naja Tehnika, 1984, No.5, pp.37-40 (in Russian)

81. Dorohov A.R., Bochagov V.N., Effects of Coatings on Boiling Heat Transfer of Mixtures and
Solutions, Izvestia Sibirskogo Otdelenija AN SSSR, ser. Engineering Sciences, 1982, No.13,
vol.3, pp.13-17 (in Russian)

82. Dulnev G.N., Zarichnjak Ju.P., Thermal Conductivity of Mixtures and Composed Materials,
Energija, Leningrad, 1974 (in Russian)

83. Dusan V., On Spreading Liquids on Solid Surfaces: Static and Dynamic Contact Lines, Ann.
Rev. Fluid Mech., vol.11 (1979), pp.371-399

84. Eva V., Asakavichjus I., Gajgalis V., Low Temperature Heat Pipes, Mosklas, Vilnius, 1982 (in
Russian)

85. Fedorov I.I., Investigation of Boiling Mechanism in Heat Pipe Wicks, in collective work: “Heat
and Mass Transfer in Flying Apparatus Engines”, Kazan’, 1978, No.2, pp.69-75 (in Russian)

86. Ferrell J.K., Alleavitch J., Vaporization Heat Transfer in Capillary Wick Structures, Chem.
Eng. Progress Symp. Series, vol.66 (1970), pp.82-91, (Russian translation: Shpilrajn
E.E.(ed.), Heat pipes, Mir, Moskva, 1972, pp.118-141)

87. Ferrell J.K., Johnson H.R., The Mechanism of Heat Transfer in the Evaporator Zone of
the Heat Pipe, Chem. Eng. Progress, Symp. Series, vol.66 (1970), (Russian translation:
Shpilrajn E.E.(ed.), Heat pipes, Mir, Moscow, 1972, pp.9-32)

88. Gaertner R.F., Westwater J.W., Population of Active Sites in Nucleate Boiling Heat Transfer,
Chem. Eng. Prog. Symp. Series, vol.56 (1960), pp.39-48

89. Gogolin V.A., Krotkov V.N., Pechaj V.A., Tovaras N.V., Danilova G.N., Borishanskaja A.V.,
Djundin V.A., Kozyrev A.A., Vahalin V.A., Protasov G.A., Heat Transfer Enhancement in

340 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Freon Shell and Tube Evaporators with Metal Coated Tubes, Holodil’naja Tehnika, 1979,
No.1, pp.26-31 (in Russian)

90. Gottzmann C.F., O’Neill P.S., Minton P.F., High Efficiency Heat Exchangers, Chem. Eng.
Prog., vol.69 (1973), No.7, pp.69-75

91. Gottzmann C.F., Wulf J.B., O’Neill P.S., Theory and Application of High Performance Boiling
Surfaces to Components of Absorption Cycle Air Conditioners, Proc. Conf. Natural Gas
Research and Technology, 1971, Chicago, Session V, Paper 3, pp.1-35

92. Griffith P., Wallis J.D., The Role of Surface Conditions in Nucleate Boiling, Chem. Eng. Prog.
Symp. Series, vol.56 (1960), pp.49-63

93. Grigorev V.A., Krohin Ju.I., Kulikov A.S., On the Issue of Determination of Under-Bubble
Liquid Film at Boiling in Capillary Channels, Trudy MEI, vol.200, 1974, pp.8-16 (in Russian)

94. Hahne E., Spindler K., Shen N., Incipience of Flow Boiling in Subcooled Well Wetting Fluids,
Proc. IX Int. Heat Transfer Conf., Jerusalem, 1990, paper 1-BO-12

95. Haider S.I., Webb R.L., A Transient Micro-Convection Model of Nucleate Pool Boiling, Int. J.
Heat Mass Transfer, vol.40 (1997), pp.3675-3688

96. Hasegawa S., Echigo R., Irie S., Boiling Characteristics and Burnout Phenomena on
Heating Surface Covered with Woven Screens, J. Nuclear Science and Technology, vol.12
(1975), pp.722-724

97. Holman J.P., Heat Transfer, McGraw-Hill, New York, 1990

98. Hsieh S.S., Weng C.J., Nucleate Pool Boiling from Coated Surfaces in Saturated R-134a
and R-407c, Int. J. Heat Mass Transfer, vol.40 (1997), pp.519-532

99. Hsu Y.Y., Graham R.W., Transport Processes in Boiling and Two-Phase Systems,
Hemisphere, 1976

100. Idel’chik I.E., Hydraulic resistance handbook, Energija, Moscow, 1975 (in Russian)

101. Imadojemu H.E., Hong K.T., Webb R.L., Pool Boiling of R-11 Refrigerant and Water on
Oxidized Enhanced Tubes, J. Enhanced Heat Transfer, vol.2 (1995), pp. 189-198

102. International Encyclopedia of Heat and Mass Transfer, eds. Hewitt G.F., Shires
G.L.,Polezhaev Y.V., CRC Press, 1996

103. Ito T., Takata Y., Kubota H., Uehara T., Makino A., Pool Boiling Heat Transfer from Porous
Surfaces to Liquid Nitrogen, Cryogenics, vol. 30 (1990), September Supplement, pp. 292-6

104. Ito T., Tanaka K., Tamari T., A Prediction of Boiling Heat Transfer Performance on Porous
Surfaces, Proc. I KSME-JSME Thermal and Fluids Eng. Conf., 1988, vol.1, pp.79-84

105. Ito T., Tanaka K., Tamari T., Prediction of Boiling Heat Transfer from Porous Surfaces, Trans.
JSME, vol.55 (1989), pp.1403-1409

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 341


Bibliography

106. Ivanovskij M.N., Privezencev V.V., Il`in Yu.A., Sidorenko E.M., Experimental Investigation of
Heat Transfer Intensity at Liquid Evaporation on Embossed Capillary Structures, Inzhynerno
- Fizicheskij Zhurnal, vol. XLVI (1984), No.4, pp.533-538 (in Russian)

107. Kelleher M.D., Egger R., Joshi Y., Modification of the Nucleate Boiling Hysteresis in the Pool
Boiling of Fluorocarbons, Proc. X Int. Heat Transfer Conf., Brighton, 1994, vol. 5, pp. 87-92,
paper 10-PB-14

108. Kim C.-J., Bergles A.E., Incipient Boiling Behavior of Porous Boiling Surfaces Used for
Microelectronic Chips, ed. Veziroglu T.N., Particulate Phenomena and Multiphase Transport,
Hemisphere, 1988, vol. 2, pp. 3-18

109. Ko S.-Y., Liu B., Yao Y.-Q., Boiling Hysteresis on Porous Metallic Coatings, Proc. II Int.
Symp. Multiphase Flow and Heat Transfer, Beijing, 1989, vol. 1, pp. 258-68

110. Konev S.V., Mitrovich J., An Explanation for the Augmentation of Heat Transfer During
Boiling in Capillary Structures, Int. J. Heat Mass Transfer, vol. 29 (1986), pp. 91-4

111. Korneev A.D., Investigation of Boiling Hydraulics and Heat Transfer in Slotted Channels
under the Conditions of Modeling of Lowered Gravitational Force, Ph. D. Thesis summary,
SMVTU, Moscow, 1977 (in Russian)

112. Korneev A.D., Investigation of Boiling Hydraulics and Heat Transfer in Vertical Slotted
Channels, Ph. D. Thesis summary, MVTU, Moscow ,1974 (in Russian)

113. Kostornov A.G., Permeable Metal Fibrous Materials, Technika, Kiev, 1983 (in Russian)

114. Kostrzewa S., Nowak B., Fundamentals of Vehicle Parts Reclaiming, WKL, Warsaw, 1986
(in Polish)

115. Kovalev S.A., Nucleate Boiling Wave Instability on Porous Surface, in “Convective Heat
Transfer. Methods and Results of Investigations”, IVTAN, pp.68-78, Moscow, 1982 (in
Russian)

116. Kovalev S.A., Derevjanko D.Ja., Mahalova M.V., Dolgincev I.I., Sinelnikov V.S., Investigation
of Forced Convection Boiling Heat Transfer Crisis for Subcooled Water on a Finned Surface,
Proc. Federal Conf. Heat and Mass Transfer – V, Minsk 1976, vol. III., part I, pp. 162-9 (in
Russian)

117. Kovalev S.A., Lenkov V.A., Mechanism of Burnout with Boiling on a Porous Surface,
Thermal Engineering, vol. 28 (1981), No. 4, pp. 201-3

118. Kovalev S.A., Lenkov V.A., On Mechanism of Boiling Crisis on a Porous Surface,
Teploenergetika, 1981, No. 4, pp. 8-11 (in Russian)

119. Kovalev S.A., Ovodkov O.A., A Study of Gas-Liquid Counterflow in Porous Media,
Experimental Thermal and Fluid Science, vol. 5 (1992), pp. 457-64

120. Kovalev S.A., Ovodkov O.A., Solovev S.L., Lenkov V.A., Liquid Boiling Heat Transfer on
Surfaces with Porous Coatings, Proc. Federal Conf. Heat Transfer – VII, Minsk, 1984, vol. 6,
part. I, pp. 104-9 (in Russian)

342 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

121. Kovalev S.A., Solovev S.L., Heat Transfer and Critical Heat Fluxes in Boiling on a Porous
Surface, Heat Transfer-Soviet Research, vol. 22 (1990), No. 3, pp. 364-75

122. Kovalev S.A., Solovev S.L., Evaporation and Condensation in Heat Pipes, Nauka, Moscow,
1989 (in Russian)

123. Kovalev S.A., Solovev S.L., Heat Transfer Model of Liquid Boiling on a Porous Surface,
Teplofizika Vysokih Temperatur, vol.22 (1984), No.6, pp.1166-71 (in Russian)

124. Kovalev S.A., Solovev S.L., Heat Transfer Model of Boiling of Liquids on a Porous Surface,
Proc. Federal Conf. Heat and Mass Transfer – VIII, pp. 28-50, Minsk, 1988 (in Russian)

125. Kovalev S.A., Solovev S.L., Heat Transfer at Boiling and Evaporation of Liquids on Porous
Surface. Proc. Federal Conf. Heat and Mass Transfer – VII, vol.2, pp.3-12, Minsk, 1984 (in
Russian)

126. Kovalev S.A., Solovev S.L., Heat Transfer and Critical Heat Fluxes at Boiling on a Porous
Surface, in collective work: “Two–Phase Flows: Heat Transfer and Hydrodynamics”, pp. 97-
108, Nauka, Leningrad, 1987 (in Russian)

127. Kovalev S.A., Solovev S.L., Ovodkov O.A., Liquid Boiling on Porous Surfaces, Proc. of
School – Seminar: Heat and Mass Transfer Processes at Two–Phase Changes and at Two-
Phase Flows,1985, pp.26-38 (in Russian)

128. Kovalev S.A., Solovev S.L., Ovodkov O.A., Liquid Boiling on Porous Surfaces, Heat
Transfer-Soviet Research, vol.19 (1987), No.1, pp.109-120

129. Kovalev S.A., Solovev S.L., Ovodkov O.A., Theory of Boiling Heat Transfer on a Capillary
– Porous Surface, Proc. 10th Int. Heat Transfer Conf., Jerusalem, 1990, vol.2, pp. 105-110,
paper 1-BO-18

130. Kovalev S.A., Shklover E.G., Heat Transfer at Boiling of Water on a Porous Surface in
a Square Channel, Teplofizika Vysokih Temperatur, vol.26 (1988), No.5, pp.918-922 (in
Russian)

131. Kraus A.D., Sixty-five Years of Extended Surface Technology (1922-1987), Appl. Mech. Rev.
vol.41 (9), 1988, pp.321-364

132. Kravchenko V.A., Ostrovskij N.Ju., Hysteresis and Transient Phenomena at Boiling of Pure
and Mixed Liquids on Sprayed Surfaces, Teploenergetika, 1988, No.6, pp.61-63 (in Russian)

133. Kravchenko V.A., Ostrovskij N.Ju., Liquid Boiling on a Heating Surface Having Cylindrical
Capillaries, Heat Transfer – Soviet Research, vol.22 (1990), No.1, pp.84-90

134. Kravchenko V.A., Ostrovskij N.Ju., Spivakov Ju.A., Heat Transfer Investigation of Boiling of
Water and Ethanol and Their Mixtures on Heating Surfaces with Porous Coating, Inzhynerno
- Fizicheskij Zhurnal., vol.47 (1984), No.5, pp.753-759 (in Russian)

135. Krohin Ju.I., Kulikov A.S., Heat Transfer Investigation of Boiling of Cryogenic Liquids in
Capillary – Porous Bodies, Trudy MEI, 1976, vol.310, pp.21-29 (in Russian)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 343


Bibliography

136. Krohin Ju.I., Kulikov A.S., Approximate Hydrodynamic Theory of Vapor Formation in
Capillary – Porous Structures, Teplofizika Vysokih Temperatur, vol.21 (1983), No.5, pp.952-
958 (in Russian)

137. Kutateladze S.S., Fundamentals of Heat Transfer Theory, Atomizdad, Moscow, 1979 (in
Russian)

138. Kuzma-Kihta Ju.A., Moskvin V.N., Sorokin D.N., Water Boiling Heat Transfer Investigation
on Surfaces with Porous Coatings for a Wide Range of pressures, Teploenergetika, 1982,
No.3, pp.53-54 (in Russian)

139. Labuncov D.A., An Approximate Theory of Heat Transfer at Developed Nucleate Boiling,
Izvestia AN SSSR, OTK, ser. Power Engineering and Transport, 1963, No.1, pp.58-71 (in
Russian)

140. Labuncov D.A., The Issues of Heat Transfer at Nucleate Boiling of Liquids, Teploenergetika,
1972, No.9, pp.14-19 (in Russian)

141. Levterov A.I., Semena M.G., Zaripov V.K., Investigation of Heat Transfer and Critical Heat
Fluxes at Nitrogen Boiling on a Heating Surface with Porous Coating, Teploenergetika,
1982, No. 4, pp.66-69 (in Russian)

142. Liang H.-S., Yang W.-J., A Remedy for Hysteresis in Nucleate Boiling Through Application
of Micrographite – Fiber Nucleation Activators, Experimental Heat Transfer, vol.9 (1996),
pp.323-334

143. Lienhard J.H., Corresponding State Correlation for the Spinodal and Homogeneous
Nucleation Limits, J. Heat Transfer, vol.104 (1982), pp.379-381

144. Lienhard J.H., Karimi A.H., Corresponding States Correlations of the Extreme Liquid
Superheat and Vapor Subcooling, J. Heat Transfer, vol.100 (1978), pp.492-495

145. Lienhard J.H., Karimi A.H., Homogeneous Nucleation and the Spinodal Limit, J. Heat
Transfer, vol.103 (1981), pp.61-64

146. Liter S.G., Kaviany M., Pool Boiling CHF Enhancement by Modulated Porous-Layer Coating:
Theory and Experiment, Int. J. Heat and Mass Transfer, vol.44 (2001), 4287-4311

147. Lisovskij V.B., Investigation of Heat Transfer at Lowered Pressure Nitrogen Evaporation
from Porous Coatings, in collective work: Heat and Mass Transfer in Systems with Porous
Elements, Minsk, 1981, pp.64-67 (in Russian)

148. Ljane P., Sui Ch., Tehver Ja., Plasma Sprayed Porous Coating Thickness Effect on Liquid
Boiling Heat Transfer, Izvestia AN ESSR, Physics. Mathematics, vol.30 (1981), No.3,
pp.276-280 (in Russian)

149. Lorenz J.J., The Effects of Surface Conditions on Boiling Characteristics, Ph.D. Thesis,
M.I.T., Dept. Mech. Eng., December 1971

150. Madejski J., Heat Transfer Theory, Szczecin Univ. Techn. Publ. House, Szczecin, 1998 (in
Polish)

344 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

151. Malyshenko S.P., Peculiarities of Boiling Heat Transfer on Surfaces with Porous Coatings,
Teploenergetika, 1991, No.2, pp.38-45 (in Russian)

152. Malysenko S.P., Andrianov A.B., Non-equilibrium Phase Transitions at Boiling on Surfaces
with Porous Coatings, IVTAN Preprint, No.1-293, 1990, pp.34 (in Russian)

153. Malyshenko S.P., Andrianov A.B., The Initial Part of the Boiling Curve on Surfaces with
Porous Coatings and Nucleation Hysteresis, Teplofizika Vysokih Temperatur, vol.25 (1987),
No.3, pp.563-572 (in Russian)

154. Malyshenko S.P., Styrikovich M.A., Heat Transfer at Pool Boiling on Surfaces with Porous
Coating, Proc. 2nd Int. Symp. Multiphase Flow and Heat Transfer, 1989, Xi’an, China, vol.1,
pp. 269-284

155. Man’kovskij O.N., Ioffe O.B., Fridgant L.G., Tolchinskij A.R., On the Mechanism of a Boiling
Process on Immersed Surfaces with Capillary – Porous Coatings, Inzhynerno-Fizicheskij
Zhurnal, vol. XXX (1976), No.2, pp.310-316 (in Russian)

156. Marto P.J., Lepere V.J., Pool Boiling Heat Transfer from Enhanced Surfaces to Dielectric
Fluids, Advances in Enhanced Heat Transfer, ASME, HTD-vol.18 (1981), pp.93-102

157. Marto P.J., Mosteller W.L., Effect of Nucleate Boiling on the Operation of Low Temperature
Heat Pipes, ASME Paper, 1969, 69-HT-24, p.9

158. Marto P.J., Wanniarachchi A.S., Pulido R.J., Augmenting the Nucleate Pool-Boiling
Charakteristics of Gewa-T Finned Tubes in R-113, Augmentation of Heat Transfer in Energy
Systems, ASME, HTD-vol.52 (1985), pp.67-73

159. McKlave J.T., Dietrich F.H., Statistics, Delen Publ. Corp., San Francisco, 1988

160. Mertz R., Groll M., Vasiliev L.L., Khalatov A.A., Kovalenko G.V., Geletuha G., Pool Boiling
from Enhanced Tubular Heat Transfer Surfaces, Proc. 11th Int. Heat Transfer Conf., Kyongju
(Korea), 1998, vol.2, pp.455-460

161. Miheev M.A., Fundamentals of Heat Transfer, Goseneregizdat, Moscow, 1956 (in Russian)

162. Mikic B.B., Rohsenow W.M., A New Correlation of Pool Boiling Data Including the Effect of
Heating Surface Characteristics, J. Heat Transfer, vol.91 (1969), pp.245-250

163. Mikic B.B., Rohsenow W.M., Bubble Growth Rates in Non-uniform Temperature Field,
Progress in Heat and Mass Transfer, 1969, vol.2, pp. 283-292

164. Mizukami K., Abe F., Futagami K., A Mechanical Model for Prediction of Boiling Incipience
Condition, Proc. the 9th Int. Heat Transfer Conf., Jerusalem, 1990, paper 1-BO-20

165. Moss R.A., Kelly A.J., Neutron Radiographic Study of Limiting Planar Heat Pipe
Performance, Int. J. Heat Mass Transfer, vol.13 (1970), pp.491-502

166. Nakayama W., Thermal Management of Electronic Equipment: A Review of Technology and
Research Topics, App. Mech. Rev., vol.39 (1986), pp1847-1468

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 345


Bibliography

167. Nakayama W., Daikoku T., Kuwahara H., Nakajima T., Dynamic Model of Enhanced Boiling
Heat Transfer on Porous Surfaces, Part I: Experimental Investigation, J. Heat Transfer,
vol.102 (1980), pp.445-450

168. Nakayama W., Daikoku T., Kuwahara H., Nakajima T., Dynamic Model of Enhanced Boiling
Heat Transfer on Porous Surfaces, Part II: Analytical Modeling, J. Heat Transfer, vol.102
(1980), pp.451-456

169. Nakayama W., Daikoku T., Nakajima T., Effects of Pore Diameter and System Pressure on
Nucleate Boiling Heat Transfer from Porous Surfaces, Advances in Enhanced Heat Transfer,
ASME, HTD-vol.18 (1981), pp.147-153

170. Nakayama W., Nakajima T., Hirasawa S., Heat Sink Studs Having Enhanced Boiling for
Cooling of Microelectronic Components, ASME Paper No. 84 –WA/HT-89, 1984

171. Neter J., Wasserman W., Whitemore G.A., Applied Statistics, Allyn and Bacon, Boston,
1998

172. Nikolaev G.P., Tokalov Ju.K., Boiling Crisis on Surfaces with Porous Coating, Inzhynerno-
Fizicheskij Zhurnal, vol. XXVI (1974), No.1, pp.5-9 (in Russian)

173. Nishikawa K., Fujita Y., Nucleate Boiling Heat Transfer and Its Augmentation, Hartnett J.P.,
Irvine, Jr., T.F. (ed.), Advances in Heat Transfer, vol. 20 (1990),
pp.1-82, Academic Press

174. Nishikawa K., Ito T., Augmentation of Nucleate Boiling Heat Transfer by Prepared Surfaces,
eds. Mizushina T., Yang W.-I., Heat Transfer in Energy Problems, 1983, Hemisphere, pp.
119-26

175. Nishikawa K., Ito T., Augmentation Performance of Boiling Heat Transfer, Research of
Effective Use of Thermal Energy, vol.1, SPEY 1, Min. Educ. Sci. Cult. Tokyo, 1982, pp.39-46

176. Nishikawa K., Ito T., Tanaka K., Augmented Heat Transfer by Nucleate Boiling at Prepared
Surfaces, Proc. ASME-JSME Thermal Eng. Conf., 1983, vol.1, pp.387-393

177. Nishikawa K., Ito T., Tanaka K., Enhanced Heat Transfer by Nucleate Boiling on a Sintered
Metal Layer, Heat Transfer – Japanese Research, vol.8 (1979), pp.65-81

178. Nischik K.P., Semena M.G., Mathematical Model of Fibrous Metal Material. 2. Pore size
distribution, Inzhynerno-Fizicheskij Zhurnal, vol.53 (1987), No. 4, pp.671-672 (in Russian)

179. O’Connor J.P., You S.M., Painting Technique to Enhance Pool Boiling Heat Transfer in
Saturated FC–72, ASME, HTD-vol. 273 (1994), pp. 11-8

180. O’Connor J.P., You S.M., Painting Technique to Enhance Pool Boiling Heat Transfer in
Saturated FC–72, J. Heat Transfer, vol. 117 (1995), pp. 387–93

181. O’Neill P.S., Gottzmann C.F., Terbot J.W., Heat Exchangers for NGL, Chem. Eng. Prog., vol.
67 (1971), No. 7, pp. 80-82

182. O’Neill P.S., Gottzman C.F., Terbot J.W., Novel Heat Exchanger Increases Cascade Cycle
Efficiency for Natural Gas Liquaefaction, Adv. Cryog. Eng., vol. 17 (1972), pp. 420-37

346 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

183. Orlov V.K., Savelev V.N., Heat Transfer Enhancement at Boiling of Cryogenic Liquids at
Lowered Pressures, Teploenergetika, 1980, No.4, pp.62-64 (in Russian)

184. Orlov V.K., Savelev V.N., Boiling Heat Transfer Investigation of Cryogenic Liquids on
Surfaces with Capillary-Porous Coating, Teploenergetika, 1980, No.8, pp.66-69 (in Russian)

185. Ornatskij A.P., Semena M.G., Timofeev V.I., Experimental Investigations of Maximum
Heat Fluxes on Flat Metal Fibrous Wicks under conditions characteristic for Heat Pipes,
Inzhynerno-Fizicheskij Zhurnal, vol. XXXV (1978), No.5, pp.782-788 (in Russian)

186. Ovodkov O.A., Kovalev S.A., Boiling Liquid and Vapor Counterflow in Porous Coatings,
Teploenergetika, 1989, No.7, pp.66-68 (in Russian)

187. Pastuszko R., Analysis of Boiling Heat Transfer on Finned Surfaces with a Capillary –
Porous Coating, Ph.D. Thesis, Dept. Mechatronics and Machinery Design, Kielce University
of Technology, Kielce, 1999 (in Polish)

188. Pastuszko R., Poniewski M.E., Boiling Heat Transfer on Two-Layer Fins, Warmeaustausch
und Erneuerbare Energiequellen, VIII Int. Symp. Tagungsmateria-len, TU Szczecin,
Szczecin-Leba, 2000, Vol.1, pp.303-310

189. Pastuszko R., Poniewski M.E., Experimental Correlation for Boiling Heat Transfer on Finned
Surfaces with Porous Covering, Proc. Int. Conf. Heat Transfer with Change of Phase, Kielce,
Poland, Sci. Papers Kielce Univ. Techn., ser. Mechanics, vol.61 (1996), pp.133-44

190. Pastuszko R., Poniewski M.E., Boiling Heat Transfer on a Two-Layer Finned Surface with a
Porous Coating, Sci. Papers of Kielce Univ. Techn., ser. Mechanics, vol.70 (1999), pp.25-38
(in Polish)

191. Pastuszko R., Poniewski M.E., Staniszewski B., Heat Transfer Experimental Investigation
on Finned Surfaces Coated with Capillary-Porous Layer, Sci. Papers of Lodz Techn. Univ.,
no.110, ser. Thermal Flowing Machines, issue 749, Lodz, 1996, pp.85-94 (in Polish)

192. Pastuszko R., Poniewski M.E., Staniszewski B., Investigations of Boiling Heat Transfer
on Finned Surfaces with Porous Coatings, Proc. 9th Nat. Symp. Heat and Mass Transfer,
Augustow 1995, vol. II, pp.163-172 (in Polish)

193. Pastuszko R., Poniewski M.E., Staniszewski B., Investigation of Boiling Heat Transfer
on Finned Surfaces Coated with a Porous Layer, Sci. Papers of Kielce Univ.Techn., ser.
Mechanics, vol.55 (1995), pp.159-168 (in Polish)

194. Pastuszko R., Poniewski M.E., Wojcik T.M., Correlations for Boiling in Fibrous Porous
Structures, Proc. 4th Minsk Int. Seminar – Heat Pipes, Pumps and Refrigerators, Minsk
(Belarus), 2000, pp. 149-55

195. Pastuszko R., Poniewski M.E., Wojcik T.M., Correlation Equations for Boiling on Flat and
Finned Surfaces with a Porous Coating, Inżynieria i Aparatura Chemiczna, 2000, no.3s,
pp.104-110 (in Polish)

196. Pate M.B., Ayub Z.H., Kohler J., Heat Exchangers for the Air Conditioning and Refrigeration
Industry: State – of – the – Art Design and Technology, Compact Heat Exchangers,
Hemisphere, 1990, pp.567-590

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 347


Bibliography

197. Pawlowski L., The Science and Engineering of Thermal Spray Coatings, J. Wiley and Sons,
Chichester, 1995

198. Peterson G.P., Ortega A., Thermal Control of Electronic Equipment and Devices, Hartnett
J.P., Irvine, Jr., T.F., (eds.), Advances in Heat Transfer, vol.20 (1990), pp.181-314, Academic
Press

199. Petruhov B.S., Genin L.G., Kovalev S.A., Heat Transfer in Nuclear Power Installations,
Energoatomizdat, Moscow, 1986 (in Russian)

200. Pihlak U.L., Tehver Ja.Ch., Porous Coating Impact on Boiling Heat Transfer, Izvestia AN
SSSR, ser. Power Engineering and Transport, 1983, No.3, pp.134-138 (in Russian)

201. Polezhaev Ju.V., Kovalev S.A., On Heat Transfer Modeling at Boiling on Porous Structures,
Teploenergetika, 1990, No. 12, s. 5-9 (in Russian)

202. Poniewski M.E., Afanasev B.A., Wojcik T.M., Heat Pipes Experimental Investigations and
Modeling, State Committee for Scientific Research, grant -9-023391-01, Final report, 1994,
Kielce Univ. of Techn. (in Polish)

203. Poniewski M.E., Wisniewski M.L., Wojcik T.M., Experimental Investigations and Modeling of
Hysteresis of Boiling Heat Transfer on Thin Porous Coverings, Proc. Int. Conf. Heat Transfer
with Change of Phase, Kielce, 1996, Sci. Papers of Kielce Univ. Techn., ser. Mechanics,
vol.61 (1996), pp.179-195

204. Poniewski M.E., Wisniewski M.L., Wojcik T.M., Heat Transfer Crisis on Porous Surfaces
– Experimental Investigations and Outline of the Maximum Point Process Raised Model,
Scientific Papers, Kielce University of Technology, ser. Mechanics, vol.67, 1999, 323-338

205. Poniewski M.E., Wisniewski M.L., Wojcik T.M., Heat Transfer Modeling of Boiling in Thin
Layer Porous Structures, Sci. Papers of Wroclaw Tech. Univ., Wroclaw 1998, No.53, part II,
pp.773-774 (in Polish)

206. Poniewski M.E., Wisniewski M.L., Wojcik T.M., Experimental and Theoretical Investigations
of Boiling Heat Transfer Hysteresis in a Porous Layer, Sci. Papers of Kielce Univ. Techn.,
ser. Mechanics, vol.70 (1999), pp.71-95 (in Polish)

207. Poniewski M.E., Wisniewski M.L., Wojcik T.M., Boiling Heat Transfer on Metal Fibrous
Porous Surfaces – Experiment, Model and Verification, eds. Bar-Cohen A., Celata G.P.,
Klausner J., Fuyita Y., Boiling 2000. Phenomena and Emerging Applications, Begel House,
2000, vol.2, pp.772-788

208. Poniewski M.E., Wisniewski M.L., Wojcik T.M., A Probabilistic Model of Boiling Heat Transfer
in Thin-Layer Porous Structures, Archives of Thermodynamics, vol.21 (2000), No.3-4,
pp.55-73

209. Poniewski M.E., Wisniewski M.L., Wojcik T.M., A Maximum Point process as a Tool for
Intralayer Boiling Crisis Modeling, Sci. Papers of Kielce Univ. Techn., ser. Mechanics, vol.70
(1999), pp.61-70 (in Polish)

348 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

210. Poniewski M.E., Wojcik T.M., Experimental Investigations of Boiling Heat Transfer on
Fibrous Porous Surface, Proc. of Sci. Conf.: Scientific Issues in Thermal Power Industry,
Publ. House of Warsaw Univ.Techn., Warsaw 1993, pp.291-297 (in Polish)

211. Poniewski M.E., Wojcik T.M., Experimental Investigation of Boiling Heat Transfer Hysteresis
on Metal Fibrous Porous Coverings, Archives of Thermodynamics, vol.20 (1999), No.1-2,
pp.93-118

212. Poniewski M.E., Wojcik T.M., Effects of Heating Surface Structural Parameters on Boiling
Heat Transfer, Sci. Papers of Kielce Univ. Techn., ser. Mechanics, vol.53 (1994), pp.287-298
(in Polish)

213. Poniewski M.E., Wojcik T.M., Afanasev B.A., Analysis of Boiling Crisis on Capillary - Porous
Coatings, Sci. Papers of Kielce Univ. Techn., ser. Mechanics, vol.55 (1995), pp.169-181 (in
Polish)

214. Poniewski M.E., Wojcik T.M., Afanasev B.A., Experimental Investigations of Boiling Heat
Transfer on Porous Surfaces, Proc. 9th Nat. Heat and Mass Transfer Symposium, Augustow
1995, vol. II, pp.193-201 (in Polish)

215. Poniewski M.E., Wojcik T.M., Afanasev B.A., Heat Pipes Experimental Investigations and
Modeling, Sci. Papers of Lodz Univ. Techn., vol.107, ser. Thermal Flowing Machines, no.710,
1995, pp.43-56 (in Polish)

216. Poniewski M.E., Wojcik T.M., Afanasev B.A., Heat Transfer and Hysteresis Phenomena
in Capillary Porous Structures, Preprints of Int. Seminar „Heat Pipes, Heat Pumps,
Refrigerators”, Minsk, Byelorussia, 1995, p.10

217. Poniewski M.E., Wojcik T.M., Afanasev B.A., Boiling Heat Transfer Hysteresis on Porous
Surfaces, Sci. Papers of Kielce Univ. Techn., ser. Mechanics, vol.55 (1995), pp.183-194 (in
Polish)

218. Poniewski M.E., Wojcik T.M., Afanasev B.A., Hysteresis of Boiling Heat Transfer on Porous
Covering, Proc. 30th Nat. Heat Transfer Conf. – vol.7, Portland, USA, 1995, ASME-HTD,
vol.309, pp.11-15

219. Poniewski M.E., Wojcik T.M., Afanasev B.A., The Heat Transfer Control Method, Polish
Patent Office, Patent no.176683, Warszawa, 1999 (in Polish)

220. Poznjak V.E., Orlov V.K. Savelev V.N., The Effect of a Heating Surface Porous Coating
on Heat Transfer Enhancement in a Condenser – Evaporator, Himicheskoe i Neftjanoe
Mashinostroenie, 1980, No.3, pp.8-9 (in Russian)

221. Poznjak V.E., Savelev V.N., Application Experience of Capillary – Porous Coatings in
Cryogenic Systems and Devices, Teploenergetika, 1990, No.12, pp.9-12 (in Russian)

222. Rannenberg M., Beer H., Heat Transfer by Evaporation in Capillary Porous Wire Mesh
Structures, Letters Heat Mass Transfer, vol.7 (1980), pp.425-436

223. Rojzen L.I., Rachickij D.G., Rubin I.R., Vertogradskaja L.M., Jubina L.M., Pypkina M.B.,
Boiling Heat Transfer of Nitrogen and Freon – 113 on Metal Porous Coatings, Teplofizika
Vysokih Temperatur, vol.20 (1982), No.2, pp.304-310 (in Russian)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 349


Bibliography

224. Rosenfeld J.H., Gernert N.J., North M.T., Internally Extended Surface Heat Pipe
Evaporators for Microelectronics Cooling, ASME-HTD,vol.273 (1994), pp.93-100

225. Rudemiller G.R., Lindsay J.D., An Investigation of Boiling Heat Transfer in Fibrous Porous
Media, Proc. IX Int. Heat Transfer Conf., 1990, vol.5, pp.159-164

226. Sasin V.Ja., Intensity of Heat Transfer in Evaporation Part of Heat Pipes, Trudy MEI – “Heat
and Mass Exchanging Processes and Apparatuses”, 1974, no.198, pp.73-79

227. Sasin V.Ja., Fedorov V.N., Sorokin A.Ja., Experimental Investigation of a Heat Pipe with
Low Temperature Boiling Liquids, Proc. of Sci. and Techn. Conf. on Results of Sci. and
Research Works in 1968 -69, Section of Power Engineering, MEI, Moscow, 1969, pp.79-84
(in Russian)

228. Sasin V.Ja., Kovalev A.N., Investigations of Thermal Conductivity of Some Materials Applied
in Heat Pipes, Trudy MEI, vol.177, 1974, pp.96-99 (in Russian)

229. Sasin V.Ja., Verba M.I., Portnov V.D., Physical Conditions of Heat and Medium Mass
Transfer in a Heat Pipe, Proc. of Sci. and Techn. Conf. on Results of Sci. and Research
Works in 1968 -69, Section of Power Engineering, MEI, Moscow, 1969, pp.85-90 (in
Russian)

230. Sato T., Matsumura H., On the Conditions of Incipient Subcooled Boiling with Forced
Convection, Bull. JSME, vol.7 (1964), No. 26, pp.392-398

231. Scurlock R.G., Enhanced Boiling Heat Transfer Surfaces, Cryogenics, vol.35 (1995), No.4,
pp.233-237

232. Semena M.G., Gershuni A.N., Zaripov V.K., Heat Pipes with Metal Fibrous Capillary
Structures, Vischa Shkola, Kiev, 1984 (in Russian)

233. Semena M.G., Kiselev Ju.F., Investigation of Vapor Generation in Gravitational Heat Pipes
with Metal Fibrous Capillary – Porous Structure, Promyshlennaja Tehnika, 1983, No.4,
pp.40-43 (in Russian)

234. Semena M.G., Kostornov A.G., Gershuni A.N., Zaripov V.K., Moroz A.L., Investigation
of Thermophysical Properties of Low Temperature Heat Pipes with Metal Fibrous Wicks,
Inzhynerno - Fizicheskij Zhurnal, vol.XXXI (1976), No.3, pp. 449-455 (in Russian)

235. Semena M.G., Kravec V.Ju., Fridrihson Ju.V., Nabochenko E.A., Peculiarities of Pressure
Impact on Heat Transfer Intensity at Water Boiling on a Porous Surface, Inzhynerno -
Fizicheskij Zhurnal, vol.62 (1992), No.6, pp.779-782 (in Russian)

236. Semena M.G., Levterov A.I., Investigation of Thermophysical Characteristics of Cryogenic


Heat Pipes with Metal Fibrous Wicks, Inzhynerno - Fizicheskij Zhurnal, vol. XXXV (1978),
No.1, pp. 48-53 (in Russian)

237. Semena M.G., Nischik A.P., Investigation of Structural Parameters of Metal Fibrous Wicks of
Heat Pipes, Inzhynerno - Fizicheskij Zhurnal, vol.XXXV (1978), No.5, pp.777-781 (in Russian)

350 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

238. Semena M.G., Zaripov V.K., Investigation of Effective Thermal Conductivity of Metal Fibrous
Wicks of Low Temperature Heat Pipes, Inzhynerno - Fizicheskij Zhurnal, vol. XXXIII (1977),
No.2, pp.255-262 (in Russian)

239. Semena M.G., Zaripov V.K., Effects of Length and Diameter of Fibers on Carcass Thermal
Conductivity of Metal Fibrous Wicks of Heat Pipes, Teploenergetika, 1977, No.9, pp.82-84
(in Russian)

240. Semena M.G., Zaripov V.K., Gershuni A.N., Investigation of Heat Transfer Limitations in a
Heating Zone of Heat Pipes with Metal Fibrous Capillary Structures, Teplofizika Vysokih
Temperatur, T. 20 (1982), No. 2, s. 317-323 (in Russian)

241. Semena M.G., Zaripov V.K., Shapoval A.A., Boiling Heat Transfer Intensity on a Surface
with Porous Coatings under Capillary Feeding Conditions, Inzhynerno - Fizicheskij Zhurnal,
vol.52 (1987), No.4, pp.592-597 (in Russian)

242. Shakir S., Thome J.R., Lloyd J.R., Boiling of Methanol – Water Mixtures on Smooth and
Enhanced Surfaces, ed. Dhir V.K., Chen J.C., Jones O.C., Multiphase Flow and Heat
Transfer, ASME, HTD, vol.47 (1985), pp.1-6

243. Shakir S., Thome J.R., Boiling Nucleation of Mixtures on Smooth and Enhanced Boiling
Surfaces, Proc. 8th International Heat Transfer Conf., 1986, San Francisco, vol. 4, pp. 2081-
2086

244. Shepherd J.W., Bartell F.E., Surface Roughness as Related to Hysteresis of Contact Angles.
III. The Systems Paraffin – Ethylene Glycol – Air, Paraffin – Methyl Cellosolve – Air and
Paraffin – Methanol – Air, J. Phys. Chem., vol. 57 (1953), pp. 458-63

245. Singh A., Mikic B.B., Rohsenow W.M., Active Sites in Boiling, J. Heat Transfer. vol.98 (1976),
pp.401-406

246. Sirotin A.G., Dvojris A.D., Ingatov L.N., Holognov V.A., Effectiveness Enhancement of Heat
Exchangers - Evaporators Caused by Application of Tubes with Porous Coatings, Gazovaja
Promyshlennost’, 1976, No.12, pp.28-32 (in Russian)

247. Smirnov G.F., Appromimate Boiling Heat Transfer Theory on Surfaces Covered with
Capillary – Porous Structures, Teploenergetika, 1977, No.9, pp.77-80 (in Russian)

248. Smirnov G.F., The Generalized Model of Boiling Heat Transfer, Proc. VIII Int. Heat Pipe
Conf., 1992, Beijing, paper A-7, pp.60-64

249. Smirnov H.F., Boiling on Coated Surfaces in Porous Structures, Journal of Porous Media,
vol.4 (2001), pp.33-52

250. Smirnov G.F., Afanas’ev B.A., Experimental Investigation of Boiling Heat Transfer in Mesh
Structures of Heat Pipes, Voprosy Radioelektroniki, ser. TRTO,1979, vol.2, pp.22-27 (in
Russian)

251. Smirnov G.F., Afanasev B.A., Investigation of Vaporization in Screen Wick-Capillary


Structures, Proc. 6th Int. Heat Pipe Conf., Advances in Heat Pipe Technology, 1982,
London, pp.405-413

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 351


Bibliography

252. Smirnov G.F., Afanasev B.A., Poniewski M.E., Boiling in Capillary-Porous Structures, Proc.
Int. Conf. Heat Transfer with Change of Phase, Kielce, 1996, Sci. Papers of Kielce Univ.
Techn., ser. Mechanics, vol.61 (1996), pp.197-220

253. Smirnov G.F., Afanasev B.A., Tjurin S.A., Majstrenko V.B., Heat Transfer and Limiting Heat
Fluxes at Vaporization in Construction Capillary Porous Structures of Heat Pipes, Proc. 7th
Int. Heat Pipe Conf., Minsk, 1990, pp.149-155

254. Smirnov G.F., Koba A.L. Afanasev B.A., The Heat Transfer by Boiling Splits, Capillaries,
Wick Structures, 1978, AIAA paper-78-461, p.8

255. Smirnov G.F., Koba A.L. Afanasev B.A., Zrodnikov V.V., Boiling Heat Transfer in Gaps,
Capillaries and Other Narrow Conditions, Proc. Federal Conf. Teplo- i Massoobmen-V,
Minsk, 1976, vol. III, part I, pp.193-197 (in Russian)

256. Smirnov G.F., Vinogradova E.N., Regimes and Limitations of Heat Transfer with Vapor
Generation Inside Capillary – Porous Structures, Izvestia Akademii Nauk SSSR, ser. Power
Engineering and Transport, 1985, No.4, pp.128-36 (in Russian)

257. Solovev A.G., Djundin V.A., Danilova G.N., Impact of a Porous Coating of a Tube Bank
Surface on Heat Transfer Intensity of Ammonia, Holodil’naja Tehnika, 1988, No.10, pp.33-36
(in Russian)

258. Solov’ev S.L., Kovalev S.A., Heat Transfer at Liquid Evaporation from the Surface of a
Capillary - Porous Body, Proc. Federal Conf. Teplo- i Massoobmen – VII, Minsk, 1984, vol.6,
pp.22-25 (in Russian)

259. Solovev S.L., Kovalev S.A., Heat Transfer at Liquid Evaporation on a Porous Surface,
Teplofizika Vysokih Temperatur, vol.22 (1984), No.3, pp.528-536 (in Russian)

260. Solovev S.L., Shklover E.G., Investigation of Forced Flow Impact on Liquid Boiling Heat
Transfer on Capillary – Porous Surface, Proc. Federal Conf. Teplo- i Massoobmen – VIII,
1988, Minsk, vol.4, pp.180-182 (in Russian)

261. Staniszewski B., Poniewski M.E., Pastuszko R., Semena M., Wojcik T.M., Experimental
Investigation and Theoretical Analysis of Methods of Cooling, Thermal Stabilization and
Thermal Control of Devices Generating Large Heat Fluxes, Final report – 1995, KBN grant –
3 3349 91 02, Kielce University of Technology, Kielce (in Polish)

262. Styrikovich M.A., Leont’ev A.I., Malyshenko S.P., On Non-volatile Additions Transfer
Mechanism at Boiling on Surfaces Coated with Porous Structure, Teplofizika Vysokih
Temperatur, vol.15 (1976), No.15, pp.998-1006 (in Russian)

263. Styrikovich M.A., Malyshenko S.P., Andrianov A.B., Non-equilibrium Phase Transitions at
Boiling on Surfaces, Proc. 9th Int. Heat Transfer Conf., vol. 2, pp. 99-104, paper 1-BO-17,
Jerusalem, 1990

264. Styrikovich M.A., Malyshenko S.P., Andrianov A.B., Konovalov S.I., Peculiarities of Boiling
on Surfaces with Non-thermal Conductive Porous Coatings, Doklady AN SSSR, ser.
Technical Physics, vol.241 (1978), No.2., pp.345-348 (in Russian)

352 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

265. Styrikovich M.A., Malyshenko S.P., Andrianov A.B., Talaev I.V., Investigation of Boiling
Processes on Porous Surfaces, Proc. Federal Conf. Teplo- i Massobmen – VII, Minsk, 1984,
pp.3-8 (in Russian)

266. Shapoval A.A., Heat Transfer at Boiling of Water and Acetone on Surfaces with Metal
Fibrous Capillary – Porous Coatings, Ph.D. Thesis summary, Academy of Sciences USSR,
Kiev, Institute of Technical Physics, 1985 (in Russian)

267. Shapoval A.A., Zaripov V.K., The Stand for Boiling Heat Transfer Investigation on Surfaces
with Porous Coatings, Zavodskaja Laboratorija, 1984, No.2, pp.54-55 (in Russian)

268. Shapoval A.A., Zaripov V.K., Semena M.G., Investigation of Water Boiling Heat Transfer
Intensity on Surfaces with Metal Fibrous Porous Coatings, Teploenergetika, 1983, No.12,
pp.65-67 (in Russian)

269. Shapoval A.A., Zaripov V.K., Semena M.G., On Boiling Heat Transfer Intensity Calculations
on Surface with Porous Coatings, Izvestia AN SSSR, Power Engineering and Transport,
1989, No.3, pp.63-68 (in Russian)

270. Shejdegger A.E., Physics of Liquid Flow through Porous Material, Gostoptechnizdat,
Moskva, 1960 (in Russian)

271. Sherkiladze I.G.(ed.), Heat Pipes for Thermo-Stabilizing Systems, Energoatomizdat,


Moskva, 1991 (in Russian)

272. Tang Y.K., Wang H.T., Solid Additives for Enhancing Nucleate Pool Boiling of Binary
Mixtures on Smooth and Gewa-T Tube, Proc. X Int. Heat Transfer Conf. 1994, Brighton,
vol.6, pp.117-122

273. Tehver Ja.Ch., Boiling on Porous Surface, Theses of Tallin Technical Univ., D. Energetics.
Electrical Eng. Mining, Tallin, 1992

274. Tehver Ja.Ch., Hysteresis Phenomena at Boiling on Porous Coatings, Teploenergetika,


1990, No.12, pp.12-14 (in Russian)

275. Tehver Ja.Ch., Influence of Porous Coating on the Boiling Burnout Heat Flux, Proc. I Baltic
Heat Transfer Conf., Göteborg, 1991, vol.2, pp. 231-242

276. Tehver Ja.Ch., Sui Ch.N., Hysteresis Phenomena at Boiling on a Porous Surface, Izvestia
AN SSSR, ser. Power Engineering and Transport, 1984, No.4, pp.163-166 (in Russian)

277. Tehver Ja.Ch., Sui Ch.N., On Boiling Crisis on Plasma Sprayed Porous Surface, Teplofizika
Vysokih Temperatur, 1992, No.3, pp.561-565 (in Russian)

278. Tehver Ja.Ch., Sui Ch.N., Transient Phenomena at Liquid Boiling on a Porous Surface,
Izvestia AN ESSR, Physics. Mathematics, vol.34 (1985), No.3, pp.303-307 (in Russian)

279. Tehver Ja.Ch., Sui Ch.N., Effects of Porous Coating Parameters on Boiling Hysteresis
Phenomena, Izvestia AN ESSR, Physics. Mathematics, vol.34 (1985), No.4, pp.413-418 (in
Russian)

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 353


Bibliography

280. Tehver Ja.Ch., Sui Ch.N., Ljane P., On Liquid Boiling Capillary Heat Transfer Hysteresis on
Surface with Porous Coating, Izvestia AN ESSR, Physics. Mathematics, vol.30 (1981), No.4,
pp.376-380 (in Russian)

281. Tehver Ja.Ch., Sui Ch.N., Temkina V.S., Heat Transfer and Hysteresis Phenomenon in
Boiling on Porous Plasma – Sprayed Surface, Experimental Thermal and Fluid Science,
vol.5 (1992), pp.714-727

282. Tehver Ja.Ch., Sui Ch.N., Temkina V.S., The Porous Coating Parameters Effect on Boiling
Hysteresis Phenomena, Paper summaries, Proc. 7th Federal Conference: Two – phase flow
in power machines and devices, Leningrad, 1985, vol. I, pp.197-199 (in Russian)

283. Tehver Ja.Ch., Sui Ch.N., Temkina V.S., Liquid Boiling Initiation on Porous Surface and Heat
Transfer Hysteresis, Preprint of AN ESSR, Informatics and Technical Physics Branch, Tallin,
1989 (in Russian)

284. Tehver Ja.Ch., Temkina V., In Situ Evaluation of Porous Coating Quality, Proc. Estonian
Acad. Sci. Eng., vol.2 (1996), No.1, pp.130-136

285. Tehver Ja.Ch., Tunik A., On Boiling on Surfaces with a Porous Coating, Izvestia AN ESSR,
Physics. Mathematics, vol.28 (1979), No.1, pp.68-72 (in Russian)

286. Tehver Ja.Ch., Tunik A., On Boiling Heat Transfer Crisis on Surfaces Covered with a Porous
Material, Izvestia AN ESSR, Physics. Mathematics, vol.26 (1977), No.2, pp.194-198 (in
Russian)

287. Tehver Ja.Ch., Tunik A., On Intralayer Boiling Heat Transfer Crisis on a Surface Covered
with Porous Material, Izvestia AN ESSR, Physics. Mathematics, vol.27 (1978), No.4, pp.433-
437 (in Russian)

288. Thome J.R., Enhanced Boiling Heat Transfer, Hemisphere, 1990

289. Tolubinskij V.I., Antonenko V.A., Ivanenko G.V., Crisis Phenomena in Boiling on Submerged
Wire Mesh-Wrapped Wall, Heat Transfer – Soviet Research, vol. 21 (1989), No. 4. pp. 531-5

290. Tolubinskij V.I., Antonenko V.A., Ivanenko G.V., Impact of Defective Contact of Porous
Structure and Heating Surface on Heat Transfer at Vapor Generation, Promyshlennaja
Teplotehnika, vol.8 (1986), No.5, vol.3-6 (in Russian)

291. Tolubinskij V.I., Antonenko V.A., Kudrickij G.R., Ivanenko G.V., Ostrovskij Ju.N., Vapor
Generation Regimes in Wicks of Low Temperature Heat Pipes, Proc. Federal Conf. Teplo- i
Massobmen VIII, Minsk 1984, vol.6, pp.26-31 (in Russian)

292. Tolubinskij V.I., Antonenko V.A., Kudrickij G.R., Ostrovskij Ju.I., On Liquid Boiling
Process Mechanism on an Immersed Surface with a Mesh Capillary – Porous Coating,
Promyshlennaja Teplotehnika, vol.6 (1984), No.3, pp.3-11 (in Russian)

293. Tolubinskij V.I., Antonenko V.A., Ostrovskij Ju.N., Shevchuk E.N., Drop Carry-Over
Phenomenon in Liquid Evaporation from Capillary Structures, Letters Heat Mass Transfer,
vol.5 (1978), pp.339-347

354 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

294. Tolubinskij V.I., Antonenko V.A., Ostrovskij Ju.,N., Shevchuk E.N., Heat Transfer Mechanism
and Limitations of Vapor Generation in Heat Pipe Evaporating Zones, Izvestia AN SSSR,
ser. Power Engineering and Transport, 1979, No.1, pp.141-148 (in Russian)

295. Tolubinskij V.I., Antonenko V.A., Ostrovskij Ju.N., Shevchuk E.N., Limiting Heat Flux
Densities at Liquid Evaporation in Wick Capillaries of Low Temperature Heat Pipes,
Teplofizika Vysokih Temperatur, vol.18 (1980), No.2, pp.367-372 (in Russian)

296. Tong W., Analysis and Modeling of Nucleate Boiling in Highly-Wetting Liquids, Ph.D. Thesis,
Dept. Mechanical Engineering, Univ. Minnesota, 1989

297. Trevin R.R., Jensen M.K., Bergles A.E., Pool Boiling from Enhanced Surfaces in Pure and
Binary Mixtures of R-113 and R-11, Proc. 10th Int. Heat Transfer Conf., Brighton (UK) –
1994, vol.5, pp.165-170, paper 10-PB-26, ed. G.F. Hewitt

298. Tsay J.Y., Yan Y.Y., Lin T.F., Enhancement of Pool Boiling Heat Transfer on a Horizontal
Water Layer through Surface Roughness and Screen Coverage, Heat Mass Transfer, vol.32
(1996), pp.17-26

299. Tunik A., Bolshakov A., Tehver Ja.Ch, Heating Surface Porous Coating Impact on Heat
Transfer Intensity at Boiling of Liquid Dielectrics, Izvestia AN ESSR, Physics. Mathematics,
vol.27 (1978), No.3, pp.364-369 (in Russian)

300. Vasilev L.L., Hydrodynamics and Heat Transfer Hysteresis Problems in Heat Pipes, in
collective work: “Heat and Mass Transfer in Systems with Porous Elements”, ITMO, Minsk,
1981, pp.3-11 (in Russian)

301. Vasilev L.L., Abramenko A.N., Kanonchik L.E., Liquid Boiling Heat Transfer on Porous
and Extended Heating Surfaces, Inzhynerno - Fizicheskij Zhurnal, vol. XXXIV (1978), No.4,
pp.741-761 (in Russian)

302. Vasilev L.L., Zhuravlev A.S., Novikov M.N., Vasilev L.L., Jr., Experimental Investigation of
Propane Boiling in Porous Structures, IV Minsk Int. Seminar – Heat Pipes, Heat Pumps,
Refrigerators, Minsk (Belarus), 2000, pp.245-255

303. Wallis G. B., One Dimensional Two Phase Flow, McGraw-Hill, 1969

304. Wang D.Y., Cheng J.G., Zhang H.J., Pool Boiling Heat Transfer from T - finned Tubes at
Atmospheric and Super – atmospheric Pressures, ASME, HTD, vol. 159, Phase Change
Heat Transfer, 1991, pp.143-147

305. Wang C.H., Dhir V.K., Effect of Surface Wettability on Active Nucleation Site Density During
Pool Boiling of Water on a Vertical Surface, J. Heat Transfer,
vol.115 (1993), pp.659-669

306. Wang C.H., Dhir V.K., On the Gas Entrapment and Nucleation Site Density During Pool
Boiling of Saturated Water, J. Heat Transfer, vol.115 (1993), pp.670-679

307. Wang B.-X., Ma J., Shi M.-H., Hysteresis Characteristics of Nucleate Pool Boiling Heat
Transfer, Proc. UK National and First European Conf. Thermal Sci., 1992, vol.1, pp.139-145

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 355


Bibliography

308. Warner D.F., Park E.L., Jr., Mayhank G., Nucleate Boiling Heat Transfer of Liquid Nitrogen
from Plasma Deposited Polymer Coated Surfaces, Int. J. Heat Mass Transfer, vol.21 (1978),
pp.137-144

309. Webb R.L., Advances in Modeling Enhanced Heat Transfer Surfaces, Proc. X Int. Heat
Transfer Conf., 1994, Brighton, UK, vol.1, pp. 445-459 (paper GK-17)

310. Webb R.L., Nucleate Boiling on Porous Coated Surfaces, Heat Transfer Eng., vol.4 (1983),
No.3-4, pp.71-82

311. Webb R.L., Principles of Enhanced Heat Transfer, Wiley Interscience, 1994

312. Webb R.L., The Evolution Enhanced Surfaces Geometries for Nucleate Boiling, Heat
Transfer Eng., vol.2 (1981), No.3-4, pp.46-69

313. Webb R.L., Chien L.H., McQuade W.F., Imadojemu H.E., Pool Boiling of Oil-Refrigerant
Mixtures on Enhanced Tubes, Proc. ASME-JSME Thermal Eng. Conf., 1995, vol.2, pp.247-
255

314. Webb R.L., Haider I., An Analytical Model for Nucleate Boiling on Enhanced Surfaces, eds.
Dhir V.J., Bergles A.E., ASME-HTD, Pool and External Boiling, 1992, pp.345-360

315. Webb R.L., Pais C., Nucleate Boiling Data for Five Refrigerants on Plain, Integral-Fin and
Enhanced Tube Geometries, Int. J. Heat Mass Transfer, vol.35 (1992), pp.1893-1904

316. Wenzel R.L., Surface Roughness and Contact Angle (comm. to the editor) J. Phys. Colloid
Chem., vol.53 (1949), pp.1466-1467

317. Wojcik T.M., The analysis of the Heating Surface Capillary – Porous Coating Parameters
Effect on Boiling Heat Transfer, Ph.D. thesis, Kielce University of Technology, Mechanical
Eng. Dept., Kielce, 1997 (in Polish)

318. Xin M. D., Chao Y. D., Analysis and Experiment of Boiling Heat Transfer on
T-Shaped Finned Surfaces, Chem. Eng. Comm., vol.50 (1987), pp.185-199

319. Xiuling Y., Hongji X., Yaozhen S., Dongji C., Yuweng Z., Hongzhang Q., Experimental
Research of Nucleate Pool Boiling Heat Transfer From Compound Enhanced Porous
Surface in Liquid Nitrogen, Cryogenics, vol.30 (1990)

320. Xiuling Y., Hongji X., Yuweng Z., Hongzhang Q., Pool Boiling Heat Transfer to Liquid
Nitrogen from Porous Metallic Coatings of Tube Bundle and Experimental Research of
Hysteresis Phenomenon, Cryogenics, vol.29 (1989), pp.460-462

321. Yamaguchi H., James D.D., Effect of Wire Meshes on Boiling Heat Transfer from a Plane
Heating Surface, Proc. the 1st World Conf. Experimental Heat Transfer, Fluid Mechanics
and Thermodynamics, Dubrovnik, 1988, pp.587-594

322. Yang S.R., Kim R.H., A Mathematical Model of Pool Boiling Nucleation Site Density in Terms
of Surface Characteristics, Int. J. Heat Mass Transfer, vol.31 (1988), pp.1127-35

323. Yilmaz S., Westwater J.W., Effect of Commercial Enhanced Surfaces on the Boiling Heat
Transfer Curve, Advances in Enhanced Heat Transfer, ASME, HTD-vol.18 (1981), pp.73-91

356 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

324. You S.M., Simon T.W., Bar–Cohen A., A Technique for Enhancing Boiling Heat Transfer with
Application to Cooling of Electronic Equipment, IEEE Trans., CHMT, vol.15 (1991), No.5,
pp.90-96

325. Zaripov V.K., Effective Conductivity Study of Heat Pipe Metal Fibrous Wicks, Ph.D. thesis
summary, Institute of Technical Thermo – Physics, AN USSR, Kiev, 1978 (in Russian)

326. Zaripov V.K., Semena M.G., Shapoval A.A., Levterov A.I., Boiling Heat Transfer Intensity on
a Surface with Porous Coatings under Free Convection Conditions, Inzhynerno - Fizicheskij
Zhurnal, vol.57 (1989), No.2, pp.181-186 (in Russian)

327. Zhang H., Chen L., Pool Boiling Heat Transfer on the Sintered Porous Surfaces, T. Nejat
Verizoglu (ed.), Particulate Phenomena and Multiphase Transport, Hemisphere, 1988, vol.1,
pp. 529-538

328. Zhang H., Zhang X., Experimental Research on Pool Boiling Heat Transfer from Spraying
Porous Surface, Preprints 7th Int. Heat Pipe Conf., Minsk, 1990, pp.187-194

329. Zhang H., Zhang Y., Hysteresis Characteristic of Boiling Heat Transfer from Powder –
Porous Surface, Advances in Phase Change Heat Transfer, Int. Academic Press, China,
1988, pp.98-103

330. Zhang H., Zhao X., Critical Heat Flux for a Thin Powder Porous Layer, Proc. the 2nd World
Conf. Experimental Heat Transfer, Fluid Mechanics and Thermodynamics, Dubrovnik, 1991,
pp.627-632

331. Zhang Y., Zhang H., Boiling Heat Transfer from a Thin Powder Porous Layer at Low and
Moderate Heat Flux, Proc. the 2nd Int. Symp. Multiphase Flow and Heat Transfer, eds.
X.J.Chen, T.N. Veziroglu, C.L. Tien, 1992, vol.1, pp.358-366

332. Zhao X., Zhang H., Experimental Study of Pool Boiling Heat Transfer from Powder Porous
Surface at Higher Heat Fluxes, Advances in Phase Change Heat Transfer, Int. Academic
Press, China, 1988, pp.236-241

333. Zuber N., On the Stability of Boiling Heat Transfer, Trans. ASME, vol. 80 (1958), pp. 711-20

334. Zuber N., Nucleate Boiling, the Region of Isolated Bubbles and the Similarity with Natural
Convection, Int. J. Heat Mass Transfer, vol.6 (1963), pp.53-78

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 357


Bibliography

358 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

Notation

a -- thermal diffusivity, m2 s –1
-- exponent
-- number of cells per area unit, m –2
A -- similarity number for boiling inside a capillary channel
-- constants (different dimensions)
Ar -- Archimedes number
b -- slot height, m
-- exponent
-- most probable pore diameter, dimensionless
-- number of dry cells per area unit, m –2
B -- geometrical coefficient, m –1
-- dimensionless constants
-- number of capillary channels feeding one vapor-filled channel
Bi -- Biot number
Bo -- Bond number
c -- exponent
-- critical area, m2
C -- constants (different dimensions)
d -- diameter, microfin characteristic dimension (substitute diameter), m
-- exponent
D -- diameter, slot width, m
-- mesh eyelet characteristic dimension, m
e -- exponent
E -- specific kinetic energy of a steam jet, J m –3
Eu -- Euler number
f -- cross-sectional area, m2
-- friction coefficient
-- functions (different dimensions)
-- ratio of surface areas, surface fraction
F -- surface area, total area, m2
Fr -- Froude number
g -- gravitational acceleration, m s –2
G -- Gibbs free energy, J
-- mass flux, kg m –2 s –1
Ga -- Galileo number
Gr -- Grashof number

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 359


Notation

h -- height, m

hlg -- latent heat of vaporization, J kg –1

H -- penetration depth, m

Im -- imaginary Bessel function, dimensionless

Ja -- Jakob number
k -- capillarity constant, m
-- exponent, proportionality factor
-- dimensionless constants
-- natural number
K -- dimensionless bubble curvature
-- dimensionless number, constants (different dimensions)
-- permeability, m2

Ka -- number of open capillaries per area unit, m –2

Kc -- similarity number specifying thermal properties of the boiling agent

Km -- similarity number

Kp -- number of vapor channels per area unit, m –2

Kp -- similarity number describing geometrical properties of the porous coating

Ks -- dimensionless number

Ks -- similarity number specifying the impact of surface tension

Kλ -- similarity number specifying the ratio of the water-soaked porous coating thermal
conductivity to the liquid thermal conductivity ratio

Kφ -- similarity number specifying the impact of the surface development

l -- distance, width, linear characteristic dimension, m


-- exponent
L -- dimensionless number
-- length, distance, characteristic dimension, m

Le -- characteristic dimension, m

m -- mass, kg
-- exponent, dimensionless constant, number of capillary channels
-- coefficient characterizing heat transfer on a fin, m –1
-- ratio of surface areas
M -- mass flow rate, kg s –1
-- the maximum point process
n -- number of capillary channels per area unit, m –2
-- exponent; number of layers
-- coefficient characterizing heat transfer on the fin face
-- linear coordinate normal to the surface, m

360 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

N -- boiling site density, m –2


-- density of openings in the heating surface, m –2
-- number of pores per area unit, m –2
-- number of menisci
-- number of fins
-- Poisson point process
Ncf -- similarity number of the constant heat flux
Np -- dimensionless number describing the impact of pressure
Nu -- Nusselt number
Nϕ -- dimensionless number describing filtering properties of the capillary-porous structure
O -- circumference, m
-- circumference per area unit, m –1
o -- exponent
p -- pressure, N m –2, Pa
-- exponent
-- significance level
-- probability, thinning function
Pe -- Peclet number
Pr -- Prandtl number
q -- heat flux, W m –2
-- heat generation per unit volume, W m –3
Q -- amount of heat, J
-- rate of heat generation, power, W
-- heat flux per unit length, W m –1
R -- radius, half of the mesh eyelet characteristic dimension, m
-- individual gas constant, J kg –1 K–1

R2 -- determination coefficient

Ra -- roughness, m
Re -- Reynolds number
-- real Bessel function, dimensionless
s -- dimensionless roughness measure
-- pitch, spacing, m
-- position coefficient
-- mean standard deviation, m
-- interfin spacing, m
sk -- scale
St -- Stanton number
T -- temperature, K (°C)
u -- exponent
-- velocity, m s –1
v -- specific volume, m3 kg –1

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 361


Notation

V -- dimensionless volume
-- volume, m3
w -- velocity, m s –1
-- number of layers, number of dimensions of spatial network
-- gamma distribution constant
-- fin width, m
We -- Weber number
x -- linear coordinate, m
-- cell
y -- linear coordinate, m
z -- coordination number
-- gamma distribution constant, m –1
Z -- Nusselt number inverse (see Fig.3.31)
-- maximum heat flux parameter (see eq. 3.40)

Greek

α -- heat transfer coefficient, W m –2 K–1


-- volumetric heat transfer coefficient, W m –3 K–1
β -- angle
-- coefficient of thermal expansion, K–1
-- coefficient of the porous layer inertial resistance, m –1
γ -- function of the pore diameter density distribution, m –1
-- dimensionless function
Γ -- mass flux per length unit, kg m –1 s –1
-- gamma distribution function
Δ -- layer thickness, m
-- height of equivalent (conventional) microfin, m
-- oscillation amplitude, m
-- increment
ΔT -- temperature difference, superheat, K, °C
ε -- porosity
ζ -- division coefficient, local friction coefficient
η -- ratio of thermal conductivities
-- dimensionless function
ϑ -- slope of saturation curve p – T, N m –2 K–1
-- temperature difference, K
θ -- wetting angle

362 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

λ -- statistical parameter, m –1
-- thermal conductivity, W m –1 K–1
-- geometric scaling coefficient
-- wavelength, m
-- intensity of Poisson process
Λ -- coefficient characterizing porous coating geometry, m –1
-- percolation threshold
µ -- absolute viscosity, N s m –2
ν -- kinetic viscosity, m2 s –1
-- intensity of drying
ξ -- correction coefficient
-- variable
π -- quantity proportional to pressure, N m –1
Π -- integral of pore dimension distribution curve
-- percolation threshold
-- specific work, specific potential energy, J m –3
ρ -- density, kg m –3
σ -- surface tension, N m –1
-- surface free energy, J m –2
τ -- time, s
-- tangential stress, N m –2
φ -- angle, radian
-- degree of dryness
Φ -- function of the boiling liquid thermal properties, W–2/3 K1/2 m5/6
-- normal distribution
χ -- dimensionless pore diameter
-- function of pore diameter distribution
ψ -- specific chemical potential, J kg –1
-- porosity function
Ψ -- thermodynamic potential, J
ω -- frequency, s –1
-- dimensionless parameter
Ω -- thermal contact resistance, K m2 W–1

Subscripts

a -- active
a.b. -- as before
act -- actual, current
am -- amplitude
bd -- breakdown

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 363


Notation

br -- binder
cal -- calorimeter
cap -- capillary
chnl -- channel
con -- convection
cph -- change of phase
cr -- critical
cycl -- cycle (of vapor formation)
c+c -- conduction and convection
d -- referring to dimension d
-- additional
del -- delay
dep -- departure
e -- equivalent, characteristic
ef -- effective
el -- electrical
entp -- entrapment
env -- environment
ev -- evaporation
exp -- experimental
ext -- external
eyt -- mesh eyelet
f -- friction
fb -- film boiling
fbr -- fiber
fin -- fin
fs -- face surface
g -- gas, vapor
grn -- grain
gth -- growth
h -- hydraulic
hp -- heat pipe
hs -- heating surface
hst -- hysteresis
htg -- heating
H -- Helmholtz instability
incip -- boiling incipience
incr -- increase
ind -- industrial
int -- initial
intcs -- intra-layer crisis

364 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved


Nucleate Boiling on Micro-Structured Surfaces

l -- liquid
lg -- liquid – gas (vapor) interface
lmt -- limit
m -- metal
-- mean, substitute
max -- maximum
micro -- microlayer
min -- minimum
ms -- meniscus
n -- nominal
nb -- nucleate boiling
n-ev -- no evaporation
opn -- open, opening
opt -- optimum
p -- pitch
-- pore or cavity
-- constant pressure
-- thinned point process
pa -- projected area
ps -- porous surface
pr -- powder
real -- real
s -- solid body
sat -- saturation
sb -- surface boiling
sbt -- substitute
sct -- suction
sg -- solid body – gas
skl -- skeleton
sl -- solid body – liquid
sm -- smooth
stat -- static
sub -- subcooling
tun -- tunnel
T -- thermal
-- Taylor instability
vol -- volume
wait -- waiting
λ -- heat conduction
σ -- surface tension
0 -- set

© 2008 M. E. Poniewski and J. R. Thome. All rights reserved 365


Notation

1 f, 2 f -- single-phase, two-phase
– -- receding
+ -- moving forward

Superscripts

cell -- cell inside the porous layer


— -- mean value
~ -- dimensionless
* -- distinctive value (specific, characteristic, modified)
-- increment with reference to saturation conditions

366 © 2008 M. E. Poniewski and J. R. Thome. All rights reserved

You might also like