0% found this document useful (0 votes)
241 views267 pages

Gao S. Capillary Mechanics 2025

The document is a book titled 'Capillary Mechanics' authored by Shiqiao Gao, Lei Jin, and Deyi Fu, published by World Scientific Publishing. It covers various aspects of capillary action, surface tension, contact angles, and related mechanical principles, providing both theoretical foundations and practical measurement methods. The authors are affiliated with prestigious institutions in China and have significant research backgrounds in related fields.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
241 views267 pages

Gao S. Capillary Mechanics 2025

The document is a book titled 'Capillary Mechanics' authored by Shiqiao Gao, Lei Jin, and Deyi Fu, published by World Scientific Publishing. It covers various aspects of capillary action, surface tension, contact angles, and related mechanical principles, providing both theoretical foundations and practical measurement methods. The authors are affiliated with prestigious institutions in China and have significant research backgrounds in related fields.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 267

CAPILLARY

MECHANICS
This page intentionally left blank
CAPILLARY
MECHANICS

Shiqiao Gao
Beijing Institute of Technology, China

Lei Jin
Beiijing Institute of Technology, China

Deyi Fu
China Electric Power Research Institute, China

World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • TAIPEI • CHENNAI
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE

Library of Congress Control Number: 2025009783

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library.

CAPILLARY MECHANICS
Copyright © 2025 by World Scientific Publishing Co. Pte. Ltd.
All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.

ISBN 978-981-98-1021-5 (hardcover)


ISBN 978-981-98-1022-2 (ebook for institutions)
ISBN 978-981-98-1023-9 (ebook for individuals)

For any available supplementary material, please visit


https://www.worldscientific.com/worldscibooks/10.1142/14222#t=suppl

Desk Editors: Nambirajan Karuppiah/Julio Hong/Amanda Yun

Typeset by Stallion Press


Email: [email protected]

Printed in Singapore
About the Authors

Gao Shiqiao, PhD, is a professor and doctoral


tutor at Beijing Institute of Technology,
China. He was awarded the Government
Special Allowance of the State Council in
1997. He has been a research fellow of the
Alexander von Humboldt Foundation since
1990. His current research interests include
nonlinear structural dynamics, MEMS/NEMS
technology and mechanics, and impact dynam-
ics. He has published more than 150 papers in
many important international and domestic academic journals and
authored eight monographs.

Jin Lei is currently an associate professor at


the School of Mechatronic Engineering, Beijing
Institute of Technology, China. Her current
research interests include electromechanical
and control technology, microelectromechani-
cal technology, and micro energy technology.
She has published over 30 papers in academic
journals and participated in the writing and
publication of six monographs.

v
vi Capillary Mechanics

Fu Deyi, PhD, is a professorate senior


engineer and senior technical expert at China
Electric Power Research Institute (CEPRI).
His research interests include wind tur-
bine measurement and certification, structural
dynamics analysis and control, fault diagnosis
and monitoring, and fatigue assessment. He is
the leader of the remote sensing expert group
of the Measuring Network of Wind Energy
Institutes (MEASNET), a member of the power
performance and mechanical load expert group of MEASNET, a
member of IEC TC88 MT13, a peer assessor of IEC RE wind energy,
and also a member of the SG551 expert group. He also serves as
a master supervisor at Beijing Institute of Technology (BIT) and
North China Electric Power University (NCEPU). He is a nationally
registered metrologist. He has published over 30 papers, obtained
16 authorized patents, and 11 software copyrights.
Contents

About the Authors v


Introduction xiii

1. Capillary and Capillary Action 1


1.1 The Phenomenon and the Function of Capillary . . 1
1.2 The Capillary Action of M/NEMS . . . . . . . . . 4
1.3 The Development of Capillary Mechanics . . . . . . 9
2. Basic Conceptions 17
2.1 State of Matter and Phase . . . . . . . . . . . . . . 17
2.1.1 State of matter . . . . . . . . . . . . . . . . . 17
2.1.2 Phase . . . . . . . . . . . . . . . . . . . . . . 18
2.2 Surface, Interface, and Bulk . . . . . . . . . . . . . 20
2.2.1 The connotation of surface, interface
and bulk . . . . . . . . . . . . . . . . . . . . 20
2.2.2 Features of interface . . . . . . . . . . . . . . 20
2.3 Surface Free Energy . . . . . . . . . . . . . . . . . . 22
2.4 The Interface Free Energy and Interfacial
Tension . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.5 Cohesion and Adhesion . . . . . . . . . . . . . . . . 27
2.6 Interface Layer and Surface Force . . . . . . . . . . 29
2.7 Chemical Potential . . . . . . . . . . . . . . . . . . 32
2.7.1 The concept of chemical potential . . . . . . 32
2.7.2 The relationship between chemical potential
and temperature pressure . . . . . . . . . . . 36

vii
viii Capillary Mechanics

2.7.3 The application of chemical potential in the


phase transition . . . . . . . . . . . . . . . . 38
2.7.4 Ideal gas chemical potential . . . . . . . . . . 39
3. Surface Tension and Interfacial Tension 41
3.1 Conception of Surface Tension . . . . . . . . . . . . 41
3.2 Molecule Meaning of Surface (Interface) Free
Energy–Surface (Interface) Tension . . . . . . . . . 45
3.3 Concept of Interfacial Tension . . . . . . . . . . . . 48
3.4 Factors Influencing Surface Tension . . . . . . . . . 49
3.4.1 Influence of the type of matter on
surface tension . . . . . . . . . . . . . . . . . 49
3.4.2 Influence of interface . . . . . . . . . . . . . 49
3.4.3 Influence of temperature . . . . . . . . . . . 51
3.4.4 Influence of density . . . . . . . . . . . . . . 54
3.4.5 Influence of pressure . . . . . . . . . . . . . . 54
3.5 Methods of Measurement of Surface and Interfacial
Tensions . . . . . . . . . . . . . . . . . . . . . . . . 55
3.5.1 Capillary rising method . . . . . . . . . . . . 56
3.5.2 Du Noüy ring method . . . . . . . . . . . . . 57
3.5.3 Maximum bubble pressure method . . . . . . 59
3.5.4 Wilhelmy plate method . . . . . . . . . . . . 62
3.5.5 Stalagmometeric method . . . . . . . . . . . 63
3.6 Methods of Measuring the Surface Tension
of Solids . . . . . . . . . . . . . . . . . . . . . . . . 66
3.6.1 Critical surface tension method . . . . . . . . 66
3.6.2 Calculating the surface tension of solid
utilizing the surface tension of high polymer
liquid or melt . . . . . . . . . . . . . . . . . . 67
3.6.3 Estimation method . . . . . . . . . . . . . . 68
3.7 Example of Surface Tension Measurement with
Interfacial Tension Instrument ZYW-200B . . . . . 69
3.7.1 Structure and working principle . . . . . . . 69
3.7.2 Preparing for the experiment . . . . . . . . . 70
3.7.3 Calibrating the instrument . . . . . . . . . . 71
3.7.4 Measurement of surface tension and
interfacial tension of liquid . . . . . . . . . . 72
3.7.5 Correcting the measured tension . . . . . . . 74
3.7.6 Note . . . . . . . . . . . . . . . . . . . . . . . 75
Contents ix

3.8 Elasticity of Liquids . . . . . . . . . . . . . . . . . . 76


3.8.1 Linearity in liquid tensile process . . . . . . . 77
3.8.2 Recoverability in liquid tensile process . . . . 77
3.8.3 The maximum tensile force of liquid on per
unit area is independent of the area . . . . . 81
3.8.4 The Hook’s law of liquid tension . . . . . . . 82
4. Contact Angle and Wetting 85
4.1 Basic Concepts . . . . . . . . . . . . . . . . . . . . 85
4.1.1 Contact angle . . . . . . . . . . . . . . . . . 85
4.1.2 Wetting . . . . . . . . . . . . . . . . . . . . . 86
4.1.3 Microscopic explanation of wetting
phenomenon . . . . . . . . . . . . . . . . . . 89
4.2 Influence Factors of Contact Angle . . . . . . . . . 91
4.2.1 Influence on contact angle by surface
roughness . . . . . . . . . . . . . . . . . . . . 92
4.2.2 Influence on contact angle by the
composition of surface materials . . . . . . . 92
4.2.3 Contact angle changing with
temperature . . . . . . . . . . . . . . . . . . 93
4.3 Contact Angle Hysteresis and Influencing
Factors . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.3.1 Influence of surface roughness . . . . . . . . 94
4.3.2 Influence of surface pollution . . . . . . . . . 94
4.3.3 Influence of deposits on solid surface . . . . . 95
4.4 Mechanics Balance Relationship of Three-Phase
Line and the Young Equation . . . . . . . . . . . . 95
4.5 Types of Wetting . . . . . . . . . . . . . . . . . . . 96
4.5.1 Bedew . . . . . . . . . . . . . . . . . . . . . . 97
4.5.2 Soak . . . . . . . . . . . . . . . . . . . . . . . 98
4.5.3 Spreading . . . . . . . . . . . . . . . . . . . . 99
4.6 Principle and Method of Measuring Contact
Angle . . . . . . . . . . . . . . . . . . . . . . . . . . 102
4.6.1 The principle of measurement for contact
angle by hypsometry . . . . . . . . . . . . . . 103
4.6.2 Photo goniometric method . . . . . . . . . . 104
4.6.3 Specular reflection method . . . . . . . . . . 105
4.6.4 Parallel beam method . . . . . . . . . . . . . 106
x Capillary Mechanics

4.6.5 Calculating contact angle from rise and


descent method by capillary tube . . . . . . 107
4.6.6 Principle of dynamometry to measure
contact angle . . . . . . . . . . . . . . . . . . 108
4.7 Measuring Method of Contact Angle Tester Based
on Photogoniometric Method . . . . . . . . . . . . 110
4.7.1 Construction compose of JYSP-180 tester . . 110
4.7.2 Installation and use of tester . . . . . . . . . 110
5. Young–Laplace Equation and Its Application 115
5.1 Young–Laplace Equation . . . . . . . . . . . . . . . 115
5.2 Jurin’s Law in Capillary . . . . . . . . . . . . . . . 122
5.3 Liquid Bridge Force between Two Plates . . . . . . 123
5.4 Falling Stream of Fluid Breaks up into Smaller
Packets . . . . . . . . . . . . . . . . . . . . . . . . . 124
6. Kelvin Equation 129
6.1 Kelvin Equation . . . . . . . . . . . . . . . . . . . . 129
6.2 Application of the Kelvin Function . . . . . . . . . 135
6.2.1 Capillary pore condense . . . . . . . . . . . . 135
6.2.2 Artificial rainfall . . . . . . . . . . . . . . . . 135
6.2.3 Calculating meniscus radius . . . . . . . . . 136
6.3 Metastability State and the Production
of a New Phase . . . . . . . . . . . . . . . . . . . . 137
6.3.1 Supersaturation vapor . . . . . . . . . . . . . 137
6.3.2 Overheating liquid . . . . . . . . . . . . . . . 138
6.3.3 Subcooled liquid . . . . . . . . . . . . . . . . 138
6.3.4 Supersaturated solution . . . . . . . . . . . . 139
7. Surface Tension Gradient and Marangoni
Effect 141
7.1 Surface Tension Gradient . . . . . . . . . . . . . . . 141
7.2 Marangoni Effect . . . . . . . . . . . . . . . . . . . 143
7.3 Phenomenon Explained by Marangoni Effect . . . . 145
7.3.1 Soap bubble . . . . . . . . . . . . . . . . . . 145
7.3.2 Tears of wine . . . . . . . . . . . . . . . . . . 146
7.3.3 Wafer drying in wet processing . . . . . . . . 147
7.4 Marangoni Convection . . . . . . . . . . . . . . . . 149
Contents xi

7.5 Flow State of Marangoni Convection . . . . . . . . 151


7.5.1 Flow state of laminar flow . . . . . . . . . . 151
7.5.2 Flow state of oscillation flow . . . . . . . . . 152
7.6 Research Method of Marangoni Effect . . . . . . . . 153
7.6.1 Linear analytical method . . . . . . . . . . . 154
7.6.2 Nonlinear analytical method . . . . . . . . . 155
7.6.3 Numerical calculation . . . . . . . . . . . . . 157
7.6.4 Displaying of Marangoni convection
field . . . . . . . . . . . . . . . . . . . . . . . 157
7.6.5 Indirect measurement method . . . . . . . . 159
7.7 Summary of Marangoni Convection Theoretical
Analysis . . . . . . . . . . . . . . . . . . . . . . . . 161
8. Capillary Dynamics 163
8.1 Hagen−Poiseuille Equation . . . . . . . . . . . . . . 163
8.2 Capillary Flow Rate . . . . . . . . . . . . . . . . . . 164
8.3 Capillary Dynamics Equation . . . . . . . . . . . . 171
8.4 Stages of Capillary Flow . . . . . . . . . . . . . . . 174
8.4.1 Phases of capillary flow . . . . . . . . . . . . 175
8.4.2 Division of the different stages . . . . . . . . 178
9. Concept and Function of Liquid Bridge 181
9.1 Concept of Liquid Bridge . . . . . . . . . . . . . . . 181
9.2 Formation of Liquid Bridge . . . . . . . . . . . . . 181
9.3 Pull-Off Force Resulted from the
Liquid-Bridge . . . . . . . . . . . . . . . . . . . . . 185
9.3.1 Concept of pull-off force . . . . . . . . . . . . 185
9.3.2 Formation mechanism of pull-off force . . . . 186
9.3.3 The analysis model of pull-off force . . . . . 186
9.3.4 The total free energy of the
liquid-bridge . . . . . . . . . . . . . . . . . . 193
9.4 The Solution on the Interface of Liquid and Vapor
of the Liquid Bridge . . . . . . . . . . . . . . . . . . 195
9.5 The Pull-Off Force of the Top Solid
in the Liquid-Bridge System . . . . . . . . . . . . . 196
9.5.1 The solution for constant wetted areas . . . . 196
9.5.2 An approximate solution for liquid with a
constant volume . . . . . . . . . . . . . . . . 197
9.5.3 An accurate solution for large gap . . . . . . 199
xii Capillary Mechanics

9.5.4 An accurate solution for the sphere-plane


liquid-bridge system with a large gap . . . . 203
9.5.5 The edge-effect of liquid-bridge . . . . . . . . 209
10. MEMS and Capillary Action 213
10.1 Capillary Action in the Processing of MEMS . . . . 213
10.1.1 Capillary adhesion in MEMS . . . . . . . . . 214
10.1.2 Control of capillary adhesion . . . . . . . . . 217
10.2 Capillary-Driven MEMS Fluid Self-Assembly . . . . 220
10.3 Dynamic Characteristics of Parallel-Plate
Structure under Capillary Force . . . . . . . . . . . 224
11. Capillary Wave 231
11.1 Phenomenon of Capillary Wave . . . . . . . . . . . 231
11.2 Dispersion Relation of the Free Surface Wave . . . 233
11.3 Analysis of the Characteristics of Free Surface
Waves and the Concept of Capillary Waves . . . . . 239

Bibliography 241
Index 247
Introduction

In recent years, micro/nano science has been continuously devel-


oping, and micro/nano technology has been continuously advanc-
ing. Numerous high-tech concepts and tangible achievements have
emerged in materials, structures, and systems. Some of them have
been applied in industrial and information-based commodity produc-
tion, while others have great application prospects and potential.
As the micro/nanoscale ranges from micrometers (10−6 meters) to
nanometers (10−9 meters), it is not only different from the traditional
macroscopic scale visible to the naked eye above millimeters but also
different from the molecular atomic scale invisible to the naked eye
at the traditional physical and chemical levels. Therefore, for a long
time, its role has been overlooked at the micro level, and its exis-
tence has been overlooked at the macro level. It was referred to as
the “neglected size world” by Wolfgang Ostwald. However, such a
size is the transition size from macro to micro, the connection size
between macro and micro, and also the interaction size of interfaces
and surfaces. At this size scale, it involves many production and daily
life issues. The connection of interfaces, surface adsorption, interface
adhesion, mechanical lubrication, friction, wear, surface pollution
and cleaning, surface corrosion, etc., are all related to the effects
of this scale. Especially when the size of machinery and structures
itself is reduced to this level, the interaction between different phases
or states of matter becomes more significant. Many of these issues
involve capillary action. Therefore, it is worth studying and gaining
a deeper understanding.

xiii
xiv Capillary Mechanics

Capillary action results from the influence of intermolecular forces


at the microscopic level and manifests as a macroscopic visible phe-
nomenon, thus spanning a large scale. Traditional capillary theory
is mostly based on macroscopic phenomenological research, such as
thermodynamic theory, energy theory, and fluid mechanics theory.
Even if it involves microscopic aspects, it only explores the influence
of highly microscopic molecular forces on the mechanism of capillary
phenomena from a qualitative perspective, rarely involving the scale
of micro and nano. However, due to the fact that capillary action
itself involves three “phases” or “states” of gas, liquid, and solid and
mainly involves surface and interface interactions, its effect is strong
at the micro and nano scales. At present, there is still a lack of spe-
cialized English books like this. From this perspective, there is both
a need for a re-understanding of capillary phenomena and capillary
effects, as well as a need for in-depth research. This book is written
from this perspective.
Capillary mechanics can be traced back to ancient times. As early
as the 13th century, people proposed the concept of capillaries and
attempted to understand the laws of capillary action from the per-
spective of blood circulation.
Capillary mechanics is also very modern. Until the 21st century,
people have been constantly exploring capillary effects in different
fields, such as capillary effects in micro and nano fields, and capillary
effects in surface science.
Capillary mechanics is very traditional. This science was born
almost contemporaneously with traditional Newtonian mechanics
and developed alongside it. Many laws are also based on Newtonian
mechanics.
Capillary mechanics is also very fashionable. Many new
interdisciplinary fields in contemporary times are either derived from
or closely related to capillary mechanics, such as nanoscience, Micro-
Electro-Mechanical systems (MEMS) science, materials science, and
interface chemistry.
The most basic principles of capillary mechanics are simple
and straightforward. The most fundamental concept of capillary
mechanics is surface tension, or surface free energy. The explanation
of capillary phenomena is often based on surface tension or surface
free energy.
Introduction xv

The laws of capillary mechanics are extremely complex and


profound. Due to the fact that molecular-level forces determine sur-
face tension or surface free energy properties and the complexity of
the interaction patterns between a pair of molecules, it is necessary
to explore the interactions between a group of molecules, whether
from the perspective of non-vector energy or vector force, which are
extremely complex. Not only is the physical process extremely com-
plex, but the mathematical form is also extremely complex. Even
from a phenomenological perspective, the laws governing the action
of a general form of matter are complex.
The capillary phenomenon is very common and also very usural.
Insert a thin glass tube into a liquid, such as water, and you will
see the liquid rise inside the glass tube. By dipping a straw in a soap
solution, you can blow out colorful bubbles, and children can conduct
such experiments.
The capillary phenomenon, however, contains extremely rich
scientific principles. From the burning of wick oil, the absorption
of water by sponge, to the growth of crops and plants, soil and
water conservation, artificial rainfall in astronomy and meteorology,
all involve basic capillary action.
The capillary action is very microscopic. Its mechanism of action
belongs to the action of intermolecular forces. And the molecules are
only in the size range of 0.1 nm, and the molecular spacing is also
only in the range of 0.1 nm. Therefore, the origin of their effects
belongs to microscopic processes.
The capillary action is also very macroscopic. The capillary
phenomenon caused by the microscopic effects of capillaries is actu-
ally macroscopic. Some may have a slow speed, but the cumulative
effect is visible to the naked eye at a macroscopic level, while others
exhibit capillary phenomena, and their macroscopic speed is quite
fast. When a nurse pricks the finger of an individual to take a blood
sample, it can clearly be seen that the blood is moving at a faster
speed in a thin test tube with the naked eye.
The capillary action is very small. Liquids can only exhibit
capillary rise in very thin tubes. And generally, its upward speed
is also very slow.
The capillary action is also grand. The so-called high mountains
and deep waters depend on the action of capillary force. The Earth
xvi Capillary Mechanics

nourishes all things and grows through the action of capillaries, while
the vapor of mountain clouds and the transformation of clouds into
rain also rely on the action of capillaries.
The capillary phenomenon is often utilized by people in their
daily lives. The tissues that wipe sweat, the corduroy sweatshirts
that absorb sweat, and the sponges that wipe the floor, all of them
utilize capillary action.
The capillary phenomenon has also attracted more attention from
ancient and modern scientists. Leonardo da Vinci of the Renais-
sance period focused on capillary phenomena, Sir Isaac Newton
studied capillary phenomena, and Thomas Young and Pierre Simon
Laplace further promoted the development of capillary mechanics. In
addition, C.F. Gauss, Johann Bernoulli, and Simeon Denis Poisson
have also made many contributions to the development of capillary
mechanics. James Thomson, Marangoni, and others further devel-
oped capillary mechanics. Lord Kelvin and others were a ground-
breaking driving force in the development of capillary mechanics.
Moreover, even the first paper by the greatest modern physicist,
Albert Einstein, explored physical problems by studying and uti-
lizing capillary phenomena.
There are many reasons why capillary phenomena and their effects
have attracted the attention of scientists from ancient and modern
times, both at home and abroad. The above characteristics are also
part of the reasons, but four of them are the most fundamental. One
reason is that it is the link and bridge between physical macro and
micro interactions. The capillary phenomenon is a unity of micro-
scopic effects and macroscopic characterization. Due to the limited
level of technology in the previous centuries, people were unable to
directly observe the underlying phenomena, especially the molecular-
level effects that could not be measured from a physical perspective.
However, due to the macroscopic phenomenon of capillaries, it can be
observed with the naked eye. Therefore, through macroscopic obser-
vation and measurement of capillary phenomena, one can indirectly
understand the laws of the microscopic material world. The second
reason is that it simultaneously involves different forms of matter,
whether from the perspective of “phase” or “state”. The gas, liq-
uid, and solid in the material world are all involved simultaneously,
making it both typical and representative, as well as reflective of
the generality of the world’s materials, especially the composition
Introduction xvii

of objects. The third reason is that the core of capillary action is


the effect of interfaces. Compared to the ontology, the interface has
both step and real variability, as well as transitional and continu-
ous properties. Therefore, its mechanism of action is relatively com-
plex. There are both stable and unstable trends, with both physical
and chemical effects. Therefore, the study of capillary phenomena
directly gave birth to modern interface science and surface science,
which are the foundations of science and technology, such as connec-
tion, lubrication, friction, cleaning, wear, adhesion, and adsorption.
The fourth reason is that capillary phenomenon is widely present in
nature and the actual production and life of people. And its under-
lying principles are complex and profound. Universality determines
people’s attention, while complexity and depth attract the attention
of scientists.
From the above, it can be seen that capillary phenomena have
always been familiar to people, and capillary action has also been
the focus of scientists. However, unfortunately, capillary mechanics
has not yet formed an independent discipline. In modern times, they
are only a branch of thermodynamics and fluid mechanics and a part
of surface science.
However, in the 21st century, micro/nano technology has
emerged and attracted global attention. Major countries around the
world regard micro/nano technology as the most likely scientific
and engineering field to achieve breakthroughs in the future.
The rise of micro/nano technology has unprecedentedly increased
human interest, cognitive level, and desire to manipulate the
world of micro/nano technology. The continuous development of
micro/nano-electro-mechanical technology and the continuous minia-
turization of micro/nano-mechanical structures have prompted
people to constantly explore deep-seated problems, and capillary
mechanics has thus been revitalized.
With improvements in manufacturing technology, the scale of
research objects continues to shrink, and the ratio of surface area
to volume of materials or gaps increases relatively. The importance
of surface force (or interfacial force) relative to volume force is greatly
enhanced, resulting in a significant increase in the importance of cap-
illary action. The importance of capillary action at the micrometer
scale is an undeniable fact, and the importance at the nanoscale is
certainly self-evident.
xviii Capillary Mechanics

At the micro/nanoscale, capillary action is a double-edged sword.


On the one hand, it has great usability, for example, people can
use capillary action to perform “capillary casting”, make “molecular
straws”, and “nano test tubes” and also perform self-assembly of
micro-devices. Capillary action has been successfully used to guide
and control fluid flow in microfluidic actuators, and people naturally
expect it to play a stronger role in nanoscale fluid manipulators. On
the other hand, at the micro/nanoscale, capillary action can bring
many unfavorable challenges that must be addressed. The capillary
adhesion effect can lead to “adhesion” or even “failure” between the
active parts of MEMS and even more severe “adhesion” or “failure”
between the active parts of nano-mechanical systems (NEMS).
It is during the continuous exploration process that people have
found that capillary mechanics has a particularly prominent impact
on the performance of micro/nano-mechanical structures, so research
on micro/nanoscale capillary interactions has received great atten-
tion. However, people’s understanding of capillary mechanics is still
limited, and there is a lack of sufficient understanding of many impor-
tant issues. Many deep-seated problems have not been well solved. In
order to better promote the study of capillary action, it is necessary
for capillary mechanics to become a relatively independent science
and to combine with the development of micro-nano technology and
science, forming a modern research field that is full of new vitality.
Chapter 1

Capillary and Capillary Action

Capillarity is a prevalent physical phenomenon observed in nature.


The ancient Chinese adage “When the clouds rise, the plinth turns
wet. When the plinth becomes wet, rain will soon follow” illustrates
the capillary phenomenon in the natural world. Most plinths are
composed of porous materials. When the humidity in the air is
elevated, it contains significant amounts of moisture. The process of
capillary condensation can lead to the saturation of plinths, serving
as an indicator that rainfall is imminent.

1.1 The Phenomenon and the Function of Capillary

Initially, we will examine two intriguing phenomena to enhance our


understanding of capillarity.
One can place a drop of mercury on a clean glass surface. The
drop will roll off without adhering to the glass. Similarly, if a clean
glass pane is immersed in mercury and subsequently removed, the
mercury will not cling to the surface of the glass. This phenomenon,
characterized by the inability of a liquid to adhere to solid surfaces, is
referred to as non-wetting. Mercury exemplifies a non-wetting liquid
in relation to glass.
When a drop of water is placed on a clean pane of glass, it adheres
to the surface and forms a thin layer. Similarly, when a clean pane of
glass is immersed in water, the surface becomes coated with a layer
of water. This phenomenon, in which a liquid adheres to the surface

1
2 Capillary Mechanics

(a) (b)

Figure 1.1. The capillary tubes dip into liquids. (a) Glass tube immersed into
water and (b) glass tube immersed into mercury.

of a solid, is referred to as wetting. Water is classified as a wetting


liquid in relation to glass.
The ability of a liquid to wet a particular type of solid may not
extend to other types of solids. For instance, water is capable of
wetting glass, whereas it does not wet paraffin. Mercury is unable to
wet glass but is capable of wetting zinc.
When a superfine glass tube is immersed in water, the water level
within the tube rises. The smaller the inner diameter of the tube,
the higher the water ascends (see Figure 1.1(a), where “s” denotes
solid, “l” denotes liquid, and “v ” denotes vapor). Conversely, when
these tubes are submerged in mercury, the mercury level within the
tube decreases. Again, the smaller the inner diameter of the tube, the
lower the mercury level drops (see Figure 1.1(b)). This phenomenon,
characterized by the ascent of a wetting liquid and the descent of a
non-wetting liquid within a superfine tube, is referred to as the “cap-
illary phenomenon”. Tubes that exhibit a pronounced capillary phe-
nomenon are known as capillary tubes. Examples of materials that
contain capillary tubes include paper, lamp wicks, gauze, soil, and
the roots and stems of plants. The liquid surface within the tube can
present either a concave or convex meniscus. When the liquid wets
the solid material of the capillary tube, the liquid surface rises and
forms a concave shape. Conversely, when the liquid does not wet the
solid material, the liquid surface drops and takes on a convex shape.
The capillary phenomenon is prevalent in both industrial
production and everyday life, with numerous illustrative examples.
Capillary and Capillary Action 3

Calcareous tufa, which is characterized by the presence of numerous


capillary tubes, can be utilized as an artificial hill, akin to a bon-
sai, facilitating moisture acquisition for the plants situated on this
artificial structure. This moisture ascent is fundamentally dependent
on the principles of capillarity. In plants, the vascular structures
within the roots and stems consist of extremely fine capillary tubes.
The capacity of plants to absorb moisture and nutrients from the
soil through their roots and stems is significantly attributed to the
capillary phenomenon and the functionality of these capillary tubes.
Additionally, the lubricating oil in machinery permeates the intersti-
tial spaces of machine components, providing lubrication through the
action of capillary tubes. Furthermore, various porous solid materi-
als, such as bricks, paper, textiles, and chalk, exhibit the ability to
absorb water due to the abundance of minute pores within these
materials, which function as capillary tubes.
The capillary effect plays a significant role in water conser-
vancy, facilitating the movement of groundwater from areas of higher
moisture to those of lower moisture. This phenomenon serves as a
vital source of water for plants. Additionally, capillary action enables
facial tissues to absorb liquids when used to wipe surfaces. Each
pore in a sponge functions similarly to a capillary tube, allowing
the sponge to absorb a substantial volume of liquid. Furthermore,
traditional corduroy sportswear effectively draws sweat away from
athletes’ skin by employing the principle of capillary action, akin to
that of a lamp wick.
In certain circumstances, capillary action can be detrimental.
For instance, the presence of excessively thin capillary tubes in the
foundation of buildings can lead to the upward movement of soil
moisture, resulting in increased indoor humidity. To mitigate this
issue, asphalt felt is often applied to the base as a preventive mea-
sure against moisture intrusion caused by capillary action. Addition-
ally, the phenomenon of water rising through capillary tubes can
significantly adversely affect agricultural production. The soil con-
tains numerous capillary tubes that facilitate the upward movement
of moisture from below the surface. To conserve subsurface moisture,
it is essential to aerate the soil and disrupt the capillary tubes at the
surface, thereby reducing moisture evaporation.
Every phenomenon possesses dual aspects. A collaborative effort
involving the University of Paris, the University of ESPCI, and
4 Capillary Mechanics

École Polytechnique seeks to transform this detrimental effect into a


beneficial application. This research group aims to manipulate micro-
plates to form three-dimensional shapes utilizing capillary action.
The process unfolds as follows: thin silicone is sculpted into various
configurations, such as flowers, triangles, or squares, after which a
droplet of water is applied to the surface. Due to the water’s ten-
dency to minimize its contact area with the surrounding air, the
droplet promptly begins to conform to the contours of the thin
plate. Following the evaporation of the water at room tempera-
ture, the curved structure gradually solidifies, ultimately achieving
its intended shape. Notably, this entire process occurs spontaneously,
requiring no further intervention after the initial application of the
droplet.
In summary, capillarity is not only prevalent in nature and daily
life, but it also demonstrates both beneficial and detrimental effects.
Understanding how to harness this phenomenon to maximize its
advantageous properties while mitigating its adverse effects is crucial
for the development of Micro-Electro-Mechanical Systems (MEMS).

1.2 The Capillary Action of M/NEMS

The aforementioned capillary phenomenon is observable at a


macroscopic level. While certain microscopic processes associated
with capillary action may not be directly observed, the resultant
macro effects can still be quantified. It is evident that the thinner the
capillary tube, the more pronounced the capillary action becomes.
Similarly, an increase in the number of capillary tubes correlates
with a heightened significance of capillary action. With advancements
in Micro/Nano-Electro-Mechanical Systems (M/NEMS) technology,
the dimensions of transformable microstructures or mobilizable
components have reached the micron and even nanometer scales.
In contrast to traditional macro functions generated by numer-
ous capillary tubes, the individual capillary action is markedly
evident in M/NEMS applications. For instance, during the fabri-
cation of surface micro-mechanical structures, enhanced capillary
interactions often cause micro-bridges and micro-trabecular struc-
tures to adhere to the substrate, thereby compromising their effi-
cacy. Additionally, in experiments assessing material constants at
Capillary and Capillary Action 5

the micro-scale, micro-bridges and micro-beams are frequently uti-


lized as specimen structures for performance testing. The presence
of capillary adhesion between the loading apparatus and the test-
ing microscale specimen can significantly affect the accuracy of the
obtained data. In micro-surface processing, when the sacrificial layer
is completely etched, deionized water is required to remove etchants
and residual materials. However, this process can easily lead to the
formation of a liquid bridge interface between two parallel surfaces
upon removal from the deionized water. Such liquid bridges gener-
ate substantial capillary forces. Notably, even in the absence of a
liquid bridge during manufacturing, humidity can facilitate the for-
mation of liquid bridges within micro-structures. The emergence of
liquid bridges is a direct consequence of capillary action, which in
turn generates relevant liquid bridge forces. For MEMS, inadequate
packaging can lead to the formation of capillary condensate at rela-
tive humidity levels of 65%. Under the influence of capillary forces, if
the restoring force of the structure is insufficient, microscale bridges
may adhere to the substrate. Therefore, it is imperative to compre-
hend the emergence and function of capillary forces and to effectively
manage and utilize these forces in practical applications.
The advancement of machining precision has progressed from
micron dimensions to the nanometer scale. Currently, the precision
of ultra-precision diffraction gratings has achieved a remarkable
accuracy of 1 nm, facilitated by diamond turning techniques. Fur-
thermore, laboratories have successfully fabricated structures such
as wires, columns, and troughs measuring less than 10 nm. With the
rapid evolution of M/NEMS process technology, M/NEMS devices
have gained widespread application. The capillary phenomenon is
present across all scales; however, its significance is particularly pro-
nounced in M/NEMS. The presence of capillarity plays a crucial
role in both the production and operational processes of these
devices.
Most designs of M/NEMS typically incorporate several
fundamental components, including micro- and nanoscale springs,
cantilever beams, films, hinges, and gears. During the manufacturing
processes of these standard structures, the final step at the wafer
level involves the release of all functional components. This process
commonly entails the etching of a sacrificial layer. Once the sacrifi-
cial layer is removed, the sections composed of permanent materials
6 Capillary Mechanics

are allowed to move freely in accordance with the intended design.


A critical aspect of the release process is the thorough cleaning of the
structures to eliminate etchants, dissolved silica (or other sacrificial
materials), microscale solid residues, and any potential remaining
by-products of the sacrificial layer. However, a significant challenge
arises during this process: the flexible sections may experience cap-
illary adhesion due to capillary forces. When the force of capillary
adhesion exceeds the mechanical restoring force of the microstruc-
ture, adherence occurs. As liquid comes into contact with adjacent
surfaces and evaporates slowly, the capillary force generated by the
liquid detergent can cause several sections to adhere to one another.
This phenomenon can be attributed to surface tension. In the context
of microfabrication technology, once the sacrificial layer is removed
and the structure is released, the structural layer is prone to adhere
to the substrate or neighboring structures. This occurrence is referred
to as release adhesion.
The presence of moisture is generally detrimental to electronic
devices, particularly to M/NEMS. During operation, if M/NEMS
devices are inadequately encapsulated, capillary condensation can
lead to the formation of liquid droplets due to the hydrophilic nature
of the materials used in these devices under conditions of high humid-
ity. Such droplets can create liquid bridges within structures, such as
micro-resonators, micro-accelerometers, and micro-gyroscopes, which
in turn generate resistance or adhesive forces that impede the relative
motion of the structures, ultimately resulting in device failure. This
phenomenon is referred to as working adhesion. Recent research con-
ducted by Sandia National Laboratories has indicated that, once the
operational conditions meet the strength requirements of M/NEMS
materials, the primary cause of device failure is the capillary adhesion
of micro/nano-mechanical structures. For instance, if a liquid bridge
due to capillary adhesion forms between the electrodes of a micro-
switch, the contact angle of the meniscus and the surface roughness
of the substrate will significantly affect the pull-in voltage. In experi-
ments involving micro- and nanoscale structures, micro/nano bridges
and micro/nano girders are commonly utilized as test specimens for
assessing material constants and functions at the microscale. The
presence of capillary adhesion between the microscale testing end
and the loading end can directly impact the accuracy of the test
data and may even lead to complete failure of the test specimen.
Capillary and Capillary Action 7

The significance of capillarity in M/NEMS is influenced by several


factors, including the diminutive size, relatively large surface area,
and the close proximity of adjacent surfaces in micro- and nanoscale
devices. The volume forces associated with mass and quality, such as
gravitational and magnetic forces, diminish rapidly with the cube of
the dimension, whereas surface forces related to surface area, includ-
ing frictional, adhesive, and van der Waals forces, decrease with
the square of the dimension. Consequently, in micro-devices, volume
forces become negligible, while surface forces emerge as predominant
factors. As a result, with the miniaturization of components, the
impact of volume forces, such as gravity, can often be disregarded;
however, the surface forces arising from contact or friction between
surfaces assume a critical role. The adhesion among structural ele-
ments significantly affects the functionality and reliability of micro
and nano machines. Given the importance of capillary adhesion in
the architecture of micro- and nanoscale devices, research into the
mechanical mechanisms of capillarity at the micro and nano scale
has proliferated in recent years.
Capillary action frequently manifests as a liquid bridge in
M/NEMS that involve mechanical and electrical components. This
phenomenon occurs when two solid surfaces are brought sufficiently
close together, allowing a small volume of liquid to be absorbed
between them. The liquid bridge phenomenon is prevalent in vari-
ous contexts and induces oligodynamic interactions between the solid
surfaces and the liquid. To separate the two solid surfaces, a force
must be exerted that exceeds the capillary force exerted by the liquid
bridge. This force is commonly referred to as the liquid bridge force
or capillary force.
In micro- and nanoscale systems, the phenomenon of liquid
bridging is observed among the structures of micro and nano devices.
For instance, a liquid bridge can form between a cantilever beam
and its substrate under conditions of high relative humidity, as seen
in miniature accelerators (or RF MEMS). Similarly, small droplets
of water can create a liquid bridge between the comb tines of a
micro-gyroscope and its base. This liquid bridge induces a cap-
illary adhesion effect between the solid surfaces and the liquid.
At the micro and nano scales, surface forces become significantly
more influential than volumetric forces. The capillary adhesion
effect can adversely affect the normal operation of RF MEMS,
8 Capillary Mechanics

micro-transducers, micro-gyroscopes, and other devices character-


ized by micro- and nanostructures. The liquid present in the liquid
bridge may either be pre-existing or may form gradually from nearly
saturated condensable vapor. Consequently, in microscopic systems,
particularly those with nanostructures, the presence of original liquid
increases the likelihood of liquid bridge formation under conditions
of elevated moisture. To mitigate the formation of liquid bridges in
microscopic or nanoscale systems, it is essential to incorporate ade-
quate vacuum desiccants during the packaging process.
The liquid bridge is a prevalent phenomenon in micro- and
nanoscale systems, particularly in measurement techniques and
equipment designed for micron- and nanosized applications, where
the liquid bridge exhibits significant effects. Currently, measurement
techniques for geometric characteristics and surface topography with
micro- and nanoscale accuracy have been developed. For instance,
advancements in Scanning Tunneling Microscopy (STM), Scanning
Probe Microscopy (SPM), and Scanning Force Microscopy (SFM)
have enabled cutting-edge operations, fabrication, and reshaping at
the atomic level in recent years.
The operational principles of measuring devices such as STM,
SPM, and SFM indicate that these instruments are susceptible to
the effects of capillary action. The SPM is fundamentally based on
the tunneling effect as described by quantum mechanics. When the
atomic-scale tip of the probe scans samples at a height of less than
1 nm, the electron clouds of the tip and the sample overlap. By apply-
ing a voltage ranging from 2 mV to 2 V, the tunneling effect facilitates
the escape of electrons from the sample, resulting in the formation of
a tunneling current. The intensity of this current exhibits a functional
relationship with the distance between the probe tip and the sam-
ple surface. Due to the uneven nature of the atomic structure at the
sample’s surface, as the probe tip traverses the material at a specified
height, the distance between the tip and the surface varies continu-
ously, leading to corresponding fluctuations in the tunneling current.
The graphical representation of these current variations can reveal
atomic-level surface patterns. In addition to the STM, a variety of
SPM techniques have been developed that exploit different interac-
tions between the probe tip and the samples to investigate surfaces
or interfaces at the nanoscale. These advanced probe microscopes
include Atomic Force Microscopy (AFM), Laser Force Microscopy
(LFM), and Magnetic Force Microscopy (MFM), all of which are
Capillary and Capillary Action 9

derived from the principles of STM. The AFM, for instance, operates
by having a microprobe glide across the sample surface, causing a
highly sensitive microcantilever beam to oscillate in response to the
surface topography. The AFM detects the displacement of the micro-
cantilever beam using optical methods or tunneling current, thereby
measuring the repulsive forces between the tip’s atomic point and the
surface atoms. Given the operational principles of probe microscopy,
there exists a significant likelihood of liquid bridges forming between
the microscope tip and the surface of the object being measured. If
small droplets of liquid are present between the microscope tip and
the sample, a liquid bridge may be established. Additionally, films of
water or other liquids can also create liquid bridges. The presence of
these liquid bridges can adversely affect the measurement accuracy
of these microscopy techniques.
Currently, capillary action presents a significant challenge in the
research and development of M/NEMS devices that has yet to be
addressed. Effectively controlling capillary action during the manu-
facturing of micro and nano devices has emerged as an issue that
warrants considerable attention.

1.3 The Development of Capillary Mechanics

The term “capillary” comes from the well-known concept “blood


capillaries”. Ibn al-Nafis (1213–1288), a prominent Arab scientist
and physiologist, was the first to propose the idea that nutrients
are obtained from the pulmonary artery through capillary circula-
tion in the 13th century. However, at that time, his understanding
of the mechanisms governing blood flow within capillaries was lim-
ited. It was not until the years 1845–1846 that German physicist
Gotthilf Heinrich Ludwig Hagen (1797–1884) and French physiolo-
gist Jean Louis Marie Poiseuille (1797–1869) conducted research that
established the principles governing blood flow in capillaries, build-
ing upon the general understanding of capillary action. Although
the concept of capillarity originates from blood capillaries, the phe-
nomenon of capillary action encompasses a variety of occurrences in
nature. In fact, capillary phenomena and their effects are prevalent
in both the natural world and everyday life.
Although the concept of capillarity was introduced in the early
13th century, it was not until the 18th century that systematic
studies of capillarity began to emerge, culminating in a more
10 Capillary Mechanics

comprehensive understanding by the late 19th century. During this


period, several prominent scientists made significant contributions to
the advancement of capillary theory. Their innovative work propelled
the development of both physics and capillary mechanics. Conse-
quently, several key theories and concepts in capillary mechanics are
named in honor of these scientists, serving as a tribute to their pio-
neering contributions. Notable examples include the Young–Laplace
equation, Gibbs phase rule, Kelvin equation, Helmholtz free energy,
and Marangoni effect.
According to the German physicist Johann Christian Poggen-
dorff (1796–1877), the capillary phenomenon was first discovered by
Leonardo da Vinci (1452–1519), who is recognized as one of the
prominent figures of the Italian Renaissance. Leonardo da Vinci
exhibited a broad interest in research and investigated numerous
significant phenomena in nature. However, the earliest documented
discovery of the capillary phenomenon is attributed to the English
scientist Francis Hawksbee (1666–1713). In 1709, Hawksbee was the
first to observe the capillary phenomenon in tubular structures and
on glass plates, noting the attraction between the glass surface and
the liquid. He found that the capillary phenomenon occurs in both
thick and thin glass tubes and concluded that its influence is present
only when the liquid comes into contact with the internal surface of
the glass tube.
In 1718, the English scientist James Jurin (1684–1750)
demonstrated that the height to which a liquid rises is solely a func-
tion of the liquid’s surface area. Additionally, he established that the
height of liquid in capillary tubes is inversely proportional to the
radius of those tubes. Jurin’s significant contributions to the under-
standing of capillary action garnered him widespread recognition in
the scientific community.
The English mathematician, physicist, and astronomer Sir Isaac
Newton posited that the phenomenon of capillarity is influenced by
the viscosity of liquids during his investigations into this property.
Nevertheless, he did not provide a comprehensive explanation of the
characteristics of capillary action in liquids.
In 1751, the German scientist Johann Segner (1704–1777)
first introduced the significant concept of surface tension in liq-
uids, attributing this phenomenon to gravitational interactions. He
endeavored to demonstrate that surface tension influences the shape
of liquid droplets. While Segner acknowledged the effect of radial
Capillary and Capillary Action 11

curvature on liquid droplets, he overlooked the impact of horizontal


curvature. His contributions to the understanding of surface tension
enhanced the comprehension of capillary action.
In 1756, the German physician Johann Gottlob Leidenfrost (1715–
1794) noted that soap bubbles exhibited a tendency to shrink. He
observed that when the inlet side of the tubule connecting to the
soap bubble was opened, the bubble would continuously contract in
an effort to expel the air. However, Leidenfrost did not ascribe this
phenomenon of shrinking to the effects of the liquid surface.
In 1787, the French mathematician and chemist Gaspard Monge
(1746–1818) applied the principle of surface tension to elucidate the
evident attractive and repulsive forces that exist between liquids and
the objects that float upon their surfaces.
In 1802, the Scottish mathematician and physicist John Leslie
(1766–1832) provided the first accurate explanation of the rise of
liquid in capillary tubes due to capillary action. He did not assume
that the direction of the attractive force between the tube and the
liquid was upward, which would facilitate the movement of the liquid.
Instead, Leslie emphasized that the primary effect of the attrac-
tive force was to increase the pressure of the fluid layer in contact
with the tube, which was significantly greater than the internal pres-
sure of the liquid. This phenomenon can be intuitively understood
as causing the fluid contact layer to spread across the surface of the
solid. Consequently, when a droplet of liquid is placed on a clean, flat
glass surface, it tends to spread out. This explains why a glass tube
becomes wet when immersed in water. The rise of the liquid contin-
ues until the weight of the water column is balanced by the surface
tension forces, at which point the liquid ceases to rise. Leslie’s the-
ory was subsequently validated through mathematical methods by
the Scottish mathematician James Ivory (1765–1842).
Thomas Young (1773–1829), an eminent English physicist,
established the theory of capillarity, which is grounded in the concept
of surface tension, in 1804. Young recognized that the contact angle
between a liquid’s surface and a solid substrate remains constant,
and he elucidated the phenomenon of capillary action based on two
foundational principles. In 1805, he was the first to propose the
equations governing solid surface tension, liquid surface tension, and
solid–liquid surface tension, collectively known as Young’s equation.
This equation serves as a fundamental formula in the analysis of the
wetting phenomenon within the framework of surface phenomena,
12 Capillary Mechanics

and it is often referred to as the “wetting equation”. Thomas Young


was a polymath in Britain and a key figure in the development of
the wave theory of light. His research spanned multiple disciplines,
including capillary action, light wave theory, sound wave theory,
hydrokinetics, shipbuilding engineering, tidal theory, the measure-
ment of gravity using pendulums, and the theory of rainbows.
Thomas Young and the French mathematician Pierre-Simon
Laplace (1749–1816) independently derived the relationship between
the additional pressure exerted beneath the curved surface of a liq-
uid and the radius of curvature, a principle known as the Young–
Laplace Equation, or simply the Laplace equation. Pierre Simon
Laplace was a distinguished mathematician, astronomer, and mem-
ber of the French Academy of Sciences. He was elected as a member
of the Collège de France in 1816 and subsequently served as its pres-
ident in 1817. Laplace was a principal figure in the development of
celestial mechanics, a co-founder of cosmogony, and a pioneer in the
field of probability theory. Although the research findings of Laplace
in capillary mechanics were remarkably similar to those of Thomas
Young in several respects, their methodologies were fundamentally
different. Laplace arrived at significant conclusions primarily through
mathematical calculations.
The research on capillarity progressed to a more significant phase
during the early 19th century. Carl Friedrich Gauss (1777–1855),
a prominent German scientist, introduced the method of energy
analysis in 1830. Gauss, renowned as a mathematician, physicist,
astronomer, and expert in geodesy, has been posthumously enshrined
in academic discourse as “the Prince of Mathematics” — an hon-
orific epithet denoting his unparalleled eminence in mathematical
sciences. He derived various equations related to capillary action and
established boundary conditions based on the principles of energy
conservation and virtual work, the latter of which was articulated
by Johann Bernoulli (1667–1748), a Swiss mathematician. In his for-
mulation, Gauss employed a concept now recognized as potential
energy. He concluded that potential energy comprises three compo-
nents: gravitational potential energy, interaction energy among fluid
particles, and the potential energy associated with the interaction
between the fluid and the solid contact surface. Furthermore, Gauss
identified deficiencies in the work of Pierre-Simon Laplace.
Capillary and Capillary Action 13

In 1831, Simeon Denis Poisson (1781–1840), a prominent French


mathematician, published a monograph addressing the phenomenon
of capillary action. He posited that the density of a liquid near its
surface undergoes rapid changes. This assertion was subsequently
validated through a series of experimental investigations. Poisson
conducted research on fluid equilibrium based on the hypothe-
sis of uniform density, ultimately concluding that the capillarity
observed under this assumption was inconsistent with theoreti-
cal predictions. Consequently, Poisson demonstrated that Laplace’s
hypothesis regarding uniform density was flawed. However, it is
important to note that Laplace’s hypothesis, which suggested that
liquid density was uniform and that attractive forces existed among
liquid molecules, proved effective within a limited scope. Further-
more, Gauss later corroborated the conclusions derived from this
hypothesis through alternative methodologies. As a result, Poisson’s
conclusions were ultimately deemed incorrect. Nevertheless, Pois-
son’s work laid the groundwork for further investigation into the
varying densities of fluids at the surface compared to those within
the bulk. In essence, Poisson’s study was fundamentally equivalent
to Laplace’s conclusions. Both fluid surface equations necessitate the
introduction of a constant, denoted as H, which must be determined
experimentally. The primary distinction lies in the fact that the con-
stant H is contingent upon either the principles of molecular forces
or the principles of surface density of the fluid. Currently, it remains
challenging to differentiate between these two principles.
James Thomson (1822–1892), the elder brother of Lord Kelvin
and an English physicist, first observed the phenomenon known as
the tears of wine in 1855. The Italian physicist Carlo Marangoni
(1840–1925) completed a doctoral dissertation on this phenomenon
in 1865, leading to its designation as the Marangoni effect. Much of
the subsequent research on this topic was conducted by the American
physicist Josiah Willard Gibbs (1839–1903), which has resulted in the
phenomenon also being referred to as the Gibbs–Marangoni effect.
The tears of wine in wineglass are related to the presence of a surface
tension gradient, which induces liquid flow toward regions of higher
surface tension.
Inspired by Gauss’s energy theory, numerous subsequent studies
have been conducted that analyze phenomena not only from the
14 Capillary Mechanics

perspective of force balance but also from the standpoint of sys-


tematic energy conservation. To date, there exist two definitions
of surface tension: one derived from energy considerations and the
other from force considerations, both of which are fundamentally
equivalent. Each of these approaches has its respective advantages
and disadvantages. The force-based method is intuitive and straight-
forward; however, it is limited by the vector nature of force, which
introduces directional considerations that complicate the analysis
of complex systems. Conversely, the energy-based method circum-
vents the issues associated with vector directionality and allows for
arithmetic summation. Nonetheless, this method is less effective for
analyzing interactions and may not be as intuitively comprehensi-
ble. Thermodynamics is fundamentally a phenomenological disci-
pline grounded in energetics. Given that capillary phenomena are
typically regarded as a subset of thermodynamics, the analytical
framework of energetics has significantly influenced advancements
in capillary mechanics. Notably, Josiah Willard Gibbs and Hermann
von Helmholtz are two of the most prominent and influential figures
in the field of thermodynamic energetics.
Gibbs established a law in 1876 that describes the relationship
among the number of phases, the number of components, and the
degrees of freedom in an equilibrium system. This principle, known
as the phase rule, articulates a universal law governing phase equi-
librium. Gibbs free energy, also referred to as the Gibbs function, is
a crucial parameter in thermodynamics, commonly denoted by the
symbol G. It is defined as the maximum work obtainable from a ther-
modynamic system during isothermal and isobaric processes, exclud-
ing work associated with volumetric changes. In other words, during
isothermal and isobaric processes, the work performed by the system
on its surroundings, apart from that due to volumetric changes, can
be equal to or less than the decrease in Gibbs free energy. Gibbs
made significant contributions to both thermodynamics and statisti-
cal physics, and his work has led to the establishment of a rigorous
and comprehensive theoretical framework in thermodynamics.
Hermann von Helmholtz (1821–1894), a prominent German
physicist and physiologist, published an influential article titled
Thermodynamics in Chemical Processes in 1882. In this work, he
distinguished between binding energy and free energy in chemical
reactions, noting that binding energy can only be converted into
Capillary and Capillary Action 15

heat, whereas free energy can be transformed into various other forms
of energy. Helmholtz derived what came to be known as the Gibbs–
Helmholtz equation from the Clausius equation. The thermodynamic
function he introduced, known as the Helmholtz free energy, serves
as a valuable tool for assessing the direction and limitations of physi-
cal and chemical processes under conditions of constant temperature
and constant volume. In 1847, he delivered a notable lecture on the
conservation of force to the German Physical Society, which signifi-
cantly enhanced his reputation within the scientific community. Dur-
ing this presentation, Helmholtz was the first to articulate the law of
energy conservation using a mathematical framework. Throughout
his career, Helmholtz conducted extensive research and made sub-
stantial contributions not only to physics but also to physiological
optics, acoustics, mathematics, and philosophy.
Lord Kelvin (1824–1907), a prominent English physicist and
inventor, made significant contributions to the study of capillary
phenomena and their implications. From a young age, Kelvin exhib-
ited remarkable intelligence and a strong dedication to his studies,
entering the preparatory school of the University of Glasgow at the
age of 10. In 1851, he was elected as a member of the Royal Society of
London and subsequently served as its president from 1890 to 1895.
Additionally, he was elected as a fellow of the French Academy of
Sciences in 1877. Kelvin conducted extensive research and made sig-
nificant contributions to various fields, including thermology, electro-
magnetics, fluid mechanics, optics, physical geography, mathematics,
and engineering applications. During his time, he garnered a distin-
guished reputation within the scientific community and received high
praise from scientists and academic institutions in Great Britain and
other Western countries. His research in thermology and electromag-
netics, along with their engineering applications, was particularly
noteworthy. In the area of capillarity, he analyzed the differing satu-
ration vapor pressures between liquids and droplets on flat surfaces,
leading to the derivation of the Kelvin equation through the appli-
cation of energy principles. The Kelvin equation serves to elucidate
phenomena such as capillary condensation, artificial rainfall, and the
radius of the meniscus.
Albert Einstein (1879–1955), the American physicist, is widely
regarded as the pioneer and founding figure of modern physics. He
is best known for his formulation of the theory of relativity and the
16 Capillary Mechanics

mass–energy equivalence principle. Notably, Einstein’s initial the-


sis focused on capillary mechanics. He completed his first thesis,
titled Obtained from the Capillary, in December 1900, which was
subsequently published in the Physics Journal Letters the following
year. In this work, Einstein derived a law of action pertaining to
intermolecular forces as they relate to capillarity.
Despite the extensive historical development of capillary
mechanics and its theoretical frameworks that inform practical
applications, it has yet to establish itself as an independent discipline.
Currently, it remains a subset or branch of thermodynamics, fluid
mechanics, or surface science. However, advancements in M/NEMS
technology, improvements in micro-nano material science, and the
ongoing miniaturization of micro- and nano-mechanical structures
have broadened the scope of applications, prompting deeper explo-
ration of fundamental issues. Through this exploration, researchers
have identified the significant influence of capillary mechanics on
micro- and nano-mechanical structures. Consequently, as a critical
surface force, capillary mechanics has garnered increased atten-
tion. Nevertheless, a comprehensive understanding of many essential
questions within capillary mechanics remains elusive, and numer-
ous unresolved issues persist. Therefore, it is imperative for capillary
mechanics to evolve into an independent discipline, integrating with
the advancements in micro-nano technology and science, to establish
a modern research field that will enhance the study of capillarity.
Chapter 2

Basic Conceptions

2.1 State of Matter and Phase

2.1.1 State of matter


The diverse substances present in the natural world are composed
of numerous microscopic particles, including molecules, ions, and
atoms. These particles engage in random thermal motion while
simultaneously interacting with one another. The attractive forces
among the particles promote aggregation, whereas thermal motion
tends to facilitate their dispersion. At specific temperature and
pressure conditions, these opposing effects reach a state of equi-
librium, resulting in the clustering of particles into a stable struc-
tural configuration, commonly referred to as a state of matter. From
a macroscopic perspective, the most prevalent states of matter on
Earth are solid, liquid, and gas. In the solid state, matter retains
a fixed volume and shape. In the liquid state, matter maintains a
constant volume but conforms to the shape of its container. In the
gaseous state, matter expands to fill any available volume.
On a microscopic level, solid materials are composed of particles
that are arranged in a regular and symmetrical order, a configuration
referred to as the crystalline state. The particles that constitute these
crystalline structures are interconnected by various types of chemical
bonds, including covalent, ionic, and metallic bonds. Consequently,
there exists a strong interaction force among the particles. These
bonded particles exhibit random vibrations around their equilibrium
positions but are unable to move away from them. As a result, solids

17
18 Capillary Mechanics

possess a stable shape and a definite volume. In contrast, while


liquids also exhibit strong interaction forces among their constituent
particles and their molecular arrangement is often similar to that of
solids, this arrangement is typically confined to very small regions
(on the order of nanometers). However, the majority of the liquid’s
volume consists of regions that are completely unordered. Liquid par-
ticles vibrate randomly around their equilibrium positions for brief
moments before relocating to new equilibrium positions, which allows
liquids to flow. The duration of these vibrations varies among differ-
ent liquids and under different environmental conditions; however,
at a given pressure and temperature, the average duration of par-
ticle vibration is constant, a phenomenon known as residence time.
For instance, the residence time of liquid metal molecules is on the
order of 10−10 seconds, while that of water molecules is approxi-
mately 10−11 seconds. For a specific liquid, an increase in temper-
ature results in a decrease in residence time, with shorter residence
times correlating to enhanced liquidity. The microscopic characteris-
tics of gases differ significantly from those of solids and liquids. The
typical distance between neighboring gas molecules is considerably
larger, resulting in diminished intermolecular forces. The motion of
gas molecules can be approximated as uniform linear motion until
they collide with other molecules or the walls of their container.
Consequently, gases expand to occupy the entirety of their container.
Unlike two immiscible liquids, two different gases can mix uniformly
without forming a distinct interface.

2.1.2 Phase
A “phase” is defined as a region within a system that exhibits
uniform chemical and physical properties. It represents a state of
matter, distinct from the term physical state. The concept of phase
emphasizes the uniformity of physical and chemical properties, and
is not restricted to a single substance. Typically, there are three
states of aggregation — liquid, solid, and gas — associated with
a given substance. Within each state of aggregation, the chemi-
cal and physical properties are consistent throughout any portion
of a pure substance, thereby categorizing each state of aggregation
as a phase. Different states of aggregation correspond to different
phases; for instance, a system exhibiting two states of aggregation
Basic Conceptions 19

will be characterized by two distinct phases. If a system exhibits two


aggregation states, it can be characterized as having two distinct
phases. For instance, at a pressure of 101.325 kPa and a tempera-
ture of 373.15 K, water and vapor achieve equilibrium and coexist
within the system. Consequently, the system comprises two phases:
the liquid phase and the gas phase of water. An interface exists
between these phases, which possess differing chemical and physi-
cal properties. As a result, the two phases can be separated using
physical methods. At the macroscopic level, the transition in the
nature of the interface is abrupt and discontinuous, allowing for clear
differentiation between the phases.
In the three states of matter, gas molecules are characterized by
their ability to mix uniformly, resulting in a single phase regardless of
the number of substances present in the system. In contrast, liquids
exhibit different behavior. When two types of liquids can mix uni-
formly, they create a homogeneous phase. Conversely, if they cannot
mix uniformly, they will separate into two distinct phases, forming
an interface. The solid state is more complex than both gas and
liquid states. Typically, solids composed of different substances are
classified as belonging to different phases.
Under varying temperature and pressure conditions, phases can
undergo transformations into one another. This phenomenon is
referred to as phase conversion, a prevalent physical process observed
in nature. Phase conversions can be categorized into first-order and
second-order conversions. During first-order phase transitions, sig-
nificant changes in volume occur, accompanied by the absorption or
release of latent heat. A common example of this phenomenon is
the transformation of water from liquid to vapor or ice under spe-
cific conditions. In contrast, second-order phase transitions involve
changes in the system that do not exhibit the aforementioned char-
acteristics; that is, there are no significant volume changes, nor is
there any absorption or release of latent heat. However, during these
transitions, certain properties, such as the heat capacity and the
coefficient of thermal expansion, may undergo abrupt changes. For
instance, some materials that exhibit electrical resistance, including
insulators, at room temperature may experience a sudden loss of
resistance when the temperature is lowered to a critical threshold.
This transition from a normal phase to a superconducting phase
exemplifies a second-order phase transition.
20 Capillary Mechanics

2.2 Surface, Interface, and Bulk

2.2.1 The connotation of surface, interface and bulk


“Surface” usually refers to the border between the gas phase and the
condensed phase (liquid phase and solid phase), including the border
between gas and liquid and the one between gas and solid. “Interface”
is the border between two condensed phases, including the border
between liquid and solid, the one between two kinds of liquid, and
the one between two kinds of solid. From the point of view of phase,
the concept of interface is more common and more general than sur-
face. The surface of liquid–gas or solid–gas can be considered as the
interface between the gas phase and the condensed phase. Therefore,
a surface is also an interface between phases. “Bulk” (body) is a term
which is different from surface or interface. In order to distinguish
it from a surface or an interface, the non-interface internal part of
a state of matter or a phase (or surface) is called bulk. Generally,
bulk indicates the same kind of non-interface phase. The physical
and chemical properties are completely uniform and consistent in
one kind of bulk.

2.2.2 Features of interface


From a macroscopic perspective, an interface is defined as a bound-
ary that separates two distinct phases and is considered to have no
thickness. However, from a microscopic perspective, the existence of
thickness is acknowledged. Regardless of whether one examines the
interface at a macroscopic or microscopic level, it possesses numerous
inherent characteristics.

2.2.2.1 Energetic properties of interface


When two phases come into contact, a contact region is established.
Within this region, the intrinsic properties of the entire system
transition from one phase to another. For stability to be achieved,
the interface must possess an interfacial free energy. However, when
considered over an extended period, this stability is relative, as abso-
lute stability does not exist. Consequently, the stability of the system
is confined to a finite duration. According to the principle of least
energy, interfacial free energy consistently seeks minimization. In a
Basic Conceptions 21

two-phase system, if the presence of the interface results in elevated


free energy, the interface will spontaneously evolve toward a state
of minimization. Similarly, the two phases will tend to achieve an
extreme degree of separation.

2.2.2.2 Transition property of surface


According to the definitions of states of matter and phases, solids,
liquids, and gases (referred to as solid phase, liquid phase, and gas
phase, respectively) exhibit distinct characteristics that are often
manifested at their surfaces and interfaces. Generally, there is no
interface between different types of gases, as gases tend to disperse
spontaneously and achieve a uniform distribution. Similarly, there is
typically no interface between two different liquids.
Therefore, the term “interface” means the one that lies between
two kinds of solids, solid and liquid, and two kinds of immiscible liq-
uids, while the term “surface” refers to the one between solid and
gas or between liquid and gas. The dynamic behavior of surfaces or
interfaces is frequently neglected in discussions concerning the sta-
bility of interfaces. In reality, for both gases and liquids, the surfaces
that separate internal bodies, as well as those between liquid and gas
phases, are characterized by constant and rapid molecular exchange.
The exchange of molecules at the surface between two phases (such as
liquid and gas or solid and gas) is referred to as surface migration or
surface transition. Analyses indicate that at 25◦ C, the average dwell
time of a water molecule at the interface between gas and water is
less than 3 ms, while the dwell time of a mercury atom is approx-
imately 5 ms. In contrast, the dwell time of a solid metal atom at
room temperature is nearly 10 s. This significant difference in tran-
sition properties between the surfaces of solids and liquids results in
distinct characteristics for each. A newly formed liquid surface can
quickly achieve thermal equilibrium; however, most solid surfaces
require a considerably longer duration to reach a complete steady
state of thermal stability. It is also noteworthy that certain solids can
sublimate directly into the gas phase. Under conditions of dynamic
thermodynamic equilibrium, the number of vapor molecules emitted
from a liquid (or solid) during evaporation (or sublimation) is equal
to the number of molecules returning to the liquid (or solid). At this
point, the vapor is referred to as saturated vapor.
22 Capillary Mechanics

2.2.2.3 Characterization of molecular in the interface area


Between two distinct phases, there exists a region characterized
by asymmetric forces acting on the molecules, which results in an
increase in interfacial free energy. When two phases come into con-
tact, a region emerges where molecular size alterations can occur.
Within this region, the system experiences a phase transition, shift-
ing from one phase to another. In the case of a non-volatile solid
interacting with an inert gas, the interface transition region is approx-
imately the thickness of a single molecule. Consequently, this leads
to a sharp transition from the perspective of boundary dynamics,
wherein solid molecules abruptly transform into gas molecules at the
interface. Conversely, this transition is not as abrupt when a pure
liquid interacts with its corresponding gas phase. In this scenario,
the transition region spans several molecular units, and the molec-
ular concentration transitions gradually from liquid density to gas
density across various density units. The subsequent chapters will
provide a more detailed analysis of the interfacial layer.

2.3 Surface Free Energy

The concepts of surface, interface, and bulk (or body) have been
delineated in the preceding section. The surface is a specific type of
interface, representing the boundary between tangible objects in the
condensed phase and the intangible gas in the gas phase. The term
“bulk” relates to the interface as well. Both surface and interface
possess inherent energy characteristics.
To effectively comprehend the concept of surface free energy, this
chapter initially presents the definitions of new surface and free
energy. When the body of an object is divided into two parts along
an interface in a vacuum or air, two new surfaces are created. These
surfaces are referred to as new surfaces.
The concept of free energy originates from thermodynamic theory.
From a thermodynamic perspective, when the temperature exceeds
absolute zero, the molecules of any substance exhibit thermal motion,
which is inherently associated with the energy of the system. Conse-
quently, all objects possess energy, specifically thermal energy. The
total thermal energy of an object comprises two components: one is
the observable thermal energy, commonly referred to as free thermal
Basic Conceptions 23

energy, and the other is potential thermal energy. Observable thermal


energy can be measured using a thermometer, whereas potential ther-
mal energy remains unobservable and reflects chemical effects. The
notion of free energy is derived from the concept of free heat. In 1882,
the German physicist and biologist Hermann von Helmholtz first
introduced the term and defined what is now known as Helmholtz
free energy.
The expression is represented as A = U − T S, where A is
Helmholtz free energy and the first letter of Arbeit in German. U
is internal energy, T is the thermodynamic temperature of the sys-
tem, and S is entropy.
Entropy is a thermodynamic property that serves as an indicator
of the energy available for performing useful work during a ther-
modynamic process. In the field of thermodynamics, the concept of
entropy is articulated through the second law of thermodynamics.
Entropy is recognized as a state function of a material system, which
exemplifies the irreversibility of thermal processes and is typically
denoted by the symbol S. In an isolated system, the changes that
occur invariably lead to an increase in entropy. From the perspective
of molecular thermal motion, the molecules within material systems
tend to transition from a state of order to one of chaos, primarily
due to thermal agitation. The increase in entropy signifies a rise in
the degree of molecular disorder. Consequently, entropy serves as a
quantitative measure of the disorder or randomness present within a
thermodynamic system. It is well established that entropy reflects the
level of molecular disorder associated with thermal motion within the
system. The term TS represents the portion of internal energy that is
incapable of performing useful work, often referred to as irreversible
work. Therefore, under specific conditions, free energy serves as an
indicator of the degree of freedom or the amount of internal energy
available for work within the system. In addition to Helmholtz free
energy, there exists another form of free energy known as Gibbs free
energy. Its expression is represented as G = H −T S, where G is Gibbs
free energy. H is the enthalpy of the system. H = U + P V , where P
is pressure and V is volume. The two types of free energy mentioned
above are typically applicable in different contexts. Gibbs free energy
is predominantly utilized in isobaric and isothermal processes, while
Helmholtz free energy is primarily employed in isochoric and isother-
mal processes. Generally, free energy is associated with the capacity
24 Capillary Mechanics

to perform useful work. Notably, in the field of physics, the term


refers to Helmholtz free energy, whereas in chemistry, it commonly
denotes Gibbs free energy.
According to the definition of Helmholtz free energy, the total
differential is

dA = dU − T dS − SdT (2.1)

According to the first law of thermodynamics, in a closed system,

dU = T dS − pdV (2.2)

Substituting equation (2.2) into (2.1) leads to

dA = −pdV − SdT (2.3)

According to the definition of Gibbs free energy, the total differ-


ential is

dG = dU + pdV + V dp − T dS − SdT (2.4)

From the first law of thermodynamics, in a closed system, it is


known that

dU = T dS − pdV (2.5)

Substituting equation (2.5) into (2.4) leads to

dG = V dp − SdT (2.6)

The bulk (or body) of an object, which represents a specific state


of matter or phase, experiences intermolecular attractions among
its constituent units, such as molecules or atoms. However, these
attractions result in an overall state of equilibrium. In the context of
thermodynamics, free energy is present within the system contain-
ing the substance. When the bulk is divided along an interface to
create two new surfaces, the units situated at the interface become
unbalanced due to their interactions. This results in a net attraction
for individual units. For one of the newly formed surfaces, the side
that remains in contact with the bulk continues to experience cohe-
sive attraction, while the opposite side is either not attracted (in a
vacuum) or experiences weak attraction from the surrounding gas
Basic Conceptions 25

(in air). To address this imbalance and achieve a new equilibrium,


a tightening effect emerges on the two new surfaces. From a force
perspective, the original cohesive attraction, which acts perpendicu-
lar to the direction of the new surface, is disrupted. Consequently, a
tension force parallel to the direction of the new surface is generated
to establish a new balance. From an energy standpoint, this tighten-
ing effect indicates that additional free energy (or new free energy) is
created on the two new surfaces compared to the free energy of the
original system. The free energy associated with the formation of the
new surfaces is referred to as surface free energy. For analytical con-
venience, the concept of specific free energy is frequently employed,
which quantifies the amount of surface free energy per unit area.
The preceding analysis and description regarding the generation
and implications of surface free energy are primarily qualitative in
nature. This qualitative understanding is instrumental in elucidating
the concept and formation mechanisms of surface free energy. How-
ever, in practical applications, a quantitative description of surface
free energy proves to be more beneficial. From a quantitative perspec-
tive, surface free energy is defined as the reversible work required to
separate one interface from the bulk material. The prior analysis indi-
cates that individual units within the bulk are subject to attractive
forces from neighboring units, resulting in a cohesive effect. When
a unit located at an interface is separated along this interface, the
cohesive forces must be overcome, which is accomplished through the
work performed by an external force. The work performed to coun-
teract the cohesive forces is subsequently stored at the interface as
potential energy, which is referred to as surface free energy. An anal-
ysis of intermolecular forces indicates that this attractive cohesive
force diminishes at a rate approximately proportional to the sixth
power of the distance for a single unit (molecule or ion). In con-
trast, for larger objects, such as semi-infinite bodies, this attractive
cohesive force decreases at a rate ranging from quadratic to quartic
power with respect to the distance. Generally, this force is character-
ized as a short-range interaction. When the separation distance falls
within the sub-micron to micron range, the force is nearly completely
attenuated. Consequently, the phenomenon commonly referred to as
interface separation is typically considered to occur within the sub-
micron to micron scale. The characterization of magnitude in rela-
tion to separation distance is relatively straightforward. In contrast,
26 Capillary Mechanics

the quantitative characterization of cohesive attraction presents a


greater degree of complexity. Consequently, the method for quanti-
tatively describing surface free energy is inherently challenging. From
this perspective, the force is considered to be perpendicular to the
surface; thus, analyses typically adopt this viewpoint when discussing
adhesive force, adhesive strength, or cohesion. It is noteworthy that
free energy is seldom analyzed from this angle. Instead, the trend in
tension and the additional (new or surplus) free energy associated
with the newly created surface are employed to describe surface free
energy. When one phase is in the gas state, its interface is defined as a
surface. Therefore, surface free energy can be articulated in terms of
a tensioning or tightening force that is parallel to the surface and acts
upon it. In light of the observed trend in tension and the additional
energy associated with the newly formed surface, it is necessary to
apply an external force to separate the interface. The work performed
by this external force is equivalent to the work required to overcome
the original cohesive forces at the interface. When surface tension
is employed to characterize the tension level of the new surface, it
follows that, from the perspective of overcoming surface tension, the
surface free energy must correspond to the work expended in the
process of separating (or forming) the new surface. The fundamental
principle underlying this analysis remains consistent with the pre-
vious discussion. The energy exerted by the bulk material on the
interface manifests as a tightening tendency at the interface. Conse-
quently, a suitable definition can be articulated as follows: specific
surface free energy (commonly referred to as surface free energy)
is defined as the work required to overcome surface tension (or the
tightening force) per unit area.

2.4 The Interface Free Energy and Interfacial Tension

The aforementioned surface represents the interface between the


condensed phase and the gas phase. The energy primarily pertains to
the interfacial free energy associated with both the condensed phase
and the gas phase. A comparable additional free energy is also present
at the interface between two condensed phases. The function of inter-
facial free energy is to stabilize the interface, which results in the
minimization of interfacial free energy and maximizes the separation
between the two phases.
Basic Conceptions 27

When two distinct condensed phases come into contact, an


interface is established. If both phases are liquid, as in the case
of two incompatible pure liquid phases, a liquid–liquid interface is
formed. Conversely, when one phase is solid and the other is liquid,
a liquid–solid interface is created. In instances where two different
solid phases come into contact, a solid–solid interface is generated.
However, due to the surface characteristics of solids, achieving close
contact between solid–solid interfaces to attain thermal equilibrium is
often challenging. Consequently, the interfaces that are typically dis-
cussed are the solid–liquid interface and the liquid–liquid interface,
which arises from the interaction of two immiscible liquids. Given
that the physical states on either side of the interface differ, the
cohesive interactions among the material molecules on each side are
also distinct. Following a similar principle of surface free energy, a
new interfacial free energy is generated. When one of the phases is
gaseous, the interfacial free energy reduces to surface free energy. In
cases where two identical liquid phases come into contact, the asym-
metry is eliminated, resulting in an interfacial free energy of zero.
Thus, the interfacial free energy serves as an indicator of the asym-
metric attraction present on both sides of the interface. Additionally,
there are scenarios in which the interfacial free energy is zero despite
the two phases being different; in such cases, the contacting phases
do not form an interface. This phenomenon can occur, for example,
when a non-wetting liquid contacts its corresponding solid or when
two different compatible liquids interact.
The phenomenon of interfacial tension can be likened to surface
tension, as it represents a force that acts to contract the interface.
This force arises from the asymmetric attraction exerted by the two
phases on either side of the interface, resulting in a tightening effect.

2.5 Cohesion and Adhesion

The pervasive existence of intermolecular forces results in the


integration of substances into various forms, commonly referred to
as the states of matter. Molecules of the same type tend to aggre-
gate, while molecules of different types can also form aggregates.
The attractive forces between identical molecules, which lead to the
cohesion of similar materials, are termed cohesion forces; these forces
operate at the material level. Conversely, the bonding effect that
28 Capillary Mechanics

occurs between two different materials due to the attractive forces


between dissimilar molecules is known as adhesion forces, which also
function at the material dimension. Adhesion effects can be catego-
rized based on various mechanisms, including mechanical adhesion,
chemical adhesion, dispersion adhesion, electrostatic adhesion, and
diffusion adhesion. Mechanical adhesion, akin to the sewing effect,
occurs when the adhesive material penetrates the pores or voids at
the interface. In contrast, chemical adhesion involves the formation
of compounds at the junction of materials, resulting in adhesion
through ionic bonds, covalent bonds, and other types of chemical
interactions. Dispersion adhesive can be characterized as a form of
adsorption, which pertains to the adhesion of two materials through
van der Waals forces. Electrostatic adhesion, on the other hand,
involves conductive materials that acquire a charge, resulting in a
charge differential at their interface. Consequently, this interaction
creates a structure analogous to a capacitor, wherein electrostatic
attraction occurs between the two plates of the capacitor-like config-
uration. When the molecules of the two materials exhibit high levels
of activity, they may undergo mutual dissolution and infiltration,
leading to the formation of a polymer. This phenomenon is referred
to as diffusion adhesion.
A phenomenon can be attributed to a multitude of forces,
exemplified by a drop of water falling onto the surface of a leaf.
The cohesive forces contribute to the formation of a water droplet,
while surface tension causes the droplet to assume a nearly spherical
shape. Additionally, adhesive forces facilitate the attachment of the
droplet to the leaf’s surface.
The concepts of cohesion and adhesion are inherently complex;
therefore, the terms cohesion work and adhesion work are frequently
employed to elucidate these phenomena.
Cohesive work, also referred to as cohesive energy, is defined as the
reversible work required to separate a unit area of material into two
distinct surfaces. This concept can be readily understood through
the principles of free surface energy. Consequently, cohesive work
is indicative of the inseparability resulting from the gravitational
forces acting between the internal molecules of the material. When an
interface is separated, it results in the formation of two new surfaces.
Therefore, the work of adhesion is expected to be twice the value of
Basic Conceptions 29

the surface free energy. If the surface free energy of the material is
γ, the adhesion work (energy) is Wc = 2γ.
The adhesion work, also referred to as adhesion energy, is defined
as the reversible work required to separate two distinct materials
per unit area along their interface. This adhesion work is not solely
dependent on the surface free energy of the individual materials; it
is also influenced by the interfacial free energy that exists between
them. This concept underscores the inseparability of the two materi-
als at the interface. When two different materials come into contact,
a new interfacial free energy is generated at the interface. Upon sep-
aration, the total free energy of the surfaces of the two newly formed
materials is the sum of their respective surface free energies. However,
the interfacial free energy dissipates upon separation. Consequently,
the net increase in energy of the system following separation is equiv-
alent to the difference between the combined surface free energies of
the two new surfaces and the original interfacial free energy. If the
surface free energy of material 1 is γ1 , the surface free energy of
material 2 is γ2 , and the interfacial free energy between them is γ12 ,
then the adhesion work (energy) is Wa = γ1 + γ2 − γ12 .

2.6 Interface Layer and Surface Force

The interface layer is a thin region that exists at the two-phase


boundary, in contact with each of the two bulk phases. Consequently,
the interface layer is characterized as three-dimensional. To
differentiate it from the three-dimensional interface layer, the two-
dimensional interface between the two phases is referred to as the
physical interface. The presence of the physical interface signifies that
there are variations in various physical quantities, including compo-
sition, on either side of the physical interface. Concerning the bulks
on both sides of the interface, this alteration can be classified as a
mutation, often referred to as a step change. However, it is important
to note that this change should not be perceived merely as a simple
mutation between the two phases; rather, it should be understood
as a continuous process of transformation. This ongoing process is
manifested at the interface layer. Given that the physical characteris-
tics and chemical properties of the interface layer are non-uniform, it
can be argued that the interface layer does not constitute a singular
30 Capillary Mechanics

phase. Current research in interface science identifies three distinct


types of interface layer models. The first model is the Gibbs surface
segmentation model. The second model is the Guggenheim interface
transition model. The final model is the physical interface model.
Essentially, the Gibbs surface segmentation model represents a dig-
ital interface layer; however, the physical significance of this inter-
face layer remains ambiguous. The Guggenheim interface transition
model is a modification of the Gibbs surface segmentation model.
This model treats the interfacial transition zone as a thermodynamic
entity, even referring to it as an interface phase. Both models men-
tioned tend to obscure the physical interface between the two phases,
leading to the perception that the macro-interface is equivalent to
the micro-interface, with each interface consisting of just one inter-
face layer. Consequently, both models exhibit certain limitations. The
physical interface type of the interface layer model first establishes
the existence of the physical interface. Furthermore, this model posits
that there are two thin interface layers that are separately connected
to the bulk materials on either side of the physical interface. These
two thin interface layers possess a thickness equivalent to the radius
of molecular interaction. Notably, the properties of the two interface
layers exhibit a sudden change on both sides of the physical interface,
a phenomenon referred to as a step change.
Molecules are influenced by both the interfacial adhesion forces
of the material and the cohesive forces at the interface. When the
interface layer is considered as a distinct entity, it is essential to
analyze its overall force by examining both the internal and external
forces acting on the surface.
As previously stated, interactions among identical molecules
result in cohesive forces, while interactions between dissimilar
molecules give rise to adhesive forces. Within a particular phase
of the bulk, cohesive forces are uniformly present throughout. At
an interface within the bulk (where the same phase is maintained),
cohesive forces manifest as a pulling force that is perpendicular to the
interface, a phenomenon that can be referred to as negative pressure.
The negative pressure exerted by the same kind of molecules per unit
area is referred to as molecular internal pressure. At the interfaces
between two different phases, the distinct types of molecules influ-
ence the interface, resulting in an adhesive pulling force (negative
pressure) that is perpendicular to the interface, which can be termed
Basic Conceptions 31

molecular external pressure. On both sides of the physical interface


layer, one side is connected to the bulk phase, while the other side is
connected to another interface layer. The side connected to the bulk
phase is influenced by molecular internal pressure, whereas the side
connected to the interface phase is affected by molecular external
pressure.
The integration of molecular internal pressure across the thickness
of the interface layer is defined as the surface internal force. Similarly,
the integration of molecular external pressure along the thickness of
the interface layer is defined as the surface external force. After per-
forming this integration and neglecting the thickness of the interface
layer, we can analyze the cohesive effects of the bulk material and the
adhesive effects of the adjacent phase at the interface. Consequently,
by disregarding the thickness of the interface layer, the surface inter-
nal force represents the attractive cohesion of the liquid bulk per unit
area of the interface layer. The surface external pressure refers to the
attractive adhesion of the adjacent phase (the interface layer of the
other phase) per unit area of the interface. For an interface layer
with negligible thickness, one side is influenced by surface internal
forces, while the other side is influenced by external surface forces.
The dimensions of both surface internal and external forces are equiv-
alent to the dimension of surface tension, as they all represent force
per unit length. However, it is important to note that both surface
internal and external forces act perpendicular to the interface, which
distinguishes them from the direction of surface tension. The surface
internal force acts toward the interior of the body, while the surface
external force is directed toward the two-phase physical interface.
Further analysis using molecular theory indicates that the internal
surface force is equivalent to half of the cohesive work and is the
same as the surface tension. That is,

W11
σ11 = = γ11 (2.7)
2
The surface external force is half of the adhesive work and is equal
to half the value of the difference between the sum of the two-phase
surface internal forces and the interfacial tension:
W12 γ11 + γ22 − γ12 σ11 + σ22 − γ12
σ12 = = = (2.8)
2 2 2
32 Capillary Mechanics

where σij is the surface force. When i = j, it is the surface internal


force; when i = j, it is the surface external force. Wij is the surface
work. When i = j, it is the cohesive work; when i = j, it is the
adhesive work. γij is surface (or interface) tension. When i = j, it is
the surface tension; when i = j, it is the interface tension.

2.7 Chemical Potential

2.7.1 The concept of chemical potential


In a simple closed system — defined as a specific amount of pure
substance or a mixture whose composition remains unchanged (with
no phase transitions or chemical changes) — only energy (in the form
of heat and work) is exchanged with the environment, without any
material exchange. Consequently, two thermodynamic variables, such
as temperature and pressure, can be utilized to determine the state of
the system. However, in open systems or systems with variable com-
position, this approach is no longer applicable. Under the condition
of constant temperature and pressure (dT = 0 and dp = 0), Gibbs
free energy, described by the formula dG = −SdT + V dp, is zero
(dG = 0). The Gibbs free energy of the system remains unchanged.
This statement is clearly incorrect for an open system, as the addi-
tion of substances to the system will result in an increase in the
system’s Gibbs free energy. For instance, when reactive substances
are added to a battery, the electrical power produced by the battery
must increase. This increase is a direct consequence of the rise in
Gibbs free energy within the system. Similarly, under conditions of
constant temperature and pressure, if a spontaneous chemical change
occurs in a closed system, dGT,p should be less than zero. This is
due to the fact that an irreversible change in the composition of the
system has taken place.
However, according to formula (2.5), since dT = 0 and dp = 0,
dGT,p = 0, which is clearly inconsistent with the practical situation.
Therefore, equation (2.5) is not applicable to the process of changing
components in a closed system. These two examples illustrate a com-
mon characteristic: when the amount of one or more components
changes in a system, this alteration will result in a change in free
energy. Therefore, for both open systems and closed systems with
varying components, it is essential to not only describe free energy
Basic Conceptions 33

using two thermodynamic variables but also to introduce the factor


that causes the change in the thermodynamic function due to the
alteration in system composition. Thus, for open systems or closed
systems with changing components, G (Gibbs free energy) not only
depends on the temperature and pressure but also depends on the
amount of each component in the system n1 , n2 , . . ., i.e.,

G = f (T, p, n1 , n2 , . . .) (2.9)

The total differential of G can be written as


     
∂G ∂G ∂G
dG = dT + dp + dn1
∂T p,n1 ,n2 ,... ∂p T,n1 ,n2 ,... ∂n1 T,p,n2,...
 
∂G
+ dn2 + · · · (2.10)
∂n2 T,p,n1 ,...

The above equation shows that the change in the Gibbs free
energy of the system is equal to the sum of the changes in G caused
separately by temperature changes dT (when the pressure and the
composition remain unchanged ), pressure changes dp (when the
temperature and the composition remain unchanged), and changes
in each component dn1 , dn2 , . . . (when the pressure, temperature,
and other components remain unchanged in addition to a par-
ticular component). The partial derivatives (∂G/∂T )p,n1 ,n2 ,... and
(∂G/∂p)T,n1 ,n2 ,... in the above equation are the same as (∂G/∂p)T
and (∂G/∂T )p of the closed system, respectively. The rates of change
in Gibbs free energy resulting from variations in temperature or pres-
sure, while maintaining a constant composition, are all equivalent.
Consequently, the rate of change corresponds to the relationship of
Gibbs free energy in a closed system where the composition remains
unchanged. That is,
 
∂G
= −S (2.11)
∂T p,n
 
∂G
=V (2.12)
∂p T,n
34 Capillary Mechanics

Then,
 
∂G
dG = −SdT + V dp + dn1
∂n1 T,P,n2 ,...
 
∂G
+ dn2 + · · · (2.13)
∂n2 T,P,n1 ,...

To simplify formula (2.13), in 1875, Gibbs proposed using the


chemical potential to replace the partial derivative (∂G/∂ni )T,p,nj=i ,
i.e.,
 
∂G
μi (T, p, n1 , n2 , . . .) = (2.14)
∂ni T,p,nj=i

In this function, nj=i means that except the component ni , others


components remain unchanged. The chemical potential, as described
by the equation above, is a function of temperature, pressure, and
the composition of the system. Therefore, the physical significance of
chemical potential is the change in Gibbs free energy (often referred
to as the Gibbs free energy change rate) that results from the addi-
tion of a unit of component I. This addition is so small that it does
not significantly alter the composition of the system under constant
temperature and pressure. The formula can be written as
   
ΔG ∂G
μi (T, p, n1 , n2 , . . .) = lim = (2.15)
Δni →0 Δni T,p,n
j=i
∂ni T,p,nj=i

The concept of chemical potential can also be understood as the


change in Gibbs free energy that would occur when 1 mole of a
substance is added to a large system. Since the system is very large,
the addition of 1 mole of material will not significantly alter the
composition of the system. A single-component system only contains
one component. Therefore, (2.14) can be written as
 
∂G
μ= = Gm (2.16)
∂n T,p

The aforementioned equation indicates that, in a single-


component system, the chemical potential is equivalent to the rate
Basic Conceptions 35

of change of Gibbs free energy when the substance is in a pure state.


This concept is also referred to as molar Gibbs free energy.
The definition of chemical potential indicates that it is a physical
parameter related to the strength properties of a system and serves
as a state function. Under conditions of constant temperature and
pressure, the value of chemical potential varies solely with changes in
composition and is independent of the total amount of the system.
After introducing chemical potential μi , (2.13) can be written as

dG = −SdT + V dp + μi dni (2.17)
i

Similarly, if the internal energy depends on the volume, entropy,


and composition of the system, then

U = f (S, V, n1 , n2 , . . .) (2.18)

The enthalpy H is the composition of the pressure and entropy of


the system. Helmholtz free energy A is decided by the temperature
and the volume. That is,

H = f (S, p, n1 , n2 , . . .) (2.19)
A = f (T, V, n1 , n2 , . . .) (2.20)

Then, using the same method as above, it is not difficult to obtain


the relationships about the chemical potential:
     
∂U ∂H ∂A
μi = = =
∂ni S,V,nj=i ∂ni S,p,nj=i ∂ni T,V,nj=i
 
∂G
= (2.21)
∂ni T,p,nj=i

Correspondingly, the total differential of its internal energy U,


enthalpy H, and Helmholtz (A) free energy can be written as

dU = T dS − pdV + μi dni (2.22)
i

dH = T dS + V dp + μi dni (2.23)
i
36 Capillary Mechanics


dA = −pdV − SdT + μi dni (2.24)
i

dG = −SdT + V dp + μi dni (2.25)
i

The four equations above are suitable for the basic equation of
chemical thermodynamics in open systems and closed systems in
which the composition changes. For a process under constant tem-
perature and pressure, we have

dGT,p = μi dni (2.26)
i

When a reversible change (phase change or chemical change) in


the composition occurs or the system reaches equilibrium, the change
in Gibbs free energy should be zero. That is,

dGT,p = 0, μi dni = 0 (2.27)
i

With an irreversible change (phase change or chemical change) in


the composition, Gibbs free energy should decrease:

dGT,p < 0, μi dni < 0 (2.28)
i

Combining the above two equations,



μi dni ≤ 0 (2.29)
i

The equation is applicable to the reversible process, while the


inequality is applicable to the irreversible process.

2.7.2 The relationship between chemical potential


and temperature pressure
The preceding discussions regarding chemical potential predomi-
nantly focus on isothermal and isobaric conditions. This empha-
sis serves to illustrate the relationship between chemical potential
Basic Conceptions 37

and the rate of change of the relative components of Gibbs free


energy under these specific conditions. However, this should not be
interpreted as an indication that there is no relationship between
chemical potential and temperature or pressure. On the contrary,
chemical potential is intricately linked to both temperature and
pressure.
According to the definition of chemical potential (2.14) and (2.12),
the relationship between the chemical potential and pressure is
   
∂G ∂G
∂μi ∂ ∂ni ∂ ∂p ∂V
= = = (2.30)
∂p ∂p ∂ni ∂ni
Since there is only one component, for the single-component sys-
tem, we have
∂V ∂V
= = Vm (2.31)
∂ni ∂n
Then,
 
∂μ
= Vm (2.32)
∂p T

That is, for pure components, under a constant temperature, the


change rate of chemical potential to pressure is equal to the molar
volume.
Similarly, from (2.14) and (2.12), the relationship between chem-
ical potential and temperature can be derived as
 
∂G  
∂μi ∂ ∂ni ∂ ∂G ∂S
∂T
= = =− (2.33)
∂T ∂T ∂ni ∂ni
For a single-component system, we have
∂μ
= −Sm (2.34)
∂T
That is, for a single-component system, under a constant pressure,
the change rate of chemical potential to temperature is equal to a
negative molar entropy.
38 Capillary Mechanics

2.7.3 The application of chemical potential in the


phase transition
If under a given temperature and pressure, the liquid in a closed
container evaporates, the evaporation tends to a certain degree of
saturation and then reaches gas–liquid equilibrium.
According to (2.32), the chemical potential can be used to
calculate the change in the Gibbs free energy of the system:

dGT,p = μi dni = μl dnl + μg dng (2.35)
i

Because the number of water liquid disappeared (dnl ) is equal to


the number of water vapor appeared (+dng ), i.e., −dnl = dng , the
above equation can be written as

dGT,p = (μg − μl )dng (2.36)

When the water irreversibly transfers into vapor, μg − μl < 0, i.e.,


μg < μl .
Thus, when the water irreversibly transfers from liquid into vapor,
it always travels from the high-chemical-potential liquid phase to
the low-chemical-potential gas phase until the water and vapor
reach the chemical potential equilibrium. The condition of gas–liquid
equilibrium is

μH2 O,l = μH2 O,g (2.37)

Extending the above results to a multi-component equilibrium


system, when it does not reach equilibrium, component B in the
prior system always travels from the high chemical potential to the
low chemical potential until all the chemical potentials are equal.
Therefore, the equilibrium condition of each phase of component B
in the multiphase system is

μαB = μβB = μγB = · · · (2.38)

In the state of saturation, since the transition between gas and


liquid phases is reversible, the change in the free energy of the system
is zero.
Basic Conceptions 39

2.7.4 Ideal gas chemical potential


The preceding analysis indicates that, under specific conditions —
such as constant temperature and pressure, constant entropy and
volume, constant pressure, constant entropy, or constant temperature
and volume — the criterion for a system to achieve phase equilib-
rium is that the chemical components across different phases possess
identical chemical potential. However, it is important to note that
this represents a general principle. A detailed characterization of the
chemical potential for particular substances and systems is necessary
for a comprehensive understanding.
For a single-component ideal gas, it is known by (2.34) that
∂μ
= Vm (2.39)
∂p
Therefore, change in chemical potential can be written as

dμ = Vm dp (2.40)

For an ideal gas, Vm = RT /p, where T is absolute temperature


and R is the gas constant. Substituting it into (2.40) leads to
RT
dμ = dp = RT d ln p (2.41)
p
If the pressure changes from p1 to p2 , the chemical potential
changes into
p2
Δμ = RT ln (2.42)
p1
This page intentionally left blank
Chapter 3

Surface Tension and Interfacial


Tension

3.1 Conception of Surface Tension

The concept and mechanism of surface free energy were introduced


in the previous chapter. Surface free energy is a form of potential
energy. There are three primary reasons for the formation of surface
free energy. The first is the creation of a new surface. The second
is the cohesive forces present in the bulk material. The third is the
asymmetry or imbalance produced by these cohesive forces, which
results in a tendency for tension on the surface. The dwell time of
surface molecules varies across different phases. In the liquid phase,
the dwell time is so short that the system quickly reaches thermo-
dynamic equilibrium. Consequently, the surface tension of the liquid
phase stabilizes rapidly. In contrast, the transfer rate of solid sur-
face molecules is relatively slow, resulting in a longer dwell time for
the solid phase. As a result, achieving thermodynamic equilibrium
in the solid phase within a short time frame is challenging. There-
fore, the macroscopic characterization of its surface tension trend is
not readily apparent. The surface free energy serves as a measure of
surface tension; however, it lacks directionality. To provide an initial
description of the directional aspect of tension, which characterizes
force, the concept of surface tension is defined. Thus, surface tension
describes the degree of tension at the surface, encompassing both
magnitude and directional characteristics. The tension exists in the

41
42 Capillary Mechanics

two-dimensional tangential direction of a surface, and its magnitude


reflects the degree of tension. As a two-dimensional force, it generally
indicates the force density, representing the tension per unit length.
Therefore, surface tension can be defined as a tension force acting
along a unit length that is parallel to the surface and perpendicular
to a specific straight line within the surface layer. This definition is
rooted in mechanics. The effect of surface tension on solid surfaces is
not as pronounced, as the time required for a solid surface to reach
thermodynamic equilibrium differs significantly from that of a liq-
uid surface. Consequently, the concept of surface tension is primarily
applied to the liquid phase, while free energy is typically used in the
context of solid surfaces.
Surface free energy is derived from surface tension, and both
concepts are utilized to describe the trends and magnitude of surface
tension. Consequently, surface tension can be defined in terms of both
mechanics, as previously mentioned, and energy. From an energy per-
spective, surface tension can be defined as the increase in surface
free energy per unit increase in area. In fact, these two definitions
are equivalent. For example, when a surface tension γ acts on the
surface, in order to overcome the surface tension by an outside force,
a work of γ · δA must be done to increase the surface area δA. This
work is stored in the surface in the form of potential energy, which
produces the new surface free energy. According to the definition of
free energy, if the new surface free energy is divided by the increased
area δA, the surface tension in mechanics definition can be easily
obtained.
There are several phenomena that can help us understand surface
tension from a mechanical perspective. These phenomena can be
categorized into two cases: statics and dynamics. A coin cannot
float on the surface of water solely due to the buoyant force, as
the mass density of the coin is significantly greater than that of
water. However, when a coin is gently placed on the water’s surface,
it does not sink, as illustrated in Figure 3.1, which represents a static
phenomenon of surface tension. In Figure 3.2, a water strider can
effortlessly move across the water’s surface using its small legs, exem-
plifying a dynamic phenomenon of surface tension. Both the floating
coin and the moving water strider are influenced by surface tension.
There are two more experiments explaining surface tension from
the mechanical standpoint. One is pulling soapy water by a piece of
Surface Tension and Interfacial Tension 43

Figure 3.1. Coin floats on water.

Figure 3.2. Water strider moves on water.

wire, and the other is putting a cylindrical needle floating on water


surface. In Figure 3.3, a soap film is formed when a piece of wire on
a wire frame is pulled slowly after the frame dipped in soapy water.
The phenomenon shown in Figure 3.3 can be described by mechanics
as follows.
If the length of the moving wire is l, the stretching force is F,
and the displacement of moving wire is Δx, the newly formed area
will be 2lΔx. If the surface tension of the upper and lower surfaces
is γ (tension on a unit length), the total force is 2γl, which should
equal the stretching force F · F = 2γl, and so γ = F/2l. From this
point of view, the mechanical meaning of surface tension during the
movement can be understood.
44 Capillary Mechanics

Figure 3.3. A metal frame pulls a soap film.

Figure 3.4. A columniform needle presses on water.

The other experiment is shown in Figure 3.4. A cylindrical needle


floats on the surface of water. Its weight is fw . The needle cannot sink
because its weight is balanced by the surface tension forces fs . fs is in
the tangent direction in which the needle touches the water and acts
on both sides of the needle. Horizontally, the surface tension effects
on the two sides are balanced out: fs sin θ − fs sin θ = 0. Vertically,
surface tension is the same as the weight of the needle. The length of
the cylinder is l. The surface tension of the water is γ. Then, fs = γl.
γ = fs /l

fw = fs cos θ + fs cos θ = 2fs cos θ

From this point of view, the mechanical meaning of a static float


object under surface tension can be also understood.
Of course, surface tension can be explained in terms of energy.
These two definitions are equivalence relations. In the experiment
of pulling soap film mentioned above, the work done by the outside
force F is F Δx. This work is stored as a potential energy in the
two surfaces of the soap film. The total area of the surfaces is 2Δxl.
This energy forms the surface free energy A: A = F · Δx. The free
energy per unit area (also called the surface free energy ratio) is
Surface Tension and Interfacial Tension 45

γ = A/2Δxl = F Δx/2Δxl = F/2l. Compared to the mechanical


characterization of surface tension above, the surface free energy γ
is the same as the surface tension γ mentioned above.

3.2 Molecule Meaning of Surface (Interface) Free


Energy–Surface (Interface) Tension

The concept and mechanism of surface free energy were introduced


in the previous chapter. Surface free energy is a form of potential
energy. There are three primary reasons for the formation of surface
free energy. The first is the creation of a new surface. The second
is the cohesive forces present in the bulk material. The third is the
asymmetry or imbalance produced by these cohesive forces, which
results in a tendency for tension on the surface. To counteract this
imbalance, a tensioning trend (or tightening) develops on the sur-
face. The fundamental internal factor of the formation of surface free
energy is the existence of attractive interaction between molecules
in the bulk (body); the fundamental external factor is the formation
of a new surface. Surface free energy is defined as “a reversible work
done to overcome the cohesion between the molecules inside the bulk
(body) in order to separate the body along a certain interface”. This
work is stored in the form of potential energy in the newly produced
surface, which is called surface free energy.
According to the molecule theory, the van de Waals force exists
between two molecules. Therefore, there is van der Waals interaction
potential between two molecules, and its expression is
Cvdw
Wvdw = (3.1)
r6
where Cvdw is the coefficient of the interaction potential by van der
Waals molecular pairs. Cvdw = Corient + Cind + Cdisp , in which
Corient is the coefficient of oriented interaction potential by Kee-
som molecular pairs, Cind is the coefficient of induced interaction
potential by Debye molecular pairs, and Cdisp is the coefficient of
dispersion interaction potential by London molecular pairs. There is
a minus in the original formula. For convenience of analysis, take the
attractive effect as positive. Therefore, the minus can be taken out
of the formula.
46 Capillary Mechanics

The magnitude of the van der Waals force between two


molecules is
∂Wvdw 6Cvdw
fvdw = − = (3.2)
∂r r7
It is along the direction of the connection of the two molecules.
According to Hamaker theory, the total interaction potential of
the object system is equal to the sum of the interaction potentials
among the molecules, so

N N
1  ij
Wsum = Wvdw (rij ) (3.3)
2
i=0 j=0(=i)

ij
where Wvdw is the interaction potential between molecule i and
molecule j, which is not affected by any other molecules; rij is the
distance between molecule i and molecule j; and N is the number of
molecules.
According to this theory, the interaction potential between a
single molecule and an infinite size of plate with certain thickness is
 l+h  ∞ 
x
W (l) = 2πCvdw ρ1 dz dx
l 0 (z + x2 )3
2
 
2πCvdw ρ1 1 1
= − (3.4)
12 l3 (l + h)3

where ρ1 is the number of molecules per unit volume (molecule den-


sity) of the plate with an infinite size and a certain thickness, h is
the thickness of plate, and l is the nearest distance between a single
molecule and the plate.
If the thickness is infinite, the interaction potential between a
single molecule and the semi-infinite body can be described as
 ∞  ∞ 
x 2πCvdw ρ1
W (l) = 2πCvdw ρ1 dz dx =
l 0 (z + x2 )3
2 12l3
(3.5)
If a unit area is used for the plate with infinite size and certain
thickness to replace the single molecule, the interaction potential of
Surface Tension and Interfacial Tension 47

two plates with infinite size and certain thickness is


  
2πCvdw ρ1 ρ2 l+h2 1 1
W (l) = − dz
12 l z 3 (z + h1 )3
 
2πCvdw ρ1 ρ2 1 1 1 1
= − − +
24 l2 (l + h2 )2 (l + h1 )2 (l + h1 + h2 )2
(3.6)

where h1 and h2 are the thicknesses of the two plates with infinite
size, respectively.
When h1 → ∞, the interaction potential between a plate with
infinite size and certain thickness and a semi-infinite body is
 
2πCvdw ρ1 ρ2 1 1
W (l) = − (3.7)
24 l2 (l + h2 )2

When h1 → ∞ and h2 → ∞, the interaction potential between


the two semi-infinite bodies is
2πCvdw ρ1 ρ2 1 A12
W (l) = 2
= (3.8)
24 l 12πl2
where A12 = π 2 Cvdw ρ1 ρ2 is the Hamaker constant quantity.
In fact, for an interface, there is still a molecular distance for two
contacting semi-infinite bulks. If l represents d11 , which is the dis-
tance between molecules on the interface surface of two semi-infinite
object bulks (bodies) whose materials are the same, the interaction
potential in the expression above will change into (absolute value is
used here for the convenience of explanation)

2πCvdw ρ1 ρ1 1 A11
W1 (d) = 2 = (3.9)
24 d11 12πd211

where A11 = π 2 Cvdw ρ21 is the Hamaker constant quantity of object 1.


This interaction potential is equal to the cohesion work of the body.
In terms of the definition of surface free energy (surface tension),
when the bulk (body) is separated along an interface, the work
for overcoming the cohesion is equal to the interaction potential.
Since two surfaces will be formed after the separation, the poten-
tial energy in each surface is equal to half of the interaction potential.
48 Capillary Mechanics

Therefore, the surface free energy (surface tension) of the object is


equal to half of the interaction potential, so
W1 (d) A11
γ1 = = (3.10)
2 24πd211
If l represents d12 , which is the distance between molecules on the
interface of the two phases, and the materials of the two semi-infinite
bodies are different, the interaction potential in equation (3.9) will be
2πCvdw ρ1 ρ2 1 A12
W12 (d) = 2 = (3.11)
24 d12 12πd212
This interaction potential energy is equal to the adhesion work
between the two phases. Since the adhesion work has the following
relation with surface tension (free energy) and interfacial tension
(free energy),

W12 = γ1 + γ2 − γ12 (3.12)

where γ12 is the interfacial tension (free energy), the interfacial


tension (free energy) can be described as
 
1 A11 A22 A12
γ12 = γ1 + γ2 − W12 = + 2 − 2 (3.13)
12π d211 d22 d12
where d11 is the distance between molecules in object 1, d22 is the dis-
tance between molecules in object 2, and d12 is the distance between
molecules on the interface of object 1 and object 2. A11 is the
Hamaker constant quantity of object 1, A22 is the Hamaker constant
quantity of object 2, and A12 is the Hamaker constant quantity of
the subject between object 1 and object 2.

3.3 Concept of Interfacial Tension

Similarly to the concept of surface free energy, interfacial free energy


exists at the interface. Similar to surface tension, there is a force at
the interface that creates a tendency for tension, known as interfacial
tension. This tension arises from the asymmetric attractive forces on
either side of the interface. The mechanics definition of interfacial
tension is as follows: interfacial tension is a force per unit line length
Surface Tension and Interfacial Tension 49

acting at the interface, which is vertical to this straight line and


parallel to the interface. The energy definition of interfacial tension
is as follows: it is the increment of interfacial free energy resulting
from an increase in per unit area of the interface. When one of the
phases is a gas, the interfacial tension will be surface tension.

3.4 Factors Influencing Surface Tension

The surface tension is a property of liquids (and, in general, solids)


that reflects their strength. Several factors influence surface ten-
sion, including the type of material, its density, and environmental
conditions such as temperature and pressure.

3.4.1 Influence of the type of matter on surface


tension
Surface tension is related to the type of matter because it arises from
the net attractive forces generated by the asymmetric cohesive effects
of surface molecules. This phenomenon depends on the attractive
forces between the molecules and their structural characteristics. For
example, a water molecule is a polar molecule, which exhibits strong
attractive forces between its molecules. Under normal atmospheric
pressure, the surface tension of water can be up to 72.75 mN · m−1
at 20◦ C. However, under the same conditions, the surface tension
of non-polar hexane molecules is only 18.4 mN · m−1 . Hydrargyrum
has a huge cohesion force. Therefore, at room temperature, hydrar-
gyrum has the greatest surface tension (γHg = 485 mN · m−1 )
among all kinds of liquids. Of course, other kinds of molten met-
als have great surface tension (generally, the data are collected in
the high-temperature molten state). For example, the surface ten-
sion of molten copper can be up to 879 mN · m−1 at 1100◦ C. The
surface tensions of some materials against that of air are listed
in Table 3.1.

3.4.2 Influence of interface


The surface tension of a liquid is typically defined as the value
measured when the liquid is in contact with air (including its vapor).
50 Capillary Mechanics

Table 3.1. Surface tension of different materials under different temperatures.

Surface tension of various liquids in dyn/cm against air


Mixture %’s are by mass dyne/cm is also called mN/m
(milli-Newton per meter) in S.I. units

Temperature Surface tension


Liquid (◦ C) γ(mN/m)

Acetic acid 20 27.6


Acetic acid (40.1%) + Water 30 40.68
Acetic acid (10.0%) + Water 30 54.56
Acetone 20 23.7
Diethyl ether 20 17.0
Ethanol 20 22.27
Ethanol (40%) + Water 25 29.63
Ethanol (11.1%) + Water 25 46.03
Glycerol 20 63
n-Hexane 20 18.4
Hydrochloric acid 17.7 M aqueous solution 20 65.95
Isopropanol 20 21.7
Mercury 15 487
Methanol 20 22.6
n-Octane 20 21.8
Sodium chloride 6.0 M aqueous solution 20 82.55
Sucrose (55%) + Water 20 76.45
Water 0 75.64
Water 25 71.97
Water 50 67.91
Water 100 58.85

Surface tension varies with changes in the properties of the matter


when it interacts with the liquid. Antonoff discovered that the inter-
facial tension between two types of liquids is equal to the difference
in their surface tensions when they are mutually saturated, even if
their mutual solubility is minimal. That is,

γ1,2 = γ1 − γ2 (3.14)

where γ1 and γ2 are the surface tensions of the two kinds of liquids
which are mutually saturated. The state of mutual saturation is that
when liquid 1 is saturated by the vapors of liquid 2, and liquid 2 is
saturated by the vapors of liquid 1. This is called the Antonoff law.
Surface Tension and Interfacial Tension 51

Table 3.2. Interfacial tension between organic liquids and water (mN · m−1 ).

Surface tension Interfacial tension

Water Organic Pure


layer liquid layer organic Calculate Measurement Temperature
Liquid σ1 σ2 liquid value value (◦ C)

Benzene 63.2 28.8 28.4 34.4 34.4 19


Aether 28.1 17.5 17.7 10.6 10.6 18
Chloroform 59.8 26.4 27.2 33.4 33.3 18
Carbon 70.9 43.2 43.4 24.7 27.7 18
tetrachloride
Pentanol

26.3 21.5 24.4 4.8 4.8 18
5%Pentanol
41.4 28.0 26.0 13.4 16.1 17
95%Benzene

The interfacial tensions between common organic liquids and water


are shown in Table 3.2.
In the interface between a liquid and a gas, surface tension is
the sum of the static attractive forces generated by the interactions
among liquid molecules. In this context, the attraction between liq-
uid molecules and air molecules can be disregarded. However, at the
interface between two different liquids, the molecules of each liquid
attract one another, which reduces the static attractive force of each
type of liquid. Consequently, the tension at the new interface is lower
than that of the greater tension of the two original surfaces. In an
extreme case, when the two types of liquids are in the same phase, the
asymmetry disappears. This leads to symmetrical attraction, result-
ing in a static attractive force of zero, as well as an interfacial tension
that is also zero.

3.4.3 Influence of temperature


Since temperature directly affects the activities of the molecules,
it directly affects surface tension. When the temperature rises, the
material will expand, the distances between molecules will extend,
and the surface tension will decrease. The linear relation between
temperature and surface tension is shown in Figure 3.5. When the
temperature is rising up to the critical point tc , the interface between
the liquid and the gas will disappear, and the surface tension tends to
52 Capillary Mechanics

Figure 3.5. γ − t curve of CCl4 .

zero. In fact, according to the analysis of the influence from density


in the following section, the rising temperature leads to a decrease in
the difference of density between the liquid phase and the gas phase.
Since the cohesion attraction is proportional to the high power of
density, when temperature rises, both the densities and the cohesion
attraction of the liquid and gas phases will be close to each other.
Then, the surface tension will be zero.
According to their research, many scholars indicate that there is
a linear relation between surface tension and temperature. They also
think that the rate slope of the linear relation is the changed value of
entropy when the surface changes within a unit area. Currently, the
descriptions of the relationship between surface tension and temper-
ature are expressed through empirical formulas, the most notable of
which is the Eötvös rule.
The Eötvös rule, named after the Hungarian physicist Loránd
(Roland) Eötvös (1848–1919), can predict the surface tension of any
kind of liquid or substance at any temperature. The prerequisites
for this prediction include knowledge of the liquid’s density, molar
mass, and critical temperature, as well as the understanding that the
surface tension approaches zero at the critical point.
Eötvös rule is based on two key assumptions:

1. Surface tension is a linear function of temperature. This


assumption holds true for most types of liquids. The relationship
Surface Tension and Interfacial Tension 53

is represented by a straight line. At the intersection of this line


and the temperature axis, surface tension reaches zero, and the
corresponding temperature at this point is known as the critical
temperature. Since the critical temperatures of various liquids dif-
fer, the Eötvös rule also illustrates the relationship between the
surface tensions of different types of liquids.
2. Data on the relationship between temperature and sur-
face tension can be represented by a curve, provided that the
density, molar mass, and critical temperature of the liquid are
known.
Let V be the molar volume and Tc be the critical temperature
of a liquid. The surface tension of the liquid is

γV 2/3 = k(Tc − T ) (3.15)

where k is the Eötvös constant value for any kind of liquid and
has a value of 2.1 · 10−7 J/(K·mol2/3 ).
Since the intersection of the curve and the temperature axis
consistently occurs 6 K (thermodynamic temperature) earlier than
the critical point in experiments, Ramsay and Shields modified
expression (3.15) to accurately describe the relationship between sur-
face tension and temperature. They achieved this by shifting the
expression 6 K to the left along the temperature axis:

γV 2/3 = k(Tc − T − 6.0) (3.16)

The molar volume V can be given by the molar mass M and the
density ρ as V = M/ρ.
γmol = γV 2/3 in the above equation can be called the molar
surface tension. Substituting it into the expression above leads to
 −2/3
M
γ=k (Tc − 6.0 − T ) (3.17)
ρ
Introducing the Avogadro constant NA , the above equation can
be rewritten as
   2/3
 M −2/3  NA
γ=k (Tc − 6.0 − T ) = k (Tc − 6.0 − T )
ρNA V
(3.18)
54 Capillary Mechanics

Table 3.3. Some liquid surface tension values at different


temperatures (mN · m−1 ).

Liquid 0◦ C 20◦ C 40◦ C 60◦ C 80◦ C 100◦ C

Water 75.64 72.75 69.56 66.18 62.61 58.85


Ethanol 24.05 22.27 20.60 19.01 — —
Toluene 30.74 28.43 26.13 23.81 21.53 19.39
Benzene 31.6 28.9 26.3 23.7 21.3 —

As John Lennard-Jones and Corner showed in 1940, by means of


statistical mechanics, the constant k  is nearly equal to the Boltz-
mann constant.
Some of the values of liquid surface tensions at different
temperatures are listed in Table 3.3.

3.4.4 Influence of density


The difference in densities on either side of the interface results in
asymmetric attractions, which contribute to the formation of surface
tension or interfacial tension. Consequently, density can influence the
magnitude of surface tension. The relation between surface tension
and density is
γ = B(ρl − ρ0 )4 (3.19)
where ρ1 is the density of the liquid phase of the liquid and ρ0 is
the density of the gas phase of the liquid. B is a constant, which has
nothing to do with the density. This equation is derived from the
parachor of interface chemistry.
When the temperature is considered in the relation between sur-
face tension and density, the equation above can be written as
γ = const(Tc − T )p (ρl − ρ0 )q (3.20)
where p and q can be obtained using “least squares” based on
experiments. Expression (3.20) is the expansion of expression (3.19).

3.4.5 Influence of pressure


The pressure of the gas phase should influence surface tension, given
the density difference between gas and liquid, as well as the static
Surface Tension and Interfacial Tension 55

attractive forces involved. Since the vapor pressure of a liquid remains


constant at a specific temperature, the effect of pressure can only be
observed by altering the pressure of air or inert gases. However, it
is important to note that air and inert gases can dissolve in and be
absorbed by the liquid. Part of the air is attracted to the surface of a
liquid. The rate of dissolution and the amount of absorption vary with
changes in pressure. Consequently, the method of altering the pres-
sure of air or inert gases to measure surface tension involves multiple
influencing factors, including dissolution, absorption, and pressure.
Some researchers have investigated the impact of pressure on surface
tension. For example, the surface tension of water is 72.82 mN·m−1
under a pressure of 0.098 MPa and 66.43 mN·m−1 under a pres-
sure of 9.8 MPa. The surface tension of benzene is 28.85 mN·m−1
under a pressure of 0.098 MPa and 21.58 mN·m−1 under a pressure
of 9.8 MPa. Therefore, surface tension decreases as pressure increases.
However, the effect of pressure on surface tension is not as pro-
nounced as that of temperature. The influence on surface tension
is minimal when the pressure changes only slightly.

3.5 Methods of Measurement of Surface


and Interfacial Tensions

There are many methods for measuring surface tension, including


the capillary rise method, Du Noüy ring method, maximum bubble
pressure method, Wilhelmy plate method, spin drop method, droplet
volume method, droplet shape analysis method, experimental ink
method, metal rod method, and stalagmometeric method.
These methods can measure surface (interface) tension in various
environments. The Du Noüy ring method is a classic technique for
measuring surface tension, even in situations where wetting is chal-
lenging. The principle involves extracting a liquid film (similar to a
soap bubble) from the liquid using a ring that is initially immersed
in the liquid while measuring the force required to lift the ring. The
maximum pressure bubble method is particularly effective for assess-
ing the variation of surface tension over time, as it measures sur-
face tension by determining the highest pressure of the bubbles. The
Wilhelmy plate method is a versatile technique, particularly well-
suited for the long-term measurement of surface tension. This method
56 Capillary Mechanics

quantifies surface tension by assessing the force exerted on a flat


plate that is positioned perpendicular to the liquid surface during
the wetting process. Additionally, it can be employed to measure
interfacial tension. The rotating droplet method is another approach
that can be utilized to determine interfacial tension, making it espe-
cially effective in scenarios involving low or very low tension. The
measured value refers to the diameter of the rotating droplet. The
droplet shape analysis method is effective for measuring interfacial
tension and surface tension, and it can be applied at very high pres-
sures and temperatures. Surface tension, or interfacial tension, of a
liquid can be determined by analyzing the geometric shape of the
droplet. The droplet volume method is particularly well-suited for
dynamically measuring interfacial tension. The measured value is
calculated by dividing the number of droplets by a specific volume
of liquid.
Due to the numerous methods available, the following presents
only a brief introduction to several commonly used measurement
techniques.

3.5.1 Capillary rising method


This is one of the simplest methods. When a clean glass capil-
lary tube is placed in a liquid, if the liquid wets the wall of the
capillary tube, due to surface tension, the liquid will rise along
the capillary tube until the rising force (2πr cos θ · γ) is balanced
by the gravity of the liquid (πr 2 ρgh), as shown in Figure 3.6,

Figure 3.6. Surface tension measured by capillary rising method.


Surface Tension and Interfacial Tension 57

i.e., 2πr cos θ · γ = πr 2 ρgh, or

ρghr
γ= (3.21)
2 cos θ

where h is the liquid rising height. γ is the surface tension, ρ is


the density of the liquid, r is the radius of the capillary, g is the
acceleration due to gravity, and θ is the contact angle. Since ρ, g,
and r are known, surface tension can be calculated by measuring the
height to which the liquid rises.
For complete wetting, θ = 0◦ , we have γ = ρghr
2 .

3.5.2 Du Noüy ring method


The Du Noüy ring method is a traditional technique used to measure
the surface tension of a liquid. In this method, a platinum wire ring
is suspended from a torque scale. As the torque wire of the scale is
gradually turned, the ring slowly lifts from the surface of the liquid,
creating a circular film of liquid, as illustrated in Figure 3.7. The force
required to raise the ring from the liquid’s surface is measured and is
directly related to the surface tension of the liquid. A measurement
system for the Du Noüy ring method is depicted in Figure 3.7.
The Du Noüy ring method consists of eight steps, as illustrated
in Figure 3.8.

Figure 3.7. Surface tension measured using the Du Noüy ring method.
58 Capillary Mechanics

Figure 3.8. The Du Noüy method and the process of measurement.

(1) When the ring is positioned above the liquid surface, the
measurement instrument registers a reading of zero.
(2) When the ring approaches the liquid surface, a slight attractive
force exists between the ring and the surface due to adhesive
attraction.
(3) The ring experiences an upward reverse force due to the slight
action of surface tension. Consequently, the pulling force mea-
sured by the instrument is negative.
(4) When the ring immerses in the liquid surface, surface tension will
dissipate. The weight of the ring overcomes buoyancy, resulting
in a downward force transmitted through the wire. Consequently,
a slight positive pull is detected by the instrument.
(5) When the ring is pulled, it will start to measure the pulling force.
(6) The force measured by the instrument increases as the ring is
pulled upward.
(7) The state indicates that the maximum force is approaching.
(8) After reaching the maximum force, the force will experience min-
imal attenuation until the liquid film is ruptured.

The data curve of the pull is shown in Figure 3.9 corresponding


to the eight steps above.
When the ring suddenly detaches from the surface film (while
maintaining the metal lever in a horizontal position), the maximum
pulling force F is equal to both the weight of the liquid being lifted,
mg, and the surface tension acting around the ring. Since the liquid
film has two sides, both inside and outside, the total circumference
Surface Tension and Interfacial Tension 59

Figure 3.9. Curve of the measurement data.

of the ring is 4πR:


F = mg = 4πR · γ (3.22)
where m is the mass of the liquid which was pulled. R is the mean
radius of the ring. The inner radius of ring is R . The radius of
platinum wire is r. The mean radius of ring is R = R + r, which is
shown in Figure 3.7. From equation (3.22), it can be obtained that
F
γ= (3.23)
4πR
Therefore, if F is measured, γ can be obtained. However, since
the pulled-up liquid is not cylindrical, equation (3.23) needs to be
multiplied by a factor f :
F
γ= ×f (3.24)
4πR
3
Many experiments indicate that the factor f is a function of RV
and Rr , where V is the volume of the liquid pulled by the ring. V
can be derived from the equation F = mg = V ρg. The value of f can
be found from Figure 3.10. The accurate value can be verified from
related monographs.

3.5.3 Maximum bubble pressure method


One of the effective methods for determining dynamic surface tension
is the “maximum bubble pressure method”, or simply the “bubble
60 Capillary Mechanics

Figure 3.10. Curve of factor f .

Figure 3.11. Equipment drawing of the maximum bubble pressure method.

pressure method”. The equipment used for this method are illus-
trated in Figure 3.11. A bubble pressure tensiometer generates gas
bubbles (e.g., air) at a constant rate and introduces them through a
capillary tube that is submerged in the sample liquid, with the radius
of the capillary already known. As the pressure inside the gas bubble
continues to increase, the maximum pressure is achieved when the
bubble takes on a completely hemispherical shape, and its radius,
Surface Tension and Interfacial Tension 61

Figure 3.12. The change in the curvature radius of a bubble.

R, corresponds precisely to the radius of the capillary, r, as depicted


in Figure 3.12. The height of liquid in the pressure meter is h when
the pressure difference between the inside and outside of the bubble
reaches its maximum, Δpmax , which can be written as
Δpmax = ρgh (3.25)
The Δpmax is proportional to surface tension of the liquid and
inversely proportional to the curvature radius, as given by
γ
Δpmax ∝ (3.26)
r
or

Δpmax = (3.27)
r
The constant quantity K is equal to 2, which is determined using
the Young–Laplace equation:

Δpmax = = ρgh (3.28)
r
or
r
γ = ρgh (3.29)
2
When two kinds of liquids, whose surface tensions are γ1 and
γ2 , are measured by the same capillary and pressure meter and the
corresponding heights of the liquids are h1 and h2 , it can be obtain
from equation (3.29) that
γ1 h1
= (3.30)
γ2 h2
Therefore, when K is unknown, the surface tension of the sample
liquid can be obtained from the liquid whose surface tension is known.
62 Capillary Mechanics

This method is widely utilized due to its advantageous charac-


teristics. For instance, the measurement process is brief and employs
simple equipment, and it does not require information about the
contact angle or the density of the liquid.

3.5.4 Wilhelmy plate method


A Wilhelmy plate is a thin device used to measure the equilibrium
surface or interfacial tension at the interface of a gas–liquid or liquid–
liquid phase. The Wilhelmy plate typically measures a few square
centimeters and is often constructed from glass or platinum, which
is roughened to ensure complete wetting. In this method, the plate
is oriented perpendicular to the interface, and the force exerted on
it is measured, as illustrated in Figure 3.13.
When a plate is immersed in a liquid, the liquid will either ascend
(in the case of a lyophilic liquid) or descend (in the case of a lyopho-
bic liquid) along the vertical wall of the plate. The Wilhelmy plate
measures the force FW exerted by the liquid on the plate, which can
be either a pull or push force, to determine the contact angle or sur-
face tension. When the system reaches equilibrium, the interface of
the gas–liquid–solid phases will achieve a mechanical balance:

Fw = γLV C · cos θ − Fb (3.31)

where γLV is the surface tension of the liquid, C is the length of line
where the plate comes into contact with the liquid surface, θ is the
contact angle, and Fb is the floatage which exists when the liquid

Figure 3.13. Wilhelmy plate.


Surface Tension and Interfacial Tension 63

exerts on the plate. Since the two surfaces are wet, C is equals to 2l.
l is the width of the plate.
If the surface tension γLV is known, equation (3.31) can be
changed into

Fw + Fb
cos θ = (3.32)
γLV · C

so as to derive the contact angle θ. If the contact angle θ is known,


equation (3.31) can be changed into

Fw + Fb
γLV = (3.33)
cos θ · C
so as to get the surface tension γLV .

3.5.5 Stalagmometeric method


The stalagmometric method involves using a stalagmometer to
measure surface tension. The stalagmometer, also known as a stac-
tometer or stalogometer, is a capillary glass tube with a widened mid-
dle section, as illustrated in Figure 3.14. The bottom of the device is
narrowed to allow the fluid to exit the tube in the form of a drop. The
design of the stalagmometer ensures that the volume of each drop
is consistent. During experiments, drops of the specific fluid flow
slowly from the tube in a vertical direction. The drops that hang
from the bottom of the tube begin to fall when their volume reaches
a maximum value, which is determined by the characteristics of the

Figure 3.14. Stalagmometer.


64 Capillary Mechanics

solution. At this moment, the weight of the drops is in equilibrium


with the surface tension. Based on Tate’s law,

mg = 2πrγ (3.34)

The drop falls when the weight (mg) is equal to the circumference
(2πr) multiplied by the surface tension (γ).The surface tension can
be calculated when the radius of the tube (r ) and the mass of the
fluid droplet (m) are known. Additionally, when the surface tension is
proportional to the weight of the droplet, a reference fluid — typically
water — can be used for comparison with the target fluid:
m1 m2
= (3.35)
γ1 γ2
In the equation, m2 and γ2 are the mass and the surface tension
of the reference fluid, while m1 and γ1 are the mass and the surface
tension of the target fluid. If we take water as a reference, then
m
γ1 = γH2 O · (3.36)
mH2 O
If the surface tension of water is known, we can calculate the
surface tension of the target fluid using the equation. The more accu-
rately the weights of the drops are measured, the more precisely we
can determine the surface tension from equation (3.36).
The equipment used in this method to measure the weight of
the liquid includes a stalagmometer, a beaker, a spherical tube, a
weighing bottle with a lid, a weighing scale with a precision of one-
thousandth, distilled water, a fluid with unknown surface tension,
and a thermometer. The steps of the experiment are as follows:
(1) Place a clean and dry stalagmometer vertically on the experi-
mental plate.
(2) Weigh the mass of the weighing bottle.
(3) Obtain some distilled water using a beaker. Cover the top of
the stalagmometer with a rubber tube. Introduce water into the
stalagmometer until the water’s surface reaches the wider middle
section of the stalagmometer.
(4) Remove the rubber ball to allow the droplet to fall freely, and
collect 20 drops in the weighing bottle.
(5) Measure the weight of the weighing bottle containing 20 drops.
Surface Tension and Interfacial Tension 65

(6) Please empty the weighing bottle and dry it for the next
measurement.
(7) Change to the target liquid and repeat Steps 2 through 6.
(8) Measure the temperature of the laboratory using a thermome-
ter, and examine the surface tension of water at the specified
temperature. Then, substitute this value into equation (3.36)
to calculate the surface tension of the target liquid at that
temperature.
The equipment used in this method for counting droplets include
a stalagmometer, a beaker, a spherical tube, a weighing bottle with
a lid, a precision balance with a thousandth of a gram accuracy,
distilled water, a fluid with unknown surface tension, and a ther-
mometer.
There are two indicator lines on the stalagmometer. The upper
line is located above the wide middle section, while the lower line is
positioned below it. The volume between these two lines is V. The
density of liquid is ρ. The mass is
m=V ·ρ (3.37)
Set the two kinds of dropping liquids in the same volume. The
number of drops n is accounted for by the volume V. Then, the mass
of each drop is
m V ·ρ
m̄ = = (3.38)
n n
The mass in equation (3.36) is replaced with m̄. Then, the surface
tension of the target liquid is
ρ nH2 O
γ1 = γH2 O · · (3.39)
ρH2 O n
The steps of this experiment are as follows:
(1) Pour distilled water into the stalagmometer until the water level
reaches the upper line. Open the stalagmometer to allow the
water to drop into the weighing bottle until the water level
reaches the lower line. Close the stalagmometer and record the
number of drops of water.
(2) Empty the weighing bottle and dry it thoroughly for the next
measurement.
66 Capillary Mechanics

(3) Change to the target liquid and repeat Steps 1 and 2.


(4) Measure the temperature of the laboratory using a thermometer
and assess the surface tension of water at the specified
temperature. Substitute this value, along with the densities of
water, into equation (3.39) to calculate the surface tension of
the target liquid.

3.6 Methods of Measuring the Surface Tension of Solids

The surface tension of a liquid is characterized by an increase in the


surface area. In contrast, the surface tension of a solid is influenced
by its surface properties. However, solids differ significantly from liq-
uids; the atoms and molecules in a solid cannot move freely as they
do in a liquid. Consequently, there is currently no direct and reli-
able method for measuring the surface tension of solids. Instead, an
indirect method is typically employed to assess the surface tension
of solids, or it can be estimated using theoretical approaches.
In the early years, Soviet scientists used a method in which a
crystal is struck with a blade. The work done by splitting the crys-
tal can be determined when the crystal splits and a new surface is
formed. For example, in this method, the measurement value of NaCl
is 150 mN · m−1 . This result is consistent with the calculation of crys-
tal lattice. Two methods which are commonly used are introduced
in the following sections.

3.6.1 Critical surface tension method


Currently, the critical surface tension method is a common method
which is widely used to measure the surface tension of solids.
In the early 1960s, Zisman and other scientists discovered that
when a series of liquids is applied to a macromolecule, the surface
tension of the liquid is proportional to the cosine of the contact angle
θ, as illustrated in Figure 3.15. The surface tensions of the series of
liquids are known, and the surface tension of the solid surface is
relatively low.
If the line in Figure 3.15 is extended to cos θ = 1(θ = 0◦ ), the cor-
responding surface tension of the liquid is the critical surface tension
γc . For example, the surface tension of polythene (γc ) is 31mN · m−1 .
Surface Tension and Interfacial Tension 67

Figure 3.15. Zisman curve of polythene surface.

Table 3.4. Critical surface tension of some kinds of solids.

Perfluorinated
dodecanoic Napht- Hexa- Poly- Poly-
Solid acid PTFE halene decane ethylene styrene PVC Nylon

γc 6 18 25 29 31 33–43 39 42–46

The physical meaning of γc is that if the surface tension of it (γ)


is smaller than γc , the liquid can wet the surface of polythene. On
the contrary, it cannot wet the solid surface. The critical surface ten-
sions of some kinds of solids are listed in Table 3.4. The γc increases
according to the increase of the surface polar.

3.6.2 Calculating the surface tension of solid


utilizing the surface tension of high polymer
liquid or melt
In general, the surface tension of a liquid decreases as the
temperature increases. The surface tension of high polymer melts
is first measured at various temperatures and then extrapolated
68 Capillary Mechanics

Figure 3.16. γ−T curve of some kinds of high polymers.

to determine the surface tension of the corresponding solid state


at a specific temperature. For example, the relationships between
the surface tension and temperature of L-PE, PDMS, and PIB are
illustrated in Figure 3.16.
According to the figure above, if the γ−T curve is extended to
20 C, γ of L-PE is 35.7 mN · m−1 , that of PIB is 34.0 mN · m−1 , and

that of PDMS is 19.8 mN · m−1 . Obviously, when using this method,


the effects of phase change on the surface tension must be noted.
Although this influence is always considered very small, the result of
this method is approximate.

3.6.3 Estimation method


A solid is composed of atoms or molecules that occupy fixed posi-
tions. In theory, if the forces within the crystal lattice of the solid
are understood, the surface tension of the solid can be calculated.
For instance, if the crystal structure of a metal and the coordination
number of the surface atoms are known, one can determine the super-
ficial energy of each atom as well as the superficial energy per unit
area (i.e., the surface tension). The coordination number of surface
Surface Tension and Interfacial Tension 69

atoms refers to the number of atoms surrounding a single atom in


the superficial layer, excluding the influence of atoms in the second
layer. In this method, the surface tension of Cu is 1.5N·m−1 , which is
consistent with the test value. However, this method is not suitable
for ionic crystals.

3.7 Example of Surface Tension Measurement with


Interfacial Tension Instrument ZYW-200B

ZYW-200B is an automatic-control instrument designed for


measuring interfacial tension and surface tension, utilizing the
Du Noüy ring method. A key feature of this instrument is its reliance
on a physical measurement approach rather than traditional chemical
methods, making it user-friendly. With ZYW-200B, users can quickly
and accurately measure the surface tension or interfacial tension of
various liquids. Additionally, the measurement results are automat-
ically displayed and recorded, and the measurement curve can be
easily stored or printed.

3.7.1 Structure and working principle


This instrument consists of a torque wire, a differential transformer, a
platinum ring, and a lifting mechanism, as illustrated in Figure 3.17.
Its operating principle is as follows.
When measuring surface tension, the platinum ring is first
immersed in the target liquid. Then, the filled glassware on the salver

Figure 3.17. Automatic tension device and its schematic diagram.


70 Capillary Mechanics

is lowered by the transmission system. As the glassware descends,


the film between the platinum ring and the liquid is stretched. Con-
sequently, the platinum ring experiences a downward force, causing
the torque wire to rotate through a lever mechanism. As a result, the
magnetic core of the differential transformer ascends while the plat-
inum ring descends. The coil of the differential transformer induces
a specific voltage. Therefore, the amount of film deformation is con-
verted into voltage, which is then translated into the corresponding
tension through computer calculations and displayed automatically.
As the film is gradually stretched, the tension increases until the film
ruptures. During this procedure, the maximum value is the measure-
ment value of the tension γ  . While this value is multiplied by the
liquid correction factor F. Then, the actual tension value γ will be
obtained by γ = γ  · F . The factor is determined by γ  , the density
of the liquid, the radius of the platinum wire, and the radius of the
platinum ring.

3.7.2 Preparing for the experiment


1. Put the instrument on a stable table board carefully. Keep it hor-
izontal by adjusting the knob on the base. Keep the room tem-
perature at 20 ± 5◦ C.
2. Open the top lip and check out the following things:

• The magnetic core of the differential transformer should be in


the normal position, where it cannot come into contact with
the wall but hangs freely.
• The torque wire should be tensioned. If not, adjust the bolts at
the two ends.
• The level arm should be kept horizontally. If not, adjust the
two ends of the torque wire.
• After checking, close the upper lip.

3. Insert the power plug into an external socket reliably grounded.


Turn on the power button and stabilize it for 30 minutes. Clean
the platinum ring and the glass. First, the platinum ring is washed
with petroleum and then with acetone. Next, the ring is heated
and dried in the oxidation flame of a gas lamp or an alcohol lamp.
When dealing with the platinum ring, avoid deforming the ring.
Surface Tension and Interfacial Tension 71

Figure 3.18. The software interface.

Then, hang the ring on a small hook of the lever arm. When it is
stabilized, start the experiment.
4. Run the program and enter the experimental preparation
interface, as shown in Figure 3.18. After inputting the parameters
and confirming, the main interface will appear. At this point, the
machine automatically becomes zero, and the adjustment range is
±10.0 mN/m. If it is in this range, the experiment can be started.
If not, the screen will automatically display a message. Open the
cover and adjust the lift rod until it reaches the adjustment range
(preferably transferred into a value of zero or very small). Then,
press the zero key to set zero.

3.7.3 Calibrating the instrument


The mass method is used to calibrate the instrument. The procedure
of calibration is described as follows.
A small piece of paper is placed on the platinum ring after the
instrument is installed. Click on the instrument calibration option in
the main interface menu. A measurement value is displayed. If the
value is not zero after stabilization, click the zero button. Then, put
a weight of 1000 mg on the paper. At this point, the value displayed
72 Capillary Mechanics

should be 81.7 mN/m (the relative error of the indicated value is


±1%). Then, click the menu for the actual value of the weight and
select 1000. If the green light is on after stabilization, click the exit
button and return to the main interface to measure the surface ten-
sion of the liquid. If the red light is on, click the calibration until
the green light is on. Click the exit button and return to the main
interface to measure the surface tension of the liquid.
Note: Generally, if the tension is in the allowable error range, the
instrument does not need to be adjusted. When the instrument is
used for the first time or is stored for a long period of time before
using, the test error is considered to be bigger. Therefore, the instru-
ment should be calibrated.

3.7.4 Measurement of surface tension and


interfacial tension of liquid
3.7.4.1 Measurement of surface tension of liquid
(1) Pour the liquid sample, which is adjusted to 25◦ C, into a beaker,
and keep the height of the liquid surface at 20–25 mm. Then, put
the beaker at the center of the tray, as shown in Figure 3.19.

Figure 3.19. The flying ring and the beaker.


Surface Tension and Interfacial Tension 73

(2) Adjust the zero. The software interface is shown in Figure 3.18.
The adjustment is done by clicking the “zero” button. This step
is only done before the first test when conducting many tests.
(3) By clicking the “up” button, the platinum ring will come into
contact with the liquid. When the ring becomes immersed in the
liquid to about 5–7 mm, the “stop” button is clicked.
(4) Click the “experiment start” button. The experiment begins and
the tray and the target liquid begin to descend. The real curve
is drawn and the maximum surface tension is displayed by the
software, as shown in Figure 3.20.
(5) When the experiment is finished, the software will end the
experiment automatically. It can also be ended by clicking the
“experiment end” button.
(6) Repeat Step 3 for the next experiment. The software of this
instrument allows one to conduct eight experiments for one sam-
ple continuously and process the data.
(7) Click the “save” button to save the experiment data. Click the
“browse data” button to check or print the data in the experi-
ment data window.

Figure 3.20. Real curve drawn by ZYW-200B.


74 Capillary Mechanics

3.7.4.2 Measurement of the interface tension of liquid (taking


the interfacial tension of petroleum products to water
as an example)
(1) Pour quantificational distilled water at 25◦ C in a beaker. Keep
the height of the liquid surface at 20–25 mm. Then, place the
beaker at the center of the tray.
(2) Lift the tray so that the platinum ring can immerse into the
liquid with a height of 5–7 mm.
(3) Pour the sample liquid, which has been adjusted to 25◦ C, into
the distilled water. The height of the sample is about 10 mm.
Note that the sample liquid cannot come into contact with the
ring, and the ring cannot come into contact with the interface of
oil–water. Keep the surface of oil–water for 30 s.
(4) Click the “experiment start” button. The experiment begins and
the tray and the target liquid begin to descend. The real curve
is drawn and the maximum surface tension is displayed by the
software.
(5) When the experiment is finished, the software will end the exper-
iment automatically. It can also be ended by clicking the “exper-
iment end” button.
(6) Repeat Step 3 for the next experiment. The software of this
instrument allows one to conduct eight experiments for one sam-
ple continuously and process the data.
(7) Click the “save” button to save the experiment data. Click the
“browse data” button to check or print the data in the experi-
ment data window.

3.7.5 Correcting the measured tension


In the Du Noüy ring method of measurement, two aspects need to
be considered:

(1) During the measurement, the ring is lifted so that the liquid
surface deforms. With the increase in the rising displacement of
the ring, the liquid deformation increases. Therefore, the radius
from the center to the break point is smaller than the mean
radius of the ring. This influence is indicated by the ratio of the
radius of the ring and the radius of the platinum.
Surface Tension and Interfacial Tension 75

(2) There is a little amount of liquid adhering to the bottom of the


ring. This influence can be expressed by a function.
According to the two situations above, the measured tension

γ should be multiplied by a correction factor F, which can be
derived from the actual tension γ ·γ = γ  ·F . According to ASTM
D971, the correction factor of the device can be calculated using
the equation

0.01452 × γ  ÷ C 2 ÷ (D − d) + 0.04534
F = 0.7250 + (3.40)
−1.679 × r ÷ R

where γ  is the displayed value (mN/m), C is the circumference


of the ring, R is the radius of the ring, D is the density of the
bottom phase at 25◦ C (g/ml), d is the density of the top phase
at 25◦ C (g/ml), and r is the radius of the platinum wire.
If the two measured phases are gas and liquid, D is the density
of liquid and d is the density of gas.

3.7.6 Note
If you want to check the historical curves, click the “open” button
under the “file menu” or click the “open curve” button. If you wish to
test a different sample liquid, click the “experiment prepare” button
to return to the parameter interface and input new parameters for
the new experiment.
During the experiment, the instrument should be positioned on a
vibration-free platform, preferably an isolation platform. It is essen-
tial to ensure that the instrument is leveled before commencing the
experiment. Once the experiment is complete, turn off the power,
remove the platinum ring, and clean the instrument. It is impor-
tant to note that interference from high-power electrical devices and
wireless communication equipment is prohibited during the experi-
ment. Additionally, the laboratory housing the instrument should be
equipped with temperature control systems to maintain a constant
temperature. Measures should also be taken to prevent contamina-
tion from harmful gases.
If the repeatability of the experiment is poor, there may be several
underlying reasons:
76 Capillary Mechanics

(1) The platinum ring and glass cup may not have been cleaned.
(2) The liquid may have expired.
(3) The magnetic core of the differential transformer may be in con-
tact with the wall.
(4) The torque wire may be loose.

3.8 Elasticity of Liquids

There are several methods to measure the surface tension of liquids,


including the capillary tube method, maximum bubble method, pull-
out method, pendant-drop method, and ring method. Among these,
the ring method is the most commonly used. There are specific stan-
dards that must be followed when employing this method. The ring
method involves immersing a ring in the liquid and then applying a
force to pull the ring upward until it separates from the liquid. The
surface tension is calculated as the maximum force divided by twice
the perimeter of the ring. From the measurement process, it can be
observed that surface tension represents the tensile limit or strength
of the liquid. When measuring the surface tension of a liquid using
the ring method, the force required to lift the ring increases with the
height of the liquid column being raised by the ring. Once the lifting
force reaches its maximum value, any further increase in the height of
the liquid column will cause the force to decrease rapidly, ultimately
becoming zero when the liquid column is broken. This measurement
process reveals that the force lifting the ring is directly proportional
to the displacement of the liquid being pulled. To study the char-
acteristics of a liquid, a lifting plate (silicon plate) can be employed
to conduct a pulling test with purified water. When the lifting plate
is completely wetted by the liquid, lifting the plate with an upward
force will generate a counteracting force from the liquid. This upward
force arises from the adhesion between the liquid and the solid plate,
as well as the cohesive forces within the liquid itself. The solid plate
must overcome this force to detach from the liquid surface. The mag-
nitude of this force can be quantified using a measuring device. When
lifting the plate upward, a specific height of liquid is drawn up, form-
ing an inward-curving liquid column. When the height of the lifted
liquid column is minimal — prior to the lifting force reaching its max-
imum value — the liquid exhibits properties akin to solid elasticity.
Surface Tension and Interfacial Tension 77

This liquid column undergoes repeated deformation through loading


and unloading, a phenomenon referred to as liquid elasticity. This
behavior is also observed during the measurement of surface tension
using the ring method.

3.8.1 Linearity in liquid tensile process


First, when doing the liquid tensile test, use the lifting plate to pull
the liquid. Test the relationship between the force for lifting the liq-
uid column and the height of the pulled-up liquid column. Before
the lifting force reaches its maximum value, the height of the lifted
liquid column is very small. If using d to represent the height of the
pulled-up liquid and f to represent the corresponding lifting force,
the lifting force for the first height ds1 is fs1 . In the same way, the lift-
ing force for the second height ds2 is fs2 , the lifting force for the nth
height dsn is fsn . By marking the points on the coordinates, the previ-
ous points (ds1 , fs1 ), (ds2 , fs2 )· · · (dsn , fsn ) appear in a perfect linear
relationship and satisfy the linearity equation f = a + kd. Upon
unloading the plate, the liquid column declines to the mth height,
the corresponding force is fjm. By marking the points on the coordi-
nates, the points (dj1 , fj1 ), (dj2 , fj2 )· · · (djm , fjm ) also appear in a
perfect linear relationship and satisfy the previous linearity equation
law, and every parameter is the same as the above, which means
the tensile force of the liquid and the height of the lifting liquid col-
umn appear to have a linear relationship. The curve between the
tensile force and displacement is shown in Figure 3.21. The silicon
plate is in the size of 17.725 mm · 17.725 mm and the liquid is puri-
fied water. The test has been done six times and shows that before
reaching the maximum force, the force and displacement appear
in a linear relationship, and the repeatability of each-time curve is
good.

3.8.2 Recoverability in liquid tensile process


The height of the lifted liquid column is very small (before the lifting
force reaches its maximum value), and the liquid above shows not
only the linearity property but also a good recoverability if the liquid
is pulled up by the plate repeatedly. Suppose in the process of pulling
up the liquid column, f0 is the force of lifting the plate and d0 is the
78 Capillary Mechanics

Figure 3.21. Force–displacement curve of silicon plate pulling up liquid.

corresponding height of the pulled-up liquid column. If the plate is


lifted continuously, the force will continue to increase. If the plate
goes downward to the previous height d0 by unloading (before the
tensile force reaches its maximum value), the liquid bridge force,
which is equal to the force on the lifting plate, will also return to
f0 . The same occurs with the pulling force when loading to the same
height. Therefore, this force is recoverable.
Further tests indicate that the liquid pulling force remains
constant when both the liquid column height and the plate area
are fixed, regardless of whether the system is in a loading or unload-
ing state. When the liquid column is subjected to repeated loading
and unloading, this behavior demonstrates that the liquid possesses
good elasticity. Additionally, when measuring the surface tension of
the liquid using the ring method, a similar phenomenon is observed.
Figure 3.22 shows the curve between the tensile force and displace-
ment by using a silicon plate with a size of 17.725 mm · 17.725 mm to
pull up purified water in the beaker repeatedly (loading and unload-
ing). To distinguish between the loading and unloading processes, a
Surface Tension and Interfacial Tension 79

Figure 3.22. Force–displacement measuring curve of pulling and recovering the


liquid repeatedly.

transmission gap is maintained during the conversion between these


two phases. This transmission gap results in an approximately hor-
izontal segment at the bend of the force curve, as illustrated in the
figure A control software can be utilized to eliminate the influence of
the test device gap and to generate the lifting and declining force–
displacement curve, as shown in Figure 3.23. In each stage (referring
to the distinct lifting and declining phases), not only is the force–
displacement relationship linear, but the fitted linear equations for
the different stages also share the same parameters through curve
fitting. Suppose the fitted equation for each stage of linearity is rep-
resented as Y = A + B × X, based on tests conducted with silicon
and purified water in a beaker. The OriginPro 7.5 software was uti-
lized to analyze the corresponding data for each stage, resulting in
the parameters of the linear equation, as presented in Table 3.5. In
the table, the parameters A and B are remarkably consistent, and
the dispersion error is minimal, indicating a strong linearity property.
The similarity of the parameters suggests a significant regularity in
80 Capillary Mechanics

Figure 3.23. Force–displacement measuring curve of pulling and recovering


the liquid repeatedly with device effect removed.

Table 3.5. Data list of each-stage linear fit equation of silicon plate pulling
purified water repeatedly.

A B

Value Error Value Error R SD

Lifting 1 0.56333 0.0193 2.54339 0.01128 0.99983 0.04041


Lifting 2 0.61912 0.01568 2.52129 0.00917 0.99989 0.03283
Declining 1 0.61053 0.02142 2.52807 0.01253 0.99979 0.04486
Declining 2 0.59351 0.02115 2.5317 0.01237 0.9998 0.04429
Average 0.5966 0.0194 2.5311 0.01133 0.99983 0.040598
Standard deviation 0.0246 0.0026 0.0093 0.00134 0.00004 0.004800

the data. Furthermore, the consistency observed in both the lifting


and declining processes demonstrates a high degree of recoverabil-
ity. As illustrated in the data curve in Figure 3.23, the tensile force
and displacement exhibit a strong linear relationship and excellent
recoverability during the processes of extension and recovery.
Surface Tension and Interfacial Tension 81

Figure 3.24. The maximum tensile force of liquid per unit area has nothing to
do with the area.

3.8.3 The maximum tensile force of liquid on per


unit area is independent of the area
Tests indicate that when the areas of the plates vary, the maximum
tensile force exhibits different values. Let the maximum tensile force
be denoted as fmax and the corresponding maximum height of the liq-
uid column be dmax . Figure 3.24 illustrates the relationship between
the maximum tensile force of the liquid and the area of the silicon
plate. The curve demonstrates a strong linear correlation between
the maximum force and the area (specifically, the area of the liquid
column), with minimal dispersion error.
The maximum force per unit area can be figured out by dividing
the maximum force of the liquid column area, as shown in Fig-
ure 3.25. In Figure 3.25, the curve is a horizontal line, which shows
that the maximum tensile force of liquid per unit area has noth-
ing to do with the area, which means it is independent of the
area.
82 Capillary Mechanics

Figure 3.25. The maximum tensile force of liquid per unit area has nothing to
do with the area.

3.8.4 The Hook’s law of liquid tension


In order to describe elasticity of liquids, the tensile force per unit
area is defined as liquid tension, which is represented by σl . That
is σl = f /A. The ratio between the height of the pulled-up liquid
column and the maximum height of the liquid column is defined as
the liquid tensile ratio, which could also be called the relative tensile
height. Use εl to represent the height and εl = d/dmax .
From Section 3.8.1, the tensile force is in a linear relationship
with the height of the pulled-up liquid column. Set the height of the
pulled-up liquid column as zero when f0 = 0, that is, pulling from
the origin of the coordinates. The tensile force f and the height d of
the pulled-up liquid column have a linear relationship as follows:
f = kd, d ≤ dmax (3.41)
When the tensile force reaches the maximum value fmax , the cor-
responding height of the liquid column is dmax . According to the
previous formula,
fmax = kdmax (3.42)
Surface Tension and Interfacial Tension 83

It could be seen from Section 3.8.3 that the maximum force of


liquid columns with different areas is in a linear relationship with the
areas of the liquid columns. Suppose the slope of the linear equation
is El and the area of the liquid column is A. Since the silicon does
not touch the liquid when the liquid column area is zero (A = 0)
and the tensile force is zero (f = 0), the linear equation between the
maximum force and the area can be written as follows:
fmax = El A (3.43)
According to equations (3.42) and (3.43), we have
A
k= El (3.44)
dmax
Substituting equation (3.44) into (3.41) leads to
d
f= El A (3.45)
dmax
Transform (3.45) into
f d
= El (3.46)
A dmax
Using the previous definition, equation (3.46) can be rewritten as
σl = El εl (3.47)
Equation (3.47) is an expression of the elasticity of liquids. Com-
pared with Hook’s law of solid materials, equation (3.47) could be
called the Hook’s law of liquid tension. In equation (3.47), El is equiv-
alent to the elasticity modulus. Therefore, it is called the elasticity
modulus of a liquid. Comparing with the elasticity of a solid, the ten-
sile stress of the liquid σl corresponds to the tensile stress of the solid
material. The tensile ratio of the liquid εl corresponds to the tensile
strain of the solid material. The elasticity modulus of the liquid El
corresponds to the elasticity modulus of the solid. It should be noted
that tensile ratio is the ratio between the height of the pulled-up
column and the maximum height of the pulled-up column, which is
different from that of the solid. El only has a relation with the liquid
property and has nothing to do with the area and material of the
lifting plate.
84 Capillary Mechanics

Figure 3.26. σL − εL test curves of silicone plates with different areas.

Since equation (3.47) can be applied to both the loading and


unloading processes, the liquid has a good linear elasticity property
when the deformation is very small before reaching the maximum
surface tension. Through a series of tests on the tensile force of the
lifting plates with different areas, the test curves of different areas
are obtained, as shown in Figure 3.26. σL − εL of the liquid appears
to exhibit very good linearity, and the stress–strain curves of silicone
plates with different areas are basically coincident.
Chapter 4

Contact Angle and Wetting

4.1 Basic Concepts

4.1.1 Contact angle


In Figure 4.1, consider a liquid droplet that falls onto a solid surface.
The liquid droplet will form a spherical crown with a defined volume.
However, the bottom area of the droplet varies depending on the
type of liquid, the properties of the solid surface, and the interac-
tions between the liquid and the solid. The stronger the adhesion
force between the liquid and solid, the larger the contact area will
be. Conversely, the weaker the adhesion force, the smaller the contact
area will be. When the contact area is larger, the angle between the
surface of the droplet and the solid surface at the three-phase bound-
ary (liquid, solid, and gas) is smaller. Conversely, when the contact
area is smaller, the angle is larger. Therefore, this angle reflects the
degree of interaction between the liquid and the solid surface, which
is referred to as the contact angle.

Figure 4.1. Contact angle.

85
86 Capillary Mechanics

The size of the contact angle depends on not only the liquid prop-
erty and the solid surface properties but also the interaction between
the two properties. According to their different properties, the con-
tact angle can be any degree from 0◦ to 180◦ . Even for the same kind
of liquid and solid, when there are differences in flatness, hardness,
surface roughness, and chemical consistency of the solid surface, the
contact angle will also be different. For analysis convenience, a spe-
cific idealized solid surface is often considered. The so-called ideal
solid surface is considered to be flat, rigid, and smooth and to have
uniform chemical properties but no contact angle hysteresis. If it does
not possess the characteristics described above, it is the actual solid
surface and cannot be called an ideal solid surface.

4.1.2 Wetting
Wetting phenomenon occurs when a liquid comes into contact and
adheres to a solid surface. Wetting is the ability of a liquid to
maintain contact with a solid surface, resulting from intermolecular
interactions (adhesive force) when the liquid and solid are brought
together. The degree of wetting (wettability) is determined by the
force balance between the adhesive force and cohesive force. When a
droplet of liquid falls on a solid surface, the adhesive force between
the liquid and solid causes the liquid droplet to spread across the
surface, while the cohesive force within the liquid causes the drop to
ball up and avoid contact with the surface. Since the contact angle
results from the combined action of the adhesive and cohesive forces,
the contact angle also reflects the degree of wetting. Figure 4.2 illus-
trates the wetting behavior of various liquids on a solid surface. The
figure indicates that a smaller contact angle corresponds to a greater
propensity for droplet spreading and a stronger adhesion between
the liquid and the solid surface.
A contact angle which is less than 90◦ (low contact angle) usu-
ally indicates that wetting of the surface is very favorable so that

Figure 4.2. Contact angle and wetting.


Contact Angle and Wetting 87

Table 4.1. The correlation between contact angle and the degree of wetting.

Strength of

Sol./Liq. Liq./Liq.
Contact angle Degree of wetting interactions interactions

θ = 00 Perfect wetting Strong Weak


00 < θ < 900 High wettability Strong Strong
Weak Weak
900 ≤ θ < 1800 Low wettability Weak Strong
θ = 1800 Perfectly non-wetting Weak Strong

the fluid will spread over a large area of the surface. Conversely, a
contact angle greater than 90◦ (high contact angle) generally signi-
fies that wetting of the surface is unfavorable, causing the fluid to
minimize contact with the surface and form a compact liquid droplet.
For water, a surface that is easily wetted is referred to as hydrophilic,
while a surface that is not easily wetted is termed hydrophobic.
Superhydrophobic surfaces exhibit contact angles greater than 150◦ ,
indicating almost no contact between the liquid droplet and the sur-
face. Table 4.1 illustrates the relationship between contact angle and
wetting.
Wetting is a common phenomenon. From the view of the
phenomenon, it is also a process which can be thought of as a pro-
cess of replacing one immiscible fluid with another on a solid surface.
Therefore, the wetting phenomenon must also be associated with
three phases. One of the phases must be a solid phase, and the other
two phases are fluid phases. Normally, wetting indicates that a gas on
the solid surface was replaced by a liquid. Therefore, solid properties
should be concerned for wetting, especially the solid interfacial prop-
erties. The essence of the wetting phenomenon is the relationship
between the solid–gas interface and the solid–liquid interface.
Since the wetting phenomenon involves solid–liquid interaction,
this phenomenon involving a certain kind of liquid can also help us
study some properties of the solid phase interface. For the interface
phenomenon, the research on the solid surface or interface in reality
is the most common one. Since this research is very difficult, com-
pared with that on liquids, people rarely know any properties about
the solid surface or interface. The wetting phenomenon is the macro-
scopic performance of microscopic molecular interactions between
solid and liquid molecules at their interface. Therefore, a study on
88 Capillary Mechanics

the wetting phenomenon will provide much indirect information and


create conditions for people to understand the properties of solid
surfaces.
Meanwhile, the wetting phenomenon can provide information
about solid and liquid interactions. Under the precondition of known
solid surface properties, a study on the wetting of unknown liquid
adhered to the known solid surface can help us indirectly understand
some surface properties of the unknown liquid. Since some kinds of
liquids are soluble in water or interact with mercury, especially with
mercury being harmful to the human body, we cannot obtain the
basic surface force properties by using the direct interaction of those
kinds of liquids with water or mercury. However, if there is no chem-
ical reaction between such liquids and the known solid, by studying
the wetting phenomenon of the liquids on a known solid surface, it is
also possible to obtain some relevant basic properties about surface
force.
In nature, there are two types of solid surfaces. Both of them
can produce an effect with liquids. However, the effect strengths of
these two types are different. Usually, they are called high-energy sur-
face and low-energy surface. To show the relative interaction energy
of a liquid with a solid surface, it must be jointly acted on with
the solid bulk itself. Metals, glass, and ceramics are usually seen as
hard solids because their ontologies are held together firmly through
internal bonds (covalent bonds, ionic bonds, metallic bonds, etc.).
To break them, a very high energy is needed. Therefore, these kinds
of solids are called high-energy solids, and their surfaces are called
high-energy surfaces. Most molecular liquids can wet high-energy
solid surfaces completely. Another kind of solid is held together by
physical forces (van der Waals force and hydrogen bonds), which have
a weak molecular lattice. Since such solids are unified through the
weaker force, which can be broken by little energy, they are called
low-energy solids, and their surfaces are called low-energy surfaces.
According to the different kinds of liquids, some solid surfaces may
be completely wetted while some others may be partially wetted
(incomplete wetting) by liquids.
The wetting phenomenon is also a common phenomenon in
nature. Many processes in animals and plants are closely related
to the wetting phenomenon. Many production processes, such as
cementing, lubrication, petrochemical, washing, welding, printing
Contact Angle and Wetting 89

and dyeing, mineral flotation, waterproofing, and paint coating, are


also closely related to the wetting phenomenon. Wetting theory is
one of the theoretical bases for these production processes.
Therefore, research on wetting should not only focus on the
phenomenon and the mechanism but also study its rules. Wetting
is an essential adhesion of liquid on the solid surface, and more
exactly, a competition between the adhesion force between liquid and
solid surfaces and the cohesion among liquids themselves. Therefore,
wetting involves the properties of liquid cohesion itself and adhesion
with solid.

4.1.3 Microscopic explanation of wetting


phenomenon
When two immiscible phases come into contact with each other,
macroscopically, only one kind of wetting phenomenon is observed.
However, its microcosmic interpretation needs to use interface layer
theory. The so-called interface layer is a thin layer which is located
near the interface (physical interface). According to the development
of surface science in the present situation, there are three types of
interface layer models. One is the Gibbs interface layer model with
a segmentation surface type. Another is Guggenheim-type interface
layer model with excessive strata. The third one is the interface layer
model with a physical interface type. The physical interface type
model affirms that there is a real physical interface phase between
two phases. Both of the two contacted phases have their own interface
sub-layers. The real physical interface was located between the two
interface layers. The thickness of the interface sub-layer is roughly
the same as the radius of molecular interaction. The changes in real
physical properties of the interface on both sides are called mutation.
Therefore, the changes in the properties of the interface layer between
these two phases are mutation. The property of interface layer and
that of inside bulk (body) are different. Whereas this difference is
not mutation but a continuous process.
According to the physical interface layer model, the region at the
juncture of liquid and solid materials is characterized by a thin liquid
layer, the thickness of which corresponds to the radius of molecu-
lar interaction. This layer is referred to as the liquid interface that
forms upon the contact between the liquid and solid phases. The
90 Capillary Mechanics

Figure 4.3. Wetting and non-wetting.

liquid interface layer is shown by a dotted line in Figure 4.3. Macro-


scopically, it is called the “adhesive layer”. The force acting on the
molecules in the liquid interface layer, or “adhesive layer”, is different
from that inside the liquid bulk (body). The molecules in this layer,
on the one side, are forced by solid-phase molecules lying on one side
of the layer, which is the adhesion force, and on the other side, experi-
ence attractional force from the liquid body lying on the other side of
the layer, which is the cohesion force. From the perspective of bound-
ary layer, the side connected by the physical interface experiences the
surface external force, and the side connected by the liquid experi-
ences the surface internal force effect. In general, these two forces
are not equal. Accordingly, the force acting on the molecules in this
liquid interface layer is not symmetrical. When the cohesion force is
greater than the adhesion force, the resultant force of the cohesion
and adhesion forces of the liquid surface layer points in the direction
toward the inside of the liquid, which tends to squeeze the molecules
of the interfacial layer as much as possible inside the liquid. At this
point of time, the solid is not wetted by the liquid. The liquid inter-
face layer has the tendency of spontaneous contractions. The surface
of the liquid has a convex shape. The contact angle for liquid–solid
is an obtuse angle. When these phenomena appear, the solid is not
wetted by the liquid. The essential reason for non-wetting is that the
cohesion force is greater than the adhesion force. When the cohesion
force is less than the adhesion force, the resultant force of the cohe-
sion and adhesion forces of the liquid interface layer points in the
direction away from the liquid (pointing toward the solid). The liq-
uid internal molecules have the tendency to be pulled into the liquid
interface layer. At this point of time, the solid is wetted by the liquid.
The liquid interface layer has the tendency to spontaneously stretch.
Contact Angle and Wetting 91

The surface of the liquid has a concave shape. The contact angle for
liquid–solid is an acute angle. When these phenomena appear, the
solid is wetted by the liquid. The essential reason for wetting is that
the adhesion force is greater than the cohesion force.

4.2 Influence Factors of Contact Angle

The size of the contact angle reflects the degree of interaction of liq-
uid and solid surfaces and also the degree of wetting. If the attraction
between the liquid and solid surfaces is very strong (such as water
drops on a strongly hydrophilic solid), the liquid (drop) will totally
spread on the solid surface. The contact angle is close to zero. If the
solid is not strongly hydrophilic, the contact angle will reach 90◦ .
In many strong hydrophilic solid surfaces, water droplets are found
with contact angles ranging from 0◦ to 30◦ . If the solid surface is
hydrophobic, the contact angle will be more than 90◦ . For superhy-
drophobic surfaces, the contact angle will reach 150◦ or even 180◦ . On
these surfaces, water droplets will simply stay on the surface, without
any substantial adhesion. These surfaces are called superhydrophobic
surfaces, which can be obtained through fluorination (Teflon). The
effect is also called “the lotus effect” because it is hydrophobic like
a lotus, even rejecting sticky liquids such as honey. Table 4.2 shows
the contact angles of different liquids and solids.

Table 4.2. Contact angle of different liquids and solids.

Liquid Solid Contact angle

Water Soda lime glass lead glass fused silica 00


Alcohol
Diethyl ether
Carbon tetrachloride
Glycerol
Acetic acid
Water Solid paraffin 107◦
Silver 90◦
Methyl iodide Soda lime glass 29◦
Lead glass 30◦
Fused silica 33◦
Mercury Soda lime glass 14◦
92 Capillary Mechanics

For different liquids and solids, the corresponding contact angle


is different. Some of them have large differences. In addition, surface
roughness, material composition, and temperature also have an
influence on the contact angle.

4.2.1 Influence on contact angle by surface roughness


Surface roughness has a direct influence on the real contact angle. It
can be described by the Wenzel model, which is shown as follows:

cos θ ∗ = γf cos θ (4.1)

where θ ∗ is the apparent contact angle, θ is the ideal contact angle,


which is also called the Young contact angle, and γf is the surface
roughness ratio, which is defined as real solid superficial area divided
by its apparent area.

4.2.2 Influence on contact angle by the composition


of surface materials
The composition of surface materials has a direct influence on the
contact angle. This relationship can be described by Cassie’s law,
which pertains to composite surfaces composed of two types, as
shown in the following:

cosθc = f1 cosθ1 + f2 cosθ2 (4.2)

where θc is the contact angle of the real composite material, θ1 is


the contact angle of material 1, f1 is the proportion of material 1,
θ2 is the contact angle of material 2, and f2 is the proportion of
material 2, f1 + f2 = 1.
For various materials, equation (4.2) can be expanded as follows:


N
cosθc = fi cosθi (4.3)
i=1

and fi = 1.
Contact Angle and Wetting 93

4.2.3 Contact angle changing with temperature


Contact angle is associated with the degree of wetting, which reflects
the molecular cohesion and the adhesive forces between molecules
at the interface. Variations in temperature can influence molecu-
lar movement, spacing, and intermolecular forces. Consequently, the
contact angle is closely linked to temperature.

4.3 Contact Angle Hysteresis and Influencing Factors

When liquid drops fall on a non-ideal real solid surface, the droplets
will unfold from their original spherical shape to a spherical crown on
the solid surface. Then, a contact angle is formed. Since this stable
state unfolded from a drop, the contact angle is called the advancing
contact angle, which is represented by θa . If we use a method to pack
up the spherical crown from the solid surface, at the beginning, the
three-phase contact wire does not move, and only the contact angle
reduces. When the contact angle has reduced to a certain degree,
the three-phase contact wire starts to shrink. This change in contact
angle from the process of taking back is called the receding contact
angle, which is represented by θr . For a real solid surface, the receding
contact angle is less than the advancing contact angle.
The phenomenon of the receding contact angle being less than
the advancing contact angle is called contact angle hysteresis. The
difference between these two contact angles is the hysteresis contact
angle. For an ideal solid surface, the receding contact angle is equal
to the advancing contact angle.
There are many factors that influence contact angle hystere-
sis. The most important factor is solid surface roughness (or non-
smoothness). Since a real surface is not very smooth, the liquid–gas
surface is suffocated during expansion. It is possible to further expand
only after attaining a maximum angle. Moreover, since a real surface
is not very smooth, receding is suffocated. It can be taken further
back only after attaining a minimum angle. Therefore, the surface
roughness has the most direct influence on contact angle hysteresis.
In addition, the presence of pollution on surfaces, the accumulation
of sediment, and three-phase line movement speed also have great
effects on contact angle hysteresis.
94 Capillary Mechanics

Figure 4.4. Advancing and receding contact angles for water on PTFE wax.

4.3.1 Influence of surface roughness


Dettre and Johnson obtained a set of classic data for water on polyte-
trafluoroethylene (the PTFE, or Teflon) wax blends through a series
of experiments, as shown in Figure 4.4. They used spray wax to
produce a rough surface and made the surface smoother by furnace
heating. In the graph, the abscissa n expresses the number of heat
treatments. The roughness decreases with an increase in the number
of heat treatments. The change in the receding contact angle θr with
the roughness is nonmonotonic.

4.3.2 Influence of surface pollution


Pollution on the surface of a liquid or solid can easily cause the hys-
teresis phenomenon. For example, if a solid plate surface has grease,
upon slowly submerging it into water, the water will be blocked by
the polluted locations. Then, it forms a large contact angle. When
water spills over grease, most oil will unfold on the surface of water.
When the receding contact angle is measured by revealing the solid
from the water, lower values will be obtained. Conducting experi-
ments by using graphite and talc, Fowkes and Harklns found that
Contact Angle and Wetting 95

the contact angle hysteresis phenomenon can in fact be eliminated


by strictly purifying the liquid and solid.

4.3.3 Influence of deposits on solid surface


Solute in liquid, such as surfactants and polyphosphate content, can
deposit a layer of film on a solid surface. In some cases, the existence
of this film can cause the hysteresis effect. Generally, once the film
is formed, it will stick on the surface.

4.4 Mechanics Balance Relationship of Three-Phase Line


and the Young Equation

When a certain kind of incompatible liquid is dropped on a liquid


surface, it presents three phases (gas α, liquid β, and liquid θ) as
illustrated in Figure 4.5. There is a triple-phase line at the triple-
phase meeting point. In equilibrium, the unit-length resultant force
of the three-phase boundary should be zero.
Their projection components in any three-phase border direction
should also be zero:

γαβ + γθβ cosθ + γαβ cosα = 0 (4.4)


γαθ cosθ + γθβ + γαβ cosβ = 0 (4.5)
γαθ cosα + γθβ cosβ + γαβ = 0 (4.6)

where α, β, and γ each represent a phase and the angle of this phase.
γij is the surface free energy (interface free energy) or surface tension
(interface tension) between the two phases. This relationship can be
shown by a similar triangle, as seen in Figure 4.6.

Figure 4.5. Three-phase line.


96 Capillary Mechanics

Figure 4.6. Neumann triangle.

The triangle is called the Neumann triangle, which meets the


following geometric condition:

α + β + θ = 2π (4.7)

This geometry relationship can be used to deduce the surface free


energy of the three-phase relationship. From this triangle, the sum
of any two sides is greater than the third side, γij < γjk + γik . This
indicates that no one energy can exceed the total of other interface
free energies.
If phase β is a rigid smooth surface, β = π. The above three
equations can become a single equation:

γSV = γSL + γLV cosθ (4.8)

where S is the rigid smooth solid surface phase, L is liquid, and V


is gas. This equation is known as the Young equation. From this
equation, we can understand that if the surface free energies of the
three phases are known, the equation can be used to determine the
contact angle. Furthermore, if one of the gas phases is replaced by a
different type of liquid, this equation remains applicable.

4.5 Types of Wetting

Generally, the wetting phenomenon is categorized into three types:


bedewing, soaking, and spreading. These three types of wetting
represent distinct methods of application. One of the most common
forms of wetting is spreading.
Contact Angle and Wetting 97

4.5.1 Bedew
The essence of bedewing is that a liquid adheres to a solid surface.
The process of bedewing means that when the liquid and solid
touch each other, a liquid–gas interface and a solid–gas interface are
changed into a solid–liquid interface.
When two phases touch each other, the intermolecular forces that
exist between them are referred to as “adhesion”. Therefore, the
essence is that the liquid adheres to the solid surface.
As shown in Figure 4.7, phase A and phase B are defined, where
phase A represents the liquid and phase B represents the solid phase.
The adhesion process occurs when phase A and phase B come into
contact with each other; the contact surface between phase A and
gas and the contact surface between phase B and gas are replaced
by the contact surface between phase A and phase B.
Before adhesion, the free surface energy of phase A is γA and that
of phase B is γB . After adhesion, these two energies will disappear
and be replaced by the interfacial free energy between A and B. The
contact area between A and B is supposed to be a united area. Under
constant temperature and pressure, the change in system free energy
caused by the adhesion process is

ΔG = γAB − γA − γB (4.9)

where γAB , γA , and γB , respectively represent interfacial free ener-


gies for the united area of A and B, A and gas, and B and gas.
According to the work–energy conservation, reduction of the free
energy is mainly used in the two-phase adhesion effect. Therefore,
the reduced value should be equal to the increased value of adhesion
work. Therefore, in terms of absolute value, the work of adhesion

Figure 4.7. Adhesion work and cohesion work.


98 Capillary Mechanics

WAB is equal to the change in the system free energy but with an
opposite sign:

WAB = −ΔG = γA + γB − γAB (4.10)

This equation verifies that if WAB ≥ 0, G ≤ 0. The adhesion


process (i.e., the bedew process) can be spontaneous. The smaller
the common interfacial tension between A and B, the greater the
WAB , the easier the adhesion process. Thus, the solid is bedewed
more easily by the liquid.
Another implication for adhesion work WAB is that it is the work
that is used to separate the unit area of phase A and phase B interface
into the interfaces of phase A and phase B with gas. Obviously, the
greater the WAB , the more firmly A and B will contact and combine
with each other, and the stronger the molecular interactions between
A and B. Therefore, the value of the adhesion work WAB reflects the
magnitude of molecular interaction between two phases.
In a special case, if the discussed two phases are the same phase,
such as phase A, the work used to separate A is not adhesion work
but cohesion work. The cohesion work for A is

wAA = γA + γA − 0 = 2γA (4.11)

The cohesion work WAA is equal to the work required for opening
and separating cylinder A, as shown in Figure 4.7. The same cohesion
work reflects not only the strength of the combination of phase A
itself but also the magnitude of molecular interaction of phase A.

4.5.2 Soak
When the clean surface of a solid is immersed in liquid, the solid–
gas interface will be changed to a solid–liquid interface. There is no
change to the surface of the liquid in such a process, which is called
soaking, as shown in Figure 4.8.
If the total area of the solid block is a unit area, since the solid
surface free energy before the immersion is γB , and the original solid
surface is changed to a liquid contact interface after immersing, its
interface free energy is γAB , and the liquid surface free energy has
no change. Therefore, under constant temperature and pressure, the
Contact Angle and Wetting 99

Figure 4.8. Soaking process.

change in system free energy caused by the process is as follows:


ΔG = γAB − γB (4.12)
The work done by the process is called soak work:
WA/B = −ΔG = γB − γAB (4.13)
According to the equation above and superficial phenomenon the-
ory, if soak work WA/B ≥ 0, G ≤ 0. The process can be spon-
taneous. For the bedewing process mentioned above, it has been
explained that, since the solid–liquid common interfacial tension is
always less than the sum of their respective surface tensions, the
adhesion work is greater than zero, so bedewing process can be
spontaneous. However, soak is different. Only when solid surface free
energy is greater than the solid–liquid common interface free energy,
WA/B is greater than zero. Only in this case can the soak process
occur spontaneously. This means that not all liquids and solids can
be soaked spontaneously.

4.5.3 Spreading
Bedew and soak are special cases. Normally, liquid spreads after
bedewing solid. The soak process always goes through the liquid sur-
face. Therefore, common soak is always accompanied by spreading.
The Neumann triangle and the Young equation predict that any one
of solid–gas surface free energy γSV , the liquid–solid interface free
energy γSL , and liquid gas–interface free energy γLV cannot be more
than the sum of the free energy of the other two surfaces. The restric-
tions are on situations for partial wetting. By breaking this restric-
tion, the conditions can be determined for the complete wetting or
total non-wetting. If γSV > γSL + γLV , it can be considered as com-
plete wetting. When γSL > γSV + γLV , it can be considered as total
100 Capillary Mechanics

non-wetting. In this case, the Young equation has no physical solu-


tion, which means the contact angle does not have a balance state
between 0◦ and 180◦ .
The spreading parameter S is a useful parameter to measure wet-
ting, which is defined as
S = γSV − (γSL + γLV ) (4.14)
When S > 0, the liquid completely wets the solid surface. And
when S < 0, it only partly wets. Substituting the Young equation
into (4.14) leads to
S = γLV (cosθ − 1) (4.15)
This equation is called the Young–Dupre equation, which has a
physical solution only when S < 0.
This is similar to the discussion on the contact angle. Some
quantitative liquid A is put on the surface of solid B or another
liquid. Then, there are two cases: one is that the liquid would form
drops on the discussed surface or a lens on another liquid; the other
is that the liquid would spread to form a liquid film on the discussed
surface or a thin film on another liquid.
Both these conditions can be called wetting phenomena. However,
their wetting degrees are different. A contact angle of θ < 90◦ usually
indicates partial wetting of the solid surface, which is called partial
wetting. θ = 0◦ means “perfect wetting”, and the liquid will spread
over the surface. Contact angles greater than 90◦ generally mean
non-wetting to the solid surface, as shown in Figure 4.9.
When θ ≤ 0◦ , liquid droplets can automatically spread over the
solid surface to form a liquid film. The liquid droplets can spread
over the solid surface.

(a) (b) (c) (d)

Figure 4.9. Spread condition of liquid on liquid or solid. (a) Wetting without
spread, (b) wetting, (c) perfect wetting spread, and (d) liquid spread to another
liquid.
Contact Angle and Wetting 101

During spreading, the solid–gas interface disappears and the


solid–liquid and liquid–gas interfaces are formed. Based on energy
change, under constant temperature and pressure, the free energy
change of system is
     
∂G ∂G ∂G
dG = dALV + dALS + dASV (4.16)
∂ALV ∂ALS ∂ASV
where dA is an area variable and the subscripts L, S, and V represent
liquid, solid, and gas, respectively. During the spreading process, the
increased area of the liquid–gas is equal to the decreased area of the
solid–gas interface. Hence,
dALV = −dASV = dALS (4.17)
Since
(∂G/∂ALV ) = γLV ; (∂G/∂ALS ) = γLS ; (∂G/∂ASV ) = γSV
(4.18)
we have
     
dG ∂G ∂G ∂G
= + − = γLV + γLS − γSV
dALV ∂ALV ∂ALS ∂ASV
(4.19)
By defining
 
dG
S=− (4.20)
dALV
where S is the spreading coefficient of the liquid over the solid surface,
we have
S = γSV − γLV − γLS (4.21)
It is the change in free energy of the system during liquid spread-
ing over the solid surface. It is also the same as the previously
defined spreading coefficient. From the perspective of energy, it has a
new meaning: the spreading coefficient is the negative energy change
which is caused by increasing unit liquid surface or solid–liquid sur-
face. If S is positive, it indicates that the system free energy can
decrease and the spreading is spontaneous. If S is negative, it indi-
cates that the system free energy needs to increase and the spreading
102 Capillary Mechanics

is not spontaneous, which can only form liquid droplets on the sur-
face. Hence, from the value of S, we can judge the ability of spreading
over the solid.
Generally, organic liquids can spread over high-energy surfaces
(metals and glass), but they cannot spread over low-energy liquids
(paraffin and polyethylene), particularly when surface tension is very
high. Presently, the most difficult to spread is the solid which has —
CF3 , secondly –CF2 –, or –CH3 and –CF2 –. The practical situation
reflects that the spreading ability of the discussed surface relates
to not only the cohesion force of molecular interactions but also the
adhesion forces of molecular interactions between the liquid and solid
surfaces.
Since the spreading process is an unbalanced process of the system
but the spreading coefficient above is based on energy or surface
free energy for the balance system, it is called traditional spreading
coefficient.
The spreading coefficient can be rewritten as
S = γSV − γLV − γSL = γSV + γLV − γSL − 2γLV = WSL − WLL
(4.22)
where WSL is the solid–liquid adhesion work and WLL is the liquid
cohesion work. When the adhesion work between the liquid and the
solid is greater than the liquid cohesion, WSL > WLL , S > 0, the
liquid can spread over the solid. When the adhesion work between
the liquid and the solid is less than the liquid cohesion, WSL < WLL ,
S < 0, the liquid cannot spread over the solid.

4.6 Principle and Method of Measuring Contact Angle

There are many measurement methods for the contact angle, includ-
ing both direct and indirect methods. By directly taking photos of
the shape of droplets or bubbles that are spreading on a solid surface,
the contact angle can be measured by assessing the angle and height
from the photos. This measurement is a direct method. Other direct
methods also include the use of specular reflection law or parallel
beams to measure the contact angle. Calculating the contact angle
from the capillary rise or decline in height by capillarity is an indi-
rect method. Indirect measurement methods also include the force
method, among others, for measuring contact angle. Currently, there
Contact Angle and Wetting 103

are many kinds of equipment for measuring the contact angle. Some
commonly used methods and principles of contact angle measure-
ment are introduced here.

4.6.1 The principle of measurement for contact


angle by hypsometry
When the volume of a drop is less than 6 μL, the influence of gravity
on its shape can be negligible, and the graph of the drop viewed
from the front is thought to be part of a standard circle. As shown in
Figure 4.10, if the height h of the liquid on the solid surface and the
diameter D of the solid contact area can be measured, the contact
angle θ can be calculated. According to a geometrical relationship,
the computational formula can be written as

2h
θ = 2 arctan (4.23)
D
where θ is the contact angle, h is the height of the liquid spherical
crown, and D is the diameter of the bottom circle of the spherical
crown.
Whether θ > 90◦ or θ < 90◦ , this equation is applicative. This
equation is deduced as follows.
For right-angled triangle ACO  , α + β = 90◦ , then

θ = 90◦ + (β − α) = (β + α) + (β − α) = 2β (4.24)

tan β = hr , so, β = arctan hr = arctan 2h


D

Figure 4.10. Hypsometry method to measure the contact angle.


104 Capillary Mechanics

Then,

2h
θ = 2 arctan (4.25)
D

4.6.2 Photo goniometric method


The tool used in this method is an isosceles right-angled protractor.
The two sides of the isosceles right-angled protractor are, respec-
tively, a and b. The protractor can be moved up and down or left
and right. Its measurement process is as follows:
1. Move down the a and b sides until they are tangent to the liquid,
as shown in Figure 4.11.
2. Continue moving down verticality until the peak of the protractor
and the border of drop are crossed at point C. Therefore, the
coordinate of the peak C can be defined, as shown in Figure 4.12.
3. Make the protractor counterclockwise and rotate around point C ;
turn δ angle until the side A intersects with the crossed point A
of the three phase surfaces, as shown in Figure 4.13. Then, θ can
be obtained.
According to a geometrical relationship,

θ = 90◦ + (β − α) = (β + α) + (β − α) = 2β

Since β = 45◦ +δ, θ = 90◦ +2δ. When the contact angle θ > 90◦ ,ide
AC of the protractor is rotated counterclockwise and δ is positive;
when the contact angle θ < 90◦ , side AC of the protractor is rotated
clockwise and δ is negative.

Figure 4.11. Tangency with both sides of protractor figure.


Contact Angle and Wetting 105

Figure 4.12. Parallel moving downward and made peak coincidence with peak
of sphere.

Figure 4.13. Left-handed rotation protractor and make solid–liquid–gas


crossing.

4.6.3 Specular reflection method


Langmuir and Schaeffer put forward through droplets a mirror to
measure the contact angle. As shown in Figure 4.14, a single beam
shines at the three-phase contact line. When and only when the
incident angle is equal to the contact angle, the observer can see
reflected light in the incident direction. Fort and Patterson improved
this method, developed it into the production, and made its accuracy
reach ±1◦ . The actual device diagram is shown in Figure 4.15. The
end of a lever which can rotate around the fixed point is equipped
with an illuminant, which can produce a single beam. At the point
where it is close to the illuminant, there is a peephole which is used to
observe the reflected light. Rotation of the lever is realized through
screw motion. The key to this operation is to make the bottom of the
tested droplets and the rotation axis of the lever coplanar. This can
guarantee that the single beam transmits to the contact line of the
droplet, which has a certain difficulty in actual practice. Good and
Ferry measured the contact angle of stainless steel in liquid hydrogen
106 Capillary Mechanics

Figure 4.14. Schematic diagram of reflected light of a single beam.

Figure 4.15. Tester of mirror reflection.

by uniting the tilting plate method and the reflection method and
made its accuracy reach ±0.5◦ . There are many advantages to this
method, such as avoiding the use of a complex optics system, simple
equipment, low cost, and ease of processing. However, it can only
measure contact angles which are less than 90◦ .

4.6.4 Parallel beam method


Allain’s parallel light method is essentially a further development of
the method of specular reflection, as shown in Figure 4.16. Paral-
lel beams entering from right above the droplet are reflected by the
gas–liquid interface around the droplet and form two angles with the
incident direction. A screen is placed at a point that has a distance
s from the measured surface. The image of the bottom of the ampli-
fying droplet appears on the screen. Its magnification depends on
s and the contact angle. The simple geometrical relationship shows
Contact Angle and Wetting 107

Figure 4.16. Light path chart of the parallel beam method.

that
 
1 D − Dd
θ= arctan (4.26)
2 2s

This method can measure contact angles which are less than 45◦
and the accuracy can reach +0.1◦ .
For surfaces capable of transmitting light, such as glass, Allain
proposed the parallel light transmission method. Parallel beams
entering right above are refracted by the gas–liquid interface around
the contact wire, enter inside the droplet, and are then refracted
once again by the liquid–solid interface into the gas phase below.
Two refracted beams and the incident direction form θ  angle. If the
screen below is placed at the point which has a distance s from the
measured surface, the inverted image of the bottom of the amplifying
droplet also appears on the screen.

4.6.5 Calculating contact angle from rise and


descent method by capillary tube
As shown in Figure 4.17, R1 is the radius of the supposed capillary
tube and R is the radius of curvature of the meniscus formed by the
liquid in a tube. Then, the contact angle θ can be calculated using
equation (4.27):

R1
cos θ = (4.27)
R
108 Capillary Mechanics

Figure 4.17. Capillary rising.

R1 can be measured from the pipe diameter and R can be obtained


from the following:

Δρ · gh = (4.28)
R
Δρ and h can be measured, and γ can be calculated from another
measurement.
Different measuring methods correspond to different principles,
and the quality of the measuring method depends on the accuracy
of measurement. With the unceasing development of optical mea-
surement technology, the measurement precision of contact angle is
also gradually improving. Since the contact angle is closely related
to surface tension, many measurement methods of contact angle are
accompanied by the measurement of surface tension. Some of the
methods above can also be used to measure the contact angle, such
as the dynamometry method.

4.6.6 Principle of dynamometry to measure


contact angle
This method is also called Wilhemly (broadly) or Wilhemly type
and was suggested by Wilhemly in 1863. The device is shown in
Figure 4.18. A test solid sheet is connected to an electronic balance
through a wire. When the sheet is not dipped into the liquid, it is
Contact Angle and Wetting 109

Figure 4.18. Wilhemly device schematic diagram. (1) tested solidsheet,


(2) wire, (3) electronic balance, (4) recorder, (5)tested unit, (6) tested liquid,
(7) removable platform, (8) hoistable platform, (9) electrical machine,
(10) support, (11) cup.

only supported by gravity, as indicated by the force on the device:

F1 = mg (4.29)

where F1 is the force when the sheet has not been dipped into the
liquid, m is the mass of the tested solid sheet, and g is acceleration
due to gravity.
When the sheet is dipped into a depth of h, it reaches a balance:

F2 = mg + cγ cos θ − ΔρgVdisp (4.30)

This represents the device when the depth is h. c is the wetting


perimeter, γ is the surface tension of the liquid, θ is the contact angle,
Δρ is the density difference between liquid and gas, h is the depth
of immersion, and Vdisp is the volume of the immersed liquid.
The difference in the force measurement device before and after
immersing the solid sheet into liquid is

ΔF = F − F = cγ cos θ − ΔρgVdisp (4.31)

The contact angle can be calculated by the measurement method.


The emergence of electronic balance greatly improved the accuracy,
which changed from the early ±1◦ to ±0.1◦ . Dynamic surface energy
analyzers of type K12, produced by a German company, are designed
based on the Wilhemly principle.
110 Capillary Mechanics

4.7 Measuring Method of Contact Angle Tester Based


on Photogoniometric Method

4.7.1 Construction compose of JYSP-180 tester


The instrument consists of an illuminant, worktable, base, magni-
fying glass, CCD camera, and liquid dropping device, as shown in
Figure 4.19.

4.7.2 Installation and use of tester


1. Install and adjust

A. Put the tester on a stable platform. Room temperature should


be constant at 20 ± 5◦ C. Adjust the leveling knob and make the
tester horizontal based on a level bubble.
B. Connect the magnifying glass to the CCD camera, and then install
them on lifting bearings.
C. Connect the CCD camera and the video capture card using a
video cable. Connect the CCD camera power.
D. Insert the power plug in the external socket which has reliable
grounding. Turn on the power switch.
E. Adjust the working condition of the tester, as follows.

Enter the video collection window. By adjusting the workbench,


lift the bearing and magnifying glass tube. The objective table will
appear in the video window. Adjust the image until it is clear. Unlock
the locking button. Turn on the camera and align the upper edge

Figure 4.19. Structure of contact angle tester.


Contact Angle and Wetting 111

Figure 4.20. Initialized states.

of the objective table with the green line in the window, as shown
in Figure 4.20. Then, tightly lock the lock button to complete the
adjustment of the equipment.
2. Use tester
A. Place the solid sample in the small objective table.
B. Utilize a dispensing apparatus to administer an appropriate vol-
ume of liquid onto a solid sample.
C. Adjust the workbench, lift bearing, and magnifying glass tube so
that the image is clear.
D. Press the “video capture” to record clear images and continue to
see if there are clearer images. If there is any, continue to press
the “video capture” to record the clear images. Click “measured”
and choose the measurement method. If a clearer image is needed,
directly click “measurement” and choose different measurement
methods.

3. Operating steps
A. As shown in Figure 4.21, move the measuring ruler up and down
or left and right
112 Capillary Mechanics

Figure 4.21. Move the ruler.

Figure 4.22. Ruler and droplet edge tangent.


Contact Angle and Wetting 113

Figure 4.23. Demarcate the high point of the droplet by ruler.

Figure 4.24. Rotate ruler to find contact angle.


114 Capillary Mechanics

Note: this time, only move horizontally or vertically the measure-


ment ruler and do not rotate it). Move the measuring ruler to the
position, as shown in Figure 4.22. Make the measuring ruler and
droplet edge tangent.
B. Lock the measuring ruler. Make it overall down until the inter-
section of the ruler and droplet edge coincide. Determine the high
coordinates of the droplet, as shown in Figure 4.23.
C. Lock the buttons up and down or left and right. Rotate the mea-
suring ruler, as shown in Figure 4.24. Make the ruler and droplet
side intersect, and then click the “ok” button to complete the
contact angle measurement.
Note: The translation and rotation test process can be accelerated
by changing the step length: directly use the “up”, “down”, “left”,
and “right” arrow keys on the keyboard to move the measuring
ruler. The “Z” and “Y” buttons are for left rotation and right
rotation, respectively.
D. Click “return” to continue with the next time.
Chapter 5

Young–Laplace Equation and Its


Application

5.1 Young–Laplace Equation

The most typical example of capillarity phenomena is the rise and


fall of liquid in a capillary tube, which results from the surface ten-
sion of the liquid. Its motivation comes from the pressure difference
between the high and low layers of liquid in the capillary tube. The
curved surface of the high layer of the liquid is related to both sur-
face tension and pressure difference. What is the relationship among
pressure, surface tension, and curved surface? In order to discuss
such a relationship among the pressure on both sides of the curved
surface, the liquid surface tension, and the geometric shape of the
curved surface, a micro-curved surface unit 1234 is considered. The
micro-surface is cut by two orthogonal normal incisions. The so-called
normal incision is the intersection line of the plane containing normal
lines and the curved surface. If two normal incisions are in the same
directions, respectively, as the maximum and minimum curvature
directions, the curvatures of curves 12 and 14 are the two principal
curvatures of the curved surface unit.
As shown in Figure 5.1, for the micro-unit 1234, curvature radius
(principal curvature) of curves 12 and 34 is r1 , the arc center angle is
dφ1 , and the arc length is a. The curvature radius (another principal
curvature) of curves 14 and 23 is r2 , the arc center angle is dφ2 , and
the arc length is b. Then the area of the micro-unit is dA = a · b =
r1 · dφ1 · r2 · dφ2 . The surface tension of the curve is supposed to be γ,

115
116 Capillary Mechanics

Figure 5.1. Micro-volume and superficial area.

and the difference in pressures acting in the normal direction on


both sides of the micro-unit is Δp (the pressure difference between
the concave side and the convex side of the curve surface). Thus,
when the curve surface has a virtual displacement δξ in the normal
direction, the increased virtual area of the micro-unit (relative to the
original area) is
δdA = δξ · dφ1 · b + δξ · dφ2 · a
= δξ(r2 dφ2 · dφ1 + r1 dφ1 · dφ2 ) (5.1)
= δξ(r1 + r2 )dφ1 · dφ2
Under the effect of pressure difference Δp, when virtual displace-
ment δξ is produced, virtual work is δW = Δp · dA · δξ. Due to the
increased virtual area (overcoming the surface tension), the increased
virtual energy on the surface is δE = δdA · γ. Based on the conser-
vation between work and energy, in the case of no other external
energy, there is δW = δE, so that
Δp · r1 r2 dφ1 dφ2 δξ = δξ(r1 + r2 )dφ1 · dφ2 (5.2)
The solution is obtained as
 
1 1
Δp = γ + (5.3)
r1 r2
This is the Young–Laplace equation, which was deduced inde-
pendently by T. Young and P.S. Laplace in 1805 and 1806, respec-
tively. Actually, the Young–Laplace equation can also be deduced
Young–Laplace Equation and Its Application 117

by mechanical equilibrium. Similarly, the micro-curved surface unit


is considered. The tension acted on the surface is supposed to be
γ, the pressure acted on the concave side of surface is Pconcave .
The convex side of surface is Pconvex , then pressure difference is
Δp = Pconcave − Pconvex . This pressure can be balanced by the
resultant force of components in normal direction of surface tension.
Components in tangential direction of γ can be balanced by itself.
The resultant force in normal direction is
dφ1 dφ2 dφ1
fs = γ · b · sin × 2 + γ · a · sin × 2 = γ · r2 dφ2 · sin ×2
2 2 2
dφ2
+ γ · r1 dφ1 sin ×2 (5.4)
2
When the micro-unit is infinitely tiny, sin(dφ1 /2) = (dφ1 /2),
sin(dφ2 /2) = (dφ2 /2),

fs = γ(r1 + r2 )dφ1 dφ2 (5.5)

The resultant force of pressure difference Δp in normal direction


is

fγ = Δp · a · b = Δp · r1 dφ1 · r2 dφ2 (5.6)

Under equilibrium condition, fs = fγ , i.e.,

γ(r1 + r2 )dφ1 · dφ2 = Δp · r1 r2 dφ1 dφ2 (5.7)

The solution is
 
1 1
Δp = γ + (5.8)
r1 r2
If the two sides of the surface are in two different phases, the
pressure of convex phase should be greater than the pressure of con-
cave phase. If convex phase is liquid and concave phase is gas, for
example, the liquid drop in the air, the pressure of liquid is greater
than the pressure of gas. If convex phase is air and concave phase is
liquid, for example, the bubble in the liquid, the pressure of gas is
greater than the pressure of liquid.
For cylinder surface, r1 → ∞, r2 = r, then the Young–Laplace
equation is changed into Δp = γ/r, where r is the radius of cylinder.
118 Capillary Mechanics

Figure 5.2. The radius for drops is r.

For sphere surface, r1 = r2 = r, then the Young–Laplace equation


is changed into Δp = 2γ/r, where r is the radius of a sphere.
This formula can also be directly deduced from a spherical
droplet. As shown in Figure 5.2, the surface area of ball with radius r
is 4πr 2 , its volume is (4/3)πr 3 , the difference of pressures which act
on spherical inside and outside is supposed to be Δp, and the surface
tension of the ball is γ. With the increase of pressure difference Δp,
the ball expands and virtual displacement δr is produced. Then, the
virtual increasing of the surface area is δA = δ(4πr 2 ) = 8πrδr. The
work for pressure difference Δp under δr is δW = Δp · 4πr 2 · δr.
The new increased virtual surface free energy for overcoming surface
tension is δE = γ · δA = γ · 8πrδr. Based on the conservation of
work–energy, these two should be the same, δW = δE. Then

Δp · 4πr 2 · δr = γ · 8πrδr (5.9)

The solution is

Δp = (5.10)
r
For flat liquid, r1 = r2 = ∞, substituting it into the above equa-
tion leads to Δp = 0.
For plane surfaces, the pressure on the two sides of the surface
is the same, i.e., there is no pressure difference. This curved surface
unit is called the additional pressure of the curved liquid surface. For
plane liquid surfaces, since there is no pressure difference, the addi-
tional pressure is zero. However, for convex liquid surfaces, since the
pressure difference is positive, the additional pressure is also positive.
Young–Laplace Equation and Its Application 119

Figure 5.3. Some phenomena can be explained by the additional pressure of


curved liquid surface.

Figure 5.4. Irregular surfaces automatically shrink into a sphere.

For concave liquid surfaces, since the pressure difference is nega-


tive, the additional pressure is negative. The total pressure inside
the liquid is less than the pressure in air. This condition is shown in
Figure 5.3.
Free drops or bubbles present as spheres (shown in Figure 5.4).
For irregular drops and bubbles, different parts of the surface have
different curvature radii, and the size and orientation of the addi-
tional pressures are different. Based on Pascal’s law in hydrostatics,
pressures inside a drop should be the same in all respects. Hence,
drops or bubbles that are acted upon by an imbalance of forces will
automatically change their shapes to spheres. The additional pres-
sures are the same, pointing inward and canceling each other out,
which makes each radius of curvature the same, forming the shape
of a sphere and balancing the system.
It shows superheated phenomena for liquid boiling. When the liq-
uid saturated vapor pressure is equal to the external pressure, the
temperature for gas–liquid phase equilibrium is the boiling point.
The phenomenon in which a liquid exceeds the boiling point with-
out boiling is known as superheating. During the process of boiling,
vaporization simultaneously proceeds in the inside and outside of the
liquid. The new formative phase is called the micro-bubble phase.
120 Capillary Mechanics

Figure 5.5. Superheated phenomena.

Since the radius of the curvature of bubbles in liquid phase is nega-


tive, there is an additional pressure Δp which points to the inside of
liquid. Thus, the bearing pressure for the bubble should be the joint
force of the outside pressure, the static pressure ph which is generated
at depth h, and additional pressure, as shown in Figure 5.5.
Only when the saturated vapor pressure is equal to the joint force,
bubbles become bigger and rise. Therefore, a liquid under superheat-
ing can boil only when the temperature is higher than its boiling
point. If the radius of new bubbles is less than 10−5 m, the super-
heated phenomena are clear. Therefore, to prevent this, zeolites and
capillaries are often added to the liquid. Air escapes from the zeo-
lites and capillaries during the process of heating, which increases
the radius of new bubbles to about 10−3 m so as to eliminate the
superheated phenomena.
This kind of additional pressure cannot be disregarded. Two
examples are as follows.

1. Cavitation phenomenon: At the beginning of the 20th century,


when the first batch of ocean ships was manufactured successfully
for a trial trip, it was found after 12 hours of trial voyage that
the propellors had become tattered and could not be used. After
years of research, finally, it was proved that the countless tiny
water bubbles caused the damage. Usually, the damage on metal
propeller by the tiny bubbles is called “cavitation”.
Young–Laplace Equation and Its Application 121

When the propeller runs up in water, countless tiny bubbles are


formed under enormous pressure. Then a sheet bubble cloud is cre-
ated. Some of them are too small to be distinguished by the naked
eye because of their minimal curvature radius. Experiments sug-
gested that by now the surrounding liquid produced great pressure
on the bubbles. This is an additional pressure, by which the liquid
membrane of the bubbles will contract with great speed and then
rupture. The pressure produced by the fracturing of the liquid
membrane can be thousands of Mbar (1 Mbar = 105 MPa). Count-
less small bubbles with such great pressure have a continuous and
intensive impact on the metal parts, which leads to damage. This
is cavitation. There are many methods to avoid cavitation, for
example, coating some material like sodium diethylene dithiocar-
bonate on the paddle surface or using the “exceeding cavitation”
method, etc.
2. Air lock: The nurses give patients injections of various kinds of
drugs. Before the injection, they must check out whether there
are some tiny bubbles in the tube. If there are some small bub-
bles, they must be removed. This is because when the blood is
mixed with little bubbles, curved liquid surfaces will be produced
in blood. While slightly adding an external pressure, the curvature
radiuses of the meniscus on the two sides of the bubbles are not
even, as shown in Figure 5.6. The force, which will block the flow
of blood, is made. Only when the additional pressure comes to a
certain value, blood can begin to flow. This is the phenomenon of
“air lock”.
When the human body changes from a high-pressure area to a
low-pressure area, the transition should be very slow because gas
solubility in liquid increases according to the increase in air pres-
sure. Therefore, in high-pressure conditions, large amounts of gas

Figure 5.6. “Air lock” phenomena in blood.


122 Capillary Mechanics

dissolve in blood and tissue fluid. If the outside pressure suddenly


decreases, the gas in the blood and tissue fluid will be released
dramatically.

5.2 Jurin’s Law in Capillary

As mentioned before, liquid rising or descending in capillary is called


capillarity. If a glass capillary is inserted in water, the liquid level in
the tube will rise; if it is inserted in mercury, the liquid level in the
tube will descend, as shown in Figure 5.7. The surface tension of the
liquid relates to the rising height h in capillary tube.
The radius of capillary tube is r and the contact angle is θ. Thus,
the curved radius of liquid is R = −(r/cos θ) and the additional
pressure is Δp = −(2γ/r) cos θ.
For concave liquid surfaces, θ < 90◦ , cos θ is positive, and Δp is
negative. In a capillary tube, the pressure below liquid surface is less
than the bottom pressure because the bottom liquid has the same
pressure with the outside liquid surface pressure, i.e., the atmospheric
pressure. The liquid in the capillary tube is pushed upwards by the
bottom pressure, causing it to rise. When the product of the rising
height h and the pressure difference between the bottom of the cap-
illary tube and the liquid surface in the capillary tube balances with

R
r

h
h

(a) (b)

Figure 5.7. Liquid rising (a) and descending (b) in capillary tube.
Young–Laplace Equation and Its Application 123

the liquid weight,


2γ cos θ
= ρgh (5.11)
r
where g is gravity acceleration and ρ is liquid density. Then,
h = 2γ cos θ/ρgr. Equation (5.11) is called the capillary rising equa-
tion. After researching this kind of effect by James Jurin in 1718, it
is also called Jurin’s Law or Jurin’s Height. It is the basic formula of
measuring the surface tension by capillary tube.
For convex level, θ > 90◦ , cos θ is negative, Δp is positive, and
the liquid surface level descends. The descending height is the same
as the result from the above equation.

5.3 Liquid Bridge Force between Two Plates

It is difficult to separate two plates which have water between them,


as shown in Figure 5.8. In the case of complete wetting, the contact
angle θ = 0◦ and the side liquid surface present as concave, as shown
in Figure 5.9. The pressure inside the liquid is less than the pressure
of the gas outside. Since r1 = δ/2 (δ is the space between two plates),
r2 → ∞, based on the Young–Laplace equation,
 
1 1 2γ
Δp = γ + = (5.12)
r1 r2 δ

Figure 5.8. Water between two plates.

Figure 5.9. For complete wetting, side liquid surface is concave.


124 Capillary Mechanics

The force which attracts the two plates together is

2γl2
f = Δpl2 = (5.13)
8
where l is the length of the plates. If the solid is irregular, it is not
easy to calculate the force. However, as long as the liquid between two
solids presents as concave, it is certain to have an attraction force.
This kind of attraction force formed by liquid is called liquid bridge
force. The existence of liquid bridges leads to a relative negative
pressure and then produces the effect of attracting the two solid
surfaces.

5.4 Falling Stream of Fluid Breaks up into Smaller


Packets

Commonly, falling stream from faucet breaks up into droplets, no


matter how smooth the stream is. This phenomenon is caused by
surface tension and is also called the Plateau–Rayleigh instability.
Considering a stream of water, as shown in Figure 5.10, no matter
how smooth the initial water flow is, it will always exhibit uneven
thickness after being subjected to some initial disturbances. This
uneven shape is manifested in two aspects geometrically. The first
aspect is the radius of the water column. The second aspect is the arc
shape of the outer surface of the water flow column. In fact, these two
aspects deal with the two principal curvatures of the outside surface
of the water column. The radius of the water column is larger in
thick areas and smaller in thin areas. The outer surface of the water
flow column presents an outward convex shape in thick areas and
an inward concave shape in thin areas. The curvature radius of the
convex shape is positive, while the curvature radius of the concave
shape is negative.
According to the Young–Laplace equation, considering only the
radius of the water flow column, the curvature radius in thick areas
is larger, and the corresponding liquid pressure is lower. Thin areas
have a smaller curvature radius corresponding to higher pressure.
The high-pressure area always tries to push the liquid toward the
low-pressure area. Therefore, liquids are easily squeezed from the thin
areas to the thick areas, causing the thin areas to become thinner
Young–Laplace Equation and Its Application 125

Figure 5.10. The uneven thickness of a stream of water after being disturbed.

and the thick areas to become thicker, resulting in neck breakage and
scattering into water droplets (beads). If only this one effect exists,
the phenomenon of water flow becoming water droplets is easy to
understand. However, from the Young–Laplace equation, it can also
be seen that, in addition to the radius of the water flow column, the
curvature radius of the outer arc surface also determines the liquid
pressure. From the curvature radius of the arc surface, it can be
inferred that in thin areas, due to their concave shape, the liquid
exhibits a relatively negative pressure to decrease the pressure, while
in thick areas, due to their convex shape, the liquid exhibits a positive
pressure to increase the pressure. This effect is exactly opposite to the
radius effect of water flow mentioned above. These two effects usually
cannot be accurately offset, and one effect is usually greater than the
other. Its size depends on the initial radius and wave number of the
water flow. The so-called wave number refers to the number of thick
and thin peaks per centimeter of water flow, and each peak along with
126 Capillary Mechanics

a valley is recorded as a number. When the wave number is at certain


values, the curvature radius of the outer surface arc dominates, and
this effect gradually weakens over time. However, when the radius
of the water column dominates, this effect continues to strengthen.
The continuously increasing result means that the water column will
disperse into water droplets.
Although a thorough understanding of how this happens requires
a mathematical development, the diagram can provide a conceptual
understanding.
This process is an unstable process of water flow column. The
occurrence of this process is also conditional. Mathematical analy-
sis shows that when the product of the wave number with the ini-
tial radius is less than unity (kR0 < 1), the unstable components,
components that grow over time, are produced. The component that
grows in the fastest rate is the one whose wave number satisfies the
equation, k · R0 ≈ 0.697, where k is the wave number and R0 is the
initial radius of the water column.
By assuming that all possible components exist initially in roughly
equal (but minuscule) amplitudes, the size of the final drops can be
predicted by determining the wave number whose component grows
the fastest. As time progresses, it is the component with the maxi-
mum growth rate that will come to dominate and eventually be the
one that pinches the stream into drops.
Certainly, the Young–Laplace function can explain some other
phenomena. If a blank has been dried, the volume will be reduced.
A solid catalyst will become harder after being flooded with water.
All of this is because the wetted solids enclose a water film, and

Figure 5.11. Hardening phenomenon.


Young–Laplace Equation and Its Application 127

the particles are close to each other. The water film will become a
slice. A concave liquid surface is formed between the particles. The
additional pressure points to gas. Thus, there is a suction effect that
compels the particles to integrate closely, as shown in Figure 5.11.
The result is that the total volume will be reduced and the particles
will cohere.
This page intentionally left blank
Chapter 6

Kelvin Equation

6.1 Kelvin Equation

From the Young–Laplace equation, the curvature of a liquid surface


has a direct influence on inside and outside pressures. The inside
pressure on a convex surface of gas phase is greater than the out-
side pressure. The inside pressure on a concave surface is less than
the outside pressure. For a large liquid plane, since the surface is
plane, the curvature is zero. Hence, there is no pressure difference
between inside and outside, which means the liquid pressure is the
same as the air pressure. However, for small droplets (ideal droplets
are spherical, while real liquid is similar to spherical), the inside
pressure is greater than the outside pressure, and the pressure dif-
ference can be expressed based on the Young–Laplace equation as
Δp = 2γ/r, where γ is surface tension and r is the radius of liquid
droplets. From this relation, the pressure of liquid droplets has an
intimate connection with the radius of the droplets. The smaller the
droplet, the greater the pressure difference. The bigger the droplet,
the smaller the pressure difference.
The evaporation, or the process of boiling in which a liquid
changes into vapor, is called gasification. Its inverse process is called
condensation. In the gasification process, material molecules escape
from the liquid to the vapor because of molecular heat. In the
condensation process, material molecules return from the vapor
to the liquid because of molecular heat. Under certain conditions
(temperature, pressure, etc.), this kind of escaping and returning

129
130 Capillary Mechanics

will reach a dynamic balance. For example, in a unit of time, the


number of molecules escaping from the liquid is the same as the num-
ber of molecules returning from the vapor. Objectively, evaporation
stops, and then the system is in a saturated state. The vapor is sat-
urated vapor, and the gas pressure is saturated vapor pressure. If a
certain amount of liquid is placed in an airtight vacuum container,
this phenomenon can be observed. Under a certain temperature, due
to molecular heat at the beginning, some of the liquid molecules can
escape from the surface of the liquid. With an increase in the number
of gas molecules, the density of vapor molecules increases. The vapor
pressure also increases. When the temperature remains at a certain
value, the gas pressure will eventually stabilize at a fixed value, and
then the gas pressure is the saturated vapor pressure at this temper-
ature. When the gas pressure reaches the saturated vapor pressure,
liquid molecules are still constantly gasifying. Water molecules in the
gas phase are also constantly condensing into liquid. Since the gasi-
fication velocity of water is equal to the vapor condensation velocity,
the liquid does not decrease, and the gas does not increase. Liquid
and gas reach equilibrium.
Put both large planar droplets and many small droplets in a closed
vessel. In addition to the evaporation-saturated process described
above, at certain temperatures and pressures, after a period of time,
the droplets gradually disappear, and the volume of the large planar
liquid increases. This phenomenon is caused by the pressure dif-
ference between the outside and inside of the liquid droplets. The
large planar liquid does not have such a pressure difference. Since
the outside and inside of the small spherical liquid droplets have a
pressure difference and the saturated vapor is closely surrounding the
droplets, the saturated vapor pressure of the small spherical droplets
is actually greater than that of the large planar liquid. The differ-
ence in saturated vapor pressure leads to the disappearance of the
small droplets. Since the existence of a pressure difference means
the existence of a potential difference, according to the principle of
minimum potential energy, if there is no other external function, the
potential energy in an airtight container should tend to be consistent.
Since the density of small droplets and planar liquid is the same, the
potential difference mainly originates from the pressure difference.
For 1 mol of liquid, its chemical potential difference can be written
 p+Δp
as Δμ1 = p Vm1 dp = V Δp = 2γV l /r, where V l is the mole
m m m
Kelvin Equation 131

volume of the liquid. The potential difference between the saturated


vapors of small liquid droplets and planar liquid is derived not only
from the vapor pressure difference but also from the change in vol-
ume. For 1 mol of vapor, its chemical potential difference can be
written as
 pr  
g RT pr
Δμ = dp = Vm Δp = RT ln (6.1)
p0 p p0
g
where Vm is the mole volume of air, R is the gas constant, T denotes
temperature in degree Kelvin, pr is the saturated vapor pressure
of liquid droplets, and p0 is the saturated vapor pressure of planar
liquid.
The Kelvin function is obtained when the chemical potential
energy of liquid is the same as its vapor chemical potential energy
and the difference in chemical potential energy between non-planar
liquid and planar liquid is the same as the difference in chemical
potential energy of its corresponding saturated vapor. Under satu-
rated conditions, since the phase change of liquid is reversible, its
total Gibbs free energy is zero. Gibbs free energy can be directly
used to obtain the Kelvin function. There are two ways to change
the planar liquid into small droplets. First, evaporate planar liquid
to saturated vapor. The saturated vapor is compressed to saturated
vapor of spherical droplets, which will condense to spherical droplets.
Second, the planar liquid can directly scatter into small spherical liq-
uid droplets. Taking 1 mol of liquid as an example, the first method
has three steps:
First step: Under constant temperature and pressure, 1 mol of
planar liquid reversibly evaporates to saturated vapor, and p0 is the
saturated vapor pressure of planar liquid. This step is a reversible
phase change. Therefore, its Gibbs free energy change is ΔGm,1 = 0.
Second step: The saturated vapor of planar liquid is compressed
into spherical liquid saturated vapor, which is an ideal compression
process of gas under constant temperature and pressure. The Gibbs
free energy change of this step is
 pr  
pr
ΔGm,2 = Vm dp = RT ln (6.2)
p0 p0
132 Capillary Mechanics

Third step: The spherical droplets of saturated vapor condense into


spherical liquid droplets. This step is a reversible phase change under
constant temperature and pressure. Its Gibbs free energy change is
zero: ΔGm,3 = 0.
The total Gibbs free energy change of this method is
 
pr
ΔGm,a = ΔGm,1 + ΔGm,2 + ΔGm,3 = ΔGm,2 = RT ln
p0
(6.3)

The second method only has one step: 1 mol of planar liquid
at p0 scatters into small liquid droplets with radius r. Since there
is an additional pressure, the scatter process occurs under constant
temperature and inconstant pressure. The liquid pressure in liquid
droplets is (p + Δp). The additional pressure is Δp = 2γ/r. If we
ignore the influence of liquid mole volume Vm l , the Gibbs free energy

change for this method is


 (p+Δp) l
l l 2γVm
ΔGm,b = Vm dp = Vm Δp = (6.4)
p r
If the density ρ and molar mass M of the liquid are known,
l = M/ρ into equation (6.4) leads to
substituting Vm
2γM
ΔGm,b = (6.5)
ρr
According to the constant chemical potential energy during phase
change at saturation, the chemical potential energy of the liquid
droplets is equal to the chemical potential energy of the saturated
vapor around them. The chemical potential energy of the planar liq-
uid is equal to the chemical potential energy of the saturated vapor
at the liquid surface. Therefore, the potential difference between the
small droplets and the planar liquid should be equal to the potential
difference between the saturated vapor around the droplets and the
saturated vapor around the planar liquid surface, which means the
two potential differences above are the same. Thus,
Δμg = Δμl (6.6)
i.e.,
  l
pr 2γVm
RT ln = (6.7)
p0 r
Kelvin Equation 133

If mole volume is substituted for mole mass M and liquid density


ρ, then Vm l = M ; substitute it in the above equation:
ρ

 
pr 2γM
RT ln = (6.8)
p0 ρr

This is the Kelvin function.


The small spherical droplets show a convex liquid surface to the
planar liquid. On the contrary, the liquid forms a concave menis-
cus surface in the pore. At this moment, the pressure inside the
liquid is less than the pressure outside the liquid. Correspondingly,
the saturated vapor pressure is less than that of the planar liquid.
As shown in Figure 6.1, when liquid forms a concave surface in a
pore, the saturated vapor pressure pr of the concave surface with a
radius r is less than the saturated vapor pressure p0 of the planar liq-
uid. Liquid pressure for the concave surface is p−2γ/r. The potential
difference between the concave liquid and the planar liquid is
 p−Δp l
l l 2γVm
Δμ = μr − μ = Vm dp = − (6.9)
p r

The potential difference between the corresponding saturated


vapor pressure and the planar saturated vapor pressure is
 pr  pr
g g RT pr
Δμ = Vm dp = dp = RT ln (6.10)
p0 p0 p p0

Figure 6.1. The concave meniscus formed by the liquid in the pores.
134 Capillary Mechanics

By the constant relation between the potential differences Δμl =


Δμg , we have

p0 2γM
RT ln = (6.11)
pr ρr

This is the Kelvin function for the capillary pore, which indicates
that the smaller the capillary, the smaller the saturated vapor pres-
sure balanced with that of the liquid.
Since the two methods lead to the same result, the change in
the total Gibbs free energy should be the same, which is ΔGm,a =
ΔGm,b . Then,
 
pr 2γM
RT ln = (6.12)
p0 ρr

The two methods are shown in Figure 6.2.


The Kelvin function shows that if the temperature is certain,
saturated vapor pressure pr of a small droplet is a function of its
radius r.
For a convex surface and small liquid droplets, r > 0, ln(pr /p0 ) >
0, which means that the saturated vapor pressure of droplets is
greater than that of planar liquid under the same temperature.

Figure 6.2. The relation between the two methods.


Kelvin Equation 135

For a concave surface and little bubbles in water, r < 0. Therefore,


 
pr 2γM
RT ln = <0 (6.13)
p0 ρr

ln(pr /p0 ) < 0, pr < p0 .


Under the same temperature, the vapor pressure of liquid around
the bubbles is less than the saturated vapor pressure of the planar
liquid surface.

6.2 Application of the Kelvin Function

The Kelvin function can explain many phenomena and can be applied
to many conditions, such as capillary pore condensation, artificial
rainfall, and calculating the radius of falcate surfaces.

6.2.1 Capillary pore condense


From the application of the Kelvin function in concave liquid
surfaces, we can find that the liquid pressure and the vapor pres-
sure of capillary pores are less than those of a planar liquid. There-
fore, under certain conditions, the gas in capillary pores is easier to
condense than that outside of the capillary pores. Hence, capillary
cohesion occurs when the vapor pressure is less than the saturated
vapor pressure P0 , which explains why, when humidity is very high,
porous media is easier to condense gas into liquid. An old saying
illustrates the principle that the condensed water around a tank indi-
cates rainy weather. Since soil and fibrous tissue have many capillary
pores and water wetted in capillaries shows a concave surface, the
planar liquid that has not been saturated begins to condense in the
capillaries. Therefore, the capillary structure of soil can help retain
water. The radius and distribution of porous solid catalysts work with
the capillary condensing function, which can be calculated using the
Kelvin function.

6.2.2 Artificial rainfall


According to the application of the Kelvin function in concave liquid
surfaces, the liquid pressure and the vapor pressure of capillary pores
136 Capillary Mechanics

are less than those of a planar liquid. The smaller the droplet radius
(size), the bigger the pressure difference. Therefore, small droplets
are easier to condense than planar liquid. From another point of
view, if the gas humidity is very large, the density and pressure of
saturated vapor are very large, and the saturated vapor of planar liq-
uid is easier to condense into liquid. In the air, if the spherical liquid
droplets are not spherical, which means the radius of spherical liquid
droplets is zero, theoretically, the pressure difference of the inside
and outside of the droplet is infinite. Therefore, it is difficult for the
gas to condense to liquid directly. However, if there is an initial liquid
droplet, then there is a radius of the liquid droplet, and correspond-
ingly there is a saturated vapor pressure. When gas pressure in the
air reaches the saturated vapor pressure, ambient gas will condense.
With the increase in clotted liquid, the radius of the liquid droplets
increases. Correspondingly, the saturated vapor pressure decreases,
which implies that the condensation is easy. Therefore, in order to
let the gas in the air condense, the initial droplet is the key. Without
the initial droplet, no matter how much the humidity is and how
large the pressure is, it is difficult to form a liquid droplet. When
there is an initial droplet, under certain condition, gas will condense
to liquid droplets. By utilizing this principle, when the sky has a
certain meteorological condition or when the steam in clouds reaches
the saturated conditions, the initial droplets can be formed by spray-
ing micro-granules into the clouds from an airplane. It can greatly
decrease the saturated degree of liquid droplets, and it becomes easy
for the steam in the clouds to condense into liquid and fall to earth.
In actual cases, the steam will not condense into the liquid phase
when it just reaches the saturated condition. The state of the steam
needs to exceed a certain condition, which is the supersaturated state.
Generally, the steam condenses into liquid in a certain supersaturated
state. The saturated state is a stable state, and the supersaturation
state is a kind of metastable state.

6.2.3 Calculating meniscus radius


According to the Kelvin function, when pressure, the saturated vapor
pressure of a planar liquid, and that around the concave liquid surface
are known, as well as the temperature, the mole volume of liquid,
and surface tension, the following can be deduced using the Kelvin
Kelvin Equation 137

function:
2γVml
r= (6.14)
RT ln ppr0

When there is a kind of liquid between two objects, the liquid


will form a liquid bridge, and the sides of the liquid bridge will form
a meniscus surface. This curved surface possesses a curvature and a
curvature radius. Utilizing the Kelvin equation, the values associated
with curvature or the radius of curvature can be determined.

6.3 Metastability State and the Production


of a New Phase

From the Kelvin function, the saturated steam pressure of small


spherical liquid or air bubbles is inversely proportional to the spher-
ical radius. The smaller the radius, the bigger the saturated vapor
pressure. In theory, when the radius is zero, the saturated vapor pres-
sure becomes infinite, which creates the problem that it is difficult to
produce a new phase. In phase-change processes, such as steam con-
densation, solidification of pure liquid substances, and crystallization
of solute in solution, particles which produce a new phase are very
tiny and the saturated vapor pressure is very large. Therefore, it is
difficult to produce a new phase in the system, and supersaturation
is subsequently induced.

6.3.1 Supersaturation vapor


The reason for the existence of supersaturated vapor is that the newly
generated tiny particles of saturated vapor are greater than those of
planar liquid. If the degree of supersaturation of the liquid is not
high and micro liquid droplets do not reach a saturated state, the
droplets cannot be generated or exist. According to the usual phase-
balance conditions, the steam should condense. This steam that does
not condense is called supersaturated steam. For example, near 0◦ C,
sometimes vapor begins to automatically condense when its pressure
reaches five times the balance of vapor pressure.
When there is dust in the steam or the inside surface of the con-
tainer is rough, these materials can act as sites of devaporation and
138 Capillary Mechanics

make liquid droplets produce or grow easily. Even if the supersatura-


tion degree of the liquid is very small, steam begins to condense. The
principle of artificial rainfall shows that, when steam in the cloud
reaches its saturation or supersaturation state, we can scatter tiny
AgI particles in the cloud. Then, the AgI particles will act as sites of
devaporation and cause the supersaturation degree of the new phase
to decrease dramatically. Then, water in the clouds easily condenses
into droplets and falls to the earth.

6.3.2 Overheating liquid


Under air pressure, when a liquid is boiling, gas not only proceeds on
the surface of the liquid but also inside the liquid. If the liquid does
not contain materials that can offer new phase seeds (little bubbles),
even if the liquid is heated to above its boiling point, it will not boil.
Therefore, based on phase equilibrium conditions, this kind of liquid
is called an overheated liquid. The reason for overheating is that new
phase seeds are difficult to generate in the liquid.
In order to prevent overheating, we often use dry enamelware or
capillaries that contain gas because these materials can store gases.
When heated, these materials release small bubbles, which help over-
come the difficulty of producing new phase seeds. Therefore, it dra-
matically decreases the degree of overheating.

6.3.3 Subcooled liquid


Under a certain pressure, when liquid materials are cooled to their
freezing point, new solid-phase particles will be generated based on
the phase equilibrium condition. However, the newly generated par-
ticles are very small, and their melting point is lower. Now, small
crystals do not reach their saturation state, so they cannot be pro-
duced automatically or exist. The liquid must continue to be cooled
to below the normal freezing point and reach the freezing point of
the small crystals. Then, the crystals will separate out. Therefore,
based on the phase equilibrium condition, this kind of liquid is called
a subcooled liquid. For example, pure water cannot freeze even if
slowly cooled to −40◦ C. Put some small crystals in the subcooled
liquid as new phase seeds; the liquid will condense quickly.
Kelvin Equation 139

6.3.4 Supersaturated solution


If a solution undergoes evaporation at a constant temperature and
specific pressure, the density of the solution will gradually increase.
When the density reaches the saturated density of ordinary crystal
solutes, according to the equilibrium condition, there should be crys-
tal precipitation. However, since tiny crystal solutes have greater sol-
ubility, the solutes of tiny crystals do not reach their saturation state.
Therefore, there is no crystal precipitation. The liquid must continue
to evaporate until it reaches its saturation state. Then, there will
be crystal precipitation. According to the equilibrium condition, this
kind of liquid is called a supersaturated solution.
During the process of crystallization, when the liquid is evapo-
rated to a certain saturation state, put some small crystals in the
crystallization system as new phase seeds. Thus, larger crystals can
be obtained.
According to thermodynamics, all supersaturated systems are not
real balanced systems and in steady state. Therefore, they are called
metastable states. However, these kinds of systems can continue to
exist without phase changes for a long period of time. It is because
under a certain condition, new phase seeds are very difficult to be
generated. For example, metal quenching involves heating a kind of
alloy product to a certain temperature. The temperature is main-
tained for a while; then, the alloy product is quickly cooled in water,
oil, or other kinds of medium. Under normal temperature, the prod-
uct can maintain its structure as it does in high temperatures. In
this way, the performance of metal products can be improved.
This page intentionally left blank
Chapter 7

Surface Tension Gradient


and Marangoni Effect

7.1 Surface Tension Gradient

The Young–Laplace equation reveals that for a concave liquid


surface, the internal pressure within the liquid is lower than the
external pressure, resulting in a state of negative pressure. This neg-
ative pressure generates a tensile force in the surrounding medium,
which correlates directly with the surface tension and also directly
with the curvature of the surface (inversely proportional to the radius
of curvature). Consequently, as surface tension increases, so does the
tensile force exerted on the surrounding medium.
For one kind of liquid, if the physical and chemical material
properties (such as material constitution, density, concentration,
and temperature) are uniform, it will form a single homogeneous
phase. However, the stringent definition of a phase complicates the
formation of a truly homogeneous phase. Distinct materials can yield
different phases, while identical materials may not necessarily result
in the same phase due to variations in density, concentration, or
temperature. Consequently, even with identical compositions, dis-
crepancies in these properties can lead to the emergence of multiple
phases. The presence of different phases is characterized by an inter-
face, which is associated with interfacial free energy and interfacial
tension. In a liquid material, any variation in density, concentration,

141
142 Capillary Mechanics

or temperature results in an overall non-uniformity, thereby giving


rise to distinct phases. If certain physical quantities, such as den-
sity, concentration, and temperature, exhibit continuous variations
within the bulk of the material, gradients will emerge — specifically,
density gradients, concentration gradients, and temperature gradi-
ents. These physical gradients subsequently induce phase gradients
and interfacial gradients, which can lead to gradients in surface ten-
sion or interfacial tension. Interfacial or surface tension gradients
may manifest within the bulk of the material or along its surface.
In a strict sense, interface tension present in the space manifests on
the different interfaces in the bulk space. For example, because of
different concentrations, a material with the same component will
have multiple phases and interfaces. Each interface corresponds to
distinct interfacial tensions, reflecting spatial changes in interfacial
tension and resulting in an interfacial tension gradient. Similarly,
surface tension gradients on a surface are primarily characterized by
differing surface tensions across various regions of the same surface.
Variations in temperature can lead to disparate temperature zones
on an identical surface, which may subsequently result in changes in
surface tension across the surface, thereby forming a surface tension
gradient.
The phenomenon of surface tension indicates that liquids with
higher surface tension exert a more substantial pulling force on their
surrounding medium compared to those with lower surface tension.
From a spatial perspective, the presence of a surface tension gra-
dient implies a variation in surface tension across different regions.
Specifically, areas characterized by elevated surface tension exhibit
a stronger cohesive force, leading to the tendency of the liquid to
migrate toward these regions while distancing itself from areas of
lower surface tension. As previously mentioned, such surface ten-
sion gradients can arise from concentration gradients. For instance,
in wine, the concentration of alcohol is not uniform throughout the
liquid, creating both a concentration gradient and a corresponding
surface tension gradient. Additionally, temperature gradients can
also induce surface tension gradients. For example, if a liquid sur-
face experiences varying temperatures, this temperature gradient will
influence the surface tension, thereby resulting in a gradient in sur-
face tension across the liquid.
Surface Tension Gradient and Marangoni Effect 143

7.2 Marangoni Effect

Given that a liquid with high surface tension exerts a stronger pull on
the surrounding liquid compared to one with low surface tension, the
existence of a gradient in surface tension will inherently induce the
flow of liquid away from areas of lower surface tension. The surface
tension gradient can be caused by a concentration gradient or a tem-
perature gradient (surface tension is a function of temperature),
which is called the Marangoni effect. This phenomenon was first
identified in the so-called “tears of wine” by physicist James Thom-
son (Lord Kelvin’s brother) in 1855. In a presentation at the Royal
Academy of Sciences in the United Kingdom, Thomson described
convection resulting from differences in surface or interfacial tension.
He demonstrated this by introducing a small quantity of alcohol onto
the surface of water, which resulted in a rapid surface movement
from the point of alcohol introduction to the surrounding area. Sub-
sequently, Italian physicist Carlo Marangoni investigated this phe-
nomenon for his doctoral dissertation at the University of Pavia,
publishing a related thesis in 1865 and conducting numerous studies
on similar effects. Although both Thomson and Marangoni attributed
the observed convection to variations in alcohol concentration affect-
ing surface tension, the phenomenon ultimately became known as
the Marangoni effect. Because most of the early treatment work per-
formed was by Willard Gibbs, sometimes it is also referred to as the
Gibbs–Marangoni effect. The reason for this effect is that liquid with
high surface tension has stronger tension than the liquid around it.
Therefore, the surface tension gradient will cause the liquid to flow
from regions with low surface tension to regions with high surface
tension. The Gibbs–Marangoni effect occurs as a result or a process.
From the view of a result, it reflects one kind of effect. As a process,
it reflects that liquid flows from regions with low surface tension to
regions with high surface tension. Since the source of the effect is
a surface tension gradient, if there is a surface tension gradient, the
Marangoni flow exists. This flow can facilitate mass and heat transfer,
forming a circulatory flow of a certain region, and is often referred to
as Marangoni convection. Change in convection speed should occur
with surface tension gradient. At the present time, mass transfer
and heat transfer have extensive functions in science and technology.
Therefore, much more research attention is focused on these.
144 Capillary Mechanics

(a) (b)

Figure 7.1. Phase boundary of the same component liquid with different con-
centrations and temperatures: (a) concentration gradient leads to interface tension
gradient; (b) temperature gradient leads to interface tension gradient.

In the context of a liquid film, the Marangoni effect refers to the


phenomenon whereby disturbances in the external environment, such
as variations in temperature and concentration, lead to a reduction in
thickness in certain areas of the liquid film. This alteration induces
a Marangoni flow, driven by a gradient in surface tension, which
facilitates the movement of the liquid toward the thinner regions
along the most favorable path. Specifically, in liquid systems, the
Marangoni effect indicates that the regions with high surface tension
have tension against the regions with low surface tension, resulting
in a liquid flow directed toward the high-surface-tension zones, as
illustrated in Figure 7.1.
The Marangoni effect encompasses a variety of applications across
different contexts. The most common examples are soap membrane
and the tears of wine effect in a wine glass. In the case of soap
bubbles, the Marangoni effect plays a crucial role in stabilizing the
soap film. Similarly, in the context of wine tears, this effect facili-
tates the movement of high-concentration wine liquid. Furthermore,
the Marangoni effect is significant in the processing of integrated cir-
cuits, particularly in the drying of silicon chip surfaces following wet-
processing procedures. The presence of liquid droplets on the silicon
surface can lead to oxidation, which may compromise the integrity
of the material. In order to avoid forming falcate liquid bridges or
droplets on the wet silicon surface, isopropyl alcohol (IPA) is dis-
persed as an aerial fog or stream on the wet silicon surface. This
application leverages the Marangoni effect to create a gradient in
surface tension, resulting in a Marangoni flow. Under the influence
of gravity, this process enables the detachment of droplets from the
silicon surface, ultimately leading to a dry silicon surface.
Surface Tension Gradient and Marangoni Effect 145

7.3 Phenomenon Explained by Marangoni Effect

7.3.1 Soap bubble


As shown in Figure 7.2, a soap bubble is a very thin film of soapy
water which forms a sphere with an iridescent surface. Soap bubbles
usually last for only a few seconds before bursting, either on their
own or by coming into contact with other objects. Soap bubbles
are often used to entertain children. They are also used in artistic
performances. Soap bubbles can help solve complex mathematical
problems of space, as they will always find the smallest surface area
between points or edges.
A soap bubble is able to exist due to the surface tension of a liquid,
typically water, which allows the surface layer to function similarly
to an elastic membrane. Soap films exhibit significant flexibility and
can generate waves in response to applied forces. However, a bubble
composed solely of a single pure liquid lacks stability and requires
the presence of a dissolved surfactant, such as soap, for stabiliza-
tion. A common misconception is that soap increases the surface
tension of water. Actually, the opposite is true. Approximately, soap
decreases one-third of the surface tension of pure water. Rather than
reinforcing bubbles, soap stabilizes them through the mechanism of
the Marangoni effect. As the soap film is stretched, the concentration
of soap at the surface diminishes, resulting in an increase in surface
tension. Consequently, soap selectively fortifies the weakest regions

Figure 7.2. A soap bubble.


146 Capillary Mechanics

of the bubble, thereby preventing excessive stretching in any particu-


lar area. Additionally, soap contributes to a reduction in evaporation
rates, which prolongs the lifespan of the bubbles, although this effect
is relatively small.
The spherical shape of a bubble is also caused by surface tension.
For a given volume, the sphere has the smallest possible surface area,
which takes on a spherical shape and minimizes the free surface of a
bubble.

7.3.2 Tears of wine


As shown in Figure 7.3, a clean and transparent glass is filled with
wine. Upon remaining stationary for a duration, it was noted that
the upper portion of the wine glass, which had not been in contact
with the wine, retained its clarity and transparency. However, there
were small droplets continuously forming and dropping back into the
liquid near and above the wine level in the glass. This phenomenon
is called tears of wine. It is most readily observed in a wine which
has a high alcohol content. It is also called wine legs.
The observed phenomenon can be attributed to the lower surface
tension of alcohol in comparison to that of water. When alcohol is
homogeneously mixed with water, areas with a lower concentration of

Figure 7.3. Tears of wine.


Surface Tension Gradient and Marangoni Effect 147

alcohol exert a greater attractive force on the surrounding fluid than


areas with a higher concentration of alcohol. The result is that the
liquid tends to flow away from regions with higher alcohol concentra-
tions. This can be easily and strikingly demonstrated by spreading a
thin film of water on a smooth surface and then allowing a drop of
alcohol to fall on the center of the film. The liquid will rush out of
the region where the drop of alcohol fell.
Wine is composed of a combination of alcohol and water, along
with dissolved sugars, acids, colorants, and flavor compounds. When
the surface of the wine comes into contact with the interior of the
glass, capillary action causes the liquid to ascend along the glass’s
surface. Both alcohol and water evaporate from this ascending film;
however, alcohol evaporates at a faster rate due to its lower boiling
point. This reduction in alcohol concentration results in an increase
in the surface tension of the liquid, which in turn facilitates the
upward movement of additional liquid from the main body of the
wine, characterized by a lower surface tension owing to its higher
alcohol content. As the wine ascends the glass, it forms droplets
resembling tears, which subsequently fall back due to gravitational
forces. This phenomenon is commonly referred to as the “tears of
wine”. Additionally, the elongated appearance of the wine as it climbs
the glass is often described as the “legs of wine”.

7.3.3 Wafer drying in wet processing


In the sequence of processes involved in the “wet process”, drying
represents the final and critical step. The drying step contributes
much to the wet process performance, perhaps even defining its
results. If drying is not well done, the chip cannot perform well.
Therefore, the drying process is the most important process in the
wetting process. A pivotal consideration is the selection of an appro-
priate drying method. Spin dryers and IPA are commonly used
today. However, with the development of micro-technology and the
requirements of increasingly high performance of chips, the size of
the structure in the chip is getting smaller and smaller, and smaller
particle sizes have come into focus. In terms of strength, the chip
should not and cannot withstand too much mechanical stress during
the processing, and residual stress is not allowed after processing.
148 Capillary Mechanics

Hence, during any process, the mechanical stress should be com-


pletely prevented from being introduced to the wafers or should be
reduced to as small as possible. Therefore, the method of spin drying
by a rotating machine is obviously not suitable. This more or less
spells the end of spin dryers for advanced applications. Although the
IPA vapor drying method is effective, it poses many safety hazards
due to the need to heat the alcohol and has high requirements for
fire prevention. Equipped with a large number of CO2 fire suppres-
sion systems, the cost is also high and generally difficult to afford.
Therefore, it is not very suitable for such dryness. Of course, there
are also high-energy ozone drying methods, but they are not very
suitable for a wide range of chip or material types. Based on analy-
sis, the only technology that already has an inherent potential for the
future is the drying technology based on the Marangoni effect. This
drying technology has been well established over the past decade
and accepted industry-wide. However, the commonly used technical
solutions for this drying technology have also reached their limits.
The low amount of IPA currently available within the process cham-
ber and the inability to change it on a recipe base requires changes
and improvements, especially for larger wafer sizes and smaller CDs.
Along with the development of etching technology, crystalline chips
need to be processed more carefully and dried thoroughly. Through
process improvements, it is possible to further enhance the drying
technology based on the Marangoni effect.
Alcohol vapor is imported into the dry chamber through atomiza-
tion. The aerosol is created by an ultrasonic oscillator, which is oper-
ated on a specific resonant frequency and primarily defined by the
mechanical length of a nozzle. By optimizing the minimum spray crit-
ical power, the best drying result can be obtained. Another way is to
adjust the drain speed. The traditional drain speed is 0.3–2.5 mm/s.
Today, the typical setting is 1–1.5 mm/s. With the availability of
different drain speeds over the cross-section, different wafer surface
area regions can be treated separately. The water columns on top
of the wafers can be drained faster to gain time. The processes run
in an ambience filled with N2 , which provides an inert atmosphere.
With the very low alcohol utilization during the process, the IPA
concentration stays far below its flame point, and this new technol-
ogy avoids safety risks. Therefore, the adapted processes can not only
avoid safety risks but also improve the drying efficiency.
Surface Tension Gradient and Marangoni Effect 149

7.4 Marangoni Convection

In microgravity environments, the significance of surface and


interface tension is markedly pronounced. The presence of a surface
tension gradient can directly induce Marangoni flow. Variations in
temperature and concentration gradients can readily alter surface
and interface tension. During the gas–liquid phase transfer process,
the exchange of materials between these two phases results in changes
to the surface tension of the liquid. When the surface tension gradi-
ent of the liquid exceeds a certain critical value, liquid regions which
are close to the interface will exhibit turbulence. This interface liq-
uid turbulence resulting from the surface tension gradient is called
Marangoni convection. For a long time, it was popularly believed that
Marangoni convection only exists on the interface and its intensity
is small. Consequently, in analyses concerning the factors influencing
actual heat and mass transfer, Marangoni convection is frequently
overlooked. The Marangoni effect is often a subject of nonlinearity
theory research. With further research on microscopic mechanisms
of heat and mass transfer, it was found that the Marangoni effect
cannot be ignored. In many heat and mass transfer processes, an
abnormal phenomenon related to the transfer speed is ascribed to
the Marangoni effect. In the fields of crystal production, pharma-
ceuticals, and metallurgy, the Marangoni effect resulting from heat
and mass transfers has an important influence on the purity and
quality of production. Therefore, using the Marangoni effect is a
very efficient way to improve transfer effect and optimize produc-
tion quality. In addition to traditional heat and mass transfer, the
research scope of Marangoni convection has been extended to vari-
ous fields. For example, in the process of deploying buffer solution
for contact lenses, adding materials that induce the Marangoni con-
vection effect can impel the corneal liquid to be refreshed and sup-
plied continuously, thereby avoiding ophthalmodynia and drying the
eyes. Lyford did research on potentiating of Marangoni effect to oil
exploitation in petroleum of porous rock. In addition, Marangoni
convection has a very important effect on many processes, such as
welding and wafer drying, and also in space labs, such as heat trans-
fer and life support systems in space, handling of material in space
without containers, crystal growth, location welding, surface liquid
fuel, gas desorption, and homogenization. Extensive experimentation
150 Capillary Mechanics

demonstrates that Marangoni convection has an important influence


on the flow state of an interface. It can promote the update of the sur-
face liquid unit and enhance the transfer speed. The transfer speeds
associated with Marangoni convection can be several times greater
than those observed in the absence of this phenomenon.
Research indicates that a surface tension gradient can lead to
the Marangoni effect. The formation of Marangoni convection also
significantly depends on the surface tension gradient of mass transfer.
The surface tension gradient determines the intensity of Marangoni
convection. Since mass transfer is accompanied by remarkable inter-
face surface tension changes, Marangoni convection caused by the
transfer also exists in many mass transfer processes, and its intensity
is quite considerable. When Marangoni convection occurs, according
to the actual measurement by Vazquez and Lu, the interface flow
speed can reach 0.1–0.6 m/s, and Marangoni convection observably
enhances the transfer speed.
Currently, experimental approaches primarily focus on measuring
the velocity of liquid surfaces or the macroscopic mass transfer
coefficient, thereby validating the impact of Marangoni convection
on macroscopic properties and mass transfer intensity. Details of
Marangoni convection are difficult to get. Essentially, Marangoni con-
vection in mass transfer is a kind of flow coupling by momentum and
mass. Through fluid computations, we can avoid the difficulties of
experimentation and resolution-solving and obtain simulation infor-
mation about the Marangoni convection field. However, most of the
simulations now depend on simplified concentration boundary condi-
tions. Therefore, the change in surface tension with concentration in
interfaces cannot be expressed, and the Marangoni convection field
information cannot be extracted. Hence, until now, researchers have
not been able to study the Marangoni effect completely, so the further
application of the effect is limited to a certain degree. However, con-
sidering the huge practical and theoretical value of the Marangoni
effect, it is highly necessary to study it further.
Marangoni convection, influenced by the induction factor acting
on the surface tension gradient, can be categorized into two distinct
types: temperature-driven and concentration-driven convection. The
flow resulting from a concentration gradient is referred to as solute
Marangoni convection, while that arising from a temperature gradi-
ent is known as heat capillary convection.
Surface Tension Gradient and Marangoni Effect 151

7.5 Flow State of Marangoni Convection

In order to characterize the Marangoni effect and convection, a


dimensionless Marangoni number, Ma, is often used. It is found from
experiments that the flow state of Marangoni convection consists of
a laminar flow and an oscillation flow. What can decide the flow
type is M a. If is less than a certain critical value, M a < M acr , its
flow state would remain as laminar flow. But if M a is greater than
a certain critical value, M a > M acr , the stable flow state would
disappear. After a transitory stage, flow acted on by surface tension
would become a kind of three-dimensional flow. The flow type can
become a periodic quantity, which is related to time and space. This
flow state is called oscillation flow.

7.5.1 Flow state of laminar flow


When a liquid or solution is introduced between two closely
positioned coaxial circular disks, a liquid bridge is established.
A temperature gradient can be applied within the space between
the upper and lower disks, aligned with the axis of the disks. In the
presence of a gravitational field, the axis of the liquid bridge should be
oriented in the direction of gravity. The environment surrounding the
liquid bridge may consist of liquid, gas, or vacuum. When one side of
the solid structure forming the liquid bridge is subjected to heating,
the temperature on that side of the liquid will increase. Given that
heat transfer is a dynamic process, various locations along the two
ends of the liquid bridge or the liquid surface will exhibit differing
temperatures. Since temperature significantly affects surface tension,
this results in variations in surface tension across different regions of
the liquid surface. Consequently, a surface tension gradient is estab-
lished in the spatial domain, which induces flow within the liquid.
If the temperature between the two disks does not exceed a certain
value and M a is less than a certain critical value, M a < M acr , the
flow would keep on being a laminar flow, as shown in Figure 7.4. The
liquid close to the free surface would flow from the high-temperature
region with a low surface tension to the low-temperature region with
a high surface tension.
If the surface tension of one kind of liquid has a minimal value
for temperature and its corresponding temperature is just right
152 Capillary Mechanics

(a) (b)

Figure 7.4. Stable flow pattern (a) top heated and (b) ring heated.

(a) (b)

Figure 7.5. Two kinds of possible steady flow spectra.

between the top and bottom disks, such as the aqueous fatty alcohol
solutions, the stream always flows from all around to the point with
the lowest surface tension, not from the high-temperature regions
to the low-temperature region. Chun found that the stable spec-
trum with a little larger Ma number would form a convective cell,
which contains two cells rotating in the same direction, as shown in
Figure 7.5(a).

7.5.2 Flow state of oscillation flow


The flow state of oscillation flow can be seen as a superposition or
synthesis flow by periodic radial and circumferential motion on a
vertical axis plane (Figure 7.6).
Surface Tension Gradient and Marangoni Effect 153

Figure 7.6. Steady flow spectrum perpendicular to the axis.

Chun and Wuest found from experiments that oscillation flow


has two oscillation patterns. A decisive role is played by the scale
ratio Ar. When Ar ≥ 0.5, the flow pattern is non-symmetrical; when
Ar ≤ 0.45, it is symmetrical. Chun observed that the flow is along the
circumference of the liquid bridge. Non-symmetrical oscillation flow
is actually an oscillation cell. Symmetrical flow has two oscillation
cells. Experimental results reported by Preisser confirmed the results
from Chun, and at the same time, they pointed out that only when
nλ = πD, oscillation exists in the liquid bridge, where λ is the cir-
cumference wavelength and n is an integer number, called modulus.
A certain scale ratio corresponds with a certain modulus. A smaller
scale ratio corresponds to a bigger modulus. When n = 1, it corre-
sponds to the non-symmetrical flow, as referred to by Chun. n = 2
represents symmetrical oscillation flow. However, if the modulus is
greater than 2, the flow state cannot be fixed. It is easy to change
from one modulus to another. Flow patterns for these two stable
modulus oscillation flows are shown in Figure 7.7. When Ar ≤ 0.5,
there exists an empirical relationship for the modulus, and the scale
ratio is nAr = 4.4.

7.6 Research Method of Marangoni Effect

Research methods for the Marangoni effect encompass both


theoretical analysis and measurement. In theoretical methods,
researchers attempt to derive corresponding control functions based
on the micro-mechanisms underlying Marangoni convection. Since
the flow is related to nonlinear solutions, obtaining them is
154 Capillary Mechanics

Figure 7.7. Two stable oscillation flows: (a) Module n = 1, nonsymmetrical


oscillation flow and (b) n = 2, symmetrical oscillation flow.

extremely challenging. Therefore, linear methods are often employed


to investigate the critical state, while nonlinear methods are used to
study conditions near the critical state. With the advancement of
computers and computational techniques, numerical methods have
become an efficient tool for solving nonlinear problems. In experi-
ments, there are generally two methods. One involves observing the
phenomenon directly, while the other involves indirectly measuring
the relevant parameters.

7.6.1 Linear analytical method


In general, a set of nonlinear partial differential equations, including
the Navier–Stokes equation, continuity equation, and mass and heat
transfer equation, can characterize the Marangoni phenomenon.
Therefore, the Marangoni phenomenon is a nonlinear process that is
mathematically challenging to analyze. Based on the linear stability
Surface Tension Gradient and Marangoni Effect 155

theory of small perturbation analysis, the nonlinear term can be


ignored for linearization, thereby reducing the difficulty of solving
the equations. Linear stability theory serves as an efficient means
for forecasting and analyzing the Marangoni phenomenon and is the
foundation for subsequent nonlinear analysis. Many researchers have
conducted extensive work on this basis. Sternling and Scriven were
the first to theoretically investigate the impact of interface phenom-
ena on mass transfer. They postulated that interface convection near
the interface is caused by the non-uniform distribution of solute at
the interface and predicted the existence of a convective structure
resembling the rotary drum convection observed in heated liquid lay-
ers. Pearson conducted research on the convective conditions driven
by surface tension using linear stability theory. Later, Brian con-
sidered the influence of the Gibbs adsorbed layer at the interface
and extended Pearson’s work. Since Brian’s boundary conditions
closely match real-world conditions, they are often used to study
the Marangoni phenomenon driven by mass transfer.

7.6.2 Nonlinear analytical method


Due to the limitations of linear analytical methods, research on the
Marangoni effect is confined to forecasting and analyzing the Ma
number. Linear theory can explain convection structures, the trans-
fer of supercritical flow types, and processes that change with time
and space. Near the critical point and in supercritical regions, the
nonlinear term in the governing equation has an effect that cannot
be ignored. Therefore, nonlinear mathematical analysis is necessary.
Since nonlinear analysis is challenging, it is often simplified as a
weak nonlinear problem, focusing only on the nonlinear behavior
near the critical point. At the same time, real-world applications
are also substantially simplified by neglecting complex boundary
conditions. Nonlinear theoretical research on Marangoni convection
primarily focuses on forecasting convection structures and analysis.
The results can be compared with observed structures to verify the
method’s validity. Scanlon and Segel were the first to investigate the
nonlinear evolution of Marangoni convection in a semi-infinite atmo-
sphere liquid–gas interface with a non-deforming interface, derived
by heat transfer. Cloot and Lebon considered a more realistic and
156 Capillary Mechanics

finite-depth liquid. In subsequent research, they found that the char-


acteristic size of the convection structure is of the same order of
magnitude as the liquid depth. Bestehorn studied the evolution of
flow patterns in Rayleigh−Bénard−Marangoni convection, revealing
the specific structure of Marangoni convection. Nonlinear analysis
by Hadji showed that there are bistable regions with hexagonal or
square patterns in supercritical regions. A flat interface is a common
assumption for boundary conditions. However, Marangoni convection
occurs at the interface, so interface deformation is a crucial charac-
teristic that plays a significant role in the initiation and evolution
of the process. Golovin considered the influence of interface defor-
mation and predicted Marangoni convection patterns derived from
temperature changes. The boundary conditions for heat transfer-
induced Marangoni convection are relatively simpler than those for
mass transfer-induced Marangoni convection, which does not need to
account for the complex materialization of the interface adsorption
layer. The application of nonlinear analysis to mass transfer-induced
Marangoni convection is limited. Bragard investigated nonlinear
problems related to Marangoni instability using weak nonlinear the-
ory when a non-deforming liquid surface is absorbing liquid. They
believed that a hexagonal flow structure is more stable than a rotary
drum structure. Ignoring the existence of the Gibbs adsorbed layer,
he studied instability problems in mass transfer-induced Marangoni
convection, which includes heat effects, using the bifurcation method.
By analyzing and combining nonequilibrium thermodynamics, Zhifa
Sun proposed that Marangoni convection derived from mass transfer
is a nonequilibrium phenomenon. It is worth emphasizing that in
nonlinear theory, the wave vector plays a decisive role in selecting
the flow pattern. However, since it does not account for the mate-
rialization of the liquid, it is difficult to experimentally validate the
theoretical results. Approximate solutions to nonlinear problems near
the critical point can be obtained. Due to the lack of real supercrit-
ical experimental data for comparison, the results cannot be easily
extrapolated. At the same time, it is challenging to explain the bifur-
cation phenomena in supercritical regions. Although the theory of
the Marangoni effect has not yet been fully developed, compared to
real-world conditions, researchers can still obtain considerable useful
information for further study.
Surface Tension Gradient and Marangoni Effect 157

7.6.3 Numerical calculation


With the development of modern computers and numerical
simulation, it is possible to solve nonlinear numerical problems
by using computer simulation. This avoids the need for analytic
solutions of complicated nonlinear mathematical equations and also
allows for the consideration of a number of real boundary condi-
tions. Boyadjiev conducted research on the process of nonlinear mass
transfer accompanied by Marangoni convection using the finite dif-
ference method. Zemei Tang researched liquid flow driven by the
Marangoni effect on the liquid–gas surface in a liquid bridge, with
the background of a floating zone, using the finite element method.
Galazka studied the influence of Marangoni flow on the growth of oxi-
dation chips. Bestehorn obtained the evolution pattern of flow struc-
ture over time, which provides direct evidence for nonequilibrium
phase changes in theory. Numerical calculations can provide detailed
flow information for flow and transfer processes, which remains an
important way to understand Marangoni flow to this day.

7.6.4 Displaying of Marangoni convection field


Appealingly, the Marangoni effect can induce instability in liquid
phase bulk flow and cause relative motion within the liquid. Since
bulk flow disturbances and interface updates occur rapidly, it is not
easy to form an ordered structure, leading to turbulence in the liquid.
Therefore, it is difficult to confirm the presence of Marangoni convec-
tion. When there is no liquid bulk flow disturbance or the disturbance
is minimal, the flow more often exists as a regularly ordered geo-
metric structure. Currently, the flow conditions with the Marangoni
effect can be observed by studying the Marangoni convection field.
Meanwhile, by comparing the Marangoni convection geometric struc-
tures obtained from experiments with those predicted by nonlinear
theoretical analytical methods, the effectiveness of these methods can
be directly proven.
When Marangoni convection occurs, by adding tracers such as
tiny aluminate powder, aluminum foil, or photochromic materials,
the path of the tracer with Marangoni convection can be observed,
and information about the Marangoni convection field can be
obtained. For example, when Schwabe conducted research on
158 Capillary Mechanics

Marangoni instability in a heated liquid layer, both polygonal and


radial flow structures of Marangoni convection could be observed by
using aluminum foil as a tracer. Since the Marangoni effect is a phe-
nomenon, the choice of tracer should not alter the materialization
(materialization is usually used in the context of making something
material or concrete, but here you might mean “manifestation” or
“appearance”) of the interface, so that the flow field can be obtained
without disturbance.
Generally, since light has no influence on the medium, Marangoni
convection at the interface can be observed using a continuous visual
optical system, such as a schlieren system. Then, the Marangoni
convection structure under critical instability conditions can be
obtained. When using optical observation methods, the gas–liquid
or liquid–liquid interface must have a visible phase boundary. This
type of phase boundary is often a horizontal static surface, a hor-
izontal surface below a strictly laminar flow, a liquid film surface
with laminar flow and free from drops, or the interface of immis-
cible phases in droplets. These conditions can help to overcome the
difficulties in observation and calculation caused by turbulent distur-
bances. It is easier to observe interface turbulence phenomena and
more convenient to perform accurate calculations using the strict
fluid mechanics functions of laminar flow.
Using the schlieren system, Orell and Westwater observed the
phase interface during the liquid–liquid extraction process of glycol,
acetic acid, and ethyl acetate and obtained an ordered flow structure.
Imaishi observed interface turbulence phenomena when aqueous solu-
tions of ethanolamine and diethanolamine absorbed CO2 in a wet
wall tower. They found that there were stable and continuous pine
needle-like flow structures. Suciu observed the turbulent state of the
phase interface in liquid–liquid extraction. Zhang conducted research
on liquid–liquid extraction triggered by bi-component spreading
through a liquid–liquid system. Additionally, Okhotsimskii con-
ducted absorbing and desorbing experiments of O2 using 16 different
organic solvents in a static container and classified interface turbu-
lence. When Agble investigated the influence of surfactants on mass
transfer in a liquid–liquid system, he used the observed schlieren
images as direct visual evidence of the impact of interface turbulence
on mass transfer and obtained a lot of information.
Surface Tension Gradient and Marangoni Effect 159

Numerous experiments have clearly demonstrated that the


Marangoni phenomenon exhibits a variety of flow shapes. The
primary flow shapes include the rotary drum and polygon struc-
tures. While the formation mechanism of simpler flow structures,
such as the rotary drum and hexagon, can be explained by weak
nonlinear theory, the mechanisms of some more complicated flow
structures remain unknown. In nonlinear theory, the complexity
increases because the wave vector, which determines the flow shape,
does not account for the material properties of the liquid. In sum-
mary, the complexity of Marangoni convection structures in different
liquids reveals the nonlinear nature of the Marangoni phenomenon.
More detailed experiments are needed to uncover its underlying laws.

7.6.5 Indirect measurement method


The appearance of the Marangoni effect in transfer processes
accelerates the renewal of the liquid’s surface, significantly enhanc-
ing the transfer rate. In both heat transfer and mass transfer, the
prominent feature of the Marangoni phenomenon surpasses that of
the mere diffusion mechanism. Consequently, by observing unusual
transfer coefficients, the Marangoni phenomenon can be indirectly
inferred, which not only circumvents the difficulty of directly observ-
ing the flow field but also allows for the investigation of the impact
of the Marangoni phenomenon on transfer efficiency, thereby aiding
in practical applications.
By measuring the macroscopic average transfer coefficient, Brian
and Imaishi assessed the influence of the Marangoni phenomenon
on mass transfer. The common characteristic of these measurements
is to investigate the impact of interface Marangoni flow on transfer
speed when the surface tension of a material changes during desorp-
tion or absorption from a solution in wet towers or similar equipment.
To eliminate the influence of gravity, mass transfer is conducted in
a vertical plane. Subsequently, the Marangoni effect is generated by
an interface tension gradient. Lu conducted research on the impact
of the Marangoni effect on liquid–gas interface transfer speeds. Tan
investigated the influence of interface turbulence on liquid layer mass
transfer and revised the permeate theory. Straub studied the impact
on heat transfer and explained that higher heat transfer rates are
160 Capillary Mechanics

associated with boiling heat transfer. Many experimental results have


not only demonstrated the enhancement of the Marangoni effect on
transfer processes but also quantified the critical Ma number and the
relationship between transfer speed and Ma number by comparing
measured transfer speeds with common transfer theory. Since the
measured macroscopic transfer coefficient is an average of transfer
coefficients and the Marangoni phenomenon is a microphenomenon
occurring at the interface, information about the microcosmic trans-
fer process cannot be directly obtained.
Laser holographic interferometer is used for microscopic quantita-
tive measurement of the Marangoni flow phenomenon and can mea-
sure the micrometric transfer coefficient. It eliminates inaccuracies
or misleading information arising from macroscopic average trans-
fer coefficients and has considerable advantages. The Mach−Zehnder
method is commonly used in interferometers. The results from this
method are interference fringes. If the interference fringes are pro-
duced by the transfer process, the transfer coefficient can be obtained
from the fringes. Shortcomings of laser holographic interferometers
include the inability to obtain direct visual evidence, expensive equip-
ment and consumables, and complex fringe patterns. Therefore, users
require higher skills. Based on the fundamentals of interferometry,
many researchers have extended this method. One classic extension
is the real-time dynamic four-wavelength mixing laser interferom-
eter proposed by Guzun-Stoica. Able suggested research combin-
ing schlieren and interference methods, utilizing the complementary
advantages of both to obtain evidence of Marangoni flow and the
corresponding transfer coefficient.
In addition to macroscopic parameters, such as the measurement
of transfer coefficients, observation of flow structure and concentra-
tion fields around the interface, and direct observation or measure-
ment of temperature fields, these are very important. They not only
provide evidence for investigating the microscopic mechanism of the
Marangoni effect but also offer practical support for understanding
microscopic mass transfer and heat transfer mechanisms. Since the
Marangoni effect can enhance transfer processes, discoveries from
microscopic experiments are crucial for improving transfer efficiency.
As a phenomenon, the Marangoni effect and flow are directly
related to interface theory and mechanical theory, and their function
can be utilized in various fields. As a factor influencing liquid
Surface Tension Gradient and Marangoni Effect 161

stability, the Marangoni effect occurring at the interface has


significant impacts on chemistry, materials engineering, fluid mechan-
ics, and space science. Current scientific development has blurred the
boundaries of traditional science disciplines. Therefore, the integra-
tion of the strengths of various branches of knowledge has become a
trend. As an intersection of chemistry, physics, mathematics, and
mechanics, studying the Marangoni effect not only enhances our
understanding of nonlinear processes but also lays the foundation
for advancements in other fields.

7.7 Summary of Marangoni Convection Theoretical


Analysis

In the transfer process, one of the important characteristics of the


Marangoni phenomenon is that it leads to fluid mechanics instability
and shapes the main flow of the liquid. Clearly, the Marangoni
phenomenon has a significant influence on the transfer. In partic-
ular, it is related to the microscopic mechanism and can enhance
our understanding of the transfer phenomenon. However, the micro-
scopic instability, the nature of macroscopic flow characterization,
and the evolution process cannot be fully explained through experi-
ments alone; theories are also necessary to enable people to perfectly
understand the essence of the Marangoni phenomenon.
This page intentionally left blank
Chapter 8

Capillary Dynamics

In previous chapters, the phenomenon of capillary rise was


introduced through the analysis of statics in both theoretical and
experimental contexts. However, the analysis only presents the final
result of the rise, without detailing the process itself. To fully
understand this phenomenon, it is essential to describe the rising
process through the lens of capillary dynamics.
Many issues are associated with capillary dynamics, including
the behavior of water, oil, and other liquids as they move through
soil, wood, and various porous materials. This also encompasses the
measurement of the density and porosity of these materials, as well
as the assessment of surface tension and the viscosity of liquids.

8.1 Hagen−Poiseuille Equation

The flow of liquid in a capillary tube can be divided into several


phases, among which the steady phase of liquid development is
particularly important. The fluid in this phase is classified as laminar
flow. Specifically, the fluid flowing through a circular tube with a uni-
form cross-section is referred to as Hagen−Poiseuille flow, which can
be directly analyzed using the Navier–Stokes equations. The Navier–
Stokes equations represent the general equations of fluid motion.
Hagen−Poiseuille flow is a specific type of flow that adheres to certain
assumptions, outlined as follows:

163
164 Capillary Mechanics

1. This flow is a steady flow and belongs to a constant flow.


Therefore, for every flow parameter, its partial derivative of time
is zero, ∂/∂t = 0.
2. The components of flow velocity in both the radial and
circumferential directions are zero; only the component along the
pipeline is non-zero, vr = vθ = 0.
3. The flow is axisymmetric and sufficiently developed. Axisymmet-
ric means ∂/∂θ = 0, and sufficient development means that the
velocity is even, ∂vz /∂z = 0.

Substituting these hypotheses into the Navier–Stokes equation


with polar coordinates instead of Cartesian coordinates leads to
 
1 d dv 1 ∂P
r = (8.1)
r dr dr η ∂z

When the pressure evenly changes uniformly along the circle tube,
∂P/∂z = Δp/l, the above equation can be changed into
 
1 d dv 1 ΔP
r = (8.2)
r dr dr η l

where v represents the flow speed along the length direction of the
tube (the z direction), r denotes the radial coordinate of the fluid
within the tube, and l indicates the length of the fluid column in
the tube. Additionally, η is the viscosity of the fluid, and ΔP is
the pressure difference between the two end surfaces of l, which is
generated by multiple forces.

8.2 Capillary Flow Rate

For a capillary tube, the Hagen−Poiseuille equation can be written as


  
1 P 1 d dv
· = r (8.3)
η l r dr dr

where p represents the total pressure difference between the two
ends of the liquid of length l in the tube. The equation above is an
Capillary Dynamics 165

ordinary differential equation. After integration,



dv 1 P 1
= r +A (8.4)
dr 2η l r

After a second integration,



1 2 P
v(r) = r + A ln(r) + B (8.5)
4η l

A and B are unknown coefficients that can be determined using the


boundary condition. According to the symmetry of the flow speed
boundary and the existence of the speed grads, A = 0. Substituting
it into expression (8.4) leads to

dv 1 P
= r (8.6)
dr 2η l

Because the radiuses of some capillary tubes are extreme small,


velocity slip may occur near the capillary wall. Suppose the slip coef-
ficient is ε. On the boundary,
  
1 2 P ∂v  1 P
v(R) = R +B =ε  =ε R (8.7)
4η l ∂r r=R 2η l

The solution is
 
ε P 1 2 P
B= R − R (8.8)
2η l 4η l

Substituting it into equation (8.5) leads to


  
1 2 P ε P 1 P
v(r) = r + R − R2 (8.9)
4η l 2η l 4η l

where R is the radius of the circular cross-section of the tube. From


expression (8.9), the flow velocity of the liquid has a relationship with
the radius r. In order to describe the velocity of the whole interface
of the fluid, the valid method is to use the concept of flow (volume
166 Capillary Mechanics

flow) and even flow velocity. Calculate the integral of flow velocity
in the section. The volume flow velocity is
 R  R 
1 | P|  2 
V̇ = v · 2πrdr = R − r 2 2πrdr
0 0 4η l
 R 
1 | P|
+ εR 2πrdr (8.10)
0 2η l
After integration,
   
1 | P| 4 1 P 3
V̇ = πR + επ4R (8.11)
8η l 8η l
Then,

π | P|  4 
V̇ = R + ε4R3 (8.12)
8η l
The average flow velocity is the volume flow through unit area,
as given by

dl V̇
= (8.13)
dt πr 2
which is also called the capillary flow rate. Substituting (8.12) into
(8.13) leads to

dl P  4 
= 2 R + 4εR3 (8.14)
dt 8r ηl

In general, the total valid drive pressure P includes three types
of pressures: the non-equilibrium atmosphere pressure PA , the pres-
sure of static water Ph , and the capillary pressure Pc . Expressions
(8.12) and (8.14) are usually called the equation of capillary flow.
This equation can be used to describe the liquid rate not only in
a straight circular capillary tube with an even section but also in
circular tubes of any shape with an even section.
The cross-sectional areas of the capillary tube, as shown in
Figure 8.1, are even, and its radius is R. The capillary tube can
be of any length and shape. A and B are the two ends of the capil-
lary tube. The end A connects with the point which has the distance
Capillary Dynamics 167

Figure 8.1. Capillary tube with arbitrary shape.

of h away from the liquid surface. The end B can be connected to


the atmospheric pressure or can draw off the air from the tube and
be sealed before the end A connects to the liquid. At the begin-
ning of the flow, the resistance is very small, the acceleration is
very great, and the change in flow velocity is very fast. This flow
includes turbulent flow and laminar flow. After a while, the flow will
change into a Poiseuille laminar flow and obey the Hagen−Poiseuille
equation.
In the equation of rate, PA can be considered a constant value;
Ph can be expressed as (Figure 8.1)

Ph = hgρ − ls gρ sin ψ (8.15)

where ls is the straight distance from A to M, ρ is the density of the


liquid, g is the acceleration due to gravity, and ψ is the angle between
the straight AM and the horizontal axis. The capillary pressure Pc is


Pc = cos θ (8.16)
R
where γ is the surface tension of the liquid and θ is the contact angle.
By substituting all the above pressures into expression (8.14), the
expression of the flow rate in the capillary tube can be obtained as


dl PA + hgρ − ls gρ sin ψ + R cos θ  
= R4 + 4εR3 (8.17)
dt 8R2 ηl
168 Capillary Mechanics

Figure 8.2. An inclined straight capillary tube.

In general, ls , ψ, ε, and θ are functions of t. Therefore, θ may be


a function of dl/dt and the pressure gradient of the liquid in the
capillary tube.
Supposing ψ, ε, and θ are all constant values in expression (8.17)
(note that when ψ is a constant value, the tube is not in an arbitrary
shape but an inclined straight tube, as shown in Figure 8.2), in this
case, ls = l. Substituting them into expression (8.16) and integrating
it leads to
 2 
r + 4εr ρg sin ψ · t
+l

PA + ρgh + 2γr cos θ PA + ρg (h − l sin ψ) + 2γ
r cos θ
= ln 2γ
ρg sin ψ PA + ρgh + r cos θ
(8.18)

The two special cases of ψ = 90◦ and ψ = 0◦ are generally con-


sidered. When ψ = 90◦ , expression (8.17) can be rewritten as
 2 
r + 4εr ρg · t
+l

  

PA + ρgh + r cos θ ρgl
=− ln 1 − 2γ
ρg PA + ρgh + r cos θ
(8.19)

In this case, the tube is erect, as shown in Figure 8.3.


Capillary Dynamics 169

Figure 8.3. An erect capillary tube.

Figure 8.4. A horizontal capillary tube.

When ψ = 0◦ , expression (8.17) can be written as

2γ  
PA + ρgh + r cos θ r 2 + 4εr t
l2 = − (8.20)

In this case, the tube is horizontal, as shown in Figure 8.4.


If the two ends of the capillary tube are opened, PA = 0,
expressions (8.19) and (8.20) can be rewritten as
 2 
r + 4εr ρg · t
+l

 
− ρgh + 2γ r cos θ ρgl
= ln 1 − (8.21)
ρg ρgh + 2γ
r cos θ
170 Capillary Mechanics

and
2γ  2 
ρgh + r cos θ r + 4εr t
l2 = − (8.22)

For entirely wetting liquid, ε = 0, and expression (8.21) is changed
into
 
−8ηl 8η (h + Δh) l
t= 2 − ln 1 − (8.23)
r ρg r 2 ρg h + Δh
where Δh is 2γ/rρg.
In the case where capillary tube is entirely wetted, when ignoring
the influence of pressure (e.g., h = 0 and the tube is horizontal)
and only considering the capillary pressure, expression (8.16) can be
changed into
dl rγ
= cos θ (8.24)
dt η 4l
After integration, we have
 
γ cos θ
l2 = rt (8.25)
η 2
Expression (8.25) is the Washburn equation or the Lucas−
Washburn equation.
Expression (8.24) can be explained as follows. The rate of liquid
penetrating the horizontal capillary tube (or a capillary tube with
a small surface area) under capillary pressure is proportional to the
radius of the capillary tube, the cosine of the contact angle, and the
ratio between the surface tension of the liquid and its viscosity and
is inversely proportional to the length of the penetrated liquid. Since
γ cos θ/2η in the Washburn equation expresses the penetration abil-
ity of liquid, it is called the penetration coefficient, or penetration
efficiency. It can express the magnitude of the rate. The penetration
coefficient shows the length that a liquid of unit radius penetrates
the capillary tube in a unit of time. The term cos θ in the penetration
coefficient indicates that the penetration has a relationship with the
material of the capillary tube.
For porous materials, the penetration rate of a liquid into the
pores can be analyzed using the total volume of penetration during
Capillary Dynamics 171

the period t. In order to simplify the calculation, suppose that the


penetration rate of some liquid into a porous material equals the
sum of n columns of capillary tubes whose radius are r1 , r1 , . . . , rn
separately.
 
2 π 1/2 2γ 1/2 3
V = πr l = 1/2 t PE + r (8.26)
2η r

Therefore, the total volume of the liquid penetrating the porous


is
  
2 π 1/2  2γ 1/2 3
V =π r l = 1/2 t PE + r (8.27)
2η r

In this expression, the case of entirely wetting is considered. At the


same time, the influence of total outside pressure is also considered.
With the derivation of (8.27) with respect to time, the penetration
volume in unit time, which is also called the rate of penetration
volume of porous material, can be obtained.

8.3 Capillary Dynamics Equation

As shown in Figure 8.5, the liquid column undergoes three distinct


types of forces during its ascent, which are capillary drive pressure
Fcap , viscous resistance on the side of the capillary tube Fvisco , and
the gravity of the liquid Fgrav .
The capillary drive pressure can be directly obtained from the
Young−Laplace equation. According to the Young−Laplace equa-
tion, the drive pressure in a capillary tube can be written as
ΔP = 2γ/R∗ = 2γ cos θ/R. The total capillary drive pressure is
Fcap = ΔP · πR2 = 2γ cos θ/R · πR2 . Here, γ is the surface tension
of the liquid, θ is the contact angle, R is the radius of the capillary
tube, and R∗ is the curvature radius of the meniscus.
The viscous resistance can be obtained from the Hagen−Poiseuille
equation and the inner friction law for Newtonian viscous fluids.
According to the inner friction law of Newton, the viscous friction
shearing force on the capillary tube wall is τ = ηdv/dr|r=R , where η
is the viscosity of the fluid, v is the velocity of the fluid and r is the
radial coordinate. It can be found that the viscous friction shearing
172 Capillary Mechanics

Figure 8.5. Liquid rising in a capillary tube.

force is proportional to the viscosity and velocity gradient. The veloc-


ity gradient on the tube wall can be obtained from Hagen−Poiseuille
equation. The Hagen−Poiseuille equation of a fluid flowing in a cir-
cular tube is
 
1 d dv 1 ∂P
r = (8.28)
r dr dr η ∂z
where P is the pressure of the fluid and z is the coordinate in the
direction of the capillary tube length, with upward considered posi-
tive. When ∂P/∂z is independent of r, integrating the above expres-
sion, for the boundary conditions dv/dr = 0, r = 0, v = 0, and
r = R, leads to
 
1 2 2 ∂P 1  ∂P  2
v= (r − R ) = (r − R2 ) (8.29)
4η ∂z 4η  ∂z 
dv 2r ∂P
= · (8.30)
dr 4η ∂z
Since the flow speed is variable along the radial direction, the aver-
age flow velocity is usually used instead of the variable flow velocity.
Capillary Dynamics 173

The average flow velocity is defined as the flow per unit area, then
 R  R  
V̇ 1 1 1  ∂P   2 
v̄ = 2
= 2
2πrvdr = 2   R − r 2 2πrdr
πR πR 0 πR 0 4η ∂z
 
R2  ∂P 
= (8.31)
8η  ∂z 
The solution is
 
 ∂P  8ηv̄
 
 ∂z  = R2 (8.32)

Substituting it into the velocity grads equation leads to


 
dv 2r  ∂P  2r 8ηv̄
= = · (8.33)
dr 4η  ∂z  4η R2
At the point of r = R on the parietal

dv  4v̄
= (8.34)
dr r=R R
Substituting this expression into the equation for the inner friction
shearing force of a viscous fluid leads to
4v̄
τ =η· (8.35)
γ
The total viscous resistance on the parietal of the capillary tube
is
Fvisco = 2πRh · τ = 8πηhv̄ (8.36)
where h is the rising height of the liquid surface. The gravity of the
fluid is
Fgrav = ρπR2 h · g (8.37)
where ρ is the mass density of the fluid and g is the acceleration due
to gravity. The resultant force of these three kinds of force is
F = Fcap − Fvisco − Fgrav (8.38)
When liquid flows in the capillary tube, not only will the velocity
change (acceleration), but also the mass of the liquid will change.
174 Capillary Mechanics

Therefore, when considering the total inertial effect, these two


changes must be considered at the same time. When setting up the
fluid dynamics equation, the momentum law is used. According to
the case of a liquid flowing in a capillary tube, its momentum law
can be written as
d(mv̄)
=F (8.39)
dt
where m is the mass of the fluid, m = πR2 hρ; v̄ is the average speed
of flow, v̄ = dh/dt; F is the resultant force of capillary drive force,
viscous resistance, and the gravity. Therefore,

d(πR2 hρḣ) 2γ cos θ


= πR2 − 8πηhḣ − πR2 hρg (8.40)
dt R
d(hḣ) 2γ cos θ 8μh
ρ = − 2 ḣ − ρgh (8.41)
dt R R
Moving the left term (called the inertial term) of the equation to
(8.41) to the right side leads to

2γ cos θ 8μh d hḣ


− 2 ḣ − ρgh − ρ =0 (8.42)
R R dt
According to this equation, from the aspect of the dynamics bal-
ance, capillary pressure is the driving force, but the viscous force,
gravity, and inertial force are resistance forces.

8.4 Stages of Capillary Flow

With the action of capillary force, a liquid will continuously rise in a


capillary tube. However, in the rising process, the liquid is influenced
not only by the capillary force. In addition to it, inertial force, viscous
resistance, and gravity will also act on the liquid. The rising of the
liquid is the result of various kinds of forces acting together. However,
at different stages, the dominant action of these forces varies. Con-
sequently, the mechanism of liquid rising is not the same throughout
the process. In the initial stage of liquid rising within a capillary tube,
the volume of liquid is very small. As a result, the viscous resistance
Capillary Dynamics 175

and gravitational forces, which are proportional to the volume of liq-


uid, are also minimal and can be disregarded. However, the capillary
pressure, which is related to the volume of liquid, and the inertial
force, which is associated with the rate of change in the volume of
liquid, cannot be ignored and will dominate. Therefore, this stage is
referred to as the pure rising stage, marking the first phase of the
rising process. As the liquid continues to rise, the volume of liquid
in the capillary tube will gradually increase. Consequently, viscous
resistance and gravitational forces will become increasingly signif-
icant, ultimately dominating the liquid’s ascent over inertial forces
and capillary pressure. This phase is referred to as the viscous-inertial
stage, which represents the second stage of the rising process. Given
that the forces in this stage are relatively greater, the dynamics of the
equation will become more complex. As the liquid rises further, in the
last stage of the rising process, the capillary drive pressure becomes
more and more close to the viscous resistance and gravity, and the
driving force and resistance become more and more close to balancing
each other. Therefore, the speed of liquid rising will become slower,
and the inertial effect will become weaker. The liquid flow tends to
the steady Poiseuille flow. The dominant force will change from mul-
tiple forces to viscous and capillary forces. Therefore, this stage is
called the pure viscous stage, which is the third stage of the whole
process of liquid rising (the last stage). For various kinds of liquid
rising flows, previous researchers have put forward several dynamic
models. However, these models are suitable for specific backgrounds
and conditions; therefore, none of them can be adapted to the whole
rising process. Each model is only adapted to one stage. Since the
dominant forces mostly exists in the second stage, and the related
mathematical expression is the most complex one, it is divided into
three stages. The temporal delineation of the three stages can be
determined based on the corresponding discriminant criteria.

8.4.1 Phases of capillary flow


8.4.1.1 Stage of inertial force action
At the instant of the capillary tube coming into contact with the
liquid, that is, the initial stage, the amount of liquid is very small,
the viscous resistance and gravity are also very small, which can
176 Capillary Mechanics

be ignored, and the dominating forces are inertial force and capil-
lary pressure in this stage. In this case, Quere put forward a model
ignoring the viscous resistance and gravity. Neglecting the viscous
resistance and gravity in expression (8.42) leads to

2γ cos θ d(hḣ)
= = ḣ2 + hḧ (8.43)
ρR dt

By solving this nonlinear differential equation, Quere gave the


rising height of capillary under a constant speed as

2γ cos θ
h=t (8.44)
ρR

This equation is adapted to the first stage of liquid rising, the


pure inertial stage.

8.4.1.2 Stage of inertial-viscous force


As the liquid continuously rises, the amount of liquid in the capillary
tube will gradually become greater. Therefore, the viscous resistance
and gravity will become more and more important, which cannot
be ignored. They will dominate the liquid rising with the inertial
force and capillary drive pressure. Considering that the gravity is
still relatively small at the beginning of this stage, which can also
be ignored, Bosanquet put forward a model ignoring the gravity. By
neglecting gravity, expression (8.42) leads to

d(hḣ)
+ ahḣ = b (8.45)
dt

where a = 8η/R2 ρ and b = 2γ cos θ/Rρ.


The solution is
 
2 2b 1 −at
h = t − (1 − e ) (8.46)
a a

This equation is adapted to the cases where the liquid surface is


not so high.
Capillary Dynamics 177

8.4.1.3 Stage of viscous force


With viscous resistance increasing, it increasingly offsets the capillary
drive pressure and the net driving force becomes smaller and smaller.
Therefore, the inertia of liquid flow becomes smaller and smaller. For
this case, Lucas and Washburn put forward a model ignoring the
inertial effect and gravity. Neglecting the influence of inertial effect
and gravity in expression (8.42) leads to
γ · R cos θ
h2 = t (8.47)

8.4.1.4 Stage of viscous force and gravity


However, when the height of the liquid surface is greater than a
specific value, the influence of gravity cannot be ignored. Fries and
Dreyer analyzed and demonstrated that the influence of gravity and
viscous forces must be considered when h > 0.1heq , where heq is the
height when the capillary pressure is the same as the static water
pressure. Washburn gave the invisible solution for the height of capil-
lary rising when considering the influence of viscous force and gravity,
which is
 
h α βh
t(h) = − − 2 ln 1 − (8.48)
β β α
Fries and Dreyer gave the visible solution, which is
α 2
h(t) = [1 + W (−e−1−β t/α )] (8.49)
β
where α = γR cos θ/4η and β = ρgR2 /8η. W (α) is the Lambert W
function, and it is the W (z) in the equation z = W (z)eW (z) .
When the capillary pressure is the same as the static water
pressure,
α 2γ cos θ
heq = = (8.50)
β ρgR
It is regrettable that the solution for the height of liquid rising —
whether in the form of an invisible or visible solution — is derived
by neglecting inertial forces due to mathematical complexities. To
date, no researcher has provided an analytical solution that simulta-
neously accounts for inertial forces, viscous forces, gravity, and cap-
illary pressure.
178 Capillary Mechanics

8.4.2 Division of the different stages


As mentioned above, the role of various kinds of forces in the total
effective drive pressure has primary and secondary points in different
stages of capillary flow. The process of capillary flow can be divided
into three major stages in terms of the role level of these forces:

1. stage of inertial force action,


2. stage of inertial force and viscous force action together, and
3. stage of viscous force action.

The second major stage can be divided into three stages, and
the whole process can be divided into five stages. It is the inertial
stage in the time range of 0–t1 (the first major stage), where t1 is
the time boundary point of the inertial stage and the viscous stage,
which can be determined using the rule that the height difference of
the liquid surface rising between the inertial phase and the viscous-
inertial phase should be smaller than 3%. t1 − t2,S is the second
stage and t2,S is the boundary point of the second stage and the third
stage, which can be determined using the rule that the flow velocities
calculated by the Quere equation and the Lucas−Washburn equation
separately should be equal. t2,S −t2,Q is the third stage and t2,Q is the
boundary point of the third phase and the fourth stage, which can
be determined using the rule that the height of liquid surface in the
Quere equation and the Lucas−Washburn equation should be equal.
t2,Q − t3 is the fourth stage and t3 is the boundary point of the fourth
stage and the viscous force stage, which can be determined using the
rule that the rising height of the liquid surface calculated separately
using the Bosanquet model and the Lucas−Washburn model should
be smaller than 3%. t3 − t→∞ is the stage of viscous force action.
For t1 , in terms of the rule mentioned above, it is

hQuere (t1 ) − hBosanquet (t1 )


0.03 = (8.51)
hQuere (t1 )

Tidying up, it leads to

0.1856 0.0232R2 ρ
t1 = = (8.52)
a η
Capillary Dynamics 179

From equation (8.46), we have


√ 
0.18 b R3 ργ cos θ
h1 = = 0.0318 (8.53)
a η2
For t2,S , when the flow speeds separately in the Quere equation
and the Lucas−Washburn are equal, the boundary point of time is
solved as
1 R2 ρ
t2,S = = (8.54)
2a 16η
For t2,Q , when the heights of liquid surface separately in the Quere
equation and the Lucas−Washburn equation are equal, as shown in
Figure 8.6,
the time point is solved as
2 R2 ρ
t2,Q = = (8.55)
a 4η
The simultaneous equations of (8.44) and (8.47) leads to
√ 
2 b R3 ργ cos θ
h2,Q = = 0.3536 (8.56)
a η2
For t3 , when the height difference of the liquid surface
rising calculated separately using the Bosanquet model and the

Figure 8.6. Time step for several equations.


180 Capillary Mechanics

Figure 8.7. Time point without dimension.

Lucas−Washburn model is smaller than 3%. We have


hLucas−Washburn (t3 ) − hBosanquet (t3 )
0.03 = (8.57)
hLucas−Washburn (t3 )
then
16.921 2.1151R2 ρ
t3 = = (8.58)
a η
Substituting it into equation (8.46) leads to
√ 
5.6429 b R3 ργ cos θ
h3 = = 0.9975 (8.59)
a η2

Ichikawa and Satoda gave the boundary point of time without


dimension, as shown in Figure 8.7:
8ηt
t∗ = at = (8.60)
ρR2

ah 16η 2 h2
h∗ = √ = (8.61)
2b ρR3 γ cos θ

where a = 8η/R2 ρ and b = 2γ cos θ/Rρ.


Chapter 9

Concept and Function of Liquid


Bridge

9.1 Concept of Liquid Bridge

When a small amount of liquid exists between the surfaces of two


closely spaced solids that can be wetted by the liquid, a connected
system is formed among the two solids and the liquid. This system
is called a liquid bridge. Liquid bridges have two main features:
transmission and connection. Transmission refers to the convection
of mass and heat transfer within the liquid. Connection refers to the
pulling force exerted by the liquid on the surfaces of the two solids.
Both the convection of mass and heat transfer within the liquid and
the pulling force on the solids are based on the surface tension acting
on the sides of the liquid.

9.2 Formation of Liquid Bridge

There are two approaches to forming a liquid bridge: one is natural,


and the other is artificial. Naturally formed liquid bridges are
common in micro- and nanoscale systems. Artificial liquid bridges
are usually created for research on Marangoni convection during mass
and heat transfer processes.
As a new mechanical and electrical system, Micro/Nano-Electro-
Mechanical Systems (M/NEMS) are widely used in both civilian
and military fields and have enormous potential. However, due to

181
182 Capillary Mechanics

the particularities of their size and processing, it is easy for liquid


bridges to form, either during processing or in the application pro-
cess. In general, liquid bridges are harmful to the performance of
M/NEMS.
The processing of MEMS structures includes two types: wetting
processing and drying processing. However, with the development of
microelectronics and MEMS technology, the wafer size has increased
significantly, from 100 to 300 mm. At the same time, the character-
istic scale of process nodes has become smaller and smaller, reaching
as low as 90 nm, sometimes even 65 nm or smaller. Additionally, the
process depth has increased, with a standard depth-to-width ratio
of 50:1 or 60:1, sometimes even 100:1. Consequently, the wafers pro-
cessed by semiconductors are becoming larger and larger, with nodes
becoming smaller and requiring greater precision. The structures are
becoming more complex, and fewer and fewer processing methods
can fulfill all these requirements. At present, the most useful method
is still wetting processing, which is expected to remain the leading
method in the next decade. The processing of wafers includes Front
End Of Line (FEOL) and Back End Of Line (BEOL). FEOL is mostly
concerned with the structure of the device, while BEOL focuses on
the connection of the device structure. The surfaces that need to be
cleaned in FEOL are Si or SiO2 , for which many methods can be
used. However, for BEOL, fewer methods can be used due to the
metal layer on the wafer.
Solid surfaces can be divided into hydrophilic and hydrophobic
types. The surface of SiO2 is hydrophilic, making it easy to wet with
a cleaning agent during the cleaning process. However, the surface
of Si is hydrophobic, which makes it difficult to clean because the
cleaning agent cannot wet the surface and remove particles. Dur-
ing processing, integrated circuits (ICs) produce dirt, which can be
classified into five types: general particles, metal shavings, organic
matter, natural oxidation matter, and micro-roughness. These types
of dirt must be cleaned off after wet processing. The cleaned wafer
must meet certain standards. To achieve these standards, a series of
chemicals are introduced during the cleanup process. The cleanup
methods include RCA cleanup, IMEC cleanup, and Ohmi cleanup.
Either before or after cleanup, some liquid (e.g., deionized water)
may remain on the wafer surface and within the processed structure.
Concept and Function of Liquid Bridge 183

The liquid residue on the surface affects wetting and adhesion. The
residual liquid within the structure will form a liquid bridge. A mild
liquid bridge can affect the performance of MEMS, while a severe
one can cause system failure due to adhesion.
Besides forming liquid bridges during the machining process,
MEMS devices are also prone to forming liquid bridges during
packaging, storage, and operation. Packaging is a crucial step for
MEMS, which involves forming the tube body, connecting the
pins (interface between macro and micro scales), and sealing the
chip. Sealing can be either vacuum sealing or non-vacuum sealing.
Currently, MEMS devices are among the most specialized for packag-
ing. Since inert gases can alter the mechanical movement of devices
during high-speed vibration, certain types of devices must operate
in a vacuum to exhibit specific functions. This is not a new issue, as
many electronic and optoelectronic devices and systems also require
a vacuum environment. However, MEMS devices differ from tra-
ditional electronic and optoelectronic devices. Besides functioning
as conductive circuits, MEMS devices can also perform mechanical
movements. Due to this movement function, they require not only
reduced humidity and damping but also reduced wear and increased
lubrication. To a certain extent, these two requirements are contra-
dictory, posing a new challenge for packaging techniques. Practices
with MEMS devices have shown that absolute vacuum packaging
is not the ideal solution. In fact, maintaining an absolute vacuum
package over a long period is not only difficult but also unneces-
sary. Therefore, non-vacuum packaging is commonly used. Liquid
bridges cannot be avoided entirely. The structures of MEMS devices
are very small, and most of them contain grooves, which act as nat-
ural capillary tubes. When the air humidity is relatively high, mois-
ture in the grooves easily condenses into droplets, forming liquid
bridges.
In order to avoid residual liquid during the process and the
formation of liquid bridges, the final step of the wetting process must
involve removing humidity or drying. Otherwise, the processed wafer
will not meet the required standards. To a certain degree, the drying
process is the crucial final step that determines the overall perfor-
mance of the entire wafer. If it is not dried properly, the wafer’s prop-
erties will suffer. Therefore, the drying process is the most important
184 Capillary Mechanics

step in the wetting process. In order to prevent the formation of


droplets and liquid bridges after packaging, a relative vacuum pack-
age is necessary. However, since it is difficult to maintain a vacuum
package for the device over a long period, it is essential to keep the
storage environment dry and sealed. Sometimes, drying and sealing
need to be processed simultaneously.
A man-made method of building a liquid bridge typically involves
using two coaxial round plates that are very close to each other. Some
liquid or solution is injected between these two plates. Usually, a tem-
perature field is applied along the axial direction of the upper and
lower plates. If in a gravity field, the axis of the liquid bridge aligns
with the direction of gravity. The environment surrounding the liq-
uid bridge can be either liquid, gas, or vacuum. When one side of
the liquid bridge, which is in contact with a solid, is heated, the tem-
perature of one end of the liquid will rise. This is a heat transfer
process in which the temperature is inconsistent at both ends and
even at different points on the liquid surface. Since temperature has
a significant effect on surface tension, the latter varies in different
positions or fields. Therefore, there is a surface tension gradient in
space. According to the Marangoni effect, liquid tends to flow from a
low-surface-tension field to a high-surface-tension field, which results
in Marangoni convection. Since surface tension decreases as the tem-
perature rises, the liquid will flow from the high-temperature field to
the low-temperature field. In the liquid-bridge system, the liquid on
the surface flows from one end to the other, forming convection. The
man-made liquid bridge can be used for research into convection.
Therefore, this type of liquid bridge is beneficial.
However, the natural formation of liquid bridges in MEMS is
usually harmful. Depending on the performance requirements, when
two solid components undergo relative movement or vibration, such
as the relative vibration between a cantilever beam and its substrate,
the liquid present between them generates an additional adhesion
force. When this adhesion force is relatively small, it directly affects
the amplitude of the vibration. If the adhesion force is relatively large,
it can cause these two solid components to lock together and result
in adhesion. Consequently, the structure becomes immobilized, and
the system fails. Since many structures in micro/nano systems rely
Concept and Function of Liquid Bridge 185

on relative movement, the presence of liquid bridges directly impacts


the system’s performance.

9.3 Pull-Off Force Resulted from the Liquid-Bridge

9.3.1 Concept of pull-off force


The force that pulls two solids together in a liquid-bridge system,
exerted by the liquid, is called the pull-off force. In fact, it is also
the force that needs to be overcome when separating the surfaces of
two solids. The reason for the pull-off force is the surface tension on
the curved liquid surface, which stems from the adhesion between
the liquid and the solid, as well as the cohesion of the liquid. When
liquid comes into contact with a solid that can be wetted, there will be
adhesion between the liquid and the solid surface. Additionally, the
liquid has cohesion, which causes a pulling effect on the two surfaces
of the solids in a liquid-bridge system. The strength of the pull is
determined by both the adhesion force and the cohesion force, but
the maximum force (peak) is limited by the weaker of the two forces.
Experiments show that when the liquid is water and the solid has a
hydrophilic surface, the adhesion force is stronger than the cohesion
force. In this scenario, if the two solid surfaces are moved to pull the
liquid, the water is pulled off first, rather than the interface between
the water and the solid. When the liquid pulls the two solid surfaces,
the force is directed along the normal to the solid surface. For parallel
plates, the direction of the force at every point in the affected zone is
the same, and the direction of the resultant force is consistent with
the direction at every point. However, for non-planar solid surfaces
(such as spherical surfaces), the normal direction of different points
in the affected zone differs, and the direction of the resultant force
depends on the vector sum of the forces at those points. The pull-
off force is the force that needs to be overcome to separate the two
solids due to the adhesive force. For the convenience of analysis,
the line is defined as the axial line of the liquid bridge, which passes
through the liquid and is perpendicular to the two solid surfaces. The
adhesive force, which acts along the axial line of the liquid bridge and
is exerted by the liquid on the solid surface, contributes to the pull-off
effect.
186 Capillary Mechanics

9.3.2 Formation mechanism of pull-off force


In a liquid-bridge system, the adhesive effect exists between the liquid
and the solid, and the liquid exhibits a cohesive effect. Therefore, the
liquid pulls the two solid surfaces together, which is the essential rea-
son for the occurrence of the liquid bridge. The pull force exists at
any scale because adhesion and cohesion exist at any scale. However,
the pull-force effect is not always obvious at every scale. If the span
(the distance between the two solid surfaces) of the liquid bridge
is relatively large or the overall size of the liquid-bridge system is
relatively large, the effect of the pull-off force is not obvious. The
reason for this is that other forces (such as volume forces) will dom-
inate. Such an effect is similar to the capillary effect; therefore, the
theory of the capillary effect is often used to analyze the pull-off
force. In fact, the characterization mechanism of the liquid bridge
can be explained by the effect of the capillary force. Analyzing the
capillary effect, the two ends of the liquid connect with the solid sur-
face, exhibiting an adhesive effect. There is surface tension on the
side surface of the liquid, and its combined effect forms a meniscus
and an inside concave surface of the liquid. This surface causes a
pressure difference between the two sides of the liquid surface (one
side is the gas outside the liquid, and the other side is the liquid
itself). Due to the concave surface, the pressure outside the liquid
(gas) is higher than that inside the liquid, creating a negative pres-
sure within the liquid. This negative pressure results in a pumping
force that pulls and attracts the surrounding medium (the upper
and lower solids and the gas on the side surface). The component of
this pumping force that is vertical to the solid surface is the pull-off
force.

9.3.3 The analysis model of pull-off force


There are two approaches to analyzing and determining the pull-off
force model: one is the thermodynamic method, and the other is the
molecular theory method. The thermodynamic method encompasses
two models: one is the capillary theory model, which is based
on the capillary pressure difference equation derived from the
Young−Laplace equation; the other is the energy theory model,
which is grounded in the principle of conservation of work and
energy.
Concept and Function of Liquid Bridge 187

9.3.3.1 Capillary theory model for liquid bridge


1. Liquid bridge between two parallel plates
As shown in Figure 9.1, when a small amount of liquid exists in a
narrow gap between two parallel solid plates, a liquid-bridge system
is formed. Since adhesive forces exist between the liquid and the two
solid surfaces, and the liquid possesses cohesive force, the two parallel
solid plates are adhered together. Consequently, an interaction and
interconnection system is established. Due to the adhesive effect of
the solids on the two ends of the liquid’s side surface, combined with
the liquid’s surface tension, a meniscus is formed. This shape results
in a pressure difference between the interior and exterior of the liquid,
with the pressure of the gas outside the liquid being higher than
that inside, thereby creating a negative pressure within the liquid.
According to the capillary theory model, the pull-off force has two
components: one is the effect of the negative pressure inside the liquid
acting on the wetted area of the solid in the direction of the liquid-
bridge axes; the other is the action of the surface tension of the
liquid’s side surface along the liquid-bridge axes at the three-phase
lines.
Assume that At and Ab represent the wetted area of the upper
and lower solid surfaces. The distance between the two plates is d.
The curvature radius of the liquid surface is r. Then, according to
the Young−Laplace equation, the pressure difference between the
two sides of the liquid surface is
 
1 1
Δp = γ + (9.1)
r1 r2

Figure 9.1. Diagram of liquid-bridge.


188 Capillary Mechanics

where γ is the surface tension of the liquid, r1 is the principal curva-


ture radius of the side curve surface of the liquid on the plane which
is vertical to the plate, and r2 is its principal curvature radius on the
plane which is parallel to the plate. For the square plate, the wetted
area is also square and the boundary is a line. Therefore, r2 = 0.
While r1 has a relation with d, the distance between the two plates.
Assume that the contact angles of the liquid with the upper and
lower plates are, respectively, θt and θb . According to the geometry
relation, d = r1 cos θt + r1 cos θb . Substituting it into equation (9.1)
leads to
γ(cos θt + cos θb )
Δp = (9.2)
d
According to the Pascal principle, this negative pressure acts not
only on the side surface of the liquid but also on the wetted surface
of the upper and lower solid surfaces. The total forces act on the
wetted areas of the upper and lower solid surfaces are, respectively,
γAt (cos θt + cos θb )
ft1 = ΔpAt = (9.3)
d
γAb (cos θt + cos θb )
fb1 = ΔpAb = (9.4)
d
The other part of the pull-off force is the component of surface
tension of the liquid side surface acting on the three-phase lines. It
is vertical to the parallel-plate surface. If lt and lb represent the cir-
cumference of the wetted area of the upper and lower solid surfaces,
respectively, the total force acting on the boundary of the wetted
areas of the upper and lower plates are, respectively,
ft2 = γlt sin θt (9.5)
fb2 = γlb sin θb (9.6)
Adding these two parts, the total pull-off force of the upper and
lower plates are, respectively,
γAt (cos θt + cos θb )
ft = ft1 + ft2 = + γlt sin θt (9.7)
d
γAb (cos θt + cos θb )
fb = fb1 + fb2 = + γlb sin θb (9.8)
d
Concept and Function of Liquid Bridge 189

These are the force and the counterforce. Therefore, ft = fb . When


the material and size of the upper and lower plates are the same,
then θt = θb = θ, At = Ab = A, lt = lb = l, and
2γA cos θ
f= + γl sin θ (9.9)
d
For non-square plates (such as circular plates), the wetted area is
not a square and the boundary line is not a straight line. Therefore,
r2 = 0. In this case, the effect of r2 should be taken into account. The
non-square surface of the solid can be analyzed similarly as above.
However, in general, it is more complex.
2. Liquid bridge of spherical plates
In the liquid-bridge system, the upper surface of the solid can also be
spherical. Assume the radius of the spherical surface is R. The two
principal curvature radiuses of the liquid meniscus are r and r ∗ . The
contact angle of liquid with the upper surface is θt . The contact angle
of liquid with the lower surface is θb . The surface tension of liquid
is γ. The angle at the position where the liquid comes into contact
with the wetted three-phase lines of the spherical surface is φ. The
distance between the bottom end of the sphere and the bottom plane
is d. Then, according to the Young−Laplace equation, the pressure
difference between the inside and outside of the meniscus side surface
of liquid is
 
1 1
Δp = γ + (9.10)
r r∗
When the size of the sphere is relatively larger than the liquid
(when R  r, r ∗  r), the effect of principal curvature radius can
be ignored. The above equation can be written as
γ
Δp = (9.11)
r
According to the geometric relation, it is known that
R(1 − cos φ) + d = r[cos θb + cos(θt + φ)] (9.12)
Then,
1 cos θb + cos(θt + φ)
= (9.13)
r R(1 − cos φ) + d
190 Capillary Mechanics

Substituting it into the pressure difference equation leads to

γ[cos θb + cos(θt + φ)]


Δp = (9.14)
R(1 − cos φ) + d

By multiplying the negative pressure by the projected area of


the wetted zone of the spherical surface onto a plane vertical to the
bottom plane, the pull-force along the axes of the liquid bridge is
determined to be

πR2 sin2 φγ[cos θb + cos(θt + φ)]


f1 = ΔpπR2 sin2 φ = (9.15)
R(1 − cos φ) + d

This is one part of the pull-off force. The other is composed of


the component of the side surface tension on the three-phase lines,
whose direction is vertical to the bottom plane. The component of
the surface tension along the liquid-bridge axes is γ sin(θt + φ). The
circumference is 2πR sin φ. Then, this part of the pull-off force is

f2 = 2πRγ sin φ sin(θt + φ) (9.16)

The total pull-off force is

πR2 sin2 φγ[cos θb + cos(θt + φ)]


f = f1 + f2 =
R(1 − cos φ) + d
+ 2πRγ sin φ sin(θt + φ) (9.17)

When the sphere is relatively large (i.e., R  r and φ → 0) and


the sphere and the bottom plane are of the same material and are in
contact with each other, d = 0 and θt = θb = θ, then

f = 4πRγ sin θ (9.18)

9.3.3.2 Capillary condensation resulting in liquid bridge


1. Capillary condensation of the structure in a system of
two parallel plates
Capillary condensation occurs not only in the gap near the
spherical-plate contact point but also within other capillary
Concept and Function of Liquid Bridge 191

structures. When two parallel plates with hydrophilic surfaces are


close to each other in a humid environment, capillary condensation
may occur, forming a liquid bridge. The relationship between the
curvature radius of the capillary condensation liquid surface and the
relative humidity can be described by the Kelvin equation:

 −1
1 1 γVm
+ = rk = (9.19)
r1 r2 R̄T ln(p0 /pr )

where r1 and r2 are the two curvature radiuses of the liquid bridge,
rk is the Kelvin radius, Vm is the mol volume, p0 is the saturation air
pressure of the liquid; for water, at 20◦ C, γVm /R̄T ≈ 0. 54 nm. When
the liquid bridge is a column, then r1 = r, r2 = 0, and if pr /p0 =
0.9 (the relative humidity is 90%), then rk ≈ 5.13 nm. Because the
distance between the two parallel plates and the radius of the liquid-
bridge surface has the relation d0 = 2r1 cos θ ≈ 2rk cos θ, if θ = 0,
the distance causing the capillary condensation is d0 = 10.23 nm.

2. Capillary condensation of the spherical-plate structure


When the spherical surface contacts the plane, a liquid bridge is
naturally formed. If the humidity of the environment is relatively
high, capillary condensation may occur, forming a liquid bridge. As
the spherical surface and the plane come into contact with each other,
the gap between them gradually increases from zero. This gap forms
a natural capillary hole of variable scale. Based on the analysis above,
when the radius of the spherical surface is very large, the curvature
radius of the side surface of the liquid within the capillary, corre-
sponding to different gaps, can be determined as

R(1 − cos φ)
r= (9.20)
2 cos θ

It is also the curvature radius of the liquid surface in the capillary


condensation. With φ gradually increasing from zero, r gradually
increases from zero. Capillary holes are formed with various possible
radiuses between the spherical surface and the plane.
According to the Kelvin equation, with a relative humidity of
pr /p0 , the Kelvin radius (average radius) causing the capillary
192 Capillary Mechanics

condensation of vapor is
2γVm
rk = (9.21)
R̄T ln pp0r
Capillary condensation occurs in the zone of φ < φr . Of course,
when the actual capillary condensation occurs, φr may be smaller
than the theoretical value, which is mentioned above, due to super-
saturation (excessive cold).

9.3.3.3 Molecular theory model of liquid bridge


The analysis of liquid cohesive force based on molecular theory is
the foundation of the liquid-bridge model within the framework of
molecular theory. According to the model of molecular forces, the
cohesive work (or interaction potential per unit area) of the liquid
can be expressed as
A11
W (d) = (9.22)
12πd2
where A11 is the Hamaker constant and d is the space between the
molecules.
The cohesive force (attractive force on unit area) can be written as
∂W (d) 2A11
f (d) = =− (9.23)
∂d 12πd3
The surface tension of the liquid can be expressed as
W (d) A11
γ= = (9.24)
2 24πd2
Substituting this surface tension into the equation of cohesive
force, it can obtained that

f (d) = − (9.25)
d
The minus indicates that it is an attractive force.
The weakest position of a liquid bridge is its neck. If it is assumed
that the area of its neck, which is vertical to the liquid-bridge axes,
is Sneck , the stretching resistance of the liquid bridge (pull-off force)
can be written as
4γSneck
F = Sneck f (d) = − (9.26)
d
Concept and Function of Liquid Bridge 193

9.3.3.4 Energy theory model of pull-off force


According to the energy theory of pull-off force, the liquid-bridge
system possesses potential energy and adheres to the principle of
conservation of work and energy. If there is a very small relative dis-
placement, denoted as δz, between the two solid surfaces, the work
done by the pull-off force on this displacement is equal to the vari-
ation, denoted as δE, in the potential energy of the liquid-bridge
system. The differential expression for this relationship is given by
δE
f= (9.27)
δz
The key to determining the pull-off force is to determine the total
potential energy of the system, especially the potential energy which
will change.

9.3.4 The total free energy of the liquid-bridge


Surface tension always leads to area contraction. Therefore, as
the area increases, the potential energy also increases. Conversely,
pressure always leads to volume expansion. Therefore, as the volume
decreases, the potential energy increases. The total free energy con-
sists of two parts: the free surface energy and the free bulk energy.
The free surface energy arises from the solid–liquid and liquid–vapor
interfaces, while the free bulk energy stems from the liquid and vapor
volumes under corresponding pressures. In order to calculate the free
energy of the liquid-bridge system, one needs to determine both the
volume of the liquid and the surface areas of the interfaces. For a
liquid bridge, the total free energy can be expressed as
n
 m

E= γi Ai − p j Vj (9.28)
i=1 j=1

The first sum term on the right-hand side represents the total
free surface energy. The second sum term represents the total free
bulk energy. Here, A denotes the interfacial area, and γ denotes
the interfacial tension. V represents the phase volume, and p repre-
sents the phase pressure. The subscripts i and j are used to indicate
the interface and the phase, respectively. Naturally, there are usu-
ally three phases: the vapor phase, the liquid phase, and the solid
194 Capillary Mechanics

Figure 9.2. A general liquid bridge.

phase. The interface refers to the contact surface between any two
of these phases, such as the interface between the vapor phase and
the liquid phase, the interface between the vapor phase and the solid
phase, and the interface between the liquid phase and the solid phase.
A liquid-bridge system is constructed by two solid phases, one liq-
uid phase, and one vapor phase. Therefore, in such a system, there
are two vapor–solid interfaces, two liquid–solid interfaces, and one
vapor–liquid interface, as shown in Figure 9.2.
If the interface area between the top solid and the liquid is Alst
(top wetted area), the interface area between the top solid and the
vapor is Avst , the interface area between the substrate solid (bottom
solid) and the liquid is Alsb (bottom wetted area), the interface area
between the substrate solid and the vapor is Avsb , the interface area
between the vapor and the liquid (i.e., the profile of the liquid) is Alvp ,
the volume of the liquid is Vl , and the volume of the vapor is Vv , the
total free energy of the liquid-bridge system, which is described by
equation (9.28), can be written as

E = γlst Alst + γvst Avst + γlsb Alsb + γvsb Avsb


+ γlvp Alvp − pl Vl − pv Vv (9.29)

Since the interfacial tension between the solid and the vapor is
very small, the effect of the solid–vapor interface in equation (9.29)
can be neglected. Then,

E = γlst Alst + γlsb Alsb + γlvp Alvp − pl Vl − pv Vv (9.30)


Concept and Function of Liquid Bridge 195

When using the energy theory to derive pull-off force, it is impor-


tant to know the total free energy. At a constant temperature, the
differentiation of the total free energy can be derived by

δE = γlst δAlst + γlsb δAlsb + γlvp δAlvp − pl δV l − pv δVv (9.31)

Since the total volume of the liquid and vapor phases is constant,
δ(Vl + Vv ) = 0, we have

δVv = −δVl (9.32)

Substituting equation (9.32) into equation (9.31) leads to

δE = γlst δAlst + γlsb δAlsb + γlvp δAlvp − (pl − pv )δVl (9.33)

For rigid solids, there is no deformation at the interfaces between


the solid and the liquid. These interfaces retain the same shapes as
the solid surface. However, the interface between the liquid and the
vapor will maintain a meniscus shape with corresponding curvatures
due to the capillary effect.

9.4 The Solution on the Interface of Liquid and Vapor


of the Liquid Bridge

On the meniscus interface of the liquid and the vapor of the liquid
bridge, there is no external force, except for the pressure of vapor
and the pressure of liquid. According to the energy theory,
δE δAlst δAlsb δAlvp
flvp = = γlst + γlsb + γlvp
δr δr δr δr
δVl
− (pl − pv ) =0 (9.34)
δr
where r is the normal coordinate of the curve meniscus surface. Con-
sider δAlst = δAlsb = 0, then
δAlvp δVl
γlvp − (pl − pv ) =0 (9.35)
δr δr
If the two principal curvature radiuses of the surface are r1 and r2 ,
according to the geometric relationships of the spatial curve surface,
196 Capillary Mechanics

we have
 
1 1
δAlvp =− + δVl (9.36)
r1 r2

Substituting equation (9.36) into equation (9.35) leads to


   
1 1 δVl
γlvp + + (pl − pv ) =0 (9.37)
r1 r2 δr
That is,
 
1 1
Δp = pv − pl = γlvp + (9.38)
r1 r2
This equation is called the Young−Laplace equation.

9.5 The Pull-Off Force of the Top Solid


in the Liquid-Bridge System

9.5.1 The solution for constant wetted areas


To obtain the pull-off force of the top solid in the liquid-bridge sys-
tem, one can use a classical derivative of the energy with respect to
the distance between the top solid and the substrate (i.e., the ver-
tical coordinate z). The principle of work–energy conservation can
be explained as follows. An incremental displacement δz in the z
direction will result in an increase in the total free energy δE of the
liquid-bridge system. According to the principle of work–energy con-
servation, this increase in energy δE should arise from the work done
by the pull-off force fa over the displacement δz.
Then, δE = fa · δz, and
δE
fa = (9.39)
δz
Substituting equation (9.33) into equation (9.39) leads to
δE δAlst δAlsb δAlvp
fa = = γlst + γlsb + γlvp
δz δz δz δz
δVl
− (pl − pv ) (9.40)
δz
Concept and Function of Liquid Bridge 197

(a) (b)

Figure 9.3. Take liquid with a piece of thin plate from a cup. (a) Scheme and
(b) photograph.

For constant interfacial areas, δAlst = δAlsb = 0. Use a piece of


thin plate to take the liquid from a cup, as shown in Figure 9.3. The
increase of the side area is
δAlvp = lt · δz · sin θt (9.41)
where θt is the contact angle of the liquid and the top solid, lt is the
perimeter of the interfacial (wetted) area Alst . Whereas δVl = Alst ·δz,
and substituting them into equation (9.40) leads to
δE
fa = = γlvp · lt · sin θt − (pl − pv ) · Alst
δz
= γlvp · lt · sin θt + Δp · Alst (9.42)
Taking γlvp = γ (the surface tension) and Δp = 2γ/rm , where rm
is the mean radius of the meniscus surface and 2/rm = 1/r1 + 1/r2 ,
we have

fa = γ · lt · sin θt + · Alst (9.43)
rm

9.5.2 An approximate solution for liquid with a


constant volume
In a general liquid-bridge system, the liquid maintains a constant
volume. That is, δVl = 0. Equation (9.40) can be written as
δE δAlst δAlsb δAlvp
fa = = γlst + γlsb + γlvp (9.44)
δz δz δz δz
198 Capillary Mechanics

To obtain the solution of equation (9.44), some geometric


relationships are needed to be analyzed and discussed. The depth–
width ratio is defined as ζ = d/w. d is the thickness of the liquid film
or the minimum width of the gap between the two solid objects. w
is the maximum width of area Alst . If the liquid bridge has a small
depth–width ratio, ζ  1 and Alst ≈ Alsb , the volume of the liquid
can be approximately expressed as

Vl ≈ d · Alst ≈ d · Alsb (9.45)

Considering the volume has a constant value, we get

δVl ≈ d · δAlst + δd · Alst = d · δAlst + δz · Alst = 0 (9.46)

Therefore,

δAlst Alst
=− (9.47)
δz d
Similarly,

δAlsb Alsb
=− (9.48)
δz d
From the geometric relationship for the meniscus surface of the
liquid bridge, as shown in Figure 9.4, the relationship between the

Figure 9.4. A circular liquid bridge.


Concept and Function of Liquid Bridge 199

thickness of the liquid film d (the distance between the two solid
objects) and the curvature radius r1 can be written as

d = r1 (cos θt + cos θb ) (9.49)

where θb is the contact angle of the liquid and the substrate solid.
Substituting equations (9.41), (9.47), and (9.48) into equation
(9.44), since γlst = −γ cos θt , γlsb = −γ cos θb , and γlvp = γ,
leads to
δE Alst
fa = = γ(cos θt + cos θb ) + γ sin θt lt (9.50)
δz d
Substituting equation (9.49) into equation (9.50) leads to

Alst
fa = γ + γ sin θt lt (9.51)
r1

9.5.3 An accurate solution for large gap


If the depth–width ratio is not small, a more general model should be
constructed. If the top solid and the substrate solid are both planes,
as shown in Figure 9.5, the increment in the vertical distance can be
expressed as

δz = δr1 (cos θt + cos θb ) (9.52)

The increment of volume should be written as

δVl = −Alvp δx + Alst δz (9.53)

Figure 9.5. A rectangular liquid bridge.


200 Capillary Mechanics

Since the total volume of the liquid maintains a constant value,


δVl = 0. Therefore,

Alst
δx = δz (9.54)
Alvp

The increment in the wetted area of the top interface can be


written as

lt Alst
δAlst = −lt · δx = − δz (9.55)
Alvp

The increment in the wetted area of the substrate interface can


be written as

lb Alsb
δAlsb = −lb · δx = − δz (9.56)
Alvp

where lt and lb are perimeters of the top interface and the substrate
interface.
The change in side area can be written as

δAlvp = lt · sin θt · δz (9.57)

Substituting equations (9.55), (9.56), and (9.57), into equation


(9.44), since γlst = −γ cos θt , γlsb = −γ cos θb , and γlvp = γ,
leads to

δE Alst
fa = = γ(lt cos θt + lb cos θb ) + γ sin θt lt (9.58)
δz Alvp

For a polygonal interface, as shown in Figure 9.6, each side area


can be written as
 π
−θt  π
2 αi lti αi
Ai = 2 tan tan−1 + r1 cos −θ
−( π −θb ) 2 2 2 2 t
2

− r1 cos β) r1 dβ (9.59)
Concept and Function of Liquid Bridge 201

Figure 9.6. A polygonal liquid bridge.

where αi is the central angle of side i and lti is the length of side i.
Integrating equation (9.32) and summing all the sides leads to
n

Alvp = Ai
i=1
n
  
αilti −1 αi
= 2 tan tan + r1 sin θt r1 (π − θt − θb )
2 2 2
i=1

2
− r1 (cos θt + cos θb ) (9.60)

For a circular wetted interface, as shown in Figure 9.4, taking the


limits αi → 0 and n → ∞ in (9.60) leads to
n

Alvp = lim Ai
αi →0(n→∞)
i=1
 2π
= [(r2 sin θt + r1 sin θt ) r1 (π − θt − θb )
0

− r12 (cos θt + cos θb )]dα


= 2πr1 [(r2 + r1 ) sin θt (π − θt − θb ) − r1 (cos θt + cos θb )]
(9.61)
202 Capillary Mechanics

Figure 9.7. A narrow, long rectangular liquid bridge.

For a rectangular interface, as shown in Figure 9.7, the side area is

b a
n

Alvp = Ai = 4 + r1 sin θt r1 (π − θt − θb )
a 2
i=1
 
2 a b
− r1 (cos θt + cos θb ) + + r1 sin θt r1 (π − θt − θb )
b 2

2
− r1 (cos θt + cos θb ) (9.62)

where a and b are the length and width of the rectangle, respectively.
For a square interface,

n

Alvp = Ai
i=1
 
= (lt + 8r1 sin θt ) r1 (π − θt − θb ) − 8r12 (cos θt + cos θb )
(9.63)

Here, for a narrow, long rectangular solid, as shown in Figure 9.7,


the interface area will contract from the long side to form a square
shape and then contract from the short side to form a circular shape.
Substituting equation (9.61) into equation (9.58), leads to

(lt cos θt + lb cos θb )Alst


fa = γ
r1 (lt + 2πr1 sin θt )(π − θt − θb ) − 2πr1 (cos θt + cos θb )
+ γ sin θt lt (9.64)
Concept and Function of Liquid Bridge 203

If the material of the top solid is the same as that of the substrate
solid, θt = θb and lt = lb ,
2lt cos θt Alst
fa = γ + γ sin θt lt
r1 (lt + 2πr1 sin θt )(π − 2θt ) − 4πr1 cos θt
(9.65)

For a liquid bridge with a small depth–width ratio, lt .  2r1 cos θt .


Equation (9.65) can be written as

2 cos θt Alst
fa = γ + γ sin θt lt (9.66)
r1 (π − 2θt )

When comparing equation (9.51) with equation (9.66), it can be


observed that these two equations are not identical. Equation (9.66)
is a more precise equation, as it takes into account the curved sur-
face of the sides during the calculation of the liquid volume. Using
the straight line distance 2r1 cos θt instead of the curve arc distance
r1 (π − 2θt ), equation (9.66) will change to equation (9.51). For a
square interface, substituting equation (9.63) into equation (9.58)
and considering lt .  2r1 cos θt , the same equation as (9.66) can be
obtained.
It must be pointed out that, as the solid–liquid interface contracts,
with its border moving away from every boundary, the shape of the
interface tends to become a circle due to surface tension.

9.5.4 An accurate solution for the sphere-plane


liquid-bridge system with a large gap
When the top solid has a spherical shape and the substrate solid
is planar, as shown in Figure 9.8(a), the situation becomes rela-
tively complicated. The conservation condition of volume can also
be written as

δVl = −Alvp δx + Alst δz = 0 (9.67)

Then,
Alst
δx = δz (9.68)
Alvp
204 Capillary Mechanics

(a) (b)

Figure 9.8. A sphere–plane liquid bridge system with a large gap. (a) Different
gaps of liquid bridge and (b) geometric relationship between sphere and plane.

When the sphere moves upward with a slight displacement δz,


the meniscus interface between the liquid and the vapor, the cur-
vature center, and the curvature radius will also undergo changes.
However, the contact angles should remain constant. Assuming that
the curvature center experiences a slight displacement to the left by
δx0 and a slight upward displacement by δy0 and that the curvature
radius increases by δr, the wetted boundary on the spherical inter-
face will contract with a displacement to the left by δxt , which can
be expressed as
δxt = δx0 + δr sin(θt + φ) − r cos(θt + φ)δφ (9.69)
where φ is called the filling angle and δφ is defined as the positive in
clockwise direction.
The vertical upward displacement can be written as
δyt = δy0 + δr cos(θt + φ) + r sin(θt + φ)δφ (9.70)
For the geometric relationships, as shown in Figure 9.8, the dis-
placement of the meniscus curve surface (or the displacement in the
left-hand direction of the horizontal apex of the profile) can be writ-
ten as
δx = δx0 + δr (9.71)
Concept and Function of Liquid Bridge 205

The wetted area of the substrate plane will contract with a fol-
lowing decrement:

δρ = δx − (1 − sin θb )δr (9.72)

Since δxt = R cos φδφ and δy0 = δr cos θb , we can obtain the
following relationships:

δz = δyt + R sin φδφ (9.73)

and

δyt = [cos θb + cos(θt + φ)]δr + r sin(θt + φ)δφ (9.74)

Substituting equation (9.74) into equation (9.73) leads to

δz = [cos θb + cos(θt + φ)]δr + [R sin φ + r sin(θt + φ)]δφ (9.75)

Substituting equation (9.69) into equation (9.71) leads to

δx = [1 − sin(θt + φ)]δr + [R cos φ + r cos(θt + φ)]δφ (9.76)

From equations (9.68), (9.72), (9.74), and (9.75), the following


solutions can be obtained:
δr C1 − C3 ξ
= (9.77)
δz C1 C2 − C3 C4
δφ C2 ξ − C4
= (9.78)
δz C1 C2 − C3 C4
δρ C5 (C1 − C3 ξ)
=ξ− (9.79)
δz C1 C2 − C3 C4

where C1 = R cos φ + r cos(θt + φ), C2 = cos θb + cos(θt + φ), C3 =


R sin φ + r sin(θt + φ), C4 = 1 − sin(θt + φ), C5 = 1 − sin θb , and
ξ = AAlvp
lst
.
The wetted area on the sphere can be written as

Alst = 2πR2 (1 − cos φ) (9.80)


206 Capillary Mechanics

The wetted area on the substrate plane can be written as

Alsb = πρ2 (9.81)

The increment of meniscus area can be described as

δAlvp = 2πR sin φδs (9.82)

where δs is the increment of the meridian of the profile.

δs = rδφ + δr(π − θt − θb − φ) (9.83)

With the relationships given in equations (9.77–9.79), deriv-


ing equations (9.80–9.82) with respect to the vertical coordinate z
leads to
δAlst 2πR2 sin φ(C2 ξ − C4 )
=− (9.84)
δz C1 C2 − C3 C4
 
δAlsb C5 (C1 − C3 ξ)
= −2πρ ξ − (9.85)
δz C1 C2 − C3 C4
δAlvp r(C2 ξ − C4 ) + (C1 − C3 ξ)(π − θt − θb − φ)
= 2πR sin φ
δz C1 C2 − C3 C4
(9.86)

Substituting equations (9.84–9.86) into equation (9.44), since


γlst = −γ cos θt , γlsb = −γ cos θb , and γlvp = γ, leads to

2πR2 sin φ(C2 ξ − C4 )


fa = γ cos θt
C1 C2 − C3 C4
 
C5 (C1 − C3 ξ)
+ 2πρ ξ − γ cos θb
C1 C2 − C3 C4
r(C2 ξ − C4 ) + (C1 − C3 ξ)(π − θt − θb − φ)
+ 2πR sin φ γ
C1 C2 − C3 C4
(9.87)

When φ tends to zero (i.e., φ → 0), the liquid-bridge force


fa also approaches zero and does not maintain a constant value.
This is consistent with practical physical phenomena. However, if an
Concept and Function of Liquid Bridge 207

approximate cylindrical side surface is used instead of an accurate


meniscus profile to calculate the side area or the volume of liquid, a
conflict will arise. This issue can be addressed as follows.
When the thickness of the liquid film is very small, the curvature
radius will be much smaller than the radius of the sphere (r  R). In
this scenario, instead of using the actual meniscus surface, an approx-
imate cylindrical surface can be employed. The side area described
by equation (9.61) (where r1  r2 ) can be approximated as
Alvp = 2πR sin φ(D + d) (9.88)
where D is the distance between the sphere and the plane and d is
the height of the spherical cap.
Neglecting the profile surface tension (the third term in equation
(9.87)), considering θb = θt = θ and ρ = R sin φ, and taking D = 0
and the limit of φ → 0, then substituting equations (9.88) and (9.80)
into equation (9.87) leads to
fa = 4πRγ cos θ (9.89)
This is the conventional equation. Since equation (9.89) is inde-
pendent of the filling angle φ, this expression shows that the liquid-
bridge force approaches a constant value when φ tends to zero. Of
course, this is in conflict with the fact that there is no pull-off force
between two completely dry surfaces.
The wetted area on the plane can be approximately written as
Alsb = πR2 sin2 φ (9.90)
Equation (9.77) can be approximately written as
δx = R cos φδφ (9.91)
The conservation condition for the volume approximately results
in
Alsb
δx = δz (9.92)
Alvp
Substituting equations (9.88), (9.90), and (9.91) into equation
(9.92) leads to
δφ tan φ tan φ
= = (9.93)
δz 2(D + d) 2R(1 − cos φ)(1 + D/d)
208 Capillary Mechanics

Taking the derivatives of equations (9.80), (9.82), and (9.90) with


respect to the vertical coordinate z and substituting equation (9.93)
into them leads to

δAlst 2πR2 sin φ tan φ


=− (9.94)
δz 2R(1 − cos φ)(1 + D/d)
δAlsb tan φ
= −πR2 2 sin φ cos φ (9.95)
δz 2R(1 − cos φ)(1 + D/d)
δAlvp
= 2πR sin φ (9.96)
δz
Substituting equations (9.94–9.96) into equation (9.44) and using
γlst = −γ cos θ, γlsb = −γ cos θ, and γlvp = γ leads to

(1 + cos φ)2
fa = πRγ cos θ + 2πRγ sin φ (9.97)
cos φ(1 + D/d)

If the surface tension is negligible, equation (9.97) can be written


as

(1 + cos φ)2
fa = πRγ cos θ (9.98)
cos φ(1 + D/d)

Although equation (9.98) is not independent of the filling angle


φ, the same form as equation (9.89) can be obtained. This is because
an approximate cylindrical surface is also used for the calculation of
the liquid volume. As D = 0 (Figure 9.8(b)) and φ → 0, equation
(9.98) can be changed into

fa = 4πRγ cos θ (9.99)

This is the same as the conventional equation (9.89).


It should be pointed out that equation (9.99) is only applicable for
large spheres and thin liquid films. Although equation (9.98) partially
modifies equation (9.99), it is only applicable for thin liquid films with
a small depth–width ratio (ζ  1).
The model built in this section is applicable not only to large
spheres and thin liquid films but also to small spheres and large
separating distances.
Concept and Function of Liquid Bridge 209

Figure 9.9. Diagram of force before wetting the edge.

9.5.5 The edge-effect of liquid-bridge


In general, a liquid-bridge system consists of a top solid, a bottom
solid, and the liquid situated between them. The liquid not only wets
the major surface of the top solid but also the profile in the vicinity
of the solid’s edge, as illustrated in Figure 9.9. In this case, the pull-
off force will be caused not only by the negative pressure FP but also
by the profile surface tension of liquid, Fγ :
F = FP + Fγ sin θ (9.100)
where FP = ΔpA is the force caused by the pressure Δp of the menis-
cus surface of the liquid. A represents the wetted area of the top solid.
Δp is determined using the Laplace equation. γ is the surface tension
of the liquid. r1 and r2 are the principal curvature radiuses of the
liquid side surface on the plane, which are vertical and parallel to
the liquid bridge. If r is the average curvature radius of the menis-
cus surface, then 2/r = 1/r1 + 1/r2 . Substituting it into the Laplace
equation leads to

Δp = (9.101)
r
Substituting (9.101) into FP = ΔpA leads to

FP = A (9.102)
r
210 Capillary Mechanics

The force caused by surface tension of the liquid can be written


as Fγ = lγ. Then, expression (9.100) can be written as

F = A + lγ sin θ (9.103)
r
where θ is the contact angle of the liquid with respect to the top
solid and l is the perimeter of the wetted area.
Actually, in most liquid-bridge systems, the liquid not only wets
the lower surface of the top solid but also its edge. Since the top solid
in a liquid bridge has some thickness, both the bottom surface and
the side surfaces of the top solid are wetted, as shown in Figure 9.10.
This will cause a force on the top solid within the liquid bridge.
Therefore, the edge effect of the liquid bridge needs to be considered
when researching the pull-off force of the top solid.
For analytical convenience, suppose the top solid has a certain
thickness (such that its profile cannot be ignored), which shares the
same characteristics as the lower surface of the top solid, including
the same material, smoothness, contact angle, and surface tension.
If only the profile of the top solid is wetted and the other surfaces do
not come into contact with the liquid (as illustrated in Figure 9.11),
then the component of the surface tension in the direction parallel

Figure 9.10. Actual wetting of liquid bridge.


Concept and Function of Liquid Bridge 211

to the profile of the top solid is


F1 = FA cos θ = lγ cos θ (9.104)
In an actual liquid-bridge system, if the liquid wets the top solid
up to its edge, it will wet both the lower surface and the profile of
the top solid. Therefore, the actual situation is not as depicted in
Figure 9.9 but as in Figure 9.10. In this case, the pull-off force acting
on the top solid in the liquid bridge includes two components: one is
caused by the wetting of the lower surface of the top solid (as shown
in Figure 9.9); the other is caused by the wetting of the profile of the
top solid (as shown in Figure 9.11). Therefore, the total pull-off force
acting on the solid is the sum of the forces depicted in Figures 9.9
and 9.10, as illustrated in Figure 9.12:
F = FP + Fγ sin θ + F1 (9.105)
or

F = A + lγ sin θ + lγ cos θ (9.106)
r
Expression (9.106) is the formula used for calculating the pull-off
force when considering the edge effect. Some experiments have been

Figure 9.11. Force diagram of wetting the profile of top solid.

Figure 9.12. Diagram for the calculating model of the pull-off force considering
the edge effect.
212 Capillary Mechanics

Table 9.1. Several results of calculated pull-off force and the measured pull-off
force (unit of force: mN; unit of scale of silicon film: mm).

Scale of
No. silicon film FP Fγ sin θ F2 F1 F Measurement

1 15.71 × 15.71 5.044 4.218 9.262 1.755 11.017 11.84


2 30 × 10 6.624 5.373 11.997 2.235 14.232 14.67
3 20 × 20 7.805 5.373 13.178 2.235 15.413 16.45
4 10 × 20 4.195 4.03 8.225 1.677 9.902 9.13
5 19.64 × 19.64 7.894 5.277 13.171 2.195 15.366 17.62
6 11.4 × 20 4.666 4.218 8.884 1.755 10.639 10.16
7 31.2 × 10 6.545 5.534 12.079 2.302 14.381 14.4
8 17.73 × 17.73 6.443 4.763 11.196 1.982 13.178 13.86
9 15.71 × 20 6.591 4.797 11.388 1.996 13.384 13.76

Remark: F2 = FP + Fγ sin θ is the pull-off force without considering the edge


effect.
F = FP + Fγ sin θ + F1 is the pull-off force considering the edge effect.
F1 = FA cos θ = lγ cos θ the force caused by the edge effect.

done to measure the pull-off force and prove the validity of expression
(9.106).
In the measurement, the contact angle of silicon and water is
67.412◦ at 18◦ C, and the surface tension of water is 72.746 mN/m
at 17◦ C. A series of data points was obtained by measuring water
wetting silicon films of various scales. The calculated data were com-
pared with the measured data in Table 9.1. As shown in Table 9.1,
the pull-off force calculated using expression (9.106) is closer to the
measured value. Therefore, the model that considers the edge effect
is valid and reasonable.
Chapter 10

MEMS and Capillary Action

Science and technology are continually advancing toward miniatur-


ization, with scales reaching the micrometer and nanometer lev-
els. In terms of machining accuracy and structural dimensions,
micrometer-level technology has become increasingly sophisticated.
Simultaneously, nanoscale processing technology and functional
devices are evolving, enabling researchers to explore atomic and
molecular scales within the realm of micro-mechanical processing
technology. Capillarity is observed at both macroscopic and micro-
scopic scales. However, capillary action is particularly pronounced
in Micro-Electro-Mechanical Systems (MEMS) due to the relatively
small volume of the devices, their relatively large surface area, and
the minimal distance between adjacent surfaces. With the rapid
development of MEMS technology, micro-electro-mechanical devices
have gained widespread application, leading to increased attention
on capillarity within MEMS.

10.1 Capillary Action in the Processing of MEMS

The processes of producing MEMS are diverse; there are more


than 110 different kinds of processes. However, only seven kinds of
processes are specially applied for some products or in some compa-
nies. All of the other kinds can be easily divided into standardized
processes. Releasing the suspended structure is a common process of
producing a MEMS device. In order to make a suspended structure,

213
214 Capillary Mechanics

some appropriate materials, a structural layer, and an etching pro-


cess of removing the materials around the structure are needed. If
the removed materials come from the substrate, it is called Bulk
MicroMachining (BMM). Otherwise, when the removed materials are
deposited, it is called Surface MicroMachining (SMM). For example,
the pressure sensors are mainly fabricated using BMM, the products
of RF MEMS are fabricated through the SMM of metals, and micro-
phones utilize the SMM technology based on dielectric materials.
If you want to use MEMS technology to process a micropump, the
multi-wafer stacked structure processed by BMM is the first choice.
The products machined by BMM have some advantages, such as high
accuracy and robust structure. However, the cost is relatively high,
and it is difficult to integrate them with CMOS. The development
trend of the process of producing MEMS is to integrate with CMOS
and develop from BMM to SMM.
During the process of producing and packaging MEMS, capillarity
will have a great influence on the properties and reliability of micro-
machines. The capillary action is important for micro-devices and
MEMS structures. In recent years, there have been more and more
studies on capillary mechanics in microscale.

10.1.1 Capillary adhesion in MEMS


Many factors influence the phenomenon of micro-machine adhesion,
mainly due to the role of surface tension, such as capillary gravi-
tation, van der Waals force, hydrogen bonds, and static electricity.
The strengths of these kinds of surface tension effects are influenced
by many factors, including the material of the structure, the sur-
face complexion of the material, relative movement, humidity, and
temperature. This section only discusses the influence of capillary
adhesion.
Etching is used in MEMS to chemically remove the layers, which
are called “sacrificial layers”, from the surface of a wafer during man-
ufacturing. Etching is a very important process module. Every wafer
undergoes many etching steps before it is completed. The etching
technique can be divided into wet etching and dry etching. The pro-
cess of wet etching uses liquid-phase (“wet”) etchants. Dry etching
is a plasma etching method, where the process involves a physical
action using plasma to attack the wafer surface. Once the sacrificial
MEMS and Capillary Action 215

Figure 10.1. Adhesion caused by water cleaning.

layer is removed, the activity structure of micro-devices can move


freely as designed. Upon release, the device is easy to attack and
influence. It will be destroyed by mechanical vibration or be stained
by pollution.
For most MEMS devices, etching is the most important procedure
because the sacrificial layer must be removed under accurate con-
trol. After release, the structure must be entirely cleaned to remove
etchants, dissolvable materials, solid remains, and other by-products.
Adhesion between the movement parts of micro-structures is a serious
problem. When the mechanical restitution force of micro-structures
is smaller than the surface adhesion force, adhesion occurs. Water-
based cleaners can promote adhesion. The liquid will form a liquid
meniscus on the hydrophilic surface when the device is taken away
from the water solution after wet etching. After evaporation and con-
traction, the micro-structure will lean onto the substrate. Finally, the
adhesion forms, as shown in Figure 10.1.
Capillary adhesion exists in traditional bulk micromachining, as
shown in Figure 10.2. The process includes deposition, dry etching,
anodic bonding, release, etc. Compared to the SMM process, this
technique can produce devices with a high depth–width ratio, which
is highly attractive for devices based on comb drivers and comb
capacitance sensors. In addition, the silicon deposited with dense
boron can be directly used as an electrode, which makes the process
easy. Firstly, the silicon film is deposited with dense boron (Fig-
ure 10.2(a)). Secondly, photolithography and dry etching are used to
create the bonding table (Figure 10.2(b)). Then, the whole structure
is made using the second photolithography (see Figure 10.2(c)).
The electrode of titanium, platinum, and gold is made on Pyrex
7740 glass through the usual process. Grooves must be etched into
the glass before sputtering so that it ensures that the electrode sur-
face lies 50 nm higher than the glass surface (Figure 10.2(d–f)). The
silicon film and the glass film are cleaned by ultrasonic waves to
reduce the pollution by solid particles and immersed in ammonia
to increase their hydrophilicity. Then, the two films are aligned and
216 Capillary Mechanics

(a) (d)

(b) (e)

(c) (f)

(g)

(h)

Figure 10.2. Process diagram of traditional bulk micromachining.

pasted. After higher-temperature and electrostatic processing, the


silicon and the glass are bonded together. Then, they are placed in
an Ethylenediamine, o-Phthalic acid, and Water (EPW) or potas-
sium hydroxide solution to erode the low-density deposits of silicon.
Finally, after soaking in methanol, they are directly placed in an oven
to release the active structure (Figure 10.2(h)). The capillary force
caused by the EPW or potassium hydroxide solution is large enough
to pull the movable plate toward the substrate.
Besides the traditional bulk silicon-dissolved wafer process, SMM
is the other important technique in MEMS processes. This process
is considered one of the key techniques in the production of MEMS
devices. Firstly, a thin film is deposited onto the substrate. Then, the
film is selectively etched to form the designed structure. At present,
some products manufactured by this MEMS technique have been
widely used, such as pressure sensors, acceleration sensors, optical
switches, and Digital Mirror Displays (DMDs). The main advantages
of SMM are as follows: it is compatible with traditional and mature
MEMS and Capillary Action 217

Figure 10.3. Liquid bridge between micro parallel plates.

Integrated Circuit (IC) fabrication techniques, and the devices man-


ufactured by surface processing are easily integrable with control
circuits.
For SMM, after the sacrificial layer is etched, the device needs
to be cleaned with deionized water to remove the etchant and any
etched materials. When the device is taken out of the deionized water,
a liquid bridge is formed between the two parallel plates, as shown in
Figure 10.3. The capillary force caused by the liquid bridge always
appears during the drying process after wet etching.
When the liquid come into contact with the adjacent surface and
slowly evaporates, the capillary force caused by the liquid cleanser
can pull the parts together. When the capillary force is large enough
to cause the adhesion, the liquid evaporation will cause a more serious
problem. The volume of the liquid is greatly reduced during evapo-
ration, which produces a large enough force to cause the collapse of
fragile suspension structures. The shrinking droplet pulls the MEMS
surface closer and closer until they touch each other and stick.
At the last step of micromachining, the micro-structure must be
rinsed and dehydrated. At the end of dehydration, the liquid disap-
pears from the chip surface. However, liquid still exists in the gap of
the micro-structure. Liquid bridges of different curvatures are formed
on the surface of the residual liquid. With the effect of surface ten-
sion, the liquid bridge force pulls the microstructure to the substrate,
which will destroy the structure or cause the moveable structure to
stick on the substrate forever.

10.1.2 Control of capillary adhesion


Capillary adhesion includes several important factors which influence
the performance of MEMS devices, such as yield rate, service life,
and reliability. These increase the scrap rate of MEMS devices and
218 Capillary Mechanics

lead to unnecessary losses. Therefore, it is important to effectively


control capillary adhesion in MEMS devices during their processing
and manufacturing.
Methods of dry etching, using non-water cleaning or drying with
supercritical carbon dioxide, can reduce or even eliminate the adhe-
sion during the releasing process. Supercritical drying is a common
method to reduce capillary force. MEMS devices are dried in liquid
dioxide carbon, which makes them reach the supercritical point by
adjusting the pressure and temperature. During the process of dry-
ing, the meniscus cannot be formed and the surfaces cannot be pulled
together. Of course, the various parts can also be chemically treated
to ensure that the surfaces of the devices do not stick even when
they come into contact with each other. The extraction by supercrit-
ical carbon dioxide can remove the solvent and release the complex
and fragile MEMS structure on the silicon. The structures which are
released and cleaned include micro-engines with a single gear, bridges
and cantilevers, pressure sensors, and comb drivers. The surface ten-
sion of the supercritical liquid can be ignored. The liquid can also
remove the solvent in the capillary gap of only a few nanometers.
In addition, the device can be dried by supercritical carbon dioxide
repeatedly without damaging its structure.
The traditional bulk silicon-dissolved wafer process can also be
improved to better control capillary adhesion. The new bulk silicon-
dissolved wafer process is shown in Figure 10.4. Firstly, the silicon
film is deposited with P ++ and a 2 μm bonding table is created using
photolithography and dry etching. At the same time, the bonding
frame is manufactured around each chip to surround it. The electrode
is formed on Pyrex 7400 glass through lithography and sputtering.
Then, the silicon and glass are aligned and electrostatically bonded
together. After bonding, the film is put into an EPW solution to
erode the low-density silicon. The liquid could not flow into the single
chip due to the sealed frames. Then, the final activating structure is
formed by double-side lithography and dry etching.
It is obvious that this kind of process can avoid residual liquid
formed in the traditional process so that it can reduce adhesion. In
addition, in this new process, the liquid bridge force does not occur
because the activating structure does not come into contact with
the liquid. Therefore, this technology increases the rate of qualified
products produced by the devices.
MEMS and Capillary Action 219

(a) (c)

(b) (d)

(e)

(f)

(g)

(h)

Figure 10.4. New BMM process.

The performance of MEMS devices is affected by their internal


environment. Additives can effectively improve the internal environ-
ment of MEMS devices. An additive is necessary for some MEMS
devices that are very sensitive to humidity and small pollution par-
ticles. Even extremely small amounts of liquid or nanoparticles can
cause the failure of the device. Adding an additive during or before
sealing MEMS devices can effectively control the adhesion. Adsor-
bents are the most important additives, which can optionally clear
small amounts of liquid and nanoparticles produced during the man-
ufacture of MEMS devices. Some adsorbents can also absorb some
other kinds of harmful materials in the package. Adsorbents can be
divided into three types: gas, liquid, and solid. Hydrogen and oxygen
molecules have been found in airtight packages of MEMS devices,
which is known to have potential hazards. The gas adsorbent can
effectively eliminate the hazard. Water is the main substance to
be absorbed by a liquid adsorbent, which is generally called a wet
adsorbent. Micro- or nano-sized solid adsorbents are more generally
used, which means that they can capture any small particles of solid
components. Adsorbents were first used in vacuum tubes to absorb
220 Capillary Mechanics

oxygen molecules which may lead to performance degradation of glow


wires. At present, adsorbents are mostly used for packaging satellite
communication modules. Of course, the same type of adsorbent can
also be used for packaging MEMS.

10.2 Capillary-Driven MEMS Fluid Self-Assembly

The self-assembly technique is a process using microscopic forces,


such as surface tension and electrostatic force, as driving forces
to make a large number of micro-devices automatically arrange
(location and orientation) from disorder to order according to some
setting rules and then couple and fix them. During this process, the
capillary-driven self-assembly is an important method. This concept
originated from molecular self-assembly in chemistry, but it is no
longer limited to the molecular scale and has been widely used in
MEMS, photonics, nano-science, near-field optics, and other fields.
Since it is difficult for traditional assembling techniques to pro-
cess complex structures and assemble high-volume components of
different materials, the techniques are unable to meet the needs of
efficiency and low cost. Therefore, the self-assembly technique has
become a hot spot in the research of device and structure assembling
technology.
At present, by using surface tension, scientists have achieved
self-assembly for multiple batches. For example, automatically
assembled LEDs have been realized on the appointed zone of
substrates and have passed the electric test. Some Chinese schol-
ars have also successfully achieved self-assembly by putting small
squares of silicon on different substrates, such as nitride silicon and
ordinary glass. Presently, they are researching self-assembling devices
with non-symmetric and complex shape structures and are trying to
manufacture some practicable structures using this technology. This
self-assembling technology has a series of advantages, such as high
accuracy and large assembly at a time, but the materials of devices
and substrates can be inconsistent. At present, this technology has
the most potential among MEMS integration technologies in terms
of practicality.
The basic principles and steps of capillary-driven MEMS self-
assembly are as follows:
MEMS and Capillary Action 221

substrate Adhesive Adhesive


layer layer
device

(a) (b)

Figure 10.5. Diagram of device, substrate, and adhesive layer.

Table 10.1. The values of interface energy on different interfaces.

Interface Surface energy (mJ·m−2 )

Water — Adhesive 46
Water — Self-assembling membrane 52
Adhesive — Self-assembling hydrophobic membrane 1

1. Produce an assembled monolayer hydrophobic membrane as the


adhesion layer with a special shape at the position of the substrate
to install the device. Other parts of the substrate should remain
hydrophilic, as shown in Figure 10.5(a).
2. Produce a self-assembly membrane with the corresponding shape
of the fixing surface of the small device, as shown in Figure 10.5(b).
3. The surface of the substrate is coated with a special hydrophobic
adhesive. When it is put into water, according to the minimum
energy principle, the free energy of the interface always tends to
its minimum. The adhesive will only attach to the hydrophobic
membrane because of its hydrophobic characteristic and the rel-
atively small interface energy between the adhesive and the self-
assembled hydrophobic membrane, as shown in Table 10.1 and
Figure 10.6.
4. A large number of micro-devices with adhesive are put into water
and shaken continuously, as shown in Figure 10.7.
5. Once the hydrophobic parts of the device come into contact with
the adhesive, they will automatically align with high precision and
then complete the self-assembly through the solidifying process of
the adhesive, as shown in Figure 10.8.
The basic model is composed of three parts: the adhesive layer
of the micro-device, the adhesive layer of the substrate, and the
222 Capillary Mechanics

Figure 10.6. Diagram of substrate coated with hydrophobic adhesive.

Figure 10.7. Diagram of micro-device put into water.

1 2 3

Figure 10.8. Devices automatically assembled with substrate.


MEMS and Capillary Action 223

Adhesive layer of micro


device Z

Y
X

adhesive

Adhesive layer of
substrate

Figure 10.9. Basic structure of model.

adhesive. The system has six degrees of freedom: three transla-


tional degrees of freedom and three rotational degrees of freedom
(Figure 10.9).
According to the basic law of thermodynamics, any closed sys-
tem in nature is always consistent with the principle of minimum
energy. As a result, a liquid tends to assume the shape with the
minimum energy. In a mono-phase system, the surface area tends to
its minimum (most kinds of liquids take the shape of a ball with-
out gravity). However, if there is an interface between the liquid
and another material, it will form a shape with a large surface. The
driving force of this process is the molecular interaction. Thus, the
difference between the surface tensions of different materials forms
the capillary force.
Self-assembly does not impose stringent criteria regarding the pre-
cision of the alignment between the dimensions and morphology of
the assembled device and the adhesive layer of the substrate. A design
layout of the substrate adhesive layer may be suitable for devices
with different sizes of the adhesive layer, so that one layout of the
substrate may be applied for the installation of more devices. On
the contrary, the devices with the same adhesive layer may be fixed
on different substrates. However, when fixing devices with different
shapes and sizes, the selection of the shape and size of the adhesive
layer should be considered. Of course, if the difference in the sizes
between adhesive layer and substrate adhesive layer of the device
224 Capillary Mechanics

is too large to meet the designed assembling precision, the determi-


nation of the limitation of the size difference of the adhesive layers
during self-aligning relies on calculations under certain conditions,
such as experiment conditions, and precision.

10.3 Dynamic Characteristics of Parallel-Plate Structure


under Capillary Force

Most core structures of MEMS devices can be simplified into a model


which is composed of an immovable bottom board and a movable
parallel plate supported by a micro elastic beam. If the humidity
in the environment where the MEMS is located is comparatively
large, the liquid between the substrate and the movable plate will
form a liquid bridge, as shown in Figure 10.10. The movable plate is
generally driven by the sine of electrostatic force with known ampli-
tude and frequency, so that it will form a dynamic liquid bridge
between the movable plate and the substrate. It is necessary to ana-
lyze the dynamic characteristics of the parallel-plate structure under
the electrostatic force and liquid bridge force because the dynamic
characteristics of the parallel-plate structure have an influence on the
performance of the whole MEMS.
The geometric structure of the liquid bridge between parallel
plates is shown in Figure 10.11. In addition to the electrostatic force,
the linear viscoelastic force, and the capillary force between the par-
allel plates will have an impact on the movable plate.
The total capillary force is the summation of the difference
between the surface tension and the surface pressure of the meniscus.

Figure 10.10. Liquid bridge between micro parallel plates.


MEMS and Capillary Action 225

r2

rn

θ r1

Rc

Figure 10.11. Geometric structure between parallel plates.

Its analytical expression is as follows:


 
1 1
Fcap = 2πrn γ − πrn2 γ − (10.1)
r2 r1
where γ is the surface tension, rn is the radius of the contour formed
when the liquid bridge comes into contact with the oscillating plate,
r1 is the principal curvature radius of the liquid meniscus surface,
and r2 is the principal curvature radius of the cross-section of the
liquid bridge and parallel plate.
The linear viscoelastic force of a liquid is a function of its dynamic
viscosity of the liquid, which is expressed as follows:
Fvis = c (η ∗ ) ż (10.2)
When all the electrostatic driving force, linear viscoelastic force
of the liquid, gravity, elastic force of the spring, and capillary force
have an impact on the movable plate, its equation can be expressed
as follows:
 
d2 z ∗ dz 2 1 1
m 2 + c(η ) + kz + 2πrn γ − πrn γ − = F0 sin ωt
dt dt r2 r1
(10.3)
where m is the mass of the movable plate, c(η ∗ ) is the damping
coefficient of the liquid linear viscoelastic force, k is the stiffness
226 Capillary Mechanics

48

36
F / m N/mm2

24

12

0.9 1.8 2.7 3.6 4.5 5.4


l / mm

Figure 10.12. Measurement curve of liquid surface tension.

coefficient of the elastic supporting beam, z is the displacement of


the movable plate, F0 is the amplitude of the driving force, and ω is
the frequency of the driving force.
In the experiment of using the Du Noüy ring method or a silicon
film to measure the surface tension of water, the relation between
pull and displacement is shown in Figure 10.12. The surface tension
refers to the peak force, so that it has the meaning of strength.
When the displacement is small enough and the pull does not
reach the peak (the ring or silicon film at the bottom), the value
of pull will linearly decrease with the decreasing displacement. It is
proved that water has an elastic characteristic under the condition
of small deformation (not reaching the strength value of the surface
tension).
In the water-stretching experiment of various areas of silicon films,
the pull is proportional to the area of the silicon film, as shown in
Figure 10.13. The experiment also indicates that for different areas,
the curves of pull–displacement on unit area are coincident.
MEMS and Capillary Action 227

Figure 10.13. Relation curve between liquid surface tension and silicon film
area.

It indicates that the capillary force is a linear function of the


oscillation amplitude z, as given by
Fcap = kcap z (10.4)
Since the curve in Figure 10.12 is a value of the pull force on
unit area, kcap is the product of the silicon area and the slope
of an approximate straight line in the elastic section, as shown in
Figure 10.12.
Substituting equation (10.4) into equation (10.3) leads to
d2 z dz
m 2
+ c(η ∗ ) + kz − kcap z = F0 sin ωt (10.5)
dt dt
The solution to the equation above is
z = u sin(ωt + ϕ) (10.6)
The amplitude is
F0 1
u= · (10.7)
k − kcap ω 2 2
(1 − ( ωn ) ) + (2 · ξ · ω 2
ωn )
228 Capillary Mechanics

The phase is
 ω

2·ξ· ωn
ϕ = arctan (10.8)
1− ( ωωn )2

k−kcap c(η∗ )
where ωn = m , ξ= √ .
2 (k−kcap )m
If the displacement is comparatively large and the pull force of
the water reaches the strength of surface tension, the elastic force
will disappear and the force between the silicon piece and the water
will be the capillary force, as shown in equation (10.1). Since the gap
between the two plates is very small, r 2 is much larger than r 1 . In
this case, πrn2 γ/r2 will tend to zero. πrn2 γ/r2 → 0. Therefore, 2πrn γ
can be ignored. According to Figure 10.11, it can be found that
h
r1 cos θ = (10.9)
2
Then, equation (10.1) can be changed into
2Aγ cos θ
Fcap = (10.10)
h
The area is A = πrn2 for a circular plate.
So, equation (10.3) can be changed into

d2 z dz k2
m + c(η ∗ ) + kz = F0 sin ωt − (10.11)
dt2 dt z + h0
where k2 = 2Aγ cos θ and h0 is the initial space between the two
plates. z = h − h0 . In this case, the system is a nonlinear system of
forced vibrations.
Before the MEMS device reaches the value of surface tension, the
stiffness coefficient of the system will be enlarged, which is called
the soft spring effect, due to the elastic action of the liquid. The
amplitude–frequency and phase–frequency characteristics of the sys-
tem will be changed. If kcap is larger than k, the elastic restoring
force is smaller than the pull force of the liquid. Thus, the system
will adhere. Therefore, when designing the support beam in MEMS
devices, the elastic restoring force should be larger than the pull force
of the liquid.
MEMS and Capillary Action 229

Figure 10.14. Liquid bridge between MEMS structures.

After the MEMS device reaches the value of surface tension, if the
elastic restoring force is smaller than Fcap , the system will adhere. If
the elastic restoring force is larger than Fcap , the system will exhibit
nonlinear vibrations. The stability and sensitivity of the system will
decrease.
To analyze the dynamic characteristics of the MEMS device under
capillary force, the liquid bridge between the MEMS structures is
shown in Figure 10.14.
The parameters of the system are as follows:
The mass of the microstructure is 0.243389 × 10−6 kg. The area of
the microstructure is 0.55 × 10−6 m2 . The amplitude of the electro-
static driving force is 5.7657 × 10−6 N and its frequency is 4123 Hz.
The initial space between the microstructures is 10 μm. In addi-
tion, the measurement value of the surface tension of the water is
73 × 10−3 N/m, and the contact angle is 79◦ .
Substituting the parameters mentioned above into equation
(10.4), the elastic coefficient of the capillary force in the elastic sec-
tion Fcap is obtained as 5.5 N/m. Therefore, if the stiffness of the
microstructure is smaller than 5.5 N/m, the system will adhere. If
the stiffness is larger than 5.5 N/m, the amplitude–frequency and
phase–frequency characteristics of the system will be changed due
to the soft spring effect. According to the analysis presented, the
capillary force will severely affect the reliability, stability, and sen-
sitivity of the micromachine structure. The capillary force must be
eliminated during the process of manufacture, sealing, and storing of
micromechanical devices.
This page intentionally left blank
Chapter 11

Capillary Wave

11.1 Phenomenon of Capillary Wave

A capillary wave is a liquid-free surface wave caused by capillary


force. When the free surface is horizontal, the fluid is in a balance
state. When some part of the fluid is disturbed, the particles of the
liquid will leave the balance position and move along the direction
which is perpendicular to the level surface. The particles will return
to the balance position due to a restoring force, such as surface ten-
sion or gravity. Then, the particles move to the other side due to
inertia. Thus, the vibration of the liquid particles is formed, and free
surface waves are generated by the vibration transmission.
Generally, free surface waves are dominated by gravity and
capillary force. Therefore, they include two components: capillary
waves and gravity waves. However, on different scales, the effects
of gravity and capillary force are different. Therefore, the types of
surface waves are different due to the different wavelengths. If the
wavelength is much larger than the wave height, the curvature of
the free surface is very small, and the component of surface ten-
sion, which is perpendicular to the surface, is so small that it can
be ignored. Only the gravity is considered. This kind of free surface
wave is called a gravity wave. On the contrary, if the wavelength is
extremely short, only surface tension is considered. Such a wave is
called a capillary wave or ripple, whose wavelength typically ranges
from one to two centimeters.

231
232 Capillary Mechanics

Figure 11.1. Ripple.

Since the performances and mechanisms of gravity and capillary


force are different, capillary waves and gravity waves have different
characteristics. The shorter the wavelength of capillary waves, the
higher the spread speed. However, the behavior of gravity waves is
contrary to that of capillary waves. Capillary waves are common in
nature. For example, ripples are a kind of capillary wave, as shown
in Figure 11.1.
From the perspective of perturbation, the most common
disturbance on the surface of water is wind. Wind generates waves,
which is the most common phenomenon. At present, from the per-
spective of interference force, there are two main explanations for
the generation of wind and waves. One is that due to the inconsis-
tent distribution of wind, part of the surface of water is affected by
pressure to generate waves. Once the waves have been produced, the
wind force has a direct effect, and the waves continue to increase.
The other explanation is based on the perspective of hydromechan-
ics. When two kinds of fluids, which are in contact with each other,
move relative to each other, the interface becomes unstable. It must
form a wave surface to maintain the balance. When there are some
little concave and convex areas on the surface of water, the airflow
rates at the points of concave and convex surfaces are different. At
the convex surface, the wind speed increases, and the wind pressure
decreases, so that a rising trend is formed. At the concave surface, the
wind speed decreases, and the pressure increases, so that a downward
trend is formed. Therefore, the water surface became more uneven.
The surface tension and gravity on the water surface produce a trend
of restitution for the surface, which forms a storm.
Capillary Wave 233

If there is no wind, the sea remains calm. When the wind speed
reaches a certain value but is still weak, capillary waves begin to
appear on the water surface. The wavelength of capillary waves is
so short and their wave height is so small that these waves can be
seen as a wrinkle on the water surface. It only exists on a thin layer
of the water surface. As the wind continues to blow, the wavelength
and wave height of the capillary waves increase, and gravity waves
are formed.

11.2 Dispersion Relation of the Free Surface Wave

The free surface wave has its frequency and a wavelength. Wave-
length can be expressed by wave number. The relation between wave
frequency and wave number is called the dispersion relation.
There is an interface tension at the interface of two kinds of fluids.
When the upper-layer fluid is gas, it is called surface tension, whose
direction is tangential to the interface of the fluid. In general, when
the fluid is in a balance state, the free surface is horizontal. The free
surface separates the upper and lower layers of the fluid. ρ and ρ
stand for the densities of the two kinds of fluid. For certain fluids,
the densities are constant and ρ > ρ . There is a hypothesis that the
two kinds of fluid are not mucous, cannot be compressed, and move
without eddies. The two layers of the fluid have a velocity potential,
φ and φ . The lower velocity and the higher velocity can be expressed
by ∇φ and ∇φ , respectively.
There are three kinds of energy in a fluid system: gravity potential
energy Vg , caused by gravity; free surface energy Vst , caused by sur-
face tension; and kinetic energy T, caused by the flow. The potential
energy Vg of gravity is easy to understand. Vg is determined by the
density ρ, acceleration of gravity g, and height z from the surface,
where z = η(x, y, t). The potential energy of gravity can be obtained
by volume integration in the directions of x, y, z, as

  η 
1
Vg = dxdy dz(ρ − ρ )gz = (ρ − ρ )g

dxdyη 2 (11.1)
0 2

where z is assumed to be zero at the even interface.


234 Capillary Mechanics

With an increase in the surface area, the energy will increase


proportionally due to the effects of surface tension. Therefore, the
surface free energy (potential energy of the surface tension) can be
expressed as
⎡ ⎤
  2  2
∂η ∂η
Vst = γ dxdy ⎣ 1+ + − 1⎦
∂x ∂y
  2  2
1 ∂η ∂η
≈ γ dxdy + (11.2)
2 ∂x ∂y

The term after the first equal sign represents the Gaspard Monge
expression method. That after the second equal mark fits the case
where the value of the derivative is very small and the surface is not
so rough.
The kinetic energy of the fluid can be written as

1
T = ρv 2 dV (11.3)
2

Using the relation v 2 = ∇φ · ∇φ = |∇φ|2 , the total kinetic energy


of the upper and lower layers of the fluid can be written as
  η  +∞
1 2 
T = dxdy dzρ |∇φ| + dzρ |∇φ |2 (11.4)
2 −∞ η

Since this kind of fluid is not compressed and has no eddies,


φ(x, y, z, t) and φ (x, y, z, t) should satisfy the Laplace equation:

∇2 φ = 0 and ∇2 φ = 0 (11.5)
 
According
 to the first Green Formula, U (ψ∇2 φ)dV = ∂U (ψ(∇φ·
n))dS − U (∇φ · ∇ψ)dV , where ∇2 = Δ is the Laplacian, ∂U is the
boundary of the zone U, and n is the normal vector of dS. If ψ = φ,
then
  
2
(φ∇ φ)dV = (φ(∇φ · n))dS − (∇φ · ∇φ)dV (11.6)
U ∂U U
Capillary Wave 235

For the problem discussed here, since ∇2 φ = 0, the left-hand side


of equation (11.6) is equal to 0. Then,
  
(φ(∇φ · n))dS = (∇φ · ∇φ)dV = |∇φ|2 dV (11.7)
∂U U U

Generally,

∂φ ∂φ ∂φ 
∇φ = i+ j+ k (11.8)
∂x ∂y ∂z

Here, we only consider the locomotion along the direction of z.


Therefore, ∇φ = ∂φ/∂zk.
In addition, because the wave height is much smaller than the fluid
height, the integration of equation (11.4) approximates to η = 0. The
boundary of ∂U in equation (11.7) is shown in Figure 11.2.
For the lower layer of the fluid, when z → −∞, ∇φ → 0 and
∇φ · n = ∂φ/∂zk · n = 0. Therefore, in fact, only when z = 0, the
integration along the boundary is not zero. Similarly, for the upper
layer of the fluid, when z → +∞, ∇φ → 0 and ∇φ · n = 0. There-
fore, again, only when z = 0, the integration along the boundary
is not zero. Then, equation (11.4) can be transformed into equation

Figure 11.2. Boundaries of the upper and lower layers of the fluid.
236 Capillary Mechanics

(11.9) using the relation in (11.7):


  
1 ∂ϕ   ∂ϕ
T ≈ dxdy ρϕ −ρϕ (11.9)
2 ∂z ∂z

In order to find the dispersion relation, consider the wave


spreading along the direction of x. The equation of the wave is

η = a cos(kx − ωt) = a cos θ (11.10)

where a is the amplitude and k is the wave number. The phase is


θ = kx − ωt. The velocity of the corresponding particle is ∂η/∂t.
Since the particle velocity expressed by the velocity potential is the
same as the particle velocity on the boundary of z = 0,
 
∂ϕ  ∂η ∂ϕ  ∂η
= and = (11.11)
∂z z=0 ∂t ∂z z=0 ∂t

When considering primarily the locomotion of the particle in the


direction perpendicular to the liquid plane, the locomotion along the
vertical direction and that along the horizontal direction are rela-
tively independent. Therefore, assume that

φ = φz · φxy (11.12)

According to equation (11.5), ∇2 φ = 0. We can get

∂2φ ∂2φ ∂2φ


∇2 φ = + 2 + 2 =0 (11.13)
∂x2 ∂y ∂z

Substituting equation (11.12) into partial differential equation


(11.13) leads to

∂ 2 φz φxy ∂ 2 φz φxy ∂ 2 φz φxy


∇2 φ = 2
+ 2
+ =0 (11.14)
∂x ∂y ∂z 2

That is,

∂ 2 φxy ∂ 2 φxy ∂ 2 φz
∇2 φ = φz + φz + φxy = 0 (11.15)
∂x2 ∂y 2 ∂z 2
Capillary Wave 237

Dividing both sides of equation (11.15) by φz φxy leads to


 2  
∂ φxy ∂ 2 φxy ∂ 2 φz
+ φxy = − 2 φz (11.16)
∂x2 ∂y 2 ∂z

When z = 0, ∂ϕ/∂z|z=0 = ∂φz /∂z|z=0 · φxy = ∂η/∂t = aω sin θ.


Therefore, φxy = c1 aω sin θ, and we can get
 2 
∂ φxy ∂ 2 φxy
+ φxy = −k2 (11.17)
∂x2 ∂y 2

At the same time, using the relation in equation (11.16), it is


known that

∂ 2 φz
− 2 φz = −k2 (11.18)
∂z

Therefore,
∂ 2 φz
− k2 φz = 0 (11.19)
∂z 2
Finally, the general expression of φz is
φz = Ae+|k|z + Be−|k|z (11.20)
If z → −∞, φ is zero. Therefore, B = 0, and
φz = Ae+|k|z (11.21)
When z = 0,
∂ϕ ∂φz
= · φxy = A |k| e+|k|z · c1 aω sin θ
∂z ∂z
= A |k| c1 aω sin θ = aω sin θ (11.22)
Therefore,
1
A= (11.23)
c1 |k|
Then,
1 +|k|z
φz = + e (11.24)
|k| c1
238 Capillary Mechanics

1 +|k|z
φ(x, y, z, t) = φz · φxy = + e ωa sin θ (11.25)
|k|
Similarly,
1 −|k|z
φ (x, y, z, t) = − e ωa sin θ (11.26)
|k|
Substituting equations (11.25) and (11.26) into equation (11.9),
integrating the equations (11.1), (11.2), and (11.9) along the x and y
directions, whose integration ranges are one wavelength, λ = 2π/k,
in the x direction and one unit length in the y direction, leads to
1
Vg = (ρ − ρ )ga2 λ (11.27)
4
1
Vst = γk2 a2 λ (11.28)
4
1  ω
2
T = (ρ + ρ ) a2 λ (11.29)
4 |k|
According to the Lagrange function, L is defined as L = T − V .
V is the algebraic sum of the gravity potential energy Vg and the
potential energy of the surface tension Vst . Then,
1 ω2
L= (ρ + ρ ) − (ρ − ρ )g − λk2 a2 λ (11.30)
4 |k|
Substituting this into the Euler−Lagrange equation,
 
d ∂L ∂L
· − =0 (11.31)
dt ∂ xi ∂xi

The variable of the generalized coordinates is only a. We can get


 

ρ − ρ γ
ω 2 = |k| g+ k2 (11.32)
ρ + ρ ρ + ρ

where g is the acceleration, ρ and ρ are the mass densities of the

two kinds of fluids, with ρ > ρ , and ρ − ρ /ρ + ρ in the first term
is called the Atwood number. This expresses the relation between
frequency and number of the wave (or wavelength), which is called
the dispersion relation.
Capillary Wave 239

11.3 Analysis of the Characteristics of Free Surface


Waves and the Concept of Capillary Waves

Wave velocity is the transmission rate of a wave. According to the


different characteristics of waves, the definitions of wave velocity are
different, such as phase velocity, group velocity, signal velocity, and
front velocity. Generally, wave velocity refers to phase velocity, which
is discussed in this section. The phase velocity of a wave is the rate
at which the phase of the wave transfers in space. This is the velocity
at which the phase of any frequency component of the wave travels.
Any given phase of the wave (for example, the peak) will travel at
the phase velocity, as shown in Figure 11.3.
As an example of the phase velocity of a wave peak, the
corresponding phase is θ = 0. Assuming that the wave peak reaches
x1 at time t1 and reaches x2 at time t2 , since kx1 − ωt1 = 0 and
kx2 − ωt2 = 0, the transformation distance per unit time is the
phase velocity of the wave peak, which can be calculated as follows:

x2 − x1 ω
C= = (11.33)
t2 − t1 k

Therefore, C can be expressed using angle frequency ω and wave


number k.
In addition, since 2π/k reflects the wavelength, the length of wave
in one period is λ = 2π/k, so k = 2π/λ. Substituting it into the

Figure 11.3. Diagram of the transformation of wave peak.


240 Capillary Mechanics

dispersion expression (11.32), it can be deduced that

gλ ρ − ρ 2πγ
C2 = + (11.34)
2π ρ + ρ λ(ρ + ρ )
where C is the wave velocity, g is the acceleration due to gravity, λ
is the wavelength, ρ and ρ are the densities of the upper and lower
layers of the fluid, and γ is the surface tension.
The first term on the right-hand side of equation (11.34) includes
g, which reflects the effect of gravity. The second term includes γ,
which reflects the effect of the surface tension. For waves with very
short wavelengths, the smaller the wavelength, the greater the influ-
ence of the surface tension on wave velocity. For waves with very
long wavelengths, the larger the wavelength, the greater the effect of
gravity on the wave velocity. When the wavelength is equal to

γ
λc = 2π (11.35)
(ρ − ρ )g

the wave velocity has a minimum value of Cmin , and



2 ρ − ρ gγ
Cmin = 2 (11.36)
ρ + ρ ρ − ρ

For short waves, since the wavelength is comparatively short, λ <


λc , the second term on the right-hand side of equation (11.34) is
dominant. The role of surface tension is remarkable. Kelvin suggested
that this kind of wave should be called a “ripple”, which is also known
as a capillary wave. For capillary waves, the shorter the wavelength,
the greater the velocity. The wave velocity decreases with an increase
in the wavelength. When λ = λc , surface tension and gravity work
on the fluctuations at the same time. This wave is called a capillary-
gravity wave. Only when λ > λc , the effect of surface tension reduces.
In this case, fluctuations are mainly caused by gravity, and the wave
is called a gravity wave.
For a two-layer fluid which is composed of air and water, if
the surface tension is γ = 0.073 N/m, the density of water is ρ =
1000 kg/m3 , the density of air is ρ = 1.299 kg/m3 , and g = 9.8 m/s2 ,
then the wavelength is λmin = 1.78 cm at the minimum wave velocity.
The minimum wave velocity is Cmin = 23.2 cm/s.
Bibliography

[Anon.] (1911). Capillary action. In Encyclopedia Britannica.


Aaronson, S. (2005). NP-complete problems and physical reality. SIGACT
News, 36(1), 30–52.
Adam, N. K. (1941). The Physics and Chemistry of Surfaces (3rd ed.).
Oxford University Press.
Adamson, A. W., and Gast, A. P. (1997). Physical Chemistry of Surfaces
(6th ed.). Wiley.
Baierlein, R. (2003). Thermal Physics. Cambridge University Press.
Ball, W. W. R. (2003). Pierre Simon Laplace (1749–1827). In A Short
Account of the History of Mathematics (4th ed.). Dover.
Bennett, C. O., and Myers, J. E. (1962). Momentum, Heat, and Mass Trans-
fer. McGraw-Hill.
Bhushan, B. (1999). Principles and Applications of Tribology. Wiley.
Bormashenko, E. (2008). Why does the Cassie-Baxter equation apply? Col-
loids Surf. A, 324, 47–50. Lin, F. (2008). Petal effect: Two major exam-
ples of the Cassie-Baxter model are the petal effect and lotus effect —
A superhydrophobic state with high adhesive force. Langmuir, 24(8),
4114–4119.
Bowden, F. P., and Tabor, D. (1954). Friction and Lubrication of Solids,
Part I. Oxford University Press.
Bowden, F. P., and Tabor, D. (1964). Friction and Lubrication of Solids,
Part II. Oxford University Press.
Boys, C. V. (1911). Soap Bubbles: Their Colours and the Forces That Mould
Them (2nd ed.).
Bush, J. W. M. (2004a). MIT lecture notes on surface tension, lecture 1.
Massachusetts Institute of Technology.
Bush, J. W. M. (2004b). MIT lecture notes on surface tension, lecture 3.
Massachusetts Institute of Technology.

241
242 Capillary Mechanics

Bush, J. W. M. (2004c). MIT lecture notes on surface tension, lecture 5.


Massachusetts Institute of Technology.
Bush, J. W. M. (2004d). Surface tension module. MIT OpenCourseWare.
Cabezas, G., Montanero, J. M., Acero, J., Jaramillo, M. A., and Fernández,
J. A. (2002). Detection of liquid bridge contours and its applications.
Meas. Sci. Technol., 13, 725–732.
Calvert, J. B. (2007). Surface tension (physics lecture notes). University of
Denver.
Chau, A., Regnier, S., Delchambre, A., and Lambert, P. (2007). Influence of
geometrical parameters on capillary forces. In Proceedings of the 2007
IEEE International Symposium on Assembly and Manufacturing (pp.
215–220).
Chow, T. S. (1998). Wetting of rough surfaces. J. Phys. Condens. Matter,
10(27), L445–L451.
Christenson, H. K. (1988). Adhesion between surfaces in unsaturated
vapors: A reexamination of the influence of meniscus curvature and
surface forces. J. Colloid Interface Sci., 121(1), 170–178.
Christenson, H. K., and Yaminsky, V. V. (1993). Adhesion and salvation
forces between surfaces in liquids studied by vapor-phase experiments.
Langmuir, 9, 2448–2454.
Comyn, J. (1997). Adhesion Science. Royal Society of Chemistry.
Dataphysics (2007). Sessile drop method. Technical Documentation.
de Gennes, P. G. (1985). Wetting: Statics and dynamics. Rev. Mod. Phys.,
57(3), 827–863.
de Gennes, P. G. (1994). Soft Interfaces. Cambridge University Press.
de Gennes, P. G., Brochard-Wyart, F., and Quéré, D. (2002). Capillary and
Wetting Phenomena — Drops, Bubbles, Pearls, Waves. Springer.
Dean, J. A. (Ed.). (1973). Lange’s Handbook of Chemistry (10th ed.).
McGraw-Hill.
Dingemans, M. W. (1997). Water Wave Propagation Over Uneven Bottoms.
World Scientific.
Ertl, G., Knözinger, H., and Weitkamp, J. (1997). Handbook of Heteroge-
neous Catalysis (Vol. 2). Wiley-VCH.
Eustathopoulos, N., Nicholas, M. G., and Drevet, B. (1999). Wettability at
High Temperatures. Pergamon.
Fan, H., and Gao, Y. X. (2001). Elastic solution for liquid-bridging-induced
microscale contact. J. Appl. Phys., 90(12), 5904–5910.
Fan, H., and Wang, G. F. (2003). Stability analysis for liquid-bridging
induced contact. J. Appl. Phys., 93(5), 2554–2558.
Ferraro, P. (1998). What breaks the shadow of the tube? Phys. Teach., 36,
542–543.
Bibliography 243

Ferraro, P. (2008). Wettability patterning of lithium niobate substrate by


modulating pyroelectric effect to form microarray of sessile droplets.
Appl. Phys. Lett., 92, 213107.
Finn, R. (1999). Capillary Surface Interfaces. American Mathematical
Society.
Fisher, L. R., and Israelachvili, J. N. (1981). Direct measurement of the
effect of meniscus forces on adhesion: A study of the applicability of
macroscopic thermodynamics to microscopic liquid interfaces. Colloids
Surf., 3, 303–319.
Gao, C. (1997). Theory of menisci and its applications. Appl. Phys. Lett.,
71, 1801–1803.
Gao, S., Jin, L., Du, J., and Liu, H. (2011). The liquid-bridge with large
gap in micro structural systems. J. Mod. Phys., 2, 404–415.
Gao, S., and Liu, H. (2008). Mechanics of Micro-Electrical-Mechanism Sys-
tems. National Defense Industry Press.
Goljan, E. (2007). Pathology (2nd ed.). Mosby Elsevier.
Good, R. J. (1992). Contact angle, wetting, and adhesion: A critical review.
J. Adhes. Sci. Technol., 6(12), 1269–1302.
Gregg, S. J., and Sing, K. S. W. (1982). Adsorption, Surface Area and
Porosity (2nd ed.). Academic Press.
Harrison, C. (Ed.). (1960). Handbook of Chemistry and Physics (42nd ed.).
CRC Press.
He, M., Blum, A. S., Aston, D. E., Buenviaje, C., and Overney, R. M.
(2001). Critical phenomena of water bridges in nanoasperity contacts.
J. Chem. Phys., 114(3), 1355–1360.
International Union of Pure and Applied Chemistry Commission on Atmo-
spheric Chemistry. (1990). Glossary of atmospheric chemistry terms.
Pure Appl. Chem., 62, 2167–2219.
International Union of Pure and Applied Chemistry Commission on Physic-
ochemical Symbols Terminology and Units. (1993). Quantities, Units
and Symbols in Physical Chemistry (2nd ed.). Blackwell Scientific Pub-
lications.
International Union of Pure and Applied Chemistry Commission on Quan-
tities and Units in Clinical Chemistry, & International Federation of
Clinical Chemistry Committee on Quantities and Units. (1996). Glos-
sary of terms in quantities and units in clinical chemistry (IUPAC-
IFCC Recommendations 1996). Pure Appl. Chem., 68, 957–1000.
https://doi.org/10.1351/pac199668040957.
Israelachvili, J. (2004). Intermolecular and Surface Forces (3rd ed.).
Academic Press.
Jang, J., Schatz, G. C., and Ratner, M. A. (2004). Capillary force in atomic
force microscopy. J. Chem. Phys., 120(3), 1157–1160.
244 Capillary Mechanics

Johnson, R. E. (1993). Wettability. In J. C. Berg (Ed.), Surface Chemistry


and Applications. Marcel Dekker.
Jurin, J. (1719). An account of some experiments shown before the Royal
Society. Philos. Trans. R. Soc. Lond., 30, 739–747.
Katchalsky, A., and Curran, P. F. (1965). Nonequilibrium Thermodynamics
in Biophysics. Harvard University Press.
Kinloch, A. J. (1987). Adhesion and Adhesives: Science and Technology.
Chapman and Hall.
Kirkness, J. P., Christenson, H. K., Wheatley, J. R., and Amis, T. C. (2005).
Application of the ‘pull-off’ force method for measurement of surface
tension of upper airway mucosal lining liquid. Physiol. Meas., 26, 677–
688.
Lamb, H. (1928). Statics, Including Hydrostatics and the Elements of the
Theory of Elasticity (3rd ed.). Cambridge University Press.
Lamb, H. (1994). Hydrodynamics (6th ed.). Cambridge University Press.
Lambert, P., Seigneur, F., Koelemeijer, S., and Jacot, J. (2006). A case
study of surface tension gripping: The watch bearing. J. Micromech.
Microeng., 16(7), 1267–1276.
Langmuir-Blodgett Instruments (2007). Surface and interfacial tension.
Technical Documentation.
Laplace, P. S. (1806). Mécanique Céleste, Supplement to the Tenth Edition.
Lauda (2007). Surfactants at interfaces. Technical Documentation.
Lee, K. S. (2008). Kinetics of wetting and spreading by aqueous surfactant
solutions. Adv. Colloid Interface Sci., 144, 54–65.
Leenaars, A. F. M., Huethorst, J. A. M., and van Oekel, J. J. (1990).
Marangoni drying: A new extremely clean drying process. Langmuir,
6, 1701–1703.
Marangoni, C. (1865). On the expansion of a drop of liquid floating in the
surface of another liquid. Published Dissertation.
Marmur, A. (1992). Modern approach to wettability: Theory and applica-
tions. In M. E. Schrader & G. Loeb (Eds.), Wettability. Plenum Press.
Marmur, A. (2003). Wetting of hydrophobic rough surfaces: To be hetero-
geneous or not to be. Langmuir, 19, 8343–8348.
McFarlane, J. S., and Tabor, D. (1950). Proc. R. Soc. Lond. A (pp. 202–
224).
McGraw-Hill. (2003). Jurin rule. In McGraw-Hill Dictionary of Scientific
and Technical Terms.
Mendoza, E. (1988). Reflections on the Motive Power of Fire — and Other
Papers on the Second Law of Thermodynamics by E. Clapeyron and R.
Carnot. Dover Publications.
Moore, W. J. (1962). Physical Chemistry (3rd ed.). Prentice Hall.
Bibliography 245

Müller, I. (2007). A History of Thermodynamics — The Doctrine of Energy


and Entropy. Springer.
Okumura, K. (2008). Wetting transitions on textured hydrophilic surfaces.
Eur. Phys. J. E, 25, 415–424.
Orr, F. M., Scriven, L. E., and Rivas, A. P. (1975). Pendular rings between
solids: Meniscus properties and capillary force. J. Fluid Mech., 67,
723–742.
Oxford University Press. (1989). Formula. In The Oxford English Dictio-
nary (2nd ed.).
Pakarinen, O. H., Foster, A. S., Paajanen, M., Kalinainen, T., Katainen,
J., Makkonen, I., Lahtinen, J., and Nieminen, R. M. (2005). Towards
an accurate description of the capillary force in nanoparticle-surface
interactions. Model. Simul. Mater. Sci. Eng., 13(7), 1175–1186.
Perrot, P. (1998). A to Z of Thermodynamics. Oxford University Press.
Petrovic, S., Robinson, T., and Judd, R. L. (2004). Marangoni heat transfer
in subcooled nucleate pool boiling. Int. J. Heat Mass Transfer, 47(23),
5115–5128.
Pfitzner, J. (1976). Poiseuille and his law. Anaesthesia, 31(2), 273–275.
Phillips, O. M. (1977). The Dynamics of the Upper Ocean (2nd ed.). Cam-
bridge University Press.
Physical Properties Sources Index: Eötvös Constant. (2008).
Phywe (2007). Surface tension by the ring method (Du Nouy method).
Technical Documentation.
Quere, D. (2008). Wetting of textured surfaces. Colloids Surf., 206, 41–46.
Reiss, H. (1965). Methods of Thermodynamics. Dover Publications.
Rowlinson, J. S., and Widom, B. (1982). Molecular Theory of Capillarity.
Clarendon Press.
Safran, S. (1994). Statistical Thermodynamics of Surfaces, Interfaces, and
Membranes. Addison-Wesley.
Salzman, W. R. (2001). Open systems. In Chemical thermodynamics. Uni-
versity of Arizona.
Schrader, M. E., and Loeb, G. I. (1992). Modern Approaches to Wettability:
Theory and Applications. Plenum Press.
Scriven, L. E., and Sternling, C. V. (1960). The Marangoni effects. Nature,
187, 186–188.
Sears, F. W., and Zemanski, M. W. (1955). University Physics (2nd ed.).
Addison Wesley.
Sharfrin, E. (1960). Constitutive relations in the wetting of low energy sur-
faces and the theory of the retraction method of preparing monolayers.
J. Phys. Chem., 64(5), 519–524.
Sherwood, L. (2007). Human physiology from cells to systems. In P. Adams
(Ed.), Human Physiology (6th ed.). Thomson Brooks/Cole.
246 Capillary Mechanics

Shinto, H., Uranishi, K., Miyahara, M., and Higashitani, K. (2002).


Wetting-induced interaction between rigid nanoparticle and plate: A
Monte Carlo study. J. Chem. Phys., 116(21), 9500–9509.
Stifter, T., Marti, O., and Bhushan, B. (2000). Theoretical investigation
of the distance dependence of capillary and van der Waals forces in
scanning force microscopy. Phys. Rev. B, 62(20), 13667–13673.
Sutera, S. P., and Skalak, R. (1993). The history of Poiseuille’s law. Annu.
Rev. Fluid Mech., 25, 1–19.
Tadmor, R. (2004). Line energy and the relation between advancing, reced-
ing and Young contact angles. Langmuir, 20, 7659–7664.
Thomson, J. (1855). On certain curious motions observable on the surfaces
of wine and other alcoholic liquors. Philos. Mag., 10, 330–333.
Thomson, W. T. (1871). On capillary attraction. Philos. Mag., 42, 448–452.
Tufillaro, N. B., Ramshankar, R., and Gollub, J. P. (1989). Order-disorder
transition in capillary ripples. Phys. Rev. Lett., 62(4), 422–425.
Van Krevelen, D. W. (1976). Properties of Polymers (2nd ed.). Elsevier.
Washburn, E. W. (1921). The dynamics of capillary flow. Phys. Rev., 17(3),
273–283.
White, H. E. (1948). Modern College Physics. van Nostrand.
Whyman, G. (2008). The rigorous derivation of Young, Cassie–Baxter and
Wenzel equations and the analysis of the contact angle hysteresis phe-
nomenon. Chem. Phys. Lett., 450, 355–359.
Wu, D., Fang, N., Sun, C., and Zhang, X. (2006). Stiction problems in
releasing of 3d microstructures and its solution. Sens. Actuat. A, 128,
109–115.
Young, T. (1805). An essay on the cohesion of fluids. Philos. Trans. R. Soc.
Lond., 95, 65–87.
Yuan, J. Y., Shao, Z., and Gao, C. (1991). Alternative method of imag-
ing surface topologies of nonconducting bulk specimens by scanning
tunneling microscopy. Phys. Rev. Lett., 67(7), 2901–2904.
Zhang, B. and Nakajima, A. (1999). Nanometer deformation caused by
the Laplace pressure and the possibility of its effect on surface tension
measurements. J. Colloid Interface Sci., 211, 114–121.
Index

A I
adhesion, 5–7, 28–31, 48, 76, 85, interfacial tension, 27, 31, 48–49, 54,
89–91, 97–99, 102, 183–186, 56, 62, 69, 74, 98–99, 141–143,
214–215, 217–219, 221 193–194

C K
capillarity, 1, 3–5, 7, 9–13, 15–16 Kelvin equation, 10, 15, 191
capillary, 55–57, 60–61, 63, 76, 102,
107, 115, 122–123, 134–135, 147, M
150, 163–178, 183, 186–187, M/NEMS, 4–7, 9, 16, 181–182
190–192, 195, 213–218, 223–225, Marangoni effect, 10, 13,
227–229 143–145, 148–151, 153, 155–161,
capillary wave, 231–233, 240 184
cohesion, 26–28, 31, 45, 47, 49, 52, Micro-Electro-Mechanical systems
89–91, 93, 98, 102, 135, 185–186 (MEMS), 4–5, 7, 182–184, 213–220,
224, 228–229
D
Du Noüy ring, 55, 57, 74, 226 N
interfacial tension(s), 51
G
Gibbs free energy, 14, 23–24, 32–36, S
38, 131, 134 state of matter, 17, 20
surface external force, 31–32
H surface free energy, 22, 25–27, 29,
Hagen–Poiseuille equation, 164, 172 41–42, 44–45, 47–48, 95, 99, 118,
Helmholtz free energy, 10, 15, 23–24, 234
35 surface internal force, 31

247
248 Capillary Mechanics

W Y
wetting, 2, 11, 55–57, 62, 86–89, 91, Young–Laplace equation, 61,
93, 96, 99–100, 109, 123, 147, 116, 118, 123–125, 129,
170–171, 182–184, 211–212 141, 171, 186–187, 189,
Wilhelmy plate, 55, 62 196

You might also like