Introduction To Cell Mechanics and Mechanobiology
Introduction To Cell Mechanics and Mechanobiology
Introduction to Introduction to
Cell Mechanics and Cell Mechanics and
Mechanobiology Mechanobiology
Christopher R. Jacobs, Hayden Huang, Ronald Y. Kwon
Jacobs Huang Kwon
biomedical engineering, bioengineering, and mechanical engineering
will also find it a useful reference. Advanced analyses and mathematical
derivations are presented as optional boxes along with solved examples in
the main text; each chapter ends with problems and annotated references.
This book contains information obtained from authentic and highly regarded sources. Every effort has been made to trace copyright holders
and to obtain their permission for the use of copyright material. Reprinted material is quoted with permission, and sources are indicated.
A wide variety of references are listed. Reasonable efforts have been made to publish reliable data and information, but the author and the
publisher cannot assume responsibility for the validity of all materials or for the consequences of their use. All rights reserved. No part of this
publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means—graphic, electronic, or mechanical,
including photocopying, recording, taping, or information storage and retrieval systems—without permission of the copyright holder.
ISBN 978-0-8153-4425-4
Published by Garland Science, Taylor & Francis Group, LLC, an informa business,
711 Third Avenue, New York, NY, 10017, USA, and 2 Park Square, Milton Park, Abingdon, OX14 4RN, UK.
15 14 13 12 11 10 9 8 7 6 5 4 3 2 1
In recent years, mechanical signals have become widely recognized as being criti-
cal to the proper functioning of numerous biological processes. This has led to the
emergence of a new field called cellular mechanobiology, which merges cell biol-
ogy with various disciplines of mechanics (including solid, fluid, statistical, com-
putational, and experimental mechanics). Cellular mechanobiology seeks to
uncover the principles by which the sensation or generation of mechanical force
alters cell function. Introduction to Cell Mechanics and Mechanobiology presents
students from a wide variety of backgrounds with the physical and mechanical
principles underpinning cell and tissue behavior.
This textbook arose from a cell mechanics course at Stanford University first
offered by two of us in 2005. Over several iterations, we taught from a set of course
notes and chapter excerpts—having found no textbook to cover the necessary
breadth of topics. Our colleagues had similar experiences teaching with the same
adhoc approach, which convinced us of the need for a comprehensive instruc-
tional tool in this area. Another reason we felt compelled to write this text is that
cell mechanics provides an excellent substrate to introduce many types of
mechanics (solid, fluid, statistical, experimental, and even computational). These
topics are traditionally covered in separate courses with applications largely
focused on engineering structures. As authors, we have varied backgrounds, but
share a common fondness for the insights mechanical engineering brings to cell
biology.
Introduction to Cell Mechanics and Mechanobiology is intended for advanced
undergraduates and early graduate students in biological engineering and bio-
medical engineering, including those not necessarily in a biomechanics track. We
do not assume an extensive knowledge in any area of biology or mechanics. We do
assume that students have a mathematics background common to all areas of
engineering and quantitative science, meaning exposure to calculus, ordinary dif-
ferential equations, and linear algebra.
The field of cell mechanics encompasses advanced concepts, such as large defor-
mation mechanics and nonlinear mechanics. We do not expect our audience to
have a strong background in the advanced mathematics of continuum mechan-
ics. Our intent is to avoid graduate-level mathematics wherever possible. In our
approach, the treatment of tensor mathematics—central to large deformation
mechanics (common in cell mechanics)—poses unique difficulties. To show sim-
plified mathematical derivations measuring mechanical parameters in the context
of living cells, we present tensors “by analogy” as matrices, rather than introduc-
ing them in a fully rigorous fashion. For example, we skip index notation entirely.
Admittedly, this approach may be less satisfying to mechanicians, which we also
consider ourselves. However, we hope that the advantages of this approach will
outweigh our oversimplifications.
The book is grouped into two parts: (I) Principles and (II) Practices. We have writ-
ten the chapters to allow instructors flexibility in presentation, depending on the
level of students and the length of the course. After introducing cell mechanics as
vi PREFACE
Online Resources
Accessible from www.garlandscience.com/cell-mechanics, Student and Instructor
Resource websites provide learning and teaching tools created for Introduction to
Cell Mechanics and Mechanobiology. The Student Resources site is open to every-
one, and users have the option to register in order to use book-marking and note-
taking tools. The Instructor’s Resource site requires registration; access is available
to instructors who have assigned the book to their course. To access the Instructor’s
Resource site, please contact your local sales representative or email science@gar-
land.com. Below is an overview of the resources available for this book. Resources
may be browsed by individual chapters and there is a search engine. You can also
access the resources available for other Garland Science titles.
For students:
• Computer simulation modules in two formats: ready-to-run simulations that simu-
late the mechanical behavior of cells and tutorial MATLAB modules on simulation
of cell behavior with the finite element method.
• Color versions of several figures are available, indicated by the figure legend in the
text.
PREFACE vii
The origin of the book is rooted in teaching from sections of outstanding books
by David H. Boal, Jonathon Howard, and Howard C. Berg. We thank Roger
Kamm, Vijay Pande, and Andrew Spakowitz, who taught some of us at various
times and have unselfishly shared course materials and handouts and, in the
case of Dr. Kamm, unpublished drafts of his own textbook. With their permis-
sion, we have incorporated their approach to some topics in Chapters 4, 5, 7, 8,
and 9 and adapted several problems into sections of our book. We are grateful
for their amazing willingness to share their intellectual product in the name of
improving the educational experience of students around the world. We also
thank reviewers Roland R. Kaunas and Peter J. Butler, who shared notes from
their own courses in cell mechanics. We are profoundly appreciative of the tire-
less work of those who have preceded us, without whom we never could have
completed this task. We thank the additional reviewers of the book, Dan Fletcher,
Christian Franck, Wonmuk Hwang, Paul Janmey, Yuan Lin, Lidan You, and
Diane Wagner, for their valuable insight and critiques of our drafts. We are also
grateful to Summers Scholl and the editorial and production teams at Garland
who took a chance on three textbook neophytes and guided us unerringly
through uncharted waters. Finally we are each deeply indebted to our families,
including Roberta, Jolene, VH, YYH, LHH, Joyce, Melody, Tae, and Cynthia.
Without your support, patience, and understanding—as this project took us
away from you on so many nights and weekends—we never could have contem-
plated this undertaking, much less completed it.
Christopher R. Jacobs
Hayden Huang
Ronald R. Kwon
This page intentionally left blank
to match pagination of print book
Detailed Contents
Probability distribution from the canonical 6.2 Measurement of forces produced by cells 160
ensemble gives Boltzmann’s law 138 Traction force microscopy measures the forces
The free energy at equilibrium can be exerted by a cell on its underlying surface 160
found using the partition function 139 Cross-correlation can be used for
The internal energy at equilibrium can be particle tracking 160
determined using the partition function 141 Determining the forces that produced a
Using the canonical approach may be displacement is an inverse problem 163
preferable for analyzing thermodynamic Microfabricated micropillar arrays can
systems 142 be used to measure traction forces directly 165
5.6 Random walks 143 Surface modification can help determine
A simple random walk can be how a cell interacts with its surroundings 166
demonstrated using soccer 143 6.3 Applying forces to cells 167
The diffusion equation can be derived Flow chambers are used for studying
from the random walk 145 cellular responses to fluid shear stress 167
Key Concepts 147 The transition between laminar and
Problems 148 turbulent flow is governed by the
Reynolds number 168
Annotated References 149
Parallel plate flow devices can be designed
for low Reynolds number shear flow 168
Chapter 6: Cell Mechanics in the Fully developed flow occurs past the
Laboratory 151 entrance length 169
6.1 Probing the mechanical behavior of cells Cone-and-plate flow can be used to
through cellular micromanipulation 151 study responses to shear 170
Known forces can be applied to cells through Diverse device designs can be used to
the use of cell-bound beads and an study responses to fluid flow 171
electromagnet 152 Flexible substrates are used for
The dependence of force on distance subjecting cells to strain 172
from the magnet tip can be calibrated Confined uniaxial stretching can lead
through Stokes’ law 152 to multiaxial cellular deformations 172
Magnetic twisting and multiple-pole Cylindrically symmetric deformations
micromanipulators can apply stresses generate uniform biaxial stretch 172
to many cells simultaneously 153 6.4 Analysis of deformation 173
Optical traps generate forces on particles Viscoelastic behavior in micromanipulation
through transfer of light momentum 153 experiments can be parameterized
Ray tracing elucidates the origin of through spring–dashpot models 173
restoring forces in optical tweezers 154 Combinations of springs and dashpots can
What are the magnitudes of forces in be used to model viscoelastic behavior 174
an optical trap? 155 Microscopy techniques can be adapted
How does optical trapping compare to visualize cells subject to mechanical
with magnetic micromanipulation? 156 loading 177
Atomic force microscopy involves the Cellular deformations can be inferred
direct probing of objects with a small from image sequences through
cantilever157 image correlation–based approaches 178
Cantilever deflection is detected using Intracellular strains can be computed
a reflected laser beam 157 from displacement fields 179
Scanning and tapping modes can be 6.5 Blinding and controls 181
used to obtain cellular topography 158 Key Concepts 182
A Hertz model can be used to estimate Problems 183
mechanical properties 158 Annotated References 184
DETAILED CONTENTS xiii
8.3 Affine networks 229 The fluid mosaic model of the cell membrane
Affine deformations assume the describes its physical properties 252
filaments deform as if they are 9.2 Phospholipid self-assembly 252
embedded in a continuum 229 Critical micelle concentration depends
Flexible polymer networks can on amphiphile molecular structure 253
be modeled using rubber elasticity 230 Aggregate shape can be understood
Anisotropic affine networks can be from packing constraints 254
modeled using strain energy approaches 233 9.3 Membrane barrier function 255
Elastic moduli can be computed The diffusion equations relate
from strain energy density 233 concentration to flux per unit area 256
Elastic moduli of affine anisotropic networks Fick’s second law shows how spatial
can be calculated from appropriate concentration changes as a function of time 257
strain energy density and angular
9.4 Membrane mechanics I: In-plane shear
distribution functions 235
and tension 259
8.4 Biomechanical function and
Thin structures such as membranes
cytoskeletal structure 236
can be treated as plates or shells 260
Filopodia are cross-linked bundles
Kinematic assumptions help describe
of actin filaments involved in cell motility 236
deformations 260
Actin filaments within filopodia can be
A constitutive model describes
modeled as elastic beams undergoing
material behavior 262
buckling 236
The equilibrium condition simplifies
The membrane imparts force on the
for in-plane tension and shear 262
ends of filopodia 238
Equilibrium simplifies in the case of
The maximum filopodium length before
shear alone 265
buckling in the absence of cross-linking is
shorter than what is observed in vivo 238 Equilibrium simplifies in the case of
equibiaxial tension 266
Cross-linking extends the maximum
length before buckling 238 Areal strain can be a measure of
biaxial deformation 267
Is the structure of the red blood cell’s
cytoskeleton functionally advantageous? 239 9.5 Membrane mechanics II: Bending 267
Thin structures can be analyzed using the In bending the kinematics are
two-dimensional shear modulus and governed by membrane rotation 268
the areal strain energy density 240 Linear elastic behavior is assumed
Sixfold connectivity facilitates resistance for the constitutive model 269
to shear 242 Equilibrium places conditions on
Fourfold connectivity does not resultant forces and moments 269
sustain shear as well as sixfold 245 Which dominates, tension or bending? 272
Key Concepts 246 9.6 Measurement of bending rigidity 272
Problems 246 Membranes undergo thermal
Annotated References 247 undulations similar to polymers 272
Membranes straighten out with tension 273
Key Concepts 275
Chapter 9: Mechanics of the Cell
Problems 275
Membrane 249
Annotated References 277
9.1 Membrane biology 249
Water is a polar molecule 249
Cellular membranes form by interacting CHAPTER 10: Adhesion, Migration, and
with water 250 Contraction 279
The saturation of the lipid tails determines 10.1 Adhesion 279
some properties of the membrane 251 Cells can form adhesions with the substrate 279
The cell membrane distinguishes inside Fluid shear can be used to measure
and outside 251 adhesion strength indirectly 281
DETAILED CONTENTS xv
B iological cells are the smallest and most basic units of life. The field of cell
biology, which seeks to elucidate cell function through better understanding
of physiological processes, cellular structure, and the interaction of cells with the
extracellular environment, has become the primary basic science for better
understanding of human disease in biomedical research. Until recently, the
study of basic problems in cell biology has been performed almost exclusively
within the context of biochemistry and through the use of molecular and genetic
approaches. Pathological processes may be considered disruptions in biochemi-
cal signaling events. The regulation of cell function by extracellular signals may
be understood from the point of view of binding of a molecule to a receptor on
the cell surface. Basic cellular processes such as cell division are considered in
terms of the biochemical events driving them. This emphasis on biochemistry
and structural biology in cell biology research is reflected in typical curricula and
core texts traditionally used for cell biology courses.
Recently, there has been a shift in paradigm in the understanding of cell func-
tion and disease primarily within the analytical context of biochemistry. In
particular, it has become well established that critical insights into diverse cel-
lular processes and pathologies can be gained by understanding the role of
mechanical force. A rapidly growing body of science indicates that mechanical
phenomena are critical to the proper functioning of several basic cell processes
and that mechanical loads can serve as extracellular signals that regulate cell
function. Further, disruptions in mechanical sensing and/or function have
been implicated in several diseases considered major health risks, such as
osteoporosis, atherosclerosis, and cancer. This has led to the emergence of a
new discipline that merges mechanics and cell biology: cellular mechanobiol-
ogy. This term refers to any aspect of cell biology in which mechanical force is
generated, imparted, or sensed, leading to alterations in cellular function. The
study of cellular mechanobiology bridges cell biology and biochemistry with
various disciplines of mechanics, including solid, fluid, statistical, experimen-
tal, and computational mechanics.
The primary goal of this introductory chapter is to motivate the study of cell
mechanics and cellular mechanobiology by: (1) demonstrating its role in basic
cellular and pathological processes; and (2) showing how cell mechanics pro-
vides an ideal framework for introducing a broad mechanics curriculum in an
integrated manner. We first present cell mechanics in the context of human dis-
ease by providing a survey of physiological and pathological processes that are
mediated by cell mechanics and can be better understood through mechanical
analyses. Next, we propose cell mechanics as an ideal substrate for introducing
principles of solid, fluid, statistical, experimental, and even computational
mechanics, and put forth the argument that cell mechanics may be the grand
challenge of applied mechanics for the twenty-first century. Finally, we present
a simple model problem: micropipette aspiration, in which a cell is partly
“sucked” into a narrow tube by a vacuum. This example will help you develop
a feeling for how cell mechanics is studied and demonstrate how a relatively
4 CHAPTER 1: Cell Mechanics as a Framework
simple approach can give important insight into cell mechanical behavior (and
how this behavior can dictate cellular function).
heart muscle rather than for functional pumping, given that the heart does not
need to pump blood in any serious manner in utero.
• Post-birth, brain development and angiogenesis all centrally involve cells’ ability to
interact with their dynamic mechanical environment.
• Cardiovascular diseases such as hypertension and heart failure often result from
long-term mechanical influences. Indeed, cardiac hypertrophy is one of the most
common responses to changes in forces. The distinction between healthy hypertro-
phy (resulting from exercise) versus pathological hypertrophy (resulting from poor
health) is still not well understood.
• The fundamental cellular processes of membrane trafficking, endocytosis and exo-
cytosis (the ways in which a cell engulfs or expels substances, respectively), micro-
tubule assembly and disassembly, actin polymerization and depolymerization,
dynamics of cell–matrix and cell–cell adhesions, chromosome segregation, kine-
tochore dynamics (such as DNA motion during cell division), cytoplasmic protein
and vesicle sorting and transport, cell motility, apoptosis (“programmed cell
death”), invasion (motion of a cell to where it is not usually located), and prolifera-
tion and differentiation (specialization of a cell to a phenotype with a particular
function) are all regulated, at least in part, by mechanical forces.
In the sections that follow, we examine a few of these examples in more detail.
Figure 1.1 Hearing occurs via mechanotransduction by the inner ear hair cell. (A) The bundle of cilia extending from the apical
surface of the cell is deflected by pressure waves in the cochlear lumen. (B) Tiny actin bundles called tip links are stretched as the cilia
deflect due to the pressure wave. (C) The tip links are attached to calcium channels or pores that open and allow calcium into the cell
where it eventually leads to a nerve impulse. (A, Courtesy of Dr. David Furness; B, from, Jacobs RA, Hudspeth AJ (1990) Symp. Quant. BioI.
55, 547–561. With permission from Cold Spring Harbor Press.)
6 CHAPTER 1: Cell Mechanics as a Framework
ions) along their concentration gradient. In the resting state, the cell keeps its
internal calcium level extremely low (<1 mM) relative to the calcium concentra-
tion outside the cell. When sound is transmitted to the inner ear, the vibrations
cause the cilia to deflect, which in turn stretches the actin filaments. This stretch-
ing creates tension that is transmitted to the channels, causing them to open. So,
when the channel is opened, calcium flows down its concentration gradient, and
the intracellular calcium concentration increases. The kinetics of signaling pro-
teins inside the cell are altered by this change in concentration, and a cascade of
biochemical events is initiated that eventually leads to a depolarization of the cell
and a nerve impulse.
As you might imagine, mechanics is very important in this process. The cilia
need to have the right mechanical characteristics to stand upright, but remain
flexible enough that they can be deflected by sound waves. The actin tip links
need to be strong enough to open the channel and to have the appropriate poly-
mer mechanics behavior so that they are stretched by cilium deflection, but are
not affected by thermal noise (recall that these are very small objects, so the
soup of molecules floating around will periodically collide with them, and can
generate some forces that need to be ignored). In this text, our goal is to build a
foundation and present a framework so that you can consider these questions
effectively.
The good news is that physical loading on your bone from staying active will pro-
tect you from losing bone, although some activities appear to be better than oth-
ers. Ballet is better than swimming, presumably because of the impact loading
involved. In fact, it has been shown that high-level athletes can actually build
bone specifically in regions of the skeleton that experience higher loading during
their sport. Despite its critical importance for human health and its status as a
compelling scientific question, the mechanism that allows bone cells (osteocytes
and osteoblasts primarily) to sense and respond to loading by coordinating the
cellular response remains basically unknown. It has been suggested that the sens-
ing mechanism might involve the cytoskeleton, focal adhesions, adherens junc-
tions, membrane channels, and even the biophysical behavior of the membrane
itself. Indeed there is evidence for each of these and many others, so it seems
likely that several cellular sensors exist, perhaps forming a redundant system.
applied mechanics experienced great growth in the past, more recently it has
undergone some degree of contraction.
Comprehensive mechanical analysis of a cell is extremely complex. There are
one-dimensional linear elements in the cytoskeleton and two-dimensional
curved shells in the cell membrane. There are also three-dimensional solids and
enormous potential for pressure effects and fluid–solid interaction. Indeed, the
overall cell is part–solid, part–liquid, something we call viscoelastic. Not only that,
but the properties of the cellular “substance” change depending on the frequency
with which forces are applied. On top of this, cellular structures are so small that
thermal and entropic effects can play an important role in their mechanics, often
requiring the analytical framework of statistical mechanics to understand their
behavior. Thus, in terms of difficult challenges in applied mechanics with poten-
tial for critical new advances and fundamental insight, it is hard to imagine a more
compelling problem than the cell. The overall picture we wish to present to you is
that much can be explained using basic mechanics, but there is still much to be
done using only slightly more advanced mechanical analysis.
Rpip
Lpro/Rpip = 1
Lpro
Rpip
Lpro/Rpip > 1
Lpro
At this point one should be able to see that geometrically the radius of the protru-
sion of the cell in the first regime is larger than Rpip.
surface tension. This surface tension is what allows some types of insects to walk
on the surface of water.
Po
MODEL PROBLEM: MICROPIPETTE ASPIRATION 13
Pcell
Lpro
Rpro is the radius of the protrusion, Lpro is the length of the protrusion into the
micropipette and Rpip is the radius of the micropipette.
For the portion of the cell that is not in the micropipette, from Equation 1.1 we
know that
2n
Pcell − Patm = , (1.2)
Rcell
where n is the surface tension of the cell. For the protrusion, or the portion of the
cell in the pipette,
2n
Pcell − Ppip = . (1.3)
Rpro
1 1
Patm − Ppip = ∆P = 2n − , (1.4)
Rpro Rcell
which relates the difference between the pressure in the surroundings and in the
pipette with the radius of the cell inside and outside of the pipette for a cell with a
given surface tension. Note that we assume the surface tension is constant through-
out the cell, even in the “crease” region where the cell contacts the micropipette,
and in the protrusion area and the cell membrane outside the micropipette.
1 1
∆P = 2n − . (1.5)
pip
R Rcell
The pressure ΔP is controlled by the user, and Rpip is also known. The cell radius
Rcell can be measured optically under a microscope, allowing one to calculate the
surface tension n. Evans and Yeung performed this experiment using different
micropipette diameters and found a surface tension of (~35 pN/μm) in neutro-
phils. This tension was found to be independent of the pipette diameter, which
supports the validity of the technique. We will see in the next section that surface
tension does not stay exactly the same if the cell is further deformed.
For a liquid drop, the surface tension will remain constant as it is aspirated into a
micropipette. In reality, cells do not behave like a perfect liquid drop. This is
because the cell membrane area increases as it is aspirated, resulting in a slight
increase in surface tension. The increase in tension per unit areal strain is given
14 CHAPTER 1: Cell Mechanics as a Framework
by what is called the areal expansion modulus. Needham and Hochmuth quanti-
fied the areal expansion modulus in neutrophils by aspirating them through a
tapered pipette (Figure 1.8). Applying progressively higher pressures, caused the
cell to advance further into the taper, increasing its surface area while main
taining constant volume (maintaining constant volume is a condition called
incompressibility). The radii at either end of the cell, Ra and Rb, the total volume
V, and the apparent surface area A were measured from the geometry. By appar-
ent we mean that the surface area of the cell is approximated to be smooth, and
we ignore small folds and undulations. The surface tension was calculated using
the Law of Laplace as
1 1
∆P = 2n − . (1.6)
Ra Rb
The “original” radius Ro of the cell was calculated from the volume (assuming the
volume remained constant) as V = 4 3 πRo3, allowing the “original” or undeformed
apparent surface area to be calculated as Ao = 4πRo2 . The areal strain ( A − Ao )/Ao
and surface tension could therefore be found for the same cell as the pressure was
increased, and the cell advanced through the taper. The surface tension was plot-
ted as function of areal strain, and the data was fitted to a line. The areal expansion
modulus was found from the slope of the line and was calculated to be 39 pN/μm.
Extrapolating the fit line to zero areal strain resulted in a resting surface tension of
24 pN/μm in the undeformed state.
(C)
(D)
MODEL PROBLEM: MICROPIPETTE ASPIRATION 15
∆P
elastic solid will not have this instability.
liquid drop: instability
1
Lpro /Rpip
Key Concepts
• The study of cellular mechanobiology bridges • Micropipette aspiration is an early and straightforward
cell biology and biochemistry with various approach for investigating cell mechanics. By modeling
disciplines of mechanics, including solid, fluid, the cell as a liquid drop, we can analyze micropipette
statistical, experimental, and computational aspiration experiments using the Law of Laplace.
mechanics. • Liquid-drop cells exhibit an instability when the radius
• A wide variety of devastating human diseases such as of the aspirated protrusion is equal to the radius of
osteoporosis, heart disease, and even cancer involve the pipette. At this point the cell cannot resist
cell mechanics in a fundamental way. increased pressure and rushes into the pipette.
• Cell mechanics is an excellent substrate for • By observing the movement of cells within the
introducing students to a wide variety of cutting-edge micropipette at the instability point, one can
approaches in mechanics. distinguish two types of cellular behavior cells: those
• Understanding the mechanical behavior of cells whose behavior is dominated by membrane tension
presents a grand challenge in theoretical, (similar to a liquid drop), and those that act as a
computational, and experimental mechanics. continuum solid.
Annotated References 17
Problems
1. As we have seen, when a liquid drop–type cell such as known as alveoli. There are around 150 million alveoli
a neutrophil is subjected to micropipette aspiration, in your lungs. During respiration the alveoli are filled
it becomes unstable. This occurs when the radius of and emptied through the action of the diaphragm and
the protrusion is equal to the radius of the pipette. intercostal muscles in the chest wall. During inhalation
For a cell whose behavior is better approximated by a the pressure outside the alveoli can drop by up to 200
continuum such as a chondrocyte, do you think there is Pa. The layer of cells that line the alveoli is constantly
a maximum pressure that can be exerted on the cell? hydrated. So, we can model them as a bubble of air
What configuration would the cell be in at this point? surrounded by water, assuming that the surface tension
of water is 70 dyn/cm. With this information, what is
2. In our analysis we ignored the friction between the radius of the alveoli? In actuality, the radius of an
the inside of the pipette and the cell wall. Is this a alveoli is about 0.2 mm. They can be this small because
reasonable assumption? Why? How might friction affect the epithelial cells secrete a protein called surfactant
the results of an aspiration experiment if it were large? that reduces the water surface tension. If fact, there is
a developmental condition known as infant respiratory
3. The classic micropipette aspiration experiment can
distress syndrome (IRDS) that occurs when insufficient
be modified to address the instability of a cell with
surfactant protein is produced. What must the surface
liquid-drop behavior so that membrane behavior can be
tension be to keep the alveoli inflated?
measured. This is done with a tapered pipette (Figure
1.8). With this approach the pipette radius changes 5. Consider two adjacent alveoli and the end of a
along its length. Derive the relationship between bronchus such that air can pass easily between them
membrane tension and the pressure in the micropipette as well as to the outside. Assume that they have the
assuming that the two sides of the cell have radii of Ra same radius and that they are in a state of equilibrium
and Rb and pipette pressures are Pa and Pb. with respect to pressure and surface tension. What
would happen if a small volume of air moved from one
4. There are other structures similar to cells that can
alveolus to the other? What must be true of surfactant
be treated with the liquid-drop model. The bronchial
as a result?
passages of the lungs terminate in small spherical sacs
Annotated References
Evans E & Yeung A (1989) Apparent viscosity and cortical tension of documenting the prevalence of, as well as the social and economic
blood granulocytes determined by micropipet aspiration. Biophys. J. costs associated with, musculoskeletal disease.
56, 151–160. An early report on micropipette aspiration can be used
Janmey PA & Miller RT (2011) Mechanisms of mechanical signal-
to determine membrane tension.
ing in development and disease. J. Cell Sci. 124, 9–18. This article
Hochmuth RM (2000) Micropipette aspiration of living cells. J. Biome- discusses what is known about the ability of cells to sense local
chanics 33, 15–22. An excellent review of the use of micropipette stiffness and includes a discussion of diseases and developmental
aspiration to characterize the fundamental behavior of cells and processes.
membranes.
Krahl D, Michaelis U, Peiper HG et al. (1994) Stimulation of bone
Huang H, Kamm RD & Lee RT (2004) Cell mechanics and mecha- growth through sports. Am. J. Sports Med. 22, 751–157. Early evi-
notransduction: pathways, probes, and physiology. Am. J. Physiol. Cell dence that loading of bone through physical activities leads to bone
Physiol. 287, C1–11. An overview of mechanosignaling with some formation.
discussion of techniques as well as an overview of cell-force interac-
Malone AM, Anderson CT, Tummala P et al. (2007) Primary cilia
tions.
mediate mechanosensing in bone cells by a calcium-independ-
Jacobs CR, Temiyasathit S & Castillo AB (2010) Osteocyte mecha- ent mechanism. Proc. Natl. Acad. Sci. USA 104, 13325–13330.
nobiology and pericellular mechanics. Annu. Rev. Biomed. Eng. 12, Early evidence that primary cilia act as mechanosensors in bone
369–400. An extensive review of how mechanics and biology inter- cells.
act at the level of the cell to regulate the skeleton in health and
Needham D & Hochmuth RM (1992) A sensitive measure of sur-
disease. Includes extensive supplemental online material.
face stress in the resting neutrophil. Biophys. J. 61, 1664–1670. An
Jacobs J (2008) The Burden of Musculoskeletal Disease in the United early demonstration of how micropipette aspiration can be used to
States. Rosemont, IL. This monograph is an extensive collection determine membrane properties.
This page intentionally left blank
to match pagination of print book
CHAPTER 2
Fundamentals in
Cell Biology
Table 2.1 The monomer subunits and resulting polymers that make up the
biochemical constituents of the cell.
Monomer Polymer
(along with other terms) can be used to quickly describe locations without having
to reference specific sites, much like north and south can be used if you are not
familiar with local landmarks.
In a similar way, there are a few terms that recur in biomedical writing with which
you should be familiar. We may not use them all in this book, but the diligent stu-
dent will want to get to know these terms. Some common ones are:
• Cell culture: the process of extracting live cells from biological tissue, and the
maintenance and growth of those cells, typically in vitro, to study the physiological
behavior of the cells under controlled conditions.
• In vitro: in a laboratory dish; not inside an organism; literally, “in glass”.
• Ex vivo: sometimes interchanged with in vitro, but occasionally referring to entire
tissues that are cultured in a laboratory dish.
• In vivo: inside a living organism, typically but not always at the natural location.
• In situ: inside a living organism, in the natural location. Growing an ear in a mouse’s
back would be an in vivo but not an in situ experiment.
• Amino acids: the fundamental building blocks (monomers) of proteins.
• DNA: deoxyribonucleic acid, the “hard copy” of genes and their regulatory
components.
• RNA: ribonucleic acid, the “working copy” of genes that assembles proteins.
• Gene: sequence of DNA that encodes a protein.
• Promoter: sequence of DNA that regulates when a particular gene is expressed.
Typically, but not always, near the gene being regulated.
• Probe: A fragment of DNA or RNA that is complementary to a specific target
sequence. The probe is able to bind to the target in a process called hybridization.
Many cellular components are polymers, assembled from subunits called mono-
Nota Bene mers (Table 2.1). The major exception to this is the most abundant constituent of
Inorganic carbon compounds. cells, water, which makes up roughly 70% of a cell’s mass. Inorganic ions are
There are several carbon compounds another exception and are critical to cell metabolism and signaling (“organic”
that are inorganic. Diamond and compounds have carbon, “inorganic” compounds do not). These components are
graphite are obvious examples. vital for the non-inert response of cells to mechanical forces. In some cases, bio-
However, there are no hard-and-fast logical polymers form the physical structure of the cell, and the biomechanical
rules to make the distinction. characterization of the cell relies on the properties of these polymers. In other
cases, the molecules are responsible for signaling. Such molecules can be impor-
tant for signaling in response to mechanical stresses (mechanosignaling), which
we discuss further in Chapter 11.
H O H H O
amino end carboxyl end
N-terminus H2N C C N C C OH C-terminus
R R
distinguishes one amino acid from another. The properties of these side chains
additionally determine the hydrophilic/hydrophobic nature of the local peptide
sequence, which is useful for determining regions of proteins that span mem-
branes and the orientation of the protein. Remarkably, there are only 20 common
amino acids (although there are several more in nature, plus some synthetic ones
that have been created in laboratories). The sequence, or order, in which these
amino acids are assembled gives rise to the remarkable diversity of structure and
function found in proteins. The amino acid sequence is formally referred to as the
primary (1°) structure of the protein.
Amino acids are asymmetric because one side has an amino group and the other
has a carboxyl group. This directionality is preserved in peptides and proteins,
with one end terminated with the amino group, the N-terminal, and the other ter-
minated with the carboxyl group, the C-terminal. When amino acids are assem-
bled, one oxygen from the carboxyl group and two hydrogens (one from each
group) form a molecule of water. This process of assembly is called dehydration
synthesis or condensation. When these bonds are broken, a molecule of water is
utilized in a reaction called hydrolysis.
Once a protein is assembled, it can fold based on the distribution of charges and
steric interactions, the latter being space available for movement based on the
size of the side groups. Some common local folding patterns include a-helices
(single spring-like structures) and b-sheets (mostly flat loops) (Figure 2.2). These
patterns are typically referred to as the secondary (2°) structure, and are mainly a
result of hydrogen bonding. The global folding pattern of the entire protein is
referred to as the tertiary (3°) structure. How proteins fold is important for their
function, because the local shape/charge distribution determines which other
molecules the proteins can interact with. The prediction of folding patterns based
on amino acid sequence is an open problem currently, with computational
models only able to handle short peptide fragments. Finally, the association of
multiple proteins is termed the quaternary (4°) structure. Molecular dynamics
computer simulations are capable of generating information regarding second-
ary and tertiary structures, but are known for being computationally expensive.
Simulation of quaternary structures, which would be a boon for pharmaceutical
design, but are also extremely computationally expensive.
The mechanics involved in folding or unfolding of proteins is a subject of intense
investigation, because such unfolding may occur in your cells. Not only does the
resistance to unfolding contribute to the “stiffness” of the molecule but there are
signaling pathways that may be activated by protein unfolding, as discussed in
later chapters. Knowing the charges and sterics of proteins and how a protein
folds is not only important in biology, but also in mechanobiology.
22 CHAPTER 2: Fundamentals in Cell Biology
carbon
nitrogen
building blocks of DNA DNA strand Figure 2.3 The various molecular
units that combine to form the
phosphate double-stranded DNA polymer are
sugar shown at different scales, from the
5′ 3′
components making up a
+ G
nucleotide to the assembled double
G G C A T
sugar base helix. (Adapted from, Alberts B, Johnson
phosphate nucleotide A, Lewis J et al. (2008) Molecular Biology
of the Cell, 5th ed. Garland Science.)
double-stranded DNA DNA double helix
3′ 3′
5′ 5′
C G G C
A T T A
T A A T
T A A
G C sugar–phosphate G C
backbone
C G C G
C G C G
A T A
G C C G
T A A T
5′ 5′
3′ 3′
hydrogen-bonded
base pairs
able to form hydrogen bonds with each other, but only in a complementary fash-
ion: C with G, and A with T (DNA) or U (RNA). Long nucleic acids can bind to each
other in the famous double-helix arrangement, but only when they have comple-
mentary sequences (Figure 2.3). Although nucleotides can have many roles in the
cell, they are primarily responsible for the storage of genetic information through
the sequence in which the bases are assembled and replicated d uring cell division
(Figure 2.4).
The forces required to unfold nucleic acid polymers are of interest for several
reasons. DNA is usually tightly coiled around proteins called histones to con-
serve space. Further, because DNA is maintained in a double helix, there is
some twisting to the molecule as it unwinds to allow access for DNA replication
or transcription (more on transcription shortly). Finally, DNA activation some-
times depends on binding or association of one part of the gene to another, and
the two parts that require interaction may be separated by a long distance. The
distance between these two parts is guided, at least in part, by how flexible the
DNA is. Understanding the mechanism by which DNA (and RNA) works
requires some comprehension of the mechanical behavior underlying these
molecules.
Each of the carbons in the sugar backbone are labeled 5′ carbon is known as the 5′ end. Genetic sequence is
numerically 1′ through 5′. Adjacent rings are bound to usually given from the 5′ end to the 3′ end in a series of
each other through a phosphate group. This phosphate single-letter codes corresponding to the bases. Many
covalently links the 5′ carbon of one monomer to the 3′ processes involving nucleic acids are also directional, as
carbon of the next. Similar to proteins, this gives a direc- seen with the polymerase chain reaction later in this
tionality to nucleotides and their chains, nucleic acids. chapter.
The end of the polymer where the phosphate binds the
Table 2.2 The 20 common amino acids and their codon sequences. Three codons
DNA replication code for the “stop” sequence. Note that there is significant redundancy.
DNA repair DNA
5′ 3′ Ala-Alanine A GCA GCC GCG GCU
3′ 5′
mRNA synthesis Arg-Arginine R AGA AGG CGA CGC CGG CGU
(transcription)
Asp-Aspartic acid D GAC GAU
mRNA
Asn-Asparagine N AAC AAU
5′ 3′
Met-Methionine M AUG
Trp-Tryptophan W UGG
entire genomic DNA strand encodes genes. These encoding regions are called
exons (“e” for encoding, Figure 2.7). Noncoding regions are called introns;
their function is not precisely known, but some intron regions are known to be
important for alternative splicing. Other intron regions are thought to be
important for DNA folding. mRNA is a compact subset of the genome that con-
tains the gene of interest, but also can be manipulated without disturbing the
original template stored in the DNA. Further, multiple mRNAs can be gener-
ated from a single DNA template, allowing for a relative amplification of pro-
tein expression at the mRNA level than if DNA were used for protein s ynthesis
directly.
transcription unit
transcription
“primary RNA transcript”
5′ capping
RNA splicing
RNA cap 3′ polyadenylation
mRNA AAAA
export
mRNA AAAA
translation
protein
By contrast with genotype, a cell’s phenotype can be It is difficult to determine when a cell has changed
much more difficult to define and is often only the result its behavior sufficiently to have differentiated into
of long-term consensus. This is confounded by the fact another cell type. These changes must, in some sense, be
that the behavior of a cell changes considerably over irreversible and permanent. Yet, in some situations cells
time as a result of changes in the extracellular environ- can lose the phenotypic markers of differentiation in a
ment. In some ways, phenotypic change and differentia- process known as dedifferentiation. Organisms that can
tion is meant to imply more enduring changes to a cell’s regenerate lost limbs are thought to be able to induce a
behavior than simple responses to external stimuli. dedifferentiation process. In some conditions osteocytes
When these stimuli persist for a sufficient time, they can can be induced to make bone. Does this mean that they
lead to changes in phenotype. For example, in bone, the have dedifferentiated into osteoblasts again? Cells often
cell responsible for forming new bone is called the oste- lose differentiation markers of their original cell type.
oblast. An osteoblast may or may not form bone in Have they dedifferentiated? These questions, although
response to biochemical and physical signals, though it they appear simple on the surface, can be scientifically
remains an osteoblast. Occasionally, an osteoblast ambiguous and are often based on functional defini-
becomes embedded in the same bone that it is forming, tions. These issues can become quite contentious and
and becomes more quiescent than an osteoblast. This vary substantially from field to field.
cell is said to have differentiated into a different cell type
called an osteocyte.
Nota Bene
Progenitor cells have the potential to differentiate into multiple different cell
types, thereby acquiring different phenotypes. A cell that has the potential to dif-
RNA undergoes maturation ferentiate into multiple cell types is known as multipotent or pluripotent.
modifications. The newly formed Pluripotent cells can differentiate into cells or all three germ layers (endoderm,
mRNA transcript often undergoes
post-transcriptional modification
mesoderm, and ectoderm), while multipotent cells can differentiate into multi-
before translation into protein. For ple, but more limited, lineages. Cells that can differentiate into any cell type in an
instance, the 3′ end of the mRNA is organism are referred to as totipotent. Progenitor cells that can differentiate into
usually supplemented with a more specialized cell types, but can also undergo cell division without differenti-
polyadenosine tail (poly(A) tail), which ating, are known as stem cells. Both embryonic stem cells (currently only obtain-
allows protein complexes to bind to able from a blastocyst) and adult stem cells (obtainable from a variety of tissues,
the mRNA transcript. These protein
including bone marrow, umbilical cord blood, fat, and muscle) are being studied
complexes can splice out the introns
of the gene to leave behind only the for their potential use in regenerative medicine and tissue engineering. One key
exons for translation. It is also challenge in working with stem cells is that once the desired cell is generated, it
possible to splice out some exons must be conditioned to work with existing tissues. For example, if a partial tissue
that the introns flank, thereby replacement is required, the cells must integrate with the host cells and exhibit
creating several protein variants from the correct strength and durability to continue functioning. Many of these charac-
the same DNA source gene. This teristics depend on cell mechanical properties, as well as mechanobiological
modification is referred to as
alternative splicing.
adaptation, and understanding and manipulating these characteristics remains
a fundamental challenge in tissue engineering.
the coding sequence (the region that encodes the amino acid sequence) for a par-
ticular gene, and begin transcription. The cell is not constrained to up- or down-
regulate (increase or decrease) transcription of all genes equally. To regulate
function and behavior, the cell needs to control transcription of specific genes
without affecting transcription of other genes. DNA-binding proteins known as
transcription factors recognize specific DNA sequences and regulate the binding
of RNA polymerase. Given the complexity of cellular function, there is a dizzying
array of transcription factors. Indeed, most genes require a complex of transcrip-
tion factors to form a protein cluster for transcripts to be produced. Finally, tran-
scription factors are themselves coded from genes, so their availability is also
subject to transcriptional regulation.
Nota Bene consider several strategies. Examination of mRNA levels would be an indication
of the activation of certain genes. One might complement mRNA profiles with
Bacteria do not have nuclei. A
protein measurements—and typically protein measurements are made after
bacterium is a prokaryotic cell—a cell
without a nucleus. Eukaryotic cells mRNA is known to change. This timing is selected because proteins come from
have a nucleus plus many of the mRNA, so one would normally expect the changes in protein concentration to
other organelles. Prokaryotic cells are lag behind the changes in mRNA, which we know only because of the central
nearly always unicellular and do not dogma.
form complex organisms. Originally
they were thought to have no Finally, there is the challenge of determining the exact mechanism by which cells
intracellular organelles, but is now sense and respond to mechanical forces. Although it is established that, stretch-
known that they do have some (such ing a cell will lead to up-regulation of mRNA and protein from certain genes, less
as ribosomes). Their DNA is often is known about how the up-regulation occurs. There are many hypotheses,
organized into a single closed loop including proteins and nucleic acids being directly unfolded by the forces, the
known as a plasmid.
cell membrane itself being flexed, and others. The central dogma provides a
much better idea of what to look for if we know roughly the pathway such up-
regulation takes.
means, such as electric fields and physical forces. These physical signals can play
a fundamental role in many aspects of cell biology and have only recently begun
to be investigated. Physical signaling and the emerging field of cellular mechano-
biology will be discussed in some detail in Chapter 11. We begin here by first con-
sidering traditional biochemical signaling.
target hormone
cells
membrane-
bound signal local
molecule mediator
bloodstream target
cell
Figure 2.10 Signaling between cells can occur through many different mechanisms. (A) Cell–cell contact may or may not be
required. (B) Autocrine and paracrine signaling occurs through the release of cell-secreted soluble chemicals. Paracrine signaling is
depicted here, in which the target and signaling cells are different. (C) Endocrine signaling involves a soluble chemical produced in the cells
of specialized organs and released into circulation. (Adapted from, Alberts B, Johnson A, Lewis J et al. (2008) Molecular Biology of the Cell,
5th ed. Garland Science.)
32 CHAPTER 2: Fundamentals in Cell Biology
effector proteins
metabolic gene regulatory cytoskeletal
enzyme protein protein
altered cell
altered altered gene
shape or
metabolism expression
movement
Intracellular calcium signaling is one of the most studied and understood of the
second messenger systems. Calcium is maintained at a very low level in the
cytoplasm. This is done either by pumping it out of the cell or by sequestering it
in the endoplasmic reticulum (sarcoplasmic reticulum in muscle cells), and to a
lesser extent mitochondria. Transient increases in intracellular calcium can
occur through several molecular mechanisms. Calcium channels in the cell
membrane can allow calcium to flow from outside the cell into the cytoplasm in
response to an extracellular signal. There are also channels to allow the seques-
tered intracellular stored calcium to be released (Figure 2.12). Calcium signal-
ing can be readily studied, by using depletion (removing calcium) or via the use
of fluorescent reporters (Figure 2.13).
Calcium channels fall into two categories, the IP3 recep- tion. It is typically finished in a matter of seconds, after
tor and the ryanodine receptor. IP3, as we have dis- which the calcium is again pumped out and seques-
cussed, is a hydrophobic second messenger released tered in preparation for the next transient. Calcium is
into the cytoplasm from the membrane. The ryanodine known to regulate the activity of several downstream
receptor is sensitive to calcium concentration itself. proteins. One of the most well-understood of these pro-
This forms a highly sensitive positive feedback loop teins is calmodulin. Calmodulin, in turn, regulates the
such that a small elevation in intracellular calcium activities of other enzymatic proteins known as, not
induces a much larger and more rapid release from surprisingly, calmodulin-dependent protein kinases or
intracellular stores. This increase in calcium concentra- CaM kinases.
tion is transient and sometimes called a wave or oscilla-
Nota Bene
Signaling cascades involving large molecules and proteins tend to be more
slower acting than second messenger systems, but have the potential for more
What is ryanodine? The ryanodine specific activity. Some complex but necessary nomenclature is required to
receptor gets its name because it can understand the dynamics of these signaling cascades. The activity of many sign-
be modulated by ryanodine, to which
ryanodine receptors show very high
aling and effector proteins is altered when they become phosphorylated, or
affinity. However, ryanodine is a experience the addition of a phosphate (PO4) group. Typically, the phosphate is
poisonous alkaloid only appearing in donated by a high-energy donor such as ATP. This is a reversible reaction that
plants, and it has no role in cell induces a conformation change, allowing a protein to switch from an inactive to
physiology. The name was given an active state. Likewise, some proteins switch from an active to an inactive
because ryanodine was used in the state when they are phosphorylated. Enzymes that phosphorylate other pro-
first purification of the receptor. It is
teins are known as kinases and those that dephosphorylate other proteins are
an accident of history.
called phosphatases (Figure 2.14). In general phosphorylation only occurs on
three specific amino acid residues within a protein: tyrosine, serine, or threo-
nine. Whereas a kinase phosphorylates another protein, a “tyrosine kinase” is
an enzyme that is restricted to adding a phosphate only to the tyrosine residue
of another protein. The diversity and complexity of protein kinase signaling is
immense. More than 500 different kinases have been identified in humans. The
target of a kinase might itself be a kinase that becomes activated upon phospho-
rylation. These are known as kinase kinases. There are also kinase kinase kinases,
but at this point it becomes so cumbersome that they begin to be referred to as
3-kinases, 4-kinases, etc.
Nota Bene As protein–protein signaling cascades progress, relatively small events, such as
Why are intracellular messengers
binding of a few receptors, can lead to larger and larger biochemical responses.
“second”? The term second This can occur through positive feedback loops at the receptor level or from cata-
messenger system stems from the lytic reactions within the cascade itself, in which a single protein can catalyze a
concept that the extracellular signal downstream reaction repeatedly, amplifying the signal. This intracellular amplifi-
represents the first “message.” The cation can be quite profound and sometimes results in cells being exquisitely sen-
cell then converts this first message sitive to small extracellular signals. Protein–protein signaling pathways are critical
to a second message by because they allow for the intricate signaling cascades that the cell needs to
transduction.
orchestrate its varied and complex functions (Figure 2.15).
signal molecule
P P
GPCR
RTK
P P
P P
Ras-GEF (Sos)
adenylyl cyclase IP3 diacylglycerol PI(3,4,5)P3
Ras
Ca2+
MAP kinase kinase kinase
cyclic AMP PDK1
calmodulin MAP kinase kinase
Figure 2.15 Intracellular signaling networks can be exceedingly complex. They can involve negative and positive feedback,
cross-talk and even redundancies. Although the details of the pathways in this schematic are beyond the scope of this course, it illustrates
most of the major signal systems at work in typical cellular regulation and allows cellular responses that are both sensitive and specific
(Adapted from, Alberts B, Johnson A, Lewis J et al. (2008) Molecular Biology of the Cell, 5th ed. Garland Science.)
Adaptor proteins do not have any enzymatic activ- when bound to each other. Alternatively, protein com-
ity themselves, but are required to bind to other proteins plexes may be subsequently transported to other regions
to allow them to become active. For instance, they can of the cell or their formation may be a signal for degrada-
bind to particular recognition sites on a protein and induce tion of the protein complex. Adaptor proteins add another
a conformational change that exposes a particular residue layer of regulation allowing additional specificity in cell
for phosphorylation. This may be necessary for the forma- signaling.
tion of complexes of multiple proteins that only function
several classes depending on how they acheive this. One such class is the receptor
tyrosine kinases. These are transmembrane proteins with an extracellular domain
separated from an intracellular domain by one or more membrane-spanning
domains. The extracellular domain is responsible for recognizing and specifically
binding to the ligand target of the receptor. Ligand binding leads to a conforma-
tional change in the intracellular domain. Receptor tyrosine kinases, like many
receptors, are actually dimers of two proteins. These dimers are only stable in the
presence of the ligand. When a stable dimer forms, tyrosines in the cytoplasmic
domains of the receptor are autophosphorylated and are thereby activated to ini-
tiate the intracellular signaling cascade. At this point, the specific signaling path-
way that is activated will depend on the particular receptor.
Another class of membrane receptors comprises the G-protein-coupled receptors
(GCPRs). They contain seven membrane-spanning domains and are each linked
to a G-protein, as the name suggests. Before ligand binding, the G-protein coupled
to the receptor is in its deactivated guanosine diphosphate (GDP)-bound form.
Upon ligand binding, the receptor acts as a guanine nucleotide exchange factor
(GEF) and allows a guanosine triphosphate (GTP) to substitute for the GDP,
thereby activating the G-protein. The activated G-protein dissociates from the
receptor and is free to diffuse and activate several downstream targets. A hydro-
phobic domain in the G-protein tends to keep it membrane-bound. Not surpris-
ingly then, the targets of the liberated G-protein are other membrane-bound
proteins. GCPRs also activate second messenger signaling systems. These include
production of IP3 and DAG through the action of the G-protein on phosphodiester-
ases and phospholipases and production of cAMP and cGMP by activation of
membrane-bound adenylyl cyclases. GPCRs can even potentiate intracellular cal-
cium increases by opening membrane calcium channels.
Integrins are a particular type of receptor that also act as mechanical linkers of the
cell to its extracellular environment (although they are not the only receptors that
do this, integrins are among the best-studied). Integrin ligands are extracellular
matrix proteins such as collagen, fibronectin, and laminin. In addition to providing
a mechanical connection between the cell and its environment, when integrins
bind they also activate intracellular signaling through integrin-associated kinases.
This signaling results in increased integrin production, recruitment and clustering
of integrins, and the eventual assembly of compact attachments known as focal
adhesions. In many cell types integrin is also an important cell survival signal as
well as a regulator of proliferation and differentiation. As we shall see, it also
appears to be important in helping the cell sense mechanical stress upon itself.
This is known as outside-in integrin signaling, in contrast with inside-out signaling.
In the case of inside-out signaling, cytoplasmic signals regulate the ability of an
integrin to bind its ligand by exposing or hiding the ligand-binding domain.
A dichroic mirror is the heart of an epifluorescence the illumination light and the emitted light from the
microscope. It is reflective below a critical wavelength specimen to travel through the same objective, but for
and transparent to light above that wavelength. It allows the emitted light to be separated and observed.
EXPERIMENTAL BIOLOGY 39
light
source beam-splitting
(or dichroic) mirror
An emission filter helps improve the image by removing light at wavelengths dif- Nota Bene
ferent from a band centered around the dye’s emission wavelength being used
Cytoskeletal and nuclear
(Figure 2.18). In this way, high-quality fluorescence images can have a virtually
fluorescence imaging. A common
black background with a very high signal-to-noise ratio. cell structure of interest is the
cytoskeleton. Actin filaments can be
Fluorophores can highlight structures visualized using antibodies against
F-actin via immunofluorescence, or
Fluorescence microscopy can be used to examine specific structures within with various compounds such as
phalloidin, (which is typically attached
cells. Certain compounds can be attached to molecules that only bind to par- to, or conjugated to, a fluorescent
ticular proteins within the cell to determine their distribution. A broad way of molecule). Phalloidin is generally
doing this is using a technique called immunofluorescence. In immunofluores- specific for actin (phalloidin staining
cence, an antibody is generated that is specific for the protein of interest. This is very fast compared with immuno
antibody (called the primary antibody) can be linked to a fluorescent molecule. fluorescence, hence it is widely used
The cell is incubated with this antibody and then washed to remove unbound despite requiring the cells to be fixed
antibodies. Upon examination using fluorescence microscopy, only the pro- as with immunofluorescence).
Unfortunately, the microtubules and
teins that were tagged with the fluorescent antibody would appear, providing intermediate filaments do not
information about the location of the protein, and to some extent its quantity. currently have widely available
Alternatively, a nonfluorescent primary antibody can be followed with a fluo- fluorescence agents and must be
rescent secondary antibody, which binds the primary antibody. This can also visualized using immunofluorescence
or transfection. The nucleus can be
visualized using several dyes; DAPI
(4′,6-diamidino-2-phenylindole) and
propidium iodide are good for
imaging fixed cells, Hoechst is a
secondary antibodies
primary antibody cell-permeant fluorescent dye.
marker Nuclear labels are generally toxic to
cells; these labels should not be used
immobilized for long-term (over days) cell tracking.
Other structures (mitochondria) can
antigen
be highlighted using various
compounds; an exhaustive listing
would be impractical here.
Figure 2.19 The use of a secondary antibody means that fluorophores or other
means of visualization only need to be linked to one antibody. Additionally, there is
potential for signal amplification, since multiple secondaries can bind to a single primary.
(Adapted from, Alberts B, Johnson A, Lewis J et al. (2008) Molecular Biology of the Cell, 5th ed.
Garland Science.)
40 CHAPTER 2: Fundamentals in Cell Biology
amplify the signal, as many secondary antibodies can attach to each primary
antibody (Figure 2.19).
it is bound to calcium. Fluorescent reporters for other ions, pH, and even voltage
have been developed.
buffer
gel + anode
buffer
42 CHAPTER 2: Fundamentals in Cell Biology
Nota Bene during migration since large proteins can fold into a tightly packed small struc-
ture. Conversely, smaller proteins that are not well packed may actually migrate
Antibodies can be used for
more slowly. In denaturing conditions, proteins acquire an unfolded linear
protein detection and
recognition. Antibodies bind to structure and migrate inversely with their molecular weight.This means gels for
specific protein fragments known as proteins in denaturing conditions (such as SDS) separate samples according to
antigens that contain a sequence of their molecular weight. It should be noted that similar restrictions apply for
interest (the epitope). A polyclonal nucleotide gels—if a DNA is in a circular plasmid configuration, it will appear
antibody is isolated from the blood of smaller than a linear strand of the same number of base pairs (actually, you will
an animal that has been injected with tend to get multiple bands). Therefore, DNA is generally linearized before, it is
a target protein. Polyclonal antibodies
run through a gel.
will bind to several epitopes on the
target protein. A monoclonal antibody The vast majority of gels use one of two polymers to form the hydrogel network.
is produced in cell culture by High-density gels of polymerized acrylamide are used to separate proteins and
immortalizing a single immune cell
from the exposed animal. Monoclonal
smaller nucleic acid polymers. There are many easy-to-use PAGE (polyacrylamide
antibodies recognize a single specific gel electrophoresis) systems and they generally hold the gels between two glass or
epitope. Polyclonal antibodies can plastic plates that force the electric current to pass directly through the gel.
amplify better (more antibodies stuck Typically, these systems hold the gel vertically with the sample moving from the
to the same target) but variability top positive electrode toward the bottom negative electrode. Higher–molecular-
each time you make more. weight nucleotides (over a few hundred base pairs) are separated in lower density
Monoclonals are more restricted in agarose gels. Agarose systems keep the gel horizontal because polymerized aga-
targeting, but the repeatability is
much higher. rose is not strong enough to support its own weight. To keep the gel hydrated, the
gel is submerged and a larger current passes through both the gel and the bathing
fluid. Although automated systems that perform gel electrophoresis in tiny capil-
lary tubes can greatly increase throughput, running agarose and SDS–PAGE gels
remains a standard procedure in most molecular biology labs.
labeled
probe
in buffer
(E)
positions
labeled
of
labeled bands
markers
Figure 2.22 Example of the steps involved to detect DNA by its radioactivity. To visualize proteins in a gel, you must first transfer
or blot them onto a membrane. Although this can be done electrophoretically, it can also be accomplished simply by soaking water with
paper towels. The membrane can then be incubated with tagged antibodies or nucleotide probes, and visualized with color or fluorescent
dyes so that the presence of a desired protein or nucleic acid fragment can be detected. (Adapted from, Alberts B, Johnson A, Lewis J et al.
(2008) Molecular Biology of the Cell, 5th ed. Garland Science.)
on the accepted name for the Southern blot, when the same approach was used
with RNA in 1977 at Stanford, it was named the northern blot. The good-natured
name-play continued with the development of the western blot for proteins. The
tradition continues with, for instance, blotting of DNA-binding proteins being
called a southwestern blot. Although there is no eastern blot, a method of blotting
lipids developed in Japan was termed far-eastern blotting.
Why are antibodies useful in experimental biology? stick to a wide variety of epitopes while not recognizing
Antibodies are proteins that have a “flexible” binding native proteins (so that the antibodies stick to unwanted
site. These proteins are typically drawn in the shape of a stuff but not to parts of your own body). In some cases,
“Y”. The binding site is usually targeted to a desired pep- this is not desirable; if you receive an organ transplant
tide sequence called an epitope, which is usually part of you do not want your immune system to attack it, but
a foreign organism (say, a surface protein on some virus). since it is technically a foreign body, you will generate
The idea is that when a foreign organism invades your antibodies against it, so immunosuppresants will gener-
body, your immune system generates antibodies that are ally be necessary). In other cases, there are mistakes—in
capable of binding key surface proteins on the foreign autoimmune syndromes, your immune system gener-
organism, thereby rendering the proteins inactive. Since ates antibodies against your own tissues, which can
there are multitudes of foreign organisms, the immune cause all sorts of complications (lupus is a classic exam-
system must be capable of generating antibodies that ple of an autoimmune dysfunction).
provides has led to it be a virtually ubiquitous tool in the modern biology lab. In
recognition of the incredible importance of PCR, Mullis received one of very few
Nobel prizes to be awarded for a technique rather than a scientific discovery. The
key to PCR is to use a natural cellular enzyme, DNA polymerase, to replicate DNA
in vitro rather than in the cell nucleus. As each new DNA strand is synthesized, it
becomes in turn a template for subsequent synthesis in a chain reaction that pro-
duces an exponential increase in DNA and gives PCR its remarkable power to
detect even a single molecule of DNA in the initial sample. At its core, PCR is made
possible by two key components: the requirement of primers for DNA polymerase
binding and the identification of thermally stable DNA polymerases.
PCR would be much less useful if it amplified all DNA equally, but in fact it can
selectively amplify a target sequence of DNA (corresponding to a particular
gene, say) and leave non-target DNA alone. This is done by exploiting a critical
property of DNA polymerase. Specifically, DNA polymerase only adds nucleo-
tides where a portion of single-stranded DNA already has some double-stranded
DNA in place. During replication double-stranded DNA is separated into two
single-strands. DNA polymerase converts each single-strand back to a double-
strand by adding complementary bases to the 3′ end of the newly synthesized
strand in a process known as extension. The fragment of complementary DNA
that binds to the DNA being replicated and allows DNA polymerase to get
started is known as a primer. During cell division these primers are made by
specialized enzymes and allow the entire DNA strand to be replicated. By con-
trast, in PCR, primers are designed to flank a specific region of DNA and amplify
only that DNA fragment. Since DNA polymerase only proceeds from the 3′ end
of the primer and adds new base pairs in the 5′ direction, the flanking primers
are directional. That is, a forward primer is designed to be complementary to a
sequence of DNA that is unique and located upstream (in the 5′ direction) of the
target. It will produce a long strand of complementary DNA starting at the for-
ward primer and extending along the DNA template. A reverse primer is designed
to bind uniquely to the complementary DNA at a point downstream of the tar-
get. The end result is that with each cycle of extension the number of fragments
produced that are flanked by the primers doubles. In short order, the exponen-
tially growing fragments vastly outnumber any other DNA sequence in the reac-
tion (Figure 2.23).
In the 1970s scientists were already using DNA polymerase to replicate DNA.
However, since DNA polymerase can only add bases to single-stranded DNA, they
had to first separate the two strands by heating to the melting or denaturing tem-
perature (typically 95°C) and allowing it to cool to around 65°C, the temperature
EXPERIMENTAL BIOLOGY 45
(A)
5′
First Cycle
etc.
DNA oligonucleotide
primers
region of
double-stranded
chromosomal
DNA to be
amplified
Figure 2.23 Schematic of DNA amplification in a typical polymerase chain reaction (PCR). The fragment flanked by the forward
and reverse primers grows exponentially with the number of amplification cycles, while the other fragments grow linearly. Therefore the
concentration of the fragment flanked by the primers quickly overwhelms all of the other species. (Adapted from, Alberts B, Johnson A,
Lewis J et al. (2008) Molecular Biology of the Cell, 5th ed. Garland Science.)
at which extension can occur (the annealing temperature). The problem was that
each time the sample was raised to the melting temperature, all of the DNA poly-
merase was denatured and needed to be re-added manually. This was very cum-
bersome and greatly limited yield. The answer came from the realization that
thermophilic bacteria thrive in high-temperature environments. Therefore, they
must have a form of DNA polymerase that can withstand high temperatures. One
of the most common forms of DNA polymerase used today is Taq polymerase
which was originally isolated from Thermus aquatiqus, a bacterium that naturally
occurs in hot springs and hydrothermal vents. With this component in place, it
was possible for modern PCR to be developed. In principle, it is simply a matter of
mixing a sample, DNA polymerase, forward and reverse primers, and an abun-
dance of DNA monomer in a tube and alternating quickly between high and low
temperatures. This is typically done in a solid-state device developed just for this
purpose known as a thermal cycler. The reaction will continue and the target is
exponentially amplified until one of the reagents is exhausted.
46 CHAPTER 2: Fundamentals in Cell Biology
for the response. When a step in a pathway is both necessary and sufficient, it is
strong evidence indeed. While this is the current strategy behind biological and
biomedical engineering approaches to mechanotransduction, it should be
acknowledged that this strategy also has weaknesses. For instance, when a new
blocker is discovered, it is impractical to test the blocker against every possible
combination of molecules. As a result, only a few controls are used to assess the
efficacy and specificity of the blocker. Blocking is also imperfect—siRNAs do not
typically completely knock down the end protein expression, rather they suppress
expression to varying degrees, depending on the cell type, design of the fragment,
and method of introducing the siRNA to the cell.
Because of the nature of signaling, it is not entirely accurate to discuss elements in
isolation. If stem cell differentiation by mechanical stimuli can be blocked by
removing compound A, and restored by overexpressing compound A in the
absence of stimuli, we might conclude that compound A, not mechanical stimuli,
is the necessary (and sufficient) ingredient for differentiation (and if we show that
mechanical stimuli up-regulate compound A, we might have the beginning of a
signaling pathway). Such a linear approach is useful in the short term, but may be
an oversimplification. For instance, there is no distinction between a major player
(engine’s spark plug) and a necessary component that is not a major player (igni-
tion key). For engineering purposes, the former is far more useful than the latter,
but often there is no way to tell which of these two is being investigated without far
more comprehensive experiments. It is possible to have two or more necessary
and sufficient pathway components, that do not have any overlap, but describe
the same process. These processes are then typically discovered to be earlier
(upstream) or later in the pathway (downstream).
that are empirical, coupled with statistical testing. Mechanistic approaches are
important, in terms of elucidating regulatory pathways. Repetition and rigorous
protocols are highly valued because of the degree of variability found among sam-
ples. By contrast, engineering approaches tend to be mathematical. They value
long-established physical principles that can be applied analytically or simulated.
To study biomechanics/mechanobiology, an appreciation for both approaches is
desirable. Not all aspects of mechanobiology are easily modeled and not all
aspects of biomechanics can be experimentally measured, but one key to
approaching this sort of research is identification of where the two areas are close
and attempting to bridge the gap between them.
Key Concepts
• Mechanobiology deals with the biological response of • Second messengers transfer information within the
cells to mechanical stimuli. cell quickly, but are broad in effect. Protein signaling
• Structurally, cell components are influenced by cascades are more specific, but act slower.
polymers. Proteins are amino acid polymers. DNA and • Microscopy is used to examine cells visually. Many
RNA are nucleic acid polymers. Polysaccharides are fluorescence techniques exist to help probe cell
sugar polymers. Lipids are fatty acid polymers. behavior.
• The central dogma of cell biology states that proteins • Gels electrophoresis is used to separate molecules
are translated from mRNA, which in turn is for analysis of various biological polymers.
transcribed from DNA. • Biological experiments tend to be reductionist,
• Cell organelles play various roles in cell physiology, mechanistic, and hypothesis-driven. Engineering
but several components are also involved in making approaches are more quantitative but complement
or modifying proteins. biological approaches.
• Receptors are proteins that span the cell membrane
and receive external signals and engage signaling
pathways inside the cell.
Problems
1. Describe a scenario whereby stretching a cell will lead two proteins, X and Y, are engaged at a particular time.
to changes in transcription of some gene, with calcium How would you experimentally determine this, given
signaling as an intermediary. Your response does not that you do not have access to a microscope sensitive
have to be based on a real pathway, but should involve enough to see the proteins themselves?
the material discussed in this chapter.
5. Here is an abstract from a 1997 article in Science (275,
2. You are examining a cell that was stained for a 1308–1311): “The small guanosine triphosphatase
component of the cytoskeleton called actin. You see a (GTPase) Rho is implicated in the formation of stress
single fiber at a location but suspect there may be two fibers and focal adhesions in fibroblasts stimulated
fibers instead. The objective you are using is a 40×, NA by extracellular signals such as lysophosphatidic acid
1.3 oil immersion. What is the closest distance between (LPA). Rho-kinase is activated by Rho and may mediate
two actin fibers at which you can still distinguish them? some biological effects of Rho. Microinjection of the
catalytic domain of Rho-kinase into serum-starved
3. A particular gene is known to be activated upon
Swiss 3T3 cells induced the formation of stress fibers
mechanical stimulation. When you assay the mRNA
and focal adhesions, whereas microinjection of the
concentration 1 hour after applying stimuli, you do
inactive catalytic domain, the Rho-binding domain, or
not see any changes. List possible reasons for this
the pleckstrin-homology domain inhibited the LPA-
observation.
induced formation of stress fibers and focal adhesions.
4. Some signaling pathways will involve multiply-linked Thus, Rho-kinase appears to mediate signals from Rho
proteins that either engage or disengage in response to and to induce the formation of stress fibers and focal
some stimuli. You are interested in examining whether adhesions.”
Problems 51
To help you better understand the abstract, please sufficient, or whether one cannot tell, the down-
note the following: Swiss 3T3 cells are a type of regulation of gene given the findings below. Note that we
fibroblast from mice; microinjection is using a special do not know where in this pathway N acts, but under
microsyringe to inject selected chemicals into cells normal conditions mechanical stimuli will also increase
directly, rather like a shot from a conventional syringe. N phosphorylation.
Do not worry about the pleckstrin-homology domain
mechanical stimuli → calcium increase → gene 1
(which means the local protein sequence is nearly
up-regulation → gene 2 down-regulation
identical to that of a protein called pleckstrin). Do not
worry about what LPA actually does, except for the (a) U nder mechanical stimulation N phosphorylation
function outlined in the abstract. was inhibited by a chemical. Gene 1 failed to up-
(a) B
riefly explain the hypothesis of the authors and regulate, but gene 2 was down-regulated.
how they tested the hypothesis. (b) Without mechanical stimuli, phoshorylated N was
(b) If you had a technique for measuring how stiff a induced by injecting phosphorylated N directly into
cell is, and you compared LPA-treated cells versus cells. Gene 2 was down-regulated.
untreated cells, which would I expect to be stiffer? (c) Under mechanical stimuli, calcium changes were
(c) Why did the authors serum-starve the cells? suppressed by a chemical. N exhibited increased
(d) If you had unlimited resources, what other controls phosphorylation, but gene 2 was not down-
would you add to what the authors did to further regulated.
test their hypothesis? 10. Discuss the following statement. “To produce one
6. If you wanted to examine the expression of a particular molecule of each possible kind of polypeptide chain,
protein, there are several molecular readouts you can 300 amino acids in length, would require more
use. Suppose you predict that in response to stretching atoms than exist in the universe.” Given the size of
a cell for an hour, the protein SSP (stretch-sensitive the universe, do you suppose this statement could
protein) will be more abundant in the cell. possibly be correct? Since counting atoms is a tricky
business, consider the problem from the standpoint
(a) H
ow can you experimentally validate your
of mass. The mass of the observable universe is
prediction?
about 1080 g, give or take an order of magnitude or
(b) S
uppose you did the experiment from (a) and
so. Assuming that the average mass of an amino acid
found that SSP increases. What would you think
is 110 Da, what would be the mass of one molecule
if a colleague told you that she also stretched the
of each possible kind of polypeptide chain 300 amino
cell but the SSP mRNA levels decreased? Does it
acids in length? Is this greater than the mass of the
matter when she assayed the cells?
universe?
7. A gene mutates so that the generated protein has
a single amino acid substitution somewhere in the 11. It is often said that protein complexes are made from
middle whereby an Asp is replaced with a Glu. (You may subunits (that is, individually synthesized proteins)
consult sources to determine what these side chains rather than as one long protein because the former is
actually are.) more likely to give a correct final structure.
(a) W
hat difference in protein structure, if any, do you (a) A
ssuming that the protein synthesis machinery
expect to result from this substitution? That is, do incorporates one incorrect amino acid for
you expect the protein to fold exactly the same each 10,000 it inserts, calculate the fraction of
way? Why or why not? bacterial ribosomes that would be assembled
(b) N
ext you grab the ends of the protein and try to correctly if the proteins were synthesized as
pull it apart. You find that the protein is “stiffer” one large protein versus built from individual
(i.e., harder to extend) with the Glu mutation. Could proteins. For the sake of calculation, assume
this affect the signaling capacity of the protein? If that the ribosome is composed of 50 proteins,
so, how? each 200 amino acids in length, and that the
subunits—correct and incorrect—are assembled
8. We are interested in the interactions of two proteins, with equal likelihood into the complete ribosome.
JP (junctional protein) and JPAP (JP-associating [The probability that a polypeptide will be made
protein). We know these proteins are linked at the cell correctly, Pc, equals the fraction correct for each
boundaries between adjacent, adhering cells. We also operation, fc, raised to a power equal to the
know that JPAP will dissociate (i.e., separate from) JP number of operations, n, Pc = (fc)n. For an error
when the cell is mechanically stimulated. We do not rate of 1/10,000, fc = 0.9999.]
know whether JP leaves the cell boundaries when (b) Is the assumption that correct and incorrect
JPAP dissociates from it. Describe an experiment using subunit assembly is equally likely true? Why or why
fluorescent-conjugated JP and JPAP to determine not? How would a change in that assumption affect
whether JP leaves the cell boundaries in response to the calculation in part (a)?
mechanical stimuli.
12. If a sample of human DNA contains 20% cysteine (C) on
9. For the following experimental results involving protein a molar basis, what are the molar percentages of A, G,
N, determine whether N phosphorylation is necessary, and T?
52 CHAPTER 2: Fundamentals in Cell Biology
13. All small intracellular mediators (second messengers) 17. A typical mammalian cell is about 1000 mm3 in volume.
are water soluble and diffuse freely through the cytosol The concentration of protein within a typical cell is
(indicate true or false, and explain why). 200 mg/ml. Using western blotting, you can detect
10 ng of a specific protein from loading 100 mg of total
14. Cells communicate in ways that resemble human protein. The specific protein you are interested in has a
communication. Decide which of the following forms molecular weight of 100 kDa.
of human communication are analogous to autocrine,
(a) H ow many cells do you need to harvest to be able
paracrine, endocrine, and synaptic signaling by cells and
to load 100 mg of total protein?
briefly say why.
(b) How many copies of a given protein per cell are
(a) A telephone conversation required to detect the western band?
(b) Talking to people at a cocktail party
(c) A radio announcement 18. After treating cells with a chemical mutagen, you isolate
(d) Talking to yourself two mutants. One carries alanine and the other carries
methionine at a site in the protein that normally carries
15. Why do signaling responses that involve changes in valine. After treating these two mutants again with the
proteins already present in the cell occur in milliseconds mutagen, you isolate mutants from each that now carry
to seconds, whereas responses that require changes in threonine at the site of the original valine. Assuming
gene expression require minutes to hours? that all mutations involved single nucleotide changes,
deduce the codons that are used for valine, methionine,
16. You want to amplify DNA between the two stretches of
threonine, and alanine at the affected site. Would you
sequence in the figure. Of the listed primers, choose the
expect to be able to find valine-to-threonine mutations in
pair that will allow you to amplify the DNA by PCR.
one step?
DNA to be amplified
5′-GACCTGTGGAAGC CATACGGGATTGA-3′
3′-CTGGACACCTTCG GTATGCCCTAACT-5′
Primers
Annotated References
Alberts B, Johnson A, Lewis J et al. (2008) Molecular Biology of the 335–350. This is the original article that describes the PCR reaction,
Cell, 5th ed. Garland Science. A comprehensive text on molecular although there are previous publications reporting its use. The tech-
biology that covers in great detail the process of transcription and nique was rapidly adopted.
translation, and the structure and function of cells. This text is a
Periasamy A (2001) Methods in Cellular Imaging. Oxford University
useful reference not only for Chapter 2, but for many parts of the
Press. A good, broad introductory reference to imaging techniques
rest of the book.
used in cell biology.
Lodish H, Berk A, Kaiser CA et al. (2007) Molecular Cell Biology, 6th
Watson JD (2001) The Double Helix: A Personal Account of the Dis-
ed. W.H. Freeman. Another good textbook on molecular biology.
covery of the Structure of DNA. Touchstone. An autobiographical
Actually, there are several texts and you may choose the one that is
explanation of the discovery of the double-helical structure of DNA
most appropriate for your level of interest.
by one of the Nobel Prize–winning team.
Mullis KB & Faloona FA (1987) Specific synthesis of DNA in vitro
via a polymerase-catalyzed chain reaction. Methods Enzymol. 155,
CHAPTER 3
I n the field of cell mechanics, there is more than one way to understand the
mechanical behavior of cells. Our review of basic solid mechanics begins with
rigid-body mechanics and goes on to consider small deformations in solid bodies.
We discuss simple loading configurations, including axial, torsion, and bending
configurations. From there we discuss some aspects of large deformation mechan-
ics, although a complete treatment is beyond the scope of this text.
result of free-body diagram analysis. For now, let us ignore deformation and
focus on problems in which the forces are simple and discrete.
the interior of the column the force resulting from the load is evenly distributed F
through the cross section, assuming the column is uniform. If we move far away
from the end, the force becomes evenly distributed throughout the cross section.
As with pressure, we can imagine a cut-plane through the column. Consider a cut- σ = F/A
plane that is perpendicular to the long axis of the column (Figure 3.8). Because
the force across this area is distributed evenly, we can apply an analogy to pres-
sure and define stress, σ, to be the force per unit area
F
σ = . (3.1)
A
Figure 3.8 Column loaded with a
This stress is considered a normal stress, because the force is acting perpendicu- distributed force F at both ends
larly across the area. Historically, tensile stresses, as illustrated in Figure 3.8, are experiences internal normal stress
assumed to be positive by convention, and compressive stresses are negative. F/A at any cut-plane oriented
perpendicular to the long axis of
Unfortunately for the analogy with pressure, the opposite sign convention is used
the cylinder.
for pressure as stress.
Nota Bene
Strain represents the normalized change in length Tensile stresses are positive. The
of an object to load convention that tensile stresses are
positive comes about because much
Just as we defined stress to be a force scaled by the area it is acting across, we can of the material-testing literature is
also define strain as a change in the length of an object, scaled by its original focused on tensile testing.
length. Consider again the column under tension from Figure 3.8. Let the initial
length of the column be denoted as L. When the load is applied, the column Nota Bene
extends by a small amount (ΔL) and the total length of the column is now L + ΔL
(Figure 3.9). Here, the strain is defined as Distinguishing material and
structural behaviors. The column is
a good example of why scaling can
∆L be so critical. Imagine that the column
ε= . (3.2)
L was twice as long initially: the amount
by which it extends under the same
load would be twice as large.
A strain associated with normal stress (that is, along the same axis as the applied However, the stress in the material
force) is called normal strain. Because strain is defined as the ratio of two would not have changed. This larger
lengths, it is inherently dimensionless. It is sometimes expressed as either per- displacement at the tip would be due
cent strain (ΔL/L × 100) or microstrain (ΔL/L × 1,000,000). Also note that the entirely to the shape of the column, in
same sign convention for stress holds for strain (tensile positive, compressive this case its length. This is an example
of the influence of geometry on
negative).
mechanical behavior and is
sometimes called the structural
behavior to distinguish it from the
The stress–strain plot for a material reveals information influence of the material properties.
about its stiffness
We have now defined stress and strain in our simple uniaxial column example.
Before we go on to more complicated examples, let us see how they relate to each ΔL
other to help us understand the behavior of a material. Our column is loaded with
axial tension and we record the deformation of the tip as a function of the load we
apply. Assume that the column is uniform in its material properties (homogene-
ous), that it behaves as an elastic material, and that its material behavior is the ΔL
ε=—
L
same in all directions (isotropic). Although the two quantities we directly measure L
are the tip displacement and the force, we calculate the stress (σ = F/A) and strain
(ε = ΔL/L) and plot them versus each other (Figure 3.10).
For a broad range of sizes and shapes of columns, this plot will exhibit the same
overall shape. For many materials, this plot is a simple line. Such materials are
known as linear (or linear elastic) materials. The slope of this line is known as the Figure 3.9 Same column from Figure
3.8 with the same load (load not
Young’s modulus of the material. Typically this proportionality constant is given
shown) will lengthen along the long
by E in the linear relationship σ = Eε, known as Hooke’s law (physics students axis of the cylinder. The ratio of the
may recognize its similarity to the one-dimensional Hooke’s law F = kx from change in length (Δ L) to the original
modeling a spring). As the load increases, a stress is eventually reached, known length (L) of the column is the strain.
58 CHAPTER 3: Solid Mechanics Primer
Nota Bene Figure 3.10 For the loading configuration (A), a stress–strain relationship can be
obtained (B). For small strains, the relationship can be a straight line; this is the linear region
Who was Young? Young’s modulus with the slope representing the stiffness, or elastic modulus. As the load increases, the
is named after the eighteenth- material will pass through the proportional limit (extent of linear region), yield point (where
century British physician Thomas deformation is not recovered upon unloading), ultimate stress (maximum stress), and failure.
Young. He conducted research in
light and vision, circulation,
mechanics, and even helped
translate some of the Rosetta Stone. as the proportional limit where it no longer behaves linearly. The stress at failure
He extended the work of Hooke, who is the ultimate stress or strength of the material. Independence from geometry is
characterized elastic behavior in a major benefit of using stress and strain rather than force and displacement.
terms of force and displacement
(F = kX) into normalized stress and
strain, whereby, the proportionality Stress and pressure are not the same thing, because stress
constant k—known as the stiffness has directionality
and a function of both material and
geometry—becomes the Young’s One key distinction between stress and pressure has to do with orientation. As we
modulus, E, and is a true material mentioned previously, the pressure at any point in a fluid does not depend on the
property. It is also sometimes orientation. For a given point inside a fluid, the vertical and horizontal pressures
referred to as the elastic modulus or
at that location are always the same. By contrast, stress depends strongly on orien-
modulus of elasticity. However, it is
only one of several possible elastic tation. In our axially loaded column example, the stress in the vertical direction is
moduli, such as the bulk modulus or F/A, yet the stress in the horizontal direction is zero, because there is no load in
shear modulus.
R A E EA
the horizontal direction. The reason behind this difference is the ability of the
molecules in a fluid to flow under load and reorient the internal force distribution.
Molecules in a solid, on the other hand, are more strongly bound to their neigh-
bors and are not free to redistribute under load. In mathematical terms, pressure
is a scalar quantity, depending (as do forces, displacements, stresses, and strains)
on the location within the material where the quantity is measured. Because
forces and displacements are vectors (these quantities are described by a magni-
tude and a direction), directionality is built into them. Stress and strain, which F
depend on the direction of forces and displacements, therefore exhibit direction-
ality-like force and displacement. However, stress and strain are even more com- τ = F/A
plicated, because they depend not only on the direction of forces and
displacements, but also on the direction of the area (or displacement) to which
they apply. To specify a component of stress, we must know the direction of the
forces relative to the orientation of the surface (the surface direction or orienta-
tion is given by the vector normal to its local area). Stress and strain are mathe-
matical quantities known as tensors.
Figure 3.11 Shear stress occurs
when the loading force is
Shear stress describes stress when forces and areas are perpendicular to the area across
perpendicular to each other which the stress acts.
Let us return to the column-loading example. Initially, the forces in the cut-plane
were perpendicular to the plane, resulting in normal stresses. What sort of
stress results if we load the column perpendicular to the long axis (that is, parallel
to the plane)? Consider again the imaginary cut-plane some distance away from
the region of load application. As before, the force is distributed evenly through
the cross section, but now the small distributed forces are lined up parallel to the
plane rather than perpendicular to it (Figure 3.11). We term this type of stress
shear stress,
F
τ = . (3.3)
A
In general, the cut-plane through the column can be in any orientation we desire.
If we select a plane that is oriented obliquely to the axis of the column, the distrib- Figure 3.12 A general cut-plane
uted forces across the surface will be a mixture of parallel and perpendicularly through a column with a normal
oriented forces. Therefore, in general stress is a mixture of normal stress and shear load will experience both shear and
normal stresses (since the vector
stress (Figure 3.12).
components can be decomposed
into components parallel and
perpendicular to the surface). This is
Shear strain measures deformation resulting from true even if the load is a mixture of
shear stress shearing and normal forces.
define the linear relationship between shear stress and shear strain with the shear
modulus, G,
τ = Gγ = Gθ (3.4)
for small θ.
δ = θR, (3.5)
M = τ 2πR 2t , (3.7)
L
MECHANICS OF DEFORMABLE BODIES 61
J =
∫ R dA = ∫ 2πR dR,
2 3
(3.14)
A 0
which is a measure of a given cross-sectional geometry to resist torsion. This
allows us to express Equation 3.13 in a more compact form,
GJθ
M= . (3.15)
L
Notice that this analysis only works for cylindrical-shaped rods. If a rod is not a cyl-
inder, if it has a square or rectangular cross section, our kinematic assumption
breaks down. Plane sections will not deform by simply sliding over one another, but
they will also become deformed (no longer remain planar). This deformation along
the axis of the rod is known as warping and occurs even for small deformations.
+ ...
62 CHAPTER 3: Solid Mechanics Primer
R J G GJ
x
z
MECHANICS OF DEFORMABLE BODIES 63
y–z-planes rotate, but they do not deform to lose their planar shape. This assump-
tion is sometimes called the “plane sections remain plane” assumption. Next, we
assume that the plane sections only rotate and do not slide relative to each other. S
In other words, we are going to assume that we can ignore the effects of shear.
Finally, we assume that as the beam deforms, the y–z-planes rotate in such a way
as to remain perpendicular to any imaginary line in the beam that was originally
oriented in the x-direction. These three assumptions serve to define fully the
deformation of every point in the beam. Furthermore, the deformed state is fully R
defined by knowing the displacement of the central axis of the beam only. The
beam becomes a simple one-dimensional problem of determining the displace-
ment in the y-direction, which we denote w(x). Next we need to determine the S – ∆S
strain in the beam as a function of the displacement w(x). ∆R
S
To accomplish this, we need to determine the amount of stretch or strain that lines
initially oriented parallel to the x-axis experience as the beam deforms. Consider a S + ∆S
very small portion of the deformed beam (Figure 3.17). Further, consider three
lines on the beam; the top surface, the bottom surface, and the central axis of the Figure 3.17 Small section of the
beam S long, ± ΔR thick.
beam. If the section is small enough, the deformed shape of each of these lines can
be approximated as sections of a circle (this is known as the osculating circle and
can be uniquely defined at every point in a curve as the best-fitting circle at that
point). Let the radius of the deformed central axis be R, the top surface R − ΔR, and
the bottom surface R + ΔR. Next, assume that the initial length of a line along the
central axis in our section of beam is S. In the deformed configuration, the central
axis is neither lengthened nor shortened. In other words, the strain at the central
axis is zero and is therefore often referred to as the neutral axis. Because the ratio of
circumference to radius is constant for similar circular sections,
S + ∆S S
= . (3.16)
R + ∆R R
∆S ∆R
= . (3.17)
S R
Notice that the quantity ΔS/S is simply the change in length of one of our imaginary
lines normalized by its initial length. That is exactly our definition of strain. Also
notice that ΔR is the distance to the imaginary line in the y-direction measured
from the neutral axis. If we position the origin of our coordinate system so that it
lies on the neutral axis, we have a simple expression for the strain in the beam,
y
ε( y ) = . (3.18)
R
So we have used the term “central axis” without really defining it. This was an
intentional oversight to simplify the explanation. It would have been more appro-
priate to refer to the neutral axis. For cross sections that are symmetric about the
central axis, the two coincide, but this need not be the case in general.
Now, notice that the radius, R, is not a constant; it is a function of the displace-
ment w(x). To formulate equations that we can solve, we must make this depend-
ency explicit. To do this, notice that the quantity 1/R is also the local curvature of
the beam typically denoted as κ. In calculus, we learned that the local slope of a
curve is given by its first derivative, but the local curvature is given by the second
derivative. Therefore strain is the product of the second derivative of the displace-
ment, and the distance from the neutral axis becomes
y d 2w
ε (y) = = yκ = y 2 . (3.19)
R dx
64 CHAPTER 3: Solid Mechanics Primer
t /2 t /2
M=
∫ dM = ∫ yσ ( y)hdy.
− t /2 − t /2
(3.20)
h
MECHANICS OF DEFORMABLE BODIES 65
σ ( y ) = Eε ( y ). (3.21)
Remarkably, with the kinematics, equilibrium, and constitutive equations, we
have completed the theoretical development for this problem. All that remains is
to combine what we know. Substituting the stress–strain relationship (Equation
3.21) into the equilibrium equation (Equation 3.20) yields
t /2
M=E
∫ yε( y)hdy.
− t /2
(3.22)
t /2
d 2w
M=E
∫
− t /2
y2
dx 2
hdy . (3.23)
Because w does not depend on y, it can come out of the integral. The end result is
a governing relationship between the moment and the curvature,
t /2
d 2w d 2w
M=E
∫
− t /2
y 2hdy
dx 2
= EI
dx 2
, (3.24)
where
t /2
I=
∫ y hdy.
− t /2
2 (3.25)
t /2 t /2
Iy =
∫ y hdy
− t /2
2
and I x =
∫ x hdx .
− t /2
2 (3.26)
The term that relates curvature to moment, EI, is a measure of the combined
effects of material stiffness and geometry to resist bending. To reflect this, it is
known as the flexural rigidity.
66 CHAPTER 3: Solid Mechanics Primer
R I E EI
d 2w M ( x )
= . (3.27)
dx 2 EI
Notice that M is a function of x and E and I are constants. Given a beam of length
L, the moment produced by the force on the end is simply the force multiplied by
the moment arm, M(x) = (x – L)F. That takes care of the right side, but what about
the boundary conditions on the left side? Clearly, the displacement must be zero,
w(0) = 0. But it is also clamped, so the slope must also be zero, dw(0)/dx = 0. We
need to solve
d 2w ( x − L)F dw(0)
= w(0)= = 0. (3.28)
dx 2 EI dx
M = (X – L)F
MECHANICS OF DEFORMABLE BODIES 67
F x3 x2
w= − L + C1 x + C 2 . (3.29)
EI 6 2
Using our boundary conditions:
w(0) = 0 ⇒ C 2 = 0
dw(0) (3.30)
= 0 ⇒ C1 = 0.
dx
Therefore, the final solution is
F x3 x2
w= − L . (3.31)
EI 6 2
d 2w M ( x ) Fw
= =− . (3.32)
dx 2 EI EI
This is a little more challenging to solve than Equation 3.28. We need a function
that returns itself when differentiated twice. This suggests
d 2w Fw
=− = −C1k 2 sin(kx ). (3.34)
dx 2 EI
w(0) = w(L) = 0
68 CHAPTER 3: Solid Mechanics Primer
F
k=± .
EI (3.36)
F F
C1 sin L = 0 or L = nπ . (3.37)
EI EI
Each of the solutions corresponding to different values of the integer, n, is a differ-
n=1 ent mode of bending (Figure 3.22).
Notice that none of the solutions to Equation 3.36 place a constraint on C1. The
n=2 etc. bending mode that corresponds to the lowest possible force is the n = 1 mode. The
force that corresponds to this mode is
Figure 3.22 Deformation patterns
corresponding to the first two π 2 EI
bending modes of the beam. Fb = , (3.38)
L2
which is known as the Euler buckling load. If the applied force is < Fb, the beam
remains straight and y(x) = 0 is the solution. It can also be shown that this solution
is stable. In other words, if the beam is deflected slightly away from y(x) = 0, it will
return to the y(x) = 0 configuration. On the other hand, if F > Fb, the y(x) = 0 solu-
tion is not stable, and the beam will buckle, taking on a deformed configuration.
This simple analysis to determine the buckling load can be used to find the EI of a
biopolymer by measuring the longest length it is able to attain.
We aim to provide you some familiarity with these tools, but the presentation is
far from complete.
As we described before, there are three parts to a problem in continuum mechan- Sx y Sx
ics: kinematics, equilibrium, and constitutive behavior. So, let us take each one
in turn.
Sx x
Equilibrium implies conditions on stress
S xz
In our discussion of stress and equilibrium, let us first introduce a notation system
for stress. In our examples, we defined stress conceptually to be the distributed y
force per unit area through a given test plane. In the example, the location of the
test plane was obvious. Now we want to create a general definition and notation x
system associated with our x-, y-, z-coordinate system. To specify the orientation
of the test plane, we will make use of the normal vector to the plane. z
Next consider the resultant force denoted by Sx (Figure 3.23). The x refers to Figure 3.23 Forces on an imaginary
the cut-plane we selected, which has a surface normal aligned in the x-direc- cut-plane in the y–z-plane.
tion. Like any vector, Sx can be broken into its three components, Sx x , Sx y , Sx z .
So, for this face, we have three potential stresses, one associated with each of
the components of Sx. To express the components of stress, we need a double
subscript notation system with the first referring to the direction of the cut-
plane normal, and the second referring to the direction of the internal force
acting through that plane. Now we have everything we need to define the stress
components in this coordinate system. For a cut-plane oriented perpendicular
to the x-direction,
Sx x (3.39)
σ xx = lim .
A→0 A
Sx y Sx (3.40)
τ xy = lim and τ xz = lim z .
A→0 A A→0 A
And for the y-cut-plane
Sy x Sy Sy (3.41)
τ yx = lim , σ yy = lim y and τ yz = lim z
A→0 A A → 0 A A → 0 A
Sz x Sz Sz (3.42)
τ zx = lim , τ zy = lim y and σ zz = lim z .
A→0 A A→0 A A→0 A
Now that we have a consistent notation system, let us see what equilibrium can
tell us about stress. Remember that the equilibrium condition is a condition on
forces. Specifically, for a nonaccelerating body, the forces must sum to zero.
Imagine a small piece of material within a general solid body, and consider the
forces on the infinitesimal element as shown (Figure 3.24). We assume that the
dimensions of the element are dx, dy, and dz respectively. Each face has one nor-
mal force and two shear forces. However, because the element is vanishingly
small, we can approximate the force on the faces far from the origin as a function
of the force on the closer face and its derivative. Taking the first term in a Taylor
series,
dSx x ( x )
Sx x ( x + dx ) = Sx x ( x ) + dx . (3.43)
dx
70 CHAPTER 3: Solid Mechanics Primer
y S yz
S yx
x
S yy
z
Now we simply sum the forces in the x-, y-, and z-directions in turn:
dSx x dSy x
∑f x
= 0 ⇒ − Sx x + Sx x +
dx
dx + −Sy x + Sy x +
dy
dy + −Sz x + Sz x +
dSz x
dz
dz
dSx y dSy y
∑f y
= 0 ⇒ − Sx y + Sx y +
dx
dx + −Sy y + Sy y +
dy
dy + −Sz y + Sz y +
dSz y
dz
dz
dSx z dSy z
∑f z
= 0 ⇒ − Sx z + Sx z +
dx
dx + −Sy z + Sy z +
dy
dy + −Sz z + Sz z +
dSz z
dz
dz .
(3.44)
Notice that the first two terms in each quantity in parentheses cancel. Next, we
can divide each row by the infinitesimal volume (dx dy dz) and simplify
Now, notice that in each term we have a differential area. In each case, this differen-
tial area does not depend on the derivative in the numerator. Therefore, we can write
MECHANICS OF DEFORMABLE BODIES 71
d Sx x d Sy x d Sz x
+ + =0
dx dydz dy dxdz dz dydz
d S x y d Sy y d Sz y
+ + =0 (3.46)
dx dydz dy dxdz dz dydz
d Sx z d Sy z d Sz z
+ + = 0.
dx dydz dy dxdz dz dydz
The differential area for each term in parentheses is the area normal to the respec-
tive force vector component. Therefore, each of these terms is simply our defini-
tion of stress. The equilibrium equations then take the following remarkably
simple form:
dσ xx dσ yx dσ zx
+ + =0
dx dy dz
dσ xy dσ yy dσ zy
+ + =0 (3.47)
dx dy dz
dσ xz dσ yz dσ zz
+ + = 0.
dx dy dz
∑M x = 0 ⇒ 2 Sy z + 2 Sz y = 0 Sx z
Sx x Sx x
or Sy z = Sz y . In terms of stress
Sx z
∫∫ σ zy dxdy =
∫∫ σ yz dxdy or σ zy = σ yz .
Szx
Likewise for the y- and z-axes, we obtain z
σ xy = σ yx and σ zx = σ xz . Szz
x
This important property of stress is that it must be Figure 3.25 The forces on a small two-dimensional
symmetric for any nonaccelerating body. Because element.
dv
ε yy = ,
dy
(3.50)
dw
ε zz = .
dz
Now we consider the shear strain. Perhaps the most logical thing to do would be
to define them as similar to normal strains. We could define
dv
ε xy = . (3.51)
dx
However, there is a problem with this approach. Strain defined in this way is not
symmetric, because in general,
dv du
≠ . (3.52)
dx dy
It will simplify things greatly to define strain as symmetric, and this can be easily
achieved by taking our definition of shear strain to be
1 dv du
ε xy = + , (3.53)
2 dx dy
which preserves the symmetry condition, because εxy = εyx, εxz = εzx, and εyz = εzy.
To clearly distinguish normal and shear strains, the symbol γ is sometimes used
for the components of the shear strain. The notation γ refers to the engineering
MECHANICS OF DEFORMABLE BODIES 73
strain defined in our pure shear and torsion examples. It differs from the contin-
uum strain ε by a factor of two, that is, γ xy = 2ε xy , γ xz = 2ε xz , and γ yz = 2ε yz .
1
ε xx = σ xx − νσ yy − νσ zz , γ xy = (1/G )τ xy
E
1
ε yy = −νσ xx + σ yy − νσ zz , γ xz = (1/G )τ xz (3.54)
E
1
ε zz = −νσ xx − νσ yy + σ zz , γ yz = (1/G )τ yz .
E
Notice that there are three material constants (E, ν, G) in Equation 3.54, though
only two of these are independent constants. The shear modulus can be expressed
in terms of Young’s modulus and Poisson’s ratio (the proof of this is left to you),
G = E/2(1 + ν).
Therefore,
1 2(1 + ν )
ε xx = σ xx − νσ yy − νσ zz , τ xy = γ xy
E E
1 2(1 + ν )
ε yy = −νσ xx + σ yy − νσ zz , τ xz = γ xz (3.55)
E E
1 2(1 + ν )
ε zz = −νσ xx − νσ yy + σ zz , τ yz = γ yz .
E E
74 CHAPTER 3: Solid Mechanics Primer
These equations can be inverted to yield the more traditional form of stress as a
function of strain:
E E
σ xx = (1 − ν )ε xx + νε yy + νε zz τ xy = γ xy
( )(
1 + ν 1 − 2ν ) 2(1 + ν)
E E
σ yy = νε xx + (1 − ν )ε yy + νε zz τ xz = γ xz (3.56)
(1 + ν )(1 − 2ν ) 2(1 + ν )
E E
σ zz = νε xx + νε yy + (1 − ν )ε zz τ yz = γ yz .
(1 + ν )(1 − 2ν ) 2(1 + ν )
σ xx ε xx
σ yy ε yy
σ zz ε zz
σ = and ε = . (3.57)
τ xy γ xy
τ xz γ xz
τ yz γ yz
σ xx
σ yy
σ
zz E
=
τ xy (1 + ν ) (1 − 2ν )
τ xz
τ yz
(1 − ν ) ν ν 0 0 0
ν (1 − ν ) ν 0 0 0 ε xx
ν ν (1 − ν ) 0 0 0 ε yy
ε zz
0 0 0
(1 − 2ν ) 0
0 or σ = Cε
2 γ xy
γ
0 0 0 0
(1 − 2ν ) 0
xz
2 γ yz
0 0 0 0 0
(1 − 2ν )
2
(3.58)
MECHANICS OF DEFORMABLE BODIES 75
and
1 −ν −ν
0 0 0
E E E
−ν 1 −ν
ε xx 0 0 0 σ xx
ε yy E E E σ
−ν −ν 1 yy
ε zz 0 0 0 σ zz or ε = Dσ
= E E E
γ xy 2(1 + ν ) τ xy
γ 0 0 0 0 0 τ
E
xz
xz
γ yz 2(1 + ν ) τ yz
0 0 0 0 0
E
0 2(1 + ν )
0 0 0 0
E
(3.59)
σ xx λ + 2µ λ λ 0 0 0 ε xx
σ yy λ λ + 2µ λ 0 0 0 ε yy
σ zz λ λ λ + 2µ 0 0 0 ε zz
=
τ xy 0 0 0 2µ 0 0 ε xy
τ xz 0 0 0 0 2µ 0 ε xz
τ yz 0 0 0 0 0 2 µ ε yz ,
where
Eν
λ =
(1 + ν ) (1 − 2ν )
and
E
µ = ,
2(1 + ν )
It is important to note that the vector notation we use is a another specified by another set of vectors x′, y′, and z′ (x,
notational system only. We arrange the components of y, z, and x′, y′, and z′ are known as basis vectors). Also,
stress and strain in a vector to make them easy to manipu- define the angle between any of these vectors to be θxx′,
late. However, stress and strain are not mathematically θxy′, etc. A rotation matrix Q can be defined such that
vectors. As you remember, a vector is a quantity that has
both magnitude and direction, like force, deformation, or
cos (θ xx ′ ) cos (θ xy ′ ) cos (θ xz ′ )
velocity. An implication of this is that a vector maintains
its magnitude and direction from one coordinate system Q = cos (θ yx ′ ) cos (θ yy ′ ) cos (θ yz ′ ) .
cos (θ zy ′ )
to another. This implies that the components of a vector cos (θ zx ′ ) cos (θ zz ′ )
must behave in a very specific way in terms of how they
relate to one another when expressed in different coordi-
nate systems. Assume that we have two (orthogonal) coor- Then any vector can be expressed in the new coordinate
dinate systems, one specified by the vectors x, y, and z and system by multiplying the vector expressed in the old
76 CHAPTER 3: Solid Mechanics Primer
coordinate system by Q. If P is a vector in the x,y,z-sys- In this text we will not be using tensors or tensor mathe-
tem then P′ in the x′,y′,z′-system is given by P′ = QP. matics. However, some of the quantities are in fact ten-
Obeying this coordinate transformation rule is the for- sors, like stress and strain. In Section 3.3 we define some
mal definition of a vector. However, our stress and strain new quantities and refer to them by their names from
“vectors” do not obey this rule. In fact, stress and strain continuum mechanics, so that you can follow up with
are more general mathematical quantities known as ten- additional reading. However, we avoid the use of any
sors. The appropriate transformation rule for tensors of tensor mathematics, and for our purposes they can be
this type requires the terms to be written in matrix nota- manipulated like matrices.
tion. For example,
σ σ xy σ xz
xx
σ = σ yx σ yy σ yz and σ′ = Q Tσ Q.
σ zx σ zy σ zz
σ xx σ xy σ xz ε xx ε xy ε xz
σ = σ σ yy σ yz and ε = ε yx ε yy ε yz
. (3.60)
yx
σ zx σ zy σ zz ε zx ε zy ε zz
Known as matrix notation, this approach is useful for mathematical manipula-
tions of stress and strain and proving new relationships. Bear in mind that the
same amount of information is contained in each notation (here, there are nine
components each for stress and strain, but three of them are redundant because
of symmetry). To illustrate one advantage of using this notation, we next derive
what are called the principal stresses and strains.
Av = ϕ v , (3.61)
where v is the eigenvector, and φ is the eigenvalue. In general, for a 3 × 3 matrix,
there are three eigenvectors/values. We also know from linear algebra that the
principal stresses are solutions of the so-called characteristic equation
as the principal directions v1, v2, and v3. We call these principal stresses and
strains because they are invariant with respect to the coordinate system. That is,
regardless of how you arrange the coordinate axes, the principal stresses and
strains for a given deformation will always be the same. Similarly, the principal
directions will not change. If you were to grab a square-shaped object at oppo-
site corners and pull, the material will deform into a kite-shape. One of the prin-
cipal directions is along the long diagonal of the kite. This remains true regardless
of which coordinate system is used.
The principal directions possess some particularly useful properties. They are
mutually perpendicular or orthogonal to each other. This means that they can
serve as a coordinate system, and it is possible to rotate our coordinate system
from our original x-, y-, and z-axes into v1, v2, and v3. When we do this, the stress
and strain matrices also change to match the new coordinate system, with the
result that the off-diagonal components of the stress matrix vanish, and the diago-
nal components are the principal stresses. In other words, using the principal
directions as our coordinate system, σ becomes
σ 1 0 0
σ =0 σ2 0 . (3.63)
0 0 σ 3
This remarkable property means that no matter how complex the state of stress,
there is always a coordinate system in which all the shear stresses vanish and only
normal stresses remain. Indeed, specifying the principal stresses and principal
directions is sufficient to fully define the state of stress. The same holds true for
strain.
2% and 1%. What directions are these strains in? To In this new coordinate system
answer this we need the principal directions. We simply
substitute back into the characteristic equation 0.02 0.0
ε= . (3.72)
ε v a = ε av a 0.0 0.01
for each εa in turn. We find that the eigenvectors (1, 1) An illustration of the principal strains can be obtained
(corresponding to ε1) and (1, –1) (corresponding to ε2). with the strain ellipse (Figure 3.27). It also shows that
Note that when you are solving for the eigenvectors, any the maximum and minimum normal strains occur in the
eigenvector can be multiplied by a scalar, and the result- principal directions.
ing vector is also an eigenvector. So, you will either need
to constrain the length or to set a component to an arbi- y normal strain
trary value (we use ones for clarity). (Alternatively, you
can construct eigenvectors of unit length.)
principal direction
Finally, notice that the eigenvectors are oriented 45
degrees from our original coordinate system. We can
define a new coordinate system rotated by 45 degrees
counterclockwise such that
x
1 1 1 1
x ′= , y ′= . (3.70)
2 1 2 −1
deformed
A
becomes a small line, da, in the deformed configuration. We can relate these two
small elements by approximating the deformation with a Taylor series expansion
and neglecting higher-order terms with the matrix equation:
da ≈ FdA , (3.73)
∂a Nota Bene
∂a x ∂a x
x Vernacular of large deformation
∂Ax ∂Ay ∂Az mechanics. Much of the notation
∂a ∂ a y ∂a y ∂a y (3.74)
and terminology of large deformation
F= = . mechanics was introduced by Clifford
∂A ∂Ax ∂Ay ∂Az Truesdell and Walter Noll (who had
∂a z ∂a z ∂a z been Truesdell’s graduate student) in
the classic 1965 text The Non-Linear
∂Ax ∂Ay ∂Az
Field Theories of Mechanics. More
recently, Noll has suggested that the
term deformation gradient, which
F is an incredibly useful quantity and forms the basis of almost all of large defor- they introduced, may be misleading
mation kinematics. Here we develop a few particularly relevant applications. and should be replaced with
transplacement gradient, which was
adopted by Truesdell and Noll in
Stretch is another geometrical measure of deformation subsequent work. However,
deformation gradient is now so
Although strain is a natural way to quantify deformation, it is far from the only deeply embedded in the literature
one. Indeed several other deformation measures exist depending on the applica- that it seems unlikely to replaced.
tion. Another popular measure of deformation is stretch. Conceptually, stretch is
simple. If something is pulled twice as long, we say it has a stretch of two. If it does
not change, it has a stretch of one. Stretch is quantified by the stretch ratio, λ, the
ratio of the deformed length divided by the original length, in contrast to strain,
which is the change in length over original length. In our infinitesimal deforma-
tion notation
∆L L + ∆L
ε= ,λ = . (3.75)
L L
|da|
λ= . (3.76)
|dA |
80 CHAPTER 3: Solid Mechanics Primer
Unlike strain, which can be normal or shear, the stretch ratio is always a measure
of the normal deformation of a prescribed differential line element. The stretch
ratio is related to the extension ratio, given by λ − 1, which is similar, but not identi-
cal, to our concept of strain. In particular, stretch is measured along a line segment
of interest and follows the line segment as it rotates and translates. In Example 3.5,
the test line that originally goes from (0, 0) to (1, 0) is deformed to a new line seg-
ment from (0, 0) to (1.015, 0.005). We decomposed the strain of this segment into
the εxx and εxy components. For stretch, we do not perform this decomposition.
Instead, we work with the entire vector and create F to describe how the line seg-
ment changes. For this particular deformation, the stretch ends up being just
above 1.015, so there is only a small difference between stretch and strain meas-
ures. However, for larger deformations, the two measures can be divergent. If
instead of (1.015, 0.005) the test line ends up at (1.5, 0.5) then εxx is 0.5 and εxy = 0.5,
but stretch is 1.58 and the extension ratio is 0.58. So although many of the math-
ematical manipulations will be similar (such as determining principal directions),
these quantities are distinct.
Stretch ratios can be defined for any particular line segment. One might choose
dA or da to be aligned with one of the coordinate axes. One particularly useful
approach is to use line segments that are aligned with the principal directions of
stretch. Neglecting rigid-body motions, the deformation in the principal direc-
tions is entirely normal (that is, there is no shear). The stretches are known as the
principal stretches (λ1, λ2, λ3). If the coordinate system is aligned with the princi-
pal directions, F takes a particular form with the principal stretches on the diago-
nals and zeros elsewhere,
λ 0 0
∂a
1
F= = 0 λ2 0 . (3.77)
∂A
0 0 λ3
The principal stretches can capture the full deformation, and formulations of
mechanical behavior in terms of principal stretches are particularly common in
large deformation constitutive modeling because of their simplicity.
E=
2
(
1 T
F F−I . ) (3.78)
Nota Bene
Right Cauchy–Green deformation The term FTF is a special quantity denoted as C = FTF.
tensor. In continuum mechanics, C
is called the right Cauchy–Green It is special because it is related to local stretch. In terms of our small test lines, it
deformation tensor. It is called “right” can be shown that da2 = (dA)C(dA). It also has interesting properties: its trace
not because it is correct, but because (sum of the diagonals) is equal to the sum of the squares of the principal stretches,
it is related to the deformation matrix and its determinant is equal to the product of the squares of the principal stretches.
of a polar decomposition when the On an intuitive level, the Green–Lagrange strain is a measure of how much the
deformation matrix is on the right of
local stretches, C, differ from one. Written out in detail in terms of deformation
the rotation matrix.
the Green–Lagrange strain is
LARGE DEFORMATION MECHANICS 81
du 1 du 2 2
dw
2
+ + +
dv
E xx =
dx 2 dx dx dx
1 du dw
2 2 2
dv dv
E yy = + + +
dy 2 dy dy dy
dw 1 du 2 2
dw .
2
+ + +
dv
Ezz =
dz 2 dz dz dz (3.79)
1 du dv 1 du du dv dv dw dw
E xy = E yx + + + +
2 dy dx 2 dx dy dx dy dx dy
1 du dw 1 du du dv dv dw dw
E xz = Ezx + + + +
2 dz dx 2 dx dz dx dz dx dz
1 dv dw 1 du du dv dv dw dw
E yz = E yx + + + + .
2 dz dy 2 dy dz dy dz dy dz
Notice that each component of strain contains a first term corresponding to small
deformation strains and an additional higher-order term that captures nonlinear
effects of large deformation. With linear homogeneous deformations, the Green–
Lagrange strain and the small strain measures are the same, because these higher-
order terms vanish.
With F in hand, we can compute the Green–Lagrange strains is aligned 45 degrees to the original coordinate
strain, E, system. In that direction the strain is 400% and in the
perpendicular direction it would be 150%. If we were to
2.75 1.25 rotate our coordinate system 45 degrees counterclock-
E=
2
(
1 T
F F−I = )
1.25
.
2.75
(3.82) wise, the new E would be
4.0 0 .
To find the principal strains, we need to find the eigen- E= (3.83)
values of E. Using the same approach as in Example 3.5, 0 1.5
we obtain principal strains of 1.5 and 4.0 and principal
directions of (1, –1) and (1, 1). The largest of the two Notice that the shear strains are zero.
F = RU, (3.84)
Nota Bene where R is what is an orthonormal matrix. That is, it has special properties such
F is also called the deformation that it is orthogonal, meaning that all of its rows or columns are pairwise orthogo-
gradient tensor. It can be nal (dot products are one if self-multiplied, zero otherwise) and it is normal, or its
manipulated to some extent like a determinant is 1. The properties of R are exactly those required for it to be a rota-
matrix, and as far as we cover in this tion matrix like Q discussed in the Advanced Materials on coordinate rotation
textbook, matrix manipulations are (Section 3.2). R represents a rigid-body rotation of the object. This part of F does
enough to get the results we need.
However, like stress and strain, F
not lead to distortions or strain, but is an integral part of measurements made if
depends on the coordinate system, the principal stretch directions are not known ahead of time. The remaining term,
because it describes a physical U, captures the distortion of the object.
process. Certain rules apply in
manipulating F, including the rotation
U sits on the right side of the decomposition of F. You can alternatively decom-
rules in the Advanced Materials on pose F into VR, where the stretch component V sits on the left side. U and V are
coordinate rotation (p. 75). not the same, because the order of matrix multiplication is important (it is not
commutative), although the rotations are the same.
In Figure 3.30, we see that a square can be rotated by 45 degrees counterclock-
wise first, then stretched to expand one side to twice the original length.
Alternatively, we can stretch the square to twice the original length, and then
rotate it by 45 degrees counterclockwise. The rotations are identical. The material
stretch is the same, but the stretch tensors are different, because in the rotate-first
case, the square is being stretched along a 45-degree line, whereas the stretch-first
case stretches the square along the horizontal axis.
Note that if we include translation in this analysis, we will have a general mathe-
matical transformation for any mechanical deformation—a general deformation
can be expressed as a single translation, a single rotation, and a single distortion.
undeformed V
R
STRUCTURAL ELEMENTS ARE DEFINED BY THEIR SHAPE AND LOADING MODE 83
1 0 0.0
U= . (3.88) 0.5 1.0 1.5 2.0 2.5
0 2
–0.5
This means that we can now solve for R, given that
F = RU,
–1.0
cosθ − sinθ 1 0
1/ 2 3
F= = . (3.89)
− 3 / 2 1 sinθ cosθ 0 2 Figure 3.31 Two boxes in the original configuration
(black) and deformed configuration (blue).
Key Concepts
• The forces and moments acting on a body that is not • Cytoskeletal proteins are relatively stiff in axial
accelerating must sum to zero. The forces and tension, but flexible in torsion and bending.
moments acting across any closed boundary must also • A column will collapse when it is exposed to
sum to zero. This gives rise to the free-body diagram. axial compression that exceeds the Euler
• For deformable bodies, stress is a normalized buckling load.
measure of force analogous to pressure. • In large deformation mechanics, the strains are not
• Strain is a normalized deformation. assumed to be infinitesimal.
• Stress and strain can be normal or shear. • The deformation gradient is the fundamental large
• Continuum mechanics gives rise to three fundamental deformation kinematic measure, and captures a
relationships: kinematics, constitutive behavior, and body’s rotation and deformation.
equilibrium. • The Green–Lagrange strain is a common large
• Euler–Bernoulli beam theory describes small deformation strain measure, although others exist
transverse deformations of slender bodies and as well.
accounts only for lengthwise stresses.
Problems
1. Let V1 be the volume of a hollow cylinder of outer 2. A cylinder is a composite of two materials, with shear
radius Ro and inner radius Ri (do not use the thin- moduli G1 and G2. The cylinder is composed of material
shell approximation) consisting of material with shear with modulus G1 from 0 ≤ R ≤ Ri and G2 from Ri ≤ R ≤ Ro.
modulus G. This cylinder is fixed at one end; a moment Derive the net effective shear modulus G for the
of M is applied to the other end, resulting in an angular composite material for torsion.
twist of θ. Let V2 be the volume of a solid cylinder of
radius ro consisting of material with shear modulus 3. If an object of arbitrary shape is completely submerged
G. Under the same loading conditions as the hollow in water without touching any other surface (such as
cylinder, the angular twist is θ/2. Derive an expression the bottom floor), is there any pressure on the surface
for ri in terms of the other parameters provided. of the object? Is there net pressure on the object (that
Problems 85
is, if you integrate the pressure over the entire surface E = 10 GPa (about the stiffness of wood), how much
of the object, is there a net force in any direction)? strain do you expect in the column when the table
Justify your response. is assembled under the minimal stress condition
derived in (a)? Neglect the weight of the column
4. You are given two materials, one with elastic modulus itself.
E1 called Mat A and the other with elastic modulus
E2 = 10E1 called Mat B. The materials are shaped into (c) The strain in part (b) is quite small, so you may
cubes of 1 m3 and loaded with 1 kN of force, evenly neglect it for this part. Someone sitting at the
across the top surface (and supported by an infinitely table is tilting their chair back and resting on the
rigid bottom surface). Estimate the strain (in terms of E1) edge of the tabletop. You model this as a block
for the following: of mass 70 kg with an arrangement as shown in
the illustration below. Assume that there is no
(a) The cube is made entirely of Mat A. friction between the block and tabletop. If the
shear modulus of the column is 5 GPa, find the
(b) The cube is made entirely of Mat B.
total displacement of the tabletop in the lateral
(c) The cube is made of 50% Mat A and 50% (horizontal) direction. Neglect the deformation of
Mat B, with the different materials layered the tabletop. Does the orientation of the column
horizontally (interface is parallel to the load). matter?
8. Assuming a model of a cell as a thin-walled sphere 17. Solve the beam equation for a beam that is free to
enclosing a pressurized fluid, derive the relationship rotate at the supports and is loaded with a single point
between the stress in the membrane and the cytoplasm load at mid-span.
pressure.
18. We solved for the buckling load assuming that the base
9. In calculating GJ for cytoskeletal proteins (Table 3.2), of the beam is free to rotate. What is the buckling load
we ignored the fact that microtubules are hollow. for the clamped configuration? This can be done with
Recalculate G and GJ for a hollow microtubule, symmetry arguments. Do not solve the beam equation.
assuming that the inner radius is 11 nm. How large
was our error? Was our order-of-magnitude estimate 19. Solve the beam equation for a beam supporting a
acceptable? distributed pressure of Q N/m that is free to rotate at
the supports.
10. The other error we made in Table 3.2 was that we
could estimate the shear modulus G with the Young’s 20. Often we are interested in the energy stored in a
modulus E. Assume an effective Poisson’s ratio of 0.3. structure because of its deformation or strain
How large was the error? Was our order-of-magnitude energy. One approach is to integrate the strain
estimate acceptable? What is the result if you assume a energy density (strain energy per unit volume) over
Poisson’s ratio of 0.0 or 0.5? the structure,
Annotated References
Fung YC (1977) A First Course in Continuum Mechanics. Prentice- Truesdell C & Noll W (1919) The Non-Linear Field Theories of Mechan-
Hall. An outstanding introduction to infinitesimal fluid and solid con- ics. Springer. A seminal book that first presented a standard unified
tinuum mechanics that includes a very lucid introduction to tensor set of concepts and notations for large deformation mechanics.
analysis. Currently out of print, but the third edition incorporates corrections
from the late Professor Truesdell’s personal copy. The history and
Malvern LE (1969) Introduction to the Mechanics of a Continuous
rationale for updating the terminology is outlined on Noll’s (cur-
Medium. Prentice-Hall. A comprehensive and classic text on large
rently at Carnegie Mellon University) website (http://www.math.
deformation mechanics. Mathematically intense and rigorous.
cmu.edu/~ wn0g/noll).
Timoshenko SP & Goodier JN (1934) Theory of Elasticity. McGraw-
Hill. A classic text in small deformation elasticity. Includes easy-to-
follow treatments of beams, plates, torsion, etc.
This page intentionally left blank
to match pagination of print book
CHAPTER 4
F = ρgAh , (4.1)
90 ChApTER 4: Fluid Mechanics Primer
Nota Bene
height = h
Incompressible fluids. In deriving density = ρ
Equation 4.1, we assumed that the
density of the water was constant, in
other words the density of the fluid is
not altered even when subjected to
large pressures. Such fluids are said to area = A
be incompressible. Most fluids
(including water) are generally
pressure = p
assumed to be incompressible under
most relevant circumstances. However,
there are cases in which pressures are
extremely large and compressibility
may need to be accounted for, even where g is the acceleration due to gravity. In this case, F acts in the direction nor-
for fluids that are normally considered mal to the bottom of the container and in a downward direction. As we discussed
incompressible. For example, die in Section 3.2, pressure is the force per unit area in the direction normal to the
casting involves the injection of molten surface on which it is acting. Given this definition, we can express Equation 4.1 in
metal into a die at extremely high terms of pressure,
pressures. This pressure causes the
liquid metal to compress, and this
compressibility must be taken into p = F /A = ρgh. (4.2)
account for in the design of the die.
Such a pressure is referred to as hydrostatic pressure, because it is the pressure
exerted by a fluid at equilibrium, resulting here from gravitational forces.
Equation 4.2 states that the hydrostatic pressure acting on the bottom of the
container is proportional to the height of the fluid. More generally, the hydro-
static pressure experienced by a cell or organism in contact with a fluid is
dependent on the height of the fluid above it. Cells and organisms may be
exposed to hydrostatic pressures that vary quite a bit. For instance, aquatic
organisms may experience different hydrostatic pressures as they swim at dif-
ferent depths. If one were to hang upside down by one’s feet (Figure 4.2), vascu-
lar cells in the feet and skull would experience a decrease and increase in
hydrostatic pressure, respectively.
h
cells
reservoir
Figure 4.4 The fluid element adjacent to the container side is subjected to forces
from the glass and the surrounding fluid. In this case, the glass exerts a force that
resists the net force imparted on the element by the surrounding forces which keeps it from
accelerating. If the glass were suddenly removed, the fluid would flow, and the column of
fluid would collapse into a puddle.
92 ChApTER 4: Fluid Mechanics Primer
with time. Between the plates, the fluid velocity varies linearly in the y-direc-
tion, with zero velocity at y = 0, and a velocity of V0 at y = h. This characteriza-
tion of the fluid velocity is called a flow profile. In the case of a linear flow
profile, a constant shear stress is applied to the upper plate to maintain this
profile. For a Newtonian fluid, this shear stress is
∂u
τ =µ , (4.3)
∂y
where τ is the shear stress, u is the fluid velocity and y is a direction perpendicular
to the direction of u. The derivative of u with respect to y is called the shear rate or
velocity gradient. The constant μ represents the dynamic viscosity, and has units
kg/ms. We will see later in Section 4.3 that Equation 4.3 comes from a more gen-
eral set of constitutive relations for Newtonian fluids. Nota Bene
Given Equation 4.3, we can determine the magnitude of the shear stress in our No-slip condition. In our example
of the linear flow profile, the fluid
chamber as contacting the plates themselves has
the same velocity as the plates. This
V0
τ =µ . (4.4) condition is called the no-slip
h condition, which stipulates that fluid
immediately adjacent to a solid
Upon inspection of Equation 4.4, we see that how fast the fluid deforms in surface cannot slip along that
response to a given shear stress is determined by the fluid viscosity. A highly vis- surface, and will travel with the same
velocity as the surface itself. An
cous fluid such as honey will undergo a much slower velocity for the same-sized everyday example of the no-slip
gap, h, in response to a given shear stress, compared with a fluid with a lower condition is that dust on fan blades
viscosity, such as water. Alternatively, it takes a lot more shear stress to move the tends to stay in place no matter how
upper plate with a desired velocity, V0, for a fluid like honey. fast the fan spins.
where Vin is the average velocity of all the fluid entering, Ain is the cross-
sectional area normal to the flow entry, and Vout and Aout are similarly defined Nota Bene
for fluid exiting (Figure 4.7). The product (V)(A) is referred to as the volumetric
Streams and rivers. An everyday
flow rate. The reason why volumetric flow rate must be conserved is that for example of conservation of mass is
incompressible fluids, where the density is assumed constant, volume is the water speed of streams and
directly proportional to mass. Conservation of mass is one of the most useful rivers. In deep regions of a channel
concepts in fluid mechanics. It explains why a pipe, a narrowing of the diame- (which have a larger cross-sectional
ter in a pipe leads to increased velocity of the fluid flowing within. As we will area), the water tends to flow slowly,
see, this relation will be critical in allowing us to solve the Navier–Stokes equa- while shallow regions tend to be
moving quickly.
tions in Section 4.3.
94 ChApTER 4: Fluid Mechanics Primer
turbulent flow
NEWTONIAN FLUIDS 95
where ρ and μ are the fluid density and dynamic viscosity, respectively, V is what Nota Bene
is called a characteristic velocity, and L is some length scale. For flow in tubes and
Transition to turbulence. Though
pipes, L is generally the diameter, and V is generally the average velocity (although
Re = 1 is often considered a transition
the radius and peak velocity may also be used). point at which flows begin to exhibit
The Reynolds number measures the relative magnitudes of inertial versus viscous turbulence, for steady pipe flows, Re
can rise as high as 2000 or so before
forces in a given flow and is widely considered one of the most important quanti-
such a transition occurs. The reason
ties in fluid mechanics. In general, a Reynolds number greater than one indicates for this is that in pipe flows, the fluid
that inertial forces, in other words how much momentum the fluid has, dominates is highly constrained, moving in a
the flow. Such inertial forces tend to drive mixing, and high Reynolds numbers are single dimension. In this instance, the
associated with turbulent flow. Reynolds numbers of less than 1 indicate that vis- changes in inertial terms tend to be
cous forces dominate. Note that at Re = 1, we have a balance between viscous and very small and mixing is more a
inertial terms, and this is sometimes referred to as the transition region, where the function of surface roughness.
flow can start developing instabilities, but is not yet fully turbulent.
∂P ∂τ
Pdydz − P + dx dydz + τ dy dxdz − τ dxdz = 0 (4.7)
∂x ∂y
∂τ
τ + — dy
∂y
dz
∂p
p + — dx
p dy ∂x
dx τ
96 ChApTER 4: Fluid Mechanics Primer
Dividing Equation 4.8 by the volume of the fluid element (dxdydz) yields
∂P ∂τ
− + = 0. (4.9)
∂x ∂y
Next, we examine the spatial dependence of the pressure and shear gradient in
the above expression. Consider the dependence of the shear gradient, ∂τ/∂y, on
x, y, and z. First, because the flow is fully developed and not changing in the
x-direction, shear stress is also not changing in the x-direction. Because we
assume no flow in the z-direction, this implies that the shear gradient can only be
a function of y,
∂τ
= f ( y ). (4.10)
∂y
Now consider the pressure gradient. Because the flow velocity is assumed to be
zero in the y- and z-directions, this implies that pressure is not a function of y and
z, and the pressure gradient can only be a function of x,
∂P
= g ( x ). (4.11)
∂x
However, Equation 4.9 states that the difference of Equations 4.10 and 4.11 must
equal zero,
− f ( y ) + g ( x ) = 0. (4.12)
The condition in Equation 4.12 can only be met if f ( y) and g(x) are both equal to
some constant, or
f ( y ) = g ( x ) = constant. (4.13)
To determine this constant, one can solve for ∂τ/∂y in Equation 4.9 and inte-
grate with respect to y to obtain
∂P
y + C1 = τ . (4.14)
∂x
Substituting in Equation 4.3 for the shear stress in the right-hand side of
Equation 4.14, we can express this relation as a function of u as
∂P ∂u
y + C1 = µ , (4.15)
∂x ∂y
where u is the velocity in the x-direction. Dividing Equation 4.15 by viscosity and
integrating with respect to y once more, we obtain the velocity profile
1 ∂P 2
y + C1 y + C 2 = u( y ), (4.16)
2 µ ∂x
boundary is exactly equal to the velocity of the solid boundary. The top and bot- Nota Bene
tom plates are stationary and u( y = 0) = 0, and u( y = h) = 0. This gives us the
Bernoulli’s equation. Consider a
final flow profile,
steady and laminar flow in which we
1 ∂P 2 inject dye, leaving a line of dye that is
( y − hy ) = u( y ). (4.17) tangent everywhere to the velocity of
2 µ ∂x the particle. Formally, we call this line
a streamline. As a particle moves
Equation 4.17 describes a parabolic velocity profile, which is commonly encoun- along a streamline, it will speed up or
tered in a variety of situations involving fluid flow. We will encounter this profile slow down due to the different forces
acting on it, such as those due to
again in our discussion of flow chambers in Section 6.3. In addition, it is straight- changes in fluid pressure and from
forward to show—using a similar approach—that within a pipe of circular cross gravity. If the flow is inviscid, in other
section, a parabolic profile is also obtained. words the fluid is assumed to have
zero viscosity, then we can relate
these parameters using a relationship
Many biological fluids can exhibit non-Newtonian behavior called Bernoulli’s equation,
In the last section we found that a Newtonian fluid will exhibit a parabolic veloc- 1 2
ity profile when subjected to pressure-driven flow between parallel plates. This P+ ρV + ρg z = constant
2
solution depended on the incorporation of the constitutive equation for
Newtonian fluids, Equation 4.3, which related shear stress and shear rate (or In the above expression, P is the local
pressure, ρ is the fluid density, V is
velocity gradient) through a constant, viscosity. Not all fluids follow the linear
the local fluid velocity, g is the
relation between shear rate (or velocity gradient) and shear stress that character- gravitational constant, and z is the
ize Newtonian fluids. The fluid here is considered non-Newtonian. For many bio- local elevation. We can see that along
logical fluid flows, viscosity is not constant, but a function of other parameters a streamline, for a given height, as
such as shear rate. pressure increases, velocity
decreases. A great way to experience
One non-Newtonian fluid is the Bingham plastic or Bingham fluid, which has this phenomenon is to hold two
been used to model the flow of biological fluids such as blood and mucus that pieces of paper close together and
exhibit solidlike behavior at low stresses but fluidlike behavior at large stresses. gently blow between them. They will
For this fluid there is a critical value of shear stress, τ0, below which the shear rate tend to move together due to the
is zero, such that increased velocity of air, which
results in a decrease in pressure
∂u between the two pieces of paper.
= 0 for τ < τ 0
∂y
(4.18)
∂u τ − τ 0
= for τ ≥ τ 0 .
∂y µ
If the shear stress is below τ0, the fluid does not deform or flow. To demonstrate
this, consider our parallel plate example in the previous section, except we substi-
tute the Newtonian fluid with a Bingham fluid. Assume that the peak level of shear
stress computed using Equations 4.13 and 4.17 is less than τ0. Here, ∂u/∂y = 0 and
u must be constant. However, we know that the velocity is zero at the plate sur-
faces by the no-slip condition and u = 0 everywhere. If this level of shear is greater
than or equal to τ0, then the fluid can flow. A classic example of the application of
the Bingham fluid model is to model paint sticking to a wall—if the shear forces
arising from gravity are not sufficiently large, the fluid stays motionless until it
dries out.
Another non-Newtonian fluid is a power-law fluid, which can be modeled using
the relationship
α
∂u
τ = β , (4.19)
∂y
where β is a constant viscous factor (not viscosity) and α is a constant scaling
exponent. Equation 4.19 can be recast as
∂u
τ = µ eff , (4.20)
∂y
98 ChApTER 4: Fluid Mechanics Primer
Bingham plastic
(C)
power law
shear thinning
where
α −1
∂u
µ eff = β . (4.21)
∂y
In this formulation, the effective viscosity of the fluid is shear rate–dependent, and
that dependency changes whether the scaling exponent is greater or less than
one. When α is greater than one, the fluid undergoes shear-thickening, in other
words the effective viscosity increases as the shear rate increases. There are not
too many everyday examples of shear-thickening solutions, but a watery mixture
of cornstarch can act as one—so much so that even though you can stir a vat of it,
you can also walk across its surface if you step quickly enough.
If the exponent is less than one, the fluid undergoes shear-thinning, where the vis-
cosity decreases as the shear rate increases. An example of a shear-thinning fluid is
ketchup, which shows its non-Newtonian behavior when one is trying to get it out of
a glass bottle. Initially, getting the ketchup to flow is difficult, but once it starts flow-
ing, it tends to flow easily. So, stirring or shaking the bottle of ketchup will shear the
fluid and temporarily decrease its viscosity and allow it to flow faster. Blood is a well-
known biological fluid that tends to exhibit shear-thinning behavior. One reason
that blood exhibits shear-thinning is that with flow the red blood cells tend to
become oriented with one another, thereby reducing viscocity. When the scaling
exponent is equal to one, we recover Newtonian fluid behavior (Figure 4.10).
phenomena such as air flow around an airplane wing, water flow in pipes, and
currents in the ocean. With regard to cellular mechanobiology, they have been
critical in shedding light on the types of mechanical forces that cells are exposed
to in vivo. For example, they have been used to predict fluid shear stresses and
pressure profiles for cells subjected to blood or interstitial fluid flow. In addition,
they are useful for modeling flow profiles for cells subjected to fluid flow in vitro,
such as within flow chambers and bioreactors.
u( x , y , z , t )
u( x , y , z , t ) = v( x , y , z , t ) . (4.22)
w( x , y , z , t )
We now consider a small rectangular-shaped fluid element moving with velocity
u(x, y, z, t). The fluid element is assumed to be aligned with the coordinate axes,
centered at point (x, y, z), and assumed to have dimensions Δx, Δy, and Δz, in the
x-, y-, and z-directions, respectively. The external forces are assumed to arise from
one of two sources; stresses imparted by the surrounding fluid acting on the faces
of the control volume, and body forces (such as that due to gravity).
First, consider the forces that act on the faces of the control volume. For simplicity,
we will begin with the forces in the x-direction. Because the fluid element has six
faces, there are six forces we must consider, one on each face: Fx(x − dx/2, y, z),
Fx(x + dx/2, y, z), Fx(x, y − dy/2, z), Fx(x, y + dy/2, z), Fx(x, y, z − dz/2), and Fx(x, y,
z + dz/2) (Figure 4.11). These forces can be computed as the product of the stress
acting on the face and the area of the face over which it is acting. In the x-direction,
∆x ∆x
Fx x − , y , z = σ xx x − , y , z ∆y ∆z
2 2
∆x ∆x
Fx x + , y , z = σ xx x + , y , z ∆y ∆z
2 2
∆y ∆y
Fx x , y − , z = σ xy x , y − , z ∆x ∆z
2 2
∆y ∆y (4.23)
Fx x , y + , z = σ xy x , y + , z ∆x ∆z
2 2
∆z ∆z
Fx x , y , z − = σ xz x , y , z − ∆x ∆y
2 2
∆z ∆z
Fx x , y , z + = σ xz x , y , z + ∆x ∆y .
2 2
100 ChApTER 4: Fluid Mechanics Primer
∆x ∂σ xx ∆x
Fx x − , y , z ≈ σ xx ( x , y , z ) − ∆y ∆z
2 ∂y 2
∆x ∂σ xx ∆x
Fx x + , y , z ≈ σ xx ( x , y , z ) + ∆y ∆z
2 ∂y 2
∆y ∂σ xy ∆y
Fx x , y − , z ≈ σ xy ( x , y , z ) − ∆x ∆z
2 ∂y 2
(4.24)
∆y ∂σ xy ∆y
Fx x , y + , z ≈ σ xy ( x , y , z ) + ∆x ∆z
2 ∂y 2
∆z ∂σ xz ∆z
Fx x , y , z − ≈ σ xz ( x , y , z ) − ∆x ∆y
2 ∂y 2
∆z ∂σ xz ∆z
Fx x , y , z + ≈ σ xz ( x , y , z ) + ∆x ∆y .
2 ∂y 2
∆x ∆x ∆y
Fxext = − Fx x − , y , z + Fx x + , y , z − Fx x , y − , z
2 2 2
(4.25)
∆y ∆z ∆z
+ Fx x , y + , z − Fx x , y , z − + Fx x , y , z +
2 2 2
dz dx
dx
Fx ⎛x + —, y, z⎞ Fx ⎛x, y + —, z⎞
⎝ 2 ⎠
⎝ 2 ⎠
dx
dy
z dz
Fx ⎛x, y, z – —⎞
⎝ 2⎠
y
x
THe NaVieR–StoKes EQuations 101
and following the substitution of the expressions in Equation 4.24 into 4.25,
∂σ ∂σ yx ∂σ zx
Fxext = xx + + ∆x ∆y ∆z. (4.26)
∂ x ∂y ∂z
The volume is also subject to a body force. If fx is the body force in the x-direction
per unit mass, and ρ is the mass density of the fluid, then the total body force is the
product of fx and the mass of the fluid in the control volume,
The partial derivatives ∂x/∂t, ∂y/∂t, and ∂z/∂t give the instantaneous change in
position of the fluid element with respect to time in the x-, y-, and z-directions,
respectively. Because the fluid element is moving with velocity vector u, then
these expressions are equivalent to the velocity of the fluid,
∂x ∂y ∂z
u= , v= , and w = , (4.29)
∂t ∂t ∂t
With the external forces and fluid element acceleration in hand, we now apply
Newton’s second law. The mass of the element is
m = ρ∆x ∆y ∆z. (4.31)
Setting the sum of forces in Equations 4.26 and 4.27 equal to the product of the
mass in Equation 4.31 and acceleration in Equation 4.30 we find
∂σ xx ∂σ yx ∂σ zx
∂x + ∂y + ∂z ∆x ∆y ∆z + f xρ∆x ∆y ∆z
(4.32)
∂v ∂v ∂v ∂v
= ρ∆x ∆y ∆z +u +v +w ,
∂t ∂x ∂y ∂z
102 ChApTER 4: Fluid Mechanics Primer
which simplifies to
∂σ xx ∂σ yx ∂σ zx ∂v ∂v ∂v ∂v
∂x + ∂y + ∂z + ρf x = ρ ∂t + u ∂x + v ∂y + w ∂z (4.33)
when we divide out the volume of the control volume, ΔxΔyΔz. Repeating these
steps for the forces and accelerations in the y- and z-directions,
∂σ xy ∂σ yy ∂σ zy ∂v ∂v ∂v ∂v
∂x + ∂y + ∂z + ρf y = ρ ∂t + u ∂x + v ∂y + w ∂z (4.34)
∂σ xz ∂σ yz ∂σ zz ∂w ∂w ∂w ∂w
∂x + ∂y + ∂z + ρf z = ρ ∂t + u ∂x + v ∂y + w ∂z . (4.35)
Equations 4.33, 4.34, and 4.35 are called Navier’s equations, named after Claude-
Louis Navier.
∂u ∂v
σ xy = σ yx = µ + (4.36)
∂y ∂x
∂v ∂w
σ yz = σ zy = µ + (4.37)
∂z ∂y
∂u ∂w
σ xz = σ zx = µ +
∂z ∂x
(4.38)
2 ∂u ∂v ∂w ∂u
σ xx = − P − µ + + + 2µ (4.39)
3 ∂x ∂y ∂z ∂x
2 ∂u ∂v ∂w ∂v
σ yy = − P − µ + + + 2µ (4.40)
3 ∂x ∂y ∂z ∂y
2 ∂u ∂v ∂w ∂w
σ zz = − P − µ + + + 2µ . (4.41)
3 ∂x ∂y ∂z ∂z
The above six equations, together with Navier’s equations, give nine equations in
total. They also introduce another unknown, pressure, making 10 unknowns in
total. We still require one more equation to make the system solvable. Because we
RHEOLOGICAL ANALYSIS 103
have used the equilibrium and constitutive relations, you may not be surprised Nota Bene
that what is missing is a statement of kinematics. Specifically, this is known as the
continuity equation and is a mathematical statement of the conservation of mass. Continuity equation. A simple
though somewhat unrigorous
As we discussed previously, for incompressible fluids, the flow rate into a fixed
analysis can be used to demonstrate
volume must be equal to the flow rate out of it, the relation between Equations 4.43
and 4.44. Consider a small, imaginary
Vin Ain = Vout Aout . (4.42) cubical volume immersed in a flow
field. The volume is aligned with the
If Ain = Aout = A, then Equation 4.42 can be rewritten as coordinate system and with
dimensions Δx, Δy, and Δz in the x-,
(∆V ) A = 0, (4.43) y-, and z-directions. The volume is
assumed to be fixed in space, with
fluid entering it at some velocity, and
where ΔV = Vout − Vin. A differential form of Equation 4.43 is leaving it at a different velocity. Let
Δu, Δv, and Δw be the changes in fluid
∂u ∂v ∂w velocity in the x-, y-, and z-directions
+ + = 0. (4.44) that occur within the volume. From
∂x ∂y ∂z Equation 4.42, we know that
∆u∆y ∆z + ∆v ∆x ∆z + ∆w ∆x ∆y = 0.
Equation 4.44 is called the differential form of the continuity equation for incom-
pressible fluids. (4.45)
Dividing Equation 4.45 by the volume
of ΔxΔyΔz, and letting the volume go
Navier–Stokes equations: putting it all together to zero, we obtain Equation 4.44.
With Navier’s equations (three total), the constitutive relations proposed by Stokes
(six total), and the continuity equation, we have 10 equations and 10 unknowns
(u, v, w, P, and the six independent components of stress). Although this is a solv-
able system of equations, it can be greatly simplified by substituting in the rela-
tions for stress in Equations 4.36–4.41 into the stress derivatives in Equations
4.33–4.35. Taking the term in parentheses in the left-hand side of Equation 4.33,
we can substitute in Equations 4.36, 4.38, 4.39, and after some manipulation and
invoking Equation 4.44, we arrive at the following expression
∂P ∂ 2u ∂ 2u ∂ 2u ∂u ∂u ∂u ∂u
− + µ 2 + 2 + 2 + Pf x = P +u +v + w . (4.46)
∂x ∂ x ∂ y ∂ z ∂ t ∂ x ∂ y ∂z
∂P ∂ 2v ∂ 2v ∂ 2v ∂v ∂v ∂v ∂v
− + µ 2 + 2 + 2 + Pf y = P +u +v +w (4.47)
∂y ∂x ∂y ∂z ∂t ∂x ∂y ∂z
∂P ∂ 2w ∂ 2w ∂ 2w ∂w ∂w ∂w ∂w
− + µ 2 + 2 + 2 + Pf z = P +u +v +w . (4.48)
∂z ∂x ∂y ∂z ∂t ∂x ∂y ∂z
Equations 4.47 and 4.48 are the Navier–Stokes equations. Together with the conti-
nuity equation, they form a set of four equations to solve for the four unknowns
u, v, w, and P.
sustain shear or recover from deformations. Within the body, tissues and organs,
as well as the cells residing within them, may be composed of both fluid and
solidlike materials. The cytoplasm of a given cell may contain a solidlike cytoskel-
etal network immersed in an aqueous environment in which numerous proteins
are densely dispersed. In this case, one can see why cells would be expected to
exhibit both solid and fluidlike mechanical behavior.
Rheology is a scientific discipline that can be broadly characterized as the study of
materials that have some capacity to flow, but which cannot be adequately
described using classical fluid mechanics. Rheology is considered a distinct
branch of continuum mechanics that bridges solid and fluid mechanics. The need
for such a discipline can be better understood by revisiting our examples of non-
Newtonian fluids; power-law fluids and Bingham plastics. Although both models
are able to capture some aspects of nonlinear fluid behaviors, they do not contain
any fundamental solidlike behaviors. No matter how you adjust the exponent in a
power-law fluid, it will always continuously deform under shear. A Bingham plas-
tic is able to resist shear stress if it is below a critical level, but here the material is
completely rigid and does not exhibit elastic behavior. Rheological methods allow
us to better understand materials that exhibit solid and fluidlike behavior, such as
cells (Figure 4.12). A subset of rheology is the study of viscoelasticity, in which
one seeks to decompose mechanical behavior into purely elastic and purely vis-
cous components. We will discuss some basic rheological approaches for investi-
gating viscoelastic substances.
σ = σ 0 cos(ωt ), (4.49)
RHEOLOGICAL ANALYSIS 105
where σ0 is the magnitude of the stress, and ω is the frequency of oscillatory load-
ing (in radians per unit time). If the material is linearly elastic, it would deform
proportionally to the stress as
ε = A cos(ω t ), (4.50)
where A is a constant. Unlike an elastic material, the stress in a purely viscous
material is not dependent on strain, but on strain rate. This is similar in concept
to a Newtonian fluid, in which shear stress is proportional to shear rate. For the
oscillatory stress described in Equation 4.49, a purely viscous material would
deform such that the time derivative of strain is proportional to stress,
ε = B sin(ω t ). (4.51)
Notice that the strain profile of a viscous material given by Equation 4.51 can be
rewritten as
ε = ε 0 cos(ω t − δ ) (4.53)
is the strain of the material, then the phase shift is δ radians. The parameter δ is
called the phase lag (Figure 4.13). The phase lag is useful because it gives a single
δ
106 ChApTER 4: Fluid Mechanics Primer
value that characterizes the degree to which the mechanical behavior is elastic
relative to that which is viscous. If the phase lag is close to zero, then the material
is behaving primarily like an elastic material. If the phase lag is close to π/2, then
the material is behaving primarily like a viscous material.
Next, consider the case in which we subject a material to oscillatory loading and
obtain the strain profile given in Equation 4.53. We can decompose this profile
into purely in-phase and out-of-phase components. First, we rewrite Equation
4.53 as a sum of cosine and sine functions using the identity
ε 0′ ε 0 cos(δ ), (4.56)
and
ε 0″ = −ε 0 sin(δ ). (4.57)
Equations 4.55, 4.56, and 4.57 demonstrate that the mechanical response given by
Equation 4.53 can be decomposed into two components, one that is in phase (the
first term, with cos(ωt)) with the driving stress, and one that is out of phase (the
second term, with sin(ωt)). The relative magnitudes of the in-phase and out-of-
phase terms are related to the magnitude of the phase lag.
We have intentionally defined the complex stress such that the real part of σ* is the
stress σ applied to the material as in Equation 4.59,
Re {σ *} = σ 0 cos(ω t ). (4.64)
where again, we have intentionally defined the complex strain such that the real
part of e* is the observed strain in Equation 4.54,
Re {ε *} = ε 0 cos(ω t − δ ). (4.66)
With the complex stress and complex strain defined, we can now define a complex
modulus as the ratio of these two quantities,
σ*
E* = . (4.67)
ε*
Substituting expressions for the complex stress and strain in Equations 4.63 and
4.65 into Equation 4.67, and following some simplification, we arrive at a compact
expression for the complex modulus,
σ 0e iω t σ 0e iω t σ
E* = i (ω t −δ )
= = 0 e iδ . (4.68)
ε 0e ε 0e iω t e −iδ ε0
Because E * is a complex number, we can define E′ as the real part of E * and E″ as
the imaginary part, in other words
E * = E ′ + iE ″ (4.69)
with
σ0
E′ = cos(δ ) (4.70)
ε0
108 ChApTER 4: Fluid Mechanics Primer
and
σ0
E″ = sin(δ ). (4.71)
ε0
E ′ is known as the elastic modulus or storage modulus. The storage modulus E ′ is
associated with the in-phase component of resistance to stress. This becomes
apparent by substituting δ = 0 into Equations 4.69, 4.70, and 4.71. Here, the com-
plex modulus is equivalent to the storage modulus, E * = E ′. Because in-phase
deformation is a characteristic of elastic materials, the storage modulus can be
considered a measure of the elastic behavior of the material. E ″ is referred to as
the damping or loss modulus, and is associated with the out-of-phase resistance
to stress. For instance, for δ = π/2, E * = iE ″. In this case, the magnitude of the
complex modulus is equal to the loss modulus. Because this phase lag is associ-
ated with viscous materials, the loss modulus can be considered a measure of the
viscous behavior of the material.
Nota Bene Before concluding this section, it is worthwhile to note that in our development,
Historical roots of complex we designated the complex modulus as E *, suggesting a modulus analogous to
modulus. The complex modulus was Young’s modulus. There is a corresponding complex shear modulus G *, which is
introduced by the German physicist actually more common in the literature because dynamic rheological measure-
Carl Gauss in the early 1800s—the ments are often made under shear. For the rest of the chapter, we will use the com-
same Gauss for whom the SI unit of
plex shear modulus G * and associated complex shear stress τ * and complex shear
magnetic strength was named.
strain γ * in our developments.
Just as we defined a complex modulus, we can also Comparing the expressions for Equations 4.75 and 4.76, it
define a complex viscosity. More specifically, the com- becomes apparent that Equation 4.75 can be rewritten as
plex viscosity can be defined as the ratio of the complex
shear stress and complex shear strain rate as G * G + iG ′ iG ′ G ″
µ* = = = + . (4.77)
iω iω ω ω
τ*
µ* = , (4.72) Letting µ* = µ′ + iµ″ , then
γ*
G″
µ′ = (4.78)
where ω
τ * = τ 0e iω t (4.73) and
G′
and µ″ = , (4.79)
ω
dγ ∗ d where μ′ is the dynamic viscosity, and μ″ is the “out-of-
γ∗ = = γ 0e i (ω t −δ ) = γ 0iωe i (ω t −δ ) . (4.74)
dt dt phase” viscosity that yields the change of the storage
modulus with respect to frequency. Upon inspection of
Equations 4.78, and 4.79, one might see that the complex
Substituting Equations 4.73 and 4.74 into 4.72, we viscosity is closely related to the complex shear modu-
obtain lus. In fact, μ′ and μ″ are the loss modulus and storage
modulus normalized by frequency, respectively. This
τ 0e iω t τ fact allows one to use these parameters to better under-
µ* = = 0 e iδ . (4.75)
γ 0iωe i (ω t −δ ) γ 0iω stand how the storage and loss moduli change with fre-
quency. For instance, assume μ′ and μ″ are relatively
constant over some frequency range, and μ′ > μ″.
Similar to Equation 4.68, the complex shear modulus Increasing the loading frequency within this range
can be computed as would result in greater increases in G ″ relative to that in
G ′, and the material would be expected to exhibit more
τ 0 iδ
G* = e . (4.76) viscous behavior and less elastic behavior with increas-
γ0 ing frequency.
α
ω πα
G * (ω ) = G0 (1 + iξ )Γ(1 − α )cos + i µω (4.80)
Φ 2
where
πα
ξ = tan (4.81)
2
and
tributor to the loss modulus. The scaling exponent determines not only how G ′
10 3 and G ″ change with frequency but the relative magnitudes of each modulus. If
G″ we assume that μ is small, we can see that as α approaches zero, ξ approaches
10 2 zero as well, and G * approaches G0, which gives the elastic component of the
10 –2 10 0 10 2 response. However, as α approaches one, ξ increases without bound, indicating
f (Hz) that the imaginary term will dominate and the material will exhibit strongly vis-
Figure 4.14 Storage and loss moduli cous effects.
as a function of frequency in cells Experiments in a variety of cell types and using various techniques for mechanical
subjected to oscillatory mechanical
loading have demonstrated that cells exhibit a scaling exponent of around 0.2–0.3
loading. Because the results are plotted
on a log–log scale, the slope of G′ gives (Figure 4.14). This power-law dependence is typical of a class of materials called
the scaling exponent α, which is soft glassy materials, which includes materials such as emulsions, slurries, and
approximately 0.2 in this case. (Adapted pastes. Soft glassy materials are characterized by their possession of some degree
from, Fabry et al., [2011] Phys. Rev. Lett. of disorder in which the discrete elements of which they are composed are entan-
87, 148102.) gled or aggregated via weak interactions. Therefore cells have often been described
in the literature as soft glassy materials.
F a Lb µ cv d = 1. (4.84)
The “1” on the right side of Equation 4.84 does not refer to its numerical value, but
to the fact that the product has no units. To do this, we decompose each of these
parameters in Equation 4.84 into their basic units and then pick the exponents so
that they all cancel out. If we replace the parameters in Equation 4.84 with their
basic units, then
a c d
kg ⋅ m b kg m
2 (m ) = 1. (4.85)
s m ⋅ s s
For this relation to hold, each unit must have a final exponent of zero. If we con-
sider the exponents for length (m), we find that the following relation must be true:
a + b − c + d = 0. (4.86)
−2a − b − d = 0. (4.88)
Equations 4.86, 4.87, and 4.88 form a system of simultaneous algebraic equations.
The system can be simplified to yield
a = −b
a = −c (4.89)
a = −d.
e g h
kg f kg m
3 (m ) = 1. (4.91)
m m ⋅ s s
Following the same steps as before, we obtain our second dimensionless term,
ρLv
ρ1µ −1L1v1 = . (4.92)
µ
F ρLv
f = 0, (4.93)
Lµv ′ µ
DIMENSIONAL ANALYSIS 113
which is a simpler functional form because it has fewer terms compared to the
original expression.
Suppose we wish to build a model of the bacterium 1000 times its normal size,
expose the model bacterium to fluid flow, and measure the drag force exerted
on the model. Recall that our second dimensionless term is ρLv/μ, which is the
Reynolds number. If we simply increase the length of the bacterium by a factor
of 1000, the Reynolds number will also increase by a factor of 1000. To scale the
experiment correctly and keep the Reynolds number constant, one could use a
fluid that has the same density but with 1000 times the viscosity and run the
bacterium model at the same velocity as the actual bacterium. We can use our
first dimensionless term to see how this would affect our measured force. In
particular, the first dimensionless parameter was F/Lμv, indicating that if L and
μ are scaled up by 1000, the drag force will be scaled up by a factor of
1000 × 1000 = 1 × 106. In this case, the measured drag must be scaled down by
the same amount—a million, not a thousand—to obtain the actual full-scale
drag acting on the bacterium.
Example 4.5: Determining the Reynolds number using a scaling relationship from the
Navier–Stokes Equations
For a one-dimensional flow, we assume that the flows in form a ratio of inertial to viscous terms, we make the
y- and z-directions are negligible, and so are changes in quotient of the left side versus the right side
those directions. The Navier–Stokes equations in x sim-
plifies to (ρU 2 /L)/(µU /L2 ) = ρUL/µ,
ρ(∂u∂t + u∂u∂x ) = ρg − ∂P ∂x + µ∂ 2u∂x 2 . which is the Reynolds number. You can make similar
ratios using the time-varying term to get the Womersley
We are interested in the viscous versus inertial terms, so parameter, and other parameters based on gravity, pres-
we ignore the time-varying term (the first term), and the sure, etc. This also explains why in pipe flows, the transi-
gravitation and pressure terms. This simplifies to tional Reynolds number can be above 1; in fully
developed, steady, one-dimensional pipe flows, the left-
ρu∂u∂x = µ∂ 2u∂x 2 . hand side is actually zero, because the term ∂u∂x is zero
and there are technically no inertial terms at all. At the
Let us use U as a characteristic velocity and L as a char- microscopic scale, however, there are imperfections in
acteristic length scale. Then, ∂u∂x can be estimated as the pipe surface, which leads to non-zero v and w, as
U/L, and, ∂2u∂x2 can be estimated as (U/L)/L = U/L2. well as changing u. Those perturbations can (with
The inertial terms (associated with momentum) are on increased velocity) induce turbulence, but usually at Re
the left, which simplifies to rU 2/L, and the viscous terms much greater than 1, more typically around 1000–2000.
are on the right, which simplifies to μU/L2. If we want to
Key Concepts
• A fluid differs from a solid in that it takes on the • The Navier–Stokes equations results from combining
shape of the container in which it is placed. equilibrium (conservation of momentum), constitutive
• Hydrostatics is the study of pressure in a stationary (Newtonian behavior), and kinematic (compatibility
fluid, and exerts pressure equally in all directions. assumption) relationships. They describe the behavior
of a very wide class of problems in fluid mechanics.
• Newtonian fluids are those in which the velocity
gradient is proportional to shear stress. The • For non-Newtonian fluids there is a nonlinear
proportionality constant is called the dynamic relationship between shear stress and velocity
viscosity. gradient. Power law fluids and Bingham plastics are
• Laminar flow involves ordered or aligned streamlines. two examples.
In contrast, turbulent flows involve tortuous • Rheology is the study of materials with some capacity
streamlines and mixing. The dimensionless quantity to flow. Viscoelastic materials exhibit both solid- and
Reynolds number (Re) is a measure of velocity and fluidlike behavior. The complex modulus can be used
governs the transition from laminar to turbulent flow. to describe viscoelastic behavior. It may depend
• Generally, Reynolds numbers much less than one are nonlinearly on frequency, and power-law relationships
dominated by viscosity and are laminar. Reynolds can be effective descriptions of complicated
numbers much greater than one are dominated by viscoelasticity.
inertia and are turbulent. The specific Reynolds • Dimensional analysis, scaling, and estimation are
number at which the transition occurs depends on methods that can provide insight into functional
the specific geometry. At the cellular scale, flow is relationships when a specific equation is not
typically laminar. available. Often these relationships are sufficient to
solve important problems.
• The velocity profile of laminar Newtonian flow for
simple geometries can sometimes be solved in • Dimensional analysis exploits the requirement for units
closed form. to be consistent in order to arrive at potential functional
relationships and novel dimensionless numbers.
• If viscosity is small enough to be neglected, the flow is
termed inviscid and is governed by Bernoulli’s equation.
116 ChApTER 4: Fluid Mechanics Primer
Problems
1. Determine the flow profile in a pipe with circular cross canceled out). So in the example, we can eliminate the
section, with inner radius r0, using differential analysis, time component of velocity and acceleration, showing
assuming pressure-driven flow with the same flow them to be redundant with respect to time. If we
conditions as we used for the parallel-plate problem. eliminate acceleration, we get the rank = 2, and time
Determine the ratio of the peak velocity to the average and velocity are left.
velocity. Is this ratio the same or different from that of (a) Based on this, go back to the fluid drag problem
the parallel-plate solution? presented in the text and determine the unit matrix
for that scenario, and show that the rank is 2. What
2. Show that for a Newtonian fluid, the complex shear
other base parameter groups could we use? (Hint: in
modulus and complex viscosity simplify to pure fluid
general, the dependent variable, force in this case,
values under an oscillatory input.
is typically not used in the base parameter groups,
3. Plot the flow profiles of a power-law fluid where β = 1 to avoid having it show up in multiple places in
and α = 0.5 or 2. You may select the plate gap and peak the function, thereby making it hard to isolate. It is
velocity for convenience, but make sure you specify partly for this reason some people prefer leaving the
them in your response. dependent variable outside the function.)
(b) Using a second base parameter group, derive the
4. Using the Navier–Stokes equations, derive the shear dimensionless form of the function. Based on this
stress acting on the bottom plate of a parallel-plate new set, if you decrease the length by a factor of
setup, with gap h, if there is both a pressure gradient 1000, keep the viscosity the same, keep the velocity
(dp/dx = C) and an upper plate moving with velocity V0, the same, increase the density by 1000, and multiply
with the bottom plate stationary. The Newtonian fluid the end force by 1000, can you still scale the
between the plates has density ρ and viscosity μ. experiment correctly? Does that make sense? Why?
5. If a cell can be treated as a fluid-filled bag with the fluid 8. Suppose you are considering a spring-dashpot system
having viscosity about 10 times that of water, determine as a model to describe some process. The spring
whether the flow within a cell during micropipette behavior can be described by Hooke’s law, F = kx, where
aspiration is laminar, turbulent, or transitional. You may x is a displacement, F is the force, and k is the spring
assume the pressure gradient is applied so that the cell constant. The Newtonian dashpot behavior is described
takes about 1 min to enter the pipette completely, at by F = μv, where v is the velocity, and μ is the viscous
roughly uniform velocity. coefficient (not viscosity). You wish to determine
the characteristic frequency of the system. Using
6. Determine a relationship for the radii of blood vessels dimensional analysis, find the dimensionless parameters
that branch out from a parent vessel of radius R to two governing this frequency.
progeny branches of radii R1 and R2 (not necessarily
equal) such that the shear stress resulting from laminar 9. Dimensional analysis is frequently used in fluid
blood flow on the inside of these vessels remains simulation. Let us say you are given a smooth sphere of
constant. Ignore fluid effects at the bifurcation itself. You radius r, suspended in a fluid of density ρ and viscosity
may assume that blood is Newtonian and incompressible. μ with a spring of constant k, under gravity g. You
stretch the spring out and release it to let the system
7. In the text, we discussed the number of base oscillate. Use dimensional analysis to determine the
parameters and units. A unit is independent if it cannot relationship of the “half-life” time of damping to the
be expressed in terms of the other units. To derive the other parameters. This half-life damping time refers to
number of independent units, one can create a matrix the time it takes the oscillations to reach one-half of the
based on the units at the top of each column and the initial amplitude when the system is initialized.
parameters beside each row. Example:
10. How much power do you give off in body heat? That
is, your presence in a room will warm the room from
time 0 1
1 your body heat. If I were to replace you with a lamp of
velocity −1 some wattage to achieve the same warming rate, what
acceleration 1 −2 wattage bulb would I need?
11. Each cell in your body is very roughly the same density
Where the first column is meters and the second as water (in reality, a bit higher, but ignore that for this
column is seconds. So time has no meters units question). If half of a typical body weight is in matrix
and a single time component (t = s1). The velocity (bones, cartilage, etc.), how many cells does a typical
has a meters component and a 1/s component adult have?
(v = m/s = m1s−1), and similarly for acceleration. The
rank of the matrix tells you the number of independent 12. How much force could your biceps muscle generate
units, and if you perform Gaussian elimination, you can if it were directly connected to a weight? Note that in
derive terms that can serve as the parameters on which the body, the bicep is actually levered to your forearm
you will base the dimensional analysis (these are not bones with the elbow as a fulcrum.
Annotated RefeRences 117
Annotated References
Fabry B, Maksym GN, Butler JP et al. (2001) Scaling the microrheol- Pritchard PJ (2011) Introduction to Fluid Mechanics. John
ogy of living cells. Phys. Rev. Lett. 87, 1481–2. This article describes Wiley. This textbook on fluid mechanics covers in much more
how cells may be characterized as soft, glassy materials by analysis mathematical detail topics including hydrostatics, differential
of the frequency response of beads attached the cells. analysis, mass conservation, dimensional analysis, and the
Navier–Stokes equations.
Kamm R (2001) Molecular, Cellular, and Tissue Biomechanics. Lec-
ture notes from course number 20.310, Massachusetts Institute Stamenovic D, Suki B, Fabry B, et al. (2004) Rheology of airway
of Technology. This course introduced the concept of scaling smooth muscle cells is associated with cytoskeletal contractile
analysis and estimation methods applied to cell mechanics. It stress. J. Appl. Physiol. 96, 1600–1605. This article uses the power-
was the inspiration for several of the authors to undertake law presented in Equation 4.17 and shows an application of rheo-
this text, which has been heavily influenced by it as a result. logical analysis to studying cell response.
Kollmannsberger P & Fabry B (2011) Linear and nonlinear rheology Vogel S (1996) Life in Moving Fluids. Princeton University Press. This
of living cells. Annu. Rev. Mater. Res. 41, 75–97. A review of the rheo- book has a minimally mathematical description of biological fluids,
logical findings as applied to cells. This article contains many refer- with more focus as the organisms scale. It covers many key con-
ences to trace how this sort of modeling developed. cepts about biology and fluids in an accessible way.
This page intentionally left blank
to match pagination of print book
CHAPTER 5
Statistical Mechanics
Primer
principle: within a closed system, any form of energy can be transformed into
another form, but the total energy of the system remains constant. In thermody
namics, we are concerned with several forms of energy, one of the most important
being internal energy. For now, we will simply consider it as a form of energy that
is the sum of multiple types of energy, one of the primary types being potential
energy.
In mechanics, potential energy is defined as the capacity to do work. It is a par
ticularly important form of energy in analyzing mechanical systems because of
the principle of minimum total potential energy. This principle states that when a
structure is subject to mechanical loading, it shall deform in such a way as to min
imize the total potential energy in reaching equilibrium. The implications of this
principle are that it does not matter if a structure is relatively simple (such as a
single polymer) or relatively complex (such as an entangled network of one mil
lion polymers), potential energy can take the form of a single scalar quantity that
represents the mechanical state of the structure. In addition, by finding the con
figuration that minimizes the potential energy, we can determine the equilibrium
configuration of the structure under a given mechanical load.
Let us consider an example using springs to demonstrate how potential energy
can be used to determine the equilibrium state of a mechanical system. Con
sider a Hookean spring. The force required to separate the ends of the spring by a
distance x is given by F = k1(x − x1), with a spring constant k1 and equilibrium
length x1. If we have a second Hookean spring, we know that F = k2(x − x2). In
this case, the potential energies of the two springs are W1 = ½k1(x − x1)2 and
W2 = 0.5k2(x − x2)2. If we link the ends of both springs together so that their lengths
are the same, and pull them by their ends so they begin to stretch, the total poten
tial energy is
because their lengths are the same (Figure 5.2). As the length x is changed, the
total potential energy of our spring system also changes. We can see that although
this system contains multiple bodies (that is, two springs), the potential energy
gives a single (scalar) representation of the mechanical state of this system. This
quantity has a useful physical meaning, as it is equal to the work this system is
capable of performing.
Now, suppose we allow the system to equilibrate without applying any external
forces—we just let the system relax. We know that each spring has its own equilib
rium length, however, now that their ends are joined, the two-spring system has
its own equilibrium length. As we mentioned earlier, the principle of minimum k1
total potential energy states that we can find the equilibrium state by finding the spring 1:
configuration that minimizes the total potential energy. For our spring system,
this means finding the length x at which Equation 5.1 is minimized. We can find k2
this length by taking the derivative of W with respect to x, and setting this equation spring 2:
to zero:
k1
dW/dx = k1( x − x1 ) + k2 ( x − x 2 ) = 0. (5.2)
two-spring
Solving for x, we get system:
k2
x = (k1 x1 + k2 x 2 )/(k1 + k2 ), (5.3)
x
which is the equilibrium length of our system.
Figure 5.2 Schematic of a two-
Note that finding the equilibrium length via the principle of minimum total poten spring system. The equilibrium length
tial energy gives the same answer as that obtained by performing a force balance. of the system, x, can be found by
We know that in equilibrium, without any external forces, there should be no net minimizing the potential energy.
122 CHApTER 5: Statistical Mechanics Primer
internal forces. Therefore any compression from one spring should balance the
extension from the other. Here, the equilibrium length x can be computed as
where we use a negative sign on the second spring to indicate that it is mechani
cally opposing the first spring (the second spring is compressing if the first spring
is extending and vice versa). As expected, Equation 5.4 is identical to Equation 5.3.
The reason performing a force balance is equivalent to potential energy minimi
zation is that fundamentally, forces are the gradient of potential energy. Forces are
balanced (net forces are zero) when there is no gradient of potential energy, which
occurs at maxima or minima. Energy maxima are technically equilibrium points
but are unstable, so they are omitted from the discussion (the classic example is
an inverted pendulum). While performing a force balance and potential energy
minimization produce equivalent results, in analyzing complex mechanical sys
tems, the latter approach is often preferred due to mathematical simplicity.
Potential energy has the added advantage that it allows us to marry the analytical
machinery of continuum mechanics with that of statistical mechanics, as we will
see in the next section.
F = EA∆L/L. (5.5)
Nota Bene The potential energy input into the rod to achieve this displacement is
Microscopic strain energy.
Microscopically, by strain energy we W = ½F ∆L. (5.6)
mean that when we displace the tip
of the rod a bit, all of the molecules In the rod example we considered the material response from the force and dis
or atoms making up the rod are placement at the tip to be distributed throughout the material in the rod, in the
separated or deformed a bit more form of stress and strain. The potential energy input into the rod by the force at the
from their equilibrium distances and tip is analogously distributed through the material in the rod. For each small vol
shapes, resulting in spring-like ume in the rod, a small element of potential energy is stored that we denote as dW.
potential energy storage throughout
the rod.
Thus, the total potential energy stored in the rod is the product of the volume of
the rod and this small element of energy
W = LAdW . (5.7)
If we equate the potential energy input at the tip with the internal potential energy
stored, we get,
dW = ½F /A ∆L/L = ½σ ε. (5.8)
The internal potential energy is strain energy, because it is the energy stored in the
form of strain (similar to the way the potential energy of a spring depends on its
stretch). The internal potential energy per unit volume is the strain energy density.
INTERNAL ENERGY 123
We showed for an axially loaded rod, the strain energy density can be found as
one-half of the product of stress and strain. In general, it can be shown that the
strain energy density is one half of the component-wise product of stress and
strain
dW = ½(σ 11ε11 + σ 12ε12 + σ 13ε13 + σ 21ε 21 + σ 22ε 22
(5.9)
+ σ 23ε 23 + σ 31ε 31 + σ 32ε 32 + σ 33ε 33 ).
resulting in increases or decreases in internal energy. This will allow us to ignore con
tributions to the internal energy that (1) do not change (or which change relatively
little) compared with the reference state, or (2) are largely decoupled from changes
in configuration. To illustrate this, in many of our calculations, we will assume that
changes in internal energy arise solely out of changes in potential energy, and ignore
changes in internal energy arising from alterations in kinetic energy.
5.2 ENTROPY
Entropy is directly defined within statistical mechanics
Our focus now turns to another important thermodynamic quantity, entropy.
You may have heard the vague explanation that entropy is related to disorder.
Nota Bene Many people have heard the “messy room” analogy: a messy room is more dis
ordered than a clean room, and so has higher entropy. One reason why the con
Entropy and the first law of cept of entropy can be difficult to grasp is that no direct relation for entropy
thermodynamics. There are
exists within thermodynamics. Instead of a direct definition, an incremental
classically three laws of
thermodynamics: definition relates the change in entropy ΔS to a change in heat Δq of a system. In
particular, for a constant temperature (isothermal) and reversible process (one
(1) Conservation of energy. that can be reversed without changing the system or its surroundings), the
(2) The entropy of an isolated system
change in entropy is defined as
(one that does not exchange
energy or mass with the outside)
∆S = ∆q/T , (5.10)
never decreases.
(3) As the temperature of a system
reaches absolute zero, its entropy where T is the absolute temperature and Δq is the amount of heat absorbed. We
reaches a minimum. see that this is an indirect, incremental definition of entropy, as it only relates the
change in entropy to thermodynamic quantities, rather than directly defining
The incremental definition of entropy
is a consequence of the first law.
entropy itself.
Specifically, the change in internal In contrast, within statistical mechanics, entropy is directly defined. This defini
energy, ΔW is equal to the change in tion of entropy was developed by Ludwig Boltzmann and was one his most impor
“entropic” energy SΔT minus the
mechanical work done by the
tant contributions. In particular, the Boltzmann definition of entropy is
system.
S = kB ln Ω. (5.11)
Nota Bene Here, kB is the Boltzmann constant and is equal to 1.38 × 10−23 J/K. Ω is defined as
the density of states and is equal to the number of microstates for a given mac
Ludwig Boltzmann’s tombstone. rostate. Each of these terms—density of states, microstates, and macrostates—will
Ludwig Boltzmann (1844–1906) was
be discussed in detail next.
an Austrian physicist who made
seminal contributions to thermody
namics and statistical mechanics. He
was a full professor by age 25 and an Microstates, macrostates, and density of states can be
advocate of the atomic theory of exemplified in a three-coin system
matter when it was unpopular. His
tombstone is inscribed with his Before formally defining microstates, macrostates, and density of states, we present
famous equation (Figure 5.3). an example to gain some intuition for these quantities. Consider a set of three coins
placed inside a container (such as a coffee can): a nickel, a dime, and a quarter. We
designate the nickel as coin 1, the dime as coin 2, and the quarter as coin 3. When
we shake the container, this flips all three coins simultaneously. Each of the coins
can land heads (denoted as h) or tails (denoted as t). Whether each coin lands heads
or tail is random (mathematically, this is a random variable).
When we shake the container, we consider each microstate to be a possible out
come of a shaking event specified by whether each coin comes up heads or tails.
If we were to shake the can, and the nickel comes up heads, the dime comes up
tails, and the quarter comes up heads, this microstate, which we denote as m, can
be written as
m = hth
ENTROPY 125
where the ith letter designates the outcome (that is, h or t) of coin number i. We
say that this system’s current microscopic state, or microstate, is “hth”
Given that there are three coins and two possible outcomes per coin, there are
eight possible microstates. These individual microstates, denoted by mx, are:
m1 = hhh
m2 = hht
m3 = thh
m4 = hth
m5 = htt
m6 = tht
m7 = tth
m8 = ttt.
We now turn our attention to macrostates. Imagine a case where in which we are
not able to observe whether each coin lands heads or tail, but only some property
that is dependent on the number of coins that lands heads. For instance, imagine
that we place a small elf into the coffee container with the coins and cover the top
of the coffee can with a lid (Figure 5.4). We tell the elf inside to count the number
of coins that lands heads after each shake of the container, and yell out that num
ber. We shake the container, and after the dime and the nickel (but not the quar
ter) lands heads, we hear a diminutive voice yell out, “two!” We shake the container
again, and after all the coins land heads, we hear “three!”
We consider this number, the total number of heads, to be representative of the
current system’s macroscopic state, or macrostate. It is considered a macroscopic
value because we cannot observe the exact configuration of the microscopic con
stituents that led to this macrostate (in other words, whether each coin landed
heads or tails). Rather, we are only able to observe some macroscopic property. If
126 CHApTER 5: Statistical Mechanics Primer
we define W(mx) to be the number of heads in microstate mx, then for each of the
eight microstates, the corresponding system macrostate is
m1 = hhh: W (m1 ) = 3
m2 = hht: W (m2 ) = 2
m3 = thh: W (m3 ) = 2
m4 = hth: W (m4 ) = 2
m5 = htt: W (m5 ) = 1
m6 = tht: W (m6 ) = 1
m7 = tth: W (m7 ) = 1
m8 = ttt: W (m8 ) = 0.
Notice that W can range from 0 to 3, and that the number of microstates associ
ated with each macrostate is different. In particular, the macrostate correspond
ing to W = 3 has only one microstate associated with it (hhh), whereas the
macrostate corresponding to W = 2 has three microstates associated with it (hht,
thh, and hth). The specific microstates associated with each macrostate are:
W = 3 : hhh
W = 1: htt,tht,tth
W = 0 : ttt.
We can now introduce the density of states (sometimes called the multiplicity of
states), Ω(W), as the total number of microstates associated with each macrostate
W. In our three-coin system, we know that for W = 3 (three heads), there is only one
such microstate (hhh). In this case, Ω(W = 3) = 1. For W = 2, there are three such
microstates: hht, thh, and hth. Therefore, Ω(W = 2) = 3. Similarly, Ω(W = 1) = 3 and
Ω(W = 0) = 1.
ENTROPY 127
is small when Ω(W) is small. Note that the minimum number of microstates asso
ciated with a particular macrostate is one. Therefore, the lowest possible value of
Ω is 1; in this case, S = kBln(1) = 0.
Ψ = W − TS. (5.12)
The principle of minimum free energy states that a closed system (namely, a sys
tem which can exchange energy in the form of heat or work, but not matter, with
its surroundings) at constant temperature will spontaneously adapt itself to lower
its free energy. A thermodynamic equilibrium is reached when free energy is min
imized. Just as potential energy minimization can be used to find the equilibrium
state of mechanical systems, free energy minimization can be used to find the
equilibrium state of thermodynamic systems.
Upon inspection of Equation 5.12, it becomes evident that for a given body, the
free energy can be minimized in two distinct ways: through decreasing the inter
nal energy W, or increasing the entropy S. The particular influence of energy ver
sus entropy in the process of free energy minimization is determined by the
relative magnitudes of W and S, as well as the temperature T. For example, at zero
temperature, the entropy term vanishes, and the free energy is equal to the inter
nal energy. Conversely, at very high temperatures, the TS term will dominate the
free energy. When the entropic contribution to the free energy TS is comparable
to the energetic contribution W (when W = TS.), a transition occurs between
energy-dominated and entropy-dominated behavior.
One question that may arise is if we have two principles of energy minimization,
one based on potential energy and one based on free energy, how do we know
FREE ENERGY 129
which one is appropriate for a given system? Also, would these two principles give
contradictory behaviors? For everyday (nonmicroscopic) structures such as a Nota Bene
steel beam in a bridge, we know that the principle of minimum potential energy
dictates its deformation under loading. However, we could also consider the beam Helmholtz and Gibbs free energy
are used for similar purposes but
to be a thermodynamic body subject to the laws of thermodynamics, which state
in different contexts. Free energy
that its deformation should be dictated by the principle of minimum free energy. denoted by Ψ here is known as the
So which one is correct? Helmholtz free energy. As the
difference between energy and
The answer is they are both correct. It is important to point out that the principle
entropy, it is often thought of as a
of potential energy minimization is not contradicted by the principle of free measure of the work that a system
energy minimization; rather, it is encompassed in it. In particular, energetic effects could do at a constant temperature
for everyday structures are so dominant that they mask any entropic influences on and volume. However, there are other
their behavior. Mathematically, this can be expressed as W TS; the free energy “thermodynamic potentials” that are
is equivalent to the internal energy. Remember from Section 5.1 that the primary useful in various contexts. For
components of internal energy are potential energy and kinetic energy; however, instance, the enthalpy is defined as
the heat transferred to an isolated
in many cases we can assume that for the microscopic structural components of system at constant pressure and is
cells such as polymers and membranes, kinetic energy is generally not dependent defined as W + pV. The Gibbs free
on their mechanical configurations. The same is true for everyday structures as energy is often used in chemistry. It is
well. Therefore, changes in internal energy can be attributed solely to changes in the maximum work that can be
potential energy. In this case, for bodies in which W TS, free energy minimiza obtained from a closed system at
tion is equivalent to potential energy minimization. For brevity, we will refer to constant volume, W − TS + pV.
systems in which W TS as mechanical systems; systems in which W ~ TS or
W TS will be referred to as thermodynamic systems.
Nota Bene
Temperature-dependence of end-to-end length in polymers Zero temperature limit. We could
arises out of competition between energy and entropy consider our treatment of solid
mechanics in Chapter 3 to be the
We now explain the concept of the competing influences between energy and zero temperature limit. This does not
entropy as a system spontaneously minimizes its free energy. Let us revisit our literally mean that the cell is frozen to
absolute zero. Rather, it is a common
example of a polymer subject to thermal fluctuations. If we were to observe a sus way of expressing that for the
pended actin polymer over time at room temperature, it would not be completely particular process we are interested
rigid, rather, it would continuously wiggle. Recall that the tendency of the polymer in, the entropic contribution TS is
to wiggle is due to random molecular forces similar to those that cause Brownian negligible as compared with the
motion. In the same way that increasing the temperature increases the small fluc energetic contribution W.
tuations in position of a suspended particle during Brownian motion, increasing
the temperature increases the tendency of the polymer to take on an increasingly
wiggly configuration (Figure 5.5).
To develop some intuition for the competition between energy and entropy in this
temperature-dependent phenomenon, we present an example of a model poly
mer subject to thermal fluctuations. Note that the formal mathematical frame
work for this introductory example will be given in subsequent sections. We first
need to define two lengths (Figure 5.6). The contour length, L, is the length of the
polymer if it were completely straight. The end-to-end length, R, which, as the
name implies, is the length of a straight line between one end of the polymer and
the other, and depends on the actual configuration of the polymer. If the polymer
is straight, R = L; in all other cases, R < L. Therefore, R is always less than or equal
to L but greater than or equal to 0. If we were to observe a thermally fluctuating
R<L
MiCroCanoniCal EnsemBle 131
The entropy is proportional to ln Ω, so the higher the value of Ω, the higher the
entropy. Therefore, polymer macrostates with end-to-end lengths of R < L are
associated with higher values of entropy than the polymer macrostate R = L.
Recall that equilibrium is achieved when the free energy is minimized, and that
increasing the entropy decreases the free energy. In our example, the smaller the
end-to-end length, the higher the entropy, and we therefore say that entropically,
a coiled configuration is more favorable than a straight configuration.
Now let us look at the effect of internal energy on the end-to-end length. Consider
the same polymer model as before, but this time imagine that instead of the links
being connected by freely rotating hinges, we now straighten out the polymer and
attach rotational springs between each link such that the springs are at their equi
librium position when adjacent links are straight (180° between the links) (Figure
5.8). The addition of these springs is a theoretical construct that represents an
increase in internal energy if the polymer is bent. In this case, bending the poly
mer requires work, which becomes stored in the springs as strain energy. Thus,
the polymer macrostate associated with R = L is associated with zero internal
energy (W = 0), while all other macrostates are associated with non-zero internal
energy (W > 0). Because the free energy is decreased when the internal energy is
reduced, we say that energetically, a straight configuration is favorable.
Now we know that entropically, a bent or coiled configuration is favorable in our
model polymer, while energetically, a straight configuration is favorable. So the
question becomes, what configuration will it tend to adopt? The answer is it depends
on the temperature. If we look at the expression for free energy, we can see that the
influence of energy or entropy on the free energy is determined by the temperature.
At lower temperatures, internal energy dominates the free energy. The polymer will
favor macrostates that are energetically favorable and will tend to be straighter. At
higher temperatures, entropy dominates the free energy; the polymer will favor
macrostates that are entropically favorable and will tend to be more wiggly.
In the next several sections, we will quantify this behavior by formulating a rela
tionship that relates the equilibrium configuration of the polymer with tempera
ture and its microscopic properties, such as the number of links and the energy of
bending. We will perform the same calculation using two distinct analytical
approaches: the microcanonical ensemble and the canonical ensemble.
W = N h. (5.14)
with the same internal energy W. To apply the microcanonical ensemble to our
model polymer, we consider it to be thermally and mechanically isolated from the Nota Bene
surroundings. We consider a constant energy ensemble of microstates, with each
microstate within the ensemble having the same energy W. The ensemble can be Are all microstates within an
ensemble equally likely to occur?
considered to be the collection of all microstates associated with the macrostate
The assertion that all microstates
having energy W. Within such an ensemble, each microstate is equally likely to within an ensemble are equally likely
occur. Therefore, if there are Ω microstates at a particular energy W, each micro to occur is actually an assumption. It
state within the ensemble occurs with the probability of p = 1/Ω. Our goal is to is sometime called the ergodic
calculate S(W), or the entropy for this ensemble of microstates having energy W. hypothesis or the fundamental
postulate of statistical mechanics
or the equal a priori probability
Entropy can be calculated via combinatorial enumeration postulate. At its core, we assume that
of the density of states we are considering the system for a
long enough period of time that all
Because entropy depends on the density of states, we must first calculate Ω(W), or possible microstates in the ensemble
the density of states Ω for each ensemble with energy W. We can find this using occur for equally long periods.
combinatorics. If we have N possible hairpin sites, the total number of configura
tions can be found using the bionomial coefficient. The binomial coefficient is
Nota Bene
m m!
n = n !(m − n )! , (5.15) Simplification of large factorials.
Factorials often appear in statistical
mechanics because calculating
which gives the number of ways in which a group of n items can be chosen from a microstates involves combinatorics.
set of m. In our problem, we can use the binomial coefficient to find the total num Factorials of large numbers are
ber of microstates in which Nh hairpins occur, given N total sites difficult to calculate. However, an
approximation can often be
sufficient. For instance, notice that:
N N!
Ω( N h ) = = .
N h N h !( N − N h )! n n
(5.16) ln n! = ∑i =1
ln i ≈
∫ ln x dx ,
Now, we know that the entropy S = kB lnΩ, or 1
entropy S
hairpin.
0
internal energy W
Nota Bene would behave at thermal equilibrium. In particular, we know from thermody
namics at thermal equilibrium,
Thermodynamic equations of
state imply relations between 1 ∂S (W )
state variables. The expression = . (5.20)
1 ∂S T ∂W
= is called an equation of
T ∂W Combining Equation 5.19 and 5.20, we obtain
state because it gives a relation
between state variables
1 ∂S (W )
(macroscopic variables that describe =
the thermodynamic state of a T ∂W
system). Equations of state can be
derived from the fundamental kB W 1 W
=− ln − ln N −
thermodynamic relation N
d W = Td S − P d V + µ d N ,
W (5.21)
where P is pressure, μ is chemical kB
potential, and N is the number of = − ln
W
particles. In particular, dW can N −
equivalently be written as
kB
∂W ∂W 1
dW = dS + dV
∂S V ,N ∂V S ,N =− ln N .
− 1
W
∂W
+ dN ,
∂N S ,V Solving for W yields
fraction of sites
few hairpins occur. At high temperature,
entropy dominates, and hairpins occur
1
– more frequently. In the limit of infinitely
2 high temperature, half of the sites
contain a hairpin.
temperature T
What happens when you decrease the temperature? In the limit of T = 0, we
know that
Nh 1
lim = lim = 0. (5.24)
T →0 N T →0
e kBT
+1
Because there are no hairpins, W = 0. In other words, in the low-temperature
limit, the free energy is dominated by the internal energy, and the free energy will
be minimized when internal energy is zero. This occurs when no sites contain a
hairpin (for instance, when the polymer is “straight”). The transition between
these two limits is sigmoidal, as can be seen in Figure 5.11.
∂S Nh
= −kB ln N h − ln ( N − N h ) = −kB ln .
∂N h N − N h (5.28)
Plugging Equations 5.27 and 5.28 into W and solving for Nh/N,
Nh 1
= , (5.29)
N
e kBT
+1
which is equivalent to Equation 5.23.
equilibrium). Both the system of interest and heat bath have associated micro
states and internal energies. If we were to consider a polymer (the system of inter
est) placed in a very large chamber filled with an ideal gas (the heat bath), a system
microstate would be a particular configuration of the polymer, and a bath micro
state would be a particular state of the gas in which each gas particle had a par
ticular position and velocity. Both microstates would be associated with a
particular value of internal energy, one for the polymer, and one for the ideal gas.
Let the subscript “s” denote the system of interest and “b” denote the bath. In
addition, let Qs(ms) be the internal energy of the system microstate ms, and
Qb(mb) be the internal energy of the bath microstate mb. If we idealize the system
and bath together as a single entity, we can consider an ensemble of different
combinations of system and bath microstates such that the total internal energy
(the sum of the internal energy of the system and the bath) is Wtot. In this man
ner, by considering an ensemble of microstates with a fixed value of energy, this
approach is similar to the microcanonical ensemble example presented previ
ously in this chapter.
Consider the case in which the system is in microstate ms. The system’s internal
energy is Qs(ms), and the internal energy of the heat bath is Wtot − Qs(ms). An
important realization (and a key step in the derivation) is that the probability for
the system to be in microstate ms is proportional to the number of heat bath
microstates with energy Wtot − Qs(ms). This can be expressed as
(
p(ms ) ∝ Ω b Wtot − Qs (ms ) , )
(5.30)
where Ωb is the bath density of states. We can write the right-hand side in the
above expression as a function of entropy. In particular, S = kB lnΩ, so the right-
hand side of Equation 5.30 can be written as
Sb(Wtot −Qs (ms ))
Ω b (Wtot − Qs (ms )) = e kB
. (5.31)
Combining Equations 5.30 and 5.31, we find
Sb(Wtot −Qs (ms ))
p(ms ) ∝ e kB
. (5.32)
Next, require the temperature of our system to remain constant. We can accom
plish this by making the assumption that the heat bath is very large, such that
changes in energy and entropy of the system do not affect the energy and entropy
of the bath. Mathematically, this implies that for any system microstate ms, Qs/
Wtot 1, allowing us to perform a Taylor expansion. In particular, for very small
displacements Δx between two points x and x0, it can be shown that a “suitably
smooth” function f(x) can be approximated as
f ( x ) ≈ f ( x 0 ) + ∆xf ′( x ), (5.33)
when Δx/x0 1.
In our case, we can Taylor-expand Sb (assuming x0 = Wtot, x = Wtot − Qs, and Δx = −
Qs) using Equation 5.33 as
∂S b
( )
Sb Wtot − Qs (ms ) ≈ Sb (Wtot ) −
∂W
Qs (ms ). (5.34)
∂S b 1
= , (5.35)
∂W Tb
138 CHApTER 5: Statistical Mechanics Primer
where Tb is the temperature of the heat bath, Equation 5.34 can be rewritten as
Qs (ms )
Sb (Wtot − Qs (ms )) = Sb (Wtot ) − . (5.36)
Tb
Combining Equations 5.32, 5.36, we finally find
Sb (Wtot ) Qs (ms ) −Qs (ms )
− 1 (5.37)
p(ms ) ∝ e kB kBT
= e kBT
,
Z
where in the right-hand-most relation in Equation 5.37 we have turned the pro
portionality for the probability into an equality by introducing a normalization
factor,
−Qs (ms )
Z= ∑ ms
e kBT
. (5.38)
Note that the summation in Z is over all possible discrete system microstates ms,
and that we have lumped the exp(Sb(Wtot)/kB) term within Z, as it does not depend
on ms.
Although we have thus far only considered discrete microstates, we will often con
sider systems that possess a continuous distribution of microstates. In this case, Z
can equivalently be written as
−Qs (m s )
Z=
∫
ms
e kBT
dms . (5.39)
The probability distribution in Equation 5.37 and normalization factor in
Equations 5.38 and 5.39 are the key relations of the canonical ensemble. Equation
5.37 is called Boltzmann’s distribution, and Equations 5.37 and 5.38 are called the
canonical partition function.
Ω(W ) = e S kB = e ST kBT
, (5.41)
∫
Z = e − β Ψ(W )dW
(5.43)
Z= ∑e − βΨ(W )
. (5.44)
W
Collectively, our analysis shows that Z can be computed in two distinct ways,
using Equation 5.39 or Equation 5.43. The main difference between these two
expressions is that in Equation 5.43, the exponential term contains the free energy
Ψ instead of the microstate energy Q, and that the integral is over all macroscopic
energies W instead of microstates ms.
For very large systems, integrals or sums of the form in Equation 5.43 and Equation
5.44 can be approximated using the saddle point method, as long as the quantity
being integrated or summed over is an extensive quantity. An extensive quantity is
one that scales linearly with the number of particles in the system, whereas an
intensive quantity is one that does not depend on the size of the system. For
instance, mass is an extensive quantity, while temperature is an intensive quan
tity. Free energy is an extensive quantity, so the above expression can be approxi
mated using saddle point integration. In particular,
−Ψmin
Z =e kBT
, (5.45)
where Ψmin is the free energy evaluated at which Ψ(W) is a minimum. Rearrang
ing this expression,
Equation 5.46 has important implications, in that it reveals that the partition func
tion has greater utility than just serving as a normalization constant for the
Boltzmann distribution. In particular, we know that for closed systems, the free
energy is minimized at thermal equilibrium. Therefore, Equation 5.46 implies
that to compute this value of free energy at equilibrium for a given system, one
merely needs to compute the partition function using Equation 5.38.
Statistical mechanics deals with determining the over summation. It also implies that for integrals of the
all behavior of systems consisting of many bodies and/ form
or degrees of freedom. Within treatments in statistical
mechanics, it is often useful to make mathematical approx
imations when the number of degrees of freedom is very ∫
S = e Ny ( x )dx ≈ e Nymax
∑ p(m)Q(m) = Z ∑ e
1 − βQ (m )
Q = Q(m), (5.47)
m m
where β = 1/(kBT). Note that in Equation 5.47, the brackets denote the expected
value based on an average over all microstates. Equation 5.47 can equivalently be
written as
∂ 1 ∂ 1 ∂Z
∑ ∂β e ∑e
1 − βQ (m ) − βQ (m )
=− =− =− . (5.48)
Z Z ∂β Z ∂β
m m
142 CHApTER 5: Statistical Mechanics Primer
Therefore,
∂ ln Z
W = Q =− . (5.50)
∂β
This shows that, like the free energy, the internal energy at equilibrium can also be
found from the partition function. Again, we are able to find the equilibrium inter
nal energy without having to enumerate the density of states.
We now know that if we calculate the partition function, we can calculate Ψ and
W at equilibrium. If we want to know S, it can be calculated easily, because
Ψ = W − TS. Thus, we can specify the thermodynamic state of the system at equi
librium from its microscopic behavior, as long as we can calculate the partition
function Z. In the next section, we revisit our model polymer to demonstrate use
of these relations.
Q= ∑n ,
i =1
i (5.51)
where ni equals one if a hairpin occurs at hairpin site ni, or zero if a hairpin does
not occur. Combining the microstate energy in Equation 5.51 with the Boltzmann
distribution, we obtain an expression for the probability of a given microstate in
thermal equilibrium,
N
1 − β ∑ ni (5.52)
p(n1 ,n2 ,..., nN ) = e i =1 ,
Z
where Z is the discrete partition function given in Equation 5.37
N
−β ∑ ni
Z= ∑e i =1
. (5.53)
m
To compute Z, we need to perform a sum over all possible microstates. To do this,
we perform N summations over each individual hairpin site,
N N
−β ∑ ni 1 1 1 −β ∑ ni
Z= ∑e
m
i =1
= ∑ ∑ ...∑ e
n1 = 0 n2 = 0 nN = 0
i =1
. (5.54)
The summation in the exponential in the rightmost relation in Equation 5.52 can
be written out explicitly as
1 1 1
Z= ∑ ∑ ...∑ e
n1 = 0 n2 = 0 nN = 0
− β n1 − β n2
e ... e − β nN . (5.55)
RanDom WalKs 143
Now, because n1, n2, . . . nN are independent variables, we can separate the sums
1 1 1
Z= ∑
n1 = 0
e − β n1 ∑
n2 = 0
e − β n 2 ... ∑e
nN = 0
− β nN
.
(5.56)
In Equation 5.56, each of the sums are numerically equivalent. Therefore, can
rewrite the above expression as
N
1
Z =
∑e − β n
= zN , (5.57)
n =0
where
1
z= ∑e
n =0
− β n
= 1 + e−β (5.58)
Nota Bene
is termed the single partition function. With the partition function, we can calcu
Partition functions for systems
late the free energy at equilibrium as of non-interacting particles can
be calculated from single
ln Z N ln(1 + e − β ) (5.59) partition functions. In Equation
Ψ=− =−
β β
5.57 we saw the partition function
was calculated as
and the equilibrium internal energy as Z = ZN,
∂ ln Z e−β N where Z is the single partition
W = = Nε −β
= β . (5.60) function. This will be the case for any
∂β 1+ e e +1
system of N identical particles that do
not interact. For instance, in the case
Note that Equation 5.60 is identical to Equation 5.22, which we derived through of the hairpinned polymer, the N
the calculation of S(W) in the microcanonical ensemble (in conjunction with hairpin sites are identical (they all can
thermodynamic relations). Although we performed the same calculation in two undergo the same range of motion)
different ways, we can see that mathematics were quite different, with the use of and they do not interact (whether
the canonical approach avoiding the need to use combinatorics to compute the one site contains a hairpin is not
density of states. For this reason, the canonical ensemble may often be preferable influenced by whether a hairpin
occurs elsewhere.)
for analyzing thermodynamic systems.
will end up at location r? Note that if the player has taken more backward steps
than forward steps, r is negative, and vice versa.
If the player takes n steps, and for each step they can only move backwards or for
wards, then there are 2n different paths that can be taken (although the player may
end up in the same position with many different paths). The number M of those
paths in which only n+ of the steps are forward can be calculated using the bino
mial coefficient from combinatorics. To see why the binomial coefficient can be
used, imagine that the “objects” are the coin flips, and that the “picks,” or the
selection of an object, are how many heads (or positive steps) occur. Then, if there
are n total flips and n+ of them come out heads, M must be
n
M = . (5.61)
n+
Our goal is to calculate the probability of the player being at point r after n flips
and steps. We have the following relationships, n = n+ + n− and r = b(n+ − n−),
where n− is the number of backward steps. Combining these yields
r
n+
b. (5.62)
n+ =
2
And,
n
n!
M (n , r ) = n + r = .
b n+ r r (5.63)
n−
2 b! b!
2 2
The probability that the player will end up at position r after n steps (p(n, r)) is
simply the number of ways or “paths” the player could use to get to point r divided
by the total number of paths the player might take. M is the total number of ways
to get to r and, as we noted above, there are 2n total paths available to the player.
Therefore, the probability is
M (n , r ) 1 n!
p(n , r ) = = n .
2n 2 n+ r n− r
(5.64)
b! b!
2 2
Equation 5.64 is the exact expression for the probability that the player will be at
a position r on the field after n steps of size b. To get a feel for what the random
walk distribution looks like for different step numbers, we calculate p(n, r) for
various values of n in Figure 5.13. Note that the appearance of the distribution
begins to resemble a normal distribution as n gets larger. This is perhaps not
unexpected, as the position r results from the sum of several independent coin-
flipping events, with the probability of heads or tails the same at each step (in
other words a 50/50 chance). The central limit theorem states that for a suffi
ciently large number of steps, r will be approximated by a normal distribution,
with this approximation becoming better as the number of steps are increased.
This is an important result, as in many cases we are interested in the behavior of
random walks in the limit as n becomes very large. We can rewrite Equation 5.64
in the limit of large n in the form of a normal or Gaussian distribution,
1 2
2 nb 2
p(n , r ) = e −r . (5.65)
2πnb 2
Equation 5.65 gives the probability distribution for a one-dimensional random
walk for n steps of length b in the limit of large n; the full derivation for Equation
RanDom WalKs 145
p(n,r )
p(n,r )
p(n,r )
0.4 0.4 0.4 0.4
n=8 n = 16
1.0 1.0
0.8 0.8
0.6 0.6
p(n,r)
p(n,r)
0.4 0.4
0.2 0.2
0.0 0.0
−8 −6 −4 −2 0 2 4 6 8 −16−14 −12−10 −8 −6 −4 −2 0 2 4 6 8 10 12 14 16
Figure 5.13 Parametric plots of Equation 5.65. As n is increased, the distribution can be increasingly approximated as a normal or
Gaussian distribution. For simplicity we have assumed b=1.
Nota Bene
5.65 is in Section 7.4. It can be shown that for a random walk of n steps of size b,
the root mean-square position is Recurrence in random walks is
dependent on the dimension of
r2 = b n. the walk. Although it is beyond our
(5.66)
scope here, it can be shown that the
random walk example we have used
The derivation of Equation 5.66 is left as an exercise for you. Combining Equation here is recurrent. This means that
5.65 and 5.66, we obtain an alternative expression for the random walk probability eventually, the player will visit every
distribution, point on the playing field, even if the
field is infinitely wide. In the context of
1 −r 2 / r 2 games of chance, this property of
p(n , r ) = e . (5.67)
2π r 2
recurrence is sometimes called the
Gambler’s Ruin. The idea is that a
gambler’s funds at any point in time
The diffusion equation can be derived from the random walk can be considered analogous to the
soccer player’s position on the field.
At the molecular level, the process of diffusion is governed by the random motion The flip and step of the soccer player
of particles. It is perhaps not surprising that the behavior of a random walk scaled is the same as each spin of the wheel,
up to the continuum level can be described by the diffusion equation. This rela roll of the dice, or hand of cards. No
tionship was first demonstrated by Albert Einstein in a 1905 paper he published matter how large the gambler’s initial
while working in the Swiss patent office. He demonstrated that Brownian motion, bankroll (so long as it is finite), he or
she will eventually go bust, assuming
the random motion of small particles suspended in a fluid, results in a rate of dif the bets are always of the same
fusion related to the velocity of the individual particles. His theoretical results fit amount, even if the odds were fair
the observed data so well, that they provided convincing support for the atomic (instead of favoring the house, as it
theory of matter before there was direct evidence of the existence of atoms or usually does). Interestingly, and
molecules. perhaps surprisingly, the property of
recurrence of random walk processes
The one-dimensional diffusion equation is known as Fick’s second law (this law is can be shown to be limited to one-
covered in more detail in Chapter 9.3), and is given as and two-dimensional situations only.
In three dimensions (or higher), the
dC d 2C walker will not visit all available
=D 2, (5.68) coordinates. Random walks of these
dt dx
types are known as transient.
146 CHApTER 5: Statistical Mechanics Primer
where C is the concentration and D is the diffusion coefficient. We now show that
this equation can be derived from a random walk process.
First, let’s recast our statement of the random walk problem slightly. Consider a
particle that is constrained to move in only one dimension. We define its initial
position to be at r = 0. After some time t, we are interested in the probability that
the particle will end up at some new location r. In recasting our random walk, we
replace the number of steps n with time t, which we will initially treat as a discrete
variable. In this case, we are able to assess where the particle is at discrete time
points. In regards to the particle position, we will treat r as a continuous variable.
Thus, the particle can move arbitrary distances. To allow for steps of any distance,
not just steps to the right and left of a specified increment (for instance, we do not
use a fixed step size b), we need to introduce another probability function, p(x).
This function gives the probability distribution for a particle making a step (posi
tive or negative) of distance x in any time step. We can now state the new problem
in terms of the probability distribution of finding the walker at a certain position,
r, at time t + 1 in terms of the distribution at time t. Specifically,
∞
p(t + 1, r ) =
∫ p(x)p(t , r − x)dx. (5.69)
−∞
This type of relationship is known as a recursion relationship. It defines p at one
point in time in terms of its distribution at the prior time. Next we can perform a
Taylor series expansion for the p(n, r − x) term in the integral about r to obtain
∞
∫
dp(t , r ) 1 2 d 2 p(t , r )
p(t + 1, r ) = p( x ) (pt , r ) − x + x + O 3 ( x ) dx
d r 2 dr 2
−∞
∞ ∞ ∞
dp(t , r ) 1 d 2 p(t , r )
=
∫
−∞
p( x )p(t , r )dx −
∫
−∞
p( x ) x
dr
dx +
2 ∫
−∞
p( x ) x 2
dr 2
dx + O 3 ( x ),
(5.70)
where O3(x) represents terms of third order or higher in x. Because p(t, r) and its
derivatives do not depend on x, these terms can be removed from the integrals. In
this case, Equation 5.70 can be rewritten as
∞ ∞ ∞
dp(t ,r ) 1 d 2 p(t ,r )
∫
p(t + 1, r ) = p(t ,r ) p( x )dx −
−∞
dr ∫
−∞
p( x ) x dx +
2 dr 2 ∫ p(x)x dx + O (x).
−∞
2 3
(5.71)
Next, we seek to simplify each of the terms on the right-hand side of Equation 5.71.
For the first term, we know that probability distributions possess the property that
integrals over their entire range must be unity. Therefore, for the first term,
∞
∫
p(t ,r ) p( x )dx = p(t ,r ). (5.72)
−∞
The second term in Equation 5.71 can be simplified if we restrict the probability
distribution, p(x), to be isotropic. In other words, there is no bias in the probabil
ity of a particle to make a step in one direction or another, p(x) = p(−x). This con
dition also implies that there is no net motion of the particles, owing, for example,
to convection should there be flow in the underlying medium. This restriction
implies that the integral in the second term is zero, thus
∞
dp(t ,r )
dr ∫ p(x)xdx = 0. (5.73)
−∞
Key ConCepts 147
Finally, we assume that the higher-order terms from the Taylor series expansion
can be neglected, which can be shown to be valid for some small region around r
as long as p is a continuous function. The net result is that
∞
1 d 2 p(t ,r )
p(t + 1,r ) − p(t ,r ) =
2 dr 2 ∫ p(x)x dx.
2
(5.74)
−∞
Now, consider the right-hand side in Equation 5.74. It is the integral over all pos
sible positions of the particle of the position squared times the probability of being
in that position. This is the definition of the average squared position, a term that
we encountered earlier in this chapter. In Section 7.4, we show that for a random
walk of n steps with fixed step size b, 〈r2〉 = nb2 (see Equation 5.66). It can be simi
larly shown that for a random walk with variable step size but with an average
squared displacement of Δr2 per step,
r 2 = n∆r ,
2
(5.75)
where n is the number steps. Combining Equations 5.74 and 5.75, we get
t ∆r 2 d 2 p(t , r )
p(t + 1, r ) − p(t , r ) = . (5.76)
2 dr 2
Now, in the limit of a large number of particles, all behaving according to this
equation, the concentration of particles at a given location is proportional to the
probability of finding a single particle at that position. Therefore, we can replace
P(n,r) by the concentration C. Also, if we divide the equation by Δt, the left-hand
side becomes the time derivative of concentration in the continuous limit and
we have
dC n∆r 2 d 2C
= . (5.77)
dt 2∆t dr 2
n∆r 2 (5.78)
D= .
2∆t
We will build on our treatment of random walks in subsequent chapters. We will
use the random walk to better understand the behavior of polymers in Section
7.4. We will also learn how confining diffusion into a two-dimensional mem
brane can greatly enhance the kinetics of diffusion-limited biochemical reaction
in Section 9.1.
Key Concepts
• Within statistical mechanics, entropy is directly behavior, whereas the influence of entropy increases
defined as S = kB lnΩ. Ω is the density of states gives with increasing temperature. The competing nature of
the number of microstates associated with a given energy and entropy in equilibrium behavior is
value of a macrostate. captured by a thermodynamic potential called the
• Thermal equilibrium of thermodynamic bodies is free energy, defined as Ψ = W − TS.
governed by two competing phenomena: internal • The principle of minimum total potential energy states
energy, and entropy. At low temperatures, internal that when a structure is subject to mechanical
energy–driven phenomena dominate a body’s loading, it shall deform in such a way as to minimize
148 CHApTER 5: Statistical Mechanics Primer
the total potential energy. The principle of minimum probability of a given microstate is given by the
free energy states that a closed system at constant Boltzmann distribution. The partition function is a
temperature will spontaneously adapt itself to lower normalization factor for the Boltzmann
its free energy. Just as potential energy minimization distribution, and is related to the minimum value of
can be used to find the equilibrium state of free energy. We can specify the thermodynamic state
mechanical systems, free energy minimization can be of the system at equilibrium by calculating the
used to find the equilibrium state of thermodynamic partition function without enumerating the density
systems. of states.
• Different analytical approaches can be used to • Random walks are a class of mathematical problems
determine behavior of thermodynamic systems at in which at discrete points in time a “walker” moves
equilibrium, such as the microcanonical ensemble in space a specified distance. Through its use we can
and the canonical ensemble. In the microcanonical relate certain macroscopic behavior of
ensemble, we analyze microstates at constant energy. thermodynamic bodies to microscopic properties.
Thermal equilibrium can be obtained from relations • When the behavior of random walks is scaled up to
derived from the microcanonical ensemble. the continuum level, their behavior is described by
• In the canonical ensemble, we analyze microstates in the diffusion equation.
a system held at constant temperature. The
Problems
1. We know that the mechanical behaviors of everyday Assume that the amount of this bias is a small number,
rod-like objects, such as steel beams and jump ropes, . Now, what is the probability of each macrostate?
can be predicted through the principle of minimum
total potential energy, but not soft structures such as 3. Consider a strand of DNA which only contains two
biopolymers, whose behavior is significantly influenced different nucleotides, A or T. Using the binomial
by entropy. We know that a transition between energy- coefficient, calculate the number of different possible
and entropy-dominated behavior occurs when W = TS. sequences if the strand is N nucleotides long, and only
Given the statistical mechanics definition of entropy, N1 of the nucleotides are T.
we can estimate a characteristic energy at which this
4. Consider 100 people in a building. Every time a person
transition occurs as kB (why would this be the case?).
moves up one floor, it takes them 1kBT of energy.
Calculate this value of energy assuming a temperature
Assume that given enough time the people distribute
of 37°C.
themselves according to Boltzmann’s distribution.
Now, consider a length of rope and a strand of DNA. How many people are on each floor on average? How
Calculate the strain energy stored in a jump rope tall does the building need to be to accommodate this
bent 180 degrees with a constant radius of curvature average distribution of people? How tall would it need
of 1 m, assuming it can be modeled as an elastic to be if there were 10 people? And if there were 1000
cylindrical beam with a Young’s modulus of 100 MPa, people?
and a radius of 0.5 cm. Do the same for a strand of
5. Spring is in the air and a mosquito is in your apartment.
DNA with the same curvature. Again, assume that the
You are watching it fly and land on a set of stairs.
strand of DNA can be modeled as an elastic cylindrical
Assuming that it needs 1kBT of energy to fly up from
beam, assuming a Young’s modulus of 1.9 GPa and a
one step to the next, what would a Boltzmann’s
radius of 1 nm. How do these values compare to the
distribution predict for the probability of finding it on
characteristic transition energy?
any given step, if there are 10 steps in total?
2. Two dice are rolled, each of which can generate
6. Consider a molecule that has a “home” position and a
a number between 1 and 6. Consider this a
linear restoring force that is proportional to the distance
statistical system. How many microstates are there for
that it is away from this position. In other words, the
this system? How many macrostates are there? What is
molecule acts like it is attached to the tip of a spring.
the density of states, Ω, for each macrostate? What
The restoring force is F = −kx, where x is the distance
is the probability of each macrostate if the dice are
from the home position and k is the spring constant.
rolled many times?
The potential energy is a function of position and can be
Now suppose that the “fair” dice are substituted with expressed as
a pair of “loaded” dice. In the loaded dice there is a
small weight placed under the 1 spot. This makes the
1 2
probability of getting a 6 slightly higher than 1/6, and W = kx .
the probability of getting a 1 slightly lower than 1/6. 2
AnnotateD ReferenCes 149
We can treat each molecular position of the molecule, 8. Imagine the same box as in Problem 7, except the
x, as a microstate of the system. In this case, assuming partition is constructed differently. Specifically, it does
thermal equilibrium, the probability of finding the not slide; however, energy (for instance, heat) is allowed
molecule in any given position is given by the to transfer freely between the two compartments.
Boltzmann distribution. What is the average energy? How would you expect the temperatures of the two
Hint: a table of integral identities may be useful. gases to be related at equilibrium? Formalize this
Note: your finding is known as the principle of result by showing that entropy is maximized when the
equipartition of energy and holds for the energy temperatures of the two gases are equal.
associated with any parameter that has an associated
9. Consider a mechanosensitive ion channel that can be in
energy that varies with the square of the parameter. For
one of two states: open or closed. Define the energy of
example, it also works for kinetic energy, which varies
the channel using a configuration parameter σ that can
with the square of speed.
be equal to 0 (closed) or 1 (open). The energy is
7. In this problem, we show that for an isolated system,
equilibrium is attained by maximizing the entropy (this E = σε open + (1 − σ )ε closed − τ∆A
is the second law of thermodynamics and is equivalent where εopen and εclosed are the energies of the channel
to stating that equilibrium is attained by minimizing in the open and closed configurations, τ is the tension
the free energy, assuming constant energy). Imagine applied to the channel, and ΔA is the change in area
a box that is isolated from the external environment, of the channel in going from closed to open. Write
meaning that the total energy of the contents within expressions for the partition function, the probability of
the box is constant. The box contains a partition the channel being in an open state, and average energy.
that divides it into two subspaces. The partition is
What is the probability that the channel is open,
constructed such that it can slide freely along the length
assuming that τ = 1, 2, 3, 4, and 5 pN/nm, Δτ = −5kBT,
of the box, similar to a plunger in a syringe; however,
and ΔA = 10 nm2?
it does not allow energy to transfer between the two
subspaces. We now fill both sides of the partition with 10. Consider a two-dimensional model polymer composed
some amount of gas. The partition slides around until of n = 4 rigid segments, with neighboring segments
it comes to rest at its equilibrium position. How would connected via a rotating joint. Assume that the angle
you expect the pressures to be related at equilibrium? between adjacent links θ can only take discrete values:
Now formalize this intuition, by showing that entropy 0° (hairpin), 90°, or 180° (straight between links).
is maximized when the pressures are equal. Hint: we Calculate the value of the density of states for two
know that the total entropy of the system is different end-to-end lengths R = 2 and R = 4. Calculate
the entropy for these same values of R. Which value of
Stot = S1( N 1 ,V1 , E1 ) + S2 ( N 2 ,V2 , E2 ) R is entropically more favorable?
You will need the thermodynamic definition of pressure 11. Consider the “90° polymer” depicted in Figure 5.6. Write
to complete the derivation, a MatlaB module that analyzes a polymer made up of
N segments of unit length and which determines the
p ∂S number of configurations possible for each end-to-end
= . length between 0 and N.
T ∂V E ,N
Annotated References
Berg HC (1993) Random Walks in Biology. Princeton University Press. Pande VS (2006) Graduate Statistical Mechanics. (lecture notes
This book, first released in 1983 and expanded ten years later, is an from course number Chem 275, Stanford University, Stanford, CA.)
excellent source for examples of random walks in biology, including The general structure of this chapter is based on a one-semester
diffusion. course on graduate statistical mechanical developed by Dr Pande
at Stanford University. The treatments of specific topics in this chap-
Chandler D (1987) Introduction to Modern Statistical Mechanics.
ter, including enumeration of states via the three-coin system, and
Oxford University Press. A concise introduction to elementary sta-
analysis of the two-level noninteracting system via the microcanon-
tistical mechanics.
ical and canonical ensemble, are based on his course notes, which
Dogic Z, Zhang J, Lau AWC et al. (2004) Elongation and fluctuations are currently in development as a formal text.
of semiflexible polymers in a nematic solvent. Phys. Rev. Lett. 92,
Phillips R, Kondev J & Theriot J (2009) Physical Biology of the Cell.
125503. Reports the formation of hairpin defects in actin filaments
Garland Science. This is an excellent text that gives many examples
(similar to that in our model polymer) suspended in anisotropic
of insights that can be made into biological phenomena through
solutions of aligned rod-like macromolecules.
statistical mechanics. (Note: Problem 6 is adapted from the text of
McQuarrie DA (2000) Statistical Mechanics. University Science Books. Section 5.5.2 in Physical Biology of the Cell; Problem 7 is adapted
A widely used text in introductory statistical mechanics courses. from Section 5.5.2; Problem 8 is adapted from Section 7.1.2.)
This page intentionally left blank
to match pagination of print book
CHAPTER 6
Cell Mechanics in the
Laboratory
Nota Bene In magnetic micromanipulation, localized forces are applied to cells by adhering
ferrous beads, typically on the order of micrometers in diameter, to the cell sur-
Magnetic beads. The name
face, and then applying electromagnetic forces to the beads. In some cases, the
“magnetic beads” is commonly used
to refer to beads used in magnetic beads are introduced into the cell interior by endocytosis. An electromagnet can
micromanipulation experiments. The be constructed by winding wire around a metal core. Shaping the tip of the core to
beads themselves are not usually a chisel (typically, a few hundred micrometers across) results in a strong magnetic
magnetic; they are typically iron- field that is generally laterally homogeneous with a strength that depends only on
bearing beads. distance from the tip.
A representative system for magnetic micromanipulation is shown in Figure 6.1.
Cells that have been seeded in a dish and that have had ferrous beads bound to
their surface are placed on a temperature-controlled stage of an inverted micro-
scope. An electromagnet is held above the dish with the tip immersed in the media.
A cell with a bead attached is located under the microscope, and the tip of the mag-
net is positioned near it. Current is passed through the electromagnet, generating
a magnetic field that causes the bead to be pulled toward the magnet tip. The force
applied is determined by the distance from the bead to the magnet tip. Bead move-
ments are determined microscopically (Figure 6.2) with particle-tracking algo-
rithms (see Section 6.2). To study the behavior of specific molecules, so-called
functionalized (surface modification is explained at the end of Section 6.2.) beads
can be used that have been coated with the protein of interest, which can engage
specific receptors of interest on the cell surface.
Figure 6.2 The bead is the dark round object left of center, with the lines drawn in
to make it easier to visualize bead motion. The shadow on the right is the
electromagnet tip. Cells are visible in the background, but typically during experiments the
light is increased to white out the cells and obtain clearer tracking of the bead.
Nota Bene
Constant force is applied to the
magnetic beads. Because force is
where F is the force, μ is the dynamic viscosity, D is the bead diameter, and V is dependent on the distance from the
the terminal velocity. By performing this procedure on beads at various distances bead to the magnet tip, the force
from the electromagnet tip, one can estimate the bead force as a function of dis- applied to a cell-bound bead actually
tance. When using this approach, it is important that flow remain laminar, as increases as the bead is pulled in the
Stokes’ formula is only applicable at very low Reynolds numbers (see Section 4.2). direction in the magnet. In most
cases, the bead displacement is
small enough that the change in
Magnetic twisting and multiple-pole micromanipulators force is can be neglected.
can apply stresses to many cells simultaneously
Although magnetic micromanipulation is useful for investigating the mechanical
behavior of single cells, it is often desirable to apply mechanical loads to many
cells simultaneously. Magnetic twisting is an alternative method that uses mag-
netic beads to apply forces to many simultaneous beads. An electromagnet is
used to magnetize beads temporarily in one direction with a powerful pulse. This Nota Bene
is quickly followed by a pulse in a perpendicular direction, causing a twisting
movement to be applied to the beads to align them with the new field (Figure Magnetic bead types. In magnetic
micromanipulation experiments,
6.3). As the beads twist, their individual magnetic fields add together to produce having the bead lose its
a net magnetic field that is rotated from the original orientation. A magnetome- magnetization is desirable if repeated
ter, which is simply a wire coil attached to a sensitive amplifier, is used to meas- measurements are being taken. That
ure the magnitude of the rotated magnetic field. It is possible to obtain the is, you want to apply some force to
relationship between torque and rotation, in an average sense, for hundreds of the bead, release it, wait some time,
beads simultaneously. and then apply the same force. Here
you want to use paramagnetic beads,
which lose most or all of their
Optical traps generate forces on particles through transfer magnetization once the magnetic
of light momentum field breaks. A paramagnetic material
is also used to construct the
Light is another way to impart forces to cell-bound beads. Perhaps you recall that electromagnet so that you can
quickly control the applied magnetic
photons are massless, but nonetheless carry momentum. Light is able to impart
field. Materials that retain a
momentum to matter with which it interacts. This fundamental property of light is significant portion of the
exploited with high-power lasers to build optical traps (also termed optical magnetization are referred to as
tweezers). ferromagnetic. In general,
paramagnetic materials respond
To understand how an optical trap works, consider an transparent particle at the more weakly to magnetic fields, but
center of a focused laser beam (Figure 6.4). The index of refraction of the particle their ability to de-magnetize is better
is higher than the surrounding environment, such that as light passes through the than ferromagnetic materials.
particle, the beams change direction or bend. The particle is redirecting the light
beam and therefore changing the light’s momentum. By conservation of momen-
tum, there must be a force acting on the particle. The bending of the beams leads
to a force gradient that pushes the bead toward the focal point and into the center
objective
of the beam, where the intensity is highest. There is also a scattering force that
Nota Bene results in a force along the direction of the laser beam. The result of these com-
bined forces is that the particle becomes “trapped” within the focused laser beam
A few key figures of the optical
in a location just downstream of the focal point. If the bead is displaced relative to
trap. In the late 1960s, Arthur Ashkin
from Bell Labs pioneered the use of the beam, the forces will act in a restorative manner to move the bead back into
light to manipulate the position of the trapping location, creating a stable equilibrium point.
tiny particles of matter. Eventually
this developed into modern-day
optical trap systems. It is also the
Ray tracing elucidates the origin of restoring forces in
basis of the work of Steven Chu, who optical tweezers
used it to cool atoms by selectively
trapping slow-moving ones, allowing To quantify the origins of the lateral restoring forces, we can use a technique called
faster-moving ones to escape. Dr. Chu ray tracing. It is useful for determining such forces, as long as the particle is large
received a Novel Prize for Physics in compared with the wavelength of light. Consider a spherical bead trapped within
1997 and became United States a laser beam, as in Figure 6.5. When the beam is unfocused it resembles a cylin-
Secretary of Energy in the Obama der. We further assume it to have a constant intensity in the axial direction, but a
administration.
radially decreasing intensity profile in the lateral direction, with the highest inten-
sity in the center.
If we follow ray 1 in Figure 6.5, it enters the left side of the bead and is deflected
rightward. In this rightward deflection, there is transfer of momentum from the
beam to the bead that imparts a force on the bead that is primarily leftward (F1).
Ray 2 is located opposite ray 1 and equivalently imparts a primarily rightward
force (F2). Because ray 2 is equal in intensity to ray 1, F2 is equal in magnitude to
F1, resulting in no net lateral force, if the bead is centered. However, if we displace
the bead to the left, the intensity of ray 2 will be less than that of ray 1 owing to the
radial gradient of light intensity. In this case, the magnitude of F1 is greater than
that of F2, resulting in a net rightward restorative force that pushes the bead back
to the beam center. Because the forces are the result of the intensity gradient, they
are referred to as gradient forces.
Although the lateral gradient force example demonstrates why the bead becomes
trapped in the lateral direction, what about the axial direction? To understand this
phenomenon, consider a focused laser beam as in Figure 6.6. The beam forms a
cone that converges at the focal point. If we follow a ray entering the left side of a
spherical bead located upstream of the focal point, it is deflected rightward upon
entering the bead, and again upon exiting the bead. There is a change of momen-
tum as it enters and exits the bead, because the light entering the bead is not
purely vertical, the shift of the rays toward the centerline means that the exiting
PROBING THE MECHANICAL BEHAVIOR OF CELLS THROUGH CELLULAR MICROMANIpULATION 155
1 2 1 2
F1 F2 F1 F2
net net
force force
maximum
intensity
light has a larger vertical momentum component than the entering light. By sum-
ming the difference in momentum vectors, one can see that there is a net vertical
change in momentum, which is transferred to the bead as an axial force in the
downstream direction. A similar analysis reveals a net vertical change in momen-
tum in the upstream direction if the bead were downstream of the focal point,
resulting in an upstream restorative force. When the lateral and axial restoring
forces are combined with the downstream scattering force, the result is that the
bead becomes trapped just downstream of the focal point.
net
force
156 CHAPTER 6: Cell Mechanics in the Laboratory
n 2
Rb3 b − 1
n I nm
F1 = 2π m . (6.3)
c nb 2
n + 2
m
The gradient force is dependent on the gradient of the light intensity and not the
absolute intensity. For a typical experiment, the forces are relatively small (on the
order of tens of piconewtons, although some modern configurations can achieve
higher forces).
Recall that the generalized Lorenz–Mie theory and Raleigh scattering can be
used when the wavelength of the beam is much smaller and larger than the beam
diameter, respectively. Often the sizes of particles of interest may be of the same
order of magnitude as the wavelength. For example, green light has a wavelength
of 510 nm; most cellular targets of interest for trapping range from a few microm-
eters to 250 nm. Particles much smaller than this are very hard to trap, because
they are difficult to image using standard optical techniques and generate much
smaller forces. Therefore for many particle sizes and beam wavelengths, the rela-
tionship between force and particle/beam properties is not very well under-
stood. In this case, forces generated by the trap can be determined empirically.
One approach is to capture a bead of known size, and move the trap at increasing
velocity until the capture is lost. Quantifying the fastest velocity before the bead
is lost, allows the maximum force generated by the trap to be calculated using
Stokes’ law.
that they can be manipulated directly without the need for beads. However, for
some biomechanical studies, optical trapping may not be the best method, as the
forces are generally low. Additionally, force-based manipulation with optical trap-
ping requires feedback technology, because forces are generated based on the rela-
tive position of the bead and the laser beam. Another important difference is that an
optical trap, as the name suggests, is a “trap.” That is, there is a relatively small equi-
librium point and a large gradient of restoration forces around that point. The forces Nota Bene
generated magnetically affect a far greater spatial region and are more homogene-
ous. This allows many beads to be manipulated together, although the manipula- The technical challenges of
bead-based studies. Although
tions are not independent. Finally, it is not clear whether magnetic fields have bead-based studies can lead to
effects on biological tissues, but it is well known that focused lasers can heat biologi- powerful insights, there are important
cal material, and with sufficient power can destroy tissue locally. In fact, this effect technical challenges and concerns
has been exploited to generate what are called laser scissors or laser scalpels, allow- that need to be kept in mind. For
ing one to disrupt parts of a cell selectively. instance, there can be significant
variability in bead–cell adhesiveness.
Often, bead–cell adhesion tends to
Atomic force microscopy involves the direct probing of objects improve slowly over time with the
with a small cantilever bead eventually being internalized to
the cell. Therefore, it may be difficult
Atomic force microscopy (AFM) is an approach for characterizing cells that allows to distinguish the mechanical
for the direct probing of the cells, using a small cantilever. There is no need to behavior of the cell from the
differences in adhesion. Indeed, the
attach a bead or other foreign substance to the cell surface. AFM works by tracking
very act of adhesion may induce an
the displacements of a cantilevered beam brought into contact with the cell or unexpected biological response.
other object of interest. Modern designs are able to measure both the shape of a
cell and its mechanical properties. If one were to imagine a finger as an AFM can-
tilever, it is easy to see how a range of physical attributes of an unknown object
could be inferred, such as mechanical stiffness (by pushing down on the object
and determining how much resistance the object offers), surface geometry (by
running one’s finger over the surface), and spatial heterogeneity in mechanical
stiffness (by probing or poking in several places).
The central component of an atomic force microscope is a cantilever that has a
small tip at the free end. The cantilever and tip are typically constructed of silicon
or a silicon compound. A typical tip has a radius of curvature on the order of
nanometers, which is quite sharp, even at the scale of a cell. In probing a speci-
men, the cantilever is lowered so that the tip contacts the cell surface. As the tip
pushes into the specimen, the cantilever deflects. We saw in Section 3.2 (see
Equation 3.31) that for the case of a linear elastic beam of length L clamped at one
end, EI is the flexural rigidity, the displacement w of the cantilever at any point x
along the beam is a function of the force at the tip, F,
F x3 x2
w= − L . (6.4)
EI 6 2
We can see from Equation 6.4 that the deflection at the tip (x = L) is propor-
tional to the tip force. The displacement at the tip can be related to the force
through an effective spring constant. Once this constant is determined, the
force being applied at the cantilever tip can be readily calculated from the tip
deflection, allowing for simultaneous quantification of deflection and force. In
this way, force-displacement curves can be generated by positioning the tip
over the surface of the specimen and measuring the deflection of the cantilever
as the tip is lowered into the specimen (this mode of AFM use is sometimes
referred to as force readout mode). Such curves are conceptually similar to the
force-deformation curves we discussed in Section 3.2 (see Figure 3.1) and can
be used to infer mechanical properties.
Nota Bene a reflective area on the back of the cantilever and the lateral motion of the beam is
measured a large distance away (Figure 6.7). A detector is placed in the path of the
Hertz contact. Heinrich Hertz reflected beam, such that a laser spot is generated on the detector. As the cantilever
introduced the solution for
bends, the laser spot on the detector changes its location. In general, the longer the
frictionless contact between two
spheres in 1881 by assuming that the laser beam path, the more the laser spot will move for a given deflection of the can-
contact force could be represented tilever, and the smaller the deflections are that can be detected. Generally, force
by an integral of point forces applied sensitivities using AFM are on the order of piconewtons or less.
over a small patch or region of
contact. With this solution, the
contact between a sphere and a Scanning and tapping modes can be used to obtain cellular
plane was determined by letting the topography
radius of one of the spheres
approach infinity. Similar approaches In addition to characterizing mechanical properties and behavior, AFM can be
can be used to determine the contact used for specimen visualization, in particular for determining surface topogra-
force due to a conical or cylindrical phy. There are several modes whereby this can be achieved. Scanning mode is
indenter.
used to generate a topographical profile of the surface being examined. Here, the
tip is lowered until it just touches the surface of the specimen, then the tip is
scanned, or dragged, across the specimen. The deflection of the cantilever at each
point is used to generate a height profile. To prevent specimen damage, in many
cases a feedback mechanism is incorporated to lower/raise the tip to maintain a
constant force. If the material is homogeneous, then the deflection is g enerally a
good representation of the surface geometry. For specimens with heterogeneous
mechanical properties, the topography may be reflective of the surface topogra-
phy as well as the local stiffness. If we were to consider a specimen that had a flat
surface but spatially heterogeneous stiffness, a stiffer region may appear “higher”
than a more compliant region.
A second AFM mode is tapping mode, whereby the tip is actively vibrated up and
down (typically by a piezoelectric actuator coupled to the AFM tip holder) close
to its resonant frequency, which is usually on the order of 10–100 kHz. The oscil-
lating tip is lowered to the specimen surface until it begins to experience van der
Waals forces, resulting in a reduction of the magnitude of the vibration. Similar to
Nota Bene scanning mode, during tapping mode operation, feedback is typically used to
Half-space. A half-space is a
maintain the cantilever oscillation amplitude constant. A major advantage of
mathematical abstraction. It is refers tapping mode over scanning mode is that when the tip contacts the surface, it
to region of space that is infinitely has sufficient oscillation amplitude to overcome “sticking” or adhesion forces
deep and wide, but has a flat surface. between the tip and the sample. It also reduces the chances that a tip will become
It is a reasonably good assumption fouled or coated in cellular debris because, ideally, the tip never comes into
for a cell, because on the scale of an physical contact with the cell. One important consideration with AFM is the
AFM tip, the surface of a cell is nature of the probes, which are sometimes unable to detect overhangs or very
virtually infinite.
steep walls.
Example 6.1: Estimating the elastic modulus of a cell using the Hertz model
Using the Hertz model, estimate the elastic modulus of a you can only obtain about two significant figures) yields
cell using υ = 0.5 (incompressible) and υ = 0.3 (typical of 75 kPa for incompressible materials and 90 kPa for bio-
many biological tissues). The radius of the probe tip is logical tissues. You can see that the choice of Poisson’s
50 nm. ratio does not make a huge difference in our estimate,
which is admittedly on the high side (cells typically in
Using Figure 6.9, we estimate the force for a 100 nm
the 10–30 kPa range by AFM) if you are not targeting
indentation (from 175 nm to 75 nm along the tip
stress fibers.
approach line) to be 1000 pN = 1 nN. Using the formula
E = 100,000(1 − υν) (actually a bit more, but realistically
200 300 400 to the left shows increasing force to deform the
rupture force (pN) cell, such that at 1 nN, the tip has advanced about
1000 75 nm into the cell. The darker line is the retraction
of the tip. At about 175 nm, the adhesion of the tip
to the cell is broken and there is a sudden
0 “zeroing” of the force. The force magnitude was
collected for a number of cells, and the histogram
of that adhesive-rupturing force is shown in the
−1000
inset. (From, Liu J, Weller GE and Zern B (2010)
0 100 200 300 Proc. Natl. Acad. Sci. USA, 107).
distance (nm)
160 CHAPTER 6: Cell Mechanics in the Laboratory
loading. The images contain unique local patterns of intensity that reflect the
shape and brightness of the fluorescent objects within that locale. These local
intensity patterns can be indentified with image correlation, and changes in posi-
tion and/or shape used infer the deformations that gave rise to these changes.
In an ideal scenario, a bead will be a perfect sphere, which appears as a perfect
circle on our image. We can then locate the center of the sphere, and thus the
bead, exactly. If we repeat the same measurement after the bead moves, we can
obtain the exact displacement vector of the bead. In the laboratory, however,
beads are not perfect spheres and the images will be limited by resolution, illumi-
nation, refraction, and other artifacts. Cross-correlation allows us to determine
locations, and therefore displacements, with high precision by using a mathemat-
ical algorithm to fit two profiles against each other.
To illustrate this approach, consider a cell seeded with k fluorescent beads (Figure
6.11). We obtain fluorescence images of each bead before, during, and after load-
ing. Each image can be represented by an array of numerical intensities that we
assume, for simplicity, are stored in a square array known as the image intensity
function, I(i, j) with 2n + 1 pixels in each direction such that
− n ≤ i , j ≤ n. (6.6)
Nota Bene
brightest pixel can jump from place to place dramatically. To address this prob-
lem, a statistical image correlation technique can be applied that is actually so
Image distortion. Although the lack precise it can determine the location of the bead with sub-pixel resolution.
of shape change in the intensity
distribution simplifies our analysis, To determine the translation in the intensity pattern between images, we first
methods exist for correlating intensity define a template image K(i, j). The template image is the ideal intensity pattern of
patterns even when they become a bead, which for simplicity we locate at (0, 0). This template could be obtained
distorted.
from the theoretical distribution for an ideal bead, or an actual bead could be
imaged before deformation. In either case, we assume that K(i, j) is 2m + 1 pixels
in each dimension, with m < n. With K defined, a statistical cross-correlation can
Nota Bene
be calculated as a measure of the degree of similarity between K and I on a pixel-
Image texture correlation. An by-pixel basis. A typical statistical correlation measure is
alternative approach for obtaining
discrete displacements that does not m m
require the use of markers such as
beads is image texture correlation.
C (i , j ) = ∑ ∑ I (i + x , j + y)K (x , y),
x =− m y =− m
(6.7)
This approach relies on the matching
of local image texture, in other words
the unique spatial variation in which is a matrix convolution of I and K. The resulting matrix, C(i, j) is known as
intensity surrounding a pixel of the cross-correlation field. It gives the degree of similarity between the intensity
interest, between an undeformed pattern surrounding the pixel (i, j) and K. The i and j that maximize C(i, j) give the
and deformed image. In particular, location at which the intensity pattern of I most closely matches that of K. This
pixels from a subset domain within location is assumed to be the “correct” location of the bead in I. If the bead is
the undeformed image are selected,
originally at (0, 0), then the distance of the bead displacement can then be calcu-
and used as the template for a
cross-correlation calculation within lated from i and j. The location of this peak in C can be computed with an accu-
the deformed image. If the deformed racy better than the size of a pixel by calculating the centroid of the
image is distorted, a transformation cross-correlation field,
(such as a stretch or shear) may need
to be applied to the template before
performing the cross-correlation ∑iC(i , j)
i,j
∑ jC(i , j)
i,j
calculation. The displacement is again ic = , jc = . (6.8)
the one that results in a maximum
in the cross-correlation.
∑C(i , j)
i,j
∑C(i , j)
i,j
MEASUREMENT OF FORCES pRODUCED BY CELLS 163
(B) (C)
1.75 × 10–3
1.00 × 106
164 CHAPTER 6: Cell Mechanics in the Laboratory
In the above expression, rx and ry are the x and y component of r and r is | r |. E and
ν are Young’s modulus and Poisson’s ratio of the elastic substrate, respectively.
Computing the above expression for all n displacements gives a simultaneous
equation between m force vectors and n displacement vectors. The goal then
becomes to find the m force vectors that give rise to the n displacement vectors
(Figure 6.13). Although methods for solving the problem are outside the scope of
the text, there are nice reviews in the literature.
forces are active—all of unknown magnitudes, being convolutions in frequency space become multiplica-
applied at unknown locations, being exerted in tions. Fourier transforms should be used with caution,
unknown directions. Consider a bead exposed to two as boundary conditions in frequency space are not
different unknown forces. If the bead moves, it could be clear-cut. Typically, there are no traction forces outside
because of either force or both forces acting in different the cell, but many solutions using Fourier transforms
amounts. To constrain the problem, multiple beads are will generate forces near the border of, but typically out-
required. However, the problem then becomes mathe- side of, the cell.
matically cumbersome because each bead gets an equa-
tion u = GF for each force being applied. Furthermore, d
(A)
is generally unknown, because the force locations are
not proscribed.
bead
To address this there are sometimes logical places to
assume forces are being applied, such as points of cel-
lular adhesion that can be determined by immunocyto-
chemistry. A more general approach is to create a
regularly spaced grid for the locations of force applica-
tion. If the grid has high enough density, a fairly com-
plete picture of the force field can be obtained.
Mathematically, for each grid location a displacement u (B)
and a force F is known. From this one can construct G.
This approach involves large arrays of displacements,
positions, and components of G, for which optimization
u F
routines are applied to extract the traction forces. This
technique is particularly sensitive to small errors in dis- d
placement because the force–displacement relation-
ship drops rapidly with distance away from the location
of the force. Another method is to use Fourier trans-
forms to convert the problem into frequency space. This Figure 6.15 A single force and single displacement
allows the inverse problem to be solved directly, because example. The force and displacement are co-linear.
F x3 x2
w= − L . (6.11)
EI 6 2
166 CHAPTER 6: Cell Mechanics in the Laboratory
Because we only measure the displacement at the tip δ at x = L, the above expres-
sion can be simplified to
3EI
F = 3 δ. (6.12)
L
However, cells typically do not encounter glass, plastic, or silicon in vivo. How the
cells interact with these artificial surfaces is an important consideration. The pro-
cess of surface modification can be a powerful way to modify the artificial material
not only to make the material more physiological or biocompatible, but also to gain
insight into the role of various properties or even specific proteins and molecules.
Techniques are available to alter roughness, charge, surface energy, hydrophobic-
ity, and other physical properties. A particularly useful approach is to functionalize
a surface, which is the process of coating it with functional groups or whole pro-
teins. These can be extracellular matrix proteins (such as fibronectin or collagen)
but they can also be protein fragments, peptides, or antibodies against epitopes of
a particular functional interest. Because the cell’s interaction with the extracellular
matrix can be highly dynamic and time-dependent, selecting the right incubation
time is critical. For small beads, a time interval 15 minutes and 1 hour is often used.
If the incubation time is too brief, the beads will not have sufficient time to bind;
too long, and the cells can endocytose (engulf) the beads, causing loss in receptor
specificity.
This process of functionalization can be combined with the photolithographic
microfabrication techniques described above. Microcontact printing or micropat-
terning involves using a soft polymer such as PDMS (polydimethylsiloxane) to Nota Bene
“print” a protein onto a surface with a microscopic pattern. A silicon master is first A functional group. A functional
created by coating a chip with a photoresist that can be removed optically on a group is the region or moiety of a
microscopic scale. The resulting pattern is then acid-etched into the silicon. protein that is responsible for its
Multiple PDMS stamps can then be created from the master. The stamp is then particular biochemical characteristic
or activity.
dipped into a protein or other solution and pressed onto glass or another surface.
Microfabrication techniques are used in “lab-on-a-chip” technologies using min-
iature channels, pumps, valves, etc. The small scale of these devices provides
enormous advantages in terms of small sample size, high throughput, accelerated
assay time, and even portability.
flow separates
low-
pressure
region
Nota Bene
Let us look at an approximation that can be used for a quick estimate of the Dynamic versus kinematic
flow device dimensions. Consider a system using cell culture medium as the viscosity. Viscosity is sometimes a
fluid (ρ = 1000 kg/m3, μ = 1 × 10-3 kg/ms). The Reynolds number is then Vh/ confusing term, as the concept of
(1 × 10-6 m2/s), where V is some characteristic velocity and h is the height of viscous resistance is associated with
the flow chamber. For Re < 1, we then need Vh < 1 × 10-6 m2/s. Next, consider both dashpots (which we will cover in
the shear stress, τ = μ du/dy. We can approximate this as τ ~ μV/h, so long as Section 6.4) and fluids, but these terms
have different units and definitions. In
the width is large compared with h. A typical shear stress for physiological this text, we refer to dashpots as
modeling of endothelial cells is about 1 Pa. So, V/h = 1000 s-1. Assume we having a viscous friction coefficient
express V in m/s and h in m to keep units simple, then Vh < 10-6 and V = 1000 h. whereas fluids have viscosity.
This means, h2 < 10-9 m, or h < 3 × 10-5 m = 30 μm, which is just about accept- In fluid mechanics, there is a further
able, seeing as cells are typically at least a few micrometers (10-6 m) large. distinction between the dynamic
Although a chamber this short can be difficult to manufacture, generally, the viscosity (which is viscosity that is
defined in Section 4.2, represented by
chamber is designed with the shortest height possible. Calculations based on
μ, and has units of kg/ms) and the
actual flow profiles and relaxing the Reynolds number requirement (to 10–100) kinematic viscosity, which is the
can improve this range somewhat. dynamic viscosity normalized by the
density of the fluid (υ = μ/ρ, with units
of m2/s). The reason for using the
Fully developed flow occurs past the entrance length kinematic viscosity is that it simplifies
some calculations (such as Reynolds
Another consideration of shear flow devices is the entrance length. Laminar flow
number), can be written with three
profiles require space to develop, and the entrance length is the size of this terms instead of four, and represents
region. It changes depending on the velocity and dimensions of the chamber. For a useful ratio that eliminates density
a circular pipe, the entrance length is about 0.06 times the Reynolds number (mass, specifically) from comparisons.
multiplied by the diameter, 0.06 × Re × d. When locating cells to examine, it is That is, denser fluids might appear
best to avoid the entrance length to avoid unexpected or unpredictable shear more viscous because of inertial
stresses. The cross-wise flow profile (perpendicular to both the height and the effects. For this chapter, viscosity will
refer to dynamic viscosity.
flow direction) will also develop over space; this is generally not desired, as the
170 CHAPTER 6: Cell Mechanics in the Laboratory
EC monolayer
EC
air, 5% CO2
microscope
reservoir
step
a b c d
known as Couette flow. A treadmill used in lieu of the upper plate of a parallel
plate chamber will generate a linear flow profile, which mitigates problems with
entrance lengths and creates a simpler relationship between the moving wall
velocity and shear stress.
fixed plate
F = kx , (6.13)
with the spring constant k a measure of the stiffness of the element. Dashpots are
viscous damper elements that obey Newton’s fluid law
F = ηv , (6.14)
174 CHAPTER 6: Cell Mechanics in the Laboratory
0.4
0.2
0.0
0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50
time (s)
ANALYSIS OF DEFORMATION 175
0.5
0.0
0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50
time (s)
The total displacement is the sum of the displacement of the single dashpot and
the Kelvin–Voigt body, x(t) = x1 + x2. Thus, to obtain x(t), we can compute x1(t)
and x2(t) individually, and then add them together to get the displacement of the
entire body. For the single dashpot, we know that x1 = (F/η1)t. For the Kelvin–Voigt
body the total force across the body is equal to the sum of the force within the
spring and dashpot
dx 2
F = kx 2 + η . (6.16)
dt
F − kt
x 2 (t ) = . (6.17)
x 1 − e 2
η
x2 x1
176 CHAPTER 6: Cell Mechanics in the Laboratory
displacement
the recoil is not to zero producing 1.5
residual deformation or strain. This
response profile to a step increase and
decrease in force is a reasonable 1.0
approximation of real-cell response to
magnetic micromanipulation.
0.5
0.0
0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50
time (s)
This can be verified by direct substitution, with initial condition x(0) = 0. Summing
the displacements of the single dashpot and the Kelvin–Voigt body,
)
Nota Bene
F F
x (t ) = t +
− kt
Lumped parameter models. The η . (6.18)
η1 k 1 − e 2
models of the type we have been
discussing are sometimes called
rheological models, because they The behavior of this model is depicted in Figure 6.25. It is very useful, because it
originated from the analysis of is the simplest model that captures both viscoelastic behavior as well as plastic
viscous fluids (see Section 4.4). They
behavior (permanent deformation). To understand this phenomenon, consider
are also sometimes referred to as
lumped parameter models, because what happens as t goes to infinity. The first term (the displacement of the single
they collect spatially distributed dashpot) increases without bound at a constant rate of F/η1. The second term (the
behaviors and lump them into simple displacement of the Kelvin–Voigt body) gives a displacement at t = infinity of F/k.
discrete elements. For the lumped For finite time, the Kelvin–Voigt body never reaches F/k, but approaches it asymp-
parameter model in Figure 6.25, all totically. Finally, it is worthwhile noting that the exponential term, −kt/η2, is
the elastic behavior was captured in sometimes written as −t/τ, where τ = η2/k is called the time constant. The time
a single equivalent spring stiffness
constant is a measure of how long it takes the exponentially decaying quantity in
parameter (k).
the parentheses to decrease by 1/e, or about 63%, of the initial amplitude.
objective lens
T=0
Flow
10 µm
T = TP T = Te
and
v ( x , y ) = v a x + v b y + vc , (6.20)
where ua, ub, uc, va, vb, and vc are unknown constants. To solve for ua, ub, and uc,
we form the set of simultaneous equations
ua x1 + ub y1 + uc = u1
ua x 2 + ub y 2 + uc = u2 (6.21)
ua x 3 + ub y 3 + uc = u3 ,
which results in three equations and three unknowns. The constants va, vb, and vc
can be solved for using a similar approach. We segment the image into triangles,
at which the vertices are locations for which we have displacement data. Then, to
estimate the displacement at an arbitrary point, we select the three vertices mak-
ing up the triangle that encloses the point.
du
ε xx =
dx
dv
ε yy = (6.22)
dy
1 du dv
ε xy = + .
2 dy dx
180 CHAPTER 6: Cell Mechanics in the Laboratory
Nota Bene
For large deformations, the Green–Lagrange strains can be computed as
dx 2 = dXCdX , (6.24)
U ν−I νλ = 0, U ν= λ ν
BLINDING AND CONTROLS 181
for each λ. We find that the eigenvectors are (1,1) (corre- Typically, the vectors would be expressed in unit form,
sponding to λ = 3) and (1, − 1) (corresponding to λ = 2). but we do not do so here to avoid the use of square roots.
Note that when you are solving for the eigenvectors, you Our result is also consistent with our findings in Example
will not generally be sufficiently constrained to get the 3.6 in that the principal stretch directions are the same as
exact values; so you can guess a value for one component. the principal strain directions.
However, our presence might also influence the participants, because we know
who received the placebo and who received the vitamins. Therefore, we may act—
perhaps unconsciously—more pleasantly toward those who received vitamins. To
avoid this error, we would have an assistant interview the participants about their
headaches; this assistant could not know who received which substance. The
patients and their treatments would be encoded so that nobody involved in the
experiment knows who received what. This approach is a double-blind design,
and is the most rigorous method for uncovering true influences.
Sometimes blinding does not work or is impractical. If we were to perform an
operative procedure to relieve pain, the people receiving the operation would
clearly know they had it, compared with patients who were left untreated. To
account for this, sometimes people use what is called a sham. This is a special
control whereby a procedure is mimicked, but a key therapeutic step is omitted.
So in this situation, the control group would have an operation, but without
whatever manipulation was thought to be analgesic. A sham surgery raises
important ethical concerns in human studies, but is a common strategy for ani-
mal investigations.
The use of controls and blinding applies to many of the techniques that are cov-
ered in this chapter. For instance, if you perform an experiment using magnetic
twisting by attaching magnetic beads to cells and then applying a torque, you can-
not simply obtain a readout (stiffness, upregulation of some gene, activation of
some channel, etc.) and decide whether the cells are responding or not. You need
to compare the readout to cells that do not have a similarly applied torque.
However, the addition of the beads may have caused some changes in cell base-
line. The better control is to have a sham in which you add beads but do not torque
the beads. Some investigators will go so far as to complete the sham by adding
beads to cells, placing the cells in the magnetic twisting device, and turning on the
equipment, but will not actually apply the torque. Alternatively, the experiment
can be performed with nonmagnetic beads. In this way, vibrations from the
equipment, stresses and temperature changes from handling and placing the dish
on the apparatus, the magnetic field itself, etc., are all eliminated as potential con-
founding influences.
Key Concepts
• A variety of techniques allow us to measure cell • Beads embedded in a flexible substrate allow
forces and manipulate cells. substrate deformations to be determined. An inverse
• There are various modes of magnetic problem is solved to obtain the applied forces.
micromanipulation (pulling vs twisting). • Cells have both solid and viscous behavior. Models
• Optical traps work by deflection of light and offer that incorporate both behaviors have been developed
contrast to magnetic methods by manipulating bead to describe cellular deformation.
displacement rather than force. Magnetic forces can • There are diverse methods for applying fluid shear
be applied to a large number of beads stress to cells. Consideration must be given not only
simultaneously. to maintaining biological compatibility, but to control
• In atomic force microscopy cells can be probed by of turbulence, entrance length, and flow development.
application of a cantilevered tip against the cell. The • Stretched samples in which strain levels are large can
deflection of the cantilever can be related to the force be analyzed by deformation gradient mathematics. Data
being applied. are generally acquired at discrete points, which are then
• Microfabrication is useful for assaying small numbers interpolated to yield displacement at desired locations.
of cells. Micropillars can be used to measure cell– • Experiments are most useful when only one factor, or
substrate forces based on the deflection of the pillars, one set of factors, is being changed at a time. Use of
which are modeled as beams. controls, blinding, and shams can help ensure the
• Traction-force microscopy is another way to measure readout is most representative of the response to the
cell–substrate forces. factor(s) being examined.
PROBLEMS 183
Problems
1. Show that a series of n springs in parallel, with spring where the value in the ith row and jth column is the
constants k1, k2, . . ., kn, where ki does not necessarily intensity for pixel (i,j).
equal kj for i ≠ j with a single dashpot of viscous
coefficient η can be reduced to a single spring and You have determined that the template image (K)
dashpot in parallel. should be
cell undergoing deformation described by F, and diameter as the membrane. As the piston moves up
that λ2 is the minimum stretch under the same and down, the media in which the cells are immersed
conditions. will also flow back and forth across the cells. Assume
the media has viscosity μ and density ρ and that the
10. Use dimensional analysis to obtain an expression for cells are linearly elastic with elastic modulus 10 kPa
the drag force on a sphere under laminar flow. You must undergoing 5% stretch. Determine the ratio of stretch-
identify the parameters that govern the force and then based stresses to that of fluid shear stresses. You may
obtain a dimensionless expression. Next, show that model the cells as 50 μm × 50 μm × 5 μm rectangular
the obtained expression is consistent with Stokes’ law solids with the smallest dimension being the height. An
(Equation 6.1). order of magnitude ratio is sufficient.
11. Consider the case in which you are devising a shear- 13. Unlike the rounded tips described in this chapter,
based technique to estimate the adhesion force of cells atomic force microscopes may also be equipped with
to a substrate. You have an inlet into a cylindrically sharp tips that are typically pyramidal in shape (for
symmetric outflow as shown below. example, the Berkovich tip). Owing to their geometry,
these tips have modified relationships between the
Q force, E, the indentation depth, and the geometry of
the tip as given by the Hertz model. To model AFM
with a pyramidal tip, it can be useful to use a scaling
relationship instead of trying to get an exact solution.
r (a) First, derive a scaling relationship for the force and
indentation depth of a spherical tip atomic force
microscope. You may assume the radius of the bead
h (R) is very large compared with the radius of the
area being indented (a), and that the radius of the
area being indented is very large compared with
The dashed line represents the cell monolayer. The flow
the indentation depth (δ). Further, you may assume
into the inlet is at volume flow rate Q (m3/s) across
only the linear elastic modulus applies (E). To solve
cross-sectional area A. The height of the outflow region
this, use an energy balance by calculating the strain
is h, and you may assume that h < r and that the flow
energy of the deformation and equating that with
is laminar. At the inlet, you pump fluid at a constant Q,
the work done by pushing the tip in. Because this is
and at the outlet, collect all the cells that detach and
a scaling relation, you may ignore constants.
exit the outflow region around the perimeter. You then
(b) Now, assume the tip is a pyramid instead. Assume
count the cells that have detached. Relate the number
the angle of the tip is α, the indentation depth is
of detached cells to the maximum adhesion force of
δ, and the cell has elastic modulus E (continue to
the cell, assuming that the adhesion force per cell is
ignore shear effects). You may find it easier to start
uniform, that cell–cell interactions can be neglected,
by assuming the tip is actually a cylinder with a
and that each cell is 10 μm in radius.
rounded end with radius a (and assuming that δ is
12. Consider the case in which you are given a stretch much less than a), and then realize that the angle α
device consisting of a membrane that is 10 cm in is related to the “width” of the indenter. Once you
diameter. The membrane is stretched by a piston derive the scaling relationship, you should see that
pushing on its bottom surface, as depicted in Figure there is a mathematical difference between the
6.21. You may assume the piston is nearly the same two types of indentation.
Annotated References
Alessandrini A & Facci P (2005) AFM: a versatile tool in biophysics. Chen CS, Mrksich M, Huang S, Whitesides GM & Ingber DE (1997)
Meas. Sci. Technol. 16, R65–R92. This journal review article summa- Geometric control of cell life and death. Science 276, 1425–1428.
rizes many aspects of biological studies using AFM. This paper describes the use of micropatterning to determine
whether adhesion area or spreading area is more important for
Brown TD (2000) Techniques for mechanical stimulation of cells
maintaining cell viability.
in vitro: a review. J. Biomech. 33, 3–14. This journal review article
describes several different methods for applying stretching and Chien S (2008) Effects of disturbed flow on endothelial cells. Ann.
shear to cells and tissues. Biomed. Eng. 36, 554–562. This summarizes findings of how shear
flow affects endothelial biology in a way that implicates disturbed
Cao J, Donell B, Deaver DR, et al. (1998) In vitro side-view imag-
flows in the development of atherosclerosis.
ing technique and analysis of human T-leukemic cell adhesion
to ICAM-1 in shear flow. Microvasc. Res. 55, 124–37. This paper Goubko CA & Cao X (2009) Patterning multiple cell types in co-
describes the use of side-view imaging to obtain quasi-3D images cultures: a review. Mat. Sci. Eng. C. 29, 1855–1868. This review
of cells during rolling adhesion. article briefly summarizes some basic, micropatterning and then
ANNOTATED REFERENCES 185
describes the more recent pattern generation for use in multiple sham surgery without insertion of the scope. The patients were
cell patterns. kept unaware of whether they had the surgery. Interestingly, the
control group experienced as much improvement as the surgery
Hoffman BD & Crocker JC (2009) Cell mechanics: dissecting the group.
physical responses of cells to force. Annu. Rev. Biomed. Eng. 11,
259–288. This paper reviews rheological properties of cells with dis- Munevar S, Wang Y & Dembo M (2001) Traction force microscopy of
cussion of various techniques and what the results from different migrating normal and H-ras transformed 3T3 fibroblasts. Biophys J.
techniques may have in common. 80 4, 1744–1757. This article describes traction force microscopy
results applied to cell analysis.
Huang H, Kamm RD & Lee RT (2004) Cell mechanics and mecha-
notransduction: pathways, probes, and physiology. Am. J. Physiol. Neuman KC & Block SM (2004) Optical trapping. Rev. Sci. Instrum. 75,
Cell Physiol. 287, C1–C11.This journal review article outlines a 2787–2809. This paper describes the state of optical trapping and
few experimental methods for studying cell mechanics and some the necessary components for achieving it.
underlying bases for the various approaches taken.
Roy P, Rajfur Z, Pomorski P, Jacobson K (2002) Microscope-based
Janmey PA & McCulloch CA (2007) Cell mechanics: integrating techniques to study cell adhesion and migration. Nat. Cell Biol. 4,
cell responses to mechanical stimuli. Annu. Rev. Biomed. Eng. 9, E91–6. This journal review article discusses several microscopy-
1–34. This journal review article that discusses mostly theoretical based techniques to characterize cell adhesion structures and
and sensory mechanisms for understanding cell mechanics and forces, including FRET and traction force microscopy.
mechanotransduction, but is also notable for presenting a nice Sotoudeh M, Jalali S, Usami S, Shyy JY, Chien S (1998) A strain device
summary in Table 1 that compares the elastic modulus by different imposing dynamic and uniform equi-biaxial strain to cultured cells.
techniques and cells with extensive references. Ann. Biomed Eng. 2, 181–189. This article discusses the design and
Kuo SC (2001) Using optics to measure biological forces and implementation of a stretch device used for cell mechanics studies,
mechanics. Traffic 2, 757–763. This journal review article summa- with validation of the exerted strains.
rizes some key findings for using optical traps and microrheology to Tan JL, Tien J, Pirone DM, et al. (2003) Cells lying on a bed of
study biological systems. microneedles: an approach to isolate mechanical force. Proc. Natl.
Landau LD & Lifshitz EM (1970) Theory of Elasticity, vol. 7, 2nd ed. Acad. Sci. USA 100, 1484–1489. This article describes the use of
Pergamon Press. This book, part of the Course of Theoretical Phys- micropillars (they call it microneedles) to determine traction forces
ics, contains many formulae that are of relevance when trying to using beam-bending analysis.
characterize deformations in solid objects. It does have newer edi- Wang JH & Lin JS (2007) Cell traction force and measurement
tions and the other books in the series cover different topics. methods. Biomech. Model. Mechanobiol. 6, 361–371. This paper
Legant WR, Miller JS, Blakely BL et al. (2010) Measurement of describes traction-force microscopy and a few different techniques
mechanical tractions exerted by cells in three-dimensional matri- for analysis of the results to extract the traction field.
ces. Nat. Meth. 7, 969–971. This is a recent journal paper that Wang N, Tolić-Norrelykke IM, Chen J et al. (2002) Cell prestress. I.
describes an approach for extending traction force microscope stiffness and prestress are closely associated in adherent contrac-
from a two-dimensional monolayer to cells imbedded in a three- tile cells. Am. J. Physiol. Cell Physiol. 282, C606–C616. This journal
dimensional hydrogel. article characterizes cell structures (including those based on the
tensegrity hypothesis) using information from a variety of tech-
Liu J, Weller GE, Zern B, Ayyaswamy PS, Eckmann DM, Muzykantov
niques, including traction force and magnetic twisting.
VR, Radhakrishnan R (2010) Computational model for nanocarrier
binding to endothelium validated using in vivo, in vitro and atomic Weibel DB, Diluzio WR & Whitesides GM (2007) Microfabrication
force microscopy experiments. Proc. Natl. Acad. Sci. USA 38, 16530– meets microbiology. Nat. Rev. Microbiol. 5, 209–218. This journal
16535. This article uses computational methods to calculate the review article summarizes different micropatterning techniques
free energy of binding a functionalized “nanocarriers” to endothe- and their applications to biological systems.
lial cells.
Young EW & Simmons CA (2010) Macro- and microscale fluid flow
Moseley JB, O’Malley K, Petersen NJ et al. (2002) A controlled trial of systems for endothelial cell biology. Lab Chip 10, 143–160. This
arthroscopic surgery for osteoarthritis of the knee. N. Engl. J. Med. paper discusses shear flow design and analysis and presents prin-
347, 81–88. This is a journal article describes a study in humans ciples for shear flow design at the microscale by taking keys from
of arthroscopic knee surgery in which the controls received a macroscale shear flow devices.
This page intentionally left blank
to match pagination of print book
PART II: praCtiCES
This page intentionally left blank
to match pagination of print book
CHAPTER 7
Mechanics of Cellular
Polymers
W e now turn our attention to one of the most important structural compo-
nents of cells: biopolymers. In Section 2.1, we learned that polymers are
linear molecules made up of repeating structures known as subunits. Biopolymers
possess a diverse range of functions within cells. For example, DNA and RNA are
polymers of nucleotides whose primary function is the storage and passage of
genetic information. In Chapter 8, we will learn more about the cytoskeleton, an
interconnected network of polymers that plays a critical role in a variety of func-
tions, such as maintenance of cell shape, mechanical force generation, and intra-
cellular transport. In many cases, the mechanical behavior of these biopolymers
is critical in allowing them to perform the diverse functions they serve.
In this chapter, we focus on obtaining a quantitative understanding of the mechan-
ics of biopolymers. We first describe the molecular structure of three important
cytoskeletal polymers: microfilaments, microtubules, and intermediate filaments.
We follow this with a model of polymerization dynamics in microfilaments and
microtubules. Next, we consider how to characterize a polymer as being flexible
or stiff and discuss three different polymer models: the ideal chain, the freely
joined chain (FJC), and the wormlike chain (WLC). Through this development, it
is our hope that you will have a better understanding of how the mechanics of
biopolymers may dictate their capacity to perform their biological functions.
37 nm
ATP
(ADP when
in filament)
− end
− end
25 nm
(A) (B) (C)
Figure 7.1 G-actin and F-actin structures. G-actin possesses a (+) and a (–) end as well as binding site for ATP. F-actin has a double-helix-
like structure consisting of two strands that spiral around the axis of the polymer, with each helical loop repeating every 37 nm. (Adapted
from, Alberts B et al. (2008) Molecular Biology of the Cell, 5th ed. Garland Science. Photo courtesy of Roger Craig.)
50 nm
MT polymerization is affected by polarity and GTP/GDP binding
The polymerization kinetics of tubulin are dictated by its asymmetric nature and
its ability to bind guanosine tiphosphate/guanosine diphosphate (GTP/GDP,
analogous to ATP/ADP in actin). Like F-actin, MTs are polar, possessing a rap-
idly polymerizing (+) end and slowly polymerizing (−) end. Once polymerized,
GTP hydrolyzes to GDP. As tubulin polymerizes, there is a region near the
polymerizing end of GTP-tubulin known as the GTP cap (Figure 7.5). Only GTP
tubulin tends to become polymerized because polymers of GDP tubulin are
unstable and easily return to monomeric form. The GTP cap can be thought of − end
as preventing the depolymerization of the entire MT.
microtubule
Periodically, the GTP cap is lost owing to impediments to further polymeriza-
tion or simple random fluctuations. When this happens, the GDP MT cata-
β tubulin heterodimer
strophically depolymerizes until the GTP cap can be reestablished. Microtubules
are constantly shifting from a state of slow polymerization to rapid depolymeri- α (= microtubule subunit)
zation in a process known as dynamic instability. Dynamic instability results in
MTs constantly extending and retracting. This is thought to be important for Figure 7.3 Structure of a MT. Each MT
consists of 13 protofilaments joined
the ability of the MTs to find and attach to the kinetochore during mitosis. At
together in a hollow tubular structure.
other times, cytoplasmic MTs tend to organize into a network with their (−) MT subunits consist of a heterodimer of
ends near the nucleus and their (+) ends pointing outward toward the cell α- and β-tubulin. (Adapted from, Alberts
membrane. B et al. [2008] Molecular Biology of the
Cell, 5th ed. Garland Science.)
Figure 7.4 MTs during mitosis. When cells divide, MTs are responsible for segregating
chromosomes and pulling them to the poles of the cell. They do this by forming two spindles
and attaching to the kinetochores.
192 CHAPTER 7: Mechanics of Cellular Polymers
Table 7.1 Intermediate filaments by type, size, chromosome, and distribution Note that there are multiple subtypes for
each class; keratins, for example, consist of 20 different type 1 filaments. (From Omary MB, Coulombe PA & McLean WH.
(2004) N. Engl. J. Med. Copyright Massachusetts Medical Society.)
Cytoplasmic
Keratins I 40–64 17 Epithelium (keratins Form obligate 1:1 heteropolymers with
9–20); hair (keratins type II; protect from mechanical and
Ha1–Ha8) nonmechanical forms of stress
II 50–68 12 Epithelium (keratins Form obligate 1:1 heteropolymers with
1–8); hair (keratins type I; protect from mechanical and
Hb1–Hb8) nonmechanical forms of stress
Vimentin III 55 10 Mesenchyme Involved in vascular tuning and wound
repair in mice
Desmin III 53 2 All muscle May be important for mitochondrial
positioning and integrity
Glial fibrillary acidic III 52 17 Astrocytes Also found in hepatic stellate cells
protein
Peripherin III 54 12 Peripheral neurons Found in enteric neurons; may be
required for development of a subset of
sensory neurons
Syncolin III 54 1 Muscle (mainly Interacts with α-dystrobrevin
skeletal and cardiac)
Neurofilaments IV 61 (light), 90 8 (light), 8 Central nervous Form obligate 5:3:1 (light:medium: heavy)
(light, medium, and (medium), (medium), 22 system heteropolymers
heavy chains) 110 (heavy) (heavy)
α-Internexin IV 61 10 Central nervous May partly compensate for peripherin in
system peripherin-null mice
Nestin IV 240 1 Neuroepithilial Is also an early developmental marker, as
found in the pancreas
Synemin IV 180 (α) 15 All muscle Found at lower levels than desmin; also
and 150 (β) (ß-isoform mainly in found in astrocytes; probably indentical
(two splice striated muscle) to desmuslin
variants)
Nuclear
Lamins A and C V 62–78 1 Nuclear lamina Arise from a single, differentially sliced
gene
Lamins B1 and B2 V 62–78 5, 19 Nuclear lamina Arise from two different genes
Other
Phakinin (CP49) Orphan 46 3 Lens Forms beaded filaments; deletion in mice
causes lens defect
Filensin (CP115) Orphan 83 20 Lens Forms beaded filaments; deletion in mice
causes lens defect
194 CHAPTER 7: Mechanics of Cellular Polymers
NH2 COOH
(A)
α-helical region in monomer
NH2 COOH
(B)
coiled-coil dimer
NH2 COOH
48 nm
COOH NH2
NH2 COOH
(C)
COOH NH2
NH2 COOH
staggered tetramer of two coiled-coil dimers
(D)
(E)
10 nm
Figure 7.6 Intermediate filament structure. The monomer (A) forms a dimer with another monomer to form a coiled-coiled structure
(B). Two dimers are then staggered to form a nonpolar tetramer (C). The tetramers associate in a staggered manner (D), allowing them to
form the final helical filament structure (E). (Adapted from, Alberts B et al. (2008) Molecular Biology of the Cell, 5th ed. Garland Science.)
d Apoly +1
= kon Apoly [G ] − koff Apoly+1 . (7.2)
dt
When the polymer is not changing length, the time derivative of [Apoly+1],
d[Apoly+1]/dt, is equal to zero. When this occurs, Equation 7.2 implies that
Apoly+1 koff
[G ] = K where K = . (7.4)
Apoly kon
The constant K is known as the dissociation constant of the reaction. It is also
known as the critical concentration, as it is the concentration of subunit at which
the polymer is neither growing nor shrinking.
A plot of Equation 7.5 can be seen in Figure 7.8. If [G] > K, dL is positive, the poly-
dt
mer is growing, whereas if [G] < K, dL is negative, the polymer is shrinking. The
dt
196 CHAPTER 7: Mechanics of Cellular Polymers
K [G]
shrinking
only point at which dL = 0, in other words when the polymer is neither growing
dt
nor shrinking, occurs when [G] = K. This confirms that the critical concentration
is the only subunit concentration at which the polymer does not change length.
KT KD [G]
shrinking
198 CHAPTER 7: Mechanics of Cellular Polymers
Nota Bene but the (+) end will be polymerizing. The addition of subunits on one end and
subtraction at the other is known as treadmilling. Once D-form subunits are lost
Polymerization force. The peak
from the shrinking (−) end, they release ADP/GDP and bind ATP/GTP to become
force that a single fixed polymerizing
filament can exert on a surface has T-form subunits and are recycled to the growing (+) end. Intracellular concentra-
been derived through tions of G-actin and tubulin are typically within the range of subunit concentra-
thermodynamic considerations to be tions at which treadmilling occurs.
kT k [G ]
F = ln on .
δ koff 7.3 PERSISTENCE LENGTH
Using this relation, it has been
estimated that the peak force for We now shift our focus from polymer structure and kinetics to mechanical behav-
F-actin is 9 pN for a typical ior. In analyzing the mechanical behavior of polymers, the appropriate choice of
concentration of G-actin of 50 mM. model is determined in large part by whether the behavior is energy- or entropy-
This is many times the force that one dominated. We learned in Section 7.1 that different biopolymers can have very
myosin motor can generate. different molecular structures. One consequence of this is that under the influ-
ence of thermal forces, these polymers will have very different tendencies to bend
and coil. Consider a strand of DNA approximately 10 μm in length. If we grab it at
the ends, straighten it, and then release it in into a solution at room temperature,
L = 10 mm we would see the strand of DNA begin to take on a convoluted configuration, with
many undulations forming along its length. Now, we do the same for a MT 10 μm
microtubule in length. The MT does not coil. In fact, it may tend to resemble a fairly straight rod
(Figure 7.11).
actin The tendency for these two polymers to take on such different configurations in
the presence of thermal forces can be attributed to differences in energetic and
DNA entropic influences in equilibrium. The molecular structure of MTs results in a
much higher increase in potential energy during bending than that of DNA.
Figure 7.11 Representative The behavior of the MTs is energy-dominated taking on a straight configura-
configurations for a 10-μm segment tion, as this reduces the internal energy and therefore the free energy.
of a MT, actin, and DNA polymer at Conversely, the behavior of DNA is entropy-dominated. The free energy is
room temperature. Under the
reduced by increasing entropy, which occurs when the DNA strand takes on a
influence of thermal forces, the
tendency to take on a coiled versus coiled configuration.
straight configuration varies for each of
the polymers. Persistence length gives a measure of flexibility in a thermally
fluctuating polymer
In this section, we introduce the persistence length, which is a characteristic
length scale that gives a measure of flexibility in a thermally fluctuating poly-
mer. Consider a continuous polymer of contour length L undergoing thermal
fluctuations, as in Figure 7.12. We define a quantity s that runs from zero to L
and gives a parameterization by which each point on the polymer can be iden-
tified. We define the orientation at each point θ(s) as the angle the polymer
makes with an imaginary horizontal line. At s = 0, the polymer end is fixed such
that θ(0) = 0.
Now consider the case where we monitor θ(s) over time, and draw the probability
distribution for θ. If we were to draw the distribution for small s (near the fixed
end), then we would expect that the chances for the polymer to have an orienta-
tion different from θ = 0 would be very small, and therefore the distribution would
be very sharp, with a peak at θ = 0. If we were to find the average cosine of the
angle, we would find that 〈cos θ〉 ≅ 〈cos 0〉 ≅ 1, where the brackets indicate the
time average. In contrast, at larger s, we would expect a much higher chance for
the polymer to have a different orientation than θ = 0, so the distribution would be
wider (Figure 7.13). The distribution becomes wider and wider for larger s until,
at some point, the distribution becomes essentially uniform. Beyond this point,
the polymer orientation is effectively random (i.e., the orientation becomes
uncorrelated from the fixed end), and so 〈cos θ〉 ≅ 0. Therefore 〈cos θ〉 decreases
from 1 to 0 as s gets larger and larger. In fact, we now show that this decrease
occurs exponentially.
Consider the following approximation for the derivative of some function f (s),
df f (s + ∆s ) − f (s )
≅ (7.6)
ds ∆s
derived from the first two terms of a Taylor series approximation. Equation 7.6 is a
good approximation for small values of Δs, as long as f (s) is reasonably smooth. In
our case, f (s) = 〈cosθ ′(s)〉 , where θ ′(s) = θ(s) − θ(0), and
df cos ( ∆θ ′(s ))cos (θ ′(s )) − sin ( ∆θ ′(s ))sin (θ ′(s )) − cos (θ ′(s ))
≅ . (7.9)
ds ∆s
Again, using the fact that Δθ ′(s) and θ ′(s) are independent, we can use the identity
〈ab〉 = 〈a〉 〈b〉, and
df cos ( ∆θ ′(s )) cos (θ ′(s )) − sin ( ∆θ ′(s )) sin (θ ′(s )) − cos (θ ′(s ))
≅ . (7.10)
ds ∆s
However, because Δθ ′(s) and θ ′(s) are equally likely to be negative or positive
(symmetric about zero, odd functions like sines average to zero),
df cos ( ∆θ ′(s )) − 1
≅ f (s ). (7.12)
ds ∆s
Nota Bene
Three-dimensional orientation
Notice that the term in the parentheses is a constant, not dependent on s, and in
correlation function. If the motion
of the polymer is three-dimensional, addition it will generally be negative, because 〈cos[Δθ ′(s)]〉 is generally less than 1.
then the orientation correlation We can rewrite the above expression as
function is
df
−s ≅ −Cf (s ). (7.13)
cos ∆θ ( s ) = e
2p
. ds
A solution to this equation is
The different normalization takes into
account the fact that if the polymer is −s
allowed to move in three dimensions, (7.14)
p
it can bend in two directions cos ∆θ (s ) = e ,
orthogonal to its long axis, whereas if
it is constrained to move in two where for the rest of the chapter we will use the notation Δθ(s) = θ ′(s), 〈cos
dimensions, it can only bend in one Δθ(s)〉 is called the orientation correlation function, C = 1/ℓp, and ℓp is a nor-
direction. Therefore in two malization factor known as the persistence length. From the above expression,
dimensions, the orientation it is easy to see that the persistence length gives a characteristic length scale
correlation function will decay
half as fast.
over which the orientations of a thermally undulating polymer become mostly
uncorrelated.
Calculate the length of each segment such that the change The corresponding lengths for actin and DNA are 1.5 μm
in angle between the two ends of each segments is, on and 5-nm, respectively. Physically, this means that on
average, 25 degrees. Assume the persistence lengths for average, a 1.5-μm span of a thermally fluctuating actin
actin and DNA are 15 μm and 50 nm, respectively. polymer would have a difference in angle of 25 degrees
from one end to the other, whereas in DNA this would
Using Equation 7.14 we can solve for s by taking the
occur over a 5-nm span. These dimensions make some
natural logarithm of both sides and isolating s to obtain
sense considering the functions of each polymer; actin
s = −ℓp ln(〈cosΔθ(s)〉). filaments must have the capacity to span significant
lengths within the cell, whereas DNA must have the
We know that a change in angle of 25 degrees corre- capacity to be tightly coiled within the nucleus.
sponds to a value for the orientation correlation function,
〈cosΔθ(s)〉, of 0.9. Using this value, we obtain
s = 0.1ℓp.
Because the beam was bent π radians, we can express Equation 7.15 in terms of a
general bend angle θ as
EIθ
Q(θ ) = . (7.16)
2R
Using the fact that the bend angle can be expressed in terms of arc length, θ = s/R,
we can rewrite Equation 7.16 as
EIθ 2
Q(θ ) = . (7.17)
2s
Equation 7.17 describes the internal energy of a beam of arc length s subject to a
constant curvature such that the bend angle is θ radians. We now assume we are
given a polymer of length s that can be modeled as an elastic beam. If we were to
immerse the polymer within a constant temperature heat bath, we could use
Boltzmann’s distribution to find the probability of finding a polymer with bend
angle θ as
1 −Q (θ )/kBT
p(θ ) = e . (7.18)
Z
To compute the partition function Z, we must integrate the exponential term over
two angles, θ and ϕ, to allow for bending in three dimensions. Specifically, Z can
be computed as
2π π
Z=
∫ ∫e
0 0
−Q (θ )/kBT
dφ sinθ dθ ,
(7.19)
where we have taken the integral with respect to a differential element of solid
angle dϕsinθdθ. We now seek to quantify the average amount of polymer curva-
ture. We can do this by computing 〈θ 2〉 as
2π π
1
θ 2
=
Z ∫ ∫e −Q (θ )/kBT
θ 2dφ sinθ dθ . (7.20)
0 0
For small angles, it is left to you to show that the solution to this integral is
2kBTs
θ2 = . (7.21)
EI
Now, we relate the above expression to the orientation correlation function. We
can do this using a Maclaurin series expansion. The Maclaurin series for cos x is
1 2 1 4
cos x = 1 − x + x ... (7.22)
2 24
Using this expansion and a small-angle assumption, we can write the orientation
correlation function as
∆θ 2 (s ) ∆θ 2 (s )
cos ∆θ (s ) ≈ 1 − = 1− . (7.23)
2 2
202 CHAPTER 7: Mechanics of Cellular Polymers
The term in brackets in the right-most relation in Equation 7.23 is the average
value of the difference in angle (squared) between two points separated by a dis-
tance s along the polymer. If the polymer is assumed to be a constant curvature
beam of arc length s, then we can substitute in our expression for 〈θ2〉 found in
Equation 7.23. In this case,
kBT (7.24)
cos ∆θ (s ) ≈ 1 − s.
EI
Using the approximation e−x ≈ 1 − x in the limit of small x, we can write our
expression for the orientation correlation function found in Equation 7.24 for
small s as
−s
−s
cos ∆θ (s ) = e p ≅ 1 − . (7.25)
p
Relating Equations 7.24 and 7.25, we can see that the persistence length ℓp is
related to flexural rigidity as
EI
p ≡ . (7.26)
kBT
R= ∑r . i (7.27)
i =1
We would like to say something about the average end-to-end length. However, if
we average the random vector R, we get 0 due to symmetry. In particular, there is
no preferred spatial orientation for our polymer, so any vector R that occurs can
b ri
204 CHAPTER 7: Mechanics of Cellular Polymers
also occur with equal probability as −R. Nonetheless, we can say something about
the average magnitude of R by examining its square
R2 = R ⋅ R
N
N
=
∑ i =1
ri ⋅
∑ r
j =1
j
(7.28)
N N
= ∑∑ r ⋅ r
i =1 j =1
i j .
We know that for the dot product of two vectors |a⋅b| = |a| |b| cos θab, where |a| and
|b| are the lengths of a and b, and θab is the angle between a and b. If we substitute
this expression into Equation 7.28, we obtain
N N
R2 = ∑∑b
i =1 j =1
2
cosθij
(7.29)
N N
=b 2
∑ ∑ cosθ
i =1 j =1
ij .
Furthermore, we know that the direction of any given segment is independent of
all of the other segments. The value of cosθab between any two segments will range
from −1 to 1. So, the average of cos θab will be 0, except when computed on the
same bond (a = b). In this case θab = 0 and cos 0 = 1. An equivalent statement is
〈cos θij〉 = 0 for i ≠ j, and 〈cos θij〉 = 1 for i = j, or 〈cos θab〉 = δij. Therefore,
n n
∑∑δ
Nota Bene
R 2 = b2 ij = nb 2 . (7.30)
Kronecker delta. The Kronecker
i =1 j =1
delta δij is a compact notation used
to signify the following relation: This is the mean square end-to-end length of a three-dimensional chain com-
i = j posed of n segments of size b. Notice that just as in the random walk from Section
1, if
δ ij = 5.6, the quantity 〈R2〉 scales linearly with n. Only now, instead of n representing
0, if i ≠ j
the number of steps in the walk, it represents the number of links in the chain.
Equation 7.32 is for a one-dimensional random walk. We now seek to extend this
relation to three dimensions. To see the connection between one and three dimen-
sions, let us start by defining the end-to-end distance for our three-dimensional
random walk in the coordinate directions, R = Rxex + Ryey + Rzez (where ex, ey, and
ez are unit vectors in the x-, y-, and z-directions). The steps in each direction are
independent. So, the mean square end-to-end distance is the sum of the mean
square end-to-end distances along each of the three coordinate directions, or
R2
Rx2 = Ry2 = Rz2 =
3
(7.34)
2
nb
= .
3
〈r 2 〉 = 〈 Rx 2 〉 = 〈 Ry 2 〉 = 〈 Rz 2 〉 = 〈R 2 〉/3. (7.35)
3 2
2〈R2 〉
p1d (n , Rx ) = p1d (n , Ry ) = p1d (n , Rz ) = e −3 Rx . (7.36)
2π〈R 〉
2
32
3 (7.37)
−3( Rx2 + Ry2 + Rz2 ) 2 R 2
= e
2π R 2
32
3 −3 R 2 2 R 2
= e .
2π R 2
206 CHAPTER 7: Mechanics of Cellular Polymers
Nota Bene We can express Equation 7.37 in terms of the number of segments n and the Kuhn
length b by substituting the relation 〈R2〉 = nb2. We obtain
Excluded volume interactions. In
the ideal chain, we assume that 3/2
segments can overlap. In other 3 2
/2 nb 2
p3d (n ,R ) = e −3R , (7.38)
words, two segments can occupy the 2πnb 2
same space at the same time. In
reality, this cannot occur. Models that
restrict segment overlap are said to which is the three-dimensional probability distribution for an ideal chain consist-
account for excluded volume ing of n segments of length b to have an end-to-end vector of R. Because the ideal
interactions. One such model that
accounts for excluded volume effects
chain is based on the above Gaussian distribution, it is also referred to as the
is based on the self-avoiding random Gaussian chain.
walk, a walk that cannot cross a point
it has traced previously.
∑ n
∑ ln 2 + 1 − s
n
− ln + s +
ln(p(n , R)) = −n ln(2) + ln(n !) 2
s =1 s =1
n+ R n−R n
− ln ! − ln ! . (7.39) = −n ln(2) + ln(n !) − 2 ln !
2 2 2
Now, if a, b, and c are positive integers such that a ≥ b, it n
R /2 +s
can be shown that ((a + b)/c)! can be expressed as
b/c
− ∑
s =1
2
ln
n
+ 1 − s
a+b
∏ a
a 2
!= ! + s , (7.40)
c c s =1
c
n
= −n ln(2) + ln(n !) − 2 ln ! (7.44)
and ((a − b)/c)! can be written as 2
a 2s
a−b ! R /2 1+
c
!= b/c
c . (7.41) − ∑
ln n ,
2s 2
∏s =1
a
c
+ 1 − s
s =1 1 − +
n n
The third term in Equation 7.39 can be written as where we have divided the numerator and denominator
by n/2 in the last term in the last line.
R /2
n+ R n!
∏ 2 + s
n We now invoke our large n approximation. In particular,
ln ! = ln
2 2 in the last term of Equation 7.44, any term with n in
s =1
(7.42) the denominator will go to zero in the limit of large
R/2
n. Because ln(1 + a) ≅ a for |a| 1, we can approxi-
n
= ln ! +
2 ∑ s =1
n
2
ln + s
mate the logarithm in the last term of Equation 7.44 as
Using Equation 7.45 and the identities Because ln(a) = b and a = eb are equivalent statements,
a
Equation 7.48 can be rewritten as
∑ s = a(a + 1)/2
s =1
(7.46)
1 n! − −
R2 R2
p(n , R) ≅ n e 2n ≅ Ce 2n , (7.49)
2 (n/2)!(n/2)!
and
a
∑1 = a ,
s =1
(7.47) where
1 n!
C= .
Equation 7.44 can be expressed as 2n n ! n ! (7.50)
2 2
n
ln(p(n , R)) ≅ −n ln(2) + ln(n !) − 2 ln !
2 We can obtain a simplified expression for C by observing
R /2 that it is a normalization constant such that
∑ n − n
4s 2
−
∞
∫
s =1
p(n , R)dR = 1. (7.51)
−∞
n
≅ −n ln(2) + ln(n !) − 2 ln !
2
In this case, we obtain
R /2 R /2
∑ s + n ∑1
4 2
− ∞ R2
∫e
n −
s =1 s =1 C =1 2n dR = 1 2πn . (7.52)
n −∞
≅ −n ln(2) + ln(n !) − 2 ln ! (7.48)
2
Combining Equations 7.49 and 7.52, we obtain Equation
7.32, which shows its equivalency to Equation 7.31 in the
R R limit of large n.
+ 1
4 2 2 R
− +
n 2 n
2
n R
≅ −n ln(2) + ln(n !) − 2 ln ! − .
2 2n
This proportionality arises due to the facts that (1) the integral of p3d(n, R) over
some range gives the exact probability for the polymer to have a end-to-end
vector within that range, and (2) this probability is equal to the number of poly-
mer configurations divided by the total number of all configurations. Using
208 CHAPTER 7: Mechanics of Cellular Polymers
3 R2
S (n , R ) = − kB 2 + S0 , (7.56)
2 nb
where we have turned the proportionality into an equivalency by accounting for
any terms that do not depend on R in the constant S0. Given an expression for
entropy, we can calculate the free energy for a chain composed of n segments with
an end-to-end vector R as
3 R2
Ψ(n ,R ) = −TS (n ,R ) = kBT 2 + Ψ0 , (7.57)
2 nb
where Ψ0 = −TS0 does not depend on R.
∂Ψ(n , R ) 3kBT
Fx = = Rx
∂Rx nb 2
∂Ψ(n , R ) 3kBT
Fy = = Ry (7.58)
∂R y nb 2
∂Ψ(n , R ) 3kBT
Fz = = Rz
∂Rz nb 2
or,
3kBT
F= R. (7.59)
nb 2
We previously remarked that the Kuhn length b is related to the persistence length.
In Section 7.6 we will show that this relation is
b = 2 p . (7.60)
IDEAL CHAIN 209
So, we can rewrite Equation 7.59 in terms of the contour length L = nb and the
persistence length as
3kBT
F= R
nb 2
3kBT
= R (7.61)
(nb)b
3kBT
= R.
2 L p
Equation 7.61 gives the force–displacement relationship for an ideal chain. On
inspection, a couple of interesting things can be seen. First, for any non-zero
temperature and non-zero R, the force is non-zero as well. This is perhaps some-
what counterintuitive, because the polymer consists of freely rotating segments
that are unable to store energy. However, these forces are not as mysterious as
they may seem. Physically, the capacity for the polymer to exert a force if the ends
are separated arises out of random collisions from the thermal environment sur-
rounding it, as well as the tendency for the polymer to take on a probabilistically
favorable state. In particular, the maximum number of polymer conformations
can occur with a zero end-to-end distance, and this state is the most entropically
favorable. Straightening the polymer out reduces the entropy and therefore
requires force.
Another interesting aspect about Equation 7.61 is that the force F is linear in R, so
pulling the ends apart with a distance R produces the same force as an elastic
spring with a spring constant of (3kBT)/(2Lℓp). This relationship has been referred
to as describing an entropic spring. The resistance to deformation is derived
entirely due to entropy and depends on temperature. We can see that the stiffness
of the spring is proportional to temperature, so increasing T increases the stiff-
ness. This is in contrast to most engineering materials such as steel, which become
more compliant with increasing temperature.
The tendency for polymers with large contour lengths to tend toward ideal chain
behavior can be better understood by examining the physical meaning of the
persistence length. We know that at length scales longer than the persistence
length, the orientation correlation between two points of a thermally fluctuating
polymer vanishes. Physically, this means that at this length scale, entropic influ-
ences become dominant over energetic influences. We also know that if a poly-
mer is longer than its persistence length, then for every span of polymer
approximately equal to this length, the orientations of the polymer become
roughly uncorrelated. At very long length scales, the polymer behaves as if it were
composed of many independently fluctuating chain segments, with each seg-
ment being on the order of the persistence length in size. The behavior of most
polymers will be dominated by entropy and tend toward that of an ideal chain
when the contour length is significantly greater than the persistence length.
In modeling such cases using an ideal chain, a Kuhn segment is not necessarily a
single monomer, but rather a span of polymer roughly equivalent to the persis-
tence length in size.
R=b ∑u . i (7.62)
i =1
To simplify the development (and without losing any generality) we will consider
a chain that is being extended in the z-direction. We can simplify the math by
expressing R in spherical coordinates about the z-axis with θi being the angle with
the z-axis and ϕi being the azimuth (Figure 7.16) as
e x ⋅ ui = sinθi cosφ i
e z ⋅ ui = cosθi ,
where ex, ey, and ez are unit vectors in the x-, y-, and z-directions.
FREELY JOINTED CHAIN (FJC) 211
y = b·sin f sin θ
x = b·sin f cos θ y
b·
sin
θ f
Using these relations, the end-to-end distance along the z-axis can be written as
Rz = e z ⋅ R
n
=b ∑e ⋅ u
i =1
z i
(7.64)
n
=b ∑ cosθ .
i =1
i
θ1,2,. . . ,n and ϕ1,2,. . .,n. The Boltzmann probability associated with the microstate
having energy Q(θ1,2,. . . ,n,ϕ1,2,. . . ,n) is
n
1 κ ∑ cosθi (7.66)
p(θ1 ,θ 2 ,....θ n ,φ 1,φ 2 ,... ,φ n) = e i =1 ,
Z
where
Fz b 2 Fz p
κB = = . (7.67)
kBT kBT
(7.68)
where we have performed the integral with respect to a differential element of
solid angle sinθ dθ dϕ because the bond angles were defined in spherical coordi-
nates. We can rearrange the order of integration in Equation 7.68 to perform the
integrals for each segment in turn as
2π π 2π π
Z=
∫ ∫
φ 1= 0 θ1 = 0
eκ cosθ 1 sinθ1dθ1dφ 1
∫ ∫e
φ 2=0 θ2 =0
κ cosθ 2
sinθ 2dθ 2dφ 2 …
2π π
∫ ∫e
φ n = 0 θn = 0
κ cosθn
sinθ ndθ ndφ n .
(7.69)
Z= ∏∫ ∫e
i =1 φ i = 0 θi = 0
κ cosθi
sinθi dθi dφ i = z n , (7.70)
where
2π π
z=
∫ ∫e
φ =0 θ =0
κ cosθ
sinθdθdφ (7.71)
is the single partition function. To compute z, we first integrate out θ as
π
z = 2π
∫e κ cosθ
sinθdθ . (7.72)
θ =0
Next, we change variables by letting ρ = cos θ and dρ = −sin θ dθ. Equation 7.72
can be evaluated as
1
eκ − e −κ sinhκ
z = −2π eκρ dρ = 2π
∫ κ
= 4π
κ
. (7.73)
−1
where sinh κ is the hyperbolic sine, sinh κ = (1/2)(eκ − e−κ).
FREELY JOINTED CHAIN (FJC) 213
∫
Rz = p(Θ)Rz dΘ , (7.74)
Θ
where for simplicity, we have used Θ in Equation 7.74 to denote integration over
all possible bond angles θ1,2,. . .,n and ϕ1,2,. . .,n. Substituting our expression for p
obtained in Equation 7.68 into Equation 7.74 and using the relation
n
n
∑
κ b cosθi
κ ∑ cosθi i =1 κRz (7.75)
e i =1
=e b =e b ,
we obtain
κRz
1
Rz =
Z
e
∫ b RzdΘ. (7.76)
Θ
Similar to our approach in calculating internal energy in the canonical ensemble,
we can rewrite Equation 7.76 as a logarithm of Z with the following manipulation
∂ bz
κR
b
Rz =
Z ∫
Θ
∂κ
e RzdΘ
κRz
b ∂
=
Z ∂κ
e
∫
Θ
b R z dΘ
(7.77)
b ∂Z
=
Z ∂κ
∂ ln Z
=b .
∂κ
Using Equation 7.77, we can compute the mean extension as
∂ ln Z
Rz = b
∂κ
∂
= bn [ln(sinh κ ) − lnκ + ln 4π] (7.78)
∂κ
1
= bn cothκ − .
κ
Recall that force is embedded inside κ. Equation 7.78 gives a force-displacement
relationship for the FJC. The expression is a little different than those with which
we are familiar, with because it is an implicit relationship. In other words, although
it does produce a force for every displacement and vice versa, we do not have an
explicit expression for force. However, it is still an entirely valid relationship.
coth(x) − 1/x
0.5
0.4
0.3
0.2
0.1
0.0
0 1 2 3 4 5
x
e x + e−x 1
= − . (7.79)
e x − e−x x
A plot of the Langevin function can be seen in Figure 7.18. Notice that as x gets
very large, ℒ approaches one. This means that even as the force grows indefinitely,
the end-to-end length of the polymer cannot exceed the contour length, nb. Thus,
the FJC overcomes one of the most critical limitations of the ideal chain. When the
value of x is small, the slope of ℒ(X) approaches 1/3. In this limiting case we can
approximate ℒ(x) to be x/3 and
κ F nb 2
Rz ≈ nb = z , (7.80)
3 3kBT
which is that obtained for the ideal chain in Equation 7.61. When the extension
(and force) are very small, the FJC model predicts the same force–displacement
behavior as the ideal chain.
∂a(s)
—
∂s
a(s)
the tangent vector (the first derivative of a(s)) to have a unit magnitude for all
values of s
2
∂a(s )
= 1. (7.81)
∂s
Alternatively, the tangent vector ∂a(s)/∂s is a unit vector.
We now specify the constitutive relations for the model. In the WLC, the chain is
assumed to resist bending deformation through increases in potential energy
with curvature similar to an elastic beam. In Section 3.2, we found the bending
energy for a beam is given by
L 2
EI ∂ 2a(s )
Q(a(s )) =
∫
2 ∂s 2
ds , (7.82)
0
where the term in parentheses is a vector that describes the local curvature. We
can express Equation 7.82 in terms of persistence length. In particular, in Section
7.3 we showed that for a curvy elastic beam subject to thermal fluctuations, its
persistence length is related to its flexural rigidity as ℓp = EI/kBT. We can recast
Equation 7.82 in terms of persistence length as
L 2
k T ∂ 2a(s )
Q(a(s )) = B p
2 ∫
∂s 2
ds. (7.83)
0
We first calculate the probability of looping for very short chain of independently fluctuating segments (each seg-
operator distances. We know that the smaller the contour ment having length on the order of ℓp), and the free
length relative to its persistence length, the more we must energy becomes dominated by the entropy. Taking this
account for energetic influences. We use the WLC model. into account, for very long operator distances, we model
Assuming an operator distance of L loops into a circle of the polymer as an ideal chain. The probability distribu-
radius R, the energy of looping for a WLC is tion for an ideal chain is given by Equation 7.38. A loop
occurs when R = 0, and so
L 2
kBT p 1 kBT p L 2π 2kBT p
Qloop =
2 ∫ ds =
R 2 R 2
=
L
, (7.84)
3
ploop =
3/2
3
=
3/2
, (7.85)
2πnb 2 2πLb
0
where in the right-most relation we made use of the fact
that R = L/2π. We know from the Boltzmann distribution where we made use of the fact in the right-most relation
that the higher the value for Qloop, the lower the proba- that L = nb. In this case, ploop → 0 as L → ∞.
bility of looping, ploop. Thus, ploop → 0 as L → 0.
Our models predict that the probability of looping goes
Now let us compute the probability for looping for very to zero for very short and long operator distances. In
long operator distances. As the contour length becomes both cases, repression would be expected to decrease
longer and longer, the polymer begins to resemble a (causing lac mRNA levels to increase).
L 2 L
k T ∂ 2a(s ) ∂a(s )
Qtot (a(s )) = Q(a(s )) − Fz Rz (a(s )) = B p
2 ∫2
∂s
ds − Fz
∂s ∫
⋅ e z ds.
0 0
(7.87)
Now that we have the internal energy of the system for each configuration a(s), we
can compute the probability of each curve using the Boltzmann distribution as
where
Z=
∫e −Qtot ( a ( s ))/kBT
da. (7.89)
∀a
In principle, we can compute the average extension under the applied force as
Rz =
∫ p(a(s))R da z (7.90)
∀a
∂ ln Z
Rz = kBT . (7.91)
∂Fz
kBT 1 Rz
−2
Rz 1 (7.92)
Fz = 1 − − + .
p 4 L 4 L
Differences in the WLC and FJC emerge when they are fitted
to experimental data for DNA
The differences in the force–extension behavior of the FJC and the WLC can be
demonstrated by fitting the two models to force–extension curves for DNA, such
as in Figure 7.20. The measurements were made by attaching one end of a DNA
molecule to a glass surface and the other end to a magnetic bead. The bead was
pulled with a known force Fz, and the extension Rz was measured optically. The
data were fitted to the WLC and FJC using the relation b = 2ℓp = 106 nm. Upon
inspecting the model fits, there are two noticeable trends. First, at large forces, the
extension asymptotically approaches the contour length. Second, the experimental
data is fit better by the WLC, especially at forces greater than approximately 0.1
pN. The WLC predicts that more force is required to extend the polymer to a given
length than the FJC. The underlying reason for this can be understood by consid-
ering the thermal fluctuations in a polymer as waves of different wavelengths
superimposed on top of each other. In the FJC, wave-like fluctuations are con-
strained to be of wavelength b or higher, because they cannot be shorter than the
length of each link. This constraint is not present in the WLC. Additional force is
required to smooth out these short wavelength fluctuations present in the WLC,
but not the FJC.
L
∂a(s )
R=
∫ ∂s
ds. (7.93)
0
L L
∂a(s ) ∂a(s ′)
R2 = R ⋅ R =
∫
0
∂s
ds ⋅
∂s ′ ∫
ds ′
0
L L
∂a(s ) ∂a(s ′)
=
∫∫
0 0
∂s
ds ⋅
∂s ′
ds ′ (7.94)
L L
∂a(s ) ∂a(s ′)
=
∫∫ ∂s
⋅
∂s ′
dsds ′.
0 0
Recall that in our parameterization of the WLC, the tangent vector ∂a(s)/∂s was
constrained to be a unit vector. Therefore the magnitude of the product in the
brackets is one and Equation 7.94 can be written as
L L
R 2
=
∫ ∫ cos ∆θ s −s′ dsds ′ , (7.95)
0 0
where Δθs−s′ denotes the angle between the tangent vectors at s and s′. In
Section 7.3 we found that for a thermally fluctuating polymer, the average dif-
ference in angle between tangent vectors is given by the orientation correlation
function, which decays exponentially such that 〈cos(Δθ(s))〉 = e−s/ℓp. We assume
that on average, the difference in angle between any two points on the chain
will decay similarly as the distance between the two point is decreased, or in
other words,
L L
∫∫e (
− s − s ′ )/ p
R2 = dsds ′. (7.96)
0 0
Key Concepts 219
It is left as an exercise to show that the solution to the double integral in Equation
7.96 is
−L L
R 2 = 2 2p e p − 1 + . (7.97)
p
Consider the case where the contour length is much greater than the persis-
tence length, L ℓp in Equation 7.98. The exponential term in the brackets goes
to zero, and assuming the L/ℓp term is much greater than one, then
L
R 2 ≈ 2 p 2 = 2 L p . (7.98)
p
If we compare this quantity to the mean squared end-to-end length of nb2 for the
FJC, we obtain a relation between Kuhn length and persistence length,
b = 2 p , (7.100)
which shows the origins of Equation 7.60. Equation 7.100 implies that a discrete
chain with Kuhn length b has a persistence length of b/2 in the limit of b L.
Key Concepts
• The cellular cytoskeleton primarily consists of three dominated by entropy. The probability distribution
biopolymers: MFs, MTs, and intermediate filaments. function for the ideal chain is based on the Gaussian
The subunits for MFs and MTs are actin monomers approximation to the random walk.
and tubulin dimers, respectively. Intermediate • For thermodynamic systems, force is equivalent to
filaments can be composed of different proteins the gradient of free energy. Separating the ends of an
depending on their location within the body. ideal chain reduces the entropy, and thus requires
• There is a highly dynamic balance between the G force.
and F forms of actin. Owing to G-actin’s polarity and • The mechanical behavior of polymers will tend
the ability for bound ATP to undergo hydrolysis, toward that of an ideal chain when the contour
polymerization kinetics can be very different at the length is significantly greater than the persistence
(+) and (−) ends. Treadmilling occurs when D-form length.
subunits are lost from the shrinking (−) end, become
• The FJC addresses a key limitation of the ideal chain,
T-form subunits, and are recycled to the growing
which is unphysical behavior for large forces. The FJC
(+) end.
is similar to the ideal chain at low extensions, but
• The persistence length gives a characteristic length gives more realistic behavior at long extensions.
scale over which the orientations of a thermally
• The WLC does not have joints but rather treats the
undulating polymer become mostly uncorrelated. A
polymer like a flexible beam with elastic energy.
polymer can be classified as being flexible, semi-
Compared with the FJC, the WLC predicts that more
flexible, or stiff depending on the relation of its
force is required to extend the polymer a given
persistence length to its contour length
length.
• In the ideal chain, all conformational changes in
internal energy are ignored. It is often used for
modeling flexible polymers whose behavior is
220 CHAPTER 7: Mechanics of Cellular Polymers
Problems
1. Using the one-dimensional Gaussian approximation to 8. What is the slope of the Langevin function, L, near
the random walk and step size of b = 1, show that the zero?
root mean square displacement 〈R2〉1/2, is the square
root of the number of steps taken. You may use integral 9. What is the force generated in polymers of spectrin,
identities. actin, and tubulin (ℓp = 15, 15 × 103, and 2 × 106 nm,
respectively) with a 100-nm contour length predicted
2. Compare the root mean square end-to-end distance by the WLC model when the polymer is extended to
〈R2〉1/2 in three dimensions of strands of spectrin, actin, 50 nm?
and MTs (ℓp = 15, 15 × 103, and 2 × 106 nm, respectively)
with 200-nm contour length ( L). 10. What is the effective (tangent) spring constant as a
function of average displacement for a flexible polymer
−s
using the WLC model? What is the approximation in
2 p
3. Show that in three dimensions cos θ ( s) = e . the limit as displacement approaches zero? And when
displacement approaches the contour length? Hint: the
4. What is the force generated in polymers of spectrin, tangent stiffness is the slope of the force–displacement
actin, and tubulin (ℓp = 15, 15 × 103, and 2 × 106 nm, curve.
respectively) with a 100-nm contour length predicted
by the ideal chain model when the polymer is extended 11. What is the effective (secant) spring constant as a
to 50 nm? And to 150 nm? Why is this not realistic? function of average displacement for a flexible polymer
Assume a temperature of 300 K in your calculations. using the WLC model? What is the approximation in
the limit as displacement approaches zero? And when
5. Assuming a persistence length for DNA of 50 nm, displacement approaches the contour length? Hint: the
determine the force at which the ideal chain and the secant stiffness is the slope of a line from the origin to
FJC differ by 10% in the extension length, assuming they the point on the force–displacement curve.
are used to model the same strand of DNA with a 1-μm
contour length. Which model predicts a longer polymer 12. Assume a persistence length, ℓp = 15, 15 × 103, and
at this force? Assume a temperature of 300 K. 2 × 106 nm for spectrin, actin, and MTs, respectively. For
2π π filaments 1-cm long at T = 300 K, what is the effective
2k BTS
∫ ∫ θ p(U)dφ sinθ dθ =
6. Show that θ 2 = 2 spring stiffness (tangent) at zero displacement and
EI when fully extended?
0 0
assuming that the probability follows a Boltzmann 13. Consider a 30-µm length of DNA such as might be found
distribution and the energy is the strain energy of a in a virus. What force is required to stretch the DNA
beam bent to an angle θ. Hint: the integral you obtain is to an end-to-end displacement x = 10, 20, and 25 µm
quite challenging. Rather than attacking it directly, try at 300 K using first a FJC model and secondly a WLC
the mathematical trick we used for the FJC and WLC. model? Assume ℓp = 50 nm. Comment on the results of
Namely, show that 〈θ2〉 can be expressed in terms of the the two models.
derivative of ln(Z) with respect to E.
14. Conduct a numerical comparison of the FJC and WLC
7. What is the force generated in polymers of spectrin, models. Plot the force–extension behavior for a 1-cm
actin, and tubulin (ℓp = 15, 15 × 103, and 2 × 106 nm, long polymer with a 0.1-mm persistence length with
respectively) with a 100-nm contour length predicted by each model. Is one always higher or lower than the
the FJC model when the polymer is extended to 50 nm? other? Does this make sense? Comment on how the
And to 150 nm? How is this more realistic than the Ideal two compare in the short and long limits. Repeat with
chain model? Hint: the relationship between force and a 1-cm persistence length and a 10-cm persistence
displacement for the FJC model is implicit (you cannot length. How does persistence length affect how the
plug in a value of 〈R〉 and calculate F.) Instead, solve models compare?
it numerically to two significant figures by guessing L L
∫∫e (
2 − s − s ′ ) / p
values of F and interpolating (or a more sophisticated 15. Show that the solution of R = dsds′ is
approach if you like). You can use MATLAB or another 0 0
programming approach. Alternatively, plot the force– −L
2l p2 e lp − 1 + L .
displacement relationship and estimate the point lp
from the graph. Comment on your results.
Annotated References 221
Annotated References
Boal D (2001) Mechanics of the Cell. Cambridge University Press. An operator. J. Molec. Biol. 267, 21–29. This article shows dependence
excellent text on cell mechanics with many in-depth treatments of of lac promoter repression on operator distances.
the polymer models covered in this chapter. Parts of the relation of
Omary MB, Coulombe PA & McLean WH (2004) Intermediate fila-
persistence length to Kuhn length were based on developments in
ments and their associated diseases. N. Engl. J. Med. 351, 2087–
this text.
2100. A nice review of intermediate filament biology. Table 1 in this
Bustamante C, Marko JF, Siggia ED & Smith S (1994) Entropic elastic- chapter was obtained from this study.
ity of l-phage DNA. Science 265, 1599– 1600. This paper gives the
Phillips R, Kondev J & Theriot J (2009) Physical Biology of the Cell.
force extension behavior for the WLC and shows the differences in
Garland Science. This textbook gives a more in-depth treatment of
the WLC and freely jointed chain when fit to force–extension behav-
the lac repressor example as well as many interesting applications
ior for DNA.
of the polymer models developed in this chapter.
Howard J (2001) Mechanics of Motor Proteins and the Cytoskel- Rubenstein M & Colby RH (2003) Polymer Physics. Oxford University
eton. Sinauer Associates. A great introduction to cytoskeletal Press. An excellent introduction to polymer physics. Derivation for
polymerization and mechanics. Students interested in learn- the Gaussian approximation to the random walk is based on the
ing more about polymerization kinetics and force generation by treatment developed in this text.
polymerization are referred to this text. The derivation of persis-
tence length, the relation of persistence length to flexural rigid- Spakowtiz AJ (2008) Polymer Physics (lecture notes, from course
ity, as well as some parts of the relation of persistence length to number ChemEng 466, Stanford University, Stanford, CA). Several
Kuhn length were based on the treatments of these topics in of the treatments in this chapter were based on excellent course
this text. notes developed by Dr. Spakowitz for a graduate course on polymer
physics at Stanford University. These treatments include develop-
Muller J, Oehler S & Muller-Hill B. (1996) Repression of lac promoter ments for the ideal chain, the freely jointed chain, and the WLC, as
as a function of distance, phase, and quality of an auxillary lac well as the lac repressor example.
This page intentionally left blank
to match pagination of print book
CHAPTER 8
Polymer Networks and
the Cytoskeleton
explicitly account for all the individual polymers in a given network. However, in
most cases, the numbers of degrees of freedom in physiological networks are too
large even for the world’s most powerful supercomputers.
An alternative approach is to represent the network as an effective continuum to
reduce the number of degrees of freedom. What we mean by an effective contin-
uum is that we equate the mechanical contributions of all the discrete polymers
within the network to the mechanical behavior of some equivalent continuum.
Then, the mechanical state of the network can be described through the analytical
framework of continuum mechanics.
Consider an imaginary cube-shaped polymer network. The length of each side of
the network is L. The network is filled with randomly aligned polymers that are
cross-linked to one another (Figure 8.1). We apply a very small uniaxial deforma-
tion to the network in the x-direction by fixing one side of the cube so that it does
not translate in the x-direction, and subjecting the opposite side of the cube to a
uniaxial normal stress σ. The lengths in the y- and z-directions are assumed to be
unconstrained, such that the network can still expand in the y- and z-directions.
Upon deformation, the new length in the x-direction is L1.
Next, we replace the polymer network with an isotropic continuum with Young’s
modulus E = Lσ/(L − L1). The deformation of the entire network under stress σ
can be described by a single homogeneous strain of (L − L1)/L. This simple exam-
ple demonstrates that by assigning appropriate constitutive behavior to the con-
tinuum, we are able to capture the mechanical behavior of the network without
explicitly modeling each individual filament. In the following sections, our overall
focus will be to form effective continuum models for the mechanical behavior of
E
σ
L
x L x
L L1 – L
L ε=—
L
L1
SCALING APPROACHES 225
polymer networks. In doing so, we will seek to identify relationships between the
effective mechanical behavior of networks and their microscopic properties, such
as polymer stiffness, length, density, orientations, and other properties.
F x3 x2
w= − L .
EI 6 2
226 CHAPTER 8: Polymer Networks and the Cytoskeleton
F
F
− FL3
w= .
3EI
This relation can be used as the basis of the scaling relationship for the mid-span
displacement in our unit cell, δ, because
FL3 FL3
δ~ 4
~ , (8.1)
EbR EbR 4
where F is some load applied to the subunit, Eb is the elastic modulus of the beam,
and the moment of inertia, I, scales as R 4. We have omitted many constants from
this expression, but those are fixed values with which we are not concerned. We
want to see how the two parameters are related, not an actual magnitude. We can
rearrange Equation 8.1 to solve for force as
E b R 4δ
F~ . (8.2)
L3
We now wish to calculate the apparent stress applied to the unit cell and the
apparent strain that the unit cell experiences. To do this, we consider the case in
which surrounding the unit cell is a hypothetical “cube” (Figure 8.3). The length
F F
F F
SCALING APPROACHES 227
of each side of the cube is L. The apparent stress is the total force applied across a
surface of the cube divided by the area L2, and the apparent strain is the total
deflection of the cube divided by the side length L. This apparent stress can be
found by dividing the force in Equation 8.2 by the area of the face of the unit cell
over which the force is acting,
F E b R 4δ
σ~ ~ . (8.3)
L2 L5
The apparent strain can be found as the displacement δ divided by the unde-
formed length of the unit cell L,
δ
ε~ . (8.4)
L
Using Equations 8.3 and 8.4, the apparent elastic modulus of the unit cell can be
found by dividing the apparent stress by the apparent strain,
σ R4
E~ ~ Eb 4 . (8.5)
ε L
Recall that volume fraction is the fraction of the volume taken up by polymer
material. We can recast Equation 8.5 in terms of this quantity, which we will
denote as ρvol. For the unit cell, the volume of the beams scale as ~LR2, whereas
the volume of the unit cell scales as L3. In this case,
LR 2 R 2
ρvol ~ ~ 2. (8.6)
L3 L
E ~ E b ρ 2vol. (8.7)
Equation 8.7 implies that when deformation of the cellular solid occurs due to
bending, the apparent macroscopic modulus scales linearly with the polymer
modulus but quadratically with the polymer volume fraction. This is an interest-
ing result, as it suggests that the stiffness of the network is more dependent on
changes in the polymer concentration rather than on the stiffness of the polymers
themselves.
E b R 2δ
F~ , (8.8)
L
E ~ E b ρ vol. (8.9)
228 CHAPTER 8: Polymer Networks and the Cytoskeleton
Nota Bene The derivations of Equations 8.8 and 8.9 are left as student exercises. Equation 8.9
implies that when deformation of the cellular solid occurs due to axial strain, the
Experimental measurements of effective elastic modulus of the network scales proportionally to both the cytoskel-
cytoskeletal volume fraction. We
etal modulus and the polymer volume fraction. By assuming a different mecha-
have seen that the cellular solid
model can predict very different nism of deformation at the microstructural level, we find different dependencies
scaling behaviors depending on of the network stiffness on the polymer volume fraction.
whether the deformation is assumed
to occur due to member bending or
axial deformation. One may ask how
The stiffness of tensegrity structures scales linearly with
these two models compare with member prestress
experimental investigations of the
relationship between cell stiffness We conclude this section with a discussion of tensegrity structures. The term
and cytoskeletal polymer volume tensegrity, which is a contraction of tensional integrity, was originally developed
fraction, and subsequently, which by Buckminster Fuller to describe a structural principle that had been invented by
model is more “correct”. Kenneth Snelson. Tensegrity structures are developed from linear structural ele-
Unfortunately, a conclusive answer to ments that are designed to withstand either tension only (such as strings or cables)
this question has yet to be attained,
owing in large part to the technical
or compression only (such as rigid struts). A critical aspect of a tensegrity struc-
difficulties associated with measuring ture is that none of the elements experience bending moments. This lack of bend-
cytoskeletal volume fraction within a ing moment is typically achieved architecturally by using freely rotating joints at
live cell. all connections. The resulting structure derives its stability from preload or pre-
stress (Figure 8.4).
Donald Ingber first suggested that the cytoskeleton might function as a tensegrity
structure. In this model, large-diameter microtubules act as rigid compression
elements, and the microfilaments and intermediate filaments act as flexible ten-
sile elements. There are some compelling arguments why a tensegrity structure
would be advantageous. Tensegrity structures possess the capacity to deform sig-
nificantly without large deformations of the individual components. That is, the
tension and compression elements rearrange themselves in response to some
load, but the individual elements do not need to stretch or compress by as much
as the entire structure. Such elemental rearrangements also underlie the capacity
for tensegrity structures to translate a deformation in one location into an equally
large (or larger) displacement some distance away, termed action at a distance.
However, a consensus of opinion on whether the cytoskeleton functions as a
tensegrity structure has not yet been formed, and this topic remains a controver-
sial one within cell mechanics.
Scaling relationships for tensegrity structures have been formed. Though its deriva-
tion is outside the scope of this book, it can be shown that in the existence of a pre-
stress, the effective network elastic modulus scales (under certain conditions) as
Fp ρ vol
E~ , (8.10)
r2
where Fp is the contractile force within the members generating the prestress.
Equation 8.10 implies that the stiffness of a tensegrity network scales linearly
with the polymer volume fraction, which is perhaps not surprising, given that
(1) the elements within tensegrity structures cannot sustain bending, and (2) a
similar scaling occurs in the axially loaded cellular solid model. Equation 8.10
also implies that the network stiffness scales linearly with the prestress of the
network.
One reason for the controversy over the tensegrity the tensegrity hypothesis. Nuclear motion has been
hypothesis is that there is evidence both in support of observed in response to a small bead being pulled far
and against this notion. In contrast to the tensegrity from the nucleus when engaging certain receptors, but
model, it is generally accepted that microtubules and not others. These results are interpreted as supporting
actin filaments cannot be classified as strictly compres- the tensegrity model’s action-at-a-distance concept.
sion- and tension-bearing elements, respectively. Prestress in actin filaments has been demonstrated using
Extension of lamellipodia occurs by axial compression of a combination of photobleaching and laser-scalpel tech-
actin fibers, and microtubules in tension are, in part, nology. The tensegrity hypothesis will likely continue to
responsible for pulling chromosomes apart during mito- be controversial for many years to come.
sis. On the other hand, there is evidence supporting
a ( A , t ) = F (t ) A + c ( t ) , (8.11)
230 CHAPTER 8: Polymer Networks and the Cytoskeleton
Nota Bene
where F(t) is the linear transformation, and c(t) is the rigid-body translation. Note
that if F or c are functions of A, F = F(A,t) or c = c(A,t), then the deformation in
Validity of the affine Equation 8.11 ceases to be affine. Therefore, affine transformations are sometimes
approximation. The validity of the referred to as homogeneous.
affine approximation for different
classes of networks, and in particular, Under an affine deformation, the polymers deform as if they are embedded in a
for networks composed of semi- continuum. Consider an imaginary cubic network subjected to a uniaxial load,
flexible polymers such as the
and assume that the filaments undergo a deformation that is affine. For all fila-
cytoskeleton, is an active area of
research. The conditions under which ment segments oriented similarly between two cross-link points, the stretch is the
deviations from affine behavior occur same. Similarly, at all cross-links in which the two cross-linked polymers are simi-
are a complex problem that depends larly oriented, the change in angle between each cross-linked pair will also be the
on several factors, such as the same. Importantly, the affine assumption greatly simplifies making a constitutive
microstructural details of the model as it really does not matter how many filaments are in the network, because
network, the applied load, and the they are assumed to be all deforming in the same manner by the deformation
length scale being considered.
prescribed by F and c.
3kBR 2
S (R ) = − + S0 , (8.12)
2nb 2
where S0 is a constant.
For a single chain, the change in free energy associated with a change in initial
end-to-end vector R = [X,Y,Z] to a new end-to-end vector r =[x,y,z] is
3kBT
∆Ψ = −T(S (r ) − S (R )) = + (r 2 − R 2 ). (8.13)
2nb 2
Also recall from Equation 7.30 that the mean square end-to-end length for the ideal
chain is ∙R2∙ = nb2. This implies that if the network was synthesized by a sequential
process of polymerization followed by cross-linking, then the mean end-to-end
length for each chain would be roughly nb2, assuming that the polymers were given
time to equilibrate before cross-linking, and that the density of the polymer solution
was low enough that substantial inter-polymer volume exclusion interactions did
AFFINE NETWORKS 231
not occur. Because the chains in the network are randomly oriented, this implies
that the mean squared end-to-end distances in the x-, y-, and z-directions must
equal each other, and therefore,
R2 nb 2
R 2
x = R 2
y = R 2
z = = . (8.14)
3 3
Next, we deform the network such that the lengths in the x-, y-, and z-directions
are now lx, ly, and lz. The deformation of the network is given by
rx = λ x Rx
ry = λ y Ry (8.15)
rz = λz Rz ,
and
lx ly l
λx = , λy = , λz = z (8.16)
Lx Ly Lz
are the stretch ratios in the x-, y-, and z-directions. In this treatment, we will
assume the deformations are homogeneous (and affine). This means that the
stretch of the individual chains between cross-links are spatially independent,
and can be described by the stretch of the bulk network (or, in other words, by the
relations above). Therefore, on average, the mean squared lengths of the chains in
the x-, y-, and z-directions after deformation will be
rx2 = λ x2 Rx2
∆Ψ =
3kBT
2nb 2
(
r 2 − R2 )
=
3kBT 2
2nb 2
(
rx + ry2 + rz2 − Rx2 + Ry2 + Rz2 )
(8.18)
3k T
(
= B 2 λ x2 Rx2 + λ y2 Ry2 + λz2 Rz2 − Rx2 − Ry2 − Rz2
2nb
)
=
2
(
kBT 2
λ x + λ y2 + λz2 − 3 , )
where we have used the relation ∙X 2∙ = ∙Y 2∙ = ∙Z 2∙ = nb2/3. The total average change
in free energy for the entire network ∙ΔΨnet∙ can be computed as the product of the
change in free energy for a single polymer, the number of polymers per unit vol-
ume ρn (which is also called the number density), and the total volume of the
network V = LxLyLz as
ρ nVkBT 2
∆Ψ net =
2
(
λ x + λ y2 + λz2 − 3 . ) (8.19)
232 CHAPTER 8: Polymer Networks and the Cytoskeleton
Now, consider the case of uniaxial stretching in the x-direction. We fix one end,
and load the opposite end such that the deformed length in the x-direction is
l x = λ Lx. (8.20)
In the y- and z-directions, there is contraction due to Poisson’s effect. Now, for
simplicity, assume that the network as a whole is incompressible. We can calcu-
late the stretch in the y- and z-directions. In particular, if we equate the volumes of
the undeformed and deformed networks,
Lx Ly Lz = λLx λ y Ly λz Lz . (8.21)
For Equation 8.21 to hold,
1
λ y = λz = . (8.22)
λ
Therefore,
ρnVkBT 2 2
∆Ψ net = − λ + − 3 . (8.23)
2 λ
We can find the force in the x-direction as
∂∆Ψ net 1
fx = = ρnAkBT λ − 2 , (8.24)
∂l x λ
where A = LyLz is the undeformed cross-sectional area (the proof of this is left as
an exercise). Now, we can find the axial stress in the x-direction as
fx 1
σ = = ρnkBT λ − 2 . (8.25)
A λ
In the limit of small strains, it can be shown that
1
λ − 2 ≈ 3ε . (8.26)
λ
σ = 3ρnkBT ε . (8.27)
The results of our analysis show that if we assume small deformations and that
the network is linearly elastic, isotropic, and incompressible, then the Young’s
modulus of our network is E = 3ρnkBT. There are several things worth noting.
First, the stiffness of the network scales linearly with the density of filaments,
ρn. This scaling of stiffness with density is similar to the scaling obtained in
Section 8.2, in which we assumed that all deformations in our lattice of beams
were axial, but not similar to the scaling obtained when we assumed the defor-
mations occurred because of bending. Perhaps this is not unexpected, because
in rubber-like networks, the ideal chains behave as entropic springs, which do
not generate force when bent. Second, we can see that the stiffness of the net-
work increases with increasing temperature, unlike most engineering materi-
als. This occurs because in a network of entropic springs, the stiffness of the
network is derived entirely from entropic costs of straightening out the indi-
vidual polymers. In particular, the polymers become straighter and have fewer
configurations available. Increasing the temperature increases the cost of this
reduction of entropy, and as such the material’s elastic modulus effectively
increases.
AFFINE NETWORKS 233
σ = Cε (8.28)
where σ and ε are 6 × 1 vectors. Therefore, in general, C is a 6 × 6 matrix. However,
Nota Bene
it turns out that C must also be symmetric, so even for a fully anisotropic linearly
elastic material there are only 21 independent elastic moduli, Symmetry of C. The reason that C
is symmetric may not be immediately
clear. Consider a complicated
C11 C21 C13 C14 C15 C16
composite material with lots of fibers
C21 C22 C23 C24 C25 C26 aligned in the z-direction. In this
anisotropic case, why would the
C31 C32 C33 C31 C32 C33
stress in the x-direction resulting
C = ,
(8.29) from a strain applied in z-direction be
equal to the stress in the z-direction
C411 C42 C43 C44 C56 C46
when the same strain is applied in
C51 C52 C53 C54 C55 C56 the x-direction? One requirement for
C C62 C63 C64 C65 C66 an elastic material it that it has a
61
strain energy function
σε cεε ∂ 2w
where Cij = Cji. In this section, we demonstrate a calculation of these 21 elastic w(ε ) = = . So, C = ,
2 2 ∂ε 2
moduli for a network of anisotropically aligned elastic rods undergoing an affine
which is symmetric, because the
deformation. However, before we do this, we will first discuss computation of
order or differentiation does not
elastic moduli from the strain energy density, because this approach will be useful matter.
in several developments in following sections.
σ = Eε . (8.30)
Combining Equations 8.30 and 8.31, we can see that the strain energy density w
can be formulated in terms of E and ε as
1 2
w= Eε . (8.32)
2
234 CHAPTER 8: Polymer Networks and the Cytoskeleton
Equation 8.32 implies that if we know the relationship between strain energy den-
sity and strain, we can compute the Young’s modulus E for the rod. In particular,
we see that E can be calculated by differentiating the strain energy density as
∂ 2w ∂ 2 1 2 Eε 2
=
∂ε 2 ∂ε 2
(8.33)
∂E ε
=
∂ε
= E.
In this simple example, we assumed that the rod was composed of an isotropic lin-
early elastic material. However, it can be shown that this relation holds for fully ani-
sotropic linearly elastic materials as well. Consider a continuum that is subjected to
some infinitesimal, homogeneous deformation described by the strain tensor. This
deformation results in a stress state described by the corresponding stress tensor.
Similar to the approach of determining the Young’s modulus by differentiating the
strain energy density with respect to strain, it can be shown that, more generally, all
21 independent elastic moduli of C can be computed as
∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w
∂ε xx ∂ε xx ∂ε xx ∂ε yy ∂ε xx ∂ε zz ∂ε xx ∂ε xy ∂ε xx ∂ε yz ∂ε xx ∂ε xz
∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w
∂ε yy ∂ε xx ∂ε yy ∂ε yy ∂ε yy ∂ε zz ∂ε yy ∂ε xy ∂ε yy ∂ε yz ∂ε yy ∂ε xz
∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w
∂ε zz ∂ε xx ∂ε zz ∂ε yy ∂ε zz ∂ε zz ∂ε zz ∂ε xy ∂ε zz ∂ε yz ∂ε zz ∂ε xz
C= .
∂w ∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w
2
∂ε xy ∂ε xx ∂ε xy ∂ε yy ∂ε xy ∂ε zz ∂ε xy ∂ε xy ∂ε xy ∂ε yz ∂ε xy ∂ε xz
∂w ∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w
2
∂ε yz ∂ε xx ∂ε yz ∂ε yy ∂ε yz ∂ε zz ∂ε yz ∂ε xy ∂ε yz ∂ε yz ∂ε yz ∂ε xz
∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w ∂ 2w
∂ε xz ∂ε xx ∂ε xy ∂ε yy ∂ε xz ∂ε zz ∂ε xz ∂ε xy ∂ε xz ∂ε yz ∂ε xz ∂ε xz
(8.34)
Equation 8.34 is a very useful result, as it implies that we can determine the effec-
tive elastic moduli of the network as long as we can formulate the functional
dependence of strain energy density with strain.
Example 8.1: The cytoskeletal modulus determined from strain energy density
If you examine the cytoplasm of a very thin cell, in a C11 C12 C13 0 0 0
region far from the nucleus, you might see the major fila- C12 C11 C13 0 0 0
ments of the cytoskeleton lying horizontally, in loose lay-
ers. If you are given the strain energy density of such a C13 C13 C33 0 0 0
section as, C= 0 0 0 C44 0 0
0 0 0 0 C44 0
w = 0.5C11(𝜀xx2 + 𝜀yy2) + 0.5C33(𝜀zz2) + C12𝜀xx𝜀yy + C13𝜀zz
(𝜀xx + 𝜀yy) + 0.5C44(𝜀xy2 + 𝜀yz2) + 0.25(C11 − C12) 𝜀xz2, 0 0 0 0 0 C11 − C12
2
calculate the stress–strain relationship and determine
This form of a material is associated with transverse iso
the general type of symmetry, if any, that exists in this
tropy, which means that the material has one set of prop-
model.
erties in one longitudinal direction and another set in
Using the tensor from Equation 8.35, we can fill in the any transverse direction within a plane perpendicular to
components as, the longitudinal direction.
AFFINE NETWORKS 235
nx
n = ny . (8.35)
nz
We can construct an angular probability density function, ω(n), which gives the
distribution of filaments over the unit sphere (which we will annotate as S2 based
on convention). The angular probability density function has the property that
∫ ω(n)dS = 1,
S2
(8.36)
which normalizes the total probability to be one. Next, consider our network
under an imposed load. Recall that the deformation is assumed to be affine.
The rods will freely rotate relative to one another and undergo some degree of
stretch. For an elastic rod with Young’s modulus Erod and volume Vrod, the total
strain energy in deforming the rod by stretching it by a factor of λ is
Vrod Erod
Wrod (λ ) = (λ − 1)2 . (8.37)
2
The degree to which each filament undergoes axial deformation will depend on
the network deformation and the orientation of the filament. In the case of
small deformations, the stretch for a filament oriented in the direction n under
an imposed strain of 𝜀 is approximately
For a network with rods oriented with angular probability density ω(n), the total
strain energy of the network is
Vrod Erod N
Wnet (ε ) =
∫W
S2
rod (λ (ε , n))ω (n)NdS =
2 ∫ (n ε n − 1) ω (n)dS ,
S2
T 2
(8.39)
where N is the total number of rods. The strain energy density can be calculated by
dividing the strain energy by the undeformed volume
ρvolErod
Wnet (ε ) =
2 2
∫
(nT εn − 1)2 ω (n)dS , (8.40)
S
236 CHAPTER 8: Polymer Networks and the Cytoskeleton
1–5 µm
–0.2 µm
Figure 8.5 Cytoskeletal structure within filopodia. (A, top) Electron micrograph of the actin cytoskeleton within a lamellipodium
of a cell. A dense network of actin filaments can be seen. Several filopodia can be seen protruding from the edge. (A, bottom) Close-up
of a filopodium. The aligned bundle of actin filaments within the filopodium can be seen. (B) Schematic of several important proteins
within filopodium. Note that the actin filaments are bundled together by the cross-linking protein fascin. (A, from, Mejillano MR et al.
(2004). Cell.)
that the persistence length of F-actin is on the order of ~10 μm. Filopodia are
much shorter than this, ~1 μm in length. Because the persistence length of actin Nota Bene
is an order of magnitude larger than the average length of the filaments within Image-based modeling of
the filopodia, we neglect entropic contributions and analyze the filaments as cytoskeletal networks. The
elastic beams. Since we are interested in the buckling behavior of filopodia, we angular distributions of filaments for
seek a relation between buckling loads and beam mechanical properties and several cases of cytoskeletal
networks have been quantified using
geometry. Such a relation was found in Section 3.2 when we analyzed the buck- image-processing analysis. Such
ling behavior of an elastic beam of length L with one end fixed, and with an axial angular distributions could be
force F applied at the free end (Figure 8.6). We found that for such a beam it will incorporated into the model
buckle if F exceeds the buckling force (Equation 3.38) described in Equation 8.40 to
simulate the mechanical behavior of
π 2 EI such networks based on realistic
Fbuckle = , (8.41) microstructures obtained from
4 L2 imaging.
where E is the Young’s modulus of the beam, and I is the moment of inertia.
Within polymer networks such as the cytoskeleton, an occurs through the cross-links such that bending and
individual polymer will typically have many cross-links stretching of polymers occurs between cross-links only.
along its span. However, near the ends of the polymer, the Because the free end is ”dangling”, it will likely not undergo
ends will be “dangling”: in other words, there will be a free stretch or bending. In Equation 8.40, for simplicity we
segment near the ends of the polymer that is not bounded assume that the free ends satisfy the affine deformation
on both sides by a cross-link. This free segment contrib- assumption. Though this is likely unphysiological, the
utes to the density of the network but in general, not its contributions of these free ends to the total strain energy is
strain energy. The reason is that the load transmission small.
238 CHAPTER 8: Polymer Networks and the Cytoskeleton
r
The maximum filopodium length before buckling in the
absence of cross-linking is shorter than what is observed
in vivo
Using Equation 8.42 and our estimate of Fmem, we can now calculate the maxi-
mum filopodium length before buckling occurs. We first assume that there is no
Figure 8.7 Protrusion of membrane cross-linking within the bundle. Assume that there are n = 30 actin filaments
by a cylindrical bundle. The force within the bundle, and each filament has a radius of Ractin = 3.5 nm and Young’s
exerted on the protrusion by the
membrane is proportional to the radius
modulus of Eactin = 1.9 GPa. Each filament will experience a compressive force
of the cylinder. Fmem/n, so the maximum filopodium length before buckling occurs is
π 2 Eactin I actin
Lnocl =
4 Fbuckle
πR 4actin (8.43)
π 2 Eactin
= 4 .
Fmem
4
n
Plugging in the appropriate values gives length before buckling occurs, Lnocl, is
Lnocl = 0.57 μm. Therefore, for a filopodium containing 30 actin filaments that are
not cross-linked, it can only grow to be 0.57 μm long before it will buckle. This
length is substantially shorter than the 1–5 μm long filopodia observed in vivo,
implying that in the absence of cross-linking, filopodia do not possess adequate
mechanical integrity to extend. Next, let us examine how this length is altered
when we take cross-linking into account.
( )
2
nπR 2actin = π nRactin . (8.44) Figure 8.8 Two different
approximations for analyzing the
filopodium. In the first (left) we assume
The term in parenthesis on the right-hand side of Equation 8.44 yields the radius no cross-linking. In the second (right) we
of a cylinder with the same cross-sectional area of n actin filaments, which is the assume that the bundle is tightly
effective radius of the bundle, cross-linked, resulting in behavior similar
to that of a single, large cylinder.
πR 4bundle
I bundle =
4
(8.46)
( ).
4
π nRactin
=
4
π 2 Eactin I bundle
Lcl =
Fbuckle
(8.47)
( )
4
π nRactin
π Eactin
2
= 4 .
4Fmem
Substituting the appropriate values gives Lcl = 3.2 μm. In other words, a highly
cross-linked filopodium, can grow to be 3.2 μm before buckling. Comparing
Equations 8.43 and 8.47, we find that the ratio of Lcl/Lnocl is n . This indicates that
the more filaments that are in a filopodium, the more pronounced the difference
in the buckling lengths between the cross-linked and un-cross-linked filopodia.
What can be learned from our analysis? We have already learned that filopodia of
1–5 μm are commonly observed in vivo. Our analysis show that without cross-
linking, filopodia will buckle before they are able to reach these lengths. However,
if they are highly cross-linked, they are much more suited to resist buckling at that
length. Filopodia much longer than Lcl (on the order of tens of micrometers) have
been observed on occasion. This suggests that other mechanisms may act in con-
cert with cross-linking to stabilize the actin bundles within filopodia, such as con-
straints on transverse deformations by the membrane sheath surrounding the
bundle.
(A) must be able to squeeze through tiny capillaries (many of which have diameters
that are smaller than the cells themselves), and then return to their original
shape. This behavior depends in large part on the cytoskeletal mechanics of
these cells.
In red blood cells, the cytoskeletal network is bound to the membrane in a two-
dimensional network. The main constituent of the network is a polymer called
spectrin. The spectrin polymers are bound together at distinct vertices, or “junc-
tions”. At each vertex is a junctional complex consisting of several proteins
(including F-actin) that serve to cross-link the spectrin polymers together, as
well as anchor the cytoskeletal network to the membrane. There are also protein
complexes along the length of the spectrin polymer that anchor it directly to the
membrane. Typically, six spectrin polymers radiate from each junction. Such a
junction is considered to exhibit sixfold connectivity. Fourfold connectivity, or
four spectrin polymers radiating from each junction, has also been observed
(B)
Q1 (Figure 8.9).
The question we seek to address is what are the structural and functional impli-
cations of these different connectivities? As the analytical and numerical
tools for understanding the mechanics of the red blood cell’s cytoskeleton
become more sophisticated, the impact of such distinct microstructures on the
mechanical behavior of red blood cells is becoming better understood.
However, we can also learn a lot from simple analyses. In the following section,
we estimate the mechanical properties of a model red blood cell’s cytoskeleton
from the mechanical behavior of its polymers and its microstructure. By seeing
how these properties change with sixfold and fourfold connectivity, we also
can see why, for red blood cells, sixfold connectivity may be functionally
advantageous.
∂u ∂v
γ = +
∂y ∂y
δ
= (8.48)
h
= tan γ
≈γ,
where the last line is the small angle approximation. Equation 8.48 implies that for
small shear deformations, the (engineering) shear strain is simply the angle
formed with the vertical, γ. The shear modulus G relates the shear stress τ to (engi-
neering) shear strain γ,
τ = Gγ . (8.49)
BIOMECHANICAL FUNCTION AND CYTOSKELETAL STRUCTURE 241
h Ks γ
In a manner analogous to Equation 8.33, the strain energy density for an isotropic
linearly elastic material under shear strain γ is
1 2 (8.50)
w= Gγ ,
2
implying that
∂ 2w (8.51)
G= .
∂γ 2
If we assume that the depth d is very small, we can treat the block as a two-
dimensional structure. Instead of a shear stress (with units of force per area), we
can define a shear force per unit length ns such that
ns = τd , (8.52)
assuming that the shear stress is constant through the depth. The shear force per
unit length can then be related to the engineering shear strain as
ns = K sγ , (8.53)
Similar to Equation 8.50 the areal strain energy density, or the strain energy per
unit undeformed area (as opposed to volume) is
1
Wa = K sγ 2 (8.55)
2
and the areal strain energy density can be differentiated to obtain the shear mod-
ulus as
∂ 2Wa (8.56)
Ks = .
∂γ 2
242 CHAPTER 8: Polymer Networks and the Cytoskeleton
2
R
Figure 8.11 Schematic h = R02 − 0
demonstrating justification for 2
choice of unit cell. For our analysis, (8.57)
we assume that the unit cell is an 3 R0
equilateral triangle with spring = .
constant ksp/2 (A). The spring constant 2
is ksp/2 when this unit cell is
superimposed with similar unit cells
If we now displace the top vertex by a small amount 𝛿 such that the triangle under-
(B). The result is a sixfold network with goes shear strain γ (Figure 8.12), then
spring constant ksp (C).
δ
tan γ =
h
(8.58)
2δ
= .
3 R0
3 R0γ
δ = . (8.59)
2
R0
BIOMECHANICAL FUNCTION AND CYTOSKELETAL STRUCTURE 243
Under this deformation, the left and right diagonal springs lengthen and shorten,
respectively, while the bottom spring is unchanged. The deformed length R of the
left diagonal spring can be found geometrically,
2
R
R = h2 + 0 + δ
2
2
3 R0 R0
2
= + + δ
2 2
(8.60)
= R 20 + R0δ + δ 2
R 20δ R 20δ 2
= R 20 + +
R0 R 20
δ δ2
= R0 1 + + 2.
R0 R 0
δ
R ≈ R0 1 + . (8.61)
R0
2
δ δ δ2
1 + 2 R = 1 + R + 4 R 2
0 0 0
(8.62)
δ
≈1+ .
R0
2
δ
R ≈ R0 1 +
2 R0
(8.63)
δ
≈ R0 + .
2
This implies that, to first order, the left diagonal spring increases in by 𝛿/2. It can
be shown similarly that the right diagonal spring shortens by the same amount.
We know that the change in strain energy for a spring with constant k, original
length R0 and new length R is
k( R − R0 )2
Wsp = .
2 (8.64)
The total change in strain energy is the sum of the change in strain energy for all
three springs. Assuming a spring constant of ksp/2 for our network, this expression
244 CHAPTER 8: Polymer Networks and the Cytoskeleton
becomes
1 ksp 1 ksp
(
( R0 + δ 2) − R0 ) (
( R0 − δ 2) − R0 )
2 2
= + +0
2 2 2 2 (8.65)
kspδ 2
= .
8
The areal strain energy density can be found by dividing W by the undeformed
area of the triangle,
W left
wa =
Asp
kspδ 2
= 8
1
R0h (8.66)
2
kspδ 2
= 8 .
1 3 R0
R0
2 2
We can write the areal strain energy density in terms of γ by substituting our rela-
tionship between 𝛿 and γ found in Equation 8.59 into Equation 8.66, which becomes
3kspγ 2 (8.67)
wa = .
8
∂ 2wa
Ks =
∂γ 2
(8.68)
3ksp
= .
4
Upon inspection of Equation 8.68, there are several interesting things to note.
First, there is no explicit dependence of the shear modulus on the number, den-
sity, or volume fraction of filaments. This is because sixfold connectivity necessi-
tates a fixed relationship between filament length L and the filament density. In
particular, if we model each spectrin polymer as a rod with length L and cross-
sectional area A, and we assume that the cytoskeletal network/membrane has a
depth d, then the volume fraction scales as
In other words, the volume fraction scales linearly with (1/L). Because ksp ~ 1/L,
this implies that Ks ~ ρ. Another interesting observation is that our prediction for
Ks compares remarkably well with experiments. Substituting b = 30 nm, T = 300
K, L = 200 nm for spectrin yields Ks = 0.9 μN/m. Experimental measurements of
Ks of red blood cell cytoskeletons in which the membrane was removed showed
BIOMECHANICAL FUNCTION AND CYTOSKELETAL STRUCTURE 245
average values of the shear modulus to be Ks = 2.4 μN/m. This is roughly three
times our predicted value, but still is in agreement to an order of magnitude. What
are some potential sources of error? One may be our approximation of spectrin as
an ideal chain. Remember that when R approaches L, the stiffness of a real chain
will increase, while the stiffness of an ideal chain will not. Therefore, the Gaussian
approximation is best when L ⪢ R, whereas for the red blood cell cytoskeleton, L
is only a little more than twice that of R.
if we ignore higher-order terms. What this means is that, to first order, the left
and right springs do not change length. This implies that the change in strain Figure 8.13 A simple model of
energy will be effectively zero, and so Ks = 0. This result arises because the stiff- fourfold symmetry. For fourfold
ness in shear comes from springs changing in length. However, if the springs do symmetry we utilize a simple square
not change length, then the network cannot resist shear, and so it has no effec- network of springs undergoing shear by
displacing the top surface by a small
tive stiffness. The result of this analysis is that sixfold connectivity is advanta- amount δ.
geous for red blood cells compared with fourfold connectivity, because sixfold
connectivity allows resistance to shear. Such resistance is important in allowing
red blood cells to squeeze through tiny capillaries and return to their original
shape.
Key Concepts
• Polymer network models aim to capture the • Filament cross-linking can have a substantial impact
aggregate behavior of large numbers of individual on the mechanical behavior of a network independent
filaments. They can be used to model cell behavior of filament properties and density.
under different underlying assumptions. • Strain energy can be used to calculate the moduli for
• Even if explicit relationships between mechanical affine materials, even if they are not isotropic.
properties and network parameters cannot be found, Anisotropic network behavior can be estimated from
scaling relationships can be identified. The cellular filament orientation.
solids model predicts a squared modulus–density • The geometric arrangement of polymers can alter the
relationship for bending dominated filament loading resistance of the macrostructure to deformation.
and a linear one for axial loading. Sixfold symmetric (triangular) networks tend to be
• Tensegrity models have no bending and rely on stiffer in shear than fourfold (square) symmetric
tensile and compressive elements forming a stable networks.
structure. They predict linear scaling of modulus with
both density and pre-load. They also predict the
action-at-a-distance phenomenon.
Problems
R
r
(b) W hat happens to the Young’s modulus if we 8. Discuss the following statements as they support or
increase the temperature? Why? oppose tensegrity as a model.
(c) Now, repeat your answer to (a) and (b), but this (a) The notion of action at a distance applies to a
time assume that the polymers have a long simple beam embedded in a wall. Push down
persistence length compared with their contour on the beam at the free end and there must
length. Model them as a network of rods with E be a reaction at the embedded end. If the wall
of 1 GPA and diameter of 10 nm. Assume that the is flexible enough, you will see substantial
junctions are freely rotating pin joints. deformations at considerable distance from the
(d) For your answer in (c), how would you expect the applied load.
Young’s modulus of the material to scale with the (b) If you sever a single actin filament within a living
modulus of the polymer and volume fraction of cell, the filament clearly retracts, and the local
the polymer. Only present a verbal argument. No structure of the cell changes within a short
calculations are required. time. However, within a fixed cell, severing an
(e) How would you expect the shear modulus of the actin filament does not result in any dramatic
material to scale with the modulus of the polymer changes.
and volume fraction if the junctions were not
pinned (they are bound such that the angles the 9. Assuming a linearly elastic isotropic material with
polymers make with each other is fixed). three-dimensional deformations (small deformations),
write the strain energy density in terms of strains and
7. Examine the network response of fourfold versus constants E and Poisson’s ratio.
sixfold connectivity for a normal stress, instead of shear
stress. Is there an advantage to one versus the other 10. Using an approach similar to that in the bending-
under normal stress? You may assume the normal dominated cellular solids model, derive the
stress results in similar strains for both cases. relations in Equations 8.8 and 8.9.
Annotated References
Boal D (2001) Mechanics of the Cell. Cambridge University Press. An Mogilner A & Rubinstein B (2005) The physics of filopodial protru-
excellent source for detailed analyses of red blood cell cytoskeletal sion. Biophys. J. 89, 782–795. Examines the mechanics and spatial-
mechanics. The analysis of sixfold and fourfold connectivity was temporal dynamics of filopodia. The analysis of filopoidal buckling
based on developments in this text. was based on treatments in this study.
Byers TJ & Branton D (1985) Visualization of the protein associa- Rubenstein B & Colby RH (2003) Polymer Physics. Oxford University
tions in the erythrocyte membrane skeleton. Proc. Natl. Acad. Sci. Press. Gives a detailed treatment of the elasticity of rubber net-
USA 82, 6153–6157. This references contains excellent images of works.
the erythrocyte spectrin cytoskeleton, including fourfold and sixfold Satcher RL & Dewey CF (1996) Theoretical estimates of mechani-
connectivity. cal properties of the endothelial cell cytoskeleton. Biophys. J. 71,
Evans E & Yeung A (1989) Apparent viscosity and cortical tension of 109–118. Applies cellular solid theory to estimate the elastic moduli
blood granulocytes determined by micropipette aspiration. Biophys. of the cytoskelton. This article details some of the scaling analysis
J. 56, 151–160. Gives membrane tension estimates for use in filopo- and experimental support used in the cellular solids model.
dia buckling analysis. Shao EY & Hochmuth RM (1996) Micropipette suction for measuring
Kamm RD (2005) Molecular, Cellular, and Tissue Biomechanics (lec- piconewton forces of adhesion and tether formation from neutro-
ture notes from course number 20.310, Massachusetts Institute of phil membranes. Biophys. J. 71, 2892–2901. This article gives esti-
Technology, Cambridge, MA). Several of the treatments in this chap- mates of force required to form filopodium-like membrane tethers
ter were based on course notes developed by Dr. Kamm for this within neutrophils.
graduate course. Specific material includes the scaling analysis of Stamenović D & Coughlin MF (1999) The role of prestress and
cellular solids and several problems. architecture of the cytoskeleton and deformability of cytoskeletal
Kwon RY, Lew AJ & Jacobs CR (2008) A microstructurally informed filaments in mechanics of adherent cells: a quantitative analysis.
model for the mechanical response of three-dimensional actin net- J. Theor. Biol. 201, 63–74. An excellent review of scaling analyses
work. Comput. Meth. Biomech. Biomed. Eng. 11, 407–418. Gives for cellular solids and tensegrity structures. The scaling relationship
details of the nonanisotropic affine network model, as well as an for tensegrity structures was obtained from this study.
extension of the model that accounts for some degree of non-affine Wang N, Naruse K, Stamenović D et al. (2001) Mechanical behav-
deformations. ior in living cells consistent with the tensegrity model. Proc. Natl.
Acad. Sci. U. S. A. 98, 7765–7770. This journal article details one
Mejillano MR, Kojima S, Applewhite DA et al. (2004) Lamellipodial
of the most compelling experiments that supports the tensegrity
versus filopodial mode of the actin nanomachinery: pivotal role
hypothesis.
of the filament barbed end. Cell 118, 363–373. This reference
assesses the role of various actin-binding proteins on the forma- Warren WE & Kraynik AM (1997) Linear elastic behavior of a low-
tion of lamellipodia and filopida and proposes a model whereby density kelvin foam with open cells. J. Appl. Mech. 64, 787–794. This
lamellipodial versus filopodial organization can be selected via journal article provides some very comprehensive mathematical
regulation of these proteins. analysis of cellular solids.
This page intentionally left blank
to match pagination of print book
CHAPTER 9
Mechanics of the Cell
Membrane
T he cell membrane is much more than a passive “bag” separating the cyto-
plasm from the environment. It is a heterogeneous, regulated barrier that
allows for both active and passive transport of substances between the inside and
outside of the cell. Its mechanical integrity is fundamental to its barrier function.
Bending and stretching of the membrane is centrally involved in exocytosis and
vesicle budding, as well as in fusion and viral invasion. It also incorporates struc-
tures for interacting with the extracellular matrix, other cells, and various com-
pounds in solution. Proteins trapped in the membrane allow for signaling between
the inside and outside of the cell. Understanding the biology and mechanics of
the membrane can foster an appreciation of the complexity of this fascinating cel-
lular component. In this chapter, we discuss the cell membrane’s structural
organization, how the two-dimensionality of the membrane limits diffusion, and
how the barrier function can be understood. We end with a formal treatment of
the cell membrane’s mechanical function.
CH3
252 CHAPtER 9: Mechanics of the Cell Membrane
internal fluid volume, known as the cytoplasm. However, the plasma mem-
brane is not a total barrier, and certain molecules can diffuse across it. This
semipermeable behavior is critical for its proper function. The barrier proper-
ties of the plasma membrane are partly a result of the thin hydrophobic layer in
the interior tail region of the bilayer. Charged or polar molecules are less likely
to cross this region. The plasma membrane also forms a barrier against large
molecules. The barrier function of the plasma membrane is further increased
by the activity of flippases, which are proteins that concentrate negatively
charged phosphatidyl serine to the inner leaflet, forming an additional charged
boundary. Permeability of the cell membrane to a given substance is hard to
predict, but is generally a function of size, charge, and solubility. Water can
cross the plasma membrane and does so by osmosis, the relatively slow transfer
from the low solute concentration side (hypotonic) to the high solute concen-
tration side (hypertonic). In general the cell’s cytoplasm is hypertonic. Water
movement across the plasma membrane is also facilitated by aquaporins, small
protein complexes that form tiny holes in the membrane to regulate the pas-
sage of water molecules.
To calculate the entropy we assume that in this state, the amphiphiles are suffi-
ciently dispersed to behave like an ideal gas. The entropy per molecule for an ideal
gas is
5 ρ 3
Sideal = kB − ln 3/ 2
, (9.2)
2 (2πmkBT )
where ρ is the number of molecules per unit volume, m is the mass of each mol-
ecule, and is Planck’s constant. Now, we can write the free energy per molecule
in the aqueous state as
5 ρ3
Ψ = W − TSideal = γ intlcnc − kBT − ln 3/ 2
. (9.3)
2 (2πmkBT )
Remember that this free energy in the aqueous state was constructed with the
condensed state defined as the reference state with zero free energy. So, the tran-
sition between the condensed and aqueous states occurs when Ψ = 0, or when
5 ρ 3
γ intlcnc = kBT − ln 3/ 2
. (9.4)
2 (2πmkBT )
254 CHAPtER 9: Mechanics of the Cell Membrane
The density ρ at this transition gives the critical density of molecules at which
aggregation occurs. Solving for ρ, we obtain
γ int lc nc
5/ 2 − k T (9.5)
ρ = A e B
,
where
Although the mass m depends on nc, for physiological values of γint, lc, and nc, the
exponential term typically decreases much faster than A with nc. Therefore, we can
see that for constant values of γint and lc, increasing the number of carbon atoms nc
decreases the aggregation density. Intuitively, this seems reasonable, as having
longer tails would increase the interfacial energy of the molecule with water and
make it more energetically unfavorable to remain in the aqueous state. Similarly,
phospholipids with two tails have a lower aggregation density compared with their
single-tailed counterparts, owing to their higher interfacial energy. What is per-
haps not as intuitive is the profound impact that increasing nc can have on lower-
ing ρ. The predicted exponential scaling of ρ with nc, has, in fact, been observed
experimentally. Specifically, increasing nc by x atoms decreases ρ by a factor of ex.
If we were to extend the hydrocarbon chain by just two atoms (let us say from 10 to
12), this would result in an almost tenfold decrease in the aggregation density.
16γ int lc
= e8 ≈ 3000
5/ 2 −
k T
First, we must calculate the mass of the longer lipid. We e B
4πR 2 (9.7)
n=
Ah
MEMBRANE BARRIER FUNCTION 255
4πR 3
n= (9.8)
3V
from the volume. Setting Equations 9.7 and 9.8 equal, we see that for a micelle
with a given radius R, Vc and Ah must be related as
Now, let l = lcnc be the hydrocarbon chain length. We assume that the interior of
the micelle cannot contain voids, owing to the enormous energy that creation of a
vacuum would require. This implies that the radius of the micelle must be less
than or equal to the hydrocarbon chain length, R ≤ l. Equation 9.9 becomes
Ah /Ae ≥ 3, (9.10)
n = wh/Ah (9.11)
Ah = 2Vc /t . (9.13)
The thickness must be less than or equal to twice the hydrocarbon chain length, or
t ≤ 2l . (9.14)
In this case,
Vc /l = A0 ≤ Ah . (9.15)
This indicates that when the effective area of the hydrophobic region approaches
that of the head region, based on packing, a bilayer shape is preferred.
Section 5.6. The connection between random molecular processes (the ran-
dom walk) and the net macroscopic or continuum-level behavior (the diffusion
equation) was one of Albert Einstein’s many important contributions. Owing to
entropy and Brownian motion, there are many applications of random walks in
biology. In this section we are going to apply the continuum level description
to understand how molecular transport works at the level of the cell and the
role of the membrane in that process. We begin by revisiting Fick’s first and
second laws of diffusion.
captured. At a position x, let the mean time of capture be The above equation describes the wait time for a single
Tc(x). Let us start with the discrete random walk and particle. To generalize to higher dimensions, you can
then extend it to a differential form. Assume that I wait a replace the second term (d2Tc/dx2) with the Laplacian of
small time Δt, the molecule will be at x + b or x − b with appropriate dimension. To solve this equation one needs
equal probability. We can write this recursive relationship boundary conditions. For this example, let us assume
that the absorber is a sphere of radius a located within a
Tc ( x ) = ∆t + 0.5(Tc ( x + b) + Tc ( x − b)).
larger impermeable sphere of radius b (a b). On the
surface of the absorber, the wait time is obviously zero
Rearranging, (Tc(a) = 0). At the second boundary, Tc has zero gradi-
ent. We leave it as an exercise to show that integrating
over the known volume the mean wait time is
0 = 2∆t + (Tc ( x + b) − Tc ( x )) − (Tc ( x ) − Tc ( x − b))
b3
Tc = .
3Da
2∆t 1 1
0= + (Tc ( x + b) − Tc ( x )) − (Tc ( x ) − Tc ( x − b)).
b b b The mean capture time is
Letting b approach zero, we obtain b2 b
Tc = ln
2D a
2∆t 1 dTc ( x ) dTc ( x − b)
0= + − .
b2 b dx dx for a particle in a two-dimensional surface with a small
circular absorber of radius a. Therefore, as the distance
Letting b approach zero once more increases, diffusion limits on kinetics increases much
more slowly if the reaction is confined to a membrane
1 d 2Tc
0= + . than if it needed to occur in free space.
D dx 2
∂ 2C ∂ 2C ∂ 2C
∇ 2C = + + . (9.21)
∂x 2 ∂y 2 ∂ z 2
258 CHAPtER 9: Mechanics of the Cell Membrane
1 ∂ 2 ∂C a
R = 0. I = − DC 0 4πR 2 = −4πDC 0a.
R 2 ∂R ∂R R2
with boundary conditions C(∞, t) = C∞ and initial condi- has a rough idea what D is, one can design an experi-
tions C(R,0) = C0(R) the initial photobleach profile. ment to time the FRAP.
Typically,
Let us work a numerical example. Assume that αP0C0 = 1
C ( R , 0) = C 0e −T ϕ ( R ) (fluorescence units), D = 10−11 m2/s, and Tφ(0) = 0.1.
Determine the half-life of the photobleached region
relative to F(0), if the photobleaching region is a disk of
where T is the time interval for the bleaching and φ(R) radius 5 μm.
is a scaled excitation intensity from the laser (φ = 0 for
R > w, where w is the disc radius, and φ = constant for Because Tφ(0) 1, we can apply the solution for F(t),
R ≤ w. The constant depends on the laser power, a which gives us
scaling factor, and the characteristic size of the laser
beam.) 1
1−
F (t ) = 4tD .
However, what we really want is the diffusion constant 20 1 + 2
D, and what we really measure is the fluorescence w
profile
Substituting the provided numbers,
∫
F (t ) = αϕ ( R)C( R ,t )dA ,
F (t ) = 1 −
1
,
(20 + 32t )
where α is a scaling factor that further corrects for
imaging attenuation and the integral is over the pho- 1 19
with t in seconds. We can quickly get F(0) = 1 − =
tobleached area. The solution for F(t) can be found 20 20
with Fourier transforms and series analysis. If one (this is not much photobleaching at all). To get to half of
uses the assumption that Tφ(0) is much less than this value, we solve
unity (slight bleaching) then the solution can be
expressed as 39 1
= 1−
40 (20 + 32t )
Tϕ (0)
1−
F (t ) = (α P0C 0 ) t , for t. One gets t = 5/8 s, or a little over half a second. This
21 + is very fast and unlikely to be something that would be
τd
useful using FRAP as modeled. One can either increase
the intensity of illumination to photobleach out the sam-
where P0 is the laser power and τd = w2/4D is the charac- ple more (in which case the expression no longer holds,
teristic time for diffusion. By measuring F(t), one can and one needs to use a different approach) or use a
obtain D for the protein of interest. Alternatively, if one smaller photobleaching area.
Nota Bene With this as our motivation, the goal of this section is to develop the equations
that describe the mechanics of the continuum behavior of the cell membrane. As
Differential geometry is needed
this section is mathematics-heavy, we will consider systems of increasing levels of
to express the mathematics of
shells. The reason we restrict most of complexity to develop some degree of intuitive understanding before presenting
our analyses to plates is that in a the general governing equations.
“full-blown” analysis, equations must
be developed in a curved system,
rather than the “ordinary” planar
Thin structures such as membranes can be treated as
Euclidian coordinate systems with plates or shells
which you are probably familiar. The
mathematical study of curved The defining characteristic of a membrane is its very thin structure relative to
coordinate systems is known as the overall size of the cell. The cell membrane lipid bilayer has a thickness of
differential geometry. It is remarkably approximately 7 nm. The cell itself is typically on the order of micrometers. By
similar in its use in shell mechanics restricting the dimensionality of the problem, certain kinematic assumptions
and general relativity. At its heart is the can be made that will greatly simplify the problem from that of general three-
definition of the covariant derivative,
which is the derivative along the
dimensional continuum mechanics. In other words, we can approximate the
curved tangents of the shell’s surface. membrane as a curved two-dimensional structure, or, from Chapter 3, a shell.
It is defined in terms of Christoffel You might be more familiar with a plate, or a flat two-dimensional structure. In
symbols. They represent an additional this section, for simplicity, we will assume that the domain or subdomain of the
term in the derivative that occurs problem is flat (that is, a plate). One exception is that in certain cases we will
because of the curved coordinate consider spherical shapes (shapes that can be described by radii in the x- and
space and greatly simplify complex
y-directions).
calculations. However, they are quite
beyond the scope of this treatment.
The mathematics to describe behavior in curved coor- coordinates, and when they occur in pairs, a summa-
dinate systems was developed to a great extent by tion is implied. The square of the magnitude of a vector
Einstein. He needed this description to cast the formu- v would be denoted as vivi = v1v1 + v2v2 + v3v3 = vxvx +
lation of general relativity theory in which space is vyvy + vzvz. This is a very powerful and compact way to
actually curved owing to the gravitational effect of represent complex vector and tensor manipulations.
mass. He developed a notational system (known as Any advanced treatment of continuum mechanics
inidical notation) in which coordinates are denoted beyond the level used in this text will generally adopt
with numbers (1, 2, or 3) rather than letters (x, y, and z). this notation out of expediency.
In his system, indices are introduced to represent the
As before, our continuum mechanics analysis involves three different parts, kin-
ematics, constitutive equations, and equilibrium. For the material model we are
again going to assume generalized Hookean behavior. We will consider equilib-
rium in distinct parts, one for the in-plane forces and one for the out-of-plane
forces. Let us begin our discussion of membrane mechanics by making kinematic
assumptions that take advantage of the thin shell assumption.
These kinematic assumptions suggest a particular way of expressing the deforma- Nota Bene
tion of the shell. In particular, it makes sense to talk about the deformation due to
The Kirchhoff hypothesis. The
motion in the plane of the shell (we assume this is the x–y-plane) and motion due to
shell kinematic assumptions are also
transverse displacements and the subsequent rotation of the normals (Figure 9.5). known as the Kirchhoff hypothesis.
This is analogous to our separation of axial deformation and bending deformation The lines perpendicular to the shell
when we considered a linear element. We will denote the total deformation as utot surface are known as directors.
and vtot in the x- and y-directions, respectively. They are simply the sum of the in-
plane deformation and the displacement due to rotation of the shell’s mid-surface.
Our kinematic assumption amounts to just assuming that the deformations take a
particular form that is a simple extension of the beam kinematics,
dw
u tot ( x , y , z ) = u( x , y ) − z
dx
dw (9.22)
ν tot ( x , y , z ) = ν ( x , y ) − z
dy
w tot ( x , y , z ) = w( x , y ).
From these, we calculate strain from Equations 3.48, 3.49, and 3.52
du d 2w 1 du dν d 2w
ε xx = − z 2 ε xy = + − z
dx dx 2 dy dx dxdy
dν d 2w (9.24)
ε yy = − z 2 ε xz = 0
dy dy
ε zz = 0 ε yz = 0.
In this development we have assumed that the defor- not be a valid assumption, particularly for cells. One
mations are small enough that our infinitesimal strain more advanced approach, known as von Kármán shell
measures are adequate. However, what may not be theory, assumes infinitesimal strains, but moderate
obvious is that we have also assumed that the rotation rotations. This introduces an extra quadratic terms in
of the normals is also infinitesimal. Often, this may the strains.
εip εθ Notice that the in-plane strain components (εxx, εyy, εxy) contain two parts, one
constant through the thickness, and one that varies linearly through the thickness.
These correspond to the strains due to in-plane motion of the normal lines, εip, and
strains due to rotations of the normal lines, εq (Figure 9.6). The in-plane strains are
edge +
related to tension and shear within the plane of the shell. The linear strains are
related to out-of-plane bending. Therefore, shells have three different deforma-
tional modes, in-plane tension (or compression), in-plane shear, and bending. The
total deformation is simply the linear superposition or sum of these three modes.
Figure 9.6 Decompositon of
deformation modes. As a result of our
kinematic assumption, the deformation A constitutive model describes material behavior
of a shell can be separated into a
component that is constant through the As in the past, we are going to assume a generalized Hookean material response
thickness, εip, and a component that (Equation 3.58). Owing to the simplified kinematics, it takes on a particularly sim-
varies linearly through the thickness, εθ. ple form:
σ xx = 2 µε xx + λ (ε xx + ε yy )
σ yy = 2 µε yy + λ (ε xx + ε yy ) (9.25)
τ xy = 2 µε xy .
Nota Bene The equilibrium condition simplifies for in-plane tension and shear
The two meanings of the term
“membrane.” As we briefly When we considered kinematics, we talked about the deformation modes.
mentioned at the end of Section 3.4, Similarly, we can decompose our treatment of equilibrium by considering the dif-
in structural mechanics terms, a ferent loading modes in turn. Specifically, it is helpful to consider in-plane load-
curved two-dimensional structure ing separately from bending. Starting with in-plane loading, this mode is restricted
that can support bending moments to be within the plane of the shell locally. Although the shell may bend, we assume
is called a shell. A curved two-
that there is no energy, stress, or moment associated with this bending. This is
dimensional structure that cannot
withstand bending is called a strange. What does it mean to say we will allow bending, but ignore bending
membrane. Unfortunately, this is a stresses? Actually, this is not as unusual as you might think. Consider a thin plastic
very different use of the word from bag like you might get from the grocery store. The plastic in these bags is so thin
its biological meaning. So be alert to (about 30 μm) that they bend with the slightest touch. Although there is bending,
the context in which these terms are the resultant moments are so small that they can effectively be ignored. In addi-
used. Mechanical analysis of tion, since the lipid bilayer acts as a two-dimensional fluid, the two leaflets can
membranes can be thought of as a
generalization of the analysis of
slide over each other further reducing the resistance to bending.
strings to two dimensions and finds We will consider two particular cases of in-plane loading, shear and tension. Let
application in the analysis of rubber us start our consideration of equilibrium by applying it to an imaginary free body.
sheets and drums.
An infinitesimal element of the membrane subjected to in-plane forces is illus-
trated in Figure 9.7.
As we learned in Chapter 3, we need to apply the equilibrium conditions to
forces rather than to stresses. To accomplish this, we must define resultant
forces in terms of stress. The resultant forces are the equivalent force that results
MEMBRANE MECHANICS I: IN-PLANE SHEAR AND TENSION 263
Nxx
Nyy
from the net action of the stresses acting on each edge of the element. Remember Nota Bene
that stress is force per unit area. So, just as in the rod or column in Section 3.2,
Pressure assumption. In this
the resultant force on each edge is the integral of the stress over the area of the
analysis we only consider the
edge. However, there are some important differences. First, because the shell is situation in which the pressure is in
two-dimensional, we must introduce the relevant x- and y-components denoted the −z-direction ( px = py = 0). We can
by subscripts. Secondly, rather than deal with the total force resultant on an always construct a coordinate
edge, N, it is more convenient to divide by the edge width, b, such that n = N/b. system with x and y aligned in the
The concept of resultant force per unit length (width) can be thought of as a plane of the membrane. We are
generalization of surface tension that we discussed in Section 1.3. We define the ignoring the effect of rotation of the
normals, which would introduce what
resultant forces per length to be
is known as a follower force and
h /2 h /2 h /2 greatly complicates the math.
∫ ∫ ∫τ
However, as long as the rotations are
nxx = σ xx dz nyy = σ yy dz nxy = xy dz. (9.26) small, we can assume that the
− h /2 − h /2 − h /2
pressure acts only in the z-direction.
Note that sometimes (particularly in the structural mechanics literature) these are
called the stress resultants.
With these definitions, we are now ready to derive the equilibrium conditions by
constructing, as usual, a free-body diagram. If we consider an element of the
membrane (Figure 9.8) that is not accelerating, equilibrium tells us that the
pz
dw d2w dw d2w
—+— dx —+— dy
dx dx2 dy dy2
x y
dw d2w dw d2w
— + — dx — + — dy
dx dxdy dy dxdy
dnxx dnyy
nxx + — dx nyy + — dy
dx dnxy dnyx dy
nxy + — dx nyx + — dy
dx dy
264 CHAPtER 9: Mechanics of the Cell Membrane
resultant forces must sum to zero. This situation is slightly more complicated,
because the membrane is not flat and the force resultants therefore change not
only with respect to dx and dy, but also dz.
Figure 9.8 is more complex than we are used to. Owing to the curvature of the
membrane, we now have to consider the local slopes. At the x = 0 edge, the
slope of the surface in the y-direction is dw/dy. At the x = dx edge, we must also
account for the rate of change in the slope. On this edge, the slope of the sur-
face in the y-direction is even more complicated, dw/dy + (d2w/dy2) dy. For the
summations in the x- and y-directions, we ignore this effect, however, for the
summation in the z-direction, it cannot be ignored. If you are careful, you
should find,
dnxx dnyx
∑ f x ⇒ − nxx dy + nxx + dx dy − nyx dx + nyx + dy dx = 0
dx dy
dnyy
dy dx − nxy dy + nxy +
dnxy
∑ f y ⇒ − nyy dx + nyy + dx dy = 0
dy dx
dw dnxx dw d 2w
∑ fz ⇒ − nxx dy + nxx +
dx + dx 2 dx
dx dy
dx dx
(9.27)
dw dnxy dw d 2w
− nxy dy + nxy + dx dy + dx
dy dx dy dxdy
dw dnyy dw d 2w
− nyy dx + nyy + dy dx + dy
dy dy dy dy 2
dw dnyx dw d 2w
− nyx dx + nyx + dy dx + dy + pzdxdy = 0.
dx dy dx dxdy
dnxx dnxy
∑ fx = 0 ⇒ + =0
dx dy
dnyx dnyy
∑ fy = 0 ⇒ + =0
dx dy (9.28)
dw dw dw dw
d nxx + nxy d nxy + nyy
dx dy dx dy
∑ fz = 0 ⇒ + + pz = 0.
dx dy
The first two equations of Equation 9.28 are analogous to the stress equilibrium
condition we derived in Equation 3.46, applied to a two-dimensional membrane
structure. However, the third equation is new and shows how the force equilib-
rium condition in the transverse direction relates membrane tension to pressure.
We can further simplify by writing out the derivatives
d 2w d 2w d 2w
nxx 2
+ 2nxy + nyy 2 + pz = 0. (9.29)
dx dxdy dy
We can express Equation 9.29 in terms of curvature, just like we did for the beam,
except that now there are curvatures in different directions.
where
d 2w d 2w d 2w
κ xx = 2
, κ xy = , κ yy = . (9.31)
dx dxdy dy 2
dnxx dnxy
+ =0 Nota Bene
dx dy
Moment balance. When we applied
dnxy dnyy the equilibrium equations, we did not
+ =0 (9.32) consider the implications of the
dx dy
moment summation. Particularly
useful is the observation that the
d 2w d 2w d 2w
nxx + 2n xy + n yy + pz = 0. in-plane shear resultants must be
dx 2 dxdy dy 2 symmetric, nxy = nyx, which is a
consequence of the requirement
To gain a deeper understanding of this equation, we will take a closer look at this that the moments about the z-axis
expression and elaborate it for two special cases, the case of planar shear and sum to zero.
equibiaxial tension.
n − nyy
n xx = nxx − xx
2
n + nyy (9.33)
n yy = nyy − xx
2
n xy = nxy .
The equilibrium condition (Equation 9.33) is not very informative. It requires that
a case of pure shear can only exist if the shear resultant is constant. However, we
can apply the constitutive equation to gain some insight, τxy = γxy. We can reformu-
late this in terms of stress resultants. Because nxy = τxyh
nxy = Gγ xy h
(9.34)
= K Sγ xy .
We have introduced a new constant KS = Gh called the membrane shear modulus,
which has units of force per unit length.
What are typical values of the membrane shear modulus for cells? The cell mem-
brane of a red blood cell has a typical value of KS = 6 × 10−6 – 9 × 10−6 N/m = 6–9
pN/μm. Why is it so small? As we have described, the lipid bilayer is formed of
molecules that are tightly bound to the layer, but easily move within the plane of
266 CHAPtER 9: Mechanics of the Cell Membrane
d 2w 1
2
= κ xx =
dx Rx
(9.36)
d 2w 1
= κ yy = .
dy 2 Ry
Therefore,
1 1
pz = n + . (9.37)
Rx Ry −
If the membrane is spherical in shape, we recover Equation 1.1 for the oil-drop
model of the cell,
2n
+ p z = 0. (9.38)
R
Notice that the law of laplace is typically expressed as 2n = pz , because the
R
curvature and pressure are oriented in different directions such that pz = –p.
MEMBRANE MECHANICS II: BENDING 267
∆A = (1 + ε )2 L2 − L2 (9.39)
and
∆A/A = (1 + ε )2 − 1 = 2ε + ε 2 2ε (9.40)
by assuming that the term involving the square of strain will be in general much
smaller than the others and can be neglected.
We can now define the material property areal expansion modulus for equibiaxial
tension in terms of the tension and areal strain,
n
KA = . (9.41)
(∆A/A)
Eh
KA = . (9.42)
2(1 − ν )
Typical values of the area expansion modulus for lipid bilayers are in the range of
KA = 0.1−1.0 N/m. The cell membrane of red blood cells, for example, has an
areal expansion modulus of approximately KA = 0.45 N/m (= 450,000 pN/μm).
This value is many orders of magnitude greater than the membrane shear modu-
lus. The bilayer is often treated as effectively inextensible. Typically, a lipid bilayer
membrane can only tolerate 4–6% areal strain before rupture. The large resist-
ance to areal change can be attributed to the changes in energy associated with
exposing the hydrophobic core of the lipid bilayer to water as the spacing between
the individual molecules is increased. Experimental approaches for measuring
KA are discussed in Section 9.6.
spectrin ankyrin How could the red blood cell maintain its characteristic bi-concave structure
rather than acting as a flaccid bag of cytoplasm?
The answer to this paradox is that cells derive their bending stiffness from
underlying cytoskeletal structures that support the bilayer. In the red blood cell,
a network of the polymer spectrin exists directly under the bilayer. It is bound to
the bilayer through periodic inclusions of ankyrin, a linker protein that attaches
membrane integral membrane to membrane-spanning proteins (Figure 9.9). As we discussed in Section 8.4,
protein band 3 the spectrin network takes on a characteristic triangular–symmetric form
Figure 9.9 Complexes of (Figure 9.10).
membrane-spanning proteins and Another way in which cells develop bending rigidity is by covering themselves
linker proteins are able to provide
with a coating of proteoglycans known as the glycocalyx. The glycocalyx can be
the cell membrane bending
stiffness. as thick as 0.5 μm. It is made up of long bottle-brush-like molecules that are
highly bifurcated and covered with a high negative charge density. This causes
intermolecular repulsion in the glycocalyx and tends to attract water molecules.
The net effect is a tendency of the glycocalyx to be highly resistant to bending. In
terms of bending, the bilayer tends to function primarily as a chemical barrier
with the structural integrity being dominated by the cytoskeleton and glycoca-
spectrum lyx (Figure 9.11).
dw dw
u = −z υ = −z . (9.43)
dx dy
From this we can calculate the strains from their continuum definitions, Equations
3.48, 3.49, and 3.52,
∂u ∂ 2w
ε xx = =− 2z
∂x ∂x
Figure 9.11 Cortical actin stiffens
the cell membrane. In this fluorescent ∂v ∂ 2w
confocal micrograph, the actin forms a ε yy = =− 2z (9.44)
∂v ∂y
dense band around the periphery of the
cell. See www.garlandscience for a color
1 ∂u ∂ v ∂ 2w
version of this figure. (Courtesy of ε xy = + =− z.
Andrew Baik, from the laboratory of X. 2 ∂y ∂ x ∂x ∂y
Edward Guo.)
As before, we can use our definition of curvature to obtain
ε xx = κ xx z ε yy = κ yy z ε xy = κ xy z. (9.45)
MEMBRANE MECHANICS II: BENDING 269
σ xx = 2 µε xx + λ (ε xx + ε yy ) = 2 µzκ xx + λz(κ xx + κ yy )
h /2 h /2
Nota Bene
mxx ( x , y ) =
∫
− h /2
σ xx ( x , y , z )z dz m yy ( x , y ) =
∫
− h /2
σ yy ( x , y , z )z dz
The units of the moment
resultants are the same as for
h /2 force. Remember that for our
x y
270 CHAPtER 9: Mechanics of the Cell Membrane
surface.
Now we are finally ready to apply the equilibrium equations. In the membrane anal-
ysis, we used force equilibrium to derive the governing equations and moment equi-
librium about the z-axis to show a symmetry condition. For plate bending, we need
to add the moment equilibrium about the x- and y-axes. Also, force equilibrium in
the x- and y-directions does not apply because there are no forces in those directions.
So, let us start with the force equilibrium in the z-direction. From Figure 9.14 we can
deduce the force equilibrium condition in the z-direction. Some simple manipula-
tion (dividing by dx and dy and canceling terms) implies
dqx dqy
+ + pz = 0. (9.49)
dx dy
dmxx dm yx
∑ Mx = 0 ⇒ + − qx = 0
dx dy
(9.50)
dm yy dmxy
∑ My = 0 ⇒ + − q y = 0.
dy dx
pz mxx
dqx dqy
x qx + — dx qy + — dy y
dx dy
dmxx dmyy
mxx + — dx myy + — dy
dx dy
dmyx
myx + — dy
dmxy dy
mxy + — dx
dx
MEMBRANE MECHANICS II: BENDING 271
These are the equilibrium equations for a plate. They are a coupled set of first-
order differential equations. You may remember that it is often possible to
replace a coupled set of lower-order differential equations with a single higher-
order differential equation. Indeed, this is the case here. We can eliminate the
transverse shear components to find a second-order equation in the moments.
It is left as an exercise for you to show that Equations 9.49 and 9.50 can be com-
bined to yield
d 2mxx d 2mxy d 2m yy
2
+2 + + p z = 0. (9.51)
dx dxdy dy 2
This is the classical equilibrium equation for a plate in terms of the distributed
moments and pressure.
Now we have all the components we need to find the governing equations for
plate bending. First, we combine our definition of moments (Equation 9.47) with
our constitutive law (Equation 9.46). Combining the integrals yields
h /2
mxx =
∫σ
− h /2
xx zdz = K B (κ xx + νκ yy )
h /2
(9.52)
m yy =
∫
− h /2
σ yy zdz = K B (κ yy + νκ xx )
h /2
1 − ν2
mxy =
∫
− h /2
σ xy zdz = K B
1+ν
κ xy ,
where
Eh 3
KB = . (9.53)
12(1 − ν 2 )
We can now insert this new expression for moments in terms of curvature into our
equilibrium Equation 9.51 to yield
d 2κ xx d 2κ xy d 2κ yy
pz = K B +2 + (9.54)
dx 2
dxdy dy 2
or in terms of displacement
d 4w d 4w d 4w
pz = K B 4 + 2 2 2 + . (9.55)
dx dx dy dy 4
This is the fourth-order differential equation describing plate bending relating the
transverse displacement to applied pressure.
This completes our development and derivation of the plate-bending equation.
However, let us try to get an intuitive feel for bending at a cellular level. The bend-
ing stiffness, KB, of lipid bilayers such as the red blood cell membrane is on the
order of 10−19 N/m. This is an exceedingly low value. Remember that KA was on
the order of 0.1 or 1 N/m. Of course, the units of these quantities are different.
Nonetheless, there are about 20 orders of magnitude between typical bending and
expansion forces. Bending forces in a bilayer are small even compared with shear.
Recall the shear stiffness, KS, is on the order of 10−6 N/m. Bending remains 13
orders smaller than this. We knew this was the case (although perhaps not how
272 CHAPtER 9: Mechanics of the Cell Membrane
d 2w d 2w d 4w d 4w d 4w
n 2 + 2
+ KB 4 + 2 2 + + pz = 0. (9.56)
dx dy dx dx dy dy 4
KB
1 ⇒ tension dominates
nλ 2
(9.57)
KB
1 ⇒ bending dominates
nλ 2
For a typical cell we might have KB = 10−18 N/m, n = 5 × 10−5 N/m, and λ = 1 μm.
Therefore, KB/[nλ2] = 0.02, indicating that, in-plane effects are more important
than bending.
polymer makes with an imaginary horizontal line at each point s, and Δθ (s)=
θ (s) − θ (0) (see Equation 7.14). In particular, we found that
−s
(9.58)
p
cos ∆θ (s ) = e .
EI
p = . (9.59)
kBT
p ∼ be 3kB T , (9.61)
AkBT π 2 /b 2 + n/K B
Aproj (n) = A − ln . (9.62)
8πK B π 2 /A + n/K B
274 CHAPtER 9: Mechanics of the Cell Membrane
Note that Aproj goes to zero when kBT/KB goes to zero, indicating that no wrinkles
will occur in the limit of zero temperature or infinite bending rigidity. Aproj also
goes to zero in the limit as n/KB gets arbitrarily large.
From this we can compute the amount of tension required to increase the pro-
jected area from a tensionless reference state,
Ak B T 1 + nA/π 2 K B
Aproj (n) − Aproj (0) = ln (9.63)
8πK B 1 + nb 2 π 2 /π 2 K B
or
k BT
∆A/A = ln(1 + nA/π 2 K B ), (9.64)
8πK B
where ΔA = Aproj(n) − Aproj(0), and we have assumed that A >> b2, the square of the
intermolecular spacing. The above expression gives the “effective” areal strain
associated with smoothing out thermal undulations under a given tension n. This
expression can be superimposed with the zero temperature areal strain given by
the areal expansion modulus (that is, the strain associated with increasing the
intermolecular distance),
k BT
∆A/A = ln(1 + nA/π 2 K B ) + n/K A . (9.65)
8πK B
An alternative approach to measuring the areal expansion where [Ci] is the molar concentration of each species
modulus is using a small-bore pipette. One can determine present. From the law of Laplace one can relate posmotic
the osmotic pressure across the membrane as a function to n, and the areal extension can be determined from the
of solute concentration and temperature as amount of membrane drawn into the pipette.
Key Concepts
• The amphiphilic nature of cell membrane • The continuum mechanical behavior of the membrane
phospholipids allows them to self-assemble into the can be decomposed into in-plane and bending
bilayer. The packing behavior of the phospholipids components. In-plane behavior can be further
contributes to the structures formed. decomposed into equibiaxial and shear responses.
• Microdomains within the membrane enhance • Derived moduli, including KA, KS, and KB, characterize
chemical kinetics. the areal, shear, and bending stiffnesses-respectively.
• Saturated and unsaturated fatty acid chains • The quantity KB/nλ2 is indicative of tension-dominated
contribute to determining membrane properties. or bending-dominated behavior.
• The fluid mosaic model describes the two- • The bilayer is relatively flexible in shear and bending
dimensional fluid behavior of the bilayer. but can be stiff with respect to areal tension.
• The membrane is semipermeable and has a critical • Entropic membrane undulations can be described
barrier function that can be described with the statistically and are related to continuum properties.
diffusion equation.
Problems
1. (a) In one dimension, the mean squared displacement (a) If you had n channels in the cell surface, where n is
of a particle undergoing a random walk varies small, what is the total current flux into the cell?
with the square root of t (time). How does the (b) If the entire sphere is acting as a sink with current
mean squared displacement vary with time in two flux I0, then the total current flux if you had n
dimensions? And in three dimensions? Why? channels—for any n—is I = I0/[1 + πb/na]. Is this
(b) L et us say you defined the diffusion velocity as the consistent with your answer from (a)? Justify your
root mean squared displacement divided by time, response.
for a given diffusion coefficient. If you do this, what (c) Using the information in (b), design a cell with
happens as t approaches 0? How can you tell if radius b = 10 μm with n channels (sinks) of radius
experimental data you are analyzing are affected a = 1 nm so that its current flux through all of the
by this problem? channels is half of the current flux if the entire cell
surface was a sink (that is, find n). You may neglect
2. For typical small molecules in water, the diffusion the local curvature in calculating the total area
coefficient is D ≃ 10−5 cm2/s. Small ions typically pass of the channels. Next, find the fraction of the cell
through channels diffusively. You can assume that the surface area that is covered with channels for this
channels are only wide enough for a single ion to pass to occur. What does this result tell you about the
through at a given time. ability of a cell to acquire needed compounds by
(a) E
stimate the time it takes for a small ion to cross diffusion through channels?
the channel, assuming there is no interference
from other ions, and the channel has the same 4. Consider the transport of a solute through a thin
length as the thickness of a cell membrane. blood vessel. The blood vessel has inner radius R1, the
(b) N
ow assume someone is building a scaled-up endothelial layer increases the radius to R2, and then
macro-cell whose radius is about 1 m. How the outer layer (matrix) further increases the radius
long would it take the same small ion to cross to R3. If the diffusion coefficient of the endothelial
the macro-channel? Is diffusion an effective layer for the solute is D1 and the diffusion coefficient
mechanism for large-scale processes? of the matrix is D2, find the total (steady-state) solute
flux out of the blood vessel, assuming the blood
3. Consider a hole of radius a in an infinitely large wall in concentration is effectively constant at C0 inside the
space. If the concentration far away from one side of the blood vessel and is zero everywhere outside (the solute
disk is C0, and is zero everywhere on the other side, the is immediately consumed). Keep in mind you need to
current flux of freely diffusing molecules passing through use the cylindrical form of the diffusion equation(s).
this hole is given by I = 4DaC0. This can be a model for
a single channel in a membrane (assuming the radius, 5. Using the same setup as in Problem 4, model the
a, is much smaller than the radius, b, of the cell itself). blood vessel as a hollow cylinder of inner radius R1
We have derived the current flux of freely diffusing and outer radius R3 but consisting of only one material
molecules passing through a sphere as I = 4πDC0b. of diffusion coefficient D, with the same flux out of
276 CHAPtER 9: Mechanics of the Cell Membrane
the vessel. Find D in terms of D1, D2, and the given increases to C0 > 0. You may assume that inside the cell,
parameters. at x → ∞, the concentration C = 0 forever.
(a) Write Fick’s second law of diffusion for this setup
6. You are developing a drug delivery system consisting of
for t > 0.
a small sphere of inner radius R, wall thickness h, and
(b) Define a new variable α so that
wall diffusion coefficient D. Assume that the internal drug
concentration is constant at C and the concentration x
α = .
outside the sphere is zero everywhere (used up 4 Dt
immediately) for the time interval we are interested in.
(a) C
alculate the total drug flux out of the sphere. Rewrite the equation you derived in (a) in terms of
(b) N
ext, assume you have a flat wall of the same α and show that the equation can be expressed as
thickness h and the same surface area as the inside ∂ 2C ∂C
of the sphere in the original problem statement. = 2α = 0
∂α 2 ∂α
Further, assume the concentrations and diffusion
coefficients are analogous to the spherical setup. You may need to use the chain rule.
Calculate the total drug flux through the wall. (c) W rite the boundary conditions for the new
(c) F
ind a criterion for h and R in terms of the other equation in (b) using the same constraints you
parameters such that the difference between determined for (a).
your answers for parts (a) and (b) is less than 1%. (d) Find the solution for the differential equation in (b),
Would something the size of a typical cell with the using the boundary conditions you found in (c). You
membrane being the “wall” qualify, based on your may want to express the solution in terms of the
criterion? error function (erf(x)).
(e) What happens at the limit of the solute flux as t
7. Assume the cell membrane has a fixed thickness h and
approaches infinity? Is it the same as the steady-
a diffusion coefficient D for some solute of interest.
state diffusive flux? Why or why not?
You may model the cell membrane locally as a flat wall.
(f) Someone makes the following statement:
The external concentration of the solute is C and is 0
“The solution derived for the time-dependent
everywhere inside the cell (consumed immediately).
concentration profiles in (d) suggest that if I have
(a) D
evelop an expression for the concentration inside two materials of different diffusion coefficients,
the membrane at steady state, making sure to with everything else being the same as in the
identify the boundary conditions you used. Sketch original problem statement, then the concentration
this concentration profile. profiles in the two materials will be identical, only
(b) A
t some time after this steady state has been shifted in time. That is, it will never be the case that
achieved, the cell changes the diffusion coefficient for t > 0 one material will exhibit a concentration
to a new D′ < D (for example, by closing some profile that the other will never exhibit.” Support or
previously open channels). You then wait for a new disprove this statement.
steady state to be achieved. List what you expect
to change with the new diffusion coefficient D′, 11. Is it possible to have water climb 1 km in a capillary tube
assuming every other parameter remains the same. on Earth, using only surface tension? You may control the
diameter of the tube (which does not have to be constant
8. You are given a finite number of identical (finite-volume) in space), the tube’s angle to the vertical (but only the
tanks separated by geometrically identical membranes. vertical height counts), and how deep the water the end
You may set the initial concentration of the solute in each is submerged (but only the height above the surface
tank (and assume that the concentration is always uniform counts). You may also assume that water is a continuum
inside the tank itself) as well as the diffusion coefficient so that you can make the diameter arbitrarily small
of each membrane (but once you have chosen the D for without worrying about molecular effects (or even that
a membrane, it must remain fixed forever). Configure the the diameter may be smaller than a water molecule). If
tanks and membranes such that the tank with the highest your answer is yes, what is the shape of the tube?
initial solute concentration will receive a flux of the solute
at some point in time (before equilibrium). 12. If you have a water droplet and compress it slightly,
resulting in an ellipsoid (technically an oblate spheroid),
b3 b3 and release the compressing force, show that the
9. Show that Tc = is the solution for Tc =
3Da 3Da droplet tends to return to a spherical shape and not to
for the conditions specified in Example 9.2. distend further, using free-body diagram arguments.
10. We wish to model the cell locally as a flat wall of 13. The lipid bilayer consists of fatty acids that have varying
thickness h and diffusion coefficient D. Because we degrees of hydrogenation. Recall that fully saturated
are looking very closely at the cell, we will treat it as lipid chains will be relatively straight (zig-zag, actually).
a semi-infinite media, such that the cell starts at x = 0 Do you expect unsaturated lipid chain regions to have
and increases to infinity. Initially, the concentration higher or lower viscosity compared with the saturated
of some solute is zero everywhere. At time t = 0, the region? How about regions containing primarily partially
concentration of the solute outside the cell suddenly hydrogenated fatty acids?
Annotated References 277
14. Show that 20. In the text, the claim was made that the bending
stiffness, KB, of the bilayer was much smaller than
dw dw dw dw for the cell, because of the support provided by the
d nxx + nxy d nxy + nyy
dx dy dx dy cytoskeleton. Conduct a literature search and write a
+ + pz = short report (no more than a quarter to half of a page)
dx dy
to support or refute this claim.
d2w d2w d2w
nxx 2
+ 2nxy + nyy + pz 21. Suppose you have a 1 μm cell with a membrane
dx dxdy dy 2
thickness of 10 nm made of an incompressible material
with a Young’s modulus of 108 Pa. Estimate the change
15. In the text we mentioned typical strains at rupture for
in pressure necessary to inflate the cell from an initial
a bilayer. Using reasonable values given in the text,
collapsed state in which the aspect ratio (height/
estimate the typical range of maximum membrane
diameter) is 0.2, to one in which the aspect ratio is one
tension and cytosolic pressure a cell of diameter 1 μm
with the large diameter approximately equal to 10 μm.
could withstand.
(This is roughly equivalent to the geometry change
16. Consider the cortex of a red blood cell. Develop a associated with a red blood cell going from its normal
model to predict the membrane mechanics of a two- state to spherical.) Assume that bending stiffness
dimensional array of cross-linked spectrin polymers dominates over tension.
arranged in a simple rectangular lattice. The lattice
22. Consider a plate experiencing a moment in one
spacing is Le. Each one of the polymers has a persistence
direction (Mxx = M Myy = Mxy = 0).
length ℓp, contour length of L, and diameter d.
This is similar to the beam-bending problem we
Using optical tweezers, researchers have found that the
analyzed in Chapter 3. In that case we found that
persistence length of spectrin is 10 nm. Furthermore,
the contour length of each spectrin molecule is 200 nm,
and Le is approximately 70 nm. d2w
M = EI .
(a) A
ssuming that the deformations are small, dx 2
what regime (ideal chain, semiflexible polymer,
continuum mechanics) are these polymers in Does the plate theory produce the same equation for
and why? What model would be appropriate to this special case? If not, why is it different? What further
describe their behavior? assumption would be needed for the plate theory to
(b) N
ow derive an expression for the membrane area reproduce the beam theory?
expansion modulus KA for the model. Eh3
23. Show that KB = .
17. Show that τxy = Gγxy = E/(2(1 + ν))γxy and KS = Gh = 12(1 − ν2 )
Eh/(2(1 + ν)).
24. Show that, taken together, the moment–curvature
18. Derive an expression for the areal expansion modulus relation (Equation 9.52) and moment equilibrium
for a membrane in terms of the membrane thickness, equation (Equation 9.51) result in the classical plate
Young’s modulus, and Poisson’s ratio. bending equation in terms of displacements,
19. Show that Equations 9.49 and 9.50 can be combined to d2κ xx d2κ xy d2κ yy
yield the equilibrium equation for a plate. pz = KB +2 +
dx 2 d xd y dy 2
Annotated References
Axelrod D, Koppel DE, Schlessinger J et al. (1976) Mobility measure- Evans E & Rawicz W (1990) Entropy-driven tension and bend-
ment by analysis of fluorescence photobleaching recovery kinetics. ing elasticity in condensed-fluid membranes. Phys. Rev. Lett.
Biophys. J. 16, 1055–1069. This research article covers the theoreti- 64, 2094–2097. Gives expression for “effective” areal strain
cal basis underlying fluorescence recovery after photobleaching in associated with smoothing out thermal undulations under a given
order to characterize diffusion in 2-dimensional systems. tension.
Berg HC (1983) Random Walks in Biology. Princeton University Press. Helfrich W (1975) Out-of-plane fluctuations of lipid bilayers. Z. Natur-
Presents a wide range of biological problems that can be analyzed forschung. C 30, 841–842. Early work demonstrating that out-of-
through the random walk. Specifically our treatment of diffusion is plane fluctuations of membranes lead to a decrease in effective
motivated from this book. area and alters stretching elasticity. Gives expression for projected
surface area for a given tension n and bending rigidity.
Boal D (2001) Mechanics of the Cell. Cambridge University Press. An
excellent text on cell mechanics with many in-depth treatments of Kamm RD (2005) Molecular, Cellular, and Tissue Biomechanics
shell mechanics. Particularly useful for this chapter is the statistical (lecture notes from course number 20.310, Massachusetts Insti-
treatment of membrane behavior. tute of Technology, Cambridge, MA). Several of the treatments in
278 CHAPtER 9: Mechanics of the Cell Membrane
this chapter were based on these graduate course notes. Specific Singer SJ & Sackmann E (1972) The fluid mosaic model of the
material included in-plane and bending behavior of plates and structure of cell membranes. Science 175, 720–731. One of
shells and several problems. the earliest descriptions of the fluid mosaic model of the cell
membrane.
Kooppel DE, Axelrod D, Schlessinger J et al. (1976) Dynamics of fluo-
rescence marker concentration as a probe of mobility. Biophys J. 16, Timoshenko S & Woinowsky-Krieger S (1959) Theory of Plates and
1315–1329. An early description of FRAP used to measure lateral Shells. Engineering Society Monographs. A defining text on the
diffusion in the cell membrane. mathematics of plate and shell mechanics.
Simons K & van Meer G (1988) Lipid sorting in epithelial cells. Bio-
chemistry 27, 6197–6202. An early description on microdomains in
lipid membranes.
CHAPTER 10
Adhesion, Migration,
and Contraction
T hus far, our focus has primarily been on cell biomechanics; in other words,
the study of the mechanical behavior of cells. In these treatments, we used
theoretical and experimental mechanics to understand better the mechanical
behavior of cells or their structural constituents. Ultimately, we are interested in
using this knowledge to understand how mechanics governs cellular function.
We now make a conceptual turn, shifting our focus from cell biomechanics to
cellular mechanobiology; in other words, aspects of biology in which mechani-
cal force is generated, imparted, or sensed, leading to alterations in biological
function. Because mechanics plays a critical role in many different physiologi-
cal and pathological processes, a complete treatment of cellular mechanobiol-
ogy is more than we can address here. In this chapter, we aim to give the reader
a firm foundation upon which to build by focusing on three vital cellular pro-
cesses in which mechanics dictates biological function: adhesion, migration,
and contraction.
10.1 ADHESION
Cells can form adhesions with the substrate
We begin this chapter with a discussion of cell–substrate adhesion, which is vital
for many biological processes, including tissue cohesion, repair, inflammatory
responses, and growth. Adhesion is also important for engineering applications,
such as seeding new constructs with cells for implantation.
Among the best-characterized cell–substrate adhesive molecules are integrins.
As we first described in Section 2.2, integrins are extracellular matrix protein
receptors that are tightly bound to the plasma membrane. They are dimers com-
posed of α and β subunits, each of which is found in one of several isoforms. Each
isoform is denoted with subscripts. The different isoform dimerizations allow for
the extracellular component of the integrins to bind to different extracellular
matrix components such as collagen or fibronectin. For example, α5β1-integrin is
composed of an α5 subunit and a β1 subunit and is a fibronectin receptor, specifi-
cally targeting the arginine–glycine–aspartate (RGD) peptide sequence in
fibronectin. The intracellular components of integrins are often structurally cou-
pled to the cytoskeleton in an indirect manner through one or more intermedi-
ary molecules. The intracellular component of the integrin may also be a major
site of enzymatic activity.
Integrins can be assembled into discrete adhesive plaques called focal adhe-
sions. As they are assembled, an intracellular molecular complex is also created.
This complex consists of many proteins, each of which has various roles in intra-
cellular biochemical signaling and forms structural links with the actin cytoskel-
eton (Figure 10.1). It is currently unclear what, if any, advantage is associated
with multiple focal attachments rather than a more distributed scheme. However,
their formation and destruction are highly regulated. It has been shown that
microtubules can disrupt focal adhesions locally at microtubule leading tips. It is
280 CHAPTER 10: Adhesion, Migration, and Contraction
(B)
extracellular
fibronectin matrix
α β
integrin
cell membrane
Ras cytoplasm
SOS
Grb2
FAK
Src
lin
xil Kinase
Pa Pa
in
Tal in Zyxin xil
lin P130
ul culin in
nc in
lin
V s
Vi n
Ta
Te
focal adhesion
thought that that the various components of focal adhesions play different roles
in adhesion, activation, and maintenance of mechanical competence, although
the exact roles of each component are still being defined. Because they play
important roles in both structural integrity and intracellular signaling, they are a
major candidate for mechanotransduction sensors (covered in Section 11.2).
ADHESION 281
where V0 is the velocity at a height h above the surface and μ is the fluid viscosity. Nota Bene
For a pressure-driven parallel plate configuration, Equation 10.1 becomes
Plate washing. An easy method for
6 µQ indirectly measuring adhesion
τ = , (10.2) strength is plate washing. In this
bh 2
approach, the same numbers of cells
are plated in multiple wells and
where Q is the volume flow rate, b is the chamber width, and h is the chamber allowed to attach for some fixed time
height. Different shear stresses can be applied to the same cell population by (typically short, such as 30 min). The
increasing the flow rate with time, or by constructing a chamber with a spatially wells are then washed several times,
varying geometry, such as one whose width changes along the direction of flow. and the remaining adhered cells in
each well are counted.
(C)
(D)
at the interface of two different phases. At the edge of the liquid drop, there will be
three surface tensions between the solid, liquid, and gas phases: one tangent to
the solid–gas interface (nSG), one tangent to the solid–liquid interface (nSL), and
one tangent to the liquid–gas interface (nLG). Assuming the drop is at equilibrium,
we can perform a horizontal force balance at the liquid-drop edge,
We now seek to calculate the energy required to “lift” the drop from the surface.
Physically, surface energy is the energy required to generate a new surface, and
arises due to the disruption of intermolecular bonds when new surfaces are cre-
ated. For a liquid, surface tension (which has units of force per unit length) and
the surface energy density (surface energy per unit area, which also has units of
force per unit length) are identical. The energy density of adhesion is
where J = JLG and n = nLG. Equation 10.5 is sometimes referred to as Young’s equa-
tion or the Young–Dupré equation. It implies that the energy density required to
“lift” the drop from its surface is dependent on only two factors, the surface ten-
sion n, and the angle that the membrane makes with the surface.
It is possible to derive a relation similar to Equation 10.5 by considering the internal
force within membrane receptors directly. It can be shown that the relationship
between adhesion energy density J and membrane tension n is
ρbFR L
= n(1 + cosθ ), (10.6)
2
where ρb is the receptor bond density, FR is the force that would rupture the recep-
tor–ligand bond, and L is the critical bond length at which the receptor is stretched
with force FR. Notice that this is essentially a reformulation of Young’s equation
(Equation 10.5).
From the form of Equation 10.6 we can make some predictions about the capacity
for cells to control the degree to which they spread. Consider a cell that is at adhe-
sive equilibrium as a half-dome on a flat surface. We can treat the cell as a liquid
drop with contact angle π/2 and membrane tension n = J0.
Equation 10.5 implies that, the adhesion energy density is J = J0. One might sup-
pose that to spread out more toward a flat pancake shape, the cell would be
required to increase the adhesion energy density indefinitely, but this is not the
case. In particular, if we increase the adhesion energy density by a factor of q, then
Equation 10.6 predicts that
q = (1 + cosθ ). (10.8)
This implies that q is at most two, which occurs when θ is equal to zero. In other
words, doubling the energy adhesion density is sufficient to spread the cell to the
maximum extent allowed by the membrane tension. This example suggests that a
cell may only need to adjust its adhesion energy density by small amounts to pro-
duce substantial changes in spreading (for instance, by adjusting its receptor
bond density as in Equation 10.6).
284 CHAPTER 10: Adhesion, Migration, and Contraction
Equation 10.5 was derived using a combination of a force Combining Equations 10.9, 10.10, and 10.11, and solving
balance and energy considerations. One can also derive Equation 10.12, leads directly to Equation 10.3.
this relation solely using energy considerations. Consider
a droplet to be a truncated sphere of radius R, parame-
terized by the angle θ (Figure 10.4). A “cone” with vertex
angle θ will truncate the part of the sphere that we use to
model the drop, resulting in a contact angle of θ as well.
θ
Using our relation for energy density of adhesion given
in Equation 10.4, the energy of the drop can be written as
where
ALG
ASL = πR sin θ (10.10)
2 2
θ
is the area of the flat surface with which the drop is in ASL
contact and
Figure 10.4 A droplet of water approximated as a
ALG + 2πR (1 − cosθ ) (10.11)
2 truncated sphere. The angle of truncation, or the
vertex angle of the cone, is θ. This angle is also the contact
angle of the droplet on the surface. The area of the solid–
is the total area of the “dome shaped” liquid–gas interface. liquid interface is “underneath” the droplet and the area of
To obtain Equation 10.3, we can use the fact that at equi- the liquid–gas interface is the “upper surface” of the
librium there is no change in energy with respect to ASL, droplet.
∂E
= 0. (10.12)
∂ ASL
n fa
ADHESION 285
As we have done in the past, we could create a free-body diagram for Figure 10.3.
However, we have already analyzed a more general case. Note that this situation is
entirely analogous to a plate bending owing to a distributed pressure. Recall
Equation 9.37,
d 2w d 2w d 4w d 4w d 4w
n 2 + 2
+ KB 4 + 2 2 + + Fa = 0, (10.13)
dx dy dx dx dy dy 4
d 2w d 4w
n 2 + K B 4 + Fa = 0. (10.14)
dx dx
We can model the adhesion force with simple linear spring-like behavior such
that the force in a single bond is given by
Fm
y 0 < y < lm
Fb = lm , (10.15)
0 y < 0, y > lm
where Fb is the force for an individual bond, Fm is the maximum force generated
by the bond before bond breakage, lm is the length of the bond at the maximum
force, and w is the displacement (Figure 10.6). If w > lm, then the bond has bro-
ken and is no longer capable of sustaining force. Next, we observe that the total
adhesion force is simply the sum of the individual bond forces:
Fa = nb Fb , (10.16)
where nb is the area density of bonds attaching the cell to the surface. We can use
energy density of adhesion by defining
nb Fmlm
J = , (10.17)
2
the work done in extending all the bonds in a unit area from zero length to their
maximum length.
w
lm
286 CHAPTER 10: Adhesion, Migration, and Contraction
To determine the energy density of adhesion, we first spring. For our example, the spring constant k is the slope
use the relationship that work equals force times dis- (Fm/lm) and x is the maximum length of the spring lm.
tance and that this work goes entirely into the adhesion Substituting yields W = 0.5 × Fm × lm. Because this is for a
energy density. Force, however, is not constant. It varies single bond, we can convert it to the desired energy density
with the length according to Equation 10.15. Therefore by multiplying by the number of bonds per unit area nb,
the work for a single bond is the integral of the force–
displacement curve, or W = 0.5 × k × x2 for a classical W = 0.5 × nb × Fm × lm = J.
F 2J y 2J
Fa = nb m y = = 2 y (10.18)
lm lm lm lm
by using appropriate substitutions. Substituting into Equation 10.14, we obtain an
expression for adhesion in terms of y alone, with the remaining terms based on
physical parameters. Although this model is fairly explicit and can be (and has
been) used to model peeling, it relies on knowing parameters that are typically
difficult to measure. For this reason, many studies of adhesion rely on strict com-
parisons from direct force measurements rather than fitting to parameterized
models. Also, bond strength is known to depend on the level of prestress, which
this treatment ignores.
Wadh ∝ Ja 2 , (10.20)
where J is the adhesion energy density. Assuming that the strain energy is bal-
anced by the adhesion energy, then Wdef ~ Wadh. We can relate the right handed
sides of the expression in Equations 10.19 and 10.20 as
Ea 5 (10.21)
∝ Ja 2 ,
R2
ADHESION 287
which simplifies to
JR 2 (10.22)
a∝ 3 .
E
This is only a scaling relationship, and is only valid for a relatively small cellular
deformation occurring upon adhesion. Still, it allows us to predict some interest-
ing scaling behavior. Consider what happens to adhesion contact if the cell radius
decreases by a factor of two, but all other parameters remain constant. A reduc-
tion in R by a factor of two would result in an approximately 1.6-fold reduction in
adhesion contact radius. If the elastic modulus increases by a factor of two the
adhesion contact radius decreases by approximately 1.3-fold. We can see that the
contact radius of the cell is more sensitive to changes in its radius than changes in
its stiffness, a somewhat unexpected result.
neutrophils
endothelial layer
(A) rolling time < 1 min (B) rolling time > 4 min Figure 10.10 Neutrophil tethering at
various fluid shear stresses on a
surface coated with P-selectin.
(A) and (C) show the tether formation
<1 min of attachment, and (B) and
(D) show tethering after >4 min of
attachment, so the tethers are more
2 dynes/cm2 fully developed (From, Ramachandran
V et al. (2004) Proc. Natl. Acad. Sci. USA
with permission from the National
Academy of Sciences.)
flow
(C) (D)
8 dynes/cm2
As we saw in Section 7.2, the kinetics of this reaction can be described quantita-
tively through the so-called law of mass action. Specifically, the assumption,
which can be obtained from statistical mechanics by calculating the rate at which
reactants would randomly collide, is that the rate at which a reaction occurs is
linearly related to the rate constant and the concentration of the reactants. In our
example of two reactants, the forward reaction occurs at a rate of
k+ [L][R] (10.24)
k−[B], (10.25)
The dissociation constant has the unique property that when the concentration of
ligand is equal to Kd, half of the receptors will be bound. This can be better under-
stood by computing the fraction of bound receptors as
[L]
[B]
[B] [B] [L]
%bound = = = . (10.28)
[B]+[R] [B] [L] +[R] [L] [L]+K d
[B] [B]
290 CHAPTER 10: Adhesion, Migration, and Contraction
percentage bound
exactly half of the receptors will be in
the bound state.
50
Kd [L]
Notice that when the [L] = Kd, %bound = 0.5. A plot of Equation 10.28 can be seen
in Figure 10.11.
t
ADHESION 291
the rate of dissociation, k−, and does not depend on k+. Next, he assumed that
force has an exponential influence on bond rupture such that the rate of dissocia-
tion in the presence of a force F across the bond is
σF
(10.29)
k− = k e0 kBT
− ,
where k−0 is the rate of dissociation in the absence of force, kB is Boltzmann’s con-
stant, T is absolute temperature, and σ is a constant that characterizes the influ-
ence of force. Notice that the form of this relationship is very similar to the
Boltzmann probability derived in Section 5.5. Because the denominator in the
exponent has units of energy, σF must also have units of energy, and therefore σ
must have units of length. Indeed, it can be shown that σ is the bond extension
that is associated with its minimum potential energy configuration. Equation
10.29 can also be expressed as
F
kBT (10.30)
k− = k e , FB =
0 FB
− .
σ
d [B]
= − k − [ B ]. (10.31)
dt
The solution of this differential equation is an exponential of the form
[B(t )] = [B(0)]e− k t . −
(10.32)
There are actually two exponentials in Bell’s model, one describing the influence of
force on the rate of dissociation, and another describing the time evolution of [B].
1 2 3 4
shear stress (dynes/cm2)
10.2 MIGRATION
Cell migration can be studied in vitro and in vivo
Cell migration is a fundamental aspect of cellular behavior that is central to pro-
cesses of immunity, regeneration, repair, inflammation, and cancer, to name a
few. In vitro assays, such as tracking motility or movement through a polymer gel,
are powerful reductionist approaches. However, in vivo strategies are becoming
more common. One elegant study was performed on heart transplant patients. Of
the people who had heart failure significant enough to require transplantation,
there were eight males who received hearts from female donors. After the male
patients died (mostly within a year, some after nearly 2 years), their female hearts
were assayed for the Y chromosome. It was found that a substantial fraction of
organ cells (up to 10%, depending on the type of cell) were from the male recipient
(turning the organ into a “chimeric tissue”, Figure 10.14). Although the source of
those cells is not known, such changes can influence other readouts, including
migration.
FPO
Figure 10.14 A section of heart
showing the presence of a Y
chromosome. The bright spot (see
arrow) is a Y chromosome detected by
in situ hybridization in a female heart
transplanted into a male patient. (From
Quaini F et al. (2002). New Engl. J. Med.
346. With permission from the
Massachusetts Medical Society.)
MIGRATION 293
WASP/Scar
ADF/cofilin
11 Pool of ATP–actin
ATP–actin
bound to profilin
ADP–actin
9 ADF/cofilin severs
Profilin and depolymerizes
ADP–actin filaments
Capping protein
Arp2/3 complex
10 Profilin catalyzes
exchange of ADP
for ATP
Figure 10.16 Stages for protrusion of the cell membrane in lamellipodia. In this model, external stimuli (1) result in the
activation of GTPases (2), which activate Wiskott–Aldrich syndrome protein (WASP). (3). This complex activates the Arp2/3 complex (4),
resulting in the formation of a new actin branch on a pre-existing actin filament. This newly formed actin branch grows by
polymerization at the barbed end of the filament (5) from a pool of profilin-bound actin monomers, pushing the membrane forward (6).
Elongation of actin filaments is terminated by capping protein (7), which binds to the growing ends of polymering actin filaments.
Older ADP-bound filaments (8) are severed by actin-depolymerizing factor (ADF)/cofilin (9). Depolymerized actin undergoes
dissociation of ADP and binds ATP (10), upon which ATP–actin re-binds profilin, renewing the pool of ATP-bound actin monomers
available for assembly (11). (Adapted from Pollard TD (2003) Nature. 442.).
branching actin network. Multiple sites for actin extension permit a relatively uni-
form growth pattern during protrusion to be achieved. We will see in the next sec-
tion that although it is generally accepted that actin branches serve to push the
membrane forward by polymerization, the exact physical mechanism by which
this may occur is still relatively unclear.
Brownian
fluctuation
of membrane
Nota Bene
Somebody stole my actin! Listeria
monocytogenes is an intracellular
the local cell membrane to undergo small thermal fluctuations. If a pool of free actin bacterium responsible for the disease
listeriosis, the most common cause
monomers is available at the leading edge, when the membrane shifts forward, space
of death associated with food-borne
is temporarily generated that allows polymer extension. This polymerization pre- pathogens. As an intracellular
vents the membrane from shifting back. As this ratcheting process is repeated, the bacterium, it resides within the
resulting actin polymerization leads to the development of a protrusion. cytoplasm of another cell, remaining
relatively hidden from the immune
An alternative hypothesis that has been advanced is related to the Brownian ratchet system. To maximize its capacity to
in that the actin filaments themselves rather than the membrane are undergoing move within the cell, the bacteria has
thermal fluctuations. The filaments randomly fluctuate in end-to-end length. developed a mechanism whereby it
When the end-to-end length decreases slightly (through bending), space is tempo- “hijacks” the actin polymerization
rarily generated that allows for polymer extension (Figure 10.18). When the fila- machinery in the cell and rides on a
ment straightens, this generates a protrusive force that extends the membrane. wave of polymerizing actin that it
induces. This polymerizing actin can
often be viewed as a “comet-tail”
Cell motion can be directed by external cues structure behind the bacterium. The
polymerizing actin propels the
We have considered the process of cell migration as if it were an aimless process. bacterium to the membrane and
However, it is clear that, cell migration can be directed to specific locations. generates a protrusion. If the infected
Chemical compounds that direct cell migration are called chemoattractants. If cell is adjacent to another cell, the
bacterium can then enter the second
there is a chemical gradient in a chemoattractant, cells may sense these gradients cell by endocytosis or a similar
and direct their migration toward the source. One of the most prominent exam- mechanism. This method of
ples of directed cell migration toward a chemoattractant is the targeting of patho- propagation has the advantage that
gens by neutrophils. The pathogens release compounds that the neutrophil the bacterium can travel from cell
recognizes as foreign. Upon activation, the neutrophils crawl about, targeting to cell without ever leaving the
local pathogens. Upon contacting the pathogens, the neutrophils engulf them, intracellular environment.
removing the source of chemoattractant.
d 2 = S 2 Pt + C1 + C 2e −2t / P , (10.34)
where the constants C1 = −C2 = −2S2P2 are found by applying the initial condi-
tions, 〈d 2〉 = 0 at t = 0, and d〈d 2〉/dt = 0 at t = 0 (the cell initially has no bias). In
two dimensions,
( (
d 2 = S 2 Pt − P 2 1 − e −t / P . )) (10.35)
d 2 = 2S 2 Pt = 2Dt , (10.36)
where D = S2P has the units of m2/s and can be considered to be an effective “dif-
fusion coefficient” of the cell.
To determine the values of S and P experimentally, one can acquire multiple cell
paths for different cells with various time intervals, and compute an average mean
square distance for each time interval. With these data established, the speed and
persistence can be calculated using standard least squares fitting algorithms. In
this approach, it is assumed that every cell is migrating under similar mathemati-
cal patterns. Some examples of speed and persistence relationships for various
cell types are presented in Figure 10.21.
10
neutrophil
leukocytes
1
0.01 0.1 1 10 100
speed (µm/min)
298 CHAPTER 10: Adhesion, Migration, and Contraction
Now we seek to demonstrate that cell 2 has a leftward bias. (0.8, –0.5)
We do this by calculating the change in angle the cell
makes at each successive time interval. In doing so, we Figure 10.22 Migration paths for two different cells. The
define the angle relative to a line parallel to the line seg- values of x in the plot refer to cell positions obtained at time
ment associated with the previous time iteration. Leftward intervals separated by a constant, δ t. The (x, y) coordinates of
changes in orientation result in negative angles, and right- each point is given in parentheses.
ward changes result in positive angles. For cell 2, we find
10.3 CONTRACTION
In the final section of this chapter, we discuss the generation of intracellular
contractile forces. When we talk about cellular contraction, one of the first
examples that comes to mind is muscle cells, which possess specialized struc-
tures for the generation of contractile forces. However, non-muscle cells also
possess the capacity to generate contractile force, such as during locomotion, or
in the generation of traction forces. In this section, we give overviews of the
molecular structures within muscle and non-muscle cells that allow for con-
tractile force generation, as well as experimental measurements of force genera-
tion in both cases. We also present molecular-based mathematical models of
contraction that allow us to gain mechanistic insight into a range of observed
phenomena.
CONTRACTION 299
sarcomere
thick
filament
M-line
pointed end
barbed end
Figure 10.23 A schematic of a sarcomere, the basic contractile unit of a muscle fiber. The z-discs (which are commonly
visible in pictures of muscles) anchor the actin filaments. The myosin bundle (thick filament) with the myosin heads sticking out
(in opposite directions on either end) is caged inside the actin (thin) filaments. (Adapted from, Bray D (2000) Garland Science,
New York.)
300 CHAPTER 10: Adhesion, Migration, and Contraction
ejection and a higher systolic pressure. This led him to the realization that the
tension a sarcomere is able to generate must increase with its initial stretch or
length (Figure 10.24).
( F + a)(v + b) = k , (10.37)
where F is force, v is velocity, and a, b, and k are constants. This equation captures
critical aspects of the force–velocity relationship. As force increases, velocity
decreases. There is a maximum force at zero velocity and, vice versa, a maximum
shortening velocity at zero force. Between these two extremes the relationship is
hyperbolic and has been shown to provide a remarkably accurate fit to experi-
mental data.
Interestingly, Hill did not arrive at his famous equation by making measurements
of force and shortening velocity directly. He actually made measurements of the
heat liberated by muscle as it shortened. These measurements were based on his
force
Maximum force generation occurs
during isometric contraction, or when
the muscle is held to a fixed length.
Maximum contraction velocity occurs
with zero force.
length velocity
maximum
contraction
velocity
F
Rs
L0
L c(=L 0)
Rb
“straight” morphology. However, if the release strain is large such that it exceeds
the stress fiber pre-strain, the fiber will exhibit a “wiggly” or “buckled” morphol-
ogy. An alternative approach to demonstrate pre-stress in stress fibers is to sever a
stress fiber using a high-powered laser, or disrupt a focal adhesion anchoring a
stress fiber. In both cases, the free end of the stress fiber can be observed to con-
tract, again demonstrating that the stress fiber is under a pre-strain.
and the pre-strain as 1 − λf. It is noteworthy that λf can be calculated in an alterna-
tive manner that does not require the determination of Rs. Consider the case
where we know Rb and the stretch ratio of the substrate, λs. λs is related to Rs and
Rb as
Rs
λs = . (10.42)
Rb
λf = T λs . (10.44)
Using this approach, the stress fiber pre-strain of cultured human endothelial cells
has been estimated to be as high as 25%, although it is highly heterogeneous. Nota Bene
Myosin variants. There is no single
Myosin cross-bridges generate sliding forces within myosin protein, but rather a huge
superfamily. There are over 40
actin bundles different genes and about 100
different protein products that are
In the previous section, we discussed various structures in muscle and non- grouped into 18 major variants
muscle cells that allow for the generation of contractile forces. In these cases, known as classes. This immense
although the structures were distinct in molecular makeup and organization, the diversity is generally the result of
same basic molecular mechanism was responsible for driving contraction: myo- various splice variants. Myosin II is
sin movement along actin. Actin and myosin together form a cellular force gen- the type found in muscle; the tails of
eration system that is critical to an amazingly wide variety of cells and tissues. this myosin allows it to form polymer
filaments. Other myosins remain as
Given that both muscle cells and non-muscle cells possess the capacity to gener- monomers and are known as
ate contractile forces using actin and myosin, the question arises, how does this unconventional myosins because
occur? Although it is well established that myosin functions as a molecular motor myosin was first identified in muscles.
in converting chemical energy into mechanical work by moving along actin, many Most of the variability is found in the
details have yet to be elucidated. However, the cross-bridge model in which collec- tail region, associated with binding to
tions of myosin molecules link adjacent actin polymers together provides a different cargos or assembling into
fibers, whereas the head regions tend
remarkably accurate model of actin–myosin interactions both within and outside
to be much more conserved. All
sarcomeres. myosins move toward the + or
In the cross-bridge model, myosin molecules go though several distinct stages or barbed end of the actin polymer,
except myosin VI, which moves
configurations (Figure 10.27). The process is cyclical in nature and at completion backward!
the myosin has returned to its initial configuration. In the first stage, myosin is
tightly bound to actin, and its ADP/ATP (adenosine diphosphate/adenosine
triphosphate) binding domain is not occupied. This is known as the rigor state,
because no sliding or motion can occur. To leave the rigor state, myosin must bind
a molecule of ATP. When this occurs, myosin releases from actin, and ATP is
ATP
ATP
rigor state
ADP•Pi
ADP
ADP Pi
304 CHAPTER 10: Adhesion, Migration, and Contraction
hydrolyzed into ADP. This releases chemical energy, allowing myosin to change
conformation into what is known as a cocked configuration. The hallmark of this
configuration is that the myosin head has moved toward the (−) or barbed end of
the actin polymer. In the third stage, myosin rebinds to the actin, but because it is
in the cocked configuration, it rebinds at a site more toward the (−) end. In the
fourth stage, rebinding of actin induces the myosin to release its phosphate. In
this process, it also undergoes the power stroke. It returns to the uncocked con-
figuration, and in the process drags the actin molecule toward the center of the
myosin cross-bridge. The power stroke is where mechanical work is produced. In
the fifth and final stage, the spent molecule of ADP is released. This returns the
myosin to the rigor state.
Nota Bene An interesting aspect of actin–myosin contraction is that the myosin heads are not
able to bind just anywhere along the polymer. In fact, each actin subunit has only
Rigor mortis occurs with lack of one binding site for the myosin heads. As we learned in Section 7.1, actin subunits
ATP. A familiar phenomenon occurs
polymerize in a helical structure. The pitch of this helix has been measured to be
when ATP generation is suddenly
halted in the body. Actin–myosin Δ = 36 nm; in other words, the polymer completes one full turn of the helix for every
interactions will be halted in one of 36 nm along its length. Interestingly, the distance that the myosin molecule stretches
two steps: just before the power when it is cocked in preparation for a power stroke, δ, is only around 5 nm. Because
stroke, or just after the power stroke, binding sites can only occur once in each full twist, it would be impossible for a
before the myosin head can myosin molecule to reach far enough forward to grab another binding site each
disengage. In either case, there may time through the cross-bridge cycle. A simple explanation for this apparent paradox
be some residual contraction,
depleting the available ATP stock,
is that many myosin molecules must work together to produce sliding. At any given
followed by a failure of relaxation point in time, most are not engaged with actin at all, whereas a few are bound and
because the actin–myosin tugging on the actin molecule, much like the legs of a millipede.
interactions are locked in place. An
example of this occurring is during The implication of this cooperative behavior is that any given myosin molecule,
rigor mortis after death. The body spends most of its time bound to ATP and, as a result, unbound from actin. For
may feel stiff because of the example, let the average time per cycle that the myosin is not bound to actin be
contraction of muscles that are toff, and the time that myosin is bound to actin be ton. We might estimate the frac-
subsequently locked in place owing tion of time myosin is engaged (ton/ttotal), a quantity we will call the duty ratio,
to exhaustion of the available ATP
supply.
rduty, by assuming that is it proportional to the ratio of δ and Δ,
δ 5 nm
rduty = = = 0.14. (10.45)
∆ 36 nm
However, even this seems to require much more frequent binding to actin than is
actually observed. In fact, myosin seems to be quite a “lazy” molecule. The ration-
ale for this statement comes from considering the force generation capacity of
entire actin–myosin bundles. The total tension should be approximately the num-
ber of cross-bridges multiplied by the force generated by each cross-bridge. The
number of cross-bridges can be estimated, and the time-average force per cross-
bridge is a function of ton and toff,
However, the force generated when myosin is not engaged to actin must be zero
(<Foff > = 0). Therefore,
t on
F = Fon = rduty Fon (10.47)
t on + t off
CONTRACTION 305
or
δ t on F
rduty = = = . (10.48)
∆ t on + t off Fon
δ+
compressed
δ–
306 CHAPTER 10: Adhesion, Migration, and Contraction
[F] Combining this with our expression for duty ratio and making the observation
that the total stroke distance δ = δ+ + δ-, we obtain
Ft
F = rduty Fon
δ km
approximately = (δ + − δ − )
hyperbolic ∆ 2
km (10.50)
= (δ + + δ − )(δ + − δ − )
2∆
vmax v km 2 k m 2
= δ+ − δ− ,
Figure 10.29 The force–velocity 2
∆ 2∆
F+ F−
relationship predicted by the
power-stroke model. Although not
strictly hyperbolic, for physiological where the first term is the positive elastic force, F+, and the second term is the
values of the physical parameters, the negative drag force, F−.
power-stroke model is similar to the
prediction of Hill’s equation. Let us consider the form of Equation 10.50 in a bit more detail and see if we can
relate it to sliding velocity. The first term depends on the distance that the myosin
heads move forward when they are cocked. As we described, most of the myosin
head groups are already bound with ATP and therefore cocked at any point in
time. Therefore, there is no reason to expect that this would depend on velocity.
On the other hand, the second term depends on the distance the heads over-
shoot the zero force position before releasing, in other words the distance the
myosin springs get compressed during each cycle. It would be reasonable for this
quantity to depend on sliding velocity. In fact, this compression distance should
increase linearly with velocity. If we introduce the release time as a new param-
eter tr, we obtain,
km
F = F+ − (vt r )2 . (10.51)
2∆
Notice that the first term is the maximum positive force the myosin is capable of
generating. It is a constant, depending only on structural parameters. Also, note
that the maximum force occurs when the velocity is zero, consistent with the
concept of a stall force, at which point the second term (the drag force) is also
zero. The power-stroke model is consistent with some important features of the
force–velocity behavior we have observed, a maximum force at zero velocity and
a maximum contraction velocity that occurs when the force is zero. However, the
drag predicted with this model increases with the square of the sliding velocity.
Upon inspection of Equation 10.51, we can see that this prediction is not consist-
ent with the predictions of the Hill model, which does not have a square depend-
ence on velocity. To resolve this issue, let us again consider the parameter Δ, the
distance separating the myosin-binding sites on actin. As we alluded to previ-
ously, the minimum possible value of Δ is the physical spacing of the sites, which
we will now term ds. From the duty ratio argument, we expect Δ to exceed ds by
some amount, but the myosin heads and the actin-binding sites must align at
some point. Therefore,
∆ = nds , (10.52)
where n is an integer. To estimate what n might be, consider that each binding site
has an effective “sweet spot” of width wbs. If, in order to bind, the myosin head
needs to be within this width, the time available for binding is inversely related to
velocity such that the time available to form a bond with the binding site is
wbs
t bs = . (10.53)
v
CONTRACTION 307
Further, we assume that while the myosin is in the sweet spot, it will bind actin at
a constant rate of k+ such that
d[M ]
= − k + [ M ]. (10.54)
dt
Note that Equation 10.54 is entirely analogous to Equation 10.31, which described
the time evolution of adhesive bonds in the Bell model. As before, the solution to
Equation 10.54 is an exponential of the form
[ M ](t ) = [ M ](0)e− k t . +
(10.55)
From the perspective of the individual myosin molecule, the probability of transi-
tioning from an unbound to a bound state in the time available for binding is
p(t bs ) = 1 − e − k+ t bs
w
(10.56)
− k+ bs
=1−e v .
The probability that a myosin head will bind in the time available is an exponen-
tial. To find the total probability over the period, we must integrate from 0 to tbs.
t bs
∫k e ( −e ) = 1 − e
− k+ t t bs − k+ t − k+ t bs
p(t bs ) = + dt = . (10.57)
0
0
You can see from this expression that if the velocity is very high, this probability
gets quite small, and the myosin heads would pass by many potential binding
spots without binding. In this limiting case Δ ⪢ d and n is large. In other words,
relatively few of the binding sites are occupied. Because each binding site is sepa-
rated from the next by a distance δ, and the distance between an occupied site is
Δ, then the probability of a given head being bound is also given by
1 ds
p(t bs ) = = . (10.58)
n ∆
ds
∆= w (10.59)
− k+ bs
1−e v
and
F =
km 2
2∆
( 2
δ+ − δ − )
kmδ +2 δ2
= 1 − −2 (10.60)
2∆ δ+
kmδ +2 v 2t −2
wbs
− kon
= 1−e v
1 − .
2ds δ +2
Equation 10.59 gives behavior that resembles the hyperbolic relationship observed
experimentally by Hill. To see this, consider some realistic values for the parame-
ters. Molecular measurements of myosin predict a value of km of roughly 5 pN/nm
and t− = 0.6 ms. As discussed, molecular structures suggest a binding-site spacing
of ds = 36 nm and myosin stroke distance of δ+ = 5 nm. The term k+ would change
308 CHAPTER 10: Adhesion, Migration, and Contraction
Key Concepts
• Cell adhesion involves multiple proteins interacting to • Cell migration involves a series of coordinated
form structures such as focal adhesions in cell–matrix processes, including protrusion, adhesion,
adhesion and similar adhesive plaques with cell–cell displacement/contraction, and release/retraction.
adhesion. Brownian ratchet processes occur at the interface of
• Adhesiveness is difficult to model precisely without polymerizing actin and membrane, and depend on
knowing the number and energy of receptors. thermal fluctuations forming a gap for the insertion
However, lumped parameter models based on energy of a new monomer.
of adhesion and surface tension can be quite • Actin–myosin interactions generate cellular force in
informative. an ATP-dependent fashion. They are responsible for
• During inflamation, neutrophils attach to vessel walls force generation in smooth and striated muscles as
in stages characterized by endothelial cell activation, well as non-muscle cells in stress fibers and during
selectin-mediated rolling, neutrophil activation, and cellular motility.
integrin-mediated firm adhesion. • The Hill equation describes the inverse hyperbolic
• The Bell model describes the effect of force on bond relationship between force and velocity.
kinetics with an exponential increase in off-rate with • Cross-bridge models describe the cyclic series of
force. states that myosin goes through in converting ATP to
• Catch bonds exhibit increased adhesion with load. force. The duty ratio of myosin is the fraction of time
Catch-slip bonds exhibit decreased adhesion once a that it is engaged with actin, which is surprisingly low.
certain threshold force is reached. • The power-stroke model describes the actin–myosin
• Cell migration can be observed in vivo but is more interaction mechanics at a molecular level by treating
easily quantified in vitro. The net migration of a cell myosin as a simple spring that is shortened during
is the result of the speed and persistence. They cocking. It predicts force–velocity relationships in
can be quantified individually from cellular terms of cocking distance, drag, and the size and
trajectories. frequency of actin-binding sites.
Problems
1. During cell migration, a cell does work that is typically 2. Consider a cell attached to a surface, as shown below.
dissipated into viscous losses within and outside the Other than the flattened contact region, the cell is
cell. In vitro, the outside of the cell is typically much less spherical with radius R. You have deployed a shear
viscous than the cytoplasm, so the losses outside the stress device to determine the adhesion strength of the
cell can be neglected. Treat the cell as a thin pancake cell to the surface, by gradually increasing the shear
of height h and radius R, and treat the cytoplasm as force applied to the cell. The adhesion force per unit
an incompressible Newtonian fluid of viscosity μ and area is denoted Fa, exerting its force roughly uniformly
density ρ. The cell attaches to the substrate with an over the adhesion area, which is approximately Ra in
adhesion force per unit area of Fa (you may assume this diameter. The shear stress device consists of a chamber
is uniform across the cell’s basal surface). If the velocity with an upper platform moving with velocity v, which
of the cell v is constant, derive an expression for v in is increased very slowly so that the flow pattern can be
terms of the other parameters and determine whether considered fully developed Couette flow at any time.
the cell would move faster or slower if it were to thin by The chamber height is h. At some critical velocity vc the
a small amount (h decreases). cell detaches.
ProbLems 309
(A)
(B)
Annotated References
Ananthakrishnan R & Ehrlicher A (2007) Forces behind cell move- Huxley AF & Niedergerke R (1954) Structural changes in muscle
ment. Int. J. Biol. Sci. 3(5), 303–317. This journal review article pre- during contraction: interference microscopy of living muscle fibres.
sents a nice overview of cell adhesion and migration as they work Nature 173, 971–973. Early description of muscle striations and how
together to move cells. they change with contraction.
Bell GI (1978) Models for the specific adhesion of cells to cells. Johnson KL (1985) Contact Mechanics. Cambridge University Press.
Science 200, 618–627. The original presentation of Bell’s model of A good reference for Hertz contact and other contact problems in
the effect of force on adhesion bonds. mechanics.
Boal D (2002) Mechanics of the Cell. Cambridge University Press. Kamm RD and Lang M. Molecular, Cellular, and Tissue Biomechan-
Boal’s text covers many aspects of cell mechanics mathematically, ics (course material from course number 20.310, Massachusetts
including some of the adhesion derivations presented in this chapter. Institute of Technology, Cambridge, MA available online at http://
Costa KD, Hucker WJ & Yin FC (2002) Buckling of actin stress fibers: ocw.mit.edu). This course covered basic principles of biomechan-
a new wrinkle in the cytoskeletal tapestry. Cell Motil. Cytoskel. 52, ics, including scaling analysis and engineering analysis, as applied
266–274. This article is the source for the method of determining to biological molecules, cells and tissues. Much of the class material
stress fiber pre-stress by culturing on a flexible substrate. served as inspiration for many topics throughout this chapter. Some
of the lecture materials can be found online in OpenCourseWare.
Deguchi S, Ohashi T & Sato M (2005) Evaluation of tension in actin
bundle of endothelial cells based on preexisting strain and tensile Kendall K (2001) Molecular Adhesion and Its Applications. Kluwer
properties measurements. Mol. Cell. Biomech. 2, 125–33. Direct Academic/Plenum Publishers. This textbook presents many funda-
measurement of stress fiber pre-stress using cantilevers. mentals of adhesion with a section on cell adhesion. Some of the
basic laws of adhesion are presented in this book.
Dembo M, Torney DC, Saxman K & Hammer D (1988) The reaction-
limited kinetics of membrane-to-surface adhesion and detachment. Lauffenburger D & Linderman JJ (1993) Receptors: Models for Bind-
Proc. R. Soc. Lond. B. 234, 55–83. Some of the first clear evidence ing, Trafficking and Signaling. Oxford University Press. This text cov-
suggesting that catch bonds might exist. ers both biology and mathematics of cell interactions, migration,
adhesion and signaling. The speed–persistence relationship and
Dillard DA & Pocius AV (2002) The Mechanics of Adhesion. Elsevier membrane peeling presented in this chapter is based, in part, on
Science. Dillard’s text provides some mathematical background the material in this book.
in adhesion. It complements but also overlaps some of Kendall’s
material. Dillard’s material includes the discussion of the JKR Maheshwari G & Lauffenburger DA (1998) Deconstructing (and
described in Section 10.1. reconstructing) cell migration. Micro. Res. Tech. 43, 358–368 This
journal article reviews many aspects of cell migration, including
Evans EA (1985) Detailed mechanics of membrane-membrane chemical gradients, path tracing, and the various steps involved in
adhesion and separation I&II. Biophys. J. 48, 175–192. An early cell motion.
report of micropipette cell–cell adhesion experiments and peeling
analysis. Marshall BT, Long M, Piper JW et al. (2003) Direct observation
of catch bonds involving cell-adhesion molecules. Nature 423,
Evans EA & Calderwood DA (2007) Forces and bond dynamics in cell 190–193. More recent direct observations of the existence of catch
adhesion. Science 316, 1148–1153. A review of the role of forces in bonds.
cellular adhesion.
Pollard TD, Blanchoin L & Dyche Mullins R (2000) Molecular mecha-
Finger EB, Puri KD, Alon R et al. (1996) Adhesion through L-selectin
nisms controlling actin filament dynamics in nonmuscle cells. Annu.
requires a threshold hydrodynamic shear. Nature 379, 266–268. A
Rev. Biophys. Biomol. Struct. 29, 545–576 Pollard’s review article
quantitative characterization of cell rolling via selectin binding.
discusses various aspects of actin remodeling and dynamics in the
Hill AV (1938) The heat of shortening and dynamics constants of context of the leading edge of a cell undergoing migration.
muscles. Proc. R. Soc. Lond. B 126, 136–195. This elegant paper
Pollard TD (2002) The cytoskeleton, cellular motility and the reduc-
describes the original Hill experiments describing measurements
tionist agenda. Nature 422, 741–745. This insight article (similar to
of changes in temperature of 1 × 10−3 °C. It is the seminal presen-
a focused review) discusses what’s known about the role of the
tation of the Hill model.
cytoskeleton in describing cell mobility.
Holmes JW (2006) Teaching from classic papers: Hill’s model of
muscle contraction. Adv. Physiol. Edu. 10, 67–72. A description of Ramachandran V, Williams M, Yago T, et al. (2004) Dynamic altera-
a simulation-based teaching module that takes students through tions of membrane tethers stabilize leukocyte rolling on P-selectin.
Hill’s original reasoning process, eventually allowing them for for- Proc. Natl. Acad. Sci. USA 101, 13519–13524. This research article
mulate his model. describes the behavior of leukocytes as they contact a surface
coated with P-selectin. The cell velocities and tether formations are
Howard J (2001) Mechanics of Motor Proteins and the Cytoskeleton. discussed.
Sinauer Associates. A great introduction to cytoskeletal polymeriza-
tion and mechanics. This book provides a solid foundation of the Tanner K, Boudreau A, Bissell MJ & Kumar S (2010) Dissecting
mechanics of the cytoskeleton and motor protein force generation. regional variations in stress fiber mechanics in living cells with laser
Much of the development for the cross-bridge model, as well as nanosurgery. Biophys. J. 99, 2775–2783. This article describes the
some problems, are adapted from this text. use of high-power lasers to disrupt stress fibers.
Huxley AF (1957) Muscle structure and theories of contration. Tees DF & Goetz DJ (2003) Leukocyte adhesion: an exquisite bal-
Prog. Biophys. Biophys. Chem. 7, 255–318. Original presentation of ance of hydrodynamic and molecular forces. News Physiol. Sci. 18,
the power-stroke model of force generation in terms of molecular 186–190. Review of balance and interplay between the forces of
parameters. flow and adhesion in leukocytes.
CHAPTER 11
Cellular
Mechanotransduction
flow
sensor
second messenger
molecules, streaming potentials, and fluid shear stress. As we will see, cells in a
variety of tissues sense and respond to fluid flow, even in tissues in which trans-
port of fluid is not a primary function.
Nota Bene
Vascular endothelium experiences blood-flow-mediated
Lining cells respond and adapt to shear stress
enhance their function. The
capacity for lumen-lining endothelial The vascular system is one of the earliest systems in which fluid flow was impli-
and epithelial cells to respond to flow cated as an important signal for regulating cell behavior. This is due in large part
indicates that these vessels are not
merely a collection of pipes and
to the fact that endothelial cells (cells lining the vessels) exhibit dramatic,
tubes meant for transport. They are observable changes in morphology in response to fluid shear stress. In particu-
dynamic, responsive systems of lar, when exposed to steady flow, they typically adopt a morphology that is
different cell types that respond to aligned in the direction of flow (Figure 11.2), although this can be dependent
external stimuli. This adds to the on the location of the cells within the vasculature. Endothelial cells isolated
growing body of evidence that the from heart valves tend to realign perpendicular to the direction of shear stress
body has few (or perhaps no) passive
(Figure 11.3). In vitro studies investigating the responses of endothelial cells to
components.
flow have revealed the capacity for fluid shear to not only regulate all morphology,
MECHANICAL SIGNALS 313
but also to regulate the levels of a variety of other regulatory molecules such as
vasodilators/vasoconstrictors and growth factors. Recirculant, or “disturbed,”
flows have been found to result in monolayer disruptions and alterations in cell
division that have been linked to sites of atherosclerotic plaque formation in
humans (Figure 11.4). This suggests that the onset of atherosclerosis may be
linked to flows that are unsteady. This hypothesis has ample support from in
vitro and in vivo experiments,. Disturbed flows tend to occur near vascular
bifurcations, and these bifurcations are where plaques tend to form. Fluctuating,
fluid shear stress may contribute to the development and progression of
atherosclerosis.
Flow 24 h
force
w
fluid flo
compression
tension
force
MECHANICAL SIGNALS 315
within the lacunar–canalicular system that drives flow from areas of compres-
sion to areas of tension (Figure 11.5), exposing osteocytes to fluid shear stress.
An important distinction of osteocyte flow sensing compared with luminal flow
sensing is that although luminal flow (in blood vessels or within kidney tubules)
can be readily observed, experimental measurement of bone flow is very diffi-
cult. This is due to the technical challenges associated with the microscopic
fluid spaces of the lacunar–canalicular system (~0.1–1 μm). However, the
recent use of fluorescence recovery after photobleaching (FRAP) has provided
some insight (Figure 11.6). In this approach, a fluorescent tracer is injected
into the animal and allowed to equilibrate within the lacunar–canalicular sys-
tem. A bone surface is exposed, and a laser scanning confocal microscope is
used to photobleach a single lacuna. Infilling of unbleached tracer molecules
from surrounding lacunae and canaliculi results in gradual recovery of fluo-
rescence in the photobleached lacuna. In the presence of flow, infilling is
enhanced by convective transport (the motion of tracer carried by the fluid).
We can determine whether flow is occurring by comparing transport rates in
the presence or absence of loading. Further, the convective velocity can be
estimated by mathematical modeling. Several experiments in cultured bone
cells have demonstrated that they sense and respond to levels of fluid flow
predicted to occur in habitual activities of daily living. Fluid flow within the
musculoskeletal system is not limited to bone. Like bone, articular cartilage is
a mechanosensitive tissue that is well known to adapt its properties to the
loading it experiences. Within cartilage, chondrocytes are embedded in an
extracellular matrix consisting primarily of proteoglycans, collagens, and
water. With mechanical loading, a variety of physical signals are generated
including hydrostatic pressure, matrix strain, and fluid flow. Several experi-
ments in cultured chondrocytes as well as in perfused cartilage and tissue-
engineered constructs indicate that fluid flow is a potent regulator of cartilage
metabolism.
Fluid flow exposes cells to several different physical sig- where Q is the flow rate, μ is the fluid viscosity, b is the
nals, and it is often desirable to distinguish which of chamber width, and h is the chamber height. Equation
these signals are driving flow-induced cellular responses. 11.1 implies that one can subject cells to the same shear
Flow exposes cells to convective chemotransport, in stress but different flow rates by adjusting the viscosity
other words the carriage of nutrients and/or signaling of the flow media (this can be achieved by adding neu-
molecules by the fluid motion. It also exposes cells to tral dextran to the media). The same shear stress would
fluid shear stress. Because these signals are coupled, one result from a doubling of the fluid viscosity and halving
question is how do we determine whether cells are being of the flow rate, or a halving of the viscosity and a dou-
stimulated by chemotransport or fluid shear? bling of the flow rate. If cell responsiveness increases
with flow rate but not shear stress the cells are respond-
One way to isolate the effects chemotransport in parallel
ing to chemotransport. Alternatively, if the responsive-
plate flow chambers is to alter shear stress parametri-
ness of cells increases with shear stress but not flow
cally while keeping flow rate constant, and vice versa.
rate, it suggests that the cells are responding to fluid
Recall from Equation 10.2 that, in these chambers, shear
shear.
stress is related to flow rate as
6 µQ
τ = (11.1)
bh 2
Sensation of deformation within elastic, hollow organs such as the lung and blad-
der is critical for their proper function. Within the lung, breathing generates cyclic
matrix stretch. In vitro, stretch has been shown to regulate lung cell growth,
remodeling of the cytoskeleton, activation of signaling molecules, and phospho-
lipid secretion. Within the bladder, during the storage phase, stretch stimulates
afferent neurons. Bladder epithelium is also sensitive to stretch, as release of
adenosine triphosphate (ATP), acetylcholine, and nitric oxide occurs under
mechanical stimulation. Elucidation of these mechanosensitive pathways may
provide potential pharmacological targets for future treatment of lung or bladder
dysfunction.
Because cells must be immersed in culture media during To isolate the effects of substrate strain versus fluid flow,
substrate stretch experiments, some degree of fluid investigators exploited the dependence of strain on sub-
movement (sometimes referred to as parasitic flow) is strate thickness during four-point bending. In particular,
unavoidable. Careful attention must be paid to minimize the beam-deflection equation for four-point bending is
parasitic flow, as it can severely confound interpretation
td
of experimental results. A classic example can be found ε= ( L − 1.33α ), (11.2)
in the field of bone mechanotransduction. Bone is a rela- α
tively stiff material, so the strains it is subjected to are where ε is strain on the substrate surface, t is the sub-
relatively small: on the order of hundreds to thousands strate thickness, d is the displacement, L is the length
of microstrains (abbreviated as με). Currently, the con- between inner supports, and α is the length between
sensus of the scientific community is that these small the inner and outer supports. By seeding cells on slides
strains are generally insufficient to stimulate bone cells of different thicknesses, investigators were able to apply
in vitro, but this view was not always accepted. For many different strains while keeping displacement rate (and
years, strains of one thousand με were believed to be suf- parasitic flow) constant, and vice versa. They found that
ficient to stimulate bone cells because in vitro cell- the cells were insensitive to changes in strain, but were
straining systems were based on bending (Figure 11.9A). highly sensitive to the displacement rate. This shows
Because bending resulted in a large displacement of the that cells were responding to fluid flow rather than sub-
substrate, these systems generated a significant amount strate strain. Because of these and other similar studies,
of parasitic fluid flow. In particular, the faster the dis- alternative approaches to cell stretching that minimize
placement through the media, the higher the levels of parasitic flows have been developed (Figure 11.9B).
fluid flow to which the cells were subjected.
(A)
(B)
Figure 11.9 Modalities for inducing substrate strain in vitro. (A) Cells are seeded on a substrate subjected to four-point
bending. The cells undergo a large degree of movement within the culture medium, exposing them to parasitic flow. (B) Cells are
subjected to substrate strain in a manner that reduces parasitic flow.
fixed tectorial membrane Figure 11.10 Inner ear hair cells. The
hair cells of the inner ear are responsible
for hearing. They transduce vibrations of
the cochlear basilar membrane into
stereocilia nerve impulses.
It is also worth noting that, by virtue of the topic, this chapter is more biological
and less quantitative than the others in this text.
substrate
fluid flow
substrate
stress. There is also evidence that cadherins, the primary component of adherens
junctions, and vascular endothelial growth factor receptor 2 (VEGFR-2), the
receptor for VEGF, forms a complex with PECAM-1 that is activated by fluid shear.
actin-binding
protein
(C)
membrane from shear stresses. So far, studies investigating the role of the glycoca-
lyx in flow sensing have focused primarily on endothelial cells.
Though the role of the glycocalyx in flow sensing has been studied primarily in the
context of endothelial mechanotransduction, membrane-associated proteins
and glycoproteins have also been proposed as facilitators of mechanotransduc-
tion in bone. Within the lacunar–canicular system, osteocyte processes are sur-
rounded by a pericellular matrix that is believed to be similar in composition to
the endothelial glycocalyx (Figure 11.14). In the presence of fluid flow, drag forces
are generated on the pericellular matrix that are transmitted to the cell surface.
Theoretical calculations predict that these forces are much higher than those gen-
erated by fluid shear alone.
Nota Bene Although evidence for cytoskeletal mechanosensing is relatively scarce, the role
of the membrane in mechanosensing is much better developed. The membrane is
Membrane-spanning regions. an ideal site, as it contains specific structures that confer sensitivity to the envi-
Membrane-spanning regions are
commonly α-helices and can often
ronment while maintaining the integrity of the intracellular components. Recall
be identified by looking for a series of that the bilayer represents a hydrophobic layer of a specific thickness (about
hydrophobic amino acids of the 7 nm). Any protein that contains a hydrophobic region, known as membrane-
correct length from a genetic spanning regions, of the same thickness can become embedded within the mem-
sequence. This can be a critical step brane. Proteins such as receptors that have binding or activation domains on both
in estimating a protein’s structure the extracellular and intracellular sides of the cell can facilitate signaling through
and function from sequence
the cell membrane by conformational changes that occur when a ligand binds or
information alone. In the case of
multiple membrane-spanning dissociates from one side. Analogously, changes in the physical and chemical
domains, the protein will fold back on properties of the lipid bilayer after exposure to mechanical force can play a critical
itself, with each membrane transition role in supporting mechanotransduction by potentially inducing conformational
forming intracellular or extracellular changes in membrane proteins.
loops. If a membrane-spanning
protein has an odd number of Perhaps one of the best-studied mechanisms of mechanosensing, membrane
membrane-spanning domains, the tension, is well known to contribute to increased probability of ion channel open-
carboxy (C)- and amino (N)-terminal ing. Stretch-activated channels have been studied in detail in membrane patches
ends of the protein will be on or fragments of bilayers anchored to the end of a micropipette. Pipette pressure
opposite sides of the membrane. can be changed to regulate the areal strain in the patch, while opening and closing
Likewise, if the protein has an even
number of membrane-spanning
of the channels is assayed by measuring electrical conductance.
domains, the C and N termini will be Other than tension, alterations in other membrane physical properties may be
on the same side of the membrane, involved in mechanosensing. Shear-induced alterations in the membrane may
either intracellularly or extracellularly.
initiate mechanotransduction through enhancing/inhibiting interactions
between signaling molecules. Several studies have demonstrated that membrane
fluidity can be altered by fluid flow. Membrane fluidity refers to the changes in
MECHANOSENSING ORGANELLES AND STRUCTURES 325
membrane viscosity that may arise from local ionic transients coupled with aggre-
gation of proteins. Alterations in membrane fluidity may induce conformational
changes in transmembrane proteins, or enable a change in the rate of interaction
of membrane proteins. In particular, because proteins with membrane-spanning Nota Bene
domains float within the cell membrane, they can be thought of as being confined Small GTPases regulate
within a two-dimensional space. This confinement can greatly alter their chemi- cytoskeletal behavior. G-proteins
cal kinetics. Proteins are much more likely to encounter and interact with each are a particular type of GTPase.
other if they are constrained to the two-dimensional space of the membrane than Another type of GTPase linked to
if they must find each other in the full three-dimensional cytoplasm. Mechanically many mechanobiological functions
includes the small GTPases,
induced changes in membrane fluidity may serve to alter normal interactions consisting of Rho, Rac, Rap, and
within this space, or allow aggregation of new groups of proteins that breach the others. These GTPases are activated
typical compartmentalization. when bound to a GTP. Knocking out,
or mutating, various GTPases leads to
Finally, membrane-bound proteins may also act as direct mechanical receptors alterations in cell morphology, actin
independent of changes in membrane physical properties. Fluid shear has been organization, migratory capability,
shown to be sufficient to activate heterotrimeric G-proteins reconstituted into spreading, and adhesion. Though
liposomes, even in the absence of any other potential mechanotransducing mol- these molecules have been
ecules. As described briefly in Section 2.2, G-protein-coupled receptors are mem- implicated in the control of several
brane-spanning receptors that are linked to G-proteins. When activated, the pathways related to motility and
adhesion, the role of these small
receptor undergoes a conformational change that allows the guanosine triphos-
GTPases in mechanotransductory
phate (GTP) to exchange for the guanosine triphosphate (GDP) on the G-protein, pathways is not well understood.
resulting in activation of the G-protein.
0.2 µm 0.1 µm
326 CHAPTER 11: Cellular Mechanotransduction
Nota Bene
the membrane, tiny invaginations of the membrane known as caveolae (Latin for
“small cave”) occur. Both forms of lipid raft have been demonstrated to mediate
Membrane structure. Lipid rafts mechanotransduction in a variety of cell types, although the specific mechanism
are not the only structure within the
has not been established. However, similar to the mechanisms described above in
cell membrane. Barriers within the
membrane can form from interaction the membrane, it has been proposed that mechanical force could alter the phys-
with the actin cytoskeleton. The part icochemical properties of the lipid rafts, leading to the direct activation of indi-
of the actin cytoskeleton that vidual signaling molecules residing within them, or facilitating interactions
underlies and supports the between molecules.
membrane is sometimes referred to
as the cortical cytoskeleton. It is
periodically attached to the
membrane via anchoring proteins. 11.3 INITIATION OF INTRACELLULAR SIGNALING
Also, there are large cell–cell
signaling structures such as synapses In the first section we provided a survey of the types of mechanical stimulus
and gap junctions and cell–cell that cells are exposed to, and in the second section we discussed structures in
mechanical connections known as
cells that transmit loads from the environment to mechanically sensitive mol-
desmosomes. There are also cell–
substrate interactions, such as focal ecules. This section focuses on the final event of mechanotransduction, load-
adhesions and hemidesmosomes, induced protein conformational changes and the generation of an intracellular
which are responsible in part for signaling cascade. In particular, we discuss two potential mechanisms by which
attaching the cell to a surface. mechanical force can be transduced into a biochemical signal at the molecular
level, opening of mechanosensitive ion channels, and exposure of cryptic bind-
ing sites.
unfavored
hidden cryptic
binding site
expose new binding sites at which chemical reactions can occur (Figure 11.17).
Certain structural motifs exist that may facilitate force-induced conformational
changes in a predictable and force-dependent manner. Some proteins that
mechanically link integrins to the cytoskeleton (and would likely be exposed to
large forces during fluid shear and/or cell stretching) have been found to possess
repeating sequences of structural motifs or modules (such as α-actinin and talin).
The mechanical stability of these modules would dictate the sequence in which
the protein unravels under force, with the more unstable modules unfolding
before the stable ones. These intermodule differences in mechanical stability
could result in sequential exposure of cryptic binding sites with increasing load
magnitudes. A molecule that undergoes such sequential conformational changes
under increasing loads could allow the cell to sense force magnitude.
F ∆x
k = k0 e , (11.11)
kBT
where k0 is the unfolding rate of an unloaded protein, F is the force, and Δx can be
related to the extension of the protein with unfolding but is more correctly termed
the effective energy barrier width.
photon
FRET
CFP YFP
Nota Bene one another (typically, FRET occurs at distances of 10 nm or less). By monitoring
changes in FRET efficiency (the ratio of acceptor–donor emission intensity) in a
Enzymes strain their substrates. cell under mechanical load, one can probe whether mechanical forces can be
It is well known that enzymes exert
strain on their substrates upon
transformed into a conformational change in the candidate protein of interest
binding, thereby catalyzing the (Figure 11.18).
underlying reaction. Straining the
enzymatic substrate could therefore
potentially regulate enzyme activity.
In this way, mechanical force has
11.4 ALTERATION OF CELLULAR FUNCTION
been hypothesized to modulate
enzymatic activity independently of After the initial molecular mechanotransduction event, an intracellular biochem-
exposure of cryptic binding sites. ical signaling cascade must be initiated if these conformational changes are ulti-
However, this mechanism has been mately to lead to alterations in cellular function. In Section 2.2, we provided a
relatively unexplored. survey of intracellular signaling mechanisms and pathways. Many of these have
been shown to be activated by mechanical stimulation. In general, the investiga-
tion of specific signaling pathways is driven, in large part, by physiological rele-
vance for a given cell type. In vascular endothelial mechanotransduction, we
might be interested in a pathway that ultimately leads to regulation of vascular
tone. In bone cells, we might be interested in a pathway that has been shown to
lead to increased mineralization. Many detailed reviews of mechanically regu-
lated intracellular signaling pathways are available (many of which are specific for
certain cell or tissue types); we will not try to recapitulate them here. Rather, we
focus on responses to mechanical stimuli that appear common across the many
cell types. We note that, the material presented is somewhat general, but these are
very active areas of research.
second messenger molecule, capable of eliciting a wide range of effects that can Nota Bene
affect many different downstream signaling molecule pathways.
Intracellular calcium and cell
Intracellular Ca2+ signaling can be visualized in real time through the use of fluo- mechanics. A variety of processes
rescent Ca2+ indicators (calcium ionophores or chelators). In vitro, such dyes can associated with the generation or
be loaded into cells relatively easily, and, when combined with fluorescent imag- resistance to mechanical force
ing, allow the visualization of the initiation and propagation of intracellular Ca2+ depend on intracellular calcium.
Muscle contraction depends on
waves in response to mechanical stimuli. Mechanical activation of intracellular
calcium interacting with the troponin/
calcium signaling, in general, occurs very quickly (generally on the order of sec- tropomyosin system for actin–myosin
onds after the initial mechanical stimulus), and the calcium waves can spread contraction to occur. Further, calcium
throughout the cell rapidly (on the order of milliseconds). Mechanically induced is required for certain types of cell
intracellular Ca2+ elevation is thought to be mediated by the rapid opening of adhesion, such as the calcium-
mechanically sensitive calcium channels in the cell membrane, or release of cal- dependent adhesion molecule
cium from intracellular Ca2+ stores. Once elevated, intracellular calcium then cadherin.
goes on to activate other, downstream pathways.
Nitric oxide, inositol triphosphate, and cyclic AMP, like Ca2+, are
second messenger molecules implicated in mechanosensation
Although intracellular calcium signaling is the most well-studied of the second Nota Bene
messenger systems, several other second messenger molecules have been Ionophores and chelators. An
shown to exhibit rapid changes soon after exposure to mechanical stimulation. ionophore allows the transfer of ions
Recall from Section 2.2 that second messengers can be broadly classified into through a hydrophobic barrier such
one of three categories: hydrophobic molecules associated with the cell mem- as the bilayer, in which they would
normally be insoluble. Ionophores
brane, dissolved gases, and molecules that do not freely cross lipid membranes. typically encase the ion in a polar
Like Ca2+, cyclic adenosine monophosphate (cAMP) is a second messenger mol- interior while exposing a hydrophobic
ecule that falls into the latter category. cAMP is synthesized from ATP by a pro- exterior to the outside. A chelator
cess that is catalyzed by adenylyl cyclase. Activation of adenylyl cyclase is typically forms multiple stable bonds
generally associated with activation of G-protein-coupled receptors. Levels of with metal ions, inactivating them
cAMP can also be regulated by cyclic nucleotide phosphodiesterases, which from their normal function or effect.
It is derived from the Greek word for
degrade cAMP. cAMP has been shown to be rapidly regulated after exposure to
“lobster claw”, Chelè.
mechanical loading in a variety of cell types. Adenylyl cyclase exists in several
different isoforms whose expression can be highly tissue-specific. The activity of
many of these isoforms can be modulated by Ca2+, allowing cross-talk between
the Ca2+ and cAMP pathways.
Recall from Section 2.2 that inositol triphosphate (IP3) is a second messenger
molecule associated with the cell membrane. It is synthesized by hydrolysis of
the molecule phosphatidylinositol 4,5-bisphosphate (PIP2) by phospholipase
C, a phospholipid that is localized in the plasma membrane. Upon cleavage of
PIP2 to IP3 by phospholipase C, IP3 diffuses to the endoplasmic reticulum and,
after binding to IP3 receptors, initiates intracellular calcium release. Several
studies have implicated IP3 in mediating mechanically stimulated intracellular
Ca2+ release.
The dissolved gas nitric oxide can be a potent second messenger because of its
ability to quickly diffuse though the cytoplasm and across lipid membranes.
Shear-induced nitric oxide production in endothelial and bone cells has been
widely observed. Production of nitric oxide is catalyzed by nitric oxide synthase
(NOS). In mammals, the endothelial isoform of NOS, eNOS (also known as
NOS-3) is a primary regulator of vascular tone. In particular, eNOS-derived
nitric oxide has been shown to be a potent vasodilator, relaxing smooth muscle
within blood vessels.
It is noteworthy to point out that second messenger molecules like those described
above have often been used as primary outcome measures for assessing the
mechanosensory role of a particular protein. In general, implicating a molecule as
having a mechanosensory function by inhibitory strategies can be more straight-
forward when assaying a cellular response that occurs fairly rapidly after sti
mulation. This is because after long exposures to loading, numerous molecular
332 CHAPTER 11: Cellular Mechanotransduction
Nota Bene
interactions often occur, and thus it becomes harder to interpret whether effects
on mechanosensation caused by inhibition of a particular molecule are because
Parallel activation of signaling the molecule in question has a direct mechanosensory function, or whether it is
pathways. Although several studies
merely involved somewhere “downstream” in the signaling pathway.
have implicated calcium in mediating
mechanosensing in a variety of cells,
in many cases these cells retain the
capacity to exhibit some
Mitogen-activated protein kinase activity is altered after
mechanically induced responses exposure to mechanical stimulation
even after complete inhibition of
loading-induced intracellular Ca2+ Downstream of second messengers, protein–protein signaling cascades involv-
activation. This suggests that there ing phosphorylation and dephosphorylation can be critical (see Section 2.2). A
are several signaling mechanisms particularly prominent signaling mechanism often activated within minutes
acting in parallel with the Ca2+ after exposure to mechanical stimulation is the group of the mitogen-activated
response.
protein kinases (MAP kinases). MAP kinase phosphorylation in response to most
(including mechanical) stimuli can be highly dynamic, occurring soon after
stimulation and remaining in the phosphorylated state for many minutes. One
well-studied MAP kinase is extracellular-signal-regulated kinase 1 and 2
(ERK1/2), which regulates cell growth and differentiation. Another is c-jun
Nota Bene N-terminal kinase (JNK), which is a so-called stress-activated protein kinase.
Note that “stress-activated”, in this context, refers to systemic stresses and not
MAP kinases and cell growth. As
the name suggests, the MAP kinases necessarily mechanical ones (such as, heat shock or chemical shock). Because
can be activated by pro-proliferative JNK is implicated in apoptosis and inflammation, it is particularly of interest in
signals (mitogens). Though MAP the study of chronic conditions such as atherosclerosis, which is likened to a
kinase activation is associated with slow-acting inflammation (similar to how metal rusting is a slow combustion
cell growth and division, in several process).
cases mechanically stimulated cells
may exhibit rapid MAP kinase One interesting aspect of MAP kinases is that they are capable of directly activat-
activation but no downstream ing transcription factors, the DNA-binding proteins that control gene transcrip-
alteration in proliferation. The factors tion. ERK1/2 is known to activate the transcription factor Elk1. JNK has been
determining whether mechanically shown to activate several transcription factors, including c-Jun, Elk1, SMAD4,
induced MAP kinase activation leads
to tissue growth are unclear. This is
ATF2, and NFAT1. By possessing the capacity to be quickly activated and to
an important issue to clarify because, directly regulate transcription factors, MAP kinases are ideally suited to partici-
in some cases, such growth is pate in pathways involved in early gene expression. Indeed, several components
undesirable. When the smooth of the MAP kinase pathway have been implicated in the expression of primary
muscle cells lining the vasculature response genes, in other words genes whose expression is altered soon after stim-
are exposed to altered mechanical ulation and that do not require de novo protein synthesis.
stresses, they may proliferate in
response, leading to an undesirable
thickening of the vascular wall. Mechanically stimulated cells exhibit prostaglandin release
Another well-characterized response of cells to mechanical stimulation is prosta-
glandin release. Prostaglandins are lipid compounds that are enzymatically
derived from fatty acids and mediate several cellular functions. There are several
types of prostaglandin, of which prostaglandin E2 (PGE2) is arguably the most
well studied in cellular mechanotransduction. PGE2 is synthesized from arachi-
donic acid, which is derived from membrane phospholipids. This arachidonic
acid is converted to prostaglandin G2 and subsequently prostaglandin H2 by the
enzyme cyclooxygenase (COX). A final isomerization step converts prostaglandin
H2 into the biologically active PGE2. COX exists in constitutive (COX-1) and induc-
ible (COX-2) isoforms and is considered the rate-limiting enzyme in the PGE syn-
thesis process. Several studies have demonstrated that mechanical stimulation
activates COX2 gene expression, elevates COX-2 protein levels, and induces PGE2
release into the extracellular environment. Upon its release, PGE2 may initiate
signaling cascades in an autocrine fashion or in other cells by binding PGE2
receptors on the cell surface, such as the receptor for PGE2 (EP2).
cell morphology. It has been widely demonstrated that when cells are subjected to
substrate stretch or fluid shear (generally over an extended period, on the order of
hours or sometimes days), cells may actively remodel their cytoskeletons and
change their gross shape. As described in Section 11.1, endothelial cells subjected
to fluid flow have been observed to align in the direction of flow. These cells have
also been demonstrated to align perpendicular to the direction of uniaxial stretch.
Other cells, such as smooth muscle cells and fibroblasts may exhibit alignment
parallel or perpendicular to major stretch directions, depending on the type of cell
used, the plating conditions, and the exact nature of the stretch (such as percent-
age strain and applied frequency).
In some cases, cells may exhibit no gross changes in morphology but may undergo
extensive cytoskeletal remodeling. Mechanically induced formation of actin stress
fibers has been observed in many cells. The stress fibers can exhibit a preferential
alignment relative to the primary direction of stretch or flow. It has been proposed
that these alignments may serve to minimize intracellular stresses (Figure 11.19).
(B)
blood pressure, fibrosis may be induced. In this case, the heart can become much
stiffer, resulting in increased workload for the heart muscles.
Key Concepts
• The sensing of mechanical signals by cells, or cellular adhesions, the nucleus, and primary cilia have all
mechanotransduction, is critical to many aspects of been implicated as potential sites of
physiology and understanding disease. Four distinct mechanosensing.
phases are involved: conversion of tissue or organ- • Mechanosensitive ion channels are a major class of
level loads into cell-level physical signals; force mechanosensitive molecules. They are known to
detection by mechanosensitive molecules undergoing respond to membrane tension as well as to thinning
a conformational change; activation of due to hydrophobic mismatch.
intracellular signaling systems; and altered cell
• Other proteins can change their enzymatic potential
metabolism.
by buckling, unfolding, and exposure of cryptic
• Many cellular-level physical signals have been shown binding sites. FRET is a powerful way to detect such
to be potent regulators of cell metabolism, including changes fluorescently.
fluid flow, stretch, and pressure.
• Second messenger signaling including IP3, cAMP, and
• Specialized excitable cells mediate our senses of Ca2+ signaling have been shown to be activated by
touch and hearing. They are some of the best- mechanical signals. This leads to activation of protein
understood mechanosensing cells and often have signaling cascades such as MAPK signaling, and
highly sensitive cellular structures capable of intercellular signaling such as PGE2. Ultimately, these
exquisite sensitivity. cascades lead to altered gene expression,
• In nonexcitable cells, structures such as the modification of the extracellular matrix, and changes
glycocalyx, cell membrane, cytoskeleton, focal in cell viability.
Problems
1. Estimate the force required to remove a membrane- Would that be sufficient to overcome the activation
bound protein from the bilayer if it has a cylindrical energy of a channel?
hydrophobic region that is 2 nm in diameter and 5 nm in
length. You can do this by assuming that the mechanical 3. Figure 11.11 depicts a primary cilium undergoing
work done by the pulling force is equal to the change in bending. From the figure, estimate the magnitude of
free energy caused by exposing the hydrophobic domain deflection and strain in the membrane surrounding
to water. You will also need to assume that 5 nm of the cilium at its base. You may assume that the
displacement is sufficient to remove the protein. cilium diameter is 200 nm. Finally, use the result
from Problem 8.4 to estimate what force applied
2. Consider the bending hair bundle from Example 11.1. to the tip would produce this magnitude
How much energy was put into deflecting the tip? of bending.
Annotated References 335
4. One of the roles of the cytoskeleton is to connect gradient? How does this compare with the channel
the nucleus to the rest of the cell. This force activation energy? Assume a typical resting potential
transmission is thought to be involved in cellular of −70 mV.
mechanotransduction. If a cell is exposed to an
externally imposed deformation, describe how you 6. In Example 11.3, the change in free energy for a
would expect the nucleus to deform in response, if (a) channel moving from a closed to an open configuration
the cytoskeleton is removed, (b) the cell cytoplasm acts was estimated to be 14kBT. Using the Boltzmann
like a continuum solid without a cytoskeleton per se, equation, predict the fraction of the time that the
or (c) the cytoskeleton acts like a tensegrity structure, channel would be open owing only to thermal
with normal cytoplasm. Specifically, how would the fluctuations. How would this change if you account for
magnitude of the nuclear deformation compare with the work done by membrane tension as the channel
that of the cell overall? opens? Assume that the cell is at its rupture or lytic
strain (3%) and a typical areal expansion modulus, KA ,
5. Ion channels are sometimes gated by the passage of 0.5 N/m.
of ions themselves. Specifically, they can lose
conductance as ions flow through them. One putative 7. For a bilayer with an areal expansion modulus
mechanism is that the ions moving through the KA = 1.0 N/m, what would the stiffness in the transverse
channel supply the energy to close the channel. direction be? In other words, for a force applied through
How much energy is associated with a monovalent the thickness, what is the slope of the force–deflection
ion moving through the channel along its electrical curve?
Annotated References
Anderson RG (1998) The caveolae membrane system. Annu. Rev. tion examined from a genetics perspective, particularly in sensory
Biochem. 67, 199–225. Gives a comprehensive review of the cell cells.
biology of caveolae.
Eyckmans J, Boudou T, Yu X & Chen CS (2011) A hitchiker’s guide to
Birukov KG, Birukova AA, Dudek SM et al. (2002) Shear stress- mechanobiology. Dev. Cell 21, 35–47. A summary of mechanotrans-
mediated cytoskeletal remodeling and cortactin translocation duction mechanisms, primarily focused on non-sensory cells.
in pulmonary endothelial cells. Am. J. Respir. Cell Mol. Biol. 26,
453–464. This journal research article describes the response of Gefen A (2011) Cellular and Biomolecular Mechanics and Mechano-
pulmonary endothelial cells to applied shear stress. The study biology. An edited text describing recent advancements in cell and
also describes the effects of certain GTPases and early response molecular mechanics and mechanobiology.
mechanoresponsive pathways to shear. Hamill OP & Martinac B (2001) Molecular basis of mechanotransduc-
Brown TD, Bottlang M, Pedersen DR & Banes AJ (1998) Loading tion in living cells. Physiol. Rev. 81, 685–740. A review of molecular
paradigms—intentional and unintentional—for cell culture mecha- mechanotransdcution mechanisms with a focus on the membrane
nostimulus. Am. J. Med. Sci. 316, 162–168. A numerical analysis of and channels; the primary source for the material on hydrophobic
parasitic flows in bending-based systems for studying the strain mismatch and channel-membrane coupling.
response in vitro. Jacobs CR, Temiyasathit S & Castillo AB (2010) Osteocyte mecha-
Chalfie M (2009) Neurosensory mechanotransduction. Nat. Rev. nobiology and pericellular mechanics. Annu. Rev. Biomed. Eng. 12,
Mol. Cell. Biol. 10, 44–52. Comprehensive review of molecular and 369–400. Provides a comprehensive review of mechanosensory
cellular mechanisms of mechanosensing in sensory cells. structures and mechanisms identified in bone cells.
Chancellor TJ, Lee J, Thodeti CK & Lele T (2010) Actomyosin ten- Knothe Tate ML, Steck R, Forwood MR & Niederer P (2000) In vivo
sion exerted on the nucleus through Nesprin-1 connections influ- demonstration of load-induced fluid flow in the rat tibia and its
ences endothelial cell adhesion, migration, and cylic strain-induced potential implications for processes associated with functional
reorientation. Biophys. J. 99, 115–123. Provides recent evidence of adaptation. J. Exp. Biol. 203, 737–745. Describes quantification of
the role of nucleus loading and nuclear matrix proteins in mecha- loading-induced flow in bone.
notransduction.
Kooppel DE, Axelrod D, Schlessinger J et al. (1976) Dynamics of
Corey DP & Hudspeth AJ (1979) Response latency of vertebrate hair fluorescence marker concentration as a probe of mobility. Biophys.
cells. Biophys. J. 26, 499–506. Early data on the molecular mecha- J. 16, 1315–1329. An early description of fluorescence recovery
nism of hair cell mechanosensing, particularly the incredibly fast after photobleaching used to measure lateral diffusion in the cell
response time. membrane.
DeBakey ME, Lawrie GM & Glaeser DH (1985) Patterns of athero- Kung C (2005) A possible unifying principle for mechanosensation.
sclerosis and their surgical significance. Ann. Surg. 201, 115–131. Nature 436, 647–654. Provides a concise summary of the roles of
This journal article presents an analysis of common atherosclerotic mechanosensitive ion channels in touch sensation and hearing.
lesion development sites as noted in clinical examinations. The arti-
Malone AM, Batra NN, Shivaram G et al. (2007) The role of actin
cle further discusses the clinical aspects of atherosclerosis, includ-
cytoskeleton in oscillatory fluid flow-induced signaling in MC3T3-
ing classification, progression and recurrence of disease.
E1 osteoblast. Am. J. Physiol. Cell Physiol. 292, C1830–C1836. Dem-
Ernstrom GG & Chalfie M (2002) Genetics of sensory mecha- onstrated differential cytoskeletal remodeling in cells exposed to
notransduction. Annu. Rev. Genet. 36, 411–53. Mechanotransduc- static and dynamic fluid flow.
336 CHAPTER 11: Cellular Mechanotransduction
Nauli SM, Alenghat FJ, Luo Y et al. (2003) Polycystins 1 and 2 mediate evidence that fluid flow is a critical cell-level physical signal in bone
mechanosensation in the primary cilium of kidney cells. Nat. Genet. mechanobiology.
33, 129–137. A discussion of the polycystins and their putative role
in primary-cilium-based mechanosensing. Reilly GC, Haut TR, Yellowley CE et al. (2003) Fluid flow induced PGE2
release by bone cells is reduced by glycocalyx degradation whereas
Nishiyama M, Shimoda Y, Hasumi M et al. (2010) Microtubule depo- calcium signals are not. Biorheology 40, 591–603. Supplies some of
lymerization at high pressure. Ann. N. Y. Acad. Sci. 1189, 86–90. A the only evidence that the cellular glycocalix is critical for mecha-
demonstration that high hydrostatic pressure can induce microtu- nosensing in bone cells.
bule depolymerization in vitro.
Simons K & van Meer G (1988) Lipid sorting in epithelial cells. Bio-
Nonaka S, Tanaka Y, Okada Y et al. (1998) Randomization of left-right chemistry 27, 6197–6202. An early description on microdomains in
asymmetry due to loss of nodal cilia generating leftward flow of lipid membranes.
extraembryonic fluid in mice lacking KIF3B motor protein. Cell 95,
p. 829–837. Seminal study demonstrating that flow generated by Tabouillot T, Muddana HS & Butler PJ (2011) Endothelial cell mem-
nodal cilia is critical for left–right determination. brane sensitivity to shear stress is lipid domain dependent. Cell.
Mol. Bioeng. 4, 169–181. Direct evidence that microdomains in the
Olsen B (2005) Nearly all cells in vertebrates and many cells in inver- bilayer, including rafts and caveolae, are modified with fluid shear
tebrates contain primary cilia. Matrix Biol. 24, 449–450. An editorial stress and are potentially involved in mechanosensing.
on the ubiquitous nature of primary cilia and their potential physi-
ological function. Vogel V & Sheetz M (2006) Local force and geometry sensing regu-
late cell functions. Nat. Rev. Mol. Cell. Biol. 7, 265–275. Provides a
Owan I, Burr DB, Turner CH et al. (1997) Mechanotransduction in concise review of potential mechanisms by which force-induced
bone: osteoblasts are more responsive to fluid forces than mechan- conformational changes may lead to exposure of cryptic binding
ical strain. Am. J. Physiol. 273 (3 Pt 1), C810–815. Provides evidence sites.
that osteoblasts subjected to four-point bending were responsive to
fluid flow rather than substrate strain. Wang Y, McNamara LM, Schaffler MB & Weinbaum S (2007) A model
for the role of integrins in flow induced mechanotransduction in
Qin YX, Lin W & Rubin C (2002) The pathway of bone fluid flow as osteocytes. Proc. Natl Acad. Sci. USA 104, 15941–15946. Describes
defined by in vivo intramedullary pressure and streaming potential a proposed mechanism for mechanosensing of flow by integrins.
measurements. Ann. Biomed. Eng. 30, 693–702. Some of the earliest
Abbreviations
Chapter 1 ε strain
γ shear strain ε strain vector
σ tensile stress εa axial strain
τ shear stress εt transverse strain
A deformed surface area εxy, εyx, εxz components of strain
A o undeformed surface area θ angle of twist
d membrane thickness θxx rotation angles
ΔP Patm – Ppip κ local curvature
FP resultant force due to pressure λ stretch ratio
Ft resultant force due to tension ν Poisson ratio
k spring constant ϕ eigenvalue
kB Boltzmann’s constant σ stress
Lpro protrusion length σ stress vector
n surface tension σyx, σzx, σzy components of stress
Patm environmental pressure τ shear stress
Pcell cellular pressure a acceleration
Pi internal pressure a deformed vector
Po external (outside) pressure A area
Ppip pipette pressure A undeformed vector
R radius of pressure vessel C FTF, right Cauchy-Green deformation tensor
Ra cell radius at one end D compliance matrix
Rb cell radius at other end E Young’s modulus
R o undeformed radius e strain vector
Rpip pipette radius E Green–Lagrange strain
Rpro protrusion radius F force
V volume F deformation gradient
G shear modulus
Chapter 3 h beam width
α angular acceleration I second moment of inertia
γ shear strain I identity
δ transverse displacement J polar moment of inertia
ΔL column extension k spring constant
ΔR change in radius of curvature kB Boltzmann’s constant
ΔS change in beam length L length
LIST OF VARIABLES AND UNITS 339
M moment L length
m mass m mass
n surface tension n surface tension
P pressure P pressure
P′ transformed vector Re Reynolds number
P original vector t time
Q rotation matrix u fluid velocity
R radius x, y, z spatial coordinates
RU or VR rotation and stretch components of polar
decomposition Chapter 5
Sx resultant force
γ shear strain
SXx components of Sx σ stress vector
SXy components of Sx τ shear stress
SXz components of Sx A cross-sectional area
u, v, w displacements β 1/kBT
v eigenvector b distance traveled in each step, Kuhn length
v1, v2, and v3 principal directions D diffusion coefficient
w beam displacement energetic cost per hairpin
x spring deformation E Young’s modulus
x deformed vector k spring constant
X undeformed vector kB Boltzmann’s constant
x, y, z spatial coordinates L contour length
m microstate
Chapter 4 n surface tension
α constant scaling exponent n, n+ number of flips, number of flips that come
β constant viscous factor out heads
δ phase lag N number of particles
γ shear strain Nh number of hairpin sites
γ* complex shear strain p(m) probability of microstate m
μ viscous coefficient q heat
μeff effective viscosity Qs(ms) energy of microstate ms
ω frequency R end-to-end length
ρ density S entropy
σ stress vector t time
τ* complex shear stress T temperature
τ shear stress V volume
ξ structural damping coefficient Y helmholtz free energy
E elastic modulus or storage modulus z single partition function
E* complex modulus Z partition function
g acceleration due to gravity
G* complex shear modulus Chapter 6
h height α angle
–
i imaginary unit, √–1 γ shear strain
k spring constant δ phase lag
kB Boltzmann’s constant ε strain
340 LIST OF VARIABLES AND UNITS
V volume S entropy
w beam displacement t thickness
x, y, z spatial coordinates T temperature
T time
Chapter 9 utot, vtot total deformation
α scaling factor w transverse displacement
γ shear strain W energy
γint interfacial energy W density of states
ε strain x, y, z spatial coordinates
θ angle
λ length Chapter 10
ρ number of molecules per unit volume γ shear strain
σ stress vector Δt time interval
τ shear stress Δ pitch of polymerized actin helical structure
𝜑(R) excitation intensity θ angle that a membrane makes with a
surface
Ψ free energy
λb stretch ratio of substrate
Planck’s constant
μ fluid viscosity
A area
n membrane tension
c concentration
ρ density
D diffusion constant
σ stress vector
E Young’s modulus
τ shear stress
EI flexural rigidity
a constant
G shear modulus
A area
J flux
ds binding-site spacing
k spring constant
D diffusion constant
kB Boltzmann’s constant
E Young’s modulus or energy
KA areal expansion modulus
F force
KB bending stiffness
F− negative drag force
KS shear stiffness
F+ positive elastic force
l hydrocarbon chain length
h height
lc average bond length between carbon
atoms J adhesion energy density
ℓp persistence length k spring constant
L length/dimension k− unbinding rate constant
m mass of molecule k+ binding rate constant
M, m moments kB Boltzmann’s constant
n force kd equilibrium constant
n surface tension kon rate of reaction
nc number of carbon atoms in chain KB bending modulus
nc number of molecules n surface tension
N number nb area density of bonds
P pressure p probability
P0 laser power R radius or end-to-end length
q heat S speed
R radius t time
342 LIST OF VARIABLES AND UNITS
toff average time per cycle that myosin is not σ stress vector
bound to actin τ shear stress
ton time per cycle that myosin is not bound to ΔA change in area
actin
Δx effective energy barrier width
tr release time
A area
T temperature or tortuosity
d displacement
v or V velocity
EB bulk modulus
w displacement
F force
wbs width of binding site ‘sweet spot’
Fb buckling load
W density of states
k spring constant
W work
kB Boltzmann’s constant
Wadh adhesion energy
KA areal expansion modulus
Wdef strain energy associated with deformation
n surface tension
x maximum length of spring
P pressure
x, y, z spatial coordinates
Q flow rate
xsh shortening distance
R radius
T temperature
Chapter 11 W density of states
γ shear strain W work
ε strain x, y, z spatial coordinates
Index
flexural rigidity 65, 200–202 fourfold connectivity 245 green fluorescent protein (GFP) 40
flippases 252 fractures 19–20 Green’s function 164
flow chambers 167–168 fragility fractures 6 Green–Lagrange strain 80
fluid flow Frank–Starling law 299–300 GTP see guanosine triphosphate
diverse device designs 171–172 FRAP see fluorescence recovery after guanine (G) 22–24
mechanotransduction 311–313, photobleaching guanine nucleotide exchange factor
315–316 free-body diagrams 12, 263–264, 270 (GEF) 36
parallel plate 104 free energy guanosine diphosphate (GDP) 36, 191,
fluid mechanics 89–117 critical micelle concentration 194, 197
dimensional analyses 110–115 253–254 guanosine triphosphate (GTP) 36, 191,
fluid statics 89–92 ideal chain polymers 207–209 192, 194, 197, 325
key concepts 115 microcanonical ensembles 135–136
Navier–Stokes equations 98–103 partition functions 139–141 hair cells 5–6, 319
Newtonian fluids 92–98 statistical mechanics primers hairpin polymers 132–134, 142–143
rheological analyses 103–110 128–131, 135–136, 139–141 half-space 158–159, 164
fluid mosaic model 252 freely joined chains (FJC) 189, hammer bone 319
fluids 210–214, 217–218 hearing 5–6
incompressible fluids 90 frequency-dependent changes 108–110 heart 4, 292, 299–300
mass conservation 93–94 FRET see Förster resonance energy Helmholtz free energy 128, 129
Newtonian fluids 92–98 transfer hemidesmosomes 192–193
non-Newtonian behavior 97–98 friction coefficients 173–174 hemodynamic forces 6
fluid shear fully developed flow 169–180 Hertz contact models 158–159
adhesion 281 functional groups 167, 168 high hydrostatic pressure 90
flow chambers 167–168 Fura-2 fluorescent dyes 33, 40 Hill equations 300–301
mechanotransduction 316, 317 fusion, membrane 8 Hodgkin–Huxley model 305
fluorescence 33, 38–40 Hooke’s law 37, 57–58, 73, 76,
conformation changes 329–330 G-actin 189–190, 195 173–174
cross-correlation 160–161, 162 gap junctions 287 hormones 31
Delaunay triangulation 178, 179 gating see channels human disease frameworks 4–8
morphological changes 333 gauge pressure 91 Huxley, Aldous/Andrew 305
fluorescence recovery after Gaussian approximations 204–205, hydrolysis 21, 197–198
photobleaching (FRAP) 206–207 hydrophilic domains 250, 251
258–259, 315 Gaussian chains 206 hydrophobic domains 250, 251, 327,
fluorochrome 161 GCPRs see G-protein-coupled 328
flux 256, 258 receptors hydrostatic pressure 89–90, 91, 92,
focal adhesions 36 GDP see guanosine diphosphate 317–318
force GEF see guanine nucleotide exchange hyper/hypotonic concentrations 252
adhesion 281–282, 290–291 factor
atomic force microscopy 159 gel electrophoresis 41–43 ideal chain polymers 189, 203–210,
contraction 299–302, 303–308 genes, definition 20, 25 213–214
freely joined polymer chains genetics, in situ studies 48 image-based modeling 236, 237
211–213 genomes 27 image correlation-based approaches
hydrostatic pressure 89–90 genomic data, bioinformatics 49 178–179
ideal chain polymers 208–209 genotypes 27–29 image distortion 162
laboratory cell mechanics 152–156, GFP see green fluorescent protein image texture correlations 162
160–173 Gibbs free energy 129 incompressibility, micropipette
mechanotransduction 321–322, glycoproteins 323–324 aspiration 14
332–333 Golgi apparatus 29 incompressible fluids 90
membranes 269–272 G-protein-coupled receptors (GCPRs) indenters 158–159
optical traps 154, 156 36 indices of refraction 37
power-stroke models 305, 306 gradients inductible cyclooxygenase (COX-2)
statistical mechanics primers deformation 180–181 enzyme 332
119–120 ideal chain polymers 208–209 inertia 65, 239
worm-like polymer chains 216–217 large deformations 78–79, 80–82 infinitesimal strain 72
Förster resonance energy transfer Newtonian fluids 93 inflammation 287, 288, 289
(FRET) 329–330 optical traps 154, 156 informatics 49
forward model 163 grand canonical ensembles 127 inner ear hair cells 319
forward primers 44 gravity 4, 89–90 inorganic compounds 20
Index 347
inositol triphosphate (IP3) 32, 36, lab-on-a-chip technologies 167 MAP see mitogen-activated protein
331–332 laboratory cell mechanics 151–185 kinase activity
in-plane shear and tension 259–267 blinding 181–182 mass action law 288–290
inside-out signaling 36 cellular micromanipulation 151–159 mass conservation 93–94
in situ studies 20, 48 controls 181–182 matrices 76, 316, 324, 333–334
integration 49, 92 deformation analyses 173–181 Maxwell bodies 174–175, 176–177
integrins 36, 192–193, 287 force application 167–173 mechanical loading 177, 178
intensive quantities 140 force measurements 160–167 mechanical stress 56
intermediate filaments 192–193, 194 key concepts 182 mechanosensitive channel of large
internal energy lac repressors 215–216 conductance (MscL) 326, 327
equilibria 141–142 lacunae 313, 314, 315 mechanotransduction 311–336
mechanical state changes 123–124 Lamé constants 75 adhesion 280
microcanonical ensembles 135–136 lamellipodia 293, 294 function alterations 330–334
partition functions 141–142 laminar flow 94–95, 168, 169–170 hearing 5–6
statistical mechanics primers laminin 192–193 intracellular signaling 326–330
120–124, 133–136, 141–142 Langevin functions 213–214 key concepts 334
interstitial fluid flow 313, 314, 315 Laplace law 12–13, 14 mechanical signals 311–318
intracellular strains 179–180 large deformations organelles 318–326
inverse problems 163–165 gradients 78–79, 80–82 steps 311, 312
in vitro studies 20, 292, 318 mechanics 78–83 structures 318–326
in vivo studies 287, 288, 292 rotation 82–83 medial fractures 19–20
ion channels see channels strain 80–82 melanoma cells 297
ionophores 331 stretch 79–80, 82–83 melting temperatures 44–45
IP3 see inositol triphosphate large factorials 133 member prestresses 228–229
isometric contractions 300 laser beams 157–158 membranes 249–278
isotropic hydrostatic pressure 91 lateral fractures 19–20 adhesion 284
Law of Laplace 12–13, 14 barrier functions 255–259
jellyfish 40 law of mass action 288–290 bending 267–274
JNK see c-jun N-terminal kinase lenses, definition 36 biology 249–252
Johannsen, Wilhelm 27–28 leukocytes 297 bowtie formations 172
see also neutrophils contraction 301–302
Kelvin-Voigt bodies 174–176 ligands 8, 30, 31, 32, 288–290 definition 84
key concepts light microscopes 37–38 elasticity 269, 301–302
adhesion 308 light momentum 153–154 fusion 8
cell biology fundamentals 50 linear elastic behavior 269 in-plane shear and tension 259–267
contraction 308 linear scaling 227–228 key concepts 275
fluid mechanics primers 115 linings 7, 312, 313 mechanotransduction 324–326,
frameworks 16 lipids 250, 251–255, 325–326, 332 327, 328
laboratory cell mechanics 182 liposomes 250 micropipette aspiration 11, 12
mechanotransduction 334 liquid-drop models 11–12, 282–284 phospholipid self-assembly 252–255
membranes 275 liquid drops 16 polymer networks 238
migration 308 Listeria bacteria 7, 295 receptors 30
polymer cellular mechanics 219 loading modes 83–84 second messengers 33
polymer networks 246 locomotion see migration stretch devices 172–173
solid mechanics primers 84 Lorenz-Mie theory 156 traction forces 160
statistical mechanics primers 147–148 loss moduli 108–110 two meanings 262
kinases 34, 332 low hydrostatic pressure 90 Merkel cells 320
kinematics 60, 62–63, 71–73 lumen 10, 312, 313 messenger RNA (mRNA) 25, 26,
membrane deformations 260–262 lumped parameter models 176 29–30
membrane rotation 268 lungs 7, 297, 313, 317 metalloproteinases 333–334
viscosity 169 lysosome 29 metastasis 8
kinetics lytic limits 328 MFs see microfilaments
adhesion 288–290 micelles 250, 252–255
polymerization 194–198 macrophages 297 microcanonical ensembles 127,
protein unfolding 329 macroscopic systems 125–126, 127 131–136
Kronecker delta 204 macrostates 124–126, 127–128 microcontact printing 167
Kuhn lengths 203, 206, 208–209, magnetic micromanipulation 151–159 microfabricated micropillar arrays
218–219 magnetic twisting 153 165–166
348 Index
signaling pathways 30–31, 35–36, storage moduli 107–110 tension 11, 12, 259–267, 272–274, 327
311–318, 326–330, 332 strain 73–74 tensors 59
silly putty 104, 105 adhesion 286–287 tetanic contractions 300
similitude experiments 113–114 affine polymer networks 233–236 texture correlations 162
single partition functions 143 anisotropic affine polymer TFM see traction force microscopy
single pipettes 281, 282 networks 233 thermal cyclers 45
sinks 256 biaxial membrane deformation 267 thermodynamic systems
site-specific recombinase (SSR) 48 deformable body networks 57–65, canonical approaches 142–143
sixfold connectivity 240, 242–245 68, 71–73, 76–78 equations of state 134
skeletal muscle contraction 300, 301 elastic deformations 122–123 first law 124
skeleton 4 elastic moduli 233–234 free energy minimization 128–129
sliding forces 303–308 equilibria 123 ideal chain polymers 208–209
small deformations strain 72 flexible substrates 172 Thermus aquatiqus 45
small inhibitory RNA (siRNA) 47–48 large deformations 80–82 thin-film deposition 165
smooth muscle cells 297, 299, mechanotransduction 316–317, 318 thin-walled cylinders 60
316–317 polymer networks 227–228 thioredoxin channels 327
sodium dodecyl phosphate (SDS) thin structure analyses 240–241 three-coin systems 124–126
41, 42 streams 93 three-dimensional orientation
sodium dodecyl sulphate– stress 73–74 correlation functions 200
polyacrylamide gel deformable body networks 56–60, thymine (T) 22–24
electrophoresis (SDS-PAGE) 64–65, 69–71, 76–78 time 256–259, 296–297
41, 42 fibers 301–304 torsion 60, 61
solid cylinder torsion 61 flow chambers 167–168 totipotent cell types 28
solid mechanics 53–87 mechanotransduction 312–314, 316 touch sensation 320
deformable body networks 55–78 membranes 263 traction force 160–161, 163, 164–166
free-body diagrams 53–55 moment resultants 269 traction force microscopy (TFM)
key concepts 84 polymer networks 228–229 160–161, 163, 164–165
large deformations 78–83 protein kinase 332 transcription 25, 26, 27, 28–29
loading modes 83–84 strain energy 123 transduction 5–6, 31, 32
rigid-body mechanics 53–55 symmetry 71 transient intercellular adhesions 287,
shape modes 83–84 stretch 79–80, 82–83, 172–173, 288, 289
structural elements 83–84 317, 324 transient receptor potential (TRP)
Southern blotting 42–43 striated muscle cells 299 channels 327
spectrin polymers 240, 268 structural elements, solid mechanics transition regions 95
speed see velocity primers 83–84 translation 25, 26, 27
spring systems 121, 131, 173–177, 285 substrates 160, 172, 279–280, 318, 330 translocation 293
SSR see site-specific recombinase sugars 24 transverse strains 68
stable intercellular adhesions 287, surface tension 11–15, 249, 282–284 treadmilling 197–198
288, 289 symmetric deformations 172–173 TRP see transient receptor potential
statistical mechanics 119–149 channels
canonical ensembles 136–143 T-actin 197 trusses 84
entropy 124–128 tapered micropipette aspiration 14 tubulin 191, 197
free energy 128–131 tapping atomic force microscopy tumors 8
internal energy 120–124 modes 158 turbulence 94–95, 168
key concepts 147–148 Taxol anti-cancer drug 191 tweezers 153–157
microcanonical ensembles 131–136 Taylor series 69–70, 146–147, 199 twitch forces 300
random walks 143–147 temperature two-dimensional shear moduli
steady flow 92–93 Brownian ratchets 294–295 240–241
stem cells 28 hairpin number 134–135 two-level systems 132, 142–143
stereocilia 319 membranes 272–273 two-pore-domain potassium channel
stiffness persistence length 198–200 protein 327
deformable body networks 57–58 polymerase chain reaction 44–45 two-spring systems 121
mechanotransduction 317 polymers 120, 198–200 tyrosine kinases 34
membranes 268 statistical mechanics primers 129–131
polymer networks 202–203, 228–229 thermal cyclers 45 ultimate stress/strength 58
stirrup bone 319 see also thermodynamic systems uniaxial force 172, 301–302
Stokes’ law 152–153 tensegrity 228–229 uniform biaxial stretches 172–173
Index 351
universal solvents 249 viscoelastic cells 9, 173–177 white light illumination 37–38
uracil (U) 22–24 viscoelastic materials 104–108 worm-like chains (WLC), polymers
viscometers 170–171 189, 214–219
vascular endothelium 312–313, 314 viscosity 104–106, 169, 173–174
vector notation 74–75 Voigt notation 74–75 Y chromosomes 292
velocity volume 14, 93, 94, 206, 225–228 yellow fluorescent protein (YFP)
contraction 300–301, 305, 306 40, 330
migration 296–297 walls, definition 84 Young’s equation 284
Newtonian fluids 93, 94–95 warping 61 Young’s moduli 233, 234, 237, 267
power-stroke models 305, 306 water 89–90, 249, 250
vesicles 29 wavelengths 37, 38 zero temperature limits 129
viability of cells 334 white blood cells 11, 287
viruses 8 see also neutrophils
This page intentionally left blank
to match pagination of print book