An Introduction To Structural Mechanics For Architects
An Introduction To Structural Mechanics For Architects
An Introduction To Structural Mechanics For Architects
Elías Cueto
David González
An Introduction
to Structural
Mechanics for
Architects
Structural Integrity
Volume 4
Series editors
José A. F. O. Correia, Faculty of Engineering, University of Porto, Porto, Portugal
Abílio M. P. De Jesus, Faculty of Engineering, University of Porto, Porto, Portugal
Advisory editors
Majid Reza Ayatollahi, School of Mechanical Engineering, Iran University of
Science & Technology, Tehran, Iran
Filippo Berto, Department of Mechanical and Industrial Engineering, Faculty of
Engineering, NTNU, Trondheim, Norway
Alfonso Fernández-Canteli, Faculty of Engineering, University of Oviedo,
Gijón, Spain
Matthew Hebdon, Virginia State University, Virginia Tech, Blacksburg, VA, USA
Andrei Kotousov, School of Mechanical Engineering, University of Adelaide,
Adelaide, SA, Australia
Grzegorz Lesiuk, Faculty of Mechanical Engineering, Wrocław University of
Science and Technology, Wrocław, Poland
Yukitaka Murakami, Faculty of Engineering, Kyushu University, Higashiku,
Fukuoka, Japan
Hermes Carvalho, Federal University of Minas Gerais, Belo Horizonte, Minas
Gerais, Brazil
Shun-Peng Zhu, University of Electronic Science and Technology of China,
Chengdu, Sichuan, China
The Structural Integrity book series is a high level academic and professional series
publishing research on all areas of Structural Integrity. It promotes and expedites
the dissemination of new research results and tutorial views in the structural
integrity field.
The Series publishes research monographs, professional books, handbooks,
edited volumes and textbooks with worldwide distribution to engineers,
researchers, educators, professionals and libraries.
Topics of interested include but are not limited to:
– Structural integrity
– Structural durability
– Degradation and conservation of materials and structures
– Dynamic and seismic structural analysis
– Fatigue and fracture of materials and structures
– Risk analysis and safety of materials and structural mechanics
– Fracture Mechanics
– Damage mechanics
– Analytical and numerical simulation of materials and structures
– Computational mechanics
– Structural design methodology
– Experimental methods applied to structural integrity
– Multiaxial fatigue and complex loading effects of materials and structures
– Fatigue corrosion analysis
– Scale effects in the fatigue analysis of materials and structures
– Fatigue structural integrity
– Structural integrity in railway and highway systems
– Sustainable structural design
– Structural loads characterization
– Structural health monitoring
– Adhesives connections integrity
– Rock and soil structural integrity
An Introduction to Structural
Mechanics for Architects
123
Elías Cueto David González
Aragón Institute of Engineering Research Department of Mechanical Engineering
University of Zaragoza University of Zaragoza
Zaragoza Zaragoza
Spain Spain
This Springer imprint is published by the registered company Springer International Publishing AG
part of Springer Nature
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
To Susana and Mamen, for their patience.
Preface
This book contains a rather concise compilation of the contents of our lectures on
structural mechanics for the students of the second course of architecture at the
School of Engineering and Architecture of the University of Zaragoza. The contents
are intended to be taught during a semester comprising some 15 weeks and 3 h per
week. The student is assumed to have been exposed to introductory university
courses on physics and mathematics, in particular vector and tensor algebra and
differential calculus.
We believe that most of the literature devoted to structural mechanics has a
profound engineering bias. We decided to write this book just because we believe
that the approach followed herein is rather new. This is not a classical book on
strength of materials nor a classical introductory book on structural mechanics.
Instead, it follows a historical perspective and introduces the reader to the funda-
mental problem of architecture (at least, from the mechanical point of view),
namely avoiding bending. It was not until the Industrial Revolution that architects
began to have broad access to iron and steel, i.e., to materials with the same strength
in tension and compression. Until that moment, the history of architectural masters
was that of people trying to avoid bending. This is at the origin of trusses and also
of buttresses in Gothic cathedrals. This book follows this historical pipeline, by
treating the difficulties associated with stone arches as inverted cable structures
subjected to pure compression, then moving to bowstring structures and trusses,
then frames, and finally carefully introducing bending and its associated
complexity.
But we embark on this quest comme il faut, starting from the theory of elasticity,
and introducing properly the various hypotheses that give shape to strength of
materials as a collection of models derived from elasticity, rather than a collection
of engineering recipes with significant professional interest but no apparent links
between them. We repeat throughout the book the famous aphorism by G. Box: no
model is true, but some are useful. This book could be considered, therefore, as a
well-justified collection of lies.
vii
viii Preface
We therefore begin by a rapid review of the theory of elasticity. To this end, the
reader may find useful the appendix on index (also known as Einstein) notation.
This is because strength of materials is no more than a collection of simplifying
hypothesis established on top of the theory of elasticity that we find this approach
rigorous, yet very useful.
We then begin by studying cable structures, the only ones capable of bearing
traction loads in ancient times. It is precisely Hooke’s finding on the equivalence
between a stone arch and an inverted cable that gave rise to the possibility of
studying, from a scientific standpoint, the problem of finding the form of an arch.
Nevertheless, we incorporate in this chapter not only classical results, but also the
most recent developments in the field of masonry structures.
Bowstring structures emerge then as a natural continuation of arches and also a
seamless introduction to the world of trusses. With the eruption of industrialization
and the availability of first iron and then steel, the possibility of designing structures
subjected to bending arose, thus opening the door for sophisticated frame structures.
Every chapter introduces some worked examples. Some others are left deliber-
ately open so as to allow the reader to further investigate the contents of the book.
None of them are especially difficult, so the interested reader may find a solution in
various, well-known references. It is precisely this (re-)search that will enable the
reader to become truly a master of the field.
We hope the reader will find the book useful and apologize in advance for the
possible mistakes it may contain, despite the effort paid in successive lectures.
We are indebted to the many individuals who have contributed, directly or indi-
rectly, to this book. Particularly, Prof. Alfonso F. Canteli transmitted to us the
tradition, possibly coming from the ETH Zurich, of employing the lower bound
theorem and the stress-field method to achieve a profound understanding of the
behavior of structures. This book takes advantage of this tradition and employs
graphic statics throughout. We truly believe that this method can be employed
successfully to teach undergraduate students and transmit a very intuitive under-
standing of how structures work.
Professor Manuel Doblaré infected us with the passion for rigor, and we have
tried to be the best disciples in this hard task. Sections devoted to beams are
undoubtedly influenced by his work in the field, following the inspiring tradition
of the late Prof. Juan C. Simó.
We must thank, in sum, the faculty at the continuum mechanics division of the
School of Engineering and Architecture of the University of Zaragoza, because it is
by means of sharing good moments over a cup of coffee or around a blackboard that
we finally learn a way of doing things. We are indebted particularly to Profs. Icíar
Alfaro, Begoña Calvo, Fany Peña, and Miguel Ángel Martínez for their help and
support through the years.
Finally, thanks to our students, the interaction with whom provided the most
efficient means of determining the best practices to reach the goal we pursue.
ix
Contents
xi
xii Contents
Abstract This chapter revisits the basics of the theory of elasticity upon which
strength of materials is based. The level is kept to the minimum that will enable
reader to understand subsequent chapters.
Although in many good books strength of materials is developed after some of the
basics of classical Newtonian mechanics (equilibrium, basically, is all that you need
most of the times), it is best understood if we introduce, at least briefly, some basic
concepts of continuum mechanics. This will allow the reader to understand the very
basic hypotheses on which this discipline is based. The price to pay is a little bit of
complexity, but, trust us, you will feel much more comfortable from Chap. 2 onwards.
We will begin by acknowledging that the type of structures we, as architects
or engineers, are interested in, are examples of a continuous medium. Even if we
know that matter is discontinuous (it is formed of elementary particles, atoms an
molecules), at the scale of a building it appears to the human eye as continuous.
Therefore, it will be helpful to consider our structure as an homogeneous portion
of R3 . This is true in general, even if our material is clearly heterogeneous, like
reinforced concrete, for instance. We will be interested in its overall macroscopic
properties, rather than those of their components. Only very advanced theories that
fall outside the scope of this book make effective use of the heterogeneities of the
material through, normally, complex multiscale simulations on powerful computers.
We will also accept other fundamental hypothesis like Saint Venant’s principle,
which states that “the difference between the effects of two different but statically
equivalent loads becomes very small at sufficiently large distances from load” [14]. In
addition, we will restrict ourselves to static-load processes and isothermal conditions.
Under these assumptions, the question is: it is necessary to begin describing the
plethora of physical phenomena taking place in a structure, but …how? In this book
we decided to do it by describing first the way a structure moves: displacements.
There are plenty of good books that begin by introducing a concept coined as stress,
but it is unnatural at this moment, needs some advanced mathematics and, above all, it
is not as visible as the displacements! Displacements are a very intuitive concept and
will provide a useful means to introduce the remaining physical processes occurring
in a structure.
1.1 Kinematics
1.1.1 Movement
If we accept to begin the study of a structure by the description of its movement (as
already mentioned, this is a rather arbitrary choice, and many books prefer to begin
with stresses), let us first consider an arbitrary solid and call it Ω. To place its points
in space, we choose an arbitrary system of reference:
with i = 1, 2, 3, that assigns to each point P in the solid its coordinates ϕ0 (P) ≡ X.
This ϕ0 function is called the reference configuration of the solid and represents, in
brief, the shape of the solid at rest, prior to the deformation.
After the application of the loads, the solid will inevitably deform (hopefully, in
an almost imperceptible manner, if we speak about structures). Therefore, it will be
necessary to define, at each instant in time, a new configuration that evolves therefore
continuously of the type
with t ∈ I ∈ R the considered time interval. As you will have noticed, we are denoting
by capital letters, i.e., X, magnitudes in the reference undeformed configuration,
while small letters like x denote magnitudes in the current updated configuration.
The situation can therefore be sketched as in Fig. 1.1. This convention is almost
universal and can be found in every book of continuum mechanics, except, probably,
the oldest ones.
It is important to know if a movement φ, the set of all configurations in time, is
regular or invertible, i.e., if ∀t, ∃φ−1
t : ϕt (Ω) → ϕ0 (Ω). This means that nothing
catastrophic has occurred, like discontinuities in the displacement field (also known
as cracks). Describing this type of phenomena is well beyond the scope of this book.
x3
x1
x2
ϕt1 (Ω)
ϕt1
φt1
X3
ϕ0
X2
Ω 1
X
ϕ0 (Ω)
ϕt2
φt2
ϕt2 (Ω)
Fig. 1.1 Time-dependent solid configurations. Do not confuse configurations (denoted usually by
the Greek letter ϕ) with movements (denoted by φ)
It is important to note, however, that not all of the structures of practical interest
can be analyzed under the assumption of small displacements. For instance,
Santiago Calatrava, the famous architect, devoted his doctoral dissertation
to the foldability of frames [4], see Fig. 1.2. The Spanish architect Emilio
Pérez Piñero also devoted the majority of his life to this type of structure. The
structural behavior of this type of frame changes drastically depending on the
configuration, from stable to a mechanism or vice versa.
4 1 An Introduction to Continuum Mechanics
Fig. 1.2 A view of the City of Sciences of Valencia, by S. Calatrava. L’hemisfèric, on the left, has
a foldable structure
∂ 2 φt
a : I × Ω → R3 , a(t, x) = (t, x).
∂t 2
However, as mentioned before, we will accept that the loading process is quasi-
static and therefore that accelerations are negligible. Phenomena like earthquakes,
for instance, are also beyond the scope of this work.
Once the movement has been defined, it is necessary to ascertain what is the difference
between a rigid solid and a deformable one. Rigid solids also move. Therefore,
1.1 Kinematics 5
∂φti ∂xi
F : I × Ω ∈ R3 → R3 × R3 , F(t, x) ≡ Fi j = (t, x) = (t, x). (1.1)
∂X j ∂X j
1 Following the previous criterion, many books employ capital letters for indices of magnitudes
defined in the reference configuration, and therefore F ≡ Fi J , i, J = 1, 2, 3.
6 1 An Introduction to Continuum Mechanics
First of all, it is of very high importance to define the displacement field, as the
distance of every point from their reference configuration to the actual one:
It is worth noting that we speak of a field since the displacement takes a different
(vector, in this case) value at each of the infinite number of points of the solid and
at each point in time, so we should employ the notation u = u(x, t), see Fig. 1.3.
Sometimes, when there is no risk of confusion, we omit the dependence on x and t.
x2
ϕt (P )
ϕt (Ω)
u(t, P )
X3
ϕ0 (P )
X2
1
X
ϕ0 (Ω)
1.1 Kinematics 7
where a comma “,” indicates a derivative with respect to the subsequent variable. We
therefore obtain a very clear interpretation: the deformation gradient is composed of:
• The contribution of a movement as rigid solid, i.e., without deformation. This part
is represented by the Kronecker-delta term δi j .
• The displacement gradient, not to be confused with the deformation gradient,
h i j = u i, j , which is a second-order tensor, too.
In turn, the displacement gradient of the solid can be decomposed as the sum of a
symmetric part and an antisymmetric one, viz.,
1 1
Fi j = δi j + u i, j = δi j + (u i, j + u j,i ) + (u i, j − u j,i ) = δi j + εi j + ωi j , (1.2)
2 2
that gives rise to the definition of the Cauchy’s strain tensor, also known as the
infinitesimal strain tensor.
1
εi j = (u i, j + u j,i ), (1.3)
2
giving rise to the so-called polar decomposition of the deformation gradient.
In vector notation, ∇u = ω + ε. It is common to denote ε = ∇ s u, to indicate
that it is the symmetric part of the displacement gradient.
Note that various strain measures are possible. For instance, the so-called Green–
Lagrange strain tensor is defined as:
1 T 1
E= (F F − I) = ∇ s u + (∇u · ∇u T ).
2 2
and ⎛ ⎞ ⎛ ⎞
ε11 ε12 ε13 εx εx y εx z
ε ≡ εi j = ⎝ ε21 ε22 ε23 ⎠ = ⎝ ε yx ε y ε yz ⎠ .
ε31 ε32 ε33 εzx εzy εz
This leads to the most frequent expression for the components of the Cauchy strain
tensor:
∂u x
εx = ,
∂x
∂u y
εy = ,
∂y
∂u z
εz = ,
∂z
1 ∂u y ∂u x
εx y = ε yx = + ,
2 ∂x ∂y
1 ∂u z ∂u x
εx z = εzx = + ,
2 ∂x ∂z
1 ∂u z ∂u y
ε yz = εzy = + .
2 ∂y ∂z
It is important to note that not every imaginable displacement field is valid from
a physical point of view (and especially under the conditions established before).
In particular, for this theory to be valid, every field defined in the solid must be
continuous and single-valued. Continuity refers to the fact that nothing catastrophic,
such as a crack, occurs. These can be imagined as the result of an interior point
splitting in two, thus giving rise to two new free surfaces. This type of phenomena
cannot be cast easily in the just introduced kinematic framework.
1.1 Kinematics 9
∂uy
uy + ∂y
dy
Y Y
∂ux
ux ux + ∂x
dx
dy dy
uy
dx X dx X
∂ux
ux + ∂y
dy
Y Y
ux
dy
dy
∂uy
uy + ∂x
dx
uy
X X
dx dx
The field must also be single-valued. Otherwise the results will be simply non-
sense: imagine for a moment a solid having, at the same time, two different displace-
ment fields. This is equivalent to obtaining a displacement field by integrating the
strains. Equivalently, observe that there are six strain-displacement relations, while
the number of components of the displacement field is three. For the resulting system
of equations to have a unique solution, it is enough to satisfy that:
R = ∇ × (∇ε) = 0,
where R is the so-called Saint Venant or compatibility tensor. Under large strain
conditions, an equivalent condition can be established for the deformation gradient,
∇ × F = 0.
1.2 Dynamics
Of course, once the description of the movement has been completed, it is important
to study the origins of these movements in terms of forces. We will consider the
possibility of volumetric forces b(x, y, z) (such as the force exerted by gravity,
inertial accelerations, …) and punctual ones, which we will call t̄. Here, the bar
indicates a prescribed known value.
10 1 An Introduction to Continuum Mechanics
Imagine for a moment that we split the solid into two parts, by cutting it through a
plane, as in Fig. 1.5. Then, in the (infinitesimal) neighborhood of every point at the
newly-created free surfaces, a force per unit area will appear. This is called the stress
vector and measured in pressure units. These forces per unit area are equilibrated
at both sides of the slicing plane, but they have, in principle, an arbitrary direction,
dependent solely on the applied forces b and t̄ (they are not, in principle, parallel nor
perpendicular to the plane, …). What is remarkable about this fact is that we can split
the solid into two pieces by employing an infinite number of planes passing through
a given point. These planes will be defined by their normal vectors, n. For each value
of n, a different distribution of stress vectors t will be obtained. It seems that the
stress is something that shows a different aspect depending on the point from which
it is viewed. We will soon demonstrate that this is precisely one of the characteristics
of a second-order tensor.
To ascertain the relationship between the chosen plane (defined by its normal, n) and
the resulting stress vector, consider Fig. 1.6. Here we see a tetrahedron whose three
faces are aligned with the coordinate axes, while the fourth is oriented in an arbitrary
direction. At every face, a stress vector appears. If we equilibrate all these vectors
along the axes’ directions, we obtain:
Among the four faces of the tetrahedron, we can establish the following relation-
ship by simply projecting onto the coordinate planes:
d Sx = d Sn n x , d S y = d Sn n y , d Sz = d Sn n z ,
such that
j
tin = ti n j .
What we observe in this equation is that a new magnitude with two indexes
j
appears, ti . We will see it in most textbooks as:
⎛ ⎞
σx τ yx τzx
σ = ⎝ τx y σ y τzy ⎠ .
τx z τ yz σz
tn
ty n
tx
tz
X
12 1 An Introduction to Continuum Mechanics
σ is precisely the Cauchy stress tensor, named after Augustin Luis Cauchy, who
first established the relationship between the orientation of the cutting plane and the
resulting stress vector:
If we come back for a moment to the case of large displacements and strains,
several stress tensors can be defined. The first possibility is to relate the force
applied in the deformed configuration per unit area in the undeformed con-
figuration. This gives rise to the so-called first Piola–Kirchhoff (or nominal)
stress tensor, P = J σ F −T , where J = det F is the volume change during
deformation.
The second Piola–Kirchhoff stress tensor is defined as S = J F −1 σ F −T .
Note that while P is not symmetric, S is.
To clearly characterize the stress state of a point, it is very useful to look for some
direction n in which only normal stresses appear (those in the diagonal of the tensor).
This is equivalent to the classical procedure of diagonalizing a matrix. To this end,
consider the eigenvalue problem:
σi j − σδi j = 0,
which can easily be solved through the solution of a cubic equation expressed in terms
of the three invariants of the stress tensor (values that do not change, regardless of
the chosen normal n, thus representing a valid measure of the stress state in the solid)
σ 3 − I1 σ 2 + I2 σ − I3 = 0,
with
I1 = σ x + σ y + σ z ,
σ τ σ τ σ τ
I2 = y zy + x zx + x yx ,
τ yz σz τ x z σz τx y σ y
and
σx τ yx τzx
I3 = τx y σ y τzy .
τx z τ yz σz
1.2 Dynamics 13
To complete the picture, we need to identify what is the relationship between the
applied loads and the stresses developed by the solid. This relationship comes from
equilibrium. Consider the elementary volume of Fig. 1.7. By the simple application
of equilibrium of forces along each coordinate axes—remember the force per unit
area units of stresses—we have:
∂σx ∂τ yx
(σx + d x − σx )dydz + (τ yx + dy − τ yx )d xdz
∂x ∂y
∂τzx
+ (τzx + dz − τzx )d xd y + bx ρ d xd ydz − ρ d xd ydz ax = 0,
∂z
with ρ the specific weight of the solid and ax the acceleration along x-axis. After
discarding high-order terms, we have:
∂σx ∂τ yx ∂τx x
+ + + ρbx = ρax . (1.4)
∂x ∂y ∂z
σz + dσz
dx τzy + dτzy
τzx + dτzx σx
τxy
τyx τyz + dτyz
σy τxz
τxz + dτxz σy + dσy
τyx + dτyx
dz τyz τxy + dτxy
σx + dσx
τzx Y
τzy
σz
X
dy
A similar equation can be obtained by applying equilibrium along each one of the
coordinate axes.
The same can be done by applying equilibrium of moment with respect to an
arbitrary point. However, this will lead simply to:
σi j = σ ji , (1.5)
or, in other words, to the symmetry of the stress tensor. Remember that this does not
hold for the first Piola–Kirchhoff stress tensor, for instance, and is only valid under
small strain and displacement assumptions.
Equation (1.4) can be written alternatively as:
∇ · σ + ρb = ρa.
So far, we have found what are the equations governing the relationship between
displacements and strains (compatibility equations) and equations governing the
relationship between external loads and stresses (equilibrium equations). To close
the circle, the reader will readily notice that we need to find the relationship between
strains and stresses. That is what we call constitutive equations. This relationship
depends on the particular material of interest and must be obtained experimentally.
A general form for this equation would be:
i.e., any suitable form that relates stresses with the movement experienced by the
solid, possibly time dependence, temperature, time and, of course, the particular
material point under consideration, X, since the material may be nonhomogeneous.
We emphasize that the behavior of the material can be history-dependent, and that is
why we have included the dependence on the history of the movement, φt0 →t . This
is the typical case of plasticity, for instance. If this behavior is not history-dependent,
it can take a simpler form of the type:
which explicitly draws our attention to the fact that the behavior depends on strain,
but also on strain rate, that is, the velocity at which we impose deformations. This
1.3 Constitutive Equations 15
is the typical example of viscous materials. Simplifying further, the material can be
simply elastic, in which case its strain state depends solely on the strain state:
P = P(F, t),
i.e., the stress state (here represented by the first Piola–Kirchhoff stress tensor)
depends solely on the strain state, here represented by the deformation gradient
F.
where values affected by the superscript (·)0 indicate some nonnull, reference value
for the variable of interest. Here, it is important to note that, due to the fact that
both strain and stress are second-order tensors, their constitutive relationship exists
in the form of a fourth-order tensor, C ≡ Ci jkl . Note that we employ, as in many
specialized books in the field, sans serif fonts for fourth-order tensors, to distinguish
from boldface second-order tensors.
Equation (1.7) receives the name Duhamel–Neumann’s law and first established
the origin of thermoelastic stresses in solids.
It is important to note that, for elastic solids, there is a strain density functional
from which we can derive the constitutive law. The existence of such a functional
guarantees that the stress-strain relationship does not depend on the history of the
movement. In this case, this functional takes the form:
1 1
V − V0 = Ci jkl (εi j − εi0j )(εkl − ε0kl ) − γi j (Θ − Θ 0 )(εi j − εi0j ) + (σi j )0 (εi j − εi0j ),
2 2
such that
∂V
σi j = .
∂εi j
16 1 An Introduction to Continuum Mechanics
by simply neglecting thermal dependence, if the problem can be cast under isothermal
conditions, and also neglecting reference values.
In vector notation,
σ = C : ε,
where “:” indicates the twice-contracted tensor product, i.e., a product that, applied
to matrices, generates a scalar, and applied to this fourth-order times second-order
tensor product, generates a second-order tensor.
In the case of large strain settings, the so-called hyperelastic materials are such
that:
∂V
S= .
∂E
In this case, the second Piola–Kirchhoff stress tensor and the Green–Lagrange
strain tensor E are said to be conjugate.
1.3.2 Symmetries
∂2 V ∂2 V
= ,
∂εi j ∂εkl ∂εkl ∂εi j
or, equivalently:
Ci jkl = Ckli j ,
Hooke was the first to establish the linear relationship between strain and stress, so
this very simple constitutive equation was named after him. Its most frequent form
is:
σi j = λθδi j + 2Gεi j , (1.9)
with θ = εii = εx + ε y + εz . Here, λ and G are the two Lame’s coefficients, defined
as a function of the Young’s modulus E and Poisson coefficient ν as
Eν E
λ= , G= .
(1 + ν)(1 − 2ν) 2(1 + ν)
Every constitutive law has its limits. For instance, no existing material presents linear
behavior independent of the level of stress. Some of them plastify, which means that
they acquire permanent deformations, or simply break in a brittle manner. However,
it is not evident how to know the precise stress state under which this will occur. Note
that stresses are indeed a tensorial magnitude, so several questions arise: do all six
different stress tensor components contribute equally to the failure of the material?
Does this occur at the same value?
Of course, the answers are no and no. The only way to ascertain this is by means
of costly experimental tests. For instance, if we subject a steel bar to a traction
test, it shows a load-displacement curve shown in Fig. 1.8. We first notice a clearly
linear part of the curve that, at some 15 kN, begins to loose linearity and, for small
increments in load, experiences large increments in displacement. This is the plastic
regime of the curve. The last part of the curve (the vertical tray at the end) represents
the register of the machine when the material is already broken and does no longer
provides any useful information.
For a single stress state it is therefore easy to determine the precise value at which
the constitutive equations is no longer valid. But what happens if we experience a
complex non-homogeneous stress state? Many researchers (Tresca, Rankine, et al.)
devoted efforts to try to understand this phenomenon and to supply a practical answer
to professionals.
Among their findings, we can mention:
• They found that the limits of Hooke’s law do not depend on hydrostatic pressure,
i.e., that pieces subject to stress states, such that the principal stresses σ I = σ I I =
σ I I I are all equal, do not produce any yielding.
18 1 An Introduction to Continuum Mechanics
Applied force, kN
10
• They also found that plastification depends solely on the second invariant of the
deviatoric stress tensor, in other words, on:
(σ I − σ I I )2 + (σ I I − σ I I I )2 + (σ I I I − σ I )2
J2 = .
6
Von Mises,2 who in 1913 established a failure criterion that has been widely
accepted since then for metals (particularly, steel and aluminium). This criterion
states that, for a complex stress state, the material fails when:
(σ I − σ I I )2 + (σ I I − σ I I I )2 + (σ I I I − σ I )2 = 2σe2 ,
where σe is the uniaxial stress at which the material also fails. The interest relies
precisely on the ease of determining this value with a single laboratory experiment.
In this way, Von Mises established a stress indicator to compare with the yield
stress σe arising from the lab tests. This stress, in an arbitrary coordinate system,
reads:
1
σVM
2
= (σ11 − σ22 )2 + (σ22 − σ33 )2 + (σ33 − σ11 )2 + 6(σ23
2
+ σ312
+ σ122
) ,
2
(1.10)
so that we simply need to verify that, under our design conditions, it holds that:
σVM ≤ σe ,
2 Do not confuse with Ludwig Heinrich Edler von Mises, the famous Austrian economist advocate
of the free market, who was actually his brother. We refer to Richard Edler von Mises (1883–1953).
1.4 Navier Equations for the Elasticity Problem 19
So far, we have explored all the equations arising in the definition of the elastic
problem, even their limits of validity. We have described the kinematic equations,
the dynamics and the constitutive equations. And, somehow inherently, we have
accepted that the load variables, b and t̄, are the data in the problem. However, we
have not even defined the unknowns of the problem.
One possibility is to decide that the stresses are the most important unknowns in
the problem (in view of the just introduced von Mises criterion) and to substitute
them in the kinematic equations with the help of constitutive equations. This leads
to the so-called Beltrami–Michell3 form of the elasticity problem. This is however
of little practical interest, in general, even though, at the end of this book, we will
develop a method for the practical analysis of structures based on this approach.
The most frequent form of the elastic problem is the so-called Navier form. In it,
we decide to choose the displacements as the unknowns of the problem. Therefore,
if we substitute the displacement-strain Eq. (1.3) into the constitutive Eq. (1.8) and
then into the equilibrium Eq. (1.6), we arrive at:
1
Ci jkl (u k,l + u l,k ) , j + ρbi = ρai in Ω,
2
u i = ū i on δΩiu ,
1
Ci jkl (u k,l + u l,k )n j = t¯i on δΩit ,
2
u i (x, 0) = ū i (x)
on Ω,
u̇ i (x, 0) = v̄i (x)
where all magnitudes affected by a bar indicate the prescribed values (boundary
and initial values for the problem). If the considered material is homogeneous (its
constitutive tensor is constant everywhere in Ω), we have:
1
Ci jkl (u k,l j + u l,k j ) + ρbi = ρai ,
2
and if it were, in addition, isotropic:
3 You can find a very interesting NASA report on the topic at https://ntrs.nasa.gov/archive/nasa/
casi.ntrs.nasa.gov/19950012494.pdf.
20 1 An Introduction to Continuum Mechanics
which is precisely the expression of the Navier form of the elastic problem. In vector
notation:
(λ + G)∇(∇ · u) + G∇ 2 u + ρb = ρa. (1.11)
This resulting equation is complex. Except in some very simple and academic
cases, no solution is known for it, even if existence and uniqueness theorems exist
that prove the validity of the approach followed therein. Nowadays, powerful compu-
tational methods, such a the finite element method, exist that are based on somewhat
similar forms of the problem. However, architects and engineers have been forced
historically to deal with these complex equations to design buildings. Their approach
is, however, a different one.
Traditionally, engineers have tried to simplify to some extent the Navier equations.
This has been done by accepting some simplifying assumptions and, at the same time,
working in a somewhat different manner. This gave rise to what we know as strength
of materials or mechanics of materials. Surprisingly, strength of materials has nothing
to do with the strength of the materials! It is a technological discipline composed by
a collection of hypotheses that allows one to simplify the elastic problem so as to be
solved (at least, approximately) easily and fast. To be employed in architectural and
engineering practice, in sum.
Very often, strength of materials is studied before the theory of elasticity, and
this hides the true hypothesis behind the resulting formulas, making it a sometimes
awkward collection of technical recipes with no apparent connection among them.
This is why we decided to begin this book by a (fast, very fast) overview of the
theory of elasticity, that will enable us to understand all the mechanisms behind the
well-known formulas and will allow us to dimension our structural members. This
fascinating journey begins in the next chapter.
1.4 Navier Equations for the Elasticity Problem 21
Worked Examples
(a) Compute the deformation gradient, its inverse and its transpose.
(b) Obtain the deformation gradient as a function of the displacement field.
(c) For this movement, and taking into account the small displacements and small
strains hypothesis, determine the strain and rotation tensors.
(d) Particularizing the strain tensor expression for the point with Cartesian coordi-
nates P = [0, 0, 0]T , compute the principal strains and their associated principal
directions.
(e) Demonstrate that the volume increment is independent of the coordinate system
used to express the strain tensor.
(f) Given the value of the Young’s modulus E = 2.1 · 106 and Poisson’s coefficient
ν = 0.3, compute the Cauchy stress tensor σ(P) at point P = [0, 0, 0]T .
(g) Compute the principal stresses and their associated principal directions at P.
Solution 1.1. (a) Applying the definitions introduced in this chapter, it is clear that:
⎛ ⎞
2X 0 −1
∂xi
Fi j (X, Y, Z ) = = ⎝ 0 3Y 2 0 ⎠ .
∂X j 2X 2Y 0
so that:
⎛ ⎞
0 −1√ √ 1
√
⎜ 3 y+2 z− (y+2) 2 z− (y+2) ⎟
3 3 2 3 2
∂X j ⎜ ⎟
(F −1
)i j (x, y, z) = =⎜
⎜0
√1 0 ⎟
⎟
∂xi ⎝
3 3 (y+2)2
⎠
−1 √−2 1
3 3 y+22
22 1 An Introduction to Continuum Mechanics
(b) We assume the origin of the coordinate system to be located at the point (0, 0, 0).
By taking into account that the reference and deformed configurations use Carte-
sian coordinate systems, we have
⎛ ⎞
X2 − Z − X
u = ⎝ Y3 − 2 − Y ⎠.
X2 + Y 2 − Z
so that: ⎛⎞
2X 0 −1
Fi j = δi j + h i j = ⎝ 0 3Y 2 0 ⎠ ,
2X 2Y 0
(d) The strain tensor particularized at point [0, 0, 0]T is given by:
⎛ ⎞
−1 0 − 21
ε(P) = ⎝ 0 −1 0 ⎠ ,
− 21 0 −1
−1 −3
εI = , ε I I = −1, ε I I I = .
2 2
To determine the principal directions, i.e., the eigenvectors, we must solve the
following system of equations:
(εi j − εδi j )n i = 0
n 21 + n 22 + n 23 = 1,
24 1 An Introduction to Continuum Mechanics
i.e., the mandatory unity norm of each eigenvector. This gives rise to the following
diagonalized strain tensor:
⎛ 1 ⎞
−2 0 0
εprinc = ⎝ 0 −1 0 ⎠ ,
0 0 − 23
T T
where n I = √12 , 0, − √12 , n I I = (0, 1, 0)T and n I I I = √12 , 0, √12 .
Finally, it is easy to see that:
(e) The volume increment per unit volume or, equivalently, the trace of the principal
strain tensor is:
ΔV 1 3
= ε I + ε I I + ε I I I = − − 1 − = −3,
V 2 2
that is the same result obtained in a different system of reference, viz.,
ΔV
= ε X + εY + ε Z = −1 − 1 − 1 = −3.
V
Obviously, this same result can be obtained in any system of reference. This is
the meaning behind the name invariants of the strain tensor.
(f) Applying the Lamé constitutive equations, we can find the Cauchy stress ten-
sor expression for the point P, where we know the strain tensor ε. The Lamé
constitutive equation is:
where
Eν
λ=
(1 + ν)(1 − 2ν)
and
E
2G =
1+ν
and
2G = 1.615 · 106 kp/cm2 .
(g) The stress tensor at [0, 0, 0]T has been obtained in (f):
⎛ ⎞
−5251 0 −807.5
σ(P) = ⎝ 0 −5251 0 ⎠ · 103 kp/cm2 .
−807.5 0 −5251
Solving the following system of equations for each principal strain, we obtain
the principal directions, i.e., the eigenvectors of the system:
(σi j − σδi j )n i = 0,
where the three components of each principal direction must satisfy again the
unity norm condition of every eigenvector
n 21 + n 22 + n 23 = 1.
Therefore:
⎛ ⎞
−4.4435 0 0
σ princ =⎝ 0 −5.251 0 ⎠ · 106 kp/cm2 ,
0 0 −6.0585
26 1 An Introduction to Continuum Mechanics
T T
where n I = √12 , 0, − √12 , n I I = (0, 1, 0)T and n I I I = √12 , 0, √12 . The
same directions that the same obtained for the principal strains. This follows
from the linearity of constitutive equations.
1.2. Two blocks, made of different materials but with the same dimensions 50 ×
50 × 50 cm3 , are placed between two infinitely rigid plates. These materials have
elastic, linear and isotropic behaviors with various Young’s moduli and Poisson
coefficients. Determine the minimum value of the forces P1 and P2 that allow the
top plate to remain horizontal, and allow, at the same time, the blocks to come into
contact. We assume that the bottom plate is fixed.
The initial distance between blocks is d = 1 cm, and there is no restriction for the
movement along the perpendicular direction to the plane of the problem. Neglect the
friction between various surfaces, and assume no temperature variations. Assuming
a uniform stress state for each block (produced by the parallel descent of the top
plate and the absence of friction), determine the:
(a) Condition to be fulfilled by both normal strains along x so as to produce contact
between the blocks.
(b) Condition to be fulfilled by both normal strains along y that allow the top plate
to remain horizontal.
(c) Stress and strain tensors for each block.
(d) P1 and P2 that make the contact between blocks possible, while keeping the top
plate horizontal (Fig. 1.9).
Data: E 1 = 1.1E6 kp/cm2 , ν1 = 0.34, E 2 = 7E5 kp/cm2 , ν2 = 0.33 and d =
1 cm.
50 cm 50 cm 1 cm 50 cm 50 cm
P1 P2
y
50 cm
x E1 , ν1 E2 , ν2
Fig. 1.9 Sketch for Example 2. Dashed lines reflect the final, deformed configuration of the top
plate and the two blocks
1.4 Navier Equations for the Elasticity Problem 27
Solution 1.2. (a) Due to the descent of the top plate that remains always horizontal,
each block will increase its horizontal dimension by magnitude ΔL (i)
x , i = 1, 2.
This magnitude depends, obviously, on the particular constitutive material of
each block. Therefore, for them to come into contact, the necessary condition
will be:
ΔL (1) ΔL (2)
x
+ x
= d = 1 cm.
2 2
In the development of this condition, we have assumed that the center of gravity
of each block remains at its original position.
If we elaborate a little bit, the just obtained expression, by expressing the length
increments in terms of the strain at each block, we have
ε(1)
x · 50 ε(2) · 50
+ x =1
2 2
so that
2
ε(1) (2)
x + εx = = 0.04,
50
which is precisely the condition sought to be satisfied by both normal strains.
(b) If the upper plate remains horizontal, it is clear that we must satisfy ΔL (1)
y =
(2)
ΔL y .
In others words, since L (1) (2)
y = L y , it must hold
ε(1) (2)
y = εy ,
• Block 2: ⎛ ⎞
σx(2) 0 0
σ 2 = ⎝ 0 σ (2)
y 0 ⎠
0 0 σz(2)
28 1 An Introduction to Continuum Mechanics
⎛ (2) ⎞
εx 0 0
ε2 = ⎝ 0 ε(2)
y 0 ⎠
0 0 ε(2)
z
In x and z directions, there is no external force applied to any block. This implies
that:
σ 1 · n = t = 0.
We thus arrive to a much more detailed expression for the stress tensors, while
the strain tensors remain unknown for the moment:
⎛ ⎞
0 0 0
σ 1 = ⎝0 σ (1) ⎠
y 0 ,
0 0 0
and ⎛ (1) ⎞
εx 0 0
ε1 = ⎝ 0 ε(1)
y 0 ⎠,
0 0 ε(1)
z
and ⎛ (2) ⎞
εx 0 0
ε2 = ⎝ 0 ε(2)
y 0 ⎠.
0 0 ε(2)
z
ε(1) (2)
x + εx = 0, 04,
1.4 Navier Equations for the Elasticity Problem 29
and
ε(1) (2)
y = εy ,
and
σ (2)
ε(2)
y
y = .
E2
E 1 (2)
σ (1)
y = σ .
E2 y
and ν2 (2)
ε(2)
x =− σ = ε(2)
z .
E2 y
ν1 (1) ν2 (2)
− σ − σ = 0.04.
E1 y E2 y
0, 04 · E 2
σ (2)
y =− = −41791.04 kp/cm2 .
ν 1 + ν2
Finally,
E 1 (2)
σ (1)
y = σ = −65671.64 kp/cm2 .
E2 y
30 1 An Introduction to Continuum Mechanics
(d) To find the value of forces P1 and P2 that allow the blocks to come into contact
while keeping the top plate horizontal, we impose an equilibrium of forces along
the vertical direction first, see Fig. 1.10,
Fy = 0,
so that
P1 + P2 = σ (1) (2)
y · 50 · 50 + σ y · 50 · 50.
50 cm 25 cm 25 cm 1 cm 25 cm 25 cm 50 cm
P1 P2
(1) (2)
Fy Fy
Fig. 1.10 Equilibrium of the top plate under the action of the two forces and the reactions coming
from the two blocks
1.4 Navier Equations for the Elasticity Problem 31
This provides us with the first equation that will allow determining the value of
both forces:
P1 + P2 = 268656716.4 kp.
To impose the horizontal position of the top plate in the deformed configuration,
we impose the equilibrium of moments, so that no resulting rotation appears,
Mz (Q) = 0,
where Q is arbitrarily chosen as the point where P1 is applied (to make it disappear
from the second resulting equation and make the final solution simpler—other choices
will obviously produce the same result):
Abstract This chapter introduces the basic aspects underlying the structures made
of cables.
2.1 Introduction
In the first chapter, we have just seen that the elastic problem is complex. So complex
that we still do not know the solution for a general problem, but only for simple
cases of little practical interest. The discipline we are going to begin to study takes
a different route. As will readily be seen, a new magnitude is to be defined that will
greatly help to alleviate our burden. This is the concept of internal force.
Cables are usually represented as a straight or curved line, depending on the type
of load they support, see Fig. 2.1, with hinges at each end (they are often supported
so that rotations are allowed, see Fig. 2.2, although the true reasons will become
clear when we study beams). This is also typical of this discipline: structures are
represented by a simplified diagram, by employing signs whose meaning will be
explained throughout this book. From now on, a cable will be a thin line, given its
relative slenderness. Even if they are actually thick, such as, for instance, cables in
bridges, see Fig. 2.3, its area to length ratio is always very low.
Solving the Navier Equation (1.11) for cable structures is indeed very difficult. Even
direct numerical simulation with the most recent codes on the market is not self-
evident. However, cable structures are in general easy to understand under some
classical hypotheses:
1. Cables do not support any form of bending, due to their slenderness. Therefore,
under the action of a load such as P, see Fig. 2.1, they acquire an equilibrium
configuration that depends on the type of load they are supporting. This is in
where A is the area of the cross section of the cable and is assumed to be constant.
N (x) is the axial internal force and can depend, in its most general case, on the
x, axial, coordinate of the cable.
Once the axial force has been determined, every magnitude can thus be obtained
by simply applying the simplified form of the constitutive and kinematic equations
seen in the previous chapter:
σ(x) N (x)
εx (x) = = , (2.2)
E AE
where εx is the only non-vanishing component of the strain tensor, and E is the
Young’s modulus of the material under consideration. Once the strain has been deter-
mined, we can obtain the displacement at any point by just integrating the kinematic
equations, viz.,
x=l x=l
N (x)
Δl = εx (x)d x = d x.
x=0 x=0 AE
If, as usual, both the area of the cross section and the Young’s modulus are constant,
we obtain: x=l
1
Δl = N (x)d x. (2.3)
AE x=0
2.2 Main Hypotheses 35
Fig. 2.2 Support apparatus for cable structures often allow free rotation
What we have done in a very subtle way is the usual form of work in this
discipline. Observe that we have defined a new magnitude: the internal force.
This may seem a waste of time, since the problem is already complex enough.
However, throughout this book (and, in general, any book devoted to strength
of materials), we will operate in this same manner: from the geometry and
loads of the structure, we will compute the distribution of internal forces and,
from them, we will compute stresses, then strains and, finally, displacements.
This can seem a weird form of working, but it is, in general, much simpler
than the integration of Eq. (1.11).
36 2 Cable Structures
Fig. 2.3 Pont de l’Assut de l’Or, Valencia, Spain (S. Calatrava), a structure that employs cables
The true advantage of this way of working is that internal forces can be determined
without solving the Navier equations. In fact, there are some simple cases—called
isostatic—in which simple force-equilibrium considerations will suffice to obtain
the internal forces acting on our structure. Structures formed by one single cable will
always be isostatic. In the example of Fig. 2.1, imagine for a second that we cut the
cable, as we did in the previous chapter, isolating a region around the applied load
P, see Fig. 2.4.
Since the internal forces must follow the geometry of the cable, by hypothesis 1
above, we know in advance the direction of the loads and can establish the equilibrium
2.3 Equilibrium in Cables 37
N1 N2
P
N1
P
N2
Fig. 2.4 Equilibrium of the load in the structure of Fig. 2.1. This is the so-called free body diagram.
On the right is the closed-force polygon ensuring equilibrium. To obtain the force polygon, just
draw every force one after the other, until a closed polygon appears. The particular ordering is not
relevant for the result
N1
R1
N1
N1
N1
N2
P
N1
The so-called Cremona diagram of Fig. 2.4 allows us to analyze the influence of the
rise in the resulting internal forces, see Fig. 2.7. Thus, changing the rise from f to
f /2, produces an increase in the internal forces.
N1 N1
f /2
P
f /2
N2 N2
Fig. 2.7 Influence of the rise in the resulting internal forces in the cable. The new Cremona diagram
is depicted in dashed lines. Notice the increase in the traction of both cables, N1 and N2
P
N1
N2
P P N2
N1
f
N3
N2 P
N3
advantage of working with internal forces is precisely the possibility of, by applying
simple equilibrium equations, obtaining every magnitude in the cables. For instance,
consider the case with two loads, see Fig. 2.8.
Here, we notice that the Cremona diagram introduces internal force N2 twice. In
fact, since equilibrium is applied at the nodes (those where hinges have been placed),
N2 appears twice: once for the node at the left, once for the node at right. Of course,
both equilibrate (in fact, N2 is an internal force) so that only externally applied forces
from the diagram (i.e., applied loads plus reactions at supports) are not cancelled in
the Cremona diagram.
40 2 Cable Structures
We now consider the case of a cable loaded with two loads acting along arbitrary
directions, see Fig. 2.9. The case can be further generalized in a straightforward
manner, see [16].
First, consider two loads of 10 and 15N , respectively. The procedure begins by
computing the resultant force, see Fig. 2.9, left. For the design of the structure, we
can choose either the slope of the cable or the rise, f¯. We put a bar on the f , since
this will not be the final rise of the cable. In this case, we fix f¯, although the case of
choosing the slope at the hinges is identical. Once chosen, this allows us to determine
the slope of the cable near the anchoring points, see the dashed line.
The intermediate part of the cable will have a slope that we can find easily from
the Cremona diagram, by just joining the intermediate points, between the end of N1
and the end of the 10N force, see the red arrows N2 . The resultant shape of the cable
10N
f¯
N3
10N N2
R
15N
R
15N
N1
Fig. 2.9 A structure with two arbitrary loads. Concept of resultant cable
10N
15N
2.3 Equilibrium in Cables 41
is shown in Fig. 2.10. Notice the slope of the intermediate section, parallel to the red
force N2 . What we have obtained is actually the funicular polygon of the loads.
When there are three or more loads acting along arbitrary directions, of course,
they need not cross at a single point. In that case, the previous procedure can be
generalized by computing the resultant of the two first loads, and then computing
the resultant between the one just computed and the third load, and so on.
The case of a cable subjected to a distributed load is even more interesting, although
it can be considered as a generalization of the previous cases, see Fig. 2.11. This can
be seen as a particular case in which a series of infinitesimal loads are applied to
the cable. Therefore, instead of a sequence of straight lines, the configuration of the
cable will be a curve, and hence the name of funicular curve rather than polygon.
In this case, the procedure is identical to the previous section. First, compute the
resultant force but note that this type of load acts per unit length, i.e., their units will
be force per unit length (usually, kN/m) in horizontal projection, not per unit cable
length. The resultant, see Fig. 2.12, will have a value q · and will be applied at the
center of gravity of the load distribution (imagine the case of a triangular distribution,
for instance). In this case, the center of gravity coincides with the middle of the span,
since the distribution is uniform. The naive approach to the problem will thus be to
divide the distributed load into small, concentrated loads that will “discretize” the
curve, but will render a close approximation to it.
The resulting funicular curve for a cable under constant, distributed load is a
parabola. This can be easily demonstrated. Let us imagine a cable portion of elemen-
tary length ds, see Fig. 2.13.
By applying equilibrium along horizontal and vertical directions, respectively, we
find,
q
42 2 Cable Structures
N1
q·
N2
q·
ds N + dN
dy N
dp
dx
x
where θ and θ + dθ represent, respectively, the angle with the horizontal where n
and N + d N are applied, respectively (not included in Fig. 2.13 for clarity). This is
equivalent to
From Eq. (2.4), we notice that the horizontal projection of N is actually constant over
the cable, i.e., N cos θ = H = const. In turn, from Eq. (2.5), we obtain, by noting
that tan θ = y ,
d(N sin θ) = d(H tan θ) = H dy = dp,
or, equivalently,
1
dy = dp,
H
from where,
1 dp
y = ,
H dx
which is the differential equation of the so-called funicular curve.
Note that, by virtue of Eq. (2.4), if the horizontal projection of the internal
force in the cable, H , is constant, the steeper you design your cable, the greater
internal forces it will suffer.
y = y(x, C1 , C2 ),
with C1 and C2 as two constants whose precise value is to be found with the help
of the attachment conditions. If the curve is subjected to a distributed load, then
dp = q · ds, and by integrating the funicular curve,
q
dy = d x,
H
which gives, in turn,
q 2
y= x + C1 x + C2 ,
2H
which is in fact the equation of a parabola. For the coordinate axis defined in Fig. 2.14,
we have,
q 2
y= x −
2
. (2.6)
2H 4
This is the curve being described by suspension bridges, such as the Golden Gate
bridge, see Fig. 2.15, for instance. In this case, if we neglect the self-weight load, the
main forces come from the hangers, vertical cables hanging from the load-bearing
cable.
44 2 Cable Structures
y
Fig. 2.15 Golden gate bridge in San Francisco, U.S.A., a typical example of a suspension bridge
Of particular importance, for reasons that will become evident in the next chapter,
is the funicular curve resulting when the cable is subject to a load proportional to its
length (like its own weight, for instance, even if we usually neglect it).
In that case, dp = q ds instead of dp = q d x, as in the previous section. Therefore,
dp = q ds = q 1 + y 2 d x,
2.5 Catenary Curve 45
dy q
= d x,
1+ y 2 H
H q
y= cosh x + C1 + C2 .
q H
This curve is known as a catenary, and is the curve described by cables along railways,
for instance.
The difference between the parabola of Eq. (2.6) and this catenary curve is
subtle. Both curves are very similar in shape, although the catenary has its
center of gravity slightly lower.
Cables, as for any structure, are subject to limitations in both their ultimate limit state
(i.e., capable of producing a catastrophic failure) and serviceability-limit state (i.e.,
capable of making the structure deflect or vibrate up to a limit in which it looses its
practical interest). For the first case, it suffices to ensure that the stress in the cable,
σ, is lower than the yield stress of the material, σe , sometimes also denoted as σ y . To
avoid large deformations, it is a practical criterion, see [16], to limit the stress due to
permanent loads to a value
σ ≤ 0.45 σe .
As seen before, the internal force, and therefore the stress, is higher where the
slope of the cable is maximal. This occurs, generally, near the anchoring points. It is
precisely there where we must ensure the safety of the structure.
46 2 Cable Structures
It is important to note, however, that even if the internal force is maximal where
the slope is higher, a shallow cable will lead, in general, to higher internal forces.
Recall Fig. 2.7, for instance, where a decrease of the slenderness ratio, / f , produced
an important increase in the internal force, as can be readily noticed in the Cremona
diagram of the structure.
If the displacement is judged to be too large, and the span cannot be reduced, then,
the solution comes from reducing the slenderness ratio by augmenting the rise f , or
by reducing the displacement by augmenting the area of the cable, recall Eq. (2.3).
It must be highlighted, however, that in general one single cable is not able to
provide with enough stiffness to the structure, so that a change in the variable
load does not produce a large displacement. Variable loads will always pro-
duce large changes in the shape of the cable. See [16] for some practical
examples.
In order to avoid such drastic changes in shape, other solutions must be taken into
account:
1. One simple, but very practical, solution is to increase the weight of the structure, to
make the relative importance of variable loads smaller. This naive or brute-force
solution is indeed very practical and has been applied successfully frequently,
especially in the construction of bridges.
1 Seehow BBC covered the re-opening of the bridge after costly modifications in it: http://news.
bbc.co.uk/2/hi/uk_news/england/1829053.stm.
2.7 Deflections in the Structure 47
Fig. 2.17 Example of a structure made of a sequence of cable beams. This sequence gives rise to
a sort of cable membrane. José de Yarza, architect. Zaragoza, Spain
48 2 Cable Structures
Fig. 2.18 Cable stiffening of the glass facade at Gare de Montparnasse, Paris, France. Baudoin,
Cassan de Marien, López and Saubot, architects
Due to the lightness of cable structures, the Irish engineer P. Rice began in the
1980s to employ cables to stiffen glass structures. This practice is nowadays very
common in many airports, railways stations and any other public building with large
glass facades, see Figs. 2.18 and 2.19.
2.7 Deflections in the Structure 49
Worked Examples
2.1 Given the loads applied to the cable structure represented in Fig. 2.20, compute
the reaction forces, the total cable length and the final height of point C, h C .
2.5 m 5m 2.54 m
1 kN
1.5 m
A
1.875 m
hC
2 kN
D
B
Solution 2.1 To obtain the reaction forces at A and D, it is necessary verify global
equilibrium on the cable structure. Therefore, the sum of the vertical reactions in A
and D are equal to 1 kN + 2 kN = 3 kN, so that
V A + VD = 3 kN.
H A = HD ,
M(D) = 2 × 2.5 + 1 × 7.5 + H A × 1.5 − V A × 10 = 0.
The system has three equations with four unknown variables (V A , VD , H A and
H D ), so we need another equation. This will be obtained taking a look to Fig. 2.22,
where the slope of the cable is known. If, in a cable, there is no other force than an
axial internal force, it must act along the axis of the cable, thus giving a relationship
between V A and H A .
The relative position between points A and B is known, so
VA 1.875
= .
HA 2.5
2.7 Deflections in the Structure 51
HA
1kN
VA HD
2kN
VD
RA
NAB
This is precisely the fourth equation needed to solve the equilibrium system of
equations.
After solving, we find H A = 2.08 kN, V A = 1.56 kN, H D = 2.08 kN and VD =
1.43 kN.
The modulus of the reaction at support A will obviously be R A = H A2 + V A2 =
2.60 kN = N AB , identical to the internal force in the first section of the cable, AB.
NAB 1kN
α
β
NBC
N BC · cos β = N AB · cos α = H A ,
5
L BC = = 5.17 m.
cos β
L = L AB + L BC + L C D
= 1.8752 + 2.52 + 5.17 + (3.225 − 1.5)2 + 2.52 = 11.34 m.
2.2 If the rise of the cable in Fig. 2.24 is 9 m, compute, for a uniform load q =
300 kp/m,
(a) Equation of the parabola adopted by the cable.
(b) Maximum and minimum values of the axial internal force in the cable.
Solution 2.2 (a) We begin by imposing equilibrium for a portion of the cable around
the lowest point of the cable, which we will denote hereafter O, see Fig. 2.25.
The moment equilibrium gives:
x
M(C) = 0; q · x · − NO · h = 0
2
2
so that h = 2N
qx
O
, that is a te equation of a parabola with vertex at O, precisely the
geometry adopted by the cable under the load q.
The horizontal force equilibrium gives:
Fx = 0; NC · cos θC = N O .
2.7 Deflections in the Structure 53
90 m
y
9m
kp
q = 300 m
NC
y θC
C(x, y)
h
O
NO
x
q = 300 kp/m
Fig. 2.25 Equilibrium at a portion of the cable from the bottom point
This expression indicate that the horizontal component of the axial force is the same
for all points of the cable—something that we have just introduced in previous
sections. By imposing vertical equilibrium,
Fy = 0; NC · sin θC = q · x.
NC = N O2 + q 2 x 2 ,
where NC is the axial force at each point of the cable, irrespective of its position.
54 2 Cable Structures
Since the supports are placed at the same height above the lowest point of the
cable, the bottom point of the cable will be in the middle of the span, and the rise is
therefore known: h = 9 m, so that
q · x2
y=h= = 0.004 · x 2
2 · NO
The minimum axial force will take place at the lowest point of the cable,
Abstract This chapter introduces the basic aspects of structures composed of arches.
3.1 Introduction
Soon after human beings began to deal with distances greater than those that can
be spanned by a tree trunk, the major problem of architecture saw the light. The
materials at hand were of a very particular nature. For instance, stone blocks can
resist (very well, in fact) compressive stresses, but without any type of mortar, they
simply detach from each other when subject to traction. Granite, for instance, resists
up to 180 MPa [12], but granite blocks simply do not resist any load if subject to
traction, or a loss of stability, leading to a catastrophic failure under complex load
states.
Therefore, the problem of architecture soon became that of designing stable struc-
tures, rather than that of designing resistant structures. Actual stresses in stone build-
ings are so low that resistance was not actually the problem. We are not aware of any
building that collapsed due to excessive stress, but many (thousands, probably; we
will never know) failed due to lack of stability.
In this chapter, we will study how to solve this problem, which was of primary
importance until the 19th century, when the Industrial Revolution introduced iron
and the steel as building materials. For the first time, steel as a building material had
the same characteristics in traction and compression, thus changing completely the
way of designing structures.
Without aiming to be very rigorous, we can say that the Romans, for whatever
reason, decided that the “perfect” shape for an arch was a semicircle, see Fig. 3.1.
© Springer International Publishing AG, part of Springer Nature 2018 55
E. Cueto and D. González, An Introduction to Structural Mechanics for Architects,
Structural Integrity 4, https://doi.org/10.1007/978-3-319-72935-0_3
56 3 Arches
They mastered arches so well that could build the Pantheon in Rome from 27 B.C.
to 14 A.D., with a diameter of 43 m, and a perfectly spherical shape.
Later, Romanesque architecture also employed semicircular arches profusely,
although most of the Roman mastery of construction was lost in this epoch. How-
ever, the use of arches was not fully understood until the scientific analysis of Hooke
(1635–1703), who was the first to find an explanation for the good load-bearing capa-
bilities of Roman arches. See Hooke’s portrait in Fig. 3.2. In the painting, Hooke is
represented with a chain in his hands, in memory of, precisely, his findings about the
shape of arches.
In his work “A Description of Helioscopes, and Some Other Instruments”, Robert
Hooke wrote the solution to the shape of the perfect arch (“a problem which no
Architectonick Writer hath ever yet attempted, much less performed”) in the form
of an anagram, see Fig. 3.3.
The solution was written as:
abcccddeeeeefggiiiiiiiiillmmmmnnnnnnooprrsssttttttuuuuuuuux.
After Hooke’s death, the executor was allowed to make public the correct ordering
in the phrase:
3.2 A Brief Historical Perspective 57
Fig. 3.2 A portrait of Robert Hooke manipulating a chain. Rita Greer, 2009
Fig. 3.3 Anagram written by Hooke in 1676 containing the solution to the problem of the perfect
shape of an arch
Besides having included an extra “e” and an extra “i”, the meaning of the anagram
was “As hangs the flexible line, so but inverted will stand the rigid arch”.
The idea is simple, yet appealing. If a cable supports loads in pure traction, its
inverted image will support loads in compression, see Fig. 3.4. This extends in a
straightforward manner to arches subjected to distributed loads, see Fig. 3.5, where
we discover that the “perfect” shape for an arch under distributed load is a parabola.
58 3 Arches
P
P
f
Fig. 3.4 Hooke’s solution to the problem of a perfect arch by just inverting the shape of the
equivalent cable
q
Fig. 3.5 An arch with a distributed load. Following Hooke, the solution will be an arch of parabolic
shape
If Hooke’s solution to the problem is correct, why did Romanesque arches never
fall and are still bearing loads today, despite being semicircular? The answer is
easy, see Fig. 3.6. In fact, the dimensions of the stone blocks used are forgiving.
To guarantee stability, quarry workers once produced big blocks. If this is the case,
the funicular curve is contained entirely within the geometry of the stone arch, thus
guaranteeing compression somewhere within the block. But if we design a thin arch,
see Fig. 3.7, the possibility arises for the funicular curve to exceed the limits of the
3.2 A Brief Historical Perspective 59
Fig. 3.7 Load bearing for a thin Roman arch. Funicular curve (red, dashed line) of the load and
true geometry. Note that hinges are produced where the funicular curve crosses the geometry of the
arch, thus causing a loss of stability
Fig. 3.8 Experimental design of the Sagrada Familia Cathedral in Barcelona by Antonio Gaudi.
The picture has been turned upside-down so as to appreciate the design of the arches. The small
sand bags in the picture are estimated so as to be proportional to the true loads in the structure
arch. Where the funicular curve crosses the arch a hinge appears, with up to three
being possible. With three hinges, the structure is actually a mechanism, and any
horizontal load will result in a collapse.
60 3 Arches
Therefore, what actually happened is that Romans and artisans during the Middle
Ages learned by doing. And they learned that, for an arch to be stable, its dimen-
sions must be large enough. Without knowing it, what they were doing was actually
enabling the space for the funicular curve to develop entirely within the geometry
of the arch. Probably many arches have fallen since the time of early Romans, but
those that have survived are really stable.
This discussion, far from being old-fashioned, as one might think, is really impor-
tant. Materials of substantial architectural importance, such as concrete, for instance,
behave not very differently than stone. In fact, the tensile stress that concrete can bear
is 10–15% the value of its compressive stress, and is often neglected for practical
dimensioning.
Note that, as mentioned in Chap. 2, if permanent loads predominate in the dimen-
sioning, a catenary curve should be preferred as the shape of the arch. This motivates
the chain in Hooke’s portrait (it suggests that Hooke was actually giving shape to
an arch). This same technique was used by Antonio Gaudi to study the arches in the
Sagrada Familia Cathedral in Barcelona (when visiting, do not miss the exhibition
in the crypt of his model of the structure and the loads applied to each arch, see
Fig. 3.8).
After the historical perspective presented in the previous section, we are now in the
position to establish the modern hypothesis commonly accepted for the design of
masonry structures, in particular, arches [10]. It is generally accepted that masonry
structures can bear compressive internal forces very well and that, in general, the
resulting stresses are one or two orders of magnitude below their limit values. Failure
by compression is in fact a very rare event.
In addition, masonry structures are not able, in general, to resist internal traction
forces. Traditionally, this has been simply neglected. Furthermore, because com-
pressive forces are predominant, slipping between stone blocks or bricks occurs also
very rarely. So, masonry arches are dimensioned to accommodate solely the axial,
compressive forces.
In view of this, we will accept the following hypothesis:
1. The arch possesses an infinitely great limit to stress under compression, i.e.,
σe+ = ∞.
2. Masonry arches do not have any tensile capacity, i.e., σe− = 0.
3. Failure by slipping between blocks or bricks is impossible,
where symbols + and − have been used to indicate compression and traction, respec-
tively. This convention is, nevertheless, purely arbitrary.
3.4 Limit Analysis of Masonry Arches 61
The design process for masonry (stone, brick, adobe) arches under the hypothesis
1–3 above is limited by the fact that, under pure equilibrium conditions, there is no
unique solution to the problem. This can readily be understood by noting that there
exist two limit funicular curves for every arch, see Fig. 3.9.
Observe that, in both limit cases as depicted in red and blue, respectively, the
keystone of the arch will present a hinge at the extrados and intrados, respectively.
Any other funicular curve between these two is equally valid, since it provides a
purely compressive line of thrust and, under hypothesis 1 above, will lead to a stable
arch. Practical methods exist for the design of limit-thrust lines, such as the one
proposed by an anonymous author but commonly attributed to Fuller [11].
The limit analysis of masonry arches, therefore, cannot be achieved by simple
equilibrium considerations, since an infinite number of solutions exist, as we have
just seen. This is caused by the hyperstatic or statically indeterminate character of
these structures. The history of arch design in the 19th century is the quest for the
“true” thrust line in the arch.
Moseley (1843, see [11] and references therein) established that the true thrust line
is the one with a minimal-force value (a sort of naive principle of minimum potential
energy). However, Poncelet (1852) suggested that, to find the right solution, it is
necessary, as in every statically indeterminate structure, for the rest of equations of
the elastic problem, particularly the constitutive equations, to come into play. Spanish
architect Eduardo Saavedra (1860) was however the first, soon before Castigliano, to
state that, to find the right solution for the thrust line in arches, the complete elastic
problem must be solved first.
Heyman, however, was the first to study systematically the effect of the perturba-
tions in arches [10]. By perturbations, we mean the effects of formwork removal, see
Fig. 3.10. After removal, the loads begin to act on the arch, so that it suffers a descent
that causes a horizontal thrust on the springer (lowest voussoir on each side of an
arch). Inevitably, three hinges form at the keystone and the springers, see Fig. 3.11.
The arch with these three hinges is now statically determinate.
It is extremely important at this point is to recall here the lower bound theorem
of the theory of structures.
Fig. 3.12 A wooden, three-hinged arch for the construction of a pedestrian bridge at Mazcuerras,
Cantabria, Spain a A general view of the bridge. b Hinge at the top. c Hinge at the supports
Under this framework, therefore, it is clear that the design of an arch is limited
to the verification that, under any possible load, the thrust line (funicular curve of
loads) passes within the intrados and extrados of the arch. We will then obtain a sort
of envelope of funicular curves that will provide us with the right shape for the arch.
64 3 Arches
Fig. 3.13 Musée d’art moderne et d’art contemporain, also known as MAMAC, in Nice, France.
Yves Bayard and Henri Vidal, architects. Henri Vidal also acted as structural engineer, his primary
degree, prior to becoming an architect
The dependence of the behavior of the arch on the perturbations, i.e., on the way
formwork is removed, or in the relative stiffness of the terrain below the foundation,
for instance, is always troublesome and somehow unpredictable. That is the reason
why most architects once chose statically determinate arches to design their struc-
tures. How to design a statically determinate arch? Simple: just force the thrust line to
3.5 Modern Arch Structures 65
Fig. 3.14 Model of the shell roof of the British Museum. Beatriz Moya, M.Sc. dissertation. School
of Engineering and Architecture, Universidad de Zaragoza, 2017
Fig. 3.15 Shell under purely compressive internal forces. Miguel Angel Campos, M.Sc. disserta-
tion. School of Engineering and Architecture, Universidad de Zaragoza, 2014
pass through three chosen points by placing three hinges at these points. Thus, three-
hinged arches became a very popular structural typology that continues to capture
undeniable interest, see Fig. 3.12.
Of course, there is in reality no strict need for a true mechanical device that allows
the two half-arches to rotate freely—this is the strict definition of a hinge. In general,
and especially for reinforced concrete structures, a simple abrupt change of section
is enough for the structure to behave as hinged.
66 3 Arches
Fig. 3.16 A pavilion vault. Tower of the old building of the Technical University of Vienna
Fig. 3.18 Principal stresses of the four hypars forming the famous umbrella by Felix Candela.
Notice how the third principal stress (relative to bending, or out-of-plane stress) does not appear in
the finite element model of the shell
Fig. 3.19 Structure of the Oceanogràfic in Valencia, Spain. F. Candela and S. Calatrava, architects.
The shell is formed by the intersection of five hypars
68 3 Arches
dimensional, rather than plane, two-dimensional structures, until 2007 [2]. Indeed,
it is not self-evident to find the funicular surface, given the shape of the plant of the
building and an interval in height for every node in the model. See Fig. 3.14 for a
simplified model of the dome of the British Museum following this theory.
Under the so-called Thrust Network Analysis (TNA) theory, a network of com-
pression forces is sought so as to guarantee equilibrium within an interval of pre-
scribed intrados–extrados surfaces. This is a powerful technique that guarantees
structures bearing load under purely compressive forces resulting extremely thin
shells, see Fig. 3.15.
There exist, however, some surfaces whose structural behavior is well-known
now for some centuries. One of these is the so-called groin vault, the result of the
intersection of two mutually perpendicular barrel vaults. The pavilion vault can
be seen as the dual of the groin vault, see Fig. 3.16. In this case, arches rely on
the arrises (intersections of the four barrel vaults, known as spindles). Following
the just introduced theory, arches rest on the arrises, exerting a thrust that is in
principle equilibrated (at least for self-weight, permanent loads; not for wind loads,
for instance) under purely compressive internal forces.
Another example of an interesting surface, well-known due for its inherent struc-
tural properties, is that of hyperbolic paraboloids, hypars, for short, see Fig. 3.17.
Fig. 3.20 A groin vault with cables to avoid excessive horizontal thrusts. Cividale del Friuli, Udine,
Italy
3.6 Shells, Vaults and Domes 69
Fig. 3.21 A tied-arch structure. Top, sketch of the structure. Bottom, an example: the Third Mil-
lennium Bridge in Zaragoza, Spain
70 3 Arches
3.7 Arch-Cables
As already analyzed, the load in an arch or a barrel vault always implies horizontal
thrust in the springer. This is the reason for the need, in some cases, of attaching a
cable to absorb these forces, see Fig. 3.20.
This idea leads us to the concept of bowstring or tied-arch bridges (this concept has
its greatest utility in the design of bridges), see Fig. 3.21. In the right support, a new
symbol is employed that represents a type of support (in many books it also include
one, two or even three wheels between the two horizontal lines) which indicate that it
can move horizontally and rotate freely, but that the vertical displacement is impeded.
The goal is that the horizontal thrust is absorbed by the cable, so that no horizontal
force is transmitted to the soil or to the remainder of the structure.
This is also the key main structural element in the roof of the new London
Heathrow Terminal 5 airport, see Fig. 3.22.
Taking this idea to the limit, a very efficient way of designing cantilevered struc-
tures is by joining an arch and a cable or, more properly, a tie and a strut, see Fig. 3.23.
This is the basic constructive element of the famous Hong Kong and Shanghai
Bank building by N. Foster, see Fig. 3.24. The previous bowstring structure and the
strut-and-tie concept will naturally lead us to the fabulous world of trusses in the
following chapter.
Exercises
3.1. Estimate the position of the hinges after formwork removal for a horseshoe
arch, see Fig. 3.25.
3.2. Do the same for a Gothic arch. Estimate the necessary span-to-thickness ratio
for this type of arch. Analyze the differences between them in a purely structural
sense.
3.3. Figure 3.26 represents a staircase for access to the Carcassonne castle walls.
As can be seen, it consists of two semi-arches, both supported by a board column
where they meet in the middle. With the help of graphical statics, draw a sketch of
the catenary curve of both arches and estimate the width of the central support such
that the resulting force lies within its section (thus avoiding failure).
3.4. Investigate Blondel’s rule (1682) in the dimensioning of arches and try to justify
its validity. This was a practical recipe employed at that time by master architects and
which proved to work quite well, judging by the number of cathedrals still standing.
3.5. Justify the moon-like shape of the arch of the Maria Pia bridge in Porto,
Fig. 3.27. Is there any structural reason for this particular shape?
Fig. 3.27 Maria Pia bridge in Porto, Portugal. In the background, the Sao Joao bridge is also visible
Chapter 4
Trusses
Abstract This chapter introduces the basic aspects underlying structures made of
trusses.
4.1 Introduction
We have seen that cable structures can be considerably stiffened if additional bars are
added, thus preventing excessive displacements under variable loads, see Fig. 4.1.
This simple idea is at the origin of trusses, although not from a historical perspec-
tive, see Fig. 4.2, where a sort of truss lacking some diagonals is depicted. This is
typical of some well-known historic buildings. They can also be viewed as a means
to avoid mechanisms, see Fig. 4.3. The addition of the diagonal bar is precisely what
prevents the structure from collapsing as a mechanism.
The truss can also be seen as an evolution of the bowstring or arch-cable structure.
Let us assume that the deck of the structure bears a critical load so that we decide to
hang it from the arch, see Figs. 4.4 and 4.5. If, in addition, we add some diagonal bars
to prevent large motion due to variable loads, we arrive at a classical truss structure.
Trusses, similar to cables and arches, bear only traction and compression internal
forces. That is why they became very popular, especially after the Industrial Rev-
olution in the 17–18th centuries and the widespread availability of steel, Fig. 4.6.
Despite their conceptual simplicity, trusses continue to be widely employed both in
industry and architecture, notably for buildings including long spans.
4.2 Hypothesis
P P
N1
N3
N2
Fig. 4.2 A sort of primitive truss with a clear lack of diagonal bars. This type of truss is called a
queen-post truss
Fig. 4.3 A truss used as a means to avoid collapse through a mechanism. The dashed line represents
the mechanism-type of displacement experienced by the structure in the absence of the diagonal
bar
(a)
(b)
Fig. 4.4 The truss as an evolution of a tied arch or bowstring structure. a After the addition of the
vertical struts and b after the addition of the diagonals
the triangular configuration of bars that this hypothesis is, to a very high degree of
accuracy, true.
Therefore, in the design and analysis of trusses we will accept the following
hypothesis:
1. Trusses will be composed of straight bars.
2. We will consider joints between bars as being hinges (even if they are not literally
so).
3. A load is assumed to be applied at the nodes only (otherwise it will inevitably
produce bending).
Under these very basic hypothesis, trusses will bear loads by developing axial
internal forces only, and therefore Eqs. (2.1)–(2.3) still hold. To see confirmation of
this, consider the structure in Fig. 4.8.
This structure was analyzed under the hypothesis of free rotation at the nodes
and again under the rigid-node assumption, just to see the difference in the results.
78 4 Trusses
Fig. 4.5 Railway bridge at Mandelieu-La Napoule, France, over the Siagne River (1862), a typical
example of a bowstring bridge. The arches were added to an existing bridge to reinforce the structure
in 1936
These are summarized in Table 4.1. As can be seen the differences are minimal, often
under 1%.
Of course, if nodes must rotate freely, so do the supports. Symbols in Fig. 4.8, for
instance, indicate that rotation is allowed (left symbol) and that, in addition, horizon-
tal displacement is free (right symbol). Examples of these supports are also shown
in Fig. 4.9. These supports closely resemble the small symbols that we normally
include in our models, but in fact they do not need to be so similar. To behave as a
pinned support, it suffices that, to a reasonable extent, the support allows a certain
degree of rotation. The symbol on the right of the structure of Fig. 4.8 is a roller.
It must allow some degree of horizontal displacement (see the right support of the
truss in Fig. 4.10).
Consider for a moment the truss in Fig. 4.11. It is similar to the one in Fig. 4.3, but we
have added one more bar. Now, every node has three bars arriving to it. It is impossible
to find a suitable funicular polygon to determine every force. An analytic approach to
4.3 Statically Determinate and Indeterminate Trusses 79
the problem will give the same result: with only two equilibrium equations (vanishing
sum of forces in vertical and horizontal directions, for instance) it is impossible to
determine three unknowns. The situation could be even worse if we consider two
pinned supports, for instance. In that case not even the reactions at the support could
have been determined by equilibrium considerations.
We know from Chap. 1 that we have enough equations to solve the problem.
There is no question at all! However, it is important to known whether we must make
them all work or not. Up to now, due to the particular mechanisms of bearing loads,
masonry arches have been able to develop statically determinate configurations,
as noticed by Heyman [10]. With one single cable, there is always a way to find
their internal forces by means of a simple Cremona diagram. This is not the case,
however, for cable beams, as we have seen. Now, for trusses, this will not be the case,
in general, and it is interesting to have a tool to know in advance the degree of static
indeterminacy of the truss.
By isolating a particular region around a node of the truss and analyzing the
resulting free-body diagram, we have noticed that we are able to find two forces at
most. This is equivalent to the two force equilibrium equations mentioned before. In
principle, there should be a third one: the equilibrium of momenta. This is equivalent
to making all the forces cross at a certain point of space (otherwise, they will provoke
80 4 Trusses
rotations). But they always cross at one point, precisely the node. Therefore, there is
a simple rule to know in advance if a truss is statically determinate or not:
• if n reactions + n bars < 2n nodes , the systems is unstable (a mechanism),
• if n reactions + n bars = 2n nodes , the systems is statically determinate and,
• if n reactions + n bars > 2n nodes , the systems is statically indeterminate.
See Fig. 4.12 for an example of this simple rule. Note that in the case of Fig. 4.12a
the structure is actually stable but critical: any perturbation in the horizontal direction
will activate the collapse mechanism. Under the forces depicted in the figure, how-
ever, the truss is in equilibrium, with vertical bars bearing an internal force of value
P. We see, in addition, that there are various sources of indeterminacy. While in
4.3 Statically Determinate and Indeterminate Trusses 81
(a)
(b) (c)
Fig. 4.8 a An example truss. b An old pinned support apparatus. c Old roller supports
Fig. 4.12c this indeterminacy is provoked by one extra (often called superabundant)
bar, in Fig. 4.12d, the origin of this issue is the presence of two pinned supports, and
therefore four reactions.
For the time being, we will concentrate on the limit analysis of statically determinate
trusses. The indeterminacy is much more complex, and two large families of methods
exist for solving these cases. These are beyond the scope of this book, even if the
solution has been mentioned before: to add constitutive and compatibility equations
to the well-known equilibrium equations.
The first way to analyze trusses, already employed in cables and arches, is to employ
the Cremona diagram. The method is sketched with the help of the example in
Fig. 4.13.
82 4 Trusses
Table 4.1 Results for the truss structure in Fig. 4.8 with and without the hypothesis of hinges at
the nodes
Bar # N w/hinges (kp) N w/o hinges (kp)
1 0.000E+00 3.73E−01
2 1.50E+02 1.50E+02
3 1.50E+02 1.50E+02
4 0.00E+00 3.73E−01
5 −2.00E+02 −2.00E+02
6 2.12E+02 2.11E+02
7 −1.50E+02 −1.49E+02
8 7.07E+01 7.03E+01
9 −1.00E+02 −9.97E+01
10 7.07E+01 7.03E+01
11 −1.50E+02 −1.49E+02
12 2.12E+02 2.11E+02
13 −2.00E+02 −2.00E+02
14 −1.50E+02 −1.50E+02
15 −2.00E+02 −2.00E+02
16 −2.00E+02 −2.00E+02
17 −1.50E+02 −1.50E+02
Fig. 4.9 Examples of pinned supports. Left: Brussels train station. Right: Madrid Airport Terminal 4
4.4 Limit Analysis of Statically Determinate Trusses 83
Fig. 4.10 An examples of a roller support (the one on the right of the central support—the one of
the left is pinned). Paris metro line near Glaciere station. Boulevard Auguste Blanqui
To begin with, we must ensure that the entire structure is not able to rotate as a
rigid solid. This is guaranteed if the three external forces that act on the structure,
P, R1 and R2 meet at one single point. The line of action of P is known, and so it
is the one of R2 , since the support is a roller and therefore is not able to react with
an horizontal component. R1 will certainly have two components, and they can be
found by forcing it to pass through the crossing point of R2 and P, see Fig. 4.13. On
the right of the figure the resulting Cremona diagram is shown. In it, P is known,
both in direction and magnitude. We know only the lines of action of R1 and R2 ,
from the sketch in Fig. 4.13 (right), but this is enough to, by tracing parallels at both
ends of P, determine their intersection, thus fixing the modules of R1 and R2 .
Note that the above procedure is possible just because the reaction R2 is vertical
(i.e., because its support is a roller). If the support is changed to a pinned one,
then the structure would be statically indeterminate, and the line of action of
R2 would be unknown. The Cremona diagram would have no solution.
Once the reactions have been determined, the procedure continues node by node,
by imposing equilibrium or, in other words, a closed-force polygon. For instance, for
the rightmost node in the structure, we have the Cremona diagram shown in Fig. 4.14.
84 4 Trusses
P P
P P P P
P P P P
(c) 3+6 > 2×4, statically indeterminate (d) 4+5 > 2×4, statically indeterminate
(internal). (external).
If the structure is statically determinate, we will not encounter any problem by pro-
ceeding node by node, always looking for some node with only two unknown forces.
Once all internal forces have been determined, the limit analysis of the structure is
simply to verify that
Ni
σi = ≤ σe , i = 1, 2, . . . , n bars ,
Ai
4.4 Limit Analysis of Statically Determinate Trusses 85
R2
P
R1
R1 R2
R2
N2
R2 N1
N2
Fig. 4.14 Cremona diagram for the rightmost node in Fig. 4.13
holds. Conversely, if one wants to design the structure, to determine the necessary
area of the cross section for each bar, it suffices to compute
Ni
Ai ≥ , i = 1, 2, . . . , n bars .
σe
Very often, for aesthetic criteria, all areas are taken as equal (and therefore the design
will be governed by the bar subjected to the maximum Ni ).
where moments have been taken with respect to the pinned support. Interior angles in
the triangles are assumed to be 60◦ . We have, therefore, three equations to determine
R1H , R1V and R2 . Again, as in the graphical analysis, if we have a pinned support at
the support on the right, then a fourth unknown R2H would appear, therefore making
the system of equations undetermined.
Once the three reactions have been found, we can continue node by node by
equilibrating their respective internal forces with the externally applied ones. Again,
for the node with a roller support,
FH = 0 → N2 + N1 cos 60 = 0,
FV = 0 → R2 − N2 sin 60 = 0,
where both N1 and N2 have been (arbitrarily) assumed to act towards the right. A
minus sign in the result will indicate that they actually act in the reverse direction.
Observe that it is nonsense to try to apply the equilibrium of moments again for
each node. This is due to the fact that the implied forces cross by definition at the
node, thus vanishing the moment trivially.
P
N5
N4
R1 R2 R1 N3
Consider again the truss in Fig. 4.13 and assume we want to obtain the value of
the internal forces N3 , N4 and N5 in Fig. 4.15. By performing an imaginary cut (red,
dashed line in the figure), and keeping only the left portion of the structure, we can
equilibrate solely this part of the structure.
Notice how reaction forces R2 and loads such as P do not intervene in this equi-
libration procedure. Now it is mandatory to equilibrate both forces and moments,
since N3 , N4 and N5 do not cross at a single point. Therefore, we have,
FH = 0 → R1H + N5 + N4 cos 60 + N3 = 0,
FV = 0 → R1V − N4 sin 60 = 0,
M = 0 → N4 cos 15 × (/2) + N5 cos 15 = 0,
where moments have been taken with respect to the pinned support (taking advantage
of the fact that two forces pass through this support). Notice also that the length of
each bar is /2. Again, we have three unknowns and three equations. Statically
indeterminate trusses will lead at some point to a system of three equations with
more than three unknowns.
As already commented in previous sections, structures only must bear the forces
they are designed for, but, notably, their deflections must be small enough so as to
not compromise their serviceability. Therefore, it is of utmost importance to have a
method for easy determination of these displacements. Several such methods exist,
such as Mohr’s method. But perhaps the most useful one, and undoubtedly the one
that will have more implications in the solution of statically indeterminate structures,
is the Principle of Virtual Work.
88 4 Trusses
To begin with, let us consider a single truss bar (axial rod) under a distributed load
per unit length, p, see Fig. 4.16.
As already seen from Eq. (2.2), for traction–compression processes, we have
σ (x) N (x)
εx (x) = = ,
E AE
so that
N (x) = AEεx (x).
where A is the cross-sectional area of the bar and E the Young’s modulus of the
material. The product AE is often referred to as the axial stiffness of the bar. As
always, x is assumed to be the longitudinal axis of the bar. If we isolate it, as we did
repeatedly, but noteworthy in the method of sections, a portion of the bar of length d x
(see the detail in Fig. 4.16, right), the only forces acting on it are precisely this axial
internal force that suffers some variation d N between both ends of the differential
segment and the external load pd x (even if we rarely consider this type of load). In
this case,
N (x) + d N − N (x) = − pd x,
and therefore,
d
(AEεx (x)) = − p.
dx
Since
du x
εx (x) = ,
dx
we arrive at
d du x
AE = − p, (4.1)
dx dx
which is the differential equation governing the displacement of the bar. Obviously,
the first possibility to determine the displacement law of the structure is to integrate,
bar by bar, this differential equation. But imagine the process of doing it, for instance,
for the Eiffel tower! This is simply out of reach for many structures of practical inter-
est. The procedure we are about to develop enables one to obtain one particular nodal
displacement without the need of integrating the differential equation throughout the
entire structure.
To that end, let us multiply both sides of Eq. (4.1) by an arbitrary function ψ,
du x
ψ AE = −ψ p,
dx
and then let us integrate both sides all along the bar,
4.5 Serviceability Limit-State Analysis of Trusses 89
Nb
p pdx
N (x) + dN
N (x)
Na
Fig. 4.16 Equilibrium in a single truss bar. On the right, equilibrium of a differential portion of the
bar
du x
ψ AE dx = − ψ pd x.
0 dx 0
ψ
where εx represents some virtual deformation coming from the arbitrary “displace-
ment” field ψ. This is precisely the expression of the Principle of Virtual Displace-
ments:
ψ
|ψ N |0 − εx N d x = − ψ pd x. (4.2)
0 0
where we have grouped on the left-hand side the term relative to external loading and
on the right-hand side the contribution of internal forces to the virtual work. Here,
the sole requirement for the virtual-force field is to be in equilibrium. Of course, it is
“virtual” in the sense that it does not correspond to the true solution of the structure.
90 4 Trusses
It is precisely Eq. (4.3) that is the one we will use for the determination of dis-
placements at a given point in the structure. The freedom to choose the external
loading allows us to do so.
Hence, to calculate the displacement (respectively, rotation) of a given node in
the truss, we will follow these steps:
1. Apply one single external force (respectively, moment) of unity value, at the
node whose displacement is sought, say Q.
2. This virtual force will be oriented in the direction in which we are interested
(typically, vertical).
3. Under these assumptions, Eq. (4.3) will look like
N Nψ
1 · uQ = εx N ψ d x = d x,
nbars 0 nbars 0 AE
where u Q is precisely the sought displacement at node Q and nbars the number
of bars in the structure. Note that, due to the freedom provided by the Principle
of Virtual Forces, we have chosen p ψ = 0.
Note that there is one exception to the criterion of “one single virtual load”.
If the original structure is symmetric, and is subjected to symmetric loading,
it is often preferable to apply two symmetric loads so as not to destroy the
symmetry. In that case, the fact of having two loads must be taken into account,
of course, since the sought displacement will be affected by a factor of two.
Note also that the application of the Principle of Virtual Work needs, for the
solution of the structure (determination of internal forces), two things: one for the
original, actual loading, and a second one for the virtual loading. This last item is,
in general, very simple, since the applied load is composed in general by one single
force. Once the practitioner has acquired some dexterity, it will be evident for him
or her how best to solve the resulting system.
As an example of the just proposed technique, consider the following structure:
We are going to compute the vertical displacement at point E of the structure.
To that end, we must first analyze the structure of Fig. 4.17 and obtain its internal
forces:
√ √
3 3
NC D = − P NE D = − P
2 2
√ √
3 1 3 3
NC E = P NE B = + P
2 2 4
√ √
1 3 3
N AE = + P N AC = − P
2 4 2
4.5 Serviceability Limit-State Analysis of Trusses 91
P/2 A E B P/2
P
VA VB
HA
A E B
√
5 3
NDB =− P
6
Once the true internal forces have been determined, we proceed to establish the
virtual loading for the structure. In this case, of course, it will consist of a single load
at node E, see Fig. 4.18.
In this case, the internal forces result in
ψ ψ ψ ψ ψ
√
N AC = NC D = −NC E = −N E D = N D B = − 3/3,
and √
N AE = N E B = 3/6,
where, again, we highlight the virtual character of these forces by using a superscript
ψ. Therefore, the vertical (downwards) displacement at node E will be
√ √ √ √ √ √
1 3 3 3 3 3 5 3
u v (E) = P+ P+ P+ P+ P
AE 3 2 2 2 2 6
√ √ √ √
3 1 3 1 3 3 20 + 3P
+ + P+ + P = .
6 2 4 2 4 6AE
92 4 Trusses
Fig. 4.19 Influence of the orientation of the diagonals. Note how a change in their orientation
changes drastically the sign in the internal forces. Red lines indicate traction while blue, dashed
lines indicate compression
So far, we have studied how to ensure an adequate behavior so as not to reach limit
states. It is interesting, however, to develop some intuitive knowledge about how
trusses work and how their different parameters affect their serviceability.
For instance, it is easy to verify that the horizontal internal forces in the bars are
roughly proportional to the span of the truss. This means that doubling the span also
doubles these forces. Conversely, halving the rise will lead to doubling the forces
again.
Indeed, it is easy to understand from the Cremona diagram that internal forces in
the truss only depend on the direction of the bars. Therefore, for constant external
loads, doubling both the rise and the span of the truss leads to exactly the same
Cremona diagram.
4.6 Criteria for the Design of Trusses 93
While in general the upper chord will be subjected to compression and the lower one
to tension, the orientation of diagonals completely determines its internal forces. This
is more easily seen in the so-called Howe (or N-) trusses,1 see Fig. 4.19. In it, tension
forces have been represented by a red line with a thickness roughly proportional to
the intensity of the force. Compression forces are represented by a dashed, blue line,
with thickness proportional again to the intensity of the force.
It is worth noting also that the force in the diagonals is inversely proportional to
the internal angle, θ , see Fig. 4.19. Therefore, the smaller the angle, the bigger the
internal force. However, approaching θ to 90◦ makes the force in the struts (vertical
bars) greater.
1 Actually first proposed by Andrea Palladio in 1570 in his four books on architecture.
94 4 Trusses
There are many various designs of trusses (most of them patented during the 19th
century), with different characteristics, but whose overall behavior obeys the just-
introduced criteria.
Apart from Howe trusses, just described, the so-called Long system (or X diago-
nals), see Fig. 4.20, is very popular. Note that this type of truss is statically indeter-
minate, and therefore will require some more sophisticated method of analysis than
the ones considered herein.
Another very popular type of trusses is the so-called K-diagonal truss, see
Fig. 4.21. Despite its more or less complex appearance, it is statically determinate.
During the 19th century, the Poncelet truss became very popular due to its very
light appearance, and it was profusely employed in train stations across Europe, see
Fig. 4.22.
In general, albeit obvious, care must be taken when reversing the sign of the loads.
This is very often the case for suction-wind loads. Therefore, nothing guarantees that
a diagonal bar will always be under traction. This is dangerous since compression
could eventually lead to buckling, something that is beyond the scope of this book,
but that must definitely be taken into account.
work together just by putting secondary beams between them. This is the system
pursued in the Charles de Gaulle Airport in Paris, for instance, see Fig. 4.23.
A second, more elaborate, possibility, is to construct the three-dimensional struc-
ture by creating a lattice, see Fig. 4.24.
96 4 Trusses
Worked Examples
4.1 Given the truss in Fig. 4.25, compute the internal forces N6 and N19 , taking into
account that the vertical and horizontal dimensions of the bars are L.
Solution 4.1 First of all, and with an eye towards the minimization of work, note
that the structure is symmetric, and so is the load. Therefore, we will have,
P
RL = R R = .
2
(This same result could have been obtained by enforcing the vertical equilibrium of
loads plus the equilibrium of moments.)
To avoid the enforcement of equilibrium equations node by node, until we arrive
at bar 6, we can employ the method of sections. We thus perform an imaginary cut
through the structure (dashed line in the figure) and, by keeping only the right part
of the structure, we proceed to enforce global equilibrium for this part (Fig. 4.26).
Enforcing the equilibrium of moments at points A and B, respectively, we obtain,
P
Mz (A) = · 5L + N6 · L = 0,
2
P
Mz (B) = · 6L + N19 · L = 0.
2
P
1 2 3 4 5 6 7 8 9 10 11 12
24 23 22 21 20 19 18 17 16 15 14 13
RL RR
B
N6
NAB
N19
A
P
2
2 kN
2 kN 2 kN
H
1 kN 1 kN
2m
F G
1m
A E
B C D
RA RE
2m 2m 2m 2m
5
N6 = − P (compression),
2
N19 = 3P (traction).
One can arrive, in this case, at exactly the same result, by enforcing force equi-
librium.
4.2 Determine the vertical displacement of node C in the truss in Fig. 4.27.
Solution 4.2 (a) We begin by enforcing equilibrium in the structure globally. This
will allow us to determine the reaction forces in the supports.
Given the symmetry of the structure and the lack of horizontal loads, vertical
reactions can be found by simply noting that the sum of vertical forces is 1 + 2 +
2 + 2 + 1 = 8 kN, so that
8kN
R A = RE = = 4 kN.
2
By next applying the equations of (internal) equilibrium at the nodes of the truss,
we are in the position to find the internal forces at each bar. In this case, the application
of equilibrium of moments is simply nonsense, since every force acting on a node
passes inevitably through it, thus generating no moment. Therefore, we must proceed
by isolating and equilibrating nodes for which at most two forces are unknown.
Otherwise, we will not be able to find their values.
98 4 Trusses
4 kN
3·2 1 √
FY = 0 ; + N AF · √ = 0 ; N AF = −3 5 kN (Compression),
2 5
2
FX = 0 ; N AB + N AF · √ = 0 ; N AB = 6 kN (Traction).
5
A
6kN NBC
B
2 √
FX = 0 ; √ NC F + N F H + 3 5 = 0,
5
1 √
FY = 0 ; √ − NC F + N F H + 3 5 = 2,
5
H
α α
√
2 5 NHG
F
G
NCH
NC H = 2 kN (Traction).
By simply imposing the symmetry of the structure, we find the rest of forces in
the structure (Fig. 4.31).
(b) To determine the vertical displacement at point C, we need to define an appro-
priate virtual loading system. It will be composed of a single load, of unity value,
applied at point C, precisely the one whose displacement is to be computed. For that
purpose, we consider the system in Fig. 4.32.
By using the same methodology as in the previous section, we obtain
ψ ψ 1
R A = RE = kN.
2
Equilibrium at node A is shown in Fig. 4.33:
√
ψ 5
N AF =− kN (Compression),
2
100 4 Trusses
2m
F G
1m
A E
B C D
ψ ψ
RA RE
1 kN
2m 2m 2m 2m
Fig. 4.32 Virtual system to compute the vertical displacement of node C in Example 4.2
ψ
N AB = 1 kN (Traction).
ψ
N B F = 0 kN.
α B
A
ψ
NAB
0.5 kN
A
1 kN ψ
B NBC
4.7 Space Trusses 101
F α
√
α α
5
2
ψ
A NCF C
B
thus giving
√ √
1 5 5 1 √
1 · u y (C) = 4 · 12 + 2 · 15 + 2 · 10 +4 = 52 + 25 5 m.
EA 2 2 EA
Chapter 5
Beams (I): Geometric Description and
Internal Forces
Abstract This chapter introduces the basic aspects underlying structures made of
beams.
5.1 Introduction
P P
M
h
Let us assume that the internal force passes at a distance e from the center of
gravity of the cross section. By adding a pair of mutually equilibrating forces at the
center of gravity (in red and blue in Fig. 5.3), the equilibrium is guaranteed. It is
clear then that the original situation is equivalent to having an axial internal force N
(in red in the picture) plus a bending moment M = N · e.
In what follows we introduce a rigorous mathematical description of the geometry
of a beam that will help us to introduce, also in a very rigorous way, the series of
internal forces and moments that a beam may suffer.
5.2 Geometrical Description of a Beam 105
For the geometrical description of a beam, we follow classical references in the field
such as [1, 13] or [19]. For these authors, a beam is an example of a Cosserat medium,
i.e., a flexible line whose points join the centers of gravity of all the beam’s sections
that can rotate around these center points. More precisely, a beam will be described
by:
1. A curve ϕ(t) that joins the centers of gravity of all the beam’s cross sections. We
will employ a parameter t to travel along this line:
ϕ : I ⊂ R → R3 .
Of course, this parameter will nearly always be the arc-length parameter of the
curve.
2. A family of flat cross sections A(t).
It is then clear
that the volume of the bar will be the union of all the cross-
sectional areas, A(t). Also, the directrix coincides with the line of centroids, as
t
just introduced, so that
xd A = ϕ(t)A(t).
A(t)
Very often, intrinsic coordinates are employed in the description of the move-
ment, see Fig. 5.4. These correspond to the so-called Frenet–Serret trihedron. In this
classical mathematical parametrization of curves, the first axis (tangent) is always
oriented along the tangent of the curve, the second one (normal) is oriented towards
the derivative of the curve with respect to the arc-length parameter t and the third
one (binormal) results from the cross product of both. These form, in any case, an
orthonormal axis of reference.
A(t)
y
x
106 5 Beams (I): Geometric Description and Internal Forces
where n(t) is the vector normal to the section of the bar and t the stress vector, which
is a function of the particular point in the section, x.
providing a much more useful expression of the internal force vector. In this
case, N is the already well-known axial internal force, while Vy and Vz are the
shear internal forces. Many textbooks also employ the notation Vy = Q y and
Vz = Q z .
In a similar manner, we will define the vector of internal moments as the result of
integrating over the section the stress vector at each point, multiplied by its distance
to the centroid of the section, i.e.,
1 When there is no risk of confusion, we will speak of “internal forces” to refer to both internal
forces and moments.
5.3 Internal Forces and Moments 107
M(t) = (x − ϕ(t)) × td A = (x − ϕ(t)) × σnd A
A(t) A(t)
⎛ ⎞
σx
= (x − ϕ(t)) × ⎝τx y ⎠ d A,
A(t) τx z
with x an arbitrary point of the beam’s cross section and ϕ(t) the point of the line
of centroids corresponding to A(t).
Again, when the identity n = e1 holds, the vector of internal moments will be:
⎞ ⎛ ⎞ ⎛
σx Mx
M(t) = (x − ϕ(t)) × σnd A = (x − ϕ(t)) × ⎝τx y ⎠ d A = ⎝ M y ⎠ ,
A(t) A(t) τx z Mz
In sum, the complete picture of internal forces in a bar can be sketched as:
T1 = N = σx d A,
A(t)
T2 = Vy = τx y d A,
A(t)
T3 = Vz = τx z d A,
A(t)
M1 = M x = (τx z x y − τx y x z )d A,
A(t)
M2 = M y = σx x z d A,
A(t)
M3 = M z = − σx x y d A.
A(t)
These internal forces are depicted in Figs. 5.5 and 5.6, respectively. Note the right-
hand convention for the rotation of the moments.
108 5 Beams (I): Geometric Description and Internal Forces
Vy
Vz
y
N
x
Mz
y
My
x
Mx
We have already mentioned the fact that introducing a new variable, the internal
forces, in the problem, constitutes in fact a true advantage for the solution of some
problems. This will appear clear in subsequent chapters of this book. For the time
being, one of these advantages is the possibility of drawing internal force and moment
diagrams. These will allow us to obtain an overall description of the degree of loading
of each member of the structure.
To obtain these very useful diagrams, it is enough to proceed as in Sect. 4.4.3
and apply the method of sections, by enforcing equilibrium on arbitrary portions of
the structure (say, from x = 0 to x = ). This will provide us with an equilibrium
equation for internal forces of the type
B
A
with pe the vector of external applied loads in the bar portion x ∈ [0, ], i.e.,
pe = ds Xd A + ds td∂ A,
0 0
A(t) ∂ A(s)
with X the force per unit volume and t the force per unit length applied to the bar in
the portion x ∈ [0, ]. ∂ A(s) represents the boundary (external surface) of the bar in
this region. Note the vector character of these equations.
For the moment diagram, we proceed in exactly the same manner, so as to obtain:
ds (x − ϕ()) × Xd A + ds (x − ϕ()) × td∂ A
0 A(s) 0 ∂ A(s)
+ (x − ϕ()) × td A + (x − ϕ()) × td A = 0. (5.2)
A(0) A()
It is then clear that, knowing the internal forces at some point in the structure
(reactions in one of the supports, for instance) will allow us to apply successively
Eqs. (5.1)–(5.2) to obtain, directly from the external loads, and by equilibrium con-
siderations only, the internal force and moment diagrams.
Note that, in this case and always where a free, cantilevered beam end appears,
it is not necessary to begin by determining the value of reactions at the sup-
ports. This is due to the fact that a free beam end (with no punctual load nor
prescribed moment) has no internal forces or moments. Thus, we can begin
our calculations by assuming all internal forces to be identically zero at B.
110 5 Beams (I): Geometric Description and Internal Forces
RAH B
MA
RAV A
where we have taken A as the reference point with respect to which we compute
moments. This is, of course, an arbitrary choice. The sign for the moments has been
taken, also in an arbitrary manner, as positive in the counter-clockwise sense. The
method will work equally well with completely opposite sign conventions.
Once we have determined the reactions at the support points, we can proceed
by the method of sections to equilibrate arbitrary portions of the structure. Thus,
by cutting the structure at a point at an arbitrary distance from the free end, B, we
unveil (see Fig. 5.9) three internal forces: N (x), V (x) and M(x). We highlight the
dependence of these internal forces on the particular, but arbitrary, position x. These
represent the effect of the just eliminated portion of the structure on the conserved
one.
By applying again equilibrium equations on the portion of length x, we obtain:
5.4 Diagrams of Internal Forces and Moments 111
B B
A A
Fig. 5.10 Internal force diagrams. Left, bending moments. Right, shear force
FH = 0, → N (x) = 0,
x
FV = 0, → V (x) = pd x = px,
0
x
x2
Mx = 0 → M(x) − pxd x, → M(x) = p ,
0 2
where an abuse of notation has been made by naming x as both the variable to
integrate and the integration limits. Note that we have just determined that all the
internal forces are positive. This means nothing more than that we have correctly
assumed their sense.
It is worth noting that, since we are interested in computing the bending moment
at the cutting point x, moments must be computed with respect to precisely this point,
and not with respect to an arbitrary point, as in the global-equilibrium enforcement
for the computation of the reactions.
Note also that we have just determined that the diagram of shear internal force
varies linearly with x, while the diagram of bending moments varies quadratically
with x, see Fig. 5.10.
In the diagrams in Fig. 5.10, we include a small symbol indicating the sense
of both internal forces. These represent what is physically happening at the
beam. For instance, in the bending moment diagram, Fig. 5.10, left, the symbol,
composed by a small portion of the bar and two mutually equilibrating arrows
showing the sense of bending, indicates that the bending moment is actually
applying traction at the upper fiber of the beam and compression to the lower
one. How do we know the precise sense of the arrows? It is simple: just take
a glimpse to the equilibrium of the beam portion in Fig. 5.9 and notice that,
on the left side of the cut, there is a counter-clockwise bending moment. Place
an equivalent arrow on the left side of the small rectangle and complete the
symbol with the opposite arrow on the right.
Moments applying traction on the upper fiber of a beam are traditionally
considered negative, but this is only an old convention, and the contrary is
equally valid. We prefer to indicate clearly the sense of the forces and moments
by employing these small symbols.
112 5 Beams (I): Geometric Description and Internal Forces
Consider now what happens if, instead of a distributed load p, a single punctual load
P is acting at point B in the structure, see Fig. 5.11. We begin by determining the
reactions at the clamped support A.
In this case,
FH = 0, → R AH = 0,
FV = 0, → R AV = P,
M A = 0 → M A = P.
By employing again the method of sections, see Fig. 5.12, we arrive at:
FH = 0, → N (x) = 0,
FV = 0, → V (x) = P,
Mx = 0 → M(x) = P x,
B B
A A
Fig. 5.13 Internal force diagrams. Left, bending moments. Right, shear force
5.4 Diagrams of Internal Forces and Moments 113
In this second example, we can readily see that the shear internal force is constant
(and equal to P), while the bending moment varies linearly with x. See the resulting
diagrams in Fig. 5.13.
Again, we have plotted internal-force diagrams at the top of the beam. The same
tradition that indicates that bending moments are positive if they apply traction
on the upper fiber of the beam indicates that this diagram should be placed
on the “positive” side of the beam, i.e., in the upper part of it. Nevertheless, it
is nonsense to attribute any particular sign to the diagram, since it should be
valid for any of them. Therefore, you can place the diagram above or below
the beam, but always accompanied by the symbol of the moments that actually
take place.
In this case, the change from a distributed load per unit length to a punctual load
has necessitated also a change in the force diagrams. The bending moment diagram
has changed from a quadratic one (Fig. 5.10) to a linear one (Fig. 5.13), while the
shear-force diagram has changed from linear to constant. Of course, this behavior is
completely general and will happen every time we consider a punctual load in our
structure.
MM
0.85
5.6 Frames
A frame is a structure composed of one or more beams and columns or side walls,
see Fig. 5.16. Essentially, their functioning is the same as an arch whose funicular
polygon does not pass through the bars, thus causing bending.
Three-hinged arches or frames (see also Fig. 3.12) are widely used due to their
static determinacy. For such a system, the funicular polygon must pass through the
three hinges, and it is therefore well-known in advance, see Fig. 5.17.
To analyze this frame, first note that the number of unknown reactions at the
supports (two at A, two at E) will not allow, in principle, to determine them through
Fig. 5.16 A sequence of frames for a concert hall at Palacio de las Artes Reina Sofía. Valencia,
Spain. S. Calatrava, architect
5.6 Frames 115
B D
C
h
A E
HA HE
VA VE
Fig. 5.17 An example of three-hinged frame. In blue, the funicular polygon of the distributed load,
that must pass through the three hinges. When the funicular polygon extends beyond the bar sections
(almost everywhere in this case), bending appears
where we have taking moments with respect to point A; so that the line of action of
reactions V A , H A and HE passes through point A and therefore produces no moment.
We have three equations for four unknowns, so we need one more equation. This
will come from the hinge at C:
2
MC = 0, → q + HE h − VE = 0.
4 2
This last equation takes C (the hinge) as the reference point and includes forces and
moments coming from one of the bars starting at C. If we include the whole structure,
we will obtain an equation linearly dependent on Eq. (5.5).
From the moments equilibrium equation at A we obtain
116 5 Beams (I): Geometric Description and Internal Forces
VE = q ,
2
so that, see Eq. (5.4),
VA = q ,
2
which also verifies the symmetry of the structure and its applied loads. From the
moment equilibrium equation at C we obtain
1 2 2
HE = H A = q −q = 0.
h 4 4
Once the reactions have been determined, we can begin to obtain the internal
force diagrams by applying the method of sections. To that end, a suitable option is
to begin by performing an imaginary cut into the bar A − B, see Fig. 5.18.
Under these assumptions, we have:
FH = 0, → N (x) = −V A = −q ,
2
FV = 0, → V (x) = H A ,
Mx = 0 → M(x) = H A x,
where positive moments have been assumed to rotate in the counter-clockwise sense,
but exactly the opposite convention could have been adopted as well. Note also that
the lines of action of V A , V (x) and N (x) pass through the cutting point, so that they
produce no moment from x. These expressions are valid for any x ∈ [0, h], so for
any point in the bar A − B.
HA
VA
5.6 Frames 117
h
A
HA
VA x
To obtain the diagrams along the bar B − C, we perform a second imaginary cut
at an arbitrary position x.2 See Fig. 5.19 for more details.
In this case,
FH = 0, → N (x) = −H A ,
FV = 0, → V (x) + q x = V A ,
x2
Mx = 0 → M(x) = q − V A x + H A h.
2
It is worth noting that now both the internal axial force N (x) and the shear force
V (x) adopt a different direction than those of the bar A − B. Internal forces are
always relative to local coordinate axis, i.e., they depend on the orientation of the
particular bar.
Since the just-obtained diagrams are valid for any point x ∈ [0, /2], and taking
into account that the structure is symmetric and is subjected to symmetric loads, there
is no need to continue our quest for the internal-force diagrams. We have completed
our task. The only work remaining is to plot these diagrams, see Fig. 5.20.
2 Note that we should have taken a second coordinate x , different from x. Again, when there is no
risk of confusion, we adopt x as the default nomenclature for our through-the-bar coordinate.
118 5 Beams (I): Geometric Description and Internal Forces
B D
C
A E
Fig. 5.20 Resulting bending-moments diagram for the structure in Fig. 5.17
Fig. 5.21 A segment of Vierendeel beam at Incheon airport, Seoul. Note that the U-shaped, three-
dimensional truss is properly a truss. However, the segment in the center of the picture has no
diagonals, thus developing strong bending moments at the nodes
120 5 Beams (I): Geometric Description and Internal Forces
Worked Examples
5.1 Determine the internal force diagrams for this structure (Fig. 5.22).
100 kp
B
C
1m
500 kp/m
1m
A
2m
Solution 5.1 At the clamped support C, reaction forces can be obtained by enforcing
equilibrium:
Rx = −500 kp (←)
R y = 100 kp (↑)
If we perform a virtual cut in the bar A − B, at an arbitrary location where the load
q is applied, say x ∈ [0, 1], the resulting free-body diagram is depicted in Fig. 5.23.
For these forces, we enforce equilibrium, see Fig. 5.23, so as to yield
x2
M(x) = −500 · kp · m, ∀x ∈ [0, 1].
2
5.7 Vierendeel Beams 121
500 kp/m
x
A
A second virtual cut in the bar A − B, this time on a section after the load q,
x ∈ [0, 1], provides the free-body diagram depicted in Fig. 5.24.
Again, we enforce equilibrium. First, in the vertical direction, so as to obtain:
V (x)
x
500 kp/m
1m
Finally, a third cut along B − C, see Fig. 5.25, is performed. In this case
(Figs. 5.26, 5.27 and 5.28):
1m
500 kp/m
1m
A
x
1m
500 kp
1m
A
2m
500 kp
1m
0 kp A
2m
5.7 Vierendeel Beams 123
950 kp · m
750 kp · m
B
750 kp · m
C
1m
250 kp · m
1m
0 kp · m A
2m
5.2 Determine the distribution of internal forces for the structure in Fig. 5.29.
Solution 5.2 As we have done repeatedly, we first equilibrate the whole structure
so as to determine the reactions at A:
RxA = 0 kp,
R yA = 0 kp,
A
C
2m
500 kp/m
B
2m D 2m E
R Cy = 2000 kp (↑).
Since no horizontal load exists, the sum of vertical reactions must equilibrate the
vertical loads, 2000 kp. By taking moments at A, for instance, we find
R Cy · 2 = 500 · 4 · 2.
Performing now a virtual cut at an arbitrary position along A − B, first in the zone
where no load exists, we conclude that no internal force exists along A − B.
Given that C − D possesses two hinges at the ends (it is, therefore, a cable), it
will suffer only axial internal forces. This force is identical, obviously, to the reaction
at support C.
To determine the internal forces at B − D we perform a new cut, as sketched in
Fig. 5.30.
From the horizontal equilibrium of forces, we have
x2
M(x) = 500 · kp · m, ∀x ∈ [0, 2].
2
500 kp/m
M (x)
N (x)
B
V (x)
x
5.7 Vierendeel Beams 125
To obtain the internal forces in the bar D − E we perform again a virtual cut as
depicted in Fig. 5.31. By equilibrating the forces in the horizontal directions, we
obtain:
V (x)
E
N (x)
M (x)
x
x2
M(x) = 500 · kp · m, ∀x ∈ [0, 2].
2
The resulting internal-force diagrams are depicted in Figs. 5.32, 5.33 and 5.34.
A 2000 kp
C
2m
0 kp 2000 kp
B 0 kp
2m D 2m E
Fig. 5.32 Resulting axial internal-force diagram for the structure in Example 5.2
126 5 Beams (I): Geometric Description and Internal Forces
A
2m C 0 kp
1000 kp
0 kp
B 0 kp
D E
1000 kp
2m 2m
A
C 0 kp · m
2m
1000 kp · m
0 kp · m
B 0 kp · m
2m D 2m E
Exercise
q
A
B C
Chapter 6
Beams (II). Normal Stresses
6.1 Introduction
We have repeatedly insisted throughout the book that the classical analysis of struc-
tures involves the determination of intermediate variables not initially present in the
Navier formulation of the elasticity problem. These variables are, of course, internal
forces and moments. They are appealing since they can be determined in many cases
from equilibrium alone and, notably, they simplify the subsequent determination of
stresses, deformations and displacements, the ultimate goal of our analysis.
However, as in previous chapters, the reader will notice how, to determine normal
stresses—remember, the σx component in our stress tensor σ—some simplifications
or hypothesis need to be taken into account. These will be the objective of this chapter.
The Euler–Bernoulli–Navier1 theory is the most useful and practical theory for long
beams. By “long” we mean bars whose length is more than, say, fifteen times the
average size of the cross-sectional area.
1 Named in honor of Jacob Bernoulli (1654–1705, who first discovered that the curvature of an
elastic beam at any point is proportional to the bending moment at that point), his nephew Daniel
Bernoulli (1700–1782, the first to formulate the differential equation of motion of a vibrating beam),
Leonhard Euler (1707–1783, who later validated their theory) and, of course, Claude-Louis Navier
(1785–1836), who first established the zero line of normal stress in 1819 and established in 1826
the elastic modulus as a property of materials independent of the second moment of area.
© Springer International Publishing AG, part of Springer Nature 2018 127
E. Cueto and D. González, An Introduction to Structural Mechanics for Architects,
Structural Integrity 4, https://doi.org/10.1007/978-3-319-72935-0_6
128 6 Beams (II). Normal Stresses
Coming back to the geometrical description of a bar—recall Fig. 5.4—, we have, see
Fig. 6.1, two main hypotheses.
1. A first kinematic hypothesis: every cross section moves as if it were a rigid solid.
2. A second kinematic hypothesis: cross sections initially perpendicular to the
beam’s axis (such as A(t) in Fig. 6.1) remain perpendicular after bending.
3. A static hypothesis: normal stresses σ y and σz , as well as shear stress τ yz are
negligible.
Hypotheses 1 and 2 are easily verified experimentally. See Fig. 6.2 for an example
in which the rigid solid-like movement of flat cross sections of the beam is clear.
Hypothesis 1 is equivalent to neglecting ε y , εz and γ yz components of the strain
tensor, meaning that there is no distortion of the cross section. For the physical
interpretation of these strain terms, the reader is referred to Fig. 1.4. This is also
equivalent to neglecting the Poisson effect in the section [9]. In other words, in spite
of the simultaneous existence of traction and compression stresses in the section—
as we will readily discover—the effect of the Poisson coefficient of contraction
and expansion, respectively, along perpendicular directions, is also negligible. Even
simpler: a rectangular cross section will always remain rectangular and of the same
dimensions. Remember always that we accept the hypothesis of small strains!
Hypothesis 3 is, surprisingly, inconsistent with hypothesis 1—recall Eq. (1.9)—
but it can be experimentally measured that these stresses are at least two orders of
magnitude lower than σx , and therefore they can be neglected if the bars show a
moderate radius of curvature.
A(t)
y
x
6.2 Euler–Bernoulli–Navier Bending Model 129
Fig. 6.2 Experimental verification of the kinematic hypothesis of the Euler–Bernoulli–Navier beam
theory. Notice how the black, straight thick lines remain straight after the application of the load
6.2.2 Kinematics
Let us translate the just introduced hypothesis into equations, useful for the design
and analysis of structures subjected to bending. We restrict ourselves to straight
bars of constant cross section. Other, more elaborate cases, can easily be treated
numerically by means of, for instance, finite element methods [8].
The starting point for our developments will be hypothesis 1, i.e., that every cross
section moves under bending as a rigid solid. It is a common practice, as already
mentioned, to take the center of gravity of the section as the origin of our system
of coordinates, see Fig. 6.1. The line joining every section’s center of gravity will
therefore define our x axis in a natural way. Under this rationale, the movement of
any point in the beam will follow, under bending, the following expression:
u = uo − r × θ, (6.1)
Here, uo is the vector of displacements for the center of gravity of the section, the
point (x, 0, 0)). In turn, θ(x) is the rotation vector of the section.
For the time being, we are considering bending alone, so that no torsion phenom-
ena exists. This implies that we will restrict ourselves to the case in which θx = 0.
Navier hypothesis (number 2) states that the rotation of the cross section is equal
to the rotation of the beam’s axis. Under the small displacement assumption, accepted
throughout this book, this angle of rotation will be approximated by its tangent and
the tangent as the slope of the beam’s axis. In other words:
du oz du oy
θy = − , θz = .
dx dx
Let us do some gymnastics at this moment. Consider again Eq. (6.1), and introduce
it into the expression of the Cauchy strain tensor, Eq. (1.3). This gives:
du ox dθ y dθz
εx = +z −y ,
dx dx dx
o
1 du y
εx y = − θz ,
2 dx
1 du oz
εx z = + θy ,
2 dx
ε y = εz = ε yz = 0.
du oz du oy
= −θ y , = θz ,
dx dx
we arrive at:
εx = εn + zκ y − yκz , (6.2)
εx y = 0,
εx z = 0,
ε y = εz = ε yz = 0,
with
du ox dθ y d 2uo dθz d 2 u oy
εn = , κy = = − 2z , κz = = ,
dx dx dx dx dx2
d 2 (y+u )
y y d2u 2
κz = dx2
= dx2 ∼ d uy ,
3 =
2 23
2 2 dx2
d(y+u y ) du y
1+ dx
1 + dx
with κz = 1/ρz the curvature of the deformed beam’s axis and ρz its radius of cur-
vature. Remember that we restrict ourselves to straight bars.
6.2.3 Stresses
To obtain the expression for the stresses in the bars, let us introduce in Eq. (6.2) the
constitutive equations for an elastic solid, i.e., Eq. (1.9).
The stress state in the bar can thus be considered as given by the simple one-
dimensional state:
σx = Eεx = Eεn − E yκz + E zκ y , (6.3)
with τx y = τx z = 0.
Again, a contradiction appears, since it is well-known that shear stresses are not
vanishing in the cross section. It is thus necessary to introduce in the model an
additional static hypothesis (the fourth one) to make the model consistent with the
experimental facts. This will be done in the next chapter, the one in which we will
study the form of shear stresses for beams under bending processes.
The final objective of this chapter is to obtain a convenient way to establish the
relationship between internal forces and stresses. Otherwise, all the work developed
so far would have been useless, and the establishment of internal forces as a new
variable in the problem is nothing but an exercise of style with no practical relevance.
Thus, coming back to the definition of internal forces in Sect. 5.3, we have,
132 6 Beams (II). Normal Stresses
N = σx d A = (Eεn − E yκz + E zκ y )d A
A A
= E εn (x)A(x) + κ y (x)S y (x) − κz (x)Sz (x)) ,
where A(x) represents the area of the cross section at a distance x along the beam’s
axis, and S y , Sz the first area moments of this cross section, i.e.,
Sy = zd A, Sz = yd A.
A A
But, since we have chosen the center of gravity of the section as the origin of coor-
dinates, these vanish:
Sy = zd A = Sz = yd A = 0,
A A
so that
du ox
N = E Aεn = E A , (6.4)
dx
or, equivalently,
N
εn = ,
AE
where AE, the product of the area of the cross section times the Young’s modulus
of the material is often referred to as the axial stiffness of the bar.
Following a similar procedure, we can obtain the expression of the internal bend-
ing moments in terms of the strain field, viz.:
M y (x) = zσx d A = E εn (x)S y (x) + κ y (x)I y (x) − κz (x)I yz (x) ,
A
Mz (x) = − yσx d A = −E εn (x)Sz (x) + κ y (x)I yz (x) − κz (x)Iz (x) ,
A
where, this time, I y , I yz , and Iz represent the second area moments of the cross
section,
Iy = z 2 d A, Iz = y 2 d A, I yz = yzd A,
A A A
and again
Sy = zd A = Sz = yd A = 0,
A A
6.2 Euler–Bernoulli–Navier Bending Model 133
since we have established our origin of coordinates at the section’s center of gravity.
Therefore,
d 2 u oz d 2 u oy
M y (x) = E κ y (x)I y − κz (x)I yz = −E I y + I yz , (6.5)
dx2 dx2
d 2 u oz d 2 u oy
Mz (x) = −E κ y (x)I yz − κz (x)Iz = E I yz + Iz , (6.6)
dx2 dx2
1 M y Iz + Mz I yz 1 Mz I y + M y I yz
κy = , κz = .
E I y Iz − I yz
2 E I y Iz − I yz
2
If we extend the previous procedure for the expression of shear internal forces,
we arrive at:
Vy = τx y d A = 0,
A
and, of course,
Vz = τx z d A = 0.
A
Obviously, this is consistent with a general case of loading such as the one
in Fig. 5.13, for instance. This motivates again the correction of the model that
will be done in the next chapter. Remember the famous aphorism by George
Box: “all models are wrong, but some are useful” (Cf. Box, G. E. P. (1976),
Science and Statistics, Journal of the American Statistical Association, 71:
791–799). This model constitutes a paradigmatic example of this assertion.
Equations (6.5) and (6.6) represent an example of constitutive equations for the
Euler–Bernoulli–Navier model. Substituted into the equilibrium equations give rise
to:
d 2 u ox
EA + px = 0
dx2
d 4 θz d 4θy dm z
E Iz 4 + E I yz + + py = 0
dx dx4 dx
d 4θy d 4 θz dm y
E Iy 4
+ E I yz 4 − + pz = 0, (6.7)
dx dx dx
134 6 Beams (II). Normal Stresses
1 M y Iz + Mz I yz 1 Mz I y + M y I yz
κy = , κz = . (6.8)
E I y Iz − I yz
2 E I y Iz − I yz
2
N M y Iz − Mz I yz Mz I y + M y I yz
σx = − z+ y,
A I y Iz − I yz
2 I y Iz − I yz
2
which relates the normal stress σx and the internal forces in the beam. This is,
therefore, the final objective of our quest and the true interest in defining internal
forces as a variable in the problem.
N My Mz
σx = − z+ y.
A Iy Iz
If, in addition, the structure lies in a plane, and so does its loading, then the
situation becomes even simpler, so as to give:
N My
σx = − z.
A Iy
In this last case, with solely one moment Mα , (α = y/z), the maximum value for
the normal stress and therefore the critical point of the section (at least in absence
of other stresses) will happen at the point farthest from the center of gravity of the
section, i.e., for maximal β, with β = z/y, so as to give:
Mα(α=y/z) Mα(α=y/z)
σxmax = β max (β = z/y) = ,
Iα(α=y/z) Wα(α=y/z)
6.2 Euler–Bernoulli–Navier Bending Model 135
Fig. 6.3 An example of plane frames built with doubly-symmetric profiles. Fish market in Luanco,
Spain
y
1 2 100 kp
3 cm
3 300 kp z
10 cm
4m 2m
Iα(α=y/z)
Wα(α=y/z) = ,
|β (β = z/y)|
max
known as the elastic section modulus. This modulus, along with the section area,
first and second area moments, are usually included in tables by steel-profile manu-
facturers so as to simplify the work of structural designers.
Let us solve a simple example of structural design. Consider the beam in Fig. 6.4,
with rectangular section of dimensions 10 × 3 cm2 .
136 6 Beams (II). Normal Stresses
Let us determine the location of the greatest stress in the structure (and therefore
the point at which the failure will begin, according to the classical elastic design
of structures). As we did previously, we begin by finding the reaction forces at the
supports in points 1 and 2:
Fx = 0 ⇒ Rx1 − 300 = 0 ⇒ Rx1 = 300 kp
Fy = 0 ⇒ R y1 + R y2 − 100 = 0 R y1 = −50 kp
⇒
Mz = 0 ⇒ 4 · R y2 − 6 · 100 = 0 R y2 = 150 kp
Once determined, these reactions forces will help to find the expression of the
internal force diagrams. The axial internal force will be constant throughout the bar,
since there is no applied force per unit length:
Vy (x = 0) = 50 kp ⇒ V1 = 50 kp,
Graphically, these internal forces look as depicted in Figs. 6.5, 6.6 and 6.7.
1 2 3
1 2 3
100 kp
6.2 Euler–Bernoulli–Navier Bending Model 137
The interest of internal-force diagrams comes from the fact that they make it
possible to determine, at a glance, the critical section in the structure. In this case, we
notice clearly that this point is point 2. It is subjected to N = 300 kp (in compression)
and M = 200 kpm. For the time being, we omit the effect of the shear forces, which
will be studied in the next chapter.
The geometric characteristics of the cross section of the bar are:
1 10 · 33
A = b · h = 10 · 3 = 30 cm2 , Iz = b · h3 = = 22.5 cm4 ,
12 12
so that the distribution of normal stresses will be:
N Mz −300 −20000
σx = − y= − y = −10 + 888.89y kp/cm2 .
A Iz 30 22.5
This means that the maximum stress in traction will take place along the upper fiber
of the beam:
σx,trac
max
= −10 + 888.89 · 1.5 = 1323.33 kp/cm 2 ,
while the maximum stress in compression will take place along the lower fiber of
the section, namely,
σx,comp
max
= −10 + 888.89 · (−1.5) = −1343.33 kp/cm 2 .
As can be noticed from Fig. 6.8, while the stresses produced by bending moments
alone (labelled as σx (M)) vanish at an y coordinate corresponding to that of the
center of gravity of the cross section—in other words, at the x-axis—the addition of
the stresses produced by the axial internal force has caused upward displacement of
the neutral axis.
138 6 Beams (II). Normal Stresses
y
−10 kp 1333.3 kpm 1323.3 kpm
3 cm
σx (N ) σx (M ) σx (N + M ) −1343.3 kpm
Fig. 6.8 Normal stresses at the critical section of the structure of Fig. 6.4 (side view). Diagrams are
not to scale. Note that the sum of the stresses produced by the axial internal force plus the contribution
of the bending moment gives rise to the actual stress distribution in the section (rightmost diagram
in the figure)
The neutral axis is concept of utmost importance and is defined as the locus
of the points of the cross section that have no normal stress:
N M y Iz − Mz I yz Mz I y + M y I yz
− z+ y = 0.
A I y Iz − I yz
2 I y Iz − I yz
2
It can be noticed that this axis passes through the center of gravity of the
section if there is no axial internal force acting on it.
The neutral axis of the section drives the regions of tension and compression stresses.
Sometimes it is necessary to ensure that the section is exclusively subjected to com-
pressive stresses. This is the case, for instance, in foundations, where it is important
that no tension appears, otherwise detachment of the foundation from the soil can
occur. In concrete structures, a material that does not support tension, it is also impor-
tant to ensure compression, although this is rarely achieved in practical cases, leading
to the need for steel bar reinforcement.
Let us introduce the eccentricity of the bending moments as e y = −Mz /N , ez =
M y /N . Let us now substitute the axial force (applied at the section’s center of gravity)
and the (possibly) two bending moments by the statically equivalent axial force N
applied at a position (e y , ez ) in the section-centered coordinate system.
We can then put the expression of the neutral axis in terms of the eccentricities,
so that:
6.2 Euler–Bernoulli–Navier Bending Model 139
1 ez Iz + e y I yz e y I y − ez I yz
− z− y = 0,
A I y Iz − I yz
2 I y Iz − I yz
2
where i α is the radius of gyration of the section along the α-axis, defined as:
iα = Iα /A, α = y, z.
The problem thus reduces to finding the pairs (e y , ez ) that produce pure tension
or compression within the section.
To further explain this concept, let us determine the core of a rectangular section,
see Fig. 6.9.
To obtain the location of the core, it is enough to determine the eccentricity of
a load that produces the neutral axis to fall along each side of the rectangle. In this
case: ⎫
A = bh ⎪
⎬ Iy 1 3
12 b h b2
I y = 12 b h ⇒ i y = A = 1 bh 3 = 122
2
1 3
⎪ bh
bh 3 ⎭ i z2 = IAz = 12bh = 12
h
Iz = 12
1
Should the neutral axis be located along the bottom side of the rectangle, we have:
ey h ez ez = 0
1+ − + b2 z = 0 ∀y ⇒
h2 2 e y = h6
12 12
Equivalently, for the neutral axis located along the right vertical edge of the
rectangle:
b
140 6 Beams (II). Normal Stresses
h/3
h
z
ey ez b ez = b6
1+ y+ − = 0 ∀z ⇒
h2 b2 2 ey = 0
12 12
By applying symmetry, we obtain the shape of the core, depicted in Fig. 6.10.
In Eq. (4.2) we established the principle of virtual work for trusses. Now, we are in
the position of establishing a much more general expression that includes bending.
Even if the model is not yet complete, i.e., we have not yet defined the relationship
between shear stresses and their corresponding internal forces, these are well-known
to have little influence on the final displacement of the structure, if important bending
moments are present.
The most general expression for this principle is:
L
ψ
σi j εi j d V = dx (σx εψx + τx y γxψy + τx z γxψz )d A, (6.9)
0
V
Of course, this term, equivalent to the virtual work produced by internal forces, must
equal the work done by external forces, i.e.:
ψ ψ
ρbi u i d V + ti u i d A
V δV
⎡ ⎤ ⎡ ⎤
L L
⎣ ψ ψ ψ
= ρbi u i d A⎦ d x + ⎣ ti u i dl ⎦ d x + ti u i d A,
0 A 0 δA A f −Ai
where A f and Ai represent, respectively, the final and initial cross sections of the
bar, where possibly some punctual forces or moments could be applied. b represent
the body forces per unit volume and t the applied tractions on the boundary of the
beam.
142 6 Beams (II). Normal Stresses
Worked Examples
6.1. We consider bars whose cross section is depicted in Fig. 6.11, for the structure
in Example 5.2. Determine
(a) the critical section (i.e., the one that will fail the first) for the loads in Fig. 5.29,
(b) the normal stress distribution for that section,
(c) the neutral axis of that section.
30 cm
z
2 cm
20 cm
Solution 6.1. (a) Section D of bars B − D and D − E is the critical section of the
structure, as one can readily notice from the observation of the diagrams in the
previous chapter. At that section, we have:
Mz = −1000 kp · m,
and
Vy = −1000 kp (in bar D − E).
(b) Since the axial internal force vanishes, normal stresses take the form
Mz
σx = − · y.
Iz
1
Iz = 30 · 303 − 16 · 263 = 21565, 333 cm4 .
12
Thus,
6.3 Principle of Virtual Work for Bending Beams 143
100000 kp · cm
σ= · y = 4, 63 · y kp/cm2 .
21565, 33 cm4
(Fig. 6.12).
(c) The neutral axis is the locus of the points satisfying σx = 0, so that we have in
this case,
σx = 0 ⇔ y = 0 cm.
y
69.55 kp/cm2
30 cm
σx (Mz )
Fig. 6.12 Normal stresses at the critical section of the structure, point D
6.2. Determine the distribution of normal stresses for the section in Fig. 6.13, by
assuming that the following internal forces are acting on it:
hG
z
80 mm
6 mm
80 mm
144 6 Beams (II). Normal Stresses
Solution 6.2. As a first step, we determine the first and second area moments of the
section, A, I y y Iz , to be able to compute stresses properly.
• Area. It is easy to see that the area can be computed as a sum of rectangular areas:
• Center of mass. Prior to determining the value of the second area moments, it is
necessary to find the position of the axes or, in other words, the center of gravity of
the section. For the z coordinate, its position is straightforward, given the symmetry
of the section.
For the y coordinate, remember that the center of gravity is such that it vanishes
the first area moments of the section,
i
y · Ai
yG = G ,
Ai
where i is an index over the various rectangles composing the section. In this case,
there are only two rectangles. Taking moments from the upper edge of the section,
we have
80 · 9 · 4.5 + 71 · 6 · 44.5
yG = = 19.37 mm = 1.937 cm = h G ,
1146
which is the distance from the center of mass to the upper edge of the section.
• Second area moments. The second area moments of a section composed of the
addition of several simpler geometries is just the sum of the respective moments
with respect to the same axis. However, sometimes it is necessary to apply the
Steiner theorem so as to obtain this second area moment from the global axis of
the section and not the one of the simpler geometry (rectangle) itself:
1 1
Iy = 9 · 803 + 71 · 63 = 385278 mm4 = 38.52 cm4 ,
12 12
1 1
Iz = 80 · 93 + 80 · 9 · (19.37 − 4.5)2 + 6 · 713 + 6 · 71 · (44.5 − 19.37)2
12 12
= 4860 + 159204.16 + 178955.5 + 269026.19 = 612045.86 mm4
= 61.20 cm4 .
• Given the symmetry of the cross section, the cross moment vanishes, I yz = 0.
6.3 Principle of Virtual Work for Bending Beams 145
y y
8.726 −1582.4
z x
80 mm
−103.82
σx (My )
103.82 [kp/cm2 ]
x
With these values, we are in the position of computing the stress distribution:
N Mz My
σx = − ·y+ · z,
A Iz Iy
z = 31.47 · y − 0.33.
146 6 Beams (II). Normal Stresses
If y = 0 cm ⇒ z = −0.33 cm.
Compression
z L.N.
Traction
Chapter 7
Beams (III). Shear Stresses
7.1 Introduction
It has already been mentioned that the hypotheses giving rise to the Euler–Bernoulli–
Navier beam model have some inconsistencies. Particularly, they predict no shear
stresses, something that is well known to be false.
To better understand this apparent lack of coherence and to motivate the proper
way to solve it, let us consider a beam under non-uniform bending moments, see
Fig. 7.1. Let us isolate a portion of it. Of course, every possible portion should be in
equilibrium, as is well-known from the first chapter of this book.
The 1 − 2 portion of the beam is obviously in equilibrium as can be noticed by
glancing at to the stress distribution. However, apply this rationale again and imagine
isolating a portion 3 − 4, see Fig. 7.2.
In this case, it is evident that there is an out-of-balance stress distribution that
would eventually result in a force uniformly accelerating the 3 − 4 portion to the
right. There is no possible equilibrium unless we postulate some kind of shear stresses
(τxy ) as the ones depicted in Fig. 7.3. In this chapter, we will study how to solve these
inconsistencies in an elegant (and rigorous!) way. This is usually achieved by means
of the Collignon theory of shear stresses.
1 2 σx (M (1 )) σx (M (2 ))
2
1 x
Fig. 7.1 A beam subjected to non-uniform bending moments. We isolate the central portion
obtained by slicing along planes 1 − 2 and verify equilibrium in it
3 4
x
Fig. 7.2 The 1 − 2 portion of the beam is isolated again, and a new portion obtained after slicing
by the 3 − 4 plane is put under scrutiny
1 Édouard Charles Romain Collignon (1831–1913) was a French engineer who worked for the
Russian railways. You can consult his 1865 work “Théorie élémentaire des poutres droites:
ponts métalliques, pont américains, combles” in the internet archive https://archive.org/details/
thorielmentaire00collgoog.
2 Jourawski, D. J. “Sur le résistance dun corps prismatique et dúne Piece Composée en Bois ou on
Tôle de Fer à une Force Perpendiculaire à leur Longeur.” Annales des Ponts et Chaussées. Vol. 12.
1856.
7.2 Collignon Theory of Shear Stresses in Beams 149
7.2.1 Hypothesis
Let us derive, in the plane bending case (one bending moment, say Mz ), the expression
for these shear stresses. For a more general case in which both Mz and My act in the
cross section, a simple superposition of their effects will be enough under the linearity
assumptions established so far.
Consider again a differential slice of a beam of length dx, as depicted in Fig. 7.5.
As a consequence of the applied load py , the bending moment evolves from Mz at
the left section to Mz + dMz at the right one. Equivalently, the shear internal force
changes from Vy to Vy + d Vy . Simple equilibrium considerations allow us to obtain
the following very useful equivalences:
Fy = 0 → Vy + d Vy − Vy + py dx = 0,
150 7 Beams (III). Shear Stresses
Mz Mz + dMz
Vy
dx
in other words,
d Vy
= −py ,
dx
and, similarly,
dx
M(x=0) = 0 → Mz + dMz − Mz + py dx + (Vy + d Vy )dx = 0.
2
By neglecting higher-order derivatives, we obtain,
dMz
= −Vy .
dx
It is important to clearly understand what these formulas mean. The first one,
d Vy
= −py ,
dx
states that the presence of a vertical load causes the shear internal force to
vary along the beam (i.e., along x). Otherwise it would stay constant along the
beam. The second one,
dMz
= −Vy ,
dx
states that the shear internal force is responsible of the evolution of the bending
moment. If no shear internal force appears in the beam, the bending moment
remains constant.
7.2 Collignon Theory of Shear Stresses in Beams 151
Let us consider now the equilibrium of a portion of the beam above the center of
gravity, see Fig. 7.6.
If we verify equilibrium in the blue-dashed region, we have, see Fig. 7.7,
F= σx dA.
A
Mz
σx = − y,
Iz
y y
py
A
x z
dx b(y)
Fig. 7.6 Equilibrium of a differential portion dx of a circular cross-section beam under transverse
loading py per unit length (positive along the positive direction of the y-axis). In this case, only a
portion above the center of gravity, at a generic position y, is equilibrated. The cross-sectional area
is A, while the area of the dashed portion is A
py
V V + dV
F F + dF
R = τxy bdx
dx x
Fig. 7.7 Equivalent forces due to shear stresses. Note the presence of a resultant force R = τxy bdx.
Do not confuse V with Vy , the internal shear force in the section. V is the resultant of the shear
stresses on A , see Fig. 7.6. Conversely, F is not the axial internal force, but the result of integrating
σx on A
152 7 Beams (III). Shear Stresses
so that
Mz Mz
F =− ydA = − Sz ,
A Iz Iz
dF Sz dMz Sz
=− = Vy = τxy b(y),
dx Iz dx Iz
1 Sz
τxy = Vy .
b Iz
If, as a practical example of this theory, we calculate the shear stress distribution
on a rectangular section, like the one in Fig. 7.4, we have,
h/2 h/2
y2 b h2
Sz = ydA = b ydy = b = −y ,
2
A y 2 y 2 4
S (y)I −S (y)I S (y)I −S (y)I
τxy (y) = 1
Vz y Iy zIz −Iz2 yz + Vy z IyyIz −Iy2 yz ,
b(y)
yz yz
For a general case, in which both normal and shear stresses exist, recall the
definition of the equivalent, von Mises stress: Eq. (1.10),
σVM = σx2 + 3τxy
2.
If the cross section of the beam is thin-walled, see Fig. 7.10, it is commonly accepted
that we can assume the arc-length parameter, s, parameterizes the shear stress.
z
154 7 Beams (III). Shear Stresses
In this case, we assume that, given the small thickness value, s, shear stresses are
approximately constant at each s-coordinate, so that:
where Sy/z (s, s0 ) represent the first area moment with respect to y or z, respectively,
between s0 and s. s0 is usually taken as a free end of the cross section, so that
Sy/z (s0 ) = 0, see Fig. 7.11.
In this formula, it is worth noting that the minus sign comes from the fact that s0
takes its minimum value where y is maximum and vice versa. For doubly-symmetric
cross section profiles, this formula reduces to the more familiar expression:
−1 Sy (s, s0 ) Sz (s, s0 )
τxs (s) = Vz + Vy .
e(s) Iy Iz
Given the results above, it is common to define a new magnitude coined as shear
flow as
q = τxs e(s),
such that:
z s2
The value of the shear flow at any point can easily be determined for its value at
another point, see Fig. 7.12, given that:
s2 s1 s2
Sy2 = zeds = zeds + zeds = Sy1 + Sy1→2 ,
s0 s0 s1
such that:
Sy1→2 = Sy (s2 , s1 )
The shear center is the point of the cross section at which an internal shear force could
be applied without causing torsional deformation—twisting—of the cross section.
This concept is particularly important in C-shaped profiles which are, for instance,
widely employed in lightweight construction. The shear center does not coincide, in
general, with the center of gravity of the cross section, see Fig. 7.14.
The application of a shear force out of the shear center is equivalent to a shear
force at the shear center plus a twisting moment
156 7 Beams (III). Shear Stresses
z
d
Center of gravity
Shear center
Mx = V · d .
In some cases, the determination of the position of the shear center is straightfor-
ward. For instance, for cross sections formed by rectangular segments intersection
at a given point, this is precisely the shear center, see Fig. 7.15. For a cross section
possessing symmetry axes, the shear center must lie within these axis. Therefore, is
the profile is doubly-symmetric, the position of the shear center coincides with the
origin of coordinates or center of gravity, see Fig. 7.16.
7.4 Deep Beams. Timoshenko Beam Theory 157
Deep beams—those with a span-to-depth ratio lower than, say, three— are very often
employed to construct walls, such as in precast concrete structures, for instance, see
Fig. 7.17. In this case, the hypothesis of “long” beams, established for the Euler–
Bernoulli–Navier theory, is not satisfied. Of particular importance is the fact that flat
cross sections do not stay perpendicular to the beam’s line of centroids, nor stay flat.
To understand this assertion, try this simple exercise: take a thick book and hold it
tight, so that the friction between pages allows us to consider it as “continuous”,
Fig. 7.18 An intuitive example of shear deformation of the cross section. The dashed line represents
the line of centroids, x, while the thick solid line represents the cross section of the “beam”. The
effect of the shear force on the section makes the section deviate from its initial perpendicular
position
rather than a mere juxtaposition of paper sheets. Then, bend it as in Fig. 7.18. You
will readily notice how bending provokes the cross section to abandon its initial
perpendicular position to the (dashed) line of “centroids”.
These structures are particularly efficient when the walls span several stories. In
Fig. 7.21, an example is depicted of a deep beam with several openings in which their
positioning has been optimized by means of the lower bound theorem.
Stepan Timoshenko3 established the theory that carries his name by adding the cross
section the possibility to rotate around the line of centroids due to shear stresses
(Fig. 7.22).
Therefore, the main contribution of Timoshenko’s work is the addition of a rotation
degree of freedom θz , independent of uy (respectively, a rotation θy , independent of
uz ). We enumerate the hypothesis necessary to define Timoshenko’s theory:
1. A first kinematic hypothesis: every cross section moves as if it were a rigid solid.
3 Stepan Prokopovych Tymoshenko (1878–1972) was a Ukrainian engineer who moved to the U.S.
He joined the University of Michigan in 1927 and Stanford in 1936. His theories are, therefore,
relatively recent. His books on the theory of elasticity have been translated into some 36 languages.
160 7 Beams (III). Shear Stresses
Slabs
with uy0 , uz0 the displacements at the cross section’s center of gravity.
As a consequence of the above hypothesis, the strain component γxy does no
longer vanish, i.e.,
duy0
γxy = − θz (x) = βy (x).
dx
This has a direct consequence, in turn, over the value of the shear stress, that would
be
τxy = Gβy (x).
so that
τxy
2
dA = Vy βy .
A G
with
Acy = κy A.
κy = 1
fcy
is known as the Timoshenko shear coefficient, and
A Sz2
fcy = dA.
Iz2 A b2
We thus arrive at the following constitutive equations for the Timoshenko theory:
162 7 Beams (III). Shear Stresses
Vy
βy = ,
GAcy
and
d θz Mz
= .
dx EIz
These constitutive equations give rise to the following model for the Timoshenko
beam:
d duy0
GAcy − θz + py = 0, (7.1)
dx dx
where py is the load per unit length pointing along the y direction, and
d d θz
EIz + Vy + mz = 0,
dx dx
or, equivalently,
d d θz duy0
EIz + GAcy − θz + mz = 0, (7.2)
dx dx dx
where mz is the bending moment per unit length, a type of loading found very rarely
(think, for instance, of the possibility of heating a beam on one side only, so that
it suffers dilation along the heated side, while the others remains at the reference
temperature).
Remember that analogue equations exist for the xz plane.
7.4 Deep Beams. Timoshenko Beam Theory 163
Worked Examples
7.1. Determine, for the critical cross section of Example 5.1, the distribution of shear
stresses, by following the just-developed theory of thin-walled profiles. We consider
a shear force Vy = −100 kp acting on the section, which is represented again in
Fig. 7.23.
30 cm
z
2 cm
20 cm
Solution 7.1. The distribution of shear stresses is given, as just developed, by:
q
τxs = ,
e(s)
where
−Vy
q= Sz (s).
Iz
We begin by computing the first area moment of the section between the points
A − B in Fig. 7.24,
s s s
SzA→B = Sz (A; B) = ydA = e · yds = 2 · 14ds = 28 · s,
0 0 0
Vy A→B 100
qA→B = qA − · Sz =0+ · 28 · s = 0.13 · s kp/cm,
Iz 21565.333
B A
C
z
Vy = −100 kp
we proceed in an analogous way for the vertical segment of the cross section, between
B − C:
s s s
SzB→C = Sz (B; C) = ydA = e · yds = 2 · (13 − s)ds = 26 · s − s2 .
0 0 0
Therefore,
Vy B→C 100
qB→C = qB − · Sz = 1, 3 + · (26 · s − s2 )
Iz 21565.333
= 1.3 + 0.12056 · s − 0.004637 · s2 kp/cm,
0.65 y
0.65 0.65
B A
0.65
C
1.042 z 1.042
0.65
0.65 0.65
[kp/cm2 ]
0.65
7.2. For the critical section of Example 6.2, compute the shear-stress distribution
predicted by assuming a thin-walled profile. We consider a shear force Vy = 300 kp
acting on the profile of Fig. 7.26.
hG
z
80 mm
6 mm
80 mm
166 7 Beams (III). Shear Stresses
E
A B
z Vy = 300 kp
C
q
τxs = ,
e(s)
where
−Vy
q= Sz (s).
Iz
Therefore:
Vy A→B 300
qA→B = qA − · Sz =0− · 1.34 · s = −6.56 · s kp/cm.
Iz 61.21
Remember that the balance of shear flows at the joints of rectangular segments of
the section, joint joint
q in = q out ,
7.4 Deep Beams. Timoshenko Beam Theory 167
B
such that, qBD = qAB
B
+ qEB
B
= 2 · (−26.24) = −52.48 kp/cm.
Proceeding in an analogous way between B − D:
s s s
SzB→D = Sz (B; D) = ydA = e · yds = 6 · (10.37 − s)ds
0 0 0
= 62.22 · s − 3 · s2 mm3 = 0.62 · s − 0.3 · s2 cm3 .
Thus:
Vy B→C
qB→D = qB − · Sz
Iz
300
= −52.48 − · (0.62 · s − 0.3s2 ) kp/cm.
61.20
In this case, the resulting distribution is quadratic. Particularizing at the beginning
and the end of the segment, we have, given that e = 6 mm = 0.6 cm,
The exercise ends by plotting the resulting distribution on the section, see
Fig. 7.28.
29.15
A
E
87.46 B 29.15
z
C
90.10
[kp/cm2 ]
0 D
Chapter 8
Beams (IV). Torsion
8.1 Introduction
Torsion occurs rarely in buildings. However, there are cases in which it may occur, and
therefore those affected elements must be carefully designed to bear these internal
moments. Torsion occurs, for instance, in spandrel beams that support beams at points
other than the pillars, see Fig. 8.1 or in floor beams firmly welded to girders, in the
case of steel structures, for instance.
A careful analysis of the internal forces and moments in this structure reveals that
bending moments in the b − e beam give rise to a different type of internal moment
in a − b − c and d − e − f , see Fig. 8.2.
In this chapter we will introduce the most common models for torsion, namely,
the Saint Venant and Prandtl models.
In the development of the Saint Venant model, we will assume the following kine-
matic hypothesis:
1. Each cross section rotates as a rigid solid (see Fig. 8.3).
2. The rate of twist, i.e., the angle rotated by the bar per unit length, κx , is constant.
3. Cross sections are free to warp in the x direction (along the longitudinal axis of
the bar), but this warping is the same for all cross sections, see Fig. 8.4. In other
words: u x = u x (y, z).
c
f
Fig. 8.1 An example of beams suffering torsion. Beams a − b − c and c − d − e are subjected to
torsion due to the bending moment at b − e
c
f
Fig. 8.2 Internal moments in the structure of Fig. 8.1. Red lines in the diagrams represent torsion
moments, while blue lines represent bending moments. Pinned supports are here assumed to allow
bending but to resist torsion (no free rotation around the local x-axis of each bar)
ux
ux ux
Fig. 8.4 Warping of a cross section. This warping movement causes the cross section to loose its
initial flatness. According to Saint Venant, this movement must be superimposed on the rigid-solid
one in Fig. 8.3 to satisfy the kinematic assumptions of his model
u x = u x (y, z),
u y = −θx (x)(z − z T ),
u z = θx (x)(y − yT ).
Once we have postulated the form of the displacement field, by resorting to the
definition of Cauchy’s strain tensor, see Eq. (1.3), we have:
εx = 0,
ε y = 0,
εz = 0,
∂u x dθx
γx y = − z,
∂y dx
172 8 Beams (IV). Torsion
(y, z)
u(y, z)
Mx
Center of torsion (yT , zT )
z
Fig. 8.5 Rotation of an arbitrary point around the center of torsion, according to hypothesis #1
∂u x dθx
γx z = + y,
∂z dx
γx z = −θx + θx = 0.
Note that, since warping is assumed to be identical for every section, no normal
stress will be produced. This has been also validated experimentally for regions far
from supports imposing restrictions to warping, as mentioned before. By resorting
to constitutive equations, Eq. (1.7), we have:
σx = 0,
σ y = 0,
σz = 0,
τ yz = 0,
∂u x dθx
τx y = G − z ,
∂y dx
∂u x dθx
τx z =G + y .
∂z dx
Where the aforementioned hypothesis could not be verified, the so-called theory of
non-uniform torsion must be applied. This theory falls beyond the scope of this book
and will not be detailed herein. In practice, a detailed three-dimensional analysis
should be preferred, probably with the help of finite element software, for instance.
If we substitute the just-obtained stresses into the equilibrium equations (1.6), we
arrive at:
8.2 Saint Venant Model 173
∂σx ∂τ yx ∂τx z
+ + = 0, → ∇ 2 u x = 0,
∂x ∂y ∂z
2
∂τx y ∂σ y ∂τ yz ∂ θx ∂ 2 θx
+ + = 0, → −G z =0 → (x) = 0, ∀z,
∂x ∂y ∂z ∂x 2 ∂x 2
2
∂τx z ∂τ yz ∂σz ∂ θx ∂ 2 θx
+ + = 0, → −G y =0 → (x) = 0, ∀y.
∂x ∂y ∂z ∂x 2 ∂x 2
These last two expression prove that the rotation of the cross section is a linear
function of x, the longitudinal coordinate of the bar. In turn, the first one imposes u x
to be an harmonic function of x. Therefore, a suitable hypothesis for the form of θx
could be
θx = κx x + B,
such that
dθx
κx = = constant.
dx
In these circumstances, it is a common strategy to assume that the warping takes
the form
u x = θx · φ(y, z).
∇ 2 φ(y, z) = 0.
This is equivalent to assume that the stress tensor takes the form:
⎛ ⎞
∂φ
0 Gθx ∂y
− z Gθx ∂φ
∂z
+y
⎜ ⎟
⎜ ⎟
σ = ⎜ Gθx ∂φ − z 0 0 ⎟ (8.1)
⎝ ∂y ⎠
Gθx ∂φ
∂z
+y 0 0
In other words, along the lateral surface of the bar we must enforce:
∂φ ∂φ
t = σ · n = Gθx − z ny + + y n z = 0,
∂y ∂z
thus giving
dφ
= zn y − yn z .
dn
We follow a similar procedure for the right and left surfaces in Fig. 8.6 so as to
obtain:
Vy = τx y d A = 0,
A
Vz = τx z d A = 0.
A
These expressions are satisfied identically due to the harmonic character of φ(y, z).
However, a more interesting result is obtained after imposing the internal equilibrium
of stresses with the externally imposed twisting moment:
∂φ ∂φ
MT = τx y z − τx z y d A = Gθx y +y −z −z dy dz,
A A ∂z ∂y
which is equivalent to
∂φ ∂φ
MT = Gθx y −z + y 2 + z 2 dy dz = Gθ J,
∂z ∂y
A
J
8.2 Saint Venant Model 175
where J is the so-called torsion constant. It can be readily noticed that J depends
only on the shape of the cross section. The product G J is known as the torsional
stiffness of the section. This comes from the fact that:
MT
θx = . (8.2)
GJ
The general procedure to solve the Saint Venant torsion model for a given bar
thus comprises the following steps:
1. Obtain the warping function by solving the Laplace problem
∇ 2 φ(y, z) = 0 in A,
dφ
dn
= zn y − yn z on ∂ A.
u x = u x (y, z),
u y = −θx (x)(z − z T ),
u z = θx (x)(y − yT ),
and
∂φ
τx y = Gθ −z ,
∂y
∂φ
τx z = Gθ +y .
∂z
One last ingredient in the Saint Venant theory is to express it in the form of
a boundary value problem (amenable to be solved by finite element methods, for
instance). This can be achieved by introducing the constitutive Eq. (8.2) arising from
Saint Venant theory into the differential equation of equilibrium of twisting moments,
so as to give:
176 8 Beams (IV). Torsion
∇ · σ = 0,
or, equivalently,
σi j, j = 0,
For the particular case of torsion, a simple choice is enough to satisfy these require-
ments. Indeed, check the validity of the following choice:
σx = σ y = σz = τ yz = 0,
∂ψ ∂ψ
τx y = G J , τx z = G J . (8.3)
∂z ∂y
1 Do not confuse the greek letter φ, i.e., phi, employed for the Saint Venant warping function, with
ψ, psi, Prandtl’s function.
8.3 Prandtl Model 177
∇ 2 ψ = constant.
By noting that Eqs. (8.1) and (8.3) represent the same magnitude, we have:
∂ψ ∂φ
= − z,
∂z ∂y
∂ψ ∂φ
=− +y ,
∂y ∂z
∇ 2 ψ = −2.
∂ψ ∂ψ
t = σ · n = Gθx ny + n z = 0,
∂z ∂y
or, equivalently
dψ
= 0.
ds
In other words,
ψ = constant
along the outer surface of the bar, ∂ A. s represents here the coordinate along this
surface. A simple yet practical solution consists in taking
ψ = 0 on Γ = ∂ A.
178 8 Beams (IV). Torsion
At the end sections of the bar, we impose the equilibrium of the externally applied
moments with the internal distribution of stresses, as we did in the previous section
for the Saint Venant theory when looking for the warping function:
∂ψ ∂ψ
MT = τx y z − τx z y d A = Gθx −y −z dy dz,
A A ∂y ∂z
so that
∂ψ ∂ψ
J =− y +z dy dz.
A ∂y ∂z
Equivalently,
∂ yψ ∂zψ
J =− + − 2ψ dy dz.
A ∂y ∂z
such that
MT = Gθx 2 ψd A
A
To obtain the Prandtl function, therefore, we must solve the following boundary-
value problem:
∇ 2 ψ = −2 in A
ψ = 0 on ∂ A.
∂ψ ∂ψ
τx y = Gθx , τx z = −Gθx .
∂z ∂y
Finally, the displacement field can be obtained by obtaining the strain field
from the stress field and then by integration.
Given the form of the Prandtl function stated here, it sometimes interpreted as a
sort of bubble over the section whose volume
ψd A
A
ψ(y, z) ≈ ψ(z).
If we thus try to solve the Prandtl problem under this new assumption,
∇ 2 ψ = −2 in A,
ψ = 0 on ∂ A.
∇ 2 ψ(z) = −2 in A, → ψ(z) = −z 2 + c1 z + c2 .
Fig. 8.7 Effect on the torsional stiffness of a closed (a) or open (b) thin-walled profile. You can
make a simple experiment by just taking the cardboard tube from a paper towel roll. Try to twist it
(without causing any local buckling), and you will see no perceptible rotation angle. On the contrary,
cut it longitudinally with a pair of scissors and notice the drastic decay of stiffness. In b you will
notice the warping of the section
8.3 Prandtl Model 181
such that
e2
c1 = 0, c2 = ,
4
thus giving
e2
ψ(z) = − z − 2
.
4
giving finally
1 3
J= be ,
3
with b the height of the section and e the (necessarily small) profile thickness. This
also opens the possibility of an easy determination of the rotation rate of the profile:
MT 3MT
θx = = .
GJ Gbe3
182 8 Beams (IV). Torsion
∂ψ ∂ψ
τx y = Gθx = −2Gθx z, τx z = −Gθx = 0.
∂z ∂y
with i the number of rectangular profiles employed to build the whole cross section,
see Fig. 8.10.
In this case, the angle rotated per unit length will be identical for every segment
of the profile (otherwise no compatibility could occur between them) and therefore,
MT i
θxi = ,
G Ji
8.3 Prandtl Model 183
y y
z z
such that
MT Ji
MT i = .
J
Since
It is therefore evident that the maximum stress in the thin-walled profile occurs,
surprisingly, where the thickness is maximal! A practical consequence of this fact is
that one should design profiles by juxtaposition of rectangular profiles of identical
thickness. Otherwise, yield will occur where the thickness is maximal.
For closed profiles, the stress distribution is known to follow different patterns. If
the thickness e is small, the stress distribution can be approximated to a reasonable
extent as being constant throughout the thickness, see Fig. 8.11.
If the thickness varies along the perimeter of the profile, from equilibrium con-
siderations it is evident that
τ1 e1 d x = τ2 e2 d x,
184 8 Beams (IV). Torsion
so that
τ1 e1 = τ2 e2 .
We can therefore define a sort of torsion stress flow, in an analogous way as we did
for shear stresses, such that
qT = τxs · e = constant,
where the coordinate s has been preferred, since in general the profile cross section
will not be oriented along the reference axis y, z. It is now evident that
Mt = qT · r ds.
Given that
1
qT · r ds = A,
2
we arrive at
Mt
qT = .
2A
We are now in the position of determining the point of maximum stress within
the section. To this end, it is enough to recall that
qT = (τxs )i · ei ,
8.3 Prandtl Model 185
so that
qT Mt
(τxs )i = = .
ei 2 Aei
Therefore, we find that the maximum stress will be at the point of minimum thickness
eimin , something more intuitive that those conclusions obtained for open, thin-walled
profiles in the previous section.
To conclude, the torsional rigidity can now be evaluated as
4 A2
J = ds ,
e
4 A2 e
J= ,
Stotal
with Stotal the length of the perimeter of the section, see Fig. 8.12.
Stotal
z
186 8 Beams (IV). Torsion
Worked Examples
8.1. Use Saint Venant’s model to study a circular cross section under torsion and to
compute its shear-stress distribution (Fig. 8.13).
MT
α
z
Solution 8.1. (a) Remember that circular sections do not present warping (i.e.,
φ(y, z) = 0) since the number of planes of symmetry is infinite. This symmetry
would be broken if warping appears.
In a Saint Venant context the following Laplace problem must to be solved to
obtain the warping function φ(y, z)
∇ 2 φ(y, z) = 0 in A,
dφ
= zn y − yn z on ∂ A.
dn
Therefore, for this example,
dφ
= zn y − yn z = R cos α sin α − R sin α cos α = 0.
dn
Summarizing,
∇ 2 φ(y, z) = 0 in A,
dφ
= 0 on ∂ A,
dn
In other words, for a circular cross section the torsion constant is the polar moment
of inertia, I0 .
(c) The rate of twist has been defined as
MT MT 2
θ= = .
GJ G π R4
This rate of twist depends on the inverse of the section radius to the power of four,
so to reduce the rate of twist (without any change in the material) it is enough to
increase the radius of the cross section.
(d) Finally, to obtain the stress field we recall that
∂φ
τx y = Gθ − z = −Gθz,
∂y
∂φ
τx z = Gθ + y = Gθy.
∂z
8.2. Use Prandtl model to study a elliptic cross section under torsion and compute
the resulting shear stress distribution (Fig. 8.15).
MT
Solution 8.2. When the function that describes the geometry of the boundary of
the cross section has a constant Laplacian, it is easy to obtain the Prandtl function
by scaling this function by a constant. For this particular case, the expression of an
ellipse satisfies this condition,
y2 z2
ψ= + 2 − 1 c,
a2 b
∇ 2 ψ = −2,
so that
a 2 b2 a 2 b2 y2 z2
c=− 2 ⇒ ψ = − + 2 −1 .
a + b2 a 2 + b2 a2 b
MT M
θ= = T .
GJ G2 A(y,z) ψd A
The warping function and the displacement field can be found by integrating the
shear stress distribution, i.e., the above equations. It is then clear that in this case we
will face a non-vanishing warping (Fig. 8.16).
τ
MT
Fig. 8.16 Shear stress distribution for a ellipsoidal cross section under torsion
8.3. Let us consider a bar, clamped in its initial and end sections A and B, respec-
tively. Let us consider two different sections along the bar. On the left, a solid circular
cross section with radius r1 = 50 mm is assumed. On the right part of the bar, a hollow
cross section with internal radius r2 = 40 mm, see Figs. 8.17 and 8.18, is considered.
Compute the distance d where we must apply a twisting moment MT to obtain the
same reaction forces MT (A) = MT (B) at both clamped supports.
MT
x
A B
625 mm 625 mm
Fig. 8.17 Fixed bar with non uniform cross section under torsion
190 8 Beams (IV). Torsion
y y
r1 r1
z z r2
Fig. 8.18 Sections for the fixed bar with non uniform cross section under torsion
A B
T T
d
Fig. 8.19 Resulting torsion diagram for the structure in Fig. 8.17
π 4
Isolid = 50 = 9817477.04 mm4 ,
2
π
Ihollow = (504 − 404 ) = 5796238.45 mm 4 .
2
The angle twisted between A and B must vanish thanks to the clamped supports, so,
625 d 1250
T T T
φ A→B = 0 = dx + dx + d x.
0 G · Isolid 625 G · Ihollow d G · Ihollow
Therefore,
Chapter 9
Statically Indeterminate Structures
Abstract This chapter introduces the methods for the analysis of statically indeter-
minate structures.
9.1 Introduction
Throughout this book, we have studied both statically determinate and indeterminate
structures. Particularly in Chap. 4, we treated this concept for trusses. In the most
recent ones, we have introduced very intuitive and powerful methods based on the
lower bound theorem. However, they do not provide us with a detailed solution for
the structures at hand.
At this point, it is time to study in a systematic way the various methods available
for the analysis of statically indeterminate structures. Since this might take a long
time, we have preferred to introduce only the most classical method for manual
analysis: the flexibility method (sometimes referred to as the method of consistent
deformations for reasons that will become apparent soon). Stiffness methods will
only be covered very briefly. For analysis with the help of a computer, stiffness
methods are always preferred, in particular the method of finite elements [8]. As you
will notice hereafter, flexibility methods are intuitive, but less systematic—always
some training will be welcomed by the analyst—while stiffness methods are very
systematic—and thus very suitable for computer implementation—but less intuitive.
In fact, one can readily lose track of the physical meaning of the resulting equations
with this type of methods.
But, whatever method you choose, the first step will always be to know in advance
if a structure is statically determinate or not.
To know to what extent a given structure is statically indeterminate and, most impor-
tantly, the degree of indeterminacy it presents, many authors employ simple formulas.
Thus, if a structure is statically determinate, it is just because we can find the value
of internal forces by resorting to equilibrium equations only.
Following the definition above, some simple formulas can help to ascertain the
degree of indeterminacy of a structure. For instance, in a three-dimensional by n bars
bars intersecting at n nodes nodes or joints, we will have three force equilibrium equa-
tions per node (3 × n nodes ) and three moment equilibrium equations (also 3 × n nodes ).
On the other hand, the number of unknowns is equal to six internal forces and
moments per bar (thus, 6 × n bars ), plus the number of reactions at the supports,
n reactions , which varies greatly with the design of the structure. The degree of inde-
terminacy (also known as the degree of hyperstaticity, doh) would be in this case:
where H represents the number of hinges present in the structure (since each hinge
can be considered as an additional equation of the type M = 0 or V = 0).
For two-dimensional structures, this reduces to:
since there are only three equilibrium equations per node and also three internal
forces (two forces and one moment, in fact) per bar.
These equations, though often useful, do not take into consideration additional
circumstances that can arise in a general situation, such as symmetry, for instance.
In a symmetric structure, symmetry provides an additional number of equations
by imposing the equality of internal forces on both sides of the symmetry plane.
The number of these additional equations strongly depends on the structure under
consideration.
If a structure possesses a doh > 0, it is said to be hyperstatic, while, if it verifies
doh = 0, it is said to be isostatic. Both are examples of stable structures, in the
sense that they can bear loads. However, if doh < 0, the structure is unstable and is
indeed a mechanism, unable to support any type of load without catastrophic failure.
Examples of such types of structures are given in Fig. 9.1.
The structure in Fig. 9.1b deserves some comments. Despite having do f = 0, it
is not a classical isostatic structure. This is so since, in its undeformed configuration,
9.2 Degree of Indeterminacy 193
P P
P P
P P
N N
N N
Fig. 9.2 Analysis of equilibrium in a critical structure. Equilibrium is possible only in a deformed
configuration
it can not achieve equilibrium. Since there is a hinge under the point of application
of the load, no bending moment can appear, and thus only axial internal forces could
occur, see Fig. 9.2 (left).
Therefore, a straightforward application of force equilibrium in the vertical direc-
tion would lead to
Fv = 0, → P = 0,
The number of equations in the problem to solve will always be the same. We always
have kinematic equations, cf. Eq. (1.3), equilibrium equations, see Eq. (1.6) and,
finally, constitutive equations, Eq. (1.7). These equations could eventually take some
simplified forms, if we accept any of the just presented models, but they always
constitute the same three families of equations. When a structure is statically deter-
minate (or isostatic), the only difference is that equilibrium equations are uncoupled
from the others, thus making it possible to find internal forces and external reactions
without invoking the rest of the equations.
Therefore, to fully solve the resulting problem, taking into account that consti-
tutive equations establish a relationship between static and kinematic variables, we
have only two options. Either we begin by selecting (internal or external) forces as
unknowns, or we begin by selecting displacements as unknowns. The former pos-
sibility gives rise to the family of flexibility (or compatibility) methods, while the
second leads to the family of stiffness (or equilibrium) methods.
F, σ = f (HU ).
ε, u = g(HU ).
9.3 Methods for the Analysis of Indeterminate Structures 195
Stiffness methods enforce the available equations the other way around:
1. Select an appropriate number of degrees of freedom (DoF) of the structure so
that they accurately govern the displacement of the (at least, theoretically) infi-
nite number of degrees of freedom of the structure, considered as a continuum
medium. Again, this operation depends strongly on the selected method in the
family of stiffness methods.
2. By invoking compatibility equations, obtain an expression for the kinematic vari-
ables as a function of DoF,
ε, u = h(DoF).
F, σ = (DoF).
4. Finally, by imposing equilibrium (hence the name of the method), obtain the value
of the selected DoF.
5. Eventually, come back to your previous expressions so as to obtain ε, u, F
and σ.
Stiffness methods are adequate for computer analysis of structures, since they are
less intuitive but very systematic. The finite element method is the classical example
of this family of methods, but also the slope-deflection method [15], for instance,
belongs to this family.
To easily understand the rationale behind flexibility methods, let us consider a simple
example, see Fig. 9.3. This structure will resent the reactions at the supports shown
in Fig. 9.4.
A simple counting of the number of reactions, bars, nodes and hinges will provide:
n bars = 1, n reactions = 4, n hinges = 0 and n nodes = 2, so that
196 9 Statically Indeterminate Structures
R2 R3
Therefore, the first step should consist of choosing the so-called “hyperstatic
unknown”. In general, our experience indicates that working with cantilever-like
structures is easy and very convenient, so we decide to take reaction R3 as the
“hyperstatic” one, and call it F hereafter. From now on, we proceed as if there were
no sliding support in the structure, by substituting its equivalent force, F for it. Note
that we treat F as a known, but it is not (yet). The problem to analyze will be,
therefore, the one in Fig. 9.5.
A simple static analysis of this isostatic equivalent structure reveals that
q2
R1 = q − F, M = F,
2
so that
( − x)2
M(x) = F( − x) − q .
2
R2 F
9.4 Flexibility or Compatibility Methods 197
The x-axis has been assumed to be centered at the clamped (left) support. These
constitute the equilibrium equations mentioned in Sect. 9.3.1, step number 2: F =
f (HU ).
By applying the principle of virtual work or any other available method, we can
notice that the displacement at the right end of the structure will be
1 F3 q4
u y (x = ) = − + .
EI 3 8
If you observe carefully the resulting equation, it is of the form u = g(HU ), i.e., a
constitutive equation.
The only remaining equation is therefore a compatibility equation. But, what does
a compatibility equation mean in this context? The term “compatibility” should be
understood here as a means of restoring the structure to its initial state. In other
words: in Fig. 9.3 there is a sliding support at the right end of the beam. In the
isostatic equivalent structure of Fig. 9.5, there is no sliding support, so, to enforce
the “compatibility”, with the initial design, we must enforce u y (x = ) = 0:
1 F3 q4
u y (x = ) = − + = 0,
EI 3 8
which provides
3q
F= .
8
This last compatibility enforcement provides the value of the hyperstatic unknown
and enables us, finally, to obtain the expressions for the reactions, internal forces and
moments in the structure.
A B
198 9 Statically Indeterminate Structures
B
MBB
(c)
Internal forces arising in the analysis of the structure in Fig. 9.6 can be readily
decomposed into three distinct contributions (and this is always the case, not only
for this simple problem!):
1. Internal forces and moments provoked by the external load q, if we assume that
the end nodes of the structure do not rotate, see Fig. 9.7a. These are called the
fixed-end moments and will be denoted by a “0” superscript.
2. Internal forces and moments caused by the rotation of node A alone, without any
external load nor rotation on node B.
3. Idem for node B if node A is assumed to be “clamped”.
Every moment follows a one-to-one relationship with the rotation produced at its
own node, so that,
MI I = k I I θI , I = A, B,
M I J = M I0J + k I I θ I + k I J θ J ∀ I, J = A, B.
9.5 Stiffness or Equilibrium Methods 199
4E I
kI I = ,
L
and
2E I
kI J = ,
L
so that
0 4E I 2E I
M AB M AB θ
= + L L
· A .
MB A M B0 A 2E I
L
4E I
L
θB
The slope deflection method1 imposes equilibrium directly at each node whose rota-
tion has been considered as a degree of freedom. For the structure in Fig. 9.6, we
have therefore, assumed that there is no bending moment in nodes A nor B,
0 4E I
0 M AB 2E I
θA
= + 2EL I L
· ,
0 M B0 A L
4E I
L
θB
1 The slope deflection method is the work of George A. Maney, who developed it in 1914 [15],
although some authors attribute it to the German professor Otto Mohr.
200 9 Statically Indeterminate Structures
pL 2 4E I
0
2E I
θA
= 12
+ 2EL I L
· ,
4E I
θB
2
0 − pL L L
12
where moments and angles have been considered positive if they act in the counter-
clockwise direction. The direct solution of the system gives:
pL 3
θA = − ,
24E I
pL 3
θB = .
24E I
We are now in the position of verifying that, for these values, it holds
M AB = M B A = 0.
Note that when there are chord rotations (in frames, for instance), additional
equilibrium equations must be established for these degrees of freedom. We will not
detail this method that itself deserves an entire book. Instead, we will briefly describe
others in the sections that follow.
The moment distribution method was developed by the North American engineer
Hardy Cross2 solves the equilibrium equations in an iterative way. In 1930, in the
absence of any computer, and for large systems of equations, it was considerably
easier to solve them by iterative methods rather than by a direct one, as in the slope-
deflection method. This is at the origin of the popularity of the method until the
appearance of the direct stiffness method and, later on, the finite element method.
If we come back to the constitutive equations, it is clear that they can be expressed
alternatively as:
0
M AB M AB kAA k AB
= + · θ A + · θB .
MB A M B0 A kB A kB B
The Cross method operates on moments rather than angles (though the differences
are minimal, since they are related by the stiffnesses, k I J . It is clear then:
0
M AB M AB MAA μB A · M A A
= + + ,
MB A M B0 A MB B μ AB · M B B
2 H. Cross proposed the moment distribution method in 1930 [6]. He was a professor at the University
of Illinois at Urbana-Champaign. His relevant article was reviewed by 38 referees and is most likely
the most reviewed paper ever! Reviewers comments comprised some 146 pages.
9.5 Stiffness or Equilibrium Methods 201
where
kI J
μJ I =
kI I
kB A 0
M B(1)A = M B0 A − M = M B0 A − μ AB M AB
0
.
k A A AB
(1) kAA 0
M AB = M AB
0
− M = 0,
k A A AB
−M B0 A − k B A θ(1) M (1)
θ(2)
B =
A
= − BA .
kB B kB B
rate of the method is fast, in practice, this method has been abandoned nowadays,
and it is studied solely from a historical perspective. The finite element method [8],
for instance, is preferred in industry and architecture as the ubiquitous method for the
computation of internal forces by employing a computer. A thorough analysis of the
finite element method is nevertheless beyond the scope of this book. The interested
reader is referred to the plethora of good books available on the subject.
The previous methods have been introduced in a matrix form, even if some of them
were developed long before matrices acquired popularity.3 As long as the number
of chord rotations is significant in a given structure, these methods become less and
less intuitive and amenable to a computer implementation.
On 6 November 1959, M.J. Turner, who was at that time the head of the Struc-
tural Dynamics Unit at Boeing, presented in a conference at Aachen, Germany, the
first paper on a new method coined the Direct Stiffness Method. Surprisingly, the
development of matrix methods of structural analysis originated in the aeroelastic
community, mainly in aviation companies. This conference paper is now lost, so that
the first paper on the topic is considered to be [20].
The direct stiffness method aims at developing a completely general technique,
in which the internal forces can be expressed as functions of the“element” degrees
of freedom. At this time, the word “element” should not yet be understood in the
“finite element” sense, but as a sort of “ingredient” of the structure, typically, a bar.
Therefore, the final objective is to find an expression of the type:
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
Ni k11 k12 k13 k14 k15 k16 ui
⎜ Vi ⎟ ⎜k21 k22 k23 k24 k25 k26 ⎟ ⎜ vi ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ Mi ⎟ ⎜k31 k32 k33 k34 k35 k36 ⎟ ⎜ θi ⎟
⎜ ⎟=⎜ ⎟ · ⎜ ⎟,
⎜ N j ⎟ ⎜k41 k42 k43 k44 k45 k46 ⎟ ⎜u j ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ V j ⎠ ⎝k51 k52 k53 k54 k55 k56 ⎠ ⎝v j ⎠
Mj k61 k62 k63 k64 k65 k66 θj
where i and j represent the initial and final nodes of the considered bar. Here, the
internal forces Ni , Vi and Mi are expressed in a completely general framework
as a function of the bar’s degrees of freedom, its axial displacement u i , its shear
displacement vi and its nodal rotations θi , with i = 1, 2, the initial and end nodes of
the bar.
At this moment, it is well-known that, for a bar of constant cross section A, second
area moment I , length L, and made of material with a Young’s modulus E, this matrix
is:
3 Professor
Carlos Felippa [7] attributes the invention of matrices to Cayley at Cambridge or to
Grassmann in Germany, around 1858, but their use was not common before 1930.
9.6 The Direct Stiffness Method 203
⎛ ⎞ ⎛ ⎞
⎛ ⎞ EA
L
0 0 − ELA 0 0 ui
Ni ⎜ ⎟ ⎜v ⎟
⎜ Vi ⎟ ⎜ 0
12E I
L3
− 6E
L2
I
0 − 12E
L3
I 6E I
L2 ⎟ ⎜ i⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ Mi ⎟ ⎜ 0 − 6E I 4E I
0 − 6E I 2E I ⎟ ⎜ θi ⎟
⎜ ⎟=⎜ L2 L L2 L ⎟ · ⎜ ⎟, (9.1)
⎜ N j ⎟ ⎜− E A 0 EA ⎟ ⎜u ⎟
⎜ ⎟ ⎜ L 0 0 0 ⎟ ⎜ j⎟
⎝ Vj ⎠ ⎜ L
⎟ ⎜ ⎟
⎝ 0 − 12E L3
I
− 6E
L2
I
0 12E I
L3
− 6E
L2
I ⎠ ⎝v j ⎠
Mj
0 6E I
L2
2E I
L
0 − 6E
L2
I 4E I
L
θj
P = K · δ.
or, equivalently,
P = L T · P.
The transpose sign is only due to a convention in which the transformation matrix L
is usually assumed to be the transpose of the one just introduced.
Therefore, if we consider one particular constitutive equation as
P = K · δ ,
L T P = K · L T δ,
from where
P = (L K L T )δ,
K = L K LT .
to express the relationship between each bar’s degrees of freedom δ and the struc-
ture’s degrees of freedom, u. This compatibility enforcement is nowadays known as
assembly of the structure’s stiffness matrix.
The direct stiffness method arises from a completely discrete origin, given by
Eq. (9.1). To obtain this equation, standard calculations are made, possibly by
employing flexibility methods. Soon after the post-World War II period, the com-
munity realized that the direct stiffness method can be obtained from a completely
continuum setting starting from the differential equation governing beam-bending
phenomena. Moreover, mathematicians soon realized that, in a 1943 paper, Courant
had employed a mesh of triangles to solve, starting from variational principles, vibra-
tion problems [5]. This readily opened the possibility to develop matrix analysis
methods with a sound physical basis. Nowadays, the resulting method, coined as the
finite element method, is the ubiquitous method in engineering analysis.
Without the aim to be completely exhaustive (a simple search on Amazon.com
for the term “finite elements” produced some 6,100 results on December 2017), we
will briefly describe the method here for completeness. The interested reader can
consult some very good introductory references such as [8], for instance.
Let us consider first, for the sake of simplicity, the differential equilibrium equation
of a truss, Eq. (4.1),
d du x
AE = − p.
dx dx
This equation constitutes what is commonly known as the strong form of the truss
problem, if completed with adequate boundary conditions. The weak form will be
obtained by first multiplying both sides of the equation by an arbitrary function w
(the so-called weight function), and integrating along the length of the bar, so that
l l
d du
w AE dx + wp d x = 0, ∀w.
0 dx dx 0
In order to simplify a little bit the previous expression, recall the form of the derivative
of a product of functions,
d df dw df d dw
(w f ) = w + f =⇒ w = (w f ) − f,
dx dx dx dx dx dx
that, after integration, provides
9.7 The Finite Element Method 205
l l
df d dw
w dx = (w f )d x − f d x,
0 dx 0 dx 0 dx
or, equivalently,
l
df dw
w d x = (w f )|0 − f d x,
0 dx 0 dx
such that
⎛ ⎞
l
⎜ du ⎟ dw du
⎜w A E ⎟ − AE d x + wp d x = 0, ∀w with w(l) = 0.
⎝ d x⎠
0 dx dx 0
σx 0
Simplifying a bit,
dw du
AE d x = (w At¯)x=0 + wp d x = 0, ∀w with w() = 0. (9.2)
0 dx dx 0
Note that the last expression coincides with the principle of virtual work devel-
oped in Eq. (4.3). Albeit it has been obtained by a very different method, both
expressions are exactly the same.
The problem thus consists in finding a sufficiently smooth function u(x) such
that it satisfies the essential boundary conditions (those related to displacements)
that satisfies Eq. (9.2). This solution is not always easy to find. In fact, we have
transformed a differential equation into an integro-differential problem, but many
difficulties remain the same. What is the interest in all these developments? Precisely
the possibility to approximate the solution to the problem within a prescribed degree
of accuracy. Let us explain this a little bit further.
Polynomials are well-known to have nice approximation properties. Therefore,
let us try to approximate the solution (the displacement field) by means of the best
available polynomials,
206 9 Statically Indeterminate Structures
(1) (2)
(1) (2)
Fig. 9.8 Approximation of the displacement field within each finite element. Here, a bar is split
into two elements, (1) and (2), and a polynomial approximation is assumed within each one
where e indicates that this approximation is made within element e, see Fig. 9.8.
Each bar will be split into one or more elements of finite size (hence the name
finite elements) and a polynomial approximation of the displacement field will be
adopted within each one of these regions. Here, αi represent the coefficients for each
monomial, to be chosen so as to guarantee that the resulting displacement field is
continuous.
To guarantee that the integral Eq. (9.2) exists, the displacement field must be
continuous,
An easy way of satisfying this condition is to take the αi coefficient as the nodal
values of the displacement or, in other words, to make the polynomials take unity
value at the element nodes. It is also well-known that, for the finite element method
to provide an accurate solution, it must be able to reproduce, at least piece-wise,
constant strain fields within each element. Therefore, the simplest polynomial to
satisfy this condition needs to be at least linear,
By putting the nodes at the beginning and the end of the element, we will have the
same number of nodal values as parameters in the polynomial approximation. In
other words:
α0e
u (x) = 1 x
e
e = p(x)α .
e
α1
p(x)
αe
9.7 The Finite Element Method 207
le
Fig. 9.9 A simple linear
finite element
xe1 xe2
1 e 2
such that
αe = (M e )−1 d e .
such that:
N e = N1e (x) N2e (x) = p(x)(M e )−1
x2e −x1e 1 1
= 1x = e x2e − x x − x1e ,
1 1 le l
N Ie (x J ) = δ I J ,
N1e (x)
1
N2e (x)
xe1 e xe2
Fig. 9.10 A simple linear finite element
n ne
n ne
u (x J ) =
e
N Ie (x J )u eI = δ I J u eI = u eJ .
I =1 I =1
This means that the predicted values for the displacement at both ends of the element
coincide with the nodal displacement values.
In this same spirit, we can compute the derivatives of the just-approximated dis-
placement field,
du d d Ne e d N1e e d N2e e
= (N e d e ) = d = u + u .
dx dx dx dx 1 dx 2
that must be expressed in matrix form, of course, by:
du e u e
d N2e
= ddNx1 1 = Be d e .
dx dx u e2
Proceeding in exactly the same way, one can envisage a quadratic element, for
instance, by increasing the polynomial degree of the approximation, see Fig. 9.11.
Obviously, this quadratic element possesses three nodes (since a parabola is described
by three points!).
For bars subjected to bending moments, the differential equation was presented in
Eq. (6.7). Consider, for the sake of simplicity of the exposition, the two-dimensional
case of a bar subjected to only one moment. In this case, the main novelty with
respect to the just explained truss-bar element is the presence of derivatives up to
order four (it was order two for trusses). Therefore, by following the same procedure
introduced before, we will need to apply integration by parts twice, so as to give
d 2w d 2u y dw
E I dx = wpd x + m̄ + (w s̄)|Γs , ∀w,
Ω dx2 dx2 Ω dx Γm
9.7 The Finite Element Method 209
N3e (x)
1 2 3
Fig. 9.11 A quadratic finite element
with m I the nodal moments (note that m I = m(x I )). Hermite polynomials take the
form
1 e
Nu1 = (1 − ξ)2 (2 + ξ), Nθ1 = (1 − ξ)2 (1 + ξ),
4 8
1 e
Nu2 = (1 + ξ)2 (2 − ξ), Nθ2 = (1 + ξ)2 (1 − ξ),
4 8
with ξ = 2xe
− 1, such that −1 ≤ ξ ≤ 1. Again, a fundamental characteristic of the
finite element method is that it holds
Nu I (x J ) = δ I J and that Nθ I (x J ) = δ I J .
210 9 Statically Indeterminate Structures
Once this Hermite approximation has been substituted into the weak form of the
problem, we arrive at a discrete, algebraic systems of equations
Kd = f + r
where
d2 Ne
Ke = E I B e T B e d x, with B e = ,
Ωe dx2
d NeT
f =
e eT
N pd x + (N eT
s̄)|Γs + m̄ .
Ωe dx
Γm
f eΩ
f eΓ
This value is exactly the same value predicted by classical, discrete matrix analysis
procedure. In other words, for Euler–Bernoulli–Navier beams, the finite element is
exact and not a mere approximation of the solution.
The reader could at this point ask his- or herself about the benefit of performing all
this complex mathematical analysis to arrive to the solution predicted some 50 years
earlier by classical matrix procedures. Indeed, the benefit consists of the possibility of
extending this procedure not to the case of bars, but to any other differential equation,
including the noteworthy equations of linear elasticity, Eq. (1.11). For instance, in
Fig. 9.12 a finite element model of an arch that employs three-dimensional linear
finite element is shown. More sophisticated models are also equally possible, see
Fig. 9.13, where a shell finite element model of the Los Manantiales restaurant at
Xochimilco, Mexico, by Félix Candela, is shown.
The finite element method is nowadays the ubiquitous method for the simulation
of a wide range of physical phenomena, ranging from solid to fluid mechanics,
spanning various scales, boundary conditions and loading types. This technique has
exceeded by far what we intended with this book. We highly recommend that the
reader learn as much as possible about it, since we are confident that it will continue
to be the reference technique for sophisticated computations in the field of structural
mechanics.
9.7 The Finite Element Method 211
Fig. 9.12 Three-dimensional finite element model of an arch subjected to self weight
Fig. 9.13 Finite element model of the shell roof for the Los Manatiales restaurant in Xochimilco,
Mexico. Félix Candela, arch. M. Sc. Thesis by T. Izuel, Universidad de Zaragoza
212 9 Statically Indeterminate Structures
Worked Examples
9.1. Consider the following statically indeterminate truss under the loads depicted
in Fig. 9.14. The bar AB is subjected to an increment of temperature ΔT = 20 ◦ C
and was mistakenly manufactured shorter than needed. The magnitude of this defect
is δ AB = 0.01 cm. In addition, the support at D undergoes a displacement to the
right u D = 1 mm due to a failure in the soil under the foundation. Determine the
internal forces in the truss by considering E = 2.1 · 106 kp/cm2 , A = 25 cm2 , α =
1.2 · 10−5 ◦ C−1 .
1T
B C
2m
A D
2m
Solution 9.1. (a) As mentioned in the previous sections, the first step in the analysis
of a structure is always to determine its degree of static indeterminacy. This will
enable us to know the type of problem we are facing. In this particular case, we
have 6 bars −2 × 4 nodes +4 reaction forces on the supports. So this system is
statically indeterminate with a degree 2 of indeterminacy. A careful observation
of the structure reveals that one of the sources of indeterminacy is one “extra”
support (there are two reactions at A and another two at D, while there are only
three global-equilibrium equations). The other comes from the bars B D and AC,
crossing over each other.
(b) We consider an equivalent determinate system, like the one in Fig. 9.15. This
choice is obviously arbitrary, but we must take into account the nature of the
indeterminacy. In other words, we could have chosen any other reaction in A or
D, or the force in bar AC, but never two reactions or two internal forces in any
9.7 The Finite Element Method 213
P
B C
2m
X
A D
HA Y
VA VD
2m
bar.
We have named these hyperstatic unknown variables as X and Y , respectively.
Now imposing equilibrium conditions for the free-body diagram depicted in
Fig. 9.15,
X X
Fx = 0 ⇒ H A = Y + P + √ − √ = Y + P ⇒ H A = Y + P,
2 2
X X
Mz (A) = 0 ⇒ 2VD − 2P + 2 √ − 2 √ = 0 ⇒ VD = P,
2 2
X X
Fy = 0 ⇒ V A + V D + √ − √ − Q = 0 ⇒ V A + V D = Q ⇒ V A = Q − P.
2 2
X
Fx = 0 ⇒ N AD = Y − √ ,
2
214 9 Statically Indeterminate Structures
X
Fy = 0 ⇒ NC D = −P − √ .
2
NAC
X
−P − √
2
N AC X √
Fy = 0 ⇒ √ − P + √ = 0 ⇒ N AC = P 2 + X,
2 2
N AC N AC
Fx = 0 ⇒ N BC + √ = 0 ⇒ N BC = − P + √ .
2 2
B X
P −P − √
2
α
X
NAB
9.7 The Finite Element Method 215
X
Fy = 0 ⇒ N AB = −Q − √ .
2
(c) To determine the value of redundant forces, let us consider two different virtual
systems associated with the equivalent statically determinate truss.
By solving this first virtual system, we find (Fig. 9.19):
B C
2m
A D
I
HA 1
VAI I
VD
2m
P = 0; Q = 0; Y = 1; X = 0,
so that,
I
N AB = 0; N AC
I
= 0; N BC
I
= 0; N AD
I
= 1; NCI D = 0.
For the second virtual system, coined as I I , see Fig. 9.20, we have,
P = 0; Q = 0; Y = 0; X = 1,
so that,
−1 −1 −1 −1
II
N AB = √ ; N AC
II
= 1; N BC
II
= √ ; N AD
II
= √ ; NCI ID = √ .
2 2 2 2
216 9 Statically Indeterminate Structures
2m
1
A D
II
HA
VAII II
VD
2m
For this exercise, it is necessary to take into account that, for bar AB, its increment
in length is due not only to its internal force, but also to its temperature increment
and the defective length,
N AB · L AB
Δ AB = + α · ΔT · L AB + δ eAB .
EA
In the application of the principle of virtual work to the virtual system I , we
must take into account that the real horizontal displacement to the right of point
D will be the actual movement of this point, or, in other words,
5
X 200
0.1 cm = u x (D) = NiI · ΔL i = Y − √ · · 1,
i=1
2 EA
X 200
0.1cm = Y − √ · ⇒ 26.25 · 103 kp
2 52.5 · 106
X X
= Y − √ ⇒ Y = 26.25 · 103 + √ .
2 2
9.7 The Finite Element Method 217
Applying again the principle of virtual work to the second virtual system, I I ,
the relative displacement between nodes B and D must be equal to the length
increment of the bar B D (in other words: this relative displacement must be
“admissible”), so,
√
ψ −X · 2 · 200
FN · δN = 1 · δB + 1 · δD = δB + δD = ΔB D = .
N
EA
Therefore:
√
−X · 2 · 200
E A
X 200 −1
= −10−4 − √ + αΔT 200 + 0, 01 · √
2 EA 2
√ √
(1000 2 + X )200 2
+ · [1]
EA
(−1000 − √X )
(Y − √X2 ) −1 2 −1
+ 200 · √ + 200 · √
EA 2 EA 2
(−1000 − 2 )X
√
−1
+ 200 · √ .
EA 2
Solving the system with two equations and two unknown variables X and Y , we
have that
N AB = −13078.08 kp (Compression),
N AC = 5767.28 kp (Traction),
N AD = 26250 kp (Traction),
N BC = −4078.08 kp (Compression),
N B D = X = 4353.07 kp (Traction),
NC D = −4078.08 kp (Compression).
9.2. Consider the following statically indeterminate structure under the loads
depicted in Fig. 9.21. Compute the bending-moment diagram for the structure.
B C
h
A D
Solution 9.2. (a) The first step will be, as always, to determine the degree of static
indeterminacy of the structure and to know in advance if it is internal or external.
9.7 The Finite Element Method 219
In this case; we have 3 unknown internal forces in 3 beams and 6 reaction forces
at the supports to determine, so we have 15 unknown variables. In the other hand,
we have 3 equations per node (and a total of 4 nodes) and 2 hinges, where we know
that bending moments vanish. Therefore, this system is statically indeterminate
of degree 15 − 14 = 1 (and this is clearly due to one redundant external reaction).
(b) We consider an equivalent isostatic system depicted in Fig. 9.22. Solving the
external equilibrium to obtain the reaction forces, we find that.
h
M M
A M M D
h h
qL qL
2 2
L
M
H A = HD = ,
h
qL
V A = VD = .
2
M A = M.
qL qx2
M (x) = 2
x − 2
x
B C
qL2
8
x
M (x) = M (1 − h
)
x
A D
M M
Fig. 9.23 Bending-moment diagrams for the equivalent isostatic system to the initial structure for
Exercise 9.2
(c) The virtual system associated with the redundant variable is shown in Fig. 9.24.
This gives rise to the bending-moment diagram shown in Fig. 9.25.
1 1
A 1 1 D
h h
L
9.7 The Finite Element Method 221
B C
x
M I (x) = 1 − h
x
A D
1 1
ψ
L
FN · u N = Mzψ · κz d x + · · ·
N b 0
Therefore,
1 h x x 2M h 2x x2
1 · θD = 0 = 2 · M 1− 1− dx = 1− + 2 dx
EI 0 h h EI 0 h h
2M h 2Mh
= h−h+ = ⇐⇒ M = 0.
EI 3 3E I
B C
qL2
8
A D
222 9 Statically Indeterminate Structures
9.3. Compute the vertical displacement of node B, assuming that all the bars
depicted in Fig. 9.27 are subjected to an increment of temperature ΔT .
3L
2
√
60o
A
C
Solution 9.3. (a) Again, the first step is to determine the degree of indeterminacy
of the structure. For this truss we have 3 bars −2 × 3 nodes +4 reaction forces
on the supports. So this system is statically indeterminate of degree one (this is
clearly due to one redundant external reaction).
(b) We consider an equivalent statically determinate system, shown in Fig. 9.28,
where X is the redundant unknown variable.
B
3L
2
√
60o
A C
X
H A = X (→); V A = VC = 0.
Obtaining now the internal forces by imposing equilibrium at each node of the
truss,
N AC = −X (Compression),
N AB = N BC = 0.
We suggest two different virtual systems. The first one is aimed at computing
the displacement at point B, while the second one will serve to compute the
hyperstatic variable X . Solving this virtual system I (see Fig. 9.29), the reaction
forces on the supports turn out to be
1
H A = 0 V A = VC = (↓)
2
3L
2
√
60o
A C
1
I
N AB = N BC
I
= √ (Compression)
3
224 9 Statically Indeterminate Structures
−3
I
N AC = √ (Traction)
3
Solving this second virtual system, I I , (see Fig. 9.30), the reaction forces on the
supports turn out to be
H A = 1(→) V A = VC = 0.
3L
2
√
60o
A C
1
II
N AC = −1 (Compression)
Finally, by applying the compatibility condition, i.e., by noting that in the real
system δCx = 0,
X = αE AΔT,
and therefore,
√ √ √
2 3 3 −X L 2 3
δB = αLΔT − + αLΔT ⇒ δ B = αLΔT.
3 6 EA 3
Appendix A
Indicial Notation
3
ai xi = p.
i=1
At this point, we can use the convention usually referred to as Einstein notation
or Einstein summation convention, which implies summation over a set of indexed
terms in a formula when an index appears twice in a single term along the variational
rank of the indexes (1 to 3 in the above equation).
Therefore,
ai xi = p
• Norm of a vector: √ √
||u|| = u·u= ui ui
• Derivative of a function:
n
∂f ∂f
df = d xi = d xi = f ,i d xi
i=1
∂ x i ∂ xi
Let us define the Kronecker Delta δi j as a symbol1 whose values can be,
1 if i = j
δi j =
0 if i = j
Solution A.1. • Indexes i, j appear repeated on the equation, so we must sum them
from 1 to 3. Using the definition of the Kronecker Delta function, we have
δ11 δ11 + δ12 δ12 + δ13 δ13 + δ21 δ21 + δ22 δ22 + δ23 δ23 + δ31 δ31 +
+δ32 δ32 + δ33 δ33 = δ11 δ11 + δ22 δ22 + δ33 δ33 = 1 + 1 + 1 = 3
ε111 ε111 + ε112 ε112 + ε113 ε113 + ε121 ε121 + ε122 ε122 + ε123 ε123 +
+ε131 ε131 + ε132 ε132 + ε133 ε133 = ε123 ε123 + ε132 ε132 =
= 1 · 1 + (−1) · (−1) = 2
For i = 2,
ε211 ε211 + ε212 ε212 + ε213 ε213 + ε221 ε221 + ε222 ε222 + ε223 ε223 +
+ε231 ε231 + ε232 ε232 + ε233 ε233 = ε213 ε213 + ε231 ε231 =
= (−1) · (−1) + 1 · 1 = 2
εi jk εi jk = 6
= a11 a22 a33 + a21 a32 a13 + a31 a12 a23 − a11 a32 a23 − a12 a21 a33 − a13 a22 a31
= ε123 a11 a22 a33 + ε231 a21 a32 a13 + ε312 a31 a12 a23 + ε132 a11 a32 a23 + ε213 a12 a21 a33
• If, in the definition of determinant of a matrix, we change the first column by the
unit vectors ei , the second component by v, and the third one by w, the vector
product is given by the following indicial expression
(v × w) = εi jk ei v j wk ,
(v × w)i = εi jk v j wk ,
230 Appendix A: Indicial Notation
î ĵ k̂
v × w = v1 v2 v3 = îv2 w3 + k̂v1 w2 + ĵv3 w1 − k̂v2 w1 − îv3 w2 − ĵv1 w3 =
w1 w2 w3
εi jk êi v j wk .
(v × w)i = εi jk v j wk .
References
1. Betsch P, Steinmann P (2002) Frame-indifferent beam finite elements based upon the geomet-
rically exact beam theory. Int J Numer Methods Eng 54(12):1775–1788
2. Block P, Ochsendorf J (2007) Thrust network analysis: a new methodology for three-
dimensional equilibrium. J Int Assoc Shell Spatial Struct 48(3):167–173
3. Bonet J, Wood RD (2008) Nonlinear continuum mechanics for finite element analysis. Cam-
bridge University Press, Cambridge
4. Calatrava S (1981) Zur Faltbarkeit von Fachwerke. PhD thesis, ETH Zurich. https://www.
research-collection.ethz.ch/handle/20.500.11850/137273
5. Courant R (1943) Variational methods for the solution of problems of equilibrium and vibra-
tions. Bull Am Math Assoc 42:2165–2186
6. Cross H (1930) Analysis of continuous frames by distributing fixed-end moments. In: Proceed-
ings of the American society of civil engineers, ASCE, pp 919–928
7. Felippa CA (2001) A historical outline of matrix structural analysis: a play in three acts. Comput
Struct 79(14):1313–1324
8. Fish J, Belytschko T (2007) A first course in finite elements. Wiley, New Jersey
9. Han SM, Benaroya H, Wei T (1999) Dynamics of transversely vibrating beams using four
engineering theories. J Sound Vibr 225(5):935–988
10. Heyman J (1982) The masonry arch. Ellis Horwood
11. Huerta S (2004) Arcos, bóvedas y cúpulas. Geometría y equilibrio en el cálculo tradicional de
estructuras de fábrica, Instituto Juan de Herrera
12. Huerta S (2008) The analysis of masonry architecture: a historical approach: to the memory of
professor henry j. cowan. Architect Sci Rev 51(4):297–328
13. Jelenić G, Crisfield MA (1999) Geometrically exact 3d beam theory: implementation of a
strain-invariant finite element for statics and dynamics. Comput Methods Appl Mech Eng
171(1):141–171
14. Love AEH (1927) A treatise on the mathematical theory of elasticity. Cambridge University
Press, Cambridge
15. Maney GA (1915) Studies in engineering. University of Minnesota, Minneapolis
16. Muttoni A (2011) The art of structures. EPFL Press, Lausanne
17. Muttoni A, Schwartz J, Thürlimann B (1996) Design of concrete structures with stress fields.
Springer Science & Business Media, Berlin
18. Reddy JN (2008) An introduction to continuum mechanics. Cambridge University Press, Cam-
bridge
19. Simo JC, Vu-Quoc L (1988) On the dynamics in space of rods undergoing large motions – a
geometrically exact approach. Comput Methods Appl Mech Eng 66(2):125–161
20. Turner MJ, Clough RJ, Martin HC, Topp LJ (1956) Stiffness and deflection analysis of complex
structures. J Aeronaut Sci 23(9):805–823
F
Field, 6 M
First area moment, 132, 152 Masonry structures, 60
First Piola–Kirchhoff stress tensor, 12, 14, Mechanics of materials, 20
15 Membrane, 47
Fixed-end moments, 198 Method of consistent deformations, see also
Flexibility method, 191, 194 Flexibility method
Force polygon, 37 Method of sections, 86, 96, 103, 108, 116
Foundation, 138 Moment distribution method, 200
Frame, 114 Movement, 2
three-hinged, 114 invertible, 2
Free body diagram, 37, 79 regular, 2
Funicular curve, 41, 43
Funicular polygon, 41, 103
N
G Navier form of the elastic problem, 19
Girder, 169 Neutral axis, 137
Green-Lagrange strain tensor, 16 Nominal stress tensor, 12
Groin vaults, 68
O
H Ordinary differential equation, 179
Hollow profile, 156
Hooke’s law, 17
Hypar, see also Hyperbolic paraboloids P
Hyperbolic paraboloids, 68 Pavilion vaults, 66, 68
Hyperstatic structures, 47, 61, 192 Permanent loads, 45
Hypothesis of small displacements, 2 Perturbations
Hypothesis of small strains, 7 arches, 61, 64
Polar decomposition, 7
Prandtl function, 176
I
Principle of Virtual Displacements, 89
Internal force, 33, 106
Principle of Virtual Forces, 89
axial, 34
Principle of Virtual Work, 87, 140
diagrams, 108, 111, 112
shear, 106
Internal moment
bending, 107 R
diagrams, 108 Radius of gyration, 139
torsion, 107 Rise, 33
twisting, 107 Rotation tensor, 7
Internal static indeterminacy, 192
Invariants of tensor, 12
Isostatic structures, 36, 192, 194 S
Saint Venant’s principle, 1
Saint Venant tensor, see also Compatibility
K tensor
Kronecker delta tensor, 7 Second Piola–Kirchhoff stress tensor, 12, 16
Index 235
T
Thin-walled profiles, 153 W
Thrust Network Analysis (TNA), 68 Warping function, 177
Tie, 71 Weak form, 204
Timoshenko shear coefficient, 161
Torsion, 169
center of torsion, 171 Y
constant, 175 Yield stress, 45