Vanucci P Tensor Algebra and Analysis For Engineers With App
Vanucci P Tensor Algebra and Analysis For Engineers With App
com
Contemporary Mathematics and Its Applications:
Monographs, Expositions and Lecture Notes
Series Editor
M Zuhair Nashed (University of Central Florida)
Editorial Board
Guillaume Bal (University of Chicago) Palle Jorgensen (University of Iowa)
Gang Bao (Zhejiang University) Marius Mitrea (University of Missouri
Liliana Borcea (University of Michigan) Columbia)
Raymond Chan (The Chinese University of Otmar Scherzer (University of Vienna)
Hong Kong) Frederik J Simons (Princeton University)
Adrian Constantin (University of Vienna) Edriss S Titi (Texas A&M University)
Willi Freeden (University of Kaiserslautern) Luminita Vese (University of California,
Charles W Groetsch (The Citadel) Los Angeles)
Downloaded from www.worldscientific.com
This series aims to inspire new curriculum and integrate current research into texts. Its aims and
main scope are to publish:
– Cutting-edge Research Monographs
– Mathematical Plums
– Innovative Textbooks for capstone (special topics) undergraduate and graduate level courses
– Surveys on recent emergence of new topics in pure and applied mathematics
– Advanced undergraduate and graduate level textbooks that may initiate new directions and
new courses within mathematics and applied mathematics curriculum
– Books emerging from important conferences and special occasions
– Lecture Notes on advanced topics
Monographs and textbooks on topics of interdisciplinary or cross-disciplinary interest are
particularly suitable for the series.
Published
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.
Printed in Singapore
Preface
Downloaded from www.worldscientific.com
v
vi Preface
them completing the theoretical part through new results and proofs. All
the exercises are entirely developed and solved at the end of the book in
order to provide the reader with thorough support for his learning.
In Chapter 1, vectors and points are introduced and also, with a
small anticipation of some results of the second chapter, applied vectors
are visited. Chapter 2 is completely devoted to the algebra of second-
rank tensors and the succeeding Chapter 3 to that of fourth-rank tensors.
Intentionally, these are the only two types of tensors introduced in the book:
They are the most important tensors in mechanics, and they allow us to
represent deformation, stress, and the constitutive laws. I preferred not to
introduce tensors in an absolutely general way but to go directly to the
most important tensors for applications in mechanics; for the same reason,
the algebra of other tensors, namely of third-rank tensors, is not presented
in this primer text.
The analysis of tensors is done using first-differential geometry of
curves, in Chapter 4, for differentiation and integration with respect to
only one variable, then introducing the differential operators for fields and
deformations, in Chapter 5.
Then, a generalization of second-rank tensor algebra and analysis in
the sense of the use of curvilinear coordinates is presented in Chapter 6,
where the notion of metric tensor, co- and contravariant components, and
Christoffel’s symbols are introduced.
Finally, Chapter 7 is entirely devoted to an introduction to the
differential geometry of surfaces. Classical topics such as the first and second
Preface vii
ix
Downloaded from www.worldscientific.com
xi
Downloaded from www.worldscientific.com
Preface v
About the Author ix
Acknowledgments xi
List of Symbols xvii
xiii
xiv Contents
2.11 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.12 Reflexions . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2.13 Polar decomposition . . . . . . . . . . . . . . . . . . . . . 51
2.14 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3. Fourth-Rank Tensors 57
3.1 Fourth-rank tensors . . . . . . . . . . . . . . . . . . . . . . 57
3.2 Dyads, tensor components . . . . . . . . . . . . . . . . . . 58
3.3 Conjugation product, transpose, symmetries . . . . . . . . 59
3.4 Trace and scalar product of fourth-rank tensors . . . . . . 62
3.5 Projectors and identities . . . . . . . . . . . . . . . . . . . 63
3.6 Orthogonal conjugator . . . . . . . . . . . . . . . . . . . . 66
3.7 Rotations and symmetries . . . . . . . . . . . . . . . . . . 67
3.8 The Kelvin formalism . . . . . . . . . . . . . . . . . . . . 69
Downloaded from www.worldscientific.com
:= : definition symbol
|: such that
∃!: exists and is unique
R: set of real numbers
E: ordinary 3D Euclidean space
V: vector space of translations, associated with E
Lin(V): vector space of second-rank tensors
Lin(V): vector space of fourth-rank tensors
x, y, z etc.: scalars (elements of R)
p, q, r etc.: points (elements of E)
u, v, w etc.: vectors (elements of V)
u = p − q etc.: vector difference of two points of E
L, M, N etc.: second-rank tensors (elements of Lin(V))
L, M, N etc.: fourth-rank tensors (elements of Lin(V))
u · v etc.: scalar product of vectors
L · M etc.: scalar product of second-rank tensors
u × v etc.: cross product of vectors
u ⊗ v etc.: dyad (second-rank tensor) of vectors
u = |u|: norm of a vector
L = |L|: norm of a tensor
|p − q|, |u − v|, |L − M|: distance of points, vectors or tensors
xvii
xviii List of Symbols
q − p, p, q ∈ E.
∀p, q ∈ E, ∃! v ∈ V| q − p = v.
∀v ∈ V, v : E → E| q = v(p) → q = p + v.
q = p + v =p1 + v = q1 ,
whose geometric meaning is depicted in Fig. 1.1. Unlike the difference, the
sum of two points is not defined and is meaningless.
We define the sum of two vectors u and v as the vector w such that
1
2 Tensor Algebra and Analysis for Engineers
q = v(p) = p + v,
then
r = u(q) = q + u = w(p),
see Fig. 1.2, which shows that the above definition actually coincides with
the parallelogram rule and that
u + v = v + u,
as is obvious, for the sum over a vector space commutes. It is evident that
the sum of more than two vectors can be defined iteratively, summing up
a vector at a time to the sum of the previous vectors.
The null vector o is defined as the difference of any two coincident
points:
o := p − p ∀p ∈ E;
v + o = v ∀v ∈ V.
Points and Vectors 3
In fact,
∀p ∈ E, v + o = v + p − p → p + v + o = p + v ⇐⇒ v + o = v.
w := ki vi , ki ∈ R, i = 1, . . . , n.
The n + 1 vectors w, vi , i = 1, . . . , n, are said to be linearly independent
if there does not exist a set of n scalars ki such that the above equation is
satisfied and are said to be linearly dependent in the opposite case.
ω : V × V → R,
ω(u, v) = u · v.
The properties of bilinearity prescribe that, ∀u, v ∈ V and ∀α, β ∈ R,
u · (αv + βw) = αu · v + βu · w,
(αu + βv) · w = αu · w + βv · w,
while symmetry implies that
u · v = v · u ∀u, v ∈ V.
v · v > 0 ∀v ∈ V, v · v = 0 ⇐⇒ v = o.
u · v = 0.
1 We adopt here and in the following the Einstein notation for summations: All the times
when an index is repeated in a monomial, then the summation
with respect to that index,
called the dummy index, is understood, e.g. ki vi = i ki vi . We then say that the index
i is saturated. If a repeated index is underlined, then it is not a dummy index, i.e. there
is no summation.
2 The scalar product ω(u, v) is also indicated as < u, v >.
4 Tensor Algebra and Analysis for Engineers
Thanks to the properties of the scalar product, we can define the Euclidean
norm of a vector v as the nonnegative scalar, denoted equivalently by v
or |v|:
√
v = |v| := v · v.
d(u, v) := |u − v| = |v − u|.
Two points or two vectors are coincident if and only if their distance is null.
The unit sphere S of V is defined as the set of all the vectors whose
norm is one:
S := {v ∈ V| v = 1}.
There is a general way to define a basis for a vector space of any kind. We
limit the introduction of the concept of basis to the case of V only, which
is of interest in classical mechanics. Generally, a basis B of V is any set of
Downloaded from www.worldscientific.com
B = {e1 , e2 , e3 }.
v = vi ei = v1 e1 + v2 e2 + v3 e3 .
ei · ej = δij ,
u · v = ui ei · vj ej = ui vj δij = ui vi = u1 v1 + u2 v2 + u3 v3 .
6 Tensor Algebra and Analysis for Engineers
In particular, it is
v · ei = vk ek · ei = vk δik = vi , i = 1, 2, 3.
So, the Cartesian components of a vector are the projection of the vector
on the three vectors of the basis B; such quantities are the director cosines
of v in the basis B. In fact, if θ is the angle formed by two vectors u and
v, then
u · v = u v cos θ.
uv = u cos θ,
uv = c v.
So,
u u u·v u·v
c= cos θ → cos θ = 2 ⇒ cos θ = .
v v v uv
We remark that while the scalar product, being an intrinsic operation,
does not change with a change of basis, the components vi of a vector are
not intrinsic quantities, but they are basis-dependent: A change of the basis
makes the components change. The way this change is done is introduced
in Section 2.11.
A frame R for E is composed of a point o ∈ E, the origin, and a basis
B of V:
R := {o, B} = {o; e1 , e2 , e3 }.
The use of a frame for E is useful for determining the position of a point
p, which can be done through its Cartesian coordinates xi , defined as the
components in B of the vector p − o:
xi := (p − o) · ei , i = 1, 2, 3.
Of course, the coordinates xi of a point p ∈ E depend upon the choice of o
and B.
Downloaded from www.worldscientific.com
We introduce now a set of definitions, concepts, and results that are widely
used in physics, especially in mechanics. For that, we need to anticipate
some results that are introduced in the next chapter, namely that of cross
product, in Section 2.9, and of complementary projector, in Exercise 2,
Chapter 2. This slight deviation from the good rule of consistent progression
in stating the results is justified by the fact that, actually, the matter
presented hereafter is still that of vectors. The reader can, of course, come
back to the topics of this section once they have studied Chapter 2.
We call applied vector vp a vector v associated to a point p ∈ E. In
physics, the concept of applied vector3 is often employed, for example, to
represent forces4 . We define the resultant of a system of n applied vectors
vip as the vector
n
R := vip .
i=1
We define the moment of an applied vector vp about a point o, called the
center of the moment, the vector
Mo := (p − o) × vp ,
and the resultant moment of a system of n applied vectors vip about a point
o the vector
n
(pi − o) × vip .
Downloaded from www.worldscientific.com
Mro :=
i=1
Also, in this case, the resultant moment does not change when R and o1 −o2
are parallel vectors, but not exclusively, as another possibility is that R = o:
For the systems of applied vectors with null resultant, the resultant moment
is invariant with respect to the center of the moment.
An interesting relation can be found if the two members of the last
equation are projected onto R, which gives
Mro1 · R = Mro2 · R : (1.2)
The projection of the resultant moment onto the direction of R does not
depend upon the center of the moment.
A particularly important case of the system with a null resultant is that
of a couple, which is composed of two opposite vectors v and −v, which
are applied to two points p and q:
vp = −vq .
Of course, by definition, R = o for any couple and, as a consequence, the
resultant moment Mr of a couple, called the moment of the couple and
simply denoted by M, is independent of the center of the moment (that is
why the index denoting the center of the moment is omitted): Referring to
Fig. 1.6,
M = (p − o) × vp + (q − o) × vq = (p − o) × v − (q − o) × v
= ((p − o) − (q − o)) × v = (p − q) × v.
If u ∈ S| u × v = o, then
bc := |(I − u ⊗ u)(p − q)| = |p − q| sin θ
10 Tensor Algebra and Analysis for Engineers
Mra × R = o ∀a ∈ A.
a = kR, k ∈ R ⇒
Proof. Existence: we need at least a point a ∈ E| MR
Mra × R = o. From Eq. (1.1), ∀o ∈ E, we get
Mro × R
o−a= ,
R2
Points and Vectors 11
Mra × R = o.
Mro × R
a=o− ∈ A.
R2
So, because Mr does not change when calculated with respect to the points
of an axis parallel to R, A is the axis passing through a and parallel to R
whose equation is
Mro × R
p =a+t R = o− + t R, t ∈ R.
R2
Downloaded from www.worldscientific.com
In this equation, the left-hand side and the first term on the right-hand side
are null by the definition of central axis. Because (a−q)×R is perpendicular
to R and R = o by hypothesis, the left-hand side is null if and only if
a = q ⇒ Aˆ = A.
Mro · R = 0 ∀o ∈ E.
the same time Mra · R = 0 and Mra × R = o, so the only possibility is that
Mra = o ∀a ∈ A,
i.e. in this case, A is the axis of points that make the resultant moment
vanish.
Two systems of applied vectors are equivalent if they have the same
resultant R and the same resultant moment Mro about any center o ∈ E.
The equivalence does not depend upon the center o. In fact, by Eq. (1.1),
if two systems have the same R and the same Mro1 , with o1 a given point,
then also Mro2 will be the same ∀o2 ∈ E.
point of E.
Proof. By construction, R is the same for the two systems; moreover, for
the equivalent system (resultant plus couple), it is
M + (o − o) × R = M.
So, if the couple has a moment M = Mro , the two systems are equivalent.
R = o, Mro = o ∀o ∈ E.
1.5 Exercises
4. Prove that
u · v = 0 ⇐⇒ |u − v| = |u + v| ∀u, v ∈ V.
5. Prove the linear forms representation theorem: Let ψ : V → R be a
linear function. Then, ∃! u ∈ V such that
ψ(v) = u · v ∀v ∈ V.
6. Consider a point p and two noncollinear vectors u, v ∈ S at p. Show
that a vector w is the bisector of the angle formed by u and v if and
only if w · u = w · v.
7. Show that in the case of systems composed of coplanar or parallel
applied vectors with R =o, Mro · R = 0 ∀o ∈ E.
8. Prove that any system of applied vectors with R = o is equivalent to
a couple.
Downloaded from www.worldscientific.com
Second-Rank Tensors
Downloaded from www.worldscientific.com
1 We consider, for the time being, only second-rank tensors that constitute a very
important set of operators in classical and continuum mechanics. In the following, we
also introduce fourth-rank tensors.
15
16 Tensor Algebra and Analysis for Engineers
(u ⊗ v)w := v · w u ∀w ∈ V.
B 2 = {ei ⊗ ej , i, j = 1, 2, 3},
Downloaded from www.worldscientific.com
L = Lij ei ⊗ ej , i, j = 1, 2, 3,
where Lij s are the nine Cartesian components of L with respect to B 2 . Lij s
can be calculated easily:
(u ⊗ v)ij = ei · (u ⊗ v) ej = u · ei v · ej = ui vj . (2.2)
2 Insome texts, the dyad is also called the tensor product; we prefer to use the term dyad
because the term tensor product can be ambiguous, as it is used to denote the product
of two tensors, see Section 2.3.
Second-Rank Tensors 17
basis B:
⎡ ⎤
L11 L12 L13
L = ⎣L21 L22 L23 ⎦ .
L31 L32 L33
(L1 L2 )v := L1 (L2 v) ∀v ∈ V.
Downloaded from www.worldscientific.com
L1 L2 = L2 L1 ;
however, by the same definition of the identity tensor and of tensor product,
IL = LI = L ∀L ∈ Lin(V).
For any tensor L ∈ Lin(V), there exists just one tensor L , called the
transpose of L, such that
u · Lv = v · L u ∀u, v ∈ V. (2.5)
u · Lv = v · L u = u · (L ) v ⇒ (L ) = L.
L
ij = ei · L ej = ej · (L ) ei = ej · Lei = Lji .
while
u · (AB)v = Bv · A u = v · B A u ⇒ (AB) = B A .
Moreover,
A tensor L is symmetric ⇐⇒
L = L .
Lij = Lji .
L = −L .
Second-Rank Tensors 19
L = Ls + La ,
with
L + L
Ls = ∈ Sym(V)
2
and
L − L
La = ∈ Skw(V)
2
so that, finally,
tr : Lin(V) → R,
called the trace, such that
tr(a ⊗ b) = a · b ∀a, b ∈ V.
20 Tensor Algebra and Analysis for Engineers
For the same definition that has been given without making use of any basis
of V, the trace of a tensor is a tensor invariant, i.e. a quantity extracted
from a tensor that does not depend upon the basis.
Linearity implies that
It is just linearity to give the rule for calculating the trace of a tensor L:
trL = trL,
A · B = tr(A B).
L · (αM + βN) = αL · M + βL · N,
(αL + βM) · N = αL · N + βM · N,
L · M = M · L,
L · L > 0 ∀L ∈ Lin(V), L · L = 0 ⇐⇒ L = O.
These properties give the rule for computing the scalar product of two
tensors A and B:
A · B = Aij (ei ⊗ ej ) · Bhk (eh ⊗ ek ) = Aij Bhk (ei ⊗ ej ) · (eh ⊗ ek )
= Aij Bhk tr[(ei ⊗ ej ) (eh ⊗ ek )] = Aij Bhk tr[(ej ⊗ ei )(eh ⊗ ek )]
= Aij Bhk tr[ei · eh (ej ⊗ ek )] = Aij Bhk ei · eh ej · ek
= Aij Bhk δih δjk = Aij Bij .
Second-Rank Tensors 21
As in the case of vectors, the scalar product of two tensors is equal to the
sum of the products of the corresponding components. In a similar manner,
or using Eq. (2.4)1 , it is easily shown that, ∀a, b, c, d ∈ V,
(a ⊗ b) · (c ⊗ d) = a · c b · d = ai bj ci dj ,
trL = I · L ∀L ∈ Lin(V).
and the distance d(L, M) of two tensors L and M the norm of the tensor
Downloaded from www.worldscientific.com
difference:
d(L, M) := |L − M| = |M − L|.
Ldev := L − Lsph
so that
L = Lsph + Ldev .
We remark that
1
trLsph = trL trI = trL ⇒ trLdev = 0,
3
i.e. the deviatoric part is a traceless tensor. Let A, B ∈ Lin(V), then
1 1
Asph · Bdev = trA I · Bdev = trA trBdev = 0, (2.9)
3 3
i.e. any spherical tensor is orthogonal to any deviatoric tensor.
22 Tensor Algebra and Analysis for Engineers
The sets
1
Sph(V) := Asph ∈ Lin(V)| Asph = trAI ∀A ∈ Lin(V) ,
3
Dev(V) := Adev ∈ Lin(V)| Adev = A − Asph ∀A ∈ Lin(V)
form two subspaces of Lin(V); the proof is left to the reader. For what is
proved above, Sph(V) and Dev(V) are two mutually orthogonal subspaces
of Lin(V).
ω :V ×V ×V →R
is a skew trilinear form if ω(u, v, ·), ω(u, ·, v), and ω(·, u, v) are linear forms
on V and if
Proof. In fact, let u = αv + βw, then for any skew trilinear form ω,
because of Eq. (2.10) applied to the permutation of the positions of the two
u and the two w.
Second-Rank Tensors 23
It is evident that the set of all the skew trilinear forms is a vector space
and that we denote by Ω, whose null element is the null form ω0 ,
ω0 (u, v, w) = 0 ∀u, v, w ∈ V.
3 The proof of this statement is rather involved and outside of our scope; the interested
reader is referred to the classical textbook by Halmos on linear algebra, Section 31 (see
the bibliography). The theory of the determinants is developed in Section 53.
4 More precisely, det L is the function that associates a scalar with each tensor (Halmos,
Section 53). We can, however, for the sake of practice, identify det L with the scalar
associated with L, without consequences for our purposes.
24 Tensor Algebra and Analysis for Engineers
so that
λbL k ω a (u, v, w) = λbL ω b (u, v, w) = ωL
b
(u, v, w)
= k ωL
a
(u, v, w) = λaL k ω a (u, v, w) ⇐⇒ λaL = λbL ,
which proves that det L does not depend upon the skew trilinear form but
only upon L.
The definition given for det L allows us to prove some important
properties. First of all,
det O = 0;
in fact, ∀ω ∈ Ω,
Moreover, if L = I, then
if and only if
det I = 1. (2.12)
det(a ⊗ b) = 0. (2.13)
In fact, if L = a ⊗ b, then
where P3 is the set of all the permutations π of {1, 2, 3} and the i,j,k s are
the components of Ricci’s alternator5 :
⎧
⎨ 1 if {i, j, k} is an even permutation of {1, 2, 3},
i,j,k := 0 if {i, j, k} is not a permutation of {1, 2, 3},
⎩
−1 if {i, j, k} is an odd permutation of {1, 2, 3}.
The above rule for det L coincides with that for calculating the
Downloaded from www.worldscientific.com
determinant of the matrix whose entries are Lij s. This shows that, once a
basis B for V is chosen, det L coincides with the determinant of the matrix
representing it in B and, finally, that
det L = L11 L22 L33 + L12 L23 L31 + L13 L32 L21
(2.15)
− L11 L23 L32 − L22 L13 L31 − L33 L12 L21 .
tr2 L − trL2
I1 = trL, I2 = , I3 = det L. (2.18)
2
Using the theorem of Binet, Theorem 7, along with Eqs. (2.12) and (2.19),
we get
1
det L−1 = .
det L
Equation (2.19) applied to L−1 , along with the uniqueness of the inverse,
gives immediately that
(L−1 )−1 = L,
while
B−1 A−1 = B−1 A−1 AB(AB)−1 = (AB)−1 .
The operations of transpose and inversion commute:
L (L )−1 = I = L−1 L = I = (L−1 L) = L (L−1 ) ⇒
(L−1 ) = (L )−1 := L− .
Second-Rank Tensors 27
Lv = λv, (2.20)
(L − λI)v = o. (2.21)
Downloaded from www.worldscientific.com
det(L − λI) = 0;
However, this is not the case for the eigenvectors that are generally different,
as a numerical example can show.
Developing the Laplace’s equation, it is easy to show that it can be
written as
6 The proof of the spectral theorem is omitted here; the interested reader can find a proof
of it in the classical text by Halmos, p. 155, see the suggested texts.
Second-Rank Tensors 29
I1 = λ1 + λ2 + λ3 , I2 = λ1 λ2 + λ2 λ3 + λ3 λ1 , I3 = λ1 λ2 λ3 .
L = λI.
∇v (v · Lv) = o, v ∈ S.
Lv = λv
and the constraint |v| = 1. The above equation is exactly the one defining
the eigenvalue problem of L: The stationary values (i.e. the maximum
and minimum) of ω hence correspond to two eigenvalues of L and the
directions v, whereupon the stationarity coincides with the respective
eigenvectors.
Two tensors A and B are said to be coaxial if they have the same normal
basis BN , i.e. if they share the same eigenvectors. Let u be an eigenvector
of A, relative to the eigenvalue λA , and of B, relative to λB . Then,
of the vectors of BN , and for the linearity of tensors, we have finally proved
the following.
Theorem 13. The spectrum of any tensor W ∈ Skw(V) is {0} and the
Downloaded from www.worldscientific.com
Proof. This theorem states that zero is the only real eigenvalue of any
skew tensor and that its multiplicity is 1. In fact, let w be an eigenvector
of W relative to the eigenvector λ. Then,
λ2 w2 = Ww · Ww = w · W Ww = −w · WWw
= −w · W(λw) = −λw · Ww = −λ2 w2 ⇐⇒ λ = 0.
Ww = o (2.22)
The last equation also shows the way the isomorphism is constructed:
In fact, using Eq. (2.22), it is easy to check that if w = (a, b, c), then
⎡ ⎤
0 −c b
⎣
w = (a, b, c) ⇐⇒ W = c 0 −a⎦ . (2.23)
−b a 0
A(W) := {u ∈ V| Wu = o}.
Second-Rank Tensors 31
1
u·u= W·W (2.24)
2
is satisfied only by w and by its opposite −w. Because both these
vectors belong to A(W), choosing one of them corresponds to choosing an
orientation for E, see the next section. We always make our choice according
to Eq. (2.23), which fixes once and for all the isomorphism between V
and Skw(V) that corresponds to any vector w with one and only one
axial tensor W and vice versa, any skew tensor W with a unique axial
vector w.
It is worth noting that the above isomorphism between the vector spaces
V and Skw(V) implies that to any linear combination of vectors a and b
Downloaded from www.worldscientific.com
W2 u = WWu = w × (w × u) = w · uw − w · wu
(2.27)
= −(I − w ⊗ w)u ⇒ W2 = −(I − w ⊗ w);
1
WW = − |W|2 (I − w ⊗ w), (2.29)
2
while Eq. (2.24) can be generalized to any two axial couples w1 , W1 and
w2 , W2 :
1
w 1 · w2 = W1 · W2 .
2
The proof of these two last properties is rather easy and left to the reader.
We define the cross product of two vectors a and b the vector
a × b = Wa b,
Downloaded from www.worldscientific.com
a × b = (a2 b3 − a3 b2 , a3 b1 − a1 b3 , a1 b2 − a2 b1 ).
It is immediate to check that such a result can also be obtained using Ricci’s
alternator,
a × b = ijk aj bk ei , (2.30)
(αa + βb) × u = αa × u + βb × u,
u × (αa + βb) = αu × a + βu × b.
In fact, the first equation above is a consequence of Eq. (2.25), while the
second one is a simple application to axial tensors of the same definition of
tensor.
Three important results concerning the cross product are stated by the
following theorems.
Second-Rank Tensors 33
a × b = o.
a × b = Wa b = o ⇐⇒ b = ka, k ∈ R,
a × b · a = a × b · b = 0. (2.31)
Downloaded from www.worldscientific.com
Proof.
a × b · a = Wa b · a = b · Wa a = −b · Wa a = −b · o = 0,
a×b · b = Wa b · b = b · Wa b = −b · Wa b ⇐⇒ a × b · b = 0.
(b ⊗ a − a ⊗ b) ∈ Skew(V).
Then,
(b ⊗ a − a ⊗ b)(a × b) = a · a × b b − b · a × b a = 0
a × b = −b × a ∀a, b ∈ V. (2.32)
This property and, again, Theorem 16 lets us derive the formula for the
double cross product:
u × (v × w) = −(v × w) × u
= −(w ⊗ v − v ⊗ w)u = u · w v − u · v w. (2.33)
u × v · w = Wu v · w = −v · Wu w = −v · u × w = w × u · v, (2.34)
and similarly,
u × v · w = v × w · u.
Using this last result, we can obtain a formula for the norm of a cross
Downloaded from www.worldscientific.com
(a × b) · (a × b) = a × b · (a × b) = (a × b) × a · b
= −a × (a × b) · b = (−a · b a + a2 b) · b = b · (a2 I − a ⊗ a)b
(2.35)
= a2 b · (I − ea ⊗ ea )b = a2 b2 eb · (I − ea ⊗ ea )eb
= a2 b2 (1 − cos2 θ) = a2 b2 sin2 θ → |a × b| = ab sin θ.
Figure 2.1: Geometrical meaning of the cross and mixed products before (left) and after
(right) the application of a tensor L on the vectors u, v, w.
Second-Rank Tensors 35
β(u, v, w) = u × v · w
Lu × Lv · Lw = det L u × v · w. (2.36)
Following the interpretation given above for the absolute value of the mixed
product, we can conclude that | det L| can be interpreted as a coefficient of
volume expansion7 cf. again Fig. 2.1. A geometrical interpretation can then
be given to the case of a non-invertible tensor, i.e. of det L = 0: It crushes
a prism into a flat region (the three original vectors become coplanar, i.e.
linearly dependent).
The adjugate of L is the tensor
Downloaded from www.worldscientific.com
L∗ := (det L)L− .
e1 × e2 · e3 = 1,
e1 × e2 · e3 = −1.
7 This result is classical and fundamental for the analysis of deformation in continuum
mechanics.
8 It is evident that this is also true for one- and two-dimensional vector spaces.
36 Tensor Algebra and Analysis for Engineers
We2 = e1 ⊗ e3 − e3 ⊗ e1 ,
We3 = e2 ⊗ e1 − e1 ⊗ e2 .
2.11 Rotations
v̂ = Rv,
ˆ · v̂ = Ru · Rv = u · R Rv = u · v ⇐⇒ R R = I = RR .
u
R ∈ Orth(V) ⇐⇒ R−1 = R .
Theorem 18. Each tensor R ∈ Orth(V) has the eigenvalue ±1, with +1
for rotations.
Ru · Ru = λ2 u2 = u2 → λ2 = 1.
We must now prove that there exists at least one real eigenvector λ. To this
end, we consider the characteristic equation
f (λ) = λ3 + k1 λ2 + k2 λ + k3 = 0,
9 From the condition R R = I and through Eq. (2.16) and the theorem of Binet, we
the orientation of the space, like mirror symmetries do, see Section 2.12.
38 Tensor Algebra and Analysis for Engineers
f (λ1 ) = 0.
But
2 (λ2 ) + 2 (λ2 ) = 1
det R = 1 ⇐⇒ λ1 = 1.
This result actually means that the jth column of R is composed of the
components in the basis B of the vector êj of B̂. Because the two bases
are orthonormal, such components are the director cosines of the axes of B̂
with respect to B.
Second-Rank Tensors 39
with ϕ the rotation’s amplitude and W the axial tensor of the rotation
axis w.
⎡ ⎤
0 −w3 w2
W = ⎣ w3 0 −w1 ⎦,
−w2 w1 0
⎡ ⎤
w12 w1 w2 w1 w3
⎣
w ⊗ w = w1 w2 w22 w2 w3 ⎦ ,
w1 w3 w2 w3 w32
Downloaded from www.worldscientific.com
⎡ ⎤
cos ϕ − sin ϕ 0
⎣
R = sin ϕ cos ϕ 0⎦ . (2.43)
0 0 1
Second-Rank Tensors 41
Moreover,
⎡ ⎤ ⎡ ⎤
0 −1 0 0 0 0
W = ⎣ 1 0 0⎦ , w ⊗ w = ⎣0 0 0⎦,
0 0 0 0 0 1
⎡ ⎤
−1 0 0
W2 = −(I − w ⊗ w) = ⎣ 0 −1 0⎦.
0 0 0
Hence,
⎡ ⎤ ⎡ ⎤
1 0 0 0 −1 0
I + sin ϕW + (1 − cos ϕ)W2 = ⎣0 1 0⎦ + sin ϕ ⎣1 0 0⎦
0 0 1 0 0 0
⎡ ⎤ ⎡ ⎤ (2.44)
Downloaded from www.worldscientific.com
−1 0 0 cos ϕ − sin ϕ 0
+ (1 − cos ϕ) ⎣ 0 −1 0⎦ = ⎣ sin ϕ cos ϕ 0⎦ = R.
0 0 0 0 0 1
ei = R−1 e
ˆ i = R e
ˆi = Rhk
(êh ⊗ êk )êi = Rhk δki e
ˆh ,
i.e.
u
ˆk = Rki ui → û = R u.
i.e.
L̂mn = Rmi Rnj Lij = Rmi Lij Rjn → L̂ = R LR.
Downloaded from www.worldscientific.com
Rw = w
and then normalizing it, while the rotation amplitude ϕ can be found using
(2.39) along with (2.27): Because the trace of a tensor is invariant, we get
trR − 1
trR = 3 + (1 − cos ϕ)tr(−I + w ⊗ w) = 1 + 2 cos ϕ → ϕ = arccos .
2
It is interesting to consider the geometrical meaning of Eq. (2.39). For this
purpose, we apply Eq. (2.39) to a vector u, see Fig. 2.3:
u1 = R1 u.
u12 = R2 u1 = R2 R1 u.
Let us now suppose that we change the order of the rotations: R2 first and
then R1 . The final result will be the vector
u21 = R1 R2 u. (2.45)
Because the tensor product is not symmetric (i.e. it does not have the
commutativity property), generally,11
u12 = u21 .
In other words, the order of the rotations matters: Changing the order of the
rotations leads to a different final result. An example is shown in Fig. 2.4.
11 We have seen, in Theorem 12, that two tensors commute ⇐⇒ they are coaxial, i.e. if
they have the same eigenvectors. Because the rotation axis is always a real eigenvector
of a rotation tensor, if two tensors operate a rotation about different axes, they are not
coaxial. Hence, the rotation tensors about different axes never commute.
44 Tensor Algebra and Analysis for Engineers
w = u + v = v + u. (2.46)
This difference, which is a major point in physics, comes from the difference
in the operators: vectors for the displacements and tensors for the rotations.
Any rotation can be specified by the knowledge of three parameters.
This can be easily seen from Eq. (2.39): The parameters are the three
components of w that are not independent because
w = |w| = w12 + w22 + w32 = 1
(i) Physical angles: The rotation axis w is given through its spherical
coordinates ψ, the longitude, 0 ≤ ψ < 2π, and θ, the colatitude, 0 ≤
θ ≤ π, see Fig. 2.5, the third parameter being the rotation amplitude ϕ.
Then,
w2
w = (sin θ cos ψ, sin θ sin ψ, cos θ) → θ = arccos w3 , ψ = arctan ,
w1
Second-Rank Tensors 45
Figure 2.7: Euler’s rotations, as seen from the respective axes of rotation.
⎡ ⎤
1 0 0
Rθ = ⎣0 cos θ − sin θ⎦ ,
0 sin θ cos θ
Roθ = (R−1 −1
ψ ) Rθ Rψ = Rψ Rθ Rψ .
Second-Rank Tensors 47
Roϕ = (R−1 −1 −1 −1
ψ ) (Rθ ) Rϕ Rθ Rψ = Rψ Rθ Rϕ Rθ Rψ .
ˆ = Ru.
u
Downloaded from www.worldscientific.com
ˆ = Roϕ u,
u
u = Roθ u,
u = Rψ u.
Finally,
R = Roϕ Roθ Rψ = Rψ Rθ Rϕ R
θ Rψ Rψ Rθ Rψ Rψ = Rψ Rθ Rϕ ,
i.e. the global rotation tensor is also equal to the composition of the
three rotations, in the order of execution, if the three rotations are
expressed in their own particular bases. This result is general and not
bounded to the Euler’s rotations nor to three rotations.
48 Tensor Algebra and Analysis for Engineers
⎡ ⎤
cos ψ cos ϕ − sin ψ sin ϕ cos θ − cos ψ sin ϕ − sin ψ cos ϕ cos θ sin ψ sin θ
⎢ ⎥
R=⎢
⎣sin ψ cos ϕ + cos ψ sin ϕ cos θ − sin ψ sin ϕ + cos ψ cos ϕ cos θ − cos ψ sin θ ⎥
⎦.
sin ϕ sin θ cos ϕ sin θ cos θ
ˆ = R u = R
u
ϕ Rθ Rψ u
L̂ = R LR = R
ϕ Rθ Rψ LRψ Rθ Rϕ .
R = Rα Rβ Rγ
with
⎡ ⎤ ⎡ ⎤
1 0 0 cos β 0 − sin β
Rα = ⎣0 cos α − sin α⎦ , Rβ = ⎣ 0 1 0 ⎦,
0 sin α cos α sin β 0 cos β
⎡ ⎤
cos γ − sin γ 0
Rγ = ⎣ sin γ cos γ 0⎦,
0 0 1
so finally,
⎡ ⎤
cos β cos γ − cos β sin γ − sin β
⎢ ⎥
R=⎢
⎣ cos α sin γ − sin α sin β cos γ cos α cos γ + sin α sin β sin γ − sin α cos β ⎥
⎦.
sin α sin γ + cos α sin β cos γ sin α cos γ − cos α sin β sin γ cos α cos β
Second-Rank Tensors 49
Let us now consider the case of small rotations, i.e. |ϕ| → 0. In such a case,
sin ϕ ϕ, 1 − cos ϕ 0,
and
R I + ϕW,
Ru (I + ϕW)u = u + ϕw × u, (2.47)
i.e. by a skew tensor and not by a rotation tensor. The term (1−cos ϕ)W2 u
has disappeared, as it is a higher-order infinitesimal quantity, and the term
ϕw × u is orthogonal to u. Because ϕ → 0, the arc is approximated by
its tangent, the vector ϕw × u, see Fig. 2.8. Applying to Eq. (2.47) the
Downloaded from www.worldscientific.com
u1 = R1 u = (I + ϕ1 W1 )u = u + ϕ1 w1 × u,
u21 = R2 u1 = (I + ϕ2 W2 )u1 = u1 + ϕ2 w2 × u1
= u + ϕ1 w1 × u + ϕ2 w2 × u
+ ϕ1 ϕ2 w2 × (w1 × u).
precise, small rotations also do not commute.12 However, for small rota-
tions, ϕ1 ϕ2 is negligible with respect to ϕ1 and ϕ2 : In this approximation,
small rotations commute. We remark that approximation (2.47) gives, for
the displacements, a law that is quite similar to that of the velocities of the
points of a rigid body:
v = v0 + ω × (p − o).
This is quite natural because
dϕ
ω=,
dt
i.e. a small amplitude rotation can be seen as a rotation made with finite
angular velocity ω in a small time interval dt.
2.12 Reflexions
Downloaded from www.worldscientific.com
Let us consider now tensors S ∈ Orth(V) that are not rotations, i.e. such
that det S = −1. Let us call S an improper rotation. A particular improper
rotation whose all eigenvalues are equal to −1 is the inversion or reflexion
tensor:
SI = −I.
The effect of SI is to transform any basis B into the basis −B, i.e. with all the
basis vectors changed in orientation (or, equivalently, to change the sign of
all the components of a vector). In other words, SI changes the orientation
of the space. This is also the effect of any other improper rotation S that
can be decomposed into a proper rotation R followed by the reflexion SI 13 :
S = SI R. (2.48)
Let n ∈ S, then
SR = I − 2n ⊗ n (2.49)
is the tensor that operates the transformation of symmetry with respect to
a plane orthogonal to n. In fact,
SR n = −n, SR m = m ∀m ∈ V| m · n = 0.
12 This can happen for some vectors all the times when w1 · u = w2 · u, like for the case
of a vector u orthogonal to both w1 and w2 ; however, this is no more than a curiosity,
it has no importance in practice.
13 The application of Binet’s theorem shows immediately that det S = −1, while
while by the same definition of trace and through Eqs. (2.13) and (2.17),
Su = SI Ru = −Ru,
i.e. it changes the orientation of the rotated vector; this is not the case
when the same improper rotation transforms the vectors of a cross product:
The rotated vector result of the cross product does not cause a change in
orientation, i.e. the cross product is insensitive to a reflexion. That is why, to
be precise, the result of a cross product is not a vector but a pseudo-vector:
It behaves like a vector apart from the reflexions. For the same reason, a
scalar result of a mixed product (scalar plus cross product of three vectors)
is called a pseudo-scalar because in this case, the scalar result of the mixed
product causes a change in sign under a reflexion, which can be checked
easily.
L = U2 .
52 Tensor Algebra and Analysis for Engineers
L = ωi ei ⊗ ei
U2 = L.
V2 = L,
Downloaded from www.worldscientific.com
v = (U − λI)e,
we get
Uv = −λv ⇒ v = o ⇒ Ue = λe
Ve = λe ⇒ Ue = Ve
u · F Fu = (Fu) · (Fu) ≥ 0,
Second-Rank Tensors 53
F = RU = VR.
R = FU−1 ,
R is unique too.
Now, let F = VR be a left polar decomposition of F; by the same
procedure, we get
√
FF = V2 → V = FF ,
R = V−1 F.
Existence: Let
√
U= F F,
R = FU−1 .
all the eigenvalues of U are strictly positive), by the theorem of Binet, also
det R > 0. Then,
R R = (FU−1 ) (FU−1 ) = U−1 F FU−1 = U−1 U2 U−1
= I ⇒ R ∈ Orth(V)+ .
Now, let
V = RUR ,
then V ∈ Sym(V) and is positive definite, see Exercise 2.14, and
VR = RUR R = RU = F,
which completes the proof.
Downloaded from www.worldscientific.com
2.14 Exercises
1. Prove that
Lo = o ∀L ∈ Lin(V).
2. Prove that, if a straight line r has the direction of u ∈ S, then the tensor
giving the projection of a vector v ∈ V on r is u ⊗ u (the orthogonal
projector), while the one giving the projection on a direction orthogonal
to r is I − u ⊗ u (the complementary projector), see Fig. 2.9.
3. For any α ∈ R, a, b ∈ V and A, B ∈ Lin(V), prove that
(αA) = αA , (A + B) = A + B , (a ⊗ b)A = a ⊗ (A b).
4. Prove that
L + O = L ∀L ∈ Lin(V).
5. Prove that
trI = 3, trO = 0.
tr(AB) = tr(BA).
L · M = L · M, LM · N = L · NM = M · L N.
v · Lw = L · v ⊗ w ∀v, w ∈ V, L ∈ Lin(V).
10. Prove that Sym(V) and Skw(V) are orthogonal, i.e. prove that
A · B = 0 ∀A ∈ Sym(V), B ∈ Skw(V).
A · L = A · Ls ,
B · L = B · La .
1
λsph = trL
3
and that any u ∈ S is an eigenvector.
24. Prove that the eigenvalues λdev of Ldev are given by
λdev = λ − λsph ,
where λ is an eigenvalue of L.
Chapter 3
Fourth-Rank Tensors
Downloaded from www.worldscientific.com
57
58 Tensor Algebra and Analysis for Engineers
(ei ⊗ ej ⊗ ek ⊗ el )(ep ⊗ eq )
= (ek ⊗ el ) · (ep ⊗ eq )(ei ⊗ ej ) = δkp δlq (ei ⊗ ej ). (3.1)
Any fourth-rank tensor can be expressed as a linear combination (the
canonical decomposition):
L = Lijkl ei ⊗ ej ⊗ ek ⊗ el , i, j = 1, 2, 3,
where Lijkl s are the 81 Cartesian components of L with respect to B 4 . Lijkl s
are defined by the operation
(ei ⊗ ej ) · L(ek ⊗ el ) = (ei · ej ) · (Lpqrs ep ⊗ eq ⊗ er ⊗ es )(ek ⊗ el )
= (ei ⊗ ej ) · (Lpqrs δrk δsl ep ⊗ eq )
= Lpqrs δrk δsl δip δjq = Lijkl .
The components of a tensor dyad can be computed without any difficulty:
A ⊗ B = (Aij ei ⊗ ej ) ⊗ (Bkl ek ⊗ el ) = Aij Bkl ei ⊗ ej ⊗ ek ⊗ el
⇒ (A ⊗ B)ijkl = Aij Bkl
so that, in particular,
((a ⊗ b) ⊗ (c ⊗ d))ijkl = ai bj ck dl .
Concerning the identity of Lin(V),
Iijkl = (ei ⊗ el ) · I(ek ⊗ el ) = (ei ⊗ ej ) · (ek ⊗ el ) = ei · ek ej · el
= δik δjl ⇒ I = δik δjl (ei ⊗ el ⊗ ek ⊗ el ).
Fourth-Rank Tensors 59
Moreover,
Using this result and Eq. (3.1), we can determine the components of a
product of fourth-rank tensors:
Downloaded from www.worldscientific.com
For any two tensors A, B ∈ Lin(V), we call the conjugation product the
tensor A B ∈ Lin(V) defined by the operation
so that
= M · (B ⊗ A)L ⇒ (A ⊗ B) = B ⊗ A,
while, cf. Exercise 7, Chapter 2,
M · (A B) L = L · (A B)M
= L · AMB = A L · MB = M A L · B
= (M A L) · (B ) = L AM · B = AM · LB
= M · A LB = M · (A B )L ⇒ (A B)
= A B .
The property
(AB) = B A
can be proved in the same manner as the analogous property of the second-
rank tensors.
A tensor L ∈ Lin(V) is symmetric ⇐⇒ L = L . It is then evident that
L = L ⇒ Lijkl = Lklij ,
which are relations called major symmetries. These symmetries number 36
on the whole so that a symmetric fourth-rank tensor has 45 independent
components. Moreover,
A B = (A B) = A B ⇐⇒ A = A , B = B ,
A ⊗ B = (A ⊗ B) = B ⊗ A ⇐⇒ B = λA, λ ∈ R.
Let us now consider the case of a L ∈ Lin(V), such that
LA = (LA) ∀A ∈ Lin(V).
Fourth-Rank Tensors 61
Lijkl = Ljikl ,
which are relations called left minor symmetries: A tensor L having the
left minor symmetries has values in Sym(V). On the whole, the left minor
symmetries number 27. Finally, consider the case of a L ∈ Lin(V), such
that
LA = L(A ) ∀A ∈ Lin(V);
Lijkl = Lijlk ,
Downloaded from www.worldscientific.com
which are relations called minor right symmetries, whose total number
is also 27. It is immediate to recognize that if L has the minor right
symmetries, then
LW = O ∀W ∈ Skw(V).
We say that a tensor has minor symmetries if it has both the right and left
minor symmetries; the total number of minor symmetries is 45 because, as
can be easily checked, some of the left and right minor symmetries are the
same, so finally a tensor with the minor symmetries has 36 independent
components.
If L ∈ Lin(V) has major and minor symmetries, then the number of
independent symmetry relations is actually 60 (some minor and major
symmetries coincide), so in such a case, L depends upon 21 independent
components only. This is the case of the elasticity tensor.
Finally, the six Cauchy–Poisson symmetries1 are those of the type
Lijkl = Likjl .
The rule for computing the scalar product is hence always the same as was
already seen for vectors and second-rank tensors: All the indexes are to be
saturated.
In complete analogy with vectors and second-rank tensors, we say that
A is orthogonal to B ⇐⇒
A · B = 0,
1 1 1
Asph := trA I = I · A I = (I ⊗ I)A = Ssph A,
3 3 3
where
1
Ssph := I⊗I
3
is the spherical projector, i.e. the fourth-rank tensor that extracts from any
A ∈ Lin(V) its spherical part. Moreover,
where
Ddev := I − Ssph
is the deviatoric projector, i.e. the fourth-rank tensor that extracts from
any A ∈ Lin(V) its deviatoric part. It is worth noting that
I = Ssph + Ddev .
We remark that
Ssph = (Ssph ) .
We introduce now the tensor Is , restriction of I to A ∈ Sym(V). It can be
introduced as follows: ∀A ∈ Sym(V),
1
A= (A + A )
2
and
1 1
A = IA = (IA + IA ) = (Iijkl Akl + Iijkl Alk )(ei ⊗ ej ⊗ ek ⊗ el );
2 2
because A = A , there is insensitivity to the swap of indexes k and l, so
1
A= (Iijkl Akl + Iijlk Alk )(ei ⊗ ej ⊗ ek ⊗ el )
2
Downloaded from www.worldscientific.com
1
= (δik δjl + δil δjk )Akl (ei ⊗ ej ⊗ ek ⊗ el ).
2
Then, if we admit the interchangeability of indexes k and l, i.e. if we
postulate the existence of the minor right symmetries for I, then I = Is ,
with
1
Is = (δik δjl + δil δjk )(ei ⊗ ej ⊗ ek ⊗ el ).
2
It is apparent that
Iijkl
s
= Iklij
s
,
i.e. Is = (Is ) but also that
1
Iijkl
s
= (δil δjk + δik δjl ) = Ijikl
s
,
2
i.e. Is has also the minor left symmetries; in other words, Is has the major
and minor symmetries, like an elasticity tensor, while this is not the case
for I. In fact,
Iijkl = Ijilk = δik δjl = δil δjk = Ijikl = Iijlk .
Because Ssph and Ddev operate on Sym(V), it is immediate to recognize
that it is also
Ddev = Is − Ssph ⇒ Is = Ssph + Ddev .
It is worth noting that
(Ddev ) = (Is − Ssph ) = (Is ) − (Ssph ) = Is − Ssph = Ddev .
Fourth-Rank Tensors 65
sph 1 1
Dijkl
dev
= Iijkl
s
− Sijkl = (δik δjl + δil δjk ) − δij δkl →
2 3
1 1
Ddev = (δik δjl + δil δjk ) − δij δkl (ei ⊗ ej ⊗ ek ⊗ el ).
2 3
We remark that the result (2.9) implies that Ssph and Ddev are orthogonal
projectors, i.e. they project the same A ∈ Sym(V) into two orthogonal
subspaces of V, Sph(V) and Dev(V).
The tensor Ttrp ∈ Lin(V) defined by the operation
Ttrp A = A ∀A ∈ Lin(V)
trp
Tijkl = (ei ⊗ ej ) · Ttrp (ek ⊗ el ) = (ei ⊗ ej ) · (el ⊗ ek ) = δil δjk .
1
Ssym A = (A + A ) ∀A ∈ Lin(V),
2
1
Wskw A = (A − A ) ∀A ∈ Lin(V).
2
Also, Ssym and Wskw are orthogonal projectors because they project the
same A ∈ Lin(V) into two orthogonal subspaces of Lin(V): Sym(V) and
Skw(V), see Exercise 10, Chapter 2.
We prove now two properties of the projectors: ∀A ∈ Lin(V),
1 1
(Ssym + Wskw )A = (A + A ) + (A − A ) = A
2 2
= IA ⇒ Ssym + Wskw = I. (3.5)
Then,
1 1
(Ssym − Wskw )A = (A + A ) − (A − A ) = A
2 2
= Ttrp A ⇒ Ssym − Wskw = Ttrp . (3.6)
66 Tensor Algebra and Analysis for Engineers
and
so that
Ddev U = UDdev .
Downloaded from www.worldscientific.com
We ponder now how to rotate a fourth-rank tensor, i.e. what are the
components of
L = Lijkl ei ⊗ ej ⊗ ek ⊗ el
Lpqrs = Rpi
Rqj Rrk Rsl Lijkl .
R = R R,
2 Here,
the symbol R indicates the orthogonal conjugator of R and not the set of real
numbers.
68 Tensor Algebra and Analysis for Engineers
w = Uw,
L = ULU = (U U)L = UL, (3.8)
L = (U U)L(U U) = ULU .
Finally, we say that L ∈ Lin(V) or L ∈ Lin(V) is invariant under an
orthogonal transformation U if
ULU = L, ULU = L;
right multiplying both terms by U or by U and through Eq. (3.7), we get
that L or L are invariant under U ⇐⇒
UL = LU, UL = LU,
i.e. ⇐⇒ L and U or L and U commute. This relation allows, for example,
the analysis of material symmetries in anisotropic elasticity.
If a tensor is invariant under any orthogonal transformation, i.e. if the
previous equations hold true ∀U ∈ Orth(V), then the tensor is said to be
isotropic. A general result3 is that a fourth-rank tensor L is isotropic ⇐⇒
there exist two scalar functions λ, μ such that
LA = 2μA + λtrA I ∀A ∈ Sym(V).
The reader is referred to the book by Gurtin (see references) for the proof
of this result and for a deeper insight into isotropic functions.
3 Actually,this is quite a famous result in classical elasticity, the Lamé’s equation, defining
an isotropic elastic material.
Fourth-Rank Tensors 69
σ = Eε.
⎧ ⎫ ⎧ ⎫
⎪ σ1 = σ11 ⎪ ⎪ ε1 = ε11 ⎪
⎪ σ2 = σ22 ⎪ ⎪ ε2 = ε22 ⎪
⎨ ⎬ ⎨ ⎬
σ3 =√σ33 ε3 =√ε33
{σ} = , {ε} = .
⎪ σ4 = √2σ23 ⎪ ⎪ ε4 = √2ε23 ⎪
⎪ σ5 = 2σ31 ⎪ ⎪ ε5 = 2ε31 ⎪
⎩ √ ⎭ ⎩ √ ⎭
σ6 = 2σ12 ε6 = 2ε12
is equivalent to the tensor form of the Hooke’s law, and all the operations
can be done by the aid of classical matrix algebra,6 e.g. the computation
of the inverse of E, the compliance tensor.
An important operation is the expression of tensor U in Eq. (3.8) in
the Kelvin formalism; some tedious but straightforward operations give the
result:
⎡ √ √ √ ⎤
2 2 2
U11 U12 U13 2U12 U13 2U13 U11 2U11 U12
⎢ ⎥
⎢ 2 2 2 √ √ √ ⎥
U21 U22 U23 2U22 U23 2U23 U21 2U21 U22
⎢ ⎥
⎢ 2 2 2 √ √ √ ⎥
U31 U32 U33 2U32 U33 2U33 U31 2U31 U32
[U ] = ⎢ √ √ √
⎥.
⎢ 2U U 2U22 U32 2U23 U33 U23 U32 + U22 U33 U33 U21 + U31 U23 U31 U22 + U32 U21 ⎥
21 31
⎢ ⎥
√ √ √
⎢ 2U U 2U32 U12 2U33 U13 U32 U13 + U33 U12 U31 U13 + U33 U11 U31 U12 + U32 U11 ⎥
31 11
⎣ ⎦
√ √ √
2U11 U21 2U12 U22 2U13 U23 U12 U23 + U13 U22 U11 U23 + U13 U21 U11 U22 + U12 U21
[U ][U ] = [U ] [U ] = [I],
i.e. that [U ] is an orthogonal matrix in R6 . Of course,
[R] = [U ]
is the matrix that in the Kelvin formalism represents the tensor R = U .
The change of basis for σ and ε are hence done through the relations
{σ } = [U ]{σ}, {ε } = [U ]{ε},
which when applied to Eq. (3.9) give
{σ} = [E]{ε} → [U ] {σ } = [E][U ] {ε } → {σ } = [U ][E][U ] {ε },
i.e. in the basis B ,
{σ } = [E ]{ε },
where
[E ] = [U ][E][U ] = [R] [E][R]
is the matrix representing E in B in the Kelvin formalism. Though it is
possible to give the expression of the components of [E ], they are so long
that they are omitted here.
6 Mehrabadi and Cowin have shown that the Kelvin formalism transforms second- and
fourth-rank tensors on R3 into vectors and second-rank tensors on R6 (M. M. Mehrabadi
and S. C. Cowin: Eigentensors of linear anisotropic elastic materials. Q. J. Mech. Appl.
Math., 43, 15–41, 1990).
Fourth-Rank Tensors 71
7 G.Verchery: Les invariants des tenseurs d’ordre 4 du type de l’élasticité, Proc. Colloque
EUROMECH 115, 1979.
8 A detailed presentation of the method can be found in the work by Vannucci:
3.10 Exercises
14. Prove the results in Eqs. (3.5) and (3.6) using the components.
15. Show that
Ssph · Ssph = 1,
Ddev · Ddev = 5,
Ssph · Ddev = 0.
16. Make explicit the orthogonal conjugator SR of the tensor SR in Eq.
(2.49).
17. Using the polar formalism, it can be proved that the material symmetry
conditions in plane elasticity are all condensed into the equation
R0 R1 sin 4(Φ0 − Φ1 ) = 0;
determine the different types of possible elastic symmetries.
Downloaded from www.worldscientific.com
75
76 Tensor Algebra and Analysis for Engineers
tensors:
v = v(t) : [a, b] → V,
L = L(t) : [a, b] → Lin(V),
L = L(t) : [a, b] → Lin(V).
Mathematically, a curve is a function that lets correspond to a real value t
(the parameter) in a given interval, an element of a space: E, V, Lin(V),
or L(V).
|v(t)|
lim = 0,
t→t0 |g(t)|
and we write
v(t) = o(g(t)) for t → t0 .
A similar definition can be given for a curve of tensors of any rank. We then
say that the curve v is differentiable in t0 ∈]a, b[ ⇐⇒ ∃v ∈ V such that
v(t) − v(t0 ) = (t − t0 )v (t0 ) + o(t − t0 ).
We call v (t0 ) the derivative1 of v at t0 . Applying the definition of
derivative to v , we define the second derivative v of v and recursively
all the derivatives of higher orders. We say that v is of class Cn if it is
continuous with its derivatives up to the order n; if n ≥ 1, v is said to be
smooth. A curve v(t) of class Cn is said to be regular if v = o ∀t. Similar
definitions can be given for curves in E, Lin(V), and Lin(V), thus defining
derivatives of points and tensors. We remark that the derivative of a curve
in E, defined as a difference of points, is a curve in V (we say, in short,
that the derivative of a point is a vector). For what concerns tensors, the
derivative of a tensor of rank r is a tensor of the same rank.
Let u, v be curves in V, L, M curves in Lin(V), L, M curves in Lin(V),
and α a scalar function, with all of them defined and at least of class C1
1 The dv
derivative is also written as , v,t or also as v̇, with the last symbol being
dt
usually reserved, in physics, to the case where t is time. For the sake of brevity, we omit
to indicate the derivative of v at t0 as v (t0 ), writing simply v .
Tensor Analysis: Curves 77
on [a, b]. The same definition of the derivative of a curve gives the following
results, whose proof is left to the reader:
(u + v) = u + v ,
(αv) = α v + αv ,
(u · v) = u · v + u · v ,
(u × v) = u × v + u × v ,
(u ⊗ v) = u ⊗ v + u ⊗ v ,
(L + M) = L + M ,
(αL) = α L + αL ,
(Lv) = L v + Lv ,
Downloaded from www.worldscientific.com
(LM) = L M + LM ,
(L · M) = L · M + L · M ,
(L ⊗ M) = L ⊗ M + L ⊗ M ,
(L M) = L M + L M ,
(L + M) = L + M ,
(αL) = α L + αL ,
(LL) = L L + LL ,
(LM) = L M + LM ,
(L · M) = L · M + L · M .
R; the equation
⎧
⎨ r1 = r1 (t)
r(t) = ri (t)ei = r1 (t)e1 + r2 (t)e2 + r3 (t)e3 → r = r2 (t)
⎩ 2
r3 = r3 (t)
It is noted that the choice of the parameter is not unique: The equation
p = p[τ (t)] still represents the same curve p = p(t) through the change of
parameter τ = τ (t).
The definition given above for the derivative of a curve of points p = p(t)
in t = t0 is equivalent to the following one2 (probably more familiar to the
reader):
Downloaded from www.worldscientific.com
(L ) = L ,
(L ) = L ,
80 Tensor Algebra and Analysis for Engineers
while for any invertible tensor L, we have (we state the following results
without proof3 )
(L−1 ) = −L−1 L L−1 ,
(det L) = det L tr(L L−1 ) = det L L · L−1 = det L L · L− .
Let Q(t) : R → Orth(V) be a differentiable function. We call spin tensor
the tensor S(t) defined as
S(t) := Q (t)Q (t).
Then, we have the following.4
Theorem 24 (Characterization of the spin tensor). S(t) ∈
Skw(V) ∀t ∈ R.
QQ = I ⇒ (QQ ) = Q Q + QQ = I = O ⇒ QQ = −Q Q ,
so
S = (Q Q ) = QQ = −Q Q = −S.
Because
r(t) = p(t) − o, r (t) = (p(t) − o) = p (t),
3 The interested reader can find these proofs in the text by Gurtin, see the suggested
texts.
4 The spin tensor and the following result are of importance in kinematics: If t is time
and Q(t) ∈ OrthV + , then the axial vector of S(t) is ω(t), the angular velocity.
Tensor Analysis: Curves 81
we also get
Downloaded from www.worldscientific.com
t
p(t) = p(a) + p (t∗ ) dt∗ .
a
The length σ of the polygonal line whose vertices are the points r(ti ) is
hence
n
σ = |r(ti ) − r(ti−1 )|.
i=1
:= sup σ .
σ
whence
b
≤ |r (t)|dt. (4.3)
a
Because r (t) is continuous on [a, b], ∀ε > 0 ∃δ > 0 such that |t − t| < δ ⇒
|r (t) − r (t)| < ε. Let t ∈ [ti−1 , ti ] and σmax < δ, which is always possible
Downloaded from www.worldscientific.com
|r (t)| ≤ |r (t) − r (ti )| + |r (ti )| < ε + |r (ti )|,
whence
ti
ti t
i
|r (t)|dt <
|r (ti )|dt + ε(ti − ti−1 ) = r (ti )dt + ε(ti − ti−1 )
ti−1 ti−1 ti−1
t t
i i
≤ r (t)dt + (r (ti ) − r (t))dt + ε(ti − ti−1 )
ti−1 ti−1
≤ |r(ti ) − r(ti−1 )| + 2ε(ti − ti−1 ).
Summing up over all the intervals [ti−1 , ti ], we get
b
|r (t)|dt ≤ σ + 2ε(b − a) ≤ + 2ε(b − a),
a
Theorem 26. The length of a curve does not depend upon its
parameterization.
For a plane curve y = f (x), we can always put t = x, which gives the
parametric equation
or in vector form,
r(t) = t e1 + f (t) e2 ,
We define the tangent vector τ (t) to a regular curve p = p(t) as the vector
p (t)
τ (t) := .
|p (t)|
By the definition of the derivative, this unit vector is always oriented as
the increasing values of t; hence, the straight line tangent to the curve in
p0 = p(t0 ) has the equation
q(t̄) = p(t0 ) + t̄ τ (t0 ).
If the curvilinear abscissa s is chosen as a parameter for the curve through
the change of parameter s = s(t), we get
p (t) p [s(t)] 1 dp(s) ds(t) dp(s)
τ (t) = = = = → τ (s) = p (s).
Downloaded from www.worldscientific.com
|p (t)|
|p [s(t)]| s (t) ds dt ds
(4.6)
So, if the parameter of the curve is s, the derivative of the curve is τ , i.e.
it is automatically a unit vector. The above equation, in addition, shows
that the change of parameter does not change the direction of the tangent
because it is only a scalar, the derivative of the parameter’s change, that
multiplies the vector. Nevertheless, in general, a change of parameter can
change the orientation of the curve.
Because the norm of τ is constant, its derivative is a vector orthogonal
to τ , see Eq. (4.1). That is why we call principal normal vector to a curve
the unit vector
τ (t)
ν(t) := . (4.7)
|τ (t)|
ν is defined only on the points of the curve where τ = o, which implies
that ν is not defined on the points of a straight line. This simply means
that there is not, among the infinite unit normal vectors to a straight line,
a normal with special properties, a principal one, uniquely linked to τ .
Unlike τ , whose orientation changes with the choice of the parameter, ν
is an intrinsic local characteristic of the curve: It is not affected by the choice
of the parameter. In fact, by the same definition, ν does not depend upon
the reference frame; then, because the direction of τ is also independent of
the parameter’s choice, the only factor that could affect ν is the orientation
of the curve, which depends upon the parameter. But a change in the
orientation affects, in (4.7), both τ and the sign of the increment dt so
Tensor Analysis: Curves 85
that τ (t) = dτ /dt does not change, nor does ν, which is hence an intrinsic
property of the curve.
The vector
τ × ν · β = 1,
plane (ν, β) the normal plane, and the plane (β, τ ) the rectifying plane,
see Fig. 4.4. The osculating plane is particularly important: If we consider
a plane passing through three nonaligned points of the curve, when these
points become closer and closer, still remaining on the curve, the plane
tends to the osculating plane. The osculating plane at a point of a curve is
the plane that better approaches the curve near the point. A plane curve
is entirely contained in the osculating plane, which is fixed.
The principal normal ν is always oriented toward the part of the space
with respect to the rectifying plane where the curve is; in particular, for
a plane curve, ν is always directed toward the concavity of the curve. To
show that, it is sufficient to prove that the vector p(t + ε) − p(t) forms with
so that
1 dτ |τ (t)|
c(s) = |τ (s)| = = . (4.8)
|p (t)| dt |p (t)|
so that, finally,
|p × p |
c= . (4.9)
|p |3
Applying this last formula to a plane curve p(t) = (x(t), y(t)), we get
|x y − x y |
c= 3 ,
(x2 + y 2 ) 2
and if the curve is given in the form y = y(x) so that the parameter t = x,
then we obtain
|y |
c= 3 .
(1 + y 2 ) 2
This last formula shows that if |y | 1, then
c |y |.
This result is fundamental to the linearized (infinitesimal) theory of beams,
plates, and shells.
dβ
so that is necessarily parallel to ν. We then set
ds
dβ
= ϑν,
ds
which is the second Frenet–Serret formula. The scalar ϑ(s) is called the
torsion of the curve in p = p(s). So, we see that the variation in β per unit
length is a vector parallel to ν and proportional to the torsion of the curve.
We can now find the variation in ν per unit length of the curve:
dν d(β × τ ) dβ dτ
= = ×τ +β× = ϑ ν × τ + c β × ν,
ds ds ds ds
so finally
dν
= −c τ − ϑ β,
ds
which is the third Frenet–Serret formula: the variation in ν per unit length
of the curve is a vector of the rectifying plane.
The three formulae of Frenet–Serret (discovered independently by
J. F. Frenet in 1847 and by J. A. Serret in 1851) can be condensed into the
symbolic matrix product
⎧ ⎫ ⎡ ⎤⎧ ⎫
⎨τ ⎬ 0 c 0 ⎨τ ⎬
ν = ⎣ −c 0 −ϑ ⎦ ν .
⎩ ⎭ ⎩ ⎭
β 0 ϑ 0 β
The matrix in the equation above is called the matrix of Cartan, and it is
skew.
90 Tensor Algebra and Analysis for Engineers
We have introduced the torsion of a curve in the previous section, with the
second formula of Frenet–Serret. The torsion measures the deviation of a
curve from flatness: If a curve is planar, it belongs to the osculating plane,
and β, which is perpendicular to the osculating pane, is hence a constant
vector. So, its derivative is null, and by the second Frenet–Serret formula,
ϑ = 0.
Conversely, if ϑ = 0 everywhere, β is a constant vector, and hence, the
osculating plane does not change and the curve is planar. So, we have that
a curve is planar if and only if the torsion is null ∀p(s).
Using the Frenet–Serret formulae in the expression of p (s), we get a
formula for the torsion:
dp ds
Downloaded from www.worldscientific.com
so that, finally,
p × p · p
ϑ=− .
|p × p |2
1 1
p(s + ε) − p(s) = ε p + ε2 p + ε3 p + o(ε3 )
2 6
1 2 1 3
= ετ + ε cν + ε (cν) + o(ε3 )
2 6
1 2 1 3
= ετ + ε cν + ε (c ν − c2 τ − c ϑβ) + o(ε3 )
2 6
1
⇒ (p(s + ε) − p(s)) · β = − ε3 c ϑ + o(ε3 ).
6
The above dot product determines if the point p(s + ε) is located, with
respect to the osculating plane, on the side of β or on the opposite one, see
Fig. 4.6: If following the curve for increasing values of s, ε > 0, the point
passes into the semi-space of β from the opposite one, because 1/6 c ε3 > 0,
it will be ϑ < 0, while in the opposite case, it will be ϑ > 0.
This result is intrinsic, i.e. it does not depend upon the choice of the
parameter, hence of the positive orientation of the curve; in fact, ν is
intrinsic, but changing the orientation of the curve, τ , and hence β, change
in orientation.
6 The word osculating comes from the latin word osculo, which means to kiss; an
osculating sphere or circle or plane is a geometric object that is very close to the curve,
as close as two lovers are in a kiss.
92 Tensor Algebra and Analysis for Engineers
1
(p(s) − qs ) · ν = − = −ρ,
c
c ρ
(p(s) − qs ) · β = − 2 = ,
c ϑ ϑ
and finally,
ρ
qs = p + ρ ν − β, (4.12)
ϑ
so the center of the sphere belongs to the normal plane; the sphere is not
defined for a plane curve. The quantity ρ is the radius of curvature of the
curve, which is defined as
1
ρ= .
c
The radius of the osculating sphere is
2
ρ
ρs = |p − qs | = ρ2 + .
ϑ
The intersection between the osculating sphere and the osculating plane at
the same point p is the osculating circle. This circle has the property of
sharing the same tangent in p with the curve, and its radius is the radius of
curvature, ρ. From Eq. (4.12), we get the position of the osculating circle
center q:
q = p + ρ ν. (4.13)
An example can be seen in Fig. 4.7, where the osculating plane, circle, and
sphere are shown for a point p of a conical helix.
Tensor Analysis: Curves 93
Figure 4.7: Osculating plane, circle, and sphere for a point p of a conical helix.
Downloaded from www.worldscientific.com
ρ c
= − 2 = 0,
ϑ c ϑ
i.e. when the curvature is constant.
For any plane curve γ(s), the center of the osculating circle q describes
a curve δ(σ) that is called the evolute of γ(s) (s and σ are curvilinear
abscissa). A point q of the evolute is then given by Eq. (4.13). We call
involute of a curve γ(s) a curve μ(σ) whose evolute is γ(s). We call the
envelope of a family of plane curves ϕ(s, κ), κ ∈ R being a parameter, a
curve that is tangent in each of its points to the curve of ϕ(s, κ) passing
through that point.
Let us consider the evolute δ(σ) of a curve γ(s); the tangent to δ(σ) is
the vector, cf. Eq. (4.13),
dq dq ds
τδ = = .
dσ ds dσ
But, cf. again Eq. (4.13) and the Frenet–Serret formulae,
dq dp dρ dν dρ dρ
= + ν +ρ =τ + ν −ρ c τ = ν,
ds ds ds ds ds ds
94 Tensor Algebra and Analysis for Engineers
so
dq dρ ds
τδ = = ν.
dσ ds dσ
Because
dq = |ν| = 1,
dσ
then
dρ ds dρ dσ
=1 ⇒ =
ds dσ ds ds
and
τδ = ν.
The evolute, δ(σ), of γ(s) is hence the envelope of its principal normals
Downloaded from www.worldscientific.com
ν(s).
This result helps us in finding the equation of the involute μ(σ) of a
curve γ(s); let p = p(s) be a point of γ(s); then, if b ∈ μ(σ), it must be
that
(b − p) · ν = 0,
with f = f (s) a scalar function of s; we remark that b = b(s), i.e. the arc
length s of γ(s) is the parameter also for μ(s), but in general, σ =s. Upon
differentiation, we get
1 + f (s) = 0 ⇒ f (s) = a − s, a ∈ R.
The curvature, c(s), and the torsion, ϑ(s), are the only differential param-
eters that completely describe a curve. In other words, given two functions
c(s) and ϑ(s), then a curve exists with such a curvature and torsion (we
remark that there are no conditions bounding these parameters). This is
proved in the following.
Theorem 27 (Bonnet’s theorem). Given two scalar functions c(s) ∈C1
and ϑ(s) ∈C0 , there always exists a unique curve γ ∈C3 whose curvilinear
abscissa is s, curvature c(s), and torsion ϑ(s).
Proof. Let
⎛ ⎞
τ
e = ⎝ν ⎠
β
be the column vector whose elements are the vectors of the Frenet–Serret
basis. Then,
de(s)
= C(s)e(s), (4.14)
ds
96 Tensor Algebra and Analysis for Engineers
with
⎡ ⎤
0 c(s) 0
C(s) = ⎣ −c(s) 0 −ϑ(s) ⎦
0 ϑ(s) 0
the matrix of Cartan. Adding the initial condition
⎛ ⎞
e1
e(0) = ⎝ e2 ⎠,
e3
we have a Cauchy problem for the basis e(0). As known, such a problem
admits a unique solution, i.e. we can associate to c(s) and ϑ(s) a family of
bases e(s) (that are orthonormal because if one of them were not so, the
Cartan’s matrix should not be skew). Call τ (s) the first vector of the basis
Downloaded from www.worldscientific.com
p(s) is the curve we are looking for (it depends upon an arbitrary point
p0 , i.e. upon an inessential rigid displacement). In fact, because |τ | = 1, s
is the curvilinear abscissa of the curve. Then, it is sufficient to write the
Frenet–Serret equations identifying them with system (4.14).
The coordinates of a point p(s) close to p0 in the basis (τ (0), ν(0), β(0))
are hence
1
p1 (s) = s − c2 (0)s3 + o(s3 ),
6
1 1
p2 (s) = c(0)s2 + c (0)s3 + o(s3 ),
2 6
1
p3 (s) = − c(0)ϑ(0)s3 + o(s3 ).
6
The projections of the curve onto the planes of the Frenet–Serret basis hence
have, close to p0 (i.e. retaining the first non-null term in the expressions
above), the following equations:
• On the osculating plane,
⎧
⎨p1 (s) = s,
Downloaded from www.worldscientific.com
4.12 Exercises
1. Using the same definition of the derivative of a curve, prove the relations
in Section (4.2).
2. Prove the relations in Eq. (4.2).
3. The curve whose polar equation is
r = a θ, a ∈ R,
is an Archimedes’ spiral, Fig. 4.9(a). Find its curvature c(θ) and its
length (θ), and prove that any straight line passing through the origin
is divided by the spiral in segments of constant length 2π a (that is
why the Archimedes’ spiral is used to record disks).
4. The curve whose polar equation is
Downloaded from www.worldscientific.com
r = aebθ , a, b ∈ R,
(a) (b)
two successive intersections with the helix. Prove then the Bertrand’s
theorem: A curve is a cylindrical helix if and only if the ratio c/ϑ =
const. Finally, prove that for the above circular helix, there are two
constants A and B such that
p × p = Au(θ) + Be3 ,
with
u = sin ωθe1 − cos ωθe2 ;
then, find A and B.
Figure 4.10: The involute of the circle and its evolute, the circle.
100 Tensor Algebra and Analysis for Engineers
7. Find the equation of the cycloid, i.e. of the curve that is the trace of
a point of a circle of radius r rolling without slipping on a horizontal
axis, see Fig. 4.12. Calculate the length of the cycloid for a complete
round of the circle, determine its curvature, and show that the evolute
of the cycloid is the cycloid itself (Huygens, 1659).
8. The planar curve whose parametric equation is
the curve along which an object moves, under the influence of friction,
when pulled on a horizontal plane by a line segment attached to a
tractor that moves at a right angle to the initial line between the object
and the puller at an infinitesimal speed, see Fig. 4.13. Show that the
length of the tangent to the tractrix between the points on the tractrix
itself and the axis x is constant ∀t, calculate the length of the curve
between t1 and t2 , calculate the curvature of the tractrix, and finally
show that its evolute is the catenary.
10. For the curve whose cylindrical equation is
r = 1,
z = sin θ,
find the highest curvature and determine whether or not it is planar.
11. Let p = p(t) be the path of a moving particle of mass m, with t being the
time. Define the velocity and the acceleration of p as, respectively, the
first and second derivative of p with respect to t. Decompose these two
vectors in the Frenet–Serret basis and interpret physically the result.
Recalling the second Newton’s principle of mechanics, what about the
forces on p?
Downloaded from www.worldscientific.com
103
104 Tensor Algebra and Analysis for Engineers
dv
= gradv n.
dn
Let ψ be a scalar of vector field of class C2 at least. Then, the laplacian
Δψ of ψ is defined by
Δψ := div(gradψ).
By the linearity of the trace, and hence of the divergence, we see easily that
the laplacian of a vector field is the vector field whose components are the
laplacian of each corresponding component of the field. A field is said to be
harmonic on Ω if its laplacian is null ∀p ∈ Ω.
The definitions given above for the differentiable field, gradient, and
class C1 can be repeated verbatim for a deformation f (p) : Ω → E.
but also,
(ϕv)(p + u) = ϕ(p + u)v(p + u) = (ϕ + gradϕ · u + o(u))
× (v + gradv u + o(u))
= ϕv + ϕgradv u + gradϕ · u v + o(u)
= ϕv + (ϕ gradv + v ⊗ gradv)u + o(u),
so comparing the two results, we get
grad(ϕv) = ϕ gradv + v ⊗ gradv.
(iii) In the same way,
(v · w)(p + u) = v · w + grad(v · w)u + o(u),
but also,
(v · w)(p + u) = v(p + u) · w(p + u) = (v + gradv u + o(u))
· (w + gradw u + o(u))
= v · w + v · (gradw u) + (gradv u) · w + o(u)
= v · w + ((gradw) v + (gradv) w) · u + o(u),
whence, by comparison of the two results,
grad(v · w) = (gradw) v + (gradv) w.
1 Forthe sake of brevity, we omit to indicate the point p, e.g. we simply write ϕ for ϕ(p)
and gradϕ for gradϕ(p).
Tensor Analysis: Fields 107
Another important result2 relating the gradient and the curl of a vector
field is as follows.
so that
1
(curlv) × v = (gradv)v − gradv2 ,
2
whence we obtain the thesis.
The proof of the following properties of the gradient are left to the reader
as an exercise:
grad(v · w) = (gradw)v + (gradv)w + v × curlw + w × curlv,
grad(u · v w) = (u · v)gradw + (w ⊗ u)gradv + (w ⊗ v)gradu, (5.2)
2
gradv · gradv = div((gradv)v − (divv)v) + (divv) .
so finally, comparing the last three results (all the subscripts are dummy
indexes, so their denomination is inessential),
div(v × w) = w · curlv − v · curlw.
The divergence has also the following properties:
div(gradv ) = grad(divv),
div((gradv)v) = gradv · gradv + v · grad(divv), (5.3)
div(ϕLv) = ϕL · gradv + ϕv · divL + Lv · gradϕ,
whose proofs can be a good exercise for the reader.
The relations of the gradient and divergence with the curl are given by
the following.
Theorem 31. Let ϕ and v be the scalar and vector fields of class C2 .
Downloaded from www.worldscientific.com
Then:
(i) div(curlv) = 0,
(ii) curl(gradϕ) = o.
Proof. (i) Using again the Ricci’s alternator to represent the cross product,
div(curlv) = div(ijk vk,j ei ) = ijk vk,j divei + ijk vk,ji = ijk vk,ji
= v3,21 + v1,32 + v2,13 − v2,31 − v3,12 − v1,23 = 0.
(ii) In a similar manner,
curl(gradϕ) = ijk ϕ,kj ei = ϕ32 + ϕ,13 + ϕ,21 − ϕ,23 − ϕ,31 − ϕ,12 = 0.
The following theorem gives an interesting relation between the curl of a
vector and the divergence of its axial tensor.
Theorem 32 (Curl of an axial vector). Let w be a differentiable vector
field and W its axial tensor field. Then,
curlw = −divW.
The way the curl of a curl3 is computed is given by the following theorem.
Proof. Using properties (iv) and (v) of Theorem 30, along with the first
of Eq. (5.3), ∀a = const. ∈ V, we get
− div((gradv) a)
= (gradv) · grada + a · div(gradv) − gradv · grada
− a · div(gradv)
= a · (div(gradv) − div(gradv)) = a · (grad(divv) − Δv),
whence, by comparison,
The proof of the following properties of the curl can be obtained using the
above results, and it is a good exercise:
3 This relation is useful in fluid mechanics for writing the vorticity equation.
Tensor Analysis: Fields 111
This lemma is fundamental for proving the three forms of the Gauss
theorem, which is of paramount importance in many fields of mathematical
physics.
whence, by comparison,
ϕn dA = gradϕ dV.
∂Ω Ω
but also,
tr gradv dV = tr(gradv)dV = divv dV,
Ω Ω Ω
so by comparison,
v · n dA = divv dV.
∂Ω Ω
Tensor Analysis: Fields 113
but also,
div(L a)dV = (divL) · a + L · grada dV = a · divL dV,
Ω Ω Ω
The following classical theorems on fields are recalled here without proof.
curlv = o ⇐⇒ v = gradϕ,
gradf = f,i ei ,
gradv = vi,j ei ⊗ ej ,
divv = vi,i ,
divL = Lij,j ei , (5.6)
Δf = f,ii ,
Δv = Δvi ei = vi,jj ei ,
Downloaded from www.worldscientific.com
curlv = (v3,2 − v2,3 )e1 + (v1,3 − v3,1 )e2 + (v2,1 − v1,2 )e3 .
∂· ∂· ∂· ∂·
∇ := ei = e1 + e2 + e3 , (5.7)
∂xi ∂x1 ∂x2 ∂x3
gradf = ∇f,
divv = ∇ · v,
curlv = ∇ × v,
Δf = ∇2 f.
5 In
what follows, and also in the following sections, f, v, L are, respectively, scalar, vector,
and tensor fields.
116 Tensor Algebra and Analysis for Engineers
or conversely,
x1 = ρ cos θ,
x2 = ρ sin θ, (5.9)
x3 = z.
p − o = ρeρ + zez ,
and the rotation tensor transforming the Cartesian basis, {e1 , e2 , e3 }, into
the cylindrical one, {eρ , eθ , ez }, is
⎡ ⎤
cos θ − sin θ 0
⎢ ⎥
Q = ⎣ sin θ cos θ 0⎦,
0 0 1
Tensor Analysis: Fields 117
so the relations between the vectors of the Cartesian and the cylindrical
basis are
The question is: How can we express the differential operators in the
(moving) frame {p; eρ , eθ , ez }? To this end, we can proceed as follows: From
Eq. (5.8),
⎧ x x
⎪ f,1 = f,ρ 1 − f,θ 2 ,
⎪ ρ ρ2
∂ρ ∂θ ∂z ⎨
f,i = f,ρ + f,θ + f,z → x2 x1
⎪ f,2 = f,ρ ρ + f,θ ρ2 ,
(5.11)
∂xi ∂xi ∂xi
⎪
⎩
f,3 = f,z .
1
gradf = f,ρ eρ + f,θ eθ + f,z ez .
ρ
finally6
1
gradv = vρ,ρ (eρ ⊗ eρ ) + (vρ,θ − vθ )(eρ ⊗ eθ ) + vρ,z (eρ ⊗ ez )
ρ
1
+ vθ,ρ (eθ ⊗ eρ ) + (vθ,θ + vρ )(eθ ⊗ eθ ) + vθ,z (eθ ⊗ ez )
ρ
1
+ vz,ρ (ez ⊗ eρ ) + vz,θ (ez ⊗ eθ ) + vz,z (ez ⊗ ez ),
ρ
or, in matrix form,
⎡ 1 ⎤
v (vρ,θ − vθ ) vρ,z
⎢ ρ,ρ ρ ⎥
⎢ 1 ⎥
gradv = vθ,ρ (vθ,θ + vρ ) vθ,z .
⎢ ρ ⎥
⎣ 1 ⎦
vz,ρ vz,θ vz,z
ρ
By the definition of divergence, we get immediately
1
divv = vρ,ρ + (vθ,θ + vρ ) + vz,z . (5.14)
ρ
Now, from Eq. (5.6), we see that divL is the vector whose components are
the divergence of the rows of the matrix representing L. So, we need first
6 Though straightforward, the details of the calculations for this formula, as for the
following ones, are particularly long and tedious, which is why they are omitted here;
however, they are a good exercise for the reader.
Tensor Analysis: Fields 119
⎧
⎪L11 = − sin θ(Lρθ cos θ − Lθθ sin θ) + cos θ(Lρρ cos θ − Lθρ sin θ),
⎪
⎪
⎪L12 = cos θ(Lρθ cos θ − Lθθ sin θ) + sin θ(Lρρ cos θ − Lθρ sin θ),
⎪
⎪
⎪L13 = Lρz cos θ − Lθz sin θ,
⎪
⎪
⎪L21 = − sin θ(Lθθ cos θ + Lρθ sin θ) + cos θ(Lθρ cos θ + Lρρ sin θ),
⎪
⎨
LCart = QLcyl Q → L22 = cos θ(Lθθ cos θ + Lρθ sin θ) + sin θ(Lθρ cos θ + Lρρ sin θ),
⎪
⎪
⎪L23 = Lθz cos θ + Lρz sin θ,
⎪
⎪
⎪L31 = Lzρ cos θ − Lzθ sin θ,
⎪
⎪
⎪L32 = Lzθ cos θ + Lzρ sin θ,
⎪
⎪
⎩
L33 = Lzz ,
Downloaded from www.worldscientific.com
then, applying Eqs. (5.10) and (5.14) in Eq. (5.6)3 for the vectors vi =
(Li1 , Li2 , Li3 ), i = 1, 2, 3, we get, through long but standard operations
and after putting θ = 0 to obtain the components of divL in the basis
{eρ , eθ , ez },
1
divL = ((ρLρρ ),ρ + Lρθ,θ − Lθθ ) + Lρz,z eρ
ρ
1
+ Lθρ,ρ + (Lθθ,θ + Lρθ + Lθρ ) + Lθz,z eθ
ρ
1
+ ((ρLzρ ),ρ + Lzθ,θ ) + Lzz,z ez .
ρ
x1 x2 x1 ρ − x1 ρ,1 x2 2x2 ρρ,1
f,11 = f,ρ − f,θ 2 = f,ρ1 + f,ρ − f,θ1 2 + f,θ
ρ ρ ,1 ρ ρ2 ρ ρ4
x1 x2 x1 ρ2 − x21 x1 x2 x2 2x1 x2
= f,ρρ − f,ρθ 2 + f,ρ − f ,ρθ − f,θθ + f,θ
ρ ρ ρ ρ3 ρ ρ2 ρ2 ρ4
sin θ cos θ sin2 θ sin2 θ sin θ cos θ
= f,ρρ cos2 θ − 2f,ρθ + f,ρ + f,θθ 2 + 2f,θ ,
ρ ρ ρ ρ2
120 Tensor Algebra and Analysis for Engineers
x2 x1 x2 ρ − x2 ρ,2 x1 2x1 ρρ,2
f,22 = f,ρ + f,θ 2 = f,ρ2
+ f,ρ + f,θ2 2 − f,θ
ρ ,2
ρ ρ ρ 2 ρ ρ4
x2 x1 x2 ρ2 − x22 x2 x1 x1 2x1 x2
= f,ρρ + f,ρθ 2 + f,ρ + f,ρθ + f,θθ − f,θ
ρ ρ ρ ρ3 ρ ρ2 ρ2 ρ4
sin θ cos θ cos2 θ cos2 θ sin θ cos θ
= f,ρρ sin2 θ + 2f,ρθ + f,ρ + f,θθ − 2f,θ ,
ρ ρ ρ2 ρ2
f,33 = f,zz .
1 1
Δf = (ρf,ρ ),ρ + 2 f,θθ + f,zz .
ρ ρ
Downloaded from www.worldscientific.com
1 1 1
Δv = (ρvρ,ρ ),ρ + 2 vρ,θθ + vρ,zz − 2 (vρ + 2vθ,θ ) eρ
ρ ρ ρ
1 1 1
+ (ρvθ,ρ ),ρ + 2 vθ,θθ + vθ,zz − 2 (vθ − 2vρ,θ ) eθ
ρ ρ ρ
1 1
+ (ρvz,ρ ),ρ + 2 vz,θθ + vz,zz ez .
ρ ρ
Finally, injecting Eqs. (5.10), (5.12), and (5.13) into Eq. (5.6)7 gives
1 1
curlv = vz,θ − vθ,z eρ + (vρ,z − vz,ρ )eθ + ((ρvθ ),ρ − vρ,θ ) ez .
ρ ρ
(5.15)
or conversely,
x1 = r cos θ sin ϕ,
x2 = r sin θ sin ϕ,
x3 = r cos ϕ.
1 1
gradf = f,r er + f,ϕ eϕ + f,θ eθ ,
r r sin ϕ
122 Tensor Algebra and Analysis for Engineers
1 1 1
gradv = vr,r er ⊗ er + (vr,ϕ − vϕ )er ⊗ eϕ + vr,θ − vθ er ⊗ eθ
r r sin ϕ
1
+ vϕ,r eϕ ⊗ er + (vϕ,ϕ + vr )eϕ ⊗ eϕ
r
1 1
+ vϕ,θ − vθ cot ϕ eϕ ⊗ eθ
r sin ϕ
1
+ vθ,r eθ ⊗ er + vθ,ϕ eθ ⊗ eϕ
r
1 1
+ vθ,θ + vr + vϕ cot ϕ eθ ⊗ eθ ,
r sin ϕ
or in matrix form,
⎡ ⎤
1 1 1
⎢ vr,r r (vr,ϕ − vϕ ) vr,θ − vθ
Downloaded from www.worldscientific.com
⎥
r sin ϕ
⎢ 1 1 1 ⎥
gradv = ⎢ vϕ,r (vϕ,ϕ + vr ) vϕ,θ − vθ cot ϕ ⎥,
r r sin ϕ
⎣ 1 1 1 ⎦
vθ,r vθ,ϕ vθ,θ + vr + vϕ cot ϕ
r r sin ϕ
1 2 1
divv = (r vr ),r + ((vϕ sin ϕ),ϕ + vθ,θ ),
r2 r sin ϕ
1 2 1 1 Lϕϕ + Lθθ cot ϕ
divL = 2 (r Lrr ),r + Lrϕ,ϕ + Lrθ,θ − + Lrϕ er
r r r sin ϕ r r
1 2 1 1 1
+ (r Lϕr ),r + Lϕϕ,ϕ + Lϕθ,θ + Lrϕ
r2 r r sin ϕ r
cot ϕ
+ (Lϕϕ − Lθθ ) eϕ
r
1 2 1 1
+ 2
(r Lθr ),r + Lθϕ,ϕ + Lθθ,θ
r r r sin ϕ
1 cot ϕ
+ Lrθ + (Lϕθ + Lθϕ ) eθ ,
r r
1 1 f,θθ
Δf = 2 (r2 f,r ),r + 2 + (f,ϕ sin ϕ),ϕ ,
r r sin ϕ sin ϕ
Tensor Analysis: Fields 123
2vr,r vr,ϕϕ − 2vϕ,ϕ vr,ϕ − 2vϕ
Δv = vr,rr + + + 2
r r2 r tan ϕ
1 vr,θθ 2vr
+ 2 − 2vθ,θ − 2 er
r sin ϕ sin ϕ r
2vϕ,r vϕ,ϕϕ+2vr,ϕ vϕ,ϕ−vϕ cot ϕ
+ vϕ,rr+ + +
r r2 r2 tan ϕ
1 vϕ
+ 2 2 (vϕ,θθ−2vθ,θ cos ϕ)− 2 eϕ
r sin ϕ r
2vθ,r vθ,ϕϕ 2vϕ,θ 1
+ vθ,rr+ + 2 + vθ,ϕ+
r r sin ϕ r2 tan ϕ
1 vθ,θθ vθ
+ 2 +2vr,θ − 2 2 eθ ,
r sin ϕ sin ϕ r sin ϕ
Downloaded from www.worldscientific.com
1
curlv = ((vθ sin ϕ),ϕ − vϕ,θ ) er
r sin ϕ
1 1
+ vr,θ − (rvθ ),r eϕ
r sin ϕ r
1
+ ((rvϕ ),r − vr,ϕ ) eθ .
r
5.8 Exercises
α
ii. vortex: v = eθ ;
ρ
α
iii. doublet: v = 2 (cos θeρ + sin θeθ ).
ρ
10. A flow with curlv = o is said to be irrotational; check that the flows in
the previous exercise are irrotational.
Chapter 6
Curvilinear Coordinates
Downloaded from www.worldscientific.com
6.1 Introduction
All the developments in the previous chapters are intended for the case
where algebraic and differential operators are expressed in a Cartesian
frame, i.e. with rectangular coordinates. The points of E are thus referred
to a system of coordinates taken along straight lines that are mutually
orthogonal and with the same unit along each one of the directions of the
frame. Though this is a very important and common case, it is not the only
possibility, and in many cases, non-rectangular coordinate frames are used
or arise in the mathematical developments (a typical example is that of the
geometry of surfaces, see Chapter 7). A non-rectangular coordinate frame
is a frame where coordinates can be taken along non-orthogonal directions
or along some lines that intersect at right angles but that are not straight
lines, or even when both of these cases occur. This situation is often denoted
in the literature as that of curvilinear coordinates; the transformations to
be done to algebraic and differential operators in the case of curvilinear
coordinates is the topic of this chapter.
125
126 Tensor Algebra and Analysis for Engineers
z3
x3
x2
z2
o
x1
z1
exist between the two sets of coordinates. The distance between two points
p, q ∈ E is1
s = (p − q) · (p − q) = (xpk − xqk )(xpk − xqk ),
1 The distance between two points p and q is still defined as the Euclidean norm of (p−q),
dx = dxi ei = dz ei ;
∂z k
introducing the vector gk ,
∂xi
gk := ei , (6.5)
∂z k
we can write
dx = dz k gk .
We see hence that a vector dx can be expressed as a linear combination of
the vectors gk ; these form therefore a basis, called the local basis. Generally,
gk ∈/ S, and it is clearly tangent to the lines z j = const. This can be seen
in Fig. 6.2 for a two-dimensional case:
xi (z 1 , z 2 + Δz 2 , z 3 ) − xi (z 1 , z 2 , z 3 ) 2
dx = lim Δx = lim Δz ei
Δx→0 Δx→0 Δz 2
∂xi
= ei dz 2 = g2 dz 2 .
∂z 2
Then,
∂xi ∂xj ∂xi ∂xj
gk · gl = ei · ej = k l δij = gkl , (6.6)
∂z k ∂z l ∂z ∂z
2 The notion of co- and contravariant components is detailed in the next section.
3 As usually done in the literature, we indicate the metric tensor by g, i.e. a lowercase
letter, though it is a second-rank tensor, not a vector.
4 The differential dx is a vector because it is the difference of two infinitely close points;
z1=const.
g1 g2
x z2=const.
z 2)
, z 2+
x(z ,
x (z 1
e2
1 z2 ) o e1
Figure 6.2: Tangent vectors to the curvilinear coordinates lines.
i.e. the components of the metric tensor g are the scalar products of the
Downloaded from www.worldscientific.com
re =g2
e er=g1
e2
r=c
st.
ons
on
t.
=c
e1
x2 x2
v2
vx2 vx2
z2 z2
v2 v v
z1 z1 v1
2 v1 2
1 1
vx1 x1 vx1 x1
Figure 6.4: Contravariant (left) and covariant (right) components of a vector in a plane.
Still referring to the planar case in Fig. 6.4, if the Cartesian components5
of v are v = (v1x , v2x ), we get
1
v = h(v1x sin α2 − v2x cos α2 ), v1 = v1x cos α1 + v2x sin α1 ,
2 (6.7)
v = h(−v1 sin α1 + v2 cos α1 ),
x x
v2 = v1x cos α2 + v2x sin α2 ,
and conversely,
x
v1 = v 1 cos α1 + v 2 cos α2 , v1x = h(v1 sin α2 − v2 sin α1 ),
(6.8)
v2x = v 1 sin α1 + v 2 sin α2 , v2x = h(−v1 cos α2 + v2 cos α1 ),
with
1
h= .
sin(α2 − α1 )
5 Inthe following, we use the superscript x to indicate a Cartesian component: vix is the
ith Cartesian component of v ∈ V and Lx ij the ijth Cartesian component of L ∈ Lin(V).
130 Tensor Algebra and Analysis for Engineers
x2
z2
x2 p
z2
z1
2
z1
1
x1 x1
Figure 6.5: Relation between Cartesian and contravariant components.
It is apparent that the Cartesian coordinates are at the same time co-
and contravariant. Still, on a planar scheme, we can see how to pass from
Downloaded from www.worldscientific.com
a system of coordinates to another one, cf. Fig. 6.5. For a point p, the
Cartesian coordinates (x1 , x2 ) are related to the contravariant ones by
x1 = z 1 cos α1 + z 2 cos α2 ,
x2 = z 1 sin α1 + z 2 sin α2 ,
and conversely,
z 1 = h(x1 sin α2 − x2 cos α2 ),
z 2 = h(−x1 sin α1 + x2 cos α1 ).
So, differentiating, we get
∂x1 ∂x1
= cos α1 , = cos α2 ,
∂z 1 ∂z 2
∂x2 ∂x2
= sin α1 , = sin α2
∂z 1 ∂z 2
and
∂z 1 ∂z 1
= h sin α2 , = −h cos α2 ,
∂x1 ∂x2
∂z 2 ∂z 2
= −h sin α1 , = h cos α1 .
∂x1 ∂x2
Injecting these expressions into Eqs. (6.7) and (6.8) gives
∂x1 ∂x2
v1 = v1x 1
+ v2x 1 , ∂xk x
∂z ∂z → vi = v (6.9)
∂x 1 ∂x 2 ∂z i k
v2 = v1x 2 + v2x 2 ,
∂z ∂z
Curvilinear Coordinates 131
and
∂z 1 ∂z 1
v 1 = v1x + v2x ,
∂x1 ∂x2 ∂z i x
→ vi = v . (6.10)
2 x ∂z
2
x ∂z
2 ∂xk k
v = v1 + v2 ,
∂x1 ∂x2
Now, if we calculate
∂z i x
ghi v i = ghi v
∂xk k
from Eq. (6.3) and by the chain rule,6 we get
∂xj ∂xj ∂z i x
ghi v i = v
∂z h ∂z i ∂xk k
∂xj ∂xj x ∂xj ∂xk x
= h vk = δjk vkx = v = vh ,
Downloaded from www.worldscientific.com
∂z ∂xk ∂z h ∂z h k
i.e. we obtain the rule of lowering of the indices for passing from contravari-
ant to covariant components:
vh = ghi v i .
Introducing the inverse7 to ghi as
∂z h ∂z i
g hi = , (6.11)
∂xk ∂xk
we get, again using the chain rule,
∂xk x ∂z h ∂z i ∂xk x
g hi vi = g hi vk = v
∂z i ∂xj ∂xj ∂z i k
∂z h ∂xk x ∂z h ∂z h x
= vk = δjk vkx = v = vh ,
∂xj ∂xj ∂xj ∂xk k
which is the rule of raising of the indices for passing from covariant to
contravariant components:
v h = g hi vi .
6 The reader can easily see that, in practice, the chain rule allows us to handle the
derivatives as fractions.
7 To prove that the contravariant components g pq are the inverse of the covariant ones,
gpq , is direct:
∂z p ∂z q ∂xj ∂xj
g pq gpq = = δjk δjk = 1.
∂xk ∂xk ∂z p ∂z q
132 Tensor Algebra and Analysis for Engineers
∂z i ∂z i ∂xk x ∂xk x
vi = v = v = δkl vkx ,
∂x l ∂xl ∂z i k ∂xl k
i.e.
∂z i
vkx = vi , (6.12)
∂xk
which is the converse of Eq. (6.9). In a similar way, we get the converse of
Eq. (6.10):
∂xk i
vkx = v. (6.13)
∂z i
Downloaded from www.worldscientific.com
Let us now calculate the norm v of a vector v; starting from the Cartesian
components and using the last two results,
√ ∂z i ∂z j ∂z i ∂z j
v = v · v = vkx vkx = vi vj = vi vj = g ij vi vj ,
∂xk ∂xk ∂xk ∂xk
or also,
√ ∂xk i ∂xk j ∂xk ∂xk i j
v = v · v = vkx vkx = v v = vv = gij v i v j
∂z i ∂z j ∂z i ∂z j
and even,
√ ∂z i ∂xk j ∂z i ∂xk
v= v·v = vkx vkx = vi v = vi v j
∂xk ∂z j ∂xk ∂z j
= δ ij vi v j = vi v i .
Through Eq. (6.13) and by the definition of the tangent vectors to the lines
of curvilinear coordinates, in Eq. (6.5), for a vector v, we get
∂xi
v = vix ei = v k ei = v k gk .
∂z k
We see hence that the contravariant components are actually the com-
ponents of v in the basis composed of gk s, the tangents to the lines of
curvilinear coordinates. In a similar manner, if we introduce the dual basis
Curvilinear Coordinates 133
moreover,
∂z h ∂xj ∂z h ∂xj ∂z h ∂xi ∂z h
g h · gk = ei · ej = δ ij = = = δ hk ,
∂xi ∂z k ∂xi ∂z k ∂xi ∂z k ∂z k
and by the symmetry of the scalar product,
δhk := gh · gk = gk · gh = δ kh .
The last equations define the orthogonality conditions for the g vectors.
Using these results and Eq. (6.15), we also have
v k = δ kh v h = gk · v h gh = gk · v = gk · vh gh = g kh vh ,
vk = δkh vh = gk · vh gh = gk · v = gk · v h gh = gkh v h ,
thus finding again the rules of raising and lowering of the indices.
What was done for vectors can be transposed, using a similar approach,
to tensors. In particular, for a second-rank tensor L, we get
∂z i ∂z j x
Lij = L ,
∂xh ∂xk hk (6.16)
∂xh ∂xk x
Lij = L
∂z i ∂z j hk
for the contravariant and covariant components, respectively, while we can
also introduce the mixed components:
∂z i ∂xk x
Li j = L ,
∂xh ∂z j hk
(6.17)
j ∂xh ∂z j x
Li = L .
∂z i ∂xk hk
134 Tensor Algebra and Analysis for Engineers
Conversely,
∂xh ∂xk ij
Lxhk = L ,
∂z i ∂z j
∂z i ∂z j
Lxhk = Lij ,
∂xh ∂xk
(6.18)
∂xh ∂z j i
Lxhk = L ,
∂z i ∂xk j
∂z i ∂xk j
Lxhk = L .
∂xh ∂z j i
From Eq. (6.18) and by the same definitions of gij , Eq. (6.3), and g ij , Eq.
(6.11), we get
∂xi ∂xj hk
L = Lxij ei ⊗ ej = L ei ⊗ ej = Lhk gh ⊗ gk
∂z h ∂z k
and
∂z h ∂z k
L = Lxij ei ⊗ ej = Lhk ei ⊗ ej = Lhk gh ⊗ gk .
∂xi ∂xj
∂xi ∂z k h
L = Lxij ei ⊗ ej = L ei ⊗ ej = Lhk gh ⊗ gk
∂z h ∂xj k
and
∂z k ∂xi k
L = Lxij ei ⊗ ej = L ei ⊗ ej = Lhk gh ⊗ gk .
∂xj ∂z h h
We see hence that a second-rank tensor can be given with four different
combinations of coordinates; even more complex is the case of higher-order
tensors, which will not be treated here.
Curvilinear Coordinates 135
∂z i
Still, by Eqs. (6.3) and (6.11) and applying the chain rule to δji = j ,
∂z
we get
∂xk ∂xk ∂xh ∂xk
gij = = δhk ,
∂z ∂z
i j ∂z i ∂z j
∂z i ∂z j ∂z i ∂z j
g ij = = δhk ,
∂xk ∂xk ∂xh ∂xk
(6.20)
∂z i ∂xk
δ ij = δhk ,
∂xh ∂z j
∂xh ∂z j
δi j = δhk .
∂z i ∂xk
So, applying Eq. (6.18) to the identity tensor, we get
∂xi ∂xj hk
I = δij ei ⊗ ej = I ei ⊗ ej = I hk gh ⊗ gk ,
Downloaded from www.worldscientific.com
∂z h ∂z k
but by Eqs. (6.16) and (6.20),
∂z h ∂z k
I hk = δij = g hk ,
∂xi ∂xj
so finally,
I = g hk gh ⊗ gk .
or also,
ϕ = ϕ(xj (z k )),
8 The term spatial here refers to differentiation with respect to spatial coordinates, which
Comparing this result with Eq. (6.17)1 , we see that the first member
actually corresponds to the components of a mixed tensor field, which is
the gradient of the vector field v, that we write as
∂v h
v h;k = + Γhkl v l , (6.24)
∂z k
where the functions
∂z h ∂ 2 xm
Γhkl = (6.25)
∂xm ∂z k ∂z l
h
are the Christoffel symbols. We immediately see that Γkl = Γhlk . The
quantity v ;k is the covariant derivative of the contravariant components
h
6.5 Exercises
show that the lines z 1 = const. and z 2 = const. are confocal ellipses and
hyperbole, determine the axes of the ellipses in terms of the parameter
c, discuss the limit case of ellipses that degenerate into a crack, and
determine its length. Finally, find g, g1 , and g2 .
8. Determine the co- and contravariant components of a tensor L in
cylindrical coordinates.
R0
(z)
x2
R
h
x3
x1 R
x2
x1
Figure 6.6: Curve in a plane and on a cone.
140 Tensor Algebra and Analysis for Engineers
g np;h = gnp;h = 0.
Chapter 7
Surfaces in E
Downloaded from www.worldscientific.com
141
142 Tensor Algebra and Analysis for Engineers
Downloaded from www.worldscientific.com
Figure 7.1: General scheme of a surface and of the tangent space at a point p.
∂u ∂v
F,U = f,u + f,v ,
∂U ∂U
∂u ∂v
F,V = f,u + f,v ,
∂V ∂V
or denoting by Jϑ the Jacobian of ϑ,
Downloaded from www.worldscientific.com
F,U f,u
= [Jϑ ] ,
F,V f,v
This result shows that the regularity of the surface, condition (7.1), the
tangent plane, and the tangent space vector do not depend upon the
parameterization of Σ. From the last equation, we also get
x1 = ϕ(u),
γ(u) : ϕ(u) > 0 ∀u ∈ G. (7.4)
x3 = ψ(u),
is the trace of a surface of revolution of the curve γ(u) around the axis x3 .
A general parameterization of such a surface is
⎧
⎨ x1 = ϕ(u) cos v,
f (u, v) : G × (−π, π] → E| x = ϕ(u) sin v, (7.5)
⎩ 2
x3 = ψ(u).
Downloaded from www.worldscientific.com
so that
Figure 7.2: Surfaces of revolution: (from left) sphere, catenoid, pseudo-sphere, and
hyperbolic hyperboloid.
A ruled surface (also named a scroll) is a surface with the property that
through every one of its points, there is a straight line that lies on the
surface. A ruled surface can be seen as the set of points swept by a moving
straight line. We say that a surface is doubly ruled if through every one of
its points, there are two distinct straight lines that lie on the surface.
Any ruled surface can be represented by a parameterization of the form
where γ(u) is a regular smooth curve, the directrix, and λ(u) is a smooth
curve. Fixing u = u0 gives a generator line f (u0 , v) of the surface; the
vectors λ(u) =o describe the directions of the generators. Some important
examples of ruled surfaces are:
Downloaded from www.worldscientific.com
• Cones: For these surfaces, all the straight lines pass through a point,
the apex of the cone, choosing the apex as the origin, then it must be
λ(u) = kγ(u), k ∈ R →
f (u, v) = vγ(u);
γ(u) = (cos 2u, sin 2u, 0), λ(u) = (cos u cos 2u, cos u sin 2u, sin u).
Figure 7.3: Ruled surfaces: (from left) elliptical cone, elliptical cylinder, helicoid, and
Möbius strip.
I(w1 , w2 ) = w1 · gw2 ,
Downloaded from www.worldscientific.com
where3
f ·f f ·f
g = ,u ,u ,u ,v
f,v · f,u f,v · f,v
is precisely the metric tensor g of Σ, cf. Eq. (6.6). In fact, f,u and f,v are
the tangent vectors to the coordinate lines on Σ, i.e. they coincide with the
vectors gk s.
I(w1 , w2 ) is the first fundamental form (or simply the first form) of
f (u, v). If w1 = w2 = w = af,u + bf,v , then
I(w) = w2 = a2 f,u
2
+ 2abf,u · f,v + b2 f,v
2
.
• Metric on Σ: ∀ds ∈ Σ,
ds2 = ds · ds = I(ds),
so if
then
ds2 = f,u
2
du2 + 2f,u · f,v du dv + f,v
2
dv 2 . (7.8)
and hence, see Eq. (7.3), if we call w = (u , v ) the tangent vector to γ,
Downloaded from www.worldscientific.com
w1 · w2 I(w1 , w2 )
cos θ = = .
|w1 ||w2 | I(w1 ) I(w2 )
• Area of a small surface on Σ: Let f,u du and f,v dv be two small vectors
on Σ, forming together the angle θ that are the transformed through4
f : Ω → Σ, of two small orthogonal vectors du, dv ∈ Ω; then, the area dA
of the parallelogram determined by them is
dA = |f,u du × f,v dv| = |f,u × f,v |du dv = f,u 2 f 2 sin2 θdu dv
,v
= f,u 2 f 2 (1 − cos2 θ)du dv = 2 f 2 − f 2 f 2 cos2 θdu dv
f,u
,v ,v ,u ,v
= f,u 2 f 2 − (f · f )2 du dv = det gdu dv.
,v ,u ,v
4 Forthe sake of conciseness, from now on, we indicate a surface as the function f : Ω → Σ,
with f = f (u, v), (u, v) ∈ Ω ⊂ R2 and Σ ⊂ E.
Surfaces in E 149
√
The term det g is hence the dilatation factor of the areas; recalling
Eq. (6.4), we see that the previous expression has a sense ∀f (u, v), i.e.
for any parameterization of the surface.
Let f : Ω → Σ be a regular surface, {f,u , f,v } the natural basis for Tp Σ, and
N ∈ S the normal to Σ defined as in (7.2). We call the map of Gauss of Σ
the map ϕΣ : Σ → S that associates to each p ∈ Σ its N : ϕΣ (p) = N(p).
To each subset σ ⊂ Σ, the map of Gauss associates hence a subset σS ⊂ S,
Fig. 7.4 (e.g. the Gauss map of a plane is just a point on S).
We want to study how N(p) varies at the varying of p on Σ. The idea
is that the change of N(p) on Σ is related to the curvature of the surface.5
For this purpose, we calculate the change in N per unit length of a curve
Downloaded from www.worldscientific.com
γ(s) ∈ Σ, i.e. we study how N varies along any curve of Σ per unit of length
of the curve itself; that is why we parameterize the curve with its arc length
s.6 Let N = Ni (u, v)ei ; then, if τ ∈ S is the tangent to the curve,
dN dNi (u(s), v(s)) ∂Ni du ∂Ni dv
= ei = + ei
ds ds ∂u ds ∂v ds
dN
= ∇Ni · τ ei = (ei ⊗ ∇Ni )τ = (∇N) τ = .
dτ
x3 x3
(p)
N(p)
p S
N(p)
o x2 S o x2
x1 x1
Figure 7.4: The map of Gauss.
5 Forcurves, the curvature is linked to the change in τ , but for surfaces this should not
be meaningful, as τ is not unique ∀p ∈ Σ while N is.
6 Actually, it is also possible to introduce the following concepts more generally for
any parameterization of the curve; anyway, for the sake of simplicity, we just use the
parameter s in the following.
150 Tensor Algebra and Analysis for Engineers
(7.11)
For any two vectors w1 , w2 ∈ Tp Σ, we define the second fundamental form
of a surface, denoted by II(w1 , w2 ), the bilinear form
II(w1 , w2 ) := I(LW (w1 ), w2 ).
Proof. Because I and LW are linear, it is sufficient to prove the thesis for
the natural basis {f,u , f,v } of Tp Σ, and by the symmetry of I, it is sufficient
to prove that
I(LW (f,u ), f,v ) = I(f,u , LW (f,v )),
i.e. that
I(−N,u , f,v ) = I(f,u , −N,v )
and in the end that
N,u · f,v = f,u · N,v .
For this purpose, we recall that
N · f,u = 0 = N · f,v .
So, differentiating the first equation by v and the second one by u, we get
N,v · f,u = −N · f,uv = N,u · f,v . (7.12)
Surfaces in E 151
B = B .
Proof.
γ(s) = γ(u(s), v(s)) → τ (s) = γ (s) = f,u u + f,v v ;
therefore, τ = (u , v ) in the natural basis and
κ(s) = γ (s) = f,u u + f,v v + f,uu u2 + 2f,uv u v + f,vv v 2 ,
and finally, by Eqs. (7.2) and (7.14),
κN (s) = γ (s) · N(s) = B11 u2 + 2B12 u v + B22 v 2 = τ · Bτ
= II(τ , τ ).
Now, if s = s(t) is a change in parameter for γ, then
γ (t) = |γ (t)|τ (t),
Downloaded from www.worldscientific.com
the quadratic form in the previous equation gets its maximum, κ1 , and
minimum, κ2 , values, and in such a basis,
X = κi ui ⊗ ui .
We call κ1 and κ2 the principal curvatures of Σ in p and u1 , u2 the principal
directions of Σ in p, see Fig. 7.5.
We call the Gaussian curvature K the product of the principal
curvatures:
K := κ1 κ2 = det X.
By Eq. (7.15) and the theorem of Binet, it is also
det B
K= . (7.17)
det g
We define the mean curvature H of a surface8 f : Ω → Σ at a point p ∈ Σ
the mean of the principal curvatures at p:
κ1 + κ 2 1
H := = trX.
2 2
Of course, a change in parameterization of a surface can change the
orientation, cf. Section 7.1, which induces a change in N into its opposite
one and, by consequence, in the sign of the second fundamental form and
hence in the normal and principal curvatures. These last are hence defined
to less the sign, and the mean curvature too, while the principal directions,
umbilicality, flatness, and Gaussian curvature are intrinsic to Σ, i.e. they
do not depend on its parameterization.
8 The
concept of mean curvature of a surface was introduced for the first time by Sophie
Germain in her celebrated work on the elasticity of plates (1815).
154 Tensor Algebra and Analysis for Engineers
x3
N
u2 Tp
p u1
o x2
Downloaded from www.worldscientific.com
dN(p)
= −κλ λ (7.18)
dλ
if and only if λ is a principal direction; κλ is the principal curvature relative
to λ.
dN
· N = 0; (7.19)
dλ
moreover,
⎡ ⎤⎧ ⎫
0 0 0 ⎨ λu ⎬
dN
= ∇N λ = ⎣ 0 0 0 ⎦ λv = N,u λu + N,v λv . (7.20)
dλ ⎩ ⎭
N,u N,v 1 0
Surfaces in E 155
II(λ)
α=− = −κλ .
I(λ)
Contrarily, if we assume Eq. (7.21), as before, we get α = −κλ , and to end,
we just need to prove that λ is a principal direction. From Eqs. (7.20) and
(7.21), we get
λu N,u + λv N,v = −κλ (λu f,u + λv f,v ).
Projecting this equation onto f,u and f,v gives the two equations
Lλu + M λv = κλ (Eλu + F λv ),
(7.22)
M λu + N λv = κλ (Eλu + Gλv ),
with the symbols E, F, G, L, M , and N defined in Notes 3 and 7 and used
here for the sake of conciseness. Let w = (wu , wv ) ∈ Tp Σ and consider the
function
ζ(w, κλ ) = II(w) − κλ I(w);
it is easy to check that ζ, ∂w
∂ζ
u
∂ζ
and ∂w v
take zero value for w = λ0 , with
λ0 the eigenvector of the principal direction relative to κλ , which gives the
system of equations
⎧
⎪ II(λ0 ) − κλ I(λ0 ) = 0,
⎪ ∂II(λ0 )
⎨ ∂II(λ0 )
− κλ = 0,
⎪ ∂wu ∂wu
⎪ ∂II(λ0 ) ∂II(λ0 )
⎩ − κλ = 0.
∂wv ∂wv
156 Tensor Algebra and Analysis for Engineers
Proof. For the sake of simplicity and without loss of generality, we can
always chose a parameterization f (u, v) of the surface such that p = f (0, 0).
Expanding f (u, v) into a Taylor’s series around (0, 0), we get the position
of a point q = f (u, v) ∈ Σ in the neighborhood of p (though not indicated
for the sake of brevity, all the derivatives are intended to be calculated at
(0, 0)):
1
f (u, v) = f,u u + f,v v + (f,uu u2 + 2f,uv uv + f,vv v 2 ) + o(u2 + v 2 ).
2
The distance with sign d(q) of q ∈ Σ from the tangent plane Tp Σ is the
projection onto N, i.e.:
1
d(q) = (f,uu u2 + 2f,uv uv + f,vv v 2 ) · N + o(u2 + v 2 )
2
1
= (B11 u2 + 2B12 uv + B22 v 2 ) + o(u2 + v 2 ),
2
or, equivalently, once we set w = uf,u + vf,v ,
d(q) = II(w, w) + o(u2 + v 2 ). (7.23)
Surfaces in E 157
If p is an elliptic point, the principal curvatures have the same sign because
K = κ1 κ2 > 0 ⇒ the sign of II(w, w) does not depend upon w, i.e. upon
the tangent vector. As a consequence, the sign of d(q) does not change with
w ⇒ ∀q ∈ U, Σ is on the same side of the tangent plane Tp Σ.
Proof. The proof is identical to that of the previous theorem until Eq.
(7.23); now, if p is a hyperbolic point, the principal curvatures have opposite
signs, and by consequence d(q) changes of sign at least two times in any
neighborhood U of p ⇒, there are points q ∈ U lying in half-spaces on the
opposite sides with respect to the tangent plane Tp Σ.
Downloaded from www.worldscientific.com
This is also the case for planar points: e.g., the point (0, 0, 0) is a planar
point for both the surfaces
z = x4 + y 4 , z = x3 − 3xy 2 ,
but in the first case, all of the surface is on one side of the tangent plane,
while it is on both sides for the second case (the so-called monkey’s saddle),
see Fig. 7.7.
Figure 7.6: Elliptic (left), hyperbolic (center), and parabolic (last two on the right)
points.
158 Tensor Algebra and Analysis for Engineers
2
Consequently, B22 = N · f,vv = 0 ⇒ det B = −B12 : The points of Σ are
hyperbolic or parabolic. Namely, the parabolic points are those with
f,u × f,v
B12 = N · f,uv = · f,uv = 0
|f,u × f,v |
⇐⇒ (γ × λ + vλ × λ) · λ = γ × λ · λ = 0.
We remark that the ruled surfaces made of parabolic points have null
Gaussian curvature everywhere: K = 0.
Let us consider ruled surfaces having only parabolic points; then, we
have the following.
Theorem 48. For a ruled surface f (u, v) = γ(u) + vλ(u), the following
are equivalents:
(i) γ , λ, λ are linearly dependent;
(ii) N,v = o.
Proof. Condition (ii) implies that N does not change along a straight line
lying on the ruled surface ⇒ f,u × f,v = γ × λ + vλ × λ does not depend on
v as well. This is possible ⇐⇒ γ × λ and λ × λ are linearly dependent,
i.e. ⇐⇒
(γ × λ) × (λ × λ) = (λ × λ · γ )λ − (λ × λ · λ)γ = (λ × λ · γ )λ = o,
i.e. when λ, λ , and γ are coplanar, which proves the thesis.
We say that a ruled surface is developable if one of the conditions of
Theorem 48 is satisfied. A developable surface is a surface that can be
Surfaces in E 159
then, the ruled surface of the tangents to γ is the surface f (u, v) : G×R → Σ
defined by
Figure 7.8 shows the ruled surface of the tangents to a cylindrical helix.
c(u) 0
B= ;
0 ϕ(u)ψ (u)
• the Gaussian curvature K:
det B c(u)ψ (u)
K = det X = = .
det g ϕ(u)
Therefore, the points of Σγ where c(u) and ψ (u) have the same sign are
elliptic, but hyperbolic otherwise.9 Parabolic points correspond to inflexion
points of γ(u) if c(u) = 0 or to points with horizontal tangent to γ(u) if
ψ (u) = 0.
As an example, let us consider the case of the pseudo-sphere, Eq. (7.6).
Then,
u
ϕ(u) = sin u, ψ(u) = cos u + ln tan .
2
Some simple calculations give
1 | tan u|
ψ (u) = − sin u + , c(u) = − ;
sin u | cot u|
as a consequence,
c(u)ψ (u) (− sin u + sin1 u )| tan u|
K= =− = −1.
ϕ(u) sin u| cot u|
Finally, K = const. = −1, which is the reason for the name of this surface.
Theorem 49. The lines of curvature of a surface are the solutions to the
differential equation
X21 u2 + (X22 − X11 )u v − X12 v 2 = 0.
it is
N = N,u u + N,v v = −LW (γ ),
hence f (u, v) is developable ⇐⇒ LW (γ ) × γ = o, i.e. when γ is a
principal direction.
The curve in Eq. (7.24) is called the ruled surface of the normals.
Let p be a non-planar point of a surface f : Ω → Σ and v1 , v2 two
vectors of Tp Σ. We say that v1 and v2 are conjugated if II(v1 , v2 ) = 0.
The directions corresponding to v1 and v2 are called conjugated directions.
Hence, the principal directions at a point p are conjugated; if p is an
umbilical point, any two orthogonal directions are conjugated.
The direction of a vector v ∈ Tp Σ is said to be asymptotic if it is
autoconjugated, i.e. if II(v, v) = 0. An asymptotic direction is hence a
direction where the normal curvature is null. In a hyperbolic point, there
Downloaded from www.worldscientific.com
The conical curves of Dupin are the real curves in Tp Σ whose equations are
II(v, v) = ±1, v ∈ S.
Let {u1 , u2 } be the basis of the principal directions. Using polar coordi-
nates, we can write
v = ρeρ , eρ = cos θu1 + sin θu2 .
Therefore,
II(v, v) = ρ2 II(eρ , eρ ) = ρ2 κN (eρ ),
and the conicals’ equations are
ρ2 (κ1 cos2 θ + κ2 sin2 θ) = ±1.
Surfaces in E 163
Figure 7.9: The conical curves of Dupin (from left): elliptic, hyperbolic, and parabolic
points.
κ1 ξ 2 + κ2 η 2 = ±1.
Downloaded from www.worldscientific.com
• Elliptical points: The principal curvatures have the same sign → one of
the conical curves is an ellipse, the other one the null set (actually, it is
not a real curve).
• Hyperbolic points: The principal curves have opposite signs → the conical
curves are conjugated hyperbolae whose asymptotes coincide with the
asymptotic directions.
• Parabolic points: At least one of the principal curvatures is null → one
of the conical curves degenerates into couple of parallel straight lines,
corresponding to the asymptotic direction, the other one is the null set.
u = z 1, v = z 2
∂gj ∂gi
= f,ji = f,ij = j , i, j = 1, 2.
∂z i ∂z
∂gi
· gh = Γhij i, j, h = 1, 2.
∂z j
Moreover, by Eq. (7.14),
∂gi
· N = f,ij · N = Bij i, j = 1, 2,
∂z j
and by Eqs. (7.10) and (7.11),
∂N
· gj = −LW (gi ) · gj = −Xgi · gj = −Xji , i, j = 1, 2,
∂z i
while, because N ∈ S, then from Eq. (4.1),
∂N
·N=0 ∀i = 1, 2.
∂z i
10 The proof is rather cumbersome and it is omitted here; in many texts on differential
geometry, the Christoffel symbols are just introduced in this way, as the projection of
the derivatives of vectors gi s onto the same vectors, i.e. as the coefficients of the Gauss
equations.
Surfaces in E 165
⎧
⎪ Γ1 g + Γ2 g = 1 ∂g11 ,
⎨ 11 11 11 21
2 ∂z 1
(7.26)
⎪ 1
⎩ Γ11 g12 + Γ211 g22 = ∂g12 − 1 ∂g11 ;
∂z 1 2 ∂z 2
⎧
⎪ Γ1 g + Γ2 g = 1 ∂g11 ,
⎨ 12 11 12 21
2 ∂z 2
(7.27)
⎪ 1
⎩ Γ12 g12 + Γ212 g22 = 1 ∂g22 ;
2 ∂z 1
⎧
⎪ Γ1 g + Γ2 g = ∂g12 − 1 ∂g22 ,
⎨ 22 11 22 21
∂z 2 2 ∂z 1
(7.28)
⎪ 1
⎩ Γ22 g12 + Γ222 g22 = 1 ∂g22 .
2 ∂z 2
The determinant of each one of these systems is simply det g = 0 → it is
possible to express the Christoffel symbols as functions of the gij s and of
their derivatives, i.e. as functions of the first fundamental form (the metric
tensor).
B11 = g11 X11 + g12 X21 , B12 = g11 X12 + g12 X22 ,
which on injecting into the previous equation gives
g11 det X = Γ111 Γ212 + Γ211 Γ222 + Γ211,2 − Γ112 Γ211 − Γ212 Γ212 − Γ212,1 . (7.29)
Setting equal to zero the coefficient of g1 , a similar expression can also be
obtained for g12 . Because g is positive definite, it is not possible that g11 =
g12 = 0. So, remembering that K = det X and the result of the previous
section, we see that it is possible to express K through the coefficients of
the first fundamental form and of its derivatives.
structures composed of prestressed membranes. Also, it can be shown that a soap film,
when not bounding a closed region, takes the form of a minimal surface.
Surfaces in E 167
Figure 7.10: The minimal surfaces of (from left) Enneper, Costa, and Schwarz.
ϕ,u (u, v, t) = f,u (u, v) + t h(u, v)N,u (u, v) + t h,u (u, v)N(u, v),
ϕ,v (u, v, t) = f,v (u, v) + t h(u, v)N,v (u, v) + t h,v (u, v)N(u, v).
If the first fundamental form of f is represented by the metric tensor g,
we look for the metric tensor gt representing the first fundamental form of
ϕ(u, v, t) ∀t:
g11
t
= ϕ,u · ϕ,u = g11 + 2t h f,u · N,u + t2 (h2 N,u
2
+ h2,u ),
g12
t
= ϕ,u · ϕ,v = g12 + t h(f,u · N,v + f,v · N,u ) + t2 (h2 N,u · N,v + h,u h,v ),
g22
t
= ϕ,v · ϕ,v = g22 + 2t h f,v · N,v + t2 (h2 N,v
2
+ h2,v ),
g11
t
= g11 − 2t h B11 + t2 (h2 N2,u + h2,u ),
g12
t
= g12 − 2t h B12 + t2 (h2 N,u · N,v + hu h,v ),
g22
t
= g22 − 2t h B22 + t2 (h2 N2,v + h2,v ),
whence
so that
We can now calculate the area A(t) of the simple region Rt = ϕ(u, v, t)
corresponding to the subset Q:
At = det g(1 − 4thH) + o(t2 )dudv.
Q
dA
=− 2hH det gdudv.
dt t=0 Q
dAt
Theorem 53. A surface f : Ω → Σ is minimal ⇐⇒ = 0 ∀R ⊂
dt t=0
Σ and for each normal variation.
The meaning of this theorem justifies the name of minimal surfaces: These
are the surfaces that have the minimal area among all the surfaces that
share the same boundary.
Surfaces in E 169
7.16 Geodesics
along a curve.
13 The operator that gives the projection of w onto a vector orthogonal to N ∈ S, i.e.
This result shows that in a geodesic, the parameter is always the natural
parameter s. Let γ(s) be a curve on Σ parameterized by the arc length s.
We call the geodesic curvature of γ(s) the function
κg := Dγ τ · (N × τ ),
where τ = γ ∈ S is the tangent vector to γ. Because N × τ ∈ S lies in
Tγ Σ, the component of τ orthogonal to Tγ Σ gives a null contribution to
κg , so we can also write
κg = τ · (N × τ ).
u ⇒ g12 (u, v) = g12 (0, v) ∀u. Moreover, let θ be the angle between the
curve α, i.e. between the coordinate line f (0, v), whose tangent vector is
f,v (0, v), and the geodesic γv (u), whose tangent vector at (0, v) is γv (u).
π
Then, θ = because γv (0) = N(α(v))×τ (v). As a consequence, g12 (0, v) =
2
0 ⇒ g12 (u, v) = 0 ∀(u, v) ∈ Ω. Finally, setting g22 = g,
1 0
g= ,
0 g
2
g,uu g,u
K = det X = − + 2.
2g 4g
Theorem 57. Geodesics are the curves of minimal distance between two
points of a surface.
Observing that p = (up , 0), q = (uq , 0), we remark that |uq − up | is exactly
the length of γ between p and q because γ is parameterized with its arc
length.
172 Tensor Algebra and Analysis for Engineers
There is another, direct and beautiful, way to show that geodesics are the
shortest path lines: the use of the method of calculus of variations.14
The length (p, q) of a curve γ(t) ∈ Σ between two points p and q is given
by the functional (7.9); it depends upon the first fundamental form, i.e.
1 2
upon the metric tensor g on σ. For the sake of conciseness, let w = (w,t , w,t )
be the tangent vector to the curve γ(w1 , w2 ) ∈ Σ. Then,
q q
√
(p, q) = I(w)dt = w · gwdt.
p p
The curve γ(t) that minimizes (p, q) is the solution to the Euler–Lagrange
equations
d ∂F ∂F d ∂F ∂F
− =o → − = 0, k = 1, 2,
dt ∂w,t ∂w dt ∂w,t
k ∂wk
where
Downloaded from www.worldscientific.com
√ i wj .
F (w, w,t , t) = w · gw = gij w,t ,t
It is more direct, and equivalent, to minimize J 2 (t), i.e. to write the Euler–
Lagrange equations for
Φ(w, w,t , t) := F 2 (w, w,t , t) = gij w,t
i j
w,t .
Therefore,
∂Φ j
= 2gjk w,t ,
∂w,t
k
∂Φ ∂ghj h j
= w w , j, h, k = 1, 2.
∂w k ∂wk ,t ,t
d ∂Φ j dgjk j j ∂gjk l j
= 2 g jk w + w = 2 g jk w + w w ,
dt ∂w,t
k ,tt
dt ,t ,tt
∂wl ,t ,t
14 The reader is referred to texts on the calculus of variations for an insight into this
matter, cf. the suggested texts. Here, we just recall the fundamental fact to be used in
the proof concerning geodesics: Let
b
J(t) = F (x, x , t)dt
a
be a functional to be minimized by a proper choice of the function x(t) (in the case of
the geodesics, J = (p, q)); then, such a minimizing function can be found as a solution
to the Euler–Lagrange equations
d ∂F ∂F
− = o.
dt ∂x ∂x
Surfaces in E 173
j ∂gjk h j 1 ∂ghj h j
gjk w,tt + w,t w,t − w w = 0, j, h, k = 1, 2,
∂w h 2 ∂wk ,t ,t
which can be rewritten as
j 1 ∂gjk ∂ghk ∂ghj h j
gjk w,tt + + − w,t w,t = 0, j, h, k = 1, 2.
2 ∂wh ∂wj ∂wk
Multiplying by g lk , we get
j 1 ∂gjk ∂ghk ∂ghj h j
g lk gjk w,tt + g lk + − w,t w,t = 0, j, h, k, l = 1, 2,
2 ∂wh ∂wj ∂wk
g lk gjk = δlj
These are the differential equations whose solution is the curve of minimal
length between two points of Σ; comparing these equations with those of a
geodesic of Σ, Eq. (7.31), we see that they are the same: The geodesics of
a surface are hence the curves of minimal distance on the surface.
The Christoffel symbols of a plane are all null; as a consequence, the
geodesic lines of a plane are straight lines. In fact, only such lines have a
constant derivative.
Through systems (7.26)—(7.28), we can calculate the Christoffel sym-
bols for a revolution surface, Eq. (7.5), which are all null excepted
2 ϕ
Γ12 = , Γ122 = −ϕ ϕ ,
ϕ
dxi dxi
A2 = r2,α2 = ,
dα2 dα2
which are the so-called Lamé’s parameters. We remark that along the lines
of curvatures, i.e. the lines αi = const., i = 1, 2, which in short from now
on, we call the lines αi , it is
ds1 = A1 dα1 ,
ds2 = A2 dα2 ,
and hence,
ds1
λ1 = = A1 e1 ,
dα1
(7.33)
ds2
λ2 = = A2 e2
dα2
are the vectors tangent to the lines of curvature. Let
1 1
e1 = r,α , e2 = r,α , e3 = e1 × e2 (= N); (7.34)
A1 1 A2 2
these three vectors form the orthonormal (local) natural basis e =
{e1 , e2 , e3 }. We always make the choice of α1 , α2 such that e3 is always
directed toward the convex side of Σ if the point is elliptic or parabolic or
toward the side of the centers of negative curvature if the point is hyperbolic.
v = v1 e1 + v2 e2 + v3 e3 ,
∂e3
= −κi λi , i = 1, 2,
∂λi
∂e3 Ai
= ei ,
∂αi Ri
with
1
Ri = −
κi
the (principal) radius of curvature along the line αi . The minus sign in the
previous equation is due to the choice made above for orienting e3 = N,
which gives always N = −ν, with ν the principal normal to the line αi .
This result can also be obtained directly, see Fig. 7.11:
de3
e3(p) e3(q)
p q lin
e
i
Ri
o
Figure 7.11: Variation of N = e3 along a line of curvature.
176 Tensor Algebra and Analysis for Engineers
lim (q − p) = λi = Ai ei .
q→p
∂e3 Ai
= ei . (7.35)
∂αi Ri
Implicitly, in this last proof, we have used the theorem of Rodrigues because
we have assumed that de3 is parallel to λi , as it is, because line αi is a line
of curvature.
We now move on to determine the changes in e1 and e2 ; for this purpose,
we remark that
∂r,α1 ∂2r ∂ 2r ∂r,α2
= = = ,
∂α2 ∂α2 ∂α1 ∂α1 ∂α2 ∂α1
so by Eq. (7.34), we get
∂(A1 e1 ) ∂(A2 e2 )
= . (7.36)
∂α2 ∂α1
∂ej
Let us study now ∂αi ; as |ej | = 1, j = 1, 2,
∂ej
· ej = 0 ∀i, j = 1, 2. (7.37)
∂αi
Because e1 · e2 = 0,
∂e1 ∂(e1 · e2 ) ∂e2 ∂e2
· e2 = − e1 · = −e1 · .
∂α1 ∂α1 ∂α1 ∂α1
By Eq. (7.36), we get
∂e2 1 ∂(A1 e1 ) 1 ∂A2
= − e2 ,
∂α1 A2 ∂α2 A2 ∂α1
Surfaces in E 177
which when inserted into the previous equation gives, by Eq. (7.37),
∂e1 1 ∂(A1 e1 ) 1 ∂A2
· e2 = − · e1 + e2 · e1
∂α1 A2 ∂α2 A2 ∂α1
A1 ∂e1 1 ∂A1 1 ∂A1
=− · e1 − e1 · e1 = − .
A2 ∂α2 A2 ∂α2 A2 ∂α2
Then, because e1 · e3 = 0,
∂e1 ∂(e1 · e3 ) ∂e3 ∂e3
· e3 = − e1 · = −e1 · ,
∂α1 ∂α1 ∂α1 ∂α1
and by Eq. (7.35),
∂e3 A1
= e1 ,
∂α1 R1
so finally,
Downloaded from www.worldscientific.com
∂e1 A1
· e3 = − .
∂α1 R1
Again, through Eqs. (7.36) and (7.37), we get
∂e1 1 ∂(A2 e2 ) 1 ∂A1 A2 ∂e2
· e2 = · e2 − e1 · e2 = · e2
∂α2 A1 ∂α1 A1 ∂α2 A1 ∂α1
1 ∂A2 1 ∂A2
+ e2 · e2 =
A1 ∂α1 A1 ∂α1
and also, by Eq. (7.35),
∂e1 ∂(e1 · e3 ) ∂e3 ∂e3 A2
· e3 = − e1 · = −e1 · = − e1 · e2 = 0.
∂α2 ∂α2 ∂α2 ∂α2 R2
The derivatives of e2 can be found in a similar way, and resuming, we have
∂e1 1 ∂A1 A1
=− e2 − e3 ,
∂α1 A2 ∂α2 R1
∂e1 1 ∂A2
= e2 ,
∂α2 A1 ∂α1
∂e2 1 ∂A1
= e1 ,
∂α1 A2 ∂α2
(7.38)
∂e2 1 ∂A2 A2
= − e1 − e3 ,
∂α2 A1 ∂α1 R2
∂e3 A1
= e1 ,
∂α1 R1
∂e3 A2
= e2 .
∂α2 R2
178 Tensor Algebra and Analysis for Engineers
gives
∂ A1 1 ∂A1 ∂ A2 1 ∂A2
− e1 − − e2 = 0,
∂α2 R1 R2 ∂α2 ∂α1 R2 R1 ∂α1
which to be true needs that the two following conditions be identically
satisfied:
∂ A1 1 ∂A1
− = 0,
∂α2 R1 R2 ∂α2
(7.39)
∂ A2 1 ∂A2
− = 0.
∂α1 R2 R1 ∂α1
The above equations are known as the Codazzi conditions. Let us now
consider the other identity
∂ 2 e1 ∂ 2 e1
= ;
∂α1 ∂α2 ∂α2 ∂α1
again using Eq. (7.38), with some standard operations, this identity can be
transformed to
∂ 1 ∂A2 ∂ 1 ∂A1 A1 A2
+ + e2
∂α1 A1 ∂α1 ∂α2 A2 ∂α2 R1 R2
∂ A1 1 ∂A1
+ − e3 = 0.
∂α2 R1 R2 ∂α2
Again, for this equation to be identically satisfied, each of the expressions in
square brackets must vanish, which gives two more differential conditions,
Surfaces in E 179
but of which only the first one is new, as the second one corresponds to
Eq. (7.39)1 . The new condition is hence
∂ 1 ∂A2 ∂ 1 ∂A1 A1 A2
+ + = 0, (7.40)
∂α1 A1 ∂α1 ∂α2 A2 ∂α2 R1 R2
∂ 2 e2 ∂ 2 e2
=
∂α1 ∂α2 ∂α2 ∂α1
does not add any independent condition, which can be easily checked. The
meaning of the Gauss–Codazzi conditions, Eqs. (7.39) and (7.40), is that
of compatibility conditions: Only when these conditions are satisfied by
functions A1 , A2 , R1 , and R2 , then such functions represent the Lamé’s
Downloaded from www.worldscientific.com
7.18 Exercises
with
γ(u) = (cos(u − α), sin(u − α), −1), λ(u) = (cos(u + α), sin(u + α), 1).
Show that:
• for α = 0, one gets a cylinder with equation x21 + x22 = 1;
π
• for α = , one gets a cone with equation x21 + x22 = x23 ;
2
π
• for 0 < α < , one gets a hyperbolic hyperboloid with equation
2
x21 + x22 x23
− = 1.
cos2 α cot2 α
8. Calculate the metric tensor of a sphere of radius R, write its first
fundamental form, and determine the area of the sector of a surface
between the longitudes θ1 and θ2 and the length of the parallel at the
latitude π/4 between these two longitudes.
9. Prove that the surface defined by
cos v sin v sinh u
f (u, v) : Ω = R × (−π, π] → E| f (u, v) = , ,
cosh u cosh u cosh u
183
184 Tensor Algebra and Analysis for Engineers
Chapter 1
185
186 Tensor Algebra and Analysis for Engineers
Chapter 2
1. ∀u, L, Lu = L(u + o) = Lu + Lo ⇐⇒ Lo = o.
2. u ∈ S → (u ⊗ u)v = v cos θu; (I − u ⊗ u)v = v − v cos θu, which is
orthogonal to u : (I − u ⊗ u)v · u = v · u − v · u u · u = o, as u ∈ S.
3. (i) ∀a, b ∈ V, a·(αA)b = (αA) a·b, and by the linearity of the scalar
product, a · (αA)b = αa · Ab = αA a · b ⇒ (αA) = αA .
(ii) ∀a, b ∈ V, a · (A + B)b = (A + B) a · b, and by linearity of the
scalar product and of tensors, a · (A + B)b = a · Ab + a · Bb =
A a · b + B a · b = (A + B )a · b ⇒ (A + B) = A + B .
(iii) ∀u, v ∈ V u · (a ⊗ b)Av = u · (a ⊗ b)(Av) = (a ⊗ b) u · (Av) =
(b⊗a)u·(Av) = A (b⊗a)u·v = A (a·u b)·v = a·u A b·v =
((A b) ⊗ a)u · v = u · ((A b) ⊗ a) v = u · (a ⊗ A b)v.
4. By linearity and the definition of O : ∀u ∈ V, (L + O)u = Lu + Ou =
Lu ⇐⇒ L + O = L.
5. (i) trI = tr(δij ei ⊗ ej ) = δij tr(ei ⊗ ej ) = δij ei · ej = δij δij = δii = 3.
(ii) trL = tr(L + O) = trL + trO ⇐⇒ trO = 0.
6. tr(AB) = tr((Aij ei ⊗ ej )(Bhk eh ⊗ ek )) = Aij Bhk tr((ei ⊗ ej )(eh ⊗ ek ))
= Aij Bhk ej · eh tr(ei ⊗ ek ) = Aij Bhk ej · eh ei cot ek = Aij Bhk δjh δik =
Solutions to the Exercises 187
Aij Bji ; in a similar way, we prove that tr(BA) = Bij Aji ; because i, j
are dummy indexes, Aij Bji = Bij Aji ⇒ tr(AB) = tr(BA).
7. (i) L ·M = tr((L ) M ) = tr(LM ) = tr(M L) = M·L = L·M.
(ii) LM · N = tr((LM) N) = tr(M L N) = M · L N;
|
= tr(N(LM) ) = tr((NM )L )
= tr(L (NM ))
= L · NM .
8. (i) (a⊗ b)(c⊗ d) = ((a⊗ b)(c⊗ d))ij ei ⊗ ej = (a⊗ b)ik (c⊗ d)kj ei ⊗ ej
= ai bk ck dj ei ⊗ ej = b · c ai dj ei ⊗ ej = b · c a ⊗ d.
(ii) A(a⊗b) = A(a⊗b)ij ei ⊗ej = Aik (a⊗b)kj ei ⊗ej = Aik ak bj ei ⊗ej
= (Aa)i bj ei ⊗ ej = (Aa) ⊗ b.
9. L · u ⊗ u = tr(L (u ⊗ u)) = tr((L u) ⊗ u) = L u · u = u · Lu.
10. A = A , B = −B ⇒ A · B = A · B = A · (−B) = −A · B ⇐⇒
Downloaded from www.worldscientific.com
A · B = 0.
11. (i) A = A ⇒ A · L = A · (Ls + La ) = A · Ls + A · La = A · Ls .
(ii) B = −B ⇒ B · L = B · (Ls + La ) = B · Ls + B · La = B · La .
12. (i) A · (BCD) = tr(A BCD) = tr((B A) CD) = (B A) · (CD).
(ii) A · (BCD) = (BCD) · A = tr((BCD) A) = tr(A(D C B )) =
tr((AD )(C B )) = tr((C B )(AD )) = tr((BC) (AD )) =
BC · AD = AD · BC.
13. L ∈ Sym(V) ⇒ L · W = 0, as already proved. Now, if L · W =
0 ∀W ∈ Skw(V), suppose L ∈ / Sym(V) ⇒ L = Ls + La ⇒ L · W =
L · W + L · W = L · W = 0; if in Skw(V), we chose W = La , we
s a a
get W · La = La · La = 0 ⇐⇒ La = O ⇒ L ∈ Sym(V).
14. I2 = 12 (tr2 L − trL2 ) = 12 (Lii Ljj − Lij Lji )
= L11 L22 + ⎡ L11 L33 + L22 L ⎧− L⎫12 L
⎤33 ⎧21 −⎫L13 L31 − L23 L32 .
0 −a3 a2 ⎨ b1 ⎬ ⎨ c1 ⎬
15. a×b·c = ⎣ a3 0 −a1 ⎦ b2 · c2 = a2 b3 c1 −a3 b2 c1 +a3 b1 c2 −
⎩ ⎭ ⎩ ⎭
−a2 a1 0 b3 c3
⎡ ⎤
0 −a3 a2
a1 b3 c2 + a1 b2 c3 − a2 b1 c3 = det ⎣ a3 0 −a1 ⎦ .
−a2 a1 0
16. Let L1 = L2 be two distinct inverse tensors of L; then, L1 L = I =
L2 L ⇒ L1 L − L2 L = O ⇒ (L1 − L2 )L = O ∀L ⇐⇒ L1 − L2 = O ⇒
L1 = L2 .
17. (a ⊗ b)ij = ai bj ; it is then sufficient to write the matrix representing
(a ⊗ b) and to compute its determinant.
188 Tensor Algebra and Analysis for Engineers
Chapter 3
12. ∀L ∈ Lin(V), L = Lsph + Ldev and Lsph = 13 trL I →, just one number
is sufficient to determine Lsph ⇒ dim(Sph(V)) = 1. Then, Ldev =
L − Lsph is determined by five numbers: dim(Dev(V)) = dim(Lin(V) −
Sph(V)) = 6 − 1 = 5.
13. (i) Ssph Ssph = 13 I ⊗ I 13 I ⊗ I = 19 (I ⊗ I) = 19 I · I I ⊗ I = 13 I ⊗ I =
Ssph .
(ii) Ddev Ddev = (Is − Ssph )(Is − Ssph ) = Is − 2Ssph + Ssph Ssph =
Is − Ssph = Ddev .
(iii) Ssph Ddev = Ssph (Is − Ssph ) = Ssph − Ssph Ssph = Ssph − Ssph = O.
sym
14. (i) Sijkl = (ei ⊗ ej ) · Ssym (ek ⊗ el ) = (ei ⊗ ej ) · ek ⊗el +e 2
l ⊗ek
1
= 2 (δik δjl + δil δjk ), Wijkl = (ei ⊗ ej ) · W (ek ⊗ el )
skw skw
= (ei ⊗ ej ) · ek ⊗el −e
2
l ⊗ek
= 12 (δik δjl − δil δjk ),
sym
→ Sijkl + Wijkl skw
= δik δjl = Iijkl ⇒ Ssym + Wdev = I.
sym trp
(ii) Sijkl − Wijkl = δil δjk = Tijkl
skw
⇒ Ssym − Wdev = Ttrp .
15. (i) Ssph · Ssph = 3 I ⊗ I · 3 I ⊗ I = 19 tr4 ((I ⊗ I) (I ⊗ I))
1 1
Chapter 4
1. (i) (u(t) + v(t)) − (u(t0 ) + v(t0 ) = (t − t0 )(u + v) + o(t − t0 ), and also,
u(t) − u(t0 ) + v(t) − v(t0 ) = (t − t0 )u + (t − t0 )v + o(t − t0 ) ⇒
(u + v) = u + v .
The proof that (L + M) = L + M and that (L + M) = L + M
can be done in a similar way.
(ii) We indicate, in short, α(t) = α, v(t) = v, α(t0 ) = α0 , v(t0 ) =
v0 , α (t0 ) = α0 , v (t0 ) = v0 → αv−α0 v0 = (αv)0 (t−t0 )+o(t−t0 );
moreover, α = α0 + α0 (t − t0 ) + o(t − t0 ), v = v0 + v0 (t − t0 )
+ o(t − t0 ) ⇒ αv − α0 v0 = (α0 + α0 (t − t0 ) + o(t − t0 ))(v0 + v0 (t −
t0 ) + o(t − t0 )) − α0 v0 = α0 v0 (t − t0 ) + v0 o(t − t0 ) + α0 v0 (t − t0 )
+ α0 v0 (t − t0 )2 + v0 (t − t0 )o(t − t0 ) + α0 o(t − t0 ) + α0 (t − t0 )o(t − t0 )
+ o(t − t0 )2 = (α0 v0 + α0 v0 )(t − t0 ) + o(t − t0 ), so by comparison,
(αv) = α v + αv .
By the same technique, one can easily prove the differentiation
rule for all the product-like quantities: (u · v) , (u × v) , (u ⊗ v) ,
(αL) , (Lv) , (LM) , (L · M) , (L ⊗ M) , (L M) , (αL) ,
(LL) , (LM) , (L · M) .
2. The proof is the same for all the cases; we just write that for v(t):
Using the two properties shown in the previous exercise, we get
Solutions to the Exercises 191
v(t) = vi (t)ei ⇒ v (t) = (vi (t)ei ) = vi (t)ei + vi (t)ei = vi (t)ei because
ei does not depend on t.
2
3. (i) p(θ) = (aθ cos θ, aθ sin θ) → c = 2+θ2 3 .
√ √ a(1+θ ) 2
(ii) = a2 (θ 1 + θ2 + ln(θ + 1 + θ2 )).
(iii) Let pi , pi+1 be two consecutive intersection points of the spiral (i
denotes the order of the intersection point) with a straight line
passing through the origin and inclined at θ; their distances from
the origin are ri = a(θ+2πi), ri+1 = a(θ+2π(i+1)) ⇒ |pi+1 −pi | =
ri+1 − ri = 2πa that does not depend upon θ.
4. (i) r = a ebθ ⇒ r = 0 ⇐⇒ θ → +∞, for b < 0, −∞ for b > 0.
1
(ii) p(θ) = (a ebθ cos θ, a ebθ sin θ) → c = a ebθ √ .
√ 1+b2
(iii) = b 1 + b2 e .
a bθ
(iv) ϑ = − a2 +bb
2.
(v) Let pi , pi+1 be two points, the intersection of the same generatrix of
the cylinder with the helix, i.e. for, say, θ + i 2π 2π
ω and θ + (i + 1) ω ⇒
d = (pi+1 − pi ) · e3 = 2πb.
(vi) By definition, a curve is a helix ⇐⇒ τ ·e3 = const; differentiating
gives τ · e3 + τ · e3 = τ · e3 = 0 ⇒ by the first equation of
Frenet–Serret, cν · e3 = 0 ⇒ ν · e3 = 0 ⇒ β is tangent to the
192 Tensor Algebra and Analysis for Engineers
2
(iii) f = ma = mv̇τ + mv ρ ν = fτ τ + fν ν; fτ = mv̇ is the tangential
force, which is responsible for the change in the scalar velocity;
2
fν = mvρ is the centripetal force, which is responsible for the path
change.
Chapter 5
⎡ ⎤⎧ ⎫
0 v1,2 − v2,1 v1,3 − v3,1 ⎨ u1 ⎬
= ⎣ v2,1 − v1,2 0 v2,3 − v3,2 ⎦ u2 ⇒ curlv
⎩ ⎭
v3,1 − v1,3 v3,2 − v2,3 0 u3
⎧ ⎫
⎨ v3,2 − v2,3 ⎬
= v1,3 − v3,1 .
⎩ ⎭
v2,1 − v1,2
7. (i) v(p) = v(p0 ) + ω × (p − p0 ) ⇒ ∃Wω ∈ Skw(V)| v(p) =
v(p0 ) + Wω (p − p0 ), with Wω the axial tensor of ω. Moreover,
by the definition of gradient, v(p) = v(p0 ) + (gradv)(p − p0 ) ⇒
Wω = gradv ⇒ gradv = gradv+gradv 2 + gradv−gradv
2 = Wω =
gradv+gradv
−Wω ⇐⇒ 2 = O ⇒ gradv = gradv−gradv 2 ⇒
gradv−gradv
v(p) = v(p0 ) + 2 (p − p0 ) so, by the definition of curl,
v(p) = v(p0 ) + 12 curlv × (p − p0 ), and comparing the two results,
we get ω = 12 curl.
Downloaded from www.worldscientific.com
Chapter 6
⎡ ⎤
1 0 0
1. Setting ρ = z 1 , θ = z 2 , z = z 3 , by Eq. (6.3), we get g = ⎣ 0 ρ2 0 ⎦ ⇒
0 0 1
ds = ghk dz dz = dρ + ρ dθ + dz .
h k 2 2 2 2
1 2 3
2. ⎡ r = z , θ = z⎤ , ϕ = z , proceeding in a similar way, we get
Setting
1 0 0
g = ⎣ 0 r2 sin2 ϕ 0 ⎦ ⇒ ds = ghk dz h dz k
0 0 r2
2
= dr2 + r2 sin ϕdθ2 + r2 dϕ2 .
3. For cylindrical coordinates, cf. Exercise 1, ds = dρ2 + ρ2 dθ2 + dz 2 ,
√ for a curve on a circular cylinder, ρ = R ⇒ dρ = 0 ⇒ ds =
and
R2 dθ2 + dz 2 ; if the equation of the helix is p(θ) = R cos θe1 +
Downloaded from www.worldscientific.com
1 2 1
4 4 + π + π arcsinh 2 R ∼ 1.324R.
π
(ii) 2 R = 2πR0 ⇒ R0 = R4 ⇒ h =
π
R2 − R02 = 15 4 R; ρ(z) =
R0
z, z = h
2π θ ⇒ ρ(θ) = R0
2π θ ⇒ ρ(θ) = 8π θ, z(θ) =
R
√h √
15 15
8π Rθ ⇒ dρ = 8π dθ, dz = 8π R dθ; equation
R
√ of the
conical helix: p(θ) = ρ(z) √ cos θe 1 + ρ(z) sin θe 2 + 15θe3 =
R
θ cos θe + θ sin θe + 15θe ⇒ ds = dρ 2 + ρ2 dθ 2 + dz 2 =
8π √ 1 2 3
√ 2π
2 2 2 2 = 2
8π dθ √+ θ dθ + 15dθ 8π 16 + θ dθ ⇒ = 0 ds =
R R
2π 2 dθ =
R
8π 0 16 + θ
R 1
√
θ 2π
2
8π 2 θ 16 + θ + 8arcsinh 4 0 ∼ 1.324R.
5. (i) Referring to Fig. 6.5 and by Eq. (6.3), x1 = z 1 cos α1 + z 2 cos α2 ,
1 cos(α2 − α1 )
x2 = z 1 sin α1 + z 2 sin α2 ⇒ g = .
cos(α2 − α1 ) 1
(ii) By Eq. (6.5), g1 = cos α1 e1 + sin α1 e2 , g2 = cos α2 e1 + sin α2 e2 .
(iii) z 1 = h(x1 sin α2 − x2 cos α2 ), z 2 = h(−x1 sin α1 + x2 cos α1 ), h =
1 1 sin α2 cos α2 2
sin(α2 −α1 ) ⇒ by Eq. (6.14), g = sin(α2 −α1 ) e1 − sin(α2 −α1 ) e2 , g =
sin α1 cos α1
− sin(α 2 −α1 )
e1 + sin(α 2 −α1 )
e2 .
1 cos α1 sin α2 −sin α1 cos α2
(iv) g1 · g = sin(α2 −α1 ) = 1,
− sin α1 cos α2 +sin α2 cos α1
g2 · g 2 = sin(α2 −α1 ) = 1,
Solutions to the Exercises 197
g2
x2
g2
g1
2
1
x1
g1
Downloaded from www.worldscientific.com
L22 = ρ12 Lx11 sin2 θ − (Lx12 + Lx21 ) sin θ cos θ + Lx22 cos2 θ ,
L23 = −Lx13 sin θ + Lx23 cos θ,
L31 = Lx31 cos θ + Lx32 sin θ,
L32 = −Lx31 sin θ + Lx32 cos θ,
L33 = Lx33 .
(ii) The covariant components can alternatively be found by Eq. (6.16)2
or, using the results of the previous point, by Eq. (6.19)2 ; by this
latter way, using the result of Ex. 1, we get easily L11 = L11 , L12 =
ρ2 L12 , L13 = L13 , L21 = ρ2 L21 , L22 = ρ4 L22 , L23 = ρ2 L23 , L31 =
L31 , L32 = ρ2 L32 , L33 = L33 .
9. (i) First, the covariant components: By Eq. (6.16)2 , setting z 1 =
r, z 2 = θ, z 3 = ϕ, we get:
L11 = Lx11 cos2 θ sin2 ϕ + (Lx12 + Lx21 ) sin θ cos θ sin2 ϕ
+ (Lx13 + Lx31 ) cos θ sin ϕ cos ϕ + Lx22 sin2 θ sin2 ϕ
Downloaded from www.worldscientific.com
1 0 0
1 ⎢ 1 ⎥
gpq ⇒ g cont
= ⎣ 0 r sin ϕ 0 ⎦ ⇒
2 2
1
0 0 r2
11 12 13 21 22
L = L11 , L = r2 Lsin 12
2 ϕ, L 2 , L
= Lr13 = r2 Lsin
21
2 ϕ, L =
23 31 32 33
r 4 sin4 ϕ , L = r4 sin2 ϕ , L = r2 , L = r4 sin2 ϕ , L = r4 .
L22 L23 L31 L32 L33
i j
(iii) By Eq. (2.7)2 ⇒ Lxhh = ∂x ∂z ∂z
h ∂xk
Lij δhk = g ij Lij .
∂xh ∂z j i j i
(iv) By Eq. (2.7)3 ⇒ Lxhh = ∂z i ∂xk L j δhk = δi L j = L i .
i
i
(v) By Eq. (2.7)4 ⇒ Lxhh = ∂x ∂z ∂xk
L j δ = δ i j Li j = Ljj .
h ∂zj h i mhk
1 hm ∂gmk 1 ∂z ∂z ∂ ∂xp ∂xp ∂ ∂xp ∂xp
11. 2 g l + ∂gml
k − ∂gkl
m = 2 ∂xp ∂xp ∂z l ∂z m ∂z k + ∂z k ∂z m ∂z l −
∂z
∂z ∂z
h
∂z m ∂ 2 xp ∂xp ∂xp ∂ 2 xp ∂ 2 xp ∂xp
∂ ∂xp ∂xp
∂z m ∂z k ∂z l = 12 ∂z ∂xp ∂xp ∂z l ∂z m ∂z k + ∂z m ∂z l ∂z k + ∂z k ∂z m ∂z l +
∂xp ∂ 2 xp ∂ 2 xp ∂xp ∂ 2 xp ∂xp h
∂z m ∂xp ∂ xp
2
∂z m ∂z k ∂z l − ∂z m ∂z k ∂z l − ∂z m ∂z l ∂z k = ∂z ∂xp ∂xp ∂z m ∂z k ∂z l =
2
∂z h ∂ xp
∂xp ∂z k ∂z l = Γkl .
h
1 ϕm ∂gθm
Γϕθθ = 2 g ∂θ + ∂θ − ∂z m
∂gθm ∂gθθ
= − 12 g ϕϕ ∂g∂ϕ = − sin ϕ cos ϕ,
θθ
rϕ = Γϕr , Γrθ = Γθr , Γϕθ = Γθϕ , and the other Γij are null.
Γϕ ϕ θ θ θ θ k
c2 1
(iii) g11 = g22 = 2 (cosh 2z − cos 2z 2 ) ⇒ g 11 = g 22 =
2
c2 (cosh 2z 1 −cos 2z2 ) , and the other components are null
Downloaded from www.worldscientific.com
⇒ Γ111 = 12 g 1m ∂g ∂z
1m
1 + ∂g1m
∂z 1 − ∂g11
∂z m = 12 g 11 ∂g 11
∂z 1
1
= cosh sinh 2z
,
2z 1 −cos2z2
1 1m ∂g1m
1
Γ12 = 2 g ∂z 2 + ∂g2m
∂z 1 − ∂g12
∂z m = 12 g 11 ∂g11
∂z 2
sin 2z 2
= cosh 2z 1 −cos 2z 2 ,
= ρ1 (ρ f,ρ ),ρ + ρ12 f,θθ + f,zz , which is the same already found in
Section 5.6.
(ii) Δf = ∂r ∂
g rk ∂z
∂f
k + Γrrj g jk ∂z
∂f ∂
k + ∂θ g θk ∂z
∂f
k + Γθθj g jk ∂z
∂f
k +
∂ϕ g
∂ ϕk ∂f
∂z k + Γϕ ϕj g ∂z k = ∂r g ∂r + Γθr g ∂z k + ∂θ g
jk ∂f ∂ rr ∂f θ rk ∂f ∂ θθ ∂f
∂θ +
∂2f 1 ∂2f 2 ∂f
Γθθϕ g ϕk ∂z
∂f
k + ∂ϕ g ∂ϕ + Γϕr g ∂z k = ∂r 2 +
∂ ϕϕ ∂f ϕ rk ∂f
r 2 sin2 ϕ ∂θ 2 + +
2
r ∂r
cot ϕ r12 ∂ϕ 1 ∂ f 1 2 1 f,θθ
r 2 ∂ϕ2 = r 2 (r f,r ),r + r 2 sin ϕ sin ϕ + (f,ϕ sin ϕ),ϕ ,
∂f
+
which is to be compared to the one given innpSection 5.7.
np
15. (i) By Eqs. (6.11), (6.25), and (6.29), g;h = ∂g
∂z h
+ Γnhr g rp
n p n 2 p 2
+ Γphr g nr , g np = ∂z ∂z n ∂z p
∂xk ∂xk , Γhr = ∂xm ∂z h ∂z r , Γhr = ∂xt ∂z h ∂z r
∂ xm ∂z ∂ xt
np n p n p n r p
⇒ g;h = ∂x∂ k ∂z ∂z ∂z ∂ ∂z ∂z ∂ ∂xm ∂z ∂z
∂xh ∂xk + ∂xk ∂xk ∂z h + ∂xm ∂z r ∂z h ∂xq ∂xq
∂z p ∂ ∂xt ∂z n ∂z r ∂z n ∂ ∂xm ∂z p ∂z p ∂ ∂xt ∂z n
+ ∂xt ∂z r ∂z h ∂xs ∂xs = ∂xm ∂z h ∂xq ∂xq + ∂xt ∂z h ∂xs ∂xs =0
Solutions to the Exercises 201
Chapter 7
f f
⎧ ,u ,v
⎨ x1 = cosh u cos v,
2. Catenoid: f (u, v) : x2 = cosh u sin v, Meridians: v = const.; if, for
⎩
x3 = u;
⎧
⎨ x1 = cosh u,
example, v = 0 ⇒ x2 = 0, is a catenary in the plane (x1 , x3 ).
⎩
x3 = u,
⎧ ⎫ ⎧ ⎫
⎨ sinh u cos v ⎬ ⎨ − cosh u sin v ⎬
f,u = sinh u sin v , f,v = cosh u cos v ⇒g=
⎩ ⎭ ⎩ ⎭
1 0
cosh2 u 0
.
0 cosh2 u
⎧ ⎫
⎨ − cosh u cos v ⎬
f,u × f,v = − cosh u sin v , |f,u × f,v | = cosh2 u ⇒ N =
⎩ ⎭
sinh u cosh u
⎧ ⎫
⎨ − cos v ⎬
1
cosh u ⎩ − sin v ⎭ ;
sinh u
⎧ ⎫ ⎧ ⎫
⎨ − sinh u sin v ⎬ ⎨ cosh u cos v ⎬
f,uv = f,vu = sinh u cos v , f,uu = cosh u sin v ,
⎩ ⎭ ⎩ ⎭
0 0
⎧ ⎫
⎨ − cosh u cos v ⎬
−1 0
f,vv = − cosh u sin v ⇒ B = ⇒ K = − cosh1 4 u .
⎩ ⎭ 0 1
0
202 Tensor Algebra and Analysis for Engineers
⎧
⎨ x1 = sin u cos v,
3. Pseudo-sphere: f (u, v) : x = sin u sin v, Meridians: v =
⎩ 2
x3 = cos u + ln tan 2 . u
⎧
⎨ x1 = sin u,
const.; if, for example, v = 0 ⇒ x2 = 0, is a tractrix
⎩
x3 = cos u + ln tan 2 , u
⎩ ⎭
sin u cos u
⎧ 2
⎫
⎨ − cos u cos v ⎬
1 2
| cos u| ⎩ − cos u sin v ⎭ ;
sin u cos u
⎧ ⎫ ⎧ ⎫
⎨ − sin u cos v ⎬ ⎨ − cos u sin v ⎬
f,uu = − sin u sin v , f,uv = f,vu = cos u cos v ,
⎩ cos u ⎭ ⎩ ⎭
− cos u − sin 2u 0
⎧ ⎫
⎨ − sin u cos v ⎬ − sincos
2
u
0
u| cos u|
f,vv = − sin u sin v ⇒ B = cos2 u sin u
⇒
⎩ ⎭ 0 | cos u|
0
K = −1.
4. Cone: f (u, v) = vγ(u) ⇒ f,u = vγ , f,v = γ → N =o ⇐⇒ v = 0 and
γ =αγ , i.e. everywhere except at the apex of the cone and on straight
lines tangent to γ.
5. The most ⎧general equation of the hyperbolic hyperboloid is
⎨ x1 = a(cos u − v sin u),
f (u, v) : x2 = b(sin u + v cos u), a, b, c ∈ R ⇒ f (u, v) = γ(u) + vλ(u),
⎩
x3 = c v,
with γ(u) = a cos ue1 + b sin ue2 , λ(u) = −a⎧sin ue1 + b cos ue2 + ce3 ⇒
⎨ x1 = a(cos u0 − v sin u0 ),
f (u, v) is a ruled surface. Fixing u = u0 ⇒ x2 = b(sin u0 + v cos u0 ),
⎩
x3 = c v,
equation of a bundle of straight lines belonging to f (u, v) as well as
Solutions to the Exercises 203
⎧
⎨ x1 = a(cos u0 − v sin u0 ),
x = b(sin u0 + v cos u0 ), The angle formed by the two straight lines
⎩ 2
x3 = −c v.
2 2 2 2 2
of the two sets is θ = arccos aa2 sin u0 +b cos u0 −c
.
sin2 u0 +b2 cos2 u0 +c2 ⎧
⎨ x1 = u,
6. x3 = x1 x2 ; setting u = x1 , v = x2 ⇒ f (u, v) : x2 = v, is of the
⎩
x3 = u v,
type f (u, v) = γ(u) + vλ(v), with γ(u) = ue1 , λ(u) = e2 + ue3 .
⎧ ⎧
⎨ x1 = u0 , ⎨ x1 = u,
The straight lines x2 = v, and x2 = v0 , belong of course to
⎩ ⎩
x3 = u0 v, x3 = u v0 ,
f (u, v); they form the angle θ = arccos √ u02v0 2 ; θ = π2 ⇐⇒ 0 =
(1+u0 )(1+v0 )
v0 = 0, i.e. at (0, 0, 0).
7. (i) γ(u) = (cos u, sin u, −1), λ(u) = (cos u, sin u, 1) ⇒ f (u, v) =
Downloaded from www.worldscientific.com
(cos u, sin u, 2v − 1), which is of the form f (u, v) = γ1 (u) + vλ1 (u),
with γ1 (u) = (cos u, sin u, −1), λ1(u) = (0, 0, 2) = const. ⇒ f (u, v)
being a cylinder whose Cartesian equation is x21 + x22 = 1.
(ii) γ(u) = (sin u, − cos u, −1), λ(u) = (− sin u, cos u, 1) ⇒
f (u, v) = (2v − 1)(− sin u, cos u, 1), which is of the form f (u, v) =
γ2 (u) + vλ2 (u), with γ2 (u) = (sin u, − cos u, −1), λ1(u) =
(−2 sin u, 2 cos u, 2) = −2γ2 (u) ⇒ f (u, v) being a cone whose
Cartesian
equation is x21 + x22 = x23 .
x1 = (1 − v) cos(u − α) + v cos(u + α),
(iii) ⇒
x2 = (1 − v) sin(u − α) + v sin(u + α), x3 = −(1 − v) + v,
x1 = cos u cos α − sin u sin α(2v − 1),
x2 = sin u cos α + cos u sin α(2v − 1), x3 = 2v − 1.
sin α
Change in parameter w = cos − 1) ⇒
α (2v ⎧
⎧
⎨ 1x = cos α(cos u − w sin u), ⎨ x1 = a(cos u − w sin u),
x = cos α(sin u + w cos u), ⇒ x = a(sin u + w cos u), a =
⎩ 2 ⎩ 2
x3 = w cos α
sin α , x3 = cw,
cos α, c = cos α
sin α , which is the parametric equation of a hyperbolic
x21 x22 x23 x21 +x22
hyperboloid with Cartesian equation a2 + a2 − c2 =1→ cos2 α −
x23
cot2 α= 1.
8. (i) A sphere of radius R : x21 + x22 + x23 = R2 ⇒ using the
spherical coordinates θ = u, ϕ = v⎧for expressing the xi s, we get
⎨ x1 = R cos θ sin ϕ,
the parametric equation f (u, v) : x = R sin θ sin ϕ, ⇒ f,u =
⎩ 2
x3 = R cos ϕ,
204 Tensor Algebra and Analysis for Engineers
√R
(θ − θ1 ).
2 2
2 2 2 2 2
cos v sin v sinh u 1 sinh u cosh u
9. (i) x21 +x22 +x23 = cosh 2 u + cosh2 u + cosh2 u = cosh2 u + cosh2 u = cosh2 u =
u = 0,
Referring to Eq. (7.7), ϕ(u) = R, ψ(u) = u ⇒ Eq. (7.32) is ⇒
v = 0,
u(t) = αt + α1 ,
with α, α1 , β, β1 = const. If α1 = β1 = 0, we
v(t) = βt + β1 ,
⎧ the geodesic γ(t) passing
get ⎧ through (R, 0, 0) for t = 0 ⇒ γ(t) :
⎨ 1x = R cos v(t), ⎨ 1 R cos(βt),
x =
x2 = R sin v(t), ⇒ x = R sin(βt), ⇒ equation of a helix if
⎩ ⎩ 2
x3 = u(t), x3 = αt,
α, β = 0, of a circle (cross section) if α = 0, β = 0, and of a straight
line on the cylinder (generatrix) if α = 0, β = 0.
Downloaded from www.worldscientific.com
Index
A
Downloaded from www.worldscientific.com
207
208 Tensor Algebra and Analysis for Engineers
curves of points, 75
theorem, 39
curvilinear abscissa, 83
Euler–Lagrange equations, 172
curvilinear coordinates, 125–126
evolute, 93
cusp, 97
cycloid, 100
F
cylinders, 146
cylindrical basis, 117 first fundamental form, 147
flux theorem, 113
D fourth-rank tensor, 57
Frenet–Serret formula, 88, 91
deformation, 103
Frenet–Serret local basis, 85
derivative, 76
derivative of a point, 76
G
determinant, 22–23
developable surfaces, 158 Gauss condition, 179
deviatoric part, 21 Gauss equations, 164
deviatoric projector, 63 Gauss theorem, 112
diffeomorphism, 143 Gauss’ basis, 163
differentiable function, 80 Gauss–Codazzi compatibility
differentiable vector field, 109 conditions, 174
differential geometry, 184 Gauss–Weingarten equations, 165
differential operators, vi, 121 Gaussian curvature, 153
dilatation factor of the areas, 149 general parameterization, 144
directional derivative, 103 generator line, 146
director cosines, 6, 67 geodesic curvature, 170
distance, 4, 21 geodesics, vii, 169
divergence, 104 global rotation, 47
divergence lemma, 111 gradient, 104
divergence of a tensor field, 104 gradient of the vector field, 137
divergence of products, 107 Green’s formula, 114
Index 209