Appendix
1 Linear Al gebra
1.1 R
n
a s a vector space
As a set, R
n
is the collection of all n-tuple (x
1
; :::; x
n
), x
k
2R. Equivalently, R
n
can be defined
as the Cartesian product of n copies of R as R
n
= R ×···×R. A tuple u = (x
1
; :::; x
n
) has
a double life, as a point in the Cartesian space with coordinates x
k
, and as a vector u~ with
the tail at the origin and head at u. For this reason, R
n
can be considered as a set of points
or a set of vectors .
u
x
1
x
n
x
2
x
2
x
n
u~
x
1
We ne ed two more structures on R
n
to make it a vector space, i .e. , the scalar multipli-
cation, and the vector addition. Let c 2R be a scalar, and u~ = (x
1
; :::x
n
); v~ = (y
1
; :::; y
n
) two
arbitrary vectors. We define
c u~ = (cx
1
; :: :; cx
n
); u~ + v~ = (x
1
+ y
1
; :::; x
n
+ y
n
):
The geometry of these two o perations are shown in Fig.1.
cu~ ; c < 0
u~
cu~ ; c > 0
v~
u~ + v~
u~ ¡v~
u~
Figure 1.
1
It is seen that R
n
is closed under the defined operations, that is, for any two vectors u~ ;
v~ in R
n
, and c
1
; c
2
2R, vector c
1
u~ + c
2
v~ is in R
n
. Moreover, the following properties hold for
any vectors u~ ; v~ ; w~ 2R
n
and constant c 2R:
i. u~ + v~ = v~ + u~
ii. c (u~ + v~) = c u~ + c v~
iii. (c
1
+ c
2
)u~ = c
1
u~ + c
2
u~
iv. (u~ + v~) + w~ = u~ + (v~ + w~ )
R
n
has n standard unit vectors e^
1
; :::; e^
n
defined below
e^
1
=
0
B
B
B
B
B
B
@
1
0
·
·
·
0
1
C
C
C
C
C
C
A
; e^
2
=
0
B
B
B
B
B
B
@
0
1
·
·
·
0
1
C
C
C
C
C
C
A
; :::; e^
n
=
0
B
B
B
B
B
B
@
0
0
·
·
·
1
1
C
C
C
C
C
C
A
:
The set fe^
k
g
k=1
n
is a basis for R
n
in the sense that any vector u~ =(x
1
; :::; x
n
) can be uniquely
represented as
u~ = x
1
e^
1
+ ···+ x
n
e^
n
:
Remark. The magnitude or norm of a vector u~ = (x
1
; :: :; x
n
) is defined as follows
ku~ k=
X
k=1
n
x
k
2
s
:
A vector is called unit or a direction vector if its no rm is equal 1. We usually denote unit
vector by notation ^.
1.2 Dot and Cross products
The dot product of two arbitrary vectors u~ = (x
1
; :::; x
n
); v~ = (y
1
; :: :; y
n
) is defined as
u~ ·v~ =
X
k=1
n
x
i
y
i
:
In particular, e^
i
·e^
j
= δ
i;j
where δ
i;j
=
1 i = j
0 i =/ j
.
Problem 1. Show that the dot product enjoys the following properti es for any vectors u~ ; v~ ; w~ and for
arbitrary constant c
i. (u~ + v~) ·w~ = u~ ·w~ + v~ ·w~ ,
ii. u~ ·v~ = v~ ·u~ ,
iii. (c u~ ) ·v~ = c (u~ ·v~)
Problem 2. Show the following relations
a) ku~ k
2
= u~ ·u~
b) ku~ + v~ kku~ k+ kv~ k
c) ku~ + v~ k
2
¡ku~ ¡v~ k
2
= 4 u~ ·v~
Problem 3. The Cauchy inequality is as follows
u~ ·v~ ku~ kkv~ k :
2 Appendix
Try to prove the inequality by the following method: ku~ + t v~ k
2
0 for all t 2R. Expand ku~ + t v~ k
2
in
terms of t and conclude the inequality. By this inequality, one can define the angle between two nonzero
vectors u~ ; v~ as follows
cos(θ) =
u~ ·v~
ku~ kk v~ k
:
By t he above equality, one can write (u~ ; v~) = ku~ kkv~ kcos(θ). Note that if u~ ·v~ = 0 for non-zero vectors
u~ ; v~, then cos(θ) = 0.
Problem 4. Show that if u~
1
; :::; u~
m
are mutually orthogonal, that is, if u~
i
·u~
j
= 0 for i =/ j then
ku~
1
+ · ··+ u~
m
k
2
= ku~
1
k+ ···+ ku~
m
k
2
:
There is a standard product in R
3
called cross or external product. For u~ = (x
1
; y
1
; z
1
);
v~ = (x
2
; y
2
; z
2
), the cross product is defined as follows
u~ ×v~ =
e^
1
e^
2
e^
3
x
1
y
1
z
1
x
2
y
2
z
2
= (y
1
z
2
¡z
1
y
2
)e^
1
+ (z
1
x
2
¡x
1
z
2
)e^
2
+ (x
1
y
2
¡y
1
x
2
)e^
3
:
Note that u~ ×v~ is a vector, while their dot p roduct is a scalar.
Problem 5. Show the relation u~ ×v~ = ¡v~ ×u~ .
Problem 6. Show that u~ ×v~ is perpendicular to u~ and v~, that is,
(u~ ×v~) ·u~ = (u~ ×v~) ·v~ = 0:
Problem 7. Show the identity
ku~ ×v~ k
2
= ku~ k
2
ku~ k
2
¡ju~ ·v~ j
2
;
and conclude the relation
ku~ ×v~ k= ku~ kkv~ ksin(θ);
where θ is the angle between u~ ; v~ in [0; π].
By the above problem, we can write
u~ ×v~ = ku~ kkv~ ksin(θ) n^;
where n^ is the unit vector perpendicular to the plane containing u~ ; v~, that is, n^ =
u~ ×v~
ku~ ×v~ k
.
A = jju~ ×v~ jj
θ
u~ ×v~
u~
v~
A
v~ ×u~
Problem 8. Show the following relation for any three vectors u~ ; v~ ; w~
u~ ·(v~ ×w~ ) = v~ ·(w~ ×u~ ):
Problem 9. If v~ ; w~ are orthogonal, show the following relation
u~ ×(v~ ×w~ ) = (u~ ·w~ )v~ ¡(u~ ·v~)w~ :
Use this result and relax the condition v~ ; w~ to be orthogonal. Use the formula and determine the
conditions the following rela tion holds
u~ ×(v~ ×w~ ) = (u~ ×v~) ×w~ :
1 Linear Algebra 3
1.3 Subspaces and direct sum
Definition 1. Tow vectors u~ ; v~ in R
n
are called linearly dependent if there is a scalar c
such that u~ = c v~ or v~ = c u~ . A vector u~ is linearly dependent on vectors v~
1
; :::; v~
m
if there are
scalars c
1
; :: :; c
m
such that
u~ = c
1
v
1
~ + ···+ c
m
v~
m
:
Vectors v~
1
; :: :; v~
m
in R
n
are linearly independent if the linear combination
c
1
v~
1
+ ···+ c
m
v~
m
= 0;
implies c
1
= ···= c
m
= 0.
Problem 10. Vectors e^
1
; :::; e^
n
are linearly independent in R
n
. Show that any n + 1 vectors of R
n
are
linearly dependent.
Let fv~
1
; ::: ; v~
d
g fo r d n be a set of linearly independent vectors i n R
n
. The span of
vectors in the given set is the set of all possible linear combinations of v~
1
; :::; v~
d
, i.e.,
spanfv~
1
; :::; v~
d
g= fc
1
v~
1
+ ·· ·+ c
d
v~
d
; c
k
2Rg:
Note that R
n
is itself equal to spanfe^
1
; :::; e^
n
g.
Proposition 1. V
d
:= spanfv~
1
; :::; v~
d
g is closed under the vector addition and scalar multi-
plication of R
n
. For this reason, V
d
is called a linear subspace of R
n
.
Definition 2. Let V be a linear subspace of R
n
. The dimension of V is the maximum
number of linearly independent vectors in V.
Example 1. Technically speaking, R
m
is not a subspace of R
n
for m < n, however, if we
interpret R
m
as spanfe^
1
; :::; e^
m
g where each e^
j
is a vector in R
n
, then R
m
is a linear subspace
of R
n
.
If V is a linear subspaces of R
n
, then its orthogonal subspace V
?
is defined as follows
V
?
= fw~ 2R
n
; w~ ·v~ = 0; v~ 2Vg:
Obviously, V
?
is a linear subspace of R
n
equipped with the vector addition and scalar
multiplication operations.
Problem 11. Show that if V =spanf(1; 1;0); (0; 1;1)g, then V
?
is the one dimensional subspace spanned
by (1; ¡1 ; 1).
Problem 12. Find the orthogonal subspace of V = f(1; 0; 1)g in R
3
.
Suppose U;V are two subspaces of R
n
and U \V = f0g. The direct sum U V is defined
as follows
U V = fc
1
u~ + c
2
v~; u~ 2U; v~ 2Vg:
Problem 13. If V is a subspace of R
n
, show that R
n
= V V
?
.
Problem 14. Let V be an arbitrary subspace of R
n
. Show that every vector u~ in R
n
can be represented
uniquely as u~ = c
1
v~ + c
2
w~ fo r v~ 2V and w~ 2V
?
.
4 Appendix
1.4 Matrices and line ar mappings
Definition 3. A mapping f : R
n
! R
m
is called linear if for any constants c
1
; c
2
and any
vectors u~ ; v~ 2R
n
, the following relation holds
f(c
1
u~ + c
2
v~) = c
1
f(u~ ) + c
2
f(v~):
Problem 15. If f : R
n
!R
m
is linear then f (0) = 0.
Proposition 2. A linear mapping f: R
n
!R
m
can be represented by a m ×n matrix.
Proof. Remember that a matrix A= [a
ij
]
m×n
is a structure of n-columns of vectors belonging
R
m
, i.e., A = [A
1
··jA
n
], where A
k
2 R
m
. The acti on of A to e^
k
is defined by the re lation
A (e^
k
) = A
k
. Now define A
f
as
A
f
= [f(e^
1
)jf(e^
2
)··jf(e^
n
)]:
It is simply seen that for arbitrary vector u~ 2R
n
, the following relation holds f(u~ ) = A
f
(u~ ) .
Problem 16. Let T : R
2
!R
2
has the matrix representation A
2×2
=
1 1
0 1
in the standard basis. Find
the matrix representation of T in the basis v~
1
=
1
1
; v~
2
=
1
¡1
.
Problem 17. Prove that a matrix 2 ×2 maps any parallelogram to a parallelogram.
Problem 18. Verify that the the matrix R
θ
=
cos(θ) ¡sin(θ)
sin(θ) cos(θ)
rotates vectors in the plane by θ degree
counter-clockwise. Verify that R
θ
1
R
θ
2
= R
θ
1
+θ
2
and conclude that R
θ
R
¡θ
is the identity matrix
1 0
0 1
.
Definition 4. Let f:R
n
!R
m
be a linear mapping. The kernel (or null space) of f, denoted
by ker(f) (or just N
f
) is a set of all vectors n~ of R
n
such that f(n~ ) = 0 2 R
m
. The image
of f denoted by I m(f) is the set of all vectors w~ 2R
m
such that w~ = f(v~) for some v~ 2R
n
.
Proposition 3. The kernel of a linear mapping f : R
n
! R
m
is a vector subspace of R
n
.
The image of f is a vector subspace of R
m
.
Problem 19. Prove the proposition.
Theorem 1. Assume that f : R
n
!R
m
is a linear mapping. The following relation holds
n = dim ker(f) + dim Im(f): (1)
Problem 20. Let S denote the orthogonal subspace of ker(f). Show that dim S = dim Im(f) and
conclude f (S) = Im(f ).
1.5 Linear m appings from R
n
to R
n
1.5.1 Determinant
Let f: R
n
!R
n
be a linear mapping, and let C be a unit cube constructed on e^
k
; k = 1; :::;
n. The image of C under f, that is f(C), is a parallelogram. In fact, every vector u~ 2C is
represented by the linear combination
u~ = c
1
e^
1
+ ·· ·+ c
n
e^
n
;
1 Linear Algebra 5
for 0 c
k
1, and thus
f(u~ ) = c
1
f
1
~
+ ···+ c
n
f
n
~
;
where f
k
= f (e^
k
). The set fc
1
f
1
~
+ ···+ c
n
f
n
~
g for 0 c
k
1 is a parallelogram constructed on
f
1
~
; :: :; f
~
n
; see Fig. 2 .
e^
1
f
~
1
= f (e^
1
)
f
~
2
= f (e^
2
)
e^
2
Figure 2.
Definition 5. Let f : R
n
!R
n
be a linear mapping, and let C be the unit cube constructed
on fe
k
^ g
k=1
n
. The determinant of f denoted by det(f) is the algebraic volume of parallelogram
f(C). The algebraic volume is t he signed volume with positive or negative signs.
Example 2. In R
2
, the determinant of A =
a
11
a
12
a
21
a
22
is defined by the following formul a
det
a
11
a
12
a
21
a
22
= a
11
a
22
¡a
12
a
21
: (2)
It is simply verified that jdet(A)j= jjA(e^
1
)jjjjA(e^
2
)jjsin(θ), where θ is the angel between two
columns of A.
If det(f) = 0, then the volume degenerates, that means vectors f
~
1
; :::; f
~
n
are linearly
dependent. If det(f ) < 0, then f changes the standard orientation of the basis fe^
k
g
k=1
n
(remember the standard rotations in R
2
and R
3
), for example f (x; y)=(y; x) with the matrix
representation
A = [f(e^
1
)jf (e^
2
)] =
0 1
1 0
;
changes the standard rotation. Let A
f
= [f
~
1
··jf
~
n
] be the representation of f: R
n
! R
n
in
the standard basis fe^
k
g
k=1
n
. The determinant det(A
f
) satisfies th e following properties:
i. If f
1
~
; ·· ·; f
~
n
are linearly dependent then det[A] = 0
ii. det[f
~
2
jf
~
1
j···jf
n
~
] = ¡det[A
f
]. In general any switch b etween column i a nd j multiple
the determinant by the factor (¡1)
i+j
.
iii. det[c f
~
1
jf
2
~
··jf
n
~
] = c det(A
f
)
iv. det[c
1
f
~
1
+ c
2
f
~
k
jf
~
2
j···jf
~
n
] = c
1
det[A
f
] for any k = 2; :::; n.
By the above properties, it is seen that if f ; g: R
n
!R
n
are two linear mappings then
det(AB) = det(A) det(B)
6 Appendix
Problem 21. Verify directly the above claim for 2 ×2 matrices.
1.5.2 Injective and surjective ma pping s
Definition 6. A linear mapping f : R
n
! R
m
is called one-to-one or injective if equality
f(u~ ) = f(v~) implies u~ = v~. A linear mapping f: R
n
! R
m
is called onto or surjective, if
R
m
= f(R
n
).
Problem 22. A linear mapping f : R
n
!R
n
is one to one if and only if ker(f) = ;, and if and only if
it is onto.
Problem 23. Let f: R
n
! R
m
be a linear mapping. Show tha t if m > n, then f can not be onto, if
m < n then f can not be one-to-one.
If f: R
n
! R
n
is one-to-one (and then onto), the mapping f
¡1
: R
n
! R
n
is called the
inverse mapping of f if the following relation holds
ff
¡1
= f
¡1
f = Id;
where Id is the identity mapping on R
n
. The identity mapping has the matrix representation
diag(1; :::; 1), where diag(1; :::; 1) has 1 on the main diagonal and zero everywhere else. Note
that Id(u~ ) = u~ for any vector u~ .
Problem 24. If f: R
n
!R
n
is a one to on e linear map, show that f
¡1
is also a one to one linear map.
1.5.3 Eigenvalues and Eigenvectors
A vector v~ 2 R
n
¡ f0g is called a n eigenvector of a linear mapping f : R
n
! R
n
if there is
a scalar λ such that f(v~) = λv~. It is seen that if v~ is an eigenvector, then vector w~ = tv~ for
arbitrary scalar t is also an eigenvector. Accordingly, one can define an eigendirection of f
that is spanfv~ g:= ft v~ ; t 2Rg; see Fig. 3.
v~
λv
~
= f (v
~
)
f
Figure 3.
Example 3. The vector v~ = (1; 1) is an eigenvector of the matrix A =
1 1
¡1 3
with the
eigenvalue λ = 2, because
1 1
¡1 3
1
1
= 2
1
1
. Matrix A =
1 1
¡1 3
has only one eigenvector
and matrix A =
2 3
¡4 ¡5
has two eigenvectors v~
1
= (1; ¡1) and v
2
~ = (3; ¡4) with eigenvalues
λ
1
= ¡1 and λ
2
= ¡2 respectively. The rotation matri x R
θ
=
cos(θ) ¡sin(θ)
sin(θ) cos(θ)
has no (real)
eigenvector for θ =/ 0; 2π. Recall that R
θ
rotates vectors counterclockwise by θ-angle. The
identity matrix I
2×2
=
1 0
0 1
has infinitely many eigenvectors. In fact, every vector in R
2
is
an eigenvector of Id
2×2
with eigenvalue λ = 1.
1 Linear Algebra 7
Proposition 4. If f: R
n
! R
n
has n distinct eigenvalues λ
1
; :::; λ
n
, then their associated
eigenvectors v~
1
; :::; v~
n
are linearly independent.
Problem 25. Prove the proposition.
If v~ is an eigenvector of a linear mapping f with eigenvalue λ, then (f ¡λId)v~ = 0, and
since v~ is nonzero, v~ must belong to the kernel of f ¡λId. Let A
f
is a matrix representation
of f, then the following relation holds
det(A
f
¡λId) = 0:
The above equation, which is an algebraic equation of λ, is called the characteristic equation
of f. If A =
a
11
a
12
a
21
a
22
, the characteristic equation is as follows
λ
2
¡tr(A) λ + det(A) = 0; (3)
where tr(A) (read trace A) is equal to a
11
+ a
22
.
Problem 26. Show that if A
2×2
has a r epeated eigenva lue λ with two linearly independent eigenvectors
then all vectors of R
2
is an eigenvector of A.
Problem 27. Let A be a 2 ×2 matrix. P rove that the f ollowing statements are equivalent
i. A is invertible.
ii. Two columns of A are linearly independent.
iii. The determinant of A is non-zero.
iv. No eigenvalue of A is zero.
Problem 28. If λ
1
; λ
2
are two eigenvalues of A
2×2
, show that det(A) = λ
1
λ
2
and tr(A) = λ
1
+ λ
2
.
Problem 29. If Q
2×2
is an invertible matrix, show the following relations
tr(Q
¡1
AQ) = tr(A), det(Q
¡1
AQ) = det(A):
1.5.4 Symmetric mappings and Jordan forms
Definition 7. A linear mapping f : R
n
! R
n
is called symmetric if the following equality
holds for arbitrary vectors u~ ; v~ 2R
n
:
f(u~ ) ·v~ = u~ · f (v~):
Theorem 2. If the linear mapping f : R
n
! R
n
is sy mmetric, then there are n mutually
orthogonal eigenvectors v~
1
; :: :; v~
n
for f. Moreover, all eigenvalues of f are real.
Problem 30. Assume v~
1
; v~
2
are two eigenvectors of a symmetric mapping f. Show t ha t hv~
1
; v~
2
i= 0.
If f: R
n
! R
n
has n linearly independent eigenvectors v~
1
; :::; v~
n
, then R
n
can be
decomposed by the direct sum R
n
= V
1
···V
n
where V
k
= spanfv~
k
g. The restriction of f
to each V
k
is a linear mapping f
k
: V
k
!V
k
, and thus we can decompose f as the direct s um
f = f
1
··· f
n
. With this interpretation, every vector v~ 2R
n
has a unique representation
v~ = c
1
v~
1
+ ···+ c
n
v~
n
, and thus f(v~) is
f(v~) = f
1
(c
1
v~
1
) + ···+ f
n
(c
n
v~
n
) = c
1
λ
1
v~
1
+ ·· ·+ c
n
λ
n
v~
n
8 Appendix
Definition 8. A linear mapping f: R
n
! R
n
is called positive definite if for any nonzero
vector v~, the following inequality holds
f(v~) ·v~ > 0:
A negative definite linear mapping is defined similarly .
It is simply seen that if A
f
= [a
ij
] is the matrix representation of the positive definite
mapping f in the standard basis, then a
ii
> 0 for i = 1; :::; n. Moreover, all real eigenvalues
of A
f
must be positive.
Problem 31. Let f: R
n
!R
n
be a symmetric mapping. A necessary and sufficient condition that f be
po sitive definite is that its all eigenvalues are positive.
If Q
n×n
is an invertible matrix, then two matrices B = Q
¡1
AQ and A are called similar.
It is seen that A;B have same characteristic polynomial as the following argument justifies it:
det(Q
¡1
AQ ¡λId) = det Q
¡1
(A ¡λId)Q = det(Q
¡1
) det(A ¡λId) detQ = det(A ¡λId):
Proposition 5. Suppose A
n×n
has n linearly independent eigenvectors v~
1
; :::; v~
n
. Then the
following relation holds
Q
¡1
AQ = diag(λ
1
; :: :; λ
n
);
where for Q = [v~
1
j·· ·jv~
n
], and λ
1
; :: :; λ
n
are associated eigenvalues (not necessarily distinct).
Matrix diag(λ
1
; :::; λ
n
) is called the Jordan form of A.
Problem 32. Prove the ab ove proposition.
Fig. 4 shows the relation between eigenvectors of A and Q
¡1
AQ .
Eigenvectors ofEigenvectors of A
v~
1
v~
2
Q
¡1
AQ
e
2
~
e
1
~
Figure 4.
Problem 33. Let A =
¡4 ¡3
3 2
. Find the matrix Q
¡1
AQ.
If a matrix A
n×n
has n repeated eigenvalues λ with only one eigenvector, then the
Jordan form of A
n×n
is a diagonal matrix with λ on the main diagonal and 1 on the upper
diagonal. For example, for a 3 × 3 matrix with repeated eigenvalue λ, the Jordan form
is
0
B
B
@
λ 1 0
0 λ 1
0 0 λ
1
C
C
A
. If a matrix 2 × 2 has two complex eigenvalues λ = σ ± i!, its Jordan form is
σ ¡!
! σ
.
Problem 34. For a n ×n matrix A show det(A) =
Q
k=1
n
λ
k
, and conclude that A is one-to-one if and
only if A do es not have a zero eigenvalue.
1 Linear Algebra 9
1.6 Linear equations
Let f : R
n
! R
m
be a linear mapping, and let b
~
2 R
m
be an arbitrary vector. The linear
equation f(u~ ) = b
~
is solvable if and only if b
~
2Im(f). If f: R
n
!R
n
is a linear one-to-one
mapping, then equation f(u~ ) = b
~
is simply solved for u~ = f
¡1
(b
~
). If dim ker(f) > 0 and u~ is a
solution to the equation, then for any vector n~ 2ker(f), v~ = n~ + u~ is also a solution. In this
context, vectors in ker(f) are called the homogeneous solutions of f (u~ ) = 0.
Problem 35. Suppose n m and f: R
n
!R
m
is a linear mapping. Show that if dim ker(f) = n ¡m,
then the linear equation f (u~ ) = b
~
is solvable for any b
~
2R
m
. What if dim ker(f) > n ¡m? If n > m and
dim ker(f) = n ¡m, show that the equation f (u~ ) = b
~
has infinitely many solutions.
Problem 36. Let A =
2 6
1 3
. For what values of b
~
2R
2
, the equation Au~ = b
~
is solvable? Verify that
the solutions of the equation Au~ =
4
2
has the form u~ = t
¡3
1
+
2
0
for t 2(¡1; 1).
Problem 37. Let f: R
n
!R
n
be a linear mapping and let u
1
~ ; u~
2
be two solutions to equation f(u~ ) = b
~
.
Show that u~
1
¡ u~
2
2 ker(f), and conclude every solution to the equation can be represented by n~ + u~
where n~ 2ker(f ).
Problem 38. Let f : R
n
!R
n
be a linear mapping and suppose u~
1
is a solution to f(u~ ) = b
~
1
and u~
2
is
a solution to f(u~ ) = b
~
2
. Show that u~
1
+ u~
2
is a solution to f(u~ ) = b
~
1
+ b
~
2
.
As we saw above, equation f(u~ ) = b
~
is solvable if b
~
2Im(f). The following problem a nswer
the solvability of a linear equation f(u~ ) = b
~
by the aid of the transpose of f. Remember that
f
t
: R
m
!R
n
is the transpose of f : R
n
!R
m
if the following equality holds for any u~ 2R
n
,
and v~ 2R
m
f(u~ ) ·v~ = u~ ·f
t
(v~):
Problem 39. Let f: R
n
!R
m
be a linear mapping. Show ker(f
t
) = [Im(f )]
?
. Co nclude that the linear
equation f (u~ ) = b
~
is solvable if hb
~
; n~ i= 0 for all n~ 2ker(f
t
). Also show
dim ker(f
t
) ¡dim ker(f) = m ¡n:
Problem 40. Find ker(f
t
) of the matrix A =
2 6
1 3
and verify that b
~
=
4
2
is orthogonal to ker(f
t
).
2 Fun ctions o f se veral variables
2.1 Topology of R
n
Fo r p = (x
1
; ·· ·; x
n
) 2R
n
, the Euclidean norm kpk is defined as
kpk= x
1
2
+ ·· ·+ x
n
2
p
;
and if q = (y
1
; :: :; y
n
), the Euclidean distance is defin ed as follows
kp ¡ qk= (x
1
¡y
1
)
2
+ ···+ (x
n
¡y
n
)
2
p
:
An immediate result of the above definitions is t he convergence of sequences in R
n
.
10 Appendix
Definition 9. A sequence (p
m
)
m=1
1
is called convergent to a if
lim
m!1
kp
m
¡ak= 0:
Proposition 6. A sequence p
m
converges to a if and only if each coordinate of p
m
converges
to its associate coordinate of a.
An open ball of radius r centered at a 2R
n
is defined as
B
r
(a) = fp 2R
n
; kp ¡ak< rg:
a
r
Problem 41. If p
n
!a then any ball B
r
(a) contains infinitely many points of the sequence.
A set D R
n
is called open if for any point a 2D, there is r > 0 such that B
r
(a) D.
A set D R
n
is called bounded if there is r > 0 such that D B
r
. If D is an open set then
its complement D
c
is closed. The complement set D
c
is defined as
D
c
= fp 2R
n
; p 2Dg:
If D is a set, its closure, cl(D) is the smallest closed set containing D, and bnd(D) denotes
the boundary set of D. A point a is called a boundary point of a set D if for any r > 0, the
following relation holds
B
r
(a) \D =/ ;; B
r
(a) \D
c
=/ ;:
The above statements means that any ball centered at a crosses both D and D
c
. We have
bnd(B
r
) = fp; kpk= rg, and cl(B
r
) = fp; kpkrg, and B
r
c
= fq; kqk> rg.
Problem 42. A set D R
n
is closed if and only if any convergent sequence (p
n
); p
n
2D converges in D.
Problem 43. Consider the set A =
1
n
; n = 1; 2; 3; ···
. Determine if A is open or closed.
Problem 44. Let D R
n
be any set, show that bnd(D) is closed.
Problem 45. Show that the set A =
y; 0 < y <
1
x
; x > 0
is open. Find bnd(A).
Problem 46. If D
1
; D
2
are open sets show that D
1
[D
2
and D
1
\D
2
are open. Repeat the argument
if D
1
; D
2
are closed.
2 Functions of several variables 11
2.2 Straight lines and planes in R
3
The parametric e quation of a straight line passi ng through a point p
0
= (x
0
; y
0
; z
0
) and
parallel to a vector r~ = (a; b; c) is p
0
~ + t r~ , or equivalently
x(t) = x
0
+ at; y(t) = y
0
+ bt; z(t) = z
0
+ ct:
If a =/ 0; b =/ 0; c =/ 0, then we can rewrite the equation as follows
x ¡x
0
a
=
y ¡y
0
b
=
z ¡z
0
c
:
(p~ ¡p
0
~ )jjr~
r~
p~ ¡ p
0
~
p
0
p
Similarly, the equation of a straight plane in R
3
passing through a given point p
0
= (x
0
;
y
0
; z
0
), and perpendicular to a given vector n~ = (a; b; c) is n~ ·(p~ ¡ p
0
~ ) = 0, or equivalently
a(x ¡x
0
) + b(y ¡y
0
) + c(z ¡z
0
) = 0;
or equivalently ax + by + cz = d, for some constant d.
The intersection of two planes in R
3
can be empty or a line depending on their position
to each other. For example, two planes P
1
: a
1
x + b
1
y + c
1
z = d
1
and P
2
: a
2
x + b
2
y + c
2
z = d
2
intersect if their normal vectors n~
1
= (a
1
; b
1
; c
1
) and n~
2
= (a
2
; b
2
; c
2
) are not parallel to each
other. In this case, the intersection line will be parallel to n~
1
×n~
2
n~
1
×n~
2
=
i
^
j
^
k
^
a
1
b
1
c
1
a
2
b
2
c
2
= (b
1
c
2
¡c
1
b
2
; c
1
a
2
¡a
1
c
2
; a
1
b
2
¡b
1
a
2
);
12 Appendix
Hence, the equation of the intersection line is
x ¡x
0
b
1
c
2
¡c
1
b
2
=
y ¡y
0
c
1
a
2
¡a
1
c
2
=
z ¡z
0
a
1
b
2
¡b
1
a
2
;
where (x
0
; y
0
; z
0
) is a point on the intersection of P
1
; P
2
.
Example 4. Find the inte rsection of two plane P
1
: 2x + y = 1, P
2
: y ¡z = ¡1.
Solution. The associated normal vectors of two plains are n
~
1
= (2; 1; 0), n
~
2
= (0; 1; ¡1), and they are
not parallel. The intersection line of two plains are in the direction of
n~
1
×n~
2
= (¡ 1; 2; 2):
Obviously, the point p
0
= (0; 1; 2) lies on both plains, and thus the intersection line equation is
x
¡1
=
y ¡1
2
=
z ¡2
2
:
2.3 Scalar functions
A function f w ith the domain D
f
R
n
is called a scalar function if Im
f
R. For example,
a mapping that measures the temperature of ea ch point of a room is a s calar function.
The graph of a scalar function y = f(x
1
; :::x
n
) is the s etf(x
1
; :::x
n
; y)g R
n+1
where
(x
1
; :::; x
n
) 2 D
f
, the domain of f. The graph of a function z = f(x; y) is the surface
f(x; y; f(x; y))g. Fo r example, function z = x
2
+ y
2
is a paraboloid in the (x; y; z)-space
The set ff(x; y) = cg for a fixed c is called the level set of f with valu e c. For example,
the level sets of f(x; y) = x
2
+ y
2
, is the set of circles of radius c
p
centered at the origin. A
level set is also called an implicit function, for example x
2
+ y
2
+ z
2
= c
2
, that is a sphere of
radius c in R
3
.
2 Functions of several variables 13
Definition 10. A scalar function f : D !R has a limit L at a 2D if for any sequence p
m
2D,
p
m
=/ a, the convergence p
m
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
R
n
a implies f(p
m
)
!
!
!
!
!
!
!
!
!
!
!
!
R
L, that is, for any " > 0, there is δ > 0
such that if 0 < kp ¡ak< δ then jf (p) ¡Lj< ".
It is simply seen that function f (x; y) =
xy
x
2
+ y
2
, does not have a limit at (0; 0). In fact, for
sequence (x
m
; y
m
) =
¡
1
m
; 0
, the limit is 0. For s equence (x
m
; y
m
) =
¡
0;
1
m
, the limit is again
0, but the limit is
1
2
for sequence
¡
1
m
;
1
m
.
Problem 47. Determine if the following functions have a limit at (0; 0)
a) f =
x
2
y
x
2
+ y
2
b) f =
sin(x) + sin(y )
x + y
Definition 11. A scalar function f: D !R is called continuous at a 2D if for any sequence
(p
m
), p
m
! a, the sequence f (p
m
) converges to f(a). The statement is equivalent t o the
following: for any " > 0, there is δ > 0 such that jf(p) ¡f(a)j< " for all p 2B
δ
(a).
Problem 48. Sup pose f: R
n
!R is co ntinuous. Show that for any open interval J, the set f
¡1
(J) = fp;
f(p) 2J g is open.
Problem 49. Show that the following function is not continuous at (0; 0).
f(x; y) =
8
<
:
x
2
y
x
4
+ y
2
(x; y) =/ (0; 0)
0 (x; y) = (0; 0)
:
2.4 Vecto r functions
Let I R be an interval. A mapping f: I !R
n
is called a vector valued function. A vector
valued function f is usually denoted by f(t) = (f
1
(t); :::; f
n
(t)) for parameter t in I.
14 Appendix
Example 5. The image of mapping f(t) = (cos(t); sin(t)) is a unit circle in the plane (x; y)
as it satisfies the relation [x(t)]
2
+ [y(t)]
2
= 1. The image of mapping f(t) = (cos t; sin t; t) is
a helix in R
3
as shown below.
A vector function f(t) has a limit at t
0
if and only if all co o rdinate functions f
k
(t) has a
limit at t
0
. Similarly, f is continuous if its all coordinate functions are continuous. We write
lim
t!t
0
f(t) = L
~
, if lim
t!t
0
f
k
(t) = l
k
for all k = 1; :: :; n, and L = (l
1
; :::; l
n
). A vector function is also
called a one dimensional parametric function.
The derivative of vector function is defined coordinate-wise, i.e.,
f
0
(t) = (f
1
0
(t); :::; f
n
0
(t)):
Fo r a fixed t
0
2I, f
0
(t
0
) is the tangent vector on the curve f(I) at t
0
as long as f
0
(t
0
) exists.
If f
0
(t
0
) does not exist, we say f is singular at t
0
. For example, function f(t) =
t
2
3
; t
is
singular at t = 0.
2 Functions of several variables 15
If f (t) denotes the trajectory of a particle in R
n
, f
0
(t
0
) is called the velocity vector of
that pa rtic le, and kf(t
0
)k is equal to its speed. The parametric representation of a curve
provides us with more information than its image. For example, the image of mapping
f(t) = (cos(!t); sin(!t)) is a unit circle for all values of ! =/ 0, however, if f(t) represents the
trajectory of a particle, the speed of the particle would be a function o f !, as the relation
kf
0
(t)k = ! justifies it. The following figure shows the velocity vectors for ! = 0.5; 1; 2
respectively from left to right.
2.5 Parametric mappings
A mapping f : D
f
R
n
!R
m
is called a parametric mapping if m > 1. For example if n = 2
and m = 3, mapping f(t; s) = (f
1
(t; s); f
2
(t; s); f
3
(t; s)) defines a parametric surface in R
3
.
s
x
z
y
t
16 Appendix
Parametric mappings are ve ry common fo r representing complex surfaces. For example,
a torus is represented by the following parametric mapping
f(θ; φ) = ((c + a cos(θ))cos(φ); (c + a cos(θ)) sin(φ); a sin(θ))
where c > a are some constants. The shape is shown below.
Figure 5.
The restriction of f (t; s) to a curve in the (t; s) plane is mapped to a curve in the surface
f(t; s). For example, γ: φ = θ/6 is a straight line in (φ; θ) plane, and it is mapped under f
to the following space curve
¡(θ) = ((c + a co s(θ))cos(θ/6); (c + a cos(θ)) sin(θ/6); a sin(θ)):
The following parametric ma ppings are respectively representations of a sphere and a cylinder
S: (sin(θ) cos(φ); sin(θ) sin(φ); cos(θ)); C: (cos(θ); sin(θ); z)
The concept of limit and continuity is coordinate-wise as well, that is, f has a limit at p 2D
f
if each coordinate function has a limit at p 2D
f
.
3 Derivatives
3.1 Derivatives of sc alar functions
3.1.1 Partial derivatives.
Let D
f
be a n open set in R
n
. For a scalar function f: D
f
!R, the partial derivative @
k
f at
p 2D
f
is defined by the following limit
@
k
f(p) = lim
t!0
f(p + te^
k
) ¡ f(p)
t
; (4)
as long as the limit exists. Fo r a two variable function f(x; y), the partial derivatives @
x
f ;
@
y
f at p = (a; b) are defined respectively as follows
@
x
f(a; b) = lim
t!0
f(a + t; b) ¡ f (a; b)
t
;
@
y
f(a; b) = lim
t!0
f(a; b + t) ¡f(a; b)
t
;
3 Derivatives 17
as long as the limits exist.
Remark 1. For the sake of simplicity, we use the flat notations @
x
; @
y
in this book instead
of the standard ones
@
@x
;
@
@y
. Another notation for the partial derivative is f
x
; f
y
for
@f
@x
;
@f
@y
.
Remark 2. Similarly, we can d efi ne partial derivative functions @
x
f ; @
y
f in the open set
D
f
, the domain of f as
@
x
f(x; y) = lim
t!0
f(x + t; y) ¡ f(x; y)
t
;
@
y
f(x; y) = lim
t!0
f(x; y + t) ¡f(x; y)
t
;
for (x; y) 2D
f
.
Remark 3. The existence of partial derivatives of a function at a point doe s not guarantees
the continuity of the function at that point. Consider the following function
f(x; y) =
(
xy
x
2
+ y
2
(x; y) =/ (0; 0)
0 (x; y) = (0; 0)
:
Even though, @
x
f(0; 0); @
y
f(0; 0) exist and are equal zero, the function is not continuous at
the origin. However, f(x; y) must be continuous and differentiable with respect to x at (a; b)
in order that @
x
f(a; b) exists. Similarly, f(x; y) must be continuous and differentiable with
respect to y a t (a; b) in order that @
y
f(a; b) exists.
3.1.2 Interpretations of partial derivatives
Like a single variable function, there are two related interpretations of partial derivatives.
Consider a two variable function f (x; y) defined on an open set D
f
. Fix a point (a; b) 2
D
f
, and consider the horizontal line parallel to x-axis passing through (a; b). The partial
derivative @
x
f(a; b ) measures the rate of change of f at (a; b) along the horizontal line, and
similarly, @
y
f(a; b) measures the rate of change of f at (a; b) along the vertical line passing
through (a; b). For exam ple, the rate of change of function f(x; y) = x
2
+ y
2
p
at (1; 1) along
x-axis is
@
x
f(1; 1) =
x
x
2
+ y
2
p
(1;1)
=
1
2
p
:
The slope of tangent lines to the surface of f(x; y) are expressed in terms of partial deriva-
tives. The projection of line (a + t; b), for t 2(¡c; c) for some c > 0 on the graph of z = f(x; y)
is a curve of the following form
¡
1
(t) = (a + t; b; f(a + t; b)):
This space curve passes through (a; b; f(a; b)) at t = 0. It is simply seen that @
x
f(a; b) is
equal to the slope of tangent line to ¡
1
(t) in the (x; z)-plane at t = 0:
d¡
1
dt
t=0
= (1; 0; @
x
f(a; b))
18 Appendix
Similarly, @
y
f(a; b) is equal to the slope of tangent line to the curve ¡
2
(t) = (a; b + t;
f(a; b + t)) at t = 0 in the (y; z)-plane. The following figure shows the g raph of function
f(x; y) = x
2
+ y
2
p
, and the projection of γ
1
(t) = (1 + t; 1) on it:
¡
1
(t) =
¡
1 + t; 1; 2 + t
2
+ 2t
p
:
The black line is the tangent to the space curve ¡
1
at t = 0. The slope of the tangent line, is
the tangent of the ang le the line makes with the horizontal line γ
1
, that is,
m =
d
dt
2 + t
2
+ 2t
p
j
t=0
=
1
2
p
:
Note that
d¡
1
dt
(0) is th e tangent vector to the curve ¡
1
at time 0
d¡
1
dt
(0) =
1; 0;
1
2
p
:
A similar argument holds for @
y
f, that is, if ¡
2
is the projection of γ
2
(t) = (1; 1 + t) on the
graph of f , then
d¡
2
dt
(0) =
0; 1;
1
2
p
:
Vectors
d¡
1
dt
(0);
d¡
2
dt
(0) are both tangent to the g raph of f at (1; 1; 2
p
) and thus the plane
span
d¡
1
dt
(0);
d¡
2
dt
(0)
is the tangent plane to the surface of f at (1; 1; 2
p
). The algebraic
equation o f the tangent plane is derived by the air of n~
n~ = v~
1
×v~
2
=
¡
1
2
p
; ¡
1
2
p
; 1
;
3 Derivatives 19
and thus the algebraic of the tangent plane is derived as
¡
1
2
p
(x ¡1) ¡
1
2
p
(y ¡1) + z ¡ 2
p
= 0:
Remark 4. For a general function z = f(x; y), two principa l tangent lines at p = (a; b) are
v~
1
= (1; 0; @
x
f(a; b)); v~
2
= (0; 1; @
y
f(a; b));
and thus n~ , the normal vector to the graph of f at p is
n~ = (¡@
x
f(p); ¡ @
y
f(p); 1):
Accordingly, the algebraic equation on the tangent plane at p is
¡@
x
f(a; b) (x ¡a) ¡@
y
f(a; b) (y ¡b) + z ¡f(a; b) = 0:
3.1.3 Chain rule
Let f be a scalar function on D
f
R
2
, and a ssume that x; y are functions of another variable,
say t, i.e., x = x(t); y = y(t). In the final analysis, f(x; y) is a function of t and thus the
ordinary derivative
df
dt
can exist.
df
dt
(t) = lim
h!0
f(x(t + h ); y(t + h)) ¡ f (x(t); y(t))
h
:
Proposition 7. Assume that f is differentiable with respect to x and y , and moreover the
partial derivatives are continuous. If x(t); y(t) are differentiable functions of t, then the
following equality that is called chain rule holds
df
dt
= @
x
f(x(t); y(t))
dx
dt
+ @
y
f(x(t); y(t))
dy
dt
: (5)
There is an important interpretation for the above fo rmula. First, note that γ:(x(t); y(t))
defines a parametric curve in the (x; y)-plane, and thus f (x(t); y(t)) can be considered as the
restriction of f to γ. Also, we can consider γ as the path of a particle moving in the (x; y)-
plane. Therefore, relation (5) states the rate of change of the value of that particle along γ.
Fo r example, if f is the density distribution function in the plane, then relation (5) states
how fast or slow the density of a particle changes when it moves along path γ.
f(γ(0))
f(γ(t
1
))
γ(t)
f(γ(t
2
))
x
y
20 Appendix
On the other hand, since the graph of f is a surface in R
3
, f(γ(t)) is the projection of
γ(t) in the surface as shown in the following figure. With this interpretation, relation (5)
defines the slope of tangent to the curve at any instance of time.
Let us consider again equality (5) and rewrite it as follows
@
x
f
dx
dt
+ @
y
f
dy
dt
=
@
x
f
@
y
f
·
0
@
dx
dt
dy
dt
1
A
:
Note that
0
B
B
@
d x
dt
dy
dt
1
C
C
A
is just the tangent vector of the curve γ = (x(t); y(t)), i.e., γ
0
(t). Vector
@
x
f
@
y
f
!
is called the gradient of f and is denoted by grad(f) or rf. Therefore, equality (5)
can be rewritten as
df
dt
= rf · γ
0
(t);
and for this reason,
df(γ(t))
dt
is called also the derivative of f along γ(t). The chain rule can
be extended to higher dimensions. For example, if f(x; y) is a differentiable function with
respect to x; y and x = (x(t; s)); y = y(t; s) are differentiable functions then
@
t
f = @
x
f@
t
x + @
y
f@
t
y;
@
s
f = @
x
f@
s
x + @
y
f@
s
y:
Problem 50. If u = u(t; x), a nd x = x(t), find
du
dt
.
Problem 51. If u = f(x ¡2t) find @
t
u and @
x
u.
3.1.4 Directional derivative
Partial derivatives are just spe cial cases of a mo re general derivative called the directional
derivative. Assume that a direction vector v^ = (v
1
; v
2
) is given (a direction vector is a unit
vector), and f: D
f
! R is a given continuous function. T he directional derivative of f at
(a; b) 2D
f
along v^ is defined by the following limit
@
v^
f(a; b) = lim
t!0
f(a + tv
1
; b + tv
2
) ¡ f(a; b)
t
;
3 Derivatives 21
as long as the limit exists. If so, then @
v^
f(a; b) measures the rate of change of f a t (a; b)
along v^. Obviously if v^= (1; 0) then @
v^
f(a; b) = @
x
f(a; b) and if v^= (0; 1), it would be equal
to @
y
f(a; b).
Proposition 8. If @
x
f ; @
y
f are continuous at (a; b), that is,
lim
(x;y)!(a;b)
@
x
f(x; y) = @
x
f(a; b); lim
(x;y)!(a;b)
@
y
f(x; y) = @
y
f(a; b);
then
@
v^
f(a; b) = rf(a; b) ·v^
The continuity in the above proposition is crucial, for example, consider the following
function
f(x; y) =
8
<
:
x
2
y
x
2
+ y
2
(x; y) =/ (0; 0)
0 (x; y) = (0; 0)
:
If v^=
1
2
p
;
1
2
p
, then
@
v^
(0; 0) = lim
t!0
t
3
2 2
p
t
2
t
=
1
2 2
p
;
however, rf(0; 0) =
0
0
, and thus rf (0; 0) · v^ = 0. The reason is that @
x
f ; @
y
f are not
continuous at (0; 0). To see this, let us find @
y
f for (x; y) =/ 0 as
@
y
f(x; y) =
x
2
(x
2
¡y
2
)
(x
2
+ y
2
)
2
;
and observe that @
y
f does not have even a limit at (0; 0). Note a lso that the directional
derivative of a function f is a special case of the chain rule rf ·γ
0
(t).
Problem 52. For the following function
f(x; y) =
8
<
:
x
2
y
x
4
+ y
2
(x; y) =/ (0; 0)
0 (x; y) = (0; 0)
show that @
r^
f(0; 0) exists for any direction r^ but partial derivatives are not continuous at (0; 0).
3.1.5 Gradient
Consider the le vel curve defined by γ: f (x; y) = c. If γ( t) = (x(t); y(t)) is a parametrization
of this curve then rf ·γ
0
(t) = 0 for any t as long as rf is a continuous vector function. This
relation means that rf(γ(t)) is always perpendicular to γ(t). For example, the level curve
x
4
+ 2x
2
y + x
2
+ y
2
= 1;
22 Appendix
has the gradient
rf =
4x
3
+ 4xy + 2x
2x
2
+ 2y
!
:
The following figure shows a few of gradient vectors f on the level curve
As it is observed, gradient vectors are perpendicular to level curves. On the other hand, if
n^ = (n
1
; n
2
) is the direction vector at a point on the level curve, the the directional derivative
@
n~
f of f is rf ·n^, and since n^ =
rf
krf k
(as long as the rf =/ 0), we obtain
@
n^
f = krf k:
Therefore, the magnitude of f at a point measures the rate of change of f along the normal
direction on the level curve. This result is extremely useful to maximize (or minimize) a
scalar function. The procedure is as follows. To maximize f (x; y), we fix an initial point
p
0
= (x
0
; y
0
). The next point p
1
is obtained by the following relation
p
1
= p
0
+ λ
rf(p
0
)
krf(p
0
)k
;
where λ > 0 is a small value. Geometrically, that mean we take a step of length λ along the
direction rf (p
0
). Iterating this procedure, that is,
p
n+1
= p
n
+ λ
rf(p
n
)
krf(p
n
)k
;
converges the maximum point of f a s long such a local or global maximum exists, and if f
satisfies some other verifiable conditions.
In above, we used frequently the operator nabla r=
0
B
B
@
@
1
·
·
·
@
n
1
C
C
A
. It is applied to differentiable
function as rf =
0
B
B
@
@
1
f
·
·
·
@
n
f
1
C
C
A
. We study this operator in more detail later in this app endix.
3 Derivatives 23
Problem 53. Show the following relations
a) r(f + g) = rf + rg.
b) r(kf ) = k rf , k 2R.
c) r(fg) = f rg + g rf .
3.1.6 Derivative and differential
Let f: D
f
R
n
!R be a s calar function defined on an open set D
f
. T he derivative of f at
p
0
is the linear mapping D
p
0
f: R
n
!R such that the following relation holds for any h
~
2R
n
lim
h
~
!0
f(p
0
+ h
~
) ¡ f(p
0
) ¡D
p
0
f(h
~
)
kh
~
k
= 0 (6)
Obviously if f is diffe rentiable at p
0
then it is continuous at that point. Moreover, such
a linear mapping must be unique . Note that if a function f is differentiable at a point a
then directional derivatives along any direction at a exist. The reverse holds only if partial
derivatives are continuous at a.
Problem 54. Verify that the above definition is compatible with the usual definition for sin gle variable
functions.
Problem 55. Show that if f is differentiable at p
0
it must be continuous at that point. Show also that
its derivative (the linear mapping) is unique.
Proposition 9. Assume that f has continuous partial derivatives at p
0
, that is,
lim
p!p
0
@
k
f(p) = @
k
f(p
0
);
for k = 1; :::; n, then D
p
0
f exists and has the matrix representation
D
p
0
f = [@
1
f(p
0
); ···; @
n
f(p
0
)]:
According to the above proposition, if f has continuous partial derivatives at p
0
then for
any h
~
2R
n
, the following relation holds.
D
p
0
f(h
~
) = rf (p
0
) ·h
~
:
Fo r this reason, some texts write D
p
0
f = rf (p
0
) that we should keep in mind that D
p
0
f is
a 1 ×n matrix, while rf (p
0
) is a n ×1 vector.
24 Appendix
Example 6. For example, function
f(x; y) =
8
<
:
x
2
y
x
4
+ y
2
(x; y) =/ (0; 0)
0 (x; y) = (0; 0)
has directio nal derivatives in all directions at the origin, howeve r, the function is not dif-
ferentiable at this point since it is not even continuous at the origin (why?). On the other
hand, functionf(x; y) = x
2
+ y
2
has continuous partial derivatives @
x
f = 2x; @
y
f = 2y and for
any p
0
= (x
0
; y
0
), we have
D
p
0
f = 2[x
0
; y
0
]:
Let us verify definition (6) for f at p
0
= (1; ¡1). For arbitrary h
~
= (h
1
; h
2
), we have
lim
h
~
!0
(1 + h
1
)
2
+ (¡1 + h
2
)
2
¡2 ¡2(h
1
¡h
2
)
h
1
2
+ h
2
2
p
= lim
h
~
!0
h
1
2
+ h
2
2
h
1
2
+ h
2
2
p
= lim
h
~
!0
kh
~
k= 0:
An immediate result of definition (6) is the linear approximation formula. If f: D
f
!R
is continuously differentiable at p
0
2D
f
, that is
lim
p!p
0
D
p
f = D
p
0
f ;
then
f(p) f(p
0
) + D
p
0
f(p~ ¡p~
0
);
or equivalently
f(p) f(p
0
) + rf (p
0
) ·(p~ ¡ p~
0
):
Fo r functions of two variables, the above formula reads
f(x; y) = f(x
0
; y
0
) + @
x
f (x ¡x
0
) + @
y
f (y ¡y
0
):
Note that the right hand side is the equation of tangent plane at (x
0
; y
0
; f (x
0
; y
0
)):
T (x; y) = (x
0
; y
0
) + @
x
f (x ¡x
0
) + @
y
f (y ¡y
0
):
Theorem 3. (Mean Value Theorem) Assume that f: D
f
! R is continuously dif-
ferentiable everywhere in D
f
. Fix a point p
0
2 D
f
. Then for any point p 2 D
f
, there is
ξ 2tp + (1 ¡t)p
0
, t 2(0; 1) such that
f(p) = f(p) + rf (ξ) ·(p~ ¡ p~
0
):
Problem 56. Prove the theorem. H int: define g(t)= f(tp + (1 ¡t)p
0
) and apply the mean value theorem
for the single variable function g(t); t 2[0; 1].
3 Derivatives 25
Definition 12. The total differential of f: D
f
!R at p
0
is defined by the following formula
df(p
0
) = @
1
f(p
0
) dx
1
+ ···+ @
n
f(p
0
) dx
n
:
Remember that for a single variable function y = y(t), the differential dy is defined by
the relation dy(t
0
) = y
0
(t
0
) dt. Fig.(6) b elow shows this relation geometrically.
dt
dy(t
0
)
t
0
y = y(t)
y(t
0
)
t
0
dt
dy = y
0
(t
0
) dt
y
t
T (x )
Figure 6.
Similarly for a 2-variable function z = f (x; y), dz is defined by the f o llowing relation and
its geometry is represented in Fig. 7.
dz = @
x
fdx + @
y
fdy:
(x
0
; y
0
)
dz = @
x
fdx + @
y
fdy
dy
dx
x
y
f(x
0
; y
0
)
z
Figure 7.
3.1.7 Critical points and local max and min
Let D
f
R
n
be a n open set. A point a 2D
f
is called a local min (or max) of f: D
f
!R if
there is a ball B
δ
(a) such that f(p) f(a) (alternatively f(p) f(a )) for all p 2 B
δ
(a). If
f is differentiable at a, a nd if a is a local min or max, then D
a
f is a zero mapping. To see
26 Appendix
this, suppose a is a local min, choose an arbitrary direction vector h
~
, and write
0 = lim
t!0
f(a + th
~
) ¡ f(a) ¡D
a
f(th
~
)
kth
~
k
= lim
t!0
f(a + th
~
) ¡ f (a) ¡tD
a
f(h
~
)
jtj
:
Fo r t > 0, we have
D
a
f(h
~
) = lim
t!0
f(a + t h
~
) ¡ f(a)
t
0
Fo r t < 0, we have
D
a
f(h
~
) = lim
t!0
f(a + t h
~
) ¡ f(a)
t
0;
and thus D
a
f(h
~
) = 0 for arbitrary h
~
, and thus D
a
f is a zero mapping (meaning rf(a) is a
zero vector).
Definition 13. (Critical point) A point a is called a critical point of a function f if either
D
a
f does not exist or D
a
f is a zero mapping. If D
a
f is a zero mapping, a can be a local min,
local max, a saddle point or non of them.
In order to determine the type of a critical point in terms of min, ma x or saddle, we
need the notion o f second derivatives. Second order partial derivatives @
ij
f are defined as
@
ij
f = @
i
(@
j
f). We have the following theorem.
Theorem 4. Let D
f
R
n
be an open set, and f: D
f
!R . Furthermore assu me that @
ij
f is
continuous, then @
ij
f = @
ji
f.
Theorem 5. Assume that f : D
f
!R is continuously differentiable of order 2 on open set D
f
.
Fix a 2D
f
, then there is ξ 2tp + (1 ¡t)a for some t 2(0; 1) such that the following relation
holds for any for any p 2D
f
f(p) = f(a) + rf(a) ·(p~ ¡a~ ) + (H
f
(ξ)(p~ ¡a~)) ·(p~ ¡a~ );
where H
f
is the Hessian matrix of f defined as H
f
= [@
ij
f]
i;j
.
Corollary 1. Assume that f is second order continuously differentiable function and rf(a)=
0, then a is a local min if H
f
(a) is a positive definite matrix, a is a local max if H
f
(a)
is a negative definite matrix, and a saddle point if H
f
(a) has eigenvalues with opposite signs.
The standard example of above three cases is f = x
2
+ y
2
; f = ¡ x
2
¡y
2
, and f = x
2
¡ y
2
as shown below
3 Derivatives 27
Note that H
f
(a) is a symmetric matrix, and thus it has n orthogonal eigenvectors, and
the Jordan form diag(λ
1
; :::; λ
n
). Therefore if all eigenvalues of H
f
(a) is positive then a is a
local min, if its all eigenvalues are negative then a is local max, a nd if there are some po sitive
and some negative eigenvalues, it is a saddle point. If there is at least one zero eigenvalue,
or equivalently det(H
f
(a)) = 0, then a may not be any of these types.
Problem 57. Find all critical point s of the function f(x; y)= x
3
+ y
3
+
9
2
x
2
¡
3
2
y
2
+6x and classify them.
3.2 Derivative of non-s calar mappings
3.2.1 Jacobi matrix
Let D
f
R
n
be an open set and f: D
f
!R
m
a continuous map. f is differe ntiable at a 2D
f
if there is a linear mapping D
a
f: R
n
!R
m
that satisfies the following relation
lim
h
~
!0
kf(a + h
~
) ¡ f (a) ¡D
a
f(h
~
)k
kh
~
k
= 0:
Equivalently, f = (f
1
; :: :; f
m
) is differentiable at a if and only if each coordinate function f
k
is differentiable at a. On the other hand, the derivative of f
k
at a is
D
a
f
k
= [@
1
f
k
(a); :::; @
n
f
k
(a)];
and since each f
k
is defined on R
n
, we obtain
D
a
f =
0
B
B
B
B
B
B
@
@
1
f
1
(a) @
2
f
1
(a) ··· @
n
f
1
(a)
@
1
f
2
(a) @
2
f
2
(a) ··· @
n
f
2
(a)
·
·
·
·
·
·
·
·
·
·
·
·
@
1
f
m
(a) @
2
f
m
(a) ::: @
n
f
m
(a)
1
C
C
C
C
C
C
A
:
The above matrix is called the Jacobi matrix of f at a, denoted also by J
f
(a). Let γ(t) be
a smooth curve passing through a at t = 0. This curve is mappe d into R
m
by f as f(γ(t)).
In the final analysis, f(γ(t)) is a vector valued function and therefore, we have
df(γ(0))
dt
= D
γ(0)
f(γ~
0
(0)):
28 Appendix
Since γ
0
(0) is the tangent vector on γ(t) at t = 0, D
γ(0)
f(γ~
0
(0)) is the tangent vector on
f(γ( t)) at t = 0. For example, for f(x; y) = (x
2
¡y
2
; x
2
+ y
2
) we have
D
(1;1)
f =
2 ¡2
2 2
:
If γ(t) = (e
¡t
; e
t
) (passing through (0; 0) at t = 0), we have γ
0
(0) =
¡1
1
and accordingly,
D
γ(0)
f(γ
0
(0)) =
¡4
0
. Note that
d
dt
f(γ(t))j
t= 0
=
d
dt
(e
¡2t
¡e
2t
; e
¡2t
+ e
¡2t
)j
t= 0
=
¡4
0
:
3 Derivatives 29
Theorem 6. Assume that the Jacobi matrix of a mapping f: D
f
R
n
!R
n
is invertible at
a. Then there is a neighborhood B
δ
(a) such that f is one to one on B
δ
(a).
Note that det(J
f
(a)) measures the volume of a parallelogram constructed on columns of
J
f
(a). In other word, if C is a unit cube made at a, then det(J
f
(a)) is equal to the volume
of parallelogram J
f
(a)(C). Hence, if det(J
f
(a)) =/ 0, there is a neighborhood B
δ
(a) such that
f is one to one on B
δ
(a).
3.2.2 Smooth surfaces
Remember that a smooth space curve is represented by a curve map γ(t) such that γ
0
(t) is
nonzero. Geometrically, this condition means that γ(t) always admits a tangent vector that
varies continuously along γ.
Definition 14. A surface S in R
3
is called smooth if S has a nonzero normal (perpendicular)
vector at all points on S.
Let f (t; s) 2 R
3
be parametric surfa ce. Consider an arbitrary point f (t
0
; s
0
) on S.
Coordinate line γ
1
(t) = (t + t
0
; s
0
) is mapped on S as ¡
1
(t) = f(t + t
0
; s
0
). The tangent vector
to this space curve is just ¡
1
0
(t
0
) = @
t
f(t
0
; s
0
). Similarly, the coordi nate line γ
2
(s) = (t
0
; s + s
0
)
is mapped as ¡
2
(s)= f (t
0
; s +s
0
) and the tangent vector is ¡
2
0
(s)=@
s
f(t
0
; s
0
). Notice that both
¡
1
0
; ¡
2
0
are tangent to S and therefore n~ := ¡
1
0
ס
2
0
is perp endicular to S a s long as it is nonzero
n~ =
i
^
j
^
k
^
@
t
x (t
0
; s
0
) @
t
y(t
0
; s
0
) @
t
z(t
0
; s
0
)
@
s
x (t
0
; s
0
) @
s
y(t
0
; s
0
) @
s
z(t
0
; s
0
)
=/ 0:
Let us see the result for a smooth f unction z = f (x; y). The graph of function is
φ(x; y) = (x; y; f(x; y));
and thus @
x
φ = (1; 0; @
x
f) and @
y
φ = (0; 1; @
y
f). The normal vector n~ is
n~ = @
x
φ ×@
y
φ = (¡@
x
f ; ¡@
y
f ; 1):
Example 7. For p
0
= (0; 1; 1) on the surface of f (x; y) = x
2
+ y
2
, consider coordinate lines
γ
1
(x) = (x; 1); γ
2
(y) = (0; y). We have f (γ
1
) = (x; 1; 1 + x
2
), f(γ
2
) = (0; y; y
2
), see the figure
shown below. Respectively, the tangent vectors are as T
1
= (1; 0; 0), T
2
= (0; 1; 2). The normal
vector n~ to the surface at p
0
is
n~ = T
1
×T
2
= (0; ¡2; 1);
and therefore the equation of tangent plane is ¡ 2(y ¡1) + z ¡1 = 0.
30 Appendix
Fo r a surface represented by the implicit function S: f (x; y; z) = 0, the normal vector n~
is derived by the f o llowing procedure. Consider an arbitrary space curve γ = (x(t); y(t); z(t))
on S, that is, f(x(t); y(t); z(t)) = 0. The chain rule s tates
df
dt
(γ) = rf (γ(t)) ·γ
0
(t) = 0:
Since γ
0
(t) is tangent to S, then rf is perpendicular to γ(t) if rf is nonzero. On the
other hand, since γ is arbitrary, then γ
0
belongs to the tangent plane on S, and thus rf is
perpendicular t o S. Therefore, we obtain n~ as
n~ = (@
x
f ; @
y
f ; @
z
f):
Problem 58. Write the equation of curve formed by the intersection of the unit sphere x
2
+ y
2
+ z
2
= 4
and the plane x + y + z = 1.
3.3 Implicit function theorem
An implicit function f(x; y) = 0 defines generally a planar curve in the (x; y) p lane. If
y = y(x), then by the chain rule, we can write
dy
dx
= ¡
@
x
f
@
y
f
:
However, there is no guarante e in general that y could be solved in terms of x or x could
be solved in t erms of y. Question is this: is there any function y = g(x) defined on an open
interval I such that f(x; g(x)) = 0 for all x 2I. The following theorem answers the question.
Theorem 7. (implicit function theorem) Suppose implicit function f(x; y) = 0 satisfies
the following conditions
i. there is a point p
0
= (x
0
; y
0
) such that f(x
0
; y
0
) = 0,
ii. there is an open ball B
"
(p
0
) such that f has continuous partial derivat ives on it,
3 Derivatives 31
iii. and that @
y
f(x
0
; y
0
) =/ 0,
then there is an open interval I = (x
0
¡δ; x
0
+δ), and a function y = g(x) such that y
0
= g(x
0
),
and f(x; g(x)) = 0 for all x 2I.
Example 8. Consider the following function
e
xy
+ x + sin(y) = 1:
The function defines a planar curve which is shown below in Fig.8.
-2 0 2
-4
-2
0
2
4
x
=
0
.
49
x
= 1
.
96
Figure 8.
The slopes at x = ¡0.49 and x = 1.96 are infinity, which implies @
y
f = 0 at those points
according to the formula y
0
= ¡
@
x
f
@
y
f
. Now fix the point p
0
= (0; 0) on the curve. As it is seen
from the figure, there is an explicit function y = g(x) for x 2(¡0.49; 1.96) such that
e
x g(x)
+ x + sin(g(x)) = 1: (7)
The result can be generali zed for functions f: R
n
!R as follow s.
Theorem 8. Assume t hat implicit function f (x
1
; x
2
; :::; x
n
) = 0 satisfies the following
conditions
i. there is a point a = (a
1
; a
2
; :::; a
n
) such that f(a
1
; ·· ·; a
n
) = 0,
ii. There is an open ball B
"
(a) such that f has continuous partial derivatives on it,
iii. and that @
n
f(a) =/ 0,
then there exists a ball B
δ
at a
0
= (a
1
; ···; a
n¡1
) and a function g: B
δ
(a
0
) ! R such that
a
n
= g(a
1
; ·· ·; a
n¡1
) and f(x
1
; x
2
; :::; g(x
1
; :: :; x
n¡1
)) = 0 for all (x
1
; :::; x
n¡1
) 2B
δ
(a
0
).
32 Appendix
4 Integrals of mutivariable functi ons
4.1 Line integrals
Let γ: (a; b) !R
n
be a smooth curve map, that is γ
0
(t) =/ 0. The length of γ(a; b) is defined
by the following integral
L =
Z
a
b
jγ
0
(t)jdt:
This definition coincides the intuitive notion of the length. For example, if the image of
γ(t) is a curved metal wire, and if we straightened the wired, we get the same length if we
calculate the length by the above integration. In the above definition, γ
0
(t) is the tangent
vector on γ at t. The quantity dl = jγ
0
(t)jdt is ca lle d the differential arc length.
Now let D R
n
be an open set and let γ be a smooth curve in D. Assume that f: D !R
is a continuous function. The integral of f along γ is defined by the follow ing integral
I =
Z
a
b
f(γ(t))dl =
Z
a
b
f(γ(t)) jγ
0
(t)jdt:
Geometrically this integral is equal to the surface area constructed on the base γ(t) an the
height f (γ(t)). In parti cular if f = 1, the above integral gives the arc length of γ. If f denote
the density function of a metal wire represented by γ, the integral denote the total mass of
the wire.
4 Integrals of mutivariable functions 33
Problem 59. Consider the metal wire in the shape of the semi-circle x
2
+ y
2
= 1, y 0. If the density
of the wire is given by ρ = k(1 ¡ y) for a constant k, find the center of mass of the wire.
4.2 Inte grals over a bounded do main of R
2
Now, let R: [a; b] ×[c; d] be a rectangle and f: R R
2
!R be a continuous function. Similar
to the integrals of single variable function, we can define the following double integral
I =
Z Z
R
f(x; y )dA;
by the aid of a n infinite sum called the Riemann sum as
I = lim
n; m!1
X
j=1
m
X
i=1
n
f(x
i
; y
j
) j
ij
j:
Here
ij
is a partition of R, j
ij
j is the area of the rectangle ij and (x
i
; y
j
) is an arbitrary
point in
ij
. We have the following theorem. Geometrically, I denote the volume constructed
on the base R, and height f .
Theorem 9. (Fubini) Assume that the f is continuous (or piecewise continuous) in R = [a;
b] ×[c; d]. Then we have
Z
R
f(x; y) dA =
Z
a
b
Z
c
d
f(x; y) dy
dx =
Z
c
d
Z
a
b
f(x; y) d x
dy:
Now, assume that D R
2
is a closed bounded domain and assume that f: D ! R is
continuous. We can inscrib e D inside a rectangle R and extend f on R as follows
f
~
(x; y) =
f(x; y) (x; y) 2D
0 (x; y) 2D
:
Even though f
~
may be discontinuous, we have the following fact
Z Z
D
f(x; y) dA =
Z Z
R
f
~
(x; y) dA:
Problem 60. Let I = [a; b] and assume that f (s; x) and
@f
@x
(s; x) are continu ou s in I × I. Use the
fundamental theorem of calculus and prove the following formula
d
dx
Z
x
0
x
f(s; x) ds = f(x; x) +
Z
x
0
x
@
@x
f(s; x) ds:
In the pro of yo u may need to pass the limit inside the integral.
4.3 Change of variables in multiple integrals
Of most important techniques to calculate a double integral (and also triple integrals) is the
change of variable technique. Let D R
2
be a bounded domain, and assume f: D !R is a
continuous function. The goal is to calculate the integral
I =
Z Z
D
f(x; y) dA:
34 Appendix
Now assume that there is a one to one mapping ': R R
2
! D, where R is rectangle [a;
b] ×[c; d]. If so, then we can calculate I in terms of an integral over R. The advantage is that
the integration over rectangular domains are much more simpler than general domains. The
procedure is as follows. Let '(u; v) = (x; y), where (u; v) 2R. See the following figure. The
differential area dA in (x; y)-plane in terms of the differential area dS in the (u; v)-plane is
dA = jdet(J
'
)jdS;
where J
'
is the Jac o bi matrix of the one to one transformation '.
u
v y
x
R
D
'
dS
dA
Accordingly, we have the fo llowing formula called the change of variable technique
Z Z
D
f(x; y) dA =
Z Z
R
f('(u; v)) jJ
'
(u; v)jdu dv:
Problem 61. Find the domain D formed by the transforming rectangle [1; 2] × [1; 2 ] under the
transformation
x = u/v
y = uv
:
Calculate the following integral
Z Z
D
e
y/x
p
e
xy
p
dA:
4.4 Surface integrals over a surface in R
3
Let S be a smooth surface in R
3
with the representation
'(u; v) = (x(u; v); y(u; v); z(u; v));
where (u; v) 2D R
2
, and assume that f is a continuous functions defined on S: f: S !R.
We want to calculate the following integral
I =
Z Z
S
f(x; y; z) dA;
where dA is the differential area of the surface S. Here we again transform the integral over
S as an integral over D as follows. Note that
dA = jj@
u
' ×@
v
'jjdS ;
4 Integrals of mutivariable functions 35
where dS is a differential area in D. Re member that @
u
' and @
v
' are tangent vectors on
S and k@
u
' × @
v
'k is equal to the area of parallelogram constructed on vectors @
u
'; @
v
'.
Therefore, we can write
Z Z
S
f(x; y; z)dA =
Z Z
D
f('(u; v)) jj@
u
' ×@
v
'jjdS:
Problem 62. If f = f (x; y) is a smooth function defined on the bounded set D R
2
, show that the
area of the surface associated to u is
A =
Z
D
1 + j@
x
f j
2
+ j@
y
f j
2
q
dx dy:
Problem 63. We show that the change of variable formula is independent of the transformation.
a) Assume ': D !S and : D
1
!S are two one to one transformations with image S. Define the
map '~ := '
¡1
: D
1
!D. Verify
jj@
t
×@
s
jj= jj@
u
' ×@
v
'jjjj@
t
'~ ×@
s
'~jj:
b) Now show
Z
D
f('(u; v)) jj@
u
' ×@
v
'jjdu dv =
Z
D
1
f( (s; t))j@
t
×@
s
jdsdt:
We frequently use the polar and spherical coordinates f o r integ rals. In polar coordinate,
the transformation is defin ed by the relations '(r; θ) = (r cosθ; r sinθ). The area differential
dS in this case is
dS =
cosθ sinθ
¡r sinθ r cosθ
dr = rdr:
If f(x; y) is a function defined in disk B
a
, a disk of radius a cente red at the origin, its double
integral in the polar coordinate is
Z
B
a
f(x; y)dA =
Z
0
2π
Z
0
a
f(r cosθ; r sinθ) rdr:
Problem 64. Show the following inequality
π
4
(1 ¡e
¡a
2
)
Z
0
a
Z
0
a
e
¡x
2
¡y
2
ddxdy
π
4
(1 ¡e
¡2a
2
);
and conclude
lim
a!1
Z
0
a
Z
0
a
e
¡x
2
¡y
2
dA =
π
4
:
Use the above result and nd
I =
Z
0
1
e
¡x
2
dx:
In the spherical coordinate, the transformation is
'(ρ; φ; θ) = (ρ cosφ s inθ; ρ sinφ sinθ; ρ cosθ):
The volume differential in this coordinate is
dV =
sinθ cosφ sinθ sinφ cosθ
ρ c o sθ cosφ ρ cosθ sinφ ¡ρ sinθ
¡ρ sinθ sinφ ρ sinθ cosφ 0
= ρ
2
sinθ:
36 Appendix
If f(x; y; z) is defined, for example, in a sphere of radius a, its integral on this sphere is equal
to
Z
S
a
f(x; y; z)dS = a
2
Z
0
2π
Z
0
π
f(a sinθ cosφ; a sinθ sinφ; a c o sθ) sinθdφ:
5 Calcul us of vector fields
5.1 Vecto r field
Let D R
n
be an open set. A vector field is a mapping f: D !R
n
. For e a ch p 2D, f(p) is
a vector in R
n
, and thus we can interpret this mapping as an assignment a vector f(p) to
every point p 2D, that is, p 7! f (p). This assignment is called a vector field.
A vector field p 7!f(p) is continuous if the association varies continuously with respect to
p, that is, the any change from p to an adjacent point q, the vector f (p) continuously varies
to f (q). Mathematically speaking, this is equivalent to the continuity of f as a mapping
on D. Remember that f = (f
1
; :::; f
n
) is a continuous mapping if and only if ea ch scalar
function f
k
is a continuous function. Similarly, a vector field p 7!f (p) is called c o ntinuously
differentiable if and only if all its coordi nate functions are continuously differentiable.
A vector fields models many physical phenomena. For example, an electrical charge q
located at the origin, generates an electrical field in the space as
E(r) =
q
4π"
0
r~
kr~ k
3
;
where "
0
is the permittivity constant of the space. This field is a force field an is completely
similar to the gravitational field generated by a mass M:
g(r) = GM
r~
kr~ k
3
;
where G is the universal constant.
5.2 Vecto r fields and differential equations
The theory of ordinary differential equations can be f o rmulated in terms of vector fields. In
fact, the system
8
>
>
>
>
<
>
>
>
>
:
dx
1
dt
= f
1
(x
1
; :::; x
n
)
·
·
·
dx
n
dt
= f
n
(x
1
; :: :; x
n
)
:
defines a vector field f = (f
1
; :::; f
n
) on some domain D, and a parametric curve γ(t) =
(x
1
(t); :::; x
n
(t)) such that the tange nt vector on γ at each instant of time t
0
that is γ
0
(t
0
)
coincides with the vector assigned to point γ(t
0
). In other word, f(γ(t)) = γ
0
(t) for all t in
an open interval. For example, the following system
x
0
= ¡y
y
0
= x
;
5 Calculus of vector fie lds 37
defines vector field f = (¡y; x). As we know, the trajectory of a point p
0
: (x
0
; y
0
) according
to the above system is the line
γ(t) = (x
0
cost ¡y
0
sint; x
0
sint + y
0
cost):
It is simply seen that
γ
0
(t) = (¡x
0
sint ¡ y
0
cost; x
0
cost ¡ y
0
sint);
that is equal to f (γ(t)). Notice that γ(t) is just the rotation mapping applied to p
o
γ(t) =
cost ¡sint
sint cost
x
0
y
0
;
that coincides with the geometry of vector field f
3
.
5.3 Gradient field
A vector fiend p 7! f(p) is called a potential, conservative or just gradient field if there is
a scalar function φ such that f = ¡rφ. The negative sign is just for historic reason. For
example, the electric field E(r) or the gravitational field g(r) are po tential. It is simply seen
that E(r) = ¡rφ, where φ is the following scalar function:
φ(r) =
q
4π"
0
1
kr~ k
:
Potential fields satisfies very nice properties that we study in sequel. In the following figure,
three vector fields are shown: f
1
= (x; y), f
2
= (¡x; ¡y), and f
3
= (¡y; x).
1
.
0
0
.
5 0
.
0 0
.
5 1
.
0
1
.
0
0
.
5
0
.
0
0
.
5
1
.
0
1
.
0
0
.
5 0
.
0 0
.
5 1
.
0
1
.
0
0
.
5
0
.
0
0
.
5
1
.
0
1
.
0
0
.
5 0
.
0 0
.
5 1
.
0
1
.
0
0
.
5
0
.
0
0
.
5
1
.
0
It is seen that f
1
; f
2
are potential field with potentials φ
1
= ¡
1
2
x
2
¡
1
2
y
2
, φ
2
=
1
2
x
2
1
2
y
2
. The
filed f
3
is not potential, i.e., there is no potential function φ such that f
3
= rφ.
Problem 65. Prove the ab ove claim, that is, show there is no scalar function φ such that f
3
= rφ.
The fo rce field generated by a potential is also called conservative. To see the reason, let
us write the second Newton's law for a un it mass as
8
<
:
d
2
x
dt
2
= p(x; y)
d
2
y
dt
2
= q(x; y)
;
38 Appendix
where the eld f = (p; q) is potential generated by a potential function φ(x; y). Let γ(t) be
the trajectory of a particle initially located at (x
0
; y
0
). The n we have
γ
00
(t) = f(γ(t)) = ¡rφ(γ(t)):
Let us define the energy along γ(t) as follows
E(t) =
1
2
kγ
0
(t)k
2
+ φ(γ(t )):
The derivative of E along γ(t) is then
dE
dt
= γ
00
(t) · γ
0
(t) +
d
dt
φ(γ(t)):
We have
d
dt
φ(γ(t)) = rφ(γ(t)) · γ
0
(t);
and thus
dE
dt
= γ
00
(t) · γ
0
(t) + rφ · γ
0
(t) = (γ
00
+ rφ) · γ
0
= 0:
Therefore, the derivative of the energy function along a trajectory γ(t) is zero, in other world,
the energy is conserved along the trajectory of the particle.
5.4 Divergence, curl and Laplacian
Two important operations on smooth vector fields are divergence and curl. The divergence
of a field f = (f
1
; :::; f
n
) at a point p is de fined by the following relation
div f(p) =@
1
f
1
(p) + ···+ @
n
f
n
(p)
It is simply seen that the divergence of a vector field f is equal to the dot product of
nabla r and field f as div f = r:f. It is also equal to the trace of matrix D
p
f. Intuitively
speaking, div f(p) measure the flow of net flux passing th rough p. If div f(p) > 0, point p
acts like a source that emits or generate flow. All points in filed f
1
= (x; y) are source points
since div f (p) = 2. If div f(p) < 0, point p acts like a sink that absorbs or attracts flow to
itself. All point s in field f
2
= (¡x; ¡y) are sink since div f (p) = ¡2. If div f(p) = 0, then the
net flow passing through p is zero, that the net amount o f incoming flow is equal to outgoing
flow. All points in field f
3
= (¡y; x) are of th is type.
Problem 66. If φ is a smooth scalar function and let f be a smooth vector field, show the following
formula
div (φf) = φ div f + f ·rφ (8)
Another important operation regarding a vector field in R
3
is the curl of the field at a
point. The curl of a vector field f = (f
1
; f
2
; f
3
) at p is de fined by the following relation
cur(f)(p) =
0
B
B
@
i
^
j
^
k
^
@
x
@
y
@
z
f
1
f
2
f
3
1
C
C
A
(p) = (@
2
f
3
¡@
3
f
2
; @
3
f
1
¡@
1
f
3
; @
1
f
2
¡@
2
f
1
)(p):
5 Calculus of vector fie lds 39
Symbolically, we can write the curl in terms of r as r×f. Since curl(f) is a vector at every
point, it also defines a ne w vector field in its domain. At each point p, curl(f )(p) measures the
rotation of vector field f in three directions: 1) compo nent @
2
f
3
(p) ¡@
3
f
2
(p) that measures
the rotation of f at p around x-axis, 2) component @
3
f
1
(p) ¡ @
1
f
3
(p) that measures the
rotation of f at p around y-axis, 3) and @
1
f
2
(p) ¡@
2
f
1
(p) that measure the rotation around
z-axis . For example, curl (¡y; x) = 2 k
^
at all points, that means the filed rotates around z-
axis with constant speed at a ll points. This rotation is e vident from the figure of the field.
Assume that φ is a smooth scalar function defined on an open subset D of R
n
. The
Laplacian of φ is defined by the following relation
φ = @
11
φ + ···+ +@
nn
φ:
It is simply seen that φ =div (grad φ).
Problem 67. Consider field f = (x
2
¡y
2
; y
2
¡z
2
; z
2
¡x
2
).
a) Find r× f at the point (1; 2; 3).
b) Verify that (r× f): i
^
is equal to r×(0; y
2
¡z
2
; z
2
¡1).
Problem 68. Assume that φ is a smooth scalar function defined in R
3
. Show r×rφ = 0.
Problem 69. Show the following relations for smooth elds f ; g in R
3
and smooth function φ
a) r×(φf) = φr× f ¡f ×(rφ).
b) r:(f × g) = g:(r× f) ¡ f(r× g).
Problem 70. If f is a smooth vector field, what is r: (r× f)?
Problem 71. Show the following relation
r×(r× f ) = r(r:f) ¡ f ;
where f = (∆f
1
; f
2
; f
3
) for a smooth vector field f = (f
1
; f
2
; f
3
).
5.5 Line integrals i n vector elds
Assume that the f: D R
n
!R
n
is a continuous field, and C is a smooth curve in D. We
define a method for the integral of f along curve C. For this, we take a parametrization of
C as γ: (a; b) !D, where γ(a; b) = C. The desired integral is defined as fo llows
Z
C
fdc =
Z
a
b
f(γ(t)) · γ
0
(t) dt: (9)
It is seen that the integral is independent of parametrization, that is, if γ
1
: (c; d) ! D is
another parametrization of C, then
I =
Z
c
d
f(γ
1
(t)) ·γ
1
0
(t) dt:
Problem 72. Assume that γ: (a; b) ! D, γ
1
: (c; d) !D are two smooth parametrizations of C. Show
the relation
Z
a
b
f(γ(t)) · γ
0
(t)dt =
Z
c
d
f(γ
1
(t)) · γ
1
0
(t) dt:
40 Appendix
If f is a force field, then the integral of f along a curve is called the work done by f . For
example, let γ(a) = p
0
, γ(b) = p
1
, then I measure the total work done by f to move a particle
from point p
0
to point p
1
along path γ. It is is interesting to note that if f is gradient, i.e.,
f = rφ for some potential φ, then the integral is independent of path along a particle moves
from p
0
to p
1
. This important fact is shown below
Z
a
b
rφ(γ(t)) ·γ
0
(t) dt =
Z
a
b
(γ(t)) = φ(γ(b)) ¡ φ(γ(a)) = φ(p
1
) ¡ φ(p
0
):
The path independence property of a field implies that the integral of field over any closed
curve is zero, that is, if γ is a clo sed curve then
I
γ
f(γ(t)) · γ
0
(t) dt = 0: (10)
We have the following theorem.
Theorem 10. Assume f: D R
n
!R
n
is a smooth vector field. If the integral of f over all
closed curves in D is zero, then f is a conservative field on D.
Problem 73. Let the condition of the above theorem holds. Fix p
0
2D. For any p 2D define φ as
φ(p) =
Z
0
t
f(γ(t)) · γ
0
(t) dt;
where γ(t) is an arbitrary smooth curve in D such that γ(0) = p
0
, and γ(t) = p.
a) Verify that φ is independent of path γ
b) Show that φ is a potential for f. For example, for two dimensional field f = (p(x; y); q(x; y) ),
ver ify the relation
lim
h!0
φ(x + h; y) ¡ φ(x; y)
h
= p(x; y); lim
h!0
φ(x; y + h) ¡ φ(x; y)
h
= q(x; y):
Problem 74. A smooth field f = (p; q) in R
2
is called exact if for all (x; y) 2R
2
the following relation
holds
@
y
f(x; y) = @
x
q(x; y):
a) Show that f is conservative.
b) Consider eld f =
¡y
x
2
+ y
2
;
x
x
2
+ y
2
defined everywhere in R
2
except t he origin. Show the following
relation
@
y
¡y
x
2
+ y
2
= @
x
x
x
2
+ y
2
:
Now, consider closed curve γ(t) = (cos t; sin t), t 2[0; 2π], calculate the line integral of f along γ.
The result is nonzero. Does it contradict the fact claimed in (a)?
5.6 Surface integrals in ve ctor fields
Let S be a surface in R
3
parameterized by a smooth map ¡(t; s) where (t; s) 2D for some
open domain D. Note that ¡ is smooth if @
t
¡ ×@
s
¡ is nonzero for all points in D. Let smoo th
field f = (f
1
; f
2
; f
3
) given in R
3
. The integral of f on the surface S is defined by the followin g
integral
I =
ZZ
S
f ·n^ dA ;
5 Calculus of vector fie lds 41
where n^ is the unit normal vector on S. This i ntegral measures the total flux passing outward
through surface S. As it is shown in the following figure, the tangential component of f
never leaves the surface (due to the fact f · n^ = 0), and only the normal component of f
contributes in the total flux. For example, the flow of water through a window measured in
m
3
sec
is a physical model for the flux.
n^
f
ds
normal component
of f
tangent component
of f
Remark 5. In some branches of mathematics, the term flux is considered as a vector and
not scalar. We will see this notion in the book for the mathematical modeling of heat flow
through a conductive media.
Fo r the parametrization ¡(t; s), we have
n^ dA = (@
t
¡ ×@
s
¡) dt ds;
and therefore, the total flux is expressed in terms of double integral
I =
ZZ
D
f(¡(t; s)) ·(@
t
¡ ×@
s
¡) dt ds:
Theorem 11. (divergence) Assume D R
3
is an open set with smooth (piecewise)
boundary. If f is a smooth vector field in cl(D), the following relation holds
ZZZ
D
div f dv =
ZZ
bnd(D)
f ·n^ dA: (11)
Note that If div f = 0 inside D, then the net amount of flow passing throu g h bnd(D)
is zero. Accordingly, the divergence of a field f in R
3
at a point p can be defined by the
following formula
div (f)(p) = lim
r!0
1
Vol(B
r
(p))
ZZ
bnd(B
r
(p))
f ·n^ dA : (12)
The above formula coincides with our previous statement about the physical interpretation
of divergence operator at a point, that div (f )(p) measures the net flux of f passing through
point p.
Problem 75. Let f = (x; y; z). Verify the formula (12) at the origin.
Example 9 . Consider the identity field f (x; y; z) = (x; y; z), and let B be the closed unit
ball in R
3
centered at the origin. The left hand side of formula (11) reads
ZZZ
B
div fdV = 3
ZZZ
B
dV = 4π: (13)
42 Appendix
The unit normal to bnd(B) is n^= (x; y; z), and then the right hand side of f ormula (11 ) reads
ZZ
bnd(B)
(x
2
+ y
2
+ z
2
) dA =
ZZ
bnd(B)
dA = 4π: (14)
Problem 76. Assume f: D R
3
!R
3
a smooth field. Use the divergence th eorem and show
ZZZ
D
f dV =
ZZ
bnd(D)
@
n^
f ds; (15)
where @
n^
f is the directional derivative of f in direction n^.
Problem 77. If D is a ball of radius R in R
n
, use the divergence theorem and show
Vol(D) =
R
n
A(D);
where Vol(D) is the volume of D and A(D) is the surface area of D. (Hint: consider the function
φ =
P
k
x
k
2
).
Problem 78. Let f be a smooth field in R
n
such that
jf(r)j
1
(1 + jjrjj)
n+1
:
Show that
Z
R
n
div (f) = 0 :
Proposition 10. (Integration by parts) Let D R
3
be an open set and f ; g: D !R are
smooth functions. We have
ZZZ
D
(@
x
f) g dV =
ZZ
bnd(D)
fgn
1
dA ¡
Z
D
f (@
x
g) dV ; (16)
where n
1
is the first component of n^ = (n
1
; n
2
; n
3
i at bnd(D). Similar relations hold for the
derivatives with respect to other components y; z.
The following proposition generalizes the ab ove result.
Proposition 11. Let D be a domain in R
3
and bnd(D) is smooth. If f is a smooth vector
field in D and g is a smooth function on D then
ZZZ
D
g div (f ) dV =
ZZ
bnd(D)
g (f ·n^) dA ¡
ZZZ
D
f ·rg dV : (17)
Problem 79. Prove the ab ove proposition by the aid of divergence theorem.
Problem 80. By the above proposition show
ZZZ
D
gfdV =
ZZ
bnd(D)
g @
n^
fdA ¡
ZZZ
D
rf ·rg dV ;
and conclude the following relation called the Green's formula:
ZZZ
D
(g f ¡f g) dV =
ZZ
bnd(D)
[g @
n^
f ¡ f@
n^
g] dA:
5 Calculus of vector fie lds 43
Let S be a smooth surface in R
3
with smooth boundary bnd(S). If f is a smooth field
in R
3
, the following relation is called th e Stoke's theorem:
ZZ
S
(r× f) ·n^ dA =
I
bnd(S)
f ·T
^
dl;
where T
^
is the unit tangent vector on curve bnd(S) and d` is the differential length of that
curve.
Problem 81. Use the Stoke's theorem and show
(r× f (p)) ·i
^
= lim
r!0
1
πr
Z
0
2π
f(γ(t)) · γ
0
(t) dt;
where γ(t) = p + (r cost; r sint; 0). Derive similar formula for the second and third components, i.e.,
(r× f (p)) · j
^
, and (r× f(p)) ·k
^
. Notice that f ·T
^
measures the rotation of f along curve bnd(S).
Problem 82. Let f = (¡y + x; ¡x+z;¡z + y). Verify the relations of the previous problem at the origin.
Problem 83. Show the following relations
a)
ZZ
S
φ curl(f ) ·n^ dA =
ZZ
S
(f ×rφ) ·n^ dA +
I
bnd(S)
φf ·T
^
dl:
b)
ZZZ
D
φ div (f ) dV =
ZZ
S
φf ·n^ dA ¡
ZZZ
D
f ·rφ dV
c)
ZZZ
D
g ·curl(f ) dV =
ZZ
S
(f × g) ·n^ dA +
ZZZ
D
f ·curl(g) dV
We have the following theorem
Theorem 12. If r×f = 0 everywhere, then f is gradient, i.e., there is a potential function
φ such that f = ¡rφ. If div (f ) = 0 everywhere, then there is a vector field g such that
f = curl(g).
6 Orthogonal curvilinear coordin at es
In many applications, it is convenient to use an orthogonal coo rdinate rather than the Carte-
sian one. The form of differential operators in a general orthogonal curvilinear coordinate is
discussed below.
6.1 Unit vectors
Let (q
1
; q
2
; q
3
) be a curvilinear coordinate system, and assume (q
1
; q
2
; q
3
)
!
!
!
!
!
!
!
!
T
(x; y; z) is a one
to one transformation to the Cartesian s pace. We first define the unit vectors q^
1
; q
2
^ and q
3
^
in the directions o f q
1
; q
2
and q
3
, and by the aid of that define the orthogonality notion. If
we consi der the restriction of T to q
1
as a one parameter map γ(q
1
), then the unit vector q^
1
can defined as
q^
1
: =
γ
0
(q
1
)
kγ
0
(q
1
)k
=
@
q
1
T
k@
q
1
T k
;
44 Appendix
and since
T (q
1
; q
2
; q
3
) = (x(q
1
; q
2
; q
3
); y(q
1
; q
2
; q
3
); z(q
1
; q
2
; q
3
));
we have
@
q
1
T = (@
q
1
x; @
q
1
y; @
q
1
z)
Similarly, we can define q^
2
; q^
3
as
q^
2
=
@
q
2
T
k@
q
2
T k
; q^
3
=
@
q
3
T
k@
q
3
T k
:
A coordinate system (q
1
; q
2
; q
3
) is called orthogonal if q^
1
; q^
2
and q^
3
are mutually orthogonal,
i.e., hq^
i
; q^
j
i= δ
ij
.
6.1.1 Polar, cylindrical and spherical coordinates
The polar coordinate (r; θ) is defined by the transformation T (r; θ) = (r cosθ; r sinθ) for
r = [0; 1) and θ 2[0; 2π). The transformation is one to one everywhere in the domain except
at the origin. Since @
r
T = cosθ i
^
+ sin θ j
^
, we obtain
r^ = cosθ i
^
+ sin θ j
^
:
Similarly, @
θ
T = ¡r sin θ i
^
+ r cos θ j
^
, and thus
θ
^
= ¡ sin θ i
^
+ cos θ j
^
:
θ
^
r^
The cylindrical transformation is defined by transformation
T (r; θ; z) = (r cosθ:r sinθ; z)
The unit vectors are derived as
r^ = cos θ i
^
+ sin θ j
^
; θ
^
= ¡ sin θ i
^
+ cos θ j
^
; z^= k
^
:
The spherical coordinate is defined by the transformations
T (ρ; φ; θ) = (ρ cosφ sinθ; ρ sinφ sinθ; ρ cosθ);
for ρ 2[0; 1); φ 2[0; 2π); θ = [0; π]. The mapping is not one to one at ρ = 0; θ = 0; π. The unit
vectors are
ρ^= sinθ cosφ i
^
+ sinθ sinφ j
^
+ cosθ k
^
; (18)
φ
^
= ¡sinφ i
^
+ cosφ j
^
; (19)
θ
^
= cosθ cosφ i
^
+ cos θ sinφ j
^
¡sin θ k
^
: (20)
6 Orthogonal curvilinear coordinates 45
6.2 Nabla r in an orthogonal coordinate
Let f be a smooth scalar function given in an orthogonal coordinate system (q
1
; q
2
; q
3
). We
can write the gradient of f in this system as
rf = f
1
q^
1
+ f
2
q^
2
+ f
3
q^
3
;
that implies f
1
= hrf ; q^
1
i, f
2
= hrf ; q^
2
i, and f
3
= hrf ; q^
3
i. Let us calculate f
1
for example.
rf is coordinate f ree, and thus we can replace it s f o rm in the Cartesian coordinate, i.e.,
rf = @
x
f i
^
+ @
y
f j
^
+ @
z
f k
^
in the a ssociated dot product and derive
hrf ; q^
1
i=
1
k@
q
1
T k
h@
x
f i
^
+ @
y
f j
^
+ @
z
f k
^
; @
q
1
x i
^
+ @
q
1
y j
^
+ @
q
1
z k
^
i:
By the orthogonality, we obtain
hrf ; q^
1
i=
1
k@
q
1
T k
(@
x
f@
q
1
x + @
y
f@
q
1
y + @
z
f@
q
1
z) =
1
k@
q
1
T k
@
q
1
f:
We obtain similar forms for f
2
; f
3
, that are,
hrf ; q^
2
i=
1
k@
q
2
T k
@
q
2
f ; hrf ; q^
3
i=
1
k@
q
3
T k
@
q
3
f
and therefore the operator r in (q
1
; q
2
; q
3
) is
r=
q^
1
k@
q
1
T k
@
q
1
+
q^
2
k@
q
2
T k
@
q
2
+
q^
3
k@
q
3
T k
@
q
3
:
6.2.1 Polar, cylindrical and spherical coordinate
Applying the formula obtained above for polar, cylindrical and spherical sy stems gives respe c-
tively
Polar: r= r^ @
r
+
1
r
θ
^
@
θ
:
Cylindrical: r= r^ @
r
+
1
r
θ
^
@
θ
+ k
^
@
z
:
Spherical: r= ρ^@
ρ
+
1
ρ sin θ
φ
^
@
φ
+
1
ρ
θ
^
@
θ
:
To calculate the Laplacian operator := r:r in (q
1
; q
2
; q
3
) system, we do as follows:
∆: =r:r=
q^
1
k@
q
1
T k
@
q
1
+
q^
2
k@
q
2
T k
@
q
2
+
q^
3
k@
q
3
T k
@
q
3
;
q^
1
k@
q
1
T k
@
q
1
+
q^
2
k@
q
2
T k
@
q
2
+
q^
3
k@
q
3
T k
@
q
3
:
Here we need @
i
(q^
j
) for i; j = 1; 2; 3. It turns out that the derivatives of q^
k
always lies in the
normal plane to it, for example,
@
q
2
(q^
1
) = α q^
2
+ β q^
3
:
The reason is for the relation hq^
i
; q^
i
i= 1 and thus h@
q
j
(q^
i
); q^
i
i= 0.
46 Appendix
Problem 84. In polar coordinate, s how the following relations
@
θ
r^ = θ
^
; and @
θ
θ
^
= ¡r^; (21)
and conclude
f = @
rr
f +
1
r
@
r
f +
1
r
2
@
θθ
f: (22)
Problem 85. In cylindrical coordinate show the relation
f = @
rr
f +
1
r
@
r
f +
1
r
2
@
θθ
f + @
zz
f:
Problem 86. In spherical coordi nate show the following relations
@
φ
ρ^= sin(θ) φ
^
and @
θ
ρ^= θ
^
; (23)
@
φ
φ
^
= ¡sin(θ) ρ^¡cos( θ) θ
^
and @
θ
φ
^
= 0; (24)
@
φ
θ
^
= cos(θ) φ
^
and @
θ
θ
^
= ¡ρ^: (25)
and conclude
f =
1
ρ
2
@
ρ
(ρ
2
@
ρ
f) +
1
ρ
2
sin
2
θ
@
φφ
f +
1
ρ
2
sin θ
@
θ
(sinθ@
θ
f):
7 Fun ction Series
Working in a function vector space where functions play the role of well-known vectors in
R
n
, requires to study sequences whose elements are functions. These type of sequences are
a natural generalization of numeric sequences that we supp o se the reader is familiar with.
7.1 The different notion s of convergence
We assume that the reader is familiar with numeric sequences. Here we consider sequences
and series w hos e elements are functions. Let f
n
: [a; b] !R for n = 1; 2; ::: be a sequence of
continuous functions. We say f
n
converges pointwise in [a; b] to f, and write f
n
!f if
lim
n!1
jf
n
(x) ¡f(x)j= 0;
for all a 2[a; b]. In other word, if a
n
= f
n
(x) for a fixed x, then (a
n
) as a numeric sequence
converges to value b = f(x). This n o tion of convergence is equivalent to the following: for
every x 2[a; b] and eve ry " > 0, there is an integer N
0
= N
0
(x; ") > 0 such that
8n N
0
)jf
n
(x) ¡f(x)j< ":
A function sequence f
n
converges uniformly to f in [a; b] if for for any " > 0, th ere is and
integer N = N (") such that
8n N
0
) max
x2[a;b]
jf
n
(x) ¡f(x)j< ":
The norm of a continuous function in a closed interval [a; b] is defined as
kf k= max
x2[a;b]
jf (x)j:
7 Function Series 47
Fo r this reason, the uniform convergence is usually written as
lim
n!1
kf
n
¡f k= 0:
Remark 6. The pointwise a nd uniform convergence are two different types of convergence
and they may should not be considered as equivalent. In fact, if f
n
! f uniformly, th en
f
n
!f pointwise, while the converse is not generally true. For example, consider the s equence
of functions f
n
(x) = x
n
in [0; 1]. The sequence converges pointwise to f(x) =
1 x = 1
0 otherwise
.
However, it is seen
max
x2[0;1]
jx
n
¡f(x)j>
1
2
;
and thus f
n
does not converges uniformly to f . In fact, f
n
does not converge uniformly to
any function.
Problem 87. If a funct ion se quence (f
n
) converges pointwise or uniformly, then its limit function is
unique.
Problem 88. If a function sequence (f
n
) converges uniformly to f in [a; b] then it converges pointwise
to f as well.
Problem 89. Assume tha t f
n
: [a; b] !R is a sequence of continuous functions converges uniformly to
f(x). Show that f(x) is continuous.
There are other notions of convergence that we study in the book, for example, the
convergence in norm as
lim
n!1
Z
a
b
jf
n
(x) ¡f(x)jdx = 0:
Fo r example, f
n
(x) = x
n
in [0; 1] co nverges to f (x) =
1 x = 1
0 otherwise
or f(x) 0 or f (x) =
(
1 x =
1
2
0 otherwise
. As we see, the limit function is not unique in the usual sense, and we have to
do something to re medy this situation.
Problem 90. If (f
n
) converges uniformly to f in [a; b], it converges to f in the following sense
Z
a
b
jf
n
(x) ¡ f(x)jdx !0:
Problem 91. Assume that a function sequence (f
n
) of continuous functions on [a; b] converges in norm
to f as
Z
a
b
jf
n
(x) ¡ f (x)jdx !0
Show that
lim
n!1
Z
a
b
f
n
(x) dx =
Z
a
b
f(x) dx:
7.2 δ-sequence
A function sequence f
n
(x) is called a Dirac δ-sequence fu nc tion at x = 0 if fo r any continuous
bounded function g(x) at x = 0, we have
lim
n!1
Z
g(x) f
n
(x) dx = g(0):
48 Appendix
Fo r example, the sequence
f
n
(x) =
(
n
2
¡
1
n
< x <
1
n
0 otherwise
;
is a δ-sequence. If g(x) is a ny bounded continuous f unction at x = 0, then
min
x2
¡
1
n
;
1
n
g(x)
Z
¡
1
n
1
n
g(x)
n
2
dx max
x2
¡
1
n
;
1
n
g(x);
and thus
lim
n!1
Z
¡
1
n
1
n
g(x)
n
2
dx = g(0):
7.3 Differentiatio n and integrati on of function sequences
Let (f
n
) be a function sequence in (a; b) or [a; b]. If f
n
! f pointw ise or uniformly, can we
claim
Z
a
b
f
n
(x)dx !
Z
a
b
f(x)dx
when n !1? Let us see an important example:
f
n
(x) =
8
>
>
>
>
<
>
>
>
>
:
2
2n
x x 2
0;
1
2
n
2
n
(2 ¡2
n
x) x 2
1
2
n
;
1
2
n¡ 1
0 otherwise
:
The graph for some n are shown in Fig.9. Each function f
n
is a triangle with the base
0;
1
2
n¡ 1
and the height 2
n
such that the area under each triangle is equal 1 independent of n.
4
2
1
1
2
8
1
4
1
8
Figure 9.
Therefore, we have for all n:
Z
0
1
f
n
(x)dx = 1:
7 Function Series 49
On the other hand, it is seen f
n
converges to f 0 on [0; 1] (why?) and thus
Z
0
1
f
n
(x)dx 9
Z
0
1
f(x)dx:
In other word,
lim
n!1
Z
f
n
(x) dx =/
Z
lim
n!1
f
n
(x)
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
{z
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
}
f(x)
dx:
Problem 92. Assume that (f
n
), f
n
: I !R is a sequence of continuous functions converging uniformly
to f . Show
lim
n!1
Z
a
b
f
n
(x) dx =
Z
a
b
f(x) dx:
The following theorem gives a sufficient condition for passing the limit inside an integral.
Theorem 13. (dominant convergence) Assume that (f
n
) is a sequence of continuous
functions converging pointwise to a function f in (a; b)(the interval may be finite or infinite).
If there is a function g(x) such that
jf
n
(x)jg(x); 8x 2(a; b);
and
Z
a
b
g(x) dx < 1;
then
lim
n!1
Z
a
b
f
n
(x) dx =
Z
a
b
f(x) dx:
We also n eed the following form of the above theorem.
Corollary 2. Assume that f (t; x) is a continuous function defined on (c; d) × (a; b), and
furthermore
f(t; x) g(x); 8t 2(c; d)
such that
Z
a
b
g(x) dx < 1
Then function
F (t) =
Z
a
b
f(t; x) dx;
is continuous in (c; d).
Problem 93. Prove the corollary by the aid of Theorem (13).
The same story of differentiation is the same. If a sequence of differentiable functions
(f
n
) converges pointwise or even uniformly to f , then there is no guarantee that f
n
0
!f
0
. As
a simple example, sequence f
n
(x) =
1
n
sin(nx) defined on (0; π) converges uniformly to the
constant function f (x) = 0. However, f
n
0
(x) = cos(nx) which is not a convergent sequence.
50 Appendix