Mathematical Analysis
Mathematical Analysis
Introduction
in which the author tries to explain why studying this note is useful, and
gives fatherly advice on how to do so.
In a sense, mathematical analysis can be said to be about continuity.
The epsilon–delta arguments that you meet in a typical calculus course
represent the beginnings of mathematical analysis. Unfortunately, too
often these definitions are briefly presented, then hardly used at all
and soon forgotten in the interest of not losing too many students and
because, frankly, it is not that important in elementary calculus. As
mathematics becomes more abstract, however, there is no way to pro-
ceed without a firm grounding in the basics. Most PDEs, for example,
do not admit any solution by formulas. Therefore, emphasis is on differ-
ent questions: Does a solution exist? If so, is it unique? And if so, does
it depend on the data in a continuous manner? When you cannot write
up a simple formula for the solution of a PDE, you must resort to other
methods to prove existence. Quite commonly, some iterative method is
used to construct a sequence which is then shown to converge to a solu-
tion. This requires careful estimates and a thorough understanding of
the underlying issues. Similarly, the question of continuous dependence
on the data is not a trivial task when all you have to work with is the
existence of a solution and some of its properties.
How to read these notes. The way to read these notes is slowly.
Because the presentation is so brief, you may be tempted to read too
Version 2004–01–15
Elements of mathematical analysis 2
much at a time, and you get confused because you have not properly
absorbed the previous material. If you get stuck, backtrack a bit and
see if that helps.
The core material is contained in the first four sections — on met-
ric spaces, completeness, compactness, and continuity. These sections
should be read in sequence, more or less. The final two sections, one
on ordinary differential equations and one on the implicit and inverse
function theorems, are independent of each other.
The end of a proof is marked with in the right margin. Sometimes,
you see the statement of a theorem, proposition etc. ended with such
a box. If so, that means the proof is either contained in the previous
text or left as an exercise (sometimes trivial, sometimes not — but
always doable, I hope). If a proof is not complete, then this is probably
intentional — the idea is for you to complete it yourself.
Metric spaces
in which the basic objects of study are introduced, and their elementary
properties are established.
Most of mathematical analysis happens in some sort of metric space.
This is a set in which we are given some way to measure the distance
d(x, y) between two points x and y. The distance function (metric) d
has to satisfy some simple axioms in order to be useful for our purposes.
Later, we shall see that a metric space is the proper space on which to
define continuity of functions (actually, there is a more general concept
— that of a topological space — that is even more appropriate, but we
shall not need that level of abstraction here).
d(x, x) = 0
d(x, y) > 0 if x 6= y
d(x, y) = d(y, x) (symmetry)
d(x, z) ≤ d(x, y) + d(y, z) (the triangle inequality)
Version 2004–01–15
3 Elements of mathematical analysis
Before moving on to the examples, we shall note that the triangle in-
equality can easily be generalised to more than three elements. For
example, two applications of the triangle inequality yields the inequal-
ity
d(x, w) ≤ d(x, y) + d(y, w) ≤ d(x, y) + d(y, z) + d(z, w).
In fact, it is not difficult to prove the general inequality
n
X
d(x0 , xn ) ≤ d(xk−1 , xk )
k=1
2 Examples.
R or C with d(x, y) = |x − y|.
qP
n
Rn or Cn with d(x, y) = kx − yk (where kxk = j=1 |xj |2 ).
Any set X with d(x, x) = 0 and d(x, y) = 1 whenever x 6= y. This is
called a discrete metric space.
Version 2004–01–15
Elements of mathematical analysis 4
4 Examples.
qP
n
Rn or Cn with kxk = j=1 |xj |2 .
Pn
Rn or Cn with kxk = j=1 |xj |.
n n
R or C with kxk = max{|xj | : 1 ≤ j ≤ n}.
The space l∞ consisting of all bounded sequences x = (x1 , x2 , . . .) of
real (or complex) numbers, with kxk = max{|xj | : 1 ≤ j ≤ ∞}.
Note that x ∈ Bε (x), and that Bε (x) may consist of no points other
than x itself (if, for example, X is given the discrete metric and ε ≤ 1).
It should also be noted that B̄ε (x) is not necessarily the closure of Bε (x)
(for example, with the discrete metric, B1 (x) = {x} while B̄1 (x) = X).
Version 2004–01–15
5 Elements of mathematical analysis
8 Exercise. Show that no sequence can have more than one limit.
As a result of the above exercise, we can talk about the limit of a
convergent sequence, and write limn→∞ xn for the limit. Though a se-
quence can have only one limit, a non-convergent sequence can have
many limit points:
Version 2004–01–15
Elements of mathematical analysis 6
(a) F is closed,
(b) the complement X \ F is open,
(c) for every x ∈ X, if every neighbourhood of x has a nonempty
intersection with F , then x ∈ F ,
(d) F contains every limit point of every sequence contained in F .
Version 2004–01–15
7 Elements of mathematical analysis
Completeness
in which we already encounter our first Theorem.
Version 2004–01–15
Elements of mathematical analysis 8
and since 0 < K < 1 it is then clear that (xn ) is a Cauchy sequence,
and hence convergent since X is complete. Let x be the limit.
We need to prove that x is a fixed point. We know that xn is an ap-
proximate fixed point when n is large, in the sense that d(f (xn ), xn ) →
0 when n → ∞ (because d(f (xn ), xn ) = d(xn+1 , xn ) ≤ K n d(x1 , x0 )).
We perform a standard gymnastic exercise using the triangle inequality:
f (x) is close to f (xn ) = xn+1 which is close to x. More precisely:
Thus, for any ε > 0 we can use the above inequality with a sufficiently
large n to obtain d(f (x), x) < ε, and since ε > 0 was arbitrary, we must
have d(f (x), x) = 0. Thus f (x) = x, and x is indeed a fixed point of f .
It remains to prove the uniqueness of the fixed point. So, assume x
and y are fixed points, that is, f (x) = x and f (y) = y. Then
and since 0 < K < 1 while d(x, y) ≥ 0, this is only possible if d(x, y) =
0. Thus x = y, and the proof is complete.
In many applications of the fixed point theorem, we are given a
function which is not a contraction on the entire space, but which is
so locally. In this case, we need some other condition to ensure the
existence of a fixed point. In the following very useful case, it turns out
that the proof of the Banach fixed point theorem can be adapted.
Version 2004–01–15
9 Elements of mathematical analysis
K < 1 is so that
Km
d(xn , xm ) < d(x1 , x0 )
1−K
provided x0 , . . . , xn are defined. With m = 0, this becomes
1
d(xn , x0 ) < d(f (x0 ), x0 ) ≤r
1−K
using the assumption. Thus xn ∈ B̄r (x0 ), and therefore we can define
xn+1 . By induction, then, xn is defined and in B̄r (x0 ) for all n. The
proof that this sequence converges to a limit which is a fixed point is
just like before.
Version 2004–01–15
Elements of mathematical analysis 10
diam A = sup{d(x, y) : x, y ∈ A}
24 Exercise. The above proof has several gaps and details left out.
Identify these, and fill them in.
Compactness
in which we define a most useful property of metric spaces such as closed and
bounded intervals.
Version 2004–01–15
11 Elements of mathematical analysis
(a) X is compact,
(b) every open cover of X contains a finite cover of X,
(c) every set of closed subsets of X with the finite intersection prop-
erty has nonempty intersection,
(d) X is totally bounded and complete.
Proof: We prove (d) ⇒ (c) ⇒ (a) ⇒ (d). The proof of the equivalence
(b) ⇔ (c) will be left as an exercise. (Hint: F is a set of closed sets
with the finite intersection property but with empty intersection if and
only if {X \ F : F ∈ F } is an open cover with no finite subset which is
also a cover.)
(d) ⇒ (c): Assume X is totally bounded and complete, and let F
be a set of closed subsets of X with the finite intersection
Sn property. If
ε > 0, by the total boundedness we may write X = k=1 B̄ε (xk ). Let
Gk = {F ∩ B̄ε (xk ) : F ∈ F }. At least one of the families G1 , . . . , Gn has
the finite intersection property (exercise: prove this). Clearly each set
in Gk has diameter at most 2ε.
So far we have proved: Every set F of closed sets with the finite
intersection property has a refinement G (by which we mean a family
of closed sets, also with the finite intersection property, so that for every
F ∈ F there exists some G ∈ G with G ⊆ F ), each of whose members
has diameter no larger than some prescribed positive number.
Let now εk & 0. Let F0 = F and, for k = 1, 2, . . . let Fk be a
refinement of Fk−1 , each of whose members has diameter at most εk .
Version 2004–01–15
Elements of mathematical analysis 12
Version 2004–01–15
13 Elements of mathematical analysis
ω = inf{sup F : F ∈ F 0 }.
34 Corollary. R is complete.
Version 2004–01–15
Elements of mathematical analysis 14
Continuity
in which we, at last, study the continuous functions, without which the study
of metric spaces would be a fruitless and boring activity. As an application, we
consider the problem of moving a differentiation operator under the integral
sign.
Version 2004–01–15
15 Elements of mathematical analysis
(a) f is continuous;
(b) for each open subset V ⊆ Y , f −1 [V ] is open;
(c) for each closed subset F ⊆ Y , f −1 [F ] is closed.
Version 2004–01–15
Elements of mathematical analysis 16
Proof: Let ε > 0. For every x ∈ X there is some δ(x) > 0 so that
ρ(f (ξ), f (x)) < ε whenever d(ξ, x) < δ(x).SBy the compactness of X
n
there are x1 , . . . , xn ∈ X so that X = j=1 Bδ(xj )/2 (xj ). Let δ =
min{δ(x1 ), . . . , δ(xn )}/2.
Now, if ξ, x ∈ X, then x ∈ Bδ(xj )/2 (xj ) for some j. If furthermore
d(ξ, x) < δ then ξ ∈ Bδ(xj ) (xj ) as well, and so
ρ(f (ξ), f (x)) ≤ ρ(f (ξ), f (xj )) + ρ(f (xj ), f (x)) < ε + ε = 2ε.
Version 2004–01–15
17 Elements of mathematical analysis
and
! Z
1 Z b Z x b
∂f ∂f
(ξ, y) dξ dy − (x0 , y) dy
x − x0 x0 ∂x a ∂x
a
1 Z b Z x
∂f ∂f
= (ξ, y) − (x0 , y) dξ dy
x − x0 a x0 ∂x ∂x
Z bZ x
1 (ξ, y) − ∂f (x0 , y) dξ dy
∂f
≤
|x − x0 | a x0 ∂x ∂x
Z bZ x
1
< ε dξ dy
|x − x0 | a x0
= |b − a|ε
Version 2004–01–15
Elements of mathematical analysis 18
uniform continuity is not enough, but if you can show, for every ε > 0,
the existence of some δ > 0 so that
∂f
(x, y) − ∂f (x0 , y) < εg(y)
∂x ∂x
Version 2004–01–15
19 Elements of mathematical analysis
Version 2004–01–15
Elements of mathematical analysis 20
Version 2004–01–15
21 Elements of mathematical analysis
kf k∞ = sup{|f (x)| : x ∈ X}
Version 2004–01–15
Elements of mathematical analysis 22
|f (y) − f (x)| ≤ |f (y) − fn (y)| + |fn (y) − fn (x)| + |fn (x) − f (x)| < 3ε
since |f (y)−fn (y)| ≤ kf −fn k∞ < ε (and similarly, |fn (x)−f (x)| < ε).
Proof: The simple idea of the proof is to use the Banach fixed point
theorem on the function
Z t
Φ(x)(t) = x0 + f (τ, x(τ )) dτ
0
Version 2004–01–15
23 Elements of mathematical analysis
kΦ(x0 ) − x0 k∞ ≤ (1 − LT )r.
Version 2004–01–15
Elements of mathematical analysis 24
kAxk
kAk = sup .
x∈Rn \{0} kxk
(Much of what follows works just as well if the Euclidean spaces Rn are
replaced by Banach spaces, and L(X, Y ) is the space of bounded linear
operators from a Banach space X to a Banach space Y .)
Version 2004–01–15
25 Elements of mathematical analysis
kf (x + ξ) − f (x) − Aξk
lim = 0.
ξ→0 kξk
∂fi
Aij = (x).
∂xj
Show that the converse does not hold, for example by considering the
function x x
p 1 2 x 6= 0,
f (x) = x21 + x22
0 x = 0.
Proof: We prove only the hard part, leaving the rest as an exercise.
What we shall prove is the following: If f has continuous first order
Version 2004–01–15
Elements of mathematical analysis 26
Xn n
X Z 1
= xk gk (0) + xk [gk (γk (t)) − gk (0)] dt
k=1 k=1 0
The first sum on the last line is the desired linear function of x; it is
Df (0)x. To show that this is really the Fréchet derivative, we must
show the second sum is o(kxk). But if ε > 0 is given we can find δ > 0
so that kgj (x) − gj (0)k < ε whenever δ > 0. Clearly, for such an x and
0 ≤ t ≤ 1 we have kγk (t)k ≤ kxk < δ, and so each of the integralsP in
the second sum has norm < ε; hence their sum has norm < ε |xj |
and so it is o(kxk) as kxk → 0.
We leave the rest of the proof as an exercise. To show that f is C 1 ,
note that once we know that f is differentiable, the matrix elements
of Df are the partial derivatives of the components of f (in fact the
columns of Df are the functions gj ).
Version 2004–01–15
27 Elements of mathematical analysis
and the proof is complete. (Exercise: Write the argument out more
carefully, dealing properly with all the epsilons and deltas.)
We can now state and prove the implicit function theorem. First,
however, let us consider the simple case of single variables. Clearly, the
curve in Figure 1 is not the graph of a function. Nevertheless, some
part surrounding the point (x0 , y0 ) is the graph of a function y = g(x).
This illustrates the fact that we can only expect to be able to show the
existence of a function g with F (x, g(x)) = 0 locally, that is, in some
neighbourhood of x0 . The trouble spots seem to be where the curve has
a vertical tangent or, equivalently, a horizontal normal. A normal vector
is given by ∇F = (∂F/∂x, ∂F/∂y), so the trouble spots are recognised
by ∂F/∂y = 0.
We return to the general case of functions of several variables. If
x ∈ Rm and y ∈ Rn , we may write the vector (x1 , . . . , xm , y1 , . . . , yn ) ∈
Rm+n as (x, y). If F is a function of (x, y), we write Dy F (x, y) ∈
L(Rn , Rn ) as Dy F (x, y)η = DF (x, y)(0, η). We then note that the
chain rule applied to the equation F (x, g(x)) = 0 yields Dx F (x, y) +
Dy F (x, g(x))Dg(x) = 0, so if Dy F (x, g(x)) is invertible, we find
Version 2004–01–15
Elements of mathematical analysis 28
Version 2004–01–15
29 Elements of mathematical analysis
Version 2004–01–15
Elements of mathematical analysis 30
where the matrices A and B satisfy inequalities of the form kAk < C
and kI −Bk < 1/2 (the latter comes from the inequality kDy H(x, y)k <
1/2, and the former is just the boundedness of Dx F in a neighbour-
P∞ of 0). But
hood then B is invertible with kB −1 k < 2 (because B −1 =
k
k=0 (I − B) ), and we have
Version 2004–01–15
31 Elements of mathematical analysis
Index
Version information
Content-wise, this document is identical to the version labeled 1999–
04–30. But the source file has been upgraded to newer LATEX standards
in order to allow a LATEX rerun so I can take advantage of Type 1 fonts
and, most importantly, pdfTEX. Line and page breaking has changed a
little bit as a result.
Version 2004–01–15