Paper II Analysis I PDF
Paper II Analysis I PDF
(Mathematics), SEM- I
Paper - II
ANALYSIS – I
PSMT102
Note- There will be some addition to this study material. You should download it again after few
weeks.
CONTENT
1. Differentiation of Functions of
Several Variables
2. Derivatives of Higher Orders
3. Applications of Derivatives
4. Inverse and Implicit Function
Theorems
5. Riemann Integral - I
6. Measure Zero Set
***
SYLLABUS
Unit I. Euclidean space
Cauchy-Schwarz inequality,
properties of the norm function .
(ref. M. Spivak)
1.
2. ,and
3.
1. If is a constant function,then .
(ref. M. Spivak).
(ref. W. Rudin)
5.5 Summary
5.0 OBJECTIVES
After reading this unit you should be able to
5.1 INTRODUCTION
You have seen how to extend the concepts of limit and continuity to functions between
metric spaces. Another important concept is differentiation. If we try to apply this to
functions between metric spaces, we encounter a problem. We realise that apart from the
distance notion, the domain and codomain also need to have an algebraic structure. So, let us
consider Euclidean spaces like Rn, which have which have both metric and algebraic
structures. Functions between two Euclidean spaces are what we call functions of several
variables.
In this chapter we shall introduce the concept of differentiability of a function of several
variables. The extension of this concept from one to several variables was not easy. Many
different approaches were tried before this final one was accepted. The definition may seem a
little difficult in the beginning, but as you will see, it allows us to extend all our knowledge of
derivatives of functions one variable to the several variables case. You may have studied
these concepts in T. Y. So, here we shall try to go a little deeper into these concepts, and deal
with vector functions of several variables.
f ( a h) f ( a )
We say that f is differentiable at a R, if the limit, lim exists.
h 0 h
f ( a h) f ( a )
In that case, we say that the derivative of f at a, f1(a) = lim .......(5.1)
h 0 h
So, we take the limit of the ratio of the increment in f(x) to the increment in x. Now, when our
function is defined on Rn, the increment in the independent variable will be a vector. Since
division by a vector is not defined, we cannot write a ratio similar to the one in (5.1). But
(5.1) can be rewritten as
f ( a h) f ( a )
lim [ − f1(a) ] = 0, or
h 0 h
lim [ ] = 0, or
h 0
where the “remainder” r(h) is so small, that tends to zero as h tends to zero.
For a fixed a, f(a), and f1(a) are fixed real numbers. This means, except for the remainder,
r(h), (5.2) expresses f(a + h) as a linear function of h. This also helps us in “linearizing” f. We
say that for points close to a, the graph of the function f can be approximated by a line. Thus,
f1(a) gives rise to a linear function L from R to R.
L: R R, h f1(a).h, which helps us in linearizing the given function f near the given
Definition 5.1 Suppose E is an open set in Rn, f : E Rm, and a E. We say that f is
differentiable at a, if there exists a linear transformation T : Rn Rm, such that
lim =0 .........................(5.4)
h 0
ii) Since E is open, , such that B(a, r) E. We choose h, such that h < r, so that
a+h E.
iii) The norm in the numerator of (5.4) is the norm in Rm, whereas the one in the denominator
is the norm in Rn.
iv) The linear transformation T depends on the point a. So, when we have to deal with more
than one point, we use the notation, Ta, Tb, and so on.
We have seen that in the one variable case, the derivative defines a linear function,
For a given point a, the linear transformation Ta is called the total derivative of f at a, and is
denoted by f1(a) or Df(a). We can then write
Example 5.1 : Consider f: Rn Rn, f(x) = a + x, where a is a fixed vector in Rn . Find the
total derivative of f at a point p Rn, if it exists.
Solution : Now, f(p + h) – f(p) = h. So, if we take T to be the identity transformation from
Rn to Rn, then we get
lim 0.
h 0
Comparing this with 5.5, we conclude that the identity transformation is the total derivative
of f at the point p.
Example 5.2 : Find the total derivative, if it exists, for f : R2 R2, f(x, y) = (x2, y2), at a point
a = (a1, a2).
f(a + h) – f(a) = (
=( )
= (2 )+( )
= +( )
Now that we have defined the total derivative, let us see how many of the results that we
know about derivatives of functions of a single variable, hold for these total derivatives.
Proof : Suppose f has two derivatives, T1 and T2 at a, and let T = T1 – T2. Let h Rn,
Then th 0 as t 0.
Since T1 is a total derivative of f at a,
Thus,
Therefore, +
+ .
Example 5.3 : i) Find the total derivative f1(a), if f : Rn Rm , f(x) = c, where c is a fixed
vector in Rm and a Rn.
lim lim 0.
h 0 h 0
We have defined the total derivative of a function as a linear transformation. Now we prove a
result about linear transformations which we may use later.
p = (p1, p2, ..., pn), and . Suppose {e1, e2, ..., en} is the standard basis for Rn. Choose
M, where M = ....... + .
If x = (x1, x2, ..., xn) is such that < , then |xi – pi| < for i = 1, 2, ..., n.
Also, < =
| | +| | + ....... +| |
< ( +.......+ )
For functions of a single variable, we know that differentiability implies continuity. The next
theorem shows that this holds for functions of several variables too.
lim 0.
h 0
<
Choose Then
<( )
By Proposition 5.1, Tp is continuous at 0, and Tp(0) = 0. So, there exists such that
< .
With your knowledge of functions of one variable, you would expect that the converse of
Theorem 5.2 does not hold. That is, continuity does not imply differentiability. The following
example shows that it is indeed so.
Example 5.4 : Consider the function f : R R2, f(x) = (|x|, |x|). We shall show that f is
continuous at 0, but is not differentiable there.
|x| < .
Hence, f is continuous at x = 0.
Now, (1, 1) and ( −1, −1) are two distinct points in R2, and B((1, 1), 1) B((−1, −1), 1) = .
. .............................(5.8)
Similarly, taking h = − , we get that T(1) B((−1, −1), 1). But this contradicts the fact
that B((1, 1), 1) and B(( −1, − 1), 1) are disjoint.
i = 1, 2, ..., m. These fis are called coordinate functions of f. Similarly, a linear transformation
Theorem 5.3 : Let f = (f1,f2, ...,fm) : Rn Rm, and p Rn. f is differentiable at p, if and only
if each fi, 1 m is differentiable at p.
lim = 0, where {e1, e2, ..., em} is the standard basis of Rm,
h 0
0 lim
h 0
= lim
h 0
lim + lim = 0.
h 0 h 0
We know that the derivative of a function of one variable denotes the rate at which the
function value changes with change in the domain variable. In the case of functions of several
variables, change in the domain vector variable means a change in any or all of its
components. But if we consider change in only one component and study the rate at which
the function value changes, we get what is known as the partial derivative of the function.
Corresponding to each component of the variable, there will be a partial derivative. Here is
the formal definition.
Definition 5.2 Let f : E Rm, where E Rn. Let x = (x1, x2, ..., xn) be an interior point of E.
Then for every i, i = 1, 2, ..., n, the limit
Remark 5.2 : i) If a function f has partial derivatives at every point of the set E, we say that f
has partial derivatives on E.
ii) It is clear from the definition that a partial derivative can be defined at an interior point of
E, and not on its boundary.
iii) If a function has a partial derivative at a point, its value depends on the values of the
function in a neighbourhood of that point. So, if the function values outside this
neighbourhood are changed, it does not affect the value of the partial derivative.
Example 5.5 : Find the partial derivative of the function, f(x, y, z) = xyz + x2z.
Solution : This is a real-valued function. You are already familiar with the partial
differentiation of such a function.
Example 5.6 : Find the partial derivatives of the function, f : R3 R2, f(x, y, z) = (xy, z2), if
they exist.
You must have observed that the partial derivatives of a vector function are formed by taking
the partial derivatives of its coordinate functions. In fact we have the following theorem,
which establishes the connection between differentiability of a vector-valued function and the
existence of partial derivatives of its coordinate functions
Theorem 5.5 : Let E be an open subset of Rn, and f : E Rm. Suppose f = (f1,f2, ...,fm) is
differentiable at p E. Then the partial derivatives exist for i = 1, 2, ..., m, j = 1, 2, ..., n.
That is,
= T( ).
Hence the limits exist, and (p) exists for all i = 1, 2, ..., m.
Since j was arbitrary, we conclude that (p) exists for all i = 1, 2, ..., m, j = 1, 2, ..., n.
If f : E Rm, where E is an open subset of Rn, and if f is differentiable at p E , then using
Theorem 5.5, the matrix of the linear transformation T can be written as
This m n matrix is called the Jacobian matrix of f at p, and is denoted by [f’(p)] or [Df(p)].
If m = n, the determinant of the Jacobian matrix is called the Jacobian of f at p, and is denoted
by .
T(x) = [f’(p)] .
Tp(h) = [ ] .
So, we can say that the total derivative Tp of a real-valued function is given by
Tp (h) = f(p) h.
Definition 5.3 : Let f : E R, where E is an open subset of Rn. Let u be a unit vector in Rn,
and p E. If lim exists, then it is called the directional derivative of f at p in
t 0
Solution : i) The unit vector u in the given direction is ( ). Hence the required
= lim
t 0
= lim = lim =5 .
t 0 t 0
Example 5.9 : Find the directional derivatives, if they exist, in the following cases:
does not exist. If either u1 or u2 is zero, we get the standard basis vectors, (1, 0) and (0, 1).
Thus, the directional derivatives in these two directions exist, and are equal to one. In any
other direction, the derivative does not exist. Note that the directional derivative in the
direction (1, 0) is fx, and that in the direction (0, 1) is fy. Thus, this function has both the
partial derivatives at (0, 0).
ii) == == = 1/ .
Thus, Duf(0, 0) = 1/ .
In fact, if we take u = (cos , sin ), then we can show that f has directional derivative at (0, 0)
in the direction of u, whatever be . That is, the directional derivatives of f at (0, 0) exist in
all directions. But you can easily show that this function is not continuous at (0, 0) by using
the two-path test. Recall, that you need to show that the limits of f, at (0, 0) along two
different paths are different. Then by Theorem 5.2 we can conclude that f is not
differentiable at (0, 0).
This example shows that the existence of all directional derivatives at a point does not
guarantee differentiability there. But we have the following theorem:
Let u be any unit vector in Rn, and take h = tu. Then h 0, as t 0. Therefore,
lim . This means,
t 0
lim . That is,
t 0
Since u was an arbitrary unit vector, we conclude that the directional derivatives of f at p
exist in all directions.
Now, if u = (u1, u2, ..., un), T(u) = T( u1e1 + u2e2 + ... + unen), where {e1, e2, ..., en} is the
standard basis of Rn. Therefore, by (5.5),
T(u) = u1T(e1) + u2T(e2) + ... + unT(en)
= u1 + u2 + ... + un
= f(p) u
(5.6 ) gives an easy way to find a directional derivative of a differentiable function, if its
partial derivatives are known. For example, if f(x, y) = x2 + y2, then fx and fy at (1, 2) are 2
and 4, respectively. So, the directional derivative of f at (1, 2) in the direction 2i – 3j is given
by (2i + 4j) = .
Theorem 5.8 : Let E be an open subset of Rn, and f : E Rm, f = (f1,f2, ...,fm). If all the
partial derivatives, Djfi(x) of all the coordinate functions of f exist in an open set containing
a, and if each function Djfi is continuous at a, then f is differentiable at a.
Proof : In the light of Theorem 5.3, it is enough to prove this theorem for the case m = 1. So,
we consider a scalar function f from Rn to R, all whose partial derivatives Djf are continuous
at a. Since E is open, for a given > 0, we can find r > 0, such that the open ball,
B(a, r) , and || x – a || < r | Djf(x) Djf(a) | < /n, for j = 1, 2, ... , n. ..................(5.7)
Now, suppose h = (h1, h2, ... , hn), ||h|| < r. Let v0 = 0, v1 = h1e1, v2 = v1 + h2e2, ... ,
f(a + vj) – f(a + vj − 1) = hjDjf(a + vj – 1 + hjej) , for some (0, 1). Then, using (5.7), we
can write
Thus, f is differentiable at a.
Definition 5.4 : A function f : E Rm, f = (f1,f2, ...,fm), where E is an open subset of Rn,
is said to be continuously differentiable, or, a C1 function, if Djfi is continuous on E for
all j, j = 1, 2, ..., n, and for all i, i = 1, 2, ..., m.
f(x, y) =
Exercises:
5.5 SUMMARY
In this unit we have extended the concept of differentiation from functions of one variable
to functions of several variables. Apart from the total derivatives, we have also defined
partial derivatives, and directional derivatives. We have proved that differentiability
implies the existence of all partial and directional derivatives at a point, but the converse
is not true. As in the case of functions of one variable, we prove that differentiable
functions are continuous, but not vice versa. We have also derived a sufficient condition
for differentiability in terms of the partial derivatives.
6
Unit Structure
6.0 Objectives
6.1 Introduction
6.5 Summary
6.0 OBJECTIVES
After reading this chapter, you should be able to
6.1 INTRODUCTION
In the last chapter you have seen how functions of several variables are differentiated. Now
we shall start by discussing how a composite function of two differentiable functions can be
differentiated. The Jacobian matrix introduced in the last chapter proves useful in this.
As we have already mentioned in Chapter 5, this m n matrix, called the Jacobian matrix, is
denoted by [Df(p)]. The kth row of this matrix is the gradient vector, fk(p), and the jth
column is the image of ej under the linear transformation Djf(p).
Thus, the Jacobian matrix of f is formed by all first order partial derivatives of f. This means,
we can write the Jacobian matrix of any function, all of whose partial derivatives exist. As we
have noted earlier, the existence of partial derivatives does not guarantee differentiability. So,
even when a function is not differentiable we would be able to write its Jacobian matrix,
provided all its partial derivatives exist.
|| (p)(h)|| = || || || = |,
since
|| ej || = 1, 1 n.
Cauchy-Schwartz inequality for inner products says that | u v | || u || || v||. Using this we
get || (p)(h)|| = || h || .
If we take M = , then
Theorem 6.1 (Chain Rule) : Let f and g be two differentiable functions, such that the
composite function f g is defined in a neighbourhood of a point a Rn. Suppose g is
differentiable at a, g(a) = p, and f is differentiable at p. Then f g is differentiable at a, and
Proof : If h is such that || h || is small, then a + h will belong to the above neighbourhood of
a, in which f g is defined. Now, since g is differentiable at a,
where Ea(h) 0, as h 0.
f(g(a + h)) – f(g(a)) = (g(a)) (a)(h) + || h ||[ (g(a)) Ea(h) + Ep(k)], if h 0. ..(6.3)
To complete the proof we need to show that the vector in the square brackets in (6.3) tends to
zero, as h tends to zero.
0, as h 0, since h 0 . ....(**)
Using (*) and (**), we can say that the term in the square brackets in (6.3) tends to zero as
h 0. Therefore,
–
0 as h 0.
Here the product on the right hand side is matrix multiplication. If y = g(x), and z = f(y),
comparing the entries in the matrices in (6.3), we get
Example 6.1 : Write the matrices for , and for the following functions, and
evaluate them at the point (2, 5). f(x, y) = (x + y, x + y , 2x + 3y), g(u, v) = (x, y) = (u2, v3).
2 2
D )= .
You can now easily verify that D(f g) (2, 5) = [D(f(4, 125)] [D(g(2, 5))].
Example 6.2 : Find partial derivatives of all possible orders for the function,
f(x, y, z) = (x2y2, 3xy3z, xz3).
Solution : Since f is a polynomial function, we do not have to worry about the existence of
partial derivatives. We get
Then, fxx = (2y2, 0, 0), fxy = = = (4xy, 9y2z, 0), fxz = (0, 3y3, 3z2).
Differentiating fy, we get fyx = (4xy, 9y2, 0), fyy = (2x2, 18xyz, 0), and fyz = (0, 9xy2, 0).
Then differentiating fz we get fzx = (0, 3y3, 3z2), fzy = (0, 9xy2, 0), and fzz = (0, 0, 6xz).
These are all possible second order derivatives of f. Proceeding in this way, we can also get
fxyz = (0, 9y2, 0), fyxz = (0, 0, 0), fzzz = (0, 0, 6x), and so on. There will be 27 third order
partial derivatives of f. See if you can get the remaining.
You know that fxy and fyx differ in the order in which f is differentiated with respect to the
variables x and y. These two derivatives have come out to be equal in Example 6.2. But you
may have seen examples of scalar functions of several variables, for which the two may not
be the same. Here is an example, to jog your memory.
Example 6.3 : Consider this function f from R2 to R, f(x, y) = for (x, y) (0, 0),
and f(0, 0) = 0. You can easily check that
fy(h, 0) = lim = h.
k 0
Thus, the mixed partial derivatives of this function both exist, but are not equal.
Remark 6.1 : If f is a function from Rn to R, the partial derivative of f with respect to the ith
variable, xi, is denoted by Dif, and the partial derivative of Dif with respect to xj , that is,
Dj(Dif) is denoted by Djif.
The following theorem gives a sufficient condition for the two mixed partial derivatives of a
function to be equal. Since the behaviour of a vector-valued function is decided by the
behaviour of its coordinate functions, it is enough to derive this sufficient condition for a
scalar function. Without loss of generality, we state the theorem for a function of two
variables.
Theorem 6.2 : Let f : R2 R, such that the partial derivatives, D1f, D2f, D12f and D21f exist
on an open set S in R2. If (a, b) S, and D12f and D21f are both continuous at (a, b), then
D12f(a, b) = D21f(a, b).
Proof : We choose positive real numbers, h and k, which are small enough so that the
rectangle with vertices (a, b), (a + h, b), (a, b + k), (a + h, b + k) lies within S.
Now we can write (h, k) = G(a + h) – G(a). Since G is defined in terms of f, and since f has
all the necessary properties, G is continuous on [a, a + h], and is differentiable in (a, a + h).
So, we apply the Mean Value Theorem for functions of a single variable to G, and get
G(a + h) – G(a) = h (c), for some c (a, a + h). Now (x) = D1f(x, b + k) – D1f(x, b). So,
we write (h, k) = G(a + h) – G(a) = h[D1f(c, b + k) – D1f(c, b)].
Now D1f (c, y) is a differentiable function of one variable with derivative equal to D 21f. So
applying MVT to D1f(c, y) on the interval [b, b + k], we get
We now write (h, k) = [f(a + h, b + k) – f(a, b + k)] – [f(a + h, b) – f(a, b)], and define
H(y) = f(a + h, y) – f(a, y), so that (h, k) = H(b + k) – H(b). Using the same arguments
that we used for G, we apply MVT to H, and then to D2f(x, p), we get
From (6.4) and (6.5) we get D21f(c, d) = D12f(q, p). Since D12f and D21f are continuous, taking
the limit as (h, k) (0,0), we get D12f(a, b) = D21f(a, b).
As we have mentioned earlier, the conditions of this theorem are sufficient, and not
necessary. In fact, the continuity of just one of the mixed partial derivatives is also sufficient
to guarantee equality. Functions whose partial derivatives are continuous play an important
role in Calculus. We classify these functions as follows:
In the next chapter we shall see that a Ck function, that is a function, all whose partial
derivatives of order up to k are continuous, can be approximated by means of a polynomial of
order k. We shall also discuss the technique to find the maximum and minimum values of a
function belonging to class C’’.
The Mean Value Theorem (MVT) is an important theorem in Calculus. It is used as a tool to
derive many other results. In the last section we have used it in the proof of Theorem 6.2. In
this section we shall see if it also holds good for functions of several variables. But first, let
us recall the one-variable case.
MVT (single variable): If f : [a, b] R is continuous on [a, b], and differentiable on (a, b),
then there exists c (a, b), such that
f(a + h) – f(a) = h .
Example 6.4 : Consider f : [0, 2 ] R2, f(t) = (cost, sint). This function is continuous on
[0, 2 ] and differentiable on (0, 2 ). Now, f(2 ) – f(0) = (1, 0) – (1, 0) = (0, 0).
(t) = ( − sint, cost). For the extension of MVT to hold, we must have
f(2 ) – f(0) = 2 (c) for some c in (0, 2 ). So, we should have (0, 0) = 2 ( sinc, cosc).
But this is impossible, since sinc and cosc both cannot be zero.
So, the extension of MVT in its stated form does not hold. But there is a way around this
difficulty. A slightly modified version of MVT does hold true for all functions of several
variables. We now state and prove this modified theorem for functions from Rn to Rm. As a
special case of this theorem you will realize that MVT holds for real-valued functions of
several variables.
Theorem 6.3 : (Mean Value Theorem) Let f : S Rm, where S is an open subset of Rn.
Suppose f is differentiable on S. Let x and y be two points in S, such that the line segment
joining x and y, L(x, y) = {tx + (1 t)y | 0 1}, also lies in S. Then for every a Rm,
there is a point z S, such that
a {f(y) – f(x)} = a { (z)(y x)} ................................(6.6)
Before we start the proof, let us understand the geometry involved. Let u = y – x. Then x + tu
gives us a point on the line segment L(x, y), if 0 1. Since S is open, we can find a
> 0, such that S, and S. See Fig. 6.1, in which we show the situation
when n = 2. The point p is on the extension of L(x, y) and is equal to x + (1 + )u. Similarly
the point q is also on the extension of L(x, y), and is equal to x – u for some > 0.
p p
Y
S
X
Figure 6.1
Thus we get a > 0, such that x + tu S for every t . Now we start the formal
proof.
Thus, we can apply MVT for functions of a single variable, and get
Remark 6.2 : i) (6.6) is true for all x, y in S, such that the line segment joining x and y is also
in S. This means, if S is a convex open set in Rn, then (6.6) will be true for all x, y in S.
Theorem 6.4 : Let f : S Rm, where S is an open connected subset of Rn. Suppose f is
differentiable on S, and = 0 for every p S. Then f is a constant function on S.
Proof : The set S is polygonally connected, since it is open and connected. Let x and y be
two points in S. Then x and y are joined by line segments L1, L2, L3, ... , Lr, lying entirely in
S. Suppose Li is a line segment joining pi and pi+1, 1 r, p1 = x, and pr+1 = y.
a {f(pi+1) – f(pi)} = a , zi Li
= 0, since = 0.
This means,
(6.8) is true for every a in Rm. So, in particular, it is true for f(y) – f(x). Thus,
Since x and y were any arbitrary points in S, we have thus proved that f is a constant function
on S.
Exercises :
1) Find the partial derivatives, D1f, D2f, D12f and D21f at (0, 0) , if they exist, for the
following function f from R2 to R.
f(x, y) = y , if (x, y) (0, 0), and f(0, 0) = 0.
2) If u(x, y) = x +y2, x(t) = 3t2 + 4, and y(t) = sin2t, find (t) and (t).
3) If u(x, y) = x – 2y + 3, x = r + s + t, y = rs + t2, find ur, us and ut at (1, 2, 4).
4) Let f : R2 R2, and g : R3 R2 be two vector functions, defined as:
f(x, y) = (sin(2x + y), cos(x + 2y)),
g(r, s, t) = (2r – s – 3t, r2 – 3st).
i) Write the Jacobian matrices for f and g. If h is the composite function, f g,
compute the Jacobian matrix of h at the point (1, 0, - 2).
5) If f is a function from R2 to R, and D1f = 0 at all points, show that f is independent of
the first variable. If D1f = D2f = 0 at all points, show that f is a constant function.
6.5 SUMMARY
In this chapter we have derived the chain rule for differentiation of composite of two
functions. We have also seen that the Jacobian matrix for the composite function is the
product of the Jacobian matrices of the two given functions. We have defined higher order
partial derivatives of functions of several variables. We have seen functions, whose second
order mixed partial derivatives depend on the order of the variables with respect to which the
function is differentiated. On the other hand, we have derived sufficient conditions for such
mixed partial derivatives to be equal. Finally, through an example we have seen that the
Mean Value Theorem cannot be extended to all vector functions. We have proved a restricted
form of the MVT for vector functions. Of course, MVT does extend to scalar-valued
functions of several variables. As a result of MVT we have proved that a function defined on
an open connected set is constant, if its derivative is uniformly zero over its domain.
7
APPLICATIONS OF DERIVATIVES
Unit Structure
7.0 Objectives
7.1 Introduction
7.5 Summary
7.0 OBJECTIVES
After reading this chapter, you should be able to
7.1 INTRODUCTION
In the two previous chapters we have discussed differentiation of scalar and vector functions
of several variables. Now we shall tell you about some applications of derivatives. In your
study of functions of one variable you have seen that a major application of the concept of
derivatives is the location of maxima and minima of a function. This knowledge is very
crucial for curve tracing. Here we shall see how the derivatives help us in locating the
extreme values of a real-valued function of several variables. But before we do that, we are
going to discuss Taylor’s theorem and Taylor’s expansions, which help us approximate a
function with the help of polynomials. This knowledge will help us derive some tests for
locating and classifying the extreme points of a function.
Theorem 7.2 (Taylor’s theorem for real functions of one variable): Let f be a real-valued
function defined on the open interval (p, q). Suppose f has derivatives of all orders up to and
including n +1 in (p, q). Let a be any point in (p, q). Then for any x (p, q),
xa ( x a) 2 ( x a) n ( x a) n 1
f(x) = f(a) + (a) + (a) + ... + (a) + (c),...(7.1)
1! 2! n! (n 1)!
Proof: We now define a new function g on [a, x], or [x, a], according as a < x, or x < a, by
( x y) ( x y) 2 ( x y) n
g(y) = f(y) + (y) + (y) + ... + (y) + A, ....(7.2)
1! 2! n!
where A is a constant, chosen so as to satisfy g(x) = g(a). We can easily write the expression
for A by using this condition. We leave this to you as an exercise. See Exercise 1).
Using the properties of f, we can see that g satisfies all the conditions of Rolle’s theorem on
its domain. Thus, we can conclude that there exists a point c (a, x), (or (x, a)) such that
(c) = 0. Now, differentiating (7.2), we see that
( x y) 2 ( x y ) ( n 1)
(y) = (y) − (y) + (x − y) (y) − (x − y) (y) + (y) − ... −
2! (n 1)!
( x y) n
(y) + (y) – (n + 1) A.
n!
f ( n1) ( y )
= [ (n + 1)A].
n!
f ( n1) (c)
Hence, (c) = [ − (n + 1)A] = 0.
n!
f ( n1) (c)
This means that A =
n!
f(x) = g(x) =
xa ( x a) 2 ( x a) n ( x a) n 1
g(a) = f(a)+ (a) + (a) + ... + (a) + (c),
1! 2! n! (n 1)!
xa ( x a) 2 ( x a) n ( x a) n 1
f(x) = f(a) + (a) + (a) + ... + (a) + (c),...(7.3)
1! 2! n! (n 1)!
The infinite series in (7.3) is convergent under the given conditions, and is called the Taylor
series of f about a.
xa ( x a) 2 ( x a) n
Pn(x) = f(a) + (a) + (a) + ... + (a) is called the nth Taylor
1! 2! n!
( x a) n 1
polynomial of f about a, and Rn(x) = (c), is called the remainder.
(n 1)!
We now state Taylor’s theorem for functions of two variables, and then find Taylor
expansions of some functions.
Theorem 7.3 (Taylor’s theorem for f: R2 R): Let f be a real-valued Cn+1 function on an
open convex set E R2. Let (a, b) E. Then for any (x, y) E,
where h = x – a, k = y – b, and (c, d) is some point on the line segment joining (a, b)
We are not going to prove this theorem. But, note the following points:
1. Recall that f is Cn+1 means f has continuous partial derivatives of all orders n + 1.
This ensures that all the relevant mixed partial derivatives are equal.
2. E is convex. This guarantees that the line segment joining any two points of E, lies in
E, the domain of f.
Example 7.1: Find the Taylor expansions of the following functions about the given points
up to the third order.
fx = 3x2 + 2y2 – 3y + 4, fy = 4xy – 3x, fxx = 6x, fxy = 4y – 3, fyy = 4x, fxxx = 6, fxxy = 0,
fxyy = 4, fyyy = 0, and all higher partial derivatives are zero. Calculating all these partial
derivatives at (1, 2), we write
Now, R3 involves all fourth order derivatives, and therefore is zero. Hence,
R3 = (h )4sin(2c + 3d), where (c, d) is some point on the line segment joining (0, 0)
and (h, k).
We are now going to state Taylor’s theorem for real-valued functions of n variables. For this,
let us first take a close look at the Taylor expansion of a function of two variables.
If we take the variables to x1, x2, instead of x and y, take (a, b) to be (a1, a2), and (h, k) to be
f(a1 + h, a2 + h2) = f(a1, a2) + ( )f(a1, a2) + f(a1, a2) + ...
= )f(a1, a2) .
Similarly,
+ D222f(a1, a2)h23
= )f(a1, a2) .
You must have noticed that we have added the mixed partial derivative terms, for example,
D12f and D21f, or D112f , D121f, and D211f. We could do this, since f ensures that that
these partial derivatives are equal. Now we state Taylor’s theorem for real-valued functions
of several variables.
Theorem 7.4 : Let f : E R, where E is a convex open subset of Rn. Further, let
a = (a1, a2, ..., an) E, h = (h1, h2, ..., hn) Rn, such that a + h D. If f Cm, then
where ... , take values from the set {1, 2, ..., n}, and the inner summation in (7.5) is
taken over all possible such k-tuples.
Further, the remainder Rm-1(c) = f(c) . This sum is taken over all
possible m-tuples (i1, i2, ..., im), where i1, i2, ..., im take values from {1, 2, ..., n},and c is some
point on the line segment joining a and a + h.
This theorem is used to approximate a given function by a polynomial. In the next section we
shall use it to derive conditions for locating and classifying extreme points of a function.
A local minimum (or relative minimum) is defined in a similar manner. You will agree that
the function f : R5 R, f(x1, x2, x3, x4, x5) = x12 + x22 + x32 + x42 + x52, clearly has a local
minimum at (0, 0, 0, 0, 0). Can you find an example of a function with a local maximum?
Definition 7.2 : A point a Rn is called a saddle point of a function f : Rn R, if every ball
B(a, r), r > 0, contains points x, such that f(x) f(a), and also other points y, such that f(y)
f(a).
In general, it is not easy to spot the local maximum or local minimum merely by observation.
For differentiable functions we can derive tests to locate these values. You know that in the
case of a differentiable function of a single variable, the derivative vanishes at an extreme
point. We have a very similar test for the location of extreme points of a function of n
variables, as you can see in the next theorem.
Proof : Since f has a local maximum at a, r > 0, such that x B(a, r) f(x) f(a).
gi(x) = f(a1, a2, ..., ai – 1, x, ai+1, ..., an). Since f(a) is the local maximum value of f, gi(ai) is the
maximum value of gi. If (a) exists, then (ai) also exists, and the two are equal. By
applying the first derivative test for functions of one variable to gi, we get
(a) = (ai) = 0.
An exactly similar proof will help us conclude that (a), if it exists, is equal to zero, even
when a is a local minimum of f.
Thus, if f has a local extremum at a, and all the partial derivatives exist at a, then f(a) = 0.
As in the case of functions of one variable, the condition in theorem 7.5 is a necessary one,
and is not sufficient. That is, if all the partial derivatives of a function at a point a are zero,
we cannot say that a is a local maximum or local minimum point. It may be neither.
An example is the function f : R2 R, f(x, y) = 1 – x2 + y2. Here fx = - 2x, and fy = 2y. So,
fx(0, 0) = 0 and fy(0, 0) = 0. But you can see clearly, that f has a maximum in the direction of
the x-axis, and a minimum in the direction of the y-axis at (0, 0). This means, f has neither a
minimum, nor a maximum at (0, 0). In fact (0, 0) is a saddle point for this function.
Theorem 7.5, tells us to look for extreme points among the critical points of a function. We
shall now see how to classify these points as local maxima, local minima, or saddle points.
This involves second order partial derivatives. This is to be expected, since in one variable
functions too, we have a second derivative test to classify stationary points. The proof of the
test for several variables involves quadratic forms. You have studied them in T. Y. B. A. /B.
Sc. We start with a definition and recall the relevant results.
Definition 7.4 : If A = (aij) is a real symmetric n x n matrix, and x = (x1, x2, ..., xn) Rn,
then Q(x) = is called a quadratic form associated with A.
It may not be very easy to get the eigen values. But we have an easier way to decide.
A principal minor of a square matrix, A, is the determinant of the matrix obtained by taking
the first k rows, and the first k columns of A, 1 n.
If all the principal minors are positive, then the associated quadratic form is positive definite.
If the principal minors are alternately positive and negative, starting with a negative minor for
k = 1, then the associated quadratic form is negative definite.
If a principal minor of order k is negative, when k is an even number, then Q(x) takes both
positive and negative values.
We now use these facts about quadratic forms to derive the second derivative test. A
definition first.
A = H(x) = .
If a Rn , the first order Taylor formula for f about a gives us the value of f(a + h) for small
values of ||h|| as
||h||2 |E(a, )| = |
Theorem 7.6 : If f is a function from Rn to R, and has continuous second order partial
derivatives in a ball B(a; r) around a stationary point a of f, then
i) This value will be positive for all h, if H(a) is positive definite. Hence,
f(a + h) – f(a) > 0 for all h, such that 0 < ||h|| < r. This tells us that f(a + h) f(a) for
every h B(a; r), that is , a is a relative minimum point of f.
The argument for proving ii) and iii) are exactly similar, and we are sure you can write those.
.Remark 7.2 : i) If an even principal minor, that is a principal minor of even order is
negative, then the point is a saddle point.
Go through the following examples carefully, they illustrate our discussion here.
Example 7.2: Locate and classify the stationary points of the functions given by
H((1, 1)) = . The principal minors are 6, and 27. Both are positive, and hence f
has a local minimum at (1, 1).
fx = 0 xy – y + 1 = 0, and fy = 0 x(x - 1) = 0 x = 0, or x = 1.
Therefore, H((0, 1)) = . det(H(0, 1)) = - 1 < 0. Hence, (0, 1) is a saddle point.
Example 7.3 : Locate and classify the stationary points of f(x, y, z) = i) xyz ,
We have indicated the procedure. We are sure now you will be able to get fxz, fyy, and fzz.
Evaluating these second order partial derivatives at the stationary points, we find,
You will find that (0, 0, 0) is a degenerate stationary point, and (2 , 2, − 2) is a saddle point.
iii) fx = 2x – y, fy = - x + z3, fz = 3yz2 – 6. Equating these to zero, we get (1, 2, 1) as the
stationary point. Check that H((1, 2, 1)) = , and the principal minors are 2, -
Exercises:
x x
1) Find the stationary points of f(x, y) = i) ii) (x + y)exy.
x y2 4
2
x y2 4
2
Look at these situations: i) A rectangular cardboard sheet is given. We have to make a closed
box out of it. What is the maximum volume that is possible?
ii) Temperature varies on a metal surface according to some formula. Where do the
maximum and minimum temperature occur on the surface?
In both these problems we have to maximize or minimize a certain function: volume in the
first case, and temperature in the second. So these are max-min. Problems. But there is a
difference between these and the problems considered in the last section. Here, an additional
constraint or condition is imposed. The given cardboard sheet has a fixed area. The
maximum/minimum temperature points are to be on the given surface.
In this section we shall see how such problems are solved. A very useful method was
developed by Joseph Louis Lagrange. This method gives a necessary condition for the
extreme points of a function. We now state the theorem and then illustrate its use through
some examples.
Theorem 7.7 : Let f : Rn R, and f C1. Suppose g1, g2, . . ., gm (m < n) are functions
belonging to C1, which vanish on an open set E in Rn. If a E is an extreme point of f, and if
(a), (a), . . . , (a) are independent vectors in Rn, then there exist real numbers, ,
, . . . , , such that
Dif(a) + Dig1(a) + Dig2(a) + . . . + Digm(a) = 0, i = 1, 2, . . . , n.
When we want to find the extreme values of a function f : Rn R, f C1, subject to some
constraints, g1(x1, x2, . . . ,xn) = 0, g2(x1, x2, . . . ,xn) = 0, . . . , gm(x1, x2, . . . ,xn) = 0, where
m < n, we set up the n equations
These n equations, along with the m equations, g1(x1, x2, . . . ,xn) = 0, g2(x1, x2, . . . ,xn) = 0, .
. . , gm(x1, x2, . . . ,xn) = 0, are then solved to get the values of the n + m unknowns, x1, x2, . . .
,xn, , , . . . , . The solutions x = (x1, x2, . . . ,xn) are the stationary points, and contain the
extreme points of f .
, ,..., are called Lagrange’s Multipliers. We use one multiplier for each
constraint.
To analytically classify these stationary points into local maximum, minimum, or saddle, is a
very complicated process. It is usually easier to look at the physical or geometrical aspect of
the problem to arrive at any conclusion. We now solve a few problems, so that the entire
process is clear to you.
Example 7.4 : Find the dimensions of the box with maximum volume that can be made with
a cardboard sheet of size 12 cm2.
Solution : If the dimensions of the box are x, y, z cms, then its volume V = xyz c. cms. And
surface area is 2(xy + yz + xz) sq. cms. Here we have to maximize V, subject to a constraint
2(xy + yz + xz) = 12, or (xy + yz + xz) = 6. So, f(x, y, z) = xyz, and
g(x, y, z) = xy + yz + xz – 6. Hence,
f(x, y, z) + g(x, y, z) = 0
f x + gx = 0 yz + (y + z) = 0, fy + gy = 0 xz + (x + z) = 0, f z + gz = 0 xy +
(x + y) = 0.
xyz = (xy + xz) = (xy + yz) = (xz + yz). If = 0, then V = 0, which is the minimum
volume. If 0, then xy + xz = xy + yz = xz + yz. That is, x = y = z (unless, of course, x =
y = z = 0).
Example 7.5 : Find the extreme values of the function given by f(x, y, z) = 2x + y + 3z,
subject to x2 + y2 = 2, x +z = 5.
Solution : Let g1(x, y, z) = x2 + y2 – 2 = 0, and g2(x, y, z) = x + z – 5 = 0. Then
=0
Example 7.6 : Find the minimum distance of a point on the intersection of the planes,
g2(x, y, z) = x + 3y + z – 2 = 0.
=0
fx + g1x + g2x = 0 2x + + =0
fy + g1y + g2y = 0 2y + +3 =0
(0, 1/2, 1/2). The distance of this point from the origin is .
Geometrically, the constraints are equations of two planes. There is no maximum to the
distance of a point on their line of intersection from the origin. So, the stationary point is a
minimum point.
x2 y2
1) Find the extreme values of the function f(x, y) = xy on the surface = 1.
8 2
x y
2) Find the extreme values of z = on the unit circle in the xy-plane.
2 3
3) Find the distance of the point (10, 1, − 6) from the intersection of the planes,
x + y + 2z = 5 and 2x – 3y + z = 12.
7.5 SUMMARY
In this chapter we have introduced Taylor’s theorem for functions of several variables. We
have also seen how to get Taylor polynomials of a given order for a given function. Of
course, to be able to do this, the function must have continuous partial derivatives of higher
orders.
We have then discussed the location of maxima and minima of a real-valued function of
several variables. This has tremendous applications in diverse fields of study. In particular,
we have proved that the extreme points of a function are located among the points at which
the gradient vector of the function is zero. That is, the points at which all the first order partial
derivatives are zero. The classification of these points into maxima, minima, or saddle points
depends on the signs of the principal minors of the Hessian matrix.
We pointed out that there are some situations, where we need to find the extreme values
subject to certain constraints. Such problems, and the method of tackling them is also
discussed, and illustrated through some examples.
8
INVERSE AND IMPLICIT FUNCTION THEOREMS
Unit Structure
8.0 Objectives
8.1 Introduction
8.4 Summary
8.0 OBJECTIVES
After reading this chapter, you should be able to
state and prove Inverse Function Theorem for functions of several variables
check if some simple functions are locally invertible
state and prove Implicit Function Theorem for functions of several variables
8.1 INTRODUCTION
In this chapter we introduce two very important theorems. You have not come across these
theorems even for functions of a single variable. In each case, we shall first discuss the single
variable case, and then extend the concept to functions of several variables. A word of
caution : these theorems are not easy. To help you understand them better, we are going to
prove some smaller results, and then use them in the proof of the theorems. Do study this
chapter carefully and we are sure you would have no difficulty in digesting the concepts.
The inverse function theorem is a very important theorem in Calculus. You may be familiar
with its one dimensional version. Before we introduce the theorem for functions from Rn to
Rn, we shall recall some results about functions of one variable:
1) If f : [a, b] R is continuous, and f(c) > 0 for some c (a, b), then such that
(c ) (a, b), and f(x) > 0 (c ). In other words, we can always find a
neighbourhood of the point c, in which f(x) has the same sign as f(c).
c (a, b), then using 1) we can prove that such that f is an injective function on
(c ) (a, b). Further, f-1: f(c ) (c ) is differentiable at f(c) ,
The statement in 2) is the inverse function theorem. Note that we do not know whether the
inverse of f exists on [a, b]. But what this theorem tells us, is that if , then f is
“locally invertible” at c. For example, we know that the function f : [0, 2 ] R, f(x) = sinx
does not have an inverse. But is a continuous function, and .
So, the theorem says that f is locally invertible at . That is, we can find a neighbourhood
N of , such that f restricted to N has an inverse. Check that f is injective when restricted
to N = ( ), and hence has an inverse on N.
We shall now see if this theorem extends to functions of several variables. Let us start with a
definition.
Definition 8.1 : Let f : E Rn, where E Rn. If f C1, f is said to be locally invertible at
a E, if there exists a neighbourhood N1 of a, N1 E, and a neighbourhood N2 of f(a), such
that f(N1) = N2, f is injective on N1, and f-1 : N2 N1 is a C1 function.
We shall soon state and prove the inverse function theorem. In the proof, we are going to use
some minor results. You have already studied some in the earlier chapters of this course.
Next we state and prove one other result, which will be useful to us.
Theorem 8.1 : Let f = (f1 f2, . . . , fn) : E Rn, where E is an open set in Rn. Suppose f C1.
If the Jacobian of f, J(a) 0 for some a E, then f is injective on a neighbourhood of a in E.
Proof : If X1, X2, . . . , Xn E, we consider a point X = (X1, X2, . . . , Xn) , whose first
n coordinates are the coordinates of X1, the next n are the coordinates of X2, and so on. We
define a function, j, such that
Now, the function j, being an n×n determinant, is a polynomial of its n2 entries, and each
entry, is a continuous function, since f C1. Thus, j is a continuous function on its
domain. We write A = (a, a, . . . , a). Then j(A) = det[Djfi(a)] = J(a) 0. Now, since f C1,
all the entries of j(A) are continuous, and hence, j(A) is also continuous. The continuity of
j(A) ensures that there exists a neighbourhood N of A, such that j(X) 0 , if X N.
Then, using the Mean Value Theorem for scalar fields (See Remark 6.2 ii).), we get
fi(x) − fi(y) = fi(ci) (x − y) fi(ci) (x − y) = 0 for some ci on the line segment joining
x and y. So, if x – y 0, then fi(ci) = 0 for some ci on the line segment joining x and y, that
is, in the neighbourhood Na, since Na is convex. This means, Djfi(ci) = 0 for every j, 1
. Thus, if C = (c1, c2, . . . , cn), then j(C) = det[Djfi(ci)] = 0. But this contradicts
(8.1). So, we conclude that x – y = 0, which proves that f is injective on Na.
Remark 8.1 : i) A function may not be injective on its entire domain. But if its Jacobian is
non-zero at a point, then it is injective on a neighbourhood of that point. In other words, it is
locally injective.
ii) If the Jacobian is non-zero, then the linear transformation Df, which represents the
derivative of f, is non-singular, and hence, is a linear isomorphism.
Example 8.1 : a) Consider the function f(x, y) = (excosy, exsiny). This function is not
injective, since f(x, 0) = f(x, 2 ). But,
Here we have a function, which is locally injective at every point of its domain, but is not
injective on the domain.
b) Consider the function f(x, y) = (x3, y3), defined on R2. The Jacobian of this function is
zero at (0, 0). But the function is locally invertible at (0, 0). In fact, it is an invertible
function.
Theorem 8.2 (The Inverse Function Theorem): Let f = (f1, f2, . . . , fn) C1, f: E Rn , where
E is an open set in Rn. Let T = f(E). Suppose J(a) 0 for some a E. Then there exists a
unique function f-1 from Y to X, where X is open in E, Y is open in T, such that
Also, f(a) f(B(a, r)). Therefore, there exists a > 0, such that B(f(a), ) f(B(a, r)).
Take X = f-1(B(f(a), )), and Y = B(f(a), ). Then X and Y satisfy i), ii), iii) and iv) in the
statement of the theorem.
To prove the last assertion v) in the statement, we have to show that all the partial derivatives
of all the component functions of f-1 are continuous on Y. For this we first define the function
j(X) = det[Djfi(xi)] , as in Theorem 8.1. Here X = (X1, X2, . . . , Xn). Then, as before, there is a
neighbourhood Na of a, such that j(X) 0, whenever each Xi Na. We can assume that the
neighbourhood N Na. This ensures that j(X) 0, whenever each Xi .
1 1
f ( y tei ) f ( y)
Now we first prove that Dif-1 exists on Y. Let y Y, and consider ,
t
where ei is the ith coordinate vector, and t is a scalar. Let x = f-1(y), and = f-1(y + tei). Then
f m ( x ' ) f m ( x) x' x
fm(xm) , m = 1, 2, . . . , n. Here xm is a point on the line
t t
segment joining x and .
So, we get a system of n equations (for the n values of m). The left hand side of an equation
in this system is 1, if m = i, otherwise it is 0. The right hand side is of the form
The determinant of this system of linear equations is j(X), which we know is non-zero. Hence
x 'j x j
we can solve it by Cramer’s rule and get the variables as the quotient of two
t
determinants. Then, as t tends to zero, approaches x, and hence, each xm also approaches x.
The determinant in the denominator, j(X) = det[Djfi(xi)] then approaches J(x), the Jacobian
x 'j x j
of f at x, which is again non-zero. Thus, as t tends to zero, the limit of exists. That
t
1 1
f ( y tei ) f ( y)
is, lim exists. Thus, Dif-1(y) exists for all i, and for all y in Y.
t 0 t
We have obtained the partial derivatives of the components of f-1 as quotients of two
determinants. The entries in these determinants are partial derivatives of the components of f,
which are all continuous. Since a determinant is a polynomial of its entries, we conclude that
the partial derivatives of f-1 are continuous on Y.
Example 8.2 : Show that the function f: R2 R2, f(x, y) = (2xy, x2 – y2) is not invertible on
R2, but is locally invertible at every point of E = {(x, y) | x > 0}. Also find the inverse
function at one such point.
Solution : Here f(1, 1) = f( − 1, − 1) = (2, 0). Therefore f is not injective, and hence is not
invertible on R2. On the other hand, if (x, y) E, then
u u2
Suppose f(x, y) = (u, v). If (x, y) E, then y = , and v = x2 . Therefore,
2x 4x 2
4 2 2 v v2 u2
2 v v 2 u 2 1/2
4x - 4x v – u = 0. Thus, x = , and x = ( ) ,
2 2
y = u(2v + 2 )−1/2
If x2 + y2 = 0, find . You must have done exercises like this in your under-graduate
dy
classes. Here, we take f(x, y) = x2 + y2, and find fx = 2x, and fy = 2y. Then = 2x/2y = x/y.
dx
Of course, y cannot be zero.
While doing this exercise, actually we have used a theorem, the implicit function theorem. To
recall, in this setting, a function which can be written as y = g(x), is called an explicit
function, and one which can be expressed only as f(x, y) = 0, is called an implicit function.
The implicit function tells us that under certain conditions, we can express an implicit
dy
function as an explicit one, and then we can use this expression to find .
dx
In this section we are going to discuss this implicit function theorem for functions of several
variables. Before we state and prove the general case, we first prove the case for functions
involving only two variables, x and y.
Theorem 8.3 : Let f be a real-valued C1 function, defined on the product , where and
are two intervals in R. Let (a, b) , and f(a, b) = 0, but fy(a, b) 0. Then there
exists an interval I in R, containing a, and a C1 function g : I R, such that g(a) = b, and
Thus, h is a C1 function, with a non-zero Jacobian at (a, b). Therefore, by the inverse function
theorem, Theorem 8.2 , we can conclude that h is locally invertible at (a, b). Let u = ( )
be the local inverse of h. You will agree that (x, y) = x for all x and y in R. That is,
u(x, y) = (x, (x, y)) for all x and y in R. We now define g as, g(x) = (x, 0), and show that
it has all the required properties.
Now, since h(a, b) = (a, 0), u(a, 0) = (a, b). This means, (a, 0) = b. Thus, g(a) = b.
Also, (x, 0) = h(u(x, 0)) = h(x, (x, 0)) = h(x, g(x)) = (x, f(x, g(x))). This implies that
f(x, g(x)) = 0.
Since u is a C1 function, g is also C1. Differentiating f(x, g(x)) = 0 with respect to x using
chain rule, we get D1f(x, g(x)) + D2f(x, g(x)) (x) = 0, and thus,
D1 ( f ( x,g ( x))
(x) = , since D2f(x, g(x)) 0.
D2 f ( x, g ( x))
Basically, this theorem tells us that under certain conditions, the relation f(x, y) = 0, between
x and y can be explicitly written as y = g(x).
Remark 8.2 : If instead of fy(a, b) 0, we take the condition fx(a, b) 0, then we can
express x as an explicit function of y.
Solution : Note that f(1,1) = 0, and fy = 3y2 – 2x = 1 at (1, 1). Further, f is a C1 function on R2.
Therefore, we can apply Theorem 8.3, and conclude that there exists a unique function g,
3x 2 2 y
defined on a neighbourhood of 1, such that g(1) = 1. Also, (x) = in this
3 y 2 2 x
neighbourhood.
Example 8.4 : Check whether Theorem 8.3 can be applied at all points, where
f(x, y) = x2 – y2 = 0.
Solution : x2 – y2 = 0 is true at points (0, 0), (1, 1),(1, −1), ( −1, 1), and ( −1, −1). fy = −2y,
and fx = 2x. At the point (0, 0), fx and fy are both zero, and hence we cannot apply the
theorem. At all the remaining points, the function satisfies all the conditions of Theorem 8.3,
and hence it can be applied. You will agree that at each of these points, we will get either
g(x) = x, or g(x) = − x.
Theorem 8.4 : Let f be a real-valued C1 function, defined on an open set, U, in Rn. Let
a = (a1, a2, ... , an-1) Rn-1, such that (a, b) U, f(a, b) = 0, and Dnf(a, b) 0. Then there
exists a unique C1 function g, defined on a neighbourhood N of a, such that g(a) = b, and
Proof : We consider a function h : U Rn−1 R, defined by h(x, y) = (x, f(x, y)). If we write
h = (h1, h2, ... , hn), then hi(x, y) = xi, for 1 i n – 1, and hn(x, y) = f(x, y). Therefore, the
Jacobian matrix of h is given by
Jh = .
The determinant of this matrix is Dnf, which is non-zero. Therefore, we can apply the inverse
function theorem (Theorem 8.2), and conclude that h is locally invertible at (a, b). If u is the
local inverse of h, and we write u = (u1, u2), then you will see that u1(x, y) = x for all (x, y).
Thus, u(x, y) = (x, u2(x, y)) for all (x, y). We now define g(x) = u2(x, 0), and show that this
has the required properties.
Also, (x, 0) = h(u(x, 0)) = h(x, (x, 0)) = h(x, g(x)) = (x, f(x, g(x))). This implies that
f(x, g(x)) = 0.
Solution : We note that f(0, −2, 0) = 0, and D2f = 2y = − 4 at (0, −2, 0). So, applying the
implicit function theorem, there exists the required neighbourhood of (0, −2, 0). In fact, you
can check that in the neighbourhood, N = B((0, − 2, 0), 1), we can express the function as
y = − (4 – x2)1/2 .
Here are some exercises that you should try :
1) Determine whether the following functions are locally invertible at the given points :
2) For each of the following functions, show that the equation f(x, y, z) = 0 defines a
continuously differentiable function z = g(x, y), in a neighbourhood of the given point:
That brings us to the end of this chapter. We hope you have studied the concepts carefully,
and have understood them.
In this chapter we have discussed two very important theorems: the inverse function theorem,
and the implicit function theorem. The proofs of these theorems are a little complicated. So
we have tried to go step by step from functions of one variable to functions of many
variables.
The Inverse Function Theorem: gives the conditions under which a function, even though not
invertible on its domain, is seen to be locally invertible. The Jacobian of the function being
non-zero at a point ensures the local invertibility of the function in a neighbourhood of that
point.
The Implicit Function Theorem: gives the conditions, under which an implicit relationship
between variables can be expressed in an explicit manner. Here, again, the Jacobian plays an
important role.
1
1
RIEMANN INTEGRAL - I
Unit Structure :
1.1 Introduction
1.2 Partition
1.3 Riemann Criterion
1.4 Properties of Riemann Integral
1.5 Review
1.6 Unit End Exercise
1.1 INTRODUCTION
1.2 PARTITION
Refinement :
Definition : Let A be a rectangle in n and f : A be a bounded
function and P be partition of A for each sub-rectangles of the
partition.
ms f inf f x : x S
g .l.b.of f on xs 1 , xs
Ms f sup f x : x S
l.u.b.of f on xs 1 , xs
where S 1, 2,...., n
Since ms M s we have L f , p U f , p
Let U f inf U f , p
L f sup L f , p
Theorem :
Let P and P be partitions of a rectangle A in n . If P
refines P then show that L f , p L f , P and U f , P U f , p .
Proof :
Let a function f : A is bounded on A P & P* are two
partition of A and P is retinement to P.
L f , p ms f V s
s p
ms f V s ms f V s1 .... V sk
ms1 f V s1 ..... ms f V sk
k
L f , p L f , p1
Now, M s f sup f x ; x S
sup f x ; x Si
Ms f Ms f i 1,..., K
i
4
U f , p ms f V s
s p
Theorem :
Let P1 & P2 be partitions of rectangle A & f : A be
bounded function. Show that L f , P2 U f , P1 &
L f , P1 f , P2 .
Proof :
Let a function f : A be a bounded find P1 & P2 are any
two partition of A.
Let P P1 P2
P is a refinement of both P1 & P2
U f , P U f , P1 ……….. (I)
U f , P U f , P2 ……….. (II)
L f , P L f , P1 ……….. (III)
L f , P L f , P2 ……….. (IV)
We get U f , P1 U f , P L f , P L f , P2 .
Hence U f , P1 L f , P2
Similarly, U f 2 , P2 U f , P L f , P L f , P1 .
Hence, U f , P2 L f , P1
Theorem :
Let a function f : A be bounded on A then for any
0, a partition P on A such that U f , P U f and
L f , P L f
5
Proof :
Let a function f :A be bounded on A
U f inf U f , P and L f sup L f , P for any 0,
partitions P1 & P2 of A such that U f , P1 U f &
L f , P2 L f .
Proof :
Let a function f : A is bounded.
U f inf U f , P
L f sup L f , P
Let f be integrable of A
U f L f
for any 0, a partition P on A such that U f , p U f 2
and L f , p L f 2 .
U f , p U f 2 & L f , p L f 2 .
U f , p L f , P U f 2 L f 2 .
U f , p L f
Conversely,
Let for any 0, a partition P on A such that
U f , p L f , P .
U P, f U f U f L f L f L f , P
6
Since U f , P U f o,
U f L f o
and L f L f , P o
we have, o U f L f
Since is arbitrary, U f L f
f is integrable over A.
Example 1
Let A be a rectangle in n and f : A be a constant
function. Show that f is integrable and f C.V A for some C .
A
Solution :
f x Cx A
f is bounded on A
Let P be a partition of A
ms f inf f x ; x s C
M s f sup f x ; x s C
L f , P ms f V S C V S CV A
S S
U f , P M s f V S C V S CV A
S S
U f L f CV A
f is integrable over A.
by Reimann criterion, 0 s.t.
f C.V A for some C .
A
Example 2 :
Let F : 0,1 X 0,1
oif xisrational
f x, y
1if xisirrational
Solution :
Let P be a partition of 0,1 0,1 into S subport of P.
7
L f , P ms f V S 0
S
U f , P M s f V S 1
S
U f 1, L f 0
U f L f
f is not integrable 0,1 0,1
Proof :
Since f is integrable over A.
by Riemann Criterion, a partition P of A.
Such that U f , P L f , P ……… (I)
inf g x sup f x
x A x A
P is refines P, we have
L f , P L f , P U f , P U f , P
U f , P L f , P U f , P L f , P
8
Now
U g , P U f , P
Ms g Msij f V sij
d
ij
i 1
Msij g Msij f u
2
n
U g , P U f , P u V Sij
d
i 1 j 1
2n
Let V sup V Sij U g , P U f , P uV d 2 nu.v
d
1 1
…….
i 1 j 1
(II)
U g , P1 L g , P1 U f , P1 d 2 n u L f , P1 d 2 n
d 2n u V
2
d 2 u
n
2 d 2n 1 u 2 2
U g , P1 L g , P1
Note that g U g , P1 U f , P1 d 2n u
A
L f , P1 d 2n u
2
d 2n u
L f , P1 n 1
2 d 2 u
9
L f , P1
2 2
L f , P
1
f
A
Now g L g , P L f , P 2
A
U f , P
f f
2
A A
f inf U f , P
A
g f
2
A A
i) ms f ms g ms f g and
ii) M s f M s g M s f g
Deduce that
L f , P L g , P L f g , P and
U f g, P U f , P U g, P
Solution :
Let P be a partition of A and S be a Subrectangle
ms f inf f x ; x S
ms f f x x S
10
Similarly ms g g x x S
ms f ms g f x g x x S
ms f ms g is lower bound of
f x g x ; x S f g x ; x S
ms f ms g is lower bound of
f x g x ; x S f g x ; x S
m f m g inf f g x ; x S
s s
ms f g
ms f ms g ms f g
ii) Ms f sub f x ; x s
Ms f f x x s
Similarly Ms g g x x S
Ms f Ms g f x g x x S
Ms f Ms g is upper bound of
f x g x ; x S f g x ; x S
Ms f Ms g sup f g x ; x S Ms f g
Ms f Ms g Ms f g
Hence,
L f , P L g , P Ms f Ms g V S
s p
Ms f g V S
s p
L f g , P
L f , P L g, P L f g, P
U f , P U g , P Ms f Ms g V S
s
Ms f g V S
s
U f g , P
U f , P U g , P U f g , P Proved.
11
Proof :
Let P be any partition of A then
U f g , P L f g , P U f , P U g , P L f , P L g , P
U f , P U g , P L f , P L g , P …………………….. (I)
f is integrable.
L f , P1 L f , P* ; U f , P1 U f , P* & L g , P2 L f , P* ;
U g , P2 U g , P* ………………………………………….. (IV)
2 U f , P1 L f , P1 U f , P* L f , P*
2 U g , P2 L g , P2 U g , P* L g , P* ……………….. (V)
f f , P 2 .
A
1
12
U f , P3 f
2
A
U g , P4 g
2
A
Let P P1 P2 P3 P4 .
Then f f , P 2 L f , P 2
A
1
Similarly g L g , P 2
A
U f , P f and U g , P g 2
2
A A
f g L f , P L g , P L f g , P f g
A A A
U f g, P
U f , P U g, P
f g
2 2
A A
f g
A A
f g f g f g
A A A A A
Proof :
Let C
Case 1
Let 0 and suppose C 0 .
Let P be a partition of A and S be a subrectangle of P.
13
M s Cf sup Cf x ; x S
sup Cf x ; x S
C sup f x ; x S
CMs f
Similarly,
ms Cf Cms f
U Cf , P Ms Cf v S C Ms f v S
S S
CU f , P
Similarly L Cf , P CL f , P
f is integrable for above 0, a partition P of A such that
U f , P L f , P C
U Cf , P L Cf , P CU f , P CL f , P
C U f , P L f , P
C C
C
By Riemann Criteria.
Cf is integrable
for 0, a partition P of A such that
C f C f CL f , P L Cf , P
C
A A
Cf U Cf , P
A
CU f , P C f
C
A
f Cf C f C f
C C
A A A A
Cf C f
A A
Case II
Now suppose C 0
Let P be a partition of A and S be any subrectangle in P.
Ms Cf CMs f and
14
ms Cf CMs f
L Cf , P CU f , P and
U Cf , P CL f , P
f is integrable for above 0, a partition P of A such that
U f , P L f , P
C
U Cf , P L Cf , P CL f , P CU f , P
C U f , P L f , P
C
C
By Riemann Criteria Cf is integrable.
for 0, a partition P of A such that C f Cf C f .
A A A
Cf C f
A A
Example 3:
Let f , g : A R be integrable & suppose f g show that
f g .
A A
Solution :
By definition f inf U f , P and g inf U g , P .
A A
Example 4:
If f : A is integrable show that if is integrable and
f
A A
f .
Solution :
Suppose f is integrable first we have to show that f is integrable.
L f , P ms f V S
S
M s f ms f V S M s f ms f V S
P P
U f , P L f , P
U f , P L f , P U f , P L f , P
By Riemann criteria
f is integrable over .
Now F inf U f , P
P
A
inf M s f V S
P
S P
inf M s f V S
P
inf M s f V S
P
P
16
inf M s f V S
P
inf U f , P
f f
A A
Example 5:
Let f : A and P be a partition of A show that f is
integrable iff for each sub-rectangle S the function f s which consist
f s .
of f restricted to S is integrable and that in this case f
A S S
Suppose f : A is integrable.
Let P be a partition of A & S be a sub-rectangle in P.
k
Msi f msi f V S
i 1
U f S
,P L f
S
,P
By Riemann Criterion
f is integrable.
S
U f s
, PS L f
s
, PS k ………………………………. (II)
S P
1 1
Ms1 f ms1 f V S 1
S P S1PS1
U f s , PS1 L f s , PS1
S P
k
S P
k , k
Let 0
f
S P S
S k L f S , PS
SP
m1s f V S
S P S 1PS1
f S
k ms1 f V S 1
S P S S1P1
L f , P1 f U f , P1
A
M f V S
s1
1
S1P1
M s1 f V S 1
S P S 1P1
U f S , PS f S C
S P S
k
S P
f S C f f S
S P A S P
f f S
A S P S
Example 6:
Let f : A be a continues function show that f is
integrable on A.
Solution :
Let f : A be a continuous function to show that f is
integrable.
A is compact.
n S
n
f x f y V A
S is compact
f is continuous
f attains its bound in S.
k
U f , P L f , P Msi f msi f V Si
i 1
k
f xi f yi V Si
i 1
k
k
V Si V Si
i 1 V A V A V A
V A
V A
1.5 REVIEW
1 1 1
VIII) A function f ; 0,1 is defined as f x n 1
n x n 1
3 3 3
where n
f 0 0
1
show that f is R-integrable on 0,1 & calculate f x dx .
0
X) A function f ; a, b is continuous on a, b f x 0
b
x a, b and f x dx 0 show that f x 0 x a, b .
a
21
2
MEASURE ZERO SET
Unit Structure :
2.1 Introduction
2.2 Measure zero set
2.3 Definition
2.4 Lebesgue Theorem (only statement)
2.5 Characteristic function
2.6 FUBIN’s Theorem
2.7 Reviews
2.8 Unit End Exercises
2.1 INTRODUCTION
Definition :
A subset ‘A’ of n said to have measure ‘O’ if for every
0 there is a cover U1 ,U 2 .... of A by closed rectangles such that
the total volume v Ui .
i 1
Theorem :
A function ‘f’ is Riemann integrable iff ‘f’ is discontinuous
on a set of Measure zero.
22
Example 1:
1) “Counting Measure” : Let X be any set and M P X the set of
all subsets : If E X is finite, then E E if E X is
infinite, then E
2) “Unit mass to x0 - Dirac delta function” : Let X be any set and
M P X choose x0 X set.
E 1if x0 E
0if x0 E
Example 2:
Show that A has measure zero if and only if there is countable
collection of open rectangle V1 ,V2 ,.... such that A Vi and
V vi .
Solution :
Suppose A has measure zero.
For 0, countable collection of closed rectangle V1 ,V2 ,....
such that A Vi and
i 1
V V 2 .
i 1
i
Then A vi ui
i 1 i 1
and V ui V ui 2V vi
i 1 i 1 i 1
2 v ui 2
i 1 2
Then A ui vi and
i 1 i 1
V vi V ui .
i 1 i 1
Solution :
Let A a1 ,...., am be finite subset of n .
Let 0, ai ai1 , ai 2 ,....., ain and
1 n
1
1 n
1
1
n
n
Then V Vi i 1 i 1
i 1 2 2
Clearly ai Vi for 1 i m
m m m
1 1
A Vi and
i 1
V Vi
i 1 i 1 2 i 1
i 1 2
i 1
2
Example 4:
If A A1 A2 A3 .... and each Ai has measure zero, then
show that A has measure zero.
Solution :
Let 0 and A A1 A2 .... with each Ai has measure zero.
24
By closed rectangle such that V u 2 ,i 1, 2,....
i 1
ii i
Example 5:
Let A n be a Rectangle show that A does not have
measure zero. But A has measure zero.
Proof :
Suppose A has measure zero.
A is a rectangle in n
V A 0
and V u .
i
A is compact
A ui .
i 1
Let P be partition of A that contains all the vertices all ui ' si 1 to
k. Let S1 , S 2 ,...., S n denote the subrectangle of partitions.
V A V S j V ui V ui
n k
j 1 i 1 i 1
25
Example 6:
Let A n with A . Show that A does not measure zero.
Solution :
Let A n , with A
Let x A
r 0 , such that B x, r A, But
B x, r y A; y x r
n
y A; yi xi r
i 1
Example 7:
Show that the closed interval a, b does not have measure
zero.
Solution :
Suppose ui i 1 be a cover of a, b by open intervals.
a, b is compact this open cover has a finite subcover.
n
Let u ui
i 1
a, b ui for some i
a, b u j for i j
Which is not possible
u is connected
u is an open interval say u c, d Then as a, b u c, d
V ui d c b a
Example 8:
If A 0,1 is the union of all open intervals ai , bi such that
each rational number in (0,1) is contained in some ai , bi . If
T bi ai 1 then show that the boundary of A does not have
i 1
measure zero.
Solution :
We first show that A 0,1 \ A
Note that A A \ A
A is open A A
Also Q 0,1 A
Q 0,1 A
0,1 A
But A 0,1 A 0,1
A 0,1
A 0,1 \ A
27
Let 1 T 0
ui 1 T
i n; ai , bi i 1 cover 0,1 and sum of lengths
Note that ui ;1
2.3 DEFINITION
V u
i 1
i
Remark :
1) If A has content O, then A clearly has measure O.
2) Open rectangles can be used instead of closed rectangles in the
definition.
Example 9:
If A is compact and has measure zero then show that A has
content zero.
Solution :
Let A be a compact set in n
Suppose that A has measure zero
a cover u1 , u2 ,.... of A such that V u for every 0 .
i 1
i
V u V u
i 1
i
i 1
i
Example 10 :
Give one example that a set A has measure zero but A does
not have content zero.
Solution :
Let A 0,1 Q
Then A is countable
A has measure zero
Now to show that A does not have content zero.
Let ai , bi ;1 i n be cover of A
A ai , bi .... an , bn
A a1 , b1 .... an , bn
But A 0,1
n
ai , bi 1
i 1
a , b 12
i 1
i i
Example 11:
Show that an unbounded set cannot have content zero.
Solution :
Let A n be an unbounded set.
To show that A does not have content zero
Suppose A has content zero for 0, finite cover of closed
k k
rectangles ui i 1 of A such that A ui and V u .
k
i
i 1 i 1
Example 12:
f : A is non-negative and f 0
A
where A is rectangle,
Solution :
For n , An x A; f x 1 n
Note that x A, f x 0 x A; F x 0
f is non-negative}
n
x A; f x 1 A n
n 1 n 1
U f , P n
Let S be a subrectangle in P
if S An M s f 1 n
clearly S P; S An covers An and
1 1
n V S M f V S M f n
S P S P
s s
f , P n
V S
S An
s p
* Oscillation o f , a of ‘f’ at a
for 0 , Let M a, f , sup f x ; x A& x a
m a, f , inf f x ; x A& x a
The oscillation o f , a of f at a defined by
o f , a lim M a, f , m a, f ,
o
30
Theorem :
Let A be a closed rectangle and let f : A be a bounded
function such that O f , x for all x A show that there is a
partition P of A with U f , P L f , P V A .
Proof :
Let x A U f , x lim M x, f , m x, f ,
O
V S
S P
V A
Theorem :
Let A be a closed rectangle and C A . Show that the
function c : A is integrable if and only if C has measure zero.
Proof :
To show that C : A is integrable iff C has measure
zero.
Theorem :
Let A be a closed rectangle and C A
32
A
c 0.
Proof :
C A be a bounded set with measure zero.
To show that c 0
A
A
c O
Fubini’s Theorem
Statement : Let A n and B n be closed rectangles and let
f : A B be integrable for x A , Let g x : B be defined by
g x y F x, y and let
x L g x L f x, y dy
B B
u x U g x U f x, y dy
B B
Then and are integable on A and f L L f x dy dx
A B A A B
f u x dx U f x, y dy dx
A B A A B
Proof :
Let PA be a partition of A and PB be a partition of B. Then
P PA , PB is a partition of A B
Let S A be a subrectangle in PA and S B be a subrectangle in PB
Then by definition,
S S A S B is a subrectangle in P
L f1 P ms f V S
S P
m
S B PB
s A sB f V S A S B
m s A sB f V S B V S A …………………. (I)
S A PA SB PB
For x S A , ms A sB
f M s gx B
For x S A,
m
S B PB
s A sB V S A V S B msB g x V S B
L g x , PB L g x L x
B
L x , PA ……………………………………… (II)
34
Now U f , P M S f V s
S P
S A PA
M S ASB f V S A S B
S B PB
M f V S B V S A …………….. (IV)
S AS B
S A PA S B PB
For x S A , M S S f M S g x
A B B
For x S A ,
M
S B PB
S AS B f V S B M
S B PB
SB g x V S B
u g x , PB u g x x
B
M u xV S
S A PA
SA A
u x , PA ……………………………………….. (V)
U f , P U u x , PA ……………………………. (VI)
By (III) & (VI)
L f , P L x , PA u L x , PA
u x , PA U f , P ………………………… (VII)
Also
L f , P L x , PA L x , PA u x , PA …………. (VIII)
f is integrable
sup L f , P inf U f , P f
P P
AB
sup L x , PA inf u x , PA f
PA PB
AB
x is integrable
35
f x L f x, y dx ………………………. (IX)
B
AB A A
sup L L x , PA inf U u x , PA f
PA PA
AB
u x is integrable.
f u x dx U f x, y dx
B
AB A A
Hence Proved
Remark :
The Fubini’s theorem is a result which gives conditions under
which it is possible to compute a double integral using interated
integrals, As a consequence if allows the under integration to be
changed in iterated integrals.
f L f x, y dxdy
B
AB B
U f x, y dxdy
A
B
Example 13:
Using Fubini’s theorem show that D12 f D21 f if D12 f and
D21 f are continuous.
Solution :
Let A R and f : A continuous
T.P.T D12 f D21 f
Suppose D12 f D21 f
D12 f D21 f x, g 0
A
Let A a, b c, d
By Fubini’s Theorem
d b
D 21 f x, y D 21 f x, y dxdy
A c a
d
D21 f x, y D12 f x, y
A A
D21 f D12 f x, y 0
A
Example 14:
Use Fubini’s Theorem to compute the following integrals.
1 1 x 2
I
dy.dx
1)
0 0
1 x2 y 2
Solution :
1 1 x 2
I
dy.dx
0 0
1 x2 y 2
1 1 x 2
dx
dy
0 0
1 x2 y2
1 y
1 1 x 2
dx tan 1
1 x 1 x 2 0
2
0
1
1
dx. .
0 1 x2 4
37
1
4 0
dx
1 x2
2
1
log x 1 x
4 0
log x 1
4
1 1
x 2
I dy sin dx
2
ii)
0 y
Solution :
C x, y ; y x 1, 0 y 1
By Fubini’s Theorem
1 1
x 2
I sin dxdy
2
0 y
1 x
x 2
sin dxdy
2
0 0
1
x 2 x
sin y dx
2 0
0
1
x 2
xsin dx
2
0
x2 x 1
Put t,
2 t 0
2
2 x
dx dt
2
dt
xdx
2 2
I sin t sin tdt cos t 0 2
dt 1 1
0
0
0 1
1 1
38
2.7 REVIEWS
1. If B A and A has measure zero then show that & has measure
zero.
2. Show that countable set has measure zero.
3. If A is non-empty open set, then show that A is not of measure
zero.
4. Give an example of a bounded set C if measure zero but C does
not have measure zero.
5. Show by an example that a set A has measure zero but A does
not have content zero.
6. Prove that a1 , b1 .... an , bn does not have content zero if ai bi
for each i .
7. If C is a set of content zero show that the boundary of C has
content zero.
8. Give an example of a set A and a bounded subset C of A measure
zero such that c does not exist.
A
f x, y dxdy f x, y dydx
a x a x
2 2
dy
sin x
12. Use Fubini’s theorem, to compute dx
0 0
x y
39