Analysis 3
Analysis 3
Jaydeb Sarkar
Statistics and Mathematics Unit
Indian Statistical Institute, Bangalore
LATEXed by
Trishan Mondal
Soumya Dasgupta
Aaratrick Basu
Priyatosh Jana
Bikram Halder
Saraswata Sensarma
Deepta Basak
ii
Contents
Credits ii
Contents iii
Notation vii
Lecture 1 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Review: Rn as a vector space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Linear Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.4 Analytic ideas in Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
Lecture 2 5
2.1 Distance and Topology in Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Limits and Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Lecture 3 9
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2 Properties of Continuous functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
Lecture 4 13
4.1 More examples of Continuous maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.2 Uniform Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.3 Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
Lecture 5 17
5.1 Matrix representation of the derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.2 Further properties of differentiable functions . . . . . . . . . . . . . . . . . . . . . . . . 18
Lecture 6 21
6.1 Partial Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
6.2 Geometric Meaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
6.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
6.4 Higher Order Partial Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
6.5 Clairaut’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Lecture 7 25
7.1 Schwarz Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
7.2 Partial and Total Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Lecture 8 27
8.1 A kind of converse of Theorem 7.2.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
8.2 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
iii
Lecture 9 31
9.1 Directional Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
9.2 Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
9.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Lecture 10 35
10.1 Extension of MVT to Several Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
10.2 More Partials and Chain Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Lecture 11 39
11.1 Chain Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
11.2 Laplacian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
11.3 Extrema of a function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
Lecture 12 45
12.1 Hessian Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
12.2 Positive Definite, Negative Definite, Semi Definite Matrices . . . . . . . . . . . . . . . . 46
Lecture 13 49
13.1 Taylor’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Lecture 14 53
14.1 Compact subsets of Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
14.2 Inverse Function Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Lecture 15 59
15.1 Inverse function theorem: Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
15.2 Implicit Function Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
15.3 Solving systems of equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Lecture 16 63
16.1 Riemann-Darboux Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Lecture 17 65
17.1 Properties of Riemann-Darboux Integration . . . . . . . . . . . . . . . . . . . . . . . . 65
17.2 Iterated Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Lecture 18 69
18.1 Fubini’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
Lecture 19 73
19.1 Integration over Bounded Domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
19.2 Two Elementary Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
Lecture 20 77
20.1 Fubini’s Theorem on Elementary Regions . . . . . . . . . . . . . . . . . . . . . . . . . . 77
20.2 Change of Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
Lecture 21 81
21.1 Change of Variables (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Lecture 22 85
22.1 Curves and Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Lecture 23 89
23.1 Line Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
iv
Lecture 24 93
24.1 Planes and Normals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
24.2 Surface and Surface Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
24.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Lecture 25 97
25.1 Tangent Plane Of G(f ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
25.2 Surface Area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
Lecture 26 101
26.1 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
26.2 Surface Integral over Scalar fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
26.3 Surface Integral over a Vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Lecture 27 105
27.1 Conservative Vector Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
27.2 Green’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
Lecture 28 111
28.1 Green’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
28.2 Gauss Divergence Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
28.3 Stokes’ Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Index 117
v
vi
List of Notation
∇f Gradient of f
∇f (x) Gradient of f at x
G(f ) Graph of a f : R2 → R
Hf (x) Hessian of f at x
∂ ∂f ∂2f
fxi xj (a), , Partial derivative of f with respect to the ith and j th coordi-
∂xj ∂xi ∂xi ∂xj
nates at a
χS Indicator function of S
⟨x, y⟩ Euclidean inner product of x and y
R(S) Space of all Riemann-Darboux integrable functions over S
Int(S),
Z S◦ Interior of a set S ⊆ Rn
f dV Riemann-Darboux integral of f over S
S
∆f (x) Laplacian of f at x
vii
lim f Limit of f as x tends to a
x→a
′
S
Z Z Set of limit points of S ⊆ Rn
f, f Line integral of f over C = ran(γ)
Zγ C Z
F⃗ · d⃗r, F⃗ · d⃗r Line integral of F⃗ over C = ran(γ)
Zγ C
On , O Open set in Rn
viii
Lecture 1
1.1 Introduction
We will talk about n-variable calculus, that is, calculus on Rn . Recall the following:
Rn = R × R · · · × R = {x = (x1 , . . . , xn ) | xi ∈ R ∀ i = 1, 2, . . . , n}
| {z }
n times
• Analysis on R consisted of ideas like open sets, compact sets, convergence, limits, differentiability,
integrability etc.
• Rn is an n-dimensional inner product space over R, with the standard orthonormal basis
{ej }nj=1
Extending the analytic ideas to Rn exploiting the algebraic structure is the matter of this course,
which further gives way to differential geometry.
1
1.4 Analytic ideas in Rn
f : Rn → R m
∀ x ∈ Rn , f (x) = a, a ∈ Rm
It turns out that linear maps are useful in understanding most other ‘nice’ functions, and so we now
look at these in more detail.
Let L be a linear map from Rn to Rm . Consider the domain {αx + y | α ∈ R}, that is, the line
through y in the direction of x. The image under L is,
{αLx + Ly | α ∈ R}
which is the line through Ly in the direction of Lx. Hence, L maps lines to lines.
Exercise. Is the converse also true?
L(x) = xL(1) ∀x ∈ R
Therefore,
L(R, R) := {set of all linear maps from R to R} ↔ R
Now consider the general case; let L : Rn → Rm be a linear map. If we fix the bases {ej }nj=1 of Rn
and {ei }m m
i=1 of R , L is determined uniquely by the equations
m
X
Lej = aij ei
i=1
and hence,
L ↔ (aij )m×n ∈ Mm,n (R)
n
! 12
X
∥x∥ = x2i ∀ x ∈ Rn
i=1
n
! 21
X
d(x, y) = ∥x − y∥ = (xi − yi )2 ∀ x, y ∈ Rn
i=1
2
Lecture 1
|x + y| ≤ |x| + |y|
Proof. We have,
2
∥x + y∥ = ⟨x + y, x + y⟩
2 2
= ∥x∥ + 2 ⟨x, y⟩ + ∥y∥
2 2
≤ ∥x∥ + ∥y∥ + 2∥x∥∥y∥ (Cauchy Schwarz inequality)
=⇒ ∥x + y∥ ≤ ∥x∥ + ∥y∥
The following is a technical result, which can be thought of as an analogue of the Lipschitz condition
for linear maps on Rn . It hints towards continuity of linear maps, and we will see that it is indeed so
later.
Theorem 1.4.3
Let L : Rn → Rm be a linear map. There is M > 0 such that
∥Lx∥ ≤ M ∥x∥ ∀ x ∈ Rn
Pn
Proof. We have, for x = i=1 xi ei ,
n
X
∥Lx∥ = xi Lei
i=1
n
X
≤ |xi |∥Lei ∥ (Triangle inequality)
i=1
n
! 12
X 2
=⇒ ∥Lx∥ ≤ ∥x∥ ∥Lei ∥ (Cauchy Schwarz inequality)
i=1
P 12
n 2
Taking M = i=1 ∥Lei ∥ , we get the result.
3
1.4 Analytic ideas in Rn
4
Lecture 2
We wish to extend the notions of (R, |·|) to (Rn , ∥·∥). We already know from Lecture 1 that the
triangle inequality holds,
∥x + y∥ ≤ ∥x∥ + ∥y∥
Br (a) = {x ∈ Rn | ∥x − a∥ < r}
Example 2.1.1
5
2.1 Distance and Topology in Rn
Π i : Rn → R
x = (x1 , . . . , xn ) 7→ xi
Theorem 2.1.1
Let {xm }m∈N ∪ {x} ⊆ Rn . Then,
xm −→ x
⇐⇒ Πi (xm ) −→ Πi (x) ∀ i ∈ {1, 2, . . . , n}
Exercise. Show that a set C ⊆ Rn is closed iff ∀ {xm }m∈N ⊆ C with xm −→ x for some x ∈ Rn , we
have x ∈ C.
Exercise. Show that:
(1) Arbitrary union of open sets is open.
(2) Finite intersection of open sets is open.
(3) Arbitrary intersection of closed sets is closed.
(4) Finite union of closed sets is closed.
(5) Any finite subset of Rn is closed.
6
Lecture 2
Example 2.1.2
Example 2.1.3
For S = [0, 1) ∪ (1, 2] ∪ {5}, ∂S = {0, 1, 2, 5} but the set of limit points is {0, 1, 2}.
(5) ∂S is closed.
Qn
(6) Let {Oi }ni=1 ⊆ P(R) and define O = i=1 Oi . If Oi ’s are open (closed), O is open (closed).
7
2.2 Limits and Continuity
Note: Any function defined on S is vacuously continuous at an isolated point a by our definition.
8
Lecture 3
3.1 Introduction
We denote the set of limit points of S ⊆ Rn by S ′ . Let f : S → Rm and a ∈ S. f is continuous at a
iff for all ε > 0, there is δ > 0 such that
lim ∥f (a + h) − f (a)∥ = 0
h→0
Theorem 3.1.1
Let S ⊆ Rn , a ∈ S ′ , b ∈ Rm , f : S → Rm . The following are equivalent:
(i) lim f = b
x→a
Note: If a ∈ S, we can take b = f (a) and get analogous results for continuity at a.
9
3.2 Properties of Continuous functions
Theorem 3.2.1
Let S ⊆ Rn , f : S → Rm . The following are equivalent:
(1) f is continuous on S.
(2) ∀ {xp } ⊆ S with xp −→ a ∈ S, we have f (xp ) −→ f (a)
(3) (Assuming S is open) f −1 (O) is open for all O ⊆ Rm open.
(4) (Assuming S is open) f −1 (C) is closed for all C ⊆ Rm closed.
• (3) ⇐⇒ (4)
This is true as if g : X → Y then for all A ⊆ Y , g −1 (Y \ A) = X \ g −1 (A).
• (1) ⇐⇒ (2)
True by 3.2.1
• (1) =⇒ (3)
Let O ⊆ Rm be open, and without loss of generality, f −1 (O) ̸= ϕ.
Let a ∈ f −1 (O) so that f (a) ∈ O. Hence, there is r > 0 such that Br (f (a)) ⊆ O. By continuity,
there is δ > 0 such that
Hence, O is open.
• (3) =⇒ (1)
Fix a ∈ S and let ε > 0. As Bε (f (a)) is open in Rm , we have f −1 (Bε (f (a))) is open in Rn .
But a ∈ f −1 (Bε (f (a))), and so, there is δ > 0 such that
This theorem gives us a huge simplification. Recall that x −→ y iff Πi (x) −→ Πi (y) for all i.
Now consider some {xp } with xp −→ a. We have
That is, f is continuous iff f is continuous coordinate wise! Hence, for talking about continuity, it is
enough to discuss real valued functions rather than f : Rn → Rm for arbitrary m.
10
Lecture 3
3.3 Examples
Example 3.3.1
f : R2 \ {(0, 0)} → R
2xy
f (x, y) = 2
x + y2
Consider the line L1 defined by y = 0 and approach (0, 0) from the right (x −→ 0+ ). We
have,
1
f L ≡ 0 =⇒ lim f = lim f 0, =0
1 (x,y)→0 along L1 n→∞ n
Now consider the line L2 defined by x = y. We have,
1 1
f L2
≡ 1 =⇒ lim f = lim f , =1
(x,y)→0 along L2 n→∞ n n
Note: The approach in the above example is often useful for showing non-existence of limits, or that
a function is not continuous.
Example 3.3.2
x3
We wish to compute lim . We have, for all (x, y) ̸= (0, 0),
(x,y)→(0,0) x2 + y 2
x3 x3
≤ = |x| ≤ ∥(x, y)∥
x2 + y 2 x2
x3
lim =0
(x,y)→(0,0) x2 + y 2
Example 3.3.3
sin x2 + y 2
Consider lim . We have (x2 + y 2 ) −→ 0 as (x, y) → (0, 0), and hence, we
(x,y)→(0,0) x2 + y 2
get
sin x2 + y 2 sin t
lim 2 2
= lim =1
(x,y)→(0,0) x +y t→0 t
11
3.3 Examples
(ii) limx→a f g = αβ
f α
(iii) If β ̸= 0, lim =
x→a g β
(iv) If f ≤ h ≤ g for some h : S → R and α = β, then lim h = α
x→a
Note: Similar results hold for continuity as well, using which we get the next examples.
12
Lecture 4
Example 4.1.1
0, (x, y) = (0, 0)
f is clearly continuous on R2 \ {(0, 0)} so we need only check the limit at (0, 0). We have,
xy 1 x2 + y 2 1
0≤ p ≤ p = ∥(x, y)∥
x2 + y 2 2 x2 + y 2 2
Exercise. Show that any linear map is continuous. (Hint: Use that the norm is continuous.)
Example 4.1.2
f :D→R
1
f (x, y) = x sin
y
Clearly f is continuous on D. We also have,
1
0 ≤ x sin ≤ ∥(x, y)∥
y
13
4.2 Uniform Continuity
4.3 Derivatives
We will use the following notation for the sake of brevity:
(1) On denotes an open set in Rn , and we omit the n if it is clear from context.
(2) A function f : Rn → Rm has components f = (f1 , f2 , . . . , fm )
14
Lecture 4
Theorem 4.3.1
Let f : On → Rm be differentiable at a ∈ On . The derivative Df (a) is unique.
Proof. Let L = Df (a) and L1 be any linear map from Rn to Rm such that
∥f (a + h) − f (a) − L1 h∥
lim =0
h→0 ∥h∥
̸ L. Then, there is h0 ∈ Rn such that ∥h0 ∥ = 1 and Lh0 ̸= L1 h0 . Consider the map
Suppose L1 =
n
h : R → R given by h(t) = th0 . Then, by the triangle inequality,
Taking the limit as t → 0, both terms on the right go to 0 by definition, and hence
which clearly contradicts the assumption. So, the derivative Df (a) is unique.
4.4 Examples
Example 4.4.1
f (a + h) − f (a) − Oh
lim =0
h→0 ∥h∥
where O denotes the zero linear map. Hence, f is differentiable on On and Df (a) = O for
any a ∈ On .
Example 4.4.2
L(a + h) − La − Lh
lim =0
h→0 ∥h∥
At this point, we are faced with the problem of actually computing derivatives of non-trivial maps. A
priori, it is not even clear if functions that are made of various differentiable functions of one variable,
say f (x, y, z) = (x2 ey z, y 3 sin(xy) cos z), are differentiable! We will perform a series of reductions
that will answer such basic questions about differentiability of functions and even provide techniques
to compute the derivatives.
15
4.4 Examples
16
Lecture 5
We have figured out that the reasonable definition of the derivative of a map f : On → Rm at a point
a ∈ On is Df (a) = L where
f (a + h) − f (a) − Lh
lim =0
h→0 ∥h∥
We now proceed to perform a series of reductions that will actually answer the question of
differentiability of functions and also provide techniques to compute the derivative.
Theorem 5.1.1
Let f : On → Rm and a ∈ On . Then, f is differentiable at a iff fi is differentiable at a for all
i. In that case, we have
Df1 (a)
Df2 (a)
Df (a) =
..
.
Dfm (a)
Proof. Assume that f is differentiable at a and let L = Df (a). For all i = 1, 2, . . . , m we set
Li = Πi ◦ L. Further, let f˜i (h) = fi (a + h) − fi (a) − Li h, so that
But then, f˜i (h) ≤ ∥f (a + h) − f (a) − Lh∥ which proves that fi is differentiable at a and has
derivative Dfi (a) = Li , for all i.
Assume that fi is differentiable at a for all i and define
Df1 (a)
Df2 (a)
L=
..
.
Dfm (a)
f˜i (h)
lim = 0 ∀i
h→0 ∥h∥
f (a + h) − f (a) − Lh
=⇒ lim =0
h→0 ∥h∥
17
5.2 Further properties of differentiable functions
Taking the limit as x → a, the first term on the right goes to 0 by the definition of Df (a) and the
second term goes to zero as any linear map is continuous. Hence, by the squeeze theorem,
g◦f
Proof. Let A = Df (a) and B = Dg(b) where b = f (a). For x, y in sufficiently small neighbourhoods
of a, b respectively, we consider the maps
18
Lecture 5
Now,
Brf (x) Brf (x)
lim =B lim =0
x→a ∥x − a∥ x→a ∥x − a∥
The other term requires some more analysis. Fix ε > 0. There is δ > 0 such that ∥rg (y)∥ < ε∥y − b∥
for all y with 0 < ∥y − b∥ < δ. By continuity of f at a, there is δ1 > 0 such that ∥f (x) − f (a)∥ < δ
for all x with 0 < ∥x − a∥ < δ1 . Hence,
and so
rg (f (x))
lim =0
x→a ∥f (x) − f (a)∥
We now note,
∥f (x) − f (a)∥ = ∥rf (x) + A(x − a)∥ ≤ ∥rf (x)∥ + MA ∥x − a∥
where we get MA > 0 by Theorem 1.4.3. Hence, we finally have
19
5.2 Further properties of differentiable functions
20
Lecture 6
Considering the maps hi : R → Rn such that hi (t) = a + tei , we have ηi = f ◦ hi . Thus, by Chain
Rule, if f is differentiable, all its partial derivatives exist.
21
6.3 Examples
6.3 Examples
We now lay out some interesting and instructive examples, which would illustrate some general
results about partial derivatives.
Example 6.3.1
∂f ∂f
= 3x2 + y cos xy and = 4x3 + x cos xy
∂x ∂y
Example 6.3.2
Definition 6.3.1
A function f : O → R is called C k (O) if all the k th order partial derivatives exist and are
continuous.
We can define higher order partial derivatives similarly. Please note that the order of differentiation
matters in general. For starters, of fxi xj and fxj xi , one may exist while the other may not. Also,
even if both exist, they may not necessarily be equal over the entire domain. We leave it to the
reader to find an example of the former case, and provide an example for the latter.
Example 6.4.1
Consider f : R2 → R given by
xy(x2 −y 2 )
(
x2 +y 2 (x, y) ̸= (0, 0)
f (x, y) =
0 (x, y) = (0, 0)
In this case, it is easy to show (Exercise!) that both fxy and fyx exist, but
22
Lecture 6
Example 6.4.2
However, in many well-behaved cases, we will find fxy = fyx . For instance, consider f : R2 → R
given by f (x, y) = sin(x) + ey + xy. Show that fxx = fyy = 1 over R2 .
Often, the dependence of the partial derivative on the order of differentiation is the exception rather
than the rule. We now develop a sufficient condition for fxy = fyx to hold.
Proof. Without loss of generality, we take (a, b) = (0, 0) ∈ O2 . As O2 is open, we choose a box
[h, 0] × [0, k] ⊂ O2 . Now, we have
We define
1
F (h, k) = (f (h, k) − f (0, k) − f (h, 0) + f (0, 0))
hk
Thus, by the above result, we have
Now we proceed for the proof in earnest. Define f1 (x) = f (x, k) − f (x, 0), which is continuous on
[0, h] and differentiable on (0, h). Thus, by Lagrange’s Mean Value Theorem, there exists c1 ∈ (0, h)
(depending upon both h and k), such that, h(f1′ (c1 )) = (f1 (h) − f1 (0)), i.e.
1
∴ fx (c1 , k) − fx (c1 , 0) =
(f (h, k) − f (0, k) − f (h, 0) + f (0, 0)) = kF (h, k)
h
1
=⇒ F (h, k) = (fx (c1 , k) − fx (c1 , 0))
k
Again, define f2 (y) = fx (c1 , y), which again satisfies all conditions for the Mean Value Theorem.
Thus, there exists c2 ∈ (0, k) such that k(f2′ (c2 )) = (f2 (k) − f1 (0)), which gives, F (h, k) = fxy (c1 , c2 ).
Repeating this entire construction, we can find (c′1 , c′2 ) ∈ [0, h] × [0, k] such that F (h, k) = fyx (c′1 , c′2 ).
Thus, fxy (c1 , c2 ) = fyx (c′1 , c′2 ). But, 0 < c1 , c′1 < h and 0 < c2 , c′2 < k. Thus, as (h, k) can be made
arbitrarily small, taking (h, k) → 0, we have
By the continuity of fxy and fyx , we have fxy (0, 0) = fyx (0, 0) .
In particular, fxy = fyx for C 2 functions over a given domain. In the next lecture, we sharpen this
result slightly, and relate the partial derivatives to the total derivative.
23
6.5 Clairaut’s Theorem
24
Lecture 7
Proof. Just as before, we take (a, b) = (0, 0). From the proof of Clairaut’s Theorem, we have
F (h, k) = fxy (c1 , c2 ) for some 0 < c1 < h and 0 < c2 < k. By continuity of fxy at (0, 0), for any
ϵ > 0 there exists hϵ , kϵ > 0, such that
But then | F (h, k) − fxy (0, 0) |< ϵ ∀ (u, v) ∈ [0, hϵ ] × [0, kϵ ], that is, F is continuous at (0, 0) with
limit fxy (0, 0) at (0, 0).
As fy (h, 0) exists for h sufficiently small, limk→0 F (h, k) exists for h sufficiently small. Thus, by
continuity of F at (0, 0),
This is a slightly more useful version of Clairaut’s Theorem. However, in many applications (say,
partial differential equations), we work with C 2 or even C ∞ functions, in which case both of these
hold automatically.
Exercise: Formulate and prove a similar result for higher order derivatives. In particular, provide a
sufficient condition for f : On → R so that
∂nf ∂nf
=
∂xi1 ∂xi2 . . . ∂xim ∂xσ(i1 ) ∂xσ(i2 ) . . . ∂xσ(im )
25
7.2 Partial and Total Derivatives
Theorem 7.2.1
Consider a function f = (f1 , f2 , . . . , fm ) : On → Rm differentiable at a ∈ On . Then all the
∂fi
partial derivatives ∂xj
exist at a. In particular, for f differentiable at a, we have,
∂fi
(Df )(a) = Jf (a) = (a)
∂xj m×n
Proof. Without loss of generality, we take m = 1, and let a = (a1 , a2 , . . . , an ). Fix an arbitrary index
i ∈ {1, 2, . . . , n}. We define ηi : [ai − ϵ, ai + ϵ] → Rn , defined by
As On is open and ηi is continuous, we can find ϵ small such that f ([[ai − ϵ, ai + ϵ]]) ⊆ On ⊆ Rn .
Evidently, ηi is differentiable and (Dηi ) = [0, . . . , 1, . . . , 0]t = eti over [ai − ϵ, ai + ϵ]. Now, by the
definition of partial derivatives, D(f ◦ ηi )(ai ) = fxi (a).
Again, by chain rule, as f is differentiable at a, D(f ◦ ηi )(ai ) = fxi (a) exists, and
As the index i was arbitrary to begin with, this completes the proof.
This theorem proves that differentiability of a function implies the existence of its partial derivatives,
and gives the form of the derivative in the standard basis. But it is often quite elaborate and laborious
to prove that a function is differentiable, whereas computation of the partial derivatives is much
more straightforward. In the next lecture, we formulate a sufficient condition for differentiability of a
function based on its partial derivatives.
26
Lecture 8
n
X ∂f
(0) ∀ h ∈ Rn
L = Jf (a) = fx1 (0) · · · fxn (0) and Lh = hi
i=1
∂xi
Claim
1
|f (h) − f (0) − Lh| → 0 as h → 0
∥h∥
27
8.2 Examples
which implies,
n n
X ∂f X ∂f
f (h) − f (0) − hi (0) = f (ĥi ) − f (ĥi−1 ) − hi (0)
i=1
∂xi i=1
∂xi
n
X ∂f ∂f
= hi ĥi−1 + ci ei − hi (0)
i=1
∂xi ∂xi
| {z }
By MVT, as explained below
defined by, ηi (t) = f (ĥi−1 + tei ). Clearly, ηi is differentiable on (0, hi ) and continuous on [0, hi ].
Then, by Mean Value Theorem,
ηi (hi ) − ηi (0) = ηi′ (ci )hi = fxi (ĥi−1 + ci ei )hi (for some ci ∈ (0, hi ))
| {z } | {z }
f (ĥi ) f (ĥi−1 )
Now, observe that, as h → 0, ĥi−1 + ci ei → 0 which in turn implies, fxi (ĥi−1 + ci ei ) → fxi (0).
Therefore,
n
1 1 X ∂f ∂f
|f (h) − f (0) − Lh| = hi ĥi−1 + ci ei − hi (0)
∥h∥ ∥h∥ i=1 ∂xi ∂xi
n
1 X ∂f ∂f
≤ |hi | ĥi−1 + ci ei − (0) (Triangle inequality)
∥h∥ i=1 ∂xi ∂xi
n
X ∂f ∂f
≤ ĥi−1 + ci ei − (0) (as ∥h∥ ≥ |hi | ∀ i)
i=1
∂xi ∂xi
−→ 0 as h → 0
With Theorem 8.1.1, computation of derivative is much easier when we are in favorable situation.
Note that,
∂fi
(i) If f is differentiable at a then ∂xj (a) exists for all i, j and Df (a) = Jf (a).
∂fi
(ii) If ∂xj is continuous at a then f is differentiable and Df (a) = Jf (a).
The gap between (i) and (ii) is the continuity of partials, which is removable.
8.2 Examples
We conclude the lecture with some instructive examples.
28
Lecture 8
Take,
x2 + y 2 sin √ 1 , (x, y) ̸= 0
f (x, y) = x2 +y 2
0, (x, y) = 0
Then,
!
2 2 1
|f (x, y) − f (0, 0)| = x + y sin p
x2 + y 2
2
≤ (x2 + y 2 ) = ∥(x, y)∥
implies that f is continuous at (0, 0). For all (x, y) ̸= (0, 0), the partials of f ,
! !
1 x 1
fx (x, y) = 2x sin p −p cos p
x2 + y 2 x2 + y 2 x2 + y 2
! !
1 y 1
fy (x, y) = 2y sin p −p cos p
x2 + y 2 x2 + y 2 x2 + y 2
f (t, 0) − f (0, 0)
fx (0, 0) = lim =0
t→0 t
f (0, t) − f (0, 0)
fy (0, 0) = lim =0
t→0 t
Also,
1 h p 1
√ f (h, k) − f (0, 0) − 0 0 = h2 + k 2 sin √ ≤ ∥(h, k)∥
h2 + k 2 k h2 + h2
which shows that, f is differentiable at (0, 0) and Df (0, 0) = 0 0 . But, fx and fy are not
continuous at (0, 0)!
So, even if a function is differentiable at some point, its partials may still not be continuous there!
Take,
( 4
y
x 3 sin x , x ̸= 0
f (x, y) =
0, x=0
• Show that,
1. f is differentiable on R2 .
2. fx and fy exist and continuous on O2 = (x, y) ∈ R2 : x ̸= 0 .
3. fx is not continuous at (0, y) for all y ̸= 0.
29
8.2 Examples
Example 8.2.3
Jf (x, y, z) = Df (x, y, z)
30
Lecture 9
f (a + tu) − f (a)
(Du f ) (a) := lim
t→0 t
the rate of change of f at a in the direction of u, provided the limit exists.
∂f
Note that (Dei f ) (a) = ∂xi (a), i.e., the partial derivatives are directional derivatives along the
standard basis vectors.
ϕ f
t a + tu f (a + tu)
η(t)=f (a+tu)
So,
n
X ∂f
η ′ (0) = ui (a) = (Du f ) (a)
i=1
∂xi
31
9.2 Gradient
9.2 Gradient
which tells us, The steepest slope is achieved when θu ∈ {0, π}, i.e., when u points along or opposite
(∇f ) (a)
to the direction of (∇f ) (a) that means max (Du f ) (a) is attained at u = (direction
∥(∇f ) (a)∥
| {z }
provided ∥(∇f )(a)∦=0
of the steepest slope). Hence, we get the following theorem,
Theorem 9.2.1
Let f : On → R be a differentiable function at a ∈ On and suppose (∇f ) (a) ̸= 0 then the
vector (∇f ) (a) points in the direction of the greatest increment of f at a with the greatest
rate ∥(∇f ) (a)∥
9.3 Examples
Example 9.3.1
Therefore, (Du f ) (1, 1, 0) = − √13 and hence maximum value of (Du f ) (1, 1, 0) is
∥(∇f ) (1, 1, 0)∥ along the unit vector ⟨0, 0, 1⟩.
Example 9.3.2
Take,
x2 y
(
x2 +y 2 (x, y) = (0, 0)
f (x, y) =
0 (x, y) ̸= (0, 0)
x2 y
then |f (x, y) − f (0, 0)| = x2 +y 2 ≤ |y| ≤ ∥(x, y)∥ implies that f is continuous at (0, 0). Now,
32
Lecture 9
f (tu) − 0
(Du f ) (0, 0) = lim
t→0 t
1 tu21 u2
= lim ·
t→0 t 1
2
= u1 u2 ̸= 0 (Because, u is an unit vector)
u1
If we assume f to be differentiable then (∇f ) (0, 0)· = 0 ̸= u21 u2 which is a contradiction!
u2
Example 9.3.2 shows that, the existence of all partial and directional derivatives at a point fails to
imply differentiability at that point.
33
9.3 Examples
34
Lecture 10
f (b) − f (a) = (∇f )(c) · (b − a) = ⟨fx1 (c), . . . , fxn (c)⟩ · ⟨(b1 − a1 ), . . . , (bn − an )⟩
η f
[0, 1] On R
f ◦η
defined by η(t) = (1 − t)a + tb. This function is differentiable, and its derivative is
b1 − a1
η ′ (t) = ...
bn − an
35
10.2 More Partials and Chain Rules
By applying the standard Mean Value Theorem to the composition f ◦ η, there exists t0 ∈ (0, 1) such
that,
f (η(1)) − f (η(0)) = (f ◦ η)′ (t0 ) = Df (η(t0 )) · Dη(t0 )
Expanding the dot product, we have,
b1 − a 1
b2 − a 2
f (b) − f (a) = fx1 (η(t0 )) fx2 (η(t0 )) · · · fxn (η(t0 )) .
..
bn − a n
f (b) − f (a) = ⟨fx1 (η(t0 )), · · · , fxn (η(t0 ))⟩ · ⟨(b1 − a1 ), · · · , (bn − an )⟩
Hence, there exists c = η(t0 ) ∈ La,b such that f (b) − f (a) = (∇f )(c) · (b − a).
f g
On Om Rp
• g ◦ f = ((g ◦ f )1 , (g ◦ f )2 , . . . , (g ◦ f )p )
• g = (g1 , g2 , . . . , gp )
• f = (f1 , f2 , . . . , fm ).
Then, the (i, j)th entry of both sides of (10.1) would become,
∂f1
∂xj(a)
∂f2
∂(g ◦ f )i ∂xj(a)
∂gi ∂gi ∂gi
(a) = ∂y (f (a)) ∂y2 (f (a)) ··· ∂ym (f (a)) .
∂xj ..
1
∂fm
∂xj (a)
36
Lecture 10
m
∂ (g ◦ f )i X ∂gi ∂fk
(a) = (b) · (a)
∂xj ∂yk ∂xj
k=1
where, b = f (a).
m
∂zi X ∂zi ∂yk
(x) = (y) · (x)
∂xj ∂yk ∂xj
k=1
Remark. In addition, the chain rule can be applied in the context of a composition map with respect
to a parameter t. For functions f : On → R, η : O1 → On and z = f ◦ η as shown below,
η f
O1 On R
η f
t (x1 (t), . . . , xn (t)) f (x1 (t), . . . , xn (t))
z=f ◦η
n
df X ∂f dxk
= ·
dt ∂xk dt
k=1
37
10.2 More Partials and Chain Rules
38
Lecture 11
g◦f
On f
Om g Rp
where b = f (a). This can be rewritten in a slightly more suggestive form by introducing new variables,
yk = fk (x1 , . . . , xn )
zi = gi (y1 , . . . , ym )
This form of the chain rule is reminiscent of the one-variable chain rule.
Example 11.1.1
df
Let, f (x, y, z) = xy 2 z and x = t, y = et , z = 1 + t, we want to calculate dt in two ways.
39
11.2 Laplacian
= (2t2 + 4t + 1)e2t
11.2 Laplacian
The Laplacian operator plays a fundamental role in analyzing the behavior of functions and fields in
multidimensional spaces. It quantifies the overall rate of change and spatial variations of a function,
providing valuable insights into its properties and behavior.
Observe that,
n
X ∂2f
∆f =
i=1
∂x2i
∂ ∂ ∂f ∂f
= ,..., . ,...,
∂x1 ∂xn ∂x1 ∂xn
= ∇ · ∇f
∂x ∂x
= cos θ, = −r sin θ
∂r ∂θ
∂y ∂y
= sin θ, = r cos θ
∂r ∂θ
We want to express fxx and fyy in terms of partial derivatives of f in polar coordinates. Notice that,
∂f ∂f ∂x ∂f ∂y
= +
∂r ∂x ∂r ∂y ∂r
40
Lecture 11
∂f ∂f ∂f
i.e., = cos θ + sin θ
∂r ∂x ∂y
Differentiating once more with respect to r, we have,
∂2f
∂ ∂f ∂f
= cos θ + sin θ
∂r2 ∂r ∂x ∂y
∂ ∂f ∂ ∂f
= cos θ + sin θ
∂r ∂x ∂r ∂y
2
∂ 2 f ∂y
2
∂ f ∂x ∂ 2 f ∂y
∂ f ∂x
= cos θ + + sin θ +
∂x2 ∂r ∂y∂x ∂r ∂x∂y ∂r ∂y 2 ∂r
2
∂2f ∂2f ∂2f
∂ f
= cos θ cos θ + sin θ + sin θ cos θ + sin θ
∂x2 ∂y∂x ∂x∂y ∂y 2
= cos θ [cos θfxx + sin θfxy ] + sin θ [cos θfxy + sin θfy ]
Hence, we get,
∂2f
= cos2 θfxx + sin2 θfy + sin 2θfxy
∂r2
∂2f
Similarly, we can find the expression for ∂θ 2 ,
∂2f
= −r (cos θfx + sin θfy ) + r2 sin2 θfxx + r2 cos2 θfy − r2 sin 2θfxy
∂θ2
Combining the above two result we can write,
∂2f 1 ∂f 1 ∂2f
∆f = fxx + fyy = + · + ·
∂r2 r ∂r r2 ∂θ2
41
11.3 Extrema of a function
∂z ∂ ∂z ∂u ∂ ∂z ∂v ∂ ∂z ∂u ∂ ∂z ∂v
= 2y + 2xy + +3 +
∂u ∂u ∂u ∂x ∂v ∂u ∂x ∂u ∂v ∂x ∂v ∂v ∂x
2
∂ 2 z ∂v
2
∂ z ∂u ∂ 2 z ∂v
∂z ∂ z ∂u
= 2y + 2xy + + 3 +
∂u ∂u2 ∂x ∂v∂u ∂x ∂u∂v ∂x ∂v 2 ∂x
∂2z ∂2z ∂2z ∂2z
∂z 1
= 2y + 2xy 2xy 2 + 3 +3 · +3 2
∂u ∂u ∂v∂u 2xy ∂u∂v ∂v
Hence, we get,
∂2z
= 2yzu + 4x2 y 2 zuu + 6xyzuv + 6xyzvu + 9zvv
∂x2
Any point at which f attains a local(global) maxima (or minima) is called extremum point of
that function. In plural, it is called Extrema.
Theorem 11.3.1
Let, f : On → R is differentiable at a ∈ On . If a is a local extremum, then
(∇f ) (a) = 0
∂f
Proof. Fix i ∈ {1, 2, . . . , n}. We want to show ∂x i
= 0. For this set, ϕi : (ai − ϵ, ai + ϵ) → R defined
by
ϕi (t) = f (a1 , . . . , ai−1 , t, ai+1 , . . . , an )
Notice that, dϕ
dt = fxi (a). Since a is local extremum of f , we can say that ai is a local extremum of
i
∂f (a)
ϕi . So, dϕ
dt = 0, which means, ∂xi = 0. We can do this for all i and hence, (∇f )(a) = 0.
i
42
Lecture 11
Question. When we did calculation for local extremum for the functions f with one variable, we
used to evaluate the stationary points by calculating, f ′ (x) = 0. Then we used to check the second
derivative in order to know whether the stationary point is local minima or maxima or saddle point.
For multivariate case also, we need 2nd order derivative to know the behavior of the stationary point.
Now what could be 2nd order total derivative?
Answer. For this purpose we will introduce Hessian Matrix in next class.
43
11.3 Extrema of a function
44
Lecture 12
It is important to note that for any function f that is twice continuously differentiable (f ∈ C 2 ), its
Hessian matrix Hf is symmetric, meaning that Hf = Hf t .
Example 12.1.1
∇f = sin 2x + 2xy, x2 + 2y ∈ R2
Now let’s introduce some notation. Given A = (aij )n×n ∈ Mn (R) and x ∈ Rn we denote QA (x)
by,
45
12.2 Positive Definite, Negative Definite, Semi Definite Matrices
a11 a12 ··· a1n x1
a21
a22 ··· a2n x2
= x1 x2 ··· xn .
.. .. .. ..
.. . . . .
a31 a32 ··· a3n xn
n X
X n
= aij xi xj
i=1 j=1
⟨Ax, x⟩ ≥ 0 ∀ x ∈ Rn \ {0}
Example 12.2.1
1. In is positive definite because for any vector x ∈ Rn \ {0} the inner product
2
⟨In x, x⟩ = ∥x∥ > 0
is strictly positive.
2. For any matrix A ∈ Mn (R), if there exists a matrix B ∈ Mn (R) such that A = B t B,
we can examine the inner product ⟨Ax, x⟩ for any x ∈ Rn \ {0}.
Let’s compute this inner product
⟨Ax, x⟩ = ⟨B t Bx, x⟩
= xt B t Bx
= (Bx)t (Bx)
2
= ∥Bx∥ for all x ∈ Rn \ {0}.
46
Lecture 12
It is worth noting that classifying positive definite matrices becomes more challenging for higher
dimensions (n > 2). However, for 2 × 2 matrices, we can easily determine it from the next theorem.
Theorem 12.2.1
a b
Let A = ∈ M2 (R) be symmetric. Then,
b c
Proof. We have ⟨Ah, h⟩ = ht Ah for any vector h. Now, consider a non-zero vector x = (x1 , x2 ) ∈
R2 \{(0, 0)} with x2 ̸= 0. Without loss of generality, we can scale x as x = (x, 1) for some x ∈ R.
Then, we have,
⟨Ax, x⟩ = ax2 + 2bx + c > 0 ∀ x ∈ R
If x2 = 0, we can choose x = 1 0 (after scaling). Then, ⟨Ax, x⟩ = a. Therefore, we can summarize
the conditions as follows,
A is Positive Definite
47
12.2 Positive Definite, Negative Definite, Semi Definite Matrices
Similarly, we can derive the conditions for negative definite and indefinite matrices as,
A is Negative Definite
⇐⇒ a < 0 and ax2 + bx + c < 0 ∀ x ∈ R
⇐⇒ a < 0 and (2b)2 − 4ac < 0
⇐⇒ a < 0 and ac − b2 > 0
A is Indefinite
⇐⇒ ax2 + bx + c < 0 for some x ∈ R
and ax2 + bx + c > 0 for some x ∈ R
⇐⇒ (2b)2 − 4ac > 0
⇐⇒ ac − b2 < 0
Lemma 12.2.1
a1 (x) a2 (x)
Let, a ∈ On , A(x) = . Suppose, A is continuous at a (i.e., ai ’s are continuous
a2 (x) a3 (x)
at a). Then, A is Positive Definite at a would imply that A is Positive Definite in a
neighborhood of a.
Proof. A(a) is Positive Definite, i.e., a1 (a) > 0 and a1 (a)a3 (a)−a2 2 (a) > 0. As a1 (x) and a1 (x)a3 (x)−
a22 (x) are polynomial of continuous functions, we can find an ϵ > 0 such that both are positive in
Bϵ (a).
48
Lecture 13
f k+1 (c)
ra,k = (x − a)k+1
(k + 1)!
We introduce the following notation for the sake of clarity in the multivariate Taylor expansion. Let
Z
α = (α1 , α2 , · · · , αn ) ∈ n≥0 , and define
Pn
• |α| = i=1 αi
• α! = α1 !α2 ! · · · αn ! (product of coordinate-wise factorials)
∂ |α|
• ∂α = α1
(αth derivative)
∂x1 · · · ∂xn αn
n
X ∂
• ∇·h= hi
i=1
∂xi
The last definition when iterated gives,
X m!
(∇ · h)m = hα ∂ α
α!
|α|=m
49
13.1 Taylor’s Theorem
where,
X 1 α
ra,k (h) = (∂ f )(a + ch)hα for some c ∈ (0, 1)
α!
|α|=k+1
t a + th f (a + th)
η
Claim
= (∇ · h)2 f (a + th)
Proceeding with induction on the order of the derivative, we get η (m) (t) = (∇ · h)m f (a + th) for all
0 ≤ m ≤ k + 1 which is our claim.
with
η ′ (0) η (k) (0)
p0,k (1) = η(0) + + ··· + (13.2)
1! k!
50
Lecture 13
and
η (k+1) (c)
r0,k (c) = (13.3)
(k + 1)!
Substituting η (m) (t) in (13.1) we have,
X 1
f (a + h) = (∂ α f )(a)hα + ra,k (h)
α!
|α|≤k
Then,
(i) f (a) is a local maximum if fxx (a) < 0 and det(Hf (a)) > 0
(ii) f (a) is a local minimum if fxx (a) > 0 and det(Hf (a)) > 0
Find the critical points and discuss the nature of the function
f (x, y) = x3 − 6x2 − 8y 2
Solution. Setting ∇f (x, y) = 0, i.e., (fx , fy )(x, y) = 0, we get the system of equations
3x2 − 12x = 0 and − 16y = 0
whose solution set is {(0, 0), (4, 0)} implying that (0, 0), (4, 0) are critical points.
The 2nd derivatives are,
fxx (x, y) = 6x − 12, fyy (x, y) = −16 and fxy (x, y) = 0
51
13.1 Taylor’s Theorem
Now, we compute the determinant of the hessian at these points to tell their nature. For
(0, 0),
−12 0
det(Hf (0, 0)) = > 0 and fxx (0, 0) = −12 < 0
0 −16
So, f has a local maximum at (0, 0). And at (4, 0),
12 0
det(Hf (4, 0)) = <0
0 −16
52
Lecture 14
Observe that a compact subset of Rn is always closed. To see this, note that every sequence
{xn } ⊆ K, where K is a compact subset of Rn that converges to some x ∈ Rn has a convergent
subsequence {xnk } that converges to the same x. Since K is compact, we can say that x ∈ K. So,
the convergent sequence {xn } converges to a point in K. Hence, K is closed.
More is true. A compact subset of Rn is bounded too. Assume that a compact subset K ⊆ Rn is
not bounded. Note that, a subset of Rn is bounded iff it is contained inside an open ball. Since K
is unbounded, we can get a sequence {xm } ⊆ K with ∥xm ∥ > m, which doesn’t have a convergent
subsequence. This shows that K is not compact that contradicts our assumption.
Therefore, a compact subset of Rn is closed and bounded. What about the converse?
Theorem 14.1.1
A closed and bounded box in Rn is compact.
n
Y
Proof. We take a closed and bounded box K := [ai , bi ] ⊆ Rn . Fix i ∈ [n]. Consider a sequence
i=1
{xm } ⊆ K. We take its projection on the ith coordinate, i.e., {Πi (xm )} ⊆ [ai , bi ]. By Bolzano-
Weierstrass Theorem, it has a convergent subsequence {Πi (xm )t } ⊆ [ai , bi ] converging to αi ∈ [ai , bi ].
\n
Consider the intersection of inverse images of all such convergent subsequences, Π−1
i {Πi (xm )t },
i=1
which is a convergent subsequence of {xm } converging to α = (α1 , . . . , αn ) ∈ K.
Proof. =⇒ Done!
53
14.2 Inverse Function Theorem
⇐= Since, K is bounded, it is contained in a closed box, i.e., there exists r > 0 such that K ⊆ [−r, r]n .
So, Theorem 14.1.1 implies that all sequences in K has a convergent subsequence, which must
converge in K because K is closed. Hence, K is compact.
Theorem 14.1.3
Let f : On → Rm be a continuous map. Then f sends compact sets to compact sets.
Proof. Let K ∈ On be compact. Take a sequence {xk } ⊆ K with a convergent subsequence {xkt } ⊆ K
converging to x ∈ K. Then, {f (xk )} is sequence in f (K) with convergent subsequence {f (xmt )}
converging to f (x). The last statement about convergence follows from the continuity of f . This
shows that f (K) is compact.
Proof. By Theorem 14.1.3, f (K) is compact. So f is bounded which implies sup f, inf f exist! Since
K K
f (K) is closed, they must exist inside f (K).
Lemma 14.2.1
Let On ⊆ Rn be open and convex. Suppose f : On → Rn be a C 1 function. If ∃ M > 0 such
that
∂fi
sup (x) ≤ M for all i, j
x∈On ∂xj
Proof. Pick x, y ∈ On and i ∈ [n]. Then using Mean Value Theorem, we can get ci ∈ Lx,y such that
The last inequality follows from the inequality |xi − yi | ≤ ∥x − y∥ which holds for all i.
54
Lecture 14
i.e., locally, the derivative of the inverse is the matrix inverse of the derivative.
Proof .
We call L = Df (a) which is given to be invertible and take g(x) :=
L−1 f (x). Then,
which contradicts the definition of derivative. Note that det Jf (a) ̸= 0. So, by continuity det Jf (x) ̸= 0
for all x ∈ U (we may shrink U if necessary). Hence, Df (x) is invertible for all x ∈ U . Again by
continuity, for all x ∈ U (we may shrink U if necessary),
∂f ∂f 1
(x) − (a) ≤ 2
∂xj ∂xj 2n
| {z }
δij
Claim
For all x, y ∈ U ,
1
∥f (x) − f (y)∥ ≥ ∥x − y∥
2
55
14.2 Inverse Function Theorem
Where (14.1) follows from the Triangle inequality. So, we get the claim! It shows that f is injective.
Next we look at the compact set ∂U ⊆ U . Since a ̸∈ ∂U , we can say f (x) ̸= f (a) for all x ∈ ∂U . So,
by continuity of f and compactness of ∂U , we can find a d ∈ R≥0 such that
∥f (x) − f (a)∥ ≥ d ∀ x ∈ ∂U
Figure 14.2
Claim
Since, inf U g cannot occur at the boundary ∂U , but must occur in U , there exists x0 ∈ U ◦ such that
∂g
∇g(x0 ) = 0 i.e., (x0 ) = 0 ∀j
∂xj
56
Lecture 14
At x0 , we get,
n
X ∂fi
(yi − fi (x0 )) (x0 ) = 0 ∀j
i=1
∂xj
t
∂fi
=⇒ ∂xj (x0 ) (y − fi (x0 )) = 0
| {z }
Df (x0 )t
Claim
−1
f −1 ≡ (f |V ) : W → V is continuous.
Claim
f −1 is differentiable.
1
f −1 (y0 + h) − f −1 (y0 ) − A−1 h = 0
lim
h→0 ∥h∥
1
f −1 (y) − f −1 (y0 ) − A−1 (y − y0 ) = 0
⇐⇒ lim (14.3)
y→y0 ∥y − y0 ∥
1 −1
ϕ f −1 (y) − f −1 (y0 ) = 0
lim A
h→0 ∥h∥
1
ϕ f −1 (y) − f −1 (y0 ) = 0
⇐⇒ lim
y→y0 ∥y − y0 ∥
57
14.2 Inverse Function Theorem
But,
ϕ f −1 (y) − f −1 (y0 ) ϕ f −1 (y) − f −1 (y0 ) f −1 (y) − f −1 (y0 )
= · =0
∥y − y0 ∥ ∥f −1 (y) − f −1 (y0 )∥ ∥y − y0 ∥
| {z } | {z }
0 ≤2
58
Lecture 15
x = r cos θ y = r sin θ
F : R>0 × R → R2
F (r, θ) = (x, y)
∂x ∂y
= cos θ = sin θ
∂r ∂r
∂x ∂y
= −r sin θ = r cos θ
∂θ ∂θ
cos θ −r sin θ
=⇒ JF (r, θ) = =⇒ det JF (r, θ) = r
sin θ r cos θ
and hence, det JF (r, θ) is non-zero in the domain we chose. The inverse function theorem then
guarantees that we can express (r, θ) as a C 1 function of (x, y), locally. Further, we also have
−1 −1
−1 1 r cos θ r sin θ
DF (x, y) = DF (F (x, y)) =
r − sin θ cos θ
In other words,
∂r ∂r
= cos θ = sin θ
∂x ∂y
∂θ 1 ∂θ 1
= − sin θ = cos θ
∂x r ∂y r
59
15.2 Implicit Function Theorem
In one variable, we could differentiate functions given as y = f (x). We have developed several variable
calculus far enough that we can now do the same for functions given as xn = f (x1 , . . . , xn−1 ). But
what about functions that are given as f (x, y) = 0, that is, implicitly in both variables? We will now
discuss an important theorem that will allow us to take derivatives of such functions as well without
needing to solve the equation for the dependent variable.
Example 15.2.1
∂F
y = f (x) for a differentiable f ⇐⇒ b ̸= 0 ⇐⇒ ̸= 0
∂y
∂F
In other words, ∂y ̸= 0 ⇐⇒ F (x, f (x)) = 0 for some differentiable f .
Example 15.2.2
Note that ∂F ∂F 1
∂y = 2y ̸= 0 for all y ̸= 0. Hence, for F (x0 , y0 ) ̸= 0, ∂y ̸= 0, there exists a C f
defined in a neighbourhood of x0 such that F (x, f (x)) = 0 for all x in that neighbourhood.
Note
Consider F (x, y) = 0 and y = f (x) for some C 1 f such that F (x, f (x)) = 0. We have,
F (x, f (x)) = 0
dF
=⇒ =0
dx
∂F ∂F dy
=⇒ + =0 (Chain rule)
∂x ∂y dx
∂F
dy ∂x
=⇒ = − ∂F
dx ∂y
Hence, the condition ∂F∂y ̸= 0 is necessary for differentiating functions defined implicitly.
Keeping the feeling of the above examples in mind, we now discuss the full theorem that
shows that it is also sufficient.
60
Lecture 15
for some C 1 function g, from the definition of Fe. Now, consider the map
Π 2 : Rn × Rm → Rm
(X, Y ) 7→ Y
Then, Π2 ◦ Fe = F . So we get,
61
15.3 Solving systems of equations
x2 + 2y 2 + z 2 + w = 6
2x3 + 4y 2 + z + w2 = 9
We wish to know whether (z, w) can be expressed as a function of (x, y) locally. Construct
the function F : R4 → R2 as F = (f1 , f2 ) where
f1 (x, y, z, w) = x2 + 2y 2 + z 2 + w − 6
f2 (x, y, z, w) = 2x3 + 4y 2 + z + w2 − 9
Hence, by the implicit function theorem, there is a neighbourhood of (1, −1) such that on
that neighbourhood, (z, w) = f (x, y) for some C 1 f .
62
Lecture 16
We now introduce some notation. Fix i ∈ {1, 2, . . . , n}. We define a partition of the ith interval
[ai , bi ] as
Pi : ai = ai,0 < ai,1 < · · · < ai,ni = bi
and the intervals of its partition as
and set
n
B(t 1 ,t2 ,...,tn )
= Bα = [x1,t1 −1 , x1,t1 ] × · · · × [xn,tn −1 , xn,tn ] = I1,t1 × · · · × In,tn
Note
1. B n = Bαn
S
α∈Λ(P )
2. Vol(B n ) = Vol(Bαn )
P
α∈Λ(P )
Theorem 16.1.1
Let f be a bounded function over B n . Let P, Pe ∈ P(B n ) and Pe ⊃ P . Then
63
16.1 Riemann-Darboux Integration
Proof. Note that L(f, Pe) ≤ U (f, Pe) follows directly from the fact that mα (Pe) ≤ Mα (Pe) ∀Pe ∈ P[a, b],
where mα = inf Bαn f and Mα = supBαn f .
We denote B(A) = {f : A → R | supA |f | < ∞} as the set of all bounded functions over A for any
A ⊆ Rn .
We have L(f, P ) ≤ U (f, P ′ ) for all P, P ′ ∈ P(B n ) by taking the common refinement Pb = P ∪ P ′ .
Hence, Z Z
f≤ f
Bn Bn
Z
At this point, the notation f (x1 , . . . , xn ) dx1 · · · dxn does not indicate repeated integration, but
Bn
we will see that it represents repeated integration for “nice” functions.
64
Lecture 17
Thus,
Z Z
0= f− f= inf U (f, P ) − sup L(f, P )
P ∈P(B n ) P ∈P(B n )
Hence for all ϵ > 0, there exists a partition P ∈ P(B n ) such that U (f, P ) − L(f, P ) < ϵ.
⇐= Conversely, assume that for every ϵ > 0, there exists a partition P ∈ P(B n ) such that
Z Z Z
U (f, P ) − L(f, P ) < ϵ. We want to show that f = f . Since U (f, P ) ≥ f and L(f, P ) ≤
Z
f , it follows that for all ϵ > 0,
Z Z
0≤ f− f <ϵ
Z Z
This implies f= f , showing that f ∈ R(B n ).
65
17.1 Properties of Riemann-Darboux Integration
C(B n ) ⊂ R(B n )
Proof. Let f ∈ C(B n ). Since f is uniformly continuous, for any ϵ > 0, there exists δ > 0 such that
for all x, y ∈ B n with ∥x − y∥ < δ, we have,
ϵ
|f (x) − f (y)| < (17.1)
2 Vol(B n )
| {z }
Call it ϵ̃
Let P ∈ P(B ) be a partition such that ∥P ∥ < δ. For each α ∈ Λ(P ), let aα ∈ Bαn . Then ∥x−aα ∥ < δ
n
for all x ∈ Bαn . It follows from the uniform continuity condition (17.1) that,
Since, (17.2) holds for all α ∈ Λ(P ), aα ∈ Bαn and for all x ∈ Bαn we have,
f (aα ) − ϵ̃ ≤ mα ≤ Mα ≤ f (aα ) + ϵ̃
Thus, U (f, P ) − L(f, P ) < ϵ, and since ϵ is arbitrary, we conclude that f ∈ R(B n ).
Now, let’s consider an important question: Does an analogue of the Fundamental Theorem of Calculus
exist in higher dimensions? Z
In one dimension (n = 1), we have the useful relationship df = f , which aids in computa-
[a,b] ∂[a,b]
tion.
However, this relationship becomes less practical in higher dimensions. For
Z instance, in n = 1, the
X
continuous counterpart to a sum an is the one-dimensional integral f . Similarly, in n = 2,
B1 Z
X
the continuous analogue to a double sum amn is the two-dimensional integral f.
B2
66
Lecture 17
Example 17.2.1
Recall the relation between total limit and iterated limits in double sequence,
Theorem
For a double sequence {amn } if limm,n→∞ amn exists and limm→∞ amn exists for all n, then
lim lim amn = lim amn
n→∞ m→∞ m,n→∞
An important analogue of the above result is Fubini’s Theorem. Computation of the Darboux integral
is typically a challenging task. However, Fubini’s Theorem offers a valuable approach that simplifies
the computation by utilizing iterated integrals.
Visualization
We look at slice functions along each axis, which enables us to simplify computations and apply
Fubini’s Theorem for efficient evaluation of multivariable integrals.
Consider a function f : B 2 → R. For each x ∈ [a1 , b1 ] we define a slice function fx : [a2 , b2 ] → R
given by fx (y) = f (x, y) for all y ∈ [a2 , b2 ]. This function is obtained by fixing x and slicing along
the y-axis at that x-coordinate. Then an iterated integral becomes
Z Z !
fx (y) dy dx
[a1 ,b1 ] [a2 ,b2 ]
The question arise whether this quantity is invariant under the interchange of x and y, i.e., we may
slice f along x−axis at y to obtain fy : [a1 , b1 ] → R given by fy (x) = f (x, y) for every x ∈ [a1 , b1 ]
and want to investigate the equality of
Z Z ! Z Z ! Z
? ?
fx (y) dy dx = fy (x) dx dy = f (17.3)
[a1 ,b1 ] [a2 ,b2 ] [a2 ,b2 ] [a1 ,b1 ] B2
67
17.2 Iterated Integrals
In this context, we observe that a partition P ∈ P(B 2 ) can be decomposed into the partitions of the
individual coordinates. Specifically, we have P = P1 × P2 for the two coordinate intervals [a1 , b1 ] and
[a2 , b2 ], and the corresponding indexing sets satisfy Λ(P ) = Λ(P1 ) × Λ(P2 ).
Q
(
1 if x = 12 , y ∈ ∩ [0, 1]
f (x, y) =
0 otherwise
Qc ∩ [0, 1]
(
1 1 if y ∈
fx ≡ 0 for all x ̸=
2
and f 12 (y) =
0 if y ∈ Q ∩ [0, 1]
Dirichlet Function
and y−slice,
(
For y ∈ Q ∩ [0, 1], fy (x) =
1 if x = 21
0 otherwise
For y ∈ Qc ∩ [0, 1], fy ≡ 0
[0, 1]. Hence, fx (y) dy dx doesn’t exist. Yet f ∈ R(B 2 ). To see this, we fix ϵ > 0
I I
and consider the partition P = P1 × P2 where,
(
P1 : 0 < 12 − ϵ < 1
2 +ϵ<1
P2 : 0 < 1
( )
1 1 1 1
So, P = 0, − ϵ × I , − ϵ, + ϵ × I , − ϵ, 1 × I .
2 2 2 2
| {z } | {z } | {z }
Bα1 Bα2 Bα3
Then mα1 = mα2 = mα3 = 0, Mα1 = Mα3 = 0 and Mα2 = 1, which implies U (f, P ) −
L(f, P ) = 2ϵ < 3ϵ. This shows that f ∈ R(B 2 ). Again L(f, P ) = 0 for all P ∈ P(B 2 ) and
hence, Z
f =0
B2
68
Lecture 18
Throughout this section, we take x ∈ B m and y ∈ B n to represent the point (x, y) ∈ B m+n . For a
bounded function f ∈ B(B m+n ), we define the slice functions
It is worth noting that fx ∈ B(B n ) and fy ∈ B(B m ). For a fixed x ∈ B m , we can compute the
lower and upper integrals of fx over B n , denoted as f (x) and f (x) respectively. Similarly, we can
compute the lower and upper integrals of fy over B m for fixed y ∈ B m . These are given by,
Z Z
f (x) = fx (y) dV (y) and f (x) = fx (y) dV (y)
Bn Bn
69
18.1 Fubini’s Theorem
Corollary
Corollary
If f ∈ C(B n ), then all possible slice functions are continuous and hence Riemann Integrable.
Thus, multidimensional integral becomes the iterated one-dimensional integrals,
Z Z Z Z Z
f= ··· f dx1 dx2 · · · dxn−1 dxn
Bn
Proof of the Fubini’s Theorem. Let P = P m × P n be a partition of B m+n . Then, we can express
the lower sum L(f, P ) as follows,
X
m+n
L(f, P ) = mα(P ) Vol Bα(P )
α(P )∈Λ(P )
X X
n m
= m(α(P m ),α(P n )) Vol Bα(P n) Vol Bα(P ) m
For each x ∈ B m and α(P n ) ∈ Λ(P n ), let mα(P n ) (x) = inf y∈Bα(P
n
n)
fx (y). It follows that mα(P n ) (x) ≥
m
m(α(P m ),α(P n )) for every x ∈ Bα(P m ) . Consequently, we have,
X
n
lP m ≤ mα(P n ) Vol Bα(P n)
α(P n )∈Λ(P n )
Z
= L(fx , P n ) ≤ fx
Bn
m
Taking infimum over all x ∈ Bα(P m ) , we obtain,
Z
lP m ≤ inf
m
fx
x∈Bα(P m )
Bn
70
Lecture 18
= m
inf f (x) = mα(P m )
x∈Bα(P m)
Question: Will the function be Riemann integrable if all the slices are Riemann integrable?
We will address this question later. In the meantime, let’s conclude this lecture with an example.
Example 18.1.1
71
18.1 Fubini’s Theorem
72
Lecture 19
Definition 19.1.1
Given Ω bounded, let B n ⊇ Ω. For f ∈ B(Ω), define
(
f (x) if x ∈ Ω
f˜B n (x) =
0 if x ∈ B n \ Ω
An immediate question that arises now is: If B n ⊇ Ω and Bˆn ⊇ Ω then will it be true that
It is in fact true that (I) =⇒ (II), but we won’t cover the proof here.
Definition 19.1.2
Z
Let f ∈ B(Ω). We say that f ∈ R(Ω) if f˜B n exists for some B n ⊇ Ω, and in this case we
Bn
define Z Z
f := f˜B n
Ω Bn
For example a line segment in Rn is of content zero, provided n > 1. We then have the following
theorems:
73
19.2 Two Elementary Regions
Theorem 19.1.1
(i) Let f ∈ B(B n ) and let D = {x ∈ B n | f is not continuous at x} be the set of disconti-
nuities of f , if D is of content zero, then f ∈ R(B n ).
(ii) If S is a content zero set then int (S) = ∅.
Now that we know how to integrate on arbitrary domains, the next question that comes to our mind
is, does there exist a Fubini’s theorem for integration over arbitrary sets? Before that we define
elementary regions.
Similarly a set Ω ⊆ R2 is x-simple/type II if there exists functions ψ1 , ψ2 ∈ B([c, d]) such that
0
x
74
Lecture 19
75
19.2 Two Elementary Regions
76
Lecture 20
In the previous lecture we extended integration over boxes to over what we called elementary regions.
This lecture explores the extention of Fubini’s theorem to integration over such sets, and talk about
some applications thereof. Towards the end, we discuss the celebrated change of variables formula of
multivariable calculus.
Theorem 20.1.1
Let f ∈ R(Ω) where Ω ⊆ R2 is a bounded elementary domain.
Z φ2 (x)
(1) If Ω = {(x, y) | a ≤ x ≤ b, and φ1 (x) ≤ y ≤ φ2 (x)} and if f (x, y) dy exists for
φ1 (x)
all x ∈ [a, b] then !
Z Z b Z φ2 (x)
f dA = f (x, y) dy dx
Ω a φ1 (x)
when Ω is x-simple.
Proof. There exists c, d ∈ R such that Ω ⊆ [a, b] × [c, d] = B 2 . We know f˜ ∈ R(B 2 ) where
(
˜ f (x, y) if (x, y) ∈ Ω
f (x, y) =
0 if (x, y) ∈ B 2 \ Ω
Z φ2 (x)
Since f˜ ∈ R(B 2 ), and since f (x, y) dy exists for fixed x, hence
φ1 (x)
77
20.1 Fubini’s Theorem on Elementary Regions
Then we get
! !
Z b Z d Z b Z φ2 (x)
=⇒ f˜(x, y) dy dx = f (x, y) dy dx
a c a φ1 (x)
!
Z Z b Z φ2 (x)
(∗)
=⇒ f˜ dA = f (x, y) dy dx
B2 a φ1 (x)
!
Z Z b Z φ2 (x)
=⇒ f dA = f (x, y) dy dx
Ω a φ1 (x)
where (∗) follows from Theorem 18.1.1. An analogous argument works for the x-simple regions.
Example 20.1.1
Example 20.1.2
Z
2
Let B = [0, π] × [− π2 , π2 ], and we want to evaluate the integral sin(x + y) dA.
B2
Z Z Z
sin(x + y) dA = sin x cos y dA + sin y cos x dA
B2 B2 B2
Z π ! Z π
! Z
2 π Z 2 π
= cos y dy sin x dx + sin y dy cos x dx
−π
2 0 −π
2 0
=4
Example 20.1.3
Z
Let Ω be the region bounded by y = 1 and y = x2 , and we want to find x2 y dV . We can
Ω
write Ω as a y-simple region as follows:
Ω = {(x, y) | −1 ≤ x ≤ 1, and x2 ≤ y ≤ 1}
78
Lecture 20
Example 20.1.4
Z
Compute f dA where f : [0, 1]2 → R is given by
[0,1]2
(
x if y ≤ x2
f (x, y) = ∀ (x, y) ∈ [0, 1]2
y if y > x2
Then f |Ω1 and f |Ω2 are both Riemann integrable, and while f |y=x2 is not continuous, the
set {(x, x2 ) | 0 ≤ x ≤ 1} is of content zero. Hence f is integrable, and we can make sense of
writing the given integral as a sum
Z Z Z
f dA = f dA + f dA
[0,1]2 Ω1 Ω2
Example 20.1.5
Z π Z π
2 sin y
2
Compute dy dx using Fubini’s theorem.
0 x y
We consider the region Ω = (x, y) | 0 ≤ x ≤ π2 , x ≤ y ≤ π2 ⊆ R2 can be written as a x-simple
region as follows: n π o
Ω = (x, y) | 0 ≤ y ≤ , 0 ≤ x ≤ y
2
This shows that
Z Z π2 Z π2
sin y sin y
dA = dy dx
Ω y 0 x y
Z π2 Z y
sin y
= dx dy
0 0 y
Z π2
= sin y dy
0
=1
79
20.2 Change of Variables
Here we effectively compensate for the change of variable by introducing the scale change factor of
φ′ . As we have seen, the scale change factor at a point for a transformation on Rn is given by the
detereminant of the Jacobian matrix at that point. Thus, this theorem has the following natural
extension to Rn :
Although it is not too hard to get a feel for the theorem from its applications, the proof is quite long
and technical, and thus omitted. We recommend the interested and daring readers to have a look
at page 67 of Calculus on Manifolds by Michael Spivak. In the next lecture, we will discuss some
applications of this result.
80
Lecture 21
In the previous lecture, we extended our notion of Riemann integration over boxes to elementary
regions. Additionally, we had discussed the change of variable formula for multivariable calculus
(Theorem 20.2.2). We now try to motivate the use of the same with some rather important applications.
81
21.1 Change of Variables (Continued)
Example 21.1.2
n 2 2
o
Compute the area of Ω = (x, y) | x 3 + y 3 < 1 . Consider the function
then clearly φ([0, 1] × [0, 2π]) = Ω. Also φ is injective and C 1 , but we have
Next up, consider a change of coordinates in R3 , from Cartesian system to the spherical co-ordinate
system. Just like the previous case, this comes in handy when dealing with functions and regions
which are spherically symmetric. One canonical example may be its use in the theory of central
forces in physics.
82
Lecture 21
φ : O3 → R3 , φ(r, ϕ, θ) = (x, y, z)
This gives det(Jφ (r, ϕ, θ)) = r2 sin ϕ, which is non-vanishing in the given domain. As φ
is injective and C 1 , we can use Theorem 20.2.2 to transform from Cartesian to spherical
coordinates. We provide a simple example for some clarity.
We leave it to the reader to do a similar analysis for the cylindrical co-ordinate system. This formula
also finds extensive use in probability theory, where it is commonly referred to as the change of
density formula (see, for instance A First Course in Probability by Sheldon Ross, or any introductory
probability book for that matter). Hopefully, we have demostrated to the reader the central role
Theorem 20.2.2 plays in analysis of several variables, enough to convince him to actually read the
proof! We will now depart from the general study of functions and integrals in Rn , and delve into
the theory of curves and surfaces, dealing primarily with R3 .
83
21.1 Change of Variables (Continued)
84
Lecture 22
Example 22.1.1
1. Let γ : [0, 1] → R2 be γ(t) = (1 − 2t, 2 + t). Clearly γ is C 1 and γ ′ (t) = (−2, 1) ̸= 0 and
thus γ is smooth.
2. γ : [0, 2π] → R2 given by γ : t 7→ (r cos t, r sin t) where r > 0 is constant. Then
γ ′ (t) = (−r sin t, r cos t) ̸= 0 ∀ t ∈ [0, 2π], thus γ is smooth.
γ : [0, nπ] → R3
t 7→ (r cos t, r sin t, ct)
Then γ ′ (t) = (−r sin t, r cos t, c) ̸= 0 and hence γ is smooth. The path of γ is called a
helix.
4. γ : [−1, 1] → R2 given by γ(t) = (|t|, t), then γ is not C 1 , hence it is not smooth.
5. γ : [0, 1] → R2 given by γ(t) = (0, t2 ), then even though γ is C 1 , but it is not smooth as
γ ′ (0) = 0.
6. γ : [0, 2π] → R2 given by γ : t 7→ (r cos t, r sin t), then path of γ is given by
85
22.1 Curves and Surfaces
[a, b]
Definition 22.1.4
Let γ : [a, b] → Rn be a C 1 curve, then
(i) ∥γ ′ (t)∥ := speed of γ at time t.
Z b
(ii) ∥γ ′ (t)∥dt := arc length of γ.
a
Let’s try to look at more natural how equation (ii) in the previous definition gives us the arc length
of a curve γ.
Let γ : [a, b] → Rn and let P := a = t0 < t1 < · · · < tm = b be a partition of the interval [a, b]. Now
define
Xm
ℓ(γ, P) = ∥γ(ti−1 ) − γ(ti )∥
i=1
Definition 22.1.5
A curve γ : [a, b] → Rn is rectifiable or said to have arc length if
which is equivalent to saying that for all ε > 0, there exists δ > 0, such that
Theorem 22.1.1
Z b
For a piecewise smooth curve γ : [a, b] → Rn it is rectifiable and ℓ(γ) = ∥γ ′ (t)∥dt.
a
Remark. Rectifiable curve ̸⇒ piecewise smooth, counter example: Cantor’s function (popularly
called the Devil’s staircase).
86
Lecture 22
Theorem 22.1.2
Let γ : [a, b] → Rn be a rectifiable parametrized curve and let γ̃ = γ ◦ φ, where φ is a strictly
increasing surjective and continuous function, then γ̃ is rectifiable and ℓ(γ) = ℓ(γ̃).
Theorem 22.1.3
Let γ : [a, b] → Rn be a smooth curve, then there exists a parametrization φ such that
∥γ̃ ′ (s)∥ = 1 for all s ∈ [c, d].
87
22.1 Curves and Surfaces
88
Lecture 23
89
23.1 Line Integrals
Z
Question. Given a scalar field f : On → R and γ ≡ C be a curve, we want to define f . But
C
exactly how we can do this?
Answer. C is a curve, so it is bounded subset of Rn . How about thinking of Riemann Integration?
For n ≥ 2, C is content zero in Rn . This does not make any sense! The right way is as following.
Let, γ : [a, b] → Rn be a smooth curve (or piecewise smooth) and C := ran(γ) (on other words path
of γ). Let, f ∈ B(C). Given P ∈ P[a, b], P : a = t0 < t1 < · · · < tm = b.
m
X
U (f, P) = Mi · si
i=1
Xm
L(f, P) = mi · si
i=1
Figure 23.1: Curve C
The above expressions are same as upper and
lower Riemann sum respectively. This opens up
“The Pandora’s box!”.
Theorem 23.1.1
Let, γ be a “Rectifiable” smooth(or piecewise smooth) and C = ran(γ) and f ∈ B(C). Then,
1. f ∈ C 0 (C) =⇒ f ∈ R(C)
2.
m
X Z
f ∈ R(C) ⇐⇒ lim f (ζi )si exist and equal to f.
||P||→0 C
i=1
Proof. Exercise.
90
Lecture 23
We have resolved the problems for Scalar field. What about vector fields?
Suppose a particle moves a distance d under a constant force F , then work done by the force is
⃗
F d cos θ = F⃗ · d.
If the force was not constant throughout the path
then how can we calculate work done by that
force? Consider the case where F is the vector
field (Force in this case) defined over a curve
(path)γ. Here, γ : [a, b] → Rn and C = ran(γ).
So, work done throughout the whole path will
be, Z
F⃗ · d⃗r
C
Z b
Which is equal to, F⃗ (γ(t)) · ∇γ(t) dt
a Figure 23.3: F⃗ is the vector field over the curve γ
Now we will look into some examples.
Example 23.1.1
91
23.1 Line Integrals
Z 1
=⇒ Work done = (−t5 , t6 , −t7 ) · (2t, −3t2 , 4t3 ) dt
0
31
=−
88
Z
F⃗ · d⃗r = lim R(F, P) (if the limit exists)
C ∥P∥→0
92
Lecture 24
Theorem 24.0.2
Let, f : On → R be a C 1 function and γ be a piecewise smooth C 1 curve on On , joining two
points A and B. Then Z
∇f · d⃗r = f (B) − f (A)
c
Which means the above line integral is “independent of parametrization”.
Proof .
Let, r be a parametrization of the curve γ and r : [a, b] → On such that
r(a) = A and r(b) = B.
Evaluating the LHS gives,
Z Z b
(23.1)
∇f · d⃗r = ∇f (r(t)) · r′ (t) dt
c a
n
!
Z b X ∂f ′
= × ri (t) dt
a i=1
∂ri
Z b Figure 24.1: Paths from
d
= (f (r(t))) dt A to B
a dt
= f (r(b)) − f (r(a))
= f (B) − f (A)
“Work done by a conservative force always depends on the starting point and the end
point, not on the path followed by the particle.”
We know if the force is conservative then we can define potential energy U as F⃗ = −∇U . Work done
by the force is simply the change of potential.
93
24.2 Surface and Surface Integrals
Equation of the plane: Consider an arbitrary point P⃗ = ⟨x, y, z⟩ on the plane then,
⟨x − x0 , y − y0 , z − z0 ⟩ · ⟨a, b, c⟩ = 0
∴ Q⃗1 × Q⃗2 = cN
⃗
n o
So, P⃗0 + r1 Q⃗1 + r2 Q⃗2 | r1 , r2 ∈ R describes the same plane as (24.1).
η
Let, η be a map defined on (−ε, ε) that maps t 7−→ r(uo = 0 + t, v0 ). Clearly, η defines smooth curve
η̃
on S. Similarly, we can define η̃ on (−ε, ε) that maps t 7−→ r(u0 , v0 + t).
Thus, η̃ also defines a smooth curve on S.
From the Figure 24.2 we can see η and η̃ are
the curves. ru (u0 , v0 ) = dη
dt t=0 , which gives the
tangent of the curve η at point (u0 , v0 ) along x
axis.
Similarly, for η̃, rv (u0 , v0 ) gives the tangent of
η̃ along y axis. The vectors ru (u0 , v0 ), rv (u0 , v0 )
spanned together to form a plane. This plane is
known as Tangent Plane.
Since r is C 1 , both r⃗u and r⃗v are continuous
and hence r⃗u × r⃗v is continuous. Also, r⃗u × r⃗v is
along the normal vector of S at r(u0 , v0 ) which
follows from the previous section.
94
Lecture 24
24.3 Examples
Let f : O2 → R be a C 1 function then the graph of f is G(f ) = {(x, y, f (x, y)) : (x, y) ∈ O2 }.
Under the conditions O2 is bounded and partial derivatives of f is bounded, we want to find
a parametrization of this Surface G(f ).
Answer. Here, we use the trivial parametrization r : O2 → R3 that is r(x, y) = (x, y, f (x, y)).
Clearly, r is one - one. Now, ru (u, v) = (1, 0, fu ) and rv (u, v) = (0, 1, fv ).
So, ru × rv = (−fu , −fv , 1) ̸= ⃗0. So, it is a parametrization of the surface.
95
24.3 Examples
Let f, g be C 1 functions on [0, b]. Consider the curve t 7−→ (0, f (t), g(t)) is a C 1 curve. If we
rotate the curve with respect to z axis, we must get a surface. Parametrization of the surface
is given by,
r(u, v) = (f (u) cos v, f (u) sin v, g(u)) ; u ∈ [0, b], v ∈ [0, 2π]
96
Lecture 25
2x
Example 25.1.1 (Equation of Tangent and Normal to z = f (x, y) = y − y 2 at (1, 1, 1))
1.99
Example 25.1.2 (Use Tangent Plane to approximate (1.99)2 − 1.01 )
Solution. Consider z = x2 − xy = f (x, y). This describes a surface. Now consider P = (2, 1, 2)
be the point on the surface. Here,⟨−fx , −fy , 1⟩ = ⟨−3, −2, 1⟩. So, equation of Tangent plane
at P is,
z = 2 + 3(x − 2) + 2(y − 1)
The given expression can be approximated as, (by putting value of x, y in the above equation
of tangent plane) z(1.99, 1.01) ≈ 1.99. ■
Our next goal is to calculate area of different surfaces. We will start with very basic example, that is,
area of a plane.
97
25.2 Surface Area
−−→
OP0 = ⟨a0 , b0 , c0 ⟩
−−→
OP1 = ⟨a1 , b1 , c1 ⟩
−−→
OP2 = ⟨a2 , b2 , c2 ⟩
= ∥rx × ry ∥ × Area(Bα )2
98
Lecture 25
Will will do the above integration over any bounded set Ω as we have done in Riemann integration
chapter. Over a bounded set Ω area of S is given by,
Z q
Area(S) = 1 + fx2 + fy2 dA (25.3)
Ω
The General method for finding surface area is described by the following theorem.
Theorem 25.2.1
Let R ⊆ R2 be a region. r : R → R3 be the parametrization of the surface S. Then,
Z
Area(S) = ∥ru × rv ∥ dA
R
99
25.2 Surface Area
100
Lecture 26
26.1 Examples
Recall in previous lecture, we discussed how to find area of a surface over a bounded region. Now we
will look towards some examples.
101
26.2 Surface Integral over Scalar fields
Z X
f dS := lim f (r(xα ))∥ru (xα ) × rv (xα )∥ Area(Bα2 )
S ∥P∥→0
α∈Λ(P)
Z
= (f ◦ r)(u, v)∥ru × rv ∥ dA (26.1)
R
As we have noticed in the case of Line Integral of over a scalar field the integration do not
depend on the parametrization of the path. In this case also if we have different two parametrization
r, r̃ for the surface S then there exist a one-one continuous function φ so that the following diagram
commutes, (in other words r = r̃ ◦ φ̃)
S
r r̃
R φ R̃
We can show that the integration in (26.1) is also same if we replace r by r̃. In other words the
surface Integral over a surface is independent of its parametrization (See Figure 26.3).
We want to evaluate Z
(x2 + y 2 + z 2 )dS
S
n p o
where, S = (x, y, z) | z = x2 + y 2 , 0 ≤ z ≤ 1
p
Solution. Let,R = (x, y)|x2 + y 2 ≤ 1 be the region and r(x, y) = x, y, x2 + y 2 is the
x y
parametrization of the surface S. Here, rx = 1, 0, √ 2 2 and ry = 0, 1, √ 2 2 .
x +y x +y
Z Z
∴ (x2 + y 2 + z 2 ) dS = (f ◦ r) · ∥rx × ry ∥ dA
S R
s
x2 y2
Z
= (f ◦ r) 1 + 2 2
+ 2 dA
R x +y x + y2
√ Z
= 2 2 (x2 + y 2 ) dA
R
√ Z 1 Z 2π 3
=2 2 r dθ dr (It’s the polar substitution)
0 0
√
= 2π
102
Lecture 26
So we formalize the notion of “Orientation” with the following definition. Thought it is only for the
surfaces in R3 . Later in Differential Geometry, you will see this notion in a more general way (for
manifolds).
103
26.3 Surface Integral over a Vector field
104
Lecture 27
Example 27.1.1
⃗r : O2 → R3
(x, y) 7→ (x, y, f (x, y))
Then we have
⃗rx × ⃗ry
⃗n(x, y) =
∥⃗rx × ⃗ry ∥
Unless otherwise mentioned this will be our standard orientation of the normal vector field.
Z
Usually computation of F⃗ · dS
⃗ is complicated, let us look at some examples to gain more familiarity.
S
105
27.1 Conservative Vector Fields
Example 27.1.2
Z
We already know that ∇f · dr = f (B) − f (A), now a natural question that arises is
C
Theorem 27.1.1
Let F⃗ be a vector field over On , the following are equivalent:
1. F⃗ is conservative.
Z
2. F⃗ · dr = 0, for all closed and piecewise smooth curve C.
C
Z Z
3. F⃗ · dr = F⃗ · dr, for all curves C1 and C2 with same initial and end points.
C1 C2
∂f
fx ≡ =P
∂x
∂f
fy ≡ =Q (27.1)
∂y
∂f
fz ≡ =R
∂z
106
Lecture 27
î ĵ k̂
∇ × F⃗ := ∂
∂x
∂
∂y
∂
∂z
P Q R
Then expanding this out and using the relations (27.1) and others we get that ∇ × F⃗ = 0. So, we
have proved that if F⃗ is conservative then ∇ × F⃗ = 0.
Remark. Thus, a necessary condition for a vector field to be conservative is that, its curl should be
the zero vector field.
Example 27.1.3
f (x, y) = xy − 3x + g(y)
∂f
But then using ∂y = x + 2 we get
∂f
x + g ′ (y) = = x + 2 ⇒ g ′ (y) = 2
∂y
Therefore taking f (x, y) = xy − 3x + 2y gives us a potential function for the vector field F⃗ .
Remark. This approach works for all F⃗ such that ∇ × F⃗ = 0 and the domain is convex.
Example 27.1.4
−y
Let F⃗ (x, y) = x2 +y 2
x
, x2 +y 2 = (P, Q) (say) on R2 \ {0}. Then we have ∂P
∂y = ∂Q
∂x , but we
will show that F⃗ is not conservative. Consider the curve
then
Z Z 2π
F⃗ · dr = F⃗ (γ(t)) · γ ′ (t) dt
C 0
Z 2π
= (− sin t, cos t) · (− sin t, cos t) dt
0
Z 2π
= dt
0
= 2π
Z
But C is clearly a closed curve, hence by Theorem 27.1.1 we must have F⃗ · dr = 0.
C
(Contradiction!)
107
27.2 Green’s Theorem
Figure 27.3: Examples of simply connected and not simply connected region
Figure 27.4: You break up the region C with the hole into two regions without holes C1 and C2 .
108
Lecture 27
Z Z Z
P dx + Q dy = P dx + Q dy + P dx + Q dy
C C̃1 C̃2
Z Z
= (Qx − Py ) dA + (Qx − Py ) dA
R1 R2
Z
= (Qx − Py ) dA
R
109
27.2 Green’s Theorem
110
Lecture 28
Proof .
(for Simple region)
Let R = {(x, y) | a ≤ x ≤ b, φ1 (x) ≤ y ≤ φ2 (x)}
be a simple region. Here C = C1 ∪ V2 ∪ C2 ∪ V1
is the curve bounding the region along anti-
clockwise direction (as shown in Figure 28.1).
Now,
Z Z b Z φ2 (x)
∂P ∂P
− dA = − dy dx
R ∂y a φ1 (x) ∂y
Z b
=− (P (x, φ2 (x)) − P (x, φ1 (x))) dx
a
Figure 28.1: A simple region
111
28.1 Green’s Theorem
Z Z
∂P
=⇒ P dx = − dA
C R ∂y
Z Z
∂Q
By similar mechanism we can show Q dy = dA. The rest follows from here.
C R ∂y
Example 28.1.1
Z ZZ
P dx + Q dy = (2y + 2y) dA (Green’s Theorem)
C [0,1]2
Z 1Z 1
= 4y dy dx
0 0
=2
Figure 28.2: ∂([0, 1]2 )
If we try to calculate the integral directly, we will end up getting
same result.
Area of a closed Region. Let R (simply connected) be a closed region and C = ∂R be the
curve enclosing the region. Using Green’s Theorem we get,
xdy − ydx
Z Z Z Z
Area(R) = dA = xdy = −ydx =
R C C C 2
x2 y2
Example 28.1.2 (Area inside the ellipse: a2 + b2 = 1)
Solution. Parametrization of ellipse x = a cos t, y = b sin t where t ∈ [0, 2π). Using the above
application of Green’s Theorem we can write,
Z Z 2π
Area = xdy = ab cos2 tdt = πab
C 0
Proof. Let D be an open and connected domain. F⃗ = ⟨P, Q⟩ is defined over D. Also let P0 = ⟨x0 , y0 ⟩
be a fixed point in the domain D and P1 = ⟨x, y⟩ ∈ D be a variable point. C be a smooth curve
joining P0 and P1 . Define
Z
φ(x, y) = F⃗ · d⃗r
C
112
Lecture 28
Since, D is open set, so we must get an open ball centered at P1 contained in D. Take a point
P1′ = ⟨x1 , y⟩ inside that open ball such that x1 < x. Let C1 be a smooth curve from P0 to P1 and C2
be a line segment from P1′ to P1 . So, C1 ∪ C2 defines a smooth curve from P0 to P1 .
C2
P1′ (x1 , y) P1 (x, y)
C1
C
P0 (x0 , y0 )
Z
As F⃗ · d⃗r is path independent We can write,
C
Z Z
φ(x, y) = F⃗ · d⃗r + F⃗ · d⃗r
C1 C2
Now we take the partial derivative of both sides of this equation with respect to x. The first integral
does not depend on the variable x since C1 is the path from P0 (x0 , y0 , z0 ) to P1′ (x1 , y, z) and so
partial differentiating this line integral with respect to x is zero.
Z Z
∂φ ∂ ⃗ ⃗
= F · d⃗r + F · d⃗r
∂x ∂x C C
Z 1 2 Z
∂ ⃗ ∂ ⃗
= F · d⃗r + F · d⃗r
∂x C1 ∂x C2
| {z }
=0
Z Z x
∂ ∂
F⃗ · d⃗r = ⟨P (t, y), Q(t, y)⟩ · ⟨1, 0⟩ dt
∂x C2 ∂x x
Z x1
∂
= P (t, y) dt
∂x x1
= P (x, y) [Fundamental Theorem of calculus]
Theorem 28.1.3
Let D be a simply connected domain in R2 and F⃗ is a C 1 vector field on D. Then F⃗ is
conservative iff ∇ × F⃗ = 0 on D.
113
28.2 Gauss Divergence Theorem
Just like FTC, the behavior over a volume is fully determined by the behavior at the boundary.
Proof of this theorem is beyond our reach. But we can see the proof for simple cases.
Proof. (For a simple case) Consider D = {(x, y, z) | φ1 (x, y) ≤ z ≤ φ2 (x, y), (x, y) ∈ [a, b] × [c, d]}.
(Exercise.) Complete the proof!
Example 28.2.1
Solution. Notice that S is open surface. We want to use Gauss Theorem 28.2.1. So we need a
close surface. Let S1 be the surface x2 + y 2 ≤ 1. Then S ⊔ S1 is a closed surface.
Z Z
F⃗ · dS
⃗= ∇ · F⃗ dV
S⊔S1 x2 +y 2 ,z 2 ≤1,z≥0
Z
= dV
x2 +y 2 ,z 2 ≤1,z≥0
2π
=
3
Parametrization of the surface S1 = (x, y, 0) | x2 + y 2 = 1 . So, rx × ry = ⟨1, 0, 0⟩ × ⟨0, 1, 0⟩.
Z Z Z
F⃗ · dS
⃗= x + y, z , x 2 2
· ⟨0, 0, 1⟩ dA = x2 dA
S1 x2 +y 2 ≤1 x2 +y 2 ≤1
Z 2π Z 1
π
= r3 cos2 θ dr dθ =
0 0 4
Z
⇒ ⃗ = 11π
F⃗ · dS
S 12
114
Lecture 28
Example 28.3.1
Z
Compute F⃗ · d⃗r, where C : x2 + y 2 = 9, z = 4 and F⃗ = ⟨−y, x, xyz⟩.
C
Z Z
F⃗ · d⃗r = (∇ × F⃗ ) · dS
⃗
C S
Z
= (∇ × F⃗ ) · ⟨0, 0, −1⟩ dA
x2 +y 2 ≤1,z=4
Z
= −2 dA
x2 +y 2 ≤1,z=4
= −18π
Stoke’s Theorem is the R3 −analogue of Green’s Theorem 28.1.1. If we take the third component of
F⃗ to be zero, i.e., R = 0, then Stoke’s Theorem 28.3.1 gives us back Green’s Theorem 28.1.1.
There is a generalized version of Stokes’ theorem. Just for information the theorem is stated
below.
• If Ω is an oriented n-manifold (with boundary) and ω is a differential form ((n − 1) form).
Then integral of ω over the boundary ∂Ω of the manifold Ω is given by,
Z Z
ω= dω
∂Ω Ω
115
28.3 Stokes’ Theorem
116
Index
L Lipschitz 3
Linear map 2
117