0% found this document useful (0 votes)
202 views15 pages

Problems On Sums and Integrals

This document discusses techniques for manipulating sums and integrals, including: 1) Swapping the order of finite sums, which is permissible if the sums converge absolutely but not conditionally. 2) Extending this idea to infinite sums and integrals, requiring absolute convergence for rearranging orders. 3) Examples of applying these techniques, such as using divisor sums to evaluate a specific sum, and theorems governing double sums.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
202 views15 pages

Problems On Sums and Integrals

This document discusses techniques for manipulating sums and integrals, including: 1) Swapping the order of finite sums, which is permissible if the sums converge absolutely but not conditionally. 2) Extending this idea to infinite sums and integrals, requiring absolute convergence for rearranging orders. 3) Examples of applying these techniques, such as using divisor sums to evaluate a specific sum, and theorems governing double sums.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 15

PROBLEMS ON SUMS AND INTEGRALS

Contents
1 Swapping sums 1
1.1 Finite sums . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Absolute and conditional convergence . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2 Riemann integral 4
2.1 Compactness and improper integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Mesh sums . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3 Discretization and inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

3 Lebesgue integrals 5
3.1 Advantages of Lebesgue integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.2 Riemann integrals and Lebesgue integrals . . . . . . . . . . . . . . . . . . . . . . . . 6
3.3 Swapping double integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.4 Interchanging limits and Lebesgue integrals . . . . . . . . . . . . . . . . . . . . . . . 7

4 Techniques for introducing more sums 7


4.1 Taylor series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4.2 Eliminating fractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.3 Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

5 Problems 11

1 Swapping sums
1.1 Finite sums
Here is an example that many of you might already know from high school math contests.
Example 1. Let n be a positive integer. Prove that
X jnk 1
ϕ(k) = n(n + 1).
k 2
k≥1

Proof. The key idea is to rewrite the floor as a sum involving divisors:
X jnk X X XX
ϕ(k) = ϕ(k) 1= ϕ(k).
k
k≥1 k≥1 k|m k≥1 k|m
k≤n m≤n

Thus we’re computing the sum of ϕ(k) over several pairs of integers (k, m) for which k | m, m ≤ n.
For example, if n = 6, the possible pairs (k, m) are given be the following table:
 

 (1, 1) (1, 2) (1, 3) (1, 4) (1, 5) (1, 6)  



 (2, 2) (2, 4) (2, 6) 



(3, 3) (3, 6)
 
(k, m) ∈

 (4, 4) 

(5, 5)

 


 


(6, 6)

1
Nominally, we’re supposed to be summing by the rows of this table (i.e. fix k and run the sum over
corresponding m). However, by interchanging the order of summation we can instead consider this
as a sum over the rows: if we instead pick the value of m first, we see that

XX n X
X
ϕ(k) = ϕ(k).
k≥1 k|m m=1 k|m
m≤n
P
Using the famous fact d|n ϕ(d) = n, we conclude
n X n
X X 1
ϕ(k) = m = n(n + 1).
2
m=1 k|m m=1

Here
P one has the idea that one can “swap Pthe order of summation”: even though there is a
single initially, by rewriting it as a double and then swapping the order, we are able to solve
the problem.
The goal of this lecture is to try and push this idea to allow us to do similar calculations over
both infinite sums and integrals. Because of the introduction of infinity, things become a little more
complicated and some more care is necessary. So, in the first part of the lecture we will address
conditions on which rearranging the order of summation or integration is permissible. After that
we will see several applications.

1.2 Absolute and conditional convergence


P
Let n an be an infinite series of complex numbers; then its limit is defined as
N
!
X X
an := lim an .
N →∞
n n=1

Note that this depends on the order of the terms: if we permute the sequence, the limit might
change! This is weird and bad since we would want “infinite addition” to be commutative, so we
want a way to avoid this behavior. This is accomplished by using the so-called notion of absolute
convergence.
P P
Definition 2. If ak converges, we say it converges absolutely if |ak | < ∞, and converges
conditionally otherwise.
P
Theorem 3 (Rearrangement okay iff absolutely convergent). Let an be a convergent series of
complex numbers.
P
(a) If an is absolutely convergent, it is invariant under permutation of the terms (the sum will
still converge, and the limit remains the same).
P
(b) If an is conditionally convergent and an are real numbers, then there exists a permutation
of the terms for which the sum converges to 2018.

Thus, any time befP


ore you try to rearrange the series, you must check first that it’s absolutely
convergent. With two signs the statement reads:

2
Theorem 4 (Fubini for doubly-indexed infinite sums). Let am,n ∈ C. If any of the three quantities
! !
X X X X X
|am,n |, |am,n | , |am,n |
(m,n)∈N2 m n n m

are convergent, then


! !
X X X X X
am,n = am,n = am,n
(m,n)∈N2 m n n m

and all three series are convergent.


Corollary 5 (Tonelli for doubly-indexed infinite sums). Let am,n ∈ R≥0 . Then
! !
X X X X X
am,n = am,n = am,n
(m,n)∈N2 m n n m

where we allow the possibility all three diverge.


Here is the classic example.
Example 6 (Putnam 2016 B6). Evaluate
∞ ∞
X (−1)k−1 X 1
.
k k2n +1
k=1 n=0

Proof. Before anything else, the sum is absolutely convergent since we have
X 1 X
−2
X 1 π2
< k = · 2 < ∞.
k(k · 2n + 1) 2n 6
k≥1 k≥1 n≥0
n≥0

Thus we may swap the order of summation freely.


We use d = k · 2n + 1 ≥ 2 as the summation variable, so that the sum in question is
X1 X (−1)k−1
.
d k
d≥2 k
∃n:d−1=k·2n

1
Now we claim that the inner sum is exactly d−1 . Indeed, if d − 1 = 2r m with m odd, then the sum
is
r
(−1)m−1 (−1)2m−1 (−1)2 m−1
 
1 1 1 1
+ + ··· + = − − ··· − r
m 2m 2r m m 1 2 2
1
= r
2 m
1
= .
d−1
Consequently, the final answer is
 
X 1 X 1 1
= − = 1.
d(d − 1) d−1 d
d≥2 d≥2

3
2 Riemann integral
So far all of this is fair-game on high school. We’ll now move into the realm of calculus.
Definition 7. A tagged partition P of [a, b] consists of a partition of [a, b] into n intervals, with a
point ξi in the nth interval, denoted

a ≤ x0 < x1 < x2 < · · · < xn ≤ b and ξi ∈ [xi−1 , xi ] ∀ 1 ≤ i ≤ n.

The mesh of P is the width of the longest interval, i.e. maxi (xi − xi−1 ).

Theorem 8 (Riemann integral). Let f : [a, b] → C be continuous. Then the definition


n
Z b !
X
f (x) dx = lim f (ξi )(xi − xi−1 )
a P tagged partition
mesh P →0 i=1

is well-defined (and finite).

There are a bunch of remarks I want to make about this result.

2.1 Compactness and improper integrals


We won’t prove the definition of the Riemann integral works out, but we will mention that its proof
hinges crucially on:
Fact 9. The interval [a, b] is compact, so continuous functions f : [a, b] → C behave well. In partic-
ular, f is bounded, and “uniformly continuous”.

This fact is false for open (or unbounded) intervals: consider the function 1/ x on (0, 1), for
example. This gives rise to the notion of “improper integrals”, such as
Z 1
1
√ dx.
0 x

As written, this does not officially make sense as a Riemann integral, since f (x) = √1 is not a
x
function on [0, 1]. Rather, we implicitly mean
Z 1
1
lim √ dx
ε→0+ ε x

since
R 1 −1f (x) is well-defined on [ε, 1]. In this case, there is no guarantee the limit exists; for example
0 x dx = ∞.
Similarly, it’s possible to set endpoints at ∞ by e.g.
Z ∞ Z B
f (x) := lim f (x)
−∞ B→∞ −B

for example.

4
2.2 Mesh sums
Sometimes, you will find that a sum can be written in such a way that it corresponds to the mesh
of a Riemann integral. In that case, one is very happy, because then it turns the entire sum into a
single integral!
Example 10. Evaluate  
1 1 1
lim + + ··· + .
n→∞ n+1 n+2 2n
Proof. Write as
n
!
1X 1
lim k
.
n→∞ n 1+ n
k=1
1
R1 1
Then, this is a mesh sum for f (x) = 1+x over [0, 1]. Thus by definition it approaches 0 1+x dx =
log 2.

2.3 Discretization and inequalities


If asked to prove an identity or inequality about integrals, it is often possible to revert back to
a discrete sum, a technique called discretization. For example, suppose one wishes to prove the
Cauchy-Schwarz inequality in the form
Z b 2 Z b  Z b 
2 2
f (x)g(x) dx ≤ f (x) dx g(x) dx
a a a

for continuous functions f, g : [a, b] → R. By taking meshes, it is sufficient to prove


!2 ! !
1X 1X 1X
f (ai )g(ai ) dx ≤ f (ai )2 dx 2
g(ai ) dx .
n n n
i i i

which is of course just the classical Cauchy-Schwarz from high school.


In practice, aside from discretization, most integral inequalities on competitions will really just
use the following fact:

Lemma 11 (The obvious inequality). Let f, g : [a, b] → R be continuous functions. If f (x) ≤ g(x)
for all x ∈ [a, b] then
Z b Z b
f (x) dx ≤ g(x) dx.
a a

3 Lebesgue integrals
3.1 Advantages of Lebesgue integrals
Unfortunately, Riemann integrals are terrible. In order to properly state theorems about interchang-
ing order of summation, it’ll be Rmuch more convenient to proceed with the LebR esgue integral,
b
which I will generally denote by X to distinguish it from the Riemann integral a .
Defining the Lebesgue integral is much more involved, because it involves a bunch of measure
theory, so I won’t define what it is (but those of you taking 18.175 will find out really soon).
However, I’ll at least mention the following reasons it’s appreciated.

5
• Better theorems P about
R b swappingR b P limits and sums. For example, for the Riemann
integral, swapping n a fn and a n fn requires uniform convergence, which is a pretty
strong condition (although it’ll be true for Taylor series, which is a frequent use case for us).

• Improper integrals can be handled natively. You can write (0,1) √1x dx and R exp(−x2 ) dx
R R

and it makes sense, unlike for the Riemann case where one has to use an improper integral.

• More versatile. Although we won’t encounter any, some functions thatRwere previously not
Riemann integrable can now be assigned values. The classic example is [0,1] 1Q = 0.

3.2 Riemann integrals and Lebesgue integrals


Of course, it’d be really silly if there wasn’t some guarantee that the Riemann integrals and Lebesgue
integrals agree.
The rules for converting a Riemann and Lebesgue integral are as follows:

• For continuous functions f : [a, b] → C, the Riemann integral and Lebesgue integrals coincide.
So proper Riemann integrals work out of the box.

• For continuous nonnegative functions f : (a, b) → R≥0 on an open (or half-open) interval
where one needs improper integrals, the improper Riemann integral and Lebesgue integrals
coincide (where we allow the possibility that the integrals are both +∞). Here, a = −∞ and
b = +∞ are allowed too.
Rd
• For general f : (a, b) → C, if the partial integrals c |f | dx are bounded for any [c, d] ⊂ (a, b)
then we can also swap as above.

On the other hand, if your signs are all over the place, then there isn’t hope in general
R ∞ sin xof converting
improper Riemann integrals to Lebesgue ones. A famous textbook example is 0 x dx which in
fact is not covered by Lebesgue integration.

3.3 Swapping double integrals


I’ll state this in the full generality, though we’ll only use it in the cases where the “σ-finite measure
spaces” are N (corresponding to infinite sums) or sub-intervals of R (corresponding to Riemann
integrals).

Theorem 12 (Fubini). Let X and Y be “σ-finit R eR measure spaces”. LRet fR : X × Y → C be


cRontinuous (or just “measurable”). If any of X Y |f (x, y)| dy dx, Y X |f (x, y)| dx dy,
X×Y |f (x, y)| d(x, y) are finite, then we have
Z Z  Z Z  Z
f (x, y) dy dx = f (x, y) dx dy = f (x, y) d(x, y).
X Y Y X X×Y

Corollary 13 (Tonelli). Let X and Y be “σ-finite measure spaces”. Let f : X × Y → R≥0 be


continuous (or just “measurable”) and nonnegative. Then
Z Z  Z Z  Z
f (x, y) dy dx = f (x, y) dx dy = f (x, y) d(x, y)
X Y Y X X×Y

where we allow the possibility that all three are +∞.

6
Remark 14. • If X = N and Y = N, then this corresponds to the double sums we stated earlier.
PR RP
• If X = N and Y ⊂ R is an interval, then this states that and can be swapped.

• Note that if X and Y are finite closed intervals and f : X × Y → C is continuous, then
hypotheses of Fubini are automatically satisfied, since X ×Y is compact. The situation where
X and Y are open/infinite is more slippery, although in most cases we’ll have nonnegativity
and then Tonelli will save us.
Tonelli’s theorem (together with the result that even improper Riemann integrals are okay with
nonnegative functions) means that whenever you have nonnegative functions, you can proceed no
holds barred — everything works beautifully. In other words nonnegative =⇒ euphoria.

3.4 Interchanging limits and Lebesgue integrals


You can read this off of the results on sums, but we’ll state them here since they have names.

Theorem 15 (Dominated convergence theorem). Let fn : I → C be a seqRuence of continuous


functions on an interval I ⊆ R. Assume that |fn (x)| ≤ g(x) for all x, where I g(x) < ∞ (i.e. g is
integrable). Then limn fn (x) is integrable and
Z Z
lim fn (x) dx = lim fn (x) dx.
n→∞ I I n→∞

Theorem 16 (Monotone convergence theorem). Suppose that fn : I → R≥0 is a sequence of


continuous functions on an interval I ⊆ R which are also nonnegative. Assume further that
fn (x) ≤ fn+1 (x) for n ∈ N, x ∈ I. Then
Z Z
lim fn (x) dx = lim fn (x) dx
n→∞ I I n→∞

where the value of any of these integrals is allowed to be infinite.

4 Techniques for introducing more sums


4.1 Taylor series
Some common ones:
X xn x2 x3
exp(x) = =1+x+ + + ... ∀x∈R
n! 2! 3!
n≥0
X xn x2 x3
log(1 − x) = − = −x − − − ... ∀ |x| < 1
n 2 3
n≥1
1 X
= xn = 1 + x + x2 + . . . ∀ |x| < 1
1−x
n≥0
X (−1)n x2n+1 x3 x5
arctan(x) = =x− + − ... ∀ |x| < 1.
2n + 1 3 5
n≥0

There is a nice theorem about Taylor series in general:

7
Theorem 17 (Convergence of Taylor series). Let f be an analytic function. Within its radius of
convergence, the Taylor series for f will

• converge absolutely for any x as a series of complex numbers, and

• converge uniformly on any compact sub-interval, as a series of functions (i.e. it is compactly


convergent).

I mention uniform convergence here since it’s actually strong enough to allow swapping inte-
gration even for the Riemann integral. Here’s the definition:
Definition 18. A sequence of functions Fn : [a, b] → C is said to converge uniformly to the function
F : [a, b] → C if
lim sup |Fn (x) − F (x)| = 0.
n→∞ x∈[a,b]
P Pn
A series n fn converges uniformly if its partial sums Fn = k=1 fk do.
But we’ll be mostly using Lebesgue integrals anyways.
So, whenever you have an analytic function on a closed interval, all the summation results work
fine! Here is a very famous example.
Example 19. Compute Z 1
log x log(1 − x) dx.
0
There is some subtlety here since this integral looks like it might be improper! Fortunately,
it’s not quite, since limx→0+ log(x) log(1 − x) = 0, and in this way we can actually regarding
log(x) log(1 − x) as a proper integral on [0, 1].

Proof. Switch to Lebesgue integration. The integral is then


Z X xn
I=− log x dx
[0,1] n
n≥1
X1Z 1
=− xn log x dx (by Tonelli)
n 0
n≥1
X1 x=1
n+1 (n + 1) log x − 1
=− x · (integration by parts)
n (n + 1)2 x=0
n≥1
X1 x=1
n+1 (n + 1) log x − 1
=− x ·
n (n + 1)2 x=0
n≥1
X 1
=
n(n + 1)2
n≥1
X1 1 1

= − − .
n n + 1 (n + 1)2
n≥1

1 PN 1 π2
The N th partial sum of this is equal to 1 − N +1 − n=1 (n+1)2 which gives 2 − 6 as N → ∞.

Remark
R1 n 20 (An application of Feynman’s trick). In my original notes, I had obtained the identity
1
0 x log x dx = − (n+1)2 using integration by parts. In class it was pointed out that Feynman’s

8
trick, more descriptively called “differentiating under the integral sign”, gives a shorter way to
prove this. Start by writing Z 1
1
xn dx =
0 n + 1
and then treat n ∈ R as a parameter. This allows one to differentiate both sides with respect to n,
yielding
Z 1
d n d 1
x dx =
0 dn dn n + 1
Z 1
1
=⇒ xn log x dx = − .
0 (n + 1)2

See http://www.math.uconn.edu/~kconrad/blurbs/analysis/diffunderint.pdf for more de-


tails on this trick.

4.2 Eliminating fractions


The following seemingly obvious statement is surprisingly useful.

Lemma 21 (Denominator → integral). For any real number s > −1 we have


Z
1
= ts ds.
s+1 (0,1)
P n
As a simple use case, let’s suppose we were given n xn for some |x| < 1 and wanted to figure
out what function it was (without knowing anything about log in advance). We can write
X xn X Z
= xn tn−1 dt
n
n n [0,1]
XZ
= x(xt)n−1 dt
[0,1]
Zn X
= x(xt)n−1 dt
[0,1] n
Z
x
= dt
[0,1] 1 − xt
= [− log(1 − xt)]1t=0
= [− log(1 − xt)]1t=0 = − log(1 − x).

Let’s also see a solution to the earlier double sum.


Example 22 (Putnam 2016 B6). Evaluate
∞ ∞
X (−1)k−1 X 1
.
k k2n+1
k=1 n=0

9
Proof. Check conditional convergence of the double sum in the same way as before. Thus we apply
Fubini freely:
∞ ∞ ∞ ∞ Z
X (−1)k−1 X 1 X (−1)k−1 X n
n
= tk2 dt
k k2 + 1 k
k=1 n=0 k=1 n=0 [0,1]
∞ X ∞ n
!
(−t2 )k
Z X
= − dt
[0,1] k
n=0 k=1
∞ ∞
Z Z !
n 2n
X Y
= log(1 + t2 ) dt = log (1 + t ) dt
[0,1] n=0 [0,1] n=0
Z   Z
1
= log dt = − log(1 − t) dt = 1.
[0,1] 1−t [0,1]

4.3 Fourier series


If f : R → C is continuous with period 1, then
N
X
f (x) = lim am exp(2πimx).
N →∞
m=−N

The Fourier coefficients am are given by


Z 1
am = f (x) exp(−2πimx) dx.
0

We again have convergence results:

Theorem 23. Let f : [0, 1] → C be periodic.

(a) The Fourier series converges uniformly provided f is continuously differentiable (this can be
weakened to “absolutely continuous”, but we won’t need that level of generality).
P
(b) The Fourier series converges absolutely as long as m∈Z |am | < ∞.

Example 24. If f : R → C is continuously differentiable with period 1, and α is an irrational number,


then Z 1
f (α) + · · · + f (nα)
lim = f (x) dx.
n→∞ n 0
P
Proof. Just write f (x) = m am exp(2πimx). Then note that for m 6= 0, if we let z = exp(2πimα)
then
z1 + z2 + · · · + zn z(1 − z n )
= →0
n n(1 − z)
as long 6 1, which holds since z is not a root of unity. This leaves just the contribution form
R 1 as z =
a0 = 0 f (x) dx.

In general, Fourier-type sums are good things to keep an eye out for, even if they don’t explicitly
come from Fourier series. For example, given a complex polynomial p(z) (or even a series):
 2πik 
• The discrete sum n−1
P
k=0 p e n extracts the coefficients with indices divisible by n,

10
R 2π it
• the integral t=0 p(e ) dt = 2π · p(0) extracts the constant term of the polynomial,
and so on. This is related to complex analysis, in which it turns complex differentiable functions
C → C are exactly the same as complex analytic functions, which means you can go nuts with all
sorts of beautiful results such as Cauchy’s theorem.

5 Problems
1. Evaluate the improper integral
1
log(1 − x)
Z
dx.
0 x

2. Determine the value of the improper integral


Z ∞
x
dx.
0 ex − 1

R∞
3. (a) Show that that min(a, b) = 0 1≤a (t)1≤b (t) dt for any nonnegative real numbers a, b ≥
0. (What do you think 1≤c (t) means?)
(b) Show that if r1 , . . . , rn are nonnegative reals and x1 , . . . , xn are real numbers then
n X
X n
min(ri , rj )xi xj ≥ 0.
i=1 j=1

R1 R1
4. For each continuous function f : [0, 1] → R let I(f ) = 0 x2 f (x) dx and J(f ) = 0 xf (x)2 dx.
Find the maximum value of I(f ) − J(f ) over all such functions f .

5. Compute  
1 1 1
lim √ +√ + ··· + √ .
n→∞ 4n2 − 12 4n2 − 22 4n2 − n2

6. Let a and b be real numbers with a < b, and let f and g be continuous functions from [a, b]
Rb Rb
to (0, ∞) such that a f (x) dx = a g(x) dx but f =
6 g. For every positive integer n, define
b
(f (x))n+1
Z
In = dx.
a (g(x))n
Show that I1 , I2 , I3 , . . . is an increasing sequence with limn→∞ In = ∞.

7. Let a0 , a1 , . . . , an , x be real numbers, where 0 < x < 1, satisfying


a1 a2 an
a0 + + 2
+ ··· + = 0.
1+x 1+x+x 1 + x + x + · · · + xn
2

Prove that for some 0 < y < 1 we have

a0 + a1 y + a2 y 2 + · · · + an y n = 0.

11
8. Find
n n
1 XX a
lim .
n→∞ n a + b2
2
a=1 b=1

9. Evaluate the following:


Z ∞
x3 x5 x7 x2 x4 x6
  
x− + − + ··· 1 + 2 + 2 2 + 2 2 2 + · · · dx.
0 2 2·4 2·4·6 2 2 ·4 2 ·4 ·6

10. Show that Z 1 X


x−x dx = n−n .
0 n≥1

11. RSuppose that f is a function on the interval [1, 3] such that −1 ≤ f (x) ≤ 1 for all x and
3
1 f (x) dx = 0. Determine the largest possible value of
Z 3
f (x)
dx.
1 x

12. Let f : R → R be continuous and satisfy f (x) ≥ 1 for all x. Suppose that
f (x)f (2x) . . . f (nx) ≤ 2018n2019
for every positive integer n and x ∈ R. Must f be constant?
R∞ 2 √
13. Show that −∞ e−x dx = π.

14. A rectangle in R2 is called great if either its width or height is an integer. Prove that if a
rectangle X can be dissected into great rectangles, then the rectangle X is itself great.
15. Compute
X 2k
.
k≥0
52k + 1

16. Determine the value of


" #
21/n 22/n 2n/n
lim + + ··· + .
n→∞ n + 1 n + 12 n + n1

17. For any continuous function f : [0, 1] → R let


Z 1 Z 1
µ(f ) = f (x) dx, Var(f ) = (f (x) − µ(f ))2 dx, M (f ) = max |f (x)|.
0 0 0≤x≤1

Show that if f, g : [0, 1] → R are continuous functions then


Var(f g) ≤ 2 Var(f )M (g)2 + 2 Var(g)M (f )2 .

12
18. For m ≥ 3, a list of m

3 real numbers aijk (where 1 ≤ i < j < k ≤ m) is said to be area
definite for Rn if the inequality
X
aijk · Area(4Ai Aj Ak ) ≥ 0
1≤i<j<k≤m

holds for every choice of m points A1 , . . . , Am in Rn . For example, the list of four numbers
a123 = a124 = a134 = 1, a234 = −1 is area definite for R2 . Prove that if a list of m 3 numbers
is area definite for R2 , then it is area definite for R3 .
19. Prove that 1
n  
!
n(n+1)
Y n √
lim = e.
n→∞ k
k=0

20. Let f : [0, 1] → R be continuous. Show that


Z 1Z 1 Z 1
|f (x) + f (y)| dx dy ≥ |f (x)| dx.
0 0 0

21. Let f : R≥0 → R be a strictly decreasing continuous function such that limx→∞ f (x) = 0.
Prove that Z ∞
f (x) − f (x + 1)
dx
0 f (x)
diverges.
22. A rectangular prism X is contained within a rectangular prism Y .
(a) Is it possible the surface area of X exceeds that of Y ?
(b) Is it possible the sum of the 12 side lengths of X exceeds that of Y ?
23. For a, b, c > 0 prove that
1 1 1 4 4 4 12 12 12
+ + + + + ≥ + + .
a b c a+b b+c c+a 3a + b 3b + c 3c + a

24. Define a function w : Z → Z as follows. For |a|, |b| ≤ 2, let w(a, b) be as in the table shown;
otherwise, let w(a, b) = 0.
b
w(a, b) −2 −1 0 1 2
−2 −1 −2 2 −2 −1
−1 −2 4 −4 4 −2
a 0 2 −4 12 −4 2
1 −2 4 −4 4 −2
2 −1 −2 2 −2 −1
For every finite nonempty subset S of Z × Z, prove that
X
A(S) := w(s − s0 ) > 0.
(s,s0 )∈S×S

13
25. Evaluate n
Y  1 + xn+1 x
lim .
x→1− 1 + xn
n≥0

26. Suppose that f : [0, 1]2 → R is continuous. Show that


Z 1 Z 1 2 Z 1 Z 1 2
f (x, y)dx dy + f (x, y)dy dx
0 0 0 0
Z 1Z 1 2 Z 1Z 1
≤ f (x, y) dx dy + [f (x, y)]2 dx dy.
0 0 0 0

27. For
h √eac
h positive integer k, let A(k) be the number of odd divisors of k in the interval
1, 2k . Evaluate:

X A(k)
(−1)k−1 .
k
k=1

14
MIT OpenCourseWare
https://ocw.mit.edu/

18.A34 Mathematical Problem Solving (Putnam Seminar)


Fall 2018

For information about citing these materials or our Terms of Use, visit: https://ocw.mit.edu/terms.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy