Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Natural Numbers: A Supplement to the Book of Rudin

The purpose of this article is to fill up the logical details of section 2.1 of the book “Prin-
ciples of Mathematical Analysis, third edition, Rudin.” Only those mathematical concepts
and techniques about number systems which are used frequently in the later of that book
are introduced. Also some proofs of theorems in chapter one and two of the textbook are
improved to make them more rigorous.
We assume the naı̈ve set theory as background here. Readers must be familiar with
those concepts, operations and some basic properties about sets, such as Cartesian prod-
ucts, index sets, De Morgan’s laws, the axiom of choice, etc.
Usually we suppose the existence of N, the set of all natural numbers, deduce some
properties about N, and then construct the set Z of all integers from N, the set Q of all
rational numbers from Z, and the set R of all real numbers from Q. Nonetheless, it is
logically correct if we start from assuming the existence of R, define a specific subset of R
to be N, and deduce those important properties about N. Here we go.
Theorem 1 (Theorem 1.19 of textbook). There is a unique ordered field, denoted by R,
with the least upper bound property up to isomorphism; that is, if pF, `, ¨, †q is also a such
field, then there is a 1-1 onto function : R Ñ F such that

px ` yq “ pxq ` pyq, pxyq “ pxq ¨ pyq, and x † y implies pxq † pyq,

for all x, y P R.
Moreover, such between R and F is unique.
Here a field is a set endowed with the common arithmetical operators, such as addition,
subtraction, multiplication and division, which satisfy some basic laws. The adjective
ordered means that there is a relation † on this field, compatible with the arithmetical
operators, such that we can compare the magnitude of elements in this field.
Definition 1. A subset S of R is called bounded above if there is ↵ P R such that s § ↵
for all s P S; here we say that such ↵ is an upper bound of S. Similar for “bounded below”
and “lower bound”.
Furthermore, an upper bound ↵ of S is called the least upper bound or supremum of
S provided ↵ satisfies one of the following equivalent statements.
(i) If † ↵, then can not be an upper bound of S.
(ii) If is an upper bound of S, then ↵ § .
(iii) For every with † ↵, there is s P S such that

† s § ↵.

(iv) For each ✏ ° 0, there is s P S such that

↵ ´ ✏ † s § ↵.

1
Notice that by (ii), if the supremum of S exists, then it is unique. In this case, we denote
the least upper bound of S by sup S. Similar for the greatest lower bound or infimum of
S, denoted by inf S.
In the theorem, that “R has the least upper bound property” means that “for every
non-empty bounded above subset S of R, sup S exists.” The least upper bound property
shows the continuum of R, unlike Q, in which there are infinite holes among rational
numbers, like a sieve. See Example 1.1 in the textbook. Also, note that the least upper
bound property implies that “if S is a non-empty bounded below subset of R, then inf S
exists.” The proof is referred to Exercise 1.5 in the textbook.
Remark. sup S and inf S may not belong to S even if they exist. For example, the open
internal p0, 1q has supremum 1 and infimum 0, but both of them are not in p0, 1q.

System of Natural Numbers


Definition 2. We say that a subset S of R is inductive if it satisfies
(i) 1 P S;

(ii) for any s P S, s ` 1 is also in S.


It is clear that the set R•1 – tr P R : ì
r • 1u is inductive, and thus we have an inductive
proper subset of R in hand. Let N – S, where the intersection runs over all inductive
subsets of R.
Note that so far we still have no ideal about what N looks like, or what kind of elements
in R belong to N. All those properties of N we used a lot in ordinary mathematics will be
proved in the following.
Proposition 1. N is inductive.
Proof. Since 1 belongs to all inductive subsets of R, 1 is also in N. For each s P N, s is in
all inductive subsets of R by the definition of N, and so is s ` 1, whence s ` 1 belongs to
N.
Theorem 2. [Principle of Mathematical Induction.] Suppose that S is a subset of
N such that (i) 1 P S; and (ii) n P S implies n ` 1 P S. Then S “ N.
Proof. Notice that S Ñ N by assumption. Then the theorem follows from the definition of
N because S is an inductive subset of R.
Definition 3. For any n P N, we define Jn to be the set tk P N : k § nu of all natural
numbers smaller or equal to n.
We can see that J1 “ t1u because R•1 is inductive. However we do NOT presume
something like Jn “ t1, 2, . . . , nu, and we won’t use those unsubstantiated properties,
which are regarded as being trivial in the ordinary mathematics, in the following proofs.

2
Proposition 2. For any n P N, Jn`1 zJn “ tn ` 1u.
(
Proof. Let S – k P N : Jk`1 zJk “ tk ` 1u . We claim that (i) 1 P S; and (ii) n P S
implies n ` 1 P S. Note that N is a subset of R and hence 1 ` 1 is an element of R.
Hereafter we denote 1 ` 1 by 2.
Assume (i) does not hold. Then there is ↵ P N such that 1 † ↵ † 2. Since R•1 is an
inductive subset of R, by the definition of N, every element in N is larger or equal to 1. So
there is no element m P N such that m ` 1 “ ↵; otherwise ↵ “ m ` 1 • 2. Then it is easy
to show that Nzt↵u is still inductive, and thus N Ñ Nzt↵u by the definition of N again, a
contradiction. It is similar to verify (ii); that is, assume that there is ↵ P Jn`2 zJn`1 which
is not n ` 2, and then use the induction hypothesis to obtain a proper inductive subset of
N and deduce a contradiction.
Therefore S “ N by mathematical induction, and the proposition follows.

Proposition 3. Let m, n P N. If m ° n, then m • n ` 1.

Proof. Since m ° n, either m P Jn`1 zJn or m ° n ` 1, both of which imply m • n ` 1


because of Jn`1 zJn “ tn ` 1u.

Theorem 3. [Principle of Strong Induction.] Suppose that S is a subset of N such


that (i) 1 P S; and (ii) Jn Ñ S implies n ` 1 P S. Then S “ N.

Proof. We define S̄ to be the set tn P N : Jn Ñ Su. Then we have the following.

• 1 P S̄ since 1 P S and hence J1 “ t1u Ñ S.

• If n P S̄, then Jn Ñ S by the definition of S̄. Hence n ` 1 P S by (ii), wherefore


Jn`1 “ Jn Y tn ` 1u(why?) is contained in S. So n ` 1 P S̄.

Therefore S̄ “ N by mathematical induction, whence for every n P N, Jn is contained in


S, and hence n P Jn belongs to S, which implies that N Ñ S. The proof is completed.

Theorem 4. [Well-Ordering Principle.] Every non-empty subset of N contains its


minimum.

Proof. Suppose that S is a subset of N in which there is no minimum. Let A “ NzS. Then
we have the following.

• 1 R S; or 1 would be the minimum of S (why?), whence 1 P A.

• If Jn Ñ A for some n, then n ` 1 R S; otherwise, since every element in N which is


smaller or equal to n belongs to A, and thus any element in S is larger or equal to
n ` 1, n ` 1 would be the minimum of S. Hence n ` 1 P A.

So by strong induction, A “ N and hence S “ H because of S Ñ N. The proof is


completed.

3
Theorem 5. [Pigeonhole Principle.] For any m, n in N with m ° n, and for any
function f : Jm Ñ Jn , there are a, b P Jm such that a ‰ b but f paq “ f pbq.
Proof. Since m ° n implies that m • n ` 1 and thus we can consider the restriction of
functions on Jn`1 , we only need to prove the case of functions from Jn`1 into Jn . We will
use the principle of mathematical induction, but from now on, we don’t describe the subset
of N explicitly anymore.
When n “ 1, we have the only one function f : t1, 2u Ñ t1u mapping both 1 and 2
to 1, whence the desired property is satisfied. Suppose that for some n P N, any function
f : Jn`1 Ñ Jn meets the required condition. Given a function g : Jn`2 Ñ Jn`1 . We divide
the proof into cases.
Case 1. If the image of g is Jn , then we consider the restriction of g on Jn`1 , and thus,
by induction hypothesis, g satisfies the condition.
Case 2. If n ` 1 has at least two preimages, it’s done.
Case 3. If n ` 1 has exact one preimage, which just happens to be n ` 2, then it is easy
to see that the restriction of g on Jn`1 can be regarded as a function, say g 1 , from Jn`1 to
Jn . Thus by induction hypothesis, the required property holds for g 1 and hence for g.
Case 4. Assume that n ` 1 has exact one preimage which is not n ` 2. We define the map
h : Jn`1 Ñ Jn`1 by $
&n ` 1,
’ if k “ gpn ` 2q,
hpkq “ gpn ` 2q, if k “ n ` 1,

%
k, otherwise,
and consider the composition h ˝ g as the following diagram.
Jn`2 Jn`1 Jn`1

n`2 g n`1 h n`1


n`1 .. ..
. .
..
. gpn ` 2q gpn ` 2q
g ´1 pn ` 1q .. ..
. .
..
. 1 1
1

It is easy to verify that h is a 1-1 onto function and that h ˝ g is a kind of functions we
dealt with in case 2 or 3. So there are a, b P Jn`2 such that a ‰ b but h ˝ gpaq “ h ˝ gpbq,
whence, by the 1-1 property of h, gpaq “ gpbq, as required.
The proof is completed.

4
Besides the principle of mathematical induction and its variations, there is still an im-
portant mathematical technique called recursive definition or inductive definition, which
had been used in the proofs of Theorem 2.8 and 2.43 implicitly. Although people some-
times confuse it with the axiom of countable choice, the principle of recursive definition is
di↵erent from that axioms and can be proved without supports of any kind of axiom of
choice. Here we only state the principle of recursive definition and demonstrate the way
how to use it formally in the following example and later in the proof of that “infinity
if and only if Dedekind-infinity.” For more details and a proof of principle of recursive
definition, refer to Section 1.8 of “Topology, 2nd Edition, by James Munkres”.
Theorem 6. [Principle of Recursive Definition.] Suppose that A is set and a1 is a
member of A. Also suppose that ⇢ is a function which assigns each function from Jn into
A for some n P N, to an element of A; that is, let A be the set

tf : f is a function from Jn into A for some n P Nu,

and let ⇢ be a function from A to A. Then there exists a unique function h : N Ñ A such
that

hp1q “ a1 , (1)
` ˘
and hpnq “ ⇢ h|Jn´1 for all n P N with n ° 1, (2)

where h|Jn´1 denotes the restriction of h on Jn´1 .


Example. To understand the meaning of the principle of recursive definition clearly, we
consider the famous sequence tFi u8
i“1 of integers, called Fibonacci sequence, which satisfies
that

F1 “ 1, F2 “ 1,
and Fn “ Fn´1 ` Fn´2 for all n • 3.

The principle of recursive definition says that if for any finite sequence a1 , a2 , a3 , . . . , an of
chosen elements in A, which is represented by the function f : Jn Ñ A, we have a definite
rule, say ⇢, to decide the next element — in the case of Fibonacci sequence ⇢ maps the
sequence tf1 u of a single term to the term f1 in that sequence, and maps the finite sequence
tf1 , f2 , . . . , fn´2 , fn´1 u to the sum fn´2 ` fn´1 of the last two terms when n • 3 — then
for any initial condition as in Eq.(1), we will obtain a unique and determinate infinite
sequence by continuing such process.

Arithmetic of Number Systems


Lemma 1. If n is a member of N with n ° 1, then n ´ 1 is also a member of N.
Proof. If n ´ 1 is not a member of N, it is not hard to prove that Nztnu is an inductive
subset of N, which contradicts to the principle of mathematical induction.

5
Proposition 4. As a subset of R, N is closed under addition and multiplication. Further-
more, for all m, n P N with n † m, m ´ n is also in N.

Proof. By the inductive property of N in R, it is easy to verify by mathematical induction


that N is closed under addition and multiplication.
Assume that there are m, n P N with n † m but m ´ n R N. Then the set

tn P N : there is m P N such that m ° n but m ´ n R N.u

is non-empty and hence contains its minimum, say no , by well-ordering principle. Let mo
belong to N such that mo ° no but mo ´ no R N. Note that no ‰ 1 by above lemma.
However by above lemma again, mo ´ 1 and no ´ 1 both are in N, mo ´ 1 ° no ´ 1 but
pmo ´ 1q ´ pno ´ 1q “ mo ´ no R N, whence no ´ 1 is also an element in above set, which
contradicts to the minimality of no . The proof is completed.

Corollary 1. For all m, n P N, the function m: Jn Ñ Jm`n zJm defined by k fiÑ k ` m is


well-defined and bijective.

Proof. Exercise.

Definition 4. We define Z to be the set

tr P R : r “ 0, r P N or ´ r P N.u

and call each element in Z an integer; also define Q to be the set


!m )
“ mn´1 : m P Z, n P Z and n ‰ 0.
n
and call each element in Q a rational number.

Corollary 2. (i ) Z is an additive subgroup of R.

(ii ) Z is closed under the multiplication.

(iii ) Q is a subfield of R.

Proof. Exercises.

Theorem 7. [Archimedean Property.] For all x, y P R with x ° 0, there is n P N such


that nx ° y.

Proof. If no such n exists, then y is an upper bound of the set E “ tnx : n P Nu. Hence
sup E exists by the least upper bound property of R. By the definition of supremum, we
have m P N such that
sup E ´ x † mx § sup E.
However, the left inequality implies that pm ` 1qx ° sup E, a contradiction.

6
Lemma 2. (i) For all x P R, there is n P Z such that n § x † n ` 1.

(ii ) For all x, y P R with y ´ x ° 2, there is n P Z such that x † n † y.

Proof. (i) Exercise. Hint: Use the Archimedean property and well-ordering principle to
prove the case of x • 1 at first.
(ii) follows from (i).

Proposition 5. Q is dense in R; that is, for all x, y P R with x † y, there is q P Q such


that x † q † y.

Proof. By the Archimedean property, there is n P N such that npy ´ xq ° 2; and hence by
the previous lemma, there is m P Z satisfying

nx † m † ny.

Thus
m
x† † y,
n
m
and n
is the desired rational.

At Most Countable Sets


In this section we will try to complete those logical details omitted in section 2.1 of the
book of Rudin, and we will adopt the same notations and terminologies as ones in the
textbook. Recall that two sets A and B are said to be equivalent, denoted by A „ B,
provided there is a 1-1 and onto function between them.

Proposition 6. For all n, m P N, Jn „ Jm if and only if n “ m.

Proof. This proposition follows from pigeonhole principle.

Proposition 7. Let n P N. If S is a proper subset of Jn , then either S “ H or S „ Jm


for some m † n. Moreover, we have the following.

(i ) Every subset of any finite set is finite.

(ii ) Every finite set is not equivalent to its proper subsets.

Proof. We prove it by induction on n. It is clear for n “ 1 because J1 “ t1u. Assume that


the desired result holds for some Jn with n P N. Let S be a proper subset of Jn`1 . We
verify the following two cases.
Case 1. If n ` 1 R S, then S is subset of Jn because of Jn`1 zJn “ tn ` 1u. Hence by
induction hypothesis, one of the following happens: S “ Jn , S “ H or S „ Jk for some
k † n. All of them lead to the required result.

7
Case 2. Suppose n ` 1 P S. Since S is a proper subset of Jn`1 and Jn`1 zJn “ tn ` 1u,
Jn zS is non-empty, say k P Jn zS. Observe that the map h : Jn`1 Ñ Jn`1 defined by
$
&n ` 1, if l “ k,

hplq “ k, if l “ n ` 1,

%
l, otherwise,

is a 1-1 onto function and that the image of S is a subset of Jn . See the following diagram.
Jn`1 Jn`1

n`1 h n`1
n n
.. ..
. .
k k
.. ..
. .
1 1

So S is equivalent to a non-empty subset of Jn , wherefore S „ Jm for some m § n † n ` 1


by induction hypothesis.
Next we prove (i) and (ii).
(i) Given a finite set A and its subset B Ñ A. Since A is finite, there is a 1-1 correspon-
dence ' between A and Jn for some n P N, which also induces a bijection between B
and a subset of Jn , as the following diagram. Therefore either B “ H or B „ Jm for
some m § n, whence B is finite.
'
A Jn
Ö

'|B
B pBq

(ii) Given a finite set A and its proper subset B à A. Similar to the proof of (i), we
have that A „ Jn and that either B “ H or B „ Jm for some m † n. By previous
proposition, A and B can not be equivalent.

8
Proposition 8. Finite union of finite sets is finite.
Remark. Here a finite collection of sets is a collection of sets whose members are indexed
by means of the elements of Jn for some n P N. We usually use the notation tAi uni“1 instead
of tAi uiPJn .
Proof. Suppose that tAi uki“1 is a finite collection of finite sets. We prove this proposition
by induction on k. î
When k “ 1, that 1i“1 Ai î “ A1 is finite is clear.
Assume that when k “ n, ni“1 Ai is finite for each collection tAi uni“1 of any finite sets.
Given a collection tAi un`1
i“1 of finite sets. By the fact that Jn`1 “ Jn Y tn ` 1u, it is easy
to verify that
˜ ¸ ˜ ¸ ˜ ¸
§
n`1 §n §n ´§
n ¯
Ai “ Ai Y An`1 “ Ai Y An`1 z Ai .
i“1 i“1 i“1 i“1
în
By induction hypothesis,
în we have that i“1 Ai is finite and hence that there isî a 1-1 onto
function f1 : Ja Ñ i“1 Ai for some a P N. By previous proposition, An`1 z p ni“1 Ai q is
either empty or equivalent to Jb for some b P N. If that set is empty,
în then it’s done. Or
Ja`b zJa Ñ An`1 z p i“1 Ai q. And thus we
by Corollary 1, there is a 1-1 onto function f2 : î
n`1
can
în`1 obtain a 1-1 onto function from J a`b onto i“1 Ai by combining f1 and f2 , whence
i“1 Ai is finite. The proof is completed.

Corollary 3. Suppose that S is an infinite set and that f : S Ñ Jn is a function for some
n P N. Then there is m P Jn such that f ´1 pmq is infinite.
î
Proof. Because of S “ ni“1 f ´1 piq, by the previous proposition, the corollary follows.
Proposition 9. Any infinite set contains a countably infinite subset. Moreover, a set is
infinite if and only if it is equivalent to one of its proper subsets.
Remark. In a common proof of showing that any infinite set contains a countably infinite
subset, we usually just describe how to process the infinite steps of picking up elements,
without saying the principle of recursive definition explicitly, as following.
First fixed a point s1 P S. If s1 , s2 , . . . , sk are chosen, since S is infinite, we can
pick up an element of S, say sk`1 , which is not one of s1 , s2 , . . . , sk . Continue
this process, we obtain a sequence tsn u8 n“1 of distinct elements of S, which is
also a countably infinite subset of S.
But there are some minor logical details missing in such proof. For each time when
s1 , s2 , . . . , sk have been chosen, we have no ideal about how to choose the next specific
element — picking up an arbitrary element blindly can not let us write the sequence defi-
nitely. Here the axiom of choice intervenes. It convinces us that there is always a unique
determinate option to choose. On the other hand, every time after choosing the next term,
we still obtain a finite sequence. How do we know that we can get a complete infinite

9
sequence eventually? In such situation, we use the principle of recursive definition to make
sure that the recursive process always gives us the same result wherever we try to start
over.
Remark. A set is called Dedekind-infinite if it satisfies one of the following equivalent
statements.

(i) That set is equivalent to one of its proper subsets.

(ii) That set contains a countably infinite subset.

Notice that proving above equivalence does not need the support of any kind of axiom of
choice. Also note that the Dedekind-infinity is slightly di↵erent from infinity. Although
Dedekind-infinity always infers infinity, the converse does not hold if we do not have some
extra presumptions — the below proof works because we use the axiom of choice. For more
details about Dedekind-infinite sets, see the Wikipedia page of Dedekind-infinite set.
Proof. Let S be an infinite set. Consider the set
(
S – tf |f is a function from Jn to S for some n P N.

For each f P S , define Sf – Szprange of f q. Since S is infinite and for each f the range of
f is finite (why?), Sf is non-empty, and hence, by the axiom of choice, there is a function
§
⇢: S Ñ Sf Ñ S
f PS

such that ⇢pf q P Sf “ Szprange of f q. By the principle of recursive definition, given s1 P S,


we have a unique function h : N Ñ S such that

hp1q “ s1 ,
` ˘
and hpnq “ ⇢ h|Jn´1 for all n P N with n ° 1.

It is easy to show that h is 1-1, and thus its image is a countably infinite subset of S.
Next we show that a set is infinite if and only if it is equivalent to one of its proper
subsets. Observe that if a set is equivalent to its proper subset, then it can not be finite
by Proposition 7(ii). Conversely, let S be an infinite set. Then S contains a countably
infinite subset, say S̄ – ts1 , s2 , . . . , sn , . . .u. We define a function g : S Ñ S by
#
sk`1 , if s “ sk for some k P N,
gpsq “
s, otherwise.

See the following diagram.

10
S S
g
s s

s1 s1
s2 s2
s3 s3
.. s4
.
..
.

It is easy to show that g is 1-1. Therefore S is equivalent to the image of g, which is a


proper subset of S. The proof is completed.
npn`1q
Lemma 3. (i ) For each n P N, 2
is also in N.

(ii ) For every n P N, there exists a unique m P N Y t0u such that

mpm ` 1q pm ` 1qpm ` 2q
†n§ .
2 2
Proof. Exercise. Hint: Both can be proved by induction on n.
For every n P N and for any set S, we define the Cartesian product of n copies of S,
denoted by S n , to be the set of all functions from Jn into S. Since J2 “ t1, 2u, we have a
natural 1-1 correspondence between S 2 and S ˆ S by considering the following map,

S2 SˆS
Q
Q

p p1q, p2qq

for each function from J2 “ t1, 2u into S. Thus we won’t distinguish the Cartesian
product of 2 copies of S and the set of all ordered pairs whose entries in S and treat them
in the same way. Similar for J3 “ t1, 2, 3u, J4 “ t1, 2, 3, 4u and so on.

Proposition 10. N „ N ˆ N. Moreover N „ Nn for all n P N.

11
Proof. With the aid of previous lemma, it is easy to verify that the following functions are
well-defined and that both of them are inverses of each other: the function f : N ˆ N Ñ N
defined by
pm ` n ´ 2qpm ` n ´ 1q
f pm, nq – `m
2
(see the following diagram; f gives us the counts) and the function g : N Ñ N ˆ N defined
by ˆ ˙
lpl ` 1q pl ` 2qpl ` 1q
gpkq – k ´ , ´k`1 ,
2 2
where l is an element of N Y t0u such that
lpl ` 1q pl ` 1qpl ` 2q
†k§ .
2 2
Therefore, N and N ˆ N are equivalent.

1 p1, 1q 2 p1, 2q 4 p1, 3q 7 p1, 4q . . .


3 p2, 1q 5 p2, 2q 8 p2, 3q p2, 4q . . .
6 p3, 1q 9 p3, 2q p3, 3q p3, 4q . . .
10 p4, 1q p4, 2q p4, 3q p4, 4q . . .
.. .. .. .. ...
. . . .

It can be proved by induction on n that N „ Nn for all n P N. We leave it as an exercise


to the readers.
Remark. There is a brilliant proof of N „ N ˆ N by considering the function ' : N ˆ N Ñ N
defined by 'pm, nq “ 2m´1 p2n´1q. But verifying that f is 1-1 and onto needs the knowledge
about integer factorizations, which is far beyond the scope of this article.

Advanced Topics
We may define 2 to be 1 ` 1, 3 to be 1 ` 1 ` 1 “ 2 ` 1, 4 to be 1 ` 1 ` 1 ` 1 “ 3 ` 1 and so on.
But such process is endless and impractical. There is a highly efficient way, based on the
below theorem and called positional numeral system, to denote any possible whole number
by using finite symbols. The usual Arabic numeral system used in elementary mathematics
is actually a positional numeral system with base ten. For more details about positional
notation, see chapter one of “Number Theory, George E. Andrews.”
Theorem 8. [Basis Representation Theorem.] Let b be a natural number larger than
1. For each n P N, there exist a unique k P N Y t0u and a unique function ∞f : Jk Y t0u Ñ
Jb´1 Yt0u, where J0 is defined to be empty set, such that f pkq ‰ 0 and n “ jPJk Yt0u f pjqbj .
Here we define that
b0 “ 1, b1 “ b, and bn “ b ¨ bn´1 for every n • 2.

12
Proof. See the page “Basis Representation Theorem” on Pr8fWiki.
As mentioned before, in the formal mathematics, we usually presume the existence
of natural numbers, use Peano axioms to describe the behaviour of the set of all natural
numbers, then enlarge the set of all natural numbers to a principal integral domain whose
elements are called integers, next consider the field of fractions of that integral domain
and call every element in the field a rational number, and finally construct the system of
real numbers via Cauchy sequences of rational numbers or Dedekind cuts. For a complete
introduction on whole procedure, see “Foundations of Analysis, Edmund Landau.”

September 20, 2019 by Ghi-Hong, Dan.

13

You might also like