You are on page 1of 54

1

Appendix B

Constructing R

In this appendix, N = {0, 1, 2, 3, 4, . . .} is the set of all natural numbers.


B.1

From N to Z

Theorem (B.1.1)
1. If a, b, c N, then a + c = b + c if and only if a = b.
2. If a, b N and a b, then exists a unique c N such that a = b + c.
3. If a, b, c N and c 6= 0, then a c = b c if and only if a = b.

4. a + b = b + a.

Theorem (B.1.2)

Define on N N the relation :

(a, b) (c, d)

if and only if

a + d = b + c.

Then is an equivalence relation.


Proof

For transitivity, let (a, b) (c, d) and (c, d) (e, f ). Then a + d = b + c

and c + f = d + e. Adding the two equations, we get


a + d + c + f = b + c + d + e.

Then applying the cancellation law in Theorem (B.1.1), we get a + f = b + e, and


hence (a, b) (e, f ).
Definition (B.1.3)

Each equivalence class in N N by the equivalence relation

is called an integer. The integer (a, b)/ is denoted by [a b], and the set of
all the integers (the equivalence classes) is denoted by Z.
Example

For any a, b, c N, (a, b) (a + c, b + c). Hence,

[a b] = [(a + c) (b + c)].
Theorem (B.1.6)

Suppose a, b N.

If a > b, then (a, b) [n 0] for a unique positive natural number n.

If a = b, then (a, b) [0 0].


If a < b, then (a, b) [0 n] for a unique positive natural number n.

For the first and third cases, use Theorem (B.1.1) (2).
After this theorem, we see that {(n, 0) : n N} {(0, n) : n N} {(0, 0)} is a
complete set of equivalence class representatives, that is, each equivalence class
has exactly one representative in this set.
We now define the addition of integers and also an ordering of the integers.
Definition (B.1.7)

1. An order relation on Z is defined by : [a b] [c d] if and only if


a + d b + c.
2. The negative of the integer [a b], written as [a b], is defined as the integer
[b a].
3. Given any two integers [a b], [c d], their sum, written as [a b] + [c d],
is defined as [(a + c) (b + d)]. Their product, written as [a b] [c d], is
[(ac + bd) (ad + bc)].

The above definition are for equivalence classes and so we have to check, in every
case, the said relation does not depend on the particular representatives chosen.

For instance, to ensure that the relation [a b] [c d] is well-defined, we have


to show the following. If [a0 b0] = [a b] and [c0 d0] = [c, d], then
[a b] [c d] if and only if [a0 b0] [c0 d0]. Indeed, we have a0 + b = b0 + a
and c0 + d = d0 + c. The relation [a b] [c d] implies a + d b + c, which
means for a unique n N, a + d + n = b + c. Hence
a0 + d0 + a + d + n = a0 + d0 + b + c = b0 + a + c0 + d = b0 + c0 + a + d, and so (by
applying Theorem (B.1.1) (2)), a0 + d0 + n = b0 + c0. Thus, a0 + d0 b0 + c0, and
[a0 b0] [c0 d0].
With the same a, a0, . . . etc., we find that [a b] [c d] = [a0 b0] [c0 d0] since
we easily verify
(ac + bd)(a0d0 + b0c0) = (ad + bc)(a0c0 + b0d0).

Why the sum and product are defined in such a way?


Note, if we think of [a b] as a b in Z, then
(a b) + (c d) = (a + c) (c + d) and (a b)(c d) = (ac + bd) (ad + bc).

Theorem (B.1.11)
Proof

is a total ordering on Z.

It is straightforward to see that is reflexive and antisymmetric. To

prove transitivity, let [a b] [c d] and [c d] [e f ]. Therefore


a + d b + c and c + f d + e. Adding the two inequalities, we get
a + d + c + f b + c + d + e, and hence a + f b + e, as desired.

is a total ordering, since given any [a b] and [c d], we have either


a + d b + c or b + c a + d. Thus, either [a b] [c d] or [c d] [a b]
holds.
Exercise

Show that for any two integers x, y, x y if and only if y x.

(Be careful, x and y are of the form [a b] and [c d].)


Theorem (B.1.13)

The addition and multiplication defined on Z is

commutative and associative, that is, for any integers x, y, z,


x + y = y + x,

(x + y) + z = x + (y + z),

xy = yx,

(xy)z = x(yz).

Furthermore, multiplication is distributive over addition:


x(y + z) = xy + xz.

These follows from the same properties in N. As an illustration, we prove


distributivity. Let x = [a b], y = [c d], z = [e f ]. Then
x(yz) = [ab][(ce+df )(cf +de)] = [(a(ce+df )+b(cf +de))(a(cf +de)+b(ce+df ))]
and

(xy)z = [(ac+bd)(ad+bc)][ef ] = [((ac+bd)e+(ad+bc)f )((ac+bd)f +(ad+bc)e)].

10

The right hand sides of the above two equalities are identical. This proves the
distributive property.
Theorem (B.1.14)

For each integer x, x + (x) = [0 0].

Where are the natural numbers in Z?


There is a copy of N in Z!
Theorem (B.1.17)

Let e : N Z be given by e(n) = [n 0] for all n N.

Then e has the following properties.


1. e is one-to-one.

11

2. For any a, b N, e(a) e(b) if and only if a b. (This means that e is an


order isomorphism, the order in N is preserved in e(N).)

3. For any a, b N, e(a+b) = e(a)+e(b), and e(ab) = e(a)e(b). (This shows that
both the additive structure and the multiplicative structure of N are preserved
in e(N).)

With these properties, we say that N and e(N) are order isomorphic, and we can
regard e(N) as a copy of N. We write n instead of e(n) = [n 0] for each natural
number n. In this way, we find that N is a subset of Z.
Proof

We prove, as an illustration, that the multiplicative structure is

12

preserved. Let a, b N. Then


e(ab) = [ab 0] = [a 0] [b 0] = e(a) e(b).

Now the set of all integers Z together with the addition defined has the following
properties.
1. The addition is commutative and associative.
2. There exists a unique integer,namely, the integer [0 0] = 0 such that for any
integer x, x + 0 = 0 + x = x.
3. For any integer u, there exists a unique integer v such that u + v = v + u = 0.

13

(Note, the integer v depends on u. In fact, v = u.)


We say that (Z, +) is an abelian group. In general, if G is a non-empty set with a
binary operation defined on it (such as addition) and satisfies the above three
properties, we say that G is an abelian group. Without the property of
commutativity, G is simply called a group.
Note

N is not a group under addition. Why?

Exercise

(B.1.18)

Show that every integer can either be written as n or n

for a unique natural number n. Furthermore n = n if and only if n = 0.


Exercise (B.1.19)

Prove the following facts about arithmetic in Z. Let

14

a, b, c Z.
1. a + 0 = a.
2. a 0 = 0.
3. a 1 = a.
4. (1) a = a.
5. a + c = b + c a = b.
6. If c 6= 0 and a c = b c, then a = b.

15

7. a b > 0 if and only if a > 0, b > 0 or a < 0, b < 0.


Definition (B.1.20)

We now define subtraction on Z by a b = a + (b) for

all a, b Z. Be careful, subtraction is neither commutative nor associative.


Exercise

Show that multiplication in Z is also distributive over subtraction.

Exercise

Show that if m, n N then m n = [m n]. Also for any

a, b Z, a b is the unique integer x which satisfies b + x = a.


B.2

From Z to Q

The extension from N to Z allows us to solve equations of the form b + x = a for


any integers a and b. But the frequently encountered linear equation 2x = 1 is still

16

not soluble in Z. In order to be able to solve linear equations like ax = b where


a, b are integers, we have to find ways to divide 1 by any non-zero integer. In the
extension from N to Z, we think of integers as m n where m and n are natural
numbers. So to extend Z further to cope with the need of division, we think of
a/b where a and b are integers, and b 6= 0.
The idea of extending Z to Q is quite similar to the case of going from N to Z.
We will define an equivalence relation on Z Z and then consider Q as the
quotient set. There will be a copy of Z in this quotient set.
Theorem (B.2.1)

Let H = Z (Z\{0}) = {(a, b) : a, b Z, b 6= 0}. Define on

H the relation by (a, b) (c, d) if and only if ad = bc. Then is an


equivalence relation.

17

The verification of the required properties, that is, reflexive, symmetric and
transitive is quite straightforward. But, be careful, so far, we have only defined
addition, subraction and multiplication of integers, no division of integers is
allowed in the argument.
Definition (B.2.2)

The equivalence classes of H are called rational numbers.

The set of all the rational numbers is denoted as Q. The equivalence class (a, b) is
denoted as [a/b].
Why do we exclude pairs of the form (a, 0) from H?
(Division by 0?)
Exercise

Show that [a/b] = [(a)/(b)]. Hence each rational number is of the

18

form [a/b] where b is a positive integer.


Definition (B.2.5)
1. The negative of the rational number [a/b], written as [a/b], is the rational
number [(a)/b].
2. If a 6= 0, the reciprocal of [a/b] is the rational number [a/b]1 = [b/a].
3. Given any two rational numbers [a/b], [c/d], their sum is [(ad + bc)/(bd)] and
their product is [(ac)/(bd)].
In these definitions, we need to check that they are good, that is, the definitions
do not depend on the representatives chosen from each equivalence class.

19

Theorem (B.2.8), (B.2.10)

Both the addition and multiplication defined on

Q are commutative and associative.


Theorem (B.2.9)

Q contains an additive identity and additive inverses.

That is, for any x Q, we have


x + [0/1] = x,
x + (x) = [0/1].
Remember each x Q is of the form [a/b] where a, b Z and b 6= 0.
Theorem (B.2.11)
inverses.

Q contains a multiplicative identity and multiplicative

20

For any x Q, x [1/1] = x.


For any x Q such that x 6= [0/1], x x1 = [1/1].

Theorem (B.2.12)
Proof

Multiplication distributes over addition in Q.

Let x = [a/b], y = [c/d], z = [e/f ] be any three rational numbers.


x(y + z) = [a/b][(cf + de)/(df )] = [(a(cf + de))/(b(df ))].

xy+xz = [(ac)/(bd)]+[(ae)/(bf )] = [(acbf +bdae)/(bdbf )] = [(acf +dae)/(bdf )].


The right hand sides of the above are identical. Hence x(y + z) = xy + xz.

21

Definition (B.2.13)

The set of the positive rational numbers is

Q+ = {r Q : r = [a/b] for some a > 0, b > 0}.

We note that r Q+ if and only if every (c, d) r must have c and d both
positive or both negative. (Use Exercise (B.1.19)7.)
Theorem (B.2.15) (Order Axiom)

Q+ is closed under addition and

multiplication. That is, if x, y Q+, then x + y Q+, x y Q+.


Definition (B.2.16) Define a relation on Q as follows: For x, y Q, x y if
and only if either x = y or (x) + y Q+.

22

Theorem (B.2.16)
Proof

is a total ordering on Q.

To prove that is transitive. Let x y, y z. May assume

x 6= y, y 6= z. Then (x) + y Q+ and (y) + z Q+. Adding these yields


(x) + z Q+. Hence x z.
To prove asymmetry, let x y and y x. If x 6= y, then (x) + y Q+ and
(y) + x Q+. Adding these, we get 0 = (x) + y + (y) + x Q+, which is a
contradiction.
Given two rational numbers x = [a/b], y = [c/d], we find that
(x) + y = [(ad + bc)/(bd)] and (y) + x = [(cb + ad)/(bd)]. If x 6= y
( ad 6= bc) , then either (x) + y Q+ or (y) + x Q+. Hence, either x < y

23

or y < x.
The next theorem shows that Q contains a copy of Z.
Theorem (B.2.18)

Define a function f : Z Q by f (n) = [n/1] for all n Z.

Let a, b Z.
1. f is one-to-one (but not onto).
2. f (a) f (b) if and only if a b.
3. f (a) = f (a).
4. f (a + b) = f (a) + f (b).

24

5. f (ab) = f (a)f (b).


Proof
1. For any a, b Z, f (a) = f (b) implies [a/1] = [b/1]. Hence, (a, 1) (b, 1), a = b.
2. If a b and a 6= b, we have (f (a))+f (b) = [a/1]+[b/1] = [(a)/1]+[b/1] =
[(a + b)/1] Q+. Thus f (a) f (b).
Conversely, f (a) f (b) f (a) = f (b) or (f (a)) + f (b) Q+. In the latter
case a + b > 0, that is b > a. In the former case, since f is one-to-one, we
have a = b. In any case, we have a b.
5. f (ab) = [(ab)/1] = [a/1][b/1] = f (a)f (b).

25

Remark

1. The properties 1, 4 and 5 shows that the subset f (Z) = {[n/1] : n Z} in Q


is isomorphic to Z, that is, apart from being in one-to-one correspondence with
Z, the subset has the same additive and multiplicative structure as Z. There is
no real structural difference between Z and f (Z). The property 2 further shows
that the ordering in Z is also preserved in f (Z). In this sense, we say that Z is a
subset of Q. (The actual meaning is that the copy f (Z) of Z is a subset of Q.)

2. The set Q with addition is an abelian group. Also, the set Q := Q\{0} is an
abelian group under multiplication. The two operations, namely, addition and
multiplication satisfy further the distributive laws. We say that Q is a field. It is

26

further called an ordered field with the endowed ordering.

Exercise (B.2.19)

For any a, b Q, a 6= 0, the equation ax = b has the

unique solution x = ba1 Q.


We adopt the common notation [a/b] =

a
b

for rational numbers.

Apart from being able to divide by non-zero numbers, Q has another important
property that Z does not possess, namely, Q is dense.
Theorem

Q is dense. That is, for any two distinct rational numbers a and b,

say a < b, there exists a rational number c such that a < c < b. Hence there
are infinitely many such c.

27

Proof

c = a 21 + b 21 is a rational number and it satisfies a < c < b. To

prove a < c, let a = [d/e], b = [f /g], and we may assume further that
e > 0, g > 0. Then
a + c = [(ef dg)/(2eg)].
This belongs to Q+ is a consequence of the fact that
a + b = [(ef dg)/(eg)] Q+. Hence a < c. The other inequality c < b is
proved by the same argument.
B.3

From Q to R

Definition (B.3.1)
such that

A Dedekind cut x of Q is a nonempty proper subset of Q

28

1. If a x and b Q such that b < a, then b x.


2. If a x, then there is a c x for which a < c.

Exercise (B.3.3)

Let a Q. Show that both {b Q : b < a} and

{b Q : b3 < 2} are Dedekind cuts.


Intuitively, a Dedekind cut is the set of all rational numbers lying on the left hand
side of some point (not including the point itself) on the real number line. But be
careful, not all Dedekind cuts can be written in the form {b Q : b < a}. Only
those cuts at rational numbers can be written in such form.
Exercise (B.3.2)

29

1. A Dedekind cut x has no greatest element. (Reason: (B.3.1)(2))

2. A Dedekind cut x has an upper bound. (Proof. Let v Q\x. Then v is an upper
bound of x, otherwise there exists u x such that v < u. Then by B.3.1(1),
v x, a contradiction.)

Definition (B.3.5)

A Dedekind cut of the rationals Q is called a real number.

The set of real numbers is denoted by R.

Definition (B.3.6)

Let x be a Dedekind cut of Q.

30

x is a rational cut if for some b Q,


x = {a Q : a < b}.

x is an irrational cut if it is not a rational cut.


Since each Dedekind cut is a subset of Q, we can define the relation on the set
of all cuts by . More precisely, for cuts x, y, we write x y if x y. Therefore
is a partial ordering. But this is indeed a total ordering. Suppose x 6= y and
there is a v Q such that v y\x, then, by the argument in (B.3.2)(2), v is an
upper bound of x. According to Definition (B.3.1)(1), y contains every element of
x. Hence, x y, x y.

31

For each rational number a, we let a be the Dedekind cut {b Q : b < a} in R.


So, a corresponds to a rational cut.
We now define the addition and multiplication of real numbers.
Definition (B.3.11)

Let x, y R. Then x + y is defined as the cut

{c Q : c = a + b for some a x and some b y}.

Denote the above set by S. We can see that S is really a cut.


First, S is clearly non-empty. Since the cut x has an upper bound, u, say, and y
has an upper bound, v, say, then u + v is an upper bound of S. So S is a proper
subset of Q.

32

Let c = a + b S and let d be a rational number such that d < c. Then


d = (a c + d) + b. Notice that (a c + d) is a rational number less than a,
hence by the definition of the cut x, (a c + d) x. Thus, d S. Also, since
a x (respectively b y), there exists e x (respectively f y) such that a < e.
(respectively b < f ). Then the rational number g = e + f belongs to S and
c = a + b < g. With all these, we showed that S is a cut: x + y is a real number.
Exercise

Show that if x and y are rational cuts, so is x + y.

Theorem (B.3.13)

Addition on R is commutative and associative.

Furthermore, 0 is an additive identity on R, that is, x + 0 = x for any x R.


Proof

Commutativity and associativity of addition on R follow from the same

33

properties on Q. Now for any cut x, we show that x + 0 = x. First we recall that

0 = {b Q : b < 0}.

For any element c x + 0, c = u + v, where u is a certain rational number in x


and v is a negative rational number. Hence, c is a rational number less than u,
which lies in x. Thus, by definition of the cut x, c x. In other words, x + 0 x.
Conversely, let w x. By definition of the cut x, there exists a rational number
h x such that w < h. The rational number ` = w h is negative, and hence
belongs to 0. Now w = h + ` x + 0. Thus, x x + 0. Combining the two
inequalities, we get x + 0 = x.

34

The negative of x
In order to be able to solve the equation x + y = 0, we need to define x, the
negative of x.
First if x is an irrational cut, we define
x = {u : u Q\x}.
Let S denote the set on the right hand side. We first show that S is really a cut.
Clearly S is non-empty since x is a proper subset of Q. Also S 6= Q since x is
non-empty: for any b x, b 6 S. It remains to confirm the two properties 1. and
2. in Definition (B.3.1).

35

For 1. let a S and b Q such that b < a. Then b Q and b > a. Note
that b cannot be an element of x, otherwise, a would also be an element of x,
contradicting the fact that a S. Hence, b Q\x. By the definition of
S, b S.
For 2. let a S. Then a 6 x. Therefore a is greater than any element v x,
otherwise, by the fact that Q is totally ordered and Definition (B.3.1)1., we would
have a x, a contradiction.
Now we invoke the assumption that x is not a rational cut. So, x is not consisting
of all the rational numbers which are less than a. That means, there is another
rational number, w, say, such that w < a and w
/ x. Then w Q\x, which
implies w S, and a < w. Hence property 2. in Definition (B.3.1) holds.

36

In the above argument, we see that the assumption of x being an irrational cut
was used only in the proof of 2. When x is a rational cut, say x = {z Q : z < c},
where c is a certain rational number, the argument in 2. is still valid as long as a
is not equal to c (but the argument breaks down when a = c). Therefore we
define the negative of x when it is a rational cut as above by
x = {u : u Q\x}\{c} = {v Q : v < c}.
Verify that this again defines a cut (which is also rational).
From this, it is easy to prove the following.
Exercise (B.3.16)

For any x R, (x) = x.

37

Theorem (B.3.17)

For any x R, x + (x) = 0. (Thus, every element of

R has an additive inverse.)


Proof

We first consider the case that x is a rational cut:


x = {v Q : v < c} for a certain rational number c.

Then x = {u Q : u < c}. Hence

x+(x) = {a+b : a x, b (x)} = {a+b : a < c, b < c} {w Q : w < 0} = 0


Conversely, for any w 0, w Q, w < 0, we can write
w = (c + w 21) + (c + w 21),

38

where the number in the first bracket is a rational number less than c (hence
belongs to x) and the second is a rational number less than c (hence belongs to
x). Thus, w x + (x).
Next we consider an irrational cut x. Then x = {v : v
/ x}, and
x + (x) = {u + v : u x, v
/ x}.
Recall that v
/ x v > u for all u x. Hence u + v < 0 and x + (x) 0.
Conversely, let w 0. So w Q, w < 0. Take any v
/ x. The rational numbers
v > v + w > v + 2w > v + 3w > v + 4w > v + 5w > decreases to minus
infinity. So eventually, for some k, v + kw will be less than a certain element of x.
Let k 1 be the first natural number such that v + kw x. Then

39

v + (k 1)w
/ x. Now
w = (v + kw) + ((v + (k 1)w)).
The first number on the right hand side belongs to x, while the second one, by
definition, belongs to x. Hence w x + (x). This proves that 0 x + (x).
Definition (B.3.8)

The set of positive real numbers in defined to be


R+ = {x R : 0 x, x 6= 0}.

Exercise (B.3.9)

Show that

R+ = {x R : 0 x} = {x R : x contains some positive rational numbers}.

40

Solution

If x contains 0, then by Definition ((B.3.1)1., x contains all rational

numbers less than 0. Hence 0 x. Furthermore, 0 x but 0


/ 0, thus 0 6= x.

Conversely, if x R+, then 0 x and 0 6= x. Thus all rational numbers less


than 0 belong to x, and there exists a rational number r x\0. Such rational
number r must be greater than or equal to zero. If r = 0, we are done. Or else,
0 < r, and since x is a cut, 0 x. Thus, x R+ 0 x, as desired.

To prove that the two sets above are equal, we notice that, since x is a cut, x
contains a positive rational number implies x contains all rational numbers less
than this number. In particular, x contains 0. Conversely, by Definition (B.3.1)1.,
if x contains 0, then x contains some positive rational number.

41

Theorem (B.3.18)

For every real number x, exactly one of the following is

true.
x R+
x = 0
x R+

Definition (B.3.19)

If x, y R+, then we define x y, the product of x and y

to be the set of all c Q such that c < ab for some a x and some b y for
which a > 0, b > 0.

42

Exercise (B.3.20)

Show that x y defined above is really a cut and it is

positive.
Solution

First we verify the two property stipulated in Definition (B.3.1). If

u xy, then there exist positive rational numbers a x, b y such that u < ab.
For any rational number v < u, we have v < ab and hence v xy.
Next, let u < ab as above. Since a x, there exists d x such that a < d.
Similarly, there exists e y such that b < e. Then u < ab < de and Definition
(B.3.1)2. is satisfied.
Finally, that xy is positive follows immediately from Exercise (B.3.9) since 0 < ab,
0 xy.

43

Defintion (B.3.21)

Define the multiplication on R as follows. Let x, y R.

If x, y R+, then xy is defined as in Definition (B.3.20).


If x = 0 or y = 0, then xy = 0.
If x R+ and y R+, then xy = (x(y)).
If x R+ and y R+, then xy = ((x)y).
If x, y R+, then xy = (x)(y).

Since the multiplication of positive real numbers are easily seen to be

44

commutative and associative, the same hold for multiplication of all real numbers.
This involves case by case checking according to Definition (B.3.21).
Theorem (B.3.23)

Multiplication on R is commutative and associative.

Furthermore, the real number 1 = {x Q : x < 1} is a multiplicative identity


on R: For all x R, x 1 = x.
Proof

We prove that 1 is a multiplicative identity on R.

First consider x R+. Let c x 1. Then c < ab for some positive rational
numbers a x, b 1. Since 0 < b < 1, we find that c < a and hence c x. This
shows x 1 x.
Conversely, for any r x, there exists a positive rational number u x such that

45

r < u. Again, there exists a positive rational number v x such that u < v. Now
let w be the positive rational number u/v, which is less than 1. Hence, w 1
and r < u = vw. This shows that r x 1. Combining the two parts, we have
x 1 = x for x R+.
In case x = 0, then by Definition (B.3.21), x 1 = 0 = x.
If x R+, then, again, by Definition (B.3.21),
x 1 = ((x) 1) = (x) = x,
as desired. This completes the proof of the Theorem (B.3.22).
Theorem (B.3.23)

Multiplication distributes over addition in R, that is, for

46

any x, y, z R, we have x(y + z) = xy + xz.


Question: How to define the reciprocal of non-zero real numbers?
Definition For any positive real number x, we define the reciprocal of x to be

{r Q : r < c1},
if x is the rational cut

x1 =
{r Q : r < c},

{r Q : r < a1 for ALL positive a x}, if x is an irrational cut.


When x R+ we define x1 to be (x)1.
We need to show that x1, as defined above, is really a cut. In the case that x is
a positive rational cut, this is clear, since c is a positive rational number, so is c1.

47

Suppose x is a positive irrational cut and let u x1. We may assume that u > 0.
Any rational number v < u is, by definition, again in x1. It remains to show that
there exists a rational number w > u, w x1. Consider the rational numbers
1
1
1
1
u+ >u+ >u+ >u+
> .
2
4
8
16
If none of these belongs to x1, that means, for each k > 0, there exists a
k
corresponding positive rational number ak x such that a1
. This
k u+2

inequality is equivalent to
(u + 2k )1 ak .
By Definition (B.3.1)1., (u + 2k )1, and all the rational numbers less than this
belong to x. This shows that indeed x = {v Q : v < u1}, a rational cut. This

48

contradicts our assumption. Hence, there exists a positive rational number of the
form u + 2k which belongs to x1. Thus property 2. in Definition (B.3.1) is
verified.
Here we notice that, x1 is also positive when x R+.
Obviously, when x R+, x1 is also a cut and x1 R+.
Theorem (B.3.24)
Proof

For any non-zero real number x, we have x x1 = 1.

We prove this when x is a positive real number. The other case is easy.

Let z be any element of x x1. Then by definition, z < cd where c is a certain


positive rational number in x and d is a certain positive rational number in x1.

49

Hence, d < a1 for any positive rational number a x. Therefore, in particular,


d < c1, and so dc < 1. Thus, z < 1 and hence, z 1.
Conversely, we prove 1 x x1. Let z 1. Without loss of generality, we may
assume that z is positive. So 0 < z < 1.
We first prove the following auxiliary result. Let x be a positive real number. For
any given positive (small) rational number , there exist positive rational
numbers c x and d
/ x such that d c < d.
This result can be seen as follows. Since x is positive, there exists c1 x with
c1 > 0. There also exists d1 Q\x, 0 < c1 < d1. If d1 c1 < d1, then
c = c1, d = d1 will be a solution. Otherwise consider the rational number

50

1
2 (c1

+ d1) which lies half way between c1 and d1. We change c1 to this number

and then denoted it as c2 if it is in x (and set d2 = d1), or else, change d1 to this


number and rename it as d2, and take c2 = c1. Repeating this consideration on
the rational number 12 (c2 + d2) and call this c3 if it is in x, otherwise call it d3. We
notice that this procedure produces the rational numbers
0 < c1 < c2 < c3 < c4 < < d4 < d3 < d2 < d1, where each ci lies in x and
each di lies in Q\x. Furthermore,
0 < di ci = 2i+1(d1 c1).
Hence as i tends to infinity, di ci goes to zero. For sufficiently large i, we have
di ci < c1 < di,

51

as desired.
We now continue our proof of 1 x x1, by showing that z < uv for some
positive rational numbers u x, v x1. By the above result, let c x and
d Q\x such that 0 < c < d, d c < (1 z)d. Since d Q\x, we have d > w
for all w x. So, d1 x1. Taking now u = c and v = d1, we find that
z < uv, and this finishes the proof.
Theorem

For any real numbers a and b, a 6= 0, the equation ax = b has the

unique solution x = a1b.


As in the previous cases where we extend from N to Z and from Z to Q, we can
see that the function g : Q R given by g(r) = r for each r Q is one-to-one,

52

and also preserves the addition and multiplication in Q, that is


g(r + s) = g(r) + g(s),

g(rs) = g(r)g(s) for all r, s Q.

Hence g(Q) is copy of Q lying in R. By abuse of language, we say that Q lies in R.


Z distinguishes itself from N by the property that it is an abelian group, in which
subtractions can be performed. Q distinguishes itself from Z by the property that
it is a field, in which division by non-zero numbers can be carried out, and also Q
is dense. The extension of Q to R adds the following.
Theorem (B.3.10)-Least upper bound property

Any non-empty subset of

R which is bounded above has a least upper bound in R.

53

Proof

First, for any two cuts x and y, their union is also a cut which is indeed

equal to max{x, y}.


Now let A R be bounded non-empty. Let b = xAx. It is easy to see that b
satisfies the two requirements in Definition (B.3.1). Also b is non-empty, and not
equal to Q since A is bounded above. Hence b is also a cut (that is a real
number). Clearly, for every x A, x b. So b is an upper bound of A. If c is any
upper bound of A, then x c for every x A, hence xAx c. This shows that
b c, and b is the least upper bound of A.
As a consequence of the above theorem, every Cauchy sequence in R converges.

The real number system is still inadequate in the sense that, many polynomial

54

equations, such as the simple quadratic equation x2 + 1 = 0, is not solvable. To


remedy this, we can further extend the real number system to the field of the
complex numbers: {a + bi : a, b R}. In the field of complex numbers, any
polynomial equation can be completely solved. We omit the details of this.

You might also like