Lemma 2.11 (Gauß 1801) :
∀
n
∈
N
:
n
=
∑
d
|
n
φ
(
d
)
{\displaystyle \forall n\in \mathbb {N} :n=\sum _{d|n}\varphi (d)}
.
Proof 1 :
We use the Möbius inversion formula, proven below without using this lemma, and lemma 2.10.
We have
∑
d
|
n
φ
(
d
)
=
S
φ
(
n
)
{\displaystyle \sum _{d|n}\varphi (d)=S_{\varphi }(n)}
and hence
φ
=
μ
∗
S
φ
{\displaystyle \varphi =\mu *S_{\varphi }}
by the Möbius inversion formula. On the other hand,
φ
(
n
)
=
μ
∗
I
1
(
n
)
{\displaystyle \varphi (n)=\mu *I_{1}(n)}
by lemma 2.10.
Hence, we obtain
μ
∗
S
φ
=
μ
∗
I
1
{\displaystyle \mu *S_{\varphi }=\mu *I_{1}}
, and by cancellation of
μ
{\displaystyle \mu }
(the arithmetic functions form an integral domain) we get the lemma.
◻
{\displaystyle \Box }
Proof 2 :
We use the converse of the Möbius inversion formula, proven below without using this lemma, and lemma 2.10.
Since
φ
(
n
)
=
μ
∗
I
1
(
n
)
{\displaystyle \varphi (n)=\mu *I_{1}(n)}
by lemma 2.10, we obtain from the converse of the Möbius inversion formula that
I
1
(
n
)
=
S
φ
(
n
)
{\displaystyle I_{1}(n)=S_{\varphi }(n)}
.
◻
{\displaystyle \Box }
Proof 3 :
We prove the lemma by double counting.
We first note that there are
n
{\displaystyle n}
many fractions of the form
m
n
{\displaystyle {\frac {m}{n}}}
,
1
≤
m
≤
n
{\displaystyle 1\leq m\leq n}
.
We now prove that there are also
∑
d
|
n
φ
(
d
)
{\displaystyle \sum _{d|n}\varphi (d)}
many fractions of this form. Indeed, each fraction
m
n
{\displaystyle {\frac {m}{n}}}
,
1
≤
m
≤
n
{\displaystyle 1\leq m\leq n}
can be reduced to
b
d
{\displaystyle {\frac {b}{d}}}
, where
gcd
(
b
,
d
)
=
1
{\displaystyle \gcd(b,d)=1}
.
d
{\displaystyle d}
is a divisor of
n
{\displaystyle n}
, since it is obtained by dividing
n
{\displaystyle n}
. Furthermore, for each divisor
d
{\displaystyle d}
of
n
{\displaystyle n}
there exist precisely
φ
(
d
)
{\displaystyle \varphi (d)}
many such fractions by definition of
φ
{\displaystyle \varphi }
.
◻
{\displaystyle \Box }
Proof 4 :
We prove the lemma by the means of set theory.
Define
S
n
,
d
:=
{
1
≤
l
≤
n
|
gcd
(
d
,
l
)
=
1
}
{\displaystyle S_{n,d}:=\{1\leq l\leq n|\gcd(d,l)=1\}}
. Then
S
n
,
d
=
{
d
k
|
1
≤
k
≤
n
/
d
,
gcd
(
k
,
n
)
=
1
}
=
d
S
n
/
d
,
1
{\displaystyle S_{n,d}=\{dk|1\leq k\leq n/d,\gcd(k,n)=1\}=dS_{n/d,1}}
. Since
|
S
n
/
d
,
1
|
=
φ
(
n
/
d
)
{\displaystyle |S_{n/d,1}|=\varphi (n/d)}
and
{
1
,
…
,
n
}
{\displaystyle \{1,\ldots ,n\}}
is the disjoint union of the sets
S
n
,
d
,
d
|
n
{\displaystyle S_{n,d},d|n}
, we thus have
n
=
∑
d
|
n
|
S
n
,
d
|
=
∑
d
|
n
d
φ
(
n
d
)
{\displaystyle n=\sum _{d|n}|S_{n,d}|=\sum _{d|n}d\varphi \left({\frac {n}{d}}\right)}
.
◻
{\displaystyle \Box }
The next theorem comprises one of the most important examples for a multiplicative function.
Theorem 2.12 (Euler 1761) :
Euler's totient function is multiplicative.
Proof 1 :
We prove the theorem using double counting (due to Kronecker ).
By definition of
φ
{\displaystyle \varphi }
, there are
φ
(
m
)
φ
(
n
)
{\displaystyle \varphi (m)\varphi (n)}
sums of the form
k
m
+
l
n
,
1
≤
k
≤
m
,
1
≤
l
≤
n
{\displaystyle {\frac {k}{m}}+{\frac {l}{n}},1\leq k\leq m,1\leq l\leq n}
,
where both summands are reduced. We claim that there is a bijection
{
k
m
+
l
n
|
1
≤
k
≤
m
,
1
≤
l
≤
n
,
k
m
,
l
n
reduced
}
→
{
r
m
n
|
1
≤
r
≤
r
m
n
reduced
}
{\displaystyle \left\{{\frac {k}{m}}+{\frac {l}{n}}{\big |}1\leq k\leq m,1\leq l\leq n,{\frac {k}{m}},{\frac {l}{n}}{\text{ reduced}}\right\}\to \left\{{\frac {r}{mn}}{\big |}1\leq r\leq {\frac {r}{mn}}{\text{ reduced}}\right\}}
.
From this would follow
φ
(
m
)
φ
(
n
)
=
φ
(
m
n
)
{\displaystyle \varphi (m)\varphi (n)=\varphi (mn)}
.
We claim that such a bijection is given by
k
m
+
l
n
↦
n
k
+
m
l
mod
m
n
n
m
{\displaystyle {\frac {k}{m}}+{\frac {l}{n}}\mapsto {\frac {nk+ml\mod mn}{nm}}}
.
Well-definedness: Let
k
m
{\displaystyle {\frac {k}{m}}}
,
l
n
{\displaystyle {\frac {l}{n}}}
be reduced. Then
k
m
+
l
n
=
k
n
+
l
m
mod
m
n
n
m
{\displaystyle {\frac {k}{m}}+{\frac {l}{n}}={\frac {kn+lm\mod mn}{nm}}}
is also reduced, for if
p
|
(
n
m
)
{\displaystyle p|(nm)}
, then without loss of generality
p
|
n
{\displaystyle p|n}
, and from
p
|
(
k
n
+
l
m
−
c
n
m
)
{\displaystyle p|(kn+lm-cnm)}
follows
p
|
l
{\displaystyle p|l}
or
p
|
m
{\displaystyle p|m}
. In both cases we obtain a contradiction, either to
gcd
(
m
,
n
)
=
1
{\displaystyle \gcd(m,n)=1}
or to
l
n
{\displaystyle {\frac {l}{n}}}
is reduced.
Surjectivity: Let
r
m
n
{\displaystyle {\frac {r}{mn}}}
be reduced. Using the Euclidean algorithm, we find
a
,
b
∈
N
{\displaystyle a,b\in \mathbb {N} }
such that
a
n
+
b
m
=
1
{\displaystyle an+bm=1}
. Then
r
a
n
+
r
b
m
=
r
{\displaystyle ran+rbm=r}
. Define
k
=
r
a
mod
m
{\displaystyle k=ra\mod m}
,
l
=
r
b
mod
n
{\displaystyle l=rb\mod n}
. Then
k
n
+
l
m
=
(
r
a
+
t
m
)
n
+
(
r
b
+
s
n
)
m
≡
r
mod
m
n
{\displaystyle kn+lm=(ra+tm)n+(rb+sn)m\equiv r\mod mn}
.
Injectivity: Let
k
n
+
l
m
≡
k
′
n
+
l
′
m
mod
m
n
{\displaystyle kn+lm\equiv k'n+l'm\mod mn}
. We show
k
=
k
′
{\displaystyle k=k'}
; the proof for
l
=
l
′
{\displaystyle l=l'}
is the same.
Indeed, from
k
n
+
l
m
≡
k
′
n
+
l
′
m
mod
m
n
{\displaystyle kn+lm\equiv k'n+l'm\mod mn}
follows
k
n
≡
k
′
n
mod
m
{\displaystyle kn\equiv k'n\mod m}
, and since
gcd
(
m
,
n
)
=
1
{\displaystyle \gcd(m,n)=1}
,
n
{\displaystyle n}
is invertible modulo
m
{\displaystyle m}
, which is why we may multiply this inverse on the right to obtain
k
≡
k
′
mod
m
{\displaystyle k\equiv k'\mod m}
. Since
1
≤
k
,
k
′
≤
m
{\displaystyle 1\leq k,k'\leq m}
, the claim follows.
◻
{\displaystyle \Box }
Proof 2 :
We prove the theorem from the Chinese remainder theorem .
Let
n
=
p
1
k
1
⋯
p
r
k
r
{\displaystyle n=p_{1}^{k_{1}}\cdots p_{r}^{k_{r}}}
. From the Chinese remainder theorem, we obtain a ring isomorphism
Z
/
n
→
Z
/
p
1
k
1
×
⋯
×
Z
/
p
r
k
r
{\displaystyle \mathbb {Z} /n\to \mathbb {Z} /p_{1}^{k_{1}}\times \cdots \times \mathbb {Z} /p_{r}^{k_{r}}}
,
which induces a group isomorphism
(
Z
/
n
)
×
→
(
Z
/
p
1
k
1
)
×
×
⋯
×
(
Z
/
p
r
k
r
)
×
{\displaystyle (\mathbb {Z} /n)^{\times }\to \left(\mathbb {Z} /p_{1}^{k_{1}}\right)^{\times }\times \cdots \times \left(\mathbb {Z} /p_{r}^{k_{r}}\right)^{\times }}
.
Hence,
|
(
Z
/
n
)
∗
|
=
∏
j
=
1
r
|
(
Z
/
p
j
k
j
)
∗
|
{\displaystyle \left|(\mathbb {Z} /n)^{*}\right|=\prod _{j=1}^{r}\left|\left(\mathbb {Z} /p_{j}^{k_{j}}\right)^{*}\right|}
, and from
∀
m
∈
N
:
φ
(
m
)
=
|
(
Z
/
m
)
∗
|
{\displaystyle \forall m\in \mathbb {N} :\varphi (m)=\left|(\mathbb {Z} /m)^{*}\right|}
follows the claim.
◻
{\displaystyle \Box }
Proof 3 : We prove the theorem from lemma 2.11 and induction (due to Hensel ).
Let
m
,
n
∈
N
{\displaystyle m,n\in \mathbb {N} }
such that
gcd
(
m
,
n
)
=
1
{\displaystyle \gcd(m,n)=1}
. By lemma 2.11, we have
m
=
∑
e
|
m
φ
(
e
)
{\displaystyle m=\sum _{e|m}\varphi (e)}
and
n
=
∑
d
|
m
φ
(
d
)
{\displaystyle n=\sum _{d|m}\varphi (d)}
and hence
m
n
=
φ
(
m
)
φ
(
n
)
+
φ
(
m
)
∑
d
|
n
,
d
<
n
φ
(
d
)
+
φ
(
n
)
∑
e
|
m
,
e
<
m
φ
(
e
)
+
∑
d
|
n
,
e
|
m
d
<
n
,
e
<
m
φ
(
d
)
φ
(
e
)
{\displaystyle mn=\varphi (m)\varphi (n)+\varphi (m)\sum _{d|n,d<n}\varphi (d)+\varphi (n)\sum _{e|m,e<m}\varphi (e)+\sum _{d|n,e|m \atop d<n,e<m}\varphi (d)\varphi (e)}
.
Furthermore, by lemma 2.11 and the bijection from the proof of theorem 2.8,
m
n
=
∑
f
|
m
n
φ
(
f
)
=
∑
e
|
m
,
d
|
n
φ
(
e
d
)
{\displaystyle mn=\sum _{f|mn}\varphi (f)=\sum _{e|m,d|n}\varphi (ed)}
.
By induction on
e
d
,
e
n
,
m
d
<
m
n
{\displaystyle ed,en,md<mn}
we thus have
m
n
=
φ
(
m
n
)
+
φ
(
m
)
∑
d
|
n
φ
(
d
)
+
φ
(
n
)
∑
e
|
m
φ
(
e
)
+
∑
e
|
m
,
d
|
n
e
<
m
,
d
<
n
φ
(
e
)
φ
(
d
)
{\displaystyle mn=\varphi (mn)+\varphi (m)\sum _{d|n}\varphi (d)+\varphi (n)\sum _{e|m}\varphi (e)+\sum _{e|m,d|n \atop e<m,d<n}\varphi (e)\varphi (d)}
.
◻
{\displaystyle \Box }
Proof 4 : We prove the theorem from lemma 2.11 and the Möbius inversion formula.
Indeed, from lemma 2.10 and the Möbius inversion formula, we obtain
φ
=
μ
∗
I
1
{\displaystyle \varphi =\mu *I_{1}}
,
which is why
φ
{\displaystyle \varphi }
is multiplicative as the convolution of two multiplicative functions.
◻
{\displaystyle \Box }
Proof 5 : We prove the theorem from Euler's product formula.
Indeed, if
m
=
P
κ
{\displaystyle m=P^{\kappa }}
and
n
=
Q
ι
{\displaystyle n=Q^{\iota }}
and
gcd
(
m
,
n
)
=
1
{\displaystyle \gcd(m,n)=1}
, then
P
∩
Q
=
∅
{\displaystyle P\cap Q=\emptyset }
and hence
∏
p
|
n
(
p
κ
p
−
p
κ
p
−
1
)
∏
q
|
n
(
q
ι
−
q
ι
q
−
1
)
=
φ
(
m
n
)
{\displaystyle \prod _{p|n}(p^{\kappa _{p}}-p^{\kappa _{p}-1})\prod _{q|n}(q^{\iota }-q^{\iota _{q}-1})=\varphi (mn)}
.
◻
{\displaystyle \Box }
Theorem 2.15 (Möbius inversion formula) :
Let
f
{\displaystyle f}
be an arithmetical function and define
S
f
(
n
)
:=
∑
d
|
n
f
(
d
)
=
f
∗
E
(
n
)
{\displaystyle S_{f}(n):=\sum _{d|n}f(d)=f*E(n)}
.
Then
f
=
μ
∗
S
f
=
∑
d
|
n
μ
(
d
)
S
f
(
n
d
)
{\displaystyle f=\mu *S_{f}=\sum _{d|n}\mu (d)S_{f}\left({\frac {n}{d}}\right)}
.
Proof :
By lemma 2.14 and associativity of convolution,
μ
∗
S
f
=
μ
∗
f
∗
E
=
μ
∗
E
∗
f
=
δ
∗
f
=
f
{\displaystyle \mu *S_{f}=\mu *f*E=\mu *E*f=\delta *f=f}
.
◻
{\displaystyle \Box }
Proof 1 :
We prove the theorem from lemma 2.10 and the fact that
φ
{\displaystyle \varphi }
is multiplicative.
Indeed, let
p
{\displaystyle p}
be a prime number and let
k
∈
N
{\displaystyle k\in \mathbb {N} }
. Then
φ
(
p
k
)
=
p
k
−
p
k
−
1
{\displaystyle \varphi (p^{k})=p^{k}-p^{k-1}}
, since
φ
(
p
k
)
=
(
μ
∗
I
1
)
(
p
k
)
=
∑
j
=
0
k
μ
(
p
j
)
p
k
−
j
{\displaystyle \varphi (p^{k})=(\mu *I_{1})(p^{k})=\sum _{j=0}^{k}\mu (p^{j})p^{k-j}}
by lemma 2.10. Therefore,
φ
(
n
)
=
∏
j
=
1
r
(
p
j
k
j
−
p
j
k
j
−
1
)
=
n
∏
j
=
1
r
{\displaystyle \varphi (n)=\prod _{j=1}^{r}\left(p_{j}^{k_{j}}-p_{j}^{k_{j}-1}\right)=n\prod _{j=1}^{r}}
,
where the latter equation follows from
n
∏
j
=
1
r
(
1
−
1
p
j
)
=
p
1
k
1
⋯
p
r
k
r
∏
j
=
1
r
(
1
−
1
p
j
)
=
∏
j
=
1
r
p
j
k
j
(
1
−
1
p
j
)
=
∏
j
=
1
r
(
p
j
k
j
−
p
j
k
j
−
1
)
{\displaystyle {\begin{aligned}n\prod _{j=1}^{r}\left(1-{\frac {1}{p_{j}}}\right)&=p_{1}^{k_{1}}\cdots p_{r}^{k_{r}}\prod _{j=1}^{r}\left(1-{\frac {1}{p_{j}}}\right)\\&=\prod _{j=1}^{r}p_{j}^{k_{j}}\left(1-{\frac {1}{p_{j}}}\right)\\&=\prod _{j=1}^{r}\left(p_{j}^{k_{j}}-p_{j}^{k_{j}-1}\right)\end{aligned}}}
.
◻
{\displaystyle \Box }
Proof 2 :
We prove the identity by the means of probability theory.
Let
n
∈
N
{\displaystyle n\in \mathbb {N} }
,
n
=
p
1
k
1
⋯
p
r
k
r
{\displaystyle n=p_{1}^{k_{1}}\cdots p_{r}^{k_{r}}}
. Choose
Ω
=
{
1
,
…
,
n
}
{\displaystyle \Omega =\{1,\ldots ,n\}}
,
F
=
2
Ω
{\displaystyle {\mathcal {F}}=2^{\Omega }}
,
P
(
A
)
:=
|
A
|
n
{\displaystyle P(A):={\frac {|A|}{n}}}
. For
j
∈
{
1
,
…
,
r
}
{\displaystyle j\in \{1,\ldots ,r\}}
define the event
E
p
j
:=
{
1
≤
k
≤
n
|
p
j
|
n
}
{\displaystyle E_{p_{j}}:=\{1\leq k\leq n|p_{j}|n\}}
. Then we have
P
(
E
p
1
¯
∩
⋯
∩
E
p
r
¯
)
=
φ
(
n
)
n
{\displaystyle P\left({\overline {E_{p_{1}}}}\cap \cdots \cap {\overline {E_{p_{r}}}}\right)={\frac {\varphi (n)}{n}}}
.
On the other hand, for each
J
=
{
j
1
,
…
,
j
l
}
⊆
{
1
,
…
,
r
}
{\displaystyle J=\{j_{1},\ldots ,j_{l}\}\subseteq \{1,\ldots ,r\}}
, we have
P
(
E
p
j
1
∩
⋯
∩
E
p
j
1
)
=
P
(
{
1
≤
k
≤
n
|
∏
j
∈
J
p
j
|
k
}
)
=
1
∏
j
∈
J
p
j
=
∏
j
∈
J
P
(
E
p
j
)
{\displaystyle {\begin{aligned}P\left(E_{p_{j_{1}}}\cap \cdots \cap E_{p_{j_{1}}}\right)&=P\left(\left\{1\leq k\leq n{\big |}\prod _{j\in J}p_{j}|k\right\}\right)\\&={\frac {1}{\prod _{j\in J}p_{j}}}=\prod _{j\in J}P\left(E_{p_{j}}\right)\end{aligned}}}
.
Thus, it follows that
E
p
1
,
…
,
E
p
r
{\displaystyle E_{p_{1}},\ldots ,E_{p_{r}}}
are independent. But since events are independent if and only if their complements are, we obtain
φ
(
n
)
n
=
P
(
E
p
1
¯
∩
⋯
∩
E
p
r
¯
)
=
P
(
E
p
1
¯
)
)
⋯
P
(
E
p
r
¯
)
=
∏
j
=
1
r
(
1
−
1
p
j
)
{\displaystyle {\frac {\varphi (n)}{n}}=P\left({\overline {E_{p_{1}}}}\cap \cdots \cap {\overline {E_{p_{r}}}}\right)=P\left({\overline {E_{p_{1}}}})\right)\cdots P\left({\overline {E_{p_{r}}}}\right)=\prod _{j=1}^{r}\left(1-{\frac {1}{p_{j}}}\right)}
.
◻
{\displaystyle \Box }
Proof 3 :
We prove the identity from the Möbius inversion formula and lemmas 2.9 and 2.10.
But by the Möbius inversion formula and since by lemma 2.10
S
φ
=
I
1
{\displaystyle S_{\varphi }=I_{1}}
,
∑
d
|
n
μ
(
d
)
n
d
=
μ
∗
S
φ
(
n
)
=
φ
(
n
)
{\displaystyle \sum _{d|n}\mu (d){\frac {n}{d}}=\mu *S_{\varphi }(n)=\varphi (n)}
.
◻
{\displaystyle \Box }
Proof 4 :
We prove the identity from the inclusion–exclusion principle .
Indeed, by one of de Morgan's rules and the inclusion–exclusion principle we have for sets
A
1
,
…
,
A
r
⊆
S
{\displaystyle A_{1},\ldots ,A_{r}\subseteq S}
|
⋂
j
=
1
r
A
j
|
=
|
S
∖
⋃
j
=
1
r
(
S
∖
A
j
)
|
=
|
S
|
−
|
⋃
j
=
1
r
(
S
∖
A
j
)
|
=
|
S
|
+
∑
∅
≠
J
⊆
{
1
,
…
,
r
}
(
−
1
)
|
J
|
|
⋂
j
∈
J
(
S
∖
A
j
)
|
=
∑
J
⊆
{
1
,
…
,
r
}
(
−
1
)
|
J
|
|
⋂
j
∈
J
(
S
∖
A
j
)
|
{\displaystyle {\begin{aligned}\left|\bigcap _{j=1}^{r}A_{j}\right|&=\left|S\setminus \bigcup _{j=1}^{r}\left(S\setminus A_{j}\right)\right|\\&=|S|-\left|\bigcup _{j=1}^{r}\left(S\setminus A_{j}\right)\right|\\&=|S|+\sum _{\emptyset \neq J\subseteq \{1,\ldots ,r\}}(-1)^{|J|}\left|\bigcap _{j\in J}\left(S\setminus A_{j}\right)\right|\\&=\sum _{J\subseteq \{1,\ldots ,r\}}(-1)^{|J|}\left|\bigcap _{j\in J}\left(S\setminus A_{j}\right)\right|\end{aligned}}}
,
where we use the convention that the empty intersection equals the universal set
S
{\displaystyle S}
.
Let now
n
=
p
1
k
1
⋯
p
r
k
r
{\displaystyle n=p_{1}^{k_{1}}\cdots p_{r}^{k_{r}}}
, and define
S
=
{
m
|
1
≤
m
≤
n
}
{\displaystyle S=\{m|1\leq m\leq n\}}
and
A
j
:=
{
l
∈
S
|
p
j
⧸
|
l
}
{\displaystyle A_{j}:=\{l\in S|p_{j}\not |l\}}
for
1
≤
j
≤
r
{\displaystyle 1\leq j\leq r}
. Since
φ
(
n
)
=
|
⋂
j
=
1
r
A
j
|
{\displaystyle \varphi (n)=\left|\bigcap _{j=1}^{r}A_{j}\right|}
,
we then have
φ
(
n
)
=
∑
J
⊆
{
1
,
…
,
r
}
(
−
1
)
|
J
|
|
⋂
j
∈
J
(
S
∖
A
j
)
|
{\displaystyle \varphi (n)=\sum _{J\subseteq \{1,\ldots ,r\}}(-1)^{|J|}\left|\bigcap _{j\in J}\left(S\setminus A_{j}\right)\right|}
.
But for each
J
⊆
{
1
,
…
,
r
}
{\displaystyle J\subseteq \{1,\ldots ,r\}}
, we have
|
⋂
j
∈
J
(
S
∖
A
j
)
|
=
{
1
≤
l
≤
n
|
∀
j
∈
J
:
p
j
|
l
}
=
{
1
≤
l
≤
n
|
(
∏
j
∈
J
p
j
)
|
l
}
=
n
∏
j
∈
J
p
j
{\displaystyle {\begin{aligned}\left|\bigcap _{j\in J}\left(S\setminus A_{j}\right)\right|&=\left\{1\leq l\leq n{\big |}\forall j\in J:p_{j}|l\right\}\\&=\left\{1\leq l\leq n{\big |}\left(\prod _{j\in J}p_{j}\right)|l\right\}={\frac {n}{\prod _{j\in J}p_{j}}}\end{aligned}}}
.
It follows
φ
(
n
)
=
n
∑
J
⊆
{
1
,
…
,
r
}
(
−
1
)
|
J
|
1
∏
j
∈
J
p
j
{\displaystyle \varphi (n)=n\sum _{J\subseteq \{1,\ldots ,r\}}(-1)^{|J|}{\frac {1}{\prod _{j\in J}p_{j}}}}
,
and since
∏
m
=
1
r
(
1
−
1
p
m
)
=
∑
J
⊆
{
1
,
…
,
r
}
(
−
1
)
|
J
|
1
∏
j
∈
J
p
j
{\displaystyle \prod {m=1}^{r}\left(1-{\frac {1}{p_{m}}}\right)=\sum _{J\subseteq \{1,\ldots ,r\}}(-1)^{|J|}{\frac {1}{\prod _{j\in J}p_{j}}}}
,
the theorem is proven.
◻
{\displaystyle \Box }