Let polynomials $f,g$ be represented as

\[f(x)=a_mx^m+a_{m-1}x^{m-1}+\dots+a_0,\tag1\] \[g(x)=b_nx^n+b_{n-1}x^{n-1}+\dots+a_0.\tag2\]

Define $h=fg$ by

\[h(x)=c_{m+n}x^{m+n}+c_{m+n-1}x^{m+n-1}+\dots+c_0.\tag3\]

Comparing coefficients, we deduce that for $0\le k\le m+n$, there is

\[c_k=\sum_{\substack{0\le r\le m\\0\le s\le n\\ r+s=k}}a_r b_s.\tag4\]

In particular, $c_0=a_0b_0$ and $c_{m+n}=a_mb_n$. In this article, we will be using these formulae to deduce a variety of results concerning polynomial factorizations.

Reducibility of polynomials $\mathbb Z[x]$

In general rings $R$, a polynomial $h\in R[x]$ is said to be irreducible if it cannot be expressed as a product of two other polynomials in $R[x]$, each having a degree strictly less than that of $h$. By invoking the well-ordering property of natural numbers, we immediately deduce that every polynomial in $R[x]$ is a product of irreducible factors in $R[x]$.

To study the reducibility of polynomials in $\mathbb Z[x]$, we focus on primitive polynomials, whose coefficients do not possess a common divisor. Every polynomial $\mathbb Z[x]$ is an integer of some primitive polynomial. It should also be noticed that primitiveness is preserved under multiplication:

Let $f$ and $g$ in (1) and (2) be primitive. Suppose there is a prime $p$ dividing all the $c_k$'s. Because $f$ and $g$ are primitive, there exists unique integers $0\le\mu\le m$ and $0\le\nu\le n$ satisfying

\[p\vert \gcd(a_0,a_1,\dots,a_{\mu-1}),\quad p\nmid a_\mu,\] \[p\vert \gcd(b_0,b_1,\dots,b_{\nu-1}),\quad p\nmid b_\nu.\]

Plugging this result into (4), we have

\[c_k\equiv\sum_{\substack{\mu\le r\le m\\ \nu\le s\le n\\r+s=k}}a_rb_s\pmod p,\]

Setting $k=\mu+\nu$, this becomes

\[c_{\mu+\nu}\equiv a_\mu b_\nu\not\equiv 0\pmod p,\]

which is a direct contradiction to our assumption that $p\vert c_{\mu+\nu}$. Therefore, $h$ must be primitive. In general, we have

Lemma 1: Let $f,g\in\mathbb Z[x]$ and $h=fg$ satisfy the expansions (1), (2), and (3). Then

\[\gcd(a_0,a_1,\dots,a_m)\gcd(b_0,b_1,\dots,b_n)=\gcd(c_0,c_1,\dots,c_{m+n}).\tag5\]

From this property, we find that for polynomials with integral coefficients, their reducibility in $\mathbb Z[x]$ is equivalent to the reducibility in $\mathbb Q[x]$:

Lemma 2: A polynomial in $\mathbb Z[x]$ is reducible over $\mathbb Z[x]$ if and only if it is reducible over $\mathbb Q[x]$.

Proof. Reducibility in $\mathbb Z[x]$ trivially implies reducibility in $\mathbb Q[x]$. For the converse, without loss of generality, assume $H\in\mathbb Z[x]$ is primitive and is factored as $H=FG$ for some $F,G\in\mathbb Q[x]$. Choose integers $M,N$ such that $f(x)=M\cdot F(x),g=M\cdot G(x)\in\mathbb Z[x]$ and let $h(x)=MN\cdot H(x)$. Using the notations of (1), (2), and (3), it follows from Lemma 1 that

\[MN=\gcd(a_0,a_1,\dots,a_m)\gcd(b_0,b_1,\dots,b_n).\]

This indicates that when

\[f_1(x)={M\over \gcd(a_0,a_1,\dots,a_m)}\cdot F(x),\tag6\] \[g_1(x)={N\over \gcd(b_0,b_1,\dots,b_n)}\cdot G(x),\tag7\]

we have $f_1,g_1\in\mathbb Z[x]$ and $H=f_1g_1$, so $H$ is reducible over $\mathbb Z[x]$ as well.

From the arguments in Lemma 2, we can also relate the divisibility of monic polynomials in $\mathbb Q[x]$ and $\mathbb Z[x]$.

Lemma 3: For monic $H\in\mathbb Z[x]$, if $H=FG$ for some monic $F,G\in\mathbb Q[x]$, then $F,G\in\mathbb Z[x]$.

Proof. Using the notations in the proof of Lemma 2, because the leading coefficient of $H$ is the product of leading coefficients of $f_1$ and $g_2$, we must have

\[M=\gcd(a_0,a_1,\dots,a_m),\quad N=\gcd(b_0,b_1,\dots,b_n)\]

in (6) and (7), so $f_1=F$ and $g_1=G$, which indicates that $F,G\in\mathbb Z[x]$.

These three results are all known as Gauss's lemma. In the rest of the article, we will use these results to produce a wide class of irreducible polynomials.

Eisenstein's criterion

If $h\in\mathbb Z[x]$ and $p$ is any prime satisfying $p\vert c_k$ for $0\le k<m+n$ and $p\nmid c_{m+n}$, then $p\vert c_0=a_0b_0$, so by Euclid's lemma, we assume without loss of generality that $p\vert a_0$. This means there exists some $1\le\mu\le m$ satisfying

\[p\vert\gcd(a_0,a_1,\dots,a_{\mu-1}),\quad p\nmid a_\mu,\]

By (4), we have $c_\mu\equiv a_\mu b_0\pmod p$. Because $\mu<m+n$, we must have $p\vert b_0$, so $p^2\vert c_0$. The contrapositive of this result becomes a sufficient condition to ensure the irreducible for a given polynomial with integral coefficients:

Theorem 1 (Eisenstein's criterion): For $f(x)=a_mx^m+a_{m-1}x^{m-1}+\dots+a_0\in\mathbb Z[x]$, if there exists some prime $p$ satisfying

\[p\vert\gcd(a_0,a_1,\dots,a_{m-1}),\quad p\nmid a_m,\quad p^2\nmid a_0,\]

then $f$ is irreducible over $\mathbb Z[x]$.

Irreducibility of cyclotomic polynomials

In this section, we prove that every cyclotomic polynomial $\Phi_n(x)$ is irreducible over $\mathbb Q[x]$. When $n=p$ is a prime, observe that

\[\Phi_p(x)=x^{p-1}+x^{p-2}+\dots+1={x^p-1\over x-1},\]

so we have

\[\Phi_p(x+1)={(x+1)^p-1\over x}=x^{p-1}+\sum_{m=1}^{p-2}\binom p{m+1}x^m+p.\]

Thus, it follows from Eisenstein's criterion that $\Phi_p(x+1)$, hence $\Phi_p(x)$, is irreducible over $\mathbb Z[x]$.

To prove the irreducibility of $\Phi_n(x)$ for general $n$, we need to invoke some algebraic number theory

Algebraic numbers and minimal polynomials

In general, a complex number $\alpha$ is algebraic if and only if it is a zero of some nonconstant polynomial in $\mathbb Q[x]$. Under more restrictions, this polynomial can be uniquely determined from $\alpha$:

Theorem 2: For each algebraic number $\alpha$, there exists a unique monic polynomial $f\in\mathbb Q[x]$ satisfying

  1. $f(\alpha)=0$.
  2. For each $g\in\mathbb Q[x]$ satisfying $g(\alpha)=0$, $f$ divides $g$.
  3. $f$ is irreducible over $\mathbb Q[x]$.

Such $f$ is called the minimal polynomial of $\alpha$.

Proof. By definition, the set

\[S_n=\lbrace f\in\mathbb Q[x]:f\text{ monic}\wedge \deg f=n\wedge f(\alpha)=0\rbrace\]

cannot be empty for all $n\in\mathbb N$. Let $d$ be the smallest positive integer for which $S_d\ne\varnothing$.

Let $f\in S_d$. We first show that $f$ satisfies the above conditions and then prove that it is unique.

If $g\in\mathbb Q[x]$ is nonconstant and satisfies $g(\alpha)=0$, there exists $q,r\in\mathbb Q[x]$ satisfying $g(x)=q(x)f(x)+r(x)$ and $0\le\deg r<d$, so we have $r(\alpha)=0$. If $d_1=\deg r>0$, then $S_{d_1}\ne\varnothing$ will cause a contradiction, so we must have $r\equiv0$. Hence, $f$ divides $g$.

If $f=gh$ for some nonconstant $g,h\in\mathbb Q[x]$, then we must have $g(\alpha)=0$ or $h(\alpha)=0$. Then $S_{d_2}\ne\varnothing$ for some $0<d_2<d$, which also causes a contradiction. Hence, $f$ is irreducible.

For uniqueness, if $f,g\in S_d$ are distinct, then $(f-g)(\alpha)=0$ and $0<d_3=\deg(f-g)<d$, so $f-g\in S_{d_3}$, which is yet another contradiction.

We will prove the irreducibility of $\Phi_n(x)$ by showing that it is the minimal polynomial of each $n$'th primitive root of unity.

Minimality of cyclotomic polynomials

Let $\omega$ be a primitive $n$'th root of unity and $f\in\mathbb Q[x]$ be its minimal polynomial. Then clearly $f(x)\vert \Phi_n(x)$, and $f(x)=\Phi_n(x)$ will follows from

Lemma 4: If $\gcd(m,n)=1$, then $f(\omega^m)=0$.

Proof. Because $\omega^m$ is also a primitive root of unity and $m$ is a product of primes that do not divide $n$, we reduce our consideration to the case when $m=p$ is a prime not dividing $n$.

Because $\Phi_n(x)$ is monic, it follows from Theorem 2 that is some monic $g\in\mathbb Q[x]$ satisfying

\[\Phi_n(x)=f(x)g(x).\tag8\]

By Lemma 3, we find that $f,g\in\mathbb Z[x]$. This allows us to play around $f,g$ using congruences. If $f(\omega^p)\ne 0$, it follows from $\Phi_m(\omega^p)=0$ and (8) that $g(\omega^p)=0$. This indicates the existence of some monic $h\in\mathbb Q[x]$ for which

\[g(x^p)=f(x)h(x),\tag9\]

and $h$, by Lemma 3, is also an element belonging to $\mathbb Z[x]$. When $g(x)=\sum_ju_jx^j$, because $(a+b)^p=a^p+b^p$ holds in fields of characteristic $p$ and $t^p\equiv t\pmod p$ for all $t\in\mathbb Z$, we have

\[g(x^p)\equiv\sum_j(u_jx^j)^p\equiv\left(\sum_ju_jx^j\right)^p=[g(x)]^p\pmod{p\mathbb Z[x]},\]

This means when interpreting in $\mathbb F_p[x]$, there is $g(x^p)=[g(x)]^p$. Let $f_1 \in\mathbb Z[x]$ be an irreducible divisor of $f$ in the sense of $\mathbb F_p[x]$. Then it follows from (9) that $f_1$ divides $g^p$, hence dividing $g$ in the sense of $\mathbb F_p[x]$, so by (8), $\Phi_n$ is divisible by $(f_1)^2$ in the sense of $\mathbb F_p[x]$.

Because $x^n-1=\Phi_n(x)\phi(x)$ for some $\phi\in\mathbb Z[x]$, we conclude that $(f_1)^2$ divides $x^n-1$ in the sense of $\mathbb F_p[x]$. This means that $\varphi(x)=x^n-1$ has a repeated root in the algebraic closure of $\mathbb F_p$, so there is some $\alpha\in\overline{\mathbb F_p}$ such that $\varphi'(\alpha)=n\alpha^{n-1}=0$, but because $\alpha\ne0$ and $p\nmid n$, we arrive at a contradiction. This indicates that $g(\omega^p)\ne0$, so $f(\omega^p)=0$.

Conclusion

In this article, we began our investigation from the arithmetical property of the coefficients of products of two polynomials. By analyzing the common divisors of the coefficients, we derived Gauss's lemma and Eisenstein's criterion. Subsequently, we combined these results with tools from algebraic number theory to prove that every cyclotomic polynomial is irreducible over $\mathbb Q[x]$.