Hostname: page-component-848d4c4894-wzw2p Total loading time: 0 Render date: 2024-06-13T02:36:52.667Z Has data issue: false hasContentIssue false

CHARACTERISATION OF PRIMES DIVIDING THE INDEX OF A CLASS OF POLYNOMIALS AND ITS APPLICATIONS

Published online by Cambridge University Press:  01 April 2024

ANUJ JAKHAR*
Affiliation:
Department of Mathematics, Indian Institute of Technology (IIT) Madras, Chennai, India e-mail: anujjakhar@iitm.ac.in
Rights & Permissions [Opens in a new window]

Abstract

Let ${\mathbb {Z}}_{K}$ denote the ring of algebraic integers of an algebraic number field $K = {\mathbb Q}(\theta )$, where $\theta $ is a root of a monic irreducible polynomial $f(x) = x^n + a(bx+c)^m \in {\mathbb {Z}}[x]$, $1\leq m<n$. We say $f(x)$ is monogenic if $\{1, \theta , \ldots , \theta ^{n-1}\}$ is a basis for ${\mathbb {Z}}_K$. We give necessary and sufficient conditions involving only $a, b, c, m, n$ for $f(x)$ to be monogenic. Moreover, we characterise all the primes dividing the index of the subgroup ${\mathbb {Z}}[\theta ]$ in ${\mathbb {Z}}_K$. As an application, we also provide a class of monogenic polynomials having non square-free discriminant and Galois group $S_n$, the symmetric group on n letters.

Type
Research Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction and statements of results

Let $K = {\mathbb Q}(\theta )$ be an algebraic number field with $\theta $ in the ring ${\mathbb {Z}}_K$ of algebraic integers of K and let $f(x)$ of degree n be the minimal polynomial of $\theta $ over the field ${\mathbb Q}$ of rational numbers. Let $d_K$ denote the discriminant of K and $D_f$ the discriminant of the polynomial $f(x)$ . It is well known that $d_K$ and $D_f$ are related by the formula

$$ \begin{align*}D_f = [{\mathbb{Z}}_K : {\mathbb{Z}}[\theta]]^2d_K.\end{align*} $$

We say that $f(x)$ is monogenic if ${\mathbb {Z}}_K = {\mathbb {Z}}[\theta ]$ , or equivalently, if $D_f = d_K$ . In this case, $\{1, \theta , \ldots , \theta ^{n-1}\}$ is an integral basis of K and K is a monogenic number field. A number field K is called monogenic if there exists some $\alpha \in {\mathbb {Z}}_K$ such that ${\mathbb {Z}}_K = {\mathbb {Z}}[\alpha ]$ .

The determination of monogenity of an algebraic number field is one of the classical and important problems in algebraic number theory. An arithmetic characterisation of monogenic number fields is a problem due to Hasse (see [Reference Hasse6]). Gaál’s book [Reference Gaál5] provides some classifications of monogenity in lower degree number fields. Using Dedekind’s Index Criterion, Jakhar et al. [Reference Jakhar, Khanduja and Sangwan8] gave necessary and sufficient conditions for ${\mathbb {Z}}_K = {\mathbb {Z}}[\theta ]$ when $\theta $ is a root of an irreducible trinomial $x^n+ax^m+b \in {\mathbb {Z}}[x]$ having degree n, providing infinitely many monogenic trinomials. Jones [Reference Jones9] computed the discriminant of the polynomial $f(x) = x^n+a(bx+c)^m \in {\mathbb {Z}}[x]$ with $1\leq m < n$ and proved that when $\gcd (n, mb) = 1$ , there exist infinitely many values of a such that ${\mathbb {Z}}_K = {\mathbb {Z}}[\theta ]$ where $K = {\mathbb Q}(\theta )$ and $\theta $ has minimal polynomial $f(x)$ . He also conjectured that if $\gcd (n,mb) = 1$ and a is a prime number, then the polynomial ${x^n+a(bx+c)^m \in {\mathbb {Z}}[x]}$ is monogenic if and only if $n^n + (-1)^{n+m}b^n(n-m)^{n-m}m^ma$ is square-free. Recently, Kaur and Kumar [Reference Kaur and Kumar12] proved that this conjecture is true. Jones [Reference Jones11] gave infinite families of number fields K generated by a root $\theta $ of an irreducible quadrinomial, quintinomial or sextinomial for which ${\mathbb {Z}}_K = {\mathbb {Z}}[\theta ].$ He also proved in [Reference Jones10] that if $\theta $ is a root of an irreducible polynomial of the type ${f(x) = x^p-2ptx^{p-1}+p^2t^2x^{p-2}+1 \in {\mathbb {Z}}[x]}$ and p is an odd prime with $p\nmid t$ , then ${\mathbb {Z}}_K\neq {\mathbb {Z}}[\theta ].$

Let $K = {\mathbb Q}(\theta )$ be an algebraic number field where $\theta $ has minimal polynomial $f(x) = x^n+a(bx+c)^m$ over ${\mathbb Q}$ with $1\leq m<n$ . We characterise all the primes dividing the index of ${\mathbb {Z}}[\theta ]$ in ${\mathbb {Z}}_K$ . As an application, we provide necessary and sufficient conditions for ${\mathbb {Z}}_K = {\mathbb {Z}}[\theta ].$ We also establish a more general result confirming [Reference Jones9, Conjecture 4.1]. Further, we give a class of monogenic polynomials of prime degree q having non square-free discriminant and Galois group isomorphic to the symmetric group $S_q$ . In some examples, we determine the index $[{\mathbb {Z}}_K:{\mathbb {Z}}[\theta ]]$ as well.

Throughout the paper, $D_f$ will stand for the discriminant of $f(x) = x^n +a(bx+c)^m$ with $1\leq m<n$ . Jones [Reference Jones9, Theorem 3.1] proved that the discriminant $D_f$ is given by

(1.1) $$ \begin{align} D_f = (-1)^{\binom{n}{2}} c^{n(m-1)}a^{n-1}[c^{n-m}n^{n} + (-1)^{m+n} a b^{n} m^{m}(n-m)^{n-m} ]. \end{align} $$

We prove the following result.

Theorem 1.1. Let $K= {\mathbb Q}(\theta )$ be an algebraic number field with $\theta $ in the ring ${\mathbb {Z}}_K$ of algebraic integers of K having minimal polynomial $f(x) = x^n+a(bx+c)^m$ , $1\leq m < n$ , over $\mathbb Q$ . A prime factor p of the discriminant $D_f$ of $f(x)$ does not divide $[{\mathbb {Z}}_{K} : {\mathbb {Z}}[\theta ]]$ if and only if p satisfies one of the following conditions:

  1. (i) when $p \mid a$ , then $p^2 \nmid ac$ ;

  2. (ii) when $p\nmid a$ , $p\mid b$ , $p\mid c$ , then $m=1$ and $p^2\nmid c$ ;

  3. (iii) when $p \nmid ac$ and $p\mid b$ with $j \geq 1$ as the highest power of p dividing n, then either $p\mid b_1$ and $p\nmid c_2$ or p does not divide $b_1[(ac^m)b_{1}^{n}+(-c_2)^{n}]$ , where

    $$ \begin{align*}b_1 = \frac{mabc^{m-1}}{p}, \quad c_2 = \frac{1}{p}[ac^m+(-ac^m)^{p^j}];\end{align*} $$
  4. (iv) when p does not divide $ab$ and $p\mid c$ , then $m=1$ and either $p \mid b_2 $ with $p\nmid c_1 $ or p does not divide $b_2[(ab ) b_2 ^{n -1} + (-c_1)^{n - 1}]$ , where

    $$ \begin{align*}b_2= \frac{1}{p}[ab+(-ab)^{p^{l}}], \quad c_1= \frac{ac}{p} \quad\text{and}\quad n-1 = p^ls', p\nmid s';\end{align*} $$
  5. (v) when p does not divide $abc$ and $p\mid m$ with $n = s'p^k,~m = sp^k,~p \nmid \gcd (s', s)$ , then the polynomials

    $$ \begin{align*}x^{s'} + a(bx+c)^s \quad\text{and}\quad \frac{1}{p}\bigg[pt(bx+c)^m-\sum\limits_{j=1}^{p^k-1}\binom{p^k}{j}(x^{s'})^{p^k-j}(a(bx+c)^s)^{\kern1.2pt j}\bigg]\end{align*} $$
    are coprime modulo p, where $t\in {\mathbb {Z}}$ is an integer such that $a = a^{p^k} + pt$ ;
  6. (vi) when $p\nmid abcm$ , then $p^2$ does not divide $D_f$ .

The following corollary is immediate. It extends the main results of [Reference Jones9].

Corollary 1.2. Let $K = {\mathbb Q}(\theta )$ and $f(x) = x^n + a(bx+c)^m$ be as in Theorem 1.1. Then ${\mathbb {Z}}_{K} = {\mathbb {Z}}[\theta ]$ if and only if each prime p dividing $D_f$ satisfies one of the conditions (i)–(vi) of Theorem 1.1.

If we take $\gcd (n,mb) = 1$ and $c=1$ , then conditions (ii)–(v) of Theorem 1.1 are not possible. So in the special case when $c=1$ and $\gcd (n,mb) = 1$ , the above corollary provides the main result of [Reference Kaur and Kumar12] stated below. This gives infinite families of monogenic polynomials and establishes a more general form of [Reference Jones9, Conjecture 4.1].

Corollary 1.3 [Reference Kaur and Kumar12].

Let $f(x) = x^n + a(bx+1)^m \in {\mathbb {Z}}[x]$ be a monic irreducible polynomial of degree n with $\gcd (n,mb) = 1$ . Then ${\mathbb {Z}}_K = {\mathbb {Z}}[\theta ]$ if and only if each prime p dividing $D_f$ satisfies either (i) $p\mid a$ and $p^2\nmid a$ or (ii) $p\nmid a$ and $p^2\nmid D_f$ .

The following proposition follows readily from the proof of Theorem 1.1(vi) and is of independent interest.

Proposition 1.4. Let $f(x) = x^q+a(bx+c)^m\in {\mathbb {Z}}[x]$ , $1\leq m < q$ , be an irreducible polynomial of prime degree. If there exists a prime p such that p divides $D_f$ and $p^2\nmid D_f$ with $p\nmid abcm$ , then the Galois group of $f(x)$ is $S_q$ .

The following result is an immediate consequence of Corollary 1.3 and Proposition 1.4. It provides a class of monogenic polynomials having non square-free discriminant and Galois group equal to a symmetric group.

Corollary 1.5. Let m be a positive odd integer and $f(x) = x^q+a(bx+1)^m\in {\mathbb {Z}}[x]$ be a polynomial having prime degree $q\geq 3$ with $q\nmid b$ . If $a \not \in \{0, \pm 1\}$ and $D_f/a^{q-1}$ are square-free numbers, then $f(x)$ is a monogenic polynomial having Galois group $S_q$ .

The following example is an application of Theorem 1.1, Corollary 1.3 and Proposition 1.4. In this example, $K={\mathbb Q}(\theta )$ with $\theta $ a root of $f(x)$ .

Example 1.6. Let p be a prime number. Consider $f(x) = x^p+p(x+1)^{p-1}$ . Note that $|D_f| = p^p(p^{p-1}-(p-1)^{p-1})$ . Using Proposition 1.4, it is easy to check that the Galois group of $f(x)$ is $S_p$ . By Corollary 1.3, ${\mathbb {Z}}_K = {\mathbb {Z}}[\theta ]$ if and only if $p^{p-1}-(p-1)^{p-1}$ is square-free. We now compute $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ for $p<20.$ For $p=2,3,7,11,17$ , it can be verified that the number $p^{p-1}-(p-1)^{p-1}$ is square-free; and hence ${\mathbb {Z}}_K = {\mathbb {Z}}[\theta ]$ . Next we calculate the exact value of $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ corresponding to $p = 5, 13$ and $19$ .

  1. (i) For $p = 5$ , it can be easily checked that $D_{f} = 5^5\cdot 3^2\cdot 41.$ In view of Theorem 1.1(i), $5$ does not divide $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ . Also, $3$ divides $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ and $41$ does not divide $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ by Theorem 1.1(vi). Since $ D_{f} $ = $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]^2\cdot d_K$ , where $d_K$ is the discriminant of K, we see that $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ is $3$ when $p = 5$ .

  2. (ii) Consider $p = 13$ . One can verify that $D_{f} = 13^{13}\cdot 5^2\cdot 7\cdot 67\cdot 109\cdot 157\cdot 229\cdot 313$ . By Theorem 1.1(i), $13$ does not divide $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ . Also in view of Theorem 1.1(vi), $5$ divides $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ and the primes $7, 67, 109, 157, 229, 313$ do not divide $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ . Since the exact power of $5$ dividing $D_{f}$ is $2$ , $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]=5$ .

  3. (iii) When $p = 19$ , then one can check that the prime factorisation of $D_{f}$ is given by $19^{19}\cdot 7^3\cdot r$ with r a square-free number. Arguing as above, $19$ and each prime p dividing r do not divide $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ and $7$ divides $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ . Therefore, $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]=7$ .

2 Proof of Theorem 1.1

In what follows, while dealing with a prime number p, for a polynomial $h(x)$ in $ {\mathbb {Z}}[x]$ , we shall denote by $\bar {h}(x)$ the polynomial over ${\mathbb {Z}}/p{\mathbb {Z}}$ obtained by interpreting each coefficient of $h(x)$ modulo p.

We first state the following well-known theorem. The equivalence of assertions (i) and (ii) of the theorem was proved by Dedekind (see [Reference Cohen2, Theorem 6.1.4], [Reference Dedekind3]). A simple proof of the equivalence of assertions (ii) and (iii) is given in [Reference Jakhar, Khanduja and Sangwan7, Lemma 2.1].

Theorem 2.1. Let $f(x) \in {\mathbb {Z}}[x]$ be a monic irreducible polynomial having the factorisation $\bar {g}_1(x)^{e_1} \cdots \bar {g}_{t}(x)^{e_{t}}$ modulo a prime p as a product of powers of distinct irreducible polynomials over ${\mathbb {Z}}/p{\mathbb {Z}}$ with each $g_i(x) \in {\mathbb {Z}}[x]$ monic. Let $K={\mathbb Q}(\theta )$ with $\theta $ a root of $f(x)$ . Then the following statements are equivalent:

  1. (i) p does not divide $[{\mathbb {Z}}_K:{\mathbb {Z}}[\theta ]]$ ;

  2. (ii) for each i, either $e_i =1 $ or $\overline g_i (x)$ does not divide $\overline M(x)$ where

    $$ \begin{align*}M(x) = \frac{1}{p}(f(x) - g_1 (x)^{e_1} \cdots g_{t} (x)^{e_{t}} );\end{align*} $$
  3. (iii) $f(x)$ does not belong to the ideal $\langle p, g_i(x)\rangle ^2$ in ${\mathbb {Z}}[x]$ for any i, $1\leq i\leq t$ .

The next lemma (see [Reference Jakhar, Khanduja and Sangwan7, Corollary 2.3]) is easily proved using the binomial theorem.

Lemma 2.2. Let $k\geq 1$ be the highest power of a prime p dividing a number $n=p^{k}s'$ and c be an integer not divisible by p. If $\bar {g}_1(x)\cdots \bar {g}_r(x)$ is the factorisation of $x^{s'}-\bar {c}$ into a product of distinct irreducible polynomials over ${\mathbb {Z}}/p{\mathbb {Z}}$ with each $g_i(x)\in \mathbb Z[x] $ monic, then

$$ \begin{align*}x^n-c=(g_1(x)\cdots g_r(x)+pH(x))^{p^k}+pg_1(x)\cdots g_r(x)T(x)+p^2U(x)+c^{p^k}-c\end{align*} $$

for some polynomials $H(x), T(x), U(x)\in \mathbb Z[x]$ .

Proof of Theorem 1.1.

Let p be a prime dividing $D_f$ . In view of Theorem 2.1, p does not divide $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ if and only if $f(x) \not \in \langle p, g(x) \rangle ^2$ for any monic polynomial $g(x) \in {\mathbb {Z}}[x]$ which is irreducible modulo p. Note that $f(x) \not \in \langle p,g(x)\rangle ^2 $ if $\bar g(x) $ is not a repeated factor of $\bar {f}(x) $ . We prove the theorem case by case.

Case (i): $p\mid a$ . In this case, $f(x) \equiv x^n \pmod p$ . Clearly, $f(x) \in \langle p, x\rangle ^2$ if and only if $p^2$ divides $ac^m$ ; consequently, $p\nmid [{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ if and only if $p^2\nmid ac$ .

Case (ii): $p \nmid a$ and p divides both b and c. In this situation, $f(x) \equiv x^n \pmod p$ and it is easy to see that $f(x) \in \langle p, x\rangle ^2$ if and only if $p^2$ divides $c^m$ . Therefore, $p\nmid [{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ if and only if $p^2\nmid c^m$ , that is, $m=1$ and $p^2\nmid c$ .

Case (iii): $p\nmid ac$ and $p\mid b$ . As $p\mid D_f$ , it is clear from (1.1) that $p\mid n$ . Write $n = p^j s', ~p\nmid s'$ . By the binomial theorem,

$$ \begin{align*} f(x) \equiv x^n + ac^m \equiv (x^{s'} + ac^m)^{p^{j}}\pmod p. \end{align*} $$

Let $\bar {g}_1(x) \cdots \bar {g}_{t}(x)$ be the factorisation of $h(x) = x^{s'} + ac^m$ over ${\mathbb {Z}}/p{\mathbb {Z}}$ , where $g_i(x) \in {\mathbb {Z}}[x]$ are monic polynomials which are distinct and irreducible modulo p. Write $h(x)$ as $g_1(x)\cdots g_t(x) + pH(x)$ for some polynomial $H(x) \in {\mathbb {Z}}[x]$ . Applying Lemma 2.2 to $h(x)$ and keeping in view that

$$ \begin{align*}f(x) = h(x^{p^{j}}) + a(bx)^m+\binom{m}{1}a(bx)^{m-1}c + \cdots + \binom{m}{m-1}a(bx)c^{m-1}\end{align*} $$

with $p\mid b$ , we see that

(2.1) $$ \begin{align}\hspace{-5pt} f(x)&=\bigg(\prod\limits _{i=1}^{t} g_i(x)+pH(x)\bigg)^{p^j}+pT(x)\prod\limits _{i=1}^{t} g_i(x)+p^2U(x) + ac^m +(-ac^m)^{p^j} + ma(bx)c^{m-1} \end{align} $$

for some polynomials $T(x), U(x) \in {\mathbb {Z}} [x].$ As $j\geq 1$ , the first three summands on the right-hand side of (2.1) belong to $\langle p, g_i(x) \rangle ^2$ for each i, $1\leq i\leq t$ . So $f(x) \in \langle p, g_i(x) \rangle ^2$ for some i, $1\leq i\leq t$ , if and only if $mabc^{m-1}x + ac^m +(-ac^m)^{p^j} = p(b_1 x +c_2)$ does so. Clearly, $p(b_1 x +c_2 )$ belongs to $\langle p, g_i(x) \rangle ^2$ for some i if and only if either p divides both $b_1 $ , $c_2 $ or $p\nmid b_1$ and the polynomials $\bar b_1 x +\bar c_2 ,~ x^n +\overline {ac^m} $ have a common root. One can easily check that the polynomials $\bar b_1 x + \bar c_2$ and $x^n + \overline {ac^m}$ have a common root if and only if $( -{\bar c_2}/{\bar b_1} ) ^{n }= -\overline {ac^m}$ , that is, if and only if $p\mid [(-ac^m) b_1 ^{n} -(-c_2)^{n}] $ . Hence, $f(x) \not \in \langle p, g_i(x) \rangle ^2$ for any i if and only if either $p\mid b_1 $ and $p\nmid c_2 $ or p does not divide $b_1 [(ac^m) b_1 ^{n} +(-c_2)^{n}]$ . This proves the theorem in case (iii) by virtue of Theorem 2.1.

Case (iv): $p\nmid ab$ and $p\mid c$ . In this case, $\overline {f}(x) = x^m(x^{n-m}+\overline {ab^m})$ . If $m\geq 2,$ then x is a repeated factor and it is easy to check that $f(x) \in \langle p, x\rangle ^2$ , that is, p always divides $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ by Theorem 2.1. So, assume now that $m =1.$ By (1.1), ${p\mid (n-1)}$ , say ${n-1 = p^l s'}$ with $p\nmid s'$ . Write $x^{s'} + ab = g_1(x)\cdots g_t(x) + pH(x)$ , where $g_1(x), \ldots , g_{t}(x)$ are monic polynomials which are distinct as well as irreducible modulo p and ${H(x) \in {\mathbb {Z}}[x]}$ . Applying Lemma 2.2 to $h(x)$ = $x^{s'} + ab$ , we can write ${f(x)= x (x^{n-1} +ab) +ac}$ as

(2.2) $$ \begin{align} f(x)=x\bigg[\bigg(\prod\limits_{i=1}^{t}g_i(x)+pH(x)\bigg)^{p^l}+pT(x)\prod\limits_{i=1}^{t}g_i(x)+ p^2U(x) + ab+(-ab)^{p^l}\bigg]+ac, \end{align} $$

where $T(x), U(x)$ belong to ${\mathbb {Z}} [x]$ . Note that $x,\bar {g}_1(x) , \ldots , \bar {g}_t(x)$ are distinct irreducible factors of $\bar f(x)$ . Since $l\geq 1$ , the first three summands inside the square bracket on the right-hand side of (2.2) belong to $\langle p, g_i(x) \rangle ^2$ for each i, $1\leq i\leq t$ . So $f(x) \in \langle p, g_i(x) \rangle ^2$ for some i, $1\leq i\leq t$ , if and only if $(ab +(-ab)^{p^l})x + ac = p(b_2 x +c_1)$ does so. Clearly, the polynomial $p(b_2 x +c_1 )$ belongs to $\langle p, g_i(x) \rangle ^2$ for some i if and only if either p divides both $b_2 $ , $c_1 $ or $p\nmid b_2 $ and the polynomials $\bar b_2 x +\bar c_1 ,~ x^{n-1} +\overline { ab} $ have a common root. The polynomials $\bar {b}_2 x + \bar {c}_{1}$ and $x^{n - 1} + \overline {ab}$ have a common root if and only if $( -{\bar c_1}/{\bar b_2} ) ^{n -1} = -\overline { ab} $ . Thus, $f(x) \in \langle p, g_i(x) \rangle ^2$ for some i if and only if either p divides both $b_2 $ , $c_1 $ or $p\nmid b_2 $ and $p\mid [(-ab ) b_2 ^{n -1} - (-c_1)^{n - 1}]$ . So we conclude that p does not divide $[{\mathbb {Z}}_K:{\mathbb {Z}}[\theta ]]$ if and only if $m=1$ and either $p \mid b_2 $ with $p\nmid c_1 $ or p does not divide $b_2[(ab ) b_2 ^{n -1} + (-c_1)^{n - 1}]$ . This proves the theorem in case (iv).

Case (v): $p \nmid abc$ and $p \mid m$ . As $p \mid D_f$ , p divides n in view of (1.1). Write $n = s'p^{k}$ , $m \kern1.3pt{=}\kern1.3pt sp^{k}$ with $p\kern1.3pt{\nmid}\kern1.3pt \gcd (s', s)$ so that $f(x) \kern1.3pt{=}\kern1.3pt (x^{s'})^{p^k} \kern1.3pt{+}\kern1.3pt a(bx+c)^{sp^k}$ . Set ${h(x) \kern1.3pt{=}\kern1.3pt x^{s'} \kern1.3pt{+}\kern1.3pt a(bx+c)^s}$ . Let $t\in {\mathbb {Z}}$ be an integer such that $a = a^{p^k} + pt$ . Then one can easily check that ${f(x) \equiv h(x)^{p^{k}}}$ (mod p). Let $h(x) \equiv g_1(x)^{d_1}\cdots g_{t}(x)^{d_{t}} \pmod p$ be the factorisation of $h(x)$ into a product of irreducible polynomials modulo p with $g_{i}(x) \in {\mathbb {Z}}[x]$ monic and $d_i> 0$ . Write

$$ \begin{align*} f(x) &= h(x)^{p^{k}} +pt(bx+c)^m-\sum\limits_{j=1}^{p^k-1}\binom{p^k}{j}(x^{s'})^{p^k-j}(a(bx+c)^s)^{\kern1.2pt j}. \end{align*} $$

Now $f(x) = (g_1(x)^{d_1} \cdots g_{t}(x)^{d_{t}})^{p^{k}} + pM(x)$ for some $M(x) \in {\mathbb {Z}}[x]$ . Since $k> 0$ , by Theorem 2.1, p does not divide $[{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]$ if and only if $\overline {M}(x)$ is coprime to $\bar {h}(x)$ , which holds if and only if the polynomial

$$ \begin{align*}\frac{1}{p}\bigg[pt(bx+c)^m-\sum\limits_{j=1}^{p^k-1}\binom{p^k}{j}(x^{s'})^{p^k-j}(a(bx+c)^s)^{\kern1.2pt j}\bigg]\end{align*} $$

is coprime to $h(x)$ modulo p. This proves the theorem in case (v).

Case (vi): $p\nmid abcm$ . Since $p\mid D_f$ and $p\nmid abcm$ , it follows from (1.1) that $p\nmid n(n-m)$ . Let $\beta $ be a repeated root of $\bar {f}(x) = x^n+\bar {a}(\bar {b}x+\bar {c})^m$ in the algebraic closure of ${\mathbb {Z}}/p{\mathbb {Z}}$ . Then

(2.3) $$ \begin{align} \bar f(\beta)=\beta^n+\bar a (\bar b\beta+\bar c)^m= \bar 0; \quad \bar f'(\beta)=\bar n\beta^{n-1}+ \bar {m}\bar{a} \bar b (\bar b\beta + \bar{c})^{m-1}= \bar 0. \end{align} $$

On substituting $\bar {n}\beta ^{n-1} = -\bar {m}\bar {a} \bar b (\bar b\beta + \bar {c})^{m-1}$ in the first equation of (2.3), we see that

$$ \begin{align*} (b\beta + c)^{m-1}(ab(n-m)\beta + nac) \equiv 0 \pmod p. \end{align*} $$

Observe that $(b\beta + c) \not \equiv 0 \pmod p$ , otherwise $\beta = \bar {0}$ in view of the first equation of (2.3) which is not possible as $p\nmid ac$ . Therefore, keeping in mind that $p\nmid abcn(n-m)$ ,

(2.4) $$ \begin{align} \beta \equiv -\frac{nc}{b(n-m)} \pmod p \end{align} $$

is the unique repeated root of $\bar {f}(x)$ in ${\mathbb {Z}}/p{\mathbb {Z}}$ and it can be easily checked that $\beta $ has multiplicity $2$ . Assuming that $\beta $ is a positive integer satisfying (2.4), we can write

$$ \begin{align*} f(x) &= (x-\beta+\beta)^n + a(b(x-\beta + \beta)+c)^m,\\ &= \sum\limits_{k=0}^{n}\binom{n}{k}\hspace{1.5pt}\beta^{n-k}(x-\beta)^k + a\bigg(\sum\limits_{k=0}^{m}\binom{m}{k}(b\beta + c)^{m-k}b^k(x-\beta)^k\bigg),\\ &= (x-\beta)^2g(x) + f'(\hspace{2pt}\beta)(x-\beta)+f(\hspace{2pt}\beta), \end{align*} $$

where $f'(x)$ is the derivative of $f(x)$ and

$$ \begin{align*}g(x) = \sum\limits_{k=2}^{n}\binom{n}{k}\hspace{1.5pt}\beta^{n-k}(x-\beta)^{k-2} + a\bigg(\sum\limits_{k=2}^{m}\binom{m}{k}(b\beta + c)^{m-k}b^k(x-\beta)^{k-2}\bigg) \end{align*} $$

is in ${\mathbb {Z}}[x]$ . Then

(2.5) $$ \begin{align} \bar{f}(x) = (x-{\beta})^2\bar{g}(x), \end{align} $$

where $\bar {g}(x) \in ({\mathbb {Z}}/p{\mathbb {Z}})[x]$ is separable. Write $g(x) = g_1(x)\cdots g_t(x) + ph(x)$ , where $g_1(x), \ldots , g_t(x)$ are monic polynomials which are distinct as well as irreducible modulo p and $h(x) \in {\mathbb {Z}}[x]$ monic. Therefore, we can write

$$ \begin{align*}f(x) = (x-\beta)^2\bigg(\prod\limits_{i=1}^{t}g_i(x) + ph(x)\bigg) + f'(\beta)(x-\beta) + f(\beta).\end{align*} $$

So, by Theorem 2.1, p does not divide $[{\mathbb {Z}}_K:{\mathbb {Z}}[\theta ]]$ if and only if $\overline {M}(x)$ is coprime to $x-\beta $ , where

$$ \begin{align*}M(x) = \frac{1}{p}[p(x-\beta)^2h(x)+(x-\beta)f'(\beta)+f(\beta)],\end{align*} $$

that is, $f(\beta ) \not \equiv 0 \pmod p^2.$ By (2.4), since $p\nmid abcmn(n-m)$ , we see that $f(\hspace{2pt}\beta ) \not \equiv 0 \pmod p^2$ if and only if $(n^nc^{n-m}+(-1)^{n+m}b^n(n-m)^{n-m}m^ma) \not \equiv 0 \pmod p^2.$ This final case completes the proof of the theorem.

3 Proof of Proposition 1.4

The following two results on Galois groups will be used in the proof of Proposition 1.4.

Theorem 3.1 [Reference Bishnoi and Khanduja1, Theorem 2.1].

Let $f(x)\in \mathbb {Z}[x]$ be a monic irreducible polynomial of degree n, having a root $\theta $ . Let p be a rational prime which is ramified in ${\mathbb Q}(\theta )$ . Suppose that $f(x)\equiv (x-c)^2\phi _2(x)\cdots \phi _r(x) \pmod p$ , where $(x-c), \phi _2(x), \ldots , \phi _r(x)$ are monic polynomials over ${\mathbb {Z}}$ which are distinct and irreducible modulo p. Then the Galois group of $f(x)$ over ${\mathbb Q}$ contains a nontrivial automorphism which keeps $n-2$ roots of $f(x)$ fixed.

Lemma 3.2 [Reference Filaseta and Moy4, Lemma 2].

Let $f(x)$ be an irreducible polynomial of degree $n\geq 2$ . If the Galois group of $f(x)$ over ${\mathbb Q}$ contains a transposition and a p-cycle for some prime $p>n/2$ , then the Galois group is $S_n.$

Proof of Proposition 1.4.

Let $\alpha $ be any root of $f(x)$ , so that $[{\mathbb Q}(\alpha ):{\mathbb Q}] = q.$ By the fundamental theorem of Galois theory, the Galois group of $f(x)$ , say $G_f$ , contains a subgroup whose index is q. By Lagrange’s theorem, q divides the order of $G_f$ . So, by Cauchy’s theorem, $G_f$ has an element of order q. Hence, $G_f$ contains a q-cycle. Now we show that $G_f$ contains a transposition. By hypothesis, there exists a prime p such that $p\mid D_f$ and $p\nmid abcm$ . As in (2.5) in the proof of Theorem 1.1(vi), $f(x) \equiv (x-\beta )^2g_1(x)\cdots g_t(x)$ (mod p), where $x-\beta , g_1(x), \ldots , g_t(x)$ are monic polynomials over ${\mathbb {Z}}$ which are distinct and irreducible modulo $p.$ Also, if $K ={\mathbb Q}(\theta )$ with $\theta $ a root of $f(x)$ , then keeping in mind the hypothesis $p^2\nmid D_f$ and the relation $D_f = [{\mathbb {Z}}_K : {\mathbb {Z}}[\theta ]]^2d_K$ , we see that $p \mid d_K$ . Hence, p is ramified in K. Therefore, by Theorem 3.1, the Galois group of $f(x)$ contains a transposition. Hence, by Lemma 3.2, the Galois group is $S_q.$

Footnotes

The author is thankful to IIT Madras for NFIG grant RF/22-23/1035/MA/NFIG/009034.

Dedicated to Professor Sudesh Kaur Khanduja

References

Bishnoi, A. and Khanduja, S. K., ‘A class of trinomials with Galois group ${S}_n$ ’, Algebra Colloq. 19(1) (2012), 905911.CrossRefGoogle Scholar
Cohen, H., A Course in Computational Algebraic Number Theory (Springer-Verlag, Berlin–Heidelberg, 1993).CrossRefGoogle Scholar
Dedekind, R., ‘Über den Zusammenhang zwischen der Theorie der Ideale und der Theorie der höheren Kongruenzen’, Götttingen Abh. 23 (1878), 123.Google Scholar
Filaseta, M. and Moy, R., ‘On the Galois group over $\mathbb{Q}$ of a truncated binomial expansion’, Colloq. Math. 154 (2018), 295308.CrossRefGoogle Scholar
Gaál, I., Diophantine Equations and Power Integral Bases: Theory and Algorithms, 2nd edn (Birkhäuser/Springer, Cham, 2019).CrossRefGoogle Scholar
Hasse, H., Zahlentheorie (Akademie-Verlag, Berlin, 1963).CrossRefGoogle Scholar
Jakhar, A., Khanduja, S. K. and Sangwan, N., ‘On prime divisors of the index of an algebraic integer’, J. Number Theory 166 (2016), 4761.CrossRefGoogle Scholar
Jakhar, A., Khanduja, S. K. and Sangwan, N., ‘Characterisation of primes dividing the index of a trinomial’, Int. J. Number Theory 13(10) (2017), 25052514.CrossRefGoogle Scholar
Jones, L., ‘A brief note on some infinite families of monogenic polynomials’, Bull. Aust. Math. Soc. 100 (2019), 239244.CrossRefGoogle Scholar
Jones, L., ‘On necessary and sufficient conditions for the monogenity of a certain class of polynomials’, Math. Slovaca 72(3) (2022), 591600.CrossRefGoogle Scholar
Jones, L., ‘Infinite families of monogenic quadrinomials, quintinomials and sextinomials’, Colloq. Math. 169 (2022), 110.CrossRefGoogle Scholar
Kaur, S. and Kumar, S., ‘On a conjecture of Lenny Jones about certain monogenic polynomials’, Bull. Aust. Math. Soc., to appear. Published online (21 November 2023).CrossRefGoogle Scholar