This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The Galois group of x2p+bxp+cpx^{2p}+bx^{p}+c^{p} over \mathbb{Q}

Akash Jim Princeton University
304 Washington Road
Princeton, NJ 08540
[email protected]
 and  Thomas Hagedorn Department of Mathematics and Statistics
The College of New Jersey
2000 Pennington Road
Ewing, NJ 08618
[email protected]
Abstract.

We prove an irreducibility criterion for polynomials of the form h(x)=x2m+bxm+c1F[x]h(x)=x^{2m}+bx^{m}+c_{1}\in F[x] relating to the Dickson polynomials of the first kind DpD_{p}. In the case when F=F=\mathbb{Q}, mm is a prime p>3p>3, and c1=cpc_{1}=c^{p}, for cc\in\mathbb{Q}, we explicitly determine the Galois group of dh=Dp(x,c)+bd_{h}=D_{p}(x,c)+b, which is Aff(𝔽p)\mathrm{Aff}(\mathbb{F}_{p}) or CpC(p1)/2Aff(𝔽p)C_{p}\rtimes C_{(p-1)/2}\vartriangleleft\mathrm{Aff}(\mathbb{F}_{p}), and the Galois group of hh, which is C2×Aff(𝔽p),Aff(𝔽p)C_{2}\times\mathrm{Aff}(\mathbb{F}_{p}),\mathrm{Aff}(\mathbb{F}_{p}), or C2×(CpC(p1)/2)C2×Aff(𝔽p)C_{2}\times(C_{p}\rtimes C_{(p-1)/2})\vartriangleleft C_{2}\times\mathrm{Aff}(\mathbb{F}_{p}).

Key words and phrases:
Galois group, irreducible, trinomial, Dickson polynomial, power compositional
2020 Mathematics Subject Classification:
12F10, 12E05, 11R09

1. Introduction

Let FF be a field, f(x)F[x]f(x)\in F[x] a polynomial, and let K/FK/F be the splitting field of f(x)f(x) over FF. If f(x)f(x) is a separable polynomial, the central result of Galois theory is that there is a bijective correspondence between the subgroups of G=Gal(K/F)G=\mathrm{Gal}(K/F), the automorphisms of KK that fix FF, and the subfields LKL\subset K containing FF. The group GG contains much information about the field extension K/FK/F. In particular, Galois proved (see [1, VI, Thm. 7.2]) that the roots of f(x)f(x) can be found via the usual arithmetic operations and nnth roots precisely when GG is a solvable group.

When F=F=\mathbb{Q}, the Galois group of f(x)f(x) has been determined for a number of classes of polynomials of small degree. For example, the Galois group of the trinomial f(x)=x2k+bxk+c[x]f(x)=x^{2k}+bx^{k}+c\in\mathbb{Q}[x] has been determined in the case of some small kk. The case k=3k=3 was addressed in the work of [2] and [3], and the case k=4k=4 was solved by [4], in the case when c=1c=1 or a rational square. In [5], Jones solves the case when kk is a prime p>3p>3, c=1c=1, and bb is an integer with |b|3|b|\geq 3. In this paper, we expand on these results and determine the Galois group of the trinomial polynomial f(x)f(x) in the case when k>3k>3 is prime and cc is a pp-th power in \mathbb{Q}. In the process of determining this Galois group, we are also able to determine the Galois group of a related family of polynomials formed from the Dickson polynomials.

In [6], Dickson introduced a set of polynomials that often give automorphisms of the finite field 𝔽qr\mathbb{F}_{q^{r}}. The Dickson polynomials of the first kind are defined by

D1(t,n)\displaystyle D_{1}(t,n) =t\displaystyle=t
D2(t,n)\displaystyle D_{2}(t,n) =t22n\displaystyle=t^{2}-2n
Dk(t,n)\displaystyle D_{k}(t,n) =tDk1(t,n)nDk2(t,n) for k>2.\displaystyle=tD_{k-1}(t,n)-nD_{k-2}(t,n)\text{ for }k>2.

Dk(t,n)D_{k}(t,n) is a degree kk polynomial and the next few polynomials are:

D3(t,n)\displaystyle D_{3}(t,n) =t33nt\displaystyle=t^{3}-3nt
D4(t,n)\displaystyle D_{4}(t,n) =t44nt2+2n2\displaystyle=t^{4}-4nt^{2}+2n^{2}
D5(t,n)\displaystyle D_{5}(t,n) =t55nt3+5n2t\displaystyle=t^{5}-5nt^{3}+5n^{2}t
D6(t,n)\displaystyle D_{6}(t,n) =t66nt4+9n2t22n3\displaystyle=t^{6}-6nt^{4}+9n^{2}t^{2}-2n^{3}
D7(t,n)\displaystyle D_{7}(t,n) =t77nt5+14n2t37n3t\displaystyle=t^{7}-7nt^{5}+14n^{2}t^{3}-7n^{3}t

A closed-form expression for Dk(t,n)D_{k}(t,n) is given by:

Dk(t,n)=i=0k2kki(kii)(n)itk2i.{\displaystyle D_{k}(t,n)=\sum_{i=0}^{\left\lfloor{\frac{k}{2}}\right\rfloor}{\frac{k}{k-i}}{\binom{k-i}{i}}(-n)^{i}t^{k-2i}\,.}

We first establish a reducibility criterion that relates to the Dickson polynomials. It generalizes Theorem 1.1(1) of Jones [5], who considered the case when h[x]h\in\mathbb{Z}[x], mm is odd, c=1c=1, and |b|3|b|\geq 3.

Theorem 1.1.

Let FF be a field and m>1m>1. The polynomial h(x)=x2m+bxm+cF[x]h(x)=x^{2m}+bx^{m}+c\in F[x] is reducible if and only if one of the following conditions holds:

  1. (1)

    f(x)=x2+bx+cf(x)=x^{2}+bx+c is reducible; or

  2. (2)

    For some prime pp divisor of mm, there exist n,tFn,\,t\in F with c=npc=n^{p} and b=Dp(t,n)b=-D_{p}(t,n); or

  3. (3)

    4m4\mid m and there exist n,tFn,t\in F with c=16n4c=16n^{4} and b=4D4(t,n)b=4D_{4}(t,n).

Remark 1.2.

If F=F=\mathbb{Q}, Theorem 1.1 allows one to determine the reducibility of x2m+bxm+c[x]x^{2m}+bx^{m}+c\in\mathbb{Q}[x] by using only the Rational Root Test to test for a rational root of the polynomials dh(x)=Dp(x,c)+bd_{h}(x)=D_{p}(x,c)+b in (2) and D4(x,n)bD_{4}(x,n)-b in (3).

When pp is a prime and h(x)=x2p+bxp+cp[x]h(x)=x^{2p}+bx^{p}+c^{p}\in\mathbb{Q}[x], we will see that the Galois group of hh is closely related to the Galois group of the polynomial

dh(x)=dp,b,c(x)=Dp(x,c)+b,d_{h}(x)=d_{p,b,c}(x)=D_{p}(x,c)+b,

where pp is a prime and b,cFb,c\in F. When hh is irreducible, we prove in Theorem 1.3 that dh(x)d_{h}(x) is irreducible and we explicitly classify the Galois groups of hh and dhd_{h}. Theorem 1.3 generalizes Theorem 1.1(2) of Jones [5], who considered the case when h[x]h\in\mathbb{Z}[x], c=1c=1, and |b|3|b|\geq 3.

Theorem 1.3.

Let p>3p>3 be prime and assume h(x):=x2p+bxp+cp[x]h(x):=x^{2p}+bx^{p}+c^{p}\in\mathbb{Q}[x] is irreducible. Then dhd_{h} is irreducible and the Galois groups of dhd_{h} and hh are determined as follows:

  1. (1)

    If b24cp(1)p(p1)/2p2b^{2}-4c^{p}\notin(-1)^{p(p-1)/2}p\mathbb{Q}^{2}, the Galois group of dhd_{h} is the affine group of 𝔽p\mathbb{F}_{p}, Aff(𝔽p)CpCp1\mathrm{Aff}(\mathbb{F}_{p})\simeq C_{p}\rtimes C_{p-1}, and the Galois group of hh is Aff(𝔽p)×C2\mathrm{Aff}(\mathbb{F}_{p})\times C_{2}.

  2. (2)

    If p1mod4p\equiv 1\bmod 4 and b24cpp2b^{2}-4c^{p}\in p\mathbb{Q}^{2}, the Galois groups of dhd_{h} and hh are identically Aff(𝔽p)\mathrm{Aff}(\mathbb{F}_{p}).

  3. (3)

    If p3mod4p\equiv 3\bmod 4 and b24cpp2b^{2}-4c^{p}\in-p\mathbb{Q}^{2}, then the Galois group of dhd_{h} is CpC(p1)/2Aff(𝔽p)C_{p}\rtimes C_{(p-1)/2}\vartriangleleft\mathrm{Aff}(\mathbb{F}_{p}), and the Galois group of hh is (CpC(p1)/2)×C2(C_{p}\rtimes C_{(p-1)/2})\times C_{2}.

Remark 1.4.

In [5], the Galois group of dhd_{h} in Theorem 1.3 (when c=1c=1) is described using Hol(Cn)\mathrm{Hol}(C_{n}), the Holomorph of CnC_{n}. It is defined as the semi-direct product:

Hol(Cn):=CnAut(Cn)\mathrm{Hol}(C_{n}):=C_{n}\rtimes\mathrm{Aut}(C_{n})

Then the Galois group of hh is Hol(C2p)\mathrm{Hol}(C_{2p}) in part (1) and Hol(Cp)\mathrm{Hol}(C_{p}) in part (2). Part (3) does not appear as part of [5, Thm 1.1(2)] as b24>0b^{2}-4>0 by its assumption that |b|3|b|\geq 3 but the hypothesis of part (3) is that b240b^{2}-4\leq 0.

Example 1.

Let p=5p=5, and consider h(x)=x103x5+32h(x)=x^{10}-3x^{5}+32. By Theorem 1.1, hh is irreducible. By Theorem 1.3, hh has Galois group Aff(𝔽5)×C2\mathrm{Aff}(\mathbb{F}_{5})\times C_{2} and

dh(x)=x510x3+20x3d_{h}(x)=x^{5}-10x^{3}+20x-3

has Galois group Aff(𝔽5)\mathrm{Aff}(\mathbb{F}_{5}).

Remark 1.5.

The case of p=2p=2 is a sub-case of the quartic, which has long been solved. Here, the Galois group of hh is C4C_{4}, and the Galois group of dhd_{h} is C2C_{2}.

Remark 1.6.

The case of p=3p=3 has been solved by Awtrey, Buerle, and Griesbach using resolvents as Example 4.1 of [3], over a general field. This generalizes Harrington and Jones’s [2]. In the case when p=3p=3, the statement of Theorem 1.3 is true, and can be derived from the work of [3]. (The Galois group must either be C6(C3C1)×C2C_{6}\simeq(C_{3}\rtimes C_{1})\times C_{2} or D6S3×C2Aff(𝔽3)×C2D_{6}\simeq S_{3}\times C_{2}\simeq\mathrm{Aff}(\mathbb{F}_{3})\times C_{2}, which can easily be distinguished by the degree of the extension.) For technical reasons, our proof of Theorem 1.3 does not work when p=3p=3.

Remark 1.7.

The Dickson polynomials of the first kind appear here because of a special case of Waring’s identity, [7]:

βk+β1k=Dk(β+β1,ββ1)\beta^{k}+\beta_{1}^{k}=D_{k}(\beta+{\beta_{1}},\beta{\beta_{1}})

The case of n=1n=1, with an arbitrary kk, is known as the kkth Vieta-Lucas polynomial of tt [8]. These polynomials are used in the work of Jones [5] in the case of polynomials x2m+Axm+1[x]x^{2m}+Ax^{m}+1\in\mathbb{Z}[x].

2. Some Preliminaries

Let FF denote an arbitrary field. NFK(α)\mathrm{N}^{K}_{F}(\alpha) and TrFK(α)\mathrm{Tr}^{K}_{F}(\alpha) will denote the norm and trace of αK\alpha\in K with respect to FF. Let f(x)=fb,c(x)f(x)=f_{b,c}(x) denote the monic trinomial x2+bx+cF[x]x^{2}+bx+c\in F[x], and denote its roots by α,α1\alpha,\alpha_{1} and its determinant by Δ:=b24c\Delta:=b^{2}-4c.

A key theorem in establishing irreducibility of power compositional polynomials (cf., e.g., [2], [4]) is Capelli’s Theorem (see Theorem 22 of [9] for proof):

Theorem 2.1 (Capelli’s Theorem).

Let f(x),g(x)F[x]f(x),g(x)\in F[x] and assume f(x)f(x) is irreducible with root α\alpha. Then f(g(x))f(g(x)) is reducible in F[x]F[x] if and only if the polynomial g(x)αg(x)-\alpha is reducible in F(α)[x]F(\alpha)[x]. Moreover, if g(x)αg(x)-\alpha has the prime factorization g(x)α=Ci=1rhi(x)eig(x)-\alpha=C\prod_{i=1}^{r}h_{i}(x)^{e_{i}}, with CF(α)C\in F(\alpha) and irreducible polynomials hi(x)F(α)[x]h_{i}(x)\in F(\alpha)[x], then

f(g(x))=C~i=1rNFF(α)(hi(x))ei,f(g(x))=\tilde{C}\prod_{i=1}^{r}\mathrm{N}_{F}^{F(\alpha)}(h_{i}(x))^{e_{i}},

where C~F\tilde{C}\in F and NFF(α)(hi(x))\mathrm{N}_{F}^{F(\alpha)}(h_{i}(x)) are irreducible.

Two well known consequences of this theorem are:

Theorem 2.2 (Capelli).

[1, VI, Thm. 9.1] Let aF,na\in F,n\in\mathbb{N}. xnax^{n}-a is irreducible in F[x]F[x] if and only if the following conditions hold:

  1. (1)

    aFpa\notin F^{p} for all primes pnp\mid n; and

  2. (2)

    If 4n4\mid n, a4F4a\notin-4F^{4}.

Corollary 2.3.

For pp prime, xpαF[x]x^{p}-\alpha\in F[x] is reducible if and only if βF\exists\beta\in F with α=βp\alpha=\beta^{p}.

Thus, if f(xp)f(x^{p}) is reducible, either f(x)f(x) is reducible, or there exists βF(α)\beta\in F(\alpha) with f(βp)=0f(\beta^{p})=0 and NFF(α)(xβ)h(x)\mathrm{N}^{F(\alpha)}_{F}(x-\beta)\mid h(x).

Define p(x):=h(x)/N(xβ)p(x):=h(x)/\mathrm{N}(x-\beta). The polynomial p(x)p(x) is symmetric in β\beta and its conjugate N(β)/βF(α)\mathrm{N}(\beta)/\beta\in F(\alpha), so it can be expressed in terms of the symmetric polynomials. By expanding and equating coefficients, we will obtain the reducibility criterion of 1.1.

3. A Reducibility Criterion for x2m+bxm+cF[x]x^{2m}+bx^{m}+c\in F[x]

In this section, we prove Theorem 1.1, which gives a reducibility criteria for x2m+bxm+cF[x]x^{2m}+bx^{m}+c\in F[x]. Theorem 1.1 extends Jones’s reducibility criterion [5, Thm. 1.1(1)], which considered the case when c=1c=1. The proof of Theorem 1.1 is essentially the same as Jones’ proof in the case c=1c=1. It will follow as a corollary from Theorem 3.1, which gives a relationship of the coefficients of

f(x):=x2+bx+cF[x].f(x):=x^{2}+bx+c\in F[x].

when f(x)f(x) is irreducible. Recall from the introduction that Dm(t,n)D_{m}(t,n) are the Dickson polynomials.

Theorem 3.1.

Let f(x):=x2+bx+cF[x]f(x):=x^{2}+bx+c\in F[x] be irreducible with a root αF(Δ)F\alpha\in F(\sqrt{\Delta})\setminus F, where Δ:=b24c\Delta:=b^{2}-4c. Then αF(Δ)m\alpha\in F(\sqrt{\Delta})^{m}, for some positive integer mm if and only if there exists some n,tFn,t\in F with c=nmc=n^{m} and b=Dm(t,n)b=-D_{m}(t,n).

Before proving Theorem 3.1, we need to establish two theorems about polynomials of the form f(xm)f(x^{m}). Given variables β,β1\beta,\beta_{1}, define:

t:=β+β1, n:=ββ1t:=\beta+{\beta_{1}},\text{\qquad}n:=\beta{\beta_{1}}

For mm\in\mathbb{N}, define

Ψm,β(x):=xmβmxβ\Psi_{m,\beta}(x):=\frac{x^{m}-\beta^{m}}{x-\beta}

and define the polynomial pm(x):=Ψm,β(x)Ψm,β1(x)p_{m}(x):=\Psi_{m,\beta}(x)\Psi_{m,\beta_{1}}(x). For m>0m>0, let ama_{m} be the coefficient of xm1x^{m-1} in pm(x)p_{m}(x)

Theorem 3.2.

The ama_{m} are given by a1=1a_{1}=1, a2=ta_{2}=t, and the recursive formula

am+2=tam+1nama_{m+2}=ta_{m+1}-na_{m}

and pm(x)p_{m}(x) is expressible in terms of {ai}1im\{a_{i}\}_{1\leq i\leq m} by:

pm(x)=i=1m1aix2m1i+amxm1+i=1m1aminixm1ip_{m}(x)=\sum_{i=1}^{m-1}a_{i}x^{2m-1-i}+a_{m}x^{m-1}+\sum_{i=1}^{m-1}a_{m-i}n^{i}x^{m-1-i}

(where the outer sums are defined to be 0 in the case of m=1m=1).

Proof.

We first compute pm(x),amp_{m}(x),a_{m} from their definition.

pm(x)\displaystyle p_{m}(x) =Ψm,β(x)Ψm,β1(x)=(i=0m1βm1ixi)(j=0m1β1m1jxj)\displaystyle=\Psi_{m,\beta}(x)\Psi_{m,\beta_{1}}(x)=(\sum_{i=0}^{m-1}\beta^{m-1-i}x^{i})(\sum_{j=0}^{m-1}{\beta_{1}}^{m-1-j}x^{j})
=l=02(m1)(i+j=l0i,jm1βm1iβ1m1j)xl\displaystyle=\sum_{l=0}^{2(m-1)}\Big{(}\sum_{\begin{subarray}{c}i+j=l\\ 0\leq i,j\leq m-1\end{subarray}}\beta^{m-1-i}{\beta_{1}}^{m-1-j}\Big{)}x^{l}

The coefficient of the xm1x^{m-1} term is

am=i+j=m10i,jm1βm1iβ1m1j=i=0m1βm1iβ1i.\displaystyle{a_{m}=\sum_{\begin{subarray}{c}i+j=m-1\\ 0\leq i,j\leq m-1\end{subarray}}\beta^{m-1-i}{\beta_{1}}^{m-1-j}=\sum_{i=0}^{m-1}\beta^{m-1-i}{\beta_{1}}^{i}}.

We can then verify the base cases explicitly:

p1(x)\displaystyle p_{1}(x) =Ψ1,β(x)Ψ1,β1(x)=11=1=a1\displaystyle=\Psi_{1,\beta}(x){\Psi_{1,\beta_{1}}(x)}=1\cdot 1=1=a_{1}
p2(x)\displaystyle p_{2}(x) =Ψ2,β(x)Ψ2,β1(x)=(x+β)(x+β1)\displaystyle=\Psi_{2,\beta}(x){\Psi_{2,\beta_{1}}(x)}=(x+\beta)(x+\beta_{1})
=x2+tx+n\displaystyle=x^{2}+tx+n
=a1x2+a2x1+a1n1x0\displaystyle=a_{1}x^{2}+a_{2}x^{1}+a_{1}n^{1}x^{0}

With regard to the recursive relation of the (am)(a_{m}), we may compute:

tamnam1=(β+β1)i+j=m10i,jm1βm1iβ1m1j ββ1i+j=m20i,jm2βm2iβ1m2j=i=0m1βmiβ1i+i=0m1βm1iβ1i+1i=0m2βm1iβ1i+1=i=0m1βmiβ1i+β10βm=i=0mβmi=am+1\begin{split}ta_{m}-na_{m-1}&=(\beta+{\beta_{1}})\sum_{\begin{subarray}{c}i+j=m-1\\ 0\leq i,j\leq m-1\end{subarray}}\beta^{m-1-i}{\beta_{1}}^{m-1-j}\\ &\text{\hskip 36.135pt}-\beta{\beta_{1}}\sum_{\begin{subarray}{c}i+j=m-2\\ 0\leq i,j\leq m-2\end{subarray}}\beta^{m-2-i}{\beta_{1}}^{m-2-j}\\ &=\sum_{i=0}^{m-1}\beta^{m-i}{\beta_{1}}^{i}+\sum_{i=0}^{m-1}\beta^{m-1-i}{\beta_{1}}^{i+1}-\sum_{i=0}^{m-2}\beta^{m-1-i}{\beta_{1}}^{i+1}\\ &=\sum_{i=0}^{m-1}\beta^{m-i}{\beta_{1}}^{i}+\beta_{1}^{0}{\beta}^{m}=\sum_{i=0}^{m}\beta^{m-i}=a_{m+1}\qquad\end{split}

And with regard to the expression of pmp_{m}, we assume for induction that the identity holds for pmp_{m} and expand pm+1(x)p_{m+1}(x):

pm+1(x)\displaystyle p_{m+1}(x) =l=02m(i+j=l0i,jmβmiβ1mj)xl\displaystyle=\sum_{l=0}^{2m}\left(\sum_{\begin{subarray}{c}i+j=l\\ 0\leq i,j\leq m\end{subarray}}\beta^{m-i}{\beta_{1}}^{m-j}\right)x^{l}\allowdisplaybreaks[3]
=l=m+12m(i+j=l1i,jmβmiβ1mj)xl+i+j=m0i,jmβmiβ1mjxm\displaystyle=\sum_{l=m+1}^{2m}\Big{(}\sum_{\begin{subarray}{c}i+j=l\\ 1\leq i,j\leq m\end{subarray}}\beta^{m-i}{\beta_{1}}^{m-j}\Big{)}x^{l}+\sum_{\begin{subarray}{c}i+j=m\\ 0\leq i,j\leq m\end{subarray}}\beta^{m-i}{\beta_{1}}^{m-j}x^{m}\allowdisplaybreaks[3]
 +l=0m1(i+j=l0i,jm1βmiβ1mj)xl\displaystyle\text{\hskip 72.26999pt}+\sum_{l=0}^{m-1}\Big{(}\sum_{\begin{subarray}{c}i+j=l\\ 0\leq i,j\leq m-1\end{subarray}}\beta^{m-i}{\beta_{1}}^{m-j}\Big{)}x^{l}\allowdisplaybreaks[3]
=x2l=m12m2(i+j=l0i,jm1βm1iβ1m1j)xl+am+1xm\displaystyle=x^{2}\sum_{l=m-1}^{2m-2}\Big{(}\sum_{\begin{subarray}{c}i+j=l\\ 0\leq i,j\leq m-1\end{subarray}}\beta^{m-1-i}{\beta_{1}}^{m-1-j}\Big{)}x^{l}+a_{m+1}x^{m}
 +ββ1l=0m1(i+j=l0i,jm1βm1iβ1m1j)xl\displaystyle\text{\hskip 72.26999pt}+\beta{\beta_{1}}\sum_{l=0}^{m-1}\Big{(}\sum_{\begin{subarray}{c}i+j=l\\ 0\leq i,j\leq m-1\end{subarray}}\beta^{m-1-i}{\beta_{1}}^{m-1-j}\Big{)}x^{l}
=n(i=1m1aminixm1i+amxm1)+am+1xm\displaystyle=n\Big{(}\sum_{i=1}^{m-1}a_{m-i}n^{i}x^{m-1-i}+a_{m}x^{m-1}\Big{)}+a_{m+1}x^{m}
 +x2(amxm1+i=1m1aix2m1i)\displaystyle\text{\hskip 72.26999pt}+x^{2}\Big{(}a_{m}x^{m-1}+\sum_{i=1}^{m-1}a_{i}x^{2m-1-i}\Big{)}
=i=1mam+1inixmi+am+1xm+i=1maix2m+1i\displaystyle=\sum_{i=1}^{m}a_{m+1-i}n^{i}x^{m-i}+a_{m+1}x^{m}+\sum_{i=1}^{m}a_{i}x^{2m+1-i}

The following theorem will be a tool to study polynomials of the form f(xm)f(x^{m}) and prove Theorem 3.1.

Theorem 3.3.

Use the same notation as in Theorem 3.2, and define the polynomial g(x):=(xβm)(xβ1m)g(x):=(x-\beta^{m})(x-\beta_{1}^{m}). Then g(xm)=x2mDm(t,n)xm+nmg(x^{m})=x^{2m}-D_{m}(t,n)x^{m}+n^{m}, where DmD_{m} is the mmth Dickson polynomial of the first kind.

Proof.
g(xm)=(xmβm)(xβ1m)=(xβ)Ψm,β(x)(xβ1)Ψm,β1(x)=(x2tx+n)pm(x)=(x2tx+n)(i=1m1aix2m1i+amxm1+i=1m1aminixm1i)=a1x2m+(nam1tam+nam1)xm+a1nm=x2m+(2nam1tam)xm+nm\begin{split}g(x^{m})&=(x^{m}-\beta^{m})(x-\beta_{1}^{m})=(x-\beta)\Psi_{m,\beta}(x)(x-\beta_{1})\Psi_{m,\beta_{1}}(x)\\ &=(x^{2}-tx+n)p_{m}(x)\\ &=(x^{2}-tx+n)\left(\sum_{i=1}^{m-1}a_{i}x^{2m-1-i}+a_{m}x^{m-1}+\sum_{i=1}^{m-1}a_{m-i}n^{i}x^{m-1-i}\right)\\ &=a_{1}x^{2m}+(na_{m-1}-ta_{m}+na_{m-1})x^{m}+a_{1}n^{m}\\ &=x^{2m}+(2na_{m-1}-ta_{m})x^{m}+n^{m}\\ \end{split}

We define

bm=bm(β,β1)=2nam1tam,b_{m}=b_{m}(\beta,\beta_{1})=2na_{m-1}-ta_{m},

Then g(xm)=x2m+bmxm+nmg(x^{m})=x^{2m}+b_{m}x^{m}+n^{m}. We observe now that:

tbm+1nbm=t(2namtam+1)n(2nam1tam)=2n(tamnam1)t(tam+1nam)=2nam+1tam+2=bm+2\begin{split}tb_{m+1}-nb_{m}&=t(2na_{m}-ta_{m+1})-n(2na_{m-1}-ta_{m})\\ &=2n(ta_{m}-na_{m-1})-t(ta_{m+1}-na_{m})\\ &=2na_{m+1}-ta_{m+2}\\ &=b_{m+2}\end{split}

Also, b1=tb_{1}=-t and b2=2nt2b_{2}=2n-t^{2}. (Note: we define a0:=0a_{0}:=0 so that b1=tb_{1}=-t, in agreement with the direct expansion of g(xm)g(x^{m}).) So indeed bm=Dm(t,n)b_{m}=-D_{m}(t,n), and as claimed, g(xm)=x2mDm(t,n)xm+nmg(x^{m})=x^{2m}-D_{m}(t,n)x^{m}+n^{m}. ∎

We now use Theorem 3.3 and Capelli’s Theorem to study polynomials of the form f(xm)f(x^{m}) and prove Theorem 3.1.

Proof of Theorem 3.1.

If there exists βF(Δ)\beta\in F(\sqrt{\Delta}) with βm=α\beta^{m}=\alpha, xmα=(xβ)Ψm(x)x^{m}-\alpha=(x-\beta)\Psi_{m}(x) is a valid factorization of xmαF(Δ)[x]x^{m}-\alpha\in F(\sqrt{\Delta})[x]. We now identify the symbol β1\beta_{1} from Theorems 3.2 and 3.3 with the conjugate of β\beta, namely N(β)/βF(Δ)\mathrm{N}(\beta)/\beta\in F(\sqrt{\Delta}). From Capelli’s Theorem (and the multiplicativity of norms), this factorization induces the F[x]F[x]-factorization

x2m+bxm+c=N(xβ)N(Ψm,β(x))=(x2tx+n)pm(x)x^{2m}+bx^{m}+c=\mathrm{N}(x-\beta)\mathrm{N}(\Psi_{m,\beta}(x))=(x^{2}-tx+n)p_{m}(x)

Equating coefficients with the expression f(xm)=x2mDm(t,n)xm+nmf(x^{m})=x^{2m}-D_{m}(t,n)x^{m}+n^{m} from Theorem 3.3 gives the desired equality:

b=Dm(t,n),c=nmb=-D_{m}(t,n),c=n^{m}

with

t=Tr(β),n=N(β)Ft=\mathrm{Tr}(\beta),n=\mathrm{N}(\beta)\in F

Conversely, if n,tF\exists n,t\in F with the given conditions, then by Theorem 3.3,

f(xm)=(x2tx+n)pm(x)f(x^{m})=(x^{2}-tx+n)p_{m}(x)

By Capelli’s Theorem, (x2tx+n)(x^{2}-tx+n) is the product of the norms of some irreducible polynomials in F(α)=F(Δ)F(\alpha)=F(\sqrt{\Delta}) that divide xmαx^{m}-\alpha. α\alpha is quadratic, so x2tx+n=N(xβ)x^{2}-tx+n=\mathrm{N}(x-\beta) for some (xβ)(xmα)(x-\beta)\mid(x^{m}-\alpha), and α=βmF(Δ)m\alpha=\beta^{m}\in F(\sqrt{\Delta})^{m}. ∎

The criterion for reducibility, Theorem 1.1, follows as a corollary.

Proof of Theorem 1.1.

We use the notation common to Theorems 1.1 and 3.1. If ff is reducible, so is h=f(xm)h=f(x^{m}). If ff is irreducible, from Capelli’s Theorem, hh is reducible if and only if xmαx^{m}-\alpha is reducible in F(α)F(\alpha), where α\alpha is a root of ff. From Theorem 2.2, this occurs if and only if αF(α)p\alpha\in F(\alpha)^{p} for some pmp\mid m prime, or α4F(α)4\alpha\in-4F(\alpha)^{4} and 4m4\mid m. From Theorem 3.1, αF(α)p\alpha\in F(\alpha)^{p} if and only if there exist n,tFn,t\in F with c=npc=n^{p} and b=Dp(t,n)b=-D_{p}(t,n). If α=4β4\alpha=-4\beta^{4} for some βF(α)\beta\in F(\alpha), denote the conjugate of β\beta over FF by β1\beta_{1}, with α1=4β14\alpha_{1}=-4\beta_{1}^{4}. Then

x4+bx2+c\displaystyle x^{4}+bx^{2}+c =(x2α)(x2α1)\displaystyle=(x^{2}-\alpha)(x^{2}-\alpha_{1})
=(x2+4β4)(x2+4β14)\displaystyle=(x^{2}+4\beta^{4})(x^{2}+4\beta_{1}^{4})
=x4+4(β4+β14)x2+16β4β14\displaystyle=x^{4}+4(\beta^{4}+\beta_{1}^{4})x^{2}+16\beta^{4}\beta_{1}^{4}
=x4+4D4(Tr(β),N(β))+16(N(β))4\displaystyle=x^{4}+4D_{4}(\mathrm{Tr}(\beta),\mathrm{N}(\beta))+16(\mathrm{N}(\beta))^{4}

By Waring’s identity (Remark 1.7). And indeed t:=Tr(β)Ft:=\mathrm{Tr}(\beta)\in F, n:=N(β)Fn:=\mathrm{N}(\beta)\in F. Conversely, if such t,nt,n exist, we observe that

x8+bx4+c=x8+4D4(t,n)+16n4=x8+(4t416nt2+8n2)x4+16n4=(x4+(2t24n)x24n2)(x4(2t24n)x24n2)\begin{split}x^{8}+bx^{4}+c&=x^{8}+4D_{4}(t,n)+16n^{4}\\ &=x^{8}+(4t^{4}-16nt^{2}+8n^{2})x^{4}+16n^{4}\\ &=(x^{4}+(2t^{2}-4n)x^{2}-4n^{2})(x^{4}-(2t^{2}-4n)x^{2}-4n^{2})\\ \end{split}

So f(x4)f(x^{4}) and therefore f(xm)f(x^{m}) are reducible. ∎

Remark 3.4.

We note that the criterion of Theorem 1.1 is very similar to parts (vi) and (vii) of Theorem 6 of Schinzel [10]. It is quite possible that Theorem 1.1 is encompassed by the results in [10] on the reducibility of trinomials. However, those results utilize elliptic curves and the proof of Theorem 1.1 does not.

4. Properties of h=x2p+bxp+cp[x]h=x^{2p}+bx^{p}+c^{p}\in\mathbb{Q}[x] and dh=Dp(x,c)+bd_{h}=D_{p}(x,c)+b

We now restrict ourselves to the case when F=F=\mathbb{Q}, m=p>3m=p>3 is prime, and the constant term of ff is the form cpc^{p}. With

f(x)\displaystyle f(x) =x2+bx+cp\displaystyle=x^{2}+bx+c^{p}
h(x)\displaystyle h(x) =f(xp)=x2p+bxp+cp\displaystyle=f(x^{p})=x^{2p}+bx^{p}+c^{p}

we assume in this section and the next that h(x)h(x) is irreducible. By Theorem 1.1(1), Δ:=b24cp\Delta:=b^{2}-4c^{p} is not a rational square and the splitting field of ff is (Δ)\mathbb{Q}(\sqrt{\Delta}). Also, by Theorem 1.1(2), the irreducibility of hh implies that bDp(t,c)b\neq-D_{p}(t,c) for any tt\in\mathbb{Q}. Hence the polynomial dh:=Dp(x,c)+bd_{h}:=D_{p}(x,c)+b has no rational roots.

Let ζn\zeta_{n} be a primitive nnth root of unity and let

Fn=(ζn)F_{n}=\mathbb{Q}(\zeta_{n})

be the nnth cyclotomic field. We will denote the splitting field of hh by KK and the splitting field of dhd_{h} by LL. We denote the roots of ff by α,α1\alpha,\alpha_{1}. We then choose β\beta so that βp=α\beta^{p}=\alpha and β1:=cβ\beta_{1}:=\frac{c}{\beta} (and indeed β1p=cpβp=cpα=α1\beta_{1}^{p}=\frac{c^{p}}{\beta^{p}}=\frac{c^{p}}{\alpha}=\alpha_{1}). hh has 2p2p roots, which are precisely

{ζpiβ}ip{ζpiβ1}ip\{\zeta_{p}^{i}\beta\}_{i\in\mathbb{Z}_{p}}\cup\{\zeta_{p}^{i}\beta_{1}\}_{i\in\mathbb{Z}_{p}}

We claim now:

Lemma 4.1.

(ζpiβ)=(ζpiβ1)\mathbb{Q}(\zeta_{p}^{i}\beta)=\mathbb{Q}(\zeta_{p}^{-i}\beta_{1}), and moreover, these are the only roots of h(x)h(x) in this extension of \mathbb{Q}.

Proof.

By construction ζpiβ1=ζpicβ=cζpiβ\zeta_{p}^{-i}\beta_{1}=\zeta_{p}^{-i}\frac{c}{\beta}=\frac{c}{\zeta_{p}^{i}\beta}, so (ζpiβ)=(ζpiβ1)\mathbb{Q}(\zeta_{p}^{i}\beta)=\mathbb{Q}(\zeta_{p}^{-i}\beta_{1}). Suppose that some other root of hh lies in (ζpiβ)\mathbb{Q}(\zeta_{p}^{i}\beta), without loss of generality, ζpjβ\zeta_{p}^{j}\beta. Then ζpij(ζpiβ)\zeta_{p}^{i-j}\in\mathbb{Q}(\zeta_{p}^{i}\beta). If iji\neq j, Fp(β)F_{p}\subset\mathbb{Q}(\beta). This is impossible because [Fp:]=p1[F_{p}:\mathbb{Q}]=p-1 and [(β):]=2p[\mathbb{Q}(\beta):\mathbb{Q}]=2p, but p12pp-1\nmid 2p given that p>3p>3. So i=ji=j, and ζpjβ=ζpiβ\zeta_{p}^{j}\beta=\zeta_{p}^{i}\beta, which is one of the given roots after all. ∎

As a corollary, there are no double roots: αα1\alpha\neq\alpha_{1} else ff is reducible, and the only pairs of roots which generate the same extension are of the form ζpiβ,ζpiβ1\zeta_{p}^{i}\beta,\zeta_{p}^{-i}\beta_{1}, whose ppth powers are α,α1\alpha,\alpha_{1}, respectively.

Now denote Bi:=(ζpiβ)=(ζpiβ1)B_{i}:=\mathbb{Q}(\zeta_{p}^{i}\beta)=\mathbb{Q}(\zeta_{p}^{-i}\beta_{1}) and :={Bi}ip\mathcal{B}:=\{B_{i}\}_{i\in\mathbb{Z}_{p}}, with ||=p|\mathcal{B}|=p. Clearly, [Bi:]=2p[B_{i}:\mathbb{Q}]=2p. Also, ζpiβ+ζpiβ1Bi\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1}\in B_{i}. We show that it is a root of dhd_{h}.

Lemma 4.2.

dh(ζpiβ+ζpiβ1)=0d_{h}(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1})=0.

Proof.

By factoring ff, we see that:

x2p+bxp+cp=f(xp)=(xp(ζpiβ)p)(xp(ζpiβ1)p)x^{2p}+bx^{p}+c^{p}=f(x^{p})=(x^{p}-(\zeta_{p}^{i}\beta)^{p})(x^{p}-(\zeta_{p}^{-i}\beta_{1})^{p})

And by Theorem 3.3:

f(xp)=x2pDp(ζpiβ+ζpiβ1,c)xp+cpf(x^{p})=x^{2p}-D_{p}(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1},c)x^{p}+c^{p}

Equating coefficients, Dp(ζpiβ+ζpiβ1,c)=b-D_{p}(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1},c)=b, i.e., dh(ζpiβ+ζpiβ1)=0d_{h}(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1})=0. ∎

We now define the fields Di:=(ζpiβ+ζpiβ1)D_{i}:=\mathbb{Q}(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1}), with DiBiD_{i}\subset B_{i}, and define

𝒟:={Di}ip\mathcal{D}:=\{D_{i}\}_{i\in\mathbb{Z}_{p}}

A priori, we do not know that the fields DiD_{i} are distinct, but for now, it suffices that the distinct symbols DiD_{i} are in bijective correspondence with p\mathbb{Z}_{p} and \mathcal{B}. Before examining the fields DiD_{i}, it will be useful to prove that:

Lemma 4.3.

Gal(K/)\mathrm{Gal}(K/\mathbb{Q}) acts transitively and equivalently on \mathcal{B}, on 𝒟\mathcal{D}, and on the roots of dhd_{h}.

Proof.

Let σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}). σ\sigma permutes the roots of hh, and from Lemma 4.1, it must be that σ(ζpiβ)Bj\sigma(\zeta_{p}^{i}\beta)\in B_{j} for some jj. Then σ(Bi)=Bj\sigma(B_{i})=B_{j}, and σ\sigma acts on \mathcal{B}. Gal(K/)\mathrm{Gal}(K/\mathbb{Q}) is transitive on the roots of irreducible hh, so its action on \mathcal{B} is also transitive. σ\sigma maps the pair of roots in BiB_{i} to the pair of roots in BjB_{j}, so the action on \mathcal{B} is equivalent to the action on the set of pairs of roots of hh. Then:

σ(ζpiβ+ζpiβ1)=ζpjβ+ζpjβ1\sigma(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1})=\zeta_{p}^{j}\beta+\zeta_{p}^{-j}\beta_{1}

So σ(Di)=Dj\sigma(D_{i})=D_{j}. Therefore, σ\sigma acts equivalently on the roots of dd, on 𝒟\mathcal{D}, and on \mathcal{B}, according to the correspondence ζpiβ+ζpiβ1DiBi\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1}\in D_{i}\subset B_{i}, and these equivalent actions of Gal(K/)\mathrm{Gal}(K/\mathbb{Q}) are transitive. ∎

We can now prove the following properties about the fields DiD_{i}.

Lemma 4.4.

dh(x)=Dp(x,c)+bd_{h}(x)=D_{p}(x,c)+b is irreducible, so Di[x]/(dh(x))D_{i}\simeq\mathbb{Q}[x]/(d_{h}(x)). Also, Bi=Di(Δ)B_{i}=D_{i}(\sqrt{\Delta}), and Di=DjD_{i}=D_{j} if and only if i=ji=j.

Proof.
x2p+bxp+cp=f(xp)=(xpα)(xpα1)=(xp(ζpiβ)p)(xp(ζpiβ1)p)x^{2p}+bx^{p}+c^{p}=f(x^{p})=(x^{p}-\alpha)(x^{p}-\alpha_{1})=(x^{p}-(\zeta_{p}^{i}\beta)^{p})(x^{p}-(\zeta_{p}^{-i}\beta_{1})^{p})

By Theorem 3.3, b=Dp(ζpiβ+ζpiβ1,c)b=-D_{p}(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1},c), so ζpiβ+ζpiβ1Bi\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1}\in B_{i} is a root of dhd_{h}. Thus, [Di:]p[D_{i}:\mathbb{Q}]\leq p, and since DiBiD_{i}\subset B_{i}, [Di:]2p[D_{i}:\mathbb{Q}]\mid 2p. Thus, [Di:][D_{i}:\mathbb{Q}] could be 11, 22, or pp. [Di:]1[D_{i}:\mathbb{Q}]\neq 1 because this would give a rational root for dhd_{h}, a contradiction.

Now assume temporarily that [Di:]=2[D_{i}:\mathbb{Q}]=2. Then Di(Δ)BiD_{i}(\sqrt{\Delta})\subset B_{i} is an extension of either degree 2 or degree 4. 42p4\nmid 2p given p>3p>3, so [Di(Δ):]=2[D_{i}(\sqrt{\Delta}):\mathbb{Q}]=2, and Di=(Δ)D_{i}=\mathbb{Q}(\sqrt{\Delta}). By the transitive action of the Galois group, all DiD_{i} are isomorphic and are thus quadratic extensions. So each root of dhd_{h} is a root of a quadratic factor, and dhd_{h} factors as a product of quadratics. But this would make deg(dh)=p\deg(d_{h})=p even, a contradiction. So [Di:]=p[D_{i}:\mathbb{Q}]=p after all, and consequently, dh(x)d_{h}(x) is irreducible with Di[x]/(dh(x))D_{i}\simeq\mathbb{Q}[x]/(d_{h}(x)), as claimed.

Then ΔDi\sqrt{\Delta}\notin D_{i} because 2p2\nmid p, so [Di(Δ):]=2p=[Bi:][D_{i}(\sqrt{\Delta}):\mathbb{Q}]=2p=[B_{i}:\mathbb{Q}] and Di(Δ)BiD_{i}(\sqrt{\Delta})\subset B_{i}, so Di(Δ)=BiD_{i}(\sqrt{\Delta})=B_{i}. If iji\neq j but Di=DjD_{i}=D_{j}, then Bi=BjB_{i}=B_{j}, a contradiction of Lemma 4.1. So Di=DjD_{i}=D_{j} if and only if i=ji=j, as desired. ∎

As a technical lemma, we must now note that:

Lemma 4.5.

K=Fp(β)K=F_{p}(\beta). [K:]=2p(p1)[K:\mathbb{Q}]=2p(p-1) if ΔFp\sqrt{\Delta}\notin F_{p} and [K:]=p(p1)[K:\mathbb{Q}]=p(p-1) if ΔFp\sqrt{\Delta}\in F_{p}.

Proof.

Fp(β)=B0(ζp)F_{p}(\beta)=B_{0}(\zeta_{p}) certainly contains all the roots of hh. Conversely, the roots of hh include β\beta and ζpβ\zeta_{p}\beta, so the splitting field is at least (ζp,β)=Fp(β)\mathbb{Q}(\zeta_{p},\beta)=F_{p}(\beta). Thus, K=Fp(β)K=F_{p}(\beta).

[Bi:]=2p[B_{i}:\mathbb{Q}]=2p and [Fp:]=p1[F_{p}:\mathbb{Q}]=p-1, and gcd(2p,p1)=2\gcd(2p,p-1)=2 because pp is odd. Thus, p(p1)[K:]2p(p1)p(p-1)\mid[K:\mathbb{Q}]\mid 2p(p-1). If ΔFp\sqrt{\Delta}\in F_{p}, then [Fp:(Δ)]=p12[F_{p}:\mathbb{Q}(\sqrt{\Delta})]=\frac{p-1}{2} is coprime to [B:(Δ)]=p[B:\mathbb{Q}(\sqrt{\Delta})]=p. Thus [K:]=2[K:(Δ)]=2p12p=p(p1)[K:\mathbb{Q}]=2[K:\mathbb{Q}(\sqrt{\Delta})]=2\cdot\frac{p-1}{2}\cdot p=p(p-1). If ΔFp\sqrt{\Delta}\notin F_{p}, then [Fp:(Δ)]=p1[F_{p}:\mathbb{Q}(\sqrt{\Delta})]=p-1 is again coprime to [B:(Δ)]=p[B:\mathbb{Q}(\sqrt{\Delta})]=p, and similarly [K:]=2[K:(Δ)]=2(p1)p=2p(p1)[K:\mathbb{Q}]=2[K:\mathbb{Q}(\sqrt{\Delta})]=2\cdot(p-1)\cdot p=2p(p-1). ∎

Remark 4.6.

As (Δ)/\mathbb{Q}(\sqrt{\Delta})/\mathbb{Q} and Fp/F_{p}/\mathbb{Q} are Galois extensions, an automorphism σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}) acts as an automorphism of (Δ),Fp\mathbb{Q}(\sqrt{\Delta}),F_{p}.

Using the lemmas of this section, we are now able to specify the actions of σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}) in terms of its action on D0,ΔD_{0},\sqrt{\Delta}, and ζp\zeta_{p}. We define ϵσ=σ(Δ)/Δ=±1\epsilon_{\sigma}=\sigma(\sqrt{\Delta})/\sqrt{\Delta}=\pm 1.

Theorem 4.7.

If ΔFp\sqrt{\Delta}\notin F_{p}, then there exists a bijection between Gal(K/)\mathrm{Gal}(K/\mathbb{Q}) and 𝒟×{±1}×{ζpi}ip×\mathcal{D}\times\{\pm 1\}\times\{\zeta_{p}^{i}\}_{i\in\mathbb{Z}_{p}^{\times}} given by σ(σ(D0),ϵσ,σ(ζp))\sigma\mapsto(\sigma(D_{0}),\epsilon_{\sigma},\sigma(\zeta_{p})). If ΔFp\sqrt{\Delta}\in F_{p}, then there exists a bijection between Gal(K/)\mathrm{Gal}(K/\mathbb{Q}) and 𝒟×{ζpi}ip×\mathcal{D}\times\{\zeta_{p}^{i}\}_{i\in\mathbb{Z}_{p}^{\times}} given by σ(σ(D0),σ(ζp))\sigma\mapsto(\sigma(D_{0}),\sigma(\zeta_{p})).

Proof.

σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}) permutes the roots of hh. Because all roots of hh are expressible in terms of β\beta and ζp\zeta_{p} (as β1=c/β\beta_{1}=c/\beta), σ\sigma is determined entirely by σ(β)\sigma(\beta) and σ(ζp)\sigma(\zeta_{p}).

If ΔFp\sqrt{\Delta}\notin F_{p}, there are 2p2p choices for σ(β)\sigma(\beta) and p1p-1 choices for σ(ζp)\sigma(\zeta_{p}), so |Gal(K/)|2p(p1)|\mathrm{Gal}(K/\mathbb{Q})|\leq 2p(p-1). But |Gal(K/)|=2p(p1)|\mathrm{Gal}(K/\mathbb{Q})|=2p(p-1), so all choices must correspond to distinct elements of the Galois group. The action σ(β)\sigma(\beta) determines (σ(B0),ϵσ)(\sigma(B_{0}),\epsilon_{\sigma}), and vice-versa, and the actions on B0,D0B_{0},D_{0} are also equivalent. Therefore, σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}) corresponds exactly to a choice of (Di,±1,ζpj)(D_{i},\pm 1,\zeta^{j}_{p}). ∎

If ΔFp\sqrt{\Delta}\in F_{p}, σ(ζp)\sigma(\zeta_{p}) determines σ(Δ)\sigma(\sqrt{\Delta}). Therefore, (σ(ζp),σ(D0))(\sigma(\zeta_{p}),\sigma(D_{0})) determines σ(β)\sigma(\beta) and thus σ\sigma. There are pp choices for σ(D0)\sigma(D_{0}) and p1p-1 choices for σ(ζp)\sigma(\zeta_{p}), so |Gal(K/)|p(p1)|\mathrm{Gal}(K/\mathbb{Q})|\leq p(p-1). But |Gal(K/)|=p(p1)|\mathrm{Gal}(K/\mathbb{Q})|=p(p-1), so all choices must correspond to distinct elements of the Galois group. Therefore, σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}) corresponds exactly to a choice of (Di,ζpj)(D_{i},\zeta_{p}^{j}). ∎

5. The Galois groups of hh and dhd_{h}

As in Section 4, we assume that

h(x)=f(xp)=x2p+bxp+cp,h(x)=f(x^{p})=x^{2p}+bx^{p}+c^{p},

that h(x)h(x) is irreducible and let dhd_{h} denote the polynomial dh=Dp(x,c)+bd_{h}=D_{p}(x,c)+b. The assumption on hh shows that dhd_{h} is irreducible.

Theorem 5.1.

Let LL be the splitting field of dhd_{h} over KK. Then K=L(Δ)K=L(\sqrt{\Delta}) and for all ii,

Di(ζp+ζp1,Δ(ζpζp1))LD_{i}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1}))\subset L
Proof.

LL contains all of the roots of dhd_{h}, which splits in KK, so DiLKD_{i}\subset L\subset K for each ii. Since ζpiβ+ζpiβ1Di\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1}\in D_{i}, each term is in LL. Then LL contains

(ζpβ+ζp1β1)+(ζp1β+ζpβ1)(ζp0β+ζp0β1)=(β+β1)(ζp+ζp1)β+β1=ζp+ζp1\frac{(\zeta_{p}\beta+\zeta_{p}^{-1}\beta_{1})+(\zeta_{p}^{-1}\beta+\zeta_{p}\beta_{1})}{(\zeta_{p}^{0}\beta+\zeta_{p}^{-0}\beta_{1})}=\frac{(\beta+\beta_{1})(\zeta_{p}+\zeta_{p}^{-1})}{\beta+\beta_{1}}=\zeta_{p}+\zeta_{p}^{-1}

as well.

Now choose an arbitrary σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}) which fixes LL. σ\sigma acts trivially on 𝒟\mathcal{D} and \mathcal{B} and acts on Δ\sqrt{\Delta} either trivially or by conjugation. If σ(Δ)=Δ\sigma(\sqrt{\Delta})=\sqrt{\Delta}, then σ(ζpiβ)=ζpiβ\sigma(\zeta_{p}^{i}\beta)=\zeta_{p}^{i}\beta, so σ(ζp)=ζp\sigma(\zeta_{p})=\zeta_{p}. Thus:

σ(Δ(ζpζp1))\displaystyle\sigma(\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})) =σ(Δ)(σ(ζp)(σ(ζp))1)\displaystyle=\sigma(\sqrt{\Delta})(\sigma(\zeta_{p})-(\sigma(\zeta_{p}))^{-1})
=Δ(ζpζp1)\displaystyle=\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})

Similarly, if σ(Δ)=Δ\sigma(\sqrt{\Delta})=-\sqrt{\Delta}, then σ(ζpiβ)=ζpiβ1\sigma(\zeta_{p}^{i}\beta)=\zeta_{p}^{-i}\beta_{1}, so σ(ζp)=ζp1\sigma(\zeta_{p})=\zeta_{p}^{-1}. Thus,

σ(Δ(ζpζp1))\displaystyle\sigma(\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})) =σ(Δ)(σ(ζp)(σ(ζp))1)\displaystyle=\sigma(\sqrt{\Delta})(\sigma(\zeta_{p})-(\sigma(\zeta_{p}))^{-1})
=Δ(ζp1ζp)\displaystyle=-\sqrt{\Delta}(\zeta_{p}^{-1}-\zeta_{p})
=Δ(ζpζp1)\displaystyle=\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})

Because Δ(ζpζp1)\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1}) is fixed by the subgroup fixing LL, Δ(ζpζp1)L\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})\in L. Finally, L(Δ)KL(\sqrt{\Delta})\subset K contains each Bi=Di(Δ)B_{i}=D_{i}(\sqrt{\Delta}), so K=L(Δ)K=L(\sqrt{\Delta}). ∎

From this point, we will need to construct automorphisms using Theorem 4.7, rather than considering arbitrary automorphisms as we have done in the previous section. We can prove:

Theorem 5.2.

The splitting field LL of dhd_{h} is D0(ζp+ζp1,Δ(ζpζp1))D_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})).

To help to understand the proof of Theorem 5.2, Figures 1,  23 provide diagrams of the relevant field inclusions in each case. All of the boxed fields are Galois over \mathbb{Q}, and the fields marked ``p×"``p\times" are conjugates.

Proof.

D0(ζp+ζp1,Δ(ζpζp1))LD_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1}))\subset L from Theorem 5.1. Since

K=B0(ζp)=D0(ζp,Δ)D0(ζp+ζp1,Δ(ζpζp1),Δ)K=B_{0}(\zeta_{p})=D_{0}(\zeta_{p},\sqrt{\Delta})\subset D_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1}),\sqrt{\Delta})

and Δ\sqrt{\Delta} is of degree at most 2, we see that that

[K:L][K:D0(ζp+ζp1,Δ(ζpζp1))]2[K:L]\leq[K:D_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1}))]\leq 2

Recall that Fp=(ζp)F_{p}=\mathbb{Q}(\zeta_{p}) is the cyclotomic field. There are then 3 cases:

  1. (1)

    ΔFp\sqrt{\Delta}\notin F_{p}; or

  2. (2)

    ΔFp\sqrt{\Delta}\in F_{p} and p1mod4p\equiv 1\bmod 4; or

  3. (3)

    ΔFp\sqrt{\Delta}\in F_{p} and p3mod4p\equiv 3\bmod 4.

Case (1): Consider the case of ΔFp\sqrt{\Delta}\notin F_{p}. From Theorem 4.7, there exists an element of σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}) defined by (σ(D0),ϵσ,σ(ζp))=(D0,1,ζp1)(\sigma(D_{0}),\epsilon_{\sigma},\sigma(\zeta_{p}))=(D_{0},-1,\zeta_{p}^{-1}). Then

σ(ζpiβ+ζpiβ1)=σ(ζp)iσ(β)+σ(ζp)iσ(β1)=ζpiβ1+ζpiβ\sigma(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1})=\sigma(\zeta_{p})^{i}\sigma(\beta)+\sigma(\zeta_{p})^{-i}\sigma(\beta_{1})=\zeta_{p}^{-i}\beta_{1}+\zeta_{p}^{i}\beta

for all ii, so σ\sigma fixes all roots of dhd_{h} and therefore LL. The trivial element also fixes LL, so by the Galois correspondence, [K:L]2[K:L]\geq 2. So [K:L]=2[K:L]=2, and L=D0(ζp+ζp1,Δ(ζpζp1))L=D_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})) after all, and is of degree 122p(p1)=p(p1)\frac{1}{2}\cdot 2p(p-1)=p(p-1), using Lemma 4.5.

Case (2): Now suppose that ΔFp\sqrt{\Delta}\in F_{p} and p1mod4p\equiv 1\bmod 4. Then [(ζp+ζp1):]=p12[\mathbb{Q}(\zeta_{p}+\zeta_{p}^{-1}):\mathbb{Q}]=\frac{p-1}{2} is even, so (Δ)(ζp+ζp1)\mathbb{Q}(\sqrt{\Delta})\subset\mathbb{Q}(\zeta_{p}+\zeta_{p}^{-1}) by the Galois correspondence. Thus

D0(ζp+ζp1,Δ(ζpζp1))=D0(ζp+ζp1,(ζpζp1),Δ)B0(ζp)=KLD_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1}))=D_{0}(\zeta_{p}+\zeta_{p}^{-1},(\zeta_{p}-\zeta_{p}^{-1}),\sqrt{\Delta})\supset B_{0}(\zeta_{p})=K\supset L

and in fact K=L=D0(ζp+ζp1,Δ(ζpζp1))K=L=D_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})), which by Lemma 4.5 is of degree p(p1)p(p-1).

Case (3): Finally, we suppose that ΔFp\sqrt{\Delta}\in F_{p}, but p3mod4p\equiv 3\bmod 4. Then [(ζp+ζp1):]=p12[\mathbb{Q}(\zeta_{p}+\zeta_{p}^{-1}):\mathbb{Q}]=\frac{p-1}{2} is odd, so Δ(ζp+ζp1)\sqrt{\Delta}\notin\mathbb{Q}(\zeta_{p}+\zeta_{p}^{-1}) by the Galois correspondence (else (Δ)(ζp+ζp1)\mathbb{Q}(\sqrt{\Delta})\subset\mathbb{Q}(\zeta_{p}+\zeta_{p}^{-1})). From Theorem 4.7, there exists an element σGal(K/)\sigma\in\mathrm{Gal}(K/\mathbb{Q}) defined by (σ(D0),σ(ζp))=(D0,ζp1)(\sigma(D_{0}),\sigma(\zeta_{p}))=(D_{0},\zeta_{p}^{-1}), and consequently σ(Δ)=Δ\sigma(\sqrt{\Delta})=-\sqrt{\Delta} because Δ(ζp+ζp1)\sqrt{\Delta}\notin\mathbb{Q}(\zeta_{p}+\zeta_{p}^{-1}), the fixed field of conjugation in FpF_{p}. Then as in case (1), σ\sigma fixes all roots of dhd_{h} and therefore LL, so [K:L]2[K:L]\geq 2. Then [K:L]=2[K:L]=2, and L=D0(ζp+ζp1,Δ(ζpζp1))=D0(ζp+ζp1)L=D_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1}))=D_{0}(\zeta_{p}+\zeta_{p}^{-1}), after all, and is of degree p(p1)2\frac{p(p-1)}{2}, using Lemma 4.5. ∎

Remark 5.3.

The choice of the automorphism σ\sigma fixing LL follows Jones’ construction in [5, Section 3, p. 6].

\boxed{\mathbb{Q}}p×Dip\times D_{i}(Δ)\boxed{\mathbb{Q}(\sqrt{\Delta})}Fp\boxed{F_{p}}L\boxed{L}p×Bip\times B_{i}K\boxed{K}pp2222ppp1p-1p1p-1p1p-1222p2p
Figure 1. The field diagram when ΔFp\sqrt{\Delta}\notin F_{p}
\boxed{\mathbb{Q}}p×Dip\times D_{i}(Δ)\boxed{\mathbb{Q}(\sqrt{\Delta})}p×Bip\times B_{i}(ζp+ζp1)\boxed{\mathbb{Q}(\zeta_{p}+\zeta_{p}^{-1})}p×Di(ζp+ζp1)p\times D_{i}(\zeta_{p}+\zeta_{p}^{-1})Fp\boxed{F_{p}}K\boxed{K}pp2222ppp14\frac{p-1}{4}p14\frac{p-1}{4}pp2222pp
Figure 2. The field diagram when ΔFp\sqrt{\Delta}\in F_{p} and p1mod4p\equiv 1\bmod 4
\boxed{\mathbb{Q}}(Δ)\boxed{\mathbb{Q}(\sqrt{\Delta})}p×Dip\times D_{i}p×Bip\times B_{i}(ζp+ζp1)\boxed{\mathbb{Q}(\zeta_{p}+\zeta_{p}^{-1})}Fp\boxed{F_{p}}L\boxed{L}K\boxed{K}pp2222ppp12\frac{p-1}{2}p12\frac{p-1}{2}22p12\frac{p-1}{2}ppp12\frac{p-1}{2}pp22
Figure 3. The field diagram when ΔFp\sqrt{\Delta}\in F_{p} and p3mod4p\equiv 3\bmod 4.

We are almost ready to prove Theorem 1.3, but first, we must recall the following fact about cyclotomic fields:

Lemma 5.4.

[1, VI, Thm 3.3] Let pp be an odd prime and let (1p)=(1)p(p1)/2(\frac{-1}{p})=(-1)^{p(p-1)/2} be the quadratic Legendre symbol. Let K=((1p)p)K=\mathbb{Q}\left(\sqrt{(\frac{-1}{p})p}\,\right). Then K(ζp)K\subset\mathbb{Q}(\zeta_{p}) and KK is the only quadratic extension of \mathbb{Q} that is a subfield of (ζp)\mathbb{Q}(\zeta_{p}).

Corollary 5.5.

Δ(ζp)\sqrt{\Delta}\in\mathbb{Q}(\zeta_{p}) if and only if Δ(1)p(p1)/2p2\Delta\in(-1)^{p(p-1)/2}p\mathbb{Q}^{2}.

Finally, we can prove Theorem 1.3.

Proof of Theorem 1.3.

Recall that we are studying irreducible polynomials hh of the form x2p+bxp+cpx^{2p}+bx^{p}+c^{p}. We note that for any irreducible hh of this form, Δ:=b24c\Delta:=b^{2}-4c must fall into exactly one of the 3 cases given. And from Lemmas 4.4 and 5.2, dhd_{h} is also irreducible with splitting field L=(β+β1,ζp+ζp1,Δ(ζpζp1))L=\mathbb{Q}(\beta+\beta_{1},\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})). We will use throughout the notation of Di:=(ζpiβ+ζpiβ1)D_{i}:=\mathbb{Q}(\zeta_{p}^{i}\beta+\zeta_{p}^{-i}\beta_{1}). Note throughout that while we will be examining specific elements of Gal(K/)\mathrm{Gal}(K/\mathbb{Q}), we will be considering their actions on the roots of dd, which is equivalent to the canonical actions of their images in Gal(L/)\mathrm{Gal}(L/\mathbb{Q}) under the division map.

Proof of Case 1. If b24c(1)p(p1)/2p2b^{2}-4c\notin(-1)^{p(p-1)/2}p\mathbb{Q}^{2}, then ΔFp\sqrt{\Delta}\notin F_{p} by Corollary 5.5. From above, this means that L=D0(ζp+ζp1,Δ(ζpζp1))L=D_{0}(\zeta_{p}+\zeta_{p}^{-1},\sqrt{\Delta}(\zeta_{p}-\zeta_{p}^{-1})) is an extension of degree p(p1)p(p-1) which does not contain Δ\sqrt{\Delta}.

Consider now the elements σ,τGal(K/)\sigma,\tau\in\mathrm{Gal}(K/\mathbb{Q}) defined using Theorem 4.7 by:

(σ(D0),ϵσ,σ(ζp))=(D0,1,ζpr)(τ(D0),ϵτ,τ(ζp))=(D1,1,ζp)\begin{split}(\sigma(D_{0}),\epsilon_{\sigma},\sigma(\zeta_{p}))&=(D_{0},1,\zeta_{p}^{r})\\ (\tau(D_{0}),\epsilon_{\tau},\tau(\zeta_{p}))&=(D_{1},1,\zeta_{p})\\ \end{split}

Where ζpr\zeta_{p}^{r} is a generator of the group of ppth roots of unity. By inspection, σ(Di)=Dri\sigma(D_{i})=D_{ri} and τ(Di)=Di+1\tau(D_{i})=D_{i+1}. Let SpS_{p} be the symmetric group on the set p={0,,p1}\mathbb{Z}_{p}=\{0,\dots,p-1\}. The Galois group of dhd_{h} therefore contains the elements

τ1=(1rr2rp2)Sp,τ2=(0 1 2p1)Sp\begin{split}\tau_{1}=&\ (1\ r\ r^{2}\ \cdots\ r^{p-2})\in S_{p},\\ \tau_{2}=&\ (0\ 1\ 2\ \cdots\ p-1)\in S_{p}\end{split}

The elements τ1\tau_{1}, τ2\tau_{2} generate Aff(𝔽p)\mathrm{Aff}(\mathbb{F}_{p}), whose size is p(p1)p(p-1). As |Gal(L/)|=p(p1)|\mathrm{Gal}(L/\mathbb{Q})|=p(p-1), Gal(L/)Aff(𝔽p)\mathrm{Gal}(L/\mathbb{Q})\simeq\mathrm{Aff}(\mathbb{F}_{p}). And Gal((Δ)/)C2\mathrm{Gal}(\mathbb{Q}(\sqrt{\Delta})/\mathbb{Q})\simeq C_{2}, so the Galois group of hh is

Gal(K/)Gal(L/)×Gal((Δ/)Aff(𝔽p)×C2\mathrm{Gal}(K/\mathbb{Q})\simeq\mathrm{Gal}(L/\mathbb{Q})\times\mathrm{Gal}(\mathbb{Q}(\sqrt{\Delta}/\mathbb{Q})\simeq\mathrm{Aff}(\mathbb{F}_{p})\times C_{2}

Proof of Case 2. If p1mod4p\equiv 1\bmod 4 and b24cp2b^{2}-4c\in p\mathbb{Q}^{2}, then ΔFp\sqrt{\Delta}\in F_{p} by Corollary 5.5. From above, L=KL=K is the splitting field of dhd_{h} and hh, and |Gal(K/)|=|Gal(L/)|=p(p1)|\mathrm{Gal}(K/\mathbb{Q})|=|\mathrm{Gal}(L/\mathbb{Q})|=p(p-1). Now consider elements σ,τGal(K/)\sigma,\tau\in\mathrm{Gal}(K/\mathbb{Q}) defined using Theorem 4.7 by:

(σ(D0),σ(ζp))=(D0,ζpr)(τ(D0),τ(ζp))=(D1,ζp)\begin{split}(\sigma(D_{0}),\sigma(\zeta_{p}))&=(D_{0},\zeta_{p}^{r})\\ (\tau(D_{0}),\tau(\zeta_{p}))&=(D_{1},\zeta_{p})\\ \end{split}

For ζpr\zeta_{p}^{r} a generator. σ\sigma conjugates ΔFp\sqrt{\Delta}\in F_{p}\setminus\mathbb{Q}, so σ(ζpiβ)=ζpriβ1\sigma(\zeta_{p}^{i}\beta)=\zeta_{p}^{ri}\beta_{1}, and σ(Di)=Dri\sigma(D_{i})=D_{-ri}. ζpr\zeta_{p}^{-r} is a generator because

(rp)=(1p)(rp)=(1)(1)=1\left(\frac{-r}{p}\right)=\left(\frac{-1}{p}\right)\left(\frac{r}{p}\right)=(1)(-1)=-1

So σ\sigma will be a p1p-1-cycle. And again, τ(Di)=Di+1\tau(D_{i})=D_{i+1}. The Galois group of dhd_{h} therefore contains the elements

τ1\tau_{1} == (1(1 r-r (r)2(-r)^{2} \ldots (r)(p2)(-r)^{(p-2)} ))
τ2\tau_{2} == (0(0 11 22 \ldots p1p-1 ))

As in Case 1, the elements τ1\tau_{1}, τ2\tau_{2} generate Aff(𝔽p)\mathrm{Aff}(\mathbb{F}_{p}), and |Aff(𝔽p)|=p(p1)=|Gal(K/)|=|Gal(L/)||\mathrm{Aff}(\mathbb{F}_{p})|=p(p-1)=|\mathrm{Gal}(K/\mathbb{Q})|=|\mathrm{Gal}(L/\mathbb{Q})|, so the Galois groups of dh,hd_{h},h are identically

Gal(K/)Gal(L/)Aff(𝔽p)\mathrm{Gal}(K/\mathbb{Q})\simeq\mathrm{Gal}(L/\mathbb{Q})\simeq\mathrm{Aff}(\mathbb{F}_{p})

Proof of Case 3. If p3mod4p\equiv 3\bmod 4 and b24cp2b^{2}-4c\in-p\mathbb{Q}^{2}, then ΔFp\sqrt{\Delta}\in F_{p} by Corollary 5.5. From above, |Gal(K/)|=p(p1)|\mathrm{Gal}(K/\mathbb{Q})|=p(p-1), but |Gal(L/)|=p(p1)2|\mathrm{Gal}(L/\mathbb{Q})|=\frac{p(p-1)}{2}. Again we consider elements σ,τGal(K/)\sigma,\tau\in\mathrm{Gal}(K/\mathbb{Q}) defined using Theorem 4.7 by:

(σ(D0),σ(ζp))=(D0,ζpr)(τ(D0),τ(ζp))=(D1,ζp)\begin{split}(\sigma(D_{0}),\sigma(\zeta_{p}))&=(D_{0},\zeta_{p}^{r})\\ (\tau(D_{0}),\tau(\zeta_{p}))&=(D_{1},\zeta_{p})\\ \end{split}

For ζpr\zeta_{p}^{r} a generator. σ\sigma again conjugates ΔFp\sqrt{\Delta}\in F_{p}\setminus\mathbb{Q}, so σ(ζpiβ)=ζpriβ1\sigma(\zeta_{p}^{i}\beta)=\zeta_{p}^{ri}\beta_{1}, and σ(Di)=Dri\sigma(D_{i})=D_{-ri}. However, we see that ζpr\zeta_{p}^{-r} is of order p12\frac{p-1}{2} because

(rp)=(1p)(rp)=(1)p(p1)/2(1)=1\left(\frac{-r}{p}\right)=\left(\frac{-1}{p}\right)\left(\frac{r}{p}\right)=(-1)^{p(p-1)/2}(-1)=1

So σ\sigma will be two p12\frac{p-1}{2}-cycles, the square of the p1p-1-cycle generated by r\sqrt{-r}. And again, τ(Di)=Di+1\tau(D_{i})=D_{i+1}. The Galois group of dhd_{h} therefore contains the elements τ=τ1τ2,τ3Sp\tau=\tau_{1}\tau_{2},\tau_{3}\in S_{p}, where τ1,τ2\tau_{1},\tau_{2} are the (p1)/2(p-1)/2 cycles

τ1\tau_{1} == (1(1 r-r (r)2(-r)^{2} \ldots (r)(p3)/2(-r)^{(p-3)/2} ))
τ2\tau_{2} == (1(-1 rr (r)2-(-r)^{2} \ldots (r)(p3)/2-(-r)^{(p-3)/2} ))

and τ3\tau_{3} is the pp cycle

τ3\tau_{3} == (0(0 11 22 \ldots p1p-1 ))

The elements τ\tau, τ3\tau_{3} generate the normal subgroup CpC(p1)/2Aff(𝔽p)C_{p}\rtimes C_{(p-1)/2}\vartriangleleft\mathrm{Aff}(\mathbb{F}_{p}), whose size is p(p1)2=|Gal(L/)|\frac{p(p-1)}{2}=|\mathrm{Gal}(L/\mathbb{Q})|. Thus, the Galois group of dhd_{h} is

Gal(L/)CpC(p1)/2Aff(𝔽p)\mathrm{Gal}(L/\mathbb{Q})\simeq C_{p}\rtimes C_{(p-1)/2}\vartriangleleft\mathrm{Aff}(\mathbb{F}_{p})

Now ΔL\sqrt{\Delta}\notin L and L(Δ)=KL(\sqrt{\Delta})=K as in Case 1, so

Gal(K/)Gal(L/)×C2(CpC(p1)/2)×C2\mathrm{Gal}(K/\mathbb{Q})\simeq\mathrm{Gal}(L/\mathbb{Q})\times C_{2}\simeq(C_{p}\rtimes C_{(p-1)/2})\times C_{2}

Remark 5.6.

The automorphisms σ,τ\sigma,\tau we choose in the proof of Theorem 1.3 are again as in Jones’ [5, Section 3, p. 6].

Remark 5.7.

In the cases of p=2p=2 and p=3p=3 (the solutions to which we mention in Section 1), our methods of proof fail beginning at Lemma 4.1, where we use the fact p12pp-1\mid 2p. The issue is essentially that ζ2=1\zeta_{2}=-1\in\mathbb{Q} for the case of p=2p=2 and that if ΔF3\sqrt{\Delta}\in F_{3}, then (Δ)=F3\mathbb{Q}(\sqrt{\Delta})=F_{3} in the case of p=3p=3. These issues both invalidate the distinctness of the BiB_{i} and DiD_{i}, which is key to Lemma 4.4, on which all of our subsequent work regarding dhd_{h} and its Galois group is based.

Remark 5.8.

From Corollary 1.2 of [5], there exist an infinite number of polynomials of the form of case 1 of Theorem 1.3.

References

  • [1] Serge Lang “Algebra” Springer, 2002 DOI: 10.1007/978-1-4613-0041-0
  • [2] Joshua Harrington and Lenny Jones “The Irreducibility of Power Compositional Sextic Polynomials and Their Galois Groups” In Mathematica Scandinavica 102.2, 2017 DOI: https://doi.org/10.7146/math.scand.a-25850
  • [3] Chad Awtrey, James R. Buerle and Hanna Noelle Griesbach “Field Extensions Defined by Power Compositional Polynomials” In Missouri Journal of Mathematical Sciences 33.2, 2021, pp. 163–180 DOI: 10.35834/2021/3302163
  • [4] Malcolm Hoong Wai Chen, Angelina Yan Mui Chin and Ta Sheng Tan “Galois groups of certain even octic polynomials” In Journal of Algebra and its Applications, 2022 DOI: https://doi.org/10.1142/S0219498823502638
  • [5] Lenny Jones “Monogenic Reciprocal Trinomials and Their Galois Groups” In Journal of Algebra and its Applications, 2020 DOI: https://doi.org/10.1142/S0219498822500268
  • [6] L.E. Dickson “The analytic representation of substitutions on a power of a prime number of letters with a discussion of the linear group I, II” In Annals of Mathematics 11.1, 1897, pp. 65–120\bibrangessep161–183
  • [7] R. Lidl, G.L. Mullen and G. Turnwald “Dickson polynomials” Longman Group, 1993
  • [8] A.. Horadam “Vieta Polynomials” In Fibonacci Quarterly, 2002, pp. 223–232
  • [9] Andrzej Schinzel “Polynomials with Special Regard to Reducibility” Cambridge University Press, 2000
  • [10] Andrjez Schinzel “On reducible trinomials” In Dissertationes Mathematicae, 1993
  • [11] Chad Awtrey and Peter Jakes “Subfields of Solvable Sextic Field Extensions” In North Carolina Journal of Mathematics and Statistics 4, 2018, pp. 1–11
  • [12] Biswajit Koley and A. Reddy “Survey on irreducibility of trinomials”, 2020 arXiv:2012.07568 [math.HO]