1 Introduction

Let p be a prime number, and let \(1+p\mathbb {Z}_p\) denote the pro-p group of principal units of the ring of p-adic integers \(\mathbb {Z}_p\)—namely, \(1+p\mathbb {Z}_p=\{1+p\lambda \,\mid \,\lambda \in \mathbb {Z}_p\}\). An oriented pro-p group is a pair \((G,\theta )\) consisting of a pro-p group G and a morphism of pro-p groups \(\theta :G\rightarrow 1+p\mathbb {Z}_p\), called an orientation of G (see [29]; oriented pro-p groups were introduced by Efrat [7], with the name “cyclotomic pro-p pairs”). An oriented pro-p group \((G,\theta )\) gives rise to the continuous G-module \(\mathbb {Z}_p(\theta )\), which is equal to \(\mathbb {Z}_p\) as an abelian pro-p group, and which is endowed with the continuous G-action defined by

$$\begin{aligned} g\cdot \lambda =\theta (g)\cdot \lambda \qquad \text {for all }g\in G\text { and }\lambda \in \mathbb {Z}_p(\theta ). \end{aligned}$$

An oriented pro-p group \((G,\theta )\) is said to be Kummerian if the following cohomological condition is satisfied: for every \(n\ge 1\) the natural morphism

$$\begin{aligned} \textrm{H}^1(G,\mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta ))\longrightarrow \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z}), \end{aligned}$$
(1.1)

induced by the epimorphism of continuous G-modules \(\mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\twoheadrightarrow \mathbb {Z}/p\) is surjective (see [11])—here we consider \(\mathbb {Z}/p\) as a trivial G-module. Moreover, the oriented pro-p group \((G,\theta )\) is said to be 1-cyclotomic if the above cohomological condition is satisfied also for every closed subgroup of G—namely, the natural morphism (1.1) is surjective also with H instead of G, and the restriction \(\theta \vert _H:H\rightarrow 1+p\mathbb {Z}_p\) instead of \(\theta \) for all closed subgroups H of G (in [25, 26] a 1-cyclotomic oriented pro-p group is called a “1-smooth” oriented pro-p group). This cohomological condition was considered first by J. Labute, who showed ante litteram that for every Demushkin group G there exists precisely one orientation which completes G into a Kummerian oriented pro-p group, namely the orientation induced by the dualizing module of G (see [13]).

In case of trivial orientations, 1-cyclotomicity translates into a purely group-theoretical statement. Namely, an oriented pro-p group \((G,\textbf{1})\)—where \(\textbf{1}:G\rightarrow 1+p\mathbb {Z}_p\) denotes the orientation which is constantly equal to 1—is 1-cyclotomic if, and only if, the abelianization of every closed subgroup of G is a free abelian pro-p group. Pro-p groups satisfying this group-theoretic condition are called absolutely torsion-free pro-p groups, and they were introduced by Würfel [36].

The main goal of this work is to produce new examples of pro-p groups which no orientations can turn into a 1-cyclotomic oriented pro-p group.

Theorem 1.1

Let G be a pro-p group with pro-p presentation

$$\begin{aligned} G =\left\langle \, x,y_1,\ldots ,y_{d_1},z_1,\ldots ,z_{d_2}\,\mid \, r_1=r_2=1 \,\right\rangle , \end{aligned}$$
(1.2)

where \(d_1,d_2\) are two positive odd integers, and either:

  1. (1.1.a)

    \(d_1+d_2\ge 4\) and

    $$\begin{aligned}\begin{aligned} r_1&=[x,y_1][y_2,y_3]\cdots [y_{d_1-1},y_{d_1}], \\ r_2&=[x,z_1][z_2,z_3]\cdots [z_{d_2-1},z_{d_2}]; \end{aligned} \end{aligned}$$
  2. (1.1.b)

    or p is odd and

    $$\begin{aligned}\begin{aligned} r_1&=y_1^p[y_1,x][y_2,y_3]\cdots [y_{d_1-1},y_{d_1}],\\ r_2&= z_1^p[z_1,x][z_2,z_3]\cdots [z_{d_2-1},z_{d_2}]. \end{aligned}\end{aligned}$$

Then there are no orientations \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) such that the oriented pro-p group \((G,\theta )\) is 1-cyclotomic.

It is worth underlining that the pro-p groups described in Theorem 1.1 are amalgamated free pro-p products of two Demushkin groups—the subgroup generated by \(x,y_1,\ldots ,y_{d_1}\) and the subgroup generated by \(x,z_1,\ldots ,z_{d_2}\)—, with pro-p-cyclic amalgam, generated by x. Despite Demushkin groups and their free pro-p products are some of the (extremely few) examples of pro-p groups which are known to give rise to 1-cyclotomic oriented pro-p groups, the presence of a pro-p-cyclic amalgam is enough to lose 1-cyclotomicity.

Oriented pro-p groups satisfying 1-cyclotomicity have great prominence in Galois theory. Given a field \(\mathbb {K}\), let \({\bar{\mathbb {K}}}_s\) and \(\mathbb {K}(p)\) denote respectively the separable closure of \(\mathbb {K}\), and the compositum of all finite Galois p-extensions of \(\mathbb {K}\). The maximal pro-p Galois group of \(\mathbb {K}\), denoted by \(G_{\mathbb {K}}(p)\), is the maximal pro-p quotient of the absolute Galois group \({{\,\textrm{Gal}\,}}({\bar{\mathbb {K}}}_s/\mathbb {K})\) of \({\mathbb {K}}\), and it coincides with the Galois group of the Galois extension \(\mathbb {K}(p)/\mathbb {K}\). Detecting maximal pro-p Galois groups among pro-p groups, are crucial problems in Galois theory. Already the pursuit of concrete examples of pro-p groups which do not occur as maximal pro-p Galois groups of fields is already considered a very remarkable challenge (see, e.g., [1, 3, 4, 24, 33]).

The maximal pro-p Galois group \(G_{\mathbb {K}}(p)\) of a field \(\mathbb {K}\) containing a root of 1 of order p gives rise to the oriented pro-p group \((G_{\mathbb {K}}(p),\theta _{\mathbb {K}})\), where

$$\begin{aligned} \theta _{\mathbb {K}}:G_{\mathbb {K}}(p)\longrightarrow 1+p\mathbb {Z}_p \end{aligned}$$

denotes the pro-p cyclotomic character (see Example 2.4 below). By Kummer theory, the oriented pro-p group \((G_{\mathbb {K}}(p),\theta _{\mathbb {K}})\) is 1-cyclotomic (see [13, p. 131] and [11, Sect. 4])—in case \(p=2\) we need to assume further that \(\sqrt{-1}\in \mathbb {K}\). Therefore, a pro-p group which cannot complete into a 1-cyclotomic oriented pro-p group does not occur as the maximal pro-p group of a field containing a root of 1 of order p—and hence neither as the absolute Galois group of any field (see, e.g., [24, Remark 3.3]). Hence, the following corollary may be deduced directly from Theorem 1.1.

Corollary 1.2

A pro-p group G as in Theorem 1.1 does not occur as the maximal pro-p Galois group of any field containing a root of 1 of order p (and also \(\sqrt{-1}\) if \(p=2\)). Hence, G does not occur as the absolute Galois group of any field.

In the recent past, other cohomological properties have been used to study maximal pro-p Galois groups—and to find examples of pro-p groups which do not occur as maximal pro-p Galois groups. By the Norm Residue Theorem—proved by M. Rost and V. Voevodsky, with the contribution by Ch. Weibel, see [12, 34]—one knows that if \(\mathbb {K}\) is a field containing a root of 1 of order p, the \(\mathbb {Z}/p\) -cohomology algebra \(\textbf{H}^\bullet (G_{\mathbb {K}}(p),\mathbb {Z}/p\mathbb {Z})\), endowed with the cup-product

$$\begin{aligned} \textvisiblespace \smallsmile \textvisiblespace :\textrm{H}^m(G_{\mathbb {K}}(p),\mathbb {Z}/p\mathbb {Z})\times \textrm{H}^n(G_{\mathbb {K}}(p),\mathbb {Z}/p\mathbb {Z})\longrightarrow \textrm{H}^{m+n}(G_{\mathbb {K}}(p),\mathbb {Z}/p\mathbb {Z}), \end{aligned}$$

is quadratic, i.e., its ring structure is completely determined by the 1st and the 2nd cohomology groups (see, e.g., [22, Sect. 2]). Moreover, it was shown by E. Matzri that if \(\mathbb {K}\) is a field containing a root of 1 of order p, then \(G_{\mathbb {K}}(p)\) satisfies the triple Massey vanishing property (see [9] and references therein)—for an overview on Massey products in Galois cohomology see [19]. These two cohomological properties were used to find examples of pro-p groups which do not occur as maximal pro-p Galois groups of fields containing a root of 1 of order p, for example in [4, Sect. 8] and in [19, Sect. 7].

We prove that the pro-p groups described in Theorem 1.1 cannot be ruled out as maximal pro-p Galois groups employing the above two cohomological obstructions.

Proposition 1.3

Let G be a pro-p group as in Theorem 1.1.

  1. (i)

    The \(\mathbb {Z}/p\)-cohomology algebra \(\textbf{H}^\bullet (G,\mathbb {Z}/p\mathbb {Z})\) is quadratic.

  2. (ii)

    The pro-p group G satisfies the cyclic p-Massey vanishing property—namely, the p-fold Massey product

    $$\begin{aligned} \langle \underbrace{\alpha ,\ldots ,\alpha }_{p\text { times}}\rangle \end{aligned}$$

    contains 0 for every \(\alpha \in \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\).

  3. (iii.a)

    If G is as in (1.1.a), then G satisfies the 3- and the strong 4-Massey vanishing property.

  4. (iii.b)

    If G is as in (1.1.b) and \(p>3\) then G satisfies the 3- and the strong 4-Massey vanishing property.

(We recall the basic notions on Massey products in Galois cohomology in Sect. 6.1 below.) Hence, Corollary 1.2 provides brand new examples of pro-p groups which do not occur as maximal pro-p Galois groups of fields containing a root of 1 of order p, and as absolute Galois groups. Moreover, we remark that the relations which define the pro-p groups described in Theorem 1.1 are rather “elementary”—just elementary commutators of generator times, possibly, the p-power of a generator—, unlike the examples provided in [1, 4, 19, 24], where the relations involve higher commutators.

Finally, we focus on the Minač–Tân pro-p group, i.e., the pro-p group G with pro-p presentation

$$\begin{aligned} G=\langle \,x_1,\ldots ,x_5\,\mid \,[[x_1,x_2],x_3][x_4,x_5]=1\,\rangle . \end{aligned}$$

In [19, Sect. 7], Minač and Tân showed that G does not satisfy the 3-Massey vanishing property, and thus it does not occur as the maximal pro-p Galois group of any field containing a root of 1 of order p. We prove that G cannot complete into a 1-cyclotomic oriented pro-p group.

Theorem 1.4

Let p be an odd prime. Then there are no orientations turning the Minač–Tân pro-p group into a 1-cyclotomic oriented pro-p group.

Theorem 1.4 has been proved independently by I. Snopce and P. Zalesskiĭ (unpublished). Theorem 1.4 provides a negative answer to the question posed in [29, Remark 3.7]—namely, the Minač–Tân pro-p group may be ruled out as a maximal pro-p Galois group of a field containing a root of 1 of order p (and thus as an absolute Galois group) in a “Massey-free” way.

Altogether, 1-cyclotomicity of oriented pro-p groups provides a rather powerful tool studying maximal pro-p Galois groups, and it succeeds in detecting pro-p groups which are not maximal pro-p Galois groups when other methods fail, as underlined above. We believe that further investigations in this direction will lead to new obstructions for the realization of pro-p groups as maximal pro-p Galois group.

Actually, Theorem 1.4, and the main result in [33] (see in particular [33, p. 1907]), may lead to the suspect that 1-cyclotomicity is a more restrictive condition in comparison with the vanishing of Massey products. Thus, we formulate the following conjecture.

Conjecture 1.5

Let \((G,\theta )\) be an oriented pro-p group, such that \({{\,\textrm{Im}\,}}(\theta )\subseteq 1+4\mathbb {Z}_2\) if \(p=2\). If \((G,\theta )\) is 1-cyclotomic, then the pro-p group G satisfies the 3-Massey vanishing property; if moreover G is finitely generated, then G satisfies the strong n-Massey vanishing property for every \(n\ge 3\).

After the publication on the arXiv of an earlier version of this paper, A. Merkurjev and F. Scavia proved the first statement of Conjecture 1.5—see [16, Theorem 1.3]—; while, on the other hand, there are 1-cyclotomic oriented pro-2 groups \((G,\theta )\) such that \({{\,\textrm{Im}\,}}(\theta )\subseteq 1+4\mathbb {Z}_2\), where G is not finitely generated and does not satisfy the strong 4-Massey vanishing property—see [15, Theorem 1.6]. In particular, [16, Theorem 1.3] implies Theorem 1.4 (see also [16, Remark 6.3]).

2 Oriented Pro-p Groups and Cohomology

2.1 Notation and Preliminaries

Throughout the paper, every subgroup of a pro-p group is tacitly assumed to be closed with respect to the pro-p topology. Therefore, sets of generators of pro-p groups, and presentations, are to be intended in the topological sense.

Given a pro-p group G, we denote the closed commutator subgroup of G by \(G'\)—namely, \(G'\) is the closed normal subgroup generated by commutators

$$\begin{aligned}{}[h,g]= h^{-1}\cdot h^g=h^{-1}\cdot g^{-1}hg,\qquad g,h\in G. \end{aligned}$$

The Frattini subgroup of G is denoted by \(\Phi (G)\)—namely, \(\Phi (G)\) is the closed normal subgroup generated by \(G'\) and by p-powers \(g^p\), \(g\in G\) (cf., e.g., [5, Proposition 1.13]). A minimal generating set of G gives rise to a basis of the \(\mathbb {Z}/p\mathbb {Z}\)-vector space \(G/\Phi (G)\), and conversely (cf., e.g., [5, Proposition 1.9]).

Finally, we denote the abelianization \(G/G'\) of G by \(G^{{{\,\textrm{ab}\,}}}\). Throughout the paper, we will make use of the following straightforward fact.

Fact 2.1

Let G be a finitely generated pro-p group. Then a subset \(\{x_1,\ldots ,x_d\}\) of G is a minimal generating set of G if, and only if, the subset \(\{x_1G',\ldots ,x_dG'\}\) of \(G^{{{\,\textrm{ab}\,}}}\) is a minimal generating set of the abelian pro-p group \(G^{{{\,\textrm{ab}\,}}}\).

2.2 Oriented Pro-p Groups

Let G be a pro-p group. An orientation \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) is said to be torsion-free if p is odd, or if \(p=2\) and \({{\,\textrm{Im}\,}}(\theta )\subseteq 1+4\mathbb {Z}_2\). Observe that one may have an oriented pro-p group \((G,\theta )\) where G has non-trivial torsion and \(\theta \) torsion-free (e.g., if \(G\simeq \mathbb {Z}/p\) and \({{\,\textrm{Im}\,}}(\theta )=\{1\}\)).

A morphism of oriented pro-p groups \((G_1,\theta _1)\rightarrow (G_2,\theta _2)\), is a homomorphism of pro-p groups \(\phi :G_1\rightarrow G_2\) such that \(\theta _1=\theta _2\circ \phi \) (cf. [29, Sect. 3, p. 1888]).

Within the family of oriented pro-p groups one has the following constructions. Let \((G,\theta )\) be an oriented pro-p group.

  1. (a)

    If N is a normal subgroup of G contained in \({{\,\textrm{Ker}\,}}(\theta )\), one has the oriented pro-p group \((G/N,\theta _{/N})\), where \(\theta _{/N}:G/N\rightarrow 1+p\mathbb {Z}_p\) is the orientation such that \(\theta _{/N}\circ \pi =\theta \), with \(\pi :G\rightarrow G/N\) the canonical projection.

  2. (b)

    If A is an abelian pro-p group (written multiplicatively), one has the oriented pro-p group \( A\rtimes (G,\theta )=(A\rtimes G,{{\tilde{\theta }}})\), with action given by \(gag^{-1}=a^{\theta (g)}\) for every \(g\in G\), \(a\in A\), where the orientation \({{\tilde{\theta }}}:A\rtimes G\rightarrow 1+p\mathbb {Z}_p\) is the composition of the canonical projection \(A\rtimes G\rightarrow G\) with \(\theta \).

2.3 Kummerianity and 1-Cyclotomicity

Let \((G,\theta )\) be an oriented pro-p group. Observe that the G-action on the G-module \(\mathbb {Z}_p(\theta )/p\mathbb {Z}_p(\theta )\) is trivial, as \(\theta (g)\equiv 1\bmod p\) for all \(g\in G\). Thus, \(\mathbb {Z}_p(\theta )/p\mathbb {Z}_p(\theta )\) is isomorphic to \(\mathbb {Z}/p\) as a trivial G-module.

An oriented pro-p group \((G,\theta )\) comes endowed with the distinguished subgroup

$$\begin{aligned} K_\theta (G)=\left\langle \,{}^{g}h\cdot h^{-\theta (g)}\,\mid \,g\in G,\,h\in {{\,\textrm{Ker}\,}}(\theta )\,\right\rangle \end{aligned}$$

(cf. [11, Sect. 3]). The subgroup \(K_\theta (G)\) is normal in G, and it is contained in both \({{\,\textrm{Ker}\,}}(\theta )\) and \(\Phi (G)\). On the other hand, \(K_\theta (G)\supseteq {{\,\textrm{Ker}\,}}(\theta )'\), so that \({{\,\textrm{Ker}\,}}(\theta )/K_\theta (G)\) is an abelian pro-p group. Moreover, if \(\theta \) is a torsion-free orientation, \(G/{{\,\textrm{Ker}\,}}(\theta )\simeq {{\,\textrm{Im}\,}}(\theta )\) is torsion-free, and thus either trivial or isomorphic to \(\mathbb {Z}_p\). Hence, the epimorphism \(G\twoheadrightarrow G/{{\,\textrm{Ker}\,}}(\theta )\) splits, and since \(ghg^{-1}\equiv h^{\theta (g)}\bmod K_\theta (G)\) for every \(g\in G\) and \(h\in {{\,\textrm{Ker}\,}}(\theta )\), one concludes that

$$\begin{aligned} \left( G/K_{\theta }(G),\theta _{/K_{\theta }(G)}\right) \simeq \dfrac{{{\,\textrm{Ker}\,}}(\theta )}{K_\theta (G)}\rtimes \left( G/{{\,\textrm{Ker}\,}}(\theta ),\theta _{/{{\,\textrm{Ker}\,}}(\theta )}\right) \end{aligned}$$

(cf., e.g., [30, Sect. 2.2, Eq. (2.1)]).

One has the following result relating the subgroup \(K_\theta (G)\) and the surjectivity of the maps (1.1) (cf. [11, Theorem 7.1], see also [30, Proposition 2.6]).

Proposition 2.2

Let \((G,\theta )\) be an oriented pro-p group with \(\theta \) a torsion-free orientation. The following are equivalent.

  1. (i)

    The natural map

    $$\begin{aligned} \textrm{H}^1(G,\mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta ))\longrightarrow \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z}), \end{aligned}$$

    is surjective for every positive integer n.

  2. (ii)

    The quotient \({{\,\textrm{Ker}\,}}(\theta )/K_\theta (G)\) is a free abelian pro-p group.

If an oriented pro-p group \((G,\theta )\) with torsion-free orientation satisfies the above two equivalent properties, then it is said to be Kummerian. Moreover, \((G,\theta )\) is said to be 1-cyclotomic if \((H,\theta \vert _H)\) is Kummerian for every subgroup \(H\subseteq G\).

Remark 2.3

The original definition of 1-cyclotomic oriented pro-p group requires only that for every open subgroup U of G, the oriented pro-p group \((U,\theta \vert _U)\) is Kummerian (cf. [29, Sect. 1]). By a continuity argument, this is enough to imply that the oriented pro-p group \((H,\theta \vert _H)\) is Kummerian also for every subgroup \(H\subseteq G\) (cf. [29, Corollary 3.2]).

If \((G,\textbf{1})\) is an oriented pro-p group with \(\textbf{1}:G\rightarrow 1+p\mathbb {Z}_p\) the orientation constantly equal to 1, then \(K_{\textbf{1}}(G)=G'\), and by Proposition 2.2\((G,\theta )\) is Kummerian if, and only if, \(G/G'={{\,\textrm{Ker}\,}}(\textbf{1})/K_{\textbf{1}}(G)\) is a free abelian pro-p group (cf. [11, Example 3.5–(1)]). Hence, \((G,\textbf{1})\) is 1-cyclotomic if, and only if, \(H/H'\) is a free abelian pro-p group for every subgroup \(H\subseteq G\), i.e., G is absolutely torsion-free (cf. [25, Remark 2.3]).

2.4 Examples

Example 2.4

Let \(\mathbb {K}\) be a field containing a root of 1 of order p, and also \(\sqrt{-1}\) if \(p=2\). Then the pro-p cyclotomic character \(\theta _{\mathbb {K}}\) of \(G_{K}(p)\)—induced by the action of \(G_{\mathbb {K}}(p)\) on the roots of 1 of p-power order contained in \(\mathbb {K}(p)\)—has image contained in \(1+p\mathbb {Z}_p\). Observe that \({{\,\textrm{Im}\,}}(\theta _{\mathbb {K}})=1+p^f\mathbb {Z}_p\), where \(f\in \mathbb {N}\cup \{\infty \}\) is maximal such that \(\mathbb {K}\) contains a root of 1 of order \(p^f\) (if \(f=\infty \), we set \(p^\infty =0\)). In particular, \(\theta _{\mathbb {K}}\) is a torsion-free orientation. The module \(\mathbb {Z}_p(\theta _{\mathbb {K}})\) is called the 1st Tate twist of \(\mathbb {Z}_p\) (cf., e.g., [20, Definition 7.3.6]).

For the convenience of the reader, here we recall J. Labute’s argument to show that the oriented pro-p group \((G_{\mathbb {K}}(p),\theta _{\mathbb {K}})\) is Kummerian—and thus also 1-cyclotomic, as every subgroup \(H\subseteq G_{\mathbb {K}}(p)\) is the maximal pro-p Galois group of an extension of \(\mathbb {K}\), with pro-p cyclotomic character \(\theta _{\mathbb {K}}\vert _H\)—as it is presented in [13, p. 131] (where the module \(\mathbb {Z}_p(\theta _{\mathbb {K}})\) is denoted by \(I=I(\chi ')\)). For every \(n\ge 1\) one has an isomorphism of continuous \(G_{\mathbb {K}}(p)\)-modules

$$\begin{aligned} \mathbb {Z}_p(\theta _{\mathbb {K}})/p^n\mathbb {Z}_p(\theta _\mathbb {K})\simeq \mu _{p^n}=\left\{ \,\zeta \in \mathbb {K}(p)\,\mid \,\zeta ^{p^n}=1\,\right\} . \end{aligned}$$

Let \(\mathbb {K}^\times \) and \(\mathbb {K}(p)^\times \) denote the multiplicative groups of units of \(\mathbb {K}\) and \(\mathbb {K}(p)\) respectively. By Hilbert 90, the short exact sequence of continuous \(G_{\mathbb {K}}(p)\)-modules

(2.1)

induces a commutative diagram

where the left-side vertical arrow \(\pi _n\) and the central vertical arrow are induced by the \(p^{n-1}\)th power map \(\textvisiblespace ^{p^n}:\mathbb {K}(p)^\times \rightarrow \mathbb {K}(p)^\times \), and the right-side vertical arrow is induced by the epimorphism of continuous \(G_{\mathbb {K}}(p)\)-modules \(\mathbb {Z}_p(\theta _{\mathbb {K}})/p^n\mathbb {Z}_p(\theta _{\mathbb {K}})\twoheadrightarrow \mathbb {Z}/p\mathbb {Z}\). Since the map \(\pi _n\) is surjective, also the other vertical arrows are surjective.

Example 2.5

Let G be a free pro-p group. Then the oriented pro-p group \((G,\theta )\) is 1-cyclotomic for any orientation \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) (cf. [29, Sect. 2.2]).

Example 2.6

Let G be an infinite Demushkin group (cf., e.g., [20, Definition 3.9.9]). By [13, Theorem 4], G comes endowed with a canonical orientation \(\chi :G\rightarrow 1+p\mathbb {Z}_p\) which is the only one completing G into a 1-cyclotomic oriented pro-p group. In particular, if \(d=\dim (\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z}))\) is even (which is always the case if \(p\ne 2\)), then G has a presentation

$$\begin{aligned} G=\left\langle \,x_1,\ldots ,x_d\,\mid \,x_1^{p^f}[x_1,x_2]\cdots [x_{d-1},x_d]=1\,\right\rangle , \end{aligned}$$

with \(f\ge 1\) (\(f\ge 2\) if \(p=2\)). In this case \(\chi (x_2)=(1-p^f)^{-1}\) and \(\chi (x_i)=1\) for \(i\ne 2\).

Example 2.7

Let \((G,\theta )\) be an oriented pro-p group, with \(\theta \) a torsion-free orientation. The oriented pro-p group \((G,\theta )\) is said to be \(\theta \) -abelian if the subgroup \(K_\theta (G)\) is trivial and if \({{\,\textrm{Ker}\,}}(\theta )\) is a free abelian pro-p group—in this case G is a free abelian-by-cyclic pro-p group, i.e.,

$$\begin{aligned} G\simeq {{\,\textrm{Ker}\,}}(\theta )\rtimes \dfrac{G}{{{\,\textrm{Ker}\,}}(\theta )} \end{aligned}$$

(cf. [30, Remark 2.2]). In other words, G has a presentation

$$\begin{aligned} G=\left\langle \,x_0,x_i\,\mid \,i\in I,\, x_i^{x_0}=x_i^{\theta (x_0)^{-1}},[x_i,x_j]=1\,\forall \,i,j\in I\,\right\rangle , \end{aligned}$$

for some set of indices I, and \(\theta (x_i)=1\) for all \(i\in I\) (cf. [22, Proposition 3.4]). A \(\theta \)-abelian oriented pro-p group \((G,\theta )\) is Kummerian by Proposition 2.2, as by definition \(K_\theta (G)\) is trivial and \({{\,\textrm{Ker}\,}}(\theta )\) is a free abelian pro-p group. Moreover, if H is a subgroup of G, then one has

$$\begin{aligned} H\simeq (H\cap {{\,\textrm{Ker}\,}}(\theta ))\rtimes \frac{H}{{{\,\textrm{Ker}\,}}(\theta \vert _H)} \end{aligned}$$

(cf. [30, Remark 2.4]), so that the oriented pro-p group \((H,\theta \vert _H)\) is \(\theta \vert _H\)-abelian, and thus Kummerian, and consequently \((G,\theta )\) is 1-cyclotomic.

One has the following result to check whether an oriented pro-p group is Kummerian (cf. [30, Propositions 2.6, 3.6]).

Proposition 2.8

Let \((G,\theta )\) be an oriented pro-p group, with \(\theta \) a torsion-free orientation. Then \((G,\theta )\) is Kummerian if, and only if, there exists a normal subgroup N of G such that \(N\subseteq {{\,\textrm{Ker}\,}}(\theta )\cap \Phi (G)\), and the quotient \((G/N,\theta _{/N})\), is a \(\theta _{/N}\)-abelian oriented pro-p group. If such a normal subgroup N exists, then \(N=K_\theta (G)\).

2.5 Kummerianity and 1-Cocyles

Let \((G,\theta )\) be an oriented pro-p group. Recall that for \(n\in \mathbb {N}\cup \{\infty \}\), a 1-cocycle \(c:G\rightarrow \mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\) is a continuous map satisfying

$$\begin{aligned} c(gh)=c(g)+{\overline{\theta (g)}}c(h)\qquad \text {for every }g,h\in G, \end{aligned}$$
(2.2)

where \(\overline{\theta (g)}\) denotes the image of \(\theta (g)\) modulo \(p^n\). From (2.2) one deduces

$$\begin{aligned} c([g,h])=\overline{\theta (gh)^{-1}}\left( c(g)(1-\overline{\theta (h)})-c(h)(1-\overline{\theta (g)})\right) . \end{aligned}$$
(2.3)

For \(n\in \mathbb {N}\cup \{\infty \}\), every element of \(\textrm{H}^1(G,\mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta ))\) is represented by a 1-cocycle \(c:G\rightarrow \mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\). The following result is due to J. Labute (cf. [13, Proposition 6]).

Lemma 2.9

Let \((G,\theta )\) be a finitely generated oriented pro-p group with torsion-free orientation, and let \(\mathcal {X}=\{x_1,\ldots ,x_d\}\) be a minimal generating set of G. The following are equivalent.

  1. (i)

    \((G,\theta )\) is Kummerian.

  2. (ii)

    For all \(n\in \mathbb {N}\cup \{\infty \}\) and for any sequence \(\lambda _1,\ldots ,\lambda _d\) of elements of \(\mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\) there exists a continuous 1-cocycle \(G\rightarrow \mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\) satisfying \(c(x_i)=\lambda _i\) for all \(i=1,\ldots ,d\).

Proposition 2.10

Let G be a finitely generated pro-p group, and let \((G,\theta )\) be a Kummerian oriented pro-p group with torsion-free orientation. If N is a normal subgroup of G such that \(N\subseteq {{\,\textrm{Ker}\,}}(\theta )\) and the restriction map

$$\begin{aligned} \textrm{res}_{G,N}^1:\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\longrightarrow \textrm{H}^1(N,\mathbb {Z}/p\mathbb {Z})^G \end{aligned}$$

is surjective, then also \((G/N,\theta _{/N})\) is Kummerian.

In order to prove Proposition 2.10 we need the following fact, whose proof—rather straightforward—is left to the reader.

Fact 2.11

Let G be a finitely generated pro-p group, and let \((G,\theta )\) be an oriented pro-p group with torsion-free orientation.

  1. (i)

    If \(c:G\rightarrow \mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\) is a continuous 1-cocycle, with \(n\in \mathbb {N}\cup \{\infty \}\), then \(c^{-1}(0)\cap {{\,\textrm{Ker}\,}}(\theta )\) is a normal subgroup of G.

  2. (ii)

    Let \(N\subseteq G\) be a normal subgroup satisfying \(N\subseteq {{\,\textrm{Ker}\,}}(\theta )\), with canonical projection \(\pi :G\rightarrow G/N\). For \(n\in \mathbb {N}\cup \{\infty \}\) one has the following:

    1. (a)

      a continuous 1-cocycle \(c:G\rightarrow \mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\) satisfying \(c\vert _N\equiv 0\) induces a continuous 1-cocycle \({\bar{c}} :G/N\rightarrow \mathbb {Z}_p(\theta _{/N})/p^n\mathbb {Z}_p(\theta _{/N})\) such that \(c={\bar{c}}\circ \pi \);

    2. (b)

      a continuous 1-cocycle \({\bar{c}} :G/N\rightarrow \mathbb {Z}_p(\theta _{/N})/p^n\mathbb {Z}_p(\theta _{/N})\) induces a continuous 1-cocycle \(c:G\rightarrow \mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\) satisfying \(c\vert _N\equiv 0\) and \(c={\bar{c}}\circ \pi \).

Proof of Proposition 2.10

Set \({\bar{G}}=G/N\) and \({\bar{\theta }}=\theta _{/N}\). For every \(n\ge 1\), the canonical projection \(\pi :G\rightarrow {\bar{G}}\) induces the inflation maps

$$\begin{aligned} \begin{aligned} f_n&:\textrm{H}^1({\bar{G}},\mathbb {Z}_p({\bar{\theta }} )/p^n\mathbb {Z}_p({\bar{\theta }} ))\longrightarrow \textrm{H}^1(G,\mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )), \\ f&:\textrm{H}^1({\bar{G}},\mathbb {Z}/p\mathbb {Z})\longrightarrow \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z}), \end{aligned}\end{aligned}$$
(2.4)

which are injective by [20, Proposition 1.6.7]. Also, the epimorphisms (respectively, of continuous \({\bar{G}}\)-modules and continuous G-modules) \(\mathbb {Z}_p({\bar{\theta }} )/p^n\mathbb {Z}_p({\bar{\theta }} )\rightarrow \mathbb {Z}/p\mathbb {Z}\) and \(\mathbb {Z}_p(\theta )/p^n\rightarrow \mathbb {Z}/p\mathbb {Z}\) induce, respectively, the morphisms

$$\begin{aligned} \begin{aligned} \tau _n^N&:\textrm{H}^1({\bar{G}},\mathbb {Z}_p(\theta )/p^n)\longrightarrow \textrm{H}^1({\bar{G}},\mathbb {Z}/p),\\ \tau _n&:\textrm{H}^1(G,\mathbb {Z}_p(\theta )/p^n)\longrightarrow \textrm{H}^1(G,\mathbb {Z}/p).\end{aligned}\end{aligned}$$
(2.5)

Altogether, by [20, Proposition 1.5.2] one has the commutative diagram

Since \((G,\theta )\) is Kummerian, \(\tau _n\) is surjective for every \(n\ge 1\). Given \({\bar{\beta }}\in \textrm{H}^1({\bar{G}},\mathbb {Z}/p\mathbb {Z})\), \({\bar{\beta }}\ne 0\), our goal is to find \(\alpha \in \textrm{H}^1({\bar{G}},\mathbb {Z}_p({\bar{\theta }} )/p^n\mathbb {Z}_p({\bar{\theta }} ))\) such that \({\bar{\beta }}=\tau _n^N(\alpha )\).

Set \(\beta ={\bar{\beta }}\circ \pi =f({\bar{\beta }})\). Then \(\beta :G\rightarrow \mathbb {Z}/p\) is a non-trivial continuous homomorphism such that \({{\,\textrm{Ker}\,}}(\beta )\supseteq N\). By hypothesis, the morphism \( N/N^p[G,N]\rightarrow G/\Phi (G)\) induced by the inclusion \(N\hookrightarrow G\), and dual to \(\textrm{res}_{G,N}^1\), is injective. Thus, one may find a minimal generating set \(\mathcal {X}\) of G such that \(\mathcal {Y}=\mathcal {X}\cap N\) generates N as a normal subgroup of G. By Lemma 2.9, there exists a continuous 1-cocycle \(c:G\rightarrow \mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta )\) satisfying

$$\begin{aligned} c(x)\equiv \beta (x)\mod p\mathbb {Z}_p(\theta )\qquad \text {for every }x\in \mathcal {X}\end{aligned}$$

—i.e., \(\tau _n([c])=\beta \), where \([c]\in \textrm{H}^1(G,\mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta ))\) denotes the cohomology class of c—and moreover \(c(x)=0\) for every \(x\in \mathcal {Y}\). Therefore, by Fact 2.11–(i), the restriction

$$\begin{aligned} c\vert _N:N\longrightarrow \mathbb {Z}_p(\theta )/p^n\mathbb {Z}_p(\theta ) \end{aligned}$$

is the map constantly equal to 0. By Fact 2.11–(ii), c induces a continuous 1-cocycle

$$\begin{aligned} {\bar{c}}:{\bar{G}} \longrightarrow \mathbb {Z}_p({\bar{\theta }})/p^n\mathbb {Z}_p({\bar{\theta }}) \end{aligned}$$

such that \({\bar{c}}\circ \pi =c\), and \([c]=f_n([{\bar{c}}])\), where \([{\bar{c}}]\in \textrm{H}^1({\bar{G}},\mathbb {Z}_p({\bar{\theta }})/p^n\mathbb {Z}_p({\bar{\theta }}))\) denotes the cohomology class of \({\bar{c}}\). Altogether, one has

$$\begin{aligned} f({\bar{\beta }})=\beta =\tau _n([c])=\tau _n\circ f_n([\bar{c}])=f\circ \tau _n^N([{\bar{c}}]).\end{aligned}$$

Since f is injective, one obtains \({\bar{\beta }}=\tau _n^N([{\bar{c}}])\). \(\square \)

Remark 2.12

Proposition 2.10 may be proved also in a purely group-theoretic way, see [3, Remark 3.9].

3 The \(\mathbb {Z}/p\mathbb {Z}\)-Cohomology of G

The purpose of this section is to prove the first statement of Proposition 1.3, and more in general to describe the \(\mathbb {Z}/p\mathbb {Z}\)-cohomology algebra \(\textbf{H}^\bullet (G,\mathbb {Z}/p\mathbb {Z})\) with G as in Theorem 1.1.

3.1 Degree 1 and 2

Let G be a pro-p group. We set the subgroup \(G_{(3)}\) of G as follows:

$$\begin{aligned} G_{(3)}={\left\{ \begin{array}{ll} G^p[G,G'] &{} \quad \text {if }p\ne 2,\\ G^4(G')^2[G,G'] &{} \quad \text {if }p=2, \end{array}\right. }\end{aligned}$$

i.e., \(G_{(3)}\) is the third term of the p-Zassenhaus filtration of G (cf., e.g., [23, Sect. 3.1]). In particular, \(G_{(3)}\) is a normal subgroup of the Frattini subgroup \(\Phi (G)\), and the quotient \(\Phi (G)/G_{(3)}\) is a p-elementary abelian pro-p group—and thus also a \(\mathbb {Z}/p\)-vector space.

Recall that the cohomology group \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) is equal to the group of pro-p group homomorphisms from G to \(\mathbb {Z}/p\), namely, one has

$$\begin{aligned} \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})=\textrm{Hom}(G,\mathbb {Z}/p\mathbb {Z})\simeq (G/\Phi (G))^*, \end{aligned}$$
(3.1)

where \(\textvisiblespace ^*\) denotes the \(\mathbb {Z}/p\)-dual (cf., e.g., [32, Chap. I, Sect. 4.2]). Thus, the dimension of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) is equal to the cardinality \(\textrm{d}(G)\) of any minimal generating set of G. On the other hand, the dimension of \(\textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z})\) is equal to the number \(\textrm{r}(G)\) of defining relations of G (cf. [32, Chap. I, Sect. 4.3]). Moreover, if both \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) and \(\textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z})\) are finite, and if the cup-product yields an epimorphism \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})^{\otimes 2}\twoheadrightarrow \textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z})\), one has an isomorphism of elementary abelian p-groups

(3.2)

(cf. [17, Theorem 7.3]). For further properties of the cohomology of pro-p groups we refer to [32, Chap. I, Sect. 4] and to [20, Chap. III, Sect. 9].

3.2 Amalgams

Henceforth, G will denote a pro-p group as in Theorem 1.1. Set

$$\begin{aligned} \begin{aligned} G_1&= \langle \,x,y_1,\ldots ,y_{d_1}\,\mid \,x^{\epsilon p}[x,y_1]\cdots [y_{d_1-1},y_{d_1}]=1\,\rangle ,\\ G_2&= \langle \,x,z_1,\ldots ,z_{d_2}\,\mid \,x^{\epsilon p}[x,z_1]\cdots [z_{d_2-1},z_{d_2}]=1\,\rangle , \end{aligned} \end{aligned}$$

with \(\epsilon =0,1\) depending on whether we are considering case (1.1.a) or (1.1.b). Then \(G_1,G_2\) are Demushkin groups, and G is the amalgamated free pro-p product

$$\begin{aligned} G=G_1\amalg _X^{{\hat{p}}}G_2, \end{aligned}$$
(3.3)

with amalgam the subgroup \(X\subseteq G_1,G_2\) generated by x. Observe that \(X\simeq \mathbb {Z}_p\), as X has infinite index in both \(G_1,G_2\), and subgroups of infinite index of Demushkin groups are free pro-p groups (cf. [32, Chap. I, Sect. 4.5, Example 5–(b)]). Therefore, the amalgamated free pro-p product is proper, i.e., \(G_1,G_2\subseteq G\) (cf. [31]).

3.3 Quadratic Cohomology

Let

$$\begin{aligned} \mathcal {B}=\left\{ \,\chi ,\,\varphi _1,\,\ldots ,\,\varphi _{d_1},\,\psi _1,\,\ldots ,\,\psi _{d_2}\,\right\} \end{aligned}$$

be the basis of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})=\textrm{Hom}(G,\mathbb {Z}/p\mathbb {Z})\) dual to \(\mathcal {X}=\{x,y_1,\ldots ,z_{d_2}\}\)—i.e.,

$$\begin{aligned}\begin{aligned}&\chi (w)={\left\{ \begin{array}{ll} 1&{}\text {\quad if }w=x \\ 0 &{}\quad \text {if }w=y_i,z_j \end{array}\right. }\qquad \text {and}\\&\varphi _i(w)={\left\{ \begin{array}{ll} \delta _{i,i'}&{}\quad \text {if }w=y_{i'} \\ 0 &{}\quad \text {if }w=x,z_j, \end{array}\right. }\qquad \psi _j(w)={\left\{ \begin{array}{ll} \delta _{j,j'}&{}\quad \text {if }w=z_{j'} \\ 0 &{}\quad \text {if }w=x,y_i, \end{array}\right. } \end{aligned}\end{aligned}$$

for every \(1\le i,i'\le d_1\) and \(1\le j,j'\le d_2\) (cf. (3.1)). With an abuse of notation, we may consider the subsets \(\mathcal {B}_1=\{\chi ,\varphi _1,\ldots ,\varphi _{d_1}\}\), \(\mathcal {B}_2=\{\chi ,\psi _1,\ldots ,\psi _{d_2}\}\), and \(\mathcal {B}_X=\{\chi \}\), as bases of \(\textrm{H}^1(G_1,\mathbb {Z}/p\mathbb {Z})\), \(\textrm{H}^1(G_2,\mathbb {Z}/p\mathbb {Z})\), and \(\textrm{H}^1(X,\mathbb {Z}/p\mathbb {Z})\) respectively.

Proposition 3.1

The algebra \(\textbf{H}^\bullet (G,\mathbb {Z}/p\mathbb {Z})\) is quadratic.

Proof

As stated in Sect. 3.2, \(G=G_1\amalg _{X}^{{\hat{p}}}G_2\) is a proper amalgamated free pro-p product. Since \(\mathcal {B}_X\subseteq \mathcal {B}_1,\mathcal {B}_2\), the restriction maps

$$\begin{aligned} {{\,\textrm{res}\,}}^1_{G_i,X}:\textrm{H}^1(G_i,\mathbb {Z}/p\mathbb {Z})\longrightarrow \textrm{H}^1(X,\mathbb {Z}/p\mathbb {Z}),\qquad \text {with }i=1,2, \end{aligned}$$

are surjective. Moreover, \(\textrm{H}^2(X,\mathbb {Z}/p\mathbb {Z})=0\), as \(X\simeq \mathbb {Z}_p\), and thus \({{\,\textrm{Ker}\,}}({{\,\textrm{res}\,}}_{G_i,X}^2)=\textrm{H}^2(G_i,\mathbb {Z}/p\mathbb {Z})\) for both \(i=1,2\). On the other hand, \(\textrm{H}^1(G_1,\mathbb {Z}/p\mathbb {Z})\) and \(\textrm{H}^1(G_2,\mathbb {Z}/p\mathbb {Z})\) are generated by \(\chi \smallsmile \varphi _1\) and \(\chi \smallsmile \psi _1\) respectively, as \(G_1,G_2\) are Demushkin groups (cf., e.g., [20, Proposition 3.9.16]), and thus

$$\begin{aligned}{} & {} {{\,\textrm{Ker}\,}}({{\,\textrm{res}\,}}_{G_i,X}^2) \\{} & {} =\textrm{H}^2(G_i,\mathbb {Z}/p\mathbb {Z})={{\,\textrm{Ker}\,}}({{\,\textrm{res}\,}}_{G_i,X}^1)\smallsmile \textrm{H}^1(G_i,\mathbb {Z}/p\mathbb {Z}),\qquad \text {with }i=1,2, \end{aligned}$$

as \({{\,\textrm{res}\,}}_{G_1,X}^1(\varphi _1)=0\) and \({{\,\textrm{res}\,}}_{G_2,X}^1(\psi _1)=0\). Finally, Demushkin groups are well-known to yield a quadratic \(\mathbb {Z}/p\mathbb {Z}\)-cohomology algebra, while \(\textbf{H}^\bullet (X,\mathbb {Z}/p\mathbb {Z})\) is obviously quadratic, as \(X\simeq \mathbb {Z}_p\). Therefore, we may apply [28, Theorem B], so that also \(\textbf{H}^\bullet (G,\mathbb {Z}/p\mathbb {Z})\) is quadratic. \(\square \)

We describe now more in detail the structure of \(\textbf{H}^\bullet (X,\mathbb {Z}/p\mathbb {Z})\). By duality—cf. [17, Theorem 7.3] and (3.2)—the set \(\{\chi \smallsmile \varphi _1,\chi \smallsmile \psi _1\}\) is a basis of \(\textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z})\), and in \(\textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z})\) one has the relations

$$\begin{aligned} \chi \smallsmile \varphi _{i'}=\chi \smallsmile \psi _{j'}=\varphi _i\smallsmile \psi _j=0 \end{aligned}$$
(3.4)

for all \(1\le i,i'\le d_1\) and \(1\le j,j'\le d_2\), with \(i',j'\ne 1\), and

$$\begin{aligned} \begin{aligned} \varphi _{i}\smallsmile \varphi _{i'}={\left\{ \begin{array}{ll} (-1)^\epsilon \chi \smallsmile \varphi _1 &{} \quad \text {if }2\mid i=i'-1, \\ 0&{}\quad \text {otherwise}, \end{array}\right. }\\ \qquad \psi _{j}\smallsmile \psi _{j'}={\left\{ \begin{array}{ll} (-1)^\epsilon \chi \smallsmile \psi _1 &{} \quad \text {if }2\mid j=j'-1, \\ 0&{}\quad \text {otherwise} \end{array}\right. } \end{aligned}\end{aligned}$$
(3.5)

(see also [23, Sect. 3.2]).

Finally, one has an exact sequence

figure a

(cf. [28, p. 653]). Since \(\textrm{H}^2(X,\mathbb {Z}/p\mathbb {Z})=\textrm{H}^3(G_i,\mathbb {Z}/p\mathbb {Z})=0\) for both \(i=1,2\), one has \(\textrm{H}^3(G,\mathbb {Z}/p\mathbb {Z})=0\), and thus by quadraticity also \(\textrm{H}^n(G,\mathbb {Z}/p\mathbb {Z})=0\) for all \(n\ge 3\).

Remark 3.2

It is well-known that if a pro-p group has non-trivial torsion, then its nth \(\mathbb {Z}/p\)-cohomology group is non-trivial for every \(n>0\); hence, G is torsion-free.

4 Proof of Theorem 1.1 Case (1.1.a)

Let G be a pro-p group as defined in Theorem 1.1, with defining relations as in (1.1.a)—namely,

$$\begin{aligned} G=\langle \,x,y_1,\ldots ,y_{d_1},z_1,\ldots ,z_{d_2}\,\mid \,r_1=r_2=1\,\rangle ,\end{aligned}$$

with \(d_1+d_2\ge 4\) and

$$\begin{aligned}\begin{aligned} r_1&= [x,y_1]\cdots [y_{d_1-1},y_{d_1}],\\r_2&=[x,z_1]\cdots [z_{d_2-1},z_{d_2}]. \end{aligned}\end{aligned}$$

Without loss of generality, we may assume that \(d_1\ge 3\).

4.1 Kummerianity

Let \(G_1,G_2\) be the two Demushkin groups as in Sect. 3.2, with \(\epsilon =0\). By Example 2.6, if

$$\begin{aligned} \theta _1:G_1\longrightarrow 1+p\mathbb {Z}_p\qquad \text {and} \qquad \theta _2:G_2\longrightarrow 1+p\mathbb {Z}_p \end{aligned}$$

are two torsion-free orientations completing respectively \(G_1\) and \(G_2\) into Kummerian oriented pro-p groups, then necessarily \(\theta _1(x)=\theta _1(y_1)=\cdots =\theta _1(y_{d_1})=1\), and analogously \(\theta _2(x)=\theta _2(z_1)=\cdots =\theta _1(z_{d_2})=1\).

Proposition 4.1

Let \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) be a torsion-free orientation. Then the oriented pro-p group \((G,\theta )\) is Kummerian if, and only if, \(\theta \) is constantly equal to 1.

Proof

If \(\theta \equiv \textbf{1}\), then \((G,\textbf{1})\) is Kummerian if, and only if, the abelianization \(G^{{{\,\textrm{ab}\,}}}\) is a free abelian pro-p group. But this is easily verified, as clearly \(G^{{{\,\textrm{ab}\,}}}\simeq \mathbb {Z}_p^{d_1+d_2-1}\).

Conversely, suppose that \((G,\theta )\) is Kummerian. Let \(N_1\) and \(N_2\) denote the normal subgroups of G generated as normal subgroups by \(z_1,\ldots ,z_{d_2}\) and \(y_1,\ldots ,y_{d_1}\) respectively. Then \(G/N_1\simeq G_1\) and \(G/N_2\simeq G_2\). Moreover, Proposition 2.10 implies that \((G/N_i,\theta _{/N_i})\) is Kummerian for both \(i=1,2\). Since \(G/N_i\simeq G_i\) for both i, Example 2.6 and the argument before the statement of the proposition imply that the torsion-free orientations \(\theta _{/N_1}\) and \(\theta _{/N_2}\) are constantly equal to 1. Hence, also \(\theta \) is constantly equal to 1, as \(\theta (w)=\theta _{/N_1}(wN_1)\) for every \(w\in G_1\), and analogously \(\theta (w)=\theta _{/N_2}(wN_2)\) for every \(w\in G_2\). \(\square \)

Therefore, if G may complete into a 1-cyclotomic oriented pro-p group, then necessarily G is absolutely torsion-free. In order to prove Theorem 1.1 in case (1.1.a), we aim at exhibiting an open subgroup H of G, of index \(p^2\), whose abelianization \(H^{{{\,\textrm{ab}\,}}}\) has non-trivial torsion.

4.2 The Subgroup U

Set \(u=y_3^p\), \(t_0=z_1^{-1}y_3\), and \(t_h=t_0t_0^{y_3}\cdots t_0^{y_3^h}\) for all \(h=0,\ldots ,p-1\). A straightforward computation shows that

$$\begin{aligned} z_1^{h}=y_3^h\cdot (t_0^{-1})^{y_3^{h-1}}\cdots (t_0^{-1})^{y_3}\cdot t_0^{-1}=y_3^ht_{h-1}^{-1} \end{aligned}$$
(4.1)

for all \(h=0,\ldots ,p-1\).

Let \(\phi _G:G\rightarrow \mathbb {Z}/p\) be the homomorphism of pro-p groups defined by \(\phi _G(y_3)=\phi _G(z_1)=1\) and \(\phi _G(x)=\phi _G(y_i)=\phi _G(z_j)=0\) for all \(i=1,2,4,\ldots ,d_1\) and \(j=2,\ldots ,d_2\), and set \(U={{\,\textrm{Ker}\,}}(\phi )\). Then U is an open subgroup of G of index p, generated as a normal subgroup by the subset

$$\begin{aligned} \mathcal {X}=\left\{ \, u,\, x,\, t_0,\, y_i,\,z_j\,\mid \, i=1,2,4,\ldots ,d_1,\,j=2,\ldots ,d_{2} \right\} , \end{aligned}$$

and \(G/U=\{U,y_3U,\ldots ,y_3^{p-1}U\}\).

Lemma 4.2

The subset

$$\begin{aligned} \mathcal {Y}_U=\left\{ \,u,\,x,\,y_2,\,t_h,\,y_i^{y_3^h},\,z_j^{y_3^h}\,\mid \,i=1,4,\ldots ,d_1,\,j=2,\ldots ,d_{2},\,h=0,\ldots ,p-1\,\right\} \end{aligned}$$

of U is a minimal generating set of U as a pro-p group.

Proof

Since U is normally generated by \(\mathcal {X}\) and \(G/U=\{U,\ldots ,y_3^{p-1}U\}\), U is generated as a pro-p group by the set \(\{w^{y_3^h}\mid w\in \mathcal {X},h=0,\ldots ,p-1 \}\). Also, U is subject to the relations

$$\begin{aligned} r_1^{y_3^h}= & {} \left[ x^{y_3^h},y_1^{y_3^h}\right] \cdots \left[ y_{d_1-1}^{y_3^h},y_{d_1}^{y_3^h}\right] =1, \end{aligned}$$
(4.2)
$$\begin{aligned} r_2^{y_3^h}= & {} \left[ x^{y_3^h},z_1^{y_3^h}\right] \cdots \left[ z_{d_2-1}^{y_3^h},z_{d_2}^{y_3^h}\right] =1, \end{aligned}$$
(4.3)

with \(h=0,\ldots ,p-1\).

Consider the abelianization \(U^{{{\,\textrm{ab}\,}}}\). Since the only factor in (4.2) which does not lie in \(U'\) is \([y_2^{y_3^h},y_3]\), the relation (4.2) implies that \([y_2^{y_3^h},y_3]\in U'\) as well, and therefore

$$\begin{aligned} y_2^{y_3^h}\equiv y_2\mod U'\qquad \text {for all }h=0,\ldots ,p-1. \end{aligned}$$

Analogously, the only factor in (4.3) which does not lie in \(U'\) is \([x^{y_3^h},z_1^{y_3^h}]\), so that the relation (4.2) implies that \([x^{y_3^h},z_1^{y_3^h}]\in U'\) as well. Hence, one has

$$\begin{aligned}\begin{aligned}{}[x,z_1]\equiv 1\bmod U'\;&\Rightarrow \; x^{y_3t_0^{-1}}\equiv x\bmod U'\\&\Rightarrow x^{y_3}\equiv x^{t_0}\bmod U',\\ \left[ x^{y_3},z_1^{y_3}\right] \equiv 1\bmod U' \;&\Rightarrow \; (x^{y_3})^{(z_1^{y_3})}=x^{y_3^2(t_0^{-1})^{y_3}}\equiv x^{y_3}\bmod U'\\&\Rightarrow \; x^{y_3^2}\equiv x^{t_1}\bmod U', \end{aligned}\end{aligned}$$

and so on. Thus,

$$\begin{aligned} x^{y_3^h}\equiv x^{t_{h-1}}\mod U'\qquad \text {for all }h=1,\ldots ,p-1. \end{aligned}$$

Altogether, \(U^{{{\,\textrm{ab}\,}}}\) is the free abelian pro-p group generated by the cosets \(\{wU'\mid w\in \mathcal {Y}_U\}\), so that Fact 2.1 yields the claim. \(\square \)

Now set \(U_1=G_1\cap U\) and \(U_2=G_2\cap U\). Then \(U_1,U_2\) are open subgroups of \(G_1,G_2\), respectively, of index p, and thus they are again Demushkin groups, on \(2+p(d_1-1)\) and \(2+p(d_2-1)\) generators respectively (cf. [6]). In particular, the defining relation of \(U_1\) is

$$\begin{aligned} s_1= \prod _{h=p-1}^0\left( \left[ y_{4}^{y_3^h},y_{5}^{y_3^h}\right] \cdots \left[ y_{d_1-1}^{y_3^h},y_{d_1}^{y_3^h}\right] \left[ x^{y_3^h},y_{1}^{y_3^h}\right] \right) [y_2,u]=1, \end{aligned}$$
(4.4)

while the defining relation of \(U_2\) is

$$\begin{aligned} \begin{aligned} s_2&=\prod _{h=p-1}^0\left( \left[ z_{2}^{z_1^h},z_{3}^{z_1^h}\right] \cdots \left[ z_{d_2-1}^{z_1^h},z_{d_2}^{z_1^h}\right] \right) [x,z_1^p] \\&=\prod _{h=p-1}^0\left( \left[ z_{2}^{y_3^ht_{h-1}^{-1}},z_{3}^{y_3^ht_{h-1}^{-1}}\right] \cdots \left[ z_{d_2-1}^{y_3^ht_{h-1}^{-1}},z_{d_2}^{y_3^ht_{h-1}^{-1}}\right] \right) [x,ut_{p-1}^{-1}] =1. \end{aligned} \end{aligned}$$
(4.5)

Also, from the relations (4.4)–(4.5) and from (4.1), one computes

$$\begin{aligned} \begin{aligned} x^{y_3}&=x^{z_1t_0}=x^{t_0}([z_{d_2},z_{d_2-1}]\cdots [z_3,z_2])^{t_0},\\ x^{y_3^2}&= x^{t_1}([z_{d_2},z_{d_2-1}]\cdots )^{t_1} \left( \left[ z_{d_2}^{y_3},z_{d_2-1}^{y_3}\right] \cdots \right) ^{t_0^{-1}t_1},\\ x^{y_3^3}&= x^{t_2}([z_{d_2},z_{d_2-1}]\cdots )^{t_2} \left( \left[ z_{d_2}^{y_3},z_{d_2-1}^{y_3}\right] \cdots \right) ^{t_0^{-1}t_2} \left( \left[ z_{d_2}^{y_3^2},z_{d_2-1}^{y_3^2}\right] \cdots \right) ^{t_1^{-1}t_2}, \end{aligned} \end{aligned}$$
(4.6)

and so on. In fact, the two relations (4.4)–(4.5)—with the \(x^{y_3^h}\)’s replaced using (4.6)—are all the defining relations we need to get U, as shown in the following.

Lemma 4.3

The pro-p group U has \(\textrm{r}(U)=2\) defining relations.

Proof

Since \(\textrm{H}^n(G,\mathbb {Z}/p\mathbb {Z})=0\) for every \(n\ge 3\) (cf. Sect. 3.3) and \([G:U]=p\), one has \(\textrm{H}^n(U,\mathbb {Z}/p\mathbb {Z})=0\) for every \(n\ge 3\) as well (cf. [20, Proposition 3.3.5]). Moreover, one has

$$\begin{aligned} \textrm{r}(U)-\textrm{d}(U)+1=p\left( \textrm{r}(G)-\textrm{d}(G)+1\right) \end{aligned}$$
(4.7)

(cf. [20, Proposition 3.3.13]). By definition, \(\textrm{r}(G)=2\) and \(\textrm{d}(G)=1+d_1+d_2\), while \(\textrm{d}(U)=3+p(d_1+d_2-2)\) by Lemma 4.2. Therefore, from (4.7) one computes \(\textrm{r}(U)=2\). \(\square \)

4.3 The Subgroup H

Let \(\phi _U:U\rightarrow \mathbb {Z}/p \) be the homomorphism of pro-p groups defined by \(\phi _U(y_1)\), \(\phi _U(y_1^{y_3})=-1\), and \(\phi _U(w)=0\) for any other element w of \(\mathcal {Y}_U\), and put \(H={{\,\textrm{Ker}\,}}(\phi _U)\). Then H is an open subgroup of U of index p. Set \(v=y_1\). Since \(U/H=\{H,vH,\ldots ,v^{p-1}H\}\), H is the pro-p group (non-minimally) generated by

$$\begin{aligned} \mathcal {X}_{H}=\left\{ \,v^p,\,\left( vy_1^{y_3}\right) ^{v^h},\,w^{v^h}\,\mid w\in \mathcal {Y}_U,\,w\ne v,y_1^{y_3},\, h=0,\ldots ,p-1\,\,\right\} , \end{aligned}$$

and subject to the 2p relations \(s_1^{v^h}=1\) and \(s_2^{v^h}=1\), with \(h=0,\ldots ,p-1\). We claim that the abelianization \(H^{{{\,\textrm{ab}\,}}}\) yields non-trivial torsion.

Proposition 4.4

The abelian pro-p group \(H^{{{\,\textrm{ab}\,}}}\) is not torsion-free.

Proof

Since all the elements of \(\mathcal {Y}_U\) showing up in the last terms of the equalities (4.6) belong to H, one deduces that \(x^{y_3^h}\equiv x\bmod H'\) for all \(h=0,\ldots ,p-1\).

Now, each factor of \(s_2\)—cf. (4.5)—is a commutator of elements of H, and thus the relations \(s_2^{v^h}=1\) yield trivial relations in \(H^{{{\,\textrm{ab}\,}}}\). On the other hand, every factor of \(s_1\)—cf. (4.4)—but \([x,y_1]\) and \([x^{y_3},y_1^{y_3}]\), is a commutator of elements of H. From (4.4) one obtains

$$\begin{aligned} \left[ x^{y_3},y_1^{y_3}\right] [x,y_1]\equiv \left[ x,v^{-1}(vy_1^{y_3})\right] [x,v]\equiv [x,v^{-1}][x,v]\equiv 1\mod H', \end{aligned}$$
(4.8)

as \(vy_1^{y_3}\in H\). Altogether, \(H^{{{\,\textrm{ab}\,}}}\) is the abelian pro-p group (non-minimally) generated by the set \(\mathcal {X}_{H^{{{\,\textrm{ab}\,}}}}=\{wH'\,\mid \,w\in \mathcal {X}_H\}\), and subject to the p relations

$$\begin{aligned} \left[ x^{v^h}H',v^{-1}H'\right] \left[ x^{v^h}H',vH'\right] =H',\qquad \text {with }h=0,\ldots ,p-1, \end{aligned}$$

as \(U/H=\{H,vH,\ldots ,v^{p-1}H\}\). From these relations, one deduces the equivalences:

$$\begin{aligned} \begin{aligned} x^{v^2}&\equiv \left( x^v\right) ^2\cdot x^{-1}\mod H'\qquad \text {with }h=1,\\ x^{v^3}&\equiv \left( x^{v^2}\right) ^2\cdot \left( x^v\right) ^{-1}\equiv \left( x^v\right) ^3\cdot x^{-2} \mod H'\qquad \text {with }h=2,\\&\quad \vdots \\ x^{v^{p-1}}&\equiv \left( x^{v^{p-2}}\right) ^2\cdot \left( x^{v^{p-3}}\right) ^{-1}\equiv \left( x^v\right) ^{p-1}\cdot x^{2-p} \mod H'\qquad \text {with }h=p-2,\\ x^{v^{p}}&\equiv \left( x^{v^{p-1}}\right) ^2\cdot \left( x^{v^{p-2}}\right) ^{-1}\equiv \left( x^v\right) ^{p}\cdot x^{1-p} \mod H'\qquad \text {with }h=p-1.\\ \end{aligned} \end{aligned}$$

But \(x^{v^p}\equiv x\bmod H'\), as \(v^p\in H\), and thus from the last of the above equivalences one obtains

$$\begin{aligned} x\equiv (x^v)^px^{1-p}\mod H'\;\Longrightarrow \;(x^v)^px^{-p}\equiv (x^vx^{-1})^p\equiv 1\mod H'. \end{aligned}$$
(4.9)

Altogether, \(H^{{{\,\textrm{ab}\,}}}\) is the abelian pro-p group minimally generated by

$$\begin{aligned} \mathcal {Y}_{H^{{{\,\textrm{ab}\,}}}}=\left\{ \,v^pH',\, xH', \, x^vH',\,\left( vy_1^{y_3}\right) ^{v^h}H',\,w^{v^h}H'\,\mid \, h=0,\ldots ,p-1\,\,\right\} , \end{aligned}$$

where \(w\in \mathcal {Y}_U\smallsetminus \{v,y_1^{y_3},x\}\), and subject to the relation \(((xH')^{-1}\cdot x^vH')^p=H'\)—in particular, \(H^{{{\,\textrm{ab}\,}}}\) is isomorphic to \(\mathbb {Z}_p^{2+p+p^2(d_1+d_2-2)}\times \mathbb {Z}/p\mathbb {Z}\). \(\square \)

5 Proof of Theorem 1.1 Case (1.1.b)

Let p be an odd prime, and let G be a pro-p group as defined in Theorem 1.1, with defining relations as in (1.1.b)—namely,

$$\begin{aligned} G=\langle \,x,y_1,\ldots ,y_{d_1},z_1,\ldots ,z_{d_2}\,\mid \,r_1=r_2=1\,\rangle ,\end{aligned}$$

with

$$\begin{aligned}\begin{aligned} r_1&= y_1^p[y_1,x]\cdots [y_{d_1-1},y_{d_1}],\\r_2&=z_1^p[z_1,x]\cdots [z_{d_2-1},z_{d_2}]. \end{aligned}\end{aligned}$$

5.1 Kummerianity

Let \(G_1,G_2\) be the two Demushkin groups as in Sect. 3.2, with \(\epsilon =1\). By Example 2.6, if

$$\begin{aligned} \theta _1:G_1\longrightarrow 1+p\mathbb {Z}_p\qquad \text {and} \qquad \theta _2:G_2\longrightarrow 1+p\mathbb {Z}_p \end{aligned}$$

are two torsion-free orientations completing respectively \(G_1\) and \(G_2\) into Kummerian oriented pro-p groups, then necessarily \(\theta _1(y_1)=\cdots =\theta _1(y_{d_1})=1\), and analogously \(\theta _2(z_1)=\cdots =\theta _1(z_{d_2})=1\), while \(\theta _1(x)=\theta _2(x)=(1-p)^{-1}\).

Proposition 5.1

An orientation \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) completes G into a Kummerian oriented pro-p group \((G,\theta )\) if, and only if,

$$\begin{aligned} \theta (x)=(1-p)^{-1}\qquad \text {and}\qquad \theta (y_i)=\theta (z_j)=1 \end{aligned}$$

for all \(i=1,\ldots ,d_1\) and \(j=1,\ldots ,d_2\).

Proof

Suppose that \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) is the orientation defined as above, and pick arbitrary p-adic integers \(\lambda ,\lambda _i,\lambda '_j\in \mathbb {Z}_p\) for \(1\le i\le d_1\) and \(1\le j\le d_2\). The assignment \(x\mapsto \lambda \), \(y_i\mapsto \lambda _i\) and \(z_j\mapsto \lambda '_j\) for every ij yields a well-defined continuous 1-cocycle \(c:G\rightarrow \mathbb {Z}_p(\theta )\), as (2.3) implies that

$$\begin{aligned}\begin{aligned} c(r_1)&= c(y_1^p)+c([y_1,x])+c([y_2,y_3])+\cdots +c([y_{d_1-1},y_{d_1}])\\&= p\cdot \lambda _1+ \theta (x)^{-1}(\lambda _1(1-\theta (x))-0)+0+\cdots +0 \\ {}&=0 \end{aligned}\end{aligned}$$

and

$$\begin{aligned}\begin{aligned} c(r_2)&= c(z_1^p)+c([z_1,x])+c([z_2,z_3])+\cdots +c([z_{d_2-1},z_{d_2}])\\&= p\cdot \lambda '_1+ \theta (x)^{-1}(\lambda '_1(1-\theta (x))-0)+0+\cdots +0 \\ {}&=0. \end{aligned}\end{aligned}$$

Therefore, \((G,\theta )\) is Kummerian by Lemma 2.9.

Conversely, suppose that \((G,\theta )\) is Kummerian. Let \(N_1\) and \(N_2\) denote the normal subgroups of G generated as normal subgroups by \(z_1,\ldots ,z_{d_2}\) and \(y_1,\ldots ,y_{d_1}\), respectively. Then \(G/N_1\simeq G_1\) and \(G/N_2\simeq G_2\). Moreover, Proposition 2.10 implies that \((G/N_i,\theta _{/N_i})\) is Kummerian for both \(i=1,2\).

Since \(G/N_i\simeq G_i\) for both i, Example 2.6 and the argument before the statement of the proposition imply that \(\theta _{/N_1}(y_1N_1)=\cdots =\theta _{/N_1}(y_{d_1}N_1)=1\), and analogously \(\theta _{/N_2}(z_1N_2)=\cdots =\theta _{/N_2}(z_{d_2}N_2)=1\), while \(\theta _{/N_1}(xN_1)=\theta _{/N_2}(xN_2)=(1-p)^{-1}\). Hence, \(\theta \) is as defined above, as \(\theta (w)=\theta _{/N_1}(wN_1)\) for every \(w\in G_1\), and analogously \(\theta (w)=\theta _{/N_2}(wN_2)\) for every \(w\in G_2\). \(\square \)

Henceforth, \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) will denote the orientation as in Proposition 5.1.

5.2 The Subgroup H

Let \(\phi _1:G_1\rightarrow \mathbb {Z}/p\oplus \mathbb {Z}/p\) and \(\phi _2:G_2\rightarrow \mathbb {Z}/p\oplus \mathbb {Z}/p\) be the homomorphisms of pro-p groups defined by

$$\begin{aligned} \begin{aligned}&\phi _1(x)=\phi _2(x)=(1,0),\\ {}&\phi _1(y_1)=\phi _2(z_1)=(0,1),\\ {}&\phi _1(y_i)=\phi _2(z_j)=(0,0)\text { for }i,j\ge 2. \end{aligned} \end{aligned}$$
(5.1)

Put \(U_1={{\,\textrm{Ker}\,}}(\phi _1)\) and \(U_2={{\,\textrm{Ker}\,}}(\phi _2)\), and also

$$\begin{aligned} t=z_1^{-1}y_1,\qquad u=x^p,\qquad v=y_1^p,\qquad w=z_1^p. \end{aligned}$$

Then \(U_1\) is an open normal subgroup of \(G_1\) of index \(p^2\), and likewise for \(U_2\) and \(G_2\)—note that by [6] both \(U_1\) and \(U_2\) are Demushkin groups.

Finally, put \(N_1={{\,\textrm{Ker}\,}}(\theta \vert _{U_1})\), \(N_2={{\,\textrm{Ker}\,}}(\theta \vert _{U_2})\), and let T be the subgroup of G generated by t. Observe that \(N_1\) and \(N_2\) are free pro-p groups, as they are subgroups of infinite index of Demushkin groups (cf. [32, Chap. I, Sect. 4.5, Example 5–(b)]), while \(T\simeq \mathbb {Z}_p\) as G is torsion-free (cf. Remark 3.2).

Let H be the subgroup of G generated by \(U_1\), \(U_2\) and T, and let M be the subgroup of H generated by \(N_1\), \(N_2\) and T. Observe that \(M\subseteq {{\,\textrm{Ker}\,}}(\theta )\). Our aim is to show that the oriented pro-p group \((H,\theta \vert _H)\) is not Kummerian. For this purpose, we need the following.

Lemma 5.2

  1. (i)

    \(M=N_1\amalg N_2\amalg T\).

  2. (ii)

    M is a normal subgroup of H, and \(H\simeq M\rtimes X^p \)

  3. (iii)

    One has an isomorphism of p-elementary abelian groups

    $$\begin{aligned} \frac{G}{\Phi (G)}\simeq \frac{X^p}{X^{p^2}}\times \frac{N_1}{N_1^p[N_1,U_1]}\times \frac{N_2}{N_2^p[N_2,U_2]}\times \frac{T}{T^p}. \end{aligned}$$
    (5.2)

Proof

Consider the pro-p tree \(\mathcal {T}\) associated to the amalgamated free pro-p product (3.3). Namely, \(\mathcal {T}\) consists of a set vertices \(\mathcal {V}\) and a set of edges \(\mathcal {E}\), where

$$\begin{aligned}\begin{aligned} \mathcal {V}&=\{\,hG_1,hG_2\,\mid \,h\in G\,\}=G/G_1\,\dot{\cup }\, G/G_2,\\ \mathcal {E}&=\{\,hX\,\mid \,h\in G\,\}=G/X, \end{aligned}\end{aligned}$$

and it comes endowed with a natural G-action, i.e.,

$$\begin{aligned} \begin{aligned} g.(hG_1)=(gh)G_1\qquad&\text {for every } g\in G,\, hG_1\in G/G_1\subseteq \mathcal {V}\\ g.(hG_1)=(gh)G_2 \qquad&\text {for every }g\in G,\,hG_2\in G/G_2\subseteq \mathcal {V},\\ g.(hX)=(gh)X\qquad&\text {for every }g\in G,\,hX\in G/X=\mathcal {E}. \end{aligned}\end{aligned}$$
(5.3)

Pick \(g\in M\) and \(hX\in \mathcal {E}\). Then \(g.hX=hX\) if, and only if, \(g\in hXh^{-1}\), i.e., \(g=hx^\lambda h^{-1}\) for some \(\lambda \in \mathbb {Z}_p\). Since \(M\subseteq {{\,\textrm{Ker}\,}}(\theta )\), it follows that

$$\begin{aligned} 1=\theta (g)=\theta \left( hx^\lambda h^{-1}\right) =\theta (x)^\lambda =(1-p)^{\lambda }, \end{aligned}$$
(5.4)

and therefore \(\lambda =0\), as \(1+p\mathbb {Z}_p\) is torsion-free. Hence, the subgroup M intersects trivially the stabilizer \(\textrm{Stab}_{G}(hX)\) of every edge \(hX\in \mathcal {E}\). By [14, Theorem 5.6], M decomposes as free pro-p product as follows:

$$\begin{aligned} M=\left( \coprod _{\omega \in \mathcal {V}'}\textrm{Stab}_{M}(\omega )\right) \amalg F, \end{aligned}$$
(5.5)

where F is a free pro-p group, and \(\mathcal {V}'\subseteq \mathcal {V}\) is a continuous set of representatives of the space of orbits \(M\backslash \mathcal {V}\). Clearly, the vertices \(G_1\) and \(G_2\) belong to different orbits, thus in the decomposition (5.5) one finds the two factors

$$\begin{aligned} \begin{aligned} \textrm{Stab}_M(G_1)=\{\,g\in M\,\mid \,gG_1=G_1\,\}=M\cap G_1, \\ \textrm{Stab}_M(G_2)=\{\,g\in M\,\mid \,gG_2=G_2\,\}=M\cap G_2. \end{aligned}\end{aligned}$$

Since \(N_1\subseteq M\cap G_1\subseteq {{\,\textrm{Ker}\,}}(\theta )\cap G_1=N_1\), one has \(\textrm{Stab}_M(G_1)=N_1\), and analogously \(\textrm{Stab}_M(G_2)=N_2\). Therefore, from (5.5) one obtains

$$\begin{aligned} M=N_1\amalg N_2\amalg \left( \coprod _{\omega \in \mathcal {V}'\smallsetminus \{G_1,G_2\}}\textrm{Stab}_{M}(\omega )\amalg F\right) . \end{aligned}$$
(5.6)

It is straightforward to see that \(t\notin N_1\amalg N_2\). Since M is generated as pro-p group by \(N_1\), \(N_2\) and t, the right-side factor in (5.6) is necessarily T, and this proves (i).

To prove (ii), we need only to show that \(uMu^{-1}=M\), as \(H=\langle \,u,M\,\rangle \). Since \(N_1\) is normal in \(U_1\), and \(u\in U_1\), then \(uN_1u^{-1}=N_1\)—analogously, \(uN_2u^{-1}=N_2\). Now, observe that the integer

$$\begin{aligned} (1-p)^p-1=\left( 1-\left( {\begin{array}{c}p\\ 1\end{array}}\right) p+\left( {\begin{array}{c}p\\ 2\end{array}}\right) p^2-\cdots -p^p\right) -1 \end{aligned}$$

is divisible by \(p^2\) but not by \(p^3\), so we put \((1-p)^p=1+p^2\lambda \), with \(\lambda \in 1+p\mathbb {Z}_p\). From the relation \(r_1=1\) one deduces

$$\begin{aligned} y_1^{x}=y_1^{1-p}\cdot \left( [y_2,y_3]\cdots [y_{d_1-1},y_{d_1}]\right) ^{-1}, \end{aligned}$$
(5.7)

and by iterating (5.7) p times, one obtains \( y_1^{u}=y_1^{(1-p)^p} n_1\) for some \(n_1\in N_1'\)—for this purpose, observe that for every \(\nu \ge 0\) and \(i\ge 1\), the triple commutator

$$\begin{aligned} \left[ y_1^\nu ,[y_i,y_{i+1}]\right] =\left[ y_i^{y_1^\nu },y_{i+1}^{y_1^\nu }\right] ^{-1}\cdot [y_i,y_{i+1}] \end{aligned}$$

belongs to \(N_1'\), as \(y_i^{y_0^\nu }\in N_1\). Analogously, \( z_1^{u}=z_1^{(1-p)^p}n_2\) for some \(n_2\in N_2'\). Altogether,

$$\begin{aligned} t^{u}=(z_1^{-1}y_1)^{u}=z_1^{u}y_1^{u}= n_2^{-1}\cdot w^{-p\lambda }\cdot t\cdot v^{p\lambda }\cdot n_1, \end{aligned}$$
(5.8)

which belongs to M—here we replaced \(z_1^{-(1-p)^p}=w^{-p\lambda }\cdot z_1^{-1}\) and \(y_1^{(1-p)^p}=y_1\cdot v^{p^\lambda }\). Hence, \(M\unlhd H\). Finally, by definition \(H=M\cdot X^p\), and moreover

$$\begin{aligned} M\cap X^p\subseteq {{\,\textrm{Ker}\,}}(\theta )\cap X^p=\{1\}, \end{aligned}$$

so that \(H=M\rtimes X^p\). This completes the proof of (ii).

Finally, by (i) and (ii) one has the isomorphism of p-elementary abelian groups

$$\begin{aligned} \begin{aligned} M/\Phi (M)&\simeq N_1/\Phi (N_1)\times N_2/\Phi (N_2)\times T/T^p \\ H/\Phi (H)&\simeq X^p/X^{p^2}\times M/M^p[M,H]. \end{aligned} \end{aligned}$$
(5.9)

From (5.8) one has that \([T,X^p]\subseteq \Phi (M)\), and since \(H=MX^p\), \(U_1=N_1X^p\), and \(U_2=N_2X^p\), form (5.9) one deduces (iii). \(\square \)

5.3 The Subgroup H and Kummerianity

Proposition 5.3

The oriented pro-p group \((H,\theta \vert _H)\) is not Kummerian.

Proof

Let N be the normal subgroup of H generated as a normal subgroup by \(N_1,N_2\), and set \({\bar{H}}=H/N\). Then \(N\subseteq {{\,\textrm{Ker}\,}}(\theta \vert _H)\), and clearly \({\bar{H}}\) is finitely generated. Moreover, by duality the restriction map \(\textrm{res}_{H,N}^1:H^1(H,\mathbb {Z}/p\mathbb {Z})\rightarrow H^1(N,\mathbb {Z}/p\mathbb {Z})^H\) is surjective, as by Lemma 5.2 one has

$$\begin{aligned} N/N^p[N,H]\simeq N_1/N_1^p[N_1,U_1]\times N_2/N_2^p[N_2,U_2], \end{aligned}$$

which embeds in \(H/\Phi (H)\). In particular, \(\{uN,tN\}\) is a minimal generating set of \({\bar{H}}\). Thus, by Proposition 2.10 if the oriented pro-p group \(({\bar{H}},{\bar{\theta }})\) is not Kummerian—where \({\bar{\theta }}=(\theta \vert _H)_{/N}:{\bar{H}}\rightarrow 1+p\mathbb {Z}_p\) is the orientation induced by \(\theta \vert _H\)—then also \((H,\theta \vert _H)\) is not Kummerian.

By (5.8), in H one has that \([t,u^{-1}]\equiv 1\bmod N\), and thus \({\bar{H}}\) is abelian. Moreover,

$$\begin{aligned} {\bar{\theta }}(uN)=\theta (u)=(1-p)^p \qquad \text {and}\qquad {\bar{\theta }}(tN)=\theta (t)=1, \end{aligned}$$

so that \({{\,\textrm{Ker}\,}}({\bar{\theta }})=\langle tN\rangle \). Therefore, the subgroup \(K_{{\bar{\theta }}}({\bar{H}})\) is generated by

$$\begin{aligned} \left( t^{-\theta (u)}utu^{-1}\right) N=t^{p^2\lambda }N. \end{aligned}$$

Thus, the quotient \({{\,\textrm{Ker}\,}}({\bar{\theta }})/K_{{\bar{\theta }}}({\bar{H}})=\langle tN\rangle /\langle tN\rangle ^{p^2}\) is not torsion-free, and by Proposition 2.2, \(({\bar{H}},{\bar{\theta }})\) is not Kummerian. \(\square \)

This completes the proof of Theorem 1.1 case (1.1.b).

Remark 5.4

If \(d_1=d_2=1\), case (1.1.b) of Theorem 1.1 is a particular case of [3, Proposition 6.5].

6 Massey Products

6.1 Massey Products in Galois Cohomology

Here we recall briefly what we need in order to prove Proposition 1.3. For a detailed account on Massey products for pro-p groups, we direct the reader to [8, 19, 35].

Let G be a pro-p group. For \(n\ge 2\), the n-fold Massey product on \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) is a multi-valued map

$$\begin{aligned} \underbrace{\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\times \cdots \times \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})}_{n\text { times}}\longrightarrow \textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z}). \end{aligned}$$

For \(n\ge 2\), given a sequence \(\alpha _1,\ldots ,\alpha _n\) of elements of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) (with possibly \(\alpha _i=\alpha _j\) for some \(1\le i<j\le n\)), the (possibly empty) subset of \(\textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z})\) which is the value of the n-fold Massey product associated to the sequence \(\alpha _1,\ldots ,\alpha _n\) is denoted by \(\langle \alpha _1,\ldots ,\alpha _n\rangle \). If \(n=2\), then the 2-fold Massey product coincides with the cup-product, i.e., for \(\alpha _1,\alpha _2\in \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) one has

$$\begin{aligned} \langle \alpha _1,\alpha _2\rangle =\{\alpha \smallsmile \alpha _2\}\subseteq \textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z}). \end{aligned}$$
(6.1)

A pro-p group G is said to satisfy:

  1. (a)

    the n -Massey vanishing property (with respect to \(\mathbb {Z}/p\mathbb {Z}\)) if for every sequence \(\alpha _1,\ldots ,\alpha _n\) of elements of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\), \(\langle \alpha _1,\ldots ,\alpha _n\rangle \ne \varnothing \) implies \(0\in \langle \alpha _1,\ldots ,\alpha _n\rangle \);

  2. (b)

    the strong n-Massey vanishing property (with respect to \(\mathbb {Z}/p\mathbb {Z}\)) if for every sequence \(\alpha _1,\ldots ,\alpha _n\) of elements of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\), the condition on the cup-products

    $$\begin{aligned} \alpha _1\smallsmile \alpha _2=\alpha _2\smallsmile \alpha _3=\cdots =\alpha _{n-1}\smallsmile \alpha _n=0 \end{aligned}$$
    (6.2)

    implies \(0\in \langle \alpha _1,\ldots ,\alpha _n\rangle \) (cf. [21, Definition 1.2])—we remind that the triviality condition (6.2) is satisfied whenever \(\langle \alpha _1,\ldots ,\alpha _n\rangle \ne \varnothing \), cf., e.g., [19, Sect. 2];

  3. (c)

    the cyclic p-Massey vanishing property if for every element \(\alpha \in \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\), the p-fold Massey product \(\langle \alpha ,\ldots ,\alpha \rangle \) contains 0.

Remark 6.1

Given a sequence \(\alpha _1,\ldots ,\alpha _n\) of elements of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\), if an element \(\omega \) of \(\textrm{H}^2(G,\mathbb {Z}/p\mathbb {Z})\) is a value of the n-fold Massey product \(\langle \alpha _1,\ldots ,\alpha _n\rangle \), then

$$\begin{aligned} \omega +\alpha _1\smallsmile \beta \in \langle \alpha _1,\ldots ,\alpha _n\rangle \qquad \text {and}\qquad \omega +\alpha _n\smallsmile \beta \in \langle \alpha _1,\ldots ,\alpha _n\rangle \end{aligned}$$

for any \(\beta \in \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) (cf. [19, Remark 2.2]).

In [18, Theorem 8.1], J. Minač and N.D. Tân proved that the maximal pro-p Galois group of a field \(\mathbb {K}\) containing a root of 1 of order p (and also \(\sqrt{-1}\) if \(p=2\)) satisfies the cyclic p-Massey vanishing property. The proof of the last property for a pro-p group G as in Theorem 1.1 is rather immediate.

Proof of Proposition 1.3–(ii)

By Proposition 4.1 and Proposition 5.1, G may complete into a Kummerian oriented pro-p group with torsion-free orientation. Hence, G satisfies the cyclic p-Massey vanishing property by [27, Theorem 3.10]. \(\square \)

6.2 Massey Products and Unipotent Upper-Triangular Matrices

Massey products for a pro-p group G may be translated in terms of unipotent upper-triangular representations of G as follows. For \(n\ge 2\) let

$$\begin{aligned} \mathbb {U}_{n+1}=\left\{ \left( \begin{array}{ccccc} 1 &{} a_{1,2} &{} \cdots &{} &{} a_{1,n+1} \\ {} &{} 1 &{} a_{2,3} &{} \cdots &{} \\ &{}&{}\ddots &{}\ddots &{} \vdots \\ {} &{}&{}&{}1&{} a_{n,n+1} \\ {} &{}&{}&{}&{}1 \end{array}\right) \mid a_{i,j}\in \mathbb {Z}/p \right\} \subseteq \textrm{GL}_{n+1}(\mathbb {Z}/p\mathbb {Z}) \end{aligned}$$

be the group of unipotent upper-triangular \((n+1)\times (n+1)\)-matrices over \(\mathbb {Z}/p\). Then \(\mathbb {U}_{n+1}\) is a finite p-group. Moreover, for \(1\le h,l\le n+1\) let \(E_{h,l}\) denote the \((n+1)\times (n+1)\) matrix with the (hl)-entry equal to 1, and all the other entries equal to 0.

Now let \(\rho :G\rightarrow \mathbb {U}_{n+1}\) be a homomorphism of pro-p groups. Observe that for every \(h=1,\ldots ,n\), the projection \(\rho _{h,h+1}:G\rightarrow \mathbb {Z}/p\) of \(\rho \) onto the \((h,h+1)\)-entry is a homomorphism, and thus it may be considered as an element of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\). One has the following “pro-p translation” of a result of W. Dwyer which interprets Massey product in terms of unipotent upper-triangular representations (cf., e.g., [11, Lemma 9.3]).

Proposition 6.2

Let G be a pro-p group, and let \(\alpha _1,\ldots ,\alpha _n\) be a sequence of elements of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\), with \(n\ge 2\). Then the n-fold Massey product \(\langle \alpha _1,\ldots ,\alpha _n\rangle \):

  1. (i)

    is not empty if, and only if, there exists a morphism of pro-p groups \({\bar{\rho }}:G\rightarrow \mathbb {U}_{n+1}/\textrm{Z}(\mathbb {U}_{n+1})\) such that \({\bar{\rho }}_{h,h+1}=\alpha _h\) for every \(h=1,\ldots ,n\);

  2. (ii)

    vanishes if, and only if, there exists a morphism of pro-p groups \(\rho :G\rightarrow \mathbb {U}_{n+1}\) such that \(\rho _{h,h+1}=\alpha _h\) for every \(h=1,\ldots ,n\).

We recall that

$$\begin{aligned} \textrm{Z}(\mathbb {U}_{n+1})=\left\{ \,I_{n+1}+aE_{1,n+1}\,\mid \,a\in \mathbb {Z}/p\mathbb {Z}\,\right\} \simeq \mathbb {Z}/p\mathbb {Z}. \end{aligned}$$

We use this fact to prove statements (iii.a)–(iii.b) of Proposition 1.3. First of all, let G be as in Theorem 1.1, and let \(\alpha _1,\ldots ,\alpha _n\) be a sequence of elements of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\). Keeping the same notation as in Sect. 3.3, for \(h=1,\ldots ,n\) one has

$$\begin{aligned} \alpha _h=\alpha _h(x)\cdot \chi +\sum _{i=1}^{d_1}\alpha _h(y_i)\cdot \varphi _i+\sum _{j=1}^{d_2}\alpha _h(z_j)\cdot \psi _j. \end{aligned}$$

Therefore, for \(h=1,\ldots ,n-1\) one obtains

$$\begin{aligned} \alpha _h\smallsmile \alpha _h=S_h\cdot (\chi \smallsmile \varphi _1)+S'_h\cdot (\chi \smallsmile \psi _1), \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} S_h=&(\alpha _h(x)\alpha _{h+1}(y_1)-\alpha _{h}(y_1)\alpha _{h+1}(x))\\&+(-1)^\epsilon \sum _{2\mid i} (\alpha _h(y_i)\alpha _{h+1}(y_{i+1})-\alpha _{h}(y_{i+1})\alpha _{h+1}(y_i)), \\ S'_h =&(\alpha _h(x)\alpha _{h+1}(z_1)-\alpha _{h}(z_1)\alpha _{h+1}(x))\\&+(-1)^{\epsilon }\sum _{ 2\mid j} (\alpha _h(z_j)\alpha _{h+1}(z_{j+1})-\alpha _{h}(z_{j+1})\alpha _{h+1}(z_j)), \end{aligned}\end{aligned}$$

with \(\epsilon =0\) if G is as in (1.1.a), and \(\epsilon =1\) if G is as in (1.1.b). If the sequence \(\alpha _1,\ldots ,\alpha _n\) satisfies condition (6.2), then one has \(S_h=S'_h=0\) for \(h=1,\ldots ,n-1\), as \(\{\chi \smallsmile \varphi _1,\chi \smallsmile \psi _1\}\) is a basis of \(\textrm{H}^2(G,\mathbb {Z}/p)\).

From now on, we will assume that \(p>3\) while considering a pro-p group G as in (1.1.b), unless stated otherwise.

6.3 3-Fold Massey Products

We are ready to prove the following.

Proposition 6.3

A pro-p group G satisfies the 3-Massey vanishing property in the following cases:

  1. (a)

    if G is as in (1.1.a);

  2. (b)

    if G is as in (1.1.b) and \(p>3\).

Proof

Let \(\alpha _1,\alpha _2,\alpha _3\) be a sequence of elements of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) satisfying (6.2). Then \(S_1=S_1'=S_2=S_2'=0\) (cf. Sect. 6.2). Our goal is to construct a morphism \(\rho :G\rightarrow \mathbb {U}_4\) such that \(\rho _{1,2}=\alpha _1\), \(\rho _{2,3}=\alpha _2\), \(\rho _{3,4}=\alpha _3\).

For every \(w\in \mathcal {X}\) set

$$\begin{aligned} A(w)=I+\alpha _1(w)E_{1,2}+\alpha _2(w)E_{2,3}+\alpha _3(w)E_{3,4}\in \mathbb {U}_4, \end{aligned}$$

where I denotes the \(4\times 4\) identity matrix. If G is as in (1.1.a), then one computes

$$\begin{aligned} \begin{aligned} C&=[A(x),A(y_1)]\cdots [A(y_{d_1-1}),A(y_{d_1})] \\ {}&= I+E_{1,4}\left( \alpha _1(y_1)\alpha _2(x)\alpha _3(y_1)+\sum _{2\mid i}\alpha _1(y_i)\alpha _2(y_{i+1})\alpha _3(y_i)\right) \\ C'&=[A(x),A(z_1)]\cdots [A(z_{d_2-1}),A(z_{d_2})] \\ {}&= I+E_{1,4}\left( \alpha _1(z_1)\alpha _2(x)\alpha _3(z_1)+\sum _{2\mid j}\alpha _1(z_j)\alpha _2(z_{j+1})\alpha _3(z_j)\right) ; \end{aligned}\end{aligned}$$
(6.3)

while if G is as in (1.1.b), then one computes

$$\begin{aligned} \begin{aligned} C&=A(y_1)^p[A(y_1),A(x)]\cdots [A(y_{d_1-1}),A(y_{d_1})] \\&= I+E_{1,4}\left( \alpha _1(x)\alpha _2(y_1)\alpha _3(x)+\sum _{2\mid i}\alpha _1(y_i)\alpha _2(y_{i+1})\alpha _3(y_i)\right) \\ C'&=A(z_1)^p[A(z_1),A(x)]\cdots [A(z_{d_2-1}),A(z_{d_2})] \\&= I+E_{1,4}\left( \alpha _1(x)\alpha _2(z_1)\alpha _3(x)+\sum _{2\mid j}\alpha _1(z_j)\alpha _2(z_{j+1})\alpha _3(z_j)\right) . \end{aligned}\end{aligned}$$
(6.4)

—observe that the exponent of \(\mathbb {U}_4\) is p, as \(p>4\), and thus \(A(y_1)^p=A(z_1)^p=I\).

In both cases, \(C,C'\in \textrm{Z}(\mathbb {U}_4)\), and therefore the assignment \(w\mapsto A(w)\) for every \(w\in \mathcal {X}\) yields a morphism \({\bar{\rho }}:G\rightarrow \mathbb {U}_4/\textrm{Z}(\mathbb {U}_4)\) satisfying \({\bar{\rho }}_{h,h+1}=\alpha _h\) for \(h=1,2,3\). Thus, \(\langle \alpha _1,\alpha _2,\alpha _3\rangle \ne \varnothing \) by Proposition 6.2.

Moreover, if \(C=C'=I\) then the same assignment yields a morphism \(\rho :G\rightarrow \mathbb {U}_4\) with the desired properties. In particular, by (6.3)–(6.4) one has \(C=I\) if \(\alpha _1(w)=\alpha _3(w)=0\) for every \(w=y_1,\ldots ,y_{d_1}\), or for every \(w=y_2,\ldots ,y_{d_1}\) and \(w=x\); and analogously \(C'=I\) if \(\alpha _1(w)=\alpha _3(w)=0\) for every \(w=z_1,\ldots ,z_{d_{d_2}}\), or for every \(w=z_2,\ldots ,z_{d_2}\) and \(w=x\).

On the other hand, if \(C\ne I\) then \(\chi \smallsmile \varphi _1=\pm {{\,\textrm{trg}\,}}(r_1 G_{(3)})\) belongs to \(\langle \alpha _1,\alpha _2,\alpha _3\rangle \), and analogously if \(C'\ne I\) then \(\chi \smallsmile \psi _1=\pm {{\,\textrm{trg}\,}}(r_2 G_{(3)})\) belongs to \(\langle \alpha _1,\alpha _2,\alpha _3\rangle \) (cf. [19, Lemma 3.7])—here the sign depends on whether the relations are as in (1.1.a) or in (1.1.b). Now, if \(\alpha _h(y_i)\ne 0\) for some \(h=1,3\) and \(i\in \{2,\ldots ,d_1\}\), then

$$\begin{aligned} \chi \smallsmile \varphi _1=\alpha _h\smallsmile \beta \qquad \text {for some }\beta \in \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z}). \end{aligned}$$

Analogously, if \(\alpha _h(z_j)\ne 0\) for some \(h=1,3\) and \(j\in \{2,\ldots ,d_2\}\), then

$$\begin{aligned} \chi \smallsmile \psi _1=\alpha _h\smallsmile \beta \qquad \text {for some }\beta \in \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z}). \end{aligned}$$

Moreover, if \(\alpha _h(x)\ne 0\) for some \(h=1,3\), then

$$\begin{aligned} \chi \smallsmile \varphi _1=\alpha _h\smallsmile \beta \qquad \text {and}\qquad \chi \smallsmile \psi _1=\alpha _h\smallsmile \beta ' \end{aligned}$$

for some \(\beta ,\beta '\in \textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\). Therefore, Remark 6.1 implies that if \(C\ne I\) or \(C'\ne I\) then \(0\in \langle \alpha _1,\alpha _2,\alpha _3\rangle \) anyway. \(\square \)

Remark 6.4

If \(p=3\) and G as in (1.1.b), then G does not satisfy the 3-Massey vanishing property. Indeed, set \(\alpha _1=\alpha _3=\varphi _1+\psi _1\), and \(\alpha _2=\varphi _1\). Then

$$\begin{aligned} \alpha _1\smallsmile \alpha _2=\alpha _2\smallsmile \alpha _3=\pm (\varphi _1\smallsmile \psi _1)=0. \end{aligned}$$

It is easy to see that one may construct a morphism of pro-p groups \({\bar{\rho }}:G\rightarrow \mathbb {U}_4/\textrm{Z}(\mathbb {U}_4)\) such that \({\bar{\rho }}_{1,2}={\bar{\rho }}_{3,4}=\alpha _1\) and \({\bar{\rho }}_{2,3}=\alpha _2\)—and thus \(\langle \alpha _1,\alpha _2,\alpha _1\rangle \ne \varnothing \) by Proposition 6.2—but, on the other hand, one may not construct a morphism of pro-p groups \(\rho :G\rightarrow \mathbb {U}_4\) satisfying \(\rho _{1,2}=\rho _{3,4}=\alpha _1\) and \(\rho _{2,3}=\alpha _2\)—so that \(0\notin \langle \alpha _1,\alpha _2,\alpha _1\rangle \) by Proposition 6.2.

6.4 4-Fold Massey Products

Proposition 6.5

A pro-p group G as in Theorem 1.1 satisfies the strong 4-Massey vanishing property.

Proof

Let \(\alpha _1,\ldots ,\alpha _4\) be a sequence of four elements of \(\textrm{H}^1(G,\mathbb {Z}/p\mathbb {Z})\) satisfying (6.2). Our goal is to construct a homomorphism of pro-p groups \(\rho :G\rightarrow \mathbb {U}_5\) such that \(\rho _{h,h+1}=\alpha _h\) for \(h=1,\ldots ,5\), so that the claim follows by Proposition 6.2.

Let I denote the identity matrix of the group \(\mathbb {U}_5\). For every \(w\in \mathcal {X}=\{x,y_1,\ldots ,z_{d_2}\}\) set

$$\begin{aligned} A(w)=\left( \begin{array}{ccccc} 1 &{}\alpha _1(w)&{}0&{}0&{}0 \\ &{}1&{}\alpha _2(w)&{}0&{}0 \\ {} &{}&{}1&{}\alpha _3(w)&{}0 \\ {} &{}&{}&{}1&{}\alpha _4(w) \\ {} &{}&{}&{}&{}1 \end{array} \right) \in \mathbb {U}_5. \end{aligned}$$

Moreover, put

$$\begin{aligned}\begin{aligned} C&=(c_{hl})= A(y_1)^{\epsilon p}\cdot [A(x),A(y_1)]^{(-1)^\epsilon }\cdots \left[ A(y_{d_1-1}),A(y_{d_1})\right] ,\\ C'&=(c'_{hl})= A(z_1)^{\epsilon p}\cdot [A(x),A(z_1)]^{(-1)^\epsilon }\cdots \left[ A(z_{d_2-1}),A(z_{d_2})\right] . \end{aligned}\end{aligned}$$

We will consider the matrix C as a function of the matrices \(A(x),\ldots ,A(y_{d_1})\), and the matrix \(C'\) as a function of the matrices \(A(x),A(z_1),\ldots ,A(z_{d_2})\).

Since \(p\ge 5\), the exponent of the p-group \(\mathbb {U}_5\) is p, and thus \(A(y_1)^p=A(z_1)^p=I\). Moreover, for every \(w,w'\in \mathcal {X}\), the \((h,h+1)\)-entry of \([A(w),A(w')]\) is 0 for every \(h=1,\ldots ,4\), and thus also \(c_{h,h+1}=c'_{h,h+1}=0\). Moreover, for \(h=1,2,3\) one has \(c_{h,h+2}=S_h\) and \(c'_{h,h+2}=S'_h\)—which are equal to 0 by (6.2).

We split the proof in the analysis of the following three cases. Our aim is to modify suitably the matrices A(w)—without modifying the \((h,h+1)\)-entries with \(h=1,\ldots ,4\)—to obtain \(C=C'=I\).

Case 1. Suppose first that:

  1. (1.a)

    \(\alpha _2(x)=\alpha _2(y_i)=0\) for all \(2\le i\le d_1\); or

  2. (1.b)

    \(\alpha _3(x)=\alpha _3(y_i)=0\) for all \(2\le i\le d_1\).

Since \(S_1=S_2=S_3=0\) by (6.2), one has

$$\begin{aligned} \alpha _1(x)\alpha _2(y_1)=\alpha _2(y_1)\alpha _3(x)= & {} 0, \end{aligned}$$
(6.5)
$$\begin{aligned} \alpha _2(x)\alpha _3(y_1)=\alpha _3(y_1)\alpha _4(x)= & {} 0, \end{aligned}$$
(6.6)

respectively, in case (1.a) and in case (1.b). Applying (6.5)–(6.6), one computes

$$\begin{aligned}{}[A(y_1),A(x)]={\left\{ \begin{array}{ll} I+\left( \alpha _3(y_1)\alpha _4(x)-\alpha _3(x)\alpha _4(y_1)\right) E_{3,5} &{}\quad \text {in case (1.a)}, \\ I+\left( \alpha _1(y_1)\alpha _2(x)-\alpha _2(x)\alpha _1(y_1)\right) E_{1,3} &{}\quad \text {in case (1.b)}, \end{array}\right. }\end{aligned}$$

and

$$\begin{aligned}{} & {} [A(y_i),A(y_{i+1})]=\\{} & {} \quad {\left\{ \begin{array}{ll} I+\left( \alpha _3(y_i)\alpha _4(y_{i+1})-\alpha _3(y_{i+1})\alpha _4(y_i)\right) E_{3,5} &{}\quad \text {in case (1.a)}, \\ I+\left( \alpha _1(y_i)\alpha _2(y_{i+1})-\alpha _2(y_{i+1})\alpha _1(y_i)\right) E_{1,3} &{}\quad \text {in case (1.b)}, \end{array}\right. }\end{aligned}$$

for \(i=2,4,\ldots ,d_1-1\). Altogether, one has \(C=I+S_3E_{3,5}\) in case (1.a) and \(C=I+S_1E_{1,3}\) in case (1.b), so that in both cases \(C=I\) by (6.2).

Analogously, if \(\alpha _2(x)=\alpha _2(z_j)=0\) for all \(2\le j\le d_2\), or if \(\alpha _3(x)=\alpha _3(z_j)=0\) for all \(2\le j\le d_2\), then \(C'=I\). This completes the analysis of case 1.

Case 2. Now suppose that \(\alpha _1(x)=\alpha _4(x)=\alpha _1(y_i)=\alpha _4(y_i)=0\) for all \(2\le i\le d_1\). Since \(S_1=S_2=S_3=0\) by (6.2), one has

$$\begin{aligned} \alpha _1(y_1)\alpha _2(x)=\alpha _3(x)\alpha _4(y_1)=0. \end{aligned}$$
(6.7)

Then one computes

$$\begin{aligned} \begin{aligned}&[A(y_1),A(x)] = I+\left( \alpha _2(y_1)\alpha _3(x)-\alpha _2(x)\alpha _3(y_1)\right) E_{2,4}+\alpha _2(x)\alpha _3(y_1)\alpha _4(y_1)E_{2,5},\\&[A(y_i),A(y_{i+1})] =I+\left( \alpha _2(y_i)\alpha _3(y_{i+1})-\alpha _2(y_{i+1})\alpha _3(y_i)\right) E_{2,4}, \end{aligned} \end{aligned}$$

where we apply (6.7) to obtain the first equality, and in the second one i runs through the even positive integers between 2 and \(d_1-1\). If \(\alpha _2(x)\alpha _3(y_1)\alpha _4(y_1)=0\) then it is straightforward to see that \(C=I+S_2E_{2,4}=I\). Otherwise, \(\alpha _2(x)\ne 0\), so that (6.7) implies that \(\alpha _1(y_1)=0\). In this case, set

$$\begin{aligned} {\tilde{A}}=I-\alpha _3(y_1)\alpha _4(y_1)E_{3,5}. \end{aligned}$$

Then

$$\begin{aligned} \left[ \tilde{A},A(x)\right] =I-\alpha _2(x)\alpha _3(y_1)\alpha _4(y_1)E_{2,5}, \end{aligned}$$

and

$$\begin{aligned}\begin{aligned} \left[ A(y_1){\tilde{A}},A(x)\right]&= \underbrace{\left[ A(y_1),[{\tilde{A}},A(x)]\right] }_{=I}\left[ {\tilde{A}},A(x)\right] [A(y_1),A(x)]\\&= I+\left( \alpha _2(y_1)\alpha _3(x)-\alpha _2(x)\alpha _3(y_1)\right) E_{2,4}. \end{aligned} \end{aligned}$$

Therefore, replacing \(A(y_1)\) with \(A(y_1){\tilde{A}}\) yields \(c_{2,4}=S_2=0\) and \(C_{hl}=0\) for \(h<l\), i.e., \(C=I\).

An analogous argument yields \(C'=I\)—after replacing suitably the matrix \(A(z_1)\) if needed—if \(\alpha _1(x)=\alpha _3(x)=\alpha _1(z_j)=\alpha _3(z_j)=0\) for all \(1\le j\le d_2\). This completes the analysis of case 2.

Case 3. Finally, if none of the above two assumptions on the triviality of the values \(\alpha _h(x)\) and \(\alpha _h(y_i)\), with \(2\le i\le d_1\), hold true, then

  1. (3.a)

    there are \(w,w'\in \{x,y_2,\ldots ,y_{d_1}\}\)—possibly \(w=w'\)—such that \(\alpha _1(w)\ne 0\) and \(\alpha _2(w')\ne 0\), or

  2. (3.b)

    there are \(w,w'\in \{x,y_2,\ldots ,y_{d_1}\}\)—possibly \(w=w'\)—such that \(\alpha _4(w)\ne 0\) and \(\alpha _3(w')\ne 0\).

Suppose we are in case (3.a). If \(w=x\) or \(w=y_i\) with i odd, set

$$\begin{aligned} {\tilde{A}}={\left\{ \begin{array}{ll} I+\frac{c_{1,4}}{\alpha _1(w)}E_{2,4}&{}\quad \text {if }w\in \{\,x,y_3,\ldots ,y_{d_1}\,\}\\ I-\frac{c_{1,4}}{\alpha _1(w)}E_{2,4}&{} \quad \text {if }w\in \{\,y_i\,\mid \,i\text { is even}\,\}, \end{array}\right. } \end{aligned}$$

and replace \(A(y_1)\) with \(A(y_1){\tilde{A}}\), if \(w=x\), or \(A(y_{i-1})\) with \(A(y_{i-1}){\tilde{A}}\) if \(w=y_i\) with i odd, or \(A(y_{i+1})\) with \(A(y_{i+1}){\tilde{A}}\), if \(w=y\) with i even. After the replacement, one has \(c_{hl}=0\) for \(h< l\le h+2\), and for \((h,l)=(1,4)\). Then, set

$$\begin{aligned} {\tilde{A}}'={\left\{ \begin{array}{ll} I+\frac{c_{2,5}}{\alpha _1(w')}E_{3,5}&{} \quad \text {if }w'\in \{\,x,y_3,\ldots ,y_{d_1}\,\}\\ I-\frac{c_{2,5}}{\alpha _1(w')}E_{3,5}&{} \quad \text {if }w'\in \{\,y_i\,\mid \,\text { is even}\,\}, \end{array}\right. } \end{aligned}$$

and replace \(A(y_1)\) with \(A(y_1){\tilde{A}}'\), if \(w=x\), or \(A(y_{i-1})\) with \(A(y_{i-1}){\tilde{A}}'\) if \(w=y_i\) with i odd, or \(A(y_{i+1})\) with \(A(y_{i+1}){\tilde{A}}'\), if \(w=y\) with i even. After this further replacement, one has \(c_{hl}=0\) for \(h< l\le h+3\). Finally, set

$$\begin{aligned} {\tilde{A}}''={\left\{ \begin{array}{ll} I+\frac{c_{1,5}}{\alpha _1(w)}E_{2,5}&{} \quad \text {if }w\in \{\,x,y_3,\ldots ,y_{d_1}\,\}\\ I-\frac{c_{1,5}}{\alpha _1(w)}E_{2,5}&{} \quad \text {if }w\in \{\,y_i\,\mid \,i\text { is even}\,\}, \end{array}\right. } \end{aligned}$$

and replace \(A(y_1)\) with \(A(y_1){\tilde{A}}''\), if \(w=x\), or \(A(y_{i-1})\) with \(A(y_{i-1}){\tilde{A}}''\) if \(w=y_i\) with i odd, or \(A(y_{i+1})\) with \(A(y_{i+1}){\tilde{A}}''\), if \(w=y\) with i even. After this last replacement, one has \(C=I\).

Now suppose we are in case (3.b). If \(w=x\) or \(w=y_i\) with i odd, set

$$\begin{aligned} {\tilde{A}}={\left\{ \begin{array}{ll} I-\frac{c_{2,5}}{\alpha _4(w)}E_{3,4}&{} \quad \text {if }w\in \{\,x,y_3,\ldots ,y_{d_1}\,\}\\ I+\frac{c_{2,5}}{\alpha _4(w)}E_{3,4}&{} \quad \text {if }w\in \{\,y_i\,\mid \, i\text { is even}\,\}, \end{array}\right. } \end{aligned}$$

and replace \(A(y_1)\) with \(A(y_1){\tilde{A}}\), if \(w=x\), or \(A(y_{i-1})\) with \(A(y_{i-1}){\tilde{A}}\) if \(w=y_i\) with i odd, or \(A(y_{i+1})\) with \(A(y_{i+1}){\tilde{A}}\), if \(w=y\) with i even. After the replacement, one has \(c_{hl}=0\) for \(h< l\le h+2\), and for \((h,l)=(2,5)\). Then, set

$$\begin{aligned} {\tilde{A}}'={\left\{ \begin{array}{ll} I-\frac{c_{1,4}}{\alpha _3(w')}E_{1,3}&{} \quad \text {if }w'\in \{\,x,y_3,\ldots ,y_{d_1}\,\}\\ I+\frac{c_{1,4}}{\alpha _3(w')}E_{1,3}&{} \quad \text {if }w'\in \{\,y_i\,\mid \,i\text { is even}\,\}, \end{array}\right. } \end{aligned}$$

and replace \(A(y_1)\) with \(A(y_1){\tilde{A}}'\), if \(w=x\), or \(A(y_{i-1})\) with \(A(y_{i-1}){\tilde{A}}'\) if \(w=y_i\) with i odd, or \(A(y_{i+1})\) with \(A(y_{i+1}){\tilde{A}}'\), if \(w=y\) with i even. After this further replacement, one has \(c_{hl}=0\) for \(h< l\le h+3\). Finally, set

$$\begin{aligned} {\tilde{A}}''={\left\{ \begin{array}{ll} I-\frac{c_{1,5}}{\alpha _1(w)}E_{1,4}&{} \quad \text {if }w\in \{\,x,y_3,\ldots ,y_{d_1}\,\}\\ I+\frac{c_{1,5}}{\alpha _1(w)}E_{1,4}&{} \quad \text {if }w\in \{\,y_i\,\mid \, i\text { is even}\,\}, \end{array}\right. } \end{aligned}$$

and replace \(A(y_1)\) with \(A(y_1){\tilde{A}}''\), if \(w=x\), or \(A(y_{i-1})\) with \(A(y_{i-1}){\tilde{A}}''\) if \(w=y_i\) with i odd, or \(A(y_{i+1})\) with \(A(y_{i+1}){\tilde{A}}''\), if \(w=y\) with i even. After this last replacement, one has \(C=I\).

Moreover, if none of the above two assumptions on the triviality of the values \(\alpha _h(x)\) and \(\alpha _h(z_j)\), with \(2\le j\le d_2\), hold true, the same argument produces suitable matrices \(A(z_1),\ldots ,A(z_{d_2})\) such that the matrix \(C'\) is the identity matrix. This concludes the analysis of case 3.

Altogether, the assignment \(w\mapsto A(x)\) for every \(w\in \mathcal {X}\)—with the matrices A(w)’s suitably modified in case of need—yields a homomorphism of pro-p groups \(\rho :G\rightarrow \mathbb {U}_5\) with the desired properties. \(\square \)

We believe that the answer to the following questions is positive.

Question 6.6

  1. (a)

    Let G be as in (1.1.a). Does G satisfy the strong n-Massey vanishing property for every \(n\ge 3\)?

  2. (b)

    Let G be as in (1.1.b). Does G satisfy the strong n-Massey vanishing property for every \(3\le n<p\)?

7 The Minač–Tân Pro-p Group

We focus now on the Minač–Tân pro-p group

$$\begin{aligned} G=\langle \,x_1,\ldots ,x_5\,\mid \,r=1\,\rangle \qquad \text {with }r=\left[ [x_1,x_2],x_3\right] [x_4,x_5]. \end{aligned}$$

Using Proposition 6.2, one may show that G does not satisfy the 3-Massey vanishing property (cf. [19, Example 7.2]). Our aim is to show that G cannot complete into a 1-cyclotomic oriented pro-p group with torsion-free orientation.

7.1 Kummerianity and 1-Cyclotomicity

Proposition 7.1

Let G be the Minač–Tân pro-p group, and let \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) be a torsion-free orientation. Then the oriented pro-p group \((G,\theta )\) is Kummerian if, and only if, \(x_4,x_5\in {{\,\textrm{Ker}\,}}(\theta )\), and:

  1. (a)

    \(x_3\in {{\,\textrm{Ker}\,}}(\theta )\); or

  2. (b)

    \(x_1,x_2\in {{\,\textrm{Ker}\,}}(\theta )\).

Proof

Let \(c:G\rightarrow \mathbb {Z}_p(\theta )\) be an arbitrary continuous 1-cocycle, and set \(c(x_i)=\lambda _i\) for \(i=1,\ldots ,5\). Applying (2.2)–(2.3) one computes \(c(r)=c([[x_1,x_2],x_3])+c([x_4,x_5])\), and

$$\begin{aligned} c([[x_1,x_2],x_3])= & {} \theta (x_1x_2)^{-1}\left( \theta (x_3)^{-1}-1\right) \left( \lambda _1(1-\theta (x_2))-\lambda _2(1-\theta (x_1))\right) ,\nonumber \\ c([x_4,x_5])= & {} \theta (x_4x_5)^{-1}\left( \lambda _4(1-\theta (x_5))-\lambda _5(1-\theta (x_4))\right) . \end{aligned}$$
(7.1)

On the other hand, \(c(r)=0\) as \(r=1\).

Suppose that \((G,\theta )\) is Kummerian. Then by Lemma 2.9, we may prescribe arbitrary values to \(\lambda _1,\ldots ,\lambda _5\). If \(\lambda _4=1\) and \(\lambda _i=0\) for \(i\ne 4\), from (7.1) and from the fact that \(c(r)=0\) one obtains \(0=1\cdot (1-\theta (x_5))\), and thus \(\theta (x_5)=1\). Analogously, if \(\lambda _5=1\) and \(\lambda _i=0\) for \(i\ne 5\), one deduces \(\theta (x_4)=1\). Finally, if \(\lambda _4=\lambda _5=0\) from (7.1) one obtains

$$\begin{aligned} 0=c(r)=\theta (x_1x_2)^{-1}\left( \theta (x_3)^{-1}-1\right) \left( \lambda _1(1-\theta (x_2))-\lambda _2(1-\theta (x_1))\right) , \end{aligned}$$

and the arbitrariness of \(\lambda _1,\lambda _2\) implies that \(\theta (x_3)=1\) or \(\theta (x_1)=\theta (x_2)=1\).

Conversely, suppose that \(x_4,x_5\in {{\,\textrm{Ker}\,}}(\theta )\), and at least one of the hypothesis (i)–(ii) holds true. Then for any choice for \(\lambda _4,\lambda _5\), by (7.1) one has \(c([x_4,x_5])=0\). On the other hand, one has

$$\begin{aligned} c([[x_1,x_2],x_3])={\left\{ \begin{array}{ll} 0\cdot (\lambda _1(1-\theta (x_2))-\lambda _2(1-\theta (x_1)))=0 &{}\quad \text {if }x_3\in {{\,\textrm{Ker}\,}}(\theta ), \\ \left( \theta (x_3)^{-1}-1\right) (\lambda _1\cdot 0-\lambda _2\cdot 0)=0 &{}\quad \text {if }x_1,x_2\in {{\,\textrm{Ker}\,}}(\theta ). \end{array}\right. }\end{aligned}$$

Altogether, any choice for \(\lambda _1,\ldots ,\lambda _5\) yields a well-defined continuous 1-cocycle \(c:G\rightarrow \mathbb {Z}_p(\theta )\), and thus \((G,\theta )\) is Kummerian by Lemma 2.9. \(\square \)

Now consider the subgroup H of G generated by \(x_3,x_4,x_5\) and by \(y=[x_1,x_2]\). Then H is subject to the relation

$$\begin{aligned} r=[y,x_3][x_4,x_5]=1. \end{aligned}$$

If \((G,\theta )\) is a 1-cyclotomic oriented pro-p group, with \(\theta \) a torsion-free orientation, then \(=(H,\theta \vert _H)\) is Kummerian. Therefore, if \(c':H\rightarrow \mathbb {Z}_p(\theta \vert _H)\) is a continuous 1-cocycle, applying (2.2)–(2.3) one obtains

$$\begin{aligned} \begin{aligned} 0=c'(r)&=c'([y,x_3])+c'([x_4,x_5])\\ {}&=\theta (yx_3)^{-1}\left( c'(y)(1-\theta (x_3))-c'(x_3)(1-\theta (y))\right) +0\\&=\theta (yx_3)^{-1}c'(y)(1-\theta (x_3)), \end{aligned} \end{aligned}$$

as \(\theta (x_4)=\theta (x_5)=1\) by Proposition 7.1, and \(y\in G'\subseteq {{\,\textrm{Ker}\,}}(\theta )\). Since \(c'(y)\) may be arbitrarily chosen by Lemma 2.9, one deduces \(\theta (x_3)=1\). This proves the following.

Lemma 7.2

Let G be the Minač–Tân pro-p group, and let \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) be a torsion-free orientation. If the oriented pro-p group \((G,\theta )\) is 1-cyclotomic then \(x_3,x_4,x_5\in {{\,\textrm{Ker}\,}}(\theta )\).

Moreover, if \((G,\theta )\) is 1-cyclotomic we may suppose without loss of generality that \(x_2\in {{\,\textrm{Ker}\,}}(\theta )\), too. Indeed, let \(v_p:\mathbb {Z}_p\rightarrow \mathbb {N}\) denote the p-adic valuation, and let \(k\ge 1\) be such that \({{\,\textrm{Im}\,}}(\theta )=1+p^k\mathbb {Z}_p\).

Suppose first that \(v_p(\theta (x_2)-1)=k\) and \(v_p(\theta (x_1)-1)>k\), and set \(z=x_2x_1\). Then \(\{z,x_2,x_3,x_4,x_5\}\) is a minimal generating set of G, \(v_p(\theta (z)-1)=k\), and G is subject to the relation

$$\begin{aligned} \left[ [z,x_2],x_3\right] [x_4,x_5]=1, \end{aligned}$$

as \([x_2x_1,x_2]=[x_1,x_2]\). Hence, we may assume \(v_p(\theta (x_1)-1)=k\).

Consequently, there exists \(\lambda \in \mathbb {Z}_p\) such that \(\theta (x_2)=\theta (x_1)^\lambda \). Now set \(z=x_1^{-\lambda }x_2\). Then \(\{x_1,z,x_3,x_4,x_5\}\) is a minimal generating set of G, \(\theta (z)=\theta (x_2)\theta (x_1)^{-\lambda }=1\), and G is subject to the relation

$$\begin{aligned} \left[ [x_1,z],x_3\right] [x_4,x_5]=1, \end{aligned}$$

as \([x_1,x_1^{-\lambda } x_2]=[x_1,x_2]\).

Therefore, from now on \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) will denote a torsion-free orientation satisfying \(x_2,\ldots ,x_5\in {{\,\textrm{Ker}\,}}(\theta )\).

7.2 The Subgroup U

Put \(u=x_1^p\) and \(t=x_1^{-1}x_3\). Let \(\phi :G\rightarrow \mathbb {Z}/p\) be the homomorphism defined by \(\phi (x_1)=\phi (x_3)=1\) and \(\phi (x_i)=0\) for \(i=2,4,5\), and let U be the kernel of \(\phi \). Then U is a normal subgroup of G of index p, and it is generated as a normal subgroup of G by \(\{u,t,x_2,x_4,x_5\}\). In fact, U is generated as a pro-p group by the set

$$\begin{aligned} \mathcal {X}_U=\left\{ \,u,\, t^{x_1^h},\,x_2^{x_1^h},\,x_4^{x_1^h},\,x_5^{x_1^h}\,\mid \,h=0,\ldots ,p-1\,\right\} , \end{aligned}$$

as \(G/U=\{U,x_1U,\ldots ,x_1^{p-1}U\}\). We need to find a subset of \(\mathcal {X}_U\) which minimally generates U as a pro-p group.

Proposition 7.3

The set

$$\begin{aligned} \mathcal {Y}_U=\left\{ \,t,\,x_2,\,x_2^{x_1},\, t^{x_1^h},\,x_4^{x_1^h},\,x_5^{x_1^h}\,\mid \,h=0,\ldots ,p-1\,\right\} , \end{aligned}$$

is a minimal generating set of U as a pro-p group. Moreover, the abelian pro-p group \(U^{{{\,\textrm{ab}\,}}}\) is not torsion-free.

Proof

The subgroup U is the pro-p group generated by \(\mathcal {X}_U\) and subject to the p-relations \(r^{x_1^h}=1\), \(h=0,\ldots ,p-1\). Since \(x_3=x_1t\), one computes

$$\begin{aligned} \begin{aligned} {[}[x_1,x_2],x_3]&= [x_1,x_2]^{-1}\cdot [x_1,x_2]^{x_3}\\&=[x_2,x_1]\cdot \left[ x_1,x_2^{x_1}\right] ^{t}\\&=x_2^{-1}\cdot x_2^{x_1}\cdot \left( \left( x_2^{x_1^2}\right) ^{-1}x_2^{x_1}\right) ^{t}. \end{aligned} \end{aligned}$$
(7.2)

From (7.2), and from the relation \(r=1\), one deduces the equivalence

$$\begin{aligned} \left( x_2^{x_1^2}\right) ^{-1}\cdot \left( x_2^{x_1}\right) ^2\cdot x_1^{-1}\equiv 1\mod U', \end{aligned}$$
(7.3)

as \([x_4,x_5]\in U'\) and \(t\in U\).

Hence, \(U^{{{\,\textrm{ab}\,}}}\) is the abelian pro-p group generated by \(\mathcal {X}_{U^{{{\,\textrm{ab}\,}}}}=\{wU'\,\mid \,w\in \mathcal {X}_U\}\) and subject to the p relations induced by the equivalences \(((x_2^{x_1^2})^{-1}(x_2^{x_1})^2 x_1^{-1})^{x_1^h}\equiv 1\bmod U'\), namely

$$\begin{aligned} \begin{aligned} x_2^{x_1^2}&\equiv \left( x_2^{x_1}\right) ^2 x_1^{-1}\mod U',\qquad \text {for }h=0,\\ x_2^{x_1^3}&\equiv \left( x_2^{x_1^2}\right) ^2 \left( x_1^{x_2}\right) ^{-1}\equiv \left( x_2^{x_1}\right) ^3 x_1^{-2}\mod U',\qquad \text {for }h=1,\\&\quad \vdots \\ x_2^{x_1^{p}}&\equiv \left( x_2^{x_1^{p-1}}\right) ^2 \left( x_1^{p-2}\right) ^{-1}\equiv \left( x_2^{x_1}\right) ^p x_1^{1-p}\mod U',\qquad \text {for }h=p-2,\\ x_2^{x_1^{p+1}}&\equiv \left( x_2^{x_1}\right) ^2\cdot x_1^{-1}\equiv \left( x_2^{x_1}\right) ^{p+1} x_1^{-p}\mod U',\qquad \text {for }h=p-1. \end{aligned} \end{aligned}$$
(7.4)

On the one hand, from (7.4) one deduces that the coset \(x_2^{x_1^h}U'\) is generated by \(x_2U'\) and \(x_2^{x_1}U'\) for every \(h=2,\ldots ,p-1\), so that \(\mathcal {Y}_{U^{{{\,\textrm{ab}\,}}}}=\{wU'\,\mid \,w\in \mathcal {Y}_U\}\) generates \(U^{{{\,\textrm{ab}\,}}}\) as an abelian pro-p group. On the other hand, from the equivalences with \(h=p-2\) and \(h=p-1\) one deduces that

$$\begin{aligned}\begin{aligned} \left( x_2^{x_1}\right) ^p x_1^{1-p}\left( x_2^{u}\right) ^{-1}&\equiv \left( x_2^{x_1}\right) ^p x_1^{1-p-1}\equiv \left( x_2^{x_1} x_1^{-1}\right) ^p\equiv 1\mod U',\\ \left( x_2^{x_1}\right) ^{p+1} x_1^{-p}\left( x_2^{ux_1}\right) ^{-1}&\equiv \left( x_2^{x_1}\right) ^{p+1-1} x_1^{-p}\equiv \left( x_2^{x_1} x_1^{-1}\right) ^p\equiv 1\mod U', \end{aligned} \end{aligned}$$

as \(x_2^u\equiv x_2\bmod U'\); therefore, they yield equivalent relations in \(U^{{{\,\textrm{ab}\,}}}\). Altogether, \(U^{{{\,\textrm{ab}\,}}}\) is the abelian pro-p group minimally generated by \(\mathcal {X}_{U^{{{\,\textrm{ab}\,}}}}\) and subject to the relation

$$\begin{aligned} \left( (x_2U')^{-1}\cdot x_2^{x_1}U'\right) ^p=1. \end{aligned}$$

Hence, \(U^{{{\,\textrm{ab}\,}}}\) is not torsion-free, and \(\mathcal {Y}_U\) is a minimal generating set of U by Fact 2.1. \(\square \)

From Proposition 7.3, one deduces that G is not absolutely torsion-free, and thus the oriented pro-p group \((G,\textbf{1})\) is not 1-cyclotomic.

7.3 1-Cyclotomicity and the Minač–Tân Pro-p Group

We are ready to prove Theorem 1.4.

Proof

Suppose for contradiction that there exists a torsion-free orientation \(\theta :G\rightarrow 1+p\mathbb {Z}_p\) such that the oriented pro-p group \((G,\theta )\) is 1-cyclotomic. Then by Sect. 7.1, we may assume without loss of generality that \(x_2,\ldots ,x_5\in {{\,\textrm{Ker}\,}}(\theta )\), while \(\theta (x_1)\ne 1\) by Sect. 7.2. Set \(\lambda \in p\mathbb {Z}_p\smallsetminus \{0\}\) such that \(\theta (x_1)=1+\lambda \).

Consider the oriented pro-p group \((U,\theta \vert _U)\), and set \(K=K_{\theta \vert _U}(U)\), \({\bar{U}}=U/K\). Our goal is to show that the oriented pro-p group \(({\bar{U}},(\theta \vert _U)_{/K})\) is not \((\theta \vert _U)_{/K}\)-abelian, so that \((U,\theta \vert _U)\) is not Kummerian by Proposition 2.8, and thus \((G,\theta )\) is not 1-cyclotomic.

Since \(K\subseteq \Phi (U)\), by Proposition 7.3 the set \(\mathcal {Y}_{{\bar{U}}}=\{wK\,\mid \,w\in \mathcal {Y}_U\}\) is a minimal generating set of \({\bar{U}}\). Now, since \(\theta (t)=\theta (x_1)=(1+\lambda )^{-1}\), one has \(w^t\equiv w^{1+\lambda }\bmod K\) for every \(w\in U\). Therefore, from (7.2), and from the fact that \([x_4,x_5]\in {{\,\textrm{Ker}\,}}(\theta \vert _U)'\subseteq K\), one obtains

$$\begin{aligned}{}[x_1,x_2]^{-1}\left( [x_1,x_2]^{x_1}\right) ^t\equiv [x_1,x_2]^{-1}\left( [x_1,x_2]^{x_1}\right) ^{(1+\lambda )^{-1}}\equiv 1\mod K, \end{aligned}$$

and consequently

$$\begin{aligned} \begin{aligned}{}[x_1,x_2]^{x_1}&\equiv [x_1,x_2]^{1+\lambda }\mod K,\\ [x_1,x_2]^{x_1^2}&\equiv [x_1,x_2]^{(1+\lambda )^2}\mod K,\\&\quad \vdots \\ [x_1,x_2]^{x_1^{p-1}}&\equiv [x_1,x_2]^{(1+\lambda )^{p-1}}. \end{aligned} \end{aligned}$$
(7.5)

Set

$$\begin{aligned}\mu =(1+\lambda )^0+(1+\lambda )^1+\cdots +(1+\lambda )^{p-1}=\dfrac{(1+\lambda )^p-1}{\lambda }.\end{aligned}$$

Then \(\mu \ne 0\) (as \(\lambda \ne 0\)), and \(p\mid \mu \). Since \([x_1,x_2]=(x_2^{x_1})^{-1}x_2\), replacing the coset \(x_2^{x_1}K\) with the coset \([x_1,x_2]K\) in \(\mathcal {Y}_{{\bar{U}}}\) yields another minimal generating set—let us call it \(\mathcal {Y}'_{{\bar{U}}}\)—of \({\bar{U}}\). Now, from (7.5) one obtains

$$\begin{aligned} \begin{aligned}{}[u,x_2]&=[x_1,x_2]^{x_1^{p-1}}\cdots [x_1,x_2]^{x_1}\cdot [x_1,x_2]\\&\equiv [x_1,x_2]^{(1+\lambda )^{p-1}}\cdots [x_1,x_2]^{1+\lambda }\cdot [x_1,x_2]\mod K\\&\equiv [x_1,x_2]^{\mu }\mod K \end{aligned} \end{aligned}$$

—observe that \([x_1,x_2]^{x_i^h}\in {{\,\textrm{Ker}\,}}(\theta \vert _U)\) for every h, and thus all such elements commute modulo K. Therefore, one has the relation

$$\begin{aligned} \left( [x_1,x_2]K\right) ^\mu =[uK,x_2K] \end{aligned}$$

between elements of the minimal generating set \(\mathcal {Y}'_{{\bar{U}}}\), and by [11, Theorem 8.1] this relation prevents the oriented pro-p group \(({\bar{U}},(\theta \vert _U)_{/K})\) from being Kummerian—and thus also \((\theta \vert _U)_{/K}\)-abelian. \(\square \)

From Theorem 1.4, we obtain a new family of pro-p groups which cannot complete into 1-cyclotomic oriented pro-p groups.

Corollary 7.4

Let G be the pro-p group with presentation

$$\begin{aligned} G=\left\langle x_1,\ldots ,x_n,x_{n+1},x_{n+2} \,\mid \, \left[ \left[ \ldots [[x_1,x_2],x_3],\ldots x_{n-1}\right] ,x_n\right] [x_{n+1},x_{n+2}]=1\right\rangle , \end{aligned}$$

with \(n\ge 3\). Then G cannot complete into a 1-cyclotomic oriented pro-p group with torsion-free orientation.

Proof

Set \(y=[\ldots [x_1,x_2],\ldots x_{n-2}]\), and let H be the subgroup of G generated by \(\{y,x_{n-1},\ldots ,x_{n+2}\}\). Then

$$\begin{aligned} H=\langle \,y,x_{n-1},\ldots ,x_{n+2}\,\mid \,[[y,x_{n-1}],x_n][x_{n+1},x_{n+2}]\,\rangle \end{aligned}$$

is isomorphic to the Minač–Tân pro-p group, and hence it cannot complete into a 1-cyclotomic oriented pro-p group with torsion-free orientation by Theorem 1.4. \(\square \)

The following question remains open (cf. [2, Example 3.2]).

Question 7.5

Is the Minač–Tân pro-p group G a Bloch–Kato pro-p group? Namely, is the \(\mathbb {Z}/p\mathbb {Z}\)-cohomology algebra of every closed subgroup of G a quadratic algebra?