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

Inducing spectral gaps for the cohomological
Laplacians of SLn()\operatorname{SL}_{n}(\mathbb{Z}) and SAut(Fn)\operatorname{SAut}(F_{n})

Piotr Mizerka [email protected] Institute of Mathematics of the Polish Academy of Sciences, Śniadeckich 8, 00-656, Warsaw, Poland
Abstract.

The technique of inducing spectral gaps for cohomological Laplacians in degree zero was used by Kaluba, Kielak and Nowak [KKN21] to prove property (T) for Aut(Fn)\operatorname{Aut}(F_{n}) and SLn()\operatorname{SL}_{n}(\mathbb{Z}). In this paper, we adapt this technique to Laplacians in degree one. This allows to provide a lower bound for the cohomological Laplacian in degree one for SLn()\operatorname{SL}_{n}(\mathbb{Z}) for every unitary representation. In particular, one gets in that way an alternative proof of property (T) for SLn()\operatorname{SL}_{n}(\mathbb{Z}) whenever n3n\geq 3.

1. Introduction

Cohomological Laplacians were introduced by Bader and Nowak [BN20] in the context of vanishing of group cohomology with unitary coefficients. Such vanishing of the first cohomology is known to be equivalent to Kazhdan’s property (T) and an algebraic characterization for it was found by Ozawa [Oza16] for finitely generated groups. It turns out that property (T) can be described in terms of a specific degree zero Laplacian Δ\Delta: Ozawa showed that a group GG has property (T) if and only if Δ2λΔ\Delta^{2}-\lambda\Delta is a sum of squares in the real group ring G\mathbb{R}G for some positive spectral gap λ\lambda. Using this characterization, Kaluba, Kielak, Nowak, and Ozawa [KNO19, KKN21] were able to show property (T) for SLn()\operatorname{SL}_{n}(\mathbb{Z}) and Aut(Fn)\operatorname{Aut}(F_{n}) for n3n\geq 3 and n5n\geq 5 respectively. Estimating the spectral gap λ\lambda from below has another advantage of prividing a lower bound for κ(G,S)\kappa(G,S), the Kazhdan constant associated to a generating set SS of a group GG. It is known that GG has property (T) if and only if κ(G,S)>0\kappa(G,S)>0 and this inequality does not depend on the generating set. For a symmetric generating set S=S1S=S^{-1}, the lower bound for Kazhdan constants is given by the formula (see [BdlHV08, Remark 5.4.7] and [KN18])

2λ|S|κ(G,S),\sqrt{\frac{2\lambda}{|S|}}\leq\kappa(G,S),

where λ\lambda is a positive constant such that Δ2λΔ\Delta^{2}-\lambda\Delta is a sum of squares.

We focus on providing lower bounds for the spectral gap of Δ1\Delta_{1}, the one-degree Laplacian. The existence of such a positive spectral gap for a group GG is equivalent to vanishing of the first cohomology of GG with unitary coefficients and reducibility of the second (reducibility means that the images of the differentials in the chain complex of Hilbert spaces computing the cohomology for a given unitary representation are closed), see [BN20, The Main Theorem]. Once a positive spectral gap is found for Δ1\Delta_{1}, it turns out to be a lower bound for the Ozawa’s spectral gap, see Remark 2.2. We acknowledge that it has been recently announced by U. Bader and R. Sauer that the reducibility condition is a consequence of vanishing of cohomology one degree lower, see [BS23].

Recently, M. Kaluba, P. W. Nowak and the author showed that for Δ1\Delta_{1}, the first cohomological Laplacian of SL3()\operatorname{SL}_{3}(\mathbb{Z}), and λ=0.32\lambda=0.32, the matrix Δ1λI\Delta_{1}-\lambda I is a sum of squares ([KMN24, Theorem 1.1]). In this paper we show how to use a slightly stronger statement for SL3()\operatorname{SL}_{3}(\mathbb{Z}) to induce the spectral gap for Gn=SLn()G_{n}=\operatorname{SL}_{n}(\mathbb{Z}), whenever n3n\geq 3. Our technique works as well for the case Gn=SAut(Fn)G_{n}=\operatorname{SAut}(F_{n}). We remark however that in such a case we were unable to obtain such a spectral gap for any particular nn. On the other other hand, in his preprint [Nit22] M. Nitsche shows property (T) for SAut(F4)\operatorname{SAut}(F_{4}). This, in combination with the work of Bader and Nowak [BN20] and the announcement of Bader and Sauer, would yield the desired gap for the Laplacian in degree one for SAut(F4)\operatorname{SAut}(F_{4}). Nevertheless, this will still not allow us to induce the spectral gap since we would need a spectral gap for a specific summand of the Laplacian of SAut(F4)\operatorname{SAut}(F_{4}).

Our induction technique is inspired by the idea of decomposing the generating set of GnG_{n} into three parts: square, adjacent, and opposite, see [KKN21]. This allows us to decompose Δ1\Delta_{1} for any GnG_{n} into the sum of the appropriate parts. Denoting the adjacent part by Adjn\operatorname{Adj}_{n} for any n3n\geq 3, and the generating set for GnG_{n} by 𝒮\\altmathcal{S}_{n}, the main theorem can be stated as follows.

Theorem 1.1 (cf. Corollary 5.5).

Suppose AdjmλI|𝒮\operatorname{Adj}_{m}-\lambda I_{|\altmathcal{S}_{m}|} is a sum of squares in 𝕄|𝒮×𝒮(Gm)\mathbb{M}_{|\altmathcal{S}_{m}|\times|\altmathcal{S}_{m}|}(\mathbb{R}G_{m}). Then

Adjnn2m2λI|𝒮\\displaystyle\operatorname{Adj}_{n}-\frac{n-2}{m-2}\lambda I_{|\altmathcal{S}_{n}|}

is a sum of squares in 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}) for any nmn\geq m.

Bounding the spectral gap for the adjacent part Adj3\operatorname{Adj}_{3} of SL3()\operatorname{SL}_{3}(\mathbb{Z}), we get the following

Corollary 1.2 (cf. Corollary 6.2).

Let Δ1\Delta_{1} be the cohomological Laplacian in degree one of Gn=SLn()G_{n}=\operatorname{SL}_{n}(\mathbb{Z}). Then Δ1λI|𝒮\\Delta_{1}-\lambda I_{|\altmathcal{S}_{n}|} is a sum of squares for λ=0.217(n2)\lambda=0.217(n-2).

To our knowledge, the lower bounds for the spectral gaps we obtained are the first examples of explicit computations for Δ1\Delta_{1} for an infinite family of property (T) groups.

This article is organized as follows. In section 2 we introduce Fox derivatives which provide a setting to calculate one-degree Laplacians and define the positivity in the group ring setting in terms of sums of squares. In the next section we introduce the presentions for GnG_{n}: using the elementary matrices for Gn=SLn()G_{n}=\operatorname{SL}_{n}(\mathbb{Z}) and Nielsen transvections for Gn=SAut(Fn)G_{n}=\operatorname{SAut}(F_{n}). Next, we decompose the Laplacian in degree one into the three parts mentioned above. In section 5 we show Theorem 1.1 using the technique of symmetrization and apply this result in the section 6 to obtain a lower bound for the spectral gap of Δ1\Delta_{1} for Gn=SLn()G_{n}=\operatorname{SL}_{n}(\mathbb{Z}) whenever n3n\geq 3.

Acknowledgements

The author would like to express his sincere thanks to Prof. Dawid Kielak for many valuable suggestions concerning the induction technique. I am also greatly indebted to Dr Marek Kaluba for his collaboration on the substance of section 5.2, as well as his careful reading of the whole contents and many valuable remarks.

This research was supported by National Science Centre, Poland SONATINA 6 grant Algebraic spectral gaps in group cohomology (grant agreement no. 2022/44/C/ST1/00037).

2. Fox derivatives and sums of squares

We describe the model of Δ1\Delta_{1} which appears in Theorem 1.1. This can be achieved by means of Fox derivatives. Further details can be found in the papers of Fox and Lyndon, [Fox53, Fox54, Lyn50, LS77].

Let G=s1,,sn|r1,,rmG=\langle s_{1},\ldots,s_{n}|r_{1},\ldots,r_{m}\rangle be a finitely presented group. The Fox derivatives sj\frac{\partial}{\partial s_{j}} are the endomorphisms of Fn\mathbb{Z}F_{n}, the group ring of the free group Fn=s1,,snF_{n}=\langle s_{1},\ldots,s_{n}\rangle, given by the following conditions:

1sj\displaystyle\frac{\partial 1}{\partial s_{j}} =0,\displaystyle=0,
sisj\displaystyle\frac{\partial s_{i}}{\partial s_{j}} =δij,\displaystyle=\delta_{ij},
(uv)sj\displaystyle\frac{\partial(uv)}{\partial s_{j}} =usj+uvsj,\displaystyle=\frac{\partial u}{\partial s_{j}}+u\frac{\partial v}{\partial s_{j}},

where δij\delta_{ij} denotes the Kronecker delta. The Jacobian of GG is then the matrix J=[risj]𝕄m×n(G)J=\left[\frac{\partial r_{i}}{\partial s_{j}}\right]\in\mathbb{M}_{m\times n}(\mathbb{Z}G) obtained by quotienting the Fox derivatives into the group ring G\mathbb{Z}G. Denoting by dd the vertical vector [1sj]𝕄n×1(G)\left[1-s_{j}\right]\in\mathbb{M}_{n\times 1}(\mathbb{Z}G), the one-degree Laplacian Δ1\Delta_{1} associated to the presentation G=s1,,sn|r1,,rmG=\langle s_{1},\ldots,s_{n}|r_{1},\ldots,r_{m}\rangle is defined by

Δ1=Δ1++Δ1,\Delta_{1}=\Delta_{1}^{+}+\Delta_{1}^{-},

where Δ1+=JJ\Delta_{1}^{+}=J^{*}J and Δ1=dd\Delta_{1}^{-}=dd^{*}, and, for any matrix MM with entries in G\mathbb{Z}G, the matrix MM^{*} is the composition of the matrix transposition with the *-involution on G\mathbb{Z}G given by gg1g\mapsto g^{-1}.

Definition 2.1.

We say that a matrix M𝕄k×k(G)M\in\mathbb{M}_{k\times k}(\mathbb{R}G) is a sum of squares if there exist matrices M1,,Ml𝕄k×k(G)M_{1},\ldots,M_{l}\in\mathbb{M}_{k\times k}(\mathbb{R}G) such that M=M1M1+MlMlM=M_{1}^{*}M_{1}+\ldots M_{l}^{*}M_{l}.

It follows from [BN20, Lemma 14] that Δ1±\Delta_{1}^{\pm}, and therefore Δ1\Delta_{1} as well, are sums of squares.

By saying that a matrix M𝕄k×k(G)M\in\mathbb{M}_{k\times k}(\mathbb{R}G) has a positive spectral gap we mean that there exists a positive λ\lambda such that MλIkM-\lambda I_{k} is a sum of squares.

As noted in [KMN24, Lemma 2.1], it turns out that we have some freedom in choosing the relations which build up Δ1+\Delta_{1}^{+}. Denote

Ji=[ris1risn]J_{i}=\left[\frac{\partial r_{i}}{\partial s_{1}}\cdots\frac{\partial r_{i}}{\partial s_{n}}\right]

and put ΔI+=iIJiJi\Delta_{I}^{+}=\sum_{i\in I}J_{i}^{*}J_{i} for any subset of indices I{1,,m}I\subseteq\{1,\ldots,m\}. Then, once ΔI++Δ1𝕄n×n(G)\Delta_{I}^{+}+\Delta_{1}^{-}\in\mathbb{M}_{n\times n}(\mathbb{R}G) has a spectral gap then Δ1=ΔI++Δ{1,,m}I++Δ1𝕄n×n(G)\Delta_{1}=\Delta_{I}^{+}+\Delta_{\{1,\ldots,m\}\setminus I}^{+}+\Delta_{1}^{-}\in\mathbb{M}_{n\times n}(\mathbb{R}G) has at least the same spectral gap.

Computing lower bounds for sepctral gaps for the first Laplacians provides the bounds for the Ozawa’s spectral gap:

Remark 2.2.

Suppose the matrix CΔI++Δ1λIC\Delta_{I}^{+}+\Delta_{1}^{-}-\lambda I is a sum of squares for some subset of relator indices I{1,,m}I\subseteq\{1,\ldots,m\} and a positive constant CC and let Δ=dd\Delta=d^{*}d. Then Δ2λΔ=d(CΔI++Δ1λI)d\Delta^{2}-\lambda\Delta=d^{*}\left(C\Delta_{I}^{+}+\Delta_{1}^{-}-\lambda I\right)d is a sum of squares as well (note that dΔI+dd^{*}\Delta_{I}^{+}d is always zero). Note that in the case the generating set 𝒮\altmathcal{S} and the set of its inverses, 𝒮\altmathcal{S}^{-1}, are disjoint, then Δ\Delta is the zero degree Laplacian |𝒮𝒮|\altmathcal{S}|-\sum_{s\in\altmathcal{S}}s for the symmetric generating set 𝒮𝒮\altmathcal{S}\cup\altmathcal{S}^{-1}.

3. Presentations of SLn()\operatorname{SL}_{n}(\mathbb{Z}) and SAut(Fn)\operatorname{SAut}(F_{n})

In this section, we introduce the presentations of SLn()\operatorname{SL}_{n}(\mathbb{Z}) and SAut(Fn)\operatorname{SAut}(F_{n}) which we shall use to perform the Fox calculus and prove property (T) for these groups. The presentations can be found in [CRW92] and [Ger84] for SLn()\operatorname{SL}_{n}(\mathbb{Z}) and SAut(Fn)\operatorname{SAut}(F_{n}) respectively. For the latter, we provide an equivalent presentation which is easy to obtain from [Ger84].

The presentation for SLn()\operatorname{SL}_{n}(\mathbb{Z}) is given below.

SLn()=Eij,1ijn|\displaystyle\operatorname{SL}_{n}(\mathbb{Z})=\big{\langle}E_{ij},1\leq i\neq j\leq n\big{|} [Eij,Ekl],il,jk,(i,j)(k,l),\displaystyle[E_{ij},E_{kl}],i\neq l,j\neq k,(i,j)\neq(k,l),
[Eij,Ejk]Eik1 for distinct i,j,k,\displaystyle[E_{ij},E_{jk}]E_{ik}^{-1}\text{ for distinct }i,j,k,
(E12E211E12)4.\displaystyle\left(E_{12}E_{21}^{-1}E_{12}\right)^{4}\big{\rangle}.

The generator EijE_{ij} represents the elementary matrix with 11 at the (i,j)(i,j)-th entry and the diagonal and zeroes elsewhere. Using Remark 2.2, we shall exclude the relation (E12E211E12)4\left(E_{12}E_{21}^{-1}E_{12}\right)^{4} from further considerations.

The generators of SAut(Fn)\operatorname{SAut}(F_{n}) are denoted by λij\lambda_{ij} and ρij\rho_{ij} for 1ijn1\leq i\neq j\leq n and are called left and right Nielsen transvections respectively. The relator set contains commutators and pentagonal relations as well, along with other relations which we shall not consider, again by Remark 2.2. We only present the commutator and pentagonal relations therefore. These relations are defined for distinct indices i,j,k,l{1,,n}i,j,k,l\in\{1,\ldots,n\} as follows:

[λi,j,ρij],[λij,λkl],[ρij,ρkl],[λij,ρkl],\displaystyle[\lambda_{i,j},\rho_{ij}],\quad[\lambda_{ij},\lambda_{kl}],\quad[\rho_{ij},\rho_{kl}],\quad[\lambda_{ij},\rho_{kl}],
[λij,λkj],[ρij,ρkj],[λij,ρik],[λi,j,ρkj],\displaystyle\quad[\lambda_{ij},\lambda_{kj}],\quad[\rho_{ij},\rho_{kj}],\quad[\lambda_{ij},\rho_{ik}],\quad[\lambda_{i,j},\rho_{kj}],
λik±[λjk±,λij1],ρik±[ρjk±,ρij1],λik±[ρjk,λij],ρik±[λjk,ρij].\displaystyle\quad\lambda_{ik}^{\pm}[\lambda_{jk}^{\pm},\lambda_{ij}^{-1}],\quad\rho_{ik}^{\pm}[\rho_{jk}^{\pm},\rho_{ij}^{-1}],\quad\lambda_{ik}^{\pm}[\rho_{jk}^{\mp},\lambda_{ij}],\quad\rho_{ik}^{\pm}[\lambda_{jk}^{\mp},\rho_{ij}].

The transvection generators correspond in SAut(Fn)\operatorname{SAut}(F_{n}) to the following automorphisms of Fn=s1,,snF_{n}=\langle s_{1},\ldots,s_{n}\rangle111We take in SAut(Fn)\operatorname{SAut}(F_{n}) the group operation order given by composition of automorphisms, i.e. αβ=αβ\alpha*\beta=\alpha\circ\beta. Note that this is different than the convention of Gersten [Ger84]. In his paper the convention is: αβ=βα\alpha*\beta=\beta\circ\alpha. We had to reverse the order in pentagonal relators for that account.:

λij(sk)={sjsi if k=i,skotherwise,ρij(sk)={sisj if k=i,skotherwise.\displaystyle\lambda_{ij}(s_{k})=\begin{cases}s_{j}s_{i}&\text{ if $k=i$,}\\ s_{k}&\text{otherwise,}\end{cases}\quad\rho_{ij}(s_{k})=\begin{cases}s_{i}s_{j}&\text{ if $k=i$,}\\ s_{k}&\text{otherwise.}\end{cases}

4. Decomposition into square, adjacent and opposite part

Let n3n\geq 3 be an integer and GnG_{n} be SLn()\operatorname{SL}_{n}(\mathbb{Z}) or SAut(Fn)\operatorname{SAut}(F_{n}). We introduce the following decompositions of the Laplacians Δ1±\Delta_{1}^{\pm} of GnG_{n}:

Δ1±=Sqn±+Adjn±+Opn±.\displaystyle\Delta_{1}^{\pm}=\operatorname{Sq}_{n}^{\pm}+\operatorname{Adj}_{n}^{\pm}+\operatorname{Op}_{n}^{\pm}.

Our decomposition is inspired by [KKN21, pp. 543 – 545].

The generators of GnG_{n} take form ωij\omega_{ij} for 1ijn1\leq i\neq j\leq n and ω\omega denoting either elementary matrices or Nielsen transvections. Then, for any 3mn3\leq m\leq n, there exists a natural embedding GmGnG_{m}\hookrightarrow G_{n} given by ωijωij\omega_{ij}\mapsto\omega_{ij}. From now on, when talking about the relations of GnG_{n}, we shall restrict ourselves to commutator and pentagonal relations only.

Let CnC_{n} be the (n1)(n-1)-simplex with the vertices {1,,n}\{1,\ldots,n\} and let F(Cn)F(C_{n}) denote the set of its faces, that is the subsets of the set {1,,n}\{1,\ldots,n\}. Denoting by 𝒮\\altmathcal{S}_{n} the generator set of GnG_{n} and by Free(𝒮\\operatorname{Free}(\altmathcal{S}_{n}) the free group on 𝒮\\altmathcal{S}_{n}, one can define the map

ϕ:Free(𝒮\𝒞\\displaystyle\phi:\operatorname{Free}(\altmathcal{S}_{n})\longrightarrow F(C_{n})

by sending an element of Free(𝒮\\operatorname{Free}(\altmathcal{S}_{n}) to the face on the indices of the generators occurring in that element, i.e.

ϕ(ωi1j1ωikjk)={i1,j1,,ik,jk}.\phi(\omega_{i_{1}j_{1}}\ldots\omega_{i_{k}j_{k}})=\{i_{1},j_{1},\ldots,i_{k},j_{k}\}.

In order to define Sqn±\operatorname{Sq}_{n}^{\pm}, Adjn±\operatorname{Adj}_{n}^{\pm} and Opn±\operatorname{Op}_{n}^{\pm} we distinguish from F(Cn)F(C_{n}) the following three subsets of faces

Edgn\displaystyle\operatorname{Edg}_{n} ={fF(Cn)||f|=2},\displaystyle=\{f\in F(C_{n})||f|=2\},
Trin\displaystyle\operatorname{Tri}_{n} ={fF(Cn)||f|=3},\displaystyle=\{f\in F(C_{n})||f|=3\},
Tetn\displaystyle\operatorname{Tet}_{n} ={fF(Cn)||f|=4},\displaystyle=\{f\in F(C_{n})||f|=4\},

that is, Edgn\operatorname{Edg}_{n}, Trin\operatorname{Tri}_{n} and Tetn\operatorname{Tet}_{n} constitute respectively the sets of edges, triangles, and tetrahedra of CnC_{n}.

The decomposition Δ1=Sqn+Adjn+Opn\Delta_{1}^{-}=\operatorname{Sq}_{n}^{-}+\operatorname{Adj}_{n}^{-}+\operatorname{Op}_{n}^{-} is obtained as follows:

(Sqn)s,t\displaystyle\left(\operatorname{Sq}^{-}_{n}\right)_{s,t} =(Δ1)s,t for ϕ(s)ϕ(t)Edgn and 0 otherwise,\displaystyle=\left(\Delta^{-}_{1}\right)_{s,t}\text{ for }\phi(s)\cup\phi(t)\in\operatorname{Edg}_{n}\text{ and }0\text{ otherwise},
(Adjn)s,t\displaystyle\left(\operatorname{Adj}^{-}_{n}\right)_{s,t} =(Δ1)s,t for ϕ(s)ϕ(t)Trin and 0 otherwise,\displaystyle=\left(\Delta^{-}_{1}\right)_{s,t}\text{ for }\phi(s)\cup\phi(t)\in\operatorname{Tri}_{n}\text{ and }0\text{ otherwise},
(Opn)s,t\displaystyle\left(\operatorname{Op}^{-}_{n}\right)_{s,t} =(Δ1)s,t for ϕ(s)ϕ(t)Tetn and 0 otherwise.\displaystyle=\left(\Delta^{-}_{1}\right)_{s,t}\text{ for }\phi(s)\cup\phi(t)\in\operatorname{Tet}_{n}\text{ and }0\text{ otherwise}.

We remark that the group ring elements dSqndd^{*}\operatorname{Sq}^{-}_{n}d, dAdjndd^{*}\operatorname{Adj}^{-}_{n}d, and dOpndd^{*}\operatorname{Op}^{-}_{n}d are equal to the elements Sqn\operatorname{Sq}_{n}, Adjn\operatorname{Adj}_{n}, and Opn\operatorname{Op}_{n} as defined in [KKN21].

The decomposition Δ1+=Sqn++Adjn++Opn+\Delta_{1}^{+}=\operatorname{Sq}^{+}_{n}+\operatorname{Adj}^{+}_{n}+\operatorname{Op}^{+}_{n} is defined, in turn, by summing over specific relators of GnG_{n}.

(Sqn+)s,t\displaystyle\left(\operatorname{Sq}^{+}_{n}\right)_{s,t} =ϕ(r)Edgn(rs)rt,\displaystyle=\sum_{\phi(r)\in\operatorname{Edg}_{n}}\left(\frac{\partial r}{\partial s}\right)^{*}\frac{\partial r}{\partial t},
(Adjn+)s,t\displaystyle\left(\operatorname{Adj}^{+}_{n}\right)_{s,t} =ϕ(r)Trin(rs)rt,\displaystyle=\sum_{\phi(r)\in\operatorname{Tri}_{n}}\left(\frac{\partial r}{\partial s}\right)^{*}\frac{\partial r}{\partial t},
(Opn+)s,t\displaystyle\left(\operatorname{Op}^{+}_{n}\right)_{s,t} =ϕ(r)Tetn(rs)rt.\displaystyle=\sum_{\phi(r)\in\operatorname{Tet}_{n}}\left(\frac{\partial r}{\partial s}\right)^{*}\frac{\partial r}{\partial t}.

Note that (Adjn+)s,t\left(\operatorname{Adj}^{+}_{n}\right)_{s,t} and (Opn+)s,t\left(\operatorname{Op}^{+}_{n}\right)_{s,t} can be non-zero only if s=ts=t or ϕ(s)ϕ(t)\phi(s)\cup\phi(t) belongs to Trin\operatorname{Tri}_{n} and Tetn\operatorname{Tet}_{n} respectively. We shall call Sqn±\operatorname{Sq}_{n}^{\pm}, Adjn±\operatorname{Adj}_{n}^{\pm} and Opn±\operatorname{Op}_{n}^{\pm} the square, adjacent and opposite part respectively.

Example 4.1.

Let n=3n=3 and let us focus on G3=SL3()G_{3}=\operatorname{SL}_{3}(\mathbb{Z}). We have six generators of G3G_{3}, namely the elementary matrices E12E_{12}, E21E_{21}, E13E_{13}, E31E_{31}, E23E_{23}, and E32E_{32}. The relator set RR we take into account contains the following 1818 relators:

[E12,E13],[E13,E12],[E12,E32],[E32,E12],\displaystyle[E_{12},E_{13}],[E_{13},E_{12}],[E_{12},E_{32}],[E_{32},E_{12}],
[E13,E23],[E23,E13],[E21,E23],[E23,E21],\displaystyle[E_{13},E_{23}],[E_{23},E_{13}],[E_{21},E_{23}],[E_{23},E_{21}],
[E21,E31],[E31,E21],[E31,E32],[E32,E31],\displaystyle[E_{21},E_{31}],[E_{31},E_{21}],[E_{31},E_{32}],[E_{32},E_{31}],
[E12,E23]E131,[E13,E32]E121,[E21,E13]E231,\displaystyle[E_{12},E_{23}]E_{13}^{-1},[E_{13},E_{32}]E_{12}^{-1},[E_{21},E_{13}]E_{23}^{-1},
[E23,E31]E211,[E31,E12]E321,[E32,E21]E311.\displaystyle[E_{23},E_{31}]E_{21}^{-1},[E_{31},E_{12}]E_{32}^{-1},[E_{32},E_{21}]E_{31}^{-1}.

The simplex C3C_{3} is a triangle. Its faces contain three edges – more precisely, Edg3={{1,2},{1,3},{2,3}}\operatorname{Edg}_{3}=\{\{1,2\},\{1,3\},\{2,3\}\}, and only one triangle, i.e. Tri3={{1,2,3}}\operatorname{Tri}_{3}=\{\{1,2,3\}\}. There are no tetrahedra in C3C_{3}.

112233{E13,E31}\{E_{13},E_{31}\}{E23,E32}\{E_{23},E_{32}\}{E12,E21}\{E_{12},E_{21}\}
Figure 1. The simplex C3C_{3} with the assignments of the generators of SL3()\operatorname{SL}_{3}(\mathbb{Z}) to its edges given by ϕ\phi.

The function ϕ\phi takes the following values on the generators (the situation depicted in Figure 1):

ϕ(E12)=ϕ(E21)={1,2},ϕ(E13)=ϕ(E31)={1,3},ϕ(E23)=ϕ(E32)={2,3},\displaystyle\phi(E_{12})=\phi(E_{21})=\{1,2\},\quad\phi(E_{13})=\phi(E_{31})=\{1,3\},\quad\phi(E_{23})=\phi(E_{32})=\{2,3\},

and is constant on the relators – for every rRr\in R one has ϕ(r)={1,2,3}\phi(r)=\{1,2,3\}. Thus, denoting by (i,j)(k,l)(i,j)(k,l) the element (1Eij)(1Ekl)(1-E_{ij})(1-E_{kl})^{*}, we get

Sq3=[(1,2)(1,2)0(1,2)(2,1)0000(1,3)(1,3)00(1,3)(3,1)0(2,1)(1,2)0(2,1)(2,1)000000(2,3)(2,3)0(2,3)(3,2)0(3,1)(1,3)00(3,1)(3,1)0000(3,2)(2,3)0(3,2)(3,2)]\displaystyle\operatorname{Sq}_{3}^{-}=\begin{bmatrix}(1,2)(1,2)&0&(1,2)(2,1)&0&0&0\\ 0&(1,3)(1,3)&0&0&(1,3)(3,1)&0\\ (2,1)(1,2)&0&(2,1)(2,1)&0&0&0\\ 0&0&0&(2,3)(2,3)&0&(2,3)(3,2)\\ 0&(3,1)(1,3)&0&0&(3,1)(3,1)&0\\ 0&0&0&(3,2)(2,3)&0&(3,2)(3,2)\end{bmatrix}

and

Adj3=[0(1,2)(1,3)0(1,2)(2,3)(1,2)(3,1)(1,3)(3,2)(1,3)(1,2)0(1,3)(2,1)(1,3)(2,3)0(1,3)(3,2)0(2,1)(1,3)0(2,1)(2,3)(2,1)(3,1)(2,1)(3,2)(2,3)(1,2)(2,3)(1,3)(2,3)(2,1)0(2,3)(3,1)0(3,1)(1,2)0(3,1)(2,1)(3,1)(2,3)0(3,1)(3,2)(3,2)(1,2)(3,2)(1,3)(3,2)(2,1)0(3,2)(3,1)0].\displaystyle\operatorname{Adj}_{3}^{-}=\begin{bmatrix}0&(1,2)(1,3)&0&(1,2)(2,3)&(1,2)(3,1)&(1,3)(3,2)\\ (1,3)(1,2)&0&(1,3)(2,1)&(1,3)(2,3)&0&(1,3)(3,2)\\ 0&(2,1)(1,3)&0&(2,1)(2,3)&(2,1)(3,1)&(2,1)(3,2)\\ (2,3)(1,2)&(2,3)(1,3)&(2,3)(2,1)&0&(2,3)(3,1)&0\\ (3,1)(1,2)&0&(3,1)(2,1)&(3,1)(2,3)&0&(3,1)(3,2)\\ (3,2)(1,2)&(3,2)(1,3)&(3,2)(2,1)&0&(3,2)(3,1)&0\end{bmatrix}.

It is apparent that Op3\operatorname{Op}_{3}^{-} is the zero matrix. Since ϕ\phi equals the face {1,2,3}\{1,2,3\} on all relators, we have

(Adj3+)s,t=rR(rs)(rt)=(Δ1+)s,t\left(\operatorname{Adj}_{3}^{+}\right)_{s,t}=\sum_{r\in R}\left(\frac{\partial r}{\partial s}\right)^{*}\left(\frac{\partial r}{\partial t}\right)=\left(\Delta_{1}^{+}\right)_{s,t}

for any two generators s,ts,t. In particular, Sq3+\operatorname{Sq}_{3}^{+} and Op3+\operatorname{Op}_{3}^{+} are the zero matrices.

Example 4.2.

This time, let n=4n=4 and G4=SAut(F4)G_{4}=\operatorname{SAut}(F_{4}). The matrices get significantly bigger for this case and for this account we shall restrict ourselves to the demonstration of the function ϕ\phi only. The following 2424 transvections consitute the generators of G4G_{4}:

λ12,λ13,λ14,λ21,λ23,λ24,λ31,λ32,λ34,λ41,λ42,λ43,\displaystyle\lambda_{12},\lambda_{13},\lambda_{14},\lambda_{21},\lambda_{23},\lambda_{24},\lambda_{31},\lambda_{32},\lambda_{34},\lambda_{41},\lambda_{42},\lambda_{43},
ρ12,ρ13,ρ14,ρ21,ρ23,ρ24,ρ31,ρ32,ρ34,ρ41,ρ42,ρ43.\displaystyle\rho_{12},\rho_{13},\rho_{14},\rho_{21},\rho_{23},\rho_{24},\rho_{31},\rho_{32},\rho_{34},\rho_{41},\rho_{42},\rho_{43}.

Since there are already relatively many relations of G4G_{4}, we will not list them all. Instead, we comment on the values that the function ϕ\phi takes on them. The simplex C4C_{4} is a tetrahderon. The faces of C4C_{4} contain six edges, four triangles and one tetrahedron:

Edg4={{1,2},{1,3},{1,4},{2,3},{2,4},{3,4}}\displaystyle\operatorname{Edg}_{4}=\{\{1,2\},\{1,3\},\{1,4\},\{2,3\},\{2,4\},\{3,4\}\}
Tri4={{1,2,3},{1,2,4},{1,3,4},{2,3,4}}\displaystyle\operatorname{Tri}_{4}=\{\{1,2,3\},\{1,2,4\},\{1,3,4\},\{2,3,4\}\}
Tet4={{1,2,3,4}}.\displaystyle\operatorname{Tet}_{4}=\{\{1,2,3,4\}\}.

The function ϕ\phi takes the following values on the generators (see Figure 2):

ϕ(λ12)=ϕ(λ21)=ϕ(ρ12)=ϕ(ρ21)={1,2},\displaystyle\phi(\lambda_{12})=\phi(\lambda_{21})=\phi(\rho_{12})=\phi(\rho_{21})=\{1,2\},
ϕ(λ13)=ϕ(λ31)=ϕ(ρ13)=ϕ(ρ31)={1,3},\displaystyle\phi(\lambda_{13})=\phi(\lambda_{31})=\phi(\rho_{13})=\phi(\rho_{31})=\{1,3\},
ϕ(λ14)=ϕ(λ41)=ϕ(ρ14)=ϕ(ρ41)={1,4},\displaystyle\phi(\lambda_{14})=\phi(\lambda_{41})=\phi(\rho_{14})=\phi(\rho_{41})=\{1,4\},
ϕ(λ23)=ϕ(λ32)=ϕ(ρ23)=ϕ(ρ32)={2,3},\displaystyle\phi(\lambda_{23})=\phi(\lambda_{32})=\phi(\rho_{23})=\phi(\rho_{32})=\{2,3\},
ϕ(λ24)=ϕ(λ42)=ϕ(ρ24)=ϕ(ρ42)={2,4},\displaystyle\phi(\lambda_{24})=\phi(\lambda_{42})=\phi(\rho_{24})=\phi(\rho_{42})=\{2,4\},
ϕ(λ34)=ϕ(λ43)=ϕ(ρ34)=ϕ(ρ43)={3,4}.\displaystyle\phi(\lambda_{34})=\phi(\lambda_{43})=\phi(\rho_{34})=\phi(\rho_{43})=\{3,4\}.
11223344{λ12,λ21,ρ12,ρ21}\{\lambda_{12},\lambda_{21},\rho_{12},\rho_{21}\}{λ13,λ31,ρ13,ρ31}\{\lambda_{13},\lambda_{31},\rho_{13},\rho_{31}\}{λ14,λ41,ρ14,ρ41}\{\lambda_{14},\lambda_{41},\rho_{14},\rho_{41}\}{λ23,λ32,ρ23,ρ32}\{\lambda_{23},\lambda_{32},\rho_{23},\rho_{32}\}{λ24,λ42,ρ24,ρ42}\{\lambda_{24},\lambda_{42},\rho_{24},\rho_{42}\}{λ34,λ43,ρ34,ρ43}\{\lambda_{34},\lambda_{43},\rho_{34},\rho_{43}\}
Figure 2. The simplex C4C_{4} with the assignments of the generators of SAut(F4)\operatorname{SAut}(F_{4}) to its edges given by ϕ\phi.

Let i,j,k,l4\leq i,j,k,l\leq 4 be pairwise distinct. The values of ϕ\phi on the relators are the following:

ϕ([λij,ρij])={i,j},ϕ([λij,λkl])=ϕ(ρij,ρkl)=ϕ([λij,ρkl])={1,2,3,4},\displaystyle\phi([\lambda_{ij},\rho_{ij}])=\{i,j\},\quad\phi([\lambda_{ij},\lambda_{kl}])=\phi(\rho_{ij},\rho_{kl})=\phi([\lambda_{ij},\rho_{kl}])=\{1,2,3,4\},
ϕ([λij,λkj])=ϕ([ρij,ρkj])=ϕ([λij,ρik])=ϕ([λi,j,ρkj])\displaystyle\phi([\lambda_{ij},\lambda_{kj}])=\phi([\rho_{ij},\rho_{kj}])=\phi([\lambda_{ij},\rho_{ik}])=\phi([\lambda_{i,j},\rho_{kj}])
 =ϕ(λik±[λjk±,λij1])=ϕ(ρik±[ρjk±,ρij1])\displaystyle\qquad\qquad\quad\text{ }=\phi(\lambda_{ik}^{\pm}[\lambda_{jk}^{\pm},\lambda_{ij}^{-1}])=\phi(\rho_{ik}^{\pm}[\rho_{jk}^{\pm},\rho_{ij}^{-1}])
 =ϕ(λik±[ρjk,λij])=ϕ(ρik±[λjk,ρij])={i,j,k}.\displaystyle\qquad\qquad\quad\text{ }=\phi(\lambda_{ik}^{\pm}[\rho_{jk}^{\mp},\lambda_{ij}])=\phi(\rho_{ik}^{\pm}[\lambda_{jk}^{\mp},\rho_{ij}])=\{i,j,k\}.

4.1. Sum of squares decomposition for square and opposite part

It turns out that the square and opposite parts behave well with respect to sum of squares decompositions. The precise statement is given below.

Lemma 4.3.

The matrices Sqn\operatorname{Sq}^{-}_{n} and Opn++2Opn\operatorname{Op}^{+}_{n}+2\operatorname{Op}^{-}_{n} are sums of squares.

Proof.

For any 1ijn1\leq i\neq j\leq n, define di,j𝕄|𝒮\×(Gn)d^{i,j}\in\mathbb{M}_{|\altmathcal{S}_{n}|\times 1}(\mathbb{R}G_{n}) to be the column vector indexed by the generators of GnG_{n} with the only non-zero entries being 1s1-s, indexed by the generators ss satisfying ϕ(s)={i,j}\phi(s)=\{i,j\}. That Sqn\operatorname{Sq}^{-}_{n} is a sums of squares follows from the decomposition below:

Sqn=1ijndi,j(di,j).\operatorname{Sq}^{-}_{n}=\sum_{1\leq i\neq j\leq n}d^{i,j}\left(d^{i,j}\right)^{*}.

We have already observed that both Opn+\operatorname{Op}^{+}_{n} and Opn\operatorname{Op}^{-}_{n} (in this case, just by the definition) have vanishing (s,t)(s,t)-entries whenever ϕ(s)ϕ(t)Trin\phi(s)\cup\phi(t)\in\operatorname{Tri}_{n} or the generator tt has the same indices as ss but with the reversed order. In addition, the diagonal vanishes for Opn\operatorname{Op}^{-}_{n} as well. This is not the case for Opn+\operatorname{Op}^{+}_{n}. However, since Opn+\operatorname{Op}^{+}_{n} is a sum of squares by its definition, it follows that the same applies to its diagonal part. On the other hand, the non-diagonal parts of Opn+\operatorname{Op}^{+}_{n} and 2Opn2\operatorname{Op}^{-}_{n} cancel out – for distinct 1i,j,k,ln1\leq i,j,k,l\leq n we have:

(Opn+)αij,βkl\displaystyle\left(\operatorname{Op}^{+}_{n}\right)_{\alpha_{ij},\beta_{kl}} =([αij,βkl]αij)[αij,βkl]βkl+([βkl,αij]αij)[βkl,αij]βkl\displaystyle=\left(\frac{\partial[\alpha_{ij},\beta_{kl}]}{\partial\alpha_{ij}}\right)^{*}\frac{\partial[\alpha_{ij},\beta_{kl}]}{\partial\beta_{kl}}+\left(\frac{\partial[\beta_{kl},\alpha_{ij}]}{\partial\alpha_{ij}}\right)^{*}\frac{\partial[\beta_{kl},\alpha_{ij}]}{\partial\beta_{kl}}
=2(1αij)(1βkl).\displaystyle=-2(1-\alpha_{ij})(1-\beta_{kl})^{*}.

5. Symmetrization of the adjacent part

In this section we show how to get the lower bound for spectral gap of Adjn=Adjn++Adjn\operatorname{Adj}_{n}=\operatorname{Adj}_{n}^{+}+\operatorname{Adj}_{n}^{-} for GnG_{n} once we assume that the lower bound for the corresponding spectral gap for GmG_{m} is known for some nm3n\geq m\geq 3. Our method is inspired by the symmetrization method used in [KKN21].

5.1. General setting for symmetrization

Note that the symmetric group Symk\operatorname{Sym}_{k} acts by automorphisms on GkG_{k} by sending the generator ωi,j\omega_{i,j} to ωσ(i),σ(j)\omega_{\sigma(i),\sigma(j)} for any permutation σSymk\sigma\in\operatorname{Sym}_{k}. This induces an action of Symk\operatorname{Sym}_{k} on Gk\mathbb{R}G_{k} by automorphisms which preserves the *-involution. Using this action we can define the action on 𝕄|𝒮×𝒮(Gk)\mathbb{M}_{|\altmathcal{S}_{k}|\times|\altmathcal{S}_{k}|}(\mathbb{R}G_{k}) by the formula

(σA)s,t=σ(Aσ1(s),σ1(t)),\left(\sigma A\right)_{s,t}=\sigma\left(A_{\sigma^{-1}(s),\sigma^{-1}(t)}\right),

for any matrix AA over Gk\mathbb{R}G_{k} indexed by the generators 𝒮\altmathcal{S}_{k}. This is also the action by *-algebra automorphisms (the multiplication being the group ring matrix multiplication). Thus, for any matrix A𝕄|𝒮×𝒮(Gk)A\in\mathbb{M}_{|\altmathcal{S}_{k}|\times|\altmathcal{S}_{k}|}(\mathbb{R}G_{k}) and σSymk\sigma\in\operatorname{Sym}_{k}, the matrix σA\sigma A is a sum of squares provided AA was a sum of squares.

5.2. Invariance of Laplacians

Let us show first that the adjacent Laplacians Adjk+\operatorname{Adj}_{k}^{+} and Adjk\operatorname{Adj}^{-}_{k} for GkG_{k} are Symk\operatorname{Sym}_{k}-invariant. The statement for Adjk\operatorname{Adj}^{-}_{k} is straightforward:

(σAdjk)s,t\displaystyle\left(\sigma\operatorname{Adj}_{k}^{-}\right)_{s,t} =σ((Adjk)σ1(s),σ1(t))=σ((1σ1(s))(1σ1(t)))\displaystyle=\sigma\left(\left(\operatorname{Adj}^{-}_{k}\right)_{\sigma^{-1}(s),\sigma^{-1}(t)}\right)=\sigma\left(\left(1-\sigma^{-1}(s)\right)\left(1-\sigma^{-1}(t)\right)^{*}\right)
=(1s)(1t)=(Adjk)s,t.\displaystyle=(1-s)(1-t)^{*}=\left(\operatorname{Adj}^{-}_{k}\right)_{s,t}.

Let us show the invariance of Adjk+\operatorname{Adj}_{k}^{+}. Let a=|𝒮a=|\altmathcal{S}_{k}| be the number of generators of GkG_{k} and bb be the number of its relations rr such that ϕ(r)Trik\phi(r)\in\operatorname{Tri}_{k}. Suppose moreover that this set of relations is invariant with respect to the action of Symk\operatorname{Sym}_{k}, that is, for all σSymk\sigma\in\operatorname{Sym}_{k} and rr satisfying ϕ(r)Trik\phi(r)\in\operatorname{Tri}_{k}, the word σ(r)\sigma(r) is another relator. Notice that each permutation σ\sigma from Symk\operatorname{Sym}_{k} defines permutations of {1,,a}\{1,\ldots,a\} and {1,,b}\{1,\ldots,b\} (we fix the order of generators 𝒮\altmathcal{S}_{k} and relations rr such that ϕ(r)Trik\phi(r)\in\operatorname{Tri}_{k}). Having this, we can endow both products (Gk)a\left(\mathbb{R}G_{k}\right)^{a} and (Gk)b\left(\mathbb{R}G_{k}\right)^{b} with the action of Symk\operatorname{Sym}_{k} as follows:

σ(ξ1,,ξa)\displaystyle\sigma\left(\xi_{1},\ldots,\xi_{a}\right) =(σξσ1(1),,σξσ1(a)),\displaystyle=\left(\sigma\xi_{\sigma^{-1}(1)},\ldots,\sigma\xi_{\sigma^{-1}(a)}\right),
σ(ξ1,,ξb)\displaystyle\sigma\left(\xi_{1},\ldots,\xi_{b}\right) =(σξσ1(1),,σξσ1(b)).\displaystyle=\left(\sigma\xi_{\sigma^{-1}(1)},\ldots,\sigma\xi_{\sigma^{-1}(b)}\right).

Finally, denote by JJ the Jacobian map from (Gk)a\left(\mathbb{R}G_{k}\right)^{a} to (Gk)b\left(\mathbb{R}G_{k}\right)^{b} given by the relations rr such that ϕ(r)Trik\phi(r)\in\operatorname{Tri}_{k}, that is Adjk+=JJ\operatorname{Adj}_{k}^{+}=J^{*}J. It turns out that the invariance of Adjk+\operatorname{Adj}_{k}^{+} follows from the equivariance of JJ:

Lemma 5.1.

For any σSymk\sigma\in\operatorname{Sym}_{k} we have σAdjk+=Adjk+\sigma\operatorname{Adj}_{k}^{+}=\operatorname{Adj}_{k}^{+} provided JJ is equivariant.

Proof.

Suppose JJ is equivariant. Let σSymk\sigma\in\operatorname{Sym}_{k}. Since we have ordered the generators in 𝒮\altmathcal{S}_{k}, we can identify them with the corresponding indices from the set {1,,a}\{1,\ldots,a\}. Thus, we must show that (σAdjk+)i,j=(Adjk+)i,j\left(\sigma\operatorname{Adj}_{k}^{+}\right)_{i,j}=\left(\operatorname{Adj}_{k}^{+}\right)_{i,j} for any 1i,ja1\leq i,j\leq a.

It is easy to check that the equivariance of JJ is equivalent to the following condition: Jy,σ1(x)=σ1Jσ(y),xJ_{y,\sigma^{-1}(x)}=\sigma^{-1}J_{\sigma(y),x} for any 1xa1\leq x\leq a and 1yb1\leq y\leq b. Using this relationship, we get

(σAdjk+)i,j\displaystyle\left(\sigma\operatorname{Adj}_{k}^{+}\right)_{i,j} =σ(l(Jl,σ1(i))Jl,σ1(j))=σ(lσ1((Jσ(l),i)Jσ(l),j))\displaystyle=\sigma\left(\sum_{l}\left(J_{l,\sigma^{-1}(i)}\right)^{*}J_{l,\sigma^{-1}(j)}\right)=\sigma\left(\sum_{l}\sigma^{-1}\left(\left(J_{\sigma(l),i}\right)^{*}J_{\sigma(l),j}\right)\right)
=l(Jσ(l),i)Jσ(l),j=(Adjk+)i,j.\displaystyle=\sum_{l}\left(J_{\sigma(l),i}\right)^{*}J_{\sigma(l),j}=\left(\operatorname{Adj}_{k}^{+}\right)_{i,j}.

The following lemma is the key tool to prove the equivariance of JJ.

Lemma 5.2.

For any σSymN\sigma\in\operatorname{Sym}_{N}, s{s1,,sN}s\in\{s_{1},\ldots,s_{N}\} and rr a word in the free group FNF_{N} generated by {s1,,sN}\{s_{1},\ldots,s_{N}\}, the following holds:

r(σ(s))=σ((σ1(r))s),\frac{\partial r}{\partial(\sigma(s))}=\sigma\left(\frac{\partial\left(\sigma^{-1}(r)\right)}{\partial s}\right),

where the action of SymN\operatorname{Sym}_{N} on FNF_{N} is given by τ(si)=sτ(i)\tau(s_{i})=s_{\tau(i)} for any τSymN\tau\in\operatorname{Sym}_{N}.

Proof.

We prove the assertion by induction on the word length of the relation rr. Note first that both sides of the equation in question vanish for r=1r=1. Suppose rr is a single generator ss^{\prime} (if rr is the inverse of a generator, the proof is analogous). In the case σ(s){s,(s)1}\sigma(s)\notin\{s^{\prime},(s^{\prime})^{-1}\} both sides of the equation in question vanish. If σ(s)=s\sigma(s)=s^{\prime}, then

r(σ(s))=ss=1=σ(ss)=σ((σ1(r))s),\frac{\partial r}{\partial(\sigma(s))}=\frac{\partial s^{\prime}}{\partial s^{\prime}}=1=\sigma\left(\frac{\partial s}{\partial s}\right)=\sigma\left(\frac{\partial\left(\sigma^{-1}(r)\right)}{\partial s}\right),

while if σ(s)=(s)1\sigma(s)=(s^{\prime})^{-1}, then

r(σ(s))=s(s)1=s=σ(s1s)=σ((σ1(r))s).\frac{\partial r}{\partial(\sigma(s))}=\frac{\partial s^{\prime}}{\partial(s^{\prime})^{-1}}=-s^{\prime}=\sigma\left(\frac{\partial s^{-1}}{\partial s}\right)=\sigma\left(\frac{\partial\left(\sigma^{-1}(r)\right)}{\partial s}\right).

Suppose now the assertion follows for all words of length k1k\geq 1. Take a word rr of length k+1k+1. Then r=uvr=uv for some words uu and vv of length at most kk. Using the induction assumption, we get

r(σ(s))\displaystyle\frac{\partial r}{\partial(\sigma(s))} =u(σ(s))+uv(σ(s))\displaystyle=\frac{\partial u}{\partial(\sigma(s))}+u\frac{\partial v}{\partial(\sigma(s))}
=σ((σ1(u))s+σ1(u)(σ1(v))s)\displaystyle=\sigma\left(\frac{\partial\left(\sigma^{-1}(u)\right)}{\partial s}+\sigma^{-1}(u)\frac{\partial\left(\sigma^{-1}(v)\right)}{\partial s}\right)
=σ((σ1(uv))s)=σ((σ1(r))s).\displaystyle=\sigma\left(\frac{\partial\left(\sigma^{-1}(uv)\right)}{\partial s}\right)=\sigma\left(\frac{\partial\left(\sigma^{-1}(r)\right)}{\partial s}\right).

Corollary 5.3.

The Jacobian homomorphism J:(G)a(G)bJ:(\mathbb{R}G)^{a}\rightarrow(\mathbb{R}G)^{b} is an equivariant map.

Proof.

By Lemma 5.2, for any 1xa1\leq x\leq a and 1yb1\leq y\leq b, we get

Jy,σ1(x)=(ry)(σ1(sx))=σ1((σ(ry))sx)=σ1Jσ(y),x.\displaystyle J_{y,\sigma^{-1}(x)}=\frac{\partial(r_{y})}{\partial\left(\sigma^{-1}(s_{x})\right)}=\sigma^{-1}\left(\frac{\partial\left(\sigma(r_{y})\right)}{\partial s_{x}}\right)=\sigma^{-1}J_{\sigma(y),x}.

5.3. Adjacent part symmetrizes well

In this section we show how to express Adjn±\operatorname{Adj}_{n}^{\pm} in terms of Adjm±\operatorname{Adj}_{m}^{\pm} and the action of the symmetric group Symn\operatorname{Sym}_{n}, for any nm3n\geq m\geq 3. This allows us to induce spectral gaps for Adj\operatorname{Adj} parts (cf. Corollary 5.5).

For any n3n\geq 3 and θTrin\theta\in\operatorname{Tri}_{n}, we denote by Adjθ±𝕄|𝒮\×𝒮\(Gn)\operatorname{Adj}^{\pm}_{\theta}\in\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}) the embedding of the adjacent part given by the traingle θ\theta of CnC_{n}, i.e., for any generators ss and tt of GnG_{n}, we have

(Adjθ)s,t={(1s)(1t)if ϕ(s)ϕ(t)=θ0if ϕ(s)ϕ(t)θ,(Adjθ+)s,t=ϕ(r)=θ(rs)rt.\displaystyle\left(\operatorname{Adj}_{\theta}^{-}\right)_{s,t}=\begin{cases}(1-s)(1-t)^{*}&\text{if }\phi(s)\cup\phi(t)=\theta\\ 0&\text{if }\phi(s)\cup\phi(t)\neq\theta\end{cases},\quad\left(\operatorname{Adj}_{\theta}^{+}\right)_{s,t}=\sum_{\phi(r)=\theta}\left(\frac{\partial r}{\partial s}\right)^{*}\frac{\partial r}{\partial t}.

Since it shall be clear from the context, given mnm\leq n, we shall also denote by Adjm±\operatorname{Adj}_{m}^{\pm} the canonical embedding of Adjm±𝕄|𝒮×𝒮(Gm)\operatorname{Adj}_{m}^{\pm}\in\mathbb{M}_{|\altmathcal{S}_{m}|\times|\altmathcal{S}_{m}|}(\mathbb{R}G_{m}) into 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}):

(Adjm)s,t\displaystyle\left(\operatorname{Adj}_{m}^{-}\right)_{s,t} ={(1s)(1t)if ϕ(s)ϕ(t){1,,m}0if ϕ(s)ϕ(t){1,,m},\displaystyle=\begin{cases}(1-s)(1-t)^{*}&\text{if }\phi(s)\cup\phi(t)\subseteq\{1,\ldots,m\}\\ 0&\text{if }\phi(s)\cup\phi(t)\nsubseteq\{1,\ldots,m\}\end{cases},
(Adjm+)s,t\displaystyle\left(\operatorname{Adj}_{m}^{+}\right)_{s,t} =ϕ(r){1,,m}(rs)rt.\displaystyle=\sum_{\phi(r)\subseteq\{1,\ldots,m\}}\left(\frac{\partial r}{\partial s}\right)^{*}\frac{\partial r}{\partial t}.

Due to the same reason, we decided to remove the grading nn from the notation Adjθ±\operatorname{Adj}_{\theta}^{\pm}.

Lemma 5.4.

For nm3n\geq m\geq 3, one has

σSymnσ(Adjm±)=m(m1)(m2)(n3)!Adjn±.\sum_{\sigma\in\operatorname{Sym}_{n}}\sigma\left(\operatorname{Adj}_{m}^{\pm}\right)=m(m-1)(m-2)(n-3)!\operatorname{Adj}_{n}^{\pm}.
Proof.

The idea of the proof is to transfer the action of the symmetric group Symn\operatorname{Sym}_{n} from the matrices over group rings to the simplex CnC_{n} and apply the invariance of Adjm±\operatorname{Adj}_{m}^{\pm} under the action of Symm\operatorname{Sym}_{m}. The transfer is possible due to the following relationships:

(1) σ(Adjθ±)=Adjσ(θ)±,\sigma\left(\operatorname{Adj}_{\theta}^{\pm}\right)=\operatorname{Adj}_{\sigma(\theta)}^{\pm},

holding for every permutation σSymn\sigma\in\operatorname{Sym}_{n} and a triangle θTrin\theta\in\operatorname{Tri}_{n}. Let us show (1) for Adj+\operatorname{Adj}^{+} only since the proof for Adj\operatorname{Adj}^{-} is a direct application of its definition and the action of Symn\operatorname{Sym}_{n} on 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}). Pick two generators ss and tt of GnG_{n}. Then, it follows by Lemma 5.2 that

(σ(Adjθ+))s,t\displaystyle\left(\sigma\left(\operatorname{Adj}_{\theta}^{+}\right)\right)_{s,t} =σ(ϕ(r)=θ(rσ1(s))rσ1(t))\displaystyle=\sigma\left(\sum_{\phi(r)=\theta}\left(\frac{\partial r}{\partial\sigma^{-1}(s)}\right)^{*}\frac{\partial r}{\partial\sigma^{-1}(t)}\right)
=ϕ(r)=σ(θ)(rs)rt=Adjσ(θ)+.\displaystyle=\sum_{\phi(r)=\sigma(\theta)}\left(\frac{\partial r}{\partial s}\right)^{*}\frac{\partial r}{\partial t}=\operatorname{Adj}_{\sigma(\theta)}^{+}.

Note that, for any k3k\geq 3, we can express Adjk±\operatorname{Adj}_{k}^{\pm} by summing its projections to the triangles of CkC_{k}:

Adjk±=θTrikAdjθ±.\displaystyle\operatorname{Adj}_{k}^{\pm}=\sum_{\theta\in\operatorname{Tri}_{k}}\operatorname{Adj}_{\theta}^{\pm}.

Applying (1) and the orbit-stabilizer theorem applied to the action of Symk\operatorname{Sym}_{k} on the triangles Trik\operatorname{Tri}_{k}, we get

(2) Adjk±=σSymk1|Stab{1,2,3}|Adjσ({1,2,3})±=1(k3)!σSymkσ(Adj3±).\displaystyle\operatorname{Adj}_{k}^{\pm}=\sum_{\sigma\in\operatorname{Sym}_{k}}\frac{1}{|\operatorname{Stab}\{1,2,3\}|}\operatorname{Adj}_{\sigma(\{1,2,3\})}^{\pm}=\frac{1}{(k-3)!}\sum_{\sigma\in\operatorname{Sym}_{k}}\sigma\left(\operatorname{Adj}_{3}^{\pm}\right).

Let τiSymn\tau_{i}\in\operatorname{Sym}_{n}, i=1,,(nm)!i=1,\ldots,(n-m)! be the representatives of the cosets Symn/Symm\operatorname{Sym}_{n}/\operatorname{Sym}_{m}. Applying (2) first for k=nk=n and then for k=mk=m, we get

(n3)!Adjn±\displaystyle(n-3)!\operatorname{Adj}^{\pm}_{n} =σSymnσ(Adj3±)\displaystyle=\sum_{\sigma\in\operatorname{Sym}_{n}}\sigma\left(\operatorname{Adj}^{\pm}_{3}\right)
=iτiτSymmτ(Adj3±)\displaystyle=\sum_{i}\tau_{i}\sum_{\tau\in\operatorname{Sym}_{m}}\tau\left(\operatorname{Adj}^{\pm}_{3}\right)
=(m3)!iτiAdjm±.\displaystyle=(m-3)!\sum_{i}\tau_{i}\operatorname{Adj}^{\pm}_{m}.

We conclude the proof by applying Symm\operatorname{Sym}_{m}-invariance of Adjm±\operatorname{Adj}^{\pm}_{m} (cf. Lemma 5.1 and Corollary 5.3):

(n3)!Adjn±\displaystyle(n-3)!\operatorname{Adj}^{\pm}_{n} =(m3)!iτiAdjm±\displaystyle=(m-3)!\sum_{i}\tau_{i}\operatorname{Adj}^{\pm}_{m}
=(m3)!iτi1m!τSymmσ(Adjm±)\displaystyle=(m-3)!\sum_{i}\tau_{i}\frac{1}{m!}\sum_{\tau\in\operatorname{Sym}_{m}}\sigma\left(\operatorname{Adj}^{\pm}_{m}\right)
=1m(m1)(m2)σSymnσ(Adjm±).\displaystyle=\frac{1}{m(m-1)(m-2)}\sum_{\sigma\in\operatorname{Sym}_{n}}\sigma\left(\operatorname{Adj}^{\pm}_{m}\right).

For any k3k\geq 3, let Adjk=Adjk++Adjk\operatorname{Adj}_{k}=\operatorname{Adj}_{k}^{+}+\operatorname{Adj}_{k}^{-}. We have the following

Corollary 5.5 (cf. Theorem 1.1).

The matrix

Adjnn2m2λI|𝒮\\displaystyle\operatorname{Adj}_{n}-\frac{n-2}{m-2}\lambda I_{|\altmathcal{S}_{n}|}

is a sum of squares in 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}), provided AdjmλI|𝒮\operatorname{Adj}_{m}-\lambda I_{|\altmathcal{S}_{m}|} is a sum of squares in 𝕄|𝒮×𝒮(Gm)\mathbb{M}_{|\altmathcal{S}_{m}|\times|\altmathcal{S}_{m}|}(\mathbb{R}G_{m}).

Proof.

As in the case of Adjm±\operatorname{Adj}^{\pm}_{m}, let us also use the notation I|𝒮I_{|\altmathcal{S}_{m}|} for the canonical embedding of I|𝒮𝕄|𝒮×𝒮(Gm)I_{|\altmathcal{S}_{m}|}\in\mathbb{M}_{|\altmathcal{S}_{m}|\times|\altmathcal{S}_{m}|}(\mathbb{R}G_{m}) into 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}). We show first that

(3) (n2)!m(m1)I|𝒮\=σSymnσI|𝒮.\displaystyle(n-2)!m(m-1)I_{|\altmathcal{S}_{n}|}=\sum_{\sigma\in\operatorname{Sym}_{n}}\sigma I_{|\altmathcal{S}_{m}|}.

Note that the right-hand side of the expression above is a multiplicity of the identity matrix. Let ω0\omega_{0} be 11 in the case Gn=SLn()G_{n}=\operatorname{SL}_{n}(\mathbb{Z}) and 22 in the case Gn=SAut(Fn)G_{n}=\operatorname{SAut}(F_{n}). Then each summand σI|𝒮\sigma I_{|\altmathcal{S}_{m}|} contributes |𝒮ω|\altmathcal{S}_{m}|=m(m-1)\omega_{0} neutral elements to the diagonal which yields n!m(m1)ω0n!m(m-1)\omega_{0} neutral elements altogether. Since the diagonal of I|𝒮\I_{|\altmathcal{S}_{n}|} has |𝒮\\\ω|\altmathcal{S}_{n}|=n(n-1)\omega_{0} neutral elements, the multiplication factor is equal to (n2)!m(m1)(n-2)!m(m-1).

Suppose AdjmλI|𝒮\operatorname{Adj}_{m}-\lambda I_{|\altmathcal{S}_{m}|} is a sum of squares. Combining Lemma Lemma 5.4 and (3), we conclude that the following matrix is a sum of squares in 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}):

m(m1)(m2)(n3)!Adjn(n2)!m(m1)λI|𝒮\.m(m-1)(m-2)(n-3)!\operatorname{Adj}_{n}-(n-2)!m(m-1)\lambda I_{|\altmathcal{S}_{n}|}.

Dividing the expression above by (n3)!m(m1)(n-3)!m(m-1), we get the assertion. ∎

6. Application to Gn=SLn()G_{n}=\operatorname{SL}_{n}(\mathbb{Z})

In this section, let Gn=SLn()G_{n}=\operatorname{SL}_{n}(\mathbb{Z}) for any n3n\geq 3. Using the algorithm for estimating lower bounds for spectral gaps for matrices over group rings described in [KMN24], we were able to obtain the following

Lemma 6.1.

The matrix Adj30.217I6\operatorname{Adj}_{3}-0.217I_{6} is a sum of squares in 𝕄6×6(SL3())\mathbb{M}_{6\times 6}(\mathbb{R}\operatorname{SL}_{3}(\mathbb{Z})).

The replication details of the computation justifying the result above have been desribed in subsection 6.1.

Corollary 6.2 (cf. Corollary 1.2).

Let Δ1\Delta_{1} be the cohomological Laplacian in degree one of GnG_{n}. Then Δ1λI|𝒮\\Delta_{1}-\lambda I_{|\altmathcal{S}_{n}|} is a sum of squares in 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}) for λ=0.217(n2)\lambda=0.217(n-2).

Proof.

Lemma 6.1 together with Corollary 5.5 show that Adjn0.217(n2)I|𝒮\\operatorname{Adj}_{n}-0.217(n-2)I_{|\altmathcal{S}_{n}|} is a sum of squares in 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}). It follows by Lemma 4.3 that the matrices Sqn\operatorname{Sq}_{n}^{-} and Op~n=Opn++2Opn\widetilde{\operatorname{Op}}_{n}=\operatorname{Op}_{n}^{+}+2\operatorname{Op}_{n}^{-} are sums of squares. From this we conclude that Δ10.217(n2)I|𝒮\\Delta_{1}-0.217(n-2)I_{|\altmathcal{S}_{n}|} is a sum of squares in 𝕄|𝒮\×𝒮\(Gn)\mathbb{M}_{|\altmathcal{S}_{n}|\times|\altmathcal{S}_{n}|}(\mathbb{R}G_{n}):

Δ10.217(n2)I|𝒮\=12(2(Adjn0.217(n2)I|𝒮\)+Op~n+Opn++2Sqn).\Delta_{1}-0.217(n-2)I_{|\altmathcal{S}_{n}|}=\frac{1}{2}\left(2\left(\operatorname{Adj}_{n}-0.217(n-2)I_{|\altmathcal{S}_{n}|}\right)+\widetilde{\operatorname{Op}}_{n}+\operatorname{Op}_{n}^{+}+2\operatorname{Sq}_{n}^{-}\right).

6.1. Replication details for SL3()\operatorname{SL}_{3}(\mathbb{Z})

The arguments for converting the numerical sum of squares approximation to the actual one have been described in subsection 3.2 of [KMN24]. Further details concerning finding the numerical approximation can be also found in [KMN24].

We provide a Julia [BEKS17] code to replicate the result of Lemma 6.1. The replication has been described at [KM22], in the ”README.md” file, in the section marked with the arXiv identifier of this preprint. Note that, in order to access the version of the repository corresponding to this article, switching to the branch marked with its arXiv identifier is necessary. The script ”SL_3_Z_adj.jl” contains the replication code for the result of Lemma 6.1. We encourage, however, to use the precomputed solution and run the script ”SL_3_Z_adj_cert.jl” contained in the ”sl3_adj_precom” folder. It should take approximately 11 minute on a standard laptop computer to run the certification script providing the rigorous mathematical proof (the script containing the whole computation takes in turn about 22 hours). The hermitian-squared matrices from the decomposition obtained in Lemma 6.1 have entries supported on the ball of radius 22 with respect to the word-length metric defined by the generating set 𝒮\altmathcal{S}_{3} consisting of six elementary matrices E12E_{12}, E13E_{13}, E21E_{21}, E23E_{23}, E31E_{31}, and E32E_{32}.

References

  • [BdlHV08] Bachir Bekka, Pierre de la Harpe, and Alain Valette. Kazhdan’s property (T), volume 11 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2008.
  • [BEKS17] Jeff Bezanson, Alan Edelman, Stefan Karpinski, and Viral B Shah. Julia: A fresh approach to numerical computing. SIAM review, 59(1):65–98, 2017.
  • [BN20] Uri Bader and Piotr W. Nowak. Group algebra criteria for vanishing of cohomology. J. Funct. Anal., 279(11):108730, 18, 2020.
  • [BS23] Uri Bader and Roman Sauer. Higher kazhdan property and unitary cohomology of arithmetic groups, 2023. https://arxiv.org/abs/2308.06517.
  • [CRW92] Marston Conder, Edmund Robertson, and Peter Williams. Presentations for 33-dimensional special linear groups over integer rings. Proc. Amer. Math. Soc., 115(1):19–26, 1992.
  • [Fox53] Ralph H. Fox. Free differential calculus. I. Derivation in the free group ring. Ann. of Math. (2), 57:547–560, 1953.
  • [Fox54] Ralph H. Fox. Free differential calculus. II. The isomorphism problem of groups. Ann. of Math. (2), 59:196–210, 1954.
  • [Ger84] S. M. Gersten. A presentation for the special automorphism group of a free group. J. Pure Appl. Algebra, 33(3):269–279, 1984.
  • [KKN21] Marek Kaluba, Dawid Kielak, and Piotr W. Nowak. On property (T) for Aut(Fn)\operatorname{Aut}(F_{n}) and SLn()\operatorname{SL}_{n}(\mathbb{Z}). Ann. of Math. (2), 193(2):539–562, 2021.
  • [KM22] Marek Kaluba and Piotr Mizerka. LowCohomologySOS, 2022. https://github.com/piotrmizerka/LowCohomologySOS.
  • [KMN24] Marek Kaluba, Piotr Mizerka, and Piotr W. Nowak. Spectral gap for the cohomological laplacian of SL3()\operatorname{SL}_{3}(\mathbb{Z}). Experimental Mathematics, 0(0):1–6, 2024.
  • [KN18] Marek Kaluba and Piotr W. Nowak. Certifying numerical estimates of spectral gaps. Groups Complex. Cryptol., 10(1):33–41, 2018.
  • [KNO19] Marek Kaluba, Piotr W. Nowak, and Narutaka Ozawa. Aut(𝔽5){\operatorname{Aut}}(\mathbb{F}_{5}) has property (T)(T). Math. Ann., 375(3-4):1169–1191, 2019.
  • [LS77] Roger C. Lyndon and Paul E. Schupp. Combinatorial group theory. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 89. Springer-Verlag, Berlin-New York, 1977.
  • [Lyn50] Roger C. Lyndon. Cohomology theory of groups with a single defining relation. Ann. of Math. (2), 52:650–665, 1950.
  • [Nit22] Martin Nitsche. Computer proofs for property (t), and sdp duality, 2022. https://arxiv.org/abs/2009.05134.
  • [Oza16] Narutaka Ozawa. Noncommutative real algebraic geometry of Kazhdan’s property (T). J. Inst. Math. Jussieu, 15(1):85–90, 2016.