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

Traceless Projector on (Mixed) Tensor Products

Abstract

We describe how traceless projection of tensors of a given rank can be constructed in a closed form. On the way to this goal we invoke the representation theory of the Brauer algebra and the related Schur-Weyl dualities. The resulting traceless projector is constructed from purely combinatorial data involving Young diagrams. By construction, the projector manifestly commutes with the symmetric group and is well-adapted to restrictions to GLGL-irreducible tensor representations. We develop auxiliary computational techniques which serve to take advantage of the obtained results for applications. The proposed method of constructing traceless projectors leads to a particular central idempotent in the semisimple regime of the Brauer algebra.

\MHInternalSyntaxOn\MHInternalSyntaxOff

  

Construction of the traceless projection of tensors
via the Brauer algebra

D. V. Bulgakova, Y. O. Goncharov1,2, T. Helpin2

dvbulgakova@gmail.com        yegor.goncharov@gmail.com        thomas.helpin@gmail.com

1 Service de Physique de l’Univers, Champs et Gravitation, Université de Mons – UMONS,
20 Place du Parc, B-7000 Mons, Belgique

2 Institut Denis Poisson, Université de Tours, Université d’Orléans, CNRS,
Parc de Grandmont, 37200 Tours, France

1 Introduction

Numerous models in classical and quantum field theory assume a vector bundle (i.e. fields taking values in a vector space), with fibers endowed with a metric which is preserved by local (fiber-wise) symmetry transformations. From the algebraic point of view, the space of multiples of the metric forms a trivial representation of the transformations in question. In this respect, given a tensor power of the vector bundle, subspaces obtained by evaluation of scalar products between certain pairs of tensor components (i.e. traces of a tensor) are also preserved. According to Wigner, elementary physical constituents of the theory correspond to irreducible representations of the underlying symmetry group/algebra, i.e. those which contain no non-trivial proper subspaces preserved by symmetry transformations. Thus, in a theory with an invariant metric, tensor fields which represent the elementary physical degrees of freedom are necessarily traceless.

This work is motivated by the purely engineering question of constructing the traceless projection of a tensor in a closed form. More precisely, given a vector space with a non-degenerate metric (either symmetric or skew-symmetric), can one describe the traceless subspace of any tensor power of this space in a systematic manner? Despite the apparent simplicity of the question, computational complexity of solving the traceless Ansatz starting from a general linear combination of a tensor and its single, double, triple, etc., traces grows drastically with the rank of a tensor. In the current work we propose a shortcut based on the natural algebraic structures and the related representation theories of the underlying symmetry algebras.

With a non-degenerate metric on a vector space VV of dimension NN, either symmetric or skew-symmetric, one canonically identifies a subgroup G(N)GL(N)G(N)\subset GL(N) of invertible linear transformations of the vector space which preserve the metric – either the orthogonal group O(N)O(N) or the symplectic group Sp(N)Sp(N), respectively. Due to considerable similarity between the two cases, they are presented in parallel with the use of the same notations.

Representation theory of the classical Lie groups GL(N)GL(N) and G(N)G(N) acting on tensor products VnV^{\otimes n} is well understood since the seminal works by H. Weyl [1]. This classical knowledge is supplemented by a complementary piece – the centralizer algebras 𝔅n(N)\mathfrak{B}_{n}(N) and n(N)𝔅n(N)\mathfrak{C}_{n}(N)\subset\mathfrak{B}_{n}(N) commuting with the action of G(N)G(N) and GL(N)GL(N). With any above pair of the mutually centralising algebras at hand, representation theories of the two meet at the same space VnV^{\otimes n}, which gives rise to remarkable interplay between their irreducible representations. This constitutes the subject of the so-called Schur-Weyl duality reviewed in Section 2. In particular, there is a one-to-one correspondence between the irreducible representations of GL(N)GL(N) (respectively, G(N)G(N)) and n(N)\mathfrak{C}_{n}(N) (respectively, 𝔅n(N)\mathfrak{B}_{n}(N)) occurring in VnV^{\otimes n}, as well as between the branching rules for the embeddings n(N)𝔅n(N)\mathfrak{C}_{n}(N)\subset\mathfrak{B}_{n}(N) and G(N)GL(N)G(N)\subset GL(N).

The algebra 𝔅n(N)\mathfrak{B}_{n}(N) is conveniently realised as a homomorphic image in End(Vn)\mathrm{End}\big{(}V^{\otimes n}\big{)} of a unital associative algebra introduced by R. Brauer in [2], referred to as the Brauer algebra Bn(εN)B_{n}(\varepsilon N) (with ε=1\varepsilon=1 for G(N)=O(N)G(N)=O(N) and ε=1\varepsilon=-1 for G(N)=Sp(N)G(N)=Sp(N)). The action of elements of Bn(εN)B_{n}(\varepsilon N) on VnV^{\otimes n} is very intuitive and consists of permutations of factors in the tensor product, as well as evaluations of scalar products on certain pairs among them. Permutations form a basis in the symmetric group algebra 𝔖nBn(εN)\mathbb{C}\mathfrak{S}_{n}\subset B_{n}(\varepsilon N), whose homomorphic image in End(Vn)\mathrm{End}\big{(}V^{\otimes n}\big{)} gives n(N)\mathfrak{C}_{n}(N). The action of Bn(εN)B_{n}(\varepsilon N) (and thus, of 𝔖n\mathbb{C}\mathfrak{S}_{n}) admits a convenient diagrammatic representation described in Section 3. The interplay between the classical Lie groups and the algebras Bn(εN)B_{n}(\varepsilon N) and 𝔖n\mathbb{C}\mathfrak{S}_{n} can be summarised via the following diagram111Here we reproduce the diagram similar to the one presented in [3].:

Refer to caption
Figure 1: Schur-Weyl dualities for the classical Lie groups GL(N)GL(N) and G(N)G(N).

In our approach, we use the representation theory of Bn(εN)B_{n}(\varepsilon N) to write down the universal traceless projector in a closed form (see Theorem 4.1), totally avoiding the problem of solving any systems of linear equations. We use the term “universal” to emphasize the manifest commutation of the traceless projector in question with the action of 𝔖n\mathbb{C}\mathfrak{S}_{n} (equivalently, with n(N)\mathfrak{C}_{n}(N)).

Taking into account the interplay between representations of GL(N)GL(N) and G(N)G(N), and thus of 𝔅n(N)\mathfrak{B}_{n}(N) and n(N)\mathfrak{C}_{n}(N), we also construct the reduced form of the traceless projector via the restriction of the universal traceless projector to a specific representation of GL(N)GL(N) (see Theorem 4.3). Recall that irreducible representations of G(N)G(N) can be realised as traceless projections of irreducible representations of GL(N)GL(N) (see, e.g., [4, Section 3.2] or [5, Chapter 10]). Technically this implies commutativity of the two operations: i) subtracting traces, ii) projecting onto a GL(N)GL(N)-irreducible subspace of VnV^{\otimes n} by applying primitive idempotents in 𝔖n\mathbb{C}\mathfrak{S}_{n} (for example, Young symmetrisers). This is what one commonly encounters in applications concerning irreducible G(N)G(N)-tensors: the major attention is paid to an appropriate symmetrisation of a tensor, while a compatible projection on the traceless subspace is postponed or even assumed implicitly. The universal traceless projector constructed in this work makes commutativity of the two operations manifest.

Note that the problem of projecting VnV^{\otimes n} onto traceless irreducible representations of G(N)G(N) in terms of the Brauer algebra was solved in [4]. The form of projectors presented therein clearly distinguishes the subtraction of traces and the projection onto an irreducible representation of GL(N)GL(N). The part which is responsible for subtracting traces is specific for each particular primitive idempotent in 𝔖n\mathbb{C}\mathfrak{S}_{n}, and the two do not commute. While in our approach, the reduced traceless projector comes from the universal traceless projector, and thus commutes with the projector onto irreducible representations on GL(N)GL(N). This uniformises the projectors in [4] and in addition allows one to utilise whatever primitive idempotents in 𝔖n\mathbb{C}\mathfrak{S}_{n}.

The traceless projector presented in Theorem 4.1 has the form of a product of linear polynomials in a particular element AnBn(εN)A_{n}\in B_{n}(\varepsilon N), which commutes with 𝔖n\mathbb{C}\mathfrak{S}_{n}. The element AnA_{n} represents the sum of traces among all pairs of indices and is remarkable for its capability of distinguishing the traceless subspace of VnV^{\otimes n} via its representation-theoretic properties combined with the Schur-Weyl-type duality for G(N)G(N): i) AnA_{n} is proportional to identity on irreducible representations of 𝔖n\mathbb{C}\mathfrak{S}_{n} which are subspaces in irreducible representations of Bn(εN)B_{n}(\varepsilon N) in VnV^{\otimes n} (see Lemma 3.3), ii) the kernel of AnA_{n} is exactly the traceless subspace of VnV^{\otimes n} (see Lemma 3.1). Therefore, given a non-zero eigenvalue α\alpha of AnA_{n}, the element (1α1An)\left(1-\alpha^{-1}\,A_{n}\right) annihilates the corresponding traceful subspace in VnV^{\otimes n}. Taking the product over all non-zero eigenvalues, one arrives at an element in Bn(εN)B_{n}(\varepsilon N)

Pn=α0(eigenvalues ofAn)(11αAn),P_{n}=\prod_{\begin{array}[]{c}\scriptstyle\alpha\neq 0\\ \scriptstyle(\text{eigenvalues of}\;\;A_{n})\end{array}}\left(1-\frac{1}{\alpha}\,A_{n}\right)\,, (1.1)

which annihilates the subspaces with non-zero eigenvalues, and acts by identity on the ones with the zero eigenvalue, i.e. the traceless ones. As a result, application of PnP_{n} projects VnV^{\otimes n} onto the traceless subspace. Note that one does not need to diagonalize AnA_{n}, but only to construct the set of its eigenvalues. The latter are obtained explicitly from the set of skew-shape Young diagrams arising from the branching rules for irreducible representations of Bn(εN)B_{n}(\varepsilon N) in VnV^{\otimes n} upon their restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n} (see Proposition 3.4). A particular useful property about the factorised form (1.1) consists in a straightforward simplification of PnP_{n} when restricted to an irreducible representations of GL(N)GL(N) in VnV^{\otimes n}, such that one needs to take into account only specific eigenvalues of AnA_{n} (see Theorem 4.3).

We also aim at adapting our construction to numerical applications. While the combinatorial data (Young diagrams and the derived numeral ingredients) is constructed directly from the related definitions, expanding the factorised expression (1.1) requires optimisation. This problem is partially resolved due to the alternative quasi-additive form of the universal traceless projector (see Corollary 4.5) given as a polynomial expression in AnA_{n} whose degree is smaller than that in (1.1). Additionally, in Section 5 we formulate the known bijection between the centraliser algebra of 𝔖n\mathbb{C}\mathfrak{S}_{n}, Cn(δ)Bn(δ)C_{n}(\delta)\subset B_{n}(\delta) (for all δ\delta\in\mathbb{C}), in terms of ternary bracelets, and propose a technique which allows one to expand (1.1) successively by expressing the left multiplication by AnA_{n} as a differential operator on the space of ternary bracelets. Besides, the relation to ternary bracelets allows us to derive the criterion for commutativily of Cn(δ)C_{n}(\delta) and of its particular sub-/quotient algebras. An extensive amount of examples demonstrating the computational efficiency of our construction are given in the companion Mathematica notebook which follows the narrative of this article. The notebook operates with a newly developed package [6] designed for studies and applications222For applications to tensor calculus the package is linked to the xAct bundle [7] via the xBrauer extension. of the Brauer algebra.

The choice of Bn(εN)B_{n}(\varepsilon N) and its representation theory is not at all unique for the purpose of constructing traceless projectors. While Schur-Weyl-type dualities relate the representation theories of G(N)G(N) and Bn(εN)B_{n}(\varepsilon N) on the finite-dimensional space VnV^{\otimes n}, one has a duality of another type – the Howe duality [8] – when one considers an infinite-dimensional subspace of particularly (anti)symmetrised tensors in the tensor algebra UT(V)U\subset T(V). In this case the subalgebra of transformations of End(T(V))\mathrm{End}(T(V)) which centralises the action of G(N)G(N) is generated by a Lie algebra 𝔡(r)\mathfrak{d}(r) (where the number r2r\geqslant 2 is specified by UU): 𝔡(r)=𝔰𝔭(2r)\mathfrak{d}(r)=\mathfrak{sp}(2r) when G(N)=O(N)G(N)=O(N) and 𝔡(r)=𝔬(2r)\mathfrak{d}(r)=\mathfrak{o}(2r) when G(N)=Sp(N)G(N)=Sp(N). Simple G(N)G(N)-modules are singled out by highest-weight conditions for 𝔡(r)\mathfrak{d}(r). The projection on the subspace of highest-weight vectors is performed by the so-called extremal projectors [9], [10] which can be constructed for any reductive Lie algebra [11] (see [12] for review). In Section 4.3 we analyse the structure of extremal projectors from the point of view of traceless projection of tensors. We show that the extremal projector for 𝔡(r)\mathfrak{d}(r) is divisible by the extremal projector for 𝔰𝔩(r)\mathfrak{sl}(r) (Lemma 4.6), which is reminiscent of the structure of projectors in [4]. An interesting open question would be to find a universal traceless projector which projects the whole space UU onto its traceless counterpart. Let us mention that our approach, with the representation theory of Bn(εN)B_{n}(\varepsilon N) at hand, applies in the context of the fixed rank nn, and is uniform for tensors of any symmetry.

As a by-product of the technique applied for constructing the traceless projection of VnV^{\otimes n}, the expression (1.1) for PnBn(δ)P_{n}\in B_{n}(\delta) (with δ\delta\in\mathbb{C}) gives a particular central idempotent in the Brauer algebra when the latter is semisimple (the result presented in Theorem 6.1). Namely, the central idempotent in question is one of the splitting idempotents described in [13] (see also references therein for an overview of the problem of constructing central idempotents in Bn(δ)B_{n}(\delta)).

The presentation is organised as follows. In Section 2 we recall some basic facts about decomposition of tensor products into irreducible representations upon the action of the metric-preserving group and its centraliser algebra, and formulate the related Schur-Weyl dualities. This results in fixing the properties of the sought traceless projector. In Section 3 we introduce the diagrammatic representation of the Brauer algebra and its action on tensors. Therein we mention some properties of the irreducible representations of Bn(εN)B_{n}(\varepsilon N) and consider their decomposition into irreducible 𝔖n\mathbb{C}\mathfrak{S}_{n}-summands upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n}. In Section 4 we give explicit expressions for the traceless projection and its restrictions to irreducible representations of GL(N)GL(N). In Section 5 we introduce auxiliary techniques which make the obtained expressions for traceless projectors accessible for applications. As an outcome of the proposed approach, in Section 6 we relate the traceless projector to a particular central idempotent in Bn(δ)B_{n}(\delta) when the latter is semisimple. Section 7 serves as a summary where we sum up all the introduced ingredients and construct the traceless projector on V4V^{\otimes 4} step by step. Technical proofs and some examples are given in the Appendix section.

2 Metric-preserving groups on tensor products
and their centraliser algebras

In this section we recall some basic material about the classical Lie groups GL(N)GL(N) and G(N)G(N) acting on tensor products, with the related Schur-Weyl dualities. Apart from merely fixing the notations, our aim here consists in motivating the usage of the Brauer algebra.

Tensor representations of classical Lie groups.

To a finite-dimensional \mathbb{C}-vector space VV of dimension dimV=N\dim V=N, one associates canonically the group GL(N)End(V)GL(N)\subset\mathrm{End}(V) of all invertible linear transformations of this space. Any pair of non-zero vectors are related by a GL(N)GL(N)-transformation, which means that VV is a simple GL(N)GL(N)-module333Throughout the paper we make no distinction between the terms representation and module, which are equivalent when one considers the action of a group/algebra on a vector space. In this context, a simple (respectively, semisimple) module corresponds to an irreducible (respectively, completely reducible) representation.. The set End(V)\mathrm{End}(V) is naturally a GL(N)GL(N)-module under the adjoint action: FRFR1F\mapsto RFR^{-1} for FEnd(V)F\in\mathrm{End}(V) and RGL(N)R\in GL(N).

If one fixes a basis {ea}V\{e_{a}\}\subset V (a=1,,Na=1,\dots,N), any vector is identified with an element of N\mathbb{C}^{N} (the set of its components): v=eavav=e_{a}v^{a}. Here, by the Einstein’s convention, for a pair of indices denoted by the same letter (one up and one down) one performs summation. To any FEnd(V)F\in\mathrm{End}(V) one associates a square matrix FabF^{a}{}_{b} via F(eb)=eaFabF(e_{b})=e_{a}\,F^{a}{}_{b} (with a slight abuse of notation, we will use the same letter for an element of a vector space and its components with respect to a particular basis). There exists a unique, up to a factor, GL(N)GL(N)-invariant linear function on End(V)\mathrm{End}(V), which is called trace, which reads

trF=Faafor any FEnd(V).\mathrm{tr}\,F=F^{a}{}_{a}\quad\text{for any $F\in\mathrm{End}(V)$.} (2.1)

If VV is additionally equipped with a non-degenerate bilinear form (metric) ,\langle\,\cdot,\cdot\,\rangle, symmetric or anti-symmetric, one identifies the group of metric-preserving linear transformations G(N)GL(N)G(N)\subset GL(N):

for anyRG(N)holdsR(v),R(w)=v,w.\text{for any}\quad R\in G(N)\quad\text{holds}\quad\langle R(v),R(w)\rangle=\langle v,w\rangle\,. (2.2)

In the case of a symmetric metric one has G(N)=O(N)G(N)=O(N) (the orthogonal group), while an anti-symmetric metric leads to G(N)=Sp(N)G(N)=Sp(N) (the symplectic group) with N2N\in 2\mathbb{N}. As a G(N)G(N)-module VV is simple. Component-wise, the metric gives rise to a non-degenerate matrix gab=ea,ebg_{ab}=\langle e_{a},e_{b}\rangle, whose inverse we denote by gabg^{ab}: one has gab=±gbag_{ab}=\pm g_{ba} and gacgcb=δabg^{ac}g_{cb}=\delta^{a}{}_{b}. The metric fixes a particular isomorphism between VV and its dual space VV^{*}, which allows transitions between vectors and co-vectors by raising and lowering the indices of the components:

va=gabvbor, equivalentlyva=gabvb.v_{a}=g_{ab}v^{b}\quad\text{or, equivalently}\quad v^{a}=g^{ab}v_{b}\,. (2.3)

To this end, instead of having separately VV and VV^{*} it is sufficient to work only with the space VV.

One can construct the nn-fold tensor product Vn=VVnV^{\otimes n}=\underbrace{V\otimes\dots\otimes V}_{n} of the space VV, with the basis ea1eane_{a_{1}}\otimes\dots\otimes e_{a_{n}},

so that anyTVncan be decomposed asT=ea1eanTa1an.\text{so that any}\;\;T\in V^{\otimes n}\;\;\text{can be decomposed as}\;\;T=e_{a_{1}}\otimes\dots\otimes e_{a_{n}}\,T^{a_{1}\dots a_{n}}\,. (2.4)

In this case we will say that TT has rank nn. The space VnV^{\otimes n} is equipped with a G(N)G(N)-invariant non-degenerate scalar product in a canonical way:

v1vn,w1wn=v1,w1vn,wn.\langle v_{1}\otimes\dots\otimes v_{n},w_{1}\otimes\dots\otimes w_{n}\rangle=\langle v_{1},w_{1}\rangle\dots\langle v_{n},w_{n}\rangle\,. (2.5)

The latter allows one to identify the pairs of dual operators F,FEnd(Vn)F,F^{*}\in\mathrm{End}(V^{\otimes n}) by the requirement

T1,FT2=FT1,T2(for all T1,T2Vn).\left<T_{1},FT_{2}\right>=\left<F^{*}T_{1},T_{2}\right>\quad\text{(for all $T_{1},T_{2}\in V^{\otimes n}$).} (2.6)

The trace (2.1), in combination with the isomorphism VVV\cong V^{*}, can be viewed as a map tr(g):V2\mathrm{tr}^{(g)}:V^{\otimes 2}\to\mathbb{C}. As a straightforward generalisation, one considers linear maps trij(g):VnVn2\mathrm{tr}^{(g)}_{ij}:V^{\otimes n}\to V^{\otimes n-2} (with 1i<jn1\leqslant i<j\leqslant n) which evaluate the scalar product between the two vectors at the iith and jjth positions:

trij(g)(v1vn)=vi,vjv1vivjvn.\mathrm{tr}^{(g)}_{ij}(v_{1}\otimes\dots\otimes v_{n})=\langle v_{i},v_{j}\rangle\;v_{1}\otimes\dots\otimes\bcancel{v_{i}}\otimes\dots\otimes\bcancel{v_{j}}\otimes\dots\otimes v_{n}\,. (2.7)

Component-wise, trij(g)\mathrm{tr}^{(g)}_{ij} acts on (2.4) as ta1angaiajta1aiajant^{a_{1}\dots a_{n}}\mapsto g_{a_{i}a_{j}}\,t^{a_{1}\dots a_{i}\dots a_{j}\dots a_{n}}. A tensor TVnT\in V^{\otimes n} is called traceless if for all 1i<jn1\leqslant i<j\leqslant n the corresponding traces vanish, i.e. trij(g)T=0\mathrm{tr}^{(g)}_{ij}T=0.

The space VnV^{\otimes n} (for all n1n\geqslant 1) is a GL(N)GL(N)-module, with an element RGL(N)R\in GL(N) applied to each tensor component:

R(v1vn)=Rv1Rvn.R(v_{1}\otimes\dots\otimes v_{n})=Rv_{1}\otimes\dots\otimes Rv_{n}\,. (2.8)

In terms of the decomposition over the basis (2.4), Ta1anRa1b1RanTb1bnbnT^{a_{1}\dots a_{n}}\mapsto R^{a_{1}}{}_{b_{1}}\dots R^{a_{n}}{}_{b_{n}}\,T^{b_{1}\dots b_{n}} . As soon as G(N)GL(N)G(N)\subset GL(N), VnV^{\otimes n} is a G(N)G(N)-module as well. In the sequel, whenever VnV^{\otimes n} is considered as a GL(N)GL(N)-module and a G(N)G(N)-module in parallel, we will call it a module without specifying the Lie group.

Due to the classical result of Weyl, the module VnV^{\otimes n} is semisimple: it decomposes into a direct sum of simple GL(N)GL(N)-modules V(μ)V^{(\mu)} (on the left) or simple G(N)G(N)-modules D(λ)D^{(\lambda)} (on the right): respectively,

VnμΣ^n,N(V(μ))gμandVnλΛ^n2f,Nf=0,,n2(D(λ))hλ.V^{\otimes n}\cong\bigoplus_{\mu\in\hat{\Sigma}_{n,N}}\big{(}V^{(\mu)}\big{)}^{\oplus g_{\mu}}\quad\quad\text{and}\quad\quad V^{\otimes n}\cong\bigoplus_{\begin{array}[]{c}\scriptstyle\lambda\in\hat{\Lambda}_{n-2f,N}\\ \scriptstyle f=0,\dots,\lfloor\tfrac{n}{2}\rfloor\end{array}}\big{(}D^{(\lambda)}\big{)}^{\oplus h_{\lambda}}\,. (2.9)

Here x\lfloor x\rfloor denotes the integer part of xx, while the index sets Σ^n,N\hat{\Sigma}_{n,N} and Λ^l,N\hat{\Lambda}_{l,N} are described in (3.34), (3.35) and (3.36).

In order to recall the structure of irreducible components D(λ)VnD^{(\lambda)}\subset V^{\otimes n}, note that the property (2.2) translated to the tensor

E=eaebgabE=e_{a}\otimes e_{b}\,g^{ab} (2.10)

implies that the \mathbb{C}-span of any tensor power EfE^{\otimes f} is a trivial G(N)G(N)-module and, as a consequence, the subspace EfVn2fVnE^{\otimes f}\otimes V^{\otimes n-2f}\subset V^{\otimes n} is invariant with respect to the action of G(N)G(N). Further, invariant subspaces in EfVn2fE^{\otimes f}\otimes V^{\otimes n-2f} correspond exactly to invariant subspaces in Vn2fV^{\otimes n-2f}, which are either traceless or proportional to EE.

Centraliser algebras.

A powerful tool for studying the decomposition of VnV^{\otimes n} into simple components is the centralizer algebra. The latter is identified as the maximal subalgebra in End(Vn)\mathrm{End}(V^{\otimes n}) whose elements commute with any transformation of the Lie group. The prior fact about the centraliser algebras in question is that they are semisimple because the group action is [1, Theorem 3.5.B]. As a representation of two mutually-commuting algebras of transformations, VnV^{\otimes n} decomposes into a multiplicity-free direct sum of pairs of simple modules: one of the Lie group, the other of its centraliser algebra. More in detail, denote n(N)\mathfrak{C}_{n}(N) and 𝔅n(N)\mathfrak{B}_{n}(N) the centraliser algebras for the action of the groups GL(N)GL(N) and G(N)G(N) respectively, and let 𝔏(μ)\mathfrak{L}^{(\mu)} and 𝔐n(λ)\mathfrak{M}_{n}^{(\lambda)} stand for the simple modules over n(N)\mathfrak{C}_{n}(N) and 𝔅n(N)\mathfrak{B}_{n}(N) respectively. Then VnV^{\otimes n} admits the following decompositions, with respect to (GL(N),n(N))\big{(}GL(N),\mathfrak{C}_{n}(N)\big{)}- and (G(N),𝔅n(N))\big{(}G(N),\mathfrak{B}_{n}(N)\big{)}-actions:

VnμΣ^n,NV(μ)𝔏(μ)andVnλΛ^n2f,Nf=0,,n2D(λ)𝔐n(λ).V^{\otimes n}\cong\bigoplus_{\mu\in\hat{\Sigma}_{n,N}}V^{(\mu)}\otimes\mathfrak{L}^{(\mu)}\quad\quad\text{and}\quad\quad V^{\otimes n}\cong\bigoplus_{\begin{array}[]{c}\scriptstyle\lambda\in\hat{\Lambda}_{n-2f,N}\\ \scriptstyle f=0,\dots,\lfloor\tfrac{n}{2}\rfloor\end{array}}D^{(\lambda)}\otimes\mathfrak{M}_{n}^{(\lambda)}\,. (2.11)

By comparing (2.9) and (2.11), gμ=dim𝔏(μ)g_{\mu}=\dim\mathfrak{L}^{(\mu)} and hλ=dim𝔐n(λ)h_{\lambda}=\dim\mathfrak{M}_{n}^{(\lambda)}, i.e. the elements of the latter module count the multiplicity of the former (and vice versa).

Realisation of centraliser algebras.

The efficiency of applications of the centraliser algebras in question is due to the possibility of parametrising their elements via a “somewhat enigmatic algebra” [1], whose action on VnV^{\otimes n} is diagrammatic and takes care of certain adjacency relations among the factors VV in VnV^{\otimes n}, treating them as “atomary” and totally ignoring its “intrinsic” vector-space structure. From the fact that G(N)GL(N)G(N)\subset GL(N) one gets the inclusion n(N)𝔅n(N)\mathfrak{C}_{n}(N)\subset\mathfrak{B}_{n}(N), so we concentrate on the latter algebra while the former will be treated via its embedding.

The algebra 𝔅n(N)\mathfrak{B}_{n}(N) is conveniently realised as a homomorphic image of an associative algebra Bn(εN)B_{n}(\varepsilon N) – the Brauer algebra [2] (here and in what follows ε=1\varepsilon=1 for G(N)=O(N)G(N)=O(N) and ε=1\varepsilon=-1 when G(N)=Sp(N)G(N)=Sp(N)). The corresponding ε\varepsilon-dependent homomorphism 𝔯\mathfrak{r} is surjective, and it is also injective for NnN\geqslant n when ε=1\varepsilon=1 and N2nN\geqslant 2n when ε=1\varepsilon=-1 [14]. The action of Bn(εN)B_{n}(\varepsilon N) on VnV^{\otimes n} is very intuitive (see Section 3.1), so instead of working with the centraliser algebra 𝔅n(N)\mathfrak{B}_{n}(N) directly, we will address to its elements via the 𝔯\mathfrak{r}-image of the Brauer algebra Bn(εN)B_{n}(\varepsilon N). In this respect, instead of the simple 𝔅n(N)\mathfrak{B}_{n}(N)-modules 𝔐n(λ)\mathfrak{M}^{(\lambda)}_{n} we will use the (isomorphic) simple Bn(εN)B_{n}(\varepsilon N)-modules Mn(λ^)M^{(\hat{\lambda})}_{n}, where λλ^\lambda\mapsto\hat{\lambda} is an ε\varepsilon-dependent involutive operation (3.36).

The algebra Bn(εN)B_{n}(\varepsilon N) contains the symmetric group algebra 𝔖n\mathbb{C}\mathfrak{S}_{n} as a subalgebra, and the 𝔯\mathfrak{r}-image of the latter gives exactly n(N)\mathfrak{C}_{n}(N) [1]. So in the sequel we work in terms of 𝔖nBn(εN)\mathbb{C}\mathfrak{S}_{n}\subset B_{n}(\varepsilon N), and use the simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules L(μ^)L^{(\hat{\mu})} instead of 𝔏(μ)\mathfrak{L}^{(\mu)} (with the same map μμ^\mu\mapsto\hat{\mu} as for the Bn(εN)B_{n}(\varepsilon N)-modules above).

Summarising the above, VnV^{\otimes n} is a (GL(N),𝔖n)\big{(}GL(N),\mathbb{C}\mathfrak{S}_{n}\big{)}-bimodule or a (G(N),Bn(εN))\big{(}G(N),B_{n}(\varepsilon N)\big{)}-bimodule. The two decompositions in (2.11) are rewritten, respectively, as follows:

VnμΣ^n,NV(μ)L(μ^)andVnλΛ^n2f,Nf=0,,n2D(λ)Mn(λ^).V^{\otimes n}\cong\bigoplus_{\mu\in\hat{\Sigma}_{n,N}}V^{(\mu)}\otimes L^{(\hat{\mu})}\quad\quad\text{and}\quad\quad V^{\otimes n}\cong\bigoplus_{\begin{array}[]{c}\scriptstyle\lambda\in\hat{\Lambda}_{n-2f,N}\\ \scriptstyle f=0,\dots,\lfloor\tfrac{n}{2}\rfloor\end{array}}D^{(\lambda)}\otimes M_{n}^{(\hat{\lambda})}\,. (2.12)

The interplay between simple GL(N)GL(N)- and 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules, as well as G(N)G(N)- and Bn(εN)B_{n}(\varepsilon N)-modules, in the decompositions (2.12) is known as Schur-Weyl duality. It was originally established for the former decomposition [1], and then adapted to the case of G(N)G(N) through the seminal works [2, 15], see also [16] (Weyl utilised the elements of the (GL(N),𝔖n)\big{(}GL(N),\mathbb{C}\mathfrak{S}_{n}\big{)}-duality to study the decomposition of the G(N)G(N)-module VnV^{\otimes n} [1]).

Projection to a simple GL(N)GL(N)-component V(μ)V^{(\mu)} (respectively, G(N)G(N)-component D(λ)D^{(\lambda)}) is obtained by fixing an element in the corresponding simple module L(μ^)L^{(\hat{\mu})} (respectively, Mn(λ^)M_{n}^{(\hat{\lambda})}). A particular well-known use of the classical Schur-Weyl duality consists in application of primitive idempotents in 𝔖n\mathbb{C}\mathfrak{S}_{n} to VnV^{\otimes n} (e.g. Young symmetrisers [17, Chapter 4] or orthogonal primitive idempotents [18, 19]) in order to obtain irreducible GL(N)GL(N)-modules. Primitive orthogonal idempotents in 𝔅n(N)\mathfrak{B}_{n}(N) in terms of the Brauer algebra were constructed in [4].

Before going into technical details about the algebras 𝔖nBn(εN)\mathbb{C}\mathfrak{S}_{n}\subset B_{n}(\varepsilon N) and their representation theory, let us declare

The main goal of the present work. We aim at constructing the universal traceless projector 𝔓n\mathfrak{P}_{n}, which is specified by the following natural requirements. First of all, it is an idempotent in End(Vn)\mathrm{End}(V^{\otimes n}), i.e. 𝔓n𝔓n=𝔓n\mathfrak{P}_{n}\mathfrak{P}_{n}=\mathfrak{P}_{n}. Second, for any TVnT\in V^{\otimes n}, for all 1i<jn1\leqslant i<j\leqslant n:

(P1)trij(g)(𝔓nT)=0(the image is traceless),(P2)iftrij(g)T=0,then𝔓nT=T(among traceless projectors, 𝔓n has maximal rank),(P3)for anys𝔖n,𝔓n𝔯(s)T=𝔯(s)𝔓nT(𝔓n preserves symmetries of a tensor).\begin{array}[]{ll}(\mathrm{P}1)&\mathrm{tr}^{(g)}_{ij}(\mathfrak{P}_{n}T)=0\quad\text{(the image is traceless),}\\ (\mathrm{P}2)&\text{if}\quad\mathrm{tr}^{(g)}_{ij}T=0\,,\quad\text{then}\quad\mathfrak{P}_{n}T=T\quad\text{(among traceless projectors, $\mathfrak{P}_{n}$ has maximal rank),}\\ (\mathrm{P}3)&\text{for any}\;\;s\in\mathfrak{S}_{n}\,,\quad\mathfrak{P}_{n}\mathfrak{r}(s)\,T=\mathfrak{r}(s)\mathfrak{P}_{n}\,T\quad\text{($\mathfrak{P}_{n}$ preserves symmetries of a tensor).}\end{array} (2.13)

The prior positive fact is that, given nn and NN, the traceless projector satisfying (P1)(\mathrm{P}1) and (P2)(\mathrm{P}2) exists and is unique [1, Theorem 5.6A] (see also [5, Chapter 10]). Let us briefly note that the proof is purely representation-theoretic and simply relies on the fact that in a representation space equipped with a non-degenerate invariant scalar product, for any invariant subspace its orthogonal complement is an invariant subspace as well. If one takes the invariant subspace WVnW\subset V^{\otimes n} of all tensors proportional to the metric, then its orthogonal complement WW^{\prime} is exactly the subspace of traceless tensors. The fact that 𝔓n\mathfrak{P}_{n} performs an orthogonal projection assures that (𝔓n)=𝔓n(\mathfrak{P}_{n})^{*}=\mathfrak{P}_{n}.

The properties (P1)(\mathrm{P}1), (P2)(\mathrm{P}2) imply that 𝔓n\mathfrak{P}_{n} is proportional to identity on the simple G(N)G(N)-submodules in VnV^{\otimes n}: it acts by identity on those simple modules which occur in WW^{\prime} and annihilates all the others. As a result, 𝔓n𝔅n(N)\mathfrak{P}_{n}\in\mathfrak{B}_{n}(N). Note also that permutations of tensor factors preserves the decomposition VnWWV^{\otimes n}\cong W\oplus W^{\prime}. From the fact that 𝔓n\mathfrak{P}_{n} is proportional to identity on each direct component and decomposing them into simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules, one concludes that the property (P3)(\mathrm{P}3) holds as a consequence of (P1)(\mathrm{P}1) and (P2)(\mathrm{P}2). Nevertheless we choose to keep it explicit because it will serve as a starting point in our construction in Section 3.2.

To this end, one naturally arrives at the idea of application of the Brauer algebra to the construction of 𝔓n\mathfrak{P}_{n}. The properties (2.13) can be verified in a systematic manner, which leads to an element PnBn(εN)P_{n}\in\ B_{n}(\varepsilon N) such that 𝔓n=𝔯(Pn)\mathfrak{P}_{n}=\mathfrak{r}(P_{n}). This result is presented in Theorem 4.1.

3 The Brauer algebra

3.1 Diagrammatic representation and action on tensors

The algebra of Brauer diagrams.

In order to introduce the Brauer algebra Bn(δ)B_{n}(\delta) (for any δ\delta\in\mathbb{C}) it is convenient to start by recalling the diagrammatic representation of the symmetric group. To any element s𝔖ns\in\mathfrak{S}_{n} one associates a permutation diagram: two rows of nn vertices (nodes) aligned horizontally and placed one above another, with the iith node in the upper row and the s(i)s(i)th node in the lower row joined by a line.

s=(12343142)[Uncaptioned image],t=(12344123)[Uncaptioned image]s=\ \begin{pmatrix}1&2&3&4\\ 3&1&4&2\end{pmatrix}\mapsto\raisebox{-0.45pt}{\includegraphics[width=60.0pt,height=60.0pt]{tau1Int.pdf}}\quad,\hskip 28.45274ptt=\ \begin{pmatrix}1&2&3&4\\ 4&1&2&3\end{pmatrix}\mapsto\raisebox{-0.45pt}{\includegraphics[width=60.0pt,height=60.0pt]{tau2Int.pdf}} (3.1)

In order to calculate the product stst of two elements s,t𝔖ns,t\in\mathfrak{S}_{n} diagrammatically, one places the diagram ss below tt and identifies the upper nodes of the former with the lower nodes of the latter. Then straightening the lines gives the diagram associated with the element stst:

st=[Uncaptioned image]=[Uncaptioned image]s\,t\;=\;\raisebox{-0.45pt}{\includegraphics[width=60.0pt,height=56.0pt]{tau1tau2.pdf}}\;=\;\raisebox{-0.45pt}{\includegraphics[width=60.0pt,height=40.0pt]{tau1tau2f.pdf}} (3.2)

In the sequel we make no distinction between the elements of 𝔖n\mathfrak{S}_{n} and the corresponding diagrams. Passing to 𝔖n\mathbb{C}\mathfrak{S}_{n} consists in allowing linear combinations of diagrams treated as basis vectors.

As a vector space, the Brauer algebra Bn(δ)B_{n}(\delta) is obtained by extending the basis of permutation diagrams to Brauer diagrams. Namely, for the same set of nodes as for the permutation diagrams, each node is required to be an endpoint of exactly one line. For example, one can consider the following two Brauer diagrams b1b_{1} and b2b_{2} for n=5n=5.

b1=[Uncaptioned image],b2=[Uncaptioned image]b_{1}\,=\,\raisebox{-0.45pt}{\includegraphics[width=60.0pt,height=40.0pt]{b1.pdf}}\;,\hskip 28.45274ptb_{2}\,=\,\raisebox{-0.45pt}{\includegraphics[width=60.0pt,height=40.0pt]{b2.pdf}} (3.3)

By a straightforward combinatorial computation one finds dimBn(δ)=(2n)!2nn!=(2n1)!!\dim B_{n}(\delta)=\frac{(2n)!}{2^{n}\,n!}=(2n-1)!!.

The lines joining an upper node with a lower node will be referred to as vertical lines, while the lines joining a pair of nodes in the same row will be referred to as arcs. We will say that a diagram contains f0f\geqslant 0 arcs if there is ff arcs in either of its rows. By construction, 𝔖n\mathbb{C}\mathfrak{S}_{n} is embedded in Bn(δ)B_{n}(\delta) via diagrams containing 0 arcs.

The product of two diagrams b1b2b_{1}b_{2} is parametrised by a variable δ\delta\in\mathbb{C} and defined in a similar fashion as before. Place b1b_{1} below b2b_{2}, identify the upper nodes of the former with the lower nodes of the latter. If 0\ell\geqslant 0 is the number of loops, the resulting vector b1b2b_{1}b_{2} is obtained by omitting the loops, straightening the lines and multiplying the so-constructed diagram by δ\delta^{\ell}.

b1b2=[Uncaptioned image]=δ[Uncaptioned image]b_{1}\,b_{2}\;=\;\raisebox{-0.45pt}{\includegraphics[width=60.0pt,height=55.0pt]{b1b2.pdf}}\hskip 8.5359pt\;=\;\hskip 8.5359pt\delta\;\raisebox{-0.45pt}{\includegraphics[width=60.0pt,height=40.0pt]{b1b2f.pdf}} (3.4)

It can be proven that the so-defined product in Bn(δ)B_{n}(\delta) is associative.

From the definition of the product of Brauer diagrams it follows that the number of arcs in a diagram can not be reduced via multiplication by other diagrams. Let J(f)Bn(δ)J^{(f)}\subset B_{n}(\delta) denote the span of all diagrams with at least ff arcs. These subspaces are two-sided ideals in Bn(δ)B_{n}(\delta) which form a chain of embeddings:

Bn(δ)=J(0)J(1)J(n2).B_{n}(\delta)=J^{(0)}\supset J^{(1)}\supset\dots\supset J^{\left(\lfloor\frac{n}{2}\rfloor\right)}\,. (3.5)

The above chain is the major tool in studying the algebra Bn(δ)B_{n}(\delta) since the seminal works [15, 14]. Note also the simple fact that

Bn(δ)/J(1)𝔖n.B_{n}(\delta)/J^{(1)}\cong\mathbb{C}\mathfrak{S}_{n}\,. (3.6)

Generators and relations.

The algebra Bn(δ)B_{n}(\delta) is generated by the following set of diagrams sis_{i} and did_{i} (i=1n1i=1\dots n-1):

si=[Uncaptioned image],di=[Uncaptioned image]s_{i}\;=\;\raisebox{-0.4pt}{\includegraphics[width=130.0pt,height=50.0pt]{si.pdf}}\;,\quad\quad d_{i}\;=\;\raisebox{-0.4pt}{\includegraphics[width=130.0pt,height=50.0pt]{di.pdf}} (3.7)

The generators verify the following set of defining relations444Equivalently, one can define the algebra Bn(δ)B_{n}(\delta) as generated by si,dis_{i},d_{i} (i=1,,n1i=1,\dots,n-1) modulo the defining relations (3.8)-(3.11), and prove that it is isomorphic to the algebra of Brauer diagrams introduced above.:

si2\displaystyle s_{i}{}^{2} =1,di=2δdi,disi=sidi=di,\displaystyle=1\,,\quad d_{i}{}^{2}=\delta\,d_{i}\,,\quad d_{i}s_{i}=s_{i}d_{i}=d_{i}\,, (3.8)
sisj=sj\displaystyle s_{i}s_{j}=s_{j} si,disj=sjdi,didj=djdifor|ij|2,\displaystyle s_{i}\,,\quad d_{i}s_{j}=s_{j}d_{i}\,,\quad d_{i}d_{j}=d_{j}d_{i}\quad\text{for}\quad|i-j|\geqslant 2\,, (3.9)
sisi+1si\displaystyle s_{i}s_{i+1}s_{i} =si+1sisi+1,didi+1di=di,di+1didi+1=di+1,\displaystyle=s_{i+1}s_{i}s_{i+1}\,,\quad d_{i}d_{i+1}d_{i}=d_{i}\,,\quad d_{i+1}d_{i}d_{i+1}=d_{i+1}\,, (3.10)
sidi+1di\displaystyle s_{i}d_{i+1}d_{i} =si+1di,di+1disi+1=di+1si,\displaystyle=s_{i+1}d_{i}\,,\quad d_{i+1}d_{i}s_{i+1}=d_{i+1}s_{i}\,, (3.11)

while any other relation in Bn(δ)B_{n}(\delta) is a result of a composition of the relations (3.8)-(3.11). The elements sis_{i} satisfy the known relations for simple transpositions in the symmetric group 𝔖n\mathfrak{S}_{n}, so linear combinations of their products constitutes the symmetric group algebra 𝔖n\mathbb{C}\mathfrak{S}_{n}.

We also introduce the following elements, for all possible i<ji<j:

sij=[Uncaptioned image],dij=[Uncaptioned image]s_{ij}\;=\;\raisebox{-0.4pt}{\includegraphics[width=140.0pt,height=50.0pt]{sij.pdf}}\;,\quad\quad d_{ij}\;=\;\raisebox{-0.4pt}{\includegraphics[width=140.0pt,height=50.0pt]{dij.pdf}} (3.12)

Action on VnV^{\otimes n}.

Application of Bn(δ)B_{n}(\delta) in the context of Schur-Weyl-type dualities for the group G(N)G(N) takes place for integer values of the parameter δ=εN\delta=\varepsilon N, where ε=+1\varepsilon=+1, NN\in\mathbb{N} when G(N)=O(N)G(N)=O(N) and ε=1\varepsilon=-1, N2N\in 2\mathbb{N} when G(N)=Sp(N)G(N)=Sp(N). To describe the left555In the present work we consider the left action of Bn(εN)B_{n}(\varepsilon N). In the context of bi-modules, it is also common to consider the right action defined as 𝔯op(b)=𝔯(b)\mathfrak{r}_{\mathrm{op}}(b)=\mathfrak{r}(b^{*}), where bb^{*} is the diagram obtained by reflecting bb with respect to the horizontal middle line. In this case, 𝔯op(b1)𝔯op(b2)=𝔯op(b2b1)\mathfrak{r}_{\mathrm{op}}(b_{1})\mathfrak{r}_{\mathrm{op}}(b_{2})=\mathfrak{r}_{\mathrm{op}}(b_{2}b_{1}). action of any diagram bBn(εN)b\in B_{n}(\varepsilon N) on VnV^{\otimes n}, denote 𝚒(b)\mathtt{i}(b) the minimal number of intersections among its lines. Then the transformation 𝔯(b)\mathfrak{r}(b) is applied to v1vnv_{1}\otimes\dots\otimes v_{n} as follows (see Theorem 2.10 in [20]): place the tensor factors in the upper nodes of bb and perform i) permutations by following the passing lines from top to bottom, ii) contractions of pairs joined by arcs in the upper row, iii) insertion of E=eaebgabE=e_{a}\otimes e_{b}\,g^{ab} at each pair of the lower nodes joined by an arc; the result is multiplied by ε𝚒(b)\varepsilon^{\mathtt{i}(b)}. For example, for the diagrams b1,b2b_{1},b_{2} introduced above one has:

𝔯(b1)v1v2v3v4v5=εeav1ebecedgacgbdv2,v4v3,v5,𝔯(b2)v1v2v3v4v5=εv2eaebecedgabgcdv1,v5v3,v4.\begin{array}[]{l}\mathfrak{r}(b_{1})\,v_{1}\otimes v_{2}\otimes v_{3}\otimes v_{4}\otimes v_{5}=\varepsilon\,e_{a}\otimes v_{1}\otimes e_{b}\otimes e_{c}\otimes e_{d}\;g^{ac}g^{bd}\langle v_{2},v_{4}\rangle\langle v_{3},v_{5}\rangle\,,\\ \mathfrak{r}(b_{2})\,v_{1}\otimes v_{2}\otimes v_{3}\otimes v_{4}\otimes v_{5}=\varepsilon\,v_{2}\otimes e_{a}\otimes e_{b}\otimes e_{c}\otimes e_{d}\;g^{ab}g^{cd}\langle v_{1},v_{5}\rangle\langle v_{3},v_{4}\rangle\,.\end{array} (3.13)

The action of the generators sis_{i}, did_{i} is particularly simple:

𝔯(si)v1vivi+1vn=εv1vi+1vivn,𝔯(di)v1vivi+1vn=vi,vi+1v1Evn.\begin{array}[]{l}\mathfrak{r}(s_{i})\,v_{1}\otimes\dots\otimes v_{i}\otimes v_{i+1}\otimes\dots\otimes v_{n}=\varepsilon\,v_{1}\otimes\dots\otimes v_{i+1}\otimes v_{i}\otimes\dots\otimes v_{n}\,,\\ \mathfrak{r}(d_{i})\,v_{1}\otimes\dots\otimes v_{i}\otimes v_{i+1}\otimes\dots\otimes v_{n}=\langle v_{i},v_{i+1}\rangle\,v_{1}\otimes\dots\otimes E\otimes\dots\otimes v_{n}\,.\end{array} (3.14)

One can check that the above transformations (3.14) verify the relations (3.8)-(3.11), and hence 𝔯\mathfrak{r} is indeed the homomorphism of algebras. As was pointed out, the homomorphism is injective when NnN\geqslant n for ε=1\varepsilon=1 and N2nN\geqslant 2n for ε=1\varepsilon=-1 [14]. When it is not the case, the non-trivial kernel is generated by particular anti-symmetrisations of tensor factors [21, Theorem 3.7] (in this relation, see also [1, Theorems 2.17.A and 6.1.B]).

The diagrammatic representation of Bn(εN)B_{n}(\varepsilon N), when translated to the action on tensors by the homomorphism 𝔯\mathfrak{r}, elucidates the operations on the components of the tensor products (or, simply, indices). Due to this convenience, in the sequel we will often consider VnV^{\otimes n} directly as a module over Bn(εN)B_{n}(\varepsilon N) and treat the two notations for the action of xBn(εN)x\in B_{n}(\varepsilon N) on TVnT\in V^{\otimes n}𝔯(x)T\mathfrak{r}(x)T and x(T)x(T) – on equal footing.

The operation ()(\,\cdot\,)^{*} defined in (2.6) is an involutive anti-automorphism (or anti-involution) of the algebra End(Vn)\mathrm{End}(V^{\otimes n}). There is a natural involutive anti-automorphism of the Brauer algebra given by flipping each diagram with respect to the horizontal middle line. As a matter of a simple check, the 𝔯\mathfrak{r}-image of a flipped diagram coincides with the adjoint to the 𝔯\mathfrak{r}-image of the initial diagram. Thus, with a slight abuse of notation,

for any diagram bBn(εN)b\in B_{n}(\varepsilon N) denote its flip bb^{*}. Then one has 𝔯(b)=𝔯(b)\mathfrak{r}(b^{*})=\mathfrak{r}(b)^{*}. (3.15)

Note that for any s𝔖ns\in\mathfrak{S}_{n}, s=s1s^{*}=s^{-1}. The action of ()(\,\cdot\,)^{*} on linear combinations of diagrams is defined by linearity.

3.2 The centraliser of 𝔖n\mathbb{C}\mathfrak{S}_{n}

Two diagrams b,bBn(δ)b,b^{\prime}\in B_{n}(\delta), are said to be conjugate, which is denoted bbb\sim b^{\prime}, if there exists an element s𝔖ns\in\mathfrak{S}_{n} (a monomial in generators sis_{i}) such that b=sbs1b^{\prime}=sbs^{-1}. Two Brauer diagrams are conjugate iff they are related by a permutation of pairs of vertically aligned nodes.

Define the average of bBn(δ)b\in B_{n}(\delta) as follows:

γb=s𝔖nsbs1.\gamma_{b}=\sum_{s\in\mathfrak{S}_{n}}sbs^{-1}\,. (3.16)

Let C𝔖n(b)C_{\mathfrak{S}_{n}}(b) denote the centralizer of a diagram bBn(δ)b\in B_{n}(\delta) in 𝔖n\mathfrak{S}_{n}, i.e. the set of elements r𝔖nr\in\mathfrak{S}_{n} such that rbr1=brbr^{-1}=b. The cardinalities of centralisers of two conjugate diagrams coincide, so one has

γb=|C𝔖n(b)|bbb.\gamma_{b}=\big{|}C_{\mathfrak{S}_{n}}(b)\big{|}\,\sum_{b^{\prime}\sim b}b^{\prime}\,. (3.17)

The subalgebra Cn(δ)C_{n}(\delta) spanned by all averages forms the centraliser of 𝔖n\mathbb{C}\mathfrak{S}_{n} in Bn(δ)B_{n}(\delta):

for any s𝔖nholdsus=suu=γbfor some bBn(δ).\text{for any }s\in\mathfrak{S}_{n}\;\;\text{holds}\;\;us=su\quad\Leftrightarrow\quad u=\gamma_{b}\;\;\text{for some $b\in B_{n}(\delta)$}\,. (3.18)

The proof of the above implication from left to right is straightforward, while for the opposite implication note that for the decomposition over the diagram basis in Bn(δ)B_{n}(\delta),

u=bBn(δ)cbb,u=\sum_{b\in B_{n}(\delta)}c_{b}\,b\,, (3.19)

the condition us=suus=su for all s𝔖ns\in\mathfrak{S}_{n} is equivalent to csbs1=cbc_{sbs^{-1}}=c_{b} for each diagram bb in the above decomposition.

The principal building block of 𝔓n\mathfrak{P}_{n}.

The significance of the algebra Cn(δ)C_{n}(\delta) for the purpose of constructing the universal traceless projector is due to (P3)(\mathrm{P}3) in (2.13), so it is natural to expect that

𝔓n=𝔯(Pn)for somePnCn(δ).\mathfrak{P}_{n}=\mathfrak{r}(P_{n})\quad\text{for some}\quad P_{n}\in C_{n}(\delta)\,. (3.20)

In this respect, one can look for PnP_{n} as a linear combination of averages (3.16) and fix the coefficients from the requirement 𝔯(J(1)Pn)=0\mathfrak{r}(J^{(1)}P_{n})=0 (this procedure was analysed in [13] for the construction of particular idempotents in Bn(δ)B_{n}(\delta), see Section 6). We choose a different direction and invoke representation theory of the Brauer algebra, focusing on the following normalised average:

An=12(n2)!γd1=1i<jndijCn(N)J(1),(An)=An.A_{n}=\frac{1}{2(n-2)!}\,\gamma_{d_{1}}=\sum_{1\leqslant i<j\leqslant n}d_{ij}\in C_{n}(N)\cap J^{(1)}\,,\quad(A_{n})^{*}=A_{n}\,. (3.21)
Lemma 3.1.

The action of AnA_{n} on VnV^{\otimes n} is diagonalisable. The subspace KerAnVn\mathrm{Ker}\,A_{n}\subset V^{\otimes n} is exactly the space of traceless tensors, while non-zero eigenvalues of AnA_{n} are in ε\varepsilon\mathbb{N}.

Proof.

We concentrate on the case G(N)=O(N)G(N)=O(N) (with ε=1\varepsilon=1), which is instructive and transparent at the same time. The case G(N)=Sp(N)G(N)=Sp(N) (with ε=1\varepsilon=-1) utilises the same idea, but is more involved technically, so we postpone it to Appendix B.1. Fix a real structure in VV such that ,\langle\,\cdot,\cdot\,\rangle (as defined in (2.5)) has Euclidean signature, and denote VV_{\mathbb{R}} the real subspace.

The action of dijd_{ij} on decomposable tensors is given by:

𝔯(dij)v1vn=vi,vjgaiajv1eaieajvn,\mathfrak{r}(d_{ij})\,v_{1}\otimes\dots\otimes v_{n}=\langle v_{i},v_{j}\rangle\,g^{a_{i}a_{j}}v_{1}\otimes\dots\otimes e_{a_{i}}\otimes\dots\otimes e_{a_{j}}\otimes\dots\otimes v_{n}\,, (3.22)

so for two tensors T1,T2VnT_{1},T_{2}\in V_{\mathbb{R}}^{\otimes n} one has

<T1,dij(T2)>=<trij(g)T1,trij(g)T2>=<dij(T1),T2>(for any i<j),\big{<}T_{1},d_{ij}(T_{2})\big{>}=\big{<}\mathrm{tr}^{(g)}_{ij}T_{1},\mathrm{tr}^{(g)}_{ij}T_{2}\big{>}=\big{<}d_{ij}(T_{1}),T_{2}\big{>}\quad\text{(for any $i<j$)}\,, (3.23)

where the trace operation was defined in (2.7). Thus, 𝔯(dij)\mathfrak{r}(d_{ij}) (for any i<ji<j) is self-adjoint, so it is diagonalisable with respect to some orthonormal basis. Since the scalar product is positive-definite, (3.23) implies that the eigenvalues of 𝔯(dij)\mathfrak{r}(d_{ij}) are non-negative. As a result, the action of An=1i<jndijA_{n}=\sum_{1\leqslant i<j\leqslant n}d_{ij} on VnV^{\otimes n} is self-adjoint, and therefore diagonalisable with respect to some orthonormal basis. With this at hand, it is a simple exercise to prove that its eigenvalues are non-negative as well.

Let us show that TVnT\in V_{\mathbb{R}}^{\otimes n} is traceless iff An(T)=0A_{n}(T)=0. One way is simple: any traceless tensor is in KerAn\mathrm{Ker}\,A_{n}. Other way around, assume An(T)=0A_{n}(T)=0 and use (3.23) to write

0=<T,An(T)>=1i<jn<trij(g)T,trij(g)T>.0=\big{<}T,A_{n}(T)\big{>}=\sum_{1\leqslant i<j\leqslant n}\big{<}\mathrm{tr}^{(g)}_{ij}T,\mathrm{tr}^{(g)}_{ij}T\big{>}\,. (3.24)

Each term on the right-hand-side is non-negative, so the whole sum vanishes only if each term vanishes individually, which is the case only when trij(g)T=0\mathrm{tr}_{ij}^{(g)}T=0. Taking into account that VVV\cong\mathbb{C}\otimes V_{\mathbb{R}}, the above arguments clearly extend to VnV^{\otimes n}. ∎

Our next goal consists in describing the properties of AnA_{n} from the point of view of the representation theory of Bn(εN)B_{n}(\varepsilon N). In particular, the latter will provide the eigenvalues of AnA_{n} on VnV^{\otimes n}.

3.3 Simple 𝔖n\mathbb{C}\mathfrak{S}_{n}- and Bn(εN)B_{n}(\varepsilon N)-modules in VnV^{\otimes n}

Combinatorial and algebraic properties of simple 𝔖n\mathbb{C}\mathfrak{S}_{n}- and Bn(εN)B_{n}(\varepsilon N)-modules are expressed via partitions of integers, which leads inevitably to Young diagrams and makes it natural to start by recalling some related basics.

Partitions and Young diagrams.

A rr-partition λ\lambda of an integer nn\in\mathbb{N}, denoted as λn\lambda\vdash n, is a weakly decreasing sequence of positive integers λ=(λ1,,λr)\lambda=(\lambda_{1},\dots,\lambda_{r}) such that |λ|=i=1rλi=n|\lambda|=\sum_{i=1}^{r}\lambda_{i}=n. If each λi\lambda_{i} is even, the partition λ\lambda is said to be even. Partitions are conveniently identified with the Young diagrams: arrays of squares placed at matrix entries (i,j)(i,j), such that for a given rr-partition λ\lambda, 1ir1\leqslant i\leqslant r and 1jλi1\leqslant j\leqslant\lambda_{i}. To any partition λ=(λ1,,λr)\lambda=(\lambda_{1},\dots,\lambda_{r}) of nn one constructs the dual partition λ=(λ1,,λλ1)\lambda^{\prime}=(\lambda_{1}^{\prime},\dots,\lambda^{\prime}_{\lambda_{1}}) by transposing the corresponding Young diagram. For convenience, partitions and the corresponding Young diagrams are identified. For example, partitions of n=4n=4 are:

(4)=\Yboxdim9pt\yng(4)(3,1)=\Yboxdim9pt\yng(3,1)(22)=\Yboxdim9pt\yng(2,2)(2,12)=\Yboxdim9pt\yng(2,1,1)(14)=\Yboxdim9pt\yng(1,1,1,1)\left(4\right)=\Yboxdim{9pt}\yng(4)\qquad\left(3,1\right)=\Yboxdim{9pt}\yng(3,1)\qquad\left(2^{2}\right)=\Yboxdim{9pt}\yng(2,2)\qquad\left(2,1^{2}\right)=\Yboxdim{9pt}\yng(2,1,1)\qquad\left(1^{4}\right)=\Yboxdim{9pt}\yng(1,1,1,1) (3.25)

with two even partitions (4)(4) and (2,2)(2,2) among them. For dual partitions one has (4)=(14)(4)^{\prime}=(1^{4}), (3,1)=(2,12)(3,1)^{\prime}=(2,1^{2}) and (22)=(22)(2^{2})^{\prime}=(2^{2}). For completeness, one also considers the partition of zero (the empty partition) \varnothing, with the empty set of boxes for the corresponding Young diagram. By definition, the empty partition is even.

The set of all Young diagrams is weakly ordered by inclusion: λμ\lambda\subset\mu implies that any box of λ\lambda is also present in μ\mu. For a pair λμ\lambda\subset\mu define the skew-shape Young diagram μ\λ\mu\backslash\lambda as a set-theoretical difference of the corresponding Young diagrams. We set by definition |μ\λ|=|μ||λ||\mu\backslash\lambda|=|\mu|-|\lambda|. For example,

givenμ=\Yboxdim9pt\yng(4,2,2,1)andλ=\Yboxdim9pt\yng(2,1),one hasμ\λ=\Yboxdim9pt\young(××,×,,)\text{given}\quad\mu=\Yboxdim{9pt}\yng(4,2,2,1)\quad\text{and}\quad\lambda=\Yboxdim{9pt}\yng(2,1)\,,\quad\text{one has}\quad\mu\backslash\lambda=\Yboxdim{9pt}\young(\times\times\leavevmode\nobreak\ \leavevmode\nobreak\ ,\times\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ ) (3.26)

To each box of a (skew-shape) Young diagram at a position (i,j)(i,j) we associate its content c(i,j)=jic(i,j)=j-i. In this respect we define the content of any Young diagram as a sum:

c(λ)=(i,j)λc(i,j).c(\lambda)=\sum_{(i,j)\in\lambda}c(i,j)\,. (3.27)

One defines the content of a skew-shape Young diagram to be c(μ\λ)=c(μ)c(λ)c(\mu\backslash\lambda)=c(\mu)-c(\lambda). For example, for the contents of the skew shape in (3.26) one has

μ\λ=\Yboxdim9pt\young(××23,×0,21,3)c(μ\λ)=1.\mu\backslash\lambda=\Yboxdim{9pt}\young(\times\times\scriptstyle{2}\scriptstyle{3},\times\scriptstyle{0},\scriptstyle{\scalebox{0.4}[1.0]{$-$}2}\scriptstyle{\scalebox{0.4}[1.0]{$-$}1},\scriptstyle{\scalebox{0.4}[1.0]{$-$}3})\quad\Rightarrow\quad c(\mu\backslash\lambda)=-1\,.

Littlewood-Richardson rule.

The \mathbb{N}-span of Young diagrams is endowed with the structure of an associative commutative monoid with the unit element given by \varnothing. For the product of two diagrams λ,ν\lambda,\nu we shall write

λν=μcλνμμ.\lambda\,{\scriptstyle\otimes}\,\nu=\sum_{\mu}c^{\mu}_{\lambda\nu}\,\mu\,. (3.28)

The structure constants cλνμc^{\mu}_{\lambda\nu} are referred to as Littlewood-Richardson coefficients. They are calculated via the Littlewood-Richardson rule, which admits a number of equivalent ways to formulate it in terms of semi-standard tableaux [22]. Let us recall that a tableau of shape μ\λ\mu\backslash\lambda is any map which associates a positive integer to each box of μ\λ\mu\backslash\lambda. A tableau is called semi-standard if the numbers in each row (respectively, column) form a weakly (respectively, strongly) increasing sequence. We will say that a partition ν=(ν1νr)\nu=(\nu_{1}\geqslant\dots\geqslant\nu_{r}) is a weight of a semi-standard tableau 𝗍(μ\λ)\mathsf{t}(\mu\backslash\lambda) if the latter contains exactly νi\nu_{i} occurrences of the entry ii. To any tableau 𝗍\mathsf{t} one associates a row word w(𝗍)w(\mathsf{t}) by reading the entries of boxes along each line from left to right proceeding from the bottom line to the top one. A word is called Yamanouchi word (equivalently, Littlewood-Richardson word or reverse lattice word) if any its suffix contains at least as many 11’s as 22’s, at least as many 22’s as 33’s, etc. For example,

forμ\λ=\Yboxdim8pt\young(××,×,)and the weightν=\Yboxdim8pt\young(,)one has exactly two semistandard tableaux𝗍1=\Yboxdim8pt\young(××11,×1,2),𝗍2=\Yboxdim8pt\young(××11,×2,1),such that the row words are Yamanouchi wordsw(𝗍1)=2111andw(𝗍1)=1211.\begin{array}[]{l}\text{for}\;\;\mu\backslash\lambda=\Yboxdim{8pt}\young(\times\times\leavevmode\nobreak\ \leavevmode\nobreak\ ,\times\leavevmode\nobreak\ ,\leavevmode\nobreak\ )\;\;\text{and the weight}\;\;\nu=\Yboxdim{8pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ )\\ \text{one has exactly two semistandard tableaux}\;\;\mathsf{t}_{1}=\Yboxdim{8pt}\young(\times\times\scriptstyle{1}\scriptstyle{1},\times\scriptstyle{1},\scriptstyle{2})\,,\;\;\mathsf{t}_{2}=\Yboxdim{8pt}\young(\times\times\scriptstyle{1}\scriptstyle{1},\times\scriptstyle{2},\scriptstyle{1})\,,\\ \text{such that the row words are Yamanouchi words}\;\;w(\mathsf{t}_{1})=2111\;\;\text{and}\;\;w(\mathsf{t}_{1})=1211\,.\end{array} (3.29)

With this at hand, one arrives at the following definition of Littlewood-Richardson coefficients:

cλνμis the number of semi-standard tableaux of the shapeμ\λand weightνwhose row word is a Yamanouchi word.\begin{array}[]{ll}c^{\mu}_{\lambda\nu}&\text{is the number of semi-standard tableaux of the shape}\;\;\mu\backslash\lambda\;\;\text{and weight}\;\;\nu\\ \hfill&\text{whose row word is a Yamanouchi word.}\end{array} (3.30)

If either λμ\lambda\not\subset\mu or |ν||μ||λ||\nu|\neq|\mu|-|\lambda|, one puts cλνμ=0c^{\mu}_{\lambda\nu}=0. It appears that cλνμ=cνλμc^{\mu}_{\lambda\nu}=c^{\mu}_{\nu\lambda}, so cλνμ0c^{\mu}_{\lambda\nu}\neq 0 also implies νμ\nu\subset\mu. From the above example one obtains c(2,1),(3,1)(4,2,1)=2c^{(4,2,1)}_{(2,1),(3,1)}=2.

We will also need an equivalent definition of Littlewood-Richardson coefficients based on the jeu de taquin. Recall that a corner of a Young diagram is any box with no other boxes on the right and below. An inside corner of a skew-shape μ\λ\mu\backslash\lambda (with λμ\lambda\subset\mu) is a corner of λ\lambda which is not a corner of μ\mu. Any chosen inner corner of a semi-standard tableau 𝗍\mathsf{t} (thought as an empty box) can be removed by the following sliding process: at any step consider the neighbour(s) on the right and below the empty box, then slide the smallest one into the empty box, while if the two are equal, slide the one below. The process continues until the empty box becomes a corner of μ\mu, and is removed afterwards. The resulting tableau is again semi-standard, so one can repeatedly perform the sliding process until the shape of the semi-standard tableau becomes a Young diagram. The whole process is called jeu de taquin, and the resulting semi-standard tableau 𝗍\mathsf{t}^{\prime} is the same for any order of processing the inner corners. It is called the rectification of 𝗍\mathsf{t}, 𝗍=Rect(𝗍)\mathsf{t}^{\prime}=\mathrm{Rect}(\mathsf{t}). For example, for the tableau 𝗍1\mathsf{t}_{1} in (3.29) one has

𝗍1=Rect(\Yboxdim8pt\young(××11,×1,2))=\Yboxdim8pt\young(111,2),namely,\Yboxdim8pt\young(××11,×1,2)\Yboxdim8pt\young(×111,××,2)\Yboxdim8pt\young(×111,×,2),\Yboxdim8pt\young(×111,×,2)\young(×111,2,×)\young(×111,2),\Yboxdim8pt\young(×111,2)\young(1×11,2)\young(11×1,2)\young(111×,2)\young(111,2).\begin{array}[]{ll}\mathsf{t}^{\prime}_{1}=\mathrm{Rect}\left(\Yboxdim{8pt}\young(\times\times\scriptstyle{1}\scriptstyle{1},\times\scriptstyle{1},\scriptstyle{2})\right)=\Yboxdim{8pt}\young(\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\scriptstyle{2})\,,\;\;\text{namely,}&\Yboxdim{8pt}\young(\times\times\scriptstyle{1}\scriptstyle{1},\times\scriptstyle{1},\scriptstyle{2})\;\;\rightarrow\;\;\Yboxdim{8pt}\young(\times\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\times\times,\scriptstyle{2})\;\;\rightarrow\;\;\Yboxdim{8pt}\young(\times\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\times,\scriptstyle{2})\,,\\ \hfill&\Yboxdim{8pt}\young(\times\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\times,\scriptstyle{2})\;\;\rightarrow\;\;\young(\times\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\scriptstyle{2},\times)\;\;\rightarrow\;\;\young(\times\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\scriptstyle{2})\;\;,\\ \hfill&\Yboxdim{8pt}\young(\times\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\scriptstyle{2})\;\;\rightarrow\;\;\young(\scriptstyle{1}\times\scriptstyle{1}\scriptstyle{1},\scriptstyle{2})\;\;\rightarrow\;\;\young(\scriptstyle{1}\scriptstyle{1}\times\scriptstyle{1},\scriptstyle{2})\;\;\rightarrow\;\;\young(\scriptstyle{1}\scriptstyle{1}\scriptstyle{1}\times,\scriptstyle{2})\;\;\rightarrow\;\;\young(\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\scriptstyle{2})\,.\end{array} (3.31)

One can check that for the other tableau in (3.29) one has the same Rect(𝗍2)=𝗍1\mathrm{Rect}(\mathsf{t}_{2})=\mathsf{t}^{\prime}_{1}.

For any Young diagram ν\nu denote 𝖤(ν)\mathsf{E}(\nu) to be the semi-standard tableau of weight ν\nu, i.e. such that each iith row is filled with ii. Then we arrive at the following equivalent definition of Littlewood-Richardson coefficients:

cλνμis the number of semi-standard tableaux of the shapeμ\λand weightνwhose rectification is 𝖤(ν).\begin{array}[]{ll}c^{\mu}_{\lambda\nu}&\text{is the number of semi-standard tableaux of the shape}\;\;\mu\backslash\lambda\;\;\text{and weight}\;\;\nu\\ \hfill&\text{whose rectification is $\mathsf{E}(\nu)$.}\end{array} (3.32)

By comparing the two examples (3.29) and (3.31), one can verify that both definitions (3.30) and (3.32) lead to the same result c(2,1)(3,1)(4,2,1)=2c^{(4,2,1)}_{(2,1)\,(3,1)}=2.

Let us consider a configuration obtained after a number of sliding processes during the jeu de taquin applied to a semi-standard tableau of a shape μ\λ\mu\backslash\lambda, and let us keep the empty boxes at the end of each sliding process. Then one has a chain of three diagrams τσμ\tau\subset\sigma\subset\mu, where μ\σ\mu\backslash\sigma is the set of empty boxed resulting from the sliding processes, σ\τ\sigma\backslash\tau is a semi-standard tableau, and τ\tau is the set of empty boxes not involved in the performed sliding processes. Let us say that a box of μ\σ\mu\backslash\sigma is an addable corner if adding it to σ\sigma leads to a Young diagram. Note that each addable corner results from a sliding process applied to an inner corner, and that each particular sliding process is invertible. Let us define the reverse sliding process for any addable corner: at any step consider the neighbour(s) on the left and above the empty box, then slide the greater one into the empty box, while if the two are equal, slide the one above. The process continues until the empty box becomes an inner corner. In this respect, any chain τσμ\tau\subset\sigma\subset\mu, with a semi-standard skew-shape diagram σ\τ\sigma\backslash\tau, can be considered as an intermediate configuration of the jeu de taquin, with both types of slidings possible. Upon exhausting direct sliding processes, the unique terminal configuration (the rectification) with τ=\tau=\varnothing was considered above. On the other hand, starting from the terminal configuration of the shape μ\ν\mu\backslash\nu, with the semi-standard tableau 𝖤(ν)\mathsf{E}(\nu), and going backwards by different sequences of reverse slidings until σ=μ\sigma=\mu leads to different semi-standard tableaux 𝗍(init)\mathsf{t}^{(\mathrm{init})} of different shapes μ\λ(init)\mu\backslash\lambda^{(\mathrm{init})}. Define μ𝜐ν\mu\slashdiv\nu to be the set of so obtained diagrams λ(init)\lambda^{(\mathrm{init})}. By construction, the terminal configuration 𝖤(ν)=Rect(𝗍(init))\mathsf{E}(\nu)=\mathrm{Rect}\big{(}\mathsf{t}^{(\mathrm{init})}\big{)} is the same for all 𝗍(init)\mathsf{t}^{(\mathrm{init})}, so according to the definition (3.32), the set μ𝜐ν\mu\slashdiv\nu contains such diagrams λ\lambda that cλνμ0c^{\mu}_{\lambda\nu}\neq 0, and only them. For example, keeping empty boxes upon constructing Rect(𝗍1)\mathrm{Rect}(\mathsf{t}_{1}) in (3.31) and applying different sequences of reverse sliding processes gives three skew-shape diagrams:

\Yboxdim8pt\young(111×,2×,×)rev. slides\Yboxdim8pt\young(×××1,11,2),\Yboxdim8pt\young(××11,×1,2),\Yboxdim8pt\young(××11,×2,1),\Yboxdim8pt\young(×111,×2,×),thus\Yboxdim8pt\young(,,)𝜐\young(,)={\young(,,),\young(,),\young()}.\Yboxdim{8pt}\young(\scriptstyle{1}\scriptstyle{1}\scriptstyle{1}\times,\scriptstyle{2}\times,\times)\;\;\xrightarrow{\text{rev. slides}}\;\;\Yboxdim{8pt}\young(\times\times\times\scriptstyle{1},\scriptstyle{1}\scriptstyle{1},\scriptstyle{2})\,,\;\;\Yboxdim{8pt}\young(\times\times\scriptstyle{1}\scriptstyle{1},\times\scriptstyle{1},\scriptstyle{2})\,,\;\;\Yboxdim{8pt}\young(\times\times\scriptstyle{1}\scriptstyle{1},\times\scriptstyle{2},\scriptstyle{1})\,,\;\;\Yboxdim{8pt}\young(\times\scriptstyle{1}\scriptstyle{1}\scriptstyle{1},\times\scriptstyle{2},\times)\,,\quad\text{thus}\quad\Yboxdim{8pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ )\slashdiv\young(\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ )=\left\{\young(\leavevmode\nobreak\ ,\leavevmode\nobreak\ ,\leavevmode\nobreak\ )\,,\;\young(\leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ )\,,\;\young(\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ )\right\}\,. (3.33)

Simple Bn(εN)B_{n}(\varepsilon N)- and 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules in VnV^{\otimes n}.

The simple (left) Bn(δ)B_{n}(\delta)-modules Mn(λ)M^{(\lambda)}_{n} are indexed by particular Young diagrams λ\lambda with |λ|=n2f|\lambda|=n-2f, with f=0,,n2f=0,\dots,\lfloor\frac{n}{2}\rfloor [23, Corollary 3.5]. We are interested in singling out only those of them which occur in VnV^{\otimes n} in the context of Schur-Weyl-type duality (2.12), when δ=εN\delta=\varepsilon N. To do so, for a given ll\in\mathbb{N} consider the set

Λl,N={λl:λ1+λ2N(for the dual partition λ) when G(N)=O(N), andλ1N/2when G(N)=Sp(N)}.\begin{array}[]{rcl}\Lambda_{l,N}=&\big{\{}\,\lambda\vdash l\;\;:&\lambda_{1}^{\prime}+\lambda_{2}^{\prime}\leqslant N\;\;\text{(for the dual partition $\lambda^{\prime}$) when $G(N)=O(N)$, and}\\ \hfill\hfil&\hfill\hfil&\lambda_{1}\leqslant N/2\;\;\text{when $G(N)=Sp(N)$}\big{\}}\,.\end{array} (3.34)

Then a Bn(εN)B_{n}(\varepsilon N)-module Mn(λ)M^{(\lambda)}_{n} appears in VnV^{\otimes n} iff λΛn2f,N\lambda\in\Lambda_{n-2f,N} for some f=0,,n2f=0,\dots,\lfloor\tfrac{n}{2}\rfloor [23, 24, 16].

The restrictions on Young diagrams in Λl,N\Lambda_{l,N} can be interpreted as the absence of certain Bn(εN)B_{n}(\varepsilon N)-modules in VnV^{\otimes n}, which occurs due to non-injectivity of 𝔯\mathfrak{r}. It is known that 𝔯\mathfrak{r} is injective for NnN\geqslant n when G(N)=O(N)G(N)=O(N) and for N2nN\geqslant 2n when G(N)=Sp(N)G(N)=Sp(N) [14], and this is exactly where the constraints trivialise. In the sequel, unless otherwise is specified, simple Bn(εN)B_{n}(\varepsilon N)-modules will be viewed as submodules in VnV^{\otimes n} (with n2n\geqslant 2 and N=dimVN=\dim V).

We will also need some basic facts about the representation theory of 𝔖nBn(εN)\mathbb{C}\mathfrak{S}_{n}\subset B_{n}(\varepsilon N). Namely, that simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules L(μ)L^{(\mu)} are indexed by Young diagrams with |μ|=n|\mu|=n. Those of them which occur in VnV^{\otimes n} carry a label from the following set:

Σn,N={μn:μ1Nfor ε=1, andμ1Nfor ε=1},\begin{array}[]{rcl}\Sigma_{n,N}=&\big{\{}\,\mu\vdash n\;\;:&\mu_{1}^{\prime}\leqslant N\;\;\text{for $\varepsilon=1$, and}\\ \hfill\hfil&\hfill\hfil&\mu_{1}\leqslant N\;\;\text{for $\varepsilon=-1$}\big{\}}\,,\end{array} (3.35)

where ε\varepsilon distinguishes between the two ways for permutations to act on VnV^{\otimes n} according to (3.14).

The sets Λ^n2f,N\hat{\Lambda}_{n-2f,N} and Σ^n,N\hat{\Sigma}_{n,N} in (2.9) and (2.12) are obtained from Λn2f,N\Lambda_{n-2f,N} and Σn,N\Sigma_{n,N} via element-wise application of the following ε\varepsilon-dependent mapping:

for any partition λ,λ^=λ when ε=1 (in particular, when G(N)=O(N)), andλ^=λ when ε=1 (in particular, when G(N)=Sp(N)).\begin{array}[]{rl}\text{for any partition $\lambda$,}&\text{$\hat{\lambda}=\lambda$ when $\varepsilon=1$ (in particular, when $G(N)=O(N)$), and}\\ \hfill\hfil&\text{$\hat{\lambda}=\lambda^{\prime}$ when $\varepsilon=-1$ (in particular, when $G(N)=Sp(N)$).}\end{array} (3.36)

Restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n}.

Upon restriction to the subgroup G(N)GL(N)G(N)\subset GL(N), the simple GL(N)GL(N)-modules decompose into a direct sum of simple G(N)G(N)-modules. Similarly, the simple Bn(εN)B_{n}(\varepsilon N)-modules decompose into a direct sum of simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules upon restriction to the subalgebra 𝔖nBn(εN)\mathbb{C}\mathfrak{S}_{n}\subset B_{n}(\varepsilon N). The two restrictions are related via comparing the two decompositions (2.12) of the same space VnV^{\otimes n} (see, e.g., [21, Lemma 4.2]):

ifV(μ)ν^even,ν1N(D(λ))c¯νλμ(upon GL(N)G(N))thenMn(λ^)ν^even,ν1N(L(μ^))c¯νλμ(upon Bn(εN)𝔖n),and vice versa.\begin{array}[]{l}\displaystyle\text{if}\quad V^{(\mu)}\cong\bigoplus_{\begin{array}[]{c}{\scriptstyle\hat{\nu}\;\;\text{even}\,,}\\ {\scriptstyle\nu^{\prime}_{1}\leqslant N}\end{array}}\big{(}D^{(\lambda)}\big{)}^{\oplus\overline{c}^{\mu}_{\nu\lambda}}\quad\text{(upon $GL(N)\downarrow G(N)$)}\\ \text{then}\displaystyle\quad M^{(\hat{\lambda})}_{n}\cong\bigoplus_{\begin{array}[]{c}{\scriptstyle\hat{\nu}\;\;\text{even}\,,}\\ {\scriptstyle\nu^{\prime}_{1}\leqslant N}\end{array}}\big{(}L^{(\hat{\mu})}\big{)}^{\oplus\overline{c}^{\mu}_{\nu\lambda}}\quad\text{(upon $B_{n}(\varepsilon N)\downarrow\mathbb{C}\mathfrak{S}_{n}$)}\,,\quad\text{and {\it vice versa}}.\end{array} (3.37)

When μΛ^n,NΣ^n,N\mu\in\hat{\Lambda}_{n,N}\subset\hat{\Sigma}_{n,N} one has c¯νλμ=cνλμ\overline{c}^{\mu}_{\nu\lambda}=c^{\mu}_{\nu\lambda} (the Littlewood-Richardson coefficients): this is the case when the Littlewood restriction rules apply [25]. The branching rules on the left-hand-side of (3.37) are extensively presented in the literature [25, 26, 27, 28]. In particular, note the combinatorial approach proposed in [29, Theorem 4.17 and Remark 4.19], where in the case μΛ^n,N\mu\notin\hat{\Lambda}_{n,N} the multiplicities c¯νλμ\overline{c}^{\mu}_{\nu\lambda} are defined via additional NN-dependent constraints on the tableaux in the definition (3.30).

In this respect, the subset of labels Λn,NΣn,N\Lambda_{n,N}\subset\Sigma_{n,N} will be called Littlewood-admissible. By applying Schur-Weyl duality (2.12) in this case, one can detect the occurrence L(μ)Mn(λ)L^{(\mu)}\subset M^{(\lambda)}_{n} upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n}. In more detail, for any integer f0f\geqslant 0 define the following set of Young diagrams (the closure):

clN(f)(λ)={μλ:μΣn,Nand|μ\λ|=2f,cνλμ0for some evenνwith|ν|=2f}.\mathrm{cl}^{(f)}_{N}(\lambda)=\left\{\mu\supset\lambda\;\;:\;\;\mu\in\Sigma_{n,N}\;\;\text{and}\;\;|\mu\backslash\lambda|=2f\,,\;\;c^{\mu}_{\nu\lambda}\neq 0\;\;\text{for some even}\;\;\nu\;\;\text{with}\;\;|\nu|=2f\right\}\,. (3.38)

When Bn(εN)B_{n}(\varepsilon N) is semisimple (i.e. when Nn1N\geqslant n-1 for ε=1\varepsilon=1 and N2(n1)N\geqslant 2(n-1) for ε=1\varepsilon=-1 [14]), the constraints on the size of diagrams in (3.38) trivialise when λΛn2f\lambda\in\Lambda_{n-2f} with f1f\geqslant 1. Occurrence of a particular simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-module in a given simple Bn(εN)B_{n}(\varepsilon N)-module Mn(λ)M^{(\lambda)}_{n} can be analysed via the following lemma.

Lemma 3.2.

Let Mn(λ)M^{(\lambda)}_{n} be a simple Bn(εN)B_{n}(\varepsilon N)-module, with |λ|=n2f|\lambda|=n-2f. Upon its restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n},

ifL(μ)Mn(λ)(with|μ|=n)thenμclN(f)(λ).\text{if}\;\;L^{(\mu)}\subset M^{(\lambda)}_{n}\;\;(\text{with}\;\;|\mu|=n)\;\;\text{then}\;\;\mu\in\mathrm{cl}^{(f)}_{N}(\lambda)\,.

The converse is also true whenever one of the possibilities hold:

  • i)

    μΛn,N\mu\in\Lambda_{n,N} (i.e. μ\mu is Littlewood-admissible),

  • ii)

    Bn(εN)B_{n}(\varepsilon N) is semisimple.

Proof.

We need to recall some basic facts about the standard Bn(δ)B_{n}(\delta)-modules Δn(λ)\Delta^{(\lambda)}_{n}, where δ\delta\in\mathbb{C} (see [3, formula (2.5)] and references therein, more detailed discussion is postponed to Section 6). Standard modules are labelled by all partitions λn2f\lambda\vdash n-2f for f=0,,n2f=0,\dots,\lfloor{\frac{n}{2}\rfloor}. Each Δn(λ)\Delta^{(\lambda)}_{n} is an indecomposable module (but not necessarily simple when δ\delta\in\mathbb{Z}), such that Mn(λ)Δn(λ)/KM^{(\lambda)}_{n}\cong\Delta^{(\lambda)}_{n}/K for the maximal proper submodule KK. In the semisimple regime of the Brauer algebra standard modules are simple, so Δn(λ)Mn(λ)\Delta^{(\lambda)}_{n}\cong M^{(\lambda)}_{n}.

The decomposition of standard modules into simple summands upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n} was described in [30, Theorem 4.1], which implies L(μ)Δn(λ)L^{(\mu)}\subset\Delta^{(\lambda)}_{n} iff cνλμ0c^{\mu}_{\nu\lambda}\neq 0 for some even ν\nu. Moreover, for the Littlewood-Richardson coefficient to be non-zero, λμ\lambda\subset\mu and |ν|=|μ||λ|=2f|\nu|=|\mu|-|\lambda|=2f. The additional restrictions on the number of rows/columns in the definition of clN(f)(λ)\mathrm{cl}^{(f)}_{N}(\lambda) reflect the restrictions on the 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules appearing in VnV^{\otimes n} according to the classical Schur-Weyl duality [1].

To finish the proof, for μΛn,N\mu\in\Lambda_{n,N} the multiplicities of L(μ)L^{(\mu)} in the branching rules (3.37) are Littlewood-Richardson coefficients. ∎

Eigenvalues of AnA_{n}.

In what follows we make use of the results of [24] (see also [3] in order to cover all δ\delta\in\mathbb{Z}). Consider the following set of pairwise-commuting elements known as Jucys-Murphy elements xkx_{k} (k=1,,nk=1,\dots,n):

x1=εN12andxk=εN12+j=1k1(sjkdjk)fork2.x_{1}=\frac{\varepsilon N-1}{2}\quad\text{and}\quad x_{k}=\frac{\varepsilon N-1}{2}+\sum_{j=1}^{k-1}(s_{jk}-d_{jk})\quad\text{for}\quad k\geqslant 2\,. (3.39)

Among the applications of the latter in the context of the present work, we mention that they can be used for constructing the maximal commutative subalgebra in Bn(εN)B_{n}(\varepsilon N), whose elements are diagonalisable on Bn(εN)B_{n}(\varepsilon N)-modules in VnV^{\otimes n}. We concentrate on the following central element

XB=k=1nxk=nεN12+XSAnwhereXS=i<jsijand (3.21) was used.X_{B}=\sum_{k=1}^{n}x_{k}=n\,\frac{\varepsilon N-1}{2}+X_{S}-A_{n}\,\quad\text{where}\quad X_{S}=\sum_{i<j}s_{ij}\quad\text{and \eqref{eq:master_class} was used.} (3.40)

Its value on the simple Bn(εN)B_{n}(\varepsilon N)-module Mn(λ)M^{(\lambda)}_{n} is (n2f)εN12+c(λ)(n-2f)\,\frac{\varepsilon N-1}{2}+c(\lambda). The term XSX_{S} is a sum of Jucys-Murphy elements in the algebra 𝔖n\mathbb{C}\mathfrak{S}_{n} [31, 19]: it is central in 𝔖n\mathbb{C}\mathfrak{S}_{n} and proportional to identity on any simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-module L(μ)L^{(\mu)} with the coefficient c(μ)c(\mu) (a possible way to check this is to apply (3.40) to Mn(μ)L(μ)M^{(\mu)}_{n}\cong L^{(\mu)} taking into account that dij(Mn(μ))=0d_{ij}\big{(}M^{(\mu)}_{n}\big{)}=0 for μn\mu\vdash n, see [24]).

Define the following set of skew-shape diagrams:

Λ¯n,N(f)=λΛn2f,NμclN(f)(λ)μ\λ.\overline{\Lambda}^{(f)}_{n,N}=\bigcup_{\lambda\in\Lambda_{n-2f,N}}\bigcup_{\mu\in\mathrm{cl}^{(f)}_{N}(\lambda)}\mu\backslash\lambda\,. (3.41)

Variations of the following lemma are known in the literature (see, e.g., the proofs in [24, Theorem 2.6] or [3, Proposition 4.2] and references therein), nevertheless we give the proof to keep the narrative self-contained.

Lemma 3.3.

Let Mn(λ)M^{(\lambda)}_{n} be a simple Bn(εN)B_{n}(\varepsilon N)-module, and let a simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-module L(μ)L^{(\mu)} occur in the decomposition of Mn(λ)M^{(\lambda)}_{n} into irreducible summands upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n}. Then μ\λΛ¯n,N(f)\mu\backslash\lambda\in\overline{\Lambda}_{n,N}^{(f)} and

for anyvL(μ),An(v)=((εN1)f+c(μ\λ))v.\text{for any}\quad v\in L^{(\mu)}\,,\quad A_{n}(v)=\big{(}(\varepsilon N-1)\,f+c(\mu\backslash\lambda)\big{)}\,v\,. (3.42)
Proof.

The relation (3.40), together with the fact that XBX_{B} and XSX_{S} are both proportional to identity on L(μ)Mn(λ)L^{(\mu)}\subset M^{(\lambda)}_{n}, implies that AnA_{n} is proportional to identity on L(μ)L^{(\mu)} as well. According to the structure of the decomposition of Mn(λ)M^{(\lambda)}_{n} upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n} (see Lemma 3.2), one has λμ\lambda\subset\mu, so c(μ)c(λ)=c(μ\λ)c(\mu)-c(\lambda)=c(\mu\backslash\lambda), and the eigenvalue in the assertion is a direct consequence of (3.40) for XBX_{B} (respectively, XSX_{S}) restricted to Mn(λ)M^{(\lambda)}_{n} (respectively, to L(μ)L^{(\mu)}). ∎

Together with the fact that VnV^{\otimes n} decomposes as a direct sum of simple Bn(εN)B_{n}(\varepsilon N)-modules, the above lemma provides an alternative proof (along with Lemma 3.1) that AnA_{n} is diagonalisable on VnV^{\otimes n}. In order to describe the eigenvalues, define the set

spec×(An)={(εN1)f+c(μ\λ):f=1,,n2,μ\λΛ¯n2f,N(f)}ε.\mathrm{spec}^{\times}(A_{n})=\big{\{}(\varepsilon N-1)\,f+c(\mu\backslash\lambda)\;\;:\;\;f=1,\dots,\lfloor\tfrac{n}{2}\rfloor\,,\quad\mu\backslash\lambda\in\overline{\Lambda}^{(f)}_{n-2f,N}\big{\}}\cap\varepsilon\mathbb{N}\,. (3.43)
Proposition 3.4.

Any non-zero eigenvalue of AnA_{n} on VnV^{\otimes n} is contained in spec×(An)\mathrm{spec}^{\times}(A_{n}). When Bn(εN)B_{n}(\varepsilon N) is semisimple, any element of spec×(An)\mathrm{spec}^{\times}(A_{n}) is a non-zero eigenvalue of AnA_{n} on VnV^{\otimes n}.

Proof.

Lemma 3.2 gives the necessary condition for occurrence of a simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-module L(μ)L^{(\mu)} in Mn(λ)M^{(\lambda)}_{n} upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n}, while Lemma 3.3 gives the corresponding eigenvalue αμ\λ\alpha_{\mu\backslash\lambda} of AnA_{n}. According to Lemma 3.1 one has: i) zero eigenvalues correspond to the components D(μ^)Mn(μ)D^{(\hat{\mu})}\otimes M^{(\mu)}_{n} with μΛn,N\mu\in\Lambda_{n,N} in (2.12), and thus to f=0f=0 in (3.42), ii) non-zero eigenvalues are in ε\varepsilon\mathbb{N}. Therefore, if αμ\λ\alpha_{\mu\backslash\lambda} is a non-zero eigenvalue of AnA_{n} on VnV^{\otimes n}, then necessarily it is in spec×(An)\mathrm{spec}^{\times}(A_{n}). When Bn(εN)B_{n}(\varepsilon N) is semisimple, Lemma 3.2 becomes a criterion, so any element in (3.43) comes from L(μ)Mn(λ)L^{(\mu)}\subset M^{(\lambda)}_{n} for some λΛn2f,N\lambda\in\Lambda_{n-2f,N} (f1f\geqslant 1) and μclN(f)(λ)\mu\in\mathrm{cl}^{(f)}_{N}(\lambda). ∎

Note that imposing intersection with ε\varepsilon\mathbb{N} is substantial in the definition of spec×(An)\mathrm{spec}^{\times}(A_{n}) in (3.43) in order to reduce the number of elements which are not eigenvalues of AnA_{n}. For example, consider N=4N=4 and ε=1\varepsilon=1, and take λ=(22)Λ4,4\lambda=(2^{2})\in\Lambda_{4,4}, μ=(24)cl4(2)(λ)\mu=(2^{4})\in\mathrm{cl}^{(2)}_{4}(\lambda). One has (N1)f+c(μ\λ)=2(N-1)f+c(\mu\backslash\lambda)=-2 which is not an eigenvalue of AnA_{n} due to Lemma 3.1 (in other words, L(μ)Mn(λ)L^{(\mu)}\not\subset M^{(\lambda)}_{n}). To this end, let us note that the relevant fact here is that μ\mu is not Littlewood-admissible. A more detailed discussion is postponed to Section 4.2 in relation to traceless projection of simple GL(N)GL(N)-modules.

Remark (on the structure of Bn(εN)B_{n}(\varepsilon N)-modules via Schur-Weyl duality).

In the semisimple regime of Bn(εN)B_{n}(\varepsilon N) one can prove Lemma 3.1 within the representation theory of the Brauer algebra (see Appendix B.2). While outside the semisimple regime (N<n1N<n-1 for ε=1\varepsilon=1 and N<2(n1)N<2(n-1) for ε=1\varepsilon=-1, see [32]) the following corollary of Lemmas 3.1 and 3.3 seems to be hard to prove without addressing to tensorial representations via Schur-Weyl duality (see Appendix B.3 for proof and recall the notion of standard Bn(δ)B_{n}(\delta)-modules in the proof of Lemma 3.2).

Corollary 3.5.

Fix N1N\geqslant 1, ε=±1\varepsilon=\pm 1 and n2n\geqslant 2 for Bn(εN)B_{n}(\varepsilon N). Take any λΛn,N\lambda\in\Lambda_{n,N} and μclN(f)(λ)\mu\in\mathrm{cl}^{(f)}_{N}(\lambda) for 2f=n|λ|2f=n-|\lambda| such that L(μ)Mn(λ)L^{(\mu)}\subset M^{(\lambda)}_{n} upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n} and denote αμ\λ\alpha_{\mu\backslash\lambda} the corresponding eigenvalue in Lemma 3.3. Then the following assertions hold:

  • i)

    αμ\λ0\alpha_{\mu\backslash\lambda}\geqslant 0,

  • ii)

    αμ\λ=0\alpha_{\mu\backslash\lambda}=0 iff λ=μ\lambda=\mu.

As a result, for the standard module Δn(λ)\Delta^{(\lambda)}_{n} the following condition is sufficient to conclude that L(μ)L^{(\mu)} does not occur in the simple head 666To recall the notion of a simple head for the situation in question, note that a standard Bn(δ)B_{n}(\delta)-module is indecomposable, so the action of Bn(δ)B_{n}(\delta) admits a block upper-triangular form. The simple head is then a simple module obtained by projection to the lowest block on the diagonal. of Δn(λ)\Delta^{(\lambda)}_{n}:

eitherαμ\λ<0orαμ\λ=0for|λ|<n.\text{either}\quad\alpha_{\mu\backslash\lambda}<0\quad\text{or}\quad\alpha_{\mu\backslash\lambda}=0\;\;\text{for}\;\;|\lambda|<n\,. (3.44)

The above Corollary does not hold outside the domain of applicability of Schur-Weyl duality. For example, consider B6(δ)B_{6}(\delta) with δ=12\delta=\tfrac{1}{2}\notin\mathbb{Z}. For non-integer values of the parameter the algebra is semisimple [23, 32], so L(3,2,1)M6(2)L^{(3,2,1)}\subset M^{(2)}_{6} (upon restriction to 𝔖6\mathbb{C}\mathfrak{S}_{6}) by Lemma 3.2, while A6(L(3,2,1))=0A_{6}\big{(}L^{(3,2,1)}\big{)}=0 by Lemma 3.3.

4 Traceless projectors

4.1 Traceless projection of VnV^{\otimes n}

Consider the following element in the centralizer of 𝔖n\mathbb{C}\mathfrak{S}_{n} in Bn(εN)B_{n}(\varepsilon N):

Pn=αspec×(An)(11αAn)Cn(εN),and define𝔓n=𝔯(Pn)𝔅n(N).P_{n}=\prod_{\alpha\in\mathrm{spec}^{\times}(A_{n})}\left(1-\frac{1}{\alpha}\,A_{n}\right)\in C_{n}(\varepsilon N)\,,\quad\text{and define}\quad\mathfrak{P}_{n}=\mathfrak{r}(P_{n})\in\mathfrak{B}_{n}(N)\,. (4.1)

Due to (3.15) and (3.21), the property (𝔓n)=𝔓n(\mathfrak{P}_{n})^{*}=\mathfrak{P}_{n} is manifest. The main result is summarised in the following theorem.

Theorem 4.1.

𝔓n\mathfrak{P}_{n} is the universal traceless projector on VnV^{\otimes n}, which satisfies the properties (2.13).

Proof.

To prove that 𝔓n\mathfrak{P}_{n} is a projector, consider the decomposition of VnV^{\otimes n} into a direct sum of irreducible Bn(εN)B_{n}(\varepsilon N)-modules Mn(λ)M^{(\lambda)}_{n}, and then decompose each of them into a direct sum of 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules L(μ)Mn(λ)L^{(\mu)}\subset M^{(\lambda)}_{n}. AnA_{n} is block-diagonal with respect to this decomposition with possible non-zero eigenvalues described by Proposition 3.4. The zero eigenvalue exactly marks the traceless subspace of VnV^{\otimes n}, which is due to Lemma 3.1. So by design of (4.1), it annihilates all 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules marked by non-zero eigenvalues of AnA_{n} (there is no problem if some elements of spec×(An)\mathrm{spec}^{\times}(A_{n}) do not show up in VnV^{\otimes n}). As a result,

AnPn(Vn)=0,thereforePn(Vn)2=Pn(Vn),A_{n}P_{n}(V^{\otimes n})=0\,,\quad\text{therefore}\quad P_{n}{}^{2}(V^{\otimes n})=P_{n}(V^{\otimes n})\,, (4.2)

so the properties (P1)(\mathrm{P}1) and (P2)(\mathrm{P}2) of (2.13) hold. The property (P3)(\mathrm{P}3) of (2.13) is manifest due to the fact that PnCn(εN)P_{n}\in C_{n}(\varepsilon N). ∎

Let us illustrate the application of Theorem 4.1 in some simple cases. In Examples 1 and 2 we consider G(N)=O(N)G(N)=O(N), so ε=1\varepsilon=1.

Example 1 (one-dimensional space).

Let us verify that for N=dimV=1N=\dim V=1 the projector trivialises: 𝔓n=0\mathfrak{P}_{n}=0 for all n2n\geqslant 2 (or, equivalently, that Pn(Vn)=0P_{n}(V^{\otimes n})=0). Note that in this case for any TVnT\in V^{\otimes n}, dij(T)=Td_{ij}(T)=T (for all 1i<jn1\leqslant i<j\leqslant n), and hence

An(T)=n(n1)2T.A_{n}(T)=\frac{n(n-1)}{2}\,T\,. (4.3)

Thus, in order to prove that Pn(T)=0P_{n}(T)=0 it suffices to show that n(n1)2spec×(An)\frac{n(n-1)}{2}\in\mathrm{spec}^{\times}(A_{n}). But this is directly what expression (4.3) says. Let us cross-check this straightforward conclusion by computing the eigenvalues of AnA_{n} from the representation-theoretic point of view. Constraints on the lengths of columns leaves the only possibility λ=\lambda=\varnothing for nn even and λ=\Yboxdim9pt\yng(1)\lambda=\Yboxdim{9pt}\yng(1) for nn odd, and ν=(2fmax)\nu=(2f_{\mathrm{max}}) with fmax=n2f_{\mathrm{max}}=\lfloor\frac{n}{2}\rfloor. On the right-hand side of λν\lambda\,{\scriptstyle\otimes}\,\nu there is the only partition μ=(n)\mu=(n) which belongs to cl1(fmax)(λ)\mathrm{cl}_{1}^{(f_{\mathrm{max}})}(\lambda), so

\ytableausetupmathmode,boxsize=1.1em,centertableauxμ\λ={ytableau}0&1\none[]n1(for n even)andμ\λ=\ytableausetupmathmode,boxsize=1.1em,centertableaux{ytableau}×&1\none[]n1(for n odd)c(μ\λ)=n(n1)2.\ytableausetup{mathmode,boxsize=1.1em,centertableaux}\mu\backslash\lambda=\ytableau{\scriptstyle}\scriptstyle{0}&\scriptstyle{1}\none[\scriptstyle{\cdots}]\scriptstyle{n\scalebox{0.4}[1.0]{$-$}1}\quad\text{(for $n$ even)}\quad\text{and}\quad\mu\backslash\lambda=\ytableausetup{mathmode,boxsize=1.1em,centertableaux}\ytableau{\scriptstyle}\times&\scriptstyle{1}\none[\scriptstyle{\cdots}]\scriptstyle{n\scalebox{0.4}[1.0]{$-$}1}\quad\text{(for $n$ odd)}\quad\Rightarrow\quad c(\mu\backslash\lambda)=\tfrac{n(n-1)}{2}\,.

Example 2 (lower-rank projectors).

In the sequel we assume N2N\geqslant 2. Let us derive the basic well-known example: the traceless projector for n=2n=2. In this case spec×(A2)\mathrm{spec}^{\times}(A_{2}) is constituted by the only value f=1f=1 in (3.43) for λ=\lambda=\varnothing and ν=\Yboxdim9pt\young()\nu=\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ ):

μ=\Yboxdim9pt\young()=\young()μ\λ=\young(01),α=(N1)+c(μ\λ)=N,\mu=\varnothing\,{\scriptstyle\otimes}\,\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ )=\young(\leavevmode\nobreak\ \leavevmode\nobreak\ )\quad\Rightarrow\quad\mu\backslash\lambda=\young(\scriptstyle{0}\scriptstyle{1})\,,\quad\alpha=(N-1)+c(\mu\backslash\lambda)=N\,,

so one immediately arrives at the expected result

P2=11NA2=[Uncaptioned image]1N[Uncaptioned image]P_{2}=1-\frac{1}{N}\,A_{2}=\,\raisebox{-0.4pt}{\includegraphics[width=15.0pt,height=20.0pt]{Id2.pdf}}\,-\frac{1}{N}\ \raisebox{-0.4pt}{\includegraphics[width=15.0pt,height=20.0pt]{d1.pdf}} (4.4)

The case n=3n=3 is slightly more cumbersome to construct from scratch (by solving the traceless Ansatz for a tensor), but is still elementary from the point of view of B3(N)B_{3}(N). As in the previous example, spec×(A3)\mathrm{spec}^{\times}(A_{3}) is obtained for f=1f=1 in (3.43) (again, ν=\Yboxdim9pt\young()\nu=\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ )), with

λ=\Yboxdim9pt\young(),μ={\Yboxdim9pt\young()\Yboxdim9pt\young(,)μ\λ={\Yboxdim9pt\young(×12),α=(N1)+3\Yboxdim9pt\young(×1,1),α=(N1)+0\lambda=\Yboxdim{9pt}\young(\leavevmode\nobreak\ )\,,\quad\mu=\left\{\begin{array}[]{l}\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ )\\ \Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ )\end{array}\right.\quad\Rightarrow\quad\mu\backslash\lambda=\left\{\begin{array}[]{ll}\Yboxdim{9pt}\young(\times\scriptstyle{1}\scriptstyle{2})\,,&\alpha=(N-1)+3\\ \Yboxdim{9pt}\young(\times\scriptstyle{1},\scriptstyle{\scalebox{0.4}[1.0]{$-$}1})\,,&\alpha=(N-1)+0\end{array}\right. (4.5)

The operator (4.1) takes the form

P3\displaystyle P_{3} =(11N1A3)(11N+2A3)\displaystyle=\left(1-\frac{1}{N-1}\,A_{3}\right)\left(1-\frac{1}{N+2}\,A_{3}\right) (4.6)
=12N+1(N1)(N+2)A3+1(N1)(N+2)(A3)2\displaystyle=1-\frac{2N+1}{(N-1)(N+2)}\,A_{3}+\frac{1}{(N-1)(N+2)}\,(A_{3})^{2}
=[Uncaptioned image]N+1(N1)(N+2)([Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image])+1(N1)(N+2)([Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image])\displaystyle=\scalebox{0.8}{$\raisebox{-0.4pt}{\includegraphics[width=15.0pt,height=20.0pt]{Id3.pdf}}$}-\frac{N+1}{(N-1)(N+2)}\,\Bigl{(}\ \scalebox{0.8}{$\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_1.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_2.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_3.pdf}}\ $}\Bigr{)}\,+\frac{1}{(N-1)(N+2)}\,\ \Bigl{(}\scalebox{0.8}{$\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_4.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_5.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_6.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_7.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_8.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_9.pdf}}$}\Bigr{)}\,

Note that the denominators in the above expression are singular at N=1N=1, which is not a problem according to Theorem 4.1: one simply omits the factor with the eigenvalue which turns to 0 upon putting N=1N=1, which brings us back to Example 1.

In the following two examples we consider G(N)=Sp(N)G(N)=Sp(N), so ε=1\varepsilon=-1.

Example 3 (lower-rank projectors).

Let us reproduce the projector in the obvious case n=2n=2. There is the only element in spec×(An)\mathrm{spec}^{\times}(A_{n}) obtained for f=1f=1 from λ=\lambda=\varnothing and ν=\Yboxdim9pt\young()\nu=\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ ) :

μ=\Yboxdim9pt\young()=\young()μ\λ=\young(01),(N+1)f+c(μ\λ)=N,\mu=\varnothing\,{\scriptstyle\otimes}\,\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ )=\young(\leavevmode\nobreak\ \leavevmode\nobreak\ )\quad\Rightarrow\quad\mu\backslash\lambda=\young(\scriptstyle{0}\scriptstyle{1})\,,\;\;-(N+1)\,f+c(\mu\backslash\lambda)=-N\,,

so one immediately arrives at the expected form of the projector:

P2=1+1NA2=[Uncaptioned image]+1N[Uncaptioned image]P_{2}=1+\frac{1}{N}\,A_{2}=\,\raisebox{-0.4pt}{\includegraphics[width=15.0pt,height=20.0pt]{Id2.pdf}}\,+\frac{1}{N}\ \raisebox{-0.4pt}{\includegraphics[width=15.0pt,height=20.0pt]{d1.pdf}} (4.7)

For n=3n=3 suppose N>2N>2 (the case N=2N=2 will be considered below), then there are two diagrams contributing to spec×(An)\mathrm{spec}^{\times}(A_{n}) arising from λ=\Yboxdim9pt\young()\lambda=\Yboxdim{9pt}\young(\leavevmode\nobreak\ ) and ν=\Yboxdim9pt\young()\nu=\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ ) (f=1f=1):

λν=\Yboxdim9pt\young(,)+\young(),soμ={\Yboxdim9pt\young()\Yboxdim9pt\young(,)μ\λ={\Yboxdim9pt\young(×12),α=(N+1)+3\Yboxdim9pt\young(×1,1),α=(N+1)+0\lambda\,{\scriptstyle\otimes}\,\nu=\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ )+\young(\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ )\,,\;\;\text{so}\quad\mu=\left\{\begin{array}[]{l}\Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ )\\ \Yboxdim{9pt}\young(\leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ )\end{array}\right.\quad\Rightarrow\quad\mu\backslash\lambda=\left\{\begin{array}[]{ll}\Yboxdim{9pt}\young(\times\scriptstyle{1}\scriptstyle{2})\,,&\alpha=-(N+1)+3\\ \Yboxdim{9pt}\young(\times\scriptstyle{1},\scriptstyle{\scalebox{0.4}[1.0]{$-$}1})\,,&\alpha=-(N+1)+0\end{array}\right. (4.8)

The operator (4.1) takes the form

P3\displaystyle P_{3} =(1+1N2A3)(1+1N+1A3)\displaystyle=\left(1+\frac{1}{N-2}\,A_{3}\right)\left(1+\frac{1}{N+1}\,A_{3}\right) (4.9)
=1+2N1(N2)(N+1)A3+1(N2)(N+1)(A3)2\displaystyle=1+\frac{2N-1}{(N-2)(N+1)}\,A_{3}+\frac{1}{(N-2)(N+1)}\,(A_{3})^{2}
=[Uncaptioned image]+N1(N2)(N+1)([Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image])+1(N2)(N+1)([Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image])\displaystyle=\scalebox{0.8}{$\raisebox{-0.4pt}{\includegraphics[width=15.0pt,height=20.0pt]{Id3.pdf}}$}+\frac{N-1}{(N-2)(N+1)}\,\Bigl{(}\ \scalebox{0.8}{$\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_1.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_2.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_3.pdf}}\ $}\Bigr{)}\,+\frac{1}{(N-2)(N+1)}\,\ \Bigl{(}\scalebox{0.8}{$\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_4.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_5.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_6.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_7.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_8.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_9.pdf}}$}\Bigr{)}\,

Example 4 (two-dimensional space, arbitrary rank).

Consider a two-dimensional space N=2N=2. In this case each index set Λn2f,2\Lambda_{n-2f,2} is constituted by the single partition (1n2f)(1^{n-2f}), so Λ¯n,2(f)\overline{\Lambda}^{(f)}_{n,2} is constituted by (nf,f)\(1n2f)(n-f,f)^{\prime}\backslash(1^{n-2f}). The corresponding eigenvalue is

(N+1)f+c(μ\λ)=(nf+1)f,sospec×(An)={(nf+1)f:f=1,,n2}.-(N+1)\,f+c(\mu\backslash\lambda)=-(n-f+1)f\,,\quad\text{so}\quad\mathrm{spec}^{\times}(A_{n})=\{-(n-f+1)f\;:\;f=1,\dots,\lfloor\tfrac{n}{2}\rfloor\}\,. (4.10)

Note that for n=2n=2 one has spec×(A2)={2}\mathrm{spec}^{\times}(A_{2})=\{-2\}, which leads immediately to (4.7) above. Note that spec×(An)\mathrm{spec}^{\times}(A_{n})\subset-\mathbb{N}, which is in agreement with Proposition 3.4. Let us consider n=3n=3 in detail, such that 𝔯\mathfrak{r} is not injective. Due to the restrictions on the number of columns, spec×(A3)={3}\mathrm{spec}^{\times}(A_{3})=\{-3\}, and the only factor constituting the projector (4.9) survives:

P3=(1+13An)=[Uncaptioned image]+13([Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image])P_{3}=\left(1+\frac{1}{3}\,A_{n}\right)=\scalebox{0.8}{$\raisebox{-0.4pt}{\includegraphics[width=15.0pt,height=20.0pt]{Id3.pdf}}$}+\frac{1}{3}\,\Bigl{(}\ \scalebox{0.8}{$\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_1.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_2.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_3.pdf}}\ $}\Bigr{)} (4.11)

By a direct calculation one observes that vanishing of traces of the 𝔯(P3)\mathfrak{r}(P_{3})-image of a tensor is not due to straightforward cancellation of all terms, but rather due to the fact that anti-symmetrization of tree two-dimensional vectors vanishes identically, so dijP3Ker(𝔯)d_{ij}P_{3}\in\mathrm{Ker}(\mathfrak{r}) for all 1i<j31\leqslant i<j\leqslant 3.

Remark.

At this stage it becomes clear that the main computational difficulty resides in expanding the factorised formula (4.1) diagram-wise. A technique which allows one to partially circumvent the latter difficulty is presented in Section 5.

4.2 Traceless projection of simple GL(N)GL(N)-modules.

The projector (4.1) is constructed in a way to take into account all simple G(N)G(N)-modules occurring in VnV^{\otimes n}. However in applications to tensor calculus in physics one often starts with a certain GL(N)GL(N)-module V(μ)VnV^{(\mu)}\subset V^{\otimes n} and performs its traceless projection. To mention a number of examples, in application to higher-spin theory and strings, symmetric tensor fields serve as a primer [33, 34] (see [35] for application of the Brauer algebra in the context of Fronsdal fields). One can also find extensive studies of fields of arbitrary symmetry types (the mixed-symmetry fields), see [36, 37, 38, 39] and references therein.

Restriction of 𝔓n\mathfrak{P}_{n}.

In this section we will adapt the construction of the universal traceless projector (4.1) for obtaining traceless projection of each simple GL(N)GL(N)-module. In this respect one ignores the metric ,\langle\,\cdot,\cdot\,\rangle for a moment and takes the first decomposition in (2.12) as a starting point. A particular choice of G(N)G(N) is nevertheless anticipated by fixing one of the two ways for permutations to act on VnV^{\otimes n} regarding the choice of ε=±1\varepsilon=\pm 1 (recall the definition of 𝔯(si)\mathfrak{r}(s_{i}) in (3.14)). A simple GL(N)GL(N)-module V(μ)V^{(\mu)} (with some μΣ^n,N\mu\in\hat{\Sigma}_{n,N}) can be realised by applying a primitive idempotent I(μ^)𝔖nI^{(\hat{\mu})}\in\mathbb{C}\mathfrak{S}_{n} to VnV^{\otimes n}, such that

𝔯(I(μ^))VnV(μ),and𝔖nI(μ^)L(μ^)(with respect to left multiplication in 𝔖n).\mathfrak{r}\big{(}I^{(\hat{\mu})}\big{)}V^{\otimes n}\cong V^{(\mu)}\,,\quad\text{and}\quad\mathbb{C}\mathfrak{S}_{n}I^{(\hat{\mu})}\cong L^{(\hat{\mu})}\quad\text{(with respect to left multiplication in $\mathbb{C}\mathfrak{S}_{n}$)}\,. (4.12)

As a particular well-known example of a primitive idempotent, one can take any Young symmetriser Y𝗍(μ^)𝔖nY_{\mathsf{t}(\hat{\mu})}\in\mathbb{C}\mathfrak{S}_{n} (𝗍(μ^)\mathsf{t}(\hat{\mu}) being a standard Young tableau of shape μ^\hat{\mu}). Let us also mention that aside from the set of Young symmetrisers, which are primitive idempotents but not orthogonal ones, the complete set of primitive orthogonal idempotents is available for 𝔖n\mathbb{C}\mathfrak{S}_{n} (see [40] and references therein, see also [41]).

Returning the metric into consideration, we will aim at explicit realisation of the following idea. Traceless projection of V(μ)V^{(\mu)} means restriction to G(N)G(N) and projection to the submodule D(μ)V(μ)D^{(\mu)}\subset V^{(\mu)} (see, e.g., [4, Section 3.2]). With the second decomposition in (2.12) at hand and applying Lemma 3.1, traceless projection of V(μ)V^{(\mu)} implies annihilating all simple Bn(εN)B_{n}(\varepsilon N)-modules which are labelled by Λn2f,N\Lambda_{n-2f,N} with f1f\geqslant 1 and which contain a module L(μ^)L^{(\hat{\mu})}.

Define the set of diagrams which parametrises all simple Bn(εN)B_{n}(\varepsilon N)-modules where a given 𝔖n\mathbb{C}\mathfrak{S}_{n}-module L(ρ)L^{(\rho)} can occur by Lemma 3.2: for any ρΛn,N\rho\in\Lambda_{n,N} set

ΛN(ρ)=evenνρ,|ν|=2f,f0(ρ𝜐ν)Λn2f,N,\Lambda^{(\rho)}_{N}=\bigcup_{\begin{array}[]{c}{\scriptstyle\text{even}\;\;\nu\subset\rho\,,}\\ {\scriptstyle|\nu|=2f\,,\;\;f\geqslant 0}\end{array}}(\rho\slashdiv\nu)\cap\Lambda_{n-2f,N}\,, (4.13)

where one makes use of the reverse of the jeu de taquin described in Section 3.1. Note that ρΛN(ρ)\rho\in\Lambda^{(\rho)}_{N}, and for any other σΛN(ρ)\sigma\in\Lambda^{(\rho)}_{N} holds |σ|<|ρ||\sigma|<|\rho|. Joining the above sets over all possible labels of 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules in VnV^{\otimes n} one reconstructs the labels of all Bn(εN)B_{n}(\varepsilon N)-modules in VnV^{\otimes n}, so

Λn,N=ρΣn,NΛN(ρ)(in general, ΛN(ρ1)ΛN(ρ2)).\Lambda_{n,N}=\bigcup_{\rho\in\Sigma_{n,N}}\Lambda^{(\rho)}_{N}\quad\text{(in general, $\Lambda^{(\rho_{1})}_{N}\cap\Lambda^{(\rho_{2})}_{N}\neq\varnothing$).} (4.14)

Continuing along these lines, we consider the following subsets specρ×(An)spec×(An)\mathrm{spec}^{\times}_{\rho}(A_{n})\subset\mathrm{spec}^{\times}(A_{n})

specρ×(An)={(εN1)f+c(ρ\σ):f=1,,n2,σΛN(ρ),|ρ||σ|=2f}ε.\mathrm{spec}^{\times}_{\rho}(A_{n})=\big{\{}(\varepsilon N-1)f+c(\rho\backslash\sigma)\;:\;f=1,\dots,\lfloor{\tfrac{n}{2}\rfloor}\,,\;\;\sigma\in\Lambda^{(\rho)}_{N}\,,\;\;|\rho|-|\sigma|=2f\big{\}}\cap\varepsilon\mathbb{N}\,. (4.15)

The following proposition is a restricted version of Proposition 3.4 and is proven along the same lines.

Proposition 4.2.

Let μΣn,N\mu\in\Sigma_{n,N}, so V(μ^)L(μ)VnV^{(\hat{\mu})}\otimes L^{(\mu)}\subset V^{\otimes n}. Any non-zero eigenvalue of AnA_{n} on V(μ^)L(μ)V^{(\hat{\mu})}\otimes L^{(\mu)} is contained in specμ×(An)\mathrm{spec}^{\times}_{\mu}(A_{n}). Conversely, any element of specμ×(An)\mathrm{spec}^{\times}_{\mu}(A_{n}) is an eigenvalue of AnA_{n} on V(μ^)L(μ)V^{(\hat{\mu})}\otimes L^{(\mu)} if one of the following conditions hold:

  • i)

    μΛn,N\mu\in\Lambda_{n,N} (i.e. μ\mu is Littlewood-admissible),

  • ii)

    Bn(εN)B_{n}(\varepsilon N) is semisimple.

Consider the reduced operator

Pn(μ)=αspecμ×(An)(11αAn),and define𝔓n(μ^)=𝔯(Pn(μ)).P_{n}^{(\mu)}=\prod_{\alpha\in\mathrm{spec}^{\times}_{\mu}(A_{n})}\left(1-\frac{1}{\alpha}\,A_{n}\right)\,,\quad\text{and define}\quad\mathfrak{P}^{(\hat{\mu})}_{n}=\mathfrak{r}(P_{n}^{(\mu)})\,. (4.16)

The above formula can not be applied directly in the only case μ=(1n)\mu=(1^{n}) where specμ×(An)=\mathrm{spec}^{\times}_{\mu}(A_{n})=\varnothing. This corresponds to totally anti-symmetric tensors V(1n)VnV^{(1^{n})}\subset V^{\otimes n} when G(N)=O(N)G(N)=O(N) and totally symmetric tensors V(n)VnV^{(n)}\subset V^{\otimes n} when G(N)=Sp(N)G(N)=Sp(N), which are automatically traceless. In this particular case we set by definition Pn(μ)=1P_{n}^{(\mu)}=1. All in all, the property (𝔓n(μ^))=𝔓n(μ^)\big{(}\mathfrak{P}_{n}^{(\hat{\mu})}\big{)}^{*}=\mathfrak{P}_{n}^{(\hat{\mu})} (for all μΣn,N\mu\in\Sigma_{n,N}) is manifest. The whole paragraph is summarised by the following result, which reflects the commonly utilised fact that the two operations – specific symmetrization of indices of a tensor and subtracting traces – can be performed separately and in any order relatively to one another.

Theorem 4.3.

Let V(μ)VnV^{(\mu)}\subset V^{\otimes n} be a simple GL(N)GL(N)-module (with μΣ^n,N\mu\in\hat{\Sigma}_{n,N}). Upon restriction to V(μ)V^{(\mu)} one has 𝔓n(μ)|V(μ)=𝔓n|V(μ)\mathfrak{P}_{n}^{(\mu)}\big{|}_{V^{(\mu)}}=\mathfrak{P}_{n}\big{|}_{V^{(\mu)}}. In particular, for a primitive idempotent I𝔖nI\in\mathbb{C}\mathfrak{S}_{n} such that 𝔯(I)VnV(μ)\mathfrak{r}(I)V^{\otimes n}\cong V^{(\mu)}, the operator

n(μ)=𝔓n𝔯(I)=𝔓n(μ)𝔯(I)(equivalently,n(μ)=𝔯(I)𝔓n=𝔯(I)𝔓n(μ))\mathfrak{I}_{n}^{\prime(\mu)}=\mathfrak{P}_{n}\mathfrak{r}(I)=\mathfrak{P}_{n}^{(\mu)}\mathfrak{r}(I)\quad\text{(equivalently,}\;\;\mathfrak{I}_{n}^{\prime(\mu)}=\mathfrak{r}(I)\mathfrak{P}_{n}=\mathfrak{r}(I)\mathfrak{P}_{n}^{(\mu)}\text{)} (4.17)

projects VnV^{\otimes n} onto a simple module isomorphic to D(μ)D^{(\mu)} when μΛ^n,N\mu\in\hat{\Lambda}_{n,N} (i.e. μ^\hat{\mu} is Littlewood-admissible), or annihilates it otherwise. In addition, if I=II^{*}=I, then (n(μ))=n(μ)\big{(}\mathfrak{I}_{n}^{\prime(\mu)}\big{)}^{*}=\mathfrak{I}_{n}^{\prime(\mu)}.

Remark.

Note that Young symmetrisers are not self-adjoint with respect to ()(\,\cdot\,)^{*}. To have the latter property at hand one should take the orthogonal primitive idempotents in 𝔖n\mathbb{C}\mathfrak{S}_{n} [18, 19, 40] (called also Young seminormal units [41] or Hermitian Young operators [42]).

The construction in question admits the following straightforward generalisation to the case of a direct sum of simple GL(N)GL(N)-modules (possibly with certain multiplicities) in VnV^{\otimes n}. For Σn,N\mathcal{I}\subset\Sigma_{n,N} define spec×(An)=μspecμ×(An)\mathrm{spec}^{\times}_{\mathcal{I}}(A_{n})=\bigcup_{\mu\in\mathcal{I}}\mathrm{spec}^{\times}_{\mu}(A_{n}), construct

Pn()=αspec×(An)(11αAn)and set𝔓n(^)=𝔯(Pn())\quad P^{(\mathcal{I})}_{n}=\prod_{\alpha\in\mathrm{spec}^{\times}_{\mathcal{I}}(A_{n})}\left(1-\frac{1}{\alpha}\,A_{n}\right)\quad\text{and set}\quad\mathfrak{P}^{(\hat{\mathcal{I}})}_{n}=\mathfrak{r}\big{(}P^{(\mathcal{I})}_{n}\big{)} (4.18)

(where ^\hat{\mathcal{I}} denotes the mapping μμ^\mu\mapsto\hat{\mu} applied element-wise). The following corollary is a simple consequence of Theorem 4.3

Corollary 4.4.

Consider Σ^n,N\mathcal{I}\subset\hat{\Sigma}_{n,N} and a GL(N)GL(N)-module

V()=μ(V(μ))cμVn(cμ1).V^{(\mathcal{I})}=\bigoplus_{\mu\in\mathcal{I}}\big{(}V^{(\mu)}\big{)}^{\oplus c_{\mu}}\subset V^{\otimes n}\quad(c_{\mu}\geqslant 1)\,. (4.19)

Then upon restriction to V()V^{(\mathcal{I})} one has 𝔓n()|V()=𝔓n|V()\mathfrak{P}_{n}^{(\mathcal{I})}\big{|}_{V^{(\mathcal{I})}}=\mathfrak{P}_{n}\big{|}_{V^{(\mathcal{I})}}.

In particular, as soon as a tensor product of GL(N)GL(N)-modules V(μ1),V(μ2)V^{(\mu_{1})},V^{(\mu_{2})} decomposes into a direct sum with the aid of the Littlewood-Richardson rule applied to μ1μ2\mu_{1}\,{\scriptstyle\otimes}\,\mu_{2}, namely

V(μ1)V(μ2)σ|μ1|+|μ2|cμ1μ2σ0,σ1N(V(σ))cμ1μ2σV(|μ1|+|μ2|),V^{(\mu_{1})}\otimes V^{(\mu_{2})}\cong\bigoplus_{\begin{array}[]{c}{\scriptstyle\sigma\,\vdash\,|\mu_{1}|+|\mu_{2}|}\\ {\scriptstyle c^{\sigma}_{\mu_{1}\mu_{2}}\neq 0\,,\;\;\sigma^{\prime}_{1}\leqslant N}\end{array}}\big{(}V^{(\sigma)}\big{)}^{\oplus c^{\sigma}_{\mu_{1}\mu_{2}}}\;\;\subset\;\;V^{\otimes(|\mu_{1}|+|\mu_{2}|)}\,, (4.20)

the reduced traceless operator is constructed as 𝔓|μ1|+|μ2|()\mathfrak{P}_{|\mu_{1}|+|\mu_{2}|}^{(\mathcal{I})}, with the index set \mathcal{I} constituted by labels σ\sigma of the modules occurring on the right-hand-side of (4.20) (see Example 8 in Appendix A).

Example 5 (totally symmetric O(N)O(N)-tensors).

For the fixed partition μ=(n)\mu=(n) one constructs spec(n)×(An)\mathrm{spec}^{\times}_{(n)}(A_{n}) for the skew-shape diagrams μ\λ\mu\backslash\lambda for all λ(n)𝜐(2f)\lambda\in(n)\slashdiv(2f), f=1,,n2f=1,\dots,\lfloor{\frac{n}{2}\rfloor}. This leads to

\ytableausetupmathmode,boxsize=1.3em,centertableauxμ\λ={ytableau}×&\none[]×f\none[]n1spec(n)×(An)={(N+2(nf1))f:f=1,,n2}.\ytableausetup{mathmode,boxsize=1.3em,centertableaux}\mu\backslash\lambda=\ytableau{\scriptstyle}\times&\none[\scriptstyle{\cdots}]\times{\scriptstyle f}\none[\scriptstyle{\cdots}]{\scriptstyle n-1}\quad\Rightarrow\quad\mathrm{spec}_{(n)}^{\times}(A_{n})=\left\{\big{(}N+2\ (n-f-1)\big{)}\,f\;:\;f=1,\dots,\lfloor\tfrac{n}{2}\rfloor\right\}\,.

Hence the reduced projector (4.16) takes the form

Pn(n)=f=1n2(1An(N+2(nf1))f).P_{n}^{(n)}=\displaystyle{\prod_{f=1}^{\lfloor\tfrac{n}{2}\rfloor}}\left(1-\frac{A_{n}}{\big{(}N+2\ (n-f-1)\big{)}f}\right)\,. (4.21)

In the next section we will rewrite the above expression in terms of the Lie algebra 𝔰𝔩(2)\mathfrak{sl}(2) which arises in the context of Howe duality for symmetric tensors. This will allow us to rewrite the expression in the expanded form.

Example 6 (maximally-antisymmetric hook O(N)O(N)-tensors).

For the partition μ=(2,1n2)\mu=(2,1^{n-2}) one constructs spec(2,1n2)×(An)\mathrm{spec}^{\times}_{(2,1^{n-2})}(A_{n}) for the skew shape diagrams μ\λ\mu\backslash\lambda with λ(2,1n2)𝜐(2)\lambda\in(2,1^{n-2})\slashdiv(2), which leads to the only possibility:

\ytableausetupmathmode,boxsize=1.3em,centertableauxμ\λ={ytableau}×&1×\none[ . . . ]×2nspec(2,1n2)×(An)={Nn+2},soPn(2,1n2)=1AnNn+2.\ytableausetup{mathmode,boxsize=1.3em,centertableaux}\mu\backslash\lambda=\ytableau{\scriptstyle}\times&{\scriptstyle 1}\\ \times\\ \none[\vbox{ \scriptsize\hbox{.}\hbox{.}\hbox{.}\kern-0.75pt}]\\ \times\\ {\scriptstyle 2-n}\\ \quad\Rightarrow\quad\mathrm{spec}_{(2,1^{n-2})}^{\times}(A_{n})=\big{\{}N-n+2\,\big{\}}\,,\;\;\text{so}\quad P_{n}^{(2,1^{n-2})}=1-\dfrac{A_{n}}{N-n+2}\;. (4.22)

Note that in order for the GL(N)GL(N)-module in question to be present in VnV^{\otimes n}, one assumes Nn1N\geqslant n-1, so the denominator is non-singular. Moreover, in accordance with Proposition 3.4 the expression in the denominator is always positive. The case of a generic hook (m,1k)(m,1^{k}) is considered in Appendix A.

Quasi-additive form of the universal projector 𝔓n\mathfrak{P}_{n}.

From the point of view of applications, the most convenient form of the universal traceless projector would be a sum of averages (i.e. elements of Cn(εN)C_{n}(\varepsilon N), recall the comment at the end of Section 3.2). To this end, along with theoretical transparency of the factorised form (4.1) and its convenience for restrictions to GL(N)GL(N)-modules, the other side of the coin is that regarding applications the expression (4.1) is quite far from being optimal – first of all due to the necessity to express the powers (An)p(A_{n})^{p} as combinations of averages. The latter problem is partially resolved with the aid of reduced traceless projectors described in Theorem 4.3. Our goal consists in summing up the latter and reconstruct the universal traceless projector as a polynomial in AnA_{n} of a smaller degree than that of (4.1).

First, we note that the traceless subspace of V(μ)VnV^{(\mu)}\subset V^{\otimes n} is non-zero only if μΛ^n,N\mu\in\hat{\Lambda}_{n,N} (μ^\hat{\mu} is Littlewood-admissible). By restricting our attention to the latter set of GL(N)GL(N)-modules, we construct the following element in Bn(εN)B_{n}(\varepsilon N). Let z(μ)z^{(\mu)} denote the central Young symmetriser associated to a simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-module indexed by μn\mu\vdash n [43], then set

P~n=μΛn,NPn(μ)z(μ).\tilde{P}_{n}=\sum_{\mu\in\Lambda_{n,N}}P^{(\mu)}_{n}z^{(\mu)}\,. (4.23)

Note that due to Proposition 4.2, each Pn(μ)P^{(\mu)}_{n} in the above formula is constructed with the minimal possible number of factors since each element in specμ×(An)\mathrm{spec}^{\times}_{\mu}(A_{n}) is an eigenvalue of AnA_{n}. Also, one has the property (P~n)=P~n(\tilde{P}_{n})^{*}=\tilde{P}_{n}. Indeed, (z(μ))=z(μ)(z^{(\mu)})^{*}=z^{(\mu)} (see (5.27) and the comment below), and each Pn(μ)Cn(εN)P^{(\mu)}_{n}\in C_{n}(\varepsilon N). The 𝔯\mathfrak{r}-image of (4.23) gives the sought resummation of the reduced projectors of Theorem 4.3.

Corollary 4.5.

The universal traceless projector admits the following form:

𝔓n=𝔯(P~n).\mathfrak{P}_{n}=\mathfrak{r}\big{(}\tilde{P}_{n}\big{)}\,. (4.24)
Proof.

Central Young symmetrisers form a decomposition of unity:

1=μnz(μ)Pn=μnPnz(μ).1=\sum_{\mu\vdash n}z^{(\mu)}\quad\Rightarrow\quad P_{n}=\sum_{\mu\vdash n}P_{n}z^{(\mu)}\,. (4.25)

Each z(μ)z^{(\mu)} is the sum of orthogonal idempotents whose left ideal in 𝔖n\mathbb{C}\mathfrak{S}_{n} is isomorphic to L(μ)L^{(\mu)}. So, by Theorem 4.3, 𝔓n𝔯(z(μ))=𝔓n(μ^)𝔯(z(μ))\mathfrak{P}_{n}\mathfrak{r}\big{(}z^{(\mu)}\big{)}=\mathfrak{P}^{(\hat{\mu})}_{n}\mathfrak{r}\big{(}z^{(\mu)}\big{)}. For V(μ^)V^{(\hat{\mu})} such that μΛn,N\mu\notin\Lambda_{n,N}, D(μ^)D^{(\hat{\mu})} is not present in the decomposition of the latter upon restriction to G(N)G(N), and hence the corresponding traceless projection vanishes identically, so one has 𝔓n(μ^)𝔯(z(μ))=0\mathfrak{P}^{(\hat{\mu})}_{n}\mathfrak{r}\big{(}z^{(\mu)}\big{)}=0 in this case. Therefore, from (4.25) on obtains

𝔓n=𝔯(Pn)=μΛn,N𝔓n(μ^)𝔯(z(μ))=𝔯(P~n).\mathfrak{P}_{n}=\mathfrak{r}(P_{n})=\sum_{\mu\in\Lambda_{n,N}}\mathfrak{P}^{(\hat{\mu})}_{n}\,\mathfrak{r}\big{(}z^{(\mu)}\big{)}=\mathfrak{r}\big{(}\tilde{P}_{n}\big{)}\,. (4.26)

Remark.

The examples of traceless projectors given in the companion Mathematica notebook utilise the expression (4.23), which is the one implemented in [6].

4.3 Tracelessness in the context of Howe duality

In the context of the group G(N)G(N) acting on tensors, aside from the Schur-Weyl-type duality which concerns representation theory of Brauer algebras, there is another well-known duality which is important through its application in field theories – namely, the Howe duality [8] (which is often referred to as “oscillator realisation” in the physics literature, see, e.g., [44, 45, 46] for applications). The latter relates representations of the classical group GG to those of an algebra 𝔄\mathfrak{A} via a bimodule where the actions of the two mutually centralise each other. In particular, a decomposition of a (G,𝔄)(G,\mathfrak{A})-bimodule reminiscent to (2.12) takes place, with finite-dimensional simple GG-modules in the left slot and a simple 𝔄\mathfrak{A}-modules in the right slot. The main difference with the Schur-Weyl-type dualities consists in considering the infinite-dimensional subspace in the tensor algebra T(V)T(V) where the actions of GG and 𝔄\mathfrak{A} meet (instead of the finite-dimensional component VnV^{\otimes n}). The algebra 𝔄\mathfrak{A} is generated by transformations which form a Lie algebra 𝔡(r)\mathfrak{d}(r) for some r1r\geqslant 1: 𝔡(r)=𝔰𝔭(2r)\mathfrak{d}(r)=\mathfrak{sp}(2r) for G(N)=O(N)G(N)=O(N) and 𝔡(r)=𝔬(2r)\mathfrak{d}(r)=\mathfrak{o}(2r) for G(N)=Sp(N)G(N)=Sp(N).

Totally symmetric traceless G(N)G(N)-tensors.

We start by revisiting Example 5 (with G(N)=O(N)G(N)=O(N)) which is a good starting point to introduce the main ideas. The space of totally symmetric tensors (of arbitrary rank) is isomorphic to the space [y]\mathbb{C}[y] of polynomials in NN variables y={ya:a=1,,N}y=\{y_{a}\;:\;a=1,\dots,N\}. Rank-nn symmetric tensors are isomorphic to the subspace of degree-nn homogeneous polynomials which we denote [y]n\mathbb{C}[y]_{n}:

T=ea1eanta1ansym(Vn)T(y)=ta1anya1yan[y]nT=e_{a_{1}}\otimes\dots\otimes e_{a_{n}}t^{a_{1}\dots a_{n}}\in\mathrm{sym}(V^{\otimes n})\quad\mapsto\quad T(y)=t^{a_{1}\dots a_{n}}\,y_{a_{1}}\dots y_{a_{n}}\in\mathbb{C}[y]_{n}

(abusing notation, we denote the tensor and the corresponding polynomial by the same letter). The trace is the same for any pair of indices, so one has tr(g)T(y)=gbctbca3anya3yan\mathrm{tr}^{(g)}T(y)=g_{bc}t^{bc\,a_{3}\dots a_{n}}\,y_{a_{3}}\dots y_{a_{n}}, which is conveniently expressed via the following second-order differential operator:

for𝖾+=12gabyaybone has𝖾+T(y)=n(n1)2tr(g)T(y).\text{for}\quad\mathsf{e}_{+}=\frac{1}{2}\,g_{ab}\frac{\partial}{\partial y_{a}}\frac{\partial}{\partial y_{b}}\quad\text{one has}\quad\mathsf{e}_{+}\,T(y)=\tfrac{n(n-1)}{2}\,\mathrm{tr}^{(g)}T(y)\,.

If one additionally considers the quadratic operator 𝖾=12gabyayb\mathsf{e}_{-}=-\tfrac{1}{2}\,g^{ab}\,y_{a}y_{b}, then for the action of 𝖾𝖾+\mathsf{e}_{-}\mathsf{e}_{+} on the polynomials one recognizes the action of AnA_{n} on tensors:

AnT2𝖾𝖾+T(y).A_{n}T\quad\mapsto\quad-2\,\mathsf{e}_{-}\mathsf{e}_{+}\,T(y)\,. (4.27)

The commutator of the two operators 𝖾,𝖾+\mathsf{e}_{-},\mathsf{e}_{+} gives 𝗁=(N2+yaya)\mathsf{h}=-\big{(}\tfrac{N}{2}+y_{a}\frac{\partial}{\partial y_{a}}\big{)}, and all together they form the Lie algebra 𝔰𝔩(2)\mathfrak{sl}(2):

[𝖾+,𝖾]=𝗁,[𝗁,𝖾±]=±2𝖾±.\left[\mathsf{e}_{+},\mathsf{e}_{-}\right]=\mathsf{h}\,,\quad\left[\mathsf{h},\mathsf{e}_{\pm}\right]=\pm 2\,\mathsf{e}_{\pm}\,. (4.28)

As far as symmetric tensors form the irreducible GL(N)GL(N)-module [y]nV(n)\mathbb{C}[y]_{n}\cong V^{(n)}, we make use of the reduced projector Pn(n)P_{n}^{(n)} (4.21). First, note that 𝗁T(y)=(N2+n)T(y)\mathsf{h}\,T(y)=-(\tfrac{N}{2}+n)\,T(y) on tensors from [y]n\mathbb{C}[y]_{n}. Next, substitution (4.27) leads to the following operator acting on polynomials:

Pn(n)P𝔰𝔩(2)=f1(1𝖾𝖾+(𝗁+f+1)f),P_{n}^{(n)}\quad\mapsto\quad P_{\mathfrak{sl}(2)}=\prod_{f\geqslant 1}\left(1-\frac{\mathsf{e}_{-}\mathsf{e}_{+}}{(\mathsf{h}+f+1)f}\right)\,, (4.29)

which coincides with the form of extremal projector for 𝔰𝔩(2)\mathfrak{sl}(2) presented in §7 of Chapter 3 in [11] (see also [12, 47] for a review). The infinite product acts on [y]\mathbb{C}[y] by consecutive application of factors and truncates at fmax=n2f_{\mathrm{max}}=\lfloor\tfrac{n}{2}\rfloor when restricted to [y]n\mathbb{C}[y]_{n}. As a result, traceless rank-nn tensors can be viewed as the subspace of highest-weight vectors of weight (N2+n)-\big{(}\frac{N}{2}+n\big{)} in the 𝔰𝔩(2)\mathfrak{sl}(2)-module [y]\mathbb{C}[y].

The extremal projector P𝔰𝔩(2)P_{\mathfrak{sl}(2)} admits the following (equivalent) additive form

P𝔰𝔩(2)=P+(c)=f0(1)ff!1j=1f(𝗁+c+j)𝖾f𝖾+f,withc=1.P_{\mathfrak{sl}(2)}=P_{+}(c)=\sum_{f\geqslant 0}\frac{(-1)^{f}}{f!}\dfrac{1}{\displaystyle{\prod_{j=1}^{f}(\mathsf{h}+c+j)}}\,\mathsf{e}_{-}^{f}\mathsf{e}_{+}^{f}\,,\quad\text{with}\;\;c=1\,. (4.30)

The products 𝖾f𝖾+f\mathsf{e}_{-}^{f}\mathsf{e}_{+}^{f} can be directly mapped to elements An(f)A_{n}^{(f)} (f=1,,n2f=1,\dots,\lfloor{\frac{n}{2}\rfloor}) which generalise An=An(1)A_{n}=A_{n}^{(1)} to the case of ff arcs:

An(f)=12ff!(n2f)!γd1d3d2f1Cn(δ)J(f)An(f)12f2f!𝖾f𝖾+f.A_{n}^{(f)}=\frac{1}{2^{f}f!(n-2f)!}\,\gamma_{d_{1}d_{3}\dots d_{2f-1}}\in C_{n}(\delta)\cap J^{(f)}\quad\Rightarrow\quad A_{n}^{(f)}\mapsto-\frac{1}{2^{f-2}f!}\,\mathsf{e}_{-}^{f}\mathsf{e}_{+}^{f}\,. (4.31)

Translated to Bn(N)B_{n}(N), the expression (4.30) gives the expanded form of the traceless projector Pn(n)P_{n}^{(n)} bypassing direct computations of the powers (An)p(A_{n})^{p} and restricting them to V(n)V^{(n)}. Rederivation of (4.30) in other frameworks can be found in the literature [48, 49].

Note that the same construction applies in the case when the metric is skew-symmetric, with G(N)=Sp(N)G(N)=Sp(N). Symmetric tensors are automatically traceless in this case which is reflected in trivialisation of the trace operator 𝖾+=0\mathsf{e}_{+}=0, as well as 𝖾=0\mathsf{e}_{-}=0. The only non-trivial operator 𝗁\mathsf{h} constitutes the Lie algebra 𝔬(2)\mathfrak{o}(2).

Mixed-symmetry G(N)G(N)-tensors.

In relation to the above example, let us mention the well-known way of realizing tensorial mixed-symmetry G(N)G(N)-modules D(ρ)VnD^{(\rho)}\subset V^{\otimes n} via homogeneous polynomials and differential operators acting on them (as before, we start with the case G(N)=O(N)G(N)=O(N)). Consider the space of polynomials [𝒚]\mathbb{C}[\boldsymbol{y}] in the variables 𝒚={yai:a=1,,N,i=1,,r}\boldsymbol{y}=\{y_{a}^{i}\;:\;a=1,\dots,N\,,\;i=1,\dots,r\}. For the rank-nn tensors of the symmetry type ρ=(ρ1,,ρr)\rho=(\rho_{1},\dots,\rho_{r}) (with |ρ|=n|\rho|=n) one considers the subspace [𝒚]ρ\mathbb{C}[\boldsymbol{y}]_{\rho} of degree-nn homogeneous polynomials which are also degree-ρi\rho_{i} homogeneous in each subset of variables {yai:a=1,,N}\{y_{a}^{i}\;:\;a=1,\dots,N\} (for each fixed ii). In other words, the elements of [𝒚]ρ\mathbb{C}[\boldsymbol{y}]_{\rho} are rank-nn tensors with rr enumerated groups of symmetrised indices, each iith group carrying ρi\rho_{i} indices.

From the polynomial variables and their derivatives one constructs the following set of differential operators which constitute the Lie algebra 𝔰𝔭(2r)\mathfrak{sp}(2r):

𝖪ij=11+δijgabyaiybj,𝖯ij=11+δijgabyaiybj,𝖧i=j(N2δji+yaiyaj).\mathsf{K}^{ij}=-\frac{1}{1+\delta_{ij}}\,g^{ab}\,y_{a}^{i}y_{b}^{j}\,,\;\;\mathsf{P}_{ij}=\frac{1}{1+\delta_{ij}}\,g_{ab}\,\frac{\partial}{\partial y_{a}^{i}}\frac{\partial}{\partial y_{b}^{j}}\,,\;\;\mathsf{H}^{i}{}_{j}=-\big{(}\dfrac{N}{2}\delta^{i}_{j}+y_{a}^{i}\frac{\partial}{\partial y_{a}^{j}}\big{)}\,. (4.32)

The operators 𝖧ij\mathsf{H}^{i}{}_{j} form the subalgebra 𝔤𝔩(r)𝔢𝔰𝔩(r)𝔰𝔭(2r)\mathfrak{gl}(r)\cong\mathfrak{e}\oplus\mathfrak{sl}(r)\subset\mathfrak{sp}(2r), where the center 𝔢\mathfrak{e} is spanned by the multiples of the Euler operator i=1r𝖧ii\sum_{i=1}^{r}\mathsf{H}^{i}{}_{i}. When r=1r=1 one recognises the above example of symmetric tensors for 𝔰𝔩(2)𝔰𝔭(2)\mathfrak{sl}(2)\cong\mathfrak{sp}(2). The algebra (4.32) centralises the action of O(N)O(N) on [𝒚]\mathbb{C}[\boldsymbol{y}] and is well-known in the context of Howe duality [8]. The polynomials T(𝒚)[𝒚]ρT(\boldsymbol{y})\in\mathbb{C}[\boldsymbol{y}]_{\rho} which constitute the irreducible O(N)O(N)-module D(ρ)D^{(\rho)} satisfy the highest-weight conditions for the dual algebra 𝔰𝔭(2r)\mathfrak{sp}(2r):

𝖧iTi(𝒚)\displaystyle\mathsf{H}^{i}{}_{i}\,T(\boldsymbol{y}) =(N2+ρi)T(𝒚)(no summation over i),\displaystyle=-\big{(}\frac{N}{2}+\rho_{i}\big{)}\,T(\boldsymbol{y})\;\;\text{(no summation over $i$)}\,, (4.33)
𝖧iTj(𝒚)\displaystyle\mathsf{H}^{i}{}_{j}\,T(\boldsymbol{y}) =0(for all i<j),\displaystyle=0\;\;\text{(for all $i<j$)}\,, (4.34)
𝖯ijT(𝒚)\displaystyle\mathsf{P}_{ij}\,T(\boldsymbol{y}) =0(for all ij).\displaystyle=0\;\;\text{(for all $i\leqslant j$)}\,. (4.35)

The algebra of the above operators is a Borel subalgebra 𝔥𝔤+𝔰𝔭(2r)\mathfrak{h}\oplus\mathfrak{g}_{+}\subset\mathfrak{sp}(2r): homogeneity degrees ρi\rho_{i} enter the eigenvalues of the Cartan operators 𝖧ii𝔥\mathsf{H}^{i}{}_{i}\in\mathfrak{h}, while the rest of the constraints represent annihilation of the highest-weight vectors by the positive root operators from 𝔤+\mathfrak{g}_{+}. In particular, 𝔤+=𝔤ˇ+𝔱\mathfrak{g}_{+}=\check{\mathfrak{g}}_{+}\oplus\mathfrak{t}: the subalgebra 𝔤ˇ+\check{\mathfrak{g}}_{+} is constituted by the operators entering (4.34) (which are positive root operators in 𝔰𝔩(r)𝔤𝔩(r)\mathfrak{sl}(r)\subset\mathfrak{gl}(r)), while 𝔱\mathfrak{t} is the abelian ideal in 𝔤+\mathfrak{g}_{+} constituted by the trace operators in (4.35). From the point of view of tensor components, the constraints (4.35) simply mean that the tensor is traceless with respect to any pair of indices. The constraints (4.34) imply that symmetrisation of indices in the iith group with any index in the jjth group is zero whenever i<ji<j. The latter if often referred to as Young property as soon as it manifests itself for the images of the Young projector Y𝗍0(ρ)VnY_{\mathsf{t}_{0}(\rho)}V^{\otimes n}, where each row in the standard Young tableau 𝗍0(ρ)\mathsf{t}_{0}(\rho) forms a sequence of consecutive integers.

The space of highest-weight vectors in any 𝔰𝔭(2r)\mathfrak{sp}(2r)-module (and thus the solution of the constraints (4.33)-(4.35)) can be obtained via application of the corresponding extremal projector, which exists and is unique for any simple Lie algebra [11] (see references therein and [12] for the historical review). The projector (4.29) is the simplest one of a kind. According to the general scheme, extremal projector for 𝔰𝔭(2r)\mathfrak{sp}(2r) is written in a form of an ordered product

P𝔰𝔭(2r)=βΔ(𝔤+)Pβ(cβ).P_{\mathfrak{sp}(2r)}=\prod_{\beta\in\Delta(\mathfrak{g}_{+})}^{\longrightarrow}P_{\beta}(c_{\beta})\,. (4.36)

Each factor Pβ(cβ)P_{\beta}(c_{\beta}) is given by the series (4.30) constructed from the operators {𝖾±β,𝗁β}\{\mathsf{e}_{\pm\beta},\mathsf{h}_{\beta}\} forming a 𝔰𝔩(2)\mathfrak{sl}(2)-triple for each positive root βΔ(𝔤+)\beta\in\Delta(\mathfrak{g}_{+}) (recall (4.28)), with

cβ=12φΔ(𝔤+)φ(𝗁β).c_{\beta}=\frac{1}{2}\sum_{\varphi\in\Delta(\mathfrak{g}_{+})}\varphi(\mathsf{h}_{\beta})\,. (4.37)

To be more specific, 𝖾β\mathsf{e}_{\beta} (respectively, 𝖾β\mathsf{e}_{-\beta}) is either 𝖧ij\mathsf{H}^{i}{}_{j} (respectively, 𝖧ji\mathsf{H}^{j}{}_{i}) with i<ji<j or 𝖯ij\mathsf{P}_{ij} (respectively, 𝖪ij\mathsf{K}^{ij}) with iji\leqslant j, and 𝗁β=[𝖾β,𝖾β]\mathsf{h}_{\beta}=[\mathsf{e}_{\beta},\mathsf{e}_{-\beta}]. The order of factors comes from the normal ordering of the positive roots (see, e.g., §4 of Chapter 1 in [11]), while the whole extremal projector is insensitive to a particular choice of normal ordering.

The case G(N)=Sp(N)G(N)=Sp(N) is considered along the same lines, with the Lie algebra 𝔬(2r)\mathfrak{o}(2r) realised by the same operators (4.32) except 𝖯ii=0\mathsf{P}_{ii}=0 and 𝖪ii=0\mathsf{K}^{ii}=0 due to the skew symmetry of the metric. With this remark at hand, root decomposition stays the same as for the case 𝔡(r)=𝔰𝔭(2r)\mathfrak{d}(r)=\mathfrak{sp}(2r), so simple Sp(N)Sp(N)-modules are singled out by the constraints (4.33)-(4.35). The expression for the extremal projector (4.36) applies also for 𝔡(r)=𝔬(2r)\mathfrak{d}(r)=\mathfrak{o}(2r). As a result, we can formulate the following lemma.

Lemma 4.6.

The extremal projector P𝔡(r)P_{\mathfrak{d}(r)} is divisible on the right by the extremal projector P𝔰𝔩(r)P_{\mathfrak{sl}(r)}:

P𝔡(r)=P𝔱P𝔰𝔩(r),P_{\mathfrak{d}(r)}=P_{\mathfrak{t}}P_{\mathfrak{sl}(r)}\,, (4.38)

which reflects the possibility of consecutive implication of the constraints (4.34) and (4.35).

Proof.

The fact that 𝔤ˇ+𝔤+\check{\mathfrak{g}}_{+}\subset\mathfrak{g}_{+} is the subalgebra and 𝔱𝔤+\mathfrak{t}\subset\mathfrak{g}_{+} is the ideal allows one to have all the factors with the trace operators on the left. By direct computation777Recall that for any 𝗁𝔥\mathsf{h}\in\mathfrak{h} the value of any root β(𝗁)\beta(\mathsf{h}) is obtained via the commutator [h,𝖾β]=β(𝗁)𝖾β[h,\mathsf{e}_{\beta}]=\beta(\mathsf{h})\,\mathsf{e}_{\beta}. one finds that for any βΔ(𝔤ˇ+)\beta\in\Delta(\check{\mathfrak{g}}_{+}) there is φΔ(𝔱)φ(𝗁β)=0\sum_{\varphi\in\Delta(\mathfrak{t})}\varphi(\mathsf{h}_{\beta})=0. As a result, for the roots in question, cβc_{\beta} in (4.37) are replaced by cβ=12φΔ(𝔤ˇ+)φ(𝗁β)c^{\prime}_{\beta}=\frac{1}{2}\sum_{\varphi\in\Delta(\check{\mathfrak{g}}_{+})}\varphi(\mathsf{h}_{\beta}), so the corresponding factors contain only the 𝔰𝔩(r)\mathfrak{sl}(r) data. Combining this fact with the comment below (4.36) about the ordering of factors proves the assertion. ∎

The operator P𝔰𝔩(r)P_{\mathfrak{sl}(r)}, applied to the space [𝒚]\mathbb{C}[\boldsymbol{y}], resolves the constraints (4.34), while the trace constraints (4.35) are resolved by P𝔱P_{\mathfrak{t}}. The factorised form of the projector (4.38) is reminiscent to the form of the projectors presented in [4], where projection to a simple G(N)G(N)-module D(ρ)VnD^{(\rho)}\subset V^{\otimes n} is performed in two steps: i) projection onto a simple GL(N)GL(N)-module V(ρ)V^{(\rho)}, and ii) subtraction of traces. Note that P𝔱P_{\mathfrak{t}} itself is not a traceless projector on [𝒚]\mathbb{C}[\boldsymbol{y}]. An interesting related problem would be to look for an analog of the universal traceless projector presented in Theorem 4.1: a projector constructed from the operators 𝖪ij\mathsf{K}^{ij}, 𝖯ij\mathsf{P}_{ij} and 𝖧ii\mathsf{H}^{i}{}_{i} which maps the whole space [𝒚]\mathbb{C}[\boldsymbol{y}] (and any irreducible GL(N)GL(N)-module V(ρ)[𝒚]V^{(\rho)}\subset\mathbb{C}[\boldsymbol{y}] in particular) onto its traceless subspace and commutes with any projector to a simple GL(N)GL(N)-module (so, with the extremal projector P𝔰𝔩(r)P_{\mathfrak{sl}(r)} in particular).

As a concluding remark, for a fixed ρn\rho\vdash n one can associate particular elements of Bn(εN)B_{n}(\varepsilon N) to the operators888The action of (𝖪ij)f(𝖯ij)f(\mathsf{K}^{ij})^{f}(\mathsf{P}_{ij})^{f}, when translated to Bn(εN)B_{n}(\varepsilon N), does not correspond to an element from Cn(εN)C_{n}(\varepsilon N). So in the framework of Howe-duality, with certain tensor components being a priori symmetrised in [𝒚]ρ\mathbb{C}[\boldsymbol{y}]_{\rho} (with ρn\rho\vdash n), controlling commutation with the whole group of permutations and, as a consequence, with projectors on simple GL(N)GL(N)-modules, is not manifest. (𝖪ij)f(𝖯ij)f(\mathsf{K}^{ij})^{f}(\mathsf{P}_{ij})^{f} and rewrite the extremal projector P𝔡(r)P_{\mathfrak{d}(r)} as an element in Bn(εN)B_{n}(\varepsilon N). Then the 𝔯\mathfrak{r}-image of the latter will reproduce the projector (ρ)\mathfrak{I}^{\prime(\rho)} of Theorem 4.3, where for the idempotent in 𝔖n\mathbb{C}\mathfrak{S}_{n} one takes the Young projector Y𝗍0(ρ^)Y_{\mathsf{t}_{0}(\hat{\rho})} (recall Young property, see the paragraph below (4.34)). While the two ways of constructing the same projector are two sides of the same coin, working with the extremal projector P𝔡(r)P_{\mathfrak{d}(r)} appears to be hard at the level of computations. The advantage of our approach is due to a convenient “condensed” way of expanding the factorised form of the traceless projector (4.1) in terms of the elements in Cn(εN)C_{n}(\varepsilon N), avoiding diagram-wise computations. This technique is presented in the forthcoming section.

5 AnA_{n} as a second-order differential operator on [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})]

The factorised formula for the traceless projector (4.1) is extremely useful for presentation and elucidating its properties. Nevertheless, expanding and calculating the powers of AnA_{n} (3.21) becomes a hard computational problem already for relatively small ranks (e.g., n=5,6,n=5,6,\dots) when performed at the level of diagrams constituting the conjugacy classes. We propose a technique which allows one to circumvent this problem by performing the calculations at the level of conjugacy classes, without decomposing them into single diagrams. The results of this section hold for any value δ\delta\in\mathbb{C} of the parameter of Bn(δ)B_{n}(\delta).

5.1 Parametrisation of bases in Cn(δ)C_{n}(\delta)

To avoid diagram-wise computations we make use of the fact that AnA_{n} is an element of the algebra Cn(δ)C_{n}(\delta), which implies that any power (An)p(A_{n})^{p} can be decomposed over a basis in Cn(δ)C_{n}(\delta). As it was mentioned in Section 3.2, any maximal independent set among the averages γb\gamma_{b} (for all bBn(δ)b\in B_{n}(\delta)) forms the basis in Cn(δ)C_{n}(\delta). In order to parametrise it we consider an equivalent reformulation of the one described in [50] (see also [13]). Namely, the bases in Cn(δ)C_{n}(\delta) are in one-to-one correspondence with a particular subset of so-called ternary bracelets. A ternary bracelet is an equivalence class of non-empty words over the ternary alphabet 𝒜={𝐧,𝐬,𝐩}\mathcal{A}=\{\mathbf{n},\mathbf{s},\mathbf{p}\} related by cyclic permutations and inversions, i.e. can be viewed as a word with its letters written along a closed loop without specifying the direction of reading. In the sequel, we will write [w][w] to denote a bracelet containing a representative ww (a word over 𝒜\mathcal{A}). For the reverse of ww we will write I(w)I(w), so according to the definition of bracelets one has [w]=[I(w)][w]=[I(w)]. The length of a bracelet is defined as the length of any among its representatives, which is written as |w||w|.

Denote 𝔟(𝒜)\mathfrak{b}(\mathcal{A}) the set of non-empty ternary bracelets with the same number of occurrences of the letters 𝐧\mathbf{n} and 𝐬\mathbf{s} (which is allowed to be 0), with the additional requirement that for any representative, if there is a pair of letters 𝐧\mathbf{n} (respectively, 𝐬\mathbf{s}), there is necessarily the letter 𝐬\mathbf{s} (respectively, 𝐧\mathbf{n}) in between. For example, [𝐬],[𝐧𝐧𝐬𝐬]𝔟(𝒜)[\mathbf{s}],[\mathbf{n}\mathbf{n}\mathbf{s}\mathbf{s}]\notin\mathfrak{b}(\mathcal{A}) and

[𝐧𝐬𝐩𝐩]=[𝐬𝐩𝐩𝐧]=[𝐧𝐩𝐩𝐬]=[𝐩𝐩𝐧𝐬]=[𝐬𝐧𝐩𝐩]=[𝐩𝐧𝐬𝐩]=[𝐩𝐬𝐧𝐩]𝔟(𝒜),|𝐧𝐬𝐩𝐩|=4.[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}]=[\mathbf{s}\mathbf{p}\mathbf{p}\mathbf{n}]=[\mathbf{n}\mathbf{p}\mathbf{p}\mathbf{s}]=[\mathbf{p}\mathbf{p}\mathbf{n}\mathbf{s}]=[\mathbf{s}\mathbf{n}\mathbf{p}\mathbf{p}]=[\mathbf{p}\mathbf{n}\mathbf{s}\mathbf{p}]=[\mathbf{p}\mathbf{s}\mathbf{n}\mathbf{p}]\in\mathfrak{b}(\mathcal{A})\,,\quad|\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}|=4\,. (5.1)

Consider the polynomial algebra [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})], i.e. the \mathbb{C}-span over

the basis monomials[w1][wr](r1,with all[wj]𝔟(𝒜))and1.\text{the basis monomials}\quad[w_{1}]\dots[w_{r}]\quad(r\geqslant 1\,,\;\;\text{with all}\quad[w_{j}]\in\mathfrak{b}(\mathcal{A}))\quad\text{and}\quad 1\,. (5.2)

We will write [w]m=[w][w]m[w]^{m}=\underbrace{[w]\dots[w]}_{m} for brevity, as well as 𝐚m=𝐚𝐚m\mathbf{a}^{m}=\underbrace{\mathbf{a}\dots\mathbf{a}}_{m} for a letter 𝐚𝒜\mathbf{a}\in\mathcal{A}. The degree of a monomial is defined as the sum of the lengths of the bracelets, deg([w1][wr])=|w1|++|wr|\deg\left([w_{1}]\dots[w_{r}]\right)=|w_{1}|+\dots+|w_{r}| (where by definition one puts deg(1)=0\deg(1)=0), and we denote by [𝔟(𝒜)]n[𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n}\subset\mathbb{C}[\mathfrak{b}(\mathcal{A})] the subset of polynomials of a given degree nn.

Consider the linear map Φ:Bn(δ)[𝔟(𝒜)]n\Phi:B_{n}(\delta)\to\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n} defined on any single diagram bBn(δ)b\in B_{n}(\delta) as follows:

  • 1)

    label each line of the diagram bb by a letter from 𝒜\mathcal{A}: the arcs in the upper (respectively, lower) row by 𝐧\mathbf{n} (respectively, 𝐬\mathbf{s}), the vertical lines by 𝐩\mathbf{p};

  • 2)

    identify the upper nodes with the corresponding lower nodes and straighten the obtained loops, which results in a set of bracelets [w1],,[wr]𝔟(𝒜)[w_{1}],\dots,[w_{r}]\in\mathfrak{b}(\mathcal{A});

  • 3)

    define Φ(b)=1n![w1][wr]\Phi(b)=\frac{1}{n!}\,[w_{1}]\dots[w_{r}].

For example, for the diagram b1b_{1} of Section 3.1 one has the above sequence of transformations

[Uncaptioned image][Uncaptioned image]{[Uncaptioned image]nsp,[Uncaptioned image]ns}15![𝐧𝐬𝐩][𝐧𝐬]\raisebox{-0.45pt}{\includegraphics[scale={0.7}]{b1.pdf}}\mapsto\raisebox{-0.45pt}{\includegraphics[scale={0.8}]{b1nsp.pdf}}\mapsto\Bigg{\{}{\raisebox{-0.35pt}{\includegraphics[scale={0.18}]{b1nsp_a.pdf}}\put(-39.0,22.0){$\mathbf{n}$}\put(-8.0,5.0){$\mathbf{s}$}\put(-39.0,-12.0){$\mathbf{p}$}\,,\,\raisebox{-0.45pt}{\includegraphics[scale={0.18}]{b1nsp_b.pdf}}\put(-45.0,0.0){$\mathbf{n}$}\put(-8.0,0.0){$\mathbf{s}$}\Bigg{\}}}\mapsto\frac{1}{5!}\,[\mathbf{nsp}][\mathbf{ns}]

The following result is a mere reformulation of [50, Theorem 2.11] (see also [13, Proposition 9] in terms of bracelets.

Proposition 5.1.

The map Φ\Phi is constant on the classes of conjugate elements in Bn(δ)B_{n}(\delta), so for any diagram bBn(δ)b\in B_{n}(\delta) the element Φ(γb)=n!Φ(b)\Phi(\gamma_{b})=n!\,\Phi(b) is a monic monomial in [w1][wr][𝔟(𝒜)]n[w_{1}]\dots[w_{r}]\in\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n}. The restriction Φ|Cn(δ)\Phi\big{|}_{C_{n}(\delta)} is the isomorphism of linear spaces.

Note that for a permutation s𝔖ns\in\mathfrak{S}_{n} one has Φ(γs)=[𝐩λ1][𝐩λr]\Phi(\gamma_{s})=[\mathbf{p}^{\lambda_{1}}]\dots[\mathbf{p}^{\lambda_{r}}]. Without loss of generality, λ1λr\lambda_{1}\geqslant\dots\geqslant\lambda_{r}, so one arrives at a rr-partition of nn, which reflects the well-known fact that classes of conjugate elements of 𝔖n\mathfrak{S}_{n} are in one-to-one correspondence with partitions of nn.

The bijection described in Proposition 5.1 allows one to construct the basis in Cn(δ)C_{n}(\delta) labelled by monic monomials in [𝔟(𝒜)]n\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n} [50, Corollary 2.12] (see also [13, Lemma 11]). Namely, by inverting Φ(γb)=ζ\Phi(\gamma_{b})=\zeta we introduce the following linear map eζ=γbCn(δ)e_{\zeta}=\gamma_{b}\in C_{n}(\delta) (bBn(δ)b\in B_{n}(\delta) is fixed modulo conjugacy equivalence).

Then the set{eζ:ζ[𝔟(𝒜)]n(monic monomials)}Cn(δ)is the sought basis.\text{Then the set}\;\;\left\{e_{\zeta}\;:\;\zeta\in\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n}\;\;\text{(monic monomials)}\right\}\subset C_{n}(\delta)\;\;\text{is the sought basis.} (5.3)

For example, one has the following basis in C3(N)C_{3}(N) parametrised by monomials 𝔟[𝔟(𝒜)]3\mathfrak{b}\in\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{3}:

e[𝐩]3=6[Uncaptioned image]=γ1Φ(γ1)=[𝐩][𝐩][𝐩],e[𝐩𝐩][𝐩]=2([Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image])=γs1Φ(γs1)=[𝐩2][𝐩],e[𝐩𝐩𝐩]=3([Uncaptioned image]+[Uncaptioned image])=γs1s2Φ(γs1s2)=[𝐩3],e[𝐧𝐬][𝐩]=2([Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image])=γd1=2A3Φ(γd1)=[𝐧𝐬][𝐩],e[𝐧𝐬𝐩]=[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]=γd1s2Φ(γd1s2)=[𝐧𝐬𝐩].\begin{array}[]{rll}e_{[\mathbf{p}]^{3}}&=6\,\ \raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{Id3.pdf}}=\gamma_{1}&\Phi(\gamma_{1})=[\mathbf{p}][\mathbf{p}][\mathbf{p}]\,,\vspace{0.2cm}\\ {e_{[\mathbf{pp}][\mathbf{p}]}}&=2\,\Bigl{(}\ \raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{perm132.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{perm321.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{perm213.pdf}}\ \Bigr{)}=\gamma_{s_{1}}&\Phi(\gamma_{s_{1}})=[\mathbf{p}^{2}][\mathbf{p}]\,,\vspace{0.2cm}\\ e_{[\mathbf{ppp}]}&=3\,\Bigl{(}\ \raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{perm312.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{perm231.pdf}}\ \Bigr{)}=\gamma_{s_{1}s_{2}}&\Phi(\gamma_{s_{1}s_{2}})=[\mathbf{p}^{3}]\,,\vspace{0.2cm}\\ e_{[\mathbf{ns}][\mathbf{p}]}&=2\,\Bigl{(}\ \raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_1.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_2.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_3.pdf}}\ \Bigr{)}\,=\gamma_{d_{1}}=2A_{3}&\Phi(\gamma_{d_{1}})=[\mathbf{n}\mathbf{s}][\mathbf{p}]\,,\vspace{0.2cm}\\ e_{[\mathbf{nsp}]}&=\,\ \raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_4.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_5.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_6.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_7.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_8.pdf}}+\raisebox{-0.4pt}{\includegraphics[width=25.0pt,height=20.0pt]{b3_9.pdf}}=\gamma_{d_{1}s_{2}}&\Phi(\gamma_{d_{1}s_{2}})=[\mathbf{n}\mathbf{s}\mathbf{p}]\,.\\ \end{array} (5.4)

The basis (5.3) allows us to express the left regular action of AnA_{n} in Cn(δ)C_{n}(\delta) in terms of a linear operator Δ\Delta on [𝔟(𝒜)]n\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n}:

Aneζ=ξ(monic monomials in[𝔟(𝒜)]n)eξΔξ=ζeΔ(ζ).A_{n}*e_{\zeta}=\sum_{\xi\;\text{(monic monomials in}\;\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n}\text{)}}e_{\xi}\,\Delta^{\xi}{}_{\zeta}=e_{\Delta(\zeta)}\,. (5.5)

The above formula serves as a definition which allows one to construct Δ\Delta by evaluating the products of Brauer diagrams in the left-hand-side of (5.5). For example, by direct computation one finds the left regular action of A3A_{3} on the above basis vectors of the above example, which in turn fixes the action of Δ\Delta on [𝔟(𝒜)]3\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{3}:

A3e[𝐩]3=3e[𝐧𝐬][𝐩],Δ([𝐩]3)=3[𝐧𝐬][𝐩],A3e[𝐩𝐩][𝐩]=e[𝐧𝐬][𝐩]+2e[𝐧𝐬𝐩],A3e[𝐩𝐩𝐩]=3e[𝐧𝐬][𝐩],A3e[𝐧𝐬][𝐩]=δe[𝐧𝐬][𝐩]+2e[𝐧𝐬𝐩],A3e[𝐧𝐬𝐩]=e[𝐧𝐬][𝐩]+(δ+1)e[𝐧𝐬𝐩],\begin{array}[]{rlcl}A_{3}*e_{[\mathbf{p}]^{3}}&=3\,e_{[\mathbf{ns}][\mathbf{p}]}\,,&\hbox{\multirowsetup$\Rightarrow$}&\Delta\big{(}[\mathbf{p}]^{3}\big{)}=3\,[\mathbf{ns}][\mathbf{p}]\,,\\ A_{3}*e_{[\mathbf{pp}][\mathbf{p}]}&=e_{[\mathbf{ns}][\mathbf{p}]}+2\,e_{[\mathbf{nsp}]}\,,&&\Delta\big{(}[\mathbf{pp}][\mathbf{p}]\big{)}=[\mathbf{ns}][\mathbf{p}]+2\,[\mathbf{nsp}]\,,\\ A_{3}*e_{[\mathbf{ppp}]}&=3\,e_{[\mathbf{ns}][\mathbf{p}]}\,,&&\Delta\big{(}[\mathbf{ppp}]\big{)}=3\,[\mathbf{ns}][\mathbf{p}]\,,\\ A_{3}*e_{[\mathbf{ns}][\mathbf{p}]}&=\delta\,e_{[\mathbf{ns}][\mathbf{p}]}+2\,e_{[\mathbf{nsp}]}\,,&&\Delta\big{(}[\mathbf{ns}][\mathbf{p}]\big{)}=\delta\,[\mathbf{ns}][\mathbf{p}]+2\,[\mathbf{nsp}]\,,\\ A_{3}*e_{[\mathbf{nsp}]}&=e_{[\mathbf{ns}][\mathbf{p}]}+(\delta+1)\,e_{[\mathbf{nsp}]}\,,&&\Delta\big{(}[\mathbf{nsp}]\big{)}=[\mathbf{ns}][\mathbf{p}]+(\delta+1)\,[\mathbf{nsp}]\,.\end{array} (5.6)

Our next goal consists in describing the action of Δ\Delta directly in terms of bracelets, which will allow us to treat (5.5) other way around and to read off the left action of AnA_{n} on Cn(δ)C_{n}(\delta) without addressing to diagram computations.

5.2 Left regular action of AnA_{n} on Cn(δ)C_{n}(\delta) via bracelets

Derivation of bracelets over 𝒜¯\bar{\mathcal{A}}.

Consider the polynomial algebra generated by bracelets over the extended alphabet 𝒜¯=𝒜𝒜˙𝒜¨\bar{\mathcal{A}}=\mathcal{A}\cup\dot{\mathcal{A}}\cup\ddot{\mathcal{A}}, where 𝒜˙={𝐬˙,𝐩˙}\dot{\mathcal{A}}=\{\dot{\mathbf{s}},\dot{\mathbf{p}}\}, 𝒜¨={𝐬¨}\ddot{\mathcal{A}}=\{\ddot{\mathbf{s}}\}. Consider the (linear) derivation map \partial which acts via Leibniz rule: on any monomial [w1][wk][w_{1}]\dots[w_{k}] as

([w1][wp])=j=1p[w1]([wj])[wp],\partial\big{(}[w_{1}]\dots[w_{p}]\big{)}=\sum_{j=1}^{p}[w_{1}]\dots\partial\big{(}[w_{j}]\big{)}\dots[w_{p}]\,, (5.7)

and on each bracelet [w]=[𝐚1𝐚][w]=[\mathbf{a}_{1}\dots\mathbf{a}_{\ell}] (𝐚j𝒜¯\mathbf{a}_{j}\in\bar{\mathcal{A}} for all j=1,,j=1,\dots,\ell) as

[𝐚1𝐚]=j=1[𝐚1(𝐚j)𝐚].\partial[\mathbf{a}_{1}\dots\mathbf{a}_{\ell}]=\sum_{j=1}^{\ell}[\mathbf{a}_{1}\dots\partial(\mathbf{a}_{j})\dots\mathbf{a}_{\ell}]\,. (5.8)

To fix \partial, we set

(𝐬)=𝐬˙,(𝐩)=𝐩˙,(𝐬˙)=𝐬¨,and(𝐧)=(𝐩˙)=(𝐬¨)=0(any bracelet where 0 occurs is put to 0).\begin{array}[]{l}\partial(\mathbf{s})=\dot{\mathbf{s}}\,,\;\;\partial(\mathbf{p})=\dot{\mathbf{p}}\,,\;\;\partial(\dot{\mathbf{s}})=\ddot{\mathbf{s}}\,,\\ \text{and}\quad\partial(\mathbf{n})=\partial(\dot{\mathbf{p}})=\partial(\ddot{\mathbf{s}})=0\;\;\text{(any bracelet where $0$ occurs is put to $0$)}\,.\end{array} (5.9)

In other words, the letters 𝐧\mathbf{n}, 𝐩˙\dot{\mathbf{p}} and 𝐬¨\ddot{\mathbf{s}} are constants with respect to \partial.

Define the set 𝔟(𝒜¯)𝔟(𝒜)\mathfrak{b}(\bar{\mathcal{A}})\supset\mathfrak{b}(\mathcal{A}) by extending 𝔟(𝒜)\mathfrak{b}(\mathcal{A}) by all possibilities to substitute the undotted letters 𝐬\mathbf{s}, 𝐩\mathbf{p} at some positions by their dotted counterparts in 𝒜¯\bar{\mathcal{A}}. For example, the bracelet [𝐧𝐬𝐩][\mathbf{nsp}] gives rise to the following bracelets [𝐧𝐬˙𝐩],[𝐧𝐬𝐩˙],[𝐧𝐬˙𝐩˙],[𝐧𝐬¨𝐩],[𝐧𝐬¨𝐩˙]𝔟(𝒜¯)[\mathbf{n}\dot{\mathbf{s}}\mathbf{p}],[\mathbf{n}\mathbf{s}\dot{\mathbf{p}}],[\mathbf{n}\dot{\mathbf{s}}\dot{\mathbf{p}}],[\mathbf{n}\ddot{\mathbf{s}}\mathbf{p}],[\mathbf{n}\ddot{\mathbf{s}}\dot{\mathbf{p}}]\in\mathfrak{b}(\bar{\mathcal{A}}). As a result, [𝔟(𝒜¯)]\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})] contains all images of [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})] upon consecutive application of \partial. The algebra [𝔟(𝒜¯)]\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})] is bi-graded:

[𝔟(𝒜¯)]=p0qp[𝔟(𝒜¯)]p(q),\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]=\bigoplus_{p\geqslant 0}\bigoplus_{q\leqslant p}\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{(q)}_{p}\,, (5.10)

where a monomial [w1][wr][𝔟(𝒜¯)]p(q)[w_{1}]\dots[w_{r}]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{(q)}_{p} has the total length pp (i.e. |w1|++|wr|=p|w_{1}|+\dots+|w_{r}|=p) and carries the total amount qq of dots above the letters. For small qq the degree (q)(q) will be indicated by qq times the symbol \prime, and [𝔟(𝒜¯)]n(0)=[𝔟(𝒜)]n\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{(0)}_{n}=\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n}. In the sequel, omitting one of the bi-degrees of a component in (5.10) will imply the direct sum over all possible values of the omitted component. For example, [𝐩˙],[𝐧𝐬˙],[𝐧𝐬˙][𝐩][𝔟(𝒜)¯][\dot{\mathbf{p}}],[\mathbf{n}\dot{\mathbf{s}}],[\mathbf{n}\dot{\mathbf{s}}][\mathbf{p}]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A})}]^{\prime}, [𝐧𝐬¨],[𝐧𝐬¨𝐩],[𝐧𝐬˙][𝐩˙𝐩][𝔟(𝒜¯)]′′[\mathbf{n}\ddot{\mathbf{s}}],[\mathbf{n}\ddot{\mathbf{s}}\mathbf{p}],[\mathbf{n}\dot{\mathbf{s}}][\dot{\mathbf{p}}\mathbf{p}]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{\prime\prime} and [𝐧𝐬˙𝐩],[𝐧𝐬¨𝐩],[𝐧𝐬˙][𝐩˙][𝔟(𝒜¯)]3[\mathbf{n}\dot{\mathbf{s}}\mathbf{p}],[\mathbf{n}\ddot{\mathbf{s}}\mathbf{p}],[\mathbf{n}\dot{\mathbf{s}}][\dot{\mathbf{p}}]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]_{3}. The next lemma is a simple consequence of the definition of \partial and the structure of 𝔟(𝒜¯)\mathfrak{b}(\bar{\mathcal{A}}).

Lemma 5.2.

The map \partial carries the bi-degree ()01\left({}^{1}_{0}\right),

:[𝔟(𝒜¯)]p(q)[𝔟(𝒜¯)]p(q+1),\partial:\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{(q)}_{p}\to\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{(q+1)}_{p}\,, (5.11)

and for any [w][𝔟(𝒜¯)]p(q)[w]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{(q)}_{p} there is pq+1[w]=0\partial^{p-q+1}[w]=0.

For example,

for[𝐧𝐬][𝐩][𝔟(𝒜¯)]3(0)one has([𝐧𝐬][𝐩])=[𝐧𝐬˙][𝐩]+[𝐧𝐬][𝐩˙][𝔟(𝒜¯)]3,2([𝐧𝐬][𝐩])=[𝐧𝐬¨][𝐩]+2[𝐧𝐬˙][𝐩˙][𝔟(𝒜¯)]3′′,3([𝐧𝐬][𝐩])=3[𝐧𝐬¨][𝐩˙][𝔟(𝒜¯)]3′′′,4([𝐧𝐬][𝐩])=0.\begin{array}[]{ll}\text{for}\;\;[\mathbf{n}\mathbf{s}][\mathbf{p}]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{(0)}_{3}\;\;\text{one has}&\partial\big{(}[\mathbf{n}\mathbf{s}][\mathbf{p}]\big{)}=[\mathbf{n}\dot{\mathbf{s}}][\mathbf{p}]+[\mathbf{n}\mathbf{s}][\dot{\mathbf{p}}]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{\prime}_{3}\,,\\ \hfill&\partial^{2}\big{(}[\mathbf{n}\mathbf{s}][\mathbf{p}]\big{)}=[\mathbf{n}\ddot{\mathbf{s}}][\mathbf{p}]+2\,[\mathbf{n}\dot{\mathbf{s}}][\dot{\mathbf{p}}]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{\prime\prime}_{3}\,,\\ \hfill&\partial^{3}\big{(}[\mathbf{n}\mathbf{s}][\mathbf{p}]\big{)}=3\,[\mathbf{n}\ddot{\mathbf{s}}][\dot{\mathbf{p}}]\in\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{\prime\prime\prime}_{3}\,,\\ \hfill&\partial^{4}\big{(}[\mathbf{n}\mathbf{s}][\mathbf{p}]\big{)}=0\,.\\ \end{array} (5.12)

Trace operation.

To introduce the final ingredient for construction of Δ\Delta, we consider the following [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})]-linear operation:

τ:[𝔟(𝒜¯)]′′[𝔟(𝒜)],\tau:\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{\prime\prime}\to\mathbb{C}[\mathfrak{b}(\mathcal{A})]\,, (5.13)

which is defined via the following rules. To formulate them, we accept a number of notations: i) we will write, for example, [𝐚v][\mathbf{a}v] or [𝐚u𝐛v][\mathbf{a}u\mathbf{b}v] to specify particular letters 𝐚,𝐛𝒜¯\mathbf{a},\mathbf{b}\in\bar{\mathcal{A}} in a bracelet, with the subwords u,vu,v being either empty or containing only letters from 𝒜\mathcal{A}, ii) we will write |w|𝐚|w|_{\mathbf{a}} for the number of occurrences of the letter 𝐚\mathbf{a} in ww iii) we will say that ww is fit if it is either empty or [w]𝔟(𝒜)[w]\in\mathfrak{b}(\mathcal{A}), and if each occurrence of 𝐬\mathbf{s} (if any) is followed by an occurrence of 𝐧\mathbf{n} at some position on the right. In the following expressions, the (sub)words 𝐩˙u,𝐩˙v\dot{\mathbf{p}}u,\dot{\mathbf{p}}v on the left-hand-sides are assumed to have u,vu,v fit unless else is specified:

τ:[𝐬¨u]2δ[𝐬u],\displaystyle\tau:[\ddot{\mathbf{s}}u]\mapsto 2\delta\,[\mathbf{s}u]\,, (5.14)
τ:[𝐬˙u𝐬˙v]2([𝐬u𝐬I(v)]+[𝐬u][𝐬v]),\displaystyle\tau:[\dot{\mathbf{s}}u\dot{\mathbf{s}}v]\mapsto 2\,\big{(}[\mathbf{s}u\mathbf{s}\,I(v)]+[\mathbf{s}u][\mathbf{s}v]\big{)}\,, τ:[𝐬˙u][𝐬˙v]2([𝐬u𝐬v]+[𝐬u𝐬I(v)]),\displaystyle\quad\tau:[\dot{\mathbf{s}}u][\dot{\mathbf{s}}v]\mapsto 2\,\big{(}[\mathbf{s}u\mathbf{s}v]+[\mathbf{s}u\mathbf{s}\,I(v)]\big{)}\,, (5.15)
τ:[𝐩˙u𝐬˙v][𝐩u𝐬I(v)]+[𝐩u][𝐬v],\displaystyle\tau:[\dot{\mathbf{p}}u\dot{\mathbf{s}}v]\mapsto[\mathbf{p}u\mathbf{s}\,I(v)]+[\mathbf{p}u][\mathbf{s}v]\,, τ:[𝐩˙u][𝐬˙v][𝐩u𝐬v]+[𝐩u𝐬I(v)],\displaystyle\quad\tau:[\dot{\mathbf{p}}u][\dot{\mathbf{s}}v]\mapsto[\mathbf{p}u\mathbf{s}v]+[\mathbf{p}u\mathbf{s}\,I(v)]\,, (5.16)
τ:[𝐩˙u𝐩˙v]{[𝐧u𝐬I(v)],if u,v are fit,[𝐧u][𝐬v],|u|𝐬>|u|𝐧,\displaystyle\tau:[\dot{\mathbf{p}}u\dot{\mathbf{p}}v]\mapsto\left\{\begin{array}[]{ll}[\mathbf{n}u\mathbf{s}I(v)]\,,&\text{if $u,v$ are fit}\,,\\ \left[\mathbf{n}u][\mathbf{s}v\right]\,,&|u|_{\mathbf{s}}>|u|_{\mathbf{n}}\end{array}\right.\,, τ:[𝐩˙u][𝐩˙v][𝐧u𝐬I(v)].\displaystyle\quad\tau:[\dot{\mathbf{p}}u][\dot{\mathbf{p}}v]\mapsto[\mathbf{n}u\mathbf{s}I(v)]\,. (5.19)
Lemma 5.3.

The rules (5.14)-(5.19) are correct and define τ\tau unambiguously. Namely, i) the monomials entering the left-hand-sides of the rules form a [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})]-basis999In other words, any element in [𝔟(𝒜¯)]′′\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{\prime\prime} is given by a unique combination of the left-hand-sides of (5.14)-(5.19) with the coefficients in [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})]. in [𝔟(𝒜¯)]′′\mathbb{C}[\mathfrak{b}(\bar{\mathcal{A}})]^{\prime\prime}, ii) if there are several representatives of a kind, the rules nevertheless lead to the same result.

For the proof see Appendix B.4.

We are in a position to express the operator Δ\Delta defined in (5.5) as a second-order differential operator on bracelets (see Appendix B.5 for proof).

Theorem 5.4.

The operator Δ\Delta defined in (5.5) is given by

Δ=12τ2.\Delta=\frac{1}{2}\,\tau\circ\partial^{2}\,. (5.20)

For example, consider the algebra C3(δ)B3(δ)C_{3}(\delta)\subset B_{3}(\delta). As a matter of consistency let us check that A3e[𝐩]3=A36=3e[𝐧𝐬][𝐩]A_{3}*e_{[\mathbf{p}]^{3}}=A_{3}*6=3\,e_{[\mathbf{ns}][\mathbf{p}]}. Indeed,

2[𝐩]3=6[𝐩˙]2[𝐩]Δ([𝐩]3)=3[𝐧𝐬][𝐩].\partial^{2}[\mathbf{p}]^{3}=6\,[\dot{\mathbf{p}}]^{2}[\mathbf{p}]\quad\Rightarrow\quad\Delta\left([\mathbf{p}]^{3}\right)=3\,[\mathbf{n}\mathbf{s}][\mathbf{p}]\,. (5.21)

For the rest one has:

2[𝐩𝐩][𝐩]=2[𝐩˙𝐩˙][𝐩]+4[𝐩˙𝐩][𝐩˙]Δ([𝐩𝐩][𝐩])=[𝐧𝐬][𝐩]+2[𝐧𝐬𝐩],2[𝐩𝐩𝐩]=6[𝐩˙𝐩˙𝐩]2[𝐧𝐬][𝐩]=[𝐧𝐬¨][𝐩]+2[𝐧𝐬˙][𝐩˙]2[𝐧𝐬𝐩]=[𝐧𝐬¨𝐩]+2[𝐧𝐬˙𝐩˙]\begin{array}[]{lcl}\partial^{2}\left[\mathbf{p}\mathbf{p}\right]\left[\mathbf{p}\right]=2\,\left[\dot{\mathbf{p}}\dot{\mathbf{p}}\right]\left[\mathbf{p}\right]+4\,\left[\dot{\mathbf{p}}\mathbf{p}\right]\left[\dot{\mathbf{p}}\right]&\hbox{\multirowsetup$\Rightarrow$}&\Delta\big{(}\left[\mathbf{p}\mathbf{p}\right]\left[\mathbf{p}\right]\big{)}=\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]+2\,\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\,,\\ \partial^{2}\left[\mathbf{p}\mathbf{p}\mathbf{p}\right]=6\,\left[\dot{\mathbf{p}}\dot{\mathbf{p}}\mathbf{p}\right]&&\Delta\big{(}\left[\mathbf{p}\mathbf{p}\mathbf{p}\right]\big{)}=3\,\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\,,\\ \partial^{2}\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]=\left[\mathbf{n}\ddot{\mathbf{s}}\right]\left[\mathbf{p}\right]+2\,\left[\mathbf{n}\dot{\mathbf{s}}\right]\left[\dot{\mathbf{p}}\right]&&\Delta\big{(}\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\big{)}=\delta\,\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]+2\,\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\,,\\ \partial^{2}\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]=\left[\mathbf{n}\ddot{\mathbf{s}}\mathbf{p}\right]+2\,\left[\mathbf{n}\dot{\mathbf{s}}\dot{\mathbf{p}}\right]&&\Delta\big{(}\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\big{)}=\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]+(\delta+1)\,\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\,,\\ \end{array} (5.22)

which reproduces the result (5.6) obtained by direct diagram-wise computation. From the third line in (5.22) one obtains (A3)2=δA3+e[𝐧𝐬𝐩](A_{3})^{2}=\delta\,A_{3}+e_{[\mathbf{nsp}]}, which gives a shortcut towards the final expression in (4.6).

Cardinality |C𝔖n(b)||C_{\mathfrak{S}_{n}}(b)| via symmetries of Φ(γb)\Phi(\gamma_{b}).

In the previous paragraphs we developed a technique to express the left regular action of AnA_{n} on the basis eζCn(δ)e_{\zeta}\in C_{n}(\delta) (where ζ[𝔟(𝒜)]\zeta\in\mathbb{C}[\mathfrak{b}(\mathcal{A})] are monic monomials) avoiding diagram-wise computations. Nevertheless, in order to apply the final expanded form of PnP_{n} (4.1) to a tensor, the basis elements eζ=γbe_{\zeta}=\gamma_{b} should be expressed as sums of conjugate Brauer diagrams (3.17). In this respect it is convenient to rescale the basis eζe_{\zeta} and use the normalised sums

eˇζ=1|C𝔖n(b)|γb,\check{e}_{\zeta}=\frac{1}{|C_{\mathfrak{S}_{n}}(b)|}\,\gamma_{b}\,, (5.23)

which are sums of conjugate diagrams with the overall coefficient 11 (recall (3.17)). In particular, An=eˇ[𝐧𝐬][𝐩]n2A_{n}=\check{e}_{[\mathbf{ns}][\mathbf{p}]^{n-2}}. It appears that the coefficient |C𝔖n(b)||C_{\mathfrak{S}_{n}}(b)| can be expressed in terms of certain symmetry properties of the representatives in bracelets, which allows one to avoid diagram computations again.

Consider the group of cyclic permutations \mathbb{Z}_{\ell} which acts on the words of the length \ell. Recall that I(w)I(w) denotes the inversion of a word ww. Given any word ww of length \ell, define its turnover stabilizer:

S(w)={h:eitherh(w)=worh(w)=I(w)}.S(w)=\left\{h\in\mathbb{Z}_{\ell}\;\;:\;\;\text{either}\;\;h(w)=w\;\;\text{or}\;\;h(w)=I(w)\right\}\,. (5.24)

Note that both conditions in (5.24) are simultaneously satisfied by an element hh\in\mathbb{Z}_{\ell} only if w=I(w)w=I(w), i.e. the word is inversion-invariant. In the case of representatives of bracelet from 𝔟(𝒜)\mathfrak{b}(\mathcal{A}), inversion-invariant words appear only as representatives of [𝐩𝐩][\mathbf{p}\dots\mathbf{p}]. In this case |S(𝐩𝐩)|=\big{|}S(\underbrace{\mathbf{p}\dots\mathbf{p}}_{\ell})\big{|}=\ell. In the other cases, when 𝐧\mathbf{n} and 𝐬\mathbf{s} occur in a bracelet from 𝔟(𝒜)\mathfrak{b}(\mathcal{A}), none of its representatives is inversion-invariant, so a permutation hh\in\mathbb{Z}_{\ell} satisfies at most one of the conditions in (5.24).

As a side remark, note the following properties concerning the definition of the stabilizer (5.24). It can happen that the conditions in (5.24) are satisfied only for the trivial transformation: for example, in the case of the bracelet [𝐧𝐬𝐩][\mathbf{n}\mathbf{s}\mathbf{p}]. Another observation is that the existence of a transformation hh\in\mathbb{Z}_{\ell} such that h(w)=wh(w)=w (respectively, h(w)=I(w)h(w)=I(w)) does not imply necessarily the existence of hh^{\prime}\in\mathbb{Z}_{\ell} such that h(w)=I(w)h^{\prime}(w)=I(w) (respectively, h(w)=wh^{\prime}(w)=w): as an example, consider [𝐧𝐬𝐩𝐧𝐬𝐩][\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{n}\mathbf{s}\mathbf{p}] (respectively, [𝐧𝐬𝐧𝐩𝐬𝐩][\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}]).

For a bracelet [w]𝔟(𝒜)[w]\in\mathfrak{b}(\mathcal{A}), the cardinality |S(w)||S(w)| does not depend on particular choice of a representative, so the function st([w])=|S(w)|\mathrm{st}\big{(}[w]\big{)}=|S(w)| is well-defined on bracelets.

For a monomial in [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})] define its turnover stability index as follows:

st([w1]m1[wr]mr)=j=1rst([wj])mj!.\mathrm{st}\big{(}[w_{1}]^{m_{1}}\dots[w_{r}]^{m_{r}}\big{)}=\prod_{j=1}^{r}\mathrm{st}\big{(}[w_{j}]\big{)}\,m_{j}!\,. (5.25)

The following lemma gives a convenient method of calculating the coefficient in (3.17) (see Appendix B.6 for proof).

Lemma 5.5.

For any diagram bBn(δ)b\in B_{n}(\delta) holds |C𝔖n(b)|=st(Φ(γb))|C_{\mathfrak{S}_{n}}(b)|=\mathrm{st}\big{(}\Phi(\gamma_{b})\big{)}. Thus, the change of basis (5.23) is written as

eˇζ=1st(ζ)eζ.\check{e}_{\zeta}=\frac{1}{\mathrm{st}(\zeta)}\,e_{\zeta}\,. (5.26)

(Non)-commutativity of Cn(δ)C_{n}(\delta).

Restriction to the subalgebra 𝔖nBn(δ)\mathbb{C}\mathfrak{S}_{n}\subset B_{n}(\delta) is the central ingredient in the proposed construction of the traceless projector. The algebra Cn(δ)C_{n}(\delta) centralises 𝔖n\mathbb{C}\mathfrak{S}_{n} in Bn(δ)B_{n}(\delta) and vice versa. Along the same lines as explained in Section 2 about the centraliser algebras, Cn(δ)C_{n}(\delta) takes care of the multiplicities in the decomposition of a simple Bn(δ)B_{n}(\delta)-module into simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules: this is because Cn(δ)C_{n}(\delta) contains all intertwiners between equivalent 𝔖n\mathbb{C}\mathfrak{S}_{n}-submodules. By applying Schur’s lemma, Cn(δ)C_{n}(\delta) is commutative iff multiplicities in the decompositions in question are bounded by 11.

In order to analyse commutativity of Cn(δ)C_{n}(\delta) we make use of the flip operation ()(\,\cdot\,)^{*} (3.15) (defined as a linear map for any δ\delta\in\mathbb{C}), such that Cn(δ)C_{n}(\delta) is the algebra with an anti-involution. Then by [51, Lemma 2.3], the algebra Cn(δ)C_{n}(\delta) is commutative iff for any uCn(δ)u\in C_{n}(\delta) holds uu=uuu^{*}u=uu^{*}. To analyse the latter condition we will make use of the basis (5.3) and introduce an involution on bracelets which commutes with the isomorphism in Proposition 5.1. Namely, upon Bn(δ)bbB_{n}(\delta)\ni b\mapsto b^{*} the arcs in the upper (respectively, lower) row are placed to the lower (respectively, upper) row, while vertical lines remain vertical. Then by construction of the map Φ\Phi, the image Φ(b)\Phi(b^{*}) is obtained from Φ(b)\Phi(b) by mapping 𝐧𝐬\mathbf{n}\mapsto\mathbf{s}, 𝐬𝐧\mathbf{s}\mapsto\mathbf{n} in each bracelet101010In other words, one extends the involutive map 𝐧𝐬\mathbf{n}\mapsto\mathbf{s}, 𝐬𝐧\mathbf{s}\mapsto\mathbf{n} on 𝒜\mathcal{A} to the automorphism of the monoid of words over 𝒜\mathcal{A}, and then to [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})] by linearity. (see [50, Lemma 2.14]). Abusing notation once again,

for anyξ[𝔟(𝒜)]defineξas obtained by𝐧𝐬,𝐬𝐧.\text{for any}\quad\xi\in\mathbb{C}[\mathfrak{b}(\mathcal{A})]\quad\text{define}\quad\xi^{*}\quad\text{as obtained by}\quad\mathbf{n}\mapsto\mathbf{s}\,,\;\;\mathbf{s}\mapsto\mathbf{n}. (5.27)

With (5.27) at hand, Cn(δ)C_{n}(\delta) is commutative iff eξeξ=eξeξe_{\xi}e_{\xi^{*}}=e_{\xi^{*}}e_{\xi} for any monic monomial ξ[𝔟(𝒜)]n\xi\in\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n}. The analysis of the latter condition constitutes the proof of the following lemma (see Appendix B.7 for proof).

Lemma 5.6.

For any δ\delta\in\mathbb{C} the algebra Cn(δ)Bn(δ)C_{n}(\delta)\subset B_{n}(\delta) is commutative for n5n\leqslant 5, and is non-commutative for n6n\geqslant 6. More precisely, for n6n\geqslant 6 and fmax=n2f_{\mathrm{max}}=\lfloor\tfrac{n}{2}\rfloor:

  • i)

    the quotient algebra Cn(δ)/J(f)C_{n}(\delta)/J^{(f)} is commutative for f=1,2f=1,2 and is non-commutative for f=3,,fmaxf=3,\dots,f_{\mathrm{max}};

  • ii)

    the subalgebra Cn(δ)J(f)C_{n}(\delta)\cap J^{(f)} is commutative for f=fmaxf=f_{\mathrm{max}} and is non-commutative for f=0,,fmax1f=0,\dots,f_{\mathrm{max}}-1.

As a particular example, Cn(δ)/J(1)Cn(δ)𝔖nC_{n}(\delta)/J^{(1)}\cong C_{n}(\delta)\cap\mathbb{C}\mathfrak{S}_{n}, which is known to be the center of 𝔖n\mathbb{C}\mathfrak{S}_{n}. For any s𝔖ns\in\mathfrak{S}_{n} one has Φ(γs)=[𝐩λ1][𝐩λr]\Phi(\gamma_{s})=[\mathbf{p}^{\lambda_{1}}]\dots[\mathbf{p}^{\lambda_{r}}], so Φ(γs)=Φ(γs)\Phi(\gamma_{s})=\Phi(\gamma_{s})^{*}. Since ()(\,\cdot\,)^{*} commutes with the isomorphism in Proposition 5.1, one has γs=(γs)\gamma_{s}=(\gamma_{s})^{*}. Therefore, for any any central element u𝔖nu\in\mathbb{C}\mathfrak{S}_{n} holds u=uu^{*}=u and 𝔯(u)=𝔯(u)\mathfrak{r}(u)^{*}=\mathfrak{r}(u). In particular, for central Young symmetrisers (utilised in (4.23)) one has (z(μ))=z(μ)(z^{(\mu)})^{*}=z^{(\mu)}.

6 Splitting idempotent

Traceless subspace as a quotient space.

In order to introduce the subject of this section, let us recall that the subspace of traceless tensors can be identified with equivalence classes Vn/WV^{\otimes n}/W (of VnV^{\otimes n} modulo WW, the subspace of tensors proportional to metric). If one denotes the canonical embedding ι:WVn\iota:W\rightarrow V^{\otimes n} and the projection π:VnVn/W\pi:V^{\otimes n}\rightarrow V^{\otimes n}/W, then the existence of a traceless projection is equivalent to existence of a section σ:Vn/WVn\sigma:V^{\otimes n}/W\rightarrow V^{\otimes n}, such that the following short exact sequence splits:

0WιVn\ext@arrow0359\arrowfill@--π\ext@arrow3095\arrowfill@--leftarrowfill:existsσVn/W0,whereπσ=id,and for𝔓n=σπone hasVn=W𝔓nVn.\begin{array}[]{l}0\xrightarrow{\quad}W\xrightarrow{\quad\iota\quad}V^{\otimes n}\mathrel{\raise 2.36806pt\hbox{$\ext@arrow 0359\arrowfill@\relbar\relbar\rightarrow{\phantom{\text{exists}\;\;\sigma}}{\pi}$}\kern-152.9203pt\lower 2.36806pt\hbox{$\ext@arrow 3095\arrowfill@\relbar\relbar\rightarrow_{l}eftarrow_{f}ill:{\text{exists}\;\;\sigma}{\phantom{\pi}}$}}V^{\otimes n}/W\xrightarrow{\quad}0\,,\;\;\text{where}\;\;\pi\circ\sigma=\mathrm{id}\,,\\ \text{and for}\;\;\mathfrak{P}_{n}=\sigma\circ\pi\;\;\text{one has}\;\;V^{\otimes n}=W\oplus\mathfrak{P}_{n}V^{\otimes n}\,.\end{array} (6.1)

In this context, the projector 𝔓n\mathfrak{P}_{n} is referred to as splitting idempotent of the above short exact sequence. By Theorem 4.1, 𝔓n\mathfrak{P}_{n} is be expressed as a 𝔯\mathfrak{r}-image of an element PnBn(εN)P_{n}\in B_{n}(\varepsilon N) (4.1).

In this section we are interested in the properties of the element PnP_{n} itself, as an operator in the left regular Bn(δ)B_{n}(\delta)-module (i.e. its action on the vector space Bn(δ)B_{n}(\delta) via left multiplication). We assume the semisimple regime for Bn(δ)B_{n}(\delta) by allowing δ\{0}\delta\in\mathbb{C}\backslash\{0\}, while for δ\{0}\delta\in\mathbb{Z}\backslash\{0\} we assume δ2(n1)\delta\leqslant-2(n-1) and δn1\delta\geqslant n-1 (for the semisimplicity criterion see [32] and references therein).

PnP_{n} as a splitting idempotent.

In analogy with the traceless projection of VnV^{\otimes n}, which is annihilated by the action of J(1)J^{(1)}, one can be interested in constructing the “traceless” projection of the left regular Bn(δ)B_{n}(\delta)-module. Due to the equivalence of the left and right regular modules of Bn(δ)B_{n}(\delta) via ()(\,\cdot\,)^{*}, the sought projector should be a central idempotent P¯nBn(δ)\bar{P}_{n}\in B_{n}(\delta) which provides the decomposition Bn(δ)=J(1)P¯nBn(δ)B_{n}(\delta)=J^{(1)}\oplus\bar{P}_{n}B_{n}(\delta) with P¯nBn(δ)𝔖n\bar{P}_{n}B_{n}(\delta)\cong\mathbb{C}\mathfrak{S}_{n} (direct product of algebras). In other words, one looks for a splitting idempotent for the following exact sequence similar to the one in (6.1):

0J(1)Bn(δ)\ext@arrow0359\arrowfill@--π¯\ext@arrow3095\arrowfill@--leftarrowfill:existsσ¯Bn(δ)/J(1)(𝔖n)0,whereπ¯σ¯=id,and forP¯n=σ¯π¯one hasBn(δ)=J(1)P¯nBn(δ).\begin{array}[]{l}0\xrightarrow{\quad}J^{(1)}\xrightarrow{\quad}B_{n}(\delta)\mathrel{\raise 2.36806pt\hbox{$\ext@arrow 0359\arrowfill@\relbar\relbar\rightarrow{\phantom{\text{exists}\;\;\bar{\sigma}}}{\bar{\pi}}$}\kern-151.50594pt\lower 2.36806pt\hbox{$\ext@arrow 3095\arrowfill@\relbar\relbar\rightarrow_{l}eftarrow_{f}ill:{\text{exists}\;\;\bar{\sigma}}{\phantom{\bar{\pi}}}$}}B_{n}(\delta)/J^{(1)}\,(\cong\mathbb{C}\mathfrak{S}_{n})\xrightarrow{\quad}0\,,\;\;\text{where}\;\;\bar{\pi}\circ\bar{\sigma}=\mathrm{id}\,,\\ \text{and for}\;\;\bar{P}_{n}=\bar{\sigma}\circ\bar{\pi}\;\;\text{one has}\;\;B_{n}(\delta)=J^{(1)}\oplus\bar{P}_{n}B_{n}(\delta)\,.\end{array} (6.2)

Such idempotent exists and is unique [13].

Theorem 6.1.

Let δ\{0}\delta\in\mathbb{C}\backslash\{0\} and nn\in\mathbb{N} are such that Bn(δ)B_{n}(\delta) is semisimple. Then P¯n\bar{P}_{n} is constructed as (4.1) by dropping all restrictions on the numbers of rows/columns in the Young diagrams in the definitions (3.34), (3.35), (3.38) and (3.41).

Proof.

First, let us show that P¯nZ(Bn(δ))\bar{P}_{n}\in Z(B_{n}(\delta)). By construction, P¯nCn(δ)\bar{P}_{n}\in C_{n}(\delta), so it commutes with 𝔖nBn(δ)\mathbb{C}\mathfrak{S}_{n}\subset B_{n}(\delta). We will prove that P¯nJ(1)=J(1)P¯n=0\bar{P}_{n}J^{(1)}=J^{(1)}\bar{P}_{n}=0, which will imply in turn that P¯n\bar{P}_{n} commutes also with any diagram from J(1)J^{(1)}.

The fact that Bn(δ)B_{n}(\delta) is semisimple means that the left regular Bn(δ)B_{n}(\delta)-module decomposes as a direct sum of simple Bn(δ)B_{n}(\delta)-modules Mn(λ)M^{(\lambda)}_{n} indexed by λn2f\lambda\vdash n-2f for all f=0,,n2f=0,\dots,\lfloor\tfrac{n}{2}\rfloor (each module occurring at least once). Upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n}, occurrence of simple 𝔖n\mathbb{C}\mathfrak{S}_{n}-modules L(μ)Mn(λ)L^{(\mu)}\subset M^{(\lambda)}_{n} is described by Lemma 3.2, which becomes the criterion due to simplicity of the standard modules (with all restrictions on the size of Young diagrams omitted).

The eigenvalue 0 of AnA_{n} occurs only for f=0f=0 in (3.43). This is obvious for non-integer values of δ\delta due to Lemma 3.3, while for integer ones the bounds of the semisimple regime of Bn(δ)B_{n}(\delta) allow one to apply the alternative proof of Lemma 3.1, which is presented in Appendix B.3. By construction, left multiplication of Bn(δ)B_{n}(\delta) by P¯n\bar{P}_{n} annihilates exactly the subspaces with non-zero eigenvalues of AnA_{n}, so P¯nJ(1)=0\bar{P}_{n}J^{(1)}=0. To prove that J(1)P¯n=0J^{(1)}\bar{P}_{n}=0 as well, note that (J(1))=J(1)(J^{(1)})^{*}=J^{(1)}. The fact that (An)=An(A_{n})^{*}=A_{n} implies that (P¯n)=P¯n(\bar{P}_{n})^{*}=\bar{P}_{n}, so (J(1)P¯n)=P¯nJ(1)=0(J^{(1)}\bar{P}_{n})^{*}=\bar{P}_{n}J^{(1)}=0. As an immediate consequence, (P¯n)2=P¯n(\bar{P}_{n})^{2}=\bar{P}_{n} because AnJ(1)A_{n}\in J^{(1)}.

To sum up, P¯n\bar{P}_{n} is a central idempotent. The fact that the eigenvalue 0 of AnA_{n} happens only for f=0f=0 implies that P¯nBn(δ)=P¯n𝔖n𝔖n\bar{P}_{n}B_{n}(\delta)=\bar{P}_{n}\,\mathbb{C}\mathfrak{S}_{n}\cong\mathbb{C}\mathfrak{S}_{n}, so P¯n\bar{P}_{n} splits the short exact sequence (6.2). ∎

7 Summary: the traceless projector for n=4n=4

Let us sum up the proposed approach of constructing the traceless projector by constructing the splitting idempotent P¯nBn(δ)\bar{P}_{n}\in B_{n}(\delta) step by step. Throughout this section the parameter δ0\delta\neq 0 is assumed to be generic complex, such that the algebra Bn(δ)B_{n}(\delta) is semisimple. To get the traceless projector (4.1), one simply puts δεN\delta\mapsto\varepsilon N and ignores the elements of spec×(A4)\mathrm{spec}^{\times}(A_{4}) which either vanish or carry the sign ε-\varepsilon.

Eigenvalues of A4A_{4}.

To arrange the set of eigenvalues let us make use of the fact that

spec×(A4)=spec(4)×(An)spec(3,1)×(An)spec(2,2)×(An)spec(2,1,1)×(An),\mathrm{spec}^{\times}(A_{4})=\mathrm{spec}^{\times}_{(4)}(A_{n})\cup\mathrm{spec}^{\times}_{(3,1)}(A_{n})\cup\mathrm{spec}^{\times}_{(2,2)}(A_{n})\cup\mathrm{spec}^{\times}_{(2,1,1)}(A_{n})\,, (7.1)

so one uses (4.15) defined in Section 4.2. One has

spec(4)×(A4)\displaystyle\mathrm{spec}_{(4)}^{\times}(A_{4}) ={δ+4, 2(δ+2)},spec(3,1)×(A4)={δ,δ+2}\displaystyle=\{{\delta+4\,,\,2(\delta+2)}\},\qquad\mathrm{spec}_{(3,1)}^{\times}(A_{4})=\{{\delta\,,\,\delta+2}\}\qquad (7.2)
spec(2,2)×(A4)\displaystyle\mathrm{spec}_{(2,2)}^{\times}(A_{4}) ={δ2, 2(δ1)},spec(2,1,1)×(A4)={δ2},\displaystyle=\{{\delta-2\,,\,2(\delta-1)}\},\qquad\mathrm{spec}_{(2,1,1)}^{\times}(A_{4})=\{{\delta-2}\}\,,

A this level one can have zeros for particular non-zero integer values δ{4,2,1,2}\delta\in\{-4,-2,1,2\}, which mark exactly the non-semisimple regime of Bn(δ)B_{n}(\delta). Also for δ=1\delta=1, spec(2,2)×(A4)\mathrm{spec}_{(2,2)}^{\times}(A_{4}) and spec(2,1,1)×(A4)\mathrm{spec}_{(2,1,1)}^{\times}(A_{4}) contain a negative element to be excluded from the consideration. Nevertheless we proceed by assuming that all the above elements enter spec×(A4)\mathrm{spec}^{\times}(A_{4}). In order to obtain the additive form of the projector we now need to expand the product of factors (1α1A4)\left(1-\alpha^{-1}A_{4}\right) over all αspec×(A4)\alpha\in\mathrm{spec}^{\times}(A_{4}). To do so we take advantage of the algorithm described in section (5).

Action of A4A_{4} on C4(δ)C_{4}(\delta) and expansion of the factorised form of P4P_{4}.

The second derivative of the elements in [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})] relevant for the construction of P4P_{4} reads:

2[𝐧𝐬][𝐩][𝐩]=[𝐧𝐬¨][𝐩][𝐩]+4[𝐧𝐬˙][𝐩˙][𝐩]+2[𝐧𝐬][𝐩˙][𝐩˙],2[𝐧𝐩𝐬𝐩]=[𝐧𝐩𝐬¨𝐩]+4[𝐧𝐩𝐬˙𝐩˙]+2[𝐧𝐩˙𝐬𝐩˙],2[𝐧𝐬][𝐩𝐩]=[𝐧𝐬¨][𝐩𝐩]+4[𝐧𝐬˙][𝐩˙𝐩]+2[𝐧𝐬][𝐩˙𝐩˙],2[𝐧𝐬][𝐧𝐬]=2([𝐧𝐬¨][𝐧𝐬]+[𝐧𝐬˙][𝐧𝐬˙]),2[𝐧𝐬𝐩][𝐩]=[𝐧𝐬¨𝐩][𝐩]+2([𝐧𝐬˙𝐩˙][𝐩]+[𝐧𝐬˙𝐩][𝐩˙]+[𝐧𝐬𝐩˙][𝐩˙]),2[𝐧𝐬𝐧𝐬]=2([𝐧𝐬¨𝐧𝐬]+[𝐧𝐬˙𝐧𝐬˙]),2[𝐧𝐬𝐩𝐩]=[𝐧𝐬¨𝐩𝐩]+2([𝐧𝐬˙𝐩˙𝐩]+[𝐧𝐬˙𝐩𝐩˙]+[𝐧𝐬𝐩˙𝐩˙]).\begin{array}[]{lcl}\partial^{2}\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]=\left[\mathbf{n}\ddot{\mathbf{s}}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]+4\,\left[\mathbf{n}\dot{\mathbf{s}}\right]\left[\dot{\mathbf{p}}\right]\left[\mathbf{p}\right]+2\,\left[\mathbf{n}\mathbf{s}\right]\left[\dot{\mathbf{p}}\right]\left[\dot{\mathbf{p}}\right]\,,&&\partial^{2}\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]=\left[\mathbf{n}\mathbf{p}\ddot{\mathbf{s}}\mathbf{p}\right]+4\,\left[\mathbf{n}\mathbf{p}\dot{\mathbf{s}}\dot{\mathbf{p}}\right]+2\,\left[\mathbf{n}\dot{\mathbf{p}}\mathbf{s}\dot{\mathbf{p}}\right]\,,\\ \partial^{2}\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right]=\left[\mathbf{n}\ddot{\mathbf{s}}\right]\left[\mathbf{p}\mathbf{p}\right]+4\,\left[\mathbf{n}\dot{\mathbf{s}}\right]\left[\dot{\mathbf{p}}\mathbf{p}\right]+2\left[\mathbf{n}\mathbf{s}\right]\left[\dot{\mathbf{p}}\dot{\mathbf{p}}\right]\,,&&\partial^{2}\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]=2\,(\left[\mathbf{n}\ddot{\mathbf{s}}\right]\left[\mathbf{n}\mathbf{s}\right]+\left[\mathbf{n}\dot{\mathbf{s}}\right]\left[\mathbf{n}\dot{\mathbf{s}}\right])\,,\\ \partial^{2}\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]=\left[\mathbf{n}\ddot{\mathbf{s}}\mathbf{p}\right]\left[\mathbf{p}\right]+2\,(\,\left[\mathbf{n}\dot{\mathbf{s}}\dot{\mathbf{p}}\right]\left[\mathbf{p}\right]+\left[\mathbf{n}\dot{\mathbf{s}}\mathbf{p}\right]\left[\dot{\mathbf{p}}\right]+\left[\mathbf{n}\mathbf{s}\dot{\mathbf{p}}\right]\left[\dot{\mathbf{p}}\right]\,)\,,&&\partial^{2}\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]=2\,(\left[\mathbf{n}\ddot{\mathbf{s}}\mathbf{n}\mathbf{s}\right]+\left[\mathbf{n}\dot{\mathbf{s}}\mathbf{n}\dot{\mathbf{s}}\right])\,,\\ \partial^{2}\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]=\left[\mathbf{n}\ddot{\mathbf{s}}\mathbf{p}\mathbf{p}\right]+2\,(\left[\mathbf{n}\dot{\mathbf{s}}\dot{\mathbf{p}}\mathbf{p}\right]+\left[\mathbf{n}\dot{\mathbf{s}}\mathbf{p}\dot{\mathbf{p}}\right]+\left[\mathbf{n}\mathbf{s}\dot{\mathbf{p}}\dot{\mathbf{p}}\right])\,.\\ \end{array}

Evaluating the trace τ\tau (5.14)-(5.19) on the previous expressions and using Theorem 5.4 yields:

Ane[𝐧𝐬][𝐩][𝐩]=δe[𝐧𝐬][𝐩][𝐩]+4e[𝐧𝐬𝐩][𝐩]+e[𝐧𝐬][𝐧𝐬],\displaystyle A_{n}*e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}=\delta\,e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}+4\,e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}\,, (7.3)
Ane[𝐧𝐬][𝐩𝐩]=δe[𝐧𝐬][𝐩𝐩]+4e[𝐧𝐬𝐩𝐩]+e[𝐧𝐬][𝐧𝐬],\displaystyle A_{n}*e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right]}=\delta\,e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right]}+4\,e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}\,,
Ane[𝐧𝐬𝐩][𝐩]=(δ+1)e[𝐧𝐬𝐩][𝐩]+e[𝐧𝐬][𝐩][𝐩]+e[𝐧𝐬𝐩𝐩]+e[𝐧𝐩𝐬𝐩]+e[𝐧𝐬𝐧𝐬],\displaystyle A_{n}*e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}=\left(\delta+1\right)\,e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}\,,
Ane[𝐧𝐬𝐩𝐩]=(δ+1)e[𝐧𝐬𝐩𝐩]+e[𝐧𝐬][𝐩𝐩]+e[𝐧𝐬𝐩][𝐩]+e[𝐧𝐩𝐬𝐩]+e[𝐧𝐬𝐧𝐬],\displaystyle A_{n}*e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}=\left(\delta+1\right)\,e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}\,,
Ane[𝐧𝐩𝐬𝐩]=δe[𝐧𝐩𝐬𝐩]+2e[𝐧𝐬𝐩][𝐩]+2e[𝐧𝐬𝐩𝐩]+e[𝐧𝐬][𝐧𝐬],\displaystyle A_{n}*e_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}=\delta\,e_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}+2\,e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}+2\,e_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}+e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}\,,
Ane[𝐧𝐬][𝐧𝐬]= 2δe[𝐧𝐬][𝐧𝐬]+4e[𝐧𝐬𝐧𝐬],\displaystyle A_{n}*e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}=\,2\delta\,e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}+4\,e_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}\,,
Ane[𝐧𝐬𝐧𝐬]= 2(δ+1)e[𝐧𝐬𝐧𝐬]+2e[𝐧𝐬][𝐧𝐬].\displaystyle A_{n}*e_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}=\,2(\delta+1)\,e_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}+2\,e_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}\,.

Before expanding the factorized form of the projector we need to express the above expression in the normalized basis eˇζ=1st(ζ)eζ\check{e}_{\zeta}=\dfrac{1}{\mathrm{st}(\zeta)}\,e_{\zeta} described in Lemma 5.5. The turnover stability index of each basis element of [𝔟(𝒜)]4\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{4} reads:

st([𝐧𝐬][𝐩][𝐩])=4,st([𝐧𝐬][𝐩𝐩])=4,st([𝐧𝐬𝐩][𝐩])=1,st([𝐧𝐬𝐩𝐩])=1,st([𝐧𝐩𝐬𝐩])=2,st([𝐧𝐬][𝐧𝐬])=8,st([𝐧𝐬𝐧𝐬])=4.\begin{array}[]{llllc}\mathrm{st}(\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right])=4\,,&\mathrm{st}(\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right])=4\,,&\mathrm{st}(\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right])=1\,,&\mathrm{st}(\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right])=1\,,\\ \mathrm{st}(\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right])=2\,,&\mathrm{st}(\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right])=8\,,&\mathrm{st}(\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right])=4\,.&\\ \end{array}

Therefore, in the normalized basis the relations (7.3) transform into :

Aneˇ[𝐧𝐬][𝐩][𝐩]=δeˇ[𝐧𝐬][𝐩][𝐩]+eˇ[𝐧𝐬𝐩][𝐩]+2eˇ[𝐧𝐬][𝐧𝐬],\displaystyle A_{n}*\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}=\delta\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}+\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}+2\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}\,, (7.4)
Aneˇ[𝐧𝐬][𝐩𝐩]=δeˇ[𝐧𝐬][𝐩𝐩]+eˇ[𝐧𝐬𝐩𝐩]+2eˇ[𝐧𝐬][𝐧𝐬],\displaystyle A_{n}*\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right]}=\delta\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right]}+\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}+2\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}\,,
Aneˇ[𝐧𝐬𝐩][𝐩]=(δ+1)eˇ[𝐧𝐬𝐩][𝐩]+4eˇ[𝐧𝐬][𝐩][𝐩]+eˇ[𝐧𝐬𝐩𝐩]+2eˇ[𝐧𝐩𝐬𝐩]+4eˇ[𝐧𝐬𝐧𝐬],\displaystyle A_{n}*\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}=\left(\delta+1\right)\,\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}+4\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}+\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}+2\,\check{e}_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}+4\,\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}\,,
Aneˇ[𝐧𝐬𝐩𝐩]=(δ+1)eˇ[𝐧𝐬𝐩𝐩]+4eˇ[𝐧𝐬][𝐩𝐩]+eˇ[𝐧𝐬𝐩][𝐩]+2eˇ[𝐧𝐩𝐬𝐩]+4eˇ[𝐧𝐬𝐧𝐬],\displaystyle A_{n}*\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}=\left(\delta+1\right)\,\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}+4\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right]}+\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}+2\,\check{e}_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}+4\,\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}\,,
Aneˇ[𝐧𝐩𝐬𝐩]=δeˇ[𝐧𝐩𝐬𝐩]+eˇ[𝐧𝐬𝐩][𝐩]+eˇ[𝐧𝐬𝐩𝐩]+4eˇ[𝐧𝐬][𝐧𝐬],\displaystyle A_{n}*\check{e}_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}=\delta\,\check{e}_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}+\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}+\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}+4\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}\,,
Aneˇ[𝐧𝐬][𝐧𝐬]= 2δeˇ[𝐧𝐬][𝐧𝐬]+2eˇ[𝐧𝐬𝐧𝐬],\displaystyle A_{n}*\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}=\,2\delta\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}+2\,\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}\,,
Aneˇ[𝐧𝐬𝐧𝐬]= 2(δ+1)eˇ[𝐧𝐬𝐧𝐬]+4eˇ[𝐧𝐬][𝐧𝐬],\displaystyle A_{n}*\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}=\,2(\delta+1)\,\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}+4\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}\,,

Expanding pairs of factors of the projector sequentially starting from the right leads to the additive form of the projector

P4=1+ζmonic monomials in[𝔟(𝒜)]4aζeˇζP_{4}=1+\sum_{\zeta\;\;\text{monic monomials in}\;\;\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{4}}a_{\zeta}\,\check{e}_{\zeta} (7.5)

with

a[𝐧𝐬][𝐩][𝐩]=δ2(δ+4)4(δ2)δ(δ+2)(δ+4),a[𝐧𝐬][𝐩𝐩]=4(δ2)δ(δ+2)(δ+4),a[𝐧𝐬𝐩][𝐩]=δ+3(δ2)(δ+2)(δ+4),a[𝐧𝐬𝐩𝐩]=1(δ2)(δ+2)(δ+4),a[𝐧𝐩𝐬𝐩]=2(δ2)δ(δ+4),a[𝐧𝐬][𝐧𝐬]=δ(δ+3)+6(δ2)(δ1)(δ+2)(δ+4),a[𝐧𝐬𝐧𝐬]=3δ+2(δ2)(δ1)(δ+2)(δ+4).\begin{array}[]{rclrcl}a_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}&=&-\,\dfrac{\delta^{2}(\delta+4)-4}{(\delta-2)\delta(\delta+2)(\delta+4)}\,,&a_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\mathbf{p}\right]}&=&\dfrac{4}{(\delta-2)\delta(\delta+2)(\delta+4)}\,,\\ a_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}&=&\dfrac{\delta+3}{(\delta-2)(\delta+2)(\delta+4)}\,,&a_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\mathbf{p}\right]}&=&-\,\dfrac{1}{(\delta-2)(\delta+2)(\delta+4)}\,,\\ a_{\left[\mathbf{n}\mathbf{p}\mathbf{s}\mathbf{p}\right]}&=&-\,\dfrac{2}{(\delta-2)\delta(\delta+4)}\,,&a_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}&=&\dfrac{\delta(\delta+3)+6}{(\delta-2)(\delta-1)(\delta+2)(\delta+4)}\,,\\ \lx@intercol\hfil a_{\left[\mathbf{n}\mathbf{s}\mathbf{n}\mathbf{s}\right]}=-\,\dfrac{3\delta+2}{(\delta-2)(\delta-1)(\delta+2)(\delta+4)}\,.\hfil\lx@intercol\end{array}

The expressions of the splitting idempotent for n=5n=5 and n=6n=6 can be found in the appendix of [13], and in the Mathematica notebook joined with this article where there is also the expression for n=7n=7.

Acknowledgements

We are grateful to X. Bekaert and N. Boulanger for valuable comments and advices during preparation of the manuscript. Y.G. and D.B. are much indebted to O. V. Ogievetsky for enlightening and instructive discussions concerning diagram algebras and their representation theory. The work of Y.G. is supported by a joint grant “50/50” UMONS – Université François Rabelais de Tours.

Appendix A Examples of traceless projections

Example 7 (arbitrary hook tensors).

For the partition μ=(m,1,,1nm)\mu=(m,\underbrace{1,...,1}_{n-m}) with m2m\geqslant 2 one constructs specμ×(An)\mathrm{spec}^{\times}_{\mu}(A_{n}) by reconstructing the skew shape diagrams μ\λ\mu\backslash\lambda (with |μ\λ|=2f|\mu\backslash\lambda|=2f) using the reverse jeu de taquin. The only starting tableaux are

\ytableausetupmathmode,boxsize=0.9em,centertableaux{ytableau}1&1\none[]1×\none[]××\none[ . . . ]××(with 2f entries ‘1’, for all f=1,,m2),\ytableausetup{mathmode,boxsize=0.9em,centertableaux}\ytableau{\scriptstyle}\scriptstyle{1}&\scriptstyle{1}\none[\scriptstyle{\cdots}]\scriptstyle{1}\times\none[\scriptstyle{\cdots}]\times\\ \times\\ \none[\vbox{ \scriptsize\hbox{.}\hbox{.}\hbox{.}\kern-0.75pt}]\\ \times\\ \times\\ \qquad\text{(with $2f$ entries `1', for all $f=1,\dots,\lfloor\tfrac{m}{2}\rfloor$),}

which leads to the following possibilities:

\ytableausetupmathmode,boxsize=0.9em,centertableauxμ\λ={ytableau}×&×\none[]×\none[]×\none[ . . . ]×(for all f=1,,m2)andμ\λ={ytableau}×&\none[]×\none[]××\none[ . . . ]×(for all f=1,,m12)\ytableausetup{mathmode,boxsize=0.9em,centertableaux}\mu\backslash\lambda=\ytableau{\scriptstyle}\times&\times\none[\scriptstyle{\cdots}]\times\none[\scriptstyle{\cdots}]\\ \times\\ \none[\vbox{ \scriptsize\hbox{.}\hbox{.}\hbox{.}\kern-0.75pt}]\\ \times\\ \\ \quad\text{(for all $f=1,\dots,\lfloor\tfrac{m}{2}\rfloor$)}\qquad\text{and}\qquad\mu\backslash\lambda=\ytableau{\scriptstyle}\times&\none[\scriptstyle{\cdots}]\times\none[\scriptstyle{\cdots}]\\ \times\\ \times\\ \none[\vbox{ \scriptsize\hbox{.}\hbox{.}\hbox{.}\kern-0.75pt}]\\ \times\\ \quad\text{(for all $f=1,\dots,\lfloor\tfrac{m-1}{2}\rfloor$)}
so,spec(μ)×(An)={f(N+2(mf))n}{f(N+2(m1f))}\text{so,}\quad\mathrm{spec}_{(\mu)}^{\times}(A_{n})=\left\{f\,\big{(}N+2\ \left(m-f\right)\big{)}-n\right\}\ \cup\ \left\{f\,\big{(}N+2\ \left(m-1-f\right)\big{)}\right\} (A.1)

Then, the reduced projector (4.16), is given by

Pn(μ)=f=1m2(11f(N+2(mf))nAn)f=1m12(11f(N+2(m1f))An).P_{n}^{(\mu)}=\displaystyle{\prod_{f=1}^{\lfloor\tfrac{m}{2}\rfloor}}\left(1-\frac{1}{f(N+2\ (m-f))-n}\,A_{n}\right)\displaystyle{\prod_{f=1}^{\lfloor\tfrac{m-1}{2}\rfloor}}\left(1-\frac{1}{f(N+2\ (m-1-f))}\,A_{n}\right)\,. (A.2)

Example 8 (The metric-affine conformal Weyl tensor)

In a Riemannian geometry (,g)(\mathcal{M},g) the affine connection \nabla on the tangent bundle of the manifold \mathcal{M} is the Levi-Civita connection associated to the metric gg. The Riemann tensor is defined with respect the Levi-Civita connection which depends on the metric only and as such we refer to it as the metric Riemann tensor. From the symmetry of the metric Riemann tensor on can easily conclude that it is a simple GLGL-module V(2,2)V^{(2,2)}. More in detail, for Rab,cd=gamRmb,cdR_{ab,cd}=g_{am}R^{m}{}_{b,cd} one has

Rab,cd+Rba,cd=Rab,cd+Rab,dc=0,andRab,cd+Rbc,ad+Rca,bd=0(the Bianchi identity).R_{ab,cd}+R_{ba,cd}=R_{ab,cd}+R_{ab,dc}=0\,,\quad\text{and}\quad R_{ab,cd}+R_{bc,ad}+R_{ca,bd}=0\quad\text{(the Bianchi identity)}\,. (A.3)

The trace decomposition of the Riemann tensor corresponds to the restriction of the representation V(2,2)V^{(2,2)} to the (local) orthogonal group. For the spacetime dimensions N4N\geqslant 4, according to the Littlewood restriction rules [25], this leads to the metric Weyl tensor (the totally traceless projection D(2,2)D^{(2,2)}), the traceless Ricci tensor (single-trace projection D(2)D^{(2)}) and the scalar curvature (the double trace D(0)D^{(0)}), recall (3.37). The expression for the metric Weyl tensor in the above decomposition is very well known [52] (see also [53, Chapter II]). For the lower dimensions N=2,3N=2,3 the metric Weyl tensor vanishes identically (as the module D(2,2)D^{(2,2)} does [1], recall Theorem 4.3). Besides, the Littewood-Richardson restriction rule [25] does not apply in this case, and the branching rules are no more expressed in terms of the Littlewood-Richardson coefficients (recall (3.37)). For N=3N=3 both the Ricci tensor and scalar curvature are present (i.e. the Riemann tensor can be expressed in terms of the Ricci tensor and the scalar curvature). For N=2N=2 only the scalar curvature enters the decomposition, despite D(2)D^{(2)} is a non-trivial module in this case (this is due to the fact that locally, any metric in two dimensions is conformally flat).

In a metric-affine geometry (,¯,g)(\mathcal{M},\bar{\nabla},g) [54] the tangent bundle of the manifold \mathcal{M} is endowed with a supplementary affine connection ¯\bar{\nabla}. In full generality this connection does not preserve the metric under parallel transport ¯g0\bar{\nabla}g\neq 0 and has torsion: ¯XY¯YX[X,Y]0\bar{\nabla}_{X}Y-\bar{\nabla}_{Y}X-[X,Y]\neq 0 for two vector fields XX and YY. The non-metric Riemann tensor associated to ¯\bar{\nabla} enjoys the only symmetry:

R¯a,b,cd+R¯a,b,dc=0whereR¯a,b,cd=gamR¯m.b,cd\bar{R}_{a,b,cd}+\bar{R}_{a,b,dc}=0\,\quad\text{where}\quad\bar{R}_{a,b,cd}=g_{am}\bar{R}^{m}{}_{b,cd}\;. (A.4)

Thus, the GLGL-irreducible components are given by the following plethysm, which is computed by applying the Littlewood-Richardson rule (3.28):

\Yboxdim9pt\yng(1)\yng(1)\yng(1,1)=\yng(3,1)+\yng(2,2)+ 2\yng(2,1,1)+\yng(1,1,1,1),where one assumes N4.\Yboxdim{9pt}\yng(1)\,\,{\scriptstyle\otimes}\,\,\yng(1)\,\,{\scriptstyle\otimes}\,\yng(1,1)\,=\yng(3,1)\,+\,\yng(2,2)\,+\,2\;\yng(2,1,1)\,+\,\yng(1,1,1,1)\;,\quad\text{where one assumes $N\geqslant 4$.} (A.5)

Upon restriction to O(N)O(N), the right-hand-side of (A.5) contains a significantly larger variety of components than in the metric case. Along with the lines of the present work, let us focus on the traceless part. It is natural to define the non-metric Weyl tensor as the totally traceless projection (with respect to the metric gg) of the non-metric Riemann tensor:

W¯a,b,cd+W¯a,b,dc=0,gabW¯a,b,cd=gabW¯a,c,bd=gabW¯c,a,bd=0.\bar{W}_{a,b,cd}+\bar{W}_{a,b,dc}=0\,,\quad g^{ab}\bar{W}_{a,b,cd}=g^{ab}\bar{W}_{a,c,bd}=g^{ab}\bar{W}_{c,a,bd}=0\,. (A.6)

Its explicit expression can be obtained by applying the projector 𝔓4\mathfrak{P}_{4} whose step-by-step construction is described in Section 7. For the lower dimensions N=2,3N=2,3 one ignores the components whose traceless part vanishes identically. Namely, when N=2N=2 none of the Young diagrams on the right-hand-side of (A.5) has a non-trivial traceless projection (recall Theorem 4.3), so the non-metric conformal Weyl tensor vanishes identically, which is reminiscent of the metric case. For N=3N=3, the right-hand-side of (A.5) (and thus the non-metric Riemann tensor) contains one GL(3)GL(3)-irreducible component (3,1)(3,1) with a non-zero traceless projection. The latter can be constructed via Corollary 4.4: for the index set ={(3,1),(2,2),(2,12)}\mathcal{R}=\big{\{}(3,1),(2,2),(2,1^{2})\big{\}} one constructs spec(3,1)×(A4)={3,5}\mathrm{spec}^{\times}_{(3,1)}(A_{4})=\{3,5\}, spec(2,2)×(A4)={1,4}\mathrm{spec}^{\times}_{(2,2)}(A_{4})=\{1,4\}, spec(2,12)×(A4)={1}\mathrm{spec}^{\times}_{(2,1^{2})}(A_{4})=\{1\}, so spec×(A4)={1,3,4,5}\mathrm{spec}^{\times}_{\mathcal{R}}(A_{4})=\{1,3,4,5\} and one acts by the following operator on the components R¯a,b,cd\bar{R}_{a,b,cd}:

P4()=(1A4)(113A4)(114A4)(115A4).P^{(\mathcal{R})}_{4}=\big{(}1-A_{4}\big{)}\big{(}1-\tfrac{1}{3}\,A_{4}\big{)}\big{(}1-\tfrac{1}{4}\,A_{4}\big{)}\big{(}1-\tfrac{1}{5}\,A_{4}\big{)}\,. (A.7)

Choosing the semi-additive form of the projector described in Corollary 4.5 will yield the same result in a much more economical way: one uses the operator P4(3,1)z(3,1)P_{4}^{(3,1)}z^{(3,1)}, where

P4(3,1)\displaystyle P_{4}^{(3,1)} =(113A4)(115A4)\displaystyle=\big{(}1-\tfrac{1}{3}\,A_{4}\big{)}\big{(}1-\tfrac{1}{5}\,A_{4}\big{)}
=eˇ[𝐩][𝐩][𝐩][𝐩]+215eˇ[𝐧𝐬][𝐧𝐬]+115eˇ[𝐧𝐬𝐩][𝐩]13eˇ[𝐧𝐬][𝐩][𝐩].\displaystyle=\check{e}_{\left[\mathbf{p}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}+\dfrac{2}{15}\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{n}\mathbf{s}\right]}+\dfrac{1}{15}\,\check{e}_{\left[\mathbf{n}\mathbf{s}\mathbf{p}\right]\left[\mathbf{p}\right]}-\dfrac{1}{3}\,\check{e}_{\left[\mathbf{n}\mathbf{s}\right]\left[\mathbf{p}\right]\left[\mathbf{p}\right]}.

In [55, Section 4] the authors choose to define the non-metric Weyl tensor by utilising the non-metric Riemann tensor (as well as the corresponding symmetric Ricci tensor and scalar curvature) in the standard expression of the metric Weyl tensor. The resulting tensor is however not traceless. To justify this definition it is asserted that no modification of the expression of the metric Weyl tensor satisfying the properties (A.6) exists. However, our analysis suggests that the sought modification does exist and is obtained via application of the traceless projector 𝔓4()\mathfrak{P}^{(\mathcal{R})}_{4} to the non-metric Riemann tensor. In particular, the non-metric Weyl tensor (A.6) is non-zero whenever N3N\geqslant 3.

Appendix B Proofs

B.1 Proof of Lemma 3.1 continued

Proof.

Let us consider the case of an antisymmetric metric on VV. With the definition of an adjoint operator (3.23) at hand, one can check that 𝔯(dij)=𝔯(dij)\mathfrak{r}(d_{ij})^{*}=\mathfrak{r}(d_{ij}). To analyse the diagonal structure of the latter we make use of the canonical isomorphism VUUV\cong U^{*}\oplus U, such that φ+u,ψ+v=φ(v)ψ(u)\langle\varphi+u,\psi+v\rangle=\varphi(v)-\psi(u). This induces block structure on the matrices of ,\langle\,\cdot,\cdot\,\rangle and any FEnd(V)F\in\mathrm{End}(V):

,(0𝟙𝟙0),F(ABCD)F(DtBtCtAt),\langle\,\cdot,\cdot\,\rangle\sim\left(\begin{array}[]{ccc}0&\vline&\mathbb{1}\\ \hline\cr-\mathbb{1}&\vline&0\end{array}\right)\,,\quad F\sim\left(\begin{array}[]{ccc}A&\vline&B\\ \hline\cr C&\vline&D\end{array}\right)\quad\quad\Rightarrow\quad\quad F^{*}\sim\left(\begin{array}[]{ccc}D^{t}&\vline&-B^{t}\\ \hline\cr-C^{t}&\vline&A^{t}\end{array}\right)\,, (B.1)

where ()t(\cdot)^{t} denotes the transpose of a matrix. The same structure holds when one considers the metric and operators on VnV^{\otimes n}, with the only difference that depending on nn the metric is either symmetric or anti-symmetric: for any S,TVnS,T\in V^{\otimes n} one has S,T=()nT,S\langle S,T\rangle=(-)^{n}\langle T,S\rangle. Set r=n(mod 2)r=n\,(\mathrm{mod}\,2) and consider the space

𝒯n(V)=VnVn2Vr(by definition, V0=),\mathcal{T}_{n}(V)=V^{\otimes n}\oplus V^{\otimes n-2}\oplus\dots\oplus V^{\otimes r}\quad\text{(by definition, $V^{\otimes 0}=\mathbb{C}$),} (B.2)

The generalisation of the above block structure for ,\langle\,\cdot,\cdot\,\rangle and FEnd(𝒯n(V))F\in\mathrm{End}\big{(}\mathcal{T}_{n}(V)\big{)} reads

,(0𝟙()r𝟙0),F(ABCD)F(Dt()rBt()rCtAt)\langle\,\cdot,\cdot\,\rangle\sim\left(\begin{array}[]{ccc}0&\vline&\mathbb{1}\\ \hline\cr(-)^{r}\mathbb{1}&\vline&0\end{array}\right)\,,\quad F\sim\left(\begin{array}[]{ccc}A&\vline&B\\ \hline\cr C&\vline&D\end{array}\right)\quad\quad\Rightarrow\quad\quad F^{*}\sim\left(\begin{array}[]{ccc}D^{t}&\vline&(-)^{r}B^{t}\\ \hline\cr(-)^{r}C^{t}&\vline&A^{t}\end{array}\right) (B.3)

Fix a pair i<ji<j. The space 𝒯n(V)\mathcal{T}_{n}(V) is preserved by trij(g)\mathrm{tr}^{(g)}_{ij}, so in the sequel we identify trij(g)\mathrm{tr}^{(g)}_{ij} with its restriction to 𝒯n(V)\mathcal{T}_{n}(V). This allows us to have trij(g)\mathrm{tr}^{(g)*}_{ij} on 𝒯n(V)\mathcal{T}_{n}(V), such that 𝔯(dij)=trij(g)trij(g)\mathfrak{r}(d_{ij})=-\mathrm{tr}^{(g)*}_{ij}\mathrm{tr}^{(g)}_{ij}. Upon a suitable reordering of the basis elements in 𝒯n(V)\mathcal{T}_{n}(V) one has the following block structure of trij(g)\mathrm{tr}^{(g)}_{ij}:

trij(g)(0B()rB0)trij(g)(0()rBtBt0),trij(g)trij(g)(BtB00BtB),\mathrm{tr}^{(g)}_{ij}\sim\left(\begin{array}[]{ccc}0&\vline&B\\ \hline\cr(-)^{r}B&\vline&0\end{array}\right)\quad\quad\Rightarrow\quad\quad\mathrm{tr}^{(g)*}_{ij}\sim\left(\begin{array}[]{ccc}0&\vline&(-)^{r}B^{t}\\ \hline\cr B^{t}&\vline&0\end{array}\right)\,,\quad\quad\mathrm{tr}^{(g)*}_{ij}\mathrm{tr}^{(g)}_{ij}\sim\left(\begin{array}[]{ccc}B^{t}B&\vline&0\\ \hline\cr 0&\vline&B^{t}B\end{array}\right)\,, (B.4)

where BtBB^{t}B is block-diagonal with respect to the decomposition of 𝒯n(V)\mathcal{T}_{n}(V) into direct summands. The transformation BtBB^{t}B is diagonalisable by an orthogonal matrix SS (i.e. such that St=S1S^{t}=S^{-1}). Therefore, BtBB^{t}B is also diagonalisable by the contragredient transformation (S1)t=S(S^{-1})^{t}=S. As a result, both non-zero blocks of 𝔯(dij)=trij(g)trij(g)\mathfrak{r}(d_{ij})=-\mathrm{tr}^{(g)*}_{ij}\mathrm{tr}^{(g)}_{ij} (one being a transformation of UU, and the other of UU^{*}) are brought to the same diagonal form, all the eigenvalues being non-positive. The same conclusion is easily extended to the action of AnA_{n} on VnV^{\otimes n}. ∎

B.2 Alternative proof of Lemma 3.1 in the semisimple regime of Bn(εN)B_{n}(\varepsilon N)

Proof.

Recall that the semisimple regime for Bn(εN)B_{n}(\varepsilon N) occurs for Nn1N\geqslant n-1 when ε=1\varepsilon=1 and N2(n1)N\geqslant 2(n-1) when ε=1\varepsilon=-1. Let us analyse the set in (3.43) without intersecting it with ε\varepsilon\mathbb{N} and show that it nevertheless coincides with (3.43). Denote αμ\λ\alpha_{\mu\backslash\lambda} the eigenvalue from Lemma 3.3 which corresponds to L(μ)Mn(λ)L^{(\mu)}\subset M^{(\lambda)}_{n}.

For ε=1\varepsilon=1 let us concentrate on the minimal possible value of the content c(μ\λ)c(\mu\backslash\lambda). First, note that because λ1|λ|=n2f\lambda^{\prime}_{1}\leqslant|\lambda|=n-2f, in the semisimple regime λ1+f(n2f)+fN\lambda^{\prime}_{1}+f\leqslant(n-2f)+f\leqslant N when f1f\geqslant 1. So the minimal possible content among skew-shapes μ\λclN(f)(λ)\mu\backslash\lambda\in\mathrm{cl}^{(f)}_{N}(\lambda) is achieved when μ\mu is obtained via adding two columns of ff boxes to the first and the second columns of λ\lambda. In this case c(μ\λ)=(λ1+λ2+f2)fc(\mu\backslash\lambda)=-(\lambda^{\prime}_{1}+\lambda^{\prime}_{2}+f-2)f, and the corresponding eigenvalue αμ\λ=(N+1fλ1λ2)f\alpha_{\mu\backslash\lambda}=(N+1-f-\lambda^{\prime}_{1}-\lambda^{\prime}_{2})f. Assuming that αμ\λ0\alpha_{\mu\backslash\lambda}\leqslant 0 one gets λ1+λ2+2fN+1+f\lambda^{\prime}_{1}+\lambda^{\prime}_{2}+2f\geqslant N+1+f, and on the other hand λ1+λ2+2fn\lambda^{\prime}_{1}+\lambda^{\prime}_{2}+2f\leqslant n. By combining the two estimates one gets Nn1fn2N\leqslant n-1-f\leqslant n-2, which is in contradiction with the semisimplicity of Bn(N)B_{n}(N).

For ε=1\varepsilon=-1 we look for the maximal possible value of the content c(μ\λ)c(\mu\backslash\lambda). Note that λ1|λ|=n2f\lambda_{1}\leqslant|\lambda|=n-2f, so in the semisimple regime λ1+2fN2+1N\lambda_{1}+2f\leqslant\frac{N}{2}+1\leqslant N when f1f\geqslant 1 and N2N\in 2\mathbb{N}. So the maximal possible content among skew-shapes μ\λclN(f)(λ)\mu\backslash\lambda\in\mathrm{cl}^{(f)}_{N}(\lambda) is achieved when μ\mu is obtained by adding a row of 2f2f boxes to the first row of λ\lambda. In this case c(μ\λ)=(2λ1+2f1)fc(\mu\backslash\lambda)=(2\lambda_{1}+2f-1)f, and the corresponding eigenvalue αμ\λ=(N+1fλ1λ2)f\alpha_{\mu\backslash\lambda}=(N+1-f-\lambda^{\prime}_{1}-\lambda^{\prime}_{2})f. Assuming that αμ\λ0\alpha_{\mu\backslash\lambda}\geqslant 0 one gets N2λ1+f1\frac{N}{2}\leqslant\lambda_{1}+f-1, and on the other hand λ1+2fn\lambda_{1}+2f\leqslant n. By combining the two estimates one gets N2n1fn2\frac{N}{2}\leqslant n-1-f\leqslant n-2, which is in contradiction with the semisimplicity of Bn(N)B_{n}(-N).

To this end, we conclude that the only possibility to have αμ\λ=0\alpha_{\mu\backslash\lambda}=0 occurs for f=0f=0, so λ=μ\lambda=\mu. This corresponds to D(λ^)Mn(λ)D^{(\hat{\lambda})}\otimes M^{(\lambda)}_{n} in the decomposition (2.12) with |λ|=n|\lambda|=n, which are annihilated by all dijd_{ij}. So KerAn\mathrm{Ker}\,A_{n} is indeed the subspace of traceless tensors. ∎

B.3 Proof of Corollary 3.5

According to the conditions in the assertion, Schur-Weyl duality applies, so for L(μ)Mn(λ)L^{(\mu)}\subset M^{(\lambda)}_{n} the first part of the assertion holds by Lemmas 3.1 and 3.3. For the second part consider L(μ)Δn(λ)L^{(\mu)}\subset\Delta^{(\lambda)}_{n} upon restriction to 𝔖n\mathbb{C}\mathfrak{S}_{n}, such that

αμ\λ=(δ1)|μ||λ|2+c(μ\λ).\alpha_{\mu\backslash\lambda}=(\delta-1)\,\frac{|\mu|-|\lambda|}{2}+c(\mu\backslash\lambda)\,. (B.5)

The above expression can be interpreted as AnL(μ)=0A_{n}L^{(\mu)}=0 along the following lines. Choose a basis in Δn(λ)\Delta^{(\lambda)}_{n} such that any element of Bn(δ)B_{n}(\delta) is represented by a block-upper-triangular matrix in End(Δn(λ))\mathrm{End}\big{(}\Delta^{(\lambda)}_{n}\big{)}, with the composition factors Mn(σi)M^{(\sigma_{i})}_{n} (with 1ir1\leqslant i\leqslant r) appearing on the diagonal (see [3, paragraph below Corollary 2.3]). There is exactly one matrix block with σ1=λ\sigma_{1}=\lambda, while for the others (when i1i\neq 1) holds |σi|>|λ||\sigma_{i}|>|\lambda|. By [3, Proposition 4.2], one has

(δ1)|σi||λ|2+c(σi\λ)=0.(\delta-1)\,\frac{|\sigma_{i}|-|\lambda|}{2}+c(\sigma_{i}\backslash\lambda)=0\,. (B.6)

The embedding L(μ)Δn(λ)L^{(\mu)}\subset\Delta^{(\lambda)}_{n} is realised via combinations of vectors from modules L(μ)Mn(σip)L^{(\mu)}\subset M^{(\sigma_{i_{p}})}_{n} (for a subset of indicies 1i1<<ir1\leqslant i_{1}<\dots<i_{\ell}\leqslant r). By Lemma (3.3), the eigenvalue of AnA_{n} on each embedding is

(δ1)|μ||σi|2+c(μ\σi)=(δ1)|μ||λ|2+c(μ\λ)=αμ\λ(\delta-1)\frac{|\mu|-|\sigma_{i}|}{2}+c(\mu\backslash\sigma_{i})=(\delta-1)\frac{|\mu|-|\lambda|}{2}+c(\mu\backslash\lambda)=\alpha_{\mu\backslash\lambda} (B.7)

(where we have used (B.6)). Therefore AnA_{n} is proportional to identity on any embedding L(μ)Δn(λ)L^{(\mu)}\subset\Delta^{(\lambda)}_{n} with the same eigenvalue αμ\λ\alpha_{\mu\backslash\lambda} (B.5). Due to Lemma 3.1, L(μ)L^{(\mu)} can not appear in Mn(λ)M^{(\lambda)}_{n} if αμ\λ<0\alpha_{\mu\backslash\lambda}<0 or αμ\λ=0\alpha_{\mu\backslash\lambda}=0 with λμ\lambda\neq\mu, and thus any its occurrence in Δn(λ)\Delta^{(\lambda)}_{n} is factored out due to M(λ)Δn(λ)/KM^{(\lambda)}\cong\Delta^{(\lambda)}_{n}/K.

B.4 Proof of Lemma 5.3

Proof.

The diversity of expressions on the left-hand-sides of the rules (5.14)-(5.19) which are necessary to fix τ\tau, is essentially limited by 𝔟(𝒜)\mathfrak{b}(\mathcal{A})-linearity. Representatives of a particular form on the left-hand-sides of the rules (5.14), (5.15) are always accessible via cyclic permutations. Taking into account the inversion, the same conclusion is valid for the rules in (5.16), (5.19) due to the following simple facts.

Fact 1.

For [𝐩˙u]=[𝐩˙I(u)][\dot{\mathbf{p}}u]=[\dot{\mathbf{p}}I(u)] either uu or I(u)I(u) is fit because [𝐩u]𝔟(𝒜)[\mathbf{p}u]\in\mathfrak{b}(\mathcal{A}).

Fact 2.

For [𝐩˙u𝐬˙v]=[𝐩˙I(v)𝐬˙I(u)][\dot{\mathbf{p}}u\dot{\mathbf{s}}v]=[\dot{\mathbf{p}}I(v)\dot{\mathbf{s}}I(u)] either uu or I(v)I(v) is fit. Indeed, since [𝐩u𝐬v]𝔟(𝒜)[\mathbf{p}u\mathbf{s}v]\in\mathfrak{b}(\mathcal{A}), either one of the bracelets [u][u], [v][v] is in 𝔟(𝒜)\mathfrak{b}(\mathcal{A}) or one of them is empty.

Fact 3.

For [𝐩˙u𝐬˙v]=[𝐩˙I(v)𝐬˙I(u)][\dot{\mathbf{p}}u\dot{\mathbf{s}}v]=[\dot{\mathbf{p}}I(v)\dot{\mathbf{s}}I(u)] either [u],[v]𝔟(𝒜)[u],[v]\notin\mathfrak{b}(\mathcal{A}) or [u],[v]𝔟(𝒜)[u],[v]\in\mathfrak{b}(\mathcal{A}). In the former case either |u|𝐧>|u|𝐬|u|_{\mathbf{n}}>|u|_{\mathbf{s}} or |v|𝐧>|v|𝐬|v|_{\mathbf{n}}>|v|_{\mathbf{s}}, while in the latter u,vu,v or I(u),I(v)I(u),I(v) are both fit. Note that the admissible representatives on the left-hand-sides of the rules (5.16) are fixed unambiguously.

Representatives on the left-hand-sides of (5.14)-(5.19) are sufficient to define τ\tau. Let us check that the definition is independent of a particular choice among admissible representatives.

  • 1)

    For the rule (5.14), the only alternative representative is [𝐬¨I(u)][\ddot{\mathbf{s}}I(u)]. Application of (5.14) leads to τ([𝐬¨I(u)])=2δ[𝐬I(u)]\tau\big{(}[\ddot{\mathbf{s}}I(u)]\big{)}=2\delta\,[\mathbf{s}I(u)]. But the latter equals to τ([𝐬¨u])\tau\big{(}[\ddot{\mathbf{s}}u]\big{)} by inversion of the representative combined with a cyclic permutation.

  • 2)

    For the first rule in (5.15), the equivalent representatives are [𝐬˙v𝐬˙u]=[𝐬˙I(u)𝐬˙I(v)]=[𝐬˙I(v)𝐬˙I(u)][\dot{\mathbf{s}}v\dot{\mathbf{s}}u]=[\dot{\mathbf{s}}I(u)\dot{\mathbf{s}}I(v)]=[\dot{\mathbf{s}}I(v)\dot{\mathbf{s}}I(u)]. By direct application of (5.15) one calculates τ([𝐬˙v𝐬˙u])=2([𝐬˙v𝐬˙I(u)]+[𝐬˙v][𝐬˙u])\tau\big{(}[\dot{\mathbf{s}}v\dot{\mathbf{s}}u]\big{)}=2\big{(}[\dot{\mathbf{s}}v\dot{\mathbf{s}}I(u)]+[\dot{\mathbf{s}}v][\dot{\mathbf{s}}u]\big{)}. But [𝐬˙v𝐬˙I(u)]=[𝐬˙u𝐬˙I(v)][\dot{\mathbf{s}}v\dot{\mathbf{s}}I(u)]=[\dot{\mathbf{s}}u\dot{\mathbf{s}}I(v)] (the two representatives are related by inversion followed by a cyclic permutation), which reproduces τ([𝐬˙u𝐬˙v])\tau\big{(}[\dot{\mathbf{s}}u\dot{\mathbf{s}}v]\big{)}. Other alternatives are analysed along the same lines.

    For the second rule in (5.15), one has to analyse [𝐬˙v][𝐬˙u][\dot{\mathbf{s}}v][\dot{\mathbf{s}}u], together with the alternatives [𝐬˙I(u)][\dot{\mathbf{s}}I(u)] and [𝐬˙I(v)][\dot{\mathbf{s}}I(v)] for each factor. By direct application of (5.15) one gets τ([𝐬˙v][𝐬˙u])=2([𝐬˙v𝐬˙u]+[𝐬˙v𝐬˙I(u)])\tau\big{(}[\dot{\mathbf{s}}v][\dot{\mathbf{s}}u]\big{)}=2\big{(}[\dot{\mathbf{s}}v\dot{\mathbf{s}}u]+[\dot{\mathbf{s}}v\dot{\mathbf{s}}I(u)]\big{)}. But [𝐬˙v𝐬˙u]=[𝐬˙u𝐬˙v][\dot{\mathbf{s}}v\dot{\mathbf{s}}u]=[\dot{\mathbf{s}}u\dot{\mathbf{s}}v] (by a cyclic permutation) and [𝐬˙v𝐬˙I(u)]=[𝐬˙u𝐬˙I(v)][\dot{\mathbf{s}}v\dot{\mathbf{s}}I(u)]=[\dot{\mathbf{s}}u\dot{\mathbf{s}}I(v)] (by composition of inversion and a cyclic permutation), which leads to τ([𝐬˙u][𝐬˙v])\tau\big{(}[\dot{\mathbf{s}}u][\dot{\mathbf{s}}v]\big{)}. Consider also τ([𝐬˙u][𝐬˙I(v)])=2([𝐬˙u𝐬˙I(v)]+[𝐬˙u𝐬˙v])\tau\big{(}[\dot{\mathbf{s}}u][\dot{\mathbf{s}}I(v)]\big{)}=2\big{(}[\dot{\mathbf{s}}u\dot{\mathbf{s}}I(v)]+[\dot{\mathbf{s}}u\dot{\mathbf{s}}v]\big{)} which equals τ([𝐬˙u][𝐬˙v])\tau\big{(}[\dot{\mathbf{s}}u][\dot{\mathbf{s}}v]\big{)} directly. Other alternatives are analysed along the same lines.

  • 3)

    For the first rule in (5.19), let u,vu,v are fit. Then for the alternative representative one has τ([𝐩˙v𝐩˙u])=[𝐧v𝐬I(u)]=[𝐧u𝐬I(v)]\tau\big{(}[\dot{\mathbf{p}}v\dot{\mathbf{p}}u]\big{)}=[\mathbf{n}v\mathbf{s}I(u)]=[\mathbf{n}u\mathbf{s}I(v)]. Else, let |u|𝐬>|u|𝐧|u|_{\mathbf{s}}>|u|_{\mathbf{n}}. Then one checks τ([𝐩˙I(u)𝐩˙I(v)])=[𝐧I(u)][𝐬I(v)]=[𝐧u][𝐬v]\tau\big{(}[\dot{\mathbf{p}}I(u)\dot{\mathbf{p}}I(v)]\big{)}=[\mathbf{n}I(u)][\mathbf{s}I(v)]=[\mathbf{n}u][\mathbf{s}v].

    For the second rule in (5.19) one considers τ([𝐩˙v][𝐩˙u])=[𝐧v𝐬I(u)]=[𝐧u𝐬I(v)]\tau\big{(}[\dot{\mathbf{p}}v][\dot{\mathbf{p}}u]\big{)}=[\mathbf{n}v\mathbf{s}I(u)]=[\mathbf{n}u\mathbf{s}I(v)].

B.5 Proof of Theorem 5.4

Proof.

The main idea consists in analysing the following formula which follows from the definition of the average (3.16): for any diagram bBn(δ)b\in B_{n}(\delta)

Anγb=γ(Anb).A_{n}\gamma_{b}=\gamma_{(A_{n}b)}\,. (B.8)

The product Anb=1i<jndijbA_{n}b=\sum_{1\leqslant i<j\leqslant n}d_{ij}b consists in summing over all possibilities of placing an arc at a pair of lower nodes of bb, which is mimicked by imposing Leibniz rule in the definition of \partial. Without loss of generality, we assume a convenient representative bb in the conjugacy class, such that all bracelets in Φ(γb)\Phi(\gamma_{b}) come from mutually non-intersecting cycles, and such that each independent cycle is either a cycle permutation, or, if bJ(1)b\in J^{(1)}, there are no intersections among the vertical lines, while each arc in the upper or lower row is incident to a pair of nodes (i,i+1)(i,i+1) (for in1i\leqslant n-1) or (1,n)(1,n) (for example, see (B.21)). The independent cycles will be schematically illustrated by rectangular blocks, with only particular significant lines specified explicitly. For example,

[Uncaptioned image][Uncaptioned image],[Uncaptioned image][Uncaptioned image],\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{csym.pdf}}\rightarrow\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{theo3_s_2.pdf}}\;,\hskip 56.9055pt\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{csym1.pdf}}\rightarrow\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{theo3_s_1.pdf}}\;, (B.9)
[Uncaptioned image][Uncaptioned image],[Uncaptioned image][Uncaptioned image]\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{theo3_diag_1.pdf}}\rightarrow\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{theo3_diag_1_rep.pdf}}\;,\hskip 28.45274pt\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{theo3_diag_2.pdf}}\rightarrow\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{theo3_diag_2_rep.pdf}} (B.10)

There are two major cases how one can place the arc upon multiplication by dijd_{ij}: either it joins two nodes within a single cycle, or the two nodes belong to two different cycles. Only the bracelets coming from these blocks will be affected by multiplication by dijd_{ij}, so all other blocks can be ignored while analysing each particular i<ji<j. This property is manifested in the definition of τ\tau as a [𝔟(𝒜)]\mathbb{C}[\mathfrak{b}(\mathcal{A})]-linear operation. We will work in terms of particular representatives in the bracelets, and read off the resulting representatives coming form the initial ones. To fix a representative, we will highlight the starting point by a circle around a node, such that one starts reading along the adjacent line (if there is a diagram attached below, one ignores it). We always imply the word uu read off by following the lines hidden behind the first block, while for the second block (if any) the word is vv. Additional arrows entering each block fix the direction of reading which leads to the corresponding word. Passing the block in the opposite direction leads to I(u)I(u) instead of uu and I(v)I(v) instead of vv. For example, from the following schema one reads off the representative [𝐧u𝐬I(v)][\mathbf{n}u\mathbf{s}I(v)]:

[Uncaptioned image] (B.11)

Each letter in a word parametrising a bracelet corresponds to a line with two endpoints. Upon multiplication dijbd_{ij}b, put 0, 11 or 22 dots above a letter in each representative depending on how many endpoints are occupied by the (i,j)(i,j)-arc attached to the lower row of bb. It is clear that 𝐧\mathbf{n} can never acquire a dot, 𝐩\mathbf{p} can acquire at most 11 dot, while 𝐬\mathbf{s} can acquire up to 22 dots. This is exactly encoded in the definition of \partial by putting (𝐧)=(𝐩˙)=(𝐬¨)=0\partial(\mathbf{n})=\partial(\dot{\mathbf{p}})=\partial(\ddot{\mathbf{s}})=0. Without loss of generality, one can consider the following schemas for the possibilities of attaching the arc.

  • 1)

    The arc is attached to another arc, which leads to a cycle:

    τn([𝐬¨u])=[Uncaptioned image]=δ[Uncaptioned image]=δ[𝐬u],\displaystyle\tau_{n}\big{(}\left[\ddot{\mathbf{s}}u\right]\big{)}\;\;=\;\;\hskip 2.84544pt\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{theo3_a_p1.pdf}}\;\;=\;\;\delta\;\raisebox{-0.4pt}{\includegraphics[scale={0.4}]{theo3_a_1.pdf}}\;\;=\;\;\delta\,[\mathbf{s}u]\,, (B.12)

    so one arrives at the rule (5.14).

  • 2)

    The arc is attached to one of the endpoints of two particular arcs of bb. One sums over the four possibilities in this case:

    τ([𝐬˙u𝐬˙v])=[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]= 2[Uncaptioned image]+ 2[Uncaptioned image]= 2[𝐬u𝐬I(v)]+2[𝐬u][𝐬v],\begin{array}[]{ll}\tau\big{(}\left[\dot{\mathbf{s}}u\dot{\mathbf{s}}v\right]\big{)}&=\;\;\hskip 2.84544pt\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_b_p1.pdf}}\;+\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_b_p2.pdf}}\;+\\ \hfill&\,\hskip 17.07182pt\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_b_p3.pdf}}\;+\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_b_p4.pdf}}\\ \lx@intercol=\;2\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_b_r2.pdf}}\;+\;2\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_b_r1.pdf}}\;=\;2\,[\mathbf{s}u\mathbf{s}I(v)]+2\,[\mathbf{s}u][\mathbf{s}v]\,,\hfil\lx@intercol\end{array} (B.13)

    so one reproduces the first rule in (5.15).

  • 3)

    Next, let us consider the case when one endpoint of the arc is attached to a passing line, while the other one occupies one of the endpoints of a particular arc in bb. Then one sums over the two possibilities, where the structure of the representatives assumes that uu is fit

    τ([𝐩˙u𝐬˙v])=[Uncaptioned image]+[Uncaptioned image]=[Uncaptioned image]+[Uncaptioned image]=[𝐩u𝐬I(v)]+[𝐩u][𝐬v],\begin{array}[]{ll}\tau\big{(}\left[\dot{\mathbf{p}}u\dot{\mathbf{s}}v\right]\big{)}&=\hskip 2.84544pt\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_d_p1.pdf}}\;+\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_d_p2.pdf}}\\ \lx@intercol=\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_e_r2.pdf}}\;+\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_d_r2.pdf}}=[\mathbf{p}u\mathbf{s}I(v)]\;+\;[\mathbf{p}u][\mathbf{s}v]\,,\hfil\lx@intercol\end{array} (B.14)

    so one recovers the first rule in (5.16).

  • 4)

    Finally, the arc can occupy the lower endpoints of two passing lines. First, suppose that uu is fit (so is vv), which leads to

    τ([𝐩˙u𝐩˙v])\displaystyle\tau\big{(}\left[\dot{\mathbf{p}}u\dot{\mathbf{p}}v\right]\big{)} =[Uncaptioned image]=[Uncaptioned image]=[𝐧u𝐬I(v)],\displaystyle=\hskip 2.84544pt\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_f_p1fit.pdf}}\;=\;\hskip 2.84544pt\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_f_r1fit.pdf}}\;=\;[\mathbf{n}u\mathbf{s}I(v)]\,, (B.15)

    in agreement with the first rule in (5.19) for the case of a fit representative. In the other case, when neither uu nor vv is not fit, with |u|𝐬>|u|𝐧|u|_{\mathbf{s}}>|u|_{\mathbf{n}}, one has

    τ([𝐩˙u𝐩˙v])\displaystyle\tau\big{(}\left[\dot{\mathbf{p}}u\dot{\mathbf{p}}v\right]\big{)} =[Uncaptioned image]=[Uncaptioned image]=[𝐧u][𝐬v],\displaystyle=\;\hskip 2.84544pt\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_f_p1.pdf}}\;=\;\hskip 2.84544pt\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_f_r1.pdf}}\;=\;[\mathbf{n}u][\mathbf{s}v]\,, (B.16)

    which is again in agreement with the first rule in (5.19).

We are left with the cases when the arc connects two independent cycles.

  • 5)

    If the arc occupies the endpoints of two particular arcs in bb, one sums over four possibilities

    τ([𝐬˙u][𝐬˙v])=[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]= 2[Uncaptioned image]+ 2[Uncaptioned image]= 2[𝐬u𝐬v]+ 2[𝐬u𝐬I(v)],\begin{array}[]{ll}\tau\big{(}\left[\dot{\mathbf{s}}u\right]\left[\dot{\mathbf{s}}v\right]\big{)}=&\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_c_p3.pdf}}\;+\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_c_p4.pdf}}\\ \hfill&+\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_c_p1.pdf}}\;+\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_c_p2.pdf}}\\ \lx@intercol=\hskip 2.84544pt\,2\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_c_r2.pdf}}\,+\,2\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_c_r1.pdf}}\;=\;2\,[\mathbf{s}u\mathbf{s}v]\;+\;2\,[\mathbf{s}u\mathbf{s}I(v)]\,,\hfil\lx@intercol\end{array} (B.17)

    which reproduces the second rule in (5.15).

  • 6)

    If the arc occupies the passing line in one cycle and an endpoint of an arc in the other cycle of bb, one sums over two possibilities (uu is fit)

    τ([𝐩˙u][𝐬˙v])=[Uncaptioned image]+[Uncaptioned image]=[Uncaptioned image]+[Uncaptioned image]=[𝐩u𝐬v]+[𝐩u𝐬I(v)],\begin{array}[]{ll}\tau\big{(}\left[\dot{\mathbf{p}}u\right]\left[\dot{\mathbf{s}}v\right]\big{)}=&\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_e_p1.pdf}}\;+\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_e_p2.pdf}}\\ \lx@intercol=\hskip 2.84544pt\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_e_r1.pdf}}\,+\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_e_r2.pdf}}\;=\;[\mathbf{p}u\mathbf{s}v]\;+\;[\mathbf{p}u\mathbf{s}I(v)]\,,\hfil\lx@intercol\end{array} (B.18)

    in agreement with the second rule in (5.16).

  • 7)

    Finally, if the arc occupies two passing lines in two independent cycles of bb (with both u,vu,v fit) one has

    τ([𝐩˙u][𝐩˙v])\displaystyle\tau\big{(}\left[\dot{\mathbf{p}}u\right]\left[\dot{\mathbf{p}}v\right]\big{)} =[Uncaptioned image]=[Uncaptioned image]=[𝐧u𝐬I(v)],\displaystyle=\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_g_p1fit.pdf}}\;=\;\hskip 2.84544pt\,\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{theo3_f_r1fit.pdf}}\;=\;[\mathbf{n}u\mathbf{s}I(v)]\,, (B.19)

    which coincides with the second rule in (5.19).

Due to Lemma 5.3, the considered cases are sufficient to prove the assertion. ∎

B.6 Proof of Lemma 5.5

Proof.

Let us first prove the assertion for bBn(δ)b\in B_{n}(\delta) such that ζ=Φ(γb)\zeta=\Phi(\gamma_{b}) is a single bracelet. Let us show that upon a convenient choice of bb among the conjugate diagrams, any tC𝔖n(b)t\in C_{\mathfrak{S}_{n}}(b) can be written as t=cmrpt=c^{m}r^{p} for some m=0,,n1m=0,\dots,n-1 and p=0,1p=0,1, where

c=[Uncaptioned image],r=[Uncaptioned image]c\;=\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{csym.pdf}}\;,\hskip 85.35826ptr\;=\;\raisebox{-0.4pt}{\includegraphics[scale={0.6}]{rsym.pdf}} (B.20)

(the fact that c=rc=r for n=2n=2 does not lead to any problem). Let us say that two lines in a diagram are adjacent if there exists a pair of vertically aligned nodes which are endpoints of these lines, i.e. these endpoints are identified by the step 2 in the definition of the map Φ\Phi in Section 5.1. For any s𝔖ns\in\mathfrak{S}_{n}, conjugation bsbs1b\to sbs^{-1} preserves the adjacency relation, which is exactly the property encoded by the bracelet Φ(γb)\Phi(\gamma_{b}) (by placing letters along an unoriented circle, one exactly defines the nearest neighbours). If a conjugation transformation preserves the diagram (with tC𝔖n(b)t\in C_{\mathfrak{S}_{n}}(b)), then the result of such transformation can be described as follows: each passing line takes the place of another passing line and each upper (respectively, lower) arc takes the place of another upper (respectively, lower) arc. Without loss of generality, take the convenient diagram in the conjugacy class: b=cb=c when bJ(1)b\notin J^{(1)}, while for bJ(1)b\in J^{(1)} take

b=[Uncaptioned image].b=\hskip 2.84544pt\raisebox{-0.4pt}{\includegraphics[scale={0.7}]{sym.pdf}}\;. (B.21)

In view of preservation of adjacency of lines, the particularly chosen diagram bb can by preserved by adjoint transformations composed only by cyclic permutations of nodes and inversion.

With this at hand, let us establish the bijection between the centraliser C𝔖n(b)C_{\mathfrak{S}_{n}}(b) and the turnover stabiliser S(wb)S(w_{b}) for a particular representative wbζw_{b}\in\zeta. Namely, to read off the word wbw_{b} (as described in steps 1, 2 in the definition of the map Φ\Phi), start at the upper right node and follow the adjacent edge. With a slight abuse of notation, define the action of permutations cc and rr on words of the length nn: c(𝐚1𝐚n)=𝐚n𝐚1𝐚n1c(\mathbf{a}_{1}\dots\mathbf{a}_{n})=\mathbf{a}_{n}\mathbf{a}_{1}\dots\mathbf{a}_{n-1} and r(𝐚1𝐚n)=cI(𝐚1𝐚n)=𝐚1𝐚n𝐚2r(\mathbf{a}_{1}\dots\mathbf{a}_{n})=cI(\mathbf{a}_{1}\dots\mathbf{a}_{n})=\mathbf{a}_{1}\mathbf{a}_{n}\dots\mathbf{a}_{2}. Then it is straightforward that cmrpC𝔖n(b)c^{m}r^{p}\in C_{\mathfrak{S}_{n}}(b) iff cmrpS(wb)c^{m}r^{p}\in S(w_{b}), which proves the assertion. ∎

B.7 Proof of Lemma 5.6

Proof.

One checks directly that for n5n\leqslant 5 any ξ[𝔟(𝒜)]n\xi\in\mathbb{C}[\mathfrak{b}(\mathcal{A})]_{n} verifies ξ=ξ\xi^{*}=\xi. As a result, any uCn(δ)u\in C_{n}(\delta) verifies u=uu^{*}=u, so Cn(δ)Bn(δ)C_{n}(\delta)\subset B_{n}(\delta) is commutative, and so is any Cn(δ)/J(f)C_{n}(\delta)/J^{(f)} and Cn(δ)J(f)C_{n}(\delta)\cap J^{(f)}. To show that Cn(δ)C_{n}(\delta) is non-commutative for any n6n\geqslant 6 take ζn=[𝐧𝐬𝐧𝐩𝐬𝐩][𝐩]n6\zeta_{n}=[\mathbf{nsnpsp}][\mathbf{p}]^{n-6} such that ζn=[𝐬𝐧𝐬𝐩𝐧𝐩][𝐩]n6ζn\zeta_{n}^{*}=[\mathbf{snspnp}][\mathbf{p}]^{n-6}\neq\zeta_{n}, and compare AneζnA_{n}e_{\zeta_{n}} with eζnAne_{\zeta_{n}}A_{n}. To do so, notice that eζnAn=(Aneζn)e_{\zeta_{n}}A_{n}=(A_{n}e_{\zeta_{n}^{*}})^{*}, so we make use of Theorem 5.4 and compare Δ(ζn)\Delta(\zeta_{n}) with (Δ(ζn))\big{(}\Delta(\zeta_{n}^{*})\big{)}^{*}. It is simple to check that Δ(ζn)=2(δ+1)ζn+\Delta(\zeta_{n})=2(\delta+1)\,\zeta_{n}+\dots and (Δ(ζn))=2δζn+\big{(}\Delta(\zeta_{n}^{*})\big{)}^{*}=2\delta\,\zeta_{n}+\dots, so Δ(ζn)(Δ(ζn))\Delta(\zeta_{n})\neq\big{(}\Delta(\zeta_{n}^{*})\big{)}^{*} for any δ\delta\in\mathbb{C}.

Let us prove the second part. For the point (i), note that bracelets which parametrise Cn(δ)/J(f)C_{n}(\delta)/J^{(f)} with f=1,2f=1,2 contain no more that one letter 𝐧\mathbf{n} (equivalently, 𝐬\mathbf{s}), which means that they are self-dual with respect to ()(\cdot)^{*}, so the quotient algebra is commutative. Now, for f3f\geqslant 3 we aim at comparing An(2)eζnA^{(2)}_{n}e_{\zeta_{n}} and eζnAn(2)e_{\zeta_{n}}A^{(2)}_{n}, where An(2)A^{(2)}_{n} is the double-arc generalisation of AnA_{n} defined in (4.31). With no analog of Theorem 5.4 available at this point, we analyse diagrammatic expressions with the aid of the following formula, which follows from the definition of the average (3.16) (see the proof of Theorem 5.4):

for any uCn(δ) and for any diagram bBn(δ) one hasuγb=γubandγbu=γbu.\text{for any $u\in C_{n}(\delta)$ and for any diagram $b\in B_{n}(\delta)$ one has}\quad\quad u\gamma_{b}=\gamma_{ub}\quad\text{and}\quad\gamma_{b}\,u=\gamma_{bu}\,. (B.22)

When applied to the products in question, one fixes a representative in eζne_{\zeta_{n}} and attaches two arcs to the lower/upper row of bb in all possible ways. By collecting only the basis vectors for ζn\zeta_{n} and ξn=[𝐧𝐬𝐩]2[𝐩]n6\xi_{n}=[\mathbf{nsp}]^{2}[\mathbf{p}]^{n-6} one can obtain

An(2)eζn=(δ2+2δ)eζn+(2+δ)eξn+,eζnAn(2)=(δ2+δ+2)eζn+2(δ+1)eξn+,A^{(2)}_{n}e_{\zeta_{n}}=(\delta^{2}+2\delta)\,e_{\zeta_{n}}+(2+\delta)\,e_{\xi_{n}}+\dots\,,\quad e_{\zeta_{n}}A^{(2)}_{n}=(\delta^{2}+\delta+2)\,e_{\zeta_{n}}+2(\delta+1)\,e_{\xi_{n}}+\dots\,, (B.23)

so An(2)eζneζnAn(2)A^{(2)}_{n}e_{\zeta_{n}}\neq e_{\zeta_{n}}A^{(2)}_{n} for any δ\delta\in\mathbb{C}.

For the point (ii), note that bracelets which parametrise Cn(δ)J(fmax)C_{n}(\delta)\cap J^{(f_{\mathrm{max}})} contains no more that one letter 𝐩\mathbf{p}, which means that they are self-adjoint, so the algebra is commutative. Now, for ffmax1f\leqslant f_{\mathrm{max}}-1 we consider ζ~n=[𝐧𝐬𝐧𝐩𝐬𝐩][𝐧𝐬]l[𝐩]r\tilde{\zeta}_{n}=[\mathbf{nsnpsp}][\mathbf{ns}]^{l}[\mathbf{p}]^{r} (with r=n2fmaxr=n-2f_{\mathrm{max}} and r+l+6=nr+l+6=n) and aim at comparing An(fmax1)eζ~nA^{(f_{\mathrm{max}}-1)}_{n}e_{\tilde{\zeta}_{n}} and eζ~nAn(fmax1)e_{\tilde{\zeta}_{n}}A^{(f_{\mathrm{max}}-1)}_{n}. Similarly to the previous case, we collect only the basis elements for ζ~n\tilde{\zeta}_{n} and ξ~n=[𝐧𝐬𝐩]2[𝐧𝐬]l[𝐩]r\tilde{\xi}_{n}=[\mathbf{nsp}]^{2}[\mathbf{ns}]^{l}[\mathbf{p}]^{r}, which gives

An(fmax1)eζ~n=δ(δ2+2δ+8)eζ~n+(δ2+2δ+8)eξ~n+,eζ~nAn(fmax1)=δ(δ2+δ+4)eζ~n+2(δ2+δ+2)eξ~n+,A^{(f_{\mathrm{max}}-1)}_{n}e_{\tilde{\zeta}_{n}}=\delta(\delta^{2}+2\delta+8)\,e_{\tilde{\zeta}_{n}}+(\delta^{2}+2\delta+8)\,e_{\tilde{\xi}_{n}}+\dots\,,\quad e_{\tilde{\zeta}_{n}}A^{(f_{\mathrm{max}}-1)}_{n}=\delta(\delta^{2}+\delta+4)\,e_{\tilde{\zeta}_{n}}+2(\delta^{2}+\delta+2)\,e_{\tilde{\xi}_{n}}+\dots\,, (B.24)

so An(fmax1)eζ~neζ~nAn(fmax1)A^{(f_{\mathrm{max}}-1)}_{n}e_{\tilde{\zeta}_{n}}\neq e_{\tilde{\zeta}_{n}}A^{(f_{\mathrm{max}}-1)}_{n} for any δ\delta\in\mathbb{C}. ∎

References